Title
stringlengths
12
257
Annotation
stringlengths
101
3.94k
PDF
stringlengths
38
45
Latex
stringlengths
1
261k
Title: Data storage, processing and visualisation for the ATCA
Abstract: We present three Virtual Observatory tools developed at the ATNF for the storage, processing and visualisation of ATCA data. These are the Australia Telescope Online Archive, a prototype data reduction pipeline, and the Remote Visualisation System. These tools were developed in the context of the Virtual Observatory and were intended to be both useful for astronomers and technology demonstrators. We discuss the design and implementation of these tools, as well as issues that should be considered when developing similar systems for future telescopes.
https://export.arxiv.org/pdf/astro-ph/0601354
\twocolumn[ \begin{changemargin}{.8cm}{.5cm} \begin{minipage}{.9\textwidth} \vspace{-1cm} \small{\bf Abstract:} We present three Virtual Observatory tools developed at the ATNF for the storage, processing and visualisation of ATCA data. These are the Australia Telescope Online Archive, a prototype data reduction pipeline, and the Remote Visualisation System. These tools were developed in the context of the Virtual Observatory and were intended to be both useful for astronomers and technology demonstrators. We discuss the design and implementation of these tools, as well as issues that should be considered when developing similar systems for future telescopes. \medskip{\bf Keywords:} astronomical data bases: miscellaneous --- methods: data analysis \medskip \medskip \end{minipage} \end{changemargin} ] \small \section{Introduction} The so-called data explosion in astronomy promises exciting new scientific developments, but brings with it many technical challenges, in collecting, storing, transporting, processing and visualising data. Virtual Observatories (VO) have developed to meet some of these technical challenges. Falling under the broad area of e-science (which incorporates other scientific domains facing similar challenges, such as genetics and particle physics) the aim of Virtual Observatory research is to provide the tools necessary for dealing with this data. The Australian Virtual Observatory (Aus-VO)\footnote{{\tt http://www.aus-vo.org}} was started in 2003 with the aim of both contributing to the international VO effort, and developing tools of use to Australian astronomers. Australia has many areas of expertise (for example radio astronomy) and it makes sense to focus our efforts on providing modern tools for working in these areas. In this context the ATNF decided to develop a range of tools for storing, processing and visualising data from the Australia Telescope Compact Array. The aim was that these tools would be useful to astronomers now, and at the same time let us explore the technology that would be necessary for developing software for future telescopes such as the Square Kilometre Array (SKA). This paper is based on a talk given at the ASA Annual Meeting in Sydney in July 2005. After giving some background about the ATCA and the Virtual Observatory, we discuss three tools developed at the ATNF over the last few years. Firstly the Australia Telescope Online Archive which contains all of the data collected so far by the ATCA; secondly a prototype data reduction pipeline for ATCA data; and finally the Remote Visualisation System for viewing large datasets. \section{The ATCA} The Australia Telescope Compact Array (ATCA) is an east-west earth-rotation synthesis interferometer, with six 22 m antennas on a 6 km baseline. It has been in operation at Narrabri since 1990. The telescope can observe at 6 bands with wavelengths 20 cm, 12 cm, 6 cm, 3 cm, 1 cm and 3 mm. Each antenna observes two frequencies simultaneously. There are six bandwidths available on Frequency 1 (128, 64, 32, 16, 8 and 4 MHz) and two bandwidths on Frequency 2 (128 and 64 MHz). The telescope produces $\sim 0.5$ GB of raw data per day, and this is likely to increase significantly with future telescope upgrades. In the rest of this section we outline the existing systems for archiving, processing and visualising ATCA data. \subsection{Data archiving}\label{s_da} Since the commencement of operation of the ATCA in June 1990, a complete record of all data observed from the telescope has been maintained offline at the telescope site, mostly recorded on CD. In conjunction with this, the ATNF maintained a record of the project proposals for observations on the telescope --– the {\it Projects database} --- and a short form of the observation parameters for each day’s observing --– the {\it Positions database}. After a proprietary period of 18 months, in which the observing team has sole access to the data obtained in an observation, the data is made publicly available. Astronomers can search for observations on the ATNF webpage, submit the details of the observation data required via e-mail and have a CD containing the data prepared for them at nominal cost. \subsection{Data processing} ATCA Data processing (reduction) is generally performed with one of the standard radio data reduction packages; most commonly Miriad \citep{sault95}, but also AIPS\footnote{Astronomical Image Processing System (AIPS), http://www.cv.nrao.edu/aips.} and AIPS++\footnote{Astronomical Image Processing System (AIPS++), http://aips2.nrao.edu.}. After loading, editing and calibrating the data, the resulting product is an intensity map referred to as the dirty image. At this stage a deconvolution algorithm, usually a variant of {\sc clean} \citep{hogbom74}, is required to produce the final image. At each stage in the process there are a range of parameters that can be set to control the type of processing performed. Both general parameters (such as calibration strategy, {\sc clean} method and type of data editing) as well as fine-grained parameters (such as calibration solution interval, number of {\sc clean} iterations and median filter size) need to be modified to obtain the best results. Hence data processing is typically a highly interactive process. There is an existing system operating at the telescope {\tt CAONIS} designed for on-the-fly imaging of ATCA data. However the design of this system makes it difficult to port to current Linux systems. \subsection{Data visualisation} The final images from the ATCA are usually visualised using tools such as Miriad or {\tt kvis} \citep{gooch96}. These are well established tools that cover many of the visualisation requirements of ATCA observers. As mentioned in the previous section, the {\tt CAONIS} system which runs at Narrabri also allows basic visualisation of images. \section{The Virtual Observatory} The umbrella organisation for Virtual Observatory work is the International Virtual Observatory Alliance (IVOA). The IVOA was formed in June 2002 with a mission to \begin{quotation}{\it \noindent facilitate the international coordination and collaboration necessary for the development and deployment of the tools, systems and organizational structures necessary to enable the international utilization of astronomical archives as an integrated and interoperating virtual observatory.} \end{quotation} The IVOA is a collaboration between over 15 member countries including Australia. The focus so far has been developing the standards required for interoperability between software developed and data produced in all areas of astronomy. Another significant aim is to develop the infrastructure required (networks and organisations) for the large scale storage and distribution of astronomical data. The IVOA working groups address a range of issues such as grid and web services, data modelling and standards for the data access. There are also four interest groups \begin{itemize} \item Applications IG \item Astronomy Grid IG \item Data Curation IG \item Theory IG \end{itemize} which focus on the requirements of particular application domains. The aim of the Australian Virtual Observatory (Aus-VO) is to provide distributed, uniform interfaces to the data archives of Australia's major observatories and the archives of simulation data. Aus-VO is a collaboration between many Australian institutions, including the Universities of Melbourne, Sydney, New South Wales and Queensland, Monash University, Swinburne University of Technology, the Australian National University and Mount Stromlo Observatory, the Victorian Partnership for Advanced Computing, the ATNF and the AAO. There are a range of VO projects underway in Australia, including the development of data archives and software for HIPASS \citep{meyer04}, RAVE \citep{siebert04}, 2QZ \citep{croom04} and SUMSS \citep{bock99}. The initial focus of most of these projects has been to make data from Australian projects widely available within the international community, in a VO compliant format. In addition there are several projects investigating novel methods for astronomical data mining and data analysis, for example \citet{rohde05} apply machine learning techniques to catalogue crossmatching. The Melbourne University group has also been setting up infrastructure such as a registry for Australian web services and data archives. \section{The Australia Telescope \\ Online Archive} In June 2003, a joint project between the ATNF and the CSIRO ICT Centre was commenced to make the ATNF archive data available online as the Australia Telescope Online Archive (ATOA). This was planned as a new data resource for astronomers, as well as the foundation for the development of online data processing systems to make the raw data more accessible to non-expert users (see Section \ref{s_pipe}). The construction of the ATOA required the copying of the offline archive (at the time, $\sim 2700$ CDs, totalling $\sim 1.7$ TB) from the telescope site to Canberra where the online archive was to be developed, creating a meta-database describing the data, and making a web front-end to search and download the data. The database consists of two parts. The first is the raw data from the telescope (RPFITS files) which is stored as normal files on the host system. In addition there is a relational database which stores all the metadata (discussed in Section \ref{s_meta}). The `vital statistics' of the ATOA are shown in Table \ref{t_atoa}. The current rate of growth of the archive is $\sim 0.5$ Gb/day. However this is likely to increase significantly in the future as new instruments come online. To maintain an growing archive (rather than a static one) it is necessary to ensure the RPFITS files are stored in a readily accessible way (currently on a RAID system) that is easily distributed over a number of drives. Also, that the database itself is easy to update in a robust manner. The ATOA was made publicly available in December 2004 and can be accessed from {\tt http://atoa.atnf.csiro.au}. \begin{table}[ht] \begin{center} \caption{ATOA Statistics}\label{t_atoa} \begin{tabular}{lr} \hline Projects & 2261 \\ Files & 57147 \\ Sources & 128111 \\ Metadata size & $\sim 4$ Gb \\ RPFITS data size & $\sim 2$ Tb \\ Growth rate & $\sim 0.5$ Gb/day \\ \hline \end{tabular} \end{center} \medskip \end{table} \subsection{Metadata}\label{s_meta} Metadata is simply data which describes other data, for example the project code or the name of the primary calibrator. The meta-database for the ATOA consists of three main parts; the contents of the original ATNF online Projects database, metadata describing the observation that is extracted directly from the raw data files produced by the telescope’s software, and metadata inferred from all of the available data sources to assist in the automation of reducing the telescope’s raw data to images. The types of metadata used in the ATOA are summarised in Table \ref{t_meta}. \begin{table}[ht] \begin{center} \caption{Metadata in the ATOA}\label{t_meta} \begin{tabular}{ll} \hline Metadata source & Examples \\ \hline Projects database & proposal; observer name\\ & country; institution \\ Positions database$^*$ & source names \& positions \\ & observing band; receivers \\ RPFITS files & scans; polarisations \\ & array configuration \\ Inferred & calibrator names \& roles\\ & calibrator--target matches \\ \hline \end{tabular} \end{center} $^*$ The positions database is included in our data model, and some of the metadata is used to reconstruct the observation metadata. However, it is not actually loaded into the ATOA. \medskip\\ \end{table} Most of the metadata available in the ATNF Positions database is also available from the metadata in the raw data files, and is finer-grained, since the Positions data is a daily summary, while the file metadata is available for each telescope pointing. The inferred metadata in the ATOA is `value added' information that is automatically determined from the existing metadata, for example the calibration role of each source (primary calibrator, secondary calibrator, target, etc). This is discussed further in Section \ref{s_roles}. \subsection{A data model for the ATCA}\label{s_dm} A data model is a comprehensive scheme describing how data is to be represented, for manipulation by humans or computer programs. Data models are critical for planning how data will be organised within a database as they describe all the relationships between the different entities. A section of our data model for the ATCA is shown in Figure \ref{f_dm}. We now briefly explain the UML (Unified Modelling Language) notation used in the data model. Each box contains an entity (e.g. {\tt Scan}) that has been identified in the metadata. Each entity has attributes (e.g. {\tt restFreq}), each of which are of a specified data type (e.g. {\tt float}). Associated entities are connected to each other with lines, which also specify the cardinality of the relationship. For example \\ \begin{minipage}{8cm} \begin{center} \resizebox{7cm}{!}{\includegraphics{uml.eps}} \end{center} \end{minipage} should be read as {\bf ``A scan has 0 or more spectral windows. A spectral windows has 1 or more scans.''} The development of a data model that covers the whole of astronomy is an ongoing project within the international VO community. We have contributed this data model to the IVOA Data Model WG as an example of a data model for radio astronomy. For more information on this topic, see the IVOA Data Modelling website\footnote{{\tt http://www.ivoa.net/twiki/bin/view/IVOA/IvoaDataModel}}. The ATOA archive database structure is created directly from the definitions in the ATOA data model. Parts of the data model contain information for specific database implementations so that all of the implementation-specific parts of the database creation are handled in this process. The data model in Figure \ref{f_dm} corresponds to the part of the data model that describes the metadata contained directly in the archive RPFITS files. The data model for the inferred data is available from the ATOA web pages\footnote{{\tt http://www.atnf.csiro.au/computing/web/atoa/implementation.html}}. \subsection{Implementation} The ATOA web interface was implemented as a Java (ver. 1.4.2)\footnote{{\tt http://java.sun.com}} application and is hosted using the Apache Tomcat (ver. 5.0.28)\footnote{{\tt http://tomcat.apache.org}} web container. Relational database services are provided by an Oracle 9i instance running on the same machine. A web based interface was chosen so as to maximise interoperability and provide easy access to users. For example, RPFITS files may simply be downloaded by constructing a suitable URL for the ATOA file server. This allows files on the server to be downloaded by a Web browser, by command-line programs that allow users to fetch the data referred to by a URL, or by application programs using libraries that allow a URL to be opened in a similar way to a file on a local file system. The user interface centres around two main web pages: the query page which allows users to specify criteria for selecting RPFITS files from the archive, and a results page which provides the means for users to inspect the metadata of matching files and download particular files if desired. The results page initially presents the user with a broad, global view of the query results in tabular form listing details such as file name, file size, principal investigator and array configuration. The user may also interactively `drill-down' for a more detailed view of any file in the list. RPFITS files can be downloaded individually or in batches. As mentioned in Section \ref{s_da} ATCA data has a proprietary period of 18 months, in which it is only accessible to members of the project team. Authorised access is supported for data within the proprietary access period. If a user wishes to access proprietary data they must first go through a manually verified authentication process after which a password is issued to the Principal Investigator for that project. In the future we plan to replace this authentication and authorisation method with a streamlined system that links the new ATNF proposal system, OPAL\footnote{{\tt http://opal.atnf.csiro.au/}}, and its authentication database to the ATOA. Users will then be given access to proprietary data by using their OPAL credentials based on the projects they are associated with in the OPAL system The ATOA web server and database are hosted on a Dell PowerEdge 750 running Debian Linux 3.0. The host has a 2.8GHz Pentium 4 processor, 2GB of RAM and is attached to a 3 terabyte Apple Xserve RAID for archive storage. \section{A data processing pipeline framework}\label{s_pipe} The data products in radio astronomy are often less accessible to the non-expert than those in other domains such as optical astronomy. It requires a reasonably high level of domain expertise to process the raw data and produce an image. Obviously for carrying out detailed scientific analysis it would be necessary to develop this expertise, or collaborate with a radio astronomer. However in an era of multiwavelength astronomy, astronomers expect to download and compare data from a variety of telescopes, at a variety of wavelengths. With this in mind we have developed an automatic pipeline for people who want to quickly inspect the data in the ATOA, to see if it was suitable for further processing. One of the aims of this project was to test the viability of `driving' the pipeline using the metadata discussed in Section \ref{s_meta}. In other words the pipeline should make decisions about what kind of processing to do --- both on a general (e.g. continuum/spectral line) and specific level (e.g. number of {\sc clean} iterations). In this section we discuss the development of extra metadata required to driving the pipeline, in particular the calibration process. We then outline our prototype pipeline which can process single pointing continuum data from the ATOA and is available for testing at {\tt http://atoa.atnf.csiro.au/test}. \subsection{Metadata for automatic \\ processing}\label{s_pmeta} The metadata in the Project and Position databases, while providing information about which astronomical sources have been recorded in an observing session, does not (in general) provide any information about the role that the observer intended the source to play in the observation (eg. primary calibrator, target source). This would be relatively easy to record in a new system, but as we are dealing with existing data we had to infer the roles of sources. Another problem for automatic processing is the grouping of data into valid `observations'. An expert would typically choose an appropriate subset of files from the archive to image. However, a non-radio astronomer may choose an subset that contains files that should be imaged separately, or files that contain data that should be ignored entirely. Although it is impossible to deal with all cases, our aim was to have the pipeline group together the selected data in such a way that an image could be made in at least $80\%$ of cases. A wide range of observation types can be recognised and characterised using the meta-database but are not yet processed by the prototype pipeline (e.g. millimetre and spectral line observations). In the following section we discuss how we assign the source roles within an observation, and the algorithm we used to match target sources with the appropriate calibrators. \subsection{Determining source roles}\label{s_roles} While matching target sources with their calibrators would be straightforward for an astronomer it is a challenge for an automatic system. In a typical (simple) observing session the primary calibrator is recorded for a short period at the start or end of the observing session; and alternating pointings are made to the secondary, and the source of interest, or target. However there is a great variety of different ways that the observer can choose to structure their observations. If an observation contains more than one target, the targets may share, or have distinct, secondary calibrators, depending on their separation in the sky. There may be several secondary calibrators for each target, and the same source may be used for primary and secondary calibration. In addition, some observers use secondary calibrators that are not in the list of recommended calibrators, and that list has itself changed over time. In order to classify the sources in an observing session the following metadata is used \begin{itemize} \item The locations and names of sources extracted from the raw telescope data \item The times and durations of the source pointings \item The names and locations of the four primary calibrators commonly used at the ATCA \item A recent ATCA catalogue of recommended secondary calibrators \item Names of sources extracted from project titles \item A pre-assembled list of possible calibrator sources \end{itemize} Once the source roles have been determined, the proximity in the sky and the proximity in observation time of the targets and their secondary calibrators are used to match targets with their respective calibrator(s). For each target pointing, a weight is calculated for each secondary calibration pointing made within two hours of observation of the target pointing: \begin{displaymath} w_{t,s} = \Sigma_{P} \Sigma_{S} e^\frac{-(3a/a_{max})^2}{2} e^\frac{-(3\Delta t / \Delta t_{max})^2}{2} \end{displaymath} \begin{displaymath} \Delta t_{t,s} < \Delta t_{max} \end{displaymath} where $S$ is the set of candidate secondary calibrators, $a$ is the angular separation between the target and the secondary calibrator, $a_{max}$ is the maximum desirable separation between the target and the secondary calibrator (and is a function of the observing frequency band). $\Delta t$ is the separation of the time midpoints of the target and calibrator pointings and $\Delta t_{max}$ is the maximum desirable time separation (two hours). The summation is over all pointings at a target ($\Sigma_P$) and all secondary calibrators within two hours of a target pointing ($\Sigma_S$). The $w_{t,s}$ are used to select suitable secondary calibrators for the respective targets from the calibrators whose $w_{t,s}$ weights dominate for a particular target. This procedure constructs the metadata required for continuum imaging at centimetre wavelengths. The algorithm works well in general, but there are some problematic cases, for example where the target is a source from the secondary calibrator catalogue. \subsection{Implementation} The underlying processing of the ATCA data is carried out using the Glish scripting language in AIPS++. The ATOA imaging Web Services interface was constructed using the Apache Axis tools (ver. 1.1)\footnote{{\tt http://ws.apache.org/axis}}, and interfaces to the processing scripts through a Perl (ver. 5.4.8)\footnote{{\tt http://www.perl.com}} script that deals with the control of the execution of the Glish scripts. The pipeline client is written using Python (ver. 2.3)\footnote{{\tt http://python.org}}, and the SOAPpy web services tools (ver. 0.11.3)\footnote{{\tt http://pywebsvcs.sourceforge.net}}. There were some minor, but difficult to find, problems in interoperation between the SOAPpy tools and Apache Axis; the data structures used in the web services calls are possibly more complicated than had been previously used between the two web services implementations. Documentation in both was not as informative or complete as it might have been. The pipeline web services can be configured to run directly on the server host, or be directed to run on other machines through a batch queuing system, since some stages in the pipeline can run for several CPU minutes. We used the OpenPBS Batch Queuing System (ver. 2.3)\footnote{{\tt http://www.openpbs.org}} for queue management, but unfortunately it has no mechanism for reporting job completion to another program. After processing for a web service completes, the batch job doing the processing sends a completion message to the program invoked by the web service that controls the execution of the processing for the service. However, at this point, the batch processing system has not yet transferred the job's output data back to the pipeline server. The control program then polls the PBS batch queue at five second intervals to ensure that the batch job has completed. The raw data from the ATOA, all the intermediate files from the data processing, the log files, and the resulting images are stored temporarily on the pipeline server. The first web service call made by a pipeline client reserves a private location for storage, and requests a lifetime for the storage. The pipeline server has a configurable maximum lifetime, and the stored data will be deleted after this time expires. Only clients who have the name of the storage area (a randomly generated string) can access it. There is no quota on the storage use of any individual temporary storage area. However, a quota may be imposed on the total amount of storage available to all active storage areas. The ATOA and pipeline web services return a URL for the generated images to the end-user's system. This allows the URL to be passed on to the Remote Visualisation System (see Section \ref{s_rvs}) image viewing system so that the image can be viewed online while it is still resident on the pipeline server. Figure \ref{f_arch} show the overall system architecture, in particular how the ATOA, pipeline and RVS interact. \section{The Remote Visualisation \\ System}\label{s_rvs} The Remote Visualisation System (RVS) was designed to enable visualisation of and interaction with large astronomical images in the context of the VO. As opposed to other VO image displays, such as CDS Aladin \citep{fernique04}, RVS does not require the user to download the data to the client machine. Furthermore it provides rendering of image cubes, such as spectral-line cubes created from ATCA data. The RVS server accepts FITS images - which can be compressed - through local {\tt file://} URLs and remote http or ftp access. The data should be co-located with or at least be available to the server on high bandwidth connection, while it places no such requirements on the client. Only minimal data transfer to the client is necessary and this is independent of the size of the source data set. The server-side architecture is distributed to enable workload sharing and extensibility. RVS makes use of several software components: CORBA to make it distributed, AIPS++ as the image rendering component and Java for the web services and client. The architecture of the RVS system is shown in Figure \ref{rvs_arch}. RVS is exposed through a web service interface using the standard Web Service Description Language (WSDL ver. 1.1)\footnote{{\tt http://www.w3.org/TR/wsdl}}. This can easily be integrated into custom applications. Several client applications make use of the web service interface; the RVSViewer - a traditional image viewer, a thumbnail service - providing preview images and a session viewer. The session viewer connects to an existing RVSViewer via a key. Multiple instances can be run at the same time, making it a possible to use it as a conferencing tool where people can observe and interact with the data. The ATOA pipeline re-uses the existing RVSViewer client by passing it the file location of the output image. RVS is not specific to the ATOA or prototype pipeline and there are plans to use it for all ATCA archives. It has been successfully tested on images and data cubes from various surveys and has good performance on large datasets. For example a 1.5 Gb data cube from the Galactic All-Sky Survey (GASS) \citep{mcclure05} takes about one minute to load. Compare this with downloading the full cube from say the U.S. to Australia which could take $\sim 1-2$ hours. For more information and direct access to RVS, see {\tt http://www.atnf.csiro.au/vo/rvs/}. \section{Discussion} The ATOA has been public since December 2004. We hope that it will encourage the reuse of ATCA data for projects other than those it was originally intended for. The framework used for the ATOA could easily be extended to include data from other telescopes and can be updated as additional metadata is required. The most significant improvement of the ATOA over existing online archives (such as NRAO\footnote{{\tt http://archive.nrao.edu/archive/e2earchive.jsp}} and MAST\footnote{{\tt http://archive.stsci.edu/}}) is the data delivery mechanism. Most existing archives do not support on demand delivery of data over the web, instead requiring the user to submit a form requesting files that then have to be transfered to a publicly accessible ftp site or to other media (such as CD) for physical delivery. In the ATOA, the batch downloading of multiple files is handled by a streaming TAR or ZIP archiving algorithm that performs dynamic archiving as files are streamed over the web, requiring no additional disk space on the server for these operations. In developing the ATCA data model and considering the type of metadata required for automatic processing we identified several new metadata types that would be useful to store in the RPFITS files. As a result the following fields have been added to the RPFITS files and will be available in all future ATCA data: \begin{itemize} \item four calibrator codes \begin{description} \item[C] (standard phase calibrator) \item[F] (primary flux calibrator) \item[B] (bandpass calibrator) \item[P] (pointing calibrator) \end{description} \item Pointing offsets \item Weather data: added rain gauge and phase rms and difference \item Attenuator settings at start of scan \item Subreflector position \item Correlator configuration \item Scan type \item Coordinate type \item Line mode \item CACAL counter \end{itemize} These will help both automatic processing systems and astronomers assess the data quality in the observations they are interested in. A full e-logbook system will be used in the future as currently the logs are all stored on paper at the telescope and hence are not easily accessible to ATOA users. We have developed a prototype pipeline for processing of raw data for single-pointing continuum images. This is attached to the ATOA to provide an improved service for users of the ATOA. At this stage the image quality is suitable for previewing the data in archive to see if it is of interest. Further manual processing would then be required to obtain images of scientific quality. A significant challenge in developing the ATOA and the prototype pipeline were integrating pre-existing software with modern software tools. For example, the Glish scripting language has no web service libraries and so an extra layer had to be developed between the data processing level and the web services. If re-implementing from scratch, a language such as Python would be a better alternative for developing the pipeline. In developing these tools we have started to explore the techniques necessary for astronomical software development in the VO era. This is essential for future telescopes and surveys that Australia will produce. Making access to existing Australian data as easy as possible will maximise its use in the international community. \section*{Acknowledgments} The authors would like to acknowledge the software development done on the RVS project, primarily by Anil Chandra and also by Praveena Tokachichu. The ATNF side of the prototype pipeline and ATOA development was managed by Neil Killeen and Jessica Chapman. Vince McIntyre contributed extensively to all three projects, in particular in setting up the hardware required. A number of ATNF staff put in significant effort to get the ATOA set up, in particular Robin Wark, Bob Sault and Mark Wieringa. Warwick Wilson and Mark Wieringa implemented the changes to add extra metadata to the RPFITS files. From the ICT Centre, Robert Power made the initial data model designs, the ATOA data loader software and ATOA query front end. Geoff Squire and Bella Robinson made significant contributions to the prototype pipeline.
Title: The Araucaria Project. The Distance to the Local Group Galaxy IC 1613 from Near-Infrared Photometry of Cepheid Variables
Abstract: We have measured accurate near-infrared magnitudes in the J and K bands of 39 Cepheid variables in IC 1613 with well-determined periods and optical VI light curves. Using the template light curve approach of Soszy{\'n}ski, Gieren and Pietrzy{\'n}ski, accurate mean magnitudes were obtained from these data which allowed to determine the distance to IC 1613 relative to the LMC from a multiwavelength period-luminosity solution in the optical VI and near-IR JK bands, with an unprecedented accuracy. Our result for the IC 1613 distance is $(m-M)_{0} = 24.291 \pm 0.014$ (random error) mag, with an additional systematic uncertainty smaller than 2%. From our multiwavelength approach, we find for the total (average) reddening to the IC 1613 Cepheids $E(B-V) = 0.090 \pm 0.007$ mag,which is significantly higher than the foreground reddening of about 0.03 mag,showing the presence of appreciable dust extinction inside the galaxy. Our data suggest that the extinction law in IC 1613 is very similar to the galactic one.Our distance result agrees, within the uncertainties, with two earlier infrared Cepheid studies in this galaxy of Macri et al. (from HST data on 4 Cepheids), and McAlary et al. (from ground-based H-band photometry of 10 Cepheids), but our result has reduced the total uncertainty on the distance to IC 1613 (relative to the LMC) to less than 3%. With distances to nearby galaxies from Cepheid infrared photometry at this level of accuracy, which are currently being obtained in our Araucaria Project, it seems possible to significantly reduce the systematic uncertainty of the Hubble constant as derived from the HST Key Project approach, by improving the calibration of the metallicity effect on PL relation zero points, and by improving the distance determination to the LMC.
https://export.arxiv.org/pdf/astro-ph/0601309
\newcommand{\up}[1]{\ifmmode^{\rm #1}\else$^{\rm #1}$\fi} \newcommand{\zdot}{\makebox[0pt][l]{.}} \newcommand{\upd}{\up{d}} \newcommand{\uph}{\up{h}} \newcommand{\upm}{\up{m}} \newcommand{\ups}{\up{s}} \newcommand{\arcd}{\ifmmode^{\circ}\else$^{\circ}$\fi} \newcommand{\arcm}{\ifmmode{'}\else$'$\fi} \newcommand{\arcs}{\ifmmode{''}\else$''$\fi} \title{The Araucaria Project. The Distance to the Local Group Galaxy IC 1613 from Near-Infrared Photometry of Cepheid Variables. \footnote{Based on observations obtained with the NNT telescope at ESO/La Silla for programs 074.D-0318(B) and 074.D-0505(B) } } \author{Grzegorz Pietrzy{\'n}ski} \affil{Universidad de Concepci{\'o}n, Departamento de Fisica, Astronomy Group, Casilla 160-C, Concepci{\'o}n, Chile} \affil{Warsaw University Observatory, Al. Ujazdowskie 4, 00-478, Warsaw, Poland} \authoremail{pietrzyn@hubble.cfm.udec.cl} \author{Wolfgang Gieren} \affil{Universidad de Concepci{\'o}n, Departamento de Fisica, Astronomy Group, Casilla 160-C, Concepci{\'o}n, Chile} \authoremail{wgieren@astro-udec.cl} \author{Igor Soszy{\'n}ski} \affil{Universidad de Concepci{\'o}n, Departamento de Fisica, Astronomy Group, Casilla 160-C, Concepci{\'o}n, Chile} \affil{Warsaw University Observatory, Al. Ujazdowskie 4, 00-478, Warsaw, Poland} \authoremail{soszynsk@astro-udec.cl} \author{Fabio Bresolin} \affil{Institute for Astronomy, University of Hawaii at Manoa, 2680 Woodlawn Drive, Honolulu HI 96822, USA} \authoremail{bresolin@ifa.hawaii.edu} \author{Rolf-Peter Kudritzki} \affil{Institute for Astronomy, University of Hawaii at Manoa, 2680 Woodlawn Drive, Honolulu HI 96822, USA} \authoremail{kud@ifa.hawaii.edu} \author{Massimo Dall'Ora} \affil{INAF-Osservatorio Astronomico di Capodimonte, via Moiariello 16, I-80131 Naples, Italy} \authoremail{dallora@na.astro.it} \author{Jesper Storm} \affil{Astrophysikalisches Institut Potsdam, An der Sternwarte 16, D-14482 Potsdam, Germany} \authoremail{jstorm@aip.de} \author{Giuseppe Bono} \affil{INAF-Osservatorio Astronomico di Roma, via Frascati 33, 00040 Monte Porzio Catone, Italy} \authoremail{bono@mporzio.astro.it} \keywords{distance scale - galaxies: distances and redshifts - galaxies: individual(IC 1613) - stars: Cepheids - infrared photometry} \section{Introduction} Cepheid variables are the most important standard candles to calibrate the first rungs of the extragalactic distance ladder, out to some 30 Mpc. As young stars, Cepheids tend to lie in dusty regions in their spiral or irregular parent galaxies. As a consequence, Cepheid distances derived from the period-luminosity (PL) relation in optical photometric bands are quite sensitive to a precise knowledge of the total reddening, foreground and intrinsic, of their parent galaxy. While the galactic foreground reddening towards any direction in the sky is usually well established, particularly in directions far away from the galactic equator, the correct assessment of the reddening produced by dust extinction {\it intrinsic to the host galaxy} is usually a difficult task, and in most work on Cepheid distances based on optical data such an intrinsic contribution to the reddening has simply been ignored. Just for this one particular reason, it is clear that more accurate Cepheid distances to galaxies can be derived in near-infrared passbands, where dust absorption is small as compared to visual wavelengths, and the distance results become increasingly independent of errors in the assumed total reddenings. Efforts along these lines have started in the early eighties with the pioneering work of McGonegal et al. (1982), and Welch et al. (1985). Yet, an important obstacle to carry out accurate Cepheid distance work in the infrared has been, until very recently, the lack of well-calibrated fiducial PL relations in the near-infrared JHK bands. This problem has now been solved by the work of Persson et al. (2004) who provided such well-calibrated relations for the LMC. Very recently, Gieren et al. (2005a) have also provided well-calibrated PL relations in the JHK bands for Milky Way Cepheids, which agree with the corresponding LMC relations when an improved version of their infrared surface brightness technique (Gieren, Fouqu{\'e} \& G{\'o}mez, 1997, 1998) is used. In the {\it Araucaria Project}, started by our group some time ago (Gieren et al. 2005c), we have conducted surveys for Cepheid variables in a number of galaxies in the Local Group, and in the more distant Sculptor Group in order to investigate the effect of environmental properties on the PL relation, and to improve the accuracy of Cepheids as distance indicators. While we are discovering Cepheids in optical photometric bands, where these stars are rather easy to detect due to their relatively large amplitudes and typical light curve shapes (e.g. Pietrzynski et al. 2002, 2004), the main goal of the program is to undertake near-infrared follow-up imaging of selected subsamples of Cepheids in our target galaxies to obtain accurate reddening information, and thus to obtain more accurate distances than what is possible from optical (VI) data alone. Near-infrared PL relations from such Cepheids with existing information on their periods, and V and/or I light curves can be obtained very economically because it is possible to obtain accurate mean JHK magnitudes for these stars from just one single-phase observation from the template light curve approach of Soszynski, Gieren and Pietrzynski (2005). The success of this approach was recently demonstrated in the case of the Sculptor galaxy NGC 300 (Gieren et al. 2005b). For this galaxy, a combination of the PL relations obtained in the optical VI and infrared JK bands has allowed to determine a distance which is practically unaffected by any remaining uncertainty on reddening. It has also been shown in that paper that from the combined optical/near-infrared approach a total uncertainty as small as 3 \% can be obtained for the Cepheid distance (as measured relative to the LMC) for such a relatively nearby (2 Mpc) galaxy. In the present paper, we are applying the same approach to the Local Group dwarf irregular galaxy IC 1613. IC 1613 is a very important galaxy in the Araucaria Project because of the very low metallicity of its young stellar population close to -1.0 dex (Skillman et al. 2003), making it the lowest-metallicity galaxy in our sample. It is therefore a key object in our effort to determine the effect of metallicity on the Cepheid PL relation, and on other stellar distance indicators, like blue supergiant stars (Kudritzki et al. 2003). A first survey for Cepheid variables in IC 1613 was carried out by Sandage (1971) who used photographic images previously obtained by Baade. Modern work on the Cepheid PL relation in IC 1613, in the optical V and I bands, has been carried out by the OGLE Project (Udalski et al. 2001) who has discovered many new, previously unknown Cepheids in this irregular galaxy. More recently, Antonello et al. (2006) have extended this work to the B and R bands. From the work of the OGLE group, it could be established that the slope of the PL relation in optical bands is identical to the slope observed for the more metal-rich LMC Cepheids, arguing for a metallicity- independent slope of the PL relation. In the near-infrared, a pioneering paper on the distance of IC 1613 from H-band photometry of 10 Cepheids was published by McAlary et al. (1984) already 20 years ago; however, the uncertainty on this distance result was rather large due to the technical difficulties to obtain accurate IR photometry at those times, and the lack of an accurate calibrating PL relation. Much more recently, Macri et al. (2001) determined a near-infrared Cepheid distance to IC 1613 from H-band photometry of four variables obtained with HST/NICMOS. The accuracy of this determination suffers, however, from the very small number of stars used in the PL solution. A main goal of the present study was to derive {\it truly accurate near-infrared PL relations for IC 1613}, based on a large number of well-observed and well-selected Cepheids (see section 3.2.), and this way reduce the current uncertainty on the distance to IC 1613 to the very small level of 3-5\% we have achieved in our previous study of NGC 300. We have organized this paper in the following way: in section 2, we describe the observations, reductions and calibration of our data; in section 3, we present the calibrated infrared mean magnitudes of the Cepheids in our selected fields in IC 1613, and determine the distance and reddening; in section 4, we discuss our results; and in section 5, we summarize the main results of this work, and present some conclusions. \section{Observations, Data Reduction and Calibration} \subsection{Optical data} Our infrared observations of IC 1613 (see next section) were obtained about four years after the OGLE-II optical observations (Udalski et al. 2001) of this galaxy. This long gap in time made it necessary to improve the periods of the Cepheids in order to calculate accurate $<K>$ and $<J>$ mean magnitudes from single-phase infrared observations with the method of Soszynski et al. (2005). For this purpose, three new V-band observations of IC 1613 with the 1.3 m Warsaw telescope located at Las Campanas Observatory were secured in September 2005. This telescope is equipped with a mosaic 8k x 8k detector, with a field of view of 35 x 35 arcmin and a scale of 0.25 arcsec/pixel. Preliminary data reductions (i.e. debiasing and flatfielding) were done with the IRAF \footnote{IRAF is distributed by the National Optical Astronomy Observatories, which are operated by the Association of Universities for Research in Astronomy, Inc., under cooperative agreement with the NSF.} package. The point-spread function photometry was obtained with the DAOPHOT and ALLSTAR programs, in an identical way as described in Pietrzy{\'n}ski et al. (2002). Our photometry was then transformed to the standard system using the OGLE-II list of carefully calibrated stars in this galaxy (Udalski et al. 2001). \subsection{Infrared data} The near-infrared data presented in this paper were collected with the ESO NTT telescope on La Silla, equipped with the SOFI infrared camera. In the setup we used (Large Field), the field of view was 4.9 x 4.9 arcmin, with a scale of 0.288 arcsec / pixel. The gain and readout noise were 5.4 e / ADU and 0.4 e, respectively. The data were obtained under two observational programs: 074.D-0318(B), 074.D-0505(B) (PI: Pietrzy{\'n}ski) as part of the Araucaria Project. Alltogether, six different, slightly overlapping fields were observed through J and Ks filters. Their location is shown in Figure 1, and the equatorial coordinates of their centers are given in Table 1. Single deep J and Ks observations of our six fields were obtained under excellent seeing conditions during three different photometric nights. On these nights, we also observed a large number of photometric standard stars from the UKIRT system (Hawarden et al. 2001). In order to account for the rapid sky level variations in the infrared domain, the observations were performed with a dithering technique. In the Ks filter, we obtained six consecutive 10 s integrations (DITs) at a given sky position and then moved the telescope by about 20 arcsec to a different random position. Integrations obtained at 65 different dithering positions resulted in a total net exposure time of 65 minutes in this filter, for a given field. In the case of the J filter, in which the sky level variations are less pronounced than in K, two consecutive 20 s exposures were obtained at each of 25 dithering positions, which corresponded to a total net exposure of about 17 minutes, for any given field. Sky subtraction was performed by using a two-step process implying the masking of stars with the XDIMSUM IRAF package in an analogous manner as described in Pietrzy{\'n}ski and Gieren (2002). Then the single images were flatfielded and stacked into the final images. PSF photometry was obtained using DAOPHOT and ALLSTAR, following the procedure described in Pietrzy{\'n}ski, Gieren, and Udalski (2002). In order to derive the aperture corrections for each frame, about 7-10 relatively isolated and bright stars were selected, and all neighbouring stars were removed using an iterative procedure. Finally, we measured the aperture magnitudes for the selected stars with the DAOPHOT program using apertures of 16 pixels. The median from the differences between the aperture magnitudes obtained this way, and the corresponding PSF magnitudes, averaged over all selected stars was finally adopted as the aperture correction for a given frame. The rms scatter from all measurements was always smaller than 0.02 mag. In order to accurately transform our data to the standard system a large number (between 8 and 15) of standard stars from the UKIRT system (Hawarden et al. 2001) was observed under photometric conditions at a variety of airmasses, together with our six science fields. The standard stars were chosen to have colors bracketting the colors of the Cepheids in IC 1613. The aperture photometry for our standard stars was performed with DAOPHOT using the same aperture as for the calculation of the aperture corrections. Given the relatively large number of standard stars we observed on each night, the transformation coefficients were derived for each night. The accuracy of the zero points of our photometry was determined to be about 0.02 mag. Since our six fields overlap (see Fig. 1), we were able to perform an internal check of our photometry comparing the magnitudes of stars located in the common regions. In every case, the independently calibrated magnitudes agree within 0.02 - 0.03 mag in both, K and J filters. Unfortunately, we are not aware of any other deep near-infrared JK images obtained for IC 1613, so an external check of our photometry is not possible. However, the magnitudes of the bright stars in our fields can be compared with the 2MASS photometry. Fig. 2 presents the difference between our K magnitudes and J-K colors, and the corresponding 2MASS data for common bright stars. Before calculating these differences, we transformed our photometry, which had been calibrated onto the UKIRT system, to the 2MASS system using the equations provided by Carpenter (2001). In spite of the relatively large uncertainties of the 2MASS data for the fainter stars in Fig. 2, it is appreciated that there is no evident zero point offset either in K or in J-K, supporting the conclusion that both datasets are well calibrated, within 0.02-0.03 mag. The pixel positions of the stars were transformed to the equatorial coordinate system using Digital Sky Survey (DSS) images. For this purpose, we used the algorithm developed and used in the OGLE project (Udalski et al. 1998). The accuracy of our astrometric transformations is better than 0.3 arcsec. \section{Results} \subsection{The Cepheid mean K and J band magnitudes} In the six NTT/SOFI fields observed in this project, 39 objects from the Cepheid list presented by Udalski et at. (2001) were identified. It is worth noticing that most of the (few) long-period OGLE-II Cepheids in IC 1613 are located in our fields. Thanks to the depth of our infrared photometry, we were able to detect Cepheids with periods down to about 2 days. In Table 2, we present the journal of individual observations of these 39 Cepheid variables. Most of them were observed only once. However, several objects located in the overlapping areas were observed twice. Before deriving the mean K and J magnitudes of the Cepheids from these data, we tried to improve the periods given by Udalski et al. (2001) using our new three-epoch optical observations, which are given in Table 3. For the long-period Cepheids (P larger than about 5 days), we could indeed improve the periods with these new data. However, for the variables with shorter periods the time elapsed between two sets of observations was too large in order to unambiguously count the number of elapsed cycles. For these Cepheids, we adopted the OGLE II periods from Udalski et al. (2001). The mean magnitudes were obtained from the template light curve method of Soszy{\'n}ski, Gieren and Pietrzy{\'n}ski (2005) which uses the V-band phases of the individual near-IR observations, and the light curves amplitudes in V and I to calculate the differences of the individual single-phase magnitudes to the mean magnitudes in J and K. For a detailed description of this technique, the reader is referred to this paper. It has been demonstrated by these authors that the mean K and J magnitudes of Cepheids can be derived from just one random-phase observation with an accuracy of 0.02-0.03 mag provided that high-quality optical and infrared data, and periods are available for the stars. In Table 4, we present the final intensity mean JK magnitudes of the 39 Cepheids in our fields with their estimated uncertainties and their adopted periods. The last column contains remarks on some of the variables. V2, V6, etc., correspond to the numbering system introduced by Sandage (1971). \subsection{Selection of the final sample} In Fig. 3, we show the optical V vs. V-I color-magnitude diagram of IC 1613 obtained from the OGLE II data (Udalski et al. 2001) on which the locations of the Cepheids observed in the present study are marked. In Fig. 4, we display the PL relations in the K and J bands which we obtain from the data of all the 39 Cepheids in Table 4. While the data define tight PL relations in both bands, there are some objects which clearly deviate from the bulk of the Cepheids in these diagrams, and which need individual discussion. These stars are indicated with open circles in Fig. 4. Star 13682 is most probably not a Cepheid variable (Sandage 1971, Antonello et al. 1999). Udalski et al. (2001) supported this conclusion from the position of this star on the V, V-I CMD (see Fig. 3), and its abnormal location on the optical PL relations on which 13682 appears much brighter than other Cepheids with similar periods. This proves also true for its near-IR magnitudes in Fig. 4. We therefore exclude this object for the distance determination. Besides star 13682, the variables 13709, 11743, 12068, 8173 and 2771 are also very significantly brighter than other Cepheids with similar periods. The two latter Cepheids with their very short periods are almost certainly first overtone pulsators. Due to the detection limit in our present near-infrared photometry we would not see fundamental mode pulsators at this very short period of about 1.3 days. The three other over-bright Cepheids are probably blended by relatively bright stars. These Cepheids are also located above the Cepheid PL relations in the V and I bands (Udalski et al. 2001). In Fig. 3, variable 11743 appears close to the red edge of the instability strip, while the heavily blended Cepheid 13709 lies outside the strip, supporting the blending hypothesis. For star 12068, unfortunately no V-I color is available. The remaining two clearly deviating stars, 10421 and 17473, were already classified as Population II Cepheids by Udalski et al. (2001). Indeed, these stars are located about 2 magnitudes below the IR Cepheid PL relations (see Fig. 4), which fully supports the conclusion about the Pop. II nature of these objects. In the light of the arguments presented above, we decided to reject all these eight objects from the final sample of Cepheids used for the distance determination. Finally, we would like to comment on the Cepheid designated as 7647. Udalski et al. (2001) suspected this star to be a heavily blended Cepheid. Indeed, as can be seen in Fig. 3, this star is located bluewards to the instability strip, suggesting that this variable is blended with a very bright blue star. In the infrared, Cepheid 7647 appears with normal flux and colors. This finding is consistent with the presence of a blue, unresolved companion star, which does contaminate the optical, but not the near-infrared photometry. We therefore retain this Cepheid in our final list of stars for the distance solution in the infrared. We remark that an omission of this star from the final sample would not significantly alter the results we will present below. The errors of the mean K-band magnitudes for Cepheids with log P (days) $<$ 0.5 become large due to a) the relatively low accuracy of the K-band photometry for such a faint stars ( $K > 20.5$ mag), and b) the increasingly uncertain mean magnitude corrections for these stars, caused by their relatively noisy optical light curves, and less accurate periods. We obtained linear regressions to the PL relations in the J and K bands for the whole sample, including the faintest stars, and for the subsamples limited to the Cepheids with log P (days) $>$ 0.5, finding very good agreement (to better than 1 $\sigma$) between the corresponding solutions. However, since the inclusion of the shortest-period Cepheids in the solutions does increase the noise significantly, we decided to adopt log P (days) = 0.5 as a lower cut-off period for our solutions. This way, our final samples in J and K still comprise some 20 Cepheids with excellent photometry, which is sufficient for a very accurate determination of the distance to IC 1613. \subsection{Determination of the distance and reddening} The least-squares fits to the mean magnitudes of the Cepheids from our carefully selected final list yield the following slopes of the PL relations: $ -3.117 \pm 0.044$ in J and $-3.148 \pm 0.053 $ in K. The stated errors are $1 \sigma$ uncertainties. These values agree very well with the slopes of the PL relations for the LMC Cepheids in these bands derived by Persson et al. (2004) (-3.153 and -3.261 in J and K, respectively), and are consistent with the LMC values within the combined uncertainties. We therefore calulated the zero points of the Cepheid J and K band PL relations in IC 1613 by adopting the corresponding slopes from Persson et al. (2004). This yields the following results: \\ J = -3.153 log P + (22.187 $\pm$ 0.040) \\ K = -3.261 log P + (21.827 $\pm$ 0.045) \\ The adopted linear regressions to our K and J Cepheid data are shown in Fig. 5. Before calculating, from the determined zero points, the relative distance of IC 1613 with respect to the LMC, we need to convert our PL relation zero point magnitudes calibrated onto the UKIRT system (Hawarden et al. 2001) to the NICMOS system on which the corresponding LMC zero points were calibrated (Persson et al. 2004). According to Hawarden et al. (2001) there are just zero point offsets between the UKIRT and NICMOS systems (e.g. no color dependence) in the J and K filters, which amount to 0.034 and 0.015 mag, respectively. After adding these offsets, and assuming an LMC true distance modulus of 18.5 mag (see next section for discussion on this assumption), we derived the following distance moduli for IC 1613: 24.385 (J), and 24.306 mag (K) . The corresponding distance moduli in the optical V (24.572 mag) and I (24.488 mag) bands had been previously calculated from the OGLE-II data by Udalski et al. (2001) adopting the linear LMC Cepheid P-L relations (Udalski et al. 1999, Udalski 2000). With the values of the distance moduli of IC 1613 derived in four different bands, providing the large coverage in wavelength from 0.5-2.2 microns, we can compute the reddening, and true distance modulus of the galaxy very accurately. Adopting the extinction law of Schlegel et al. (1998), and following the approach we have developed in the study of NGC 300 (Gieren et al. 2005b), we fit a straight line to the relation $(m-M)_{0} = (m-M)_{\lambda} - A_{\lambda} = (m-M)_{\lambda} - E(B-V) * R_{\lambda} $. The best least squares fit to this relation yields: \\ $(m-M)_{0} = 24.291 \pm 0.014$ $ E(B-V) = 0.090 \pm 0.007$ From Fig. 6, it is appreciated that the true distance modulus, and the total reddening of IC 1613 are indeed very well determined from the available distance moduli in the different photometric bands. \section{Discussion} In the following, we will discuss the various assumptions we made, and possible systematic errors which could affect our distance determination of IC 1613. Almost certainly, the largest contribution to our total error budget comes from the current uncertainty of the distance to the LMC. Since this problem has been extensively discussed in the recent literature (e.g. Benedict et al. 2002; Walker 2003), we will not focus on this discussion here. The value of 18.50 mag for the true LMC distance modulus adopted in this paper assures to have our distance results on the same scale as the results from the HST Key Project team (Freedman et al. 2001) and our own previous distance studies in the course of the Araucaria Project (Pietrzynski et al. 2004, Gieren et al. 2004, Gieren et al. 2005b). The adopted fiducial slopes of the J- and K-band Cepheid PL relations from the work of Persson et al. (2004) are based on about 100 LMC Cepheids with periods bracketting those of the IC 1613 Cepheids used in our present study. The Persson et al. infrared PL relations clearly represent the most accurate determination of these relations currently available in the literature. From the recent work of Gieren et al. (2005a), there is evidence that the infrared Cepheid PL relations in the Milky Way agree with the corresponding LMC relations, within the combined $1 \sigma$ uncertainties. In Gieren et al. (2005b), we found that the slopes of the Persson et al. PL relations in J and K do also provide excellent fits to the Cepheid near-IR data in NGC 300, with its slightly more metal-rich young population than the one in the calibrating LMC (Urbaneja et al. 2005). From the present study of IC 1613, we now see that the slopes of the LMC near-IR PL relations give an excellent fit to the metal-poor population of Cepheids in IC 1613, too. This indicates that on the one hand, use of the Persson et al. LMC PL relations does not introduce any significant systematic error in our current determination of the IC 1613 distance; on the other hand, it strongly suggests that in the near-infrared domain, as in the optical domain, the slope of the Cepheid PL relation is independent of metallicity in the wide range from about -1.0 dex up to solar abundance. This empirical finding is in good agreement with the model predictions of Bono et al. (1999) who have found that both, the zero-point and the slope of the K-band PL relation depend only marginally on metal abundance. They found that the predicted slope in K is 3.19 $\pm$ 0.09 for the LMC, and 3.27 $\pm$ 0.09 in the SMC, compatible with both the empirical value determined by Persson et al. (2004) for the LMC, and with a zero change in the slope of the K-band PL relation when going from LMC (-0.3 dex) metallicity to the SMC (-0.7 dex) metal abundance. Finally, it is worthwhile to notice that the adoption of the slightly non-linear P-L relations for Cepheids in the LMC, as advocated by Tammann and Reindl 2002, and more recently Ngeow et al. 2005, would practically have no influence on the results presented in this paper. Indeed, as has already been stated in Ngeow et al. (2005), such an effect would introduce a change less than 3 percent for the derived distance modulus, which is in the order of the one $\sigma$ error of our current determination. In order to check this out more carefully, we used the Ngeow et al. P-L relations in both optical and infrared bands for LMC Cepheids with periods longer than 10 days as fiducial relations and re-calculated the distance moduli in the VIJK bands. This exercise resulted in revised distance moduli to IC 1613 which in all bands were consistent within one $\sigma$ with our original results obtained by using the Cepheid P-L relations of Udalski (2000) and Persson et al. (2004) for the LMC. The possible non-linearity of the LMC P-L relation, and the associated slight change of its slope for the long-period Cepheids, is therefore not a significant problem in the context of our current distance work. Yet, it will be very important to improve on the slope for the long-period LMC Cepheid P-L relation by using very accurate and homogeneous new data. We are currently involved in a project to obtain such new data in the V and I bands. The sample of Cepheids used for our present distance determination to IC 1613 is relatively large, making our distance result invulnerable to the problem of an inhomogeneous filling of the instability strip which is ideally required in such studies. We suspect that the main reason for the difference of 0.14 mag between our current distance result for IC 1613 and the one obtained by Macri et al. (2001) is the small number of Cepheids available for their study (4), which does not guarantee a homogeneous filling of the instability strip and can cause a relatively large systematic offset of the derived distance modulus from the true value. Therefore, we consider our present result as consistent with the HST-based result of Macri et al. (2001). The location of the Cepheids of our final sample in the CMD in Fig. 3 shows that they do indeed cover the instability strip quite homogeneously. Moreover, the period range for the PL solution is very wide and rather uniformly covered with stars-we chose our IR fields in such a way as to optimize the period distribution of the Cepheids in these fields. Applying different cut-off periods to our sample (e.g. log P = 0.5, 0.7, 1), we always reproduce the zero point results to within one $\sigma$. From this we conclude that our choice for the cut-off period does not affect our final results in any significant way. The most important source of uncertainty while using the optical data {\it alone} is the interstellar reddening. Our present study shows that most of the reddening to the IC 1613 Cepheids is actually contributed in the galaxy itself, which explains the overestimation of the distance to IC 1613 in previous studies from optical data which had only used the very small foreground extinction to IC 1613 to make the reddening correction. Using infrared data, and in particular K-band photometry where the reddening is by an order of magnitude smaller than in the optical bands, the error due to reddening is minimized to a practically insignificant level of about 0.01 mag. From the fact that our new value of E(B-V) yields very consistent distances from the PL relations in all optical and infrared bands, we can also conclude that the extinction law in IC 1613 is not significantly different from the Milky Way law of Schlegel et al. (1998). This is the same conclusion we had already reached in the case of NGC 300 (Gieren et al. 2005b). Another contribution to the error budget comes from the effect of unresolved companions on the Cepheid magnitudes. The few strongly blended Cepheids in our sample were easily detected from their positions on the multiband PL relations, and on the CMD, and were discarded from our further analysis (see section 3.2.). As we extensively discussed in our previous papers (e.g. Gieren et al. 2004, 2005b; Bresolin et al. 2005, Pietrzynski et al. 2004) the blending effect was found to be very small in the cases of NGC 300, and of NGC 6822. In the paper of Bresolin et al. (2005), we were able to demonstrate from HST/ACS images that those Cepheids in NGC 300 which we had identified as strongly blended in the ground-based photometry were indeed the ones with the brightest nearby companions. In that paper, it was shown that the effect of unresolved companion stars on the Cepheids which constituted the final sample was less than 2 percent. Given that IC 1613 is located at less than half the distance of NGC 300, and has a much smaller stellar density, it is reasonable to assume that the effect of blending due to unresolved companion stars on its distance is even smaller than in the case of NGC 300, and does not contribute in a significant way to the systematic uncertainty of our present result. While it now seems well established that the slopes of the Cepheid P-L relations in optical and near-infared bands do not depend, within our current detection sensitivity, on metallicity over a very broad range of this parameter ( -1 $<$ [Fe/H] $<$ 0 ; see previous discussion), a possible metallicity dependence of their {\it zero points} is still under discussion (Sakai et al. 2004; Storm et al. 2004; Pietrzynski et al. 2004; Pietrzynski and Gieren 2005). In particular, due to the fact that up to now very few galaxies have been {\it exhaustively} surveyed for Cepheids in the infrared, it is currently not possible to draw any firm conclusion about the potential variation of the infrared PL relation zero points with metallicity. Soon, once the data for all target galaxies observed in the course of the Araucaria Project will be analyzed, we should be able to put tighter constraints on this open question, and if needed calibrate the metallicity dependence of PL relation zero points in both optical and infrared domains with high precision. \section{Summary and conclusions} We have measured accurate NIR magnitudes in the J and K bands for 39 Cepheid variables in the Local Group galaxy IC 1613 with well-determined periods and optical (VI) light curves. Mean magnitudes in J and K were derived for these variables using the single-phase approach of Soszy{\'n}ski et al. (2005). After carefully cleaning the Cepheid list from blended objects, Population II variables and overtone pulsators, we have determined accurate PL relations. Fits to these observed relations were made using the slopes of the LMC relations determined by Persson et al. (2004), which gave an excellent representation of the IC 1613 data, providing for the first time solid evidence that the slope of the Cepheid PL relation is independent of metallicity down to the low metallicity of -1.0 dex of the IC 1613 young population in the near-infrared domain, too. This is in agreement with the theoretical predictions of Bono et al. (1999). By combining the zero points of the J and K band PL relations in our study with the ones derived by Persson et al. for the LMC, we derive relative distance moduli of IC 1613 with respect to the LMC in both bands. Combining these infrared moduli with the distance moduli previously derived by Udalski et al. (2001) in V and I, we determine the total (average) reddening of the Cepheids in IC 1613, and the true distance modulus of this galaxy with an unprecedented accuracy. For the reddening, we find E(B-V) = 0.090 $\pm$ 0.007 mag, and for the true distance modulus of IC 1613 from our multiwavelength approach we obtain 24.291 $\pm$ 0.014 mag (random error). As in the case of our study of NGC 300 with the same method, we find evidence that there is a significant contribution to the total reddening from dust absorption {\it intrinsic} to IC 1613, which had been neglected in the previous Cepheid distance work on this galaxy. The excellent fit of the distance moduli to the assumed galactic extinction law suggests that the interstellar extinction in this small irregular galaxy follows closely the galactic law. We show that our derived Cepheid distance is very insensitive to systematic uncertainties caused by the fiducial PL relations used in our fits, possible inhomogeneous filling of the instability strip by our Cepheid sample, and problems with blending of the variables. Any remaining influence of the uncertainty of reddening on our distance result is negligible. All these possible sources of error contribute less to the total systematic uncertainty of our result than the two dominant sources of error, which are the zero points of our JK photometry ($\pm$ 0.03 mag), and the distance to the LMC, which we have {\it adopted} as 18.50 mag, and whose current uncertainty seems in the order of $\pm$ 0.10 mag. Our distance determination for IC 1613 is in reasonable agreement with the previous determination of Macri et al. (2001) from HST H-band photometry of four Cepheids, 24.43 $\pm$ 0.08 mag. We attribute the 0.14 mag difference mainly to the small number of stars available to Macri et al. in their study. Our new distance determination is also in very good agreement with the very early infrared work of McAlary et al. (1984); these authors had obtained a distance modulus of 24.31 $\pm$ 0.12 from H-band data of 10 Cepheids in IC 1613. A change of their assumed reddening of 0.03 mag to our larger value found in this study still produces excellent agreement of their result with ours. The distance to IC 1613 was also determined by Dolphin et al. (2001) using HST data, and employing a number of distance indicators (TRGB, red clump stars, RR Lyrae stars, Cepheids). We note that their Cepheid sample was very small, which can clearly lead to spurious results. Udalski et al. 2001 have observed an order of magnitude larger sample of Cepheids in this galaxy and showed that all these different distance indicators yield consistent distances to this galaxy. Those distance measurements are all in very good agreement with the distance of IC 1613 obtained in this study if the revised reddening of 0.09 mag found in this study, and a LMC true distance modulus of 18.5 mag are assumed. As a final conclusion, we have produced in this work a determination of the Cepheid distance to IC 1613 whose random error is of the order of 1\%, and estimated systematic error (excluding the uncertainty of the adopted LMC distance) is in the order of 2\%. This accuracy, when combined with distance determinations of similar accuracy we pretend to obtain for most of the other galaxies of the Araucaria Project, should enable us to pin down the metallicity dependence of the PL relation zero points in the different optical and near-infrared photometric bands with the 1-2\% accuracy needed to produce a significant improvement in the determination of the Hubble constant from distance determinations to galaxies in the nearby field from their Cepheid populations, which is the approach used in the HST Key Project of Freedman et al. (2001). Our work in the Araucaria Project should therefore strongly contribute, in the very near future, to make best use of the past work of the Key Project team. \acknowledgments We would like to thank the anonymous referee for his interesting suggestions which helped to improved this paper. We gratefully acknowledge the generous allocation of observing time by ESO to our distance scale projects. We also appreciate the excellent staff support at the telescope at ESO/La Silla where these data were obtained. WG and GP gratefully acknowledge financial support for this work from the Chilean Center for Astrophysics FONDAP 15010003. Support from the Polish KBN grant No 2P03D02123 and BST grant for Warsaw University Observatory is also acknowledged. \clearpage \begin{deluxetable}{c c c} \tablewidth{0pc} \tablecaption{Coordinates of the Centers of the Observed SOFI/NTT Fields in IC 1613} \tablehead{ Field & \colhead{RA} & \colhead{DEC} } \startdata C1 & 01:04:58.9 & 02:05:48.5 \nl C2 & 01:04:56.6 & 02:04:13.2 \nl C3 & 01:04:59.6 & 02:09:23.9 \nl C4 & 01:04:37.2 & 02:05:58.8 \nl C5 & 01:04:35.1 & 02:09:29.4 \nl C6 & 01:05:03.0 & 02:09:44.4 \nl \enddata \end{deluxetable} \clearpage \begin{deluxetable}{ccccccc} \tablewidth{0pc} \tablecaption{Journal of the Individual J and K Observations of IC 1613 Cepheids} \tablehead{ \colhead{ID} & HJD (J) & J & $\sigma_{\rm J}$ & HJD (K) & K & $\sigma_{\rm K}$} \startdata 11446 & 53215.86015 & 16.997 & 0.009 & 53215.80516 & 16.492 & 0.009 \\ 11446 & 53315.64411 & 17.400 & 0.016 & 53315.58668 & 16.822 & 0.016 \\ 10421 & 53315.64411 & 19.551 & 0.060 & 53315.58668 & 18.877 & 0.071 \\ 1987 & 53370.57443 & 17.938 & 0.019 & 53370.52660 & 17.593 & 0.024 \\ 736 & 53315.72831 & 17.833 & 0.018 & 53315.67116 & 17.260 & 0.018 \\ 7647 & 53370.57443 & 18.144 & 0.020 & 53370.52660 & 17.716 & 0.027 \\ 13738 & 99999.99999 & 99.999 & 9.999 & 53215.88146 & 18.165 & 0.032 \\ 13738 & 53315.56228 & 18.763 & 0.043 & 53315.50158 & 18.249 & 0.046 \\ 13682 & 99999.99999 & 99.999 & 9.999 & 53215.88146 & 15.763 & 0.007 \\ 13682 & 53315.56228 & 16.760 & 0.010 & 53315.50158 & 15.751 & 0.007 \\ \enddata \end{deluxetable} \clearpage \begin{deluxetable}{cccc} \tablewidth{0pc} \tablecaption{Journal of the Individual V band Observations of IC 1613 Cepheids} \tablehead{ \colhead{ID} & HJD & V & $\sigma_{\rm V}$} \startdata 11446 & 53620.76797 & 18.385 & 0.009 \nl 11446 & 53621.75534 & 18.422 & 0.008 \nl 11446 & 53621.82741 & 18.412 & 0.008 \nl 10421 & 53620.76797 & 21.632 & 0.079 \nl 10421 & 53621.75534 & 21.648 & 0.069 \nl 1987 & 53620.76797 & 19.764 & 0.019 \nl 1987 & 53621.75534 & 19.815 & 0.018 \nl 1987 & 53621.82741 & 19.829 & 0.019 \nl 736 & 53620.76797 & 19.593 & 0.017 \nl 736 & 53621.75534 & 19.671 & 0.018 \nl 736 & 53621.82741 & 19.675 & 0.023 \nl 7647 & 53620.76797 & 19.110 & 0.014 \nl 7647 & 53621.75534 & 19.080 & 0.012 \nl 7647 & 53621.82741 & 19.100 & 0.012 \nl 13738 & 53620.76797 & 20.035 & 0.020 \nl 13738 & 53621.75534 & 20.115 & 0.021 \nl 13738 & 53621.82741 & 20.109 & 0.023 \nl \enddata \end{deluxetable} \clearpage \begin{deluxetable}{cccccccccccc} \tablewidth{0pc} \tablecaption{Final Intensity Mean J and K Magnitudes of IC 1613 Cepheids} \tablehead{ \colhead{OGLE ID} & P & $\log{P}$ & $\langle{J}\rangle$ & $\sigma_J$ & $\langle{K}\rangle$ & $\sigma_K$& remarks} \startdata 11446 & 41.87 & 1.62194 & 17.114 & 0.009 & 16.605 & 0.009& V20\\ 10421 & 29.19 & 1.46529 & 19.694 & 0.060 & 19.029 & 0.071&PII, V47\\ 1987 & 25.398 & 1.40480 & 17.715 & 0.019 & 17.405 & 0.024&V11\\ 736 & 23.469 & 1.37049 & 17.745 & 0.018 & 17.256 & 0.018&V2\\ 7647 & 16.488 & 1.21716 & 18.056 & 0.020 & 17.688 & 0.027&blend\\ 13738 & 16.420 & 1.21537 & 18.590 & 0.043 & 18.066 & 0.028&V18\\ 13682 & 14.317 & 1.15585 & 16.815 & 0.010 & 15.818 & 0.005&not Cepheid, V39\\ 17473 & 13.154 & 1.11906 & 99.999 & 9.999 & 20.669 & 0.251 &PII\\ 7664 & 10.4390 & 1.01866 & 18.996 & 0.030 & 18.555 & 0.048&V16\\ 926 & 9.4286 & 0.97445 & 19.109 & 0.028 & 18.736 & 0.036&V6\\ 879 & 9.2130 & 0.96440 & 19.156 & 0.046 & 18.689 & 0.054&V25\\ 13808 & 7.572 & 0.87921 & 19.591 & 0.092 & 19.160 & 0.058&\\ 13759 & 7.3403 & 0.86571 & 19.272 & 0.074 & 18.832 & 0.112&V7\\ 13709 & 6.741 & 0.82872 & 18.480 & 0.041 & 17.739 & 0.020&blend\\ 5037 & 6.3175 & 0.80055 & 19.824 & 0.065 & 19.484 & 0.127&\\ 11604 & 5.8191 & 0.76486 & 19.685 & 0.051 & 19.133 & 0.097&\\ 13780 & 5.5771 & 0.74641 & 19.973 & 0.137 & 19.256 & 0.069&V9\\ 11831 & 5.0269 & 0.70130 & 19.902 & 0.063 & 19.532 & 0.111&\\ 8146 & 4.5630 & 0.65925 & 20.306 & 0.075 & 19.730 & 0.122&\\ 14287 & 4.365 & 0.63998 & 99.999 & 9.999 & 19.951 & 0.142 &\\ 12109 & 4.1364 & 0.61662 & 20.128 & 0.079 & 19.344 & 0.093&\\ 13784 & 4.0657 & 0.60914 & 99.999 & 9.999 & 19.459 & 0.096 &V10\\ 11743 & 3.8953 & 0.59054 & 19.348 & 0.035 & 18.795 & 0.047&blend, V53\\ 8127 & 3.8444 & 0.58483 & 20.636 & 0.097 & 20.267 & 0.159&\\ 2240 & 3.0733 & 0.48760 & 20.941 & 0.138 & 20.221 & 0.147&V35\\ 18349 & 2.8700 & 0.45788 & 99.999 & 9.999 & 19.639 & 0.102 &V29\\ 19024 & 2.8418 & 0.45359 & 99.999 & 9.999 & 19.859 & 0.154 &\\ 12068 & 2.781 & 0.44420 & 20.013 & 0.060 & 19.339 & 0.091&blend\\ 2760 & 2.7123 & 0.43334 & 21.101 & 0.127 & 20.651 & 0.203&\\ 10804 & 2.6629 & 0.42535 & 21.041 & 0.137 & 20.404 & 0.197&V48\\ 12526 & 2.6310 & 0.42012 & 99.999 & 9.999 & 20.982 & 0.358 &\\ 7322 & 2.3378 & 0.36881 & 21.125 & 0.141 & 21.145 & 0.328&\\ 6128 & 2.2578 & 0.35369 & 20.712 & 0.114 & 20.140 & 0.174&\\ 8782 & 2.0930 & 0.32077 & 21.013 & 0.137 & 20.673 & 0.255&\\ 5996 & 2.0682 & 0.31559 & 20.888 & 0.183 & 20.508 & 0.284& V60\\ 2389 & 2.0286 & 0.30720 & 20.837 & 0.115 & 20.363 & 0.175&\\ 13481 & 1.678 & 0.22479 & 99.999 & 9.999 & 21.226 & 0.392 &\\ 2771 & 1.3290 & 0.12352 & 99.999 & 9.999 & 20.193 & 0.150 &FO\\ 8173 & 1.3103 & 0.11737 & 20.587 & 0.099 & 20.020 & 0.130&FO\\ \enddata \end{deluxetable} \clearpage \begin{deluxetable}{cccccc} \tablewidth{0pc} \tablecaption{Reddened and Extinction-Corrected Distance Moduli for IC 1613 in Optical and Near-Infrared Bands} \tablehead{ \colhead{Band} & $V$ & $I$ & $J$ & $K$ & $E(B-V)$ } \startdata $m-M$ & 24.572 & 24.488 & 24.385 & 24.306 & -- \nl ${\rm R}_{\lambda}$ & 3.24 & 1.96 & 0.902 & 0.367 & -- \nl $(m-M)_{0}$ & 24.277 & 24.309 & 24.302 & 24.273 & 0.090 \nl \enddata \end{deluxetable}
Title: The broadband afterglow of GRB 030328
Abstract: We here report on the photometric, spectroscopic and polarimetric monitoring of the optical afterglow of the Gamma-Ray Burst (GRB) 030328 detected by HETE-2. We found that a smoothly broken power-law decay provides the best fit of the optical light curves, with indices alpha_1 = 0.76 +/- 0.03, alpha_2 = 1.50 +/- 0.07, and a break at t_b = 0.48 +/- 0.03 d after the GRB. Polarization is detected in the optical V-band, with P = (2.4 +/- 0.6)% and theta = (170 +/- 7) deg. Optical spectroscopy shows the presence of two absorption systems at z = 1.5216 +/- 0.0006 and at z = 1.295 +/- 0.001, the former likely associated with the GRB host galaxy. The X-ray-to-optical spectral flux distribution obtained 0.78 days after the GRB was best fitted using a broken power-law, with spectral slopes beta_opt = 0.47 +/- 0.15 and beta_X = 1.0 +/- 0.2. The discussion of these results in the context of the "fireball model" shows that the preferred scenario is a fixed opening angle collimated expansion in a homogeneous medium.
https://export.arxiv.org/pdf/astro-ph/0601293
\title{The broadband afterglow of GRB 030328} \classification{95.75.De; 95.75.Fg; 95.75.Hi; 95.85.Kr; 98.70.Rz} \keywords {gamma-ray bursts; astronomical observations: visible; photometry; spectroscopy; polarimetry} \author{E. Maiorano}{ address={INAF-IASF, Bologna, Italy} } \author{N. Masetti}{ address={INAF-IASF, Bologna, Italy} } \author{E. Palazzi}{ address={INAF-IASF, Bologna, Italy} } \author{S. Savaglio}{ address={The Johns Hopkins Univ., Baltimore, USA} } \author{E. Rol}{ address={Univ. of Leicester, UK} } \author{P.M. Vreeswijk}{ address={ESO-Santiago, Chile} } \author{E. Pian}{ address={INAF-Oss. Astron. Trieste, Italy} } \author{P.A. Price}{ address={Univ. of Hawaii, Honolulu, USA} } \author{B.A. Peterson}{ address={Australian National Univ., Weston, Australia} } \author{M. Jel\'{i}nek}{ address={IAA-CSIC, Granada, Spain} } \author{S.B. Pandey}{ address={IAA-CSIC, Granada, Spain} } \author{M.I. Andersen}{ address={AIP, Potsdam, Germany} } \author{A.A. Henden}{ address={US Naval Observatory, Flagstaff, USA} } \section{Introduction} GRB 030328 was a long, bright GRB detected on 2003 Mar 28.4729 UT, by the FREGATE, WXM, and SXC instruments onboard {\it HETE-2}, and rapidly localized with sub-arcminute accuracy (Villasenor et al. 2003). About $\sim$1 hour after the GRB, its optical afterglow has been detected by the 40-inch Siding Spring Observatory (SSO) telescope (Peterson \& Price 2003). A study of the X--ray afterglow of GRB 030328 was performed by Butler et al. (2005) using {\it Chandra} data. We report here on the study of the optical afterglow emission of GRB 030328 made, within the GRACE\footnote{GRB Afterglow Collaboration at ESO; see {\tt http://www.gammaraybursts.org/grace/}} collaboration, performed with 7 different optical telescopes. A more detailed presentation of these data will appear in Maiorano et al. (2005). \section{Observations} Optical $UBVRI$ data of the GRB 030328 Optical Transient (OT), for a total of 130 photometry points, were acquired at the 40-inch SSO (Australia), 1m ARIES (India), 2.5m NOT (Spain), 1.54m ESO Danish, 2.2m ESO/MPG, ESO VLT-$Antu$ (Chile) and 1m USNO-FS (USA) telescopes. A series of six 10-min optical spectra was obtained at ESO-Paranal with VLT-{\it Antu} starting 0.59 d after the GRB. The Grism 300V was used with a nominal spectral coverage of 3600--8000 \AA~and a spectral dispersion of 2$\farcs$7 \AA/pixel. Linear polarimetry $V$-band observations were acquired between 0.66 and 0.88 d after the GRB at VLT-{\it Antu}. Five complete imaging polarimetry cycles were performed. \section{Results} \subsection{Photometry} In Fig. 1 (left) we plot our photometric measurements together with those reported in the GCN circulars\footnote{{\tt http://gcn.gsfc.nasa.gov/gcn/gcn3\_archive.html}}. For the cases in which no error was reported, a 0.3 mag uncertainty was assumed. The $UBVRI$ zero-point calibration was performed using the photometry by Henden (2003). The optical data of Fig. 1 (left) were corrected for the Galactic foreground reddening assuming $E(B-V)$ = 0.047 mag (Schlegel et al. 1998). The GRB 030328 host galaxy emission in the $BVRI$ bands was computed from the data of Gorosabel et al. (2005) and subtracted from our optical data set. The best fit of the $R$-band data (in Fig. 1, left) is obtained using a smoothly broken powerlaw (Beuermann et al. 1999), with temporal indices $\alpha_1 = 0.76 \pm 0.03$ and $\alpha_2 = 1.50 \pm 0.07$ before and after a break occurring at $t_b = 0.48 \pm 0.03$ d from the GRB trigger, and with $s = 4.0 \pm 1.5$ the parameter modeling the slope change rapidity. This best-fit curve describes well the data in the other bands also (see Fig. 1, left). This means that the decay of the OT can be considered as achromatic. From the measured jet break time we can compute the jet opening angle value for GRB 030328 which is, following Sari et al. (1999), $\theta_{jet} \sim 3\fdg2$. \subsection{Broadband analysis} By using the available information, we have constructed the optical-to-X--ray spectral flux distribution of the GRB 030328 afterglow at the epoch 0.78 days after the GRB, that is, the time with best broadband photometric coverage. As Fig. 1 (right) shows, the best descriptions are a single powerlaw (dashed line) with a spectral index $\beta_{\rm X-opt}$ = 0.83$\pm$0.01, or a broken powerlaw with $\beta_{\rm opt}$ = 0.47$\pm$0.15 and assuming $\beta_{\rm X}$ = 1.0$\pm$0.2 from Butler et al. (2005). However, the single powerlaw description of the broadband afterglow does not fit any of the synchrotron fireball scenarios (Sari et al. 1998, 1999). Instead, in the broken powerlaw description, which means that the synchrotron cooling frequency $\nu_{\rm c}$ lies between the optical and X--ray bands, we obtain that the GRB 030328 afterglow broadband evolution is consistent with a jet-collimated expansion in a homogeneous medium with fixed opening angle (M\'esz\'aros \& Rees 1999) and with an electron distribution index $p$ = 2. Assuming a negligible host absorption and using the optical and X--ray spectral slopes above, we obtain for $\nu_{\rm c}$ the value $5.9 \times 10^{15}$ Hz, which places this frequency in the ultraviolet band. \subsection{Spectroscopy} Figure 2 shows the spectrum of the GRB 030328 OT. Most of the significant features can be identified with Fe {\sc ii}, Mg {\sc ii}, Al {\sc ii} and C {\sc iv} absorption lines in a system at a redshift of $z$ = 1.5216 $\pm$ 0.0006. These lines are associated with the circumburst gas or interstellar medium in the GRB host galaxy. A lower redshift absorption system at $z$ = 1.295 $\pm$ 0.001 is also found: for it, only two lines can be identified (Fe {\sc ii} $\lambda$2600 and the unresolved Mg {\sc ii} $\lambda$$\lambda$2796, 2803 doublet). Its detection indicates the presence of a foreground absorber. \subsection{Polarimetry} After correcting for spurious field polarization, we found $Q_{\rm OT} = 0.029 \pm 0.008$ and $U_{\rm OT} = -0.004 \pm 0.008$. The fit of the data with the relation of Di Serego Alighieri (1997) yielded for the OT a linear polarization $P = (2.4 \pm 0.6)$ \% and a polarization angle $\theta = 170^{\circ} \pm 7^{\circ}$, corrected for the polarization bias (Wardle \& Kronberg 1974). In order to check whether variations of $P$ and $\theta$ occurred during the polarimetric run, we also separately considered each of the 5 single polarimetry cycles. Although with lower S/N, $P$ and $\theta$ are consistent with being constant across the whole polarimetric observation run.
Title: A genetic algorithm for the non-parametric inversion of strong lensing systems
Abstract: We present a non-parametric technique to infer the projected-mass distribution of a gravitational lens system with multiple strong-lensed images. The technique involves a dynamic grid in the lens plane on which the mass distribution of the lens is approximated by a sum of basis functions, one per grid cell. We used the projected mass densities of Plummer spheres as basis functions. A genetic algorithm then determines the mass distribution of the lens by forcing images of a single source, projected back onto the source plane, to coincide as well as possible. Averaging several tens of solutions removes the random fluctuations that are introduced by the reproduction process of genomes in the genetic algorithm and highlights those features common to all solutions. Given the positions of the images and the redshifts of the sources and the lens, we show that the mass of a gravitational lens can be retrieved with an accuracy of a few percent and that, if the sources sufficiently cover the caustics, the mass distribution of the gravitational lens can also be reliably retrieved. A major advantage of the algorithm is that it makes full use of the information contained in the radial images, unlike methods that minimise the residuals of the lens equation, and is thus able to accurately reconstruct also the inner parts of the lens.
https://export.arxiv.org/pdf/astro-ph/0601124
\date{} % \pagerange{\pageref{firstpage}--\pageref{lastpage}} \pubyear{2005} \label{firstpage} \begin{keywords} gravitational lensing -- methods:~data analysis -- dark matter -- galaxies:~clusters:~general \end{keywords} \section{Introduction} The deflection of light caused by a gravitational lens and the amplifying and distorting effect thereof on the images of background sources, provide us with a means to measure the total mass of the lensing object. If only, of course, that it is possible to ``invert'' such a lensing system, i.e. to infer the mass distribution of the lens given the positions and shapes of a set of lensed images of background sources and the redshifts of the lens and the sources. The inversion of gravitational lensing systems is interesting in its own right, since it puts constraints on the spatial dark matter distribution of the lensing objects and thus helps constrain dark matter physics \citep{n04,d04}, but it may also contribute to cosmology. Good reconstructions of lensing systems may help to constrain the density parameter and the redshift evolution of the dark energy \citep{y01,s04,m05}. The use of the gravitational lensing effect to measure masses of point-mass lenses (or ``stars'') was envisaged by \citet{r64} and \citet{l64}. The idea of using simple parametric models for extended lenses, such as galaxies or clusters of galaxies, can be traced back to \citet{dr80}, who use a King model to estimate the mass distribution of the galaxy that produces the two images of QSO~0957+561A,~B. Since then a host of parametric inversion methods has been developed and applied to observed strong lensing systems, such as the ring cycle method of \citet{k89} or the maximum entropy method of \citet{w94}. \citet{kn93} fitted a bimodal lensing potential, constructed from two elliptical pseudo-isothermal potentials, images of the cluster A370, obtained from the ground under excellent seeing conditions. More elaborate parametric models for lensing clusters associate a simple mass distribution, e.g. a power-law of radius, to each galaxy and to the cluster as a whole. The many parameters that define the lens model are then determined by a $\chi^2$ fit to the observed images \citep{tkd98}. \citet{b05} suggest using a gravitational lens as a cosmic magnifying glass to recover structural details about distant sources. These authors assume a parametric form for the lens mass distribution and use a genetic algorithm to non-parametrically reconstruct the surface brightness distribution of the source. If only a few sources are being lensed or if the sources do not sufficiently cover the caustics, parametric methods are clearly the preferred approach. However, the multitude of arcs and distorted images visible in massive clusters (e.g. \citealt{s87,lh88}), which are observed routinely now at high spatial resolution with the Hubble Space Telescope (e.g. \citealt{br05}), contain a wealth of information and call for more flexible and model independent inversion methods. Using the pixelation method \citep{sw97,a98}, the lens plane is divided into a static grid. The mass in each grid cell and the source positions are estimated so as to construct a solution that best reproduces the observed image positions, subject to regularizing constraints that ensure a smooth mass distribution for the lens that stays close to the luminosity distribution. \citet{t00} perform a multipole-Taylor expansion of the two-dimensional lensing potential, the coefficients of which are determined by a $\chi^2$-fit to the observed images. In \citet{d05}, a non-parametric inversion technique, called {\sc slap}, is presented and applied to an HST image of the cluster Abell 1689 \citep{d05b,k02} that, like the pixelation method, makes use of a grid division of the mass distribution of the lens. This time, however, the grid is dynamic:~it is refined iteratively where the mass density is large. Recently, \citet{d05c} extended {\sc slap} to {\sc wslap} in order to also take into account information in the weak lensing regime. Thus, {\sc slap} and {\sc wslap} are fast and versatile tools for inverting observed lens systems. However, both methods assume the background galaxies to be point sources, which may lead to an over-estimation of the central mass density of the lens in order to focus the images into very compact sources and to physically implausible regions with negative mass density in the lens plane. Adding weak-lensing information alleviates the dependence of the solution on this minimization threshold. {\sc wslap} can make use of quadratic programming to avoid unphysical negative mass densities for the lens. However, this limits the analysis to observables that are linear functions of the lens mass density, such as image positions, and does not allow to incorporate e.g. surface brightness information, which depends non-linearly on the lens mass density. \citet{bra05} also used weak and strong lensing data to invert the X-ray cluster RX~J1347.5$-$1145. Their method evaluates the gravitational potential on a non-dynamic grid. The best fitting gravitational potential is then constructed non-parametrically by minimising a $\chi^2$ function, starting from a parametric priorsolution. The ideal non-parametric lens inversion algorithm {\em (i)} should be free of any assumptions regarding the mass distribution of the lens or the luminosity distributions of the sources, {\em (ii)} should not depend on any prior on the lens mass distribution or any regularisation scheme that could bias the solution, {\em (iii)} should not produce unphysical, i.e. negative, mass densities {\em (iv)} should be free of any uncontrollable parameters, {\em (v)} should be easily extendible to any kind of data, both in the strong and weak lensing regimes, without having to change the inner workings of the algorithm or having to worry about features like continuity or differentiability of the objective function that is extremised. These conceptual issues are the main motivation for this paper, rather than computational speed. Genetic algorithms do an excellent job at fulfilling all these constraints. In this paper, we describe and test a new non-parametric lens inversion technique. The technique makes use of a dynamic grid on which the mass distribution of the lens is approximated by a weighted sum of basis functions and of a genetic algorithm to determine the unknown weights. In the strong lensing regime, where multiple images of each source are available, the following data are offered to the algorithm:~the redshifts of the sources and the lens, and the observed positions of the images. The genetic algorithm generates solutions that satisfy only one minimal constraint:~the back-projected images of a single source should overlap as well as possible in the source plane. We briefly describe the relevant background to gravitational lens systems and genetic algorithms in Section \ref{sec:back}. The details of the inversion method are given in Sect. \ref{sec:invert}. We discuss a number of tests to which we subjected the method in Sect. \ref{sec:sim}. Finally, our conclusions are summarized in Sect. \ref{sec:conc}. \section{Background} \label{sec:back} \subsection{The lens equation} In the thin lens approximation, the lens equation relates viewing directions $\Vec{\theta}$, that are defined in the lens plane, to positions $\Vec{\beta}$ in the source plane:~ \begin{equation} \Vec{\beta}(\Vec{\theta}) = \Vec{\theta} - \frac{D_{ds}}{D_s}\;\Vec{\hat{\alpha}}(\Vec{\theta}) \mcm \label{eq_lenseqn} \end{equation} with $\Vec{\hat{\alpha}}$ the deflection angle, $D_s$ the distance between the observer and the source, and $D_{ds}$ the distance between the lens and the source. The gravitational bending of light rays, described by the deflection angle $\Vec{\hat{\alpha}}$, depends on the viewing direction $\Vec{\theta}$, the mass distribution of the lens and the distance between the lens and the observer. Here and in the following, distances should be interpreted as angular-diameter distances. For simplicity, we will adopt a standard CDM cosmology, with a matter density $\Omega=1$ and a Hubble parameter $H_0 = 70$~km~s$^{-1}$~Mpc$^{-1}$, in which the angular-diameter distance between an observer at redshift $z_1$ and a source at redshift $z_2$ is given by \begin{equation} D(z_1,z_2) = \frac{2c}{H_0} \frac{1}{1+z_2} \left(\frac{1}{\sqrt{1+z_1}}- \frac{1}{\sqrt{1+z_2}} \right) \mpt \end{equation} We will always assume that the redshifts of the lens and of the source(s) are known to the observer. The lens equation projects the images back onto their respective sources in the source plane. When a gravitational lens produces multiple images of a single source, one can use the lens equation to find, for each image, the corresponding region in the source plane. Since all images correspond to a single source, all back-projected images have to coincide. \subsection{The Plummer lens} We first describe the gravitational lens effect caused by a Plummer sphere \citep{p11} at a distance $D_d$ from the observer. The projected density distribution of a Plummer sphere with total mass $M$ and angular scale-length $\theta_P$ as a function of angular distance $\theta$ is given by \begin{equation} \Sigma(\theta) = \frac{M}{\pi D_d^2}\frac{\theta_P^2}{(\theta^2+\theta_P^2)^2}. \label{eq_sigma_plummer} \end{equation} This mass distribution leads to the following lens equation: \begin{equation} \Vec{\beta}(\Vec{\theta}) = \Vec{\theta}-\frac{D_{ds}}{D_s D_d}\frac{4 G M}{c^2}\frac{\Vec{\theta}}{\theta^2+\theta_P^2} \end{equation} if the coordinate system in the lens plane is centered on the Plummer sphere. As a first step towards inverting a given lens system, we write the (unknown) projected mass distribution of the lens as a sum of Plummer mass distributions, of the form given by eq. (\ref{eq_sigma_plummer}). We chose the Plummer mass distribution as basis function because it is well-behaved at all radii and yields a finite total mass. The deflection angle is then simply the sum of the deflection angles caused by each individual Plummer distribution. For $N$ individual Plummer lenses, this yields the following lens equation: \begin{equation} \Vec{\beta}(\Vec{\theta}) = \Vec{\theta} - \frac{D_{ds}}{D_s D_d} \frac{4 G}{c^2} \sum_{i = 1}^N \frac{\Vec{\theta}-\Vec{\theta}_{s,i}}{|\Vec{\theta}-\Vec{\theta}_{s,i}|^2+\theta_{P,i}^2} M_i \mcm \label{eq_lenseqn_mplum} \end{equation} with $\Vec{\theta}_{s,i}$ the position of the centre of a Plummer distribution in the lens plane, $M_i$ its mass, and $\theta_{P,i}$ its angular scale-length. A given set of $R$ points in the lens plane, $\Vec{\theta}_k,\,k = 1 \ldots R$, is related to a corresponding set of $R$ points in the source plane by a matrix equation \citep{d05}. Indeed, let $\Theta$ be a vector of length $2R$, containing the coordinates of the points in the image plane, in which $x$ and $y$ components alternate. Similarly, $B$ is a vector of length $2R$ which will contain the coordinates of the corresponding points in the source plane. The masses $M_i$ of the Plummer distributions that make up the mass distribution of the lens are stored in an $N$ dimensional column vector $M$. The lens equation can then be rewritten as \begin{equation} B = \Theta - \gamma M \mcm \end{equation} with $\gamma$ a $2R \times N$ matrix whose components are given by: \begin{eqnarray} \gamma_{2k-1,l} &=& \frac{D_{ds}}{D_d D_s} \frac{4 G}{c^2} \frac{(\Vec{\theta}_k-\Vec{\theta}_{s,l})_x}{|\Vec{\theta}_k-\Vec{\theta}_{s,l}|^2+\theta_{P,l}^2} \nonumber \\ \gamma_{2k,l} &=& \frac{D_{ds}}{D_d D_s} \frac{4 G}{c^2} \frac{(\Vec{\theta}_k-\Vec{\theta}_{s,l})_y}{|\Vec{\theta}_k-\Vec{\theta}_{s,l}|^2+\theta_{P,l}^2} \mpt \end{eqnarray} The problem of inverting a gravitational lens system is thus transformed into the problem of finding the vector $M$, given the matrices $\Theta$ and $\gamma$. \subsection{Genetic algorithms} \label{genalg} With genetic algorithms, one tries to breed good solutions to a given problem. A central concept is the genome, which is an encoded representation of a possible solution. Usually, the genome will encode the parameters of a specific model. For a particular genome, there has to be some kind of measure of how adequate it fits the data. This value is usually called the fitness of the genome. The algorithm starts with a random set of genomes: the population. From this population, a new one will be created using the following procedure: \begin{itemize} \item For each genome, the fitness is calculated. \item A new set of genomes is created by combining and copying genomes of the current population. Selection of genomes in this reproduction step should favor genomes with a better fitness. \item Finally, mutations are introduced in the new population to ensure genetic variety. \end{itemize} When creating the new population, the best genome is often copied without mutations. This approach is often referred to as elitism and ensures that the best member of the new population will perform at least as well as the fittest member of the old population. Thus, generation after generation, one tries to breed increasingly better solutions to a problem. A complete overview of genetic programming techniques can be found in \cite{koza:book}. \section{The inversion method}\label{sec:invert} In the following, we discuss two key features of our inversion method: the use of a dynamic grid in the lens plane on which the Plummer lenses are positioned (which defines the matrix $\gamma$) and the genetic algorithm employed to breed the best approximation to the projected mass distribution of the lens (i.e., the vector $M$). \subsection{The dynamic grid} The procedure starts with a square grid, large enough to encompass the projected mass density of the lens. At first, this area is uniformly subdivided in square grid cells. At the centre of each grid cell, a Plummer mass distribution is positioned. The width of each Plummer distribution is set proportional to the side of its grid cell. We tested which proportionality factor allows to best reproduce a wide range of mass densities and found that a value of 1.7 yields a good trade-off between smoothness and dynamic range. The same scale factor was subsequently used in our lens inversion simulations. The genetic algorithm (see subsection \ref{sub:gen}) then breeds, for this given grid, the best solution $M$. Given this first approximation of the total mass density, a new grid is constructed by further subdividing grid cells that contain a large fraction of the total mass or that reside in areas with large density gradients. This way, the new grid will allow a better approximation of the mass density, without wasting resources on areas which contain little mass or detail. With each cell of this new grid a Plummer distribution is associated and the individual masses are determined by the genetic algorithm, as before. In our implementation, this procedure of refining the grid is repeated unless the number of grid cells exceeds one thousand. Fig. \ref{fig1} illustrates the procedure. At first, a uniform grid is used. With this grid, a first estimate of the distribution is found and this is used to create a new grid. The figure shows a few additional mass density estimates on which new grids are based. \subsection{The genetic algorithm} \label{sub:gen} \subsubsection{Genome and fitness} The goal of the genetic algorithm is to determine good values for the masses of the individual Plummer distributions which are laid out according to a specific grid. Therefore, the genome in our genetic algorithm will encode the masses of these Plummer distributions. For a specific set of Plummer masses, we need to define a way to evaluate how good the corresponding solution is. Since we are working in the strong lensing regime, it is assumed that the gravitational lens system produces multiple images of one or more sources. If one would project these images back to the source plane using the exact lens equation for the lens under study, one would find that back-projected images of the same source will overlap perfectly. For this reason, the degree in which back-projected images of the same source overlap will be used to determine how good the suggested solution actually is. The way this is implemented is as follows. For a given solution of the mass distribution of the lens, the images of a single source are projected back to the source plane. The areas occupied by each image are surrounded by rectangles: two examples are shown in Fig. \ref{fig2}. Corresponding corners of the rectangles are connected with imaginary springs. Consider two rectangles, each enclosing a backprojected image. For corresponding corners, the distance is calculated in absolute units, for example in units of arcminutes or arcseconds. In a previous step, a length scale was calculated as the average of the lengths of the sides of all the rectangles belonging to a specific source. The distance between two corresponding corners is then divided by this length, yielding a dimensionless distance $d$. The ``potential energy'' for this pair of corners is then simply $d^2$. Repeating this for the other three corners and adding together the energies then gives the potential energy of these two rectangles. For a specific source, this procedure is then done for all pairs of backprojected images and the sum of these potential energies is then the potential energy contribution of this source. The fitness value of a given lens solution is the sum of the potential energies of all sources. It is important to take into account the scaling of the rectangles when calculating the potential energy values. Comparing the left and right parts of Fig. \ref{fig2}, it is clear that the left situation definitely corresponds to a better overlap, while on an absolute scale the potential energy of the right situation will be the lower one. For this reason, we express distances between corners of rectangles relative to the size of the rectangles, or, in other words, relative to the size of the source. As was mentioned above, the genome represents the masses of the individual Plummer distributions. To be more precise, the genome only represents the relative contribution of each Plummer distribution: each Plummer mass is represented by a dimensionless, integer number between 0 and 1000. These numbers are stored in the vector $M$ and the matrix product \begin{equation} \Theta' = \gamma M \end{equation} is calculated. For the dimensionless masses to be converted into real masses, the vector $M$ needs to be multiplied with a factor $\mu$, bringing the lens equation in the form \begin{equation} B = \Theta - \mu \Theta'. \end{equation} Since $\Theta$ and $\Theta'$ are constant column matrices, it is an easy and computationally inexpensive task to find, for a given $M$, the factor $\mu$ that maximises the fitness, or, in other words, for which the back-projected images of the sources coincide best. The value of the fitness of that particular situation is then considered to be the fitness of the genome. \subsubsection{Reproduction and mutation}\label{sub:rep} In our implementation, a population of 250 genomes is used. Based on the many simulations we did (see below), 250 genomes has always led to good solutions within an acceptable amount of time. To obtain a new population, some genomes are copied from the original population while others are obtained by merging two genomes. The procedure of merging two genomes consists of a few steps which are illustrated in Fig. \ref{fig3}. At first, the values between 0 and 1000 of each genome are multiplied with their best $\mu$ value to obtain the true Plummer masses they represent. Then, for each Plummer distribution, the procedure selects at random the mass from one of the two genomes. Finally, these values are rescaled to integer numbers in such a way that the largest number is 500. When the new population is complete, mutations are introduced in some genomes. In early generations, some values are simply changed to a random number between 0 and 1000. It is for this reason that the previous step rescaled the Plummer masses to a maximum value of 500. This way, a random change of the value will also allow a considerable increase in mass for that Plummer distribution. When the best fitness values of successive generations start to converge, a new mutation rule is adopted. In this case, random integer numbers in the interval $[-200,200]$ are generated and added to some of the genomes' values. Resulting values which are negative or larger than $1000$, are set to zero or $1000$ respectively. The first mutation rule makes sure that a large range of mass densities can be inspected. When the algorithm starts to converge near a good solution, the second mutation rule assures that the algorithm can more closely approach the best solution. \subsubsection{Stopping criterion} The algorithm can be stopped if the fitness of the best genome ceases to improve significantly. We use the following stopping criterion : if the fitness of the last generation, denoted by {\tt new\_fit} fulfills the constraint $|${\tt new\_fit}$-${\tt old\_fit}$| <$ {\tt new\_fit}/50, with {\tt old\_fit} the fitness of 150 generations ago, the algorithm is stopped. We inspected that raising the factor of 50, so that the code runs longer, does not significantly change the solution, i.e. that the solution has converged. Just for safety, we implemented an upper limit of 15000 on the number of generations but all inversions we tried so far converged after less than 5000 generations. \subsection{Averaging multiple solutions} The genetic algorithm uses a random initial population, selects genomes at random and introduces random mutations. Because of this, multiple applications of the inversion procedure for a specific set of images will in general yield slightly different solutions. These solutions are all equally acceptable: in all cases, the back-projected images of a single source coincide very well with each other and with the true position of the source. Given this variety of possible solutions, it is interesting to calculate the average of a set of solutions. This averaging procedure will enhance the common characteristics of all the individual lens solutions while suppressing random fluctuations. One can also calculate the standard deviation of these individual solutions. This will identify the regions in which the solutions agree as well as the regions in which there is a lot of uncertainty about the mass density. Averaging the solutions will also increase the smoothness of the retrieved mass density. What determines the convergence of the fitness value (see section \ref{genalg}) is the amplitude of the mutations. Once the difference between the best possible lens solution and the best genome becomes comparable to or smaller than the mutation amplitude, the best genomes of subsequent generations merely scatter around some lowest achieved fitness value. Lowering the mutation amplitude when the fitness starts to converge and averaging a few tens of independent solutions both help to get as close as possible to the best possible solution. Being able to create an averaged solution is an attractive feature of our approach, but it would be of little use if the resulting mass density would not be a good solution of the inversion problem (or a worse solution than the individual solutions). Using simulations (see Sect. \ref{sec:sim}), we found that the averaged solution is indeed also a good solution, with a very high fitness, and in many cases even does a better job than many of the individual solutions. This is because the random mutations that occur during the reproduction process of the genomes, cause the best solution to oscillate around the ``true'' solution. Since averaging a set of solutions suppresses these random fluctuations, the averaged solution can be a more faithful realisation of the true solution than any of the individual solutions. Also, the inversion of a gravitational lens is clearly an ill-posed problem so it's no great supprise that multiple solutions exist. For these reasons, the averaged solution is definitely a very acceptable one. \section{Simulations}\label{sec:sim} We conducted many simulations in order to test the validity of our approach. Having full knowledge of the original lens as well as of the original sources, we can easily check the accuracy with which lenses can be reconstructed in ideal circumstances, i.e. when the redshifts of the lens and the sources are known exactly. The mass distributions of the lenses in these simulations were created by randomly adding a number of Plummer distributions. The number of sources, their positions and redshifts were also chosen at random. A wide variety of these gravitational lens systems were used to test the algorithm. In the example that we present below, a lens with mass of the order of $10^{15}\,M_\odot$ was positioned at $z = 0.45$ while the redshifts of the sources were sampled from a uniform distribution in the interval $[1.2,4.0]$. In Fig. \ref{fig4}, we show the mass distribution and the positions and shapes of the sources. The total mass of the lens within a radius of $1.5${\arcmin}, which is slightly further out than the position of the outermost image, is $0.95\times 10^{15} \,M_\odot$ and the number of sources in this simulation is $15$. This configuration was used to generate the images shown in the left panel of Fig. \ref{fig5}, which in turn serves as input for the inversion algorithm. The resolution of this image is $1024 \times 1024$ pixels. Critical lines and caustics for a source at redshift $z=2.5$ are presented in the right panel of Fig. \ref{fig5}, in which the source positions are also indicated. The genetic algorithm then constructs a lens solution that projects images of a single source onto overlapping regions in the source plane. For this particular simulation, the fitness converged for a grid containing about 400 Plummer mass distributions. As explained before, multiple applications of the inversion algorithm yield different solutions. Still, each solution manages to produce overlapping back-projected images and, while this is in no way enforced by the algorithm, the positions of the back-projected images are very close to the true source positions. After applying the inversion routine $25$ times and averaging the individual solutions, we obtained the final solution presented in the left panel of Fig. \ref{fig6}. This figure shows a striking resemblance to the left panel of Fig. \ref{fig4}. Clearly, the mass distribution of the lens is retrieved with very high accuracy. The fitness values of the 25 individual solutions and of the averaged solution are shown in the left panel of Fig. \ref{fig7}. Since averaging the individual solutions suppresses random generation-to-generation fluctuations, which can even prevent the solutions from further lowering the fitness value, and enhances their common traits, the averaged solution outperforms each individual solution. When the images of Fig. \ref{fig5} are projected back onto the source plane, we obtain the situation shown in the right panel of Fig. \ref{fig6}. The back-projected images overlap very well and are close to the true source positions. The critical lines and caustics of the averaged solution for a source at redshift $z=2.5$ are shown in the right panel of Fig. \ref{fig7}, which can be compared with the right panel of Fig. \ref{fig5}. Again, the resemblance is striking. In the left panel of Fig. \ref{fig8}, we show the absolute value of the difference between the mass distributions of the input lens and of the averaged solution. In the right panel of Fig. \ref{fig8}, the standard deviation of the 25 individual solutions is presented. The first quantity is a measure for the quality of the fit, the second measures the disagreement between the individual solutions. For a circularly symmetric lens, only the total mass enclosed within the radius of the outermost image can be determined. The lens employed in the simulation is not spherically symmetric but one can still surmise that we do not have a very good handle on the mass outside the outermost image. In Fig. \ref{fig9}, we show the circularly averaged density profiles of the input lens and of the averaged solution. As expected, both agree excellently with each other within the inner $\sim 1.5${\arcmin}, which is about the position of the outermost image. Outside that radius, the density is no longer well constrained by the data and the profile of the best lens solution drops below that of the input lens. For the averaged solution, the mass enclosed within a radius of 1.5{\arcmin} is $0.96 \times 10^{15}\,M_\odot$ which can be compared with the input lens, which comprises a mass of $0.95 \times 10^{15}\,M_\odot$ within the same radius. From this simulation and the many others we have ran, we conclude that given enough observational constraints, our method succeeds in inferring the mass distribution of a lens given the redshifts of the lens and the sources and the positions and shapes of the images. If the sources sample the caustics sufficiently well, the density profile of the lens can be reconstructed with great accuracy out to the outermost image. A very important feature of our method is that it does not minimise residuals of the lens equation, like e.g. {\sc slap}. Because tangential images are larger than radial ones, this residual is dominated by the tangential arcs and, as a consequence, methods that make use of it are less sensitive to the information contained in the radial images. This is clear from e.g. \citet{d05b}, where the non-parametrically reconstructed density profile of the cluster A1689 shows a central decline which the authors contribute to this effect. Our method is insensitive to the size of the images and thus makes full use of the information contained in the radial images. This is reflected in Fig. \ref{fig9}, which shows that the central density of the input lens is very accurately retrieved. \section{Discussion and conclusion}\label{sec:conc} The procedure described and illustrated above is a non-parametric method for inverting gravitational lenses, making no a priori assumptions regarding the shape of the lens. We only impose the condition that an acceptable solution must be able to map images of the same source onto overlapping regions in the source plane. The procedure only requires that one can identify which images correspond to the same source. In particular, no information about the sizes of the sources needs to be provided. The size of the grid on which the algorithm will determine the mass distribution of the lens needs to be specified. However, because a single solution can be obtained relatively fast, it is an easy task to try a variety of sizes until one is obtained which generates a lens with a good fitness value. A multi-resolution grid with a few hundred cells is usually sufficient to represent any plausible lens mass distribution. The implementation used for the simulation in this paper employs a population of $250$ genomes. Based on a suite of simulations, this value proved to be sufficiently large to to yield good solutions within an acceptable amount of time. The calculations were done in a distibuted manner, using sixteen Intel \textregistered{} Xeon \texttrademark{} 2.4 GHz processors of a computer cluster. Depending on the number of sources, creating a single solution may require several hours. To give a specific example, the $25$ solutions used in the simulation were created in four days. The simulation discussed in this article, together with the many others we performed, indicate that our inversion technique successfully solves the lens inversion problem. The reconstructed sources lie close to their true positions and their shapes are retrieved quite accurately as well. The procedure determines the mass of the lens very accurately. The averaging procedure guarantees the removal of random fluctuations and yields a smooth mass density very close to the true solution. Note that because the masses of the individual Plummer distributions are always represented by positive numbers in the genome, no negative mass densities will be produced. Of course, the quality of the reconstruction depends on the quality of the information at hand. The algorithm depends on the availability of multiply imaged sources, which identifies the relevant area in our procedure as the area within the outermost caustic. When this area is sampled well by the sources, we can expect a good reconstruction of the mass density, as indicated by the simulation described previously. A major advantage of the algorithm is that it makes full use of the information contained in the radial images, unlike methods that minimise the residuals of the lens equation, and is thus able to accurately reconstruct also the inner parts of the lens. Another advantage of a genetic algorithm is the ease with which the fitness criterion can be specified. One simply has to devise a way to associate a fitness value with a specific genome, without worrying about features like continuity or differentiability. E.g., to test whether information about the amplifying effect of the gravitational lens can improve the lens reconstruction, we added a contribution to the fitness value:~the maximum brightness values in the back-projected images of a single source should lie as close as possible to each other. Several simulations indicated that augmenting the procedure in this fashion does not improve the final result since the $\Vec{\beta}(\Vec{\theta})$ mapping was already very well approximated anyway. Not requiring surface brightness information not only reduces the computational cost, but also makes the method less sensitive to noise in the images. Incorporating information about the shear field at larger radii outside the gravitational lens is straightforward. For each genome the shear components $\gamma_1(\Vec{\theta})$ and $\gamma_2(\Vec{\theta})$ can be calculated and compared with the observed values. The expression for the fitness can be changed so as to penalise genomes with large differences between the measured and the model shear. Further improvements of the genetic algorithm, such as making the mutation amplitude depend automatically on the convergence of the fitness, are easily implemented and are left as future research. An application to real data will be presented in a subsequent paper. \section*{Acknowledgments} We would like to thank Prof. Philippe Bekaert and Tom Van Laerhoven of the Expertise Centre for Digital Media for granting us access to the computer cluster and for taking care of the related practical issues. We also would like to thank the anonymous referee for the valuable remarks and suggestions. They very much improved the content and presentation of this paper. \bsp \label{lastpage}
Title: Percolation Galaxy Groups and Clusters in the SDSS Redshift Survey: Identification, Catalogs, and the Multiplicity Function
Abstract: We identify galaxy groups and clusters in volume-limited samples of the SDSS redshift survey, using a redshift-space friends-of-friends algorithm. We optimize the friends-of-friends linking lengths to recover galaxy systems that occupy the same dark matter halos, using a set of mock catalogs created by populating halos of N-body simulations with galaxies. Extensive tests with these mock catalogs show that no combination of perpendicular and line-of-sight linking lengths is able to yield groups and clusters that simultaneously recover the true halo multiplicity function, projected size distribution, and velocity dispersion. We adopt a linking length combination that yields, for galaxy groups with ten or more members: a group multiplicity function that is unbiased with respect to the true halo multiplicity function; an unbiased median relation between the multiplicities of groups and their associated halos; a spurious group fraction of less than ~1%; a halo completeness of more than ~97%; the correct projected size distribution as a function of multiplicity; and a velocity dispersion distribution that is ~20% too low at all multiplicities. These results hold over a range of mock catalogs that use different input recipes of populating halos with galaxies. We apply our group-finding algorithm to the SDSS data and obtain three group and cluster catalogs for three volume-limited samples that cover 3495.1 square degrees on the sky. We correct for incompleteness caused by fiber collisions and survey edges, and obtain measurements of the group multiplicity function, with errors calculated from realistic mock catalogs. These multiplicity function measurements provide a key constraint on the relation between galaxy populations and dark matter halos.
https://export.arxiv.org/pdf/astro-ph/0601346
\title{Percolation Galaxy Groups and Clusters in the SDSS Redshift Survey: Identification, Catalogs, and the Multiplicity Function} \author{ Andreas A. Berlind, \altaffilmark{1,2} Joshua Frieman, \altaffilmark{2} David H. Weinberg, \altaffilmark{3} Michael R. Blanton, \altaffilmark{1} Michael S. Warren, \altaffilmark{4} Kevork Abazajian, \altaffilmark{4} Ryan Scranton, \altaffilmark{5} David W. Hogg, \altaffilmark{1} Roman Scoccimarro, \altaffilmark{1} Neta A. Bahcall, \altaffilmark{6} J. Brinkmann, \altaffilmark{7} J. Richard Gott III, \altaffilmark{6} S.J. Kleinman, \altaffilmark{7} J. Krzesinski, \altaffilmark{8,9} Brian C. Lee, \altaffilmark{10} Christopher J. Miller, \altaffilmark{11} Atsuko Nitta, \altaffilmark{7} Donald P. Schneider, \altaffilmark{12} Douglas L. Tucker, \altaffilmark{13} Idit Zehavi, \altaffilmark{14,15} for the SDSS collaboration } \keywords{cosmology: large-scale structure of universe --- galaxies: clusters} \altaffiltext{1}{Center for Cosmology and Particle Physics, New York University, New York, NY 10003, USA; aberlind@cosmo.nyu.edu} \altaffiltext{2}{Center for Cosmological Physics and Department of Astronomy and Astrophysics, University of Chicago, Chicago, IL 60637; frieman@fnal.gov} \altaffiltext{3}{Department of Astronomy, The Ohio State University, Columbus, OH 43210; dhw@astronomy.ohio-state.edu} \altaffiltext{4}{Theoretical Division, Los Alamos National Laboratory, Los Alamos, NM 87545} \altaffiltext{5}{Physics and Astronomy Department, University of Pittsburgh, Pittsburgh PA, 15260} \altaffiltext{6}{Department of Astrophysical Sciences, Princeton University, Princeton NJ, 08544} \altaffiltext{7}{Subaru Telescope, 650 N A'ohoku Pl., Hilo, HI 96720} \altaffiltext{8}{Apache Point Observatory, P.O. Box 59, Sunspot, NM 88349} \altaffiltext{9}{Mt. Suhora Observatory, Cracow Pedagogical University, ul. Podchorazych 2, 30-084 Cracow, Poland} \altaffiltext{10}{Lawrence Berkeley National Lab, Berkeley CA 94720} \altaffiltext{11}{Cerro-Tololo Inter-American Observatory, NOAO, Casilla 603, La Serena, Chile} \altaffiltext{12}{Department of Astronomy and Astrophysics, Pennsylvania State University, University Park, PA 16802} \altaffiltext{13}{Fermi National Accelerator Laboratory, MS 127, PO Box 500, Batavia, IL 60510} \altaffiltext{14}{Steward Observatory, University of Arizona, 933 N. Cherry Ave., Tucson AZ 85721} \altaffiltext{15}{Deptartment of Astronomy, Case Western Reserve University, Cleveland, OH 44106} \section{Introduction} \label{intro} Galaxies are gregarious by nature. Bright galaxies typically reside in groups or clusters, surrounded by less luminous neighbors. Interactions within the group or cluster environment may have important effects on the star formation history, morphology, dynamics, and other properties of member galaxies. Characterizing the relation between galaxy properties and their group environment is thus a key step in understanding galaxy formation and evolution. At the density thresholds often used to identify groups, most members should belong to the same, gravitationally bound dark matter (DM) halo.\footnote{Throughout this paper, we use the term ``halo'' to refer to a gravitationally bound structure with overdensity $\rho/\bar{\rho}\sim200$, so an occupied halo may host a single luminous galaxy, a group of galaxies, or a cluster. Higher overdensity concentrations around individual galaxies of a group or cluster constitute, in this terminology, halo substructure, or ``sub-halos''.} Recent approaches to describing the relation between galaxies and DM focus on galaxy populations of DM halos as a function of halo mass. Specifically, the bias of a particular class of galaxies can be characterized by its Halo Occupation Distribution (HOD), which specifies the probability distribution $P(N|M)$ that a halo of mass $M$ contains $N$ such galaxies, together with relations describing the relative spatial and velocity distributions of galaxies and dark matter within halos (\citealt{berlind_weinberg_02} and references therein). A well defined group catalog with well understood properties can play a central role in the empirical determination of this relation. This paper presents a group and cluster catalog defined from the Sloan Digital Sky Survey (SDSS, \citealt{york_etal_00}). While this catalog is useful for many purposes, our overriding objective is to obtain a well understood measurement of the group multiplicity function (the space density of groups as a function of richness), with the goal of determining the HOD in the high mass regime \citep{peacock_smith_00,berlind_weinberg_02,marinoni_hudson_02,kochanek_etal_03,lin_etal_04}. With this objective in mind, we have adopted a simple group-finding algorithm, friends-of-friends in redshift space \citep{huchra_geller_82}, and carried out extensive tests on realistic mock catalogs in order to assess its performance and optimize parameter choices. We apply the group-finding algorithm to volume-limited samples of galaxies so that the resulting group statistics characterize the clustering of well defined populations of galaxies. Galaxy clusters have been the focus of study since they were first seen on optical photographic plates \citep{shapley_ames_26}. \citet{zwicky_37} pioneered the study of clusters as dynamical objects by using imaging and spectroscopy of the Coma cluster to estimate its mass. However, the most influential pioneering work on clusters was done by \citet{abell_58}, who assembled the first large sample of galaxy clusters. The Abell catalog of rich galaxy clusters \citep{abell_58,abell_etal_89} was created by eyeball identification in the Palomar Observatory Sky Survey and it spawned numerous follow-up studies. \citet{devaucouleurs_71} shifted focus to poorer systems by studying nearby groups of galaxies. \citet{gott_turner_77b} made the first measurement of the group multiplicity function using the \citep{turner_gott_76} catalog of groups selected based on the projected surface density of galaxies. With the advent of large redshift surveys, group identification became three dimensional and thus less subject to projection effects. Group-finding in redshift space was pioneered by \citet{huchra_geller_82} and \citet{geller_huchra_83}, using the Center for Astrophysics (CfA) redshift survey. Subsequent versions of the CfA redshift survey were used to identify groups by various authors \citep{nolthenius_white_87,ramella_etal_89,moore_etal_93,ramella_etal_97}. Other redshift surveys that spawned group catalogs were the Nearby Galaxies Catalog \citep{tully_87}, the ESO Slice Project \citep{ramella_etal_99}, the Las Campanas Redshift Survey (LCRS) \citep{tucker_etal_00}, the Nearby Optical Galaxy Sample (NOG) \citep{giuricin_etal_00}, the Southern Sky Redshift Survey (SSRS) \citep{ramella_etal_02}, the 2dF redshift survey \citep{merchan_zandivarez_02,eke_etal_04a,yang_etal_05}, and even the high redshift DEEP2 survey \citep{gerke_etal_05}. There have been several efforts to detect clusters in the SDSS to date, most of them using the photometric data rather than the redshift data. \citet{annis_etal_99} developed the maxBCG technique, where Brightest Cluster Galaxy (BCG) candidates are identified based on their colors and magnitudes and other cluster members are selected from nearby galaxies that have the colors of the E/S0 ridgeline. \citet{kim_etal_02} developed a hybrid matched filter (HMF) technique that assumes a radial profile for clusters and convolves the data with that filter. \citet{goto_etal_02} developed the cut-and-enhance (CE) method, which selects overdensities of galaxies that have similar colors. All these techniques were applied to the early SDSS commissioning data \citep{bahcall_etal_03,goto_etal_02}. \citet{lee_etal_04} identified compact groups by looking for small and isolated concentrations of galaxies in the SDSS Early Data Release (EDR; \citealt{stoughton_etal_02}). Cluster searches in the SDSS redshift survey have also been carried out. \citet{goto_etal_05} used a friends-of-friends algorithm (though with linking lengths that do not scale with the changing number density of galaxies due to the flux limit) to identify clusters in the SDSS Data Release 2 (DR2; \citealt{abazajian_etal_04}). \citet{merchan_zandivarez_05} used a friends-of-friends algorithm to identify groups in the SDSS Data Release 3 (DR3; \citealt{abazajian_etal_05}). \citet{weinmann_etal_05} used the \citet{yang_etal_05} algorithm to identify groups in SDSS DR2. \citet{miller_etal_05} developed the C4 algorithm for finding clusters in redshift space and also applied it to the SDSS DR2. The C4 algorithm looks for concentrations of galaxies in a seven-dimensional position and color space. It takes advantage of the color similarity of cluster member galaxies and thus minimizes contamination due to projection. However, some correlations are built into the method, and modeling it in order to understand the properties of the resulting cluster catalog requires a complete model of the galaxy population (including colors and luminosities). Our method complements the C4 catalog by applying a simple and easily modeled algorithm to volume-limited samples with homogeneous properties. In \S~\ref{data} we describe the SDSS data that we use. In \S~\ref{mocks} we describe the mock catalogs that we use to optimize our group-finder and to estimate uncertainties for our measured group statistics. In \S~\ref{groupfinder} we outline our group-finding algorithm and choice of parameters. We present a detailed discussion of tests with mock catalogs in the Appendix, with the key points summarized in the main text. We discuss incompleteness in our group catalogs due to fiber collisions and survey edges in \S~\ref{incompleteness}. The group catalogs are published in electronic tables and their contents are described in \S~\ref{catalog}. Finally, in \S~\ref{multiplicity}, we present our measured group multiplicity function. We will use this to constrain the HOD in future work. We summarize our results in \S~\ref{summary}. \section{Data} \label{data} \subsection{SDSS} The SDSS is a large imaging and spectroscopic survey that is mapping two-fifths of the Northern Galactic sky and a smaller area of the Southern Galactic sky, using a dedicated 2.5 meter telescope \citep{gunn_etal_06} at Apache Point, New Mexico. The survey uses a photometric camera \citep{gunn_etal_98} to scan the sky simultaneously in five photometric bandpasses \citep{fukugita_etal_96,smith_etal_02} down to a limiting $r$-band magnitude of $\sim22.5$. The imaging data are processed by automatic software that does astrometry \citep{pier_etal_03}, source identification, deblending and photometry \citep{lupton_etal_01,lupton_05}, photometric calibration \citep{hogg_etal_01,smith_etal_02,tucker_etal_05}, and data quality assessment \citep{ivezic_etal_04}. Algorithms are applied to select spectroscopic targets for the main galaxy sample \citep{strauss_etal_02}, the luminous red galaxy sample \citep{eisenstein_etal_01}, and the quasar sample \citep{richards_etal_02}. The main galaxy sample is approximately complete down to an apparent $r$-band Petrosian magnitude limit of $<17.77$. Targets are assigned to spectroscopic plates using an adaptive tiling algorithm \citep{blanton_etal_03a}. Finally, spectroscopic data reduction pipelines produce galaxy spectra and redshifts. We use the large-scale structure sample \texttt{sample14} from the NYU Value Added Galaxy Catalog (NYU-VAGC; \citealt{blanton_etal_04a}) as our primary galaxy sample. Galaxy magnitudes are corrected for Galactic extinction \citep{schlegel_etal_98} and absolute magnitudes are k-corrected \citep{blanton_etal_03b} and corrected for passive evolution \citep{blanton_etal_03c} to rest-frame magnitudes at redshift $z=0.1$. A significant fraction of the sample that we use was made publicly available with the SDSS Data Release~3 \citep{abazajian_etal_05}. \begin{table} \begin{center} \centerline{\small Table~1. Volume-limited Sample Parameters} \begin{tabular}[t]{lccccc} \tableline \tableline Name & $\zmin$ & $\zmax$ & $<\Mr$ & $\Ng$ & $\ng$ \\ \tableline $Mr20$ & 0.015 & 0.100 & -19.9 & 57138 & 0.00673 \\ $Mr19$ & 0.015 & 0.068 & -19.0 & 37820 & 0.01396 \\ $Mr18$ & 0.015 & 0.045 & -18.0 & 18895 & 0.02434 \\ \tableline \label{tab:samples} \end{tabular} \end{center} Note---Absolute magnitude thresholds listed are for $\zmax$. $\ng$ is in units of $\hden$. \end{table} The galaxy redshift sample has an incompleteness due to the mechanical restriction that spectroscopic fibers cannot be placed closer to each other than their own thickness. This fiber collision constraint makes it impossible to obtain redshifts for both galaxies in pairs that are closer than $55''$ on the sky. In the case of a conflict, the target selection algorithm randomly chooses which galaxy gets a fiber \citep{strauss_etal_02}.\footnote{In cases where a target galaxy fiber collides with a target quasar fiber, priority is always given to the quasar, but such collisions only constitute $\sim 5\%$ of all cases.} Spectroscopic plate overlaps alleviate this problem to some extent, but fiber collisions still account for a $\sim 6\%$ incompleteness in the main galaxy sample. Since this incompleteness is most severe in regions of high galaxy density, it is necessary to correct for it in studies of groups and clusters. We correct for fiber collisions by giving each collided galaxy the redshift of its nearest neighbor on the sky (usually the galaxy it collided with), and we show in \S~\ref{incompleteness} that this procedure is adequate for our purposes. Putting collided galaxies at the redshifts of their nearest neighbors will cause some nearby galaxies to be placed at high redshift, artificially making their estimated luminosities very high. Since the abundance of highly luminous galaxies is low, this contamination can become a significant fraction of all highly luminous galaxies. For this reason, we also give collided galaxies the magnitudes (in addition to the redshifts) of their nearest neighbors. The resulting luminosity distribution is thus unbiased. There is some additional incompleteness due to bright foreground stars blocking background galaxies, but this is at the $\sim 1\%$ level. In order to limit the effects of incompleteness on our group identification, we restrict our sample to regions of the sky where the completeness (ratio of obtained redshifts to spectroscopic targets) is greater than $90\%$. Our final sample covers 3495.1 square degrees on the sky and contains 298729 galaxies. \subsection{Volume-limited Samples} In this and subsequent papers, we are primarily interested in using galaxy groups to constrain the properties of galaxies as a function of their underlying dark matter halo mass. It is therefore important that the population of galaxies constituting the groups is homogeneous within the sample volume. For this reason, we construct volume-limited subsamples of the full SDSS redshift sample that are each complete in a specified redshift range down to a limiting $r$-band absolute magnitude threshold. We construct each sample by choosing redshift limits $\zmin$ and $\zmax$, and only keeping galaxies whose evolved, redshifted spectra would still make the redshift survey's apparent magnitude and surface brightness cuts at the limiting redshifts of the sample. Since the apparent magnitude limit of the redshift sample varied across the sky in the commissioning phases of the survey, we cut the $r$-band magnitude limit from $\sim17.77$ back to 17.5. This more conservative limit is uniform across the sky. We construct three such volume-limited samples. Figure~\ref{fig:vollim} shows these samples in the luminosity-redshift plane. Each dot in the figure shows a galaxy in the SDSS redshift survey. The sharp cutoff curve along the lower-right part of the plot shows our $r=17.5$ apparent magnitude limit. We select three redshift ranges for our volume-limited samples: $0.015-0.1$, $0.015-0.068$, and $0.015-0.045$. These samples are complete down to absolute $r$-band magnitudes of $\Mr<-19.9$, $-19$, and $-18$, respectively.\footnote{All absolute magnitudes are quoted for $\Omegam=0.3$, $\Omegal=0.7$, and a value of the Hubble constant $h \equiv H_0 / 100\Hunits = 1$. For other values of $H_0$, one should add $5\log h$ to the quoted absolute magnitudes.} We refer to these samples as $Mr20$, $Mr19$, and $Mr18$, henceforth. Regions of the plot that make it into these three samples are shown in blue, green, and red, respectively. The limiting absolute magnitude of each sample changes slightly with redshift due to the passive evolution corrections applied to galaxy luminosities: as a galaxy is moved to the outer edge of a given volume-limited sample, its luminosity increases somewhat, allowing lower redshift galaxies to make it into the sample at lower luminosities than they do at higher redshifts. We choose the first limiting redshift of $\zmax=0.1$ because this yields the largest possible volume-limited sample (largest number of galaxies). We choose lower redshift samples in order to probe galaxy populations less luminous than $\Lstar$. We use a lower redshift limit of $0.015$ for all three samples to alleviate some of the problems associated with obtaining accurate photometry of nearby highly extended galaxies. The redshift limits, luminosity thresholds at $\zmax$, number of galaxies, and space densities of these samples are listed in Table~1. Figure~\ref{fig:skyvollim} shows a Hammer (equal area) projection (in equatorial coordinates) of sample $Mr20$. Points represent galaxies in the sample. The curve shows the location of the Galactic plane. The figure illustrates the patchy and non-uniform nature of the sample footprint on the sky, which has irregular edges, as well as multiple holes. This irregularity exacerbates systematic errors due to edge effects. We deal with incompleteness due to edge effects in \S~\ref{incompleteness}. Figure~\ref{fig:slicevollim} shows an equatorial slice through sample $Mr20$. The slice is $4^\circ$ thick and each point shows the RA and redshift of a galaxy in the sample. Prominent in this projection of the data is the the giant supercluster at $z\sim0.08$ at the left end of the Sloan Great Wall of Galaxies, which extends from longitude 132 degrees (at $z\sim0.05$) to longitude 210 degrees (at $z\sim0.08$) (See \citealt{gott_etal_05}). \section{Mock Catalogs} \label{mocks} Our main scientific motivation for constructing group catalogs from the SDSS data requires that identified groups most closely resemble systems of galaxies that occupy a common dark matter halo. Moreover, it is important that we statistically quantify the degree to which our groups do not satisfy this criterion. For both these reasons, it is imperative that we use mock galaxy catalogs that are constructed by populating dark matter halos in N-body simulations with mock galaxies. The N-body simulations must satisfy two basic conditions: they must contain a large enough volume to fit our largest volume-limited sample, $Mr20$, and they must resolve the smallest mass halos that can host a galaxy in our least luminous volume-limited sample, $Mr18$. HOD fits to the SDSS two-point correlation function of galaxies suggest that the minimum dark matter halo mass that can host a galaxy of luminosity $\Mr\sim -18$ is approximately $2\times 10^{11}\hMsun$ \citep{zehavi_etal_05,tinker_etal_05}. Requiring that a halo contain at least forty dark matter particles to be resolved means that we need N-body simulations with particle masses less than $5\times 10^{9}\hMsun$. We use a series of N-body simulations of a $\Lambda$CDM cosmological model, with $\Omegam=0.3$, $\Omegal=0.7$, $\Omegab=0.04$, $h\equiv H_0/(100~\mathrm{km~s}^{-1}~\mathrm{Mpc}^{-1})=0.7$, $n_s=1.0$, and $\sigma_8=0.9$. This model is in good agreement with a wide variety of cosmological observations (see, e.g., \citealt{spergel_etal_03,tegmark_etal_04b, abazajian_etal_05b}). Initial conditions were set up using the transfer function calculated for this cosmological model by CMBFAST \citep{seljak_zaldarriaga_96}. The simulations were run at Los Alamos National Laboratory (LANL) using the Hashed-Oct-Tree (HOT) code \citep{warren_Salmon_93}. We use a total of six independent simulations of varying size and resolution, which we refer to as \texttt{LANL1-6}. The size of box $\Lbox$, number of particles $N_\mathrm{p}$, and resulting particle mass $m_\mathrm{p}$ for each simulation are listed in Table~2. The gravitational force softening is $\epsilon_{\rm grav}=12\hkpc$ (Plummer equivalent). \begin{table*} \begin{center} \centerline{\small Table~2. Mock Catalog Parameters} \begin{tabular}{l|cccc|cccc} \tableline \tableline & \multicolumn{4}{c|}{N-body} & \multicolumn{4}{c}{HOD} \\ Mock & Name & $\Lbox$ & $N_\mathrm{p}$ & $m_\mathrm{p}$ & $\Mmin$ & $\Mcut$ & $M_1$ & $\alpha$ \\ & & ($\hmpc$) & & ($10^9\hMsun$) & ($10^{11}\hMsun$) & ($10^{13}\hMsun$) & ($10^{12}\hMsun$) & \\ \tableline \texttt{LANL1.Mr20} & \texttt{LANL1} & $384$ & $1024^3$ & 4.39 & 10.0 & --- & 25.0 & 1.1 \\ \texttt{LANL1.Mr20b} & & & & & 9.08 & 1.14 & 12.3 & 0.9 \\ \texttt{LANL1.Mr19} & & & & & 3.7 & --- & 8.2 & 1.0 \\ \texttt{LANL1.Mr18} & & & & & 1.9 & --- & 3.4 & 0.9 \\ \tableline \texttt{LANL2.Mr20} & \texttt{LANL2} & $384$ & $1024^3$ & 4.39 & 10.0 & --- & 25.0 & 1.1 \\ \texttt{LANL2.Mr20b} & & & & & 9.08 & 1.14 & 12.3 & 0.9 \\ \texttt{LANL2.Mr19} & & & & & 3.7 & --- & 8.2 & 1.0 \\ \texttt{LANL2.Mr18} & & & & & 1.9 & --- & 3.4 & 0.9 \\ \tableline \texttt{LANL3.Mr20} & \texttt{LANL3} & $384$ & $1024^3$ & 4.39 & 10.0 & --- & 25.0 & 1.1 \\ \texttt{LANL3.Mr20b} & & & & & 9.08 & 1.14 & 12.3 & 0.9 \\ \texttt{LANL3.Mr19} & & & & & 3.7 & --- & 8.2 & 1.0 \\ \texttt{LANL3.Mr18} & & & & & 1.9 & --- & 3.4 & 0.9 \\ \tableline \texttt{LANL4.Mr20} & \texttt{LANL4} & $400$ & $1280^3$ & 2.54 & 10.0 & --- & 25.0 & 1.1 \\ \texttt{LANL4.Mr20b} & & & & & 9.08 & 1.14 & 12.3 & 0.9 \\ \texttt{LANL4.Mr19} & & & & & 3.7 & --- & 8.2 & 1.0 \\ \texttt{LANL4.Mr18} & & & & & 1.9 & --- & 3.4 & 0.9 \\ \tableline \texttt{LANL5.Mr20} & \texttt{LANL5} & $543$ & $1024^3$ & 12.4 & 10.0 & --- & 25.0 & 1.1 \\ \texttt{LANL5.Mr20b} & & & & & 9.08 & 1.14 & 12.3 & 0.9 \\ \tableline \texttt{LANL6.Mr20} & \texttt{LANL6} & $768$ & $1024^3$ & 35.1 & 10.0 & --- & 25.0 & 1.1 \\ \tableline \label{tab:mocks} \end{tabular} \end{center} \end{table*} We identify halos in the dark matter particle distributions using a friends-of-friends algorithm with a linking length equal to $0.2$ times the mean interparticle separation. We then populate these halos with galaxies using a simple model for the HOD of galaxies more luminous than a luminosity threshold. Every halo with a mass $M$ greater than a minimum mass $\Mmin$ gets a central galaxy that is placed at the halo center of mass and is given the mean halo velocity. A number of satellite galaxies is then drawn from a Poisson distribution with mean $\Nsat = ((M-\Mmin)/M_1)^\alpha$, for $M\geq\Mmin$. These satellite galaxies are assigned the positions and velocities of randomly selected dark matter particles within the halo. In order to construct mock catalogs for each of our three volume-limited samples $Mr20$, $Mr19$, and $Mr18$, we select sets of values for the parameters $\Mmin$, $M_1$, and $\alpha$ that yield the observed \citet{zehavi_etal_05} galaxy-galaxy correlation functions for these samples. These HOD parameter values are similar to the best-fit values given by \citet{zehavi_etal_05} (they are slightly different because the model for $\Nsat$ was different in that paper). We refer to these sets of mock catalogs with the suffixes \texttt{.Mr20}, \texttt{.Mr19}, and \texttt{.Mr18}. In addition to these mock catalogs, we construct a set of catalogs for the $Mr20$ sample using an alternative HOD model, where the mean number of satellites in a halo of mass $M$ is $\Nsat = \mathrm{exp}[-\Mcut/(M-\Mmin)] (M/M_1)^\alpha$, for $M>\Mmin$ (also used by \citealt{tinker_etal_05}). We fix the value of the slope $\alpha$ to 0.9, which is lower than that for the \texttt{.Mr20} mocks, and we choose values for the remaining HOD parameters that yield the observed \citet{zehavi_etal_05} correlation function of $\Mr<-20$ galaxies. We refer to these sets of mock catalogs with the suffix \texttt{.Mr20b}. The values for all mock HOD parameters are listed in Table~2. We construct ten realizations of each mock catalog listed in Table~2 by using different random number generator seeds when we (a) draw a number of satellite galaxies for each halo from a Poisson distribution of mean $\Nsat$, and (b) select random dark matter halo particles to give their positions and velocities to these satellite galaxies. The dispersion among the ten realizations for one mock catalog therefore represents the scatter among possible observed states for a given halo distribution and HOD model. We now have a set of mock catalogs containing galaxies in real space and in the cubical geometry of the N-body simulations. We refer to these as our ``real-space cube mocks''. We create a redshift-space version of these catalogs by assuming the distant observer approximation and aligning the line-of-sight along one of the axes of the simulation cubes. We use the mock galaxies' peculiar velocities to move them along the line-of-sight into redshift space. We refer to the resulting mock catalogs as our ``redshift-space cube mocks''. We use these real-space and redshift-space cube mocks to determine optimal parameters for our group-finding algorithm. We summarize this determination in \S\ref{groupfinder} and discuss details in the Appendix. For the purpose of studying the effects of SDSS incompleteness on our measured groups, as well as for obtaining estimates of the uncertainty in our measured group multiplicity function, we also require mock catalogs that have the same geometry as our SDSS volume-limited samples. The total volume of our largest sample, $Mr20$, is approximately $210^3\hvol$, which is more than six times smaller than any of our mock cubes. However, the SDSS geometry is highly irregular (as seen in Fig.~\ref{fig:skyvollim}) and can only be fully embedded in a cube of much larger volume than the survey itself. The $Mr20$ sample, for example, has a maximum extent of $\sim600\hmpc$ when both the North and South Galactic portions are included. In order to carve this sample geometry out of our mock catalogs, we create mock cubes with eight times larger volume by tiling each mock cube $2\times2\times2$. Since the N-body simulations used to construct the mocks were run with periodic boundary conditions, we can tile the cubes without having density discontinuities at the boundaries. We set the center of this tiled cube to be the origin and put galaxies into redshift space using the line-of-sight component of their peculiar velocities. We then compute RA, DEC, and redshift coordinates for every mock galaxy in the tiled cube. Finally, we only keep galaxies whose coordinates on the sky would place them in regions of the SDSS survey that have completeness greater than $90\%$, and whose redshifts lie within the redshift limits of the specific volume-limited sample we are constructing mock catalogs for. Since the volume of each simulation cube is at least six times larger than our largest volume-limited sample $Mr20$, we try to carve out as many independent volumes with the $Mr20$ geometry as possible without too much overlap. We do this by performing many sets of three rotations (one around each Cartesian axis) and testing how much overlap the resulting catalogs have with each other (i.e., how many common mock galaxies do they share). With the right combination of rotation angles, we can carve out two $Mr20$ mock catalogs that share fewer than $3\%$ of their galaxies with each other, but we cannot obtain more without significant overlap. We create two such independent mock catalogs, with the correct SDSS geometry, from every one of the ten HOD realizations of the mock cubes listed in Table~2, except for the \texttt{LANL6.Mr20} mock. This procedure yields 200 mock catalogs for the $Mr20$ sample (5 N-body simulations $\times$ 2 HOD models $\times$ 10 HOD realizations $\times$ 2 mocks per simulation cube), and 80 mock catalogs each for the $Mr19$ and $Mr18$ samples (4 N-body simulations $\times$ 1 HOD model $\times$ 10 HOD realizations $\times$ 2 mocks per simulation cube). The final step in creating mock SDSS catalogs is to incorporate the fiber collision constraint. We use a friends-of-friends algorithm to identify groups of mock galaxies that are linked together by the $55''$ minimum angular separation of fibers. We then select ``collided'' mock galaxies (whose redshifts will be unknown) in each such collision group in a way that minimizes the number of such galaxies. For example, if a collision group contains three galaxies in a row, where the first is closer than $55''$ from the second and the second is closer than $55''$ from the third, but the first is more than $55''$ from the third, we will always select the middle galaxy to be the collided one. In cases where multiple choices yield the same number of collided galaxies, we select randomly (e.g., in collision groups with only two galaxies). This procedure is designed to mimic the tiling code that assigns spectroscopic fibers to SDSS target galaxies \citep{blanton_etal_03a}. If we perform this operation on the \texttt{.Mr20} catalogs we end up with only $\sim 3\%$ of mock galaxies being tagged as collided. This is about half the fraction of SDSS galaxies in our $Mr20$ sample that don't have measured redshifts due to fiber collisions. The reason for this discrepancy is that galaxies in the $Mr20$ volume-limited sample do not only collide with each other; they also collide with galaxies more luminous than $\Mr\sim-20$ at redshifts higher than the sample limit $z=0.1$ and galaxies less luminous than $\Mr\sim-20$ at lower redshifts. Most of these additional galaxies that can collide with a given galaxy in $Mr20$ are uncorrelated background or foreground galaxies. It is therefore sufficient to model them as a background screen of galaxies on the sky that have an angular correlation function equal to the mean for all SDSS galaxies. For this purpose, we use the very large volume \texttt{LANL6.Mr20} cube mock. We use \texttt{LANL6.Mr20} to construct a ``screen'' catalog with the correct SDSS angular geometry and a variable outer redshift limit, and superpose it onto each of our \texttt{.Mr20}, \texttt{.Mr19}, and \texttt{.Mr18} mock catalogs. We then allow all galaxies to collide with each other and keep track of collided mock galaxies. We set the outer redshift limit of the screen catalog to the value that results in $\sim 6\%$ of mock galaxies being tagged as collided. We find that we need approximately seven times more galaxies in the screen catalog than in the mocks in order to achieve this collided fraction. Using this approach we construct three versions of every mock catalog described above: a version with no fiber collisions applied (``true'' version), a version where collided galaxies have no redshifts and are dropped out of the mock catalog altogether (``uncorrected'' version), and a version where collided galaxies are assigned the redshift of the galaxy they collided with (``corrected'' version). These mock catalogs allow us to test the effects of fiber collisions on our measured group multiplicity function (discussed in \S~\ref{incompleteness}.) \section{Group-Finding Algorithm} \label{groupfinder} We wish to identify galaxy groups primarily in order to measure the group multiplicity function and use it to constrain the HOD of galaxies as a function of galaxy properties. This goal places a number of demands on the group-finding algorithm: (1) It should identify galaxy systems that occupy the same dark matter halos with the least possible merging of different halos into the same group and the least possible splitting of individual halos into multiple groups. (2) It should produce a group multiplicity function that is unbiased with respect to the halo multiplicity function. (3) It should be simple and well-defined so that the statistical and systematic uncertainty in the measured group multiplicity function can be accurately characterized. (4) It should use only the spatial positions of galaxies in redshift space to identify groups, and not galaxy properties such as color or luminosity. These requirements point to an algorithm that uniquely identifies density enhancements in redshift space. We adopt the simple and well understood friends-of-friends (FoF) algorithm, where galaxies are recursively linked to other galaxies within a specified linking volume around each galaxy. The FoF algorithm has several attractive features. First, for a given linking volume (usually specified by one linking length in real space and two linking lengths in redshift space), FoF produces a unique group catalog. Second, it does not assume or enforce any particular geometry for groups (e.g., spherical), but rather identifies structures that are approximately enclosed by an isodensity surface whose density is monotonically related to the linking lengths. Third, the algorithm satisfies a nesting condition: all the members of a group identified with one set of linking lengths are also members of the same group identified using larger linking lengths. The FoF algorithm has been used extensively to identify dark matter halos in N-body simulations (e.g., \citealt{davis_etal_85}) and has been shown to produce halo catalogs with mass functions that are close to universal (within $\sim 20\%$) for a wide range of epochs and cosmological models \citep{jenkins_etal_01}. FoF has also been the most used algorithm for identifying galaxy groups in redshift surveys \citep{huchra_geller_82,geller_huchra_83,nolthenius_white_87,ramella_etal_89, moore_etal_93,ramella_etal_97,ramella_etal_99,tucker_etal_00,giuricin_etal_00, ramella_etal_02,merchan_zandivarez_02,eke_etal_04a}, though alternative methods have also been used (see e.g., \citealt{tully_87,marinoni_etal_02,gerke_etal_05, yang_etal_05}). These FoF studies all used the same basic algorithm, but differed in their choices for linking lengths and in their methods for dealing with the varying density of galaxies inherent in flux-limited surveys. We use the basic \citet{huchra_geller_82} algorithm, where two galaxies are linked to each other if both their transverse and line-of-sight separations are smaller than a given pair of projected and line-of-sight linking lengths, respectively. Specifically, two galaxies $i$ and $j$ with angular separation $\theta_{ij}$ and redshifts $z_i$ and $z_j$, have a projected separation $D_{\perp,ij}$ and a line-of-sight separation $D_{\parallel,ij}$ (both in $\hmpc$) given by \footnote{We use these simple equations, rather than the exact formulae for the redshift-distance and angular diameter-distance relations because, at $z=0.1$ (the outer limit of our sample), the difference between these formulae is less than $1\%$.} \begin{eqnarray} D_{\perp,ij} & = & (c/H_0)(z_i+z_j)~\mathrm{sin}(\theta_{ij}/2), \\ D_{\parallel,ij} & = & (c/H_0)|z_i-z_j|. \end{eqnarray} The two galaxies are then linked to each other if \begin{equation} D_{\perp,ij} \leq \bperp~\ng^{-1/3} \end{equation} and \begin{equation} D_{\parallel,ij} \leq \bpar~\ng^{-1/3}, \end{equation} where $\ng$ is the mean number density of galaxies, and $\bperp$ and $\bpar$ are the projected and line-of-sight linking lengths in units of the mean intergalaxy separation. Since we use volume-limited samples of SDSS galaxies, $\ng$ is constant throughout the sample volumes, and thus the linking lengths are also constant. The resulting linking volume around each galaxy is very similar to a cylinder, oriented along the line-of-sight, whose radius is equal to the projected linking length and whose height is equal to twice the line-of-sight linking length. It is not a perfect cylinder because its radius increases with redshift, making it slightly wider at the far end than at the near end, and its bases are slightly curved. However, for the small linking lengths considered here, a cylinder is a good approximation. The FoF algorithm works recursively, whereby a galaxy is linked to all its ``friends'', which are in turn linked to their ``friends'', etc., to yield a unique group of galaxies. \subsection{Choice of Linking Lengths} The most important ingredient of our group-finding algorithm is our choice for the linking lengths $\bperp$ and $\bpar$. If the linking lengths are too small, then the group-finder will break up single halos into multiple groups. If the linking lengths are too large, then different halos will be fused together into single groups. There are no values for the linking lengths that will work perfectly for every halo, even in real space. In redshift space this problem becomes substantially worse, since redshift-space distortions both move halos and elongate them along the line-of-sight, often causing them to overlap with each other. The right choice of linking lengths depends on the purpose for which groups are being identified. If we require a group catalog that is highly inclusive and groups together every galaxy inhabiting the same halo, then we will use larger linking lengths than if our goal is to minimize contamination by galaxies that come from different halos. For our purposes, we wish to obtain a balance between being inclusive and reducing contamination, while producing groups that have an unbiased multiplicity function. In order to find the right combination of linking lengths, we use the mock galaxy catalogs described in \S~\ref{mocks}. Specifically, we use the real- and redshift-space cube mocks, which are constructed by applying simple HOD models to the \texttt{LANL1} and \texttt{LANL4} N-body simulations. Since we know which mock galaxies occupy the same dark matter halos, we can evaluate how well a particular choice of linking lengths recovers features of the halo population. The mocks that we use here have a cubical geometry, and we assume the distant observer approximation when we put mock galaxies into redshift space. We use the full cubical mocks rather than those with the correct SDSS geometry because the full mocks have a much larger volume and thus better statistics. Moreover, our goal is to find the best linking lengths for any redshift survey, and we will deal with systematic effects specific to our SDSS sample geometry separately. The FoF algorithm that we use is therefore slightly different from the one outlined above, in that the linking volume is a perfect cylinder (i.e., $D_{\perp,ij}$ is simply the projected distance between two mock galaxies). We run the FoF group-finder on the mock catalogs for a grid of linking length values, and we study the properties of the resulting group catalogs. Specifically, we investigate four features of the recovered group distribution: (1) the group multiplicity function compared to the ``true'' halo multiplicity function; (2) The relation between the number of galaxies in a halo $\Ntrue$ and the number of galaxies in its associated group $\Nobs$; (3) The distribution of projected group sizes as a function of group richness compared to the ``true'' distribution of projected halo sizes as a function of halo multiplicity; (4) The distribution of group velocity dispersions as a function of group richness compared to the ``true'' distribution of halo velocity dispersions as a function of halo multiplicity. We check how each set of linking lengths performs in the above four tests, for each of the four HOD model mock cubes (\texttt{.Mr20, .Mr20b, .Mr19, .Mr18}). In the case of each HOD model, we average results over the 10 HOD realizations described in \S~\ref{mocks} and over the \texttt{LANL1} and \texttt{LANL4} N-body simulations. We do this procedure for groups that are identified in both real space (for which there is only one linking length), and redshift space. These tests are described in detail in the Appendix. Here we summarize the main results. In real space, a linking length choice of $b=0.2$ yields galaxy groups with ten or more members that pass all four tests listed above. Groups with $N<10$ show systematic deviations in abundance, multiplicity, projected sizes, and velocity dispersions from the corresponding halos with $N<10$. The choice of $b=0.2$ is not surprising, given that the same linking length was used to identify halos in the N-body simulations. It is also not surprising that the group-finding fails the tests for small groups, where adding or losing a couple of galaxies makes a large fractional difference to the group size. The threshold of $N\sim 10$ is independent of the underlying dark matter halo mass. This means that we can push the regime in which the groups are reliable to lower mass systems by using a lower luminosity sample (where each halo will contain more galaxies). Of course, the change of luminosity threshold comes at the expense of statistical power, since low luminosity samples have smaller volumes than high luminosity samples. The number of groups in a volume-limited sample scales roughly with the number of galaxies, and a luminosity threshold near the characteristic luminosity $L_*$ maximizes this number. In redshift space the situation is more complicated. No set of transverse and line-of-sight linking lengths is able to produce groups that pass all four tests listed above, even for large size groups. Figure~\ref{fig:linkinglengths.hod20} summarizes our tests for the \texttt{.Mr20} HOD model mocks. Results for the other HOD models are similar and are shown in the Appendix. The figure shows regions (shaded) of the two-dimensional linking length space ($\bpar$ vs. $\bperp$) that pass each of our four tests. \subsubsection{Multiplicity Function} The dark and thin shaded region in Figure~\ref{fig:linkinglengths.hod20}, labeled $n(N)$, shows linking lengths that pass the group multiplicity function test. In other words, these linking lengths yield mock group catalogs whose multiplicity functions are unbiased relative to the ``true'' input halo multiplicity function, in the regime $N\geq 10$. In this case, ``unbiased'' means that the shape of the multiplicity function is on average the same as the ``true'' shape and its amplitude is within $10\%$ of the ``true'' amplitude. Linking length values that lie along the upper boundary of the shaded region (e.g, the values $\bperp=0.11$, $\bpar=1.5$) yield multiplicity functions that are $10\%$ too high in amplitude, whereas values that lie along the lower boundary yield multiplicity functions whose amplitudes are $10\%$ too low. These results show that an increase in either linking length generally leads to an increase in the multiplicity function for $N\geq 10$. This increase is compensated for by a corresponding decrease in the abundance of isolated (i.e., $N=1$) and low $N$ groups. The shaded region appears to be close to horizontal only because the vertical axis is highly compressed with respect to the horizontal axis. \subsubsection{$\Ntrue$ vs. $\Nobs$} The group multiplicity function is an average statistic showing the abundance of all groups as a function of $N$. It is therefore possible, in principle, for it to be unbiased relative to the halo multiplicity function, without the relation between individual halo multiplicities and their recovered group multiplicities being correct. For this reason, we also require that the group-finder yield an unbiased relation between the multiplicity of individual halos, $\Ntrue$, and their recovered groups, $\Nobs$. In order to check this, we must match input halos to recovered groups in a one-to-one way. There are many ways to do this matching, and no one way is more correct than another. For example, a halo can be associated with the group that contains most of its galaxies, or the group that contains its central galaxy, or the group whose centroid is closest to the halo center. We associate each halo to the group that contains its central galaxy. When two or more halos are matched to the same group, we choose the halo that shares the largest number of common galaxies with the group. Halos that are not associated with any group are considered ``undetected,'' and groups that are not associated with any halo (because they don't contain any halo central galaxies) are considered ``spurious''. The light (and green) shaded region in Figure~\ref{fig:linkinglengths.hod20} that roughly tracks and is slightly wider than the $n(N)$ region shows linking lengths that pass the $\Ntrue$ vs. $\Nobs$ test. In other words, these linking lengths yield mock group catalogs with an unbiased median relation between $\Ntrue$ and $\Nobs$ for associated halos and groups, in the regime $N\geq 10$. We consider the relation to be unbiased if its slope is within $10\%$ of unity. Linking length values that lie along the upper boundary of the shaded region yield associated halos and groups with a median relation $\Ntrue=1.1\Nobs$, whereas values that lie along the lower boundary yield the relation $\Ntrue=0.9\Nobs$. As expected, most linking lengths that pass the multiplicity function test also pass the $\Ntrue$ vs. $\Nobs$ test. This breaks down, however, for values of $\bperp$ greater than 0.16-0.17. \subsubsection{Projected Sizes} The (blue) shaded region in Figure~\ref{fig:linkinglengths.hod20}, labeled ``Projected sizes'', shows linking lengths that pass the projected sizes test. These linking lengths yield mock groups with an unbiased median relation between rms projected size and group multiplicity $N$, in the regime $N\geq 10$. We consider the relation to be unbiased if it is within $10\%$ of the ``true'' relation between median rms projected halo size and halo multiplicity. This shaded region is roughly vertically oriented because the projected linking length $\bperp$ affects the projected sizes of groups much more than the line-of-sight linking length $\bpar$. Clearly, increasing $\bperp$ leads to galaxy groups with larger projected sizes. The shaded region is not completely vertical, however, because increasing $\bpar$ also leads to larger projected size groups, albeit in a much less sensitive way. \subsubsection{Velocity Dispersions} The (red) shaded region in Figure~\ref{fig:linkinglengths.hod20}, labeled ``Velocity dispersions'', shows linking lengths that pass the velocity dispersion test. These linking lengths yield mock groups with an unbiased median relation between group velocity dispersion and group multiplicity $N$, in the regime $N\geq 10$. We consider the relation to be unbiased if it is within $10\%$ of the ``true'' relation between median halo velocity dispersion and halo multiplicity. This shaded region is roughly horizontally oriented because the line-of-sight linking length $\bpar$ affects the velocity dispersions of groups much more than $\bperp$. Clearly, increasing $\bpar$ leads to galaxy groups with larger velocity dispersions. The shaded region is not completely horizontal, because changing $\bperp$ also affects the velocity dispersions of groups, though not consistently in the same sense. \subsubsection{Our Adopted Linking Lengths} It is obvious from Figure~\ref{fig:linkinglengths.hod20} that no combination of FoF linking lengths passes all four tests listed above. We can choose linking lengths that successfully recover the abundance and projected sizes, or the abundance and velocity dispersions of groups as a function of multiplicity, but not all three simultaneously. We can also choose linking lengths that successfully recover both the projected sizes and velocity dispersions of groups as a function of multiplicity, but since the multiplicity function of such groups is incorrect, the overall size and velocity dispersion distributions will also be incorrect. This failure to recover all features of groups in redshift space is a fundamental shortcoming of the FoF group-finder when applied to redshift space. Given that most redshift-space group-finding algorithms operate on very similar principles, i.e., they identify overdense regions that are elongated along the line-of-sight, it is likely that this shortcoming is shared by other group-finders as well. To our knowledge, no group-finder has been shown to pass all four of the tests considered here for a single choice of parameters. Figure~\ref{fig:linkinglengths.hod20} shows that in order to recover groups with unbiased velocity dispersions, the line-of-sight linking length must be substantially larger than the mean intergalaxy separation. With $\bpar$ that large, groups are bound to be linked together along the line-of-sight. The only way to then obtain groups with the correct multiplicity function is to have a transverse linking length small enough that galaxies in the outer parts of halos are not included in the recovered groups. The resulting groups bear little physical resemblance to their parent halos. If, on the other hand, we recover groups with unbiased projected sizes, then the groups will be missing some of their fastest moving galaxies and this decrease in multiplicity will be compensated by including as group members a few galaxies in the infall regions of halos. These groups are much more physically similar to their parent halos. For this reason, we choose to sacrifice velocity dispersions, rather than projected sizes, when selecting values for the FoF linking lengths. Figure~\ref{fig:linkinglengths.hod20} shows the linking length values that we adopt and use in this paper (yellow star). These values are \begin{equation} \bperp=0.14, \qquad \bpar=0.75 ~. \end{equation} Our mock catalog tests show that the FoF algorithm with these linking lengths finds galaxy groups with $N\geq 10$ that have: (1) an unbiased multiplicity function; (2) an unbiased median relation between the multiplicities of groups and their associated halos; (3) a spurious group fraction of less than $\sim 1\%$; (4) a halo completeness (fraction of halos that are associated one-to-one with groups) of more than $\sim 97\%$; (5) the correct projected size distribution as a function of multiplicity; (6) a velocity dispersion distribution that is $\sim 20\%$ too low at all multiplicities. These results hold for all of the mock catalogs that we have used (see results for other HOD models in the Appendix) and are thus not very sensitive to the HOD model assumed or to the specific realization of the underlying density field. We note that our adopted group-finder only has the above properties when dark matter halos are defined using a FoF algorithm with a linking length of 0.2 times the mean interparticle separation, since that was the definition used to construct our mock catalogs. A different halo definition (such as FoF using a different linking length, or a spherical overdensity halo-finder) will result in a different optimal group-finder. Previous FoF group analyses have used different linking lengths. For example, \citet{eke_etal_04a} adopt $\bperp=0.13, \bpar=1.43$ in their analysis of groups in the 2dF Galaxy Redshift Survey (2dFGRS; \citealt{colless_etal_01}). With a similar transverse linking length but much larger line-of-sight linking length than used here, this parameter combination yields unbiased projected sizes and velocity dispersions, but it overpredicts the abundances of halos by $20-30\%$ at large multiplicities (see Figure~\ref{fig:linkinglengths.hod20}). These groups are thus poorly suited to our primary objective of using group abundances as a cosmological test. \citet{yang_etal_05} and \citet{weinmann_etal_05} use a group-finder that assumes a mass, radius, and velocity dispersion for each preliminary group and then includes or discards galaxies from the group based on these assumed properties (similar to a matched filter technique). This method might, in principle, be able to simultaneously recover groups with unbiased abundances, projected sizes, and velocity dispersions - at the expense of model independence - but this remains to be tested. \section{Incompleteness} \label{incompleteness} There are two main sources of incompleteness that will affect the richnesses of groups, and hence the multiplicity function, in our SDSS group catalogs: fiber collisions and survey edges. Both these effects will prevent galaxies from being included in some groups, and thus cause the richness of these groups to be underestimated. These sources of incompleteness and their effects on the measured group multiplicity function must be accounted for. \subsection{Fiber Collisions} \label{fibcols} Fiber collisions cause an incompleteness that grows with the surface density of galaxies and is thus especially important in group and cluster studies. Moreover, the surface density in groups is likely a function of group richness. The mean surface density of a group of richness $N$, mass $M$, and radius $R$ scales like $\Sigma \sim N/R^2 \sim N/M^{2/3}$. For a power-law relation between mean richness and halo mass $N\sim M^\alpha$, the surface density is $\Sigma \sim N^{1-2/3\alpha}$. This scaling relation is clearly a crude approximation, but it illustrates that the incompleteness due to fiber collisions likely varies with group richness and can thus affect both the amplitude and slope of the multiplicity function. We use the 100 \texttt{LANL1-5.Mr20} mock catalogs (5 N-body simulations $\times$ 10 HOD realizations $\times$ 2 mocks per simulation cube) to assess the impact of fiber collisions on the group multiplicity function. We apply the group-finder described in \S~\ref{groupfinder} to the ``uncorrected'' and ``true'' versions of these mock catalogs and measure the resulting multiplicity functions. Figure~\ref{fig:nbodyfibcol} shows these multiplicity functions averaged over all the mock catalogs. The figure shows that dropping collided galaxies from the sample lowers the amplitude of the multiplicity function by more than 10\% and also slightly changes its slope. The amplitude drops because some groups in each richness bin lose galaxies and are thus shifted to lower $N$ bins. There are also some groups from higher $N$ bins that are shifted into these bins, but their number is smaller than the number of groups lost because the abundance of groups drops steeply with increasing $N$. \citet{zehavi_etal_05} show that the effect of fiber collisions on the galaxy two-point correlation function can be successfully corrected for by including each collided galaxy at the redshift of its nearest neighbor. We apply the same correction to our mock catalogs to produce a set of ``corrected'' mocks. Figure~\ref{fig:nbodyfibcol} shows that this correction works very well in the regime $N\geq 10$, and we therefore adopt it for our group identification. \subsection{Survey Edges} \label{edges} Groups that are identified near the edges of a given sample could be missing galaxies that are located just outside the sample. Similar to fiber collisions, edge effects always shift groups from higher to lower richness. Moreover, large and extended groups have a higher probability of being affected by edges than do small and compact groups because they can straddle an edge while being further away from it. Edge effects are most severe when the ratio of a sample's surface area to its enclosed volume is high. Figure~\ref{fig:skyvollim} shows that the SDSS sample has a highly irregular footprint on the sky, which implies a high surface-to-volume ratio. Edge effects are, therefore, potentially severe in our samples. When the SDSS survey is complete and the gap in the North Galactic cap is filled in, edge effects will be much less important. We can measure the effects of edges using our mock catalogs, since we know what galaxies lie on the other side of edges. For every group identified in our \texttt{LANL1-5.Mr20} mock catalogs, we determine how many galaxies are missing due to edges. An edge can lie either in the perpendicular direction, or along the line-of-sight due to a sample's redshift limits. The solid curve in the right panel of Figure~\ref{fig:nbodyedgestats} shows the fraction of mock groups that are missing one or more galaxies due to edges, as a function of group richness $N$. The affected fraction climbs from 10\% to 40\% as $N$ goes from 5 to 50. Edges clearly affect a large fraction of high richness groups in our sample, but counting a group as affected if it loses only a single galaxy is a very conservative test. It makes more sense to calculate the fraction of groups that lose a fixed fraction of their galaxies, rather than just a single galaxy. The dashed curve in the same panel shows the fraction of groups that lose 25\% or more of their galaxies. The affected fraction defined this way is $\sim10\%$, roughly independent of richness. Figure~\ref{fig:nbodyedge} shows the effect of edges on the multiplicity function (blue curve). The effect of edges on the abundance of mock groups grows from zero at $N=2$ to approximately 20\% at $N=50$. It is, therefore, very important to correct for edges, since they systematically change the shape of the multiplicity function and, hence, the derived HOD. We measure the shortest distance of every galaxy from the survey edges by laying down points around each galaxy at successively larger radii and checking if they also lie within our sample volume. The smallest radius at which points fall outside the sample volume is the distance of the galaxy from the edge. Any group that contains at least one galaxy within a linking length from the edge, whether it is a projected linking length in the tangential direction or a line-of-sight linking length in the redshift direction, is potentially affected, since there could be galaxies on the other side that would be linked to the same group. One possible way to deal with edges is to throw out all such groups. This is a very conservative solution, since it ensures that all groups in our final sample are uncontaminated by edges. However, it is tricky to estimate the new effective volume of the sample, which is necessary for measuring the multiplicity function. Moreover, the effective volume for large groups will be smaller than that for small groups. Another possibility is to keep all groups, but somehow correct the multiplicities of those that are potentially affected by edges. This solution has the advantage that no groups are lost, but it is once again difficult to estimate the effective volume of the sample, even if all multiplicity corrections are exactly right. A third possibility is to reject all groups whose centers lie less than a minimum distance from the edge. This correction has the advantage that it produces an unbiased sample and it is simple to estimate the new effective volume. However, it is important to use the correct minimum distance. If it is too small, then the correction will not work for the largest groups; if it is too big, then we will unnecessarily reduce our sample size. The left panel of Figure~\ref{fig:nbodyedgestats} shows the fraction of mock groups that are missing one or more galaxies due to edges, as a function of the distance from the group centroid to the edge. The fraction drops from 20\% at 100 Kpc to 5\% at 500 Kpc and less than 1\% at 1 Mpc. It does not go to zero at larger distance because there are groups with high velocity dispersion that can be far from the edge and still have galaxies within a linking length of the outer or lower redshift limit of our sample. This figure suggests that if we set the minimum distance to 500 Kpc in the tangential direction and 500 km/s in the redshift direction, we should eliminate most groups that are affected by edges. We make this correction on our mock group catalogs, and the number of groups in the resulting catalog is reduced by $\sim 22\%$ on average. We estimate the new effective volume of each group catalog by scaling the original volume by the fraction of groups that survive the edge cut. This estimate, though not exactly accurate, is simple to make and adequate for our purposes. Figure \ref{fig:nbodyedge} shows that this correction results in a multiplicity function that is unbiased due to edges (dashed red curve). Our mock catalog tests show that we can deal with survey edges effectively if we measure the multiplicity function after eliminating all groups whose centers (estimated as the centroids of their member galaxy positions) lie less than 500 Kpc from an edge in the tangential direction or less than 500 km/s from an edge in the radial direction. Applying this edge cut to the $Mr20$, $Mr19$, and $Mr18$ SDSS group catalogs reduces the numbers of groups by 22.0\%, 30.2\%, and 41.1\%, respectively. Our measurement of the multiplicity function for these samples includes this correction, though the group catalogs that we present include all groups. \section{Group and Cluster Catalog} \label{catalog} We apply our group-finding algorithm to the three volume-limited samples described in \S~\ref{data} and get three group catalogs. The fractions of ungrouped, isolated galaxies are 43.7\%, 41.2\%, and 39.8\% for the $Mr20$, $Mr19$, and $Mr18$ samples, respectively. The fractions of galaxies grouped in pairs are 19.1\%, 18.3\%, and 17.9\%. The remaining 37.2\%, 40.6\%, and 42.3\% of galaxies are in groups of three or more members. Samples $Mr20$, $Mr19$, and $Mr18$ contain a total of 4107, 2684, and 1357 groups with richness $N\geq3$, respectively. Figure~\ref{fig:groupslice} shows an equatorial slice with groups identified from sample $Mr20$. The slice is $4^\circ$ thick and each point shows the RA and redshift of a group with $N\geq3$. A comparison of this figure to Figure~\ref{fig:slicevollim} shows that groups and clusters trace the large-scale structure of galaxies, as expected. Larger groups are preferentially located in higher density regions, whereas smaller groups are more uniformly distributed. It is striking that the majority of very large groups reside within the large supercluster at $z=0.08$. Figure~\ref{fig:groupslice2} shows the same slice, but with points representing the positions of member galaxies in $N\geq3$ groups. A visual inspection of the figure shows that group velocity dispersions, which are responsible for the finger-of-God effect, are largest in the most luminous groups. For each group, we compute an unweighted group centroid, which consists of a group right ascension, declination, and mean redshift. We compute a total group luminosity that is the sum of luminosities of its member galaxies. Since we are dealing with volume-limited samples, the luminosity of a given group in samples $Mr20$, $Mr19$, $Mr18$, only counts galaxies with absolute magnitudes brighter than -19.9, -19, -18, respectively. For example, for the $Mr20$ sample, the total group absolute magnitude is \begin{equation} \Mrtot = -2.5 \mathrm{log}\left( \sum_{i=1}^{N} 10^{-0.4M_{\band{0.1}{r},i}} \right), \end{equation} and it is equivalent to integrating the galaxy luminosity function within the group from $\Mr=-19.9$ to $-\infty$. Note that we compute these group absolute magnitudes using the altered absolute magnitudes for galaxies that do not have measured redshifts due to fiber collisions (see \S~\ref{data}). We also compute a total group color, which is simply defined as $\grgrp = \Mgtot - \Mrtot$. We compute a group one-dimensional velocity dispersion given by \begin{equation} \sigv = \frac{1}{1+\bar{z}}\sqrt{\frac{1}{N-1}\sum_{i=1}^{N} (cz_i - c\bar{z})^2}, \end{equation} and an rms projected group radius given by \begin{equation} \Rproj = \sqrt{\frac{1}{N}\sum_{i=1}^{N} r_i^2}, \end{equation} where $r_i$ is the projected distance between each member galaxy and the group centroid. In the three portions of Table~3, we present the groups and clusters with $N\geq3$, selected from samples $Mr20$, $Mr19$, and $Mr18$. For each group, we list a group ID (column 1); the (J2000) right ascension and declination of the group centroid (columns 2, 3); the mean redshift of the cluster (column 4); the group richness $N$ (column 5); the total $r$-band absolute magnitude of the group, $\Mrtot$ (column 6); the total color of the group, $\grgrp$ (column 7); the line-of-sight velocity dispersion of the group, $\sigv$ (column 8); the projected rms radius of the group $\Rproj$ (column 9); the perpendicular distance of the group center from the survey edge $\redge$ (column 10). The groups in each portion of Table~3 are ranked in decreasing order of richness $N$. We show the first few rows of each portion of the table in the text and make the entire table available in the electronic version of the journal, as well as at \texttt{http://cosmo.nyu.edu/aberlind/Groups}. In Table~4, we present the member galaxies of the groups listed in Table~3. For each galaxy we list the ID of the group to which it belongs (column 1); the (J2000) right ascension and declination (columns 2, 3); the redshift (column 4); the $r$-band absolute magnitude $\Mr$\footnote{Galaxies without measured redshifts due to fiber collisions are assigned the absolute magnitude of their nearest neighbor, as described in \S~\ref{data}.} (column 5); the $\gr$ color (column 6); a fiber collision flag that is equal to 0 if the galaxy has its own measured redshift and 1 if it has been given the redshift of its nearest neighbor (column 7); the perpendicular distance of the galaxy from the survey edge $\redge$ (column 8). As before, we show the first few rows of each portion of Table~4 in the text and make the entire table available in the electronic version of the journal, as well as at \texttt{http://cosmo.nyu.edu/aberlind/Groups}. \begin{table*} \begin{center} \centerline{\small Table~3. Group and Cluster Catalogs for Samples $Mr20$, $Mr19$, and $Mr18$} \smallskip \begin{tabular}{rrrccccccr} \tableline \tableline ID & RA & DEC & $\bar{z}$ & $N$ & $\Mrtot$ & $\grgrp$ & $\sigv$ & $\Rproj$ & $\redge$ \\ & (deg) & (deg) & & & & & (km/s) & ($\hmpc$) & ($\hmpc$) \\ \tableline \cutinhead{$Mr20$} 33974 & 239.580740 & 27.312343 & 0.08797 & 132 & -25.920 & 0.946 & 723.7 & 1.371 & 17.7 \\ 16089 & 247.172589 & 40.164633 & 0.03057 & 97 & -25.468 & 0.891 & 661.1 & 1.318 & 89.3 \\ 8817 & 358.535971 & -10.372017 & 0.07405 & 61 & -25.190 & 0.921 & 736.0 & 0.734 & 17.9 \\ 14552 & 183.450292 & 59.266666 & 0.09386 & 51 & -24.861 & 0.808 & 338.3 & 1.079 & 22.9 \\ 12289 & 159.824898 & 4.987457 & 0.06815 & 51 & -24.859 & 0.899 & 661.4 & 1.161 & 47.1 \\ 3025 & 195.700154 & -2.627141 & 0.08183 & 49 & -24.805 & 0.911 & 377.1 & 1.247 & 57.9 \\ 20593 & 169.362355 & 54.469262 & 0.06907 & 49 & -24.831 & 0.906 & 426.4 & 1.202 & 35.5 \\ \cutinhead{$Mr19$} 9501 & 246.963120 & 40.182569 & 0.03009 & 197 & -25.839 & 0.886 & 588.7 & 1.317 & 88.2 \\ 4915 & 10.447791 & -9.381301 & 0.05543 & 95 & -25.068 & 0.927 & 572.4 & 0.981 & 38.8 \\ 4634 & 329.333792 & -7.765802 & 0.05727 & 86 & -25.016 & 0.724 & 564.0 & 0.677 & 52.5 \\ 10986 & 14.231949 & -0.655097 & 0.04378 & 86 & -24.944 & 0.935 & 385.4 & 1.076 & 5.2 \\ 5585 & 351.303515 & 14.909898 & 0.04113 & 83 & -24.622 & 0.871 & 496.8 & 1.045 & 53.2 \\ 3709 & 214.187113 & 1.962572 & 0.05333 & 81 & -24.902 & 0.887 & 368.3 & 1.160 & 42.9 \\ 11585 & 18.686704 & 0.254973 & 0.04442 & 68 & -24.704 & 0.903 & 386.8 & 0.744 & 27.0 \\ \cutinhead{$Mr18$} 4792 & 247.062059 & 40.107520 & 0.03011 & 311 & -25.934 & 0.865 & 584.2 & 1.300 & 90.5 \\ 2748 & 351.183638 & 14.580962 & 0.04128 & 152 & -25.057 & 0.903 & 446.6 & 1.014 & 72.3 \\ 6984 & 173.640705 & 49.042739 & 0.03270 & 65 & -24.086 & 0.918 & 526.2 & 0.533 & 45.7 \\ 1968 & 220.146510 & 3.491413 & 0.02680 & 54 & -23.853 & 0.946 & 274.1 & 0.506 & 23.6 \\ 5607 & 14.274495 & -0.247149 & 0.04303 & 52 & -24.066 & 0.915 & 309.0 & 0.760 & 13.0 \\ 5948 & 18.760997 & 0.307893 & 0.04326 & 49 & -24.108 & 0.876 & 264.9 & 0.659 & 26.5 \\ 5692 & 51.279369 & -0.496506 & 0.03664 & 48 & -23.871 & 0.870 & 246.1 & 0.802 & 44.6 \\ \tableline \label{tab:groups} \end{tabular} \end{center} Note---The rest of the table can be found in the electronic version of the ApJ, or at \texttt{http://cosmo.nyu.edu/aberlind/Groups} \end{table*} \begin{table*} \begin{center} \centerline{\small Table~4. Member Galaxies of Groups and Clusters for Samples $Mr20$, $Mr19$, and $Mr18$} \smallskip \begin{tabular}{lcrccccc} \tableline \tableline groupID & RA & DEC & $z$ & $\Mr$ & $\gr$ & fibcol & $\redge$ \\ & (deg) & (deg) & & & & & ($\hmpc$) \\ \tableline \cutinhead{$Mr20$} 14 & 196.769894 & -0.039161 & 0.08086 & -20.168 & 0.945 & 1 & 72.3 \\ 14 & 196.799107 & -0.024688 & 0.08051 & -20.498 & 0.918 & 0 & 72.3 \\ 14 & 196.788454 & -0.029741 & 0.08086 & -20.168 & 0.945 & 1 & 72.3 \\ 14 & 196.779246 & -0.038656 & 0.08086 & -20.168 & 0.945 & 0 & 72.3 \\ 15 & 197.264020 & -0.053520 & 0.07962 & -20.302 & 0.457 & 0 & 72.4 \\ 15 & 197.207327 & 0.047123 & 0.07987 & -19.950 & 0.895 & 0 & 72.4 \\ 15 & 197.165432 & 0.102322 & 0.08016 & -20.467 & 0.872 & 0 & 72.4 \\ \cutinhead{$Mr19$} 1 & 169.180550 & -0.213320 & 0.03917 & -19.355 & 0.752 & 0 & 13.5 \\ 1 & 169.195964 & -0.100215 & 0.03898 & -19.315 & 0.584 & 0 & 13.5 \\ 1 & 169.387065 & -0.187503 & 0.03999 & -20.762 & 0.967 & 0 & 13.5 \\ 5 & 199.555960 & -0.148218 & 0.04825 & -19.267 & 0.321 & 0 & 65.9 \\ 5 & 199.656619 & -0.226944 & 0.04731 & -19.705 & 0.960 & 0 & 65.9 \\ 5 & 199.665084 & -0.175183 & 0.04708 & -20.975 & 0.976 & 1 & 65.9 \\ 5 & 199.679052 & -0.178932 & 0.04708 & -20.975 & 0.976 & 0 & 65.9 \\ 5 & 199.671638 & -0.173772 & 0.04708 & -20.975 & 0.976 & 1 & 65.9 \\ \cutinhead{$Mr18$} 1 & 194.342587 & -0.630508 & 0.02247 & -18.821 & 0.744 & 1 & 57.7 \\ 1 & 194.353591 & -0.622488 & 0.02247 & -18.821 & 0.744 & 0 & 57.7 \\ 1 & 194.313130 & -0.657646 & 0.02295 & -18.837 & 0.894 & 0 & 57.7 \\ 2 & 169.180550 & -0.213320 & 0.03917 & -19.355 & 0.752 & 0 & 13.4 \\ 2 & 169.195964 & -0.100215 & 0.03898 & -19.315 & 0.584 & 0 & 13.4 \\ 2 & 169.387065 & -0.187503 & 0.03999 & -20.762 & 0.967 & 0 & 13.4 \\ 2 & 169.300864 & -0.189302 & 0.03972 & -18.203 & 0.819 & 0 & 13.4 \\ \tableline \label{tab:members} \end{tabular} \end{center} Note---The rest of the table can be found in the electronic version of the ApJ, or at \texttt{http://cosmo.nyu.edu/aberlind/Groups} \end{table*} \clearpage \section{Multiplicity Function} \label{multiplicity} With group catalogs in hand, we can now measure the group multiplicity function. The differential group multiplicity function, $\ngrpN$, is defined as the number density of groups in bins of richness $N$, where richness bins can have a width of unity or more. Before computing $\ngrpN$, we must make the corrections for incompleteness described in \S~\ref{incompleteness}. Though the catalogs presented in \S~\ref{catalog} already include the fiber collision correction, we also compute the multiplicity function from an alternate $Mr20$ group catalog that does not include this correction in order to see the magnitude of the correction. Figure~\ref{fig:groupmultfibedge} shows this uncorrected multiplicity function, as well as the multiplicity function that includes the fiber collision correction. The figure shows that applying the correction boosts the amplitude of the multiplicity function, just as it did in our mock tests in \S~\ref{incompleteness}. Figure~\ref{fig:groupmultfibedge} also shows the effect on the multiplicity function of applying the edge correction described in \S~\ref{incompleteness}. This effect is small, typically less than 5\%, though it is larger in individual bins at high $N$, where the number of groups is small. \begin{table}[t] \begin{center} \centerline{\small Table~5. Group Multiplicity Function for $Mr20$ Sample} \begin{tabular}[t]{lccc} \tableline \tableline $\Nmin$--$\Nmax$ & $\ngrpN$ & $\signgrpN$ & $\signgrpN$ (Poisson) \\ \tableline 3--3 & $2.290\times 10^{-4}$ & $1.110\times 10^{-5}$ & $5.881\times 10^{-6}$ \\ 4--4 & $1.054\times 10^{-4}$ & $4.890\times 10^{-6}$ & $3.990\times 10^{-6}$ \\ 5--5 & $4.909\times 10^{-5}$ & $4.181\times 10^{-6}$ & $2.723\times 10^{-6}$ \\ 6--6 & $3.263\times 10^{-5}$ & $4.465\times 10^{-6}$ & $2.220\times 10^{-6}$ \\ 7--7 & $1.962\times 10^{-5}$ & $1.979\times 10^{-6}$ & $1.722\times 10^{-6}$ \\ 8--8 & $1.496\times 10^{-5}$ & $2.250\times 10^{-6}$ & $1.503\times 10^{-6}$ \\ 9--9 & $1.118\times 10^{-5}$ & $2.398\times 10^{-6}$ & $1.299\times 10^{-6}$ \\ 10--10 & $8.906\times 10^{-6}$ & $1.502\times 10^{-6}$ & $1.160\times 10^{-6}$ \\ 11--11 & $5.139\times 10^{-6}$ & $1.292\times 10^{-6}$ & $8.810\times 10^{-7}$ \\ 12--12 & $4.223\times 10^{-6}$ & $8.632\times 10^{-7}$ & $7.986\times 10^{-7}$ \\ 13--13 & $3.780\times 10^{-6}$ & $7.200\times 10^{-7}$ & $7.555\times 10^{-7}$ \\ 14--14 & $2.565\times 10^{-6}$ & $1.283\times 10^{-6}$ & $6.224\times 10^{-7}$ \\ 15--15 & $2.873\times 10^{-6}$ & $9.335\times 10^{-7}$ & $6.587\times 10^{-7}$ \\ 16--16 & $2.868\times 10^{-6}$ & $1.165\times 10^{-6}$ & $6.581\times 10^{-7}$ \\ 17--17 & $1.361\times 10^{-6}$ & $6.868\times 10^{-7}$ & $4.533\times 10^{-7}$ \\ 18--18 & $1.358\times 10^{-6}$ & $4.131\times 10^{-7}$ & $4.530\times 10^{-7}$ \\ 19--19 & $1.209\times 10^{-6}$ & $5.133\times 10^{-7}$ & $4.273\times 10^{-7}$ \\ 20--21 & $9.817\times 10^{-7}$ & $3.079\times 10^{-7}$ & $3.851\times 10^{-7}$ \\ 22--24 & $6.039\times 10^{-7}$ & $2.253\times 10^{-7}$ & $3.020\times 10^{-7}$ \\ 25--28 & $3.401\times 10^{-7}$ & $9.522\times 10^{-8}$ & $2.266\times 10^{-7}$ \\ 29--30 & $9.061\times 10^{-7}$ & $4.483\times 10^{-7}$ & $3.699\times 10^{-7}$ \\ 31--34 & $3.398\times 10^{-7}$ & $7.501\times 10^{-8}$ & $2.265\times 10^{-7}$ \\ 35--42 & $1.699\times 10^{-7}$ & $6.455\times 10^{-8}$ & $1.602\times 10^{-7}$ \\ 43--61 & $6.360\times 10^{-8}$ & $2.982\times 10^{-8}$ & $9.801\times 10^{-8}$ \\ \tableline \label{tab:mult20} \end{tabular} \end{center} Note---$\ngrp$ and $\signgrpN$ are in units of $\hden$. \end{table} We must calculate errorbars for the multiplicity function in order to use it to constrain the HOD. We use our mock catalogs for this purpose. Specifically, we compute fractional errors from the dispersion among 10 independent mock catalogs for the $Mr20$ sample (\texttt{LANL1-5.Mr20} mocks $\times$ 1 HOD realization $\times$ 2 mocks per simulation cube), and 8 mock catalogs for each of the $Mr19$ and $Mr18$ samples (\texttt{LANL1-4.Mr19}/\texttt{LANL1-4.Mr18} mocks $\times$ 1 HOD realization $\times$ 2 mocks per simulation cube). Note that we do not use multiple HOD realizations because the underlying halo populations themselves would not be independent. Before computing errors, we correct each mock catalog for fiber collisions and edge effects in the same way as in the data. The computed errors thus implicitly include any contribution from these correction procedures. The SDSS multiplicity function shown in Figure~\ref{fig:groupmultfibedge} becomes very noisy at high richness because the abundance of groups drops with $N$ and the figure uses richness bins with a width of unity. It makes more sense to increase the bin width with $N$ so as to beat down the noise. Moreover, since we calculate errorbars for the multiplicity function using our mock catalogs, each richness bin must contain enough mock groups so that an errorbar can be reliably estimated. We choose richness bins for each group catalog so that each bin contains at least eight SDSS groups and twenty mock groups (among all mock catalogs used). At low multiplicities, the bin width is always unity because there are many groups with low $N$. At higher multiplicities, however, the richness bins grow wider in order to satisfy these criteria. The bin widths for samples $Mr20$, $Mr19$, and $Mr18$, are listed in the first columns of Tables~5, 6, and~7, respectively. Once a richness bin is defined, the abundance of groups in that bin, $\ngrpN$, is simply the number of groups having richnesses within the bin, divided by the sample volume and divided by the bin width. The values of $\ngrpN$ are listed in the second columns of Tables~5, 6, and~7. We use the same richness bins to compute the abundance of mock groups for each independent mock catalog, and we compute errors, $\signgrpN$, in the SDSS multiplicity function by measuring the dispersion among the mock multiplicity functions. These errors are listed in the third columns of Tables~5, 6, and~7. Finally, we also compute Poisson errors for the SDSS $\ngrpN$, which we list in the fourth columns of Tables~5, 6, and~7. In some of the highest multiplicity bins, the Poisson errors are larger than the mock errors. In these cases, the mock errors are likely underestimated and it is best to use the Poisson errors in their place. \begin{table}[t] \begin{center} \centerline{\small Table~6. Group Multiplicity Function for $Mr19$ Sample} \begin{tabular}[t]{lccc} \tableline \tableline $\Nmin$--$\Nmax$ & $\ngrpN$ & $\signgrpN$ & $\signgrpN$ (Poisson) \\ \tableline 3--3 & $4.514\times 10^{-4}$ & $2.872\times 10^{-5}$ & $1.545\times 10^{-5}$ \\ 4--4 & $1.889\times 10^{-4}$ & $1.201\times 10^{-5}$ & $9.996\times 10^{-6}$ \\ 5--5 & $1.085\times 10^{-4}$ & $9.323\times 10^{-6}$ & $7.575\times 10^{-6}$ \\ 6--6 & $6.292\times 10^{-5}$ & $8.977\times 10^{-6}$ & $5.769\times 10^{-6}$ \\ 7--7 & $5.027\times 10^{-5}$ & $5.465\times 10^{-6}$ & $5.157\times 10^{-6}$ \\ 8--8 & $2.856\times 10^{-5}$ & $2.434\times 10^{-6}$ & $3.887\times 10^{-6}$ \\ 9--9 & $1.853\times 10^{-5}$ & $2.832\times 10^{-6}$ & $3.131\times 10^{-6}$ \\ 10--10 & $1.534\times 10^{-5}$ & $2.799\times 10^{-6}$ & $2.849\times 10^{-6}$ \\ 11--11 & $1.534\times 10^{-5}$ & $2.577\times 10^{-6}$ & $2.849\times 10^{-6}$ \\ 12--12 & $1.164\times 10^{-5}$ & $2.236\times 10^{-6}$ & $2.482\times 10^{-6}$ \\ 13--13 & $8.994\times 10^{-6}$ & $2.135\times 10^{-6}$ & $2.181\times 10^{-6}$ \\ 14--14 & $7.936\times 10^{-6}$ & $2.105\times 10^{-6}$ & $2.049\times 10^{-6}$ \\ 15--15 & $5.819\times 10^{-6}$ & $1.186\times 10^{-6}$ & $1.755\times 10^{-6}$ \\ 16--16 & $5.819\times 10^{-6}$ & $1.718\times 10^{-6}$ & $1.755\times 10^{-6}$ \\ 17--18 & $5.819\times 10^{-6}$ & $1.318\times 10^{-6}$ & $1.755\times 10^{-6}$ \\ 19--20 & $2.380\times 10^{-6}$ & $5.168\times 10^{-7}$ & $1.122\times 10^{-6}$ \\ 21--23 & $2.292\times 10^{-6}$ & $5.243\times 10^{-7}$ & $1.101\times 10^{-6}$ \\ 24--26 & $1.587\times 10^{-6}$ & $4.621\times 10^{-7}$ & $9.164\times 10^{-7}$ \\ 27--32 & $7.054\times 10^{-7}$ & $2.228\times 10^{-7}$ & $6.109\times 10^{-7}$ \\ 33--38 & $7.054\times 10^{-7}$ & $3.069\times 10^{-7}$ & $6.109\times 10^{-7}$ \\ 39--51 & $3.256\times 10^{-7}$ & $4.634\times 10^{-8}$ & $4.151\times 10^{-7}$ \\ 52--86 & $1.209\times 10^{-7}$ & $3.602\times 10^{-8}$ & $2.529\times 10^{-7}$ \\ \tableline \label{tab:mult19} \end{tabular} \end{center} Note---Same units as Table~5. \end{table} \begin{table}[t] \begin{center} \centerline{\small Table~7. Group Multiplicity Function for $Mr18$ Sample} \begin{tabular}[t]{lccc} \tableline \tableline $\Nmin$--$\Nmax$ & $\ngrpN$ & $\signgrpN$ & $\signgrpN$ (Poisson) \\ \tableline 3--3 & $7.311\times 10^{-4}$ & $6.909\times 10^{-5}$ & $4.000\times 10^{-5}$ \\ 4--4 & $3.436\times 10^{-4}$ & $3.325\times 10^{-5}$ & $2.742\times 10^{-5}$ \\ 5--5 & $1.948\times 10^{-4}$ & $2.200\times 10^{-5}$ & $2.065\times 10^{-5}$ \\ 6--6 & $1.248\times 10^{-4}$ & $1.629\times 10^{-5}$ & $1.652\times 10^{-5}$ \\ 7--7 & $1.182\times 10^{-4}$ & $1.546\times 10^{-5}$ & $1.608\times 10^{-5}$ \\ 8--8 & $5.686\times 10^{-5}$ & $9.917\times 10^{-6}$ & $1.116\times 10^{-5}$ \\ 9--9 & $3.284\times 10^{-5}$ & $5.340\times 10^{-6}$ & $8.477\times 10^{-6}$ \\ 10--10 & $3.066\times 10^{-5}$ & $5.777\times 10^{-6}$ & $8.191\times 10^{-6}$ \\ 11--11 & $2.626\times 10^{-5}$ & $8.403\times 10^{-6}$ & $7.581\times 10^{-6}$ \\ 12--13 & $1.423\times 10^{-5}$ & $1.629\times 10^{-6}$ & $5.580\times 10^{-6}$ \\ 14--15 & $8.756\times 10^{-6}$ & $1.443\times 10^{-6}$ & $4.378\times 10^{-6}$ \\ 16--17 & $1.203\times 10^{-5}$ & $1.761\times 10^{-6}$ & $5.132\times 10^{-6}$ \\ 18--23 & $3.647\times 10^{-6}$ & $7.402\times 10^{-7}$ & $2.825\times 10^{-6}$ \\ 24--31 & $2.188\times 10^{-6}$ & $6.091\times 10^{-7}$ & $2.188\times 10^{-6}$ \\ 32--152 & $1.447\times 10^{-7}$ & $1.673\times 10^{-8}$ & $5.627\times 10^{-7}$ \\ \tableline \label{tab:mult18} \end{tabular} \end{center} Note---Same units as Table~5. \end{table} Figure~\ref{fig:groupmult} shows the SDSS multiplicity functions for the three volume-limited samples, along with the mock errorbars for the $Mr20$ sample. Though we measure and show the multiplicity function down to a multiplicity of $N=3$, our tests with mock catalogs have shown that it is only unbiased with respect to the true halo multiplicity function for $N\geq 10$. When using this measured multiplicity function to constrain the HOD, we must either only use bins with $N\geq 10$, or attempt to calibrate the relation between the measured group multiplicity function and the true halo multiplicity function at lower values of $N$. The central curve of Figure~\ref{fig:nbodymultbxybz}, discussed in the Appendix, effectively provides this calibration for $Mr20$ and the cosmology adopted in our mock catalogs. The multiplicity functions shown in Figure~\ref{fig:groupmult} appear to be close to power-law relations. In order to test this, we perform a simple power-law fit to each multiplicity function in the regime $N\geq 10$. We use only the diagonal errors of the full covariance matrix (i.e., the errors listed in Tables~5, 6, and ~7). We find that all three multiplicity functions are well-fit by power-law relations, with best-fit slopes of $-2.72\pm0.16$, $-2.48\pm0.14$, and $-2.49\pm0.28$ for the $Mr20$, $Mr19$, and $Mr18$ samples, respectively. \section{Summary and Discussion} \label{summary} We have used a simple friends-of-friends algorithm to identify galaxy groups in volume-limited samples of the SDSS redshift survey. We have selected FoF linking lengths that are best at grouping together galaxies that occupy the same dark matter halos. We based this choice on extensive tests with mock galaxy catalogs, which we constructed by populating halos in N-body simulations with galaxies. The result of our mock tests is that no combination of perpendicular and line-of-sight linking lengths can yield groups that successfully recover all aspects of the parent halo distribution, even for large richness systems. Specifically, FoF cannot identify groups that simultaneously have unbiased abundances, projected sizes, and velocity dispersions. The ideal group-finding parameters for a given study depend on its scientific objectives. Given our objective of using the multiplicity function to constrain the HOD, it makes sense to sacrifice velocity dispersions and obtain groups with unbiased abundances and projected sizes. Our choice of linking lengths results in a group catalog that, for groups of ten or more members, has an unbiased multiplicity function, an unbiased median relation between the multiplicities of groups and their parent halos, an unbiased projected size distribution as a function of multiplicity, and a velocity dispersion distribution that is $\sim 20\%$ too low for all multiplicities. We correct for fiber collisions and survey edge effects and present three SDSS group catalogs (for three different volume-limited samples) and their measured multiplicity functions. It is important to recognize that our adopted group finder has the above properties only for halos defined using FoF with a linking length of 0.2 times the mean interparticle separation, since this is how halos were identified in our mock catalogs. A different halo definition (such as FoF with a different linking length, or spherical overdensity halos) would require a different set of optimal group-finding parameters. This is not a problem as long as the same halo definition is used consistently. For example, an HOD measured from these group catalogs will hold for this halo definition, and any theoretical model should use the same halo definition to compare its predictions to the measured HOD. We chose this particular halo finder because it has been widely used and tested, and the properties of the resulting halo distribution (e.g., mass function) are well understood. The groups and clusters that we present here are intended to be systems of galaxies that belong to the same virialized dark matter halo. We can test whether these systems are virialized by computing crossing times for the groups and checking if they are sufficiently less than the Hubble time. We define the crossing time divided by the hubble time as \begin{equation} \frac{\tcross}{\tH} = \frac{(\Rrms/\hmpc)}{(\sigv/100\kms)}, \end{equation} where $\Rrms$ is the one-dimensional group radius, which is equal to the projected (two-dimensional) radius, $\Rproj$, divided by the square root of two. We correct for the velocity dispersion bias revealed in our mock tests by applying a 20\% upward correction to all group velocity dispersions, and we compute $\tcross/\tH$ for all groups. We find that, for all three group catalogs, the median value of $\tcross/\tH$ is $\sim0.15$, and 80\% of all groups have values less than $\sim0.29$. These numbers can be interpreted in terms of the spherical infall model \citep{gunn_gott_72,gott_turner_77a}, or other analytic or numerical models. However, at a first glance, the numbers are encouraging and suggest that most of our groups are likely virialized systems. The group and cluster catalogs presented here are well-suited for testing many of the predictions and assumptions made by galaxy formation models regarding the relationship between galaxies and their underlying dark matter halos. We will investigate several of these issues in subsequent papers. \acknowledgments We thank Zheng Zheng and Jeremy Tinker for their help with choosing HOD parameters for constructing mock catalogs and Luis Teodoro for his help with making the mock catalogs. AAB acknowledges support by NSF grant AST-0079251 and the NSF Center for Cosmological Physics, while at the University of Chicago, and by NASA grant NAG5-11669, NSF grant PHY-0101738, and a grant from NASA administered by the American Astronomical Society, while at New York University. AAB also acknowledges the hospitality of the Aspen Center for Physics, where some of this work was completed. MRB and DWH acknowledge support by NSF grant AST-0428465. DHW acknowledges support by NSF grant AST-0407125. JRG acknowledges support by NSF grant AST-0406713. Portions of this work were performed under the auspices of the U.S. Dept. of Energy, and supported by its contract \#W-7405-ENG-36 to Los Alamos National Laboratory. Funding for the SDSS and SDSS-II has been provided by the Alfred P. Sloan Foundation, the Participating Institutions, the National Science Foundation, the U.S. Department of Energy, the National Aeronautics and Space Administration, the Japanese Monbukagakusho, the Max Planck Society, and the Higher Education Funding Council for England. The SDSS Web Site is http://www.sdss.org/. The SDSS is managed by the Astrophysical Research Consortium for the Participating Institutions. The Participating Institutions are the American Museum of Natural History, Astrophysical Institute Potsdam, University of Basel, Cambridge University, Case Western Reserve University, University of Chicago, Drexel University, Fermilab, the Institute for Advanced Study, the Japan Participation Group, Johns Hopkins University, the Joint Institute for Nuclear Astrophysics, the Kavli Institute for Particle Astrophysics and Cosmology, the Korean Scientist Group, the Chinese Academy of Sciences (LAMOST), Los Alamos National Laboratory, the Max-Planck-Institute for Astronomy (MPA), the Max-Planck-Institute for Astrophysics (MPIA), New Mexico State University, Ohio State University, University of Pittsburgh, University of Portsmouth, Princeton University, the United States Naval Observatory, and the University of Washington. \section*{Appendix} In this appendix, we describe the mock catalog tests that help us choose optimal FoF parameters. Since our primary goal for identifying groups is to measure the group multiplicity function and use it to constrain the HOD, we clearly require our FoF algorithm to produce groups that have an unbiased multiplicity function with respect to the true halo multiplicity function. In addition, we require an unbiased relation between the multiplicities of groups and their associated halos. Finally, we would like our groups to have unbiased projected size and velocity dispersion distributions as a function of multiplicity. We create a grid of FoF linking lengths and check how each set of linking lengths performs in the above tests, for each of the four HOD model mock cubes (\texttt{.Mr20, .Mr20b, .Mr19, .Mr18}). In the case of each HOD model, we average results over the 10 HOD realizations described in \S~\ref{mocks} and over the \texttt{LANL1} and \texttt{LANL4} N-body simulations. Before focusing on redshift space, we briefly examine how well FoF recovers the true multiplicity function in real space, since this represents the best possible case (any group finder will almost certainly perform worse in redshift space). We apply FoF to the real-space cube mocks using a single linking length (the linking volume around each mock galaxy is a sphere), and investigate how the recovered multiplicity function varies with the value of this linking length. In particular, we compare the mock group multiplicity functions to the input halo multiplicity functions that were used to construct the mock catalogs. Figure~\ref{fig:nbodymultbxy} shows this comparison for the \texttt{.Mr20} mocks. The bottom panel of the figure shows the logarithm of the ratio of group to halo multiplicity function, and the horizontal solid line therefore denotes the ``unbiased'' case. The figure reveals that, at large $N$, the group multiplicity function has an unbiased shape that is independent of the choice of linking length (at least for the range of linking lengths shown). The amplitude, however, is dependent on the linking length used, with larger linking lengths leading to a higher abundance of groups at large $N$. A linking length of $b=0.2$ (in units of the mean intergalaxy separation) yields a group multiplicity function with an unbiased amplitude at large $N$. This is not surprising given that the same value was used to identify dark matter halos in the N-body simulations while constructing mock catalogs. At low $N$, the multiplicity function is highly biased, both in shape and amplitude. The abundance of groups relative to halos at a given multiplicity $N$ decreases when FoF splits these halos into smaller groups or merges them to form larger groups. This decrease is countered by an increase due to the merging of smaller halos or the splitting of larger halos. The balance between these competing effects determines whether the multiplicity function is biased or not. For linking lengths near $b=0.2$, merging dominates over splitting, which means that group abundances at a given multiplicity are mainly determined by a balance between halos at that $N$ merging to yield larger groups and smaller halos merging to replenish the lost groups. However, this balance breaks at $N=1$ because, while FoF merges $N=1$ halos (i.e., isolated galaxies) to form larger groups, there are no smaller halos that can merge to replenish $N=1$ groups. The abundance of $N=1$ groups is therefore necessarily less than that of $N=1$ halos (it can only be more if the linking length is so small - approximately $b\sim0.1$ - that single galaxy groups splinter off in large numbers from larger halos). Since most galaxies live in $N=1$ halos ($\sim70\%$ in these mock catalogs), merging a small fraction of them to form larger groups will fractionally increase the abundance of larger $N=2, 3, 4$, etc. groups significantly. This is seen in Figure~\ref{fig:nbodymultbxy}: the abundance of $N=1$ groups is lower than that of halos by $\sim20\%$ for $b=0.2$, causing the abundance of $N=2$ and $N=3$ groups to be $\sim50\%$ higher. Only for $N>10$ does the group abundance settle down and become unbiased. This behavior is a fundamental limitation of the FoF algorithm, and it has the consequence that group abundances can only be trusted for large multiplicity groups. In redshift space, group finding is much more challenging because finger-of-god distortions stretch groups along the line-of-sight, making it more likely that single halos will be split into multiple groups and that neighboring halos will be merged into the same groups. Figure~\ref{fig:nbodyslice} illustrates these effects by showing the performance of FoF in a small slice through a single mock catalog (one HOD realization of the \texttt{LANL4.Mr20} mock catalog). The top-left panel shows the mock galaxies in real space, with each $N>4$ halo denoted by a unique color. The bottom-left panel shows the same galaxies in redshift space, where the line-of-sight is oriented along the $z$-axis of the mock cube. Large open circles have radii equal to the halo virial radii and are centered at the halo centers in real space, and the galaxy centroids in redshift space. We run our adopted FoF group-finder (described in \S~\ref{groupfinder}) on the redshift-space mock and denote each resulting $N>4$ group with a unique color in the bottom-right panel. Finally, we show the group galaxies' real-space positions in the top-right panel. Large dotted circles are centered at the group centroids and have virial radii that are estimated by assuming a halo mass function and a monotonic relation between group multiplicity and mass. A visual comparison of the real- and redshift-space panels reveals many of the failure modes of FoF group-finding in redshift space. The halo denoted by green in the left-side panels is fairly well recovered by FoF as the group denoted by green in the right-side panels. However, a couple of halo galaxies are missed in group finding, such as the one whose velocity moved it the furthest away from the center of the halo. Most of the galaxies in the halo denoted by blue are linked together in the same group, also denoted by blue. However, many galaxies that do not belong to the ``blue'' halo are also linked to the same group. This is seen clearly in the top-right panel, where seven of the ``blue'' group galaxies' real-space positions place them well outside the halo. A similar thing occurs to the halos and corresponding groups denoted by magenta and cyan. Most of the galaxies in the large ``red'' halo are recovered correctly into the ``red'' group, but there are some galaxies added to this group that do not belong to the ``red'' halo, as well as a few galaxies that do belong to that halo, but have splintered off into a different group (denoted by dark green). Despite these imperfections, there is clearly a substantial correspondence between the groups identified by FoF and the true population of halos in this slice. We now examine the relative multiplicity functions of groups and halos when the groups are identified in redshift space. If we use the same linking length in transverse and line-of-sight directions, finger-of-god distortions will cause halos to be split into multiple small groups along the line-of-sight. This is demonstrated by the dashed curve in Figure~\ref{fig:nbodymultbxybz}, which shows the multiplicity function of groups identified with a single linking length of $b=0.2$. The abundance of groups is vastly underestimated for $N\gtrsim 5$, and the effect grows with $N$ because richer halos have higher velocity dispersions. We therefore need to use different linking lengths in the line-of-sight and perpendicular directions. We apply FoF to our redshift-space cube mocks for a grid of perpendicular and line-of-sight linking lengths and find that we can recover an unbiased multiplicity function at large $N$ for the right combinations of linking lengths. Figure~\ref{fig:nbodymultbxybz} shows one such combination ($\bperp=0.14$, $b_z=0.75$) and demonstrates how the group multiplicity function changes with the line-of-sight linking length $b_z$. Generally, larger linking lengths in either direction lead to a higher abundance of groups at large $N$. We record all linking length combinations that yield unbiased multiplicity functions in the large $N$ regime and show the successful parameter space in Figure~\ref{fig:linkinglengths.hod20}, as discussed in \S~\ref{groupfinder}. Recovering an unbiased multiplicity function does not guarantee that the one-to-one relation between the multiplicities of halos and their recovered groups is also unbiased. We therefore also investigate this relation. As described in \S~\ref{groupfinder}, we associate each halo to the recovered group that contains the halo's central galaxy. Groups that contain central galaxies from more than one halo are associated with the halo with which they share the largest number of galaxies. Halos that end up not being associated with any group are considered ``undetected,'' and groups that are not associated with any halo (i.e., they contain no halo central galaxies) are considered ``spurious''. Once we have associated mock groups one-to-one with their parent halos, we can look at the relation between the halo and group multiplicities (i.e., $\Ntrue$ vs. $\Nobs$). In addition, we can look at the fraction of halos that are detected and the fraction of groups that are spurious. Figure~\ref{fig:nbodycompbxybz} shows how these relations depend on the line-of-sight linking length. The bottom panel of the figure shows one set of linking lengths ($\bperp=0.14$, $b_z=0.70$) that yields an unbiased median relation between $\Ntrue$ and $\Nobs$, but the scatter around this relation is large and quite asymmetric. 90\% of groups at a given $\Nobs$ are associated with halos that have up to 40\% higher and 60\% lower $\Ntrue$. Increasing the line-of-sight linking length causes groups to grow and thus biases the median $\Ntrue$ vs. $\Nobs$ relation by tilting it toward larger $\Nobs$. As before, we record all linking length combinations that yield unbiased median relations between group and halo multiplicities, and we show the successful parameter space in Figure~\ref{fig:linkinglengths.hod20}. The top panel of Figure~\ref{fig:nbodycompbxybz} shows the completeness (fraction of halos that are associated one-to-one with groups) as a function of halo multiplicity $\Ntrue$, and the middle panel shows the spurious group fraction as a function of group multiplicity $\Nobs$. Over a wide range of FoF linking lengths, the completeness for halos with $N\gtrsim 5$ is over 95\%, and the spurious fraction for groups with $N\gtrsim 5$ is less than 5\%. Increasing the line-of-sight linking length causes a drop in the halo completeness and a corresponding drop in the spurious group fraction, since more halos get linked to the same groups. For the final linking lengths that we use (see \S~\ref{groupfinder}), the halo completeness is greater than 97\% and the spurious group fraction less than 1\% for $N\gtrsim 10$. The high completeness and low spurious fraction are a result of how we associate groups to halos. Since we only require a group to have a halo's central galaxy in order to be associated with it, most groups and halos have one-to-one associations. If we used a more stringent criterion for group-halo association, for example by requiring that a group contain some minimum fraction of a halo's galaxies, then the halo completeness would be lower and the spurious group fraction higher, but the scatter in $\Ntrue$ vs. $\Nobs$ would be reduced. The three panels of Figure~\ref{fig:nbodycompbxybz}, put together, characterize the errors in the FoF group finder. Changing the definition for how groups are associated to halos does not change the errors in group-finding; it merely redistributes the errors among the three panels. In addition to requiring that our groups have unbiased abundances and multiplicities, we would also like them to have unbiased size distributions. For every group in our redshift-space cube mocks, we measure the projected rms radius and the line-of-sight velocity dispersion of galaxies. We compare these to the projected rms radii and actual velocity dispersions of halo galaxies. Figure~\ref{fig:nbodygrpstatsbxybz} shows the median, 10th, and 90th percentile projected size and velocity dispersion as a function of multiplicity for halos, compared to that for groups identified with two different line-of-sight linking lengths. Increasing the line-of-sight linking length produces groups with higher velocity dispersions, but it has less impact on the projected size distributions. The opposite is naturally true when we increase the perpendicular linking length. Linking length combinations that yield groups with unbiased abundances and projected sizes tend to yield velocity dispersions that are biased low. This is illustrated in Figure~\ref{fig:nbodygrpstatsbxybz}, which shows that the linking length combination $\bperp=0.14$, $b_z=0.7$ yields groups with velocity dispersions that are $\sim 20\%$ too low relative to halos. The line-of-sight linking length must be more than doubled to repair this bias, but then the abundances of groups would be too high. Figure~\ref{fig:linkinglengths.hod20} shows the linking length parameter space that satisfies each of the above tests. As discussed in \S~\ref{groupfinder}, there is no combination of perpendicular and line-of-sight linking lengths that yields groups with unbiased abundances, projected sizes, and velocity dispersions, even at high multiplicity. We choose to sacrifice velocity dispersions and adopt the parameters $\bperp=0.14$, $b_z=0.75$. All the above tests and resulting choice of linking lengths were done using the \texttt{.Mr20} mock catalogs. Since we plan to use our group catalog to constrain the HOD, it is vital that our choice of linking lengths does not depend sensitively on the input HOD assumed when constructing the mocks. For this reason, we repeat all the above tests with the \texttt{.Mr20b} mock catalogs, which use a different input HOD to model the same $Mr20$ sample of SDSS galaxies. The results are shown in Figure~\ref{fig:linkinglengths.hod20b}. It is clear that our adopted group finder performs equally well in both sets of mock catalogs, demonstrating that our choice of linking lengths is insensitive to the underlying HOD. It is also important to show how well our linking lengths work on lower luminosity galaxy samples, since we apply them to the SDSS $Mr19$ and $Mr18$ samples. We thus repeat our mock tests with the \texttt{.Mr19} and \texttt{.Mr18} mock catalogs and show the results in Figures~\ref{fig:linkinglengths.hod19} and~\ref{fig:linkinglengths.hod18}, respectively. The figures show that lower luminosity (higher density) samples require slightly higher line-of-sight linking lengths in order to retain unbiased multiplicity functions. However, this effect is small. When applied to the \texttt{.Mr18} mock catalogs, our adopted linking lengths yield a multiplicity function that is 10\% too low in amplitude. Overall, Figures~\ref{fig:linkinglengths.hod20}, \ref{fig:linkinglengths.hod20b}, \ref{fig:linkinglengths.hod19}, and~\ref{fig:linkinglengths.hod18} demonstrate that our choice of linking lengths is fairly robust. \def\baselinestretch{1} \bibliographystyle{apj} \bibliography{}
Title: Resolving the Stellar Outskirts of M31 and M33
Abstract: Many clues about the galaxy assembly process lurk in the faint outer regions of galaxies. The low surface brightnesses of these parts pose a significant challenge for studies of diffuse light, and few robust constraints on galaxy formation models have been derived to date from this technique. Our group has pioneered the use of extremely wide-area star counts to quantitatively address the large-scale structure and stellar content of external galaxies at very faint light levels. We highlight here some results from our imaging and spectroscopic surveys of M31 and M33.
https://export.arxiv.org/pdf/astro-ph/0601121
\articletitle[Resolving the Stellar Outskirts of M31 and M33]{Resolving the Stellar Outskirts of M31 and M33} \author{Annette Ferguson\altaffilmark{1}, Mike Irwin\altaffilmark{2}, Scott Chapman\altaffilmark{3}, Rodrigo Ibata\altaffilmark{4}, Geraint Lewis\altaffilmark{5}, Nial Tanvir\altaffilmark{6}} \affil{\altaffiltext{1}{Institute for Astronomy, University of Edinburgh, UK} \altaffiltext{2}{Institute of Astronomy, University of Cambridge, UK} \altaffiltext{3}{California Institute for Technology, Pasadena, USA} \altaffiltext{3}{Observatoire de Strasbourg, Strasbourg, France} \altaffiltext{3}{Institute of Astronomy, University of Sydney, Australia} \altaffiltext{3}{Centre for Astrophysics Research, University of Hertfordshire, UK}} \section{Introduction} The study of galaxy outskirts has become increasingly important in recent years. From a theoretical perspective, it has been realised that many important clues about the galaxy assembly process should lie buried in these parts. Cosmological simulations of disk galaxy formation have now been carried out by several groups and have led to testable predictions for the large-scale structure and stellar content at large radii -- for example, the abundance and nature of stellar substructure (e.g. Bullock \& Johnston 2005, Font et al. 2005), the ubiquity, structure and content of stellar halos and thick disks (e.g. Abadi et al. 2005, Governato et al. 2004, Brook et al. 2005) and the age distribution of stars in the outer regions of thin disks (e.g. Abadi et al. 2003). Bullock \& Johnston (2005) find that Milky Way-like galaxies will have accreted 100-200 luminous satellites during the last 12~Gyr and that the signatures of this process should be readily visible at surface brightnesses of V$\sim 30$ magnitudes per square arcsec and lower. Although traditional surface photometry at such levels (roughly 9 magnitudes below sky) remains prohibitive, star count analyses of nearby galaxies have the potential to reach these effective depths (e.g. Pritchet \& van den Bergh 1994). The requirement of a large survey area (to provide a comprehensive view of the galaxy) and moderate-depth imagery can now be achieved in a relatively straightforward manner using wide-field imaging cameras attached to medium-sized telescopes. \section{The INT WFC Surveys of M31 and M33} In 2000, we began a program to map the outer regions of our nearest large neighbour, M31, with the Wide-Field Camera equipped to the INT 2.5m. The success of this program led us to extend our survey to M33 in the fall of 2002. To date, more than 45 and 7 square degrees have been mapped around these galaxies respectively. Our imagery reaches to V$\sim$24.5 and {\sl i}$\sim$23.5 and thus probes the top 3 magnitudes of the red giant branch (RGB) in each system. The raw data are pipeline-processed in Cambridge and source catalogues are produced containing positions, magnitudes and shape parameters. The M31 survey currently contains more than 7 million sources, and the M33 survey more than 1 million. Magnitude and colour cuts are applied to point-like sources in order to isolate distinct stellar populations and generate surface density maps (see Figure 1). The faint structures visible by eye in Figure 1 have effective V-band surface brightnesses in the range 29-30 magnitudes per square arcsec. Early versions of our M31 maps have been discussed in Ibata et al. (2001), Ferguson et al. (2002) and Irwin et al. (2005). \subsection{Results for M31} The left-hand panel of Figure 1 shows the distribution of blue (i.e. presumably more metal-poor) RGB stars in and around M31. A great deal of substructure can be seen including the giant stream in the south-east, various overdensities near both ends of the major axes, a diffuse extended structure in the north-east and a loop of stars projected near NGC~205. {\sl Origin of the Substructure:} Do the substructures in M31 represent debris from one or more satellite accretions, or are they simply the result of a warped and/or disturbed outer disk? We are addressing these issues with deep ground-based imagery from the INT and CFHT, Keck-10m spectroscopy and deep HST/ACS colour-magnitude diagrams (CMDs). Our findings to date can be summarized as follows: \begin{itemize} \item{M31 has at least 12 satellites lying within a projected radius of 200~kpc. The bulk of these systems, the low-luminosity dwarf spheroidals, are unlikely to be associated with the stellar overdensities since their RGB stars are much bluer than those of the substructure (Ferguson et al. 2002).} \item{The combination of line-of-sight distances and radial velocities for stars at various locations along the giant stellar stream constrains the progenitor orbit (e.g. McConnachie et al. 2003, Ibata et al. 2004). Currently-favoured orbits do not connect the more luminous inner satellites (e.g. M32, NGC~205) to the stream in any simple way however this finding leaves some remarkable coincidences (e.g. the projected alignment on the sky, similar metallicities) yet unexplained.} \item{Deep HST/ACS CMDs reaching well below the horizontal branch reveal different morphologies between most substructures in the outskirts of M31 (Ferguson et al. 2005, see Figure 2). These variations reflect differences in the mean age and/or metallicity of the constituent stellar populations. Analysis is underway to determine whether multiple satellite accretions are required, or whether consistency can be attained with a single object which has experienced bursts of star formation as it has orbited M31. The giant stream is linked to another stellar overdensity, the NE shelf, on the basis of nearly identical CMD morphologies and RGB luminosity functions; indeed, this coupling seems likely in view of progenitor orbit calculations (e.g. Ibata et al. 2004).} \end{itemize} {\sl Smooth Structure:} The INT/WFC survey provides the first opportunity to investigate the smooth underlying structure of M31 to unprecedented surface brightnesses. We have used the dataset to map the minor axis profile from the innermost regions to $\gtrsim 55$~kpc (Irwin et al. 2005). Figure 3 shows how the combination of inner diffuse light photometry and outer star count data can be used to trace the effective {\sl i}-band surface brightness profile to $\sim 30$~magnitudes per square arcsec. The profile shows an unexpected flattening (relative to the inner R$^{1/4}$ decline) at large radius, consistent with the presence of an additional shallow power-law stellar component (index $\approx -2.3$) in these parts. This component may extend out as far as 150~kpc (Guhathakurta, these proceedings). Taken together with our knowledge of the Milky Way halo, this finding supports the ubiquitous presence of power-law stellar halos around bright disk galaxies (see also Zibetti et al. 2004, Zibetti \& Ferguson 2004). The kinematics of M31's outer regions are being probed with Keck DEIMOS spectroscopy (e.g. Ibata et al. 2005). Two surprising results have emerged from our program so far. Firstly, there is a high degree of rotational support at large radius, extending well beyond the extent of M31's bright optical disk. Secondly, the overall coherence of this kinematic component is in striking contrast to its clumpy substructured appearance in the star count maps. Further work is underway to understand the nature and origin of this rotating component. \subsection{Results for M33} The right-hand panel of Figure 1 shows the RGB map of the low mass system, M33. Although it has the same limiting absolute depth as the M31 map, the stellar density distribution is extremely smooth and regular (Ferguson et al. 2006, in preparation). To a limiting depth of $\sim 30$~magnitudes per square arcsec (readily visible by eye here), the outer regions of M33 display no evidence for stellar substructure (c.f. the simulations of Bullock \& Johnston 2005). Equally surprising, our analysis of the isophote shape as a function of radius indicates no evidence for any twisting or asymmetries. M33 appears to be a galaxy which has evolved in relative isolation. The radial {\sl i}-band profile of M33 has been quantified via azimuthally-averaged photometry in elliptical annuli of fixed PA and inclination (Figure 3). The inner parts of the profile are constructed from diffuse light photometry, whereas the outer regions are derived from RGB star counts. The luminosity profile displays an exponential decline out to $\sim 8$~kpc (roughly 4.5 scalelengths) beyond which it significantly steepens. This behaviour is reminiscent of the ``disk truncations'' first pointed out by van der Kruit in the 80's, but until now not seen directly with resolved star counts. The steep outer component dominates the M33 radial light profile out to at least 14~kpc and limits the contribution of any shallow power-law stellar halo component in M33 to be no more than a few percent of the disk luminosity (Ferguson et al. 2006, in preparation). \section{Future Work} Quantitative study of the faint outskirts of galaxies provides important insight into the galaxy assembly process. The outskirts of our nearest spiral galaxies, M31 and M33, exhibit intriguing differences in their large-scale structure and stellar content. While M31 appears to have formed in the expected hierarchical fashion, M33 shows no obvious signatures of recent accretions. Observations of the outer regions of additional galaxies are required to determine which of these behaviours is most typical of the general disk population. \begin{chapthebibliography}{<widest bib entry>} \bibitem[Abadi et al.(2003)]{abadi03} Abadi, M.~G., Navarro, J.~F., Steinmetz, M., \& Eke, V.~R.\ 2003, \apj, 597, 21 \bibitem[Abadi et al.(2005)]{abadi03} Abadi, M.~G., Navarro, J.~F., \& Steinmetz, M. \ 2005, astro-ph/0506659 \bibitem[Brook et al.(2005)]{brook05} Brook, C.~B., Veilleux, V., Kawata, D., Martel, H., \& Gibson, B.~K.\ 2005, astro-ph/0511002 \bibitem[Bullock \& Johnston(2005)]{bj05} Bullock, J.~S., \& Johnston, K.~V.\ 2005, astro-ph/0506467 \bibitem[Ferguson et al.(2002)]{ferg02} Ferguson, A.~M.~N., Irwin, M.~J., Ibata, R.~A., Lewis, G.~F., \& Tanvir, N.~R.\ 2002, AJ, 124, 1452 \bibitem[Ferguson et al.(2005)]{ferg05} Ferguson, A.~M.~N., Johnson, R.~A., Faria, D.~C., Irwin, M.~J., Ibata, R.~A., Johnston, K.~V., Lewis, G.~F., \& Tanvir, N.~R.\ 2005, ApJL, 622, L109 \bibitem[Font et al.(2005)]{font05} Font, A.~S., Johnston, K.~V., Bullock, J.~S., \& Robertson, B.\ 2005,astro-ph/0507114 \bibitem[Governato et al.(2004)]{govern04} Governato, F., et al.\ 2004, \apj, 607, 688 \bibitem[Ibata et al.(2001)]{ibata01} Ibata, R., Irwin, M., Lewis, G., Ferguson, A.~M.~N., \& Tanvir, N.\ 2001, Nature, 412, 49 \bibitem[Ibata et al.(2004)]{ibata04} Ibata, R., Chapman, S., Ferguson, A.~M.~N., Irwin, M., Lewis, G., \& McConnachie, A.\ 2004, MNRAS, 351, 117 \bibitem[Ibata et al.(2005)]{ibata05} Ibata, R., Chapman, S., Ferguson, A.~M.~N., Lewis, G., Irwin, M., \& Tanvir, N.\ 2005, ApJ, 634, 287 \bibitem[Irwin et al.(2005)]{irwin05} Irwin, M.~J., Ferguson, A.~M.~N., Ibata, R.~A., Lewis, G.~F., \& Tanvir, N.~R.\ 2005, ApJL, 628, L105 \bibitem[McConnachie et al.(2003)]{mcconn03} McConnachie, A.~W., Irwin, M.~J., Ibata, R.~A., Ferguson, A.~M.~N., Lewis, G.~F., \& Tanvir, N.\ 2003, MNRAS, 343, 1335 \bibitem[Pritchet \& van den Bergh(1994)]{pvdb94} Pritchet, C.~J., \& van den Bergh, S.\ 1994, \aj, 107, 1730 \bibitem[Zibetti \& Ferguson(2004)]{2004MNRAS.352L...6Z} Zibetti, S., \& Ferguson, A.~M.~N.\ 2004, \mnras, 352, L6 \bibitem[Zibetti et al.(2004)]{2004MNRAS.347..556Z} Zibetti, S., White, S.~D.~M., \& Brinkmann, J.\ 2004, \mnras, 347, 556 \end{chapthebibliography}
Title: A radial velocity survey of low Galactic latitude structures: III. The Monoceros Ring in front of the Carina and Andromeda galaxies
Abstract: As part of our radial velocity survey of low Galactic latitude structures that surround the Galactic disc, we report the detection of the so called Monoceros Ring in the foreground of the Carina dwarf galaxy at Galactic coordinates (l,b)=(260,-22) based on VLT/FLAMES observations of the dwarf galaxy. At this location, 20 degrees in longitude greater than previous detections, the Ring has a mean radial velocity of 145+/-5 km/s and a velocity dispersion of only 17+/-5 km/s. Based on Keck/DEIMOS observations, we also determine that the Ring has a mean radial velocity of -75+/-4 km/s in the foreground of the Andromeda galaxy at (l,b)\sim(122,-22), along with a velocity dispersion of 26+/-3 km/s. These two kinematic detections are both highly compatible with known characteristics of the structure and, along with previous detections provide radial velocity values of the Ring over the 120<l<260 range. This should add strong constraints on numerical models of the accretion of the dwarf galaxy that is believed to be the progenitor of the Ring.
https://export.arxiv.org/pdf/astro-ph/0601176
\begin{keywords} Galaxy: structure -- galaxies: interactions -- Galaxy: formation \end{keywords} \section{Introduction} With the public release of all sky surveys, many details have been gained in the structure of the outer parts of the Galactic disc. In particular, the Sloan Digital Sky Survey (SDSS) revealed the existence of a stellar structure in the anticentre direction, near the Galactic plane and slightly over the edge of the disc \citep{newberg02}. Visible as a clear main sequence at a Galactocentric distance of $18\kpc$ for $180\deg<l<225\deg$ and $|b|<30\deg$, this structure is unlike what is expected for the Galactic disc. \citet{ibata03} used the INT Wide Field Camera to show this structure is in fact present in the second and third Galactic quadrants, circling the disc in a ring-like fashion. Radial velocity measurements also revealed a kinematically cold population with a velocity dispersion of only $15-25\kms$, once more unexpected for a Galactic structure, leading to the conclusion that this so-called Ring\footnote{This structure has been called many names including the Monoceros Ring, the Galactic Anticentre Stellar Structure or GASS and the Ring. Given its extent in Galactic longitude and its presence in numerous constellations, we prefer to call it simply the Ring.} is produced by the accretion of a dwarf galaxy in the Galactic plane whose tidal arms are wrapped around the Milky Way. Subsequently, much work has been invested in trying to determine the true extent of this Ring and where possible determine its kinematics to constrain the orbit of its progenitor. \citet{rocha-pinto03} used the 2MASS catalogue to probe the distribution of M giant stars and found that the Ring may extend over the $120\deg\simlt l\simlt270\deg$ range in the Northern hemisphere and be present in the Southern hemisphere, with $|b|<35\deg$. \citet{conn05a} extended the \citet{ibata03} INT/WFC survey to show the Ring is present within 30 degrees of the Galactic plane throughout the whole second quadrant but seems to disappear at $l\sim90\deg$. More constraints on the structure were gained by the \citet{crane03} spectroscopic survey of putative Ring M-giant stars in the anticentre direction ($150\deg<l<230\deg$, $|b|<40\deg$) which confirmed that most of these stars are indeed linked to this structure. Using the simple model of a population orbiting the Milky Way in a prograde, circular orbit, they determined their sample was best fitted by a population at a Galactocentric distance of $18\kpc$ and with a rotational velocity of $220\kms$. The resulting velocity dispersion of $20\pm4\kms$ around this model is in good agreement with the velocity dispersion measured by the SDSS team \citep{yanny03}. While tracking the Ring, two other structures were discovered that do not seem to be directly related to the Monoceros Ring. Using the 2MASS catalogue, \citet{rocha-pinto04} reported the existence of a diffuse population of M giants in the direction of the Triangulum and Andromeda constellations, behind known detections of the Ring. Also using the 2MASS catalogue, our group presented evidence of a dwarf galaxy located closer than the Ring, just at the edge of the Galactic disc, in the Canis Major constellation \citep{martin04,bellazzini04}. In particular, we argued that the accretion of the Canis Major galaxy onto the Galactic disc would naturally reproduce similar features as the one observed for the Ring. However, a link between these three structures remains putative at the moment, even though current models show such a scenario is highly plausible \citep{penarrubia05,martin05,dinescu05}. To gain more insight into the nature of these structures, we have started a radial velocity survey of regions at low Galactic latitude where the Ring structure may be located. Using the AAT/2dF multi-fibre spectrograph, we first targeted the Canis Major dwarf \citep{martin05}. This survey also revealed the presence of the Ring behind the dwarf, with a Heliocentric radial velocity of $133\pm1\kms$ and dispersion of $23\pm2\kms$ at a Galactocentric distance of $19\kpc$, highly compatible with previous detections \citep{conn05b}. In this third paper of the series, we report the presence of the Ring in the foreground of the Carina dwarf galaxy from VLT/FLAMES observations of the dwarf at $(l,b)=(260\deg,-22\deg)$. Using Keck2/DEIMOS observations of regions around M31, we also determine the radial velocity of the Ring at $(l,b)\sim(122\deg,-22\deg)$ where the Ring is known to exist but where its radial velocity has not yet been measured. Section 2 presents the Ring detection in front of Carina and Section 3 deals with the detection in front of the Andromeda galaxy. Section 4 concludes this letter. In the following, all the magnitudes have been corrected for extinction using the maps from \citet{schlegel98}. We also assume that the Solar radius is $R_\odot = 8\kpc$, that the LSR circular velocity is $220\kms$, and that the peculiar motion of the Sun is ($U_0=10.00\kms, V_0=5.25\kms, W_0=7.17\kms$; \citealt{dehnen98}). Except when stated otherwise, the radial velocities, $v_{r}$, are Heliocentric radial velocities, not corrected for the motion of the Sun. \section{The Ring in front of the Carina dwarf galaxy} FLAMES observations in the Carina fields were taken from the public access ESO raw data archive\footnote{http://www.eso.org/archive} and included 14 setups centred on various fields designed to study the Carina dwarf galaxy. The total integration time of each setup was typically $15000$ seconds giving excellent signal-to-noise ($>$20:1 per 0.2\AA\ sampling element) for the Galactic foreground stars in these directions. The Carina data plus calibration sequences were downloaded and processed through the standard ESO FLAMES low resolution pipeline. At the time of using the pipeline, sky subtraction was not part of the pipeline procedure so the remaining processing steps used the FLAMES software developed for the DART project (see e.g. \citealt{tolstoy04}). Briefly, this software stacks the individual repeat spectra on the same target field; combines all the sky observations to form a master sky spectrum; and then optimally scales, shifts (if necessary) and resolution-matches the sky to the object spectrum prior to sky subtraction. The sky-subtracted spectra are then searched for CaII near infrared triplet lines and a model template CaII spectrum is used to extract velocity information. The direct imaging used here also came from the ESO archive and comprised 4 $V,I$ sequences of ESO/WFI data covering $\approx 1 \times 1$ square degrees centred on the Carina dwarf. This was processed through the standard Cambridge pipeline \citep{irwin01} to produce catalogues of magnitudes, colours and object morphological classifications. The velocity distribution of all the stars targeted with FLAMES is displayed on Figure~1. The most prominent feature is of course the peak of stars at $v_{r}\sim220\kms$ produced by Red Giant Branch stars belonging to the Carina dwarf galaxy (the primary targets of these FLAMES observations). Stars with $v_{r}<180\kms$ are expected to be Galactic stars belonging to the thin disc, thick disc and/or stellar halo. However, they seem to follow a bimodal Gaussian distribution produced by populations with central velocities and dispersions of $(\mu_A,\sigma_A)=(49\pm7\kms,33\pm2\kms)$ and $(\mu_B,\sigma_B)=(145\pm5\kms,17\pm5\kms)$ and a 9 to 1 ratio according to a maximum likelihood fit of a two component Gaussian model. Although population A has the expected characteristics of a disc-like population in the foreground, the low dispersion of population B is more puzzling. The Colour-Magnitude Diagram (CMD) of Figure~2 displays the location of stars belonging to these two populations with hollow circles for population A stars ($v_{r}<110\kms$) and filled circles for population B stars ($110<v_{r}<180\kms$). The two populations show drastically different colour-magnitude distributions, pointing at a genuine difference between them. Population A is widely spread in colour and follows, once more, the expected distribution of foreground disc stars. On the other hand population B is confined on the bluer part of our sample, with almost all stars having $(V-I)_0<1.3$. Of the three stars with $(V-I)_0>1.3$, the one with the highest $(V-I)_0$ is only just a population B star with $v_{r}=112\kms$ (the two others have radial velocities of 143 and $149\kms$). Such a sharp colour cut makes it very unlikely that population B is produced by disc stars. Among the Galactic components, only the stellar halo is expected to have such behaviour (see e.g. \citealt{conn05a} for a comparison of the disc and halo Galactic component CMDs as they appear in the Besan\c{c}on model of \citealt{robin03}). However, the velocity dispersion of the halo is $\sim100\kms$ \citep[e.g.][]{gould03}, at odds with the $17\kms$ found for population B. Since this low dispersion could be an artifact of the radial velocity cuts we used, we investigate the number of stars that would be expected for a halo population at $(l,b)=(260\deg,-22\deg)$. According to the synthetic Besan\c{c}on model of Galactic stellar populations \citep{robin03}, there should be less than 225 halo stars per square degree in the CMD region highlighted on Figure~2 and that contains most of our population B stars, independently of radial velocity. Since our FLAMES targets roughly cover 0.25 square degrees, we would expect our sample to contain less than 60 halo stars if complete. Assuming the Milky Way is surrounded by a non-rotating stellar halo with a velocity dispersion of $100\kms$, our sample should contain only $\sim15$ halo stars within $110\kms<v_{r}<180\kms$, once again, if it were complete. Since within the selection box of Figure~2, the completeness is under 5\%, it is highly unlikely that the 56 population B stars belong to the halo. Therefore, population B is most likely a non-Galactic population of stars that lie in front of the Carina dwarf galaxy, with a mean radial velocity of $145\pm5\kms$ and a velocity dispersion of $17\pm5\kms$. A direct comparison with the velocity distribution of all the stars from the model that fall in the same region of the CMD would of course be more suited for our purpose. However the model greatly overpredicts the number of thick disc stars in this region of the sky when compared to the observations. This is probably due to an overestimate of the thick disc flare in the model. Although population A is very well reproduced by the thin disc population of the model, the modeled thick disc population is twice as numerous, centred on $v_{r}\sim80\kms$ and with a high dispersion of $\sim50\kms$. Such a population is clearly not present in our data. Since the sharp $(V-I)_0<1.3$ colour limit of population B is not expected for a disc population, we prefer comparing population B with only the halo population of the model. With a detection of the Ring behind the Canis Major galaxy under the Galactic disc only 20 degrees away in Galactic longitude from the feature we detect in front of Carina, this population could be another detection of the Ring. Previous detections of the Ring have mainly relied on CMDs and especially on the main sequence of this population compared to the disc population at brighter magnitudes \citep{newberg02,ibata03,conn05a}. Unfortunately, for the Carina CMD, the red clump of the dwarf galaxy at $(V-I)_0\sim0.7$ and $V_0\sim20.5$ lies in the region of interest for detecting the Ring main sequence. However, comparison with fiducials shows the location of population B stars is not incompatible with turn-off stars from the old metal-rich population at a Galactocentric distance of $\sim20\kpc$ that is usually assumed for the Ring; even though the Carina-driven selection criteria applied to select the stars in the sample prevents a reliable comparison. The radial velocity characteristics of this population in front of Carina further supports a connection with the Ring. Indeed, the determined velocity dispersion of $17\pm5\kms$ is within the $15-25\kms$ range of the SDSS and \citet{crane03} detections and close to the $24\pm2\kms$ of the detection in the background of Canis Major. For this latter case, the higher uncertainties on the 2dF radial velocity value of each star ($\sim10\kms$) compared to those of the FLAMES derived velocities ($\sim3\kms$) may also artificially increase the dispersion. When corrected from the Solar motion, the mean radial velocity of population B, $v_{\mathrm{gsr,B}}=-65\kms$, is very close to the $v_{\mathrm{gsr,bCMa}}=-67\kms$ found only 20 degrees away behind the Canis Major dwarf galaxy. Therefore, we conclude that the non-Galactic population we have uncovered in front of the Carina dwarf galaxy is most likely part of the Ring. \section{Kinematics of the Ring in front of the Andromeda galaxy} The presence of the Ring in front of the Andromeda galaxy was first reported by \citet{ibata03} from the analysis of INT/WFC Colour-Magnitude diagrams. The CMD of their `M31-N' field is shown on the left panel of Figure~3 and the Ring is clearly visible as a main sequence that extends from $(V-i,V)_{0}\sim(1.2,22)$ to $(V-i,V)_{0}\sim(0.6,20.0)$. During our Keck/DEIMOS survey of M31 outer disc and halo substructures \citep[e.g.][]{ibata05}, we took the opportunity to target foreground Ring stars that fortuitously fall in the targeted regions. The data were reduced as in \citet{ibata05} and the CMD of all the stars with a radial velocity uncertainty lower than $10\kms$ is presented on the right panel of Figure~3. To select only probable members of the Ring structure, we construct a selection box around the Ring main sequence from the INT CMD (box C in Figure~3). Among the stars in the DEIMOS sample, 86 fall within this Ring selection box and the radial velocity distribution of these stars is displayed on Figure~4. Aside from Galactic halo and M31 disc stars at $v_{r}<-200\kms$, a peak is apparent at $\sim-70\kms$. Applying a maximum likelihood algorithm to fit with a Gaussian model those stars with $-140<v_{r}<0\kms$ that produce the peak, reveals that this population has a mean velocity of $-75\pm4\kms$ and an intrinsic dispersion of $26\pm3\kms$, corrected for the uncertainties on each measured radial velocity. A direct comparison with the radial velocity distribution of stars from the Besan\c{c}on model within the same sky region ($119\deg<l<124\deg$ and $-24\deg<b<-19\deg$) and that fall in the same CMD selection box reveals the Ring sub-sample is not incompatible with the model. Indeed, a Kilmogorov-Smirnov test yields a probability of 10 percent that the two populations are identical. However, it is not unexpected that disc stars and Ring stars should show similar behaviour since both populations are believed to orbit the Milky Way on nearly circular orbits and the small shift in distance between them does not translate into a significant difference in radial velocity. Given that our selection box is constructed to contain Ring stars, it would be surprising that all the stars of the sample belong to the Galactic disc. On the contrary, we find it more likely that we are observing mainly Ring stars, as is suggested by the relatively low velocity dispersion of $26\pm3\kms$ in the sample, at odds with the $\sim50\kms$ found in the Besan\c{c}on model within the selection box. This low dispersion is also compatible with previous detections, especially since some disc and/or halo stars certainly fall in the same radial velocity range and increase the dispersion. The mean Heliocentric velocity of $-75\pm4\kms$ which converts to a Galactocentric standard of rest radial velocity of $v_{\mathrm{gsr}}=94\pm4\kms$ is also only slightly higher than the simple circular \citet{crane03} model. As for the detection in the foreground of the Carina dwarf, the radial velocity similarities between the previous detections of the Ring and the population in front of M31 strengthen the Ring nature of our detection. \section{Summary and Conclusion} We have presented the detection of two groups of stars that lie in front of the Carina dwarf galaxy at $(l,b)=(260\deg,-22\deg)$ and in front of the Andromeda galaxy at $(l,b)\sim(122\deg,-22\deg)$ and that cannot be satisfyingly explained by known Galactic components. The proximity with known detections of the Ring that surrounds the Galactic disc makes it highly probable that they belong to the same structure. Both detections have a low velocity dispersion ($17\pm5\kms$ and $26\pm3\kms$ respectively), a characteristic value encountered in all previous detections of the Ring. With the Ring detection behind the Canis Major dwarf galaxy reported by \citet{conn05b}, the radial velocity of the Ring population is now sampled throughout the $120\deg<l<260\deg$ range, which should provide important constraints on N-body models. However, it can be directly seen in Figure~5 that currently known radial velocity values for the Ring are not exactly reproducible by the \citet{crane03} simple circular model. In fact, trying to fit all the detections in a single orbit of the progenitor proves unsatisfactory, whether the orbit is forced to be circular or allowed to be slightly elliptical. It would therefore seem that models where the Ring completely surrounds the Galactic disc with multiple tidal arms are following the right track (see e.g. \citealt{penarrubia05} and \citealt{martin05}). In addition to the new radial velocities we report in this letter, such models would highly benefit from a similar survey as the one presented in \citet{conn05a}, but this time to higher longitudes to study in more detail the morphology of the Ring in these regions and especially to add a distance constraint to the detection in front of the Carina dwarf. \section*{acknowledgements} NFM is grateful to the IoA for the kind hospitality during the months at Cambridge in which this work was mainly performed. NFM acknowledges support from a Marie Curie Stage Research Training Fellowship under contract MEST-CT-2004-504604. GFL acknowledges support from ARC DP 0343508 and is grateful to the Australian Academy of Science for financially supporting a collaboratory visit to Strasbourg Observatory. \newcommand{\mnras}{MNRAS} \newcommand{\pasa}{PASA} \newcommand{\nat}{Nature} \newcommand{\araa}{ARAA} \newcommand{\aj}{AJ} \newcommand{\apj}{ApJ} \newcommand{\apjl}{ApJ} \newcommand{\apjs}{ApJSupp} \newcommand{\aap}{A\&A} \newcommand{\aaps}{A\&ASupp} \newcommand{\pasp}{PASP}
Title: Multifrequency observations of the jets in the radio galaxy NGC 315
Abstract: We present images of the jets in the nearby radio galaxy NGC 315 made with the VLA at five frequencies between 1.365 and 5 GHz with resolutions between 1.5 and 45 arcsec FWHM. Within 15 arcsec of the nucleus, the spectral index of the jets is 0.61. Further from the nucleus, the spectrum is flatter, with significant transverse structure. Between 15 and 70 arcsec from the nucleus, the spectral index varies from 0.55 on-axis to 0.44 at the edge. This spectral structure suggests a change of dominant particle acceleration mechanism with distance from the nucleus and the transverse gradient may be associated with shear in the jet velocity field. Further from the nucleus, the spectral index has a constant value of 0.47. We derive the distribution of Faraday rotation over the inner +/-400 arcsec of the radio source and show that it has three components: a constant term, a linear gradient (both probably due to our Galaxy) and residual fluctuations at the level of 1 - 2 rad/m^2. These residual fluctuations are smaller in the brighter (approaching) jet, consistent with the idea that they are produced by magnetic fields in a halo of hot plasma that surrounds the radio source. We model this halo, deriving a core radius of approximately 225 arcsec and constraining its central density and magnetic-field strength. We also image the apparent magnetic-field structure over the first +/-200 arcsec from the nucleus.
https://export.arxiv.org/pdf/astro-ph/0601660
\label{firstpage} \begin{keywords} galaxies: jets -- radio continuum:galaxies -- galaxies: individual: NGC\,315 -- magnetic fields -- polarization -- MHD \end{keywords} \section{Introduction} \label{intro} The giant FR\,I \citep{FR74} radio source NGC\,315 was first imaged by \citet{Brid76}, who showed that it has an angular size of nearly 1$^\circ$. The extended radio structure is described in more detail by \citet{Brid79}, \citet{Fom80}, \citet{Willis81}, \citet{Jaegers}, \citet{Venturi93}, \citet{Mack97} and \citet{Mack98}. The main jet has also been imaged extensively on parsec scales \citep{Linfield81,Venturi93,Cotton99,Xu00}. X-ray emission from the first 10\,arcsec of the main jet was detected by \citet{WBH}, but no optical emission from this region has yet been reported. The source is associated with a giant elliptical galaxy at a redshift of 0.01648 \citep{Trager}, giving a scale of 0.335\,kpc/arcsec for our adopted cosmology (Hubble constant $H_0$ = 70\,$\rm{km\,s^{-1}\,Mpc^{-1}}$, $\Omega_\Lambda = 0.7$ and $\Omega_M = 0.3$). NGC\,315 is a member of a group or poor cluster of galaxies \citep{Nolthenius,Miller02} located in one of the filaments of the Pisces-Perseus supercluster \citep{Ensslin01,Huchra}. HST images \citep{Verdoes99} show a 2.5-arcsec diameter dust lane and a nuclear point source. The dust lane is associated with a disk of ionized gas which is probably in ordered rotation \citep{Noel-Storr}. CO emission, also with a line profile indicating rotation, was detected by \citet{Leon}. The inferred mass of molecular hydrogen is $(3.0 \pm 0.3) \times 10^8$\,M$_\odot$ and the cold gas is likely to be cospatial with the dust. HI absorption against the nucleus was detected by \citet{vanG89}. There is evidence for a weak, polarized broad H$\alpha$ line in the nuclear spectrum \citep{Ho97,Barth99,Noel-Storr}. Hot gas associated with the galaxy has been imaged using {\sl ROSAT} and {\sl Chandra} \citep{WB,WBH}. Within $\approx$\,90 arcsec of the nucleus, the jets in NGC\,315 are initially narrow, then expand rapidly (``flare'') and re-collimate \citep{Brid82,CLBC}. We have modelled the inner $\pm$70\,arcsec of this {\em flaring region} as a two-sided, symmetrical, relativistic flow, fitting to deep, high-resolution VLA observations at 5\,GHz in order to derive the three-dimensional distributions of velocity, proper emissivity and magnetic-field structure \citep{CLBC}. Our main conclusions are as follows. \begin{enumerate} \item The jets are inclined by $38^\circ \pm 2^\circ$ to the line of sight. \item Where they first brighten, their on-axis velocity is $\beta = v/c \approx 0.9$. They decelerate to $\beta \approx 0.4$ between 8 and 18\,kpc from the nucleus (15 -- 33\,arcsec in projection) and the velocity thereafter remains constant. \item The ratio of the speed at the edge of the jet to its value on-axis ranges from $\approx$0.8 close to the nucleus to $\approx$0.6 further out. \item The longitudinal profile of proper emissivity is split into three power-law regions separated by shorter transition zones and the emission is intrinsically centre-brightened. \item To a first approximation, the magnetic field evolves from a mixture of longitudinal and toroidal components to predominantly toroidal by 26\,kpc (48\,arcsec in projection). \item Simple adiabatic models fail to fit the emissivity variations. \end{enumerate} In the present paper, we investigate the energy spectrum of the relativistic particles in the jets of NGC\,315 in the context of the models developed by \citet{CLBC}. We use VLA observations at frequencies between 1.365 and 5\,GHz\footnote{The 5-GHz observations are those discussed by \citet{CLBC}} to derive the spectrum of the jets at resolutions of 5.5 and 1.5\,arcsec and relate the observed spectral gradients to velocity, emissivity and field structure. A separate paper (Worrall et al., in preparation) will describe the radio structure of the main jet at high resolution and its relation to new {\sl Chandra} images. We also determine the variations of Faraday rotation over the jets and test the hypothesis that these result from magnetic-field irregularities in hot, X-ray emitting plasma associated with the surrounding group of galaxies. Finally, we determine the apparent magnetic-field structure of the jets on scales larger than those covered by \citet{CLBC}. In Section~\ref{obs}, we describe the observations and their reduction. The total-intensity images are presented in Section~\ref{Images} and we use them to derive distributions of spectral index in Section~\ref{Spectra}. We then discuss the distributions of Faraday rotation (Section~\ref{Faraday}) and apparent magnetic-field structure (Section~\ref{Field}) derived from observations of linear polarization. Section~\ref{Summary} summarizes our main results. \section{Observations and images} \label{obs} \subsection{Observations} VLA data were obtained at 4.985\,GHz in the B, C/D and A/D configurations as described by \citet{Venturi93} and \citet{Cotton99}. These were supplemented by additional observations in the A and C configurations with a centre frequency of 4.860\,GHz to give complete coverage of the spatial scales accessible to the VLA in a single pointing. In order to map Faraday rotation, we observed at 1.365, 1.413, 1.485 and 1.665\,GHz in the B and C configurations of the VLA using a lower bandwidth. We also extracted observations in A configuration at 1.413\,GHz from the VLA archive.\footnote{In addition, we re-analysed the C-configuration dataset at 8.4\,GHz from \citet{Venturi93}, but poor weather during the observations precluded accurate absolute flux and polarization calibration, so we do not discuss them here.} A journal of observations is given in Table~\ref{record}. \begin{center} \begin{table} \caption{Record of VLA observations. $\nu$ and $\Delta\nu$ are the centre frequency and bandwidth, respectively, and t is the on-source integration time.} \begin{center} \begin{tabular}{clllc} \hline Config- & Date & $\nu$ & $\Delta\nu$ & t \\ uration & & (MHz) & (MHz) & (min) \\ \hline B & 1989 Apr 13 & 4985.1 & 50 & 279 \\ B & 1995 Oct 25 & 4985.1 & 50 & 396 \\ C/D & 1996 May 10 & 4985.1 & 50 & 417 \\ A/D & 1996 Oct 07 & 4985.1 & 50 & 428 \\ A & 1996 Nov 2 & 4860.1 & 100 & 586 \\ C & 1997 Jul 12 & 4860.1 & 100 & 283 \\ A & 1980 Dec 21 & 1413.0 & 25 & 473 \\ B & 2001 Mar 19 & 1365.0 & 12.5 & 108 \\ B & 2001 Mar 19 & 1413.0 & 12.5 & 108 \\ B & 2001 Mar 19 & 1485.0 & 12.5 & 109 \\ B & 2001 Mar 19 & 1665.0 & 12.5 & 107 \\ C & 2001 Jul 17 & 1365.0 & 12.5 & 61 \\ C & 2001 Jul 17 & 1413.0 & 12.5 & 61 \\ C & 2001 Jul 17 & 1485.0 & 12.5 & 65 \\ C & 2001 Jul 17 & 1665.0 & 12.5 & 57 \\ \hline \end{tabular} \end{center} \label{record} \end{table} \end{center} \subsection{Data reduction} \label{reduction} All of the data reduction was done in the {\sc aips} package. Initial amplitude and phase calibration were applied using standard methods and the flux-density scales were set using observations of 3C\,48 or 3C\,286. Standard instrumental polarization calibration was applied and the zero-points of ${\bf E}$-vector position angle were determined using observations of 3C\,138 or 3C\,286. The data for each configuration were first adjusted to a common phase centre in J2000 coordinates, imaged and self-calibrated separately. They were then concatenated in turn, starting with the widest configuration. The slight difference in centre frequency between the datasets at 4.86 and 4.985\,GHz was ignored (we show in Section~\ref{RM} that this has a negligible effect on the analysis of polarization) and we will refer to the combination as the ``5\,GHz dataset''. At this frequency, the core showed significant variability between observations (cf.\ \citealt{Lazio01}) and the flux density of the unresolved component in the larger-configuration dataset was adjusted to match that observed with the smaller configuration when both were imaged at matched resolution. No core variability was detected at lower frequencies. A further iteration of phase self-calibration was done after each combination. Our final datasets are listed in Table~\ref{Datasets}, together with the minimum and maximum spatial scales they sample. \begin{center} \begin{table} \caption{Final uv datasets. The columns are: (1) centre frequency, (2) array configurations used, (3) minimum and (4) maximum spatial scales. \label{Datasets}} \begin{tabular}{llll} \hline $\nu$ & Configurations &\multicolumn{2}{c}{Scales (arcsec)}\\ (MHz) & & Min & Max \\ 1365 & BC & 4 & 900\\ 1413 & ABC & 1.2 & 900\\ 1485 & BC & 4 & 900\\ 1665 & BC & 4 & 900\\ 4985/4860 & ABCD & 0.4 & 300\\ \hline \end{tabular} \end{table} \end{center} Our observations were designed to give the maximum sensitivity for the inner jets of NGC\,315 and were therefore taken with the pointing centre on or near the nucleus. As can be seen from Table~\ref{Datasets}, the maximum spatial scale sampled adequately at 5\,GHz is $\approx$300\,arcsec. Even at L-band, the lobe associated with the counter-jet (\citealt{Mack97} and Fig.~\ref{fig:ilow}a) is severely attenuated by the primary-beam response of the VLA and is not visible on our images. We did not recover the total flux density of the source at any frequency, so we estimated appropriate zero-spacing flux densities from the shortest-spacing visibility amplitudes. We made images at five resolutions: 45, 5.5, 2.35, 1.5 and 0.4 arcsec FWHM, using similar baseline ranges at all frequencies and weighting the data in the uv plane as required. After imaging, we made both {\sc clean} and maximum-entropy deconvolutions. Although the latter algorithm gave slightly smoother images, it introduced a significant large-scale ripple parallel to the jet axis, whereas {\sc clean} gave a flat background. We therefore show the {\sc clean} images, although we quote quantitative results only where the two deconvolution methods agree. We also compared the $I$ images made with and without zero-spacing flux densities and before and after subtraction of a local zero-level. None of these differences led to significant changes in spectral index or degree of polarization compared with the errors quoted below. After deconvolution, all of the images were corrected for primary beam attenuation. We then took averages of the $I$ images at 1.365 -- 1.665\,MHz (``mean L-band images''). Data in Stokes $Q$ and $U$ were imaged without zero-spacing flux densities and {\sc clean}ed. A first-order correction for Ricean bias \citep{WK} was applied to the images of polarized intensity $P = (Q^2+U^2)^{1/2}$ used to derive the degree of polarization $p = P/I$. The off-source noise levels at the centre of the field for the final images are given in Table~\ref{noise} (the 0.4-arcsec image at 5\,GHz is discussed in detail by Worrall et al., in preparation, and is therefore not considered further here). Note that the wide-field L-band images at a resolution of 5.5\,arcsec are significantly affected by bandwidth smearing in their outer regions, images of point sources being broadened by a factor of 2 in the radial direction at a distance of 22\,arcmin from the phase centre \citep{ObsSS}. This limitation needs to be taken into account only in the discussion of the source morphology on large scales (Section~\ref{Images}). Measurements of spectra are restricted to the inner 200\,arcsec of the field, where the effects of bandwidth smearing are $<$3\% in peak intensity or image size for any of our frequency/resolution combinations. Our estimates of Faraday rotation, which extend to larger scales, should not be affected systematically by bandwidth smearing. \begin{center} \begin{table} \caption{Image resolutions and noise levels. $\sigma_I$ is the off-source noise level on the $I$ image; $\sigma_P$ the average of the noise levels for $Q$ and $U$. The noise levels were evaluated before correction for the primary beam response and apply only at the centre of the field for corrected images. Asterisks denote the images used by \citet{CLBC} \label{noise}} \begin{tabular}{llcc} \hline $\nu$ & FWHM &\multicolumn{2}{c|}{rms noise level} \\ (GHz) & (arcsec) &\multicolumn{2}{c|}{($\mu$Jy / beam area)} \\ & &$\sigma_I$&$\sigma_P$ \\ Mean L & 45 & 200 &$-$ \\ 1.365 & 5.5 &41 & 35 \\ 1.413 & 5.5 &38 & 34 \\ 1.485 & 5.5 &37 & 33 \\ 1.665 & 5.5 &37 & 30 \\ Mean L & 5.5 & 17 &$-$ \\ 5 & 5.5 &15 & 12 \\ 5*& 2.35 & 10 & 8 \\ 1.413 & 1.5 & 36 & 36 \\ 5 & 1.5 & 10 & 10 \\ 5*& 0.40 & 13 & 7 \\ \hline \end{tabular} \end{table} \end{center} \section{Total intensity} \label{Images} A grey-scale of the large-scale radio structure of NGC\,315 at 327\,MHz \citep{Mack97} is given in Fig.~\ref{fig:ilow}(a). Our observations are sensitive only to emission from the region indicated by the box in this figure and our mean L-band image of this region at a resolution of 45\,arcsec FWHM is displayed in Fig.~\ref{fig:ilow}(b). A higher-resolution (5.5\,arcsec FWHM) image of the same area is shown in Fig.~\ref{fig:i5.5full} and a detail of the sharp bend in the main jet $\approx$20\,arcmin from the nucleus is plotted at the same resolution but on a larger scale in Fig.~\ref{fig:ibend}. Finally, the 5-GHz emission from the inner 4\,arcmin of the jets is shown in Fig.~\ref{fig:i2.35} at a resolution of 2.35\,arcsec FWHM. We refer to the NW and SE jets as the {\em main} and {\em counter-}jets, as the former is brighter at most distances from the nucleus. A striking feature of the main jet is its almost constant width between $\approx$100 and $\approx$400~arcsec from the nucleus (Fig.~\ref{fig:i5.5full}). This is the {\em collimation shoulder} identified by \citet{Willis81}; a similar feature is visible in the counter-jet, but cannot be traced out as far. The lack of expansion over such an extended region is surprising if the jets are confined solely by thermal plasma associated with the surrounding galaxy group, as a significant pressure gradient would be expected on scales of a few hundred arcsec (we argue in Section~\ref{RMorigin} that the core radius of the group-scale plasma is $\approx$225~arcsec). An alternative possibility is that the jets also respond to the ${\bf J \times B}$ forces of their own toroidal fields on scales $\ga$100\,arcsec. In Section~\ref{Field}, we show that the observed polarization structure is consistent with a dominant toroidal component (see also \citealt{CLBC}), but we cannot tell from the high-frequency synchrotron emission alone whether this component is vector-ordered or has many reversals (evidence from Faraday rotation is also inconclusive; see Section~\ref{RMorigin}). The possibility that both pressure confinement and magnetic confinement could act together to produce a collimation shoulder was discussed by \citet{BCH}, but they assumed rather different physical parameters from those we now consider appropriate for NGC\,315. Our observations show that the collimation shoulder in the main jet ends at a bright feature with a sharp edge on the side towards the nucleus, inclined by $\approx$60$^\circ$ to the jet axis. At this point (marked ``Deflection'' in Fig.~\ref{fig:i5.5full}), the flow changes direction by $\approx$8$^\circ$ and re-expands with an opening angle $\approx$10$^\circ$ (defined in terms of the jet FWHM and the angular distance from the nucleus; \citealt{Willis81}). At a similar distance, the counter-jet does not deflect significantly, its surface brightness decreases monotonically away from the nucleus and it expands less rapidly than the main jet \citep[Fig.~\ref{fig:ilow}b]{Willis81}. The brightness distributions in both jets show large-scale ``banding'' -- repeated, but irregular alternation of bright and faint regions with surface brightnesses differing by factors of 1.5 to 2 -- along their lengths on arcminute scales (Fig.~\ref{fig:i5.5full}). The brightness bands extend across both jets but their variations are slower than those in the flaring region or at the edges of the jets. These variations could, in principle, result either from periods of enhanced activity in the nucleus or from interactions between the jets and their surroundings. If they were due to fluctuations in activity in the nucleus that propagated outwards at constant velocity $\beta c$, then they would appear at projected distances $D_{\rm j}$ and $D_{\rm cj}$ in the main and counter-jets, respectively, where $D_{\rm j}/D_{\rm cj} = (1+\beta\cos\theta)/(1-\beta\cos\theta)$ and $\theta \approx$ 38$^\circ$ \citep{CLBC}. Any transverse velocity gradients or deceleration will complicate this expression and the former effect should distort the bands into arcs that are concave towards the nucleus. We see no obvious relation between the distances of the bands in the two jets for any plausible value of $\beta$ and no evidence for systematic concave curvature of the bands beyond the flaring region. Furthermore, the most prominent banding appears to be associated with regions where the jets deflect or change their collimation properties. It therefore seems more likely that the banding is associated with ongoing interactions between the jets and their surroundings, although we cannot rule out a contribution to large-scale brightness fluctuations from slow variations in the jet output. The remarkable 180$^\circ$ bend in the main jet at the West end of the source is well known from earlier observations. Our data (Fig.~\ref{fig:ilow}b) show the emission after the bend at a resolution comparable to the 610-MHz WSRT image presented by \citet{Mack97}. The brightness distribution at the first part of the bend (where the jet deflects by $\approx$100$^\circ$) shows complex structure at 5.5-arcsec resolution. A bright ridge runs along the outside edge, with a lane of reduced emission next to it (both features are labelled on Fig.~\ref{fig:ibend}). Note that the suggestion that the flow is re-energised by an intergalactic shock \citep{Ensslin01} applies to emission further downstream, after the second bend. The compact knot of emission at the SW edge of the jet just before the bend (Fig.~\ref{fig:ibend}) coincides in location and shape with the optical emission from the flattened background galaxy FGC 0110 (z = 0.021965) and appears not to be physically related to NGC\,315. The source at RA 00 57 38.710, Dec. 30 22 44.99 (J2000; labelled ``Background source'' in Fig.~\ref{fig:i2.35}a) is unresolved and has a flat spectrum. The polarized flux density and ${\bf E}$-vector position angle vary smoothly across this position, consistent with addition of an unpolarized point source to the jet emission. There is a faint optical counterpart on the Digital Sky Survey and an X-ray point source is detected at a consistent position with the {\em ROSAT} PSPC (\citealt{WB}; fig.~2). The source is likely to be a background quasar, despite its location on the projected jet axis. There is a strong, diffuse component of emission on the axis of the main jet (``On-axis enhancement'' in Fig.~\ref{fig:i2.35}a) at $\approx$225\,arcsec from the nucleus, with no obvious counterpart in the counter-jet. The 2.35-arcsec resolution images illustrate the initial flaring of the jets (discussed in detail by \citealt{CLBC} and Worrall et al., in preparation) followed by recollimation to an almost uniform diameter \citep{Willis81,Brid82}. Although similar behaviour is observed in other sources \citep{LB02a,CL}, the physical scale on which NGC\,315 flares and recollimates is unusually large. This is probably a consequence of the low external density \citep{WBH}, and we will explore this idea quantitatively elsewhere, using the conservation-law approach developed by \citet{LB02b}. The jets in NGC\,315 bend slightly as they recollimate, from a position angle of $-48.5^\circ$ close to the nucleus to $-52.8^\circ$ at distances $\ga$100\,arcsec from the nucleus. The outer isophotes of the main and counter-jets are very similar before the bend, but the counter-jet is slightly wider at larger distances. The main jet is brighter than the counter-jet {\em on-axis} at all distances from the nucleus, but the counter-jet is brighter at the {\em edge} between $\approx$100 and 200\,arcsec. In the flaring region, by contrast, the main jet is significantly brighter than the counter-jet both on-axis and at the edge of the jet \citep{CLBC}. We successfully fit the structure of the inner $\pm$70\,arcsec of the jets in NGC\,315 using an intrinsically symmetrical model in which all apparent differences between the main and counter-jets result from relativistic aberration and beaming \citep{CLBC}. Such models clearly cannot be continued to indefinitely large scales, as FR\,I jets almost always show evidence (e.g.\ bends and disruption) for asymmetric interaction with the external medium. The largest-scale structure of NGC\,315 shown in Fig.~\ref{fig:ilow}(a) is an excellent example, with jets of unequal length terminating in entirely different ways. This raises the question: how large is the region over which symmetrical, relativistic models can be applied? At 400~arcsec from the nucleus, there are clear asymmetries in both the deflection and collimation properties; these must be intrinsic, although the overall brightness asymmetry persists. There is also good evidence for interaction with the external medium at distances of 70 -- 100\,arcsec on both sides of the nucleus, where both jets bend (symmetrically, this time). It seems likely that intrinsic and relativistic effects become comparable between 100 and 200~arcsec, where the sidedness difference on-axis is the same as on small scales, but the edge value is reversed. Asymmetries in apparent magnetic-field structure, which also imply that the flow remains mildly relativistic in this region, are discussed in Section~\ref{Field}. Our working hypothesis is therefore that relativistic effects dominate the observed differences between the two jets only before the first bends, but that environmental effects become first comparable and then dominant at larger distances, although slightly relativistic bulk flow probably continues to the largest scales and may remain responsible for the generally brighter appearance of the NW jet far from the nucleus. \section{Spectra} \label{Spectra} \subsection{Accuracy} We define spectral index $\alpha$ in the sense $S(\nu) \propto \nu^{-\alpha}$. We estimated spectral indices both for individual pixels and by integration of flux density over well-defined regions. Values at 5.5-arcsec resolution were determined from weighted power-law fits to data at all five frequencies between 1.365 and 5\,GHz. At 1.5-arcsec resolution, spectral indices were calculated between 1.413 and 5\,GHz. There are three main sources of error in the estimate of $\alpha$, as follows. \begin{enumerate} \item The transfer of the amplitude scale from the primary calibrator. The errors for the four L-band frequencies are likely to be tightly correlated, since they were observed during the same periods, data being taken simultaneously at 1365 and 1413\,MHz, and at 1485 and 1665\,MHz. Consequently, the principal effect of flux-density scale transfer errors is a constant offset in spectral index. \item Residual deconvolution effects, typically on scales of 5 -- 20\,arcsec. These are approximately proportional to surface brightness. \item Thermal noise. \end{enumerate} We model the error from (ii) as 0.03 times the flux density and that from (iii) as the noise level estimated off-source (from Table~\ref{noise}), appropriately integrated. These two contributions are added in quadrature. In addition, we estimate the rms spectral-index offset due to transfer errors in the flux-density scale to be 0.02. This should be taken in addition to the errors quoted below. \subsection{Spectral-index images and tomography} \label{tomography} Figs~\ref{fig:spec}(a) and (b) show the spectral indices, $\alpha$, determined from a weighted power-law fit to data at all five frequencies between 1.365 and 5\,GHz at 5.5\,arcsec resolution and between 1.413 and 5\,GHz at a resolution of 1.5\,arcsec, respectively. It is clear from Fig.~\ref{fig:spec}(a) that there are transverse gradients in spectral index where the jets are expanding rapidly. These gradients can only be see clearly on spectral-index images where the errors are small, and for this reason both panels of Fig.~\ref{fig:spec} are blanked where the rms error in $\alpha$ is $>0.05$. In order to search for transverse variations over a larger area, we need to average along the jets. The gradients between 34.5 and 69\,arcsec from the nucleus are best displayed by averaging along radii from the nucleus and plotting the results against angle from the local jet axis, as is shown for the main and counter-jets in Fig.~\ref{fig:transspec_radial}(a). Further from the nucleus, where the jets recollimate, we have averaged along the local jet axis to derive transverse spectral-index profiles. The results are shown for two regions in each of the main and counter-jets in Fig.~\ref{fig:transspec_long}. The fluctuations in these regions are dominated by quasi-periodic deconvolution errors: this problem is particularly acute for the counter-jet at $\sim$100\,arcsec from the nucleus (Fig.~\ref{fig:transspec_long}a). The spectral index is everywhere consistent with the mean value of $\langle \alpha \rangle = 0.47$ between 70 and 160\,arcsec. Another method of displaying spatial variations of the spectrum is ``spectral tomography'' \citep{K-SR,KSetal}. This involves the generation of a set of images $I_{\rm t}({\bf r}) = I({\bf r},\nu_1) - (\nu_2/\nu_1)^{\alpha_{\rm t}} I({\bf r},\nu_2)$ for a range of values of $\alpha_{\rm t}$, where $I({\bf r},\nu)$ is the brightness at position ${\bf r}$ and frequency $\nu$. If the brightness distribution can be represented as the sum of two components with different spectral indices $I({\bf r},\nu) = S_{\rm a}({\bf r})\nu^{-\alpha_{\rm a}} + S_{\rm b}({\bf r})\nu^{-\alpha_{\rm b}}$, then the ``a'' component will disappear from the image $I_{\rm t}$ when $\alpha_{\rm t} = \alpha_{\rm a}$. We made a set of images of $I_{\rm t}$ with $\nu_1 = 1.365$\,GHz, $\nu_2 = 5$\,GHz and $\alpha_{\rm t}$ from 0.4 -- 0.7 in steps of 0.01. For $\alpha_{\rm t} \approx 0.44$, the outer edges of both jets disappear at distances from the nucleus between $\approx$22\,arcsec and $\approx$80\,arcsec. The image of $I_{\rm t}$ for $\alpha_{\rm t} = 0.44$ is shown in Fig.~\ref{fig:spec_tomo}. Further from the nucleus, $I_{\rm t}$ is close to zero across the whole width of both jets for $\alpha_{\rm t} \approx 0.47$. There is no single value of $\alpha_{\rm t}$ for which the steep-spectrum component vanishes completely in an image of $I_{\rm t}$. The main features of the spectral-index distribution are as follows. \begin{enumerate} \item Variations in the spectral index are subtle ($0.4 \la \alpha \la 0.65$ everywhere). \item The spectral index is slightly, but significantly steeper in the jet bases than elsewhere. Between 7.5 and 22.5\,arcsec from the nucleus the mean values at 5.5-arcsec resolution are 0.63 and 0.58 in the main and counter-jets, respectively (the difference between them is not significant). \item At 1.5 arcsec resolution (Fig.~\ref{fig:spec}b), the spectral index of the main jet is essentially constant for the first $\approx$15\,arcsec, with a mean spectral index $\langle\alpha\rangle = 0.61$, consistent with the value determined at lower resolution. \item Between $\approx$15 and 70 arcsec from the nucleus, the spectral index is steeper on-axis ($\alpha \approx 0.5$) than at the edges ($\alpha \approx 0.44$) in both jets. This is illustrated by the transverse profiles averaged between 34.5 and 69\,arcsec from the nucleus (Fig.~\ref{fig:transspec_radial}a). \item At 1.5-arcsec resolution, the on-axis spectral index is slightly higher between 15 and 60\,arcsec from the nucleus ($\langle\alpha\rangle = 0.55$; Fig.~\ref{fig:spec}b) than at smaller distances. $\alpha$ cannot be determined to adequate accuracy at the edges of the jet for distances $\ga$15\,arcsec at this resolution. \item The tomographic analysis shows the spectral gradient in a different way: if we subtract off a component with $\alpha_{\rm t} = 0.44$ (Fig.~\ref{fig:spec_tomo}), the emission at the edges of the jet and at large distances from the nucleus essentially vanishes. What remains (positive in Fig.~\ref{fig:spec_tomo}) corresponds to the jet bases and to a ridge of steeper-spectrum emission at larger distances. The latter is clearly visible in both jets. \item The flatter-spectrum edge first becomes detectable at $\approx$15\,arcsec from the nucleus and widens thereafter. It occupies the entire width of the jet from $\approx$70\,arcsec outwards. The transition between steeper and flatter spectrum on-axis is poorly defined. \item There is no evidence for any transverse spectral gradient at larger distances, after the jets recollimate, although the data are noisy and do not cover quite the full width of the jets (Fig.~\ref{fig:transspec_long}). \item This is confirmed by the tomographic analysis: $I_{\rm t}$ for the outer parts of the region vanishes for $\alpha_{\rm t} = 0.47$, the mean spectral index, confirming that $\alpha$ is constant within our errors. \end{enumerate} \subsection{Deprojection of the spectral-index distribution} \citet{KSetal} and \citet{K-SR} suggested that the spectral index of an {\em on-axis} component in a jet is the value of $\alpha_{\rm t}$ at which the component appears to vanish against the background of the surrounding emission (exactly as for an {\em edge} component such as that in the NGC\,315 jets; Section~\ref{tomography}) and can therefore be derived simply from a tomographic analysis. This requires an additional assumption which may not be correct, namely that when the the on-axis component is subtracted, the remaining emission has a smooth brightness distribution (this is {\em not} true for the model proposed below). For NGC\,315, our three-dimensional model of the emissivity \citep{CLBC} gives a good fit to the observed emission at 5\,GHz, so we isolated the on-axis component and measured its spectral index, as follows. \begin{enumerate} \item We first used the tomography image with $\alpha_{\rm t} = 0.44$ as a template for the on-axis component. To a reasonable approximation, this defines a cone with a half-opening angle of 13.6$^\circ$ projected on the sky. Assuming an angle to the line of sight of $\theta = 38^\circ$ \citep{CLBC}, the half-opening angle in the jet frame is $8.3^\circ$ . \item We then made a model 5-GHz image, as in \citet{CLBC} but with the emissivity set to zero within this cone, convolved it to a resolution of 5.5 arcsec and subtracted it from the observed 5-GHz $I$ image. The residual corresponds to the observed emission from within the central cone. \item We scaled the model to 1.365\,GHz assuming a spectral index of 0.44, as appropriate for the edge emission and then subtracted it from the observed 1.365-GHz image. \item Both residual images showed little emission towards the edges of the jets, implying that the model subtraction was reasonably accurate. \item Finally, we derived a spectral-index image for the on-axis component alone. This is shown in Fig.~\ref{fig:spinespec}. \end{enumerate} Fig.~\ref{fig:spinespec} shows that the spectrum of the on-axis component flattens slightly with distance from the nucleus. The mean values of $\alpha$ between 7.5 and 22.5\,arcsec are 0.60 and 0.61 for the main and counter-jets, respectively. The corresponding values for distances between 22.5 and 66.5 arcsec are 0.56 and 0.55. For the main jet, these values are consistent with the measurements at 1.5-arcsec resolution without subtraction. At 5.5-arcsec resolution in both jets, they are slightly higher than the values measured on-axis before subtraction, which include a contribution from the flatter-spectrum, off-axis component. We conclude that the structure observed in Fig.~\ref{fig:spec}(a) does not result entirely from the superposition of two components with constant, but different spectral indices. The sketch in Fig.~\ref{fig:specsketch} summarizes our results on the distribution of spectral index in the jets. \subsection{Comparison with other sources} Typical spectral indices for the flaring regions of other FR\,I radio jets are in the range 0.5 -- 0.6 (\citealt{Young} and references therein). Indeed, \citet{Young} suggested that FR\,I jets have a ``canonical'' low-frequency spectral index of $\alpha = 0.55$, but this conclusion is based primarily on lower-resolution data than we consider here. Their canonical value is intermediate between that for the jet bases in NGC\,315 and the significantly flatter spectra seen at larger distances. There are very few published studies of spectral {\em variations} in the flaring regions of FR\,I jets. Short regions of slightly steeper spectrum than the average have been detected in the bases of three other FR\,I jets: 3C\,449 \citep{K-SR}, PKS\,1333$-$33 \citep{KBE} and 3C\,31 (Laing et al., in preparation). The measurement of $\alpha = 0.7 \pm 0.2$ in 3C\,449 applies to the faint inner jets (corresponding to the innermost 5\,arcsec in NGC\,315), but differs marginally from that of the brighter nearby emission, given the large error. In PKS\,1333$-$33, the spectrum flattens from $\alpha \approx 1.0$ to $\alpha \approx 0.6$ between 10 and 35\,arcsec from the nucleus (2.4 -- 8.4\,kpc in projection); this includes both the faint inner jets and the bright part of the flaring region, as in NGC\,315. In 3C\,31, there is a steeper-spectrum region extending to $\approx$6\,arcsec in the main jet (plausibly also in the counter-jet). There is a flatter-spectrum edge on one side of the main jet in 3C\,31 (Laing et al., in preparation). Transverse variations of spectral index in the sense that the spectrum is {\em steeper} on-axis have not been reported in any other sources, but their flaring regions are smaller in both linear and angular size and are difficult to resolve. Gradients at the level of $\Delta\alpha \approx$ 0.05 -- 0.1 are also tricky to detect without data at more than two frequencies. The tendency for the jet spectrum to be flatter at the edge is in the opposite sense to the spectral gradients found on large scales in tailed radio sources \citep[Laing et al., in preparation]{KSetal,K-SR} but the latter effect occurs in completely different regions of the jets, where they merge into the tails. \subsection{Acceleration mechanisms} The X-ray emission detected by \citet{WBH} in NGC\,315 coincides with the steeper-spectrum ($\alpha =0.61$) region at the base of the main jet. The form of the synchrotron spectrum in those FR\,I jet bases which emit X-rays is now well established \citep{Hard01,Hard02,Parma03,Hard05,PW05}. It can be characterised as a broken power law with spectral indices of 0.5 -- 0.6 at radio wavelengths and 1.2 -- 1.6 in X-rays. The spectrum of NGC\,315 is consistent with this pattern, although its high-frequency slope is poorly constrained \citep{WBH}. This spectral shape is consistent with synchrotron emission from a single electron population; ongoing particle acceleration is therefore required (see also Worrall et al., in preparation). The magnetic-field strengths calculated for the on-axis emissivity model of \citet{CLBC}, assuming a minimum-pressure condition, range from 3.3\,nT at 6\,arcsec in projection to 0.44\,nT at 69\,arcsec. For electrons with Lorentz factor $\gamma$ radiating at the synchrotron critical frequency $\nu_{\rm c}$ in a magnetic field $B$, we have $\nu_{\rm c}$/Hz = 41.99 $\gamma^2$ ($B$/nT) \citep{Longair}, so at 5\,GHz, $6 \times 10^3 \la \gamma \la 1.6 \times 10^4$ and in the X-ray band at 1\,keV, $4 \times 10^7 \la \gamma \la 1.1 \times 10^8$. The flatter-spectrum edges occur where \citet{CLBC} infer substantial velocity shear across the jets. If the jets are relativistic and faster on-axis than at their edges, the approaching jet always appears more centre-brightened than the receding one. This difference is reflected in the sidedness-ratio image. The average transverse sidedness profile between 34.5 and 69\,arcsec from the nucleus is shown in Fig.~\ref{fig:transspec_radial}(b) for comparison with the spectral-index profile over the same region. The velocity profile is modelled as a truncated Gaussian function with $\beta = 0.38$ on-axis and 0.22 at the edge, although \citet{CLBC} note that the on-axis velocity may be larger ($\beta \approx$ 0.5) if the shear occurs over a narrow range of radii in the jet: this might give a better fit to the sidedness profile. The flatter-spectrum edge first becomes visible $\approx$15\,arcsec from the nucleus. This coincides to within the errors with: \begin{enumerate} \item the {\em start} of rapid deceleration, as inferred by \citet{CLBC}, 14\,arcsec in projection from the nucleus; \item the {\em end} of the region of enhanced radio and (in the main jet only) X-ray emissivity \citep[Worrall et al., in preparation]{CLBC}. \item The first point at which the observed jet/counter-jet sidedness image gives any evidence for transverse velocity gradients, 16\,arcsec from the nucleus \citep{CLBC}. \end{enumerate} The deceleration and enhanced emission regions are marked on Fig.~\ref{fig:specsketch}. Note that the detection of transverse gradients in sidedness and spectral index may be limited by resolution. The changes of spectral index observed in NGC\,315 suggest that (at least) two different electron acceleration mechanisms are required, as follows. \begin{enumerate} \item The first mechanism dominates at the base of the flaring region (the initial 15\,arcsec in NGC\,315) and may continue at a lower level on the axis of the jet to $\approx$70\,arcsec. It generates emission from radio to X-ray wavelengths and has a characteristic spectral index $\alpha \approx 0.6$ in the former band. Three pieces of evidence suggest that this mechanism is dominant where the jet is fast ($\beta \approx 0.9$). Firstly, its characteristic spectral index is observed across the whole of the jet width in NGC\,315 until the start of rapid deceleration. Secondly, in both 3C\,31 \citep{LB04} and NGC\,315 \citep{WBH}, the bright X-ray emission occurs upstream of the deceleration region. Finally, \citet{LB04} show from radio data alone that significant injection of fresh relativistic particles is required before the start of deceleration in 3C\,31 to counter-balance adiabatic losses. \item The second mechanism causes the flattening of the spectrum towards the edges of the jets observed from 15\,arcsec outwards, but eventually spreading over the entire jet width. It produces spectral indices in the range $0.44 \la \alpha \la 0.5$ for electrons emitting at radio wavelengths in NGC\,315 and appears to be associated with velocity shear across the jets. A possible candidate for the this mechanism is the shear acceleration process described by \shortcite{RD1,RD2}, but their estimates of the acceleration timescale for electrons in conditions appropriate to FR\,I jets are very long (at least if the mean free path $\sim$ gyro-radius), and it is unclear whether the process is efficient enough to influence the spectrum. \end{enumerate} \section{Faraday rotation and depolarization} \label{Faraday} \subsection{Faraday rotation} \label{RM} In order to investigate the variations of Faraday rotation along the jets of NGC\,315, we made images of rotation measure (RM) at a resolution of 5.5\,arcsec by least-squares fitting to the relation $\chi(\lambda^2) = \chi(0) + {\rm RM}\lambda^2$ (where $\chi$ is the {\bf E}-vector position angle) for all 5 frequencies between 1.365 and 5\,GHz. The fits were weighted by errors in $\chi$ derived from Table~\ref{noise}. We excluded a small region around the core which was affected by residual instrumental polarization and included only points where the rms error in position angle was $<$15$^\circ$ at all frequencies. The resulting RM image is shown in Fig.~\ref{fig:RMimages}(a). The fit to a $\lambda^2$ law is very good everywhere: two examples are shown in Fig.~\ref{fig:rmegs}. The extreme value of RM $\approx -90$\,rad\,m$^{-2}$ results in rotations of $\approx1^\circ$ between the two centre frequencies combined in our 5\,GHz dataset and $\approx$0.4$^\circ$ and $\approx$5$^\circ$ across the bands at 5 and 1.4\,GHz, respectively. The worst of these effects, rotation across the band at the lowest frequencies, results in a spurious depolarization $<$0.1\%, which is negligible compared with errors due to noise. The images of $Q$ and $U$ show little power on large spatial scales and our estimates of position angle should be reliable over the full range of distances from the nucleus shown in Fig.~\ref{fig:RMimages}(a). The area over which the RM can be determined accurately is limited primarily by the primary beam at 5\,GHz (540\,arcsec FWHM). The mean RM is $-$75.7\,rad\,m$^{-2}$. The most obvious feature of the RM image is a nearly linear gradient along the jets, as shown by the profile in Fig.~\ref{fig:RMprofiles}(a). In order to reveal smaller-scale structure in the RM, we initially fitted and subtracted a function ${\rm RM} = {\rm RM}_0 + a_x x$, where $x$ is measured along the axis from the nucleus ($a_x$ and ${\rm RM}_0$ are constants). We used an unweighted least-squares fit, as any attempt to weight by the estimated errors caused the brightest part of the main jet to be fitted at the expense of other regions. The gradient is $a_x = $ 0.025\,rad\,m$^{-2}$\,arcsec$^{-1}$ and the residual image is shown in Fig.~\ref{fig:RMprofiles}(b). This indicates that variations in RM across the jet are also significant, as can be seen more clearly in averaged profiles, particularly between 69 and 112.5\,arcsec from the nucleus (Fig.~\ref{fig:RMtrans}b). Closer to the nucleus, the gradient is barely visible (Fig.~\ref{fig:RMtrans}a), but the jets are narrower there and the profiles are consistent with the gradient measured at larger distances. The transverse variation also appears to be linear, so we fitted and subtracted a function ${\rm RM} = {\rm RM}_0 + a_x x + a_y y$, where $y$ is a coordinate transverse to the jet and $a_y$ is a constant. The gradients along and transverse to the jet axis become $a_x$ = 0.018\,rad\,m$^{-2}$\,arcsec$^{-1}$ and $a_y$ = 0.051\,rad\,m$^{-2}$\,arcsec$^{-1}$, respectively. Taken at face value, the best-fitting gradient is 0.054\,rad\,m$^{-2}$\,arcsec$^{-1}$ at an angle of 72$^\circ$ to the mean jet axis. Note, however, that the transverse gradient is essentially determined by a subset of the data from the widest parts of the main and counter-jets (Fig.~\ref{fig:RMtrans}b). Removal of the large-scale gradient leaves fluctuations in the local mean RM which appear significantly larger on the counter-jet side (Fig.~\ref{fig:RMimages}c). By definition, the signal-to-noise ratio in $I$ is lower in the counter-jet. Although this is partially offset by a higher average degree of polarization, the errors in RM are still larger than in the main jet. To evaluate the significance of the fluctuations, we considered only points with fitting errors $\leq$2.5\,rad\,m$^{-2}$ and calculated the expected errors in the means for boxes of length 30\,arcsec along the jet axis containing more than 50 such points (the errors are corrected for oversampling). The results are shown in Fig.~\ref{fig:RMprofiles}(b). The fluctuations are significant, and form an ordered pattern with a typical scale $\sim$100\,arcsec. They are larger by a factor of $\approx$2 in the counter-jet and the first bin of the main jet (within 30 arcsec of the nucleus) compared with the rest of the main jet. The residual fluctuations within the boxes (i.e.\ after subtracting the local mean) are comparable with the errors in RM except in the brightest regions close to the nucleus. We therefore made a first-order correction to the rms RM, $\sigma_{\rm RM raw}$, by subtracting the fitting error $\sigma_{\rm fit}$ in quadrature to give $\sigma_{\rm RM} = (\sigma_{\rm RM raw}^2 - \sigma_{\rm fit}^2)^{1/2}$. The profiles of $\sigma_{\rm RM raw}$ and $\sigma_{\rm RM}$ are both plotted in Fig.~\ref{fig:RMprofiles}(c), for the same selection of points as in Fig.~\ref{fig:RMprofiles}(b). The corrected profile is very uncertain, but suggests that $\sigma_{\rm RM}$ has a maximum of 2\,rad\,m$^{-2}$ close to the nucleus on the counter-jet side and may be slightly asymmetric in the sense that the rms RM is lower in the main jet than the counter-jet at the same distance from the nucleus. We can image these smaller-scale fluctuations directly only at the bright base of the main jet. There, the RM at a resolution of 1.5\,arcsec FWHM can be derived accurately from the difference between 1.413 and 5\,GHz position-angle images, using the lower-resolution data to resolve the $n\pi$ ambiguity. Fig.~\ref{fig:RMimages}(d) shows the RM at this resolution. Data are plotted only where the rms error in the fitted RM $<$2\,rad\,m$^{-2}$. Fluctuations are clearly detected: the rms is $\sigma_{\rm RM raw}$ = 2.1\,rad\,m$^{-2}$, giving $\sigma_{\rm RM} =$ 1.6\,rad\,m$^{-2}$ after making a first-order correction for fitting error, as above. This is in good agreement with the value for the innermost bin of the profile of rms RM for the main jet at lower resolution (Fig.~\ref{fig:RMprofiles}c). \subsection{Depolarization} \label{Depol} The variation of $p$ with wavelength at low resolution potentially measures fluctuations of RM across the observing beam which cannot be imaged directly with adequate sensitivity. This variation is small, and is best quantified by fitting to the first-order approximation $p(\lambda^2) \approx p(0) +p^\prime(0)\lambda^2$ where $p^\prime(\lambda^2) = dp/d(\lambda^2)$. We expect $p^\prime(0) < 0$ (depolarization) under most circumstances. The quantity $p^\prime(0)/p(0)$ is directly related to the commonly quoted depolarization ratio, but is biased in the present case, since both the gradient and the degree of polarization depend directly on a single high-frequency measurement (at 5\,GHz), so deviations in $p(0)$ and $p^\prime(0)$ are anticorrelated. We therefore tested for the presence of depolarization using the gradient $p^\prime(0)$ alone. We derived $p^\prime(0)$ by weighted least-squares fitting to images of $p$ at the 5 frequencies between 1.365 and 5\,GHz at a resolution of 5.5\,arcsec FWHM. Two sets of $p$ images were used: in the first, an estimate of the local zero-level was subtracted from the $I$ images before calculating $p = P/I$; in the second, the original $I$ images were used (any differences indicate systematic errors in the estimation of large-scale structure). The fitting weights were the inverse squares of errors in $p$ derived from the values in Table~\ref{noise} and points were only included if the errors were $<$0.3 at all frequencies. The resulting polarization gradients are very small, and there are no obvious variations. In order to determine the significance of the gradients, we measured their mean values over the main and counter-jets. The maximum scale of structure imaged accurately in total intensity is $\approx$300\,arcsec (Table~\ref{Datasets}), so we calculated the means for points between 9 and 150\,arcsec from the nucleus along the jet axis (excluding the small region around the core to avoid spurious instrumental polarization, as in Section~\ref{RM}). The mean values for the main and counter-jets derived from the zero-level corrected images were $\langle p^\prime(0)\rangle = 0.03 \pm 0.08$ and $0.16 \pm 0.16$, respectively. Without zero-level correction, the values become $0.05 \pm 0.08$ and $0.22 \pm 0.16$. We conclude that the increase of $p$ with wavelength for $\lambda \leq 0.22$\,m ($\nu \geq 1.365$\,GHz), which is in any case opposite to the expected effects of Faraday rotation, is not significant. \subsection{The origin of the rotation measure} \label{RMorigin} NGC\,315 has Galactic coordinates $l = 124.^\circ6$, $b = -32.^\circ5$. This is on the outskirts of Region A of \citet{SK}, where the majority of sources have large negative RM's ($\sim -100$\,rad\,m$^{-2}$ at the centre of the region). \citet{DK05} derived spherical-harmonic models of the Galactic RM distribution by fitting to RM's of large numbers of extragalactic sources. These models predict a Galactic contribution $\approx -47$\,rad\,m$^{-2}$ at the position of NGC\,315. The bulk of the mean RM of $-$75.7\,rad\,m$^{-2}$ is therefore likely to be Galactic in origin. \citet{SC86} determined RM variations, which they argued to be primarily Galactic, across a number of sources within Region A (their Region 1). Their plot of squared RM difference $\Delta{\rm RM}^2$ against separation suggests $\langle \Delta{\rm RM}^2\rangle^{1/2} \sim$ 10 -- 30\,rad\,m$^{-2}$ on a scale of 700\,arcsec, but with large uncertainties. The observed $\Delta{\rm RM}$ ($\approx 15$\,rad\,m$^{-2}$ on the same scale along the jets of NGC\,315; Fig.~\ref{fig:RMprofiles}a) is again consistent with a Galactic origin. The maximum change in RM across the jet ($\approx 4$\,rad\,m$^{-2}$ over 80\,arcsec; Fig.~\ref{fig:RMtrans}b) is within the range of the upper limits plotted by \citet{SC86}. We note, however, that NGC\,315 is located outside the core of Region A, so it may be that smaller values of $\langle \Delta{\rm RM}^2\rangle^{1/2}$ are appropriate. The linearity of the RM gradient along the jet and the fact that the maximum gradient is aligned neither with the jet axis nor with the minor axis of the galaxy both imply that little of the Faraday rotating medium is associated with the jets or with the host galaxy. In either case we would expect some non-linear variation with distance from the nucleus. We conclude that most of the mean RM and its linear gradient are likely to be Galactic in origin, but a significant contribution from material local to NGC\,315 is not ruled out. In particular, we cannot exclude the hypothesis that some of the RM gradient transverse to the jet results from an ordered toroidal field within or just outside the jet \citep{L81,Blandford93}. The associated position angle rotation between 1.365 and 5\,GHz is at most a few degrees and significant departures from $\lambda^2$ rotation would not be detectable even if the thermal plasma responsible for the Faraday rotation is mixed with the synchrotron-emitting material \citep{Burn66}. If local toroidal fields are solely responsible for the transverse RM gradients, then their vector directions must be the same in the main and counter-jets. The {\em residual} RM fluctuations are qualitatively very similar to those in 3C\,31 (Laing et al., in preparation) but have amplitudes that are 10 times smaller on similar angular scales. The larger-scale ($\sim$100\,arcsec) fluctuations are systematically lower on the main (approaching) jet side. The distribution of fluctuations on smaller scales is also consistent with such an asymmetry, but is not well determined. As with the transverse gradients discussed earlier, the observed position-angle rotations are too small to be sure that the RM fluctuations are due to foreground plasma, but the asymmetry between approaching and receding jets suggests an origin in a distributed magnetoionic medium surrounding the host galaxy, a possibility we now explore. The thermal plasma associated with NGC\,315 and observed using {\em Chandra} can be described by a beta model with a core radius of 1.55\,arcsec \citep{WBH}. This cannot be responsible for the RM fluctuations, which occur on far larger scales. The most likely hot plasma component to be responsible for the Faraday rotation would be associated with the poor group of galaxies surrounding NGC\,315 \citep{Nolthenius,Miller02}, but has not yet been detected in X-ray observations. Since RM fluctuations are seen in both jets, a spherical distribution is plausible. In order to make a rough estimate of the parameters of the putative group component, we took a simple model for the field structure in which cells of fixed size $l$ at radius $r$ contain randomly orientated fields $B(r)$ \citep{LD82,Felten96}. The density distribution was taken to be a beta model: $n(r) = n_0 (1+r^2/r^2_c)^{-3\beta_{\rm atm}/2}$ with $B(r) \propto n(r)^N$. We derived the variation of $\sigma^2_{\rm RM}$ along the projection of the jet axis by numerical integration, assuming that the jets have $\theta = 37.9^\circ$ everywhere. This is a similar approach to the calculation of depolarization asymmetry by \citet{GC91} and \citet{Tribble92}; our code also reproduces the analytical results of \citet{Felten96} for $\theta = 90^\circ$ and $N = 0$ or 0.5. Following \shortcite{Dolag01,Dolag06}, we assumed that $B(r) \propto n(r)$ or $N = 1$ (an empirical result derived from RM's of radio sources in and behind cluster and rich groups). Our detection of RM variations on a range of scales is qualitatively consistent with the idea that the power spectrum of the magnetic-field fluctuations is a power law \citep{Tribble91,EV03,Murgia,VE05}, but our data are too noisy and poorly sampled to constrain its functional form. \citet{Murgia} show that a single-scale model gives a very similar relation between the RM variance $\sigma^2_{\rm RM}$ and radius $r$ to that derived for the more realistic case of a power-law power spectrum provided that $l$ is interpreted as the correlation length of the magnetic field. Finally, we fitted the resulting $\sigma_{\rm RM}$ curves by eye to the profiles of RM fluctuations on different scales in Fig.~\ref{fig:RMprofiles}(b) and (c). We fixed the value of $\beta_{\rm atm} = 0.5$ and adjusted the core radius to give a reasonable fit to the profile. For both plots, we found that $r_c \approx 225$\,arcsec gave an adequate fit. Strictly speaking, $\sigma^2_{\rm RM}$ is the RM variance evaluated over a window much larger than the maximum fluctuation scale, whereas the profiles in Fig.~\ref{fig:RMprofiles} describe fluctuations over two different ranges of scale. We have therefore added the two model variances to give a rough estimate of the total (the curves shown in Fig.~\ref{fig:RMprofiles}b and c actually have the same amplitude). Note that the process of removing a linear trend from the RM profile will have suppressed some power in large-scale fluctuations, particularly transverse to the jet. The amplitude of the model variance profile is related to the central density and field, and to the correlation length \citep{Felten96}: $(n_0/{\rm m}^{-3})^2 (B_0/{\rm nT})^2 (l/{\rm kpc}) \approx 700$. This is a very rough estimate, but is enough to establish that a low-density group-scale gas component with a low magnetic field can generate the observed Faraday rotation. Plausible parameters might be $n_0 \approx$ 600\,m$^{-3}$, $B_0 \approx$ 0.015\,nT (0.15\,$\mu$Gauss) and $l \approx$ 10\,kpc. We note that the recollimation of the jets may also require a large-scale hot gas component. \section{Magnetic-field structure} \label{Field} Vectors whose magnitudes are proportional to $p$ at zero wavelength and whose directions are those of the apparent magnetic field inferred from the rotation-measure fit of Section~\ref{RM} are plotted in Fig.~\ref{fig:ivec5.5}. At higher resolution, we derived the apparent field direction by interpolating the RM image onto a finer grid and using it to correct the observed 5-GHz position angles (Fig.~\ref{fig:ivec2.35}). The apparent field structure in the flaring region (up to $\approx$70 arcsec from the nucleus) is discussed extensively by \citet{CLBC}. Here, we concentrate on larger scales, after the jets recollimate. The signal-to-noise ratio for individual points, particularly at the jet edges, is often $<$3 in linear polarization, causing them to be blanked in Figs~\ref{fig:ivec5.5} and \ref{fig:ivec2.35}. In particular, it is impossible to see whether the parallel-field edge continues to large distances in the main jet. We therefore adopted the following procedure to derive the average degree of polarization. \begin{enumerate} \item We first corrected the observed 5-GHz, 2.35-arcsec position angles for Faraday rotation using the linear model derived in Section~\ref{RM}, which is defined everywhere in the field, unlike the RM image. \item We then changed the origin of position angle to be along the jet axis, so that apparent field along or orthogonal to the jet appears entirely in the $Q$ Stokes parameter (also verifying that there is very little signal in $U$). \item We then integrated $Q$ and $I$ along the jet axis from 69 -- 113 and from 113.5 -- 157.5 arcsec from the nucleus (the same areas used for the profiles of spectral index in Fig.~\ref{fig:transspec_long}). Lack of short spacings at 5\,GHz precludes extension of this analysis to larger distances. \item Finally, we divided the results to give transverse profiles of $Q/I$. Provided that the apparent field is either along or orthogonal to the jet axis, $p = |Q/I|$. We have chosen the sign convention so that $Q > 0$ for transverse apparent field and $< 0$ for longitudinal field. \end{enumerate} The resulting profiles are shown in Fig.~\ref{fig:QIprof}. Figs~\ref{fig:ivec5.5} -- \ref{fig:QIprof} show that the apparent field configuration found in the flaring region -- transverse on-axis and longitudinal at the edges -- persists until at least 160\,arcsec in both jets. In particular, the longitudinal-field edge of the main jet is easily detected in the profiles, even though it is not clearly visible on the images. The main difference from the corresponding profiles for the flaring region (fig.~9 of \citealt{CLBC}) is that the on-axis polarization is higher in both jets at larger distances. This is a continuation of the trend in the longitudinal profile shown by \citet[their fig.~8]{CLBC}. As in the flaring region, the on-axis (perpendicular) polarization is always higher in the counter-jet, reaching levels close to the theoretical maximum of $p_0 = 0.69$ for the observed spectral index. In the main jet, $p \approx 0.4$ on-axis. Both jets are very highly polarized at their edges. \citet{CLBC} modelled the three-dimensional structure of the field in the outer parts of the flaring region as a mixture of toroidal and longitudinal components of roughly equal magnitude on-axis but with the former dominant at the edge of the jet. The differences between the two jets are attributed to relativistic aberration, so the fact that they persist after the jets recollimate suggests that there is little further deceleration in this region, despite the reversal in sidedness at the edges of the jets (Section~\ref{Images}). A decrease in the on-axis longitudinal field component, bringing the configuration closer to a purely toroidal one, would result in a polarization distribution consistent with that observed. Some longitudinal component must remain, however, otherwise $p$ would be close to $p_0$ on the axis for both jets. The very high degree of polarization at the jet edge requires the radial field component to be very small, as inferred for the flaring region by \cite{CLBC}. \section{Summary} \label{Summary} We have imaged the jets in the nearby FR\,I radio galaxy NGC\,315 with the VLA at five frequencies in the range 1.365 -- 5\,GHz and at resolutions ranging from 45 -- 1.5\,arcsec FWHM. Our total intensity observations reveal new details of the structure, particularly around the sharp bend in the main jet. The flaring regions of both jets, where they initially expand rapidly and then recollimate, show a complex and previously unknown spectral structure. Within 15\,arcsec of the nucleus, the spectral index has a uniform value of $\alpha =$ 0.61 in both jets. This region is associated with strong X-ray emission in the main jet, high radio emissivity, complex filamentary structure, and fast flow with $\beta \approx 0.9$ \citep{CLBC,WBH}. Between 15 and 70\,arcsec, the spectrum is steeper on-axis than at the edges of the jet. We have developed a novel deprojection technique which allows us to isolate two spectral components. The first (on-axis) forms a continuation of the jet base, its spectral index flattening gradually from 0.61 to 0.55. The second (at the edge of the jet) has $\alpha \approx$ 0.44 and is associated with a region where strong shear is inferred \citep{CLBC}. We speculate that two different acceleration mechanisms are involved, one associated with fast flow, dominant close to the nucleus and capable of accelerating electrons to the very high energies required to produce X-ray emission ($\gamma \sim 10^8$), the other being driven by shear and generating the flatter spectral indices seen at the edges of the jet. Both mechanisms must efficiently generate the electrons with $\gamma \sim 10^4$ which radiate at cm wavelengths. At distances $\ga$70\,arcsec, the spectral index is consistent with $\alpha \approx 0.47$ everywhere. We have imaged the variations of Faraday rotation over the jets. All of the rotation is resolved and must originate mostly in foreground material. There is no detectable depolarization. The largest contributions -- a constant term and a linear gradient -- are probably Galactic in origin. We have also detected residual fluctuations of $\approx$1 -- 2 rad\,m$^{-2}$ rms on scales $\sim$ 5 -- 100\,arcsec. The amplitude of fluctuations on scales $\ga$30\,arcsec is larger by a factor $\approx$2 for the counter-jet, consistent with an origin in magnetoionic material around the source, but not in the known X-ray-emitting halo, whose core radius is too small. We model the Faraday-rotating medium as a spherical halo with a core radius $\approx$225\,arcsec and derive an approximate value for the product $(n_0/{\rm m}^{-3})^2 (B_0/{\rm nT})^2 (l/{\rm kpc}) \approx 700$, where $n_0$ is the central density, $B_0$ the central magnetic field and $l$ is the magnetic-field correlation length. Our analysis is therefore consistent with models of Faraday rotation proposed for rich clusters (e.g.\ \citealt{CT}), but requires much lower densities and field strengths. We predict that a tenuous, group-scale halo should be detectable in sensitive X-ray observations; measurement of its density will allow us to estimate the magnetic-field strength. We have derived the apparent magnetic field direction (corrected for Faraday rotation) and degree of polarization at distances between 70 and 160\,arcsec from the nucleus. The structure is qualitatively similar to that seen in the flaring region, with transverse field on-axis and longitudinal field at the edges of both jets, but the degree of polarization on-axis is larger. The difference in polarization structure between the main and counter-jets observed in the flaring region by \cite{CLBC} persists at larger distances. This can be explained fully as an effect of differential aberration on radiation from intrinsically identical jets, as long as their velocities remain significantly relativistic on the relevant scales. The asymmetry in RM fluctuation amplitude is consistent with the jet orientation required by this analysis and the presence of a tenuous, magnetized group halo. The large angular size of the flaring region in NGC\,315 and our use of deep observations at several frequencies has allowed us to image spectral variations at a level of detail not yet achieved in any other jet. Taken together with X-ray imaging and modelling of the jet velocity field, this has given important insights into the particle acceleration mechanisms. It will be interesting to see whether our results apply to other FR\,I jets and to study the spectral variations in NGC\,315 over a wider frequency range. \section*{Acknowledgments} JRC acknowledges a research studentship from the UK Particle Physics and Astronomy Research Council (PPARC). The National Radio Astronomy Observatory is a facility of the National Science Foundation operated under cooperative agreement by Associated Universities, Inc. We thank Greg Taylor for the use of his rotation-measure code, Karl-Heinz Mack for providing the 327-MHz WSRT image, Frank Rieger for discussions on shear acceleration and the referee for helpful comments. We also acknowledge the use of the {\sc healpix} package ({\tt http://healpix.jpl.nasa.gov}) and the provision of the models of \citet{DK05} in {\sc healpix} format. \label{lastpage}
Title: A VLT-UVES spectrscopic analysis of C-rich Fe-poor stars
Abstract: Large surveys of very metal-poor stars have revealed in recent years that a large fraction of these objects were carbon-rich, analogous to the more metal-rich CH-stars. The abundance peculiarities of CH-stars are commonly explained by mass-transfer from a more evolved companion. In an effort to better understand the origin and importance for Galactic evolution of Fe-poor, C-rich stars, we present abundances determined from high-resolution and high signal-to-noise spectra obtained with the UVES instrument attached to the ESO/VLT. Our analysis of carbon-enhanced objects includes both CH stars and more metal-poor objects, and we explore the link between the two classes. We also present preliminary results of our ongoing radial velocity monitoring.
https://export.arxiv.org/pdf/astro-ph/0601253
\title{A VLT/UVES spectroscopic analysis of C-rich Fe-poor stars} \author{T. Masseron\inst{1,2}, B. Plez\inst{2}, F. Primas\inst{1}, S. Van Eck\inst{3}, \and A. Jorissen\inst{3}} \institute{ESO, Garching, Germany, \and GRAAL, Universit\'e Montpellier II, France, \and Institut d'Astronomie et d'Astrophysique, Universit\'e Libre de Bruxelles, Belgium} \section{Introduction} A subgroup of carbon stars, the CH stars, were first distinguished from other carbon stars 60 years ago (\cite{keenan42}). Their chief characteristics are : strong CH absorption lines, strong Swan bands of C$_2$, strong resonance lines of Ba~II and Sr~II (later shown to reflect genuine overabundances of s-process elements), weak lines of the iron-group elements and high proper motion. Since it was shown that all field CH stars were spectroscopic binaries (\cite{mcclure90}), the binary scenario was accepted: CH stars have been polluted by a nearby companion, formerly on the AGB, now a defunct white dwarf. But some recent results have cast doubts on this general picture. Conflicting evidences include the radial velocity monitoring of targets from the HK survey (\cite{beers92}) which show no radial velocity variations (\cite{preston01}), and the discovery of carbon but not s-process enhanced stars (e.g. \cite{sneden96}) . Furthermore, the unexpectedly high fraction of carbon-enhanced stars among metal-poor stars discovered in the HK survey (about 25\% in the metallicity range [Fe/H] $\le$ -2.5 compared to 1-2 \% among stars of higher metal abundances), make them crucial for the study of the early stages of Galaxy evolution. \begin{table} \caption{The sample and the atmospheric parameters.} \label{table} \begin{center} \leavevmode \footnotesize \begin{tabular}{l|c|c|c|c} \hline star &T$_{\rm eff}$ & log~g & [Fe/H] & $^{12}$C/$^{13}$C \\ \hline CS 22891-171 & 5100 & 1.8 & -2.31 & 5 \\ CS 22942-019 & 5100 & 2.5 & -2.48 & 7 \\ CS 22945-017* & 6400 & 4.3 & -2.60 & 4 \\ CS 22953-003 & 4600 & 1.0 & -3.27 & $>$5 \\ CS 22956-028* & 6700 & 3.5 & -2.38 & 3 \\ CS 30322-023 & 4000 & 0.0 & -3.78 & 5 \\ HD 26* & 5200 & 2.6 & -0.71 & ? \\ HD 5424* & 5000 & 3.0 & -0.41 & 4 \\ HD 24035* & 5000 & 3.7 & 0.16 & 4 \\ HD 168214 & 5200 & 3.5 & -0.03 & ? \\ HD 187861* & 4800 & 1.8 & -2.21 & 10 \\ HD 196944* & 5250 & 1.7 & -2.21 & ? \\ HD 206983 & 4550 & 1.6 & -0.93 & 4 \\ HD 207585* & 5800 & 4.0 & -0.35 & 10 \\ HD 211173 & 4800 & 2.5 & -0.17 & 13 \\ HD 218875 & 4600 & 1.5 & -0.63 & 100 \\ HD 219116 & 4800 & 1.8 & -0.45 & 7 \\ HD 224959* & 5000 & 2.0 & -2.08 & 7 \\ HE 1419-1324 & 4900 & 1.8 & -3.28 & 4 \\ HE 1410+0213* & 4550 & 1.0 & -2.38 & 4 \\ HE 1001-0243* & 5000 & 2.3 & -3.12 & 30 \\ \hline \multicolumn{5}{l}{* probably binary, from radial velocity monitoring}\\ \end{tabular} \end{center} \end{table} \section{Observations and abundance determinations} We have started a new extensive high-resolution analysis of carbon-enhanced stars, based on a sample including at the moment 86 stars, aimed at investigating the C-enrichment phenomenon over a wide range of metallicities, chosen in the Barktevicius catalog (1996), in the HK survey and in the Hamburg/ESO survey (\cite{christlieb01}). Observations were made with the VLT/UVES echelle spectrograph at high S/N and spectral resolution, as well as with the ESO 1m52/FEROS spectrograph. A few radial velocity observations were made with the OHP 1m93 ELODIE instrument.\\ In Table~\ref{table}, we present a subset of our sample for which we have derived C, N (from CN and CH bands), O (from [O~I] 630nm, when possible), and Ba and Eu abundances, from VLT observations.\\ Effective temperatures were initially estimated from photometry available in the literature, using the calibration of \cite{alonso99}. However, due to strong \ absorption by molecular CH and CN bands impacting the photometry by an unknown amount, T$_{\rm eff}$ were instead derived by forcing the abundance determined from individual Fe I lines to show no dependence on excitation potential. The gravity was determined from the ionization equilibrium of Fe~I and Fe~II. The microturbulent velocity was set by requiring no trend of Fe~I abundance with equivalent width. The observed spectra are compared to synthetic ones, computed with the "turbospectrum" package (\cite{alvaplez98}). This program uses OSMARCS atmosphere models, initially developed by \cite{gustafsson75} and later improved by \cite{plez92}; see \cite{gustafsson03} for details on recent improvements. \\ The line lists are the same as used by \cite{hill02} for atoms and for CH, C$_2$ and CN, and their various isotopes.\\ \section{Results} Figures~\ref{figcarbon}, \ref{figcarbonnitrogen}, \ref{fignitrogen}, and \ref{figeuba} present the derived abundances. Carbon is indeed enhanced in the atmospheres of most of our metal-poor targets (Fig.~\ref{figcarbon}). Excluding the 2 stars around [Fe/H] $= -3.5$ that do not show a large C enhancement, the average carbon abundance is almost constant from the CH stars ([Fe/H] $> -2.0$) to the very metal-poor carbon-enhanced stars (from [C/H]$\approx 0$ at solar Fe/H to [C/H]$\approx -0.5$ at [Fe/H]$\approx -3$). The lowest metallicity star HE 0107-5240 ([Fe/H] $=-5.3$, recently discovered by \cite{christlieb02}), shows a carbon abundance close to the solar value. This extreme star follows the trend of Fig.~\ref{figcarbon}. The nitrogen abundance, when combined to the carbon abundance and the carbon isotopic ratio, provides additional information. Figure~\ref{figcarbonnitrogen} shows the sum of C and N overabundance as a function of Fe/H. Some of the stars that did not show a large C/Fe do show a large (C+N)/Fe, as the most metal-poor star of our sample. The $^{12}$C/$^{13}$C ratio is generally low (see Table~\ref{table}), often close to the CN-cycle equilibrium value. There is a general anticorrelation between the N overabundance and $^{12}$C/$^{13}$C (Fig.~\ref{fignitrogen}), characteristic of the operation of the CN cycle. Note that the stars of our sample with low $^{12}$C/$^{13}$C and still on the main sequence (cf. log~g values in Table~\ref{table}) are binaries, supporting the mass-transfer scenario. The carbon and nitrogen enhancements are large, and whether the CN cycling happened in the stars we observe when they evolved to the giant stage, or in the companions that polluted them, remains to be determined. The combined (C+N)/Fe overabundance is much larger at lower metallicity, pointing towards a primary origin of these elements. \cite{asplund03} warns for 3D effects that may lead to overestimates of abundances derived from molecular lines in metal-poor stars, but we doubt that the trend could be erased by NLTE and 3D effects.\\ Ba is an s-process element and, as for carbon, its overabundance is explained by mass-transfer from a now extinct AGB star. The r-process element Eu is believed to originate from supernovae, although the r-process site(s) is still debated. Figure~\ref{figeuba} shows [Eu/Fe] vs. [Ba/Fe] for the stars of Table~\ref{table}. In addition to a few stars that are very enriched in both r- and s-process elements, two groups emerge: one enriched in Eu and not in Ba, and a large group of stars with [Ba/Fe] around +1, and no or little Eu enhancement. This latter group encompasses the solar metallicity Ba stars present in our sample.\\ The carbon-rich, very iron-poor stars are of various origins: (i) mass-transfer binaries polluted by an AGB or a more massive star, (ii) single stars, some enriched in s-process elements, other in r-process elements, some maybe in both. We are pursuing our analysis of more stars in our sample, and of more chemical elements, in order to provide useful constraints on their origin, and on the early chemical evolution of the Galaxy.
Title: Proper Motions of Dwarf Spheroidal Galaxies from Hubble Space Telescope Imaging. IV: Measurement for Sculptor
Abstract: This article presents a measurement of the proper motion of the Sculptor dwarf spheroidal galaxy determined from images taken with the Hubble Space Telescope using the Space Telescope Imaging Spectrograph in the imaging mode.
https://export.arxiv.org/pdf/astro-ph/0601547
\setcounter{figure}{0} \title{Proper Motions of Dwarf Spheroidal Galaxies from \textit{Hubble Space Telescope} Imaging. IV: Measurement for Sculptor.\footnote{Based on observations with NASA/ESA \textit{Hubble Space Telescope}, obtained at the Space Telescope Science Institute, which is operated by the Association of Universities for Research in Astronomy, Inc., under NASA contract NAS 5-26555.}} \author{Slawomir Piatek} \affil{Dept. of Physics, New Jersey Institute of Technology, Newark, NJ 07102 \\ E-mail address: piatek@physics.rutgers.edu} \author{Carlton Pryor} \affil{Dept. of Physics and Astronomy, Rutgers, the State University of New Jersey, 136~Frelinghuysen Rd., Piscataway, NJ 08854--8019 \\ E-mail address: pryor@physics.rutgers.edu} \author{Paul Bristow} \affil{Space Telescope European Co-ordinating Facility, Karl-Schwarzschild-Str. 2, D-85748, Garching bei Munchen, Germany \\ E-mail address: bristowp@eso.org } \author{Edward W.\ Olszewski} \affil{Steward Observatory, The University of Arizona, Tucson, AZ 85721 \\ E-mail address: eolszewski@as.arizona.edu} \author{Hugh C.\ Harris} \affil{US Naval Observatory, Flagstaff Station, P. O. Box 1149, Flagstaff, AZ 86002-1149 \\ E-mail address: hch@nofs.navy.mil} \author{Mario Mateo} \affil{Dept. of Astronomy, University of Michigan, 830 Denninson Building, Ann Arbor, MI 48109-1090 \\ E-mail address: mateo@astro.lsa.umich.edu} \author{Dante Minniti} \affil{Universidad Catolica de Chile, Department of Astronomy and Astrophysics, Casilla 306, Santiago 22, Chile \\ E-mail address: dante@astro.puc.cl} \author{Christopher G.\ Tinney} \affil{Anglo-Australian Observatory, PO Box 296, Epping, 1710, Australia \\ E-mail address: cgt@aaoepp.aao.gov.au} \keywords{galaxies: dwarf spheroidal --- galaxies: individual (Sculptor) --- astrometry: proper motion} \section{Introduction} \label{sec:intro} Shapley (1938) discovered the Sculptor dwarf spheroidal (dSph) galaxy --- the first example of this type of galaxy in the vicinity of the Milky Way --- on a plate with a 3~hour exposure time taken with the Bruce telescope. Shapely notes ``... that systems such as the Sculptor cluster may not be uncommon; their luminosity characteristics would enable them to escape easy discovery.'' Since the detection of Sculptor, astronomers have identified eight other dSphs. Sculptor is at a celestial location of $(\alpha,\delta) = (01^{\mbox{h}}00^{\mbox{m}}09^{\mbox{s}}, -33^{\circ} 42^{\prime}30^{\prime\prime})$ (J2000.0; Mateo 1998), which corresponds to Galactic coordinates of $(\ell, b)=(287\fdg 5,-83\fdg 2)$. Thus, Sculptor lies nearly at the South Galactic Pole. Kaluzny \etal\ (1995) searched for variable stars in a $15^{\prime} \times 15^{\prime}$ field centered approximately on the dSph by taking $V-$ and $I-$band images with the 1-m Swope telescope at Las Campanas Observatory over a period of more than two months. The search resulted in the identification of 226 RR~Lyr stars. The average $V-band$ magnitude of the RR~Lyr stars gives a distance modulus of $(m-M)_{V}=19.71$, which corresponds to a heliocentric distance of 87~kpc. This estimate is practically the same as that obtained by Hodge (1965); it is consistent with the estimate of Baade \& Hubble (1939), but somewhat larger than the estimate of Kunkel \& Demers (1977). This study adopts the estimate of Kaluzny \etal\ (1995) for the distance to Sculptor. Irwin \& Hatzidimitriou (1995) derives the most comprehensive set of structural parameters for Sculptor --- and seven other dSphs --- using star counts from UK Schmidt telescope plates. With a luminosity of $(1.4\pm 0.6) \times 10^{6}$~L$_{\odot}$, Sculptor is among the most luminous dSphs. Its major-axis core and limiting radii are $5.8 \pm 1.6$~arcmin and $76.5 \pm 5.0$~arcmin, respectively, which are in good agreement with the values derived by Demers, Kunkel, \& Krautter (1980). However, they differ from the values derived by Eskridge (1988a), who also uses star counts from UK Schmidt telescope plates. The discrepancy is likely due to an underestimation of the central density in the latter study for reasons that are discussed in Irwin \& Hatzidimitriou (1995). The isopleth map of Sculptor (Panel (f) of Figure~1 in Irwin \& Hatzidimitriou 1995) shows that the ellipticity of the isodensity contours increases with increasing projected radius from the center of the dSph: the ellipticity is consistent with 0 in the inner 10~arcmin and smoothly increases to a value of 0.32 in the outermost region. The study observes that Sculptor ``... looks remarkably similar to numerical simulations of dSph galaxies that are tidally distorted.'' The position angle of the major axis is $99 \pm 1$~degrees. Eskridge (1988b) finds asymmetric ``structure'' in Sculptor in an isopleth map of the difference between the stellar surface density and a fitted 2-D model. In contrast, Irwin \& Hatzidimitriou (1995) finds only the increase of the ellipticity with projected radius in a similar map. The left panel in Figure~\ref{fig:fields} shows a 30~arcmin $\times$ 30~arcmin region of the sky centered on Sculptor. The dashed ellipse is the boundary of the core. Walcher \etal\ (2003) studies the structure of Sculptor --- together with those of Carina and Fornax --- using $V$-band images taken with the MPG/ESO 2.2-m telescope at La Silla. The images have an areal coverage of 16.25~square degrees and reach a limiting magnitude of $V \approx 23.5$. The study derives major-axis core and limiting radii of $7.56 \pm 0.7$~arcmin and $40 \pm 4$~arcmin, respectively, using a King (1962) model, as did Irwin \& Hatzidimitriou (1995). The 1-$\sigma$ disagreement between the two derived core radii is perhaps larger than expected from two data sets that have many stars in common. The apparently more serious disagreement between the two limiting radii is most likely due only to a larger true uncertainty in the limiting radius caused by uncertainties in the background surface density and the poor fit of the model to the outer part of the surface density profile. Walcher \etal\ (2003) confirms that the ellipticity of the surface-density contours increases with increasing projected radius; the contours also suggest ``extensions'' from the ends of the major axis that are interpreted as tidal tails. The radial projected density profile shows a ``break'' --- a departure from the fitted King model --- at around 30~arcmin, which the study interprets as evidence for the existence of an extended stellar component. Walcher \etal\ (2003) uses the relation between the King-model tidal radius, the mass, and the perigalacticon of a Galactic satellite developed by Oh, Lin, \& Aarseth (1992) to deduce that Sculptor has a perigalacticon of 28~kpc. A recent article by Coleman \etal\ (2005) does not confirm the finding in Walcher \etal\ (2003) that Sculptor has tidal tails and an extended stellar component. Instead, the analysis of photometric data in the $V$ and $I$ bands for a $3.1^{\circ}\times3.1^{\circ}$ field shows that a King model with a limiting radius of $72.5\pm4.0$~arcmin is a satisfactory fit to the radial profile of stars that lie on the giant branch. The limiting magnitudes of the photometry are $V=20$ and $I=19$, respectively. The study notes that oversubtracting the field population in its data produces a radial profile with the smaller fitted limiting radius and significant extratidal structure found in Walcher \etal\ (2003). Using additional information from the spectroscopy of 723 stars selected from the red giant branch, Coleman \etal\ (2005) derives an upper limit of $2.3\% \pm 0.6\%$ for the contribution from stars beyond the tidal boundary to the total mass of Sculptor. The study does not find any conclusive evidence for tidal interaction between the Milky Way and Sculptor. In contrast, Westfall \etal\ (2005) does find evidence. This study uses imaging in the $M$, $T_{2}$, and $DDO51$ bands to separate member giants from foreground dwarfs in a 7.82 degrees$^{2}$ area that covers the eastward side of Sculptor including the central region. Candidate members are also selected from the region of the blue horizontal branch in the color-magnitude diagram. The selection of members is checked with spectroscopy for 147 candidates. The study finds members up to 150 arcmin from the center of Sculptor --- the spatial extent of the survey and beyond the tidal boundary if that is identified with the measured King limiting radius of 80~arcmin. Several of these stars are spectroscopically-confirmed members. The radial surface brightness profile shows a break to a shallower slope at a radius of about 60~arcmin, which resembles the radial profiles seen in simulations of satellites interacting with the Milky Way (Johnston \etal\ 1999). Thus, Westfall \etal\ (2005) argues in favor of a significant tidal interaction between the Milky Way and Sculptor. It is beyond the scope of this work to resolve the apparent conflict between Coleman \etal\ (2005) and Westfall \etal\ (2005) by judging the merits of the analyses presented in both articles. Needless to say, a disagreement exists about the effect of the Galactic tidal field on the structure of Sculptor; measuring the proper motion of the dSph may allow us to impose constraints on this effect. Armandroff \& Da Costa (1986) measured the radial velocities of 16 giants in Sculptor which average to produce a systemic heliocentric velocity of $107.4 \pm 2.0$~km~s$^{-1}$. This measurement alleviated the large uncertainty in this quantity, which existed due to mutually contradictory estimates from Hartwick \& Sargent (1978) and Richter \& Westerlund (1983). More recently, Queloz, Dubath, \& Pasquini (1995) measured the radial velocities of 23 giant stars. The sample includes 15 stars observed previously by Armandroff \& Da Costa (1986). The implied systemic heliocentric velocity of $109.9 \pm 1.4$~km~s$^{-1}$ (after excluding two stars that are likely binaries) agrees within the quoted uncertainties with the measurement of Armandroff \& Da Costa (1986). Our article adopts a mean velocity of $109.9 \pm 1.4$~km~s$^{-1}$ for calculating the space velocity of Sculptor. Queloz, Dubath, \& Pasquini (1995) finds no apparent rotation of the dSph around its minor axis. Tolstoy \etal\ (2004) measured radial velocities for 308 potential members of the dSph and find a systemic velocity of 110~km~s$^{-1}$. There is no discussion of rotation, but Figure~4 in that article shows that the velocity dispersion does not increase with radius, as would be expected if there were a net rotation larger than the central dispersion. Interestingly, the study finds that the red horizontal branch stars have a more compact spatial distribution and a smaller velocity dispersion than the older and more metal-poor blue horizontal branch stars. The two most recent photometric and spectroscopic surveys by Coleman \etal\ (2005) and Westfall \etal\ (2005) confirm the greater central concentration of the more metal-rich stars. The two studies also find no evidence for rotation. The mass-to-light ratio, $(M/L)$, of Sculptor is larger than a typical value for a Galactic globular cluster; it is, however, smaller than the $M/L$s for some other Galactic dSphs. Armandroff \& Da Costa (1986) derives a central $M/L_{V}$ of $6.0 \pm 3.1$ and Queloz, Dubath, \& Pasquini (1995) determines the somewhat larger value of $13 \pm 6$, in solar units. Both studies note that the measured $M/L$ does not imply unequivocal support for dark matter in Sculptor. The stars of Sculptor are old. Fitting isochrones to the principal sequences in the color-magnitude diagram, Da Costa (1984) finds that the majority of the stars are younger by ``2--3~Gyr than Galactic globular clusters of similar metal abundance provided the helium abundances and the CNO/Fe ratios are also similar.'' Da Costa (1984) also detects ``blue stragglers'' and estimates their age to be about 5~Gyr under the assumption that they are ``normal'' main-sequence stars, i.e., stars which did not acquire mass from a companion. No stars younger than 5~Gyr exist in Sculptor, indicating an absence of ongoing or recent star formation. However, Sculptor contains HI gas. Carignan \etal\ (1998) and later Bouchard \etal\ (2003) detect two distinct clouds of HI that are diametrically opposite to each other almost along the minor axis and 20 -- 30~arcmin from the center of the dSph. These clouds are within the tidal radius. The HI gas is very likely to be associated with Sculptor because its mean heliocentric velocity is similar to that of the dSph. Carignan \etal\ (1998) discusses mechanisms that might account for the existence of the clouds. Removing the gas from the dSph by a time-dependent tidal force due the Milky Way is one possibility. Carignan \etal\ (1998) suggests that the alignment between the proper motion vector from Schweitzer \etal\ (1995) and a line passing through the two clouds supports this hypothesis for the origin of the clouds. However, if the tidal force affects the HI, it should also affect the stars and the possible signatures of tides in the stellar component of Sculptor are either absent or inconsistent with the direction of the Schweitzer \etal\ (1995) proper motion vector. For example, the increasing ellipticity of isodensity contours with increasing projected radius could be due to tides (e.g., Johnston, Spergel, \& Hernquist 1995), but then the major axis should be along the proper motion vector. The possible ``tidal extensions'' reported by Walcher \etal\ (2003) are at the ends of the major axis. If these extensions are a continuation of the increasing ellipticity, they also argue for an orbital plane parallel to the major axis. Walcher \etal\ (2003) claims that the eastern extension bends to the south, i.e., parallel to the minor axis and so argues that the orbital plane is aligned in the north-south direction. However, this alignment is inconsistent with the increasing ellipticity being due to the tidal force since, given the large distance of Sculptor, the Sun is nearly in the orbital plane and so the increasing ellipticity should be aligned with the tidal extensions. Schweitzer \etal\ (1995) reports the first measurement of the proper motion for Sculptor: $(\mu_{\alpha},\mu_{\delta})=(72 \pm 22, -6 \pm 25)$~mas~century$^{-1}$. This value includes contributions from the motions of the Sun and LSR; this article refers to this quantity as the ``measured proper motion." The measurement derives from 26 photographic plates imaged with a variety of ground-based telescopes using either a ``blue'', B, or V filter. The earliest epoch is 1938 and the latest is 1991. The study estimates, among other quantities, the perigalacticon of the implied orbit. The best estimate ranges from 60~kpc for the ``infinite halo'' potential of the Milky Way to 78~kpc for the ``point mass'' potential. If the perigalacticon is no smaller than 60~kpc, then the Galactic tidal force has not played a significant role in the evolution of Sculptor. Numerical simulations of Piatek \& Pryor (1995) or Oh, Lin, \& Aarseth (1995) show that for a typical dSph, even with a $M/L_V$ as low as 3, a perigalacticon of 60~kpc is too large for tides to have an important effect. Motivated by the idea that some of the Galactic dSphs and globular clusters may be pieces of a tidally-disrupted progenitor satellite galaxy, several studies propose that they form ``streams'' in the Galactic halo. Lynden-Bell (1982) hypothesizes that Fornax, Leo I, Leo II, and Sculptor are members of the ``FLS stream.'' Majewski (1994) adds the newly-discovered Sextans to the FLS stream and recalculates its common plane --- naming it the ``FL$^{2}$S$^{2}$ plane.'' The FLS and the FL$^{2}$S$^{2}$ planes differ only slightly. In a more extensive study, Lynden-Bell \& Lynden-Bell (1995) infers that Sculptor may belong to one of three possible streams (see their Table~2). Stream \# 2 contains the LMC, SMC, Draco, Ursa Minor, and, possibly, Sculptor and Carina; stream \# 4a contains Sextans, Sculptor, Pal 3, and, possibly, Fornax; finally, stream \# 4b contains Sextans, Sculptor, and, possibly, Fornax. For each stream, Lynden-Bell \& Lynden-Bell (1995) calculates the expected proper motion of Sculptor. Kroupa \etal\ (2004) notes that the 11 dwarf galaxies nearest to the Milky Way form a disk with a thickness to radius ratio of $\leq$~0.15. The article argues that the distribution expected for such nearby substructure in a cold-dark-matter universe is spherical, that the observed distribution is not, and, thus, that these objects are the tidal debris from the disruption of a larger satellite galaxy. In contrast, Kang \etal\ (2005) and Zentner \etal\ (2005) find that a planar distribution of nearby Galactic satellites is actually common in numerical simulations of galaxy formation. A direct comparison between the results of the simulations and the distribution of nearby satellites finds that they are consistent. A test of the reality of streams or planar alignments is to measure the space motions of the satellites. Piatek \etal\ (2005; P05) reports a proper motion for Ursa Minor. The implied orbit for Ursa Minor is not in the plane defined by Kroupa \etal\ (2004). The proper motion also rules out membership in the stream proposed by Lynden-Bell \& Lynden-Bell (1995). The measured proper motion for Carina (Piatek \etal\ 2003; P03) does not agree well with the predictions of Lynden-Bell \& Lynden-Bell (1995), but is not precise enough to rule out membership in a stream. Piatek \etal\ (2002; P02) finds that a preliminary proper motion for Fornax is inconsistent with the predictions of Lynden-Bell \& Lynden-Bell (1995) and that its direction is also inconsistent with an orbit in the FL$^{2}$S$^{2}$ plane. Dinescu \etal\ (2004) reports an independent measurement of the proper motion of Fornax. This motion is consistent, within its uncertainty, with the predictions of Lynden-Bell \& Lynden-Bell (1995) and the direction of this motion is along the great circle defined by the FL$^{2}$S$^{2}$ plane. Dinescu \etal\ (2004) notes that the proper motion for Sculptor in Schweitzer \etal\ (1995) is inconsistent with the FL$^{2}$S$^{2}$ plane. This article reports a second independent measurement of the proper motion for Sculptor and discusses the implications of the derived space motion on the dSph-Galaxy interaction. Section~\ref{sec:data} describes observations and the data. The following section describes the analysis of the data leading to the derivation of the proper motion. Section~\ref{sec:pmm} compares the proper motion from Schweitzer \etal\ (1995) with the one reported in this article. The next section, Section~\ref{sec:orbit}, integrates and describes the orbit of Sculptor. Section~\ref{sec:disc} discusses the implications of the orbit for the importance of the Galactic tidal force on the structure and internal kinematics of Sculptor, for the star formation history, and for the membership of Sculptor in the proposed streams of galaxies and globular clusters in the Galactic halo. The final section is a summary of the main results and conclusions. \section{Observations and Data} \label{sec:data} The Hubble Space Telescope (HST, hereafter) imaged two distinct fields in Sculptor using the Space Telescope Imaging Spectrograph (STIS, hereafter) in imaging mode with no filter (50CCD). Each field contains a known quasi-stellar object (QSO, hereafter), which serves as a reference point. The left panel of Figure~\ref{fig:fields} depicts the locations of the two fields on the sky: two small squares --- one inside and the other outside of the core. The name of the field inside the core is SCL~$J0100-3341$, which derives from the IAU designation of the QSO in this field. Tinney \etal\ (1997) confirms the identity of this QSO: it is at $(\alpha, \delta) = (01^{\mbox{h}}00^{\mbox{m}}25\fs 3, -33^{\circ} 41^{\prime}07^{\prime\prime})$ (J2000.0), has a redshift $z=0.602 \pm 0.001$, and has a magnitude $B=20.4$. The observations of the SCL~$J0100-3341$ field occurred on September 24, 2000 and on September 26, 2002. At each epoch, there are three exposures at each of the eight dither pointings for the total of 24 images. The ``ORIENTAT'' angle --- the position angle of the Y axis of the CCD measured eastward from north --- is the same to within one-tenth of a degree for all of the exposures and equal to -67.5 degrees. The top-right panel in Figure~\ref{fig:fields} shows the SCL~$J0100-3341$ field. The QSO is in the cross-hair. The name of the field outside of the core is SCL~$J0100-3338$. The QSO in this field, also confirmed by Tinney \etal\ (1997), is at $(\alpha, \delta) = (01^{\mbox{h}}00^{\mbox{m}}32\fs 6, -33^{\circ} 38^{\prime}32^{\prime\prime})$ (J2000.0), has a redshift $z=0.728 \pm 0.001$, and has a magnitude $B=20.4$. $HST$ observed this field on September 13, 1999; September 28, 2000; and on September 28, 2002. At each of the three epochs, there are three exposures at each of the eight dither pointings for a total of 24 images. The ``ORIENTAT'' angle is the same to within one-tenth of a degree and equal to -69.3 degrees for all of the exposures for this field. The bottom-right panel of Figure~\ref{fig:fields} shows the SCL~$J0100-3338$ field. The QSO is in the cross-hair. Owing to its greater distance from the center of the dSph, the SCL~$J0100-3338$ field contains fewer stars than does the SCL~$J0100-3341$ field. Bristow (2004) and P05 discuss the effect of the decreasing charge transfer efficiency of the STIS CCD on astrometric measurements. If not accounted for, the decreasing charge transfer efficiency may introduce a spurious contribution to a measured proper motion. Bristow \& Alexov (2002) developed computer software which approximately restores an image taken with STIS to its pre-readout condition. All of the results that this article reports are based on images restored using the program of Bristow \& Alexov (2002). \section{Analysis} \label{sec:analysis} P02 describes our method of deriving a proper motion from images taken with \textit{HST} and containing at least one QSO. Fundamental to the method is the concept of an effective point-spread function (ePSF, hereafter), which Anderson \& King (2000) describes in detail. The subsequent two articles in this series, P03 and P05, expand and improve upon the basic method. The analysis reported here incorporates only minor new features into the method; thus, the reader should consult those earlier articles for the details. Instead, this study mentions the major elements of the method alongside figures depicting key diagnostics of the performance of the method and briefly describes the new features. \subsection{Flux Residuals} \label{sec:rf} Equation 22 in P02 defines a ``flux residual'' diagnostic, $\cal{RF}$. It is the measure of how the shape of the constructed ePSF matches the shape of an image of an object. In the case of a perfect match, ${\cal RF} =0$; if the ePSF is narrower, ${\cal RF} > 0$; otherwise, ${\cal RF} < 0$. Several factors affect the shape of the PSF for an object. 1. Type of an object. A PSF for a galaxy is generally wider than that for a star, all else being equal. 2. Color of an object. Because of diffraction and aberrations, the width of the PSF is color-dependent. 3. Tilt or curvature of the focal plane. The PSF varies with location because the CCD surface and focal plane do not coincide everywhere. 4. Thermal expansion. Because the \textit{HST} moves in and out of the Earth's shadow, its temperature is continuously changing. These changes cause the telescope to expand or contract, affecting its focal length. 5. Charge traps in the CCD. As the packets of charge representing an object move along the $Y$ axis (the direction of readout for STIS), those on its leading side fill partially each trap encountered, so that there are fewer traps available to remove charge from subsequent packets (Bristow \& Alexov 2002). This non-uniform loss of charge across the object changes its PSF. Given the aforementioned factors affecting the shape of the PSF, a plot of ${\cal RF}$ \textit{versus} the $X$- or $Y$-coordinate of an object will, in the best case, show that the points scatter around ${\cal RF}=0$. In a less desirable case, the points may show trends with $X$ or $Y$ or both. These trends signal that the true PSF varies with location. Because of the scarcity of stars in the observed fields, our method constructs a single and constant ePSF for a given field and epoch. A constant ePSF is one that does not vary with either $X$ or $Y$. Figures \ref{fig:rf-scl1} and \ref{fig:rf-scl2} show plots of ${\cal RF}$ \textit{versus} $X$ (panels in the left-hand column) and ${\cal RF}$ \textit{versus} $Y$ (panels in the right-hand column) for the SCL~$J0100-3341$ and SCL~$J0100-3338$ fields, respectively. The rows of panels from top to bottom are each one epoch, arranged in chronological order. The filled squares in a plot correspond to the QSO. Note that the number of ${\cal RF}$ values for a given object may be equal to the number of exposures --- individual images --- at a given epoch, or be less if the object is not measured in one or more exposures. No panel in Figure~\ref{fig:rf-scl1}, except for the top-left one, shows a trend between ${\cal RF}$ and $X$ or ${\cal RF}$ and $Y$. The top-left panel shows that the mean ${\cal RF}$ decreases linearly with $X$, implying that the shape of the true PSF becomes progressively narrower and more peaked than that of the constructed ePSF with increasing $X$. We are unable to trace the origin of this dependence. The values of ${\cal RF}$ for the QSO are larger than those for other objects at both epochs and are all positive, implying that the true PSF for the QSO is wider than the constructed ePSF and than that for a star. No panel in Figure~\ref{fig:rf-scl2} shows a trend as conspicuous as the one in the upper-left panel of Figure~\ref{fig:rf-scl1}. Nevertheless, the left-hand panel in the middle row does show a hint of variability of the true PSF with location. The PSF of the QSO in the SCL~$J0100-3338$ is similar to that of a star. The values of ${\cal RF}$, though still biased towards positive values, are comparable to those for bright stars. There are two reasons why the PSF of a QSO can be different from that of a star. 1. The underlying galaxy can broaden the image of a QSO. 2. The color of a typical QSO is bluer than that of a typical star. So, particularly for the unfiltered STIS imaging, the true PSF is narrower for a bluer object. Thus, depending on the interplay between the distance to a QSO and its color, the values of ${\cal RF}$ for the QSO can average more positive than, more negative than, or the same as those for a bright star. Visual inspection of the QSO in the SCL~$J0100-3341$ field shows what appears to be a single spiral arm or tidal feature extending from its image, suggesting that the underlying galaxy is indeed the cause of the large positive values of ${\cal RF}$ for this QSO. Experience with the data for other dSphs (P02, P03, and P05) has shown that trends in the ${\cal RF}$ values with position do not necessarily produce systematic errors in the positions of objects. The next section searches for such systematic errors in the position. \subsection{Position Residuals} \label{sec:rx-ry} Fitting an ePSF to the science data array of an object (the $5\times5$ array of pixels representing an object; see P02 for more detail on this array and our procedures) determines its centroid. With 24 images per field and epoch, there can be up to 24 measurements of the centroid. The actual number will be smaller than 24 if an object is flagged out from one or more images because its array is corrupted by cosmic rays or hot pixels. The dithering, rotation, and change of scale (e.g., due to ``breathing'' of the \textit{HST}) between any two images cause the centroid of an object measured in these two images to differ. Therefore, at each epoch, every field has a fiducial coordinate system that coincides with the coordinate system of the first image in chronological order. The adopted transformation from the coordinate system of each subsequent image to the fiducial system contains a linear translation, rigid rotation, and a uniform scale change. Let $(X_{0,j}^{i,k}, Y_{0,j}^{i,k})$ be the centroid of object $i$ at epoch $j$ in image $k$ transformed to the fiducial coordinate system and the mean centroid of object $i$ in the fiducial coordinate system of epoch $j$ be $(<X_{0,j}>^{i},<Y_{0,j}>^{i})$. Define position residuals, ${\cal RX}_{j}^{i,k}$ and ${\cal RY}_{j}^{i,k}$, for an object $i$ as ${\cal RX}_{j}^{i,k}=<X_{0,j}>^{i} - X_{0,j}^{i,k}$ and ${\cal RY}_{j}^{i,k}=<Y_{0,j}>^{i}- Y_{0,j}^{i,k}$. Ideally, ${\cal RX}_{j}^{i,k} = {\cal RY}_{j}^{i,k} = 0$ for all $j$ and $k$. Random noise causes ${\cal RX}_{j}^{i,k}$ and ${\cal RY}_{j}^{i,k}$ to differ from zero, but it does not cause any trends with respect to other quantities. However, systematic errors can cause such trends. Anderson \& King (2000) demonstrates that a mismatch between the true PSF and the ePSF causes ${\cal RX}_{j}^{i,k}$ and ${\cal RY}_{j}^{i,k}$ to depend on the location of a centroid within a pixel --- the pixel phase $\Phi_{x}$ or $\Phi_{y}$. By definition, $\Phi_{x,j}^{i,k} \equiv X_{0,j}^{i,k}-Int(X_{0,j}^{i,k})$ and $\Phi_{y,j}^{i,k} \equiv Y_{0,j}^{i,k}-Int(Y_{0,j}^{i,k})$, where the function $Int(x)$ returns the integer part of the variable $x$. Figure~\ref{fig:rxry-scl1} plots ${\cal RX}$ and ${\cal RY}$ \textit{versus} $\Phi_{x}$ or $\Phi_{y}$ for the SCL~$J0100-3341$ field. The plots in the panel \ref{rx-ry-9-scl1} are for the 2000 epoch and those in the panel \ref{rx-ry-10-scl1} are for the 2002 epoch. The filled squares correspond to the QSO and the dots to stars with a $S/N$ greater than 30. The plots of ${\cal RX}$ \textit{versus} $\Phi_{x}$ and ${\cal RY}$ \textit{versus} $\Phi_{y}$ in Figures~\ref{rx-ry-9-scl1} and \ref{rx-ry-10-scl1} show trends between these quantities for the QSO. Values of ${\cal RX}$ and ${\cal RY}$ tend to be negative for $\Phi_{x}$ and $\Phi_{y}$ less than about 0.5~pixel, and they tend to be positive for $\Phi_{x}$ and $\Phi_{y}$ greater than about 0.5~pixel. The points corresponding to the stars do not show these trends. The plots of the cross terms, ${\cal RX}$ \textit{versus} $\Phi_{y}$ and ${\cal RY}$ \textit{versus} $\Phi_{x}$, do not show any trends for the QSO or for the stars. These trends indicate a mismatch between the ePSF and the true PSF (Anderson \& King 2000). Both stars with $S/N > 15$ and the QSO contribute to the construction of the ePSF. Therefore, the more-extended true PSF of the QSO causes the ePSF to be wider than an ePSF constructed using only stars; in other words, the ePSF is a ``compromise'' between that of the stars and that of the QSO. An ePSF constructed using objects with $S/N > 100$ diminishes the trends in the values of ${\cal RX}$ and ${\cal RY}$ for the QSO because the shape of the ePSF is more akin to the shape of the true PSF of the QSO. However, increasing the $S/N$ threshold to 100 or more in the construction of the ePSF is undesirable because the resulting ePSF is poorly sampled because there are only a few stars with $S/N$ greater than this limit. Instead, we choose to allow the errors in the position of the QSO to remain and be reflected in a greater uncertainty for the measured proper motion for this field. Figure \ref{fig:rxry-scl2} plots ${\cal RX}$ and ${\cal RY}$ \textit{versus} $\Phi_{x}$ or $\Phi_{y}$ for the SCL~$J0100-3338$ field. Figures~\ref{rx-ry-8-scl2}, \ref{rx-ry-9-scl2}, and \ref{rx-ry-10-scl2} are for the 1999, 2000, and 2002 epochs, respectively. Only objects with a $S/N$ greater than 15 are shown. No plot shows clear evidence for trends between ${\cal RX}$ or ${\cal RY}$ and $\Phi_{x}$ or $\Phi_{y}$ for the QSO or for the stars. In this field, the true PSF of the QSO resembles that for a star, which is confirmed by Figure~\ref{fig:rf-scl2}, where the values of $\cal{RF}$ for the QSO are indistinguishable from those for stars. \section{Proper Motion of Sculptor} \label{sec:pm} At this point, there are two lists of fiducial coordinates, one for each epoch, for the SCL~$J0100-3341$ field, and three for the SCL~$J0100-3338$ field. Define the standard coordinate system to be that which moves uniformly together with the stars of Sculptor. Thus, transforming the fiducial coordinates of a star of Sculptor from different epochs into the standard coordinate system produces the same value within the measurement uncertainties. In contrast, the transformed coordinates of the QSO or any other object that is not a member of Sculptor will show uniform motion. The proper motion of Sculptor derives from the motion of the QSO in the standard coordinate system. P05 describes a procedure for deriving the motion of the QSO, and any other object that is not a member of the dSph, in the standard coordinate system from lists of fiducial coordinates at three epochs. The procedure includes a linear motion in the fitted transformations between the fiducial coordinate systems and the standard coordinate system for those objects whose $\chi^2$ calculated with zero motion is above a threshold. The SCL~$J0100-3341$ field has only two epochs, so we have modified the procedure for this case by excluding those objects with $\chi^2$ values above a threshold from the calculation of the transformations between the coordinate systems. The motion of the QSO is just the difference of the two transformed coordinates. The following two sections describe the results from applying these procedures to the two fields. \subsection{Motion of the QSO in the SCL~$J0100-3341$ field} \label{sec:pm1} The number of objects with a measured centroid is 567 and 516 in epochs 2000 and 2002, respectively. Among these, 470 are common to the two epochs. The choice for the individual $\chi^2$ that triggers fitting for uniform linear motion is 15. The multiplicative constant that ensures a $\chi^2$ of one per degree of freedom is 1.151 (see P05 for a discussion of these parameters). The transformation of the measured centroids to the standard coordinate system used in this article is \begin{eqnarray} x_{j}^{\prime\, i} &=& x_{off} + c_{1} + c_{2}(x_{j}^{i} - x_{off}) + c_{3}(y_{j}^{i} - y_{off}) \label{eq:tranx} \\ y_{j}^{\prime\, i} &=& y_{off} + c_{4} + c_{5}(x_{j}^{i} - x_{off}) + c_{6}(y_{j}^{i} - y_{off}) \label{eq:trany} \\ \sigma_{xj}^{\prime\, i}&=& \sqrt{(c_{2}\sigma_{xj}^{i})^{2} + (c_{3}\sigma_{yj}^{i})^{2}} \\ \sigma_{yj}^{\prime\, i}&=& \sqrt{(c_{5}\sigma_{xj}^{i}) + (c_{6}\sigma_{yj}^{i})^{2}}. \label{eq:tranuy} \end{eqnarray} The above represents a modification of the method described in P05, afforded here because of the greater number of stars. In the equations, $c_{1}$ through $c_{6}$ are the free parameters, $(x_{off}, y_{off}) =(512, 512)$~pixel defines the reference point for the transformation, and $(x^{i}_{j},y^{i}_{j})$ is a measured centroid of the $i$th object at the $j$th epoch which is transformed to $(x^{\prime i}_{j},y^{\prime i}_{j})$ in the standard coordinate system. Equations 10 and 11 in P05 define position residuals $RX_{j-1}^{i}$ and $RY_{j-1}^{i}$ for an object $i$ transformed to the standard coordinate system from the fiducial coordinate system of the $j$th epoch. For an ideal case, $RX_{j-1}^{i} = RY_{j-1}^{i} =0$. Figure~\ref{fig:rx-ry-scl1} shows $RX$ \textit{versus} $X$ and $RY$ \textit{versus} $Y$ for the SCL~$J0100-3341$ field. The most prominent feature is a ``step'' in $RX_{1-1}$ \textit{versus} $X$ at $X \simeq 320$~pixel. The values of $RX_{1-1}$ tend to be negative for $X$ below the step, indicating the presence of a systematic error in the $X$ coordinates whose source we are unable to trace. The values of $RX_{2-1}$ tend to be positive for $X \lesssim 320$~pixel, which is forced by the fitting procedure. An \textit{ad hoc} approach for removing the ``steps'' is to replace $x_{j}^{i}$ with $x_{j}^{i} + c_{7}$ in the Equations~\ref{eq:tranx} through \ref{eq:tranuy} when $x_{j}^{i} \leq 320$~pixels and to fit for the additional free parameter $c_{7}$. Applying this remedy removes the ``steps,'' as is shown by Figure~\ref{fig:rx-ry-scl1-c} which plots the same quantities as Figure~\ref{fig:rx-ry-scl1}. In this corrected fitting procedure, the value of the multiplicative constant that ensures $\chi^2$ of one per degree of freedom decreased to 1.123 because of the smaller residuals. The fitted value of $c_7$ is 0.019~pixel. The proper motion for this field derives from this fit. Figure~\ref{fig:rx-ry-scl1-cb} is the same as Figure~\ref{fig:rx-ry-scl1-c} except that the points are the weighted mean residuals in ten equal-length bins in $X$ or $Y$. Note the different vertical scale. The points are plotted at the mean of the coordinate values in the bin. The average residuals show no systematic trends above a level of 0.001~pixel. Figure~\ref{fig:xq-yq-scl1} shows the location of the QSO as a function of time in the standard coordinate system. The top panel shows the variation of the $X$ coordinate and the bottom panel does the same for the $Y$ coordinate. The motion of the QSO is $(\mu_{x},\mu_{y})=(0.0032 \pm 0.0032, 0.0005 \pm 0.0035)$~pixel~yr$^{-1}$. The contribution to the total $\chi^2$ from the QSO, and from any other object whose motion was fit for, is 0 because a line always passes exactly through two points. \subsection{Motion of the QSO in the SCL~$J0100-3338$ field} \label{sec:pm2} The number of objects with measured centroids is 343, 326, and 314 in epochs 1999, 2000, and 2002, respectively. Among these, 257 are common to the three epochs. The choice for the individual $\chi^2$ that triggers fitting for uniform linear motion is 15. The multiplicative constant that ensures a $\chi^2$ of one per degree of freedom is 1.176. Figures~\ref{fig:rx-ry-scl2} and \ref{fig:rx-ry-scl2-b} show position residuals, $RX$ and $RY$, as a function of position in the standard coordinate system for the SCL~$J0100-3338$ field. They are analogous to Figures~\ref{fig:rx-ry-scl1} and \ref{fig:rx-ry-scl1-cb}. From top to bottom, the rows of panels are for epochs 1999, 2000, and 2002. No panel shows unambiguous trends between $RX$ and $X$ or $RY$ and $Y$. The largest deviations of the average residuals are $RY \simeq 0.004$~pixel for $Y < 100$~pixel. Any systematic trends at the location of the QSO are on the order of 0.001~pixel. Although not shown in the figures, the plots of the cross-terms do not show trends either. Figure~\ref{fig:xq-yq-scl2} is analogous to Figure~\ref{fig:xq-yq-scl1} for the SCL~$J0100-3338$ field. Note that the slopes in the corresponding plots in Figures~\ref{fig:xq-yq-scl2}\ and \ref{fig:xq-yq-scl1}\ need not be the same because the two fields are rotated with respect to each other --- though for the fields in Sculptor the rotation is only a few degrees. The uncertainties shown for the points in Figure~\ref{fig:xq-yq-scl2} are those calculated from the scatter of the measurements about the mean for an individual epoch increased by a multiplicative factor. The introduction of this factor reduces the contribution to the total $\chi^2$ from the QSO. Without it, the contribution was 9.52. The contribution to the $\chi^2$ has approximately two degrees of freedom, which implies a 0.9\% probability of a $\chi^2$ larger than 9.52 by chance. Such a small probability likely indicates the presence of unaccounted-for systematic errors. We choose to increase the uncertainty in our fitted proper motion by multiplying the uncertainties of the mean positions at each epoch by the same numerical factor so that contribution to the total $\chi^2$ is about one per degree of freedom. Our fitting procedure calculates a value for the factor for all objects whose contribution to the total $\chi^2$ exceeds 4.6, which is expected 10\% of the time by chance. The value of the factor is 2.2 for the QSO and the uncertainty in the fitted motion of the QSO increases by essentially the same amount. The motion of the QSO is $(\mu_{x},\mu_{y})=(-0.0043\pm 0.0050, 0.0034\pm 0.0038)$~pixel~yr$^{-1}$. \subsection{Measured Proper Motion} \label{sec:pmm} Table~1 gives the measured proper motion for each field in the equatorial coordinate system and their weighted mean. Table~2 tabulates the proper motions for those objects in the SCL~$J0100-3341$ field for which it was measured. Table~3 does the same for the SCL~$J0100-3338$ field. The first line of Table~2 and Table~3 corresponds to the QSO and subsequent objects are listed in order of decreasing $S/N$. The ID number of an object is in column~1, the $X$ and $Y$ coordinates of an object in the earliest image of the first epoch (o65q09010 for SCL~$J0100-3341$ and o5bl02010 for SCL~$J0100-3338$) are in columns 2 and 3, and the $S/N$ of the object at the first epoch is in column 4. The components of the measured proper motion, expressed in the equatorial coordinate system, are in columns 5 and 6. Each value is the measured proper motion in the standard coordinate system corrected by adding the weighted mean proper motion of Sculptor given in the bottom line of Table~1. To indicate that this correction has been made, the proper motion of the QSO is given as zero. The listed uncertainty of each proper motion is the uncertainty of the measured proper motion, calculated in the same way as for the QSO, added in quadrature to that of the average proper motion of the dSph. The contribution of the object to the total $\chi^2$ is in column~7. Although column 7 is in Table 2 for the sake of symmetry with Table 3, the $\chi^2$ contributions are not meaningful. Schweitzer \etal\ (1995) reports the first measurement of the proper motion for Sculptor; this study reports an additional two independent measurements. Figure~\ref{fig:pm} compares the three independent measurements, each represented by a rectangle. A dot at the center of a rectangle is the best estimate of the proper motion. The sides of a rectangle are offset from the center by the 1-$\sigma$ uncertainties. Rectangles 1, 2, and 3 represent the measurements by Schweitzer \etal (1995), this study (field SCL~$J0100-3341$), and this study (field SCL~$J0100-3338$), respectively. The $\alpha$ components of our measurements 2 and 3 agree almost exactly and their $\delta$ components differ by only 1.4$\times$ the uncertainty of their difference. While the $\delta$ component of measurement 1 agrees with the $\delta$ components of measurements 2 and 3, the $\alpha$ component does not agree with either one. The $\alpha$ components of measurements 1 and 2 differ by 2.3$\times$ the uncertainty of their difference and those for measurements 1 and 3 differ by 2.2$\times$. Because of the large difference in the $\alpha$ components of the proper motion between the measurement from Schweitzer \etal\ (1995) and from our two fields, we choose to use the weighted average proper motion from Table~1 to determine the space velocity of Sculptor. \subsection{Galactic Rest Frame Proper Motion} \label{sec:pmgrf} The measured proper motion of the dSph contains contributions from the motion of the LSR and the peculiar motion of the Sun. The magnitude of the contributions depend on the Galactic longitude and latitude of the dSph. Removing them yields the Galactic-rest-frame proper motion --- the proper motion measured by a hypothetical observer at the location of the Sun but at rest with respect to the Galactic center. Columns (2) and (3) of Table~4 give the equatorial components, $(\mu_{\alpha}^{\mbox{\tiny{Grf}}}, \mu_{\delta}^{\mbox{\tiny{Grf}}})$, of the Galactic-rest-frame proper motion. Their derivation assumes: 220~km~s$^{-1}$ for the circular velocity of the LSR; 8.5~kpc for the distance of the Sun from the Galactic center; and $(u_\odot, v_\odot, w_\odot) = (-10.00 \pm 0.36, 5.25 \pm 0.62 , 7.17 \pm 0.38)$~km~s$^{-1}$ (Dehnen \& Binney 1998) for the peculiar velocity of the Sun, where the components are positive if $u_{\odot}$ points radially away from the Galactic center, $v_{\odot}$ is in the direction of rotation of the Galactic disk, and $w_\odot$ points in the direction of the North Galactic Pole. Columns (4) and (5) give the Galactic-rest-frame proper motion in the Galactic coordinate system, $(\mu_{l}^{\mbox{\tiny{Grf}}},\mu_{b}^{\mbox{\tiny{Grf}}})$. The next three columns give the $\Pi$, $\Theta$, and $Z$ components of the space velocity in a cylindrical coordinate system centered on the dSph. The components are positive if $\Pi$ points radially away from the Galactic axis of rotation, $\Theta$ points in the direction of rotation of the Galactic disk, and $Z$ points in the direction of the North Galactic Pole. The derivation of these components assumes 87~kpc (Kaluzny \etal\ 1995) for the heliocentric distance to and $109.9 \pm 1.4$~km~s$^{-1}$ (Queloz, Dubath, \& Pasquini 1995) for the heliocentric radial velocity of Sculptor. The last two columns give the radial and tangential components of space velocity for an observer at rest at the Galactic center. The component $V_{r}$ is positive if it points radially away from the Galactic center. Thus, at present, Sculptor is moving away from the Milky Way. \section{Orbit and Orbital Elements of Sculptor} \label{sec:orbit} Knowing the space velocity of a dSph permits a determination of its orbit for a given form of the Galactic potential. This study adopts a Galactic potential that has a contribution from a disk of the form (Miyamoto \& Nagai 1975) \begin{equation} \label{diskpot} \Psi_{\mbox{\small{disk}}}=-\frac{G M_{\mbox{\small{disk}}}}{\sqrt{R^{2}+(a+\sqrt{Z^{2}+b^{2}})^{2}}}, \end{equation} from a spheroid of the form (Hernquist 1990) \begin{equation} \label{spherpot} \Psi_{\mbox{\small{spher}}}=-\frac{GM_{\mbox{\small{spher}}}} {R_{\mbox{\small{GC}}}+c}, \end{equation} and from a halo of the form \begin{equation} \label{logpot} \Psi_{\mbox{\small{halo}}}=v^{2}_{\mbox{\small{halo}}}\ln (R^{2}_{\mbox{\small{GC}}}+d^{2}). \end{equation} In the above equations, $R_{\mbox{\small GC}}$ is the Galactocentric distance, $R$ is the projection of $R_{\mbox{\small GC}}$ onto the plane of the Galactic disk, and $Z$ is the distance from the plane of the disk. All other quantities in the equations are adjustable parameters and their values are the same as those adopted by Johnston, Sigurdsson, \& Hernquist (1999): $M_{\mbox{disk}}=1.0\times10^{11}$~M$_{\odot}$, $M_{\mbox{spher}}=3.4\times10^{10}$~M$_{\odot}$, $v_{\mbox{halo}}=128$~km~s$^{-1}$, $a=6.5$~kpc, $b=0.26$~kpc, $c=0.7$~kpc, and $d=12.0$~kpc. Figure~\ref{fig:orbit} shows the projections of the orbit of Sculptor onto the $X-Y$ (top-left panel), $X-Z$ (bottom-left panel), and $Y-Z$ (bottom-right panel) Cartesian planes. The orbit results from an integration of the motion in the Galactic potential given by Equations~\ref{diskpot}, \ref{spherpot}, and \ref{logpot}. The integration extends for 3~Gyr backwards in time and begins at the current location of Sculptor with the negative of the space velocity components given in the bottom line of columns (6), (7), and (8) of Table~4. The filled square marks the current location of the dSph, the filled star indicates the center of the Galaxy, and the two small circles mark the points on the orbit where $Z=0$ or, in other words, where the orbit crosses the plane of the Galactic disk. The large circle is for reference: it has a radius of 30~kpc. In the right-handed coordinate system of Figure~\ref{fig:orbit}, the current location of the Sun is on the positive $X$-axis. The figure shows that Sculptor is moving away from the Milky Way, is closer to perigalacticon than apogalacticon, and that it has a nearly polar orbit with a modest eccentricity. Table~5 tabulates the elements of the orbit of Sculptor. The value of the quantity is in column (4) and its $95\%$ confidence interval is in column (5). The latter comes from 1000 Monte Carlo experiments, where an experiment integrates the orbit using an initial velocity that is generated by randomly choosing the line-of-sight velocity and the two components of the measured proper motion from Gaussian distributions whose mean and standard deviation are the best estimate of the quantity and its quoted uncertainty, respectively. The eccentricity of the orbit is defined as \begin{equation} \label{eccentricity} e = \frac{(R_{a} - R_{p})}{(R_{a} + R_{p})}. \end{equation} The most likely orbit has about a 2:1 ratio of apogalacticon to perigalacticon, though the 95\% confidence interval for the eccentricity allows ratios approximately between 1.7:1 and 4:1. The orbital period of Sculptor, 2.2~Gyr, is about 50\% longer than those of Carina (1.4~Gyr; P03) and Ursa~Minor (1.5~Gyr; P05). \section{Discussion} \label{sec:disc} Knowing the orbit can help answer several questions about Sculptor, or, at least, increase the level of our understanding of this galaxy. These questions are: 1. Is Sculptor a member of a stream of galaxies? 2. Is its star formation history correlated with the orbit? 3. What is the origin of the HI clouds detected in close proximity to the dSph? 4. Does Sculptor contain dark matter? \subsection{Is Sculptor a Member of a Stream?} \label{stream} Lynden-Bell \& Lynden-Bell (1995) proposes that Sculptor may be a member of one of three possible streams: stream No.~2 (together with the LMC, SMC, Draco, Ursa Minor, and Carina); No.~4a (together with Sextans, Pal 3, and Fornax); or No.~4b (together with Sextans and Fornax). Columns (2) and (3) in Table~6 give the predicted heliocentric (i.e., ``measured'' in our terminology) proper motion in the equatorial coordinate system for Sculptor if it indeed belongs to any of the three streams. The magnitude of the proper motion vector, $\vert \vec{\mu} \vert = \sqrt {\mu_{\alpha}^{2} + \mu_{\delta}^{2}}$, and its position angle are in columns (4) and (5). For easy comparison, the corresponding quantities from our study are in the bottom line of the table. Comparing the entries shows that the predictions for streams 2 and 4a disagree significantly with our measurement. However, the prediction for stream No. 4b is closer: the magnitudes differ by 1.6$\times$ the uncertainty in their difference, while the position angles differ by 1.6$\times$. Differences of this size should occur by chance 1\% of the time. The measured proper motion based on only the three-epoch data in the SCL~$J0100-3338$ field improves the agreement with the prediction for stream 4b. Thus, while we rule out the possibility that Sculptor is a member of stream No.~2 or 4a, its membership in stream~4b is possible. Stream 4b contains both Sculptor and Fornax. The Dinescu \etal\ (2004) proper motion for Fornax is $(\mu_\alpha, \mu_\delta) = (59 \pm 16, -15 \pm 16)$~mas~cent$^{-1}$. The magnitude and position angle of the proper motion are $61 \pm 16$~mas~cent$^{-1}$ and $104\pm 15$~degrees. The prediction for stream 4b from Lynden-Bell \& Lynden-Bell (1995) is 20~mas~cent$^{-1}$ and 162~degrees. The difference between the measured and predicted proper motions would be this large or larger by chance only 0.4\% of the time. Thus, the physical reality of stream~4b is doubtful. Dinescu \etal\ (2004) argues that Fornax and Sculptor are members of the same stream that also includes Leo I, Leo II, and Sextans. Together the galaxies define the FL$^{2}$S$^{2}$ plane. If they do form a stream, their Galactic-rest-frame proper motion vectors should be aligned with the great circle passing through the galaxies. The position angle of the great circle passing through Sculptor and Fornax is about 99~degrees at the location of Sculptor and 95~degrees at Fornax. The position angle of the Galactic-rest-frame proper motion for Fornax reported by Dinescu \etal\ (2004) is $79 \pm 25$~degrees, which differs by 0.64$\times$ its uncertainty from the position angle of the great circle. If Sculptor and Fornax form a stream, then they should move in the same direction along the great circle connecting them. Thus, the proper motion of Fornax from Dinescu \etal\ (2000) implies that the position angle of the Galactic-rest-frame proper motion of Sculptor should be 99~degrees. The position angle for the proper motion of Sculptor reported here is $333 \pm 15$~degrees, which differs from the prediction by 8.4$\times$ its uncertainty. Discounting the proper motion from this study and instead using the position angle of $40 \pm 24$~degrees implied by the proper motion measured by Schweitzer \etal\ (1995) does not remove the disagreement: the position angle differs by $2.5\times$ its uncertainty from that of the great circle. We conclude that Sculptor and Fornax do not belong to the same stream. Kroupa \etal\ (2004) shows that the 11 dwarf galaxies nearest to the Milky Way are nearly on a plane, whose two poles are at $(\ell,b) = (168,-16)$~degrees and $(348,+16)$~degrees. Adopting the direction of the angular momentum vector as the pole of the orbit, then the location of the pole is \begin{equation} (\ell,b) = (\Omega+90^{\circ},\Phi-90^{\circ}). \end{equation} Because of the left-handed nature of the Galactic rotation, prograde orbits have $b < 0$ and retrograde orbits have $b > 0$. Thus, the pole of our orbit for Sculptor is $(\ell,b) = (5 \pm 16, -4 \pm 1.6)$~degrees, where the uncertainties are 1-$\sigma$ values from the Monte Carlo simulations. The galactic longitudes of the poles of the plane and orbit agree within the uncertainty, but the galactic latitudes do not. They differ by 20~degrees, which is more than 12$\times$ the uncertainty in the location of the pole of the orbit. However, there is also some uncertainty in the orientation of the plane passing through the dwarf galaxies near the Milky Way. We conclude the plane of the orbit of Sculptor is similar to the plane defined by the nearby dwarf galaxies. \subsection{The Effect of the Galactic Tidal Force on the Structure of Sculptor} \label{sec:tides} The measured ellipticity of the isodensity contours increases with projected distance from the center of Sculptor (see Figure~1 in Irwin \& Hatzidimitriou 1995), akin to surface density contours of a model dSph in the numerical simulations of Johnston, Spergel, \& Hernquist (1995; e.g., see Figure~4). If the Galactic tidal force deformed Sculptor from an initial nearly-spherical shape to its present elongated shape in the outer regions then, from our vantage point nearly in the orbit plane, the position angle of its projected major axis should be similar to --- or differ by 180 degrees from --- the position angle of the Galactic-rest-frame proper motion vector, as predicted by the numerical simulations of Oh, Lin, \& Aarseth (1995), Piatek \& Pryor (1995), or Johnston, Spergel, \& Hernquist (1995). The position angle of the projected major axis is $99 \pm 1$~degrees and the position angle of our measured Galactic-rest-frame proper motion vector is $333 \pm 15$~degrees. Allowing for the 180-degree degeneracy, the difference between the two position angles is 3.6$\times$ the uncertainty of their difference, which suggests that the Galactic tidal force has not elongated Sculptor. \subsection{Does Star Formation History Correlate with the Orbital Motion of Sculptor?} \label{sec:sfh} Da Costa (1984) imaged a 3~arcmin $\times$ 5~arcmin field located just outside of the core radius of Sculptor in three bands: $B$, $V$, and $R$. The photometry reaches the main-sequence turn off. Comparing theoretical isochrones with the distribution of stars in the color-magnitude diagram, the study finds that the majority of stars is about 2-3 Gyr younger than the galactic globular clusters of comparable metal abundance. An earlier study by Kunkel \& Demers (1977) based on $B$ and $V$ photometry extending to 0.4 magnitudes below the horizontal branch reaches a similar conclusion. The color-magnitude diagram also shows a population of ``blue stragglers,'' which might be indicative of an extended period of star formation. Da Costa (1984) concludes, however, that, if an intermediate-age stellar population exists in Sculptor, it is ``infinitesimal compared to that of the Carina system.'' Deep \textit{HST} imaging by Monkiewicz \etal\ (1999) in a single field, reaching 3 magnitudes below the main-sequence turn-off, confirms the basic picture of Sculptor uncovered by Kunkel \& Demers and Da Costa (1984). This color-magnitude diagram also reveals the presence of ``blue stragglers'' and implies an age comparable to that of the galactic globular clusters. The small number of stars in the small field made a search for an intermediate-age stellar population inconclusive. Majewski \etal\ (1999), Hurley-Keller \etal\ (1999), and Harbeck \etal\ (2001) use wide-field imaging to show the presence of two stellar populations with distinctly different metallicities ([Fe/H] = --2.3 and --1.5; Majewski \etal\ 1999). The more metal rich population is more centrally concentrated in the galaxy. Most recently, Tolstoy \etal\ (2004) confirms the above picture using wide-field imaging and spectroscopy. Spectroscopically-determined metallicities range from --2.8 to --0.9. Stars more metal-rich than --1.7 are more centrally concentrated and have a smaller velocity dispersion than the rest of the sample. However, both stellar populations are older than 10~Gyrs. The aforementioned studies show that there were at least two episodes of star formation at times more than 10~Gyr ago. Because 10~Gyrs is much longer than the orbital period of approximately 2.2~Gyr, there is no clear connection between the stimulation of star formation and processes such as the Galaxy-Sculptor tidal interaction or the effects of ram pressure. The lack of correlation could be due to the loss of all of the gas in Sculptor about 10~Gyr ago. But, surprisingly, the observations indicate that Sculptor has HI today. \subsection{HI Gas in Sculptor} \label{sec:gas} Unlike most other Galactic dSphs, Sculptor contains a detectable amount of HI. Knapp \etal\ (1978) detects three clouds of HI in the vicinity of Sculptor and speculates that one of them, with a radial velocity of 120~km~s$^{-1}$, may be associated with Sculptor, whose radial velocity at the time was uncertain. Carignan \etal\ (1998) confirms and refines this detection and puts a lower limit on the mass of HI of $3.0 \times 10^{4}$~M$_{\odot}$. Bouchard \etal\ (2003) repeats the observations over a wider field with the Parkes single-dish telescope and, over a smaller region, at higher angular resolution with the Australia Telescope Compact Array. The better data show that the HI is not associated with a background galaxy and the probability of a chance superposition of a galactic high velocity cloud is less than 2\%. These arguments, together with the agreement within 4~km~s$^{-1}$ of the radial velocity of the HI and the radial velocity of Sculptor make a strong case for the physical association of the clouds with the dSph. The gas is in two clouds. Figure~\ref{fig:gas}\ shows the distribution of HI on the sky in the direction of Sculptor based on the Australia Telescope Compact Array data from Bouchard \etal\ (2003). The asterisk marks the optical center of Sculptor and the two clouds are about 20--30~arcmin from the center, one to the northeast and one to the southwest. The masses of the clouds are $(4.1 \pm 0.2) \times 10^4$~M$_\odot$ and $(1.93 \pm 0.02) \times 10^5$~M$_\odot$, respectively. The two clouds lie nearly along the minor axis of the dSph; the orientation of Sculptor is shown in Figure~\ref{fig:gas}\ by the ellipse representing the optical boundary. Although the association between the gas and the dSph seems well-established, why Sculptor still has gas when most of the other galactic dSphs do not and the cause of the observed configuration of the gas are debated. Mayer \etal\ (2005) studies the loss of gas from dwarf galaxies in the Local Group using numerical simulations that include ram pressure stripping. It quantifies the expected result that a galaxy with a deeper potential well or with a larger perigalacticon is more likely to retain its gas. The orbit of Sculptor is similar to those of Carina and Ursa Minor, which suggests that differences in the degree of tidal shocking or ram pressure stripping are not the reason for the difference in gas retention. However, note that the 95\% confidence intervals for the perigalacticons of all three orbits are still too large to make a conclusive statement. The larger mass of Sculptor compared to Carina and Ursa Minor is the most likely reason why it was able to retain gas. Mechanisms that could affect the distribution of HI within the dSph are tidal interaction, ram pressure, and forces from supernovae or winds from young stars. Tidal interaction and ram pressure tend to spread the gas in the plane of the orbit. The two arrows in Figure~\ref{fig:gas}\ represent the Galactic-rest-frame proper motions as measured by Schweitzer \etal\ (1995; dashed) and this study (solid). The ratio of their lengths is the same as the ratio of the magnitudes of the proper motions. The dotted line is a section of a great circle that passes through Sculptor and Fornax; its position angle is 99~degrees. The Schweitzer \etal\ (1995) proper motion (dashed arrow) is nearly aligned with the line that connects the centers of the clouds. Carignan \etal\ (1998) notes this alignment and suggests that it may indicate a ``tidal'' origin for the clouds: presumably, the Galactic tidal force stretches the gas, initially centered within the dSph, into its observed distribution. The most serious problem with such a picture is that the tidal force would stretch the distribution of both the stars and the gas, which then aligns the major axes of the gaseous and stellar distributions from our perspective as an observer nearly in the plane of the orbit of Sculptor. They are not aligned, perhaps because the motion of the gas is governed by both gravitational forces and pressure gradients. As was noted in Section~\ref{sec:tides}, if the observed ellipticity of Sculptor is due to the tidal force, then the Galactic-rest-frame proper motion should be aligned with the major axis. Figure~\ref{fig:gas}\ shows that the solid arrow, our measurement, is closer to such an alignment than the dashed arrow. Because the tidal force is zero at the center of a dSph, it cannot by itself separate a single cloud centered on the dSph into two clouds. Thus, within the context of a picture in which tides have had an important effect on Sculptor, our proper motion is more plausible than that of Schweitzer \etal\ (1995). However, is our proper motion consistent with the geometry of the two clouds? We think yes. First, the two clouds are elongated in the direction of our proper motion (see particularly Figure~1 in Bouchard \etal\ 2003). The observed elongation could be due to ram pressure from the motion through a gaseous Galactic halo. Second and more speculatively, the two clouds could be due to the Rayleigh-Taylor instability when the HI in Sculptor moves through the hot and low-density gaseous halo. Or the gas could have been squeezed out perpendicular to the direction of motion by the compressive tidal shock when Sculptor crosses the Galactic disk. Expansion combined with infall of the gas forms a ring that looks like two clouds in projection. \subsection{Is there dark matter in Sculptor?} \label{sec:dm} Estimates of the $M/L_V$ for Sculptor range from $6 - 13$ and estimates of the limiting radius range from about $40 - 80$~arcmin. The estimates of $M/L$ assume that mass follows light. If this is true, then the implied mass of Sculptor must be large enough, given our orbit, to produce a tidal radius that is at least as large as the observed limiting radius. Equating the tidal radius and the limiting radius predicts a value for $M/L$, which should agree with the measured value. Also, the dSph must have a mass and, hence, $M/L$ large enough for it to have survived destruction by the Galactic tidal force on our orbit. In lieu of numerical simulations, an approximate analytical approach is to calculate the tidal radius, $r_t$, beyond which a star becomes unbound from the dSph. For a logarithmic Galactic potential, $r_t$ is given by (King 1962; Oh, Lin, \& Aarseth 1992) \begin {equation} r_t = \left(\frac{(1-e)^2}{[(1+e)^2/2e]\ln[(1+e)/(1-e)] +1} \, \frac{M}{M_G}\right)^{1/3} a. \label{eq:rtidal} \end {equation} Here $e$ is eccentricity of the orbit, $a$ is the semi-major axis ($a \equiv (R_{a}+R_{p})/2$), $M$ is the mass of the dSph, and $M_G$ is the mass of the Galaxy within $a$. Equating $r_t$ with the observed limiting radius derived by fitting a King (1966) model, $r_k$, gives a value for $M/L_V$ for a given orbit. If $r_k = 40$~arcmin, then 28\% of the orbits in Monte Carlo simulations have $M/L_V > 6$ and 10\% of the orbits have $M/L_{V} > 13$. If $r_k = 80$~arcmin, then 100\% of the orbits have $M/L_V > 13$. These results show that the global $M/L$ of Sculptor is probably larger than the measured $M/L$, if the larger of the measured limiting radii is identified with the tidal radius. The $M/L$ calculated assuming that mass follows light underestimates the true global $M/L$ if Sculptor contains dark matter that is more spatially extended than the luminous matter (e.g., Pryor \& Kormendy 1990). However, equation~\ref{eq:rtidal}\ shows that $M \propto r_t^3$, so the values for $M/L$ derived using this equation are sensitive to the measured value of the limiting radius. Until kinematic measurements definitively identify the tidal radius, an $M/L$ derived with the above argument should be treated with caution. The average of the measured values of $M/L_V$ for Galactic globular clusters is 2.3 (Pryor \& Meylan 1993). Could the true $M/L_V$ of Sculptor be similar to this average? Numerical simulations by Oh, Lin, \& Aarseth (1995) and Piatek \& Pryor (1995) show that the ratio of the limiting radius derived by fitting a theoretical King model (King 1966), $r_k$, to the tidal radius defined by Equation~(\ref{eq:rtidal}) is a useful indicator of the importance of the Galactic tidal force on the structure of a dSph. These simulations show that: if $r_{k}/r_t \lesssim 1.0$, the Galactic tidal force has little effect on the structure of the dSph; at $r_k/r_t \approx 2.0$, the effect of the force increases rapidly with increasing $r_k/r_t$; and, for $r_k/r_t \approx 3.0$, the dSph disintegrates in a few orbits. Assuming that $M/L_V = 2.3$ and $r_k = 40$~arcmin, $r_k/r_t > 2.0$ for 6\% of the orbits generated in Monte Carlo simulations. If $r_k = 80$~arcmin, the fraction is 100\%. Thus, it is possible that Sculptor could have survived for a Hubble time on its current orbit if it only contains luminous matter. We conclude that measured orbit of Sculptor does not require it to contain dark matter. \subsection{A Lower Limit for the Mass of the Milky Way} \label{sec:massofg} Sculptor is bound gravitationally to the Milky Way. The Galactocentric space velocity of the dSph imposes a lower limit on the mass of the Milky Way within the present Galactocentric radius of the dSph, $R$. Assuming a spherically symmetric mass distribution and zero for the total energy of the dSph, the lower limit for the mass of the Milky Way is given by \begin{equation} \label{mwmass} M=\frac{R\left (V_{r}^{2} + V_{t}^{2} \right )}{2G}. \end{equation} Setting $R=87$~kpc and using the values from Table~4 for $V_{r}$ and $V_{t}$, $M = (4.6\pm 2.0) \times 10^{11}\ M_{\odot}$. This lower limit is consistent with other recent estimates of the mass of the Milky Way, such as the mass of $5.4^{+0.1}_{-0.4} \times 10^{11}\ M_{\odot}$ within $R=50$~kpc found by Sakamoto, Chiba, \& Beers (2003). The Milky Way potential adopted in Section~\ref{sec:orbit} has a mass of $7.8\times 10^{11}\ M_{\odot}$ out to $R=87$~kpc. \section{Summary} \label{sec:sum} This article presents a measurement of the proper motion of Sculptor using data taken with \textit{HST} and STIS in imaging mode. Using this measurement, it derives the orbit and discusses: membership in proposed streams, tidal interaction with the Milky Way, the relation between the orbit and the star-formation history, the HI gas associated with the dSph, the dark matter content, and a lower limit on the mass of the Milky Way. The list below enumerates our findings. 1. Two independent measurements of the proper motion give a weighted mean value of $(\mu_\alpha,\mu_\delta)=(9\pm 13, 2 \pm 13)$~mas~cent$^{-1}$ in the equatorial coordinate system for a heliocentric observer. 2. Removing the contributions to the measured proper motion from the motions of the Sun and of the LSR gives a Galactic-rest-frame proper motion of $(\mu_{\alpha}^{\mbox{\tiny{Grf}}}, \mu_{\delta}^{\mbox{\tiny{Grf}}})=(-23\pm 13, 45 \pm 13)$~mas~cent$^{-1}$ in the equatorial coordinate system for an observer at the location of the Sun but at rest with respect to the Galactic center. In the Galactic coordinate system this motion is $(\mu_{l}^{\mbox{\tiny{Grf}}},\mu_{b}^{\mbox{\tiny{Grf}}}) = (11 \pm 13, -50 \pm 14)$~mas~cent$^{-1}$. 3. The radial and tangential components of the space velocity are $V_{r}=79 \pm 6$~km~s$^{-1}$ and $V_{t}=198 \pm 50$~km~s$^{-1}$, respectively, as measured by a Galactocentric observer at rest. 4. The best estimate of the orbit shows that Sculptor is approaching its apogalacticon of $R_{a}=122$~kpc on a polar orbit with eccentricity $e=0.29$. The perigalacticon of the orbit is $R_{p}=68$~kpc and the orbital period is $T=2.2$~Gyr. 5. Sculptor is not a member of streams~2 and 4a proposed by Lynden-Bell \& Lynden-Bell (1995). It could be a member of stream~4b, though the proper motion of Fornax measured by Dinescu \etal\ (2004) makes the physical reality of this stream doubtful. 6. The proper motion vectors of Sculptor and Fornax show that they cannot be members of the same stream. 7. The pole of the orbit of Sculptor is 26~degrees from the pole of the plane of the Galactic dSphs noted by Kroupa \etal\ (2004). This difference is much larger than the uncertainty in the pole of the orbit, but is probably within the uncertainty of the definition of the plane of the dSphs. 8. A comparison of the orbit of Sculptor to those of other dSphs does not provide a clear reason for why Sculptor contains HI while the others do not. The origin and distribution of HI remain puzzling. This article proposes that, while the line connecting the two clouds of HI is nearly perpendicular to our Galactic-rest-frame proper motion, some combination of ram pressure, tidal interaction, and Rayleigh-Taylor instability could produce this geometry. \acknowledgments We thank the anonymous referee for comments that helped to improve the presentation of our work. We also thank Sergei Maschenko for pointing out to us that our method for propagating uncertainties from the measured proper motions to the space velocity was incorrect. CP and SP acknowledge the financial support of the Space Telescope Science Institute through the grants HST-GO-07341.03-A and HST-GO-08286.03-A and from the National Science Foundation through the grant AST-0098650. EWO acknowledges support from the Space Telescope Science Institute through the grants HST-GO-07341.01-A and HST-GO-08286.01-A and from the National Science Foundation through the grants AST-9619524 and AST-0098518. MM acknowledges support from the Space Telescope Science Institute through the grants HST-GO-07341.02-A and HST-GO-08286.02-A and from the National Science Foundation through the grant AST-0098661. DM is supported by FONDAP Center for Astrophysics 15010003. \clearpage \setcounter{figure}{0} \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \setcounter{table}{0} \newdimen\digitwidth\setbox0=\hbox{\rm 0}\digitwidth=\wd0 \catcode`@=\active\def@{\kern\digitwidth} \begin{deluxetable}{lrr} \tablecolumns{3} \tablewidth{3.5truein} \tablecaption{Measured Proper Motion of Sculptor} \tablehead{ &\colhead{$\mu_{\alpha}$}&\colhead{$\mu_{\delta}$}\\ \colhead{Field}&\multicolumn{2}{c}{(mas cent$^{-1}$)}\\ \colhead{(1)}&\colhead{(2)}&\colhead{(3)}} \startdata SCL@J$0100-3341$ &$9\pm17$&$14\pm16$ \\ \noalign{\vspace{1pt}} SCL@J$0100-3338$ &$8\pm20$&$-27\pm25$ \\ \hline \noalign{\vspace{1pt}} Weighted mean&$9\pm13$&$2\pm13$\\ \enddata \end{deluxetable} \begin{deluxetable}{ccccccc} \tablecolumns{7} \tablewidth{5.0truein} \tablecaption{Measured Proper Motions For Objects in the SCL~$J0100-3341$ Field} \tablehead{ &X&Y& &$\mu_{\alpha}$&$\mu_{\delta}$ & \\ \colhead{ID}&\colhead{(pixels)}&\colhead{(pixels)}&\colhead{$S/N$}&\colhead{(mas @cent$^{-1}$)}& \colhead{(mas@cent$^{-1}$)} & \colhead{$\chi^2$} \\ \colhead{(1)}&\colhead{(2)}&\colhead{(3)}&\colhead{(4)}&\colhead{(5)}& \colhead{(6)} & \colhead{(7)} } \startdata 1& 459& 352& 125& $ 0 \pm 21 $& $ 0 \pm 21 $& \nodata \\ 2& 327& 740& 164& $ 2766 \pm 17 $& $ 1048 \pm 18 $& \nodata \\ 3& 588& 330& 112& $ -812 \pm 16 $& $-1714 \pm 17 $& \nodata \\ 4& 137& 627& 99& $ 180 \pm 17 $& $ -127 \pm 21 $& \nodata \\ 5& 496& 398& 10& $ -66 \pm 46 $& $ 178 \pm 72 $& \nodata \\ 6& 574& 710& 6& $ 36 \pm 93 $& $ 258 \pm 63 $& \nodata \\ \enddata \end{deluxetable} \begin{deluxetable}{ccccccc} \tablecolumns{7} \tablewidth{5.0truein} \tablecaption{Measured Proper Motions For Objects in the SCL~$J0100-3338$ Field} \tablehead{ &X&Y& &$\mu_{\alpha}$&$\mu_{\delta}$ & \\ \colhead{ID}&\colhead{(pixels)}&\colhead{(pixels)}&\colhead{$S/N$}&\colhead{(mas @cent$^{-1}$)}& \colhead{(mas@cent$^{-1}$)} & \colhead{$\chi^2$} \\ \colhead{(1)}&\colhead{(2)}&\colhead{(3)}&\colhead{(4)}&\colhead{(5)}& \colhead{(6)} & \colhead{(7)} } \startdata 1& 523& 490& 180& $ 0 \pm 24 $& $ 0 \pm 28 $& 2.10 \\ 2& 414& 177& 42& $ -188 \pm 22 $& $ -299 \pm 20 $& 0.61 \\ 3& 950& 832& 20& $ 89 \pm 31 $& $ -562 \pm 27 $& 4.25 \\ \enddata \end{deluxetable} \hoffset=-0.5truein \begin{deluxetable}{lrrrrrrrrr} \tablecolumns{10} \tablewidth{8.8truein} \tablecaption{Galactic-Rest-Frame Proper Motion and Space Velocity of Sculptor} \tablehead{&\colhead{$\mu_{\alpha}^{\mbox{\tiny{Grf}}}$}& \colhead{$\mu_{\delta}^{\mbox{\tiny{Grf}}}$}& \colhead{$\mu_{l}^{\mbox{\tiny{Grf}}}$}&\colhead{$\mu_{b}^{\mbox{\tiny{Grf}}}$}& \colhead{$\Pi$}&\colhead{$\Theta$}&\colhead{$Z$}&\colhead{$V_{r}$}& \colhead{$V_{t}$}\\ \colhead{Field}&\multicolumn{2}{c}{(mas cent$^{-1}$)} &\multicolumn{2}{c}{(mas cent$^{-1}$)}&\colhead{(km s$^{-1}$)}&\colhead{(km s$^{-1}$)}&\colhead{(km s$^{-1}$)} &\colhead{(km s$^{-1}$)}&\colhead{(km s$^{-1}$)}\\ \colhead{(1)}&\colhead{(2)}&\colhead{(3)}&\colhead{(4)} &\colhead{(5)}&\colhead{(6)}&\colhead{(7)}&\colhead{(8)}&\colhead{(9)} &\colhead{(10)}} \startdata SCL@J$0100-3341$&$-23\pm17$&$57\pm16$& $6\pm17$&$-62\pm17$& $-186\pm69$&$159\pm68$&$-107\pm8$& $82\pm7$&$254\pm66$\\ \noalign{\vspace{5pt}} SCL@J$0100-3338$&$-24\pm20$&$17\pm25$& $18\pm21$&$-23\pm25$& $-113\pm88$&$10\pm100$&$-88\pm12$& $73\pm9$&$124\pm76$\\ \noalign{\vspace{3pt}} \hline \noalign{\vspace{5pt}} Weighted mean&$-23\pm13$&$45\pm13$& $11\pm13$&$-50\pm14$& $-158\pm54$&$112\pm56$&$-101\pm7$& $79\pm6$&$198\pm50$\\ \enddata \end{deluxetable} \begin{deluxetable}{lcccc} \tablecolumns{5} \tablewidth{4.75truein} \tablecaption{Orbital elements of Sculptor} \tablehead{Quantity&Symbol&Unit&Value&95\% Conf. Interv.\\ \colhead{(1)}&\colhead{(2)}&\colhead{(3)}&\colhead{(4)}&\colhead{(5)}} \startdata Perigalacticon&$R_{p}$&kpc&$68$&$(31,83)$ \\ \noalign{\vspace{1pt}} Apogalacticon&$R_{a}$&kpc&$122$&$(97,313)$\\ \noalign{\vspace{1pt}} Eccentricity &$e$&&$0.29$&$(0.26,0.60)$\\ \noalign{\vspace{1pt}} Period&$T$&Gyr&$2.2$&$(1.5,4.9)$\\ \noalign{\vspace{1pt}} Inclination&$\Phi$&deg&86&$(83,90)$ \\ \noalign{\vspace{1pt}} Longitude&$\Omega$&deg&275&$(243,306)$ \\ \enddata \end{deluxetable} \begin{deluxetable}{lcccc} \tablecolumns{5} \tablewidth{4.5truein} \tablecaption{Predicted Proper Motion of Sculptor} \tablehead{ & \colhead{$\mu_{\alpha}$}& \colhead{$\mu_{\delta}$}& \colhead{$\vert \vec{\mu}\vert$}& \colhead{PA}\\ \colhead{Stream No.}&\multicolumn{3}{c}{(mas cent$^{-1}$)}&\colhead{(degrees)}\\ \colhead{(1)}&\colhead{(2)}&\colhead{(3)}&\colhead{(4)}&\colhead{(5)}} \startdata 2 & 51 & --80 & 95 & 147 \\ \noalign{\vspace{1pt}} 4a & 80 & --61 & 101 & 127 \\ \noalign{\vspace{1pt}} 4b & -13 & -27 & 30 & 205\\ \noalign{\vspace{1pt}} \hline \noalign{\vspace{1pt}} Our Result&$9\pm13$&$2\pm13$&$9\pm13$&$77\pm81$ \\ \enddata \end{deluxetable}
Title: Exploring the surface properties of Transneptunian Objects and Centaurs with polarimetric FORS1/VLT observations
Abstract: Polarization is a powerful remote-sensing method to investigate solar system bodies. It is an especially sensitive diagnostic tool to reveal physical properties of the bodies whose observational characteristics are governed by small scatterers (dust, regolith surfaces). For these objects, at small phase angles, a negative polarization is observed, i.e., the electric vector E oscillates predominantly in the scattering plane, contrary to what is typical for rather smooth homogeneous surfaces. The behavior of negative polarization with phase angle depends on the size, composition and packing of the scatterers. These characteristics can be unveiled by modelling the light scattering by the dust or regolith in terms of the coherent backscattering mechanism. We have investigated the surface properties of TNOs and Centaurs by means of polarimetric observations with FORS1 of the ESO VLT. TNOs Ixion and Quaoar, and Centaur Chiron show a negative polarization surge. The Centaur Chiron has the deepest polarization minimum (-1.5 - 1.4%). The two TNOs show differing polarization curves: for Ixion, the negative polarization increases rapidly with phase; for Quaoar, the polarization is relatively small (~ -0.6%), and nearly constant at the observed phase angles. For all three objects, modelling results suggest that the surface contains an areal mixture of at least two components with different single-scatterer albedos and photon mean-free paths.
https://export.arxiv.org/pdf/astro-ph/0601414
\title{Exploring the surface properties of Transneptunian Objects and Centaurs with polarimetric FORS1/VLT observations \thanks{Based on observations made with ESO Telescopes at the Paranal Observatory under programme ID 69.C-0133 and 073.C-0561 (PI: H.\ Boehnhardt)}} \author{ S.~Bagnulo \inst{1} \and H.~Boehnhardt \inst{2} \and K.~Muinonen \inst{3} \and L.~Kolokolova \inst{4} \and I.~Belskaya \inst{5} \and M.A.~Barucci \inst{6} } \institute{European Southern Observatory, Alonso de Cordova 3107, Vitacura, Santiago, Chile. \email{sbagnulo@eso.org} \and Max-Planck-Institut f\"{u}r Sonnensystemforschung, Max-Planck-Strasse 2, 37191 Katlenburg-Lindau, Germany.\\ \email{hboehnha@linmpi.mpg.de} \and Observatory, PO Box 14, 00014 University of Helsinki, Finland. \email{muinonen@cc.helsinki.fi} \and University of Maryland, College Park, MD, USA. \email{ludmilla@astro.umd.edu} \and Astronomical observatory of Kharkiv National University, 35 Sumska str., 61022 Kharkiv, Ukraine. \email{irina@astron.kharkov.ua} \and LESIA, Observatoire de Paris, 5, pl.~Jules Janssen, FR-92195 Meudon cedex, France. \email{antonella.barucci@obspm.fr} } \date{Received: 15 November 2005 / Accepted: 3 January 2006} \abstract{ Polarization is a powerful remote-sensing method to investigate solar system bodies. It is an especially sensitive diagnostic tool to reveal physical properties of the bodies whose observational characteristics are governed by small scatterers (dust, regolith surfaces). For these objects, at small phase angles, a negative polarization is observed, i.e., the electric vector $\vec{E}$ oscillates predominantly in the scattering plane, contrary to what is typical for rather smooth homogeneous surfaces. The behavior of negative polarization with phase angle depends on the size, composition and packing of the scatterers. These characteristics can be unveiled by modelling the light scattering by the dust or regolith in terms of the coherent backscattering mechanism. } {We investigate the surface properties of TNOs and Centaurs by means of polarimetric observations with FORS1 of the ESO VLT. } { We have obtained new broadband polarimetric measurements over a range of phase angles for a TNO, 50000\,Quaoar (in the $R$ Bessel filter), and a Centaur, 2060\,Chiron (in the $BVR$ Bessel filters). Simultaneously to the polarimetry, we have obtained $R$ broadband photometry for both objects. We have modelled these new observations of Quaoar and Chiron, and revised the modelling of previous observations of the TNO 28978\,Ixion using an improved value of its geometric albedo. } { TNOs Ixion and Quaoar, and Centaur Chiron show a negative polarization surge. The Centaur Chiron has the deepest polarization minimum (-1.5 -- 1.4\,\%). The two TNOs show differing polarization curves: for Ixion, the negative polarization increases rapidly with phase; for Quaoar, the polarization is relatively small ($\simeq -0.6$\,\%), and nearly constant at the observed phase angles. For all three objects, modelling results suggest that the surface contains an areal mixture of at least two components with different single-scatterer albedos and photon mean-free paths. } {} \keywords{Kuiper Belt -- polarization} \titlerunning{Polarimetry of Kuiper belt TNO objects and Centaurs} \authorrunning{S.~Bagnulo et al.\ } \section{Introduction}\label{Sect_Introduction} Transneptunian objects (TNOs) in the Kuiper Belt are considered to represent one of the oldest and possibly most original population of solar system bodies that can be observed from Earth. Centaurs are escapees from the Kuiper Belt through gravitational interaction with Neptune and the other giant planets. They may eventually become members of the Jupiter family of comets, or may be ejected from the planet region due to close encounters with the giant planets. The intense study of physical properties of TNOs and Centaurs was triggered by the advent of large telescopes on the ground: besides a large set of photometric colours, also visible and near-IR spectra of a number of objects are available now. Polarimetric observations are more scarce: except for Pluto/Charon system (that was observed unresolved, e.g., by Kelsey \& Fix \cite{KelFix73}), it was only recently that broadband polarized radiation of a TNO, the Plutino 28978\,Ixion, has been observed and modelled (Boehnhardt et al.~\cite{Boeetal04}). Polarimetry is a powerful tool to investigate the physical properties of atmosphereless bodies. At small ($\le 30\degr$) phase angles (the phase angle is the angle between the Sun and the observer as seen from the object), these objects exhibit a phenomenon of \textit{negative polarization}: the observed flux perpendicular to the plane Sun-Object-Observer (the scattering plane) minus the observed flux perpendicular to that plane, divided by the sum of the two fluxes, turns to be a negative quantity. This phenomenon, first discovered through lunar observations by Lyot (\cite{Lyot29}), escapes from common sense interpretation, since elementary physics tells that reflected electric vector $\vec{E}$ oscillates predominantly in the plane perpendicular to the scattering plane rather than in the scattering plane. Solar-system objects show two types of angular dependence of negative polarization: either a smooth phase-angle change that has the minimum at $\sim 10\degr$ (S-, C- asteroids, Moon) or a sharp surge with the minimum at $\sim 1-2\degr$ (Saturn rings, Europa, E-asteroids) (see, e.g., Rosenbush et al.~\cite{Rosetal02}). Both types of negative polarization, which also were observed in powdered laboratory samples, are currently interpreted in terms of enhanced backscattering of multiply scattered rays (Shkuratov \cite{Shkuratov89}; Muinonen \cite{Muinonen90}). Observations of negative polarization and simultaneous photometry of main-belt asteroids and other solar system bodies (see, e.g., Belskaya et al. \cite{Beletal05}; Rosenbush et al. \cite{Rosetal05}) can be modelled to infer the properties of the surface texture of these objects. Faintness of the targets was the main obstacle hampering the same kind of study in TNOs and Centaurs\footnote{Another difference in the observing and modelling techniques is that, due to the larger distance, the observed phase angle range is much smaller for TNOs and Centaurs than for main-belt asteroids}. Thanks to the advent of the large telescopes and instruments equipped with polarimetric capabilities, observations of TNOs and Centaurs are nowadays possible with signal to noise ratio comparable to that commonly reached for main-belt asteroid observations with small and middle-size telescopes. After our first polarimetric study of 28978\,Ixion (Boehnhardt et al. \cite{Boeetal04}), in this paper we present new polarimetric and photometric measurements obtained with FORS1 at the ESO Very Large Telescope (VLT) for a TNO, 50000\,Quaoar, and a Centaur, 2060\,Chiron. We also present a revised modelling of the observed polarization and photometry of 28978\,Ixion based on a determination of the geometric albedo that has been recently obtained, and that was not available at the time of our first modelling effort. \section{Target summary}\label{Sect_Targets} Criteria used for target selection are that the targets are bright enough to allow us to measure the polarization with an error bar smaller than 0.05\,\% in less than two hours telescope time. With FORS1 at the ESO VLT, this sets the $R$ magnitude limit to about 20. Another constraint is the possibility to observe the largest possible phase angle range. Complementary information on the geometric albedo and surface composition is essential for the modelling part. Moreover, it is desirable to study members of the various dynamical and taxonomic groups identified among the TNO population (Plutinos, Centaurs, classical and scattered-disk Objects). We finally selected three objects with known physical parameters (geometric albedos, colours, spectral slopes, and surface composition): Ixion, Quaoar, and Chiron. \subsection{28978\,Ixion} 28978\,Ixion, discovered in 2001, belongs to the dynamical class of Plutinos, and it is one of the largest known TNOs (400--550\,km according to Stansberry et al. \cite{Stansbe05}). The visible spectrum by Marchi et al. (\cite{Maretal03}) is featureless with a gradient $S'$ of 19.8\,\%/100\,nm. Optical and near-IR (Licandro et al. \cite{Licetal02}) spectra have been interpreted by Boehnhardt et al. (\cite{Boeetal04}) using an areal mixture of Titan tholin, amorphous carbon, water ice, and ice tholin. The same authors also present a surface model of Ixion based upon their $R$ filter polarimetry and simultaneous $R$ band photometry of the objects spanning the phase angle range 0.25\degr\ -- 1.34\degr. In that work, an $R$-band geometric albedo of 0.1 was assumed for the modelling. Here we repeat the analysis for the higher $R$-band geometric albedo now available from Spitzer observations (0.23; Stansberry, priv.\ comm.). Further details about the properties of this object are given by Boehnhardt et al.~(\cite{Boeetal04}). \subsection{50000\,Quaoar} 50000\,Quaoar is a classical disk object in the Kuiper Belt. Orbital elements and red visible colours (Fornasier et al.~\cite{Foretal04}) suggest that the object could be a member of the ``dynamically hot'' population that is supposed to have migrated to the classical disk only after formation closer to the Sun (Gomes \cite{Gomes03}). Apart from Pluto/Charon, 50000\,Quaoar is the only TNO so far for that disk-resolved photometry could be performed: HST measurements allowed to determine the overall size and geometric albedo of the object to be $1260 \pm 190$\,km, and about 0.1, respectively (Brown \& Trujillo \cite{BroTru04}). The photometric lightcurve of the object seems to be double-peaked with a period of about 17.6\,h and an amplitude of 0.13\,mag suggesting an aspherical shape of the body and/or geometric albedo variations of the surface (Ortiz et al.\ \cite{Ortetal03}). Quaoar visible spectra were obtained by Marchi et al.\ (\cite{Maretal03}) and by Fornasier et al.\ (\cite{Foretal04}). The reflectivity gradients $S'$ obtained in the two papers are not fully consistent, and their mean value is $27.6 \pm 0.3$\,\%/100\,nm. Visible spectrum appears to be featureless. Quaoar has been observed also in the near-infrared by Jewitt and Luu (\cite{JewLuu04}) at the Subaru 8\,m telescope. The complete spectrum shows a positive-slope continuum from 0.4 up to 1.3\,$\mu$m, that is considered typical for the presence of organic materials on its surface. The spectrum shows strong absorption bands at 1.5 and 2.0\,$\mu$m due to H$_2$O ice with the band at 1.65\,$\mu$m typical for the crystalline structure in the ice. A small presence of ammonia hydrate has also been supposed on the basis of the presence of faint features at 2.2\,$\mu$m (detected also by Pinilla-Alonso et al.\ \cite{Pinetal04}). This was the first time that the quality of a TNO spectrum was good enough to distinguish between crystalline and amorphous ice. The detection of crystalline ice indicates that the temperature has reached at least 110\,K (critical temperature necessary for crystallization). This object is large enough to be cryovolcanically active, and crystalline ice and ammonia hydrate might be products of this type of activity. Jewitt and Luu (\cite{JewLuu04}) suggested that Quaoar has been recently resurfaced either by impacts or by cryovolcanic outgassing or by a combination of these two processes. \subsection{2060\,Chiron} 2060\,Chiron is the first discovered (1977) and best observed Centaur. The intermediate character of the object between TNOs and comets is apparent from photometric observations that show recurrent episodes of coma activity (gas and dust) and of stellar appearance (see for example Meech \& Belton \cite{MeeBel90}, Luu \& Jewitt \cite{LuuJew90}, Bus et al.\ \cite{Busetal91}, Duffard et al.\ \cite{Dufetal02}). Chiron was observed spectroscopically in the visible region by many authors (see Barucci et al.\ \cite{Baretal03}) showing a flat spectrum with no absorption features. The reflectivity gradient $S'$, which ranges between $-0.2$ up to $2.3 \pm 0.1$\,\%/100\,nm, seems more similar to that of C-type asteroids rather than to the mean reflectance slope of cometary nuclei. The small variation on the optical reflectivity gradient could be due to dust production variation connected to episodes of recurrent cometary activity. Several spectra obtained in the near-infrared did not show any features. Only Foster et al.~(\cite{Fosetal99}) and Luu et al.~(\cite{Luuetal00}) detected a 2\,$\mu$m absorption band suggesting the presence of H$_2$O ice on Chiron's surface. Later on, Romon-Martin et al. (\cite{Rometal03}) observed Chiron again during high activity in the visible and NIR showing a flat behaviour without any spectral features finding that is compatible with the hypothesis made by Luu et al.\ (\cite{Luuetal00}) that the detection of water ice in Chiron spectra would be correlated with its cometary activity level. Such activity could cause a rain of cometary debris on its surface changing the surface mantle. The ice present on the surface, is probably mixed with dark impurities which mask the spectral bands. Nucleus properties using multi-wavelength information has revealed an about 70\,km nucleus of relatively bright geometric albedo (0.17) and moderate axis ratio (1.16) or surface albedo variations (Groussin et al.\ \cite{Groetal04}). From the large number of publications on Chiron (more than 150 to-date) several interesting properties of 2060\,Chiron have been worked out. However, a synoptic picture of the nucleus and its surface properties has not yet evolved. \section{New observations with FORS1}\label{Sect_New_Observations} Observations of Chiron and Quaoar have been obtained at the ESO VLT with the FORS1 instrument in service mode during the observing period from April to September 2004. Until June 1, 2004, FORS1 was attached at the VLT Unit Telescope 1 (Antu). After that date, FORS1 was moved to the VLT Unit Telescope 2 (Kueyen). FORS1 is a multi-mode instrument for imaging and (multi-object) spectroscopy equipped with polarimetric optics. For the present study, FORS1 has been used to measure the broadband polarization of Chiron at six different epochs in the Bessel $BVR$ filters, and to measure the broadband polarization of Quaoar at five different epochs in the Bessel $R$ filter (Sect.~\ref{Sect_Polarimetry}). In fact, one series of Chiron measurements was started on night 2004-08-05/06 and aborted after the observations in the $R$ filter because the seeing conditions were too good ($\la 0.6''$) with consequent risk of CCD saturation. The series was repeated the following night. From target acquisition images, a by-product of polarimetric observations, we could also obtain photometry in the Bessel $R$ filter (Sect.~\ref{Sect_Photometry}). Differential tracking was used for all our observations, so that long exposure time images of Chiron, obtained in polarimetric mode, could be combined altogether to search for coma activity (Sect.~\ref{Sect_Coma_Searching}). Taking advantage of the flexibility offered by the `VLT service observing mode', we distributed the observations along a few months as to obtain data points approximately equally spread over the phase angle ranges of the targets. We set precise time intervals for the execution of the observations. In presence of the Moon, the sky-background is highly polarized, hence we generally tried to avoid observations with the target close to the Moon, and with a too large fraction of lunar illumination. However, \textit{a posteriori} we found that the presence of the Moon did not jeopardize our observations. The log of the observations can be inferred from Tables \ref{Tab_Pol_Chi} to \ref{Tab_Pho_Qua}. \subsection{Polarimetry}\label{Sect_Polarimetry} To perform linear polarization measurements, a $\lambda/2$ retarder waveplate and a Wollaston prism are inserted in the FORS1 optical path (see Appenzeller \cite{App67}). The $\lambda/2$ retarder waveplate can be rotated in 22.5\degr\ steps. Stokes~$Q$ and $U$ parameters (defined as in Shurcliff \cite{Shu62}) are measured by combining the photon counts (background subtracted) of ordinary and extra-ordinary beams (\fo\ and \fe, respectively) observed at various retarder waveplate positions $\alpha$, where $\alpha$ indicates the angle between the acceptance axis of the ordinary beam of the Wollaston prism and the fast axis of the retarder waveplate. In the following, we will always work with the ratios $Q/I$ and $U/I$, and will adopt the notation: \[ \pq = \frac{Q}{I}\ \ {\rm and}\ \ \pu = \frac{U}{I} \] In the ideal case, \pq\ is obtained measuring the quantity \[ r = (-1)^k\ \frac{\fo - \fe}{\fo +\fe} \] at any retarder waveplate position $\alpha=k\,45\degr$, and \pu\ is obtained measuring the ratio $r$ at any position $\alpha=k\,45\degr + 22.5\degr$ ($k=0,\,1,\,2,\,\ldots,\,7$). The validity of this assertion can be verified e.g.\ with the help of Eq.~(1.33) of Landi Degl'Innocenti \& Landolfi (\cite{LanLan04}). In practice, there are several deviations from the ideal case. For instance, the actual retardance value of the retarder waveplate may deviate from the nominal $\pi$ value; the transmission of the ordinary and extraordinary beam are not identical, even after flat fielding correction. The effect of these (and other) sources of \textit{instrumental polarization} can be largely reduced at the first order by measuring \begin{equation} \begin{array}{rcl} \pq^{ij} &=& \frac{1}{2} \Bigg\{ \left(\frac{\fo - \fe}{\fo + \fe}\right)_{\alpha = 2(i-1) \times 45^\circ} - \left(\frac{\fo - \fe}{\fo + \fe}\right)_{\alpha = (2j-1) \times 45^\circ} \Bigg\} \\ \pu^{ij} &=& \frac{1}{2} \Bigg\{ \left(\frac{\fo - \fe}{\fo + \fe}\right)_{\alpha = 2(i-1) \times 45^\circ + 22.5^\circ} - \left(\frac{\fo - \fe}{\fo + \fe}\right)_{\alpha = (2j-1) \times 45^\circ + 22.5^\circ} \Bigg\} \\ \end{array} \label{Eq_Stokes_QU} \end{equation} where $i$ and $j$ are integers numbers\footnote{We note that in a similar formula that we reported in Boehnhardt et al.\ (\cite{Boeetal04}), the factor 2 in the indices that denote the angles is missing because of a typo.}. In the simplest case, linear polarization can be measured from the observations obtained at four angles of the retarder waveplate: \[ \begin{array}{rcl} \pq & = & \frac{1}{2} \Bigg\{ \left(\frac{\fo - \fe}{\fo + \fe}\right)_{\alpha= 0\degr} - \left(\frac{\fo - \fe}{\fo + \fe}\right)_{\alpha=45\degr} \Bigg\} \\ \pu & = & \frac{1}{2} \Bigg\{ \left(\frac{\fo - \fe}{\fo + \fe}\right)_{\alpha=22.5\degr} - \left(\frac{\fo - \fe}{\fo + \fe}\right)_{\alpha=67.5\degr} \Bigg\} \\ \end{array} \] It is convenient (and recommended in the FORS1/2 user manual) to obtain Stokes $Q$ and $U$ adding up observations obtained with the retarder waveplate at various positions: \begin{equation} \begin{array}{rcl} \pq&=&\frac{1}{m}\, \sum_{l=1}^{m} \pq^{ll} \\ [3mm] \pu&=&\frac{1}{m}\, \sum_{l=1}^{m} \pu^{ll} \;,\\ \end{array} \label{Eq_Sumpol} \end{equation} where $m$ represents the number of pairs of observations for each Stokes parameter, and $\pq^{ll}$ and $\pu^{ll}$ are obtained from Eq.~(\ref{Eq_Stokes_QU}) setting $i=j=l$. We performed simple numerical simulations to study the impact on the precision of the polarimetric measurements of a deviation of the waveplate retardance from its nominal value (180\degr). We found that using Eq.~(\ref{Eq_Sumpol}) with $m=2$, a deviation from the nominal value of the waveplate retardance as large as 5\degr, for the polarization value observed in our targets, would introduce a spurious contribution $\ll 0.01$\,\%. Figure~4.1 of FORS1/2 user manual shows that actual deviation of the retarder waveplate from the nominal value is well within 5\degr. We conclude that in our data, the effect of instrumental polarization due to the chromathism of the retarder waveplate is negligible. It should be noted that FORS1 is affected by a problem with spurious linear instrumental polarization that cannot be eliminated even by using the reduction technique explained above. This spurious polarization, due to the presence of rather curved lenses in the collimator, combined with the non complete removal of reflections by the coatings, is axially symmetric, and smoothly increases from less than 0.03\,\% on the optical axis to 0.7\,\% at an axis distance of 3 arcmin (in the $V$ band). This problem has been discovered and investigated by Patat \& Romaniello (\cite{PatRom05}). In our case, since our targets are always in the center of the field of view, the problem of the instrumental polarization can be safely ignored. The error-bar due to photon-noise on the \pq\ or \pu\ measured from a pair of observations is \begin{equation} \begin{array}{rcl} \sigma^2_{X^{ij}} & = & \left(\left(\frac{\fe}{(\Den)^2}\right)^2 \sigma^2_{\fo} + \left(\frac{\fo}{(\Den)^2}\right)^2 \sigma^2_{\fe}\right)_{\alpha=2(i-1) \times 45^\circ + \phi_0 } + \\ & & \left(\left(\frac{\fe}{(\Den)^2}\right)^2 \sigma^2_{\fo} + \left(\frac{\fo}{(\Den)^2}\right)^2 \sigma^2_{\fe}\right)_{\alpha=(2j-1) \times 45^\circ + \phi_0} \;, \\ \end{array} \label{Eq_Sigma_QU} \end{equation} where $\phi_0 = 0$ in case $X$ represents \pq\ and $\phi_0=22.5\degr$ in case $X$ represents \pu. When \pq\ and \pu\ are obtained adding up $m$ pairs as in Eq.~(\ref{Eq_Sumpol}), the error is given by \[ \sigma^2_{X} = \frac{1}{m} \sum_{l=1}^{m} \sigma^2_{X^{ll}} \] For the sake of simplicity, assuming $\fo = \fe = \cal{N}$ and also assuming $\cal{N}$ to be the same for all positions of the retarder waveplate, we obtain \begin{equation} \sigma_{X} = \frac{1}{2} \frac{1}{\sqrt{m\cal{N}}} \label{Eq_Photon_Noise} \end{equation} where $\sqrt{\cal{N}}$ represents the signal to noise ratio (SNR) measured in the individual beam for each of the $m$ exposures (in other words, $\sqrt{m\cal{N}}$ is the cumulative SNR in each beam). It appears for instance that to get an error bar on the Stokes parameter \pq\ or \pu\ of about 0.1\,\%, one should take a pair of exposures with SNR of about 500 each (integrated over the individual point-spread function area of each beam of each exposure). When multiple pairs of exposures are taken, it is useful to study the distribution of the $\pq^{ij}$ ($\pu^{ij}$) values obtained substituting $i,j = 1, 2, \ldots m$. In particular, a $\sigma$-clipping algorithm can be applied to the $\pq^{ij}$ and $\pu^{ij}$ distributions in order to ``clean'' the data, by rejecting those values that deviate more than a certain distance from the median (see Boehnhardt et al.\ \cite{Boeetal04}, and Bagnulo et al.\ \cite{Bagetal05}). Stokes~$Q$ and $U$ are usually measured with the instrument position angle = 0\degr, i.e. to have the acceptance axis of the ordinary beam of the Wollaston prism aligned to the North Celestial Meridian (and the acceptance axis of the extra-ordinary beam perpendicular to it). We then transform the Stokes parameters according to \begin{equation} \begin{array}{rcl} \pq' &=& \phantom{-}\cos(2(\Phi+\pi/2))\, \pq + \sin(2(\Phi+\pi/2))\, \pu\\ \pu' &=& - \sin(2(\Phi+\pi/2))\, \pq + \cos(2(\Phi+\pi/2))\, \pu\\ \end{array} \label{Eq_QU_Transform} \end{equation} where $\Phi$ is the angle between the direction Object-North Pole and the direction Object-Sun. This angle can be calculated applying the four parts formula to the spherical triangle defined by the object (with coordinates \alphaobj,\deltaobj), the Sun (with coordinates (\alphasun,\deltasun) and the North celestial pole: \[ \sin \deltaobj \cos(\alphasun - \alphaobj) = \cos(\deltaobj) \tan(\deltasun) - \sin(\alphasun - \alphaobj) \frac{1}{\tan(\Phi)}\;. \] This way $\pq'$ \textit{represents the flux perpendicular to the plane Sun-Object-Earth (the scattering plane) minus the flux parallel to that plane, divided by the sum of the two fluxes}. The angle of maximum polarization is obtained as \begin{equation} \Thetar = \frac{1}{2} \arctan \left(\frac{\pu'}{\pq'}\right) + \Theta_0' \end{equation} where \[ \Theta_0' = \cases {0 &{\rm if} $\pq' > 0$ and $\pu'\ge0$ \cr \pi &{\rm if} $\pq' > 0$ and $\pu'<0$ \cr \pi/2 &{\rm if} $\pq' < 0$ \cr } \;. \] or \[ \Thetar = \cases{\pi/4 &{\rm if} $\pq' = 0$ and $\pu' > 0$ \cr 3\pi/4 &{\rm if} $\pq' = 0$ and $\pu' < 0$ \cr } \] (see Landi Degl'Innocenti \& Landolfi \cite{LanLan04}.) Incidentally, it should be noted that the $\Theta_0'$ term is occasionally incorrectly neglected. Observations of Chiron were performed with the retarder waveplate at all positions between 0 and 157.5\degr\ (at 22.5\degr\ steps), i.e., setting $m=2$ in Eq.~(\ref{Eq_Sumpol}), using the broadband Bessel $B$, $V$, and $R$ filters. For Quaoar we used all positions of the retarder waveplate from 0\degr\ to 337.5\degr, i.e., setting $m=4$ in Eq.~(\ref{Eq_Sumpol}), using the Bessel $R$ filter. For each frame, exposure times $t$ were as follows. For Chiron we set $t=460, 140$, and 110\,s in the $B$, $V$, and $R$ filter, respectively. For Quaoar ($R$ filter only), we set $t = 250$\,s. Note that, for each Stokes parameter we obtained 2 pairs of exposures for Chiron, and 4 pairs of exposures for Quaoar. Therefore the total shutter time for Chiron observations was of about 60, 20, and 15\,min in the $B$, $V$, and $R$ filter, respectively, and of about 67\,min for Quaoar ($R$ filter only). For each observation of Chiron, the typical SNR accumulated in each beam was about 1700, 1600, and 1600, in the $B$, $V$, and $R$ filters respectively. For each observation of Quaoar, in the $R$ filter, we obtained a SNR accumulated in each beam of about 1400. The photon counts \fo\ and \fe\ were measured via simple aperture photometry performed on the images, obtained after bias subtraction and flatfield correction (master flat field was obtained from sky images obtained during twilight with no polarimetric optics in). More sophisticated methods based on point spread function (PSF) fitting are difficult to apply because of the star trailing due to differential tracking on the moving targets. Sky background was measured in an annulus with 5\,\arcsec\ inner radius, centered around the source, of 2\arcsec\ to 6\arcsec\ width. The errors on \fo\ and \fe\ were estimated as explained in the Sect.~3.3.5.8 of Davis (\cite{Dav87}). \fo, \fe, and their error estimates were measured for aperture values ranging from 0.6\arcsec to 4\arcsec. \pq\ and \pu\ values were found to be slightly dependent on the aperture adopted to measure \fo\ and \fe. This effect was more critical when the target was in a crowded field and/or in presence of strong background polarization by the Moon. The final value was selected as the one for which the error on \pq\ and \pu\ was minimum, i.e., usually for aperture values between 1.2\arcsec\ and 2.5\,\arcsec. \begin{table*} \caption{The observed polarization of 2060\,Chiron. The meaning of the various columns is given in the text} \label{Tab_Pol_Chi} \centering \begin{tabular}{cccclcccrrr} \hline \hline Date & Time (UT) & \multicolumn{1}{c}{Moon} & \multicolumn{1}{c}{} & \multicolumn{1}{l}{Sky} & Phase angle & Filter & \multicolumn{1}{c}{$\pq'$} & \multicolumn{1}{c}{$\pu'$} & \multicolumn{1}{c}{$\Thetar$} \\ \multicolumn{1}{c}{(yyyy mm dd)} & \multicolumn{1}{c}{(hh:mm)} & \multicolumn{1}{c}{dist.} & \multicolumn{1}{c}{FLI} & \multicolumn{1}{l}{transp.} & \multicolumn{1}{c}{(DEG)} & & \multicolumn{1}{c}{(\%)} & \multicolumn{1}{c}{(\%)} & \multicolumn{1}{c}{(DEG)} \\ \hline 2004 05 27 & 06:37 &139\degr&0.3&CLR & 3.469 &$R$&$-1.14 \pm 0.04$&$ 0.02 \pm 0.04$&$89.4 \pm 1.0$\\ 2004 05 27 & 07:03 & & & & 3.468 &$V$&$-1.20 \pm 0.04$&$ 0.13 \pm 0.04$&$86.9 \pm 0.9$\\ 2004 05 27 & 07:52 & & & & 3.467 &$B$&$-1.24 \pm 0.04$&$-0.04 \pm 0.05$&$90.9 \pm 1.1$\\[2mm] 2004 06 29 & 03:19 & 64\degr&0.4&CLR? & 1.410 &$R$&$-1.40 \pm 0.05$&$ 0.05 \pm 0.04$&$89.1 \pm 0.8$\\ 2004 06 29 & 03:44 & & & & 1.409 &$V$&$-1.35 \pm 0.05$&$-0.08 \pm 0.04$&$91.6 \pm 0.9$\\ 2004 06 29 & 04:33 & & & & 1.407 &$B$&$-1.32 \pm 0.04$&$ 0.08 \pm 0.05$&$88.2 \pm 1.2$\\[2mm] 2004 08 06 & 05:53 & 94\degr&0.7&PHO & 1.766 &$R$&$-1.40 \pm 0.04$&$ 0.00 \pm 0.04$&$89.9 \pm 0.8$\\[2mm] 2004 08 07 & 03:46 &105\degr&0.7&THN (c)& 1.828 &$R$&$-1.33 \pm 0.04$&$-0.01 \pm 0.04$&$90.2 \pm 0.8$\\ 2004 08 07 & 04:13 & & & & 1.829 &$V$&$-1.46 \pm 0.03$&$-0.14 \pm 0.03$&$92.7 \pm 0.7$\\ 2004 08 07 & 05:00 & & & & 1.832 &$B$&$-1.52 \pm 0.04$&$-0.09 \pm 0.05$&$91.8 \pm 0.9$\\[2mm] 2004 08 13 & 00:17 &167\degr&0.9&PHO (c)& 2.219 &$R$&$-1.33 \pm 0.04$&$ 0.08 \pm 0.04$&$88.3 \pm 1.0$\\ 2004 08 13 & 00:42 & & & & 2.221 &$V$&$-1.34 \pm 0.04$&$-0.08 \pm 0.04$&$91.6 \pm 0.9$\\ 2004 08 13 & 01:31 & & & & 2.223 &$B$&$-1.45 \pm 0.03$&$-0.19 \pm 0.05$&$93.8 \pm 1.1$\\[2mm] 2004 09 01 & 00:56 & 74\degr&0.5&THN? (c)& 3.327&$R$&$-1.12 \pm 0.04$&$ 0.05 \pm 0.04$&$88.7 \pm 1.1$\\ 2004 09 01 & 01:21 & & & & 3.328 &$V$&$-1.09 \pm 0.05$&$-0.08 \pm 0.04$&$92.1 \pm 1.2$\\ 2004 09 01 & 02:10 & & & & 3.330 &$B$&$-1.15 \pm 0.05$&$-0.24 \pm 0.06$&$95.9 \pm 1.6$\\[2mm] 2004 09 27 & 01:24 & 56\degr&0.4&CLR? & 4.232 &$R$&$-0.89 \pm 0.07$&$ 0.03 \pm 0.05$&$89.1 \pm 1.8$\\ 2004 09 27 & 01:49 & & & & 4.232 &$V$&$-1.05 \pm 0.05$&$-0.20 \pm 0.05$&$95.4 \pm 1.7$\\ 2004 09 27 & 02:38 & & & & 4.233 &$B$&$-0.87 \pm 0.09$&$-0.10 \pm 0.08$&$93.3 \pm 3.0$\\[2mm] \hline \end{tabular} \end{table*} \begin{table*} \caption{The observed polarization of 50000\,Quaoar. The meaning of the various columns is given in the text} \label{Tab_Pol_Qua} \centering \begin{tabular}{cccclcccrrr} \hline \hline Date & Time (UT) & \multicolumn{1}{c}{Moon} & \multicolumn{1}{c}{} & \multicolumn{1}{l}{Sky} & Phase angle & Filter & \multicolumn{1}{c}{$\pq'$} & \multicolumn{1}{c}{$\pu'$} & \multicolumn{1}{c}{$\Thetar$} \\ \multicolumn{1}{c}{(yyyy mm dd)} & \multicolumn{1}{c}{(hh:mm)} & \multicolumn{1}{c}{dist.} & \multicolumn{1}{c}{FLI} & \multicolumn{1}{l}{transp.} & \multicolumn{1}{c}{(DEG)} & & \multicolumn{1}{c}{(\%)} & \multicolumn{1}{c}{(\%)} & \multicolumn{1}{c}{(DEG)} \\ \hline 2004 04 18 & 04:43 & 47\degr&1.0&CLR & 0.952 &$R$&$-0.64 \pm 0.05$&$ 0.14 \pm 0.08$&$83.9 \pm 3.9$\\ 2004 05 13 & 07:38 & 93\degr&0.8&CLR? & 0.496 &$R$&$-0.50 \pm 0.06$&$-0.14 \pm 0.05$&$97.7 \pm 3.6$\\ 2004 05 26 & 03:37 &107\degr&0.2&PHO & 0.252 &$R$&$-0.49 \pm 0.06$&$ 0.10 \pm 0.05$&$84.3 \pm 3.5$\\ 2004 07 10 & 01:44 &118\degr&0.7&PHO (c) & 0.797 &$R$&$-0.53 \pm 0.04$&$ 0.09 \pm 0.05$&$85.3 \pm 3.1$\\ 2004 08 11 & 02:34 & 75\degr&0.8&CLR? & 1.231 &$R$&$-0.65 \pm 0.04$&$-0.01 \pm 0.06$&$90.4 \pm 2.6$\\ \hline \end{tabular} \end{table*} The results of the polarimetric measurements are given in Tables~\ref{Tab_Pol_Chi} and \ref{Tab_Pol_Qua}, that are organized as follows. Columns 1 and 2 give the epoch of the observations (date and UT time). Columns~3 and 4 give Moon angular distance and fraction of lunar illumination (FLI), respectively. Column 5 gives an estimate of the night time sky conditions: THN = thin cirrus, CLR= clear, PHO = photometric. The classification given in these Tables is based on our inspection of reduction products and of the atmospheric monitors of the observatory at \\ {\tt http://www.eso.org/gen-fac/pubs/astclim/}\\ \ \ {\tt forecast/meteo/CIRA/images/repository/lossam/} \\ and is somewhat arbitrary. Those nights when no photometric standard stars were observed, or when the zeropoints were not considered stable for the entire night, are indicated with a question mark. Sky transparency does not affect the precision of the polarization measurements, but it does affect the precision of the photometry (see Tables~\ref{Tab_Pho_Chi} and \ref{Tab_Pho_Qua}). Note that the precision of the measurements (both polarimetric and photometric) depends also on how the field of view close to the target is crowded with background objects. This situation of course changes from epoch to epoch. Observations labelled with (c) are hampered to some extent by a crowded background. In the case of Chiron observations on 2004-09-01 the situation was probably especially critical. Column 6 gives the Sun-Target-Observer angle, i.e., the target's apparent phase angle as seen at observer's location, expressed in degrees. This phase angle was obtained from the object ephemeris calculated at the JPL's solar system dynamics WWW site at {\tt http://ssd.jpl.nasa.gov}. Column~7 specifies the broadband filter used for the observations. Columns~8, 9, and 10 give \pq', \pu', and the angle of maximum polarization $\Thetar$ , respectively, after the transformation of Eq.~(\ref{Eq_QU_Transform}). It should be noted that for both Chiron and Quaoar, the measured $\pu'$ is generally consistent with zero (equivalent to the fact that $\Thetar$ is generally consistent with 90\degr). This means that the principal axes of the polarization ellipse are aligned with the coordinate axes perpendicular and parallel to the scattering plane. Furthermore, the observed $\pq'$, i.e., the flux perpendicular to the scattering plane minus the flux parallel to that plane, divided by the sum of the two fluxes, is always a \textit{negative} quantity. This means that the direction of the polarization is included in the scattering plane (this case is normally referred to as ``negative polarization''), in contrast to what is expected for a dielectric medium, i.e., that the direction of polarization be perpendicular to the scattering plane (this case is normally referred to as ``positive polarization''). Polarimetric measurements of Quaoar and Chiron are plotted against object phase angle in Figs.~\ref{Fig_Pol_Qua} and \ref{Fig_Pol_Chi}. As far as Chiron observations are concerned, it appears that the observed polarization does not depend on the filter used, at least when considering the error bars of the observations. The absolute value of the polarization increases as the phase angle decreases, perhaps reaching a minimum around phase 1.5-2.0 deg. Observations of this object at smaller phase angles were unfortunately not taken although originally scheduled in our service observing campaign. \subsection{Photometry}\label{Sect_Photometry} The object magnitudes were measured in the acquisition images obtained in the $R$ Bessel filter with no polarimetric optics in the light path. Exposure times were 5\,s for Chiron, and 30\,s for Quaoar. Since our program was aimed at polarization measurements, we did not obtain a number of observations of photometric standard stars sufficient to estimate precise zeropoints and extinction coefficients for the various nights. In most of the cases, the zeropoints were obtained from only one frame obtained during the night when our observations were executed (and not necessarily close to them in time). In some cases, no photometric standard stars were observed at all (see Tables~\ref{Tab_Pol_Chi} and \ref{Tab_Pol_Qua}). In these cases we adopted the values measured in the nearest night with similar sky conditions. For the extinction coefficient $K_R$ and the colour term $k_{VR}$ we used the values estimated for the P73 ESO observing period published at\\ {\tt http://www.eso.org/observing/dfo/quality/FORS1/}\\ \ \ {\tt qc/photcoeff/photcoeffs\_fors1.html}\\ i.e., \[ \begin{array}{rcl} K_R &=& 0.045 \pm 0.019 \\ k_{VR} &=& 0.0090 \pm 0.0017 \\ \end{array} \] For the Quaoar colour indices we adopted $V-R = 0.64 \pm 0.01$ (Tegler et al. \cite{Tegetal03}), and for Chiron $V-R = 0.37 \pm 0.03$ (Davies et al. \cite{Davetal98}). Finally, reduced magnitude $R$ of the target was obtained by taking into account the distance Sun-Object $r$ and the distance Earth-Object $\Delta$ at the epoch of the observations, and calculating the reduced magnitude \begin{equation} R = m_R -5\,\log (r)- 5\,\log (\Delta) \label{Eq_Reduced_Mag} \end{equation} where $m_R$ is the observed $R$-band magnitude. The results of our photometry are given in Tables~\ref{Tab_Pho_Chi} and \ref{Tab_Pho_Qua}. From a linear least-square fit of the data, we found that for Chiron the $H_R$ absolute magnitude at zero phase angle is $5.52 \pm 0.07$\,mag, and the slope of the opposition surge is $0.045 \pm 0.023$\,mag/deg. Previous numerous photometric observations of Chiron (for summary see Groussin et al. \cite{Groetal04}) have shown considerable variations of absolute magnitude with heliocentric distance attributed to its sporadic cometary behavior. The absolute magnitude of Chiron in 2004 is brighter as compared to the observations in 1988-1990 made at the same heliocentric distances (Bus et al.\ \cite{Busetal91}). It may indicate that we observed Chiron closely after an activity period which may have caused partial resurfacing of the object. This is why we paid a special attention to search for a coma around Chiron (see Sect.~\ref{Sect_Coma_Searching}). For Quaoar, the $H_R$ absolute magnitude at zero phase angle, obtained from a \textit{linear fit} is $2.16 \pm 0.05$\,mag, and the linear slope is $0.16 \pm 0.06$\,mag/deg. The determined value of absolute magnitude coincides with that used for Quaoar for geometric albedo determination (Brown \& Trujillo \cite{BroTru04}). Note that $H_R$ absolute magnitudes at zero phase angles and linear slopes will be re-estimated based on our modelling technique in Sect.~\ref{Sect_Modelling}. \begin{table} \caption{The observed $R$ photometry of 2060\,Chiron} \label{Tab_Pho_Chi} \centering \begin{tabular}{ccccc} \hline \hline Date & Time (UT) & Phase angle & \multicolumn{1}{c}{$m_R$} & \multicolumn{1}{c}{$R$} \\ \multicolumn{1}{c}{(yyyy mm dd)} & \multicolumn{1}{c}{(hh:mm)} & \multicolumn{1}{c}{(DEG)} & (mag) & (mag) \\ \hline 2004 05 27 & 06:23 & 3.470 & $16.64 \pm 0.04$& 5.68 \\ 2004 06 29 & 03:04 & 1.411 & $16.48 \pm 0.05$& 5.56 \\ 2004 08 06 & 05:40 & 1.766 & $16.49 \pm 0.04$& 5.54 \\ 2004 08 07 & 03:33 & 1.828 & $16.65 \pm 0.04$& 5.69 \\ 2004 08 13 & 00:02 & 2.219 & $16.55 \pm 0.05$& 5.58 \\ 2004 09 01 & 00:39 & 3.326 & $16.65 \pm 0.04$& 5.63 \\ 2004 09 27 & 01:10 & 4.232 & $16.84 \pm 0.04$& 5.74 \\ \hline \end{tabular} \end{table} \begin{table} \caption{The observed $R$ photometry of 50000\,Quaoar} \label{Tab_Pho_Qua} \centering \begin{tabular}{ccccc} \hline \hline Date & Time (UT) & Phase angle & \multicolumn{1}{c}{$m_R$} & \multicolumn{1}{c}{$R$} \\ \multicolumn{1}{c}{(yyyy mm dd)} & \multicolumn{1}{c}{(hh:mm)} & \multicolumn{1}{c}{(DEG)} & (mag) & (mag) \\ \hline 2004 04 18 & 03:59 & 0.952 & $18.60 \pm 0.05$& 2.27 \\ 2004 05 13 & 06:54 & 0.497 & $18.60 \pm 0.04$& 2.27 \\ 2004 05 26 & 02:52 & 0.252 & $18.51 \pm 0.04$& 2.19 \\ 2004 07 10 & 00:59 & 0.796 & $18.57 \pm 0.04$& 2.24 \\ 2004 08 11 & 01:49 & 1.231 & $18.73 \pm 0.04$& 2.38 \\ \hline \end{tabular} \end{table} \subsection{Search for a coma around 2060\,Chiron}\label{Sect_Coma_Searching} Visual inspection and measurements of the full-width-at-half-maximum of the Chiron and neighbouring star images did not reveal any indications for the presence of a coma around the Centaur. For a more thorough analysis we calculated the radial flux profile measured in concentric rings around the object, and compared it with the radial profile across the star trails, averaged along the trail direction and for both sides of the trail. The resulting object and star profiles are normalized to unity brightness at 'one pixel' central distance and to zero at background distance 30\,pixels from the center. Figure~\ref{Fig_Coma} shows the results for three observing dates (for other dates the presence of star trails close to the object image jeopardized the accuracy of this analysis). Since the radial profile of Chiron (solid line) is basically identical to that of the comparison stars (dotted line) or falls at slightly lower flux levels, we conclude that a coma around Chiron is not present or it is well beyond our detection limit (order of 31\,mag/arcsec) - if present at all. Hence, we assume that the polarimetry of the object is not contaminated by dust around the object. \section{Modelling of the observed polarization}\label{Sect_Modelling} \begin{table} \caption{ The best fit coherent-backscattering model parameters for Ixion, Quaoar, and Chiron. We give the single-scattering albedos $\tilde{\omega}$ and dimensionless mean free paths $k\ell$ for the dark (subscript $d$) and bright components ($b$), the weight of the dark component $w_d$, the rms values of the polarimetric fits, as well as the $R$-band absolute magnitudes $H_R$ and slope parameters $k_R$. } \label{Table_Results} \begin{tabular}{llll} \hline\hline & Ixion & Quaoar & Chiron \\ \hline $\tilde{\omega}_d$ & 0.45 & 0.35 & 0.15 \\ $\ell_d$ & 250 & 300 & 120 \\ $\tilde{\omega}_b$ & 0.80 & 0.50 & 0.60 \\ $\ell_b$ & 20 & 10 & 500 \\ $w_d$ & 0.74 & 0.46 & 0.14 \\ rms & 0.029\,\% & 0.069\,\% & 0.067\,\% \\ $H_R$ & 3.25 & 2.15 & 5.41 \\ $k_R$ & 0.12\,deg$^{-1}$& 0.11\,deg$^{-1}$& 0.041\,deg$^{-1}$\\ \hline \end{tabular} \end{table} We interpret the polarimetric and photometric phase curves of Ixion, Quaoar, and Chiron through extensive numerical simulations of coherent backscattering by Rayleigh scatterers. Note that we avoid making an assumption that the fundamental scatterers responsible for the coherent backscattering contribution would be the single particles in the regolith. With the phenomenological modeling currently including the first multipole contribution of an electric dipole, we allow for the possibility that the fundamental scatterers can be the volume and surface inhomogeneities within and on the particles, respectively. Shadowing among the regolith particles is known to contribute to the opposition effect but not to the negative polarization surge. Here, as in Boehnhardt et al. (\cite{Boeetal04}), the shadowing effect manifests itself in the residual slope parameter resulting from the combined coherent backscattering modeling on the polarimetry and photometry. The coherent-backscattering mechanism is a multiple-scattering mechanism for scattering orders higher than the first one. For a recent review of the coherent-backscattering mechanism and its relevance in asteroid studies, see Muinonen et al. (\cite{Muietal02}) and Muinonen (\cite{Muinonen04}). Note that the computational technique for coherent backscattering accounts for all orders of scattering contributing in a non-negligible way to the backscattering peaks and polarization surges. In the following we limit ourselves to illustrate the mechanism at the second order of scattering. Let us consider a semi-infinite random medium of discrete scatterers, constrained by a plane-parallel boundary with free space, and an electromagnetic plane wave incident on the random medium from the free space. Let us assume that the incident wave interacts with one of the scatterers, giving rise to first-order scattering, and that, subsequently, the first-order scattered field interacts with another scatterer, giving rise to second-order scattering. Let us assume that the field scattered by the second scatterer escapes the medium and is detected by the observer in the free space. Now there is the reciprocal sequence of scatterings in the opposite direction involving the very same two scatterers. In the exact backward scattering direction, due to the identical lengths of the propagation paths, the two reciprocal wave components interfere constructively whereas, in other directions, the interference characteristics vary. After configurational averaging, a backscattering peak results in the proximity of the backward scattering direction. Focusing in on the polarization characteristics, the interference is selective and tends to invert the linear polarization characteristics in single scattering: for positively polarizing single scattering, coherent backscattering tends to result in negative polarization. The illustration can be readily generalized to higher orders of scattering. In the modeling that follows, all relevant orders of scattering are taken into account. For the modelling of Ixion, Quaoar, and Chiron, we carried out coherent backscattering computations for a total of 360 spherical media of Rayleigh scatterers, as follows: we assumed 18 different single-scattering albedos \[ \tilde{\omega}=0.05, 0.10, \ldots, 0.90 \] and 20 different dimensionless mean free paths \[ \begin{array}{rcl} k\ell = 2\pi \ell/\lambda &=& 10, 20, 30,\ \ldots,\ 100, 120, 140,\ \ldots, \\ & & 200, 250, 300,\ \ldots,\ 400, 500 \\ \end{array} \] where $k$ and $\lambda$ are the wave number and wavelength, respectively. As to the $R$-band geometric albedos $p_R$, we adopted $p_R=0.23$ for Ixion (Stansberry, priv.\ comm.), 0.11 for Quaoar (calculated using the diameter from Brown \& Trujillo \cite{BroTru04} and our determination of the absolute magnitude), and 0.17 for Chiron (Barucci et al. \cite{Baretal04}). In our previous interpretation of Ixion polarimetric data (Boehnhardt et al. \cite{Boeetal04}), we had assumed $p_R = 0.1$. We compared the polarimetric observations against the spherical media composed of monodisperse Rayleigh scatterers with given single-scattering albedo and mean free path (two parameters). The fits were poor: for a fixed geometric albedo, the monodisperse Rayleigh-scattering model tends to result in polarizations that are substantially pronounced as compared to the polarizations observed. As in Boehnhardt et al. (\cite{Boeetal04}), we then studied a two-component Rayleigh-scattering model consisting of dark and bright scatterers. There are five parameters in such a model: two single-scattering albedos and two mean free paths, and the weight factor for the dark component (one minus the weight factor of the bright component). Fixing the geometric albedo fixes the weight factor, reducing the number of free parameters to four. After a systematic study of physically realistic combinations of the two kinds of scatterers, satisfactory fits were obtained for the polarizations of all three objects. For Ixion, the data point at phase angle 0.43\degr\ was omitted as an outlier. The model parameters and rms values of the fits are summarized in Table~\ref{Table_Results} and the actual fits are depicted in Figs.~\ref{Fig_Pol_Ixi} -- \ref{Fig_Pol_Chi}. After the polarization fits, approximate brightness fits were obtained by varying the absolute magnitude $H_R$ and slope parameter $k_R$ of a linear phase dependence multiplying the coherent backscattering contribution (see Fig.~\ref{Fig_Pho_Ixi} -- \ref{Fig_Pho_Chi}). For Chiron, the observation at 1.8\degr\ phase angle was omitted as outlier. For all three objects, the dark component shows a mean free path substantially longer than the wavelength. For Chiron, the mean free path of the bright component is also considerably longer than the wavelength. The dark components of Ixion and Quaoar resemble one another and the bright component of Chiron: this is concluded from the similarity of the parameters of the corresponding scatterers on the surfaces of the three objects. The phase curves of Ixion and Chiron resemble each other to an extent where their combined polarimetric and photometric data could be explained using a single two-component scattering model. The polarimetric observations suggest that Ixion and Chiron could have similar surface structure. It is notable that the geometric albedo ranges of Ixion and Chiron overlap, whereas Quaoar stands out as a darker object. More observational and theoretical work is required for further conclusions. \section{Discussion and Conclusions}\label{Sect_Discussion} We have presented the results of the first polarimetric observations for a Centaur and a classical disk object in the Kuiper Belt. Together with polarimetric observations of Ixion (Boehnhardt et al., \cite{Boeetal04}), a representative of the Plutino population in the belt, they give us a first idea about the behavior of linear polarization and intensity of the light reflected by surfaces of Kuiper belt objects and Centaurs. Two Kuiper Belt objects, Ixion and Quaoar, were observed practically in the same range of phase angles (0.25\degr--1.3\degr). They have shown completely different polarization-phase behaviour. For Ixion the negative polarization rapidly increases with phase angle ($-1.02 \pm 0.025$\,\%/deg), while for Quaoar the polarization degree is changed very slowly ($-0.17 \pm 0.10$\,\%/deg). Our polarimetric observations of Chiron were made for larger phase angles, 1.4\degr--4.2\degr. They are characterized by a pronounced branch of negative polarization with a minimum of 1.4--1.5\,\% at phase angles of 1.5\degr--2.0\degr. Such polarimetric characteristics are unique among Solar System bodies. For the majority of Solar System bodies the polarization at the phase angle 1\degr\ is characterized by values within 0.1--0.5\,\%. The largest values of the polarization at 1\degr\ measured for non-TNO objects were 0.83\,\% for the dark side of Iapetus (albedo = 0.05) and 0.73\,\% for Saturn Ring A (albedo = 0.75) (Rosenbush et al.\ \cite{Rosetal02}) that is noticeably smaller than the values obtained at our observations of Ixion and Chiron. Polarimetric observations of Chiron at smaller phase angles are urgently needed for a better modelling, and to compare Chiron data with the polarimetric curve of Ixion and Quaoar. Further observations would be also desirable to identify the inversion angle. Our modelling has shown that the possible way to explain observed polarization properties of KBOs and Centaur is to assume two-component surface media consisting of dark and bright scatterers. Such a model succeeds in fitting all observed polarimetric and photometric characteristics of the three objects and the two-component modelling is realistic. A more thorough theoretical study is beyond the scope of the present article and should be carried out in the nearest future. \begin{acknowledgements} The authors wish to thank D.\ Rabinowitz for sharing his unpublished photometric data of Quaoar which have helped us to verify the calibration of our data, E.~Landi Degl'Innocenti and M.\ Landolfi for their help to write Sect.~3, and O.~Hainaut for his help to identify adequate observing periods for our targets (i.e., with no or little risk of background star confusion). \end{acknowledgements}
Title: The Offline Software Framework of the Pierre Auger Observatory
Abstract: The Pierre Auger Observatory is designed to unveil the nature and the origins of the highest energy cosmic rays. The large and geographically dispersed collaboration of physicists and the wide-ranging collection of simulation and reconstruction tasks pose some special challenges for the offline analysis software. We have designed and implemented a general purpose framework which allows collaborators to contribute algorithms and sequencing instructions to build up the variety of applications they require. The framework includes machinery to manage these user codes, to organize the abundance of user-contributed configuration files, to facilitate multi-format file handling, and to provide access to event and time-dependent detector information which can reside in various data sources. A number of utilities are also provided, including a novel geometry package which allows manipulation of abstract geometrical objects independent of coordinate system choice. The framework is implemented in C++, and takes advantage of object oriented design and common open source tools, while keeping the user side simple enough for C++ novices to learn in a reasonable time. The distribution system incorporates unit and acceptance testing in order to support rapid development of both the core framework and contributed user code.
https://export.arxiv.org/pdf/astro-ph/0601016
\title{The Offline Software Framework of the Pierre Auger Observatory} \author{S. Argir{\`o}, S.L.C. Barroso, J. Gonzalez, L. Nellen, T. Paul, T.A. Porter,\\ L. Prado Jr., M. Roth, R. Ulrich and D. Veberi{\v c} \thanks{The work of J. Gonzalez and T. Paul was funded in part by the National Science Foundation. The work of T.A. Porter was funded in part by the US Department of Energy. }% \thanks{S. Argir{\'o} is with the INFN and University of Torino, S.L.C. Barroso is with the Centro Brasileiro de Pesquisas F{\'i}sicas, J.Gonzalez and T. Paul are with Northeastern University, L. Nellen is with the Universidad Nacional Aut{\'o}noma de M{\'e}xico, T. Porter is with Louisiana State University, L. Prado Jr. is with the University of Campinas, M. Roth and R. Ulrich are with the Forschungzentrum Karlsruhe, and D. Veberi{\v c} is with Nova Gorica Polytechnic. }} \begin{keywords} offline software, framework, object oriented, simulation, cosmic rays \end{keywords} \section{Introduction} \PARstart{T}{he} offline software framework of the Pierre Auger Observatory~\cite{Abraham:2004dt} provides an infrastructure to support a variety of distinct computational tasks necessary to analyze data gathered by the observatory. The observatory itself is designed to measure the extensive air showers produced by the highest energy cosmic rays ($> 10^{19}$~eV) with the goal of discovering their origins and composition. Two different techniques are used to detect air showers. Firstly, a collection of telescopes is used to sense the fluorescence light produced by excitation of nitrogen induced by the cascade of particles in the atmosphere. This method can be used only when the sky is moonless and dark, and thus has roughly a 10\% duty cycle. The second method uses an array of detectors on the ground to sample particle densities as the air shower arrives at the Earth's surface. Each surface detector consists of a tank containing 12 tons of purified water instrumented with photomultiplier tubes to detect the Cherenkov light produced when particles pass through it. The surface detector has a 100\% duty cycle. A subsample of air showers detected by both instruments, dubbed hybrid events, are very precisely measured and provide an invaluable energy calibration tool. In order to cover the full sky, the observatory will consist of two sites, one in the southern hemisphere and one in the north. The southern site is located in Mendoza, Argentina, and construction there is nearing completion, at which time the observatory will comprise 24 fluorescence telescopes overlooking 1600 surface detectors spaced 1.5~km apart on a hexagonal grid. Colorado has been selected as the location for the northern site. The requirements of this project place rather strong demands on the software framework underlying data analysis. Most importantly, the framework must be flexible and robust enough to support the collaborative effort of a large number of physicists developing a variety of applications over the projected 20 year lifetime of the experiment. Specifically, the software must support simulation and reconstruction of events using surface, fluorescence and hybrid methods, as well as simulation of calibration techniques and other ancillary tasks such as data preprocessing. Further, as the experimental run will be long, it is desirable that the software be extensible in case of future upgrades to the observatory instrumentation. The offline framework must also handle a number of data formats in order to deal with event and monitoring information from a variety of instruments, as well as the output of air shower simulation codes. Additionally, it is desirable that all physics code be ``exposed'' in the sense that any collaboration member must be able to replace existing algorithms with his own in a straightforward manner. Finally, while the underlying framework itself may exploit the full power of C++ and object-oriented design, the portions of the code directly used by physicists should not assume a particularly detailed knowledge of these topics. The offline framework was designed with these principles in mind. Implementation has taken place over the last three years, and the first physics results based upon this code were recently presented at the $29^{\rm th}$ International Cosmic Ray Conference~\cite{icrc2005}. \section{Overview} The offline framework comprises three principal parts: a collection of processing {\em modules} which can be assembled and sequenced through instructions provided in an XML file, an {\em event} structure through which modules can relay data to one another and which accumulates all simulation and reconstruction information, and a {\em detector description} which provides a gateway to data describing the configuration and performance of the observatory as well as atmospheric conditions as a function of time. These principal ingredients are depicted in figure~\ref{f:general}. These components are complimented by a set of foundation classes and utilities for error logging, physics and mathematical manipulation, as well as a unique package supporting abstract manipulation of geometrical objects. Each of these aspects of the framework is described in more detail below. \section{User code, Modules and Configuration} \label{sec:config} Experience has shown that most tasks of interest of the Pierre Auger Collaboration can be reasonably factorized into sequences of well-defined processing steps. Physicists prepare such processing algorithms in so-called {\em modules}, which they can insert into the framework via a registration macro in their code. This modular design allows collaborators to easily exchange code, compare algorithms and build up a wide variety of applications by combining modules in various sequences. Run-time control over module sequences is afforded through a {\em run controller} which invokes the various processing steps within the modules according to a set of externally provided instructions. As most of our applications do not require extremely sophisticated sequencing control at the module level, we have chosen to construct a very simple XML~\cite{xml}-based language for specifying sequencing instructions. This provides users with a tool which can be learned very quickly, but which is still rich enough to still support most common applications. Figure~\ref{f:xml} shows a simple example of the structure of a sequencing file. Parameters, cuts and configuration instructions used by modules or by the framework itself are also stored in XML files. A central directory points modules to their configuration file(s) using a local filename or URI and creates parsers to assist in reading information from these files. This directory is constructed from a {\em bootstrap} file whose name is passed on the command line at run time. The configuration mechanism can also concatenate and write a log file with all configuration data accessed during a run. The log file format identical to the {\em bootstrap} format, so the logs can be subsequently read in to reproduce a run with an identical configuration. This configuration logging mechanism may also be used to record the versions of modules and external libraries which are used during a run. To check configuration files for errors, we employ XML Schema~\cite{schema} validation throughout. This has proved successful in saving coding time for developers and users alike, and facilitates much more detailed error checking than most users are likely to implement on their own. All XML handling is based upon the Xerces~\cite{xerces} validating parser, augmented by a wrapper to simplify use and compliment functionality with features such as unit handling, expression evaluation, and casting of data in XML files to atomic types or STL containers. \section{Data Access} The offline framework provides two parallel hierarchies for accessing data: the {\em event} for reading and writing information that changes per event, and the read-only {\em detector description} for retrieving static or slowly varying information such as detector geometry, calibration constants, and atmospheric conditions. \subsection{Event} The event data structure contains all raw, calibrated, reconstructed and Monte Carlo data and acts as the principal backbone for communication between modules. As it is a communication backbone, reference semantics are used throughout to access data structures in the event and constructors are kept private to prevent accidental copying of event components. The event structure is built up dynamically as needed, and is instrumented with a simple protocol allowing modules to interrogate the event at any point to discover its current constituents. The event representation in memory is decoupled from the representation on disk. Serialization is currently implemented using the ROOT~\cite{root} toolkit, though the design is intended to allow for relatively straightforward changes of serialization machinery. A set of simple-to-use input/output modules are provided to allow users to transfer all or part of the event from memory to a file at any stage in the processing, and to reload the event and continue processing from that point onward. These modules are are built upon a a set of utilities to support the multi-format reading and writing required to deal with different raw event and monitoring formats as well as the different formats used by the AIRES~\cite{aires}, CORSIKA~\cite{corsika} and CONEX~\cite{conex} air shower simulation packages. \subsection{Detector Description} \label{sec:detector} The {\em detector description} provides a unified interface from which module authors can retrieve information about the detector configuration and performance at a particular time. The interface is organized following the hierarchy normally associated with the observatory instruments. Requests for data are passed by this interface to a registry of so-called {\em managers}, each of which is capable of extracting a particular sort of information from a particular data source. Data retrieved from the manager are cached in the interface for future use. In this approach, the user deals with a single interface even though the data sought may reside in any number of different sources. Generally we choose to store static detector information in XML files, and time-varying monitoring and calibration data in MySQL~\cite{mysql} databases. The structure of the detector description machinery is illustrated in figure~\ref{f:detector}. Note that it is possible to implement more than one manager for a particular sort of data. In this way, a special manager can override data from a general manager. For example, a user can decide to use a database for the majority of the description of the detector, but override some data by writing them in an XML file which is interpreted by the special manager. The specification of which data sources are accessed by the manager registry and in what order they are queried is detailed in a configuration file. The configuration of the manager registry is transparent to the user code. The detector description is also equipped to support a set of plug-in functions, called {\em models} which can be used for additional processing of data retrieved through the detector. These are used primarily to interpret atmospheric monitoring data. As an example, users can invoke a model designed to evaluate attenuation due to aerosols between two points in the atmosphere. This model interrogates the detector interface to find the atmospheric conditions at the specified time, and computes the attenuation. Models can also be placed under command of a {\em super-model} which can attempt various methods of computing the desired result, depending on what data are available at the specified time. \section{Utilities} The offline framework is built on a collection of utilities, including a XERCES-based XML parser, an error logger, various mathematics and physics services, testing utilities and a set of foundation classes to represent objects such as signal traces, tabulated functions and particles. The utilities collection also includes a novel geometry package, which we describe in more detail here. As discussed previously, the Pierre Auger Observatory comprises many instruments spread over a large area and, in the case of the fluorescence telescopes, oriented in different directions. Consequently there is no natural preferred coordinate system for the observatory; indeed each detector component has its own preferred system, as do components of the event such as the air shower itself. Furthermore, since the detector spans some 40~km from side to side, the curvature of the earth cannot generally be neglected. In such a circumstance, the necessity of keeping track of all the the required transformations when performing geometrical computations is tedious and error prone. This problem is alleviated in the offline geometry package by providing geometrical objects such as points and vectors which keep track of the coordinate system in which they are represented. Operations on these objects can then be written in an abstract way, independent of any particular coordinate system choice. There is no need for pre-defined coordinate system conventions, or coordinate system conversions at module boundaries. Coordinate systems themselves are defined in terms of transformations of other coordinate systems, with an ultimate root coordinate system as the foundation. In order to avoid reliance on this root coordinate system by all of the client code, a registry of pre-defined coordinate systems is provided. Furthermore, various specialized coordinate systems, such as the coordinate system of the shower or one of the telescopes, can be retrieved from different parts of the event and detector description. Locations of detector components are provided by survey teams in UTM (Universal Transverse Mercator) coordinates, which are convenient for navigation, but less so for data analysis. The geometry package therefore includes support for transformations between geodetic and Cartesian coordinates. \section{Build System and Quality Control} To help ensure code maintainability and stability in the face of a large number of contributors and a decades long experimental run, unit and acceptance testing are integrated into the offline framework build and distribution system. This sort of quality assurance mechanism is crucial for any software which must continue to develop over a timescale of many years. Our build system is based on the GNU autotools~\cite{autotools}, which provide hooks for integrating tests with the build and distribution system. A substantial collection of unit tests has been developed, each of which is designed to comprehensively test a single framework component. These unit tests are run at regular intervals, and in particular prior to releasing a new version of the software. We have employed the CppUnit~\cite{cppunit} testing framework as an aid in implementing these unit tests. We are currently in the process of developing more involved acceptance tests which will be used to verify that modules and framework components working in concert continue to function properly during ongoing development. \section{External packages} The choice of external packages upon which to build the offline framework was dictated not only by package features, support and the requirement of being open-source, but also by our best assessment of prospects for longevity. At the same time, we attempted to avoid locking the design to any single-provider solution. To help achieve this, the functionality of external libraries is often provided to the client code through wrappers or fa{\c c}ades, as in the case of XML parsing described in section~\ref{sec:config}, or through a bridge, as in the case of the detector description described in section~\ref{sec:detector}. The collection of external libraries currently employed includes ROOT~\cite{root} for serialization, Xerces~\cite{xerces} for XML parsing and validation, CLHEP~\cite{clhep} for expression evaluation and geometry foundations, Boost~\cite{boost} for its many valuable C++ extensions, and optionally Geant4~\cite{g4} for detailed detector simulations. \section{Conclusions} We have implemented an offline software framework for the Pierre Auger Observatory. It provides machinery to help collaborators work together on data analysis problems, compare results, and carry out production runs of large quantities of simulated or real data. The framework is configurable enough to adapt to a diverse set of applications, while the user side remains simple enough for C++ non-experts to learn in a reasonable time. The modular design allows straightforward swapping of algorithms for quick comparisons of different approaches to a problem. The interfaces to detector and event information free the users from having to deal individually with multiple data formats and data sources. This software, while still undergoing vigorous development and improvement, has been used in production of the first physics results from the observatory. \section*{Acknowledgment} The authors would like to thank the fearless early users of the offline framework.
Title: XMM-Newton View of the z>0 Warm-Hot Intergalactic Medium Toward Markarian 421
Abstract: The recent detection with Chandra of two warm-hot intergalactic medium (WHIM) filaments toward Mrk 421 by Nicastro et al. provides a measurement of the bulk of the "missing baryons" in the nearby universe. Since Mrk 421 is a bright X-ray source, it is also frequently observed by the XMM-Newton Reflection Grating Spectrometer (RGS) for calibration purposes. Using all available archived XMM observations of this source with small pointing offsets (<15"), we construct the highest-quality XMM grating spectrum of Mrk 421 to date with a net exposure time (excluding periods of high background flux) of 437 ks and \~15000 counts per resolution element at 21.6A, more than twice that of the Chandra spectrum. Despite the long exposure time neither of the two intervening absorption systems is seen, though the upper limits derived are consistent with the Chandra equivalent width measurements. This appears to result from (1) the larger number of narrow instrumental features caused by bad detector columns, (2) the degraded resolution of XMM/RGS as compared to the Chandra/LETG, and (3) fixed pattern noise at \lambda > 29A. The non-detection of the WHIM absorbers by XMM is thus fully consistent with the Chandra measurement.
https://export.arxiv.org/pdf/astro-ph/0601620
command. \def\hi{\ifmmode {\rm H}\,{\sc i}~ \else H\,{\sc i}~\fi} \def\kms{\rm\,km\,s^{-1}} \def\hubunits{\rm\,km\,s^{-1}\,Mpc^{-1}} \def\K{\,{\rm K}} \def\cm{{\rm cm}} \def\chandra {{\it Chandra}} \def\xmm {{\it XMM}} \def\xmmnewton {{\it XMM--Newton}} \def\fuse {{\it FUSE}} \def\cvi {\ion{C}{6}} \def\nvi {\ion{N}{6}} \def\neix {\ion{Ne}{9}} \def\nvii {\ion{N}{7}} \def\nvi {\ion{N}{6}} \def\ovi {\ion{O}{6}} \def\ovii {\ion{O}{7}} \def\oviii {\ion{O}{8}} \def\ovihv {\ion{O}{6}$_{\rm HV}$} \def\novi {N_{\rm OVI}} \def\novii {N_{\rm OVII}} \def\noviii {N_{\rm OVIII}} \def\kms {~km~s$^{-1}$} \shorttitle{XMM Observations of Mrk 421} \shortauthors{Williams et al.} \begin{document} \title{\boldmath{\it XMM--Newton} View of the $z>0$ Warm--Hot Intergalactic Medium Toward Markarian 421} \author{Rik J. Williams\altaffilmark{1}, Smita Mathur\altaffilmark{1}, Fabrizio Nicastro\altaffilmark{2,3,4}, Martin Elvis\altaffilmark{2}} \altaffiltext{1}{Department of Astronomy, The Ohio State University, 140 West 18th Avenue, Columbus OH 43210, USA} \altaffiltext{2}{Harvard--Smithsonian Center for Astrophysics, Cambridge, MA, 01238, USA} \altaffiltext{3}{Instituto de Astronom\'ia, Universidad Aut\'onomica de M\'exico, Apartado Postal 70-264, Ciudad Universitaria, M\'exico, D.F., CP 04510, M\'exico} \altaffiltext{4}{Osservatorio Astronomico di Roma, Istituto Nazionale di Astrofisica, Italy} \email{williams,smita@astronomy.ohio-state.edu} \keywords{ intergalactic medium --- X-rays: general --- cosmology: observations } \section{Introduction} Most of the baryons that comprise 4\% of the mass--energy budget of the universe are found in the intergalactic medium (IGM), primarily appearing as the Lyman--alpha ``forest'' in high--redshift quasar spectra \citep{weinberg97}. At more recent times ($z\la 2$) the process of structure formation has shock--heated the IGM to temperatures of $\sim 10^{5-7}$\,K, thus rendering the hydrogen nearly fully ionized and producing (at most) broad, extremely weak Lyman--alpha absorption \citep[e.g.][]{sembach04,richter04}. Known as the warm--hot intergalactic medium (WHIM), this phase is predicted to contain roughly half of the baryonic matter at low redshifts \citep{cen99,dave01}. Its extremely low density (typically $\delta\sim 10-100$) precludes the detection of WHIM thermal or line emission with current facilities, so the only way to directly measure these ``missing'' baryons is through far--ultraviolet and X-ray spectroscopic measurements of absorption lines from highly ionized heavy elements \citep{perna98,hellsten98,fangb02}. Several early attempts to detect these intervening WHIM absorbers in X-rays \citep[e.g.][]{fang02,mathur03,mckernan03} and more recent surveys \citep{fang05} yielded only tentative detections at best. Intervening Lyman--alpha \citep{shull96,sembach04} and \ion{O}{6} \citep[e.g.][]{savage02} absorbers had also been seen in \fuse\ and HST quasar spectra, but their ionization states and possible galactic halo origins \citep[e.g.][]{tumlinson05} are quite uncertain. These uncertainties are largely mitigated with the recent detection by \citet{nicastro05a,nicastro05b} of two X-ray absorption systems at $z=0.011$ and $z=0.027$ along the line of sight to the blazar Mrk 421. These filaments account for a critical density fraction of $\Omega_{\rm WHIM}=0.032^{+0.042}_{-0.021}$, fully consistent with the mass of the missing baryons in the local universe (albeit with large uncertainties). While future proposed missions such as \emph{Constellation--X}, \emph{XEUS}, or \emph{Pharos} (Nicastro et al., in preparation) will be able to measure $\Omega_{\rm WHIM}$ to far greater precision with detections of numerous weaker X-ray forest lines, the \citet[][hereafter N05]{nicastro05a} results present a key early confirmation of numerical predictions using observational capabilities that are \emph{currently} within our grasp --- hence any test of their correctness is of great importance. Although each of these two absorption systems was detected with high confidence through multiple redshifted X-ray absorption lines, the \emph{individual} absorption lines were generally quite weakly detected, mostly at the $2-4\sigma$ level. Moreover, while they employed high--quality \chandra\ and \fuse\ data taken during exceptionally bright outbursts of Mrk 421, the many archived \xmmnewton\ observations of this source were not included in the analysis. With roughly twice the effective area of \chandra/LETG, \xmm/RGS is in principle superior for X-ray grating spectroscopy between $\sim 10-40$\,\AA; however, its slightly worse resolution (approximately 60\,m\AA\ FWHM, versus 50\,m\AA\ for \chandra/LETG), higher susceptibility to background flares, and multitude of narrow instrumental features can present serious complications for WHIM searches. Independent confirmation of the \chandra\ results with a separate instrument like \xmm\ is thus important. While some groups have searched for WHIM features in a limited number of \xmm\ Mrk 421 spectra \citep[e.g.,][]{ravasio05}, a complete and systematic analysis has yet to be performed. Here we present a search for $z>0$ WHIM features employing all ``good'' observations of Mrk 421 available in the \xmm\ archive, and a comparison of these results to those presented by N05. \section{Data Reduction and Measurements} We searched the \xmm\ archive for all Mrk 421 Reflection Grating Spectrometer (RGS) data. Although 31 separate observations were available, 16 had pointing offsets $\Delta\theta \ga 60$\arcsec\ while the rest were offset by less than 15\arcsec. Since spectral resolution and calibration quality can degrade at large offsets, we only included those with $\Delta\theta < 15$\arcsec. One extremely short observation (0158971101, with $t_{\rm exp}=237$\,s) was also excluded to simplify the data reduction process. Using the standard \xmm\ Science Analysis System version 6.5.0 routines\footnote{See \url{http://xmm.vilspa.esa.es/sas}}, RGS1 light curves were built for the remaining 14 ``good'' observations (see Table~\ref{tab_log}), and the spectra were reprocessed to exclude periods of high background levels and coadded. These combined, filtered RGS1 and RGS2 spectra have effective exposure times of $\sim 440$\,ks and over $9\times 10^6$ combined RGS1$+$RGS2 first--order counts between $10-36$\,\AA\ with $\sim 15000$ counts per 0.06\,\AA\ resolution element in RGS1 near 21\,\AA, over twice that in the N05 Mrk~421 \chandra\ spectrum. We first used the spectral fitting program \emph{Sherpa}\footnote{\url{http://cxc.harvard.edu/sherpa/}} to fit a simple power law plus Galactic foreground absorption model to the RGS1 and RGS2 data; however, at such high spectral quality the RGS response model does not appear to be well--determined, and large residuals remained. For line measurements, we thus only considered $\sim 2$\,\AA\ windows around each wavelength of interest, using a power law to independently model the RGS1 and RGS2 continua within each window and excluding the strongest narrow detector features (with typical widths of 70\,m\AA\ or less). None of the intervening absorption lines were apparent through a visual inspection of the \xmm\ spectrum, though several of the $z=0$ lines reported by \citet{williams05} could be seen clearly. A narrow Gaussian absorption line (FWHM$=5$\,m\AA) was included in the model for each line measurement or upper limit reported by N05. When convolved with the RGS instrumental response these absorption lines appeared broadened to the RGS line spread function (LSF) widths \citep[typically FWHM$=60-70$\,m\AA;][]{denherder01}. The $2\sigma$ upper limits on all equivalent widths were then calculated (allowing the central line wavelengths to vary within the $1\sigma$ errors reported by N05). Since the shapes of the RGS1 and RGS2 instrumental responses are quite different, these limits were calculated using both a joint fit to the RGS1$+$RGS2 spectra as well as the individual RGS1 and RGS2 spectra. It should be noted that wherever one RGS unit is unusable, the total response is effectively halved, at which point it has a similar effective area to \chandra/LETG. The resulting equivalent width limits are listed in Table~\ref{tab_ew}. \section{Discussion} Figure~\ref{fig_xmmspec} shows the spectral windows used to determine upper limits on the N05 measured lines, with the data (black), continuum fit (blue), \chandra\ measurements and limits (N05; red solid and dotted lines respectively), and \xmm\ limits (green) overplotted. In all cases, the N05 measurements (or $3\sigma$ upper limits) appear to be consistent with the $2\sigma$ upper limits we have derived directly from the \xmm\ data, as shown in the figure and listed in Table~\ref{tab_ew}. The \ovii\ line at $z=0.027$ looks as though it might be visible in the spectrum, but this is most likely due to the weak instrumental feature at $\sim 22.1$\,\AA. For two lines (\nvii\ and \nvi\ at $z=0.027$) the \xmm\ $2\sigma$ upper limits are approximately equal to the N05 best--fit measurements, but since the N05 values are quite uncertain this result is still consistent. Why, then, with $2-4$ times the counts per resolution element, was \xmm\ unable to detect the intervening absorption systems seen by \chandra? Several factors appear to have been involved in this non--detection, primarily: (1) narrow instrumental features caused by bad detector columns, (2) the broader LSF as compared to \chandra/LETG, and (3) fixed--pattern noise at long wavelengths: \begin{enumerate} \item{While broad instrumental features can be taken into account by modifications to the continuum model (as in N05), it is difficult or impossible to distinguish narrow features from true absorption lines; thus, any line falling near one of the detector features shown in Figure~\ref{fig_rgsmod} can easily be lost\footnote{These response file data can be found at \url{http://www.astronomy.ohio-state.edu/$\sim$smita/xmmrsp/}}. This was responsible for the non-detection of the $z=0.011$ \ovii\ K$\alpha$ line. Although it was the strongest line reported by N05, its wavelength falls directly on a narrow RGS1 feature and within the non--functional CCD4 region on RGS2, thereby preventing this line from being detectable with either RGS. Since 18\% of the wavelength space for studying redshifted \ovii\ ($\lambda>21.6$\,\AA) toward Mrk 421 is directly blocked by these narrow features (with this number climbing to about 60\% if resolution elements immediately adjacent to bad columns are included), these bad columns present the single greatest hindrance to \xmm/RGS studies of the WHIM.} \item{Even for lines where both RGS1 and RGS2 data are available and the instrumental response appears to be relatively smooth, the lower resolution of \xmm\ contributes to the nondetectability of the weaker $z>0$ absorption lines. Figure~\ref{fig_lsfplot} shows the LSFs for both \xmm/RGS1 (solid) and \chandra/LETG assuming an unresolved line with $W_\lambda=10$\,m\AA\ at 21.6\,\AA. While the core of the RGS1 response is $\sim 20$\% broader than that of the LETG, the RGS1 LSF has extremely broad wings: only 68\% of the line flux is contained within the central 0.1\AA\ of the RGS1 LSF, as opposed to 96\% for the LETG. This reduces the apparent depth of absorption lines by about a factor of two as compared to \chandra/LETG, severely decreasing the line detectability.} \item{At long wavelengths ($\lambda\ga 29$\,\AA) strong fixed--pattern noise is apparent as a sawtooth pattern in the instrumental response, strongly impeding the detection of species such as \nvi\ and \cvi. Indeed, in these wavelength regimes (the lower two panels of Figure~\ref{fig_xmmspec}), the \nvi\ and \cvi\ absorption lines are nearly indistinguishable from the continuum.} \end{enumerate} \section{Conclusion} We have presented the highest signal--to--noise coadded \xmm\ grating spectrum of Mrk 421 to date, incorporating all available archival data. This spectrum serves as an independent check on the recent detection of two $z>0$ WHIM filaments by N05. While none of the \chandra--detected absorption lines are seen in the \xmm\ spectrum, the upper limits derived from the \xmm\ data are consistent with the equivalent widths reported by N05 (even though the \xmm\ data contain a larger number of counts), and hence do not jeopardize the validity of the \chandra\ measurement. The non--detections can be attributed primarily to narrow instrumental features in RGS1 and RGS2, as well as the inferior spectral resolution of \xmm\ and fixed--pattern noise at longer wavelengths. This underscores the extreme difficulty of detecting the WHIM, illustrates how the aforementioned (apparently small) effects can greatly affect the delicate measurement of weak absorption lines, and re--emphasizes the importance of high resolution and a smooth instrumental response function for current and future WHIM absorption line studies. \acknowledgments We thank the \xmm\ team for their efforts on this excellent mission, the helpdesk staff for their assistance with the data reduction, and the anonymous referee for helpful comments. This research is based on archival data obtained with XMM-Newton, an ESA science mission with instruments and contributions directly funded by ESA Member States and NASA. RJW is supported by an Ohio State University Presidential Fellowship, and FN acknowledges the support of NASA Long--Term Space Astrophysics Grant NNG04GD49G. \clearpage \begin{deluxetable}{lcccc} \tabletypesize{\footnotesize} \tablecolumns{5} \tablewidth{240pt} \tablecaption{\xmmnewton\ observation log \label{tab_log}} \tablehead{ \colhead{ID} & \colhead{Date} & \colhead{$t_{\rm exp}$\tablenotemark{a}} & \colhead{$t_{\rm filt}$\tablenotemark{b}} & \colhead{Rate\tablenotemark{c}} \\ \colhead{} & \colhead{} & \colhead{ks} & \colhead{ks} & \colhead{s$^{-1}$} } \startdata 0099280101 &2000 May 25 &63.8 &21.2 &15.7\\ 0099280201 &2000 Nov 01 &40.1 &34.1 &5.4\\ 0099280301 &2000 Nov 13 &49.8 &46.6 &15.3\\ 0099280501 &2000 Nov 13 &21.2 &17.8 &17.2\\ 0136540101 &2001 May 08 &38.8 &36.1 &11.7\\ 0136540301 &2002 Nov 04 &23.9 &20.5 &11.7\\ 0136540401 &2002 Nov 04 &23.9 &20.1 &13.6\\ 0136540701 &2002 Nov 14 &71.5 &62.8 &16.4\\ 0136541001 &2002 Dec 01 &70.0 &58.1 &8.3\\ 0158970101 &2003 Jun 01 &43.0 &25.3 &9.0\\ 0158970201 &2003 Jun 02 &9.0 &6.6 &9.7\\ 0158970701 &2003 Jun 07 &48.9 &29.9 &5.4\\ 0158971201 &2004 May 06 &65.7 &40.5 &19.5\\ 0162960101 &2003 Dec 10 &30.0 &17.5 &9.8\\ \hline\hline TOTAL & &572.3 &437.1 &12.2\\ \enddata \tablenotetext{a}{Total observation duration.} \tablenotetext{b}{Effective RGS1 exposure time after filtering for periods of high background levels.} \tablenotetext{c}{Average count rate in the filtered RGS1 first--order source spectral extraction region.} \end{deluxetable} \clearpage \begin{deluxetable}{lccccccc} \tabletypesize{\footnotesize} \tablecolumns{8} \tablecaption{Absorption line equivalent width measurements \label{tab_ew}} \tablehead{ \colhead{Line} & \colhead{$\lambda$\tablenotemark{a}} & \colhead{$z$\tablenotemark{a}} & \colhead{$W_{\lambda, {\rm N05a}}$\tablenotemark{a}} & \colhead{$W_{\lambda, {\rm R1}}$\tablenotemark{b}} & \colhead{$W_{\lambda, {\rm R2}}$\tablenotemark{b}} & \colhead{$W_{\lambda, {\rm R1+R2}}$\tablenotemark{b}} & \colhead{Note} \\ \colhead{} & \colhead{\AA} & \colhead{} & \colhead{m\AA} & \colhead{m\AA} & \colhead{m\AA} & \colhead{m\AA} & \colhead{} } \startdata \ion{Ne}{9}$_{K\alpha}$ &$13.80\pm 0.02$ &$0.026\pm 0.001$ &$<1.5$ &$<5.2$ &$<1.9$ &$<2.9$ &1\\ \ion{O}{7}$_{K\beta}$ &$19.11\pm 0.02$ &$0.026\pm 0.001$ &$<1.8$ &$<2.5$ &$<2.1$ &$<1.5$ &\\ \ion{O}{8}$_{K\alpha}$ &$19.18\pm 0.02$ &$0.011\pm 0.001$ &$<4.1$ &$<7.6$ &$<5.8$ &$<4.1$ &\\ \ion{O}{8}$_{K\alpha}$ &$19.48\pm 0.02$ &$0.027\pm 0.001$ &$<1.8$ &\nodata &$<3.9$ &\nodata &2\\ \ion{O}{7}$_{K\alpha}$ &$21.85\pm 0.02$ &$0.011\pm 0.001$ &$3.0^{+0.9}_{-0.8}$ &\nodata &\nodata &\nodata &2,3\\ \ion{O}{7}$_{K\alpha}$ &$22.20\pm 0.02$ &$0.028\pm 0.011$ &$2.2\pm 0.8$ &$<3.9$ &\nodata &\nodata &3\\ \ion{N}{7}$_{K\alpha}$ &$25.04\pm 0.02$ &$0.010\pm 0.001$ &$1.8\pm 0.9$ &$<3.0$ &$<6.0$ &$<4.4$ &\\ \ion{N}{7}$_{K\alpha}$ &$25.44\pm 0.02$ &$0.027\pm 0.001$ &$3.4\pm 1.1$ &$<4.3$ &$<4.2$ &$<3.5$\\ \ion{N}{6}$_{K\alpha}$ &$29.54\pm 0.02$ &$0.026\pm 0.001$ &$3.6\pm 1.2$ &$<3.8$ &$<8.7$ &$<3.4$ &\\ \ion{C}{6}$_{K\alpha}$ &$34.69\pm 0.02$ &$0.028\pm 0.001$ &$2.4\pm 1.3$ &$<5.5$ &$<5.2$ &$<4.2$ &\\ \enddata \tablenotetext{a}{\ Line wavelength, redshift, and equivalent width measurements (or $3\sigma$ upper limits) from \citet{nicastro05a}.} \tablenotetext{b}{$2\sigma$ equivalent width upper limits measured from the RGS1 only (R1), RGS2 only (R2), and joint (R1+R2) fits to the XMM--Newton spectrum, when available.} \tablecomments{ (1) A nearby chip gap in RGS1 renders this measurement unreliable, so only the RGS2 measurement was used in Figure~\ref{fig_xmmspec}; (2) Line was unmeasurable in RGS1 because of a detector feature; (3) Line was unmeasurable in RGS2 because of a detector feature. } \end{deluxetable}
Title: {Interstellar Plasma Weather Effects in Long-term Multi-frequency Timing of Pulsar B1937+21
Abstract: We report here on variable propagation effects in over twenty years of multi-frequency timing analysis of pulsar PSR B1937+21 that determine small-scale properties of the intervening plasma as it drifts through the sight line. The phase structure function derived from the dispersion measure variations is in remarkable agreement with that expected from the Kolmogorov spectrum, with a power law index of $3.66\pm 0.04$, valid over an inferred scale range of 0.2--50 A.U. The observed flux variation time scale and the modulation index, along with their frequency dependence, are discrepant with the values expected from a Kolmogorov spectrum with infinitismally small inner scale cutoff, suggesting a caustic-dominated regime of interstellar optics. This implies an inner scale cutoff to the spectrum of $\sim 1.3\times 10^9$ meters. Our timing solutions indicate a transverse velocity of 9 km sec$^{-1}$ with respect to the solar system barycenter, and 80 km sec$^{-1}$ with respect to the pulsar's LSR. We interpret the frequency dependent variations of DM as a result of the apparent angular broadening of the source, which is a sensitive function of frequency ($\propto\nu^{-2.2}$). The error introduced by this in timing this pulsar is $\sim$2.2 $\mu$s at 1 GHz. The timing error introduced by ``image wandering'' from the slow, nominally refractive scintillation effects is about 125 nanosec at 1 GHz. The error accumulated due to positional error (due to image wandering) in solar system barycentric corrections is about 85 nanosec at 1 GHz.
https://export.arxiv.org/pdf/astro-ph/0601242
\title{Interstellar Plasma Weather Effects in Long-term Multi-frequency \\ Timing of Pulsar B1937+21} \author{R. Ramachandran, P. Demorest, D. C. Backer} \affil{Department of Astronomy and Radio Astronomy Laboratory, University of California, Berkeley, CA 94720-3411, USA; \\ e-mail: ramach, demorest, dbacker@astro.berkeley.edu} \author{I. Cognard} \affil{Laboratoire de Physique et Chimie de l'Environnement, CNRS, 3A avenue de la Recherche Scientifique, F-45071 Orleans, France} \author{A. Lommen} \affil{Department of Physics and Astronomy, Franklin and Marshall College, P.O.Box 3003, Lancaster, PA 17604, USA} \keywords{ISM: general --- pulsars: general --- radio continuum: general --- scattering --- turbulence} \section{Introduction} The dispersion measure (DM) of a pulsar probes the column density of free electrons along the line of sight (LOS). Observed DM variations over time scales of several weeks to years sample structures in the electron plasma over length scales of $10^{10}$ m to $10^{12}$ m. Diffraction of pulsar signals is the result of scattering by structures on scales below the Fresnel radius, $10^{8}$ m or so. The DM as well as the scattering measure (SM) variability along the LOS to the Crab pulsar was first reported by Rankin \& Isaacman (1977), who reported that the DM variability poorly correlated with the SM variability. Helfand et al. (1980) inferred an upper limit for DM variations of a few parts in a thousand for several pulsars. In an earlier study of PSR B1937+21 Cordes et al. (1990) measured a DM change of $\Delta DM\sim 0.003$ pc cm$^{-3}$ over a period of a thousand days. The work of Phillips \& Wolszczan (1991) presented the variations of DM observed along the LOS to a few pulsars. They connected these variations to those on diffractive scales, and derived an electron density fluctuation spectrum slope of $3.85\pm0.04$ over a scale range of $10^7-10^{13}$ meters. Backer et al. (1993) report on further DM variability and show that the amplitude of the variations known at that time are consistent with a scaling by the square root of DM. Another important investigation by Kaspi et al. (1994) studied DM variations of the millisecond pulsars PSR B1937+21 and B1855+09 over a time interval of calendar years 1984 to 1993. In addition to establishing a secular variation in DM over this time interval, they also show that the underlying density power spectrum has an index of 3.874$\pm$0.011, which is close to what we would expect if the density fluctuations are described by Kolmogorov spectrum. An anomalous dispersion event towards the Crab pulsar was reported by Backer et al. (2000), where they report a DM ``jump'' as large as 0.1 pc cm$^{-3}$. In this work, we present results of various long term monitoring programs on PSR B1937+21. Our data, which includes that of Kaspi, Taylor \& Ryba (1994), spans calendar years from 1983 to 2004. These data sets have been taken with five different telescopes, the NRAO\footnote{The National Radio Astronomy Observatory (NRAO) is owned and operated by Associated Universities, Inc under contract with the US National Science Foundation.} Green Bank 42-m (140-ft) and 26-m (85-ft) telescopes, the NAIC\footnote{The National Astronomy and Ionosphere Center is operated by Cornell University under contract with the US National Science Foundation.} Arecibo telescope and the \nancay ~telescope, at frequency bands of 327, 610, 800, 1400 and 2200 MHz. After giving the details of our observations in \S\ref{sec-obsvn}, we describe our analysis methods in \S\ref{sec-anal}. This is followed in \S\ref{sec-sightline} by a discussion on the distribution of scattering material along the LOS. As we describe, the knowledge of temporal and angular broadening of the source, proper motion, and scintillation based velocity estimates enables us to at least qualitatively study the distribution of scattering matter as well as properties of its wave-number spectrum. We have measured some of the basic refractive scintillation parameters from our observations, and these are discussed in \S\ref{sec-diffref}. The frequency dependence of the refractive scintillation time scale and the modulation index indicate a caustic-dominated regime that results from a large inner scale in the spectrum. We have detected DM variations as a function of time and frequency. We determine the phase structure function of the medium with the knowledge of the time dependent DM variations, which is consistent with a Kolmogorov distribution of density fluctuations between scale sizes of about 1 to 100 A.U. These are summarized in \S\ref{sec-dmvar} and \S\ref{sec-dmfreq}. \begin{table} \begin{center} \caption[]{Template fit parameters at various frequencies.} \begin{tabular}{lcccccccc} \hline {\bf $\nu$ (MHz)} & backend & ~~$w_1$~~ & ~~$l_2$~~ & ~~$w_2$~~ & ~~$h_2$~~ & ~~$l_3$~~ & ~~$w_3$~~ & ~~$h_3$~~ \\ \hline\hline 327 & GBPP & 8.7 & 0 & 0 & 0 & 186.9 & 12.0 & 0.51 \\ 430 & ABPP & 10.3 & 7.0 & 2.5 & 0.19 & 186.8 & 12.4 & 0.53 \\ 610 & GBPP & 9.4 & 7.1 & 2.9 & 0.34 & 187.3 & 10.8 & 0.60 \\ 863 & EBPP & 8.9 & 7.7 & 5.2 & 0.38 & 187.7 & 10.9 & 0.55 \\ 1000 & GBPP & 9.2 & 8.6 & 3.9 & 0.56 & 187.9 & 11.3 & 0.54 \\ 1419 & ABPP & 8.5 & 8.5 & 3.7 & 0.37 & 187.5 & 10.1 & 0.45 \\ 1689 & EBPP & 9.1 & 9.2 & 3.9 & 0.37 & 187.4 & 10.5 & 0.40 \\ 2200 & GBPP & 9.4 & 9.8 & 3.2 & 0.29 & 187.6 & 10.4 & 0.36 \\ 2379 & ABPP & 10.0 & 9.8 & 3.0 & 0.29 & 187.1 & 10.8 & 0.33 \\ \hline \end{tabular} \label{tab:fitpars} \end{center} \end{table} PSR B1937+21 is known for its short term timing stability. However, the achievable long term timing accuracy is suspected to be seriously limited by the interstellar scattering properties. With our sensitive measurements, we are in a position to quantify these errors. In \S\ref{sec-timingerror}, we describe in detail various sources of these errors and quantify them. \section{Observations } \label{sec-obsvn} We have used five different primary data sets for this analysis. The first set is the 1984--1992 Arecibo pulse timing and dispersion measurements obtained by Kaspi et al. (1994; hereafter KTR94). Their observations were performed with their Mark~II backend (Rawley 1986; Rawley, Taylor \& Davis 1988) and later their Mark~III backend (Stinebring et al. 1992) at two different radio frequency bands, 1420 MHz and 2200 MHz. Their analysis methods are described in KTR94. The second data set is from 800-MHz and 1400-MHz observations at the NRAO 140-ft telescope in Green Bank, WV. The Spectral Processor backend, a hardware FFT device, was used. Details about the observations and analysis are contained in an earlier report on dispersion measure variability (Backer et al. 1993). The third data set consists of observations at 327 MHz and 610 MHz using the 26m (85-ft) pulsar monitoring telescope at NRAO's Green Bank, WV site. Room temperature (uncooled) receivers at the two bands are mounted off-axis. At 327 MHz the total bandwidth used was 5.5 MHz, and 16 MHz was used at 610 MHz. The two orthogonally polarized signals were split into 32 frequency channels in a hybrid analog/digital filter bank in the GBPP (Green Bank--Berkeley Pulsar Processor). Dispersion effects were removed in the GBPP in real-time with a coherent (voltage) deconvolution algorithm. At the end of the real-time processing folded pulse profiles were recorded for each frequency channel and polarization. Further details of the backend and analysis can be found in Backer, Wong \& Valanju (2000). PSR B1937+21 was observed for about two hours per day starting in mid-1995. The fourth data set comes from a bi-monthly precision timing program that includes B1937+21 at the Arecibo Observatory which we started in 1999 after the telescope upgrade. Signals at 1420 MHz and 2200 MHz were recorded using the ABPP backend (Arecibo--Berkeley Pulsar Processor), which is identical to the GBPP. Our typical observing sessions at 1420 MHz and 2200 MHz had bandwidths of 45 MHz and 56 MHz, respectively, and integration times of approximately 10 minutes per session. \begin{table} \tt \begin{center} \caption[]{Parameters of PSR B1937+21.} \begin{tabular}{ll} \hline {\bf Parameter} & {\bf value} \\ \hline\hline PSR & 1937+21 \\ RAJ (hh:mm:ss) & 19:39:38.561 (1) \\ DECJ (dd:mm:ss) & 21:34:59.136 (6) \\ PMRA (mas yr$^{-1}$) & 0.04 (20) \\ PMDEC (mas yr$^{-1}$) & -0.45 (6) \\ $f$ (Hz) & 641.9282626021 (1) \\ $\dot{f}_{-15}$ (Hz s$^{-1}$) & -43.3170 (6) \\ $\ddot{f}_{-26}$ (Hz s$^{-2}$) & 1.5 (3) \\ PEPOCH (MJD) & 47500.000000 \\ START & 45985.943 \\ FINISH & 52795.286 \\ EPHEM & DE405 \\ CLK & UTC (NIST) \\ \hline \end{tabular} \label{tab:1937par} \end{center} \end{table} The fifth data set is from a pulsar timing program that has been ongoing since 1989 October with the large decimetric radio telescope located at \nancay, France. The Nan\c cay telescope has a surface area of 7000 m$^2$, which provides a telescope gain of 1.6 K Jy$^{-1}$. Observations are performed with dual-linear feeds at frequencies 1280, 1680 and 1700 MHz. Then the signal is dedispersed by using a swept frequency oscillator (at 80 MHz) in the receiver IF chain. The pulse spectra are produced by a digital autocorrelator with a frequency resolution of 6.25 kHz. Cognard et al. (1995) describe in detail the backend setup and the analysis procedure. A small amount of additional data from the Effelsberg telescope was used in our profile analysis. At Effelsberg the EBPP backend, a copy of the GBPP/ABPP, was used. \section{Basic analysis } \label{sec-anal} We first present several results from the analysis of these data sets: a description of the frequency-dependent profile template used for timing; spin and astrometric timing parameters from high frequency data; pulse broadening, flux densities and dispersion measure as functions of time. In \S\ref{sec-sightline} we proceed to interpret these results and return to finer details regarding dispersion measure variations in \S\ref{sec-dmvar}. Our basic data set consists of average pulse profiles obtained approximately every 5 minutes in each of the radio frequency bands -- 327, 610, 800, 1420 and 2200 MHz. Figure \ref{fig:obs_summary} provides a graphical summary of observation epochs vs date. For data sets corresponding to all frequencies except 327 MHz, {\it Times of Arrival} (TOAs) were computed by cross correlating these average profiles with a template profile. The template profile at a given frequency was made by using multiple Gaussian fits to very high signal to noise ratio average profiles at that frequency; the interactive program {\it bfit}, which is based on M. Kramer's original program {\it fit} was used. These fit parameters are listed in Table 1. Col.~1 in the Table gives the radio frequency and the backend name is in col.~2. Col.~3 gives the width of component 1 ($w_1$; its location is taken to be 0 degrees and its amplitude is set to 1.0); cols.~4-6 and cols.~7-9 give the location ($l$), width ($w$) and amplitude ($h$) values for components 2 and 3, respectively. The location and width are given in units of longitudinal degrees, where 360$^{\circ}$ indicates one full rotation cycle. The results of this analysis can be compared with that of Foster et al. (1991) which are given on the line at 1000 MHz \footnote{The widths $w_1$ and $w_3$ are inverted in Table 4 of Foster et al.}. There is reasonable agreement for all values except $h_2$ which must have been erroneously entered in Table 4 of Foster et al. In our analysis templates corresponding to arbitrary frequencies are produced by spline-interpolation of the component parameters. We used the Arecibo (1420 MHz and 2200 MHz) TOAs, and the GBT 140-ft (800 MHz and 1420 MHz) TOAs to fit for pulsar spin (rotation frequency ($f$), first time derivative ($\dot{f}$), and second time derivative ($\ddot{f}$)) and astrometric (position ({\tt RAJ}, {\tt DECJ}), proper motion ({\tt PMRA}, $\mu_{\alpha}$, along right acension, and {\tt PMDEC}, $\mu_{\delta}$, along declination) parameters. All TOAs were referred to the UTC time scale kept by the National Institute of Standards and Technology (NIST) via GPS satellite comparison. We removed the effects of variable dispersion from this fitting procedure with weekly estimation of DMs and subsequent extrapolation of the dual frequency data to infinite frequency prior to parameter estimation. The nature of achromatic timing noise makes it particularly difficult to determine a precise timing model. As one adds additional higher derivatives of rotation frequency (e.g., a third derivative), the best fit parameters change by amounts much larger than the nominal errors reported by the package that we used, TEMPO. The results are listed in Table 2. The errors presented in the table incorporates the range of variation of each parameter, as additional derivative terms are included. In comparison to Kaspi, Taylor \& Ryba (1994), the derived proper motion values are marginally different. We attribute this difference to the variable influence of timing noise. An important point that needs to be stressed here is that there is no reason for us to assume that the higher derivative terms of rotation period (e.g., $\ddot{f}$ or higher) has anything to do with the radiative braking index. They are most likely dominated by some intrinsic instabilities of the star itself, or some other perturbation on the star. Extension of dispersion measurement to 327 MHz requires removal of the time-variable broadening of the intrinsic pulse profile owing to multipath propagation in the interstellar medium. We deconvolved the effect of interstellar scattering following precepts first introduced by Rankin et al. (1970). We assume that the interstellar temporal broadening is quantified in terms of convolution of a Gaussian function and a truncated exponential function. If there is only one scattering screen along the LOS, the assumption of a truncated exponential function will suffice to represent the scatter broadening. However, since the scattering may arise from material distributed all along the LOS, a more realistic representation is approximated by a truncated exponential function ``smoothed'' (convolved) with a Gaussian function. The intrinsic pulse profile was estimated by extrapolation of parameters from the higher frequency profiles. In the deconvolution procedure, we minimized the normalized $\chi^2$ value by varying the width of the Gaussian $w_g$ and the decay time scale of the truncated exponential $\tau_e$, while keeping the intrinsic pulse profile fixed. The pulse scatter broadening is quantified as $\tausc\; = \; (w_g^2 + \tau_e^2)^{1/2}$. We repeated this for average profiles obtained at every epoch to obtain the $\tausc$ measurement. In our fits, the average value of $w_g$ came to about 74 $\mu$sec, whereas the corresponding value for $\tau_e$ was about 85 $\mu$sec. The measurement of $\tausc$ versus time at 327~MHz is plotted in the Figure \ref{fig:tausc}. This quantity has a mean value of 120~$\mu$s, an RMS variation of 20~$\mu$s, and a fluctuation timescale of $\sim60$~days. We explain these variations as the result of refractive modulation of this inherently diffractive parameter in discussion below. The estimated RMS variation at the next higher frequency in our data set, 610 MHz, is $\sim$2.5 $\mu$s, using a frequency dependence of $\tausc\propto\nu^{-4.4}$. This is too small to allow fitting at this frequency band. In the strong scintillation regime, time dependent variations in the observed flux occur in two distinct regimes --- {\it diffractive} and {\it refractive}. The diffractive effects are dominated by structures smaller than the Fresnel scale, and appear on short time scales and over narrow bandwidths. In our observations diffractive modulations are strongly suppressed. On the other hand refractive effects occur on days time scales and are correlated over wide bandwidths. We have analyzed our best data sets -- the densely sampled data at 327 MHz and 610 MHz from Green Bank and at 1410 MHz from Nan\c cay -- for flux density variations as a function of time. The data are presented in Figure \ref{fig:fluxplot}. In analyzing the low frequency flux data from Green Bank, we have not adopted a rigorous flux calibration procedure. While there is a pulsed calibration noise source installed in this system, equipment changes and the nature of the automated observing have led to large gaps in the calibration record. Rather than dealing with a mix of calibrated and uncalibrated data, or lose a large fraction of the data, we decided not to apply any calibration. Instead, we normalize our data by assuming the system temperature is constant. In order to see what effect this has on our results, we did two tests. First, we analyzed observations of PSR~B1641$-$45, taken with the same system, over a similar time range. This pulsar is known to have a very long refractive timescale, $\tref >1800$~days (Kaspi \& Stinebring, 1992), so it can be used as a flux calibrator. In our data, we find it to have a modulation index, $m=0.10$. This immediately puts a upper limit of 10\% on any systematic gain and/or system temperature variations. Since modulation adds in quadrature, and we observe modulation indeces of $m\sim0.4$ for PSR~B1937+21, gain fluctuations represent at most a small fraction of the observed modulation. We also considered the possibility that gain variations could influence our measurement of $\tref $. This might happen if they occur with a characteristic timescale longer than 1~day. In order to test this, we analyzed observations of the Crab pulsar, PSR~B0531+21, again taken with the same system over the same time range. The refractive parameters of this pulsar were studied in detail by Rickett \& Lyne (1990). It makes a good comparison since it has modulation index of $m=0.4$ at 610~MHz, very similar to PSR~B1937+21. Applying the structure function analysis (see \S\ref{sec-diffref}) to this data gives $\tref =11$~days at 610~MHz, and $\tref =63$~days at 327~MHz, consistent with the previously published results and a scaling law of $\tref \propto\nu^{-2.2}$. The procedure that we have adopted to calibrate our data set from \nancay~ telescope is described in detail in Cognard et al. (1995). \section{Distribution of scattering material along the line of sight} \label{sec-sightline} Several authors have shown how the scattering parameters of a pulsar can be used to assess the distribution of scattering material along the LOS (Gwinn et al. 1993; Deshpande \& Ramachandran 1998; Cordes \& Rickett 1998). This results from the varied dependences of the scattering parameters on the fractional distance of scattering material along the LOS. PSR B1937+21 is viewed through the local spiral arm as well as the Sagittarius arm which are both potential sites of strong scattering. The parameters employed in this analysis are: the temporal pulse broadening by scattering ($\tausc$; or its conjugate parameter $\Delta\nu$, the diffractive scintillation bandwidth), the diffractive scintillation time scale ($\tdiff$), the angular broadening from scattering ($\theta_H$), the proper motion of the pulsar ($\mu_{\alpha}$, $\mu_{\delta}$), and a distance estimate of the pulsar ($D$). Let us first compare $\theta_H$ and $\tausc$ that are the result of multiple scattering along the LOS, and express them as (Blandford \& Narayan 1985) \begin{eqnarray} \tausc &=& \frac{1}{2cD}\int_0^D x(D-x)\;\psi(x)\; {\rm d}x \\ \theta_H^2 &=& \frac{4\ln2}{D^2}\int_0^D x^2\psi(x)\; {\rm d}x. \label{eq:tautheta} \end{eqnarray} In these equations, $x$ is the coordinate along the LOS, with the pulsar at $x=0$ and the observer at $x=D$. $\psi(x)$ is the mean scattering rate. If the scattering material is uniformly distributed along the LOS, then the relation between the two quantities can be expressed as $\theta_H^2=16\ln2\left(c\tausc/D\right)$. With the distance to the pulsar of 3.6~kpc to the pulsar (according to the distance model of Cordes \& Lazio 2002), and the average pulse broadening time scale of 120 $\mu$s (from the present work), we obtain an estimate of the angular broadening, $\theta_{\tau}$, of 12 mas. This is in modest agreement with the measured value of 14.6$\pm 1.8$ mas (Gwinn et al. 1993), given the uncertainty in the distance to the pulsar and the simple assumption that the scattering material is uniformly distributed along the LOS. Next, we formulate two approaches to estimation of the velocity of the LOS with respect to the scattering medium, and use these approaches to assess the location and extent of the medium. The transverse velocity of the pulsar based on the measured proper motion (Table 2) an assumed distance of $D=3.6$ kpc (Cordes \& Lazio 2002) is 9 km s$^{-1}$. This value is the velocity of the pulsar with respect to the solar system barycenter. With the assumed ``Flat Rotation Curve'' linear velocity of the Galaxy of 225 km s$^{-1}$, and the Sun's peculiar velocity of 15.6 km s$^{-1}$ in the Galactic coordinate direction of $(l,b) = (48.8^{\circ}, 26.3^{\circ})$ (Murray 1986), the transverse velocity of the pulsar in its LSR ($V_p$) is 80 km s$^{-1}$. The {\it scintillation velocity} ($\viss$), which is an estimate of the velocity of the {\it diffraction pattern} at the location of the Earth, is estimated from the decorrelation bandwidth ($\Delta\nu$) and the diffractive scintillation time scale ($\tdiff$). Gupta et al. (1994) conclude that \begin{equation} \viss\;=\;3.85\times 10^4\; \sqrt{\frac{D \; z \; \Delta \nu}{(1-z)}} \; \frac{1}{\tdiff\;\nu_{\rm GHz}} \;\;\; {\rm km\;s^{-1}} \end{equation} where $\nu_{\rm GHz}$ is the observing frequency in GHz, $D$ is in kpc, $\Delta\nu$ is in MHz, and $\tdiff$ is in seconds. The variable $z$ gives the fractional distance to the scattering screen, where $z=0$ gives the observer's position, and $z=1$ gives the pulsar's position. The value of decorrelation bandwidth is computed by the relation $\Delta \nu \; =\; 1 / (2 \pi \tausc)$. When the effective scattering screen is midway along the LOS ($z=0.5$), $\viss\;=\;V_p$, and when the screen is at the location of the pulsar ($z=1.0$), $\viss = \infty$. While doing this, an important assumption is that the pulsar proper motion is dominant over contributions from differential Galactic motion, solar peculiar velocity, and the Earth's annual orbital modulation. In the case of PSR B1937+21, this assumption is not justified. The effective scattering screen, which is located somewhere along the LOS, has a Galactic motion whose component along the LOS direction is different from that of the pulsar or the Sun. In order to correct for this effect, let us calculate the LOS velocity across the effective scattering screen at a fractional distance $z$ from the observer: \begin{equation} \vlos\;=\; 3.85\times 10^4\; \frac{\sqrt{D \; z \; (1-z)\; \Delta \nu}}{\tdiff\;\nu_{\rm GHz}} \;\;\; {\rm km\;s^{-1}} \label{eq:vlos} \end{equation} Then, let us assume that the scattering along the LOS can be adequately expressed by having a thin screen alone, at a distance of $D_s=zD$ from the observer. Then, Equation \ref{eq:tautheta} can be expressed as \begin{eqnarray} \tausc &=& \frac{\psi_{\circ}}{2c}\; D\;z \; (1-z) \\ \theta_H^2 &=& 4\;\ln 2\; (1-z)^2 \psi_{\circ} \end{eqnarray} Here, $\psi_{\circ}$ gives the mean scattering rate corresponding to the effective thin screen. Then, let us express independently the transverse velocity of the LOS across the scattering screen as \begin{eqnarray} \vec{V}_{\perp}' &=& (1-z)\;\vec{V}_e\; + \; z\vec{V}_p \;-\; \vec{V}_G (zD\hat{n}) \nonumber\\ &=& \vec{V}_E + zD\vec{\mu} - \vec{V}_G (zD\hat{n}), \label{eq:vlos2} \end{eqnarray} where $V_e$ is Earth's orbital velocity, $V_p$ is the pulsar transverse velocity in its LSR, and $V_G$ is the transverse velocity contribution from the Galactic differential motion to the screen. $V_E$ gives the contribution of the Earth's motion on the LOS velocity across the screen. Equations \ref{eq:vlos} and \ref{eq:vlos2} give two independent estimates of the line of sight velocity across the effective scattering screen and therefore allow us to solve for the value of $z$ given $D$. With $D=3.6$ kpc (Cordes \& Lazio 2002), we find $z=0.7$. The LOS velocity is 51 km s$^{-1}$. The assumed value of $\tdiff$ is 78 seconds at 327 MHz (scaled from 444 seconds at 1400 MHz of Cordes et al. 1991), and the value of $\Delta\nu$ is 0.0013 MHz calculated from $\tausc=120$ $\mu$s. To summarize, the measured value of $\theta_H=14.6\pm 1.8$ mas and the estimated value of $\theta_{\tau}$ are consistent with each other, suggesting a uniformly distributed scattering medium. On the other hand, comparison of velocity components, $V_p$, $\vlos$ and $V_{\rm los}^\prime$ suggest a thin-screen at $z\sim 2/3$. As Deshpande \& Ramachandran (1998) demonstrate, this solution is equivalent to having a uniformly distributed scattering medium! Therefore, we conclude that the line of sight to PSR B1937+21 can be described adequately by a uniformly distributed scattering matter. The Earth's orbital velocity around the Sun will modulate the observed scintillation speed, and therefore the diffractive scintillation time scale, with a one-year time scale. The amplitude of this modulation will depend on the effective $z$ of the diffracting material, and so monitoring could provide an estimate of the effective screen location. If the effective screen is close to the Earth, then the modulation is strong, and if it is located close to the pulsar, then it is negligible. Figure \ref{fig:annualmod} demonstrates this effect. The ordinate and abscissa give the LOS velocity across the effective scattering screen along the galactic longitude and latitude, respectively. For an assumed distance of 3.6 kpc, the straight line shows the expected centroid velocity of $\vlos^\prime$. The left most end of the line (origin of the plot) corresponds to $z=0$, and the right most end corresponds to $z=1$. The annual modulation of $\vlos^\prime$, shown as the two ellipses, correspond to z=0.5 and $z=2/3$. We have no way of identifying this annual modulation in our data, as we are insensitive to diffractive effects in our data set. Another measurement that could help us is the direct measurement of distance to this object by parallax measurements. Chatterjee et al. (2005, private communication), from their preliminary Very Long Baseline Array (VLBA) based parallax measurements, report that the distance to PSR B1937+21 is $2.3^{+0.8}_{-0.5}$ kpc, if they force the proper motion value to be the same as that of our timing based measurements (Table 2). In the coming year, accuracy of their measurements will improve with further sensitive observations. \section{Refractive scintillation } \label{sec-diffref} \subsection{Parameter estimation} \label{sec-diffrefparm} We determine refractive scintillation parameters from the data presented in Figure \ref{fig:fluxplot} following the structure function approach in previous studies (Stinebring et al 2000; Kaspi \& Stinebring; Rickett \& Lyne 1990). We define the structure function $D_F$ for flux time series $F(t)$ as \begin{equation} D_F(\delta t) = \frac{\langle [F(t) - F(t+\delta t)]^2\rangle}{\langle F(t) \rangle^2}, \label{eq:structflux} \end{equation} where $\delta t$ is a time delay. Since our flux measurments occur at discrete and unevenly spaced time interals, we compute the flux difference for all possible lags, then average results into logarithmically spaced bins. The flux structure function typically has a form described by Kaspi \& Stinebring et al. (1992) - a flat, noise dominated section at small lags, then a power-law increase which finally saturates at a value $D_s$ at large lags. In practice, the saturation regime may have large ripples in it, an effect of the finite length of any data set. In addition, the measured flux structure function is offset from the ``true'' flux structure function due to a contribution from uncorrelated measurement errors. At 327~MHz and 610~MHz, we estimate this noise term from the short-lag (noise regime) values. At 1410~MHz (from \nancay), we use the individual flux error bars to get the noise level. After subtracting the noise value, we fit the result to a function of the form \begin{equation} D_F(\delta t) = \left\{ \begin{array}{ll} D_s(\delta t/T_s)^\alpha, & [0<\delta t<T_s] \\ D_s, & [\delta t > T_s] \end{array} \right. \label{eqn:sffit} \end{equation} In this fit, the power law slope $\alpha$, the saturation timescale $T_s$, and the saturation value $D_s$ are all free parameters. The flux structure function data and fits are shown in Figure \ref{fig:fluxstruct}. As shown in Rickett \& Lyne (1990), the refractive parameters can be measured from the flux structure function using the following relationships: The modulation index $m$ is given by $m = \sqrt{D_s/2}$, and the refractive scintillation timescale $\tref$ is given by $D_F(\tref) = D_s/2$. All the measured parameters, including those measured by earlier investigators are summarized in Table 3. Based on a propagation model through a simple power-law density fluctuation spectrum, we expect to see refractive variations in the flux measurements on a timescale $\tref\sim0.5\theta_H D/\vlos$, where $\vlos$ is the line of sight velocity across the effective scattering screen. For the argument sake, if we assume an effective scattering screen at $z=0.5$, then $\vlos\sim 40$ km sec$^{-1}$. With $\theta_H = 14.6$ mas, the expected refractive scintillation time scale is $\sim$3 years at 327 MHz. This is more than an order of magnitude in excess of the measured value. Furthermore, if the density fluctuations in the medium are distributed according to the Kolmogorov power law distribution, then the expected frequency scaling law is $\tref\propto\lambda^{2.2}$. Our measured values indicate a significantly different scaling. Although it is consistent with $\tref\propto\lambda^{2.2}$ between 610 and 1420 MHz, it is not so between 327 and 610 MHz, where it is consistent with being directly proportional to $\lambda$. Our observed modulation index ($m$) values are also considerably larger than predicted, and show a ``flatter'' wavelength dependence, as listed in Table 3. We will address this issue in detail in the following section. \subsection{Nature of the spectrum -- inner scale cutoff} \label{sec-cutoff} The three disagreements with a simple model summarized in \S\ref{sec-diffrefparm} force us to explore a few aspects of the electron density power spectrum that may possibly explain what we observe. The effects of {\it caustics} on the observed scintillations have been explored by several earlier investigators, most notably Goodman et al. (1987) and Blandford \& Narayan (1985). In particular, if the power law scale distribution in the medium is truncated at an inner scale that is considerably larger than the diffractive scale, as they show, the observed flux variations are dominated by caustics. This is of great interest to us, as this seems to explain all the discrepancies that we note in our observed refractive parameters. For instance, as Goodman et al. (1987) show that if the inner scale cutoff is a considerable fraction of the Fresnel scale, then the observed fluctuation spectrum of flux is dominated by fluctuation frequencies that are lower than the diffractive frequencies, but significantly higher than that expected from refractive scintillation. This is what we observe. Moreover, as they note, the observed wavelength dependence of the refractive time scale, as well as the modulation index is expected to be ``shallower'' than the expected values of $\lambda^{2.2}$ and $\lambda^{-0.57}$, respectively. A ``shallow'' frequency dependence of the modulation index has been reported by others (Coles et al. 1987; Kaspi \& Stinebring 1992; Gupta et al. 1993; Stinebring et al. 2000). While Kaspi \& Stinebring (1992) find that the observed refractive quantities matched well with the predicted values for five objects, three other objects, especially PSR B0833--45, has a significantly shorter measured $\tref$ and greater modulation index than expected. This is very similar to our situation here with PSR B1937+21. Stinebring et al. (2000) concluded that the 21 objects that they analysed fell into two groups. The first group followed the frequency dependence predicted by a Kolmogorov spectrum with the inner cutoff scale far less than the diffractive scales ({\it Kolmogorov-consistent group}). The second group, which is the {\it super-Kolmogorov group}, is consistent with a Kolmogorov spectrum with an inner scale cutoff at $\sim 10^{8}$ meters. The observed modulation indices were consistently greater than that of the Kolmogorov predictions, as we have seen in our measurements of PSR B1937+21. This group includes pulsars like PSRs B0833--45 (Vela), B0531+21 (Crab), B0835--41, B1911--04 and B1933+16. An important physical property that binds them all is that, excepting one object, all objects have a strong {\it thin-screen} scatterer somewhere along the LOS. This is either a supernova remnant (or a plerion) like in the case of Vela and Crab pulsars, a HII region as in the case of B1942--03 and B1642--03, or a Wolf-Reyet star as in the case of B1933+16 (see Prentice \& ter Haar 1969; Smith 1968). Although our measurements show that pulsar PSR B1937+21 is consistent with the characteristics of the {\it super-Kolmogorov} group, as we describe in \S\ref{sec-sightline}, we find no compelling evidence for the presence of any dominant scatterer somewhere along the LOS. To summarize, while some investigators have reported agreement of the measured refractive properties with the theoretical expectations from a Kolmogorov spectrum with an infinitismally small inner scale, there are a considerable number of cases where the observed properties are significantly different from that predicted by the simple Kolomogorov spectrum. These other cases can be explained by invoking spectrum with a large inner scale cutoff, including the case where the cutoff approaches the Fresnel radius and leads to a caustic-dominated regime. From Gupta et al. (1993) and Stinebring et al. (2000), the modulation index can be specified as a function of the inner cutoff scale as \begin{equation} m\;=\;0.85\; \left( \frac{\Delta\nu}{\nu} \right)^{0.108}\; \left( \frac{r_i}{10^8{\rm m}} \right)^{0.167} D_{\rm kpc}^{-0.0294}. \end{equation} With the known value of $\Delta\nu$ at 327 MHz of 1.33 kHz, the distance to the pulsar of 3.6 kpc, and the observed modulation index of 0.39, the inner scale cutoff, $r_i$, comes to $1.3\times 10^9$ meters. \section{DM variations} \label{sec-dmvar} We turn now to the dispersion measure variations presented in Figure \ref{fig:dmvstime} that sample density variations on transverse scales much larger than those involved with diffractive and refractive effects. The most striking feature in Figure \ref{fig:dmvstime} is the large secular decline from 71.040 pc cm$^{-3}$ in 1985 to 71.033 pc cm$^{-3}$ in 1991 and then to 71.022 pc cm$^{-3}$ by late 2004. These long-term secular variations are many times greater than the RMS fluctuations of $\sim 10^{-3}$ pc cm$^{-3}$ on short time scales. An important question that arises is whether these variations are the result of a spectrum of electron-density turbulence, or whether there might be a contribution from the smooth gradient of a cloud, or clouds along the LOS. We look at this question from two angles. First we present a phase structure function analysis of the dispersion measure data and estimate a power-law index of the electron density spectrum. Then we estimate the probability that such a spectrum would produce a 22-y realization that was so strongly dominated by the large, monotonic changes mentioned above. We write the power spectrum of electron-density fluctuations as \begin{equation} P(q)\;=\; C_n^2\;q^{-\beta},\;\;\;\;\;\;\;\;\;\;\;\;[\qo<q<\qi] \end{equation} \noindent where $\beta$ is the power law index, $\qo$ and $\qi$ are the spatial frequencies corresponding to the outer and the inner boundary scale, within which this power law description is valid. $C_n^2$ is the amplitude, or strength, of the fluctuations. A quantity that is closely related to the density spectrum which can be quantified by observable variables is the phase structure function, $D_{\phi}(b)$, with $b=2\pi/q$. This is defined as the mean square geometric phase between two straight line paths to the observer, with a separating distance of $b$ between them in the plane normal to the observer's sight line. The phase structure function and the density power spectrum are related by a transform (Rickett 1990; Armstrong, Rickett, Spangler 1995), \begin{eqnarray} D_{\phi}(b)\;=\;\int_0^{\infty} 8\pi^2\lambda^2 r_e^2\;dz'\; \int_0^{\infty}q\;[1-J_0(bqz'/z)]\;dq \nonumber \\ \times P(q=0) \end{eqnarray} Here, $r_e$ is the classical electron radius (2.82$\times$10$^{-15}$ meters), $J_{\circ}$ is the Bessel function. Under the conditions that we have assumed, $D_{\phi}(b)$ is also a power law (Rickett 1990; Armstrong, Rickett \& Spangler 1995), given by \begin{equation} D_{\phi}(b)\;=\;\left(\frac{b}{b_{\rm coh}}\right)^{\beta-2} \end{equation} \noindent where $b_{\rm coh}$ is the coherence spatial scale. Dispersion measure can be written as \begin{equation} DM\;=2.410\times 10^{-16}\, \left[\frac{(\nu_1^2-\nu_2^2)}{ \nu_1^2 \nu_2^2}\right]\; \left(\frac{\Delta\phi}{f}\right)\; {\rm pc\; cm^{-3}}, \end{equation} \noindent where $\Delta\phi$ is the difference in the arrival phases $(\phi_2-\phi_1)$ of the pulse at the two barycentric radio frequencies (Hz) $\nu_1$ and $\nu_2$, with $f$ being the barycentric rotation frequency (Hz) of the pulsar. With this linear relation between DM and geometric phase difference, then structure function can be written as (KTR94) \begin{eqnarray} D_{\phi}(b_{\circ})&=&\left( \frac{2\pi}{\nu} \frac{{\rm Hz}}{ 2.410 \times 10^{-16}\;{\rm pc\;cm^{-3}}}\right)^2 \nonumber\\ && \times \langle [DM(b+b_{\circ}) - DM(b)]^2 \rangle. \label{eqn:struct} \end{eqnarray} Here, the angular brackets indicate ensemble averaging. The transformation between the spatial coordinate $b$ (and the spatial delay $b_{\circ}$) and the time coordinate $t$ (or time delay $\tau$) is simply given by $b=V_\perp t$, where $V_\perp$ is the transverse velocity of the LOS across the effective scattering screen. \begin{table} \label{tab:scintparams} \begin{center} \caption[]{Measured and expected parameters.} \begin{tabular}{c|c|c|c|c|c|c|c} \hline $\tausc$ & $\theta_H$ & $\tdiff$ & \multicolumn{2}{|c|}{$T_{\rm ref}$} & \multicolumn{2}{|c|}{$m$} & {\bf $\nu$} \\ ($\mu$s) & (mas) & (s) & observed & expected & observed & expected & (MHz) \\ \hline\hline 120$^{\dagger}$ & 14.6$^{b}$ & -- & 73 days & 3 y$^f$ & 0.33 & 0.14$^c$ & 327 \\ 38$^e$ & -- & 100$^e$ & -- & -- & -- & -- & 430 \\ -- & -- & -- & 43.9 days & 6 mon$^g$ & 0.39 & 0.2 & 610 \\ -- & -- & 260$^a$ & 3 days$^d$ & 45 days$^g$ & 0.45 & 0.33$^c$ & 1400 \\ 0.17$^e$ & -- & 444$^e$ & -- & -- & -- & -- & 1400 \\ \hline \end{tabular} \end{center} $^{\dagger}$Has a time dependent RMS variation of 20$\mu$s\\ $^a$Cordes et al. (1986) \\ $^b$Gwinn et al. (1993) \\ $^c$Romani et al. (1986); Kaspi \& Stinebring (1992) \\ $^d$Lestrade et al. (1998) give the value as 13 days\\ $^e$Cordes et al. 1990 \\ $^f$Calculated with $\tref\sim\theta_HD/2\vlos$ \\ $^g$Extrapolated with $T_{\rm ref}\propto\lambda^{2.2}$ \end{table} With the understanding that any difference in DM that we compute for a time baseline from Figure \ref{fig:dmvstime} corresponds to a point in the phase structure function, we can derive the phase structure function on the basis of Equation \ref{eqn:struct}. This is given in Figure \ref{fig:struct}. There are several important points in Figure \ref{fig:struct}. The solid line gives the best fit line for the data in the time interval of 5 days to 2000 days. The derived values of the intercept and the power law index ($\beta$) are, \begin{eqnarray} {\rm intercept} &=& 4.46\pm 0.09 \nonumber\\ \beta &=& 3.66\pm 0.04 \label{eq:fitpars} \end{eqnarray} The value of $\beta$ is remarkably close to the value expected from a Kolmogorov power law distribution ($\beta=11/3$). We are using the terminology ``intercept" only to indicate the value of $\log[D_{\phi}(\tau)]$ when $\log[time lag (days)]$ is zero. Here, a cautionary remark is warranted. Given the finite time span of our data set, and the fact that the low spatial frequencies dominate the long term variations in DM, we do not have a stationary sample of noise spectrum. We have estimated the error in each bin of the structure function as \[ \sigma_s = \frac{\sigma_{D}}{\sqrt{N_i}}, \] where $\sigma_{D}$ is the root mean square deviation with respect to the mean phase structure function value in a bin, $D_{\phi}(\tau)$, and $N_i$ is the number of ``independent'' samples in the bin. This is estimated as the smaller of $(T/\tau)$ and the actual number of samples that have gone into the estimation of $D_{\phi}(\tau)$. Here, $T$ is the time span of the data. By assuming that the transverse speed of the sightline across the effective scattering screen is $\sim$40 km sec$^{-1}$ (half of pulsar's velocity in its LSR), we can translate the delay range between which this slope is valid, to 0.2 to 50 A.U. The time delay value corresponding to the phase structure function value of unity is, by definition, the coherent diffractive time scale ($\tdiff$) at the corresponding radio frequency, with the assumption that the scattering material is uniformly distributed along the LOS. From the fit parameters given in Equation~\ref{eq:fitpars}, this delay is 180 seconds. This should be compared with the measured $\tdiff$ value of 444$ \pm $28 seconds tabulated in Table 3. If we interpret the inner scale cutoff value of $r_i \sim 1.3\times 10^9$ meters as the scale size below which the slope ($\beta$) of the density fluctuation spectrum changes to a value greater than that given in Equation \ref{eq:fitpars}, then the fact that the measured $\tdiff$ value of 444 seconds being significantly greater than 180 seconds is understandable. In the limiting case, where the slope of the density irregularity power spectrum changes to the critical value of $\beta = 4$ below the inner scale cutoff value, the expected $\tdiff$ value is about 1100 seconds. This makes it very important to measure the exact frequency dependence of the diffractive parameters like temporal scatter broadening and diffractive scintillation time scale. To the best of our knowledge, Cordes et al. (1990) show the most complete multi-frequency measurements of the diffractive scintillation parameters of this pulsars. Their measurements are not accurate enough to distinguish between such small variations in slope. While our analysis of DM variability suggests a Kolmogorov spectrum at AU scales, we are struck by the long term monotonic decrease of DM and wonder if we might be seeing the effects of smooth gradients in large scale galactic structures that are not part of a turbulent cascade. We performed a Monte Carlo simulation to investigate this. In each realization of the simulation, we generated with a different random number seed, a screen of density fluctuations. We assumed that the random fluctuations corresponding to a given spatial frequency are described by a Gaussian function, but the total power as a function of spatial frequencies is described by a single power law of index --11/3. Assuming that the screen is located at the mid point along the sight line, we let the pulsar drift with its transverse velocity, and measured the implied column density (DM) as a function of time. We developed a procedure similar to that of Deshpande (2000) to compare the observed $\dmt$ curve with the simulated ones. From the observed $\dmt$ curve, we computed the parameter $\Delta DM\;=\;[DM(t) - DM(t-\tau_{\circ})]$, where $\tau_{\circ}$ is the time delay. Our aim is to compare the distribution of this parameter in very short delays and very large delays. As we can see in Figure \ref{fig:struct}, the structure function describes a well defined slope between the delay range of $\sim$30 days to $\sim$2000 days. We defined two delay bins, 30--60 and 1300--2000 days, within which we monitored the distribution function of the quantity $\Delta DM$. From this, we could infer that the distribution at the bin of 1300--2000 days had a span of $\sim 20 \sigma_s$, where $\sigma_s$ is the RMS deviation of the distribution at the delay range of 30--60 days. That is, the $\Delta DM$ values that we see at largest delays is as high as 20 times that of the typical deviations at short delays. We performed the same procedure on the simulated set of data to quantify the likelihood of such deviations. We simulated 1024 number phase screens. Out of these 1024 screens, we found that such large deviations were possible $\sim$7\% of the times. This is perhaps not surprising, as with such a steep spectrum, it is obvious that most of the power is in large scales (smaller spatial frequencies), and hence they tend to dominate our $\Delta DM$ measurements. We conclude that while monotonic changes of this magnitude are rare, it is consistent with a turbulent cascade spectrum of density fluctutations. \section{Frequency dependence of DM} \label{sec-dmfreq} Dispersion depends on the column density of electrons. In a uniform medium radio wave propagation senses the average density in a tube whose beam waist is set by the Fresnel radius $\sqrt{z(1-z)\lambda D}$. In a turbulent medium frequency-dependent multipath propagation can expand this tube considerably. With refraction, the center of the tube wanders from the geometric LOS. Indeed there may be a number of wave propagation tubes, each with their independent relative gain. The consequence is that DM and related effects will show frequency dependence: \begin{enumerate} \item the effective DM depends on frequency. \item the DM variations at lower frequencies will be much ``smoother'' than that at higher frequencies, as the apparent angular size of the source acts as a {\it smoothing function} on the measured DM variations. \item since the apparent size of the source is larger at low frequencies, some features of the ISM that are visible at lower frequencies may be invisible at higher frequencies! \end{enumerate} We can explore these effects by assuming that the timing residuals at 327 MHz and 610 MHz, which are relative to the timing model derived at higher frequencies that included removal of DM variations, are due to DM variations. The smoothing effect of scattering could be revealed by a spectral analysis. The slow variations were removed to pre-whiten the spectrum that would otherwise be severely contaminated. The 327 MHz data was fit to a fourth order polynomial and the result was subtracted from both data sets. The two right side panels give the residual DM values after subtracting the best fit curve from the actual DM curve. The resulting spectral comparison fails to have sufficient signal to clearly demonstrate increased smoothing at 327 MHz relative to 610 MHz. Higher signal-to-noise ratio is required. The DM variations relative to long term trends in the right-hand panels of Figure \ref{fig:dmindiv} are different. An important source of systematic error that can affect our analysis here is the effect of scattering on the derived DM as a function of time at a given frequency. At 327 MHz, as we described in \S\ref{sec-anal}, we perform an elaborate procedure to fit for the scatter broadening of the pulse profile, in order to compute the ``true'' TOA of the profile. However, we do not follow this procedure at 610 MHz (or any other higher frequency). The error due to this can be quantified easily from Figure \ref{fig:tausc}. The temporal scatter broadening value varies by an RMS value of some 19.6 $\mu$s. With the wavelength dependence of $\tausc\propto\lambda^{4.4}$, the expected RMS variation at 610 MHz is 1.3 $\mu$s. The equivalent DM perturbation at 610 MHz with respect to infinite frequency is $\sim 10^{-4}$~pc~cm$^{-3}$. \section{Achievable timing accuracy} \label{sec-timingerror} In this section, we will estimate quantitatively errors introduced by various scintillation related effects. For PSR B1937+21, although a typical observation with highly sensitive telescopes like Arecibo telescope helps us achieve a TOA accuracy of a few tens of nanoseconds, the ultimate long term accuracy seems to be far greater than this. In general, it is a combination of frequency independent ``intrinsic timing nose" from the pulsar itself, and the frequency dependent effects, such as what we are addressing here. With some 18 years of data at 800, 1400 and 2200 MHz, Lommen (2002) quantifies the timing timing residual, after fitting for position, proper motion, rotation frequency and its time derivative (see also Kaspi et al. 1994). A large fraction of the left over residuals is presumably the intrinsic timing noise. As we have mentioned before, we have absorbed a good part of this by fitting for the second time derivative of the rotation frequency, $\ddot{f}$ (see Table~2). In this section, our aim is to quantify possible timing errors from various ``chromatic'' effects related to interstellar scintillation. \subsection{Fluctuation of apparent angular size} The temporal variability of pulse broadening, $\tausc$, (as shown in Figure \ref{fig:tausc}) means that even the apparent angular broadening of the source, $\theta_H$, is also changing as a function of time. Since $\tausc \propto \theta_H^2$, with the RMS variation in $\tausc$ of 19.6 $\mu$s at the radio frequency of 327 MHz, the corresponding variation in $\theta_H$ comes out to be $\sim$8\% of the mean value. This change occurs with typical time scales of $\sim$67 days, which is the time scale with which $\tausc$ changes. Since we have only one epoch of $\theta_H$ measurement, we have no way of observationally verifying the mean value or the time scale of its variation. \subsection{Image wandering and the associated timing error} Due to non-diffractive scintillation that ``steers'' the direction of rays (``refractive focussing''), the position of the pulsar is expected to change as a function of time. This is an important and significant effect, as it introduces a TOA residual as a function of time, depending on the instantaneous position of source on the sky. Several authors have investigated this effect in the past (Cordes et al. 1986; Romani et al. 1986; Rickett \& Coles 1988; Fey \& Mutel 93; Lazio \& Fey 2001). For a Kolmogorov spectrum of irregularities ($\beta = 11/3$) with infinitismally small inner scale cutoff, Cordes et al. (1986) predict the value of RMS image wandering as \begin{equation} \langle\delta\theta^2\rangle^{1/2} = 0.18\; {\rm mas} \; \left( \frac{D_{\rm kpc}}{\lambda_{\rm cm}} \right)^{-1/6} \theta_H^{2/3} \end{equation} For an assumed distance to PSR B1937+21 of 3.6 kpc, this comes to 2 mas at 327 MHz (wavelength, $\lambda = 92$ cm). The value of 2 mas is still significantly less than the apparent angular size of the source, 14.6 mas, measured by Gwinn et al. (1993). However, for a spectrum with a steeper slope or with a significantly larger inner scale cutoff (as in our case), the value of $\langle\delta\theta^2\rangle^{1/2}$ is expected to be much larger, perhaps comparable to the value of $\theta_H$. In order to estimate the timing errors introduced by this image wandering, we need an estimate of scattering measure ($SM$) and $C_n^2$ along the LOS to this pulsar. Following Cordes et al. (1991), \begin{eqnarray} SM &=& \int_0^D C_n^2(x)\;\;{\rm d}x \nonumber\\ &=& \left(\frac{\theta_H}{128\;{\rm mas}} \right)^{5/3} \nu_{\rm GHz}^{11/3} \nonumber\\ &=& 292\; \left( \frac{\tausc}{D_{\rm kpc}} \right)^{5/6} \nu_{\rm GHz}^{11/3}. \end{eqnarray} Here, $\tausc$ is specified in seconds. $SM$ is specified in units of kpc m$^{-20/3}$. Assuming a distance of 3.6 kpc, $\tausc$ = 120 $ \mu $s, $\nu$=0.327 GHz, the value of $SM$ comes to $\sim 8.8\times 10^{-4}$ kpc m$^{-20/3}$. Assuming that the scattering material is uniformly distributed along the LOS, $C_n^2\sim 2.4\times 10^{-4}$ m$^{-20/3}$. Then, for a Kolmogorov spectrum, the RMS timing residual due to the image wandering can be written as (Cordes et al. 1986) \begin{equation} \sigma_{\delta t_{\theta}} = 26.5 \;\; {\rm ns} \;\; \nu^{-49/15} D^{2/3} \left( \frac{C_n^2}{10^{-4} m^{-20/3}} \right)^{4/5}. \end{equation} With the computed value of $C_n^2$ and a distance of 3.6 kpc, this amounts to 4.8 $\mu$s at 327 MHz. Given the frequency dependence, this effect can be minimized by timing the pulsar at higher frequencies. For instance, at frequencies of 1 GHz and 2.2 GHz, this error translates to 125 and 2 nanosec, respectively. However, given the significantly large value of the inner scale cutoff, the RMS timing error that we have computed may well be a lower limit, and it is likely to be higher. Given the fact that the exact source position due to this effect is unknown at any given time, it is very difficult to compensate for this effect. \subsection{Positional errors in solar system barycentric corrections} As we saw above, due to image wandering, the apparent position of the source wanders in the sky. This introduces yet another timing error. While translating the TOA at the observatory to the solar system barycenter, we assume a position which is away from the actual apparent position at the time of observation. This introduces an error, which can be quantified as (Foster \& Cordes 1990) \begin{equation} \Delta t_{\rm bary}\;=\; \frac{1}{c}\; (\vec{r_e}\cdot\hat{n}) \; (1-z) \Delta\theta_r(\lambda). \end{equation} Here, $c$ is the velocity of light, the dot product term gives the projected extra path length travelled by the ray due to Earth's annual cycle around the Sun, and $\Delta\theta_r(\lambda)$ is the positional error due to image wandering. Of course, this term is a function of frequency, and hence the error accumulated is different at different frequencies. At 327 MHz, with an RMS image wandering angle of 2 mas, for an object at the ecliptic plane, $\Delta t_{\rm bary}\sim 2$ $\mu$s. For PSR B1937+21, this error amounts to $\sim$0.8 $\mu$s. At frequencies 1 GHz and 2.2 GHz, this error translates to 85 and 17 nanosec, respectively. \subsection{Timing error due to DM variation as a function of frequency} \label{sec-accuracy} An important issue that arises due to the frequency dependent DM variation is the timing accuracy. Pulsars like PSR B1937+21 are known for the accuracy to which one can compute the pulse TOA. Given this, one wishes to eliminate any error that is incurred due to systematic effects like what we have here. Between 327 and 610 MHz (the two curves in Figure \ref{fig:dmindiv}), the typical relative fluctuation of DM that we see is about $5\times 10^{-4}$ pc cm$^{-3}$. As we discussed before, this is purely due to the fact that the effective interstellar column length sampled at one frequency is different from that at another frequency, due to the scatter broadening of the source. At 610 MHz, this relative DM fluctuation corresponds to some 6 $\mu$s at 610 MHz. That is, at 610 MHz, typically an unaccounted residual of 6 $\mu$s is incurred due to just effective DM errors. Even if the behavior of the pulse emission is extremely stable, at low frequencies, interstellar scattering limits our timing capabilities. Due to the fact that dispersion delay goes as $\nu^{-2}$, although the above mentioned effect seems significant, one should be able to reduce it by going to higher frequencies. For instance, at 2.2 GHz, the DM-limited TOA error for PSR B1937+21 will be $\sim$0.5$\mu$s. This is not necessarily encouraging news, as a timing residual error of 0.5$\mu$s is large when compared to the accuracy that we can achieve in quantifying the TOAs (a few tens of ns) for this pulsar, given our observations with sensitive telescopes like Arecibo. To summarize, although one takes into account time dependent DM changes while analysing the data, in order to achieve high accuracy timing, it is important to correct for a frequency dependent DM term. This introduces another dimension of correction in the timing analysis. \section{Concluding remarks} We have presented in this paper a summary of over twenty years of timing of PSR B1937+21. These observations have been done over frequencies ranging from 327 MHz to 2.2 GHz with three different telescopes. Given the agreement between the measured apparent angular broadening and that estimated by the temporal broadening, and the measured proper motion velocity and that estimated by the knowledge of scintillation parameters, we conclude that the scattering material is uniformly distributed along the sightline. There are three significant discrepancies between the expected values and the measured refractive parameters. These are, \begin{enumerate} \item The measured flux variation time scale is about an order of magnitude shorter than what is expected from the knowledge of the observed apparent angular broadening. \item The flux variation time scale is observed to be directly proportional to the wavelength, whereas it is expected to vary as proportional to $\lambda^{2.2}$ (for a Kolmogorov spectrum). \item The flux modulation index is observed to have a wavelength dependence that is much ``shallower" than the expected value. \end{enumerate} These three discrepancies consistently imply that the optics is ``caustic-dominated''. This would mean that the density irregularity spectrum has a large inner scale cutoff, $1.3\times 10^9$ m. Our extrapolation of the phase structure function from the regime sampled by DM variations to the diffractive regime seems to indicate that the expected $\tdiff$ value is considerably shorter than the measured value. This is in favor of the above conclusion. Accurate measurements of frequency dependence of diffractive parameters is much needed. In general, Millisecond pulsars are known for their timing stability. Potentially, we may achieve adequate accuracy in timing some of these pulsars to understand some of the most important questions related to the gravitational background radiation, or the internal structure of these neutron stars. However, our analysis here shows that interstellar scattering could be an important and significant source of timing error. As we have shown, although PSR B1937+21 is known to produce short term TOA errors as low as 10--20 nanosec with sensitive observations, the long term error is far larger than this. After fitting for $\ddot{f}$ (which absorbes most of the achromatic timing noise), the best accuracy that we can achieve for this pulsar is $0.9\;\mu$sec at 1.4 GHz, and about 0.5 $\mu$sec at 2.2 GHz (by one of the authors, Andrea Lommen). It appears that almost all of this error can be accounted for by various effects that we have discussed in \S\ref{sec-timingerror}. In general, for millisecond pulsars with substantial DM, even if achromatic timing noise is small, interstellar medium may be a major source of timing noise. \acknowledgements We thank M. Kramer for sharing the data from the Effelsberg-Berkeley Pulsar Processor (EBPP), and S. Chatterjee and W. Brisken for sharing their VLBA based proper motion and parallax results of PSR B1937+21 prior to the publication. This work was in part supported by the NSF grant, AST--9820662.
Title: Spectra of the spreading layers on the neutron star surface and constraints on the neutron star equation of state
Abstract: Spectra of the spreading layers on the neutron star surface are calculated on the basis of the Inogamov-Sunyaev model taking into account general relativity correction to the surface gravity and considering various chemical composition of the accreting matter. Local (at a given latitude) spectra are similar to the X-ray burst spectra and are described by a diluted black body. Total spreading layer spectra are integrated accounting for the light bending, gravitational redshift, and the relativistic Doppler effect and aberration. They depend slightly on the inclination angle and on the luminosity. These spectra also can be fitted by a diluted black body with the color temperature depending mainly on a neutron star compactness. Owing to the fact that the flux from the spreading layer is close to the critical Eddington, we can put constraints on a neutron star radius without the need to know precisely the emitting region area or the distance to the source. The boundary layer spectra observed in the luminous low-mass X-ray binaries, and described by a black body of color temperature Tc=2.4+-0.1 keV, restrict the neutron star radii to R=14.8+- 1.5 km (for a 1.4-Msun star and solar composition of the accreting matter), which corresponds to the hard equation of state.
https://export.arxiv.org/pdf/astro-ph/0601689
\date{Accepted, Received} \pagerange{\pageref{firstpage}--\pageref{lastpage}} \pubyear{2005} \label{firstpage} \begin{keywords} {accretion, accretion discs -- radiative transfer -- X-rays: binaries -- stars: neutron } \end{keywords} \section{Introduction} Matter accreting on to a weakly magnetized neutron star (NS) in low mass X-ray binaries (LMXRBs) can form an accretion disc which extend down to the NS surface. A boundary layer (BL) is formed between the accretion disc and the NS surface, where a rapidly rotating matter of the disc is decelerated down to the NS angular velocity. The amount of the energy, which is generated during this process, is comparable with the energy generated in the accretion disc \citep{SS86,SS98}. There is no generally accepted theory of the BL. Two different approaches to the BL description are considered. First of them, which we will call a `classical model', considers the BL between a central star (a white dwarf or a NS) as a part of the accretion disc \citep{P77,PS79,T81,SS88,BK94,PN95,PS01}. In this model the component of velocity normal to the accretion disc plane is zero. The half-thickness of the BL is determined by the same relation, as for the accretion disc: \be \label{eq:Hbl} \Hbl\sim \frac{c_{\rm s}}{v_{\rm K}} R, \ee where $c_{\rm s}$ is a sound speed in the BL and $v_{\rm K}$ is the Keplerian velocity close to the NS surface of radius $R$. The radial extension of the BL is determined by the relation \citep{P77} \be \label{eq:hbl} \hbl\sim \frac{c^2_{\rm s}}{v^2_{\rm K}} R \sim \Hbl\frac{\Hbl}{R}. \ee In this classical model the accreting matter in the BL is decelerated in the accretion disc plane, along radial coordinate only, due to the viscosity operating within the differentially rotating BL, similarly to the accretion disc. From the observational point of view, the classical BL is a bright equatorial belt close to the NS surface. The effective temperature of the BL is higher than the maximum accretion disc effective temperature, because the BL is smaller than the accretion disc, while their luminosities are comparable. Another approach was suggested by \citet[][ hereafter IS99]{IS99}. The BL is considered as a spreading layer (SL) on the NS surface. The accreting matter diffusing along the radial direction in the accretion disc and reaching the neutron star surface gains the velocity component normal to the accretion disc plane due to the ram pressure from the accretion disc. Then the matter spirals along the NS surface towards the poles. Rotating at the NS surface, the matter is decelerated due to a turbulent friction between the rapidly rotating matter and a slowly rotating NS surface. The kinetic energy of the accreting gas is mostly deposited in two bright belts at some latitude above and below the NS equator. The width and the latitude of the belts depend on the mass accretion rate. The larger the accretion rate, the wider the belts are and the closer they are to the NS poles. At the accretion rate close to the Eddington limit ($L_{\rm BL} \sim \Ledd$) the bright belts expand all over the NS surface. The observational difference between two BL models is not significant. At low accretion rates ($\Lbl\sim 0.01 \Ledd$) the latitude of bright belts of the SL is small ($\sim$ few degrees) and the vertical extension of the SL is comparable to the classical BL thickness. At high accretion rates ($\Lbl\sim \Ledd$) the classical BL thickness is comparable to the NS radius \citep[see][]{PS01}. Therefore, the effective temperatures of these BL models are of the same order. In the approach by IS99, the NS radius is assumed to be larger that $3\Rs$ (where $\Rs=2GM/c^2$ is the Schwarzschild radius of a NS of mass $M$), but the accretion disc structure is not significantly changed up to the NS surface, and the radial velocity is always subsonic. If the radial velocity of accreting gas is supersonic at the surface (see e.g. \citealt*{PN92}, for a possibility of the supersonic radial velocity in classical BL, and \citealt*{KW91}, for the ``gap accretion'' when the NS is within the innermost stable circular orbit), some fraction of the kinetic energy (associated with a small radial velocity component) should be dissipated in an oblique shock, but most of it still remains stored in the kinetic energy of the gas rotating at the surface to be dissipated later in the SL. The gap accretion model of \citet{KW91} is rather similar to the SL, but they did not consider the fate of the accreting material and its spread over the surface. At very low accretion rate, both models should produce hard Comptonization spectra extending to $\sim100$ keV. The aim of this work is the calculation of the radiation spectra of the SLs and their comparison to the observed X-ray spectra of the BLs in the luminous LMXRBs. \section{Spreading layer model} The theory of the SL was developed by IS99 under the assumption of Newtonian gravity. They considered accretion of the pure hydrogen plasma. Here we repeat the IS99 theory for plasmas of arbitrary chemical composition taking into account general relativity (GR) corrections using the pseudo-Newtonian potential. These corrections may be important, because the maximum effective temperature of the SL, which should be smaller than the local Eddington effective temperature $\Tcr$, depends on the opacity and the gravity. The critical temperature is determined by the balance between the surface gravity and the radiative acceleration: \be \label{eq:tc} \frac{G \Mns}{\Rns^2 \sqrt{1-\Rs/\Rns}} = \frac{\sigmasb \Tcr^4}{c} \sigmae, \ee where $\sigmae = 0.2 (1+X)$ cm$^2$~g$^{-1}$ is the electron scattering opacity, $X$ is the hydrogen mass fraction, and $\sigmasb$ is the Stefan-Boltzmann constant. It is clear, that solar (or larger) helium abundance together with the GR correction will lead to higher $\Tcr$, and, therefore, to a higher maximum effective temperature of the SL. This could have important consequences for the determination of the neutron star parameters from observations. We use the pseudo-Newtonian potential in the form: \be \label{eq:fi} \Psi (r) = -c^2 \left(1 - \sqrt{1-\Rs/r}\right). \ee This potential gives the correct GR surface gravity \be \label{eq:g0} g_{\rm 0}(\Rns)= \frac{G \Mns}{\Rns^2\sqrt{1-\Rs/\Rns}}, \ee but gives the Keplerian velocity at the NS surface \be \label{eq:vk} v^2_{\rm K}(\Rns) = \frac{G \Mns}{\Rns\sqrt{1-\Rs/\Rns}}, \ee which is smaller than the correct GR value. \subsection{Main equations} Below we rewrite the IS99 equations for the SL for the pseudo-Newtonian potential (\ref{eq:fi}) and considering arbitrary abundances. We consider the dynamics of the SL on the NS surface (see Fig.~\ref{fig1}). The full hydrodynamic equations, which describe this process are as follows \citep[see for example][]{M78}. The continuity equation is \be \label{eq:cont} \frac{\partial \rho}{\partial t} + {\bf \nabla \cdot} (\rho \vecv) =0, \ee where $\rho$ is the plasma density, $\vecv$ is the vector of the gas velocity in the SL with components $\vphi$, $\vtheta$ and $v_{r}$, which are velocities of the SL along longitude, latitude and radius correspondingly. Conservation of momentum for each gas element is described by the vector Euler equation \be \label{eq:euler} \rho \frac{\partial \vecv}{\partial t} + \rho \vecv \cdot \nabla \vecv = - \nabla P + \vecf, \ee where $P=P_{\rm rad}+P_{\rm g}$ is the total pressure which is a sum of the radiation and gas pressures, and $\vecf$ is a force density. The energy equation for the gas in the SL is \begin{eqnarray} \label{eq:energy} \frac{\partial}{\partial t} \left( \frac{1}{2} \rho v^2 +\varepsilon\right) + \nabla \cdot \left[ \left(\frac{1}{2} \rho v^2 + \varepsilon + P\right){\vecv}\right] = \\ \nonumber \vecf \cdot \vecv - \nabla \cdot \vecq +Q^+. \end{eqnarray} Here $\varepsilon = \varepsilon_{\rm rad}+ \varepsilon_{\rm g}$ is the total density of internal energy, where $\varepsilon_{\rm rad}=aT^4$ is the radiation energy density and $\varepsilon_{\rm g} = (3/2) P_{\rm g}$ is the density of the internal gas energy. The first term on the right hand side is the power produces by the force density, the second is the energy, which is lost by radiation ($\vecq$ is a vector of the radiation flux), and the third is the heat, which is generated within a unit volume of the SL. Following IS99 we consider the steady state SL model in the spherical coordinate system $(r,\theta,\varphi)$, where $\theta$ is the latitude and $\varphi$ is the azimuthal angle (see Fig.~\ref{fig1}). We also assume that the SL has a small thickness (in comparison with the NS radius $\Rns$, therefore the radial coordinate $r=\Rns$), the radial velocity component is zero $v_r=0$, and it is axially symmetric (therefore, all of the derivatives $\partial / \partial \varphi$ equal to zero). In this case equations (\ref{eq:cont})--(\ref{eq:energy}) take the following form. The continuity equation is \be \label{eq:cont1} \frac {1}{R \cos \theta} \frac{\partial}{\partial \theta} (\cos \theta\ \rho \vtheta) = 0 , \ee the three components of the Euler equation are \be \label{eq:e1} -\rho\left(\frac{\vtheta^2 +\vphi^2}{R}\right) = -\frac{\partial P}{\partial r} +\fr, \ee \be \label{eq:e2} \rho \frac{\vtheta}{R} \frac{\partial \vtheta}{\partial \theta} + \rho \frac{\vphi^2}{R} \tan \theta = -\frac{1}{R}\frac{\partial P}{\partial \theta} +\ftheta, \ee \be \label{eq:e3} \rho \frac{\vtheta}{R} \frac{\partial \vphi}{\partial \theta} - \rho \frac{\vphi \vtheta}{R} \tan \theta = \fphi, \ee and the energy equation is \begin{eqnarray} \label{eq:energy1} \frac{1}{R \cos \theta} \frac{\partial}{\partial \theta} \left[ \cos\theta \ \vtheta \left(\frac{1}{2} \rho v_0^2 + \varepsilon +P\right)\right] = \\ \nonumber \ftheta\vtheta + \fphi\vphi - \frac{\partial q}{\partial r} + Q^+, \end{eqnarray} where \be \label{v0} v_0^2= \vphi^2 + \vtheta^2. \ee Here the radiation flux has only one (radial) non-zero component and its divergence is computed in the plane-parallel approximation which is a consequence of our assumption of small height of the SL. A small azimuthal component of the radiation flux arises due to the aberration, which we neglect here. It is clear that equation~(\ref{eq:e1}) can be solved independently on equations (\ref{eq:e2})--(\ref{eq:e3}) and we can consider some averaging over the layer's height. Therefore, we arrive at a one-dimensional problem. In this case instead of equations (\ref{eq:cont1})--(\ref{eq:energy1}) we obtain \be \label{eq:cont2} \frac {1}{R \cos \theta} \frac{\partial}{\partial \theta} \left( \cos \theta \int \rho \vtheta \d r\right) = 0 , \ee \be \label{eq:e22} \int \rho \vtheta \frac{\partial \vtheta}{\partial \theta} \d r + \tan \theta \int \rho \vphi^2 \d r = -\frac{\partial}{\partial \theta} \int P \d r + R \int \ftheta \d r , \ee \be \label{eq:e32} \int \rho \vtheta \frac{\partial \vphi}{\partial \theta} \d r - \tan \theta \int \rho \vphi \vtheta \d r = R \int \fphi \d r , \ee \begin{eqnarray} \label{eq:energy2} \frac{1}{R \cos \theta} \frac{\partial}{\partial \theta} \left[ \cos\theta \int \vtheta \left(\frac{1}{2} \rho v_0^2 + \varepsilon +P\right) \d r \right] = \\ \nonumber \int \ftheta\vtheta \d r + \int \fphi\vphi \d r - q + \int Q^+ \d r. \end{eqnarray} Here the integration over radius is from $R$ to $R+\hs$, where $\hs$ is the local SL thickness. We define the corresponding force densities in the next section. \subsection{Vertical averaging} \label{sec:avehei} The one-dimensional equations for the SL structure are derived using the averaging along the height at a given latitude. It means that we have to calculate all the integrals in equations (\ref{eq:cont2})--(\ref{eq:energy2}) for some model of the SL vertical structure. The simplest way is just to consider the variables averaged over the height. IS99 used a more complicated model of averaging. They constructed a simple model of the SL using assumptions that velocities $\vtheta$, $\vphi$ and the radiation flux do not depend on the height $q(r)=\mbox{const}=\sigmasb \Teff^4$. The latter suggestion means that all of the thermal energy in the SL is generated at the bottom. This model is described by the hydrostatic equilibrium equation (\ref{eq:e1}) taken in the form \be \label{eq:he} \frac{\d P}{\d m} = \geff \equiv g_0 - \frac{\vphi^2+ \vtheta^2}{R}, \ee and the radiation transfer equation in the diffusion approximation \be \label{eq:re} \frac{\d\varepsilon_{\rm rad}}{\d m} = \frac{3q}{c} \sigmae. \ee Here and below we use a new independent variable: a column mass $m$ and a new geometrical coordinate $z$, which are defined as \be \d m = \rho \d z = -\rho \d r. \nonumber \ee Coordinate $z$ has an opposite direction relative to $r$ and $z$=0 at $r=R+\hs$. We also defined the $r$ component of the force density \be \fr = -g_0 \rho. \nonumber \ee Equations (\ref{eq:he})--(\ref{eq:re}) have to be supplemented by the equation of state \be \label{eq:se} P = \frac{\rho kT}{\mu m_{\rm p}} + \frac{\varepsilon_{\rm rad}}{3}, \ee where $\mu=4/(3+5X)$ is the mean molecular weight and $m_{\rm p}$ is the proton mass. Equations (\ref{eq:he})--(\ref{eq:se}) can be solved with the simple boundary conditions $P(m=0)=0$, $T(m=0)=0$: \be P=\geff m, \ee \be \varepsilon_{\rm rad}= \frac{3q}{c} m \sigmae, \ee \be \rho = \mu m_{\rm p} \frac{\gwr}{k} \left( \frac{a c}{3q \sigmae}m^3 \right)^{1/4}, \ee \be T=\left( \frac{3q}{ac} m \sigmae \right)^{1/4}= \Teff \left( \frac{3}{4} \taue\right)^{1/4}, \ee where $\taue= m \sigmae$ is the electron scattering optical depth of the layer, and \be \gwr \equiv \geff - \grad = \geff - \frac{q}{c} \sigmae. \ee The column density $m$ is related to the geometrical depth~$z$ \be m = \frac{(\mu m_{\rm p} \gwr)^4}{4^4 \sigmae k^4} \frac{ac}{3q} z^4, \ee which gives the following dependence of temperature on height \be T=\frac{\mu m_{\rm p} \gwr}{4k} z. \ee Following IS99, we consider the values of temperature and density at the bottom of the SL $\TS$ and $\rho_{\rm S}$ as parameters. In this case the local SL thickness $\hs$ is: \be \label{eq:height} \hs = \frac{4k\TS}{\mu m_{\rm p} \gwr}. \ee We can also calculate all of the integrals in equations~(\ref{eq:cont2})--(\ref{eq:energy2}): the total surface density \be \int_{0}^{\hs} \rho \d z = m(z=\hs) = \Sigmas, \ee the pressure surface density \be \int^{\hs}_0 P \d z = \frac{1}{5} \geff \hs \Sigmas= \frac{4}{5} \frac{\geff}{\mu m_{\rm p}\gwr} \Sigmas k \TS, \ee the surface density of the internal energy \begin{eqnarray} E_{\rm int} = \int^{\hs}_0 \left(\varepsilon_{\rm rad} + \frac{3}{2} P_{\rm g}\right) \d z = \frac{3}{2} \left(\geff+\grad\right) \frac{\Sigmas \hs}{5}, \end{eqnarray} the local flux \be q=\sigmasb \Teff^4 = \frac{ac}{3\sigmae} \frac{\TS^4}{\Sigmas}, \ee and the enthalpy flux \be H= E_{\rm int} + \int^{\hs}_0 P \d z = \left( \frac{5}{2} \geff + \frac{3}{2} \grad\right) \frac{\Sigmas \hs}{5} . \ee If we take $X=1$, all these relations will be the same, as derived by IS99 with one exception: there is no potential energy of the SL in our energy equation. Thus our expression for $H$ contains a factor $5/2$ instead of $7/2$ as in IS99. Below we will show that this produces only a small quantitative differences between our and IS99 models. In the IS99 model there are two forces, which give contribution to the force density in equations (\ref{eq:cont2})--(\ref{eq:energy2}). These are the gravity force, which has only the radial component (see above) and the force arising due to the friction between the NS surface and the SL. This force is directed along the NS surface and is expressed in the IS99 model through stress $\tau$ and its azimuthal and meridional components $\tauphi$ and $\tautheta$. IS99 have parameterized it in the form: \begin{eqnarray} \tauphi&=& -\int_0^{\hs} \fphi\d z = \alphab \rhos\vphi v_0,\\ \nonumber \tautheta &=& -\int_0^{\hs} \ftheta\d z = \alphab \rhos\vtheta v_0, \end{eqnarray} where $\alphab=v_*^2/v_0^2$ is the parameter of the stress parametrization, $v_*$ is the velocity of turbulent pulsations. We should note that $\alphab$ is not the same $\alpha$ that is used in the accretion disc theory. In accretion discs, $\alpha$ (in the first approximation) is the square of the ratio of the turbulent velocity to the sound speed $\alpha = v_*^2/c_{\rm s}^2$ and can be quite high, up to 0.1--1. In the SL, the plasma velocity $v_0$ is close to the Keplerian velocity at the NS surface and is orders of magnitude larger than the sound speed. The velocity of turbulent pulsation is also limited by the radiation viscosity at the SL bottom. IS99 carefully investigated this matter and estimated $\alphab \sim 10^{-3}$. We used this value in most of the paper. IS99 have ignored the mechanical work between the SL and the NS (which accelerates the stellar rotation). In our work we use the same approximation. A fraction of the kinetic energy of the accreting gas that goes to increase the rotational energy of the NS is approximately $2 \Omegans/\Omegak$, where $\Omegans$ is the NS angular velocity and $\Omegak$ is the Keplerian angular velocity at the NS surface. As we consider a non-rotating NS and the characteristic time to increase its angular velocity is orders of magnitude larger than the characteristic time of the SL $t= \Rns/\vtheta=10\ \mbox{km}/ 10^3\ \mbox{km s}^{-1} = 0.01$ s, ignoring the mechanical work on the NS is a reasonable approximation. In this approximation, therefore, all the work due to friction between the SL and the NS transforms to heat: \be \label{eq:heat} \int_0^{\hs} Q ^+ \d z = - (\tauphi\vphi + \tautheta \vtheta) = -\tau v_0. \ee \subsection{One-dimensional model of the spreading layer} \label{sec:structure} Using relations (\ref{eq:height})--(\ref{eq:heat}) we can rewrite equations, which describe the one-dimensional SL structure. The continuity equation can be rewritten via the accretion rate as \be \label{eq:cn} \dot{M} = 4 \pi R \ \cos\theta\ \vtheta \Sigmas, \ee Therefore, the product $\cos\theta \ \vtheta \Sigmas =const$. The Euler equations are \begin{eqnarray} \label{eq:fine1} \cos\theta\ \vtheta \Sigmas \vtheta' + \frac{4}{5} \cos\theta\ \left( \frac{\geff}{\gwr}\frac{k\TS}{\mu m_{\rm p}} \Sigmas\right)' = \\ \nonumber -R\cos\theta\ \tautheta - \sin\theta\ \Sigmas \vphi^2 , \end{eqnarray} \be \label{eq:fine2} \Sigmas\vtheta(\cos\theta\ \vphi)' = -R \cos\theta \tauphi. \ee Here the prime means the derivative over $\theta$. The second term in the left hand side of equation~(\ref{eq:fine1}) is the lateral force gradient, and the terms in the right hand side are the components of the stress force and the centrifugal force. We see from equation~(\ref{eq:fine2}) that the momentum along $\varphi$ coordinate is changed due to the friction with the NS surface only. The system of equations is closed by the energy equation \be \label{eq:en} \Sigmas\vtheta \left( \frac{v_0^2}{2} + \frac{2}{5} \frac{k\TS}{\mu m_{\rm p}}\frac{5\geff+3\grad}{\gwr} \right)' = -R q . \ee The system of equations~(\ref{eq:cn})--(\ref{eq:en}) can be transformed to the three dimensionless equations for $\vphi(\theta)$, $\vtheta(\theta)$ and $\TS (\theta)$ as was done by IS99. These equations are solved with the boundary conditions at the transition zone between the accretion disc and the SL: the initial latitude, where the SL starts, is close to the NS equator $\theta_{\rm 0} \sim $ 10$^{-2}$; the initial relative deviation $\delta$ of $\vphi$ from the Keplerian velocity $\vphi=v_{\rm K}(1-\delta)$; and the initial ratio of the kinetic energy of the SL along $\theta$ coordinate and it's thermal energy $\Theta \equiv (\mu m_{\rm p} v^2_{\theta_0})/k\TS $. As was demonstrated by IS99, the solution of equations~(\ref{eq:cn})--(\ref{eq:en}) depends very little on $\theta_0$ (if $\theta_0$ sufficiently small $<0.1$, see below) and $\delta$, but strongly depend on parameter $\Theta$. We choose the solutions which are closest to the critical solution (in this solution $\vtheta$ is equal to the sound speed at the maximum latitude of the spreading layer), but slightly subsonic. The necessary critical value of parameter $\Theta$ is found by the bisection method. The distributions of $\vtheta$, the effective temperature $\Teff$, and the surface density $\Sigmas$ along the latitude $\theta$ for four models with the same accretion rate, corresponding to first model luminosity are shown in Fig.~\ref{fig2}. The first model (shown by the solid curves) is our model with the pseudo-Newtonian potential and solar hydrogen abundance $X=0.7$. The dashed curves are for our model with GR correction, but with pure hydrogen $X=1$; the dotted curves correspond to our model without GR corrections ($\Rs=0$ in the equations) with solar hydrogen abundance ($X=0.7$), while the dot-dashed curves are for the IS99 model with GR corrections and solar hydrogen abundance. It is clear that the solar abundance lead to a narrower spreading layer with a smaller surface density. A higher helium abundance as well as the the GR corrections lead to a higher effective temperature and a larger latitudinal velocity. Our model gives a slightly wider SL with slightly smaller latitudinal velocity but same effective temperature and surface density. Most calculations below were performed for our model with the GR correction and solar abundances. The surface density and the effective temperature distributions along the latitude for models with 0.1, 0.2, 0.4 and 0.8 of the Eddington luminosity are presented in Fig.~\ref{fig3}. Variations of parameter $\alphab$ lead to some changes in the SL structure. The SL column density is inversely proportional to $\alphab$, while the resulting effective temperature decreases only by about 1 per cent with decreasing of $\alphab$ by an order of magnitude. The lower boundary of the SL was taken very close to the equator, $\theta_{\rm 0} \approx 0.01-0.001$, in IS99. Formally, the accretion disc thickness is close to zero at the inner boundary, if we take the usual inner boundary condition for the component of the stress tensor $W_{r\varphi}(R_{\rm in})=0$. In the case of the accretion disc around a NS this condition is not correct, and the disc thickness at the NS surface is considerable. The luminous accretion disc half-thickness can be evaluated from the balance of the radiation force and $z$-component of gravity: \be \label{z0} z_0 = \frac{3 \sigmae}{8 \pi c} \dot{M}. \ee We calculated the SL models with the initial latitudes $\theta_{01} = \arcsin (z_0/\Rns)$ and $\theta_{02} = \arcsin (0.5 z_0/\Rns)$. The surface density and the effective temperature distributions along the latitude for the second case are shown in Fig.~\ref{fig4} . The qualitative behavior of these distributions is close to the case of small $\theta_0$ with some shift along the latitude. The main difference is the maximum possible luminosity of the SL. At this luminosity the SL reaches the NS poles. For $\theta_{02}$, the maximum possible luminosity is about 0.4 $\Ledd$, while for $\theta_{01}$ it is about 0.25 $\Ledd$. \subsection{Vertical structure of the spreading layer} \label{sec:vertical} The IS99 SL model was constructed under an assumption that the local layer velocity and the radiation flux along height is constant. It means that the SL is decelerated and energy is liberated in the infinitely thin layer at the NS surface. This is an approximation only and the velocity distribution should not be uniform and the energy should be generated at all heights. Therefore, we need a more detailed vertical model of the SL for calculating its radiation spectrum. For evaluation of the viscosity parameter $\alphab$, IS99 used classical theory of hydrodynamic BLs \citep{LL59} with the logarithmic velocity and the energy generation distribution along the height. In this case, both the energy generation rate and the velocity gradient are inversely proportional to the distance from the NS surface $z$ \be \label{eq:lgr} \frac{\d q}{\d z} \propto \frac{\d v}{\d z} \propto \frac{1}{z}. \ee It is clear that these dependencies cannot be correct in a SL. The SL has a finite thickness with a low density at the surface. However according to equation~(\ref{eq:lgr}) some amount of energy has to be generated in the surface layers. At present time, a theory of the radiation-dominated turbulent boundary layer does not exist. Thus we here can only make similar assumptions about the energy generation and velocity gradient along the height. We assume that these values are inversely proportional to the surface density measured from the NS surface: \be \label{eq:qgr} \frac{\d q}{\d m} = -A~ \frac{q_{\rm 0}}{\Sigmas-m}, \ee \be \label{eq:vgr} \frac{\d v}{\d m} = -A~ \frac{v_0}{\Sigmas-m}, \ee where $A= 2.5 \alphab^{1/2}$, $q_{\rm 0}$ and $v_0$ are the local radiation flux and the average layer velocity at a given latitude obtained from the one-dimensional model. Equations (\ref{eq:qgr}) and (\ref{eq:vgr}) are very close to the IS99 SL model assumptions (the layer is decelerated and the energy is generated at the bottom of the layer). Integration of these equations yields \be \label{eq:q} q(m)=q_{\rm 0} \left[1+A~ \ln\left(1-\frac{m}{\Sigmas}\right)\right], \ee \be \label{eq:v} v(m)=v_0 \left[ 1+A~ \ln\left(1-\frac{m}{\Sigmas}\right)\right]. \ee These equations can be used up to some critical column density \be \label{eq:mstr} m_*=\Sigmas\left(1-\exp[-A^{-1}]\right), \ee which is very close to the local surface density. The hydrostatic equilibrium equation then reads \be \label{eq:hyd} \frac{\d P_{\rm g}}{\d m} = g_{\rm 0} - \frac{v^2(m)}{\Rns} -\frac{q(m)}{c}\sigmae \ee and the radiation transfer equation is \be \label{eq:rad} \frac{1}{3} \frac{\d\varepsilon}{\d m} = \frac{q(m)}{c}\sigmae. \ee The temperature and the gas pressure distributions along the height are: \begin{eqnarray} \label{eq:tm} T(m) &=& \Teff~ \left[ \frac{3}{4}m \sigmae \ \left(1-A\left[1+ \frac{\Sigmas-m}{m} \right. \right. \right. \\ \nonumber & \times& \left. \left. \left. \ln\left(1-\frac{m}{\Sigmas} \right) \right] \right) +\frac{1}{2}\right] ^{1/4} \end{eqnarray} \begin{eqnarray} \label{eq:pg} P_{\rm g}(m) & = & g_0 m - \frac{v^2_0}{\Rns} m\left(1-2A+2A^2\right) \\ \nonumber & +& \frac{v^2_0}{\Rns}A\left(\Sigmas-m\right) \ln\left(1-\frac{m}{\Sigmas}\right)\\ \nonumber &\times& \left[A~\ln\left(1-\frac{m}{\Sigmas}\right)-2A+2\right] \\ \nonumber &- & \frac{q_0\sigmae}{c}\left[m-A\left(m+\left(\Sigmas-m\right) \ln\left(1-\frac{m}{\Sigmas}\right)\right)\right]. \end{eqnarray} These solutions are obtained using the boundary conditions at the surface $\varepsilon(m=0)=2q_0/c$ and $P_{\rm g}(m=0)=0$. At the same time, there is a disagreement between this vertically explicit model and one-dimensional model, because the velocity and the flux vertical profiles are different. We suggest, that the model can be made self-consistent, if we find a new value of the surface density $\Sigmas'$ at a given latitude, which conserves the mass flux \be \label{eq:msfl} v_0 \Sigmas= \int_0^{\Sigmas'} v(m) \d m = v_0 (1-A) \Sigmas'. \ee Therefore, the new value of the surface density $\Sigmas' = \Sigmas/(1-A)$. For $\alphab = 10^{-3}$, this gives $\Sigmas' = 1.086 \Sigmas$ and we take these values below for our calculations. There are similar disagreements for other integrals over the height in equations (\ref{eq:cont2})--(\ref{eq:energy2}). For example: \be \label{eq:ke} v_0^2 \Sigmas= \int_0^{\Sigmas'} v^2(m) \d m = v_0 (1-2A+A^2) \Sigmas'. \ee In this case, we have to take a new value of the surface density $\Sigmas' = \Sigmas/(1-2A+A^2)$, which gives $\Sigmas' = 1.171 \Sigmas$ if $\alphab = 10^{-3}$. Our vertically explicit models disagree with the one-dimensional ones by about 10 per cent. Fortunately, the emitted local spectra depend very little on the surface density of the SL. \section{Spectrum of the spreading layer} For calculation of the SL spectra we divide it into a number of rings over the latitude which have different effective temperatures $\Teff$, matter velocities $v_0$ and surface densities $\Sigmas$. We then calculate the vertically explicit model for each ring, solve the radiative transfer equation and obtain the local SL spectrum. Then we integrate local spectra from the SL surface accounting for the general and special relativity effects. \subsection{Local spectra} To calculate a vertically explicit hydrodynamical model with the radiation transfer we use standard methods for stellar atmospheres modelling \citep{M78}. Our models are obtained in the hydrostatic and the plane-parallel approximations. The effective temperatures of the considered SL models are rather high ($\sim$ 2 keV) and these models are similar to the atmospheres of bursting NSs, where Compton scattering have to be taken into account. The vertically explicit local SL model is described by the following equations: the equation of hydrostatic equilibrium (\ref{eq:hyd}), the energy generation law (\ref{eq:q}), the velocity law (\ref{eq:v}), the RTE accounting for the Compton effect using the \citet{K57} operator: \begin{eqnarray} \label{eq:rtr} \frac{\partial^2 ( f_{\nu} J_{\nu})}{\partial \tau_{\nu}^2} = \frac{k_{\nu}}{k_{\nu}+\sigmae} \left(J_{\nu} - B_{\nu}\right) - \frac{\sigmae}{k_{\nu}+\sigmae} \frac{kT}{\me c^2} \times \\ \nonumber x \frac{\partial}{\partial x} \left(x \frac{\partial J_{\nu}}{\partial x} - 3J_{\nu} + \frac{\Teff}{T} x J_{\nu} \left[ 1 + \frac{CJ_{\nu}}{x^3} \right] \right), \end{eqnarray} where $x=h \nu /k\Teff$ is dimensionless frequency, $f_{\nu}(\tau_{\nu}) \approx 1/3$ is the variable Eddington factor, $J_{\nu}$ is the mean intensity of radiation, $B_{\nu}$ is the black body (Planck) intensity, $k_{\nu}$ is the opacity due to the free-free and bound-free transitions, $\sigmae$ is the electron (Thomson) opacity, $T$ is the local electron temperature, $\Teff$ is the effective temperature of SL at a given latitude, and $C=c^2 h^2/2(k\Teff)^3$. The optical depth $\tau_{\nu}$ is defined as \be \d \tau_{\nu} = (k_{\nu}+\sigmae) \d m. \ee These equations have to be completed by the energy balance equation \begin{eqnarray} \label{eq:econs} \int_0^{\infty} k_{\nu}\left(J_{\nu} - B_{\nu}\right) \d\nu - \frac{1}{4\pi}\frac{\d q}{\d m} - \sigmae \frac{kT}{\me c^2} \times \\ \nonumber \left[ 4 \int_0^{\infty} J_{\nu} \d\nu - \frac{\Teff}{T} \int_0^{\infty} x J_{\nu} \left( 1+\frac{CJ_{\nu}}{x^3}\right) \d\nu \right]=0 \end{eqnarray} and by the ideal gas law \be \label{eq:gstat} P_{\rm g} = N_{\rm tot} kT, \ee where $N_{\rm tot}$ is the number density of all particles, as well as by the particle and charge conservation laws. We assume local thermodynamical equilibrium (LTE) in our calculations, so the number densities of all ionization and excitation states of all elements have been calculated using Boltzmann and Saha equations. For solving these equations and computing the local SL model we used the Kurucz's code {\sc ATLAS} \citep{K70,K93} modified for high temperature. All ionization states of the 15 most abundant elements are taken into consideration. The photoionization cross-sections from the ground states of all ions are calculated using {\sc phfit2} code \citep{V96}. For details see \citet{SGS02} and \citet{I03}. The code was also modified to account for Compton scattering. The scheme of calculation is the following. First, the input parameters of the local SL model are defined from the total one-dimensional SL model (see Sect. \ref{sec:structure}): the effective temperature $\Teff$, the surface gravity $g_0$, the surface density $\Sigmas$, and the local average layer velocity $v_0$. Then the analytical vertically explicit model (\ref{eq:tm}--\ref{eq:msfl}) are calculated together with the new value of surface density $\Sigmas'=\Sigmas/(1-A)$. The calculations are performed for the set of 98 column densities $m$, distributed logarithmically with equal steps from $m= 10^{-5}$ g cm$^{-2}$ to $0.99 m_{*}$. The gas pressure, which is found from equation~(\ref{eq:pg}), is not varied during the iterations. For this starting model, all number densities and the opacities at all depth points and all the frequencies (we use 300 logarithmically equidistant frequency points) are calculated. The RTE (\ref{eq:rtr}) is solved by the Feautrier method \citep{M78,ZS91,PSZ91,GS02} iteratively, because it is non-linear. Between the iterations we calculate the variable Eddington factors $f_{\nu}$ and $h_{\nu}$, using the formal solution of the RTE for three angles. Usually 5--6 iterations are sufficient to achieve convergence. We used the usual condition at the outer boundary \be \frac{\partial ( f_{\rm \nu} J_{\nu}) }{\partial \tau_{\nu}} = h_{\nu} J_{\nu}, \ee where $h_{\nu}$ is the surface variable Eddington factor, and the inner boundary condition \be \frac{\partial J_{\nu}}{\partial \tau_{\nu}} = \frac{\partial B_{\nu}}{\partial \tau_{\nu}}. \ee The outer boundary condition is found from the lack of the incoming radiation at the SL surface, and the inner boundary condition is obtained from the diffusion approximation $J_{\nu} \approx B_{\nu}$ and $q_{\nu} \approx 4\pi/3 \times \partial B_{\nu}/\partial \tau_{\nu}$. This condition is satisfied for any SL optical thickness, because the SL bottom is the NS surface. The boundary conditions along the frequency axis are \be \label{eq:lbc} J_{\nu} = B_{\nu} \ee at the lower frequency boundary, $\nu=\nu_{\rm min}=10^{14}$ Hz ($h\nu_{\rm min} \approx$ 0.03 eV $\ll k\Teff$) and \be \label{eq:hbc} x \frac{\partial J_{\nu}}{\partial x} - 3J_{\nu} + \frac{\Teff}{T} x J_{\nu} \left( 1 + \frac{CJ_{\nu}}{x^3} \right)=0 \ee at the higher frequency boundary $\nu=\nu_{\rm max}=3\ 10^{19}$~Hz ($h \nu_{\rm max} \approx$ 100 keV $\gg k\Teff$). Condition (\ref{eq:lbc}) means that at the lowest energies the true opacity dominates the scattering $k_{\nu} \gg \sigmae$, and therefore $J_{\nu} \approx B_{\nu}$. Condition (\ref{eq:hbc}) means that there is no photon flux along the frequency axis at the highest energy. The solution of the RTE (\ref{eq:rtr}) should also satisfy the energy balance equation (\ref{eq:econs}) and the surface flux condition \be \int_0^{\infty} q_{\nu} (m=0) \d\nu = \sigmasb \Teff^4. \ee We calculated the relative flux error along the depth \be \varepsilon_{q}(m) = 1 - \frac{q(m)}{\int_0^{\infty} q_{\nu} (m) \d\nu}, \ee where $q(m)$ is found from the energy generation law (\ref{eq:q}), and $q_{\nu} (m)$ is radiation flux at a given depth obtained from the first moment of the RTE \be 4\pi \frac{\partial (f_{\nu} J_{\nu})}{\partial \tau_{\nu}} = q_{\nu}. \ee Then the temperature corrections were evaluated using three different procedures. The first procedure is the integral $\Lambda$-iteration method based on the energy balance equation (\ref{eq:econs}) which was modified to account for Compton scattering. It works well in the upper layers. The second one is the modified Avrett-Krook flux correction, which uses the relative flux error and is good in deep layers. And the third one is the surface correction, which is based on the emergent flux error. See \citet{K70} for the detailed description of the methods. The iteration procedure is repeated until the relative flux error is smaller than 1 per cent, and the relative flux derivative error is smaller than 0.01 per cent. As a result we obtain the self-consistent local SL model together with the emergent spectrum of radiation. Our method of calculation was checked on the atmosphere model of bursting NS. The equations which describe the bursting atmosphere are simpler, because there is no velocity field along the surface ($v_0=0$) and the integral flux is constant along depth ($\d q/\d m=0$). We compared our model atmospheres with the most recent models of \citet*{MJR04}. The radiation spectra and the temperature structure for some models with $\Teff=2\ 10^7$~K, solar H/He abundances, and various surface gravities are shown in Fig.~\ref{fig5}. These results are in a perfect agreement with the results of \citet{MJR04}. The emergent spectra and the temperature structure for the models with the solar abundance of heavy elements are shown in Fig.~\ref{fig6}. In the surface layers, local cooling is small because of the low density, and the temperature equals the Compton temperature of radiation which is slightly higher than the effective temperature. In dipper layers, at $\taue\sim 0.1$, the cooling due to thermal emission (free-free and bound-free) becomes important (as thermal emissivity per gram is proportional to density) and the temperature decreases. At large optical depth the temperature rises again and follows the $\taue^{1/4}$ relation, typical for a grey atmosphere. At higher surface gravity (at fixed $\Teff$), the plasma density is higher, resulting in a more significant temperature dip. We see that heavy elements have rather minimal influence on the models close to the Eddington limit (lower~$g$). The comparison between the bursting NS models and different local SL models for the same $\Teff$ and effective $\log g$ is shown in Fig.~\ref{fig7}. For this SL model we use the vertical structure model, which is described in Section~\ref{sec:vertical}. We also investigated, whether the model for the vertical structure is important for the emergent spectra of local SL models. We also calculated the SL model with the constant velocity and flux derivatives: \be \frac{\d v}{\d m} = -\frac{v_0}{\Sigmas} \ee and \be \frac{\d q}{\d m} = -\frac{q_0}{\Sigmas}. \ee In this case the mass flux conservation requirement (\ref{eq:msfl}) leads to $\Sigmas'= 2 \Sigmas$. The spectrum and the temperature structure of this model are shown in Fig.~\ref{fig7} by squares. The surface temperatures of the local SL models are higher than the bursting NS model surface temperature. The reason is the non-zero flux derivative in the energy conservation equation (\ref{eq:econs}). This means that a part of the energy is released in the upper atmosphere and is heating it additionally. The smaller the surface density (i.e. the larger the flux derivative), the higher the surface temperature. But the differences in the temperature structure have very small influence on the emergent spectra. Therefore we conclude, that details of the vertical structure have negligible influence on the emergent spectrum for the optically thick models ($\Sigmas\ge 100\ \mbox{g cm}^{-2}$). It is well known that the model spectra of bursting NS close to the Eddington limit are well described by a diluted Planck spectrum with the color temperature $\Tc = \fc \Teff$ with the hardness factor $\fc$ varying in the interval 1.6--1.9 and the dilution factor $D=\fc^{-4}$. \citet{PSZ91} have derived an analytical formula for the hardness factor, which successfully describes high luminosity ($L \approx \Ledd$) burst spectra: \be \label{eq:fc} \fc = \left( 0.15 \ \ln C_1+0.59 \right)^{-4/5} C_1^{2/15} \ell^{3/20}, \ee where $C_1=(3+5X)/(1-\ell)$ and $\ell=L/\Ledd=\grad/\geff$. Equation (\ref{eq:fc}) works well also for models with heavy elements. The local spectra of the optically thick SL (with $L > 0.2 \Ledd$) are very similar to the burst spectra with corresponding parameters (see Fig. ~\ref{fig7}). The local SL are very close to the Eddington limit \be \label{eq:grgeff} \grad = \frac{q_0}{c} \sigmae \approx g_0 - \frac{v_0^2}{\Rns} = \geff. \ee For example, the distributions of the ratios $\grad/g_0$ and $\grad/\geff$ along the latitude for SL models with three different luminosities are shown in Fig.~\ref{fig8}a. Corresponding hardness factor distributions are shown in Fig.~\ref{fig8}b. The comparison of the two local SL spectra (close to equator and at higher latitude) with the diluted Planck spectra and hardness factors given by equation (\ref{eq:fc}) are shown in Fig.~\ref{fig9}a. Closer to the equator, effective gravity is low as centrifugal force is large. The gas is levitating above the NS and $\ell$ is close to unity. The energy dissipation and the effective temperature are low. Thus, $\fc$ is large and the spectrum is close to the diluted Planck. At higher latitude, the layer is decelerated, while the energy dissipation and $\Teff$ grow. However, the effective gravity grows faster reducing $\ell$ and the color correction $\fc$. The spectrum shows deviations from the diluted Planck spectrum at low energies. At high energies, the Wien part of both spectra can well be described by the diluted Planck. \subsection{Integral spectra} Now we can compute the integral total model spectrum of the SL, which is seen by a distant observer accounting for the relativistic effects such as gravitational redshift, light bending, relativistic Doppler effect and aberration. We take into account only half of the SL because another half is hidden by the accretion disc and divide the SL surface on 10 latitude rings and on 100 angles in azimuth. In a spherical coordinate system, where the accretion disc coincides with the $\theta = 0\degr$ plane, the spectrum of the SL is \citep{PG03} \be \label{eq:totsp} F_{\rm E} = \frac{\Rns^2}{D^2} \int\limits_0^{\theta_{\rm SL}} \int\limits_0^{2 \pi} \eta^3 \delta^3 I(E',\cos \alpha', \theta) \cos \alpha' \cos \theta\ \d\theta\ \d\varphi. \ee Here the observed and the emitted photon energies are connected by the relation $ E=E' \ \eta\ \delta$, where $\eta = \sqrt{1-\Rs/\Rns}$, the Doppler factor $\delta = 1/\gamma (1-\beta \cos\xi)$, $ \beta=\vphi(\theta)/c$ (here we neglected low latitudinal velocity), the Lorentz factor $\gamma =1/\sqrt{1-\beta^2}$, and $ \cos\xi = - \sin\alpha\ \sin i\ \sin\varphi/\sin\psi$. The light bending is accounted for by the relation \citep{B02} \be \cos\alpha = \frac{\Rs}{\Rns} + \eta^2 \cos\psi, \ee where $\cos\psi = \cos i \sin\theta + \sin i \cos\theta \cos\varphi$, and the relativistic aberration gives $ \cos\alpha' = \delta\cos\alpha$ \citep{PG03}. Here $i$ is the inclination angle of the NS polar axis to the line of sight, $D$ is the distance to the observer, and $\theta_{\rm SL}$ is the SL boundary. Only visible surface elements with $\cos\alpha > 0$ give contribution to the total spectrum. The emitted specific intensity $I(E',\cos\alpha', \theta)$ is taken from the computed local SL flux assuming angular dependence for the electron scattering atmosphere \be \label{eq:ang} I(E',\cos\alpha', \theta) = \frac{q_{\rm E'}(\theta)}{\pi} (0.4215+0.86775 \cos\alpha'). \ee This formula gives a good approximation to the specific intensity of the emergent radiation (see Fig.~\ref{fig10}). The total spectra of the SL model for two inclination angles, $i=0{\degr}$ and $90{\degr}$, are shown in Fig.~\ref{fig9}b. The spectra computed using the local diluted Planck spectra are shown also for comparison. The difference is very small in the high energy part ($E> 10$ keV) and more significant at lower energies ($E < 7$ keV). Dependence of spectral shape on the inclination angle is not significant (see also Fig.~\ref{fig11}a). Differences between the SL spectra, which are seen at different inclinations are comparable to the differences due to change in the SL luminosities (see Fig.~\ref{fig11}b). It is interesting, that the total spectra can also be well described by the diluted Planck spectrum. The color temperature depends slightly on the assumed turbulence parameter $\alphab$. Decreasing $\alphab$ by an order of magnitude increases $\Tc$ by 0.1 keV. \section{Comparison with observations} \label{sec:obs} In LMXRBs a weakly magnetized NS is surrounded by the accretion disc which transforms to the boundary/SL close to the NS surface. At present, about 100 LMXRBs are known. They can be divided into two different classes. The Z-sources are very luminous ($L \sim 0.1 - 1 \Ledd$) and have relatively soft, two-component spectra. Both components are close to the black body with color temperatures of about 1 keV and 2--2.5 keV. The atoll sources are less luminous ($L \sim 0.01 - 0.05 \Ledd$) and are observed in two states, the high/soft and the low/hard. In the soft state, the radiation spectra are similar to those of the Z-sources, while in the hard state they are close to the spectra of the Galactic black hole sources in the hard states \citep[e.g. Cyg X-1, see e.g.][]{P98,B00}. These hard spectra are well described by unsaturated Comptonization of soft photons in the hot ($kT \sim$ 30 -- 100 keV) optically thin ($\taue\sim $ 1) plasma. The soft component can be associated with the radiation from the accretion disc, while the hard one with the boundary/SL \citep{M84} or possibly with a corona or hot optically thin inner accretion flow \citep[see discussion in][]{DG03} in case of low-luminosity atoll sources. At high luminosities, the BL is optically thick and its effective temperature is higher than that of the accretion disc, because the BL is smaller than the accretion disc, while their luminosities are comparable. The hard component is also more variable than the soft component at the timescales from millisecond to 1000 seconds \citep*{M84,GRM03}. The Fourier-frequency resolved spectroscopy confirms that a component variable at high frequencies (and sometimes showing quasi-periodic oscillations, see \citealt{vdK00}) has a blackbody-like spectrum with the color temperature $\Tc= 2.4\pm 0.1$ keV \citep{GRM03,RG06} which is very similar for the five investigated sources. On the other hand, the variability of the soft component is very similar to the variability of the soft component of black hole sources in their soft states, which is associated with the accretion disc. Based on these arguments, we associate the hard blackbody-like component with the BL and compare our theoretical SL spectra with it. Spectra computed for one SL model together with the observed BL spectra obtained by the Fourier-frequency resolved spectroscopy \citep{GRM03,RG06} are shown in Fig.~\ref{fig12}. We see a very good agreement between theoretical spectra and the spectrum of GX 340+0 at the normal brach (at high accretion rates). The spectra of five Z- and atoll-sources (open circles) are similar to our SL spectra at high energies, but have a soft excess. This excess may be related to the emission of the classical BL, the inner part of the accretion disc. The observed spectral similarity gives us a confidence to try to determine NS parameters from the observed spectra. As we have shown above the spectrum of the SL can be represented by a diluted blackbody. The effective temperature of radiation is determined by the critical temperature from equation (\ref{eq:tc}), where the left-hand side is multiplied by $\ell$, the ratio of the local flux to the critical Eddington one (reduced due to the action of the centrifugal force). The observed color temperature is $\Tc=\fc \sqrt{1-\Rs/\Rns}\ \Tcr$, where corrections are made for spectral hardening and gravitational redshift. For the known color correction and $\ell$, the NS radius as a function of compactness $\Mns/\Rns$ can then be found from \be \label{eq:eos} \Rns= \frac{ \ell \fc^4 c^3} { 2 \sigmasb \Tc^4 \sigmae} \frac{\Rs}{\Rns} \left( 1- \frac{\Rs}{\Rns} \right) ^{3/2} . \ee Assuming $\fc=1.6-1.8$ and $\ell=0.8$, \citet{RG06} obtained constraints on the NS mass-radius relation (shown in Fig. \ref{fig13} by dotted curves). The maximum NS radius is reached for $\Rs/\Rns=2/5$: \be \label{eq:rmax} \Rns_{\max}= \frac{24.6}{1+X} \frac{ \ell}{0.8} \left( \frac{ \fc}{1.7} \right)^4 \left( \frac{ \Tc}{2.4\ \mbox{keV}} \right)^{-4} \ \mbox{km} . \ee Here instead we calculate exactly a grid of the SL model spectra, where the main input parameters are the NS mass $\Mns$ and radius $\Rns$, and the SL luminosity. The NS mass is varied from 1 to 2 $\msun$ with a step of 0.2 $\msun$, and the NS radius is varied from 10 to 24 km with a step 1 km. Only the models with $\Rns > 3 \Rs$ are considered. We take $\alphab=10^{-3}$ and luminosity of $0.4 \Ledd$, and compute spectra for four inclination angles 0, 30, 60 and 90 degrees and for three chemical compositions: pure hydrogen ($X=1$), solar abundance ($X=0.7$), and pure helium ($X=0, Y=1$). The spectra are fitted by the black body and the corresponding color temperatures are found. Models with higher He abundance have a smaller hardness factor as can be seen from equation (\ref{eq:fc}). However, the local effective temperature of the SL is higher for larger He abundance (see eq. \ref{eq:tc} and Fig. \ref{fig2}c). The higher $\Teff$ leads to a higher color temperature of the integral SL spectrum. For example, at NS radius of 13 km and mass $1.4\msun$ pure hydrogen models give color temperature of about 2.5 keV, while pure helium models produce harder spectra with $\Tc\approx 3$ keV. Contours corresponding to the color temperature equal 2.3 (right), 2.4 (central) and 2.5 keV (left) are shown on the $\Mns-\Rns$ plane (Fig.~\ref{fig13}) together with the NS models for various equations of state. These iso-temperature curves are shown for the inclination angle $i=45{\degr}$. Comparison of the observed spectra to the theoretical spectra of the SL constrains the NS radius at $13.5 \pm 1.5$ km (for pure hydrogen $X=1$ model), $14.8 \pm 1.5$ km (solar composition $X=0.7$) and $19 \pm 1.5$ km (pure helium $X=0, Y=1$) assuming the NS mass of 1.4 solar mass. For pure hydrogen and solar abundance, the permitted radii are consistent with the hard equation of state of the NS matter. If the composition is solar, but the heavier elements are able to sink, the emitted spectra would correspond to a pure hydrogen atmosphere requiring thus smaller radii. Increasing the inclination to $90{\degr}$ increases the deduced NS radii by about 10 per cent, while assuming $i=0{\degr}$, gives a 15 per cent reduction on $R$. The uncertainty in the luminosity increases the width of possible NS radii by about 50 per cent. Another source of uncertainty comes from the turbulence parameter $\alphab$. With $\alphab$ decreasing by an order of magnitude the spectrum hardens by 0.1 keV. This results in about 15 per cent decrease of the NS radius that is required to produce the observed spectra. Thus $\alphab\sim 10^{-5}$ is needed to reconcile the derived NS radii with the soft equations of state (assuming solar composition). Such a small $\alphab$ at the same time yields a very large column density of the SL and a rather long life-time of the accreting gas in the layer (of the order of 1 s, instead of 10 ms as in the model of IS99). Finally, we would like to emphasize that our method of determination of the NS radius from the SL spectrum is based on the observed color temperature of radiation alone, because the SL radiates locally at almost Eddington flux. The color temperature can be related to the effective temperature which is a function of the stellar compactness (and chemical composition) as given by equation~(\ref{eq:tc}). This method is identical to that used for the radius-expansion X-ray bursts which are believed to reach Eddington luminosity \citep*[see e.g.][]{LvPT93}. In contrast to the standard methods based on the modeling of the thermal emission from the NS surface \citep[see for example][]{vpL87,T05}, there is no need to know precisely either the area of the emitting region, or the distance to the source. As the standard method gives the apparent stellar radius at infinity, which is related to the NS parameters through \be \label{eq:rmrinf} R_{\infty} = \Rns \left( 1- \frac{\Rs}{\Rns} \right)^{-1/2} , \ee the allowed band of $\Rns$ and $\Mns$ is nearly orthogonal to that obtained from the color temperature and equation (\ref {eq:eos}) (see the almost vertical dashed curve in Fig. \ref{fig13}). Thus for a NS, where both the thermal emission from the surface (e.g. during the quiescence) and the BL emission (during the accretion phase) are observed, it would be possible to determine $\Rns$ and $\Mns$ independently. Interestingly our constraints on the NS radius are very similar to those obtained by \citet{HR05} for the thermally emitting quiescent NS X7 in the globular cluster 47 Tucanae $\Rns=14.5^{+1.8}_{-1.6}$ km. They are also consistent with the lower limit $\Rns>14$ km obtained by \citet{T05} for the isolated NS RX~J1856-3754. \section{Conclusions} We have derived the one-dimensional equations describing the SL model on a spherical NS surface from the usual hydrodynamic equations. The obtained equations are similar to those in IS99, except for the energy conservation law where we neglected the surface density of the gravitational potential energy which is of the second order in $H/R$. This difference, however, leads only to small quantative changes. We have also implemented a pseudo-Newtonian potential to account for the main general relativity corrections and considered various chemical compositions of the accreting matter. We have studied the vertical (radial) structure of the SL with different assumptions about the vertical distributions of the radiation flux and azimuthal velocity. The temperature structure and the emergent radiation spectra of the SL are computed accounting for the effect of Compton scattering. We showed that the local (at a given latitude) emergent spectra depend very little on details of the SL vertical structure in optically thick cases with $\Sigmas\gtrsim 100\ \mbox{g\ cm}^{-2}$ ($L \gtrsim 0.1 \Ledd$). These spectra can be described by the diluted Planck spectrum and are similar to the spectra of X-ray bursts with the same effective temperature and the effective surface gravity. The integral SL spectra were computed accounting for relativistic effects such as the gravitational redshift and light bending, the relativistic Doppler effect and aberration. These spectra slightly depend on the inclination angle to the line of sight and on the SL luminosity. The local effective temperature increases with latitude, while the hardness factor $\fc$ decreases. This leads to only slight variation of the color temperature on latitude. As a result, the integral spectra can also be well described by a single-temperature diluted Planck spectrum. We compared our theoretical integral SL spectra with the observed spectra of the LMXRBs BLs. The observed color temperature of 2.4 $\pm$ 0.1 keV \citep{GRM03,RG06} can be reproduced for hard equations of state of NS material. Our model constrains radii of NSs in LMXRBs to 13--16 km for a 1.4 solar mass star. Soft equations of state (smaller NS radii) can be reconciled with the observed spectra only for very low viscosity $\alphab\sim10^{-5}$. Calculation of $\alphab$ from the first principles is a challenging problem that deserves further attention. \section*{Acknowledgments} This work was supported by the Academy of Finland grants 107943 and 102181, the Jenny and Antti Wihuri Foundation, RFBR grant 05-02-17744, and the Russian President program for support of the leading science school (grant Nsh - 784.2006.2). We are grateful to M. Revnivtsev for providing us with the spectral data, and to D. G. Yakovlev and P. Haensel for the theoretical mass-radius relations for neutron and strange stars. We thank the referee for useful comments. \label{lastpage}
Title: Spitzer Reveals Hidden Quasar Nuclei in Some Powerful FR II Radio Galaxies
Abstract: We present a Spitzer mid-infrared survey of 42 Fanaroff-Riley class II radio galaxies and quasars from the 3CRR catalog at redshift z<1. All of the quasars and 45+/-12% of the narrow-line radio galaxies have a mid-IR luminosity of nuLnu(15 micron) > 8E43 erg/s, indicating strong thermal emission from hot dust in the active galactic nucleus. Our results demonstrate the power of Spitzer to unveil dust-obscured quasars. The ratio of mid-IR luminous narrow-line radio galaxies to quasars indicates a mean dust covering fraction of 0.56+/-0.15, assuming relatively isotropic emission. We analyze Spitzer spectra of the 14 mid-IR luminous narrow-line radio galaxies thought to host hidden quasar nuclei. Dust temperatures of 210-660 K are estimated from single-temperature blackbody fits to the low and high-frequency ends of the mid-IR bump. Most of the mid-IR luminous radio galaxies have a 9.7 micron silicate absorption trough with optical depth <0.2, attributed to dust in a molecular torus. Forbidden emission lines from high-ionization oxygen, neon, and sulfur indicate a source of far-UV photons in the hidden nucleus. However, we find that the other 55+/-13% of narrow-line FR II radio galaxies are weak at 15 micron, contrary to single-population unification schemes. Most of these galaxies are also weak at 30 micron. Mid-IR weak radio galaxies may constitute a separate population of nonthermal, jet-dominated sources with low accretion power
https://export.arxiv.org/pdf/astro-ph/0601485
\title{Spitzer Reveals Hidden Quasar Nuclei in Some Powerful FR II Radio Galaxies} \author{Patrick Ogle} \affil{Spitzer Science Center, California Institute of Technology, Mail Code 220-6, Pasadena, CA 91125} \email{ogle@ipac.caltech.edu} \author{David Whysong\altaffilmark{1} \& Robert Antonucci} \affil{Physics Dept., University of California, Santa Barbara, CA 93106} \altaffiltext{1}{now at NRAO, Array Operations Center, P. O. Box O, 1003 Lopezville Rd., Socorro, NM 87801-0387} \shorttitle{Hidden Quasar Nuclei} \shortauthors{Ogle et al.} \keywords{galaxies: active, galaxies: quasars, galaxies: jets, infrared: galaxies} \section{Unification of Quasars and Radio Galaxies} The nature of the energy source in active galactic nuclei (AGNs) is a fundamental problem. The basic model attributes the large luminosity of these systems to gravitational energy release in an accretion disk around a supermassive black hole. A jet may be driven by magnetic fields threading the disk \citep{bp82}. The black hole spin energy may also be tapped and converted into electromagnetic Poynting flux and particles in a relativistic jet \citep{bz77,pc90,m99,d04}. Extragalactic radio sources are categorized by their morphology as either of two types \citep{fr74}. Fanaroff-Riley (FR) type I sources are edge-darkened, while FR IIs are edge-brightened. The different morphology of FR Is indicates that they are not related to FR IIs by orientation. FR Is also have lower radio luminosities than FR IIs for a given host galaxy luminosity \citep{ol94,b96} and most have low-ionization nuclear emission region (LINER) spectra \citep{hl79}. However, not all FR Is can be characterized by low accretion power \citep{wa04,cr04}. The present paper focuses on FR IIs, which contain powerful jets with bright terminal hot spots and lobes. Furthermore, we count broad-line radio galaxies (BLRGs) as low-luminosity quasars. Quasars and narrow-line radio galaxies (NLRGs) may be unified by orientation-dependent obscuration. Radio galaxies are thought to host quasar nuclei that are obscured by circumnuclear dusty tori aligned with the radio jets \citep{ant84} . Unification of radio galaxies and quasars can therefore explain the lack of quasars viewed at large angles to the radio axis \citep{b89}. The percentage of high-redshift radio galaxies (60\% of the 3CRR FR II sample at $z>0.5$) would then indicate a torus covering fraction of $\sim 0.6$. However, there appears to be a discrepancy between the redshift distributions of quasars and radio galaxies at $z<0.5$, with a factor of $\sim 4$ more narrow-line radio galaxies than quasars \citep{s93}. Furthermore, the median projected linear size of these 'excess' radio galaxies is smaller than expected for quasars seen in the sky plane \citep{s93,w05}. The unification hypothesis may be modified to include a second population of lower luminosity, low-excitation FR II radio galaxies \citep{wj97,grw04}. Alternatively, it has been argued that the torus covering fraction may increase with decreasing radio luminosity \citep{l91}. The unification hypothesis has been qualitatively confirmed by spectropolarimetry of radio galaxies, many of which have been shown to have highly polarized broad emission lines and blue continuum, scattered from material which has a direct view of the active galactic nucleus \citep{cdb97, cot99}. Of particular note are the original discovery of highly polarized broad H$\alpha$ from the hidden quasar nucleus in 3C 234 \citep{ant84}, and the discovery of highly polarized broad H$\alpha$ in the spectrum of the powerful radio galaxy Cygnus A \citep{ocm97}. However, this method of detecting hidden quasars relies on an appropriately placed scattering region to view the otherwise hidden nucleus. Such a region is not guaranteed to exist for all radio galaxies, and thus spectropolarimetry can easily yield false negatives. Polarimetry is also ineffective at determining the luminosity of the hidden nucleus, since the scattering efficiency is usually unknown. Another way to search for hidden quasar nuclei is to observe radio galaxies in the mid-IR. If the unification hypothesis is correct, the dusty torus should serve as a crude calorimeter of the central engine \citep{mhm01,sfk04,hmb04,wa04}. Optical, UV, and X-ray photons from the quasar nucleus are absorbed by dust in the torus and the energy is re-emitted in the thermal infrared. This explains why blue, UV color-selected quasars emit 10-50\% of their luminosity in the IR \citep{spn89,hmc00}. There appears to be no connection between the bulk of this IR emission and nonthermal radio emission, except in core-dominated radio sources such as blazars. Observations of matched 3CR quasars and radio galaxies by ISO indicate similar IR luminosities, consistent with the unification picture \citep{mhm01,hmb04}. However, differences in 24 $\mu$m/ 70 $\mu$m color may indicate that mid-IR emission from the torus is anisotropic by a factor of $\le 3$ \citep{srh05}. We present {\it Spitzer} observations of a sample of 42 FR II radio galaxies and quasars selected from the 3CRR survey. The goals are to search for mid-IR emission from hidden quasar nuclei and test the ubiquity of the unification hypothesis. The {\it Spitzer} Infrared Spectrograph (IRS) combines the advantages of unprecedented sensitivity from 5-36.5 $\mu$m to measure the mid-IR continuum and spectral resolution to measure high ionization emission lines powered by hidden AGNs. In the current paper, we present evidence for hidden quasar nuclei based on mid-IR photometry extracted from the IRS spectra. We examine in detail the spectra of the subset of 14 mid-IR luminous radio galaxies which appear to contain hidden quasar nuclei. Spectra of the quasars and mid-IR weak radio galaxies and a statistical study of the complete sample will be presented in separate papers. \section{Sample} We begin by selecting a well-defined, radio flux-limited and redshift-limited sample of 55 radio galaxies and quasars from the 3CRR catalog \citep{lrl83}. We include all 3CRR sources with FR II radio morphology, a flux of $S_{178}>16.4$ Jy\footnote{The radio flux limit is 15 Jy using \cite{lrl83} flux values and 16.4 Jy on the standard \cite{b77} scale.} at 178 MHz, and a redshift of $z<1$. The original 3CRR catalog has a flux limit of 10 Jy at 178 MHz, is restricted to northern declinations ($\delta >10 \arcdeg$), and has galactic latitude $|b|>10\arcdeg$. It is the canonical low-frequency selected catalog of bright radio sources, has optical identifications and redshifts for all entries, and has been extensively observed in most wavebands. We select only sources with FR II radio morphology. We verify or update the FR classification of all sources by inspection of the latest published radio maps. Compact, steep-spectrum sources (CSSs: 3C 48, 138, 147, 286, and 309.1) with radio major axis $D<10$ kpc \citep{ffp85} are excluded from the sample because they may constitute a class of young or frustrated radio sources. Here and throughout this paper, we assume a cosmology with $H_0=70$ km s$^{-1}$ Mpc$^{-1}$, $\Omega_\mathrm{m}=0.3$, and $\Omega_\Lambda=0.7$. Size and morphology indicate that CSSs are not related to FR IIs by orientation. It is essential for our unification studies that we select a sample based on isotropic radio lobe flux, and {\it not} on optical or IR properties, so that it is unbiased by orientation-dependent selection effects. In particular, our sample includes no blazars. No sources make the flux limit only because of beamed emission from the core of the radio jet. Our sample includes quasars as well as radio galaxies, and we use the quasar subsample as a control. We aim to determine whether and which narrow-line FR II radio galaxies have mid-IR power comparable to quasars or broad-line radio galaxies of similar radio lobe flux and redshift. The 42/55 sources in our sample which we have observed with {\it Spitzer}, or which have {\it Spitzer} data in the public archive are listed in Tables 1 and 2. The 25 {\it mid-IR luminous} sources with $\nu L_\nu(15\mu\mathrm{m}) >8\times 10^{43}$ erg s$^{-1}$ (14 NLRGs and 11 quasars or BLRGs) are listed in Table 1, and the 17 {\it mid-IR weak} galaxies with $\nu L_\nu(15\mu\mathrm{m}) <8\times 10^{43}$ erg s$^{-1}$ are listed in Table 2. The reason for this particular division is explained below. Optical source classifications are based on emission line properties. Type 1 sources have directly visible broad emission lines (quasars and BLRGs), and type 2 sources (NLRGs) do not. The NLRGs are further classified using their forbidden emission lines \citep{jr97,wrb99}\footnote{Updated optical classifications are available at http://www-astro.physics.ox.ac.uk/$\sim$cjw/3crr/3crr.html.}. High-excitation galaxies (HEGs) are defined to have [O {\sc iii}] $\lambda$5007 equivalent widths of $>10$ \AA~ and [O {\sc iii}] $\lambda$5007/[O {\sc ii}] $\lambda$3727 $>1$. The sources which do not meet these criteria are classified as low-excitation galaxies (LEGs). The equivalent width criterion ensures that [O {\sc iii}] is measurable in moderate S/N spectra. However, it remains to be seen whether some sources with low [O {\sc iii}] equivalent width might have [O {\sc iii}] $\lambda$5007/[O {\sc ii}] $\lambda$3727 $>1$. In addition, we caution that [O {\sc ii}] and [O {\sc iii}] may be subject to differing amounts of extinction. \section{Observations} We observed the sources in our sample with the Infrared Spectrograph (IRS) on the {\it Spitzer} Space Telescope \citep{h04,w04}. We used the low-resolution ($R\sim 64-128$) modules Short-Low (SL) and Long-Low (LL) for accurate spectrophotometry over the wavelength range 5-36.5 $\mu$m. Wavelengths 36.5-40 $\mu$m are unusable because of low-S/N and 2nd-order bleed-through caused by filter delamination in LL 1st order (LL1). The absolute and relative flux accuracies of IRS are generally better than 10\% and 4\%, respectively, as judged from observations of bright standard stars. However, additional low-level instrumental artifacts may become important for faint sources. We used IRS in standard flux-staring mode, for 2 cycles at 2 nod positions in each of the modules SL1, SL2 (SL 1st and 2nd order), and LL2 (LL 2nd order). We executed 1 cycle at 2 nod positions for LL1, which covers the 20-36.5 $\mu$m range. A typical observation includes 240 s of on-source exposure time in each of SL1, SL2, and LL1, and 480 s in LL2, for a total of 2000 s per target (including overhead). Archived IRS data are used for 14 sources which were observed partly or in full by other investigators (with similar or longer exposure times). Nod or off-slit observations were subtracted to remove foreground emission from the telescope, zodiacal light, and interstellar medium. Spectra were then extracted from the Basic-Calibrated Datasets (BCDs), using the {\it Spitzer} IRS Custom Extraction (SPICE\footnote{ http://ssc.spitzer.caltech.edu/postbcd/spice.html} version 1.1) software and standard tapered extraction windows. The extraction window full-widths are proportional to wavelength in each order to match the diffraction-limited telescope point-spread function (SL2: $7\farcs2$ at 6 $\mu$m, SL1: $14\farcs 4$ at 12 $\mu$m, LL2: $21\farcs7$ at 16 $\mu$m, LL1: $36\farcs6$ at 27 $\mu$m). We rebinned portions of the spectra of 3C 55, 172, 220.1, 244.1, 263.1, 280, and 330 by factors of 4-8 in order to improve the S/N at short wavelengths. Spectral orders were trimmed at the edges and merged to produce final spectra. The SL and LL slits have widths of $3\farcs7$ and $10\farcs6$, respectively. Standard point-source flux calibrations (version 12.0) were applied to correct for slit and aperture losses and convert the spectra from electron s$^{-1}$ to Jy. In most cases, fluxes match to $<15\%$ across order boundaries, consistent with a point source that is well-centered in all of the slits. However, in 5 cases (3C 192, 216, 220.1, 380, and 381) SL2 fluxes are larger by 17-35\% relative to the other orders. Assuming that these mismatches owe to variable slit-loss caused by pointing errors, the orders with low flux are adjusted upward to match the orders with high flux. Order mismatches may alternatively be an indication of extended mid-IR emission. The results for a few sources with nearby neighbors in the slit should be viewed with caution. In the case of 3C 310, a nearby companion galaxy (to the east) may contribute a significant fraction of the flux ($<50\%$) in the LL1 slit. Similarly, a nearby source may potentially contribute to the LL spectra of 3C 438 (which is, however undetected at 15 $\mu$m). The SL2 spectrum of 3C 388 may be weakly affected by flux from a nearby star on the slit ($<20\%$). The northern component of the double nucleus in 3C 401 falls outside of the SL slits, but falls inside the LL slit used to measure the 15 $\mu$m flux. \subsection{Mid-Infrared and Radio Luminosities} We measure the mean $6.5-7.5$ $\mu$m and $13.0-17.0$ $\mu$m flux densities $F_\nu(7$ and 15 $\mu$m, rest) of each target (Tables 1 \& 2). All {\it Spitzer} flux densities in this paper are in observed units at a constant rest-frame wavelength defined by $\lambda_\mathrm{rest}=\lambda_\mathrm{obs}/(1+z)$, where $z$ is the redshift measured from optical emission lines and cataloged in the NASA Extragalactic Database (NED\footnote{ http://nedwww.ipac.caltech.edu}). This avoids any complication from potentially large cosmological K-corrections that could otherwise be introduced by a steep IR continuum slope or redshifted silicate absorption features. We choose to measure the mid-IR flux at 7 and 15 $\mu$m to avoid the 9.7 $\mu$m trough and the deepest part of the 18 $\mu$m silicate absorption trough. We exclude the 14.0-14.5 $\mu$m and 15.3-15.8 $\mu$m wavelength regions from our photometry, to avoid emission from Ne {\sc v} and Ne {\sc iii}. The 7 and 15 $\mu$m bands are within the Spitzer IRS bandpass for redshifts $z<1.28$. However, the archival LL data for two quasars (3C 254 and 275.1) are not yet public. We extrapolate their SL spectra to obtain $F_\nu($15 $\mu$m, rest) using $F_\nu($7 $\mu$m, rest) and the observed (relatively line-free) 5-7 $\mu$m spectral index. Radio luminosities $\nu L_\nu($178 MHz, rest) are estimated from the observed 178 MHz fluxes and K-corrected using the 178-750 MHz radio spectral index \citep{lrl83}. The sources in our sample display a large range of nearly 3 orders of magnitude in mid-IR to radio luminosity: $\nu L_\nu(15$ $\mu\mathrm{m})/\nu L_\nu(178$ MHz$)=0.8-680$ (Fig.1). This quantity is thought to reflect the relative importance of accretion luminosity and jet kinetic power dissipation. However, different size and time scales are probed by the radio (10 kpc-1 Mpc) and mid-IR (0.1-100 pc), and the radio power may be sensitive to differences in environmental conditions. For the purpose of studying quasar and radio galaxy unification, it is natural to divide the sample into the {\it mid-IR luminous} NLRGs which emit as powerfully as quasars or BLRGs, and the {\it mid-IR weak} NLRGs that do not. We adopt an empirical dividing line of $\nu L_\nu(15$ $\mu\mathrm{m})> 8 \times 10^{43}$ erg s$^{-1}$ to separate hidden quasars from mid-IR weak radio galaxies. The cutoff is set at 1/2 the luminosity of the mid-IR weakest BLRG (3C 219) to allow for some degree of anisotropy at 15 $\mu$m. Fourteen NLRGs satisfy our criterion and are thus likely to contain hidden quasar or BLRG nuclei (Table 1). Notably, all of these NLRGs are optically classified as HEGs. The 17 mid-IR weak NLRGs with $\nu L_\nu(15$ $\mu\mathrm{m})<8 \times 10^{43}$ erg s$^{-1}$ (Table 2) have mixed optical classifications, including both HEGs and LEGs. These sources have lower S/N mid-IR spectra, which will be considered in detail in a later paper. Six mid-IR weak NLRGs (including 2 HEGs and 4 LEGs) are undetected by {\it Spitzer} at 15 $\mu$m, and one is also undetected at 7 $\mu$m. \subsection{Hidden Quasar Spectra} \subsubsection{Continuum Emission} We now present {\it Spitzer} spectra of the 14 mid-IR luminous NLRGs that ostensibly contain hidden quasar nuclei (Figs. 2-4). We also plot the spectral energy distributions (SEDs) of the sources with published near-IR photometry (Fig. 5). The collected photometric data were measured in the J, H, K, L$^\prime$, and M wavelength bands from the ground \citep{ll84,llm85,srl99,sww00}. The photometric apertures range in size from $3-11\arcsec$, with preference given to the apertures that most closely match the {\it Spitzer} SL slit width. Where available, the ground-based L$^\prime$ and M-band photometry agrees with {\it Spitzer} spectrophotometry remarkably well. There is no indication of variability over the time span of 20 yr. Four of the low-redshift NLRGs (3C 33, 234, 381, and 452) have broad peaks in their $\nu L_\nu$ spectra (and SEDs) at $1.5-2.5\times 10^{13}$ Hz (12-20 $\mu$m). A maximum and spectral curvature near 20 $\mu$m are also suggestive of broad peaks in the {\it Spitzer} spectra of 3C 55, 244.1, 265, and 330. The large amplitude ($\sim 0.5-1.0$ dex) of the mid-IR bump (Fig. 5) excludes a large contribution of synchrotron emission to the mid-IR continuum of most sources. This is not surprising if the equatorial plane of the dusty torus is roughly perpendicular to the radio jet, such that jet emission is beamed away. The high redshifts of the NLRGs 3C 172, 220.1, 263.1, 268.1, and 280 preclude the identification of a mid-IR bump in the SEDs of these sources. The unusually flat, blue SED of 3C 433 may indicate a quasar viewed at {\it low} inclination (Section 3.2.2). We attribute the mid-IR continuum bump visible in most sources to thermal emission from warm or hot dust. Fitting the mid-IR peak with a single-temperature blackbody model indicates dust with a temperature of $210-225\pm 0.5$ K (Fig. 5). While this temperature characterizes the peak of the mid-IR SED, hotter dust must also be present. At frequencies greater than the peak of the SED ($2.0-7.5\times 10^{13}$ Hz), the continuum emission of most sources can be characterized using a power law with spectral index $\alpha = 1.1-2.1$ (Table 1 \& Fig. 6). This emission likely comes from a continuous distribution of dust temperature. We measure the spectral index between 7 and 15 $\mu$m, avoiding the 9.7 $\mu$m and 18 $\mu$m silicate absorption troughs. The most blue and apparently hottest mid-IR luminous NLRG is 3C 265, while the most red and coolest are 3C 55 and 3C 268.1 (Fig. 6). In comparison, some mid-IR weak sources such as 3C 310 and 3C 388 are quite blue ($\alpha \sim -0.1- +0.7$), indicating a large contribution of starlight from the host galaxy to the 7 $\mu$m continuum. The near-IR continuum shifts into the {\it Spitzer} IRS passband for the highest redshift ($z>0.7$) sources. The spectra of the NLRGs 3C 55 and 3C 265 steepen above $7.5\times10^{13}$ Hz (below 4 $\mu$m). Fitting these spectral turnovers with single-temperature blackbodies, we find emission from hot dust with temperatures of $520\pm 10$ K and $660\pm 10$ K, respectively. Altogether, the mid-IR luminous radio galaxies in our sample show emission from dust with temperatures distributed in the range 210-660 K. Hotter temperature dust (up to the sublimation temperature) may be present but not visible for radio galaxy tori viewed at high inclination \citep{pk92}. Extinction by cold foreground dust in the host galaxy may also affect the spectral index. For Galactic-type dust, $A(7,15,35~\mu\mathrm{m})/A(\mathrm{V})=(0.020,0.015,0.004)$ \citep{m00}. An extinction of $A(\mathrm{V})=100$ would steepen the 7-15 $\mu$m spectral index by $\delta \alpha= 0.6$ (Fig. 6a). The observed range in spectral index for the mid-IR luminous NLRGs is $\delta \alpha=1.0$, corresponding to $A(\mathrm{V}, 7, 15, 35~\mu\mathrm{m})=(167, 3.3, 2.5, 0.6)$ mag. Thus if reddening by a cold foreground dust screen accounted entirely for the range in mid-IR slope, the mid-IR emission could be anisotropic by factors of $f_\mathrm{A}(7,15,35~\mu\mathrm{m})\sim (22,10,1.3)$. However, these are upper limits since variations in the spectral index are also controlled by the physical temperature distribution of the visible dust. The SEDs of several sources (3C 33, 55, 172, 265, and 452) have upturns at short wavelengths, which we attribute to stellar emission from the host galaxy (Fig. 5). The wavelength of the upturn (1-5 $\mu$m) is an indicator of the relative strength of the mid-IR bump seen from our direction vs. host galaxy light, occurring at shorter wavelength for sources with a stronger mid-IR bump. This may have important consequences for understanding the K-z Hubble diagram for 3C radio galaxies, for which it has been argued that AGNs contribute a negligible fraction of the K-band flux \citep[e.g.,][]{sww00}. This may be incorrect for a few of the most luminous mid-IR sources in our sample, including 3C 234 and 3C 280 where there appears to be much emission from hot dust in the K band. Detailed spectral modeling, combined with radio orientation indicators, promises to further characterize the temperature distribution, optical depth, and inclination of the dusty torus that is thought to produce most of the mid-IR emission from hidden quasar nuclei. Such an analysis is, however, outside the scope of the present paper. \subsubsection{Silicate Absorption} The silicate absorption trough at 9.7 $\mu$m is detected in 12/14 of the mid-IR luminous NLRG spectra (Table 3 \& Figs. 2-4). The equivalent width EW$_{9.7}$ and apparent optical depth $\tau_\mathrm{9.7}$ are measured relative to a local continuum fit to either side of the trough, indicated in Figures 2-4. The optical depth is averaged over the trough bottom (rest 9.2-10.2 $\mu$m) to improve the S/N. It should be kept in mind that the apparent $\tau_\mathrm{9.7}$ is just a convenient parameterization of (and lower limit to) the total optical depth since there must also be broad-band silicate absorption of the adjacent continuum. The apparent silicate optical depths are small ($\tau_\mathrm{9.7}=0.02-0.2$), for all but 3C 55 and 3C 433. If attributed to foreground dust screens, this would indicate optical extinction of only $A_\mathrm{V}=0.2-5.1$ mag (Fig. 6b), assuming a Galactic extinction law with $A_\mathrm{V}/\tau_\mathrm{9.7}=12.3-25.6$ mag \citep{rl85,dl84}. The extinction values are clearly underestimated since they imply that the hidden nuclei in 3C 234, 265, 381, and 452, which have $\tau_\mathrm{9.7}\le 0.1$, should be reddened but directly visible at H$\alpha$. The same discrepancy between $\tau_\mathrm{9.7}$ and estimates of extinction at shorter wavelengths is seen for the hidden quasar nucleus in Cygnus A, and attributed to a radial gradient in torus dust temperature \citep{iu00}. The observed range of $\tau_\mathrm{9.7}$ may correspond to a range of equatorial silicate dust column densities in the torus, or alternatively a range of viewing angles. In this regard, more detailed modeling of the torus, including its geometric and temperature structure is clearly called for. Filling-in of the silicate troughs by silicate {\it emission} from the torus or narrow-line region (NLR) may also reduce the apparent silicate optical depths in some sources. This is predicted for an optically thick torus viewed at an intermediate or face-on inclination \citep{pk92}. Recently, strong silicate emission features were detected by {\it Spitzer} in several radio-loud (3C) and radio quiet (PG) quasars \citep{shk05,hss05}. The failure of previous attempts to observe this feature inspired torus models with large dust grain size \citep{ld93} or a spatial distribution of optically thick clumps \citep{nie02}. However, it appears that past non-detections owe to inadequate wavelength coverage to determine the underlying continuum. The NLRGs 3C 55 and 3C 433 have significantly deeper silicate troughs than other NLRGs, with $\tau_\mathrm{9.7}=0.9$ and 0.7, respectively (Fig. 6b). We suggest that their active nuclei and tori are absorbed by an additional (kpc-scale) cold dust screen in the host galaxy. As noted above, the NLRG 3C 433 is unusual in having a flat, blue continuum (similar to some of the quasars in our sample). A blue mid-IR spectrum is not necessarily at odds with deep silicate absorption features. It can be understood if this is a quasar viewed at low inclination to the jet and torus axes, but through an ($A_\mathrm{V}\sim 10$) cold dust screen. This amount of extinction would result in very little reddening at 7-15 $\mu$m ($\delta \alpha=0.05-0.14$), but would be sufficient to create the deep 9.7 $\mu$m trough (Fig. 6b) and would obscure any optical broad lines. The NLRG 3C 433 is also unique in having the only unambiguously detected 18 $\mu$m silicate trough, with equivalent width $EW_{18}=0.42 \pm 0.01$ $\mu$m and apparent optical depth $\tau_\mathrm{18}=0.07 \pm 0.03$ (averaged over 17-19 $\mu$m). The ratio of $\tau_\mathrm{18}$ to $\tau_\mathrm{9.7}$ apparent silicate trough depths is $0.10 \pm 0.04$, consistent with a $0.11$ ratio for Galactic-type silicate dust \citep{dl84}. We do not see the full 18 $\mu$m silicate trough in the spectrum of 3C 55 because of inadequate rest-wavelength coverage. \subsubsection{Forbidden Emission Lines} All of the mid-IR luminous NLRGs with high S/N spectra have forbidden emission lines from highly ionized metals, including [O {\sc iv}] $\lambda 25.89$ $\mu$m, [Ne {\sc ii}] $\lambda 12.81$, [Ne {\sc iii}] $\lambda 15.55$, [Ne {\sc v}] $\lambda 14.3$, [Ne {\sc v}] $\lambda 24.31$, [Ne {\sc vi}] $\lambda 7.65$, [S {\sc iii}] $\lambda 18.71$, [S {\sc iii}] $\lambda 33.48$, and [S {\sc iv}] $\lambda 10.51$ (Figs. 2-4). We measure the line flux and rest equivalent width of each emission line relative to the local continuum level (Table 4). Formal uncertainties are computed from the noise in the continuum to either side of the line. Upper limits are estimated for undetected emission lines, assuming they are unresolved. The large range of ionization states (especially high-ionization Ne {\sc v}, Ne {\sc vi}, and S {\sc iv}) indicates photoionization by a hidden source of far-UV photons \citep{asl99,slv02,acs04}, e.g. a quasar nucleus. Low critical densities in the range $10^3-10^6$ cm$^{-3}$ \citep{asl99} indicate that the forbidden lines arise in the NLR. There could plausibly be contributions from starburst emission to the lower-ionization emission lines such as [Ne {\sc ii}]. For 3C 55 and 3C 433, the [S {\sc iv}] $\lambda 10.51$ line has a relatively large flux even though it sits at the bottom of a deep silicate trough. This line must then arise from a region not heavily obscured by dust, such as the extended NLR. The resolving power of IRS is insufficient to measure the intrinsic emission line widths, which are therefore $<4700$ km s$^{-1}$. In order to assess the ionization level of the emission line regions, we compute several emission line ratios (Fig. 7). In particular, the [O {\sc iv}]/[Ne {\sc ii}] and [Ne {\sc v}]/[Ne {\sc ii}] ratios can be used as diagnostics of the relative contributions of AGN and starburst emission to the mid-IR emission line spectra of galaxies \citep{g98,slv02}. In the sources where O {\sc iv} , Ne {\sc v}, and Ne {\sc ii} emission are all detected (3C 33, 234, 381, and 433), we find [O {\sc iv}] $\lambda 25.89$ $\mu$m / [Ne {\sc ii}] $\lambda 12.81$ $\mu$m $>1.0$ (Fig. 7a) and [Ne {\sc v}] $\lambda 14.3$/ [Ne {\sc ii}] $\lambda 12.81$ $\mu$m $> 0.5$ (Fig. 7c), indicating a $>50\%$ AGN contribution to the emission line spectrum. The [Ne {\sc ii}] $\lambda 12.81$ $\mu$m line is relatively weak or undetected in the $z>0.2$ sources, making it difficult to apply these diagnostics. However, the large EWs of the [Ne {\sc vi}] or [S {\sc iv}] lines in 3C 55, 265, and 330 indicate that the emission line spectra of these sources are also AGN-dominated. \subsubsection{Molecular Emission} Polycyclic aromatic hydrocarbon (PAH) emission features are commonly seen in star-forming regions, starbursts, and starburst-dominated ULIRGs \citep[e.g.,][]{g98,acs04}. The only PAH feature we detect is the weak 11.3 $\mu$m line in the spectrum of 3C 33, with a flux of $2.3\pm 0.3 \times 10^{-14}$ erg s$^{-1}$ cm$^{-2}$ and equivalent width of $0.009\pm 0.003$ $\mu$m (Fig. 2). We do not detect a 6.2 $\mu$m PAH feature in 3C 33 (EW$<0.8$ $\mu$m) or any of the other mid-IR luminous NLRGs, though this spectral region is generally noisier. If present, we could not cleanly resolve the 7.7 $\mu$m PAH feature from the adjacent [Ne {\sc vi}] line, nor the 12.7 $\mu$m PAH feature from the adjacent [Ne {\sc ii}] line. Regardless, we do not see any hint of these PAH features in any source, suggesting EW$<<0.1$ $\mu$m. Therefore in most cases, neither the 7.7 or 11.3 $\mu$m PAH features can have a significant impact on the measurement of the silicate trough. The general lack of PAH features is a strong indication that the primary power source in mid-IR luminous radio galaxies is accretion power, not hot stars. It is likely that PAHs are destroyed in the torus, which is exposed to intense X-ray emission from the AGN \citep{v91}. The weak PAH emission that is present in 3C 33 may arise in star-forming regions shielded from the AGN. We detect the H$_2$ 0-0 S(3) 9.67 $\mu$m and H$_2$ 0-0 S(1) 17.03 $\mu$m pure rotational emission lines of molecular hydrogen (at the $3\sigma$ level) only in the spectrum of the NLRG 3C 433 (Fig. 2). The line fluxes are 0.6 $\pm 0.2 \times 10^{-14}$ and 1.4 $\pm 0.4 \times 10^{-14}$ erg s$^{-1}$ cm$^{-2}$ respectively (EW $=$ 0.008 and 0.013 $\mu$m). The location of the 9.67 $\mu$m line at the bottom the deep 9.7 $\mu$m silicate trough may indicate that the H$_2$ emission region is exterior to the obscuring dust screen. The H$_2$ emission lines can be produced in warm molecular gas heated either by shocks or X-ray photons from the AGN \citep{rkl02}. The ratio of S(3)/S(1) line fluxes is $0.5 \pm 0.2$, which indicates warm H$_2$ with an excitation temperature of $300\pm30$ K. We estimate a warm H$_2$ mass of roughly $2\times 10^8 M_\odot$, assuming a Boltzmann distribution of rotational level populations, an unresolved source, and negligible mid-IR extinction. \section{Discussion} \subsection{Hidden Quasar Nuclei} The high mid-IR luminosities $\nu L_\nu(15$ $\mu\mathrm{m})= 10^{44}-10^{46}$ erg s$^{-1}$ of $45 \pm 12\%$ (14/31) of the FR II NLRGs in our sample are consistent with hidden quasar or BLRG nuclei. In fact, such copious hot dust emission directly requires a hidden source of quasar-like luminosity to power it. Including the 11 quasars and BLRGs seen directly, we find that at least $60 \pm 12\%$ (25/42) of 3CRR FR II sources at $z<1$ with $S_{178}>16.4$ Jy contain powerful AGNs. The percentage of {\it mid-IR luminous} AGNs obscured by dust and therefore classified as NLRGs is $56\pm 15\%$ (14/25), corresponding to a mean torus covering fraction of 0.56 and mean torus opening half-angle of $55\pm 11 \arcdeg$. (If the 8 mid-IR weak HEGs are counted as highly obscured quasars, then the mean torus covering fraction increases to $0.67 \pm 0.14$.) Both of these numbers are consistent with the estimated 60\% mean torus covering fraction required to unify $z=0.5-1.0$ radio galaxies and quasars \citep{b89}. The receding torus model \citep{l91} predicts a larger torus covering fraction for low luminosity sources. If so, we would expect a larger mean mid-IR/radio ratio and a smaller type 1 fraction for low-redshift than for high-redshift sources. The fraction of type 1 mid-IR luminous sources is 3/9 ($0.3\pm 0.2$) at $z<0.5$ vs. 8/16 ($0.5 \pm 0.2$) at $z>0.5$. Clearly, a larger sample is needed to put meaningful constraints on the variation of torus covering fraction with redshift or luminosity. For the galaxies that contain a powerful mid-IR source, we find that there are other indications of a hidden AGN. The high-ionization, mid-IR forbidden lines such as [Ne {\sc v}], even more-so than the strong optical [O {\sc iii}] lines, are telltale signatures of a non-stellar source of FUV photons in the radio galaxies that show them. Because of the lower extinction in the mid-IR relative to the optical, it is likely that we can see these lines closer to the nucleus than optical narrow lines such as [O {\sc iii}] \citep{hss5}. At least three of the mid-IR luminous NLRGs in our sample are known to have highly polarized broad emission lines. The NLRG 3C 234 has quasar-like mid-IR luminosity and a highly polarized broad H$\alpha$ line \citep{ant84}. The NLRG 3C 265 has quasar luminosity, highly polarized UV flux, and a highly polarized broad Mg {\sc ii} line \citep{tco98}. The low-redshift NLRG 3C 33 has mid-IR luminosity comparable to the BLRG 3C 219, and highly polarized broad H$\alpha$ and H$\beta$ \citep{cot99}. We have an ongoing program to obtain optical spectropolarimetry of the rest of the sources in our sample. While the radio galaxies with highly polarized broad lines are known to contain hidden type-1 AGNs, there were previously no reliable measurements of the hidden AGN luminosities. Our mid-IR flux measurements yield rough calorimetric estimates of the hidden AGN luminosities, subject to uncertainties in SED, torus covering fraction, and any mid-IR anisotropy. \subsection{Mid-IR Weak Radio Galaxies} The majority of radio galaxies in our sample (17/31 or $55\pm 13\%$) are relatively weak mid-IR sources. It is possible that some of the AGNs are highly obscured even at 15 $\mu$m because they are viewed through a very large dust column. A mid-IR source obscured by a nearly Compton-thick ($\tau_\mathrm{e}=0.7$) disk with Galactic dust/gas ratio would have an equatorial extinction of $\sim 5$ mag (factor of 100) at 15 $\mu$m. Even in this case, any mid-IR emission above the disk might not be obscured. For example, there may be a contribution from dust in the narrow-line region (NLR), above the hole in the torus \citep[e.g. NGC 1068,][]{gpa05,bnm00}, tending to make the mid-IR emission more isotropic. For galaxies with redshift $z \le 0.22$, {\it Spitzer} can measure the flux at $\lambda = 30$ $\mu$m (rest), which should be more isotropic and less subject to extinction than the 15 $\mu$m flux. Nevertheless, 9/11 mid-IR weak galaxies in this redshift range are also weak at 30 $\mu$m, with $\nu L_\nu(30$ $\mu\mathrm{m}) <8 \times 10^{43}$ erg s$^{-1}$. The two exceptions are 3C 61.1 and 3C 123, with $\nu L_\nu(30$ $\mu\mathrm{m})=1.27 \pm 0.08$ and $1.4\pm 0.2\times 10^{44}$ erg s$^{-1}$, respectively. In comparison 3C 452, the weakest mid-IR luminous NLRG in this redshift range, has $\nu L_\nu(30$ $\mu\mathrm{m})=8.63 \pm 0.09\times 10^{43}$ erg s$^{-1}$. Therefore, reclassifying the NLRGs by their luminosity at 30 $\mu$m would only gain us an additional 2 mid-IR luminous sources. This leads us to believe that most of the mid-IR weak radio galaxies truly lack a powerful accretion disk. Relatively low accretion power suggests, but does not prove, that some FR II jets may be driven by radiatively inefficient accretion flows or black hole spin-energy \citep{bbr84, m99}. As we mentioned above, roughly half (9/17) of the mid-IR weak NLRGs are are classified as LEGs with weak optical [O {\sc iii}] emission. The [O {\sc iii}] emission in these sources may be weak because there is no strong source of UV photons to power the NLR. Qualitatively similar conclusions have been drawn by other investigators \citep[e.g.,][]{hl79, ccc00,grw04}. Alternatively, the NLR may be partly or completely obscured in LEGs \citep{hss5}. It will be important to make a more quantitative assessment of the optical and IR emission line strengths, to evaluate the extinction and determine what UV luminosity is necessary to power the NLR in mid-IR weak NLRGs. \subsection{Radio Properties and Unification} One of the major motivations for the radio galaxy and quasar unification theory is the deficit of lobe-dominant vs. core-dominant FR II quasars \citep{b89}. Relativistic beaming models predict that there should be relatively more sources where the radio jet is beamed away and the high frequency radio core is weak. Core fluxes at 5 GHz are tabulated for the 3CRR catalog by \cite{lrl03}\footnote{The online 3CRR catalog is available at http://www.3crr.dyndns.org.}. We identify the mid-IR luminous NLRGs in our sample with the missing lobe-dominant quasar population (Fig. 8). Their median core to lobe ratio is $R_\mathrm{c}=\nu L_\nu($core, 5 GHz$)/ \nu L_\nu(178$ MHz$)=5$, while the median $R_\mathrm{c}=180$ for the quasars. Conversely, all of the mid-IR luminous sources with $R_\mathrm{c}>100$ are classified as quasars. We will perform a more detailed statistical analysis when the Spitzer observations of our full sample are complete. For most mid-IR weak NLRGs, we reject the possibility that the AGN and radio core are viewed in an 'off' state while the radio lobes are still active. A 5 GHz core is detected in 13/17 of the mid-IR weak radio galaxies (Fig. 8) and 13/14 of the mid-IR luminous galaxies. Furthermore, mid-IR weak and mid-IR luminous NLRGs have comparable core to total luminosity ratios of $R_\mathrm{c} = 1-100$. Deeper 5 GHz observations of the 5 non-detected cores (in 3C 28, 153, 172, 315, and 319) will be necessary to determine whether or not they are in an off state (e.g., $R_\mathrm{c}<0.1$). We find that FR II radio morphology is not a reliable predictor of nuclear mid-IR luminosity for radio galaxies. {\it Contrary to the simple unification paradigm, not all narrow-line FR II galaxies host nuclei as powerful as quasars with matched radio lobe luminosity}. Unification theories must be modified to account for an additional population of mid-IR weak radio galaxies. Both intrinsic jet power and interaction with the interstellar and intergalactic medium are likely to be important for determining radio morphology. The break luminosity between FR I and FR II radio sources is found to increase with host galaxy optical luminosity \citep{ol94,b96}. The existence of radio sources with hybrid FR I/II morphology also points to the importance of environmental effects in determining radio morphology \citep{gkw00,gmk05}. Furthermore, most FR I radio jets are one-sided, relativistic, and narrowly collimated on sub-parsec scales, just like FR IIs, and decollimate only on kpc scales \citep[e.g. M87,][]{jbl99}. Contrary to a common misconception, not all FR I sources have radiatively inefficient nuclei. For example, Centaurus A is persuasively argued to have a powerful hidden AGN \citep{wa04}, and the BLRG 3C 120 is a well-known FR I source. Deep VLA observations of the optically luminous, 'radio-quiet' quasar E1821+643 demonstrate that it has an FR I radio morphology \citep{br01}. These and other cases \citep{ant01} demonstrate that many powerful AGNs are FR Is. The observed variation in radio morphology and a wide range in AGN radio-loudness \citep{kss89} do not necessarily require a weak coupling between jet power and accretion power, but may demonstrate that multiple factors are at work. At least five parameters may be necessary to theoretically unify all AGN types: black hole mass, black hole spin, accretion rate, radiative efficiency, and viewing angle. There is still much work ahead before we completely understand how basic physical parameters regulate the activity of supermassive black hole systems. Models that tie jet production to accretion onto a spinning black hole are particularly promising \citep[e.g.,][]{m99,hk06}. Much progress has been made in understanding the aspect-dependent appearance of AGN disks and jets, as a consequence of relativistic beaming and obscuration \citep{up95}. Our {\it Spitzer} observations confirm that many FR II radio galaxies would appear as powerful quasars if viewed from an unobscured direction (e.g. along the radio axis). However, just as many FR II radio galaxies would not. Our {\it Spitzer} observations also demonstrate that powerful radio jets may be produced even by mid-IR weak AGN. A powerful, luminous accretion disk is not always necessary to produce a highly collimated, relativistic jet. \section{Conclusions} (1.) We report on a large {\it Spitzer} spectroscopic survey of 3CRR FR II radio sources with $S_{178}>16.4$ Jy at $z<1$. We find strong mid-IR emission from $45 \pm 12\%$ (14/31) of NLRGs, which have luminosities comparable to matched BLRGs and quasars. Other indicators including high-ionization mid-IR lines and highly polarized broad emission lines confirm that some of these sources contain hidden quasar or BLRG nuclei. This demonstrates the power of {\it Spitzer} IRS for unveiling hidden quasars and estimating their luminosities. (2.) We present {\it Spitzer} spectra of the 14 mid-IR luminous radio galaxies. In most cases, the mid-IR continuum bump from 3-30 $\mu$m can be produced by a distribution of hot dust with temperatures in the range 210-$660$ K. These high temperatures are most likely maintained by hidden AGNs. The silicate absorption trough at 9.7 $\mu$m has an apparent optical depth of $\tau=0-0.2$ in most cases, consistent with dust temperatures decreasing outward from the center of a dusty torus. Two sources, 3C 55 and 433, have deeper silicate troughs which may be produced by additional cool dust in the host galaxy. (3.) However, not all FR II radio galaxies emit strongly in the mid-IR. Contrary to single-population unification schemes, the majority of narrow-line radio galaxies in our sample (17/31 or $55 \pm 13\%$) have weak or undetected mid-IR emission compared to matched quasars and BLRGs, with $\nu L_\nu(15$ $\mu\mathrm{m}) < 8 \times 10^{43}$ erg s$^{-1}$. For a few sources, this may possibly be the result of anisotropic torus emission viewed through a large column density of dust. However, it is likely that most of the weakest sources do not contain a powerful accretion disk. These may be truly nonthermal, jet-dominated AGNs, where the jet is powered by a radiatively inefficient accretion flow or black hole spin-energy rather than energy extracted from an accretion disk. \acknowledgements This work is based on observations made with the {\it Spitzer} Space Telescope, which is operated by the Jet Propulsion Laboratory, California Institute of Technology under NASA contract 1407. We have also made use of the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with NASA. Support for this research was provided by NASA through an award issued by JPL/Caltech. We thank Dave Meier, Lee Armus, Bill Reach, and the anonymous referee for their helpful input and comments on the manuscript. \begin{deluxetable}{cccccccc} \tablecaption{Mid-IR Luminous Sources} \tablewidth{0pt} \tablehead{ \colhead{3C} & \colhead{type\tablenotemark{a}} & \colhead{z} & \colhead{$F_7$\tablenotemark{b}} & \colhead{$F_{15}$\tablenotemark{c}} & \colhead{log $\nu L_{15}$ \tablenotemark{d }} & \colhead{$\alpha$ \tablenotemark{e}} &\colhead{S178\tablenotemark{f}} } \startdata 175 & QSR & 0.7700 & 6.96 $\pm$ 0.07 & 21.6 $\pm$ 0.7 & 45.83 &1.49 $\pm$ 0.04 & 19.2 \\ 196 & QSR & 0.871 & 8.0 $\pm$ 0.1 & 22.9 $\pm$ 0.6 & 45.96 &1.38 $\pm$ 0.04 & 74.3 \\ 216 & QSR & 0.6703 & 9.8 $\pm$ 0.1 & 28.7 $\pm$ 0.6 & 45.83 &1.41 $\pm$ 0.03 & 22.0 \\ 219 & BLRG & 0.1744 & 3.6 $\pm$ 0.1 & 11.2 $\pm$ 0.4 & 44.21 &1.50 $\pm$ 0.06 & 44.9 \\ 254 & QSR & 0.7361 & 6.02 $\pm$ 0.08 & \nodata & 45.56:&\nodata & 21.7 \\ 263 & QSR & 0.646 & 13.6 $\pm$ 0.1 & 29.8 $\pm$ 0.8 & 45.81 &1.03 $\pm$ 0.04 & 16.6 \\ 275.1 & QSR & 0.5551 & 3.04 $\pm$ 0.07 & \nodata & 44.76:&\nodata & 19.9 \\ 325 & QSR & 0.8600 & 1.17 $\pm$ 0.07 & 4.1 $\pm$ 0.2 & 45.20 &1.6 $\pm$ 0.1 & 17.0 \\ 380 & QSR & 0.6920 & 15.1 $\pm$ 0.1 & 40.4 $\pm$ 1.2 & 46.00 &1.29 $\pm$ 0.04 & 64.7 \\ 382 & BLRG & 0.0579 & 86.1 $\pm$ 0.3 &114. $\pm$ 2. & 44.24 &0.37 $\pm$ 0.02 & 21.7 \\ 390.3 & BLRG & 0.0561 & 56.8 $\pm$ 0.5 &164. $\pm$ 4. & 44.37 &1.39 $\pm$ 0.03 & 51.8 \\ \hline 33 & HEG & 0.0597 & 19.8 $\pm$ 0.3 & 75. $\pm$ 2. & 44.08 &1.75 $\pm$ 0.04 & 59.3 \\ 55 & HEG & 0.7348 & 4.7 $\pm$ 0.1 & 23.3 $\pm$ 0.7 & 45.82 &2.11 $\pm$ 0.06 & 23.4 \\ 172 & HEG & 0.5191 & 0.40 $\pm$ 0.04 & 1.5 $\pm$ 0.2 & 44.31 &1.7 $\pm$ 0.2 & 16.5 \\ 220.1 & HEG & 0.610 & 0.77 $\pm$ 0.05 & 2.4 $\pm$ 0.1 & 44.67 &1.5 $\pm$ 0.1 & 17.2 \\ 234 & HEG\tablenotemark{g} & 0.1848 & 86. $\pm$ 2. & 239. $\pm$ 3. & 45.59 &1.35 $\pm$ 0.03 & 34.2 \\ 244.1 & HEG & 0.4280 & 3.50 $\pm$ 0.09 & 14.4 $\pm$ 0.3 & 45.13 &1.86 $\pm$ 0.04 & 22.1 \\ 263.1 & HEG & 0.8240 & 0.63 $\pm$ 0.07 & 2.7 $\pm$ 0.1 & 44.98 &1.9 $\pm$ 0.2 & 19.8 \\ 265 & HEG & 0.8110 & 9.0 $\pm$ 0.1 & 21.1 $\pm$ 0.5 & 45.86 &1.12 $\pm$ 0.04 & 21.2 \\ 268.1 & HEG & 0.970 & $<0.5$ & 3.0 $\pm$ 0.2 & 45.17 &2.1 $\pm$ 0.2\tablenotemark{h} & 23.3 \\ 280 & HEG & 0.996 & 4.58 $\pm$ 0.09 & 13.2 $\pm$ 0.4 & 45.83 &1.39 $\pm$ 0.05 & 25.8 \\ 330 & HEG & 0.550 & 1.72 $\pm$ 0.06 & 6.4 $\pm$ 0.2 & 45.00 &1.72 $\pm$ 0.06 & 30.3 \\ 381 & HEG & 0.1605 & 11.1 $\pm$ 0.2 & 36.4 $\pm$ 0.8 & 44.65 &1.56 $\pm$ 0.05 & 18.1 \\ 433 & HEG & 0.1016 & 40.0 $\pm$ 0.4 & 98. $\pm$ 1. & 44.67 &1.18 $\pm$ 0.02 & 61.3 \\ 452 & HEG & 0.0811 & 11.3 $\pm$ 0.2 & 45. $\pm$ 1. & 44.13 &1.80 $\pm$ 0.04 & 59.3 \\ \enddata \tablenotetext{a}{Optical spectral type \citep{lrh99,jr97}.} \tablenotetext{b,c}{Flux densities (mJy) and $3\sigma$ upper limits at 7 $\mu$m (rest) and 15 $\mu$m (rest) from {\it Spitzer} IRS.} \tablenotetext{d}{~Logarithm of luminosity (erg s$^{-1}$) at 15 $\mu$m (rest). Values for 3C 254 and 3C 275.1 are extrapolated from 7 $\mu$m because the LL data are unavailable.} \tablenotetext{e}{Spectral power law index for $F_\nu\sim \nu^{-\alpha}$ from 7-15 $\mu$m (rest).} \tablenotetext{f}{Radio flux density (Jy) at 178 MHz (observed) \citep{lrl83}, multiplied by a factor of 1.09 to convert to the \cite{b77} standard flux scale.} \tablenotetext{g}{The broad H$\alpha$ line visible in total flux is entirely scattered \citep{ant84}.} \tablenotetext{h}{Spectral power law index for 3C 268.1 measured from 8-15 $\mu$m (rest).} \end{deluxetable} \begin{deluxetable}{cccccccc} \tablecaption{Mid-IR Weak Sources} \tablewidth{0pt} \tablehead{ \colhead{3C} & \colhead{type\tablenotemark{a}} & \colhead{z} & \colhead{$F_7$\tablenotemark{b}} & \colhead{$F_{15}$\tablenotemark{c}} & \colhead{log $\nu L_{15}$ \tablenotemark{d }} & \colhead{$\alpha$ \tablenotemark{e}} &\colhead{S178\tablenotemark{f}} } \startdata 61.1 & HEG & 0.1878 & 0.64 $\pm$ 0.06 & 3.0 $\pm$ 0.2 & 43.70 &2.0 $\pm$ 0.2 & 34.0 \\ 192 & HEG & 0.0597 & 1.28 $\pm$ 0.07 & 3.2 $\pm$ 0.2 & 42.71 &1.2 $\pm$ 0.1 & 23.0 \\ 274.1 & HEG & 0.4220 & 0.20 $\pm$ 0.05 & $<$0.9 &$<$43.91 &\nodata & 18.0 \\ 300 & HEG & 0.270 & 0.26 $\pm$ 0.04 & 0.7 $\pm$ 0.2 & 43.40 &1.3 $\pm$ 0.4 & 19.5 \\ 315 & HEG & 0.1083 & 0.97 $\pm$ 0.08 & 1.9 $\pm$ 0.2 & 43.01 &0.9 $\pm$ 0.2 & 19.4 \\ 388 & HEG & 0.0917 & 0.98 $\pm$ 0.06 & 0.84 $\pm$ 0.09& 42.66 &0.4 $\pm$ 0.2 & 26.8 \\ 436 & HEG & 0.2145 & 0.65 $\pm$ 0.05 & 1.5 $\pm$ 0.2 & 43.52 &1.1 $\pm$ 0.2 & 19.4 \\ 438 & HEG & 0.290 & 0.16 $\pm$ 0.04 &$<$0.45 &$<$43.27 &\nodata & 48.7 \\ \hline 28 & LEG & 0.1953 & 0.45 $\pm$ 0.06 & $<$0.30 &$<$42.74 &\nodata & 17.8 \\ 123 & LEG & 0.2177 & 1.07 $\pm$ 0.08 & 2.8 $\pm$ 0.4 & 43.81 &0.7 $\pm$ 0.2 &206.0 \\ 153 & LEG & 0.2769 & 0.29 $\pm$ 0.05 & 1.0 $\pm$ 0.2 & 43.59 &1.7 $\pm$ 0.3 & 16.7 \\ 173.1 & LEG & 0.2921 & 0.38 $\pm$ 0.04 & 0.6 $\pm$ 0.1 & 43.40 &0.6 $\pm$ 0.3 & 16.8 \\ 288 & LEG & 0.2460 & 0.40 $\pm$ 0.05 &$<$0.60 &$<$43.25 &\nodata & 20.6 \\ 310 & LEG & 0.0538 & 0.81 $\pm$ 0.07 & 0.73$\pm$ 0.1 & 41.98 &-0.1 $\pm$ 0.2 & 60.1 \\ 319 & LEG & 0.1920 & $<$0.15 &$<$0.27 &$<$42.68 &\nodata & 16.7 \\ 326 & LEG & 0.0895 & 0.67 $\pm$ 0.09 &$<$0.39 &$<$42.16 &\nodata & 22.2 \\ 401 & LEG & 0.2011 & 0.25 $\pm$ 0.05 & 0.8$\pm$ 0.2 & 43.17 &1.5 $\pm$ 0.4 & 22.8 \\ \enddata \tablenotetext{a-f}{See Table 1.} \end{deluxetable} \begin{deluxetable}{ccc} \tablecaption{Silicate Trough} \tablewidth{0pt} \tablehead{ \colhead{3C} & \colhead{EW$_{9.7}$ \tablenotemark{a}} & \colhead{$\tau_{9.7}$ \tablenotemark{b}} } \startdata 33 & -0.222 $\pm$ 0.007 & 0.14 $\pm$ 0.02 \\ 55 & -1.51 $\pm$ 0.03 & 0.9 $\pm$ 0.1 \\ 172 & $>-0.4$ & \nodata \\ 220.1 & $>-0.3$ & \nodata \\ 234 & -0.028 $\pm$ 0.006 & 0.019 $\pm$ 0.008 \\ 244.1 & -0.23 $\pm$ 0.03 & 0.18 $\pm$ 0.08 \\ 263.1 & -1.3 $\pm$ 0.2 & \nodata \\ 265 & -0.11 $\pm$ 0.03 & 0.04 $\pm$ 0.06 \\ 268.1 & -0.4 $\pm$ 0.1 & 0.2 $\pm$ 0.5: \\ 280 & -0.23 $\pm$ 0.02 & 0.16 $\pm$ 0.09 \\ 330 & -0.41 $\pm$ 0.06 & 0.1 $\pm$ 0.1 \\ 381 & -0.19 $\pm$ 0.02 & 0.07 $\pm$ 0.02 \\ 433 & -1.30 $\pm$ 0.01 & 0.71 $\pm$ 0.07 \\ 452 & -0.196 $\pm$ 0.009 & 0.10 $\pm$ 0.02 \\ \enddata \tablenotetext{a}{The 9.7 $\mu$m silicate trough (rest) equivalent width in $\mu$m.} \tablenotetext{b}{Apparent 9.7 $\mu$m silicate optical depth, averaged over 9.2-10.2 $\mu$m (rest).} \end{deluxetable} \begin{deluxetable}{ccccccccc} \tablecaption{Emission Lines\tablenotemark{a}} \tablewidth{0pt} \tablehead{ \colhead{3C} & \colhead{[Ne {\sc vi}]} & \colhead{[S {\sc iv}]} & \colhead{[Ne {\sc ii}]} & \colhead{[Ne {\sc v}]} & \colhead{[Ne {\sc iii}]} & \colhead{[S {\sc iii}]} & \colhead{[Ne {\sc v}]} & \colhead{[O {\sc iv}]} \\ \colhead{ } & \colhead{$\lambda$ 7.65} & \colhead{10.51} & \colhead{12.81} & \colhead{14.3} & \colhead{15.55} & \colhead{18.71} & \colhead{24.31} & \colhead{25.89} } \startdata 33 & 2.8(0.2)& 1.2(0.1)& 3.9(0.2)& 2.0(0.3)& 5.3(0.2)& 2.5(0.4)& 1.6(0.2)& 8.1(0.2)\\ & 0.022 & 0.012 & 0.037 & 0.019 & 0.051 & 0.028 & 0.029 & 0.159 \\ 55 &1.82(.06)& 2.2(0.3)& $<0.5$ & 1.1(0.2)& 2.0(0.3)& $<0.4$ & \nodata & \nodata \\ & 0.062 & 0.17 & $<0.01$ & 0.036 & 0.068 & $<0.01$ & \nodata & \nodata \\ 172 & $<0.1$ & $<0.2$ & $<0.3$ & $<0.6$ & $<0.6$ & $<0.8$ & \nodata & \nodata \\ & $<0.08$ & $<0.1$ & $<0.08$ & $<0.2$ & $<0.2$ & $<0.3$ & \nodata & \nodata \\ 220.1 & $<0.2$ & 0.5(0.2)& $<0.2$ & $<0.5$ &0.35(.09)& $<0.2$ & \nodata & \nodata \\ & $<0.06$ & 0.21 & $<0.05$ & $<0.1$ & 0.11 & $<0.05$ & \nodata & \nodata \\ 234 & 1.7(0.1)& 3.2(0.3)& 0.8(0.2)& 3.3(0.7)& 8.2(0.7)& $<3.$ & 3.9(0.7)& 7.5(1.0)\\ & 0.0034 & 0.0077 & 0.0022 & 0.010 & 0.027 & $<0.01$ & 0.026 & 0.056 \\ 244.1 & 0.7(0.2)& 0.9(0.2)& 1.4(0.2)& 0.6(0.4)& 0.3(0.2)&0.68(.06)& \nodata & \nodata \\ & 0.033 & 0.041 & 0.067 & 0.033 & 0.02 & 0.043 & \nodata & \nodata \\ 263.1 & $<0.6$ & $<0.8$ & 0.6(0.2)& $<0.6$ & 0.6(0.2)& \nodata & \nodata & \nodata \\ & $<0.1$ & $<0.3$ & 0.19 & $<0.2$ & 0.23 & \nodata & \nodata & \nodata \\ 265 & 0.6(0.2)& 1.6(0.4)& $<0.8$ & $<0.6$ & $<0.5$ & \nodata & \nodata & \nodata \\ & 0.012 & 0.040 & $<0.02$ & $<0.02$ & $<0.02$ & \nodata & \nodata & \nodata \\ 268.1 &0.49(.08)& $<0.4$ & $<0.4$ & $<0.8$ & $<0.2$ & \nodata & \nodata & \nodata \\ & 0.15 & $<0.1$ & $<0.1$ & $<0.4$ & $<0.04$ & \nodata & \nodata & \nodata \\ 280 & $<0.3$ & $<0.5$ & 0.3(0.1)& $<0.4$ & 0.9(0.2)& \nodata & \nodata & \nodata \\ & $<0.01$ & $<0.02$ & 0.014 & $<0.02$ & 0.050 & \nodata & \nodata & \nodata \\ 330 & 0.4(0.2)&0.64(.08)& $<0.5$ & 0.4(0.1)& 0.7(0.2)&0.19(.08)& \nodata & \nodata \\ & 0.03 & 0.076 & $<0.05$ & 0.04 & 0.09 & 0.02 & \nodata & \nodata \\ 381 & 1.3(0.1)&0.62(.07)& 0.6(0.2)& 0.7(0.3)& 1.2(.3) &1.49(.08)&0.56(.08)& 3.9(0.2)\\ & 0.019 & 0.012 & 0.011 & 0.014 & 0.025 & 0.040 & 0.026 & 0.196 \\ 433 & 3.1(0.3)& 2.2(0.2)& 1.9(0.3)& 2.7(0.3)& 5.2(0.3)& 1.2(0.8)& 2.3(0.5)& 7.9(0.4)\\ & 0.014 & 0.023 & 0.012 & 0.020 & 0.042 & 0.013 & 0.028 & 0.105 \\ 452 & $<0.5$ & 0.7(0.1)&2.11(.07)& $<0.3$ & 1.9(0.2)& 1.4(0.4)& $<0.3$ & 1.3(0.2)\\ & $<.008$ & 0.012 & 0.034 & $<.005$ & 0.032 & 0.026 & $<0.01$ & 0.057 \\ \enddata \tablenotetext{a}{Notes. For each source, emission line fluxes ($10^{-14}$ erg s$^{-1}$ cm$^{-2}$) or 2$\sigma$ upper limits are on the first line and rest equivalent widths ($\mu$m) are on the second line. Emission line rest wavelengths ($\mu$m) are at the top of each column.} \end{deluxetable}
Title: Dispersion relation for electromagnetic wave propagation in a strongly magnetized plasma
Abstract: A dispersion relation for electromagnetic wave propagation in a strongly magnetized cold plasma is deduced, taking photon-photon scattering into account. It is shown that the combined plasma and quantum electrodynamic effect is important for understanding the mode-structures in magnetar and pulsar atmospheres. The implications of our results are discussed.
https://export.arxiv.org/pdf/astro-ph/0601311
\title[Dispersion relation in strongly a magnetized plasma]{Dispersion relation for electromagnetic wave propagation in a strongly magnetized plasma} \author{G. Brodin M. Marklund, L. Stenflo and P.K. Shukla} \address{Department of Physics, Ume{\aa} University, SE--901 87 Ume{\aa}, Sweden} \pacs{52.25.Xz (Magnetized plasmas), 52.35.Mw (Nonlinear phenomena), 52.27.Ep (Electron-positron plasmas)} \section{Introduction} The quantum electrodynamical (QED) phenomenon of elastic photon--photon scattering, due to the interaction of photons with virtual electron--positron pairs, has recently received increased attention \cite{Ding1992,Moulin1999,BMS2001,Eriksson2004,Shen-etal,Shen-Yu,Bulanov-etal,JPP,% Marklund-Brodin-Stenflo,MSSBS2005,RMP}. Several papers are motivated by the desire to detect photon--photon scattering in laboratories \cite {Ding1992,Moulin1999,BMS2001,Eriksson2004}, whereas others \cite {Shen-etal,Shen-Yu,Bulanov-etal} concern phenomena that might be relevant when the laser power is further increased to produce electric fields strengths close to the Schwinger field $\sim 10^{18}\, \mathrm{V\,m^{-1}}$ \cite{Bulanov-etal}% . Up to now, however, observable effects of photon--photon scattering are likely to occur only for astrophysical systems \cite {Baring-Harding,Shaviv-etal,Marklund-Brodin-Stenflo,magnetar,Bialynicki-Birula1970,% Adler,MSSBS2005}, where the large magnetic field strengths in pulsar and magnetar environments \cite{magnetar,Duncan-Thompson,Palmer-etal} open up for QED processes to play an important role, leading to phenomena such as frequency down-shifting \cite{Adler,Bialynicki-Birula1970} and lensing \cite{Shaviv-etal}. \ The frequency down-shifting \ is a result of so called photon splitting \cite {Adler,Bialynicki-Birula1970}, which is one of the consequences of elastic photon-photon scattering, and the process may even be responsible for the radio silence of magnetars \cite{Baring-Harding}. Another QED-process of interest in pulsar and magnetar environments is pair-production \cite{Beskin-book} due to the strong field interactions, which lead to the presence of an electron-positron pair plasma in the pulsar and magnetar atmospheres. However, with a few exceptions (e.g.\ \cite{MSSBS2005,JPP}), we note that most papers considering photon--photon scattering have omitted plasma effects when considering electromagnetic wave propagation under these conditions. In the present paper we will consider electromagnetic wave propagation at an arbitrary angle to a strong external magnetic field $\mathbf{B}_{0}$, and include the QED-effects associated with that field, as well as the influence of an electron-positron pair plasma. The former effect is described within the framework of the Heisenberg--Euler Lagrangian, which constitutes an effective theory of photon--photon scattering \cite {Heisenberg-Euler,Schwinger}, and the latter contribution follows from elementary plasma theory. A comparatively general dispersion relation will be derived. It reduces to previous results in a number of limiting cases \cite{Chen-book,Bialynicki-Birula1970,MSSBS2005}. In order to determine the contribution from the pair-plasma on the propagation properties in pulsar and magnetar atmospheres, we adopt the Goldreich-Julian expression for the plasma density \cite{Beskin-book}, and evaluate the dispersion relation for field strengths in the pulsar and magnetar range, $B_{0}\sim 10^{8}-10^{10}% \,\mathrm{T}$. In the radio-wave regime it then turns out that for one of the EM-wave polarizations, the plasma effects are typically negligible as compared to the QED-effects, whereas for the other polarization, the opposite is true in most cases. Noting that important processes in pulsar and magnetar environments, e.g.\ photon splitting, typically involve both EM-wave polarizations, we will conclude that QED and plasma effects should be simultaneously included when studying radio wave propagation in such environments. \section{Derivations} If vacuum fluctuations are taken into account, such as under highly energetic conditions (e.g.\ pulsar plasmas and the next generation of laser-plasma systems), Maxwell's equation will be altered by the quantum vacuum self-interaction through the polarization \begin{equation} \mathbf{P}=2\kappa \epsilon _{0}^{2}\left[ 2(E^{2}-c^{2}B^{2})\mathbf{E}% +7c^{2}(\mathbf{E}\cdot \mathbf{B})\mathbf{B}\right] \end{equation} and magnetization \begin{equation} \mathbf{M}=2\kappa \epsilon _{0}^{2}c^{2}\left[ -2(E^{2}-c^{2}B^{2})\mathbf{B% }+7(\mathbf{E}\cdot \mathbf{B})\mathbf{E}\right] \end{equation} respectively, see e.g.\ \cite{BMS2001}. Here $\kappa =(\alpha /90\pi )(1/\epsilon _{0}E_{\mathrm{crit}}^{2})$ gives the strength of the quantum vacuum nonlinearity, $\alpha \approx 1/137$ is the fine-structure constant, $% E_{\mathrm{crit}}=m^{2}c^{3}/e\hbar \sim 10^{18}\,\mathrm{V\,m^{-1}}$ is the Schwinger critical field, $m$ is the electron rest mass, $c$ is the speed of light in vacuum, $e$ is the magnitude of the electron charge, and $\hbar $ is Planck's constant divided by $2\pi $. These corrections to Maxwell's vacuum equations are valid as long as $|\mathbf{E}|\ll E_{\mathrm{crit}}$ and $% \omega \ll \omega _{e}=mc^{2}/\hbar $, where $\omega _{e}\approx 8\times 10^{20}\,\mathrm{rad\,s^{-1}}$ is the Compton frequency. Next we Fourier decompose the electromagnetic perturbations, which have frequencies $\omega $ and wavevectors $\mathbf{k}$. Maxwell's equations together with the plasma equations of motion then yield \begin{equation} \Delta ^{ab}\delta E_{b}=0. \end{equation} using index notation. Here the matrix $\Delta ^{ab}=n^{a}n^{b}-n^{2}\delta ^{ab}+\epsilon ^{ab}$, where $n^{a}=k^{a}c/\omega $, $n=kc/\omega $ is the plasma refractive index, where $k=|\mathbf{k}|$, $\epsilon ^{ab}=\epsilon _{% \mathrm{classical}}^{ab}+\epsilon _{\mathrm{QED}}^{ab}$ is the dielectric tensor, \begin{equation} \epsilon _{\mathrm{classical}}^{ab}=\delta ^{ab}+i\omega \sum_{s}\left( \frac{% \omega _{\mathrm{p}s}}{\omega }\right) ^{2}\sigma _{s}^{ab}, \label{eq:dielectric-classical} \end{equation} \begin{equation} \epsilon _{\mathrm{QED}}^{ab}=-4\xi \left[ \delta ^{ab}+n^{a}n^{b}-n^{2}\delta ^{ab}-\frac{7}{2}b^{a}b^{b}-2(\eta ^{aij}n_{i}b_{j})(\eta ^{bkl}n_{k}b_{l})% \right] , \end{equation} $s$ denotes the plasma particle species, $\omega _{\mathrm{p}s}=(q_{s}^{2}n_{s}/\epsilon _{0}m_{s})^{1/2}$ is the plasma frequency for species $s$, $\ \xi =\kappa \epsilon _{0}c^{2}B_{0}^{2}=(\alpha /90\pi )(cB_{0}/E_{\mathrm{crit}})^{2}$ is the dimensionless QED parameter, $% b^{a}=B_{0}^{a}/B_{0}$, and \begin{equation} (\sigma _{s}^{ab})^{-1}=-i\omega \delta ^{ab}+\omega _{\mathrm{c}s}\eta ^{abj}b_{j}, \label{eq:sigmainv} \end{equation} with the cyclotron frequency $\omega _{\mathrm{c}s}=q_{s}B_{0}/m_{s}$ for species $s,$ $\delta ^{ab}$ is the Kronecker dela and $\eta _{abc}$ is the totally anti-symmetric unit tensor. From the definition (\ref{eq:sigmainv}) we obtain \begin{equation} \sigma _{s}^{ab}=\frac{i\omega }{\omega ^{2}-\omega _{\mathrm{c}s}^{2}}(\delta ^{ab}-b^{a}b^{b})-\frac{\omega _{\mathrm{c}s}}{\omega ^{2}-\omega _{\mathrm{c}s}^{2}}\eta ^{abj}b_{j}+\frac{i}{\omega }b^{a}b^{b}, \label{sigma-QED} \end{equation} and the dielectric tensor (\ref{eq:dielectric-classical}) is thus \begin{equation} \fl \epsilon _{\mathrm{classical}}^{ab}=\delta ^{ab}-\sum_{s}\left[ \frac{\omega _{\mathrm{p}s}^{2}}{\omega ^{2}-\omega _{\mathrm{c}s}^{2}}(\delta ^{ab}-b^{a}b^{b})+\frac{% i\omega _{\mathrm{p}s}^{2}\omega _{\mathrm{c}s}}{\omega (\omega ^{2}-\omega _{\mathrm{c}s}^{2})}\eta ^{abj}b_{j}+\left( \frac{\omega _{\mathrm{p}s}}{\omega }\right) ^{2}b^{a}b^{b}\right] , \label{Eps-classic} \end{equation} We note that the full dielectric tensor depends on the wavevector through the QED contribution $\epsilon _{\mathrm{QED}}^{ab}$. Freely propagating waves are characterized by the vanishing of the dispersion relation $D(\omega ,% \mathbf{k})=\mathrm{det}(\Delta ^{ab})$. Writing the QED-tensor $\epsilon _{% \mathrm{QED}}^{ab}$in matrix form, letting the $\mathbf{k}$-vector lie in the $xz$-plane, we then obtain \begin{equation} \epsilon^{ab}_{\mathrm{QED}}=-4\xi \left( \begin{array}{ccc} 1-n_{\Vert }^{2} & 0 & n_{\bot }n_{\Vert } \\ 0 & 1-n^{2}-2n_{\bot }^{2} & 0 \\ n_{\bot }n_{\Vert } & 0 & -\frac{5}{2}-n_{\bot }^{2} \end{array} \right) \label{QED-matrix} \end{equation} where $n_{\Vert }=k_{\Vert }c/\omega $, $n_{\bot }=k_{\bot }c/\omega $ and the $\mathbf{k}$-vector is written as $\mathbf{k}=k_{\bot }\widehat{\mathbf{x% }}+k_{\Vert }\widehat{\mathbf{z}}$. From (\ref{Eps-classic}) the classical contributions to $\Delta ^{ab}$ is \begin{equation} \!\!\!\!\!\!\! \left( \begin{array}{ccc} 1- \displaystyle{\sum\limits_{s}\frac{\omega _{\mathrm{p}s}^{2}}{\omega ^{2}-\omega _{\mathrm{c}s}^{2}} }% -n_{\Vert }^{2} & \displaystyle{i\sum\limits_{s}\frac{\omega _{\mathrm{p}s}^{2}\omega _{\mathrm{c}s}}{\omega (\omega ^{2}-\omega _{\mathrm{c}s}^{2})}} & n_{\bot }n_{\Vert } \\ \displaystyle{-i\sum\limits_{s}\frac{\omega _{\mathrm{p}s}^{2}\omega _{\mathrm{c}s}}{\omega (\omega ^{2}-\omega _{\mathrm{c}s}^{2})}} & 1-\displaystyle{\sum\limits_{s}\frac{\omega _{\mathrm{p}s}^{2}}{\omega ^{2}-\omega _{\mathrm{c}s}^{2}}}-n^{2} & 0 \\ n_{\bot }n_{\Vert } & 0 & 1-\displaystyle{\sum\limits_{s}\frac{\omega _{\mathrm{p}s}^{2}}{\omega ^{2}}}-n_{\bot }^{2} \end{array} \right) \label{Classical-matrix} \end{equation} The determinant of the sum of the matrixes (\ref{QED-matrix}) and (\ref {Classical-matrix}) is then evaluated to give the dispersion relation \begin{eqnarray} \fl 0 =\left( (1-n^{2})(1-4\xi )-\sum\limits_{s}\frac{\omega _{\mathrm{p}s}^{2}}{\omega ^{2}-\omega _{\mathrm{c}s}^{2}}+8\xi n_{\bot }^{2}\right) \times \nonumber \\ \fl \left[\! \left(\!\! (1-n_{\Vert }^{2})(1-4\xi )- \!\! \sum\limits_{s}\frac{\omega _{\mathrm{p}s}^{2}}{\omega ^{2}-\omega _{\mathrm{c}s}^{2}}\! \right) \!\! \left(\!\! 1+10\xi - \!\! \sum\limits_{s}\frac{\omega _{\mathrm{p}s}^{2}}{\omega ^{2}}-n_{\bot }^{2}\left( 1-4\xi \right)\!\! \right)\!\! -n_{\bot }^{2}n_{\Vert }^{2}\left( 1-4\xi \right) \! \right] - \nonumber \\ \fl \left( \sum\limits_{s}\frac{\omega _{\mathrm{p}s}^{2}\omega _{\mathrm{c}s}}{\omega (\omega ^{2}-\omega _{\mathrm{c}s}^{2})}\right) ^{2}\left( 1+10\xi -\sum\limits_{s}\frac{% \omega _{\mathrm{p}s}^{2}}{\omega ^{2}}-n_{\bot }^{2}\left( 1-4\xi \right) \right) \label{Full-DR} \end{eqnarray} The dispersion relation (\ref{Full-DR}) is the main result of the present paper. It describes wave propagation at any angle to the external magnetic field in a multi-component plasma, and it includes the QED effects associated with the external magnetic field. Thus, it applies to high frequency electromagnetic waves of any polarization, as well as electrostatic oscillations and low frequency waves, such as Alfv\'en waves. As a specific example of how the plasma dispersion relation is affected by the QED effects we consider the case of an electron--positron plasma with $\omega \sim \omega_{\mathrm{p}} \ll |\omega_{\mathrm{c}}|$. The dispersion relation relation for the ordinary mode propagating perpendicular to the background magnetic field, with strength $\sim 10^{10}\,\mathrm{T}$ is depicted in figure 1. A number of limiting cases of (% \ref{Full-DR}) have previously appeared in the literature. First, neglecting the QED-effects (i.e.\ letting $\xi \rightarrow 0$), we immediately obtain the standard dispersion relation for a cold multi-component plasma (see e.g.\ \cite{Chen-book}). Alternatively, letting $\omega _{\mathrm{p}}\rightarrow 0$, we note that the dispersion relation depends on the propagation angle relative to the magnetic field. Furthermore, we note that the indicies of refraction depend on the polarization even without a plasma. These QED-effects due to a strong external magnetic field are wellknown (often referred to as ''birefringence of vacuum''). Our dispersion relation in the limit $\omega _{\mathrm{p}}\rightarrow 0$ agrees with those of previous works, see e.g.\ \cite{Adler,Bialynicki-Birula1970}. The combined contribution from the QED-effects due to a strong magnetic field and a non-zero plasma density have previously been considered \cite{MSSBS2005} in the limit of parallel propagation and allowing for large amplitudes. Taking the limit of a small wave amplitudes in the dispersion relation (11) of reference \cite {MSSBS2005}, and letting $n_{\bot }\rightarrow 0$ in (\ref{Full-DR}) we obtain agreement with \cite{MSSBS2005}. \section{Conclusion} QED-effects associated with the external magnetic field are likely to be of importance in environments with extreme magnetic fields, in particular in the vicinity of astrophysical objects like pulsars and magnetars. For example, the radio silence of magnetars is assumed to be connected with QED-effects associated with the magnetar fields \cite{Baring-Harding}, which could reach $10^{10} - 10^{11}\,\mathrm{T}$ $\ $\ close to the surface. However, in the same environments, we also expect the presence of an electron-positron plasma \cite{Beskin-book}. Thus we evaluate (\ref{Full-DR}) with $% \sum_{s}=\sum_{e,p}$, where $e$ and $p$ denotes electrons and positrons, respectively. Considering propagation at an arbitrary angle to the magnetic field in an electron-positron plasma, letting $\omega _{\mathrm{p}e,\mathrm{p}p}\!\sim\! \omega\! \ll \! \left| \omega _{\mathrm{c}e,\mathrm{c}p}\right| $ , using $\xi\! \ll\! 1$ and noting that the factor $[\sum_{e,p}\omega _{\mathrm{p}s}^{2}\omega _{\mathrm{c}s}/\omega (\omega ^{2}-\omega _{\mathrm{c}s}^{2})]^{2}$ then becomes negligibly small (due to the approximate cancellation of the electron and positron contributions), we find from (% \ref{Full-DR}) that the dispersion relation separates in two modes that can be approximated by \begin{equation} 1-n^{2}+8\xi n_{\bot }^{2}+\frac{\omega _{\mathrm{p}}^{2}}{\omega _{\mathrm{c}}^{2}(1-4\xi )}% \approx 0 \label{Magnetosonic-DR} \end{equation} and \begin{equation} (1-n^{2})(1-4\xi ) - \left( -14\xi +\frac{\omega _{\mathrm{p}}^{2}}{\omega ^{2}}% \right) (1-n_{\Vert }^{2}) \approx 0 , \label{O-mode-DR} \end{equation} where $\omega _{\mathrm{p}}=(\omega _{\mathrm{p}e}^{2}+\omega _{\mathrm{p}p}^{2})^{1/2}$ is the total plasma frequency, and $\omega _{\mathrm{c}}$ $=eB_{0}/m$ is the magnitude of the electron (or positron) cyclotron frequency. In the vicinity of pulsars or magnetars where $\omega _{\mathrm{c}}\sim 10^{19}\! -\! 10^{21}\,\mathrm{rad\,s^{-1},}$ the last term of (\ref{Magnetosonic-DR}) is negligible unless the plasma density is extremely high. Omitting that term, the dispersion relation (\ref {Magnetosonic-DR}) is then the same as that used in \cite {Bialynicki-Birula1970} for the high phase velocity mode when considering photon-splitting. Similarly the ordinary mode described by (\ref {O-mode-DR}), reduces to the mode with the lower phase velocity of reference \cite {Bialynicki-Birula1970} when the plasma is removed. However, for the latter dispersion relation we note that a relatively modest plasma density is enough to significantly affect the propagation properties in the radio wave regime. To make a concrete estimate, we adopt the Goldreich--Julian density \begin{equation} n_{\mathrm{GJ}}=7\times 10^{15}\left(\frac{0.1}{\tau }\right)\left(\frac{B_{\mathrm{pulsar}}}{10^{8}}\right)\,\mathrm{m}^{-3} \label{Julian-Goldreich} \end{equation} where $\tau $ is the pulsar period time (in seconds) and $B_{\mathrm{pulsar}}$ the pulsar magnetic field (in tesla). The pair plasma density is expected to satisfy $% n_{e}=n_{p}=Mn_{\mathrm{GJ}}$, where $M$ is the multiplicity \cite {Beskin-book,Luo-etal}. Moderate estimates then give $M=10$ \cite{Luo-etal}. Choosing this value and letting $\tau =1\,\mathrm{s}$, we note that for magnetar field strengths, $B_{\mathrm{pulsar}}=10^{10}\,\mathrm{T}$, the term due to the plasma $\propto \omega _{\mathrm{p}}^{2}/\omega ^{2}$ in (\ref{O-mode-DR}) dominates over the term due to QED $\propto 14\xi $ for frequencies up to $% \omega \sim 10^{14} - 10^{15}\,\mathrm{rad\,s^{-1},}$ i.e.\ in the infrared regime and below. Furthermore, we note that photon splitting \cite {Adler,Bialynicki-Birula1970} as described by standard QED (i.e.\ with zero plasma density) requires that the phase velocity of the dispersion relation in (% \ref{Magnetosonic-DR}) is higher than that of (\ref{O-mode-DR}). While this is always true in the absence of a plasma, we note that for wave frequencies in the infra-red regime and below, the Goldreich--Julian density given by (\ref{Julian-Goldreich}) is enough to increase the phase velocity of the mode in (\ref{O-mode-DR}) above that of (\ref{Magnetosonic-DR}), unless we choose the period time $\tau $ extremely low. \ Thus we conclude that photon--photon splitting as described by vacuum theories is not likely to apply to magnetar atmospheres, unless the pair-production \cite {Beskin-book} responsible for the Goldreich-Julian expression is effectively suppressed. Wave cascade processes as a mechanism to explain the radio silence of magnetars \cite{Baring-Harding} could still be possible, but for densities of the order of (\ref{Julian-Goldreich}), plasma nonlinearities are likely to dominate over the pure QED effects. \section*{References}
Title: Introduction: Paleoheliosphere versus PaleoLISM
Abstract: Speculations that encounters with interstellar clouds modify the terrestrial climate have appeared in the scientific literature for over 85 years. This article introduces a series of articles that seek to give substance to these speculations by examining the exact mechanisms that link the pressure and composition of the interstellar medium surrounding the Sun to the physical properties of the inner heliosphere at the Earth.
https://export.arxiv.org/pdf/astro-ph/0601356
\newcommand\adsr{{Adv.~Space~Res.}}% \newcommand\jatp{{J.~Atmos.~Terres.~Phys.}}% \newcommand\aj{{Astron.~J.}}% \newcommand\actaa{{Acta Astron.}}% \newcommand\areps{{Ann.~Rev.~Earth \& Plan.~Sci.}}% \newcommand\araa{{Ann.~Rev.~Astron.~\& Astrophys.}}% \newcommand\apj{{Astrophys.~J.}}% \newcommand\apjl{{Astrophys.~J.~Let.}}% \newcommand\apjs{{Astrophys.~J.~Supl.}}% \newcommand\ao{{Appl.~Opt.}}% \newcommand\apss{{Astrophys.~\& Space Sci.}}% \newcommand\aap{{Astron.~\& Astrophys.}}% \newcommand\aapr{{Astron.~ \& Astrophys.~Rev.}}% \newcommand\aaps{{Astron.~ \& Astrophys. Supl.}}% \newcommand\azh{{Astron.~Zh.}}% \newcommand\baas{{Bull.~Amer.~Astron.~Soc.}}% \newcommand\caa{{Chinese Astron. Astrophys.}}% \newcommand\cjaa{{Chinese J. Astron. Astrophys.}}% \newcommand\icarus{{Icarus}}% \newcommand\jcap{{J. Cosmology Astropart. Phys.}}% \newcommand\jrasc{{JRASC}}% \newcommand\memras{{Mm.~Roy.~Astron.~Soc.}}% \newcommand\mnras{{Mon.~Not.~ Roy.~Astron.~Soc.}}% \newcommand\na{{New Astron.}}% \newcommand\nar{{New Astron.~Rev.}}% \newcommand\pra{{Phys.~Rev.~A}}% \newcommand\prb{{Phys.~Rev.~B}}% \newcommand\prc{{Phys.~Rev.~C}}% \newcommand\prd{{Phys.~Rev.~D}}% \newcommand\pre{{Phys.~Rev.~E}}% \newcommand\prl{{Phys.~Rev.~Lett.}}% \newcommand\pasa{{PASA}}% \newcommand\pasp{{Pub.~Astron.~Soc.~Pac.}}% \newcommand\pasj{{PASJ}}% \newcommand\qjras{{QJRAS}}% \newcommand\rmxaa{{Rev. Mexicana Astron. Astrofis.}}% \newcommand\skytel{{S\&T}}% \newcommand\solphys{{Sol.~Phys.}}% \newcommand\sovast{{Soviet~Ast.}}% \newcommand\ssr{{Space~Sci.~Rev.}}% \newcommand\zap{{Zeit.~Astrophy.}}% \newcommand\nat{{Nature}}% \newcommand\iaucirc{{IAU~Circ.}}% \newcommand\aplett{{Astrophys.~Lett.}}% \newcommand\apspr{{Astrophys.~Space~Phys.~Res.}}% \newcommand\bain{{Bull.~Astron.~Inst.~Netherlands}}% \newcommand\fcp{{Fund.~Cosmic~Phys.}}% \newcommand\gca{{Geochim.~Cosmochim.~Acta}}% \newcommand\grl{{Geophys.~Res. ~Lett.}}% \newcommand\jcp{{J.~Chem.~Phys.}}% \newcommand\jgr{{J.~Geophys.~Res.}}% \newcommand\jqsrt{{J.~Quant.~Spec.~Radiat.~Transf.}}% \newcommand\memsai{{Mem.~Soc.~Astron.~Italiana}}% \newcommand\nphysa{{Nucl.~Phys.~A}}% \newcommand\physrep{{Phys.~Rep.}}% \newcommand\physscr{{Phys.~Scr}}% \newcommand\planss{{Planet.~Space~Sci.}}% \newcommand\procspie{{Proc.~SPIE}}% \def\ebv{$E$(B-V)} \def\glong{$\ell$} \def\nHI{\hbox{$n$(H$^\mathrm {o }$)}} \def\nHII{\hbox{$n$(H$^\mathrm {+ }$)}} \def\Beten{\hbox{$^{ 10 }$Be}} \def\Cfourteen{\hbox{$^{ 14 }$C}} \def\tauhalf{\hbox{$\tau _ {1/2}$}} \def\Fesixty{\hbox{$^{ 60 }$Fe}} \def\Kforty{\hbox{$^{ 40 }$K}} \def\Clthirtysix{\hbox{$^{ 36 }$Cl}} \def\HI{\hbox{H$^ \mathrm {o }$}} \def\HeI{\hbox{He$^ \mathrm {o }$}} \newcommand{\deeg}{$^\circ$} \def\kms{\hbox{km s$^\mathrm {-1}$}} \def\cc{\hbox{cm$^{-3}$}} \def\cmtwo{\hbox{cm$^{-2}$}} \def\Rs{\hbox{R$_\mathrm {S}$}} \setcounter{chapter}{0} \articletitle[Introduction: Paleoheliosphere versus PaleoLISM]{Introduction: \\ Paleoheliosphere versus PaleoLISM} \chaptitlerunninghead{Paleoheliosphere versus PaleoLISM} \author{Priscilla C. Frisch} \affil{University of Chicago} \email{frisch@oddjob.uchicago.edu} \begin{keywords} Heliosphere, interstellar clouds, interstellar medium, cosmic rays, magnetosphere, atmosphere, climate, solar wind, paleoclimate \end{keywords} \section{The Underlying Query} If the solar galactic environment is to have a discernible effect on events on the surface of the Earth, it must be through a subtle and indirect influence on the terrestrial climate. The scientific and philosophical literature of the 18th, 19th and 20th centuries all include discussions of possible cosmic influences on the terrestrial climate, including the effect of cometary impacts on Earth (\nolinebreak \cite{Halley:1694}), and the diminished solar radiation from sunspots, which Herschel attributed to ``holes'' in the luminous fluid on the surface of the Sun\footnote{In this same paper Herschel commented that ``Whatever fanciful poets might say, in making the sun the abode of blessed spirits, or angry moralists devise, in pointing it out as a fit place for the punishment of the wicked, it does not appear that they had any other foundation for their assertions than mere opinion and vague surmise; but now I think myself authorized, \emph{upon astronomical principles,} to propose the sun as an inhabitable world, and am persuaded that the foregoing observations, with the conclusions I have drawn from them, are fully sufficient to answer every objection that may be made against it. '' These comments show that valuable data are not always interpreted correctly.} (\nolinebreak \cite{Herschel:1795}). The discovery of interstellar material in the 20th century led to speculations that encounters with dense clouds initiated the ice ages (\nolinebreak \cite{Shapley:1921}), and many papers appeared that explored the implications of such encounters, including the influence of interstellar material (ISM) on the interplanetary medium and planetary atmospheres (e.g. \cite{Fahr:1968,BegelmanRees:1976,McKayThomas:1978,Thomas:1978,McCrea:1975,TalbotNewman:1977,Willis:1978,ButlerNewmanTalbot:1978}). The ISM-modulated heliosphere was also believed to affect climate stability and astrospheres (e. g. \cite{Frisch:1993a,Frisch:1997,ZankFrisch:1999}). Recent advances in our understanding of the solar wind and heliosphere (e. g. \cite{WangRichardson:2005,Fahr:2004}) justify a new look at this age-old issue. This book addresses the underlying question: \begin{verse} \emph{How does the heliospheric interaction} \emph{with the interstellar medium affect the heliosphere, interplanetary medium, and Earth?} \end{verse} The heliosphere is the cavity in the interstellar medium created by the dynamic ram pressure of the radially expanding solar wind, a halo of plasma around the Sun and planets, dancing like a candle in the wind and regulating the flux of cosmic rays and interstellar material at the Earth. Neutral interstellar gas and large interstellar dust grains penetrate the heliosphere, but the solar wind acts as a buffer between the Earth and most other interstellar material and low energy galactic cosmic rays (GCR). Together the solar wind and interstellar medium determine the properties of the heliosphere. In the present epoch the densities of the solar wind and interstellar neutrals are approximately equal outside of the Jupiter orbit. Solar activity levels drive the heliosphere from within, and the physical properties of the surrounding interstellar cloud constrain the heliosphere from without, so that the boundary conditions of the heliosphere are set by interstellar material. Figure \ref{fig:1} shows the Sun and heliosphere in the setting of the Milky Way Galaxy. The answer to the question posed above lies in an interdisciplinary study of the coupling between the interstellar medium and the solar wind, and the effects that ISM variations have on the 1 AU environment of the Earth through this coupling. The articles in this book explore different viewpoints, including \emph{gedanken} experiments, as well as data-rich summaries of variations in the solar environment and paleoclimate data on cosmic ray flux variations at Earth. The book begins with the development of theoretical models of the heliosphere that demonstrate the sensitivity of the heliosphere to the variations in boundary conditions caused by the passage of the Sun through interstellar clouds (\cite{Zanketal:2006jos,PogorelovZank:2006jos}). A series of \emph{gedanken} experiments then yield the response of planetary magnetospheres to encounters with denser ISM (\cite{Parker:2006jos}). Variations in the galactic environment of the Sun, caused by the motions of the Sun and clouds through the Galaxy, are shown to occur for both long and short timescales (\cite{Shaviv:2006jos,FrischSlavin:2006jos}). The heliosphere acts as a buffer between the Earth and interstellar medium, so that dust and particle populations inside of the heliosphere, which have an interstellar origin, vary as the Sun traverses interstellar clouds. These buffering mechanisms determine the interplanetary medium\footnote{The buffering processes convert interstellar neutrals into low energy ions, which are convected outwards with the solar wind and accelerated to low cosmic ray energies that have an anomalous composition, including abundant elements with FIP$>$13.6 eV. The high energy galactic cosmic ray population incident on the heliosphere is also modulated.}. The properties of these buffering interactions are evaluated for heliosphere models that have been developed using boundary conditions appropriate for when the Sun traverses different types of interstellar clouds (\cite{Landgraf:2006jos,Moebiusetal:2006jos,FlorinskiZank:2006jos,Fahretal:2006jos}). The consequences of Sun-cloud encounters are then discussed in terms of the accretion of ISM onto the terrestrial atmosphere for dense cloud encounters, and the possibly extreme variations expected for cosmic ray modulation when interstellar densities vary substantially (\cite{Fahretal:2006jos,YeghikyanFahr:2006jos}). Radioisotope records on Earth extending backwards in time for over $\sim$0.5 Myrs, together with paleoclimate data, suggest that cosmic ray fluxes are related to climate. The galactic environment of the Sun must have left an imprint on the geological record through variations in the concentrations of radioactive isotopes (\cite{KirkbyCarslaw:2006jos}). The selection of topics in this book is based partly on scientific areas that have already been discussed in the literature. The authors who were invited to contribute chapters have previously studied the heliosphere or terrestrial response to variable ISM conditions or cosmic rays. Figure \ref{fig:1} shows the heliosphere in our setting of the Milky Way Galaxy. A postscript at the end of this chapter lists basic useful information. I introduce the term ``paleoheliosphere'' to represent the heliosphere in the past, when the boundary conditions set by the local interstellar material (LISM) may have differed substantially from the boundary conditions for the present-day heliosphere. The ``paleolism'' is the local ISM that once surrounded the heliosphere. \section[Addressing the Query: The Heliosphere for Different Interstellar Environments]{Addressing the Query: The Heliosphere and Particle Populations for Different Interstellar Environments} The solar wind drives the heliosphere from the inside, with the properties of the solar wind varying with ecliptic latitude and the phase of the 11-year solar activity cycle. The global heliosphere is the volume of space occupied by the supersonic and subsonic solar wind. Interstellar material forms the boundary conditions of the heliosphere, and the windward side of the heliosphere, or the ``upwind direction'', is defined by the interstellar velocity vector with respect to the Sun. The leeward side of the heliosphere is the ``downwind direction''. Figure \ref{fig:1} shows a cartoon of the present-day heliosphere, with labels for the major landmarks such as the termination shock, heliopause, and bow shock. In the present-day heliosphere, the transition from solar wind to interstellar plasma occurs at a contact discontinuity known as the ``heliopause'', which is formed where the total solar wind and interstellar pressures equilibrate (\nolinebreak \cite{Holzer:1989}). For a non-zero interstellar cloud velocity in the solar rest frame, the solar wind turns around at the heliopause and flows around the flanks of the heliosphere and into the downwind heliotail. Before reaching the heliopause, the supersonic solar wind slows to subsonic velocities at the ``termination shock'', where kinetic energy is converted to thermal energy. The subsonic solar wind region between the termination shock and heliopause is called the inner ``heliosheath''. The outer heliosheath lies just beyond the heliopause, where the pristine ISM is distorted by the ram pressure of the heliosphere. A bow shock, where the interstellar gas becomes subsonic, is expected to form ahead of the present-day heliosphere in the observed upwind direction of the ISM flow through the solar system. Large interstellar dust grains and interstellar atoms that remain neutral inside of the orbit of Earth, such as He, are gravitationally focused in the downwind direction. This ``focusing cone'' is traversed by the Earth every year in early December, and extends many AU from the Sun in the leeward direction (e.g. \cite{Landgraf:2000,Moebiusetal:2004,Frisch:2000amsci}). The heliotail itself extends $>10^3$ AU from the Sun in the downwind direction, forming a cosmic wake for the solar system. Of significance when considering the interaction of the heliosphere with an interstellar cloud is that neutral particles enter the heliosphere relatively unimpeded, after which they are ionized and convected outwards with the solar wind. Ions and small charged dust grains are magnetically deflected in the heliosheath around the flanks of the heliosphere (see Figure \ref{fig:1}). Space and astronomical data now confirm the basic milestones of the outer heliosphere. Voyager 1 crossed the termination shock at 94 AU on 16 December, 2004 (UT), and observed the signature of the termination shock on low-energy particle populations, the solar wind magnetic field, low-energy electrons and protons, and Langmuir radio emission (\nolinebreak \cite{Stoneetal:2005,Burlagaetal:2005,Gurnett:2005,Deckeretal:2005}). The present-day termination shock appears to be weak, with a solar wind velocity jump ratio (the ratio of upstream to downstream values) of $\sim$2.6 and a magnetic field compression ratio of $\sim$3. The magnetic wall that is predicted for the heliosphere (\nolinebreak \cite{Linde:1998,Ratkiewicz:1998}, Chapter 3 by Pogorelov and Zank) appears to have been detected through observations of magnetically aligned dust grains (\nolinebreak \cite{Frisch:2005L}), and the offset between upwind directions of interstellar \HI\ and \HeI\ (\nolinebreak \cite{Lallementetal:2005}). The compressed and heated \HI\ in the hydrogen wall region of the outer heliosheath has now been detected around a number of stars (\nolinebreak \cite{Woodetal:2005}). The present-day solar wind is the baseline for evaluating the heliosphere response to ISM variations in the following articles, so a short review of the solar wind is first presented. The remaining part of $\S$1.2 introduces the topics in the following articles in terms of the underlying query of the book. \subsection{The Present Day Solar Wind } The solar wind originates in the million degree solar corona that expands radially outwards, with a density $\sim 1/R^2_\mathrm{ S }$ where $R_\mathrm{ S }$ is the distance to the Sun, and contains both features that corotate with the Sun, and transient structures (\nolinebreak e.~g.~\cite{Gosling:1996}). The properties of the solar wind vary with the phase of the solar magnetic activity cycle and with ecliptic latitude. The best historical indicator of solar magnetic activity levels is the number of sunspots, first detected by Galileo in 1610, which are magnetic storms in the convective zone of the Sun. Sunspot numbers indicate that the magnetic activity levels fluctuate with a $\sim$11 year cycle, or the ``solar cycle'', and solar maximum/minimum corresponds to the maximum/minimum of sunspot numbers. The magnetic polarity of the Sun varies with a $\sim$22 year cycle. During solar maximum, a low-speed wind, with velocity $\sim$300--600 \kms\ and density $\sim 6-10$ particles \cc\ at 1 AU, extends over most of the solar disk. Open magnetic field lines\footnote{Open magnetic field lines are formed in coronal holes that reconnect in the outer heliosphere and contain low density and very high speed, $\sim$700 \kms, solar wind.} are limited to solar pole regions. A neutral current sheet $\sim$0.4 AU thick forms between the solar wind containing negative magnetic polarity fields and the solar wind that contains positive magnetic polarity fields. The neutral current sheet reaches its largest inclination ($\ge 70^\circ$) during solar maximum. During the conditions of solar minimum, a high speed wind with velocity $\sim$600--800 \kms\ and density $\sim$5 \cc\ is accelerated in the open magnetic flux lines in coronal holes. During mininum, the high speed wind and open field lines extend from the polar regions down to latitudes of $\leq 40$\deeg (\nolinebreak \cite{Smithetal:2003,Richardson:1995}). The higher solar wind momentum flux associated with solar minimum conditions produces an upwind termination shock that is $\sim$5--40 AU more distant in the upwind direction than during solar maximum conditions (e.g. \cite{SchererFahr:2003,ZankMueller:2003,Whang:2004}). The alignment and strength of the solar magnetic multipoles depend on the phase of the solar cycle (\cite{Bravoetal:1998}). During solar minimum conditions, the magnetic field is dominated by the dipole and hexapole moments, and the dipole moment is generally aligned with the solar rotation axis. Sunspots migrate from high to low heliographic latitudes. The magnetic poles follow the coronal holes to the solar equator as solar activity increases. During the solar maximum period, the galactic cosmic rays undergo their maximum modulation, the dipole component of the magnetic field is minimized, the quadrapole moment dominates, and the polarity of the solar magnetic field reverses (\cite{LockwoodWebber:2005}, Figure \ref{fig:2}). Over historic times, the cosmic ray modulation by the heliosphere correlates better with the open magnetic flux line coverage than with sunspot numbers (\cite{McCrackenetal:2004}). Variable cosmic ray modulation produced by a variable heliosphere may be a primary factor in both solar and ISM forcing of the terrestrial climate. The heliosphere modulation of cosmic rays is well established. John Simpson, to whom this book is dedicated, initiated a program 5 solar cycles ago in 1951 to monitor cosmic ray fluxes on Earth using high-altitude neutron detectors (\cite{Simpson:2001}). The results show a pronounced anticorrelation between cosmic ray flux levels and solar sunspot numbers, which trace the 11-year Schwabe magnetic activity cycle, and which also show that the polarity of the solar magnetic field affects cosmic ray modulation (see Figure \ref{fig:2}). The articles in this book show convincingly that the ISM also modulates the heliosphere, and the effect of the solar wind on the heliosphere must be differentiated from the influence of interstellar matter. Variations in solar activity levels are also seen over $\sim 100-200$ year timescales, such as the absence of sunspots during the Maunder Minimum in the 17th century. Modern climate records show that the Maunder Minimum corresponded to extremely cold weather, and radioisotope records show that the flux of cosmic rays was unusually high at this time (see Kirkby and Carslaw, Chapter 12). Similar effects will occur from the modulation of galactic cosmic rays by the passage of the Sun through an interstellar cloud. These temporal and latitudinal variations in the solar wind momentum flux produce an asymmetric heliosphere, which varies with time. Any possible historical signature of the ISM on the heliosphere must first be distinguished from variations driven by the solar wind itself. \subsection{Present Day Heliosphere and Sensitivity to ISM} The ISM forms the boundary conditions of the heliosphere, so that encounters with interstellar clouds will affect the global heliosphere, the interplanetary medium, and the inner heliosphere region where the Earth is located. Today an interstellar wind passes through the solar system at --26.3 \kms\ (\cite{Witte:2004}). An entering parcel of ISM takes about 20 years to reach the inner heliosphere, so that ISM near the Earth is constantly replenished with new inflowing material. This warm gas is low density and partially ionized, with temperature $T \sim$6,300 K, and densities of neutral and ionized matter of \nHI$\sim 0.2$ \cc, and \nHII$\sim 0.1$ \cc. An elementary perspective of the response of the heliosphere to interstellar pressures is given by an analytical expression for the heliopause distance based on the locus of positions where the solar wind ram pressure, $P_\mathrm{SW}$, and the total interstellar pressure equilibrate (\cite{Holzer:1989}). The solar wind density $\rho$ falls off as $\sim 1/R^2$, where $R$ is the distance to the Sun, while the velocity $v$ is relatively constant. At 1 AU the solar wind ram pressure is $P_\mathrm{SW,1AU} \sim \rho~v^2 $ so the heliosphere distance, $R_\mathrm{HP} $, is given by: \begin{eqnarray*} P_\mathrm{SW,1AU}/R_\mathrm{HP}^2 \sim P_\mathrm{B} +P_\mathrm{Ions, thermal} + P_\mathrm{Ions, ram} + P_\mathrm{Dust} + P_\mathrm{CR} \end{eqnarray*} The interstellar pressure terms include the magnetic pressure $P_\mathrm{B}$, the thermal, $P_\mathrm{Ions, thermal}$, and the ram, $P_\mathrm{Ions, ram}$, pressures of the charged gas, and the pressures of dust grains, $P_\mathrm{Dust}$, and cosmic rays, $P_\mathrm{CR}$, which are excluded by heliosphere magnetic fields and plasma. Some interstellar neutrals convert to ions through charge exchange with compressed interstellar proton gas in heliosheath regions, adding to the confining pressure. An important response characteristic is that, for many clouds, the encounter will be ram-pressure dominated, where $P_\mathrm{ram} \sim m v^2$ for interstellar cloud mass density $m$ and relative Sun-cloud velocity $v$, so that variations in the cloud velocity perturb the heliosphere even if the thermal pressures remain constant. The multifluid, magnetohydrodynamic (MHD), hydrodynamic and hybrid approaches used in the following chapters provide much more substantial models for the heliosphere, and include the coupling between neutrals and plasma, and field-particle interactions. These sophisticated models predict variations in the global heliosphere in the face of changing interstellar boundary conditions, and for a range of different cloud types. Although impossible to model a solar encounter with every type of interstellar cloud, the following articles include discussions of many of the extremes of the interstellar parameter space, including low density gas with a range of velocities, very tenuous plasma, high velocity clouds, dense ISM, and magnetized material for a range of field orientations and strengths. The discussions in these chapters extrapolate from our best theoretical understanding of the heliosphere boundary conditions today to values that differ, in some cases dramatically, from the boundary conditions that prevailed at the beginning of the third millennium in the Gregorian calendar. The Sun has been, and will be, subjected to many different physical environments over its lifetime. Theoretical heliosphere models yield the properties of the solar wind-ISM interaction for these different environments, which in turn determine the nature and properties of interstellar populations inside of the heliosphere for a range of galactic environments. These models form the foundation for understanding the significance of our galactic environment for the Earth. The interstellar parameter space is explored by Zank et al. (Chapter 2), where 28 sets of boundary conditions are evaluated with computationally efficient multifluid models. Moebius et al. (Chapter 8), Fahr et al. (Chapter 9), Florinski and Zank (Chapter 10), and Yeghikyan and Fahr (Chapter 11) also develop heliosphere models for a range of interstellar conditions. Together these models evaluate the heliosphere response to interstellar density, temperature, and velocity variations of factors of $\sim 10^{9}$, $\sim 10^{5}$, and $\sim 10^{2}$, respectively. The interstellar magnetic field introduces an asymmetric pressure on the heliosphere, affecting the heliosphere current sheet and cosmic ray modulation. Pogorelov and Zank (Chapter 3) use MHD models to probe the heliosphere response to the interstellar magnetic field, including charge exchange between the neutrals and solar wind. The resulting asymmetry provides a test of the magnetic field direction, and shows strong differences between cases where the interstellar flow is parallel, instead of perpendicular, to the interstellar magnetic field direction. Since the random component of the interstellar magnetic field is stronger, on the average, than the ordered component, particularly in spiral arm regions where active star formation occurs, a range of interstellar magnetic field strengths and orientations are expected over the solar lifetime (Shaviv, Chapter 5, and Frisch and Slavin, Chapter 6). \subsection{Planetary Magnetospheres} The Earth's magnetosphere acts as a buffer between the solar wind and atmosphere, and as such is an ingredient in understanding the effect of our galactic environment on the Earth. The decreasing solar wind density in the outer heliosphere results in an interplanetary medium around outer planets that is more sensitive to ISM variations than for inner planets, with implications for the magnetospheres of Jupiter, Neptune, and Uranus. Most topics in this book are already considered in the scientific literature, but questions about magnetosphere variations from an ISM-modulated heliosphere have received scant attention. In a quintessential \emph{gedanken} experiment, Parker explores the interaction between magnetospheres and the solar wind for variations in the interstellar density, and for inner versus outer planets (Chapter 4). \subsection{Short and Long Term Variations in the Galactic Environment} There is every reason to expect that the galactic environment of the Sun varies over geological timescales. The Sun moves through space at a velocity of 13--20 \kms, and interstellar clouds have velocities ranging up to hundreds of \kms. The Arecibo Millennium survey showed that $\sim$25\% of the mass contained in interstellar \HI, including both warm and cold ISM, is in clouds traveling with velocities $\ge$10 \kms\ through the local standard of rest (\nolinebreak \cite{HTI}). Thus Sun-cloud encounters with relative velocities exceeding 25 \kms\ are quite likely, and for a typical cloud length of $\sim$1 pc the cloud transit time would be $\sim$40,000 years. The many types of ISM traversed by the Sun during the past several million years have affected the heliosphere, the inner solar system, and the flux of anomalous and galactic cosmic rays at Earth (\cite{FrischYork:1986,Frisch:1997,Frisch:1998}). For the past $\sim$3 Myrs the Sun has been in a nearly empty region of space, the ``Local Bubble'', with very low densities of $<$10$^{-26}$ gr \cc. Within the past 44,000--140,000 years the Sun entered a flow of tenuous, partly neutral ISM, nick-named the ``Local Fluff'', with density $\sim$60 times higher (Chapter 6). This transition was accompanied by the appearance of interstellar dust and neutrals in the heliosphere, along with the pickup ion and anomalous cosmic ray populations. Galactic cosmic ray modulation was affected, providing a possible link between our galactic environment and climate. Intriguingly, the averaged cosmic ray flux at Earth, as traced by \Beten\ records, was lower in the past $\sim$135 kyrs than for earlier times (Chapter 12). Was the decrease in the galactic cosmic ray flux $\sim$135 kyrs ago caused by an increase in modulation as the Sun entered the Local Fluff? The galactic environment of the Sun also varies quite dramatically over long time scales, as discussed by Shaviv (Chapter 5). Over its 4.5 billion year lifetime, the Sun traverses spiral arm and interarm regions, with atomic densities varying from less than $10^{-26.1}$ g \cc\ to over $10^{-20.1}$ g \cc, and temperatures ranging over 7 orders of magnitude, 10--10$^7$ K. The Sun is now in low density space between the Perseus and Sagittarius spiral arms, and on the inner edge of what is known as the Orion Spur on the Local Arm. The Local Arm is not shown in Figure \ref{fig:1}, as is consistent with the usual Galaxy depictions. The Local Arm does not appear to be a grand design spiral shock (Bochkarev, 1984\nocite{Bochkarev:1984}). The Sun has a systematic motion of 13--20 km s$^{-1}$ with respect to the nearest stars, corresponding to $\sim$3--4 AU per year. The Local Interstellar Cloud (LIC) now surrounding the Sun traverses the heliosphere at $\sim$5.5 AU per year. The Sun oscillates vertically through the galactic plane once every $\sim$34 Myrs, and orbits the center of the Milky Way Galaxy once per $\sim$220 Myrs. Shaviv evaluates variations in the galactic environment of the Sun over long timescales. This bold discussion compares various geologic records of cosmic ray flux variations, based on radioisotope data that sample timescales of $\sim 10^8$ years, with models of the Milky Way Galaxy spiral arm pattern to reconstruct the timing of the Sun's passage through spiral arms. The chapter concludes that star formation in spiral arms leaves a signature on the radioisotope records of the solar system. Frisch and Slavin (Chapter 6) reconstruct short-term variations of the galactic environment of the Sun using observations of interstellar matter towards nearby stars and inside of the solar system. Radiative transfer models of the LIC show that ionization varies across this low density cloud, so that the heliosphere boundary conditions vary from radiative transfer considerations alone as the Sun traverses the LIC. Cloud transitions are predicted during the past $\sim$3 Myrs, including the departure of the Sun from the Local Bubble interior 44,000--140,000 years ago, and entry into the surrounding cloud 1000--40,000 years ago. \subsection{Interstellar Dust} The particle populations formed by the interactions between the solar wind and interstellar dust, gas, and cosmic rays are emissaries between the cosmos and inner heliosphere, varying as the Sun moves through clouds. About $\sim$1\% of the mass of the cloud surrounding the Sun is contained in interstellar dust grains. The largest of these charged grains, mass $> 10^{-13} $ g, have large magnetic Larmor radii of $>$500 AU at the heliopause for an interstellar field of $\sim$1.5 $\mu$G, and flow into the solar system. The Earth passes through the gravitational focusing cone formed by these grains early each December. The smallest charged grains, mass$< 10^{-14.5} $ g and radii$< 0.01 ~\mu$m, have Larmor radii of $\sim$20 AU, depending on the magnetic field strength and radiation field, and are deflected around the heliosheath (\cite{Frischetal:1999}). Interstellar dust grains are measured in the inner heliosphere within $\sim$5 AU of the Sun, and over the solar poles, by satellites such as Ulysses, Galileo and Cassini. Landgraf (Chapter 7) reviews the properties of the interaction between interstellar dust and the solar wind, and speculates on the changes that might be expected from an encounter with a dense interstellar cloud. Should it some day be possible to compare the ratio of large to small interstellar dust grains on the surfaces of the inner versus outer planets, it would become possible to disentangle cloud encounters from solar activity effects. At the very large end of the dust population mass spectrum we find interstellar micrometeorites, with masses $\sim$3 $\times$ 10$^{-7}$ g, open orbits, and inflow velocities greater than the 42 \kms\ escape velocity from the solar system at 1 AU. These interstellar objects, detected by radar as they impact the atmosphere, evidently originate in circumstellar disks such as that around $\beta$ Pictoris, and in the interior of the Local Bubble (\cite{Baggaley:2000,Meiseletal:2002}). These objects do not collisionally couple to the interstellar gas (\cite{GruenLandgraf:2000}), and should not vary with the type of ISM surrounding the Sun. \subsection{Particle Populations in the Inner and Outer Heliosphere} Presently, low energy interstellar neutrals, high energy galactic cosmic rays, and interstellar dust all enter the heliosphere. The characteristics of each of these populations and their secondary products are modified as the Sun transits the ISM, or the cloud ionization changes. The first ionization potential (FIP) of \HI\ is 13.6 eV. Neutral interstellar atoms with FIP$<$13.6 eV are ionized in nearly all interstellar clouds because the main source of interstellar opacity is \HI. Interstellar ions are deflected around the heliosheath, so the result is that only interstellar atoms with FIP$>$13.6 eV enter the heliosphere where they are then destroyed, primarily by charge exchange with solar wind ions. The density of interstellar neutrals in the inner heliosphere depends on the density and ionization of the surrounding cloud, the ionization (or ``filtration'') of those neutrals by the heliosheath, and the subsequent interactions with the solar wind inside of the heliosphere. Secondary products produced by solar wind interactions with interstellar neutrals inside of the heliosphere include pickup ions \footnote{The pickup ions are interstellar neutrals formed by charge exchange with the solar wind. Energetic neutral atoms are formed by energetic ions that capture an electron from a low energy neutral by charge exchange. The gravitational focusing cone contains heavy elements (mainly He) that are predominantly ionized inside of 1 AU and therefore gravitationally focused downwind of the Sun (Chapter 8). Large interstellar dust grains are also gravitationally focused (Chapter 7). The anomalous cosmic ray population is formed from pickup ions accelerated to low cosmic ray energies, $<$\nolinebreak 1 GeV, in the solar wind and at the termination shock, and then subjected to the same modulation and propagation processes as galactic cosmic rays (\cite{Jokipii:2004}).}, energetic neutral atoms, the gravitational focusing cone formed by helium (also seen in dust), and the anomalous cosmic ray population with energies $<$\nolinebreak 1 GeV. Interstellar neutrals inside of the heliosphere, and the heliosphere itself, form a coupled system that together respond to variations in the heliosphere boundary conditions. Moebius et al. (Chapter 8) model the heliosphere for several different conditions, and then probe the response of the inner heliosphere to the density of interstellar neutrals flowing into this ISM-modified heliosphere. At 1 AU, the neutral densities, particle populations derived from interstellar neutrals, and characteristics of the helium focusing cone all respond to variations in the interstellar boundary conditions. For some cases, increased neutral fluxes fall on the atmosphere of Earth (also see Yeghikyan and Fahr, Chapter 11). The velocity structure of the ISM appears to vary on subparsec scale lengths (Frisch and Slavin, Chapter 6), and these variations may in some cases result in significant modifications of the inner heliosphere, particularly the gravitational focusing cone, when all other interstellar parameters such as thermal pressure are invariant (Zank et al., Chapter 2, Moebius et al., Chapter 8). The most readily available diagnostics of the paleoheliosphere are radioisotopes, formed by cosmic ray spallation on the atmosphere, interplanetary and interstellar dust, and meteorites. Thus, the evaluation of cosmic ray modulation for various types of interstellar cloud boundary conditions is a key part of understanding the paleoclimate records that might trace the solar journey through the Milky Way Galaxy. Fahr et al. (Chapter 9) and Florinski and Zank (Chapter 10) use our understanding of galactic cosmic ray modulation in the modern-day heliosphere as a basis for making detailed calculations of the response of the paleoheliosphere, or the heliosphere as it once was, to the paleolism, or the local interstellar medium that once surrounded the Sun. The predictions of these calculations are quite intriguing. Both the termination shock compression ratio and the solar wind turbulence spectrum may vary dramatically with different environments, as mass-loading by pickup ions and the heliosphere properties vary. The problem of galactic cosmic ray modulation in an ISM-forced heliosphere is extremely important to understanding the paleoheliosphere signature in the terrestrial isotope record. Today, galactic cosmic rays (GCR) with energies $\ge 0.25$ GeV penetrate the solar system, and anomalous cosmic rays (energies $<1$ GeV) are formed from accelerated pickup ions. The cosmic ray flux at Earth is sampled by geological radioisotope records, as reviewed Kirkby and Carslaw (Chapter 12, also see Florinski and Zank, Chapter 10). Astronomical data indicates that the Sun has emerged from a region of space with virtually no neutral ISM within the past $\sim$0.4--1.5 10$^5$ years, and entered the Local Fluff (Chapter 6). The GCR modulation discontinuity that accompanied this transition may be in the geologic record, which show lower cosmic ray fluxes at Earth, on the average, for the past 135 kyrs years than the 135 kyrs before that (\cite{Christl:2004}). \subsection{Atmosphere Accretion from Dense Cloud Encounters} Harlow Shapley (1921) suggested \nocite{Shapley:1921} that an encounter between the Sun and giant dust clouds in Orion may have perturbed the terrestrial climate and caused ice ages. The discovery of interstellar \HI\ and \HeI\ inside the heliosphere was soon followed by studies of the ISM influence on the atmosphere for dense cloud conditions (\cite{Fahr:1968,BegelmanRees:1976,McKayThomas:1978,Thomas:1978,McCrea:1975,TalbotNewman:1977,Willis:1978,ButlerNewmanTalbot:1978}). Yeghikyan and Fahr (Chapter 11), evaluate the density of ISM at the Earth based on models describing the heliosphere inside of an dense cloud, and the interactions between the solar wind and ISM for these dense cloud conditions (also see Chapter 9, by Fahr et al.). These models then yield the concentration of interstellar hydrogen at the Earth, and the flow of water downward towards the Earth's surface, as a function of the dense cloud density. Significant atmosphere modifications are predicted in some cases. Enhanced neutral populations at 1 AU for a somewhat lower interstellar cloud density regime are discussed in Chapter 8, by Moebius et al. \subsection{Possible Effects of Cosmic Rays} Both solar activity cycles (Figure \ref{fig:2}) and ISM variations modulate the cosmic ray flux in the heliosphere, and Kirkby and Carslaw (Chapter 12) compare galactic cosmic ray records with paleoclimate archives. They examine sources of climate forcing such as solar irradiance and cosmic ray fluxes, and conclude that arguments in favor of cosmic ray climate forcing are strong although the mechanism is uncertain. This relation between cosmic ray flux levels and the climate is shown by radioisotope records and climate archives, such as ice cores, stalagmites, and ice-rafted debris, and for modern times, by historical records. Paleoclimate archives include terrestrial records of cosmic ray spallation in the atmosphere, as traced by radioisotopes with short half-lives (\tauhalf), e.~g.~ \Cfourteen\ (\tauhalf=5,730 yrs) and \Beten\ (\tauhalf=1.6 Myrs). Possible mechanisms linking the cosmic ray flux at 1 AU and the climate include cloud nucleation by cosmic rays, and the global electrical circuit (see Chapter 12 and \cite{RobleHaysII:1979}). The discussion in Chapter 12 provides persuasive evidence linking the surface temperature to cosmic ray fluxes at Earth. The anticorrelation between sunspot number and cosmic ray fluxes in Figure \ref{fig:2} shows the heliosphere role in cosmic ray modulation; this mechanism must have also been a prominent mechanism for relating the ISM-modulated heliosphere with the climate. Fortunately this hypothesis is also verifiable by comparing paleoclimate data with astronomical data on the timing of cloud transitions. The radioisotope records also indicate that cosmic ray fluctuations have occurred over longer timescales of many $10 ^8$ years. Shaviv compares the \Clthirtysix\ (\tauhalf $\sim$0.3 Myrs) and \Kforty\ (\tauhalf $\sim$1.3 Gyr) cosmic ray exposure records in iron meteorites (Chapter 5), but in this case to obtain cosmic ray flux increases due to the Sun's location in spiral arms where active star formation occurs. A number of studies, none convincing, have invoked the geological \Beten\ record, as a proxy for cosmic ray fluxes at Earth, to infer historical encounters with interstellar clouds. As a way of dating the Loop I supernova remnant, it was suggested that the relative constancy of \Beten\ in sea sediments precluded a strong nearby X-ray source within the past $\sim$2 Myrs (\cite{Frisch:1981}). Sonett (1992) \nocite{Sonett:1992} suggested that peaks in \Beten\ layers 35,000 and 65,000 years ago resulted from a compressed heliosphere caused by the passage of a high-velocity interstellar shock. This extreme heliosphere compression expected for a rapidly moving cloud is supported by heliosphere models (Chapter 2). Structure in the \Beten\ peaks has also been related to spatial structure in the local ISM (\cite{Frisch:1997}), and solar wind turbulence caused by mass-loading of interstellar neutrals may supply the required mechanism. Global geomagnetic excursions such as the events $\sim$32 kyr and $\sim$40 kyr ago also affect the \Beten\ record, and can not be ignored (\nolinebreak \cite{Christl:2004}). Indeed, Figure \ref{fig:2} shows the sensitivity of galactic cosmic ray fluxes on Earth to geomagnetic latitude. \section{Closing Comments} This brief summary of the scientific question motivating this book does not relay the full significance of the galactic environment of the Sun to the heliosphere and Earth; the following chapters provide deeper insights into this question. Historical and paleoclimate data show a correspondence between high cosmic ray flux levels and cool temperatures on Earth (\cite{Parker:1996}). The disappearance of sunspots for extended periods of time, such as the Maunder Minimum in the years 1645 to 1715, shows up in terrestrial radioisotope records such as \Beten\ in ice cores (Chapter 12). The solar magnetic activity cycle was present during this period, and cosmic ray modulation by the heliosphere was still evident (\cite{McCrackenetal:2004}). The \Beten\ record now extends to $\sim$10$^5$ years before present, raising the hope that encounters between the Sun and interstellar clouds can be separated from solar activity effects, and from the global signature of geomagnetic pole wandering. Sunspots have long been controversial as an influence on the terrestrial climate. Sir William Herschel carefully observed them, and postulated that diminished solar radiation at Earth during sunspot maximum affected the terrestrial climate (1801). \nocite{Herschel:1801} Prof. Langley (1876) \nocite{Langley:1876} measured the radiative heat from sunspot umbral and penumbral regions, and concluded the $<$0.1\% solar radiation decrease associated with sunspots was inadequate to affect the climate. Climate records show that the Maunder Minimum and other periods of low solar activity levels have been exceptionally cold, which implicates high cosmic ray fluxes with cold climate conditions. Solar activity levels have returned to historic highs in the past few decades (\cite{CaballeroMcCracken:2004}), and the historic correlations indicate these high levels also yield warm climate conditions. Unfortunately, these scientific conclusions also impact the politically loaded issue of global warming. The possibility that the cosmos has affected the terrestrial climate is a longtime source of speculation, with many of the first discussions focused on explaining the ``Universal Deluge". In 1694 Edmond Halley presented his thoughts to the Royal Society as to whether the "casual Shock of a Comet, or other transient Body" might instantly alter the axis orientation or diurnal rotation of the Earth, thus disturbing sea levels, or whether the impact of a comet could explain the presence of "vast Quantities of Earth and high Cliffs upon Beds of Shells, which once were the Bottom of the Sea" (\cite{Halley:1694}). Halley's speculation has resurfaced in the hypothesis that the impact of a comet led to the extinction of dinosaurs 65 Myrs ago at the Cretaceous-Tertiary boundary (\nolinebreak \cite{Alvarez:1982}). The common sense disclaimer that accompanied Halley's discussion is timeless: \emph{ ``... the Almighty generally making us of Natural Means to bring about his Will, I thought it not amiss to give this Honourable Society an Account of some Thoughts that occurr'd to me on this Subject; wherein, if I err, I shall find myself in very good Company.''} The articles in this volume show firmly that the interaction between the heliosphere and ISM depends on the detailed boundary conditions set for the heliosphere by each type of interstellar cloud encountered by the Sun, and that the galactic environment of the Sun changes over both geologically short time scales of $< 10^5$ years, and long time scales of $> 10 ^7$ years. This interaction, in turn, affects the flux of gas, dust, and energetic particles in the inner heliosphere. The discussions in this book also apply to the study of astrospheres around cool stars, which are expected to have similar properties as the heliosphere. Is the historical astrosphere of a star a factor in climate stability for planetary systems? I think so (\cite{FrischYork:1986}). If so, then the sample of $\sim$100 detected extrasolar planetary systems can be narrowed to those that are the most likely to harbor technological civilizations by evaluating the astrosphere characteristics suitable to the space trajectory of each star (\nolinebreak \cite{Frisch:1993a}). Astrospheres have now been detected towards $\sim$60\% of the observed cool stars within 10 pc (\nolinebreak \cite{Woodetal:2005}), and extensive efforts to detect Earth-sized exoplanets are underway. Perhaps some day these questions will be answered. \vspace*{0.1in} \emph{Acknowledgments:} The author thanks Dr. Clifford Lopate, of the University of New Hampshire, for providing Figure \ref{fig:2}, and Dr. Lopate thanks NSF Grant 03-39527 for supporting the research displayed in this figure. The author thanks NASA for supporting her research, including grants NAG5-13107 and NNG05GD36G. Additional support has been provided by grants NAG 5-13558 and NAG5-11999. This article will appear in the book ''Solar Journey: The Significance of Our Galactic Environment for the Heliosphere and Earth'', Springer, in press (2006), editor P. C. Frisch. \begin{table}[h] \caption[Commonly Used Terms and Acronyms.]{Commonly Used Terms and Acronyms } \begin{tabular}{ll} Object & Description \\ \hline \emph{Interstellar:}& \\ Interstellar Material, ISM & Atoms in the space between stars \\ Local Fluff or CLIC & ISM within $\sim$30 pc, density $10^{-24.3}$ g \cc \\ & CLIC=Cluster of Local Interstellar Clouds \\ Local Interstellar Cloud, LIC & The cloud feeding ISM into the solar system \\ Local Bubble, LB & Nearby ISM with density $< 10 ^{-26.1}$ g \cc\\ & \\ \emph{Heliosphere:}& \\ Solar Wind, SW & Solar plasma expanding to form heliosphere \\ & Density$\sim$5 ions \cc, velocity $\sim$450 km s$^{-1}$ \\ & at Earth \\ Neutral Current Sheet & Thin neutral region separating SW \\ & with opposite magnetic polarities \\ Heliosphere, HS & Region of space containing the solar wind \\ Termination Shock, TS & Shock where solar wind becomes subsonic \\ & TS at $\sim$94 AU on 16 December, 2004 \\ Heliosheath & Subsonic solar wind, outside TS \\ Heliosphere Bow Shock & Shock where LIC becomes subsonic \\ Focusing Cone & Gravitationally focused ISM dust \\ & and helium gas downwind of the Sun\\ & \\ \multicolumn{2}{l}{\emph{Interstellar Products in the Heliosphere:}} \\ Pickup Ions, PUI & Ions from SW-ISM charge exchange \\ Energetic Neutral Atoms & ENAs, Energetic atoms formed by \\ & charge exchange with ions \\ Cosmic Rays: & \\ \hspace{3mm}Anomalous, ACR & Accelerated pickup ions, energy $<$1 GeV \\ \hspace{3mm}Galactic, GCR & From supernova, energy $>$1 GeV at Earth\\ \hline \end{tabular} \end{table} \section*{Postscript: Definitions} The nine planets of the solar system (including Pluto as a planet) extend out to 39 AU, compared to the distance of the solar wind termination shock in the upwind direction of 94 AU. The Earth is 8.3 light minutes from the Sun, versus the $\sim$0.5 light day distance to the upwind termination shock of the solar wind. The ecliptic and galactic planes are tilted with respect to each other by $\sim$60$^\circ$, and the north ecliptic pole points towards the galactic coordinates $\ell$=96.4$^\circ$ and $b$=+29.8$^\circ$. This tilt allows the separation of large scale ecliptic and large scale galactic phenomena by geometric considerations. Acronyms are used throughout this book, and some of these are listed in Table 1. For those new to this subject, an astronomical unit, AU, is the distance between the Earth and Sun. A parsec, pc, is 206,000 AU, 3.3 light years (ly), or 3.1 $\times$ 10$^{18}$ cm. For comparison, the Earth is 8.3 light minutes from the Sun, and the nearest star, $\alpha$ Cen, is 1.3 pc from the Sun. The planet Pluto is at 39 AU, versus the 94 AU termination shock distance. \begin{chapthebibliography}{} \bibitem[{Alvarez}, 1982]{Alvarez:1982} {Alvarez}, L.~W. (1982). \newblock {Experimental evidence that an Asteroid Impact LED to the Extinction of many Species 65 Million Years Ago}. \newblock {\em NASA STI/Recon Technical Report N}, 83:33813. \bibitem[{Baggaley}, 2000]{Baggaley:2000} {Baggaley}, W.~J. (2000). \newblock {Advanced Meteor Orbit Radar Observations of Interstellar Meteoroids}. \newblock {\em \jgr}, 105:10353--10362. \bibitem[{Begelman} and {Rees}, 1976]{BegelmanRees:1976} {Begelman}, M.~C. and {Rees}, M.~J. (1976). \newblock {Can Cosmic Clouds Cause Climatic Catastrophes}. \newblock {\em \nat}, 261:298. \bibitem[{Bochkarev}, 1984]{Bochkarev:1984} {Bochkarev}, N.~G. (1984). \newblock {Large-Scale Bubble Structure of the Interstellar Medium and Properties of the Local Spiral Arm}. \newblock {In \emph{Local Interstellar Medium},} \newblock {Eds. Y.~Kondo, F.~Bruhweiler, and B.~Savage} \bibitem[{Bravo} et~al., 1998]{Bravoetal:1998} {Bravo}, S., {Stewart}, G.~A., and {Blanco-Cano}, X. (1998). \newblock {The Varying Multipolar Structure of the Sun's Magnetic Field and the Evolution of the Solar Magnetosphere Through the Solar Cycle}. \newblock {\em \solphys}, 179:223--235. \bibitem[{Burlaga} et~al., 2005]{Burlagaetal:2005} {Burlaga}, L.~F., {Ness}, N.~F., {Acu{\~n}a}, M.~H., {Lepping}, R.~P., {Connerney}, J.~E.~P., {Stone}, E.~C., and {McDonald}, F.~B. (2005). \newblock {Crossing the Termination Shock into the Heliosheath: Magnetic Fields}. \newblock {\em Science}, 309:2027--2029. \bibitem[{Butler} et~al., 1978]{ButlerNewmanTalbot:1978} {Butler}, D.~M., {Newman}, M.~J., and {Talbot}, R.~J. (1978). \newblock {Interstellar Cloud Material - Contribution to Planetary Atmospheres}. \newblock {\em Science}, 201:522--525. \bibitem[{Caballero-Lopez} et~al., 2004]{CaballeroMcCracken:2004} {Caballero-Lopez}, R.~A., {Moraal}, H., {McCracken}, K.~G., and {McDonald}, F.~B. (2004). \newblock {The Heliospheric Magnetic Field from 850 to 2000 AD Inferred from $^{10}$Be Records}. \newblock {\em \jgr (Space Physics)}, 109:12102. \bibitem[{Christl} et~al., 2004]{Christl:2004} {Christl}, M., {Mangini}, A., {Holzk{\"a}mper}, S., and {Sp{\"o}tl}, C. (2004). \newblock {Evidence for a link between the flux of galactic cosmic rays and Earth's climate during the past 200,000 years}. \newblock { \em \jatp}, 66:313--322. \bibitem[{Decker} et~al., 2005]{Deckeretal:2005} {Decker}, R.~B., {Krimigis}, S.~M., {Roelof}, E.~C., {Hill}, M.~E., {Armstrong}, T.~P., {Gloeckler}, G., {Hamilton}, D.~C., and {Lanzerotti}, L.~J. (2005). \newblock {Voyager 1 in the Foreshock, Termination Shock, and Heliosheath}. \newblock {\em Science}, 309:2020--2024. \bibitem[{Fahr}, 1968]{Fahr:1968} {Fahr}, H.~J. (1968). \newblock {On the Influence of Neutral Interstellar Matter on the Upper Atmosphere}. \newblock {\em \apss}, 2:474. \bibitem[{Fahr}, 2004]{Fahr:2004} {Fahr}, H.-J. (2004). \newblock {Global Structure of the Heliosphere and Interaction with the Local Interstellar Medium: Three Decades of Growing Knowledge}. \\ \newblock {\em \adsr}, 34:3--13. \bibitem[Fahr et~al., 2006]{Fahretal:2006jos} Fahr, Hans~J., Fichtner, Horst, Scherer, Klaus, and Stawicki, Olaf (2006). \newblock {Variable Terrestrial Particle Environments During the Galactic Orbit of the Sun}. \newblock In {\em {Solar Journey: The Significance of Our Galactic Environment for the Heliosphere and Earth, {in press}}}, page {000}. Springer. \bibitem[Florinski and Zank, 2006]{FlorinskiZank:2006jos} Florinski, Vladimir and Zank, Gary~P. (2006). \newblock {The Galactic Cosmic Ray Intensity in the Heliosphere in Response to Variable Interstellar Environments}. \newblock In {\em {Solar Journey: The Significance of Our Galactic Environment for the Heliosphere and Earth, {in press}}}, page {000}. Springer. \bibitem[{Frisch} and {York}, 1986]{FrischYork:1986} {Frisch}, P. and {York}, D.~G. (1986). \newblock Interstellar clouds near the {S}un. \newblock In {\em The {G}alaxy and the {S}olar {S}ystem}, pages 83--100. Eds. R.~Smoluchowski, J.~Bahcall, and M.~Matthews (University of Arizona Press). \bibitem[{Frisch}, 1981]{Frisch:1981} {Frisch}, P.~C. (1981). \newblock The Nearby Interstellar Medium. \newblock {\em \nat}, 293:377--379. \bibitem[{Frisch}, 1993]{Frisch:1993a} {Frisch}, P.~C. (1993). \newblock G-star Astropauses - {A} Test for Interstellar Pressure. \newblock {\em \apj}, 407:198--206. \bibitem[Frisch, 1997]{Frisch:1997} Frisch, P.~C. (1997). \newblock Journey of the {S}un. \newblock {\em http://xxx.lanl.gov/}, astroph/9705231. \bibitem[{Frisch}, 1998]{Frisch:1998} {Frisch}, P.~C. (1998). \newblock Interstellar Matter and the Boundary Conditions of the Heliosphere. \newblock {\em \ssr}, 86:107--126. \bibitem[Frisch, 2000]{Frisch:2000amsci} Frisch, P.~C. (2000). \newblock The {G}alactic Environment of the {S}un. \newblock {\em American Scientist}, 88:52--59. \bibitem[{Frisch}, 2005]{Frisch:2005L} {Frisch}, P.~C. (2005). \newblock {Tentative Identification of Interstellar Dust in the Magnetic Wall of the Heliosphere}. \newblock {\em \apjl}, 632:L143--L146. \bibitem[{Frisch} and {Slavin}, 2006]{FrischSlavin:2006jos} {Frisch}, P.~C. and {Slavin}, J.~D. (2006). \newblock {Short Term Variations in the Galactic Environment of the Sun}. \newblock In {\em {Solar Journey: The Significance of Our Galactic Environment for the Heliosphere and Earth, {in press}}}, page {000}. Springer. \bibitem[{Frisch} et~al., 1999]{Frischetal:1999} {Frisch}, P.~C., {Dorschner}, J.~M., {Geiss}, J., {Greenberg}, J.~M., {Gr\"un}, E., {Landgraf}, M., {Hoppe}, P., {Jones}, A.~P., {Kr{\"{a}}tschmer}, W., {Linde}, T.~J., {Morfill}, G.~E., {Reach}, W., {Slavin}, J.~D., {Svestka}, J., {Witt}, A.~N., and {Zank}, G.~P. (1999). \newblock {Dust in the Local Interstellar Wind}. \newblock {\em \apj}, 525:492--516. \bibitem[{Gosling}, 1996]{Gosling:1996} {Gosling}, J.~T. (1996). \newblock {Corotating and Transient Solar Wind Flows in Three Dimensions}. \newblock {\em \araa}, 34:35--74. \bibitem[{Gruen} and {Landgraf}, 2000]{GruenLandgraf:2000} {Gruen}, E. and {Landgraf}, M. (2000). \newblock {Collisional Consequences of Big Interstellar Grains}. \newblock {\em \jgr}, 105:10291--10298. \bibitem[{Gurnett} and {Kurth}, 2005]{Gurnett:2005} {Gurnett}, D.~A. and {Kurth}, W.~S. (2005). \newblock {Electron Plasma Oscillations Upstream of the Solar Wind Termination Shock}. \newblock {\em Science}, 309:2025--2027. \bibitem[{Halley}, 1724]{Halley:1694} {Halley}, E. (1724). \newblock {Some Considerations about the Cause of the Universal Deluge, Laid before the Royal Society, on the 12th of December 1694. By Dr. Edmond Halley, R. S. S.} \newblock {\em Philosophical Transactions Series I}, 33:118--123. \bibitem[{Heiles} and {Troland}, 2003]{HTI} {Heiles}, C. and {Troland}, T.~H. (2003). \newblock {The Millennium Arecibo 21 Centimeter Absorption-Line Survey. I. Techniques and Gaussian Fits}. \newblock {\em \apjs}, 145:329--354. \bibitem[{Herschel}, 1795]{Herschel:1795} {Herschel}, W. (1795). \newblock {On the Nature and Construction of the Sun and Fixed Stars. By William Herschel, LL.D. F. R. S.} \newblock {\em Philosophical Transactions Series I}, 85:46--72. \bibitem[{Herschel}, 1801]{Herschel:1801} {Herschel}, W. (1801). \newblock {Observations Tending to Investigate the Nature of the Sun, in Order to Find the Causes or Symptoms of Its Variable Emission of Light and Heat; With Remarks on the Use That May Possibly Be Drawn from Solar Observations}. \newblock {\em Philosophical Transactions Series I}, 91:265--318. \bibitem[{Holzer}, 1989]{Holzer:1989} {Holzer}, T.~E. (1989). \newblock Interaction between the Solar Wind and the Interstellar Medium. \newblock {\em \araa}, 27:199--234. \bibitem[{Jokipii}, 2004]{Jokipii:2004} {Jokipii}, J.~R. (2004). \newblock {Transport of Cosmic Rays in the Heliosphere}. \newblock In {\em AIP Conf. Proc. 719: Physics of the Outer Heliosphere}, pages 249--259. \bibitem[Kirkby and Carslaw, 2006]{KirkbyCarslaw:2006jos} Kirkby, Jasper and Carslaw, Kenneth~S. (2006). \newblock {Variations of Galactic Cosmic Rays and the Earth's Climate}. \newblock In {\em {Solar Journey: The Significance of Our Galactic Environment for the Heliosphere and Earth, {in press}}}, page {000}. Springer. \bibitem[{Lallement} et~al., 2005]{Lallementetal:2005} {Lallement}, R., {Qu{\' e}merais}, E., {Bertaux}, J.~L., {Ferron}, S., {Koutroumpa}, D., and {Pellinen}, R. (2005). \newblock {Deflection of the Interstellar Neutral Hydrogen Flow Across the Heliospheric Interface}. \newblock {\em Science}, 307:1447--1449. \bibitem[{Landgraf}, 2000]{Landgraf:2000} {Landgraf}, M. (2000). \newblock {Modeling the Motion and Distribution of Interstellar Dust inside the Heliosphere}. \newblock {\em \jgr}, 105:10303--10316. \bibitem[Landgraf, 2006]{Landgraf:2006jos} Landgraf, Markus (2006). \newblock {Variations of the Interstellar Dust Distribution in the Heliosphere}. \newblock In {\em {Solar Journey: The Significance of Our Galactic Environment for the Heliosphere and Earth, {in press}}}, page {000}. Springer. \bibitem[{Langley}, 1876]{Langley:1876} {Langley}, S.~P. (1876). \newblock {Measurement of the Direct effect of Sun-Spots on Terrestrial Climates}. \newblock {\em \mnras}, 37:5. \bibitem[Linde, 1998]{Linde:1998} Linde, T.~J. (1998). \newblock {\em A Three-Dimensional Adaptive Multifluid {MHD} Model of the Heliosphere}. \newblock PhD thesis, Univ. of Michigan, Ann Arbor. \\ \newblock {\emph{http://hpcc.engin.umich.edu/CFD/publications}}. \bibitem[{Lockwood} and {Webber}, 2005]{LockwoodWebber:2005} {Lockwood}, J.~A. and {Webber}, W.~R. (2005). \newblock {Intensities of Galactic Cosmic Rays of 1.5 GeV Rigidity versus Heliospheric Current Sheet Tilt}. \newblock {\em \jgr (Space Physics)}, 110:4102. \bibitem[{Lopate}, 2005]{Lopate:2005} {Lopate}, C. (2005). \newblock {Fifty Years of Ground Level Solar Particle Event Observations, In \emph{Solar Eruptions and Energetic Particles}}. \newblock {\em \jgr}. \bibitem[{M{\" o}bius} et~al., 2004]{Moebiusetal:2004} {M{\" o}bius}, E., {Bzowski}, M., {Chalov}, S., {Fahr}, H.-J., {Gloeckler}, G., {Izmodenov}, V., {Kallenbach}, R., {Lallement}, R., {McMullin}, D., {Noda}, H., {Oka}, M., {Pauluhn}, A., {Raymond}, J., {Ruci{\' n}ski}, D., {Skoug}, R., {Terasawa}, T., {Thompson}, W., {Vallerga}, J., {von Steiger}, R., and {Witte}, M. (2004). \newblock {Synopsis of the Interstellar He Parameters from Combined Neutral Gas, Pickup Ion and UV Scattering Observations}. \newblock {\em \aap}, 426:897--907. \bibitem[M\"obius et~al., 2006]{Moebiusetal:2006jos} M\"obius, Eberhard, Bzowski, Maciek, M\"uller, Hans-Reinhard, and Wurz, Peter (2006). \newblock {Effects in the Inner Heliosphere Caused by Changing Conditions in the Galactic Environment}. \newblock In {\em {Solar Journey: The Significance of Our Galactic Environment for the Heliosphere and Earth, {in press}}}, page {000}. Springer. \bibitem[{McCracken} et~al., 2004]{McCrackenetal:2004} {McCracken}, K.~G., {McDonald}, F.~B., {Beer}, J., {Raisbeck}, G., and {Yiou}, F. (2004). \newblock {A Phenomenological Study of the Long-Term Cosmic Ray Modulation, 850-1958 AD}. \newblock {\em \jgr (Space Physics)}, 109:12103. \bibitem[{McCrea}, 1975]{McCrea:1975} {McCrea}, W.~H. (1975). \newblock {Ice Ages and the Galaxy}. \newblock {\em \nat}, 255:607--609. \bibitem[{McKay} and {Thomas}, 1978]{McKayThomas:1978} {McKay}, C.~P. and {Thomas}, G.~E. (1978). \newblock {Consequences of a Past Encounter of the Earth with an Interstellar Cloud}. \newblock {\em \grl}, 5:215--218. \bibitem[{Meisel} et~al., 2002]{Meiseletal:2002} {Meisel}, D.~D., {Janches}, D., and {Mathews}, J.~D. (2002). \newblock {Extrasolar Micrometeors Radiating from the Vicinity of the Local Interstellar Bubble}. \newblock {\em \apj}, 567:323--341. \bibitem[{Parker}, 1996]{Parker:1996} {Parker}, E.~N. (1996). \newblock {Solar Variability and Terrestrial Climate.} \newblock In {\em The Sun and Beyond}, pages 117. \bibitem[Parker, 2006]{Parker:2006jos} Parker, Eugene~N. (2006). \newblock {Interstellar Conditions and Planetary Magnetosphere}. \newblock In {\em {Solar Journey: The Significance of Our Galactic Environment for the Heliosphere and Earth, {in press}}}, page {000}. Springer. \bibitem[Pogorelov and Zank, 2006]{PogorelovZank:2006jos} Pogorelov, Nikolai~V. and Zank, Gary~P. (2006). \newblock {The Influence of the Interstellar Magnetic Field on the Heliospheric Interface}. \newblock In {\em {Solar Journey: The Significance of Our Galactic Environment for the Heliosphere and Earth, {in press}}}, page {000}. Springer. \bibitem[{Ratkiewicz} et~al., 1998]{Ratkiewicz:1998} {Ratkiewicz}, R., {Barnes}, A., {Molvik}, G.~A., {Spreiter}, J.~R., {Stahara}, S.~S., {Vinokur}, M., and {Venkateswaran}, S. (1998). \newblock {Effect of Varying Strength and Orientation of Local Interstellar Magnetic Field on Configuration of Exterior Heliosphere}. \newblock {\em \aap}, 335:363--369. \bibitem[Richardson et~al., 1995]{Richardson:1995} Richardson, J.~D., Paularena, K.~I., Lazarus, A.~J., and Belcher, J.~W. (1995). \newblock {Radial Evolution of the Solar Wind from {IMP} 8 to {V}oyager 2}. \newblock {\em \grl}, 22:325--328. \bibitem[{Roble} and {Hays}, 1979]{RobleHaysII:1979} {Roble}, R.~G. and {Hays}, P.~B. (1979). \newblock {A Quasi-Static Model of Global Atmospheric Electricity. II - Electrical Coupling between the Upper and Lower Atmosphere}. \newblock {\em \jgr}, 84:7247--7256. \bibitem[{Scherer} and {Fahr}, 2003]{SchererFahr:2003} {Scherer}, K. and {Fahr}, H.~J. (2003). \newblock {Breathing of Heliospheric Structures Triggered by the Solar-cycle Activity}. \newblock {\em Annales Geophysicae}, 21:1303--1313. \bibitem[Shapley, 1921]{Shapley:1921} Shapley, H. (1921). \newblock Note on a Possible Factor in Changes of Geological Climate. \newblock {\em J. Geology}, 29. \bibitem[Shaviv, 2006]{Shaviv:2006jos} Shaviv, Nir~J. (2006). \newblock {Long-term Variations in the Galactic Environment of the Sun}. \newblock In {\em {Solar Journey: The Significance of Our Galactic Environment for the Heliosphere and Earth, {in press}}}, page {000}. Springer. \bibitem[{Simpson}, 2001]{Simpson:2001} {Simpson}, J.~A. (2001). \newblock {\em {The Cosmic Radiation}}, pages 117--152. \newblock Century of Space Science, Volume I (Kluwer Academic Publishers). \bibitem[{Smith} et~al., 2003]{Smithetal:2003} {Smith}, E.~J., {Marsden}, R.~G., {Balogh}, A., {Gloeckler}, G., {Geiss}, J., {McComas}, D.~J., {McKibben}, R.~B., {MacDowall}, R.~J., {Lanzerotti}, L.~J., {Krupp}, N., {Krueger}, H., and {Landgraf}, M. (2003). \newblock {The Sun and Heliosphere at Solar Maximum}. \newblock {\em Science}, 302:1165--1169. \bibitem[{Sonett}, 1992]{Sonett:1992} {Sonett}, C.~P. (1992). \newblock {A Supernova Shock Ensemble Model using Vostok $^{10}$Be Radioactivity}. \newblock {\em Radiocarbon}, 34:239--245. \bibitem[{Stone} et~al., 2005]{Stoneetal:2005} {Stone}, E.~C., {Cummings}, A.~C., {McDonald}, F.~B., {Heikkila}, B.~C., {Lal}, N., and {Webber}, W.~R. (2005). \newblock {Voyager 1 Explores the Termination Shock Region and the Heliosheath Beyond}. \newblock {\em Science}, 309:2017--2020. \bibitem[{Talbot} and {Newman}, 1977]{TalbotNewman:1977} {Talbot}, R.~J. and {Newman}, M.~J. (1977). \newblock {Encounters Between Stars and Dense Interstellar Clouds}. \newblock {\em \apjs}, 34:295--308. \bibitem[{Thomas}, 1978]{Thomas:1978} {Thomas}, G.~E. (1978). \newblock {The Interstellar Wind and Its Influence on the Interplanetary Environment}. \newblock {\em Annual Review of Earth and Planetary Sciences}, 6:173--204. \bibitem[{Vall{\'e}e}, 2005]{ValleeSA:2005} {Vall{\'e}e}, J.~P. (2005). \newblock {The Spiral Arms and Interarm Separation of the Milky Way: An Updated Statistical Study}. \newblock {\em \aj}, 130:569--575. \bibitem[{Wang} and {Richardson}, 2005]{WangRichardson:2005} {Wang}, C. and {Richardson}, J.~D. (2005). \newblock {Dynamic Processes in the Outer Heliosphere: Voyager Observations and Models}. \newblock {\em \adsr}, 35:2102--2105. \bibitem[{Whang}, 2004]{Whang:2004} {Whang}, Y.~C. (2004). \newblock {Solar Cycle Variation of the Termination Shock}. \newblock In {\em AIP Conf. Proc. 719: Physics of the Outer Heliosphere}, pages 22--27. \bibitem[{Willis}, 1978]{Willis:1978} {Willis}, D.~M. (1978). \newblock {Atmospheric Water Vapour of Extraterrestrial Origin - Possible Role in Sun-Weather Relationships}. \newblock {\em \jatp}, 40:513--528. \bibitem[{Witte}, 2004]{Witte:2004} {Witte}, M. (2004). \newblock {Kinetic Parameters of Interstellar Neutral Helium. Review of Results obtained During One Solar Cycle with the Ulysses/GAS-Instrument}. \newblock {\em \aap}, 426:835--844. \bibitem[{Wood} et~al., 2005]{Woodetal:2005} {Wood}, B.~E., {Redfield}, S., {Linsky}, J.~L., {M{\"u}ller}, H.-R., and {Zank}, G.~P. (2005). \newblock {Stellar Ly{$\alpha$} Emission Lines in the Hubble Space Telescope Archive }. \newblock {\em \apjs}, 159:118--140. \bibitem[Yeghikyan and Fahr, 2006]{YeghikyanFahr:2006jos} Yeghikyan, Ararat and Fahr, Hans~J. (2006). \newblock {Accretion of Interstellar Material into the Heliosphere and onto Earth}. \bibitem[{Zank} and {Frisch}, 1999]{ZankFrisch:1999} {Zank}, G.~P. and {Frisch}, P.~C. (1999). \newblock Consequences of a Change in the {G}alactic environment of the {S}un. \newblock {\em \apj}, 518:965--973. \bibitem[{Zank} and {M{\"u}ller}, 2003]{ZankMueller:2003} {Zank}, G.~P. and {M{\"u}ller}, H.-R. (2003). \newblock {The Dynamical Heliosphere}. \newblock {\em \jgr (Space Physics)}, 108:7--1. \bibitem[Zank et~al., 2006]{Zanketal:2006jos} Zank, Gary~P., M\"uller, Hans-R., Florinski, Vladimir, and Frisch, Priscilla~C. (2006). \newblock {Heliospheric Variation in Response to Changing Interstellar Environments}. \newblock In {\em {Solar Journey: The Significance of Our Galactic Environment for the Heliosphere and Earth, {in press}}}, page {000}. Springer. \end{chapthebibliography} \emph{Submitted 15 November 2005; accepted 17 November 2005}
Title: Two-Current-Sheet Reconnection Model of Interdependent Flare and Coronal Mass Ejection
Abstract: Time-dependent resistive magnetohydrodynamic simulations are carried out to study a flux rope eruption caused by magnetic reconnection with implication in coexistent flare-CME (coronal mass ejection) events. An early result obtained in a recent analysis of double catastrophe of a flux rope system is used as the initial condition, in which an isolated flux rope coexists with two current sheets: a vertical one below and a transverse one above the flux rope. The flux rope erupts when reconnection takes place in the current sheets, and the flux rope dynamics depends on the reconnection sequence in the two current sheets. Three cases are discussed: reconnection occurs (1) simultaneously in the two current sheets, (2) first in the transverse one and then in the vertical, and (3) in an order opposite to case 2. Such a two-current-sheet reconnection exhibits characteristics of both magnetic breakout for CME initiation and standard flare model. We argue that both breakout-like and tether-cutting reconnections may be important for CME eruptions and associated surface activities.
https://export.arxiv.org/pdf/astro-ph/0601231
\title{TWO-CURRENT-SHEET RECONNECTION MODEL OF INTERDEPENDENT FLARE AND CORONAL MASS EJECTION} \author{Y. Z. ZHANG\altaffilmark{1}, J. X. WANG\altaffilmark{1} AND Y. Q. HU\altaffilmark{2} } \altaffiltext{1}{National Astronomical Observatories, Chinese Academy of Sciences, Beijing 100012, China} \altaffiltext{2}{School of Earth and Space Sciences, University of Science and Technology of China, Hefei 230026, China} \keywords{Sun: corona $-$ Sun: coronal mass ejections (CMEs) $-$ Sun: flares $-$ Sun: magnetic fields} \section{INTRODUCTION} A number of coronal mass ejections (CMEs) showed structures consistent with the ejection of a magnetic flux rope as it has been reported by Chen et al. (1997), Wood et al. (1999) and Dere et al. (1999). Therefore, magnetic flux ropes have been presumed to be typical structures in the solar corona, and their eruptions might be closely related to solar flares and CMEs (Forbes, 2000; Low, 2001). A lot of studies, both analytical and numerical, tried to explain such eruptive phenomena (Anzer, 1978; Priest, 1988; Forbes \& Isenberg, 1991; Isenberg et al., 1993; Mikic \& Linker, 1994; Forbes \& Priest, 1995; Low, 1996; Wu et al., 1997; Antiochos et al., 1999; Chen \& Shibata, 2000; Hu \& Liu, 2000; Lin \& Forbes, 2000; Amari et al., 2000; Lin et al., 2001, Cheng, et al., 2005, T\"{o}r\"{o}k \& Kliem, 2005). Most of them were associated with a bipolar magnetic configuration and assumed that reconnection in the current sheet below the flux rope triggers the eruption by the so-called tether cutting of the field lines. However, observations always show complicated magnetic configuration and global coupling of different flux systems (see an example described by Wang et al. (2005) and a statistic analysis by Zhou et al. (2005)). Two types of models are popular in the investigation of solar eruptive phenomena: the standard flare model and the magnetic breakout model. The standard flare model for magnetic explosions in eruptive flares was first proposed by Sturrock (1966), and advanced by a lot of latter studies ( Hirayama, 1974; Heyvaerts et al., 1977; Sturrock et al., 1984; Shibata et al., 1995; Tsuneta, 1997; Shibata, 1999; Chen \& Shibata, 2000; Moore et al., 2001). Recently, Chen \& Shibata (2000) proposed an emerging flux trigger mechanism for CMEs, in which reconnection in the current sheet below the rope leads to an eruption of the CME and a cusp-shaped solar flare. All of these studies showed that a cusp structure and a two ribbon flare occur in the lower corona, and that the reconnection is tether-cutting at the internal current sheet. Another type of models is the breakout model (Antiochos et al.,1999) that involves multipolar topology and requires external magnetic reconnection to occur on the top of the sheared arcade. In their model the background field has a spherically symmetric quadrupolar configuration, rather than a simple bipolar one. Many observations have shown that CMEs and flares are often two aspects of the same eruptive event. In a recent study (Zhang, et al. 2005) we found that a double catastrophe exists for an isolated flux rope embedded in a quadrupolar background field. After the first catastrophe, the flux rope levitates in the solar corona and two current sheets coexist with the rope, a transverse one above and a vertical one below the rope. As a product of interaction between the central and overlying arcades, the transverse current sheet represents the large-scale nature of the flux system. On the other hand, the vertical current sheet is limited to the interior of the central arcade and comes from a local interaction between small-scale bipoles. The coexistence of the two current sheets differentiates the present magnetic configuration with either the configuration of magnetic breakout model, or that of standard flare model. In the absence of reconnection, the flux rope may levitate in the corona in equilibrium. The resulting magnetic configuration provides a pre-eruption magnetic topology for a potential CME and its associated surface magnetic activity, and meets the requirements of magnetic breakout and standard flare models. Once reconnection sets in across one of the two current sheets or both, an eruption of the flux rope is inevitable, which is presumably responsible for concurrence of CMEs and flares. To explore this possibility, we take one of the force-free field solutions obtained by Zhang et al. (2005) as the initial state, which is located right after the first catastrophic point, introduce resistive dissipation in the current sheets, and examine the dynamic evolution of the flux rope system. The numerical results will show both breakout and tether-cutting. We describe the time-dependent, resistive magnetohydrodynamic (MHD) equations and the solution procedures in section 2. We discuss the evolution of the flux rope system in section 3, and conclude our work in section 4. \section{BASIC EQUATIONS AND SOLUTION PROCEDURES} We use time-dependent resistive MHD simulations to study the dynamic evolution of a flux rope system in the presence of resistance. For 2.5-dimensional (2.5-D) MHD problems in spherical coordinates (r,$\theta,\varphi$), one may introduce a magnetic flux function $\psi(t,r,\theta)$ related to the magnetic field by $$ {\mathbf B} = \bigtriangledown\times \left ( \frac{\psi}{r\sin\theta}\hat{\varphi} \right ) + {\mathbf B}_\varphi, \ \ \ \ {\mathbf B}_\varphi = B_\varphi \hat{\varphi} , \eqno (1) $$ where $B_\varphi$ is the azimuthal component of the magnetic field. Then the 2.5-D resistive MHD equations are cast in the following form $$ \frac{\partial\rho}{\partial t} + \bigtriangledown\cdot(\rho {\mathbf v})=0, \eqno (2) $$ $$ \frac{\partial{\mathbf v}}{\partial t} + {\mathbf v}\cdot\bigtriangledown {\mathbf v} + \frac{1}{\rho}\bigtriangledown p + \frac{1}{\mu\rho} [L\psi\bigtriangledown\psi + {\mathbf B}_\varphi\times(\bigtriangledown \times{\mathbf B}_\varphi )] $$ $$ +\frac{1}{\mu\rho r\sin\theta}\bigtriangledown\psi \cdot(\bigtriangledown\times{\mathbf B}_\varphi )\hat{\varphi} + \frac{GM_{\odot}}{r^2}\hat{r}=0, \eqno (3) $$ $$ \frac{\partial\psi}{\partial t}+ {\mathbf v}\cdot\bigtriangledown\psi - {1 \over \mu}\eta r^{2} \sin^{2}\theta L \psi = 0, \eqno (4) $$ $$ \frac{\partial B_{\varphi}}{\partial t} + r\sin\theta\bigtriangledown \cdot \left ( \frac{B_\varphi{\mathbf v}}{r\sin\theta} \right ) + \left [ \bigtriangledown \psi\times \bigtriangledown \left ( \frac{v_\varphi}{r\sin\theta} \right ) \right ]_\varphi-\frac{1}{r sin\theta}\bigtriangledown\eta \cdot \bigtriangledown(\mu r \sin\theta B_{\varphi}) $$ $$ -{1\over \mu}\eta r \sin \theta L(r B_{\varphi} \sin \theta)= 0, \eqno (5) $$ $$ \frac{\partial T}{\partial t}+{\mathbf v}\cdot\bigtriangledown T + (\gamma-1)T\bigtriangledown\cdot{\mathbf v}-\frac{\gamma-1}{\rho}\eta {\mathbf j}^2 = 0, \eqno (6) $$ where $$ L\psi\equiv\frac{1}{r^{2}\sin^{2}\theta} \left ( \frac{\partial^{2}\psi}{\partial r^{2}}+\frac{1}{r^{2}}\frac{\partial^{2}\psi}{\partial\theta^{2}}- \frac{\cot\theta}{r^{2}}\frac{\partial\psi}{\partial\theta} \right ) , \eqno (7) $$ $$ {\mathbf j} = {1\over \mu}\bigtriangledown \times {\mathbf B} = -{1\over \mu}r\sin\theta L\psi \hat{\varphi} +{1\over \mu}\bigtriangledown \times (B_\varphi \hat{\varphi}), \eqno (8) $$ $\rho$ is the density, ${\mathbf v}$ is the flow velocity, $\mu$ is the vacuum magnetic permeability, $G$ is the gravitational constant, $M_\odot$ is the mass of the Sun, $T$ is the temperature, $\gamma$ (= 1.05) is the polytropic index, $\eta$ is the resistivity, and ${\mathbf j}$ is the current density. The computational domain is taken to be $1 \leq r \leq 30$ in the unit of $R_\odot$ ($R_\odot$ is the solar radius), $0\leq\theta\leq \pi/2$, discretized into $130 \times 90$ grid points. The grid spacing increases according to a geometrical series of common ratio 1.03 from 0.02 at the base ($r$ = 1) to 0.86 at the top ($r$ = 30), whereas a uniform mesh is adopted in the $\theta$-direction. The multistep implicit scheme (Hu 1989) is used to solve equations (2)-(6). As for the boundary conditions, we use appropriate symmetrical conditions at the pole and equator, and calculate the quantities at the top in terms of equivalent extrapolation except for $B_\varphi$ and $\psi$. The magnetic field is potential above the transverse current sheet that is below the top boundary. Therefore, $B_\varphi$ is set to be zero and $\psi$ is calculated from $j_\varphi$ = $-r\sin\theta L\psi$ = 0 at the top (see Hu et al., 2003; Hu, 2004; Zhang et al., 2005). The initial corona is assumed to be isothermal and static with $T=T_0=2\times 10^6$ K and $\rho = \rho_0 = 1.67\times 10^{-13}$ kg$\cdot$m$^{-3}$ at the coronal base, where $T_0$ and $\rho_0$ are taken to be the units for temperature and density, respectively. Taking a characteristic value of 0.01 for $\beta$, the ratio of gas pressure to magnetic pressure, leads to a characteristic value of $\psi_0$ = $(2\mu \rho_0 R T_0 R_\odot^4 / \beta )^{1/2}$ = 5.69$\times 10^{14}$ Wb, taken to be the unit of $\psi$. Other units of interest are $B_0$ = $\psi_0/R_\odot^2$ = 1.18$\times 10^{-3}$ T for field strength, $v_A$ = $B_0/(\mu\rho_0)^{1/2}$ = 2570 km$\cdot$s$^{-1}$ for velocity, $\tau_A$ = $R_\cdot /v_A$ = 271 s for time, and $j_0 = B_0/(\mu R_\odot )$ = 1.35$\times 10^{-6}$ A$\cdot$m$^{-2}$ for electric current density. We choose a force-free field solution as the initial magnetic field. This solution was obtained by Zhang et al. (2005) right after the first catastrophic point, characterized by an isolated flux rope levitating in the corona and accompanied by two current sheets, a transverse one above and a vertical one below the rope. The annular magnetic flux per radian is 0.6 in the unit of $\psi_0$, and the axial magnetic flux is 0.0416 in the unit of $\psi_0$ for the flux rope, and both of them are conserved during subsequent dynamic evolutions of the flux rope system. The magnetic energy of the initial field is 1.71, which is still larger than the energy of the associated partially open field, 1.662, by 2.9\% (see Zhang, et al.,2005). The excess energy is obviously in favor of high-speed CMEs. The initial field chosen above is in equilibrium in the ideal MHD regime, but will certainly evolve into a dynamic state once reconnection sets in across the current sheets. The temporal evolution of the whole system depends on how reconnection occurs in the two current sheets. Three cases will be treated, labelled A, B, and C hereinafter, and they differ in the sequence of reconnection. Reconnection starts simultaneously in the two current sheets in case A, first in the transverse current sheet and later on in the vertical one in case B, and in the opposite order in case C. To control the sequence of reconnection, we introduce a critical current density for each current sheet, denoted by $j_t$ for the transverse current sheet and $j_v$ for the vertical one. When the current density nearby the transverse current sheet exceeds $j_t$ or that nearby the vertical current sheet exceeds $j_v$, the resistivity of $\eta$ is set to be 0.01, and $\eta$ is set to be 0 elsewhere. Consequently, we may simply set $j_t$ larger than the initial peak current density in the transverse current sheet to delay reconnection or smaller than the initial peak current density to start reconnection across the sheet. Notice that a larger value of $j_t$ just causes a delay of reconnection, rather than prohibits it. As a mater of fact, the current density in the transverse current sheet grows with time during the rope eruption, so it may eventually exceed $j_t$ somewhere, leading to a delayed onset of reconnection in the sheet. The same is the case for the vertical current sheet. Such an expedient measure is somewhat artificial but satisfies our purpose. Through tentative calculations, we find that the initial peak current density is 5.3 in the transverse current sheet and 22.1 in the vertical current sheet. Consequently, we choose ($j_t$, $j_v$) = (5, 20) for case A, (5, 40) for case B, and (10, 20) for case C. \section{SIMULATION RESULTS} As mentioned in the previous section, we intend to discuss three cases, a simultaneous reconnection in the transverse and vertical current sheets for case A, a first reconnection in the transverse current sheet followed by a second in the vertical current sheet for case B, and a first reconnection in the vertical current sheet followed by a second in the transverse current sheet for case C. In each case, we use the height of the rope axis relative to the solar surface, $h_a$, to mark the position of the flux rope. For the initial state, we have $h_a$ = 1.70. In case A, reconnection occurs simultaneously in the transverse and vertical current sheets. Figures 1a-1c show the magnetic configuration at three separate times, along with the temperature distribution in color. Figure 1a corresponds to the initial state, and resistive dissipation is switched on in both current sheets at $t$ = 0. Since then, high temperature appears in the current sheet regions because of reconnection, and the flux rope erupts upward, as shown in Figures 1b and 1c. The rope is immediately accelerated without an initial slow rising phase as shown in Figure 2 (solid), and it gains its maximum eruption speed of 595 km$\cdot$s$^{-1}$ at about $t$ = 5 $\tau_A$, when $h_a$ reaches 2.31 (Figure 1c). Meanwhile, a cusp-shaped structure with high temperature is clearly seen in Figure 1c, a typical feature of flares. Also, a high temperature structure appears in the corona right above the cusp structure at 1.5 in height. \placefigure{fig1} \placefigure{fig2} In case B, reconnection occurs first in the transverse current sheet, and then with the growth of the current in the vertical current sheet, reconnection follows over there. Figures 3a-3c show the magnetic configuration and temperature distribution at several separate times. At $t$ = 1 $\tau_A$ when reconnection is initiated in the transverse current sheet, the temperature along the sheet rises. As shown by dashed line in Figure 2, the flux rope's speed increases with time very slowly until reconnection sets in across the vertical current sheet at about $t$ = 7 $\tau_A$ (Figure 3b). Then the flux rope undergoes a slight deceleration of short duration (about 1 $\tau_A$), followed by a quick acceleration. The rope gains its maximum speed of 670 km$\cdot$s$^{-1}$ at about $t$ = 12 $\tau_A$, when $h_a$ reaches 2.34 (Figure 3c). Similarly, a cusp-shaped structure with high temperature and a coronal high temperature structure also appear in this case. \placefigure{fig3} In case C, reconnection occurs first in the vertical current sheet, and then with the growth of the current in the transverse current sheet, reconnection follows over there. Figures 4a-4c show the magnetic configuration and temperature distribution at several separate times. At $t$ = 1 $\tau_A$ when reconnection occurs only in the vertical current sheet, the temperature along the sheet rises, as shown in Figure 4a. It can be seen from the dash-dotted profile in Figure 2 that the flux rope is accelerated before that time, slightly decelerated afterwards about 2 $\tau_A$ in duration, and then accelerated again with a much larger acceleration. The flux rope gains a maximum speed of 568 km$\cdot$s$^{-1}$ at about $t$ = 10 $\tau_A$, when $h_a$ reaches 3.0 (Figure 4c). This case differs from case B in that the cusp-shaped structure is formed much earlier: it becomes clear as early as $t$ = 4.7 $\tau_A$ (Figure 4b). And at that time the reconnection initiates in the transverse current sheet. \placefigure{fig4} In summary, magnetic reconnection causes an eruption of the flux rope and the formation of a cusp-shaped structure of high temperature in all three cases. The former is presumably a manifestation of CMEs whereas the latter characterizes a two-ribbon flare. The reconnection sequence plays a critical role in the motion of the erupting flux rope and the formation of the cusp-shaped structure. The reconnection in the transverse current sheet is apt to produce a gradual acceleration of the flux rope but a higher peak speed and has little bearing on the formation of the cusp-shaped structure. On the other hand, the reconnection in the vertical current sheet is directly responsible for the formation of the cusp-shaped structure and leads to an immediate acceleration of the flux rope. It is interesting to note that a short term deceleration occurs before the rapid acceleration caused by reconnection across the vertical current sheet, as seen in cases B and C. Presently we do not know exactly why the flux rope has such a behavior. A possible reason might be that the magnetic pressure decreases right beneath the flux rope when reconnection starts in the vertical current sheet. High resolution observations at both optical and radio bands show indications that flux systems shrink first during the impulsive phases of flares, and then explode later in the main phases of flares (Ji et al., 2004, Li \& Gan, 2005). This seems to be consistent with the simulation results of reconnection occurring in the vertical current sheet in cases B and C. More careful work needs to be done in order to judge whether this is a common behavior of flux rope dynamics in the flare impulsive phase. Incidentally, since we have not considered the background solar wind, the flux rope's speed decreases after they obtain a peak speed in all three cases. \section{Concluding Remarks} Using time-dependent resistive MHD simulations, we find solutions associated with an isolated coronal flux rope embedded in a quadrupolar background field and accompanied by a transverse current sheet above and a vertical current sheet below the rope. Reconnection may occur in the current sheets either simultaneously or one after another. The present model agrees with the breakout model (Antiochos, 1999; Lynch, et al. 2004) if reconnection is initiated in the transversal current sheet, and it returns to the standard flare model (Chen \& Shibata, 2000) if reconnection is initiated in the vertical current sheet. Nevertheless, we argue that both breakout-like external reconnections and tether-cutting internal reconnections are essential to the magnetic eruption in general. Williams et al. (2005) showed observational evidence for the presence of both tether-cutting and breakout in eruptive events. Our simulations just combine the two models together, which is probably more relevant to observations that many eruptive events occur in background fields of quadrupolar magnetic configuration (Sterling \& Moore, 2004; Sterling \& Moore, 2004; Gary \& Moore, 2004). The present magnetic configuration and the dynamical evolution shed new light on understanding the relationship between CMEs and flares, which is a topic with great interest and hot debates. More and more investigations prefer a closer and rather intrinsic association between CMEs and surface activities (see Zhang et al.a,b; Zhou et al. 2003). Zhang et al. (2001a) reported that the kinematic evolution of CMEs can be described in a three-phase scenario: the initiation phase, the impulsive acceleration phase, and the propagation phase. Furthermore, they found that following the initiation phase, the CME displays an impulsive acceleration phase, which starts almost simultaneously with the flare onset time. After the acceleration phase the CME undergoes a propagation phase. And Zhang et al. (2001b) found a halo CME that moved slowly in the initial phase, and was later on accelerated and erupted. This is consistent with our case B, in which reconnection starts first in the transversal current sheet, leading to a slow upward motion of the CME, and subsequently, because of reconnection onset in the vertical current sheet, the CME acceleration is quickened until it reaches the maximum speed. In other words, the breakout first occurs and the tether-cutting follows. However, this is just one possibility, the other two cases we work out would appear in different circumstances. Zhou et al.(2003) gave a statistic result that $59\%$ of the selected 197 halo CMEs initiate earlier than the flare onset and $41\%$ are preceded by flare onsets. The latter samples may relate to our case C. Furthermore, Zhang et al. (2001a) also found one CME that did not show an initiation phase, but was immediately accelerated to the maximum speed. This example is very similar to our case A in which reconnection occurs simultaneously in the two current sheets. Another point is worthy of mentioning as to the effect of the reconnection sequence on the maximum speed of CMEs. The flux rope, identified as the CME here, has the largest speed when reconnection starts first in the transverse current sheet. On the other hand, the maximum speed is the lowest when reconnection starts first in the vertical current sheet. This implies that the reconnection sequence may affect the maximum speed of CMEs. \acknowledgments The authors are greatly indebted to the anonymous referee for helpful comments and valuable suggestions on the manuscript. One of the authors (YZZ) thanks J.Y. Ding for kind assistance in coding and P.F. Chen for helpful discussions. The work is supported by the National Natural Science Foundation of China (10233050, 40274049) and the National Key Basic Science Foundation (TG2000078404). \clearpage \clearpage
Title: Radio Linear and Circular Polarization from M81*
Abstract: We present results from archival Very Large Array (VLA) data and new VLA observations to investigate the long term behavior of the circular polarization of M81*, the nuclear radio source in the nearby galaxy M81. We also used the Berkeley-Illinois-Maryland Association (BIMA) array to observe M81* at 86 and 230 GHz. M81* is unpolarized in the linear sense at a frequency as high as 86 GHz and shows variable circular polarization at a frequency as high as 15 GHz. The spectrum of the fractional circular polarization is inverted in most of our observations. The sign of circular polarization is constant over frequency and time. The absence of linear polarization sets a lower limit to the accretion rate of $10^{-7} M_\odot y^{-1}$. The polarization properties are strikingly similar to the properties of Sgr A*, the central radio source in the Milky Way. This supports the hypothesis that M81* is a scaled up version of Sgr A*. On the other hand, the broad band total intensity spectrum declines towards milimeter wavelengths which differs from previous observations of M81* and also from Sgr A*.
https://export.arxiv.org/pdf/astro-ph/0601474
\newcommand\degd{\ifmmode^{\circ}\!\!\!.\,\else$^{\circ}\!\!\!.\,$\fi} \newcommand{\rdm}{{\rm\ rad\ m^{-2}}} \title{Radio Linear and Circular Polarization from M81*} \author{Andreas Brunthaler\inst{1,2}, Geoffrey C. Bower\inst{3}, Heino Falcke\inst{4,5} } \institute{Joint Institute for VLBI in Europe, Postbus 2, 7990 AA Dwingeloo, The Netherlands \and Max-Planck-Institut f\"ur Radioastronomie, Auf dem H\"ugel 69, 53121 Bonn, Germany \and Radio Astronomy Laboratory, University of California, Berkeley, CA 94720, USA \and ASTRON, Postbus 2, 7990 AA Dwingeloo, The Netherlands \and Department of Astrophysics, Radboud Universiteit Nijmegen, Postbus 9010, 6500 GL Nijmegen, The Netherlands} \offprints{brunthal@mpifr-bonn.mpg.de} \date{Received 30 December 2005 / Accepted 19 January 2006} \abstract{We present results from archival Very Large Array (VLA) data and new VLA observations to investigate the long term behavior of the circular polarization of M81*, the nuclear radio source in the nearby galaxy M81. We also used the Berkeley-Illinois-Maryland Association (BIMA) array to observe M81* at 86 and 230 GHz. M81* is unpolarized in the linear sense at a frequency as high as 86 GHz and shows variable circular polarization at a frequency as high as 15 GHz. The spectrum of the fractional circular polarization is inverted in most of our observations. The sign of circular polarization is constant over frequency and time. The absence of linear polarization sets a lower limit to the accretion rate of $10^{-7} M_\odot y^{-1}$. The polarization properties are strikingly similar to the properties of Sgr A*, the central radio source in the Milky Way. This supports the hypothesis that M81* is a scaled up version of Sgr A*. On the other hand, the broad band total intensity spectrum declines towards milimeter wavelengths which differs from previous observations of M81* and also from Sgr A*. \keywords{galaxies: active, galaxies: individual: Messier Number: M81, polarization } } \section{Introduction} The nearby spiral galaxy M81 (NGC\,3031) shares many properties with the Milky Way. It is similar in type, size and mass and it also contains a nuclear radio source, M81*, that is most likely associated with a supermassive black hole. M81* has been studied extensively using Very Long Baseline Interferometry (VLBI) in the past. \citeN{BietenholzBartelRupen2000} resolved M81* into a stationary core with a one sided jet. Multi-wavelength (\citeNP{HoFilippenkoSargent1996}) and sub-millimeter observations (\citeNP{ReuterLesch1996}) showed many similarities between M81* and Sgr A*, the central radio source in our Milky Way (\citeNP{MeliaFalcke2001}). A jet model of Sgr A* has been applied to M81*, where it can reproduce the radio flux density and the size of the radio core by changing the accretion rate (\citeNP{Falcke1996}). The sizes of both radio sources show a $\sim 1/\nu$ dependency on the frequency (e.g. \citeNP{BietenholzBartelRupen2004} for M81*, and \citeNP{BowerFalckeHerrnstein2004} and \citeNP{ShenLoLiang2005} for Sgr A*). M81* is an apparent transitional object between Sgr A* and high luminosity AGN. As the brightest of the nearby low luminosity AGN (LLAGN), it is 5 orders of magnitude brighter than Sgr A* at radio wavelengths. M81* is substantially underluminous at X-ray wavelengths ($L\sim 10^{-5} L_{edd}$), yet not as much as Sgr A* ($L\sim 10^{-10} L_{edd}$). Still, it is the faintest LLAGN we can study. Furthermore, the polarization properties of M81* and Sgr A* are very similar. Sgr A* shows circular polarization in absence of linear polarization (\citeNP{BowerFalckeBacker1999}, \citeNP{BowerBackerZhao1999}, \citeyearNP{BowerWrightBacker1999}) and we detected the same behaviour in M81* (\citeNP{BrunthalerBowerFalcke2001}.) The polarization properties of Sgr A* and M81* are in contrast to the properties of most radio jets in active galactic nuclei where linear polarization often exceeds circular polarization by a large factor (e.g.~\citeNP{WardleHomanOjha1998}; \citeNP{RaynerNorrisSault2000}). The absence of linear polarized emission in Sgr A* can be explained as a consequence of the accretion flow. \citeN{BowerFalckeSault2002} investigated the long term behavior of the circular polarization in Sgr A* from archival VLA data and showed that {\it i)} the circular polarization is variable on timescales of days to months, {\it ii)} the sign of the circular polarization stayed constant over the entire time range of almost 20 years, and {\it iii)} the average spectrum of circular polarization is inverted. After the discovery of circular polarization in M81* we used the VLA to investigate the variability of the circular polarization on short timescales. We used additional archival VLA data to investigate the long term behavior of the circular polarization of M81*. \section{Observations} \subsection{Archival VLA Data} M81* and the supernova SN1993J in M81 were observed many times with the Very Long Baseline Array (VLBA) and the phased Very Large Array (VLA) over the last decade (e.g. \citeNP{BartelBietenholzRupen1994}; \citeNP{BietenholzBartelRupen2000}). The observations were typically 16 hours in duration, were made at different frequencies and involved rapid switching between M81*, SN1993J and scans roughly every hour on the extragalactic background source 0954+658 for calibration purposes. We used the VLA data from the observations on 5 November 1993 ($\nu=$ 8.4 and 15 GHz), 16 December 1993 (8.4 and 15 GHz), 29 January 1994 (4.8 and 8.4 GHz), 21 April 1994 (4.8 GHz), 29 August 1994 (4.8 and 8.4 GHz), 31 October 1994 (8.4 GHz), 23 December 1994 (8.4 GHz), and 7 April 1996 (8.4 GHz). The VLA data had 50 MHz of bandwidth in two sidebands in right (RCP) and left (LCP) circular polarization modes. Data reduction was performed with the Astronomical Image Processing System (AIPS). 3C\,48 was used as primary amplitude calibrator. Then amplitude and phase self-calibration was performed on 0954+658. This forces 0954+658 to have zero circular polarization. The amplitude calibration was transfered to M81* and SN1993J before we performed phase self-calibration on M81* and SN1993J. Finally, we mapped all three sources in Stokes I and V. Flux densities were determined by fitting an elliptical Gaussian component to the sources. \subsection{New VLA observations} In addition to the archival VLA data, we used the VLA to observe M81* on 02 March 2001 and 9 dates between 15 June and 30 August 2001. During the latter period, the minimum and maximum separations between observing dates were 2 and 15 days, respectively. The VLA was in B configuration for the first observation and C array for the remaining observations. Observations were made at 5.0, 8.4, 15 and 22 GHz with 50 MHz of bandwidth in two sidebands in RCP and LCP modes. Each observation was between two and four hours in duration. Observing and analysis was performed with AIPS and followed the procedures outlined in \citeN{BrunthalerBowerFalcke2001}. The extragalactic point source J1044+719 was used as a phase, amplitude and polarization calibrator as well as reference pointing source. The extragalactic point source J1053+704 was used to check for polarization calibration errors . 3C 286 was used as an absolute amplitude calibration source. Finally, we mapped all three sources in Stokes I, U, Q, and V. Flux densities were determined by fitting an elliptical Gaussian component to the sources. We also used the VLA to observe M81* on 09 August 2003 at 15 GHz, 22 GHz, and 43 GHz in polarimetric mode. The VLA was in A configuration with a resolution of about 50 milliarcseconds at 43 GHz. The total integration time on M81* was 8, 48, and 52 minutes at 15, 22 and 43 GHz respectively and spread over a time range of 11 hours. Phase and amplitude calibration were performed using the nearby compact source, J1056+714. Phase self-calibration was also performed on M81* to eliminate the effects of atmospheric decorrelation. Polarization leakage calibration was performed using simultaneous full track observations of J1056+714 and J1048+701. \subsection{BIMA observations} We used the Berkeley-Illinois-Maryland Association (BIMA) array to observe M81* at 3\,mm wavelength (\citeNP{WelchThorntonPlambeck1996}). Observations were made in the multiplexed polarimetric mode described in \citeN{BowerWrightBacker1999}. The receivers were tuned to sky frequencies of 82.8 GHz (lower sideband) and 86.2 GHz (upper sideband). The compact source 0954+658 was used for phase calibration. Leakage calibration was determined from observations of 3C 279 on 11 November 2003. Observations of M81* were made on 6 dates in September and October 2003 (Table~\ref{High}). BIMA observations at 230 GHz on M81* were also performed in November 2003 (Table~\ref{High}). These data were also obtained in polarimetric mode with similar observing parameters to the 3\,mm observations. Each observation was a 8 hour track. \subsection{Error Analysis} The Stokes parameter V is measured as the difference between the left- and right-handed parallel polarization correlated visibilities. Errors in circular polarization measurements with the VLA have numerous origins: thermal noise, gain errors, beam squint, second-order leakage corrections, unknown calibrator polarization, background noise and radio frequency interference. The errors caused by amplitude calibration errors, beam squint, and polarization leakage scale with the source strength, and therefore the fractional circular polarization is a more relevant indicator for the detection of circular polarization. A detailed discussion of these errors is given in \citeN{BowerFalckeBacker1999} and \citeN{BowerFalckeSault2002}. We calculated the systematic errors based on the model for the VLA for circular polarization from \citeN{BowerFalckeSault2002}. For M81*, SN1993J, and J1053+704 the errors on the fractional circular polarization in Tables~\ref{all-c} -- \ref{all-u} are separated into statistical and systematic terms, while for the calibrator sources 0954+658 and J1044+719 only the statistical error is given. The calibrator sources do not have a systematic error, since their circular polarization was assumed to be zero during the calibration. \section{Results} \subsection{Circular and Linear Polarization} The results for the archival data at 4.8, 8.4, and 15 GHz are shown in Tables~\ref{all-c}, \ref{all-x}, and \ref{all-u} respectively. We consider circular polarization as detected if the measured flux density exceeds the combined statistical and systematic errors (added in quadrature) by a factor of three. 0954+658 showed no circular polarization as expected. M81* showed circular polarization in all observations except one (16 December 1993 at 15 GHz). SN1993J showed circular polarization only in one epoch (29 August 1994 at 4.8 GHz) and no circular polarization in all other observations. The upper limits on the circular polarization of SN1993J are not very meaningful at 15 GHz due to the low flux density and unexpected high noise. The results for the new observations at 4.8, 8.4, and 15 GHz are shown in Tables~\ref{all-c}, \ref{all-x}, and \ref{all-u} respectively. The 22 GHz data and the high frequency observations on 9 August 2003 gave no useful limits on the circular polarization, mainly because the short integration time and higher systematic errors. At 4.8. 8.4, and 15 GHz, 1044+719 showed no circular polarization as expected. M81* showed circular polarization in three epochs at 4.8 GHz, all except one epoch at 8.4 GHz and four epochs at 15 GHz. The check source 1053+704 showed only circular polarization in two epochs at 15 GHz. The three {\it detections} of circular polarization in the check sources SN1993J and 1053+704 are caused by either remaining amplitude calibration errors or by a small level of circular polarization in the sources. Since both sources show no circular polarization at 8.4 GHz, the frequency band with the highest sensitivity, it is most likely that the {\it detected} circular polarization in SN1993J and 1053+704 comes from residual amplitude calibration errors. The fractional circular polarization in M81* at 15 GHz is higher by a factor of 2 and 4 than in 1053+704 in the observations on 15 June 2001 and 4 July 2001 respectively. In these two cases, the measured circular polarization of M81* is probably only partly caused by the amplitude calibration errors. In the BIMA observations at 86 and 230 GHz neither linear nor circular polarization is detected for M81* in any individual epoch. Mean linear polarization at 86 GHz is $1.2$ mJy, or 1.6\% ($3\sigma$). Mean circular polarization at 86 GHz is $2.8 \pm 0.4$ mJy, or $3.9 \pm 0.5$\%. Although this is formally a detection, it is not clear whether systematic errors are significant. Due to the low flux density, limits on the polarized flux density are not significant at 230 GHz. The 8.4 GHz data seems to be the most reliable data since M81* showed circular polarization in all epochs except one while the check sources SN1993J and 1053+704 never showed circular polarization. The light-curve of total intensity and fractional circular polarization is shown in Fig.~\ref{x-light}. The fractional circular polarization shows significant variability on timescales of a few weeks which is not correlated with the variability in the total intensity. Between 4 August 2001 and 16 August 2001, the fractional circular polarization at 8.4 GHz dropped from 0.78\% to less than 0.35\% while the total intensity showed no significant change. However it is remarkable that, despite the strong variability, the sign of the circular polarization is always positive. At the other two frequencies, the sign is also positive when circular polarization is detected in M81*. The spectrum of the fractional circular polarization is inverted ($\alpha > 0$ for $m_{c} \propto \nu^\alpha$) with values between 0.4 and 2 in most epochs when it was detected at more than one frequency. The mean spectral index between 4.8 and 8.4 GHz is 0.51, while the mean spectral index between 8.4 and 15 GHz is 1.52. Only the observation on 18 August 2001 shows a steep spectrum of $\alpha$=-0.77 between 4.8 and 8.4 GHz.. Linear polarization was not detected in the VLA observations on 9 August 2003 and the BIMA observations at 86 GHz (Fig.~\ref{lp}). The archival data was not searched for linear polarization. \subsection{Total intensity} The high frequency VLA data taken on 9 August 2003 give a simultaneous total intensity spectrum of M81* that shows the flux density decreasing from 15 to 43 GHz with a spectral index of $\sim -0.6$. In the BIMA observations M81* is clearly detected at 86 GHz in total intensity and is strongly variable on a time scale $\sim 10$ days. The mean total intensity is 71 mJy. Due to low sensitivity, M81* is only marginally detected at 230 GHz in total intensity. The mean flux density for M81* from this observation is $31 \pm 4 \pm 15$ mJy, where the first error is the statistical error and the second error is the systematic error due to decorrelation. Atmospheric decorrelation may be serious and these results may significantly underestimate the total flux density of M81* (Table~\ref{High}). Fig.~\ref{bb-spec} shows a non simultaneous broad-band spectrum of M81*. The 15 -- 43 GHz data points are from the VLA observations on 9 August 2003. The 4.8 and 8.4 GHz data points are from the observation on 19 July 2001, where the 15 GHz flux was comparable to the 15 GHz flux in the 9 August 2003 observation. The 86 GHz and 230 GHz data points are from the BIMA observations on 7 September 2003 and November 2003 respectively. Also shown are the maximal and minimal measured values at each frequency in our observations. Although the spectrum is not simultaneous it is clear that the spectrum declines towards higher frequencies. M81* underwent a flare in total intensity during the new VLA observations between June and August 2001. The peak was reached at 8.4 GHz before 15 June 2001, while the flux density continued to rise at 4.8 until 4 July 2001. At 15 GHz, the peak was reached in 30 June 2001. Fig.~\ref{flare-spec} shows the spectral indices between 4.8 and 8.4 GHz, and bewteen 8.4 and 15 GHz during this flare. The spectrum at the lower frequencies shows a smooth transition from an inverted ($\alpha\sim$ +0.7) to a steep ($\alpha\sim$~-~0.4) spectrum. At higher frequencies, the spectral index does not follow a trend and is scattered between -0.4 and +0.5. The fast change in spectral index between 4.8 and 8.4 GHz could be caused by a drop in the turnover frequency of a syncrotron self-absorpted jet from above 8.4 to below 4.8 GHz and should be accompanied by a fast expansion of the jet. This behaviour is known in other active galactic nuclei (e.g. III~Zw~2: \citeNP{BrunthalerFalckeBower2000}, \citeNP{BrunthalerFalckeBower2005}). The scatter of the spectral index between 8.4 and 15 GHz could be caused by multiple sub-flares that occur at 15 GHz. \section{Discussion} The origin of circular polarization in AGN is still not known. Several mechanisms have been proposed in the literature. Interstellar propagation effects predict a very steep spectrum (\citeNP{MacquartMelrose2000}) which is not consistent with our observations. One possible mechanism could be Faraday conversion (\citeNP{Pacholczyk1977}; \citeNP{JonesODell1977}) of linear polarization to circular polarization caused by the lowest energy relativistic electrons. \citeN{BowerFalckeBacker1999} proposed a simple model for Sgr~A* in which low-energy electrons reduce linear polarization through Faraday de-polarization and convert linear polarization into circular polarization. Faraday conversion can also affect the spectral properties of circular polarization and may lead to a variety of spectral indices, including inverted spectra (\citeNP{JonesODell1977}). In inhomogeneous sources, conversion can produce relatively high fractional circular polarization (\citeNP{Jones1988}). Gyro-synchrotron emission, can also lead to high circular polarization with an inverted spectrum and low linear polarization (\citeNP{Ramaty1969}). However, this mechanism is to some degree related and also requires that M81* and Sgr A* both contain a rather large number of low-energy electrons. Faraday conversion is also favored by \citeN{BeckertFalcke2002} and \citeN{RuszkowskiBegelman2002}. The long-term stability of the sign of circular polarization suggests that the Faraday conversion is connected to fundamental properties of the source and the material which is responsible for the conversion. It requires uniformity in the magnetic pole and accretion conditions over the observed timescales. One possible scenario is described by \citeN{ensslin2003} where the sign of the circular polarization is connected to the sense of rotation of the central engine. In this scenario M81* is expected to rotate counter-clockwise. The size of the radio emission of M81* at 8.4 GHz is $\sim 0.45$ mas, or $\sim$ 1800 AU at 4 Mpc (\citeNP{BietenholzBartelRupen1996}). The fact that M81* is depolarized at at level of $<$ 0.1$\%$ at the same frequency requires that the depolarizing material is also present at a scale of $\sim$ 1800 AU. While the polarization properties of M81* and Sgr A* are strikingly similar the total intensity spectrum seems to be different. In Sgr A*, the radio flux density rises towards the sub-mm regime (e.g. \citeNP{ZylkaMezgerWard-Thompson1995}, \citeNP{FalckeMarkoff2000}). Our measurements indicate a different trend for M81*. Although the spectrum in Fig.~\ref{bb-spec} is not simultaneous and M81* shows strong variability our measured flux densities at 86 and 230 GHz are lower than the typical flux densities at centimeter wavelengths. This is different from the results in \citeN{ReuterLesch1996} who find an inverted spectrum up to $\sim$ 100 GHz. We can not tell whether we observed M81* in a phase with unusual low millimeter emission or the \citeN{ReuterLesch1996} observations were made during an outburst at millimeter wavelengths. A simultaneous monitoring project from centimeter to millimeter wavelengths would be needed to decide this question. Linear polarization is not detected for M81* at any wavelength longward of 3.6 mm. The presence of a jet at VLBI resolution, radio synchrotron emission, and circular polarization suggest that M81* is intrinsically linearly polarized but depolarized during propagation through a magnetized plasma. The case is similar to that of Sgr A*, which is detected in linear polarization only at wavelengths shortward of 3.6 mm. For Sgr A*, the detection of linear polarization at short wavelengths provides an upper limit to the rotation measure of a few times $10^6 {\rm\ rad\ m^{-2}}$. This provides an upper limit to the accretion rate of $\sim 10^{-7} M_\odot y^{-1}$. For M81*, we find a lower limit to the rotation measure of $\sim 10^4 \rdm$ for the case of beam depolarization and $\sim 4\times 10^5 \rdm$ for the case of bandwidth depolarization, under the assumption that the intrinsic source is polarized. The lesser value could originate in the dense interstellar medium but the larger value exceeds that seen anywhere in the ISM. ADAF and Bondi-Hoyle accretion models will depolarize the source unless the accretion rate falls below $10^{-9} M_\odot y^{-1}$ (\citeNP{QuataertGruzinov2000}). However, for radiatively inefficient accretion flows, the larger of the RM limits implies a lower limit to the accretion rate of $10^{-7} M_\odot y^{-1}$. The accretion rate necessary for the X-ray luminosity is $10^{-5} M_\odot y^{-1}$. Since bandwidth depolarization effects decrease as $\lambda^3$, measurement of the linear polarization at a wavelength of 0.8 mm would increase the accuracy of the accretion rate constraint by nearly two orders of magnitude. \section{Summary \& Conclusion} We have presented VLA observations of M81* from 1994 until 2002 that show that circular polarization is present at 4.8, 8.4, and 15 GHz in absence of linear polarization. The fractional circular polarization is variable on timescales of days and months and not correlated with the total flux density of the source. The sign of the circular polarization was, if detected, at all frequencies and times always positive. The polarization properties are strikingly similar to the properties of Sgr A*, the central radio source in the Milky Way. This supports the hypothesis that M81* is a scaled up version of Sgr A*. \citeN{AitkenGreavesChrysostomou2000} and \citeN{BowerWrightFalcke2003} detected linear polarization at 230 GHz and higher frequencies that also shows variability (\citeNP{BowerFalckeWright2005}; \citeNP{MarroneMoranZhao2006}). Given the similarity between M81* and Sgr A* we expect to see also linear polarization in M81* at higher frequencies. \begin{acknowledgements} This research was partially supported by the DFG Priority Programme 1177. The National Radio Astronomy Observatory is operated by Associated Universities, Inc., under a cooperative agreement with the National Science Foundation. \end{acknowledgements} \bibliography{brunthal_refs} \bibliographystyle{aa} \appendix{} \section{Tables} \begin{table*} \begin{center} \caption{Circularly polarized flux at 4.8 GHz for M81* and calibrators. The errors on the fractional circular polarization are separated into statistical and systematic terms for the target and check source. For the calibrator source only the statistical error is given.\label{all-c}} \begin{tabular}{rrrrrr} \hline\hline Date & Source & I & $P_{c}$ & rms & $m_{c}$\\ & & [mJy] &[mJy] &[mJy] & $[\%]$\\ \hline 28 Jan. 1994 & 0954+658 & 622.2 & $<$ 0.43 &0.11 & $<$ 0.07 $\pm$ 0.02\\ & M81* & 95.0 & 0.36 &0.06 & 0.38 $\pm$ 0.06 $\pm$ 0.04\\ & SN1993J & 77.1 & $<$ 0.17 &0.05 & $<$ 0.22 $\pm$ 0.06 $\pm$ 0.03\\ \hline 21 Apr. 1994 & 0954+658 & 534.7 & $<$ 0.21 &0.07 & $<$ 0.04 $\pm$ 0.01\\ & M81* & 114.1 & 0.52 &0.05 & 0.46 $\pm$ 0.04 $\pm$ 0.03\\ & SN1993J & 62.1 & $<$ 0.22 &0.04 & $<$ 0.35 $\pm$ 0.06 $\pm$ 0.03\\ \hline 29 Aug. 1994 & 0954+658 & 597.2 & $<$ 0.18 &0.06 & $<$ 0.03 $\pm$ 0.01\\ & M81* & 109.0 & 0.21 &0.04 & 0.19 $\pm$ 0.04 $\pm$ 0.03\\ & SN1993J & 54.6 & 0.18 &0.03 & 0.33 $\pm$ 0.05 $\pm$ 0.03\\ \hline \hline 02 Mar. 2001 & 1044+719 & 1568.2 & 0.11 & 0.11 & $<$ 0.01 $\pm$ 0.01\\ & M81* & 130.5 & 0.44 & 0.10 & 0.34 $\pm$ 0.08 $\pm$0.05\\ & 1053+704 & 388.2 & 0.03 & 0.07 & $<$ 0.01 $\pm$ 0.02 $\pm$0.06\\ \hline 15 Jun. 2001 & 1044+719 & 1710.9 & 0.33 & 0.35 & $<$ 0.02 $\pm$ 0.02 \\ & M81* & 136.7 & 0.45 & 0.11 & 0.33 $\pm$ 0.08 $\pm$0.05\\ & 1053+704 & 490.1 & -0.58 & 0.21 & $<$ 0.12 $\pm$ 0.04 $\pm$0.06\\ \hline 30 Jun. 2001 & 1044+719 & 1645.4 & 0.4 & 0.32 & $<$ 0.02 $\pm$ 0.02\\ & M81* & 140.9 & 0.3 & 0.12 & $<$ 0.21 $\pm$ 0.09 $\pm$0.05\\ & 1053+704 & 497.5 & -1.41 & 0.33 & $<$ 0.28 $\pm$ 0.07 $\pm$0.07\\ \hline 04 Jul. 2001 & 1044+719 & 1687.5 & 0.01 & 0.01 & $<$ 0.01 $\pm$ 0.01\\ & M81* & 142.8 & 0.12 & 0.05 & $<$ 0.09 $\pm$ 0.04 $\pm$0.04\\ & 1053+704 & 484 & 0.01 & 0.01 & $<$ 0.01 $\pm$ 0.01 $\pm$0.05\\ \hline 19 Jul. 2001 & 1044+719 & 1645.3 & 0.08 & 0.24 & $<$ 0.01 $\pm$ 0.05\\ & M81* & 134.9 & 0.46 & 0.19 & $<$ 0.34 $\pm$ 0.14 $\pm$0.07\\ & 1053+704 & 516.8 & 0.12 & 0.12 & $<$ 0.02 $\pm$ 0.02 $\pm$0.10\\ \hline 04 Aug. 2001 & 1044+719 & 1641.8 & 0.1 & 0.31 & $<$ 0.01 $\pm$ 0.02\\ & M81* & 128.6 & 0.07 & 0.07 & $<$ 0.05 $\pm$ 0.05 $\pm$0.07\\ & 1053+704 & 564.3 & 0.14 & 0.14 & $<$ 0.02 $\pm$ 0.02 $\pm$0.10\\ \hline 16 Aug. 2001 & 1044+719 & 1636.5 & 0.06 & 0.23 & $<$ 0.01 $\pm$0.01 \\ & M81* & 122.7 & 0.44 & 0.18 & $<$ 0.36 $\pm$ 0.15 $\pm$0.07\\ & 1053+704 & 524.3 & -0.38 & 0.23 & $<$ 0.07 $\pm$0.04 $\pm$0.10 \\ \hline 18 Aug. 2001 & 1044+719 & 1607.5 & 0.01 & 0.01 & $<$ 0.01 $\pm$ 0.01\\ & M81* & 118.2 & 0.74 & 0.16 & 0.63 $\pm$ 0.14 $\pm$ 0.07\\ & 1053+704 & 532.6 & 0.14 & 0.14 & $<$ 0.03 $\pm$ 0.03 $\pm$ 0.10\\ \hline 25 Aug. 2001 & 1044+719 & 1602.7 & 0.02 & 0.15 & $<$ 0.01 $\pm$ 0.01\\ & M81* & 112.7 & 0.37 & 0.16 & $<$ 0.33 $\pm$ 0.14 $\pm$ 0.07\\ & 1053+704 & 546 & 0.13 & 0.13 & $<$ 0.02 $\pm$ 0.02 $\pm$ 0.10\\ \hline 30 Aug. 2001 & 1044+719 & 1655.7 & 0.13 & 0.13 & $<$ 0.01 $\pm$ 0.01\\ & M81* & 113.9 & 0.23 & 0.11 & $<$ 0.2 $\pm$ 0.10 $\pm$ 0.07\\ & 1053+704 & 548 & -0.08 & 0.14 & $<$ 0.01 $\pm$ 0.03 $\pm$ 0.10\\ \hline \end{tabular} \end{center} \end{table*} \begin{table*} \begin{center} \caption{Circularly polarized flux at 8.4 GHz for M81* and calibrators. The errors on the fractional circular polarization are separated into statistical and systematic terms for the target and check source. For the calibrator source only the statistical error is given.\label{all-x}} \begin{tabular}{rrrrrr} \hline\hline Date & Source & I & $P_{c}$ & rms & $m_{c}$\\ & & [mJy] &[mJy] &[mJy] & $[\%]$\\ \hline 05 Nov. 1993 & 0954+658 & 663.5 & $<$ 0.27 &0.04& $<$ 0.04 $\pm$ 0.01\\ & M81* & 110.4 & 0.59 &0.03& 0.53 $\pm$ 0.03 $\pm$ 0.04\\ & SN1993J & 62.2 & $<$ 0.07 &0.02& $<$ 0.11 $\pm$ 0.03 $\pm$ 0.03\\ \hline 16 Dec. 1993 & 0954+658 & 655.5 & $<$ 0.12 &0.04& $<$ 0.02 $\pm$ 0.01\\ & M81* & 85.7 & 0.23 &0.03& 0.27 $\pm$ 0.04 $\pm$ 0.03\\ & SN1993J & 54.7 & $<$ 0.06 &0.02& $<$ 0.11 $\pm$ 0.04 $\pm$ 0.03\\ \hline 28 Jan. 1994 & 0954+658 & 600.6 & $<$ 0.24 &0.08& $<$ 0.04 $\pm$ 0.01 \\ & M81* & 111.2 & 0.79 &0.04& 0.71 $\pm$ 0.04 $\pm$ 0.04 \\ & SN1993J & 48.9 & $<$ 0.14 &0.03& $<$ 0.29 $\pm$ 0.06 $\pm$ 0.03\\ \hline 29 Aug. 1994 & 0954+658 & 648.1 & $<$ 0.11 &0.04& $<$ 0.02 $\pm$ 0.01\\ & M81* & 102.0 & 0.35 &0.03& 0.34 $\pm$ 0.03 $\pm$ 0.04\\ & SN1993J & 34.5 & $<$ 0.08 &0.03& $<$ 0.23 $\pm$ 0.09 $\pm$ 0.03\\ \hline 31 Oct. 1994 & 0954+658 & 670.9 & $<$ 0.12 &0.04& $<$ 0.02 $\pm$ 0.01\\ & M81* & 117.2 & 0.33 &0.03& 0.28 $\pm$ 0.03 $\pm$ 0.04\\ & SN1993J & 33.5 & $<$ 0.08 &0.03& $<$ 0.24 $\pm$ 0.09 $\pm$ 0.03\\ \hline 23 Dec. 1994 & 0954+658 & 693.7 & $<$ 0.43 &0.14& $<$ 0.06 $\pm$ 0.02\\ & M81* & 76.7 & 0.52 &0.09& 0.68 $\pm$ 0.12 $\pm$ 0.07\\ & SN1993J & 30.4 & $<$ 0.22 &0.07& $<$ 0.72 $\pm$ 0.23 $\pm$ 0.05\\ \hline 07 Apr. 1996 & 0954+658 & 786.3 & $<$ 0.11 &0.04& $<$ 0.01 $\pm$ 0.01\\ & M81* & 165.8 & 1.15 &0.03& 0.69 $\pm$ 0.02 $\pm$ 0.04\\ & SN1993J & 20.5 & $<$ 0.15 &0.05& $<$ 0.73 $\pm$ 0.24 $\pm$ 0.04\\ \hline \hline 02 Mar. 2001 & 1044+719 & 1478.7 & 0.13 & 0.13 & $<$ 0.01 $\pm$ 0.01\\ & M81* & 143.8 & 0.62& 0.07 & 0.43 $\pm$ 0.05 $\pm$ 0.05\\ & 1053+704 & 524.8 & 0.35 & 0.10 & $<$ 0.07 $\pm$ 0.02 $\pm$ 0.06\\ \hline 15 Jun. 2001 & 1044+719 & 2136.1 & 0.11 & 0.11 & $<$ 0.01 $\pm$ 0.01\\ & M81* & 201.7 & 1.08 & 0.10 & 0.53 $\pm$ 0.05 $\pm$ 0.05\\ & 1053+704 & 1006.1 & 0.17 & 0.17 & $<$ 0.02 $\pm$ 0.02 $\pm$ 0.06\\ \hline 30 Jun. 2001 & 1044+719 & 1488.6 & 0.01 & 0.01 & $<$ 0.01 $\pm$ 0.01\\ & M81* & 193.9 & 0.75 & 0.11 & 0.39 $\pm$ 0.06 $\pm$ 0.05\\ & 1053+704 & 763.0 & 0.34 & 0.19 & $<$ 0.04 $\pm$ 0.02 $\pm$ 0.07\\ \hline 04 Jul. 2001 & 1044+719 & 1525.8 & 0.57 & 0.57 & $<$ 0.04 $\pm$ 0.04\\ & M81* & 175.7 & 0.93 & 0.09 & 0.53 $\pm$ 0.05 $\pm$ 0.05\\ & 1053+704 & 753.3 & 0.13 & 0.13 & $<$ 0.02 $\pm$ 0.02 $\pm$ 0.06\\ \hline 19 Jul. 2001 & 1044+719 & 1474.8 & 0.10 & 0.10 & $<$ 0.01 $\pm$ 0.01\\ & M81* & 135.1 & 0.60 & 0.09 & 0.45 $\pm$ 0.07 $\pm$ 0.05\\ & 1053+704 & 769.2 & 0.27 & 0.13 & $<$ 0.04 $\pm$ 0.02 $\pm$ 0.07\\ \hline 04 Aug. 2001 & 1044+719 & 1485.4 & 0.12 & 0.12 & $<$ 0.01 $\pm$ 0.01\\ & M81* & 107.0 & 0.84 & 0.17 & 0.78 $\pm$ 0.16 $\pm$ 0.05\\ & 1053+704 & 785.1 & 0.57 & 0.26 & $<$ 0.07 $\pm$ 0.03 $\pm$ 0.07\\ \hline 16 Aug. 2001 & 1044+719 & 1469.1 & 0.27 & 0.27 & $<$ 0.02 $\pm$ 0.02\\ & M81* & 101.4 & 0.36 & 0.16 & $<$ 0.35 $\pm$ 0.16 $\pm$ 0.07\\ & 1053+704 & 758.3 & 1.20 & 0.27 & $<$ 0.16 $\pm$ 0.04 $\pm$ 0.10\\ \hline 18 Aug. 2001 & 1044+719 & 1476.8 & 0.41 & 0.65 & $<$ 0.03 $\pm$ 0.03\\ & M81* & 98.0 & 0.40 & 0.09 & 0.41 $\pm$ 0.09 $\pm$ 0.07\\ & 1053+704 & 775.2 & 0.41 & 0.30 & $<$ 0.05 $\pm$ 0.04 $\pm$ 0.10\\ \hline 25 Aug. 2001 & 1044+719 & 1406.9 & 0.11 & 0.11 & $<$ 0.01 $\pm$ 0.01\\ & M81* & 89.2 & 0.32 & 0.09 & 0.36 $\pm$ 0.10 $\pm$ 0.05\\ & 1053+704 & 746.0 & 1.30 & 0.24 & $<$ 0.17 $\pm$ 0.03 $\pm$ 0.07\\ \hline 30 Aug. 2001 & 1044+719 & 1520 & 0.17 & 0.17 & $<$ 0.01 $\pm$ 0.01\\ & M81* & 92.3 & 0.42 & 0.16 & 0.45 $\pm$ 0.17 $\pm$ 0.07\\ & 1053+704 & 789.2 & 0.30 & 0.30 & $<$ 0.04 $\pm$ 0.04 $\pm$ 0.10\\ \hline \end{tabular} \end{center} \end{table*} \begin{table*} \begin{center} \caption{Circularly polarized flux at 15 GHz for M81* and calibrators. The errors on the fractional circular polarization are separated into statistical and systematic terms for the target and check source. For the calibrator source only the statistical error is given.\label{all-u}} \begin{tabular}{rrrrrr} \hline\hline Date & Source & I & $P_{c}$ & rms & $m_{c}$\\ & & [mJy] &[mJy] &[mJy] & $[\%]$\\ \hline 05 Nov. 1993 & 0954+658 & 664.5 & $<$ 0.26 &0.09& $<$ 0.04 $\pm$ 0.01\\ & M81* & 108.2 & 1.14 &0.18& 1.05 $\pm$ 0.17 $\pm$ 0.05\\ & SN1993J & 42.3 & $<$ 4.05 &1.35& $<$ 9.57 $\pm$ 3.19 $\pm$ 0.03\\ \hline 16 Dec. 1993 & 0954+658 & 606.4 & $<$ 0.28 &0.09& $<$ 0.05 $\pm$ 0.01\\ & M81* & 87.2 & $<$ 0.06 &0.19& $<$ 0.07 $\pm$ 0.22 $\pm$ 0.05\\ & SN1993J & 39.0 & $<$ 9.0 &3.32& $<$ 23.0 $\pm$ 8.51 $\pm$ 0.03\\ \hline \hline 02 Mar. 2001 & 1044+719 & 999.8 & 0.03 & 0.12 & $<$ 0.01 $\pm$ 0.01\\ & M81* & 114.2 & 0.81 & 0.19 & 0.71 $\pm$ 0.17 $\pm$ 0.05\\ & 1053+704 & 485.2 & 0.25 & 0.18 & $<$ 0.05 $\pm$ 0.04 $\pm$ 0.07\\ \hline 15 Jun. 2001 & 1044+719 & 2275 & 0.36 & 0.36 & $<$ 0.02 $\pm$ 0.02\\ & M81* & 188.1 & 2.16 & 0.38 & 1.15 $\pm$ 0.20 $\pm$ 0.06\\ & 1053+704 & 1354.7 & 8.28 & 0.75 & 0.61 $\pm$ 0.06 $\pm$ 0.07\\ \hline 30 Jun. 2001 & 1044+719 & 1669.9 & 0.43 & 0.43 & $<$ 0.03 $\pm$ 0.03\\ & M81* & 257.4 & 3.14 & 0.4 & 1.22 $\pm$ 0.16 $\pm$ 0.07\\ & 1053+704 & 1051.5 & 2.61 & 0.66 & $<$ 0.25 $\pm$ 0.06 $\pm$ 0.09\\ \hline 04 Jul. 2001 & 1044+719 & 1955.8 & 1.67 & 1.67 & $<$ 0.09 $\pm$ 0.09 \\ & M81* & 221.2 & 3.66 & 0.46 & 1.66 $\pm$ 0.21 $\pm$ 0.06\\ & 1053+704 & 1162.1 & 4.9 & 0.88 & 0.42 $\pm$ 0.08 $\pm$ 0.07\\ \hline 19 Jul. 2001 & 1044+719 & 1663.3 & 0.58 & 0.58 & $<$ 0.03 $\pm$ 0.03\\ & M81* & 123.3 & 1.65 & 0.68 & $<$ 1.34 $\pm$ 0.55 $\pm$ 0.09\\ & 1053+704 & 1011.9 & 1.76 & 1.05 & $<$ 0.17 $\pm$ 0.10 $\pm$ 0.12\\ \hline 04 Aug. 2001 & 1044+719 & 1663 & 0.29 & 1.14 & $<$ 0.02 $\pm$ 0.07\\ & M81* & 113.4 & 0.36 & 0.36 & $<$ 0.31 $\pm$ 0.31 $\pm$ 0.09\\ & 1053+704 & 973.4 & 6.42 & 1.92 & $<$ 0.66 $\pm$ 0.20 $\pm$ 0.12\\ \hline 16 Aug. 2001 & 1044+719 & 1876 & 0.51 & 0.51 & $<$ 0.03 $\pm$ 0.03\\ & M81* & 107 & 0.38 & 0.42 & $<$ 0.35 $\pm$ 0.39 $\pm$ 0.09\\ & 1053+704 & 1043.9 & 3.42 & 1.37 & $<$ 0.33 $\pm$ 0.13 $\pm$ 0.12\\ \hline 18 Aug. 2001 & 1044+719 & 1496.6 & 0.46 & 0.98 & $<$ 0.03 $\pm$ 0.07\\ & M81* & 79.3 & 0.2 & 0.2 & $<$ 0.25 $\pm$ 0.25 $\pm$ 0.09\\ & 1053+704 & 830.4 & 2.44 & 0.95 & $<$ 0.29 $\pm$ 0.11 $\pm$ 0.12\\ \hline 25 Aug. 2001 & 1044+719 & 1378 & 0.49 & 0.49 & $<$ 0.04 $\pm$ 0.04\\ & M81* & 82.9 & 0.47 & 0.76 & $<$ 0.56 $\pm$ 0.56 $\pm$ 0.09\\ & 1053+704 & 756.7 & 0.82 & 1.65 & $<$ 0.11 $\pm$ 0.22 $\pm$ 0.12\\ \hline 30 Aug. 2001 & 1044+719 & 1846.3 & 0.76 & 0.76 & $<$ 0.04 $\pm$ 0.04\\ & M81* & 104.7 & 0.54 & 0.53 & $<$ 0.52 $\pm$ 0.52 $\pm$ 0.09\\ & 1053+704 & 988.9 & 0.41 & 0.41 & $<$ 0.04 $\pm$ 0.04 $\pm$ 0.12\\ \hline \end{tabular} \end{center} \end{table*} \begin{table*} \begin{center} \caption{Polarized and total flux density of M81* at high frequencies unsing the VLA (15, 22, and 43 GHz) and BIMA (83, 86, and 230 GHz).~\label{High}} \begin{tabular}{rccrrrrr} \hline\hline Date & Frequency & Sideband & I & Q & U & V & m$_p$\\ &[GHz]&&[mJy] & [mJy] &[mJy] &[mJy] &$[\%]$\\ \hline 09 Aug. 2003 & 15 & & 125.9 $\pm$ 1.5 & $<$ 1.4 & $<$ 1.4 & $<$ 1.4 & $<$ 1.0 \\ & 22 & & 118.2 $\pm$ 1.3 & $<$ 1.1 & $<$ 1.1 & $<$ 1.3 & $<$ 1.0 \\ & 43 & & 66.8 $\pm$ 2.0 & $<$ 3.3 & $<$ 3.3 & $<$ 4.2 & $<$ 4.9 \\ \hline \hline 07 Sep. 2003 & 83 & lsb & 44.0 $\pm$ 2.6 & -2.0 $\pm$ 2.6 & 5.6 $\pm$ 2.6 & 0.3 $\pm$ 2.6 & \\ & 86 & usb & 41.8 $\pm$ 2.6 & -1.8 $\pm$ 2.6 & -3.4 $\pm$ 2.6 & 3.2 $\pm$ 2.6 & \\ & & avg & 42.9 $\pm$ 1.8 & -1.8 $\pm$ 1.8 & 1.1 $\pm$ 1.8 & 1.8 $\pm$ 1.8 & 5.1 $\pm$ 4.2 \\ \hline 12 Sep. 2003 & 83 & lsb & 89.4 $\pm$ 1.8 & -1.3 $\pm$ 1.8 & -2.1 $\pm$ 1.8 & 5.7 $\pm$ 1.8 & \\ & 86 & usb & 92.4 $\pm$ 1.8 & -0.8 $\pm$ 1.8 & -1.1 $\pm$ 1.8 & 2.5 $\pm$ 1.8 & \\ & & avg & 90.9 $\pm$ 1.3 & -1.1 $\pm$ 1.3 & -1.6 $\pm$ 1.3 & 4.1 $\pm$ 1.3 & 2.1 $\pm$ 1.4 \\ \hline 21 Sep. 2003 & 83 & lsb & 86.4 $\pm$ 1.5 & -0.2 $\pm$ 1.5 & -0.7 $\pm$ 1.5 & 3.8 $\pm$ 1.5 & \\ & 86 & usb & 86.9 $\pm$ 1.5 & -1.3 $\pm$ 1.5 & -0.2 $\pm$ 1.5 & 3.4 $\pm$ 1.5 & \\ & & avg & 86.7 $\pm$ 1.1 & -0.2 $\pm$ 1.1 & -0.5 $\pm$ 1.1 & 3.6 $\pm$ 1.1 & 0.6 $\pm$ 1.3 \\ \hline 06 Oct. 2003 & 83 & lsb & 70.4 $\pm$ 2.0 & 3.5 $\pm$ 2.0 & 0.7 $\pm$ 2.0 & 0.2 $\pm$ 2.0 & \\ & 86 & usb & 72.7 $\pm$ 2.0 & -0.3 $\pm$ 2.0 & 0.3 $\pm$ 2.0 & 3.5 $\pm$ 2.0 & \\ & & avg & 71.6 $\pm$ 1.4 & 1.6 $\pm$ 1.4 & 0.5 $\pm$ 1.4 & 1.8 $\pm$ 1.4 & 2.3 $\pm$ 2.0 \\ \hline 09 Oct. 2003 & 83 & lsb & 46.1 $\pm$ 1.8 & 5.1 $\pm$ 1.8 & -2.1 $\pm$ 1.8 & 1.0 $\pm$ 1.8 & \\ & 86 & usb & 45.1 $\pm$ 1.8 & -3.2 $\pm$ 1.8 & 2.7 $\pm$ 1.8 & 1.5 $\pm$ 1.8 & \\ & & avg & 45.6 $\pm$ 1.3 & 1.0 $\pm$ 1.3 & 0.3 $\pm$ 1.3 & 1.2 $\pm$ 1.3 & 2.2 $\pm$ 2.9 \\ \hline 12 Oct. 2003 & 83 & lsb & 43.9 $\pm$ 1.5 & -2.4 $\pm$ 1.5 & -1.0 $\pm$ 1.5 & 3.8 $\pm$ 1.5 & \\ & 86 & usb & 34.5 $\pm$ 1.5 & 6.1 $\pm$ 1.5 & 4.9 $\pm$ 1.5 &-1.0 $\pm$ 1.5 & \\ & & avg & 39.2 $\pm$ 1.1 & 1.9 $\pm$ 1.1 & 2.0 $\pm$ 1.1 & 1.4 $\pm$ 1.1 & 7.0 $\pm$ 2.8 \\ \hline \hline 01 Nov. 2003 & 230 & lsb & 33.7 $\pm$ 6 & -6.6 $\pm$ 6 & 1.2 $\pm$ 6 & -4.5 $\pm$ 6 & \\ & & usb & 27.5 $\pm$ 6 & 2.7 $\pm$ 6 & -19.9 $\pm$ 6 & 7.3 $\pm$ 6 & \\ & & avg & 30.6 $\pm$ 4 & -2.0 $\pm$ 4 & -9.4 $\pm$ 4 & 1.4 $\pm$ 4 & 31.4 $\pm$ 13\\ \hline \end{tabular} \end{center} \end{table*}
Title: Theoretical foundations for on-ground tests of LISA PathFinder thermal diagnostics
Abstract: This paper reports on the methods and results of a theoretical analysis to design an insulator which must provide a thermally quiet environment to test on ground delicate temperature sensors and associated electronics. These will fly on board ESA's LISA PathFinder (LPF) mission as part of the thermal diagnostics subsystem of the LISA Test-flight Package (LTP). We evaluate the heat transfer function (in frequency domain) of a central body of good thermal conductivity surrounded by a layer of a very poorly conducting substrate. This is applied to assess the materials and dimensions necessary to meet temperature stability requirements in the metal core, where sensors will be implanted for test. The analysis is extended to evaluate the losses caused by heat leakage through connecting wires, linking the sensors with the electronics in a box outside the insulator. The results indicate that, in spite of the very demanding stability conditions, a sphere of outer diameter of the order one metre is sufficient.
https://export.arxiv.org/pdf/gr-qc/0601096
\jl{6} \title[Theoretical foundations for\ldots]{Theoretical foundations for on-ground tests of \textsl{LISA PathFinder} thermal diagnostics} \author{A Lobo$^{1,2}$\footnote[3]{To whom correspondence should be addressed.}, M Nofrarias$^2$, J Ramos-Castro$^3$ and J Sanju\'an$^2$} \address{$^1$ Institut de Ci\`encies de l'Espai, {\sl CSIC}} \address{$^2$ Institut d'Estudis Espacials de Catalunya ({\sl IEEC\/}), Edifici {\sl Nexus}, Gran Capit\`a~2--4, 08034 Barcelona, Spain} \address{$^3$ Departament d'Enginyeria Electr\`onica, {\sl UPC}, Campus Nord, Edif.\ C4, Jordi Girona 1--3, 08034 Barcelona, Spain \ead{lobo@ieec.fcr.es}} \date{\today} \pacs{04.80.Nn, 95.55.Ym, 04.30.Nk} \submitto{\CQG} \section{Introduction \label{sec.1}} \lisa Pathfinder (\lpf) is an \esa mission, whose main objective is to put to test critical parts of \lisa (Laser Interferometer Space Antenna), the first space borne gravitational wave (GW) observatory \cite{bender}. The science module on board \lpf is the \lisa Test-flight Package (\ltp)~\cite{lpfall}, which basically consists in two test masses in nominally perfect geodesic motion (free fall), and a laser metrology system, which reports on \emph{residual deviations} of the test masses' actual motion from the ideal free fall, to a given level of accuracy \cite{gerhar}. In order to ensure that the test masses are not deviated from their geodesic trajectories by external (non-gravitational) agents, a so called Gravitational Reference System (GRS) is used~\cite{rita}. This consists in position sensors for the masses which send signals to a set of micro-thrusters; the latter take care of correcting as necessary the spacecraft trajectory, so that at least one of the test masses remains centred relative to the spacecraft at all times. The combination of the GRS plus the actuators is known as \emph{drag-free} subsystem\footnote{ The term \emph{drag-free} dates back to the early days of space navigation, when it was used to name a trajectory correction system designed to compensate for the effect of atmospheric drag on satellites in low altitude orbits.}. The \emph{drag-free} is of course a central component of \lisa, and needs to be operated at extremely demanding levels of accuracy. The laser metrology system should then be sufficiently precise to measure relative test mass deviations. The overall level of noise acceptable for \lisa is defined in terms of rms acceleration spectral density, and has been set to \begin{equation} S_{a,{\rm LISA}}^{1/2}(\omega)\leq 3\!\times\!10^{-15}\,\left[ 1 + \left(\frac{\omega/2\pi}{3\ {\rm mHz}}\right)^{\!\!2}\right]\, {\rm m}\,{\rm s}^{-2}/\sqrt{\rm Hz} \label{eq.1} \end{equation} in the frequency range $\quad 10^{-4}\,{\rm Hz}\leq\omega/2\pi\leq 10^{-3}\,{\rm Hz}$. This is equivalent to $S_h^{1/2}$\,$\sim$\,4$\times$10$^{-21}$ Hz$^{-1/2}$, with the same frequency dependence. Because \lpf is a \emph{technological mission}, aimed to assess the feasibility of \lisa, its ultimate goal has been relaxed to~\cite{toplev} \begin{equation} S_{a,{\rm LPF}}^{1/2}(\omega)\leq 3\!\times\!10^{-14}\,\left[ 1 + \left(\frac{\omega/2\pi}{3\ {\rm mHz}}\right)^{\!\!2}\right]\, {\rm m}\,{\rm s}^{-2}/\sqrt{\rm Hz} \label{eq.2} \end{equation} in the frequency range $1\,{\rm mHz}\leq\omega/2\pi\leq 30\,{\rm mHz}$, i.e., one order of magnitude less demanding, both in noise amplitude and in frequency band. Equation (\ref{eq.2}) gives the \emph{global} noise budget. This is naturally made up of contributions from different perturbative agents, such as temperature and magnetic field fluctuations, GRS and interferometer noise, etc. As a general rule, a requirement on the magnitude of each of the various perturbing factors is set at a 10\,\% fraction of the total. In the case of temperature fluctuations, this is equivalent to \begin{equation} S_{T}^{1/2}(\omega)\leq 10^{-4}\,{\rm K}/\sqrt{\rm Hz}\ , \quad 1\,{\rm mHz}\leq \omega/2\pi \leq 30\,{\rm mHz} \label{eq.3} \end{equation} Because temperature stability is important, a decision has been taken to place high precision thermometers in several strategic spots across the \ltp ---as part of what is called \emph{Diagnostics Subsystem}~\cite{lobo} \footnote{ The Diagnostics Subsystem of the \ltp also includes magnetometric measurements and a charged particle flux detector.}. Such high precision temperature measurements will be useful to identify the fraction of the total system noise which is due to thermal fluctuations only, and this will in turn provide important debugging information to assess the performance of the \ltp. \subsection{Temperature measurements \label{sec.1-1}} If the temperature gauges are to be sensitive to fluctuations at the level given by (\ref{eq.3}) then clearly the entire measuring device should be less noisy, typically by a factor of~10. This means that such device, which includes both the sensors \emph{and} the associated electronics, can generate a maximum level of noise of \begin{equation} S_{T, {\rm sensor}}^{1/2}(\omega)\leq 10^{-5}\,{\rm K}/\sqrt{\rm Hz}\ , \quad 1\,{\rm mHz}\leq \omega/2\pi \leq 30\,{\rm mHz} \label{eq.4} \end{equation} Research work is currently being conducted at \ieec (Barcelona, Spain) to identify the appropriate sensors and design the better suited front end electronics. But the prototype system needs of course to be tested for compliance with equation~(\ref{eq.4}). Thus, in order to do a meaningful test, the system must be sufficiently thermally isolated that the observed fluctuations in the readout data can be attributed \emph{solely} to sensor noise, rather than to a combination of it with real ambient temperature fluctuations. This means temperature fluctiations in the thermomters' placements should again be at least one order of magnitude below the target sensitivity, equation~(\ref{eq.4}), or \begin{equation} S_{T, {\rm testbed}}^{1/2}(\omega)\leq 10^{-6}\,{\rm K}/\sqrt{\rm Hz}\ , \quad 1\,{\rm mHz}\leq \omega/2\pi \leq 30\,{\rm mHz} \label{eq.5} \end{equation} It turns out that 10$^{-6}\,{\rm K}/\sqrt{\rm Hz}$ is a truly demanding temperature stability, orders of magnitude beyond the capabilities of normal thermally regulated rooms. We thus need to design a specific thermal insulator to shield the sensors from ambient temperature fluctuations during the test process. In the ensuing pages we describe in detail the insulator design. It is extremely important to stress at this point that the performance of the insulator, i.e., its ability to screen out ambient temperature fluctuations, \emph{cannot be checked} experimentally, at least under working thermal conditions in the laboratory. This is because, by definition, the insulator is the tool to check the sensing instruments, \emph{not viceversa}: we need to rely on the results of theoretical argumentation to make a decission on which is the appropriate thermal insulator for our purposes. An experimental verification of the model is only thinkable under much more extreme conditions, where external temperature fluctuations are orders of magnitude higher than the ones which will be met during the test. \section{Thermal insulator design concept \label{sec.2}} The idea of the insulator design is displayed in figure~\ref{fig.1}: an interior metal core of good thermal conductivity is surrounded by a thick layer of a poorly conductive material. The inner block ensures thermal stability of the sensors attached to it, while the surrounding substrate efficiently shields it from external temperature fluctuations in the laboratory ambient. We propose a spherical shape for the sake of simplicity of the mathematical analysis, even though this will be eventually changed to cubic in the actual experimental device due to practical feasibility issues. \subsection{Mathematical model \label{sec.2.1}} The basic assumption of the mathematical analysis we shall present is that heat flows from the interior of the insulator to the air outside, and from the latter to the interior of the insulator, only by thermal \emph{conduction}. This is a very realistic hypothesis in the context of the experiment, as radiation mechanisms are certainly negligible and convection should not play any significant role, either, since the entire body is solid, and temperature fluctuations will be small at all times anyway. Let then $T({\bf x},t)$ be the temperature at time $t\/$ of a point positioned at vector {\bf x} relative to the centre of the sphere. $T({\bf x},t)$ thus satisfies Fourier's partial differential equation \cite{carslaw} \begin{equation} \rho c_{\rm p}\,\frac{\partial}{\partial t} T({\bf x},t) = \nabla\cdot\left[\kappa\nabla T({\bf x},t)\right] \label{eq.6} \end{equation} where $\rho$, $c_{\rm p}$ and $\kappa$ are the density, specific heat and thermal conductivity, respectively, of the substrate. We shall assume these are uniform values within each of the two materials making up the insulating body, with abrupt changes in the interface. We can thus represent them as discontinuous functions of the radial coordinate, as follows: \begin{equation} \rho, c_{\rm p}, \kappa({\bf x}) = \left\{\begin{array}{ll} \rho_1, c_{{\rm p}1}, \kappa_1 \quad & {\rm if}\ \ 0\leq r < a_1 \\ \rho_2, c_{{\rm p}2}, \kappa_2 \quad & {\rm if}\ \ a_1\leq r < a_2 \end{array}\right. \label{eq.7} \end{equation} with $r\/$\,$\equiv$\,$|{\bf x}|$. Initial and boundary conditions are the following: \begin{equation} T({\bf x},t=0) = 0\ ,\quad T(r=a_2,t) = T_0(\theta,\varphi;t) \label{eq.8} \end{equation} where $\theta\/$ and $\varphi\/$ are spherical angles which define positions on the sphere's surface. The boundary temperature can be expediently expressed as a multipole expansion: \begin{equation} T_0(\theta,\varphi;t) = \sum_{lm}\,b_{lm}(t)\,Y_{lm}(\theta,\varphi) \label{eq.9} \end{equation} where $Y_{lm}(\theta,\varphi)$ are spherical harmonics, and $b_{lm}(t)$ are boundary multipole temperature components. In practice, the boundary temperature will be \emph{randomly fluctuating}, therefore $b_{lm}(t)$ will be considered \emph{stochastic} functions of time. We shall also reasonably assume them to be \emph{stationary Gaussian} noise processes with known spectral densities, $S_{lm}(\omega)$. As shown in the appendix, the frequency analysis of this problem leads to a \emph{transfer function} expression of the temperature inside the body: \begin{equation} \tilde T({\bf x},\omega) = \sum_{lm}\,H_{lm}({\bf x},\omega)\,\tilde b_{lm}(\omega) \label{eq.10} \end{equation} where \emph{tildes} (\,$\tilde{}$\,) stand for Fourier transforms, e.g., \begin{equation} \tilde T({\bf x},\omega)\equiv\int_{-\infty}^\infty\, T({\bf x},t)\,e^{-i\omega t}\,dt \label{eq.11} \end{equation} etc. If we make the further assumption that different multipole temperature fluctuations at the boundary are \emph{uncorrelated}, i.e., \begin{equation} \langle\tilde b^*_{l'm'}(\omega)\,\tilde b_{lm}(\omega)\rangle = S_{lm}(\omega)\,\delta_{l'l}\,\delta_{m'm} \label{eq.12b} \end{equation} then the spectral density of fluctuations at any given point inside the insulating body is given by \begin{equation} S_T({\bf x},\omega) = \sum_{lm}\,\left|H_{lm}({\bf x},\omega)\right|^2\, S_{lm}(\omega) \label{eq.12} \end{equation} It is ultimately the spectral density $S_T({\bf x},\omega)$ which has to comply with the requirement expressed by equation~(\ref{eq.4}). Based on knowledge (by direct measurement) of ambient laboratory temperature fluctuations, equation~(\ref{eq.12}) will provide the guidelines, as regards materials and dimensions, for the actual design of a suitable insulator jig. \section{Homogeneous boundary conditions \label{sec.3}} Thermal conditions in the laboratory are rather \emph{homogeneous}. This means that the boundary temperature fluctuations will be in practice essentially independent of the angles $\theta\/$ and $\varphi$, i.e., \begin{equation} T_0(\theta,\varphi;t) = B(t) \label{eq.13} \end{equation} and consequently the generic expansion equation~(\ref{eq.9}) includes only the \emph{monopole} term, hence \begin{equation} b_{00}(t) = \sqrt{4\pi}\,B(t) \label{eq.14} \end{equation} The temperature $T({\bf x},\omega)$ in this case will only depend on radial depth, $r$, therefore, \begin{equation} \tilde T(r,\omega) = H(r,\omega)\,\tilde B(\omega) \label{eq.15} \end{equation} with $H(r,\omega)$\,$\equiv$\,$\sqrt{4\pi}\,H_{00}({\bf x},\omega)$. According to equation~(\ref{eq.a13}) of the Appendix, this is \begin{equation} \hspace*{-0.6 cm} H(r,\omega) = \left\{\begin{array}{ll} \xi_0(\omega)\,j_0(\gamma_1r)\ , & 0\leq r \leq a_1 \\[1 em] \eta_0(\omega)\,j_0(\gamma_2r) + \zeta_0(\omega)\,y_0(\gamma_2r)\ , & a_1\leq r \leq a_2 \end{array}\right. \label{eq.16} \end{equation} This is a low-pass filter transfer function ---even though the cumbersome frequency dependencies involved in the expressions above do not make it immediately obvious. A plot of the square modulus of $H(r,\omega)$ is shown in figure~\ref{fig.2} for $r\/$\,=\,0 (red curve). The figure also shows a low-pass filter of the first order with the same frequency cut-off, $|H_{\rm 1st\ order}(\omega)|^2$\,=\,$(1+\omega^2\tau^2)^{-1}$, for conceptual comparison (blue curve). The most salient feature emerging out of the plot is the stronger drop in $H(r,\omega)$ at the high frequency tails. The latter can be easily assessed in quantitative detail, and the result is \begin{equation} |H(0,\omega)|\sim\omega\tau\,e^{-\sqrt{\omega\tau}} \label{eq.19} \end{equation} where $\tau$ is the filter's time constant ---a complicated function of the insulator's physical and geometric properties, to be discussed below. As already mentioned in the Introduction section, to test the temperature sensors and electronics we need a very strong noise suppression factor in the \ltp frequency band. A look at figure~\ref{fig.2} readily shows that high damping factors require such frequency band to lie in the filter's tails. The thermal insulator should therefore be designed in such a way that its time constant $\tau\/$ be sufficiently large to ensure that the \ltp frequencies are high enough compared to 1/$\tau\/$. The exponential drop in the transfer function shown by equation~(\ref{eq.19}) makes the filter actually feasible with reasonable dimensions. \section{Numerical analysis \label{sec.4}} In this section we consider the application of the above formalism to obtain practically useful numbers for the actual implementation of a real insulator device which complies with the needs of our experiment. First of all, a selection of an \emph{aluminum} core surrounded by a layer of \emph{polyurethane} was made. Aluminum is a good heat conductor and is easy to work with in the laboratory; polyurethane is a good insulator and is also convenient to handle, as it can be foamed to any desired shape from canned liquid. Other alternatives are certainly possible, but this appears sufficiently good and we shall therefore only make reference to this specific one. The relevant physical properties of aluminium and polyurethane are specified in table~\ref{tab.1}. \begin{table}[h!] \begin{center} \begin{tabular}{lccc} & Density & Specific heat & Thermal conductivity \\ & $\rho$ (kg\,m$^{-3}$) & $c_{\rm p}$ (J\,kg$^{-1}$\,K$^{-1}$) & $\kappa$ (W\,m$^{-1}$\,K$^{-1}$) \\[1ex] \hline \\[-1.3ex] {\sf Aluminum} & 2700 & 900 & 250 \\ {\sf Polyurethane} & 35 & 1000 & 0.04 \end{tabular} \caption{Density, specific heat and thermal conductivity of aluminium and polyurethane. Units are given in the International System. \label{tab.1}} \end{center} \end{table} Figure \ref{fig.3} plots the \emph{amplitude damping coefficient} of the insulator block, $|H(r,\omega)|$, at the lower end of the \ltp frequency band, i.e., 1 mHz, and at the interface position, $r\/$\,=\,$a_1$. Each of the curves corresponds to a fixed value of the latter, and is represented as a function of the outer radius of the insulator. This choice is useful because the sensors are implanted for test on the surface of the aluminium core, and also because at higher frequencies thermal damping is stronger. So in practice the actual damping power of the device will be the one plotted, and better at the higher frequencies in the measuring bandwidth. The figure clearly shows that the assymptotic regime of equation~(\ref{eq.19}) is quite early established. The choice of dimensions for the insulating body must of course ensure that the minimum requirement, equation~(\ref{eq.5}) is met. For this, a primary consideration is the size of the ambient temperature fluctuations in the site where the experiment is made. Dedicated measurements in our laboratory showed that \begin{equation} S_{T, {\rm ambient}}^{1/2}(\omega)\sim 10^{-1}\,{\rm K}/\sqrt{\rm Hz}\ , \quad 1\,{\rm mHz}\leq \omega/2\pi \leq 30\,{\rm mHz} \label{eq.20} \end{equation} We therefore need to implement a device such that $|H(a_1,\omega)|$\,$\leq$\,10$^{-5}$ throughout the measuring bandwidth (MBW). Suitable dimensions can then be readily read off figure~\ref{fig.3}, and various alternatives are possible, as seen. Before making a decission, however, we need to make an additional estimate of the heat leakage down the electric wires which connect the temperature sensors with the elctronics, which lies of course outside the insulator. We come to this next. \subsection{Heat leakage through connecting wires} We use a simple model, consisting in assuming the connecting wires behave as straight metallic rods which connect the central aluminum core with the electronics, placed in the external laboratory ambient. Because the polyurethane provides a very stable insulation, we can neglect the lateral flux, hence only a unidimensional heat flow needs to be considered. For this, the following equation relates the heat flux to the temperature difference between the two wires' edges: \begin{equation} \dot{Q}(t) = \kappa_{\rm wire}\,\frac{\pi R_{\rm wire}^2}{\ell_{\rm wire}}\; [T(a_2,t)-T(a_1,t)] \label{eq.21} \end{equation} where $\kappa_{\rm wire}$ is the thermal conductivity of the wire, $R_{\rm wire}$ its transverse radius, and $\ell_{\rm wire}$ its length \emph{inside} the polyurethane layer. On the other hand, the heat flux results in temperature variations in the metal core, given by \begin{equation} \dot{Q}(t) = \rho_1c_{{\rm p}1}V_1\, \frac{\partial T}{\partial t}(a_1,t) \label{eq.22} \end{equation} where $V_1$\,=\,$4\pi a_1^3/3$ is the volume of the metal core. Equating the above expressions we find \begin{equation} \kappa_{\rm wire}\,\frac{\pi R_{\rm wire}^2}{\ell_{\rm wire}}\; [T(a_2,t)-T(a_1,t)] = \rho_1c_{{\rm p}1}V_1\, \frac{\partial T}{\partial t}(a_1,t) \label{eq.23} \end{equation} For fluctuating temperatures, we can now obtain the relationship between the spectral density at the aluminium core and the ambient, due to heat conduction along the wire: \begin{equation} S_{T,{\rm wire}}^{1/2}(a_1,\omega) = |H_{\rm wire}(\omega)|\, S_{T,{\rm ambient}}^{1/2}(\omega) \label{eq.24} \end{equation} where \begin{equation} |H_{\rm wire}(\omega)|\simeq\frac{\pi}{\omega}\, \frac{\kappa_{\rm wire}\,R_{\rm wire}^2} {\rho_1c_{{\rm p}1}V_1\,\ell_{\rm wire}} \label{eq.25} \end{equation} and where the approximation has been made that the temperature fluctuations at the inner end of the wire are much smaller than those at the outer end, due to the presence of the polyurethane layer. In practice, there will be several sensors for test inside the insulator. Under the hypothesis made that no lateral heat flux is relevant, the transfer function for a bundle of $N\/$ of wires is, at most, $N\/$ times that of a single wire. Thus, \begin{equation} |H_{N{\rm wires}}(\omega)| = \frac{3N}{\omega/2\pi}\, \frac{\kappa_{\rm wire}\,R_{\rm wire}^2} {8\pi\rho_1c_{{\rm p}1}a_1^3\,\ell_{\rm wire}} \label{eq.26} \end{equation} Let us consider numerical values in this expression. We use thin copper wires ($\kappa_{\rm Cu}$\,=\,401\,Wm$^{-1}$K$^{-1}$) of radius $R_{\rm wire}$\,=\,0.1\,mm, and assume some fiducial parameters for the size of the aluminium core, $a_1$, the wire length, $\ell_{\rm wire}$, the number of connecting wires, $N$, and the frequency, $\omega/2\pi$. The following obtains: \begin{equation} \hspace*{-1.2 cm} |H_{N{\rm wires}}(\omega)| = 1.1\times 10^{-5}\, \left(\frac{N}{30}\right)\! \left(\frac{a_1}{13\ {\rm cm}}\right)^{\!-3}\! \left(\frac{\ell_{\rm wire}}{25\ {\rm cm}}\right)^{\!-1}\! \left(\frac{\omega/2\pi}{1\ {\rm mHz}}\right)^{\!-1} \label{eq.27} \end{equation} This result indicates that, for laboratory fluctuations in the level of equation~(\ref{eq.20}), leakage through wiring causes fluctuations in the temperature sensors of about 10$^{-6}$\,K/$\sqrt{\rm Hz}$, equation~\eref{eq.24}, which is compliant with the requirement of stability of equation~\eref{eq.5}. The most sensitive parameter in the above expression is the size of the metal core, and this determines the need to make it somewhat large. The length of the wires has been taken to be 25~cm, but this does not necessarily mean we need $a_2$\,=\,38~cm (assuming the radius of the aluminum core is $a_1$\,=\,13~cm), because the wires can be partly wound inside the polyurethane layer to further protect the system against leakage. In fact, this wire lengthening is an easy way to improve attenuation. As regards frequency dependence, compliance is guaranteed in the entire MBW if it is at its lower end: indeed, not only $|H_{\rm wire}(\omega)|$ decreases as $\omega^{-1}$, also ambient noise fluctuations drop below 10$^{-1}$\,K/$\sqrt{\rm Hz}$ at higher frequencies. \section{Conclusions} Temperature fluctuation measurement is very demanding in the \ltp, and subsequently \lisa, as reflected by equation~\eref{eq.4}. Accordingly, very delicate sensor and associated electronics must be designed, and of course tested in ground before boarding. However, even the best laboratory conditions are orders of magnitude worse than the above requirement, so meaningful tests of the temperature sensing system cannot be tested without suitably screening the sensors from ambient temperature fluctuations. We have addressed how this can be accomplished by means of an insulating system consisting of a central metallic core surrounded by a thick layer of a very poorly conducting material. The latter provides good thermal insulation, while the central core, having a large thermal inertia, ensures stability of the sensors' environemnt. The choice of materials is flexible, so aluminium and polyurethane, which are easily available in the market, has been adopted. Thereafter, the dimensions need to be fixed. The appropriate sensors for the needs are temperature sensitive resistors, more specifically thermistors ---also known as NTCs. It appears that, because these sensors need to be wired to external electronics, heat leakage through such wires is an effect which needs to be quantitatively assessed to prevent losses. We have analysed this problem, and concluded that it strongly depends on the central metallic core size, and imposes that it be somewhat large. Laboratory ambient temperature fluctuations, determined by dedicated \emph{in situ} measurements, are of the order of 10$^{-1}$\,K/$\sqrt{\rm Hz}$ at 1~mHz, and dropping at higher frequencies within the MBW. The required stability conditions at the sensors, attached at the core's surface, thus need an attenuation factor of 10$^{-5}$, or better. Our analysis determines that a central aluminium core of 13~cm of radius, surrounded by a concentric layer of polyurethane 15--20~cm thick, comfortably provides the needed thermal screening which guarantees a meaningful test of the sensors' performance. The results of this paper are based on modelling. Because our aim is to produce a very stable thermal environment for the temperature sensors, we cannot check \emph{experimentally} the correctness of our conclusions. We must instead rely on the validity of the hypotheses made ---essentially that heat only flows by thermal conduction--- and on the underlying physical laws which govern heat conduction. Even though there is good reason to believe that both are sufficiently accurate, unexpected behaviour e.g. at the interface between the metal core and the insulator, may partly distort the results. Direct measurements with very large temperature gradients applied across the insulating device are envisaged, and will be reported elsewhere as an auxiliary independent test of the model. \ack We want to thank Albert Tom\`as, from {\sl NTE}, for discussions on the subject of this paper. Support for this work came from Project ESP2004-01647 of Plan Nacional del Espacio of the Spanish Ministry of Education and Science (MEC). MN acknowledges a grant from Generalitat de Catalunya, and JS a grant from MEC. \appendix \section{Thermal insulator frequency response functions \label{sec.a1}} Here we present some mathematical details of the solution to the Fourier problem, equations~(\ref{eq.6})-(\ref{eq.9}). We first of all Fourier transform equations~(\ref{eq.6}) and (\ref{eq.9}): \begin{equation} i\omega\,\rho c_{\rm p}\,\tilde T({\bf x},\omega) = \nabla\cdot\left[\kappa\nabla\tilde T({\bf x},\omega)\right] \label{eq.a1} \end{equation} \begin{equation} \tilde T_0(\theta,\varphi;\omega) = \sum_{l=0}^\infty\sum_{m=-l}^l\,\tilde b_{lm}(\omega)\,Y_{lm}(\theta,\varphi) \label{eq.a2} \end{equation} Equation (\ref{eq.a1}) can be recast in the form \begin{equation} \left(\nabla^2 + \gamma_1^2\right)\,\tilde T({\bf x},\omega) = 0\ , \quad 0\leq r\leq a_1 \label{eq.a3a} \end{equation} \begin{equation} \left(\nabla^2 + \gamma_2^2\right)\,\tilde T({\bf x},\omega) = 0\ , \quad a_1\leq r\leq a_2 \label{eq.a3b} \end{equation} where $r\/$\,$\equiv$\,$|{\bf x}|$, and \begin{equation} \gamma_1^2\equiv -i\omega\,\frac{\rho_1 c_{\rm p,1}}{\kappa_1}\ ,\quad \gamma_2^2\equiv -i\omega\,\frac{\rho_2 c_{\rm p,2}}{\kappa_2} \label{eq.a4} \end{equation} To these, matching conditions at the interface\footnote{ The temperature and the \emph{heat flux} should be continuous across the interface.} and boundary conditions must be added: \begin{equation} \tilde T(r=a_1-0,\omega) = \tilde T(r=a_1+0,\omega) \label{eq.a7a} \end{equation} \begin{equation} \kappa_1\,\frac{\partial \tilde T}{\partial r}(r=a_1-0,\omega) = \kappa_2\,\frac{\partial \tilde T}{\partial r}(r=a_1+0,\omega) \label{eq.a7b} \end{equation} \begin{equation} \tilde T(r=a_2,\omega) = \tilde T_0(\theta,\varphi;\omega) \label{eq.a7c} \end{equation} Equations (\ref{eq.a3a}) and (\ref{eq.a3b}) are of the Helmholtz kind. Their solutions are thus respectively given by \begin{equation} \hspace*{-2.25 cm} \tilde T({\bf x},\omega) = \left\{\begin{array}{ll} \displaystyle \sum_{lm}\,A_{lm}(\omega)\,j_l(\gamma_1r)\,Y_{lm}(\theta,\varphi)\ , & 0\leq r \leq a_1 \\[1.7 em] \displaystyle \sum_{lm}\,\left[C_{lm}(\omega)\,j_l(\gamma_2r) + D_{lm}(\omega)\,y_l(\gamma_2r)\,\right]\, Y_{lm}(\theta,\varphi)\ , & a_1\leq r \leq a_2 \end{array}\right. \label{eq.a5} \end{equation} where $j_l\/$ and $y_l\/$ are spherical Bessel functions \cite{as72}, \begin{equation} \hspace*{-0.8 cm} j_l(z) = z^l\,\left(-\frac{1}{z}\,\frac{d}{dz}\right)^{\!\!l}\, \frac{\sin z}{z}\ ,\quad y_l(z) = -z^l\,\left(-\frac{1}{z}\,\frac{d}{dz}\right)^{\!\!l}\, \frac{\cos z}{z} \label{eq.a6} \end{equation} and the coefficients $A_{lm}(\omega)$, $C_{lm}(\omega)$ and $D_{lm}(\omega)$ are to be determined by equations~(\ref{eq.a7a})--(\ref{eq.a7c}). These can be expanded as follows, respectively: \begin{eqnarray} \sum_{lm}\,A_{lm}(\omega)\,j_l(\gamma_1a_1)\,Y_{lm}(\theta,\varphi)\ = & & \nonumber \\ \ \ =\ \sum_{lm}\,\left[C_{lm}(\omega)\,j_l(\gamma_2a_1) + D_{lm}(\omega)\,y_l(\gamma_2a_1)\,\right]\, Y_{lm}(\theta,\varphi) & & \label{eq.a8a} \end{eqnarray} \begin{eqnarray} \kappa_1\gamma_1\, \sum_{lm}\,A_{lm}(\omega)\,j'_l(\gamma_1a_1)\,Y_{lm}(\theta,\varphi)\ = & & \nonumber \\ \ \ =\ \kappa_2\gamma_2\, \sum_{lm}\,\left[C_{lm}(\omega)\,j'_l(\gamma_2a_1) + D_{lm}(\omega)\,y'_l(\gamma_2a_1)\,\right]\, Y_{lm}(\theta,\varphi) \label{eq.a8b} \end{eqnarray} \begin{eqnarray} \sum_{lm}\,\left[C_{lm}(\omega)\,j_l(\gamma_2a_2) + D_{lm}(\omega)\,y_l(\gamma_2a_2)\,\right]\, Y_{lm}(\theta,\varphi)\ = & & \nonumber \\ \ \ =\ \sum_{lm}\,\tilde b_{lm}(\omega)\,Y_{lm}(\theta,\varphi) \label{eq.a8c} \end{eqnarray} Because of the completeness property of the spherical harmonics, the above equations completely determine the coefficients $A_{lm}(\omega)$, $C_{lm}(\omega)$ and $D_{lm}(\omega)$. The result is \begin{equation} \hspace*{-2 cm} A_{lm}(\omega) = \xi_l(\omega)\,\tilde b_{lm}(\omega)\ ,\ \ C_{lm}(\omega) = \eta_l(\omega)\,\tilde b_{lm}(\omega)\ ,\ \ D_{lm}(\omega) = \zeta_l(\omega)\,\tilde b_{lm}(\omega) \label{eq.a9} \end{equation} with \begin{equation} \hspace*{-1 cm} \xi_l(\omega) = \frac{1}{\Delta_l(\omega)}\,\left[ \kappa_2\gamma_2\,j_l(\gamma_2a_1)\,y'_l(\gamma_2a_1) - \kappa_2\gamma_2\,j'_l(\gamma_2a_1)\,y_l(\gamma_2a_1)\right] \label{eq.a9a} \end{equation} \begin{equation} \hspace*{-1 cm} \eta_l(\omega) = \frac{1}{\Delta_l(\omega)}\,\left[ \kappa_2\gamma_2\,j_l(\gamma_1a_1)\,y'_l(\gamma_2a_1) - \kappa_1\gamma_1\,j'_l(\gamma_1a_1)\,y_l(\gamma_2a_1)\right] \label{eq.a9b} \end{equation} \begin{equation} \hspace*{-1. cm} \zeta_l(\omega) = \frac{1}{\Delta_l(\omega)}\,\left[ \kappa_1\gamma_1\,j_l(\gamma_2a_1)\,j'_l(\gamma_1a_1) - \kappa_2\gamma_2\,j'_l(\gamma_2a_1)\,j_l(\gamma_1a_1)\right] \label{eq.a9c} \end{equation} and \begin{eqnarray} \hspace*{-1 cm} \Delta_l(\omega) & = & \ \kappa_1\gamma_1\,j'_l(\gamma_1a_1)\,\left[ j_l(\gamma_2a_1)\,y_l(\gamma_2a_2) - j_l(\gamma_2a_2)\,y_l(\gamma_2a_1)\right]\ + \nonumber \\ & + & \ \kappa_2\gamma_2\,j_l(\gamma_1a_1)\,\left[ j_l(\gamma_2a_2)\,y'_l(\gamma_2a_1) - j'_l(\gamma_2a_1)\,y_l(\gamma_2a_2)\right] \label{eq.a10} \end{eqnarray} When the above results, equations~(\ref{eq.a9a}) through (\ref{eq.a10}), are inserted back into equation~(\ref{eq.a5}) the result stated in equation~(\ref{eq.10}) in the main text obtains, i.e., \begin{equation} \tilde T({\bf x},\omega) = \sum_{lm}\,H_{lm}({\bf x},\omega)\,\tilde b_{lm}(\omega) \label{eq.a11} \end{equation} where \begin{equation} \hspace*{-1.8 cm} H_{lm}({\bf x},\omega) = \left\{\begin{array}{ll} \xi_l(\omega)\,j_l(\gamma_1r)\,Y_{lm}(\theta,\varphi)\ , & 0\leq r \leq a_1 \\[1 em] \left[\eta_l(\omega)\,j_l(\gamma_2r) + \zeta_l(\omega)\,y_l(\gamma_2r)\,\right]\, Y_{lm}(\theta,\varphi)\ , & a_1\leq r \leq a_2 \end{array}\right. \label{eq.a12} \end{equation} For monopole only boundary conditions, equation~(\ref{eq.15}), the transfer function is \begin{equation} \hspace*{-0.6 cm} H(r,\omega) = \left\{\begin{array}{ll} \xi_0(\omega)\,j_0(\gamma_1r)\ , & 0\leq r \leq a_1 \\[1 em] \eta_0(\omega)\,j_0(\gamma_2r) + \zeta_0(\omega)\,y_0(\gamma_2r)\ , & a_1\leq r \leq a_2 \end{array}\right. \label{eq.a13} \end{equation}
Title: The X-ray properties of young radio-loud AGN
Abstract: We present XMM-Newton observations of a complete sample of five archetypal young radio-loud AGN, also known Gigahertz Peaked Spectrum (GPS) sources. They are among the brightest and best studied GPS/CSO sources in the sky, with radio powers in the range L_{5GHz}=10^{43-44} erg/s and with 4 sources having measured kinematic ages of 570 to 3000 yrs. All sources are detected, and have 2-10 keV luminosities from 0.5 to 4.8x10^{44} erg/s. In comparison with the general population of radio galaxies, we find that: 1) GPS galaxies show a a range in absorption column densities similar to other radio galaxies. We therefore find no evidence that GPS galaxies reside in significantly more dense circumnuclear environment, such that they could be hampered in their expansion. 2) The ratio of radio to X-ray luminosity is significantly higher than for classical radio sources. This is consistent with an evolution scenario in which young radio sources are more efficient radio emitters than large extended objects at a constant accretion power. 3) Taking the X-ray luminosity of radio sources as a measure of ionisation power, we find that GPS galaxies are significantly underluminous in their [OIII]_{5007 Angstrom}, including a weak trend with age. This is consistent with the fact that the Stroemgren sphere should still be expanding in these young objects. This would mean that here we are witnessing the birth of the narrow line region of radio-loud AGN.
https://export.arxiv.org/pdf/astro-ph/0601141
\date{} \pagerange{\pageref{firstpage}--\pageref{lastpage}} \pubyear{2002} \label{firstpage} \newcommand{\apj}{{ApJ}} \newcommand{\apjs}{{ApJS}} \newcommand{\apjl}{{ApJ}} \newcommand{\aj}{{AJ}} \newcommand{\aap}{{A\&A}} \newcommand{\aaps}{{A\&AS}} \newcommand{\nat}{{Nat}} \newcommand{\jetp}{{JETP}} \newcommand{\mnras}{{MNRAS}} \newcommand{\phrvl}{{PhRvL}} \newcommand{\phrc}{{PhRvC}} \newcommand{\prc}{{PhRvC}} \newcommand{\araa}{{ARA\&A}} \newcommand{\pasj}{{PASJ}} \newcommand{\pasp}{{PASP}} \newcommand{\npa}{{NuPhA}} \newcommand{\iaucirc}{{IAU circ.}} \newcommand{\aplett}{{Astrophysical Letters}} \newcommand{\rvmp}{{\it Rev. Mod. Physics}} \newcommand{\xmm}{{\it XMM-Newton}} \newcommand{\chandra}{{\it Chandra}} \newcommand{\asca}{{\it ASCA}} \newcommand{\rosat}{{\it ROSAT}} \newcommand{\einstein}{{\it Einstein}} \newcommand{\cangeroo}{{\it CANGEROO}} \newcommand{\whipple}{{\it Whipple}} \newcommand{\hegra}{{\it HEGRA}} \newcommand{\hess}{{\it HESS}} \newcommand{\smm}{{\it SMM}} \newcommand{\sax}{{\it BeppoSAX}} \newcommand{\rxte}{{\it RXTE}} \newcommand{\osse}{{\it OSSE}} \newcommand{\egret}{{\it CGRO-EGRET}} \newcommand{\integr}{{\it INTEGRAL}} \newcommand{\glast}{{\it GLAST}} \newcommand{\comptel}{{\it COMPTEL}} \newcommand{\cgro}{{\it CGRO}} \newcommand{\xspec}{{\it xspec}} \newcommand{\oiii}{\hbox{[O\,III]}} \newcommand{\msun}{{$M_{\odot}$}} \newcommand{\nh}{{$N_{\rm H}$}} \newcommand{\ep}{{e$^+$e$^-$}} \newcommand{\fluxunit}{{ph\,cm$^{-2}$s$^{-1}$}} \newcommand{\kms}{{km\,s$^{-1}$}} \newcommand{\ndot}{{\dot{N}_{UV}}} \newcommand{\NH}{{N_{\rm H}}} \newcommand{\loiii}{{L_{\rm [O III]}}} \begin{keywords} galaxies: active -- X-ray: galaxies \end{keywords} \begin{table*} \centering \begin{minipage}{\textwidth} \caption{The sample of GPS/CSO radio sources with $|b|>20$\degr\ from the \citet{pearson88} catalogue. Indicated are, (column 1) the coordinates, (column 2) redshift, $z$, (column 3) radio flux density at 5~GHz, $S_{5 \rm GHz}$ (erg s$^{-1}$), (column 4) radio luminosity at 5~GHz, $L_{5\rm GHz}$\ (erg s$^{-1}$), (column 5) the 5007~\AA \oiii\ emission line luminosity, $L_{\rm [O III]}$\ (erg s$^{-1}$) from \citet{lawrence96}, and (column 6) the kinematic age of the radio hot spots \citep[][except for B1358+624]{polatidis03} \label{tab-sample}} \begin{tabular}{@{}lllcccc@{}}\hline Source & Position & $z$ & $S_{5 \rm GHz}$\footnote{Obtained from the NASA/IPAC Extragalactic Database (NED)}& $\log L_{5\rm GHz}$ & $\log L_{\rm [O III]}$ & Kinematic Age\\ & (J2000) & & Jy & & & yr\\ \hline B0108+388 & $01^h11^m37.3^s$\ +39\degr 06\arcmin 28\arcsec & 0.668 & 1.6 & 44.0 &40.8 & $570\pm50$\\ B0710+439 & $07^h13^m38.1^s$\ +43\degr 49\arcmin 17\arcsec & 0.518 & 1.6 & 43.7 & 42.4 & $930\pm100$\\ B1031+567 & $10^h35^m07.0^s$\ +56\degr 28\arcmin 47\arcsec & 0.45 & 1.3 & 43.4 & 41.6 & $1800\pm600$\\ B1358+624 & $14^h00^m28.6^s$\ +62\degr 10\arcmin 39\arcsec & 0.431 & 1.8 & 43.5 & 41.8 & $2400\pm1000$ \footnote{ The kinematic age of B1358+624 has not been directly measured, but is based on on its size and the average size-age relation of GPS/CSO sources in \citet{polatidis03}.} \\ B2352+495 & $23^h55^m09.4^s$\ +49\degr 50\arcmin 08\arcsec & 0.238 & 1.5 & 43.0 & 41.3 & $3000\pm750$\\ \hline \end{tabular} \end{minipage} \end{table*} \section{Introduction} Ever since their discovery, it has been speculated that those compact radio sources that show convex-shaped radio spectra at cm wavelengths, may be young objects \citep{shklovsky65,blake70}. As a class, they were named Gigahertz Peaked Spectrum (GPS) sources after their characteristic radio spectrum \citep[see][for a review]{odea98}, most likely caused by synchrotron self absorption \citep[e.g.]{fanti90,snellen00a}. High resolution Very Long Baseline Interferometry (VLBI) observations have shown that these sources are typically up to a few hundred parsec in size, often exhibiting jet and/or lobe structures on two opposite sides from their central core $-$ the reason why they are also called Compact Symmetric Objects \citep[CSO; eg.][]{Wilkinson94}. The most compelling evidence that GPS/CSO are indeed young radio sources comes from VLBI monitoring observations, showing that the bright archetypal objects in this class have hot-spot propagation velocities of $\sim$0.1-0.2c \citep{owsianik98a,owsianik98b,tschager00}, indicating kinematic ages of $\sim$10$^{3}$ years. This is in contrast to speculations that GPS/CSO galaxies are small due to confinement by a particularly dense and clumpy interstellar medium (ISM) that impedes the outward propagation of the jets \citep{vanbreugel84,odea91}. Note that a large fraction of GPS sources, in particular those showing a convex radio spectrum peaking at higher frequencies than a few GHz, turn out to be identified with high redshift quasars (z$\sim$2-3). Their connection with the population of GPS/CSO galaxies at lower redshift is not clear, and they may well be a completely separate class of object which just also happen to exhibit a convex shaped spectrum \citep{snellen99}. By no means it has been established that the GPS quasars may also represent a young stage of radio source evolution. Here we present \xmm\ X-ray observations of a small but complete sample of all five GPS sources from the \citet{pearson88} sample with Galactic latitude $|b| > 20$\degr\ (Table~\ref{tab-sample}). These are among the brightest GPS/CSO galaxies in the sky. All these sources appear to be young radio loud AGN with four out of five sources having measured kinematic ages of their hot spots, indicating ages of up to $\sim3000$~yr. No kinematic age measurement for B1358+624 exists, but based on its size and the age-size measurements of GPS/CSO sources \citep{polatidis03} we estimate its age to be 2400 yr. Although the sample is limited in size, it allows us to study the correlations between their X-ray, optical and radio properties, and to assess how they compare to those of mature radio galaxies. One expects that the X-ray luminosity is largely a manifestation of the instantaneous accretion power of the central black hole, whereas the radio luminosity is expected to evolve substantially over the life time of the source \citep[e.g.]{readhead96,snellen00a}. Moreover, X-ray absorption measurements are an excellent means to probe the circum nuclear density of the GPS galaxies. Several papers on X-ray observations of GPS sources have appeared in the literature, but one should be cautious to interpret these results in terms of X-ray properties of young radio-loud AGN. A first success was obtained by \citep{odea00} with ASCA. Although they did not detect B2352+495, they obtained a firm detection of GPS galaxy B1345+125 (PKS1345+125). Furthermore, \citet{guainazzi04} detected B1404+288 (Mkn 668).\footnote{Prior to the submission of this publication we learned that Guainazzi et al. have detected a number of other GPS/CSO galaxies in X-rays, which do not overlap with our sample \citep{guainazzi05}} Both are low redshift GPS galaxies exhibiting strong optical line emission and are powerful infrared emitters, and may be not representative to the class of young radio-loud AGN (yet no reliable ages have been measured for these sources). The combined X-ray and radio observations of the GPS {\em quasar} B0738+313 presented by \citet{siemiginowska03b} show that, with its prominent kpc-scale X-ray/radio jet, it is certainly not a young radio-loud AGN. In order to easily compare our results for GPS/CSO galaxies with the X-ray properties of a large sample of AGN by \citet{sambruna99} we adapt here a cosmology with $H_0 = 75$~km s$^{-1}$ Mpc$^{-1}$ and $q_0 = 0.5$. \begin{table} \centering \caption{Log of the \xmm\ observations. The MOS exposure time and event rates refer to the average value for MOS1 and MOS2.\label{tab-obs}} {% \begin{tabular}{@{}lcccc@{}}\hline Source & Observation ID & Start Date & Exposures & Event rates \\ & & & (MOS/PN) & (MOS/PN)\\ & & d/m/y & ks & ct s$^{-1}$\\ \hline B0108+388 & 0202520101 & 09/01/2004 & 16.4/12.0 & 1.6/13.6\\ B0710+439 & 0202520201 & 22/01/2004 & 14.2/11.2 & 3.2/26.6\\ B1031+567 & 0202520301 & 21/10/2004 & 22.2/12.8 & 34.4/93.0\\ B1358+624 & 0202520401 & 14/04/2004 & 12.4/12.0 & 23.2/120.6\\ B2352+495 & 0202520501 & 25/12/2004 & 15.8/12.8 & 3.7/28.5\\ \hline \end{tabular} } NOTE -- The event rates refer to the total detector count rates, not the source count rates. \end{table} \begin{table*} \begin{center} \begin{minipage}{\textwidth} \caption{Observational properties and parameters obtained by modeling the observed X-ray spectra in the range 0.5-10~keV.\label{tab-res}} \begin{tabular}{@{}llllll@{}}\hline\noalign{\smallskip} & B0108+388 & B0710+439 & B1031+567 & B1358+624 &B2352+495 \\ \noalign{\smallskip}\hline \noalign{\smallskip} PN 0.5-10~keV source count rate ($10^{-3}$~cts\,s$^{-1}$) & $4.7\pm0.9$ & $89.5\pm3.0$ &$12.6\pm2.0$ & $44.3\pm2.7$ & $8.6\pm1.2$\\ MOS1+2 0.5-10~keV source count rate ($10^{-3}$~cts\,s$^{-1}$) & $0.71\pm0.25$ & $33.0\pm1.1$ & $4.3\pm0.5$ & $15.8\pm0.9$ & $3.4\pm0.4$\\ \noalign{\smallskip} Galactic \nh\ ($10^{20}$cm$^{-2}$) & 5.80 & 8.11 & 0.56 & 1.96 &12.4\\ \noalign{\smallskip} Normalization \footnote{Statistical errors correspond to $\Delta C=1$ (68\% confidence limits).} ($10^{-5}$ph s$^{-1}$keV$^{-1}$ cm$^{-2}$@ 1 keV) & $3.3\pm1.1$\footnote{ The 3$\sigma$ lower limit is $1.1\times10^{-5}$ph s$^{-1}$keV$^{-1}$ cm$^{-2}$. The source is detected at the $7\sigma$ level.} &$8.1\pm0.6$ & $1.2\pm0.2$ & $6.1\pm1.6$ & $1.1\pm0.2$\\ Power law slope \footnote{Brackets indicate that the power law slope was fixed to this value.} ($\Gamma$) & (1.75) & $1.59\pm0.06$ & (1.75) & $1.24\pm0.17$ & (1.75) \\ Intrinsic \nh\ ($10^{22}$cm$^{-2} $) & $57\pm20$\footnote{ The 3$\sigma$ lower limit is $1.8\times10^{23}$~cm$^{-2}$.} & $0.44\pm 0.08$ & $0.50\pm0.18$ & $3.0\pm0.7$ & $0.66\pm0.27$\\ \noalign{\smallskip} Flux (2-10 keV) \footnote{Including absorption.} ($10^{-13}$~erg\, s$^{-1}$cm$^{-2}$) & 0.50 & 4.0 & 0.51 & 4.8 & 0.41\\ Luminosity (2-10 keV)\footnote{Calculated for rest frame energies, ignoring absorption.} ($10^{44}$~erg\, s$^{-1}$) & 1.18 & 2.16 & 0.22 & 1.67 & 0.046\\ C-statistic/bins &137.5/99 & 497.4/401 & 104.2/97 & 122.9/99 & 126.3/99 \\ \noalign{\smallskip} \hline \end{tabular} \end{minipage} \end{center} \end{table*} \section{Observations, spectral analysis and results} \xmm\ \citep{jansen01} observed the five GPS/CSO sources as part of its guest observation program from January to December 2004 (Table~\ref{tab-obs}). All observations were made with the ``Thin1'' optical blocking filter. For the data reduction we used the standard \xmm\ software package SAS v6.0.0. Unfortunately several observations were plagued by a high particle background (see Table~\ref{tab-obs}). In the case of {B0108+388}, {B0710+439}, {B2325+495} we removed time intervals with a high background count rate, using cut off rates of 15.5~ct\,s$^{-1}$ and 2.5~ct\,s$^{-1}$ for resp. PN and MOS. For {B1031+567} and {B1358+624} the high background persisted throughout the observation, and we simply used all available data. The observation {B2325+495} had an intermediate background activity, so we selected time intervals with $< 40$~ct\,s$^{-1}$ and $< 5$~ct\,s$^{-1}$ for PN and MOS. For spectral extraction we used circular extraction regions with radii of 15\arcsec\ for {B1031+567} and {B1358+624}, and 25\arcsec\ for the other three sources. Background spectra were obtained from rectangular regions near the source position, but excluding regions around 35\arcsec\ of the source. We extracted spectra for the two MOS CCD cameras \citep{turner01}, and the PN camera \citep{strueder01}. For each source we combined the spectra of MOS1 and MOS2, into one spectrum, which we analysed using averaged instrumental response matrices. The potential systematic error introduced is small compared to the statistical errors, given the fact that the MOS1 and MOS2 are virtually identical instruments with similar instrumental response functions. All the five sources of the sample are detected. For the spectral analysis we employed a simple model consisting of a power law continuum and two absorption components: One represents the Galactic absorption, with an absorption column, \nh, fixed to the Galactic value of \citet{dickey90}\footnote{We extracted the absorption columns from the online ``\nh-tool'', \\ \url{http://heasarc.gsfc.nasa.gov/Tools}}. The other absorption component corresponds to the absorption column intrinsic to the host galaxy. The redshift of this component was fixed to that of the galaxy, but the absorption column density was a free parameter. The spectral analysis was done with the spectral fitting program {\it xspec} \citep{xspec}, using the absorption models of \citet{wilms00} (called {\it tbabs} and {\it ztbabs} in \xspec). The slope of the power law continuum was a free parameter for the high signal to noise spectra of {B0710+439} and {B1358+624}, but fixed to 1.75 for the other three sources whose spectra are statistically more limited. The best fit parameters, together with the inferred intrinsic X-ray luminosities between 2-10~keV are listed in Table~\ref{tab-res}. The spectra and best fit models are shown in Fig.~\ref{spectra1}. \section{Interpretation} As we are interested in whether GPS/CSO galaxies are different from other radio-loud AGN we compare their X-ray properties to those of the sample of radio-loud AGN whose X-ray properties were determined by \citet{sambruna99} from ASCA observations. We use all the sources of their broad line radio galaxies (BLRG), narror line radio galaxies (NLRG) and radio galaxies (RG) subsamples. Since \citet{sambruna99} do not quote upper limits for the non-detected intrinsic X-ray asborption components, we estimate upper limits proportional to the observed flux taking into account the errors on the detected instrinsic absorption components. In Fig.~\ref{correlations} we have set out the radio and \oiii\ luminosities of the galaxies in our sample to their X-ray luminosities, and we compare them to the \citet{sambruna99} sample. In Fig.~\ref{fig-nhd} we show the intrinsic absorption column density distribution. These figures reveal three important properties of our sample of GPS/CSO galaxies: 1) for their X-ray emission GPS/CSO galaxies are relatively radio-loud; 2) their \oiii\ emission is relatively low; 3) the column density distribution is similar to those of radio-loud AGN classified by \citet{sambruna99} as narrow line radio galaxies (NLRGs) and radio galaxies (RGs), but the absorption is on average higher than those of broad line radio galaxies. As we discuss below these three properties support the hypothesis that GPS/CSO galaxies are indeed young radio-galaxies. {% Note that there are strong indications that GPS radio galaxies have relatively low [OIII] lumininosities with respect to their radio luminosities as compared to compact steep spectrum (CSS) sources \citep{odea98}. This is probably the same trend between radio, [OIII], and X-ray luminosity that we report here, but without the X-ray luminosity as intermediary quantity. As noted in the introduction, the X-ray luminosity is the quantity that is probably the best indicator for the intrinsic power of the AGN. } \begin{table} \caption{A comparison of the X-ray derived absorption column $\NH$ and radio absorption column measurements $N_{\rm HI}$ \citep{pihlstroem03}. \label{tab-nh}} \centering \begin{tabular}{@{}lcr@{}}\hline &$\log \NH$ &$\log N_{\rm HI}$ \\\hline B0108+388 & 23.8 & 21.9 \\ B1031+567 & 21.7 & $<20.1$ \\ B1358+624 & 22.5 & 20.3\\ B2352+495 & 21.8& 20.5 \\\hline \end{tabular} \end{table} \subsection{The absorbing column density} \label{sec-nh} An alternative explanation for the small extent of the radio jets in GPS/CSO galaxies that is still often considered in the literature \citep{odea98} is that the radio jets are quenched by a high density in the vicinity of the nucleus, in other words they are ``frustrated radio sources''. It is clear from Table~\ref{tab-res} and Fig.~\ref{fig-nhd} that this is unlikely to be the case, as the intrinsic X-ray absorption is similar to other radio-loud AGN, with column densities ranging from a relatively modest $4\times10^{21}$~cm$^{-2}$\ to a considerably large $6\times10^{23}$~cm$^{-2}$. A similar conclusion was drawn by \citet{pihlstroem03} based on HI radio absorption observations of a large sample of compact radio sources, and by \citep{odea05} from upper limits to the molecular gas content in GPS sources. Note that the X-ray absorption gives more stringent constraints on the actual gas column densities than HI and molecular absorption densities, as X-ray absorption depends on the total column toward the central source, whereas radio observations only probe the neutral fraction of the gas. This is quite an important distinction since AGN are expected to create an extended ionised region, as we discuss in section~\ref{sec-oiii}. Furthermore, X-ray absorption probes the gas toward the accretion disk, whereas the radio absorption probes the neutral gas toward the radio source, which is situated at larger radii. It is therefore not surprising that for the four galaxies in our sample for which also HI absorption measurements have been made the HI column is always one to two orders of magnitude lower than the X-ray absorption column (Table~\ref{tab-nh} and Fig.~\ref{fig-nh}). Attributing the difference solely to ionisation effects would mean ionisation fractions of 90\% to 99\%. However, the ionisation fractions are likely to be lower, because a substantial part of the X-ray absorption may occur inside the central 100~pc, which is not probed by absorption toward the radio hot spots. \citet{pihlstroem03} found a strong anti-correlation between the linear size of the radio emission and the HI column density, which they use to probe the average density profile of the interstellar medium. They do not consider ionisation effects, whereas this could be an additional cause for the observed anti-correlation: small jets are associated with young AGN, which do therefore not yet have an extended narrow line emission region of ionised gas (see section~\ref{sec-oiii}). If this is the case it makes it less straightforward to derive an average interstellar medium density profile from the relation between $N_{\rm HI}$\ and linear size of the radio emission, since the growth of the emission line region also depends on the density and UV luminosity of the AGN. \subsection{The optical line emission: the case for an expanding emission line region} \label{sec-oiii} Usually a high [O III] luminosity is taken as an indication for the presence of a powerful AGN, but Fig.~\ref{correlations} indicates that GPS/CSO galaxies are relatively underluminous in [O III] compared to their radio or X-ray flux. This may not be too surprising if one takes into account that these are young radio galaxies, in which the AGN has switched on only a few thousand years ago. The reason is that it takes time to establish a large emission line region by photo-ionisation. This is best illustrated by a simple calculation, for which we assume an average interstellar medium density of 1~cm$^{-3}$\ and a typical luminosities of $\log L_{\rm \oiii} = 42$, and $\log L_{X} = 44$. The number rate of ionising photons (i.e. $> 13.6$ eV) is $\ndot \sim 10^{54}$ ph s$^{-1}$\ for a power law spectrum with photon index -1.75. Comparing this to the total number of hydrogen atoms within a typical region of 5~kpc radius \citep{baum89b}, one finds that $\sim 10^{67}$ atoms have to be ionised. In other words the central source has to shine for at least $\sim 10^{67}/10^{54} = 10^{13}$~s, or $\sim$300,000 yr before it has completely ionized a region with a radius of 5~kpc. This means that if an AGN has become only recently active it must be surrounded by a small, but rapidly expanding ionisation nebula \citep[see][for a calculation concerning the first generation of quasars]{white03}. Hence, this would imply that in GPS/CSO galaxies we see the birth of the narrow line region of radio-loud AGN. In order to see whether this is the reason that GPS/CSO galaxies are relatively underluminous in [O III] we consider a simplified model for the evolution of a Str\"omgen sphere. For an old ionisation nebula in equilibrium, the number rate of ionising photons ($\ndot$) should equal the number of recombinations, from which follows the equilibrium radius, $R_i$, of a Str\"omgen sphere: \begin{equation} R_i^3 = \frac{3 \ndot}{4\pi n_{\rm H} n_{\rm e} \alpha_{\rm H}}, \end{equation} with $\alpha_{\rm H}$\ the hydrogen recombination coefficient. The initial rapid expansion of the ionisation nebula is described by e.g. \citet{spitzer68} \begin{equation} r_i^3 = R_i^3\{1- \exp( -n_{\rm e} \alpha_{\rm H} t)\} ,\label{ifront} \end{equation} with $r_i$\ the radius of the ionisation front and $t$\ the time since the central source switched on. For small $n_{\rm e} \alpha_{\rm H} t$ we can approximate this by \begin{equation} r_i^3 \approx n_{\rm e} \alpha_{\rm H} t R^3 = \frac{3 \ndot}{4\pi n_{\rm H} } t \end{equation} The \oiii\ line emissivity is then given by \begin{eqnarray} \dot{N}_{\rm \oiii} = \frac{4\pi}{3} r_i^3 n_{\rm e} n_{\rm OIV} f_{5007 \rm\AA}\alpha_{\rm O III} =\nonumber\\ n_{\rm e} \frac{n_{\rm OIV}}{ n_{\rm H}}f_{5007 \rm\AA}\ \alpha_{\rm O III} \ndot t, \end{eqnarray} with $f_{5007 \rm\AA}$ the probability of emitting a photon at $5007$~\AA\ after recombination. Assuming the interstellar medium densities in the GPS/CSO galaxies are more or less similar, we can expect the following correlation between the \oiii\ luminosity and the number rate of ionising photons for young GPS/CSO sources of kinematic age $\tau$, provided that the age of the radio galaxy coincides with the birth of the ionisation nebula: \begin{equation} L_{\rm \oiii} \propto \tau \ndot \label{eq-scaling}. \end{equation} As a first approximation we consider whether there is a relation between the observables $L_{\rm \oiii}$, and $\tau$\ and $L_X$. However, Fig.~\ref{fig-opt-age} illustrates that the five sources in our sample do not support a simple scaling of $L_{\rm \oiii} \propto \tau L_X$. There may be various reasons why this is not the case: e.g. the interstellar medium density varies from galaxy to galaxy, their is no simple proportionality between $L_X$ and $\ndot$, due to different spectral energy distributions (different spectral slopes, or spectral breaks), and shocks induced by jet cloud interactions may provide an additional source of ionisation.\footnote{Although we do not have kinematic age estimates of {PKS1345+125} and Mkn668 \citep{guainazzi04}, reasonable values for the age in fact show these galaxies to be too bright in \oiii\ compared to the galaxies in our sample. In this case the reason is very likely that a large part of the \oiii\ emission is not related to the activity of the central nucleus, as both galaxies show evidence of recent merger activity, and are bright infrared sources.} Moreover, one of the outliers is B1358+624, for which we only have an approximate age. However, a hint of what may the prime reason for deviations from Eq.~\ref{eq-scaling} is provided by the very low \oiii\ luminosity of {B0108+388}, because this is also the source with the highest absorption column (Table~\ref{tab-res}). So the most likely reason that the optical emission is lower than expected is that the ionising UV flux is blocked by absorbing material close to the nucleus. The absorbing material is probably not the result of neutral hydrogen and helium, as, being close to the nucleus it would be ionised almost immediately, but dust grains, which may survive the extreme conditions close to the nucleus for 1000~yr to $10^6$~yr, depending on the destruction mechanisms and dust particle sizes \citep[e.g.][]{villar-martin01}. For a young source like {B0108+388} this means that an appreciable amount of dust may still enshroud the nucleus, frustrating the formation of a narrow emission line region. Note that this may also explain the relatively high neutral hydrogen column density of {B0108+388} \citep[][Fig.~\ref{fig-nh}]{pihlstroem03}; the hydrogen ionisation fraction of the inner stellar medium is likely to be low, which would also support the idea that photo-ionisation is the dominant source of ionisation. If ionisation is dominated by shocks generated by the jet, one would expect that {B0108+388} would be relatively bright in \oiii\ as its interstellar medium density is apparently high. B1358+624 deviates less from the expected relation in the right hand panel of Fig.~\ref{fig-opt-age} due to its relatively flat X-ray spectrum, which makes that the number flux of UV photons is relatively small with respect to the X-ray luminosity. However, since we do not know the broad band spectral shape, we do not want to overemphasise this. Let us now consider the possible implications of dust absorption. The optical depth of dust particles depends on dust particles cross sections ($\sigma_d = \pi r^2$ with $r$ the physical size of the particles, $r\sim0.1~\mu$m) and the column density of dust particles $N_d$. To obtain an order of magnitude estimate we assume that most dust particles consist of silicates, and that all silicon is depleted into dust. As $N_{\rm Si} \approx 4\times10^{-4} \NH$, and a typical dust particle density is $\rho = 3.5$~g cm$^{-3}$, we have $N_d \approx 10^{-13} \NH$, and the optical depth should be around $\sigma_d N_d = 3\times10^{-23} N_{\rm H}$. This means that dust particles absorb an appreciable amount of UV flux if the hydrogen column density is comparable to, or exceeds, $10^{23}$cm$^{-2}$, which is only the case for {B0108+388}. We have therefore extrapolated from the observed $L_X$\ the ionising photon luminosity $\ndot$\ using the observed spectral properties. Plotting now $\loiii$\ as a function of $\tau\ndot$ we see less scatter, certainly if we allow for dust absorption (Fig.~\ref{fig-opt-age}). In order to bring {B0108+388} on the expected relation we need a conversion of $\NH$ to dust optical depth that is 16\% of the above order magnitude estimate. Note that in a log-log plot the absorption enters linearly, since $\ln(\dot{N}_{UV-abs}) = \ln(\ndot) - \sigma_d N_d$. Hence, only {B0108+388} is likely to be significantly affected by dust absorption. The large uncertainties in extrapolating from the observed $L_X$\ to an ionising UV flux makes that we cannot use Fig.~\ref{fig-opt-age} (right panel) to prove that Eq.~\ref{eq-scaling} is an accurate description, but it makes it at least plausible that for GPS/CSO galaxies the \oiii\ emission is relatively low due to an underdeveloped narrow emission line region. This is consistent with the idea that GPS-galaxies have AGN that switched on around the same time that the radio jets were formed. Further support for the idea that GPS/CSO galaxies are in the process of creating an extended narrow line region comes from the fact that neutral hydrogen apparently extends close to the compact radio jets,% given the fact that there is strong anti-correlation between the neutral hydrogen column density and jet-size \citep{pihlstroem03}. Note that GPS/CSO galaxies have sizes of the order of a few 100~pc, whereas narrow emission line regions can extend up to 10~kpc. \subsection{The radio luminosity versus the X-ray luminosity} The X-ray emission from AGN is thought to come predominantly from the immediate vicinity of the central black hole, i.e. thermal emission from the accretion disk reprocessed by the hot plasma in its vicinity. It is unlikely that synchrotron radiation from the jet makes a dominant contribution to the X-ray band. The reason is that, given the typical magnetic fields infered from radio luminosities \citep[$>1$~mG, ][]{odea98}, the synchrotron cooling time relevant for X-ray synchrotron radiation is in the order of only 2~yr for an electron energy of 10~erg. This means that a very small fraction of the total radio jet volume could produce X-ray synchrotron emission.{ Another potential source of X-ray emission from outside the central region could be inverse Compton emission by the relativistic electron population in the jets \citep[c.f.][]{belsole05}. Although we cannot totally exclude a significant inverse Compton contribution, it seems unlikely to be the case for our sample. The reason is that the X-ray emission should in that case come from the same locations as the radio emission (the bright regions in the jets). However, we would then expect that the measured X-ray absorption columns would be more consistent with radio absorption measurements toward the jets, which is not the case (section~\ref{sec-nh}). } It is therefore reasonable to assume that, compared to the radio and \oiii\ luminosity, the X-ray luminosity is more directly related to the accretion power of the AGN. Nevertheless, the radio and optical luminosities are still indirectly related to the accretion power, given the correlations between radio, optical and X-ray luminosities \citep[e.g.][]{sambruna99}. It is, therefore, interesting that GPS/CSO galaxies seem on average radio bright compared to the X-ray luminosity (Fig.~\ref{correlations}). Given the absorption column distribution and low \oiii\ emission, we can ignore the interpretation that the radio emission is relatively bright because the interstellar medium is dense in GPS/CSO galaxies, as argued by advocates of the ``frustrated radio source'' scenario. It is, therefore, very probable that GPS/CSO galaxies are radio bright because they are young. However, within our sample no relation between age and radio over X-ray ratio can be seen (Fig.~\ref{fig-radio-age}), nor is there a correlation with column density. Nevertheless, the fact that for the sample as a whole the radio to X-ray luminosity is brighter than for other radio-loud AGN is at least qualitatively in agreement with several radio evolution models, such as \citet{fanti95,readhead96,kaiser97,alexander00} and \citet{snellen00a}. These models describe the evolution of radio jets, with radio brightnesses depending on the radial density distribution of the interstellar medium. The radio jets are relatively bright as long as the they plow through the dense regions of the galaxies, but decline as soon as they propagate outside the core of the galaxy. The difference between the various models is that some assume a density distribution described by a King profile \citep{snellen00a,alexander00}, whereas others assume a density profile falling of as power law of radius \citep{kaiser97}. As a result, the non-power law models predict that galaxies in the GPS-phase are still increasing in radio luminosity in time, until the jet has reached the core radius of $\sim 1$~kpc, after which the luminosity declines. In this case the relatively brightest phase of radio galaxies would be represented by the so-called compact steep spectrum sources (CSS), which have more extended radio jets than GPS/CSO sources. The power law density models predict that right after the jet emerges from the core region the radio emission starts to decline. Our results are inconclusive regarding the details of the early evolution of the radio luminosity. However, future X-ray observations of CSS galaxies could help to clarify the brightness evolution further, as models with a King profile predict that CSS galaxies should, on average, have a higher radio to X-ray luminosity ratio than GPS/CSO galaxies, whereas models that assume power law density profiles predict that CSS galaxies should have a smaller radio to X-ray luminosity. \section{Conclusions} We have presented \xmm\ observations of a sample of all the GPS/CSO galaxies with $|b|>20$\degr\ from the \citet{pearson88} catalog, four of which have measured kinematic time scales for the jet expansion. All five of the sources are detected by \xmm\ thereby increasing the number of X-ray detected GPS/CSO galaxies from 2 to 7. These detections allow us to compare the X-ray properties of GPS/CSO galaxies with those of other radio loud AGN. The results presented here support the hypothesis that GPS/CSO galaxies represent the young phases in the evolution of radio-loud AGN. The alternative explanation that the radio sources are compact due to confinement by an exceptionally high density of the interstellar medium in those galaxies, seems extremely unlikely in view of the low intrinsic X-ray absorption column densities, which ranges from $\NH = 4\times10^{21}$~cm$^{-2}$\ to $6\times10^{23}$~cm$^{-2}$, and has a distribution similar to other radio loud AGN. {% After submission of our manuscript, a preprint by \citet{guainazzi05} arrived at apparently different conclusions based on a sample of five different GPS galaxies observed by \chandra\ and \xmm. Only one of the five GPS galaxies in their sample has \nh $< 10^{22}$~cm$^{-2}$, whereas this is 75\%$\pm$26\% for their control sample. Note, however, that in our sample three out of five have \nh $< 10^{22}$~cm$^{-2}$. Taking both samples together, this means that four out of ten GPS galaxies, or 40\%$\pm$20\% have \nh $< 10^{22}$~cm$^{-2}$\ consistent with the control sample. The fact that the absorption columns toward GPS galaxies are consistent with those of other radio galaxies suggests that the interstellar medium densities in the cores of GPS/CSO galaxies are similar to those of other radio loud AGN.} A similar conclusion was reached by \citet{pihlstroem03} based on HI radio absorption observations, but the X-ray data provide stronger constraints, as the total column density contributes to the absorption, including ionized regions of the interstellar medium. This may contribute to the fact that in all cases the X-ray column densities are higher than the HI column densities. A difference between the radio column densities and the X-ray column densities is also that the X-ray column is measured toward the central source, whereas the HI column density is toward the jets, which extend outside the central region. If the difference between radio and X-ray column density is dominated by those geometrical effects, this would be additional evidence against the confinement scenario, since the jets have apparently been able to pierce through the dense local regions that contribute most to the X-ray absorption column. Although the $\NH$\ distribution of our sample cannot be distinguished from other radio-loud AGN, the ratio of radio to X-ray luminosity shows that GPS galaxies have a strong tendency to be relatively radio bright. This supports the view that for the same thrust of the jets younger radio sources are relatively bright \citep{kaiser97,snellen00a}. However, the data is inconclusive concerning whether GPS-galaxies represent the most radio-luminous phases in the lifes of radio-loud AGN, as would be the case if the source develops in a density profile that drops of as power law with distance \citep{kaiser97,pihlstroem03}, or whether they are still in their brightening phase. This would be the case if the interstellar medium is best described by a King profile with a relative uniform density within $\sim$1~kpc of the center and then dropping of as a power law \citep[e.g.][]{snellen00a}. In the latter case the brightest evolutionary phase of radio-loud AGN would be represented by the compact steep spectrum sources (CSS), which have more extended radio emission than GPS/CSO galaxies. A similar study to this one concerning CSS-galaxies can clarify this issue. Finally, we find that GPS/CSO galaxies are relatively weak in \oiii\ line emission. This is again in support of the idea that GPS/CSO galaxies represent the very earliest stages of the evolution of radio-loud AGN, since narrow line regions need time to build up to their equilibrium size, and young narrow line regions are therefore not as bright as fully developed ones. The narrow line region is powered by the UV flux of the central source, but also shocks induced by the expanding jets are likely to contribute to their formation. For those very young radio-loud AGN we advocate here that their emission line regions are powered by the UV flux from the central sources. A case in point is that the relatively weakest \oiii\ source, {B0108+388}, has also the highest X-ray column density, which suggests that a dusty torus is partially blocking the UV light. The low \oiii\ luminosity therefore shows that the birth of the narrow emission line region must coincide more or less with the birth of the radio jet. {% However, we caution that we only have a limited knowledge of the nature of the ionization mechanism for the [O III] line emission. i.e. both shocks from the jets, as the UV radiation from the AGN may contribute to the ionization. Moreover, compact steep spectrum (CSS) sources, probably representing a more advanced evolutionary state of radio galaxies than GPS galaxies, show that the forbidden line emission tends to be aligned with the radio jet \citep{devries99}. This phenomenon is not well understood, but it should be accounted for if one wants to build a more detailed model of the evolution of emission line nebulae in radio galaxies.} In summary, the findings from this X-ray study lends further support to the theory that GPS/CSO galaxies represent the early phases of radio-loud AGN, in which also the narrow line emission nebula is still in the early phases of its evolution. Future X-ray studies may help to further clarify the relation between age or jet-size, the extent or brightness of the emission line region and the power of the X-ray emission. It is important to include also compact steep spectrum (CSS) sources in such a study, as they are likely to represent the next phase in the evolution of radio-loud AGN. Comparing their X-ray to radio luminosity may help clarify whether the radio emission declines already during the GPS-phase, or first increases, then peaks around the CSS-phase and from then on weakens. \section*{Acknowledgments} We thank for Elisa Costantini for helpful discussions on dust grain properties. The Space Research Organization of the Netherlands is supported financially by NWO, the Netherlands Organization for Scientific Research. This research has made use of the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. \xmm\ is an ESA science mission, with instruments and contributions directly funded by the ESA member states and the USA (NASA).
Title: The Anomalous Early Afterglow of GRB 050801
Abstract: The ROTSE-IIIc telescope at the H.E.S.S. site, Namibia, obtained the earliest detection of optical emission from a Gamma-Ray Burst (GRB), beginning only 21.8 s from the onset of Swift GRB 050801. The optical lightcurve does not fade or brighten significantly over the first ~250 s, after which there is an achromatic break and the lightcurve declines in typical power-law fashion. The Swift/XRT also obtained early observations starting at 69 s after the burst onset. The X-ray lightcurve shows the same features as the optical lightcurve. These correlated variations in the early optical and X-ray emission imply a common origin in space and time. This behavior is difficult to reconcile with the standard models of early afterglow emission.
https://export.arxiv.org/pdf/astro-ph/0601350
\title{The Anomalous Early Afterglow of GRB 050801} \author{ E.~S.~Rykoff,\altaffilmark{1}, V.~Mangano,\altaffilmark{2}, S.~A.~Yost\altaffilmark{1}, R.~Sari\altaffilmark{3}, F.~Aharonian\altaffilmark{4}, C.~W.~Akerlof\altaffilmark{1}, M.~C.~B.~Ashley\altaffilmark{5}, S.~D.~Barthelmy\altaffilmark{6}, D.~N.~Burrows\altaffilmark{7}, N.~Gehrels\altaffilmark{6}, E.~G\"{o}\v{g}\"{u}\c{s}\altaffilmark{8}, D.~Horns\altaffilmark{4}, \"{U}.~K{\i}z{\i}lo\v{g}lu\altaffilmark{10}, H.~A.~Krimm\altaffilmark{6,11}, T.~A.~McKay\altaffilmark{1}, M.~\"{O}zel\altaffilmark{12}, A.~Phillips\altaffilmark{5}, R.~M.~Quimby\altaffilmark{13}, G.~Rowell\altaffilmark{4}, W.~Rujopakarn\altaffilmark{1}, B.~E.~Schaefer\altaffilmark{14}, D.~A.~Smith\altaffilmark{15}, H.~F.~Swan\altaffilmark{1}, W.~T.~Vestrand\altaffilmark{16}, J.~C.~Wheeler\altaffilmark{13}, J.~Wren\altaffilmark{16}, F.~Yuan\altaffilmark{1}, } \altaffiltext{1}{University of Michigan, 2477 Randall Laboratory, 450 Church St., Ann Arbor, MI, 48109, erykoff@umich.edu} \altaffiltext{2}{INAF-IASF, Palermo, Italy} % \altaffiltext{3}{California Institute of Technology, Pasadena, CA, 91125, USA} \altaffiltext{4}{Max-Planck-Institut f\"{u}r Kernphysik, Saupfercheckweg 1, 69117 Heidelberg, Germany} \altaffiltext{5}{School of Physics, Department of Astrophysics and Optics, University of New South Wales, Sydney, NSW 2052, Australia} \altaffiltext{6}{NASA Goddard Space Flight Center, Laboratory for High Energy Astrophysics, Greenbelt, MD 20771} \altaffiltext{7}{Pennsylvania State University, University Park, PA, 16802, USA} \altaffiltext{8}{Sabanc{\i} University, Istanbul, Turkey} \altaffiltext{9}{Istanbul University Science Faculty, Department of Astronomy and Space Sciences, 34119, University-Istanbul, Turkey} \altaffiltext{10}{Middle East Technical University, 06531 Ankara, Turkey} \altaffiltext{11}{Universities Space Research Association, 10227 Wincopin Circle, Suite 212, Columbia, MD 21044} \altaffiltext{12}{\c{C}anakkale Onsekiz Mart \"{U}niversitesi, Terzio\v{g}lu 17020, \c{C}anakkale, Turkey} \altaffiltext{13}{Department of Astronomy, University of Texas, Austin, TX 78712} \altaffiltext{14}{Department of Physics and Astronomy, Louisiana State University, Baton Rouge, LA 70803} \altaffiltext{15}{Guilford College, Greensboro, NC, 27410, USA} \altaffiltext{16}{Los Alamos National Laboratory, NIS-2 MS D436, Los Alamos, NM 87545} \keywords{gamma rays:bursts} \section{Introduction} Gamma-ray bursts (GRBs) are the most luminous explosions in the universe, but the origin of their emission remains elusive. With the launch of the \emph{Swift} $\gamma$-ray Burst Explorer~\citep{gcgmn04} in late 2004, great progress has been made in the study of the early afterglow phase of GRBs. However, only a small number of bursts have been imaged simultaneously in both the optical and X-ray bands in the first minutes after the burst~\citep{nkgpg05,qryaa05,rykaa05,bbbbc05}. In this letter, we report on the earliest detection of optical emission, starting at 21.8 seconds after the onset of GRB~050801 with the ROTSE-IIIc (Robotic Optical Transient Search Experiment) telescope located at the H.E.S.S. site in Namibia. This is the most densely sampled early lightcurve yet obtained. It does not fade or brighten significantly over the first $\sim250$ seconds, after which there is a break and the lightcurve declines in a typical power-law fashion. The \emph{Swift}/XRT also obtained early observations starting at 69 seconds after the burst onset. The X-ray lightcurve shows the same features as the optical lightcurve. These correlated variations in the early optical and X-ray emission imply a common origin in space and time. This behavior differs from that seen in GRB~050319~\citep{qryaa05}, GRB~050401~\citep{rykaa05}, and GRB~050525a~\citep{bbbbc05}. It is difficult to explain this behavior with standard models of early afterglow emission without assuming there is continuous late time injection of energy into the afterglow. \section{Observations and Analysis} \label{sec:observations} The ROTSE-III array is a worldwide network of 0.45~m robotic, automated telescopes, built for fast ($\sim 6$ s) responses to GRB triggers from satellites such as HETE-2 and \emph{Swift}. They have wide ($1\fdg85 \times 1\fdg85$) fields of view imaged onto Marconi $2048\times2048$ back-illuminated thinned CCDs, and operate without filters. The ROTSE-III systems are described in detail in \citet{akmrs03}. On 2005 August 01, \emph{Swift}/BAT detected GRB~050801 (\emph{Swift} trigger 148522) at 18:28:02.1 UT. The position was distributed as a Gamma-ray Burst Coordinates Network (GCN) notice at 18:28:16 UT, with a $4\arcmin$ radius $3\sigma$ error circle. The burst had a $T_{90}$ duration of $20\pm3\,\mathrm{s}$ in the 15-350 keV band, and consisted of two peaks separated by around 3 seconds. The position was released during the tail end of the $\gamma$-ray emission~\citep{smbbc05}. The \emph{Swift} satellite immediately slewed to the target, with the XRT beginning observations in windowed timing mode at 69 s after the start of the burst and switching to photon counting mode at 89.3 s after the trigger. ROTSE-IIIc, at the H.E.S.S. site in Namibia, responded automatically to the GCN notice, beginning its first exposure in less than 8 s, at 18:28:23.9 UT. The automated burst response included a set of ten 5-s exposures, ten 20-s exposures, and 134 60-s exposures before the burst position dropped below our elevation limit. The first set of ten exposures were taken with subframe readout mode to allow rapid sampling (3-s readout between each 5-s exposure). Near real-time analysis of the ROTSE-III images detected a $15^{th}$ magnitude source at $\alpha=13^h36^m35\fs4$, $\delta=-21\arcdeg55\arcmin42\farcs0$ (J2000.0) that was not visible on the Digitized Sky Survey red plates, which we reported via the GCN Circular e-mail exploder within 7 minutes of the burst~\citep{ryr05}. No spectroscopic redshift has been reported for this GRB, although the \emph{Swift}/UVOT detected the afterglow in all filters including the $UVW2$ filter at $188\,\mathrm{nm}$~\citep{bbhgc05}, which implies that the redshift is $\lesssim1.2$. In addition, the afterglow was dimmer than 23 mag with no evidence for a bright host galaxy~\citep{fjhww05b}. The X-ray photometry is shown in Table~\ref{tab:xray}. Time bin midpoints and durations are listed in seconds, relative to the \emph{Swift} trigger time, 18:28:02 UT. The count rate is in counts/s and the flux is in $10^{-11}\,\mathrm{erg}\,\mathrm{cm}^{-2}\,\mathrm{s}^{-1}$, for the energy range 0.2-10 keV. The X-ray data has been corrected for a hot CCD column crossing the source as well as a nearby source $30''$ away. Photon counting data from the first orbit have been corrected for pile-up. We chose a time binning that ensures a detection of at least $3.5\sigma$ for each time bin before corrections were applied. The gaps in the data are caused by earth occultation. There is no spectral variation across the lightcurve, and the $N_H$ value is consistent with the Galactic value ($7\times10^{20}\,\mathrm{cm}^{-2}$). The best-fit spectrum (with $N_H$ fixed to $7\times10^{20}\,\mathrm{cm}^{-2}$) is a power law with photon index $1.87\pm0.15$ (90\% confidence level). The relative errors for the fluxes are slightly larger than those for the count rate due to the additional systematic error from the conversion. \begin{deluxetable}{cccc} \tablewidth{0pt} \tablecaption{\emph{Swift}/XRT observations of the afterglow of GRB~050801.\label{tab:xray}} \tablehead{ \colhead{T-mid (s)} & \colhead{Duration (s)} & \colhead{Count Rate ($\mathrm{cts}\,\mathrm{s}^{-1}$)} & \colhead{Flux ($10^{-11}\,\mathrm{erg}\,\mathrm{cm}^{-2}\,\mathrm{s}^{-1}$)} } \startdata 74.1 & 10.0 & $3.46\pm0.81$ & $19.2\pm 5.7$ \\ 84.1 & 10.0 & $1.99\pm0.70$ & $11.1\pm 4.4$ \\ 111.8 & 45.0 & $1.62\pm0.39$ & $ 9.0\pm 2.7$ \\ 164.3 & 60.0 & $1.40\pm0.31$ & $ 7.8\pm 2.2$ \\ 214.3 & 40.0 & $1.83\pm0.43$ & $10.1\pm 3.0$ \\ 254.3 & 40.0 & $1.83\pm0.43$ & $10.1\pm 3.0$ \\ 291.8 & 35.0 & $2.25\pm0.51$ & $12.5\pm 3.6$ \\ 334.3 & 50.0 & $1.52\pm0.35$ & $ 8.4\pm 2.5$ \\ 396.8 & 75.0 & $1.01\pm0.24$ & $ 5.6\pm 1.6$ \\ 471.8 & 75.0 & $1.05\pm0.24$ & $ 5.8\pm 1.7$ \\ 561.8 & 105.0 & $0.69\pm0.167$ & $ 3.8\pm 1.1$ \\ 686.8 & 145.0 & $0.50\pm0.12$ & $ 2.76\pm 0.83$ \\ 859.3 & 200.0 & $0.31\pm0.08$ & $ 1.71\pm 0.55$ \\ 4346.7 & 320.0 & $0.069\pm0.021$ & $ 0.38\pm 0.14$ \\ 4856.7 & 700.0 & $0.060\pm0.013$ & $ 0.333\pm 0.092$ \\ 5715.3 & 510.0 & $0.043\pm0.014$ & $ 0.241\pm 0.086$ \\ 6357.8 & 775.0 & $0.027\pm0.009$ & $ 0.149\pm 0.055$ \\ 11249.6 & 2560.0 & $0.012\pm0.003$ & $ 0.065\pm 0.021$ \\ 17040.6 & 2550.0 & $0.0095\pm0.0028$ & $ 0.053\pm 0.018$ \\ 22814.0 & 2575.0 & $0.0069\pm0.0025$ & $ 0.038\pm 0.015$ \\ 31556.0 & 8237.1 & $0.0049\pm0.0015$ & $ 0.0275\pm 0.0096$ \\ 47812.8 & 17585.3 & $0.0033\pm0.0010$ & $ 0.0182\pm 0.0063$ \\ 366425.9 & 515983.6 & $< 0.0004$ & $< 0.00222$ \\ \enddata \tablecomments{Time bin midpoints and durations are relative to the \emph{Swift} trigger time, 18:28:02 UT.} \end{deluxetable} The optical photometry is shown in Table~\ref{tab:opt}. The ROTSE-IIIa images were bias-subtracted and flat-fielded by our automated pipeline. The flat-field image was generated from 30 twilight images. We used SExtractor~\citep{ba96} to perform the initial object detection and to determine the centroid positions of the stars. The images were processed with our custom RPHOT photometry program based on the DAOPHOT PSF-fitting photometry package~\citep{qryaa05}. The unfiltered thinned ROTSE-III CCDs have a peak response similar to an $R$-band filter. The magnitude zero-point for was calculated from the median offset of the fiducial reference stars to the USNO B1.0 $R$-band measurements to produce $C_R$ magnitudes. After the first 30 images, frames were co-added in logarithmic time bins to maintain roughly constant signal-to-noise. \begin{deluxetable}{ccc} \tablewidth{0pt} \tablecaption{ROTSE-IIIc Optical Photometry of GRB~050801.\label{tab:opt}} \tablehead{ \colhead{$t_{\mathrm{start}}$} & \colhead{$t_{\mathrm{end}}$} & \colhead{$C_R$} } \startdata 21.8 & 26.8 & $14.93\pm 0.05$\\ 29.9 & 34.9 & $14.79\pm 0.05$\\ 38.0 & 43.0 & $14.80\pm 0.04$\\ 46.1 & 51.1 & $14.91\pm 0.06$\\ 54.2 & 59.2 & $14.83\pm 0.05$\\ 62.4 & 67.4 & $14.91\pm 0.04$\\ 70.5 & 75.5 & $14.75\pm 0.04$\\ 78.6 & 83.6 & $14.87\pm 0.05$\\ 86.7 & 91.7 & $14.88\pm 0.05$\\ 94.8 & 99.8 & $14.93\pm 0.05$\\ 113.5 & 133.5 & $14.98\pm 0.03$\\ 143.3 & 163.3 & $15.09\pm 0.03$\\ 172.7 & 192.7 & $15.12\pm 0.03$\\ 203.0 & 223.0 & $15.06\pm 0.03$\\ 232.5 & 252.5 & $15.13\pm 0.04$\\ 262.3 & 282.3 & $15.21\pm 0.04$\\ 291.8 & 311.8 & $15.35\pm 0.04$\\ 321.0 & 341.0 & $15.47\pm 0.04$\\ 350.8 & 370.8 & $15.59\pm 0.03$\\ 380.3 & 400.3 & $15.70\pm 0.04$\\ 409.9 & 469.9 & $15.89\pm 0.04$\\ 479.8 & 539.8 & $16.12\pm 0.03$\\ 549.0 & 609.0 & $16.29\pm 0.04$\\ 618.2 & 678.2 & $16.31\pm 0.05$\\ 688.1 & 748.1 & $16.63\pm 0.06$\\ 757.2 & 817.2 & $16.59\pm 0.06$\\ 826.6 & 886.6 & $16.66\pm 0.07$\\ 896.3 & 956.3 & $16.75\pm 0.06$\\ 965.5 & 1025.5 & $16.93\pm 0.07$\\ 1034.9 & 1094.9 & $16.92\pm 0.09$\\ 1104.7 & 1233.9 & $16.99\pm 0.06$\\ 1243.6 & 1442.0 & $17.10\pm 0.05$\\ 1451.4 & 1650.3 & $17.39\pm 0.07$\\ 1659.7 & 1858.6 & $17.48\pm 0.07$\\ 1867.9 & 2136.8 & $17.60\pm 0.06$\\ 2146.5 & 2485.3 & $17.78\pm 0.07$\\ 2495.2 & 2832.6 & $17.88\pm 0.07$\\ 2841.9 & 3249.7 & $18.26\pm 0.11$\\ 3259.7 & 3736.8 & $18.24\pm 0.09$\\ 3745.9 & 4332.1 & $18.71\pm 0.20$\\ 4341.4 & 4956.6 & $18.49\pm 0.09$\\ 4966.5 & 5721.7 & $18.88\pm 0.12$\\ 5731.0 & 6554.7 & $18.99\pm 0.15$\\ 6564.4 & 7527.4 & $18.83\pm 0.13$\\ 7536.7 & 8619.8 & $19.63\pm 0.22$\\ 8629.6 & 10357.0 & $19.49\pm 0.16$\\ \enddata \tablecomments{Start and end times are relative to the \emph{Swift} trigger time, 18:28:02 UT.} \end{deluxetable} \section{Results} With a detection only 21.8 s after the start of the burst, this is the earliest detection of an optical counterpart of a GRB, as well as the most densely sampled early afterglow. Only four GRBs have had optical counterparts detected within the first minute, and none of these had more than two detections in the first minute. The first 250 s of the optical afterglow shows short timescale variability relative to an overall flat lightcurve. This is in stark contrast to the prompt counterpart of GRB~990123~\citep{abbbb99}, which had a very bright $9^{th}$ mag peak at 60~s after the burst onset, generally interpreted as the signature of reverse shock emission~\citep{sp99b}. This afterglow shows no evidence for reverse shock emission. Figure~\ref{fig:optandxraylc} shows a comparison of the early optical and X-ray lightcurves of GRB~050801, combined with the prompt $\gamma$-ray emission. The prompt BAT $\gamma$-ray flux densities have been extrapolated to the X-ray band [0.2-10 keV]. This extrapolation was performed with the best-fit photon index of $2.0\pm0.2$ for the time-averaged $\gamma$-ray spectrum from 20-150 keV, as in \citet{tgcmc05}. The statistical errors scaled from the BAT count rate are shown; the gray region denotes the uncertainties from the extrapolation to the X-ray regime. The X-ray flux values have been converted to flux density (Jy) using an effective frequency of $<\nu> = 6.89\times10^{17}\,\mathrm{Hz}$, the flux weighted average in the 0.2-10 keV range with the best-fit photon index $\Gamma = 1.87$. The ROTSE-III optical magnitudes have been converted to flux density assuming the unfiltered ROTSE-III images are equivalent to $R_c$, and have been approximately adjusted for Galactic extinction by 0.24 mag~\citep{sfd98}. The de-extinction does not have a significant effect on the derived spectral indices. After the break at $\sim250$ seconds, the optical lightcurve decays as $t^{-1.31\pm0.11}$, followed by a brief but significant plateau at $\sim800$ seconds. The top panel shows the ratio of optical flux to X-ray count rate for the first 7000 s, scaled to the average ratio value. The X-ray count rate rather than the X-ray flux was used to avoid the systematic error introduced when converting from count rate to flux, and is made possible by the lack of X-ray spectral evolution. The ROTSE-III observations have been co-added to match the times of the XRT integrations as closely as possible. The flux ratio is consistent with a constant value (dashed-line) with a $\chi^2$ of 15.9 (16 degrees of freedom). The break at $\sim250\,\mathrm{s}$ has no systematic change in the optical to X-ray flux ratio, and is therefore achromatic. Only three $\gamma$-ray bursts have had prompt optical detections contemporaneous with the $\gamma$-ray emission. The prompt optical counterpart of GRB~041219a~\citep{vwwfs05} was correlated with the $\gamma$-ray emission, implying a common origin. However, both GRB~990123~\citep{abbbb99} and GRB~050401~\citep{rykaa05} demonstrated a different origin for the $\gamma$-rays and the optical radiation. Although we do not have a prompt optical detection in the case of GRB~050801, we can interpolate between the high energy prompt lightcurve scaled to the X-ray band (gray band in Figure~\ref{fig:optandxraylc}) and the first X-ray detection. During this interval the high energy emission falls by a factor of $\gtrsim100$ while the optical emission is unchanged. This suggests a different origin for the prompt $\gamma$-ray emission and the early optical emission. However, the X-ray and optical afterglow of GRB~050801 do appear to arise from a similar origin after $\sim80$~s. The two lightcurves are plotted in the main panel of Figure~\ref{fig:optandxraylc}. Each lightcurve shows similar flat behavior at the early time, with a break around 250~s. \section{Discussion} In the standard fireball model of GRB afterglow emission, the spectral energy distribution of GRB afterglows can be fit by a broken power-law with spectral segments $F_\nu \propto \nu^\beta$ (for a review, see \citet{p05}). The spectral index obtained by comparing the de-extincted optical (see Figure~\ref{fig:optandxraylc}) to X-ray flux density during the second XRT integration is $\beta_{\mathrm{opt-X}} = -0.92\pm0.05$, consistent with the X-ray only spectral index of $\beta_{\mathrm{X}} = -0.87\pm0.15$ [0.2-10 keV]. To test for evolution in the broadband spectral index, we have compared the optical and X-ray lightcurves during the first 7000 s (top panel of Figure~\ref{fig:optandxraylc}). The optical to X-ray flux ratio is consistent with a constant value ($\chi^2=15.9$ with 16 degrees of freedom). Across the break at 250 s, both $\beta_{\mathrm{opt-X}}$ and $\beta_{\mathrm{X}}$ are unchanged, and therefore the break is achromatic. Furthermore, there is no evidence of a spectral change in the UVOT images~\citep{bbhgc05}, although the time resolution is insufficient to constrain the time of the break. Many X-ray lightcurves have been seen to steepen around 1000 s - 5000 s post-burst with no change in the X-ray spectral index~\citep{nkgpg05}, For the few bursts with sufficient early optical and X-ray coverage~\citep{qryaa05, bbbbc05}, this behavior has not been mirrored in the optical band. The tight correlation between the optical and X-ray emission suggests that they share the same origin in space and time. The standard fireball model of GRB afterglows can explain the behavior of the optical and X-ray lightcurve after 250~s. The observed spectral parameters and decay indices are most consistent with a fireball expanding adiabatically into a constant density medium, with the typical synchrotron frequency $\nu_\mathrm{m}$ below the optical band, and the cooling frequency $\nu_\mathrm{c}$ above the X-ray band. For example, this can be produced by the following parameters: the electron energy index $p = 2.8$; the isotropic equivalent energy $E \sim 10^{53}\,\mathrm{erg}$ at a redshift of $z \sim 0.5$; the circumburst density $n \sim 0.7\,\mathrm{cm}^{-3}$; the energy fraction in the electrons $\epsilon_e \sim 0.07$; and the energy fraction in the magnetic field $\epsilon_B \sim 0.0002$. These values of the electrons and magnetic energy are consistent with those deduced for other bursts albeit on the lower side. If the ejecta were expanding into a $1/r^2$ density profile (a so-called ``wind'' medium), the fireball model predicts a relationship between the spectral and temporal behavior that is inconsistent at the $4\sigma$ level with the observations after 250~s. We now investigate the possible explanations of the flat early lightcurve and the origin of the break at 250~s. First, any spectral transition (e.g. $\nu_m$ crossing the optical band) would fail to explain the achromatic nature of the break. Achromatic breaks observed in other afterglows have been interpreted as geometric, when the edge of a conical jet becomes visible to the observer and the jet starts to spread~\citep{hbfsk99,sgkpt99}. At 250 s, this would be the earliest such ``jet break'' detected. In the fireball model, the post jet break afterglow is expected to decay as $t^{-p}$, where $p$ is the electron energy index with $N_e \propto E^{-p}$, provided that $p>2$~\citep{sph99}. A hard electron index of $p<2$ predicts a post-jet decay even steeper than $t^{-p}$~\citep{dc01}. Therefore, the observed post-break temporal decay implies $p \le 1.3$ which predicts a significant pre-break decay~\citep{dc01} that is inconsistent with the observed pre-break flatness as well as the observed spectral index $\beta$. Therefore, the achromatic evolution of GRB~050801 cannot be explained with a jet break. We have investigated whether the early afterglow is consistent with the predictions of a structured jet viewed off axis~\citep{gk03}. In this case, it is difficult to create a sharp early break; under such conditions, the post-break evolution should track closely with the electron energy index $p$, which is inconsistent with observations as described above. Such an early break at 250~s, can perhaps be explained as the onset time of the afterglow. If the reverse shock is non relativistic (as indicated by the relatively short duration of the burst, see \citet{sari97}) then self similar expansion starts once the mass collected from the environment is a factor $\gamma$ smaller than that in the ejecta: \begin{equation} t_{{\rm afterglow}}=100\,{\rm s}\,(1+z) \left( \frac{E}{10^{53}\,{\rm erg}} \right)^{1/3} \left( \frac{n}{1\,{\rm cm^{-3}}} \right) \left(\frac{\gamma}{100} \right)^{-8/3}. \end{equation} A value of the initial Lorentz factor $\gamma$ just below a hundred would therefore be consistent with an onset time of 250~s. However, it is difficult to reconcile the flat part before 250~s as the rise of the afterglow. During the onset, since the fireball is coasting with a constant Lorentz factor, the bolometric luminosity is given by $L_\mathrm{B} \propto t^2 n$, the surface area times the density. For a constant density a sharp rise $\propto t^2$ is therefore expected. For a wind-like decreasing density, the lightcurve should be flat as observed. However, as stated before, a wind density profile seems inconsistent with the behavior after 250~s. Continuous energy injection has been suggested as a source of early X-ray light\-curve flat\-tening~\citep{nkgpg05}. This injection could be observed if the initial fireball ejecta had a range of Lorentz factors, with the slower shells catching up with the decelerating afterglow~\citep{rm98,sm00}. However, we require a very steady injection of energy to produce the observed lightcurve, flat for more than a decade in time. If we adopt this explanation, the afterglow must start before our first optical observation, implying an initial Lorentz factor of more than 200, and energy injection rate which is roughly constant over a decade in time, and which shuts off suddenly at 250~s. Flat or very slowly decaying optical lightcurves have been seen in a number of other early afterglows (eg, GRB~030418~\citep{rspaa04}, GRB~050319~\citep{qryaa05}, and GRB~041006~\citep{msmy04,ysr04}). Early X-ray lightcurves detected by \emph{Swift} are typically more complex, with rapidly fading sections and short timescale flares~\citep{nkgpg05}. The early afterglow of GRB~050801, flat in both optical and X-rays, is, so far, unique. It is inconsistent with the standard fireball model for early afterglow emission, unless continuous energy injection is involved. Further \emph{Swift} prompt GRB detections, combined with rapid follow-up by \emph{Swift} and ground-based telescopes, will provide further opportunities to explore the origin of this type of early afterglow behavior. \acknowledgements This work has been supported by NASA grants NNG-04WC41G and NGT5-135, NSF grants AST-0407061, the Australian Research Council, the University of New South Wales, and the University of Michigan. Work performed at LANL is supported through internal LDRD funding. The Palermo work is supported at INAF by funding from ASI on grant number I/R/039/04. Special thanks to Toni Hanke at the H.E.S.S. site. \newcommand{\noopsort}[1]{} \newcommand{\printfirst}[2]{#1} \newcommand{\singleletter}[1]{#1} \newcommand{\switchargs}[2]{#2#1}
Title: Radio continuum and molecular line observations of four bright-rimmed clouds
Abstract: We present the results of radio continuum and molecular line observations conducted using the Mopra millimetre-wave telescope and Australia Telescope Compact Array. These observations reveal the presence of a dense core embedded within each cloud, and the presence of a layer of hot ionised gas coincided with their bright-rims. The ionised gas has electron densities significantly higher than the critical density above which an ionised boundary layer can form and be maintained, strongly supporting the hypothesis that these clouds are being photoionised by the nearby OB star(s). From an evaluation of the pressure balance between the ionised and molecular gas, SFO 58 and SFO 68 are identified as being in a post-pressure balance state, while SFO 75 and SFO 76 are more likely to be in a pre-pressure balance state. We find secondary evidence for the presence of ongoing star formation within SFO 58 and SFO 68, such as molecular outflows, OH, H$_2$O and methanol masers, and identify a potential embedded UC HII region, but find no evidence for any ongoing star formation within SFO 75 and SFO 76. Our results are consistent with the star formation within SFO 58 and SFO 68 having been triggered by the radiatively driven implosion of these clouds.
https://export.arxiv.org/pdf/astro-ph/0601718
\title{Radio continuum and molecular line observations of four bright-rimmed clouds} \author{J. S. Urquhart\inst{1}, M. A. Thompson\inst{2}, L. K. Morgan\inst{3,4} \& Glenn J. White\inst{4,5}} \offprints{J. S. Urquhart: jsu@ast.leeds.ac.uk} \institute{ Department of Physics and Astronomy, University of Leeds, Leeds, LS2 9JT, UK \and School of Physics Astronomy \& Maths, University of Hertfordshire, College Lane, Hatfield, AL10 9AB, UK \and Centre for Astrophysics and Planetary Science, School of Physical Sciences, University of Kent, Canterbury, CT2 7NR, UK \and Green Bank Telescope, P.O. Box 2, Green Bank, WV 24944, USA \and Dept. of Physics \& Astronomy, The Open University, Walton Hall, Milton Keynes, MK7 6AA, UK \and Space Physics Division, Space Science \& Technology Division, CCLRC Rutherford Appleton Laboratory, Chilton, Didcot, Oxfordshire, OX11 0QX, UK } \date{} \abstract{}{To search for evidence of triggered star formation within four bright-rimmed clouds, SFO~58, SFO~68, SFO~75 and SFO~76.}{We present the results of radio continuum and molecular line observations conducted using the Mopra millimetre-wave telescope and Australia Telescope Compact Array. We use the \mbox{$J$=1--0} transitions of $^{12}$CO, $^{13}$CO and C$^{18}$O to trace the distribution of molecular material and to study its kinematics.}{These observations reveal the presence of a dense core ($n_{\rm{H}_2}>10^4$~cm$^{-3}$) embedded within each cloud, and the presence of a layer of hot ionised gas coincided with their bright-rims. The ionised gas has electron densities significantly higher than the critical density ($>$~25 cm$^{-3}$) above which an ionised boundary layer can form and be maintained, strongly supporting the hypothesis that these clouds are being photoionised by the nearby OB star(s). Using a simple pressure-based argument, photoionisation is shown to have a profound effect on the stability of these cores, leaving SFO~58 and SFO~68 on the edge of gravitational stability, and is also likely to have rendered SFO~75 and SFO~76 unstable to gravitational collapse. From an evaluation of the pressure balance between the ionised and molecular gas, SFO~58 and SFO~68 are identified as being in a post-pressure balance state, while SFO~75 and SFO~76 are more likely to be in a pre-pressure balance state. We find secondary evidence for the presence of ongoing star formation within SFO~58 and SFO~68, such as molecular outflows, OH, H$_2$O and methanol masers, and identify a potential embedded UC HII region, but find no evidence for any ongoing star formation within SFO~75 and SFO~76.}{Our results are consistent with the star formation within SFO~58 and SFO~68 having been triggered by the radiatively driven implosion of these clouds.} \keywords{Stars: formation -- ISM: clouds -- ISM: HII regions -- ISM: individual object: bright-rimmed clouds: SFO~58, SFO~68, SFO~75 and SFO~76 -- ISM: molecules -- Radio continuum: ISM} \authorrunning{J. S. Urquhart et al.} \titlerunning{Radio observations of four bright-rimmed clouds} \section{Introduction} From the moment they turn on OB stars begin to drive an ionisation front into the surrounding molecular material, photo-evaporating and dissipating the molecular cloud from which they have formed. The rapidly expanding HII region leads to the formation of a dense shell of neutral gas, swept up in front of the ionisation front. These dense shells, and dense neutral clumps of material, surrounding evolved HII regions have long been suspected to be regions where star formation could have been triggered \mbox{(\citealt{elmegreen1977,sandford1982})}. Bright-Rimmed Clouds (BRCs) are small molecular clouds located on the edges of evolved HII regions and are considered, due to their relatively simple geometry and isolation within HII regions, to be ideal laboratories in which to study the physical processes involved in triggered star formation. The photoionisation of the BRCs surface layers by UV photons from nearby OB stars leads to the formation of a layer of hot ionised gas, known as an \emph{Ionised Boundary Layer} (IBL), which surrounds the rim of the molecular cloud. The hot ionised gas streams off the surface of the cloud into the low density HII regions, resulting in a continuous mass loss by the cloud, known as a \emph{photo-evaporative flow} (\citealt{megeath1997}). Within the IBL the incoming ionising photon flux is balanced by recombination, with only a small fraction of ionising photons penetrating the IBL to ionise new material (\citealt{lefloch1994}), which replenishes the ionised material within the IBL lost to the photo-evaporative flow. If the IBL is over-pressured with respect to the molecular gas within the BRC, shocks are driven into the molecular gas, resulting in the compression of the cloud, and can lead to the formation of dense cores which are then triggered to collapse by the same (or a subsequent) shock front (\citealt{elmegreen1992}). The propagating shock front may also serve to trigger the collapse of pre-existing dense cores, thus leading to the creation of a new generation of stars. This method of triggered star formation is known as \emph{Radiative--Driven Implosion} (RDI) and may be responsible for the production of hundreds of stars in each HII region (\citealt{ogura2002}), and perhaps even contributing up to $\sim 15$ \% of the low-to-intermediate mass IMF (\citealt{sugitani2000}). Shocks continue to be driven into the cloud, compressing the molecular material until the internal density and pressure is balanced with the pressure of the IBL; after which the shock fronts dissipate and the cloud is considered to be in a quasi-steady state known as the cometary stage (\citealt{bertoldi1990,lefloch1994}). Once equilibrium is reached the ionisation front is unable to have any further influence on the internal dynamics of the cloud, but continues to propagate into the cloud photo-evaporating the molecular gas, which streams into the HII region, eroding the cloud and accelerating it radially away from the ionising star via the \emph{rocket effect} (\citealt{oort1955}). The mass loss resulting from photoionisation ultimately leads to the destruction of the cloud on a timescale of several million years. A search of the IRAS point source catalogue, correlated with the Sharpless HII region catalogue (\citealt{sharpless1959}) and the ESO(R) Southern Hemisphere Atlas resulted in a total of 89 BRCs being identified with an associated IRAS point source, 44 in the northern and 45 in the southern sky (\citealt{sugitani1991, sugitani1994}; collectively known as the SFO catalogue). The association of an IRAS point source located within these BRCs suggests that these clouds might contain embedded protostars. Several comprehensive studies of individual BRCs have been reported (e.g. \citealt{lefloch1997, megeath1997,yamaguchi1999,codella2001, devries2002,dobashi2002,thompson2004a}), all of which have confirmed their association with protostellar cores. However, the question of star formation occurring more widely in these objects still remains unclear, and evidence for triggered star formation within the small number of BRCs so far observed remains circumstantial and inconclusive. To address these issues we are currently conducting a complete census of the SFO catalogue; here we report the results of a detailed study of four southern BRCs. In an earlier paper (Thompson, Urquhart \& White, 2004b; hereafter Paper~I) we reported the results of a relatively low angular resolution radio continuum survey of the 45 southern BRCs taken from the SFO catalogue. In that survey we detected radio continuum emission toward eighteen BRCs. In each case the radio emission was found to correlate extremely well with both the morphology of the optical bright rim seen in the DSS~(R band) image and the Photo-Dominated Region (PDR) traced by the Midcourse Space eXperiment (MSX)\footnote{Available from the NASA/IPAC Infrared Processing and Analysis Center and NASA/IPAC Infrared Space Archive both held at http://www.ipac.caltech.edu.} 8~$\mu$m image (\citealt{price2001}), consistent with the hypothesis that these eighteen BRCs are being photoionised by the nearby OB star(s). In this paper we present high resolution molecular line and radio continuum observations toward four clouds (i.e. SFO~58, SFO~68, SFO~75 and SFO~76) from the eighteen photoionised clouds identified in Paper~I which displayed the best correlation between the optical, MIR and radio emission and appeared to possess the simplest geometry. These clouds are thus considered ideal candidates for further investigation into the applicability of the RDI star-formation mechanism. From these observations we investigate the internal and external structure of these clouds and calculate their physical parameters, which when combined with archival data will provide a comprehensive picture of star formation within these BRCs and allow us to determine whether or not it could have been triggered. The structure of this paper is as follows: in Section~2 we briefly describe the morphologies of the BRCs and the HII regions in which they are located, including a summary of any previous observations that have been reported toward them. In Section~3 we describe our observation strategy and the molecular line and radio continuum observations, followed in Section~4 by the observational results and analyses. In Section~5 we discuss the impact the ionisation front has had on the stability, morphology and future evolution of these clouds and try to evaluate the current state of star formation within each cloud and whether it could have been triggered. We present a summary and our conclusions in Section~6. \section{Description of individual BRCs} \label{sect:summary_clouds} In Figure~\ref{fig:dss_brc_images} we present a large scale DSS image of each BRC and the HII region in which they are situated. In these we have indicated the position of the ionising star(s) as identified by \citet{yamaguchi1999} and the IRAS point source with a cross and an ellipse respectively. In these images it would appear that these BRCs are dense condensations of material that are connected to an extended shell of molecular gas which surrounds the HII region that are beginning to protrude into the HII region as the ionisation front ionises the less dense material around them. An arrow has been added from the centre of each bright rim which points directly toward the ionising star(s). These images clearly show the cloud morphologies to be curved in the general direction of the ionising stars; this is especially evident for SFO~58 and SFO~68. The physical parameters relating to each HII region are presented in Table~\ref{tbl:HII_regions}; the distances and spectral types of the ionising star(s) have been taken from \citet{yamaguchi1999}, the age of each HII region has been estimated by calculating their expansion timescale as described by \citet{thompson2004a}. \begin{table*} \begin{center} \caption[HII regions containing selected BRCs]{HII regions containing selected bright-rimmed clouds.} \begin{tabular}{lccccc} \hline \hline HII & Distance & Associated & Ionising & Spectral & HII region \\ region & (kpc)& BRC & star(s)& type & age (Myr)\\ \hline RCW 32 & 0.7 & SFO~58 & HD 74804 & B0 V & 0.26 \\ RCW 62 & 1.7 & SFO~68 & HD 101131 & O6.5 N & 1.20 \\ & & & HD 101205 & O6.5 \\ & & & HD 101436 & O7.5 \\ RCW 98 & 2.8 & SFO~75 & LSS3423 & O9.5 IV &0.31 \\ RCW 105 & 1.8 & SFO~76 & HD 144918 & O7 & 0.73 \\ \hline \end{tabular} \label{tbl:HII_regions} \end{center} \end{table*} \begin{table*}[!ht] \begin{center} \caption[Parameters of the embedded IRAS point sources]{Parameters of the embedded IRAS point sources. (Upper limits are indicated by parenthesis.)} \begin{tabular}{cccccc} \hline \hline Cloud id. &IRAS id. &IRAS colour type & Luminosity (\lsun) & Spectral type & Mass (\msun) \\ \hline SFO~58 & 08435--4105 & hot cirrus & 140 & B7 V & 4.2 \\ SFO~68 & 11332--6258 & class 0/UC HII & (3400) & (B1--B2) &(10.6) \\ SFO~75 & 15519--5430 & hot cirrus & 34000 & O9.5 & 20.6 \\ SFO~76 & 16069--4858 & class 0/UC HII & 5600 & B1 & 12.2 \\ \hline \end{tabular} \label{tbl:IRAS_sources} \end{center} \end{table*} The IRAS point sources can be clearly seen to lie within the cloud, slightly behind the rim with respect to the direction of ionisation. The parameters of the IRAS point sources associated with each BRC are presented in Table~\ref{tbl:IRAS_sources}; the luminosities and classifications of the IRAS colours have been taken from \citet{sugitani1994}, the spectral types of the embedded IRAS sources have been estimated using the tables of \citet{panagia1973} and \citet{de_jager1987} assuming that the IRAS infrared luminosity is due to the presence of a single Zero Age Main Sequence (ZAMS) star, the masses have been estimated using the \mbox{L$_{\rm{star}}$ $\simeq$ M$_{\rm{star}}^{3.45}$} relationship (\citealt{allen1973}). Although the assumption that the infrared luminosity of each IRAS point source is due to a single embedded star is a rather crude approximation, it does enable an upper limit to the spectral type of the embedded star to be estimated. There are a couple of interesting points to note from Table~\ref{tbl:IRAS_sources}. Firstly, three of the four BRCs could possibly harbour Massive Young Stellar Objects (MYSOs), and secondly, two BRCs have IRAS colours consistent with the presence of UC HII regions. \subsection{SFO~58} The BRC SFO~58 is situated on the edge of the HII region RCW~32, which is located at a heliocentric distance of $\sim$~700 pc (\citealt{georgelin1973}). The ionising star of RCW~32 has been identified as HD 74804, a B0 V--B4 II star (\citealt{yamaguchi1999}, and references therein), which is located at a projected distance of 1.53~pc from the bright rim of SFO~58. \subsection{SFO~68} SFO~68 is located on the northwestern edge of RCW~62, a bright HII region located approximately \mbox{1.7 kpc} from the Sun (\citealt{yamaguchi1999}). The HII region is driven by three O stars, HD~101131, HD~101205 and HD~101436, which have the spectral types of O6.5, O6.5 and O7.5 respectively (\citealt{yamaguchi1999}). At \mbox{1.7 kpc} the projected distances between SFO~68 and the three ionising stars range between 7.4--14.9 pc, with HD 101131 providing the vast majority of the ionising flux ($>90$ \%) impinging upon the surface of SFO~68. SFO~68 has IRAS colours consistent with that of an UC~HII region, which has led to SFO~68 being included in several surveys of high-mass star forming regions, such as maser surveys (\citealt{braz1989,macleod1992,caswell1995}), and molecular line surveys (\citealt{zinchenko1995,bronfman1996}). H$_2$O, OH (\citealt{braz1989}) and 6.7~GHz methanol masers (\citealt{macleod1992,caswell1995}) have all been detected toward SFO~68, suggesting the presence of ongoing high-mass star formation within SFO~68. This is supported by the FIR luminosity, from which the presence of a B1--B2~ZAMS star embedded within the cloud can be inferred. The \vlsr~of the detected maser emission range between $-$12 and $-$17~\kms, which is in good agreement with the \vlsr~obtained from our CO observations ($-$16.7~\kms; see Section~4.1). In addition to the detected masers, a survey of candidate UC HII regions conducted by \citet{bronfman1996} detected CS(\emph{J}=2--1) emission toward the IRAS point source embedded within SFO~68. CS is a high density tracer with a critical excitation density threshold between 10$^4$--10$^5$~cm$^{-3}$, and therefore confirms the presence of dense gas coincident with the position of the IRAS point source. This CS emission has a \vlsr~of $-$15.4 \kms; similar to our CO velocity, and the velocities measured from the masers, confirms they are dynamically associated with this cloud. \citet{zinchenko1995} mapped SFO~68 using the \mbox{CS(\emph{J}=2--1)} transition as well as making single pointing observations toward the molecular peak, identified from the CS map, in the C$^{34}$S(\emph{J}=2--1) and $^{12}$CO(\emph{J}=1--0) molecular transition lines. \citet{zinchenko1995} estimated the main beam temperature and mass of the molecular cloud to be $\sim$~24~K and 425~\msun~respectively. \subsection{SFO~75} The bright-rimmed cloud SFO~75 is located on the southeastern edge of the HII region RCW~98. Embedded within SFO~75 is the IRAS point source 15519--5430, which is the most luminous in the SFO catalogue with an FIR luminosity, \emph{L}$_{\rm{FIR}}$ $\sim$ $3.4\times10^{4}$~L$_\odot$. The surface exposed to the HII region is being ionised by LSS 3423, an O9.5~IV star located 0.61~pc to the northwest. RCW~98 lies at a heliocentric distance of 2.8~kpc (\citealt{yamaguchi1999}). \subsection{SFO~76} The exciting star of RCW~105 is HD~144918, an O6 star located at a projected distance of 1.79~pc to the northwest of SFO~76 assuming a heliocentric distance of 1.8~kpc (\citealt{yamaguchi1999}). The IRAS point source 16069--4858 is located very close to the rim of the BRC and has an FIR luminosity of 5600~\lsun, and colours consistent with the presence of an UC HII region, which has led to this source being included in many of the same surveys of high-mass star forming regions as SFO~68. However, unlike SFO~68, searches for H$_2$O, OH (\citealt{braz1989,caswell1995}) and 6.7~GHz methanol masers (\citealt{walsh1997}) resulted in non-detections. CS emission has been detected toward SFO~76 (\citealt{bronfman1996}), confirming the presence of dense molecular gas coincident with the position of the IRAS point source. The FIR luminosity is consistent with the presence of a single B1 ZAMS star, supporting the identification of this cloud as a high-mass star forming region. However, the non-detection of maser emission suggests that either SFO~76 is at an earlier stage of development than SFO~68, or that it is simply in the process of forming a cluster of intermediate-mass stars. \section{Observations and data reduction} \subsection{Survey strategy} Theoretical RDI models \citep{bertoldi1989,bertoldi1990,lefloch1994} suggest that the pressure balance between the IBL and the molecular gas can be used to identify clouds in which star formation may have been triggered, or is likely to be triggered in the future. A comparison of the internal and external pressures can result in three possible scenarios depending on whether the cloud is (1) over-pressured, (2) under-pressured or (3) in approximate pressure balance with the respect to the IBL. The implications of each of these scenarios are as follows: \begin{enumerate} \item Over-pressured clouds: the ionisation front is likely to have stalled at the surface of these clouds where it will remain, unable to overcome the internal pressure of the cloud until the pressure in the IBL increases to match that of the molecular cloud. The ionisation front has no dynamical effect on these clouds and therefore is unlikely to have influenced any current, or future, star formation within these clouds. \item Under-pressured clouds: these clouds are thought to have only recently been exposed to the ionisation front, and although it is highly likely that shocks are currently being driven into these clouds, these shocks have not yet led to the equalisation of the internal and external pressures, leaving the ionisation front stalled at the surface of these clouds. These clouds are thought to be in a \emph{pre-pressure balance state}. Any current star formation present in these clouds is unlikely to have been triggered and is more likely to be pre-existing. \item Approximate pressure balance: these clouds are thought to have been initially under-pressured with respect to the ionisation front, however, shocks have propagated through the surface layers compressing the molecular gas, leading to an equalisation of the internal and external pressures, and leaving a dense core in its wake as it continues toward the rear of the cloud. These clouds are said to be in a \emph{post-pressure balance state}. Models suggest that the star formation within these clouds may have been triggered (\citealt{lefloch1994}). \end{enumerate} In order to evaluate the state of the pressure balance the physical properties of the ionised and molecular gas need to be measured. A combination of radio continuum and molecular line observations have previously been successfully used to determine the current state of several clouds (e.g. \citealt{lefloch1997,white1999,lefloch2002,thompson2004a}). To build on the models, and these previous observational studies, we have made high resolution molecular line and radio continuum observations of four BRCs. \subsection{CO observations} \label{sect:co_observations} Observations of the four BRCs were made during June 2003 in the $J$=1--0 rotational lines of $^{12}$CO, $^{13}$CO and C$^{18}$O using the Mopra millimetre-wave telescope. Mopra is a 22 metre telescope located near Coonabarabran, New South Wales, Australia.\footnote{Mopra is operated by the Australia Telescope National Facility, CSIRO and the University of New South Wales.} The telescope is situated at an elevation of 866 metres above sea level, and at a latitude of 31 degrees south. The receiver is a cryogenically cooled \mbox{($\sim$ 4 K)}, low-noise, Superconductor-Insulator-Superconductor (SIS) junction mixer with a frequency range between 85--116 GHz, corresponding to a half-power beam-width of \mbox{36--33 $\pm$ 2\arcsec} (Mopra Technical Summary version 10).\footnote{Available at http://www.narrabri.atnf.csiro.au/mopra/.} The receiver can be tuned to either single or double side-band mode. The incoming signal is separated into two channels, using a polarisation splitter, each of which can be tuned separately allowing two channels to be observed simultaneously. The receiver backend is a digital autocorrelator capable of providing two simultaneous outputs with an instantaneous bandwidth between 4--256 MHz. For these observations a bandwidth of 64 MHz with a 1024-channel digital autocorrelator was used, giving a frequency resolution of \mbox{62.5 kHz} and a velocity resolution of \mbox{0.16--0.17 km s$^{-1}$} over the \mbox{109--115 GHz} frequency range. For the $^{12}$CO and $^{13}$CO observations the second channel was tuned to \mbox{86.2 GHz} (SiO maser frequency) to allow pointing corrections to be performed during the observations. However, both bands were tuned to \mbox{109.782 GHz} for the C$^{18}$O observations in order to optimise the signal-to-noise ratio. System temperatures were between $\sim 500$--$600$ K for $^{12}$CO and $\sim 250$--$350$~K for both $^{13}$CO and C$^{18}$O depending on weather conditions and telescope elevation, but were found to be stable over the short time periods required to complete each map, varying by no more than approximately 10 \%.\footnote{With the exception of the $^{12}$CO emission observed toward SFO~58 which suffered from large system temperature variation due to poor weather rendering the $^{12}$CO map unreliable. (Only the $^{13}$CO map will be presented for this source.)} Position-switching was used to subtract sky emission. Antenna pointing checks every two hours showed that the average pointing accuracy was better than 10$^{\prime\prime}$ r.m.s.. \begin{table}[!ht] \begin{center} \caption[Summary of Mopra CO observations]{Summary of Mopra CO observations.} \label{tbl:co_observations} \begin{tabular}{lcccc} \hline \hline Isotope & Frequency & Velocity res. & Grid & Integration \\ (\emph{J}=1--0) & (GHz) & (km s$^{-1}$) & size& time (s)\\ \hline $^{12}$CO & 115.271 & 0.162 & $9\times9$ & 30 \\ $^{13}$CO& 110.201 & 0.170 &$9\times9$ & 30\\ C$^{18}$O& 109.782 & 0.170 &$3\times3$ & 120 \\ \hline \end{tabular} \label{tbl:co_line} \end{center} \end{table} The $^{12}$CO and $^{13}$CO observations consisted of spectra taken of a $9\times9$ pixel grid centred on the cometary head of each cloud, using a grid spacing of 15\arcsec. Each grid position was observed for 30 seconds, interleaved with observations at an off-source reference position for 90 seconds after each row of 9 points. The C$^{18}$O maps consist of a smaller grid of $3\times3$ points with the same spacing, and were centred on the molecular peaks identified from the $^{13}$CO maps. For each of the C$^{18}$O grid positions a total integration time of 2 minutes was used. A summary of the observational parameters is presented in Table~\ref{tbl:co_line} with the grid centres and off-source reference positions presented in Table~\ref{tbl:positions}. \begin{table*} \begin{center} \caption[Pointing centres and off-source reference positions]{Pointing centres and off-source reference positions for all four BRCs.} \label{tbl:positions} \begin{tabular}{cccccc} \hline \hline Cloud & IRAS &\multicolumn{2}{c}{Pointing centre} & \multicolumn{2}{c}{Reference position} \\ id.& id.& $\alpha$(2000) & $\delta$(2000) & $\alpha$(2000) & $\delta$(2000) \\ \hline SFO~58 & 08435-4105& 08:45:25.4 & $-$41:16:02 & 08:45:26.1 & $-$41:15:10 \\ SFO~68 & 11332-6258 & 11:35:31.9 & $-$63:14:51 & 11:24:20.5 & $-$64:09:56 \\ SFO~75 & 15519-5430 & 15:55:50.4 & $-$54:38:58 & 16:02:17.6 & $-$55:19:12 \\ SFO~76 & 16069-4858 & 16:10:38.6 & $-$49:05:52 & 16:04:09.3 & $-$48:48:24 \\ \hline \end{tabular} \end{center} \end{table*} The measured antenna temperatures, $T_A^*$, were corrected for atmospheric absorption, ohmic losses and rearward spillover, by taking measurements of an ambient load (assumed to be at 290 K) placed in front of the receiver following the method of \citet{kutner1981}. To correct for forward spillover and scattering, these data are converted to the corrected receiver temperature scale, T$_R^*$, by taking account of the main beam efficiency, \emph{B$_{eff}/F_{eff}$} $\sim0.42\pm0.02.$\footnote{Main beam efficiencies have only been accurately determined at 86, 100 and 115 GHz which have efficiencies of 0.49$\pm0.03$, 0.44$\pm0.03$ and 0.42$\pm0.02$ respectively (\citealt{ladd2005}). Interpolating from these efficiencies it is easy to show that, although the efficiency is probably not the same at 110 GHz (i.e. $^{13}$CO and C$^{18}$O ) to that at 115 GHz, any difference is smaller that the errors involved. We have therefore adopted the 115 GHz efficiency for all three CO lines.} All of the BRCs have angular diameters larger than $\sim$ 80$^{\prime\prime}$, and therefore to take account of the contribution made to the measured intensity from extended material that couples to the error beam, we have used the extended beam efficiency (i.e. $\eta_{xb}\sim0.55$) to correct the $^{12}$CO measurements. Absolute calibration was performed by comparing measured line temperatures of Orion~KL and M17SW to standard values. We estimate the combined calibration uncertainties to be no more than 10 \%. The ATNF data reduction package, SPC, was used to process the individual spectra. Sky-subtracted spectra were obtained by subtracting emission from the off-source reference position from the on-source data. A correction was made to account for the change in the shape of the dish as a function of elevation. The data have been Hanning-smoothed to improve the signal-to-noise ratio, reducing the velocity resolution to 0.32--0.34 \kms. \subsection{Radio observations} Centimetre-wave continuum observations were carried out with the Australia Telescope Compact Array (ATCA)\footnote{The Australia Telescope Compact Array is funded by the Commonwealth of Australia for operation as a National Facility managed by CSIRO.} between June 2002 and September 2003. ATCA is located at the Paul Wild Observatory, Narrabri, New South Wales, Australia. The ATCA consists of 6$\times$22 metre antennas, 5 of which lie on a 3 km east-west railway track with the sixth antenna located 3 km farther west. Each antenna is fitted with a dual feedhorn system allowing simultaneous measurements of two wavelengths, either 20 \& 13 cm or 6 \& 3.6 cm. Additionally, during a recent upgrade, 12 and 3 mm receivers were installed allowing observations at $\sim$ 20 and $\sim$ 95 GHz respectively. The 6/3.6 cm receiver system was used for the observations of SFO~58, SFO~68 and SFO~76, while observations of SFO~75 were carried out using the new 12 mm receivers. A summary of the observational parameters is presented in Table~\ref{tbl:radio_observations}. \begin{table*} \begin{center} \caption[ATCA radio observational parameters]{Observational parameters for the ATCA radio observations.} \begin{tabular}{ccccccc} \hline \hline Cloud& Wavelength&\multicolumn{2}{c}{Phase centre} & Array & Integration& Phase\\ id.&(cm)& $\alpha$(J2000) & $\delta$(J2000)& configurations& time (hrs) & calibrators\\ \hline SFO~58\dotfill&6/3.6& 08:45:25.4 & $-$41:16:02& 750/352/214 &36& 0826--373\\ SFO~68\dotfill&6/3.6& 11:35:31.9 & $-$63:14:51& 750/352/367&18& 1129--58\\ SFO~76\dotfill&6/3.6& 16:10:38.6 & $-$49:05:52& 750/352/367&18& 1613--586 \\ \hline SFO~75\dotfill&1.3& 15:55:50.4 & $-$54:38:58& 352&12& 1613-586 \\ \hline \label{tbl:radio_observations} \end{tabular} \end{center} \end{table*} The 6/3.6 cm observations of SFO~58, SFO~68 and SFO~76 were made at two different frequency bands centred at 4800 and 8309~MHz using bandwidths of 128 and 8~MHz respectively. The second frequency was observed in spectral band mode (using a total of 512 channels, giving a frequency resolution of 15.6~kHz) in order to observe the H92$\alpha$ radio recombination line, however, this proved too faint to be detected toward all three clouds. Each source was observed using three separate configurations over a twelve hour period for SFO~58, and six hours each for SFO~68 and SFO~76 (split into 8$\times$40 minute observations spread over a 12~hour period to optimise \emph{u-v} coverage). Observations of SFO~75 were made using the 12 mm receivers centred at 23569 MHz ($\sim1.3$~cm) and used a bandwidth of 128 MHz. SFO~75 was observed for a total of 12 hours in a single configuration. To correct these data for fluctuations in the phase and amplitude caused by atmospheric and instrumental effects, a phase calibrator was observed for two minutes after approximately every 40 minutes of on-source integration ($\sim$ 15 minutes for SFO~75 due to the atmosphere being less stable at higher frequencies). The primary flux calibrator, 1934--638, was observed once during each set of observations to allow for the absolute calibration of the flux density. To calibrate the bandpass the bright point source 1921--293 was also observed once during each set of observations. The phase centres, array configurations, total integration time and phase calibrators are tabulated in Table~\ref{tbl:radio_observations}. The calibration and reduction of these data were performed using the MIRIAD reduction package (\citealt{sault1995}) following standard ATCA procedures. The data were CLEANed using a robust weighting of 0.5 to obtain the same sensitivity as natural weighting, but with a much improved beam-shape and lower sidelobe contamination, with the exception of the 3.6 cm data for SFO~58 and SFO~68. Using a robust weighting of 0.5 for these two clouds resulted in poorer imaging of the large scale structure of the ionised gas surrounding their bright rims. (The smaller bandwidth used for the 3.6 cm observations results in these images being a factor of three less sensitive than the 6 cm observations.) For the 3.6 cm data for SFO~58 and SFO~68 a robust weighting of 1 was used, resulting in a slight loss of resolution but improved sensitivity. The data obtained from baselines which included the 6th antenna were found to distort the processed images (due to the large gap in \emph{u-v} coverage at intermediate baselines) and so were excluded from the final images. There are three calibration errors that need to be considered: absolute flux calibration, r.m.s pointing errors and the lack of short baselines. The uncertainties introduced by the first two of these are no more than a few percent each. The shortest baseline for all of these radio observations was 31 metres, and therefore flux loss due to short baselines is not considered to be significant. The combined uncertainty is estimated to be no more than $\sim$ 10 \%. \section{Results and analysis} In Figures \ref{fig:co_sfo58}--\ref{fig:co_sfo76} we present plots of the integrated $^{13}$CO (\emph{upper left panel}), $^{12}$CO (\emph{lower left panel}), radio continuum emission (\emph{upper right panel}) contoured over a DSS image of the BRC and surrounding region. These images reveal a strong spatial correlation between the distribution of both the molecular and ionised gas with the optical morphology of the bright rims. The molecular gas traced by the $^{12}$CO and $^{13}$CO contours displays a similar morphology, both tightly correlated within the rim of the cloud and with a steep intensity gradient decreasing toward the HII region. The steep intensity gradient is possibly a consequence of shock compression of the molecular gas which is swept up in front of the expanding ionisation front and accelerated into the cloud, consistent with the predictions of RDI. The $^{13}$CO emission reveals the presence of a dense molecular core embedded within every BRC, set back slightly from the bright rim of the cloud with respect to the direction of the ionising star(s). All of the molecular cores appear to be centrally condensed, which suggests they may be gravitationally bound, or have a gravitationally bound object embedded within them, such as a protostar. There is an interesting difference between the spatial distribution of $^{12}$CO and $^{13}$CO emission toward SFO~75. The peak of the integrated $^{12}$CO emission detected toward SFO~75 is correlated with the position of the bright rim of the cloud, and is slightly elongated in a direction parallel to the morphology of the rim, whereas the integrated $^{13}$CO emission peaks farther back within the molecular cloud and is elongated in a direction perpendicular to the rim. There are a few possible explanations that should be considered: heating of the surface layers of the cloud by the FUV radiation field, the ionisation front is preceeded by a PDR, which could sharpen the $^{12}$CO emission, or that the cloud is angled to the line of sight and that we are seeing the body of the cloud through the bright rim. The morphology of the radio and optical emission seen toward all four clouds shows the rims to be curving directly away from the ionising star(s), starting to form the typical cometary structure seen toward more evolved BRCs (e.g. the Eagle nebula). The presence of radio emission and its tight correlation with the morphology of the rim strongly supports the hypothesis that an IBL is present between the ionising star(s) and the molecular material within these BRCs. In addition to the IBL, the radio emission image of SFO~58, also reveals the presence of a compact radio source within the optical boundary of the cloud, which is coincident with the position of the molecular core identified in $^{13}$CO image, possibly indicating the presence of an UC HII region within this cloud. The radio images were analysed using the visualisation package \emph{kvis} (part of the \emph{karma} image analysis suite (\citealt{gooch1996}). The image parameters and measurements of the peak and source integrated emission are summarised in Table~\ref{tbl:image5}. \begin{table*} \caption[Summary of physical parameters derived from radio observations]{Summary of physical parameters derived from radio observation images.} \begin{center} \small \begin{tabular}{cccccccc} \hline \hline Cloud & & Restoring &Position &Peak & Integrated & Source-averaged & Image \\ id.& $\lambda$ & beam &angle & emission & emission & emission& r.m.s. noise\\ & (cm)& (arcsec)& (degrees)&(mJy/beam) & (mJy) & (mJy/beam)&(mJy/beam) \\ \hline SFO~58 & 3.6 &$27.6\times17.6$ & 2.8 &2.66 &23.1& 1.33 & 0.11\\ & 6 & $21.6\times12.6$ & 1.7 & 2.38 &38.5& 1.44 & 0.10 \\ \hline SFO~68 & 3.6 &$20.9\times15.6$ & 7.5 & 5.22 & 69.3 & 1.63 & 0.17 \\ & 6 & $21.6\times15.6$ & 10.42 & 6.36 & 151.1 & 2.97 & 0.16\\ \hline SFO~76 & 3.6 & $13.3\times10.9$ & 50.1 & 9.84 & 195.4 & 1.74 & 0.25\\ & 6 & $20.6\times17.3$ & 35.4 & 20.82 & 255.7 & 6.99 & 0.30\\ \hline SFO~75 & 1.3 & $11.5\times4.9$ & $-$9.2 & 3.39 & 30.3 & 0.53 & 0.20 \\ \hline \end{tabular} \label{tbl:image5} \end{center} \end{table*} The IRAS point sources within SFO~68, SFO~75 and SFO~76 are located slightly behind the ionisation front, with respect to the ionising star(s), and toward the centre of the bright rim, where one would expect photoionisation induced shocks to focus molecular material, and where the RDI models predict the cores to form (\citealt{lefloch1994}). The positions of the $^{13}$CO peak emission and the IRAS point source seen toward SFO~58 (Figure~\ref{fig:co_sfo58}) are not so well correlated; the IRAS point source is located approximately 1\arcmin~to the south of the molecular peak. It is possible that the IRAS point source is unrelated to the molecular core detected, but is associated with another core which has formed on the edge of the molecular cloud, however, considering the positional correlation of the $^{13}$CO core with the compact radio source and taking account of the pointing errors, size of the IRAS beam ($\sim2^{\prime}$ at 100 $\mu$m), and the positional inaccuracy of the IRAS point source, we do not consider the displacement between the IRAS point source and the centre of the $^{13}$CO core to be significant.\footnote{Comparing the positions of dense cores identified from ammonia maps of the Orion and Cepheus clouds with the positions of their associated IRAS sources \citet{harju1993} found that they could be offset from each other by up to 80$^{\prime\prime}$.} \subsection{Physical parameters of the cores} \label{sect:co_analysis} The angular size of each of the four molecular cores was estimated from the FWHM of azimuthally averaged $^{13}$CO intensity maps. Averaging all spectra within the derived angular size of each core, a source-averaged spectrum was produced for each of the three molecular lines; these spectra are presented in the \emph{lower right panels} of Figures~\ref{fig:co_sfo58}--\ref{fig:co_sfo76}. We note that the baseline for SFO~76 is poor due to the fact that the reference position was found to contain emission at a nearby velocity (i.e. $-$26.5 km s$^{-1}$) to the main line. Gaussian profiles were fitted to the core-averaged spectra of each core to determine the emission peak, FWHM line width and V$_{\rm{LSR}}$ for each spectral line. These values are listed in Table~\ref{tbl:co_data5} along with the central position of each core derived from a 2D Gaussian fit to the $^{13}$CO emission map. The measured \vlsr~of each source was compared to those reported by \citet{yamaguchi1999} and found to agree to better than 2~\kms, with the exception of SFO~76. For this source \citet{yamaguchi1999} reported a V$_{\rm{LSR}}$ of $\sim$ $-$37 \kms~compared to our measurement of $-$23~\kms. The \vlsr~obtained from our CO observations compares well with that reported by \citet{bronfman1996} from CS observations toward SFO~76 (i.e. $-$22.2~\kms). It is therefore unclear why there is such a large disagreement between the \vlsr~reported by \citet{yamaguchi1999} and that reported in this survey for SFO~76. Comparisons between the $^{12}$CO lines and other isotopomers for each core reveal no significant variations in the kinematic velocities of the emission peaks for either SFO~75 or SFO~76. However, inspection of the SFO~58 source-averaged $^{12}$CO spectrum shows evidence of a blue wing component not present in either the $^{13}$CO or C$^{18}$O spectra (this spectrum was best fitted by two Gaussian components). Comparing the source-averaged $^{12}$CO spectrum of SFO~68 to the $^{13}$CO and C$^{18}$O spectra reveals a small shift in velocity, which suggests the presence of a broad blue wing. There are several physical phenomenon that could give rise to these observed line wings such as: another cloud on the line of sight, a signature of shock compressions in the surface layers, or the presence of a protostellar outflow. In order to try to determine the source of these blue wings and to look for kinematic signatures of shocks we produced channel maps and position-velocity diagrams of the $^{12}$CO emission for each cloud. Only the position-velocity diagram produced for SFO~68 shows any evidence of a large scale velocity gradient, revealing the presence of moderate velocity wing components with a FWHM $\sim$ 8~\kms (see Figure \ref{fig:pv_sfo68}); these can either be attributed to a protostellar outflow, or could be tracing the compression/expansion motions of the surface layers of a collapsing cloud. The spatially localised wings seen in the position-velocity diagram are more suggestive of a bipolar outflow, however, our observations do not have either the mapping coverage or the angular resolution necessary to be able to spatially resolve the red and blue outflow lobes that would confirm the presence of an outflow. The nature of the blue wings is at present unclear, however, taking into account the presence of a UC HII region within SFO~58 (see Section \ref{sect:uchii_region}) and the association of SFO~68 with methanol, OH and H$_2$O masers -- both of which are strong indications of ongoing star formation -- we consider the protostellar outflow hypothesis to be the most likely. We must stress that these are only tentative detections and further observations are required to confirm the presence of protostellar outflows in these two sources. \begin{table*} \caption[Physical values derived from CO spectra]{Results of Gaussian fitting of the CO spectra observed toward the molecular cores within BRCs.} \begin{center} \begin{tabular}{ccccccc} \hline \hline Cloud id.&\multicolumn{2}{c}{Core position}&Molecular & V$_{\rm{LSR}}$ & Peak T$_R^*$ &FWHM \\ & $\alpha$ (J2000) & $\delta$ (J2000) &line & (km s$^{-1}$) & (K) & (km s$^{-1}$)\\ \hline SFO~58 &08:45:26 & $-$41:15:10 &$^{12}$CO& 2.1& 11.9& 1.3\\ && && 3.6& 33.7& 1.51\\ &&&$^{13}$CO & 3.4& 17.3 & 1.5 \\ &&&C$^{18}$O & 3.4 & 2.4& 1.2 \\ \hline SFO~68 & 11:35:31 & $-$63:14:31&$^{12}$CO& $-$17.1& 23.0& 4.0\\ &&&$^{13}$CO & $-$16.7 & 10.8& 2.5\\ &&&C$^{18}$O & $-$16.4 & 1.8&2.3\\ \hline SFO~75 & 15:55:49 & $-$54:39:13&$^{12}$CO& $-$37.5 & 29.4& 3.5\\ &&&$^{13}$CO & $-$37.5 & 16.1& 2.6 \\ &&&C$^{18}$O & $-$37.5 & 2.7& 2.3\\ \hline SFO~76 & 16:10:40 & $-$49:06:17&$^{12}$CO& $-$23.3 & 26.4& 2.7\\ &&&$^{13}$CO & $-$23.4 & 10.1& 2.1\\ &&&C$^{18}$O & $-$23.0 & 1.2& 1.5\\ \hline \end{tabular} \label{tbl:co_data5} \end{center} \end{table*} The optically thin transitions of C$^{18}$O, and the moderately optically thick ($\tau$ $<$ 1) $^{13}$CO, were used to determine the optical depth, gas excitation temperature and C$^{18}$O column density following the procedures described by \citet{urquhart2004} using the following equations \begin{equation} \frac{T_{13}}{T_{18}}=\frac{1-{\rm{e}}^{-\tau_{13}}}{1-{\rm{e}}^{-\tau_{18}}} \end{equation} \noindent where $T_{13}$ and $T_{18}$ can be either the corrected antenna or receiver temperatures, and $\tau_{13}$ and $\tau_{18}$ are the optical depths of the $^{13}$CO and C$^{18}$O transitions respectively. The optical depths are related to each other by their abundance ratio such that $\tau_{13}$ = X$\tau_{18}$, where X is the $^{13}$CO/C$^{18}$O abundance ratio. To estimate the $^{13}$CO/C$^{18}$O abundances we first need to estimate $^{12}$C/$^{13}$C abundances for all four clouds. The galactocentric distances for each source lie between 6.5 and 8.5 kpc, which were compared to the $^{12}$C/$^{13}$C gradient measured over the Galactic disk by \citet{langer1990}. This gives a $^{12}$C/$^{13}$C ratio range of between $\sim$ 45-55, and therefore a value of 50 was adopted for the $^{12}$C/$^{13}$C ratio. Assuming the abundance of $^{16}$O/$^{18}$O in all of the sources to be similar to solar system abundances (i.e. $\sim$ 500; \citealt{zinner1996}), gives a $^{13}$CO/C$^{18}$O abundance ratio of 10. The gas excitation temperature was estimated using the optically thin C$^{18}$O line in the following equation, \begin{equation} T_R^*\simeq [T_{\rm{ex}}- T_{\rm{bg}}] \tau_{18} \end{equation} \noindent where $T_R^*$ is the corrected receiver temperature and $T_{\rm{bg}}$ is the background temperature assumed to be $\simeq$ \mbox{2.7 K}. We have assumed the cores are in Local Thermodynamic Equilibrium (LTE) and can therefore be described by a single temperature (i.e. \mbox{$T_{\rm{ex}} = T_{\rm{Kin}} = T$}). The derived core temperature and optical depth are related to the column densities, $N$ (cm$^{-2}$), through the following equation, \begin{equation}\label{eq:column_density} N({\rm{C}}^{18}{\rm{O}})=2.42 \times 10^{14}\tau_{18} \left[ \frac{\Delta v ~T}{1-{\rm{exp}}(-5.27/T)} \right] \end{equation} \noindent where $\Delta v$ is the FWHM of the C$^{18}$O line, $\tau_{18}$ and \emph{T} are as previously defined. To convert the C$^{18}$O column densities to H$_2$ column densities a fractional abundance of (C$^{18}$O/H$_2$) = $1.7\times 10^{-7}$ (\citealt{goldsmith1996}) was assumed. The H$_2$ number density was calculated assuming the cores to be spherical, and that contamination effects from emission along the line of sight can be neglected. The total mass (M$_\odot$) was estimated by multiplying the volume densities of each core by the total volume using, \begin{equation} M_{\rm{core}}=6.187\times10^{25}R^3n_{\rm{H_2}}\mu m_{\rm{H}} \label{eq:mass} \end{equation} \noindent where \emph{R} is the radius of the core (pc), \emph{n$_{\rm{H_2}}$} is the molecular hydrogen number density (cm$^{-3}$), $\mu$~is the mean molecular weight (taken to be 2.3, assuming a 25 \% abundance of helium by mass), and \emph{m$_{\rm{H}}$} is the mass of a hydrogen atom. The physical parameters for the molecular cores calculated from the $^{13}$CO and C$^{18}$O data are summarised in Table~\ref{tbl:co_summary5}. We estimate the uncertainties involved in the estimate of the column density to be no more than 50~\%. In estimating the density we have considered the additional uncertainties in the distance to the cores and in the assumptions that the cores are spherical. Taking these additional uncertainties into account we estimate the mass and density calculated to be accurate to within a factor of two. \begin{table*} \caption[Summary of parameters derived from $^{13}$CO and C$^{18}$O data]{Summary of parameters derived from $^{13}$CO and C$^{18}$O data.} \begin{center} \begin{tabular}{ccccccccc} \hline \hline Cloud & \vlsr &$\tau_{18}$ & \emph{T} & Log(N$_{\rm{H_2}}$) & Log(n$_{\rm{H_2}}$) & Angular & Physical & Mass \\ id. & (\kms) & & (K) & (cm$^{-2}$) & (cm$^{-3}$) & size (\arcsec) & diameter (pc) & (M$_{\odot}$) \\ \hline SFO~58 &3.4 & 0.07 & 37.0 & 22.53 & 4.89 & 47 & 0.16 & 9.5 \\ SFO~68 & $-$16.7&0.06 & 29.4 & 22.51 & 4.36 & 68 & 0.46 & 66.4 \\ SFO~75 & $-$37.5&0.13 & 23.5 & 22.75 & 4.53 & 41 & 0.56 & 177.3\\ SFO~76 &$-$23.3 &0.05 & 26.7 & 22.26 & 4.06 & 59 & 0.52 & 48.1 \\ \hline \end{tabular} \label{tbl:co_summary5} \end{center} \end{table*} The physical parameters of the cores are similar, with the densities ranging between $\sim$ 10$^4$--10$^5$~cm$^{-3}$, the physical diameters varying between $\sim$ 0.2--0.6~pc, kinetic temperatures ranging from $\sim$~24 to 37~K, and masses from $\sim$~10--180~M$_{\odot}$. The temperature of all of the cores are significantly higher than would be expected for starless cores \mbox{(\emph{T} $\sim$ 10 K}; \citealt{evans1999}), which suggests these cores possess an internal heating mechanism, possibly a YSO, or an UC HII region (as indicated by the IRAS colours; see Table~\ref{tbl:IRAS_sources}). It is possible these cores are being heated by the surrounding FUV radiation field, however, if that were the case, we would expect to find a temperature gradient that peaked at the position of the bright rim and decreased with distance into the cloud, but this is not observed. \subsection{Physical parameters of the ionised boundary layers} \label{sect:IBL} \label{sect:radio_analysis} Using the radio emission detected toward the rims of SFO~58, SFO~68, SFO~75 and SFO~76 we can quantify the ionising photon flux impinging upon them. Making the assumption that all of the ionising photon flux is absorbed within the IBL we can determine the photon flux, $\Phi$ (cm$^{-2}$ s$^{-1}$), and the electron density, n$_{e}$ (cm$^{-3}$), using the following equations which have been modified from Equations~2 and 6 of \citet{lefloch1997} (see Paper~I for details): \begin{equation} \Phi=1.24\times10^{10}S_{\nu}T_e^{0.35}\nu^{0.1}\theta^{-2} \end{equation} \begin{equation} n_{e}=122.21\times\sqrt{\frac{S_{\nu}T_e^{0.35}\nu^{0.1}}{\eta R\theta^2}} \end{equation} \noindent where \emph{$S_{\nu}$} is the integrated flux density in mJy, $\nu$ is the frequency at which the integrated flux density is evaluated in GHz, $\theta$ is the angular diameter over which the flux density is integrated in arc-seconds, $\eta$R is the shell thickness in pc, and \emph{T$_{e}$} is the electron temperature in K. Note, an average HII~region electron temperature of \mbox{$\sim$ 10$^{4}$ K} and \mbox{$\eta$ = 0.2} (\citealt{bertoldi1989}) have been assumed. The values calculated for the ionising photon fluxes, and electron densities within the IBLs of the four clouds, are presented in Table~\ref{tbl:physical_parameters}. The main uncertainties in these values are due to flux calibration ($\sim10$~\%), the approximation of the electron temperature, $T_e\simeq10^4$~K (e.g. a difference in the electron temperature of 2000~K corresponds to an uncertainly in the photon flux of $\sim$~25~\%), and the $\eta=0.2$. The uncertainties in flux calibration and electron temperature combine to give a total uncertainty in the calculated photon fluxes and electron densities of no more than 30~\%. However, approximating $\eta=0.2$ leads to the the electron density being underestimated by at most a factor of $\sqrt{2}$, and since the uncertainty in $\eta=0.2$ dominates the others it effectively sets a lower limit for the electron density. \begin{table*}[!hbt] \caption[Summary of derived physical parameters of the IBL]{Summary of derived physical parameters of the IBL.} \begin{center} \begin{tabular}{cccccc} \hline \hline Cloud & \multicolumn{3}{c}{Photon fluxes (10$^8$ cm$^{-2}$~s$^{-1}$)} & \multicolumn{2}{c}{Electron densities $n_e$(cm$^{-3}$)}\\ id.& Predicted $\Phi_P$ & Peak $\Phi$ & Mean $\Phi$ & Peak & Mean \\ \hline SFO~58 & 20.4 & 31.9 & 19.3 & 338& 262 \\ SFO~68& 126.0 & 68.5 & 32.0 & 341 & 233 \\ SFO~75 & 448.0 & 257 & 40.0 & 839 & 332 \\ SFO~76 & 441 & 263.1& 46.5 & 1242& 526 \\ \hline \end{tabular} \label{tbl:physical_parameters} \end{center} \end{table*} Predicted ionising fluxes were calculated from the Lyman flux of the candidate ionising stars and their projected distances to the clouds following the method described in \mbox{Paper I}. Comparing the predicted fluxes to the measured fluxes, and taking into account any attenuation between the OB star(s) and the clouds, we find good agreement (to within a factor of two). In the majority of cases the measured fluxes at the surface of the BRCs are lower than the predicted flux, which is to be expected, as the predicted flux is a strict upper limit. The one notable exception to this is SFO~58 where the measured flux is a factor of one and half times greater than the predicted upper limit. There are two possible explanations for this discrepancy: the spectral type of the ionising star is incorrect, or the distance to the HII region is incorrect. We favour the former as a misclassification of the ionising star's spectral type by even half a spectral class can alter its predicted Lyman flux by up to a factor of two, which would more than account for the difference in fluxes reported here. The high resolution radio observations result in a much tighter correlation between the predicted and measured fluxes than were found with the low angular resolution data presented in Paper I. The variation between measured and predicted fluxes for the low resolution data for these four clouds ranged from a factor of a few to more than ten (in the case of SFO~58). Moreover, the analysis of the distribution of radio emission suggests that two sources (SFO~68 and SFO~76) lie in the foreground relative to the locations of the ionising star(s), and thus are located farther from the ionising star(s) than the projected distance, used to derive the predicted flux, would suggest. In these two cases the correlation could be considerably better than the factor of two quoted above. The main reason for the improved correlation between the predicted and measured fluxes is because the higher resolution observations have been able to resolve the radio emission, resulting in the detected emission being much more tightly peaked. This has allowed more accurate measurements of the flux density to be obtained, and consequently more realistic values for the ionising fluxes and electron densities to be calculated. Values calculated from the low resolution observations presented in Paper I suffer due to the large size of the synthesised beams (typically $\sim$ 90\arcsec~ and $\sim$ 60\arcsec~ for the 20~cm + 6~cm and 13~cm + 3~cm observations respectively) which dilutes the emission if the IBL is not resolved, resulting in significantly lower flux densities being measured. This explains why, in every case, the high resolution radio observations have resulted in an increase in the calculated ionising fluxes. Another reason is the vast improvement in the \emph{u-v} coverage obtained by using multiple array configurations, which allows the brightness distribution of the emission to be more accurately deconvolved from the visibility data. Therefore, although low resolution radio observations may be useful in identifying clouds that are likely to possess an IBL and are thus subject to photoionisation from the nearby OB star(s), their main use is to limited to the determination of global estimates for the physical parameters of the IBLs. The mean electron densities calculated for each cloud range between 233--526 cm$^{-3}$, considerably greater than the critical value of \emph{n}$_{\rm{e}}\sim$ 25 cm$^{-3}$ above which an IBL is able to develop around the cloud (\citealt{lefloch1994}). The excellent correlation of the radio emission with the bright-rim of the clouds strongly supports the presence of an IBL at the surface of each of these clouds, confirming their identification as potential triggered star forming regions. It is therefore clear that these clouds are being photoionised by the nearby OB star(s), however, it is not yet clear to what extent the ionisation has influenced the evolution of these clouds, and what part, if any, it has played a part in triggering star formation within these clouds. \subsection{Evaluation of the pressure balance} \label{sect:pressure_balance} \label{sect:implications_pressure_balance} In this section the results of the molecular line and radio observations will be used to evaluate the pressure balance between the hot ionised gas of the IBLs, and the cooler neutral gas within the BRCs. Following the method described in Paper~I we calculated the internal ($P_{\rm{int}}$) and external ($P_{\rm{ext}}$) pressures (N cm$^{-2}$) using, \begin{equation} P_{\rm{int}}\simeq\sigma^{2}\rho_{\rm{int}} \label{eq:internal_pressure} \end{equation} \begin{equation} P_{\rm{ext}}=2\rho_{\rm{ext}}c^2 \end{equation} \noindent where $\sigma^2$ is the square of the velocity dispersion (i.e. $\sigma^2=\langle\Delta v\rangle^2/(8\rm{ln}2$), where $\Delta v$ is the core-averaged C$^{18}$O line width (\kms)), $\rho_{\rm{int}}$ is the core-averaged density calculated in Section~\ref{sect:co_analysis}, $\rho_{\rm{ext}}$ and $c$ are the ionised gas density and sound speed (assumed to be $\sim$ 11.4 \kms) respectively. The external pressure term includes contributions from both thermal and ram pressure. The electron densities calculated in Section~\ref{sect:IBL} were used to estimate the ionised gas pressures for each cloud's IBL. The calculated internal and external pressures are presented in Table~\ref{tbl:pressure_balance}. \begin{table} \caption[Summary of the pressure balance analysis]{Evaluation of the pressure balance.} \begin{center} \label{tbl:pressure_balance} \begin{tabular}{ccc} \hline \hline Cloud & \multicolumn{2}{c}{Pressure ($P/k_B$) ($10^6$ cm$^{-3}$ K)}\\ id.& Internal ($P_{\rm{int}}$)& External ($P_{\rm{ext}}$)\\ \hline SFO~58 & 5.3 & 7.8\\ SFO~68 & 4.4 & 7.0 \\ SFO~75 & 9.1 & 26.5\\ SFO~76 & 1.2 & 15.7\\ \hline \end{tabular} \end{center} \end{table} The largest two uncertainties in the calculation of the molecular pressure are: the uncertainty in the observed line temperature, and the possibility that the C$^{18}$O derived density may be affected by depletion, either onto dust grain ice mantles, or through selective photo-dissociation. The uncertainties in the densities are thought to be no more than a factor of two, the effects of depletion and photo-dissociation are harder to quantify. However, all of the core temperatures are considerably larger than 10 K, where depletion is expected to be greatest, and given that they are embedded within the clouds, away from the ionisation front, where they are shielded from much of the ionising radiation, these effects are not thought to be significant. We estimate the molecular pressures presented to be accurate to within a factor of two. The IBL pressures are lower limits (due to the electron densities being lower limits) and taking account of the uncertainties are considered to be accurate to $\sim$ 30~\%. As discussed in Section~3.1 theoretical models (\citealt{bertoldi1989,lefloch1994}) have revealed the pressure balance to be a sensitive diagnostic that can be used to determine the evolutionary state of BRCs. Comparing the internal and external pressures calculated from the molecular line and radio observations (presented in Table~\ref{tbl:pressure_balance}), reveals that all of the clouds are under pressured with respect to their IBLs, and are thus in the process of having shocks driven into them. However, taking account of the possible factor of two uncertainty in our calculations of these parameters, it is possible that two of these clouds are in approximate pressure balance (i.e. SFO~58 and SFO~68); these clouds are likely to be in a post-pressure balance state, and it is therefore possible that any current, or imminent, star formation within these clouds could have been triggered. The remaining two clouds, SFO~75 and SFO~76, are under-pressured by factors of three and twelve respectively, with respect to their IBLs, strongly suggesting that these clouds have only recently been exposed to the HII region and that shocks are currently being driven into the surface layers of these clouds, closely followed by a D-critical ionisation front. These clouds are likely to be in a pre-pressure balance state where the shocks have not propagated very far into the surface layers; it is therefore unlikely that the molecular cores within these two clouds have been formed by RDI, but are more likely to pre-date the arrival of the ionisation front and have only recently been exposed to the HII region. Any current star formation taking place within these clouds is unlikely to have been triggered. \subsection{Compact radio source associated with SFO~58} \label{sect:uchii_region} The 6 cm radio continuum image (see \emph{upper left panel} Figure~3) of SFO~58 clearly shows the presence of a radio source positionally coincident with the $^{13}$CO core embedded within this BRC, both of which lie at the focus of the BRC where the photoionisation induced shock is expected to concentrate the majority of the mass within the cloud. The radio source and embedded $^{13}$CO core are offset from the IRAS point source, but their positional correlation hints at a possible association between the two. At a distance of 700~pc the angular size of the compact radio source ($\sim$~18\arcsec) corresponds to a physical diameter of $<$~0.06~pc, which suggested it might be a compact HII region similar to those found within other BRCs reported in \mbox{Paper I} (i.e. SFO~59, SFO~62, SFO~74, SFO~79 and SFO~85; see \citealt{urquhart2004} for a detailed investigation of SFO~79). The radio flux of the compact radio source is consistent with the presence of a single embedded B2--B3 ZAMS star. Furthermore, the correlation of the position of the radio source with that of the molecular core detected in the CO observations, both of which are located at the focus of the bright rim (see Figure~\ref{fig:co_sfo58}), offer some circumstantial support for this hypothesis. However, it is possible that the presence of the radio source is an unfortunate alignment of the cloud with an extragalactic background source. To try to determine the nature of this radio emission we attempted to calculate the spectral index ($\alpha$) of the emission using the integrated flux at both frequencies (i.e. $S_\nu \propto \nu^{\alpha}$), however, this proved inconclusive due to the poor sensitivity of the 3.6 cm map. We therefore tentatively identify this compact radio source as a possible compact HII region embedded within SFO~58. \begin{table} \caption[Physical parameters of CRS 4]{Derived parameters of the compact radio source associated with SFO~58.} \begin{center} \begin{tabular}{lc} \hline \hline Position\dotfill & $\alpha$(J2000) = 08:45:26 \\ \dotfill&$\delta$(J2000) = $-$41:15:06\\ Source size\dotfill & $<$ 18\arcsec \\ Physical diameter & $<0.06$ pc \\ 3.6 cm flux density\dotfill & 1.69 mJy\\ 6 cm flux density\dotfill & 2.28 mJy\\ log(\emph{N}$_i$)\dotfill & 44.0 photon s$^{-1}$\\ Spectral type\dotfill & ZAMS B2--B3 \\ \hline \end{tabular} \label{tbl:embedded_source} \end{center} \end{table} \section{Discussion} Whilst a full hydrodynamic analysis is beyond the scope of this paper, an insight into the potential effect that exposure to the FUV radiation field has had upon the stability of the cores can be gained from a simple static analysis. Additionally, we will estimate the lifetime of these clouds in the light of the continued mass loss through photo-evaporation by the nearby OB star(s), and evaluate its effect on future star formation within these clouds. \subsection{Gravitational stability of the cores pre- and post-exposure of the clouds to UV radiation} \label{sect:stability} To investigate the effect that the arrival of the ionisation front has on the stability of the cores we need to compare the stability of the cores while they were still embedded within their parental molecular cloud to that of the cores once exposed to the FUV radiation field. In the following analysis we implicitly assume that the cores pre-date the arrival of the ionisation front (as shown in the previous section, this is certainly likely for the cores embedded with SFO~75 and SFO~76), and that there was negligible external pressure from the surrounding molecular material. The pre-exposure stability of the cores can be estimated using the standard virial equation to derive the virial mass, $M_{\rm{vir}}$ (\msun). Comparing these masses with the core masses calculated from the CO data will give an indication of their pre-exposure stability. The virial mass can be calculated using the standard equation (e.g. \citealt{evans1999}), \begin{equation} M_{\rm{vir}}\simeq210R_{\rm{core}} \langle\Delta v\rangle^2 \end{equation} \noindent where \emph{R}$_{\rm{core}}$ is the core radius (pc) measured from the integrated $^{13}$CO maps, $\Delta v$ is the FWHM line width of the C$^{18}$O line (\kms). The results are presented in Table~\ref{tbl:virial_mass} with the core masses derived from the CO observations. Comparing the calculated masses of each core with their virial mass, it is clear that three cores were gravitationally stable against collapse while still embedded within their parental molecular cloud, however, taking the errors into account it is possible that SFO~75 was close to being unstable to gravitational collapse. Now the effect of exposure to the FUV radiation field of the HII region will be examined. This will be estimated using a modified version of the virial mass (i.e.~the Bonnor Ebert approach) that takes account of the external pressure of the surrounding medium, in this case the pressure of the IBL. Following the notation of \citet{thompson2004a}, this pressure-sensitive virial mass will be referred to as the \emph{pressurised virial mass}, $M_{\rm{pv}}$ (\msun), given as, \\ \begin{equation} M_{\rm{pv}}\simeq5.8\times10^{-2}\frac{\langle \Delta v\rangle^4}{G^{3/2}P_{\rm{ext}}^{1/2}} \end{equation} \\ \noindent where $P_{\rm{ext}}$ (N m$^{-2}$) is the pressure of the ionised gas within the IBL, \emph{G} is the gravitational constant, and $\Delta v$ is in \kms. The calculated values for the pressurised virial masses are presented in Table~\ref{tbl:virial_mass}. This equation is sensitive to the accuracy of the measured line width, and taking into account the errors involved with Gaussian fits to the spectral lines, the calculated values for the pressurised virial masses are considered to be accurate to within a factor of two (\citealt{thompson2004a}). In this case it was assumed that the external pressure acts over the entire surface of the core, not just the side illuminated by the OB star, and that the cores were pre-existing cores recently uncovered by the expansion of the ionisation front. This is a rather simplistic approach but does allow the effect that exposure to the FUV radiation has on the stability of the cores to be investigated. \begin{table} \caption[Summary of core masses: physical, virial and pressurised virial masses]{Summary of core masses as derived from the CO data as well as the virial masses, $M_{\rm{vir}}$, and pressurised virial masses, $M_{\rm{pv}}$.} \begin{center} \small \begin{tabular}{cccccc} \hline \hline Cloud& $M_{\rm{co}}$ & $M_{\rm{vir}}$ & $M_{\rm{pv}}$ & $\dot{M}$ & Lifetime \\ id.& (M$_\odot$) &(M$_\odot$) &(M$_\odot$) & (M$_\odot$ Myr$^{-1}$) & (Myr)\\ \hline SFO~58 & 9.5 & 24.2 & 10.7 & 14.9 & 0.64\\ SFO~68 & 66.4 & 255.5 & 152.2 & 58.6& 1.13\\ SFO~75 & 177.3 & 311.1 & 78.2 & 64.6& 2.75\\ SFO~76 & 48.1 & 122.8 & 18.4 & 12.6& 3.80\\ \hline \end{tabular} \label{tbl:virial_mass} \end{center} \end{table} Comparing the values for the virial and pressurised virial masses shows that exposure to the FUV radiation field dramatically reduces the mass above which the clouds become unstable against gravitational collapse. The difference between the virial and pressurised viral masses range from a factor of $\sim$~2 (i.e. SFO~58 and SFO~68) to a factor of $\sim$ 7 for SFO~76. Taking account of the errors involved in this analysis, only differences of a factor of four or larger are significant. We are therefore unable to determine if exposure to the FUV radiation has had an impact on the stability of SFO~58 or SFO~68, however, it is clear that the exposure is likely to have had a significant impact on the stability of SFO~75 and SFO~76. Moreover, comparing the pressurised virial masses of the cores with those estimated from the CO data reveals that SFO~75 and SFO~76 are both more than a factor of two more massive than the critical pressurised viral mass. It is therefore likely that the exposure of these two cores to their respective HII regions has rendered them unstable against gravitational collapse. However, detailed hydrodynamical or radiative transfer modelling (e.g. \citealt{thompson-white2004,deVries2005}) of higher signal-to-noise molecular line data are needed to investigate the presence of collapse motions in these clouds. \subsection{Eventual fate of the BRCs} Once a cloud has reached the cometary stage the shocks dissipate, however, the cloud's mass continues to be slowly eroded away as the D-type ionisation front continues to propagate into it (\citealt{lefloch1994}). In this situation the propagation of the ionisation front leads to a constant mass loss in the form of a photo-evaporated flow into the HII region (\citealt{megeath1997}). The amount of material within the boundary of a cloud is finite, and thus the effect of the ionisation is to slowly erode the limited reservoir of material available for star formation. This mass loss eventually results in the total ionisation and dispersion of the cloud, and perhaps the disruption of ongoing star formation either by disrupting any molecular cores before the accretion phase has begun, or by exposing the protostar to the FUV radiation field before the accretion phase has finished. Once the protostar has been exposed much of the surrounding envelope of molecular material becomes ionised, therefore limiting the possible size of the forming protostar (see \citealt{whitworth2004}). Therefore the mass loss rate is an important parameter that can help determine the effect photoionisation has on current, and future star formation within BRCs, and in estimating their lifetime. To evaluate the mass loss we use Equation 36 from \citet{lefloch1994} which relates the mass loss (M$_\odot$~$\rm{Myr}^{-1}$) to the ionising flux illuminating the cloud ($\Phi$ photons cm$^{-2}$ s$^{-1}$), i.e. \begin{equation} \dot{M}=4.4\times10^{-3}\Phi^{1/2}R_{\rm{Cloud}}^{3/2} \label{eqn:mass_loss} \end{equation} The globally averaged photon flux calculated in Section~\ref{sect:radio_analysis} and the cloud radii presented in Paper I were used to estimate the mass loss rate for each cloud using Equation~\ref{eqn:mass_loss}; these values are presented in Table~\ref{tbl:virial_mass} along with an estimate for the lifetime of each cloud. The BRC mass loss rates range between $\sim$ 12--59 $M_\odot$ Myr$^{-1}$, corresponding to cloud lifetimes from as little as $6\times10^5$ yr to several Myr. The accretion phase of protostar formation is known to last for several \mbox{10$^5$ yr} (\citealt{andre2000}). Therefore any ongoing, or imminent, star formation within SFO~58, SFO~68, SFO~75 and SFO~76 will be unaffected by the ionisation and mass loss, especially SFO~68 where the star formation already appears to be well developed and unlikely to be affected by the mass loss experienced by the cloud. However, the future star formation within SFO~58 may be adversely affected as the ionisation front propagates into the cloud. Although there is no evidence of any current star formation taking place within either SFO~75 or SFO~76, we have shown these clouds are likely to be undergoing RDI as well as having sufficiently long lifetimes for RDI to be a viable method of triggered star formation. \subsection{Star formation and the evolution of the BRCs} Direct evidence of whether star formation within these clouds has been triggered is not readily available, however, it is possible to investigate the probability that the star formation has been triggered by considering the circumstantial evidence. It is interesting to note that there is strong evidence for the presence of ongoing star formation within the two clouds (SFO~58 and SFO~68) that fall into the post-pressure balance cloud category, such as the molecular outflows (see Section~\ref{sect:co_analysis}), association with OH, H$_2$O and methanol masers (\citealt{braz1989,macleod1992,caswell1995}, i.e. SFO~68) and the possible association with an embedded UC HII region (Section~\ref{sect:uchii_region}, i.e. SFO~58). Moreover, the association of the UC~HII region with SFO~58 and of H$_2$O and methanol masers with SFO~68 --- which are respectively and almost exclusively associated with Class 0 protostars (\citealt{furuya2001}) and high-mass star formation (\citealt{minier2003}) --- lead us to conclude that the star formation within these two clouds is relatively recent, being no more than a few 10$^5$ years old. Contrary to the evidence of recent high-mass star formation within the post-pressure balanced clouds we find no evidence for any ongoing star formation within either of the two pre-pressure balance clouds (SFO~75 and SFO~76), short of the presence of the embedded IRAS point source. If, as suggested by the RDI models (e.g. \citealt{lefloch1994, vanhala1998}) and observations (e.g. \citealt{sugitani1991}), that rim morphologies represent an evolutionary sequence (see Figure~\ref{fig:rim_classification}), we should expect to find clouds at similar stages of evolution to exhibit similar physical parameters, and furthermore, the star formation within clouds at different evolutionary states to be at different stages of development. It is therefore useful to compare the observational results to the evolutionary sequence predicted by the models to investigate any differences in the star formation within clouds with different rim morphologies. However, we must point out that the boundary conditions of where the clouds meet the larger-scale molecular material are very important and may affect the following analysis. To emphasise the morphology of each rim a black curved line has been fitted (by eye) to the radio contours following the minimum gradient of the emission (see \emph{upper right panel} of Figures~\ref{fig:co_sfo58}--\ref{fig:co_sfo76}). The four clouds separate quite nicely into two morphological groups after comparison of the rim morphologies presented in Figure~\ref{fig:rim_classification}. SFO~58 and SFO~68 are typical of a type A rim morphology, whereas SFO~75 and SFO~76 show only slight curvature intermediate between type A and type 0. Comparing these rim morphologies with those predicted by the RDI models of \citet{lefloch1994} we find that SFO~75 and SFO~76 closely resemble the 0.036 My and SFO~58 and SFO~68 closely resemble the 0.126 My snapshots (i.e. Figure 4a and b of \citet{lefloch1994}). This comparison should be viewed with caution as the Lefloch \& Lazareff models were calculated for a simple cloud with specific ionising fluxes and so the absolute timescales are more than likely invalid for our ensemble of BRCs. However the comparison between individual clouds and their relative model ages does support our conclusion that both SFO~75 and SFO~76 are in the early stages of ionisation, having only recently been exposed to the ionising radiation of their HII region, and that SFO~58 and SFO~68 have been exposed for a significantly longer period of time. Moreover, we find that the evolutionary age of the SFO~58 and SFO~68 suggested by the models compares well with the age of the star formation indicator mentioned in an earlier paragraph (i.e. $\sim$ several 10$^5$ yr). Although we have not been able to conclusively prove that the star formation has been triggered within SFO~58 and SFO~68, we have shown that is possible, if not likely. Furthermore, we have shown there are clearly significant morphological and star formation differences between the post- and pre-pressure balance clouds in this survey, which are consistent with the predictions of RDI models. From the observations of the four clouds presented here, it seems that there is reasonably good evidence to support the suggestion that the schematic presented in the Figure~\ref{fig:rim_classification} is an evolutionary picture of the changing morphology of BRCs under the influence of photoionisation. \section{Summary and conclusions} In this paper the results of a detailed investigation of four BRCs (SFO~58, SFO~68, SFO~75 and SFO~76) are presented, including high resolution radio molecular line and continuum observations obtained with the Mopra millimetre telescope and the ATCA. The main aim is to distinguish between pre- and post-pressure balance clouds, and to evaluate to what extent the star formation within these BRCs has been influenced by the photoionisation from the nearby OB star(s). Each of the BRCs was mapped using the $^{12}$CO, $^{13}$CO and C$^{18}$O rotational transitions using the Mopra telescope. To complement the molecular line observations, high resolution radio continuum maps of all four BRCs were obtained using the ATCA. The CO observations reveal the presence of a dense molecular core within every BRC, located behind the bright rim and, in most cases, coincident with the position of the IRAS point sources. The distribution of the $^{13}$CO emission maps suggest that the cores have a spherical structure and are centrally condensed, consistent with the presence of an embedded gravitationally bound object, such as a YSO. The H$_2$ number densities and cores masses range between 3--$8\times10^4$~cm$^{-3}$ and 10--180~\msun~respectively. The core temperatures ($\sim$ 30~K) are significantly higher than expected for starless cores ($\sim$ 10~K, \citealt{clemens1991}), supporting evidence for the presence of an internal heating source, such as a protostar. The high angular resolution radio observations have confirmed the presence of an IBL surrounding the rim of all four clouds, with the detected emission displaying excellent correlation with the morphology of the cloud rim seen in the optical images. The increased resolution and improved \emph{u-v} coverage of the radio observations has resulted in significantly higher flux densities being measured, which in turn has led to higher values for the ionising fluxes, which are now more in line with the predicted fluxes than the low angular resolution observations presented in Paper I. The electron densities are all significantly higher than the critical density of 25~cm$^{-3}$ (\citealt{lefloch1994}) above which an IBL can form and be maintained. All these facts strongly support the hypothesis that these clouds are being photoionised by the nearby OB star(s). CO and radio continuum data are used to evaluate the pressure balance at the HII/molecular region interface. Comparing these values the clouds are found to fall into two categories: pre- and post-pressure balance states; SFO~75 and SFO~76 are identified as being in a pre-pressure balance state, and SFO~58 and SFO~68 are identified as being in a post-pressure balance state (taking account of the errors, see Section~\ref{sect:pressure_balance}). We draw the following conclusions from our observations: \begin{enumerate} \item Analysis has revealed clear morphological and evolutionary differences between the pre- and post-pressure balance clouds: \begin{itemize} \item The two clouds identified in this survey as being in a post-pressure balance state are also the same two identified as having a type A rim morphology and show strong evidence of ongoing high- to intermediate-mass star formation (e.g. UC HII region, masers and molecular outflows). \item The two clouds identified as being in a pre-pressure balance state have all been classified as type 0 clouds, and show no evidence of recent or ongoing star formation. \end{itemize} \item The two classifications of rim morphologies, type 0 and type A, correspond to the 0.036 Myr and 0.126 Myr snapshots from the \citet{lefloch1994} RDI model, where the time indicates the exposure time of a cloud to an ionising front. This is consistent with our conclusion that SFO~75 and SFO~76 have only recently been exposed to the ionisation front and that SFO~58 and SFO~68 have been exposed for a significantly longer period of time. Moreover, the morphological age predicted by the RDI models is similar to the estimated age of the star formation within SFO~58 and SFO~68, support the possibility that the star formation has been triggered. \item Using a simple pressure-based argument, exposure to the FUV radiation field within the HII regions is shown to have a profound effect on the stability of these cores. All of the cores were stable whilst embedded in their natal molecular clouds, however, in all cases exposure to the FUV radiation field of the HII region reduces the stability of the cores by more than a factor of two (in the case of SFO~76 by almost a factor of seven). The reduced stability leaves two clouds on the edge of being unstable to gravitational collapse, and two clouds that have masses at least a factor of two greater than the pressurised virial masses. Analysis of the pre-exposure stability indicates that it is possible that the core embedded within SFO~75 was unstable to gravitational collapse prior to being exposed to the ionisation front. From this analysis we conclude that the cores within SFO~75 and SFO~76 are both unstable to gravitational collapse. \item The radio continuum observations toward SFO~58 reveal the presence of an embedded compact radio source within the optical boundary of the bright rim. The radio source is offset by 1\arcmin~from the position of the IRAS point source, but correlates extremely well with the position of the peak of the molecular core, both of which are located at the focus of the bright rim, suggesting that the radio source is associated with the cloud. The size ($<$~0.06 pc) and integrated radio flux are all consistent with the presence of an UC HII region embedded within the molecular core. The radio flux of this source is consistent with the presence of an UC HII region excited by a single ZAMS B2--B3 star. Inspection of the core-averaged $^{12}$CO spectrum reveals evidence of a substantial blue wing, possibly indicating the the presence of a molecular outflow. This is the first tentative evidence for ongoing star formation within SFO~58. \item The physical sizes and masses of the molecular cores are typically larger than a single star might be expected to form from. The IRAS luminosities are generally much higher than the bolometric luminosities typical for individual Class 0 and Class I stars (\citealt{andre1993,chandler2000}). Additionally, evidence that two BRCs are active high-mass star forming regions is presented in this paper, which form exclusively in clusters. It is therefore highly likely that the presence of the IRAS point source indicates the presence of multiple protostellar systems rather than a single protostar. Higher resolution molecular line observations are required to investigate the possible multiplicity of these sources. \end{enumerate} \begin{acknowledgements} The authors thank the Director and staff of the Paul Wild Observatory, Narrabri, New South Wales, Australia, for their hospitality and assistance during our Compact Array and Mopra observing runs, and the Mopra support scientist Stuart Robertson for his help and advice. We would also like to thank the referee Bertrand Lefloch for his very helpful comments and suggestions. This research would not have been possible without the SIMBAD astronomical database service operated at CDS, Strasbourg, France, and the NASA Astrophysics Data System Bibliographic Services. We have made use of Digitised Sky Survey images was produced at the Space Telescope Science Institute under U.S. Government grant NAG W-2166. These images are based on photographic data obtained using the Oschin Schmidt Telescope on Palomar Mountain and the UK Schmidt Telescope. The plates were processed into the present compressed digital form with the permission of these institutions. This research has also made use of the NASA/IPAC Infrared Science Archive, which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. \end{acknowledgements} \bibliography{3417} \bibliographystyle{aa}
Title: Far Ultraviolet Spectral Images of the Vela Supernova Remnant
Abstract: We present far-ultraviolet (FUV) spectral-imaging observations of the Vela supernova remnant (SNR), obtained with the Spectroscopy of Plasma Evolution from Astrophysical Radiation (SPEAR) instrument, also known as FIMS. The Vela SNR extends 8 degrees in the FUV and its global spectra are dominated by shock-induced emission lines. We find that the global FUV line luminosities can exceed the 0.1-2.5 keV soft X-ray luminosity by an order of magnitude. The global O VI:C III ratio shows that the Vela SNR has a relatively large fraction of slower shocks compared with the Cygnus Loop.
https://export.arxiv.org/pdf/astro-ph/0601586
\title{Far Ultraviolet Spectral Images of the Vela Supernova Remnant} \author{K. Nishikida\altaffilmark{1}, J. Edelstein\altaffilmark{1}, E. J. Korpela\altaffilmark{1}, R. Sankrit\altaffilmark{1}, W. M. Feuerstein\altaffilmark{1}, K. W. Min\altaffilmark{2}, J-H. Shinn\altaffilmark{2}, D-H. Lee\altaffilmark{2},I-S. Yuk\altaffilmark{3}, H. Jin\altaffilmark{3}, K-I. Seon\altaffilmark{3}} \altaffiltext{1}{Space Sciences Laboratory, University of California, Berkeley, CA 94720} \altaffiltext{2}{Korea Advanced Institute of Science and Technology, 305-701, Daejeon, Korea} \altaffiltext{3}{Korea Astronomy and Space Science Institute, 305-348, Daejeon, Korea} \keywords{ISM: individual (Vela Supernova Remnant) -- supernova remnants -- ultraviolet: ISM} \section{Introduction} The Vela supernova remnant (SNR) has been studied in great detail due to its proximity \citep[250pc;][]{vela_distance} and its large angular diameter of $8^{\circ}$ \citep[]{VelaROSAT}. The remnant is $\sim$10,000 years old \citep{vela_age} and its overall emission is dominated by the interaction of the SN blast wave with the interstellar medium (ISM), but also has a pulsar and plerionic nebula near its center. Vela is one of two Galactic SNRs (the other being the Cygnus Loop) that has been extensively studied in the UV. The UV emission from SNRs arises primarily in shocks driven by the supernova blast wave into interstellar clouds. The shock velocities responsible for the UV emission lie in the range $\sim$50-300 km s$^{-1}$, and heat the gas to temperatures of ~$10^5 - 10^6$ K. The hot shocked gas can also be studied via absorption so long as there are suitable background continuum sources. Absorption line studies of Vela using \textit{Copernicus} showed the presence of O~{\small VI} and N~{\small V}, high ionization species expected in shocked gas, and also high velocity components of lower ionization species \citep{Jenkins1976a,Jenkins1976b}. Absorption line studies have also been carried out with \textit{IUE} \citep{Jenkins,Nichols}, \emph{HST} \citep{Jenkins1995,Jenkins1998} and \textit{FUSE} \citep{Slavin}. These studies have shown the existence of fast shocks ($\gtrsim$160 km s$^{-1}$) distributed widely but inhomogeneously across the face of the remnant, and also the variations in the dynamic pressure driving these shocks. Emission line studies of Vela have been carried out using \textit{IUE} \citep{Vela_IUE}, \textit{HUT} \citep{RaymondVela}, \textit{FUSE} \citep{ravi2,ravi} and \emph{Voyager 2} UVS \citep[henceforth BVL]{vela_voyager}. These observations, except for the ones obtained by \emph{Voyager 2}, were of regions with angular extents of order an arc-minute and probed the properties of individual shock fronts. BVL analyzed \emph{Voyager 2} spectra of a $1.5^{\circ} \times 2.0^{\circ}$ region in the northern part of Vela, which showed strong C~{\small III} and O~{\small VI} emission features. They found variations in the flux ratios on scales of 0.2$^{\circ}$. They also concluded from the overall flux ratio between the two lines that slower shocks (those unable to produce O~{\small VI}) were more prevalent in Vela than in the Cygnus Loop. \textit{FUSE} spectra of a few regions in Vela well separated from each other have strong C~{\small III} lines and relatively weak or no O~{\small VI}, and show the presence of slower shocks ($100 - 140$ km s$^{-1}$) spread over the face of the remnant \citep{raviproc}. We present spectral images of the Vela SNR in several FUV lines and FUV spectra of the entire remnant. The data were obtained with \emph{SPEAR} (The Spectroscopy of Plasma Emission from Astrophysical Radiation), also known as \emph{FIMS} (Far-ultraviolet Imaging Spectrograph). \emph{SPEAR}, launched Sepember 27, 2003 on the Korean satellite STSAT-1, is a dual-channel FUV imaging spectrograph (S channel 900 - 1150 \AA, L channel 1350 - 1750 \AA, $\lambda/\Delta\lambda\sim$550) with a large imaged field of view (S: 4.0$^{\circ} \times 4.6'$, L: 7.5$^{\circ} \times 4.3'$, spatial resolution $\sim10'$) optimized for the observation of diffuse emission \citep[see][for an overview of the instrument and mission.]{Instrument,Mission}. A large effective field of view can be obtained by sweeping across the sky. This combination of instrument properties yields a dataset that supplements data obtained in previous UV studies of the remnant. The data allow us to estimate the total flux from Vela in several FUV lines. In the following sections we present the observations (\S2), discuss the spectral images (\S3) and the total spectrum (\S4). The last section (\S5) summarizes the importance of these data for the study of SNRs. \section{Observation and Data Analysis} The Vela SNR region was observed between January 31 and February 4, 2004. The data were processed as described in \citet[]{Instrument,Mission} including rejection of data for which the attitude knowledge was poor ($>$30') or data contaminated by airglow, evident from an increased count rate at the end of each orbit ($\leq$10\% of the data). The resulting number of photons and exposure obtained over 16 orbits were (4.4$\times10^{6}$, 7214 s) and (1.8$\times10^{5}$, 6307 s) for the L channel and S channel, respectively. While the SNR region was entirely observed in the L channel, the S channel coverage was incomplete. The photon's sky coordinates ($\alpha$, $\delta$) and wavelengths ($\lambda$) were binned to 0.15$^{\circ}$ (similar to the \emph{SPEAR} imaging resolution after attitude reconstruction) and to 1 \AA, respectively and combined with exposure maps to create a 3-d count-rate data cube ($\alpha$, $\delta$, $\lambda$) for each spectral channel. We identified data contaminated by bright stars in each channel as pixels in the wavelength-integrated total count rate image that exceeded 3 times the median count rate of a relatively star-free area. The L and S channels contained 20\% and 10\% of the total pixels identified as stars, respectively. To create spectral images, the data cube was summed across the waveband of interest with star pixels removed. Star pixel ``holes'' were filled with a linear fit to adjacent pixels at the ``hole's'' declination. The undetected stars have a flux $<1.5\times 10^{-11}$ ergs s$^{-1}$ cm$^{-2}$\AA$^{-1}$ in the \emph{SPEAR} L channel. Improved stellar identification, removal and reconstruction methods are currently being developed. Net spectral images for specific emission lines, $I_{\lambda_{Net}}$, were constructed by subtracting a continuum image, $I_{cont}$, made from a similar width spectral region \emph{adjacent} to the emission line from an image, $I_{\lambda}$, made from a spectral region \emph{including} the emission line, i.e. $I_{\lambda_{Net}}=I_{\lambda}- I_{cont}$. This approach often results in an over-subtraction of the lower intensity pixels. The continuum selected from the global spectra is not suitable for lower intensity pixels because of the non-zero slope of the adjacent continuum. Therefore we apply a variable scaling factor, $f$, to the subtraction of the continuum image: $I_{\lambda_{Net}} = I_{\lambda}-f \times I_{cont}$. For each line image, the factor $f$ was set for each image pixel associated with a bin of a five-bin intensity histogram of $I_{\lambda}$ such that 70\% (i.e. $\sim$2$\sigma$) of those pixels would have a net positive flux. The resulting values of $f$ were 1.0 at high to medium intensity bins and decreased for the lowest one to three intensity bins, depending on $\lambda$, with a typically minimum value of $f=$0.6 for the lowest intensity bin. The resulting $I_{\lambda_{Net}}$ were smoothed using a 3-pixel (0.45$^{\circ}$) square median smoothing function. Diffuse spectra were derived by totaling photons and exposure over regions of interest, excluding bright star pixels, and then smoothed by a 3 \AA\ boxcar function. \section{Spectral Images} We show the C {\small IV} (1543--1557 \AA) image of Vela in Fig.~\ref{c4image}. This is the most prominent emission line in the L channel, and the figure shows the overall morphology of the FUV emission from the SNR. Images of Vela in C {\small III} (974-980 \AA), O {\small VI} (1019-1033 \AA), Si {\small IV} \& O {\small IV]} (1398--1413 \AA) are shown in Fig.~\ref{lineimages2}. Also shown for comparison is the \emph{ROSAT} All-Sky Survey (RASS) 1/4 keV image of the same region. The \emph{SPEAR} maps show that FUV emission is present over most of the Vela SNR. The C~{\small III} map in particular shows the widespread presence of radiative shocks. O {\small VI} and C {\small III} trace gas at very different ionization states. O~{\small VI} production requires a shock velocity of about 150 km/s while C~{\small III} is produced even in 80 km/s shocks. There is some overall correspondence between the two, showing where the FUV producing shocks exist, but C~{\small III} is more extended. C~{\small IV} and Si~{\small IV} \& O~{\small IV]} images show a close correlation: C~{\small IV}, Si~{\small IV}, and O~{\small IV]} have comparable ionization potentials and are roughly co-extensive in the post-shock gas. Although the FUV and X-ray emission features are not closely correlated, their extent is roughly the same. \citet{lu} attribute the X-ray emission to thermal emission from a hot, thin gas in the SNR interior. We used the plasma temperatures and emission measures in Table 1 of \citet{lu} as input for CHIANTI \citep{chianti0,chianti} and confirmed that they produce insufficient thermal FUV emission by several orders of magnitudes compared with the observed emission. The FUV emission comes from shocks that have been driven into interstellar clouds. These clouds are large enough that they have not been destroyed by the blast wave sweeping over them, and they have a high covering factor. The total FUV luminosity of Vela exceeds the total soft X-ray luminosity (see \S4). Thus, the overall radiation rate of the SNR is dominated by shocks in higher density regions traced by their FUV emission. The most prominent localized feature is an intense knot of emission near ($\alpha$, $\delta$)=(8.6 h, -42.5$^{\circ}$) that appears at all FUV wavelengths, including the soft X-ray band. A detailed comparison shows that the peak emission for each FUV wavelength and the soft X-ray are non-coincident and arranged in an arc, suggesting that the feature could be a complex and intense bow shock perhaps analogous to the X-ray ``bullet shocks,'' or ``knots,'' of similar scale. \citet{VelaROSAT} identified six (labeled A through F) of these X-ray knots and suggested that they were created by high-density ejecta shock-heating the ambient medium. The FUV images appear to be limb brightened along the north-east shell boundary and, notably, joins the X-ray knot D to the main shell in C {\small IV} and Si {\small IV}/O {\small IV]}. FUV emission is absent from the knot B region despite its similarity in X-ray intensity to knot D. This is consistent with FUV emission arising from shocked interstellar media because the knot D bullet is believed to be encountering a more dense ambient media than knot B. An FUV enhancement particularly notable in O {\small VI}, whose localized intensity peak is coincident with the Vela pulsar, and C {\small III}, which coincides with the Vela-X radio continuum nebula \citep[Figures 1 \& 2]{vela_480,vela_843}, is coincident with a soft X-ray ridge, suggestive of a shock structure. Curiously, there is little associated C {\small IV} or Si {\small IV} \& O {\small IV]} emission from the same area in comparison to the region of maximum intensity ($\alpha$, $\delta$)=(8.6 h, -42.5$^{\circ}$). This suggests that we are detecting at least two markedly different physical conditions in the region surrounding the Vela pulsar. A more detailed examination will be necessary to tell if these are related, or merely happen to lie along the same line of sight. \section{FUV Spectra} The diffuse \emph{SPEAR} FUV spectra from the entire Vela SNR region are shown in Fig.~\ref{spectra}. The spectra show strong emission lines from highly ionized atoms produced by high velocity shocks and/or hot plasmas. Detected lines include C {\small III} (977 \AA), N~{\small III} (990 \AA), O~{\small VI} (1032, 1038 \AA), O {\small IV]} and Si {\small IV} (1400, 1403 \AA\ unresolved), N~{\small IV} (1486 \AA), C {\small IV} (1548, 1550 \AA), He {\small II} (1640 \AA), and O {\small III]} (1660, 1666 \AA). The emission line-like feature near 1695 \AA\ is instrumental. Table~\ref{luminosity} shows the observed FUV line intensities averaged over the SNR. The C {\small IV}, He {\small II}, and O {\small III]} line profiles were fitted with a Gaussian line profile and a local linear background, while the C {\small III} and O~{\small VI} 1032 \AA\ lines required an additional Gaussian line profile to fit airglow lines adjacent to C~{\small III} and O {\small VI} $\lambda$1032. The O {\small VI} doublet intensity was calculated by multiplying the 1032 \AA\ line intensity by 1.5 (assuming a 2:1 ratio between $\lambda \lambda$1032, 1038) since \emph{SPEAR} does not have the spectral resolution necessary to resolve O {\small VI} $\lambda$1038 and C {\small II}$^{*}$ $\lambda$1037. For C~{\small III} and O~{\small VI}, we assumed that the average intensity calculated from areas with sky coverage applied to the entire SNR. Since the Vela SNR appears to be faint at the edges, it is likely that the intensities are slightly overestimated. The estimated systematic uncertainty in the \emph{SPEAR} sensitivity is $\sim25\%$ \citep{Instrument}. The \emph{Voyager 2} UVS measurements of O {\small VI} and C {\small III} line intensities at 28 \AA\ resolution (BVL) are in agreement within a factor of two below \emph{SPEAR} measurements in an area near ($\pm1^{\circ}$) UVS pointings. The global observed O~{\small VI}:C~{\small III} ratio is $\sim1$, consistent within a factor of about two (both above and below) reported by BVL. We apply multiplicative correction factors (shown in Table~\ref{luminosity}) to the observed global FUV luminosities, presuming a diameter of $8^{\circ}$ and distance of 250 pc, to derive dereddened luminosities. These correction factors were calculated by \citet{ravi} by assuming R=3.1 for selective extinction, E(B-V)=0.1 \citep{Vela_color}, and using the extinction curve suggested by \citet{Vela_ExtinctionCurve}. The integrated C~{\small III} or O~{\small VI} line luminosity can exceed the entire 0.1 - 2.5 keV luminosity, $2.2\times 10^{35}$ ergs s$^{-1}$ \citep{lu}, by more than an order of magnitude, confirming the importance of the FUV waveband to SNR cooling. Vela's global FUV luminosity can be compared with that of the Cygnus Loop \citep{cygnus_voyager}. Vela is about five times less luminous than Cygnus in the X-ray and in C {\small IV}, about twice as faint in O {\small VI}, and is equally as luminous in C {\small III}. The observed O~{\small VI}:C~{\small III} ratio of Cygnus is $\sim2$, illustrating the relatively large fraction of slower shocks in Vela compared with Cygnus. For Cygnus, \citet{cygnus_rocket} showed that the O {\small VI} luminosity is at least as much as the 0.1--4 keV X-ray luminosity \citep{cygnus_einstein}; the combined O~{\small VI}, C~{\small III}, C~{\small IV} luminosity was found to be ten times as large as the X-ray luminosity \citep{cygnus_voyager}. In both the Cygnus Loop and Vela, the supernova blast wave is expanding into inhomogeneous surroundings, and the radiative shocks in the denser interaction regions emit strongly in the FUV. \section{Summary and Future Work} We have presented global FUV spectral images and spectra of the Vela SNR. Our data indicate inhomogeneous shock-induced emission from the SNR surface. The images show limb brightening and knots of emission. The spectra are consistent with past FUV observations of Vela. The global FUV luminosities of emission lines can exceed the soft X-ray luminosity by an order of magnitude, providing an efficient cooling channel to the SNR. The data will be used to compare the global FUV morphology with images in other wavebands and examine how the SNR evolves and interacts with the ambient ISM. \emph{SPEAR} spectra toward previously observed regions \citep[ex.][]{RaymondVela,ravi2} will be modeled in detail to determine the physical properties of the regions. Mapping the O~{\small VI}:C~{\small III} ratio will allow us to map the distribution of shock velocities across Vela. Furthermore, the global \emph{SPEAR} spectra can be compared with those of neighboring regions to investigate Vela's association with the Gum nebula. \acknowledgements \emph{SPEAR / FIMS} is a joint project of KASSI \& KAIST (Korea) and U.C., Berkeley (USA), funded by the Korea MOST and NASA Grant NAG5-5355. We used NASA's \emph{SkyView} (http://skyview.gsfc.nasa.gov) facility and CHIANTI, a collaboration of NRL (USA), RAL (UK), U. Florence (Italy) and Cambridge (UK). \clearpage \begin{table}[htbp] \caption{Observed global FUV intensities and dereddened luminosities of the Vela SNR} \begin{tabular}{c|cccccc} Species & C~{\small III} & O~{\small VI} & C~{\small IV} & He~{\small II} & O~{\small III]} & X-ray \\ \hline\hline Observed intensity ($10^5 LU$) & 2.6 & 2.7 & 2.3 & 0.5 & 0.5 & \\ Reddening correction & 4.47 & 3.73 & 2.07 & 2.03 & 2.03 & \\ Dereddened luminosity ($10^{35}$ ergs s$^{-1}$) & 26.8 & 22.7 & 8.0 & 1.3 & 1.5 & 2.2 \\ \end{tabular} \tablecomments{Diameter of 8$^{\circ}$ and 250 pc distance assumed. The O~VI luminosity was calculated assuming a 2:1 line ratio between O~VI 1032 \AA\ and 1038 \AA\ lines. Reddening correction values are taken from \citet{ravi}. X-ray (0.1--2.5 keV) luminosity from \citet{lu}. 1 Line Unit (LU) = 1 photon s$^{-1}$ cm$^{-2}$ sr$^{-1}$ = 1.9$\times 10^{-11}$ ergs s$^{-1}$ cm$^{-2}$ sr$^{-1}$ at 1032 \AA\ and 1.3$\times 10^{-11}$ ergs s$^{-1}$ cm$^{-2}$ sr$^{-1}$ at 1550 \AA.} \label{luminosity} \end{table} \clearpage \clearpage \clearpage
Title: Theoretical Isochrones with Extinction in the K Band. II. J - K versus K
Abstract: We calculate theoretical isochrones in a consistent way for five filter pairs near the J and K band atmospheric windows (J-K, J-K', J-Ks, F110W-F205W, and F110W-F222M) using the Padova stellar evolutionary models of Girardi et al. We present magnitude transformations between various K-band filters as a function of color. Isochrones with extinction of up to 6 mag in the K band are also presented. As found for the filter pairs composed of H & K band filters, we find that the reddened isochrones of different filter pairs behave as if they follow different extinction laws, and that the extinction curves of Hubble Space Telescope NICMOS filter pairs in the color-magnitude diagram are considerably nonlinear. Because of these problems, extinction values estimated with NICMOS filters can be in error by up to 1.3 mag. Our calculation suggests that the extinction law implied by the observations of Rieke et al for wavelengths between the J and K bands is better described by a power-law function with an exponent of 1.66 instead of 1.59, which is commonly used with an assumption that the transmission functions of J and K filters are Dirac delta functions.
https://export.arxiv.org/pdf/astro-ph/0601470
\title{Theoretical Isochrones with Extinction in the $K$ Band. II. \\ $J$ -- $K$ versus $K$} \author{Sungsoo S. Kim\altaffilmark{1}, Donald F. Figer\altaffilmark{2}, and Myung Gyoon Lee\altaffilmark{3}} \altaffiltext{1}{Department of Astronomy and Space Science, Kyung Hee University, Yongin-shi, Kyungki-do 449-701, South Korea; sungsoo.kim@khu.ac.kr.} \altaffiltext{2}{Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218; figer@stsci.edu.} \altaffiltext{3}{Astronomy Program, SEES, Seoul National University, Seoul 151-742, South Korea; mglee@astrog.snu.ac.kr.} \keywords{Hertzsprung-Russell diagram --- techniques: photometric --- stars: fundamental parameters --- infrared: stars} \section{INTRODUCTION} \label{sec:introduction} Kim et al. (2005, hereafter Paper I) have calculated theoretical isochrones with extinction for some $H$ and $K$ band filters using the Padova stellar evolutionary models by Girardi et al. (2002). In Paper I, we found that the reddened isochrones of different filter pairs in $H$ and $K$ bands behave as if they follow different extinction laws, and that care is needed when applying an extinction law obtained with one filter pair to other, similar filter pairs. For example, if the extinction law for the Johnson-Glass $H$ and $K$ filters obtained by Rieke, Rieke, \& Paul (1989) is directly applied to the photometry from the {\it Hubble Space Telescope} ({\it HST}) NICMOS filters (F160W, F205W, and F222M), estimated extinction values can be in error by up to 0.3 mag for true extinction at $K$ of 6 mag or less. To reduce this error, Paper I introduced an ``effective extinction slope'' for each filter pair and isochrone model. It was also found that the extinction behavior of isochrones in the color-magnitude diagram (CMD) for filter pair F160W--F222M is highly nonlinear (i.e., the amount of extinction is not proportional to color excess) because of a significant width difference in the two filters. These problems are certainly not limited to the isochrones for filter pairs in the $H$ and $K$ bands. This problem will apply to any situation in which one applies an extinction law deduced from one filter pair to other similar filter pairs. Furthermore, the nonlinear behavior of the extinction vector in the CMD will be problematic for filter pairs with significant difference in width. In the present paper, we extend the calculations performed in Paper I to the isochrones for filter pairs in the $J$ and $K$ bands. The filters considered here are the four ground-based filters $J$, $K$ (Johnson et al. 1966), $K'$ (Wainscoat \& Cowie 1992), and $K_s$ ($K$-short; developed by M. Skrutskie; see the appendix of Persson et al. 1998), and the three NICMOS filters F110W, F205W, and F222M (transmission functions of these filters are shown in Figure~\ref{fig:filter}). Out of these seven filters, we consider five filter pairs: $J$--$K$, $J$--$K'$, $J$--$K_s$, F110W--F205W, and F110W--F222M. We adopt a Vega-based photometric system (VEGAMAG system), which uses Vega ($\alpha$ Lyr) as the calibrating star. For photometric zero points of NICMOS filters, we adopt $\langle f_\nu^{\rm Vega} \rangle$ values from the NICMOS Data Handbook (ver. 5.0): 1775~Jy for F110W, 703.6~Jy for F205W, and 610.4~Jy for F222M. For the spectra of synthetic stellar atmospheres, we adopt Kurucz ATLAS9 no-overshoot models\footnote{See NOVER files at http://kurucz.harvard.edu/grids.html.} (Kurucz 1993) calculated by Castelli et al. (1997). The metallicities of these models cover the values of [M/H] = $-2.5$ to $+0.5$. A microturbulent velocity $\xi=2\,{\rm km \, s^{-1}}$ and a mixing length parameter $\alpha =1.25$ are adopted in the present study. For the temporal evolution of effective temperature and luminosity as functions of stellar mass (i.e., stellar evolutionary tracks), we adopt the ``basic set" of the Padova models\footnote{See http://pleiadi.pd.astro.it.} (Girardi et al. 2002). We consider isochrones with a metallicity $Z$ = 0.0001, 0.001, 0.019, and 0.03. The stellar spectral library and the evolutionary tracks we adopted assume a solar chemical ratios. For more details on the magnitude system, stellar spectral library, and evolutionary tracks that we adopt here, readers are referred to Paper I. Throughout this paper, we generically refer to the atmospheric wavebands centered near 1.25, 1.65, and 2.2~\micron, as the $J$, $H$, and $K$ bands, whereas we refer to the Johnson-Glass filters (Johnson et al. 1966; Glass 1974) as the $J$, $H$, and $K$ filters. \section{Isochrones} \label{sec:isochrones} We first prepare a table of magnitudes for all spectra in ATLAS9 models in the $J$ and $K$ band filters, covering a large range in $T_{eff}$, $\log g$, and [M/H], using equations (5) and (6) of Paper I. We use this table as a set of interpolates for the $T_{eff}$, $\log g$, and $Z$ values predicted by the stellar evolution models for a given age in order to estimate synthetic isochrones. Isochrones for $A_\lambda = 0$, calculated in this way, are shown in Figures~\ref{fig:iso1}$-$\ref{fig:iso4} for four different metallicities and four ages. The color differences between filters are more prominent for the highest metallicity isochrones. In most cases, isochrones for $K'$ and $K_s$ are nearly indistinguishable, and those for F205W and F222M are quite close to each other. In general, for red giants, intrinsic color differences between the atmospheric and NICMOS filters are 0.2--0.4~mag. As an independent check of our procedure, in Figure~\ref{fig:padova} we compare our $J-K$ versus $K$ isochrones to those calculated by Girardi et al. (2002). The isochrones match nicely, except at the extremes. The discrepancy in the bright end is caused from the empirical M giant spectra that Girardi et al. (2002) added to their spectral library, and that in the faint end is by the addition of late M dwarf spectra. The discrepancies are considerable only at the top and bottom $\sim$ 1 mag of the isochrone, where only a small fraction of giants reside, or else stars are too faint for most observational situations. Magnitude transformations between $K$-band filters can be obtained from our isochrones. We find that the magnitude difference can be well fitted by a third-order polynomial for $K < 4$~mag, and by a separate second-order polynomial for $K > 4$~mag. The largest residuals from the fit are 0.012~mag for the former and 0.008~mag for the latter. The coefficients of the best-fit functions are presented in Tables~\ref{table:trans1} and \ref{table:trans2}, along with the residuals and fitting ranges. One useful way of using these tables would be to compare the magnitudes of helium-burning clump giant stars, which are rather insensitive to metallicity or age and are often used as distance indicators, observed with different photometric systems (the clump stars show a small variation with age, however; see Figer et al. 2004). We present here isochrones with $K$-band extinctions of up to 6 mag, some of which are shown in Figures~\ref{fig:red1}$-$\ref{fig:red5}. For the extinction between the $J$ and $K$ bands, we adopt a power law, \begin{equation} \label{extinction} A_\lambda = A_0 \left ( \frac{\lambda}{\lambda_0} \right )^{-\alpha}, \end{equation} where we choose $\lambda_0 = 2.2 \, \mu$m, and $A_0$ is the extinction at $\lambda_0$. When assuming that the transmission functions of the $J$ and $K$ filters are Dirac delta functions centered at 1.24 and 2.21~$\mu$m, respectively, the extinction law by Rieke et al. (1989) gives $\alpha=1.59$. However, as discussed below in this section, the apparent extinction behavior of isochrones in the CMD can differ from the actual extinction law, as a result of a nonzero width and asymmetry of the filter transmission functions. We find that $\alpha=1.66$ makes the isochrone for the $Z = 0.019$, age = $10^9$~yr model behave in the CMD as if it followed an extinction law with $\alpha=1.59$. We choose this particular isochrone for calibrating the extinction law, with the assumption that the stars used in Rieke et al. (1989) to derive their extinction law, which are the stars in the central parsec of our Galaxy, can be represented by the same metallicity and age. For the sake of comparison, isochrones in Figures~\ref{fig:red1}$-$\ref{fig:red5} have been dereddened by the amount $A_0 ( \lambda_c / \lambda_0 )^{-1.66}$, where the central wavelength of the filter $\lambda_c$ is defined by equation (8) of Paper I, and given in Table~\ref{table:lambdac}. Since we have dereddened the isochrones with the known amount of extinction at $\lambda_c$, all the dereddened isochrones with different extinction values in Figures~\ref{fig:red1}$-$\ref{fig:red5} should be coincident if the filter transmission functions were Dirac delta functions centered at $\lambda_c$. As in Paper I, the dereddened isochrones misalign significantly, and this implies that the amount of extinction inferred from a CMD is sensitively dependent on the shape of the filter transmission function. When estimating the amount of extinction from an observed CMD, one converts an observed color excess to an extinction value, following an assumed extinction law, which usually has the form of a power law. When one has photometric data from a pair of two filters, $X$ and $Y$, the amount of extinction can be estimated by \begin{eqnarray} \label{A_est} A_Y^{est} & = & \frac{(m_X-m_Y)-(m_X-m_Y)_0}{A_X/A_Y-1} \cr & = & \frac{(m_X-m_Y)-(m_X-m_Y)_0} {(\lambda_X/\lambda_Y)^{-\alpha}-1}, \end{eqnarray} where $m_X$, $m_Y$ and $\lambda_X$, $\lambda_Y$ are the magnitudes and the central wavelengths of the two filters, respectively, and subscript 0 denotes the intrinsic value. For estimating extinction from our isochrones, we first use $\alpha=1.59$. Figure~\ref{fig:adiff1} shows the difference between the inferred extinction values, using equation~(\ref{A_est}) and colors from our reddened isochrones, and the actual extinction values. Here the extinction of each isochrone has been calculated using the mean color (for $A^{est}_Y$) and magnitude (for $A_Y$) of the reddened isochrone data points having intrinsic $K$-band magnitudes between $-$6 and 0 mag. As the figure shows, the differences between estimated and actual extinction values are much larger for the NICMOS filter pairs. The largest relative difference is $\sim 24$\%, and the largest absolute difference is 1.25 mag. Note that the extinction estimates for the $Z = 0.019$ and age = $10^9$~yr model inferred from $H$ and $K$ are very close to the actual extinction values, justifying our choice of $\alpha=1.66$ for equation~(\ref{extinction}). The error bar in the figure represents the standard deviation of $A^{est}_Y-A_Y$ values. Some of the F110W isochrones show quite large deviations, as pointed out in Appendix A of Lee et al. (2001). To reduce the problems seen in Figure~\ref{fig:adiff1}, Paper I introduced an ``effective extinction slope'' $\alpha_{eff}$ for each filter pair and isochrone model, which is defined such that it better describes the extinction behavior in the CMD: \begin{equation} \label{alpha_eff} \alpha_{eff} = - \frac{ \log (1+1/b) }{ \log (\lambda_X/\lambda_Y) }, \end{equation} where $b$ is the slope of the straight line that fits the distribution of reddened magnitudes versus reddened colors, as in Figure~\ref{fig:extlaw}. This figure shows reddened $K$-band magnitudes and colors for the $Z = 0.019$ and age = $10^9$~yr isochrone (the figure only shows an isochrone data point whose intrinsic $K$ magnitude is 0, as an example). We calculate $b$ for data points of each isochrone whose intrinsic $K$ magnitudes are between $-6$ and 0~mag, and take an average for each isochrone model. Table~\ref{table:alpha_eff} shows the averages and standard deviations of $\alpha_{eff}$ values for each isochrone model. For atmospheric filters, the standard deviations of $\alpha_{eff}$ in an isochrone is generally much smaller than the differences of average $\alpha_{eff}$ values between different isochrones, while those for NICMOS filters are relatively larger. The average $\alpha_{eff}$ values range from 1.403 to 1.610, which are 15\% to 0.02\% smaller than the original $\alpha$ value we adopted for extinction, 1.66. As seen in Figure~\ref{fig:adiff3}, extinction values estimated by equation~(\ref{A_est}) with $\alpha_{eff}$ are closer to the actual values for atmospheric filters, but still deviate significantly from the actual values for NICMOS filters, because of the nonlinear extinction seen in Figure~\ref{fig:extlaw}. As pointed out in Paper I, the nonlinear extinction behavior of NICMOS filters is due to a significant difference in relative widths of the two filters: the width to central wavelength ratio $\Delta \lambda / \lambda_c$ is $\sim 0.5$ for F110W, while those for F205W and F222M are $\sim 0.3$ and $\sim 0.07$, respectively. Figure~\ref{fig:nonlin} shows the effect of the filter width by comparing the extinction behavior of six imaginary filter pairs. Filter pair $a$ represents $J$ and $K$, whose $\Delta \lambda / \lambda_c$ values are both $\sim 0.16$, and its extinction behavior in the CMD is nearly linear. On the other hand, filter pairs $b$ and $c$, which represent filter pairs F110W--F205W and F110W--F222M, show considerable nonlinearity. When the $\Delta \lambda / \lambda_c$ of the short-wavelength filter is reduced by $\sim 60$\%, however, the extinction behaves much more linearly ($d$ and $e$). This shows that the nonlinear extinction in filter pairs F110W--F205W and F110W--F222M is due to a relatively larger $\Delta \lambda / \lambda_c$ value of the F110W filter. When both filters have the same large $\Delta \lambda / \lambda_c$ values ($\sim 0.5$), the extinction becomes almost linear again ($f$). The introduction of effective extinction slopes does not alleviate the nonlinear extinction problem of NICMOS filter pairs. So in Table~\ref{table:nonlin} we provide the coefficients of best-fit third-order polynomials of the extinction curves for NICMOS filter pairs shown in Figure~\ref{fig:adiff1} so that one can accurately estimate the extinction value for NICMOS filter pairs as well. Note that we assumed $\alpha = 1.59$ for all isochrone models when estimating $A_Y^{est}$ in Figure~\ref{fig:adiff1}. As in paper I, we find that for the filters whose extinction behavior is relatively linear, the transformation of extinction values from filter $Y$ to filter $Y'$ can be obtained by \begin{equation} \label{A_trans} A_{Y'} = A_Y \left ( \frac{\lambda_{Y'}}{\lambda_Y} \right )^{-\alpha}, \end{equation} if $A_Y$ is estimated with $\alpha_{eff}$, and the original $\alpha$ value of 1.66 is used in the above equation. \section{SUMMARY} \label{sec:summary} We have calculated in a consistent way five near-infrared theoretical isochrones for filter pairs composed of $J$ and $K$ filters: $J$--$K$, $J$--$K'$, $J$--$K_s$, F110W--F205W, and F110W--F222M. We presented isochrones for a $Z$ of 0.0001--0.03 and an age of $10^7$--$10^{10}$~yr. Even in the same Vega magnitude system, near-infrared colors of the same isochrone can be different by up to $\sim 0.4$ mag at the bright end of the isochrone for different filter pairs. The difference in intrinsic colors for a red giant for atmospheric filters and the {\it HST} NICMOS filters is generally 0.2--0.4 mag. We have provided magnitude transformations between $K$-band filters as a function of color from $J$ and $K$ band filters. We also presented isochrones with $A_K$ of up to 6 mag. We found that care is needed when comparing extinction values that are estimated using different filter pairs, in particular when comparing those of atmospheric and NICMOS filter pairs: extinction values inferred using NICMOS filters can be in error by up to 1.3 mag. To alleviate this problem, we introduced an ``effective extinction slope'' for each filter pair and isochrone model, which describes the extinction-dependent behavior of isochrones in the observed CMD. We also provided a procedure to accurately estimate the extinction value for NICMOS filter pairs, whose extinction curves in the CMD are highly nonlinear. \acknowledgements We thank Jae-Woo Lee for a helpful discussion. S. S. K. was supported by the Astrophysical Research Center for the Structure and Evolution of the Cosmos (ARCSEC) of the Korea Science and Engineering Foundation through the Science Research Center (SRC) program. M. G. L. was in part supported by the ABRL (R14-2002-058-01000-0) and the BK21 program. \clearpage \clearpage \begin{deluxetable}{cclrrrrcc} \tabletypesize{\scriptsize} \tablecolumns{9} \tablewidth{0pt} \tablecaption{ \label{table:trans1}Best-Fit Coefficients for Magnitude Differences ($K < 4$~mag)} \tablehead{ \colhead{} & \colhead{Magnitude} & \colhead{} & \colhead{} & \colhead{} & \colhead{} & \colhead{} & \colhead{Residual\tablenotemark{a}} & \colhead{Fitting Range\tablenotemark{b}} \\ \colhead{Color} & \colhead{Difference} & \colhead{$Z$} & \colhead{$c_0$} & \colhead{$c_1$} & \colhead{$c_2$} & \colhead{$c_3$} & \colhead{(mag)} & \colhead{(mag $\sim$ mag)} } \startdata $J - K$ & $K'$ $-$$K$ & 0.0001 & $-0.001$ & $ 0.028$ & $-0.064$ & $ 0.057$ & $0.004$ & $-0.236 \sim 0.720$ \\ $J - K$ & $K'$ $-$$K$ & 0.001 & $-0.001$ & $ 0.032$ & $-0.039$ & $-0.036$ & $0.007$ & $-0.162 \sim 0.859$ \\ $J - K$ & $K'$ $-$$K$ & 0.019 & $ 0.002$ & $ 0.036$ & $-0.147$ & $ 0.074$ & $0.011$ & $-0.213 \sim 1.250$ \\ $J - K$ & $K'$ $-$$K$ & 0.03 & $ 0.001$ & $ 0.042$ & $-0.163$ & $ 0.083$ & $0.007$ & $-0.129 \sim 1.237$ \\ $J - K$ & $K_s$$-$$K$ & 0.0001 & $-0.000$ & $ 0.012$ & $-0.013$ & $ 0.006$ & $0.002$ & $-0.236 \sim 0.720$ \\ $J - K$ & $K_s$$-$$K$ & 0.001 & $-0.000$ & $ 0.015$ & $-0.003$ & $-0.052$ & $0.006$ & $-0.162 \sim 0.859$ \\ $J - K$ & $K_s$$-$$K$ & 0.019 & $ 0.002$ & $ 0.017$ & $-0.103$ & $ 0.051$ & $0.009$ & $-0.213 \sim 1.250$ \\ $J - K$ & $K_s$$-$$K$ & 0.03 & $ 0.001$ & $ 0.024$ & $-0.121$ & $ 0.061$ & $0.006$ & $-0.129 \sim 1.237$ \\ $J - K$ & F205W$-$$K$ & 0.0001 & $-0.030$ & $ 0.052$ & $-0.143$ & $ 0.150$ & $0.006$ & $-0.236 \sim 0.720$ \\ $J - K$ & F205W$-$$K$ & 0.001 & $-0.030$ & $ 0.056$ & $-0.095$ & $ 0.028$ & $0.005$ & $-0.162 \sim 0.859$ \\ $J - K$ & F205W$-$$K$ & 0.019 & $-0.029$ & $ 0.060$ & $-0.147$ & $ 0.074$ & $0.007$ & $-0.213 \sim 1.250$ \\ $J - K$ & F205W$-$$K$ & 0.03 & $-0.029$ & $ 0.061$ & $-0.147$ & $ 0.077$ & $0.004$ & $-0.129 \sim 1.237$ \\ $J - K$ & F222M$-$$K$ & 0.0001 & $-0.031$ & $-0.002$ & $-0.001$ & $-0.019$ & $0.005$ & $-0.236 \sim 0.720$ \\ $J - K$ & F222M$-$$K$ & 0.001 & $-0.030$ & $ 0.001$ & $-0.010$ & $-0.046$ & $0.006$ & $-0.162 \sim 0.859$ \\ $J - K$ & F222M$-$$K$ & 0.019 & $-0.028$ & $ 0.002$ & $-0.113$ & $ 0.060$ & $0.012$ & $-0.213 \sim 1.250$ \\ $J - K$ & F222M$-$$K$ & 0.03 & $-0.028$ & $ 0.009$ & $-0.134$ & $ 0.071$ & $0.007$ & $-0.129 \sim 1.237$ \\ $J - K'$ & $K$ $-$$K'$ & 0.0001 & $ 0.001$ & $-0.029$ & $ 0.068$ & $-0.061$ & $0.004$ & $-0.224 \sim 0.715$ \\ $J - K'$ & $K$ $-$$K'$ & 0.001 & $ 0.001$ & $-0.033$ & $ 0.047$ & $ 0.024$ & $0.006$ & $-0.155 \sim 0.881$ \\ $J - K'$ & $K$ $-$$K'$ & 0.019 & $-0.001$ & $-0.037$ & $ 0.142$ & $-0.070$ & $0.011$ & $-0.203 \sim 1.290$ \\ $J - K'$ & $K$ $-$$K'$ & 0.03 & $-0.001$ & $-0.042$ & $ 0.154$ & $-0.076$ & $0.006$ & $-0.123 \sim 1.278$ \\ $J - K'$ & $K_s$$-$$K'$ & 0.0001 & $ 0.001$ & $-0.017$ & $ 0.054$ & $-0.055$ & $0.003$ & $-0.224 \sim 0.715$ \\ $J - K'$ & $K_s$$-$$K'$ & 0.001 & $ 0.001$ & $-0.018$ & $ 0.038$ & $-0.017$ & $0.002$ & $-0.155 \sim 0.881$ \\ $J - K'$ & $K_s$$-$$K'$ & 0.019 & $ 0.000$ & $-0.019$ & $ 0.043$ & $-0.022$ & $0.003$ & $-0.203 \sim 1.290$ \\ $J - K'$ & $K_s$$-$$K'$ & 0.03 & $ 0.001$ & $-0.019$ & $ 0.040$ & $-0.021$ & $0.002$ & $-0.123 \sim 1.278$ \\ $J - K'$ & F205W$-$$K'$ & 0.0001 & $-0.029$ & $ 0.025$ & $-0.085$ & $ 0.099$ & $0.003$ & $-0.224 \sim 0.715$ \\ $J - K'$ & F205W$-$$K'$ & 0.001 & $-0.029$ & $ 0.024$ & $-0.054$ & $ 0.058$ & $0.004$ & $-0.155 \sim 0.881$ \\ $J - K'$ & F205W$-$$K'$ & 0.019 & $-0.031$ & $ 0.024$ & $-0.002$ & $ 0.001$ & $0.005$ & $-0.203 \sim 1.290$ \\ $J - K'$ & F205W$-$$K'$ & 0.03 & $-0.030$ & $ 0.019$ & $ 0.012$ & $-0.004$ & $0.005$ & $-0.123 \sim 1.278$ \\ $J - K'$ & F222M$-$$K'$ & 0.0001 & $-0.030$ & $-0.032$ & $ 0.067$ & $-0.080$ & $0.005$ & $-0.224 \sim 0.715$ \\ $J - K'$ & F222M$-$$K'$ & 0.001 & $-0.029$ & $-0.032$ & $ 0.032$ & $-0.012$ & $0.003$ & $-0.155 \sim 0.881$ \\ $J - K'$ & F222M$-$$K'$ & 0.019 & $-0.030$ & $-0.035$ & $ 0.036$ & $-0.015$ & $0.004$ & $-0.203 \sim 1.290$ \\ $J - K'$ & F222M$-$$K'$ & 0.03 & $-0.029$ & $-0.035$ & $ 0.031$ & $-0.013$ & $0.003$ & $-0.123 \sim 1.278$ \\ $J - K_s$ & $K$ $-$$K_s$ & 0.0001 & $ 0.000$ & $-0.012$ & $ 0.013$ & $-0.005$ & $0.002$ & $-0.232 \sim 0.718$ \\ $J - K_s$ & $K$ $-$$K_s$ & 0.001 & $ 0.000$ & $-0.015$ & $ 0.008$ & $ 0.043$ & $0.006$ & $-0.160 \sim 0.879$ \\ $J - K_s$ & $K$ $-$$K_s$ & 0.019 & $-0.002$ & $-0.017$ & $ 0.098$ & $-0.047$ & $0.009$ & $-0.209 \sim 1.289$ \\ $J - K_s$ & $K$ $-$$K_s$ & 0.03 & $-0.001$ & $-0.023$ & $ 0.113$ & $-0.055$ & $0.006$ & $-0.127 \sim 1.278$ \\ $J - K_s$ & $K'$ $-$$K_s$ & 0.0001 & $-0.001$ & $ 0.017$ & $-0.052$ & $ 0.053$ & $0.003$ & $-0.232 \sim 0.718$ \\ $J - K_s$ & $K'$ $-$$K_s$ & 0.001 & $-0.001$ & $ 0.018$ & $-0.037$ & $ 0.017$ & $0.002$ & $-0.160 \sim 0.879$ \\ $J - K_s$ & $K'$ $-$$K_s$ & 0.019 & $-0.000$ & $ 0.019$ & $-0.042$ & $ 0.022$ & $0.003$ & $-0.209 \sim 1.289$ \\ $J - K_s$ & $K'$ $-$$K_s$ & 0.03 & $-0.001$ & $ 0.018$ & $-0.039$ & $ 0.021$ & $0.002$ & $-0.127 \sim 1.278$ \\ $J - K_s$ & F205W$-$$K_s$ & 0.0001 & $-0.029$ & $ 0.041$ & $-0.133$ & $ 0.147$ & $0.005$ & $-0.232 \sim 0.718$ \\ $J - K_s$ & F205W$-$$K_s$ & 0.001 & $-0.030$ & $ 0.042$ & $-0.090$ & $ 0.075$ & $0.004$ & $-0.160 \sim 0.879$ \\ $J - K_s$ & F205W$-$$K_s$ & 0.019 & $-0.031$ & $ 0.042$ & $-0.044$ & $ 0.022$ & $0.006$ & $-0.209 \sim 1.289$ \\ $J - K_s$ & F205W$-$$K_s$ & 0.03 & $-0.031$ & $ 0.037$ & $-0.027$ & $ 0.016$ & $0.005$ & $-0.127 \sim 1.278$ \\ $J - K_s$ & F222M$-$$K_s$ & 0.0001 & $-0.030$ & $-0.014$ & $ 0.012$ & $-0.024$ & $0.005$ & $-0.232 \sim 0.718$ \\ $J - K_s$ & F222M$-$$K_s$ & 0.001 & $-0.030$ & $-0.014$ & $-0.007$ & $ 0.006$ & $0.004$ & $-0.160 \sim 0.879$ \\ $J - K_s$ & F222M$-$$K_s$ & 0.019 & $-0.030$ & $-0.016$ & $-0.008$ & $ 0.008$ & $0.005$ & $-0.209 \sim 1.289$ \\ $J - K_s$ & F222M$-$$K_s$ & 0.03 & $-0.030$ & $-0.016$ & $-0.010$ & $ 0.008$ & $0.004$ & $-0.127 \sim 1.278$ \\ F110W$-$F205W & $K$ $-$F205W & 0.0001 & $ 0.030$ & $-0.040$ & $ 0.085$ & $-0.070$ & $0.007$ & $-0.304 \sim 0.928$ \\ F110W$-$F205W & $K$ $-$F205W & 0.001 & $ 0.030$ & $-0.043$ & $ 0.053$ & $-0.010$ & $0.005$ & $-0.210 \sim 1.122$ \\ F110W$-$F205W & $K$ $-$F205W & 0.019 & $ 0.029$ & $-0.047$ & $ 0.086$ & $-0.033$ & $0.006$ & $-0.272 \sim 1.621$ \\ F110W$-$F205W & $K$ $-$F205W & 0.03 & $ 0.029$ & $-0.047$ & $ 0.086$ & $-0.034$ & $0.004$ & $-0.166 \sim 1.615$ \\ F110W$-$F205W & $K'$ $-$F205W & 0.0001 & $ 0.029$ & $-0.018$ & $ 0.048$ & $-0.043$ & $0.003$ & $-0.304 \sim 0.928$ \\ F110W$-$F205W & $K'$ $-$F205W & 0.001 & $ 0.029$ & $-0.018$ & $ 0.035$ & $-0.030$ & $0.003$ & $-0.210 \sim 1.122$ \\ F110W$-$F205W & $K'$ $-$F205W & 0.019 & $ 0.031$ & $-0.017$ & $-0.000$ & $-0.000$ & $0.005$ & $-0.272 \sim 1.621$ \\ F110W$-$F205W & $K'$ $-$F205W & 0.03 & $ 0.030$ & $-0.013$ & $-0.010$ & $ 0.003$ & $0.005$ & $-0.166 \sim 1.615$ \\ F110W$-$F205W & $K_s$$-$F205W & 0.0001 & $ 0.029$ & $-0.031$ & $ 0.078$ & $-0.067$ & $0.005$ & $-0.304 \sim 0.928$ \\ F110W$-$F205W & $K_s$$-$F205W & 0.001 & $ 0.030$ & $-0.032$ & $ 0.055$ & $-0.036$ & $0.004$ & $-0.210 \sim 1.122$ \\ F110W$-$F205W & $K_s$$-$F205W & 0.019 & $ 0.031$ & $-0.032$ & $ 0.026$ & $-0.011$ & $0.006$ & $-0.272 \sim 1.621$ \\ F110W$-$F205W & $K_s$$-$F205W & 0.03 & $ 0.031$ & $-0.027$ & $ 0.014$ & $-0.007$ & $0.005$ & $-0.166 \sim 1.615$ \\ F110W$-$F205W & F222M$-$F205W & 0.0001 & $-0.001$ & $-0.041$ & $ 0.085$ & $-0.079$ & $0.007$ & $-0.304 \sim 0.928$ \\ F110W$-$F205W & F222M$-$F205W & 0.001 & $-0.000$ & $-0.042$ & $ 0.050$ & $-0.033$ & $0.004$ & $-0.210 \sim 1.122$ \\ F110W$-$F205W & F222M$-$F205W & 0.019 & $ 0.001$ & $-0.043$ & $ 0.018$ & $-0.006$ & $0.006$ & $-0.272 \sim 1.621$ \\ F110W$-$F205W & F222M$-$F205W & 0.03 & $ 0.001$ & $-0.037$ & $ 0.005$ & $-0.002$ & $0.004$ & $-0.166 \sim 1.615$ \\ F110W$-$F222M & $K$ $-$F222M & 0.0001 & $ 0.031$ & $ 0.001$ & $ 0.001$ & $ 0.008$ & $0.005$ & $-0.324 \sim 0.955$ \\ F110W$-$F222M & $K$ $-$F222M & 0.001 & $ 0.030$ & $-0.001$ & $ 0.004$ & $ 0.021$ & $0.005$ & $-0.222 \sim 1.152$ \\ F110W$-$F222M & $K$ $-$F222M & 0.019 & $ 0.028$ & $-0.003$ & $ 0.063$ & $-0.024$ & $0.011$ & $-0.291 \sim 1.665$ \\ F110W$-$F222M & $K$ $-$F222M & 0.03 & $ 0.028$ & $-0.009$ & $ 0.076$ & $-0.029$ & $0.006$ & $-0.176 \sim 1.669$ \\ F110W$-$F222M & $K'$ $-$F222M & 0.0001 & $ 0.030$ & $ 0.021$ & $-0.033$ & $ 0.032$ & $0.005$ & $-0.324 \sim 0.955$ \\ F110W$-$F222M & $K'$ $-$F222M & 0.001 & $ 0.029$ & $ 0.022$ & $-0.013$ & $ 0.003$ & $0.003$ & $-0.222 \sim 1.152$ \\ F110W$-$F222M & $K'$ $-$F222M & 0.019 & $ 0.030$ & $ 0.024$ & $-0.017$ & $ 0.005$ & $0.004$ & $-0.291 \sim 1.665$ \\ F110W$-$F222M & $K'$ $-$F222M & 0.03 & $ 0.029$ & $ 0.024$ & $-0.014$ & $ 0.004$ & $0.003$ & $-0.176 \sim 1.669$ \\ F110W$-$F222M & $K_s$$-$F222M & 0.0001 & $ 0.030$ & $ 0.010$ & $-0.006$ & $ 0.011$ & $0.005$ & $-0.324 \sim 0.955$ \\ F110W$-$F222M & $K_s$$-$F222M & 0.001 & $ 0.030$ & $ 0.009$ & $ 0.005$ & $-0.003$ & $0.004$ & $-0.222 \sim 1.152$ \\ F110W$-$F222M & $K_s$$-$F222M & 0.019 & $ 0.030$ & $ 0.011$ & $ 0.007$ & $-0.004$ & $0.005$ & $-0.291 \sim 1.665$ \\ F110W$-$F222M & $K_s$$-$F222M & 0.03 & $ 0.030$ & $ 0.010$ & $ 0.009$ & $-0.005$ & $0.004$ & $-0.176 \sim 1.669$ \\ F110W$-$F222M & F205W$-$F222M & 0.0001 & $ 0.001$ & $ 0.038$ & $-0.076$ & $ 0.070$ & $0.007$ & $-0.324 \sim 0.955$ \\ F110W$-$F222M & F205W$-$F222M & 0.001 & $ 0.000$ & $ 0.040$ & $-0.045$ & $ 0.030$ & $0.004$ & $-0.222 \sim 1.152$ \\ F110W$-$F222M & F205W$-$F222M & 0.019 & $-0.001$ & $ 0.041$ & $-0.017$ & $ 0.005$ & $0.006$ & $-0.291 \sim 1.665$ \\ F110W$-$F222M & F205W$-$F222M & 0.03 & $-0.001$ & $ 0.036$ & $-0.004$ & $ 0.002$ & $0.004$ & $-0.176 \sim 1.669$ \\ \enddata \tablecomments{Magnitude differences are fitted to a function $[{\rm Mag\, Diff}] = c_0 + c_1[{\rm Color}] + c_2[{\rm Color}]^2 + c_3[{\rm Color}]^3$. Only the data points that have $\log T_{eff} \ge 3500$~K and $\log g \ge 0$ were considered for the fitting.} \tablenotetext{a}{The largest absolute residual.} \tablenotetext{b}{Color range where the fit is valid.} \end{deluxetable} \clearpage \begin{deluxetable}{cclrrrcc} \tabletypesize{\scriptsize} \tablecolumns{8} \tablewidth{0pt} \tablecaption{ \label{table:trans2}Best-Fit Coefficients for Magnitude Differences ($K > 4$~mag)} \tablehead{ \colhead{} & \colhead{Magnitude} & \colhead{} & \colhead{} & \colhead{} & \colhead{} & \colhead{Residual\tablenotemark{a}} & \colhead{Fitting Range\tablenotemark{b}} \\ \colhead{Color} & \colhead{Difference} & \colhead{$Z$} & \colhead{$c_0$} & \colhead{$c_1$} & \colhead{$c_2$} & \colhead{(mag)} & \colhead{(mag $\sim$ mag)} } \startdata $J - K$ & $K'$ $-$$K$ & 0.0001 & $-0.012$ & $ 0.047$ & $-0.003$ & $0.001$ & $ 0.332 \sim 0.772$ \\ $J - K$ & $K'$ $-$$K$ & 0.001 & $ 0.034$ & $-0.126$ & $ 0.127$ & $0.003$ & $ 0.338 \sim 0.893$ \\ $J - K$ & $K'$ $-$$K$ & 0.019 & $ 0.102$ & $-0.331$ & $ 0.252$ & $0.004$ & $ 0.539 \sim 0.992$ \\ $J - K$ & $K'$ $-$$K$ & 0.03 & $ 0.129$ & $-0.401$ & $ 0.291$ & $0.003$ & $ 0.559 \sim 0.987$ \\ $J - K$ & $K_s$$-$$K$ & 0.0001 & $-0.010$ & $ 0.042$ & $-0.019$ & $0.001$ & $ 0.332 \sim 0.772$ \\ $J - K$ & $K_s$$-$$K$ & 0.001 & $ 0.018$ & $-0.059$ & $ 0.055$ & $0.002$ & $ 0.338 \sim 0.893$ \\ $J - K$ & $K_s$$-$$K$ & 0.019 & $ 0.055$ & $-0.182$ & $ 0.134$ & $0.003$ & $ 0.539 \sim 0.992$ \\ $J - K$ & $K_s$$-$$K$ & 0.03 & $ 0.073$ & $-0.229$ & $ 0.160$ & $0.002$ & $ 0.559 \sim 0.987$ \\ $J - K$ & F205W$-$$K$ & 0.0001 & $-0.049$ & $ 0.082$ & $ 0.000$ & $0.001$ & $ 0.332 \sim 0.772$ \\ $J - K$ & F205W$-$$K$ & 0.001 & $ 0.026$ & $-0.203$ & $ 0.217$ & $0.004$ & $ 0.338 \sim 0.893$ \\ $J - K$ & F205W$-$$K$ & 0.019 & $ 0.127$ & $-0.490$ & $ 0.386$ & $0.007$ & $ 0.539 \sim 0.992$ \\ $J - K$ & F205W$-$$K$ & 0.03 & $ 0.173$ & $-0.601$ & $ 0.447$ & $0.004$ & $ 0.559 \sim 0.987$ \\ $J - K$ & F222M$-$$K$ & 0.0001 & $-0.026$ & $-0.023$ & $ 0.003$ & $0.001$ & $ 0.332 \sim 0.772$ \\ $J - K$ & F222M$-$$K$ & 0.001 & $-0.027$ & $-0.016$ & $-0.007$ & $0.001$ & $ 0.338 \sim 0.893$ \\ $J - K$ & F222M$-$$K$ & 0.019 & $-0.030$ & $-0.027$ & $ 0.002$ & $0.001$ & $ 0.539 \sim 0.992$ \\ $J - K$ & F222M$-$$K$ & 0.03 & $-0.035$ & $-0.020$ & $-0.002$ & $0.001$ & $ 0.559 \sim 0.987$ \\ $J - K'$ & $K$ $-$$K'$ & 0.0001 & $ 0.012$ & $-0.050$ & $ 0.004$ & $0.001$ & $ 0.328 \sim 0.750$ \\ $J - K'$ & $K$ $-$$K'$ & 0.001 & $-0.037$ & $ 0.138$ & $-0.140$ & $0.003$ & $ 0.334 \sim 0.871$ \\ $J - K'$ & $K$ $-$$K'$ & 0.019 & $-0.127$ & $ 0.401$ & $-0.300$ & $0.005$ & $ 0.543 \sim 0.975$ \\ $J - K'$ & $K$ $-$$K'$ & 0.03 & $-0.174$ & $ 0.520$ & $-0.369$ & $0.003$ & $ 0.593 \sim 0.971$ \\ $J - K'$ & $K_s$$-$$K'$ & 0.0001 & $ 0.002$ & $-0.006$ & $-0.017$ & $0.001$ & $ 0.328 \sim 0.750$ \\ $J - K'$ & $K_s$$-$$K'$ & 0.001 & $-0.018$ & $ 0.074$ & $-0.080$ & $0.003$ & $ 0.334 \sim 0.871$ \\ $J - K'$ & $K_s$$-$$K'$ & 0.019 & $-0.060$ & $ 0.184$ & $-0.143$ & $0.002$ & $ 0.543 \sim 0.975$ \\ $J - K'$ & $K_s$$-$$K'$ & 0.03 & $-0.080$ & $ 0.233$ & $-0.172$ & $0.001$ & $ 0.593 \sim 0.971$ \\ $J - K'$ & F205W$-$$K'$ & 0.0001 & $-0.038$ & $ 0.036$ & $ 0.004$ & $0.001$ & $ 0.328 \sim 0.750$ \\ $J - K'$ & F205W$-$$K'$ & 0.001 & $-0.006$ & $-0.087$ & $ 0.101$ & $0.002$ & $ 0.334 \sim 0.871$ \\ $J - K'$ & F205W$-$$K'$ & 0.019 & $ 0.040$ & $-0.202$ & $ 0.166$ & $0.003$ & $ 0.543 \sim 0.975$ \\ $J - K'$ & F205W$-$$K'$ & 0.03 & $ 0.077$ & $-0.286$ & $ 0.211$ & $0.002$ & $ 0.593 \sim 0.971$ \\ $J - K'$ & F222M$-$$K'$ & 0.0001 & $-0.013$ & $-0.074$ & $ 0.007$ & $0.001$ & $ 0.328 \sim 0.750$ \\ $J - K'$ & F222M$-$$K'$ & 0.001 & $-0.065$ & $ 0.125$ & $-0.150$ & $0.004$ & $ 0.334 \sim 0.871$ \\ $J - K'$ & F222M$-$$K'$ & 0.019 & $-0.160$ & $ 0.383$ & $-0.305$ & $0.004$ & $ 0.543 \sim 0.975$ \\ $J - K'$ & F222M$-$$K'$ & 0.03 & $-0.214$ & $ 0.513$ & $-0.380$ & $0.003$ & $ 0.593 \sim 0.971$ \\ $J - K_s$ & $K$ $-$$K_s$ & 0.0001 & $ 0.010$ & $-0.043$ & $ 0.020$ & $0.001$ & $ 0.330 \sim 0.761$ \\ $J - K_s$ & $K$ $-$$K_s$ & 0.001 & $-0.018$ & $ 0.060$ & $-0.056$ & $0.002$ & $ 0.335 \sim 0.885$ \\ $J - K_s$ & $K$ $-$$K_s$ & 0.019 & $-0.060$ & $ 0.197$ & $-0.143$ & $0.003$ & $ 0.544 \sim 0.989$ \\ $J - K_s$ & $K$ $-$$K_s$ & 0.03 & $-0.084$ & $ 0.257$ & $-0.177$ & $0.002$ & $ 0.594 \sim 0.986$ \\ $J - K_s$ & $K'$ $-$$K_s$ & 0.0001 & $-0.002$ & $ 0.006$ & $ 0.016$ & $0.001$ & $ 0.330 \sim 0.761$ \\ $J - K_s$ & $K'$ $-$$K_s$ & 0.001 & $ 0.017$ & $-0.069$ & $ 0.075$ & $0.003$ & $ 0.335 \sim 0.885$ \\ $J - K_s$ & $K'$ $-$$K_s$ & 0.019 & $ 0.054$ & $-0.165$ & $ 0.128$ & $0.002$ & $ 0.544 \sim 0.989$ \\ $J - K_s$ & $K'$ $-$$K_s$ & 0.03 & $ 0.070$ & $-0.204$ & $ 0.151$ & $0.001$ & $ 0.594 \sim 0.986$ \\ $J - K_s$ & F205W$-$$K_s$ & 0.0001 & $-0.040$ & $ 0.042$ & $ 0.019$ & $0.001$ & $ 0.330 \sim 0.761$ \\ $J - K_s$ & F205W$-$$K_s$ & 0.001 & $ 0.010$ & $-0.149$ & $ 0.168$ & $0.003$ & $ 0.335 \sim 0.885$ \\ $J - K_s$ & F205W$-$$K_s$ & 0.019 & $ 0.085$ & $-0.343$ & $ 0.276$ & $0.005$ & $ 0.544 \sim 0.989$ \\ $J - K_s$ & F205W$-$$K_s$ & 0.03 & $ 0.134$ & $-0.454$ & $ 0.337$ & $0.003$ & $ 0.594 \sim 0.986$ \\ $J - K_s$ & F222M$-$$K_s$ & 0.0001 & $-0.015$ & $-0.068$ & $ 0.024$ & $0.001$ & $ 0.330 \sim 0.761$ \\ $J - K_s$ & F222M$-$$K_s$ & 0.001 & $-0.045$ & $ 0.046$ & $-0.064$ & $0.002$ & $ 0.335 \sim 0.885$ \\ $J - K_s$ & F222M$-$$K_s$ & 0.019 & $-0.092$ & $ 0.174$ & $-0.144$ & $0.003$ & $ 0.544 \sim 0.989$ \\ $J - K_s$ & F222M$-$$K_s$ & 0.03 & $-0.122$ & $ 0.245$ & $-0.184$ & $0.002$ & $ 0.594 \sim 0.986$ \\ F110W$-$F205W & $K$ $-$F205W & 0.0001 & $ 0.054$ & $-0.071$ & $-0.000$ & $0.001$ & $ 0.454 \sim 0.961$ \\ F110W$-$F205W & $K$ $-$F205W & 0.001 & $-0.050$ & $ 0.223$ & $-0.176$ & $0.005$ & $ 0.484 \sim 1.107$ \\ F110W$-$F205W & $K$ $-$F205W & 0.019 & $-0.162$ & $ 0.460$ & $-0.278$ & $0.008$ & $ 0.724 \sim 1.252$ \\ F110W$-$F205W & $K$ $-$F205W & 0.03 & $-0.231$ & $ 0.584$ & $-0.328$ & $0.005$ & $ 0.784 \sim 1.257$ \\ F110W$-$F205W & $K'$ $-$F205W & 0.0001 & $ 0.040$ & $-0.030$ & $-0.003$ & $0.001$ & $ 0.454 \sim 0.961$ \\ F110W$-$F205W & $K'$ $-$F205W & 0.001 & $-0.002$ & $ 0.088$ & $-0.074$ & $0.002$ & $ 0.484 \sim 1.107$ \\ F110W$-$F205W & $K'$ $-$F205W & 0.019 & $-0.036$ & $ 0.150$ & $-0.097$ & $0.003$ & $ 0.724 \sim 1.252$ \\ F110W$-$F205W & $K'$ $-$F205W & 0.03 & $-0.068$ & $ 0.203$ & $-0.118$ & $0.002$ & $ 0.784 \sim 1.257$ \\ F110W$-$F205W & $K_s$$-$F205W & 0.0001 & $ 0.042$ & $-0.033$ & $-0.015$ & $0.001$ & $ 0.454 \sim 0.961$ \\ F110W$-$F205W & $K_s$$-$F205W & 0.001 & $-0.026$ & $ 0.162$ & $-0.133$ & $0.004$ & $ 0.484 \sim 1.107$ \\ F110W$-$F205W & $K_s$$-$F205W & 0.019 & $-0.095$ & $ 0.291$ & $-0.182$ & $0.005$ & $ 0.724 \sim 1.252$ \\ F110W$-$F205W & $K_s$$-$F205W & 0.03 & $-0.141$ & $ 0.370$ & $-0.214$ & $0.003$ & $ 0.784 \sim 1.257$ \\ F110W$-$F205W & F222M$-$F205W & 0.0001 & $ 0.030$ & $-0.091$ & $ 0.002$ & $0.001$ & $ 0.454 \sim 0.961$ \\ F110W$-$F205W & F222M$-$F205W & 0.001 & $-0.079$ & $ 0.218$ & $-0.186$ & $0.005$ & $ 0.484 \sim 1.107$ \\ F110W$-$F205W & F222M$-$F205W & 0.019 & $-0.190$ & $ 0.438$ & $-0.277$ & $0.008$ & $ 0.724 \sim 1.252$ \\ F110W$-$F205W & F222M$-$F205W & 0.03 & $-0.266$ & $ 0.569$ & $-0.330$ & $0.005$ & $ 0.784 \sim 1.257$ \\ F110W$-$F222M & $K$ $-$F222M & 0.0001 & $ 0.025$ & $ 0.018$ & $-0.002$ & $0.001$ & $ 0.465 \sim 1.017$ \\ F110W$-$F222M & $K$ $-$F222M & 0.001 & $ 0.027$ & $ 0.010$ & $ 0.006$ & $0.001$ & $ 0.496 \sim 1.170$ \\ F110W$-$F222M & $K$ $-$F222M & 0.019 & $ 0.025$ & $ 0.029$ & $-0.005$ & $0.001$ & $ 0.742 \sim 1.322$ \\ F110W$-$F222M & $K$ $-$F222M & 0.03 & $ 0.031$ & $ 0.023$ & $-0.002$ & $0.001$ & $ 0.806 \sim 1.327$ \\ F110W$-$F222M & $K'$ $-$F222M & 0.0001 & $ 0.011$ & $ 0.055$ & $-0.003$ & $0.001$ & $ 0.465 \sim 1.017$ \\ F110W$-$F222M & $K'$ $-$F222M & 0.001 & $ 0.069$ & $-0.103$ & $ 0.089$ & $0.004$ & $ 0.496 \sim 1.170$ \\ F110W$-$F222M & $K'$ $-$F222M & 0.019 & $ 0.122$ & $-0.208$ & $ 0.131$ & $0.004$ & $ 0.742 \sim 1.322$ \\ F110W$-$F222M & $K'$ $-$F222M & 0.03 & $ 0.152$ & $-0.259$ & $ 0.151$ & $0.003$ & $ 0.806 \sim 1.327$ \\ F110W$-$F222M & $K_s$$-$F222M & 0.0001 & $ 0.013$ & $ 0.053$ & $-0.014$ & $0.001$ & $ 0.465 \sim 1.017$ \\ F110W$-$F222M & $K_s$$-$F222M & 0.001 & $ 0.048$ & $-0.043$ & $ 0.042$ & $0.002$ & $ 0.496 \sim 1.170$ \\ F110W$-$F222M & $K_s$$-$F222M & 0.019 & $ 0.078$ & $-0.103$ & $ 0.068$ & $0.003$ & $ 0.742 \sim 1.322$ \\ F110W$-$F222M & $K_s$$-$F222M & 0.03 & $ 0.100$ & $-0.140$ & $ 0.082$ & $0.002$ & $ 0.806 \sim 1.327$ \\ F110W$-$F222M & F205W$-$F222M & 0.0001 & $-0.028$ & $ 0.083$ & $-0.001$ & $0.001$ & $ 0.465 \sim 1.017$ \\ F110W$-$F222M & F205W$-$F222M & 0.001 & $ 0.065$ & $-0.173$ & $ 0.148$ & $0.005$ & $ 0.496 \sim 1.170$ \\ F110W$-$F222M & F205W$-$F222M & 0.019 & $ 0.139$ & $-0.313$ & $ 0.200$ & $0.006$ & $ 0.742 \sim 1.322$ \\ F110W$-$F222M & F205W$-$F222M & 0.03 & $ 0.195$ & $-0.404$ & $ 0.235$ & $0.004$ & $ 0.806 \sim 1.327$ \\ \enddata \tablecomments{Magnitude differences are fitted to a function ${\rm [Mag\, Diff]} = c_0 + c_1[{\rm Color}] + c_2[{\rm Color}]^2$. Only the data points that have $\log T_{eff} \ge 3500$~K and $\log g \ge 0$ were considered for the fitting.} \tablenotetext{a}{The largest absolute residual.} \tablenotetext{b}{Color range where the fit is valid.} \end{deluxetable} \clearpage \begin{deluxetable}{ccccccc} \tablecolumns{7} \tablewidth{0pt} \tablecaption{ \label{table:lambdac}Central Wavelength $\lambda_c$ ($\mu$m)} \tablehead{ \colhead{$J$} & \colhead{$K$} & \colhead{$K'$} & \colhead{$K_s$} & \colhead{F110W} & \colhead{F205W} & \colhead{F222M} } \startdata 1.237 & 2.212 & 2.114 & 2.160 & 1.140 & 2.079 & 2.219 \\ \enddata \tablecomments{$\lambda_c$ is defined by eq. (8) of Paper I.} \end{deluxetable} \clearpage \begin{deluxetable}{lcccccc} \tabletypesize{\scriptsize} \tablecolumns{7} \tablewidth{0pt} \tablecaption{ \label{table:alpha_eff}Averages and Standard Deviations of $\alpha_{eff}$ Values} \tablehead{ \multicolumn{2}{c}{Isochrone Model} & \colhead{} & \colhead{} & \colhead{} & \colhead{} & \colhead{} \\ \cline{1-2} \colhead{$Z$} & \colhead{Age} & \colhead{$J-K$} & \colhead{$J-K'$} & \colhead{$J-K_s$} & \colhead{F110W$-$F205W} & \colhead{F110W$-$F222M} } \startdata 0.0001 & $10^7$ & 1.610$\pm$0.000 & 1.608$\pm$0.000 & 1.610$\pm$0.000 & 1.479$\pm$0.001 & 1.500$\pm$0.002 \\ 0.0001 & $10^8$ & 1.608$\pm$0.003 & 1.605$\pm$0.004 & 1.607$\pm$0.004 & 1.467$\pm$0.011 & 1.486$\pm$0.011 \\ 0.0001 & $10^9$ & 1.600$\pm$0.004 & 1.597$\pm$0.005 & 1.598$\pm$0.004 & 1.440$\pm$0.013 & 1.460$\pm$0.012 \\ 0.0001 & $10^{10}$ & 1.596$\pm$0.003 & 1.593$\pm$0.003 & 1.595$\pm$0.003 & 1.429$\pm$0.008 & 1.450$\pm$0.007 \\ 0.001 & $6.3 \times 10^7$ & 1.608$\pm$0.004 & 1.605$\pm$0.004 & 1.607$\pm$0.004 & 1.467$\pm$0.012 & 1.487$\pm$0.012 \\ 0.001 & $10^8$ & 1.605$\pm$0.006 & 1.603$\pm$0.006 & 1.604$\pm$0.006 & 1.459$\pm$0.017 & 1.478$\pm$0.017 \\ 0.001 & $10^9$ & 1.595$\pm$0.004 & 1.593$\pm$0.003 & 1.595$\pm$0.003 & 1.430$\pm$0.008 & 1.451$\pm$0.007 \\ 0.001 & $10^{10}$ & 1.591$\pm$0.005 & 1.590$\pm$0.003 & 1.592$\pm$0.003 & 1.421$\pm$0.010 & 1.444$\pm$0.008 \\ 0.019 & $10^7$ & 1.610$\pm$0.000 & 1.608$\pm$0.000 & 1.610$\pm$0.000 & 1.478$\pm$0.001 & 1.498$\pm$0.002 \\ 0.019 & $10^8$ & 1.599$\pm$0.011 & 1.599$\pm$0.009 & 1.600$\pm$0.009 & 1.448$\pm$0.028 & 1.470$\pm$0.025 \\ 0.019 & $10^9$ & 1.588$\pm$0.006 & 1.590$\pm$0.004 & 1.590$\pm$0.004 & 1.418$\pm$0.012 & 1.442$\pm$0.010 \\ 0.019 & $10^{10}$ & 1.582$\pm$0.005 & 1.586$\pm$0.003 & 1.587$\pm$0.003 & 1.406$\pm$0.011 & 1.432$\pm$0.009 \\ 0.03 & $6.3 \times 10^7$ & 1.606$\pm$0.007 & 1.605$\pm$0.006 & 1.606$\pm$0.006 & 1.465$\pm$0.019 & 1.485$\pm$0.018 \\ 0.03 & $10^8$ & 1.599$\pm$0.012 & 1.599$\pm$0.009 & 1.600$\pm$0.010 & 1.448$\pm$0.029 & 1.470$\pm$0.026 \\ 0.03 & $10^9$ & 1.586$\pm$0.006 & 1.589$\pm$0.004 & 1.590$\pm$0.004 & 1.415$\pm$0.012 & 1.440$\pm$0.010 \\ 0.03 & $10^{10}$ & 1.581$\pm$0.005 & 1.585$\pm$0.003 & 1.586$\pm$0.003 & 1.403$\pm$0.011 & 1.429$\pm$0.009 \\ \enddata \tablecomments{Data are presented in the form of average $\pm$ standard deviation. The average and standard deviation values are calculated from the data points of each isochrone whose intrinsic $K$ magnitudes are between $-6$ and 0~mag.} \end{deluxetable} \clearpage \begin{deluxetable}{lccrrrrrcrrrrr} \tabletypesize{\tiny} \tablecolumns{14} \tablewidth{0pt} \tablecaption{\label{table:nonlin}Extinction Behavior of {\it HST} NICMOS Filter Pairs} \tablehead{ \multicolumn{2}{c}{Isochrone Model} & \colhead{} & \multicolumn{5}{c}{F110W$-$F205W} & \colhead{} & \multicolumn{5}{c}{F110W$-$F222M} \\ \cline{1-2} \cline{4-8} \cline{10-14} \colhead{$Z$} & \colhead{Age} & \colhead{} & \colhead{$c_0$} & \colhead{$c_1$} & \colhead{$c_2$} & \colhead{$c_3$} & \colhead{$\sigma(A^{est})$} & \colhead{} & \colhead{$c_0$} & \colhead{$c_1$} & \colhead{$c_2$} & \colhead{$c_3$} & \colhead{$\sigma(A^{est})$} } \startdata 0.0001 & $10^7$ & & 7.07E-04 & 2.07E-01 & $-$1.08E-01 & 7.77E-03 & 0.011 & & $-$9.83E-05 & 2.67E-01 & $-$1.18E-01 & 7.01E-03 & 0.013 \nl 0.0001 & $10^8$ & & 9.78E-04 & 1.75E-01 & $-$1.05E-01 & 7.70E-03 & 0.077 & & 1.18E-04 & 2.37E-01 & $-$1.16E-01 & 7.07E-03 & 0.080 \nl 0.0001 & $10^9$ & & 1.61E-03 & 1.14E-01 & $-$1.00E-01 & 7.88E-03 & 0.082 & & 6.21E-04 & 1.84E-01 & $-$1.14E-01 & 7.41E-03 & 0.083 \nl 0.0001 & $10^{10}$ & & 1.82E-03 & 8.93E-02 & $-$9.82E-02 & 7.95E-03 & 0.049 & & 7.90E-04 & 1.62E-01 & $-$1.13E-01 & 7.53E-03 & 0.050 \nl 0.001 & $6.3 \times 10^7$ & & 9.79E-04 & 1.75E-01 & $-$1.05E-01 & 7.69E-03 & 0.082 & & 1.01E-04 & 2.38E-01 & $-$1.16E-01 & 7.05E-03 & 0.084 \nl 0.001 & $10^8$ & & 1.16E-03 & 1.58E-01 & $-$1.04E-01 & 7.77E-03 & 0.115 & & 2.30E-04 & 2.21E-01 & $-$1.16E-01 & 7.17E-03 & 0.118 \nl 0.001 & $10^9$ & & 1.75E-03 & 9.25E-02 & $-$9.86E-02 & 7.93E-03 & 0.049 & & 6.72E-04 & 1.66E-01 & $-$1.13E-01 & 7.53E-03 & 0.047 \nl 0.001 & $10^{10}$ & & 1.86E-03 & 7.52E-02 & $-$9.74E-02 & 7.98E-03 & 0.057 & & 8.48E-04 & 1.51E-01 & $-$1.12E-01 & 7.57E-03 & 0.054 \nl 0.019 & $10^7$ & & 7.77E-04 & 2.03E-01 & $-$1.08E-01 & 7.74E-03 & 0.010 & & $-$7.16E-05 & 2.64E-01 & $-$1.18E-01 & 7.01E-03 & 0.011 \nl 0.019 & $10^8$ & & 1.33E-03 & 1.36E-01 & $-$1.02E-01 & 7.82E-03 & 0.172 & & 3.57E-04 & 2.05E-01 & $-$1.15E-01 & 7.23E-03 & 0.171 \nl 0.019 & $10^9$ & & 1.93E-03 & 6.95E-02 & $-$9.71E-02 & 7.97E-03 & 0.068 & & 8.10E-04 & 1.46E-01 & $-$1.12E-01 & 7.55E-03 & 0.066 \nl 0.019 & $10^{10}$ & & 2.07E-03 & 4.13E-02 & $-$9.42E-02 & 7.93E-03 & 0.064 & & 8.75E-04 & 1.22E-01 & $-$1.09E-01 & 7.48E-03 & 0.062 \nl 0.03 & $6.3 \times 10^7$ & & 1.03E-03 & 1.72E-01 & $-$1.05E-01 & 7.73E-03 & 0.120 & & 1.43E-04 & 2.35E-01 & $-$1.16E-01 & 7.06E-03 & 0.122 \nl 0.03 & $10^8$ & & 1.30E-03 & 1.36E-01 & $-$1.02E-01 & 7.83E-03 & 0.182 & & 3.61E-04 & 2.06E-01 & $-$1.15E-01 & 7.25E-03 & 0.180 \nl 0.03 & $10^9$ & & 1.92E-03 & 6.25E-02 & $-$9.61E-02 & 7.93E-03 & 0.071 & & 8.03E-04 & 1.40E-01 & $-$1.11E-01 & 7.51E-03 & 0.068 \nl 0.03 & $10^{10}$ & & 2.04E-03 & 3.35E-02 & $-$9.28E-02 & 7.84E-03 & 0.063 & & 8.45E-04 & 1.15E-01 & $-$1.08E-01 & 7.44E-03 & 0.063 \nl \enddata \tablecomments{Coefficients of best-fit third-order polynomials for the extinction curves in Figure~\ref{fig:adiff1} for {\it HST} NICMOS filter pairs. The difference of the estimated extinction and the true extinction is fitted to a function $[A_Y^{est}-A_Y] = c_0 + c_1[A_Y^{est}] + c_2[A_Y^{est}]^2 + c_3[A_Y^{est}]^3$; $\sigma$($A^{est}$) is the average of the standard deviations of $A_Y^{est}-A_Y$ values.} \end{deluxetable} \clearpage \clearpage
Title: The Ultra Luminous X-ray sources in the High Velocity System of NGC 1275
Abstract: We report the results of a study of X-ray point sources coincident with the High Velocity System (HVS) projected in front of NGC 1275. A very deep X-ray image of the core of the Perseus cluster made with the Chandra Observatory has been used. We find a population of Ultra-Luminous X-ray sources (ULX; 7 sources with LX [0.5-7 keV] > 7x10^39 erg/s). As with the ULX populations in the Antennae and Cartwheel galaxies, those in the HVS are associated with a region of very active star formation. Several sources have possible optical counterparts found on HST images, although the X-ray brightest one does not. Absorbed power-law models fit the X-ray spectra, with most having a photon index between 2 and 3.
https://export.arxiv.org/pdf/astro-ph/0601180
\label{firstpage} \begin{keywords} galaxies: clusters: individual: Perseus - ULX - galaxies: individual: NGC\,1275 \end{keywords} \section{Introduction} The study of Ultra-Luminous X-ray sources (ULX) has been greatly expanded by the high spatial resolution and spectral grasp of the \emph{Chandra} and \emph{XMM-Newton} observatories, respectively. ULX sources (Fabbiano \& White 2003; Miller \& Colbert 2004) have 2--10~keV X-ray luminosities exceeding $10^{39}\ergps$ and are found some distance from the centres of galaxies; they are not active galactic nuclei. Their luminosity exceeds that for a $10\Msun$ black hole accreting at the Eddington limit which radiates isotropically and so have created much interest in the possibility that they contain even higher mass black holes, such as InterMediate Black Holes (IMBH) of $\sim10^3\Msun$ (Makishima et al 2000; Miller, Fabian \& Miller 2004). Alternatively they may appear so luminous because of beaming (Reynolds et al 1999; King et al 2001, Zezas \& Fabbiano 2002) or due to super Eddington accretion (Begelman 2002). ULX are most common in starburst galaxies and in very active star-forming regions, such as in the Antennae and the Cartwheel galaxy, where populations of tens of them are found (Zezas et al 2002; Gao et al 2003; Wolter \& Trinchieri 2004). In some cases variability rules out the possibility that they are just clusters of lower-luminosity objects. The origin of IMBH is unclear. They may form as a result of binary interactions in dense stellar environments (Portegies Zwart \& McMillan 2002). A comparison of IMBH ULX candidates with a number of well known stellar-mass black holes candidates (BHC; Miller et al 2004) demonstrates that the ULX are more luminous but have cooler thermal disk components than standard stellar-mass BHC. Therefore, ULX in this sample are clearly different from the sample of stellar-mass BHC and are consistent with being IMBH. Here we report on the discovery of a population of 8 point X-ray sources to the N of the nucleus of NGC\,1275, which is the central galaxy in the Perseus cluster. All exceed $10^{39}\ergps$ in X-ray luminosity, and 7 are formally ULX, if they are at the distance of the cluster. The spatial region where they lie coincides with the High Velocity System of NGC\,1275. We assume that they are part of that system. We see no other point sources (apart from the nucleus) over the body of NGC\,1275 (Fig.~1). NGC\,1275 is embedded in a complex multiphase environment. Optical imaging and spectroscopy first established the existence of two distinct emission-line system toward NGC\,1275: a low-velocity component associated with the galaxy itself at 5200 km $\rm{s^{-1}}$ and a high-velocity component at 8200 km $\rm{s^{-1}}$ projected nearby on the sky (Minkowski 1955, 1957). This latter component is associated with a small gas-rich galaxy falling into the cluster along our line of sight (Haschick, Crane \& van der Hulst 1982). A merger scenario has been proposed (Minkowski 1955, 1957). However, interaction of the low and/or high-velocity system with a third gas-rich galaxy or system of galaxies (Holtzman et al. 1992; Conselice, Gallagher \& Wyse 2001), or influences from the surrounding dense intracluster medium (ICM) (Sarazin 1988; Boroson 1990; Caulet et al. 1992) have been discussed. Deep \emph{Chandra} observations have clarified the position of the High Velocity System (HVS). The depth of the observed X-ray absorption (e.g. Fig.~1) is nor infilled by emission from hot gas projected along the line-of-sight so the HVS must lie well in front of NGC\,1275. Gillmon, Sanders \& Fabian (2004) have estimated a lower limit on the distance of the HVS from the nucleus of 57 kpc. The low- and high-velocity system are therefore not yet directly interacting. The HVS $\emph{is}$ however strongly interacting with the ICM of the Perseus cluster, which has triggered strong star formation. In this paper we describe the detailed analysis of the X-ray spatial and spectral properties of the discrete sources in the high velocity system. The paper is organized as follows: in Sect. 2 and 3 we present reduction and results from the imaging analysis and spectral analysis, respectively; in Sect. 4 we discuss the results; and Sect. 5 summarizes our findings. Throughout this paper we use a redshift of 0.018 and $H_0=70~\rm{km~s^{-1}~Mpc^{-1}}$. This gives a luminosity distance to the cluster of 80 Mpc; 1 arcsec corresponds to a physical distance of 370~pc. \section{Imaging analysis} The \emph{Chandra} datasets included in this analysis are listed in Table~\ref{tab:obs}. The total exposure time, after removing periods containing flares, is 890~ks. To prepare the data for analysis, all of the datasets were reprocessed to use the latest appropriate gain file (acisD2000-01-29gain\_ctiN0003). The datasets analysed each used an aimpoint on the ACIS-S3 CCD. The datasets were filtered using the lightcurve in the 2.5 to 7~keV band on ACIS-S1 CCD, which is a back-illuminated CCD like the ACIS-S3. The CIAO \textsc{lc\_clean} tool was used to remove periods 20 per~cent away from the median count rate for all the lightcurves. This procedure was not used for datasets 03209 and 04289 which did not include the S1 CCD, however no flares were seen in these observations on the S3 CCD. Each of the observations was reprojected to match the coordinate system of the 04952 observation. \begin{table*} \begin{tabular}{lllllll} Obs. ID & Sequence & Observation date & Exposure (ks) & Nominal roll (deg) & Pointing RA & Pointing Dec \\ \hline 3209 & 800209 & 2002-08-08 & 95.8 & 101.2 & 3:19:46.86 & +41:31:51.3 \\ 4289 & 800209 & 2002-08-10 & 95.4 & 101.2 & 3:19:46.86 & +41:31:51.3 \\ 6139 & 800397 & 2004-10-04 & 51.6 & 125.9 & 3:19:45.54 & +41:31:33.9 \\ 4946 & 800397 & 2004-10-06 & 22.7 & 127.2 & 3:19:45.44 & +41:31:33.2 \\ 4948 & 800398 & 2004-10-09 & 107.5 & 128.9 & 3:19:44.75 & +41:31:40.1 \\ 4947 & 800397 & 2004-10-11 & 28.7 & 130.6 & 3:19:45.17 & +41:31:31.3 \\ 4949 & 800398 & 2004-10-12 & 28.8 & 130.9 & 3:19:44.57 & +41:31:38.7 \\ 4950 & 800399 & 2004-10-12 & 73.4 & 131.1 & 3:19:43.97 & +41:31:46.1 \\ 4952 & 800400 & 2004-10-14 & 143.2 & 132.6 & 3:19:43.22 & +41:31:52.2 \\ 4951 & 800399 & 2004-10-17 & 91.4 & 135.2 & 3:19:43.57 & +41:31:42.6 \\ 4953 & 800400 & 2004-10-18 & 29.3 & 136.2 & 3:19:42.83 & +41:31:48.5 \\ 6145 & 800397 & 2004-10-19 & 83.1 & 137.7 & 3:19:44.66 & +41:31:26.7 \\ 6146 & 800398 & 2004-10-20 & 39.2 & 138.7 & 3:19:43.92 & +41:31:32.7 \\ \end{tabular} \caption{\emph{Chandra} observations included in this analysis. The exposure given is the time remaining after filtering the lightcurve for flares. All observations were taken with the aimpoint on the ACIS-S3 CCD. All positions are in J2000 coordinates.} \label{tab:obs} \end{table*} The 900~ks X-ray image covering the energy range $\rm{0.3-0.8~keV}$ is shown in Fig \ref{fig:HVS}. The bright NGC\,1275 nucleus is clearly seen at RA $3^h 19^m 48^s$ and Dec. +$41^o$30'42" (J2000) and the high-velocity system is seen in absorption to the north of the nucleus. The CIAO \textsc{celldetect} source detection routine was then used on the reprocessed level 2 event data to produce a preliminary list of point sources. The cell size ranges between 4 pixels to 8 pixels. This algorithm strongly depends on the local background and the detection cell in not adjustable to the size of the source. As the X-ray diffuse emission of the NGC 1275 is very strong, the source list may well include false detections in high background level regions. Therefore problematic sources embedded in such regions have been excluded in our analysis. Moreover, as mentioned above, we only included sources associated with the HVS. We have detected 8 bright sources close to the nucleus of NGC\,1275, located in the northern inner radio lobe of 3C 84. All of these source are embedded in the same region as the HVS (see Fig. \ref{fig:whole_picture}). There are no sources associated with the southern lobe (Fig. \ref{fig:whole_picture}), thus we assume these sources are associated with the HVS. Fig. \ref{fig:figure} (\emph{left}) shows the smoothed ACIS-S3 image in the 0.3--7.0 keV band, including numbered labels of all the detected sources (\emph{top}), centred on source labelled N3 (\emph{centre}) and centred on source N5 (\emph{bottom}). All the point-like sources are listed in Table~\ref{tab:positions}, showing their positions and count rates. \begin{table} \begin{center} \begin{tabular}{lcc} \hline \hline N & Position(J2000)& Count Rate \\ & & (counts $\rm{ ks^{-1}}$) \\ \hline 1.... & 03:19:48.736 +41:30:47.25 & 0.34$\pm$0.05 \\ 2.... & 03:19:48.166 +41:30:46.64 & 1.69$\pm$0.06 \\ 3.... & 03:19:48.090 +41:31:01.88 & 2.60$\pm$0.08 \\ 4.... & 03:19:47.994 +41:30:52.30 & 1.42$\pm$0.09\\ 5.... & 03:19:47.925 +41:30:47.50 & 1.19$\pm$0.09 \\ 6.... & 03:19:47.602 +41:30:47.01 & 0.74$\pm$0.06 \\ 7.... & 03:19:47.422 +41:30:51.93 & 0.95$\pm$0.08\\ 8.... & 03:19:47.214 +41:30:47.62 & 1.28$\pm$0.08\\ \hline \end{tabular} \caption{Positions of sources detected near the NGC\,1275 centre and displayed in Fig. \ref{fig:figure} (column 2) and count rate in the energy range between 0.5--7.0 keV (column 3).} \label{tab:positions} \end{center} \end{table} We have used archival \emph{HST} observations of NGC\,1275 in order to search for optical counterparts. The galaxy was imaged with the WFPC2 camera on \emph{HST} using the F814W ($\sim$ I, on 2001 November 6 with an exposure time of 1200~s) and F702W ($\sim$ R, on 1994 March 31 with an exposure time of 140~s) broad-band filters. Several coincidences between X-ray sources and optical knots of emission (F814W) can be seen in Fig. \ref{fig:figure} ({right}), showing the same regions as Fig. \ref{fig:figure} ({left}). The \emph{HST} image shows many highly absorbed features. When we compare in detail, sources N7 and N8 are located in star forming regions, while N2 and N6 have a point-like counterpart. Sources N1, N3, N4 and N5 have no optical identification. Therefore, we have found a possible correlation between compact X-ray sources and regions of vigorous star formation. The implications are discussed later. \begin{table} \begin{center} \begin{tabular}{lcccc} \hline \hline N & F${\rm _X}$/F$_{{\rm F814W}}$& F$_{\rm X}$/F$_{{\rm F702W}}$& ${\rm M_{F814W}}$ & ${\rm M_{F702W}}$ \\ & & & &\\ \hline 1.... & $>26.5$ & $>16.2$ & $>$22.6 & $>$22.1 \\ 2.... & 25.4 & 23.6 & 20.4 & 20.3 \\ 3.... & $>$18800 & ... & $>$26.8 & ... \\ 4.... & $>$1081 & $>$800 & $>$24.5 & $>$24.2 \\ 5.... & $>$123 & $>$60.6 & $>$22.6 & $>$21.9 \\ 6.... & 26.2 & 28.4 & 21.7 & 21.7 \\ 7.... & 76.7 & 51.4 & 22.3 & 21.9 \\ 8.... & 134 & 90.1 & 22.2 & 21.7 \\ \hline \end{tabular} \caption{Optical analysis. X-ray to optical ratios (columns 2 and 3) and magnitude determinations (columns 4 and 5) for the filters F814W and F702W, respectively, with the X-ray flux between 1.0--7.0~keV.} \label{tab:optical} \end{center} \end{table} In order to investigate the emission mechanism of these ULX, the X-ray to optical flux ratios have been computed between the F702W and F814W \emph{HST} broad-bands and 1.0--7.0~keV X-ray band. Preliminary processing of the raw images including corrections for the flat fielding was done remotely at the \emph{Space Telescope Science Institute} through the standard pipeline. For each frame, cosmic rays were removed by image combination, using the {\sc imcombine} routine in IRAF. After cosmic ray removal, the frames were added using task {\sc wmosaic} in STSDAS package. Photometric measurements were made with {\sc phot} task, within the NOAO package. Finally the fluxes and magnitudes have been determined using the photometric zero-point information in the header of the calibrated image files. These results are shown in Table \ref{tab:optical}, including the X-ray to optical flux ratios from the F814W and F702W broad-band filters, and the magnitude determinations from the same filters. In the cases where an optical counterpart has not been found (N1, N3, N4 and N5), the magnitudes and fluxes are just a lower limit. \section{Spectral analysis} We extracted spectra for all the detected sources close to the HVS, using extraction regions defined to include as many of the source photons as possible, but at the same time minimizing contamination from nearby sources and background. The background region was either a source-free circular annulus or several circles surrounding each source, in order to take into account the spatial variations of the diffuse emission and to minimize effects related to the spatial variation of the CCD response. For each source, we extracted spectra from each of the datasets. These spectra were summed to form a total spectrum for each source. Response and ancillary response files were created for each source in each of the observations using the CIAO \textsc{mkacisrmf} and \textsc{mkwarf} tools. The responses for a particular source were summed together, weighting according to the number of counts in each observation. The spectra were fitted using XSPEC v.11.3.2. In order to use the ${\rm \chi^{2}}$ statistic, we grouped the data to include at least 20 counts per spectral bin, before background subtraction. In spectral fitting we excluded any events with energies above 7.0 keV or below 0.5 keV. \begin{table} \begin{center} \begin{tabular}{lccr} \hline \hline N & N${_{\rm H}}$ & ${\rm \Gamma}$ & ${\rm \chi^{2}}$/d.o.f. \\ & (${\rm 10^{21}cm^{-2}}$) & & \\ \hline 1.... & 2.5$^{(a)}$ & 3.20$^{+0.23}_{-0.37}$ & 112.90/101 \\ 2.... & 2.72$^{+1.43}_{-0.87}$ & 1.78$^{+0.30}_{-0.24}$ & 101.50/109 \\ 3.... & 2.49$^{+0.40}_{-0.40}$ & 2.08$^{+0.09}_{-0.09}$ & 153.86/142 \\ 4.... & 2.05$^{+0.91}_{-0.96}$ & 2.29$^{+0.44}_{-0.28}$ & 156.24/152 \\ 5.... & 2.64$^{+1.23}_{-0.93}$ & 2.92$^{+1.44}_{-0.36}$ & 124.09/139 \\ 6.... & 3.74$^{+1.57}_{-1.39}$ & 3.51$^{+0.48}_{-0.66}$ & 102.58/92 \\ 7.... & 4.03$^{+1.78}_{-1.45}$ & 3.20$^{+1.39}_{-0.48}$ & 133.69/135 \\ 8.... & 2.66$^{+1.00}_{-0.91}$ & 2.13$^{+0.52}_{-0.25}$ & 150.81/138 \\ \hline \end{tabular} \caption{Spectral fits. (a) The column density of source N1 has been fixed due to the low count rate.} \label{tab:fittings} \end{center} \end{table} Table \ref{tab:fittings} summarizes our spectral results in terms of the absorbing column density and photon index. The sources have been modelled with an absorbed power law slope with photon index between ${\rm \Gamma=}$[1.78-5.56] and an equivalent column density of $\rm{N_H=[2.05-4.03]\times 10^{21} cm^{-2}}$. In all the cases the single component power law give satisfactory fits. The column density of source N1 has been fixed due to the low count rate. The fitted $\rm{N_H}$ values are consistent with the intrinsic absorption measured e.g. in the optical band; the value of A$\rm{_V}$=0.54 corresponds to $\rm{N_H\sim 1.1\times 10^{21}cm^{-2}}$, assuming $\rm{A_V=N_H \times 5.3 \times 10^{-22}}$ for $\rm{ R_V=3.1}$ (Bohlin et al. 1978). This value should be a lower limit to the fitted $\rm{N_H}$ value to be consistent, as is seen in Table \ref{tab:fittings}. As an example of our spectral fits, the brightest source, N3, has been fitted with a power-law with spectral index of $2.08\pm0.09$ and absorption of $\rm{N_H=2.5\pm0.4 \times 10^{21}~cm^{-2}}$ (see Fig. \ref{fig:source1_spec}). \begin{table} \begin{center} \begin{tabular}{llll} \hline \hline N & F$\rm{_{obs}}$(0.5--7.0 keV) & F$\rm{_{corr}}$(0.5--7.0 keV) & $\rm{log~L_{X}}$ \\ & erg $\rm{cm^{-2}~s^{-1}}$ & erg $\rm{cm^{-2}~s^{-1}}$ & 0.5--7.0 keV \\ \hline 1.... & 2.09 $\times 10^{-15}$ & 4.34 $\times 10^{-15}$ &39.51 \\ 2.... & 7.59 $\times 10^{-15}$ & 9.97 $\times 10^{-15}$ &39.86 \\ 3.... & 1.64 $\times 10^{-14}$ & 2.28 $\times 10^{-14}$ &40.22 \\ 4.... & 7.76 $\times 10^{-15}$ & 1.10 $\times 10^{-14}$ &39.91 \\ 5.... & 5.36 $\times 10^{-15}$ & 1.07 $\times 10^{-14}$ &39.90 \\ 6.... & 3.02 $\times 10^{-15}$ & 9.23 $\times 10^{-15}$ &39.84 \\ 7.... & 4.28 $\times 10^{-15}$ & 1.18 $\times 10^{-14}$ &39.95 \\ 8.... & 7.67 $\times 10^{-15}$ & 1.16 $\times 10^{-14}$ &39.93 \\ \hline \end{tabular} \caption{Fluxes (observed and k-corrected) and luminosities assuming a cosmological model with $\rm{H_{0}=70~km~s^{-1} Mpc^{-1}}$ and z=0.018.} \label{tab:luminosities} \end{center} \end{table} In Table \ref{tab:luminosities} we list the 0.5-7~keV flux and (absorption corrected) luminosities of the individual sources based on the best-fit power law model. The lower limit of the luminosity of point sources in the image, if at the distance of NGC\,1275, is $\rm{L_X(0.5-7.0~keV)=3.2\times 10^{39}erg~s^{-1}}$, which is already well above the Eddington limit for a neutron star binary ($\rm{L_X\sim 3 \times 10^{38}erg~ s^{-1}}$) and is also above the limit of canonical ULX, i.e. $\rm{\ge 10^{39}erg ~ s^{-1}}$. The brightest point source has a luminosity of $\rm{L_X(0.5-7.0~keV)=1.67 \times 10^{40} erg ~ s^{-1}}$, and is one of the brightest individual sources found in a galaxy. A ULX source more luminous than the entire X-ray luminosity of a normal galaxy has been found in the Cartwheel system with a luminosity of at least $\rm{L_X \sim 2-4 \times 10^{40} erg ~ s^{-1}}$ (Gao et al. 2003; Wolter \& Trinchieri 2004). They explain this luminosity with a high-mass X-ray binary source (HMXB). The high X-ray luminosity suggests either a single extremely bright source, or a very dense collection of several high $\rm{L_X}$ sources, which would be even more peculiar. Evidence of time variability might suggest that is a single high $\rm{L_X}$ source. Time variability analysis has been performed. The observations span about two years. Two data files were observed on 2002 August 8 and 10, and the other eleven data files were observed from 2004 October 4 to 2004 October 20, giving an almost daily coverage. The exposure times are between 22 and 143 ks. The data characteristics allows us determine short variation in 16 days (second period) and long-term variability of 2 years. Because of the low count rates of the sources in NGC\,1275 (see Table \ref{tab:positions}), it is very hard to search for short-term variability. We extracted light-curves, using {\sc dmextract} CIAO task for the two brightest sources (N3 and N4) (net count rate greater than 0.98 count $\rm{s^{-1}}$) binned with bin sizes of 500, 1000, 2500 and 5000 s. In both cases the points were consistent with the respective mean values and variability has not been found. Furthermore, the mean values between 2002 and 2004 are the same, including errors bars. Therefore, evidence of time variability has not been found during the whole set of observations. \section{Discussion} \emph{Chandra} has revealed significant populations of ULX in the interacting systems of the Antennae (NGC 4038/9; Zezas, Fabbiano \& Murray 2002) and the Cartwheel ring galaxy (Gao et al. 2003; Wolter \& Trinchieri 2004), where dramatic events have stimulated massive star formation. We have reported here on another example (Fig. \ref{fig:figure} left) in the HVS of NGC\,1275 which is interacting with the ICM of the Perseus cluster. The sources are spatially associated with the distribution of absorbing clouds seen in soft X-ray (Fig.~2) and optical (Fig.~3) images. Two sources (N7 and N8) are directly linked with dust knots and another two (N2 and N6) have an optical point-like counterpart (Fig. \ref{fig:figure} \emph{bottom}). Similar correspondence have been found in the Cartwheel galaxy with the outer ring (Wolter \& Trinchieri 2004) and in the Antennae galaxies with 39 X-ray sources within the WFPC2 field (Zezas et al. 2002). The optical brightness of the counterparts in the HVC are too high to be individual stars and so may be associated with young star clusters. Following the discussion of young star clusters in NCG\,1275 given by Richer et al (1993), an object of magnitude 22 corresponds to a cluster mass of about $10^6\Msun$ if its age is about $10^7\yr$. The HVC system travels at least 30~kpc in $10^7\yr$ so if a strong interaction with the core of the Perseus galaxy cluster has triggered star cluster formation in the HVC, then the clusters should have ages less than $\sim 10^8\yr$. Our interpretation of the spatial correspondence with star clusters is that the regions are especially active, indicating a real link between ULX and star-forming regions, and meaning they are young objects. However the optical limits on sources N3 and 4 rule out any association with massive clusters in those cases (the limit on the absolute magnitude is about $-8$). In M31 and the Milky Way (Grimm, Gilfanov \& Sunyaev 2003), XRB have luminosities consistent with the Eddington limit of a $\rm{\sim 2 M_\odot}$ accreting object. They produce luminosities $\rm{\sim 3 \times 10^{38}~ erg~s^{-1}}$, about one order of magnitude below the limiting luminosity in our sample ($\rm{3.2 \times 10^{39}~ erg~s^{-1}}$). It is possible that our ULX consist of at least 15 (or 130, in the case of the brightest source found) `normal' XRB clustered together, perhaps in a young star cluster. However in other objects we know that variability requires the presence of intrinsically luminous X-ray sources (e.g. M82; Griffiths et al. 2000, Kaaret et al. 2001). Alternative possibilities are that black hole sources, with masses in the range of galactic black hole binaries, are mildly beamed (Reynolds et al. 1999 and King et al. 2001). Spectral and timing features however rule out this possibility in some ULX (e.g. Strohmayer \& Mushotzky 2003). We note that compact supernova remnants sometimes have ULX luminosities (e.g. Fabian \& Terlevich 1998), but no recent supernovae have been reported for NGC\,1275 (SN1968A was to the S of the HVS; Capetti 2002). Finally, we recall the IMBH model which has spectral support from some sources (Miller et al 2004; the level of absorption in NGC\,1275 is too high for any soft excess to be observed). They may form in dense star clusters. Our optical studies have clearly shown that the ULX have very high X-Ray to optical flux ratios. X-ray selected AGN from the \emph{Rosat all sky survey} tend to have $\rm{log(F_X/F_{opt})\sim 1}$. Thus the ULX do not have the optical properties expected if their were simple extensions of AGN (IMBH, as low luminosity limit). However, low mass X-ray binaries in the Milky Way have $\rm{F_X/F_{opt}\sim 100-10000}$ (Mushotzky 2004). The results found in our system indicate that we have a mixed group of objects (see Table \ref{tab:optical}). At least 4 out of 8 sources (N3, N4, N5 and N8) have high X-ray to optical flux ratios. At least 3 out of 8 (N1, N2 and N6) have lower X-ray to optical ratios, possibly because they lie in star clusters. Our data are consistent with no significant variability, similar to the result obtained on NGC\,3256 by Lira et al. (2002). Time variability is frequently observed in ULX (e.g. IC\,342, Sugiho et al. 2001 or M51 X-1, Liu et al. 2002), arguing that most of them are single compact objects, rather than a sum of numerous lower luminosity objects in the same object. While most ULX vary, many show low amplitude variability on long time scales (e.g. the Antennae galaxies, Zezas et al. 2002), which is very different to galactic black holes. Portegies Zwart, Dewi \& Maccarone (2004) find that a persistent bright ULX requires a doner star exceeding $15\Msun$. The search for characteristic frequencies is one of the most productive way of determining the nature of the ULX. \section{Conclusions} We have described the detailed analysis of the spatial and spectral properties of the discrete X-ray sources detected with a deep \emph{Chandra} ACIS-S observation around NGC\,1275. Our results are summarized below: \begin{enumerate} \item We have detected a total of 8 sources to the north of NGC\,1275 nucleus. \item The sources are spatially coincident with the High Velocity System and thus probably associated with it. They are therefore ULX. \item Four of the sources have an optical counterpart in the I and R bands (from \emph{HST} images); two of which are point-like sources and the other two are associated with star-forming regions. \item In all the cases a single component power law gives satisfactory fits, with spectral index of $\rm{\Gamma=}$[1.78-3.51] and an equivalent column density of $\rm{N_H=[2.05-4.03]\times 10^{21}cm^{-2}}$. \item The minimum luminosity is $\rm{L_X(0.5-7.0 keV)=3.2\times 10^{39}erg~ s^{-1}}$ (source N1), which is already above the limit of canonical ULX. \item No variability was detected in the two brightest sources found. \end{enumerate} Our results add to the growing evidence that some episodes of rapid star formation lead to the production of ULX. Young, massive, star clusters may be involved in some, but not all of the sources. \section*{Acknowledgements} OGM acknowledges the financial support by the Ministerio de Educacion y Ciencia through the program AYA2003-00128 and grant FPI BES-2004-5044. ACF thanks the Royal Society for support.
Title: The Most Metal-Rich Intervening Quasar Absorber Known
Abstract: The metallicity in portions of high-redshift galaxies has been successfully measured thanks to the gas observed in absorption in the spectra of quasars, in the Damped Lyman-alpha systems (DLAs). Surprisingly, the global mean metallicity derived from DLAs is about 1/10th solar at 0<z<4 leading to the so-called ``missing-metals problem''. In this paper, we present high-resolution observations of a sub-DLA system at z_abs=0.716 with super-solar metallicity toward SDSS J1323-0021. This is the highest metallicity intervening quasar absorber currently known, and is only the second super-solar absorber known to date. We provide a detailed study of this unique object from VLT/UVES spectroscopy. We derive [Zn/H]=+0.61, [Fe/H]=-0.51, [Cr/H]=<-0.53, [Mn/H] = -0.37, and [Ti/H] = -0.61. Observations and photoionisation models using the CLOUDY software confirm that the gas in this sub-DLA is predominantly neutral and that the abundance pattern is probably significantly different from a Solar pattern. Fe/Zn and Ti/Zn vary among the main velocity components by factors of \~ 3 and ~ 35, respectively, indicating non-uniform dust depletion. Mn/Fe is super-solar in almost all components, and varies by a factor of ~ 3 among the dominant components. It would be interesting to observe more sub-DLA systems and determine whether they might contribute significantly toward the cosmic budget of metals.
https://export.arxiv.org/pdf/astro-ph/0601079
\title{The Most Metal-Rich Intervening Quasar Absorber Known\thanks{Based on the UVES observations collected during the DDT ESO prog. ID No. 274.A-5030 at the VLT/Kueyen telescope, Paranal, Chile } } \author{C. P\'eroux$^1$, V. P. Kulkarni$^2$, J. Meiring$^2$, R. Ferlet$^3$, P. Khare$^4$, J. T. Lauroesch$^5$, G. Vladilo$^6$, \& D. G. York$^7$. } \offprints{C. P\'eroux.} \institute{$^1$ European Southern Observatory, Garching-bei-M\"unchen, Germany. \email{cperoux@eso.org}\\ $^2$ Dept. of Physics and Astronomy, Univ. of South Carolina, Columbia, USA.\\ $^3$ Institut d'Astrophysique de Paris, UMR7095 CNRS, Universite Pierre \& Marie Curie, France.\\ $^4$ Dept. of Physics, Utkal University, Bhubaneswar, India.\\ $^5$ Dept. of Physics and Astronomy, Northwerstern University, Evanston, USA.\\ $^6$ Osservatorio di Trieste, Trieste, Italy.\\ $^7$ Dept. of Astronomy and Astrophysics, Univ. of Chicago, Chicago, USA. } \authorrunning{C. P\'eroux et al.} \titlerunning{The Most Metal-Rich Intervening Quasar Absorber Known} \date{Received August 17, 2005; accepted January 3, 2006} \abstract{The metallicity in portions of high-redshift galaxies has been successfully measured thanks to the gas observed in absorption in the spectra of quasars, in the Damped Lyman-$\alpha$ systems (DLAs). Surprisingly, the global mean metallicity derived from DLAs is about 1/10$^{\rm th}$ solar at 0$\la$z$\la$4 leading to the so-called ``missing-metals problem''. In this paper, we present high-resolution observations of a sub-DLA system at \zabs=$0.716$ with super-solar metallicity toward SDSS J1323$-$0021. This is the highest metallicity intervening quasar absorber currently known, and is only the second super-solar absorber known to date. We provide a detailed study of this unique object from VLT/UVES spectroscopy. We derive [Zn/H]=$+$0.61, [Fe/H]=$-$0.51, [Cr/H]=$<-$0.53, [Mn/H] = $-$0.37, and [Ti/H] = $-$0.61. Observations and photoionisation models using the CLOUDY software confirm that the gas in this sub-DLA is predominantly neutral and that the abundance pattern is probably significantly different from a Solar pattern. Fe/Zn and Ti/Zn vary among the main velocity components by factors of $\sim 3$ and $\sim 35$, respectively, indicating non-uniform dust depletion. Mn/Fe is super-solar in almost all components, and varies by a factor of $\sim 3$ among the dominant components. It would be interesting to observe more sub-DLA systems and determine whether they might contribute significantly toward the cosmic budget of metals. \keywords{Galaxies: abundances -- intergalactic medium -- quasars: absorption lines -- quasars: individual: SDSS J1323$-$0021} } \section{Introduction} Damped Lyman-$\alpha$ systems (DLAs) seen in absorption in the spectra of background quasars are selected over all redshifts independent of the intrinsic luminosities of the underlying galaxies. They have hydrogen column densities, \loghi\ $\ga$ 20.3 and are the major contributors to the neutral gas in the Universe at high redshifts (Storrie-Lombardi \& Wolfe 2000; P\'eroux \e\ 2003b). But it has been suggested that at least some of the \hi\ lies in systems with \hi\ column density below that required by the traditional DLA definition, in the ``sub-Damped Lyman-$\alpha$ Systems (sub-DLAs)'' with $19.0$ $<$ \loghi\ $<$ 20.3. The DLAs and sub-DLAs offer direct probes of element abundances over $ > 90 \%$ of the age of the Universe. Zn is a good probe of the total (gas and solid phase) metallicity, in DLAs because Zn tracks Fe in most Galactic stars with [Fe/H]$> -$3, it is undepleted on interstellar dust grains, and the lines of the dominant ionisation species Zn II are often unsaturated (e.g., Pettini et al. 1999). Abundances of depleted elements such as Cr or Fe relative to Zn probe the dust content and the relative abundances can also yield information about the nucleosynthetic processes (e.g., Pettini et al. 1997; Kulkarni, Fall, \& Truran, 1997; P\'eroux et al., 2002; Khare et al. 2004). A study of the cosmological evolution of the \hi\ column density-weighted mean metallicity in DLAs (e.g., Kulkarni \& Fall, 2002) shows surprising results. Contrary to most models of cosmic chemical evolution (e.g., Malaney \& Chaboyer, 1996; Pei, Fall \& Hauser, 1999), recent observations indicate at most a mild evolution in DLA global metallicity with redshift for 0$\la$z$\la$ 4 (Prochaska \e\ 2003; Khare et~al. 2004; Kulkarni, et~al., 2005; and references therein). Even theoretical models such as Smoothed-particle-hydrodynamics simulations (Nagamine, Springel, \& Hernquist, 2004) predict that the true DLA metallicities could be 1/3 solar at z=2.5 and higher at lower redshifts. Even at z=2.5, making a census of the predicted and observed neutral comoving densities of gas, $\Omega$, and metals, $\Omega_{\rm Z}$, one finds that most of the baryons are in the Lyman-$\alpha$ forest but its metal content is extremely low. The measured value of $\Omega_{\rm HI}$(DLA) is only a small fraction of $\Omega_{\rm baryons}$ and the DLA global mean metallicity is about 1/10$^{\rm th}$ solar. The metallicity of Lyman break galaxies is still poorly constrained; but, in any case, these objects are known to be star-forming galaxies and may not be representative of the normal galaxy population. In total, these three components account for no more than $\approx 10-15$\% of what we expect to have been produced by z=2.5 (Pettini \e\ 2003; Bouch\'e \e\ 2005). The missing metals problem in low-redshift DLAs is even more surprising since the high global star formation rate estimates at z$>$1.5 (e.g. Madau \e\ 1998) imply that higher metallicities should be expected at low redshift. It is possible that $\Omega_Z$(DLA) has been underestimated. Metal-rich DLAs could obscure quasars due to their possible high dust content (Fall \& Pei 1993). This may be the reason for the apparent low metallicity in the DLAs observed in optically selected quasars (e.g., Fall \& Pei 1993; Boiss\'e et al. 1998). Recently Vladilo and P\'eroux (2005) have shown that the fraction of high-redshift DLAs missed due to dust obscuration, could be up to 50\%, which is consistent with the results of surveys of radio selected quasars (Ellison \e\ 2001). They have estimated that at z$\sim$2.3, the real mean metallicity of DLAs could be 5 to 6 times higher than what is observed, which may help alleviate the missing metals problems. Indeed, systems at lower redshift may have significantly more dust at any given metallicity simply because the dust in these objects has had more time to process the elements. On the other hand, new lines of evidence are pointing toward lower \nhi\ quasar absorbers like Lyman Limit Systems (LLS) and sub-DLAs being more metal-rich (P\'eroux \e\ 2003a; Jenkins \e\ 2005). The dust bias, if real, is also likely to be less severe for metal-rich sub-DLAs as compared to the metal-rich DLAs due to the lower gas and therefore dust content in the former, for a constant dust to gas ratio. Thus, the obscuration bias will affect the DLAs at a lower dust-to-gas ratio as compared to the sub-DLAs. This scenario is consistent with the recent radio surveys (e.g., Vladilo \& P\'eroux 2005), but still needs to be further quantified observationally. Indeed, Zn measurements exist for only two sub-DLAs at low $z$: the marginally super-solar sub-DLA toward Q0058$+$019 with $z_{\rm abs}$=0.61, \loghi=20.08, and [Zn/H]=$+$0.08 (Pettini \e\ 2000), and the supersolar sub-DLA toward SDSS J1323$-$0021 with $z_{\rm abs}$=0.72, \loghi=20.21, and [Zn/H]=$+$0.40 (Khare et al. 2004). Although Khare et al. reported [Zn/H]=$+$0.40 for this latter absorber, the modest resolution of their MMT data could not resolve the Mg I+Zn II $\lambda$ 2026 and Cr II+Zn II $\lambda$ 2062 blends.The extent of line saturation on the derived column densities was also unclear. With the goal of addressing these issues with high-resolution data, we obtained VLT/UVES spectra of this quasar, which are presented here. These new data are essential to confidently determine the metallicity by minimising the problem of line saturation. Section 2 presents the observational set-up and data reduction process, while section 3 presents the analysis. Section 4 provides a discussion of the results. \section{Observations and Data Reduction} Spectra of SDSS J1323$-$0021 (z$_{\rm em}$=1.390; SDSS mag $g=18.49$) were acquired in service mode as Director's Discretionary Time (DDT) on 3$^{\rm rd}$ of March and 13$^{\rm th}$ of March 2005 with the high-resolution UVES spectrograph mounted on Kueyen Unit 2 VLT (D'Odorico \e\ 2000). Three exposures of length 4100 sec, 3500 sec and 4100 sec were obtained with standard 390$+$562 settings thus providing a wavelength coverage of $\sim$3300\AA-4400\AA, 4700\AA-5600\AA\ and 5800\AA-6600\AA. The data were reduced using the most recent version of the UVES pipeline to accommodate for the new format of the raw fits file (version: uves/2.1.0 flmidas/1.1.0). Master bias and flat images were constructed using calibration frames taken the closest in time to the science frames. The science frames were extracted with the ``optimal'' option. The spectrum was then corrected to vacuum heliocentric reference. The resulting spectra were combined weighting each spectrum with its signal-to-noise. The final spectra have resolution of 4.7 km s$^{-1}$ at Zn II $\lambda 2026$. The spectra were divided into 100 {\AA} regions, and each region normalised using cubic spline functions of orders 1 to 5. \section{Analysis} \begin{table*} \begin{center} \caption{Parameter fit to the $z_{\rm abs} = 0.716$ sub-DLA model. Velocities and b are in km s$^{-1}$ and N's are in cm$^{-2}$.} \label{t:fit} \begin{tabular}{l r r r r r r r r r c c} \hline\hline &Vel & b &\ion{Mg}{i} &\ion{Mg}{ii} &\ion{Fe}{ii} &\ion{Zn}{ii} &\ion{Cr}{ii} &\ion{Mn}{ii} &\ion{Ti}{ii} &[Fe/Zn] &[Mn/Fe]\\ \hline N(X) &$-$120.8 & 7.1 &... &5.63e12 &5.82e12 &... &... &3.75e11 &... &... & 0.79 \\ $\sigma$ &... &... &... &1.66e11 &4.08e11 &... &... &1.26e11 &... &... &... \\ N(X) &$-$96.9 &16.3 &... &3.65e12 &3.64e12 &2.32e12 &$<$2.81e12 &2.97e11 &2.09e11 &$-$2.65 &0.89 \\ $\sigma$ &... &... &... &1.21e11 &4.31e11 &6.53e11 &... &1.93e11 &1.27e11 &... &... \\ N(X) &$-$80.1 & 6.2 &3.16e10 &3.55e12 &3.92e12 &... &$<$1.69e12 &2.09e11 &1.86e11 &... &0.71 \\ $\sigma$ &... &... &1.15e10 &1.36e11 &3.69e11 &... &... &1.23e11 &8.41e10 &... &... \\ N(X) &$-$62.9 & 5.5 &1.20e11 &8.13e12 &4.25e12 &$<$3.10e11 &$<$4.77e11 &2.01e11 &1.07e11 &$-$1.71 &0.65 \\ $\sigma$ &... &... &1.23e10 &3.68e11 &4.22e11 &... &... &1.01e11 &8.00e10 &... &... \\ N(X) &$-$51.4 & 5.0 &$<$1.33e10 &$>$2.86e12 &5.74e12 &$<$2.56e11 &... &1.44e11 &$<$6.92e10 &$-$1.49 &0.38 \\ $\sigma$ &... &... &... &... &1.30e12 &... &... &1.19e11 &... &... &... \\ N(X) &$-$43.0 & 7.4 &4.80e11 &$>$6.01e13 &2.41e13 &1.07e12 &... &1.98e11 &1.32e11 &$-$1.49 &$-$0.11\\ $\sigma$ &... &... &2.29e10 &... &3.54e12 &5.18e11 &... &1.43e11 &9.85e10 &... &... \\ N(X) &$-$13.0 &12.5 &3.04e12 &$>$2.38e14 &1.97e14 &2.57e12 &... &2.40e12 &4.79e11 & $-$0.96 &0.07 \\ $\sigma$ &... &... &1.48e11 &... &2.54e13 &6.22e11 &... &3.14e11 &2.16e11 &... &... \\ N(X) &3.9 & 30.5 &2.36e12 &$>$1.48e14 &4.00e14 &... &... &5.33e12 &$<$4.17e11 &... &0.10 \\ $\sigma$ &... &... &2.03e11 &... &2.63e13 &... &... &5.28e11 &... &... &... \\ N(X) &12.0 & 5.0 &1.44e11 &... &... &$<$3.66e11 &$<$3.20e11 &... &$<$3.89e10 &... &0.01 \\ $\sigma$ &... &... &4.81e10 &... &... &... &... &... &... &... &... \\ N(X) & 28.3 &12.0 &2.06e12 &$>$7.14e13 &1.15e14 &1.53e12 &$<$2.37e12 &1.28e12 &3.80e11 &$-$0.97 &0.03 \\ $\sigma$ &... &... &1.25e11 &... &2.94e13 &7.60e11 &... &2.86e11 &1.53e11 &... &... \\ N(X) &44.5 &11.0 &4.76e12 &$>$2.30e13 &2.89e14 &7.76e12 &... &4.34e12 &4.37e11 &$-$1.27 &0.16 \\ $\sigma$ &... &... &4.37e11 &... &5.42e13 &1.48e12 &... &2.90e11 &1.22e11 &... &... \\ N(X) &56.6 &6.9 &8.70e11 &$>$4.62e13 &2.83e13 &$<$2.73e11 &... &6.30e11 &2.69e11 &$-$0.83 &0.33 \\ $\sigma$ &... &... &1.15e11 &... &1.82e13 &... &... &1.97e11 &9.48e10 &... &... \\ N(X) &71.8 &11.9 &1.93e12 &$>$5.09e14 &3.12e14 &2.49e12 &6.07e12 &3.62e12 &1.58e11 &$-$0.75 &0.04 \\ $\sigma$ &... &... &5.19e10 &... &4.29e13 &7.30e11 &3.24e12 &2.22e11 &1.05e11 &... &... \\ N(X) &92.7 &6.4 &2.42e12 &$>$2.99e14 &3.76e13 &7.35e12 &$<$1.54e12 &1.20e12 &2.09e11 &$-$2.14 &0.48 \\ $\sigma$ &... &... &8.51e10 &... &4.70e12 &1.49e12 &... &1.40e11 &7.87e10 &... &... \\ N(X) &119.3 &8.0 &7.45e10 &4.49e12 &2.33e12 &5.13e11 &6.01e12 &1.98e11 &1.02e08 &$-$2.19 &0.91 \\ $\sigma$ &... &... &1.32e10 &4.90e07 &2.85e11 &4.48e11 &3.11e12 &1.27e11 &1.42e07 &... &... \\ N(X) &194.1 &4.6 &2.44e10 &1.91e12 &... &... &... &... &... &... &... \\ $\sigma$ &... &... &1.00e10 &7.56e10 &... &... &... &... &... &... &... \\ \hline \end{tabular} \end{center} \end{table*} Several lines of Zn II, Cr II, Fe II, Mn II, and Ti II were detected at z$_{\rm abs}$=0.716. Fe I, Zn I, and Co II were not detected. The column densities were estimated by fitting multi-component Voigt profiles to the observed absorption lines using the program FITS6P (Welty, Hobbs, \& York 1991) that evolved from the code used by Vidal-Madjar et al. (1977). FITS6P minimizes the $\chi^{2}$ between the data and the theoretical Voigt profiles convolved with the instrumental profile. The atomic data were adopted from Morton (2003). The absorption profiles show a complex velocity structure with a total of 16 components needed. The velocity and Doppler b parameters of the various components were estimated from the Mg I, Mg II, and Fe II lines. The component at 12 km s$^{-1}$ is negligible in most species except Mg I, Zn II and Cr II. The component at 194 km s$^{-1}$ is negligible in all species but Mg II. The component at $-$121 km s$^{-1}$ is detected in Mg II, Fe II, and Mn II, but not in Mg I, Zn II, Cr II, or Ti II. The best-fit column densities in the individual components and their uncertainties were estimated assuming the same fixed b and v values for all species in a given component (Figure~\ref{f:fit}). The results of the profile fitting analysis are summarised in Table~\ref{t:fit}. Column densities in the few weak components that could not be well-constrained due to noise are marked with ``...''; their contributions to the total column densities (listed in Table~\ref{t:ab}) are negligible. As an additional check, we also estimated the total column densities using the apparent optical depth (AOD; Savage \& Sembach 1991) method for the various detected lines and obtained results consistent with those from the profile-fitting method. The agreement was within 0.01 dex for Mg I and Mn II, within 0.1 dex for Fe II and Zn II, and within 0.15 dex for Ti II. The Mg I $\lambda 2026.5$ contribution to the Zn II $\lambda 2026.1$ line was estimated using the component parameters for Mg I derived from the Mg I $\lambda 2852$ profile. This contribution, indicated by a dashed blue curve in the top left panel of Fig~\ref{f:fit}, was a small fraction of the observed strength of the $\lambda 2026$ line. The remaining part of the $\lambda 2026$ line was fitted with a 15-component model for Zn II, using the same $b$ values and velocities, but varying the column densities in the individual components. The Zn II fit thus obtained was used to estimate the Zn II contribution to the $\lambda 2062$ line. The remaining part of the $\lambda 2062$ line was fitted with a 15-component model for Cr II, using the same set of $b$ values and velocities. This was the only available estimate for Cr II, since the Cr II $\lambda \lambda$ 2056, 2066 lines lie in noisy regions and are undetected. The relatively large errors in the Cr II column densities arise from the noisy nature of the $\lambda 2062$ line. The Fe II column densities in the components at low velocities were constrained by using the Fe II $\lambda \lambda 2260, 2374$ lines, since the stronger Fe II lines are saturated. The weaker Fe II components at high positive and negative velocities were constrained in column density using the $\lambda \lambda 2374, 2382, 2600$ lines, since these components are poorly constrained by the weaker $\lambda \lambda 2260,2374$ lines. The Mg II $\lambda 2796, 2803$ profiles were fitted together, but provide only a lower limit to \mg2\ owing to saturation in the central components. The relative abundances were calculated using solar abundances from Asplund et al. (2005), adopting the mean of photospheric and meteoritic values for Mg, Ti, Cr, Fe, Zn, and the meteoritic value for Mn. One concern is that sub-DLAs may be partially ionised in H, artificially enhancing the ratio of Zn II to H I, for instance. To investigate the ionisation corrections, we used the CLOUDY software package (version 94.00, Ferland 1997) and computed photoionisation models assuming ionisation equilibrium and a solar abundance pattern. We thus obtained the theoretical column density predictions for any ionisation state of all observed ions as a function of the ionisation parameter U. Our findings confirm the observations: from a comparison of the observed and theoretical Mg II/Mg I ratios, we deduced that the gas in this sub-DLA is predominantly neutral ($\log U <- 5$) and the overall abundance pattern is probably not solar. This latter point is also clear from the relative abundances listed for each component in Table~\ref{t:fit}. It should be pointed out that there are no third ionisation stage detected in the system under study. Nevertheless, we do have Ti II, which has the same ionisation potential as H I. The fact that Ti is not suppressed compared to Fe or Mn also implies that the gas is neutral with low ionisation parameter. \begin{table} \begin{center} \caption{Summary of total abundances.} \label{t:ab} \begin{tabular}{l c l c } \hline\hline Id&$\log N_{\rm total}$&A(X/N)$_{\sun}$&[X/H]$^*$\\ \hline \ion{Mg}{i} &$13.26\pm0.01$ &... &... \\ \ion{Mg}{ii} &$>15.15$ &$-$4.47 &$>-$0.58\\ \ion{Fe}{ii} &$15.15 \pm 0.03$ &$-$4.55 &$-$0.51 $\pm 0.20$\\ \ion{Zn}{ii} &$13.43 \pm 0.05$ &$-$7.40 &$+$0.61$\pm 0.20$\\ \ion{Cr}{ii} &$<13.33$ &$-$6.37 &$<-$0.52\\ \ion{Mn}{ii} &$13.31 \pm 0.02$&$-$6.53 &$-$0.37$\pm 0.20$\\ \ion{Ti}{ii} &$12.49 \pm 0.11$&$-$7.11 &$-$0.61$\pm 0.22$\\ \hline \end{tabular} \end{center} $^*$ The error bars on [X/H] include the errors in log $N(X)$ and \loghi. \end{table} \section{Discussion and Conclusions} Table~\ref{t:ab} lists the abundances, using $\loghi=20.21^{+0.21}_{-0.18}$ (Khare \e\ 2004) that we derived from Voigt profile fitting of the damped Ly-$\alpha$ line in the publicly available HST/STIS spectrum of SDSS J1323$-$0021 (program GO 9382; PI: Rao). Rao et al. (2005) obtained $\loghi=20.54^{+0.16}_{-0.15}$ from the same data set. We use the former value since it gives a smaller residual with respect to the data and therefore regard this absorber as a sub-DLA. For either \nhi, the strength of the Zn II lines detected in our UVES spectrum implies a super-solar metallicity. Using the standard definition: $[X/H] = \log [N(X)/N(H)]_{DLA}- \log [N(X)/N(H)]_{\odot}$, we find [Zn/H]=$+$0.61. In principle, if Mg I $\lambda$ 2852 were substantially saturated, the contribution of Mg I $\lambda$ 2026 could be higher than our best-fit estimate. However, based on our apparent optical depth measurements and profile-fitting results, we estimate that it would take $\sim$8 times more total Mg~{\sc i} than the value derived from the $\lambda 2852$ line to contribute the entirely of the $\lambda 2026$ line. Such a high value of Mg~{\sc i} can be ruled out by the observed profile of the $\lambda 2852$ line. (Of course, such a scenario would also be inconsistent with the observed strength and profile of the $\lambda 2062$ line.) To understand this issue in more detail, we estimated the maximum Mg~{\sc i} in the dominant components that would still give the shape of the Mg I $\lambda$2852 profile consistent with the observed profile within the noise level in the continuum. This maximum total Mg~{\sc i} $= 2.7e13$ is about $50 \%$ larger than the best-fit value listed in Table 2. Putting this maximum Mg I model in the $\lambda 2026$ line, the corresponding total Zn~{\sc ii} needed to fit the remaining part of the $\lambda$ 2026 line would be 2.5e13, lower by $< 10 \%$ from our best-fit value. Thus [Zn/H] is at least $>+$0.59, indicating that our result would not be affected much by saturation of Mg I $\lambda$ 2852. Finally, as an additional check, we also estimated the maximum contribution of Cr II 2062 by rebinning our spectrum by factors of 10 or 20, measuring the upper limit for Cr II$\lambda$ 2056 in the rebinned spectrum. We then spread this upper limit for N(Cr II) over the 2062 profile, using the velocity model derived from the combination of the lines. We assume the same Fe/Cr ratio in all components, taking the percentage of Cr II in each components relative to total N(CrII) summed over all components to be the same as the corresponding percentage of Fe II in that component. Fitting the remaining part of the $\lambda 2062$ line with a 15-component model of Zn II, we estimated $N_{\rm Zn II} > 2.05e13$, i.e. $[Zn/H] > 0.50$. The abundances of Fe, Cr, Mn, and Ti lie in the range of $-$0.4 to $-$0.6 dex, and indicate that this absorber is not only metal-rich, but also very dusty. Using the model from Vladilo (2004), we find that 95\% of the Fe is in dust phase and the total metallicity is even slightly higher than 0.6 dex. This sightline also shows substantial reddening compared to the SDSS quasar composite ($\Delta (g-i) = 0.47$; Khare et al. 2004). Considering only the better-determined components between $-$13 and 100 km s$^{-1}$, Fe/Zn varies by a factor of $\sim3$ and Ti/Zn varies by $\sim$35. [Mn/Fe] varies by $\sim 3$, but indicates a super-solar Mn abundance with respect to Fe in all components. The relative abundance of Mn with respect to Fe is not expected to exceed the solar value as Mn, unlike Fe, has an odd atomic number. However, [Mn/Fe]$>$0 is often seen in the Galactic interstellar gas due to the stronger depletion of Fe on to dust grains. To summarise, our high resolution VLT/UVES data have allowed us to alleviate the saturation issue in Zn II and Cr II lines and therefore unambiguously prove the super-solar metallicity of the sub-DLA at $z_{\rm abs}$=0.716 toward SDSS J1323$-$0021. If the dust obscuration bias for DLAs is indeed significant, as proposed by Vladilo and P\'eroux (2005), sub-DLA systems such as the one reported here could be better probes of dusty regions with significant past star formation (e.g. Lauroesch \e\ 1996; York \e\ 2006), as similar DLA systems will be missed due to dust obscuration. On the other hand, it is also, possible to envisage a scenario where the dust obscuration is not very significant in quasar absorbers (as is indicated by the rising spectrum of gamma-ray burst afterglows). In this scenario it is possible that the sub-DLA systems may indeed be more metal-rich as compared to DLAs as indicated by observations of P\'eroux \e\ (2003a) and by the observations of super-solar metallicity in such systems presented here and in Pettini et al (2003). With the large-scale spectroscopic surveys of quasars currently underway (e.g. the SDSS, York \e\ 2000, York \e\ 2001), such metal-rich sub-DLA systems may be found in large numbers. If future observations indeed find such systems, sub-DLAs may contribute significantly to the overall global metallicity. \begin{acknowledgements} We are grateful to ESO director, Catherine Cesarsky, for time allocation to this DDT program, and to the VLT staff for carrying out our observations in service mode. VPK and JM acknowledge support from the U. S. National Science Foundation grant AST-0206197. \end{acknowledgements}
Title: The complex X-ray morphology of NGC 7618: A major group-group merger in the local Universe?
Abstract: We present results from a short {\em Chandra}/ACIS-S observation of NGC 7618, the dominant central galaxy of a nearby ($z$=0.017309, d=74.1 Mpc) group. We detect a sharp surface brightness discontinuity 14.4 kpc N of the nucleus subtending an angle of 130$^\circ$ with an X-ray tail extending $\sim$ 70 kpc in the opposite direction. The temperature of the gas inside and outside the discontinuity is 0.79$\pm$0.03 and 0.81$\pm$0.07 keV, respectively. There is marginal evidence for a discontinuous change in the elemental abundance ($Z_{inner}$=0.65$\pm$0.25,$Z_{outer}$=0.17$\pm$0.21 at 90% confidence), suggesting that this may be an `abundance' front. Fitting a two-temperature model to the ASCA/GIS spectrum of the NGC 7618/UGC 12491 pair shows the presence of a second, much hotter ($T$=$\sim$2.3 keV) component. We consider several scenarios for the origin of the edge and the tail including a radio lobe/IGM interaction, non-hydrostatic `sloshing', equal-mass merger and collision, and ram-pressure stripping. In the last case, we consider the possibility that NGC 7618 is falling into UGC 12491, or that both groups are falling into a gas poor cluster potential. There are significant problems with the first two models, however, and we conclude that the discontinuity and tail are most likely the result of ram pressure stripping of the NGC 7618 group as it falls into a larger dark matter potential.
https://export.arxiv.org/pdf/astro-ph/0601378
\title{The complex X-ray morphology of NGC 7618: A major group-group merger in the local Universe?} \author{R. P. Kraft} \affil{Harvard/Smithsonian Center for Astrophysics, 60 Garden St., MS-67, Cambridge, MA 02138} \author{C. Jones} \affil{Harvard/Smithsonian Center for Astrophysics, 60 Garden St., MS-2, Cambridge, MA 02138} \author{P. E. J. Nulsen} \affil{Harvard/Smithsonian Center for Astrophysics, 60 Garden St., MS-6, Cambridge, MA 02138} \author{M. J. Hardcastle} \affil{University of Hertfordshire, School of Physics, Astronomy, and Mathematics, Hatfield AL10 9AB, UK} \keywords{galaxies: individual (NGC 7618) - X-rays: galaxies - galaxies: ISM - groups: mergers} \section{Introduction} The complex cluster morphology seen in X-ray images and in galaxy distributions gave support to the hypothesis that large structures form hierarchically; that is, that small groups of galaxies merge to form low-mass subclusters, which then merge to form a massive rich cluster with the infalling groups aligned along large filaments. Galaxies and groups are the building blocks of the observable Universe and contain the bulk of the observable baryons. Groups are estimated to contain a significant fraction, 20-30\%, of the total matter in the Universe. Thus groups are important cosmological indicators of the distribution and properties of the dark matter. However, because they are not as luminous as clusters, they have received less study than their more massive cousins. Observations of rich clusters show many examples of pending or ongoing mergers of subclusters. In particular in X-ray cluster catalogs, about 40\% of rich clusters show substructure \citep{jon84, jon99, moh95}. Virtually all stages of cluster mergers have been thoroughly investigated with both the {\em Chandra} and XMM-Newton observatories \citep{mar00,vik01,bri04,hen04}. However, less attention has been paid to the merging of groups and the formation of low-mass clusters due to their lower X-ray luminosity and the paucity of examples in the local Universe. To our knowledge, the only nearby example of the early stages of the merger of two roughly equal mass groups is the NGC 499/NGC 507 pair \citep{kim95,kra03}. In the hierarchical scenario, group mergers represent a critical transitional phase in the formation of larger scale structure. An understanding of the group merger process is thus fundamental to understanding the growth of structure. In this paper, we report results from analysis of a short {\em Chandra}/ACIS-S observation of the nearby elliptical galaxy NGC 7618 ($z$=0.017309 or d$_L$=74.1 Mpc for WMAP cosmology \citep{spe03} - 1$''$=350 pc). The X-ray luminosity of NGC 7618 is $\sim$7$\times$10$^{42}$ ergs s$^{-1}$ in the 0.1-10 keV band, typical of groups, not isolated elliptical galaxies, although it appears to be optically isolated \citep{col01}. We find a sharp surface brightness discontinuity in the X-ray emission north of of the nucleus of NGC 7618, and an extended tail to the south. We conclude that these features are either the result of a major group-group merger with UGC 12491, a group that lies 14.1$'$ on the sky from NGC 7618 at virtually identical redshift ($z$=0.017365) \citep{ebl02}, or ram-pressure stripping due to infall of NGC 7618 into a larger gravitational potential (which may include UGC 12491). This pair has been poorly studied in large part because of its relatively low galactic latitude ($\ell=105.575$, $b=-16.909$, $A_V$=0.97, $N_H$=1.19$\times$10$^{21}$ cm$^{-2}$)). \section{{\em Chandra} and ASCA Observations} NGC 7618 was observed for 18.4 ks with Chandra/ACIS-S on December 10, 1999 (OBSID 802). The lightcurve of events in the 5.0-10.0 keV bandpass on the entire S3 chip, excluding the NGC 7618 nucleus and any point sources visible by eye, was created using 259 s bins and examined for periods of flaring background. Intervals where the rate was more than 3$\sigma$ above the mean rate were removed. There was considerable background flaring during this observations, and almost 10 ks of data were excluded. Only 8438 s of good time remained. Bad pixels, hot columns, and columns along node boundaries were also removed. We present data from the S2 and S3 chips in this paper. Absorption by foreground gas in our galaxy ($N_H$=1.19$\times$10$^{21}$ cm$^{-2}$) was included in all spectral fits. NGC 7618 was also observed by ASCA for $\sim$54 ks on July 7, 1998. UGC 12491 is contained within the FOV of the ASCA/GIS, and we use these data to measure the temperature of the gas around this galaxy and the diffuse emission between NGC 7618 and UGC 12941. Data from the SIS was not used because of its smaller field of view. \section{Analysis} An adaptively smoothed ASCA/GIS image of the NGC 7618/UGC 12491 pair (usin the CIAO program `csmooth') is shown in Figure~\ref{gisimg}. The optical and X-ray properties of both groups are summarized in Table~\ref{galtab}. Their X-ray luminosities are each $\sim$6-7$\times$10$^{42}$ ergs s$^{-1}$, and their X-ray emission extends to radii of 150-200 kpc. The luminosities and spatial extents far exceed those of typical isolated elliptical galaxies and are more representative of poor clusters or fossil groups \citep{vik99}. The central elliptical galaxies represent only a small fraction of the gravitating mass that resides in a much larger dark matter halo. No optical census of the galaxy populations around either NGC 7618 or UGC 12491 has been undertaken, but their recessional velocities differ by only 17 km s$^{-1}$. The temperature of the gas around these galaxies is $\sim$0.8 keV, typical of groups. Based on their X-ray luminosities, X-ray extents, and spatial proximity we conclude that these two objects are groups likely within $\sim$300 kpc of each other and gravitationally interacting. An adaptively smoothed, exposure corrected, background subtracted {\em Chandra}/ACIS-S image in the 0.5-2.0 keV band of NGC 7618 is shown in Figure~\ref{acisimgb}. The X-ray bright active nucleus of the host galaxy is labeled at the center. There are three unusual features to note in this image. First, the peak of the diffuse X-ray emission (shown orange in Figure~\ref{acisimgb}), and therefore presumably the peak in gas density, as well as the center of the host galaxy lie $\sim$1$'$ North of the center of the larger scale diffuse X-ray emission (shown green). An X-ray `tail' extends $\sim$3.3$'$ (69.3 kpc) South of the nucleus. The statistical significance of this `tail' can be seen in the X-ray surface brightness profiles shown in Figure~\ref{tailwedge}. These profiles were made in two 90$^\circ$ sectors to the North (green) and South (red) with the nucleus of NGC 7618 at the vertex. The surface brightness in the S sector at distances larger than 70$''$ from the nucleus is several times higher than in the North sector. Either the dark matter has a very unusual distribution, or the gas is not in hydrostatic equilibrium in the gravitational potential. Second, there is a sharp surface brightness discontinuity $\sim$41$''$ (14.4 kpc) North of the nucleus that spans $\sim$130$^\circ$. This discontinuity is delineated by the white arrows in Figure~\ref{acisimgb}. and is probably a contact discontinuity between two moving fluids. Third, on smaller scales, the X-ray emission from the gas peaks $\sim$9$''$ (3.2 kpc) to the E of the nucleus of the host galaxy as seen in Figure~\ref{xoptovl}. In the central 10 kpc of the host galaxy, the stars will dominate the gravitating mass, so there clearly is an offset between the hot gas and the gravitating mass. All of these strongly suggest that the gas has partially separated from the gravitating matter and indicate non-hydrostatic motions. There is an X-ray point source coincident with the optical nucleus, presumably a low-luminosity AGN. The point source contains 48 counts in the 0.5-5.0 keV band. Assuming a power-law spectrum with photon index 1.7 and galactic absorption, the X-ray luminosity of the active nucleus is 4.2$\times$10$^{40}$ ergs s$^{-1}$ in the 0.1-10.0 keV band (unabsorbed). The X-ray luminosity of the nucleus is roughly an order of magnitude larger than that expected based on the correlation of X-ray and radio cores \citep{can99}. This may be a considerable underestimate if the nucleus is heavily absorbed, but is typical of that found in other ``normal'' elliptical galaxies \citep{jon05}. The surface brightness profile in a 60$^\circ$ sector to the North of NGC 7618, shown in Figure~\ref{sbprof}, drops by approximately a factor of two across the discontinuity. At the sharpest region of the edge to the NE of the nucleus, the surface brightness drops by a factor of 4 over a distance of $\sim$4$''$ (1.4 kpc). The surface brightness profile in a 30$^\circ$ sector to the NE is shown in Figure~\ref{newedge}. The morphology of this discontinuity is similar to that seen in {\em Chandra} observations of cluster 'cold-fronts', although there is no evidence for a temperature discontinuity between the two moving fluids as commonly seen in clusters of galaxies \citep{vik01,mar01,maz01}. We fitted absorbed APEC models to the spectra in 90$^\circ$ sectors of two annular regions, one inside the discontinuity and one outside. The thickness of the inner and outer annuli were 33.0$''$ (11.6 kpc) and 66.4$''$ (23.2 kpc), respectively. The gas temperature does not change significantly across the discontinuity ($T_{inner}$=0.785$\pm$0.025 keV, $T_{outer}$=0.810$\pm$0.070 keV). There is marginal evidence for a jump in the elemental abundance ($Z_{inner}$=0.65$\pm$0.25, $Z_{outer}$=0.17$\pm$0.21 - all uncertainties at 90\% confidence), however. This suggests that the discontinuity could be an `abundance' front due to a sharp discontinuity in the elemental abundance, and therefore the emissivity of the gas as observed in NGC 507 \citep{kra03}. The temperature and elemental abundance structure in the gas is probably more complex than we have assumed here, but the short exposure time and limited quality of the data prevent a more detailed analysis. Fitting power laws to the surface brightness profiles interior to and exterior to the discontinuity and assuming hydrostatic equilibrium, we find a large change in the power law index across the discontinuity, $\beta_{in}$=0.30 and $\beta_{out}$=0.64. The change in $\beta$ is almost certainly not related to a change in the gravitational potential and suggestive of non-hydrostatic gas motions. We estimate the gas density on both sides of the discontinuity by deprojecting the surface brightness profile (assuming spherical symmetry) and find proton densities of 6.0$^{+1.2}_{-0.8}$ and 5.0$^{+2.4}_{-1.2}$ $\times$10$^{-3}$ inside and outside the discontinuity, respectively. The upper limit of the velocity of the gas interior to the discontinuity estimated from the maximum pressure difference is Mach 0.9 or $\sim$ 420 km s$^{-1}$ \citep{vik01}. The X-ray morphology of the ASCA/GIS image (Figure~\ref{gisimg}) suggests that the two groups reside in a larger-scale dark matter potential. A third X-ray peak is seen 8.5$'$ to the east of NGC 7618, and diffuse emission extends at least 15$'$ to the north and 10$'$ to the south. Any gas that resides in this larger scale dark matter halo should be hotter than the gas in the NGC 7618 group. We fitted the ASCA/GIS spectrum in three regions, two 7$'$ radius circles centered on each galaxy and a third region 15$'$ in radius centered between NGC 7618 and UGC 12491 excluding the two 7$'$ radius circles centered on each galaxy. The radius of the two circles centered on NGC 7618 and UGC 12491 corresponds to the 90\% encircled energy radius for the ASCA/GIS in the 1-2 keV band. Background was determined from ASCA/GIS high-latitude, blank sky observations taken from the HEASARC and generated using the FTOOL `mkgisbgd'. We fit absorbed, single temperature APEC models to each spectrum with $N_H$ frozen at the Galactic value and the elemental abundance frozen at 0.6$Z_\odot$. The results of these fits are summarized in the top half of Table~\ref{spectab}. Single temperature models are poor fits for the regions centered on NGC 7618 and UGC 12491, but provide an adequate description of the diffuse emission between the galaxies. The temperature of the diffuse gas between NGC 7618 and UGC 12491 is 2.32$^{+0.50}_{-0.34}$. We also fit two temperature models to the two spectra extracted from the regions centered on NGC 7618 and UGC 12491. As before, both the $N_H$ and the elemental abundance were frozen. For NGC 7618, we also froze the temperature of one component at 0.8 keV, the value determined from the ACIS-S spectral fitting, to reduce the number of free parameters. If this parameter is allowed to freely vary, the fit value is consistent with 0.8 keV. For UGC 12491, the temperatures of both components were allowed to freely vary. The results of these fits are summarized in the bottom half of Table~\ref{spectab}. The two temperature model provides acceptable fits in both cases, and the temperature of the hotter component is $\sim$2.3 keV. \section{Discussion} There are at least four possible explanations for the observed X-ray structures. First, the complex X-ray morphology could be the result of a radio lobe/IGM interaction. NGC 7618 is a radio source. It is detected in the NVSS with a flux density of 20 mJy, with some evidence of extension NW/SE. At higher frequencies (5 and 8 GHz) two 15-min archival VLA observations only detect a point-like core coincident with the center of the galaxy, with a flux density of 4.5 mJy at 4.9 GHz and 3.0 mJy at 8.4 GHz. However, at lower frequencies, the flux density is much higher: 1.2 Jy in the 151-MHz 6C catalog \citep{hal93}, 0.26 Jy in the 408-MHz B3 catalog \citep{fic85}, and 1.0 Jy in the 74-MHz VLA Low-Frequency Sky Survey (VLSS: http://lwa.nrl.navy.mil/VLSS/). The high flux density at low frequencies suggests that there may be a relic radio source, the aged synchrotron plasma from a more energetic phase of the active nucleus. The X-ray morphology is very different from that seen in other examples of radio lobe/ISM interactions, however. There are no obvious X-ray cavities as are commonly seen in such radio lobe/ISM interactions (e.g. \citep{mcn00,jon02,hei02}), nor is there evidence of hot, shock-heated shell that would be present if the radio lobes were expanding supersonically \citep{kra03}. It is not clear whether any of these features would be expected in the case of a relic radio source, however. A sensitive low-frequency radio map would give us a better understanding of the possible interaction between radio-emitting plasma and the hot gas. Second, it is possible that the NGC 7618 gas is oscillating, or 'sloshing', in the gravitational potential because of a recent merger/interaction with a lower mass sub-group. A dust lane has been detected in the optical host galaxy, perhaps indicative of a recent merger and adding support to this hypothesis \citep{col01}. Similar structures (on larger scales) to those presented here have been seen in X-ray observations of clusters of galaxies \citep{mar01,tit05}. In this scenario, the gas is oscillating, or `sloshing', in the dark matter potential due to a recent merger with a sub-cluster. The infall of the sub-cluster drives a shock into the gas that displaces it from the gravitating matter. The gas oscillates around the center of mass, eventually returning to hydrostatic equilibrium via viscous dissipation. Such processes can contribute a significant amount of energy to the gas and may disrupt or prevent the formation of cooling flows in clusters of galaxies. If we naively assume the gas on both sides of the discontinuity is in hydrostatic equilibrium, we find an unphysical discontinuity in the gravitating mass. The apparent mass discontinuity is the result of the non-zero acceleration of the gas inside the discontinuity. The distribution of the gas will reflect the reduced gravity. That is, the acceleration term in Euler's equation is non-zero for gas inside the discontinuity. Assuming that the core is at maximum displacement from the center, the gravitational potential energy of the gas is $U\sim G\Delta M r^{-1} \sim$7$\times$10$^{14}$ ergs gm$^{-1}$, or roughly one third of the thermal energy of the gas. There is one important difficulty with this interpretation, however. On tens of kpc scales, the emission peak of the gas and the optical galaxy (NGC 7618) are both offset (to the N) relative to the larger scale X-ray isophotes (shown green in Figure~\ref{acisimgb}). If the sloshing scenario is correct only the gas and not the galaxy/gravitating mass, should be moving relative to the large scale dark matter halo. The host galaxy should be resting at the center of the gas distribution. In addition, the peak of the X-ray emission lies to the East of the nucleus of the host galaxy, but the larger scale `sloshing' is North/South. This suggests that there are significant gas motions along perpendicular axes, and that the gas is coupled to the gravitating mass (the stars) along the North/South axis, but partially decoupled along the East/West axis. It is difficult to see how there could be the case unless the merger were off-axis. If this is the case, the merging galaxy should be detectable by an optical census. The third possibility is that the observed structures are the result of a 'near-miss' flyby of the nearby UGC 12491 group. In this scenario, the UGC 12491 group passed by NGC 7618 from the SE toward its current position to the WNW, and the complex X-ray morphology of NGC 7618 is the result of fluid motions in the gas induced by the gravitational and hydrodynamical interaction. The observed X-ray morphology is the result of the NGC 7618 core being displaced relative to the larger scale gravitational potential, and the surface brightness discontinuity represents a discontinuity in the elemental abundance. X-ray observations of groups and clusters show that the elemental abundance is strongly peaked toward the center. The displacement of the core of a group relative to its halo could create the observed structures. Hydrodynamic simulations of head-on collisions of equal mass clusters show that the close approach of the cores can create strong non-hydrostatic motions in the gas \citep{rot97}. Elongation of the dark matter potential, and thus the gas distribution, is a natural consequence of these collisions. In these simulations, the gas becomes partially separated from the gravitating matter and is not a good tracer of the dark matter distribution. The separation between UGC 12491 and NGC 7618 was probably (much) smaller in the past in order to create such a large disturbance in the NGC 7618 gas. The simulations of merging clusters show little separation of the gas from the dark matter as the two clusters initially approach each other. It is only as the cores collide/merge and subsequently separate that significant gas velocities develop. If the current positions of the two groups are, in fact, their closest approach so far, non-hydrostatic effects should just be starting to manifest themselves. Therefore, the two groups must have already passed each other at least once. There are several difficulties with this scenario, however. If the hotter gas seen in the ASCA/GIS spectrum is shock-heated group gas, the measured temperature ratio (0.8 to 2.3 keV) implies an infall/approach velocity of the two groups of $\sim$1170 km s$^{-1}$ \citep{lan89}, much larger than typical peculiar velocities. It would also be difficult to explain the large amount of hot gas ($\sim$10$^{12}$ M$_\odot$) seen in the ASCA/GIS observation in this scenario under the assumption that this component is entirely shock-heated group gas. The total (thermal) energy of the hotter component is only $\sim$1.3$\times$10$^{61}$ ergs, a reasonable value for the collison of two groups (we note that an AGN outburst could easily supply this much energy to the gas as well \citep{mcn04}). If the hot component is, in fact, shock-heated group gas, it is not bound to the group potential and a transient phenomenon, escaping as a wind. It is likely that these two groups are gravitationally bound. Ignoring angular momentum (which is not believed to be important on large scales \citep{hof86}) and dissipative forces, and assuming that the mass of each group is 10$^{13}$ M$_\odot$, the escape velocity of this pair is $\sim$550 km s$^{-1}$. If either of these groups has a velocity relative to the center of mass larger than this value, the pair will be unbound. This is larger than typical peculiar velocities of galaxies/groups, and it is reasonable to conclude that the system is bound. Significant angular momentum will decrease this value, whereas dissipative forces will increase it. The dynamic parameters of this pair are too poorly known to make a definitive statement, however. The fourth possibility is that NGC 7618 is falling into a larger gravitational potential. In this scenario, either UGC 12491 is at the center of the potential and NGC 7618 is falling into it, or both NGC 7618 and UGC 12491 are falling into a larger scale potential. The existense of the hotter ($\sim$2.3 keV) gas component in the ASCA/GIS observation (Figure~\ref{gisimg}) supports this scenario. The X-ray structures seen in the gas around NGC 7618 are then the result of ram-pressure stripping. This is the classic `cold-front' scenario discussed by \citet{vik01}. The effects of ram-pressure stripping on the hot gas atmospheres of early-type galaxies falling into clusters have been studied in {\em Chandra} observations of NGC 4472, NGC 4552 (falling into the Virgo cluster) and NGC 1404 (falling into/toward NGC 1399, the dominant member of the Fornax cluster) \citep{bil04,mac05a,mac05b}. In these cases, a sharp surface brightness discontinuity in the direction of infall and a diffuse tail in the opposite direction have been observed. The gas around NGC 7618 is being stripped by hotter, lower density gas that presumably resides in the larger/deeper dark matter halo. If this scenario is correct, either NGC 7618 is falling into UGC 12491 or both groups are merging with/falling into a larger dark matter potential in which there is no central dominant elliptical galaxy. The high temperature of the second spectral component in the ASCA analysis and the lack of an azimuthally symmetric X-ray halo around UGC 12491 (Figure~\ref{gisimg}) support the latter hypothesis. This phenomenon has in fact already been observed on a smaller scale in the Pegasus I group \citep{kra05}. Neither of the massive ellipticals in this group, NGC 7619 or NGC 7626, lie at the center of the extended X-ray halo. {\em Chandra} observations of NGC 7619 show a sharp surface brightness discontinuity to the NE (presumably the direction of infall), and an extended, ram-pressure stripped tail in the opposite direction. If we assume the hotter (2.3 keV) component is in hydrostatic equilibrium with the dark matter potential and that the gas density follows a beta-model profile with $\beta$=0.67 and $r_0$=50 kpc (typical for clusters of galaxies), the gravitating mass, $M_{grav}$, within a radius of 325 kpc of the midpoint between NGC 7618 and UGC 12491 is $\sim$5.6$\times$10$^{14}$ M$_\odot$. The gas mass, $M_{gas}$, required to account for the observed ASCA/GIS flux of the hotter component within this radius is $\sim$7$\times$10$^{11}$ M$_\odot$ for this density profile. Ignoring the stellar component, which is insignificant on these spatial scales, the baryon fraction, $f=M_{gas}/M_{grav}$ is $\sim$1.5\%. Such a low value of baryon fraction is not improbable for a 2.3 keV cluster. We speculate that this larger dark matter halo may be a `failed' cluster in which much of the gas was blown off during a cataclysmic event early in its formation (either a merger or a powerful AGN remnant). We caution, however, that the spatial resolution of the ASCA/GIS is poor and our knowledge about the morphology of the gas limited. In addition, this gas may not be in hydrostatic equilibrium, so the uncertainties on both (gas and gravitating) mass estimates are large. A moderate XMM-Newton observation of this pair could measure the temperature and morphology of the hotter component and resolve this issue. We note that in this scenario the surface brightness discontinuity north of the NGC 7618 nucleus cannot be the stagnation point between the hot gas in NGC 7618 and the gas of this putative larger scale halo. If the observed X-ray surface brightness discontinuity is in fact the stagnation point between these two gases, the temperature of the gas exterior to the discontinuity must be on the order or higher than that of the halo gas (from Bernoulli's equation). The stagnation point therefore must lie beyond the observable emission, and the surface brightness discontinuity represents a contact discontinuity between two fluids. Unless the infall is highly supersonic, the gas interior to the stagnation point should not be highly disturbed and should remain in rough hydrostatic equilibrium in the gravitational potential of NGC 7618. A deeper {\em Chandra} observation of the central regions of NGC 7618 is required to elucidate the hydrodynamics of the gas. \section{Conclusions} We have observed a sharp surface brightness discontinuity in the X-ray emission from the hot gas in the NGC 7618 group, and an X-ray `tail' extending 70 kpc in the opposite direction in an 8 ks {\em Chandra}/ACIS-S observation. Archival ASCA/GIS observations indicate the presence of a hotter (2.3 keV) component, although the morphology of this gas is poorly constrained. We conclude that there are three possible explanations for these features. First, the NGC 7618/UGC 12491 pair underwent a recent `near-miss' flyby. If this is the case, this pair is the nearest early-stage merger of two roughly equal mass groups. Second, UGC 12491 may be at the center of a cluster, and NGC 7618 is falling into it. Third, NGC 7618 and UGC 12491 are both falling into a gas poor cluster with no dominant central elliptical galaxy. Whether the observed features are the result of a group-group merger, or the infall of two groups into a larger dark matter potential, the NGC 7618/UGC 12491 pair is one of the best examples of an ongoing merger in the local Universe. Deeper X-ray observations are required to better constrain the thermodynamic parameters of the gas in the central regions and the larger scale halo. Radio observations will be critical in assessing the role of radio plasma/IGM interactions. If the observed X-ray features are the result of ram-pressure stripping or a merger interaction between the groups, the effects on the relic radio halo are likely to have been dramatic. A detailed optical census including velocities of the other galaxies in the both groups will be useful to constrain their dynamics and their relationship to the larger dark matter potential. \acknowledgements This work was supported by NASA contracts NAS8-38248, NAS8-39073, the Chandra X-ray Center, and the Smithsonian Institution. We would like to thank the anonymous referee for comments that improved this paper. \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \begin{table} {\small \begin{center} \begin{tabular}{|l|c|c|}\hline & NGC 7618 & UGC 12491 \\ \hline $m_B$ & 14.0 & 14.9 \\ \hline $M_B$ & -21.3 & -20.2 \\ \hline $z$ & 0.017309 & 0.017365 \\ \hline Distance (Mpc) & 74.1 & 74.3 \\ \hline $L_X$ (ergs s$^{-1}$) & 6.9$\times$10$^{42}$ & 6.2$\times$10$^{42}$ \\ \hline X-ray Radius & $\sim$170 kpc & $\sim$200 kpc \\ \hline \end{tabular} \caption{Summary of the X-ray and optical properties of the NGC 7618 and UGC 12491 galaxies. The X-ray luminosity is in the 0.1-10 keV bandpass (unabsorbed) within 7$'$ (146 kpc) of the nucleus. Absolute magnitudes have been corrected for extinction.}\label{galtab} \end{center} } \end{table} \clearpage \begin{table} {\small \begin{center} \begin{tabular}{|l|c|c|c|}\hline & NGC 7618 & UGC 12491 & Diffuse \\ \hline\hline \multicolumn{4}{|c|}{Single Temperature Fits} \\ \hline $k_BT$ (keV) & 1.43$^{+0.08}_{-0.14}$ & 1.32$^{+0.10}_{-0.13}$ & 2.32$^{+0.50}_{-0.34}$ \\ \hline Flux & 3.06$\times$10$^{-12}$ & 5.37$\times$10$^{-12}$ & 2.53$\times$10$^{-12}$ \\ \hline $\chi^2_\nu$ & 1.95 & 1.60 & 0.73 \\ \hline\hline \multicolumn{4}{|c|}{Two Temperature Fits} \\ \hline $k_BT_1$ (keV) & 0.80 & 0.86$^{+0.16}_{-0.09}$ & \\ \hline Flux & 2.26$\times$10$^{-12}$ & 3.69$\times$10$^{-12}$ & \\ \hline $k_BT_2$ (keV) & 2.21$^{+0.33}_{-0.67}$ & 2.26$^{+1.19}_{-0.35}$ & \\ \hline Flux & 1.48$\times$10$^{-12}$ & 2.15$\times$10$^{-12}$ & \\ \hline $\chi^2_\nu$ & 1.03 & 0.78 & \\ \hline\hline \end{tabular} \caption{Best-Fit Temperatures and Fluxes of the ASCA/GIS data in three regions. All uncertainties are 90\% confidence for one parameter of interest. Units of fluxes are ergs cm$^{-2}$ s$^{-1}$ (unabsorbed) in the 0.5-2.0 keV band.}\label{spectab} \end{center} } \end{table}
Title: The Local Effects of Cosmological Variations in Physical 'Constants' and Scalar Fields II. Quasi-Spherical Spacetimes
Abstract: We investigate the conditions under which cosmological variations in physical `constants' and scalar fields are detectable on the surface of local gravitationally-bound systems, such as planets, in non-spherically symmetric background spacetimes. The method of matched asymptotic expansions is used to deal with the large range of length scales that appear in the problem. We derive a sufficient condition for the local time variation of the scalar fields driving variations in 'constants' to track their large-scale cosmological variation and show that this is consistent with our earlier conjecture derived from the spherically symmetric problem. We perform our analysis with spacetime backgrounds that are of Szekeres-Szafron type. They are approximately Schwarzschild in some locality and free of gravitational waves everywhere. At large distances, we assume that the spacetime matches smoothly onto a Friedmann background universe. We conclude that, independent of the details of the scalar-field theory describing the varying `constant', the condition for its cosmological variations to be measured locally is almost always satisfied in physically realistic situations. The very small differences expected to be observed between different scales are quantified. This strengthens the proof given in our previous paper that local experiments see global variations by dropping the requirement of exact spherical symmetry. It provides a rigorous justification for using terrestrial experiments and solar system observations to constraint or detect any cosmological time variations in the traditional `constants' of Nature in the case where non-spherical inhomogeneities exist.
https://export.arxiv.org/pdf/gr-qc/0601056
\title{The Local Effects of Cosmological Variations in Physical 'Constants' and Scalar Fields \\ II. Quasi-Spherical Spacetimes} \author{Douglas J. Shaw} \affiliation{DAMTP, Centre for Mathematical Sciences, University of Cambridge, Wilberforce Road, Cambridge CB3 0WA, UK} \author{John D. Barrow} \affiliation{DAMTP, Centre for Mathematical Sciences, University of Cambridge, Wilberforce Road, Cambridge CB3 0WA, UK} \date{\today} \section{\protect\bigskip Introduction} Over the past few years there has been a resurgence of observational and theoretical interest in the possibility that some of the fundamental `constants' of Nature might be varying over cosmological timescales \cite% {webb}. In respect of two such `constants', the fine structure constant, $% \alpha $, and Newton's `constant' of gravitation, $G$, the idea of such variations is not new, and was proposed by authors such as Milne \cite{milne}% , Dirac \cite{dirac}, and Gamow \cite{gam} as a solution to some perceived cosmological problems of the day \cite{btip}. At first, theoretical attempts to model such variations in constants were rather crude and equations derived under the assumption that constants like $G$ and $\alpha $ are true constants were simply altered by writing-in an explicit time variation. This approach was first superseded in the case of varying $G$ by the creation of scalar-tensor theories of gravity \cite{jordan}, culminating in the standard form of Brans and Dicke \cite{bd} in which $G$ varies through a dynamical scalar field which conserves energy and momentum and contributes to the curvature of spacetime by a means of a set of generalised gravitational field equations. More recently, such self-consistent descriptions of the spacetime variation of other constants, like $\alpha $ \cite{bek, bsm}, the electroweak couplings \cite{ewk}, and the electron-proton mass ratio, $\mu $% , \cite{bm} have been formulated although most observational constraints in the literature are imposed by simply making constants into variables in formulae derived under the assumption that are constant. The resurgence of interest in possible time variations in $\alpha $ and $\mu $ has been brought about by significant progress in high-precision quasar spectroscopy. In addition to quasar spectra, we also have available a growing number of laboratory, geochemical, and astronomical observations with which to constrain any local changes in the values of $\ $these constants \cite{reviews}. Studies of the variation of other constants, such as $G$, the electron-proton mass ratio, $\mu =m_{e}/m_{pr}$, and other standard model couplings, are confronted with an array of other data sources. The central question which this series of papers addresses is how to these disparate observations, made over vastly differing scales, can be combined to give reliable constraints on the allowed global variations of $% \alpha $ and the other constants. If $\alpha $ varies on cosmological scales that are gravitationally unbound and participate in the Hubble expansion of the universe, will we see any trace of this variation in a laboratory experiment on Earth? After all, we would not expect to find the expansion of the universe revealed by any local expansion of the Earth. In Paper I \cite% {shawbarrow1}, we examined this question in detail for spherically symmetric inhomogeneous universes that model the situation of a planet or a galaxy in an expanding Friedmann-Robertson-Walker (FRW)-like universe. In this paper we relax the strong assumption of spherical symmetry and examine the situation of local observations in a universe that contains non-spherically symmetric inhomogeneity. Specifically, we use the inhomogeneous metrics found by Szekeres to describe a non-spherically symmetric universe containing a static star or planet. As in Paper I, we are interested in determining the difference (if any) between variations of a supposed 'constant' or associated scalar field when observed locally, on the surface of the planet or star, and on cosmological scales. When a `constant', $\mathbb{C}$, is made dynamical we can allow it to vary by making it a function of a new scalar field, $\mathbb{C}\rightarrow \mathbb{C}(\phi )$, that depends on spacetime coordinates: $\phi =\phi (\vec{% x},t)$. It has become general practice to combine take all observational bounds on the allowed variations of $\mathbb{C}(\phi )$. This practice assumes implicitly that any time variation of $\mathbb{C,}$ on or near the Earth, is comparable to any cosmological variation that it might experience, that is to high precision \begin{equation} \dot{\phi}(\vec{x},t)\approx \dot{\phi}_{c}(t), \label{wettcond} \end{equation}% \noindent for almost all locations $\vec{x}$, where $\phi _{c}$ is the cosmological value of $\phi $. This assumption is always made without proof, and there is certainly no \emph{a priori} reason why it should be valid. Strictly, $\phi $ mediates a new or `fifth' force of Nature. If the assumed behaviour is correct then this force is unique amongst the fundamental forces in that its value locally reflects its cosmological variation directly. In this series of papers we are primarily interested in theories where the scalar field, or `dilaton' as we shall refer to it, $\phi $, evolves according to the conservation equation \begin{equation*} \square \phi =B_{,\phi }(\phi )\kappa T-V_{,\phi }(\phi ), \end{equation*}% where $T$ is the trace of the energy momentum tensor, $T=T_{\mu }^{\mu }$, (with the contribution from any cosmological constant neglected). We absorb any dilaton-to-cosmological constant coupling into the definition of $V(\phi )$. The dilaton-to-matter coupling $B(\phi )$ and the self-interaction potential, $V(\phi )$, are arbitrary functions of $\phi $ and units are defined by $\kappa =8\pi G$ and $c=\hslash =1$. This covers a wide range of theories which describe the spacetime variation of `constants' of Nature; it includes Einstein-frame Brans-Dicke (BD) and all other, single-field, scalar-tensor theories of gravity \cite{bd, bsm, poly, posp}. In cosmologies that are composed of perfect fluids and a cosmological constant, it will also contain the Bekenstein-Sandvik-Barrow-Magueijo (BSBM) theory of varying $\alpha $, \cite{bsm}, and other single-dilaton theories which describe the variation of standard model couplings, \cite{posp}. We considered some other possible generalisations in \cite{shawbarrow1}. It should be noted that our analysis and results apply equally well to any theory which involves weakly-coupled, `light', scalar fields, and not just those that describe variations of the standard constants of physics. In first paper of this series, \cite{shawbarrow1}, we determined the conditions under which condition \ref{wettcond} would hold near the surface of a virialised over-density of matter, such as a galaxy or star, or a planet, such as the Earth, under the assumption of spherical symmetry. We chose to refer to this object as our `star'. In Paper I, matched asymptotic expansions were employed to analyse the most general, \emph{% spherically-symmetric}, dust plus cosmological constant embeddings of the `star' into an expanding, asymptotically homogeneous and isotropic spherically symmetric universe. We proved that, independent of the details of the scalar-field theory describing the varying `constant', that \ref% {wettcond} is almost always satisfied under physically realistic conditions. The latter condition was quantified in terms of an integral over sources that can be evaluated explicitly for any local spherical object. In this paper we extend that analysis, and our main result, to a class of embeddings into cosmological background universes that possess \emph{no} Killing vectors i.e. \emph{no} symmetries. The mathematical machinery that we use to do this is, as before, the method of matched asymptotic expansions, employed in \cite{shawbarrow1}, where the technical machinery is described in detail. A summary of the results obtained there can also be found in \cite{shawbarrowlett}. This paper is organised as follows: We shall firstly provide a very brief summary of the method of matched asymptotic expansions used here. In section II we will introduce the geometrical set-up that we will use. We will be working in spacetime backgrounds of Szekeres-Szafron type \cite{szek, szafron}. We describe theses particular solutions of Einstein's equations briefly in section II and then in greater detail in section III. In section IV we extend the analysis of \cite{shawbarrow1} to include non-spherically symmetric backgrounds of Szekeres-Szafron type. In section V, we consider the validity of the approximations used, and state the conditions under which they should be expected to hold. In section VI we perform the matching procedure (as outlined below), and extend the main result of \cite% {shawbarrow1} to Szekeres-Szafron spacetimes. We consider possible generalisations of our result in section VII before considering the implications of the results in the section VIII. We will employ the method of matched asymptotic expansions \cite{hinch, Death}. We solve the dilaton conservation equations as an asymptotic series in a small parameter, $\delta $, about a FRW background and the Schwarzschild metric which surrounds our star. The deviations from these metrics are introduced perturbatively. The former solution is called the \emph{exterior expansion} of $\phi $, and the latter the \emph{interior expansion} of $\phi $. The exterior expansion is found by assuming that the length and time scales involved are of the order of some intrinsic exterior length scale, $L_{E}$. Similarly in the interior expansion we assume all length and time scales we be of the of $L_{I}$, the interior length scale. Neither of the two different expansions will be valid in both regions. In general, we define $\delta :=L_{I}/L_{E}\ll 1$. This means that in general only a subset of our boundary conditions we will be enforceable for each expansion, and as a result both the interior and exterior solutions will feature unknown constants of integration. To remove this ambiguity, and fully determine both expansions, we used the formal matching procedure. The idea is to assume that both expansions are valid in some intermediate region, where length scales go like $L_{int}=L_{I}^{\alpha }L_{E}^{1-\alpha } $, for some $\alpha \in (0,1)$. Then by the uniqueness property of asymptotic expansions, both solutions must be equal in that intermediate region. This allows us to set the value of constants of integration, and effectively apply \emph{all} the boundary conditions to both expansions. A fuller discussion of this method, with examples, and its application in general relativity is given in \cite{shawbarrow1}. \section{Geometrical Set-Up} We shall consider a similar geometrical set-up to that of Paper I. We assume that the dilaton field is only weakly coupled to gravity, and so its energy density has a negligible effect on the expansion of the background universe. This allows us to consider the dilaton evolution on a fixed background spacetime. We will require this background spacetime to have the same properties as in Paper I, but with the requirement of spherical symmetry removed: \begin{itemize} \item The metric is approximately Schwarzschild, with mass $m$, inside some closed region of spacetime outside a surface at $r=R_{s}$. The metric for $% r<R_{s}$ is left unspecified. \item Asymptotically, the metric must approach FRW and the whole spacetime should tend to the FRW metric in the limit $m\rightarrow 0$. \item When the local inhomogeneous energy density of asymptotically FRW spacetime tends to zero, the spacetime metric exterior to $r=R_{s}$ must tend to a Schwarzschild metric with mass $m$ . \end{itemize} We will also limit ourselves to considering spacetimes in which the background matter density satisfies a physically realistic equation of state, specifically that of pressureless dust ($p=0$). We also allow for the inclusion of a cosmological constant, $\Lambda $. The set of all non-spherical spacetimes that satisfy these conditions is too large and complicated for us to examine fully here; and such an analysis is beyond the scope of this paper. We can simplify our analysis greatly, however, we specify four further requirements: \begin{enumerate} \item The flow-lines of the background matter are geodesic and non-rotating. This implies that the flow-lines are orthogonal to a family of spacelike hypersurfaces, $S_{t}$. \item Each of the surfaces $S_{t}$ is conformally flat. \item The Ricci tensor for the hypersurfaces $S_{t}$, ${}^{(3)}R_{ab}$, has two equal eigenvalues. \item The shear tensor, as defined for the pressureless dust background, has two equal eigenvalues. \end{enumerate} The last three of these conditions seem rather artificial; however, when the deviations from spherical symmetry are in some sense `small' we might expect them to hold as a result of the first condition. In the spherically symmetric case, condition 1 implies conditions 2, 3 and 4. In the absence of spherical symmetry, these conditions require the background spacetime to be of Szekeres-Szafron type, containing pressureless matter and (possibly) a cosmological constant. The conditions (1 - 4) combined with the background matter being of perfect fluid type provide an invariant definition of the Szekeres-Szafron class of metrics that is due to Szafron and Collins \cite% {collins, kras}. We have demanded that the `local' or interior region be approximately Schwarzschild. The intrinsic length scale of the interior is defined by the curvature invariant there: \begin{equation} L_{I}\equiv \left( \tfrac{1}{12}R_{abcd}R^{abcd}\right) ^{-1/4}=\frac{% R_{s}^{3/2}}{\left( 2m\right) ^{1/2}}. \label{invar} \end{equation}% The exterior (or cosmological) region is approximately FRW, and so its intrinsic length scale is proportional to the inverse square root of the local energy density: $1/\sqrt{\kappa \varepsilon +\Lambda }$, where $% \varepsilon $ is the matter density. In accord with current astronomical observations, we assume that this FRW region is approximately flat, and so we set our exterior length scale appropriate for the present epoch, $t=t_{0}, $ by the inverse Hubble parameter at that time: \begin{equation*} L_{E}\equiv 1/H_{0}. \end{equation*}% We can now define a small parameter by the ratio of the interior and exterior length scales: \begin{equation*} \delta =L_{I}/L_{E}. \end{equation*} \section{Szekeres-Szafron Backgrounds} In 1975 Szekeres \cite{szek} solved the Einstein equations with perfect fluid source by assuming a metric of the form: \begin{equation*} \mathrm{d}s^{2}=\mathrm{d}t^{2}-e^{2\alpha }\mathrm{d}r^{2}-e^{2\beta }\left( \mathrm{d}x^{2}+\mathrm{d}y^{2}\right) , \end{equation*}% with $\alpha $ and $\beta $ being functions of $(t,r,x,y)$. The coordinates where assumed to be comoving so that the fluid-flow vector is of the form: $% u^{\mu }=\delta _{0}^{\mu }$; This implies $p=p(t)$ and $\ $the acceleration $\dot{u}^{\mu }=0$. Szekeres assumed a dust source with no cosmological constant, $p=0$, although his results were later generalised to arbitrary $% p(t)$ by Szafron \cite{szafron} and the explicit dust plus $\Lambda $ solutions were found by Barrow and Stein-Schabes \cite{JBJSS}.\emph{\ }In general, these metrics have \emph{no} Killing symmetries \cite{bonn}. Spherically-symmetric solutions of this type with $\alpha (r,t)$ and $\beta (r,t)$ were, in fact, first discussed by Lemaоtre \cite{lem} and are usually referred to as the Tolman-Bondi models \cite{tolbondi}; much of the analysis of Paper I assumed a Tolman-Bondi background. The Szekeres-Szafron models can be divided into two classes: $\beta _{,r}=0$ and $\beta _{,r}\neq 0$. Both classes include all FRW models in their homogeneous and isotropic limit; however, only the latter 'quasi-spherical' class includes the external Schwarzschild solution. Since we want to have some part of our spacetime look Schwarzschild we will only consider the $% \beta _{,r}\neq 0$ quasi-spherical solutions. We will also limit ourselves to spacetimes with a cosmological constant, \cite{JBJSS}, so in effect the total pressure is $p=-\Lambda $. These universes contain no gravitational radiation as can be deduced from the existence of Schwarzschild as a special case which ensures a smooth matching to Schwarzschild, which contains no gravitational radiation. With these restrictions, $\alpha $ and $\beta $ are given by: \begin{eqnarray} e^{\beta } &=&\Phi (t,r)e^{\tilde{\nu}(r,x,y)}, \\ e^{\alpha } &=&h(r)e^{-\tilde{\nu}(r,x,y)}\left( e^{\beta }\right) _{,r}, \\ e^{-\tilde{\nu}} &=&\tilde{A}(r)(x^{2}+y^{2})+2\tilde{B}_{1}(r)x+2\tilde{B}% _{2}(r)y+\tilde{C}(r), \end{eqnarray}% where $\Phi (t,r)$ satisfies: \begin{equation*} \Phi _{,t}^{2}=-\tilde{k}(r)+2\tilde{M}(r)/\Phi +\frac{1}{3}\Lambda \Phi ^{2}. \end{equation*}% The functions $\tilde{A}(r)$, $\tilde{B}_{1}(r)$, $\tilde{B}_{2}(r)$, $% \tilde{C}(r)$, $\tilde{M}(r)$, $\tilde{k}(r)$ and $\tilde{h}(r)$ are arbitrary up to the relations: \begin{equation*} \tfrac{1}{4}E(r):=\tilde{A}\tilde{C}-\tilde{B}_{1}^{2}-\tilde{B}_{2}^{2}=% \tfrac{1}{4}\left[ \tilde{h}^{-2}(r)+\tilde{k}(r)\right] . \end{equation*}% The surfaces $(t,r)=const$ have constant curvature $E(r)$. We will require that the inhomogeneous region of our spacetime is localised, so that it is by some measure finite. This implies that the surfaces of constant curvature must be closed; we must therefore restrict ourselves to only considering backgrounds where $E>0$. Whenever this is the case, we always can rescale the arbitrary functions so that $E\ $can be set equal to $1$ by the rescalings \begin{eqnarray} A(r):= &&\tilde{A}(r)/\sqrt{E(r)},\;B_{1}(r):=\tilde{B}_{1}(r)/\sqrt{E(r)}% ,\;B_{2}(r):=\tilde{B}_{2}(r)/\sqrt{E(r)},\;C(r):=\tilde{C}(r)/\sqrt{E(r)}% ,e^{\nu }:=\sqrt{E}e^{\tilde{\nu}} \notag \\ k:= &&\tilde{k}(r)/E(r),\;h^{-2}:=\tilde{h}(r)^{-2}/E(r)=1-k(r),\;R(t,r):=% \Phi (r,t)/\sqrt{E},\;M(r)=\tilde{M}(r)/E^{3/2}. \notag \end{eqnarray}% These transformations can be viewed as the `gauge-fixing' of arbitrary functions. In this gauge, $R(t,r)$ is a `physical' radial coordinate, i.e. the surfaces $(t,r)=const$ have surface area $4\pi R^{2}$ and the metric becomes \begin{equation*} \mathrm{d}s^{2}=\mathrm{d}t^{2}-\frac{\left( 1+\nu _{,R}R\right) ^{2}R_{,r}^{2}\mathrm{d}r^{2}}{1-k(r)}-R^{2}e^{2\nu }\left( \mathrm{d}x^{2}+% \mathrm{d}y^{2}\right) , \end{equation*}% \noindent where $e^{-\nu }=A(r)(x^{2}+y^{2})+2B_{1}(r)x+2B_{2}(r)y+C(r)$ and $AC-B_{1}^{2}-B_{2}^{2}=\tfrac{1}{4}$, and $\nu _{,R}:=\nu _{,r}/R_{,r}$ and: \begin{equation*} R_{,t}^{2}=-k(r)+2M(r)/R+\frac{1}{3}\Lambda R^{2}. \end{equation*}% In this quasi-spherically symmetric subcase of the Szekeres-Szafron spacetimes the surfaces of constant curvature, $(t,r)=const$, are 2-spheres \cite{szek2}; however, they are not necessarily concentric. In the limit $% \nu _{,r}\rightarrow 0$, the $(t,r)=const$ spheres becomes concentric (see fig. \ref{fig1}). We can make one further coordinate transformation so that the metric on the surfaces of constant curvature, $(t,r)=const$, is the canonical metric on $S^{2}$ i.e. $\mathrm{d}\theta ^{2}+\sin ^{2}\theta \mathrm{d}\phi ^{2}$: \begin{eqnarray} x\rightarrow X &=&2\left( A(r)x+B_{1}(r)\right) , \notag \\ y\rightarrow Y &=&2\left( A(r)y+B_{2}(r)\right) , \notag \end{eqnarray}% \noindent where $X+iY=e^{i\varphi }\cot \theta /2$. This yields \begin{equation*} -\nu _{,r}|_{x.y}=\frac{\lambda _{z}(X^{2}+Y^{2}-1)+2\lambda _{x}X+2\lambda _{y}Y}{X^{2}+Y^{2}+1}=\lambda _{z}(r)\cos \theta +\lambda _{x}(r)\sin \theta \cos \varphi +\lambda _{y}(r)\sin \theta \sin \varphi , \end{equation*}% \noindent where we have defined: \begin{eqnarray} \lambda _{z}(r):= \frac{A^{\prime }}{A}, \qquad \lambda _{x}(r):= \left( \frac{2B_{1}}{A}\right) ^{\prime }A, \qquad \lambda _{y}(r):= \left( \frac{2B_{2}}{A}\right) ^{\prime }A. \end{eqnarray}% With this choice of coordinates, the local energy density of the dust separates uniquely into a spherical symmetric part, $\varepsilon _{s}$, and and a non-spherical part, $\varepsilon _{ns}$: \begin{equation*} \varepsilon =\varepsilon _{s}(t,R)+\varepsilon _{ns}(t,R,\theta ,\varphi ), \end{equation*}% \noindent where: \begin{eqnarray} \kappa \varepsilon _{s} &=&\frac{2M_{,R}}{R^{2}}, \\ \kappa \varepsilon _{ns} &=&-\frac{R\nu _{,R}}{1+\nu _{,R}R}\cdot \left( \frac{2M}{R^{3}}\right) _{,R}. \end{eqnarray}% We define $M_{,R}=M_{,r}/R_{,r}$. Following the conventions of our previous paper we write define \begin{equation*} M:=m+Z(r), \end{equation*} where $m$ is the gravitational mass of our `star'. \subsection{Exterior Expansion} As a result of the way that the inhomogeneity is introduced in these models, we want the FRW limit to be `natural' , that is for the $O(3)$ orbits to become concentric in this limit; we therefore require $\nu _{r}\sim o(1)$ as $\delta \rightarrow 0$ in the exterior. This follows from the requirement that the whole spacetime should become homogeneous in a smooth fashion in the limit where the mass of our `star' vanishes: $m\rightarrow 0$. Put another way, the introduction of our star is the only thing responsible for making the surfaces of constant curvature non-concentric. We define, as in the previous paper, dimensionless `radial' and time coordinates appropriate for the exterior by \begin{equation*} \tau =H_{0}t,\text{ \ \ }\rho =H_{0}r. \end{equation*} The \emph{exterior limit} is defined by $\delta \rightarrow 0$ with $\tau $ and $\rho $ fixed. In the exterior region we find asymptotic expansions in this limit. According to our prescription, we write \begin{equation*} H_{0}Z(\rho )\sim \frac{1}{2}\Omega _{m}\rho ^{3}+\delta ^{p}z_{1}(r)+o(\delta ^{p}), \end{equation*} and \begin{equation*} H_{0}^{-1}\lambda _{i}\sim \delta ^{s}l_{i}(\rho )+\mathcal{o}(\delta ^{s}). \end{equation*} Since $H_{0}^{-2}\left( \frac{2M}{R^{3}}\right) _{,R}\sim \mathcal{O}(\delta ^{p},\delta )$, we have that: $H_{0}^{-2}\kappa \varepsilon _{ns}\sim \mathcal{O% }(\delta ^{p+s},\delta ^{1+s})$ whereas $H_{0}^{-2}\kappa \varepsilon _{s}\sim \mathcal{O}(\delta ^{p},\delta )$. Thus, the non-spherical perturbation to the energy density is always of subleading order compared to the first order in spherical perturbation. The first-order, non-spherical, metric perturbation appears at $\mathcal{O}(\delta ^{s})$; however, since this is equivalent to a coordinate transform on $(r,\theta ,\varphi )$ and the dilaton field, $\phi $, is homogeneous to leading order in the exterior, this perturbation does not make any corrections to the dilaton conservation equation at $\mathcal{O}(\delta ^{s})$. Thus, both at leading order, and at next-to-leading order, both the energy density and the dilaton field will behave in the same way as in the spherically-symmetric Tolman-Bondi case - with the possible addition of a non-spherically symmetric vacuum perturbation to the dilaton, $\phi $, i.e. $\phi \sim \phi _{s}+\phi _{ns}+o(\delta ^{p})$ where $\phi _{s}$ is the spherically symmetric solution and $\square _{FRW}\phi _{ns}=0$. As in our previous paper, however, we are not especially interested in the exterior solution for $\phi $ beyond zeroth order, just the effect of any background variation in $\phi $ on what is measured on the surface of a local 'star'. \subsection{Interior Expansion} We define dimensionless coordinates for the interior in the same way as we did for the spherically symmetric case: \begin{equation*} T=L_{I}^{-1}(t-t_{0})\text{ and }\xi =R/R_{s}. \end{equation*} We define the \emph{interior limit} to be $\delta \rightarrow 0$ with $T$ and $\xi $ fixed, and perform out interior asymptotic expansions in this limit. To lowest order in the interior region, we write $Z\sim \delta ^{q}R_{s}\mu _{1}$, and $\lambda _{i}:=\delta _{q^{\prime }}R_{s}^{-1}b_{i}$% , where $i=\{x,y,z\}$. The condition that $\kappa \varepsilon >0$ everywhere requires $q^{\prime }\geq q$ and then, to next-to-leading order, the interior expansion of $\phi $ will be the same as it was in the spherically-symmetric Tolman-Bondi case. We can potentially include a non-spherical vacuum component for $\phi $; however, this will be entirely determined by a boundary condition on $R=R_{s}$ and the need that it should vanish for large $R$. To find the leading-order behaviour of the $\phi _{,T}$ we need to know $\phi $ at next-to-leading order. The only new case we need to consider therefore is when $q^{\prime }=q$, i.e. $\kappa \varepsilon _{ns}\sim \kappa \varepsilon _{s}$. In the spherically symmetric case we considered two distinct subclasses of the Tolman-Bondi models: the flat, $k=0 $, Gautreau-Tolman-Bondi spacetimes, \cite{gautreau, kras} and the non-flat, $k\neq 0$, Tolman-Bondi models with a simultaneous initial singularity. In Gautreau-Tolman-Bondi models the initial singularity is non-simultaneous from the point of view of geodesic observers. The latter class is the more realistic, since in the former the world-lines of matter particles stream out of the surface of our star at $R=R_{s}|_{R=R_{s}}$ i.e. $R_{,t}>0$, whereas in the simultaneous big-bang models we can demand that matter particles fall \emph{onto} this surface i.e. $R_{,t}|_{R=R_{s}}<0$. With this choice, and if $R_{s}=2m$, the non-flat models properly describe the embedding of a black hole into an expanding universe, whereas the Gautreau-Tolman-Bondi model technically describes the embedding of a white-hole in the same universe. In this paper we shall, therefore, only give the results explicitly for the non-flat case -- however, we can present a simple procedure to transform our results to the flat Gautreau case. We define \emph{\ } \begin{equation*} \eta =\left( \xi ^{3/2}-3T/2\right) ^{2/3};\text{ }R_{s}\eta =r+\mathcal{O}% (\delta ^{q},\delta ^{2/3}). \end{equation*}% From the exact solutions we find: \begin{equation*} k(\eta )=\delta ^{2/3}k_{0}\left( 1+\delta ^{q}\mu _{1}(\eta )+o\left( \delta ^{q}\right) \right) +\mathcal{O}\left( \delta ^{5/3}\right) , \end{equation*}% where \begin{equation*} k_{0}(\delta T)=\frac{2m}{R_{s}}\left( \frac{\pi }{H_{0}t_{0}+\delta T}% \right) ^{2/3}. \end{equation*}% We can remove the $\mathcal{O}(\delta ^{2/3})$ metric perturbation by a redefinition of the $T$ coordinate, $T\rightarrow T^{\ast }$: \begin{equation*} \sqrt{1-\delta ^{2/3}k_{0}}T^{\ast }=T+\int^{\xi }\frac{\sqrt{\frac{2m}{% R_{s}\xi ^{\prime }}}\left( 1-\sqrt{1-\left( \frac{\delta ^{2/3}\pi \xi ^{\prime }}{H_{0}t_{0}+\delta T}\right) }\right) }{1-\frac{2m}{R_{s}\xi ^{\prime }}}\mathrm{d}\xi ^{\prime }. \end{equation*}% To leading order we see that $T\sim T^{\ast }$. The interior expansion of the metric, for $q^{\prime }=q$, is written: \begin{equation*} \mathrm{d}s_{int}^{2}\sim R_{s}^{2}\left( j_{ab}^{(0)}(\xi )+\delta ^{q}j_{ab}^{(1)s}(\xi ,\chi )+\delta ^{q}j_{ab}^{(1)ns}(\xi ,\chi )+o(\delta ^{q})\right) \mathrm{d}x^{a}\mathrm{d}x^{b}+o(\delta ^{q}). \end{equation*}% where $j_{ab}^{(0)}$ and $j_{ab}^{(1)s}$ are given by: \begin{eqnarray} j_{ab}^{(0)}\mathrm{d}x^{a}\mathrm{d}x^{b} &=&\frac{R_{s}}{2m}\mathrm{d}% T^{\ast 2}-\left( \mathrm{d}\xi +\xi ^{-1/2}\mathrm{d}T^{\ast }\right) ^{2}-\xi ^{2}\{\mathrm{d}\theta ^{2}+\sin ^{2}\theta \mathrm{d}\varphi ^{2}\}, \label{j1eq.2} \\ j_{ab}^{(1)s}\mathrm{d}x^{a}\mathrm{d}x^{b} &=&-\frac{\mu _{1}(\chi )}{\xi ^{1/2}}\mathrm{d}\xi \mathrm{d}T^{\ast }-\frac{\mu _{1}(\chi )}{\xi }\mathrm{% d}T^{\ast 2}. \end{eqnarray}% These are the same as in the spherically symmetric case. The non-spherically symmetric perturbation is given by \begin{eqnarray} j_{ab}^{(1)ns}\mathrm{d}x^{a}\mathrm{d}y^{b} &=&2(b_{z}\cos \theta +b_{x}\cos \varphi \sin \theta +b_{y}\sin \varphi \sin \theta )\xi \mathrm{d}% \eta ^{2} \\ &&-2\xi ^{2}\left[ b_{z}\sin \theta +b_{x}\cos \varphi (1-\cos \theta )+b_{y}\sin \varphi (1-\cos \theta )\right] \mathrm{d}\theta \mathrm{d}\eta \notag \\ &&-2\xi ^{2}(1-\cos \theta )\sin \theta (b_{x}\sin \varphi -b_{y}\cos \varphi )\mathrm{d}\varphi \mathrm{d}\eta . \notag \end{eqnarray}% \noindent The spherically symmetric part of the local energy density, $% \kappa \varepsilon _{s}$ is the same as it was in the Tolman-Bondi cases: \begin{equation*} R_{s}^{2}\kappa \varepsilon _{s}=\delta ^{q}\frac{2m}{R_{s}}\frac{\mu _{1,\xi }% }{\xi ^{3/2}\eta ^{1/2}}. \end{equation*}% The non-spherically symmetric part is: \begin{equation*} R_{s}^{2}\kappa \varepsilon _{ns}=-\delta ^{q}\frac{6m}{R_{s}}\frac{\left( b_{z}\cos \theta +b_{x}\sin \theta \cos \phi +b_{y}\sin \theta \sin \phi \right) }{\xi ^{3/2}\eta ^{1/2}}. \end{equation*}% and to ensure that the energy density is everywhere positive we need $\mu _{,\eta }^{(1)}\geq 3b_{i}$. \section{Extension to quasi-spherical situations} \subsection{Boundary Conditions} We demand the same boundary conditions as before: as the physical radius tends to infinity, $R\rightarrow \infty $, we demand that the dilaton tends to its homogeneous cosmological value: $\phi (R,t)\rightarrow \phi _{c}(t)$. This can be applied to the exterior approximation. In the interior, we demand that the dilaton-flux passing out from the surface of our `star' at $% R=R_{s}$ is, at leading order, parametrised by: \begin{equation} -R_{s}^{2}\left( 1-\frac{2m}{R_{s}}\right) \left. \partial _{\xi }\phi _{0}\right\vert _{\xi =R_{s}}=2mF\left( \bar{\phi}_{0}\right) =\int_{0}^{R_{s}}\mathrm{d}R^{\prime }R^{\prime }{}^{2}B_{,\phi }(\phi _{0}(% \hat{\xi}^{\prime }))\kappa \varepsilon (R^{\prime }), \label{phiflux} \end{equation}% where $\bar{\phi}_{0}=\phi _{0}(R=R_{s})$. The function $F(\phi )$ can be found by solving the dilaton field equations to leading order in the $R<R_{s} $ region. If the interior region is a black-hole ($R_{s}=2m$) then we must have $F(\phi )=0$; otherwise we expect $F(\phi )\sim B_{,\phi }(\phi )$. Without considering the sub-leading order dilaton evolution inside our `star', i.e. at $R<R_{s}$, we cannot rigorously specify any boundary conditions beyond leading order. Despite this, we can guess at a general boundary condition by perturbing eq. (\ref{phiflux}): \begin{equation} -R_{s}^{2}\left( 1-\frac{2m}{R_{s}}\right) \left. \partial _{R}\tilde{\delta}% (\phi )\right\vert _{\xi =R_{s}}=-\left. \tilde{\delta}\left( \sqrt{-g}% g^{RR}\right) \partial _{R}\phi _{0}\right\vert _{R=R_{s}}+2\tilde{\delta}% (M)F\left( \bar{\phi}_{0}\right) +2mF_{,\phi }(\bar{\phi}_{0})\tilde{\delta}% \left( \bar{\phi}_{0}\right) +\mathrm{smaller}\;\mathrm{terms}, \label{pertbdry} \end{equation}% where $\tilde{\delta}(X)$ is the first sub-leading order term in the interior expansion of $X$; $M$ is the total mass contained inside $\xi <R_{s} $ and is found by requiring the conservation of energy; and at $t=t_{0}$ we have $M=m$. Only $\tilde{\delta}\left( \bar{\phi}_{0}\right) $ remains unknown; however, we shall assume it to be the same order as $\tilde{\delta}% (\phi )$ and see that this unknown term is usually suppressed by a factor of $2m/R_{s}$ relative to the other terms in eq. (\ref{pertbdry}). \subsection{Interior Expansion} In the spherically symmetric case we found that $\phi \sim \phi _{I}^{(0)}+\delta ^{q}\phi _{I}^{(1)}+o(\delta ^{q})$. In the non-spherical case, where $q^{\prime }=q$, we relabel $\phi _{I}^{(1)}\rightarrow \phi _{I}^{(1)s}$ and we have additional non-spherical modes: \begin{equation*} \phi \sim \phi _{I}^{(0)}(\xi ,T)+\delta ^{q}\phi _{I}^{(1)s}(\xi ,T)+\delta ^{q}\phi _{I}^{(1)z}(\xi ,T)\cos \theta +\delta ^{q}\phi _{I}^{(1)x}(\xi ,T)\sin \theta \cos \varphi +\delta ^{q}\phi _{I}^{(1)y}(\xi ,T)\sin \theta \sin \varphi +o(\delta ^{q}) \end{equation*}% where: \begin{eqnarray} -\frac{2m}{R_{s}}\left( \xi ^{3/2}\phi _{I,TT}^{(1)i}+\frac{3}{2}\phi _{I,T}^{(1)i}\right) &+&\frac{1}{\eta ^{1/2}}\left( \frac{\xi ^{5/2}}{\eta ^{1/2}}\phi _{I,\eta }^{(1)i}\right) _{,\eta }-\frac{2}{\xi ^{1/2}}\phi _{I}^{(1)i}=\frac{6m}{R_{s}}B_{,\phi }\left( \phi _{I}^{0}\right) \frac{% b_{i}\left( \eta \right) }{\eta ^{1/2}} \label{nsphieqn} \\ &+&\left( \frac{2m}{R_{s}}\right) \frac{1}{\eta ^{1/2}}F\left( \bar{\phi}% _{0}\right) \left[ \left( b_{i}(\eta )\xi \left( \frac{1+\frac{2m}{R_{s}\xi }% }{1-\frac{2m}{R_{s}\xi }}\right) \right) _{,\eta }-2b_{i}(\eta )\right] . \notag \end{eqnarray}% We can solve this order by order in $2m/R_{s},$ and to lowest order we find: \begin{eqnarray} \phi _{I}^{(1)i} &\sim &\frac{2m}{R_{s}}B_{,\phi }\left( \phi _{I}^{0}\right) \xi \int^{\eta }\mathrm{d}\eta ^{\prime }\frac{b_{i}(\eta ^{\prime })}{\xi ^{^{\prime }2}}-\frac{2m}{R_{s}}B_{,\phi }\left( \phi _{I}^{0}\right) \frac{1}{\xi ^{2}}\int_{\xi =1}^{\eta }\mathrm{d}\eta ^{\prime }\xi ^{\prime }b_{i}(\eta ^{\prime }) \\ &+&\frac{2m}{R_{s}}F\left( \bar{\phi}_{0}\right) \frac{1}{\xi ^{2}}\int_{\xi =1}^{\eta }\mathrm{d}\eta ^{\prime }b_{i}(\eta ^{\prime })\xi ^{\prime }+% \frac{C_{i}}{\xi ^{2}}+D_{i}\xi +\mathcal{O}((2m/R_{s})^{2}) \notag \end{eqnarray}% Since we are interested in finding when and where the local time variation of $\phi $ deviates from its cosmological value, we are chiefly concerned with the case $q\leq 1$. The matching condition then requires that we fix $% D_{i}$ so that in the intermediate limit we have $\phi _{I}^{(1)i}\sim \xi ^{n}$ with $n<1$. The value of $C_{i}$ should be set by a boundary condition on $R=R_{s}$. We cannot specify $C_{i}$ exactly without further information about the interior of our `star' in $R<R_{s}$. If we assume that the prescription for the sub-leading order boundary condition given above is correct then we find: \begin{eqnarray} \partial _{\xi }\phi _{I}^{(1)i}|_{\xi =1} &\sim &\frac{2m}{R_{s}}\left. \frac{b_{i}}{\eta ^{1/2}}\right\vert _{\xi =1}F\left( \bar{\phi}_{0}\right) +% \mathcal{O}((2m/R_{s})^{2}) \notag \\ \Rightarrow C_{i} &=&\frac{m}{R_{s}}B_{,\phi }\left( \phi _{I}^{0}\right) \int^{\xi =1}\mathrm{d}\eta ^{\prime }\frac{b_{i}(\eta ^{\prime })}{\xi ^{^{\prime }2}}+\tfrac{1}{2}D \notag \end{eqnarray}% From now onwards we set $C_{i}=0,$ for simplicity; even when this is not correct we do not expect the magnitude of $C_{i}$ or $C_{i,T}$ to be larger than any of the other terms in $\phi _{I}^{(1)i}$ or $\phi _{I,T}^{(1)i}$, respectively. The time-derivative of $\phi _{I}^{(1)i}$ for fixed $R$ is: \begin{equation} \phi _{I,T}^{(1)i}\sim \frac{4m}{R_{s}}B_{,\phi }\left( \phi _{I}^{0}\right) \xi \int^{\eta }\mathrm{d}\xi ^{\prime }\frac{b_{i}(\eta ^{\prime })}{\xi ^{^{\prime }5/2}}+\frac{2m}{R_{s}}B_{,\phi }\left( \phi _{I}^{0}\right) \frac{1}{\xi ^{2}}\int_{\xi =1}^{\eta }\mathrm{d}\xi ^{\prime }\frac{% b_{i}(\eta ^{\prime })}{\xi ^{1/2}}-\frac{2m}{R_{s}}F\left( \bar{\phi}% _{0}\right) \frac{1}{\xi ^{2}}\int_{\xi =1}^{\eta }\mathrm{d}\eta ^{\prime }% \frac{b_{i}(\eta ^{\prime })}{\xi ^{1/2}}+D_{,T}\xi \label{nsphitev} \end{equation}% In the next section we shall discuss what we require of the $b_{i}$ for the matching procedure to be valid. In section VI we will then use the matching conditions to find $D$ and $D_{,T}$. We could also relax the requirement that the leading-order mode in $\phi $ be spherically symmetric. At next-to-leading order these new modes would generate extra terms in $\phi _{I}^{(1)}$. In general, an $l$-pole at leading order becomes an $l+1$-pole at next-to-leading order. The magnitude of the extra time-dependence that is picked up is, however, the same each time. Hence, we restrict ourselves by taking the leading-order mode to be spherically symmetric for the time being. Note also that we can pass from the simultaneous big-bang case, to the spatially flat, `Gautreau', case by setting $k=0$ and making the transform $\eta \rightarrow \chi =\left( \xi ^{3/2}+3T/2\right) ^{2/3}$. This will also mean that $\phi _{I,T}\rightarrow -\phi _{I,T}$. \section{Validity of Approximations} All of the conditions found in Paper I for the matching of the spherically symmetric parts of $\phi $ to be possible still apply here. However, we must now satisfy some extra conditions that come from the requirement that the non-spherical parts should also be matchable. We assume that $b_{i}\left( \eta \right) \propto \eta ^{d_{i}}$ as $\eta \rightarrow \infty $ for some $d_{i}>0$. At order $\delta ^{q}$, the growing mode in the non-spherically symmetric part of the interior approximation will then grow like $\delta ^{q}\eta ^{d_{i}+1}/\xi $. In the intermediate, or matching, region we have that $\eta ,\xi \sim \delta ^{-\alpha }$ for some $\alpha \in (0,1)$. We require $\phi _{I}$ to have a valid asymptotic expansion this region. This implies that there exists some $\alpha \in (0,1)$ such that, for each $i$, we have $\alpha -q/d_{i}>0$. In the exterior we shall write $H_{0}^{-1}\lambda _{i}\sim \delta ^{p_{i}^{\prime }}l_{i}(\rho )$, where $p_{i}^{\prime }>0$ comes from the requirement that the 2-spheres of constant curvature become concentric in the exterior limit. As $\rho \rightarrow 0$ we assume that $l_{i}(\rho )\propto \rho ^{-f_{i}}$. We previously stated that $Z\sim \frac{1}{2}\Omega _{m}\rho ^{3}+\delta ^{p}z_{1}+o(\delta ^{p})$ in the exterior. We assume that as $\rho \rightarrow 0$, we have $z_{1}\propto \rho ^{-m}$. Although we did not explicitly consider the exterior expansion of $\phi $ we can now examine the behaviour of the leading-order non-spherically symmetric mode in the intermediate limit of that exterior expansion. We noted above that there will be no $\mathcal{O}(\delta ^{p_{i}^{\prime }})$ correction resulting from the $l_{i}$. The leading-order mode will therefore either go like $% \max_{i}\left( \delta ^{p+p_{i}^{\prime }}z_{1}(\rho )l_{i}(\rho )\right) $ if $p<1$ or $\max_{i}\left( \delta ^{1+p_{i}^{\prime }}(\rho )l_{i}(\rho )\right) $ otherwise, and $\rho \sim \mathcal{O}(\delta ^{1-\alpha })$ in the intermediate region. For the exterior expansion to be valid in the intermediate region we therefore require \begin{eqnarray} \max_{i}(p_{i}^{\prime } &+&(1-\alpha )(f_{i}+m))>-p\;\mathrm{if}\;p\leq 1, \notag \\ \max_{i}(p_{i}^{\prime } &+&(1-\alpha )f_{i})>-1\;\mathrm{if}\;p\geq 1. \notag \end{eqnarray} These conditions on $\alpha $ are equivalent to the following: there exists $% \alpha $ such that the interior expansion of $R^{2}\kappa \varepsilon _{ns}$ is $o(1)$ as $\delta \rightarrow 0$ for all $0<\alpha ^{\prime }<\alpha $ where $\xi ,T\sim \mathcal{O}(\delta ^{-\alpha })$, and the exterior expansion of $% R^{2}\kappa \varepsilon _{ns}$ is also $o(1)$ as $\delta \rightarrow 0$ for all $0<\alpha ^{\prime \prime }<\alpha $ where $\rho ,\tau -\tau _{0}\sim \mathcal{O}(\delta ^{1-\alpha })$. This suggests that the condition for the matching procedure to work, as far as the spherically non-symmetric modes are concerned, is simply that \begin{equation*} R^{2}\kappa \varepsilon _{ns}\ll 1\;\mathrm{everywhere.} \end{equation*}% We can also rephrase and generalise the conditions for the matching procedure to be possible w.r.t. the spherically symmetric modes (as found in \cite{shawbarrow1}) in a similar fashion: for all $\alpha \in (0,1)$, and keeping $L_{I}^{\alpha }L_{E}^{1-\alpha }(t-t_{0}),L_{I}^{\alpha }L_{E}^{1-\alpha }R$ fixed, we have $\lim_{\delta \rightarrow 0}{R^{2}\kappa \Delta \varepsilon _{s}}=o(1)$ and $\lim_{\delta \rightarrow 0}{2(m+Z)/R}=o(1)$% . We can combine our two conditions by simply replacing $\Delta \varepsilon _{s} $ by $\Delta \varepsilon $ in the above expression. Strictly speaking, since $% \alpha \in (0,1)$ (as opposed to $[0,1)$, $(0,1]$ or $[0,1]$) we can also replace $\Delta \varepsilon $ by just $\varepsilon $ since $R^{2}\kappa \varepsilon _{FRW}$ is small everywhere outside the exterior region. For Szekeres backgrounds the first of these conditions implies the second everywhere outside the interior region. Therefore, the matching procedure is certainly possible to zeroth order, if: \begin{equation*} \forall \alpha \in (0,1):\;\;\lim_{\delta \rightarrow 0}\left( R^{2}\kappa \varepsilon (R,t)\right) =o(1)\;\mathrm{and}\;\;\lim_{\delta \rightarrow 0}\left( M(R,t)/R\right) =o(1)\;\;\mathrm{with}\;\;\{L_{I}^{\alpha }L_{E}^{1-\alpha }(t-t_{0}),L_{I}^{\alpha }L_{E}^{1-\alpha }R\}\;\mathrm{% fixed}, \end{equation*}% \noindent where $M(R,t)$ is the gravitational mass inside the surface $% (t,R)=const$. Equivalently, in \emph{any} intermediate region the background spacetime is asymptotically Minkowski as $\delta \rightarrow 0$: everywhere which is not in either the interior or exterior regions can be considered to be a weak-field perturbation of Minkowski spacetime. The power of our method is that we do \emph{not} require this to be true of the interior and exterior regions. So long as this condition holds in the intermediate region, we can match the zeroth-order approximations in some region and find the circumstances under which condition (\ref{wettcond}) holds by comparing the relative sizes of the derivatives $\phi _{c,t}$ and $\phi _{I,t}^{(1)}$. \section{Matching} We rewrite the expression for the $\phi _{I}^{(1)i}$ in terms of the non-spherical part of local density: \begin{eqnarray} \delta ^{q}\phi _{I}^{(1)ns} &=&\delta ^{q}\left( \phi _{I,t}^{(1)z}\cos \theta +\phi _{I,t}^{(1)x}\sin \theta \cos \varphi +\phi _{I,t}^{(1)y}\sin \theta \sin \varphi \right) \notag \\ &\sim &-\frac{1}{3}B_{,\phi }\left( \phi _{I}^{0}\right) R\int^{r}\mathrm{d}% r^{\prime }R_{,r}\kappa \varepsilon _{ns}(r^{\prime },t)-\frac{R}{R_{s}}\hat{D}% (T,\theta ,\phi ) \notag \\ &&-\frac{1}{3}B_{,\phi }\left( \phi _{I}^{0}\right) \frac{1}{R^{2}}% \int_{R=R_{s}}^{r}\mathrm{d}r^{\prime }R_{,r}R^{3}\kappa \varepsilon _{ns}(r^{\prime },t)-\frac{1}{3}F\left( \bar{\phi}_{0}\right) \frac{1}{R^{2}}% \int_{R=R_{s}}^{r}\mathrm{d}r^{\prime }R_{,r}R^{3}\kappa \varepsilon _{ns}(r^{\prime },t) \notag \end{eqnarray}% \noindent where $\hat{D}(T,\theta ,\phi ):=D_{z}\cos \theta +D_{x}\sin \theta \cos \varphi +D_{y}\sin \theta \sin \varphi $. By examining the dilaton equations of motion in the FRW region, we can see there is a component of the leading-order $(\theta ,\varphi )$-dependent term in the exterior expansion or $\phi $ behaves like \begin{equation*} -\frac{1}{3}B_{,\phi }(\phi _{c})R\int_{\infty }^{r}\mathrm{d}r^{\prime }R_{,r}\kappa \varepsilon _{ns}(r^{\prime },t) \end{equation*}% \noindent for $R\ll H_{0}^{-1}$ and $t$ fixed. Therefore matching requires that we choose $\hat{D}$ such that \begin{eqnarray} \delta ^{q}\phi _{I}^{(1)ns} &=&\delta ^{q}\left( \phi _{I,t}^{(1)z}\cos \theta +\phi _{I,t}^{(1)x}\sin \theta \cos \varphi +\phi _{I,t}^{(1)y}\sin \theta \sin \varphi \right) \\ &\sim &-\frac{1}{3}B_{,\phi }\left( \phi _{I}^{0}\right) R\int_{\infty }^{r}% \mathrm{d}r^{\prime }R_{,r}\kappa \varepsilon _{ns}(r^{\prime },t)+\frac{1}{3}% B_{,\phi }\left( \phi _{I}^{0}\right) \frac{1}{R^{2}}\int_{R=R_{s}}^{r}% \mathrm{d}r^{\prime }R_{,r}R^{3}\kappa \varepsilon _{ns}(r^{\prime },t) \\ &&-\frac{1}{3}F\left( \bar{\phi}_{0}\right) \frac{1}{R^{2}}\int_{R=R_{s}}^{r}% \mathrm{d}r^{\prime }R_{,r}R^{3}\kappa \varepsilon _{ns}(r^{\prime },t). \notag \end{eqnarray}% The interior expansion is now fully specified to order $\mathcal{O}(\delta ^{p})$. We are interested in the behaviour of $\phi _{I,t}$ and we find \begin{eqnarray} \delta ^{q}\phi _{I,t}^{(1)ns} &\sim &\frac{2}{3}B_{,\phi }\left( \phi _{I}^{0}\right) R\int_{\infty }^{r}\mathrm{d}r^{\prime }R_{,r}R_{,t}\frac{% \kappa \varepsilon _{ns}(r^{\prime },t)}{R}+\frac{1}{3}B_{,\phi }\left( \phi _{I}^{0}\right) \frac{1}{R^{2}}\int_{R=R_{s}}^{r}\mathrm{d}r^{\prime }R_{,r}R_{,t}R^{2}\kappa \varepsilon _{ns}(r^{\prime },t) \\ &&-\frac{1}{3}F\left( \bar{\phi}_{0}\right) \frac{1}{R^{2}}\int_{R=R_{s}}^{r}% \mathrm{d}r^{\prime }R_{,r}R_{,t}R^{3}\kappa \varepsilon _{ns}(r^{\prime },t)+% \frac{1}{3}F\left( \bar{\phi}_{0}\right) RR_{,t}\kappa \varepsilon _{ns}(r,t). \notag \end{eqnarray}% This expression is valid whenever $R_{s}\gg 2m$, and the requirements for matching are satisfied. In these cases we expect $F\left( \bar{\phi}% _{0}\right) \approx B_{,\phi }\left( \phi _{I}^{0}\right) +\mathcal{O}% (2m/R_{s})$; so, approximately, we have \begin{equation*} \delta ^{q}\phi _{I,t}^{(1)ns}\sim \frac{2}{3}B_{,\phi }\left( \phi _{c}\right) R\int_{\infty }^{r}\mathrm{d}r^{\prime }R_{,r}R_{,t}\frac{\kappa \varepsilon _{ns}(r^{\prime },t)}{R}+\frac{1}{3}B_{,\phi }\left( \phi _{c}\right) RR_{,t}\kappa \varepsilon _{ns}(r,t). \end{equation*}% In the case, where $R_{s}=2m,$ and our `star' is actually a black-hole, we require $F\left( \bar{\phi}_{0}\right) $ to ensure that the $\phi $ is well-defined as $R\rightarrow 2m$. Even so, in this case, equation (\ref% {nsphitev}) will not be strictly valid, since it was derived under the assumption of $R_{s}\gg 2m$. By inspection of the dilaton evolution equation in the interior, eq. (\ref{nsphieqn}), however, we expect that $\delta ^{q}\phi _{I,t}^{(1)ns}$ near the black-hole horizon to be of similar magnitude to the RHS of eq. (\ref{nsphitev}). Combining the results of this paper with those for the spherically symmetric case we find: \begin{equation*} \phi _{I,t}-\phi _{c,t}\sim B_{,\phi }\left( \phi _{c}\right) \int_{\infty }^{r}\mathrm{d}r^{\prime }R_{,r}R_{,t}\kappa \Delta \varepsilon _{s}(r^{\prime },t)+\frac{2}{3}B_{,\phi }\left( \phi _{c}\right) R\int_{\infty }^{r}\mathrm{% d}r^{\prime }R_{,r}R_{,t}\frac{\kappa \varepsilon _{ns}(r^{\prime },t)}{R}+% \frac{1}{3}B_{,\phi }\left( \phi _{c}\right) RR_{,t}\kappa \varepsilon _{ns}(r,t). \end{equation*}% We require that $|(\phi _{I,t}-\phi _{c,t})/\phi _{c,t}|\ll 1$ for \ref% {wettcond} to hold and so ensure that local observations will detect variations of $\phi $ occurring on cosmological scales. \section{Generalisation: a conjecture} So far, we have found an analytic approximation to the values of $\phi $ and $\phi _{c,t}$ in the interior. More succinctly (although less explicitly) we can say that, to leading order in $\delta $, the values of $\phi $, $\phi _{,t}$ and $\phi _{,r}$ can all be found everywhere outside the exterior region from the approximation: \begin{equation} \phi \approx \phi _{hom}(t)+\phi _{l}(\vec{x},t), \label{phicon} \end{equation}% where $\phi _{l}$ is the solution to: \begin{equation*} \square _{sch}\phi _{l}=B_{,\phi }\kappa \Delta \varepsilon \end{equation*}% with $\Delta \varepsilon =\varepsilon (\vec{x},t)-\varepsilon _{c}(t)$, $\square _{sch}$ is the wave operator in a Schwarzschild background, and $t$ is the proper time of a comoving observer. This is solved w.r.t. the boundary conditions $\phi _{l}\rightarrow 0$ as $R\rightarrow \infty $ (where $R=0$ is the centre of our `star') and the flux out of the `star' is as given by equations (\ref{phiflux}) and (\ref{pertbdry}). The homogeneous term is \begin{equation*} \phi _{hom}(t)=\phi _{c}(t+\Delta t(\vec{x},x)) \end{equation*}% where the \emph{lag}, $\Delta t(\vec{x},t)$, is defined by: \begin{equation*} \vec{\nabla}^{2}\Delta t-\vec{v}^{\ast }\cdot \vec{\nabla}(\vec{v}^{\ast }\cdot \vec{\nabla}\Delta t)-\vec{v}^{\ast }\cdot \vec{\nabla}\Delta t(\vec{% \nabla}\cdot \vec{v}^{\ast })=-\vec{\nabla}\cdot \Delta \vec{v}, \end{equation*}% with $\nabla _{i}=\partial _{i}$, $i=\{1,2,3\}$, and $\Delta v=\vec{v}-H\vec{% x}$, where $\vec{v}$ is the velocity of the dust particles relative to $% R=\Vert \vec{x}\Vert =0$. The velocity $\vec{v}^{\ast }$ has the following properties: $\vec{v}^{\ast }=\vec{v}$ in some region that includes all the interior and excludes all of the exterior; $\vec{v}^{\ast }=\Delta \vec{v}$ everywhere else. In a general sense, the interior and exterior are two disjoint regions of total spacetimes where general-relativistic effects are non-negligible at leading order (e.g. such as when $\Vert \vec{v}\Vert \approx 1$). The interior region should be closed, and in the exterior region $\Vert \Delta \vec{v}\Vert $ is small. So, $\vec{v}^{\ast }$ should be defined in such a way that it respects all the symmetries of the spacetime and so that $\Vert \vec{v}^{ast}\Vert \ll 1$ everywhere outside the interior region. This is required to ensure that $\Delta t$, as defined above, is finite. It can be seen to come out of the matching procedure. When the background spacetime satisfies the conditions given below, the precise way in which $\vec{v}^{\ast }$ is defined does not effect the leading order behavior of $\Delta t$. For boundary conditions, we must require the flux out of $\Delta t$ out of the `star' to vanish, and require $\Delta t\rightarrow 0$ as $\Delta v\rightarrow 0$, i.e. as $R\rightarrow \infty $. This is the natural generalisation of what has been seen in the Szekeres-Szafron backgrounds $\vec{v}^{i}=R_{,t}(R,t)\delta _{R}^{i}$. In these cases the equation is just an ordinary differential equation in $R$ with solution: \begin{equation*} \Delta t=\int_{R}^{A}\mathrm{d}R^{\prime }\frac{(R_{,t}(R^{\prime },t)-HR^{\prime }+R_{,t}(R_{s},t)+HR_{s})}{1-R_{,t}^{2}(R^{\prime },t)}% +\int_{A}^{\infty }\mathrm{d}R^{\prime }\frac{(R_{,t}(R^{\prime },t)-HR^{\prime }+R_{,t}(R_{s},t)+HR_{s})}{1-(R_{,t}(R^{\prime },t)-HR)^{2}}, \end{equation*}% where $A$ is some arbitrary value of $A$ in the intermediate region, and each $A$ represents a particular choice of definition for $\vec{v}^{\ast }$. This expression is only valid to leading order in the interior and intermediate regions. To this order all choices for $A$ are equivalent. Near $R=R_{s}$, to leading order in $\delta =L_{I}/L_{E}$, this ensures that $% \mathrm{d}(t+\Delta t)\sim \mathrm{d}v$, where $v=t_{sch}+R+2m\ln (R/2m-1)$ is the advanced time coordinate and $t_{sch}$ is the standard, curvature-defined, Schwarzschild time-coordinate. The solution for $\phi _{hom}$ is then, to leading order in $\delta $, just the particular one given by Jacobson in \cite{jacobson}. We have assumed that the generalisations of the Szekeres-Szafron result for $\phi $ hold. We have only proved that this assumption holds for the subset of Szekeres-Szafron spacetimes for which the matching procedure works. Nonetheless, based on this analysis, we conjecture that \ref{phicon} provides a good numerical approximation to the value of $\phi $, and by differentiating once, to $\phi _{,t}$ and $\partial _{i}\phi $, $i=\{1,2,3\}$, near the surface of our `star', for any dust plus $\Lambda $ spacetime that can be everywhere considered to be a weak-field perturbation of either Schwarzschild, Minkowski, or FRW spacetime; that is, \begin{equation*} R^{2}\kappa \Delta \varepsilon (R,t)\ll 1,\qquad 2(M(R,t)-m)/R\ll 1 \end{equation*}% where $M(R,t)$ is the gravitational mass contained inside the surface $% (R,t)=const$. One could seek to motivate our conjecture as some sort of analytical continuation from the Szekeres-Szafron spacetimes to more general backgrounds, but such arguments would, we believe, be hard to frame in any rigorous context and are beyond the scope of the analysis in this paper. \section{Discussion} In this paper we have extended the analysis of \cite{shawbarrow1} to include a class of dust-filled spacetimes without any symmetries provided by the Szekeres-Szafron metrics. Again, we have used the method of matched asymptotic expansions to link the evolution of the dilaton field, $\phi $, in an approximately Schwarzschild region of spacetime to its evolution in the cosmological background. By these methods, we have provided a rigorous construction of what has been simply assumed about the matching procedures in earlier studies \cite{early}. We have also analysed, more fully, the conditions that we need the background spacetime to satisfy for the matching procedure to be valid, and we have interpreted these conditions in terms of their requirements on the local energy density. Finally, we have conjectured a generalisation of our result to more general spacetime backgrounds than those considered here. By combining the results found here with those of the previous paper, we conclude that, in the class of quasi-spherical Szekeres spacetimes in which the matching procedure is valid, the local time variation of the dilaton field will track its cosmological value whenever: \begin{equation} \left\vert \frac{B_{,\phi }\left( \phi _{c}\right) \int_{\infty }^{r}\mathrm{% d}r^{\prime }R_{,r}\kappa \Delta (R_{,t}\varepsilon _{s}(r^{\prime },t))+\frac{2% }{3}B_{,\phi }\left( \phi _{c}\right) R\int_{\infty }^{r}\mathrm{d}r^{\prime }R_{,r}R_{,t}\frac{\kappa \varepsilon _{ns}(r^{\prime },t)}{R}+\frac{1}{3}% B_{,\phi }\left( \phi _{c}\right) RR_{,t}\kappa \varepsilon _{ns}(r,t)}{\dot{% \phi}_{c}}\right\vert \ll 1. \label{condition1} \end{equation}% When the cosmological evolution of $\phi $ is dominated by its matter coupling: $\dot{\phi}_{c}\sim \mathcal{O}(B_{,\phi }H_{0}^{-1}\kappa \varepsilon _{c}),$ this condition is equivalent to: \begin{equation*} \left\vert H_{0}\int_{\infty }^{r}\mathrm{d}r^{\prime }R_{,r}\frac{\Delta (R_{,t}\varepsilon _{s}(r^{\prime },t))}{\varepsilon _{c}(t)}+\frac{2}{3}% H_{0}R\int_{\infty }^{r}\mathrm{d}r^{\prime }R_{,r}R_{,t}R^{-1}\frac{% \varepsilon _{ns}(r^{\prime },t)}{\varepsilon _{c}(t)}+\frac{1}{3}H_{0}R_{,t}\frac{% \varepsilon _{ns}(r,t)}{\varepsilon _{c}(t)}\right\vert \ll 1. \end{equation*} In the other extreme, when the potential term dominates the cosmic dilaton evolution, the left-hand side of the above condition is further suppressed by a factor of $B_{,\phi }(\phi _{c})/V_{,\phi }(\phi _{c})\ll 1$. As in our previous paper, \cite{shawbarrow1}, we can see that for a given evolution of the background matter density, condition (\ref{wettcond}) is more likely to hold (or will hold more strongly) when $\left\vert B_{,\phi }(\phi _{c})\kappa \varepsilon _{c}/V_{,\phi }(\phi _{c})\right\vert \ll 1$. We reiterate our previous statement that: \emph{domination by the potential term in the cosmic evolution of the dilaton has a homogenising effect on the time variation of} $\phi $. The non-spherically symmetric parts of energy density enter into the expression differently. The magnitude of the terms on the left-hand side of eq. (\ref{condition1}) is, as in the spherically symmetric case, still $% \left\langle H_{0}R\Delta R_{,t}\varepsilon /\varepsilon \right\rangle (R,t)$ where $\left\langle \cdot \right\rangle (R,t)$ represents some `average' over the region outside the surface $(R,t)=const$. We should note that, given the condition on $\kappa \varepsilon $ that has been required for matching, the leading-order contribution to $\kappa \varepsilon _{ns}$ is everywhere of dipole form and this is responsible for the special form of the average over the non-spherically symmetric terms. We can also see that, as a result of form of eqn. (\ref{condition1}), peaks in $\kappa \varepsilon _{ns}$ that occur outside of the interior region will, in the interior, produce a weaker contribution to the left-hand side of eqn. (\ref{condition1}% ) than a peak of similar amplitude in a spherically symmetric energy density $\kappa \varepsilon _{s}$. This behaviour would continue if we were also to account for higher multipole terms in $\kappa \varepsilon _{ns}$. The higher the multipole, the more `massive' the mode, and the faster it dissipates. If we are interested in finding a sufficient condition (as opposed to a necessary and sufficient one) for (\ref{wettcond}) to hold locally, then in most circumstances we will be justified in averaging over the non-spherically symmetric modes in the same way as we average over the spherically symmetric ones. In most cases, this will over-estimate rather than under-estimate the magnitude of the left-hand side of our condition, (% \ref{condition1}). This reasoning leads us to the statement that for $\dot{% \phi}(\mathbf{x},t)\approx \dot{\phi}_{c}(t)$ to hold locally it is sufficient that: \begin{equation} \mathcal{I}:=\int_{\gamma (R)}\mathrm{d}lH_{0}R^{\prime }\frac{\Delta (v\varepsilon )}{\varepsilon _{c}}\ll 1 \label{suff} \end{equation}% \noindent where $\mathrm{d}l:=\mathrm{d}rR_{,r}$, $v=R_{,t}$ is the velocity of the dust particles, $\lim_{R\rightarrow \infty }v=H_{0}R$. We make the same generalisation that we did in Paper I by taking $\gamma (R)$ to run from $R$ to spatial infinity along a past, radially-directed light-ray. In this way, we incorporate the limitations imposed by causality. We should also assume that the above expression includes some sort of average over angular directions; to be safe we could replace $\varepsilon $ by its maximum value for fixed $R$ and $t$. This sufficient condition, (\ref{suff}), is precisely the generalised condition proposed in our first paper on this issue. The inclusion of deviations from spherical symmetry, therefore, has little effect of the qualitative nature of the conclusions that were found in \cite{shawbarrow1}. If anything, we have seen that the non-spherical modes dissipate faster and, as a result, will produce smaller than otherwise expected deviations in the local time derivative of $\phi $ from the cosmological ones. On Earth we should expect, as before, that the leading-order deviation of $% \dot{\phi}$ from $\dot{\phi}_{c}$ is produced by the galaxy cluster in which we sit, and that for a dilaton evolution that is dominated by its coupling to matter, this effect gives $\mathcal{I}\approx 6\times 10^{-3}\Omega _{m}^{-1}h\ll 1$, where $\Omega _{m}\approx 0.27$ and $h\approx 0.71$. If the cosmic dilaton evolution is potential dominated then $\mathcal{I}$ is even smaller. We conclude, as before, that irrespective of the value of the dilaton-to-matter coupling, and what dominates the cosmic dilaton evolution, that \begin{equation*} \dot{\phi}(\mathbf{x},t)\approx \dot{\phi}_{c}(t) \end{equation*}% will hold in the solar system in general, and on Earth in particular, to a precision determined by our calculable constant $\mathcal{I}$. We also conclude, as before, that whenever $\mathcal{I}\ll 1$ near the horizon of a black hole, there will be no significant gravitational memory effect for physically reasonable values of the parameters \cite{memory, jacobson}. Our result relies on one major assumption: the physically realistic condition that the scalar field should be weakly coupled to matter and gravity -- in effect, the variations of 'constants' on large scales must occur more slowly than the universe is expanding and so their dynamics have a negligible back-reaction on the cosmological background metric. In this paper we have removed the previous condition of spherical symmetry at least in as far as the spacetime background is well described by Szekeres-Szafron solution. We have therefore extended the domain of applicability our general proof: that \emph{terrestrial} and \emph{solar system} based observations can legitimately be used to constrain the \emph{cosmological} time variation of supposed `constants' of Nature and other light scalar fields. \begin{acknowledgments} \bigskip We thank Tim Clifton and Peter D'Eath for discussions. D. Shaw is supported by a PPARC studentship. \end{acknowledgments}
Title: Resolving the compact dusty discs around binary post-AGB stars using N-band interferometry
Abstract: We present the first mid-IR long baseline interferometric observations of the circumstellar matter around binary post-AGB stars. Two objects, SX Cen and HD 52961, were observed using the VLTI/MIDI instrument during Science Demonstration Time. Both objects are known binaries for which a stable circumbinary disc is proposed to explain the SED characteristics. This is corroborated by our N-band spectrum showing a crystallinity fraction of more than 50 % for both objects, pointing to a stable environment where dust processing can occur. Surprisingly, the dust surrounding SX Cen is not resolved in the interferometric observations providing an upper limit of 11 mas (or 18 AU at the distance of this object) on the diameter of the dust emission. This confirms the very compact nature of its circumstellar environment. The dust emission around HD 52961 originates from a very small but resolved region, estimated to be ~ 35 mas at 8 micron and ~ 55 mas at 13 micron. These results confirm the disc interpretation of the SED of both stars. In HD 52961, the dust is not homogeneous in its chemical composition: the crystallinity is clearly concentrated in the hotter inner region. Whether this is a result of the formation process of the disc, or due to annealing during the long storage time in the disc is not clear.
https://export.arxiv.org/pdf/astro-ph/0601169
\title{ Resolving the compact dusty discs around binary post-AGB stars using N-band interferometry \thanks{Based on observations made with the Very Large Telescope Interferometer of the European Southern Observatory (program id 073.A-9002(A)), the 1.2\,m Flemish Mercator telescope at Roque de los Muchachos, Spain and the 1.2\,m Swiss Euler telescope at La Silla, Chile } } \author{ P. Deroo\inst{1} \and H. Van Winckel\inst{1} \and M. Min\inst{2} \and L.B.F.M. Waters\inst{1,2} \and T. Verhoelst\inst{1} \and W. Jaffe\inst{3} \and S. Morel\inst{4} \and F. Paresce\inst{4} \and A. Richichi\inst{4} \and P. Stee\inst{5} \and M. Wittkowski\inst{4} } \institute{Instituut voor Sterrenkunde, K.U. Leuven, Celestijnenlaan 200B, B-3001 Leuven, Belgium \and Astronomical Institute ``Anton Pannekoek'', University of Amsterdam, Kruislaan 403, 1098 SJ Amsterdam, the Netherlands \and Leiden Observatory, P.B. 9513, Leiden 2300 RA, the Netherlands \and European Southern Observatory, Karl-Scharzschild-Strasse 2, 85748 Garching, Germany \and Observatoire de la C\^ote d' Azur, CNRS-UMR 6203, Avenue Copernic, 06130 Grasse, France} \date{Received <date> / Accepted <date>} \offprints{ P. Deroo \\ \email{Pieter.Deroo@ster.kuleuven.be} } \abstract{ We present the first mid-IR long baseline interferometric observations of the circumstellar matter around binary post-AGB stars. Two objects, \object{SX Cen} and \object{HD 52961}, were observed using the VLTI/MIDI instrument during Science Demonstration Time. Both objects are known binaries for which a stable circumbinary disc is proposed to explain the SED characteristics. This is corroborated by our N-band spectrum showing a crystallinity fraction of more than 50\,\% for both objects, pointing to a stable environment where dust processing can occur. Surprisingly, the dust surrounding SX\,Cen is not resolved in the interferometric observations providing an upper limit of 11 mas (or 18 AU at the distance of this object) on the diameter of the dust emission. This confirms the very compact nature of its circumstellar environment. The dust emission around \object{HD\,52961} originates from a very small but resolved region, estimated to be $\sim$35 mas at 8\,$\mu$m and $\sim$55 mas at 13\,$\mu$m. These results confirm the disc interpretation of the SED of both stars. In \object{HD\,52961}, the dust is not homogeneous in its chemical composition: the crystallinity is clearly concentrated in the hotter inner region. Whether this is a result of the formation process of the disc, or due to annealing during the long storage time in the disc is not clear. \keywords{Stars: circumstellar matter -- Stars: AGB and post-AGB -- Stars: individual: \object{HD\,52961} and SX\,Cen -- Techniques: interferometric -- Infrared: stars } } \titlerunning{ Discs around evolved objects: \object{HD\,52961} and SX\,Cen } \authorrunning{P. Deroo et al.} \section{Introduction} The fast stellar evolution connecting the Asymptotic Giant Branch (AGB) to the Planetary Nebulae (PNe) phase is still poorly understood \citep[e.g.][]{Vanwinckel_2003}. Many detailed studies of individual transition objects (post-AGB stars) exist, but it is not clear how these objects are related by evolutionary channels. Moreover, there is general agreement that binary interactions must play a significant role in many well studied sources. Binarity is for instance invoked in the physical models to understand the observational characteristics of some spectacular geometries observed in PNe. More recently, also the geometries and kinematical structures around resolved post-AGB stars might be linked to binarity \citep[][ and references therein]{Balick_2002}. Since many uncertainties remain in our understanding of the final evolution of single stars, it is no surprise that this is even more the case when the star is a member of a binary system. Direct detection of the binary nature of central stars of resolved nebulae is often difficult due to the high obscuration. Moreover, in crossing the HR-diagram, the stars must pass the pop\,II Cepheid instability strip, in which pulsational instabilities occur. This makes radial velocity variations not a straightforward signature of orbital variations. In the sample of optically bright post-AGB stars, binaries are being detected, however, and for an overview we refer to \citet[][ and references therein]{Vanwinckel_2003}. One of the important observational characteristics of those binaries is the shape of their SED. They show a dust-excess starting near sublimation temperature, irrespective of the effective temperature of the central object and this despite the lack of a current dusty mass loss \citep[][ and references therein]{Deruyter_2005, Deruyter_2006}. Moreover, when available, the long wavelength fluxes show a black-body slope which indicates the presence of a component of large mm-sized grains. It is argued in \cite{Deruyter_2006} that in all the investigated objects, gravitationally bound dust is present, likely in a Keplerian disc. Note that only for the most famous example, the \object{Red Rectangle}, this dust emission is resolved and it shows a clear disc structure, both in the near-IR and in the visible \citep{Menshchikov_2002,Cohen_2004}. Moreover, the gaseous component was spatially resolved using mm-interferometry \citep{Bujarrabal_2003,Bujarrabal_2005}. The latter observations clearly demonstrated the Keplerian rotation of the disc. In all other cases, the presence of a disc is postulated. More detailed studies of individual cases can be found: examples are \object{89\,Her} \citep{Waters_1993}, \object{HR\,4049} \citep{Waelkens_1991a, Dominik_2003} and \object{IRAS08544-4431} \citep{Maas_2003}. Given the orbits detected so far, one of the conclusions is that it is clear that most binaries cannot have evolved along single star evolutionary tracks. The high spatial resolution of the mid-IR instrument MIDI mounted on the VLTI interferometer of ESO makes this an ideal instrument to probe the circumstellar material around these binaries for two reasons: (i) the discs are likely compact so high spatial resolution measurements are needed to resolve the discs and (ii) the discs are shown to emit a significant part of their total luminosity in the N-band. We therefore carefully selected 2 binary post-AGB stars for which there is significant indirect evidence for the presence of a stable circumstellar dust reservoir. The data presented in this contribution are taken during Science Demonstration Time to illustrate the potential of MIDI coupled to the VLTI to study the compact circumstellar environment suspected in those evolved stars. In Sect.~\ref{sect:global_characteristics}, we introduce both objects and refine the orbital parameters published in our previous papers. The observational log is presented in Sect.~\ref{sect:observations} while the reduction of the interferometric dispersed fringes is reported in Sect.~\ref{sect:reduction}. We discuss our findings in Sect.~\ref{sect:discussion} and come to our conclusions in Sect.~\ref{sect:conclusions}. \section{\object{HD\,52961} and \object{SX\,Cen}: global characteristics}\label{sect:global_characteristics} \object{SX\,Cen} and \object{HD\,52961}, having both a spectral type F-G \citep{Kholopov_1999,Shenton_1994,Waelkens_1991}, are located in the pop\,II instability strip with \object{SX\,Cen} known as a very regular RV\,Tauri star with a period of 32.9 days, while \object{HD\,52961} shows a photometric periodicity of 72 days. They are members of the chemically anomalous post-AGB stars for which the photospheric abundances of the different elements are closely linked to their condensation temperature \citep{Waelkens_1991, Vanwinckel_1992, Vanwinckel_1995, Maas_2002}. Members of this class show higher photospheric abundances for chemical elements with a lower condensation temperature. In fact, \object{HD\,52961} is one of the most extreme examples of this class of objects. It is a highly metal-poor object \citep[\hbox{[Fe/H] = -4.8, }][]{Waelkens_1991} which has more zinc than iron in absolute (!) number \citep[\hbox{[Zn/Fe] = +3.1, }][]{Vanwinckel_1992}. There is general agreement that this abundance pattern is caused by a chemical fractionation process caused by dust formation in the circumstellar environment. After decoupling of the gas and dust, reaccretion of the gas causes the observed abundance pattern. \cite{Waters_1992} proposed a scenario in which the circumstellar dust is trapped in a stable disc. The occurrence of such a disc likely implies binarity for post-AGB stars. Indeed, radial velocity measurements proved that all the extremely depleted objects are binaries \citep{Vanwinckel_1995}. In the following we refine our previously published Spectral Energy Distribution (SED) as well as the orbital elements of both objects. \subsection{Spectral energy distribution}\label{sect:SED} The SED of \object{SX\,Cen} is discussed in the literature \citep{Goldsmith_1987,Shenton_1994,Maas_2002} where a total reddening of \hbox{E(B-V) = 0.3\,$\pm$\,0.1} is found. They find a broad infrared excess starting already at K which, in combination with the confirmed binarity, is interpreted in \cite{Maas_2002} as a signature of a dusty disc, not an outflow. The SED is reproduced in the top panel of Fig. \ref{fig:sedBOTH} in which the ISO/SWS spectrum is overplotted. This spectrum shows a broad silicate emission feature at 10 $\mu$m, inside the MIDI wavelength range. Due to the low flux levels at longer wavelengths, we cannot be conclusive about the origin of the feature around 18 $\mu$m. This is possibly an artifact, although it is present in both scans. The distance of \object{SX\,Cen} can be estimated comparing the intrinsic luminosity with the integrated flux of the scaled Kurucz model. The luminosity of \object{SX\,Cen} is estimated using the period-luminosity relation derived for the LMC RV\,Tauri variables \citep{Alcock_1998} to be about \hbox{600\,$\pm$\,400\,L$_{\sun}$}. The large uncertainty is a relic of the scatter in the P-L relation. This provides a rough distance estimate for \object{SX\,Cen} of \hbox{1.6\,$\pm$\,0.5\,kpc}. For \object{HD\,52961}, we constructed a SED using IUE data (0.115 $\mu$m -- 0.320 $\mu$m), Geneva optical photometry, near-IR JHKLM photometry \citep{Bogaert_1994} and far-IR IRAS photometry. In addition, one SCUBA \citep{Holland_1999} observation was made to obtain a continuum measurement at 850 $\mu$m, providing F$_{850\mu\rm{m}} = 2.8 \pm 1.9$\,mJy with Mars as flux calibrator. As the object suffers photometric variations over time, we constructed the SED for photometric maximum only. The colour excess due to interstellar and circumstellar extinction was estimated by searching the best correspondence between the appropriate Kurucz model and the dereddened SED in the optical and UV. The SED was dereddened using the average interstellar extinction law of \cite{Savage_1979} and the Kurucz model was chosen according to the stellar parameters given in \cite{Waelkens_1991}, i.e. T$_{\rm{eff}} = 6000$\,K, log(g)$=0.5$ and [Fe/H]$=-4.5$. The result is shown in Fig.~\ref{fig:sedBOTH}, where the Geneva photometry at photometric minimum is overplotted using grey crosses. While the photometry and the Kurucz model are consistent in the UV and optical, a clear IR excess due to dust is observed at longer wavelengths. This excess distribution, in combination with the confirmed binarity \citep[][ and refinements in Sect. \ref{sect:orbital_elements}]{Vanwinckel_1995} and the lack of a current dusty mass loss, is also interpreted as evidence for a dusty disc instead of an outflow. The luminosity of \object{HD\,52961} is estimated as \hbox{1900\,$\pm$\,1300\,L$_{\sun}$} using the same period-luminisoty relation as for \object{SX\,Cen}, providing a distance of \hbox{1.4\,$\pm$\,0.5\,kpc}. In the MIDI wavelength range \hbox{(8 -- 13 $\mu$m)}, the amount of flux emitted by the stellar photosphere with respect to the total flux is only 1\,\% for \object{SX\,Cen} and 5\,\% for \object{HD\,52961}. In addition, both objects show a clear silicate resonance in emission in the N-band (see the ISO/SWS spectrum in the top panel of Fig. \ref{fig:sedBOTH} and the MIDI spectra in Fig. \ref{fig:totalspectrum}). Therefore, the MIDI instrument, providing spectrally dispersed visibilities over the N-band, is ideally suited to probe the circumstellar geometries of the dust around both objects. \subsection{Orbital elements}\label{sect:orbital_elements} We refined the orbital elements which were published already in \cite{Vanwinckel_1999} and \cite{Maas_2002} for \object{HD\,52961} and \object{SX\,Cen} respectively. Our accumulation of data is now such that we covered close to 3 (\object{HD\,52961}) and 2 (\object{SX\,Cen}) orbital cycles. The heliocentric radial velocity data folded on the orbital periods are given in Fig.~\ref{fig:orbits} and the orbital elements are listed in Table~\ref{tab:orbitalelements}. The data sampling of \object{HD\,52961} is not very extensive and in the residuals no clear modulation on the pulsational period is found. For the RV\,Tauri star \object{SX\,Cen}, the pulsational amplitude in radial velocity is significant. After pre-whitening of the orbital solution, the pulsational period of 16.46\,d is clearly recovered (Fig.~\ref{fig:pulsationSXCEN}). We cleaned the original data with a harmonic least square fit of the 16.46 days pulsation period and three harmonics. The variance reduction of the pulsation model is 81\%. After cleaning the original data with this pulsation model, we redetermined the orbital elements. The eccentric orbit was found to be significant according to the classical Lucy and Sweeney test \citep{Lucy_1971}. \begin{table} \caption{The orbital elements of the program stars. All symbols have their usual meaning. The number of measurements (N) and the number of covered orbital cycles in our monitoring program are also given.}\label{tab:orbitalelements} \begin{tabular}{llll} \hline \hline & unit & \object{HD\,52961} & \object{SX\,Cen} \\ \hline P & days & 1297 $\pm$ 7 & 592 $\pm$ 13 \\ T$_{o}$ & JD & 2448591 $\pm$ 38 & 2452107 $\pm$ 10 \\ K & km\,s$^{-1}$ & 13.3 $\pm$ 0.9 & 22.9 $\pm$ 0.5 \\ $\gamma$ & km\,s$^{-1}$ & 7.4 $\pm$ 0.5 & 19.1 $\pm$ 0.4 \\ e & & 0.22 $\pm$ 0.05 & 0.16 $\pm$ 0.02 \\ $a\,\sin i$ & AU & 1.54 & 1.23 \\ f(M) & M$_{\odot}$ & 0.29 & 0.70 \\ N & \# & 31 & 78 \\ cycles covered & & 2.9 & 2.0\\ \hline \end{tabular} \end{table} As shown in our previous papers, both objects show a long term trend in their photometric light-curve which is due to variable circumstellar reddening in the line of sight towards the object. For \object{SX\,Cen} this long term trend is periodic with a period of 615 days \citep{Oconnell_1933,Voute_1940}, very close to the orbital period. For \object{HD\,52961} it was not very clear whether the subtle effect is periodic or not. If the circumstellar material is indeed mainly stored in a disc around the objects and assuming this disc is located in the orbital plane, the inclination of the disc cannot be very small. Assuming an inclination varying between edge-on (i\,=\,90$^{o}$) up to 60$^{o}$, and a mass of the evolved component of 0.6 M$_{\odot}$, the mass of the companion varies between 0.8\,--\,1.1 M$_{\odot}$ for \object{HD\,52961} and 1.4\,--\,1.9 M$_{\odot}$ for \object{SX\,Cen}. The unseen companion is probably an unevolved main sequence star, since the lack of an UV excess and of any sign of symbiotic activity make the presence of a massive compact object very unlikely. Moreover, the orbital characteristics in period and eccentricity make it very unlikely that the companion is a post red giant as well \citep{Vanwinckel_2003}. Neither stars are filling their Roche Lobe now, but in both cases it is clear that the actual orbit is too small to accommodate a full grown AGB star. The stars must have suffered an evolutionary phase with severe binary interaction when at giant dimensions. \section{Observations}\label{sect:observations} The VLTI/MIDI interferometer \citep{Leinert_2003} was used to combine the light coming from the UT2 and UT3 telescopes. The observations of the targets, \object{SX\,Cen} and \object{HD\,52961}, were performed in three nights of Science Demonstration Time in February at a projected baseline in the range of 40 to 50 meters. A detailed log of the observations of the science targets is presented in Table \ref{tab:log}. The following observing sequence was carried out, according to the standard procedures for MIDI, and repeated for target stars and calibrators. First, acquisiton images are obtained by both telescopes independently (i.e. without beam combiner and prism) to ensure overlap of the beams, which is required for interferometric combination. Then, the MIDI beam combiner, the slit and the prism are inserted. This produces two spectrally dispersed interferometric outputs of opposite phase. The zero optical path difference (OPD) is searched for by scanning a range of a few millimeters around the expected value. When found, MIDI uses its piezo-driven mirrors to keep the fringe pattern at a fixed position within a $\approx 200$\,$\mu$m scan length, while the VLTI delay lines compensate for the drift in OPD position due to sidereal motion and for the slow component of atmospheric piston. Fringes are integrated for about 1-3 minutes. Finally, photometric data are recorded using one telescope at a time, with the same optical set-up but using chopping to subtract sky and background. To correct for optical imperfections and atmospheric turbulence, a calibrator of known diameter is measured as well. The time-lag between the measurement of this calibrator and the science object is about 30 min. However, considering the present accuracy per single visibility measurement of about 10\,\%, we can also use calibrators observed in the same mode one or two hours earlier or later \citep[see e.g.][]{Leinert_2004}. A list of the calibrator observations is given in Table \ref{tab:log}. \begin{table*} \caption{A summary of the observations with the MIDI instrument of \object{SX\,Cen} and \object{HD\,52961}. For each science target, the calibrators used to calibrate the visibility are given (the flux calibrators are given in Table \ref{tab:photcal}). The angular diameter in the Limb Darkened Disc approximation is obtained from \cite{Verhoelst_2005} (cf. http://www.ster.kuleuven.ac.be/$\sim$tijl/MIDI\_calibration/mcc.txt). The reported flux for the calibrator sources is the IRAS 12.5\,$\mu$m flux. Nomenclature: UT = Universal Time, PB = Projected Baseline and PA = Projected baseline Angle} \label{tab:log} \begin{center} \begin{tabular}{l c c c c c c c c c c} \hline \hline \multicolumn{1}{c}{night} & \multicolumn{1}{c}{science} & \multicolumn{1}{c}{UT} & \multicolumn{1}{c}{PB} & \multicolumn{1}{c}{PA} & \multicolumn{1}{c}{airmass} & \multicolumn{1}{c}{calibrator} & \multicolumn{1}{c}{UT} & \multicolumn{1}{c}{spectral} & \multicolumn{1}{c}{diameter} & \multicolumn{1}{c}{flux} \\ \multicolumn{1}{c}{yyyy/mm/dd} & \multicolumn{1}{c}{target} & \multicolumn{1}{c}{hh mm ss} & \multicolumn{1}{c}{(m)} & \multicolumn{1}{c}{(\degr )} & \multicolumn{1}{c}{}& \multicolumn{1}{c}{target} & \multicolumn{1}{c}{hh mm ss} & \multicolumn{1}{c}{type} & \multicolumn{1}{c}{(mas)} & \multicolumn{1}{c}{(Jy)} \\ \hline 2004/02/09 & \object{HD\,52961} & 02 20 02 & 39.7 & 45 & 1.2 & HD\,49161 & 02 45 52 & K4\:III & 2.44 $\pm$ 0.01 & 10.35\\ & \object{SX\,Cen} & 07 37 52 & 44.6 & 41 & 1.1 & HD\,67582 & 06 24 15 & K3\:III & 2.30 $\pm$ 0.01 & 9.33 \\ & & & & & & HD\,67582 & 07 13 39 & K3\:III & 2.30 $\pm$ 0.01 & 9.33 \\ & & & & & & HD\,107446 & 08 08 53 & K3.5\:III & 4.43 $\pm$ 0.02 & 32.42 \\ 2004/02/10 & \object{HD\,52961} & 04 13 04 & 46.1 & 46 & 1.4 & HD\,67582 & 03 46 12 & K3\:III & 2.30 $\pm$ 0.01 & 9.33 \\ & \object{SX\,Cen} & 07 37 52 & 44.6 & 41 & 1.1 & HD\,107446 & 08 16 44 & K3.5\:III & 4.43 $\pm$ 0.02 & 32.42 \\ 2004/02/11 & \object{SX\,Cen} & 08 03 25 & 43.6 & 46 & 1.1 & HD\,120404 & 08 27 38 & K7\,III & 2.96 $\pm$ 0.02 & 13.28 \\ \hline \end{tabular} \end{center} \end{table*} \section{Reduction}\label{sect:reduction} \subsection{incoherent vs coherent analysis} We used two different methods for the MIDI data reduction. The first method is based on power spectrum analysis (hereafter called incoherent analysis), while the second method reduces all frames to the same OPD and adds them coherently (hereafter called coherent analysis). For the incoherent analysis of the data, we used the MIA package (MIDI Interactive Analysis, http://www.mpia-hd.mpg.de/MIDISOFT/) developed at the Max-Planck Institut f\"ur Astronomie in Heidelberg, while for the coherent analysis we used the EWS package (Expert Work Station) developed by Walter Jaffe at the Leiden observatory \citep{Jaffe_2004}. During the incoherent analysis, we separated the different scans in those with and without fringes, where each scan is Fourier-transformed from OPD to fringe frequency space. Considering the wavelengths present in the band and the rate at which the OPD is changing, the power is calculated in the correct frequency interval. The total power of all measured scans with fringes is then averaged and an estimate of the noise is subtracted. This noise estimate is based on the frames without fringes. This provides a value of the instrumental visibility squared of each channel. Contrary to the coherent method, the major difficulty of this method is that an accurate estimate of the off-fringe noise power is needed. Since our science targets have small fluxes in the 10 $\mu$m window, a reliable estimate of the noise power is difficult to obtain and we focused during data reduction on the coherent method (see below). Our incoherent analysis was only used to check the results obtained by a coherent analysis and both methods give consistent results. \subsection{coherent analysis} We first investigated the photometric datasets. The averages of the target and sky frames are calculated and subtracted, providing a raw two-dimensional spectrum of the object. The position and width of this spectrum is determined and a spatial mask is constructed from the location and average width of the spectrum at each wavelength position. After multiplication of the detector images with this mask, the rows are added providing a one dimensional raw spectrum of the object (i.e. not corrected for the atmospheric transmission and instrumental efficiency). The spatial mask is then used to extract the information of the interferometric observations as well, in the assumption that all instrumental parameters stay the same between the interferometric and photometric observation. The two detector spectra with opposite phase, are subtracted, resulting in one interferometrically modulated spectrum. In this way, the background is reduced by approximately 90 percent. Contrary to the incoherent method which allows the summing of scans where the relative OPD is not known, the coherent method needs an accurate determination of the atmospheric delay. The large wavelength coverage of the N-band ensures that this can be accurately done by measuring the fringes in frequency space \citep[rather than in OPD space which is done in an incoherent analysis, e.g.][]{Tubbs_2004}. As a first step, the known instrumental delay is removed from each frame after which the (previously unknown) atmospheric delay is retrieved using a group delay estimation. At this point, the data is not yet fully coherent because of the instrumental phase imposed on the data (e.g. the varying index of refraction of water vapor imposes variations in phase that are not removed by a group delay fitting). These phase shifts are almost constant as a function of frequency and can be approximated as a constant phase shift over the N-band \citep{Jaffe_2004}. Finally, the data can be added coherently to obtain the final visibility amplitude and differential phase. The instrumental visibility is then calculated dividing the fringe amplitude by the non-interferometric, photometric exposures. Repeating this procedure for a calibrator enables to estimate the instrumental visibility loss and thus determining the calibrated visibility of the science object. \subsection{the data} \subsubsection{photometry}\label{sect_photometry} A raw spectrum is obtained each night by subtracting the masked target frames from the masked sky frames. This spectrum is flux calibrated and corrected for atmospheric transmission using the calibrator spectra observed during the same night. For the calibrators, the intrinsic spectra were synthetised from {\sc marcs} atmosphere models \citep[][ and further updates]{Gustafsson_1975}, using the temperature, surface gravity and angular diameter determined in \cite{Vanboekel_2004}. This approach is preferred over a Rayleigh-Jeans approximation of the calibrator spectrum, since the SiO first overtone band head is not negligible in a K giant N-band spectrum. The 12\,$\mu$m flux of this synthetic spectrum is well within 1\,$\sigma$ of the color corrected IRAS 12\,$\mu$m flux. For both objects, the absolute flux calibration has been performed with the data of 10 February only, using the calibrators listed in Table \ref{tab:photcal}. Both reduced spectra (R\,$\sim 30$) are shown in Fig.~\ref{fig:totalspectrum}, where they are compared to independent spectra taken by the ISO/SWS (R\,$\sim 248$) and the SPITZER/IRS (R\,$\sim 127$) instrument. \begin{table} \caption{A summary of the calibrators used to flux calibrate the spectrum of \object{SX\,Cen} and \object{HD\,52961}. All calibrators listed were observed on February, 10. The reported flux is the IRAS 12.5 $\mu$m flux, and the diameters are taken from \cite{Verhoelst_2005}. } \label{tab:photcal} \begin{center} \begin{tabular}{l c c c c c} \hline \hline \multicolumn{1}{c}{calibrator} & \multicolumn{1}{c}{UT} & \multicolumn{1}{c}{airmass} & \multicolumn{1}{c}{spectral} & \multicolumn{1}{c}{diameter} & \multicolumn{1}{c}{flux} \\ \multicolumn{1}{c}{target} & \multicolumn{1}{c}{hh mm} & \multicolumn{1}{c}{} & \multicolumn{1}{c}{type} & \multicolumn{1}{c}{(mas)} & \multicolumn{1}{c}{(Jy)} \\ \hline HD\,67582 & 02 37 & 1.09 & K3 III & 2.30 $\pm$ 0.01 & 9.33 \\ HD\,67582 & 03 46 & 1.07 & K3 III & 2.30 $\pm$ 0.01 & 9.33 \\ HD\,49161 & 04 43 & 1.56 & K4 III & 2.44 $\pm$ 0.01 & 10.35 \\ HD\,107446 & 08 16 & 1.24 & K3.5 III & 4.43 $\pm$ 0.02 & 32.42 \\ \hline \end{tabular} \end{center} \end{table} \subsubsection{interferometry}\label{sec:interferometric_measurement} We start this discussion with an error estimate on the observed visibilities. The main source of error is the varying overlap between the interferometric beams due to imperfect source acquisition and residual image motion \citep[see e.g.][]{Leinert_2004}. This reduces the visibility for calibrator and/or science object with an unknown amount. The shape over the N-band, however, remains the same. The visibility variation within the spectral band, is therefore much more reliable than its absolute value. To get a quantitative estimate of the absolute uncertainty on the visibility, we look at all calibrators ($\sim$\,point sources) observed during one night. If the interferometric efficiency is constant throughout the night, all calibrator measurements should yield the same instrumental visibility. In Fig. \ref{fig:instrumental_visibility}, the mean instrumental visibility of all six calibrators observed in the prism mode during the night of February 9 is plotted. The variance on the mean is overplotted. It is clear from this figure that the instrumental loss of visibility is much higher at 8 $\mu$m than at 13 $\mu$m and that the uncertainty on the absolute value of the visibility is about 15 \%. However, when calibrating the visibility of the science source using a calibrator source observed in direct concatenation, this quantitative error is an upper limit. In the following, we use an error of 15 \% on the absolute visibility, which is therefore a conservative estimate. Calibrated visibilities are obtained dividing the raw visibility by the instrumental visibility. To calibrate the measurement of \object{SX\,Cen} observed at 9 February, we used the mean instrumental visibility as obtained from the three last calibrator measurements. Unfortunately, such a mean could not be used for the other measurements. Instead, we employed the calibrator closest in time to calibrate the visibility of the science source (see Table \ref{tab:log}). The resulting calibrated visibilities are shown in Figs.~\ref{fig:meanVisSXCEN} and \ref{fig:visibilities_reduced}. \section{Discussion}\label{sect:discussion} \subsection{visibilities} Because the angle between the projected baselines is the same to within 5 degrees for both objects, no large effects due to a possible asymmetry in the source morphology are expected. Therefore, as a first-order approximation, we modelled the circumstellar environment of objects using a uniform disc. The visibility in this assumption is given by $V = 2 J_1(x)/x$, where $x=2\pi \theta B/ \lambda$ with $\theta$ the diameter of the disc and $B$ the projected baseline length. This function is smoothly increasing with wavelength, as long as $x < 1.22 \pi$. The increase is however different for various amounts in resolving power, a steeper increase is observed if the source is more resolved. Assuming a temperature distribution in the disc, with the colder dust located further away from the star than the hotter dust, the gradient decreases. For an unresolved source, the value of the visibility remains constant at unity. The disc around \object{SX\,Cen} is unresolved in all measurements even using a 45\,m baseline. The visibility is close to unity and shows a flat distribution over the passband. The mean calibrated visibility for all measurements is shown in Fig. \ref{fig:meanVisSXCEN}. Because all measurements are observed at approximately the same projected angle (ranging from 41 to 46 degrees) this means that in that particular orientation, the structure is smaller than 11 mas at 8 $\mu$m and 17 mas at 13 $\mu$m in a uniform disc approximation. Using a gaussian distribution modelling, the FWHM gives respectively 7 mas and 10 mas as upper limits. \object{HD\,52961} shows quite a different picture. For this source, the visibilities are low (see Fig.~\ref{fig:visibilities_reduced}), thus the source is clearly resolved. Immediately noted is the fact that we do not get an increase in visibility amplitude which is quite linear (expected for a uniform disc model in the observed visibility range). Instead we see a ``bump'' in the visibility pattern ranging from 9 to 12\,$\mu$m. The geometry of the disc around this object can clearly not be modelled with the same uniform disc at all wavelenghts (see also Fig. \ref{fig:angularSize}). Using a uniform disc approximation for each wavelength independently is however instructive. For each wavelength bin, we made a $\chi^2$ minimalisation between the observed visibility at both baselines and a uniform disc model. This fit is shown for three representative wavelengths in the upper panel of Fig. \ref{fig:angularSize}. The diameter of the source at all wavelength bins in a uniform disc approximation is shown in the lower panel of Fig. \ref{fig:angularSize}. The measurements at both baselines are very consistent in a uniform disc approximation for each wavelength independently (the mean reduced chi-square over the wavelength band is as low as 0.09) and provide a diameter increasing from $\sim$35 mas at 8 $\mu$m to about $\sim$55 mas at 13 $\mu$m. In a gaussian distribution modelling, the FWHM gives respectively $\sim$23 mas and $\sim$34 mas (and a mean reduced chi-square of 0.22). The increase towards longer wavelengths is consistent with a dust-distribution for which the temparture decreases further away from the star. We however note that the observed increase in size is not smooth over the wavelength band. There is an increase in size from 8 $\mu$m to 8.5 $\mu$m and onwards 11.5 $\mu$m, while in between a sort of plateau exists. We interpret this plateau as resulting from a non-homogenous distribution of the radiating silicates which contribute most in the inner regions close to the central star, thus lowering the overall size (see also the following sections). For both objects we interpret the small angular scales of the dust around the objects as another clear indication that the circumstellar dust is stored in a compact Keplerian disc around the system. \subsection{spectra} The SED of both objects shows a significant near-IR excess, indicative of a hot dust component, while there is no evidence for an ongoing dusty mass loss of those rather hot stars. In \cite{Deruyter_2006}, this is interpreted as originating from a hot inner rim near dust sublimation temperature. Because no dust can survive at higher temperatures, this dust receives head on radiation from the star and is therefore supposed to be puffed up \citep[e.g. the wall model for HR4049 elaborated by][]{Dominik_2003}. This is corraborated by the lower limit on the opening angle of the disc as seen from the star of 13$^o$ for \object{HD\,52961} and 32$^o$ for \object{SX\,Cen} \citep{Deruyter_2006}. In this model, we expect a highly centerally peaked intensity distribution which provides that the correlated spectra measured by the interferometer are dominated by the inner regions of the disc. However, because of the varying spatial resolution from 8 to 13 $\mu$m, a slope is introduced in the correlated spectrum. To determine the magnitude of this effect, a detailed modelling has to be performed, which is out of the scope of this article. For now, we assume that this effect is a smooth function of wavelength, thus having only a marginal effect on any mineralogy determination. Because \object{SX\,Cen} is unresolved at all employed baseline settings, the correlated spectrum is identical to the single telescope spectrum and thus no additional information is available for this object. However, for \object{HD\,52961}, which is clearly resolved in both measurements, the shape of the two correlated spectra is predominantly determined by the inner parts of the disc. This means that we can construct independent spectra of geometrically different areas of dust around \object{HD\,52961}. The single telescope spectrum provides the full N-band spectrum of all the dust around \object{HD\,52961}. The correlated spectra sample smaller parts of the disc. These spectra, each sampling a different geometrical part of the disc, are shown in Fig. \ref{fig:allspectraatonce}. From this figure, it is clear that the shape of the correlated spectra is quite different from the single telescope spectrum. This points to a different chemical composition of the inner part of the disc and the outer part of the disc. To quantify this, we have fitted the different spectra independantly. \subsection{silicate mineralogy}\label{sec:chemical_composition} In order to determine the mineralogy and sizes of the emitting dust grains, we made a fit to the N-band spectra using calculated emissivities of irregularly shaped, chemically homogeneous dust grains. The most important dust species causing spectral signature in the 10\,$\mu$m window are amorphous and crystalline olivine (Mg$_{2x}$Fe$_{2-2x}$SiO$_4$), amorphous and crystalline pyroxene (Mg$_{x}$Fe$_{1-x}$SiO$_3$), and amorphous silica (SiO$_2$), where $x$ determines the Mg/Fe ratio ($x=1$ for the crystalline silicates, $x=0.5$ for the amorphous silicates). The complex refractive indices for the different grain species were taken from various authors listed in Table \ref{tab:refractive_indices}. To simulate the effects of particle irregularity we employ a particular implementation of the so-called \emph{statistical approach} using a distribution of hollow spheres. This distribution is very successful in reproducing the measured absorption spectra of irregularly shaped particles \citep{Min_2003, Min_2005}. In addition to the dust species causing the feature, we also add a continuum contribution which accounts for emission by large grains and/or for the possible presence of featureless components such as metallic iron and iron sulfide. This continuum contribution is modeled using a constant mass absorption coefficient. In the 10 $\mu$m region we are mainly sensitive to the dust grains smaller than a few $\mu$m. We represent the size distribution of the particles by two different grain sizes, $0.1$ and 1.5\,$\mu$m. A similar method was successfully employed by, for example, \citet{Bouwman_2001}, \cite{Honda_2003}, \cite{Honda_2004}, \cite{Vanboekel_2004nature} and \cite{Vanboekel_2005} to fit 10\,$\mu$m emission spectra of circumstellar discs. Particles larger than a few $\mu$m contribute mainly to the continuum. In addition we assume that the thermal radiation we analyze originates from optically thin parts of the disc, which allows us to add the contributions from the various components linearly. For the emission of the outer parts of the disc, tentatively attributed to layers directly heated by the stellar flux, this is a reasonable approximation. Because the stellar radiation is incident under a high inclination, the temperature distribution in the surface layer of the disc must be very sharp and therefore, the emission in the N-band comes likely from optically thin parts. For the inner parts of the disc, the situation is more complex: a large fraction of the radiation comes from the inner rim, which has regions of both low and high opacity. The fit, using an optically thin assumption for the different contributing minerals, is therefore certainly too simplistic. We use it here as a first order estimate to show the chemical gradient of the silicates in the disc, but a detailed 2D radiative transfer model with a gradient in the physico-chemical condition of the dust grains will be needed to quantify the results. This is outside the scope of this paper. We assume all dust grains, including the ones causing the continuum, to have the same temperature distribution. Due to the limited wavelength range this temperature distribution can be represented by a single Planck curve with a characteristic temperature $T_c$. This is justified because it is very likely that the dust grains of different species are coagulated, implying thermal contact between the various components. The characteristic temperatures used in the modeling are given in Table \ref{tab:composition}. \begin{table}[!t] \begin{center} \linespread{1.3} \selectfont \begin{tabular}{lc} \hline \hline grain species & reference \\ \hline Amorphous Olivine & \cite{Dorschner1995} \\ Amorphous Pyroxene & \cite{Dorschner1995} \\ Forsterite & \cite{Servoin1973} \\ Enstatite & \cite{Jaeger1998} \\ Amorphous Silica & \cite{1960PhRv..121.1324S} \\ \hline \end{tabular} \end{center} \linespread{1} \caption{A list of the references of the complex refractive indices employed for the various grain species. } \label{tab:refractive_indices} \end{table} The abundances of the dust components are determined by using a linear least square fitting procedure with constraints on the weights to avoid negative values. The temperature of the grains and the underlying continuum is varied from 0 to 1500\,K until a best fit is obtained. The dust parameters derived from the unresolved spectrum of \object{SX\,Cen} are given in the upper row of Table~\ref{tab:composition}. The resulting best fit spectrum is shown as a dotted line in Fig.~\ref{fig:totalspectrum}. The grains in the circumstellar environment of \object{SX\,Cen} are highly crystalline and also on average relatively large compared to the interstellar grain population. This implies a large amount of dust processing in the circumstellar environment. For \object{HD\,52961}, we fit the spectrum corresponding to the inner disk (the correlated spectrum, angular size $\sim$ 20 mas) and that corresponding to the outer disk (the total disk spectrum from which the correlated spectrum is subtracted) separately. We focus here on the correlated spectrum taken with the 40\,m baseline. The resulting best fit model spectra are shown in Fig.~\ref{fig:chemical_hd52961} and the composition is given in Table~\ref{tab:composition}. The varying spatial resolution over the N-band introduces an extra slope in the spectra which was not corrected. Therefore, the characteristic temperatures ($T_c$) derived for the inner and outer disc spectra are not realistic. The influence of this slope on the determined chemical fractions is however expected marginal (see e.g. \cite{Vanboekel_2004nature}). The average composition over the disk can be derived by taking the mass weighted average of the inner (56\%) and outer (44\%) disk regions. The overall composition of the dust in \object{HD\,52961} is for $\sim$50\% crystalline, and contains $\sim$60\% $1.5\,\mu$m grains. This is considerably less than what we find in \object{SX\,Cen}. It is also clear from Table~\ref{tab:composition} that the crystalline silicates are not uniformly distributed over the disk. The inner disk has a much higher crystallinity than the outer disk. In order to fit the prominent feature around 9.5\,$\mu$m in the total and the outer disk spectra of \object{HD\,52961}, we have to add large (1.5\,$\mu$m) silica grains. We have tried several other dust components in order to explain this spectral feature, but found no spectral match using any of them. We have no explanation for the presence of these amounts of large silica grains, and thus its detection is debatable. However, its presence is also indicated from the Spitzer IRS spectrum, which shows a weak feature around 21\,$\mu$m which is naturally reproduced using large silica grains (not shown). For a full mineralogy, the broader wavelength range sampled by our Spitzer data is needed which is outside the scope of this paper. \begin{table*}[!t] \begin{center} \linespread{1.3} \selectfont \begin{tabular}{cccccccccccccc} \hline \hline Star & T$_c$ & Cryst. & Large & \multicolumn{2}{c}{Olivine [\%]} & \multicolumn{2}{c}{Pyroxene [\%]} & \multicolumn{2}{c}{Forsterite [\%]} & \multicolumn{2}{c}{Enstatite [\%]} & \multicolumn{2}{c}{Silica [\%]} \\ & 10$^2$\,K & [\%] & grains [\%] & Small & Large & Small & Large & Small & Large & Small & Large & Small & Large\\ \hline \object{SX\,Cen} & $6.8_{-0.3}^{+0.3}$ & $78_{-23}^{+17}$ & $93_{-4}^{+3} $ & - & $21_{-17}^{+24}$ & - & $1_{-1}^{+6}$ & $7_{-3}^{+3}$ & $25_{-9}^{+9}$ & - & $46_{-15}^{+12}$ & $1_{-1}^{+1}$ & $0_{-0}^{+2}$\\ \object{HD\,52961} (inner)&$12_{-2}^{+2}$ & $79_{-12}^{+10}$ & $63_{-12}^{+7} $ & $6_{-6}^{+13}$ & $4_{-4}^{+16}$ & - & $1_{-1}^{+15}$ & $31_{-3}^{+3}$ & $1_{-1}^{+7}$ & $0_{-0}^{+6}$ & $47_{-11}^{+9}$ & - & $9_{-3}^{+3}$\\ \object{HD\,52961} (outer)&$14_{-1}^{+1}$ & $19_{-3}^{+4} $ & $59_{-17}^{+18}$ & - & $5_{-5}^{+10}$ & $22_{-17}^{+16}$ & $20_{-16}^{+19}$ & $19_{-3}^{+3}$ & - & - & $0_{-0}^{+4}$ & $0_{-0}^{+1}$ & $34_{-3}^{+4}$\\ \hline \end{tabular} \end{center} \linespread{1} \caption{The composition and grain sizes of the dust in the circumstellar environments around \object{SX\,Cen} and \object{HD\,52961} as derived from our fitting procedure. The olivine and pyroxene grains are amorphous while the forsterite and enstatite are crystalline. For \object{HD\,52961} 56\% of the dust mass is in the inner disk region. It should be noted that, as explained in the text, the temperatures determined for the inner and outer disc spectra of \object{HD\,52961} are not realistic.} \label{tab:composition} \end{table*} \subsection{formation history of the disc} The composition of the dust in the circum-binary disc as a function of distance from the binary can give important clues to its formation history. In principle, the disc could have been formed by capturing a "normal" AGB wind, or through non-conservative mass transfer in an interacting binary. In the wind scenario, it is not unreasonable to assume that most dust was in the form of amorphous silicates, since this is the usual dust composition for O-rich AGB stars with a moderate to high mass loss rate \citep[see e.g.][]{Sloan_1995,Waters_1996L, Cami_2002}. In the interacting binary scenario, the dust may or may not have formed before the material entered the circum-binary disc, but in any case the thermal history of that dust would have been very different from that of the wind scenario: the grains were likely at high temperatures for a long period of time, increasing the chances of a substantial crystallisation. Therefore the wind scenario predicts a predominantly amorphous silicate composition, while the interacting binary scenario more likely produces (highly) crystalline discs. Once in the disc, both grain aggregation and crystallisation may occur. Grain aggregation is a strong function of density and thus would be most efficient in the inner disc regions. Large grains settle quickly to the mid-plane thus creating a cold mid-plane population of grains, which we believe is responsible for the millimeter continuum emission \citep[see e.g.][]{Deruyter_2006}. The inner disc reaches temperatures above the glass temperature, forcing the grains to anneal. Therefore, in both the wind and in the interacting binary scenarios the innermost disc regions are expected to be strongly crystalline. The two scenarios predict strongly differing radial gradients in crystallinity however. The present-day orbital parameters of the binary systems with circum-binary discs strongly suggest that interaction took place when the current post-AGB star was on the AGB. Therefore it seems difficult to imagine that a standard stellar wind formed the discs, and one would expect the discs on average to have much higher crystallinity than typical AGB winds. The recent spectral survey by \cite{Deruyter_2006b} indeed suggests that the circumbinary discs are much more crystalline than typical AGB outflows. One complicating factor is that at present not much is known about the composition of the dust in AGB outflows in the dust forming layers: by far most data are spatially unresolved and present the final outcome of the dust formation process in O-rich AGB outflows. Our MIDI observations indeed confirm that the inner disc region of \object{HD\,52961} is extremely crystalline, and that the outer disc regions are less so. At first glance this would suggest the wind scenario is more likely, but the orbital parameters indicate substantial AGB interaction. These first MIDI observations thus raise interesting questions: is the outer disc of \object{HD\,52961} really amorphous and what kind of disc formation scenario could lead to amorphous grains? Does this hold for all systems, or is there an orbital separation dependence? Clearly more study is required to answer these questions. \subsection{comparison with Herbig Ae/Be stars}\label{sec:comparison} In \cite{Deruyter_2006}, it is argued that the broad-band SED characteristics of the discs around binary post-AGB objects are very similar to the those of the Herbig Ae/Be group II sources. Herbig Ae/Be stars are intermediate mass pre-main sequence stars surrounded by remnant material of the star formation process. For these objects, the existence of a passive circumstellar disc is firmly established \citep[e.g.][ and references therein]{Waters_1998, Eisner_2003}. The Herbig Ae/Be stars are subdivided in two groups \citep{Meeus_2001} with the group I sources showing a rising mid-IR flux excess, while the group II sources only show a modest mid-IR excess. The difference in SED characteristics between both groups is attributed to disc geometry. The mid-IR excess of group I sources is indicative of the flaring of the outer disc, while the inner rim of the group II sources shadows the whole disc and no flaring occurs \citep{Chiang_1997, Dullemond_2001}. \cite{Vanboekel_2004} used the most recent models \citep{Dullemond_2004} of the discs around Herbig Ae/Be stars to compute the visibilities to be expected in the MIDI wavelength range. \cite{Leinert_2004} on the other hand made observations with the MIDI instrument of several of these objects. We make a comparison of the results obtained in these publications for the group II sources and our observations under the assumption of the similarity of both source geometries. Concerning the continuum radiation, the modelling performed in \cite{Vanboekel_2004} shows that the size of the disc increases more rapidly from 8 to 13 $\mu$m than the interferometric resolution decreases. This provides a visibility curve which is decreasing from 8 to 13 $\mu$m. The observations indeed show this qualitative behaviour, however some objects, e.g. HD144432, show a rather horizontal slope. This very similar behaviour is observed for \object{HD\,52961} as well. The slope of the continuum visibility does not increase as expected for a uniform disc, it is instead rather constant with wavelength. Concerning the visibility in the feature, the modelling performed by \cite{Vanboekel_2004} suggests a lowered visibility for the silicate feature than for the continuum. The disc is irradiated by the central object and therefore the disc surface is hotter than the disc midplane. Because the opacity in the silicate resonance band is higher than in the continuum, one looks less deep into the disc in the resonance. This results in the fact that in the 10 $\mu$m region, a larger region in the resonance is seen than in the continuum. The observations of Herbig Ae/Be stars show a similar qualitative behaviour, however the visibility decrease is less pronounced. In fact, \cite{Vanboekel_2004nature} finds that for three Herbig Ae stars of the sample of \cite{Leinert_2004}, there is a large radial gradient in the processing. The innermost region of the proto-planetary discs has a substantially higher crystallinity degree with a shape very similar to that of comets in our solar system, while the outer region is clearly less processed. Clearly, the homogenous distribution of dust adopted in the modelling of these discs is a very crude approximation. For \object{HD\,52961}, no visibility decrease is observed in the feature, instead an increase is seen. The spatial distribution of the dust responsible for the resonance is not homogeneously distributed, instead the hot inner region of the dust is much more crystalline than the outer parts (Sect. \ref{sec:chemical_composition}). This qualitative similarity in the distribution of the chemical species in the discs around some Herbig Ae stars and \object{HD\,52961} is surprising in the context of the completely different formation history of both. \section{Conclusions}\label{sect:conclusions} The main conclusion of our presented MIDI observations is that they prove the very compact nature of the circumstellar environments of \object{HD\,52961} and \object{SX\,Cen}. \object{SX\,Cen} is not resolved using a 45\,m baseline, which gives an upperlimit of only 18\,AU at the estimated distance. For the well resolved \object{HD\,52961}, the angular size in the N-band varies between 35 and 55\,mas in a uniform disc approximation, which translates to a size of 50 and 80\,AU. Both stars have an effective temperature in the 6000\,K range and since there is no evidence for a current dusty mass-loss we interpret these results as a very stringent proof of the existence of a stable reservoir near the star. A Keplerian disc seems the only plausible solution. The dust sublimation temperature is reached much further out than the binary orbits, hence the discs must be circumbinary. This is corroborated by the measured size of the dust-emission region around \object{HD\,52961}. Given the size of the orbits, the discs were probably formed in a poorly understood phase of strong binary interaction, when the star was at giant dimensions. Both discs are O-rich and there is no evidence for a C-rich component. They were consequently formed prior to the late AGB evolution where the stars could have changed into C-stars. The mass of the companion of \object{SX\,Cen} (1.4 -- 1.9 M$_{\odot}$) is probably within the range of C-star progenitors. We conclude that the normal single star AGB evolution was shortcut by the presence of a binary companion. Clearly the formation of a stable Keplerian disc is a key ingredient in the late evolution of both binaries. \object{SX\,Cen} is an RV\,Tauri star of photometric class b which shows a long term variability of the mean magnitude with a period similar to its orbital period probably due to variable circumstellar extinction in the line of sight during orbital motion. The inclination cannot be very small. Additional interferometric data on different projected angles will be necessary to probe the expected asymmetries. The characteristics of the dust grains seem to be very different from normal single star outflows. This is shown in the mineralogy of the silicate resonance feature which shows for both objects a highly crystalline component and a size distribution with a much stronger component of large ($> 1$\,$\mu$m) grains than what is observed in outflows of AGB stars. It is not clear whether this reflects the formation history of the disc or this is due to the longer processing time of the dust in the Keplerian discs. Our analysis of \object{HD\,52961} shows that the crystallinity is clearly concentrated in the hotter inner region of the disc. Crystallisation by annealing is very temperature dependent and a similar picture arises as what is seen in the discs around some young stellar objects: the grains in the hot inner region were subject to a much stronger processing while in the outer region remained less processed. MIDI as spectrally dispersed N-band interferometer is an ideal instrument to study the chemo-physical structure of the inner regions of these discs. \begin{acknowledgements} The authors would like to thank Jeroen Bouwman for the reduction of the SPITZER/IRS spectrum of \object{HD\,52961} and Bram Acke for the reduction of the ISO/SWS spectrum of \object{SX\,Cen}. We also like to thank the referee, K. Ohnaka, for the many valuable comments. We thank the staff of the Geneva Observatory and the staff of the Instituut voor Sterrenkunde of the K.U.Leuven for the generous award of time on the Swiss Euler telescope at La Silla and the Flemish Mercator telescope at La Roque de los Muchachos respectively. We also thank our colleagues from the Instituut voor Sterrenkunde for their contribution to the gathering of the data. P.D.~and H.V.W.~acknowledge financial support from the Fund for Scientific Research of Flanders (FWO). \end{acknowledgements} \bibliographystyle{aa}
Title: Full polarization study of SiO masers at 86 GHz
Abstract: We study the polarization of the SiO maser emission in a representative sample of evolved stars in order to derive an estimate of the strength of the magnetic field, and thus determine the influence of this magnetic field on evolved stars. We made simultaneous spectroscopic measurements of the 4 Stokes parameters, from which we derived the circular and linear polarization levels. The observations were made with the IF polarimeter installed at the IRAM 30m telescope. A discussion of the existing SiO maser models is developed in the light of our observations. Under the Zeeman splitting hypothesis, we derive an estimate of the strength of the magnetic field. The averaged magnetic field varies between 0 and 20 Gauss, with a mean value of 3.5 Gauss, and follows a 1/r law throughout the circumstellar envelope. As a consequence, the magnetic field may play the role of a shaping, or perhaps collimating agent of the circumstellar envelopes in evolved objects.
https://export.arxiv.org/pdf/astro-ph/0601098
\def\etal{et al.\ } \def\kms{km\thinspace s$^{-1}$ } \def\Lsun{L$_\odot$} \def\water{H$_2$O~} \def\Msun{M$_\odot$} \def\ms{m\thinspace s$^{-1}$} \def\percc{cm$^{-3}$} \title{Full polarization study of SiO masers at 86 GHz} \author{F.\,Herpin\inst{1}, A.\,Baudry\inst{1}, C.\,Thum\inst{2}, D.\,Morris\inst{2}$^,$\inst{3} and H.\,Wiesemeyer\inst{2}} \institute{ Observatoire Aquitain des Sciences de l'Univers, Laboratoire d'Astrodynamique, d'Astrophysique et d'A\'eronomie de Bordeaux, CNRS/INSU UMR n$^{\circ}$ 5804, BP 89, 33270, France \and IRAM, 300 rue de la Piscine, Domaine Universitaire, 38406 Saint Martin d'H\`eres, France \and Present address: Raman Research Institute, 560080 Bangalore, India } \titlerunning{SiO maser polarization} \abstract{We study the polarization of the SiO maser emission in a representative sample of evolved stars in order to derive an estimate of the strength of the magnetic field, and thus determine the influence of this magnetic field on evolved stars. We made simultaneous spectroscopic measurements of the 4 Stokes parameters, from which we derived the circular and linear polarization levels. The observations were made with the IF polarimeter installed at the IRAM 30m telescope. A discussion of the existing SiO maser models is developed in the light of our observations. Under the Zeeman splitting hypothesis, we derive an estimate of the strength of the magnetic field. The averaged magnetic field varies between 0 and 20 Gauss, with a mean value of 3.5 Gauss, and follows a $1/r$ law throughout the circumstellar envelope. As a consequence, the magnetic field may play the role of a shaping, or perhaps collimating agent of the circumstellar envelopes in evolved objects.} \textbf{Keywords.} Maser: SiO -- polarization-- survey -- stars: late-type, evolution, magnetic field \section{Introduction} The prodigious mass loss observed in numerous and widespread evolved stars make these objects the main recycling agents of the interstellar medium, and thus one of the most important objects in the Universe. Even though our knowledge of evolved stars has considerably improved over recent years, some of their main characteristics remain insufficiently understood (see the review by Herwig 2003): which mechanisms are responsible for their drastic change of geometry when evolving to the Planetary Nebula (hereafter PN) stage ? What is powering so efficiently the mass loss and could the magnetic field play a major role ? Important information about the physics and chemistry prevailing in the circumstellar envelope (hereafter {\em CSE}) of evolved stars can be retrieved from radiowave line emission of molecules, specially from maser emission (see the review by Bujarrabal 2003). These envelopes can be probed at different depths through the study of three masing molecules, OH, \water and SiO. Our current knowledge indicates that: \begin{itemize} \item OH radiation traces the outer part of the envelope, at 1000-10000 AU from the central star; \item \water molecules are located at intermediate distances, i.e. a few 100 AU; \item SiO maser emission comes from the inner regions of the envelope, between 5 to 10 AU (a few stellar radii R$_{\star}$). \end{itemize} The SiO maser emission is produced in small gas cells, and is known to be polarized. The polarization (circular or linear) and angle of the emission can be measured and thus improve our knowledge of these objects. In addition, studying the maser polarization can shed light on the maser theory itself. As explained further in this paper, several uncertainties in the theory make data interpretation often difficult, and new observational data are helpful. One of the most interesting quantities that can be derived from polarization measurements is the stellar magnetic field. According to theory (e.g. Elitzur 1996 or 2002), measurement of the maser radiation polarization can lead to an estimation of the magnetic field strength B and can reveal its spatial structure (via interferometric observations). In single dish observations, all of the maser components get smeared within the beam and only the mean value of B along the line of sight ($B_{//}$) can be derived. Only SiO masers are capable of tracing the magnetic field as close as $\sim$ 5 AU from the central star. But if SiO masers are to be used as a B-field tracer, we first need to give evidence that SiO masers are reliable B-field tracers. This requires more detailed theories than available today. Nevertheless, we tentatively derive in this work the field strength in the CSE inner layers of several evolved stars. Research on astronomical masers polarization is very active but is made difficult both by the lack of specific instrumental facilities and by the excitation and propagation of the masers themselves. Until now, numerous polarimetric observations of OH masers have been done, several of \water masers, but few of SiO maser emission. Few SiO polarimetric observations have been done with VLBI giving the very first images of the magnetic field in some objects (e.g. Kemball \& Diamond 1997 in TX Cam). Most of the early SiO studies were done in linear polarization. The first complete SiO polarimetric observations were performed by Johnson \& Clark (1975), then by Troland \etal (1979); emission was found to be typically 15-30 \% linearly polarized and to exhibit no circular polarization. Barvainis, McIntosh \& Predmore (1987), and McIntosh \etal (1989) measured circular polarization of $1-9$ \% in several stars. Circular (0-4 \%) and linear (3.7-9.7 \%) polarizations were measured in VY CMa by McIntosh, Predmore \& Patel (1994). Later, Kemball \& Diamond (1997) made the first image of the magnetic field in the atmosphere of TX Cam, measuring a circular polarization level of 5 \% with some features showing polarization up to 30-40 \%. It must be stressed that SiO, as H$_2$O, is a non-paramagnetic species. Zeeman splitting exists but the sublevels overlap; the effect is thus undetectable and hence only net polarization can be used to trace the magnetic field. The current status of our knowledge on the magnetic field strength can be summarized as follows: \begin{itemize} \item between 1000-10000 AU, $B_{//}\sim 5-20$ mG (OH masers, e.g. Kemball \& Diamond 1997, Szymczak \& Cohen 1997); \item at a few 100 AU from the star, $B_{//} \sim$ a few 100 mG (\water masers, e.g. Vlemmings, Diamond \& van Langevelde 2001, Vlemmings, van Langevelde \& Diamond 2005); \item at 5-10 AU, $B_{//} \sim 5-10$ G (SiO masers; Kemball \& Diamond 1997, in TX Cam). \end{itemize} The main purpose of this work is to measure and analyze the SiO maser polarization in terms of magnetic field strength in a representative sample of evolved stars. Our observations are presented in Section 2; they include simultaneous spectroscopic measurements of the 4 Stokes parameters. The results for individual stars are discussed in Section 3. In Section 4, we compare our data with predictions from existing SiO maser models and initiate a discussion on the validity of these models. Within the limitations of one of these models we derive the magnetic field strength and try to determine the role of the magnetic field. More broadly, a summary of the magnetic field topic in evolved stars is also given in Section 4. In Section 5 we give some concluding remarks. \section{Observations} An electromagnetic plane wave is defined by two components (horizontal and vertical): \begin{equation} \label{ } e_H(z,t)=E_H\ e^{j(\omega t-kz-\delta)} \end{equation} \begin{equation} \label{ } e_V(z,t)=E_V\ e^{j(\omega t-kz)} \end{equation} where $\delta$ is the phase difference between horizontal and vertical components. Its energy flux is described by the 4 Stokes parameters: \begin{equation} \label{ } I = <{E_H}^2> + <{E_V}^2> \end{equation} \begin{equation} \label{ } Q = <{E_H}^2> - <{E_V}^2> \end{equation} \begin{equation} \label{ } U = 2 <E_H E_V cos\ \delta> \end{equation} \begin{equation} \label{ } V = 2 <E_H E_V sin\ \delta> \end{equation} From these parameters, one deduces: \begin{itemize} \item the circular polarization rate $p_C = V/I$ \item the linear polarization rate $p_L = \sqrt{Q^2+U^2}/I$ \item the polarization angle $\chi = \frac{\arctan (U/Q)}{2}$ \end{itemize} The linear$/$circular polarization rate is sometimes called the linear$/$circular fractional polarization. \scriptsize \begin{table*} [htb] \caption{ \label{table} Stars observed in this work. The stellar type is derived from the literature as are the mass loss rates (e.g. Loup \etal 1993) and the period (e.g. AAVSO data). } {\begin{tabular}{l|c|c|c|c|c|c|c} \hline {\bf Stars} & RA & DEC & Type & $V_{LSR}$ & dM/dt & Period & Optical \\ & (J2000) & (J2000) & & [\kms] & [M$_{\odot}/$yr]& [days] & phase \\ \hline IRAS 18055-1433 & 18:08:23.20 & -14:32:43.0 & IR late-type & 180 & unknown & unknown & \\ IRAS 18158-1527 & 18:18:41.50 & -15:26:25.0 & IR late-type & 20 & unknown & unknown & \\ IRAS 18204-1344 & 18:23:17.90 & -13:42:46.0 & IR Supergiant (M8) & 45 & 4.2 $10^{-6}$& unknown & \\ W And & 02:17:32.96 & 44:18:17.8 & Mira (S6,1E-S9,2E$/$M4-M1) & -35 & 8.0 $10^{-7}$ & 395.9 & 0.62 \\ AU Aur & 04:54:15.00 & 49:54:00.3 & Mira (C6-7,3E(N0E)) & 8 & 1.1 $10^{-7}$ & 400 & 0.17 \\ NV Aur & 05:11:19.43 & 52:52:33.6 & Mira (M10) & 2 & 7.6 $10^{-6}$ & 635 & \\ R Aur & 05:17:17.69 & 53:35:10.0 & Mira (M6.5E-M9.5E) & -3 & 9.8 $10^{-7}$ & 457.5 & 0.83 \\ RU Aur & 05:40:07.89 & 37:38:10.7 & SRb (M7E-M9E) & -35 & unknown & 466.4 & 0.25 \\ TX Cam & 05:00:51.15 & 56:10:54.0 & Mira (M8-M10) & 10 & 2.5 $10^{-6}$ & 557.4 & 0.70 \\ V Cam & 06:02:32.30 & 74:30:27.1 & Mira (M7E) & 8 & 1.6 $10^{-6}$ & 522.4 & 0.91 \\ R Cnc & 08:16:33.83 & 11:43:34.6 & Mira (M6E-M9E) & 14 & 6 $10^{-7}$ & 361.6 & 0.33 \\ W Cnc & 09:09:52.63 & 25:14:53.8 & Mira (M6.5E-M9E) & 38 & 3.1 $10^{-8}$ & 393.2 & 0.76 \\ VY CMa & 07:22:58.33 & -25:46:03.2 & Red Supergiant (M3-M4II) & 15 & $10^{-5}$ & 400 & 0.37 \\ S CMi & 07:32:43.08 & 08:19:05.3 & Mira (M6E-M8E) & 52 & 4.1 $10^{-8}$ & 332.9 & 0.35 \\ R Cas & 23:58:24.79 & 51:23:19.5 & Mira (M6E-M10E) & 27 & 1.1 $10^{-6}$ & 430.4 & 0.66 \\ S Cas & 01:19:41.97 & 72:36:39.3 & Mira (S3,4E-S5,8E) & -28 & 3.1 $10^{-6}$ & 612.4 & 0.98 \\ T Cas & 00:23:14.25 & 55:47:33.3 & Mira (M6E-M9.0E) & -7 & 5.1 $10^{-7}$ & 444.8 & 0.22 \\ T Cep & 21:09:31.85 & 68:29:27.6 & Mira (M5.5E-M8.8E) & -1 & 1.4 $10^{-7}$ & 388.1 & 0.94 \\ R Com & 12:04:15.20 & 18:46:56.7 & Mira (M5E-M8EP) & -3 & $10^{-7}$ & 362.8 & 0.43 \\ S CrB & 15:21:23.96 & 31:22:02.7 & Mira (M6E-M8E) & 3 & 5.8 $10^{-7}$ & 360.2 & 0.44 \\ R Crt & 11:00:33.87 & -18:19:29.6 & SRb (M7III) & 10 & 7.5 $10^{-7}$ & 160 & \\ $\chi$ Cyg & 19:50:33.94 & 32:54:50.6 & Mira (S6,2E-S10,4E) & 10 & 5.6 $10^{-7}$ & 408 & 0.37 \\ UX Cyg & 20:55:05.40 & 30:24:53 & irregular variable (M4E-M6.5E) & 1 & 3.2 $10^{-6}$ & 565 & 0.77 \\ R Hya & 13:29:42.82 & -23:16:52.9 & Mira (M6E-M9E(TC)) & -8 & 1.4 $10^{-7}$ & 388.8 & 0.95 \\ W Hya & 13:49:02.03 & -28:22:03.0 & SRa (M7.5E-M9EP) & 42 & 8.1 $10^{-8}$ & 361 & 0.83 \\ X Hya & 09:35:30.26 & -14:41:28.5 & Mira (M7E-M8.5E) & 26 & 4.8 $10^{-8}$ & 301.1 & 0.0 \\ R Leo & 09:47:33.49 & 11:25:44.0 & Mira (M6E-M8IIIE-M9.5E) & 0 & $10^{-7}$ & 309.9 & 0.84 \\ W Leo & 10:53:34.44 & 13:42:54.4 & Mira (M5.5E-M7E) & 49 & unknown & 391.7 & 0.45 \\ R LMi & 09:45:34.28 & 34:30:42.8 & Mira (M6.5E-M9.0E) & 2 & 2.8 $10^{-7}$ & 372.2 & 0.56 \\ T Lep & 05:04:50.84 & -21:54:16.2 & Mira (M6E-M9E) & -29 & 7.3 $10^{-9}$ & 368.1 & 0.59 \\ RS Lib & 15:24:19.78 & -22:54:39.7 & Mira (M7E-M8.5E) & 7 & 1.8 $10^{-8}$ & 217.6 & 0.28 \\ Ap Lyn & 06:34:34.90 & 60:56:33.0 & Mira (M9) & -23 & 4.9 $10^{-6}$ & unknown & \\ U Lyn & 06:40:46.49 & 59:52:01.6 & Mira (M7E-M9.5E) & -10 & unknown & 433.6 & 0.13 \\ GX Mon & 06:52:46.90 & 08:25:20.0 & Mira (M9) & -9 & 5.4 $10^{-6}$ & 527 & \\ SY Mon & 06:37:31.28 & -01:23:43.6 & Mira (M6E-M9) & -57 & unknown & 422.2 & 0.18 \\ V Mon & 06:22:43.58 & -02:11:43.2 & Mira (M5E-M8E) & 5 & unknown & 341 & 0.0 \\ U Ori & 05:55:49.18 & 20:10:30.7 & Mira (M6E-M9.5E) & -38 & 2.8 $10^{-7}$ & 368.3 & 0.24 \\ RR Per & 02:28:28.73 & 51:16:21.1 & Mira (M6E-M7E) & 7 & unknown & 389.6 & 0.17 \\ S Per & 02:22.51.76 & 58:35:11.4 & SRc (M3IAE-M7) & -40 & 1.4 $10^{-6}$ & 822 & 0.58 \\ QX Pup & 07:42:16.83 & -14:42:52.1 & PN (M6) & 34 & 1.1 $10^{-4}$ & unknown & \\ Z Pup & 07:32:38.06 & -20:39:29.2 & Mira (M4E-M9E) & 4 & unknown & 508.6 & 0.74 \\ VX Sgr & 18:08:04.05 & -22:13:26.6 & Red Supergiant (M4EIA-M10EIA) & 6 & 5.5 $10^{-6}$ & 732 & 0.38 \\ AH Sco & 17:11:17.02 & -32:19:30.7 & SRc (M4E-M5IA-IAB) & -7 & $10^{-6}$ & 713.6 & 0.98 \\ RR Sco & 16:56:37.85 & -30:34:48.1 & Mira (M6II-IIIE-M9) & -28 & 1.1 $10^{-8}$ & 281.4 & 0.35 \\ R Ser & 15:50:41.74 & 15:08:01.4 & Mira (M5IIIE-M9E) & 28 & 2.6 $10^{-7}$ & 356.4 & 0.20 \\ S Ser & 15:21:39.53 & 14:18:53.1 & Mira (M5E-M6E) & 20 & $<2.2$ $10^{-7}$ & 371.8 & 0.76 \\ WX Ser & 15:27:47.30 & 19:33:48.0 & Mira (M8E) & 7 & 2.6 $10^{-6}$ & 425.1 & 0.30 \\ \hline \end{tabular}} \end{table*} \normalsize \addtocounter{table}{-1} \scriptsize \begin{table*} [htb] \caption{ \label{table} (-continued). Stars observed in this work. The stellar type is derived from the literature as are the mass loss rates (e.g. Loup \etal 1993) and the period (e.g. AAVSO data). } {\begin{tabular}{l|c|c|c|c|c|c|c} \hline {\bf Stars} & RA & DEC & Type & $V_{LSR}$ & dM/dt & Period & Optical \\ & (J2000) & (J2000) & & [\kms] & [M$_{\odot}/$yr]& [days] & phase \\ \hline IK Tau & 03:53:28.80 & 11:24:22.7 & Mira (M6E-M10E) & 35 & 4.4 $10^{-6}$ & 470 & 0.80 \\ R Tau & 04:28:18.00 & 10:09:44.8 & Mira (M5E-M9E) & 14 & 6.5 $10^{-8}$ & 320.9 & 0.68 \\ RX Tau & 04:38:14.57 & 08:20:09.4 & Mira (M6E-M7E) & -41 & $<5.8 10^{-8}$ & 331.8 & 0.13 \\ R Tri & 02:37:02.32 & 34:15:51.4 & Mira (M4IIIE-M8E) & 57 & 1.1 $10^{-7}$ & 266.9 & 0.0 \\ R UMi & 16:29:57.87 & 72:16:49.0 & SRb (M7IIIE) & -6 & unknown & 325.7 & 0.35 \\ S UMi & 15:29:34.66 & 78:38:00.2 & Mira (M6E-M9E) & -42 & unknown & 331 & 0.68 \\ R Vir & 12:38:29.95 & 06:59:19.0 & Mira (M3.5IIIE-M8.5E) & -26 & unknown & 145.6 & 0.12 \\ RS Vir & 14:27:16.39 & 04:40:41.1 & Mira (M6IIIE-M8E) & -12 & 3.8 $10^{-7}$ & 353.9 & 0.61 \\ RT Vir & 13:02:37.96 & 05:11:08.5 & SRb (M8III) & 18 & 7.4 $10^{-7}$ & 155 & 0.60 \\ S Vir & 13:33:00.11 & -07:11:41.0 & Mira (M6IIIE-M9.5E) & 12 & 4.1 $10^{-7}$ & 375.1 & 0.27 \\ \hline \end{tabular}} \end{table*} \normalsize We present here spectroscopic measurements of the 4 Stokes parameters (see Fig. 1). The observations were made with the IF polarimeter installed at the IRAM 30m telescope on Pico Veleta, Spain (Thum \etal 2003). Simultaneous measurements of I, U, Q, V allow us to calculate I, $p_L$, $p_C$ and $\chi$ for each velocity channel. The polarization angle calibration (i.e. the sign of Stokes U) was verified by observations of the Crab Nebula. Moreover, planets (polarization of planets is negligible at our frequency) have been used to check the instrumental polarization on the optical axis. The instrumental beam polarization is known to be stronger in Stokes Q and U than in Stokes V (known to be $\leq$ 2-3\%, see Thum \etal 2003, comparable to our sensitivity as stated elsewhere). If some detections from sources with weak p$_L$ are from a bad or uncertain pointing, they naturally induce a value of p$_C$ which is weaker than p$_L$. A strong instrumental polarization in Stokes V would be rather due to a bad phase tracking (the IF polarimeter works in a manner quite similar to that of an adding interferometer, and good phase tracking is essential). From several tests (Thum \etal 2003, Wiesemeyer, Thum \& Walmsley 2004), we know that polarization seen for weak SiO components with (Q,U,V)= (+ - -), (- - +) or (- + -) is instrumental polarization. We see that signature for only 3 objects (R Crt, R UMi and RT Vir). Some instrumental polarization may thus contaminate the observations of these objects. All instrumental parameters were carefully calibrated through specific procedures described in Thum \etal (2003). The error on $p_{L,C}$ is $\leq 2-3$ \%. SiO (v=1, J=2-1) line observations at 86.243442 GHz were carried out towards 57 stars in August and November 1999 with the IRAM 30m radiotelescope. The pointing was regularly checked directly on the star itself (for the vast majority of objects). In order to obtain flat baselines, we used the wobbler switching mode. The system temperature of the SIS receiver ranged from 110 to 170 K. The front-ends were the facility receivers A100 and B100, and the back-end was the autocorrelator. The lines were observed with a spectral resolution of 0.3 \kms. The integration times were 4-10 minutes using the wobbler switching. The forward and main beam efficiencies were respectively 0.92 and 0.77 at 3 mm. (Additional SiO (v=1, J=5-4) line observations at 215.596 GHz were also performed in most stars studied here; results will be reported elsewhere.) Our source sample (see Table 1) consists of 43 Miras, 7 Semi-Regular stars (hereafter {\em SR}), 2 IR late-type stars, 1 irregular variable, 3 supergiants and 1 Planetary Nebula (QX Pup) selected from our SiO maser master catalogue (Herpin \& Baudry, private communication). Coordinates and the main characteristics of the objects are given in Table 1. Nearly 60 \% of stars in this table have been observed with the HIPPARCOS satellite and have thus excellent optical positions; such positions have been adopted in our work. \scriptsize \begin{table} [htb] \caption{ \label{table} Derived parameters of the different components of the SiO maser emission profile for each star. Only the well identified components are given (distinct peak or strong wing emission separated from the bulk emission). Note that the polarization is fractional. The $\delta P$ is the rms derived from the $p_C$ plot.} {\begin{tabular}{l|c|c|c|c|c|c} \hline {\bf Source} & v$_{LSR}$ & F$_{\nu}$ & p$_{c}$ & p$_{L}$ & $\delta$p & $\chi$ \\ & [\kms] & [Jy] & & & & [$^{\circ}$] \\ \hline \hline {\bf 18055-1433} & 180.6 & 1.02 & -0.10 & 0.11 & $ 0.01$ & 50 \\ & 182.7 & 1.7 & -0.30 & 0.40 & $ 0.01$ & 30 \\ \hline {\bf 18158-1527} & 15.2 & 2.28 & 0.08 & 0.16 & 0.02 & 90 \\ & 17.5 & 1.86 & -0.10 & 0.20 & 0.02 & 50 \\ & 21.7 & 1.74 & 0.11 & 0.15 & 0.02 & 170 \\ & 24.5 & 0.72 & 0.43 & 0.49 & 0.02 & 46 \\ \hline {\bf 18204-1344} & 38.3 & 20.94 & 0.0 & 0.06 & $ 0.01$ & 170 \\ & 41.5 & 33.18 & 0.03 & 0.06 & $ 0.01$ &150 \\ & 45.0 & 28.26 & 0.0 & 0.02 & $ 0.01$ &120 \\ & 49.0 & 3.90& 0.04 & 0.08 & $ 0.01$ &110 \\ & 53.7 & 8.34 & $\pm$0.07 & 0.06 & $ 0.01$ &120 \\ \hline {\bf W And} & -38.0 & 4.98 & 0.08 & 0.20 & 0.01 & 70 \\ & -35.9 & 16.32 & -0.01 & 0.02 & 0.01 & 30 \\ & -34.0 & 9.72 & -0.07 & 0.16 & 0.01 & 175 \\ & -32.3 & 2.52 & 0.12 & 0.28 & 0.02 & 140 \\ \hline {\bf AU Aur} & 3.4 & 3.72 & -0.09 & 0.17 & $ 0.01$ & 71 \\ & 5.3 & 10.62 & 0.07 & 0.18 & $ 0.01$ & 165 \\ & 7.9 & 27.12 & $\pm$0.02 & 0.09 & $ 0.01$ & 60 \\ & 10.4 & 8.70 & 0.09 & 0.13 & $ 0.01$ & 110 \\ \hline {\bf NV Aur} & -1.2 & 4.68 & -0.10 & 0.15 & $ 0.01$ & 140 \\ & 1.7 & 17.94 & -0.06 & 0.13 & $ 0.01$ & 150 \\ & 3.0 & 14.04 & -0.04 & 0.08 & $ 0.01$ & 170 \\ \hline {\bf R Aur} & -7.5 & 12.48 & 0.0 & 0.0 & $ 0.01$ & 60 \\ & -5.6 & 31.56 & -0.09 & 0.27 & $ 0.01$ & 90 \\ & -3.7 & 36.00 & -0.13 & 0.34 & $ 0.01$ & 85 \\ & -1.2 & 38.58 & -0.07 & 0.16 & $ 0.01$ & 90 \\ & -0.1 & 13.14 & -0.12 & 0.27 & $ 0.01$ & 90 \\ \hline {\bf RU Aur} & -38.2 & 1.85 & 0.08 & 0.15 & 0.02 & 120 \\ & -34.9 & 1.05 & -0.12 & 0.40 & 0.02 & 85 \\ & -33.2 & 0.25 & -0.22 & 0.40 & 0.02 & 40 \\ & -27.1 & 0.2 & 0.0 & 0.35 & 0.02 & 175 \\ \hline {\bf TX Cam} & 5.0 & 11.52 & -0.06 & 0.16 & $ 0.02$ & 130 \\ & 6.1 & 13.68 & -0.03 & 0.08 & $ 0.02$ & 80 \\ & 8.0 & 23.46 & 0.04 & 0.20 & 0.01 & 175 \\ & 10.1 & 117.06 & -0.01 & 0.17 & 0.01 & 175 \\ & 11.3 & 26.46 & 0.03 & 0.06 & 0.01 & 170 \\ & 13.2 & 22.62 & -0.03 & 0.05 & $ 0.02$ & 50 \\ & 14.9 & 12.78 & 0.02 & 0.18 & $ 0.02$ & 185 \\ \hline {\bf V Cam} & 3.8 & 3.12 & -0.04 & 0.22 & 0.01 & 0 \\ & 7.6 & 12.72 & 0.0 & 0.04 & 0.01 & 160 \\ & 9.0 & 9.24 & 0.02 & 0.05 & 0.01 & 0 \\ \hline {\bf R Cnc} & 9.2 & 2.70 & 0.12 & 0.26 & $ 0.01$ & 100 \\ & 10.3 & 5.16 & -0.02 & 0.04 & $ 0.01$ & 160 \\ & 13.7 & 56.46 & 0.02 & 0.17 & $ 0.01$ & 90 \\ & 15.8 & 33.66 & -0.01 & 0.06 & $ 0.01$ & 180 \\ \hline {\bf W Cnc} & 33.9 & 21.18 & 0.13 & 0.32 & 0.01 & 40 \\ & 41.9 & 2.82 & -0.03 & 0.12 & 0.01 & 150 \\ \hline \end{tabular}} \end{table} \normalsize \addtocounter{table}{-1} \scriptsize \begin{table} [htb] \caption{ \label{table} (-continued). Derived parameters of the different components of the SiO maser emission profile for each star. Only the well identified components are given (distinct peak or strong wing emission separated from the bulk emission). Note that the polarization is fractional. The $\delta P$ is the rms derived from the $p_C$ plot.} {\begin{tabular}{l|c|c|c|c|c|c} \hline {\bf Source} & v$_{LSR}$ & F$_{\nu}$ & p$_{c}$ & p$_{L}$ & $\delta$p & $\chi$ \\ & [\kms] & [Jy] & & & & [$^{\circ}$] \\ \hline \hline {\bf VY CMa} & 5.2 & 553.4 & -0.01 & 0.02 & $ 0.05$ & 125 \\ & 8.5 & 973.3 & 0.0 & 0.01 & $ 0.05$ & 160 \\ & 11.3 & 1086.7 & 0.04 & 0.06 & $ 0.05$ & 110 \\ & 14.9 & 559.0 & -0.02 & 0.03 & $ 0.05$ & 80 \\ & 17.6 & 553.0 & -0.03 & 0.08 & $ 0.05$ & 160 \\ & 19.5 & 769.2 & 0.0 & 0.0 & $ 0.05$ & 160 \\ & 22.7 & 1569.0 & -0.04 & 0.13 & $ 0.05$ & 0 \\ & 38.7 & 207.3 & -0.01 & 0.01 & $ 0.05$ & 150 \\ \hline {\bf S CMi} & 50.1 & 3.61 & 0.07 & 0.36 & 0.01 & 50 \\ & 52.1 & 11.42 & 0.03 & 0.16 & 0.01 & 50 \\ & 54.9 & 3.89 & 0.0 & 0.0 & 0.02 & 75 \\ \hline {\bf R Cas} & 25.4 & 600.0 & 0.06 & 0.20 & $ 0.01$ & 25 \\ & 27.1 & 780.1 & 0.03 & 0.09 & $ 0.01$ & 70 \\ & 28.4 & 419.6 & -0.02 & 0.08 & $ 0.01$ & 90 \\ \hline {\bf S Cas} & -32.5 & 1.38 & 0.0 & 0.0 & 0.01 & 85 \\ & -30.0 & 11.52 & -0.32 & 0.52 & 0.01 & 110 \\ & -28.3 & 2.94 & 0.12 & 0.17 & 0.01 & 20 \\ & -27.2 & 4.98 & 0.05 & 0.05 & 0.01 & 20 \\ \hline {\bf T Cas} & -11.2 & 1.62 & 0.11 & 0.27 & $ 0.01$ & 110 \\ & -8.7 & 11.70 & -0.02 & 0.18 & $ 0.01$ & 100 \\ & -5.1 & 5.82 & 0.08 & 0.12 & $ 0.01$ & 140 \\ \hline {\bf T Cep} & -2.5 & 63.00 & $\pm$0.03 & 0.05 & $ 0.01$ & 100 \\ &-0.5 & 117.42 & 0.04 & 0.16 & $ 0.01$ & 140 \\ & 0.8 & 36.00& -0.04 & 0.05 & $ 0.01$ & 105\\ \hline {\bf R Com} & -4.4 & 7.50 & -0.04 & 0.10 & $ 0.01$ & 80 \\ &-3.3 & 8.82 & 0.07 & 0.27 & $ 0.01$ & 20 \\ & -1.4 & 3.42 & 0.02 & 0.35 & $ 0.01$ & 50 \\ \hline {\bf S Crb} & 0.8 & 41.40 & 0.07 & 0.17 & $ 0.01$ & 75 \\ & 2.0 & 64.44 & 0.0 & 0.22 & $ 0.01$ & 70 \\ & 4.9 & 28.62 & 0.0 & 0.35 & $ 0.01$ & 160 \\ \hline {\bf R Crt} & 4.5 & 4.02 & -0.07 & 0.13 & 0.01 & 100 \\ & 10.4 & 36.84 & 0.0 & 0.0 & 0.01 & 50 \\ & 16.4 & 4.74 & -0.08 & 0.15 & 0.01 & 120 \\ \hline {\bf $\chi$ Cyg} & 7.1 & 50.16 & -0.19 & 0.42 & $ 0.01$ & 75 \\ & 10.0 & 206.1 & 0.11 & 0.32 & $ 0.01$ & 140 \\ & 14.6 & 36.72 & 0.0 & 0.0 & $ 0.01$ & 190 \\ \hline {\bf UX Cyg} & -1.3 & 9.12 & -0.04 & 0.08 & $ 0.01$ & 100 \\ & -0.5 & 21.96 & 0.01 & 0.06 & $ 0.01$ & 110 \\ & 0.8 & 44.22 & 0.02 & 0.06 & $ 0.01$ & 175 \\ \hline {\bf R Hya} & -11.2 & 31.14 & 0.09 & 0.34 & 0.01 & 135 \\ & -10.0 & 67.08 & 0.10 & 0.34 & 0.01 & 140 \\ & -8.9 & 182.82 & 0.02 & 0.06 & 0.01 & 150 \\ & -6.6 & 146.22 & 0.05 & 0.25 & 0.01 & 130 \\ \hline {\bf W Hya} & 37.8 & 151.20 & 0.02 & 0.07 & 0.005 & 70 \\ & 41.4 & 934.38 & $\pm$0.02 & 0.03 & 0.005 & 150 \\ & 44.8 & 202.02 & 0.06 & 0.15 & 0.005 & 100 \\ \hline {\bf X Hya} & 21.8 & 4.26 & 0.01 & 0.70 & 0.01 & 80 \\ & 23.6 & 3.18 & 0.03 & 0.25 & 0.01 & 60 \\ & 28.9 & 23.28 & -0.01 & 0.08 & 0.01 & 170 \\ \hline {\bf R Leo} & -1.4 & 195.30 & 0.10 & 0.30 & $ 0.01$ & 140 \\ & 0.5 & 1110.6 & -0.09 & 0.36 & $ 0.01$ & 80 \\ \hline \end{tabular}} \end{table} \normalsize \addtocounter{table}{-1} \scriptsize \begin{table} [htb] \caption{ \label{table} (-continued). Derived parameters of the different components of the SiO maser emission profile for each star. Only the well identified components are given (distinct peak or strong wing emission separated from the bulk emission). Note that the polarization is fractional. The $\delta P$ is the rms derived from the $p_C$ plot.} {\begin{tabular}{l|c|c|c|c|c|c} \hline {\bf Source} & v$_{LSR}$ & F$_{\nu}$ & p$_{c}$ & p$_{L}$ & $\delta$p & $\chi$ \\ & [\kms] & [Jy] & & & & [$^{\circ}$] \\ \hline \hline {\bf W Leo} & 42.7 & 2.04 & 0.0 & 0.18 & 0.05 & 170 \\ & 44.6 & 2.76 & 0.0 & 0.10 & 0.05 & 100 \\ & 47.9 & 3.30 & 0.0 & 0.20 & 0.05 & 170 \\ & 54.2 & 4.62 & 0.08 & 0.42 & 0.02 & 35 \\ \hline {\bf R LMi} & 0.3 & 61.98 & 0.0 & 0.02 & 0.01 & 25 \\ & 2.2 & 52.80& 0.12 & 0.36 & 0.01 & 0 \\ & 3.5 & 34. 62 & 0.06 & 0.20 & 0.01 & 15 \\ & 4.5 & 14.16 & 0.04 & 0.08 & 0.01 & 20 \\ & 6.2 & 10.56 & 0.05 & 0.20 & 0.01 & 80 \\ \hline {\bf T Lep} & -32.0 & 15.48 & -0.11 & 0.34 & $ 0.01$ & 10 \\ & -30.1 & 10.74 & -0.06 & 0.20 & $ 0.01$ & 20 \\ & -27.1 & 42.48 & 0.03 & 0.10 & $ 0.01$ & 95 \\ & -25.4 & 7.80 & 0.10 & 0.14 & $ 0.01$ & 75 \\ \hline {\bf RS Lib} & 3.2 & 7.62 & 0.01 & 0.08 & 0.01 & 150 \\ & 6.5 & 35.94 & -0.10 & 0.15 & 0.01 & 150 \\ & 9.1 & 12.24 & -0.06 & 0.20 & 0.01 & 150 \\ \hline {\bf Ap Lyn} & -25.1 & 9.90 & 0.02 & 0.37 & $ 0.01$ & 185 \\ & -23.0 & 36.36 & 0.12 & 0.65 & $ 0.01$ & 165 \\ & -20.9 & 9.24 & 0.03 & 0.20 & $ 0.01$ & 150 \\ \hline {\bf U Lyn} & -14.0 & 12.00 & -0.02 & 0.04 & 0.01 & 45 \\ & -11.7 & 30.78 & 0.0 & 0.08 & 0.01 & 40 \\ & -4.5 & 7.19 & -0.13 & 0.24 & 0.01 & 80 \\ \hline {\bf GX Mon} & -10.4 & 29.64 & 0.12 & 0.38 & $ 0.01$ & 100 \\ & -8.5 & 21.60 & -0.03 & 0.08 & $ 0.01$ & 20 \\ & -6.4 & 10.38 & 0.07 & 0.20 & $ 0.01$ & 120 \\ & -4.9 & 7.92 & -0.02 & 0.07 & $ 0.01$ & 160 \\ & -3.0 & 4.98 & 0.01 & 0.02 & $ 0.01$ & 55 \\ \hline {\bf SY Mon} & -59.6 & 3.01 & 0.12 & 0.58 & 0.02 & 75 \\ & -56.0 & 2.52 & -0.05 & 0.15 & 0.02 & 30 \\ \hline {\bf V Mon} & 2.0 & 8.58 & 0.05 & 0.10 & 0.02 & 170 \\ & 4.8 & 3.12 & 0.05 & 0.14 & 0.02 & 80 \\ \hline {\bf U Ori} & -42.4 & 12.42 & 0.05 & 0.15 & 0.01 & 80 \\ & -39.9 & 46.26 & -0.04 & 0.16 & 0.01 & 50 \\ & -37.8 & 13.08 & 0.02 & 0.06 & 0.01 & 145 \\ & -34.2 & 32.40 & 0.13 & 0.40 & 0.01 & 170 \\ \hline {\bf RR Per} & 5.1 & 10.81 & 0.03 & 0.14 & 0.01 & 165 \\ & 7.5 & 29.46 & 0.11 & 0.43 & 0.01 & 175 \\ & 8.6 & 27.78 & 0.03 & 0.30 & 0.01 & 185 \\ & 11.1 & 13.02 & 0.10 & 0.34 & 0.01 & 180 \\ \hline {\bf S Per} & -47.4 & 7.02 & -0.03 & 0.03 & 0.02 & 75 \\ & -44.9 & 31.62 & 0.15 & 0.04 & 0.05 & 190 \\ & -43.3 & 35.70 & 0.01 & 0.04 & 0.05 & 190 \\ & -39.2 & 71.16& 0.0 & 0.03 & 0.05 & 150 \\ & -36.0 & 37.79 & $\pm$0.03 & 0.06 & 0.02 & 80 \\ \hline {\bf QX Pup} & 27.3 & 11.22 & 0.02 & 0.15 & $ 0.01$ & 120 \\ & 29.4 & 6.84 & 0.0 & 0.03 & $ 0.01$ & 70 \\ & 35.3 & 9.12 & 0.17 & 0.41 & $ 0.01$ & 120 \\ & 41.6 & 2.70 & 0.04 & 0.07 & $ 0.01$ & 10 \\ \hline {\bf Z Pup} & 3.9 & 40.98 & 0.07 & 0.23 & 0.01 & 140 \\ \hline {\bf VX Sgr} & -6.1 & 25.98 & 0.0 & 0.0 & $ 0.005$ & 140 \\ & 0.9 & 89.16 & 0.01 & 0.02 & $ 0.005$ & 130 \\ & 3.4 & 122.39 & -0.02 & 0.05 & $ 0.005$ & 10 \\ & 5.7 & 178.38 & -0.01 & 0.04 & $ 0.005$ & 10 \\ & 10.1 & 114.61 & -0.03 & 0.10 & $ 0.005$ & 145 \\ & 14.1 & 53.34 & 0.01 & 0.01 & $ 0.005$ & 145 \\ & 16.2 & 80.04 & 0.01 & 0.02 & $ 0.005$ & 160 \\ \hline \end{tabular}} \end{table} \normalsize \addtocounter{table}{-1} \scriptsize \begin{table} [htb] \caption{ \label{table} (-continued). Derived parameters of the different components of the SiO maser emission profile for each star. Only the well identified components are given (distinct peak or strong wing emission separated from the bulk emission). Note that the polarization is fractional. The $\delta P$ is the rms derived from the $p_C$ plot.} {\begin{tabular}{l|c|c|c|c|c|c} \hline {\bf Source} & v$_{LSR}$ & F$_{\nu}$ & p$_{c}$ & p$_{L}$ & $\delta$p & $\chi$ \\ & [\kms] & [Jy] & & & & [$^{\circ}$] \\ \hline \hline {\bf AH Sco} & -11.4 & 26.69 & 0.05 & 0.06 & $ 0.005$ &120 \\ & -10.0 & 48.31 & 0.01 & 0.02 & $ 0.005$ &120 \\ & -7.2 & 78.78 & 0.0 & 0.0 & $ 0.005$ &150 \\ & -4.7 & 54.74 & 0.0 & 0.02 & $ 0.005$ &20 \\ & -2.8 & 32.99 & -0.02 & 0.05 & $ 0.005$ &170 \\ \hline {\bf RR Sco} & -33.9 & 4.02 & 0.10 & 0.14 & 0.02 & 65 \\ & -30.2 & 9.48 & 0.0 & $\pm$0.07 & 0.02 & 85 \\ & -26.5 & 11.16 & 0.04 & 0.10 & 0.02 & 90 \\ & -23.2 & 2.82 & 0.06 & 0.16 & 0.02 & 55 \\ \hline {\bf R Ser} & 27.8 & 21.61 & -0.07 & 0.30 & $ 0.01$ & 40 \\ \hline {\bf S Ser} & 18.7 & 11.22 & 0.04 & 0.09 & $ 0.01$ & 180 \\ & 20.9 & 16.92 & 0.0 & 0.06 & $ 0.01$ & 100 \\ \hline {\bf WX Ser} & 4.1 & 20.34 & -0.05 & 0.15 & $ 0.01$ & 140 \\ & 6.6 & 12.18 & 0.04 & 0.15 & $ 0.01$ & 80 \\ & 9.8 & 8.41 & -0.12 & 0.30 & $ 0.01$ & 140 \\ \hline {\bf IK Tau} & 31.6 & 46.74 & 0.04 & 0.14 & 0.01 & 180 \\ & 34.7 & 288.24 & 0.05 & 0.16 & 0.01 & 170 \\ & 36.8 & 53.04 & 0.07 & 0.18 & 0.01 & 140 \\ & 39.6 & 27.84 & 0.13 & 0.36 & 0.01 & 125 \\ \hline {\bf R Tau} & 9.5 & 3.78 & -0.10 & 0.15 & $ 0.01$ & 60 \\ & 11.8 & 19.51 & 0.06 & 0.13 & $ 0.01$ & 150 \\ & 13.1 & 10.26 & 0.07 & 0.19 & $ 0.01$ & 140 \\ & 14.7 & 7.68 & -0.09 & 0.31 & $ 0.01$ & 55 \\ & 16.4 & 6.24& -0.08 & 0.19 & $ 0.01$ & 75 \\ & 19.0 & 3.24 & -0.20 & 0.44 & $ 0.01$ & 55 \\ \hline {\bf RX Tau} & -44.0 & 3.36 & -0.06 & 0.10 & 0.02 & 120 \\ & -41.0 & 10.98 & -0.01 & 0.01 & 0.02 & 100 \\ & -38.9 & 2.94 & 0.04 & 0.15 & 0.02 & 170 \\ \hline {\bf R Tri} & 56.7 & 13.62 & -0.05 & 0.24 & $ 0.01$ & 105 \\ \hline {\bf R UMi} & -6.1 & 1.98 & 0.0 & 0.25 & 0.05 & 110 \\ \hline {\bf S UMi} & -44.4 & 21.54 & 0.0 & 0.0 & 0.005 & 30 \\ & -43.3 & 30.66 & 0.0 & 0.01 & 0.005 & 20 \\ & -40.8 & 24.24 & -0.02 & 0.10 & 0.005 & 10 \\ & -40.0 & 27.66 & 0.02 & 0.07 & 0.005 & 10 \\ \hline {\bf R Vir} & -26.8 & 0.84 & 0.15 & 0.30 & 0.05 & 170 \\ & -25.7 & 2.16 & -0.10 & 0.15 & 0.02 & 10 \\ \hline {\bf RS Vir} & -15.7& 2.64 & -0.11 & 0.18 & 0.01 & 0 \\ & -13.3 & 10.62 & -0.08 & 0.32 & 0.01 & 150 \\ & -12.5 & 8.46 & 0.0 & 0.20 & 0.01 & 0 \\ & -9.1 & 7.08 & -0.06 & 0.52 & 0.01 & 160 \\ \hline {\bf RT Vir} & 9.1 & 1.20 & 0.0 & 0.02 & $ 0.02$ & 70 \\ & 13.8 & 2.22 & 0.12 & 0.19 & $ 0.02$ & 20 \\ & 15.7 & 3.12 & 0.12 & 0.20 & $ 0.02$ & 140 \\ & 18.0 & 7.98 & 0.0 & 0.06 & $ 0.02$ & 130 \\ & 20.7 & 3.48 & -0.10 & 0.12 & $ 0.02$ & 120 \\ & 22.6 & 3.03 & -0.06 & 0.11 & $ 0.02$ & 60 \\ & 27.7 & 2.34 & -0.04 & 0.10 & $ 0.02$ & 40 \\ \hline {\bf S Vir} & 10.3 & 13.62 & -0.04 & 0.27 & $ 0.01$ & 180 \\ & 12.2 & 7.79 & 0.03 & 0.06 & $ 0.01$ & 105 \\ & 13.7 & 22.52 & -0.05 & 0.19 & $ 0.01$ & 130 \\ \hline \end{tabular}} \end{table} \normalsize \section{Results: Polarization Study} \subsection{Individual results} Values of the polarization level presented here (see Table 2) are those measured for the different components within the SiO maser emission profile for each star. Examples are given in Fig.2 for a few stars. The complete Figure 2 with all the observations is available in electronic form at http://www.edpscience.org. Only the well identified components are considered (distinct peaks or strong wing emission well separated from the bulk emission, according to the noise). Some interesting cases are briefly presented below. Some profiles show isolated emission red$/$blue-shifted from the main emission which are more strongly circularly polarized (e.g. IRAS 18204-1344). These peculiar characteristics imply a different spatial origin for the main and higher/lower velocity components. Sometimes the circular polarization is regularly varying across the profile (e.g. R Leo), but sometimes not. In T Lep, the SiO emission shows two peaks linked by a plateau; the circular polarization is linearly varying across the profile from -11 to 10 \% (see Fig. 2). Several objects (e.g. S UMi, IK Tau) show the same $p_C$ pattern. The IR late-type source IRAS18158-1527 exhibits a complex profile with several well defined components, each of them differently polarized indicating a complex maser structure with probably different maser spots contributing to the whole emission. The red wing emission is highly polarized (43 \%). Such a complex multi-component maser line profile and "semi-circle", convex, $p_C$ pattern appear to be characteristic of SR objects (see other similar objects in our sample and R Crt in Fig. 2). Nevertheless, the circular polarization pattern observed in the Mira star U Lyn is a convex profile as encountered in SR objects. One of the most studied Mira star is R Leo. The profile is made of a strong emission with a blue broad line wing. Main and linewing emissions are strongly polarized (respectively negative and positive $p_C \sim 9-10$ \%). R Leo is a very well studied object exhibiting a bipolar jet throughout its envelope. The clear symmetry observed between the positive and negative circular polarization patterns in the main and wing line emissions suggests that the maser emission comes from the jet lobes. We note that the Mira star RS Lib exhibits an emission and polarization pattern similar to that observed in R Leo. R Leo and RS Lib may have the same spatial structure. \addtocounter{figure}{-1} \subsection{Analysis} The circular polarization level in several of our objects has already been measured by Kemball \& Diamond (1997) or Barvainis, McIntosh \& Predmore (1987). For TX Cam and W Hya, our results are consistent with previous observations: \begin{itemize} \item in TX Cam, the bulk of the emission is weakly circularly polarized while its wings show $p_C \sim 3-6 \%$ in good agreement with the VLBI observations of Kemball \& Diamond in 1997 who derived an average value of $p_C \sim 3-5 \%$; \item in the Semi-Regular object W Hya, the central emission is weakly polarized ($\pm$ 2 \%), while the wings and secondary peaks show $p_C$= 2-6 \%), which is consistent with the 5 \% of Barvainis, McIntosh \& Predmore (1987). \end{itemize} On the contrary, the circular polarization level we derive in VY CMa, R Cas, R Leo, and VX Sgr is different from levels measured by Barvainis, McIntosh \& Predmore (1987), respectively 1-4, 2-6, 9-10 and less than 3 \% while they found respectively 6.5, 1.5, 2.4 and 8.7 \%. This difference is significant and may be due to variability over the fifteen intervening years. Indeed time variability of the polarization remains an open question in the field. Glenn \etal (2003) have shown that the individual maser feature lifetime ranges from a few months or less to more than 2 years, i.e. the characteristic time over which the $Q$ and $U$ spectral features persist. The average linear polarization is 23 \% in Glenn \etal sample with a typical dispersion of 7\%. Cotton et al. (2004) have comparable epoch spacing and do not conclude on the variability. Our observations were repeated at intervals of a few months (August and November 1999) and the polarization tends to remain stable between the two epochs. We emphasize that we cannot spatially distinguish with a single dish radiotelescope between the different maser spots producing the SiO profile (various masers spots contribute in the various features observed at a given velocity). The whole SiO maser emission region, hence all the maser cells, lie within the 29 arcseconds of the 30m (but not necessarily with a uniform distribution) while the SiO emission covers less than 40 milliarseconds in TX Cam (Kemball \& Diamond 1997) and thus everything is beam averaged. This means that any conclusion on the geometry of the objects observed here would be much uncertain. Only global trends or global geometry can be discussed. One of the consequences of this spatial resolution problem is that if the polarization vectors are distributed isotropically around the object, the average polarization level that we measure is zero, even if the maser emission produced in each SiO cell is well polarized. A global analysis of our data in Table 2 shows the following. We find that $p_L$ varies between 0 and 70 \%, and $p_C$ between 0 and $\pm$43 \%. Hence, polarization vectors are not distributed isotropically. Emission from Mira-type objects clearly tends to have a relatively high linear ( ${<p_L>}_{Mira} \simeq 30$\%, ${<p_L>}_{SR} \simeq 11$\%) and circular polarization (${<p_C>}_{Mira} \simeq 9$\%, ${<p_C>}_{SR} \simeq 5$\%). Note that the emission from the PN QX Pup is highly polarized, and, on the contrary, maser emission from supergiants shows very weak polarization ($<p_L>=5$\%, $<p_C>=2$\%), with the exception of one maser component in S Per. Moreover, all observations show that the polarization level varies across the maser line profile (see Fig. 2), i.e. the different spectral components of the maser emission producing the profile are coming from different localizations in the SiO shell and have different polarization levels. The highest polarization level for one object can be encountered either in the main peak, or in the other components. Semi-Regular objects (RU Aur, R Crt, W Hya, S Per, AH Sco, R UMi, RT Vir) have a common circular polarization pattern with the central main emission unpolarized and other peak emission or wings being strongly polarized: a characteristic "semi-circle" (i.e. convex shape) pattern for $p_C$ is observed (see R Crt in Fig. 2). The infrared late-type star IRAS18158-1527 exhibits a similar pattern, thus suggesting that this star is a semi-regular. A group of objects (W And, NV Aur, T Cas, R Com, T Lep, IK Tau, S Ser, S UMi) shows approximately the same $p_C$ pattern (see T Lep in Fig. 2); the circular polarization varies linearly across the line profile from a positive value to a negative one (or the contrary). The only common spectral characteristic of the SiO emission from these stars is the presence of an plateau-like emission on top of which the narrow emission peaks are located. \section{Discussion} In this section, we first discuss our source sample in the frame of the 2-color diagram. Then, we briefly summarize the existing SiO maser polarization theories. Finally, we discuss our data set in this context and estimate the stellar magnetic field strength. \subsection{2-color diagram} Stars of our sample can be plotted in a [12]-[25], [25]-[60] color-color diagram (van der Veen \& Habing 1988; [12], [25] and [60] stand respectively for 12, 25 and 60 microns IRAS-fluxes). This diagram is partitioned into several regions (see Fig. 3) defined by van der Veen \& Habing as follows: Region I, oxygen-rich non variable stars without circumstellar shells; Region II, variable stars with young O-rich circumstellar shells; Region IIIa, variable stars with more evolved O-rich circumstellar shells; Region IIIb, variable stars with thick O-rich circumstellar shells; Region IV, variable stars with very thick O-rich circumstellar shells; Region V, Planetary Nebulae and non-variable stars with very cool envelopes; Region VIa, non variable stars with relatively cold dust at large distance; Region VIb, variable stars with relatively hot dust close to the star and relatively cold dust at large distance; Region VII, variable stars with more evolved C-rich circumstellar shells. On Fig. 3 are represented the linear and circular polarization level for the main SiO emission component from each star. Most of the objects in our sample fall in regions II and IIIa and do not show particular characteristics, except for S Cas (an S-type star) where the circular polarization is high. Mira-type stars are in regions I, II, IIIa and VII. IR late-type objects are in VIb and VII. The SRa semi-regular variable W Hya is in I, while the SRb stars lie in IIIa (RU Aur, RT Vir, R Crt), II (R UMi), and Src in IIIa (S Per) and VII (AH Sco). The Red Supergiants, VY CMa and VX Sgr, and the IR supergiant IRAS 18204-1344 lie in VII. Objects in Region VII do not exhibit strong polarization compared to other objects, perhaps because of their more C-rich circumstellar shells (e.g. AU Aur) or because of the presence of hot dust close to the star implying less SiO abundance and thus weaker emission, making the polarization measurement less significant. The presence of hot dust may also influence the pumping of the SiO molecules and thus the polarization level; the optically thick, hence isotropic, radiation field of hot dust can assist the collisional pumping. This could apply to UX Cyg (an irregular variable), IRAS 18055-1433 and IRAS 18158-1527 in region VIb. QX Pup in region V is a PN and exhibits strong polarized emission. Note that IRAS 18055-1433 and IRAS 18158-1527 show very strong circular polarization in their line wings. We may conjecture here that wing emission comes from more outer layers than those where the main line is excited (Herpin \etal 1998); as a consequence, the SiO cells giving rise to wing emission are less influenced by the presence of hot dust (hot dust preferably lies in the inner layers). \subsection{The SiO maser polarization theory} Since SiO is non-paramagnetic, the Zeeman splitting $g\nu_B$ ($g$ is the LandЋ factor) is much less than the Doppler width. Moreover, the degree of saturation is the ratio of the rate $R$ for stimulated emission to the loss rate $\Gamma$ (usually $\Gamma$ is approximated by the inverse radiation lifetime for a vibrational transition, $\Gamma \simeq 5$ s$^{-1}$ for SiO masers, Wiebe \& Watson 1998). Hence, if $R\geq \Gamma$, the maser is saturated. In fact, in the Orion case Plambeck \etal (2003) show that, despite the radiation beam angle is unknown, the 86 GHz SiO maser is saturated. The maser is saturated if the angle averaged intensity $J= \frac {I \Omega_b} {4\pi}$ ($\Omega_b$ is the beaming solid angle) is larger than the saturation intensity $J_S$; $J_S$ is a theoretical quantity. The saturation depends on the angle into which the radiation is beamed, but this angle is unknown (Watson \& Wyld 2001), thus $J$ cannot be directly measured (even if $I$ is measurable when the maser is resolved, the beaming angle $\Omega$ is not an observable). For more than one decade, two schools have come up against each other to explain SiO maser emission. SiO polarization theory is described in: (i) Watson (e.g. Watson \& Wyld 2001, Wiebe \& Watson 1998, Nedoluha \& Watson 1994); (ii) Elitzur (2002, 1998, 1996, 1994). The main difference between the two approaches rests in the pumping mechanisms. While anisotropic pumping associated with a weak field produces high $p_L$ and quite significant $p_C$ in Watson's model, a strong magnetic field is necessary with the more classical pumping mechanisms used in Elitzur's model. Details about both models can be found in the Elitzur's review (2002). We may summarize the main characteristics of Watson's model as follows: \begin{itemize} \item non-Zeeman effect; \item anisotropic pumping; \item no direct relation between $p_{C}$ and $B$. An estimation of B can only be derived through complete calculation of the radiative transfer. Nevertheless, when maser saturation is not important, the "thermal" spectral line equation $\frac {V}{\delta I/\delta v}=\alpha B \cos \theta$ is applicable (Fiebig \& G\"usten 1989); $I$ is the intensity with respect to Doppler velocity $v$, $\theta$ is the angle between $B$ and the line of sight, $\alpha$ is a constant; \item the Zeeman splitting parameter $g\Omega$ (in frequency units $g\Omega = 1.5 B[mG] s^{-1}$ for SiO masers) is $\simeq R$; \item saturated or unsaturated maser (saturated maser increases $p_C$); \item linear correlation between $p_L$ and $p_C$, and high $p_L$ is needed; \item intensity dependent circular polarization; \item $B$ of a few 10 mG varying as $r^{-2,3}$ throughout the envelope. \end{itemize} In contrast with Watson's work, Elitzur's model is based on the Zeeman effect and the exponential maser growth in the unsaturated phase; the polarization characteristics are preserved as the radiation is amplified into the saturated regime. This model was improved several times (Elitzur 1994, 1996, 1998) and takes into account the anisotropic pumping. The magnetic field generates circular polarization and the main pumping mechanism for the SiO maser is a "classical" radiative-collisional process. For saturated masers, a direct relation between $p_C$ and $B$ is obtained from simple calculations. The ratio $x_B$ of the Zeeman splitting $\Delta \nu_B$ to the Doppler linewidth $\Delta \nu_D$, can be determined (Elitzur 1996) from $v_{peak}$, the ratio of the Stokes parameter $V/I$ at a given peak feature: \begin{equation} \label{ } x_B=\frac {3\sqrt{2}}{16} ~ v_{peak} \cos \theta \end{equation} Following Barvainis, Mc Intosh \& Predmore (1987) and Elitzur (1996) we arbitrarily take $\theta \simeq 45^{\circ}$ \begin{equation} \label{ } \Rightarrow x_B= \frac{3}{16} ~ v_{peak} \end{equation} \begin{equation} \label{ } \Rightarrow x_B= 0.1875 ~10^{-2} ~m_C \end{equation} where $m_C$ is the polarization fraction in percentage (i.e. 100 $p_C$). Moreover: \begin{equation} \label{ } x_B= 14 g \lambda \frac {B}{\Delta v_D} \end{equation} where $g$ is the Land\'e factor with respect to the Bohr magneton, $\lambda$ the transition wavelength in cm, $B$ the field in Gauss and $\Delta v_D$ the Doppler width in km$s^{-1}$. Thus, \begin{equation} \label{ } B= 0.1875 ~10^{-2} ~\frac {\Delta v_D}{14 g \lambda} ~m_C \end{equation} With $g\simeq 10^{-3}$, $\Delta v_D =1$ \kms and $\lambda=0.2877$ cm, we derive: \begin{equation} \label{ } B\simeq 0.46 ~m_C \end{equation} This model predicts that there is no correlation between $p_C$ and $p_L $ (see also Watson \& Wyld 2001); such a mechanism leads to an inferred magnetic field of a few Gauss to 10 Gauss for the SiO maser zone, the field hence varying as $r^{-1,2}$ across the envelope. \subsection{Theory against present observations} The relevance of the two different polarization theories can be assessed via a few observational checks: \begin{itemize} \item dependence of circular polarization on intensity; \item linear correlation between $p_C$ and $p_L$; \item Zeeman effect, i.e. spectral shape of the Stokes parameter $V$; \item coherence of B strength values inferred from OH and \water observations. \end{itemize} As we do not resolve the maser emission into individual spatial components, our current data set is biased by the beam averaging. Therefore, some effects cannot be tested with our data. In the Zeeman effect case, the spectral shape of the Stokes parameter $V$ must follow an antisymmetric {\em S} curve with sharp reversal at line center (Elitzur 1996). Unfortunately the doppler width is less than the resolution of the observations and we cannot conclude. We may look for any correlation between $p_C$ and $p_L$. As shown in Figs. 2 and 3, it first appears that in all cases $p_L$ is larger than $p_C$ as is predicted by all models. More precisely, $p_C$ is noticeable if and only if $p_L$ is high. If we plot the values of $p_C$ and $p_L$ derived for all maser components in this work (Fig. 4) and make a regression fit to our data we obtain $p_C \simeq 0.015 + 0.25 ~p_L$. The circular polarization level tends to vary approximately linearly with $p_L$ in agreement with the fact that no $p_C$ is detected towards sources with a marginal $p_L$ detection. Note that in Fig. 4, four objects, S Per, IRAS18055-1433, S Cas and $\chi$ Hya, do not follow the same general trend observed for the rest of the sample. (The case of the Supergiant S Per, however, is an exception as it exhibits substantial $p_C$ while $p_L<p_C$.) This observation may in fact favour Watson's model. It must be stressed again that the beam-averaged polarization that we measure makes any conclusion uncertain. In fact, due to this averaging, we should observe no correlation at all, even if such one would exist ! As mentioned earlier (see Section 3.2) we observe relatively high circular polarization rates in several stars (${<p_C>}_{Mira} \simeq 9$\%, ${<p_C>}_{SR} \simeq 5$\%). These values are larger than those predicted from Watson's model (e.g. Nedoluha \& Watson 1994). We also have not been able to find any correlation of $p_C$ with total intensity. Finally, although we adopt the Zeeman case to derive the magnetic field strength (see Section 4.5), we cannot conclude firmly from present observations which maser theory prevails for SiO emission. \subsection{Magnetic field in AGB stars} A better knowledge of the stellar magnetic field strength is crucial to understand the last stages in the life of an Asymptotic Giant Branch (hereafter AGB) star. These stages are characterized by a high mass loss process driven by the radiation pressure; they are also influenced by the magnetic field (Palen \& Fix 2000, Blackman, Frank \& Welch 2001, and references therein). A strong magnetic field may rule the mass loss geometry; in particular, it could be the cause of a higher or lower mass-loss rate in the equatorial plane (Soker 2002), and thus determine the global shaping of these objects. But, as the direct dynamical effect of the magnetic activity is much lower than that of the wind (although in local spots the magnetic field can be dominant), the role of the magnetic field might be indirect. Moreover, the observed high mass loss rates are hardly explained by a single process and need a combination of several factors such as rotation, the presence of a companion (binary stars, with our without common envelope, exhibiting mass transfer or tidal effects are common; Soker 1997) and a magnetic field. During its quick transition to the PN stage, the AGB star will completely change its geometry: the quasi-spherical object becomes axisymmetrical, point-like symmetrical or even shows higher order symmetries (Johnson \& Jones 1991, Sahai \& Trauger 1998, Balick \& Frank 2002). The classical or generalized {\em Interacting Stellar Winds} (hereafter ISW or GISW) models (Kwok 2000, Soker \& Livio 1989, Morris 1987) try to explain this shaping, but have serious difficulties in producing complicated structures with peculiar jets or ansae (e.g., CRL 2688, Delamarter 2000) and do not fully address the origin of the wind. Furthermore, recent X-ray studies with the Chandra satellite do not completely agree with GISW predictions for temperatures (Guerrero \etal 2001). Some recent studies tend to demonstrate the importance of the magnetic field in evolved objects. Bujarrabal \etal (2001) show that for 80\% of the PPNe in their sample the fast molecular flows have too high momenta to be powered by radiation pressure only (1000 times larger in some cases) what may be explained by magnetic field. Moreover, X-ray emission found in evolved stars (e.g. H\"unch, Schmitt \& Schr\"oder 1998) may indicate the presence of a hot corona that possibly results from magnetic activity. Very recently, magnetic field was discovered for the first time in central stars of PN (Jordan \etal 2005) and estimated to kiloGauss, much stronger than what we find here from our SiO data in QX Pup. New models involving the magnetic field have been developed trying to explain the morphology changes of an object during its transition from the AGB stage to the PN stage; B plays the role of a catalyst and of a collimating agent. The most simple models are based on a moderately weak magnetic field alone ($B \simeq 1$ Gauss at the stellar surface, a few $10^{13}$ cm, i.e. at a radiis of a few AU, Soker 1998). The influence of B is stressed by the work of Smith \etal (2001) and Greaves (2002) in VY CMa. But the role of B can only be decisive when its energy density is greater than the radiative pressure, i.e. when B is greater than around 10 G close to the stellar surface in the SiO region (see Soker \& Zaobi, 2002). Arguing that such a strong field may be very unusual, Balick \& Frank (2002) explain that B alone cannot produce the observed structures, and a combination of several factors has thus to be considered (rotation, magnetic field and presence of a companion). Soker \& Harpaz (1992) first proposed a model with a weak magnetic field ($\leq 1$ G) and included a slow rotation together with the presence of a companion to transform the envelope (and lead for example to the peculiar geometry observed in NGC 6826 or NGC 6543). Even if the star were not binary, the influence of B is probably important locally (Palen \& Fix 2000). A significant magnetic field can form cold spots on the star's surface and a slow rotation of the star can then increase the field strength to build up a dipolar magnetic field varying as $1/r^{3}$ (Matt \etal 2000); such a field is stronger at the equator and may thus lead to an axisymmetrical mass loss. The main argument against the dominant influence of the magnetic field on the shaping of the circumstellar envelope is that a strong field seems to be necessary to dominate the dynamics of the gas. However, several authors (Pascoli 1985, 1992, 1997, Chevalier \& Luo 1994, Garc\'{\i}a-Segura 1997, Gurzadyan 1997, Delamarter 2000) have demonstrated the strong influence of a reasonable toro\"{\i}dal magnetic field embedded in the normal radiation-driven stellar wind ({\em Magnetic Wind Bubble} theory, hereafter {\em MWB}). This field has a strength between a few Gauss and a few 10 Gauss at a few stellar radii (the SiO region is believed to be at $\sim 10^{14}$ cm or $\sim 7-10$ AU), varies as $1/r^2$, then as $1/r$ at larger radii; therefore $B\sim1$ mG at $10^{16,17}$ cm or 700-7000 AU. These results are confirmed by the simulations of Garc\'{\i}a-Segura, Lopez \& Franco (2001) for the PN He 2-90. Even if the origin of the wind is not explained by these models, it seems clear that a magnetic field is essential to generate fast collimated outflows (Kastner \etal 2003). There are many models of magnetic jet production and collimation and some, or all of them, are applicable to various star geometries. One most interesting study was performed by Blackman, Frank \& Welch (2001) in which the magnetic field emerges from the AGB stellar core and the resulting 1G field helps to collimate the radiation-driven wind or a stronger, more anisotropic, magnetically driven wind. \subsection{Magnetic field in our sample} The exact interpretation, in terms of magnetic field, of our observations depends on the adopted specific SiO maser model (see Sections 4.2 and 4.3). From the current knowledge of the strength of the magnetic field in the OH and \water layers we expect $B_{//}$ of a few Gauss at least in the SiO maser region (see also Kemball \& Diamond 1997), i.e. at 5-10 AU from the central object. This tends to invalidate Watson's model, and furthermore tends to agree with a field varying in $r^{-1,2}$ as predicted by Elitzur's model. However, Vlemmings \etal (2005) measured the circular polarization of the \water maser emission in a few evolved stars with the VLBA observations and showed that the magnetic field is either a solar-type field (with a $r^{-2}$ field strength dependence) or a dipole magnetic field (with a $r^{-3}$ dependence) in their sample. In the following, we decide to use Elitzur's theory (Zeeman case) to infer magnetic field strength from the circular polarization levels. From equation (12), we thus calculate the mean value of the magnetic field $B_{//}$ for each star and give results in Table 3. For our sample B$_{//}$ is between 0 and 20 Gauss, with a mean value of 3.5 G. This value combined with the strength of the field in more outer layers of the envelope (OH and \water masers) agrees with a B field variation law in $1/r$, closer to Elitzur's model. As explained in the Introduction and in Section 4.4, B alone can be the main agent to shape the circumstellar envelope if its value is larger than around 10 Gauss. This means that only S Cas, RU Aur, IRAS 18055-1433 and IRAS 18158-1527 may have a {\em magnetic field ruled geometry} ($B> 10$ G in these objects). The rest of our sample shows that B is sufficiently strong to be dominant at this stage of the AGB star evolution, but it should be associated with rotation and the presence of a companion as suggested in models mentioned in Section 4.4. Despite our B measurements are beam averaged, they suggest in many cases that they are not too much (not orders of magnitude) below the critical value; local B values may exceed in many cases the critical value, and therefore participate in the shaping of the AGB envelopes. Our estimated values of B are consistent with the {\em MWB} theory (toro\"{\i}dal magnetic field) or the model of Blackman, Frank \& Welch (2001). % \begin{table} [h] \caption{ \label{table} Average magnetic field strengths derived from $p_C$.} {\begin{tabular}{lc} \hline {\bf Source} & $B_{//}$ (G) \\ \hline IRAS18055-1433 & 4.6-13.9 \\ IRAS18158-1527 & 3.7-20.0 \\ IRAS18204-1344 & 0-3.2 \\ W And & 0.4-5.9 \\ AU Aur & 0.9-4.2 \\ NV Aur & 1.9-4.6 \\ R Aur & 0-6.0 \\ RU Aur & 0-10.2 \\ TX Cam & 0.4-2.8 \\ V Cam & 0-1.9 \\ R Cnc & 0.4-5.6 \\ W Cnc & 1.4-6.0 \\ VY CMa & 0-1.9 \\ S CMi & 0-3.2 \\ R Cas & 0.9-2.8 \\ S Cas & 0-14.9 \\ T Cas & 0.9-5.1 \\ T Cep & 1.4-1.9 \\ R Com & 0.9-3.2 \\ S Crb & 0-3.2 \\ R Crt & 0-3.7 \\ $\chi$ Cyg & 0-8.8 \\ UX Cyg & 0.4-1.9 \\ R Hya & 0.9-4.6 \\ W Hya & 0.9-2.8 \\ X Hya & 0.4-1.4 \\ R Leo & 4.2-4.6 \\ W Leo & 0-3.7 \\ R LMi & 0-5.6 \\ T Lep & 1.4-5.1 \\ RS Lib & 0.4-4.6 \\ Ap Lyn & 0.9-5.6 \\ U Lyn & 0-6.0 \\ GX Mon & 0.4-5.6 \\ SY Mon & 2.3-5.6 \\ V Mon & 2.3 \\ U Ori & 0.9-6.0 \\ RR Per & 1.4-5.1 \\ S Per & 0-7.0 \\ QX Pup & 0-7.9 \\ Z Pup & 3.2 \\ VX Sgr & 0-1.4 \\ AH Sco & 0-2.3 \\ RR Sco & 0-4.6 \\ R Ser & 3.2 \\ S Ser & 0-1.9 \\ WX Ser & 1.9-5.6 \\ IK Tau & 1.9-6.0 \\ R Tau & 2.8-9.3 \\ RX Tau & 0.4-2.8 \\ R Tri & 2.3 \\ R UMi & 0 \\ S UMi & 0-0.9 \\ R Vir & 4.6-7.0 \\ RS Vir & 0-5.1 \\ RT Vir & 0-5.6 \\ S Vir & 1.4-2.3 \\ \hline \end{tabular}} \end{table} From Elitzur (1996) and our measurements ${<p_C>}_{Mira} \simeq 9$\%, ${<p_C>}_{SR} \simeq 5$\%, we can estimate (see Eq. (9) in Sect. 4.2) that $<x_B>$ is around 0.017 and $9.4$ $10^{-3}$ respectively for Mira-type objects and semi-regular variables. Moreover, according to Elitzur (1996, see Fig.2), there is no stationary physical solutions for propagation at ${\sin}^2 \theta < \frac {1}{3}$ (i.e. at ${\sin}^2 \theta < \frac {1}{3}$ the radiation is not polarized). As $x_B$ is small ($<0.02$), from Fig.2 of Elitzur (1996) we can estimate the volume of phase space in which propagation of linear polarization in a maser is possible or not ($\theta > 35.3^{\circ}$ or $\theta < 35.3^{\circ}$); we then calculate that the probability for a random magnetic axis to be aligned with a given direction (our line of sight) is better than $35.3^{\circ}$. Our present estimate is that around 18 \%. Therefore, Elitzur's model predicts that 18.4 \% of the SiO 86 GHz masers should not be linearly polarized, because such polarized masers cannot propagate if the magnetic field, although weak, is closer than $35.3^{\circ}$ to the line of sight (propagation direction). Hence non polarized maser emissions do not imply no or weak magnetic field. In our sample, roughly 13 \% of the SiO maser components have no detectable or very weak ($<3\%$) polarization. We looked without success for a possible correlation between the polarization rates and physical parameters such as the known envelope asymmetry, the presence of SiO maser high velocity linewings (see Herpin \etal 1998), or the mass loss rate. If the magnetic field plays an important role in the shaping of the object, one may expect to find a relationship between the strength of $B_{//}$ (thus $p_C$) and the geometry of the object. Unfortunately, no trend is clearly found in our data. Nevertheless it is known that radiative pressure is driving the wind in AGB objects and it is thus not surprising that we find no correlation between the B strength and a known asymmetry in our sample. Of course, our stellar sample would require new observations with sufficient spatial resolution (VLBI) to confirm the present results; the same type of study should also be conducted toward several Proto-PN and PN objects. \section{Conclusion} We have made a study of the SiO maser polarization in a representative sample of evolved stars, simultaneously measuring, for the first time, the 4 Stokes parameters. From our measurements we derive the circular and linear polarization levels and shows that, due to the beam averaging of our polarization measurements, we cannot firmly discriminate between the two dominant theories of SiO maser emission. In particular, VLBI observations of our source sample are absolutely necessary to distinguish between Zeeman or non-Zeeman theories. Nevertheless, the magnetic field strength was derived assuming Elitzur's model. $B_{\\}$ varies between 0 and 20 Gauss, with a mean value of 3.5 G. As a consequence, we suggest that the magnetic field plays a significant role in the evolution of these objects. Within the frame of the Zeeman theory the magnetic field could shape or even collimate the gas layers surrounding the AGB objects. Emission from Mira-type objects clearly tends to have a higher linear ( ${<p_L>}_{Mira} \simeq 30$\%, ${<p_L>}_{SR} \simeq 11$\%) and circular polarization (${<p_C>}_{Mira} \simeq 9$\%, ${<p_C>}_{SR} \simeq 5$\%). Basically, if there is a real correlation between $p_C$ and the strength of the magnetic field, this trend may indicate that the magnetic field may be stronger in Mira objects than in Semi-Regular variables (at least in the inner layers of the circumstellar envelope). To better understand the mechanisms at work with the magnetic field, complementary studies have to be conducted and in particular the presence of a companion has to be investigated in a large sample of objects. Of course VLBI maps of the magnetic field in these stars are essential. Another important objective is to investigate the evolution of the magnetic field and its influence during the transition from the AGB star phase to the PN stage. \acknowledgements{The authors are grateful to M. Elitzur for reading and commenting on this paper. We also thank W.D. Watson for his useful comments and suggestions. The authors are indebted to the staff of the IRAM 30m telescope who most efficiently helped during the observations and to R. Mauersberger who closely followed part of these observations. Finally, we also thank the referee for several useful comments.} \Online \addtocounter{figure}{-3} \addtocounter{figure}{-1} \addtocounter{figure}{-1} \addtocounter{figure}{-1} \addtocounter{figure}{-1} \addtocounter{figure}{-1} \addtocounter{figure}{-1}
Title: An HLLC Solver for Relativistic Flows -- II. Magnetohydrodynamics
Abstract: An approximate Riemann solver for the equations of relativistic magnetohydrodynamics (RMHD) is derived. The HLLC solver, originally developed by Toro, Spruce and Spears, generalizes the algorithm described in a previous paper (Mignone & Bodo 2004) to the case where magnetic fields are present. The solution to the Riemann problem is approximated by two constant states bounded by two fast shocks and separated by a tangential wave. The scheme is Jacobian-free, in the sense that it avoids the expensive characteristic decomposition of the RMHD equations and it improves over the HLL scheme by restoring the missing contact wave. Multidimensional integration proceeds via the single step, corner transport upwind (CTU) method of Colella, combined with the contrained tranport (CT) algorithm to preserve divergence-free magnetic fields. The resulting numerical scheme is simple to implement, efficient and suitable for a general equation of state. The robustness of the new algorithm is validated against one and two dimensional numerical test problems.
https://export.arxiv.org/pdf/astro-ph/0601640
\date{Accepted ??. Received ??; in original form ??} \pagerange{\pageref{firstpage}--\pageref{lastpage}} \pubyear{2005} \label{firstpage} \begin{keywords} hydrodynamics - methods: numerical - relativity - shock waves \end{keywords} \section{Introduction} Strong evidence nowadays supports the general idea that relativistic plasmas may be closely related with most of the violent phenomena observed in astrophysics. Most of these scenarios are commonly believed to involve strongly magnetized plasmas around compact objects. Accretion onto super-massive black holes, for example, is invoked as the primary mechanism to power highly energetic phenomena observed in active galactic nuclei, \citep{Macchetto99,ERZ02,McK05,Shapiro05}. In this respect, the formation and propagation of relativistic jets and the accretion flow dynamics pose some of the most challenging and interesting quests in modern theoretical astrophysics. Likewise, a great deal of attention has been addressed, in the last years, to the darkling problem of gamma ray bursts \citep[see for example][]{MR94,McFW99,KG02,RRD03}, whose models often appeal to strongly relativistic collimated outflows \citep{Aloy_etal00, Aloy_etal02}. Other attractive examples include pulsar wind nebulae \citep{BZAV05}, microquasars \citep{Meier03,McKG04}, X-ray binaries \citep{VRT02} and stellar core collapse in the context of general relativity \citep{Bruenn85, DFM02}. Theoretical investigations based on direct numerical simulations have paved a way towards a better understanding of the rich phenomenology of relativistic magnetized plasmas. Part of this accomplishment owes to the successful generalization of existing shock-capturing Godunov-type codes to relativistic magnetohydrodynamics (RMHD) \citep[see][and reference therein]{K99, Balsara01,dZBL03}. Implementation of such codes is based on a conservative formulation which requires an exact or approximate solution to the Riemann problem, i.e., the decay of a discontinuity separating two constant states \citep{Toro97}. In terms of computational cost, employment of exact relativistic Riemann solvers may become prohibitive due to the high degree of intrinsic nonlinearity present in the equations. This has focused most computational efforts towards the development of approximate solvers which, nevertheless, require knowledge of the exact solution, at least on some level \citep{MM03}. The presence of magnetic fields further entangles the solution, since the number of decaying waves increases from three to seven \citep{AP87,Anile89}. An exact analytical approach to the solution (which does not allow compound waves) has been recently presented in \cite{GR05}, while \cite{RMPIM05} derived a special case where the velocity and magnetic field are orthogonal. The trade-off between efficiency, accuracy and robustness of such approximate methods is still a matter of research. Solvers based on local linearization have been presented in \cite{K99} (KO henceforth), \cite{Balsara01} (BA henceforth) and \cite{KKU02}. Despite the higher accuracy in reproducing the full wave structure, these solvers rely on rather expensive characteristic decompositions of the Jacobian matrix. Conversely, the characteristic-free formulation of Harten-Lax-van Leer (HLL) of \cite{HLL83} has gained increasing popularity due to its ease of implementation and robustness. The HLL approach has been successfully applied to the RMHD equations by \citealt{dZBL03} (dZBL henceforth) as well as to the general relativistic case \citep[see for example][]{GmKT03,DLSS05} and to the investigation of extragalactic jets, see \cite{LAAM05}. Besides the computational efficiency, however, the HLL formulation averages the full solution to the Riemann problem into a single state, and thus lacks the ability to resolve single intermediate waves such as Alfv{\'e}n, contact and slow discontinuities. In \cite{MB05} (paper I henceforth) we proposed an approach that cured this deficiency by restoring the missing contact wave. The resulting scheme generalized the HLLC approximate Riemann solver by \citet{TSS94} to the equations of relativistic hydrodynamics without magnetic fields. Here, along the same lines, we propose an extension of the HLLC solver to the relativistic magnetized case. Similar work has been presented in the context of classical MHD by \cite{Gurski04} and \cite{Li05}. The new HLLC Riemann solver is implemented in the framework of the corner transport upwind (CTU) method of \cite{Colella90}, coupled with the constrained transport (CT) evolution \citep{EH88} of magnetic field. The algorithm naturally preserves the divergence-free condition to machine accuracy and is stable up to Courant number of $1$. The paper is organized as follows. The relevant equations are given in \S\ref{sec:equations}. In \S\ref{sec:hllc} we derive the new HLLC Riemann solver. Numerical tests, together with the implementation of the CTU-CT method are shown in \S\ref{sec:test}. \section{The RMHD Equations}\label{sec:equations} The motion of an ideal relativistic magnetized fluid is described by conservation of mass, \begin{equation}\label{eq:mass} \partial_\alpha (\rho u^\alpha) = 0 \;, \end{equation} energy-momentum, \begin{equation}\label{eq:mom} \partial_\alpha\Big[(\rho h + |b|^2)u^\alpha u^\beta - b^\alpha b^\beta + p\eta^{\alpha\beta}\Big] = 0 \;, \end{equation} and by Maxwell's equations: \begin{equation}\label{eq:maxwell} \partial_\alpha(u^\alpha b^\beta - u^\beta b^\alpha) = 0 \;. \end{equation} see, for example, \cite{AP87} or \cite{Anile89}. In equations (\ref{eq:mass}), (\ref{eq:mom}) and (\ref{eq:maxwell}) we have introduced the rest mass density of the fluid $\rho$, the four velocity $u^\alpha$, the covariant magnetic field $b^\alpha$ and the relativistic specific enthalpy $h$. The total pressure $p$ results from the sum of thermal (gas) pressure $p_g$ and magnetic pressure $|b|^2/2$, i.e., $p = p_g + |b|^2/2$. In what follows we assume a flat metric, so that $\eta^{\alpha\beta}= \textrm{diag}(-1,1,1,1)$ is the Minkowski metric tensor. Greek indexes run from $0$ to $3$ and are customary for covariant expressions involving four-vectors. Latin indexes (from $1$ to $3$) describe three-dimensional vectors and are used indifferently as subscripts or superscripts. The four-vectors $u^\alpha$ and $b^\alpha$ are related to the spatial components of the velocity $\A{v} \equiv (v_x, v_y, v_z)$ and laboratory magnetic field $\A{B} \equiv (B_x, B_y, B_z)$ through \begin{equation}\label{eq:four-vectors} \begin{array}{ccc} u^\alpha & = & \DS \gamma\big(1,\; \A{v}\big) \;,\\ \noalign{\medskip} b^\alpha & = & \DS \gamma\left(\A{v}\cdot\A{B} \,, \quad \frac{\A{B}}{\gamma^2} + \A{v}(\A{v}\cdot\A{B})\right)\;, \end{array} \end{equation} with the normalizations \begin{equation} u^\alpha u_\alpha = -1 \;,\quad u^\alpha b_\alpha = 0 \;, \end{equation} \begin{equation} |b|^2 \equiv b^\alpha b_\alpha = \frac{|\A{B}|^2}{\gamma^2} + (\A{v}\cdot\A{B})^2 \;, \end{equation} where $\gamma = (1 - \A{v}\cdot\A{v})^{-1/2}$ is the Lorentz factor. We follow the same conventions used in paper I, where velocities are given in units of the speed of light. Writing the spatial and temporal components of equation (\ref{eq:maxwell}) in terms of the laboratory magnetic field yields \begin{equation}\label{eq:induction} \pd{\A{B}}{t} = \nabla\times(\A{v}\times\A{B}) \;, \end{equation} \begin{equation}\label{eq:divB} \nabla\cdot\A{B} = 0 \; , \end{equation} i.e., they reduce to the familiar induction equation and the solenoidal condition. For computational purposes, equations (\ref{eq:mass})--(\ref{eq:maxwell}) are more conveniently put in the standard conservation form \begin{equation}\label{eq:rmhd_eq} \pd{\A{U}}{t} + \sum_k \pd{\A{F}^k(\A{U})}{x^k} = 0 \; , \end{equation} together with the divergence-free constraint (\ref{eq:divB}), where $\A{U} = (D, m_x, m_y, m_z, B_x, B_y, B_z, E)$ is the vector of conservative variables and $\A{F}^k$ are the fluxes along the $x^k\equiv (x,y,z)$ directions. The components of $\A{U}$ are, respectively, the laboratory density $D$, the three components of momentum $m_k$ and magnetic field $B_k$ and the total energy density $E$. From equations (\ref{eq:mass}), (\ref{eq:mom}) and the definitions (\ref{eq:four-vectors}) one has \begin{eqnarray} D & = & \rho\gamma \;, \label{eq:cons_var_D}\\ \noalign{\medskip} m_k & = & (\rho h\gamma^2 + \A{B}^2)v_k - (\A{v}\cdot\A{B})B_k \; , \label{eq:cons_var_m}\\ \noalign{\medskip} E & = & \DS \rho h\gamma^2 - p_g + \frac{\A{B}^2}{2} + \frac{\A{v}^2\A{B}^2 - (\A{v}\cdot\A{B})^2}{2} \label{eq:cons_var_E}\;, \end{eqnarray} and \begin{equation}\label{eq:fluxes} \A{F}^x(\A{U}) = \left(\begin{array}{c} Dv_x \\ \noalign{\medskip} \DS m_xv_x - B_x\frac{b_x}{\gamma} + p \\ \noalign{\medskip} \DS m_yv_x - B_x\frac{b_y}{\gamma} \\ \noalign{\medskip} \DS m_zv_x - B_x\frac{b_z}{\gamma} \\ \noalign{\medskip} 0 \\ \noalign{\medskip} B_yv_x - B_xv_y \\ \noalign{\medskip} B_zv_x - B_xv_z \\ \noalign{\medskip} m_x \end{array}\right) \;. \end{equation} Similar expressions hold for $\A{F}^y(\A{U})$ and $\A{F}^z(\A{U})$ by cyclic permutations of the indexes. Notice that the fluxes entering in the induction equation are the components of the electric field which, in the infinite conductivity approximation, becomes \begin{equation} \A{\Omega} = -\A{v}\times\A{B} \;. \end{equation} The non-magnetic case is recovered by letting $\A{B}\to 0$ in the previous expressions. Finally, proper closure is provided by specifying an additional equation of state. Throughout the following we will assume a constant $\Gamma$-law, with specific enthalpy given by \begin{equation}\label{eq:eos} h = 1 + \frac{\Gamma}{\Gamma - 1}\frac{p_g}{\rho} \;, \end{equation} where $\Gamma$ is the constant specific heat ratio. \subsection{Recovering primitive variables} \label{sec:contoprim} Godunov-type codes are based on a conservative formulation where laboratory density, momentum, energy and magnetic fields are evolved in time. On the other hand, primitive variables, $\A{V} = (\rho, \A{v}, p_g, \A{B})$, are required when computing the fluxes (\ref{eq:fluxes}) and more convenient for interpolation purposes. Recovering $\A{V}$ from $\A{U}$ is not a straightforward task in RMHD and different approaches have been suggested by previous authors: BA used an iterative scheme based on a $5\times 5$ Jacobian sub-block of the system (\ref{eq:rmhd_eq}); KO solves a $3\times 3$ nonlinear system of equations; dZBL (the same approach is also used in \cite{LAAM05}) further reduced the problem to a $2\times 2$ system of nonlinear equations. Here we reduce this task to the solution of a single nonlinear equation, by properly choosing the independent variable. If one sets, in fact, $W = \rho h \gamma^2$, $S = \A{m}\cdot\A{B}$, the following two relations hold: \begin{equation}\label{eq:c2p_E} E = W - p_g + \left(1 - \frac{1}{2{\gamma}^2}\right)|\A{B}|^2 -\frac{S^2}{2 W^2} \;, \end{equation} \begin{equation}\label{eq:c2p_m} |\A{m}|^2 = \left(W + |\A{B}|^2\right)^2 \left(1 - \frac{1}{\gamma^2}\right) - \frac{S^2}{W^2} \left(2W + |\A{B}|^2\right) \;. \end{equation} Since at the beginning of each time step $\A{m}$, $\A{B}$ and $S$ are known quantities, equation (\ref{eq:c2p_m}) allows one to express the Lorentz factor $\gamma$ as a function of $W$ alone: \begin{equation}\label{eq:lorentz} \gamma = \left(1 - \frac{S^2(2W + |\A{B}|^2) + |\A{m}|^2W^2}{(W + |\A{B}|^2)^2W^2} \right)^{-\HALF}\;. \end{equation} Using the equation of state (\ref{eq:eos}), the thermal pressure $p_g$ is also a function of $W$: \begin{equation}\label{eq:c2p_eos} p_g(W) = \frac{W - D\gamma}{\Gamma_r\gamma^2} \;, \end{equation} where $\Gamma_r = \Gamma/(\Gamma - 1)$ and $\gamma$ is now given by (\ref{eq:lorentz}). Thus the only unknown appearing in equation (\ref{eq:c2p_E}) is $W$ and \begin{equation}\label{eq:press_fun} f(W) \equiv W - p_g + \left(1 - \frac{1}{2{\gamma}^2}\right)|\A{B}|^2 -\frac{S^2}{2 W^2} - E = 0 \; \end{equation} can be solved by any standard root finding algorithm. Although both the secant and Newton-Raphson methods have been implemented in our numerical code, we found the latter to be more robust and computationally efficient and it will be our method of choice. The expression for the derivative needed in the Newton scheme is computed as follows: \begin{equation} \frac{df(W)}{dW} = 1 - \frac{dp_g}{dW} + \frac{|\A{B}|^2}{\gamma^3} \frac{d\gamma}{dW} + \frac{S^2}{W^3} \;, \end{equation} where $dp_g/dW$ is computed from (\ref{eq:c2p_eos}), whereas $d\gamma/dW$ is computed from eq. (\ref{eq:lorentz}): \begin{equation} \begin{array}{c} \DS \frac{dp_g}{dW} = \frac{\gamma(1 + Dd\gamma/dW) - 2Wd\gamma/dW} {\Gamma_r\gamma^3} \;, \\ \noalign{\bigskip} \DS \frac{d\gamma}{dW} = -\gamma^3\,\frac{2S^2(3W^2 + 3W|\A{B}|^2 + |\A{B}|^4) + |\A{m}|^2W^3}{2W^3(W + |\A{B}|^2)^3} \;. \end{array}\end{equation} Once $W$ has been computed to some accuracy, the Lorentz factor can be easily found from (\ref{eq:lorentz}), thermal pressure from (\ref{eq:c2p_eos}) and velocities are found by inverting equation (\ref{eq:cons_var_m}): \begin{equation} v_k = \frac{1}{W + |\A{B}|^2}\left(m_k + \frac{S}{W}B_k\right) \end{equation} Finally, equation (\ref{eq:cons_var_D}) is used to determine the proper density $\rho$. \subsection{The Riemann Problem in RMHD}\label{sec:riemann} In the standard Godunov-type formalism, numerical integration of (\ref{eq:rmhd_eq}) depends on the computation of numerical fluxes at zone interfaces. This task is accomplished by the (exact or approximate) solution of the initial value problem: \begin{equation}\label{eq:riemann} \A{U}(x,0) = \left\{\begin{array}{ccc} \A{U}_{L,i+\HALF} & \quad \textrm{if} \; & x < x_{i+\HALF} \;, \\ \noalign{\medskip} \A{U}_{R,i+\HALF} & \quad \textrm{if} \; & x > x_{i+\HALF} \;, \\ \noalign{\medskip} \end{array}\right. \end{equation} where $\A{U}_{L,i+\HALF}$ and $\A{U}_{R,i+\HALF}$ are assumed to be piece-wise constant left and right states at zone interface $i+\HALF$. The evolution of the discontinuity (\ref{eq:riemann}) constitutes the Riemann problem. As in classical MHD, evolution in a given direction is governed by seven equations in seven independent conserved variables. Integration along the $x$-direction, for example, leaves $B_x$ unchanged since the corresponding flux is identically zero, eq. (\ref{eq:fluxes}). The solution to the initial value problem (\ref{eq:riemann}) results, therefore, in the formation of seven waves: two pairs of magneto-acoustic waves, two Alfv{\'e}n waves and a contact discontinuity. The complete analytical solution to the relativistic MHD Riemann problem has been recently derived in closed form by \cite{GR05}. A number of properties regarding simple waves are also well established, see \cite{AP87} and \cite{Anile89}. \cite{RMPIM05} discuss the case in which the magnetic field of the initial states is tangential to the discontinuity and orthogonal to the flow velocity. General guidelines, relevant to the present work, follow below. Across a magneto-acoustic (fast or slow) shock, all components of $\A{V}$ can change discontinuously. Thermodynamic quantities (e.g., $\rho$ and $p_g$) are continuous through a relativistic Alfv{\'e}n wave (as in the classical case), but contrary to the classical counterpart, the magnetic field is elliptically polarized and the normal component of the velocity is discontinuous \citep{K97}. Through the contact mode, only density exhibits a jump while thermal pressure, velocity and magnetic field are continuous. For the special case in which the component of the magnetic field normal to a zone interface vanishes, a degeneracy occurs where tangential, Alfv{\'e}n and slow waves all propagate at the speed of the fluid and the solution simplifies to a three-wave pattern. Under this condition, the approximate solution outlined in paper I can still be applied with minor modifications, see \S\ref{sec:bx0} in this paper and \cite{MMB05}. \section{The HLLC Solver}\label{sec:hllc} The derivation of the HLL and HLLC approximate Riemann solvers has already been discussed in paper I and will not be repeated hereafter. Following the same notations, we approximate the solution to the initial value problem (\ref{eq:riemann}) with two constant states, $\A{U}^*_L$ and $\A{U}^*_R$, bounded by two fast shocks and a contact discontinuity in the middle. We write the solution on the $x/t=0$ axis as \begin{equation}\label{eq:hllc_states} \A{U}(0,t) = \left\{\begin{array}{ccc} \A{U}_L & \quad \textrm{if} & \; \lambda_L \ge 0 \;, \\ \noalign{\medskip} \A{U}^*_L & \quad \textrm{if} & \; \lambda_L \le 0 \le \lambda^* \;,\\ \noalign{\medskip} \A{U}^*_R & \quad \textrm{if} & \; \lambda^* \le 0 \le \lambda_R \;,\\ \noalign{\medskip} \A{U}_R & \quad \textrm{if} & \; \lambda_R \le 0 \;, \\ \noalign{\medskip} \end{array}\right. \end{equation} where $\lambda_L$ and $\lambda_R$ are, respectively, the minimum and maximum characteristic signal velocities and $\lambda^*$ is the velocity of the middle contact wave. The corresponding inter-cell numerical fluxes are: \begin{equation}\label{eq:hllc_flux} \A{f} = \left\{\begin{array}{ccc} \A{F}_L & \quad \textrm{if} & \; \lambda_L \ge 0 \;, \\ \noalign{\medskip} \A{F}^*_L & \quad \textrm{if} & \; \lambda_L \le 0 \le \lambda^*\;, \\ \noalign{\medskip} \A{F}^*_R & \quad \textrm{if} & \; \lambda^* \le 0 \le \lambda_R\;, \\ \noalign{\medskip} \A{F}_R & \quad \textrm{if} & \; \lambda_R \le 0 \;. \\ \noalign{\medskip} \end{array}\right. \end{equation} The intermediate fluxes $\A{F}^*_L$ and $\A{F}^*_R$ are expressed in terms of $\A{U}^*_L$ and $\A{U}^*_R$ through the Rankine-Hugoniot jump conditions: \begin{equation}\label{eq:jump_1}\begin{array}{ccc} \lambda_L \left(\A{U}^*_L - \A{U}_L\right) & = & \A{F}^*_L - \A{F}_L \;,\\ \noalign{\medskip} \lambda^* \left(\A{U}^*_R - \A{U}^*_L\right) & = & \A{F}^*_R - \A{F}^*_L \;,\\ \noalign{\medskip} \lambda_R \left(\A{U}_R - \A{U}^*_R\right) & = & \A{F}_R - \A{F}^*_R \;,\\ \noalign{\medskip} \end{array}\end{equation} where, in general, $\A{F}^*_{L,R} \neq \A{F}(\A{U}^*_{L,R})$. The consistency condition is obtained by adding the previous equations together: \begin{equation}\label{eq:consistency1} \frac{(\lambda^* - \lambda_L) \A{U}^*_L + (\lambda_R - \lambda^*) \A{U}^*_R}{\lambda_R - \lambda_L} = \A{U}^{hll} \;, \end{equation} where \begin{equation}\label{eq:hll_state} \A{U}^{hll} = \frac{\lambda_R \A{U}_R - \lambda_L\A{U}_L + \A{F}_L - \A{F}_R}{\lambda_R - \lambda_L} \,, \end{equation} is the \emph{state} integral average of the solution to the Riemann problem. Similarly, if one divides each expression in eq. (\ref{eq:jump_1}) by the corresponding $\lambda$'s on the left hand sides and adds the resulting expressions, \begin{equation}\label{eq:consistency2} \frac{\A{F}^*_L\lambda_R(\lambda^* - \lambda_L) + \A{F}^*_R\lambda_L(\lambda_R - \lambda^*)}{\lambda_R - \lambda_L} = \lambda^*\A{F}^{hll}\;, \end{equation} with \begin{equation}\label{eq:hll_flux} \A{F}^{hll} = \frac{\lambda_R\A{F}_L - \lambda_L\A{F}_R + \lambda_R\lambda_L (\A{U}_R - \A{U}_L)}{\lambda_R - \lambda_L} \,. \end{equation} being the \emph{flux} integral average of the solution to the Riemann problem. Since the sets of jump conditions across the contact discontinuity differ depending on whether $B_x$ vanishes or not, we proceed by separately discussing the two cases. In either case, the speed of the contact wave is assumed to be equal to the (average) normal velocity over the Riemann fan, i.e. $\lambda^* \equiv v^*_x$. The normal component of magnetic field, $B_x$, is assumed to be continuous at the interface, so that $B_x^* \equiv B_{x,L} = B_{x,R}$ can be regarded as a parameter in the solution. \subsection{Case $B^*_x \neq 0$}\label{sec:bx} We start by noticing that equations (\ref{eq:consistency1}) and (\ref{eq:consistency2}) provide a total of 14 relations. Six additional conditions come by imposing continuity of total pressure, velocity and magnetic field components across the contact discontinuity. This gives us a freedom of $20$ independent unknowns, $10$ per state; we choose to introduce the following set of unknowns for each state \begin{equation}\label{eq:hllc_vars} \left\{D^*, \, v_x^*,\, v_y^*,\, v_z^*,\, B_y^*,\, B_z^*,\, m_y^*,\, m_z^*,\, E^*,\, p^*\right\}\;. \end{equation} The normal component of momentum ($m_x^*$) is not an independent variable since we assume, for consistency, that \begin{equation}\label{eq:mE_rel} m_x^* = (E^* + p^*)v_x^* - \left(\A{v}^*\cdot\A{B}^*\right)B_x^* \;. \end{equation} The previous relation obviously holds between conservative and primitive physical quantities. We point out that the choice (\ref{eq:hllc_vars}) is not unique and alternative sets of independent variables may be adopted. According to the previous definitions, the state vector solution to the Riemann problem is written as \begin{equation}\label{eq:hllc_u*} \A{U}^* = \Big(D^*, m_x^*, m_y^*, m_z^*, B_y^*, B_z^*, E^*\Big)^t \;. \end{equation} while the flux vector, eq. (\ref{eq:fluxes}), becomes \begin{equation}\label{eq:hllc_f*} \A{F}^* = \left(\begin{array}{c} D^*v^*_x \\ \noalign{\medskip} \DS m_x^*v^*_x - \frac{B_x^*B_x^*}{(\gamma^*)^2} - B^*_xv^*_x \left(\A{v}^*\cdot\A{B}^*\right) + p^* \\ \noalign{\medskip} \DS m_y^*v^*_x - \frac{B_x^*B_y^*}{(\gamma^*)^2} - B^*_xv^*_y \left(\A{v}^*\cdot\A{B}^*\right) \\ \noalign{\medskip} \DS m_z^*v^*_x - \frac{B_x^*B_z^*}{(\gamma^*)^2} - B^*_xv^*_z \left(\A{v}^*\cdot\A{B}^*\right) \\ \noalign{\medskip} B^*_yv^*_x - B^*_xv^*_y \\ \noalign{\medskip} B^*_zv^*_x - B^*_xv^*_z \\ \noalign{\medskip} m^*_x \end{array}\right) \end{equation} As in paper I, we adopt the convention that quantities without the $L$ or $R$ suffix refer indifferently to the left ($L$) or right ($R$) state. The six conditions across the contact discontinuity are \begin{equation}\begin{array}{ccc} v^*_{x,L} = v^*_{x,R} \;, & v^*_{y,L} = v^*_{y,R}\;, & v^*_{z,L} = v^*_{z,R} \;, \\ \noalign{\medskip} B^*_{y,L} = B^*_{y,R} \;, & B^*_{z,L} = B^*_{z,R}\;, & p^*_L = p^*_R \;. \end{array}\end{equation} For these quantities the suffix $L$ or $R$ is thus unnecessary. From the transverse components of the magnetic field in the state consistency condition (\ref{eq:consistency1}), one immediately finds that \begin{equation}\label{eq:transv_B} B^*_{y} = B_{y}^{hll} \; ,\quad B^*_{z} = B_{z}^{hll} \; . \end{equation} Thus the transverse components the magnetic field are given by the HLL single state. Similarly, from the fifth and sixth components of the flux consistency condition (\ref{eq:consistency2}) one can express the transverse velocity through \begin{equation}\label{eq:transv_v} B^*_xv^*_y = B^*_yv^*_x - F_{B_y}^{hll} \;,\quad B^*_xv^*_z = B^*_zv^*_x - F_{B_z}^{hll} \;, \end{equation} where $F_{B_y}^{hll}$ and $F_{B_z}^{hll}$ are the $B_y$- and $B_z$- components of the HLL flux, eq. (\ref{eq:hll_flux}). Simple manipulations of the normal momentum and energy components in equation (\ref{eq:consistency1}) together with (\ref{eq:mE_rel}) yield the following simple expression: \begin{equation}\label{eq:state_eq} E^{hll} v^*_x + p^*v^*_x - B^*_x\, \big(\A{v}^*\cdot\A{B}^*\big) = m_x^{hll} \;. \end{equation} Similar algebra on the momentum and energy components of the flux consistency condition (\ref{eq:consistency2}) leads to \begin{equation}\label{eq:flux_eq} \Big[F^{hll}_E - B^*_x\,(\A{v}^*\cdot\A{B}^*)\Big]v^*_x - \left(\frac{B^*_x}{\gamma^*}\right)^2 + p^* - F^{hll}_{m^x} =0 \;. \end{equation} where $1/(\gamma^*)^2 = 1 - (v_x^*)^2 - (v_y^*)^2 - (v_z^*)^2$. Now, if one multiplies equation (\ref{eq:flux_eq}) by $v_x^*$ and subtracts equation (\ref{eq:state_eq}), the following quadratic equation may be obtained: \begin{equation}\label{eq:quadratic} a(v_x^*)^2 + bv_x^* + c = 0 \;, \end{equation} with coefficients \begin{equation}\begin{array}{ccl} a & = & F^{hll}_E - \A{B}^{hll}_\perp\cdot\A{F}^{hll}_{\A{B}_\perp} \;,\\ \noalign{\medskip} b & = & - F^{hll}_{m^x} - E^{hll} + \left|\A{B}_\perp^{hll}\right|^2 + \left|\A{F}^{hll}_{\A{B}_\perp}\right|^2 \;, \\ \noalign{\medskip} c & = & m_x^{hll} - \A{B}^{hll}_\perp\cdot\A{F}^{hll}_{\A{B}_\perp} \;. \end{array}\end{equation} In the previous expressions $\A{B}^{hll}_\perp \equiv (0,B^{hll}_y,B^{hll}_z)$, $\A{F}^{hll}_{\A{B}_\perp} \equiv (0,F^{hll}_{B_y}, F^{hll}_{B_z})$. Similar arguments to those presented in paper I lead to the conclusion that only the root with the minus sign is physically admissible. Once $v_x^*$ is known, $v_y^*$ and $v_z^*$ are readily obtained from (\ref{eq:transv_v}), $p^*$ is computed from (\ref{eq:flux_eq}), while density, transverse momenta and energy are obtained using the Rankine-Hugoniot jump conditions across each fast wave: \begin{eqnarray} D^* & = & \DS \frac{\lambda - v^x}{\lambda - v_x^*} D \;, \label{eq:D_jump} \\ \noalign{\medskip} \label{eq:transv_my} m^*_y & = &\DS \frac{-B^*_x\left[\frac{B^*_y}{(\gamma^*)^2} + (\A{v}^*\cdot\A{B}^*)v_y^*\right] + \lambda m_y - F_{m_y}}{\lambda - v^*_x} \;, \label{eq:my_jump} \\ \noalign{\medskip} \label{eq:transv_mz} m^*_z & = &\DS \frac{-B^*_x\left[\frac{B^*_z}{(\gamma^*)^2} + (\A{v}^*\cdot\A{B}^*)v_z^*\right] + \lambda m_z - F_{m_z}}{\lambda - v^*_x} \;, \label{eq:mz_jump} \\ \noalign{\medskip} E^* & = & \DS \frac{\lambda E - m_x + p^*v_x^* - (\A{v}^*\cdot\A{B}^*)B^*_x}{\lambda - v_x^*} \;. \label{eq:E_jump} \end{eqnarray} In equations (\ref{eq:my_jump}) and (\ref{eq:mz_jump}), $F_{m_y}$ and $F_{m_z}$ are, respectively, the $m_y$- and $m_z$- components of the flux, eq. (\ref{eq:fluxes}), evaluated at the left or right state. As in paper I, we have omitted the suffix $L$ or $R$ for clarity of exposition. \subsection{Case $B^*_x = 0$}\label{sec:bx0} For vanishing normal component of the magnetic field a degeneracy occurs where the Alfv{\'e}n waves and the two slow magnetosonic waves propagate at the speed of the contact discontinuity. For this case the approximate character of the HLLC solver offers a better representation of the exact solution, since the Riemann fan is comprised of three waves only. At the contact discontinuity, however, only the normal component of the velocity $v_x$ and the total pressure $p$ are continuous (KO). The remaining variables experience jumps. This only adds $2$ constraints to the $14$ jump conditions, leaving a freedom of $8$ unknowns per state. However, the transverse velocities $v_y$ and $v_z$ do not enter explicitly in the fluxes (\ref{eq:hllc_f*}) and the jump conditions can be written entirely in terms of $\{D^*, v^*_x, m^*_y, m^*_z, B^*_y, B^*_z, E^*, p^*\}$, i.e. $8$ unknowns per state. Straightforward algebra shows that the coefficients of the quadratic equation (\ref{eq:quadratic}) are now given by \begin{equation} a = F^{hll}_E \;,\quad b = - F^{hll}_{m^x} - E^{hll} \;, \quad c = m_x^{hll} \;, \end{equation} i.e., they coincide with the expressions derived in paper I. The root with the minus sign still represents the correct physical solution. Once $v_x^*$ is found, the total pressure $p^*$ is derived from \begin{equation} p = - F^{hll}_{E}v_x^* + F^{hll}_{m_x}\;, \end{equation} and the normal momentum (\ref{eq:mE_rel}) becomes \begin{equation} m_x^* = (E^* + p^*)v_x^* \;. \end{equation} The remaining quantities are easily obtained from the jump conditions: \begin{eqnarray} D^* & = & \DS \frac{\lambda - v_x}{\lambda - v_x^*} D \;, \\ \noalign{\medskip} m^*_{y,z} & = & \DS \frac{\lambda - v_x}{\lambda - v^*_x}\, m_{y,z} \;, \\ \noalign{\medskip} E^* & = & \DS \frac{\lambda E - m_x + p^*v_x^*}{\lambda - v_x^*} \;, \label{eq:E_jump2} \\ \noalign{\medskip} B^*_{y,z} & = & \DS \frac{\lambda - v_x}{\lambda - v_x^*}\, B_{y,z} \;. \end{eqnarray} \subsection{Remarks}\label{sec:remarks} The expressions derived separately in \S\ref{sec:bx} and \S\ref{sec:bx0} are suitable in the $B_x\neq0$ and $B_x\to0$ cases, respectively. Although other degeneracies may be present (see KO for a thorough discussion) no other modifications are necessary to the algorithm. Before testing the new solver, however, a few remarks are worth of notice: \begin{enumerate} \renewcommand{\theenumi}{(\arabic{enumi})} \item The solutions derived separately for $B_x \neq 0$ and the special case $B_x = 0$ automatically satisfy the consistency conditions (\ref{eq:consistency1}) and (\ref{eq:consistency2}) by construction; \item In the limit of zero magnetic field, the expressions derived in \S\ref{sec:bx0} reduce to those found in paper I; \item In the classical limit, our derivation does not coincide with the approximate Riemann solvers constructed by \cite{Gurski04} or \cite{Li05}. The reason for this discrepancy stems from the fact that both \cite{Gurski04} and \cite{Li05} assume that transverse momenta and velocities are tied by the relation $m^*_{y,z} \equiv \rho^*v_{y,z}^*$. Although certainly true in the exact solution, this assumption reduces, in the HLLC approximate formalism, the number of unknowns from $10$ to $8$ (when $B_x\neq0$) thus leaving the systems of jump conditions (\ref{eq:jump_1}) overdetermined. Should this be the case, the number of equations exceeds the number of unknowns and the integral relations across the Riemann fan inevitably break down. This explains the inconsistencies found in Li's and Gurski's derivations and further discussed in \cite{MK05}. Therefore, in the classical limit, our expressions automatically imply $m^*_{y,z} \neq \rho^*v_{y,z}^*$ and the correct expressions for the transverse velocities are still given by (\ref{eq:transv_v}), whereas transverse momenta should be derived from the jump conditions accordingly. Furthermore, contrary to Li's misconception, consistency with the jump conditions requires that the magnetic field components be uniquely determined by (\ref{eq:transv_B}) and no other choices are thus possible. \item The reader might have noticed that in the limit of vanishing $B_x$, some of the expressions given in \S\ref{sec:bx} do not reduce to the those found in \S\ref{sec:bx0}. This property also persists in the classical limit, see \cite{Gurski04}, and \cite{Li05}. The reason for this discrepancy relies on the assumption of continuity of the transverse components of magnetic field across the tangential wave $\lambda^*$: when $B_x \to 0$, a degeneracy occurs where the tangential, Alfv{\'e}n and slow waves all propagate at the speed of the fluid and the solution simplifies to a three-wave pattern. In the exact solution, the continuity of $B_y$ and $B_z$ across the tangential wave is lost since the middle state bounded by the two slow waves becomes singular. \item Lastly, we note that in both the classical and relativistic case the transverse velocities given by eq. (\ref{eq:transv_v}) become ill-defined as $B_x\to 0$. However, in the classical case, the terms involving $v^*_y$ or $v^*_z$ in the flux definitions remain finite as $B_x\to 0$. Conversely, this is not the case in RMHD for arbitrary orientation of the magnetic field as one can see, for example, using eq. (\ref{eq:transv_my}): \begin{equation} m^*_y \sim \frac{(B_z^{hll}v_x^* - F^{hll}_{B_z} ) (F^{hll}_{B_y} B^{hll}_z - F_{B_z}^{hll}B^{hll}_y)} {B_x(\lambda - v^*_x)} + O(1) \end{equation} as $B_x\to 0$. Fortunately, for strictly two dimensional flows (e.g. when $B_z = v_z = 0$) the leading order term vanishes and the singularity is avoided. In the general case, however, we conclude that more sophisticated solvers should allow the presence of rotational discontinuities in the solution to the Riemann problem. This has been done, for example, by \cite{MK05} in the context of classical MHD. \end{enumerate} \subsection{Wave Speed Estimate}\label{sec:speeds} The full characteristic decomposition of the RMHD equation (i.e. the eigenvalues and eigenvectors of the Jacobian matrix $\partial\A{F}^x/\partial\A{U}$) was extensively analyzed by \cite{AP87} and \cite{Anile89}. In the one-dimensional case the Jacobian matrix can be decomposed into seven eigenvectors associated with four magnetosonic waves (fast and slow disturbances), two Alfv{\'e}n waves and one entropy wave propagating at the fluid velocity. The eigenstructure is therefore similar to the classical case and it can be shown that the ordering of the various speeds and corresponding degeneracies are preserved \citep{Anile89}. Since the HLLC approximate Riemann solver requires an estimate of the outermost waves, the right and left-going fast shock speeds identify the necessary characteristic velocities. Thus we set \citep{Davis88}: \begin{equation}\begin{array}{c}\label{eq:wavespeeds} \lambda_L = \min\big(\lambda_-(\A{V}_L), \lambda_-(\A{V}_R)\big) \;, \\ \noalign{\medskip} \lambda_R = \max\big(\lambda_+(\A{V}_L), \lambda_+(\A{V}_R)\big) \;, \end{array} \end{equation} where $\lambda_{-}$ and $\lambda_+$ are the minimum and maximum roots of the quartic equation \begin{equation}\label{eq:eigenspeed_eq} \rho h(1-c_s^2)a^4 = (1-\lambda^2) \left[(|b|^2 + \rho h c_s^2)a^2 - c_s^2{\cal B}^2\right] \;, \end{equation} with $a = \gamma(\lambda - v_x)$, ${\cal B} = b^x - \lambda b^0$. In absence of magnetic field, both the (left and right-going) slow and fast shocks propagate at the same speed and equation (\ref{eq:eigenspeed_eq}) reduces to the quadratic equation (22) shown in paper I. When $\A{B} \neq \A{0}$, no simple analytical expression is available and solving (\ref{eq:eigenspeed_eq}) requires numerical or rather cumbersome analytical approaches. Recently, \cite{LAAM05} proposed approximate simple lower and upper bounds to the required eigenvalues. Here we choose to solve eq. (\ref{eq:eigenspeed_eq}) by means of analytical methods, where the quartic is reduced to a cubic equation which is in turn solved by standard methods. There are special cases where it is possible to handle some of the degeneracies more efficiently using simple analytical formulae: \begin{itemize} \item for vanishing total velocity, equation (\ref{eq:eigenspeed_eq}) reduces to a bi-quadratic, \begin{equation} (\rho h + |b|^2)\lambda^4 - (|b|^2 + \rho hc_s^2 + B_x^2c_s^2)\lambda^2 + c_s^2B_x^2 = 0 \end{equation} \item for vanishing normal component of the magnetic field, equation (\ref{eq:eigenspeed_eq}) yields a quadratic equation: \begin{equation} a_2 \lambda^2 + a_1 \lambda + a_0 = 0 \end{equation} with $a_2 = \rho h\big[c_s^2 + \gamma^2(1-c_s^2)\big] + {\cal Q}$, $a_1 = -2\rho h\gamma^2v_x(1-c_s^2)$, $a_0 = \rho h\big[ - c_s^2 + \gamma^2v_x^2(1-c_s^2)\big] - {\cal Q}$ and ${\cal Q} = |b|^2 - c_s^2(\A{v}_\perp\cdot\A{B}_\perp)^2$. \end{itemize} For all other cases we solve the quartic equation (\ref{eq:eigenspeed_eq}). \subsection{Positivity of the HLLC scheme}\label{sec:positivity} The set of physically admissible conservative states, $G$, identify all the $\A{U}$'s yielding positive thermal pressure $p_g$ and total velocity $|\A{v}| < 1$, according to the procedure outlined in \S\ref{sec:contoprim}. Thus the positivity of the HLLC approximate Riemann solver requires that \begin{itemize} \item both left and right intermediate states $\A{U}^*_L$ and $\A{U}^*_R$ belong to $G$; \item the first-order scheme yields updated conservative states that are in $G$. \end{itemize} Unfortunately, the mathematical proof of the positivity of the HLLC scheme presents remarkable algebraic difficulties. In absence of the singular behavior described in \S\ref{sec:remarks}, investigations have been carried at the numerical level by verifying that each intermediate state $\A{U}^*$ correspond to a primitive, physically admissible state. In all the tests presented in this paper and several others not discussed here, the scheme did not manifest any loss of positivity. However, in the general three-dimensional case when $B_x,B_y,B_z\neq0$, the terms involving $B_x$ in the expressions for the transverse momenta may become arbitrarily large as $B_x\to 0$ and a loss of positivity can be experienced. \section{Algorithm Validation}\label{sec:test} \subsection{Corner Transport Upwind for relativistic MHD}\label{sec:ctu} The RMHD equations (\ref{eq:rmhd_eq}) are evolved in a conservative, dimensionally unsplit fashion: \begin{equation}\label{eq:update} \A{U}^{n+1}_{i,j} = \A{U}^n_{i,j} + \A{\cal L}^{x,n+\HALF}_{i,j} + \A{\cal L}^{y,n+\HALF}_{i,j} \,, \end{equation} where the $\A{\cal L}$'s are Godunov operators \begin{equation}\label{eq:god_op_x} \A{\cal L}^{x,n+\HALF}_{i,j} = - \frac{\Delta t}{\Delta x_i} \left(\A{f}^{x,n+\HALF}_{i+\HALF,j} - \A{f}^{x,n+\HALF}_{i-\HALF,j}\right)\,, \end{equation} \begin{equation}\label{eq:god_op_y} \A{\cal L}^{y,n+\HALF}_{i,j} = - \frac{\Delta t}{\Delta y_j} \left(\A{f}^{y,n+\HALF}_{i,j+\HALF} - \A{f}^{y,n+\HALF}_{i,j-\HALF}\right)\,, \end{equation} and $\A{U}^n$ is the set of volume-averaged conservative variables $\A{U}^n = \Big(D, \A{m}, \bar{\A{B}}, E\Big)^n$ at time $t=t^n$. Here $\bar{\A{B}}$ denotes the zone-averaged magnetic field. For clarity of exposition we will omit, throughout the following, integer-valued subscripts $(i,j)$ and retain only the half-integer notation to denote zone edge values. The fluxes appearing in equations (\ref{eq:god_op_x}) and (\ref{eq:god_op_y}) are computed by solving, at each zone interface, a Riemann problem with suitable time-centered left and right input states. For example, we obtain $\A{f}^{y,n+\HALF}_{j+\HALF}$ as the HLLC flux with input states given by $\A{V}^{n+\HALF}_{j+\HALF,L}$ and $\A{V}^{n+\HALF}_{j+\HALF,R}$, respectively. Computation of time-centered left and right zone edge values proceeds using the corner transport upwind (CTU) of \cite{Colella90}, recently extended to relativistic hydrodynamics by \cite{MPB05} and to classical MHD by \cite{GS05}. Here we generalize the CTU approach to relativistic MHD by following a slightly different approach, although equivalent to the guidelines given in \cite{Colella90}. For the sake of conciseness, only the essential steps will be described hereafter. The unfamiliar reader is referred to the work of \cite{Colella90}, \cite{Saltzman94} and \cite{GS05} for more comprehensive derivations. In our formulation, second-order accurate left and right states are sought in the form \begin{equation}\label{eq:pred_states} \A{V}^{n+\HALF}_{i\pm\HALF,S} = \A{V}^{x,n+\HALF} \pm \frac{\delta_x\A{V}^n}{2} \, , \quad \A{V}^{n+\HALF}_{j\pm\HALF,S} = \A{V}^{y,n+\HALF} \pm \frac{\delta_y\A{V}^n}{2} \, , \quad \end{equation} where we take $S=L$ ($S=R$) with the plus (minus) sign. The slopes $\delta_x\A{V}^n$ and $\delta_y\A{V}^n$ are computed at the beginning of the time step using, for example, the monotonized central-difference (MC) limiter: \begin{equation}\label{eq:mc_lim} \delta_x q^n = s_i\min\left(2|\Delta q^n_+|, 2|\Delta q^n_-|, \frac{|q^n_{i+1} - q^n_{i-1}|}{2}\right) \,, \end{equation} where $q\in\A{V}$ and \begin{equation} \Delta q^n_\pm = \pm\left(q^n_{i\pm1} - q^n_i\right) \,,\; s_i = \frac{\sign(\Delta q^n_+) + \sign(\Delta q^n_-)}{2} \;. \end{equation} An alternative smoother prescription is given by the harmonic mean \citep{vLeer77}: \begin{equation}\label{eq:vl_lim} \delta_x q^n = \frac{2\max\left(0, \Delta q_+\Delta q_-\right)} {\Delta q_+ + \Delta q_-} \,. \end{equation} Equation (\ref{eq:mc_lim}) provides smaller dissipation at discontinuities, whereas equation (\ref{eq:vl_lim}) was found to give less oscillatory results. Interpolation in the $y$-direction is done in a similar manner. Additional forms of limiting may be adopted if necessary, see \S\ref{sec:shockflattening} and \S\ref{sec:mdlimit}. The cell- and time- centered values on the right hand sides of equations (\ref{eq:pred_states}) are computed from a Taylor expansion of the conservative variables, i.e. \begin{equation}\label{eq:cent_pred_x} \A{U}^{x,n+\HALF} \approx \A{U}^n + \frac{\Delta t}{2}\pd{\A{U}}{t} = \A{U}^n - \frac{\Delta t}{2}\left( \pd{\hat{\A{F}}^x}{x} + \pd{\A{F}^y}{y}\right) \,, \end{equation} \begin{equation}\label{eq:cent_pred_y} \A{U}^{y,n+\HALF} \approx \A{U}^n + \frac{\Delta t}{2}\pd{\A{U}}{t} = \A{U}^n - \frac{\Delta t}{2}\left( \pd{\A{F}^x}{x} + \pd{\hat{\A{F}}^y}{y}\right) \,. \end{equation} Following \cite{Colella90}, we approximate the spatial derivative in the direction normal to a zone interface (denoted with a hat) with the Hancock step already introduced in paper I, \begin{equation}\label{eq:hancock_diff} \pd{\hat{\A{F}}^x}{x} \approx \frac{ \A{F}^x\left(\A{V}^n_{i+\HALF,L}\right) - \A{F}^x\left(\A{V}^n_{i-\HALF,R}\right)}{\Delta x_i}\,, \end{equation} whereas the derivative in the tangential direction is computed in an upwind fashion using a Godunov operator: \begin{equation}\label{eq:upwind_diff} \Delta t\pd{\A{F}^y}{y} \approx -\A{\cal{L}}^{y,n} = \frac{\Delta t}{\Delta y_j} \left(\A{f}^{y,n}_{j+\HALF} - \A{f}^{y,n}_{j-\HALF}\right)\,. \end{equation} The state $\A{U}^{y,n+\HALF}$ is obtained by similar arguments by interchanging the role of normal and tangential derivatives. We would like to point out that the Godunov operators used in the predictor step involve left and right states computed at $t=t^n$ (and not at $t=t^{n+\HALF}$ as in \cite{GS05}): \begin{equation}\label{eq:init_states} \A{V}^n_{i\pm\HALF,S} = \A{V}^n \pm \frac{\delta_x\A{V}^n}{2} \;, \quad \A{V}^n_{j\pm\HALF,S} = \A{V}^n \pm \frac{\delta_y\A{V}^n}{2} \,. \end{equation} This choice still makes the scheme second-order accurate in space and time and was found, in our experience, to yield a more robust algorithm. Besides, our CTU implementation does not require a primitive variable formulation, thus offering ease of implementation in the context of relativistic hydro and MHD, where the Jacobian $\partial\A{F}/\partial\A{U}$ is particularly expensive to evaluate. Note that a total of four Riemann problems are involved in the single time step update (\ref{eq:update}). It can be easily verified that for one-dimensional flows, the corner transport upwind method outlined above reduces to the scheme presented in paper I. Finally, the choice of the time step $\Delta t$ is based on the Courant-Friederichs-Lewy (CFL) condition \citep{CFL28}: \begin{equation} \Delta t = \textrm{CFL} \times \min_{i,j}\left( \frac{ \Delta x}{\max(|\lambda^x_L|,|\lambda^x_R|)}, \frac{ \Delta y}{\max(|\lambda^y_L|,|\lambda^y_R|)}\right) \,, \end{equation} where $0<\textrm{CFL}<1$ is the Courant number and $|\lambda^x_{L,R}|$, $|\lambda^y_{L,R}|$ are the zone interface wave speeds computed in the $x$ and $y$ directions according to (\ref{eq:wavespeeds}). \subsubsection{Contrained Transport Evolution of the Magnetic Field} \label{sec:divB} It is well known that multidimensional numerical schemes do not generally preserve the solenoidal condition, eq. (\ref{eq:divB}), unless special discretization techniques are employed. In this respect, several approaches have been suggested in the context of the classical MHD equations \citep{Toth00, LdZ00} and some of them have been recently extended to the relativistic case, see dZBL. Here we adopt the constrained transport (CT) \citep{EH88} and follow the approach of \cite{BS99} for its integration in Godunov-type schemes. In the CT approach a new staggered magnetic field variable is introduced. In this representation, the components of the magnetic field are treated as area-weighted averages on the zone faces to which they are orthogonal. Thus, $B_x$ is collocated at $(i+\HALF,j)$, whereas $B_y$ at $(i,j+\HALF)$. No jump is allowed in the normal component of $\A{B}$ at a zone boundary, consistently with the well posedness of the Riemann problem presented in \S\ref{sec:riemann} and \S\ref{sec:hllc}. Transverse components may be discontinuous. In this formulation, a discrete version of Stoke's theorem is used integrate the induction equation (\ref{eq:induction}). For example, after the predictor steps (\ref{eq:cent_pred_x}) and (\ref{eq:cent_pred_y}), we update the face-centered magnetic field according to \begin{equation}\label{eq:stokes}\begin{array}{ccc} \DS B^{n+\HALF}_{x,i+\HALF} & = & \DS B^{n}_{x,i+\HALF} - \frac{\Delta t^n}{2\Delta y_j}\Big(\Omega^z_{i+\HALF, j+\HALF} - \Omega^z_{i+\HALF, j-\HALF}\Big)\;, \\ \noalign{\medskip} \DS B^{n+\HALF}_{y,j+\HALF} & = & \DS B^{n}_{y,j+\HALF} + \frac{\Delta t^n}{2\Delta x_i}\Big(\Omega^z_{i+\HALF, j+\HALF} - \Omega^z_{i-\HALF, j+\HALF}\Big)\;, \end{array} \end{equation} and similarly after the corrector step. The electromotive force $\Omega$ is collocated at cell corners and is computed by straightforward arithmetic averaging: \begin{equation}\label{eq:omega} \Omega^z_{i+\HALF, j+\HALF} = \frac{ \Omega^z_{i+\HALF,j} + \Omega^z_{i, j+\HALF} + \Omega^z_{i+\HALF,j+1} + \Omega^z_{i+1,j+\HALF}}{4} \;, \end{equation} where, $\Omega^z_{i+\HALF,j} \equiv -f^{x,n}_{B_y,i+\HALF,j}$ and $\Omega^z_{i,j+\HALF} \equiv f^{y,n}_{B_x,i,j+\HALF}$ are the $z$ components of the electric fields available at grid interfaces during the upwind step. Despite its simplicity, eq. (\ref{eq:omega}) lacks of directional bias and more sophisticated algorithms may be used to incorporate upwind information in a consistent way, see \cite{LdZ04}, \cite{GS05}. For ease of implementation we will not discuss them here. It is a straightforward exercise to verify that the $\nabla\cdot\A{B} = 0$ condition is preserved from one time step to the next one, due to perfect cancellation of terms. Notice also that, since $B_x$ is continuous at the $(i+\HALF,j)$ interface, only $\bar{B}_y$ and $\bar{B}_z$ need to be interpolated during the reconstruction procedure in the $x$-direction. A similar argument applies to $\bar{B}_x$ and $\bar{B}_z$ when interpolating along the $y$ coordinate. Since equation (\ref{eq:update}) evolves volume-averaged quantities, the zone-averaged magnetic field, $\bar{\A{B}}$, is computed at the beginning of the time step from the face-averaged magnetic fields using linear interpolation: \begin{equation}\label{eq:bx_average} \bar{B}_{x} = \frac{B_{x,i+\HALF} + B_{x,i-\HALF}}{2} \;, \end{equation} \begin{equation}\label{eq:by_average} \bar{B}_{y} = \frac{B_{y,j+\HALF} + B_{y,j-\HALF}}{2} \;. \end{equation} Equations (\ref{eq:omega}), (\ref{eq:bx_average}) and (\ref{eq:by_average}) are second-order accurate in space. \subsubsection{Summary} We summarize our CTU constrained transport algorithm by the following steps: \begin{enumerate} \renewcommand{\theenumi}{(\arabic{enumi})} \item At the beginning of the time step, form the volume averages (\ref{eq:bx_average}) and (\ref{eq:by_average}) from the face centered magnetic field. \item Compute $x$ and $y$ limited slopes by interpolating cell centered primitive variables according to eq. (\ref{eq:mc_lim}) or (\ref{eq:vl_lim}). \item\label{xpred} Make a sweep along the $x$ direction. Form left and right states using the first of eq. (\ref{eq:init_states}) with $B^{n}_{x,i+\HALF,L} = B^{n}_{x,i+\HALF,R}$ equal to the $x$ component of the face centered magnetic field; \begin{itemize} \item[-] use the Hancock step (\ref{eq:hancock_diff}) to compute the $x$ derivative in eq. (\ref{eq:cent_pred_x}) and add the resulting contribution to $U^{x,n+\HALF}$; \item[-] compute the $\A{\cal L}^{x,n}$ Godunov operator by solving Riemann problems at the $(i+\HALF,j)$ interfaces and add the resulting contribution to $U^{y,n+\HALF}_{i,j}$. \end{itemize} \item\label{ypred} Make a sweep along the $y$ direction. Form left and right states using the second in eq. (\ref{eq:init_states}) with $B^{n}_{y,j+\HALF,L} = B^{n}_{y,j+\HALF,R}$ equal to the $y$ component of the face centered magnetic field; \begin{itemize} \item[-] obtain the $\A{\cal L}^{y,n}$ Godunov operator (\ref{eq:upwind_diff}) by solving Riemann problems at the $(i,j+\HALF)$ interfaces; add the resulting contribution to $U^{x,n+\HALF}_{i,j}$. \item[-] use the Hancock step relative to the $y$ direction to compute the $y$ derivative and add it to $U^{y,n+\HALF}_{i,j}$; \end{itemize} \item Compute the time-centered area weighted magnetic field using Stoke's theorem (\ref{eq:stokes}). This concludes the predictor step. \item Make a sweep along the $x$ direction with left and right time-centered states given by the first equation in (\ref{eq:pred_states}) with $B^{n+\HALF}_{x,i+\HALF,L} = B^{n+\HALF}_{x,i+\HALF,R}$ equal to the time centered face-averaged magnetic field computed via Stoke's theorem. Obtain the $\A{\cal L}^{x,n+\HALF}$ Godunov operator. \item Repeat the previous step by sweeping along the $y$ direction. Compute the $\A{\cal L}^{y,n+\HALF}$ Godunov operator. \item Update the cell-centered conservative variables using eq. (\ref{eq:update}) and the face-averaged magnetic field using Stoke's theorem. \end{enumerate} \subsection{One-dimensional test problems}\label{sec:1d} One-dimensional problems are specifically designed to verify the ability of the algorithm in reproducing the exact wave pattern. In what follows we present four shock-tube tests, already introduced by BA and dZBL, with left and right states given in Table \ref{tab:ic}. Computations are performed on the interval $[0,1]$ and the initial discontinuity is placed at $x = 0.5$. The final integration time is $t=0.4$. Note that the constrained transport algorithm is unnecessary, since eq. (\ref{eq:divB}) is trivially satisfied in one-dimensional flows. \begin{table}\begin{center} \begin{tabular}{ccccccccc} Test & $\rho$ & $p_g$ & $v_x$ & $v_y$ & $v_z$ & $B_x$ & $B_y$ & $B_z$ \\ \hline\hline 1L & 1 & 1 & 0 & 0 & 0 & 0.5 & 1 & 0 \\ 1R & 0.125 & 0.1 & 0 & 0 & 0 & 0.5 & -1 & 1 \\ \hline 2L & 1 & 30 & 0 & 0 & 0 & 5 & 6 & 6 \\ 2R & 1 & 1 & 0 & 0 & 0 & 5 & 0.7 & 0.7 \\ \hline 3L & 1 & $10^3$ & 0 & 0 & 0 & 10 & 7 & 7 \\ 3R & 1 & 0.1 & 0 & 0 & 0 & 10 & 0.7 & 0.7 \\ \hline 4L & 1 & 0.1 & 0.999 & 0 & 0 & 10 & 7 & 7 \\ 4R & 1 & 0.1 & -0.999 & 0 & 0 & 10 & -7 & -7 \\ \hline \end{tabular} \caption{Initial conditions for the one-dimensional shock tube problems presented in the text. In all test problems we adopt a resolution of $1600$ uniform computational zone, covering the interval $[0,1]$. Integration is carried until $t=0.4$.} \label{tab:ic} \end{center}\end{table} \subsubsection{Problem 1}\label{sec:p1} The first test problem, initially proposed by \cite{vP93}, is a relativistic extension of the \cite{BW88} magnetic shock tube. In analogy with the classical case we use the ideal equation of state (\ref{eq:eos}) with specific heat ratio $\Gamma = 2$. The breakup of the initial discontinuity sets up a left-going fast rarefaction wave, a left-going compound wave, a contact discontinuity, a right-going slow shock and a right-going fast rarefaction wave. We compare, in Fig. \ref{fig:flat1}, the results obtained with the first-order HLL and HLLC solvers on $100$ uniform computational zones. The exact solution (given by the solid line) was obtained using the numerical code available from \cite{GR05}. The left going compound wave located at $x \approx 0.5$ is only visible in the numerical integration since the code used to generate the analytical solution (shown as the solid line in Fig. \ref{fig:flat1}) does not allow compound structures by construction. As expected, the HLLC Riemann solver attains sharper representation of the contact discontinuity when compared to the HLL scheme. Because of the reduced smearing in proximity of the contact wave, neighboring structures such as the compound wave on the left and the slow shock on the right can be better resolved when using the HLLC solver. Computations at different resolutions show, in fact, that the L-1 norm errors in density are reduced by roughly $20\div 30\%$ (see left panel in Fig. \ref{fig:res_1st}), with $L_1(\%)$ being, respectively, $0.53$ and $0.74$ for the HLLC and HLL solver at the highest resolution employed ($6400$ zones). Fig. \ref{fig:sod1} shows the results obtained with the second-order scheme with the MC limiter, eq. (\ref{eq:mc_lim}), and the same Courant number, $CFL = 0.8$ on $1600$ grid points. A direct comparison with the exact solution shows that all discontinuities are correctly captured and resolved on few computational zones, owing also to the presence of a compressive limiter. In this respect, our second-order HLLC scheme provides similar results to those obtained with the third-order central ENO-HLL scheme by dZBL. The L-1 norm errors computed at different resolutions with the two different solvers differ by $\approx 10\div 20\%$, see left panel in Fig. \ref{fig:res_2nd}. When compared to the more sophisticated, characteristic-based algorithm presented in BA, our results show slightly sharper representation of the right-going slow shock and the contact discontinuity. Small overshoots appear in the Lorentz factor profile at the left going compound wave and the right going slow shock. More diffusive slope limiters do not exhibit this feature. \subsubsection{Problem 2}\label{sec:p2} The resulting wave pattern for this configuration is comprised of two left-going rarefaction fans (fast and slow) and two right-going slow and fast shocks. The specific heat ratio used for this calculation is $\Gamma = 5/3$. The weak slow rarefaction located at $x\approx 0.53$ and the slow shock at $x\approx 0.86$ are separated by a contact discontinuity where the proper density changes by a factor of $\sim 7$. The velocity on either side of the contact wave is mildly relativistic, with a maximum Lorentz factor of $\approx 1.36$. The improvement offered by the HLLC Riemann solver over the HLL approach in the resolution of the contact wave is evident from Fig. \ref{fig:flat2}, where we compare the density profiles obtained with the first order schemes against the analytical solution. Computations obtained with the second-order limiter (\ref{eq:mc_lim}) show excellent agreements with the analytical profiles, see Fig. \ref{fig:sod2}. Our single-step HLLC scheme attain considerably sharper resolution than the results obtained by previous calculations. The two right-going shocks, for instance, are smeared over $\sim 3$ grid points, approximately half of the resolution shown in BA and dZBL. Moreover, the smearing of the contact wave is considerably reduced when compared to the HLL scheme in dZBL ($\sim 10$ zones vs. $\sim 14$). Similar overshoots, though, appear at the right of contact mode. The discrete L-1 errors for different grid sizes are shown in the right panel of Fig. \ref{fig:res_2nd}, where, at the maximum resolution employed ($6400$ zones) the HLLC and HLL errors reduce to $0.17 \%$ and $0.25\%$, respectively. \subsubsection{Problem 3}\label{sec:p3} The configuration for this test is similar to the previous problem, but a higher pressure jump separates the initial left and right states, see Table \ref{tab:ic}. Only the second-order scheme with the Van Leer limiter (\ref{eq:vl_lim}) and a Courant number of $0.8$ has been employed. The ideal equation of state (\ref{eq:eos}) with $\Gamma = 5/3$ is used. The ensuing wave pattern shows a stronger relativistic configuration, with a maximum Lorentz factor of $\sim 3.37$, see Fig. \ref{fig:sod3}. The presence of magnetic fields makes the problem even more challenging than its hydrodynamical counterpart (see test 3 in paper I), since the contact wave, slow and fast shocks now propagate extremely close to each other. As a result, a thin density shell sets up between the contact mode and the slow shock. The higher compression factor (more than $100$) follows from a more pronounced relativistic length contraction effect. At the resolution of $1600$ grid zones, the relative error in the density peak ($\rho_{\max} \approx 9.98$) is $1.2 \%$. A second thin shell-like structure forms between the slow and fast shocks, as can be seen in the profiles in Fig. \ref{fig:sod3}. The peaks achieved in the transverse components of velocity ($\approx -0.37$) and magnetic field ($\approx 8.95$) achieve, respectively, $87\%$ and $95\%$ of their exact values. The small shell thickness, however, still prevents a clear resolution of the two right going shocks, visible in the exact solution. This demonstrates that relativistic magnetized flows can develop rich and complex features difficult to resolve on a grid of fixed size. Similar conclusions have been drawn by previous investigators. Results obtained with the HLL solver (not shown here) indicates that the resolution attained at the contact discontinuity is equivalent. Therefore, as it was also pointed out in paper I, we conclude that, for strong blast waves where relativistic contraction effects produce closely moving discontinuities, the HLL and HLLC schemes produce nearly identical results. \subsubsection{Problem 4}\label{sec:p4} The collision of two relativistic streams is considered in the fourth test problem. The initial impact produces two strong relativistic fast shocks propagating symmetrically in opposite direction about the impact point, $x=0.5$, see Fig. \ref{fig:sod4}. Two slow shocks delimiting a high pressure region in the center follow behind. Computations are carried out with $CFL = 0.8$ and the Van Leer limiter, eq. (\ref{eq:vl_lim}). Spurious oscillations in vicinity of strong shocks are reduced by switching to the more diffusive minmod limiter, see \S\ref{sec:shockflattening}. No contact waves are present in the problem and, not surprisingly, the quality of our solution is essentially the same obtained by previous authors: the fast shocks are resolved in $2\div 3$ cells, whereas the slow shocks are smeared out over $5\div 6$ zones. Very similar patterns are observed in the work of BA and dZBL. It is well known that Godunov-type schemes suffer from a common pathology, often found in these type of problems. In the classical case, this has been recognized for the first time by \cite{Noh87}. The wall heating problem, in fact, consists in an undesired entropy buildup in a few zones around the point of symmetry. Our scheme is obviously no exception as it can be inferred by inspecting the density profile in Fig. \ref{fig:sod4}. We repeated the test with the HLL scheme and found that this pathology is worse when the HLLC scheme is used. The relative numerical undershoot in density, in fact, were found to be $\sim 5\%$ for the HLL and $\sim 12\%$ for the HLLC scheme. Since similar errors were also reported by BA, and the same conclusions have been drawn in paper I, we raise the question as to whether the degree of this pathology grows with the complexity of the Riemann solver. Future, more specific works should address this problem. \subsection{Two-dimensional test problems}\label{sec:multid} Multi-dimensional numerical computations of magnetized flows are notoriously more challenging, due to the necessity to preserve the divergence-free constraint (\ref{eq:divB}). In what follows, we consider three test problems: a cylindrical blast wave test, the interaction of a strong magnetosonic shock with a cloud and the propagation of an axisymmetric jet in cylindrical coordinates. \subsubsection{Cylindrical Blast Wave}\label{sec:p5} Cylindrical explosions in cartesian coordinates are particular useful in checking the robustness of the code and the algorithm response to different kinds of degeneracies. Here we follow the same setup adopted by KO, where the square $[-6, 6]\times[-6, 6]$ is filled with a uniform ($\rho = 10^{-4}$, $p_g=3\cdot10^{-5}$), initially static ($\A{v} = \A{0}$) medium, threaded by a constant magnetic field $\A{B} = (B_x,0)$. The circular region $\sqrt{x^2 + y^2} < 0.08$ is initialized with constant higher density and pressure values, $\rho = 0.01$ and $p_g = 1$ decreasing linearly for $0.08\le r \le 1$. We adopt the ideal equation of state (\ref{eq:eos}) with specific heat ratio $\Gamma = 4/3$. We consider two setups, corresponding to a relatively weak magnetic field $B_x = 0.1$ and a strong field $B_x = 1$. Figures \ref{fig:blast_lo} and \ref{fig:blast_hi} show the magnetic field distribution, thermal pressure and Lorentz factor for the two configurations at $t = 4$. Computations are carried using the van Leer limiter, eq. (\ref{eq:vl_lim}), together with the multidimensional limiting procedure described in \S\ref{sec:mdlimit} on $200\times200$ uniform grid zones. The Courant number is $0.4$. The expanding region is delimited by a fast forward shock propagating (nearly) radially at almost the speed of light. In the weak field case, a reverse shock delimits the inner region where expansion takes place radially. Magnetic field lines are squeezed in the $y$ direction building up a shell of higher magnetic pressure. In the $x$ direction the motion of the gas is not hindered by the presence of the field and it achieves a higher Lorentz factor ($\gamma_{\max} = 4.39$). In the strong field case, the expansion is magnetically confined along the $x$ direction and the outer fast shock has reduced amplitude. The maximum Lorentz factor is $\gamma_{\max} = 4.02$. We point out that numerical integrations for this test were possible only by locally redefining the total energy at the end of the time step: \begin{equation}\label{eq:new_E} E \rightarrow E + \frac{\bar{\A{B}}_{fa}^2 - \bar{\A{B}}_c^2}{2}\;, \end{equation} where $\bar{\A{B}}_c$ is the cell-centered magnetic field obtained after the Godunov step, whereas $\bar{\A{B}}_{fa}$ is the new magnetic field obtained by averaging the face centered values given by (\ref{eq:stokes}). Notice that equation (\ref{eq:new_E}) only redefines the energy contribution of the magnetic field that is not directly coupled to the velocity, see eq. (\ref{eq:cons_var_E}) and thus may be regarded as a first-order correction. In this respect, the energy correction we propose is the same usually adopted in CT schemes, see \cite{BS99}, \cite{Toth00}. Although this optional step results in a slight loss of energy conservation at the discretization level, it was nevertheless found to become particularly useful in problems where the magnetic pressure dominates over the thermal pressure by more than two order of magnitudes. \subsubsection{Relativistic Shock-Cloud Interaction}\label{sec:p6} The interaction of a strong relativistic fast shock with a cloud is considered on the unit square $[0,1]\times[0,1]$ in 2-D cartesian coordinates $(x,y)$. This problem has been extensively used for testing classical MHD codes see \cite[][and references therein]{DW94,Toth00}. Here we consider a relativistic extension adopting a somewhat different initial condition, with magnetic field orthogonal to the slab plane. The shock wave travels in the positive $x$-direction and is initially located at $x=0.6$. Upstream, for $x > 0.6$, the flow is highly supersonic with pre-shock values given by $(\rho, \gamma_x, p_g, B_z)_\textrm{pre} = (1, 10, 10^{-3}, 0.5)$, where $\gamma_x = (1 - v_x^2)^{-\HALF}$. In this reference frame, shocked material is at rest with values given by \begin{equation} \left(\begin{array}{c} \rho \\ \noalign{\medskip} p_g \\ \noalign{\medskip} B_z \end{array}\right)_{\textrm{post}} = \left(\begin{array}{c} 42.5942 \\ \noalign{\medskip} 127.9483 \\ \noalign{\medskip} -2.12971 \\ \noalign{\medskip} \end{array}\right) \;. \end{equation} Notice that the magnetic field carries a rotational discontinuity and the compression factor of density across the shock in not limited to $7$ (we use $\Gamma = 4/3$) as in the classical case, but achieves a much higher value ($\approx 43$). This feature is unique to relativistic flows. A circular density clump with $\rho = 10$ and radius $r = 0.15$ is placed ahead of the shock front, centered at $(x,y) = (0.8,0.5)$. Transverse velocities $v_y$ and $v_z$ and the $x$ and $y$ components of magnetic field are set to zero everywhere. We use $400\times 200$ computational zones, by assuming reflecting boundary at $y = 0.5$ and free flow across the remaining boundaries. The MC limiter, eq. (\ref{eq:mc_lim}), is employed everywhere except in proximity of strong shocks where we revert to the minmod limiter, see \S\ref{sec:shockflattening}. The Courant number is $0.4$. Shortly after the impact, the cloud undergoes strong compression with the density rising by a factor of more than $20$. The collision generates a bow fast shock propagating in the shocked material and a reverse shock is transmitted into the cloud. After the transmitted shock reaches the back of the cloud, the two bent parts of the original incident shock join back together and complicated wave pattern emerges. By $t=1$ the cloud is completely wrapped by the incident shock, and the cloud expands in the form of a mushroom-shaped shell, see upper half of Fig. \ref{fig:sc}. The solution computed with the HLL solver (lower half in Fig. \ref{fig:sc}) show similar structures, although the amount of numerical viscosity is considerably higher. Notice that, because of the assumed slab symmetry, the condition $\A{v}\cdot\A{B} = 0$ is preserved in time and the solution to the Riemann problem at each interface consists of a three wave pattern: two fast waves separated by a tangential discontinuity. In this regard, our HLLC solver provides a better approximation of the full wave structure. \subsubsection{Relativistic Jet}\label{sec:p8} As a final example, we consider the propagation of an axisymmetric jet in cylindrical coordinates $(r,z)$. The configuration adopted here corresponds to model C2-pol-1 in \cite{LAAM05}. The domain $[0,12]\times[0,50]$ (in units of jet beam) is initially filled with a static uniform distributions of density, gas pressure and magnetic field, given respectively by \begin{equation} \rho_a = 1 \,,\quad p_a = \frac{\eta v_b^2}{\Gamma(\Gamma-1)M^2 - \Gamma v_b^2} \, ,\quad B_z = \sqrt{2p_a}. \end{equation} The numerical value of $p_a$ follows from the definitions of the beam Mach number $M = v_b/c_s = 6$, jet to ambient density ratio $\eta = 10^{-2}$ and beam axial velocity $v_b=0.99$. The ideal equation of state (\ref{eq:eos}) is used with $\Gamma = 5/3$. The jet nozzle is located at the lower boundary $r \le 1$, $z=0$, where boundary conditions are held constant in time, $(\rho, v_r, v_z, B_r, B_z, p_g) = (\eta, 0, v_b, 0, B_z, p_a)$. For $r>1$ we prescribe boundary values with antisymmetric profiles for axial velocity and radial magnetic field. Symmetric profiles are imposed on the remaining quantities. This configuration corresponds to a twin counter jet propagating in the opposite direction. Outflow boundaries are imposed on all other sides, except at $r=0$ where reflecting boundary conditions are used. We employ a uniform resolution of $20$ zones per beam radius and carry integration until $t = 126$ with $CFL = 0.4$. The results are shown in Fig. \ref{fig:jet}, where we display density logarithm (upper panel), magnetic pressure (middle panel) and Lorentz factor distributions (lower panel). In each panel, the upper and lower halves show the solutions obtained with the HLLC and HLL solvers, respectively. As we already pointed out in the non magnetic case (Paper I), the HLLC integration features considerably less amount numerical diffusion as evident from the richness in small scale structures, notably in the density distribution. In fact, density is the physical quantity more sensitive to the introduction of the tangential wave in the Riemann solver. Comparing our results with those of \cite[][see their Fig. 5]{LAAM05} we can observe that our solution has a similar (or even larger) richness in fine structure details at half the resolution (20 ppb in our case, 40 ppb in their case). \section{Conclusions} An HLLC approximate Riemann solver has been developed for the relativistic magnetohydrodynamic equations. The new approach improves over the single state HLL solver in the ability to capture exactly isolated tangential and contact discontinuities. Several test problems in one and two dimensions demonstrate better resolution properties and a reduced amount of the numerical diffusion inherent to the averaging process of the single state HLL scheme. The solver is well-behaved for strictly two-dimensional flows, although applications to genuinely three-dimensional problems may suffer from a pathological singularity when the component of magnetic field normal to a zone interface approaches zero. This feature does not persist in the classical limit. Multidimensional integration has been formulated in a versatile and efficient way within the framework of the corner transport upwind (CTU) method. The algorithm is stable up to Courant numbers of $1$ and preserves the divergence-free condition via constrained transport evolution of the magnetic field. The additional computational cost and the numerical implementation in an existing relativistic MHD code are minimal. \appendix \section{} \subsection{Shock Flattening}\label{sec:shockflattening} For strong shocks, we found that the one-dimensional prescriptions (\ref{eq:mc_lim}) or (\ref{eq:vl_lim}) can still produce spurious numerical oscillations eventually leading to the occurrence of negative pressures. A weak form of flattening is introduced by replacing eq. (\ref{eq:mc_lim}) or (\ref{eq:vl_lim}) with the minmod limiter whenever a strong shock is detected. In order for the latter condition to hold, we require that both $\nabla\cdot\A{v} < 0$ and $\chi_{\min} = 0$, where $\nabla\cdot\A{v}$ is computed by central differences whereas \begin{equation} \chi_{\min} = \min\left(\chi^x_{i+1,j},\chi^x_{i,j},\chi^x_{i-1,j}, \chi^y_{i,j+1},\chi^y_{i,j},\chi^y_{i,j-1}\right) \,. \end{equation} The switches $\chi^x$ and $\chi^y$ are designed as follows \begin{equation} \chi^x_{i,j} = \left\{\begin{array}{cc} 1 & \DS \; \textrm{if}\quad \frac{p_{i+1,j} - p_{i-1,j}} {\min\left(p_{i+1,j}, p_{i-1,j}\right)} \le \epsilon \;,\\ \noalign{\medskip} 0 & \DS \; \textrm{otherwise} \;, \end{array}\right. \end{equation} \begin{equation} \chi^y_{i,j} = \left\{\begin{array}{cc} 1 & \DS \; \textrm{if}\quad \frac{p_{i,j+1} - p_{i,j-1}} {\min\left(p_{i,j+1}, p_{i,j-1}\right)} \le \epsilon \;,\\ \noalign{\medskip} 0 & \DS \; \textrm{otherwise} \;, \end{array}\right. \end{equation} where we set $\epsilon = 5$ in all computations presented in this paper. \subsection{Multidimensional Limiting} \label{sec:mdlimit} Occasionally, we found that strong shocks propagating obliquely to the grid in highly magnetized media may benefit from an additional form of limiting, based on genuinely multidimensional constraints. When needed, we enforce the maximum and minimum interpolated values in each cell $(i,j)$ to lie within the bounds provided by the four neighboring zones $(i+1,j), (i-1,j), (i,j+1), (i,j-1)$. Specifically, denote with $\hat{q}^{\max}$ and $\hat{q}^{\min}$ the maximum and minimum values of $q\in\A{V}$ in these cells. Once the limited slopes $\delta_x q$ and $\delta_y q$ have been computed according to (\ref{eq:mc_lim}) or (\ref{eq:vl_lim}), we apply the correction \begin{equation} \delta_xq \rightarrow \tau\delta_xq\;, \quad \delta_yq \rightarrow \tau\delta_yq\;, \quad \end{equation} where the multi-dimensional limiter $\tau$ is constructed as in \cite{Balsara04}: \begin{equation} \tau = \min\left(1,\psi\min\left( \frac{\hat{q}^{\max} - q}{\delta^{\max}}, \frac{q - \hat{q}^{\min}}{\delta^{\min}} \right)\right) \;, \end{equation} with $\delta^{\max} = \max(|\delta_x q|,|\delta_y q|)$, $\delta^{\min} = \min(|\delta_x q|,|\delta_y q|)$. We set $\psi = 2$ for density and magnetic field, $\psi = 3/4$ for velocity and $\psi = 1$ for thermal pressure. \label{lastpage}
Title: Seeing Star Formation Regions with Gravitational Microlensing
Abstract: We qualitatively study the effects of gravitational microlensing on our view of unresolved extragalactic star formation regions. Using a general gravitational microlensing configuration, we perform a number of simulations that reveal that specific imprints of the star forming region are imprinted, both photometrically and spectroscopically, upon observations. Such observations have the potential to reveal the nature and size of these star forming regions, through the degree of variability observed in a monitoring campaign, and hence resolve the star formation regions in distant galaxies which are too small to be probed via more standard techniques.
https://export.arxiv.org/pdf/astro-ph/0601667
command. \shorttitle{Star Formation \& Gravitational Microlensing} \shortauthors{Gil-Merino \& Lewis} \begin{document} \title{Seeing Star Formation Regions with Gravitational Microlensing} \author{Rodrigo Gil-Merino\altaffilmark{1} \& Geraint F. Lewis\altaffilmark{2}} \affil{Institute of Astronomy, School of Physics, University of Sydney, NSW 2006, Australia} \altaffiltext{1}{rodrigo@physics.usyd.edu.au} \altaffiltext{2}{gfl@physics.usyd.edu.au} \keywords{gravitational lensing -- microlensing -- star forming regions -- dark halo populations} \section{Introduction} Gravitational microlensing is now a well established technique for the investigation of the distribution of compact (dark) matter in the universe. Furthermore, it also provides a powerful tool to study unresolved sources, such as in the case of the structure of QSOs, through temporal differential magnification (e.g. Yonehara et al. 1998). From an observer's point of view, gravitational microlensing can be naturally divided in two different regimes. In the case of Galactic microlensing, the optical depth is low and a single star microlenses another star within the Galactic halo or in one of the galaxies in the Local Group (Paczy\'nski 1986a). With Extragalactic microlensing, where the light from a distant quasar shines through a closer galaxy, the optical depth is roughly unity and many stars contribute to the overall microlensing effect (Paczy\'nski 1986b). This paper considers this latter regime, were the source region is populated by a number of hot, young stars in a star forming region. Such a situation will occur in strongly lensed, multiply imaged systems, such as the multiple images seen in galaxy clusters (Mellier 1999), or the case where a isolated galaxy gravitationally lenses a more distant galaxy (i.e. Warren et al 1996). In a similar vein, Lewis \& Ibata (2001) investigated the effect of a cosmological distribution of compact objects on the surface brightness distributions of galaxies at $z$$<$0.5, considering a small microlensing optical depth ($\leq$0.04) and they determined that low-level fluctuations in surface brightness of $\sim$2\% should result. Lewis et al. (2000) extended that analysis to distant galaxies observed through galaxy clusters, assuming dark matter to be composed of compact objects. Focusing upon Abell~370 as a case study, concluding that for low-luminosity ($\sim$10$^4$L$_\odot$) stellar populations would show rapid fluctuations exceeding 10\% of the mean in the highest cases. In this contribution we address the question of what microlensing signatures should be apparent in the case of part of a galaxy which is lensed by another galaxy. In particular, if the lensed parts of the source galaxy are regions of star formation, highly dominated by young, massive stars. Such a situation was recent presented by Smith et al. (2005) who reported the discovery of a new strong gravitationally lensed system, with an elliptical galaxy acting as the lens. The lens galaxy in this system is at redshift $z=0.0345$ and the source, proposed to be a star formation region, is at $z\sim0.45$, with the arcs formed by the gravitational mirage showing `knots' of an extreme blue color of $B-I_c=1.1$ (extinction corrected). This discovery poses the idea that microlensing in the multiple images of these systems might be able to distinguish the type of source stars involved in the mirage and help in the interpretation of its nature. Within the context of gravitational lensing, a star formation region would appear as a non-uniform source, composed of a number of bright points in a more extended background. Hence, the microlensing imprint of such a source should show quite a different variability imprint from the uniform sources typically considered in gravitational microlensing experiments. The nature of this imprint is the basis of this current contribution. \section{Microlensing simulations}\label{sim} For the purpose of this study we performed microlensing simulations by means of ray-shooting techniques (Paczy\'nski 1986b, Schneider \& Weiss 1987, Kayser et al. 1986, Wambsganss 1990, Witt 1993, Lewis et al. 1993). To compute magnification patterns, one has to select certain values for the convergence ($\kappa$), which represents the gravitational potential due to matter in the beam, and the shear ($\gamma$), which is the perturbation to the beam due to the large scale distribution of matter. Typically, these parameters are drawn from a lens model for a particular system. For this study, however, representative values of $\kappa=0.55$ and $\gamma=0.55$ are employed, following Schechter et al. (2004), although other combinations would illustrate the situation equally; $kappa$ here includes also any form of compact dark matter, the effects of an smooth dark matter component are described in Schechter \& Wambsganss (2002). Since high resolution maps are required, we used a receiving field of 2 Einstein radii\footnote{ The Einstein radius is defined in the source plane as $ER=sqrt{(4GM/c^2) (D_{s}D_{ls}/D_{l})}$, where $M$ is the mass of the microlens, $D$ is the angular distance to the source (s), the lens (l) and between the lens and the source (ls), c is the velocity of light and G the gravitational constant}, covered by a $2048^2$ pixels area. The microlenses were randomly distributed and selected to have the same mass, $M_{\mu lens}=1 M_{\odot}$. Again, the selection of the mass range is arbitrary for our purposes. However, it is important to note that rather than covering a large area in the simulations, the key point remains in the resolution of the magnification patterns, because we are interested in small flux changes from pixel to pixel, so we also selected a high number of rays that resulted in over 700 per pixel on average. The next step in the simulations is introducing the effect of the source. To do this, we assumed a source plane at $z=0.5$ and two different sizes of $0.1$ and $0.5$~Einstein radii (ER), which corresponds to a physical size of $0.02$~pc and $0.1$~pc respectively at that distance for the standard $\Lambda$CDM cosmoslogy. Although star formation regions might be larger than the bigger size considered, these two examples will illustrate the different effects due to their sizes and could be seen as clumps of star formations within larger regions (compact and ultra-compact H~{\small II} regions as indicators of star formation might be $<$0.1~pc, see e.g. Giveon et al. 2005 and references therein). We also assumed that our lens plane is at $z=0.04$ (following the case of Smith et al. 2005). Depending on the stellar density of the source region, the number of stars in that region can vary from just a few up to hundreds. Considering first the $0.1$~Einstein radii region, we `built' three different sources: one containing 8 stars, another containing 80, and the last one as a uniform source, i.e., containing one star per pixel in the region (the number of stars are not representative of any particular region, and have been chosen arbitrarily). The results for the first region size are displayed in Fig.~\ref{fig1}. The upper left-hand pannel corresponds to a $\sim$1~ER$^2$ region of the original magnifcation pattern. The upper right-hand pannel is the magnification pattern convolved with a region of $0.1$~ER containing 8 stars. The lower right-pannel is the same as the previous one but containing 80 stars. The lower left-hand pannel is the magnification pannel convolved with an `uniform' source of the same physical size. In all the panels the same track has been drawn, in order to compared the synthetical light curves to each other; these are depicted in Fig.~\ref{fig3}. The light curves are $\sim$1~ER long, showing the different expected fluctuations corresponding to the different scenarios. The magnification distributions for the magnification patterns corresponding to the different panels in Fig.~\ref{fig1} are shown in Fig.~\ref{fig2}. Clearly the number of stars in the region has a significant influence on the resulting light curve; in effect, the presence of each star produces a ``shift-and-add'' to the mangification map, greatly increasing the number and overall density of caustics. This is reflected as additional peaks in the light curve. As the number of stars in increased to 80, some of the caustic structure has begun to wash out, leaving small scale fluctuations superimposed on a more gentle background, whereas the smooth source (which can be thought of as a very high density of stars) has washed out all small scale detail. In Fig.~\ref{fig3b}, for comparison, we consider a region size of $0.5$~ER containing also 8 stars, 80 stars and a `uniform' source in the same manner as in Fig.~\ref{fig1}. The corresponding track shows a completely different light curves compare to Fig.~\ref{fig3} although their positions are the same, due to the new caustic structure of the magnification maps according to the different size of the region considered. The interpretation of these figures: if the magnification pattern is convolved with a `uniform' source profile (Fig.~\ref{fig1}, lower left panel) the result is always a smoother pattern, with smooth transitions in the value of the magnification from pixel to pixel; if the source area is made of a number of point-like objects, the convolution will show many caustics slightly shifted one another, with no smooth transition between them. This translates into a rapid variability in the lightcurves of the corresponding source. Also, the size of the regions considered plays an active role in the final imprint of microlensing in the observational light curves. \section{Applications and Discussion} The application of the simulations described in the previous Section can be done in the following manner. If multiple lensed `knots' are detected in an image (see, e.g., Fig. 3c in Smith et al. 2005) and are thought to be star-forming regions, the flux will be highly dominated by young O-stars. In principle, since young massive stars are rare due to their evolutionary process, only a few are expected in these star forming knots (an ultraviolet and optical spectral atlas of the Small Magellanic Cloud includes $<$20 O-stars, see Walborn et al. 2000). Observing these areas, e.g. in the UV band, which characterizes regions of star formation, with periodic photometry, the variability of the observed light curves will be related to the number and separation of these stars present in the star forming regions (the contamination by late-type stars in the UV will be almost null). In practice, one could treat the problem statistically by simulating the observed variability and thus put limits on the amount and luminosity of young dominating stars. Knowing the luminosity of these stars accurately is important, because their masses derived by stellar evolutionary models and by stellar atmosphere models can be compared. Stellar evolution theory and initial mass function might take advantage of these results as well. Gravitational microlensing might be the only tool to `resolve' these stars in clusters of star formation, otherwise impossible to investigate in galaxies at moderate/high redshift. Spectroscopy of microlensed star-forming regions might help to put limits on their nature as well. Gravitational microlensing of broad spectral lines in QSOs has been studied theoretically by a number of authors (e.g., Abajas et al. 2002, Lewis et al. 2004, Richards et al. 2004) and used to put limits on the size of the broad line emitting regions of the QSOs. In those cases, the natural shape of a line is distorted by the complex net of caustics produced by the microlenses on the source plane. Since microlensing of the broad line region is expected when its physical size is of the order of the Einstein radius of the lens projected onto the source plane, microlensed spectral lines give an idea of such physical sizes. In the same way, observing typical spectral lines of O-stars (e.g. in the UV, O {\small V} $\lambda$1371\AA, C {\small III} $\lambda$1176\AA~in the optical, He {\small II} $\lambda$4686\AA, N {\small III}$\lambda$4634\AA~and $\lambda$4640\AA) one would expect the lines to be deformed by the presence of the caustics (due to magnifications/demagnifications), and these variations in the spectral lines might reveal the size and populations of the star forming regions. To illustrate this, we plot in Fig.~\ref{fig4} the spectra of an O-star (upper pannel) and a solar-like G-star (lower pannel), obtained from the Kurucz models database\footnote{http://garnet.stsci.edu/STIS/stis\_models.html}. For any unresolved star formation region, the (far-)UV range of the spectrum will be dominated by these O-stars. Late-type stars fluxes in UV are several orders of magnitude lower and thus they do not contribute significantly to the total luminosity and the flux distribution in the upper pannel of Fig.~\ref{fig4} might be a representative one for that part of the spectrum (different lines might be present, obviously). Microlensing affecting this part of the spectrum will only show an enhancement of the flux. However, the effect is slightly different when using optical range of the spectrum. In this case, the flux contribution due to late-type stars starts to be dominant, although O-stars flux is still significantly present. We construct a toy star formation region model, merging the spectra of the O-star and the G-star shown in Fig.~\ref{fig4}, assuming that 90\% of the total flux comes from solar-like stars and the rest is produced by early-type stars. The toy model is depicted in Fig.~\ref{fig5} (lower line, spectrum marked as 'O-star$+$G-star') showing only the 1500\AA-5500\AA ~wavelength interval. When the star formation region travels across the magnification pattern in Fig.~\ref{fig1}, late-type stars will act as a constant flux background as a whole and microlensing will affect mainly O-stars. In this way, the microlensing signature in the spectra will be a flux ratio variation between O-stars and late-type stars spectral lines. This is shown also in Fig.~\ref{fig5} (upper line, spectrum marked as 'O-star$+$G-star$+$microlensing'). There is not only an enhancement of the flux in the bluest part of the spectrum, but also a deformation of certain lines due to the different microlensing effect on the different type of stars. In Fig.~\ref{fig6} and Fig.~\ref{fig7} we repeat the procedure, but assuming a different relative flux between the two types of stars. In Fig.~\ref{fig6}, 1\% of the flux is coming from O-type stars and 99\% is from late-type stars; in Fig.~\ref{fig7} the percentage is 0.1\% and 99.9\% for early and late-type stars respectively. As shown in Fig.~\ref{fig3}, high variability is expected. Both Figures~\ref{fig2} and \ref{fig3} show that the amount of variability depends on the nature of the star formation regions (number of stars, size of the regions...). This means that comparing several consecutive spectra one would be able to statisticaly determine the relative flux variability of the spectral lines and continuum due to the presence of the caustics, and thus compare them with the expected one from the simulations, puting limits to the number and distribution of the early-type stars. A key point to note is that to detect these microlensing effects on star formation regions is the time scale of the events. Considering the lens configuration describe in Sec.~\ref{sim}, we can estimate a typical separation between microcaustics in the magnification maps in Fig.~\ref{fig3}. This separation is $\sim$0.005~ER for the upper right-hand panel, which corresponds to approximately 10$^{-3}$~pc. Assuming a transverse velocity for the source galaxy of $\sim$6000~km/s (see Kayser et al. 1986), the resulting time-scale for the events is $\sim$50 days. The time-scales of the events get shorter when the number of O-stars gets higher, although the flux variability is smaller. This means that six data points in a time period of around three months should be able to described the type of variability involved in the gravitational microlensing scenario. \section{Conclusions} We described in this contribution how to apply gravitational microlensing to the observations of unresolved extragalactic star forming regions. The discussion shows that due to the caustics configuration in the magnification maps of the region, rapid monitoring campaigns, both photometric or spectroscopic, would reveal high variability fluctuations due to the number of early-type stars. The specific amount of variability will depend on the number of stars and their distribution in the region, as well as on the exact configuration of the microlenses in the lensing galaxy. Thus, the study of a particular system requires the knowledge of a lens model to perform the right simulations and the analysis of the results should be based on a statistical approach. The advantage of the method, if these circumstances take effect, is that we might be able to investigate star formation regions which are difficult to analyse with more traditional techniques. \acknowledgments
Title: The Small-Scale Environment of Quasars
Abstract: Where do quasars reside? Are quasars located in environments similar to those of typical L* galaxies, and, if not, how do they differ? An answer to this question will help shed light on the triggering process of quasar activity. We use the Sloan Digital Sky Survey to study the environment of quasars and compare it directly with the environment of galaxies. We find that quasars (M_i < -22, z < 0.4) are located in higher local overdensity regions than are typical L* galaxies. The enhanced environment around quasars is a local phenomenon; the overdensity relative to that around L* galaxies is strongest within 100 kpc of the quasars. In this region, the overdensity is a factor of 1.4 larger than around L* galaxies. The overdensity declines monotonically with scale to nearly unity at ~1 Mpc, where quasars inhabit environments comparable to those of L* galaxies. The small-scale density enhancement depends on quasar luminosity, but only at the brightest end: the most luminous quasars reside in higher local overdensity regions than do fainter quasars. The mean overdensity around the brightest quasars (M_i < -23.3) is nearly three times larger than around L* galaxies while the density around dimmer quasars (M_i = -22.0 to -23.3) is ~1.4 times that of L* galaxies. By ~0.5 Mpc, the dependence on quasar luminosity is no longer significant. The overdensity on all scales is independent of redshift to z = 0.4. The results suggest a picture in which quasars typically reside in L* galaxies, but have a local excess of neighbors within ~0.1 - 0.5 Mpc; this local density excess likely contributes to the triggering of quasar activity through mergers and other interactions.
https://export.arxiv.org/pdf/astro-ph/0601522
\title{The Small-Scale Environment of Quasars} \author{Will~Serber\altaffilmark{1}, Neta~Bahcall\altaffilmark{1}, Brice~M\'{e}nard\altaffilmark{2}, Gordon~Richards\altaffilmark{1}\altaffilmark{3}} \altaffiltext{1}{Princeton University Observatory, Princeton, NJ 08544, USA} \altaffiltext{2}{Institute for Advanced Study, Einstein Drive, Princeton, NJ 08540, USA} \altaffiltext{3}{Department of Physics and Astronomy, The Johns Hopkins University, 3400 North Charles Street, Baltimore, MD 21218-2686} \slugcomment{Dec15 '05} \keywords{Quasars: General, Galaxies: Statistics} \section{Introduction} \label{sec.intro} For more than two decades, significant effort has been spent attempting to understand the triggering mechanism of quasar activity, as well as the relation between quasars and their host galaxies. Since \citet{bell_1969}, it has become widely accepted that quasars are fueled by accretion of gas onto super-massive black holes. Observations have shown that a number of nearby galaxies have a central black hole whose mass correlates with the luminosity of the spheroid of the host galaxy. This connection suggests that the formation of the black hole is linked to the formation of the galaxy which, in turn, is known to strongly depend on its environment. To develop a better understanding of the quasar phenomenon, it is therefore important to investigate and quantify the relation between quasars and their environments. Despite the importance of this issue, our knowledge of the quasar environment is still limited. Quasar environments have been studied on different scales ranging from those of the host galaxy to those of large scales. Such studies have provided important but controversial results regarding the environment of quasars. It has been known for more than three decades that quasars are associated with enhancements in the spatial distribution of galaxies \citep{bahcall_1969}. Studies have shown that, in the nearby universe, quasars reside in environments ranging from small to moderate groups of galaxies rather than in rich clusters (\citealt{bahcall_1991b,fisher_1996,mclure_2001}). Early observations of quasar environments \citep{stockton_1978, yee_1984, yee_1987, boyle_1988, smith_1990, ellingson_1991}, revealed a positive association of bright quasars with neighboring galaxies at a level somewhat higher than that of normal galaxies and comparable to the environment of small- to intermediate-richness groups of galaxies \citep[e.g.,][]{bahcall_1991}. Observations of quasar environment from the Hubble Space Telescope snapshot survey \citep{bahcall_1997} further support this density enhancement around bright quasars. All these observations focused on small scales, typically within $\sim 0.5\,$Mpc of the quasars, and used relatively small samples of objects. Early observations of the clustering properties of quasars themselves, as measured by the quasar auto-correlation function, suggest that quasars are significantly more strongly clustered than galaxies on scales up to $10\,$Mpc and greater \citep[e.g.,][]{shaver_1988, shanks_1988, chu_1988, chu_1989, crampton_1989}, but less clustered than rich clusters of galaxies \citep[e.g.,][]{bahcall_1991}. This finding suggests that quasars are located in high overdensity regions, more so than $L^*$ galaxies, since higher overdensity regions are clustered more strongly than lower overdensity regions \citep[e.g.,][]{bahcall_1983, kaiser_1984, bardeen_1986}. An overdense environment would indeed be expected if the quasar activity was triggered by galaxy interactions. On the other hand, new generation surveys, such as the Two Degree Field (2dF) and the Sloan Digital Sky Survey (SDSS), have given rise to different results. Using significantly larger complete samples of quasars and galaxies, these surveys have shown that on large scales, i.e. from 1 to $10\,$Mpc, the quasar-galaxy cross-correlation and the quasar auto-correlation are comparable to the correlation function of $L^*$ galaxies. This suggests, in conflict with previous results, that quasars and Active Galactic Nuclei (AGNs) inhabit environments similar to those of $L^*$ galaxies \citep[e.g.,][]{smith_1995, croom_1999, croom_2003, miller_2003, kauffmann_2004, wake_2004}. The results also suggest that the quasar correlation function does not depend significantly on either quasar luminosity or redshift within the ranges studied. Recent work on sub-Mpc scales using the SDSS to find quasar pairs suggests that the quasar-quasar auto-correlation function may be enhanced relative to the galaxy-galaxy distribution \citep{hennawi_2005}, consistent with the earlier results on small scales, as discussed above. In this paper we use the SDSS survey to determine the galaxy environment around quasars as a function of scale. The SDSS is uniquely suited for this investigation: it is the largest complete survey available of both galaxies and quasars, carried out in a well-calibrated, self-consistent manner. The data used in this study covers 4000 deg$^2$, with $\sim2\times10^3$ quasars of redshift $z\le0.4$ and ten million photometric galaxies to a magnitude limit of $i = 21$. We use these data to determine the mean galactic environment around quasars as a function of quasar luminosity and redshift. For comparison, the same analysis is then repeated to find the local environment around $10^5$ spectroscopic galaxies in the SDSS area, as well as around random positions in the survey. All the analyses are carried out using the same ten million photometric galaxies. This technique allows a direct comparison between the environment around quasars with that around random points as well as with the environment around $L^*$ galaxies, thereby minimizing potential selection effects and systematics. The outline of the paper is as follows: we discuss the data in Section~\ref{sec.data}, the analysis in Section~\ref{sec.analysis}, and the results in Section~\ref{sec.results}. The conclusions are summarized in Section~\ref{sec.conclusions}. Throughout this paper, we use a cosmological model with $H_0\,=\,70\,{\rm km^{-1}\,Mpc^{-1}}$, $\Omega_M\,=\,0.3$, and $\Omega_{\Lambda}\,=\,0.7$ for both absolute magnitudes and distance measures. All distances are measured using comoving coordinates. \section{Data} \label{sec.data} We use Sloan Digital Sky Survey (SDSS) data to determine the galactic environment of quasars and galaxies. The SDSS \citep{york_2000,stoughton_2002,pier_2003,Abazajian_2003,gunn_2005} is conducting an imaging survey of $10^4$ square degrees of the sky in five bands ($u, g, r, i, z$) \citep{fukugita_1996,gunn_1998}, followed by a spectroscopic multi-fiber survey of the brightest $10^6$ galaxies and $10^5$ quasars. The spectroscopic targets are selected from the high quality imaging data using well-defined selection criteria \citep{lupton_2001,hogg_2001,strauss_2002,richards_2002}. The drift-scan imaging survey reaches a limiting magnitude of $r < 23$ \citep{fukugita_1996,gunn_1998,lupton_2001}. The main spectroscopic survey targets galaxies to $r < 17.7$, with a median redshift of $z \sim0.15$ and a tail reaching $z \sim0.4$ (Strauss et al 2002). The spectroscopic survey of quasars , with $i < $19, reaches quasar redshifts out to $z \sim5.4$. For more details on the SDSS see the above references. The high quality imaging and spectroscopic survey of quasars and galaxies provides a unique data set for studying the environment of quasars and comparing it directly with the environment of galaxies. To do so, we use the third data release (DR3) of the SDSS spectroscopic sample of quasars \citep{schneider_2005}, selecting all spectroscopic quasars with redshift $z \le 0.4$ and $i$-band Galactic extinction corrected and k-corrected magnitude $-24.2 \le M_i \le -22.0$ \citep{richards_2002}. We set an upper limit of $-24.2$ on the luminosity to avoid bright objects that may interfere with counting nearby galaxies. After applying masks for missing fields (see the end of this section), a sample of 2028 $z \le 0.4$ quasars is used, covering an area of approximately 4000 deg$^2$. In addition to the quasars, we use a sample of spectroscopic galaxies as targets in our analysis so that we may compare the environment of quasars with that of galaxies. The spectroscopic galaxy sample used for comparison is the NYU-LSS sample 12 \citep{blanton_2003a,blanton_2003b}, which is comprised of a complete spectroscopic sample of galaxies to $i\,=\,18.5$ corrected for both Galactic extinction and k-correction, with redshifts in the range $z~\sim0.001$ to $z~\sim0.4$. After applying masks and limiting the galaxy redshift to $0.08 \le z \le 0.4$ (since there are no quasars with $z < 0.08$), a sample of $\sim10^5$ spectroscopic galaxies is available over a $2230\,$ deg$^2$ area (mostly overlapping the quasar area). Our galaxy sample has a median redshift of 0.13 and a median magnitude of $M_i = -21.3$, and our quasar sample has a median redshift of 0.32 and a median magnitude of $M_i = -22.5$. The environment of the above targets - spectroscopic quasars and spectroscopic galaxies - is then determined using the photometric galaxies from DR3 of the SDSS imaging survey. We use a sample of over 10 million galaxies with magnitude in the range $14 \le i \le 21$. For further comparison, we also repeat the environment analysis at approximately $10^3$ random positions per target using the same method and background sample of photometric galaxies. All samples were corrected using the same masks, which remove missing fields, missing stripes, and regions where the stripe boundaries extend beyond the extent of the photometric galaxy survey. These masks were created with SDSSpixel, a pixelization scheme routinely applied to the SDSS data\footnote{See http://lahmu.phyast.pitt.edu/\~{}scranton/SDSSPix/ for information on SDSSpixel}. We also remove all targets (i.e., quasars, spectroscopic galaxies, and random points) that are closer than $1\,$Mpc to the boundary or to a mask in the photometric sample in order to ensure that all targets have a complete field of photometric galaxies within the scales of interest. \section{Analysis} \label{sec.analysis} We study the environment of the spectroscopic targets (quasars and galaxies) by determining the number of photometric galaxies within different projected radii from the quasars, from $25\,$kpc to $1\,$Mpc ($h=0.7$). In order to normalize the density of photometric galaxies, we also estimate the density of photometric galaxies around a large number of random positions in the survey area. This method allows a direct comparison of the observed density around quasars and around galaxies with that found around random positions using the same observed distribution of photometric galaxies. This latter comparison yields normalized overdensities, i.e., observed density over random density, for both the quasars and the spectroscopic galaxies. This technique then allows a direct comparison of the density of galaxies around spectroscopic targets. Throughout the analysis, we treat all of these targets - quasars, spectroscopic galaxies, and random points - in exactly the same manner in order to provide a direct and straightforward comparison between the environment of quasars, galaxies, and random positions, and help minimize potential biases. For each of the targets, we determine the number of photometric galaxies within projected comoving radius bins from $25\,$kpc to $1\,$Mpc ($h=0.7$). The innermost $25\,$kpc (15.3\arcsec to 3.3\arcsec over our redshift range) is removed in all density estimations in order to avoid deblending issues at these small separations. The number of photometric galaxies observed around quasars and around spectroscopic galaxies, $\ngq$ and $\ngg$ respectively, is divided by the same number found around random points, $\ngr$; these overdensities, $\ngq/\ngr$ and $\ngg/\ngr$, are determined for each of the radii specified above. In order to account for the different redshift distributions of the $z \le 0.4$ targets (quasars and galaxies), the overdensities are determined for each individual quasar or galaxy using the mean $\ngr$ appropriate for that target's redshift. All overdensities are then averaged in the relevant redshift bins, and are investigated as a function of radius, luminosity, and redshift. Our environment estimator is less sensitive to faint galaxies at high redshift, but it is still informative as we are interested in an excess in quasar environment density relative to that of galaxies. At low redshift, the average number of photometric galaxies around quasars ranges from a few galaxies within $100\,$kpc (typically twice the number found around random points) to $\sim100\,$ galaxies within $1\,$Mpc. In order to estimate the density errors and the correlations between different scales, we have generated $10^5$ bootstrap samples of the spectroscopic quasar, galaxy and random samples, and measured the standard deviation among different realizations. As the counts of photometric galaxies are dominated by projection effects, i.e. objects uncorrelated to the spectroscopic targets, the error bars are close to Poisson errors, especially on large-scales. \section{Results} \label{sec.results} Our analysis produces normalized overdensities around quasars and around spectroscopic galaxies. We note that the normalized densities we use are defined as ratios of the galaxy counts around each target relative to that around random points (e.g. $\ngq/\ngr$); a ratio of unity implies no overdensity, i.e. the same density around quasars as around random positions. Another common definition of overdensity relative to random can be calculated using $\ngq/\ngr\,-\,1$. This can be directly obtained from the $\ngq/\ngr$ ratios provided below. The quasar overdensity is presented as a function of redshift in Figure~\ref{f.z} within each of our four standard radii. The mean overdensity around quasars is shown by the dashed lines. The overdensity is 2.12 within $0.1\,$Mpc of the quasars; i.e., the density is 2.12 times larger than the density around random points. The mean overdensity decreases monotonically with radius; it is 1.57 within $0.25\,$Mpc, 1.27 within $0.5\,$Mpc, and 1.13 within $1.0\,$Mpc of the quasars. This overdensity refers to the mean of all quasars with $-24.2 \le M_i \le -22$. The excess photometric galaxies refers to galaxies within the magnitude range $14 \le i \le 21$ (Section~\ref{sec.data}). The mean galaxy overdensity around quasars is independent of redshift for $z \le 0.4$ (Figure~\ref{f.z}). In Figure~\ref{f.over} we present the overdensity as a function of luminosity for quasars and for spectroscopic galaxies within radii of 0.1, 0.25, 0.5, and $1\,$Mpc ($h=0.7$). We find that, at all radii, the overdensity around galaxies (solid line) increases with galaxy luminosity. This is expected as brighter galaxies are located, on average, in higher density regions \citep[e.g.,][]{davis_1988, hamilton_1988, white_1988, zehavi_2004, eisenstein_2005}. The galaxy overdensity is greatest on small scales ($0.1\,$Mpc and closer), and decreases on larger scales, as expected. The luminosity-overdensity trend is steeper on small scales than on large scales. The overdensity around $L^*$ galaxies is $1.51\pm0.01$ times larger than random within a radius of $0.1\,$Mpc; the overdensity is nearly doubled, to $2.95\pm0.05$, for galaxies that are brighter by 1 magnitude. The quasar overdensity in Figure~\ref{f.over} shows an increase with quasar luminosity on the smallest scales, but only for the most luminous quasars. The trend is considerably weaker than for galaxies and becomes negligible, with nearly no dependence on quasar luminosity by $\sim0.5-1\,$Mpc scales as well as for quasars with lower luminosities (Figure~\ref{f.over}). This indicates that there is little or no correlation between quasar activity and environment on these larger scales, and that quasar activity is typically triggered by an interaction with a neighbor present within $\sim100\,$kpc from the quasar host galaxy. The mean overdensity around quasars, $<\ngq/\ngr>$, for all $z \le 0.4$ quasars brighter than $M_i \le -22$ is indicated by the horizontal dashed line in Figure~\ref{f.over}. The mean overdensity decreases with radius. The results are summarized in Table~1. \setcounter{table}{0} \begin{table} \caption{Mean Galaxy Overdensity around Quasars ($-24.2 \le M_i \le -22.0$, $z \le 0.4$) and around $L^*$ Galaxies.} \begin{tabular}{cccc} \hline \hline $R_{max}$ (Mpc)& $\frac{\ngq}{\ngr}$&$\frac{\ngl}{\ngr}$&$\frac{\ngq/\ngr}{\ngl/\ngr}$\\ \hline 0.10 & $2.12\pm0.08$ & $1.507\pm0.010$ & $1.41\pm0.06$\\ 0.25 & $1.57\pm0.03$ & $1.235\pm0.004$ & $1.27\pm0.03$\\ 0.50 & $1.27\pm0.02$ & $1.144\pm0.003$ & $1.11\pm0.02$\\ 1.00 & $1.13\pm0.01$ & $1.082\pm0.002$ & $1.05\pm0.01$\\ \hline \end{tabular} \end{table} Figure~\ref{f.over} allows us to compare the environment of quasars with the environment of galaxies of different magnitudes. We use the average overdensity around $L^*$ galaxies as a standard by which to measure the quasar environment. Since quasar luminosity is clearly physically unrelated to the galactic luminosity, it should be noted that $L^*$ galaxies are used only as an frame of reference for comparing the relative environments. We find the mean overdensity around quasars to be larger than around $L^*$ galaxies (shown by the vertical dashed line in Figure~\ref{f.over}) on all scales $<1\,$Mpc. The mean quasar overdensity is larger than the overdensity around $L^*$ galaxies by factors that range from $1.41\pm0.06$ within $0.1\,$Mpc, to $1.27\pm0.03$ within $0.25\,$Mpc, $1.11\pm0.02$ within $0.5\,$Mpc, and $1.05\pm0.01$ within $1\,$Mpc (Table~1). (Using the alternate overdensity definition, $[\ngq/\ngr\,-\,1]\,/\,[\ngl/\ngr\,-\,1]$, we find an excess ratio of $2.2\pm0.16$ within $0.1\,$Mpc decreasing to $1.6\pm0.13$ within $1\,$Mpc. The interpretation is, of course, the same.) The local density enhancement around quasars is similar to the local enhancement around $\sim2L^*$ galaxies. The overdensity around quasars relative to $L^*$ increases to a factor of $2.8\pm0.56$ closest to the quasars, at $\sim40\,$kpc (see below). On scales larger than $1\,$Mpc, the quasar overdensity becomes similar to the environment of $L^*$ galaxies. These results indicate that quasars are located in higher density regions than are $L^*$ galaxies, but that the overdensity exists mostly in regions very close to the quasars ($\lesssim0.1\,$Mpc), and is thus a very local excess. This local excess of galaxies near quasars likely plays an important role in triggering the quasar activity through mergers and other interactions. These results are found to be only weakly dependent on redshift in the $z < 0.4$ range studied here. This is shown in Figure~\ref{f.overz}, where the results, similar to Figure~\ref{f.over}, are presented for different redshift ranges. As can be seen, the trend of overdensity with quasar and galaxy luminosity remains consistent within the statistical uncertainties and no significant trend can be detected as a function of redshift. The redshift independence is further illustrated in Figure~\ref{f.overz2}, where the overdensity around quasars and the overdensity around galaxies are plotted as a function of luminosity for several redshift bins. The results show a redshift independent overdensity signal for both quasars and for galaxies. (The highest luminosity galaxies, which do show some evolution, do not affect our conclusions as we compare our quasars only to $L^*$ galaxies.) Our results for the overdensities and the comparison with the environment of $L^*$ galaxies are unaffected by redshift. The scale dependence of the galaxy overdensity around spectroscopic quasars and galaxies is presented as a function of comoving distance from the target object in Figure~\ref{f.corr}. The quasar overdensities are shown for both the brightest quasars ($M_i = -23.3$ to $-24.2$) and for fainter quasars ($M_i = -22$ to $-23.3$). We find that the quasar overdensities are larger than those of $L^*$ galaxies at all radii less than $\sim0.5\,$Mpc, but the overdensity increases substantially on smaller scales. The overdensities around the most luminous quasars are larger than around fainter quasars, mostly on small scales ($\le\,0.1\,$Mpc). The lower panel of Figure~\ref{f.corr} presents the ratio of quasar overdensity to that of $L^*$ galaxy overdensity, as a function of scale. These data illustrate the transition between the large scales, where quasars and $L^*$ galaxies inhabit similar environments, and the small scales, where quasars are located in higher overdensity regions than are $L^*$ galaxies. On these small scales, the quasar overdensity is comparable to that around $\sim2L^*$ galaxies (or brighter, for the most luminous quasars). Our results on scales greater than $0.5\,$Mpc are consistent with the recent SDSS and 2dF research on the large scale quasar-galaxy clustering \citep[e.g.,][]{smith_1995, croom_1999, croom_2003, wake_2004}, while our results on smaller scales are consistent with the early work discussed in Section~\ref{sec.intro} \citep[e.g.,][]{shaver_1988, shanks_1988, chu_1988, chu_1989, crampton_1989} as well as with recent quasar-quasar pair studies \citep{hennawi_2005}. This difference in the relative quasar environment between small ($\le\,0.5\,$Mpc) and large ($\ge\,1\,$Mpc) scales explains some of the previously contradictory results discussed in Section~\ref{sec.intro}. These results suggest a picture in which quasars reside in, on average, galaxies with a luminosity comparable to $\sim L^*$, but with a local excess of neighbors within $\sim0.1\,-\,0.5\,$Mpc. This local excess is likely associated with triggering quasar activity. The different galaxy densities observed for the bright and faint quasars cannot be due to gravitational lensing. The dark matter distribution in and around galaxies induce gravitational lensing effects which locally enlarge the sky solid angle and magnify the images of background objects. This effect, the magnification bias, can increase or decrease the density of distant sources around foreground galaxies depending on the variation of the number of sources as a function of magnitude \citep{narayan_1989}. It can thus create correlations between source luminosity and foreground galaxy density. However, the amplitude of this effect is expected to be only at the percent level \citep{menard_2002, jain_2003} and recent observations with the SDSS have confirmed these predictions \citep{scranton_2005}. This suggests that gravitational lensing does not significantly affect our density estimations. \section{Conclusions} \label{sec.conclusions} We use the SDSS data to investigate the local environment of quasars and compare it with the environment around galaxies and around random positions. This study provides a direct comparison between the environment of quasars and galaxies. We use bright quasars with $-24.2 \le M_i \le -22$ ($h=0.7$) and redshift $z \le 0.4$ (2028 quasars over $4000\,$deg$^2$) to study the density of photometric galaxies with $i = 14$ to 21 (sample of $\sim10^7$ galaxies) located within $25\,$kpc to $1\,$Mpc ($h=0.7$) of the parent quasars. We compare the results with the same analysis carried out at random positions in the SDSS survey, $\ngr$; this yields the galaxy overdensity, over random, around quasars, i.e., $\ngq/\ngr$. We investigate this overdensity as a function of quasar redshift and luminosity. The same analysis is then repeated for determining the overdensity around $10^5$ spectroscopic galaxies in the SDSS data ($i \le 18.5$) in the same redshift range ($z = 0.08$ to $0.4$). This allows a direct comparison of the quasar overdensity to the overdensity around galaxies. The overdensities are studied as a function of scale, luminosity, and redshift. This comparison of quasar environment with that of galaxies provides a self-consistent comparison of the host environments. Our results are summarized below. 1. At all radii, from $25\,$kpc to $\sim1\,$Mpc, quasars are located in higher density environments than are $L^*$ galaxies. The overdensity around quasars relative to that of $L^*$ galaxies increases with decreasing scale: the overdensity is greatest closest to the quasars. At a distance of $40\,$kpc of the quasars, the mean overdensity for the brightest quasars ($-24.2 \le Mi \le -23.3$, $z \le 0.4$) is nearly a factor of three times larger than the overdensity around $L^*$ galaxies. The mean overdensity around the brightest quasars relative to that of $L^*$ galaxies decreases with increasing scale to a value of 2.2 within $0.1\,$Mpc, 1.2 at $0.3\,$Mpc, and approaches unity at $\sim0.5 - 1\,$Mpc. On these larger scales, quasars reside in environments similar to those of $L^*$ galaxies. 2. The brightest quasars are found to be located in higher overdensity regions than are fainter quasars, especially at small separations ($< 0.1\,$Mpc). On larger scales, bright and faint quasars live in similarly dense environments. 3. The mean overdensity around quasars 0 ($<~\ngq/\ngr~> = 2.12\pm0.08$ within $0.1\,$Mpc, decreasing to $1.13\pm0.01$ within $1\,$Mpc) is independent of redshift for $z \le 0.4$. The mean overdensity is also independent of luminosity except for the brightest quasars, which are located in higher density environments. This dependence of quasar environment on luminosity, showing enhancement only for the most luminous quasars, is consistent with recent galaxy merging models of quasars \citep{hopkins_2005}. 4. The enhanced mean overdensity around quasars is observed to be a local phenomenon, affecting mostly the $\sim$ $0.1\,$Mpc region closest to the quasars. On these scales, very close to the quasars, the high overdensity of galaxies likely affects the formation and triggering of the quasar activity through mergers and other interactions. On scales of $\sim 1\,$Mpc, the quasars inhabit similar environments to those of normal $L^*$ galaxies. \section{Acknowledgements} \label{sec.acknowledgements} GTR acknowledges support from a Gordon and Betty Moore Fellowship in data intensive sciences. Funding for the creation and distribution of the SDSS Archive has been provided by the Alfred P. Sloan Foundation, the Participating Institutions, the National Aeronautics and Space Administration, the National Science Foundation, the U.S. Department of Energy, the Japanese Monbukagakusho, and the Max Planck Society. The SDSS Web site is http://www.sdss.org/. The SDSS is managed by the Astrophysical Research Consortium (ARC) for the Participating Institutions. The Participating Institutions are The University of Chicago, Fermilab, the Institute for Advanced Study, the Japan Participation Group, The Johns Hopkins University, the Korean Scientist Group, Los Alamos National Laboratory, the Max-Planck-Institute for Astronomy (MPIA), the Max-Planck-Institute for Astrophysics (MPA), New Mexico State University, University of Pittsburgh, University of Portsmouth, Princeton University, the United States Naval Observatory, and the University of Washington.
Title: ELT requirements for future observations of the Intergalactic Medium
Abstract: We summarise the science cases for an ELT that were presented in the parallel session on the intergalactic medium, and the open discussion that followed the formal presentations. Observations of the IGM with an ELT provides tremendous potential for dramatic improvements in current programmes in a very wide variety of subjects. These range from fundamental physics (expansion of the Universe, nature of the dark matter, variation of physical constants), cosmology (geometry of the Universe, large-scale structure), reionisation (ionisation state of the IGM at high redshift>6, to more traditional astronomy, such as the interactions between galaxies and the IGM (metal enrichment, galactic winds and other forms of feedback), and the study of the interstellar medium in high redshift galaxies through molecules. The requirements on ELTs and their instruments for fulfilling this potential are discussed.
https://export.arxiv.org/pdf/astro-ph/0601637
\firstsection % \section{Introduction} The advent of echelle spectrographs on 8m class telescopes since the early 1990's has revolutionised our understanding of the intergalactic medium (IGM) as observed in quasar spectra. These bright sources have smooth intrinsic spectra with broad emission lines, yet the {\em observed} spectra contain hundreds of narrow absorption lines due to intervening absorbers. The latter can be studied in great detail from the exquisite, $\rm S/N> 40$, spectra possible with UVES on VLT and HiRes on Keck. Most of the absorption in the UV is due to neutral hydrogen left over from the Big Bang, forming a forest of lines (Bahcall \& Salpter 1965; Gunn \& Peterson 1965; Lynds 1971). The weaker lines with column density $N$({H~{\sc i})$\le 10^{15}{\rm cm}^{-2}$ are traditionally called the \lq Lyman-$\alpha$ forest\rq, with the strongest lines with column density $\ge 10^{20.3}{\rm cm}^{-2}$ that show a measurable damping-wing called \lq Damped Lyman-$\alpha$ systems\rq\, (DLAs). The number of lines as a function of redshift and column-density, $d^2N/dz/dN$(H~{\sc i}), is close to a power-law $\propto~N$(H~{\sc i})$^{\beta}$ with $\beta<0$, as function of column-density, and evolves strongly with redshift as fewer lines are produced as the mean density decreases due to the expansion of the Universe, see Rauch (1998) for a review. The weaker lines originate in the filaments of the cosmic web which itself is a natural outcome of how structure forms in a dark matter dominated cosmology (Bi \etal\ 1992; Cen \etal\ 1994; Schaye 2001). The neutral hydrogen fraction is small at redshifts $\le 6$ (Gunn \& Peterson 1965) as the gas is photo-ionised and photo-heated by the UV-background, with photo-ionisation rate $\Gamma$, produced by galaxies and quasars (Haardt \& Madau 1996). At lower $z\le 2$, the forest of lines thins-out into a Lyman-$\alpha$ \lq savanna\rq\,, but the decline is slowed because $\Gamma$ also decreases as the emissivity from galaxies and QSOs drops (Theuns \etal\ 1998a; Dav\'e \etal\ 1999). Conversely at increasing $z\ge 6$, the mean density increases, but $\Gamma$ also decreases as many source have yet to form, turning the forest into a Lyman-$\alpha$ \lq jungle\rq\, which absorbs (nearly) all light, perhaps signaling the end of reionisation (Becker \etal\ 2001; Djorgovski \etal\ 2001). The forest provides a tremendous probe of how the IGM evolves in the intermediate redshift range $2\le z\le 5$, because the absorbers are only mildly non-linear and hence can be simulated reliably (Cen \etal\ 1994; Hernquist \etal\ 1996; Theuns \etal\ 1998b; Zhang \etal\ 1998; Bryan \etal\ 1999). The combination of superb data with reliable models makes it possible to constrain models and determine parameters. Stronger lines form near galaxies, with the DLAs potential proto-galaxies or proto-galactic lumps (Wolfe 1995; Haehnelt \etal\ 1998; Ledoux \etal\ 1998). Since these systems are discovered in absorption, it is a worry that even denser systems might be missed because they make the background QSO too faint to appear in a magnitude-limited survey. For a recent appraisal of this issue see Ellison \etal\ (2005 and reference therein). DLAs shield the UV-background and some fraction of the gas becomes molecular (see, e.g., Srianand this volume). The prospect of studying star formation in small systems at high redshift which are too faint to study in emission, is very exciting. Quasar spectra also contain \lq metal\rq\, lines from highly ionised species such as \cfour\,, \sifour\, and \osix\, (e.g., Cowie \etal\ 1995). These metals were synthesised in stars and managed to diffuse into the lower density surroundings, either as a result of galactic winds, or due to an early generation of population~III~ stars. The next section gives a short overview of recent results, with emphasis on opportunities for progress with the advent of new observatories. \section{IGM observations with ELT: science} We begin with a short overview of numerical simulations of the IGM, as these can be used to investigate the main limitations of current observational strategies, thereby guiding the design for new instruments. We then discuss current and future science that can be done with IGM observations. The next section summarises the corresponding requirements for an ELT. \subsection{Hydrodynamical simulations} Most of the weaker ``Lyman$-\alpha$ forest'' lines form in mildly over dense or under dense structures that can be simulated accurately (Bi \etal\ 1992; Cen \etal\ 1994). Fig.~\ref{fig:ts_fig1} displays the gas distribution in a cosmological hydrodynamical simulation at $z=3$, and shows that the gas traces the filamentary pattern that results from structure formation in a dark matter Universe. A sight line through such a density distribution will most often go through low density voids, occasionally intersecting a filament which will produce an absorption line, and even more rarely pass close to or even straight through a galaxy halo, producing a very strong absorption line. Mock spectra generated from such simulations look very similar to the real data, see, e.g., Fig.~\ref{fig:ts_fig2}. Since most of the lines are due to structures that are only mildly over dense, it is possible to simulate them quite reliably. Comparison of such simulated spectra with observed ones makes it possible to constrain the model parameters and investigate which cosmological parameters determine the line statistics. Most of the volume in the simulations is photo-heated by the UV-background after reionisation, the volume affected by shocks from structure formation is small. The temperature of the IGM affects the properties of the lines (Theuns, Schaye \& Haehnelt 2000), because the widths of the narrowest lines is restricted by thermal broadening. Detailed comparison with data allows one to constrain the thermal history of the IGM (Schaye \etal\ 2000; Ricotti \etal\ 2000; Bryan \& Machacek 2000; McDonald \etal\ 2001), because the thermal time-scales are long in the low-density IGM. This also puts constraints on reionisation if photo-heating is the dominant heating mechanism (Theuns \etal\ 2002b). The detailed line properties are also sensitive to the nature of the dark matter, and can for example constrain the mass of a putative warm dark matter particle (Croft \etal\ 1999; Viel \etal\ 2005). If the warm dark matter smoothing length is comparable to the width of filaments, then this will affect the line shape. Note that these scales become non-linear at lower $z$, making it much harder to put tight constraints on the dark matter properties. Sight lines passing close to a galaxy may be affected by non-gravitational effects such as feedback from star formation or AGN. Fig.~\ref{fig:ts_fig3} illustrates how supernova feedback causes bigger galaxies to be embedded in a hot bubble of metal enriched gas, which expands into the lower density surroundings of halos. Such sight lines will also show metal line absorption, and it is possible to compare in detail the metal-hydrogen correlation in simulations and data to infer the physical properties in the surroundings of high-$z$ galaxies (Adelberger \etal\ 2003; Pieri \etal\ 2005). ELTs will provide dramatic improvements in this subject, because the bigger collecting area will allow one to observe fainter sources, and hence allow a far finer grid of sight lines probing galaxy environs. \subsection{Fundamental Physics} High resolution high signal-to-noise spectra of high-$z$ QSOs are frequently used to test the current theories of cosmology and fundamental physics.\\ \noindent {\em CODEX:} The observed wavelength of a given absorption line changes with time due to the expansion of the Universe. The COsmic Dynamics EXperiment (CODEX) (see Molaro \etal\ , this volume) aims to measure this change directly by observing many lines in a QSO spectrum at extreme signal-to-noise and resolution, and repeat the measurement a few decades later. Such an experiment therefore also requires very high and long term stability of the spectrograph. A more indirect measure of the expansion of the Universe is by determining the CMB temperature at intermediate redshift using the fine-structure excitation lines of carbon in DLAs, which is excited by CMB photons (see \cite{srianand00}). Detecting C~{\sc i} absorption lines from the low density regions, where the collisional excitation will be sub-dominant using high S/N spectra will allow one to directly map the redshift evolution of temperature of the CMBR (R. Srianand, this volume). Some of the current theories of fundamental physics, such as SUSY, GUT and Super-string theory, allow possible space and time variations of the fundamental constants. QSO absorption lines can be used to probe the time evolution of fundamental constants. The heavy element absorption lines and H$_2$ Lyman Werner band absorptions lines are used to investigate the time-variation of the electromagnetic coupling constant $\alpha$ (see \cite{murphy03}; \cite{chand04} and Mollaro \etal\ this volume) and the proton to electron mass ratio ($\mu$)(see \cite{ivanchik05}), respectively. The available constraints based on 8m class telescopes are still much higher than those achieved by terrestrial techniques. Higher resolution (R$\ge100\,000$) and good signal-to-noise ratios ($>100$) are needed to improve the precession. The Square Kilometer Array (SKA) will measure the 21-cm line in most of the DLAs. The wavelength of the transition depends on $\alpha$, $\mu$ and the proton g-factor and can provide a combined constraint on the variation of all these fundamental constants (see \cite{curran04}). Detecting H$_2$ and weak transitions of Mn~{\sc ii}, Ni~{\sc ii} in DLAs with high signal-to-noise and resolution will allow us to lift the degeneracy between the variation of different fundamental constants that decide the shift of the 21\,cm absorption line. To avoid the systematics caused by the small-scale properties of the lines we require high resolution and high signal-to-noise to improve current constraints. To study the redshift evolution and to be able to use different sets of lines from the same system, it is of paramount importance to have a wide wavelength coverage. \subsection{Cosmology} The large-scale flux distribution can be used to infer the dark matter power spectrum (Croft \etal\ 1998; McDonald \etal\ 2000; Viel \etal\ 2004) and constrain the neutrino mass (Croft \etal\ 1999; Viel \etal\ 2005). These measurements are currently limited by uncertainties in the shape of the QSO's underlying continuum, calibration of the echelle spectrograph for high resolution spectra, and by the statistics of available spectra, and would not obviously benefit from an ELT. However, the influence of large-scale structure is also very prominent in simulations at {\em low} optical depths, below the median $\tau$. Such observations require much higher S/N than currently available, since even S/N=50 data cannot recover the median optical depth at $z\sim 2$. Observations of the forest along parallel sight lines can constrain the topology of the Universe via the Alcock-Paczynski (1979) test (e.g., Rollinde \etal\ 2003). Such a project would benefit greatly from observing fainter QSOs and bright LBGs to improve the sampling in the transverse direction to the line of sight. \subsection{Reionisation} The Lyman-$\alpha$ forest becomes increasingly opaque above $z\ge 6$, perhaps signaling the end of reionisation. If the IGM is polluted through winds from early generations of dwarf galaxies then the ionisation state of the IGM can be probed through the absorption produced by C~{\sc iv}, C~{\sc ii}, O~{\sc i} and Si~{\sc ii} (with NIR wavelength $\lambda<2.1\mu$m for $z<12.5$, 14.7, 15.1 and 15.7, respectively, e.g., Oh 2002). Given the rapidly declining space density of QSOs, Gamma Ray Bursts (GRBs) or super novae could be used as background sources. GRBs have mean afterglow fluxes of 1.5 to 0.05$\mu$J at $z=10$, 1 to 10 days after the explosion ($K_{AB}=23.6$ to 27). High resolution ($R=4\times 10^4$) and high signal-to-noise ratios ($>50$) are required to detect individual lines in the NIR. Observations of a very bright $4\mu$J GRB at $R=10^4$ and S/N=50 gives detection limits for $N$({\rm C~{\sc ii}})=$4\times 10^{12}{\rm cm}^{-2}$ and $N$(O~{\sc i})=1$\times 10^{13}{\rm cm}^{-2}$ . Pair-instability SNe of $M=140-260M_\odot$ pop.~III precursors have $K_{\rm AB}=25$ for $z=10-15$, and are also potential targets, with a possible time-lag of weeks between discovery and ELT follow-up spectroscopy (see the presentation by J. Bergeron in this volume for details). \subsection{Galaxy-Intergalactic medium interactions and metal enrichment} The metal density of carbon as inferred from C~{\sc iv} pixel optical depth analysis (Songaila 2001; Schaye \etal\ 2003) shows little evidence for evolution over the redshift range $z=2-5$, with possibly a decline by factor of two above $z=6$. It is possible that not all metals are seen. Just as most metals are in the hot intra-cluster gas at $z<1$, metals could be in hot gas resulting from galactic winds at higher $z$, thereby not producing significant C~{\sc iv} absorption (Theuns \etal\ 2002a and this volume). The shape of the UV-background, and its evolution with $z$, is the main uncertainty in converting optical depth to metallicity. Improved constraints require the detection of many more transitions to eliminate this uncertainty. What is the origin of the metals seen in the IGM? Are the metals due to galactic winds, or is some fraction the result of pop.~III stars? This important question can be addressed by correlating metals seen in absorption with the presence of galaxies (e.g., Adelberger \etal\ 2003, 2005; Pieri, Schaye \& Aguirre 2005). This can be done by probing the IGM with many sight lines, and requires obtaining spectra of fainter sources, including the brighter Lyman-break galaxies (LBGs) themselves. Current state of the art (Adelberger \etal\ 2005) is limited to measuring the mean metallicity in C~{\sc iv} as function of the galaxy's impact parameter; higher S/N should make it possible to look for metals in each individual galaxy spectrum and obtain good redshifts. A better understanding of galaxy-IGM interactions is needed to constrain how feedback from stars and AGN affects galaxy formation as a function of redshift and galaxy mass. The redshift range $z\sim 3$ is well suited for such a study, as there are many lines in the observed optical-NIR part of the spectrum suited to ground-based observations, but an ELT is required to be able to observe fainter QSOs or brighter LBGs and dramatically improve the sampling with many more sight lines. \subsection{Molecules at high $z$} Detecting H$_2$ and other molecules at high-z through their electronic transitions is important for understanding the physical conditions and astrochemistry in the interstellar medium of galaxies and protogalaxies at a very early epoch. Up to now, H$_2$ has been detected in $\sim15\%$ of DLAs (\cite{ledoux03}) and only one system shows detectable HD (Varshalovich \etal\ 2001). H$_2$ can potentially be detected from LBGs and GRB host galaxies. This will allow us to understand the interstellar medium in these early galaxies. As the Lyman Werner band absorptions of H$_2$ are expected in the \lya forest, it is important to have high resolution R = 20\,000 and signal-to-noise ($>20$) in the blue spectrum. ELTs with blue sensitive spectrographs can allow us to search for H$_2$ in fainter QSOs, GRBs and brighter LBGs. However, in the case of GRBs, H$_2$ may be in non-equilibrium and it is important to target the source as quickly as possible to be able to detect the H$_2$ lines and follow the time variation of H$_2$ column density. This will give important clues about the GRB hosts. Detecting other molecules in systems with H$_2$ is also important for the understanding of astrochemistry in low metallicity gas in the early universe. CO is not detected in DLAs and the achieved limits are close to the lowest column measured in Milky Way. As the metallicities in DLAs are low, to test N(H$_2$) vs. N(CO) relation we need to push this limit by roughly a factor 50 (see the presentation by R. Srianand in this volume). \section{IGM observations with ELT: requirements} In this section we summarise the ELT requirements that emerge out of the discussion in the parallel session on IGM. \noindent{\em {High $z>7$}}: {\bf Reionisation, metals, Lyman-$\alpha$ emitters:} Constraining IGM enrichment and its ionisation state from metal lines at $z>7$ requires observations at intermediate resolution of $R=2000$ in the NIR with S/N up to 100. An OH-line suppressor with multiple IFUs with field-of-view of several arcmin$^2$ is ideal. Targets are moderately faint QSOs and Lyman-break galaxies of $m_{AB}\sim 27$, but require an ELT larger than 30m. NIR observations at $R=10^4$ with S/N up to 100 is possible from average-luminosity GRBs and pop.~III SNe. Targets need to be found using dedicated ground and space-based telescopes. \begin{enumerate} \item[$\bullet$] NIR, R=2000, S/N=100 (bright QSOs) \item[$\bullet$] NIR, R=10000, S/N=100 (single target GRBs) \end{enumerate} \noindent{\em {Intermediate $z<7$}: {\bf Metals:}} High resolution $R=40000$ and S/N of 10000 optical spectra of bright QSOs ($z=2-5$, $m_{AB}=16-17$) and bright GRBs ($z$ up to 7, lag is 1 day, $m_{AB}=20$, S/N=100) in single target mode are required to study the distribution of metals in the IGM, and its evolution with $z$. The spectrograph should be {\em blue sensitive} and have a large wavelength range ($\lambda=3030-9300\AA$) to be able to cover a large range of transitions and constrain the ionisation corrections. The latter is the major uncertainty in inferring metallicity, so a large $\lambda$ range is essential. \begin{enumerate} \item[$\bullet$] optical, R=40000, S/N=$10^3-10^4$ (single target bright QSOs, GRBs). Blue sensitive, large $\lambda$ ($3030-9300\AA$) coverage. \end{enumerate} \noindent{\em {Lower $z<5$}: {\bf galaxy-IGM connection, UV-escape from galaxies:}} The main gain of an ELT is the possibility of observing fainter QSOs, which allows one to sample the metal distribution in the IGM, and its correlation with galaxies dramatically better by providing a much finer grid of lines along which the IGM can be probed. The QSOs and bright LBGs can be observed with optical, high $R=50000$ spectroscopy (S/N=100) to probe the distribution of metals. A detailed correlation of these metals with galaxies requires the redshift determination of the fainter LBGs (up to 0.01$L_\star$) using optical/NIR MOS of $R=2000-5000$, with multiple IFUs with a total FoV of several arcmin$^2$, centered on LBGs and QSOs. NIR is required to obtain good redshifts for the galaxies from stellar and ISM lines, since many of the UV-lines can be significantly off-set from the redshift of the stars. \begin{enumerate} \item[$\bullet$] Optical/NIR, R=2000-5000, S/N=100 (0.01$L_\star$~LBGs) with multiple IFUs, FoV several arcmin$^2$ \item[$\bullet$] Optical, R=50000, S/N=100 (bright LBGs, QSOs) \end{enumerate} Many small programmes could be started at early stages of construction if the instruments are available. 8m-class telescopes will be used to find (candidate) LBGs and Lyman-$\alpha$ emitters. \begin{acknowledgments} We wish to thank IAU for a travel grant. TT thanks PPARC for the award of an Advanced Fellowship, and J Schaye, R Bower and I Smail for comments on the draft. \end{acknowledgments}
Title: The First Scientific Results from the Pierre Auger Observatory
Abstract: The southern site of the Pierre Auger Observatory is under the construction near Malargue in Argentina and now more than 60% of the detectors are completed. The observatory has been collecting data for over 1 year and the cumulative exposure is already similar to that of the largest forerunner experiments. The hybrid technique provides model-independent energy measurements from the Fluorescence Detector to calibrate the Surface Detector. Based on this technique, the first estimation of the energy spectrum above 3 EeV has been presented and is discussed in this paper.
https://export.arxiv.org/pdf/astro-ph/0601035
\title{ The First Scientific Results from the Pierre Auger Observatory } \classification{95.85.Ry, 98.70.Sa } \keywords {Cosmic Rays, Ultra-High Energy Particles} \author{T. Yamamoto} { address={KICP, Enrico Fermi Institute, University of Chicago, 5640 S. Ellis Ave, Chicago IL 60637, USA } } \author{The Pierre Auger Observatory Collaboration}{ address={} } The Pierre Auger Observatory is the largest cosmic ray detector ever built to study the Ultra-High Energy Cosmic Rays (UHECR) with unprecedented statistics and high precision \cite{AUGER}. In particular, it is important to address whether the cosmic-ray spectrum continues beyond $10^{20}$ eV. Due to the interaction with microwave background photons, a steepening is expected around $10^{20}$ eV in the energy spectrum if the sources are distributed uniformly throughout the Universe. This conclusion is independent of the composition of the UHECR's. Recent measurements of the energy spectrum by the AGASA which used surface detector (SD) array \cite{AGASA} and the HiRes which is using fluorescence detector (FD) \cite{HyRes} have yielded conflicting results. There are serious limitations in the use of only the SD or the FD alone to measure the primary spectrum. The SD provides high event statistics with high efficiency and robust exposure estimation. The SD energy estimation, however, traditionally relies on Monte-Carlo simulations which require assumptions about the hadronic-interaction model and the primary-chemical composition. On the other hand, the FD provides a calorimetric energy measurement but the estimation of the exposure has a comparatively large uncertainty relative to the SD. Based on one year operation of a portion of the Pierre Auger Observatory, the first scientific results were released this summer concerning the upper limit of the UHE gamma ray flux \cite{Photon}, anisotropy of the arrival directions \cite{Aniso}, and the energy spectrum \cite{Spect}. The cumulative exposure, 1750 $km^2$-$sr$-$yr$, is similar to those achieved by the largest forerunner experiments. Statistical uncertainties are still too large to draw any firm conclusions ether rejecting or confirming results obtained by previous experiments. However, there is an important step achieved in these results. The Pierre Auger Observatory was designed as a hybrid detector to observe the shower particles at ground level by the SD and the associated fluorescence light generated in the atmosphere by the FD. Combining the strengths of the SD and the FD, we have developed a reliable estimate of the primary energy spectrum using the full SD exposure without making assumptions about the primary masses or hadronic model. \\ The southern site of the Pierre Auger Observatory is now under construction on an Argentinian pampa ($35^{\circ}$ S, $69^{\circ}$ W, 1400 m.asl, 875.5 g/cm$^2$). The SD consists of 1600 water Cherenkov tanks planed on a triangular 1.5 km grid covering 3000 $km^2$ area with $2\pi$ sky coverage. The construction of the Southern site is now 60\% complete. The whole area of the SD will be overlooked by an FD from 4 sites. Each FD site has 6 telescopes and each telescope has a $30^{\circ}\times28.6^{\circ}$ field of view with $1.5^{\circ}$ pixel size. Three FD sites are completed and operating now and one is under construction. The events recorded in the SD are reconstructed using the arrival time and the signal size from the shower particles reaching the detectors. The magnitude of the signal at 1 km from the shower axis, S(1000) in Vertical Equivalent Muon (VEM), is estimated from the Lateral Distribution Function fit as a size parameter of the shower \cite{LDF}. Two cosmic rays of the same energy, but incident at different zenith angles, will yield different values of S(1000) due to an attenuation of the shower in the atmosphere. This attenuation is measured by the well-established technique of the constant intensity cut (CIC) method. The principle of this method is that the nearly isotropic intensity of cosmic rays means that the integrated intensity above any given energy must be the same at all zenith angles ($\theta$ degree). One finds the S(1000) at every zenith angle that corresponds to a single primary energy by varying S(1000) at each zenith angle to obtain a fixed integral intensity. Based on this method, the zenith angle dependence of S(1000), the CIC curve is obtained as \begin{equation} S(1000)_{38^{\circ}} = \frac{S(1000)_{\theta}}{1.049+0.0097\theta - 0.00029\theta^2} \end{equation} where $S(1000)_{38^{\circ}}$ VEM is S(1000) adjusted to $\theta=38^{\circ}$. (The median zenith angle of the showers is $38^{\circ}$.) The link between $S(1000)_{38^{\circ}}$ and the primary energy can be established using data from the FD. On dark dry nights, the fluorescence signals are observed simultaneously with the SD events. The fit to the FD-energy as a function of $S(1000)_{38^{\circ}}$ is \begin{equation} log(E) = -0.79+1.06 log(S(1000)_{38^{\circ}}) \label{eq:energy} \end{equation} where $E$ is the FD-energy in EeV. The events detected by the SD are selected as follows: The estimated energy must be greater than 3 EeV because detection efficiency is saturated (nearly 100\%) above this energy. The zenith angle of the arrival direction must be smaller than 60$^{\circ}$. And the event must fall within a well-defined fiducial area. The estimate of the SD exposure is simple. The fiducial area is monitored in the trigger system so that exposure is calculated as the time integration of the aperture given by the fiducial area and the 60$^{\circ}$ zenith-angle limit. The spectrum is then obtained by dividing the number of events in given energy intervals by the exposure as shown in Figure.\ref{fig:spectrum}. The systematic uncertainty of the energy spectrum comes mainly from the energy assignment. In the estimation of the FD-energy, there are several uncertainties which include the fluorescence yield (15\%), missing energy carried by high-energy muons and neutrinos (4\%), the absolute calibration of the FD telescopes (12\%), and atmospheric condition (10\%). Overall the uncertainty of the FD-energy is about 25\%. These systematic errors will be reduced significantly in a year with completion of the FD calibration and the measurement of the fluorescence yield in laboratories . The statistical uncertainty in Equation.\ref{eq:energy} causes additional energy-dependent systematic uncertainty in the energy estimation. This uncertainty is dominant to the systematic error in the highest energy and will automatically shrink with the rapidly-increasing hybrid statistics. The total systematic error is indicated in the Figure.\ref{fig:spectrum}. It should be noted that this energy spectrum was measured in the southern sky which could differ from that of northern sky measured in the previous experiments. The energy scale based on the FD measurements is systematically lower than that from an SD analysis that uses QGSJetII simulations with proton primaries. The difference is similar to the conflicting energy scales of the HiRes and the AGASA collaborations. The exposure of the southern observatory is expected to increase by a factor of 5$\sim$7 over the next two years. With completion of the FD calibration, the statistical and systematic errors will shrink accordingly, permitting a study of spectral features and the energy scale. This work was supported in part by the Kavli Institute for Cosmological Physics through the grant NSF PHY-0114422, by NSF AST-0071235, and DE-FG0291-ER40606 at the University of Chicago. \begin{center} \end{center}
Title: Time delay of SBS 0909+532
Abstract: The time delays between the components of a lensed quasar are basic tools to analyze the expansion of the Universe and the structure of the main lens galaxy halo. In this paper, we focus on the variability and time delay of the double system SBS 0909+532A,B as well as the time behaviour of the field stars. We use VR optical observations of SBS 0909+532A,B and the field stars in 2003. The frames were taken at Calar Alto, Maidanak and Wise observatories, and the VR light curves of the field stars and quasar components are derived from aperture and point-spread function fitting methods. We measure the R-band time delay of the system from the chi-square and dispersion techniques and 1000 synthetic light curves based on the observed records. One nearby field star (SBS 0909+532c) is found to be variable, and the other two nearby field stars are non-variable sources. With respect to the quasar components, the R-band records seem more reliable and are more densely populated than the V-band ones. The observed R-band fluctuations permit a pre-conditioned measurement of the time delay. From the chi-square minimization, if we assume that the quasar emission is observed first in B and afterwards in A (in agreement with basic observations of the system and the corresponding predictions), we obtain a delay of - 45 (+ 1)/(- 11) days (95% confidence interval). The dispersion technique leads to a similar delay range. A by-product of the analysis is the determination of a totally corrected flux ratio in the R band (corrected by the time delay and the contamination due to the galaxy light). Our 95% measurement of this ratio (0.575 +/- 0.014 mag) is in excellent agreement with previous results from contaminated fluxes at the same time of observation.
https://export.arxiv.org/pdf/astro-ph/0601473
\title{Time delay of SBS 0909+532} \author{A. Ull\'an\inst{1} \and L. J. Goicoechea\inst{1} \and A. P. Zheleznyak\inst{2} \and E. Koptelova\inst{3} \and V. V. Bruevich\inst{3} \and T. Akhunov\inst{4} \and O. Burkhonov\inst{4}} \offprints{A. Ull\'an} \institute{Departamento de F\'{\i}sica Moderna, Universidad de Cantabria, Avda. de Los Castros s/n, 39005 Santander, Spain\\ \email{aurora.ullan@postgrado.unican.es, goicol@unican.es} \and Institute of Astronomy of Kharkov National University, Sumskaya 35, 61022 Kharkov, Ukraine\\ \email{zheleznyak@astron.kharkov.ua} \and Sternberg Astronomical Institute, Universitetski pr. 13, 119992 Moscow, Russia\\ \email{koptelova@xray.sai.msu.ru, bruevich@sai.msu.ru} \and Ulug Beg Astronomical Institute of Uzbek Academy of Science, Astronomicheskaya. Str. 33, 700052 Tashkent, Republic of Uzbekistan\\ \email{talat77@rambler.ru, boa@astrin.uzsci.net}} \date{Accepted January 12, 2006} \titlerunning{Time delay of SBS 0909+532} \authorrunning{A. Ull\'an et al.} \abstract{ The time delays between the components of a lensed quasar are basic tools to analyze the expansion of the Universe and the structure of the main lens galaxy halo. In this paper, we focus on the variability and time delay of the double system SBS 0909+532A,B as well as the time behaviour of the field stars. We use $VR$ optical observations of SBS 0909+532A,B and the field stars in 2003. The frames were taken at Calar Alto, Maidanak and Wise observatories, and the $VR$ light curves of the field stars and quasar components are derived from aperture and point--spread function fitting methods. We measure the $R$--band time delay of the system from the $\chi^2$ and dispersion techniques and 1000 synthetic light curves based on the observed records. One nearby field star (SBS 0909+532c) is found to be variable, and the other two nearby field stars are non--variable sources. With respect to the quasar components, the $R$--band records seem more reliable and are more densely populated than the $V$--band ones. The observed $R$--band fluctuations permit a pre--conditioned measurement of the time delay. From the $\chi^2$ minimization, if we assume that the quasar emission is observed first in B and afterwards in A (in agreement with basic observations of the system and the corresponding predictions), we obtain $\Delta \tau_{BA}$ = $-$ 45 $^{+ 1}_{-11}$ days (95\% confidence interval). The dispersion technique leads to a similar delay range. A by--product of the analysis is the determination of a totally corrected flux ratio in the $R$ band (corrected by the time delay and the contamination due to the galaxy light). Our 95\% measurement $\Delta m_{BA}$ = $m_B(t + \Delta \tau_{BA}) - m_A(t)$ = 0.575 $\pm$ 0.014 mag is in excellent agreement with previous results from contaminated fluxes at the same time of observation. \keywords{ Gravitational lensing -- Quasars: general -- Quasars: SBS 0909$+$532 -- Stars: variables: general } } \section{Introduction} The system SBS 0909+532 was discovered by Stepanyan et al. (1991). Some years later, a collaboration between the Hamburger Sternwarte and the Harvard--Smithsonian Center for Astrophysics resolved the system into a pair of quasars (A and B) with a direct $R$--band flux ratio (at the same time of observation) $\Delta m$ = $m_B - m_A$ = 0.58 mag and a separation of about 1\farcs1 (Kochanek et al. 1997). The direct $R$--band flux ratio was not consistent with the direct flux ratios at other wavelengths: $\Delta m$ = 0.31 mag in the $I$ band and $\Delta m$ = 1.29 mag in the $B$ band. From observations with the 4.2 m William Herschel Telescope, a Spanish collaboration got spectra for each component of the system. The data showed that the system consists of two quasars with the same redshift ($z_s$ = 1.377) and identical spectral distribution, supporting the gravitational lens interpretation of SBS 0909+532 (Oscoz et al. 1997). Oscoz et al. (1997) detected a \ion{Mg}{ii} doublet in absorption at the same redshift ($z_{abs}$ = 0.83) in both components, and they suggested that the absorption features were associated with the photometrically unidentified lensing galaxy. Through a singular isothermal sphere (SIS) lens model, the authors also inferred the first constraint on the time delay between the components: $|\Delta \tau_{BA}| \leq$ 140 days, where $\Delta \tau_{BA}$ is the delay of B with respect to A and the Hubble constant is assumed to be $H_0$ = 70 km s$^{-1}$ Mpc$^{-1}$. In recent years, Lubin et al. (2000) indicated the possible nature of the main deflector (early--type galaxy) and confirmed its redshift ($z_d$ = 0.830). Leh\'ar et al. (2000) reported on a program including Hubble Space Telescope (HST) observations of SBS 0909+532. They discovered the main lens galaxy between the components, which has a large effective radius, with a correspondingly low surface brightness. This lens galaxy is closer to the brightest component (A), which is not in contradiction with SIS--like lens models when the farther and fainter component (B) is stronger affected by dust extinction (see below). The colors of the lens are consistent with those of an early--type galaxy at redshift 0.83. Assuming a singular isothermal ellipsoid (SIE) model, Leh\'ar et al. predicted a time delay $\Delta \tau_{BA}$ in the range [$-$ 10, $-$ 87] days ($H_0$ = 70 km s$^{-1}$ Mpc$^{-1}$). At a given emission time, the sign "$-$" means that the corresponding signal is observed first in B and later in A. The COSMOGRAIL collaboration provided the distribution of predicted time delays of the system (Saha et al. 2005). In their histogram (Fig. 10 of Saha et al.), there are two features: the main feature is an asymmetric peak around $-$ 80 days and the secondary one is another asymmetric peak around $-$ 45 days. Therefore, if the COSMOGRAIL predictions are right, the time delay is very probably of 2--3 months (component B leading component A), but we cannot rule out a delay of about one and a half months. On the other hand, the flux ratio anomaly pointed by Kochanek et al. (1997) was confirmed and accurately studied by Motta et al. (2002) and Mediavilla et al. (2005), who reported the existence of differential extinction in the main lens galaxy. Chartas (2000) and Page et al. (2004) also studied the system in the X--ray domain. Time delays are basic tools to discuss the present expansion rate of the Universe and the structure of the main lens galaxy haloes (e.g., Refsdal 1964; Kochanek, Schneider \& Wambsganss 2004), so that variability studies are crucial. While some time delays have been measured from radio light curves (PKS 1830$-$211: Lovell et al. 1998; Q0957+561: Haarsma et al. 1999; B0218+357: Biggs et al. 1999; B1600+434: Koopmans et al. 2000; B1422+231: Patnaik \& Narasimha 2001; B1608+656: Fassnacht et al. 2002) or X--ray variability (e.g., Q2237+0305: Dai et al. 2003), an important set of delays are based on optical monitoring of gravitationally lensed quasars. Optical frames taken at Apache Point Observatory, Fred Lawrence Whipple Observatory and Teide Observatory were used to estimate a 14--month delay for the double system Q0957+561 (e.g., Pelt et al. 1996; Kundi\'c et al. 1997; Serra--Ricart et al. 1999; Ovaldsen et al. 2003). Although the time delay of this first multiple quasar has been confirmed through independent observations, the measurement is only 5\% accurate, or equivalently, there is an uncertainty of about 20 days (Goicoechea 2002). The Tel--Aviv University (TAU) group have recently determined the time delay between the two components of HE 1104$-$1805 (Ofek \& Maoz 2003). The TAU delay of HE 1104$-$1805 disagrees with the earlier estimation by Gil--Merino, Wisotzki \& Wambsganss (2002), but it is in excellent agreement with the determination by Wyrzykowski et al. (2003). Schechter et al. (1997) measured two delays for the quadruply imaged quasar PG 1115+080. The Belgian--Nordic collaboration carried out a very intense activity during the past five years. They participated in several monitoring projects and measured several time delays at optical wavelengths: B1600+434 (Burud et al. 2000), HE 2149$-$2745 (Burud et al. 2002a), RXJ 0911.4+0551 (Hjorth et al. 2002), SBS 1520+530 (Burud et al. 2002b) and FBQ 0951+2635 (Jakobsson et al. 2005). The formal accuracies of these 5 estimations range from 5 to 25\% (the 1$\sigma$ error bars vary from 4 to 24 days). Kochanek et al. (2005) also measured the time delays between the components of the quadruple quasar HE 0435$-$1223. The aim of this paper is to present $VR$ observations of SBS 0909+532 in 2003 conducted by the University of Cantabria (UC, Spain), the Institute of Astronomy of Kharkov National University (IAKhNU, Ukraine), the Sternberg Astronomical Institute (SAI, Russia) and the Ulug Beg Astronomical Institute of Uzbek Academy of Science (UBAI, Uzbekistan). We also present TAU observations of the field stars in 2003, which have been kindly made available to us. This new optical monitoring campaign was carried out at the Calar Alto Observatory (Spain), the Maidanak Observatory (Uzbekistan) and the Wise Observatory (Israel), and the frames were taken with the 1.5 m Spanish telescope, the 1.5 m AZT$-$22 telescope at Mt. Maidanak and the Wise Observatory 1 m telescope (Section 2). In Section 3, we describe the methodology to obtain the fluxes of the quasar components and the field stars. The $VR$ light curves are also shown in Section 3. Section 4 is devoted to the time delay estimation from the light curves of A and B (quasar components) in the $R$ band. Finally, in Section 5 we summarize our conclusions and discuss the feasibility of an accurate determination of the cosmic expansion rate and the surface density in the main lensing galaxy. \section{Observations} We have three different sets of frames for SBS 0909+532. The first set of optical frames cover the period between 2003 March 4 and June 2, and they are part of a UC project to test the feasibility of quasar monitoring programs through 1$-$2 m telescopes in Spain (Ull\'an 2005). These observations were made with the 1.52 m Spanish telescope at Calar Alto Observatory (EOCA), Almeria, Spain (see Ziad et al. 2005 for a site--testing on Calar Alto). The EOCA is equipped with a Tektronics 1024$\times$1024 CCD detector, which has pixels with a physical size of 24 $\mu$m, giving a 0.4 arcsec pixel$^{-1}$ angular scale. The gain is 6.55 e$^{-}$/ADU and the readout noise is 6.384 e$^{-}$. During this first monitoring, exposures in the $V$ and $R$ Johnson--Cousins filters were taken every night when clear, what makes a total of 20 observing nights. Bad weather in 2003 March and April prevented us from achieving a very dense sampling. For each monitoring night we have three consecutive frames on each filter, i.e., three 300 s exposures in the $V$ passband and three 180 s exposures in the $R$ passband. Those were the maximum exposure times to avoid saturation of selected stars in the field. In Figure 1 we show a typical frame. In this typical exposure, half a dozen bright and non--saturated stars were fitted within the field of view (FOV). Following the notation of Kochanek et al. (1997), the FOV included the gravitationally lensed quasar ("GL") and nearby field stars "a" (South), "b" (North) and "c" (West). The FOV also included two stars that were introduced by Nakos et al. (2003) and were labelled as "s1" and "s2". These two stars are placed relatively far from the gravitational lens system, and they appear close to the North--West edge of the frame (see Fig. 1). A sixth star ("x") appears close to the South--West edge of the typical frame. The second set of observations include frames in February 2003 as well as during April--May and October--November 2003. The total number of nights is 18. In this second program the images were taken with the 1.5 m AZT$-$22 telescope at Maidanak Observatory (Uzbekistan), with near diffraction--limited optics and careful thermo--stabilization, which allow for high--angular--resolution imaging. The AZT$-$22 telescope has a LN--cooled (liquid nitrogen cooled) CCD--camera, SITe- 005 CCD, manufactured in Copenhagen (Denmark). For this camera, the imaging area is split into 2000$\times$800 pixels, where the pixel size is 15 $\mu$m and the intrinsic angular scale is 0.26 arcsec pixel$^{-1}$. The frames were taken in the $R$ Bessel filter, which corresponds approximately to the $R$ Johnson--Cousins passband. The poor tracking system of this telescope allows only exposures up to 3 minutes. To obtain sufficiently high photometric accuracy, we took several frames each observation night. With respect to the rectangular FOV of the telescope, the North/South coverage was 2.5 times smaller than the East/West one, so the "s1", "s2" and "x" stars were not included within the FOV. Figure 2 shows a zoomed--in image made from one of the best frames in terms of seeing. There are two close quasar components, but the very faint galaxy is not apparent. The observations at Mt. Maidanak are part of IAKhNU, SAI and UBAI projects to follow up the variability of gravitationally lensed quasars. \begin{table} \centering \begin{tabular}{ccc} \hline\noalign{\smallskip} Observatory (Telescope) & Frames/night (Filter) & Observation Periods \\ \noalign{\smallskip}\hline\noalign{\smallskip} Calar Alto (1.5 m) & 3 $\times$ 300 s ($V$) + 3 $\times$ 180 s ($R$) & March--June \\ Maidanak (1.5 m) & (3--11) $\times$ 180 s ($R$) & February, April--May, October--November \\ Wise (1.0 m) & 1 $\times$ 420 s ($R$) & 22 unevenly distributed nights \\ \noalign{\smallskip}\hline \end{tabular} \caption{Observations of SBS 0909+532 in 2003.\label{tbl-1}} \end{table} For the past six years the TAU group have been monitoring several gravitationally lensed quasars with the Wise Observatory 1 m telescope. The targets are mainly monitored in the Johnson--Cousins $R$--band, and the frames are obtained with a cryogenically cooled Tektronix 1024$\times$1024--pixel back--illuminated CCD. The angular scale is 0.7 arcsec per pixel. This pixel scale and the median seeing ($FWHM$) of about 2\arcsec\ do not allow resolving most of the lensed objects, e.g., SBS 0909+532. However, the frames of SBS 0909+532 in 2003 are characterized by wide FOVs, which incorporate the "a--c", "s1--s2" and "x" stars. This fact permits to do differential photometry between several pairs of field stars, and thus, to test the reliability of the Calar Alto and Maidanak records. The pre--processing of the images included the usual bias subtraction, flat fielding using sky flats, sky subtraction and cosmic ray removal by using the Image Reduction and Analysis Facility (IRAF) and Munich Image Data Analysis System (MIDAS) environments. Some details about the whole observational campaign are included in Table 1 (observatories, telescopes, frames/night, filters and observation periods). \section{Photometry and $VR$ light curves} Due to the small angular separation between the two lensed components, about 1\farcs1 (Kochanek et al. 1997), the photometry of SBS 0909+532 is a difficult task. This task is also complicated by the presence of the main lensing galaxy between the components, which could make the computation of individual fluxes even harder. In general, aperture photometry does not work, so we must look for better approaches. An initial issue is to decide about the inclusion or not inclusion of a photometric model for the lensing galaxy. In principle, when computing the fluxes of SBS 0909+532 we may use a galaxy model derived from the HST images of the system. The galaxy model could also be inferred from the best images in terms of seeing. Once the relevant information on the galaxy is known, we would apply a PSF fitting method to all optical images, setting the galaxy properties to those derived from the HST or the best--quality images, and allowing the remaining parameters to vary (e.g., McLeod et al. 1998; Ull\'an et al. 2003). Magain, Courbin \& Sohy (1998) also presented an alternative task (deconvolution) that combines all the frames obtained at different epochs to determine the numerical light distribution of the lensing galaxy as well as the positions of the point--like sources (quasar components), since these parameters do not vary with time. The flux of the point--like sources are allowed to vary from image to image, which produces the light curves. However, these and other procedures have a reasonable limitation: they only work well when the galaxy light has a significant contribution to the crowded regions in the individual frames. For a very faint galaxy in a standard (i.e., not superb) frame, there is confusion between galaxy signal and noise, so the use of a given galaxy model could lead to biased fluxes of the components. The biases will depend on the quality of the image (seeing, signal--to--noise ratio, etc), which must produce artificial variability superposed to the real one. On the other hand, the use of a direct PSF fitting method (neglecting the galaxy brightness) leads to contaminated fluxes of the components. But if the galaxy is very faint, the contaminations will be small. Moreover, the variation of the quasar fluxes, seeing conditions, etc, will cause fluctuations in the contaminations, which are expected to be below the typical contamination levels. For standard frames of a quasar lensed by a very faint extended object, it is really difficult to choose between both approaches (with and without galaxy). Most of the Calar Alto and Maidanak individual frames of SBS 0909+532 do not show evidences for a galaxy brightness profile. This fact is due to the faintness of the galaxy, as we corroborate here below. If we consider an hypothetical astronomer that neglects the galaxy brightness and does direct PSF fitting (without taking into account the galaxy when doing the computation of the fluxes), it is possible to attain a rough estimation of the maximum contamination from the galaxy to the closest component A (at 0\farcs4 from the centre of the deflector). We take into account the paper about 10 lens systems by Leh\'ar et al. (2000), where, in Table 3, we can find the best available photometric and astrometric (HST) data of SBS 0909+532. The authors were able to trace the galaxy light in the $H$ passband, by measuring its position and brightness. If we use the colors in the same table, we conclude that $m_{gal} \sim$ 19 mag and $m_A \sim$ 16 mag in the $I$ band (near--IR), and $m_{gal} >$ 20.4 mag and $m_A \sim$ 16.7 mag in the $V$ optical band. Therefore, as the $R$ filter is placed just between the $I$ filter and the $V$ one, we may assume that $m_{gal} - m_A \sim$ 3.5 mag in the $R$ band. The difference of 3.5 mag is consistent with a ratio of fluxes $F_{gal}/F_A$ of about 1/25. Thus, in the case of QSO 0957+561 we found a $R$--band ratio of fluxes $F_{gal}/F_A$ of about 1/2.5 (Ull\'an et al. 2003), and now we have $F_{gal}/F_A \sim$ 1/25, what explains our unsuccessful efforts when measuring the flux of the lens galaxy in standard frames. As a result of that, in an extreme case (when direct PSF fitting leads to a magnitude $m_{A + gal}$ instead of $m_A$, i.e., all the galaxy light is included in the profile of the A component) we find a relationship: $m_A = m_{A + gal} + F_{gal}/F_A$, where the true flux (in magnitudes) $m_A$ differs from the contaminated flux through direct PSF fitting ($m_{A + gal}$) in a quantity $F_{gal}/F_A$. This maximum contamination of A would be only of 40 mmag, and the real contamination of both components will be less than our upper limit. The artificial fluctuations (caused by variable contamination) will be even smaller than the typical contamination levels, so we expect they will not play an important role in analyses of quasar variability (e.g., time delay estimates). In order to derive the light curves of the components A and B, we decide to use a direct PSF fitting method and do not consider the galaxy brightness in the fits. The key idea of this procedure is to obtain the different fluxes we are interested in by using a PSF that comes from a bright star in the field common to all frames. The point--like objects (quasar components and stars) are modelled by means of the empirical PSF. Hence, we do not use a theoretical PSF (i.e., Gaussian distribution, Lorentzian distribution, etc), but the two--dimensional profile of a star in this field (a PSF star). Apart from a PSF star, we also need a reference star to do differential photometry and to obtain relative fluxes $m_A - m_{ref}$ and $m_B - m_{ref}$. The good behaviour of the reference star is usually checked by using a control star, so the fluxes $m_{con} - m_{ref}$ are expected to agree with a constant level. Nevertheless, since the $R$--band flux ratio is discussed in Section 5, we also want to obtain a rough estimation of the contaminations from this direct technique. With this aim, a deconvolution technique (Koptelova et al. 2005) is also applied to a set of frames with good seeing and signal. The selected frames are fitted to a model including the galaxy, and thus, we are able to obtain a few clean fluxes of components A and B and compare them with the corresponding contaminated fluxes (through a direct PSF fitting). The averaged contaminations are used in Section 5. \subsection{Calar Alto frames and light curves} We adopt a model of the system including two point--like sources and a constant background. The model is fitted to each image by adjusting its 7 free parameters (two--dimensional positions of A and B, instrumental fluxes of both components and background) to minimize the sum of the square residuals, as described in McLeod et al. (1998) and Leh\'ar et al. (2000). We use windows of 64$\times$64 pixels. Each empirical PSF is a subframe of 64$\times$64 pixels around the PSF star (the "a" star in Fig. 1), while the lens system is analyzed from a subframe of the same size, but centered on the double quasar. The instrumental fluxes of the "b", "c", "s1" and "s2" stars are also inferred from 64$\times$64 pixels windows centered on them. We initially focus on the nearby field stars, and take the "b--c" stars as the control--reference objects. The "a" object is the brightest star in the "a--c" triangle, and "b" and "c" were spectroscopically identified by Kochanek et al. (1997) and Zickgraf et al. (2003): "b" is a FG star, whose spectrum includes the G--band and \ion{Ca}{ii} H--K lines, and "c" is a M3 star. The $R - I$ and $B - R$ colors of the brightest component (A) and the "a" star are similar, the colors of the faintest component (B) are close to the colors of the "b" star, and the "c" star has colors different to those of the components and the "a--b" stars (see Table 1 of Kochanek et al. 1997). On the other hand, after checking the PSFs of the three nearby field stars ("a--c"), we do not find significant differences between them. This suggests that the global shape of the PSFs around the lens system does not depend on the position and color of the point--like objects, so the PSF of the "a" star seems to be a reliable tracer of the PSF associated with any point--like object in the region of interest. As a first attempt for obtaining light curves we use the "b" and "c" stars as the control and reference objects, respectively. Unfortunately, we find clear evidences in favour of variability of the "c" star, since the three curves $m_A - m_c$, $m_B - m_c$ and $m_b - m_c$ have a similar global behaviour. This fact forces us to rule out the "c" star as a reference--control object and, thus, to take the "a" and "b" nearby field stars as the control and reference point--like sources, respectively. In the next subsection, we analyze the Maidanak--Wise fluxes $m_a - m_b$ and show that both stars ("a" and "b") are non--variable objects. This result permits to assure the good behaviour of "b". The Calar Alto fluxes $m_a - m_b$ are not included in the analysis, since most Calar Alto data disagree with the Maidanak--Wise common level of flux. We found an anomaly in the behaviour of the Calar Alto relative fluxes for widely separate stars (see below), so only the relative fluxes for neighbouring point--like objects are reliable photometric measurements. Fortunately, the comparison between the quasar components and the "b" nearby reference star seems to be a feasible approach. After applying the photometric method to the three individual frames for each filter and night (see Table 1), we obtain three different measurements of $y_A = m_A - m_b$ and $y_B = m_B - m_b$ in the $V$ and $R$ passbands for each night. To test the reliability of the instrumental fluxes of A and B, we analyse the residues in each residual frame. A residual frame is an image after subtracting the fitted background and point--like objects (PSF fitting method). More properly, we focus on the residual subframe occupied by the system, and then we estimate the residue--to--signal ratio ($R/S$) in each pixel of interest. A $R/S$ value less than 10\% is acceptable, so a subframe with at least 90\% of pixels having acceptable residues is considered to be related to reliable photometric solutions. Thus, we classify the individual fits in two categories: fits leading to $<$ 90\% of pixels having acceptable residues (bad fits, unreliable results) and good fits that are associated with reliable results ($\geq$ 90\% of pixels having acceptable residues). As a complementary test, we study the relation between the quality of the fits (in terms of post--fit residues) and two relevant parameters (image quality). The signal--to--noise at the brightest pixel of the lens system, $(S/N)_{max}$, and the seeing, $FWHM$ (in \arcsec), are the two parameters to compare with the fit quality. Some kind of correlation between good fits and good images is expected. In Figure 3 we draw the $(S/N)_{max}$--$FWHM$ plots for frames in the $R$ filter (top panel) and the $V$ filter (bottom panel). Circles and triangles represent good and bad fits, respectively. The plots in Fig. 3 indicate that the good fits correspond to images with high or moderate $(S/N)_{max}$ ($\geq$ 30). Moreover, at moderate $(S/N)_{max}$ ($\sim$ 30--50), most of the good fits seem to be associated with a relatively good seeing ($<$ 2 \arcsec). To obtain a robust photometry, we finally discard the frames corresponding to the triangles in Fig. 3. For each filter and night, if there are two or three good frames (good fits), then we get mean values of $y_A$ and $y_B$, and compute standard deviation of means as errors. We only consider relative fluxes with uncertainties $\leq$ 40 mmag. Now we plot $y_A$ (circles) and $y_B - 0.45$ mag (squares) in Figure 4 ($R$--band fluxes). If we concentrate our attention in the period with the best sampling (after day 2755), the A light curve shows a moderate decline and the B record shows a moderate rise. Indeed it seems that the "b" star is a good reference object (constant flux), since there is no zero--lag global correlation between $y_A$ and $y_B$. In Figure 5 we show the light curves $y_A$ and $y_B -$ 0.65 mag in the $V$ passband. In this case we have a total of 11 points for the A component (circles) and 10 points for the B component (squares). The $V$--band and $R$--band light curves of the A component are consistent with each other. A final moderate decline appears in both curves. The situation is more confused for the B component. The $R$--band final rise is not clearly reproduced in the $V$ band, and the $V$--band final measurements could have underestimated formal errors. We note the relative faintness of B in the $V$ band ($\Delta m \sim$ 0.8 mag), and thus, the possibility of systematic uncertainties when the PSF fitting method is applied at some epochs. The data in both optical filters are available at http://grupos.unican.es/glendama/. After presenting the records of the double quasar, we concentrate on the Calar Alto light curves of the field stars that were previously introduced by Kochanek et al. (1997) and Nakos et al. (2003), i.e., $y_a = m_a - m_b$, $y_c = m_c - m_b$, $y_{s1} = m_{s1} - m_b$ and $y_{s2} = m_{s2} - m_b$. There are no previous studies on the variability of the nearby field stars "a--c". On the other hand, the farther field stars ("s1--s2") were verified to be non--variable by using 76 Wise frames taken from 1999 December 24 to 2002 March 3 (Nakos et al. 2003). As Nakos et al. (2003) found that "s1" and "s2" seem to be useful reference stars, we check the behaviour of "s1--s2" in 2003. The PSF of the stars in the surroundings of the double quasar could slightly differ from the PSF of the "s1--s2" stars in a relatively far region. Therefore, we must be careful when obtaining the instrumental fluxes of the farther stars. To detect possible anomalies caused by a mismatch between the brightness profile of the "a" star and the PSF of "s1--s2", the light curves $y_{s1}$ and $y_{s2}$ are derived from both PSF fitting and aperture methods. The records $y_a$, $y_c$, $y_{s1}$ and $y_{s2}$ in the $R$ filter are depicted in Figure 6. To guide the eyes, we use some offsets and dashed horizontal lines and put all the relative records of each pair within a box. Filled and open symbols are associated with PSF fitting and aperture, respectively. The top box includes the $y_a$ + 2.15 mag fluxes (open squares). The second, third and fourth boxes (under the top one) correspond to the $y_c$ (filled squares), $y_{s1}$ + 2.38 mag (filled and open triangles) and $y_{s2}$ + 0.29 mag (filled and open circles) records, respectively. As most of the stars are brighter than the quasar components (A and B) and they are far from other objects, the typical formal errors in the stellar fluxes are clearly less than the typical uncertainties in the fluxes of the components (these are usually fainter and are placed in a crowded region). The stellar error bars in Fig. 6 are often smaller than the sizes of the associated symbols. When doing aperture photometry on six $R$--band Wise frames covering the first semester of 2003, we obtain a $y_a$ + 2.15 mag light curve (open triangles in the top box of Fig. 6) that disagrees with the Calar Alto trend in the overlap period (between days 2710 and 2760). In the next subsection, we show that the Wise and Maidanak brightnesses are constant and consistent with each other, so the Calar Alto values of $y_a$ are not true fluxes, but anomalous results. On the contrary, the Wise light curve $y_c$ (open circles in the second box of Fig. 6) agrees with the Calar Alto curve in the overlap period. From the Wise frames we confirm the flux level during the high--state of "c". Unfortunately, the small--amplitude variability of "c" (rms fluctuation of $\sim$ 8 mmag) cannot be confirmed from the Wise data. The rms fluctuation of the Wise fluxes ($\sim$ 9 mmag) is very similar to the Calar Alto variation, but the formal errors are relatively large ($\sim$ 10 mmag). Moreover, there are no Wise frames in 2003 May (around the day 2780) and, thus, we cannot check (via Wise data) the reliability of the Calar Alto dip in $y_c$ (80--100 mmag). However, the flux of the "c" star at day 2793 in the $V$ band confirms the existence of a transition from the low--state to the high--state, which is finished at days 2800--2810 (see the last open circle in the second box of Fig. 6). For the "s1--s2" stars, which are as far from star "b" as star "a" is, we again find a disagreement between the Calar Alto trends and the Wise records (open astroids and rhombuses in the third and fourth boxes of Fig. 6). Although aperture curves are closer to the Wise behaviours, we cannot fairly reproduce the Wise data. Some probes with the "x" star (using $y_x = m_x - m_b$) also indicate that the Calar Alto and Wise behaviours disagree. It seems that the differential photometry between widely separate stars may lead to meaningless results, and only the relative fluxes for neighbouring objects are reliable. To test this conclusion, apart from the successful results through the neighbouring stars "b" and "c", we also analyze the differential photometry between the pair "s1--s2" (see Fig. 1). The curves $m_{s2} - m_{s1} -$ 0.20 mag are depicted in the bottom box of Fig. 6: Calar Alto (filled and open star symbols) and Wise (open crosses). In the overlap period (from day 2710 to day 2760), there is a reasonable agreement between the results from both observatories, and the Calar Alto measurements seem to be quite reliable. From the Calar Alto frames, both photometric techniques are consistent with each other, but a constant flux cannot explain the observations. When we fit the data sets to a constant, our best solutions are characterized by $\chi^2 \sim$ 162 (PSF fitting) and $\chi^2 \sim$ 6 (aperture). It is a curious fact that aperture photometry on only one frame per night leads to relative fluxes in rough agreement with a constant level. However, more refined measurements (aperture or PSF fitting on several frames per night) reveal the variability of one ("s1" or "s2") or both stars. \subsection{Maidanak frames and global $R$--band light curves of SBS 0909+532} In the case of the $R$--band Maidanak observations, in order to derive the relative fluxes of the components of SBS 0909+532, we also use a direct PSF fitting. For a given frame, after to obtain a first estimate of the free parameters (initial solution), the fit is refined through an iterative procedure, which works as the CLEAN algorithm ({\O}stensen 1994). The iterative task is done with each individual image, and the solutions converge after a few cycles. For each night, we take all the available images and obtain the mean values of $y_A$ and $y_B$. From the standard deviation of the means, we also derive the errors in $y_A$ and $y_B$. In agreement with the criteria in subsection 3.1, only fluxes with errors less than or equal to 40 mmag are considered. Apart from the analysis of the lens system, using aperture photometry, we also measure $y_a$. The relative fluxes $y_a$ are depicted in Figure 7 (open star symbols). The Maidanak measurements in the first semester of 2003 and the six Wise data of $y_a$ (open triangles; see here above) are tightly distributed around $-$ 0.842 mag (solid line in Fig. 7). The rms fluctuation of the data is only of $\sim$ 6 mmag (see the dashed lines in Fig. 7), which is consistent with the typical error of the measurements. The $y_a$ results in Fig. 7 suggest that both "a" and "b" are non--variable objects. Through 2003 (first and second semesters) we do not find any evidence in favour of variability of the "a--b" stars. We show our global $R$--band light curves of SBS 0909+532 in Figure 8. The open circles (Maidanak) and filled circles (Calar Alto) are the measurements of $y_A$, whereas the open squares (Maidanak) and filled squares (Calar Alto) are the values of $y_b -$ 0.45 mag. We have 31 points for the A component (circles) and 26 points for the B one (squares). The top panel of Fig. 8 contains the results in the winter--spring of 2003 and the bottom panel of Fig. 8 includes the results in the autumn of 2003. For each component we test the existence of a bias between the Calar Alto and Maidanak fluxes, e.g., $\beta_A = y_A$ (Calar Alto) $- y_A$ (Maidanak). Very small biases of $\beta_A$ = + 15 mmag and $\beta_B$ = $-$ 30 mmag are found, and these corrections are taken into account to make the global records in Fig. 8. The biases are derived from the comparison between the Maidanak fluxes in a thirty day period (from day 2750 to day 2780) and the Calar Alto fluxes at equal or close dates (see the top panel of Fig. 8). To roughly estimate the contaminations from the direct PSF fitting technique, we take some of our best Maidanak images (in terms of seeing conditions, $FWHM \sim$ 1 $\arcsec$) in the $R$ band. A zoom--in of one of these best frames is shown in Fig. 2. Firstly, we combine the selected frames and derive a numerical model of the galaxy from a regularizing algorithm. To produce a more stable reconstruction, the real galaxy profile is assumed to be close to the Sersic profile (Koptelova et al. 2005). Our deconvolution method differs only slightly from the former deconvolution techniques by Magain, Courbin \& Sohy (1998) and Burud et al. (1998). Figure 9 presents the galaxy reconstruction obtained from the stack of the $R$--band selected frames. The box in Fig. 9 is 16\farcs6 on a side. The positions of the components are labeled with two crosses: A is on the left and B is on the right. The innermost contours are circular--elliptical rings, whereas the outermost contours show a less definite shape. Secondly, the selected frames are fitted to a photometric model that includes the galaxy brightness. Therefore, we are able to infer clean relative fluxes of A and B (without contamination by galaxy light) and to compare them with the contaminated ones (from direct PSF fitting). As result of the comparison, we report typical (averaged) contaminations of 18.8 mmag and 4 mmag for the A and B components, respectively. These very weak contaminations are in reasonable agreement with our preliminary considerations in the beginning of this section, and are taken into account in the measurement of the $R$--band flux ratio in Section 5. \section{Time delay} To calculate the time delay between both components of SBS 0909+532, we use the $R$--band brightness records corresponding to the winter--spring of the year 2003. The $R$--band records are more densely populated than the $V$-band ones. Moreover, the $R$--band time coverage in the winter--spring of 2003 (about 120 days) is longer than the time coverage in the autumn of 2003 (about 50 days). Thus we focus on the $R$--band data from day 2670 to day 2790, i.e., 22 points in the A component and 19 points in the B component (see the top panel of Figure 8). There are different number of points for component A and component B because we only consider fluxes with uncertainties below 40 mmag (see Section 3). As the B component is fainter, its photometric uncertainties are larger and the number of final data is smaller. The new light curves are characterized by a mean sampling rate of one point each six days. Once we have the data set, a suitable cross--correlation technique is required. Here we mainly use the $\chi^2$ minimization (e.g., Kundi\'c et al. 1997) and the minimum dispersion ($D^2$) method (Pelt et al. 1994, 1996). However, although other techniques are probably less robust than the $\chi^2$ and $D^2$ ones (doing a first delay measurement, without a previous empirical determination), we also tentatively explore the modified cross--correlation function (MCCF) technique (Beskin \& Oknyanskij 1995; Oknyanskij 1997). The MCCF combines properties of both standard cross--correlation functions: the CCF by Gaskell \& Spark (1986) and the DCF by Edelson \& Krolik (1988). We begin our analysis using the $\chi^2$ method, which is based on a comparison between the light curve $y_A$ (or $y_B$) and the time shifted light curve $y_B$ (or $y_A$). For a given lag, one can find the magnitude offset that minimizes the $\chi^2$ difference. From a set of lags, it can be derived a set of minima (of $\chi^2$), which permits to make a $\chi^2$ spectrum: $\chi^2$ vs lag. The best solution of the delay is the lag corresponding to the minimum of the $\chi^2$ spectrum. In general, the shifted epochs $t'_B$ (or $t'_A$) do not coincide with the unchanged epochs $t_A$ (or $t_B$), so we estimate the values of $y_A(t'_B)$ (or $y_B(t'_A)$) by averaging the A (or B) fluxes within bins centered on times $t'_B$ (or $t'_A$) with a semiwidth $\alpha$. To average in each bin, it is appropriate the use of weights depending on the separation between the central time $t'_B$ (or $t'_A$) and the dates $t_A$ (or $t_B$) in the bin. In principle, we concentrate in the interval [$-$ 90, + 90] days, which includes the predicted negative delays (see Introduction) as well as a wide range of unlikely positive delays (positive delays are inconsistent with basic observations of the system). Firstly, the curve $y_A$ and the time shifted curve $y_B$ are compared with each other (using bins in the A component). In order to work with a reasonable time--resolution, we use $\alpha$ values less than or equal to two times the mean sampling time, i.e., $\alpha \leq$ 12 days. The $\chi^2$ value roughly grows with the size of the bin, and $\chi^2 \sim$ 1 for $\alpha$ = 7$-$9 days. For $\alpha$ = 7$-$9 days, there are best solutions $\Delta \tau_{BA}$ = $+$ 46$-$48 days ($\chi^2$ = 0.97$-$0.98), and we show the corresponding spectra in Figure 10. We have drawn together the spectra for $\alpha$ = 7 days (dashed line), $\alpha$ = 8 days (solid line) and $\alpha$ = 9 days (dotted line). Apart from the main minima close to + 50 days, there are other secondary minima at negative and positive lags. In Fig. 10, two secondary minima seem to stay significant for all the bin sizes: the minima close to $-$ 50 days and the probable edge effects at + 80$-$90 days. We also compare the curve $y_B$ and the time shifted curve $y_A$, using bins in the B component. For $\alpha$ = 10 days, we obtain a best solution $\Delta \tau_{BA}$ = $-$ 44 days ($\chi^2$ = 1.15). Smaller and larger bins lead to solutions characterized by $\chi^2 <$ 0.7 and $\chi^2 \geq$ 1.2, respectively. In Figure 11, the solid line represents the spectrum for $\alpha$ = 10 days, while the dashed line represents the spectrum for $\alpha$ = 9 days and the dotted line traces the spectrum for $\alpha$ = 11 days. Main minima in the interval $-$ 40$-$50 days appear in all these cases. Unfortunately, important signals at positive lags and probable border effects at + 80$-$90 days are again included in the complex spectra. The important structures at positive lags in Figs. 10$-$11 are probably caused by artifacts in the cross--correlation, so they have no physical origin, but are due to the 10/20--day gaps and the moderate variability of the components. Therefore, taking $\alpha$ = 10 days (bins in the B component) and a negative range [$-$ 90, 0] days, we try to determine a pre--conditioned time delay. In order to derive uncertainties, we follow a simple approach. We make one repetition of the experiment by adding a random quantity to each original flux in the light curves. The random quantities are realizations of normal distributions around zero, with standard deviations equal to the errors of the fluxes. We can make a large number of repetitions, and thus, obtain a large number of $\Delta \tau_{BA}$ values. The true value will be included in the whole distribution of measured delays. From the $\chi^2$ minimization (bins in B and $\alpha$ = 10 days) and 1000 repetitions, we obtain the histograms in Figure 12. Regarding the distributions in the top panel (delays) and bottom panel (flux ratios) of Fig. 12, the main features lead to measurements $\Delta \tau_{BA}$ = $-$ 45 $^{+ 1}_{-11}$ days and $\Delta m_{BA}$ = 0.590 $\pm$ 0.014 mag (95\% confidence intervals). We note that the main delay peak is asymmetric, so 55\% of the repetitions correspond to $-$ 44$-$45 days, whereas 40\% of the repetitions correspond to values $<$ $-$ 45 days. The secondary delay peak (around $-$ 20 days) represents about 5\% of the repetitions and is associated with the secondary minima in the negative region of Fig. 11. Therefore, the distribution in the top panel of Fig. 12 permits a 95\% estimation of the time delay of SBS 0909+532. In Figure 13 (top panel), the A light curve (circles) shifted by the optimal values of the time delay and the magnitude offset (time--delay--corrected flux ratio), and the unchanged B light curve (squares) are plotted. The cross--correlation using bins in the B component ($\alpha$ = 10 days) indicates that the initial variations in the brightness of B reasonably agree with the final fluctuations in the brightness of A. The overlap for a delay of $-$ 80 days (e.g., Saha et al. 2005) also appears in the bottom panel of Fig. 13. However, this last time delay is clearly rejected by the observations, since the $\chi^2$ value is larger than 10 ($\chi^2 \sim$ 18). To confirm the results from the $\chi^2$ minimization, we also use the dispersion spectra introduced by Pelt et al. (1994, 1996). The basic idea is a combination of $y_A$ and $y_B$ into one global record for every lag $\tau$ and magnitude offset $m_0$ by taking all the values of $y_A$ as they are and shifting the values of $y_B - m_0$ by $\tau$. For each $\tau$ one can find the $m_0$ value that minimizes a dispersion estimate $D^2(\tau,m_0)$, so a dispersion spectrum $D^2(\tau)$ can be made in a direct way. We focus on the $D^2_{4,2}$ spectra that are called $D^2$ for simplicity (see Pelt et al. 1996 for details). This technique incorporates a decorrelation length ($\delta$), where $\delta$ plays a role similar to that of $\alpha$ in the $\chi^2$ method. Considering reasonable values of $\delta$ (from 7 to 11 days, see here above), we are able to make some interesting spectra. In Figure 14 we have plotted together the spectra for $\delta$ = 7 days (dashed line), $\delta$ = 9 days (solid line) and $\delta$ = 11 days (dotted line). Although there are main minima in the interval $-$ 40$-$50 days, there are also significant signals at positive lags and probable border effects at + 90 days. In the negative region of Fig. 14, a secondary minimum around $-$ 70 days appears. Using $\delta$ = 9 days and a negative range [$-$ 90, 0] days, we carry out a second pre--conditioned measurement of the time delay. The uncertainties are deduced from 1000 repetitions of the experiment (see here above), and the relevant histograms are shown in Figure 15. While the top panel contains the distribution of delays, the bottom panel traces the distribution of flux ratios. Through the distributions in Fig. 15, we obtain that $\Delta \tau_{BA}$ = $-$ 48 $^{+ 7}_{-6}$ days and $\Delta m_{BA}$ = 0.585 $\pm$ 0.020 mag (90\% confidence interval). These $D^2$ results strengthen the conclusions from the $\chi^2$ technique. A marginal measurement (10\% confidence interval) of $\Delta \tau_{BA}$ = $-$ 67 $^{+ 1}_{-2}$ days and $\Delta m_{BA}$ = 0.558 $^{+ 0.007}_{-0.008}$ mag is also possible. However, both this possibility and the $\chi^2$ result of around $-$ 20 days are probably related to the presence of gaps and the absence of strong variability in the light curves. A MCCF technique (Beskin \& Oknyanskij 1995; Oknyanskij 1997) is also explored. The MCCF is a modification of the standard cross--correlation functions (CCF and DCF). When this MCFF is applied to our data in the lag interval [$-$ 60, + 60] days, the maximum correlation coefficient (0.907) corresponds to a lag of $-$ 45 days. This last result basically agrees with the $\chi^2$ and dispersion spectra in Figs. 11 and 14. \section{Conclusions} Nowadays several groups are trying to coordinate the rich but scattered research potential in the field of gravitationally lensed quasar monitoring. The goals are to rationalize the astronomical work and to catalyze big scientific collaborations so that the astrophysics community can get a significant progress in the understanding of the central engine in lensed quasars, the structure of the lensing galaxies and the physical properties of the Universe as a whole. Some examples about that are the Astrophysics Network for Galaxy LEnsing Studies (ANGLES, http://www.angles.eu.org/), the Cosmic Lens All-Sky Survey (CLASS, http://www.aoc.nrao.edu/$\sim$smyers/class.html) and the COSmological MOnitoring of GRAvItational Lenses (COSMOGRAIL, http://www.cosmograil.org/). The University of Cantabria group (Spain), three groups of the former Soviet Union (Institute of Astronomy of Kharkov National University, Ukraine, Sternberg Astronomical Institute, Russia, and Ulug Beg Astronomical Institute of Uzbek Academy of Science, Uzbekistan) and the Tel--Aviv University group (Israel) are also carrying out a series of initiatives to better exploit the recent individual monitoring campaigns as well as to solidify some future common project. In this paper we present the first collaborative programme on the variability of the double quasar SBS 0909+532A,B. The $VR$ observations of the system and the field stars were made with three modern ground--based telescopes in the year 2003. The SBS 0909+532c star (N23210036195 in the GSC2.2 Catalogue) at ($\alpha$, $\delta$) = (09:12:53.59, +52:59:39.82) in J2000 coordinates is found to be variable, with two different levels of flux. The $VR$ gap between the low--state and the high--state is of 80--100 mmag, and the low--state lasts about one month. In the high--state the star also seems to vary, but these small--amplitude variations are not so significant as the gap between states. We want to remark the variability of this nearby star ("c" star), and to encourage colleagues to follow-up its fluctuations and identify the kind of variable source. The "c" star cannot be used as the reference object (differential photometry), because it introduces a zero--lag global correlation between the light curves of the quasar components A and B. However, the "a--b" nearby stars are non--variable sources, and we choose the "b" star as the reference candle. On the other hand, the "s1" and "s2" stars are relatively far objects, which were proposed as good references in a previous analysis (Nakos et al. 2003). However, the new $R$--band light curve $m_{s2} - m_{s1}$ reveals the variability of one ("s1" or "s2") or both stars. This variability could be either a very rare phenomenon or a consequence of doing more refined measurements (aperture or PSF fitting on several frames per night). We warn about the possible problems with this pair of stars and think it merits more attention. The point--spread function (PSF) fitting methods permit to resolve the two components of the quasar and to derive the $VR$ light curves of each component. These new $VR$ light curves represent the first resolved brightness records of SBS 0909+532. Although the $V$--band curves are interesting, the $R$--band records seem more reliable and are more densely populated. The $R$--band curves show a moderate variability through 2003, and the observed fluctuations are promising for different kinds of future studies. To estimate the time delay between the components of SBS 0909+532, we use an 120--day piece of the $R$–-band brightness records, and $\chi^2$ and dispersion ($D^2$) techniques. The cross--correlation of the two light curves (A and B) leads to complex $\chi^2$ spectra. However, assuming that the quasar emission is observed first in B and afterwards in A, or in other words, $\Delta \tau_{BA} <$ 0 (in agreement with basic observations of the system), 95\% measurements $\Delta \tau_{BA}$ = $-$ 45 $^{+ 1}_{-11}$ days and $\Delta m_{BA}$ = 0.590 $\pm$ 0.014 mag are inferred from 1000 repetitions of the experiment (synthetic light curves based on the observed records). From the $D^2$ minimization (Pelt et al. 1996) and 1000 repetitions, we also obtain 90\% measurements $\Delta \tau_{BA}$ = $-$ 48 $^{+ 7}_{-6}$ days and $\Delta m_{BA}$ = 0.585 $\pm$ 0.020 mag. The $D^2$ uncertainties are derived under the already mentioned assumption that $\Delta \tau_{BA}$ is negative. There is a clear agreement between the results from both techniques, so a delay value of about one and a half months is strongly favoured. Our light curves rule out a delay close to three months, which has been claimed in a recent analysis (Saha et al. 2005). When we measure the time delay of the system, we simultaneously derive the time--delay--corrected flux ratio (at the same emission time) in the $R$ band. This quantity, $\Delta m_{BA}$ = $m_B(t + \Delta \tau_{BA}) - m_A(t)$, is contaminated by light of the lens galaxy, and taking into account the weak contaminations of A and B (see the end of subsection 3.2), the totally corrected $R$--band flux ratio is 0.575 $\pm$ 0.014 mag. We remark that our final $R$ flux ratio is in total agreement with the rough (uncorrected by the time delay and the contamination by galaxy light) measurement by Kochanek et al. (1997): 0.58 $\pm$ 0.01 mag. To properly determine a flux ratio, one must use clean fluxes at the same emission time, i.e., fluxes at different observation times and without contamination (Goicoechea, Gil--Merino \& Ull\'an 2005). Only for particular cases (e.g., faint lens galaxy, short delay and moderate variability), it may be reasonable to use direct fluxes. In order to get a reasonably good value of $\chi^2$, we do not need to introduce a time dependent magnitude offset or a complex iterative procedure (e.g., Burud et al. 2000; Hjorth et al. 2002), i.e., only a delay and a constant offset are fitted. This is a strong point of the analysis. The agreement between the results from different techniques is another strong point. However, the new measurements have some weak points that we want to comment here. The weakest point is the relatively poor overlap between the A and B records, when the A light curve is shifted by the best solutions of the time delay and the magnitude offset (e.g., see the top panel of Fig. 13). Moreover, we carry out pre--conditioned measurements, since a negative interval [$-$ 90, 0] days is considered in the estimation of uncertainties (component B leading component A). This second weak point is related to the presence of 10/20--day gaps and the moderate variability of the components, which does not permit to fairly rule out positive delays. We nevertheless remark that the negative interval is in good agreement with the predictions by Leh\'ar et al. (2000) and Saha et al. (2005), and we find $\chi^2$ and $D^2$ minima around $-$ 45 days when the observed data and both negative and positive lags are taken into account (see Figs. 11 and 14). Of course, as any another first determination of a time delay, the 1.5--month value should be confirmed from future studies. Forty years ago, Refsdal (1964) suggested the possibility of determining the current expansion rate of the Universe (Hubble constant) and the masses of the galaxies from the time delays associated with extragalactic gravitational mirages. More recently, for a singular isothermal ellipsoid (SIE), Koopmans, de Bruyn \& Jackson (1998) found that the time delay can be cast in a very simple form, depending on basic cosmological parameters, redshifts and image positions. The relevant image positions are the positions with respect to the centre of the main lens galaxy, and the SIE delay is similar to the delay for a singular isothermal sphere (SIS). In principle, a singular density distribution is justified because a small core radius changes the time delay negligibly, and only a small core radius seems to be consistent with the absence of a faint central image (e.g., Kochanek 1996). Moreover, individual lenses and lens statistics are usually consistent with isothermal models (e.g., Witt, Mao \& Keeton 2000 and references therein), so it is common to adopt an isothermal profile. Witt, Mao \& Keeton (2000) showed that an external shear changes the simple SIS time delay in proportion to the shear strength. For two--image lenses that have a small shear and images at different distances from the centre of the lens, the shear should have a small effect on the time delay. Thus, when one has accurate measurements of image positions, redshifts and time delay, it is viable an accurate estimation of $H_0$ (using complementary information on the matter/energy content of the Universe). Very recently, Kochanek (2002) also presented a new elegant approach to the subject. He modelled the surface density locally as a circular power law, with a mean surface density $<\kappa>$ in the annulus between the images. Expanding the time delay as a series in the ratio of the thickness of the annulus to its average radius, it is derived a delay that is proportional to the SIS time delay. The zero--order expansion term consists of the SIS delay and a multiplicative factor $2(1 - <\kappa>)$. Kochanek also incorporated the quadrupoles of an internal shear (ellipsoid) and an external shear. However, for two-image lenses where the images lie on opposite sides of the lens, the delay depends little on the quadrupoles. This novel perspective is useful to infer $<\kappa>$ from observations of the lens system (time delay, image positions and redshifts) and complementary cosmological data (expansion and matter/energy content of the Universe). For SBS 0909+532, although the redshifts are very accurately known and the time delay is now tightly constrained (or at least there is a first accurate estimation to be independently confirmed), the inaccurate position of the main lens galaxy does not permit an accurate measurement the cosmic expansion rate and the surface density of the main deflector. We have $H_0 \propto \theta_B^2 - \theta_A^2$ and $1 - <\kappa> \propto (\theta_B^2 - \theta_A^2)^{-1}$, where $\theta_A$ and $\theta_B$ are the image angular positions with respect to the centre of the main lens galaxy. On the other hand, using the astrometry in Table 3 of Leh\'ar et al. (2000), it is easy to obtain $\theta_B^2 - \theta_A^2$ = 0.4 $\pm$ 0.2. Thus we conclude that the accuracy in $\theta_B^2 - \theta_A^2$ is only 50\%, indicating the necessity of new accurate astrometry of SBS 0909+532. \begin{acknowledgements} The UC members are indebted to J. Alcolea (Observatorio Astron\'omico Nacional, Spain) for generously granting permission to operate the 1.52 m Spanish telescope at Calar Alto Observatory (EOCA) in March--June 2003. This 4--month season was supported by Universidad de Cantabria funds and the Spanish Department for Science and Technology grant AYA2001-1647-C02. AU thanks the Departamento de F\'isica Te\'orica y del Cosmos de la Universidad de Granada (E. Battaner) for hospitality during the observational season. The post--observational work and the spreading of results are supported by the Department of Education and Science grants AYA2002-11324-E, AYA2004-20437-E and AYA2004-08243-C03-02. We acknowledge the use of data obtained by the SAI group headed by B. Artamonov. We are also indebted to D. Maoz and E. Ofek for providing us with the Wise frames of SBS 0909+532. APZ is grateful for the support of the Science and Technology Center of Ukraine (STCU), grant U127k. The observational work by the UBAI group (TA and OB) at Mt. Maidanak was supported by the German Research Foundation (DFG), grant 436 UZB 113/5/0-1. We are also grateful to the referee for several helpful comments. We acknowledge support by the European Community's Sixth Framework Marie Curie Research Training Network Programme, Contract No.MRTN-CT-2004-505183 "ANGLES". The GSC-II is a joint project of the Space Telescope Science Institute (STScI) and the Osservatorio Astronomico di Torino (OAT). STScI is operated by the Association of Universities for Research in Astronomy, for the NASA under contract NAS5-26555. The participation of the OAT is supported by the Italian Council for Research in Astronomy. Additional support is provided by ESO, Space Telescope European Coordinating Facility, the International GEMINI project and the ESA. \end{acknowledgements} {}
Title: The nearest young moving groups
Abstract: The latest results in the research of forming planetary systems have led several authors to compile a sample of candidates for searching for planets in the vicinity of the sun. Young stellar associations are indeed excellent laboratories for this study, but some of them are not close enough to allow the detection of planets through adaptive optics techniques. However, the existence of very close young moving groups can solve this problem. Here we have compiled the members of the nearest young moving groups, as well as a list of new candidates from our catalogue of late-type stars possible members of young stellar kinematic groups, studying their membership through spectroscopic and photometric criteria.
https://export.arxiv.org/pdf/astro-ph/0601573
command. \shorttitle{Near moving groups} \shortauthors{L\'opez-Santiago et al.} \begin{document} \title{The nearest young moving groups} \author{J. L\'opez-Santiago\altaffilmark{1,2}, D. Montes\altaffilmark{2}, I. Crespo-Chac\'on\altaffilmark{2} and M.J. Fern\'andez-Figueroa\altaffilmark{2}} \affil{\altaffilmark{1}INAF - Osservatorio Astronomico di Palermo, Piazza Parlamento 1, I-90134 Palermo, Italy} \affil{\altaffilmark{2}Departamento de Astrof\'{\i}sica y Ciencias de la Atm\'osfera, Facultad de Ciencias F\'{\i}sicas, Universidad Complutense de Madrid, E-28040 Madrid, Spain} \keywords{associations and clusters: moving groups --- stars: kinematics --- stars: stellar activity --- stars: lithium abundance --- stars: planets} \section{Introduction} \label{sec:intr} In recent years, a series of young stellar kinematic groups (clusters, associations, and moving groups) of late-type stars with similar space motion and ages ranging from 8 to 50 Myr \citep[see][and references therein]{zuc04a} has been discovered in our neighbourhood: TW~Hya, $\beta$~Pic, AB~Dor, $\eta$~Cha, $\epsilon$~Cha, Tucana and Horologium associations. In addition, several more distant young associations such as MBM~12 \citep{hea00}, Corona Australis \citep{qua01}, and possibly the group of stars with a motion similar to that of HD~141569 \citep{wei00} have also been identified. In the Galactic velocity space, they situate inside the boundaries of the \object{Local Association} (see Fig.~\ref{fig:uv}), a mixture of young stellar complexes ---~OB and T-associations~--- and clusters with different ages \citep{egg75, egg83a, egg83b, m01}. These associations of very young stars are excellent laboratories for investigations of forming planetary systems \citep{zuc04}. Nevertheless, they are generally situated at distances above 50 pc, which makes them less accessible to adaptive optics systems even on large telescopes. It is well-known that tightly bound, long-lived open clusters can account for only a few per cent of the total galactic star formation rate \citep{wie71}. Therefore, either most clusters and associations disperse very quickly after star formation has started or most are born in isolation \citep{wic03}. The existence of very young moving groups (MGs) with a few dozens of stars showing the same spectroscopic properties ---~i.e. age, metallicity, level of magnetic activity~--- is in agreement with the first explanation. Small associations of stars may be dispersed by galactic differential rotation since they are not gravitationally bounded enough, taking into account that their nucleus consist of only a few stars as in the case of the Ursa Major MG (see King et al. 2003, for a recent review) or the recently discovered AB~Doradus MG \citep{zuc04}. The location of these young MGs inside the Local Association and its proximity in the $UV$-plane can be explained as the result of the juxtaposition of several star forming bursts in adjacent cells of the velocity field (see Montes et al. 2001, and references therein) or dynamical perturbations caused by spiral waves \citep{sim04,fam05,qui05}. Thus, one expects to find groups of coeval stars with similar space motion in our neighbourhood. In 2001, the $\beta$~Pic~MG \citep{zuc01} --- a group of stars with an age of $\sim$~12~Myr \citep{zuc01, ort04} at a mean distance of $\sim$~35~pc co-moving with the well-known young star $\beta$~Pic --- was confirmed to be the closest kinematic group up to date. More recently, \citet{zuc04} have identified a new group of stars co-moving with the also well-known young star AB~Dor, at a mean distance of $\sim$~30~pc, and with an age of $\sim$~50~Myr. Nevertheless, the existence of a nearer association of a few stars was proposed by \citet{gai98} and studied in detail by \citet{fur04}, though its existence is quite controversial. Here we discuss about the fact of the nearest MGs using both spectroscopic and photometric criteria of membership for a sample of stars that includes the proposed members from the literature and our list of young cool stars possible members of young stellar kinematic groups \citep{m01, lop05}. \section{The Hercules-Lyra Association} \label{sec:herc} Based on the kinematics of young solar analogues in the solar neighbourhood, \citet{gai98} confirmed the existence of a group of four stars (marked with $\dag$ in Table~\ref{tab:MGs}) co-moving in the space towards the constellation of Hercules. Recently, \citet{fur04} has extended the sample of late-type stars of this MG up to 15 nearby (d~$<$~25~pc) candidates, proposing the name Hercules-Lyra since several members show a radiant ``{\it evenly matched}'' with this constellation. Comparing the level of chromospheric activity of the stars of his sample with that of the members of the Ursa Major Association and looking for the existence of lithium in their spectrum, he notices that several candidates of Hercules-Lyra appear to be coeval of the Ursa Major stars, for which he gives an age of $\sim$~200~Myr. On the contrary, other candidates seem to be older (e.g. HD~111395) or younger (HD~17925, HD~82443, and HD~113449), questioning the existence of Hercules-Lyra as an entity independent of the Local Association. However, he considers unlikely that the majority of his sample can originate from the Pleiades alone, or other clusters of the Local Association since ``{\it they are poorer and more distant}'' as pointed out by \citet{jef95}. Thus, he confirms ``{\it the bulk}'' of the sample --- formed by the stars HD~166, HD~96064, HD~97334, HD~116956, HD~139777, HD~139813 and HD~141272 (see his Table~1) --- to be an entity on its own. Here we discuss the possible existence of the Hercules-Lyra MG as an independent association using kinematic (space motion), spectroscopic (lithium abundance) and photometric (isochrone fitting) criteria. A total of 12 possible members (stars marked with {\it a} in Table~\ref{tab:MGs}) have been added to the initial sample of \citet{fur04} from our catalogue of {\it Late-type Stars Possible Members of Young Stellar Kinematic Groups} \citep{m01}. The candidates have been chosen by their kinematics assuming a total dispersion of \mbox{$\pm$ 6 km~s$^{-1}$} in $U$ and $V$, respectively; that is, an average position of \mbox{($U$, $V$) = (-15.4, -23.4)} km~s$^{-1}$ has been determined using the stars given by \citet{fur04}, and every star in our catalogue in a radius of \mbox{$\pm$ 6 km~s$^{-1}$} has been selected. The value of the dispersion has been chosen equal to that of the \mbox{$\sim 200$ Myrs} old Castor~MG \citep{m01}, coeval of the Hercules-Lyra Association. No restriction in the $W$ component has been imposed in this first selection. In Table~\ref{tab:MGs} we summarize the results obtained by us. From the whole sample of 27 candidates, eight stars have been discarded as members by their space motion: HD~25457, located inside the B4 subgroup (see Fig.~\ref{fig:uv}); HD~96064, HD~112733, HIP~67092, the binary system made up of the F-type star HD~139777 and HD~139813, and HD~207129 all them with a value in $W$ higher than that of the rest of the candidates (see Fig.~\ref{fig:uvw_MGs}); and HD~113449, classified as member of the AB~Dor~MG by \citet{zuc04} (see Table~\ref{tab:MGs} and $\S$~\ref{sec:abdor} for a more detailed discussion) and questioned by \citet{fur04} because of its relatively high lithium abundance. We have also studied the lithium abundance --- measured as the equivalent width of the lithium line $\lambda$6707.8~\AA, $EW$({Li~{\sc i}}) --- in each one of the candidates. The values of $EW$(Li~{\sc i}) have been taken from \citet{lop05} and compared with those of the members of well-known stellar clusters (see Fig.~\ref{fig:li_MGs}). The results appear to be consistent with an age of \mbox{150 -- 300~Myr} for seven candidates. However, several stars (HD~1466, HD~17295, \mbox{1E~0318.5-19.4}, and HD~82443) show an $EW$(Li~{\sc i}) comparable to that of the members of the Pleiades while other five (HD~37394, HD~97334B, HD~111395, HD~116956 and HD~141272) are fully depleted or have a value lower than the expected for a member of the Hercules-Lyra Association. For isochrone fitting, we have adopted pre-main sequence models from \citet{sie00}. For $T_{\rm eff} < 4000$~K, the models systematically underestimate the age when comparing with clusters of known age such as the Pleiades and IC~2391 in a $M_{\rm v}$ vs. $V-I$ diagram \citep{lop05} due to the transformation from flux to colour. Bearing this in mind, the corrected transformation adopted by \citet{lop03} and \citet{lop05b} for stars cooler than 4000~K has been used in this work. The values of $V-I$ have been taken from the Hypparcos Catalogue \citep{esa97}. The result of comparing the position of the stars with the isochrones in the colour-magnitude diagram (CMD) (Fig.~\ref{fig:VI}) is again in agreement with and age of \mbox{$\sim$ 150 - 300 Myrs}. Nevertheless, no conclusions can be inferred from the CMD alone since isochrones of more than 80~Myr converge for $V-I \le 1.8$ mag., and ages larger than 300 Myr could be adopted. From the combination of the three criteria, the total sample of candidates is reduced to 10 stars with $EW$(Li~{\sc i}) and position in the CMD compatibles with an age of \mbox{$\sim$ 200 Myrs}, which could form the bulk of the Hercules-Lyra Association, and other 15 definitively non members or with a doubtful classification (Table~\ref{tab:MGs}). The members show a deviation \mbox{($\sigma_{\rm U}$, $\sigma_{\rm V}$) = (2.46, 1.61) km s$^{-1}$} from the centre (\mbox{($U$, $V$) = (13.19, 20.64) km s$^{-1}$}) lower than that of other coeval MGs such as Castor and Ursa Major \citep[see][and references therein]{m01}. A similar dispersion (\mbox{$\sigma_{\rm W} \approx 3.4$ km s$^{-1}$}) is found in W, confirming the results in $U$ and $V$. In the same way, the shape of the MG in the velocity field is in agreement with the theory of the MGs \citep{egg65, age79, sku99, asi99b, lop05}. According to this theory, not-gravitational bounded stars formed in the same forming region and with low sigmas in $U$, $V$ and $W$ are dispersed during their rotation around the Galactic centre, inducing a particular shape in both the space and the velocity field since some of the stars fall behind while others go ahead. The Galactic potential maintains the group bounded during several hundreds of years, in spite of the initial velocity dispersion in the molecular cloud, in both the $UV$-plane and the $W$ component. \section{The AB Dor MG and subgroup B4} \label{sec:abdor} Very recently, \citet{zuc04} have identified a large group of stars with the same space motion than the well-known young K-dwarf AB~Dor (d = 15~pc), a quadruple system \citep{clo05, gui05} made up of three late-type stars --- AB~Dor~A (HD~36705), AB~Dor~Ba and AB~Dor~Bb --- and a very low mass companion which has recently been object of discussion because of the discrepancy between its dynamical mass and that predicted by evolutionary models \citep{clo05}. All the stars listed in Table~1 of \citet{zuc04} are situated inside the Local Association (see Fig.~\ref{fig:uv}) near the boundaries of the young disk stellar population \citep{egg84}, and have at least one indicator of youth. Taking the intensity of the H$\alpha$ emission line of these stars and the position in a $V-K_{\rm s}$ diagram of three M-type members of the MG into account, they estimate an age of $50\pm10$ Myr for the AB~Dor MG. Very recently, \citet{luh05} and \citet{luh05b} have showed that the components of AB~Dor should have an age of 75 -- 150 Myr based on the comparison of both their position in the $M_{\rm K}$ vs. $V-K_{\rm s}$ diagram with respect to the Pleiades and IC~2391 clusters, and the $EW$(Li~{\sc i}) of AB~Dor~A with that of rapidly rotating K dwarfs in the Pleiades. Moreover, with an age of $\sim$~100 Myr the discrepancy between observations and models for the very low-mass companion (AB~Dor~C) would disappear \citep[eg.][]{clo05}. Taking this into account, they propose an age range of 75 -- 150 Myr for all the MG. To the initial sample of \citet{zuc04}, we have added 13 stars (marked with {\it b} in Table~\ref{tab:MGs}) from our catalogue of {\it Late-type Stars Possible Members of Young Stellar Kinematic Groups} \citep{m01}. These stars have been included to both searches: a) for other members of the group and b) to show the existence of two subgroups of different ages more clearly. They have been chosen because of their kinematics, assuming a total dispersion of \mbox{$\pm 4$ km s$^{-1}$} in $U$ and $V$ respectively, around the centre of the AB~Dor~MG defined by \citet{zuc04}. We have imposed no restriction to the $W$ component for this first selection. The whole sample contains a total of 50 stars. We have compared the $EW$({Li~{\sc i}}) of every star in the sample with that of known members of young open clusters (Fig.~\ref{fig:li_MGs}), as well as their position in the $V-I$ diagram with the isochrones of \citet{sie00} (Fig.~\ref{fig:VI}). The results reveal the existence of two subgroups with stars showing different spectroscopic and photometric features, mixed in the velocity field (see Fig.~\ref{fig:uvw_MGs}). The members of the first subgroup, that includes AB~Dor and PW~And --- a very active young K2-dwarf \citep{lop03} --- show $EW$({Li~{\sc i}}) similar to that of the high-rotators in the Pleiades (upper continuous line in Fig.~\ref{fig:li_MGs}) which are above the values found in the low-rotators of IC~2602 (lower dashed line in Fig.~\ref{fig:li_MGs}). Their position between the 30 and the 80 Myr isochrones in the $V-I$ diagram, together with the first result, is compatible with an age of 30 - 50~Myr. Moreover, the stars from the sample of \citet{zuc04} belonging to this subgroup are situated above the sequence of the Pleiades in the $M_{\rm K}$ vs. $V-K_{\rm s}$ diagram in \citet{luh05}. Here we have obtained a dispersion $\sigma \approx 2$ km~s$^{-1}$ in the W component, quite similar to the one observed in other young stellar associations such as Tucana or $\epsilon$~Cha \citep[see][and references therein]{zuc04a}. For determining the dispersion we have rejected the stars BD+07~1919A and B (marked with * in Table~\ref{tab:MGs}) since their radial velocities --- used for calculating the Galactic velocity components --- have not been corrected for binarity since no orbital solution has been found in the literature. Nevertheless, although the membership of this system is not completely reliable taking the value of their $W$ component into account, it has been included in the sample as possible member because of the position of the A component in the CMD, which suggests an age of $\sim$~30~Myr. The stars in the second subgroup show features, in terms of $EW$({Li~{\sc i}}) values and position in the CMD, comparable with that of the members of the Pleiades cluster: in Fig.~\ref{fig:li_MGs} they are situated slightly above the lower envelope of the Pleiades, while in Fig.~\ref{fig:VI} they situate on the (ZAMS) 80~Myrs isochrone. Its members could be considered as part of subgroup B4, one of the four subgroups found by \citet{asi99} inside the Local Association in their study of the space motion of OB~associations using Hypparcos astrometric data. The authors find a mean age of $\sim$~150~Myr for this subgroup using information from the photometry. The higher dispersion found for the stars of this second subgroup in the velocity space (Fig.~\ref{fig:uvw_MGs}) is in agreement with the age estimated by us. On the other hand, the results about AB~Dor MG indicate that this quadruple system has indeed $\sim$~50~Myr. The value of $EW$(Li~{{\sc i}}) for AB~Dor~A is somewhat above the upper envelope of the Pleiades but not so high as the one of IC~2602 (Fig.~\ref{fig:li_MGs}). On the other hand, its ($V-I$) colour situates it between the 30 and 80 Myr isochrones (Fig.~\ref{fig:VI}). The same result is clearly visible in Fig.~1 of \citet{luh05}, where AB~Dor is situated above the lower sequence of the Pleiades in the $M_{\rm K}$ vs. $V-K_{\rm s}$ diagram. With an age of 50~Myr, the discrepancy between observations and models for the AB Dor very low-mass companion (AB Dor C) shown in \citet{clo05} continuous, although it can be solved if the very low-mass companion were indeed an unresolved binary system \citep{mar05}. \section{Discussion and conclusions} \label{sec:conc} In Table~\ref{tab:MGs} we list the stars belonging to the nearest moving groups: Hercules-Lyra Association and AB~Dor MG, and those being part of the Local Association B4 subgroup. For the Hercules-Lyra Association, a division between certain members and candidates with doubtful classification or non members has been made. In the three groups, new candidates from our catalogue of {\it Late-type Stars Possible Members of Young Stellar Kinematic Groups} \citep{m01} have been selected because of their kinematics (see $\S$~\ref{sec:herc} and $\S$~\ref{sec:abdor}). A total of 75 stars including the known members and the new candidates selected by us have been analysed. Kinematic, spectroscopic and photometric criteria have been utilized to discriminate non members from the rest of candidates of the Hercules-Lyra Association, and to distinguish between the members of the AB~Dor MG and those of the B4 subgroup. In the velocity space, Hercules-Lyra is clearly distinguishable from the rest of the sample (see Figs.~\ref{fig:uv} and~\ref{fig:uvw_MGs}). The dispersion in $U$, $V$, and $W$ is comparable with that of other coeval MGs such as Castor and Ursa Major \citep[e.g.][]{m01}, and compatible with its age (see $\S$~\ref{sec:herc}). On the other hand, AB~Dor~MG and B4 subgroup are mixed up and age-dating criteria are necessary to distinguish between the members of both groups. Nevertheless, the dispersion in $W$ for AB~Dor MG is quite smaller than the one of B4 subgroup. Age-dating criteria are also necessary to discriminate non members of Hercules-Lyra from the certain ones. The results of applying them are summarized in Table~\ref{tab:MGs}: the Hercules-Lyra Association is formed by 10 certain members situated at a mean distance of $\sim$~20~pc and show values of $EW$(Li~{{\sc i}}) (Fig.~\ref{fig:li_MGs}) and a position in the $V-I$ CMD (Fig.~\ref{fig:VI}) compatible with an age of 150~--~300~Myr; the members of AB~Dor MG are situated at a mean distance of $\sim$~30~pc and show lithium abundances typical of stars with 30~--~50~Myr (Fig.~\ref{fig:li_MGs}), which is in agreement with their position in the $M_{\rm V}$ vs. $V-I$ diagram (Fig.~\ref{fig:VI}); finally, a set of stars with $EW$(Li~{{\sc i}}) and positions in the CMD compatible with an age of 80~--~120~Myr are mixed with Hercules-Lyra and AB~Dor MG, and have been classified as other members of the Local Association B4 subgroup (see $\S$~3). Note that the age estimated using the position of the members of Hercules-Lyra in the CMD is a lower limit since the 80~Myrs isochrone is overlapped with the ZAMS for spectral types earlier than about K5. On the other hand, the age estimated using the equivalent width of the Li~{\sc i} line $\lambda$6707.8 \AA \ is more robust since the 50\% of the stars classified as members have measurements of the $EW$(Li~{{\sc i}}): the Li indicator is useful only for spectral type later than G0, but only three of the 25 candidates of the initial sample are F stars. Stars in these three subgroups form an excellent list of young cool stars for studying how planets are formed, since they cover a range of ages between 30 and 200 Myr, characteristic of the period during which the Solar System was formed, and they are close enough to be accessible to adaptive optics. In addition, they can be taken as targets for direct imaging detection of sub-stellar companions ---~brown dwarfs and extra-solar giant planets~--- \citep{neu00,mar03,mas05,low05} and for cold dust, debris disks \citep{gai04,met04,liu04,che05}. \acknowledgments This work was supported by the Universidad Complutense de Madrid and the Spanish Ministerio de Educaci\'on y Ciencia (MEC), Programa Nacional de Astronom\'{\i}a y Astrof\'{\i}sica under grants AYA2004-03749 and AYA2005-02750. ICC acknowledges support from MEC under AP2001-0475. We would like to thanks the referee for useful comments which have contributed to improve the manuscript. \clearpage \clearpage \clearpage
Title: Did Swift measure GRB prompt emission radii?
Abstract: The Swift X-Ray Telescope often observes a rapidly decaying X-ray emission stretching to as long as $ t \sim 10^3$ seconds after a conventional prompt phase. This component is most likely due to a prompt emission viewed at large observer angles $\theta > 1/\Gamma$, where $\theta\sim 0.1$ is a typical viewing angle of the jet and$\Gamma\geq 100$ is the Lorentz factor of the flow during the prompt phase. This can be used to estimate the prompt emission radii, $r_{em} \geq 2 t c/\theta^2 \sim 6 \times 10^{15}$ cm. These radii are much larger than is assumed within a framework of a fireball model. Such large emission radii can be reconciled with a fast variability, on time scales as short as milliseconds, if the emission is beamed in the bulk outflow frame, e.g. due to a random relativistic motion of ''fundamental emitters''. This may also offer a possible explanation for X-ray flares observed during early afterglows.
https://export.arxiv.org/pdf/astro-ph/0601557
\date{\today} \title{Did {\it Swift} measure GRB prompt emission radii?} \author{M. LYUTIKOV } \affil{University of British Columbia, 6224 Agricultural Road, Vancouver, BC, V6T 1Z1, Canada and Department of Physics and Astronomy, University of Rochester, Bausch and Lomb Hall, P.O. Box 270171, 600 Wilson Boulevard, Rochester, NY 14627-0171, USA } \keywords{gamma-rays: burster} \section{Introduction} Recently launched \Swift satellite \citep{Gehrels} together with a network of ground based observations have been providing scientific community with crucial information on Gamma Ray Bursts (GRBs). Besides the landmark detection of afterglows from short GRBs \citep[\eg][]{Gehr05}, \Swift has gathered crucial data on developments of GRBs at early times. This is especially important since early observations provide clues to the properties of the ejecta, like its composition, lateral distribution of energy etc. At late times the energy is mostly transfered to the forward shock, properties of which can hardly be used to probe the ejecta. A number of surprising results related to early afterglows have emerged \citep[\eg][]{Tagliaferri,Nousek,Chincarini,obrien}: (i) early, $t \leq 10^3$ s, rapidly-decaying X-ray component, (ii) X-ray flares occurring at $t \sim 10^2-10^4$ s, (iii) shallower than expected initial decay (or hump) of the afterglow. These features are common, but the light curves show a large variety. In this letter we discuss the first two mentioned effects, \ie rapidly-decaying component and X-ray flares, since both can be related to the prompt emission (as opposed to afterglow) and can thus be used to probe the ejecta and the central engine. \section{Prompt emission radii} The initial fast-decaying part of afterglows can be a ''high altitude'' prompt emission, coming from angles $\theta > 1/\Gamma$ \citep{Kumar00,Barthelmy05}, where $\theta$ is the angle between the line of sight and the direction from the center of the explosion towards an emitting point and $\Gamma$ is the Lorentz factor of the outflow. For a $\delta$-function in time prompt emission pulse, after an initial spike the observed flux should decay as $t^{-(2+\alpha)}$, where $\alpha \approx 0.5 $ is prompt emission's spectral index, \citep{Fenimore}, roughly consistent with observations. One also expects that prompt and early afterglow emission join smoothly, which seems to be generally observed \citep{obrien}. [Exceptions, like GRB050219a \citep{Tagliaferri}, may be due to interfering X-ray flares.] If we accept the interpretation of the fast decaying part as ''high altitude'' prompt emission, one can then determine radii of the prompt emission and compare them with model predictions. The currently most popular fireball model \cite[\eg][]{Piran04} relates radii of emission $r_{em}$ to the variability time scale $\delta t$ of the central source $r_{em} \sim 2 \Gamma_0^2 c \delta t$, where $ \Gamma_0 \sim 100-300$ is the initial Lorentz factor. Within the framework of the fireball model this is also the variability time scale of the prompt emission. Observationally, prompt emission shows variability on time scales as short as milliseconds, while most power is at a fraction of a second \citep{BeloStern}. Adopting $\delta t\sim 0.1 $ s, the prompt emission radius is $r_{em} \sim 6 \times 10^{13}$ cm $(\Gamma_0/100)^2$. If the emission is generated at $r_{em}$ and is coming to observer from large angles, $\theta > 1/\Gamma$, its delay with respect to the start of the prompt pulse is $t \sim (r_{em}/c) \theta^2/2$. If one can estimate $\theta$, then this can be used to measure $r_{em}$. This can be done from late ''jet breaks'', giving typically $\theta \sim 0.1$ \cite[\eg][]{Frail}. Then, for the X-ray tail of the prompt emission extending to $t \sim 1000$ seconds, the implied emission radius is $r_{em} > 6 \times 10^{15}$ cm. This is much larger than is assumed in the fireball model. To make it consistent with the fireball model and variability on short times scales, the Lorentz factor of the flow should be huge, $\Gamma_0 \geq 1000$, but this would imply that emission is strongly de-boosted, $\Gamma_0 \theta \sim 100 $. Increasing $\theta$ cannot save the day either since the required viewing angle would be $\theta \sim 1$, implying a jet moving always from an observer. Along the similar lines of reasoning, \cite{Lazzati05} estimated prompt emission radii for a particular case of GRB 050315 for which a possible jet break is identified \citep{Vaughan} . Steep decay in that case is relatively short and lasts for 100 s, giving $r_{em} > 2.5 \times 10^{14}$ cm. Note, that any observed duration of the steep decay phase provides only {\it a lower limit} on the prompt emission radius since the end of the steep decay may be related to emergent afterglow emission and not to the fact that the edge of the jet becomes visible, see Fig. \ref{GRBafter}. On the other hand, late jet break time provides an estimate of the total opening angle of the jet. In any case, GRBs with longer lasting steep decay phase, up to $10^3$ s, provide the most severe constraints on the models. Thus, the interpretation of the fast-decaying initial X-ray light curve as prompt emission seen at large angles can hardly be inconsistent with the fireball model. We should then either look for alternative possibilities to produce the fast-decaying part of the X-ray light curve \citep[\eg][]{mr01}, or consider models that advocate production of prompt emission at larger radii, see \S \ref{Conc}. \section{Fast variability from large radii} If prompt emission is produced at distances $\sim 10^{15}-10^{16}$ cm, how can a fast variability, on times scales as short as milliseconds, be achieved? One possibility, is that emission is beamed in the outflow frame, for example due to a relativistic motion of (using pulsar physics parlance) ''fundamental emitters'' \citep{lb03}. To prove this point, we consider an spherical outflow expanding with a bulk Lorentz factor $\Gamma$ with $N$ randomly distributed emitters moving with respect to the shell rest frame with a typical Lorentz factor $\gamma_{T}$. Highly boosted emitters, moving towards an observer, have a Lorentz factor $ \gamma \sim 2 \gamma_T \Gamma $ in the observer frame. If emission is generated at distances $r_{em}$, the observed variability time scale can be as short as $ \sim (r_{em}/c) /2 \gamma^2 \approx (r_{em}/c) / 8 ( \gamma_T \Gamma)^2$, so that modest values of $\gamma_T \sim 5-10 \ll \Gamma \sim 100-300 $ would suffice to produce a short time scale variability from large distances $r_{em}\sim 10^{15} - 10^{16} $ cm. The model should satisfy a number of constraints. First, the number of sub-jets directed towards an observer from viewing angles $\theta< 1/\Gamma$ should be larger than unity (in order to produce at least one true prompt emission spike), but should not be too large, otherwise prompt emission will be a smooth envelope of overlapping spikes. If a typical jet opening angle is $\theta_j$, then the number of sub-jets seen ''head-on'' from angles $< 1/\Gamma$ is \be n_{prompt} \sim { \pi N \over (\Gamma \gamma_{T} \theta_j)^2}. \ee This should be larger than $1$. The second constraint that the model should satisfy relates to the efficiency of energy conversion. Suppose that the thickness of an outflowing shell in its rest frame is $ L_{shell} \sim t_s c \Gamma$, where $t_s $ is a source activity time ($t_s \sim 30-100$ s for long bursts and $t_s \sim 1 $ s for short bursts). Suppose then that fundamental emitters operate for a time $ t_{pulse} = \eta_t L_{shell}$ in the flow frame, where $\eta_t $ is a dimensionless parameter. During this time the source can tap into energy contained within volume $ (c t_{pulse})^3$. The ratio of this volume times the number of emitters to the total volume of the shell is a measure of efficiency of energy conversion into radiation: \be \eta = { N (c t_{pulse})^3 \over r_{em}^2 \theta_j ^2 t_s c \Gamma} \ee Since tapping of energy in the volume $(c t_{pulse})^3$ is a definite upper limit on conversion efficiency, in the calculations we allow $\eta$ defined above to be slightly larger than unity. To produce light curves we calculate the intensity of emission from sub-jets that are randomly located within the shell and moving in random direction with random Lorentz factors $1<\gamma_T < \gamma_{T,max} =5 $. Each emitter is isotropic in its rest frame and is active for a random time $0<t'_{em} < \eta_T t_s c \Gamma = t_{pulse,max}$ with $\eta_T =0.5$. The observed intensity of emission from each sub-jet $\propto \delta^{3+\alpha}$ \citep{BlandfordLind}, where $\delta = 1/\gamma (1-\beta \cos \theta_{sj}) $ is a total Doppler factor including bulk and random motion, $\theta_{sj}$ is an angle between the line of sight and direction of the sub-jet motion. As the burst progresses, larger angles and more of sub-jets producing prompt emission become visible. Most of them will be seen from angles $> 1/\gamma_T$ in the bulk frame, producing a combined smooth curve overlaid with spikes. The average Doppler factor decreases with time $\delta \approx t_s \Gamma / t $ and the average flux decays as $t^{-(2+\alpha)}\approx t^{-2.5}$ for $\alpha=0.5$. In Fig. \ref{GRBafter} we plot an example of a prompt light curve in this model. \subsection{Lateral dependence of prompt emission} Variations of the decay rate from the $t^{-(2+\alpha)}$ law may be used to probe angular dependence $L(\theta_{axis})$ of the intensity of the prompt emission, where $\theta_{axis}$ is an angle between the axis of the explosion and an emitting point. More shallow decays can be due to, \eg, a structured jet, with $L\sim \theta_{axis}^{-2}$ observed outside of some core: late time emission then is coming from the more energetic core part. The effective emission intensity increases approximately as $\theta ^2 \propto t$, and will result in an observed decay $t^{-(1+\alpha)}$. Similarly, if the prompt emission is seen within a core, late emission comes from less energetic wings, giving in case of a structured jet a flux $ \propto t^{-(3+\alpha)}$. Qualitatively, the relativistic internal motion of emitters makes it ''easier'' to see the high altitude emission. To show this numerically we parameterize the {\it number density of emitters} as $n(\theta) \propto 1/(\theta^2 + \theta_0^2)$, where $\theta_0$ is an angular core radius. [There are, naturally other possible parameterizations, e.g. of intensity of each emitter]. The results are presented in Fig. \ref{GRBafterStruct}. We can also expect deviations from a simple power-law decay due to not exactly spherical form of the emitting surface. Such distortions are expected due to a development of the Kelvin-Helmholtz instability during an accelerating phase of the outflow. They won't be erased during the coasting stage due to causal disconnection of the flow separated by angles $> 1/\Gamma$. Additional complications may come from the way the data analysis is performed, \eg through a choice of initial time trigger (\cite{Zhang05}, see also \cite{Lazzati05}). \section{Origin of X-ray flares} Early X-ray light curves show complex behavior with flares and frequent changes in a temporal slope \citep[\eg][]{obrien}. Flares show very short rise and fall times, much shorter than observation time after the on-set of a GRB, while the underlying afterglow has the same behavior before and after the flare \citep{Burrows} (though there are exceptions). Both of these observations argue against a physical process in the forward shock. In addition, there is a hardening of the spectrum during X-ray flares \citep{Burrows}. In the present model we interpret X-ray flares as been due to sub-jets located at large viewing angles, $\theta > 1/\Gamma$, but directed towards an observer. Randomly located, narrow spikes are clearly seen in the model light curves, Figs. \ref{GRBafter}-\ref{GRBafterStruct}. In addition, as the flares are less de-boosted than the average high altitude outflow, they will have a harder spectrum, as observed. \section{Discussion} \label{Conc} In this letter we first point out that the interpretation of the initial fast-decaying part of the X-ray GRB light curves as a prompt emission seen at large angles, and a generic estimate of jets' opening angle allows a measurement of the radius of prompt emission, which turns out to be relatively large, $> 10^{15}$ cm. On basic grounds, $\gamma$-ray emission should be generated before the deceleration radius $r_{dec} \sim \left( { E_{iso} \over 4 \pi \rho c^2 \Gamma_{dec}^2} \right)^{1/3}\sim 10^{16}- 10^{17}$ cm, when most energy of the outflow is given to the surrounding medium (here $ E_{iso}$ is isotropic equivalent energy, $ \rho$ is density of external medium, $\Gamma_{dec}$ is Lorentz factor at $r_{dec}$). \footnote{Note, that $r_{dec}$ defined above is {\it independent} of ejecta content, contrary to the claim in \cite{zk05}, see \cite{LZK}.} The inferred emission radius is within this limit. The estimate of the emission radius is very simple, and, in some sense, generic. It can hardly be consistent with the fireball model, unless extreme assumptions are made about the parameters (\eg very large Lorentz factor). On the other hand, there are alternative models (\eg the electromagnetic model \citep{l05}, see also \cite{thompson}) that place prompt emission radii at large distances, just before the deceleration radius $r_{dec} $. Secondly, we show how models placing emission at large radii may be able to reproduce a short time scale variability of the prompt emission and explain later X-ray flares. This can be achieved if the prompt emission is beamed in the rest frame of the outflow, which may be due to an internal relativistic motion of ''fundamental emitters''. What can produce a relativistic motion in the bulk frame? It can be due, for example, to a relativistic Burgers-type turbulence (a collection of randomly directed shock waves). It is not clear how such turbulence may be generated. Alternatively, relativistic internal sub-jets can result from reconnection occurring in highly magnetized plasma with $\sigma \gg 1$, where $\sigma$ is a plasma magnetization parameter \citep{KC84}. In this case the matter outflowing from a reconnection layer reaches relativistic speeds with $\gamma_{out} \sim \sigma$ \cite[]{lu03}. Internal synchrotron emission by such jets, or Compton scattering of ambient photons, will be strongly beamed in the frame of the outflow. Note, that {\it this model does not require late engine activity } to produce flares. One of the main observational complications is that at observer times larger than the conventional prompt phase, the X-ray light curve is a sum of the tail of the prompt emission, coming presumably from internal dissipation in the ejecta, and the forward shock emission. It is not obvious how to separate the two components. For example, GRBs which do not show a fast initial decay may be dominated by the forward shock emission from early on \citep{obrien}. This uncertainty also affects estimates of the emission radius since the end of the steep decay may be related to the emergent afterglow emission and not to the jet opening angle (or observer's angle, in case of a structured jet), see Fig. \ref{GRBafter}. Another complication is that at these intermediate times, $10^3 \leq t \leq 10^4$ s, even the forward shock emission itself often does not conform to the standard afterglow models, showing flatter than expected profiles \citep[\eg][]{Nousek}. A consequence of the model is that {\it some} short GRBs may be just a single spike directed towards an observer of a long GRBs. In our model the shorter spikes are highly beamed, less frequent and produce harder emission. This can apply only to {\it some} short GRBs since as a class they are well established to have different origin than long GRBs (from non-observation of a supernova signature and coming from a distinctly different host galaxy population).
Title: Development of Gaseous Tracking Devices for the Search of WIMPs
Abstract: The Time Projection Chamber (TPC) has been recognized as a potentially powerful detector for the search of WIMPs by measuring the directions of nuclear recoils, in which the most convincing signature of WIMPs, caused by the Earth's motion around the Galaxy, appears. We report on the first results of a performance study of the neutron exposure of our prototype micro-TPC with Ar-C$_2$H$_6$ (90:10) and CF$_4$ gas of 150 Torr.
https://export.arxiv.org/pdf/astro-ph/0601568
\begin{frontmatter} \title{Development of Gaseous Tracking Devices for the Search of WIMPs} \author[KYOTO]{H. Sekiya\corauthref{cor}}, \corauth[cor]{Corresponding author. tel:+81 75 753 3868; fax:+81 75 753 3799.} \ead{sekiya@cr.scphys.kyoto-u.ac.jp} \author[KYOTO]{K. Hattori}, \author[KYOTO]{S. Kabuki}, \author[KYOTO]{H. Kubo}, \author[KYOTO]{K. Miuchi}, \author[WASEDA]{T. Nagayoshi}, \author[KYOTO]{H. Nishimura}, \author[KYOTO]{Y. Okada}, \author[KOBE]{R. Orito}, \author[KYOTO]{A. Takada}, \author[ICRR]{A. Takeda}, \author[KYOTO]{T. Tanimori}, \author[KYOTO]{K. Ueno} \address[KYOTO]{Department of Physics, Graduate School of Science, Kyoto University, Kitashirakawa, Sakyo, Kyoto, 606-8502, Japan} \address[WASEDA]{Advanced Research Institute for Science and Engineering, Waseda University, \\ 17 Kikui-cho, Shinjuku, Tokyo, 162-0044, Japan} \address[KOBE]{Department of Physics, Graduate School of Science and Technology, Kobe University, 1-1 Rokkoudai, Nada, Kobe, 657-8501, Japan} \address[ICRR]{Kamioka Observatory, ICRR, University of Tokyo,\\ 456 Higasi-mozumi, Hida-shi, Gifu, 506-1205, Japan} \begin{keyword} dark matter\sep WIMP \sep TPC\sep direction sensitive detector \PACS 95.35.+d\sep 29.40.Gx \sep 29.40.Cs \end{keyword} \end{frontmatter} \section{Introduction} \label{intro} It is considered by many that the galactic halo is composed of weakly interacting massive particles (WIMPs) as dark matter\cite{jung}. These particles could be detected directly by measuring the nuclear recoils produced by their elastic scattering off nuclei in detectors. The most convincing signature of WIMPs appears in the directions of nuclear recoils. It is provided by the Earth's large velocity through the isothermal galactic halo ($\sim$230 km/s). Hence, detectors sensitive to the direction of the recoil nucleus would have a great potential to identify WIMPs\cite{dir}. Time Projection Chambers (TPCs) with fine spacial resolutions are among such devices, and we are developing a micro TPC, which can detect three-dimensional fine tracks of charged particles\cite{miuchi}. Since the energy deposits of WIMPs to nuclei are only a few tens of keV and the range of nuclei is limited, the micro-TPC should be operated at low pressures. We also focused on the detection of WIMPs via spin-dependent(SD) interactions and are interested in operating the micro-TPC with CF$_4$\cite{NA}, because $^{19}$F has a special sensitivity to SD interactions for its unique spin structure\cite{collar}. In the present work, in order to examine the response of the micro-TPC to nuclear recoils at low pressures as a first step, we irradiated a 150 Torr Ar-C$_2$H$_6$ (90:10 mixture) gas (one of the standard gases for TPCs) and CF$_4$ with neutrons from $^{252}$Cf. The track lengths and deposited energies of Ar, C, and F recoils were investigated. \section{The micro-TPC} The prototype micro-TPC used in this measurements is shown in Fig.\ref{fig:DC}. The field cage consists of a drift cathode plane and nine 0.2 $\mu$m copper wires of 1cm pitch with connections of 10 M$\Omega$ resistor, which forms a uniform electric field in the detection volume of $10\times10\times10$ cm$^3$. The $\mu$-PIC\cite{upic} for 2-dimensional readout is $10\times10$cm$^2$ with 256 anode strips and 256 cathode strips each with a 400 $\mu$m pitch. We also used a GEM having a 10 cm$\times$10 cm$^2$ sensitive area as a sub-amplification device between the field cage and $\mu$-PIC, as illustrated in Fig. \ref{fig:DC}, which enables stable operation and avoids discharges with low HV operation of both the $\mu$-PIC and GEM. The details of this GEM are described in Refs.\cite{gem,hattori}. The output charges of $256+256$ channels are pre-amplified (0.7 V/pC) and shaped (with a gain of 7) and discriminated via ASD chips (4 channels/chip, SONY CXA3653Q)\cite{asd}. The pre-amplified signals are summed and digitized by 100 MHz 8bit flash ADCs in order to determine the deposited energy and the track direction as the waveforms hold the Bragg curve shapes. The reference threshold voltage ($0-100$mV) is commonly supplied to all the ASD chips and all discriminated digital signals are sent to the position encoding module based on FPGAs with an internal clock of 100 MHz, so that the anode and cathode coincident position (x,y) and the timing (z) are recorded in the memory module and the tracks of charged particles are reconstructed in software. The tracking performances for electrons, protons, and MIPs are reported elsewhere\cite{miuchi,hattori}. \section{Measurements and Results} As illustrated in Fig. \ref{fig:Setup}, the micro-TPC was set in a 6 mm-thick aluminum vessel of 60 cm diameter $\times$ 20cm height. In a typical run, the vessel was evacuated to $\sim8\times10^{-3}$ Torr, the SAES GETTER$^{\mbox{\scriptsize{\textcircled{\tiny R}}}}$ pump in the vessel was activated, and then the vessel was filled with Ar-C$_2$H$_6$ (90:10) or CF$_4$ gas to a pressure of 150 Torr and sealed for the duration of the measurement. For measuring the gas gain and the energy calibration, the gas was irradiated with $^{109}$Cd 22 keV and $^{133}$Ba 31.0 keV X-rays through a 1mm thick aluminum window close to the sensitive volume. We irradiated the micro-TPC with neutrons from a 1 MBq $^{252}$Cf source on the top of the vessel. Since one fission decay of $^{252}$Cf emits 3.8 neutrons and 9.7 $\gamma$-rays on average\cite{neut}, the $\gamma$-rays or neutrons detected by a $10\times10\times2$ cm$^3$ plastic scintillator were used as the event trigger. In the $\gamma$/n-triggered events, gamma events would dominate under normal gas gain ($\sim$10000) operation. Since the $dE/dx$ values of the neutron events are much larger than those of gamma events, we operated the $\mu$-PIC and GEM with a rather low gas gain (below 1000) in order to observe the nuclear recoils. In such different gas gain measurements, we fixed the anode voltage of the $\mu$-PIC and changed the voltage between the GEM electrodes. Below a gas gain of about 2000, our system was not able to measure the $^{109}$Cd 22 keV x-ray correctly due to a mismatch of the dynamic range of the ASD chips and the flash ADC; therefore, the deposited energy in low-gain operations was extrapolated from the calibrations with the high gas gain operations. We evaluated the track length as a function of the measured electron equivalent energy in the following way. \subsection{Ar-C$_2$H$_6$ 150Torr run} The drift cathode plane was supplied $-1$ kV, which gave a drift field of 60 V/cm and an electron drift speed of 4.0 cm/$\mu$s. The anode voltage of the $\mu$-PIC was fixed at 350 V. For nuclear recoil measurements, the threshold of the discriminator of the ASD chip was set to 80 mV and the measured track length of events when the GEM voltage was set to 200 V (gas gain of 3000) and 135 V (gas gain of 900) is shown in Fig. \ref{fig:ArTE}. The MC (Geant4\cite{geant}) simulated track length without consideration of the diffusion, the energy resolution, and the $dE/dx$ threshold is also indicated for a comparison. The geometry used for the simulation was in accordance with Fig. \ref{fig:DC} and Fig. \ref{fig:Setup} and, the neutron energy spectrum of the spontaneous fissions of $^{252}$Cf was assumed to be \begin{equation} \frac{dN}{dE}=\sqrt{E}\exp\left(-\frac{E}{T}\right), \end{equation} where $T=1.3$ MeV\cite{fission}. Under operation with a gas gain of 3000, electron recoils and proton (of C$_2$H$_6$) recoils were clearly observed according to their $dE/dx$. On the other hand, in the operation of the gas gain of 900, the C and Ar recoils and some proton recoils were observed due to the high $dE/dx$ threshold. \subsection{CF$_4$ 150Torr run} The drift cathode plane was supplied $-2$ kV, which gives a drift field of 120 V/cm and the electron drift speed of 12.0 cm/$\mu$s. The anode voltage of the $\mu$-PIC was fixed at 600 V. The measured track length of events when the GEM voltage was set to 215 V (gas gain of 4500) and 95 V (gas gain of 800) are shown in Fig.\ref{fig:CF4TE}. The threshold voltage of the discriminator of the ASD chip was as high as 100mV; therefore, only C and F recoils were clearly observed under operation with a gas gain of 800. \section{Discussion and Prospects} We successfully showed the nuclear recoils in 150 Torr of Ar-C$_2$H$_6$ (90:10) and CF$_4$ gases according to their $dE/dx$ by changing the detector threshold. The energy loss of protons became lower as the energy increased as opposed to the other nuclei\cite{srim}. Consequently, the proton band in Fig. \ref{fig:ArTE}(b) is truncated at the threshold set in the measurements, which corresponds to about 5 keV/400$\mu$m. In terms of $dE/dx$, the tracks in the micro TPC were much easier to detect for C, F or Ar recoils. Ultimately, our concern is the recoil direction of such nuclei below 100 keV to allow us to observe the signals of WIMPs. In order to obtain longer tracks and clear Bragg curves, higher gas gain operations at lower pressures with low-energy neutron beams are needed. The measurement of the incident neutron energy with Time-Of-Flight may also be useful to examine the quenching factor of nuclear ionization in the micro-TPC.
Title: Multi-Dimensional Simulations of the Accretion-Induced Collapse of White Dwarfs to Neutron Stars
Abstract: We present 2.5D radiation-hydrodynamics simulations of the accretion-induced collapse (AIC) of white dwarfs, starting from 2D rotational equilibrium configurations of a 1.46-Msun and a 1.92-Msun model. Electron capture leads to the collapse to nuclear densities of these cores within a few tens of milliseconds. The shock generated at bounce moves slowly, but steadily, outwards. Within 50-100ms, the stalled shock breaks out of the white dwarf along the poles. The blast is followed by a neutrino-driven wind that develops within the white dwarf, in a cone of ~40deg opening angle about the poles, with a mass loss rate of 5-8 x 10^{-3} Msun/yr. The ejecta have an entropy on the order of 20-50 k_B/baryon, and an electron fraction distribution that is bimodal. By the end of the simulations, at >600ms after bounce, the explosion energy has reached 3-4 x 10^{49}erg and the total ejecta mass has reached a few times 0.001Msun. We estimate the asymptotic explosion energies to be lower than 10^{50}erg, significantly lower than those inferred for standard core collapse. The AIC of white dwarfs thus represents one instance where a neutrino mechanism leads undoubtedly to a successful, albeit weak, explosion. We document in detail the numerous effects of the fast rotation of the progenitors: The neutron stars are aspherical; the ``nu_mu'' and anti-nu_e neutrino luminosities are reduced compared to the nu_e neutrino luminosity; the deleptonized region has a butterfly shape; the neutrino flux and electron fraction depend strongly upon latitude (a la von Zeipel); and a quasi-Keplerian 0.1-0.5-Msun accretion disk is formed.
https://export.arxiv.org/pdf/astro-ph/0601603
\title{Multi-Dimensional Simulations of the Accretion-Induced Collapse of White Dwarfs to Neutron Stars} \author{L. Dessart\altaffilmark{1}, A. Burrows\altaffilmark{1}, C.D. Ott\altaffilmark{2}, E. Livne\altaffilmark{3}, S.-Y. Yoon\altaffilmark{4}, N. Langer\altaffilmark{5} } \altaffiltext{1}{Department of Astronomy and Steward Observatory, The University of Arizona, Tucson, AZ \ 85721; luc@as.arizona.edu,burrows@as.arizona.edu} \altaffiltext{2}{Max-Planck-Institut f\"{u}r Gravitationsphysik, Albert-Einstein-Institut, Golm/Potsdam, Germany; cott@aei.mpg.de} \altaffiltext{3}{Racah Institute of Physics, The Hebrew University, Jerusalem, Israel; eli@frodo.fiz.huji.ac.il} \altaffiltext{4}{Astronomical Institute ``Anton Pannekoek'', University of Amsterdam, Kruislaan 403, 1098 SJ, Amsterdam, The Netherlands; scyoon@science.uva.nl} \altaffiltext{5}{Astronomical Institute, Utrecht University, Princetonplein 5,3584 CC, Utrecht, The Netherlands; n.langer@astro.uu.nl} \keywords{hydrodynamics -- neutrinos -- rotation -- stars: neutron -- stars: supernovae: general -- stars: white dwarfs} \section{Introduction} Stars can follow a few special evolutionary routes to form an unstable Chandrasekhar mass core. A main-sequence star of more than $\sim$8\,\mo evolves to form either a degenerate O/Ne/Mg core (Barkat et al. 1974; Nomoto 1984,1987; Miyaji \& Nomoto 1987) or a degenerate Fe core (Woosley \& Weaver 1995), which, due to photodisintegration of heavy nuclei and/or electron capture, collapses to form a protoneutron star (PNS). If an explosion ensues, the event is associated with a Type II supernovae (SN). Less massive stars end their lives as white dwarfs. White dwarfs located in a binary system may accrete from a companion and achieve the Chandrasekhar mass, triggering the thermonuclear runaway of the object and leading to Type Ia SN, leaving no remnant behind. However, a third class of objects is expected. Theoretically, massive white dwarfs with O/Ne/Mg cores, due to their high central density ($\sgreat$10$^{10}$\,g\,cm$^{-3}$), experience rapid electron capture that leads to the collapse of the core. This is accretion-induced collapse (AIC), an alternative path to stellar disruption through explosive burning, currently associated with Type Ia SN (Nomoto \& Kondo 1991). It is presently unclear what fraction of all white dwarfs will lead to AICs, but of those white dwarfs that evolve to form a Chandrasekhar-mass O/Ne/Mg core, all will necessarily undergo core collapse. One formation channel is the coalescence of two white dwarfs (Mochkovitch \& Livio 1989), with either C/O or O/Ne/Mg cores, although few such binary systems have yet been observed with a cumulative mass above the Chandrasekhar mass. There is still uncertainty as to whether such binary systems would not undergo thermonuclear runaway rather than collapse. Since the coalescence of two white dwarfs requires a shrinking of the orbit through gravitational radiation, these systems will take many gigayears to coalesce. An alternative formation mechanism is via single-degenerate systems, through a combination of high original white dwarf mass and mass and angular-momentum accretion by mass transfer from a (non-degenerate) H/He star (Nomoto \& Kondo 1991). Binary star population synthesis codes predict the occurence of the AIC of white dwarfs with a galactic rate of 8$\times$10$^{-7}$ yr$^{-1}$ to 8$\times$10$^{-5}$ yr$^{-1}$, depending, amongst other things, on the treatment of the common-envelope phase and mass transfer (Yungelson \& Livio 1998). The set of parameters leading to a Type Ia rate of 10$^{-3}$\,yr$^{-1}$ corresponds to an AIC rate of 5$\times$10$^{-5}$ yr$^{-1}$. The observed Type Ia rate of $\sim$3$\times$10$^{-3}$\,yr$^{-1}$ (Madau et al. 1998; Blanc et al. 2004; Manucci et al. 2005) would imply a galactic AIC rate of 1.5$\times$10$^{-4}$\,yr$^{-1}$. These rates are likely functions of galaxy and metallicity (Yungelson \& Livio 2000; Belczynski et al. 2005; Greggio 2005; Scannapieco \& Bildsten 2005). Based on r-process nucleosynthetic yields obtained from previous simulations of the AIC of white dwarfs, Fryer et al. (1999) inferred rates ranging from $\sim$10$^{-5}$ to $\sim$10$^{-8}$\,yr$^{-1}$. Overall, AICs are not expected to occur more than once per 20--50 standard Type Ia events; because they are intrinsically rarer, they remain to be identified and observed in Nature. Whatever their origin, their fundamental nature is to accrete both mass and angular momentum from a companion object. Rotation is, therefore, a key physical component. Prior to core collapse, differential rotation acts as a stabilizing agent for shell burning by widening its spatial extent through enhanced mixing and reducing the envelope density through centrifugal support (Yoon \& Langer 2004). Mass accretion also leads to an increase of the central density, which may rise up to a few $10^{10}$ g\,cm$^{-3}$, establishing suitable conditions for efficient electron capture on Mg/Ne nuclei. By the time of core collapse and depending on the evolutionary path followed, such white dwarfs may cover a range of masses from $\sgreat$1.35\,\mo up to $\sim$2\,\mo (2.7\,\mo) in the case of a non-degenerate (degenerate) companion, potentially well in excess of the standard Chandrasekhar limit, and possessing an initial rotational energy up to $\sim$10\% of their gravitational binding energies. At such values, the centrifugal potential leads to a deformation of the white dwarf from spherical symmetry, equipotentials and isopressure surfaces adopting a peanut-like shape in cross section for the largest rotation rates. Such structures are obtained in the 2D differentially rotating equilibrium white dwarf models constructed by Yoon \& Langer (2005, YL05; see also Liu \& Lindblom 2001). In the past, the collapse of O/Ne/Mg cores originating from stars in the 8-10\,\mo range has been studied in 1D by Baron et al. (1987ab), Mayle \& Wilson (1988), and Woosley \& Baron (1992), who showed that the shock generated at core bounce stalls rather than leading to a prompt explosion. However, Hillebrandt et al. (1984) and Mayle \& Wilson (1988) obtained delayed explosions, the former after 20-30\,ms, and the latter after $\sim$200\,ms, supposedly driven by neutrino energy deposition behind the stalled shock. Woosley \& Baron (1992) found the emergence of a sustained neutrino-driven wind, with a mass loss rate of 0.005\,\mo\,s$^{-1}$ and with an ejecta electron fraction of the order of 0.45, showing promise, modulo uncertainties, for a contribution to the enrichment of the ISM in r-process elements. Using 1D/2D SPH simulations, Fryer et al. (1999) reproduced the simulations of Woosley \& Baron (1992), confirmed that the shock stalls due to the copious neutrino losses associated with core bounce, did not find prompt explosion, and, depending on the equation of state (EOS) employed, observed a delayed explosion. They focused mostly on the early phase, prior to the neutrino-driven wind, and found an ejected mass of low-$Y_{\rm e}$ material of $\sim$0.05\,\mo, depending on adopted model assumptions. Their 2D simulations with solid-body rotation showed similar properties to their 1D equivalents, the authors attributing the small differences to the different grid resolution. This may partly stem from the essentially spherical explosion triggered just $\la$100\,ms after core bounce, ejecting the fast-rotating material in the outer mantle. Consequently, at the end of their 2D simulations, they obtain very slow rotation rates for the PNS, i.e., of $\sim$1\,s. Recently, Kitaura et al. (2005) re-inspected the collapse of the progenitor model used by Hillebrandt et al. (1984). They confirmed again the by-now well-accepted idea that no prompt explosion occurs, but instead obtain a successful, though sub-energetic, delayed explosion in spherical symmetry, powered by neutrino heating and a neutrino-driven wind that sets in $\sim$200\,ms after bounce. A primary motivation for this work is to improve upon these former investigations that assumed one-dimensionality, sphericity, and/or zero-rotation, and start instead from the more physically-consistent 2D models of YL05, thereby fully accounting for the effects of rotation on the collapse, bounce, and post-bounce evolution of the white dwarf core and envelope, as well as for the strong asphericity of the progenitor. Our study uses VULCAN/2D (Livne et al. 2004; Walder et al. 2005) to perform 2D Multi-Group Flux Limited Diffusion (MGFLD) radiation hydrodynamics simulations. As we will demonstrate, rotation plays a major role in the post-bounce evolution, making 1D investigations of such objects of limited utility. By carrying out the simulations from $\sim$30\,ms before bounce to $\sgreat$600\,ms after bounce, we capture a wide range of physical processes, including the establishment of a strong, fast, and aspherical neutrino-driven wind. We model the centrifugally-supported equatorial regions and the large angular momentum budget leading to the formation of a sizable accretion disk. Moreover, an Eulerian investigation is better-suited than a Lagrangean approach to explore the neutrino-driven wind that develops after $\sgreat$200\,ms. Finally, despite the relative scarcity of AIC in Nature, these simulations represent interesting examples for the formation of disks around neutron stars. The main findings of this work are the following: We find that the AIC of white dwarfs forms $\sim$1.4-\mo neutron stars, expelling a modest mass of a few 10$^{-3}$\,\mo mostly through a neutrino-driven wind that develops $\sgreat$200\,ms after bounce, and that they lead to very modest explosion energies of 5-10$\times$10$^{49}$erg. Accounting for the rotation and the asphericity of the progenitor white dwarfs reveals a wealth of phenomena. The shock wave generated at core bounce emerges through the poles rather than the equator, and it is in this excavated polar region that the neutrino-driven wind develops. The strong asphericity of the newly-formed protoneutron stars leads to a latitudinally-dependent neutrino flux, while the effects of rotation modify the relative flux magnitude of different neutrino flavors. Besides mass accretion by both the neutron star and mass ejection by the initial blast and the subsequent wind, we find a sizable component that survives and resides in a quasi-Keplerian disk, which obstructs the wind flow at low latitudes. This disk will be accreted by the neutron star only on longer, viscous timescales. In the next two sections, we present the two selected progenitor models in more detail; we also discuss the radiation-hydrodynamics code VULCAN/2D and the various assumptions made. In \S\ref{sect_results}, we present the simulation results, focusing on the general temporal evolution from the start until $\sgreat$600\,ms, a time by which the neutrino-driven wind has reached a steady state. We then analyse in more detail the various components of these simulations. In \S\ref{sect_pns}, we discuss the properties of the nascent neutron stars, with special attention paid to the geometry of the neutrinospheres. In \S\ref{sect_nu}, we discuss the neutrino signatures, both in terms of luminosity and energy distribution. In \S\ref{sect_disk}, we focus on the residual material lying at low latitudes, forming a quasi-Keplerian disk. In \S\ref{sect_ener}, we turn to the energetics of the explosion and describe in detail the main component of the simulations at late times, i.e., the neutrino-driven wind. In \S\ref{sect_ye}, we analyse the electron fraction of the ejected material and address the relevance of the AIC of white dwarfs for neutron-rich element pollution of the interstellar medium. In \S\ref{sect_gw}, we present the gravitational-wave signal predicted for the AIC of our white dwarf models. In \S\ref{sect_conc}, we wrap up with a discussion of the main results of this investigation and present our conclusions. \section{Initial models} \label{sect_progenitor} In this section, we present the properties of the AIC progenitors selected in our study and summarize the presentation in \S2 of YL05. The general assumption for the construction of 2D progenitor models for the AIC of white dwarf (YL05) is that the resulting structure of the object is essentially independent of its evolutionary history, the only factors that matter being the given (final) mass, angular momentum, and central density, $\rho_{\rm c}$. Additionally, the angular velocity distribution $\Omega(r,z)$, where $r$ is the cylindrical radius and $z$ is the distance to the equator, is determined self-consistently, given a number of properties identified in 1D models: 1) the role played by the dynamical shear instability, 2) the compression (and spin up) of the surface layers due to mass accretion which puts its peak angular velocity interior to the surface radius, 3) the adopted surface rotational velocity value (its fraction of the local Keplerian value), and 4) the geometry of the angular velocity profile, which we assume to be constant on cylinders, i.e., $\Omega = \Omega(r)$. Together with the pressure/density dependence $P = P(\rho)$, such rotating stars are called barotropic. Note that the criterion used in YL05 for the shear rate for the onset of the dynamical shear instability is determined using the EOS of Blinnikov et al. (1996). The 2D rotating models then correspond to equilibrium configurations iteratively found from trial density and angular velocity distributions, using the Self-Consistent-Field method (Ostriker \& Mark 1968; Hachisu 1986), under the constraint that the density $\rho$ is solely a function of the effective potential $\Psi(r,z)$, given as the sum of the gravitational potential, i.e., $$\Phi(r,z) = -G \int \frac{\rho(r',z')}{|{\bf R}-{\bf R'}|} d^3 R' \,, $$ and the centrifugal potential, i.e., $$\Theta(r) = -\int \Omega^2(r') r' dr' \,,$$ where $R' = \sqrt{r'^2 + z'^2}$ (Tassoul 2000, YL05). In this paper, we select two progenitors with masses of 1.46 \mo and 1.92 \mo and present their global characteristics in Table~\ref{tab_aic}. Both models have an initial central density $\rho_{\rm c}$ equal to 5$\times$10$^{10}$\,g\,cm$^{-3}$. The 1.46-\mo model serves as a reference for an object with a moderate initial rotational energy $T$ relative to gravitational energy $|W|$, i.e., $T/|W| = 0.0076$, and, indeed, shows only a modest initial departure from spherical symmetry, with a polar ($R_{\rm p}$) to equatorial ($R_{\rm eq}$) radius ratio $R_{\rm p}/R_{\rm eq} = 0.7$. At the other end of the white dwarf mass spectrum, the 1.92-\mo model has considerable rotational energy, with initial $T/|W| = 0.0833$, and the morphology of the star departs strongly from spherical symmetry, with $R_{\rm p}/R_{\rm eq} = 0.28$. For later reference, we also provide in Table~\ref{tab_aic} the ratio $T/|W|$ at the end of the simulations. In practice, the infall of the ambient material causes numerical difficulties soon after the start of the simulation. These difficulties were resolved by trimming the outer and low-density layers of both white dwarf progenitors. While the polar radius is hardly affected, the equatorial radius is reduced to 980\,km (down from 1130\,km) for the 1.46-\mo model and 1860\,km (down from 2350\,km) for the 1.92-\mo model. The progenitor mass is, however, reduced by less than one part in 10$^6$. Hence, we do not expect any perceptible effect on the results. \begin{deluxetable}{lccccccc} \tablewidth{8cm} \tabletypesize{\scriptsize} \tablecaption{Properties of selected AIC progenitors \label{tab_aic}} \tablehead{ \colhead{M}& \colhead{$R_{\rm p}$}& \colhead{$R_{\rm eq}$}& \colhead{$J$}& \colhead{$T$}& \colhead{$|W|$}& \colhead{$T/|W|$}& \colhead{$T/|W|$} \\ \colhead{\mo}& \colhead{km}& \colhead{km}& \colhead{erg$\cdot$s}& \colhead{erg}& \colhead{erg}& \colhead{initial}& \colhead{final} \\ \colhead{}& \colhead{}& \colhead{}& \colhead{(10$^{50}$)}& \colhead{(10$^{50}$)}& \colhead{(10$^{50}$)}& \colhead{}& \colhead{} } \startdata 1.46 & 800 & 1130 &0.160 & 0.7 & 91.97 & 0.0076 & 0.059 \\ 1.92 & 660 & 2350 &1.092 & 10.57 & 126.9 & 0.0833 & 0.262 \\ \enddata \end{deluxetable} \section{VULCAN/2D Simulation Code} \label{sect_VULCAN} The simulations discussed in this paper were performed with the Newtonian hydrodynamic code VULCAN/2D (Livne 1993), supplemented with an algorithm for neutrino transport as described in Livne et al. (2004) and Walder et al. (2005). The version of the code used here is the same as that discussed in Dessart et al. (2005) and Burrows et al. (2005), and uses the 2D Multi-Group Flux-Limited Diffusion (MGFLD) method to handle neutrino transport (see Appendix~A of Dessart et al. 2005). The MGFLD variant of VULCAN/2D is much faster than the more accurate, but considerably more costly, multi-angle $S_n$ variant. Doppler velocity-dependent terms are not included in the transport, although advection terms are. Frequency redistribution due to the subdominant process of neutrino-electron scattering is neglected. Our calculations include 16 energy groups logarithmically distributed in energy from 2.5 to 220\,MeV, take into account the electron and anti-electron neutrinos, and bundle the four additional neutrino and anti-neutrino flavors into a ``$\nu_{\mu}$'' component. VULCAN/2D uses a hybrid grid, switching from Cartesian in the inner 20\,km to spherical-polar further out. In the simulation of the 1.46-\mo model, mapped onto a 180$^{\circ}$ wedge, nearly perfect top-bottom symmetry about the equatorial plane was maintained during the pre- and post-bounce evolution. Thus, for the 1.92-\mo, we limited the computational domain to just one hemisphere. The thus reduced number of zones was used to increase the resolution. To summarize, the 1.46-\mo model uses a grid with a maximum resolution in the Cartesian inner region of 0.56\,km, and a minimum resolution of 150\,km at a maximum radius of 5000\,km, with 121 regularly spaced angular zones to cover 180$^{\circ}$. The 1.92-\mo model uses a grid with a maximum resolution in the Cartesian inner region of 0.48\,km, and a minimum resolution of 100\,km at a maximum radius of 4000\,km, with 71 regularly-spaced angular zones covering 90$^{\circ}$. A tricky part of the set up was to choose the properties of the ``ambient'' medium surrounding the AIC model. This need arises because our inputs are Lagrangean in spirit, while VULCAN/2D employs a Eulerian grid. Ideally, one would like to have material that merely occupies the space that will soon, after bounce, be replaced by the ejected material following the explosion. A key requirement is, thus, that this material have a very low pressure, to influence as little as possible the properties of the blast, but to allow for a smooth transition from circumstellar to ejected material in a given region of the Eulerian grid. To achieve this, we extended our SHEN EOS (Shen et al. 1998) down to very low densities (10\,g\,cm$^{-3}$) and low temperatures (10$^8$ K), conditions in which the medium is actually radiation-dominated and, thus, has a pressure that depends mostly on temperature. A successful choice was to adopt a density of 1000\,g\,cm$^{-3}$ and a temperature of 4$\times$10$^{8}$\,K for the ambient medium surrounding both white dwarf models. A major deficiency of the white dwarf progenitor models used here is their unknown initial thermal structure, which YL05 did not provide. Given the additional difficulty in handling low temperatures and high densities, we resorted to using a parameterized function of the local density, i.e., $$ T(r,z) = T_{\rm c} (\rho_{\rm c}/\rho(r,z))^{0.35}\,,$$ with $T_{\rm c} = 10^{10}$\,K for the 1.46-\mo model, and $T_{\rm c} = 1.3 \times 10^{10}$\,K for the 1.92-\mo model. Note that, similarly, Woosley \& Baron (1992) were forced to set a central temperature of 1.2$\times$10$^{10}$\,K at the start of their simulation. Figure~\ref{fig_init} recapitulates the basic properties of the white dwarf progenitors (left column: 1.46-\mo model; right column: 1.92-\mo model) as mapped onto our Eulerian grid. In the top panel, we show a color map of the density on which we superpose line contours of the effective potential $\Psi$ (defined above), as computed by VULCAN/2D. Our computation of the gravitational potential is based on a multipole expansion in spherical harmonics up to $l=33$. We reproduce the fundamental property of these fast-rotating white dwarf progenitors, in that the isopressure surfaces and the equipotentials coincide. In the bottom panel, we plot the initial angular velocity field. To avoid the distortion of streamlines of the infalling ambient material, the angular velocity is set to zero outside the WD progenitor. Note how, from zero, the angular velocity rises to a maximum value near the outer equatorial radius in the 1.46-\mo model (left column), while it is much higher in the center of the white dwarf in the 1.92-\mo model (right column). As we will show below, rotation has a much bigger impact on the collapse phase for the latter configuration. We also overplot line contours of the temperature for each model, following the prescription outlined in the above paragraph. Finally, we adopt an initial electron fraction of 0.5; the high initial central density permits fast electron capture which soon decreases the $Y_{\rm e}$ in the core, leading to bounce on a timescale ten times shorter than typically experienced in the collapse of the core of massive star progenitors (Woosley \& Weaver 1995; Heger et al. 2000; Woosley et al. 2002). \section{Simulation results} \label{sect_results} In this section, we discuss the general properties of the pre- and post-collapse phases for both models at the same time. We follow the post-bounce evolution of the 1.46-\mo (1.92-\mo) model for 550\,ms (780\,ms), for a total of 520\,000 (820\,000) timesteps. In Figs.~\ref{fig_seq46}--\ref{fig_seq92}, we provide entropy (saturated at 20\,k$_{\rm B}$/baryon, where k$_{\rm B}$ is Boltzmann's constant) color maps showing the key events in the evolution of both white dwarf models, with starting conditions discussed in the previous section and displayed in Fig.~\ref{fig_init}. We complement these figures with Fig.~\ref{fig_seqye} for the electron fraction evolution at three reference times (left column: 1.46-\mo model; right column: 1.92-\mo model). Note also that we overplot most figures with black arrows representing velocity vectors, whose maximum length is set to 10\% of the width (or the height) of the image. The corresponding maximum velocity is then given for each image in the figure captions - we also mention if we saturate the vector lengths. Having a large central density of 5$\times$10$^{10}$\,g\,cm$^{-3}$, both models achieve nuclear densities after the same time of $\sim$37\,ms. Differences in the bounce properties are attributable to the initial inner angular velocity distributions (Fig.~\ref{fig_init}). Compared to the 1.46-\mo model, the faster rotator has a lower maximum density at bounce (slightly shifted from the grid center to $r\sim$1\,km), i.e., 2.2 instead of 3.1$\times$10$^{14}$\,g\,cm$^{-3}$. It manifests an oblate, rather than a spherical, inner region (the inner tens of kilometers) of low entropy ($\sim$1$k_{\rm B}$/baryon; visible in red). The deleptonized material in this inner region is, however, aspherically distributed in both models, although more so in the 1.92-\mo model, with the lower $Y_{\rm e}$ material lying in a disk structure along the equatorial direction. The flatter density gradient in this direction and the longer dwell time near the neutrinosphere conspire to produce this enhanced deleptonization. At core bounce, the low-density outer parts of the white dwarf progenitor have not yet started to infall and the progenitor still retains its original shape. The collapse of the inner regions, however, creates a rarefaction wave that triggers the infall of the outer material, with a magnitude that is more pronounced along the poles due to the compact structure of the white dwarf in these directions. Although we do not observe a prompt explosion (i.e. occuring on a dynamical timescale of just a few milliseconds), the shock progresses slowly outwards without conspicuously stalling. This is slightly different from the core-collapse simulations recently published, whereby the shock {\it systematically} stalls (Burrows et al. 2005; Buras et al. 2005ab). In the equatorial direction, facilitated by the centrifugal support of the infalling material, the shock progresses outwards steadily, and is faster in the faster rotating model, reaching a few hundred kilometers after $\sim$100\,ms. Given the initial constant rotation rate on cylinders (see \S\ref{sect_progenitor}), the {\it radial} inflow of mass brings angular momentum to the equatorial ($z \sim 0$) region, which becomes more strongly supported, facilitating further the progress of the shock at low latitudes. This exacerbates the non-spherical development of the shock structure in both models. In its wake, we see a few large-scale whirls, resulting from the generation, at mid-latitudes and due to shock passage, of large vortical motions. \epsscale{1} In the polar direction, centrifugal support is absent, but the stalling of the shock is prevented by the quickly decreasing accretion rate established by the steeper density gradient, reduced polar radius, and smaller mass budget. Hence, although the shock slowly migrates along the equatorial regions, it soon traverses the surface layer of the white dwarf along the poles and escapes outwards into the ambient medium. Due to the strong asphericity of the progenitor, this occurs only 70\,ms after core bounce in the 1.92-\mo model, $\sim$30\,ms earlier than for the 1.46-\mo model. The ambient medium is then swept up by this blast, whose opening angle is constrained by that of the uncollapsed disk of the progenitor. The outflow expansion rate is larger 30-40 degrees away from the poles than right along the pole, coinciding in Fig.~\ref{fig_seqye} with the lower $Y_e$ material. The effect is large for the 1.92-\mo model, giving a butterfly shape in cross section to the faster expanding portions of the outflow. Off-axis material has more rotational kinetic energy available to convert to 2D planar ($r,z$) kinetic energy as it streams outward, reducing its rotational velocity while preserving angular momentum, thereby resulting in an enhanced acceleration compared to that of the material situated along the poles and lacking rotational energy. As the shock expands, it wraps around the disk, typically with a speed near that of the local sound speed. Nearer the pole, the outflow sweeps along the pole-facing side of the pole-excavated white dwarf, entraining surface material and effectively loading the outflow with more mass, causing unsteady fallback onto the neutron star. For the 1.46-\mo progenitor, the shock completely wraps around the low-latitude regions of the white dwarf, and finally emerges from the outer equatorial regions, as witnessed by an outward-moving entropy jump. By 250\,ms after bounce, the shock reaches a few thousand kilometers, assumes a near spherical shape, and the entire white dwarf material outside the newly-formed neutron star flows nearly radially outward. In the high-rotation (1.92-\mo) progenitor, the white dwarf possesses a lot more mass at near zero-latitude and this confines more drastically the emerging shock along the poles. As the shock migrates outwards, it opens up; it does wrap around the progenitor, but much later than when the shock escapes in the polar directions. Despite the reduced ram pressure associated with centrifugal support, the shock stalls a few hundred milliseconds after bounce along all near-equatorial directions. Within 100-200\,ms, the newly-formed neutron star has a mass of $\sim$1.4\,\mo, similar in both models despite the 0.5\,\mo difference in progenitor mass. Note that the large neutron star asphericity and the sizably lower density at the neutrinosphere for near-zero latitudes makes this mass definition ambiguous at such early times, especially in the 1.92-\mo model. Indeed, the neutron star is not clearly distinguishable from the surrounding equatorial material, so imposing either a density cut of 10$^{10-11}$\,g\,cm$^{-3}$ or a radius cut in defining the newly-born neutron star appears arbitrary when determining the residual mass. In the 1.46-\mo model, about 0.06\,\mo remains outside of the neutron star, mostly in the equatorial disk region; in the 1.92-\mo model, 0.6\,\mo is now lying in this disk-like structure. The rest of the initial mass is outflowing material, which, if selected according to an outward radial velocity discriminant of 10000\,\kms (comparable to the escape velocity at 3000\,km), reaches 4$\times$10$^{-3}$\,\mo for the 1.46-\mo model and 3$\times$10$^{-3}$\,\mo for the 1.92-\mo model. These various components are documented in more detail in \S\S\ref{sect_disk}--\ref{sect_ye}. The late-time evolution of both models is characterized by a strong neutrino-driven wind that sets in about 300\,ms after bounce, replenishing the grid with denser material (on average 10$^4$\,g\,cm$^{-3}$) and large velocities (with a maximum of 30000\,\kms along the poles). The properties of the neutrino-driven wind are very angle-dependent, the density changing by 30\% between the pole and the 40$^{\circ}$ latitude, while the radial outflow velocity varies by a factor of 3 in the 1.92-\mo model. The latitudinal dependence of the mass flux per unit solid angle is therefore dominated by a variation in asymptotic velocity. We will discuss this result in more detail in \S\ref{sect_ener}. By the time we stop the simulations, at 550\,ms and 780\,ms for the 1.46-\mo and 1.92-\mo models, all the ambient medium originally placed around the white dwarf progenitor has been swept away by the neutrino-driven wind, which occupies all the space outside the neutron star and the disk. The electron fraction of the material ejected in the original blast is close to 0.45--0.5, while subsequently, in the neutrino-driven wind, the values are lower, with a pronounced decrease towards lower latitudes (note, however, the high $Y_{\rm e}$ right along the pole; see \S\ref{sect_ye}). In Fig.~\ref{fig_rho_radslice}, we recapitulate for both models the evolution described above by showing equatorial and polar slices of the density as a function of time. Striking features are the distinct polar and equatorial surface radii, the fast infall of the inner regions to nuclear densities, the slow plowing of the shock along the equatorial directions, superseded in radial extent and velocity by the shock in the polar direction as the surface mass shells collapse in, and finally the emergence of a sequence of density kinks associated with the birth of the fast neutrino-driven wind that sweeps away the previously shocked material that did not leave the grid. Similarly, in Fig.~\ref{fig_vr_radslice}, we show slices of the radial velocity, $V_R$, along the polar (top row) and equatorial (bottom row) directions for the 1.46-\mo (left column) and 1.92-\mo (right column) models. Notice the much larger infall velocities, similar along the poles and the equator for the 1.46-\mo model, but with a strong latitudinal dependence in the 1.92-\mo model. In that model, the speed contrast between the polar and equatorial directions is $\sim$30000\,\kms. Overall, the evolution is more rapid along the poles than on the equator, with larger asymptotic velocities (30000\,\kms compared with 10000\,\kms), and with the establishment of a quasi-stationary outflow at late times along the pole. These radial slices offer a means of better interpreting the fluid velocities, depicted with vectors, in most color maps shown in this paper. We also show isodensity contours in most color maps to provide some feeling for the density distribution. Having described the general properties of the two simulations of the AIC of a 1.46-\mo and 1.92-\mo white dwarf, we now address more specific issues, covering the properties of the nascent neutron star (\S\ref{sect_pns}), the neutrino signatures (\S\ref{sect_nu}), the properties of the residual disk and the angular momentum history (\S\ref{sect_disk}), the neutrino-driven wind and the global energetics (\S\ref{sect_ener}), the electron fraction of the ejected material (\S\ref{sect_ye}), and, finally, the gravitational wave signatures (\S\ref{sect_gw}). \section{Neutron Star properties} \label{sect_pns} The white dwarf progenitors discussed in this paper, due to their evolution to high central densities, high rotational kinetic energies, and high mass, are distinctive in that their cores always collapse to form neutron stars rather than being disrupted by the explosive burning of carbon and oxygen. We find that neutron stars formed from the AIC of the progenitor white dwarfs used in this work are very aspherical (see, for example, Walder et al. 2005; Liu \& Lindblom 2001; Janka \& M\"onchmeyer 1989ab), although there are significant differences in evolution after bounce between the two models. In Fig.~\ref{fig_nusphere_46}, we show color maps of the density field 59\,ms after bounce (top row), as well as for the last time computed ($t=570$\,ms after bounce; bottom row) in the 1.46-\mo model. To render more striking the level of asphericity of the neutron star ``surface,'' we overplot the neutrinospheres $R_{\nu}(\varepsilon_{\nu},r,z)$, adopting the definition $$ \tau(R_{\nu}(\varepsilon_{\nu},r,z)) = \int_{R_{\nu}(\varepsilon_{\nu},r,z)}^\infty \kappa_{\nu}(\rho,T,Y_{\rm e}) \rho(r',z') dR' = 2/3\,,$$ where $\kappa_{\nu}(\rho,T,Y_{\rm e})$ is the combined material absorption and scattering opacity to neutrinos, and the integration is carried out along radial rays, with $R' = \sqrt{r'^2 + z'^2}$, using 30 equally spaced latitudinal directions per quadrant. Strictly, this definition is most appropriate for the photosphere/neutrinosphere in a plane-parallel atmosphere, but it gives a sense of the asphericity of the collapsed core. Note that the above expression contains a dependence both on the neutrino flavor and the neutrino energy $\varepsilon_{\nu}$. In the left column of Fig.~\ref{fig_nusphere_46}, line contours correspond to such neutrinosphere radii as a function of energy group, bracketing the peak of the neutrino energy distribution at the neutrinosphere, i.e., between 2.5 and 46\,MeV. Material opacity to neutrinos increases with the square of the energy, so that higher-energy neutrinos have larger neutrinospheres. Here, for the 1.46-\mo model, these radii vary from $\sim$30 to $\sim$120\,km along the equatorial direction, with little departure from sphericity (30\% lower values are obtained in the polar direction). In the right panel, line contours correspond to the neutrinosphere for the three different flavors at a neutrino energy ($\varepsilon_{\nu}$) of 12.5\,MeV, associated with the peak of the energy distribution at infinity. Note that this is approximate, since the neutrino energy distribution hardens with time and manifests a latitudinal dependence (see \S\ref{sect_nu}). As in standard 1D and 2D core-collapse computations, we find that the electron neutrinos decouple from matter at larger radii than the $\bar{\nu}_e$ and ``$\nu_{\mu}$'' neutrinos. Here, the former decouple at 90\,km (60\,km) along the equator (pole), the latter two further in but at a similar radius of 70\,km (50\,km). The neutrinospheres show a similar shape for all three flavors (and all energy groups), reflecting the corresponding asphericity in the density field. In the bottom-row panels, we reproduce the above for the last time in the 1.46-\mo simulation. The departure from sphericity is now considerable, with both an oblateness and a strong pinching of the neutrinospheres along the polar directions. Along the equatorial direction, the radial spacing between neutrinospheres of consecutive and higher energy groups has increased, and the lower (higher) energy groups decouple further in (out) than in the previous snapshot, with neutrinospheres between 20 and 150\,km (from 2.5 to 50\,MeV), 80 and 50\,km (for the $\nu_e$ neutrino, and $\bar{\nu}_e$/``$\nu_{\mu}$'' neutrinos, respectively). Along the polar direction, neutrinospheres of all neutrino energy groups (and all neutrino flavors at 12.5\,MeV) shown reside in a narrow range of radii between 20 and 30\,km (22 to 25\,km). Here again, the neutrinospheres depicted follow very closely the contours of density, which is the primary factor controlling the neutrino optical depth. In Fig.~\ref{fig_nusphere_92}, we duplicate Fig.~\ref{fig_nusphere_46} (note the different spatial scale) for the 1.92-\mo model, showing the same quantities both early after bounce (59\,ms) and significantly later at 775.5\,ms after bounce. There are numerous differences with the 1.46-\mo model. First, the neutrinospheres are aspherical even right after bounce (top row), with equatorial (polar) radii larger (smaller) by a factor of 2-3 compared with those in the 1.46-\mo model. All flavors reveal similar neutrinosphere locations. In the 1.46-\mo model, the later ratio of the equatorial and polar radii is 2.5:1, irrespective of energy group and flavor. This becomes 15:1 in the 1.92-\mo model, and is thus a considerable departure from sphericity; the faster rotating model has a neutrinosphere radius of just 14\,km along the pole, but 215\,km along the equator. This is the most conspicuous difference with the essentially spherical neutron stars seen in non-rotating simulations of the more standard core collapse of massive stars (Keil et al. 1996; Swesty \& Myra 2005; Buras et al. 2005a; Dessart et al. 2005), with neutrinosphere radii of the order of 20-30\,km at comparable times after core bounce. The very aspherical neutron stars formed through the AIC of a white dwarf make the determination of the neutron star mass somewhat ambiguous. Rather than taking the enclosed mass within a given spherical radius, we compute the total mass from all regions above a given mass density. For a density cut of 10$^{10}$\,g\,cm$^{-3}$, we obtain neutron star masses of 1.42\,\mo for the 1.46-\mo model, and 1.5\,\mo for the 1.92-\mo model. However, if we adopt a density cut of 10$^{11}$\,g\,cm$^{-3}$, the neutron star masses are, respectively, 1.39\,\mo and 1.30\,\mo. The higher-mass progenitor model has now a smaller neutron star mass, reflecting the strong asphericity of the density field. While we might associate the neutrinosphere with the neutron star surface and with a standard mass density of 10$^{11}$\,g\,cm$^{-3}$, such an association in the present fast rotating neutron star is inappropriate, since the neutrinospheres extend well into regions where the density is $\sles$10$^{10}$\,g\,cm$^{-3}$. These neutron star mass values are reached at 100\,ms after core bounce and remain essentially constant. The neutrino-driven wind that appears after a few 100\,ms decreases the neutron star mass at a rate of just a few 10$^{-3}$\,\mo\,s$^{-1}$ (see \S\ref{sect_ener}). Note also that the enhanced centrifugal support in the faster-rotating, higher-mass model leads to bounce at a 30\% lower maximum density compared with the 1.46-\mo model, and both a reduced and a delayed mass accretion rate along the equatorial direction compared with what would prevail in the absence of rotation. At the end of each simulation, the total neutron star angular momentum is 1.35$\times$10$^{49}$ erg$\cdot$s (1.13$\times$10$^{49}$ erg$\cdot$s) for the 1.46-\mo model, using the density cut at 10$^{10}$\,g\,cm$^{-3}$ (10$^{11}$\,g\,cm$^{-3}$) , and 4.57$\times$10$^{49}$\,erg$\cdot$s (2.79$\times$10$^{49}$\,erg$\cdot$s) for the 1.92-\mo model. However, the accretion rate at a given Eulerian radius is higher and longer-lived at smaller latitudes, because of the larger amount of mass available and the flatter density profile in those regions. Overall, the presence of a massive accretion disk in the fast rotating model complicates the definition of the neutron star mass at such early times. Evolution over minutes/hours/days will likely lead to significant accretion onto the neutron star, resulting in a much higher final mass. The final rotational to gravitational energy ratio $T/|W|$ is 0.059 for the 1.46-\mo model and 0.262 for the 1.92-\mo model. These values are large and for the latter model, large enough to cause the growth of secular and perhaps even dynamical instabilities (Tassoul 2000). Using realistic post-bounce configurations for a rotating massive star progenitor, Ott et al. (2005a) find a dynamically unstable spiral mode for $T/|W|$ as low as $\sim$0.08. Thus, it is likely that the PNS structures found here, especially for the 1.92-\mo model, would develop some non-axisymmetric instability that would cause, among other things, outward angular momentum transport. These results are significantly different from those of Fryer et al. (1999), who obtained a successful explosion $\sles$100\,ms after bounce, a PNS mass of $\sim$1.2\,\mo, and a $\sim$1\,s period at $\sim$200\,ms after core bounce. Such different conclusions stem from their adoption of slow, solid-body progenitor rotation, with most of the angular momentum stored in the outer mantle and blown away by the explosion, rather than being accreted by the PNS. \section{Neutrino signatures} \label{sect_nu} The first observational signature of an AIC explosion would be the copious emission of neutrinos immediately after core bounce. As discussed above, the properties of the bounce of the core and the nascent neutron star are close enough to those obtained in simulations of the core collapse of massive progenitors that one expects a neutrino signal with a somewhat similar evolution and character (see, for example, the predictions for the 11-\mo model of Woosley \& Weaver 1995 in Dessart et al. 2005). In Fig.~\ref{fig_nuflux}, we show the neutrino luminosities (on a log scale) for the 1.46-\mo (left panel) and 1.92-\mo (right panel) models, with distinct curves for the different neutrino flavors (solid line: $\nu_e$; dashed line: ${\bar{\nu}_e}$; dash-dotted line: ``$\nu_{\mu}$''), as well as different colors for the equatorial (black) and polar (red) directions. These are luminosities in the sense that the flux in each direction is scaled by 4$\pi R^2$ where $R$ is a spherical radius, of 250\,km for the 1.46-\mo model and 400\,km for the 1.92-\mo model (chosen to be well above the neutrinosphere for the energy at the peak of the neutrino distribution). They correspond to the total luminosity that would have obtained had the selected directional flux been the same in all directions. The temporal evolution of the various fluxes for both models is comparable. The total neutrino luminosity reaches a maximum of 5.2$\times$10$^{53}$\,erg\,s$^{-1}$, mostly due to the $\nu_e$ neutrino contribution, and decreases to $\sim$4$\times$10$^{52}$\,erg\,s$^{-1}$ at 500\,ms after bounce in the 1.46-\mo model, with a further 30\% decrease for the 1.92-\mo model. At later times, the main reason for this difference is the much lower ``$\nu_{\mu}$'' neutrino luminosity in the 1.92-\mo model. This reduction has been seen and discussed by Fryer \& Heger (2000) in the context of the collapse of rotating cores of massive progenitors. The smaller core densities (weaker bounce) achieved in models with fast rotation lead to smaller temperatures and, consequently, smaller neutrino emission, with a larger effect for the $\nu_{\mu}$ and $\nu_{\tau}$ neutrinos (grouped under the name ``$\nu_{\mu}$'' here). So, while the globally lower neutron star densities in the fast-rotating model induce a reduction in neutrino luminosity compared to the 1.46-\mo model, the same effect introduces a latitudinal variation of neutrino fluxes in the faster rotating model, with fluxes, irrespective of neutrino flavor, larger by a factor of about two along the pole than along the equator (note that at a radius of 250\,km, the difference is higher and on the order of three). This variation, not discussed by Fryer \& Heger (2000), results from the further variation of the neutrinosphere temperatures with latitude within a given rotating model. In both models, but more so in the 1.92-\mo model, the temperature gradient and the temperatures are reduced at the neutrinosphere (for a given energy group) along the equator compared to the poles, irrespective of energy group and flavor, as is clearly visible in Fig.~\ref{fig_temp}. For example, we see that the temperature on the 10$^{10}$\,g\,cm$^{-3}$ contour is 4\,MeV in the polar direction and 0.5\,MeV in the equatorial direction. Within the neutrinosphere region, we find fluid velocities that are oriented preferentially in the $z$-direction, along cylinders, illustrating the weak or absent convection that results from the stabilizing specific angular momentum profile (Fryer \& Heger 2000; Heger et al. 2000; Ott et al. 2005b; see also \S\ref{sect_disk}). Along a given angular slice, the temperature has a maximum at mid-latitudes, caused by enhanced ($\nu_e$) neutrino energy deposition in this direction. These regions offer a tradeoff, since the flux is still higher at such latitudes than along the equator, while the density is relatively higher than along the poles (see \S\ref{sect_ener}). The latitudinal variations seen in the collapsed models of AIC progenitors are extreme, and, indeed, for the slower rotation rates typically obtained for massive-star core-collapse progenitors (Heger et al. 2000), a modest anisotropy is found instead (Walder et al. 2005). We document further the neutrino signatures of the AIC of white dwarfs by showing in the top row of Fig.~\ref{fig_nufluxspec}, for the 1.46-\mo (left) and 1.92-\mo (right) models, the time-integrated neutrino emission at infinity for each flavor, as a function of neutrino energy. In the bottom three panels of each column, we show the individual neutrino distributions at three representative times of the simulations (at bounce, halfway through the simulation, and at the last simulated time). The overall flux level and hardness of the energy distribution are higher along the polar direction, the variation at a given time towards higher latitude mimicking the time evolution seen for non-rotating core collapse simulations of massive star progenitors. We compute the average neutrino energies, here defined as $$ \sqrt{\langle\varepsilon_{\nu}^2\rangle} \equiv \left[ \frac{\int d\varepsilon_{\nu} \varepsilon_{\nu}^2 F_{\nu}(\varepsilon_{\nu},R)} {\int d\varepsilon_{\nu} F_{\nu}(\varepsilon_{\nu},R)} \right]^{\frac{1}{2}}\,.$$ For the 1.46-\mo model ($R=250$\,km), we obtain similar values to within 2-3\% along the pole and along the equator with $\langle\varepsilon_{\nu_e}\rangle=10$\,MeV, $\langle\varepsilon_{{\bar\nu}_e}\rangle=15$\,MeV, and $\langle\varepsilon_{\nu_{\mu}}\rangle=24$\,MeV. For the 1.92-\mo model ($R=600$\,km), we obtain systematically lower values than for the 1.46-\mo model, and for lower latitudes. Along the equator (pole), we obtain $\langle\varepsilon_{\nu_e}\rangle=9$\,MeV (10\,MeV), $\langle\varepsilon_{{\bar\nu}_e}\rangle=14$\,MeV (16\,MeV), and $\langle\varepsilon_{\nu_{\mu}}\rangle=16$\,MeV (21\,MeV). Note that this definition of the neutrino ``average'' energy is similar to that found in Thompson et al. (2003), who used the mean intensity $J_{\nu}$ in place of the flux $F_{\nu}$. Since we are close to free-streaming regimes at the adopted radii for the peak of the neutrino energy distribution, the two are equivalent. In the 1.46-\mo model, and at late times, we see that the neutrino display is more dramatic (total flux is twice as high) and is characterised by a harder spectrum than in the 1.92-\mo model. This is due to the more compact and, thus, hotter neutrinospheres of the neutron star formed by the lower-mass white dwarf progenitor. \section{Rotation and the remnant disk} \label{sect_disk} Let us now turn our discussion to the angular momentum and angular velocity budget and profiles in our simulations. As discussed in \S\S\ref{sect_results}--\ref{sect_pns}, we find that the early neutron stars have comparable masses in the two simulations, the rest residing not so much in the outflow than in a substantial amount of ``circum-neutron star'' disk material, rotating fast, but having little outflow or inflow velocity (see Fig.~\ref{fig_vr_radslice}). In Fig.~\ref{fig_w_j}, we plot a temporal sequence of the equatorial radial profile of the angular velocity (top row) and specific angular momentum (bottom row), for the 1.46-\mo model (left column) and 1.92-\mo model (right column). In the 1.46-\mo model, the central angular velocity is $\sim$0.1\,rad\,s$^{-1}$ (or a period $P=63$\,s) at the start of the simulation (initial conditions in the YL05 progenitor), and $\sim$1000\,rad\,s$^{-1}$ ($P=6.3$\,ms) at the end the simulation, a spin-up factor of 10000. In the 1.92-\mo model, we start with a much higher angular rotation rate of $\sim$20\,rad\,s$^{-1}$ ($P=0.3$\,s), but the final values are comparable with those of the 1.46-\mo model, being $\sim$2800\,rad\,s$^{-1}$ ($P=2.2$\,ms). Thus, both simulations lead to the formation of a neutron star with a period of a few milliseconds, although we expect the neutron star formed in the 1.92-\mo model to further accrete mass and angular momentum, which may spin-up the residue to even shorter periods. The general angular velocity and specific angular momentum profiles for both models are quite similar. Despite wiggles observed in the 1.46-\mo model in the inner 10 kilometers (which we associate with slight numerical artifacts along the axis - this problem is not present in the 1.92-\mo model, whose grid covers only 90$^{\circ}$), the neutron star is close to solid-body rotation out to 30\,km, showing a steady and smooth decline with radius beyond. In all four panels, we overplot as a broken and black line the corresponding local Keplerian angular velocity, $\Omega_{\rm Keplerian}(r) = \sqrt{GM/r^3}$, where $G$ is the gravitational constant and $M$ is the mass interior to the radius (cylindrical, or spherical). (For these plots, we employ the corresponding neutron star mass.) The angular velocity or specific angular momentum profiles beyond 30\,km graze the corresponding line for the Keplerian value, being always lower by a few tens of percent. In the 1.46-\mo model, the profiles evolve significantly toward this Keplerian limit, angular momentum being gained along the equatorial direction through the radial infall of non-zero latitude material. The constancy of the angular velocity with $z$ allows a significant gain from such infall. In the 1.92-\mo model, the rotational properties along the equator are originally closer to the Keplerian values, but, accordingly, evolve little. In both models, angular momentum is transported outwards, first in material blown away by the shock wave initiated at core bounce, and then in the neutrino-driven wind. This occurs as the outflowing material wraps around the progenitor white dwarf and eventually meets along the equator. There is no ``physical'' viscosity in the code that would permit a proper modeling of the accretion disk. Mass accretion should occur in partnership with outward transport of angular momentum over a longer timescale, yet to be determined. In the 1.92-\mo model, at the end of the simulation, the near-Keplerian disk extends from 30\,km out to $\sim$1800\,km, covering a range of densities (temperatures) from 10$^{13}$\,g\,cm$^{-3}$ down to 10$^{8}$\,g\,cm$^{-3}$ (3\,MeV down to 0.1\,MeV; Fig.~\ref{fig_rho_radslice}). \section{Energetics and the neutrino-driven wind} \label{sect_ener} Given that it leads to the formation of a $\sim$1.4-\mo neutron star, the AIC of a 1.4--2.0\,\mo white dwarf is expected to result, in the case of a successful explosion, to an outflow of modest mass. Furthermore, due to the similarity with the core collapse of massive star progenitors, their explosion kinetic energy should be lower than the $\sim$10$^{51}$ erg inferred, e.g., for SN1987A (Arnett 1987). As discussed in \S\ref{sect_results} and \S\ref{sect_disk}, the total sum of the neutron star and disk masses is very close to the original progenitor mass, leaving typically a few 0.001\,\mo for the ejected material. Integrating all the mass that has left the grid over the course of the simulation, as well as all the material outflowing with a positive radial velocity greater than 10000\,\kms (which is of the order of the escape velocity at a radius of 3000\,km) we find a value of 4$\times$10$^{-3}$\,\mo for the 1.46-\mo model and 3$\times$10$^{-3}$\,\mo for the 1.92-\mo model. In Fig.~\ref{fig_energy}, we show the evolution of the corresponding gravitational (blue), thermal (cyan), and kinetic (2D planar: red; rotational: green) energies for this outflowing material as solid lines, including the total energy as a black dotted line. At the last computed time, the total energy is indeed lower than that inferred for standard core collapse. Adopting a radial velocity cut of 10000\,\kms, we find an energy at the last simulated time of 2.7$\times$10$^{49}$erg for the 1.46-\mo model and 2$\times$10$^{49}$erg for the 1.92-\mo model. Note that energy is still being pumped into the wind by the slowly decaying neutrino luminosity emanating from the neutron star; the trend of the total energy curve suggests that the total energy of the explosion will be 2-3 times higher, thus $\sim$10$^{50}$erg for the 1.46-\mo model and $\sim$5$\times$10$^{49}$erg for the 1.92-\mo model. This is over one order of magnitude smaller than the explosion energy inferred for normal core-collapse supernovae. The AIC of white dwarfs is likely to lead generically to underenergetic explosions because there is too little mass to absorb neutrinos, most of it being quickly accreted while the rest is centrifugally-supported at large radii, far beyond the region where there is a positive net gain of electron-neutrino energy. Interestingly, similar underenergetic explosions are obtained by Kitaura et al. (2005) and Buras et al. (2005b) for initial main sequence stars of 8.8\,\mo (Nomoto 1984, 1987) and 11.2\,\mo (Woosley, Heger, \& Weaver 2002). Echoing the properties of AIC progenitors, the low envelope mass and fast declining density (and, therefore, accretion rate) are key beneficial components for the success of neutrino-driven explosions, but the same properties are also why the explosion is necessarily underenergetic. The various curves also show a few dips and bumps. The first bump, most pronounced in the 1.92-\mo model, is associated with an early outflow that eventually fell back to smaller radii, while subsequent bumps are caused by episodic mass loading of the neutrino-driven wind, which sets in $\sim$200\,ms after core bounce and drives a 8$\times$10$^{-3}$\,\mo\,s$^{-1}$ (5$\times$10$^{-3}$\,\mo\,s$^{-1}$) mass loss rate in the 1.46-\mo (1.92-\mo; see also Fig.~\ref{fig_mdot}, top panel) model. The higher-mass flux in the 1.46-\mo model results from the higher neutrino luminosity, higher mean neutrino energies, and bigger opening angle of escape for the neutrino-driven wind. Interestingly, this mass flux is strongly angle-dependent, varying by a factor of a few between the pole and the angle that grazes the pole-facing side of the disk. As described in \S\ref{sect_results}, the dynamical effects of the neutrino-driven wind are to entrain the material lying along this interface, tearing the disk via Kelvin-Helmholtz shear instabilities and mass-loading the wind along the corresponding latitudes. Further in, neither this wind nor the neutrinos have an appreciable dynamical impact in driving the disk material outwards, a feature only excacerbated by the reduced neutrino flux at low latitudes. In Fig.~\ref{fig_mdot}, we show in the bottom panel the latitudinal variation of the asymptotic velocity (solid line) and the density (dotted line). Both show an overall decrease towards lower latitudes by a factor of three. The dip in density and higher values of the velocity along the pole to around 70$^{\circ}$ latitude are possibly due to wind mass loading, in combination with centrifugal support at the neutrinosphere for off-polar latitudes. In Fig.~\ref{fig_nuflux_2d}, we show at early times (top row) and at the last simulated time (bottom row) for the 1.46-\mo model (left column) and the 1.92-\mo model (right column) color maps of the total neutrino flux in the radial direction, with isodensity contours overplotted as white curves, and velocity vectors as black arrows. First, due to the history of the collapse, the neutron star is relatively devoid of overlying material in the polar direction, while for the higher-mass progenitor a massive ($\sim$0.6\,\mo), dense (10$^{6-10}$\,g\,cm$^{-3}$ ), near-Keplerian disk obstructs the neutron star at latitudes $\sles\pm$40$^{\circ}$. Given this configuration, conditioned essentially by the mass distribution of the progenitor white dwarf, the dynamical effect of a spherically-symmetric neutrino flux would be enhanced along the ``excavated'' polar direction. Indeed, we see a strong neutrino-driven wind in the polar direction that does not exist in directions within $\sim\pm$40$^{\circ}$ of the equator. However, even in the absence of this anisotropic matter distribution, Fig.~\ref{fig_nuflux_2d} reveals the strong latitudinal variation of the neutrino flux at a given Eulerian radius, a variation that is established independently of the configuration of the circum-neutron star disk material. What controls the flux geometry is the combination of two effects. First, the exceptional elongation of the neutrinospheres along the equatorial direction leads to a decoupling radius (surface) about 10 (100) times bigger for a polar observer than for an equatorial observer. The angle-dependent decoupling radius of neutrinos mitigates this result (Walder et al. 2005), but, as shown in Fig.~\ref{fig_nuflux_2d}, the latitudinal variation along different directions persists in the total neutrino flux. Similarly, Fig.~\ref{fig_nu_flux_vec} shows the anisotropy of the $\nu_e$ neutrino flux, rendered by the corresponding flux vectors. Notice how the base of the flux vectors in the high-density central regions is perpendicular to the local isodensity (or, equivalently, equipotential) contour. Second, as shown in Fig.~\ref{fig_temp}, the temperature and its radial-gradient along a given isodensity contour are both significantly lower along the equatorial direction, leading to reduced diffusive fluxes. These properties are reminiscent of the effect of gravity darkening (von Zeipel 1924) in fast rotating (non-compact) stars and the associated scaling of the radiative flux with the local effective gravity (see Owocki et al. 1996), although this may be the first time it is reported in the context of a protoneutron star (but see Walder et al. 2005). With such a polar-enhanced wind, the angular momentum loss rate is reduced, with consequences for the spin evolution of the PNS. \section{Ejecta composition} \label{sect_ye} In Fig.~\ref{fig_ye_dist}, we show for the two baseline models the electron fraction ($Y_{\rm e}$) distribution of the material in the ejecta (material outside the neutron star moving outwards with a radial velocity greater than 10000\,\kms), accounting as well for the mass loss through the outer grid radius. Such cumulative outflow amounts to 4$\times$10$^{-3}$\,\mo for the 1.46-\mo model and 3$\times$10$^{-3}$\,\mo for the 1.92-\mo model. We obtain double-peak profiles, the first blast propelling symmetric material ($Y_{\rm e}=0.5$), subsequently followed after 200\,ms by progressively neutron-rich material, i.e., $Y_{\rm e} =0.25-0.35$, in the neutrino-driven wind. Note that for these runs we enforced an upper limit of 0.5 to the computed $Y_{\rm e}$ values. Fryer et al. (1999) obtained an ejecta mass in the vicinity of 0.2\,\mo, two orders of magnitude larger than our values. Because our ejecta masses are much smaller, we find that the mass loss rates and the kinetic energies associated with the neutrino-driven wind are relatively more important for the global energetics of the AIC of white dwarfs. At late times, the asymptotic electron fraction $Y_{\rm e}^{\rm a}$ of the neutrino-driven wind varies with latitude (despite the smooth variation of other quantities at correspondingly larger distances). Material ejected within 20$^{\circ}$ of the pole has an electron fraction of $\sim$0.5, while towards the equator, this electron fraction decreases to 0.3, rising again to near 0.5 values in regions belonging to the disk (see bottom row panels in Fig.~\ref{fig_seqye}). The $Y_{\rm e}^{\rm a}$ values seen in our simulations are in fact already set when wind material leaves the vicinity of the neutrinosphere, whose properties depend on the particle trajectory under scrutiny. We know from previous studies (Qian \& Woosley 1996; Wheeler et al. 1998; Thompson et al. 2001; Pruet et al. 2005; Fr\"{o}lich et al. 2005) that the asymptotic electron fraction of the ejecta is controled by competing factors. The electron and anti-electron neutrino luminosities, modulated by the hardness of their respective energy distributions, influence the electron flavor production rates via the reactions $\nu_e$n$\rightarrow$pe$^{-}$ and $\bar{\nu}_e$p$\rightarrow$e$^+$n and, thereby, the neutron-richness of the ejecta. The expansion timescale sets the duration over which interactions between neutrinos and nucleons can take place. The starting value of the electron fraction, i.e., at the base of the outflow, is altered by the above factors and differs from the asymptotic value seen. In Fig.~\ref{fig_ye_gain}, we show a color map of the electron fraction in the inner 200\,km, highlighting the butterfly shape of the deleptonized region in cross section, in stark contrast with the corresponding near-spherical shape seen in core-collapse simulations of both rotating and non-rotating progenitors (Keil et al. 1996; Walder et al. 2005; Dessart et al. 2005). Deleptonization obtains preferentially in the vicinity of the dumbbell-shaped neutrinosphere, and stretches outwards for off-polar latitudes. Along the equator, deleptonization ceases at smaller radii due to the lower effective temperatures (tied to the neutrino fluxes). Temperature and neutrino flux are in fact intertwined, since energy deposition by neutrinos may raise the temperature locally in the so-called gain region. This is also vividly represented in Fig.~\ref{fig_ye_gain} (right) by the net gain associated with electron-neutrino energy deposition in this inner region, which also shows the same butterfly shape. Electron neutrinos emerge from the neutron star, and, due to the dumbbell neutrinosphere morphology, at much smaller radii along the poles than for off-polar latitudes. The decreasing neutrino flux (dilution) reduces this energy deposition beyond $\sim$50\,km, and even at smaller radii along the equator due to the additonal flux reduction there (Fig.~\ref{fig_nuflux_2d}). The asymptotic value of the material electron fraction is determined in the vicinity of the neutrinosphere and, therefore, is directly influenced by this configuration of the inner $Y_{\rm e}$ distribution. Along the pole, the wind carries initially low $Y_{\rm e}$ material that absorbs electron neutrinos, whose associated luminosity is one magnitude higher than that of the anti-electron neutrinos (see Fig.~\ref{fig_nuflux}), raising the electron fraction to the ceiling value of 0.5 artificially adopted in these calculations. Away from the pole, the neutrinosphere is located further out, and despite similar neutrinosphere $Y_{\rm e}$ values, the larger distance from the neutron star implies a reduced electron-neutrino luminosity and a reduced absorption of neutrinos, leading to asymptotic values of the electron fraction of only $\sim$0.25, not far from the values at the corresponding neutrinosphere. To summarize, the progressive decrease of the electron fraction (and of the entropy) away from the pole is a result of the reduced electron-neutrino luminosities in the vicinity of the latitudinal-dependent neutrinosphere radius (and the associated reduced heating and electron-capture rates). \section{Gravitational wave signature} \label{sect_gw} We estimate the gravitational wave emission from aspherical mass motions in our models via the Newtonian quadrupole formalism as described in M\"onchmeier et al. (1991). In addition, we compute the gravitational wave strain from anisotropic neutrino emission employing the formalism introduced by Epstein (1978) and developed by Burrows \& Hayes (1996) and M\"uller \& Janka (1997). Due to rapid rotation and the resulting oblateness of the core, one would expect that rotating AIC models would have significant gravitational wave signatures. However, though the contribution to the metric strain in the equatorial plane, $h_+$, of the aspherical and dynamical matter distributions is not small, we find that that of the aspherical neutrino field is larger in magnitude, though at much smaller frequencies. While we calculate that $h_+$(max) for the matter in the 1.46-\mo\ model is $\sim$5.9$\times$10$^{-22}$, with a spectrum that peaks at $\sim$430 Hz, the corresponding $h_+$(max) due to neutrinos is $\sim$4.6$\times$10$^{-21}$ (derived from the fluxes at 200 km), but at frequencies between ($\sim$0.1-10 Hz). The total energy radiated in gravitational waves is $\sim$5.7$\times$10$^{-10}$ \mo{$c^2$}. The corresponding numbers for the faster rotating and more massive 1.92-\mo\ model are $\sim$3.6$\times$10$^{-21}$ (matter), $\sim$165 Hz, $\sim$2.0$\times$10$^{-20}$ (neutrinos at 300 km), and $\sim$7.0$\times$10$^{-8}$ \mo{$c^2$}. Note that almost all of the energy is being emitted by mass motions (99.8\% in the 1.46-\mo\ model and 98.4\% in the 1.92-\mo\ model), since the power scales with the time derivative of the wave strain $h_+$, which is small for the waves emitted from the aspherical neutrino field. We compare the above numbers with those obtained by Fryer, Holz \& Hughes (2002) for an AIC model of Fryer et al.\ (1999) which was setup with a simple, solid-body rotation law and had a final T/$|$W$|$ of $\sim$0.06 (for our 1.46-\mo\ model: 0.059; see Table 1). They did not consider anisotropic neutrino emission. Our more realistic initial models yield maximum (matter) gravitational wave strains that are 1.5 to 2 orders of magnitude smaller than those predicted by Fryer, Holz \& Hughes (2002). The total energy emissions match within a factor of two since our models emit at higher frequencies. Based on our results, we surmise that gravitational waves from axisymmetric AIC events may be detected by current LIGO-class detectors if occurring anywhere in the Milky Way, but not out to megaparsec distances as suggested by Fryer, Holz \& Hughes (2002). It is, however, likely (\S 5) that at least the 1.92-\mo\ model will undergo a dynamical rotational instability leading to non-axisymmetric deformation (which can not be captured by our 2D approach) and, hence, to copious gravitational wave emission over many rotation periods, greatly enhancing detectability. \section{Discussion and conclusions} \label{sect_conc} We have presented a radiation/hydrodynamic study with the code VULCAN/2D of the collapse and post-bounce evolution of massive rotating high-central-density white dwarfs, starting from physically consistent 2D rotational equilibrium configurations (Yoon \& Langer 2005). The main results of this study are: \begin{itemize} \item The AIC of white dwarfs leads to a successful explosion with modest energy $\sles$10$^{50}$\,erg, thus comparable to the energies obtained through the collapse of O/Ne/Mg core of stars with $\sim$8-11\,\mo main sequence mass (Kitaura et al. 2005; Buras et al. 2005b). This is, however, underenergetic, by a factor of about ten, compared with the inferred value for the core collapse of more massive progenitors leading to Type II Plateau supernovae. Although less and less likely to be the engine behind most core-collapse supernova explosions, the neutrino mechanism can successfully power explosions of low-mass progenitors and AICs due to the limited mantle mass and steeply declining accretion rate. \item Due to high-mass and angular-momentum accretion, white dwarf progenitors leading to AIC can have masses of up to $\sim$2\,\mo and rotate fast, with rotational to gravitational energy ratios of up to a few percent prior to collapse. The asphericity of such white dwarfs allows the shock generated at core bounce to escape along the poles in just a few tens of milliseconds, opening a hole in the white dwarf along the poles. The blast expands and wraps around the progenitor and escapes the grid ($\sim$5000 km) within a few hundred milliseconds, at which time a neutrino-driven wind has grown in the pole-excavated region of the white dwarf. Both the original blast and the wind show strong latitudinal variations, partly constrained by the obstructing uncollapsed equatorial disk regions of the progenitor, whose centrifugal support prevents it from collapsing on a dynamical timescale. Rotation in such progenitors, thus, affects both directly and indirectly the morphology of the explosion. \item The neutron stars formed have masses on the order of 1.4\,\mo, with rotation periods close to a millisecond in the rigidly rotating inner $\sim$30\,km. The final rotational to gravitational energy ratios, for our two test cases, cover 0.06 to 0.26, the latter being large enough to grow non-axisymmetric instabilities. At the end of our simulations, the neutron stars are oblate and pinched along the poles, with polar and equatorial radii in the ratio 1:15 for the faster rotating (1.92-\mo) model. \item The morphology of the neutron star leads to a latitudinal variation of the neutrino flux, the net energy gain, and the temperature. In the faster rotating model, the ``$\nu_\mu$'' neutrino flux is reduced, while the anti-electron neutrino flux is a factor of ten lower than that of the electron-neutrino. This raises the electron fraction of the ejected material to values close to 0.5 along the poles, but to only $\sim$0.25 at lower latitudes, since the corresponding neutrinosphere is more remote and, thus, the electron-neutrino flux is smaller. This introduces a latitudinal dependence of the electron fraction of the ejected material, but more importantly allows neutron-rich material, with entropy on the order of 20-40\,$k_{\rm B}$/baryon, to escape. Thus, a low-entropy r-process might take place under these conditions. \item The high original angular momentum of the progenitor follows the mass and is, thus, found mostly in the neutron star at the end of the simulation. However, rotational energy is also given to the ejecta, which uses it to gain (planar) radial kinetic energy to escape the potential well, while the rest is found in a quasi-Keplerian disk of up to 0.5\,\mo in the 1.92-\mo model. This disk is an essential component of these AICs; it collimates the explosion and the neutrino-driven wind and also suggests a second stage of long-term accretion onto the compact remnant. \item The total ejected mass is only of the order of a few times 0.001\,\mo, with only a quarter in the form of $^{56}$Ni. The original blast and the short-lived neutrino-driven wind will lead to a considerable brightening of the object, but the small ejecta mass will quickly become optically-thin, gamma rays leaking out rather than depositing their energy to power a durable light curve. Therefore, these explosions should be underluminous and very short lived. Their appearance may also vary considerably with viewing angle, depending on the mass of the progenitor and the presence of a sizable disk in the equatorial regions. \end{itemize} This study has shown that more consistent, rotating 2D models alter considerably our understanding of accretion-induced collapse, previously obtained under the simplifying assumptions of spherical symmetry and/or zero rotation. Further improvements will come by including a consistent temperature structure for the progenitor white dwarf and by accounting for the effects of magnetic fields. Due to the rapid differential rotation, magnetic fields could be amplified considerably and result in MHD jets that might alter yet again our overall picture of accretion-induced collapse and the energetics of the phenomenon. Three-dimensional effects may also alter the protoneutron star properties presented here, since the faster rotating (1.92-\mo) model is expected to experience non-axisymmetric instabilities. \acknowledgments We acknowledge discussions with and help from Jeremiah Murphy and Casey Meakin. Importantly, we acknowledge support for this work from the Scientific Discovery through Advanced Computing (SciDAC) program of the DOE, grant number DE-FC02-01ER41184 and from the NSF under grant number AST-0504947. E.L. thanks the Israel Science Foundation for support under grant \# 805/04, and C.D.O. thanks the Albert-Einstein-Institut for providing CPU time on their Peyote Linux cluster. We thank Jeff Fookson and Neal Lauver of the Steward Computer Support Group for their invaluable help with the local Beowulf cluster. This research used resources of the National Energy Research Scientific Computing Center, which is supported by the Office of Science of the U.S. Department of Energy under Contract No. DE-AC03-76SF00098.
Title: F_D-Term Hybrid Inflation with Electroweak-Scale Lepton Number Violation
Abstract: We study F-term hybrid inflation in a novel supersymmetric extension of the SM with a subdominant Fayet-Iliopoulos D-term. We call this particular form of inflation, in short, F_D-term hybrid inflation. The proposed model ties the mu-parameter of the MSSM to an SO(3)-symmetric Majorana mass m_N, through the vacuum expectation value of the inflaton field. The late decays of the ultraheavy particles associated with the extra U(1) gauge group, which are abundantly produced during the preheating epoch, could lower the reheat temperature even up to 1 TeV, thereby avoiding the gravitino overproduction problem. The baryon asymmetry in the Universe can be explained by thermal electroweak-scale resonant leptogenesis, in a way independent of any pre-existing lepton- or baryon-number abundance. Further cosmological and particle-physics implications of the F_D-term hybrid model are briefly discussed.
https://export.arxiv.org/pdf/hep-ph/0601080
\begin{flushright} CERN-PH-TH/2006-003\\[-2pt] {\tt hep-ph/0601080}\\ January 2006 \end{flushright} \bigskip \begin{center} {\LARGE {\bf {\boldmath $F_D$}-Term Hybrid Inflation with}}\\[0.3cm] {\LARGE {\bf Electroweak-Scale Lepton Number Violation}}\\[1.5cm] {\large Bj\"orn Garbrecht$^{\, a}$ and Apostolos Pilaftsis$^{\, a,b}$}\\[0.5cm] {\em $^a$School of Physics and Astronomy, University of Manchester,}\\ {\em Manchester M13 9PL, United Kingdom}\\[0.3cm] {\em $^b$CERN, Physics Department, Theory Division, CH-1211 Geneva 23, Switzerland} \end{center} \vspace{1.5cm} \centerline{\bf ABSTRACT} \noindent We study $F$-term hybrid inflation in a novel supersymmetric extension of the SM with a subdominant Fayet--Iliopoulos $D$-term. We call this particular form of inflation, in short, $F_D$-term hybrid inflation. The proposed model ties the $\mu$-parameter of the MSSM to an SO(3)-symmetric Majorana mass $m_N$, through the vacuum expectation value of the inflaton field. The late decays of the ultraheavy particles associated with the extra U(1) gauge group, which are abundantly produced during the preheating epoch, could lower the reheat temperature even up to 1~TeV, thereby avoiding the gravitino overproduction problem. The baryon asymmetry in the Universe can be explained by thermal electroweak-scale resonant leptogenesis, in a way independent of any pre-existing lepton- or baryon-number abundance. Further cosmological and particle-physics implications of the $F_D$-term hybrid model are briefly discussed. \noindent \medskip \noindent {\small PACS numbers: 98.80.Cq, 12.60.Jv, 11.30Pb} \newpage \setcounter{equation}{0} \section{Introduction} The inflationary paradigm constitutes an ingenious theoretical framework, in which many of the outstanding problems in standard cosmology can be successfully addressed~\cite{review}. The recent WMAP data~\cite{WMAP}, compiled with other astronomical observations~\cite{MT,Lyman}, improved upon the precision of about a dozen of cosmological parameters. These include the power spectrum $P^{1/2}_{{\cal R}}$ of curvature perturbations, the spectral index $n_s$, the baryon-to-photon ratio of number densities $\eta_B$ and others. The values of these cosmological observables put severe constraints on the model-building of successful models of inflation and their theoretical parameters. For instance, one of the basic requirements for slow-roll inflation is that the so-called inflaton potential be flat. In this respect, supersymmetry (SUSY) emerges as a compelling ingredient in model-building for protecting the flatness of the inflaton potential against quantum corrections. In addition to the aforementioned element of naturalness, inflationary models would have a greater value if they were predictive and testable as well. One such predictive and perhaps most appealing scenario is the well-celebrated model of hybrid inflation~\cite{Linde}. In this model, the inflaton field $\phi$ can start its slow-roll from values well below the Planck scale $m_{\rm Pl} = 2.4\times 10^{18}$~GeV. This renders the model very predictive, in the sense that an infinite set of possible higher-dimensional non-renormalizable operators, being suppressed by inverse powers of $1/m_{\rm Pl}$, will not generically contribute significantly to cosmological observables, such as $P^{1/2}_{{\cal R}}$ and $n_s$. In the hybrid model, inflation ends through the so-called waterfall mechanism, once the field $\phi$ passes below a critical value $\phi_c$. When this happens, another field $X$ different from $\phi$, with vanishing initial value, develops a tachyonic instability and rolls fast down to its true vacuum expectation value~(VEV). Super-\linebreak symmetric realizations of hybrid inflation from $F$-terms were first analyzed in~\cite{CLLSW,DSS}, whereas hybrid inflation triggered by a dominant Fayet--Iliopoulos~(FI) $D$-term~\cite{FI} was subsequently considered in~\cite{Halyo}. In this paper we study $F$-term hybrid inflation in a novel supersymmetric extension of the Standard Model~(SM) that includes a subdominant FI $D$-term. We call this scenario for brevity, the $F_D$-term hybrid model. To account for the low-energy neutrino data, we introduce 3 singlet neutrino superfields $\widehat{N}_{1,2,3}$ that contain 3 right-handed neutrinos $\nu_{1,2,3\,R}$ and their supersymmetric scalar counterparts $\widetilde{N}_{1,2,3}$. Most importantly, the model ties the $\mu$-parameter of the Minimal Supersymmetric Standard Model~(MSSM) to an SO(3) symmetric Majorana mass $m_N$, through the VEV of the inflaton field $\phi$~\cite{PU2,Francesca}. Hence, the $F_D$-term hybrid model naturally predicts lepton-number violation at the TeV or even at the electroweak scale. In this scenario, the non-zero baryon asymmetry in the Universe (BAU), $\eta_B \approx 6.1 \times 10^{-10}$, can be explained by leptogenesis~\cite{FY,BAUpapers} and specifically by thermal electroweak-scale resonant leptogenesis~\cite{APRD,PU2}. In this paper we also study the constraints on the parameters of the $F_D$-term hybrid model that result from a reheat temperature $T_{\rm reh} \stackrel{<}{{}_\sim} 10^9$~GeV, which is necessary to avoid the well-known gravitino overproduction problem. This consideration puts severe limits on the size of the superpotential couplings of the theory, forcing them {\em all} to acquire rather suppressed values, namely to be smaller than about $10^{-5}$~\cite{SS}. To overcome this problem of unnaturalness, the presence of a subdominant FI $D$-term in the $F_D$-term hybrid model is very crucial and provides a new mechanism of relaxing dramatically the above upper limit. More explicitly, the size of the $D$-term controls the decay rates of the ultraheavy fermions and bosons associated with the extra gauge group U(1)$_X$. In the absence of the $D$-term and any other non-renormalizable interaction, these ultraheavy gauge-sector particles are absolutely stable. On the other hand, these particles are abundantly produced during the preheating epoch, thus dominating the energy density of the Universe shortly after the period of the first reheating caused by the perturbative inflaton decays. Therefore, their late decays induced by a non-vanishing $D$-term could give rise to a second reheating phase in the evolution of the early Universe. Depending on the actual size of the FI $D$-term, this second reheat temperature may be as low as 1~TeV, giving rise to an enormous entropy release that can dilute the gravitinos produced during the first reheating to an unobservable level. The paper is organized as follows: Section~\ref{FDmodel} presents the model-building aspects of the $F_D$-term hybrid model with electroweak-scale lepton-number violation. Technical details related to the radiatively-induced FI $D$-term are relegated to Appendix~A. In Section~\ref{reheat}, we estimate the reheat temperature from the perturbative inflaton decays and derive the resulting gravitino constraint on the theoretical parameters. We then discuss the non-perturbative production of the supermassive fields associated with the U(1)$_X$ gauge group during the preheating epoch and how their late decays can help to lower the reheat temperature even up to~1~TeV. Section~\ref{inflation} is devoted to inflation. Here we investigate two regimes: (i) cold hybrid inflation, where dissipative effects can be ignored, and (ii) warm hybrid inflation, where dissipative effects dominate over the expansion rate of the Universe. In Section~\ref{BAU} we illustrate how the $F_D$-term hybrid model can realize thermal electroweak-scale resonant leptogenesis. Finally, Section~\ref{conclusions} summarizes our conclusions, including a brief discussion of further possible implications of the $F_D$-term hybrid model for particle physics and cosmology. \setcounter{equation}{0} \section{The {\boldmath $F_D$}--Term Hybrid Model}\label{FDmodel} The $F_D$-term hybrid model may be defined by the superpotential \begin{eqnarray} \label{Wmodel} W & =& \kappa\, \widehat{S}\, \Big( \widehat{X}_1 \widehat{X}_2\: -\: M^2\Big)\ +\ \lambda\, \widehat{S} \widehat{H}_u \widehat{H}_d\ +\ \frac{\rho}{2}\, \widehat{S}\, \widehat{N}_i \widehat{N}_i\ +\ h^{\nu}_{ij} \widehat{L}_i \widehat{H}_u \widehat{N}_j\nonumber\\ &&+\ W_{\rm MSSM}^{(\mu = 0)}\; , \end{eqnarray} where $W_{\rm MSSM}^{(\mu = 0)}$ is the MSSM superpotential without the $\mu$ term: \begin{equation} W_{\rm MSSM}^{(\mu = 0)}\ =\ h^u_{ij}\,\widehat{Q}_i\widehat{H}_u\widehat{U}_j\: +\: h^d_{ij}\,\widehat{H}_d\widehat{Q}_i\widehat{D}_j\: +\: h_l\, \widehat{H}_d\widehat{L}_l\widehat{E}_l \; . \end{equation} In~(\ref{Wmodel}), $\widehat{S}$ is the SM-singlet inflaton superfield, and $\widehat{X}_1$ and $\widehat{X}_2$ is a chiral multiplet pair of the so-called waterfall fields which have opposite charges under the additional U(1)$_X$ gauge group. The superpotential~(\ref{Wmodel}) of the model is uniquely determined by the $R$ transformation: $\widehat{S} \to e^{i\alpha}\, \widehat{S}$, $\widehat{X}_{1,2} \to e^{\pm i\beta}\,\widehat{X}_{1,2}$, $\widehat{L} \to e^{i\alpha}\, \widehat{L}$, $\widehat{Q} \to e^{i\alpha}\, \widehat{Q}$, with $W \to e^{i\alpha} W$, whereas all other fields remain invariant under an $R$ transformation. As a consequence of the $R$ symmetry, higher-dimensional operators of the form $\widehat{X}_1 \widehat{X}_2 \widehat{N}_i \widehat{N}_i/m_{\rm Pl}$, for example, are not allowed. In addition, the model contains a subdominant FI $D$-term, $-\frac{1}{2} g\, m^2_{\rm FI}\, D$, giving rise to the $D$-term potential \begin{equation} \label{Dterm} V_D\ =\ \frac{g^2}{8}\ \Big( |X_1|^2\, -\, |X_2|^2\, -\, m^2_{\rm FI}\,\Big)^2\; . \end{equation} The FI $D$-term plays no role in the inflationary dynamics, as long as $g m_{\rm FI} \ll \kappa\, M$. In Appendix~\ref{Dappendix}, we show how a subdominant $D$-term can be generated radiatively after Planck-scale heavy degrees of freedom have been integrated out. The presence of the $D$-term is important to break an accidental discrete charge symmetry that survives after the spontaneous symmetry breaking (SSB) of the U(1)$_X$. Such a breaking is crucial to render all U(1)$_X$ gauge-sector particles unstable. As we will see in Section~\ref{reheat}, an upper limit on the size of the FI term is obtained by requiring a sufficiently low reheat temperature, e.g.~$T_{\rm reh} \stackrel{<}{{}_\sim} 10^9$~GeV, in order to suppress the gravitino abundance to an unobservable level. From~(\ref{Wmodel}) it is straightforward to read the Lagrangian of the inflationary soft SUSY-breaking sector, \begin{equation} \label{Lsoft} -\, {\cal L}_{\rm soft}\ =\ M^2_S S^*S\: +\: \Big( \kappa A_\kappa\, S X_1X_2\: +\: \lambda A_\lambda S H_u H_d\: \: +\: \frac{\rho}{2}\, A_\rho\, S \widetilde{N}_i\widetilde{N}_i\: -\: \kappa a_S M^2 S \: \ +\ {\rm H.c.}\,\Big)\,, \end{equation} where $M_S$, $A_{\kappa,\lambda,\rho}$ and $a_S$ are soft SUSY-breaking mass parameters of order $M_{\rm SUSY} \sim 1$~TeV. In the regime $|S| \gg M$ relevant to inflation, the dominant tree-level and one-loop contributions to the renormalized effective potential may be described by \begin{eqnarray} \label{VpotFD} V_{\rm inflation} & \approx & \kappa^2 M^4\ \Bigg[\, 1\: +\: \frac{1}{64\pi^2}\, \Big(\, 4\kappa^2 \: +\: 8\lambda^2\: +\: 6\rho^2\,\Big)\, \ln\Bigg(\frac{|S|^2}{M^2}\Bigg)\,\Bigg]\nonumber\\ && -\: \Big( \kappa a_S M^2 S\: +\: {\rm H.c.}\Big)\ +\ V_{\rm SUGRA}\ , \end{eqnarray} where $V_{\rm SUGRA}$ denotes the supergravity (SUGRA) correction that results from the K\"ahler potential. Assuming a minimal K\"ahler potential, the SUGRA correction of interest to us takes on the simple form~\cite{CLLSW,CP,LR} \begin{equation} \label{Vsugra} V_{\rm SUGRA}\ =\ \kappa^2 M^4\, \frac{|S|^4}{2\,m^4_{\rm Pl}}\ , \end{equation} where $m_{\rm Pl} = 2.4\times 10^{18}$~GeV is the reduced Planck mass. Possible one-loop contributions to $V_{\rm inflation}$ from $A_{\kappa,\lambda,\rho}$-terms become significant only for relatively low values of $M$, e.g. $M\stackrel{<}{{}_\sim} 10^8$~GeV, for $\kappa,\lambda,\rho \sim 1$, and may therefore be neglected. At the tree level, however, only the tadpole term $\kappa a_S M^2\, S$ may become relevant for values of $\kappa \stackrel{<}{{}_\sim} 10^{-4}$, whereas the other soft SUSY-breaking terms are negligible during inflation~\cite{SS}. We now investigate the stability of the inflationary trajectory in the presence of the Higgs doublets $H_{u,d}$ and the right-handed scalar neutrinos $\widetilde{N}_{1,2,3}$. Specifically, the initial condition for inflation is \begin{equation} \label{initial} {\rm Re}\, S^{\rm in}\ =\ |S^{\rm in}|\ \gg\ M\,,\qquad X^{\rm in}_{1,2}\ =\ 0\,,\qquad H^{\rm in}_{u,d}\ =\ 0\,,\qquad \widetilde{N}^{\rm in}_{1,2,3}\ =\ 0\; . \end{equation} At the end of inflation, one should ensure that the waterfall fields acquire a high VEV, i.e. $X^{\rm end}_{1,2}\ =\ M$, while all other fields have small VEVs, possibly of the electroweak order. To achieve this, we have to require that the Higgs-doublet and the sneutrino mass matrices stay positive definite throughout the inflationary trajectory up to the critical value $|S_c| \approx M$, whereas the corresponding mass matrix of $X_{1,2}$ will be the first to develop a negative eigenvalue and tachyonic instability close to $|S_c|$. In this way, it will be the fields $X_{1,2}$ which will first start moving away from 0 and set in to the `good' vacuum $X^{\rm end}_1\ =\ X^{\rm end}_2\ =\ M$, instead of having the other fields, e.g.~$H_{1,2}$ and $\widetilde{N}^{\rm in}_{1,2,3}$, go to a `bad' vacuum where $X^{\rm end}_{1,2}\ =\ 0$, $H^{\rm end}_{1,2}\ =\ \frac{\kappa}{\lambda}\, M$ and $\widetilde{N}^{\rm in}_{1,2,3} = \frac{\kappa}{\rho}\, M$. To see this, let us write down the mass matrix in the background Higgs-doublet field space $(H^\dagger_d\,,\ H_u )$ as \begin{equation} \label{Mdoublet} M^2_{\rm Higgs}\ =\ \left(\! \begin{array}{cc} |\lambda|^2 |S|^2 & -\,\kappa \lambda (M^2 - X_1 X_2 )\: +\: \lambda A_\lambda S \\ -\,\kappa^* \lambda^* (M^2 - X^*_1 X^*_2 ) +\: \lambda^* A^*_\lambda S^* & |\lambda|^2 |S|^2 \end{array}\!\right)\ . \end{equation} The requirement of positive definiteness may be translated into the simple condition: \begin{equation} \label{Scondition} |\lambda|\, |S|^2\ \ge\ |\kappa (M^2 - X_1 X_2 )\: -\: A_\lambda S|\ . \end{equation} From this last inequality, we may see that the condition $\lambda \stackrel{>}{{}_\sim} \kappa$ is sufficient for ending hybrid inflation to the `good' vacuum. Likewise, one obtains a condition analogous to~(\ref{Scondition}) from the sneutrino mass matrix, which amounts to having $\rho \stackrel{>}{{}_\sim} \kappa$. The above two restrictions on the superpotential couplings $\lambda$ and $\rho$ will be imposed throughout our analysis. As mentioned above, after the end of inflation, one has $X^{\rm end}_{1,2} = M$, giving rise to a high mass for the inflaton field, i.e.~$2|\kappa |^2 M^2 |S|^2$. Combining this fact with the soft SUSY-breaking tadpole $-\kappa a_S M^2 S$ and the trilinear coupling $\kappa A_\kappa S X^{\rm end}_1 X^{\rm end}_2$, one gets a VEV for the inflaton~\cite{DLS}: \begin{equation} \label{Send} v_S\ \equiv\ \langle S^{\rm end} \rangle\ =\ \frac{1}{2|\kappa|}\, \Big|\,A_{\kappa} - a_S\,\Big|\ , \end{equation} where we have neglected the VEVs of the Higgs doublets $H_{u,d}$. The VEV of $S$ induces an effective $\mu$-term and an SO(3) symmetric lepton-number-violating Majorana mass $m_N$ of the electroweak order~\cite{PU2}: \begin{equation} \label{mumN} \mu\ =\ \lambda\, v_S\;, \qquad m_N\ =\ \rho\, v_S\; . \end{equation} If $\rho$ and $\lambda$ are comparable in magnitude, then these two mass parameters are tied together and can naturally be of the TeV or even of the electroweak scale. In Sections~\ref{reheat} and~\ref{inflation}, we will derive the constraints on the key theoretical parameters $\kappa$, $\lambda$, $\rho$ and $M$ from the requirement of a low reheat temperature, $T_{\rm reh} \stackrel{<}{{}_\sim} 10^9$~GeV, and successful inflation. \setcounter{equation}{0} \section{Preheating and Second Reheating}\label{reheat} In the SUGRA framework, the reheat temperature is constrained by the fact that an overabundant amount of gravitinos may destroy, through their possible late decays, the successful predictions of Big Bang nucleosynthesis~\cite{Sarkar}. This possibility is avoided, if the gravitino abundance $Y_{3/2}$ is small enough, i.e.~$Y_{3/2} < 10^{-12}$. The latter may be translated to an upper limit on the reheat temperature, i.e.~$T_{\rm reh} \stackrel{<}{{}_\sim} 10^9$~GeV. If the gravitinos are stable, the above limit may be relaxed by one order of magnitude to $\sim 10^{10}$~GeV. This depends on whether the so-called next-to-lightest supersymmetric particle (NLSP) has a small branching fraction to hadronic decay modes~\cite{FIY}. In addition to the above upper limit, the reheat temperature $T_{\rm reh}$ is also constrained from below, depending on the mechanism of baryogenesis. Thus, for successful electroweak-scale resonant leptogenesis, a lower bound of order TeV on $T_{\rm reh}$ should be considered. In the following we will study the post-inflationary dynamics. To this end, let us define the fields: \begin{eqnarray} \label{Xpm} X_\pm \!&=&\! \frac{1}{\sqrt{2}}\, (X_1\: \pm\: X_2)\ =\ \langle X_\pm \rangle\: +\: \delta X_\pm\; ,\nonumber\\ \delta X_\pm \!&=&\! \frac{1}{\sqrt{2}}\, (R_\pm\, +\, {\rm i}I_\pm)\; . \end{eqnarray} As mentioned in the introduction, inflation ends, once the inflaton field, $\phi = \sqrt{2}\, {\rm Re}\, S$, rolls below a critical value $\phi_c \approx \sqrt{2}\, M$. Then, the waterfall regime begins, where the waterfall fields $S$ and $R_+$ evolve rapidly (we use the gauge freedom to ensure that all VEVs point to real directions). Ignoring small corrections due to a non-vanishing FI $D$-term, $m_{\rm FI}$, the VEVs of $S$ and $R_+$ oscillate around \emph{zero}, whereas $X_+$ attains its U(1)$_X$-breaking VEV, $\langle X_+ \rangle = \sqrt{2} M$. The masses of the waterfall- or $\kappa$-sector fields $\phi$ and $R_+$ are equal to $m_\kappa = \sqrt 2 \kappa M$. The inflaton $\phi$ decays predominantly into pairs of charged and neutral higgsinos, $\tilde{h}^\pm_{u,d}$, $\tilde{h}^0_{u,d}$, $\tilde{\bar{h}}^0_{u,d}$, and into pairs of right-handed Majorana neutrinos $\nu_{1,2,3\,R}$. The decay width of the inflaton is given by \begin{equation} \label{infldecay} \Gamma_\phi\ =\ \frac{1}{32\pi}\: \Big(\, 4\lambda^2\: +\: 3 \rho^2\, \Big)\: m_\kappa\; . \end{equation} It turns out that the field $R_+$ decays into the scalar SUSY partners of the aforementioned fields at the same rate. Hence, we find \begin{equation} \Gamma_\phi\ =\ \Gamma_{R_+}\ \equiv\ \Gamma_\kappa\; . \end{equation} The reheat temperature resulting from the perturbative decays of the $\kappa$-sector fields may usually be estimated by \begin{equation} \label{Treh} T^\kappa_{\rm reh}\ =\ \left( \frac{90}{\pi^2\, g_*}\right)^{1/4}\, \sqrt{\Gamma_\kappa\: m_{\rm Pl} }\ , \end{equation} where $g_* = 228.75$ is the number of the relativistic degrees of freedom in the supersymmetric model under consideration. The gravitino bound then implies that \begin{equation} \label{Tkappa} \kappa\, \left(\, \lambda^2\: +\: \frac{3}{4}\, \rho^2\, \right)\ \stackrel{<}{{}_\sim}\ 3 \cdot 10^{-15}\,\times\, \left(\frac{T^\kappa_{\rm reh}}{10^9~{\rm GeV}}\right)^2\, \left( \frac{10^{16}~{\rm GeV}}{M}\right)\; . \end{equation} If $\kappa \approx \lambda \approx \rho$, this amounts to being each individual coupling smaller than about $10^{-5}$, for $M =10^{16}$~GeV and $T^\kappa_{\rm reh} \stackrel{<}{{}_\sim} 10^9$~GeV. So far, we have only considered the post-inflationary dynamics of the $\kappa$-sector fields, $S$, $R_+$ and $I_+$, to which all the energy of the inflationary potential is stored at the onset of the waterfall regime. We now turn our attention to the $g$-sector, namely to the particles associated with the extra U(1)$_X$ gauge group. This distinction of the different fields involved after inflation is made clear in Table~\ref{spectrum}. Thus, the $g$- or U(1)$_X$ gauge- sector contains the U(1)$_X$ gauge boson $V_\mu$, the Dirac fermion $\psi_g$, which consists of the gaugino $\lambda$ and the fermionic superpartner of $X_-$, and the scalars $R_-$ and $I_-$; the field $I_-$ is a massless would-be Goldstone boson which becomes the longitudinal component of $V_\mu$. Each of the $g$-sector particles has a mass $m_g=2^{-1/2} g \langle X_+ \rangle$. In fact, during the waterfall transition, their masses evolve rapidly from 0 to $g M$. As we will see below, this rapid non-adiabatic mass variation triggers the so-called preheating mechanism, through which the $g$-sector particles can be produced in sizeable amounts. Their decays can only be induced by the presence of a non-vanishing $D$-term, which breaks explicitly a discrete charge symmetry in the $F$- and the $D$-term sectors which would remain otherwise intact even after the SSB of the U(1)$_X$. \begin{table} \begin{center} \begin{tabular}{|c|c|c|c|} \hline & & & \\ Sector & Boson & Fermion & Mass\\ & & & \\ \hline\hline & & & \\ Waterfall & $S$, $R_+$, $I_+$ & $\psi_\kappa= \left( \begin{array}{c} \frac{1}{\sqrt2} \left[ \left(1-\frac{v}{2M}\right)\psi_{X_1} +\left(1+\frac{v}{2M}\right)\psi_{X_2} \right] \\ \psi_S^\dagger \end{array} \right) $ & $\sqrt 2 \kappa M$ \\ ($\kappa$-sector) & & & \\ & & & \\ \hline & & & \\ U(1)$_X$ Gauge & $V_\mu$, $R_-$ & $\psi_g= \left( \begin{array}{c} \frac{1}{\sqrt2} \left[ \left(1+\frac{v}{2M}\right)\psi_{X_1} -\left(1-\frac{v}{2M}\right)\psi_{X_2} \right] \\ -{\rm i}\lambda^\dagger \end{array} \right) $ & $g M$ \\ ($g$-sector) & & & \\ & & & \\ \hline \end{tabular} \end{center} \caption{\em Particle spectrum of the waterfall and U(1)$_X$ gauge sectors after inflation, where $V_\mu$ denotes the U(1)$_X$ gauge boson and $\lambda$~its associate gaugino.}\label{spectrum} \end{table} To make this last point explicit, let us express the relevant $F$- and $D$-term potential in terms of the fields $X_\pm$ defined in~(\ref{Xpm}): \begin{equation} \label{VFD} V_{FD}\ =\ \frac{\kappa^2}{4}\, \Big|\,X^2_+\: -\: X^2_-\: -\: 2\,M^2\,\Big|^2\ +\ \frac{g^2}{8}\, \Big(\, X^*_+ X_-\: +\: X^*_- X_+\: -\: m^2_{\rm FI}\, \Big)^2\; . \end{equation} It is obvious that the potential $V_{FD}$ possesses an additional discrete charge symmetry under the transformation, $X_\pm \to \pm X_\pm$, if the FI mass term vanishes, $m^2_{\rm FI} = 0$. In the absence of a FI term, this symmetry will still survive even after the SSB of the U(1)$_X$ along the flat direction $\langle X_1\rangle = \langle X_2 \rangle = M$,\footnote{Observe that an analogous discrete charge symmetry also survives after SSB in the so-called $D$-term inflationary model~\cite{Halyo}, where $M = 0$ and $m_{\rm FI} \neq 0$. In this case, the waterfall fields $X_{1,2}$ transform as $X_{1,2} \to \pm X_{1,2}$, while their VEVs after inflation are $\langle X_1 \rangle = m_{\rm FI}$ and $\langle X_2 \rangle = 0$.} or equivalently when $\langle X_+ \rangle = \sqrt{2} M$ and $\langle X_-\rangle = 0$. As a consequence, the U(1)$_X$ gauge boson $V_\mu$, the scalar field $R_- = \sqrt{2}\,{\rm Re} (X_-)$ and their fermionic superpartner $\psi_g$ are all stable with a mass $g M$. This feature is highly unsatisfactory for the hybrid model without a FI term, since these particles can be produced in large numbers during the preheating process, and since they are very massive, they could dominate and so overclose the Universe at later times. The presence of the FI term $m_{\rm FI}$ breaks explicitly the above discrete charge symmetry and so provides a new decay mechanism for making these particles unstable. To leading order in the expansion parameter $m_{\rm FI}/M$, the potential $V_{FD}$ given in~(\ref{VFD}) can be minimized using the linear field decompositions \begin{equation} \label{Xdec} X_+\ =\ \sqrt{2}\,M\: +\: \delta X_+\,,\qquad X_-\ =\ \frac{v}{\sqrt{2}}\: +\: \delta X_-\ , \end{equation} where $v = m^2_{\rm FI}/(2M)$. Table~\ref{spectrum} exhibits the particle spectrum of the waterfall and U(1)$_X$ gauge sectors to leading order in $m_{\rm FI}/M$. Unlike the case of a vanishing FI $D$-term, the scalar field $R_-$ of mass $gM$ will now decay into pairs of two lighter scalars, $R_+$ and $I_+$, of mass $\sqrt{2}\,\kappa M$, assuming that $g \gg \kappa$. The strength of this coupling is given by the Lagrangian \begin{equation} \label{Lint} {\cal L}_{\rm int}\ =\ \frac{g^2 m^2_{\rm FI}}{8 M}\; R_-\, (R_+^2 + I_+^2)\; . \end{equation} The $D$-term induced decay width of the $R_-$ particle can readily be found to be \begin{equation} \label{GammaR} \Gamma_{R_-}\ =\ \frac{g^3}{128 \pi}\, \frac{m^4_{\rm FI}}{M^3}\ , \end{equation} and the same rate also holds true for the decay of $I_-$, or equivalently for the longitudinal polarization of $V_\mu$. Correspondingly, the decays of the $g$-sector fermions $\psi_g$ are induced by the Lagrangian \begin{equation} {\cal L}_{\rm int}\ =\ -\, \frac{g}{8}\ \left(\frac{m_{\rm FI}}{M}\right)^2\, \left(R_+ -{\rm i}I_+\right)\, \bar\psi_g\, \frac{1-\gamma_5}{2}\, \psi_\kappa\ +\ {\rm H.c.} \end{equation} Neglecting soft SUSY-breaking, we find that $\Gamma_{\psi_g} = \Gamma_{R_-} \equiv \Gamma_g$. If the decay rate $\Gamma_g$ of the $g$-sector particles is sufficiently low, they may dominate the energy density of the Universe at later times, eventually leading to a second reheating phase due to their out of equilibrium decays. To offer an initial estimate, consider that, after the first reheating, the energy density $\varrho_\kappa$ of the waterfall-sector fields gets distributed among their decay products and so diluted as relativistic radiation $\propto a^{-4}$, where $a$ is the usual cosmological scale factor describing the expansion of the Universe. Meanwhile, the energy density $\varrho_g$ of the ultraheavy $g$-sector particles produced via preheating scales as $\propto a^{-3}$, such that $\varrho_g/\varrho_\kappa\propto a$. Moreover, during a radiation-dominated epoch, the dependence of the Hubble expansion rate $H$ on $a$ is $H\propto a^{-2}$. Let us therefore denote with $H_{\rm reh}$ the Hubble rate at the first reheating of the Universe and $H_{\rm eq}$ the Hubble rate at the time, when $\varrho_g=\varrho_\kappa$. Then, the U(1)$_X$ gauge-sector particles will dominate the energy density of the Universe, when \begin{equation} \label{condition:domination} H_{\rm eq}\ =\ H_{\rm reh}\, \left(\frac{\varrho^0_g}{\varrho^0_\kappa}\right)^2\ \gg\ \Gamma_g\;, \end{equation} where the superscript $0$ stands for the energy density right after preheating. Note that $\varrho_g/\varrho_\kappa$ is conserved until the time of the first reheating, since both $\varrho_g$ and $\varrho_\kappa$ scale as $a^{-3}$ during this period. The $g$-sector particle production via preheating can be computed numerically~\cite{PREHEATING}, by first solving for the mode functions and then using these to calculate the Hamiltonian energy density. For the evolution of the VEVs $\langle X_1 \rangle \approx \langle X_2 \rangle$, we assume that they initially undergo strong damping due to tachyonic preheating~\cite{TACHYPREH}. This phenomenon can be mimicked by setting \begin{equation} \label{profile} \langle X_1 \rangle\ =\ \langle X_2 \rangle\ =\ \left\{ \begin{array}{lr} 0\,, & \textnormal{for}\quad t\leq -\pi/(4 \sqrt 2 \kappa M)\,,\\ \frac{1}2\,M\, [1+\sin(\sqrt 2\kappa M t)]\,, & \textnormal{for}\quad -\pi/(4 \sqrt 2 \kappa M)< t < \pi/(4 \sqrt2 \kappa M)\,,\\ M\, & \textnormal{for}\quad t\geq \pi/(4 \sqrt2 \kappa M)\,, \end{array} \right. \end{equation} More precise forms of field evolutions may be obtained using numerical simulations~\cite{TACHYPREH}. For an initial estimate, however, only the velocity of the transition is important. In Fig.~\ref{figure:preheating} we display the energy densities $\varrho_F$ and $\varrho_B$ of the $g$-sector fermions $\psi_g$ and bosons $R_-$ and $V_\mu$ (produced via preheating), normalized to the energy density $\varrho_{\rm WF} \equiv \varrho_\kappa$ carried by the waterfall-sector particles, as functions of the U(1)$_X$ gauge coupling $g$, for $\kappa = 10^{-3}$. For the given profile~(\ref{profile}) of field evolutions, these normalized energy densities depend only very weakly on $\kappa$. The above results strongly suggest that the U(1)$_X$ gauge-sector particles, $\psi_g$, $R_-$ and $V_\mu$, if sufficiently long-lived, will dominate the energy density of the early Universe. We may estimate the second reheat temperature $T^g_{\rm reh}$ caused by their late decays, by employing a formula very analogous to (\ref{Treh}). Solving this last relation for the ratio $m_{\rm FI}/M$ yields \begin{equation} \label{ratio:mFI:M} \frac{m_{\rm FI}}{M} \approx \ 8.4 \cdot 10^{-4}\times \left( \frac{0.5}{g}\right)^{3/4} \left(\frac{T^g_{\rm reh}}{10^9~{\rm GeV}}\right)^{1/2}\, \left( \frac{10^{16}~{\rm GeV}}{M}\right)^{1/4}\; . \end{equation} For second reheat temperatures of cosmological interest, i.e.~$1~{\rm TeV}\leq T^g_{\rm reh}\leq 10^9~{\rm GeV}$, we obtain the combined constraint for $M = 10^{16}$~GeV: \begin{equation} \label{FIcombined} 10^{-6}\ \stackrel{<}{{}_\sim}\ \frac{m_{\rm FI}}{M}\ \stackrel{<}{{}_\sim}\ 10^{-3}\; . \end{equation} From our discussion in this section, it is evident that the late decays of the ultraheavy U(1)$_X$ gauge-sector fields, which are copiously produced during the preheating epoch, will give rise to a second reheating phase in the evolution of the early Universe at a temperature $T^g_{\rm reh} \ll T^\kappa_{\rm reh}$. This makes the $F_D$-term hybrid model an interesting cosmological scenario that could even lead to a complete relaxation of the strict bound~(\ref{Tkappa}) on the couplings $\kappa,\ \lambda,\ \rho$. The reason is that gravitinos, which are produced very efficiently at high reheat temperatures $T^\kappa_{\rm reh}>10^9 {\rm GeV}$, will now be diluted by the large entropy release from the late decays of the $g$-sector particles. In this way, the so-called gravitino overproduction problem can be completely avoided. A~detailed study of this topic will be given elsewhere~\cite{GPP}. \setcounter{equation}{0} \section{Inflation}\label{inflation} In this section we will discuss the additional constraints on the theoretical parameters of the $F_D$-term hybrid model from the power spectrum $P_{\cal R}^{1/2}$ and the spectral index $n_s$. We distinguish two possible regimes of inflation: (i) the cold hybrid inflation (CHI), where dissipative effects on inflation are negligible, e.g.~for $\kappa ,\ \lambda ,\ \rho\ \stackrel{<}{{}_\sim} 10^{-2}$ and (ii) the warm hybrid inflation (WHI), where dissipation might dominate over the expansion rate of the Universe~\cite{AB}. \subsection{Cold Hybrid Inflation} In models of hybrid inflation, the spectral index $n_s$ may well be approximated as follows~\cite{review}: \begin{equation} \label{nS} n_s\: -\: 1\ =\ \frac{d\ln P^{1/2}_{\cal R}}{d\ln k}\ \approx\ 2\eta\; , \end{equation} where $k$ is the comoving wavenumber at the horizon exit and \begin{equation} \label{eta} \eta\ =\ m^2_{\rm Pl}\ \frac{V_{\phi\phi}}{V}\ \end{equation} is the so-called $\eta$-parameter. In~(\ref{eta}), $V$ denotes the inflationary potential, and $V_\phi = dV/d\phi$, $V_{\phi\phi} = d^2V/d\phi^2$ etc. The current WMAP data~\cite{WMAP} show a strong preference for a red-tilted spectrum, with $n_s - 1 \le 0$, implying that $V_{\phi\phi} \le 0$. The actual value is $n_s = 0.98 \pm 0.02$~\cite{Lyman}. The $T_{\rm reh}$ constraint~(\ref{Tkappa}) on the theoretical parameters imply that $\kappa,\ \lambda,\ \rho \stackrel{<}{{}_\sim} 10^{-5}$. In this case, the radiative correction to the potential becomes subdominant and may be ignored to a good approximation. The potential driving inflation simplifies considerably to \begin{equation} \label{VCHI} V_{\rm inflation}\ =\ \kappa^2 M^4\: -\: \sqrt{2}\,\kappa\,a_S M^2 \phi\: +\: \frac{1}{2}\, M^2_S\, \phi^2\: +\: \frac{\kappa^2\,M^4}{8\,m^4_{\rm Pl}}\, \phi^4\ , \end{equation} where $\phi = \sqrt{2}\, {\rm Re}\, S$ is the inflaton field canonically normalized. For $M_S < 1$~TeV, $\kappa \ge 10^{-6}$ and $M \ge 10^{15}$~GeV, the soft SUSY-breaking term $M_S$ can be omitted. The inflationary potential $V_{\rm inflation}$ of (\ref{VCHI}) generically leads to a blue-tilted spectrum, i.e.~$n_s - 1 = 2 \eta > 0$, which is slightly disfavoured by the recent WMAP data. In the following, we will concentrate on the regime where the loop correction dominates the slope of the potential, such that a negative value for $n_s - 1$ becomes possible. This possibility arises for $10^{-4} \stackrel{<}{{}_\sim}\ \kappa ,\ \lambda ,\ \rho\ \stackrel{<}{{}_\sim} 10^{-2}$. Naively, such large values of the parameters lead to a too high reheat temperature $T_{\rm reh}$, i.e.~$T_{\rm reh} \stackrel{>}{{}_\sim} 10^{10}$~GeV. However, as we have discussed in Section~\ref{reheat}, the presence of a subdominant $D$-term renders the stable U(1)$_X$ gauge-sector fields unstable, and so a large amount of entropy can be released from their late decays, leading to a $T_{\rm reh}$ which may even be as low as 1~TeV. Our results simplify considerably if one assumes that the slope of the inflationary potential given in~(\ref{VpotFD}) is dominated by the $\lambda$-dependent term. To be specific, the number of $e$-folds ${\cal N}_e$ is given by \begin{equation} \label{Nefold} {\cal N}_e\ =\ \frac{1}{m^2_{\rm Pl}}\; \int_{\phi_{\rm end}}^{\phi_{\cal N}}\, d\phi\: \frac{V}{V_\phi}\ \approx\ \frac{2\pi^2}{\lambda^2}\; \frac{\phi^2_{\cal N}}{m^2_{\rm Pl}}\ . \end{equation} Notice that at the horizon exit, it is $\phi_{\cal N} = \sqrt{{\cal N}_e/2}\, (\lambda/\pi)\, m_{\rm Pl}$ and $\phi_{\cal N} \stackrel{<}{{}_\sim} 10^{-1}\, m_{\rm Pl}$, for $\lambda \stackrel{<}{{}_\sim} 0.1$ and ${\cal N}_e = 60$. Hence inflation starts at values of $\phi_{\cal N}$ well below $m_{\rm Pl}$. In terms of the number of $e$-folds ${\cal N}_e$, the power spectrum $P^{1/2}_{\cal R}$ of the curvature perturbations may now be given by \begin{equation} \label{PRCHI} P^{1/2}_{\cal R}\ =\ \frac{1}{2\sqrt{3}\, \pi m^3_{\rm Pl}}\; \frac{V^{3/2}}{|V_\phi|} \ \approx\ \sqrt{\frac{2{\cal N}_e}{3}}\ \frac{\kappa}{\lambda}\ \Bigg( \frac{M}{m_{\rm Pl}}\Bigg)^2\ =\ 5\times 10^{-5}\ . \end{equation} Evidently, for ${\cal N}_e = 60$ and $M=10^{16}$~GeV, the parameter $\lambda$ cannot be by more than one order of magnitude larger than $\kappa$, i.e.~$\lambda \stackrel{<}{{}_\sim} 10\, \kappa$. Finally, the spectral index $n_s$ in terms of ${\cal N}_e$ may be expressed as follows: \begin{equation} \label{etaCHI} n_s\, -\, 1\ =\ -\ \frac{1}{{\cal N}_e}\ \approx\ -\,0.02\ , \end{equation} for ${\cal N}_e = 50$--60. In this CHI regime, the model predicts a red-tilted spectrum, as currently favoured by the WMAP data. \subsection{Warm Hybrid Inflation} It has been extensively argued~\cite{AB} that dissipative effects due to radiation production of massless particles during inflation may dominate over the expansion rate $H$ of the Universe. This form of inflation is known as warm inflation. Although a firm first principles derivation for the existence of a strong dissipative regime of inflation is still missing,\footnote{A detailed calculation based on a two-particle irreducible effective action in an expanding deSitter background metric would be highly preferable.} it might be worth presenting tentative results for such a possible situation, using the semi-empiric formalism on warm inflation developed in~\cite{AB}. In the framework of WHI, dissipation occurs from the radiation produced by the decays of the excited $H_u$ doublet of mass $\lambda S$. Specifically, the interactions relevant to WHI are \begin{equation} \label{Lwarm} -\, {\cal L}_{\rm int}^{\rm WHI}\ =\ |S|^2\, \bigg[\, |\lambda |^2\, |H_u|^2\: +\: |\rho |^2\, \bigg(\,\sum_{i=1}^3\, |\widetilde{N}_i|^2\bigg)\,\bigg]\ +\ \Big( h_t\, H_u\, \bar{Q}_t\, t_R\: +\: h^{\nu}_{ij}\, \bar{L}_i \tilde{h}_u \widetilde{N}_j\: +\: {\rm H.c.}\Big)\; . \end{equation} The dominant decay mode will be $H_u \to Q_t t_R$~\cite{BB}; the other possible decay channel $\widetilde{N}_j \to L_i \tilde{h}_u$ is Yukawa-coupling suppressed and kinematically allowed only when $\rho > \lambda$. Adapting the results of~\cite{AB,BB} to our model, the dominant friction term for $|S| \gg M$ is given by \begin{equation} \label{Ys} Y_S\ \approx\ \frac{\sqrt{\pi}\, \alpha_\lambda^{3/2}\, \alpha_t}{20\,\sqrt{2}}\ \phi\; , \end{equation} where $\alpha_\lambda = \lambda^2/(4\pi)$ and $\alpha_t = h^2_t/(4\pi)$. The dynamics of warm inflation is governed by the following two equations: \begin{eqnarray} \label{phiS} \ddot{\phi}\ +\ 3H\, ( 1 + r )\, \dot{\phi}\ +\ V_\phi & = & 0\; ,\\ \label{rhorad} \dot{\rho}_{\rm rad}\ +\ 4\, H\rho_{\rm rad} & = & Y_S\, \dot{\phi}^2\ , \end{eqnarray} where $r = Y_S/(3H)$, with $H^2 \approx \kappa^2 M^4/(3 m^2_{\rm Pl})$. In the strong dissipative regime where $r \gg 1$, inflation usually ends when $\rho_{\rm rad} > \rho_{\rm vac} \approx \kappa^2 M^4$. Assuming conditions of slow roll during WHI, i.e.~$\eta /r^2 \ll 1$, we may determine the number of $e$-folds by \begin{equation} \label{NeWHI} {\cal N}_e \ =\ \frac{1}{m^2_{\rm Pl}}\; \int^{\phi_{\cal N}}_{\phi_{\rm end}}\, d\phi\ \frac{(1+r)\,V }{V_\phi}\ =\ \frac{\pi \alpha_\lambda^{1/2}\, \alpha_t}{60\, \kappa}\ \frac{ \phi^3_{\cal N}}{ m_{\rm Pl}\, M^2}\ . \end{equation} In the limit $r\gg 1$, the power spectrum $P_{\cal R}^{1/2}$ due to WHI is approximately given by \begin{equation} \label{PRWHI} (P_{\cal R}^{1/2})_{\rm WHI}\ \approx\ \left( \frac{3\pi}{4} \right)^{1/4}\, \sqrt{\frac{T_{\rm rad}}{H}}\; r^{5/4}\, (P_{\cal R}^{1/2})_{\rm CHI}\ . \end{equation} The temperature $T_{\rm rad}$ associated with radiation production can be calculated from (\ref{rhorad}), by solving the approximate equation \begin{equation} \label{Trad} \rho_{\rm rad}\ =\ \frac{\pi^2}{30}\ g_*\ T^4_{\rm rad}\ \approx\ \frac{3r}{4}\ \dot{\phi}^2\; , \end{equation} where $\dot{\phi} \approx - V_\phi/(3r H)$ is evaluated at the horizon exit. Putting everything together, we find \begin{equation} \label{PRWHI2} (P_{\cal R}^{1/2})_{\rm WHI}\ \approx\ g_*^{-1/8}\, {\cal N}_e^{5/8}\, (2\kappa)^{1/4}\, \alpha_\lambda^{5/8}\, \alpha_t^{1/2}\, \left(\frac{M}{m_{\rm Pl}}\right)^{1/2}\ =\ 5 \times 10^{-5}\ . \end{equation} It is interesting to observe that WHI leads to a viable inflationary scenario even for strong couplings, e.g.~for $\kappa ,\ \alpha_\lambda,\ \alpha_t\ \sim~1$. In this case, the U(1)$_X$-breaking scale $M$ will be as low as $10^{10}$~GeV, in agreement with the earlier discussion in \cite{BB}. Obviously, it would be difficult to associate such a low scale for $M$ with gauge coupling unification. Finally, the spectral index $n_s$ in WHI is calculated in terms of ${\cal N}_e$ to be: $n_s - 1 \approx - 5/(4 {\cal N}_e) \approx -0.025$. \setcounter{equation}{0} \section{Baryon Asymmetry in the Universe}\label{BAU} As discussed in Section~\ref{reheat}, the late decays of the U(1)$_X$ gauge-sector particles may lead to a second reheating phase in the evolution of the early Universe, giving rise to a very low final reheat temperature~$T_{\rm reh}$. Depending on the size of the $D$-term, $T_{\rm reh}$ may even be as low as 1~TeV. In such a case, the BAU may be explained by thermal electroweak-scale resonant leptogenesis~\cite{APRD,PU2}. The $F_D$-term hybrid model under study can realize such a scenario even within a minimal SUGRA framework, where all soft SUSY-breaking parameters are constrained at the gauge-coupling unification point $M_X$, which can be chosen to be $M = M_X \approx 10^{16}$~GeV. Instead, electroweak baryogenesis is no longer viable in minimal SUGRA, since it requires an unconventionally large hierarchy between the left-handed and right-handed top squarks~\cite{EWBAU}. An advantageous feature of resonant leptogenesis is that the predictions for the BAU are almost independent of any pre-existing lepton- or baryon-number abundance. This kind of fixed-line attractor behaviour is a consequence of the quasi-in-thermal equilibrium dynamics governing the heavy Majorana neutrino sector. It results from the fact that the heavy neutrino decay widths can be several orders of magnitude larger than the expansion rate $H$ of the Universe. A detailed analysis of this dynamics was presented in \cite{PU2}, where single lepton-flavour and freeze-out sphaleron effects were systematically considered for the {\em first time}. In particular, it was shown that single lepton-flavour effects resulting from the Yukawa-neutrino couplings $h^{\nu}_{ij}$ can have a dramatic impact on the predictions for the BAU, enhancing its value by many orders of magnitude. From the model-building point of view, phenomenologically rich scenarios are now possible with testable implications for high-energy colliders~\cite{prodN} and low-energy observables, such as $\mu \to e\gamma$, $\mu \to eee$ and $\mu \to e$ conversion in nuclei~\cite{LFVN}. We will not reiterate all these results here, but only underline some of the key model-building aspects related to the neutrino sector of the $F_D$-term hybrid model. The $F_D$-term hybrid model contains a $3\times 3$ Majorana mass matrix $M_S$, which is SO(3) symmetric at the gauge-coupling unification point $M_X = M \approx 10^{16}$~GeV, i.e.\ $M_S = m_N {\bf 1}_3$. The parameter $m_N = \rho v_S$ is a universal Majorana mass whose natural value is of the order of the soft SUSY-breaking or the electroweak scale, i.e.~$m_N \sim M_{\rm SUSY}$ or $m_t$. The SO(3) symmetry of the heavy neutrino sector is broken explicitly by the Yukawa neutrino couplings $h^{\nu}_{ij}$. In order to explain the low-energy light neutrino data, the breaking of the SO(3) symmetry should proceed via an intermediate step, namely SO(3) should first break into its subgroup SO(2) $\simeq$ U(1)$_l$. This can be achieved by coupling all lepton doublets $L_{e,\mu,\tau}$ to the linear combination: $\frac{1}{\sqrt{2}}\, (\nu_{2R} + i \nu_{3R})$. These Yukawa couplings could be as large as the $\tau$-Yukawa coupling $h_\tau$, i.e.~$h^{\nu}_{i2} = i h^{\nu}_{i3} \sim 10^{-2}$. As a consequence of the U(1)$_l$ symmetry, the resulting light neutrino mass matrix ${\bf m}^\nu$ vanishes identically to all orders in perturbation theory. The remaining U(1)$_l$ symmetry can be broken by smaller Yukawa couplings of the order of the electron Yukawa coupling $h_e$, i.e.~$h^{\nu}_{i1} = \varepsilon_i \sim 10^{-6}$--$10^{-7}$, which arise when one couples $L_{e,\mu ,\tau}$ to $\nu_{1R}$~\cite{PUcomment}. Further breaking of the U(1)$_l$ symmetry is induced in the heavy-neutrino sector by renormalization-group and threshold effects while running $M_S$ from $M$ to $m_t$~\cite{Branco}. Thus, $M_S$ will generically modify to: $M_S = m_N {\bf 1}_3 + \Delta M_S$, where one typically has $(\Delta M_S)_{ij}/m_N \sim 10^{-5}$--$10^{-7}$. Taking the effect of U(1)$_l$-breaking parameters $(\Delta M_S)_{ij}$ and $\varepsilon_i$ into account, one obtains a light neutrino mass matrix which can comfortably accommodate the low-energy light neutrino data, e.g.~with an inverted hierarchical light neutrino spectrum~\cite{PU2}. On the other hand, the heavy neutrino sector of the $F_D$-term hybrid model consists of 3 nearly degenerate heavy Majorana neutrinos $N_{1,2,3}$ of mass $m_{N_{1,2,3}} \approx m_N$, which can give rise to successful baryogenesis through thermal electroweak-scale resonant leptogenesis~\cite{PUcomment}. \setcounter{equation}{0} \section{Conclusions}\label{conclusions} We have studied $F$-term hybrid inflation in a novel supersymmetric extension of the SM, to which a subdominant FI $D$-term was added. We called this particular form of inflation $F_D$-term hybrid inflation. The $F_D$-term hybrid model we have been analyzing in this paper ties the $\mu$-parameter of the MSSM to an SO(3) symmetric Majorana mass $m_N$, through the VEV of the inflaton field. As a consequence, the model predicts {\em naturally} lepton-number violation at the electroweak scale. In order to obtain predictions for the observables $P_{\cal R}^{1/2}$, $n_s$ and $\eta_B$ compatible with global cosmological analyses~\cite{Lyman}, as well as interesting particle-physics phenomenology that could be tested in laboratory experiments, one needs to make certain assumptions for the model of $F_D$-term hybrid inflation: \begin{itemize} \item[ (i)] Successful hybrid inflation relies on the assumption that the inflaton field is displaced from its minimum in the beginning of inflation, whereas all other non-inflaton fields have zero VEVs, according to (\ref{initial}). \item[ (ii)] The present $F_D$-term hybrid scenario utilizes a minimal K\"ahler potential, where terms of order $H^2 |S|^2$ in the potential are set to zero or assumed to be negligible. This consideration introduces some tuning in general SUGRA models with non-minimal K\"ahler potentials. \item[(iii)] In order to get a red-tilted spectrum with negative $n_s - 1$, one has to assume that the radiative corrections dominate the slope of the inflationary potential. This possibility arises for superpotential couplings: $10^{-4} \stackrel{<}{{}_\sim} \kappa,\ \lambda,\ \rho \stackrel{<}{{}_\sim} 10^{-2}$. \item[ (iv)] Even though a bare $D$-tadpole may be present as a bare parameter in the tree-level Lagrangian, we have considered here, however, the possibility that such a term is generated radiatively after heavy degrees of freedom have been integrated out. These heavy degrees of freedom are assumed to be Planck-mass chiral superfields which are oppositely charged under the U(1)$_X$ and which break explicitly the discrete charge symmetry discussed after (\ref{VFD}) and in Appendix~A. \item[ (v)] We have assumed that the coupling $\rho$ of the inflaton to neutrino superfields is SO(3) symmetric or very close to it. After the inflaton receives a VEV, one ends up with 3 nearly degenerate heavy Majorana neutrinos with masses at the electroweak scale. This enables one to successfully address the BAU within the thermal electroweak-scale resonant leptogenesis framework (see our discussion in Section~\ref{BAU}). As has also been discussed in Section~\ref{BAU}, if one assumes that the neutrino-Yukawa couplings $h^\nu_{ij}$ have a certain hierarchical structure controlled by the approximate breaking of global flavour symmetries, the model can have further testable implications for $e^+e^-$ colliders and low-energy experiments of lepton flavour and/or number violation. \end{itemize} The requirement for a sufficiently low reheat temperature $T_{\rm reh} \stackrel{<}{{}_\sim} 10^9$~GeV, which does not lead to overproduction of gravitinos, provides an important constraint on the basic theoretical parameters $\kappa$, $\lambda$ and $\rho$. The naive limits on these couplings derived from reheating due to perturbative inflaton decay are very strict, i.e.~$\kappa,\ \lambda,\ \rho \stackrel{<}{{}_\sim} 10^{-5}$. These limits may be completely avoided by considering the late decays of the U(1)$_X$ gauge-sector particles which are induced by a non-vanishing FI $D$-term $m^2_{\rm FI}$. Their decay rates depend crucially on~$m^2_{\rm FI}$. As menioned above in point~(iv) and in Appendix~A, the generation of a FI $D$-tadpole and its size may be engineered by adding Planck-scale heavy degrees of freedom to the theory and by subjecting these into extended $R$ symmetries. In this way, a phase of second reheating takes place in the evolution of the early Universe, which can lead to a significant lowering of the reheat temperature even up to 1~TeV. The $F_D$-term hybrid model with electroweak-scale lepton number violation can easily be embedded within a minimal SUGRA theory, where all soft SUSY-breaking parameters are constrained at the gauge coupling unification point $M_X$ which can be chosen to be $M \approx 10^{16}$~GeV. Instead, electroweak baryogenesis is not viable in a minimal SUGRA scenario of the MSSM. Moreover, the CP-odd soft SUSY-breaking phases required for successful electroweak baryogenesis face severe constraints from the non-observation of the electron and neutron electric dipole moments, even though the latter arise diagrammatically at the 2-loop level~\cite{CKP}. The $F_D$-term hybrid model under discussion conserves $R$-parity. The reason is that all superpotential couplings either conserve the $B-L$ number or break it by even number of units. Specifically, the operator $\widehat{S}\widehat{N}_i\widehat{N}_i$ breaks explicitly $L$, as well as $B-L$, by 2~units. Consequently, the lightest supersymmetric particle (LSP) of the spectrum is expected to be stable, thus providing a viable candidate to address the so-called Cold Dark Matter (CDM) problem. The new aspect of our model is that right-handed sneutrinos could be the LSPs, opening up new possibilities in the phenomenology of CDM and its detection. From the particle-physics point of view and in the low-energy limit where the waterfall sector has decoupled and the $\rho$-coupling neglected for simplicity, the $F_D$-term hybrid model becomes identical to the so-called Minimal Nonminimal Supersymmetric Standard Model (MNSSM) in the decoupling limit of a large tadpole~\cite{PP}. In particular, in the framework of WHI discussed in Section~\ref{inflation}, the coupling $\lambda$ can be sizeable, i.e.~$\lambda \sim 0.6$. In this case, the Higgs phenomenology of the MSSM will modify drastically, despite the decoupling of the singlet Higgs states. One striking possibility in the MNSSM is that the charged Higgs boson $H^+$ could be lighter than the SM-like Higgs boson~\cite{PP2}, thus pointing to particular collider phenomenologies~\cite{DP}. However, even within the traditional scenario of CHI, where $\kappa, \lambda \stackrel{<}{{}_\sim} 10^{-2}$, the $F_D$-term hybrid model will favour particular benchmark scenarios of the MSSM. For example, if $\lambda \gg \kappa$, the $F_D$-term hybrid model may account for a possible large value of the $\mu$-parameter. Specifically, if $\lambda = 4 \kappa$, one gets from~(\ref{Send}) the hierarchy $\mu \approx 4 M_{\rm SUSY}$, which is the so-called CPX benchmark scenario~\cite{CPX} describing maximal CP violation in the MSSM Higgs sector at low and moderate values of $\tan \beta$. A possible natural solution to the famous cosmological constant problem is expected to provide further constraints on the model building of cosmologically viable models in future. Nevertheless, the $F_D$-term hybrid model presented in this paper constitutes a first attempt towards the formulation of a minimal Particle-Physics and Cosmology Standard Model, whose validity could, in principle, be tested in laboratory experiments and further vindicated by astronomical observations. \bigskip\bigskip \subsection*{Acknowledgements} We thank Arjun Berera, Rudnei Ramos and Antonio Riotto for useful discussions. We also thank Constantine Pallis for collaboration in the early stages of this project. AP~dedicates this work to the memory of Darwin Chang, an invaluable friend and collaborator. This work is supported in part by the PPARC research grants: PPA/G/O/2002/00471 and PP/C504286/1. \newpage \def\theequation{\Alph{section}.\arabic{equation}} \begin{appendix} \setcounter{equation}{0} \section{{\boldmath $D$}--Term Engineering}\label{Dappendix} The generation and the size of a $D$-term may be engineered by adding Planck-scale heavy degrees of freedom to the theory and by subjecting these into extended $R$ symmetries. To elucidate our point, let us first consider a model augmented by a pair of oppositely charged superfields $\widehat{\overline{X}}_{1,2}$, with U(1)$_X$ charges: $Q(\widehat{\overline{X}}_2) = - Q(\widehat{\overline{X}}_1 ) = Q(\widehat{X}_1) = - Q(\widehat{X}_2 ) = 1$. The extended superpotential $W$ of our interest is \begin{equation} \label{Wdterm} W \ =\ \kappa\, \widehat{S}\, \Big( \widehat{X}_1 \widehat{X}_2\: -\: M^2\Big)\ +\ \xi\, m_{\rm Pl}\, \widehat{\overline{X}}_1\,\widehat{\overline{X}}_2\ +\ \xi_1\, \frac{ ( \widehat{\overline{X}}_1\widehat{X}_1 )^2}{2\, m_{\rm Pl}}\ +\ \xi'_1\, \frac{ ( \widehat{\overline{X}}_2\widehat{X}_2 )^2}{2\, m_{\rm Pl}}\ . \end{equation} This form of the superpotential may be enforced by the $R$ symmetry: $\widehat{S} \to e^{i\alpha}\, \widehat{S}$, $\widehat{X}_{1,2} \to e^{\pm i\beta}\,\widehat{X}_{1,2}$, $\widehat{\overline{X}}_{1,2} \to e^{i(\frac{a}{2} \mp \beta)}\, \widehat{\overline{X}}_{1,2}$, $\widehat{L} \to e^{i\alpha}\, \widehat{L}$, $\widehat{Q} \to e^{i\alpha}\, \widehat{Q}$, with $W \to e^{i\alpha} W$. As before, all remaining fields are considered to be neutral under the $R$ symmetry. Notice that the same $R$-symmetry allows for the operator $\kappa' S (\widehat{X}_1 \widehat{X}_2 )^2/m^2_{\rm Pl}$. The presence of this superpotential term can trigger shifted hybrid inflation, where the gauge symmetry U(1)$_X$ is broken along the inflationary trajectory, thereby inflating away unwanted topological defects~\cite{JKLS}. A $D$-term will now be generated after integrating out the Planck-scale superfields $\widehat{\overline{X}}_{1,2}$. The loop-induced $D$-tadpole $m^2_{\rm FI}$ is found to be \begin{equation} \label{FIdterm} m^2_{\rm FI}\ \approx\ \frac{\xi^2_1 - \xi'^2_1}{8\pi^2}\ \frac{M^4}{m^2_{\rm Pl}}\ \ln\left(\frac{m_{\rm Pl}}{M}\right)\ . \end{equation} For $M = 10^{16}$~GeV, we find that $m_{\rm FI}/M \stackrel{<}{{}_\sim} 10^{-3}$, for $\xi_1,\ \xi'_1 \stackrel{<}{{}_\sim} 0.3$. Observe that if $\xi_1 = \xi'_1$, the discrete charge symmetry discussed after (\ref{VFD}) gets restored again and $m_{\rm FI}$ vanishes identically. The size of the $D$-term may be suppressed further, if the Planck-mass chiral superfields $\widehat{\overline{X}}_{1,2}$ possess higher U(1)$_X$ charges. In general, one may assume that the U(1)$_X$ charges of $\widehat{\overline{X}}_{1,2}$ are: $Q(\widehat{\overline{X}}_2) = - Q(\widehat{\overline{X}}_1 ) = n$, where $n\ge 1$. In addition, we require for $\widehat{\overline{X}}_{1,2}$ to transform under U(1)$_R$ as follows: \begin{equation} \label{Rsymn} \widehat{\overline{X}}_{1,2}\ \to\ e^{\frac{i}2\, [a \, \mp\, (n+1) \beta ]}\; \widehat{\overline{X}}_{1,2}\; , \end{equation} while $\widehat{S}$, $\widehat{X}_{1,2}$ and all other fields transform as before. With this symmetry restriction, the superpotential reads: \begin{equation} \label{Wdtermn} W \ =\ \kappa\, \widehat{S}\, \Big( \widehat{X}_1 \widehat{X}_2\: -\: M^2\Big)\ +\ \xi\, m_{\rm Pl}\, \widehat{\overline{X}}_1\,\widehat{\overline{X}}_2\ +\ \xi_n\, \frac{ (\widehat{\overline{X}}_1)^2\, (\widehat{X}_1)^{n+1}}{2\,m^n_{\rm Pl}}\ +\ \xi'_n\, \frac{ ( \widehat{\overline{X}}_2)^2\,(\widehat{X}_2)^{n+1}}{2\,m^n_{\rm Pl}}\ . \end{equation} In this case, the loop-induced $D$-term is given by \begin{equation} \label{FIdtermn} m^2_{\rm FI}\ \approx\ \frac{\xi^2_n - \xi'^2_n }{8\pi^2}\ \frac{M^{2(n+1)}}{m^{2n}_{\rm Pl}}\ \ln\left(\frac{m_{\rm Pl}}{M}\right)\ . \end{equation} To obtain a small ratio $m_{\rm FI}/M \sim 10^{-6}$, with $\xi_n,\ \xi'_n \sim 1$, one would need $n = 5,\ 6$. Finally, it is important to remark that the loop-induced $D$-term does not lead to spontaneous breakdown of global supersymmetry. \end{appendix} \newpage
Title: SDSSJ212531.92-010745.9 - the first definite PG1159 close binary system
Abstract: The archival spectrum of SDSSJ212531.92-010745.9 shows not only the typical signature of a PG1159 star, but also indicates the presence of a companion. Our aim was the proof of the binary nature ofthis object and the determination of its orbital period.We performed time-series photometry of SDSSJ212531.92-010745.9. We observed the object during 10 nights, spread over one month, with the Tuebingen 80cm and the Goettingen 50cm telescopes. We fitted the observed light curve with a sine and simulated the light curve of this system with the nightfall program. Furthermore, we compared the spectrum of SDSSJ212531.92-010745.9 with NLTE models, the results of which also constrain the light curve solution. An orbital period of 6.95616(33)h with an amplitude of 0.354(3)mag is derived from our observations. A pulsation period could not be detected. For the PG1159 star we found, as preliminary results from comparison with our NLTE models, Teff about 90000K, log g about 7.60, and the abundance ratio C/He = 0.05 by number fraction. For the companion we obtained with a mean radius of 0.4 +/- 0.1 Rsol, a mass of 0.4 +/- 0.1 Msol, and a temperature of 8200K on the irradiated side, good agreement between the observed light curve and the nightfall simulation, but we do not regard those values as final.
https://export.arxiv.org/pdf/astro-ph/0601512
\title{SDSS\,J212531.92$-$010745.9 - the first definite PG1159 close binary system} \author{T. Nagel\inst{1} \and S. Schuh\inst{2} \and D.-J. Kusterer\inst{1} \and T. Stahn\inst{2} \and S.D. H\"ugelmeyer\inst{2} \and S. Dreizler\inst{2} \and B.T. G\"ansicke\inst{3} \and M.R. Schreiber\inst{4} } \offprints{T. Nagel} \mail{nagel@astro.uni-tuebingen.de} \institute{Institut f\"ur Astronomie und Astrophysik, Eberhard-Karls-Universit\"at T\"ubingen, Sand 1, 72076 T\"ubingen, Germany \and Institut f\"ur Astrophysik, Georg-August-Universit\"at G\"ottingen, Friedrich-Hund-Platz 1, 37077 G\"ottingen, Germany \and Department of Physics, University of Warwick, Coventry, CV4 7AL, Great Britain \and Departamento de Fisica y Meteorologia, Facultad de Ciencias, Universidad de Valparaiso, Valparaiso, Chile } \date{Received xx.xx.xx / Accepted xx.xx.xx} \abstract {} {The archival spectrum of SDSS\,J212531.92$-$010745.9 shows not only the typical signature of a PG\,1159 star, but also indicates the presence of a companion. Our aim was the proof of the binary nature of this object and the determination of its orbital period.} {We performed time-series photometry of SDSS\,J212531.92$-$010745.9. We observed the object during 10 nights, spread over one month, with the T\"ubingen 80\,cm and the G\"ottingen 50\,cm telescopes. We fitted the observed light curve with a sine and simulated the light curve of this system with the \texttt{nightfall} program. Furthermore, we compared the spectrum of SDSS\,J212531.92$-$010745.9 with NLTE models, the results of which also constrain the light curve solution. } {An orbital period of 6.95616(33)\,h with an amplitude of 0.354(3)\,mag is derived from our observations. A pulsation period could not be detected. For the PG\,1159 star we found, as preliminary results from comparison with our NLTE models, $T_{\rm eff}$\,$\sim$\,90\,000\,K, $\log g$\,$\sim$\,7.60, and the abundance ratio C/He\,$\sim$\,0.05 by number fraction. For the companion we obtained with a mean radius of $0.4\pm 0.1\,\rm R_\odot$, a mass of $0.4\pm 0.1\,\rm M_\odot$, and a temperature of 8\,200\,K on the irradiated side, good agreement between the observed light curve and the \texttt{nightfall} simulation, but we do not regard those values as final. } {} \keywords{stars: AGB and post-AGB -- white dwarfs -- binaries: close } \titlerunning{The first definite PG1159 close binary system} \authorrunning{T. Nagel at al.} \section{Introduction} PG\,1159 stars are hot hydrogen-deficient (pre-)white dwarfs with effective temperatures between 75\,000 and 200\,000\,K, and $\log g$\,=\,5.5--8.0 (Werner 2001). They are in the transition between the asymptotic giant branch (AGB) and cooling white dwarfs. Spectra of PG\,1159 stars are dominated by absorption lines of He\,{\sc ii}, C\,{\sc iv} and O\,{\sc vi}. Current theory suggests (e.g. Werner 2001) that they are the outcome of a late helium-shell flash, a phenomenon that drives the currently observed fast evolutionary rates of three well-known objects (FG~Sge, Sakurai's object, V605 Aql). Flash-induced envelope mixing produces a H-deficient stellar surface. The photospheric composition then essentially reflects that of the region between the H- and He-burning shells in the precursor AGB star. The He-shell flash forces the star back onto the AGB. The subsequent, second post-AGB evolution explains the existence of Wolf-Rayet central stars of planetary nebulae and their successors, the PG\,1159 stars. Currently, 37 PG\,1159 stars are known. Figure\,\ref{pg1159stars} shows their position in a log\,$T_{\rm eff}$-$\log g$-diagram. Two of them have been found to be binary stars. These are NGC\,246 (e.g. Bond \& Ciardullo 1999), which is a resolved visual binary, and PG\,2131+066 (Wesemael \etal 1985). Concerning the latter, it is still unclear whether it is a close binary (Paunzen \etal 1998) or a resolved visual binary with an M2V star as companion (Reed \etal 2000). \section{The spectrum of SDSS\,J212531.92$-$010745.9} The spectrum of SDSS\,J212531.92$-$010745.9 (u=17.15, g=17.54, r=17.75, i=17.79, z=17.83), taken on Sept. 6th 2002, is from the Sloan Digital Sky Survey (SDSS) archive Data Release (DR) 4. The spectrum shows significant features that are typical for PG\,1159 stars, for example the strong C\,{\sc iv} absorption lines at $4650-4700\AA$ and He\,{\sc ii} at $4686\AA$ (Fig.\,\ref{spectrum}). Furthermore, the spectrum shows features which indicate the presence of a companion. The Balmer series of hydrogen is seen in emission, H$_\alpha$ - H$_\delta$ can clearly be identified. This is probably due to a cool companion which is heated up by irradiation from the hydrogen-deficient PG\,1159 star. Figure\,\ref{spectrum} shows the observed spectrum ($t_{\rm exp}=3703\,\rm s$) of SDSS\,J212531.92$-$010745.9. Overlayed are a PG\,1159 NLTE model spectrum with $T_{\rm eff}$\,=\,90\,000\,K, $\log g$\,=\,7.60, C/He\,=\,0.05, and N/He\,=\,0.01, a blackbody model spectrum with $T$\,=\,8\,200\,K for the irradiated companion, and the sum of the two model spectra. The parameters of both stellar components are estimates obtained from a qualitative comparison of our NLTE models to the single SDSS spectrum. Detailed parameters for both stars need to be derived from a full two-component analysis of orbital phase resolved spectroscopy. The effective temperature in particular may be lower or higher by 20\,000\,K. The surface temperature of the companion's irradiated side was also constrained with \texttt{nightfall} simulations, see below. The overall shape of the observed spectrum is well fitted with the combination of a PG\,1159 star and a cool, irradiated companion, but especially the C\,{\sc iv} spectral lines of the PG\,1159 model atmosphere are not strong enough. There is another PG\,1159 star showing this phenomenon (H\"ugelmeyer \etal, in prep.), and also none of the deep absorption lines which some DO white dwarfs show can be fitted (e.g. Werner \etal 1995). The spectral signatures of an A star, as one would expect for the companion with 8\,200\,K surface temperature at the irradiated side, cannot be seen in the observation. This may be because the irradiation from the PG\,1159 leads to a temperature inversion in the upper layers of the companion's atmosphere up to $\tau_{\rm Ross}=1$, which causes the observed emission line spectrum (Barman \etal 2004). \section{Photometry of SDSS\,J212531.92$-$010745.9} \begin{table} \centering \caption{Observation log. All observations are performed with clear filter.} \begin{tabular}{rrrcc} \hline \hline \noalign{\smallskip} Date & t$_{\rm exp}[s]$& t$_{\rm cycle}[s]$ & Duration$[s]$ & Telescope\\ \noalign{\smallskip} \hline \noalign{\smallskip} 2005/09/21 & 90 & 98 & 18900 & 80\,cm\\ 2005/09/22 & 90 & 98 & 18899 & 80\,cm\\ 2005/09/23 & 90 & 98 & 21758 & 80\,cm\\ 2005/09/23 & 180 & 194& 14873 & 50\,cm\\ 2005/10/06 & 240 & 248& 10202 & 50\,cm\\ 2005/10/07 & 240 & 246& 14897 & 50\,cm\\ 2005/10/08 & 240 & 248& 9298 & 50\,cm\\ 2005/10/10 & 90 & 98 & 19852 & 80\,cm\\ 2005/10/11 & 240 & 248& 17872 & 50\,cm\\ 2005/10/18 & 90 & 98 & 16532 & 80\,cm\\ 2005/10/26 & 90 & 98 & 20095 & 80\,cm\\ \noalign{\smallskip} \hline \end{tabular}\label{tab_obs} \end{table} Photometric observations of SDSS\,J212531.92$-$010745.9 were performed during 10 nights (Tab.\,\ref{tab_obs}) using the T\"ubingen 80\,cm f/8 telescope with an \mbox{SBIG ST-7E} CCD camera and the G\"ottingen 50\,cm f/10 telescope with an \mbox{SBIG STL-6303E} CCD camera. To achieve good time resolution we chose clear filter exposures with a binning of 2x2 pixels to reduce readout time. The exposure time was t$_{\rm exp}$=90\,s for the observations with the 80\,cm telescope. In the case of the 50\,cm telescope, the exposure time was t$_{\rm exp}$=180\,s and t$_{\rm exp}$=240\,s. The observing conditions were good during the nights, considering that the telescopes are located in the cities of T\"ubingen and G\"ottingen. All images were bias and dark current corrected, then aperture photometry was performed using our IDL software TRIPP (Time Resolved Imaging Photometry Package, Schuh \etal 2003). The relative flux of the object was calculated with respect to the same two comparison stars (SDSS J212530.60-010921.0 and SDSS J212528.83-010828.5) for all nights, which were tested for stability. The resulting light curve is displayed in Fig.\,\ref{lightcurve}. To analyse the combined light curve of all nights, we used CAFE (Common Astronomical Fit Environment, G\"ohler, priv. comm.), a collection of routines written in IDL. The brightness variation is probably caused by a reflection effect. The companion is, due to the small separation, heated up on one side by irradiation from the PG\,1159 star, and the orbital motion then leads to a variable light curve. We fitted the combined light curve of all nights with a sine, achieving best results for a period of 6.95616(33)\,h (Fig.\,\ref{lightcurve}). The observed variability has a mean amplitude of 0.354(3)\,mag. To check if the observed light curve can be explained by a PG\,1159 star and an irradiated companion and for an impression of what the system geometry might look like we simulated the light curve of the binary system for an orbital period of 6.95616\,h with the program \texttt{nightfall}. Figure\,\ref{nightfall} shows the simulated and observed light curves of all nights, folded onto the orbital period. For the PG\,1159 star we assumed $T_{\rm eff}$\,=\,90\,000\,K, a mass of 0.6\,$\rm M_\odot$ and a radius of 0.1\,$\rm R_\odot$. For the companion we varied the mass from 0.1\,$\rm M_\odot$ to 0.7\,$\rm M_\odot$. We found that the observed light curve can be reproduced best with an M dwarf with an effective temperature of $3\,500\pm 150\,\rm K$, a mean radius of $0.4\pm 0.1\,\rm R_\odot$ and a mass of about $0.4\pm 0.1\,\rm M_\odot$. For the inclination of this system we obtained $70\pm 5\,^\circ$. Due to the irradiation by the PG\,1159 star the surface of the companion would be heated up to a surface temperature of 8\,200\,K, which, in combination with the PG\,1159 star, reproduces the overall shape of the observed spectrum quite well, as can be seen in Fig.\,\ref{spectrum}. The broad dip at the minimum of the light curve is well reproduced by this system configuration, too. In Table \ref{paras} we list all stellar and system parameters assumed and derived. We found that ellipsoidal variation due to geometrical deformation of the stars cannot generate the observed light curve. In the above configuration, calculated by nightfall according to the geometry in Djurasevic 1992, the equatorial radius of the M dwarf is only 4.5\,\% larger than its polar radius, and the PG\,1159 star is not affected by deformation above the numerical limit of \texttt{nightfall}. Because the object is positioned in the GW\,Vir instability strip (Fig.\,\ref{pg1159stars}), we also looked for pulsation periods below two hours in the light curve of SDSS\,J212531.92$-$010745.9. Therefore, we calculated a Lomb-Scargle periodogram (Scargle 1982) for the night with the best S/N (2005/10/26). But observational noise precludes detection of any periodicity with amplitudes below about 50\,mmag. The amplitudes of pulsating PG\,1159 stars normally are of the order of a few percent of a magnitude, and SDSS\,J212531.92$-$010745.9 is fainter than HE1429-1209, for which we recently discovered pulsation with the T\"ubingen 80\,cm telescope (Nagel \& Werner 2004). Our 80\,cm telescope might therefore just be too small to detect pulsation below 50\,mmag in this case. \begin{table} \centering \caption{Stellar and system parameters of SDSS\,J212531.92$-$010745.9, assumed (normal font) or derived from comparison with NLTE model spectra (boldface), photometric analysis (*) and nightfall simulation (italic).} \begin{tabular}{rrrr} \hline \hline \noalign{\smallskip} Parameter & PG\,1159 & Companion & System \\ \noalign{\smallskip} \hline \noalign{\smallskip} $T_{\rm eff}\,\,[\rm K]$ & {\bf $\sim$90\,000} & 3\,500$\pm$150 & \\ $T_{\rm eff,irr} \,\, [\rm K]$ & & {\it 8\,200} & \\ $\log g \,\,[\rm cm/s^2]$ & {\bf $\sim$7.6} & & \\ $m \,\,[\rm M_\odot]$ & 0.6 & {\it 0.4$\pm$0.1} & 1.0$\pm$0.1 \\ $r \,\,[\rm R_\odot]$ & 0.1 & {\it 0.4$\pm$0.1} & \\ $P_{\rm orb} \,\,[\rm h]$ & & & 6.95616(33) *\\ $\Delta m \,\,[\rm mag]$ & & & 0.354(3) * \\ $a \,\,[\rm R_\odot]$ & & & 1.85\\ $i \,\,[^\circ]$ & & & {\it 70$\pm$5}\\ \noalign{\smallskip} \hline \end{tabular}\label{paras} \end{table} \section{Conclusions} \begin{enumerate} \item The spectrum of SDSS\,J212531.92$-$010745.9 from DR4 of the Sloan Digital Sky Survey shows the signature of a PG\,1159 star plus emission from a cool irradiated companion. \item We performed time-series photometry during 10 nights with the T\"ubingen 80\,cm and the G\"ottingen 50\,cm telescopes and detected a period of 6.95616(33)\,h with an amplitude of 0.354(3)\,mag. This represents the orbital period of the binary system. Thus, SDSS\,J212531.92$-$010745.9 is the first close PG\,1159 binary without any doubts. \item From a first comparison with NLTE model spectra we derived, as preliminary results, an effective temperature of 90\,000\,K, $\log g\,\sim$\,7.60 and the abundance ratio C/He\,$\sim$\,0.05 for the PG\,1159 component. A detailed, quantitative NLTE spectral analysis of the PG\,1159 star and the irradiated companion has to be done next. We will report on the results in a subsequent paper. \item We simulated the light curve of the binary system with an orbital period of 6.95616\,h using \texttt{nightfall}. A good agreement with the observed light curve was obtained for a mean radius of $0.4\pm 0.1\,\rm R_\odot$, a mass of \,$0.4\pm 0.1\,\rm M_\odot$ and a temperature of the irradiated surface of about 8\,200\,K for the companion. \end{enumerate} To determine the system parameters more precisely, high-resolution phase-resolved spectroscopy of SDSS\,J212531.92$-$010745.9 is necessary. It should then be possible to derive both the companion's variable light contribution to the overall spectrum as well as dynamical masses for both components from radial velocity measurements of their distinct line systems. \begin{acknowledgements} We thank T.-O. Husser, R. Lutz and E. Nagel for supporting the observations. We acknowledge the use of {\texttt CAFE 5.1}, an astronomical fit enviroment, written by Eckart G\"ohler. We acknowledge the use of the \texttt{nightfall} program for the light curve synthesis of eclipsing binaries written by Rainer Wichmann (http://www.lsw.uni-heidelberg.de/$\sim$rwichman/Nightfall.html). BTG was supported by a PPARC Advanced Fellowship. \end{acknowledgements}
Title: Second-order perturbations of a zero-pressure cosmological medium: comoving vs. synchronous gauge
Abstract: Except for the presence of gravitational wave source term, the relativistic perturbation equations of a zero-pressure irrotational fluid in a flat Friedmann world model coincide exactly with the Newtonian ones to the second order in perturbations. Such a relativistic-Newtonian correspondence is available in a special gauge condition (the comoving gauge) in which all the variables are equivalently gauge invariant. In this work we compare our results with the ones in the synchronous gauge which has been used often in the literature. Although the final equations look simpler in the synchronous gauge, the variables have remnant gauge modes. Except for the presence of the gauge mode for the perturbed order variables, however, the equations in the synchronous gauge are gauge invariant and can be exactly identified as the Newtonian hydrodynamic equations in the Lagrangian frame. In this regard, the relativistic equations to the second order in the comoving gauge are the same as the Newtonian hydrodynamic equations in the Eulerian frame. We resolve several issues related to the two gauge conditions often to fully nonlinear orders in perturbations.
https://export.arxiv.org/pdf/astro-ph/0601041
\draft \twocolumn[\hsize\textwidth\columnwidth\hsize\csname @twocolumnfalse\endcsname \title{Second-order perturbations of a zero-pressure cosmological medium: \\ comoving vs. synchronous gauge} \author{Jai-chan Hwang${}^{(a)}$ and Hyerim Noh${}^{(b)}$} \address{${}^{(a)}$ Department of Astronomy and Atmospheric Sciences, Kyungpook National University, Taegu, Korea \\ ${}^{(b)}$ Korea Astronomy and Space Science Institute, Daejon, Korea \\ E-mails: ${}^{(a)}$jchan@knu.ac.kr, ${}^{(b)}$hr@kasi.re.kr } \vskip2pc] \section{Introduction} \label{sec:Introduction} The general relativistic cosmological linear perturbation theory was first developed by Lifshitz in 1946 \cite{Lifshitz-1946}. Lifshitz took the synchronous gauge condition in which the perturbations of the time-time part and the space-time part of the metric tensor are equal to zeros; this gauge condition can be taken to fully nonlinear order without losing any physical degree of freedom \cite{LL}. The synchronous gauge condition has been popular in the cosmological perturbation literature despite the complicating fact that, except for the zero-pressure case, there are remnant gauge modes for both the spatial and temporal gauge conditions. There exist other spatial and temporal gauge conditions which fix the gauge transformation property completely in general situation, thus without any remaining gauge mode \cite{Harrison-1967,Field-Shepley-1968,Nariai-1969}. This point was clarified by Bardeen \cite{Bardeen-1980,Bardeen-1988}. In a zero-pressure medium the density perturbation equation in the synchronous gauge coincides with the one in the comoving gauge \cite{Lifshitz-1946,Nariai-1969}. The density perturbation equation in the comoving gauge condition is known to resemble the Newtonian equation most closely \cite{Nariai-1969,Bardeen-1980}, and the equations coincide in the zero-pressure case \cite{Lifshitz-1946,Bonnor-1957}. Thus, in the zero-pressure case the density perturbation equation in the synchronous gauge coincides with the Newtonian one to the linear order \cite{Lifshitz-1946,Bonnor-1957}. The synchronous gauge was also used in the nonlinear perturbation studies \cite{Tomita}, and Kasai \cite{Kasai-1992} has derived second-order differential equations for density perturbation which is valid to fully nonlinear order. Although, such an equation in the synchronous gauge naturally has proper linear limit which corresponds to the Newtonian equation, it has been unclear whether such a correspondence continues to the nonlinear situation. Recently, we have successfully shown an exact relativistic-Newtonian correspondence of scalar-type perturbations to the second order based on the comoving gauge \cite{NL,second-order-CQG,second-order-PRD}. In the zero-pressure case our comoving gauge condition differs from the conventional synchronous gauge in the spatial gauge condition. In this work we will investigate the case in the original synchronous gauge. We will show that although the equations in the synchronous gauge look simpler than the ones in the comoving gauge, the variables still have remaining (spurious) spatial gauge mode to the second order. The equations in the synchronous gauge, however, are gauge invariant and can be identified as the Newtonian hydrodynamic equations in the Lagrangian frame. Whereas, the equations in the comoving gauge can be identified as the Newtonian hydrodynamic equations in the Eulerian frame. Results in Secs. \ref{sec:NL} and the Appendices are valid to fully nonlinear order in perturbations, and unless mentioned otherwise results in the remaining sections are valid to the second order in perturbations. We closely follow notations used in \cite{NL,second-order-CQG,second-order-PRD}. We set $c \equiv 1$. \section{Fully nonlinear perturbations} \label{sec:NL} The energy conservation equation and the Raychaudhury equation give \cite{covariant,second-order-CQG,second-order-PRD} \bea & & \tilde {\dot {\tilde \mu}} + \tilde \mu \tilde \theta = 0, \label{covariant-eq1} \\ & & \tilde {\dot {\tilde \theta}} + \frac{1}{3} \tilde \theta^2 + \tilde \sigma^{ab} \tilde \sigma_{ab} - \tilde \omega^{ab} \tilde \omega_{ab} + 4 \pi G \tilde \mu - \Lambda = 0, \label{covariant-eq2} \eea where $\tilde \theta \equiv \tilde u^a_{\;\; ;a}$ is an expansion scalar based on a fluid four-vector $\tilde u_a$; $\tilde \sigma_{ab}$ and $\tilde \omega_{ab}$ are the shear and the rotation tensors based on $\tilde u_a$, respectively; tildes indicate the covariant quantities and the Latin indices indicate spacetime components. An overdot with tilde is a covariant derivative along the $\tilde u_a$ vector, e.g., $\tilde {\dot {\tilde \mu}} \equiv \tilde \mu_{,a} \tilde u^a$. By combining these equations we have \bea & & \left( \frac{\tilde {\dot {\tilde \mu}}}{\tilde \mu} \right)^{\tilde \cdot} - \frac{1}{3} \left( \frac{\tilde {\dot {\tilde \mu}}}{\tilde \mu} \right)^2 - \tilde \sigma^{ab} \tilde \sigma_{ab} + \tilde \omega^{ab} \tilde \omega_{ab} - 4 \pi G \tilde \mu + \Lambda = 0. \nonumber \\ \label{covariant-eq3} \eea Equations (\ref{covariant-eq1})-(\ref{covariant-eq3}) are fully nonlinear and covariant, thus valid to all orders in perturbations. \subsection{Temporal Comoving Gauge} In this work we will {\it assume} an irrotational fluid, thus $\tilde \omega_{ab} \equiv 0$. We will consider two different gauge conditions. In both gauge conditions we will have \bea & & \tilde u_\alpha = 0, \eea due to a common temporal gauge condition together with the irrotational condition; the Greek indices indicate space components. If we introduce the spatial part of the four-vector as \bea & & \tilde u_\alpha \equiv a \left( - \hat v_{,\alpha} + \hat v^{(v)}_\alpha \right), \label{u-def} \eea where $\hat v^{(v)}_\alpha$ is a vector-type perturbation (thus transverse), the irrotational condition sets $\hat v^{(v)}_\alpha \equiv 0$ and our temporal comoving gauge sets $\hat v \equiv 0$. Since $\tilde u_\alpha = 0$ the fluid four-vector in this gauge coincides with the {\it normal} frame four-vector $\tilde n_a$ with $\tilde n_\alpha \equiv 0$. Notice that our temporal comoving gauge condition $\hat v \equiv 0$ (together with the irrotational condition) implies $\tilde u_\alpha = 0$. This {\it differs} from the ordinarily known {\it comoving} frame condition which sets $\tilde u^\alpha \equiv 0$ \cite{Taub-1978}. In our case the normalized ($\tilde u^a \tilde u_a \equiv -1$) fluid four-vector $\tilde u_a$ becomes \bea & & \tilde u_0 = - {1 \over \sqrt{- \tilde g^{00}}}, \quad \tilde u_\alpha \equiv 0; \nonumber \\ & & \tilde u^0 = \sqrt{- \tilde g^{00}}, \quad \tilde u^\alpha = - {\tilde g^{0\alpha} \over \sqrt{- \tilde g^{00}}}. \label{u} \eea In the zero-pressure case the momentum conservation equation implies $\tilde g^{00} = - 1/a^2$ where $a$ is the cosmic scale factor of the Friedmann background world model. In the ADM approach \cite{ADM}, our temporal comoving gauge $\hat v = 0$ together with the irrotational condition implies vanishing momentum vector $J_\alpha \equiv - \tilde n_b \tilde T^b_\alpha = 0$. The ADM momentum conservation equation in Eq.\ (13) of \cite{NL} gives $N_{,\alpha} = 0$ where $\tilde g^{00} \equiv -1/N^2$, thus $N = N(t)$. In another way, since the acceleration vector $\tilde a_\alpha \equiv \tilde u_{\alpha ;b} \tilde u^b = (\ln{N})_{,\alpha}$ vanishes (i.e., geodesic flow) for the zero-pressure irrotational flow, we have $N = N(t)$; see Eqs.\ (27) and (42) of \cite{NL}. Without losing generality we can set $N = a(t)$. Thus we have \bea & & \tilde g^{00} = - {1 \over a^2}. \label{g^00} \eea Thus, Eq.\ (\ref{u}) becomes \bea & & \tilde u_0 = - a, \quad \tilde u_\alpha \equiv 0; \quad \tilde u^0 = {1 \over a}, \quad \tilde u^\alpha = - a \tilde g^{0\alpha}, \label{u-pressureless} \eea which is valid to fully nonlinear order. We can show that to all orders in perturbations the fluid quantities are independent of the spatial gauge condition which could affect $\tilde g^{0\alpha}$; see the Appendix A. \subsection{Nonlinear perturbed equations} We introduce perturbations \bea & & \tilde \mu \equiv \mu + \delta \mu, \quad \tilde \theta \equiv 3 H - \kappa, \eea where $H \equiv \dot a/a$ and $\delta \equiv \delta \mu/\mu$; an overdot denotes a time derivative based on background proper-time $t$. The $\tilde \theta$ is an expansion scalar of the fluid four-vector which is the same as the normal four-vector because $\tilde u_\alpha = 0$ in our case. Using Eq.\ (\ref{u-pressureless}) we have \bea & & \tilde {\dot {\tilde \mu}} = \dot \mu \left( 1 + \delta \right) + \mu \left( \dot \delta - {1 \over a} N^\alpha \delta_{,\alpha} \right), \nonumber \\ & & \tilde {\dot {\tilde \theta}} = 3 \dot H - \left( \dot \kappa - {1 \over a} N^\alpha \kappa_{,\alpha} \right), \label{dot-mu} \eea where $N^\alpha$ is the shift vector in the ADM notation with $N^\alpha \equiv a^2 \tilde g^{0\alpha}$; the spatial indices of the ADM variables are based on $h_{\alpha\beta} \equiv \tilde g_{\alpha\beta}$. Equations (\ref{covariant-eq1}), (\ref{covariant-eq2}) give \bea & & \left( \dot \mu + 3 H \mu \right) \left( 1 + \delta \right) \nonumber \\ & & \quad + \mu \left[ \dot \delta - {1 \over a} \delta_{,\alpha} N^\alpha - \left( 1 + \delta \right) \kappa \right] = 0, \\ & & 3 \left( \dot H + H^2 \right) + 4 \pi G \mu - \Lambda \nonumber \\ & & \quad - \left[ \dot \kappa - {1 \over a} \kappa_{,\alpha} N^\alpha + 2 H \kappa - 4 \pi G \mu \delta - {1 \over 3} \kappa^2 - \tilde \sigma^{ab} \tilde \sigma_{ab} \right] \nonumber \\ & & \quad = 0. \eea The background parts give \bea & & \dot \mu + 3 H \mu = 0, \quad 3 \left( \dot H + H^2 \right) + 4 \pi G \mu - \Lambda = 0. \label{BG-eqs} \eea The perturbed parts give \bea & & \hat {\dot \delta} = \left( 1 + \delta \right)\kappa, \label{NL-eq1} \\ & & \hat {\dot \kappa} + 2 H \kappa = {1 \over 3} \kappa^2 + \tilde \sigma^{ab} \tilde \sigma_{ab} + 4 \pi G \mu \delta, \label{NL-eq2} \eea where $\hat {\dot \delta} \equiv \dot \delta - a^{-1} \delta_{,\alpha} N^\alpha$. By combining these equations we have \bea & & \hat {\ddot \delta} + 2 H \hat {\dot \delta} - 4 \pi G \mu \delta = 4 \pi G \mu \delta^2 + {4 \over 3} {(\hat {\dot \delta})^2 \over 1 + \delta} + \left( 1 + \delta \right) \tilde \sigma^{ab} \tilde \sigma_{ab}. \nonumber \\ \label{NL-eq3} \eea These equations are valid to the fully nonlinear orders in perturbations, subject only to the temporal comoving gauge condition, the zero-pressure condition, and the irrotational condition. \subsection{The synchronous gauge} Under the synchronous gauge we set $\tilde g_{0\alpha} \equiv 0$ (thus $N^\alpha \equiv 0$) using the spatial gauge condition (together with the irrotational condition), thus \bea & & \tilde {\dot {\tilde \mu}} = \hat {\dot {\tilde \mu}} = {\dot {\tilde \mu}}. \eea Thus, Eqs.\ (\ref{NL-eq1})-(\ref{NL-eq3}) simply give \bea & & \dot \delta = \left( 1 + \delta \right)\kappa, \label{NL-SG-eq1} \\ & & \dot \kappa + 2 H \kappa = {1 \over 3} \kappa^2 + \tilde \sigma^{ab} \tilde \sigma_{ab} + 4 \pi G \mu \delta, \label{NL-SG-eq2} \\ & & \ddot \delta + 2 H \dot \delta - 4 \pi G \mu \delta = 4 \pi G \mu \delta^2 + {4 \over 3} {\dot \delta^2 \over 1 + \delta} + \left( 1 + \delta \right) \tilde \sigma^{ab} \tilde \sigma_{ab}, \nonumber \\ \label{NL-SG-eq3} \eea which are valid to the fully nonlinear order. Using $\Delta \equiv \delta/(1 + \delta)$ Kasai \cite{Kasai-1992} has derived \bea & & \ddot \Delta + 2 H \dot \Delta - 4 \pi G \mu \Delta = - {2 \over 3} {\dot \Delta^2 \over 1 - \Delta} + \left( 1 - \Delta \right) \tilde \sigma^{ab} \tilde \sigma_{ab}. \nonumber \\ \eea To nonlinear order in perturbations the above equations are incomplete yet because of $\tilde \sigma^{ab} \tilde \sigma_{ab}$ term. Later we will show that these equations in the synchronous gauge differs from the equations in the comoving gauge to the second order. Furthermore, although these equations look simple, we will show that $\delta$ (thus $\Delta$ as well) and $\kappa$ still have remnant gauge modes to the second order. In Sec.\ \ref{sec:SG} we will show that to the second order the equations are gauge invariant and can be identified with the Newtonian hydrodynamic equations in the Lagrangian frame. In this regard, the equations in the comoving gauge correspond to the Newtonian hydrodynamic equations in the Eulerian frame. \section{Second-order perturbations} As the metric we take \bea & & ds^2 = - a^2 \left( 1 + 2 \alpha \right) d \eta^2 - 2 a^2 \beta_{,\alpha} d \eta d x^\alpha \nonumber \\ & & \quad + a^2 \left[ g^{(3)}_{\alpha\beta} \left( 1 + 2 \varphi \right) + 2 \gamma_{,\alpha|\beta} + 2 C^{(t)}_{\alpha\beta} \right] d x^\alpha d x^\beta, \label{metric} \eea where $\alpha$, $\beta$, $\gamma$ and $\varphi$ are spacetime dependent perturbed-order variables, and $C^{(t)}_{\alpha\beta}$ is a transverse and tracefree perturbed-order variable. Spatial indices of perturbed order variables are based on $g^{(3)}_{\alpha\beta}$, and a vertical bar indicates the covariant derivative based on $g^{(3)}_{\alpha\beta}$; $g^{(3)}_{\alpha\beta}$ could become $\delta_{\alpha\beta}$ in a flat Friedmann background. We {\it ignored} the transverse vector-type perturbation variables. We introduce $\chi \equiv a ( \beta + a \dot \gamma)$. The perturbed variables can be regarded as nonlinearly perturbed ones to any order in perturbations. To the second order, from Eqs.\ (55), (57) of \cite{NL} we have \bea & & N^\alpha = - \beta^{,\alpha}, \nonumber \\ & & \tilde \sigma^{ab} \tilde \sigma_{ab} = \bar K^\alpha_\beta \bar K^\beta_\alpha = {1 \over a^4} \left[ \chi^{,\alpha|\beta} \chi_{,\alpha|\beta} - {1 \over 3} \left( \Delta \chi \right)^2 \right] \nonumber \\ & & \quad + \dot C^{(t)\alpha\beta} \left( {2 \over a^2} \chi_{,\alpha|\beta} + \dot C^{(t)}_{\alpha\beta} \right), \label{sigma} \eea where $N^\alpha$ is evaluated to the linear order; $\bar K^\alpha_\beta$ is a tracefree part of the extrinsic curvature. We note that $\tilde \sigma^{ab} \tilde \sigma_{ab}$ is spatially gauge invariant to the second order, see Sec.\ \ref{sec:gauge-issue}. Before comparing equations in the two different spatial gauges, we {\it compare} our $\hat v$ in Eq.\ (\ref{u-def}) with the notation used in \cite{NL} to the second order in perturbations. In \cite{NL} we introduced the fluid quantities based on the normal-frame vector $\tilde n_a$ and provided the relation of fluid quantities between the energy-frame ($E$) and the normal-frame ($N$). Our fluid four-vector $\tilde u_a$ is based on the energy frame which sets $\tilde q_a \equiv 0$. The energy-frame fluid four-vector is introduced in Eq.\ (53) of \cite{NL}, and using the relations given in Eqs.\ (87), (88) of \cite{NL} we have \bea & & \tilde u_\alpha = a \left( V_\alpha^E - B_\alpha + A B_\alpha + 2 V^\beta_E C_{\alpha\beta} \right) \nonumber \\ & & \quad = a \left\{ {Q_\alpha^N \over \mu + p} - {1 \over (\mu + p)^2} \left[ \left( \delta \mu + \delta p \right) Q_\alpha^N + Q^\beta_N \Pi_{\alpha\beta} \right] \right\}. \nonumber \\ \eea Using the decomposition of the normal-frame flux vector $Q_\alpha^N \equiv (\mu + p) ( - v_{,\alpha} + v_\alpha^{(v)})$ in Eq.\ (175) of \cite{NL} and setting $v_\alpha^{(v)} \equiv 0$ we have \bea & & \hat v_{,\alpha} = v_{,\alpha} - {1 \over \mu + p} \left[ \left( \delta \mu + \delta p \right) v_{,\alpha} + v^{,\beta} \Pi_{\alpha\beta} \right]. \eea Thus, the temporal comoving gauge $v \equiv 0$ in \cite{NL} implies $\hat v = 0$ and vice versa. \subsection{The comoving gauge} \label{sec:CG} In \cite{NL,second-order-CQG,second-order-PRD} we took the temporal comoving gauge and the spatial $\gamma = 0$ gauge \bea & & v \equiv 0, \quad \gamma \equiv 0. \label{CG} \eea In this work, we call this the {\it comoving} gauge. Thus, we have $\beta = \chi/a$. The momentum conservation equation in Eq.\ (105) of \cite{NL} gives \bea & & \alpha = - {1 \over 2 a^2} \chi^{,\beta} \chi_{,\beta}. \label{alpha-CG} \eea Thus, apparently, $\alpha$ does not vanish to the second order. Later we will show that if we take $\beta = 0$ as the spatial gauge condition instead of $\gamma = 0$, we have vanishing $\alpha$. However, we prefer $\gamma \equiv 0$ as the spatial gauge condition because it fixes the spatial gauge degree of freedom completely (as long as we simultaneously take the temporal gauge which removes the temporal gauge degree of freedom completely, like our $v = 0$), see Sec.\ VI of \cite{NL}. Whereas, $\beta \equiv 0$ fails to fix the spatial gauge degree of freedom completely, thus having remaining gauge degree of freedom even after imposing the gauge condition, see Sec.\ \ref{sec:two-gauges}. In our gauge the fluid four-vector in Eq.\ (\ref{u-pressureless}) becomes \bea & & \tilde u_0 = - a , \quad \tilde u_\alpha = 0; \nonumber \\ & & \tilde u^0 = {1 \over a}, \quad \tilde u^\alpha = \frac{1}{a^2} \chi^{,\beta} \left[ \left( 1 - 2 \varphi \right) \delta^\alpha_\beta - 2 C^{(t)\alpha}_{\;\;\;\;\;\beta} \right]. \label{u-CG} \eea Thus, although we prefer to call this the temporal comoving gauge (see \cite{Bardeen-1980,Bardeen-1988}), because $\tilde u_\alpha = 0$ and $\tilde u^\alpha \neq 0$, our fluid four-vector corresponds to the normal four-vector rather than the comoving one. Using Eq.\ (\ref{sigma}), Eqs.\ (\ref{NL-eq1}), (\ref{NL-eq2}) give \bea & & \dot \delta + {1 \over a^2} \delta_{,\alpha} \chi^{,\alpha} - \kappa = \delta \kappa, \label{delta-eq-CG} \\ & & \dot \kappa + {1 \over a^2} \kappa_{,\alpha} \chi^{,\alpha} + 2 H \kappa - 4 \pi G \mu \delta \nonumber \\ & & \quad = \left( {1 \over a^2} \chi^{,\alpha|\beta} + \dot C^{(t)\alpha\beta} \right) \left( {1 \over a^2} \chi_{,\alpha|\beta} + \dot C^{(t)}_{\alpha\beta} \right). \label{kappa-eq-CG} \eea These also follow from the energy-conservation equation and the trace part of ADM propagation equation in Eqs.\ (104), (102) of \cite{NL}. In \cite{second-order-CQG,second-order-PRD} we {\it identified} to the second order \bea & & \delta \mu \equiv \delta \varrho, \quad \kappa \equiv - {1 \over a} \nabla \cdot {\bf u}, \label{identify-second-order} \eea where $\delta \varrho$ and ${\bf u}$ are Newtonian density and velocity perturbations, respectively. As we ignore the rotational mode, the velocity is of potential type with ${\bf u} = \nabla u$. Apparently, we need $\chi$ to the linear order only, and to that order we have \cite{second-order-CQG,second-order-PRD} \bea & & \nabla \chi = a {\bf u}, \label{identify-chi} \eea where we {\it assume} a flat Friedmann background world model. With these identifications of the relativistic metric and energy-momentum perturbation variables (these are equivalently gauge-invariant combinations, see Sec.\ \ref{sec:two-gauges}) with the Newtonian hydrodynamic variables, Eqs.\ (\ref{delta-eq-CG}), (\ref{kappa-eq-CG}) give \bea & & \dot \delta + {1 \over a} \nabla \cdot {\bf u} = - {1 \over a} \nabla \cdot \left( \delta {\bf u} \right), \label{delta-eq-3rd} \\ & & {1 \over a} \nabla \cdot \left( \dot {\bf u} + H {\bf u} \right) + 4 \pi G \mu \delta = - {1 \over a^2} \nabla \cdot \left( {\bf u} \cdot \nabla {\bf u} \right) \nonumber \\ & & \quad - \dot C^{(t)\alpha\beta} \left( {2 \over a} \nabla_\beta u_\alpha + \dot C^{(t)}_{\alpha\beta} \right). \label{u-eq-3rd} \eea By combining these we have \bea & & \ddot \delta + 2 {\dot a \over a} \dot \delta - 4 \pi G \mu \delta = - {1 \over a^2} {\partial \over \partial t} \left[ a \nabla \cdot \left( \delta {\bf u} \right) \right] + {1 \over a^2} \nabla \cdot \left( {\bf u} \cdot \nabla {\bf u} \right) \nonumber \\ & & \quad + \dot C^{(t)\alpha\beta} \left( {2 \over a} \nabla_\beta u_\alpha + \dot C^{(t)}_{\alpha\beta} \right), \label{density-eq-3rd} \eea which also follows from Eq.\ (\ref{NL-eq3}). Except for the presence of the gravitational waves as source terms Eqs.\ (\ref{delta-eq-3rd})-(\ref{density-eq-3rd}) are valid {\it exactly} in the Newtonian system \cite{Peebles-1980}. Although our relativistic equations are valid to the second order, the Newtonian equations are valid to fully nonlinear order. Thus, all nonvanishing higher-order perturbation terms in the relativistic case are pure general relativistic corrections. Recently, we have presented such pure general relativistic correction terms appearing in the third order perturbations in \cite{third-order}. \subsection{The synchronous gauge} \label{sec:SG} The synchronous and comoving gauge conditions correspond to taking \cite{LL} \bea & & v \equiv 0 , \quad \beta \equiv 0 ; \quad \alpha = 0. \label{SG} \eea In this work, we call this simply the {\it synchronous} gauge. Thus, we have $\dot \gamma = \chi/a^2$. If we take $v \equiv 0$ and $\beta \equiv 0$ as the temporal and the spatial gauge conditions, respectively, the momentum conservation equation gives $\alpha = 0$ to {\it all} orders in perturbations; although this is well known in \cite{LL}, we give proofs in the Appendix B. Thus, in the zero-pressure medium without rotation we can simultaneously impose the comoving ($v = 0$) and the synchronous ($\alpha = 0$) temporal gauge conditions as long as we also take $\beta \equiv 0$ as the spatial gauge condition \cite{LL}; Kasai took these conditions in his work in \cite{Kasai-1992}. The original synchronous gauge used by Lifshitz \cite{Lifshitz-1946} took $\alpha = 0$ and $\beta = 0$ as the temporal and the spatial gauge conditions, respectively. These gauge conditions are known to be {\it incomplete} in fixing both the temporal and the spatial gauge modes even to the linear order. Thus, even after imposing these gauge conditions we have remaining gauge modes present in the solutions, in general. Meanwhile, $v \equiv 0$ and $\gamma \equiv 0$ fix the temporal and spatial gauge degree of freedoms completely, thus no gauge mode is present in the solution, see Sec.\ \ref{sec:gauge-issue}. Since the original synchronous gauge implies $v = 0$ (the nonvanishing solution of $v$ is the remnant temporal gauge mode) in the zero-pressure case, we only have to pay attention to the possible presence of the spatial gauge mode. In this gauge we have $\tilde g_{00} = - a^2 = 1/\tilde g^{00}$ and $\tilde g_{0\alpha} = 0 = \tilde g^{0\alpha}$. Thus, the fluid four-vector in Eq.\ (\ref{u-pressureless}) becomes \bea & & \tilde u_0 = - a, \quad \tilde u_\alpha = 0; \quad \tilde u^0 = {1 \over a}, \quad \tilde u^\alpha = 0, \label{u-SG} \eea which can be compared with Eq.\ (\ref{u-CG}) in the comoving gauge. Thus, since $\tilde u^\alpha = 0$, our fluid four-vector corresponds to the conventionally known comoving four-vector \cite{Taub-1978}, and simultaneously normal because $\tilde u_\alpha = 0$ as well. All the statements in the above two paragraphs are valid for all perturbational orders. Using Eq.\ (\ref{sigma}), Eqs.\ (\ref{NL-SG-eq1}), (\ref{NL-SG-eq2}) give \bea & & \dot \delta - \kappa = \delta \kappa, \label{delta-eq-SG} \\ & & \dot \kappa + 2 H \kappa - 4 \pi G \mu \delta \nonumber \\ & & \quad = \left( {1 \over a^2} \chi^{,\alpha\beta} + \dot C^{(t)\alpha\beta} \right) \left( {1 \over a^2} \chi_{,\alpha\beta} + \dot C^{(t)}_{\alpha\beta} \right). \label{kappa-eq-SG} \eea These also follow from the energy-conservation equation and the trace part of ADM propagation equation in Eqs.\ (104), (102) of \cite{NL}. By combining these equations we have \bea & & \ddot \delta + 2 H \dot \delta - 4 \pi G \mu \delta = \dot \delta^2 + 4 \pi G \mu \delta^2 \nonumber \\ & & \quad + \left( {1 \over a^2} \chi^{,\alpha\beta} + \dot C^{(t)\alpha\beta} \right) \left( {1 \over a^2} \chi_{,\alpha\beta} + \dot C^{(t)}_{\alpha\beta} \right). \label{ddot-delta-eq-SG} \eea Apparently, these equations in the synchronous gauge look simpler than Eqs.\ (\ref{delta-eq-CG}), (\ref{kappa-eq-CG}) in the comoving gauge. Compared with Eqs.\ (\ref{delta-eq-CG}), (\ref{kappa-eq-CG}) in the comoving gauge, in Eqs.\ (\ref{delta-eq-SG}), (\ref{kappa-eq-SG}) we lack the convective-derivative-like terms in the left-hand-sides. By changing the time derivatives as \bea & & {\partial \over \partial t} \rightarrow {\partial \over \partial t} + {1 \over a^2} ( \nabla \chi ) \cdot \nabla = {\partial \over \partial t} + {1 \over a} {\bf u} \cdot \nabla, \label{time-derivative} \eea we can show that Eqs.\ (\ref{delta-eq-SG})-(\ref{ddot-delta-eq-SG}) become the same ones in the comoving gauge in Eqs.\ (\ref{delta-eq-CG}), (\ref{kappa-eq-CG}), and (\ref{density-eq-3rd}); in the last step of Eq.\ (\ref{time-derivative}) we used Eq.\ (\ref{identify-chi}) which is valid for a flat background. {\it If} we make the same identification of the density and velocity perturbations as in Eqs.\ (\ref{identify-second-order}), (\ref{identify-chi}), thus assuming a flat background, Eqs.\ (\ref{delta-eq-SG}), (\ref{kappa-eq-SG}) become: \bea & & \dot \delta + {1 \over a} \nabla \cdot {\bf u} = - {1 \over a} \delta \nabla \cdot {\bf u}, \label{delta-eq-SG2} \\ & & {1 \over a} \nabla \cdot \left( \dot {\bf u} + H {\bf u} \right) + 4 \pi G \mu \delta = - {1 \over a^2} \left( \nabla^\beta u^\alpha \right) \left( \nabla_\beta u_\alpha \right) \nonumber \\ & & \quad - \dot C^{(t)\alpha\beta} \left( {2 \over a} \nabla_\beta u_\alpha + \dot C^{(t)}_{\alpha\beta} \right). \label{kappa-eq-SG2} \eea Ignoring the gravitational waves, these equations can be identified as the Newtonian hydrodynamic equations in the Lagrangian frame. Although equations in the synchronous gauge look simpler than the ones in the comoving gauge, the presence of additional convective-like terms in the comoving gauge allows us to make exact (except for the gravitational waves) correspondence with the Newtonian hydrodynamic equations in the Eulerian frame \cite{second-order-CQG,second-order-PRD}. Whereas, the equations in the synchronous gauge can be identified as the Newtonian equations in the Lagrangian frame. However, the variables in the synchronous gauge still have the remnant spatial gauge mode due to incomplete fixing nature of the spatial gauge condition $\beta \equiv 0$ in that gauge. That is, to the second order, $\delta$ and $\kappa$ in the synchronous gauge have the remaining gauge modes, see Sec.\ \ref{sec:two-gauges}. Now, we can relate the variables in the synchronous ($S$) gauge to the ones in the comoving ($C$) gauge. {}From Eqs.\ (\ref{CG-SG}), (\ref{chi-GT}), and (\ref{identify-chi}) we have \bea & & \delta_S = \delta_C + \left( \int^t {1 \over a^2} \nabla \chi dt + \nabla \gamma_{S,{\rm Gauge}} \right) \cdot \nabla \delta_C, \nonumber \\ & & \kappa_S = \kappa_C + \left( \int^t {1 \over a^2} \nabla \chi dt + \nabla \gamma_{S,{\rm Gauge}} \right) \cdot \nabla \kappa_C, \label{SG-CG-sol} \eea where $\gamma_{S,{\rm Gauge}}$ is the gauge mode present to the linear order in $\gamma$; see the next section. In a flat background, from Eq.\ (\ref{identify-chi}) we have $\nabla \chi = a {\bf u}$. Notice that, even after ignoring the gauge modes $\delta_S$ and $\kappa_S$ naturally differ from $\delta_C$ and $\kappa_C$, respectively, because the final equations are different. Using Eq.\ (\ref{SG-CG-sol}), Eqs.\ (\ref{delta-eq-SG})-(\ref{ddot-delta-eq-SG}) give Eqs.\ (\ref{delta-eq-CG}), (\ref{kappa-eq-CG}), and (\ref{density-eq-3rd}). Although the variables in the synchronous gauge have remnant spatial gauge mode, somehow the equations in the synchronous gauge coincide with the Newtonian ones in the Lagrangian frame. Meanwhile, the Newtonian hydrodynamic equations have nothing to do with the gauge mode which appears only in the relativistic treatment. We can show that the situation is consistent in the synchronous gauge. {}From Eqs.\ (\ref{xi_alpha-SG}), (\ref{GT-SG}) the gauge mode of $\delta_{S,{\rm Gauge}} = \xi^\alpha \nabla_\alpha \delta_C$ is proportional to the linear-order solution of $\delta_C$; similarly, the gauge mode of $\kappa_{S,{\rm Gauge}} = \xi^\alpha \nabla_\alpha \kappa_C$ is proportional to the linear-order solution of $\kappa_C$. Thus, the behaviours of the gauge mode cannot be distinguished from the solutions to the linear order, and can be absorbed to the linear order solutions. We can also check that the gauge modes in Eq.\ (\ref{SG-CG-sol}) cancel out in Eqs.\ (\ref{delta-eq-SG}) and (\ref{kappa-eq-SG}). In this sense, Eqs.\ (\ref{delta-eq-SG}) and (\ref{kappa-eq-SG}), thus Eqs.\ (\ref{ddot-delta-eq-SG}), (\ref{delta-eq-SG2}) and (\ref{kappa-eq-SG2}) as well, are gauge-invariant. Therefore, to the second order in the synchronous gauge, although the variables have remnant gauge mode, the equations are gauge invariant; this happens because the gauge mode temporally behaves exactly like one of the physical solutions. A similar situation occurs to the linear order in the original synchronous gauge which took only $\alpha = 0 = \beta$ \cite{Lifshitz-1946}. Under these gauge conditions Lifshitz derived \bea & & \ddot \delta + 2 H \dot \delta - 4 \pi G \mu \delta = 0, \label{ddot-delta-eq-SG-lin} \eea which is the LHS of Eq.\ (\ref{ddot-delta-eq-SG}) and concides with the later derived Newtonian equation \cite{Bonnor-1957}. However, under these gauge conditions (i.e., without taking $v = 0$), $\delta$ still have the remnant gauge mode due to the incomplete fixing nature of the temporal gauge condition $\alpha \equiv 0$. It happens that the temporal behaviour of gauge mode of $\delta$ is proportional to $H$ which {\it coincides} with one of the two physical solutions, see \cite{GRG-1991}. Thus, although $\delta$ has the gauge mode Eq.\ (\ref{ddot-delta-eq-SG-lin}) is gauge invariant. In our synchronous gauge which also takes $v = 0$ the temporal gauge condition is fixed completely, but a similar situation repeats due to an incomplete fixing nature of the spatial synchronous gauge condition ($\beta \equiv 0$) now to the second order in perturbation. \section{Gauge issue} \label{sec:gauge-issue} \subsection{Gauge transformation} Under a transformation between two coordinates $\hat x^a = x^a + \tilde \xi^a$, the gauge transformation properties of all metric and energy-momentum variables to the second order are presented in Sec.\ VI of \cite{NL}. Since both the synchronous gauge and the comoving gauge take $v = 0$ we have \bea & & \tilde \xi^0 = 0. \eea This follows from Eqs.\ (234) or (238) of \cite{NL}: in the normal frame by setting $Q_\alpha = 0$ (i.e., $v = 0$) in both gauges we have $\tilde \xi^0_{\;\;,\alpha} = 0$; or in the energy frame, by setting $V_\alpha - B_\alpha + A B_\alpha + 2 V^\beta C_{\alpha\beta} = 0$ (i.e., $\hat v = 0$) in both gauges, we again have $\tilde \xi^0_{\;\;,\alpha} = 0$. Thus, without losing generality we can take $\tilde \xi^0 = 0$. To the second order, with $\tilde \xi^0 \equiv 0$, from Eqs.\ (229), (232) of \cite{NL} we have \bea & & \hat \alpha = \alpha - \alpha_{,\alpha} \xi^\alpha - \beta_{,\alpha} \xi^{\alpha\prime} - {1 \over 2} \xi^{\alpha\prime} \xi_\alpha^\prime, \nonumber \\ & & \hat \delta = \delta - \delta_{,\alpha} \xi^\alpha, \quad \hat \kappa = \kappa - \kappa_{,\alpha} \xi^\alpha, \label{GT1} \eea where $\xi^\alpha \equiv \tilde \xi^\alpha$ with the index of $\xi^\alpha$ based on $g^{(3)}_{\alpha\beta}$; a prime indicates the time derivative based on the conformal time $\eta$ with $d \eta \equiv d x^0 \equiv dt/a$. The gauge transformation property of $\kappa$ follows from the scalar nature of the expansion scalar $\tilde \theta$ with $\tilde \theta \equiv 3 H - \kappa$ where $\tilde \theta$ is based on the normal frame; for the gauge transformation of a scalar quantity, see Eq.\ (239) of \cite{NL}. To the linear order, from Eq.\ (252) of \cite{NL} we have \bea & & \hat \beta_{,\alpha} = \beta_{,\alpha} + \xi_\alpha^\prime, \quad \hat \gamma_{,\alpha} = \gamma_{,\alpha} - \xi_\alpha. \label{GT2} \eea Thus, $\chi \equiv a ( \beta + \gamma^\prime )$ is gauge invariant to the linear order, and \bea & & \alpha - \alpha_{,\alpha} \gamma^{,\alpha} + {1 \over 2} \beta_{,\alpha} \beta^{,\alpha}, \quad \delta - \delta_{,\alpha} \gamma^{,\alpha}, \quad \kappa - \kappa_{,\alpha} \gamma^{,\alpha}, \label{GI} \\ & & \alpha - \alpha_{,\alpha} \gamma^{,\alpha} - \left( \beta + {1 \over 2} \gamma^\prime \right)^{,\alpha} \gamma^\prime_{,\alpha}, \label{GI-alpha} \eea are gauge invariant to the second order. \subsection{Two gauges} \label{sec:two-gauges} In the comoving gauge, by imposing $\gamma \equiv 0$ in all coordinates (i.e., $\hat \gamma \equiv 0 \equiv \gamma$), from Eq.\ (\ref{GT2}) we have \bea & & \xi_\alpha = 0. \eea Thus, the spatial gauge transformation property is fixed completely. {}From Eq.\ (\ref{GT1}) we have \bea & & \hat \alpha = \alpha, \quad \hat \delta = \delta, \quad \hat \kappa = \kappa, \eea and each variable in this gauge has unique gauge-invariant counterpart as $\delta$ and $\kappa$ in Eq.\ (\ref{GI}) and $\alpha$ in Eq.\ (\ref{GI-alpha}). Thus, we can equivalently regard all variables in this gauge as (spatially and temporally) gauge invariant ones. {}For example, $\delta_{v,\gamma} \equiv \delta - \delta_{,\alpha} \gamma^{,\alpha}$ is a unique gauge invariant combination which is the same as $\delta$ in the $v = 0 = \gamma$ gauge conditions; for an explicit form of $\delta_{v,\gamma}$ including $v$, see Eq.\ (282) in \cite{NL}. We note that these results (i.e., values remain the same in the comoving gauge conditions, complete fixing of the gauge degrees of freedom, and presence of unique corresponding gauge-invariant variables) continue to be valid even in higher-order perturbations, \cite{NL}. Whereas, in the synchronous gauge, by imposing $\beta \equiv 0$ in all coordinates (i.e., $\hat \beta \equiv 0 \equiv \beta$), from Eq.\ (\ref{GT2}) we have \bea & & \xi_\alpha^\prime = 0. \eea Thus, even after imposing the gauge condition we have \bea & & \xi_\alpha = \xi_\alpha ({\bf x}), \label{xi_alpha-SG} \eea which is the remaining gauge mode. Thus, under the synchronous gauge, from Eqs.\ (\ref{GT1}), (\ref{GT2}) we still have \bea & & \hat \gamma_{,\alpha} = \gamma_{,\alpha} - \xi_\alpha, \nonumber \\ & & \hat \alpha = \alpha - \alpha_{,\alpha} \xi^\alpha, \quad \hat \delta = \delta - \delta_{,\alpha} \xi^\alpha, \quad \hat \kappa = \kappa - \kappa_{,\alpha} \xi^\alpha, \label{GT-SG} \eea where the transformation of $\gamma$ is valid to the linear order. In this sense variables in the synchronous gauge have remaining gauge modes even after imposing the gauge condition. $\gamma$ has the remaining spatial gauge mode even in the linear order, and the other variables have remaining gauge modes to the second order. \subsection{Transformation between the two gauges} Using the gauge transformation properties of the variables we can translate the equations and solutions in one gauge into the ones in another gauge condition. We indicate the comoving gauge and the synchronous gauge by subindices $C$ and $S$, respectively. To the linear order we have \bea & & \beta_C = \gamma_S^\prime = {1 \over a} \chi. \label{chi-GT} \eea We present three different ways to reach the transformation properties. {}First, we consider a transformation from the synchronous gauge (unhat) to the comoving gauge (hat). {}From Eq.\ (\ref{GT1}) we have \bea & & \alpha_C = - {1 \over 2} \xi^{\alpha\prime} \xi_\alpha^\prime, \nonumber \\ & & \delta_C = \delta_S - \delta_{S,\alpha} \xi^\alpha, \quad \kappa_C = \kappa_S - \kappa_{S,\alpha} \xi^\alpha. \label{GT-SG-1} \eea We need to determine the gauge transformation function $\xi^\alpha$, apparently, only to the linear order. {}From Eq.\ (\ref{GT2}) we have \bea & & \xi_\alpha = \gamma_{S,\alpha}. \eea Thus, Eq.\ (\ref{GT-SG-1}) becomes \bea & & \alpha_C = - {1 \over 2 a^2} \chi^{,\alpha} \chi_{,\alpha}, \nonumber \\ & & \delta_C = \delta_S - \delta_{S,\alpha} \gamma_S^{,\alpha}, \quad \kappa_C = \kappa_S - \kappa_{S,\alpha} \gamma_S^{,\alpha}, \label{CG-SG} \eea where we used Eq.\ (\ref{chi-GT}). Second, we consider a transformation from the comoving gauge (unhat) to the synchronous gauge (hat). {}From Eq.\ (\ref{GT1}) we have \bea & & \alpha_C = {1 \over 2} \xi^{\alpha\prime} \xi_\alpha^\prime + \beta_{C,\alpha} \xi^{\alpha\prime}, \nonumber \\ & & \delta_S = \delta_C - \delta_{C,\alpha} \xi^\alpha, \quad \kappa_S = \kappa_C - \kappa_{C,\alpha} \xi^\alpha. \label{GT-CG-1} \eea {}From Eq.\ (\ref{GT2}) we have \bea & & \xi_\alpha = - \gamma_{S,\alpha}. \eea Thus, Eq.\ (\ref{GT-CG-1}) leads to the same results in Eq.\ (\ref{CG-SG}). {}Finally, the gauge invariant combination in Eq.\ (\ref{GI}) provides a simpler derivation. {}From the gauge invariance of combinations in Eq.\ (\ref{GI}) we directly have Eq.\ (\ref{CG-SG}). Using these gauge transformation properties in Eq.\ (\ref{CG-SG}) we can derive Eqs.\ (\ref{delta-eq-SG}), (\ref{kappa-eq-SG}) from Eqs.\ (\ref{delta-eq-CG}), (\ref{kappa-eq-CG}), and vice versa. \section{Discussion} In this work we have compared the general relativistic weakly nonlinear cosmological perturbation equations in two different gauge conditions. In our previous works we have successfully shown that, except for the coupling with gravitational waves, the relativistic perturbation equations of a zero-pressure irrotational fluid coincide exactly with the Newtonian ones to the second order in perturbations. Such a relativistic-Newtonian correspondence was available in our special comoving gauge condition in which all the variables can be equivalently regarded as gauge invariant ones. In this work we have compared these results with the ones in the synchronous gauge. The case in the synchronous gauge was previously studied without noticing the similarity or difference of the equations with the Newtonian ones to the nonlinear orders. In this work we compared equations in the synchronous gauge with the ones in the comoving gauge and in the Newtonian case. Although the variables in this gauge have remnant spatial gauge modes due to the incomplete gauge fixing of the spatial gauge condition the equations are gauge invariant. Ignoring the gravitational waves, the equations in the synchronous gauge can be identified with the Newtonian hydrodynamic equations in the Lagrangian frame to the second order, whereas the equations in the comoving gauge can be identified as the Newtonian ones in the Eulerian frame. These Eulerian and Lagrangian correspondences can be understood because the fluid four-vector in our comoving gauge is in fact normal as in Eq.\ (\ref{u-CG}) whereas the four-vector in the synchronous gauge is both normal and comoving (thus Lagrangian) as in Eq.\ (\ref{u-SG}). In our way to clarify the case in the synchronous gauge we have addressed and resolved several issues related to the two gauge conditions often to fully nonlinear orders in perturbations. \subsection*{Acknowledgments} We thank Professors J. Richard Bond, Lev Kofman and Misao Sasaki for insightful and clarifying discussions. H.N. was supported by the Korea Research Foundation Grant No. R04-2003-10004-0. J.H. was supported by the Korea Research Foundation Grant No. 2003-015-C00253. \section*{Appendices} \subsection*{A. Invariance of fluid quantities} Here, we {\it show} that the fluid quantities based on the fluid four-vector in Eq.\ (\ref{u-pressureless}) do not depend on the choice of $\tilde g^{0\alpha}$ (the spatial gauge condition) to all orders in perturbations. Using a fluid four-vector $\tilde u_a$ the energy-momentum tensor is decomposed into fluid quantities as \cite{covariant,NL} \bea & & \tilde T_{ab} \equiv \tilde \mu \tilde u_a \tilde u_b + \tilde p \left( \tilde g_{ab} + \tilde u_a \tilde u_b \right) + \tilde q_a \tilde u_b + \tilde q_b \tilde u_a + \tilde \pi_{ab}; \label{Tab} \\ & & \tilde \mu \equiv \tilde T_{ab} \tilde u^a \tilde u^b, \quad \tilde p \equiv {1 \over 3} \tilde T_{ab} \tilde h^{ab}, \quad \tilde q_a \equiv - \tilde T_{cd} \tilde u^c \tilde h_a^d, \nonumber \\ & & \tilde \pi_{ab} \equiv \tilde T_{cd} \tilde h_a^c \tilde h_b^d - \tilde p \tilde h_{ab}, \label{fluid-Tab} \eea where $\tilde h_{ab} \equiv \tilde g_{ab} + \tilde u_a \tilde u_b$; we have $\tilde u^a \tilde q_a \equiv 0 \equiv \tilde u^a \tilde \pi_{ab}$, $\tilde \pi_{ab} \equiv \tilde \pi_{ba}$, and $\tilde \pi^a_a \equiv 0$. The variables $\tilde \mu$, $\tilde p$, $\tilde q_a$ and $\tilde \pi_{ab}$ are the energy density, the isotropic pressure (including the entropic one), the energy flux and the anisotropic pressure (stress) based on the fluid four-vector, respectively. Let us introduce another four-vector $\tilde U_a$ with \bea & & \tilde U_0 = - a, \quad \tilde U_\alpha \equiv 0; \quad \tilde U^0 = {1 \over a}, \quad \tilde U^\alpha = - a \tilde g^{0\alpha}_U. \label{u-second} \eea Thus, $\tilde U_a$ is subject to the same conditions as $\tilde u_a$ in Eq.\ (\ref{u-pressureless}), but with possibly different spatial gauge condition which could lead to $\tilde g^{0\alpha}_U \neq \tilde g^{0\alpha}$. The fluid quantities based on $\tilde U_a$ are similarly defined as in Eqs.\ (\ref{Tab}), (\ref{fluid-Tab}) with $\tilde U_a$ replacing $\tilde u_a$: for example, we have $\tilde \mu^U \equiv \tilde T_{ab} \tilde U^a \tilde U^b$, etc. We can easily show that if $\tilde p = \tilde q_a = \tilde \pi_{ab} = 0$ we have \bea & & \tilde \mu^U \equiv \tilde T_{ab} \tilde U^a \tilde U^b = \tilde \mu \tilde u_a \tilde u_b \tilde U^a \tilde U^b = \tilde \mu \tilde u_0 \tilde u_0 \tilde U^0 \tilde U^0 = \tilde \mu, \eea and $\tilde p^U = \tilde q_a^U = \tilde \pi_{ab}^U = 0$, and vice versa. This result is also valid to fully nonlinear order. \subsection*{B. Justification of Eq.\ (\ref{SG})} Here, we {\it show} that in a zero-pressure irrotational medium we can take the original synchronous gauge ($\alpha \equiv 0 \equiv \beta$) together with the temporal comoving gauge ($v \equiv 0$) simultaneously to all orders in perturbations. This was known in \cite{LL}, see Sec.\ 97 in \cite{LL}. Here, it is important to take $\beta \equiv 0$ as the spatial synchronous gauge although we prefer to take $\gamma \equiv 0$ as the spatial gauge condition because of the remnant gauge mode in the $\beta \equiv 0$ case. We provide two different proofs based on the ADM and the covariant formulations. We begin by taking $v \equiv 0$ and $\beta \equiv 0$ as the temporal and spatial gauge conditions, respectively. In Eq.\ (\ref{g^00}) we showed that $\tilde g^{00} = - 1/N^2 = - 1/a^2$. The spatial gauge condition $\beta = 0$ together with the irrotational condition implies $\tilde g_{0\alpha} \equiv N_\alpha = 0$. Thus, from Eq.\ (2) of \cite{NL} we have $\tilde g_{00} \equiv - a^2 ( 1 + 2 \alpha ) = - N^2 = - a^2$. This implies that we have $\alpha = 0$. Now, in the covariant approach, $v = 0$ and irrotational conditions imply $\tilde u_\alpha = 0$. Since $\tilde u_a$ is the fluid four-vector we take the energy frame, $\tilde q_a \equiv 0$. The momentum conservation equation in Eq.\ (27) of \cite{NL} gives $\tilde a_a = 0$. The spatial gauge condition $\beta = 0$ together with the irrotational condition implies $\tilde g_{0\alpha} = 0$, thus $\tilde g^{0\alpha} = 0$ as well. Since $\tilde a_\alpha \equiv \tilde u_{\alpha;b} \tilde u^b = \tilde \Gamma^0_{0\alpha} = {1 \over 2} \tilde g^{00} \tilde g_{00,\alpha}$, $\tilde a_\alpha = 0$ implies that $\tilde g_{00}$ is a function of time only. Thus, we have $\alpha = 0$. If we impose $\alpha \equiv 0$ and $\beta \equiv 0$ as the gauge condition, instead, we have non-vanishing $J_\alpha$ or $\tilde u_\alpha$, thus nonvanishing $v$. In a zero-pressure medium this nonvanishing $v$ can be identified as the remnant temporal gauge mode, which can be set equal to zero without losing physical degree of freedom.
Title: Bias-free Measurement of Giant Molecular Cloud Properties
Abstract: (abridged) We review methods for measuring the sizes, line widths, and luminosities of giant molecular clouds (GMCs) in molecular-line data cubes with low resolution and sensitivity. We find that moment methods are robust and sensitive -- making full use of both position and intensity information -- and we recommend a standard method to measure the position angle, major and minor axis sizes, line width, and luminosity using moment methods. Without corrections for the effects of beam convolution and sensitivity to GMC properties, the resulting properties may be severely biased. This is particularly true for extragalactic observations, where resolution and sensitivity effects often bias measured values by 40% or more. We correct for finite spatial and spectral resolutions with a simple deconvolution and we correct for sensitivity biases by extrapolating properties of a GMC to those we would expect to measure with perfect sensitivity. The resulting method recovers the properties of a GMC to within 10% over a large range of resolutions and sensitivities, provided the clouds are marginally resolved with a peak signal-to-noise ratio greater than 10. We note that interferometers systematically underestimate cloud properties, particularly the flux from a cloud. The degree of bias depends on the sensitivity of the observations and the (u,v) coverage of the observations. In the Appendix to the paper we present a conservative, new decomposition algorithm for identifying GMCs in molecular-line observations. This algorithm treats the data in physical rather than observational units, does not produce spurious clouds in the presence of noise, and is sensitive to a range of morphologies. As a result, the output of this decomposition should be directly comparable among disparate data sets.
https://export.arxiv.org/pdf/astro-ph/0601706
\title{Bias-free Measurement of Giant Molecular Cloud Properties} \author{Erik Rosolowsky\altaffilmark{1}} \affil{Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, MS-66, Cambridge, MA 02138} \email{erosolow@cfa.harvard.edu} \and \author{Adam Leroy} \affil{Department of Astronomy, 601 Campbell Hall, University of California at Berkeley, CA 94720} \email{aleroy@astro.berkeley.edu} \altaffiltext{1}{National Science Foundation Astronomy \& Astrophysics Postdoctoral Fellow} \keywords{ISM:clouds --- methods:data analysis --- radio lines:ISM} \section{Introduction} Over the last 15 years, it has become possible to observe molecular emission in nearby galaxies with sufficient resolution and sensitivity to distinguish individual giant molecular clouds (GMCs). The immediate goal of such studies is to determine whether (and how) the GMCs in other galaxies differ from those seen in the Solar neighborhood. The most common method used to address this question has been to use molecular-line tracers of H$_2$, in particular $^{12}$CO($1\to 0$), to compare the macroscopic properties (size, line width, and luminosity) of GMCs in other galaxies to those of Milky Way GMCs. Unfortunately, a wide variety of methods have been used to reduce data from spectral line data cubes into macroscopic GMC properties. As a result, many of the differences between GMC populations found in the literature can be attributed, at least partially, to observational artifacts or methodological differences. It is therefore difficult to assess what the real differences between GMC populations are based on the reported data in the literature. For GMCs that are either marginally resolved or marginally detected, observational biases can be severe. Figure \ref{RESBIAS} shows the variation of the measured spatial size and line width with the resolution for a model cloud. Typical Galactic GMCs have sizes of a few 10s of parsecs, comparable to the spatial resolution of many data sets used to study extragalactic GMCs \citep[e.g.~][]{vog87,ws90,wr91,is93,wil93,fu99,she00,ros03}. Figure \ref{RESBIAS} shows that when the size of the beam is comparable to the size of the object, the measured size is much higher than the true size of the object. Millimeter spectrometers and correlators often have excellent frequency resolution, so the spectral resolution bias is usually less important for GMC studies, but it can become substantial when data are binned to increase signal-to-noise. Figure \ref{EXTRAPMOMS} shows that the measured spatial size, line width, and flux of a real GMC in M~33 \citep[EPRB1 from][]{ros03} are all strong functions of the sensitivity of the data. We discuss another major source of bias, the method by which emission is decomposed into GMCs, in the Appendix. To place these biases in the the context of real molecular cloud studies, Figure \ref{DISTCOMP} shows the peak flux and angular size of typical GMCs \citep[those of][]{srby87} as a function of distance. The sensitivities and resolutions of a representative sample of molecular cloud studies have been indicated as horizontal lines in these plots. Distances to commonly observed objects have also been labeled. Figure \ref{DISTCOMP} demonstrates that most studies of extragalactic GMCs are conducted where the clouds of interest are only marginally resolved and are found at low sensitivity. Even future observations of GMCs in the Virgo cluster using the Atacama Large Millimeter Array (ALMA) will be affected by resolution and sensitivity biases. In this paper, we examine the effects of biases stemming from finite spatial resolution, spectral resolution, and sensitivity in molecular-line observations of GMCs. We recommend data analysis methods to produce a standardized set of observed cloud properties that account for these biases. Most of the methods used in this paper have been adopted piecemeal and {\it ad hoc} in previous studies. Here, we endeavor to justify our choice of methodologies and to synthesize various author's techniques for approaching the problems of molecular cloud data analysis. In Section \ref{MOMENTS}, we describe a standardized method to measure three basic properties of an emission distribution --- size, line width, and flux --- while accounting for the sensitivity and resolution of a data set. In Section \ref{PHYSICAL}, we discuss how these measurements can be transformed into physical quantities --- radius, line width, luminosity, and implied mass. Finally, we consider the effects of using interferometers to derive cloud properties in Section \ref{INTERF}. The results in these three sections are applicable to all observations of molecular clouds. In contrast, the methods used for decomposing emission into molecular clouds vary widely, and there is little basis for favoring one method over another in all cases. Hence, we defer a presentation of our decomposition algorithm to the Appendix of the paper, leaving only a brief discussion of the decomposition problem in Section \ref{DECOMPSECT}. We conclude the paper by exhibiting several examples of the application of our standardized methods to previously observed data and making recommendations for future observations (Section \ref{APPLICATIONS}). The methods described in this paper require a computer program to apply. A documented software version of the decomposition and measurement algorithms is available from the authors as an IDL package. \section{Measuring Molecular Cloud Properties \label{MOMENTS}} This section describes how to derive the spatial size, line width, and flux from a region of emission within a spectral line map (a ``data cube'') while accounting for the finite sensitivity and resolution of the data set. We use moment methods \citep[e.g.~][]{srby87}, which make full use of position and intensity information without assuming a functional form for the cloud. They are therefore robust to pathologies in the data {\it within} the cloud. Moments are, however, sensitive to the inclusion of false emission (noise) at the edge of a cloud. Including noise has the effect of artificially increasing the values of the moment. Therefore the methods outlined here should be employed in conjunction with careful signal identification so that the calculations include as little noise as possible. We assume throughout this section that the algorithm is being applied to a distribution of real emission that we label a ``cloud'' (we discuss signal identification and decomposition in the Appendix). \subsection{Moment Measurements of Size, Line Width, and Flux} This subsection describes how to apply moment methods to derive the size, line width, and flux from a distribution of emission (a ``cloud'') within a position-position-velocity data cube. The data cube consists of a number of pixels that have sizes of $\delta x$, $\delta y$, and $\delta v$ in the two spatial dimensions and the velocity dimension, respectively. The $i$th pixel in the data cube has positions $x_i$ and $y_i$, velocity $v_i$, and brightness temperature $T_i$. We assume that the cloud is contiguous and bordered by an isosurface in brightness temperature of value $T_{edge}$, so that all of the pixels in the cloud have $T > T_{edge}$ and the pixels outside the cloud have $T < T_{edge}$ or are separated from the cloud by emission with $T < T_{edge}$. We begin by rotating the spatial axes so that the $x$ and $y$ axes align with the major and minor axis of the cloud, respectively. We determine the orientation of the major axis using principal component analysis. We find the eigenvectors of the intensity-weighted covariance matrix for the cloud, \begin{equation} \nonumber \frac{1}{\sum_i T_i} \left[\begin{array}{ll} \sum_i T_i \left( x_i - \bar{x} \right)^2 & \sum_i T_i \left( x_i - \bar{x} \right) \left( y_i - \bar{y} \right) \\ \sum_i T_i \left( x_i - \bar{x} \right) \left( y_i - \bar{y} \right) & \sum_i T_i \left( y_i - \bar{y} \right)^2 \\ \end{array}\right] \mbox{ .} \end{equation} \noindent In the equations above the sum $\sum_{i}$ runs over all pixels within the cloud and $\bar{x}$ and $\bar{y}$ are the intensity weighted mean positions within the cloud (defined below). We define the new $x$ axis to lie along the eigenvector with the largest eigenvalue --- the major axis of the cloud. The $y$ axis lies perpendicular to the $x$ axis along the minor axis of the cloud. In the discussion below, $x$ refers to position along the major axis and $y$ refers to position along the minor axis. Rotating the axes in this manner yields information about the axis ratio of the cloud and allows a more careful deconvolution. This method for determining the position angle of molecular clouds has also been adopted by \citet{koda}. To measure of the size of the cloud, we compute the geometric mean of the second spatial moments along the major and minor axis. This is $\sigma_{r}$, the root-mean-squared (RMS) spatial size: \begin{equation} \sigma_{r} (T_{edge}) = \sqrt{\sigma_{maj} (T_{edge})~\sigma_{min} (T_{edge})} \end{equation} \noindent where $\sigma_{maj} (T_{edge})$ and $\sigma_{min} (T_{edge})$ are the RMS sizes (second moments) of the intensity distribution along the two spatial dimensions. We adopt this particular functional form since it has been used in previous observational studies \citep{srby87} and explored in depth by \citet{bert92} with respect to inclination, aspect ratio, and virialization. We calculate $\sigma_{maj} (T_{edge})$ and $\sigma_{min} (T_{edge})$ by: \begin{eqnarray} \sigma_{maj} (T_{edge}) &= &\sqrt{\frac{\sum_{i}^{cloud} T_i \left[ x_i - \bar{x} (T_{edge}) \right]^2}{\sum_{i}^{cloud} T_i}}, \mbox{ where} \\ \bar{x} (T_{edge}) &= &\frac{\sum_{i}^{cloud} T_i x_i}{\sum_{i}^{cloud} T_i} \mbox{~and} \\ \sigma_{min} (T_{edge}) & = & \sqrt{\frac{\sum_{i}^{cloud} T_i \left[ y_i - \bar{y} (T_{edge}) \right]^2}{\sum_{i}^{cloud} T_i}} \mbox{, where} \\ \bar{y} (T_{edge}) & = & \frac{\sum_{i}^{cloud} T_i y_i}{\sum_{i}^{cloud} T_i} \mbox{.} \end{eqnarray} \noindent In the equations above the sum $\sum_{i}^{cloud}$ runs over all pixels within the cloud. We have written each of the moments as a function of $T_{edge}$ because changing the isosurface that defines the boundary of the cloud ($T_{edge}$) will change the set of pixels included in the sum and therefore the values of the moments. Note that $\sigma_{r}$ is not the RMS distance ($d = \sqrt{x^2 + y^2}$) from the center of the cloud. Rather it is the analogous to the RMS size of the cloud along an arbitrarily chosen axis. Thus, $\sigma_{r} = \sigma_{x} = \sigma_{y}$ for a perfectly round cloud, while the RMS distance from the center for such a distribution is larger, $\sigma_{d} = \sqrt{2} \sigma_{x} > \sigma_{x}$. Also note that $\sigma_{maj} / \sigma_{min}$ is the axis ratio of the cloud and will be $\sim 1$ for round clouds and $\gg1$ for elongated or filamentary clouds. We calculate the velocity dispersion, $\sigma_v(T_{edge})$ in the same manner as the size: \begin{eqnarray} \sigma_v (T_{edge})& = &\sqrt{\frac{\sum_{i}^{cloud} T_i \left[ v_i - \bar{v} (T_{edge}) \right]^2}{\sum_{i}^{cloud} T_i}} \mbox{, where} \\ \bar{v} (T_{edge})& = & \frac{\sum_{i}^{cloud} T_i v_i}{\sum_{i}^{cloud} T_i} \mbox{ .} \end{eqnarray} \noindent The sums again run over all emission in the cloud. For a Gaussian line profile, such as that found for most clouds, the full-width half-maximum (FWHM) line width, $\Delta V (T_{edge})$ will be related to $\sigma_v (T_{edge})$ by \begin{equation} \Delta V (T_{edge}) =\sqrt{8\ln(2)}~\sigma_v (T_{edge}) \mbox{.} \end{equation} \noindent Finally, we calculate the flux of the cloud, $F_{\mathrm{CO}} (T_{edge})$ using the zeroth moment: \begin{equation} F_{\mathrm{CO}} (T_{edge}) = \sum_i T_i~\delta v~\delta x~\delta y \mbox{.} \end{equation} \noindent If $\delta x$ and $\delta y$ are in units of arcseconds, $\delta v$ in km s$^{-1}$, and $T_i$ in K, then the resulting flux will have units of K km s$^{-1}$ arcsecond$^{2}$. \subsection{Correcting for the Sensitivity Bias} \label{EXTRAPSECT} The sensitivity of a dataset influences the cloud properties derived from that data, a fact that we have emphasized in the previous section by explicitly writing the moments as functions of $T_{edge}$, the cloud boundary (usually set by the signal-to-noise ratio of the data). Figure \ref{EXTRAPMOMS} shows the variation of spatial size, line width, and flux as a function of sensitivity for a bright cloud in M33. The data for this cloud shows a substantial {\em sensitivity bias}; all of the derived properties are strong functions of the boundary isosurface ($T_{edge}$). In order to compare data sets with different sensitivities, one must account for this bias. In this section we describe a method to do this by extrapolating the measured properties of a cloud---$\sigma_{maj} (T_{edge})$, $\sigma_{min} (T_{edge})$, $\sigma_v (T_{edge})$, and $F_{\mathrm{CO}} (T_{edge})$---to those we would expect to measure for a cloud within a boundary isosurface of $T_{edge} = 0$ Kelvin (i.e., perfect sensitivity). We estimate the values of the moments at $T_{edge}=0$~K by extrapolating from higher values of $T_{edge}$. This technique was originally suggested for inferring total cloud areas by \citet{bt80} and applied to molecular cloud properties by \citet{syscw}. We calculate each of the moments for a sample of boundary temperatures, $T_{edge}$, ranging from near the peak temperature of the cloud to the lowest boundaries allowed by the data. Thus, we measure the variations of the moments as a function of the boundary temperature, $T_{edge}$, within the cloud (this is how we constructed the plot shown in Figure \ref{EXTRAPMOMS}). Below, we assume that we have measured each of the four moments for values of $T_{edge}$ ranging from $T_{min}$ (the minimum allowed by the data, that is the sensitivity limit) to $T_{max}$ (near the peak temperature of the cloud). We estimate the value of the moments at $T_{edge} = 0$~K by performing a weighted, linear least-squares fit to the measured moments. As an example, we consider $\sigma_{maj} (T_{edge})$. The data are modeled as \begin{equation} \sigma_{maj} (T_{edge}) = m~T_{edge} + \sigma_{maj} (0~\mbox{K}) \end{equation} and the fit determines the extrapolated moment, $\sigma_{maj} (0~\mbox{K})$. For the fit, each pair of data $\{T_{edge}, \sigma_{maj}(T_{edge})\}$ is assigned a weight proportional to the number of data in the cloud with $T>T_{edge}$, so that measurements of the moment using more data are weighted more heavily. Practically, this means that points to the left in Figure \ref{EXTRAPMOMS} have higher weights than those to the right. We use this linear extrapolation for $\sigma_{maj}$, $\sigma_{min}$, and $\sigma_v$, but we find that a quadratic extrapolation (including a $T_{edge}^2$ term) gives better results for the zeroth moment, $F_{\mathrm{CO}}$ (though we revert to a linear extrapolation when the extrapolated flux is lower than the measured flux). We plot the flux of a Gaussian cloud as a function of $T_{edge}$ in Figure \ref{FLUXFIG} for an uncorrected zeroth moment and the linear and quadratic extrapolations to illustrate the appropriateness of the quadratic fit to the zeroth moment. At very low sensitivities (signal-to-noise ratios near unity), the quadratic extrapolation is very noisy, but for signal-to-noise ratios of $2$ or better it does a dramatically better job of recovering the true flux of the cloud ($F/F_0 = 1$) than either the linear extrapolation or no extrapolation. Figure \ref{EXTRAPMOMS} also shows these extrapolations for a bright cloud in M~33. The result of this extrapolation is a set of four moments --- $\sigma_{maj} (0~\mathrm{K})$, $\sigma_{min} (0 ~\mathrm{K})$, $\sigma_v (0~\mathrm{K})$, and $F_{\mathrm{CO}} (0 ~\mathrm{K})$ --- that correspond to those we would expect to measure given infinite sensitivity. The values of these moments should be directly comparable even among datasets with different sensitivities (values of $T_{min}$). Note that diffuse emission surrounding a GMC may confuse this method. If one data set is measured with sensitivity sufficient to detect diffuse emission surrounding a GMC, while another lacks the sensitivity to do so then this approach may not be sufficient to correct for the sensitivity bias. This problem may be particularly acute when comparing Galactic GMCs observed with very good sensitivity to extragalactic clouds with worse signal-to-noise ratios. Interferometric data ``resolves out'' emission significantly more extended than the synthesized beam (see \S\ref{INTERF}), representing another bias against detecting diffuse emission. \citet[][]{polk}, \citet[][]{blitz85}, \citet[][]{ros03}, and \citet[][]{ler05} find evidence for diffuse emission surrounding GMCs in the Milky Way and the Local Group Galaxies M~31, M~33, and IC~10, respectively. \subsection{Correcting for the Resolution Bias} \label{BEAMDCSECT} Any astronomical data set represents the convolution of the intensity of the source with the profile of the instrument used to observe it. Care must therefore be taken in measuring sizes and line widths when the extent of the intensity distribution is comparable to the instrumental profile. In a typical spectral line data cube two profiles are important: the spatial beam and the width of a velocity channel. In this section, we describe simple corrections to account for the effects of finite spatial and spectral resolution. We ``deconvolve'' the spatial beam from the measured cloud size by subtracting the RMS beam size, $\sigma_{beam}$, from the extrapolated spatial moments, $\sigma_{maj} (T_{edge} = 0~\mathrm{K})$ and $\sigma_{min} (T_{edge} = 0~\mathrm{K})$, in quadrature --- an approach that is exact for Gaussians. The deconvolved second moment is given by: \begin{equation} \sigma_{r,dc} = \sqrt{ [ \sigma_{maj}^2 \left(0~\mathrm{ K} \right) - \sigma_{beam}^2 ]^{1/2} ~ [ \sigma_{min}^2 \left(0~\mathrm{ K} \right) - \sigma_{beam}^2 ]^{1/2}}~\mbox{ ,} \end{equation} \noindent where $\sigma_{maj} \left(0~\mathrm{ K} \right)$ and $\sigma_{min} \left(0~\mathrm{K} \right)$ are extrapolated to the 0 Kelvin isosurface as described in \S\ref{EXTRAPSECT}. This extrapolation is necessary to make this deconvolution valid: subtracting the full $\sigma_{beam}$ from the spatial moment measured for only part of the cloud will lead to an overcorrection and thus to an underestimate of the cloud size. Measuring the spatial size along the minor axis is also necessary to ensure that the cloud is indeed resolved in all dimensions. This is an advantage of the choice of axes (major/minor) described above. With sufficient signal-to-noise, this method of deconvolution provides a robust measurement of cloud size even for marginally resolved clouds. Instrumental resolution also affects the measured line width. Spectrometers measure the average intensity across a channel, rather than sampling the intensity at the center (nominal frequency) of that channel. When the width of the spectral line under consideration is comparable to the bandwidth of a single channel, the line strength varies significantly across an individual channel. In this case, the average value may differ substantially from the value at line center. The output of the spectrometer is thus a convolution of the true spectral profile with the profile of an individual channel. We account for this potential bias towards higher line widths by a simple deconvolution of the channel width from the extrapolated second moment: \begin{equation} \sigma_{v,dc}=\sqrt{{\sigma_{v}^2 \left(0~\mathrm{ K}\right)-\frac{\Delta V^2_{chan}}{2~\pi}}} \end{equation} where $\sigma_{v} \left(0~\mathrm{ K} \right)$ is the second moment of the cloud in the $v$ dimension extrapolated to 0 Kelvin as described in \S \ref{EXTRAPSECT} and $\Delta V_{chan}$ is the width of a velocity resolution element. Although the channel profiles are usually square in shape and not Gaussian, we simplify the deconvolution by approximating the channel shape as a Gaussian with integrated area equal to that of a square channel with width $\Delta V_{chan}$. For such a Gaussian, $\sigma_{chan}=\Delta V_{chan}/\sqrt{2~\pi}$. \subsection{Comparison with Other Methods} \label{compare} We use extrapolated moments to measure GMC properties rather than employing an established method from the literature. In this section, we justify our choice by comparing several methods of measuring GMC properties. We focus on the performance of these methods at marginal resolution and low signal-to-noise, conditions typical of extragalactic GMC observations. Determining the radius of a cloud is particularly difficult because GMCs are often asymmetrical with poorly defined boundaries. Several authors have devised methods to return a single characteristic size for complicated emission distributions. The intensity-weighted second moments in the spatial directions have been used in many studies \citep[e.g.~][]{srby87}, but are sensitive to both noise and convolution effects. The other commonly used method \citep[e.g.~][]{clumpfind,hc01} is to infer the radius based on the area of the cloud: \[R_e = \sqrt{\frac{A-A_{pt}}{\pi}}.\] Here $A_{pt}$ is the area of a point source that has been convolved with a beam and measured with the same signal-to-noise as the emission in the map. Finally, \citet{ws90} and \citet{tay99} adopted the size of the cloud as the mean of the deconvolved FWHMs of the emission distribution along two perpendicular directions. We compare these three methods to the extrapolated moment method presented above across a range of resolutions and sensitivities. We measure the size of a Galactic GMC using each method after convolving it to a desired resolution and adding noise to produce a particular signal-to-noise ratio. For the data, we use the $^{12}$CO data from the Rosette molecular cloud \citep{bs86}, which we clip at 2$\sigma_{RMS}$ (the RMS noise in the original data set) and integrate in the velocity dimension to produce a map of integrated intensity. For a range of sensitivities and resolutions, we convolve this map with a Gaussian beam and add noise. We measure the size of the cloud in 100 realizations of the noise for each such \{resolution, sensitivity\} pair using (a) the extrapolated moment method, (b) the moment of the data without extrapolation, (c) the area method and (d) the FWHM method. Wherever possible we corrected for the effects of beam convolution and signal-to-noise for each of the methods. The results of the analysis are shown in Figure \ref{radplot}. Figure \ref{radplot} shows the recovered radius as a function of resolution and sensitivity, with the ``true'' radius defined as that measured at very high sensitivity and very good resolution (i.e. in the original data, the top right corner of each plot). The hashed region of parameter space shows the range of parameters over which each algorithm recovers a radius within 10\% of the true value. The extrapolated moment method has the largest hashed region and so is remarkably robust, recovering the true radius over a large range of sensitivities and resolutions. Only at low sensitivity ($T_{max}/\sigma_{RMS}<5$) and marginal resolution ($\sigma_{beam}\gtrsim \sigma_0$), does the derived radii depart systematically from the true radius. Notably, the extrapolated second moment performs quite well at signal-to-noise ratios from 5 to 10 (in the integrated intensity map), values typical of extragalactic CO data sets. By contrast, the uncorrected moment method (panel b in Figure \ref{radplot}) underestimates the size at low signal-to-noise since (by construction) the uncorrected moment does not account for emission below the noise level. Similarly, the area method (panel c) shows systematic variation at both low signal-to-noise (where emission drops below the noise level) and low resolution (where the convolved area of the emission distribution grows disproportionately because of the filamentary nature of the cloud). Finally, the FWHM method (panel d) shows large systematic variations since it depends only on the location of the FWHM contour and not on the remainder of the emission distribution. The region of systematic underestimation at low signal-to-noise but reasonable resolution shows the effects of missing the diffuse emission mentioned in \S\ref{EXTRAPSECT}. The Rosette includes more CO emission at low intensities than the extrapolated moment predicts from the high intensity data. As a result, when that diffuse emission is not included in the measurement, the algorithm underestimates the true radius of the GMC. This effect is seen panels (a) and (b) --- the extrapolated and uncorrected second moments --- and is more pronounced in the uncorrected second moment, panel (b). We perform a similar experiment on recovering the line width of an emission line in noisy data. We measure the recovered line width of a Gaussian line of known width using three methods (a) the extrapolated moment method (b) an uncorrected second moment and (c) a Gaussian fit to the line. For a range of signal-to-noise levels ($T_{max}/\sigma_{RMS}$) and channel widths $\Delta V_{chan}$ we measure the recovered line width relative to the known line width. The mean values of the recovered line for 1000 realizations of the noise are plotted in Figure \ref{dvplot}. The extrapolated moment does not show the systematic variation with signal-to-noise seen in the uncorrected moment. The extrapolated moment is nearly as robust a measure as the Gaussian fit for a perfectly Gaussian line and will prove superior if the line is not Gaussian. Robust recovery of the line width using any method requires $\Delta V_{line} / \Delta V_{chan} > 2$. \subsection{Assessing Errors in GMC Properties} The formal uncertainty associated with each moment measurement is quite small. Cloud identification and extrapolation represent larger sources of uncertainty, but their effects are difficult to assess formally. We use bootstrapping methods to estimate the uncertainties in our measurements of cloud properties. The bootstrapping method determines errors by generating several trial clouds from the original cloud data. A trial cloud is generated by considering the cloud to be a collection of data $\{x_i,y_i,v_i,T_i\}$ for $i=1\dots N$, the number of points in the cloud. The data are sampled for $N$ random values of $i$, allowing for $i$ to be repeated. The properties of the cloud are measured for each trial cloud. We estimate the uncertainty from the variance of the cloud properties derived from these resampled and remeasured data sets. The final uncertainty in each property is the standard deviation of the bootstrapped values scaled up by the square root of the oversampling rate. The oversampling rate, which is usually equal to the number of pixels per beam, accounts for the fact that not all of the data in each cloud are independent. For many interferometric data sets this is an important effect, since these data can have $10$ or more pixels per beam. We compare the uncertainties produced by the bootstrapping to those derived from repeatedly adding noise to and then reanalyzing a data set. We use the bright cloud in M~33 shown in Figure \ref{EXTRAPMOMS}. We conduct 100 realizations of the data plus random noise. The resulting uncertainties in $\sigma_{maj}$, $\sigma_{min}$, $\sigma_v$, and the flux are $3\%$, $2\%$, $3\%$, and $3\%$. Repeatedly bootstrapping the same data set (adjusted to have the same final noise level) yields average uncertainties of $9\%$, $9\%$, $11\%$, and $5\%$. The bootstrapping estimates are higher for this bright cloud because they reflect both the formal uncertainty and the robustness of the result to the removal of a given piece of data. In the low signal-to-noise regime, the values for the two methods converge as noise dominates the uncertainty derived from bootstrapping --- for example, performing the same experiment in a dimmer M~33 cloud with $1/4$ the luminosity of the bright cloud and comparable noise, bootstrapping yields errors of $31\%$, $33\%$, $32\%$, and $35\%$ in the four moments while repeated realizations produces scatters of $15\%$, $27\%$, $30\%$, and $40\%$. The bootstrapping method produces a robust, believable estimate of the uncertainty in the measurement of the properties of a particular, defined cloud. It does not account for uncertainties in the assignment of emission to a cloud either as a result of noise or choice of algorithm. These uncertainties are more systematic than random in nature and may be best assessed by analyzing the emission distribution using several methods. The bootstrapping estimate may be treated as an accurate estimate of the uncertainties in the results {\em given} that one adopts the methods presented in this paper. \section{Deriving Physical Quantities from Moment Measurements \label{PHYSICAL}} In this section, we outline how to use the measured size, line width, and flux to calculate several physical quantities of interest: the effective spherical radius, the virial mass, and the luminous mass. Throughout this section we assume that clouds can be described as self-gravitating spheres with density profiles $\rho \propto r^{-1}$ and negligible support from magnetic fields or confinement by external pressure. We assume below that the data consists of observations of the \co\ \jone\ transition, in units of brightness temperature, but the method is readily adaptable to analogous data sets. \subsection{The Spherical Radius} We define a factor $\eta$ that relates the one-dimensional RMS size, $\sigma_r$, to the radius of a spherical cloud $R$: $R=\eta \sigma_r$. It is possible to derive an estimate for $\eta$ based on spherical cloud of radius $R$ with a density profile $\rho \propto r^{-\beta}$. In this model, \begin{eqnarray} \sigma_r^2 &=& \frac{\int_0^R dr \int_0^{2\pi} d\theta \int_0^{\pi} d\phi~x^2 \rho_0 r^{-\beta} r^2 \sin \phi}{\int_0^R dr \int_0^{2\pi} d\theta \int_0^{\pi} d\phi~ \rho_0 r^{-\beta} r^2 \sin \phi} \\ \sigma_r^2 &=& \frac{1}{3}~\frac{3 - \beta}{5 - \beta} R^2 \\ \mathrm{so~that~} \eta &=& \sqrt{3~\frac{5 - \beta}{3 - \beta}} \end{eqnarray} \noindent For a cloud with $\beta = 1$, $\eta = \sqrt{6}$, somewhat higher than the empirical correction of $3.4/\sqrt{\pi}$ derived by Solomon et al. (1987). The difference arises, in part, from using $^{12}$CO as a density tracer. Since $^{12}$CO emission saturates in dense regions and vanishes from low density regions, the apparent density profile in $^{12}$CO is shallower than the true density profile. Hence, an appropriate value of $\eta$ likely falls in the range between $3.4/\sqrt{\pi}$ and $\sqrt{6}$. For tracers like $^{13}$CO with higher critical densities, a different value of $\eta$ may be appropriate. Since comparison to this ``anchoring'' data set may be more important than adopting a self-consistent --- but grossly oversimplified --- model for a cloud, we recommend the Solomon et al. (1987) definition of the cloud radius, $R \approx 1.91 \sigma_{r}$. Note, that adjusting the definition of the radius renders the virial mass formula we present below inexact. \subsection{The Spatial Size, Line Width, Luminosity, and Mass} We convert the cloud properties $\sigma_r ( 0~\mathrm{K})$, $\sigma_v ( 0~\mathrm{K})$, and $F_{\mathrm{CO}} ( 0~\mathrm{K})$ to the physical quantities $R$, $\Delta V$, and $L_{\mathrm{CO}}$. For a cloud at a distance of $d$ (in parsecs), the physical radius will be \begin{equation} R[\mathrm{pc}] = \frac{R ( 0~\mathrm{K}) [\mathrm{arcsec}]}{3600} \times \frac{\pi}{180} \times d [\mathrm{pc}]~\mathrm{,} \end{equation} \noindent the FWHM line width will be \begin{equation} \Delta V = \sqrt{8\ln (2)}~\sigma_v ( 0~\mathrm{K})~\mathrm{,} \end{equation} \noindent and the luminosity of the cloud, $L_{\mathrm{CO}}$, will be \begin{eqnarray} \nonumber L_{\mathrm{CO}} [\mathrm{K~km~s}^{-1}\mathrm{~pc}^2] &=& F_{\mathrm{CO}}(0~\mathrm{K}) [\mathrm{K~km~ s}^{-1}~\mathrm{arcsec}^2] \\ \nonumber &\times& (d[\mathrm{pc}])^2\\ &\times& \left(\frac{\pi}{180\cdot 3600}\right)^2. \end{eqnarray} \noindent A particular CO luminosity, $L_{\mathrm{CO}}$, implies a mass of molecular gas, $M_{\mathrm{Lum}}$, of \begin{eqnarray} \nonumber {M_{\mathrm{Lum}}}~[{M}_{\odot}]& = &\frac{X_{\mathrm{CO}}}{2 \times 10^{20} [\mathrm{cm}^{-2}/(\mathrm{K~km~s}^{-1})]} \times 4.4~L_{\mathrm{CO}}\\ &\equiv& 4.4~X_2~L_{\mathrm{CO}} \end{eqnarray} \noindent where $X_{\mathrm{CO}}$ is the assumed CO-to-H$_2$ conversion factor. This calculation includes a factor of 1.36 (by mass) to account for the presence of helium. Including helium is important to facilitate comparison with the virial mass, which should reflect all of the gravitating mass in the cloud. We have adopted a fiducial value of the CO-to-H$_2$ conversion factor of $X_{\mathrm{CO}}=2\times 10^{20}\mbox{ cm}^{-2} (\mbox{K km s}^{-1})^{-1}$ and express changes relative to this value in terms of the parameter $X_2$. \subsection{The Virial Mass} We compute the virial masses under the assumption that each cloud is spherical and virialized with a density profile described by a truncated power law of the form $\rho \propto r^{-\beta}$ with no magnetic support or pressure confinement. As with the spherical radius correction, the exact density profile of the cloud will affect the correct form of the virial theorem mass. For $\beta = 1$, the virial mass is given by the formula (Solomon et al. 1987): \begin{equation} {M_{\mathrm{VT}}} = 189~M_{\odot}~\Delta V^2 \, R \end{equation} \noindent and more generally by \begin{equation} {M}_{\mathrm{VT}} = 125~M_{\odot}~\frac{5-2\beta}{3-\beta}~\Delta V^2 \, R~\mathrm{,} \end{equation} \noindent where $\Delta V$ is the FWHM velocity line width in \kmpers, $R$ is the radius in pc, and the cloud has a density profile of $\rho \propto r^{-\beta}$. Clouds exhibit a range of non-spherical geometries and may be supported by magnetic fields or confined by pressure. Therefore, the studying the virial parameter may be more useful than the virial mass itself. The virial parameter is a constant of order unity that characterizes deviations from the virial theorem applied to a non-magnetic cloud with no external pressure and constant density. Following \citet{mckee-vt}, we define the virial parameter as \begin{equation} \alpha = \frac{5 \sigma_v^2 R}{M_{\mathrm{Lum}} G} = \frac{5 \eta \sigma_v^2 \sigma_r}{(4.4 X_2 L_{\mathrm{CO}}) G}~\mbox{.} \end{equation} Larger-than-unity virial parameters can result from pressure confinement, while $\alpha<1$ may result from significant magnetic support. Incorrect values of the CO-to-H$_2$ conversion factor may skew the result in either direction. Finding $\alpha<2$ means that the clouds are gravitationally bound in the absence of significant magnetic support. \section{Measuring Cloud Properties from Interferometer Observations} \label{INTERF} Millimeter-wave interferometers are required to resolve even the most massive molecular clouds in galaxies beyond the Magellanic clouds (see Figure \ref{DISTCOMP}). Unfortunately, interferometers are not sensitive to spatial frequencies outside the limited region of the $(u,v)$ plane that they sample. Practically, this means interferometers do not measure the total flux from the emission distribution; and structures are resolved out, usually on large angular scales that correspond to small separations in the $(u,v)$ plane. Ideally, interferometer observations are combined with single-dish observations that supply the missing information. In practice such observations are conducted infrequently and the unknown total flux and short-spacing information is estimated using deconvolution algorithms such as CLEAN or maximum entropy \citep[see][and references therein]{tms}. \citet{she00} simulate the results of using only an interferometer to observe Galactic GMCs as if these well-studied clouds were located in M31. They find that interferometers experience significant (50\%) flux loss for their simulated observation, primarily from extended emission around clouds. However, they find that the flux loss does not change the size and line width of the cloud. \citet{hel02} examine the recovery of large-scale flux distributions from interferometer measurements in more depth and explore the effectiveness of deconvolution algorithms at low signal-to-noise. They find that deconvolution algorithms recover flux nonlinearly at low sensitivities, finding much less flux at low sensitivities than one would expect. Since much of the data on extragalactic GMCs have low signal-to-noise, this may represent an important bias. We assess the effects of interferometric biases on the methods presented here by extending the method of \citet{she00}. We use $^{12}\mathrm{CO}$ observations of three galactic GMCs: the Orion molecular complex \citep{wil05}, the Rosette Molecular Cloud \citep{bs86}, and an excerpt from the Outer Galaxy Survey of \citet{hc01} which contains the molecular clouds associated with the W3/W4/W5 \ion{H}{2} regions\footnote{A map of the original Orion data and typical maps at low resolution and sensitivity appear in the Appendix.}. We simulate observing these three molecular complexes in M31 ($D=770$ kpc) with the BIMA interferometer. We Fourier transform each plane of each data set into the $(u,v)$ domain and resample the data along the $(u,v)$ tracks that would be sampled by BIMA observations of the data provided the GMCs were in M31. The $(u,v)$ coverage reflects typical observing strategies for extragalactic clouds, such as interleaving observations of the source with calibrators and other sources. We add thermal noise and phase noise to the $(u,v)$ data, including a phase noise component with a magnitude that depends on the length of the baseline. We adjust the scale of the thermal and phase noise to produce the desired peak signal-to-noise (we find our results depend only weakly on whether the noise is thermal or phase). We invert the resulting $(u,v)$ data using the MIRIAD software package \citep{miriad} producing maps separated by 2 km s$^{-1}$, and then we deconvolve the dirty maps using a CLEAN algorithm that terminates at the 2$\sigma_{RMS}$ level. For each trial cloud, we then calculate the cloud properties using the methods of \S\S \ref{MOMENTS} and \ref{PHYSICAL}. For comparison, we compute cloud properties using the same procedure to simulate single-dish observations with signal-to-noise and effective resolution identical to the mock interferometer data. We generate the mock single-dish observations by sampling the transformed image data for an equal number of $(u,v)$ points as the interferometer data, but the points are normally distributed in the $(u,v)$ plane and one point is forced to lie at $(0,0)$ thus sampling the total power. The width of the $(u,v)$ point distribution is chosen to give a beam size similar to that of the mock interferometer data. Random thermal and phase noise is added to these data in the same fashion as for the interferometer data. Then the data are inverted using natural weighting and deconvolved in the exact same fashion as the interferometer data (though the deconvolution step has little effect). Again, we extract cloud properties using the methods of \S\S\ref{MOMENTS} and \ref{PHYSICAL}. With these simulations, we compare the cloud properties derived from interferometer data to single-dish data that are equivalent in every other fashion, thereby isolating the effects of interferometers on the derived properties. The additional biases imposed by limited resolution and sensitivity are discussed separately in \S\ref{compare}. Here we focus on mock BIMA observations of the three molecular complexes using three antenna configurations: the C array (extended), the D (compact) array, and a combination of C and D array \citep[see][for details on the configurations]{wright-configs}. The synthesized beam sizes for these configurations are $14.4''$, $6.1''$ and $8.8''$ for the D, C and C+D hybrid array observations respectively, corresponding in turn to 54, 22 and 33 parsecs at 770~kpc. We conduct 10 sample observations for each cloud in each array at a range of sensitivities ranging from $T_{peak}/\sigma_{RMS}=3$ to $>100$. Figure \ref{interf_props} shows the properties recovered by mock observations of the Orion molecular complex for each array and a range of sensitivities. The values of each properties are normalized by the value recovered by mock single-dish observations at the same sensitivity and resolution. Thus, the only difference between the four sets of properties (C, D, C+D, and single dish) is the $(u,v)$ coverage of the simulated observations. We find that the derived properties from mock single-dish observations follow the same behavior as the simulations in \S\ref{compare}. Thus, it is possible to decouple the two sets of biases -- those arising from marginal resolution and sensitivity and those arising from using interferometers -- and examine only the latter. We plot the results for Orion because these observations show the most dramatic variation of the three complexes studied, but the results are qualitatively the same for all three data sets. Orion is the most sensitive of the three to spatial filtering because it consists of three GMCs and therefore shows more structure than the other two targets. Based on the results of the mock observations, we make the following comments regarding the use of interferometer data alone in measuring cloud properties. Most of these points can be seen visually in Figure \ref{interf_props}. \begin{enumerate} \item Cloud properties measured from interferometric data are biased. The degree of bias is affected by the sensitivity of the array as well as the $(u,v)$ coverage of the observations. \item A minimum signal-to-noise of 10 is required for stable recovery of cloud properties. Below this level, errors in cloud properties, can approach 100\% for interferometer data. \item Even for intermediate signal-to-noise values ($T_{peak}/\sigma_{RMS}=10$) there are significant systematic effects on the cloud properties. The most extreme effects are on the luminosity measurement, which can be 40\% lower than a single-dish observation. This effect is much less pronounced for measurements of the line width and the radius, which show $\lesssim 10\%$ variations. This result, that the radius and line width are relatively robust to the spatial filtering of the interferometer, confirms the qualitative results of \citet{she00}. The values of derived properties are always {\em underestimated} relative to single-dish observations. \item Even at high sensitivities, the spatial filtering of interferometers affects property recovery at the 10\% level. For example, C-array observations of Orion systematically underestimate the radius of the cloud by 10\% even for very high signal to noise ratios and D-array observations underestimate the flux of the Orion by 5\% even at high sensitivity. \item For interferometer observations, the dynamical mass measurements of GMCs are more robust than the luminosity measurements. This behavior will bias interpretations of the virial parameter in extragalactic observations. Estimates of the CO-to-H$_2$ conversion factor based on the assumption that GMCs are bound or virialized are likely to {\em overestimate} $X_{\mathrm{CO}}$. \item For a given signal-to-noise value, observations with the widest range of $(u,v)$ coverage provide the most robust measurement of cloud properties. Thus, in achieving a given sensitivity, observers should favor arrays with more antennae or observations made in multiple configurations. \end{enumerate} \section{A Note on Decomposition} \label{DECOMPSECT} The choice of how to decompose an emission distribution into individual clouds may be the most important source of bias in GMC property measurements. Many different methods have been applied to identify GMCs in blended emission, the most prevalent being decomposition by eye and the application of the GAUSSCLUMPS \citep{gaussclumps} or the CLUMPFIND algorithm \citep{clumpfind}. When comparing GMCs between two data sets, care must be taken to decompose emission in a consistent way across both data sets, preferably using the same algorithm on both data sets. Furthermore, the physical values of any tuning parameters in the algorithms should be matched where possible so that both algorithms search for peaks in the emission over the same spatial scale (rather than angular or resolution-units) or intensity range. This will avoid, for example, comparing ``clumps'' in a Galactic molecular cloud to GMCs in another galaxy. A more extreme method to ensure accurate comparison is to convolve the higher (spatial) resolution data set to the spatial resolution of the other data \citep[e.g.~][]{she00}. However, this approach clearly sacrifices accuracy of the derived parameters to allow a more careful comparison between two data sets. In the Appendix to this paper we present a robust, conservative, new decomposition algorithm. This algorithm is designed to avoid creating spurious clouds from noise and to remain sensitive to non-Gaussian structures in the data. Additionally, the parameters of the algorithm are fixed to physical values rather than being determined by the data. This algorithm is designed explicitly with the goal of decomposing emission into GMCs (rather than clumps or other structures) with extragalactic data in mind. The methods for measurement of cloud properties described above --- including the sensitivity and resolution corrections --- are independent of the decomposition algorithm and are important no matter what decomposition algorithm is chosen. In order to avoid confusion between these two separate problems, we choose to describe the decomposition algorithm in the Appendix. \section{Discussion and Conclusions} \label{APPLICATIONS} We conclude the paper by applying these methods to molecular line data sets that have been previously published. In future studies, the algorithm will be used to evaluate the differences between GMC populations between galaxies. Here, we simply present an analysis designed to demonstrate the method's utility. We present the median corrections found for a large set of extragalactic (Local Group) observations and a test application of our methods to Galactic data. We use all the methods discussed in the previous sections and the decomposition algorithm discussed in the Appendix. \subsection{The Effects of Extrapolation and Deconvolution} We apply the methods outlined here to an array of data from across the Local Group and measure the properties of 110 spatially resolved clouds to estimate the typical magnitude of the sensitivity and resolution corrections for extragalactic data. We use BIMA data on M~33 and M~31 \citep{ros03,ros05}; NANTEN observations of the LMC \citep[][]{fu99}; OVRO observations of IC~10 \citep[][]{wa05}; and a SEST map of N83 in the SMC \citep[][]{bol03}. Table \ref{corrections} shows the number of clouds measured in each galaxy along with the median sensitivity and resolution corrections applied to the radius, line width, and flux. For comparison, we also measure the properties of a number of clouds in the outer Galaxy (Quadrant 2, see below) from the \citet[][]{mwco} CO survey of the Milky Way. Table \ref{corrections} includes all spatially resolved clouds with masses (derived from the CO luminosity) of $5 \times 10^4$ M$_{\odot}$ or more ($92$ of the $110$ extragalactic clouds are above this mass). The numbers quoted in Table \ref{corrections} are ``correction factors,'' \begin{equation} \frac{R_{corrected}}{R_{uncorrected}}, ~\frac{\sigma_{v,corrected}}{\sigma_{v,uncorrected}},\mbox{ and } \frac{L_{\mathrm{CO},corrected}}{L_{\mathrm{CO},uncorrected}} \mbox{ .} \end{equation} Table \ref{corrections} shows that throughout the Local Group data the corrections suggested in this paper have magnitudes of a few tens of percent. We draw several conclusions based on these data: \begin{enumerate} \item Resolution effects on the size of clouds tend to be significant --- we would overestimate cloud sizes by $\sim 40$\%\ if we did not apply a deconvolution. In the Milky Way data, this effect is much less severe. The sizes of Milky Way clouds are measured to within $5\%$ before the resolution correction. Unresolved clouds do not contribute to Table \ref{corrections}, so if the effects of the resolution bias were completely neglected this would be much larger (a naive approach would measure these clouds to have the size of the spatial beam). \item Resolution effects on the line width are negligible throughout the Local Group data. \item Sensitivity effects are also significant. Without a correction for the sensitivity bias, the size, line width, and luminosity of clouds would all be significantly underestimated. This sensitivity bias is least severe --- only about 20 -- 30\%\ --- for the line width, and most significant (and variable) for the luminosity. Sensitivity corrections to the luminosity vary from 20\% to more than 100\%. \item The resolution and sensitivity biases for the size measurement tend to cancel out, so that the completely uncorrected radius measurement is often within 10 -- 20\%\ of the corrected value. This is a happy coincidence of resolution and sensitivity within the Local Group, not evidence that sensitivity and resolution corrections are unimportant. \item The magnitude of corrections across the Local Group data are fairly uniform. This is because GMCs near the resolution limit tend to outnumber higher mass GMCs. Unresolved GMCs are not included in the analysis, so the median cloud through all the data sets appears marginally resolved. \item In order to compare extragalactic data to Galactic data (with very good sensitivity and resolution and therefore small corrections) it is crucial to correct for the sensitivity and resolution biases. \end{enumerate} \begin{center} \begin{deluxetable*}{l c c c c c c} \tablecaption{\label{corrections} Typical Corrections for Local Group Data} \tablehead{ \multicolumn{1}{l}{Galaxy} & \multicolumn{1}{c}{$N_{Clouds}$} & \multicolumn{3}{c}{Sensitivity Correction} & \multicolumn{2}{c}{Resolution Correction} \\ \multicolumn{1}{l}{} & \multicolumn{1}{c}{} & \multicolumn{3}{c}{$\overbrace{\phm{SpanningSpann}}$} & \multicolumn{2}{c}{$\overbrace{\phm{SpanningSpann}}$} \\ \multicolumn{1}{l}{} & \multicolumn{1}{c}{} & \multicolumn{1}{c}{$\frac{R_{corrected}}{R_{uncorrected}}$} & \multicolumn{1}{c}{$\frac{\sigma_{v,corrected}}{\sigma_{v,uncorrected}}$} & \multicolumn{1}{c}{$\frac{L_{CO,corrected}}{L_{CO,uncorrected}}$} & \multicolumn{1}{c}{$\frac{R_{corrected}}{R_{uncorrected}}$} & \multicolumn{1}{c}{$\frac{\sigma_{v,corrected}}{\sigma_{v,uncorrected}}$} \\} \startdata LMC & 46 & $1.4$ & $1.2$ & $1.7$ & $0.8$ & $1.0$ \\ M~31 & 28 & $1.5$ & $1.3$ & $1.6$ & $0.7$ & $1.0$ \\ IC~10 & 17 & $1.7$ & $1.3$ & $2.3$ & $0.7$ & $1.0$ \\ M~33 & 15 & $1.4$ & $1.2$ & $1.5$ & $0.7$ & $1.0$ \\ SMC & 4 & $1.1$ & $1.2$ & $1.2$ & $0.7$ & $1.0$ \\ MW\tablenotemark{a} & 107 & $1.1$ & $1.1$ & $1.4$ & $1.0$ & $1.0$ \\ \enddata \tablenotetext{a}{Quadrant 2 clouds with $M > 5 \times 10^4$ M$_{\odot}$.} \end{deluxetable*} \end{center} \subsection{Analysis of Second Quadrant CO Data} The method described in this paper has been designed with extragalactic data in mind. However, a crucial step in interpreting extragalactic measurements is to make a fair comparison with Galactic data. In this section we report some results of applying our decomposition and measurement algorithms to the survey of the second Galactic quadrant by \citet{mwco}. We compare the results of this analysis to the results by \citet[][]{hc01} and show that our analysis recovers results that are consistent with theirs. We decompose and analyze $^{12}$CO($1\to 0$) from the second quadrant \citep[Survey 17 in Table 1 of][]{mwco}. The data set covers the Galactic plane from $\ell = 70^{\circ}$ to $\ell = 210^{\circ}$ with a noise level of $0.3$ K. We measure the distance to the molecular emission using the kinematic distances by adopting a flat rotation curve with $\Theta_{\mathrm{LSR}} = 220$~km~s$^{-1}$ and $R_\odot=8.5$~kpc. We omit local emission by discarding all elements of the data cube with a kinematic distance less than 2 kpc as well as all elements in the data cube that are connected by significant emission in position or velocity space to such pixels. We apply the decomposition algorithm described in the Appendix and measured sizes, line widths, and luminosities of GMCs using the methods of \S \ref{EXTRAPSECT} and \S \ref{BEAMDCSECT}. The analysis recovers 431 clouds with resolved angular sizes and line widths located within 10 kpc of the Sun. We include the median sensitivity and resolution corrections for massive ($>5 \times 10^4$ M$_{\odot}$) clouds in Table \ref{corrections} above. Do the results from our algorithm agree with previous studies of Galactic GMCs? The data set covers the region studied by \citet{hc01} using the $45''$ resolution of the FCRAO 14m. That data set has a lower sensitivity than the \citet[][]{mwco} data, so we apply a rudimentary sensitivity correction (assuming that the GMCs are Gaussian and using their peak temperature and boundary isosurface) to their results and scale to $X_{\mathrm{CO}}=2\times 10^{20}\mbox{ cm}^{-2} (\mbox{K km s}^{-1})^{-1}$ before comparing GMC properties. We focus on a comparison of the virial parameter between the two studies --- a full treatment of the Galactic ``Larson's Laws'' is beyond the scope of this paper. We draw several conclusions from the comparison: \begin{enumerate} \item Figure \ref{virparams} shows that the virial parameters derived in our analysis largely agree with those found by \citet[][]{hc01}. Below masses of $\approx 3 \times 10^4$ M$_{\odot}$, both surveys find the same virial parameter in a given mass bin. \item Above $\approx 3 \times 10^4$ M$_{\odot}$, our analysis of the \citet[][]{mwco} data set may yield a slightly higher virial parameter, on average. This may be evidence for the diffuse emission mentioned in \S\ref{EXTRAPSECT} --- the higher sensitivity \citet[][]{mwco} data may include diffuse emission surrounding the GMCs while the \citet[][]{hc01} data may miss this effect. It may also reflect inadequacies in the simple sensitivity correction we apply to the \citet[][]{hc01} data. Resolution effects may also play a role --- the \citet[][]{hc01} data set has $\sim 5$ times better resolution than the \citet[][]{mwco} data and so the lower resolution data may tend to lump unbound clouds together. The number of clouds in the high mass bins is relatively low, so the discrepancy may not be particularly significant. \item We apply our algorithm to a small portion of the OGS data and find our corrections for resolution and sensitivity bias increase the mean virial parameter to be consistent with the results from the \citet{mwco} data. This suggests that the differences in the virial parameters for the high mass clouds in Figure \ref{virparams} may simply be methodological. \item Both catalogs of outer Galaxy clouds find an inverse relationship between luminous mass and the virial parameter, approximately $\alpha \propto M_{\mathrm{Lum}}^{-0.2}$ --- a relationship that was also observed by \citet{srby87} for inner Galaxy molecular clouds. \end{enumerate} Thus we find agreement with the results of previous studies of Galactic molecular clouds. Our methods applied to Galactic data find the same behavior observed in earlier work and we find agreement among the method applied to several data sets. \subsection{Conclusions} We have presented a method for measuring macroscopic GMC properties --- spatial size, line width, and luminosity --- from a region of emission in a spectral line data cube. This method corrects for biases from limited sensitivity and resolution and produces reliable results that are directly comparable among a wide variety of data sets. We correct for limited sensitivity via an extrapolation to a theoretical 0 Kelvin isosurface. We apply a simple quadratic deconvolution to the extrapolated values to account for resolution biases. We find that bootstrapping methods yield believable estimates of the uncertainties in the derived parameters. We present a set of suggestions for transforming the derived properties into physical quantities of interest. In the Appendix to this paper we present a new method for decomposing emission into individual GMCs. This method is conservative and robust, designed to produce robust results from low resolution/sensitivity data. In this section we have applied all of these methods to an array of extragalactic (Local Group) and Galactic data. We find that the algorithm reproduces established results for Galactic GMCs and that the sensitivity and resolution biases are potentially significant --- often $\sim 40\%$ --- for even the most recent Local Group GMC measurements. Based on this investigation of the observational biases in measuring molecular cloud properties, we note several important points that should be considered in planning observations of GMCs and interpreting the results. First, resolution and sensitivity biases can be corrected to $<10\%$ error provided the cloud has a modest peak signal-to-noise ($T_{peak}/\sigma_{RMS}\gtrsim 10$) and is marginally resolved $R_{cld} > 0.8\theta_{\mathrm{FWHM}}$, where $\theta_{\mathrm{FWHM}}$ is the full width at half maximum extent of the beam. Given current and future telescope capabilities, even the properties of extragalactic GMCs can be accurately measured. From Figure \ref{DISTCOMP}, we can see that single dish surveys can accurately study clouds more massive than $10^4~M_{\odot}$ in the Magellanic clouds and careful interferometer observations can recover cloud properties for clouds with mass $M\gtrsim 10^5~M_{\odot}$ in M31 and M33. However, interferometer observations systematically underestimate molecular cloud properties. All else being equal, interferometers can underestimate fluxes by 40\% and cloud radii by 10\% relative to single-dish observations. Line widths are largely unaffected by interferometer observations. The magnitude of the systematic bias depends on the array configuration and the sensitivity of the observations. In general, wider coverage of the $(u,v)$ plane produces better property recovery. To minimize bias, observers should favor observations from several array configurations or from arrays with many elements. If possible, the interferometer data should be supplemented with single-dish observations. Finally, the decomposition algorithm used to separate emission into physically relevant structures will systematically affect molecular cloud properties. To date, there is no algorithm that should be favored in all circumstances, but any comparative study of GMC properties should be consistent in the choice of algorithm. For example, the results of a CLUMPFIND algorithm applied to an extragalactic data set are not directly comparable to a catalog produced by a simple contouring method applied to Milky Way data. Provided the same algorithm is used across multiple data sets, referencing algorithm parameters to a common physical scale will minimize systematic differences. We will make a software version of the decomposition and measurement algorithm available as an IDL package. Further, since methodology can affect the results of GMC studies so strongly, we encourage authors working in the field to make their data available to the community after publication in order to facilitate future rigorous comparisons. \acknowledgements We are extremely grateful to Tom Dame and the Millimeter-Wave Group at the Center for Astrophysics for providing both the Orion data and the Quadrant 2 data used in the paper. We also thank the NANTEN Group at Nagoya, especially Yasuo Fukui and Akiko Kawamura, providing us with the LMC CO data. We are grateful to Fabian Walter for providing us with the OVRO IC~10 data. Alberto Bolatto, Jason Wright, Jon Swift, and Leo Blitz all offered helpful comments on drafts of the paper. We thank Ronak Shah for helping us compare our IDL version of GAUSSCLUMPS to the original implementation of the algorithm. The informed comments of an anonymous referee greatly improved the paper, particularly in encouraging us to explore the effects of interferometers. ER is grateful for support through the National Science Foundation Astronomy \& Astrophysics Postdoctoral Fellows Program (AST-0502605). This work is partially supported by NSF grant 0228963 to the Radio Astronomy Laboratory at UC Berkeley. \begin{appendix} \label{ouralg} \section{Appendix: A New Decomposition Algorithm \label{DECOMPOSITION}} The choice of what emission to identify as a GMC may be the single largest source of uncertainty in measuring and comparing GMC properties. A number of methods have been employed over the years, from simple contouring methods \citep[e.g.~][]{sss85,dect86,srby87,hc01} to fitting three-dimensional Gaussians \citep[GAUSSCLUMPS,][]{gaussclumps,gaussclumps2}, to modified watershed algorithms \citep[CLUMPFIND and its kin,][]{clumpfind,clumpfind-mod}. In this section we present a new decomposition algorithm that is designed to identify clouds at low sensitivities while avoiding the introduction of a false clouds due to noise. This algorithm consists of two parts: identifying regions of significant emission in the data set and then assigning this emission to individual ``clouds.'' \subsection{Signal Identification} We first identify regions of contiguous, significant emission in our position-position-velocity data cube. We estimate the noise in the data set by measuring the RMS intensity, $\sigma_{RMS}$, from a signal-free region of the data cube. We then construct a high-significance mask. This mask includes only adjacent channels that both have intensities above $4\sigma_{RMS}$. We expand that mask to include all emission above a lower threshold --- typically two channels above $2\sigma_{RMS}$ significance --- that is connected to the original high significance mask through pixels with $\ge 2\sigma_{RMS}$ significance. The resulting mask contains most of the significant emission in the data cube. Lowering the threshold below two channels at $\approx 1.5\sigma_{RMS}$ runs the risk of biasing the moment measurements towards high values by including false emission (noise with positive values) in the cloud. \subsection{Cloud Identification} \label{algorithm} In this section, we describe the algorithm used to decompose a region of emission into individual subsections representing the physically distinct entities in the data (``clouds''). Through the description of the algorithm, there are several parameters that can be varied to produce changes in the resulting decomposition. In general, we set these parameters using physical prior knowledge of the GMCs we are seeking to catalog. We discuss the choice of these parameters in the following section. \begin{enumerate} \item {\em Discard Small or Low Contrast Regions:} If a region is too small for us to measure meaningful properties from it, we discard the region. We require that each region has an area larger than two beam sizes, so that we can measure its size; and a velocity width of more than a single channel, so that we can measure its line width. If the intensity contrast between the peak and the edge of the region is less than a factor of two, we lack the dynamic range needed to correct the sensitivity bias and we therefore discard the region. If a region is not discarded we proceed to the next step. \item \label{xform}{\em Rescale the Data to Reduce the Effects of Substructure:} Molecular clouds contain significant substructure that confuses the decomposition of these sources. The substructure is often significantly brighter than bulk of the gas in the cloud. We rescale the data to reduce the contrast between this substructure and the cloud as a whole. The data are rescaled using the following transform: \[ T' = \left\{\begin{array}{cl} T & ;~T<T_{clip}, \\ T_{clip} [1+{\arctan (T/T_{clip}-1)}] & ;~ T\ge T_{clip}. \\ \end{array} \right. \] This transform reduces the contrast pixels with $T\gtrsim 2 T_{clip}$ while preserving the relative brightness distribution. The value of $T_{clip}$ is left as a free parameter. The transformed data are used in the decomposition algorithm. Such brightness transforms are frequently used in the decomposition algorithms used in other fields such as medical imaging \citep[e.g.~][]{cytom}. \item {\em Identify Independent Local Maxima:} We identify potential local maxima, by identifying the elements in the data cube that are larger than all their neighbors. We consider neighbors to be all data that lie within a box with side length $D_{max}$ in position and $\Delta V_{max}$ in velocity centered on the local maximum. These are our ``candidate maxima.'' The parameters $D_{max}$ and $\Delta V_{max}$ are free parameters. If we find more than one candidate maximum in a region of emission, we proceed to the next steps to further verify each maximum's independence. If we find a single candidate maximum then we label the region as a cloud and measure its properties as described in the main paper.\label{locmax} \item {\em Find Shared Isosurfaces and Reject Small Clouds or those with Smooth Mergers:} For each {\em pair} of candidate maxima we calculate the value of the highest intensity isosurface to contain both maxima. We refer to this highest shared isosurface as the {\em merge level}. Using this set of highest shared isosurfaces, we calculate three properties of interest: \begin{enumerate} \item The area uniquely associated with each maximum (i.e. the area above the merge level for that maximum). \item The antenna temperature interval between the merge level and each maximum, referred to as the {\em contrast interval}. \label{decimate} \item The fractional amount by which each of the (unextrapolated) moment values changes across each shared isosurface (i.e. $\frac{\Delta \sigma_{maj}}{\sigma_{maj}}$ for each maximum across each shared isosurface). We consider the two clouds to merge smoothly across the isosurface only if: \begin{enumerate} \item None of the second moments increase by more than 100\% for both maxima. \item No two of the second moments increase by more than 50\% for both maxima. \item The flux increases by less 200\% for both maxima. \end{enumerate} \end{enumerate} We use these three properties to pare maxima from the region. We reject maxima associated with small areas (less than two beam sizes, as for the region above) and contrast intervals less than $\Delta T_{max}$, a free parameter. Choosing $\Delta T_{max}\ge 2\sigma$ significantly reduces the effects of noise on decomposition \citep[e.g.~][]{clumpfind-mod} since noise is associated with low contrast intervals. Finally, when a pair of maxima merge smoothly across a shared isosurface, we keep the higher intensity maximum and discard the lower one. This is a conservative choice in the decomposition algorithm: unless merging the two kernels significantly alters the properties of one of the clouds associated with the separate kernels, we assume the kernels are not physically distinct. The effect of removing kernels that merge smoothly from our data set is to reduce the algorithm's sensitivity to substructure within clouds. We iterate this step until we have a set of maxima associated with the required areas and separated from each other by significant jumps in their properties. \item {\em Define Clouds Using Shared Isosurfaces:} The surviving maxima each correspond to a ``cloud.'' That cloud consists of the emission within the lowest intensity isosurface uniquely associated with the cloud. Emission that lies below this isosurface is part of a ``watershed'' shared among clouds and we do not assign it to any cloud. By not considering contested emission, i.e.~emission that could be associated with distinct local maxima, we avoid the problem of how to properly assign this emission to local maxima. \item {\em Measure Cloud Properties:} Finally, we apply the methods described in \S2 and 3 to derive spatial sizes, line widths, and luminosities for each cloud. The transformed data (Step \ref{xform}) are inverse transformed into the original brightness units for this analysis. \end{enumerate} \subsection{Using Physical Priors to Establish Algorithm Parameters} Several of the algorithm's parameters are left to the choice of the user. Without any prior knowledge of the physical objects in the data, we establish default values for these parameters that will produce a reasonable decomposition based solely on the characteristics of the data. These defaults are chosen to provide sensitivity to real substructure within the data without contamination by noise and can be regarded as the {\it minimum} appropriate values for these parameters in most cases. The free parameters in the algorithm are the brightness transform threshold ($T_{clip}$, Step \ref{xform}) position and velocity window size used in searching for local maxima ($D_{max}$ and $\Delta V_{max}$, Step \ref{locmax}), and the minimum contrast temperature ($\Delta T_{max}$, Step \ref{decimate}). Without prior knowledge of substructure in the data, no brightness transform should be applied ($T_{clip}=\infty$). The minimum values for the parameters in the search for initial local maxima are set by the resolution of the data: $D_{max}=\theta_{beam}$ and $\Delta V_{max}=\Delta V_{chan}$ or else separations between local maxima cannot be resolved. Finally, we use $\Delta T_{max}=2\sigma_{RMS}$ to prevent the noise in the data set from being recognized by the algorithm as legitimate structure and becoming the basis for the decomposition. In the main section of the paper, we developed methods that measured physical properties of molecular emission without observational bias. Ideally, the decomposition algorithm used in the analysis would also be free of observational bias. A good algorithm would, for example, decompose the same emission into similar structures regardless of the resolution and sensitivity of the observations. To progress towards this goal, we set the parameters of the decomposition algorithm to have similar values in physical units (pc, km s$^{-1}$, K) rather than data units (beam widths, channels, $\sigma_{RMS}$). In the molecular ISM, which has structure on a range of scales, the choice of the physical values for the algorithm parameters must be motivated by prior knowledge of the objects that we wish to identify. The choice of parameters for identifying GMCs would be different from the choice of parameters for identifying clumps and cores. As a population, GMCs have size scales of 10s of parsecs, line widths of several km s$^{-1}$, and brightness temperatures of $T \lesssim 10$~K \citep[e.g.~][]{srby87}. In contrast, the clumpy substructure within clouds has a size scale of 1 pc, line widths of order 1 km s$^{-1}$ and can have brightness temperatures of 30 K or higher \citep[e.g.~][]{clumpfind}. We select the parameters of the algorithm to find molecular clouds rather than the clumps within them. In particular, we set the minimum separation between local maxima as $D_{max}=15$~pc and the velocity separation to be $\Delta V_{max} =2$~km s$^{-1}$. We also fix a contrast interval in antenna temperature of $\Delta T_{max} = 1$~K rather than the data-driven value of $2 \sigma_{RMS}$. Finally, we must account for the presence of bright molecular gas substructure within molecular clouds. For example, Orion A has a few distinct regions separated by $>15$ pc with antenna temperatures in excess of 15 K \citep{wil05}. These are associated with hot molecular gas around young stars like the Trapezium cluster. In general, the typical kinetic temperature of gas in GMCs is $\sim 10$~K so brightness temperatures in excess of 10 K are usually associated with structure {\it within} molecular clouds not changes from cloud to cloud. Hence, to catalog clouds and not substructure, we must reduce the influence of the high $T_b$ data with the parameter $T_{clip}$. We use $T_{clip} = 2.5$~K to maintain the full sensitivity for the gas separating GMCs (which typically has $T_A \lesssim 5$~K in the \citet{srby87} data) while reducing the influence of brighter gas associated with a single GMC. Using $T_{clip}=2.5$~K has little influence on extragalactic data where beam deconvolution typically averages out the presence of bright substructure within the clouds. We summarize our choices of parameters in Table \ref{params} when the parameters are motivated by the data (Data-Based) or by physical assumptions about the structures being extracted (GMC Physical Priors). The resolution or the sensitivity of a data set may be sufficiently poor that the physical parameters are unattainable in the data set (e.g.~$\theta_{beam}>15~$pc, $2\sigma_{rms}> 1$~K). In this case, the decomposition must be regarded with caution as it may not be directly comparable to other data sets. The adopted values are appropriate for decomposing data sets of $^{12}$CO emission. The values of $T_{clip}$ and $\Delta T_{max}$ would be different for other tracers. \begin{deluxetable}{lcc} \tablecaption{Parameter Choices for Decomposition Algorithm\label{params}} \tablehead{ \colhead{Parameter} & \colhead{Data-Based} & \colhead{GMC Physical Priors}}\ \startdata $T_{clip}$ & $\infty$ & 2.5 K \\ $D_{max}$ & 1 beam width & 15 pc \\ $\Delta V_{max}$ & 1 channel & 2 km s$^{-1}$ \\ $\Delta T_{max}$ & $2\sigma_{RMS}$ & 1 K \\ \enddata \end{deluxetable} \subsection{Comparison with Existing Algorithms} To demonstrate that the proposed algorithm is actually an improvement on existing methods, we analyze trial data sets with known properties using this algorithm and compare the results to those from the GAUSSCLUMPS and CLUMPFIND algorithms to the trial data. We use the CLFIND algorithm implemented in MIRIAD \citep{miriad} and our own IDL implementation of the \citet{gaussclumps2} GAUSSCLUMPS algorithm, which we have compared to the standard algorithm operating on a real data set with satisfactory results. We first examine how adept the three algorithms are at decomposing a pair of blended clouds. We construct a series of data cubes with two unresolved model clouds separated in position by a variable distance. The model clouds each have a peak signal-to-noise ratio of 10 with a Gaussian line profile. For each value of the separation, we decompose 100 data cubes with different realizations of the noise using the three algorithms, using data based choices for the algorithm parameters. Figure \ref{nclplot} (left) shows the mean number of clouds recovered by each algorithm as a function of the separation (we count clouds with peak signal-to-noise larger than 5 as ``recovered''). As the trial clouds are moved farther apart, the typical number of clouds detected by each algorithm increases by 1. However, only the algorithm presented here consistently recovers a single cloud at low separations. GAUSSCLUMPS and CLUMPFIND produce false clouds from the noise. The jump in the number of clouds detected by the GAUSSCLUMPS algorithm occurs where the separation equals the resolution, whereas both the current algorithm and CLUMPFIND are able to distinguish the clouds only if their separation is over 1.5 times the resolution. GAUSSCLUMPS appears to be able to distinguish tight blends of clouds but is the most susceptible to noise. As a second test of the algorithms, we compared the number of clouds that the algorithms recover from a single cloud with a circular top-hat brightness profile ($T_A=\mbox{const. for } r<R_0;~0$ otherwise) and peak signal-to-noise of 10. The size of the cloud is varied with respect to the resolution of the data set and 100 data sets for each value of $R_0$ are decomposed by each algorithm. The non-Gaussian brightness profile confounds all of the algorithms but to varying degrees. CLUMPFIND detects an increasing number of spurious clouds as the cloud grows, suggesting that the number of false clouds grows with the volume studied. Despite the non-Gaussian profile, GAUSSCLUMPS does surprisingly well with large sources. The current algorithm, however, does the best job of detecting a single source in the presence of noise. For analyzing data with a relatively low signal-to-noise, we find the decomposition algorithm presented in this Appendix should be favored for identifying clouds. \subsection{Applying the Algorithm to the Orion-Monoceros Region} The Orion Molecular Cloud is among the best studied of all GMCs, so it is an good place to compare the results of our methods to those of previous work. For this comparison, we use the data and results of the recent, uniform survey of the entire Orion-Monoceros region by \citet{wil05}. We analyze their final data set using the methods in this paper (with physical priors for the decomposition algorithm parameters) and summarize the results in Figure \ref{orion_fig}, an integrated intensity map of the region. The decomposition results in 81 molecular clouds, most of which are associated with the Galactic plane at the top of the Figure. The results of the algorithms decomposition of the region are good: major molecular clouds are identified as single entities in the decomposition. The algorithm succeeds where other cloud identification schemes would face difficulty. Nearly all of the emission in the data is connected above a single isosurface so that simple contouring methods would identify the entire Orion-Monoceros region as a single molecular cloud, though more complicated algorithms \citep{syscw} may succeed. CLUMPFIND and GAUSSCLUMPS would isolate individual peaks and decompose clouds into their substructure --- as was intended by their design (CLUMPFIND identifies 617 clumps in the same data). The properties of the clouds agree well with the values published for a human decomposition of the emission. The results of the analysis are given in Table \ref{orion_props}. For comparison, we adopt the distances of \citet{wil05}. We compare the results of the algorithm to their results, after scaling their mass up by 10\% to account for a difference in the adopted CO-to-H$_2$ conversion factors. In several cases the masses agree quite well (to within 5\%, Orion A, Mon R2, Scissors), while other features show $\sim 30\%$ differences. These systematic discrepancies arise from differences in how emission is assigned into structures. For example, the algorithm only identifies the central region of Orion B as a molecular cloud; it does not include emission near the location of Orion East that is nominally part of the Orion B cloud because the assignment of this emission is contested with neighboring clouds. Similarly, the algorithm characterizes the Northern Filament region as five distinct clouds rather than the single large cloud that a human decomposition produced. Despite these differences, the results of the algorithm are reassuring -- in most cases, the well-known molecular clouds are identified as single clouds and there is good agreement between their derived physical properties and the results of previous studies. \begin{deluxetable}{lcccccc} \tablecaption{\label{orion_props} Properties of Major Molecular Clouds in Orion-Monoceros} \tablewidth{0pt} \tablehead{ \colhead{Name} & \colhead{Distance}\tablenotemark{a} & \colhead{$M_{\mathrm{human}}$} \tablenotemark{a} & \colhead{$M_{\mathrm{algorithm}}$} & \colhead{$R_e$} & \colhead{$\sigma_v$} & \colhead{$\alpha$} \\ & \colhead{(pc)} & \colhead{($10^4 M_\odot$)} & \colhead{($10^4 M_\odot$)} & \colhead{(pc)} & \colhead{(km s$^{-1}$)} & } \startdata Orion A & 480 & 12. & $11. \pm 0.06$ & $18.6 \pm 0.2$ & $2.9 \pm 0.1$ & $1.6 \pm 0.1$ \\ Orion B & 500 & 9.1 & $5.6 \pm 0.08$ & $12.1 \pm 0.3$ & $1.5 \pm 0.1$ & $0.6 \pm 0.1$ \\ Orion East & 120 & 0.013 & $0.013 \pm 0.001$ & $1.2 \pm 0.1$ & \nodata & \nodata \\ Mon R2 & 800 & 12. & $11. \pm 0.5$ & $25.9 \pm 1.2$ & $1.6 \pm 0.1$ & $0.7 \pm 0 .1$ \\ Crossbones & 470 & 1.9 & $0.57 \pm 0.02$ & $11.0 \pm 0.6$ & $0.8 \pm 0.1$ & $1.5 \pm 0.3$ \\ Northern Filament & 390 & 1.9 & $1.2 \pm 0.07$ & $8.6 \pm 0.7$ & $1.7 \pm 0.1$ & $2.6 \pm 0.5$ \\ \enddata \tablenotetext{a}{Adapted from Table 2 of \citet{wil05}.} \end{deluxetable} \subsection{Decomposing Orion-like Clouds in Other Galaxies} As a final test of the decomposition and analysis algorithm, we apply the algorithm to simulated observations of the Orion-Monoceros region with beam sizes and signal-to-noise values typical of extragalactic GMC observations. We simulate observations by convolving the data set to the desired resolution and adding a convolved data cube of noise which is scaled up to give the desired peak signal-to-noise ratio in the data. We perform this for a peak signal-to-noise ($S/N$) values of 10 and 30 combined with beam sizes of 10, 20 and 50 pc. The results of the decompositions are displayed in Figure \ref{decompdemo}. At coarse resolution and low sensitivity, the fainter clouds are undetectable. However, the algorithm identifies each of Orion A, Orion B, Mon R2 and the Northern Filament in at least one of the trial data sets. At high resolution (10 pc) and peak signal-to-noise ($S/N=30$), the algorithm successfully identifies the clouds with properties that are consistent, within the uncertainties determined by the algorithm, with the masses of the clouds in Table \ref{orion_props}. The systematic effects of poor resolution manifest themselves for beam sizes $\gtrsim 20$ pc where the decomposition of blended emission results in $\sim 20\%$ variations in the properties of the clouds relative to those found in the original data set for both high and low $S/N$. For a beam size of 20 pc, the clouds are not resolved since this is over twice the size of the clouds along their minor axes. Finally, at 50 pc resolution, Mon R2 is not found by the algorithm and high $S/N$ is required to distinguish Orion A and Orion B. There is a variation of 100\% (0.3 dex) in the derived physical parameters for the clouds; this variation is reflected in the estimates of the uncertainties. As expected based on the 20 pc resolution data, the clouds are not spatially resolved for a beam size of 50 pc. An accurate decomposition of emission into typical Galactic GMCs requires a beam size $\sim 20$ pc though only a modest sensitivity is required: $S/N\gtrsim 10$. \end{appendix}
Title: Tail emission from a ring-like jet: its application to shallow decays of early afterglows and to GRB 050709
Abstract: Similar to the pulsar, the magnetic axis and the spin axis of the gamma-ray burst source may not lie on the same line. This may cause a ring-like jet due to collimation of the precessing magnetic axis. We analyze the tail emission from such a jet, and find that it has a shallow decay phase with temporal index equal to -1/2 if the Lorentz factor of the ejecta is not very high. This phase is consistent with the shallow decay phase of some early X-ray afterglow detected by {\it{swift}}. The ring-like jet has a tail cusp with sharp rising and very sharp decay. This effect can provide an explanation for the re-brightening and sharp decay of the X-ray afterglow of GRB 050709.
https://export.arxiv.org/pdf/astro-ph/0601670
\title{Tail emission from a ring-like jet: its application to shallow decays of early afterglows and GRB 050709} \volnopage{Vol.0 (200x) No.0, 000--000} \setcounter{page}{1} \author{Yuan-Chuan Zou and Zi-Gao Dai} \institute{Department of Astronomy, Nanjing University, Nanjing 210093, China.\\ \email{zouyc@nju.edu.cn, dzg@nju.edu.cn}} \date{Received~~2006 January 30; accepted~~} \authorrunning{Y. C. Zou \& Z. G. Dai} \titlerunning{Tail emission of a ring-like jet} \section{Introduction}\label{intro} It is well known that pulsars originate from the core collapse of massive stars. The average angle between the spinning axis and the magnetic axis of pulsars is about 27$^\circ$ \citep{leahy91}. Similarly, the spin axis and magnetic axis of the central engine of the gamma-ray burst may not lie on one line. As the ejecta may be collimated by the magnetic axis, while the magnetic axis is processing, so the ejecta may be in a spiral shape at first\citep[][and reference therein]{fargion05}. As the diversity of the velocities of the ejecta [as assumed in the standard fireball model of gamma-ray bursts\citep{piran05}], the spiral ejecta ejected at different times will collide and merge into one whole shell at last. These collisions just produce internal shocks of gamma-ray burst. At last, these collisions make the ejecta merge into a ring-shaped jet. Even if the the ejecta is conical, the baryon-loaded region still be ring-like \citep{eichler03}. \citet{granot05} and \citet{eichler05} have analyzed afterglows from ring-like jets. It has also been used to interpret the $h\nu_{\rm peak}-E_{\rm iso}$ relation\citep{eichler04}. Tail emission plays an important role at the times when shocks disappear. The temporal index is $-(2+\beta)$ for a cone-shaped jet, where $\beta$ is the spectral index of the emission\citep{kumar00, yamazaki05}. Considering the zero point effect of time, the light curves can be steeper during a short period\citep{nousek05, zhang05, wu06}. \citet{nousek05} and \citet{zhang05} have also shown that a shallow decay with index of about $-1/2$ follows the steep decay for most X-ray afterglows. For the X-ray afterglow of short burst GRB 050709, there is an unexpected high-flux point followed by a very steep decay\citep{fox05}. These two observations can both be explained naturally by considering the tail emission of ring-like jets. In \S \ref{sec:model} we give the expressions of tail emission from a ring-like jet. In \S\S \ref{sec:shallow} and \ref{sec:050709}, the shallow decay and X-ray afterglow of GRB 050709 are analyzed respectively. At last, we summarize our results in \S \ref{sec:conclusion}. \section{Model}\label{sec:model} Considering several ring-like sub-jets emitted from the central engine, they merge into one whole ring-like jet accompanied with internal shocks. This final ring with uniform energy density and sharp edges expands with Lorentz factor $\gamma$, as sketched in Fig. \ref{fig:ring-sketch}. Assuming the radiation from the ring-like jet begins and ceases at radius $R_c$ (and correspondingly at time $t_c$) suddenly, we calculate the tail emission from high latitudes of the ring. The relation is $R_c \simeq 2\eta^2 c t$, where $\eta$ is the mean Lorentz factor of the internal shocks. The relation between the latitude angle $\theta$ and the observed time $t$ is \begin{equation} R_c(1-\cos\theta)=c(t-t_c)/(1+z), \end{equation} where $z$ is the cosmological redshift. Neglecting the depth of the ejecta and the emission from time equal arrival surface, and defining the emissivity $I'_{\nu'}$ per unit area in the comoving frame, which is uniform in the whole ring, the flux density in the observer's frame is \begin{equation} f_{\nu}(t>t_c) =\frac{ I'_{\nu'}}{4\pi D_L^2} \mathscr{D}^2 \frac{{\rm d}S}{{\rm d}t/(1+z)}, \label{eq:f_nu_general} \end{equation} where $D_L$ is the luminosity distance, $\mathscr{D}=1/[\gamma(1-\sqrt{1-1/\gamma^2} \cos\theta)]$ is the Doppler factor, and ${\rm d}S$ is the emitted area during a period ${\rm d}t$. At early times when $\theta < \theta_w$, the tail emission is the same as the case of an on-axis conical jet, which has been investigated by many authors \citep{kumar00, fan05}. There are two limiting cases: for $\theta \ll 1/\gamma$, \begin{equation} f_{\nu}(t>t_c) \propto \delta t^0, \label{eq:11} \end{equation} and for $1 \gg \theta \gg 1/\gamma$, \begin{equation} f_{\nu}(t>t_c) \propto \delta t^{-(2+\beta)}, \label{eq:12} \end{equation} where $\delta t \equiv t-t_c$. Here we consider the emission as a single power-law profile $I'_{\nu'} \propto \nu'^{-\beta}$, which is valid for the high frequency emission $\nu' > \max(\nu'_c, \nu'_m)$, where $\nu'_c$ is the cooling frequency and $\nu'_m$ is the typical frequency of synchrotron emission. In the case $\theta > \theta_w$, the width of the ring can be neglected, and the flux density \begin{equation} f_{\nu}(t>t_c) \propto \mathscr{D}^{-(2+\beta)} \frac{\sin{\theta_p} \cos{(\theta/2)}}{\sin{\theta} \sqrt{1-\left(\frac{\sin {(\theta/2)}} {\sin {\theta_p}}\right)^2}}. \label{eq:f_line} \end{equation} There are two limiting cases in which equation (\ref{eq:f_line}) can be simplified. For $\theta \ll 1/\gamma$, \begin{equation} f_{\nu}(t>t_c) \propto \delta t^{-1/2}, \label{eq:21} \end{equation} and for $1 \gg \theta \gg 1/\gamma$, \begin{equation} f_{\nu}(t>t_c) \propto \delta t^{-(5/2+\beta)}. \label{eq:22} \end{equation} \section{Shallow Decay of Early X-ray Afterglow}\label{sec:shallow} Statistics of the early X-ray afterglows has shown that there is a shallow decay phase with temporal index about $-1/2$\citep{nousek05, zhang05}. This corresponds to the case: $1/\gamma > \theta > \theta_w$, and can be described by equation (\ref{eq:21}), where the temporal index is just $-1/2$. As the general shallow decay lasts from $10^2-10^3$s to $10^3-10^4$s \citep[Fig. 1 in][]{zhang05}, this gives limits: the Lorentz factor of the emitting shell $\gamma < 7.3 (1+z)^{1/2} R_{c,16}^{1/2} \delta t_{3.5}^{-1/2}$, and the width of the ring-like jet $\theta_w < 1.4\times 10^{-2} (1+z)^{-1/2} R_{c,16}^{-1/2} \delta t_{2.5}^{1/2}$. (The conventional donation $Q=Q_k\times 10^k$ is used throughout this paper.) This implies that the shallow decay component originates from the shocked shells with low Lorentz factors, while these shocks may be formed due to ejected sub-shells with different Lorentz factors. This model can answer the following questions: Firstly, why is there a steep decay before the shallow decay appears in general case? In \citet{zhang05}, it is general that the temporal index of this steep decay is less that $-3$. The answer is that the two power law decays originate from two different emitting shells with different Lorentz factors. The steep decay corresponds to the greater Lorentz factor shell, which satisfies $1/\gamma < \theta$, and the temporal indices are $-(2+\beta)$ or $-(5/2+\beta)$ corresponding to equations (\ref{eq:12}) and (\ref{eq:22}) respectively. The steep decay may become steeper because of the zero time selection effect\citep{wu06}. Secondly, why is there no spectral evolution before and after the break time from the shallow decay phase to the steep decay phase. This is also mentioned with spectral index value $\sim -1$ in \citet{zhang05}. It is believed that, after the break, the afterglow becomes a ``normal'' afterglow. It is possible that the tail emission phase and the ``normal'' afterglow emission phase are both in the case $\nu_X > \{\nu_m,\nu_c\}$ (corresponding to the spectral index $-p/2$) and thus the spectra are the same. Thirdly, since the shallow phase and the steep phase originate from different sources, how to understand the conjunction at the break time \citep[also can be seen in Fig. 1 in][]{zhang05}? As time goes on, the case converts from $1/\gamma > \theta > \theta_w$ to $\theta > 1/\gamma > \theta_w$, and then the light curve of the tail emission decay has a temporal index $-(5/2+\beta)$. This is steeper than the ``normal'' afterglow with temporal index $\sim -1.2$. Some time later, the ``normal'' afterglow will exceed the tail emission definitely, as in the case GRB 050525a \citep[Fig. 1 in][]{nousek05} (at about 3000s, there is a steep decay). However, GRB 050315\citep{vaughan05} can be classified into the case that ``normal'' afterglow exceed the shallow tail emission before the tail emission breaks to steep phase. \section{X-ray Afterglow of GRB 050709}\label{sec:050709} GRB 050709 is a short burst with duration 0.3 s, and five points of X-ray emission after the burst were obtained by Swift and Chandra \citep{fox05}. Figure \ref{fig:050709} shows the fit by assuming that the latter four points are the tail emissions from the first point, with parameters $R_c=7.7\times 10^{16}$cm, $\gamma=15.5$, $\beta=1.1$, $\theta_p=0.5$ and $\theta_w=0.005$. As the first X-ray point occurs at time about 100s, the radius $r \simeq 2\eta^2 c t/(1+z) \simeq 5.2 \times 10^{16} \eta_{2}^2 t_2$cm, is consistent with the value of the parameter $R_c$. As a short burst has less total energy than a long burst does, the ejected shell can be decelerated quickly. The Lorentz factor at $R_c$ is $\gamma \simeq {E_{\rm{iso}}}/({\pi R_c^3 n m_p c^2}) \simeq 26 E_{\rm{iso},50}^{1/2} n_1^{-1/2} R_{c,16.5}^{-3/2}$, where the external medium density $n$ is chosen equal to 1$\rm{cm}^{-3}$ because the host is a star-forming galaxy. Therefore, the parameters chosen to fit the X-ray data are reasonable for this short burst. We can see four stages for this tail light curve: first, a horizontal phase corresponds to the case $\theta < \theta_w < 1/\gamma$; second, a shallow decay with temporal index $-1/2$ corresponds to the case $\theta_w < \theta < 1/\gamma$; third, a sharper decay with temporal index $-(2.5+\beta)$ corresponds to the case $\theta_w < 1/\gamma < \theta$; and finally, a tail cusp with sharp rising and very sharp decay, which comes from the end of the ring. We should note that the solid line doesn't fit the data very well, especially that the tail cusp of the model can't reach to the observed data. This shortage may be overcome by considering a non-uniform ring-like jet or some other mechanism. However, its unique feature, which the emission after the tail cusp decays very sharply, is consistent with the last two observed points. On the other hand, it is possible that the second, third and fifth points in this figure belong to the ``normal'' afterglow from an external shock. \section{Conclusions}\label{sec:conclusion} Enlightened from pulsars, we suggest that the magnetic axis and the spin axis of a gamma-ray burst source point to different orientations. The ejecta along the magnetic axis will form a ring finally. Gamma-ray emission will be observed if the observer locates in the solid angle of the ring. We have investigated the tail emission from a ring-like jet. We find that the early shallow decay phase and the late re-brightening of the X-ray emission of GRB 050709 can be explained. Note that the shallow decay phase is only possible in the low Lorentz factor cases. For the case $1/\gamma < \theta_w$, only the steep one appears. As the tail emission from the shells with high Lorentz factors decays very quickly, the main emissions will be dominated by the slower shells at later times. YCZ thanks helpful discussions with Jia Wang and Xuefeng Wu. This work was supported by the National Natural Science Foundation of China (grants 10233010 and 10221001).
Title: On the accuracy of the ALI method for solving the radiative transfer equation
Abstract: We solve the integral equation describing the propagation of light in an isothermal plane-parallel atmosphere of optical thickness $\tau^*$, adopting a uniform thermalization parameter $\epsilon$. The solution given by the ALI method, widely used in the field of stellar atmospheres modelling, is compared to the exact solution. Graphs are given that illustrate the accuracy of the ALI solution as a function of the parameters $\epsilon$, $\tau^*$ and optical depth variable $\tau$.
https://export.arxiv.org/pdf/astro-ph/0601341
\title{On the accuracy of the ALI method for solving the radiative transfer equation} \titlerunning{Accuracy of the ALI method} \author{L. Chevallier\inst{1} \and F. Paletou \inst{2}\thanks{Present address: Observatoire Midi-Pyr\'{e}n\'{e}es, Laboratoire d'Astrophysique (UMR 5572), 14 avenue E. Belin, F-31400 Toulouse Cedex.} \and B. Rutily \inst{1}} \institute{ Centre de Recherche Astronomique de Lyon (UMR 5574 du CNRS), Observatoire de Lyon, 9, avenue Charles Andr\'{e}, 69561 Saint-Genis-Laval cedex, France\\ \email{loic.chevallier@obs.univ-lyon1.fr} \\ \email{rutily@obs.univ-lyon1.fr} \and Observatoire de la C\^{o}te d'Azur, D\'{e}partement G. D. Cassini (UMR 6529 du CNRS), BP 4229, 06304 Nice cedex 4, France \\ \email{fpaletou@mail.ast.obs-mip.fr} } \date{Received 25 April 2003 / Accepted 30 July 2003} \abstract{We solve the integral equation describing the propagation of light in an isothermal plane-parallel atmosphere of optical thickness $\tau^*$, adopting a uniform thermalization parameter $\epsilon$. The solution given by the ALI method, widely used in the field of stellar atmospheres modelling, is compared to the exact solution. Graphs are given that illustrate the accuracy of the ALI solution as a function of the parameters $\epsilon$, $\tau^*$ and optical depth variable $\tau$. \keywords{Radiative transfer -- Methods: numerical -- Stars: atmospheres} } \section{Introduction} The solution of the radiative transfer equation (RTE) is at the heart of the stellar atmospheres modelling, since this equation has to be solved typically thousands of times in order to construct a realistic model. It is thus crucial to get a clear idea of the accuracy with which the RTE is solved, and the effect it has on the determination of the main physical quantities of the model: populations, electron density, temperature, etc. In this article, we focus on the first point checking the ALI method for solving the integral form of the RTE, since this method is nowadays at the basis of most numerical schemes used to determine the radiation field in stellar atmospheres. We recall that ``ALI'' means Accelerated (or Approximate) Lambda Iteration, the Lambda operator being defined by Eqs.~(\ref{eq_lambda})-(\ref{eq_e1}) below for the scattering law we adopt here. The ALI code used in this paper is a combination of an accelerated iterative method (with a diagonal $\Lambda$-operator) and a formal solver based on parabolic short characteristics. Recent reviews on this approach are Paletou (2001), Hubeny (2003) and section 3 of Trujillo Bueno (2003). The accuracy of our ALI code is tested while applied to a well-known problem consisting of a homogeneous, isothermal slab with isotropic and monochromatic light scattering (Sec.~\ref{sec_2}). Indeed, this idealized problem can be solved exactly, which allows for a direct comparison with the solution given by the ALI method. This problem is very simple on physical grounds but implies analytical and numerical calculations that are far from trivial. It contains the seeds of most of the difficulties met when solving the RTE in a thick, highly scattering medium. It thus provides an excellent test for numerical codes since very accurate analytical solutions are available (Sec.~\ref{sec_3}). After a brief description of our ALI code, we move to the numerical tests in Sec.~\ref{sec_4}, which is the main part of this paper. The link with previous studies on the subject (Trujillo Bueno \& Fabiani Bendicho 1995, Trujillo Bueno \& Manso Sainz 1999) is finally commented in Sec.~\ref{sec_5}. \section{The standard radiative transfer problem}\label{sec_2} This problem consists in solving the RTE in a homogeneous plane-parallel atmosphere of optical thickness $\tau^*>0$ (possibly infinite); light scattering is assumed to be isotropic and monochromatic. It is furthermore supposed that the matter is in local thermodynamical equilibrium with uniform temperature $T$ through the atmosphere. The thermal source function at any frequency is then $\epsilon B(T)$, where $\epsilon$ is the (spatially invariant) photon destruction probability per scattering and $B(T)$ the Planck function at temperature $T$ (frequency dependence is not mentioned). In the absence of any external source of radiation, this problem reduces to solving the following integral equation for the source function $S$ (Mihalas 1978): \begin{equation} \label{eq_s1} S(\tau)=\epsilon B(T)+(1-\epsilon)(\Lambda S)(\tau) \, , \end{equation} where the $\Lambda$-operator for isotropic and monochromatic scattering is \begin{equation} \label{eq_lambda} (\Lambda S)(\tau)= \frac{1}{2}\int_0^{\tau^*}E_1(\vert \tau-\tau^{\prime}\vert)S(\tau^{\prime}) \,\mathrm{d}\tau^{\prime} \, . \end{equation} Here, $E_1$ is the first exponential integral function as defined by \begin{equation} \label{eq_e1} E_1(\tau)=\int_0^1\exp(-\tau/\mu)\frac{\,\mathrm{d}\mu}{\mu}\quad(\tau>0) \, . \end{equation} We remind the reader that Eq. (\ref{eq_s1}) models the multiple scattering of photons of frequency $\nu$ assuming that 1) the scattering is monochromatic (or coherent) if $\nu$ belongs to a continuum, 2) the line profile is rectangular (Milne profile) if $\nu$ belongs to a spectral line (see, e.g., Ivanov 1973, p. 57). The solution to problem (\ref{eq_s1}) is $S(\tau)=S(\epsilon,\tau^*,\tau)B(T)$, where $S(\epsilon,\tau^*,\tau)$ satisfies the integral equation \begin{equation} \label{eq_s2} S(\epsilon,\tau^*,\tau)=\epsilon+(1-\epsilon)(\Lambda S)(\epsilon,\tau^*, \tau) \end{equation} depending on parameters $\epsilon$ and $\tau^*$. Note that this function is symmetrical about the $\tau$-mid-plane: $S(\epsilon,\tau^*,\tau)= S(\epsilon,\tau^*,\tau^*-\tau)$. This equation is the integral formulation of the RTE in our model; it specifies the {\em standard radiative transfer problem} we intend to solve analytically (Sec.~\ref{sec_3}) and numerically (Sec.~\ref{sec_4}). \section{Analytical solution of the standard problem}\label{sec_3} There are many analytical methods for solving the integral equation (\ref{eq_s2}). The classical approach, recently reviewed by Chevallier \& Rutily (2003, hereafter Paper I), involves the basic auxiliary functions of radiative transfer theory in plane-parallel geometry, namely the $H$-function for a semi-infinite space, and the $X$- and $Y$-functions for a finite slab (Chandrasekhar 1960). The $H$-function depends on the parameter $\epsilon$ and on an angular variable $\mu$, taken as positive hereafter. In addition the $X$- and $Y$-functions depend on $\tau^*$, and we have $X(\epsilon,\tau^*,\mu)\to H(\epsilon,\mu)$ and $Y(\epsilon,\tau^*,\mu)\to 0$ as $\tau^* \to +\infty$. The zero-order moments of the functions $H$, $X$, and $Y$ yield the surface values of the solution $S$ to (\ref{eq_s2}). The moment of the $H$-function is defined and given by \begin{equation} \alpha_0(\epsilon)=\int_0^1H(\epsilon,\mu)\,\mathrm{d}\mu=\frac{2}{1+\sqrt{\epsilon}} \, , \end{equation} and those of the $X$- and $Y$-functions defined as \begin{equation} \alpha_0(\epsilon,\tau^*)=\!\int_0^1 \!X(\epsilon,\tau^*\!,\mu)\,\mathrm{d}\mu\;,\quad \beta_0(\epsilon,\tau^*)=\!\int_0^1 \!Y(\epsilon,\tau^*\!,\mu)\,\mathrm{d}\mu \end{equation} are related by \begin{equation} \left[ 1-\frac{1-\epsilon}{2}\alpha_0(\epsilon,\tau^*) \right]^2 - \left[ \frac{1-\epsilon}{2}\beta_0(\epsilon,\tau^*) \right]^2=\epsilon \, . \end{equation} There is no exact expression of these moments. In a semi-infinite atmosphere, the surface value of the solution $S$ to (\ref{eq_s2}) is \begin{equation} S(\epsilon,0)=1-\frac{1-\epsilon}{2}\alpha_0(\epsilon)=\sqrt{\epsilon} \end{equation} and it is \begin{equation} \label{eq_s4} S(\epsilon,\tau^*,0)=1-\frac{1-\epsilon}{2}[\alpha_0(\epsilon,\tau^*)+\beta_0(\epsilon,\tau^*)] \end{equation} in a finite slab. As $S$ is symmetrical about the $\tau$-mid-plane, $S(\epsilon,\tau^*,\tau^*)= S(\epsilon,\tau^*,0)$. These relations were first derived by Sobolev (1957, 1958). The former result is the famous ``$\!\sqrt{\epsilon}$-law'' for semi-infinite media. The latter one is less known; it requires a table of moments $( \alpha_0, \,\beta_0 )$ for numerical applications. Such tables are available in the literature: see references in Van de Hulst (1980, p. 225-227). Very accurate surface values of the $S$-function can also be found in Paper I. The calculation of the function $S$ within the slab is discussed in detail in Paper I, which contains ten-figure tables of $S(\epsilon, \tau^*, \tau)$ for $(\epsilon, \tau^*)$ = $(0.5,2)$, $(10^{-2},20)$, $(10^{-4},2000)$ and $(10^{-8}, 2\times 10^8)$. In a half-space, the internal solution is known since the end of the 50's and it can be expressed in closed-form in terms of the $H$-function. In a finite slab, the solution involves two non-classical auxiliary functions $\zeta_+$ and $\zeta_-$, that are implicitly defined by Fredholm integral equations over $[0, 1]$. These equations can be solved very accurately, so that the solution in a finite slab is nearly as accurate as in a half-space. The accuracy is estimated at better than $10^{-10}$ for any value of $\epsilon, \tau^*$ and $\tau$, which means that the solution given in Paper I can safely be used as an accuracy test of the ALI code. The general behavior of the $S$-function is shown in Fig.~\ref{fig1}, which illustrates the Table 3 of Paper I ($\epsilon = 10^{-4}$ and $\tau^* = 2000$). It can be seen that the solution $S$ tends to 1 for large values of $\tau$ and that it drops when $\tau$ is close to the thermalization depth $1/k(\epsilon)\approx 58$ for $\epsilon=10^{-4}$, where $k(\epsilon)$ is defined in Paper I. It tends steeply to the surface value $S(0)$ as $\tau$ tends to 0, {\em with an infinite derivative at 0}. It is regrettable that the generally adopted logarithmic scale in $\tau$ obscures this essential last point, as seen when comparing the solid and dashed curves of Fig.~\ref{fig1}. The explanation lies in the fact that $\partial S/\partial(\log \tau)=\tau \,\partial S/\partial \tau\to 0$ even if $\partial S/\partial \tau\sim E_1(\tau) \to +\infty$. \section{Comparison with ALI numerical solutions}\label{sec_4} In this section we compare in detail the analytical solution described in the previous section to the one given by our ALI code. This code uses a diagonal approximate $\Lambda$-operator (Olson et al. 1986). At each iteration, a formal solver has to be used in order to calculate the transform of the source function by the $\Lambda$-operator. Inserting the definition (\ref{eq_e1}) of the $E_1$-function into Eq. (\ref{eq_lambda}) and inverting the order of integrations, the so-called formal solution to the RTE is first calculated \begin{eqnarray} \label{eq_i} \lefteqn{I(\tau ,\mu ) = \left\lbrace \! \begin{array}{ll}\displaystyle - \frac{1}{\mu } \! \int _{0}^{\tau }S(\tau^{\prime})\exp \left[ (\tau -\tau ^{\prime})/\mu \right] \,\mathrm{d}\tau ^{\prime} & \rm if\;-1\le\mu <0, \\ \displaystyle S(\tau ) & \rm if \;\mu =0, \\ \displaystyle +\frac{1}{\mu } \! \int_{\tau }^{\tau^*}S(\tau ^{\prime})\exp \left[ -(\tau ^{\prime}-\tau )/\mu \right] \,\mathrm{d}\tau ^{\prime} & \rm if\; 0<\mu \le +1. \end{array} \right. } \nonumber \\ \end{eqnarray} Then the $\Lambda$-transform of the source function is derived, since it is here the associated mean intensity \begin{equation} \label{eq_ls} (\Lambda S)(\tau )=\frac{1}{2}\int _{-1}^{+1}I(\tau ,\mu ) \,\mathrm{d}\mu \, . \end{equation} The formal solution (\ref{eq_i}) is calculated following the method of short characteristics whose basic elements can be found in Olson \& Kunasz (1987) and Kunasz \& Auer (1988). It was further improved by the implementation of monotonic interpolation for multi-dimensional applications (Auer \& Paletou 1994) and by Fabiani Bendicho \& Trujillo Bueno (1999) for three-dimensional applications with horizontal periodic boundary conditions. In the present paper, we used parabolic short characteristics. The $\mu$-integration in (\ref{eq_ls}) is performed with the help of a Gaussian quadrature. A numerical acceleration scheme is used so as to improve the rate of convergence of ALI: this is the so-called Ng-acceleration introduced in the field of radiative transfer by Auer (1987, 1991; see also Rybicki \& Hummer 1991). We have calculated the relative error \begin{equation} \label{eq_d} d(\epsilon, \tau^*,\tau)=\left| \frac{S_\mathrm{ALI}(\epsilon,\tau^*,\tau)-S(\epsilon,\tau^*, \tau)}{S(\epsilon,\tau^*,\tau )}\right| \end{equation} at various optical depths, where $S(\epsilon,\tau^*,\tau)$ is the analytical solution of Sec.~\ref{sec_3} and $S_\mathrm{ALI}(\epsilon,\tau^*,\tau)$ is the solution given by the ALI code. This error corresponds to the ``true error'' defined by Auer, Fabiani Bendicho \& Trujillo Bueno (1994), who used a finer grid to calculate $S(\epsilon, \tau^*, \tau)$. We introduce also the maximum value of $d(\epsilon,\tau^*,\tau)$ when the $\tau$-variable covers the domain $[0,\tau^*]$, viz. \begin{equation} \label{eq_dm} d_\mathrm{M}(\epsilon,\tau^*)=\max_{0\leq\tau\leq\tau^*} d(\epsilon,\tau^*,\tau) \, . \end{equation} Of course $d$ and $d_\mathrm{M}$ depend on the number of iterations $N$ performed by the ALI code during each run. Finally we define $N_\mathrm{c}$ as the number of iterations used to reach convergence, which is the smallest value of $N$ satisfying the condition $|1-d_\mathrm{M}(N)/d_\mathrm{M}(+\infty)| < \varepsilon_\mathrm{c}$, where $d_\mathrm{M}(+\infty) = d_\mathrm{M}(N=10\,000)$ and $\varepsilon_\mathrm{c}$ is arbitrarily set to $0.01$ in the present paper. The slab optical depth is discretized using a logarithmic grid, symmetric with respect to the mid-plane, with $n_{\tau}$ points per decade, including the $\tau=0$ point, the next point denoted by $\tau_\mathrm{m}$, and the last point $\tau=\tau^*/2$. The angular integration in Eq. (\ref{eq_ls}) is performed with a symmetric grid containing $n_\mu$ Gauss-Legendre points in $[0,1]$. There is no frequency integration since light scattering has been supposed monochromatic. In most of our calculations, we chose the values $\tau_\mathrm{m} = 10^{-4}$, $n_\tau=9$, and $n_\mu=5$ (i.e., values quite often adopted for stellar atmospheres modelling). Some values of $(\epsilon,\tau^*)$ may be (0.01, 20) for a continuum, ($10^{-4}, 2000$) for an ``average'' spectral line and ($10^{-8}, 2\times10^8$) for a strong spectral line. The quantities of interest are the maximum relative error $d_\mathrm{M}$ and the number of iterations to reach convergence $N_\mathrm{c}$, which depend on $\epsilon$, $\tau^*$ and numerical parameters $\tau_\mathrm{m}$, $n_\tau$, $n_\mu$ and $N$. We first study the variation of $d(\epsilon,\tau^*,\tau)$ with $\tau$. Then, we study the influence of $\epsilon$, $\tau^*$, $n_\tau$ on $d_\mathrm{M}$ and $N_\mathrm{c}$, for given $\tau_\mathrm{m}$, $n_\mu$ and for $N=N_\mathrm{c}$. \subsection{The influence of $\tau$ and $N$} Figure~\ref{fig2} shows the variation of the relative error $\tau \to d(\epsilon,\tau^*,\tau)$ for the three selected values of $\epsilon$ and $\tau^*$. It can be seen that the accuracy (i.e. maximum relative error) of our ALI code is about $5\times 10^{-3}$ for the three cases studied here. Relative error is close to this accuracy when $\tau$ is smaller than the thermalization depth $1/k(\epsilon)$ of the atmosphere (black dots on the curves), and significantly improves beyond (up to $10^{-8}$). In photon mean free path units, the thermalization depth is 6, 58 and 5774 for $\epsilon = 10^{-2}, 10^{-4}$ and $10^{-8}$ respectively. Note that the surface relative error is a good estimator of the accuracy in spectral lines, but not in the continuum. The iterative algorithm was stopped after $N=1000$ iterations. Figure \ref{fig3} shows that this number ensures convergence of the ALI code, the convergence being slower when $\epsilon\to 0$ and $\tau^* \to +\infty$. Irregular steps in these curves are due to the Ng acceleration process, here operated every four iterations. \subsection{The influence of $\tau_\mathrm{m}$} We point out that $d_\mathrm{M}$ is improved when $\tau_\mathrm{m}$ goes to 0, up to a given value where the accuracy is constant. Including the $\tau=0$ point in the grid and choosing $\tau_\mathrm{m} < 10^{-2}$, the best accuracy is warranted. Excluding the $\tau=0$ point from the grid has no influence on accuracy if we choose $\tau_\mathrm{m} < 10^{-4}$. The standard choice $\tau_\mathrm{m}=10^{-4}$ is thus correct, and this value will be adopted hereafter. \subsection{The influence of $n_\tau$ and $n_\mu$} In Fig.~\ref{fig4} are shown the variations of $d_\mathrm{M}$ in an average line as a function of $n_\tau$ and $n_\mu$. The maximum relative error $d_\mathrm{M}$ decreases with increasing number $n_\tau$ of $\tau$-grid points, and it is sensitive to the choice of the number $n_\mu$ of angular grid points up to an optimal value $n_\mu^\mathrm{(opt)}$; the latter is defined as the smallest value of $n_\mu$ for which the condition $|1-d_\mathrm{M}(n_\mu)/d_\mathrm{M}(64)| < 0.01$ holds. The accuracy does not increase with a finer $\mu$-grid. We note that $n_\mu^\mathrm{(opt)} < n_\tau$ and that we have a linear dependence of this optimal value on $n_\tau$: $n_\mu^\mathrm{(opt)} = 0.8 n_\tau + 2.7$ (dashed curve). This fit is still valid for strong lines. As seen in Fig.~\ref{fig5} (same as Fig.~\ref{fig4} for the continuum), the fit for lines cannot be applied to the continuum, for which $n_\mu^\mathrm{(opt)} > n_\tau$. It is still possible to define and calculate an optimal value for $n_\tau < 18$, using the relation $n_\mu^\mathrm{(opt)} = 1.7 n_\tau + 1.3$ (dashed curve). It appears that our ALI code is more demanding in angular resolution when solving the problem (\ref{eq_s2}) in a continuum than in a line. The results of Figs.~\ref{fig4} and \ref{fig5} are detailed in Fig.~\ref{fig6} for the three chosen values of $(\epsilon,\tau^*)$ and $n_\mu = 64$. We remark that the accuracy improves with $n_\tau$ for each couple $(\epsilon,\tau^*)$, more significantly in the continuum than in lines. Figure~\ref{fig7} gives the number of iterations $N_\mathrm{c}$ used to reach convergence for $\varepsilon_\mathrm{c} = 10^{-2}$ and $n_\mu = 64$. The number $N_\mathrm{c}$ appreciably increases with $n_\tau$ in the lines: it is indeed well known that the rate of convergence of the one-point ALI iterative scheme drops for an increasing refinement of the spatial grid (Olson et al. 1986); however improvements were already proposed (e.g., Trujillo Bueno \& Fabiani Bendicho 1995) in order to increase significantly the rate of convergence of ALI-based methods. \subsection{The influence of $\epsilon$ and $\tau^*$} The maximum relative error $d_\mathrm{M}$ and number of iterations $N_\mathrm{c}$ are shown in Figs.~\ref{fig8} and \ref{fig9} for an extended range of $(\epsilon,\tau^*)$ after the ALI code has converged ($N=10\,000$, $n_\tau=9$ and $n_\mu=5$ are fixed here). As seen in Fig.~\ref{fig8}, the accuracy hardly changes as $\epsilon\to 0$ and $\tau^*\to +\infty$, but the number of iterations needed to achieve convergence increases substantially (see Fig.~\ref{fig9}). When $\epsilon>0.1$ the accuracy no longer depends on values of $\tau^*$. The comparison of Figs.~\ref{fig6} and \ref{fig8} leads to a disagreement since the parameter $n_\mu$ is set to different values, 64 and 5 respectively. In Fig.~\ref{fig9}, we have plotted the parameter $N_\mathrm{c}$ as a function of $\epsilon$ and $\tau^*$. When $\epsilon\to 0$ and $\tau^* \to +\infty$, we note a slowing down of the convergence (already seen in Fig.~\ref{fig3}). \section{Comments on previous studies}\label{sec_5} Now we compare our results for the monochromatic scattering problem with those published by Trujillo Bueno \& Fabiani Bendicho (1995) and Trujillo Bueno \& Manso Sainz (1999). Although these two papers concern mainly the development of new iterative methods for radiative transfer applications (for the unpolarized and polarized cases respectively) they give some information on the accuracy of the numerical solutions obtained for spatial grids of increasing resolution. In Table~\ref{tab1}, good agreement is found between our values of $N_\mathrm{c}$, $d(\epsilon, \tau^*, 0)$ and those given by these authors; our surface relative error $d(\epsilon, \tau^*, 0)$ corresponds to their surface true error $T_\mathrm{e}$. The observed small discrepancies are possibly due to the different scattering laws adopted, leading to different $\Lambda$-operators. \begin{table*} \caption{Comparison of results obtained with our (ALI+Ng) code and previous ones. Our $N_\mathrm{c}$ is defined by $\varepsilon_\mathrm{c} = 0.01$, while values from other authors are based on a graphical guess $\varepsilon_\mathrm{c} \approx 0.05$. The optical thickness is $\tau^* = 2\times 10^8$. Note that in Trujillo Bueno \& Manso Sainz (1999), $N_\mathrm{c}$ values (in parenthesis) are given for a non-accelerated Jacobi scheme. These numbers have been divided by 2 in order to estimate the number of iterations when Ng acceleration is used.} \label{tab1} \centering \begin{tabular}{llllllll} \hline \noalign{\smallskip} $\epsilon,n_\tau,n_\mu$ & \multicolumn{2}{l}{JTB \& PFB (1995)} & \multicolumn{2}{l}{JTB \& RMS (1999)} & \multicolumn{3}{l}{This article} \\ \noalign{\smallskip} \hline \noalign{\smallskip} & $N_\mathrm{c}$ & $T_\mathrm{e}$ & $N_\mathrm{c}$ & surface $T_\mathrm{e}$ & $N_\mathrm{c}$ & $d_\mathrm{M}(\epsilon,\tau^*)$ & $d(\epsilon,\tau^*,0)$ \\ \noalign{\smallskip} \hline\noalign{\smallskip} $10^{-6}$, 9, 1 & 180 & $3.5\times 10^{-3}$ && & 179 & $8.6\times 10^{-2}$ & $4.1\times 10^{-3}$ \\ $10^{-12}$, 9, 1 & 1300 & $3.5\times 10^{-3}$ &&& 985 & $8.6\times 10^{-2}$ & $4.1\times 10^{-3}$ \\ $10^{-4}$, 5, 64 && & 33 (65) & $2\times 10^{-2}$ & 42 & $1.5\times 10^{-2}$ & $1.3\times 10^{-2}$\\ $10^{-4}$, 9, 64 && & 75 (150) & $3\times 10^{-3}$ & 88 & $4.5\times 10^{-3}$ & $3.9\times 10^{-3}$\\ $10^{-4}$, 18, 64 && & 175 (350) & $4\times 10^{-4}$ & 184 & $1.1\times 10^{-3}$ & $9.0\times 10^{-4}$ \\ $10^{-4}$, 36, 64 && & 400 (800) & $5\times 10^{-5}$ & 356 & $2.4\times 10^{-4}$ & $1.9\times 10^{-4}$ \\ \hline \end{tabular} \end{table*} In Fig.~\ref{fig10} it is shown how far the semi-infinite exact result $S(\epsilon,0)=\sqrt{\epsilon}$ agrees with the finite one. This comparison is useful since many authors use the $\sqrt{\epsilon}$-law as a check for their calculations in thick slabs. We plot the relative difference $\delta_1=|\,1-S(\epsilon,\tau^*,0)/\sqrt{\epsilon}\,|$ as a function of $k(\epsilon) \tau^*$, and the quantity $\exp\left(-k(\epsilon)\tau^*\right)$ which characterizes the validity of the $\sqrt\epsilon$-law (solid curve). For $k(\epsilon)\tau^* > 100$, the accuracy limit of our code is reached, which explains that the solid curve no longer fits the dots. This law is very well satisfied in lines ($k(\epsilon)\tau^* \approx 34$ in an average line) but not enough in the continuum ($k(\epsilon)\tau^* \approx 3.3$). We conclude that the $\sqrt\epsilon$-law can be used as a test for the ALI code when $k(\epsilon)\tau^* > 10$, since then $\sqrt\epsilon$ is an approximation to the surface value with an accuracy better than $10^{-4}$, as seen in Fig.~\ref{fig10}. In Trujillo Bueno \& Fabiani Bendicho (1995), the Eddington approximation is used as the reference solution for a one-point angular quadrature $n_\mu=1$ with $\mu=\pm 1/\sqrt 3$. The analytical expression of the Eddington approximation in a finite slab is: \begin{eqnarray} \label{eq_sedd} \lefteqn{S_\mathrm{E}(\epsilon,\tau^*,\tau) = 1 - (1-\epsilon) \frac{\exp\left(-\sqrt{3\epsilon}\,\tau\right) + \exp\left(-\sqrt{3\epsilon}\,(\tau^* - \tau)\right)}{1+\sqrt\epsilon+(1-\sqrt\epsilon)\exp\left(-\sqrt{3\epsilon}\,\tau^*\right)}. } \nonumber \\ \end{eqnarray} This is the exact solution of the monochromatic scattering problem when the mean intensity is calculated with the above-mentioned one-point angular quadrature. However, as is well-known, it gives only an approximation to the exact (i.e., multi-angle) solution of the full problem (\ref{eq_s1})-(\ref{eq_e1}). In other words, the true error given by Trujillo Bueno \& Fabiani Bendicho (1995) is relative to the $n_\mu=1$ monochromatic scattering problem only, it does not give information on the error that would have been got by comparing the numerical solution to the $n_\mu=1,3,5, \ldots$ problem to the exact multi-angle solution ($n_\mu=\infty$). In fact, as given in Table~\ref{tab1} for $n_\mu=1$, when the solution of the $n_\mu=1$ problem is compared to the exact multi-angle solution, we find that the maximum error for $n_\tau=9$ is $8.6\times 10^{-2}$. The latter represents the maximum relative difference between the Eddington approximation and the exact solution (Fig.~\ref{fig11}). A similar investigation, but for the two-level atom resonance-line scattering polarization problem, was carried out by Trujillo Bueno \& Manso Sainz (1999), whose Table 3 gives the surface true-error values of the fractional atomic polarization for $n_\mu=3,5,7,11, \ldots, 61$. \section{Conclusion} Our ALI code has been subjected to a wide range of tests, revealing at the same time its capabilities and its limits. Before developing these two points, we note that our conclusions are relative to the particular code we have used (based on Jacobi's method), specifically solving the standard problem (\ref{eq_s1})-(\ref{eq_e1}). The accuracy of the code is ultimately determined by the accuracy of the formal solver we have used (parabolic short characteristics). We have checked the great robustness of our code, which is certainly its most remarkable feature. It is able to solve the standard problem (\ref{eq_s1})-(\ref{eq_e1}) for a wide range of input parameters $\epsilon$ and $\tau^*$, with no important lack of performance when $\epsilon \to 0$ and/or $\tau^* \to +\infty$. However, the lowest accuracy of the ALI numerical solutions happens in the outermost layers of a star, corresponding to $\tau$ lower than the thermalization depth $1/k(\epsilon)$, these layers forming, by definition, the atmosphere of the star. The accuracy of our code is not better than, say $10^{-2}$, when we choose $n_\tau = 9$, $n_\mu = 5$ and limit the number of iterations to $N<100$, as it is currently done in stellar atmospheres modelling. To improve the accuracy of the calculations up to $\sim 10^{-3}$, the parameters $n_\tau$, $n_\mu$ and $N$ should be set to larger (but today impractical) values when solving the radiative transfer equation on a large frequency spectrum, i.e. at thousands of frequencies. Indeed we pointed out a truly noticeable improvement of the accuracy when using finer grids in $\tau$ or $\mu$. Such an observation was made easier by the use of a very accurate reference solution. Of course, increasing the level of refinement of both spatial and angular quadratures has a strong impact upon the number of iterations needed for convergence. However, to overcome this difficulty while keeping the same accuracy on the numerical solutions, methods based on Gauss-Seidel and successive over-relaxation iterations were already proposed by Trujillo Bueno \& Fabiani Bendicho (1995). Another important question is relative to the propagation of errors in a stellar atmosphere model: to what extent are the main quantities provided by the model (populations of heavy particles, electron density, pressure, etc.) sensitive to the accuracy on the RTE solution? We intend to tackle this subject in a future work by constructing an accurate -- but still very idealized -- stellar atmosphere model, in which the main quantities are first derived from an exact solution to the RTE, and then from the solution given by a ALI-based numerical method. \begin{acknowledgements} The authors wish to thank M. Ahues, A. Largillier, G. Panasenko (Numerical Analysis team of the University Jean Monnet of Saint-Etienne, France), A. Amosov (Moscow Power Engineering Institute, Russia) and J. Bergeat (Centre de Recherche Astronomique de Lyon) for some helpful discussions concerning this work. We also thank Ivan Hubeny and Javier Trujillo Bueno for their valuable comments on a previous version of our manuscript. \end{acknowledgements}
Title: The Great Observatories Origins Deep Survey VLT/FORS2 Spectroscopy in the GOODS-South Field: Part II
Abstract: We present the second campaign of the ESO/GOODS program of spectroscopy of faint galaxies in the GOODS-South field. Objects were selected as candidates for VLT/FORS2 observations primarily based on the expectation that the detection and measurement of their spectral features would benefit from the high throughput and spectral resolution of FORS2. The reliability of the redshift estimates is assessed using the redshift-magnitude and color-redshift diagrams and comparing the results with public data. 807 spectra of 652 individua targets have been obtained in service mode with the FORS2 spectrograph at the ESO/VLT, providing 501 redshift determinations. The typical redshift uncertainty is estimated to be sigma_z ~ 0.001. Galaxies have been selected adopting three different color criteria and using the photometric redshifts.The resulting redshift distribution typically spans two redshift domains: from z=0.5 to 2 and z=3.5 to 6.2. In particular, 94 B435-,V606-,i775 "dropout" Lyman break galaxies have been observed, yielding redshifts for 65 objects in the interval 3.4<z<6.2. Three sources have been serendipitously discovered in the redshift interval 4.8<z<5.5. Together with the previous release, 930 sources have now been observed and 724 redshift determinations have been carried out. The reduced spectra and the derived redshifts are released to the community through the ESO web page this http URL Large scale structures are clearly detected at z=0.666, 0.734, 1.096, 1.221, 1.300, and 1.614. A sample of 34 sources with tilted [OII]3727 emission has been identified, 32 of them in the redshift range 0.9<z<1.5.
https://export.arxiv.org/pdf/astro-ph/0601367
\newcommand{\magcir}{\ \raise -2.truept\hbox{\rlap{\hbox{$\sim$}}\raise5.truept \hbox{$>$}\ }} \newcommand{\mincir}{\ \raise -2.truept\hbox{\rlap{\hbox{$\sim$}}\raise5.truept \hbox{$<$}\ }} \title{The Great Observatories Origins Deep Survey} \subtitle{VLT/FORS2 Spectroscopy in the GOODS-South Field: Part II} \author{E. Vanzella\inst{1} \and S. Cristiani\inst{1} \and M. Dickinson\inst{2} \and H. Kuntschner\inst{3} \and M. Nonino\inst{1} \and A. Rettura\inst{5,6} \and P. Rosati\inst{5} \and J. Vernet\inst{5} \and \\ C. Cesarsky\inst{5} \and H. C. Ferguson\inst{4} \and R.A.E. Fosbury\inst{3} \and M. Giavalisco\inst{4} \and A. Grazian\inst{9} \and J. Haase\inst{3} \and L. A. Moustakas\inst{7} \and \\ P. Popesso\inst{5} \and A. Renzini\inst{8} \and D. Stern\inst{7} \and the GOODS Team } \institute{ INAF - Osservatorio Astronomico di Trieste, Via G.B. Tiepolo 11, 40131 Trieste, Italy. \and National Optical Astronomy Obs., P.O. Box 26732, Tucson, AZ 85726. \and ST-ECF, Karl-Schwarzschild Str. 2, 85748 Garching, Germany. \and Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218. \and European Southern Observatory, Karl-Schwarzschild-Strasse 2, Garching, D-85748, Germany. \and Universite' Paris-Sud 11, Rue Georges Clemenceau 15, Orsay, F-91405, France \and Jet Propulsion Laboratory, California Institute of Technology, MS 169-506, 4800 Oak Grove Drive, Pasadena, CA 91109 \and INAF - Astronomical Observatory of Padova, Vicolo dell'Osservatorio 5, I - 35122 Padova -ITALY \and INAF - Osservatorio Astronomico di Roma, Via Frascati 33, I-00040 Monteporzio Roma, Italy \thanks{Based on observations made at the European Southern Observatory, Paranal, Chile (ESO programme 170.A-0788 {\it The Great Observatories Origins Deep Survey: ESO Public Observations of the SIRTF Legacy/HST Treasury/Chandra Deep Field South.}) } } \offprints{E. Vanzella, \email{vanzella@oats.inaf.it}} \date{Received ; accepted } \abstract {} {We present the second campaign of the ESO/GOODS program of spectroscopy of faint galaxies in the GOODS-South field.} {Objects were selected as candidates for VLT/FORS2 observations primarily based on the expectation that the detection and measurement of their spectral features would benefit from the high throughput and spectral resolution of FORS2. The reliability of the redshift estimates is assessed using the redshift-magnitude and color-redshift diagrams and comparing the results with public data.} {807 spectra of 652 individual targets have been obtained in service mode with the FORS2 spectrograph at the ESO/VLT, providing 501 redshift determinations. The typical redshift uncertainty is estimated to be $\sigma_z \simeq 0.001$. Galaxies have been selected adopting three different color criteria and using the photometric redshifts. The resulting redshift distribution typically spans two redshift domains: from z=0.5 to 2 and z=3.5 to 6.2. In particular, 94 $B_{435}$-,$V_{606}$-,$i_{775}$-"dropout" Lyman break galaxies have been observed, yielding redshifts for 65 objects in the interval 3.4$<$z$<$6.2. Three sources have been serendipitously discovered in the redshift interval 4.8$<$z$<$5.5. Together with the previous release, 930 sources have now been observed and 724 redshift determinations have been carried out. The reduced spectra and the derived redshifts are released to the community through the ESO web page $\it{http://www.eso.org/science/goods/}$ \thanks{The catalog (Table~\ref{tab:tblspec}) is available in electronic form at the CDS via anonymous ftp to cdsarc.u-strasbg.fr (130.79.128.5) or via http://cdsweb.u-strasbg.fr/cgi-bin/qcat?J/A+A/}. Large scale structures are clearly detected at $z \simeq 0.666, 0.734, 1.096, 1.221, 1.300$, and $1.614$. A sample of 34 sources with tilted [O\,{\sc ii}]3727 emission has been identified, 32 of them in the redshift range 0.9$<$z$<$1.5. } {} \keywords{Cosmology: observations -- Cosmology: deep redshift surveys -- Cosmology: large scale structure of the universe -- Galaxies: evolution. } \section{Introduction} The Great Observatories Origins Deep Survey (GOODS) is a public, multi-facility project that aims to answer some of the most profound questions in cosmology: how did galaxies form and assemble their stellar mass? When was the morphological differentiation of galaxies established and how did the Hubble Sequence form? How did Active Galactic Nuclei (AGN) form and evolve, and what role do they play in galaxy evolution? How much do galaxies and AGN contribute to the extragalactic background light? Is the expansion of the universe dominated by a cosmological constant? A project of this scope requires large and coordinated efforts from many facilities, pushed to their limits, to collect a database of sufficient quality and size for the task at hand. It also requires that the data be readily available to the worldwide community for independent analysis, verification, and follow-up. The program targets two carefully selected fields, the Hubble Deep Field North (HDF-N) and the Chandra Deep Field South (CDF-S), with three NASA Great Observatories (HST, Spitzer and Chandra), ESA's XMM-Newton, and a wide variety of ground-based facilities. The area common to all the observing programs is 320 arcmin$^2$, equally divided between the North and South fields. For an overview of GOODS, see \cite{dick03}, \cite{renz03} and \cite{giava04a}. In the last five years the CDF-S has been the target of several spectroscopic campaigns (\cite{crist00}, \cite{croom01}, \cite{bunk03}, \cite{stan04}, \cite{stro04}, \cite{vanderwel04}, \cite{dick04}, \cite{szo04}, \cite{fevre05}, \cite{vanz05}). This is the second paper in a series presenting the results of the GOODS spectroscopic program carried out with the VLT/FORS2 spectrograph. For a full description of its aims we refer to the first paper (\cite{vanz05}, RUN1 hereafter). Here we recall that the ESO/GOODS spectroscopic program is designed to observe all galaxies for which VLT optical spectroscopy is likely to allow the redshift determination. The program makes full use of the VLT instrument capabilities (FORS2 and VIMOS), matching targets to instrument and disperser combinations in order to maximize the effectiveness of the observations. The magnitude limits and selection bandpasses to some extent depend on the instrumental setup being used. The aim is to reach mag~$\sim24-25$ with adequate S/N, with this limiting magnitude being in the B band for objects observed with the VIMOS LR-Blue grism, in the V band for those observed in the VIMOS LR-Red grism, and in the z band for the objects observed with FORS2. The second FORS2 spectroscopic campaign (17 masks, RUN2 hereafter) in the Chandra Deep Field South, was carried out in the period fall 2003 - early 2004 in service mode. New FORS2 observations were performed in December 2004 (6 masks, RUN3) mainly focused on color-selected Lyman break ``dropout'' targets and 5 more masks will be completed before February 2006 (RUN4) mainly dedicated to sources detected at 24$\mu$m with the {\it Spitzer Space Telescope} MIPS instrument. These data will be described in a forthcoming paper. The VIMOS spectroscopic survey in the GOODS-S field is started and will produce hundreds of redshift determinations, mainly in the redshift range 0$<$z$\leq$3.5. The paper is organized as follows. Sect. 2 describes the target selection, while Sect. 3 describes the observations and data reductions. The redshift determination is presented in Sect. 4. In Sect. 5 we discuss the data and in Sect. 6 the conclusions are presented. Throughout this paper the magnitudes are given in the AB system (\cite{oke77}) (AB~$\equiv 31.4 - 2.5\log\langle f_\nu / \mathrm{nJy} \rangle$), and the ACS F435W, F606W, F775W, and F850LP filters are designated hereafter as $B_{435}$, $V_{606}$, $i_{775}$ and $z_{850}$, respectively. We assume a cosmology with $\Omega_{\rm tot}, \Omega_M, \Omega_\Lambda = 1.0, 0.3, 0.7$ and $H_0 = 70$~km~s$^{-1}$~Mpc$^{-1}$. \section{Target Selection} Galaxies were selected as candidates for FORS2 observations primarily based on the expectation that the detection and measurement of their spectral features would benefit from the high throughput, moderately-high spectral resolution, and reduced long-wavelength fringing of FORS2 relative to other instrument options such as VIMOS. In particular, the main spectral emission and absorption features for galaxies at $0.8 < z < 2.0$ appear at very red optical wavelengths ($7000\AA < \lambda < 1\mu$m). Similarly, very faint Lyman break galaxies at $z \gtrsim 4$, selected as $B_{435}$, $V_{606}$ and $i_{775}$--dropouts from the GOODS ACS photometry, also benefit greatly from the red throughput and higher spectral resolution of FORS2. In practice, several categories of object selection criteria were used to ensure a sufficiently high density of targets to efficiently populate masks. These criteria were: \begin{enumerate} \item{Primary catalog: $(i_{775}-z_{850}) > 0.6$ and $z_{850} < 25$. This should ensure redshifts $z \gtrsim 0.7$ for ordinary early-type galaxies (whose strongest features are expected to be absorption lines), and higher redshifts for intrinsically bluer galaxies likely to have emission lines.} \item{Secondary catalog: $0.45 < (i_{775}-z_{850}) < 0.6$ and $z_{850} < 25$.} \item{Photometric-redshift sample: $1<$$z_{\rm phot}$$<2$ and $z_{850} < 25$, from \cite{moba04}.} \item{$B_{435}$, $V_{606}$ and $i_{775}$--dropouts color selected Lyman break galaxy candidates (see \cite{giava04b} and \cite{dick04}).} \item{A few miscellaneous objects, including host galaxies of supernovae detected in the GOODS ACS observing campaign.} \end{enumerate} The targets were selected from a preliminary catalog based on the v1.0 public release of the GOODS ACS images. This version includes all five epochs of the GOODS ACS data \footnote{{\it ftp://archive.stsci.edu/pub/hlsp/goods/catalog$\_$r1/}}, and is a significant improvement on the previous, 3-epoch v0.5 release that was used to select targets for the first FORS2 observations (RUN1, \cite{vanz05}). For this paper and data release, the objects observed with FORS2 have been matched to the public release ACS catalog version r1.1z, also based on the 5-epoch v1.0 ACS images. The r1.1z catalog is based on the r1.0z SExtractor run, and merely corrects errors and omissions in the r1.0z catalog files. When designing the masks, we generally tried to avoid observing targets that had already been observed in other redshift surveys of this field, namely the K20 survey of \cite{cimatti02} and the survey of X-ray sources by \cite{szo04}. 807 spectra of 652 individual targets have been extracted from the RUN2 (multiple observations have been performed, especially for the high redshift candidates). Out of these 652 targets, 178 are from the primary catalog, 117 are from the secondary catalog, 141 are from the photometric redshift selection, 94 are from the Lyman break sample, and 3 are from miscellaneous list. The remaining 119 sources have been serendipitously identified, due to: a) sources randomly in the slit other than the target or b) sources put in the slit in the situation where no targets were available or c) relatively bright objects put in the slit for the alignment of the mask. The total number of individual sources observed in the RUN1 + RUN2 is 930 (1203 spectra reduced) with 724 redshift determinations. The spectroscopic database presented here is incomplete: none of the above listed categories has been exhaustively observed, nor any GOODS subarea has been fully covered. \section{Observations and Data Reduction} \begin{table} \centering \caption{Journal of the FORS2 observations (RUN2).} \begin{tabular}{lccc} \hline \hline Mask ID & Date & exp.time (s)\\ \hline 914250 & Aug. 2003 & 17$\times$1200 \cr 905513 & Sept. 2003 & 18$\times$1200 \cr 943018 & Sept. 2003 & 12$\times$1200 \cr 924345 & Sept. 2003 & 12$\times$1200 \cr 945143 & Sept. - Oct. 2003 & 12$\times$1200 + 3$\times$1000 \cr 992438 & Oct. - Dec. 2003 & 12$\times$1200 \cr 985931 & Nov. 2003 & 12$\times$1200 + 2$\times$120 \cr 990204 & Dec. 2003 & 12$\times$1200 \cr 904509 & Dec. 2003 & 12$\times$1200 \cr 991435 & Dec. 2003 & 12$\times$1200 \cr 935030 & Dec. 2003 & 12$\times$1200 \cr 951937 & Dec. 2003 & 12$\times$1200 + 1100 + 500 \cr 960930 & Dec. 2003 & 12$\times$1200 \cr 961839 & Jan. 2004 & 12$\times$1200 \cr 932802 & Jan. 2004 & 12$\times$1200 \cr 993304 & Jan. 2004 & 12$\times$1200 \cr 951526 & Feb. 2004 & 3$\times$1200 \cr \hline \label{tab:tblobs} \end{tabular} \end{table} The VLT/FORS2 spectroscopic observations were carried out in service mode during several nights at the end of 2003 and the beginning of 2004. A summary is presented in Table~\ref{tab:tblobs}. In all cases the $300I$ grism was used as dispersing element without order-separating filter. This grism provides a scale of roughly 3.2\AA~pix$^{-1}$. The nominal resolution of the configuration was R=$\lambda/\Delta\lambda$=660, which corresponds to 13{\AA} at 8600{\AA}. The spatial scale of FORS2 is $0.126\arcsec$/pixel. The slit width was always $1\arcsec$. Dithering of the targets along the slits was applied typically with steps of 0,$\pm$8 pixels, in order to effectively improve the sky and fringe subtraction, and remove CCD blemishes. \subsection{\it Data Reduction} Data were reduced with a semi-automatic pipeline that we have developed on the basis of the MIDAS package (\cite{eso_midas}), using commands of the LONG and MOS contexts. The main procedures have been described in the previous paper (\cite{vanz05}). In the cases of multiple observations of the same source in different masks, the one dimensional spectra have been co-added weighing according to the exposure time, the seeing condition and the resulting quality of each extraction process (defects present in the CCD, object too close to the border of the slit, etc.). A visual check of the two dimensional frames has been performed (in some cases the two dimensional spectra have also been co-added, in order to improve and guide the visual inspection). We emphasize here that we opted to observe the science targets {\em without} an order-sorting filter, implying deleterious effects to the flux calibration. The second order overlap becomes important at wavelengths above $\sim$8000\AA~depending on the color of the target. In Figure~\ref{fig:test_flux} the comparison between two 1-D calibrated spectra (one blue ($i-z$)$\sim$0 and one red ($i-z$)$\sim$1) and the correspondent ACS photometry is shown. The photometric values in the $i$ and $z$ bands are marked with two filled circles and are consistent with the derived spectral behavior. For the red objects that dominate the FORS2 target selection, we felt that the improved wavelength coverage more than compensates for the partial unreliability of the flux calibration. Due to both this second order light and uncertain slit losses, we caution against using the calibrated fluxes for scientific purposes. Fluxes in the released 1-D spectra are given in units of $10^{-16}$ erg s$^{-1}$ cm$^{-2}$ \AA$^{-1}$. \section{Redshift Determination} Spectra of 652 individual objects have been extracted from RUN2. From them we have determined 501 redshifts. In the large majority of the cases the redshift has been determined through the identification of prominent features of galaxy spectra: depending on the redshift and the nature of the source the 4000\AA\ break, Ca H and K, g-band, MgII 2798-2802, AlII 3584, Ly$\alpha$, Si\,{\sc ii} 1260.4\AA, O\,{\sc i} 1302.2\AA, C\,{\sc ii} 1335.1\AA, Si\,{\sc iv} 1393.8,1402.8\AA, Si\,{\sc ii} 1526.7\AA, C\,{\sc iv} 1548.2, 1550.8\AA~in absorption and Ly$\alpha$, [O\,{\sc ii}]3727, [O\,{\sc iii}]5007, H$\beta$, H$\alpha$ in emission. The redshift estimation has been performed cross-correlating the observed spectrum with templates of different spectral types (S0, Sa, Sb, Sc, Elliptical, Lyman Break, etc.), using the $rvsao$ package in the IRAF environment. The redshift identifications are summarized in Table~\ref{tab:tblspec} and are available at the URL $\it{http://www.eso.org/science/goods/}$. In Table~\ref{tab:tblspec}, the column {\em ID} contains the target identifier, that is constructed out of the target position (e.g., $GDS~J$033206.44-274728.8) where GDS stands for {\bf G}OO{\bf D}S {\bf S}outh. The coordinates are based on the GOODS v1.1 astrometry. The v1.1 release is based upon the v1.0 SExtractor run, and merely corrects errors and omissions in the v1.0 catalog files. The cataloged sources are identical, in both number and ordering, to the v1.0 release. The columns $z_{850}$ and ($i_{775}$-$z_{850}$) list the magnitude (SExtractor ``MAG$\_$AUTO'') and the color (SExtractor ``MAG$\_$ISO'') of the sources derived from the catalog v1.1. The color has been measured through isophotal apertures defined in the $z_{850}$ band image (as done in \cite{dick04} and \cite{giava04b}). The {\em quality} flag (QF hereafter), indicates the reliability of the redshift determination. As described in the previous work (\cite{vanz05}, RUN1), the QF has been divided in three categories: ``A'', ``B'' and ``C''. An estimation of the confidence level associated to each class ``A'', ``B'' and ``C'' can be derived analyzing the FORS2 measurements in common with independent spectroscopic estimations available in literature. This has been done in the previous paper (RUN1) where 39 sources have been analyzed and in Sect. 5.1 of the present work (98 more sources, see below). In this way the sample of FORS2 measurements in common with independent spectroscopic surveys counts 137 galaxies, in which we find 0, 1 and 4 FORS2 wrong redshifts for classes ``A'', ``B'' and ``C'', respectively. In this way at 1$\sigma$ (\cite{gehrels86}) the confidence level of the ``A'', ``B'' and ``C'' categories turns out to be $\simeq$ 98$\%$, $\simeq$ 97$\%$ and 93$\%$. There are 291 objects classified with quality ``A'', 119 with quality ``B'', 91 with ``C'', and 151 with ``X'', an inconclusive spectrum. The flag "{\em class}" groups the objects for which emission line(s) (``em.''), absorption-line(s) (``abs.'') or both (``comp.'') are detected in the spectrum. In the present catalog, three sources have been classified as stars. In 30$\%$ of the cases the redshift is based only on one emission line, usually identified as [O\,{\sc ii}]3727 or Ly$\alpha$. In these cases the continuum shape, the presence of breaks, the absence of other spectral features in the observed spectral range and the broad band photometry are particularly important in the evaluation. The quality for these sources ranges from ``A'' to ``C'' depending on the additional information described above (35$\%$ of the sample with a single emission line have QF=''A'', with a mean redshift $<z>$=1.21$\pm$0.2). The {\em comments} column contains additional information relevant to the particular observation. The most common ones summarize the identification of the principal lines, the inclination of an emission line due to internal kinematics, the weakness of the signal (``faint''), the low S/N of the extracted spectrum (``noisy''), the apparent absence of spectroscopic lines (``featureless continuum''), etc. In few cases the spectrum extracted is the combination of more than one source in the slit and where possible the redshifts of the ``components'' have been estimated. In the RUN1 + RUN2 spectroscopic data, 11 sources in the GOODS-S field are not present in the ACS photometric catalog v1.1. Six of them have a redshift estimation (an example is shown in Figure~\ref{fig:MIXED}). Three out of six appear to be emission line objects whose continuum is too faint and has not been detected in the ACS catalogs. The other seven sources are outside the ACS area. \begin{table*} \centering \caption{Spectroscopic redshift catalog. $\dag$} \begin{tabular}{lcccccl} \hline \hline ID(v1.0) & $z_{850}$ & $(i_{775}-z_{850})$ & zspec &class. & Quality & comments \\ \hline GDS~J033245.99-275108.3 & 23.48 &0.47 &1.238 &em. & B &[O\,{\sc ii}]3727 \cr GDS~J033246.04-274929.7 & 26.06 &1.77 &5.787 &em. & A &LyA (faint continuum) \cr GDS~J033246.05-275444.8 & 21.49 &0.53 &0.733 &abs. & A &CaH,g-band,H$\beta$,Mg,CaFe \cr GDS~J033246.16-274752.3 & 24.46 &0.43 &1.221 &em. & B &[O\,{\sc ii}]3727 \cr \hline \multicolumn{6}{l} {$\dag$ This table is available in its entirety via $\it{http://www.eso.org/science/goods/}$.}\\ \multicolumn{6}{l} {A portion is shown here for guidance regarding its form and content.}\\ \label{tab:tblspec} \end{tabular} \end{table*} \section{Discussion} In the following, if not specified, we consider the entire FORS2 sample, including both RUN1 and RUN2. This sample is summarized in Table~\ref{tab:matrix} where the sources are divided into different selection categories (see Sec. 2) and by redshift (or whether a redshift could be determined). The distribution of the quality flags is also tabulated. \subsection{Reliability of the redshift - comparison with public data} A practical way to assess the reliability of the redshifts reported in Table~\ref{tab:tblspec} is to compare the present results with independent measurements from other surveys. In the last five years the CDF-S has been the target of several spectroscopic campaigns (the surveys with the number of redshifts in parenthesis used in the comparison are here reported: \cite{crist00} (5), \cite{croom01} (29), \cite{bunk03} (1), \cite{stan04} (3), \cite{stro04} (14), \cite{vanderwel04} (6), \cite{dick04} (1), \cite{szo04} (124), \cite{fevre05} (748), \cite{vanz05} (234)). Making use of a publicly available master compilation of all spectroscopic redshifts in the GOODS/CDF-S region (Rettura et al. in preparation, available at the URL {\it http://www.eso.org/science/goods/spectroscopy/CDFS$\_$Mastercat/} we have been able to compare our redshift determinations with the existing data in the literature. There are 98 objects in common with the present second release of the FORS2 GOODS survey (RUN2). For 87 cases out of 98 (89$\%$) the agreement is very good, with a mean difference $<z_{FORS2_{RUN2}} - z_{CDF-S}>$= 0.0042 $\pm$ 0.0095. 15 objects have a redshift determination both in RUN1 and RUN2. The distribution of the redshift differences has a median $|z_{FORS2_{RUN2}} - z_{FORS2-{RUN1}} | = 0.0002$ and a difference between the 82 and 18 percentile of $2.6 \cdot 10^{-3}$. Assuming equipartition of the redshift uncertainties between RUN1 and RUN2 we derive a typical error on the redshift determinations in the FORS2 GOODS spectroscopy of $\sigma_z \simeq 0.001$. Ten cases show ``catastrophic'' discrepancies between the RUN2 and the K20, \cite{szo04} and the VVDS surveys, i.e. $|z_{FORS2_{RUN2}} - z_{CDF-S}|$ greater than 0.08. In order to compare the redshift estimations we recall here which is the quality level adopted by other authors. In the K20 survey the QF adopted is 1, 0 or -1 if the redshift determination is solid, tentative or unconclusive, respectively. In \cite{szo04} the QF=3 indicates reliable redshift determination with unambiguous X-ray counterpart, QF=2 corresponds to a reliable redshift determination and a value of 0.0 indicates no success. QF=1 indicates the detection of $some$ feature in the spectrum (typically a single narrow emission line). QF=0.5 is used when there is a hint of some spectral feature. In the VVDS, the flags 2,3,4 are the most secure with a confidence of 75$\%$, 95$\%$ and 100$\%$ respectively. Flag 1 is an indicative measurement (50$\%$), flag 9 indicates that there is only one secure emission line, and flag 0 indicates a measurement failure with no features identified. In the following we discuss in detail each discrepant spectrum: \begin{enumerate} \item{GDS~J033232.08-274155.2. This is a discrepancy with our previous identification (RUN1) and the present one (RUN2). In the first run the redshift determination was tentative (quality ``C'', z= 0.960) and in the second run we derived z=1.393 (QF=''B''). However the co-addition of the two produces a featureless continuum, we have changed the quality to ``X''.} \item{GDS~J033217.77-274714.9. K20 and VVDS assign redshift 0.729 and 0.731, respectively (and quality 1 and 3). In the FORS2 spectrum there are three objects in the slit, the GDS~J033217.77-274714.9 is a serendipitous source at the border of the slit, its exposure time is reduced of 50$\%$ due to the dithering process. The continuum is faint and a possible emission line is detected at 7522.6\AA~interpreted to be [O\,{\sc ii}]3727 at z=1.018 (QF=''C'').} \item{GDS~J033232.18-274534.9. K20 assigns a redshift 0.332 with quality 1. The FORS2 spectrum shows [O\,{\sc ii}]3727, MgI, CaHK, g-band and the Balmer Break at z=0.523 (QF=''A'').} \item{GDS~J033239.67-274850.6. Szokoly et al. measure redshift 3.064 with quality 3. Our spectrum shows a smoothed break at $\sim$ 6000\AA~and an absorption line at 6789.0\AA, our redshift determination is tentatively z=3.885, QF=''C''. The spectrum starts at 5600\AA, if it is at redshift 3.064, the most relevant spectral features are outside the spectral coverage. We note that if the redshift is 3.064 the absorption line we measure at 6789.0\AA would be consistent with the Al\,{\sc ii} 1670.8\AA.} \item{GDS~J033240.84-275546.7. Szokoly et al. measure redshift 0.625 with quality 0.5. Our spectrum shows a featureless continuum and starts at 5790\AA, a possible emission line is detected at 8277.6\AA, we assign tentatively z=1.221 QF=''C''.} \item{GDS~J033222.44-275606.1. VVDS measure redshift 0.490 with quality 2. The FORS2 spectrum shows a tilted emission line at 7790.3\AA~and a faint-diffuse continuum. We assign tentatively z=1.090 (QF=''C''). We note that in the FORS2 spectrum the [O\,{\sc ii}]3727, [O\,{\sc iii}]5007 or H$\beta$ lines at z=0.490 have not been detected.} \item{GDS~J033225.28-275524.2. VVDS measure redshift 0.923 with quality 1. The FORS2 spectrum shows [O\,{\sc ii}]3727 (slightly tilted), CaHK, MgI and the Balmer Break at z=1.017 (QF=''A'').} \item{GDS~J033230.37-274008.5. VVDS measure redshift 1.083 with quality 2. The source shows a bright continuum and [O\,{\sc ii}]3727, MgII and the NeIII lines at z=1.357 (QF=''A'').} \item{GDS~J033230.50-275312.3. VVDS measure redshift 1.427 with quality 2. The FORS2 spectrum shows [O\,{\sc ii}]3727 (tilted), CaHK, MgII, g-band at z=1.017 QF=''A''.} \item{GDS~J033234.82-274721.9. VVDS measure redshift 0.315 with quality 3. In the FORS2 spectrum an emission line has been detected at 8632.3\AA, interpreted as [O\,{\sc ii}]3727 at z=1.316 with QF=''B''. The continuum starts at 6260\AA, and if we assume the line to be H$\alpha$ at z=0.315 the H$\beta$ and/or [O\,{\sc iii}]5007 are not present.} \item{GDS~J033242.97-274649.9. VVDS measure redshift 0.831 with quality 1. The FORS2 spectrum shows [O\,{\sc ii}]3727, CaHK, NeIII and $H\delta$ (in absorption) at z=1.036 with QF=''A''.} \end{enumerate} In summary, 7 out of 10 discrepant redshift determinations turn out to be probably correct in the FORS2 spectroscopy, all with QF better or equal to QF=``B''. Of the remaining 3 sources (all with QF=''C''), one is uncertain and two are probably wrong in the FORS2 spectroscopic identification due to the reasons described above. \subsection{Reliability of the redshifts - diagnostic diagrams} Figures~\ref{fig:z_vs_mag} and \ref{fig:i_zVSzspec} show the redshift-magnitude and the color-redshift distributions, respectively. Figure~\ref{fig:i_zVSzspec} shows the behavior for galaxies at redshift less than 2 and quality flag ``A'' and ``B''. The two populations of ``emission-line'' (star-forming) and ``absorption-line'' (typically elliptical) galaxies are clearly separated. The mean color of the absorption-line objects outline the upper envelop of the distribution, consistent but increasingly bluer than the colors of a non-evolving $L^{\star}$ elliptical galaxy (estimated integrating the spectral templates of \cite{cole80} through the ACS bandpasses). The emission-line objects show in general a bluer $i_{775}-z_{850}$ color and a broader distribution than the absorption-line sources. The broader distribution, with some of the emission-line objects entering the color region of the ellipticals, is possibly explained by dust obscuration, high metallicity or strong line emission in the $z_{850}$ band (for example emission lines [O\,{\sc iii}]5007, H$\beta$ at redshift 0.8, as measured for the source GDS~J033219.53-274111.6). \begin{table*} \centering \caption{Summary of the spectroscopic catalog as a function of the redshift bin (first column), categories (from column two to six) and serendipitously identified sources (column seven). The contribution of the different quality flags (``A'', ``B'' or ``C'') are also reported. A total of 930 spectra have been analyzed (RUN1 and RUN2).} \begin{tabular}{lllllll|c} \hline \hline z-bin & cat. 1)$_{(A,B,C)}$ & cat. 2)$_{(A,B,C)}$ & cat. 3)$_{(A,B,C)}$ & cat. 4)$_{(A,B,C)}$ & cat. 5)$_{(A,B,C)}$ & seren. & Sum \\ \hline no redshift & 57 &20 &43 & 37 &0 &49 &206 \cr stars & 4$_{(0,3,1)}$ &0$_{(0,0,0)}$ &0$_{(0,0,0)}$ & 7$_{(1,2,4)}$ &0$_{(0,0,0)}$ &3$_{(2,0,1)}$&14 \cr (0..1) & 19$_{(12,1,6)}$ &42$_{(35,6,1)}$ &23$_{(18,3,2)}$ & 2$_{(2,0,1)}$ &1$_{(1,0,0)}$ &121$_{(83,19,18)}$&208 \cr [1..2) & 193$_{(113,51,29)}$&83$_{(49,24,10)}$&115$_{(76,25,13)}$& 4$_{(2,2,0)}$ &2$_{(2,0,0)}$ &34$_{(10,15,9)}$ &431 \cr [2..3) & 0$_{(0,0,0)}$ & 1$_{(1,0,0)}$ & 0$_{(0,0,0)}$ & 0$_{(0,0,0)}$ &0$_{(0,0,0)}$ & 0$_{(0,0,0)}$ &1 \cr [3..4) & 0$_{(0,0,0)}$ & 0$_{(0,0,0)}$ & 0$_{(0,0,0)}$ &26$_{(14,7,4)}$&0$_{(0,0,0)}$ & 0$_{(0,0,0)}$ &26 \cr [4..5) & 0$_{(0,0,0)}$ & 0$_{(0,0,0)}$ & 0$_{(0,0,0)}$ &23$_{(6,8,9)}$ &0$_{(0,0,0)}$ & 2$_{(0,1,1)}$ &25 \cr [5..6) & 0$_{(0,0,0)}$ & 0$_{(0,0,0)}$ & 0$_{(0,0,0)}$ &14$_{(7,3,5)}$ &0$_{(0,0,0)}$ & 3$_{(0,1,2)}$ &17 \cr [6..7) & 0$_{(0,0,0)}$ & 0$_{(0,0,0)}$ & 0$_{(0,0,0)}$ & 2$_{(0,1,1)}$ &0$_{(0,0,0)}$ & 0$_{(0,0,0)}$ &2 \cr \hline Sum & 273 &146 & 181 &115 &3 &212 &$\bf{930}$\cr \hline \hline \label{tab:matrix} \end{tabular} \end{table*} \subsection{Redshift distribution and Large Scale Structure} The top and bottom panels of Figure~\ref{fig:zdistr} show the redshift distribution of the galaxies at redshift less than 2 and greater than 2, respectively (solid line QF ``A'' and ''B'', dotted line QF ``C''). In the following sections we discuss the redshift distribution separating the low ($z<2$) and high ($z>2$) redshift intervals. \subsubsection{The sample at $z < 2$} The redshift distribution is consistent with the criteria for the target selection (color and photometric redshift selected), with the majority of the sources having redshifts in the interval 1$<$z$<$2 (see also Table~\ref{tab:matrix}, last column). In the RUN1+RUN2, out of 181 galaxies selected via photometric redshift, 138 have a spectroscopic redshift identification and 136 with zspec$>$0.8 (115 at zspec$>$1). Table~\ref{tab:z_properties} shows the fraction of determined redshifts as a function of the spectral features identified, i.e. emission lines, absorption lines, emission \& absorption lines. The median of the redshift distribution of each class is close to 1, with a more populated tail in the redshift interval 1$<z<$2 (see top panel of Figure~\ref{fig:zdistr}). Obviously, in the presence of emission lines, it is easier to determine a redshift. As reported in Table~\ref{tab:z_properties}, 537 galaxies (including ``em.'' and ``comp'') show the [O\,{\sc ii}]3727 emission line and assuming as an extreme case that all the inconclusive redshift determinations (category ``X'') belong to the class ``abs.'', the number of ``em.'' sources is still dominant, comprising 63$\%$ of the entire target list. This is a likely reason why the majority of galaxies identified in the present work belong to the ``em.'' class. Alternatively, [O\,{\sc ii}]3727 is a classic star forming indicator and the redshift interval $1<z<2$ corresponds to the peak of the mean star formation intensity of the universe. There are 102 galaxies identified with absorption lines only (``abs.'' class, mainly Ca H and K, MgII 2798-2802) in the range of redshift between 0.3-2.0. 28 sources out of 102 with only absorption features detected have been serendipitously-observed, the redshift distribution of this sample peaks at z=0.68$\pm$0.2. Six galaxies have been identified at redshift $\sim$2. These sources show the Mg\,{\sc ii} 2798,2802\AA~in absorption (in three cases the [Fe\,{\sc ii}] 2344,2383\AA~absorption lines are also present), five of them (GDS~J033241.84-274657.1 QF=''B'', GDS~J033240.06-274755.4 QF=''A'', GDS~J033228.17-274648.4 QF=''C'', GDS~J033240.27-274949.7 QF=''C'' and GDS~J033233.84-274520.5 QF=''C'') have been discovered in the RUN2 and have blue colors ($i_{775}-z_{850}<0.6$). Two examples of 2D spectra are shown in Figure~\ref{fig:GAL_Z2} and the composite one-dimensional spectrum is shown in the right panel of Figure~\ref{fig:stack1p61}. For these sources the [O\,{\sc ii}]3727 emission, if present, is out of the spectral range, at 11180\AA. The source GDS~J033233.85-274600.2 is an elliptical galaxy at $z$=1.91 already discussed by \cite{cimatti04}, and has been observed in the RUN1. \begin{table*} \centering \caption{Fractions of sources in the redshift interval 0$<$z$<$2 with different spectral features (RUN1 + RUN2 without stars). The fractions of the different categories observed in Sect.~2 are also shown.} \begin{tabular}{lccc|cccccc|c} \hline \hline Spectral class &$(z_{median})_{-1\sigma}^{+1\sigma}$&$z_{min}$&$z_{max}$& cat.1) & cat.2) & cat.3) & cat.4) &cat.5) & cat.-1) (seren.)&Sum\\ \hline \cr emission & (1.13)$_{-0.74}^{+1.33}$ & 0.067 & 1.621 & 117 & 92 & 123 & 3 & 2 & 104 & 441\cr \cr absorption & (1.00)$_{-0.67}^{+1.22}$ & 0.337 & 1.998 & 51 & 15 & 5 & 2 & 1 & 28 & 102\cr \cr em. \& abs. & (1.02)$_{-0.67}^{+1.29}$ & 0.382 & 1.380 & 44 & 18 & 10 & 2 & 0 & 22 & 96\cr \\ \hline Sum & & & & 212 & 125 & 138 & 7 & 3 & 154 & $\bf{639}$ \cr \hline \label{tab:z_properties} \end{tabular} \end{table*} 441 sources belong to the ``em.'' class (they are dominated by emission lines, mainly [O\,{\sc ii}]3727), many of them entering the so-called ``spectroscopic desert'' up to z=1.621. It is interesting to note that 133 galaxies out of 138 with redshift and photo-z selected show the [O\,{\sc ii}]3727 emission line. 96 sources have been classified as intermediate between ``em.'' and ``abs.'' classes, where both emission and absorption lines with an evident 4000\AA~break are present. \subsubsection{Large Scale Structure} The presence in the CDF-S of large scale structure (LSS) at $z<2$ is indicated by the peaks in the redshift distribution (see Figure~\ref{fig:LSS}). To assess the significance of these structures we follow a procedure similar to that adopted by \cite{gilli03}, who observed features in their X-ray source redshift distribution. We have distributed the sources (the ``signal distribution'') in the velocity domain ($V~=~c~ln(1+z)$, so that $dV=\frac{c~dz}{1+z}$) and smoothed with a Gaussian filter with $\sigma_{S}=300~km/s$ (the typical error in the redshift determination). We have then smoothed the observed distribution with a Gaussian filter with $\sigma_{S}=15000~km/s$ and considered this as the background distribution. We have searched for possible redshift peaks in the signal distribution, computing a signal-to-noise ratio defined as $SNR$ = ($\frac{S-B}{B^{0.5}}$), where $S$ is the number of sources in a velocity interval of fixed width $\Delta V=2000~km/s$ and $B$ is the number of background sources in the same interval. Adopting the threshold $SNR>$5 we have found 6 peaks in the signal distribution (indicated with an arrows in the Figure~\ref{fig:LSS}). In order to estimate the expected fraction of possibly ``spurious'' peaks arising from the background fluctuations, we have simulated 10$^{5}$ samples of the same size of the observed distribution and randomly extracted from the smoothed background distribution and applied our peak detection method to each simulated sample. The result is that, with the adopted threshold, the average number of spurious peaks due to background fluctuations is 0.06. Of the simulated samples, 5.7$\%$ show one spurious peak, 0.1$\%$ show two spurious peaks, and only one simulation (out of $10^{5}$) has three spurious peaks. None of the simulated samples have four or more spurious peaks. \begin{table} \centering \caption{Peaks detected in the FORS2 source redshift distribution, sorted by increasing redshift. The signal and background distributions are smoothed with $\sigma_{S}=300km/s$ and $\sigma_{B}=15000km/s$, respectively. Together with the central redshift of each peak, the number of sources N in each peak and the probability (determined on 10$^{5}$ simulations) to detect spurious peaks arising from the background distribution with a $SNR$ equal or greater than the $SNR$ value measured in the signal distribution.} \begin{tabular}{ccccc} \hline \hline z & N & SNR & Prob. \\ \hline 0.666 & 22 & 8.6 & 4.5$\times$10$^{-4}$ \cr 0.734 & 40 & 16.6 & $<$1$\times$10$^{-5}$ \cr 1.096 & 42 & 8.0 & 1.9$\times$10$^{-3}$ \cr 1.221 & 47 & 9.7 & 2.2$\times$10$^{-4}$ \cr 1.300 & 35 & 7.4 & 4.2$\times$10$^{-3}$ \cr 1.614 & 20 & 11.1 & 7.0$\times$10$^{-5}$ \cr \hline \label{tab:LSS} \end{tabular} \end{table} The six source peaks detected by our procedure are listed in Table~\ref{tab:LSS}, where for each peak we give the average redshift, the number of objects (N) in the peak and the probability (derived from the 10$^{5}$ simulations) of observing a spurious peak with the SNR equal or greater than the measured SNR of the peak detected in the signal distribution. The peaks at $z \sim 0.734$ and $z \sim 0.666$ are already known (\cite{cimatti02}, \cite{gilli03}, \cite{fevre04}). The other four indications of large scale structures in the CDF-S have been identified at redshift 1.096, 1.221, 1.300 (also described by \cite{adami05}) and 1.614. We note that other two peaks have been detected with a $SNR\sim$4.5 at redshift 1.040 and 1.382. In the current spectroscopic catalog (RUN1 + RUN2) 20 galaxies at $z \sim 1.61$ have been discovered (Figure~\ref{fig:groupz1.61} shows an example of the z$\sim$1.61 galaxies discovered in the RUN2). The number of sources increase if we consider other surveys: \begin{enumerate} \item{the observations of \cite{gilli03} who found a peak in the redshift distribution of X-ray sources at z=1.618 (5 galaxies);} \item{the three galaxies at z$\sim$1.61 (\cite{cimatti02}, \cite{cimatti04}) which are passively evolving early type galaxies} \item{at least 5 more galaxies in the third FORS2 run (from our preliminary reduction);} \end{enumerate} Fig.~\ref{fig:space_distr} shows the spatial distribution of the galaxies at $z ~\approx~ 1.61$ using both the present work and data from the literature. The current sample contains 28 galaxies apparently distributed in a non-uniform way, the majority of them have been detected in the upper part of the field and 3 pairs have an angular separation below 4 arcseconds ($\sim$ 30 kpc at $z \sim 1.61$). At redshift 1.61, the ACS $B_{435}$ band is sampling the 1667\AA~rest-frame UV radiation. As reviewed by \cite{kenni98}, one can estimate the SFR from the rest-frame UV luminosity density $L_{\nu}$ in the range 1500-2500~\AA~using the following relation: SFR(M$_{\odot}$yr$^{-1}$) = 1.4 $\times$ 10$^{-28}$ $L_{\nu}$ (ergs s$^{-1}$ Hz$^{-1}$) for a Salpeter IMF, covering the range 0.1 to 100M$_{\odot}$. This relation applies only to galaxies with continuous star formation over time scales of 10$^{8}$ years or longer. We have estimated the rest-frame luminosity density $L_{\nu}$ (in ergs s$^{-1}$ Hz$^{-1}$) of the 20 galaxies at z=1.61 identified in the current FORS2 spectroscopic campaign, using the apparent $B_{435}$ AB magnitude (the SExtractor ``mag$\_$auto'', \cite{ber96}) and the luminosity distance. The final luminosity is $L_{\nu,o}$ = $L_{\nu}$ $\times$ 10$^{0.4~Av}$, where $Av$ represents the amount of dust extinction. Adopting no extinction ($Av$=0), we obtain a lower limit for the mean star formation rate of $<$SFR$>$4$\pm$2 M$_{\odot}$yr$^{-1}$. Assuming $Av$=1 or $Av$=2 the $<$SFR$>$ increase from 10$\pm$5 M$_{\odot}$yr$^{-1}$ to 24$\pm$14 M$_{\odot}$yr$^{-1}$, respectively. The rest frame composite spectrum of twenty galaxies at z=1.61 is shown in the left panel of Figure~\ref{fig:stack1p61}. The [O\,{\sc ii}]3727 line and the Mg\,{\sc ii} 2798,2802\AA~and [Fe\,{\sc ii}] 2344,2383\AA~are clearly evident. \subsection{The Lyman break galaxies} 116 sources in the FORS2 RUN1 and RUN2 belong to the class 4), i.e., objects selected to be at high redshift by Lyman break color criteria. It is important to divide the first and second run in order to characterize the success rate. As already discussed in the previous paper (\cite{vanz05}), in the first FORS2 run 14 candidate dropouts were observed, and only one was confirmed at z=5.83. Another five were found to be stars and the remaining sources had inconclusive spectra. The photometric selection of the dropouts galaxies in the first FORS2 run was based on an incomplete photometric dataset (first three epochs photometry). In the following, we consider only the results from RUN2, for which dropout candidates were selected from the full (five epochs) ACS photometry. 94 Lyman break galaxy candidates selected by the $B_{435}$,$V_{606}$ and $i_{775}$-dropout criteria were observed in RUN2. The redshift distribution measured for 65 of these galaxies is shown in the lower panel of Figure~\ref{fig:zdistr}. The 75$\%$, 70$\%$ and 70$\%$ of the observed $B_{435}$, $V_{606}$ and $i_{775}$-dropout color selected candidates have a redshift estimation. The sources with inconclusive redshift determination are in general too faint or without evident spectral features. 100$\%$ of the $B_{435}$-dropouts with a measured redshift have been confirmed to be at redshift between 3.4 and 4.6, 90$\%$ of the $V_{606}$-dropouts with a measured redshift are in the range 4.4 and 5.6, and 93$\%$ of the $i_{775}$-dropouts with a measured redshift are at redshift greater than 5.2 (one source is a probable star). \begin{table*} \centering \caption{Fraction of confirmed dropout candidates in the second FORS2 run (RUN2), ``Nobs.'' indicates the number of sources observed. Four serendipitously-observed high redshift sources are also reported.} \begin{tabular}{lccc|ccc} \hline \hline classes (Nobs.) & confirmed high-z (*)& confirmed low-z & no redshift & (*) ``em.''$_{(A,B,C)}$ &(*) ``abs.''$_{(A,B,C)}$ &(*) ``em.''+''abs.''$_{(A,B,C)}$ \\ \hline $B_{435}$-drop (44) & 33 (3.418$<$z$<$4.597)& 0 & 11 &12$_{(8,3,1)}$ &20$_{(9,5,6)}$ &1$_{(1,0,0)}$\cr $V_{606}$-drop (30) & 19 (4.400$<$z$<$5.554)& 2 (z$<$1.4) & 9 &13$_{(5,5,3)}$ & 6$_{(0,2,4)}$ &0$_{(0,0,0)}$\cr $i_{775}$-drop (20) & 13 (5.250$<$z$<$6.200)& 1 (star) & 6 & 8$_{(3,4,1)}$ & 5$_{(0,0,5)}$ &0$_{(0,0,0)}$\cr Serend. & 3 (4.838$<$z$<$5.541) & - &-& 3$_{(0,0,3)}$ & 0$_{(0,0,0)}$ &0$_{(0,0,0)}$\cr \hline \label{tab:high-z} \end{tabular} \end{table*} Table~\ref{tab:high-z} (and Table~\ref{tab:matrix}) summarize the success rate as a function of redshift, quality flag, class and selection criteria. Columns 5, 6 and 7 of Table~\ref{tab:high-z} show the fraction of the confirmed high redshift galaxies and the ``class'' flag that is related to the features detected in the redshift determination. Beyond redshift 5, if no spectral lines are present, the main features indicating the high redshift nature of the source are: the break in the continuum due to galactic and intergalactic absorption blueward 1215.8\AA, and the flatness of the continuum redward the 1215.8\AA. Figure~\ref{fig:panoramic_high_z} shows the two-Dimensional collection of the 18 galaxies at redshift greater than 5 discovered in the RUN2 and Figure~\ref{fig:z6p097} shows the one-dimensional spectrum of the galaxy GDS~J033223.84-275511.6 at z=6.097. In the top of the figure the spectrum of the sky (not flux calibrated) is shown together with the response curves of the ACS filters $i_{755}$ and the $z_{850}$. In some cases the Ly$\alpha$ is in emission (marked with a circle) and the break of the continuum is evident. Five sources show only the continuum break (a solid segment marks the possible position of the break). The mean value of the observed $i_{755} - z_{850}$ for this sample increases with increasing redshift. The presence of the Lyman emission line, however, can affect significantly the resulting color of the galaxy, introducing a scatter in the blue or in the red directions. For example in the case of the source GDS~J033218.92-275302.7, the strong Ly$\alpha$ line at z=5.554 produces an $i_{755} - z_{850}$ = 0.625. Similarly, in the case of the source GDS~J033223.84-275511.6, the intense Ly$\alpha$ line falls in the $z_{850}$ band, producing an $i_{755} - z_{850}$ $>$ 4. The source GDS~J033217.96-274817.0 is an $i_{775}$-dropout candidate. The FORS2 spectrum is the superposition of two sources. In Figure~\ref{fig:idrop_blue} the one and two dimensional spectra and the color ACS image of the sources are shown. One source (GDS J033217.95-274817.5) is clearly blue ($i_{775}-z_{850}$=-0.25) with respect the $i_{775}$-dropout candidate ($i-z$=1.18). The one dimensional spectrum shows a break at $\sim$ 7800\AA~and the flatness shape redward the break. Collapsing $\sim$ 100 columns below and beyond the 7800\AA~break, the two resulting profiles are shifted of $\sim$ 0.4 arcsecond, consistently with the separation of the two sources measured in the ACS image. Interpreting this break due to the high redshift nature of the $i_{775}$-dropout source, the redshift is $\sim$ 5.4 with QF=''C''. We note that the uncertainty of the position of the break is increased by the presence of the sky absorption A-band at $\sim$7600\AA. \subsection{Galaxies showing a tilted [O\,{\sc ii}]3727 line} The current FORS2 spectroscopic catalog contains a sample of sources showing a spatially resolved [O\,{\sc ii}]3727 line with a characteristic ``tilt'' indicative of a high rotation velocity. Table~\ref{tab:OII_TILT} lists the 34 sources sorted by increasing redshift, the majority of them belong to the interval 1$<$z$<$1.5. Figure~\ref{fig:OIItilted1} show an example of the two dimensional spectra of the galaxies and the sky lines. The [O\,{\sc ii}]3727 line is marked with a circle. As discussed in the previous paper (\cite{vanz05}) the resolution of the FORS2 spectra favor the detection of high velocity rotational systems. Moreover a not optimized orientation of the slit suggest that in general the ``true'' maximum velocities may be significantly higher than the measure value. As an example, we have analyzed the velocity field of object J033227.73-275451.8 and estimate the stellar mass from the multi-wavelength dataset. We first traced the centroid of the [O\,{\sc ii}]$\lambda\lambda$3726,3729 emission line doublet along the spatial position. Since the resolution of our spectrum is too low to resolve the doublet, we fixed the ratio between the two components to 1 and we checked that the results were fairly insensitive to this assumption. We then compared this measured rotation curve with a set of synthetic rotation curves for which the velocity rises linearly up to one disk scale length and is flat at larger radii. This step takes into account the inclination of the disk with respect to the line of sight ($i=53\pm3 \deg$), the disk scale length ($r_d=0.405 \pm 0.045 \arcsec$) (derived from the morphological analysis, Rettura et al., in preparation), the slit misalignment with respect to the galaxy major axis ($22\pm2 \deg$), the width of the slit (1") and the seeing ($0.7\arcsec$) (this method is similar to that of \cite{boehm04}). The best fit rotation curve gives a rotation velocity of $355\pm50$ {\it km$s^{-1}$}. Using prescription from \cite{bosch02}, this implies a total halo mass of $1\pm0.4 \times10^{12}M\sun$. Dedicated spectroscopic observations specifically designed for the dynamical mass estimation (higher spectral resolution, optimized slit orientation, etc.), should be performed/preferred in order to decrease the uncertainties. We have used the full optical (HST/ACS B,V,i,z), near infrared (VLT/ISAAC J,Ks) to mid-infrared (Spitzer/IRAC $3.6 \mu, 4.5 \mu, 5.8 \mu, 8.0 \mu$) data to study the Spectral Energy Distribution (SED) of the same galaxy. We have adopted $1.5~arcsec$ radius aperture-corrected to $3.5~arcsec$ radius photometry to account for different instrumental PSFs. We have compared the observed SED with a set of template computed with P\'EGASE.2 models (\cite{RV97}) via $\chi^{2}$ minimization technique (the Salpeter IMF has been assumed). A more detailed description of the multi-wavelength cataloging and the fitting SED technique used here will be presented in Rettura et al. (in preparation) on a larger sample. We have calculated the errors for the mass estimate by sampling the full probability distribution in the parameters-space. Results of the SED fit are shown in Figure~\ref{fig:Vrot} right panel. We find a best-fit stellar mass of $2.6\times10^{10}~M_{\odot}$ with a 1$\sigma$ confidence interval between 1.7-3.7$\times10^{10}~M_{\odot}$. The comparison between this estimation and the dynamical halo mass produces a stellar mass over halo mass ratio of $f*=0.026$. This result is consistent with the estimations performed by \cite{conselice05} on a large sample of disk galaxies at $z\leq$1.1, where they find a wide range of $f*$ values (0.004 $\leq f* \leq $ 2). \begin{table} \centering \caption{Sample of 34 galaxies (RUN1 + RUN2) with tilted [O\,{\sc ii}]3727.} \begin{tabular}{lcccc} \hline \hline GDS ID & zspec& class & quality \\ \hline GDS J033254.87-275456.0 & 0.125 &em. & C \cr GDS J033237.54-274838.9 & 0.665 &em. & A \cr GDS J033215.88-274723.1 & 0.896 &em. & A \cr GDS J033227.66-275437.4 & 0.963 &em. & A \cr GDS J033227.73-275451.8 & 0.966 &comp. & A \cr GDS J033249.73-275517.4 & 0.981 &em. & A \cr GDS J033234.56-275543.6$\dag$& 0.983 &em. & A \cr GDS J033222.44-275606.1$\dag$& 1.090 &em. & C \cr GDS J033226.03-274856.0 & 1.016 &em. & A \cr GDS J033230.50-275312.3 & 1.017 &comp. & A \cr GDS J033225.28-275524.2 & 1.017 &comp. & A \cr GDS J033235.72-275615.4 & 1.033 &em. & A \cr GDS J033233.71-274210.2 & 1.043 &em. & B \cr GDS J033234.42-275405.7 & 1.088 &comp. & A \cr GDS J033225.86-275019.7 & 1.095 &em. & A \cr GDS J033246.71-274556.0 & 1.095 &em. & B \cr GDS J033247.42-274711.1 & 1.098 &em. & A \cr GDS J033215.23-274437.8 & 1.109 &em. & B \cr GDS J033223.18-274921.5 & 1.110 &em. & B \cr GDS J033216.28-274447.6 & 1.183 &em. & C \cr GDS J033216.26-274703.3 & 1.219 &em. & A \cr GDS J033238.01-275408.2$\dag$& 1.243 &em. & B \cr GDS J033224.94-275020.2 & 1.294 &em. & B \cr GDS J033205.67-274253.5 & 1.296 &em. & A \cr GDS J033232.42-274150.1$\dag$& 1.296 &em. & B \cr GDS J033232.47-274151.5$\dag$& 1.296 &em. & B \cr GDS J033213.21-274158.0 & 1.297 &em. & B \cr GDS J033240.94-274427.5 & 1.298 &comp. & A \cr GDS J033244.35-275506.4$\dag$& 1.305 &em. & A \cr GDS J033230.71-274617.2 & 1.307 &em. & A \cr GDS J033234.82-274721.9$\dag$& 1.316&em. & B \cr GDS J033239.66-275406.3 & 1.343 &em. & A \cr GDS J033240.08-275532.6 & 1.461 &em. & A \cr GDS J033229.06-275542.8 & 1.469 &em. & A \cr \hline \multicolumn{4}{l} {$\dag$ possible tilted line.}\\ \label{tab:OII_TILT} \end{tabular} \end{table} \section{Conclusions} As a part of the Great Observatories Origins Deep Survey, a large sample of galaxies in the Chandra Deep Field South has been spectroscopically targeted. After the RUN1 (\cite{vanz05}) and RUN2 (present work) a total of 930 objects with $z_{850} \mincir 26.8$ have been observed with the FORS2 spectrograph at the ESO VLT providing 724 redshift determinations. From a variety of diagnostics the measurement of the redshifts appears to be precise (with a typical $\sigma_z \simeq 0.001$) and reliable. The reduced spectra and the derived redshifts are released to the community ($\it{http://www.eso.org/science/goods/}$). They constitute an essential contribution to reach the scientific goals of GOODS, providing the time coordinate needed to delineate the evolution of galaxy masses, morphologies, and star formation, calibrating the photometric redshifts that can be derived from the imaging data at 0.36-8$\mu$m and enabling detailed studies of the physical diagnostics for galaxies in the GOODS field. \begin{acknowledgements} We are grateful to the ESO staff in Paranal and Garching who greatly helped in the development of this programme. The work of DS was carried out at the Jet Propulsion Laboratory, California Institute of Technology, under a contract with NASA. We thank the ASI grant I/R/088/02 (SC, MN, EV). \end{acknowledgements}
Title: The Paths of Quintessence
Abstract: The structure of the dark energy equation of state phase plane holds important information on the nature of the physics. We explain the bounds of the freezing and thawing models of scalar field dark energy in terms of the tension between the steepness of the potential vs. the Hubble drag. Additionally, we extend the phase plane structure to modified gravity theories, examine trajectories of models with certain properties, and categorize regions in terms of scalar field hierarchical parameters, showing that dark energy is generically not a slow roll phenomenon.
https://export.arxiv.org/pdf/astro-ph/0601052
\title{The Paths of Quintessence} \author{Eric V.\ Linder} \affiliation{Berkeley Lab, University of California, Berkeley, CA 94720} \date{\today} \section{Introduction} \label{sec.intro} The discovery of the acceleration of the cosmic expansion has thrown physics and astronomy research into a ferment of activity, from a search for fundamental theories to investigation of predictions relating models and the cosmological dynamics, to development of astrophysical surveys yielding improved measurements. The acceleration, or more generally the expansion history of the scale factor evolution with time, $a(t)$, can be equivalently treated by a dark energy pressure to density, or equation of state, ratio $w(a)$ \cite{linjen}. One model, Einstein's cosmological constant, predicts $w=-1$ at all times, but generically the dark energy phenomenon has dynamics, a time varying $w(a)$. It is important to note that the current epoch of accelerated expansion is very different from the early epoch of inflation. We know a priori that dark energy does not completely dominate the universe now and we do not know a priori that dark energy obeys a slow roll approximation (in fact we will see it is unlikely to). In these senses, dark energy is a more challenging phenomenon than inflation. We are faced with three ``Goldilocks'' problems: 1) dark energy is dynamically important, but not fully dominant, with $\sim75\%$ of the total energy density today, 2) the universe is accelerating, as from a component with $w\lesssim-0.8$, but the field responsible may not be slowly rolling (unless fine tuned) as it would if nearly completely potential dominated, and 3) if it's not the cosmological constant, what happened to the cosmological constant? To discover whether the physics is the cosmological constant, and to distinguish between alternate theories, requires measurement of the possible dynamics. Caldwell \& Linder \cite{caldlin} (Paper 1) investigated the dynamics in the phase plane of $w$-$w'$, where $w'=dw/d\ln a$, for canonical scalar field dark energy, or quintessence, finding that this reveals important clues to the nature of the new physics. In such a plane, the time or scale factor variable is a parameter along the paths of dynamics. They found distinct structure, categorizing fields into those that at early times are locked by Hubble friction into a cosmological constant like state, and then move away from this (thaw) as the dark energy dominates, and those that initially roll due to the steepness of the potential but later approach the cosmological constant (freeze). In this article some of these results are put on a firmer footing, examined in greater detail, and extended to models beyond canonical scalar fields, including modified gravity theories. \section{General Dynamics} \label{sec:scadyn} We begin with a brief, explicit derivation of the key dynamics equation, Eq.~(\ref{eq:wp}). A scalar field $\phi$ possesses both potential energy $V(\phi)$ and kinetic energy $(1/2)\dot\phi^2$. The Lagrangian density is of the form \beq {\mathcal L}=(1/2)g_{\mu\nu}\partial^\mu\phi\partial^\nu\phi-V(\phi), \eeq for a canonical, minimally coupled scalar field in a metric $g_{\mu\nu}$. The equation of motion for the field, the Klein-Gordon equation, \beq \ddot\phi+3H\dot\phi+V_{,\phi}=0, \label{eq:kg} \eeq where $H=\dot a/a$ is the Hubble parameter, follows from functional variation of the Lagrangian. (Spatial inhomogeneities should be negligible on subhorizon scales, but also see \S\ref{sec:inhom}.) The energy-momentum tensor is generated through Noether's theorem and one can identify the energy density and pressure: \beq \rho=(1/2)\dot\phi^2+V(\phi)\quad;\quad p=(1/2)\dot\phi^2-V(\phi). \eeq It will be convenient to invert these and write the potential and kinetic energies in terms of $\rho$ and $w$: \beq V=(1/2)(1-w)\rho\quad;\quad (1/2)\dot\phi^2=(1/2)(1+w)\rho. \label{eq:vk} \eeq Note that $w$ and $\rho$ are both functions of time, or scale factor. To obtain an equation for the variation $w'$, take the derivative of $V$ with respect to $\phi$, \beq \vp=\dot V/\dot\phi=(1/2)[(1-w)\dot\rho-\rho\dot w]/[(1+w)\rho]^{1/2}. \label{eq:vdot} \eeq Employing the continuity equation $\dot\rho=-3H\rho(1+w)$ and $H=d\ln a/dt$, one obtains \beq w'=-3(1-w^2)-(1-w)(1+w)^{1/2}\sqrt{(3/8\pi) \Omega_\phi(a)}\frac{M_P \vp}{V}, \label{eq:wp} \eeq where $\Omega_\phi(a)=8\pi\rho/(3H^2M_P^2)$ is the dimensionless dark energy density, and $M_P$ is the Planck mass. \subsection{Distinguishing $\Lambda$ \label{sec:barrier}} We see that $w'$ is related to the nearness of the equation of state ratio to the cosmological constant value, i.e.\ $1+w$, and the inverse of the characteristic field scale of the potential, $\vp/V$ (sometimes phrased as a slow roll parameter, $M_P|\vp/V|\ll1$). If $w$ is readily distinguished from $-1$, then we know the dark energy is not a cosmological constant, regardless of the value of the time variation $w'$. More difficult is the case where $\eps=1+w$ is a small quantity. Then it will be quite important to determine whether $w'$ is zero or not. Eq.\ (\ref{eq:wp}) guides us in following the dynamics in the $w$-$w'$ phase space. The first point to notice is that the reciprocal of the characteristic field scale is not generically a small parameter useful for a ``slow roll'' approximation. Figure \ref{fig:wwpscale} shows curves of constant field scale \beq \Phi=V/(-\vp), \eeq in the $w$-$w'$ plane. Only a tiny sliver of the phase space, plus a small hump, satisfies the slow roll approximation; unless one is exceedingly close to the cosmological constant behavior there is substantial dynamics in the field. When $1+w\ll1$, the second term in Eq.~(\ref{eq:wp}) dominates and $w'>0$ (creating the ``humps''), while for less negative $w$ the first term can dominate and drive $w'<0$. This driving occurs closer to $w=-1$ for larger $\Phi$. Within the large-$\Phi$ hump, the dark energy looks similar to a cosmological constant. Suppose one conjectures some physics limiting the size of the field scale, equivalently leading to avoidance of flatness in the potential. An upper bound on $\Phi$ in the scalar field behavior would impose a barrier around the cosmological constant $\Lambda$, saying that the scalar field dynamics must be distinguishable from $\Lambda$, if it is not $\Lambda$. The second panel of Fig.~\ref{fig:wwpscale} zooms in to illustrate this barrier for $\Phi<M_P$, which rules out any freezing field, and any thawing field with $1+w<0.004$. So any scalar field theory with field scales barred from being transPlanckian are distinguishable from $\Lambda$ at this precision. If the restriction uses, say, $\Phi<M_P/\sqrt{8\pi}$ (i.e.\ the Planck mass is defined in terms of Newton's constant as $G_N=(8\pi M_P^2)^{-1}$ rather than $G_N=M_P^{-2}$ as above) then the limit becomes $1+w>0.1$. The physical origin for the conjectured limit on the characteristic field scale is not clear, but the implications are important enough to consider the possibility. Since the field scale is related to the inverse of the flatness of the potential, then physics that perturbs a flat direction in the potential would give this effect. In some supersymmetric models, loop corrections generate a logarithmic tilt $V\sim\phi^n\ln(\phi/\mu)$ \cite{witten,damourmukhanov}. This would give $\Phi\approx\phi/n$ (or $\phi\,\ln(\phi/\mu)$ for $n=0$), and restricting $\phi<M_P$ (for the effective field theory to be valid), provides the limit $\Phi\lesssim{\cal O}(M_P)$. However, the generic breakdown of slow roll is independent of any $\Phi$ upper limit conjecture and we do not consider the latter further. Hierarchical parameters to replace slow roll are discussed in \S\ref{sec:slow}. \subsection{Driving and dragging \label{sec:kglines}} Returning to the Klein-Gordon equation, we can understand behavior in the $w-w'$ phase space by first a general and then a specific analysis of the terms. Writing Eq.~(\ref{eq:kg}) as $\ddot\phi+3H\dot\phi=-\vp$, we can require $-\vp\ge0$, reflecting that the field rolls down the potential to large $\phi$. Using Eq.~(\ref{eq:vk}) for $\dot\phi$, we write this condition in terms of $w$, $w'$ as $w'\ge-3(1-w^2)$. The boundary defines the null line $\vp=0$ discussed further below. Similarly, writing the Klein-Gordon equation as $\ddot\phi+\vp=-3H\dot\phi$ and again using that the field rolls to larger values with time implies $-3H\dot\phi\le0$. This reduces to the condition $w\ge-1$. If we flip the direction of the field motion and potential slope, i.e.\ the field rolls down the potential to smaller $\phi$, then these conditions remain. (I.e.\ $\dot V=\dot\phi\vp$ is still negative, and so the transition from Eq.~(\ref{eq:vdot}) to Eq.~(\ref{eq:wp}) flips the sign of the second term on the right hand side of Eq.~(\ref{eq:wp}), canceling the reversed sign of $\vp$.) However, if we make the field move up the potential then we have the relations $w'\le-3(1-w^2)$ and $w<-1$, in the phantom region, i.e.\ the boundary lines just continue smoothly through $w=-1$. Finally, we can move $\ddot\phi$ to the right hand side to obtain $3H\dot\phi+\vp=-\ddot\phi$. This divides the $w-w'$ plane into regions where the field is accelerating or decelerating, with the boundary being the coasting behavior $\ddot\phi=0$. This line corresponds to the condition $w'=3(1+w)^2$. Larger values of $w'$ arise from a field accelerating down the potential, while smaller values come from a field decelerating (this motion of the field should not be confused with the accelerating expansion of the universe, which can hold for either region). These three boundaries -- the null line $w'=-3(1-w^2)$, coasting line $w'=3(1+w)^2$, and phantom line $w=-1$ -- define the broad characteristics of the phase plane. To investigate the dynamics further, we must examine the dominance of the different terms in the equation of motion (\ref{eq:kg}). The first term is the acceleration of the field, the second term a friction term, or Hubble drag, due to the expansion of the universe, and the third is a driving term due to the steepness of the potential. We define \beqa X&=&\frac{\ddot\phi}{H\dot\phi} \\ Y&=&\frac{\ddot\phi}{\vp}\ . \eeqa Figure~\ref{fig:kgterms} shows curves of constant $X$, $Y$ in the $w-w'$ plane. Note that $X=Y=0$ corresponds to an epoch of coasting in the scalar field dynamics, $\ddot\phi=0$, as discussed above. This is nongeneric, as the field would need to be finely tuned to neither accelerate due to the slope of the potential nor decelerate due to the Hubble drag, but be perfectly balanced. Indeed, the dynamics of scalar fields in Paper 1 avoid this region, causing the split into the distinct thawing and freezing regions, respectively above and below this line. In the accelerating field region, the friction term is the major determinant of behavior initially, as the field evolves away from a frozen (cosmological constant-like) state in the matter dominated epoch. The upper boundary of the thawing region is given by $X=3/2$, where this value follows directly from the exponent of the expansion history, $t\sim a^{3/2}$ for matter domination. Thus fields that thaw during matter domination begin to move along the $X=3/2$ line (see discussion, and Figure 2, in Paper 1). We can translate any acceleration to friction ratio $X$ to a $w-w'$ behavior through \beqa w'&=&2X(1+w)+3(1+w)^2=(1+w)(3+2X+3w) \nonumber \\ &\approx& 2X(1+w), \eeqa where the last approximation is for $1+w\ll1$. Note that the linear boundaries used in Paper 1 were (good) approximations to the general parabolic behavior. The value $X=0$ gives the coasting line and $X=-3$ gives the null line. Thus, the upper thawing boundary $X=3/2$ corresponds to $w'\approx 3(1+w)$ for $1+w\ll1$. If the rolling field then enters a region where the potential slope is shallower (as usually happens), then the field will accelerate less and curve toward the $\ddot\phi=0$ line. Since today the field cannot have rolled so far that $\Omega_{\rm de}>0.8$, the dynamical track remains within the thawing region $1+w<w'<3(1+w)$, i.e.\ $1/2<X<3/2$. For potentials that steepen as the field rolls down, e.g.\ PNGB models with the field starting near the top of the potential, the tracks instead lie above the $X=3/2$ or $w'=3(1+w)$ line. The PNGB potential also steepens more rapidly for small symmetry scales $f$, and indeed, as mentioned in Linder \cite{osc}, PNGB models roughly follow $w'=F(1+w)$, where $F$ is proportional to the inverse of $f$. (Also see \cite{kalopersorbo} for discussion of PNGB models, fine tuning, and slow roll.) In the decelerating field region, the steepness of the potential impacts the freezing. As the potential becomes shallower, the friction is more effective. In the limit of a flat potential, one obtains the dynamics track given by the null curve, $\vp=0$, in Fig.~\ref{fig:kgterms}. As given in Paper 1, this corresponds to $w'=-3(1-w^2)$, and the skating model of \cite{lincurv,liddleskate}. In terms of the friction and driving terms, $X=-3$ and $Y=\infty$. The lower boundary of the freezing region lies along the $Y=1$ ($X=-3/2$) line, equivalent to $w'=3w(1+w)$. (See \S\ref{sec:track} for a rationale.) We will later see that this line also has physical significance. The general relation between the acceleration to steepness ratio $Y$ and the $w-w'$ track is \beq w'=3(1+w)\left[w+\frac{1-Y}{1+Y}\right]=3w(1+w)+3(1+w)\frac{1-Y}{1+Y}. \eeq Again we have a parabolic behavior. The coasting line has $Y=0$, and the null line corresponds to $Y\to\infty$. One could use either variable $X$ or $Y$ throughout the phase plane, since \beq X=-3\frac{Y}{1+Y}\quad ; \quad Y=-\frac {X}{X+3}, \eeq but this somewhat obscures the physics of friction and steepness. (That said, we note that the thawing/freezing boundaries are fairly symmetric in $X$, with the outer boundaries at $X=\pm3/2$ and the inner ones at $X\approx\pm1/2$.)\footnote{Recall from Paper 1 that the upper bound on the freezing region is not sharply defined, and extends somewhat above the $w'=w(1+w)$ line shown for convenience in this paper.} Note that it is not legitimate to assume tracking behavior (where the equation of state is constant and related to the dominant component's equation of state) to impose limits on regions in the phase plane, as for example \cite{chiba,scherrer} did to try to tighten the constraints of Paper 1. For one thing, not all freezing models need start as trackers. Secondly, just because the most negative value of $w'$ lies above some boundary curve does not ensure that the entire trajectory remains above the curve. Most importantly, the tracking approximation breaks down as the dark energy first becomes significant, so it is inapplicable for much of the observable dynamical history. \section{More Specific Dynamics \label{sec:more}} In addition to analyzing general behavior through the Klein-Gordon equation terms, we can investigate the properties of the phase plane or specific track families in terms of other variables. These could include working from the equation of state $w-w'$ relation directly or the cosmic expansion history $a(t)$. While not quite as insightful as the physics motivated driving and drag terms, they can highlight interesting properties. \subsection{Mocker models \label{sec:mocker}} Consider a model with dynamics given by $w'=Cw(1+w)$. Note that while such an equation forms the boundaries of the freezing region, freezing models do not follow such a trajectory but rather are almost orthogonal to such tracks (at least initially). So we are talking about fundamentally different models. The behavior of the dark energy equation of state and density follow \beqa w(a)&=&-1+\left[1-\frac{w_0}{1+w_0}a^C\right]^{-1} \\ \rho_{\rm de}(a)&=&\rho_{\rm de}(1)\,\left[(1+w_0)a^{-C}-w_0\right]^{3/C}, \eeqa where $w_0$ is the equation of state today. In the past, $a\ll1$, the component will act like additional nonrelativistic matter, with $w=0$, while in the future it will approach a cosmological constant. Since such dark energy sometimes looks like dark matter and sometimes like a cosmological constant, we call this a ``mocker'' model. These are basically what are known as ``quartessence'' models (see \cite{makleroliveirawaga} for an overview), of which the Chaplygin gas is one example. However, we develop them directly from the phase space dynamics rather than an ansatz for the pressure, so the dynamical behavior is more general. For example, a constant pressure model could be a cosmological constant, or could be a mocker with $C=3$. Note $C=3$ gives precisely the expression for the lower boundary of the freezing region (or is $Y=1$ or $X=-3/2$ in the notation of \S\ref{sec:kglines}). We name the $w'=3w(1+w)$ curve the constant pressure line (also see \cite{scherrer} in the context of barotropic fluids). Such combined behavior models are heir to all the usual problems of trying to unify dark matter and dark energy, e.g.\ growth instabilities of density perturbations \cite{sandvik,amendola0509099}. Analysis of perturbations requires knowledge of the full theory, however. Merely from the phase plane dynamics, though, we can see trouble arising for such unified models. As $C$ gets smaller, the model moves along its trajectory more quickly, acting less like dark matter except at very early times, $1+z\gg[-w_0/(1+w_0)]^{1/C}$. Conversely, as $C$ gets larger, acceleration of the cosmic expansion occurs later and the model becomes a poorer fit to a host of (purely geometric) observations such as supernova distances and the distance to the CMB last scattering surface. \subsection{Relation to parameterized $w(a)$ \label{sec:parw}} The approach taken in this article is to examine dark energy dynamics directly in the phase plane $w-w'$, where a time variable runs along each trajectory. It is useful to see the relation of standard parametrizations in terms of the temporal behavior, i.e.\ $w(a)$, to this approach. The standard two parameter function $w(a)=w_0+w_a(1-a)$ was shown by Linder \cite{linprl,linpr} to provide an excellent approximation to exact solutions of the Klein-Gordon equation in a wide variety of models. In this ansatz we have $w'=-aw_a=w-w_\infty$, where the high redshift equation of state $w_\infty=w_0+w_a$. This describes a straight line of slope 1 in the $w-w'$ plane, and can be rewritten as as $w'=(1+w)-(1+w_\infty)$. In particular, if $w_\infty=-1$ we have exactly the behavior of thawing models (lying along lower bound of that region). From Fig.~2 of Paper 1, we see as well that many tracking models that fit present data (i.e.\ $\Omega_{\rm de}\sim0.7$ and $w<-0.8$) are reasonably well described, on average, by a line of slope unity. Of course this approximation will break down in the future, as the field freezes more fully, turning toward the cosmological constant; at the same time this $w(a)$ ansatz loses validity as it moves toward ever more negative $w$. However, since data only exist toward the past, we see why the $w_a$ parametrization is an excellent approximation. To retain boundedness for both the past and present, as well as to allow more dramatic dynamics (essentially slopes other than unity in $w-w'$), one could use the ``e-fold'' model of \cite{eospar} or ``kink'' model of \cite{corakink}. Both utilize four parameters for their description. The e-fold model has a more transparent translation to $w'=dw/d\ln a$ since it also uses dynamics in terms of $\ln a$: \beq w(a)=w_f+\frac{\Delta w}{1+(a/a_t)^\tau}, \eeq where $\tau$ is the transition rapidity in units of units of e-folds $\ln a$, $a_t$ is the transition scale factor, $w_f$ is the asymptotic future value, and $\Delta w=w_p-w_f$ is the difference between asymptotic past and future values. In the $w-w'$ phase plane we have \beq w'=-\tau(w-w_f)\left(1-\frac{w-w_f}{\Delta w}\right). \eeq Note that as we found in the Klein-Gordon equation analysis of \S\ref{sec:kglines} the equation for $w'$ is quadratic in $w$. We can identify several special cases. If the asymptotic future state is deSitter ($w_f=-1$), then $w'=\tau(\Delta w^{-1}-1)(1+w) +(\tau/\Delta w)\,w(1+w)$. This looks like the sum of a thawing model and a model in the freezing region, i.e.\ the dark energy can be viewed as the sum of two components. If we further take $\Delta w=1$, then we remove the thawing component and end up with $w'=\tau\,w(1+w)$ -- a mocker model with $w_p=0$ and $w_f=-1$. Starting instead with an asymptotic past state of $w_p=-1$ gives $w'=\tau\,(1+w)(w-w_f)/(-1-w_f)$. In the limit $w_f\to\infty$ (i.e.\ not worrying about the region where there is no data) this gives a thawing model $w'=\tau(1+w)$. Thus the four parameter e-fold ansatz is also quite versatile. The rapidity parameter is directly related to both the slope of and the velocity along the phase space trajectory, and ties in with the steepness of the scalar field potential, as we saw in \S\ref{sec:kglines} with the PNGB models where the slope was proportional to the inverse of the symmetry scale $f$. Finally, one could invert the situation and go from a parametrization in the phase plane to derive the function $w(a)$. For example, a track $w'=A(1+w)+B(1+w)^2$, which we have seen is a common form, implies \beq 1+w=(1+w_0)/\left[(1+x)a^{-A}-x\right], \eeq where $x=(B/A)(1+w_0)$ defines the present along the trajectory (equivalently the dimensionless matter density $\om$ today). Note that while the trajectory has two parameters, the equation of state has three parameters since we must have a parameter running along the track. At high redshift, if $A>0$ then $w\to-1$ and we have a thawing model, asymptotically independent of $B$. If $A<0$ then $w(z\gg1)=-1+(-A/B)$ and $w'\to0$, i.e.\ it begins like a tracking model. It reaches a minimum $w'_{\rm min}=-A^2/(4B)$ at $w_\star=-1-A/(2B)=[-1+w(z\gg1)]/2$; that is, the trajectory is a parabola from its tracking value of the equation of state to its future, cosmological constant value. A mocker model is the special case $C=-A=B$. For completeness, we give the dark energy density: \beq \rho_{\rm de}(a)=\rho_{\rm de}(1)\,(1+x-xa^A)^{3/B}. \eeq Of particular interest is the ``leveling'' model where, in loose physical analogy to the inflationary power spectrum tilt $n-1$ being driven to zero by large numbers of e-folds of expansion, the equation of state tilt $1+w$ is driven to zero by the deSitter expansion as the energy density approaches a certain constant value, $\rho_f$. That is, take $1+w=D[\rho_{\rm de}(a)-\rho_f]$. This is equivalent to the above parabolic model with $A=-3D\rho_f$ and $B=-3$. Another interesting parabolic track is the coasting line $w'=3(1+w)^2$. This corresponds to not a leveling but a tilting, with $1+w=(1+w_0)\,[\rho_{\rm de}(1)/\rho_{\rm de}(a)]$, so $w$ is tilted away from $-1$ as the energy density decreases. \subsection{Acceleration and jerk \label{sec:jerk}} One could leave behind the physics of the accelerating phenomenon and instead use variables in terms of the acceleration itself, though this seems less appealing. The deceleration parameter \beq q=-a\ddot a/\dot a^2=\frac{1}{2}+\frac{3}{2}w\Omega_{de}(a), \eeq and the jerk \beq j=a^2\, \dddot a/\dot a^3=1-\frac{3}{2}\Omega_{de}(a)\,[w'-3w(1+w)]. \eeq We also have $j=q+2q^2-q'$. Note that interpreting $q$ and $j$ or $q'$ as Taylor expansions of the expansion is of strictly limited use (since observations span $\Delta\ln a\sim{\cal O}(1)$) and can be dangerous \cite{noqexp}. Furthermore, there is the same ambiguity there was with using pressure as a variable. We note that any model where it touches the constant pressure line $w'=3w(1+w)$ has $j=1$; this is equivalent to $X=-3/2$ in the notation of \S\ref{sec:kglines}. The two standard special cases of $j=1$ lie at the ends of this line: an Einstein-de Sitter universe with $w=0$ ``dark energy'' and a $\Lambda$CDM universe with cosmological constant dark energy. The constant pressure line is also related to the adiabatic sound speed of the dark energy. (Note this is not the true sound speed of perturbations arising from the microphysics of whatever the dark energy is.) The adiabatic sound speed \beq c_a^2=\frac{\dot p}{\dot\rho}=w-\frac{1}{3}\frac{w'}{1+w}, \label{eq:ca} \eeq and we see it vanishes for $w'=3w(1+w)$. On the null line, the adiabatic sound speed equals the speed of light (the same as the true sound speed for a canonical scalar field). Models below the null line would need to have $c_a^2>1$. \section{Beyond Scalar Fields \label{sec:beyond}} For the cosmic expansion dynamics we can always define an effective equation of state even if the accelerating mechanism is not a scalar field \cite{linjen}, through \beq w_{\rm eff}=-1-\frac{1}{3}\frac{d\ln\delta H^2}{d\ln a}, \label{eq:weff} \eeq where $\delta H^2=(H/H_0)^2-\om a^{-3}$ is the unknown part of the Hubble parameter, that not due to matter. So it is of interest to see to what extent the dynamical behaviors we have discussed carry over to the $w_{\rm eff}-w'_{\rm eff}$ plane. That is, are freezing and thawing behaviors more general than for scalar fields, and do the null and coasting lines still play a role? Due to the diversity of possible accelerating physics we do not present a general analysis of these important questions but rather calculate some specific cases. \subsection{Scalar-tensor gravity \label{sec:scatens}} Scalar-tensor theories modify the Einstein-Hilbert action with both an additional scalar field and a coupling to Ricci scalar curvature $R$. These are of great interest as a comparison in tests of general relativity and also because gravitational theories involving a nontrivial function of the scalar curvature can be transformed to scalar-tensor theories. See \cite{scatensreview} for a general introduction. We consider coupling of a general form in the scalar field, but linear in the curvature. So the general relativistic $R/(8\pi G)\to F(\phi)R$. One then obtains the usual Friedmann expansion equations, with extra terms giving an effective dark energy density \cite{bmp} \beq \rho_{\rm ST}=V(\phi)+(1/2)H^2(q-1)(q+5)F_\phi^2+3H^2[(8\pi G)^{-1}-F], \label{eq:rhost} \eeq where $q$ is the deceleration parameter, $F_\phi=dF/d\phi$, and the last term involves the change of the gravitational strength from Newton's constant $G$. From this density one can then define $\delta H^2=8\pi G\rho_{\rm ST}/(3H^2)$ (note that we use the usual $G$ here since the deviation is absorbed into $\rho_{\rm ST}$ as mentioned above). From this the equation of state $w_{\rm eff}$ and its variation $w'_{\rm eff}$ can be calculated. A key quantity will be $F/F_\phi^2\equiv\omega_{JBD}$. This is the Jordan-Brans-Dicke parameter and its inverse must be very small according to solar system tests. Expanding $F$ about the present, \beqa F(a)&\approx&(8\pi G)^{-1}-(1-a)F_\phi \dot\phi/(aH) \\ &\approx&(8\pi G)^{-1}-z(1-q)F_\phi^2. \eeqa So the ratio of the second to the first term ($\sim\omega_{JBD}^{-1}$) is small, and gravity is nearly Einsteinian. But this means that the first term in $\rho_{\rm ST}$ dominates (unless $8\pi GV/H^2\ll1$, but then it doesn't affect the expansion and there is no acceleration). Thus, the restriction of the scalar-tensor theory by solar system constraints means that its effective equation of state must be very close to a cosmological constant -- within $\sim\omega_{JBD}^{-1}$. Since $\omega_{JBD}^{-1}<2.5\times 10^{-5}$, this would be rather challenging to distinguish from a cosmological constant with cosmological observations! One possible loophole is if the solar system limits on the scalar coupling should not be applied to a cosmological situation because of the different spacetime backgrounds with very different scalar curvatures. This arises for example in chameleon scenarios \cite{chameleon}. The most stringent cosmological bounds on varying $G$ arise from primordial nucleosynthesis and give $\omega_{JBD}^{-1}\lesssim 3\times10^{-3}$ \cite{gvarycos} (but see \cite{nsvary}). Calculation of the effective phase plane parameters finds \cite{bmplinder} \beqa w_{\rm eff}(z=0)&=&-1+0.46/\omega_{JBD} \\ w'_{\rm eff}(z=0)&=&-0.36/\omega_{JBD}. \eeqa While even with only the cosmological bounds on $\omega_{JBD}$ these are quite close to the cosmological constant in the phase plane, it is interesting to note that the current values lie along $w'=0.78w(1+w)$, in the freezing region. Indeed its trajectory is a freezing one, with scalar-tensor theories asymptotically attracted to general relativity \cite{damourst,rboost}, and to $w=-1$. One last thing to note, however, is that because scalar-tensor theories possess anisotropic stress, the growth of density perturbations will be modified from the quintessence case (see, e.g., \cite{stgrowth} and references therein). \subsection{Braneworld cosmology and $H^\alpha$ \label{sec:brane}} In a braneworld cosmology \cite{dgp,deffayet}, effective acceleration appears due to a weakening of gravity on large scales as it ``leaks'' from our brane into a higher dimensional bulk. The Friedmann expansion equation becomes \beq H^2-H/r_c=8\pi G\rho_m/3, \eeq where $r_c$ is the crossover distance and $\rho_m$ the matter density. The effective equation of state due to the modification is $w_{\rm eff}= -[1+\om(a)]^{-1}$ \cite{lueeos}. Its trajectory in the phase plane is plotted in Fig.~\ref{fig:bw}. Note that it looks like a freezing model, and will indeed approach a cosmological constant in the asymptotic future. The position along the trajectory is a time variable, so taking the present to be, say, $\om=0.2$ would extend the solid curve in Fig.~\ref{fig:bw} slightly further (since the figure uses $\om=0.3$). We also see why it is so well approximated by a $w_0-w_a$ model, as discussed in \S\ref{sec:parw}. (Recall, however, that $w_a$ is actually defined at $z=1$, not $z=0$, to give the best physics fit \cite{linprl}.) We can further generalize the modification to $\delta H^2\sim H^\alpha$ \cite{dvaliturner}. Then we find \beqa w_{\rm eff}&=&-\left[1-\frac{\alpha}{\alpha-2}\om(a)\right]^{-1} \\ w'_{\rm eff}&=&3w(1+w)[1-(2/\alpha)(1+w)]. \label{eq:wpalpha} \eeqa Recall that the braneworld case above corresponds to $\alpha=1$. For acceleration today (with $\om=0.3$), we require $\alpha<1.57$; for $w<-0.8$ today we require $\alpha<0.91$. Note that all $H^\alpha$ modified gravity models will look similar (one does require $\alpha<2$ for a negative equation of state at early times). See \S\ref{sec:track} for discussion of their tracking behavior. In particular, they all approach the cosmological constant along $w'=3w(1+w)$, what was called the constant pressure line for scalar fields. Their tracks must always lie between $w'=3w(1+w)$ and the $w'=0$ axis. When $\alpha<0$, the trajectories switch to the phantom regime with $w<-1$, but the bounds still hold. \section{Polytropic Dark Energy \label{sec:poly}} An interesting, if phenomenological, way of obtaining acceleration is to modify the Friedmann expansion equation but keeping a pure matter universe. While this leaves open important questions about its relation to fundamental theory and the growth of density perturbations, we can investigate some general aspects of the effective equation of state dynamics. Consider general functions of the matter density (sometimes referred to as barotropic models \cite{scherrer}) \beq H^2=(8\pi G/3)\,g(\rho)=(8\pi G/3)\,[\rho+f(\rho)]. \eeq The quantity $f(\rho)$ will act like an effective dark energy. Using Eq.~(\ref{eq:weff}) we can define \beqa w_{\rm eff}&=&-1+\frac{d\ln f}{d\ln\rho} \\ w'_{\rm eff}&=&-3\,\frac{d^2\ln f}{d\ln\rho^2}, \eeqa and identical relations hold for the total equation of state $w_{\rm tot}$ and its variation $w'_{\rm tot}$ upon substituting $g$ for $f$. The simplest example of such a modification is $f\sim\rho^n$, the Cardassian model of \cite{freeselewis}. From the above equations we see that it corresponds to a constant equation of state $w=-1+n$. If we require $w<-0.9$ (as observations favor for a constant equation of state), then $n<0.1$; unfortunately $\rho^{1/10}$ does not obviously appear to be a natural modification of the Friedmann equation. We can investigate the phase space dynamics by relating $w'$ to $w$: \beq w'=3w(1+w)-3\frac{\rho^2}{f}\frac{d^2f}{d\rho^2}. \eeq Immediately we see that whether the effective dark energy lies below the freezing region or not depends on the sign of $d^2f/d\rho^2$. An equivalent question is whether the effective pressure is decreasing or increasing with time (since the $w'=3w(1+w)$ line is that of constant pressure). The model will follow the mocker model track $w'=3w(1+w)$ if $f=A+B\rho$. This form is equivalent to a redefinition of $\om$, e.g. $\om\to\om(1+B)$, plus a cosmological constant $A$. As such it has a nonzero minimum in its effective potential, allowing it to reach the freezing boundary. It is important to remember that the analysis in this paper and Paper 1 applies to the dynamics of the dark energy itself. Trajectories in the $w_{\rm tot}-w'_{\rm tot}$ plane convolve the matter and dark energy components, mixing the dynamics and so not giving rise to the clear differentiations and regions found. This is why Ref.~\cite{scherrer} appears to find a violation of the freezing bound and even null bound for some barotropic models; those are actually phantom models in the dark energy phase space, but are dragged by the matter to $w_{\rm tot}>-1$. We make this more explicit later in this section. For a richer dynamical behavior we propose a class of modifications of the Friedmann expansion equation we call polytropic models \cite{freese02}. Here \beq H^2/(8\pi G/3)=g(\rho)=\rho\,[1+(\rho/\rhos)^{-\alpha}]^\beta. \label{eq:polyg} \eeq At densities much greater than some crossover value $\rhos$, e.g.\ at high redshift, the Friedmann equation is standard. At low densities, the expansion is modified, with $w_{\rm tot}$ asymptotically approaching $-\alpha\beta$. If we want a future deSitter state, we could choose $\beta=1/\alpha$. For just the effective dark energy equation of state, the value in the past is $w_{\rm eff}=-\alpha$, and in the future of course it dominates so $w_{\rm eff}=-\alpha\beta$. Figure~\ref{fig:polyw} illustrates the phase plane dynamics. The first panel takes $\beta=1/\alpha$, so that the future state is deSitter. Models with $\alpha<1$ have $w\ge-1$ and act like freezing models, starting from a constant $w=-\alpha$ and today (marked by crosses) lying in the freezing region, before heading toward the cosmological constant. Phantom models have $\alpha>1$ and act like mirror images of freezing models, even to lying within the phantom freezing region today. In the second panel, we fix $\beta=1/2$. While the models start at the same phase space point as the previous models with the same $\alpha$, now their endpoints are different. Indeed for $\alpha<2/3$ the acceleration of the expansion is a temporary phase. Furthermore, the trajectories with $\alpha\lesssim1.5$ do not lie in the freezing regime and both regular and phantom models have $w'>0$. Such polytropic models without a deSitter future will be clearly distinguishable from both freezing and thawing quintessence. Can we give some physics motivation for the polytropic form, aside from its simplicity and proper asymptotic behavior? When $\beta=0$, there is no modification; when $\beta=1$ we have the power law modification of the Cardassian case, with $n=1-\alpha$, hence a constant $w=-\alpha$. There are some motivations for power law modification from higher dimension theories (for $n<1$ see \cite{freeselewis}, the nonaccelerating $n=2$ arises in Randall-Sundrum brane scenarios \cite{randallsundrum}). When $\beta=1/2$, the modification is similar to that from a Chaplygin gas \cite{chaplygingas}, as we see below; this has claimed motivation from Born-Infeld actions and brane solutions \cite{bento}. So at least the polytropic form unifies different prescriptions for dark energy. Note that as $\beta$ increases from zero, for fixed $\alpha$, the size of the ``hump'' in the trajectory decreases and the future value of $w$ moves back toward the initial value. At $\beta=1$ the trajectory collapses to a point at $w=-\alpha$, $w'=0$. For even larger $\beta$, the hump is flipped (i.e.\ the sign of $w'$ changes) and again increases in size, with the future value of $w$ drawing away to more negative values. When we plot the same models as in the first panel of Fig.~\ref{fig:polyw}, but in the $w_{\rm tot}-w'_{\rm tot}$ plane, in Fig.~\ref{fig:polywtot}, we see that the models that were phantom in the effective dark energy lie in the region $w'_{\rm tot}<3w_{\rm tot}(1+w_{\rm tot})$. Furthermore, when $\alpha>2$ they can even lie below what was the null line $w'=-3(1-w^2)$. In a nice analysis of barotropic models, Scherrer \cite{scherrer} noted something similar (his barotropic models are a function of an arbitrary component density not necessarily matter density). For a barotropic perfect fluid the adiabatic sound speed is the physically relevant sound speed for perturbations (but not in the quintessence case, or in a general multicomponent case), and Scherrer's bound of $w'<3w(1+w)$ holds for $c_a^2>0$ (cf.\ Eq.~\ref{eq:ca} here). In general, however, this is not a violation of the bounds of this article and Paper 1 because it occurs only when the adiabatic assumption holds, e.g.\ when viewing the total equation of state, not the properties of a (non-adiabatic) effective dark energy. Suppose, however, we chose to fit the dark energy component itself, rather than full energy density entering the Friedmann equation, by the polytropic form Eq.~(\ref{eq:polyg}). This is somewhat strange to do, since then the effective dark energy contains a matter-like part, in addition to the pure matter density, and the polytrope was designed to be the modified Friedmann equation as a whole. If we do so, though, then the dark energy equation of state phase plane (and not the total equation of state) is represented by the curves in Fig.~\ref{fig:polywtot}. Moreover, the curve with $\alpha=2$, $\beta=1/2$ is the trajectory of the Chaplygin gas. The generalized Chaplygin gas with pressure $p\sim-\rho^{-\alpha_{gcg}}$ corresponds to polytropic dark energy with $\alpha=\alpha_{gcg}+1$, $\beta=1/\alpha$. Whenever $\alpha\beta=1$ (cf.\ \cite{gondolofreese}) we have mocker behavior with $w'=3\alpha w(1+w)$. As stated above, however, taking the dark energy itself to be polytropic means hiding both a cosmological constant (if $\beta=1/\alpha$) and a spurious matter density within the dark sector. \section{Hierarchy parameters} \label{sec:slow} Returning to more general dynamics, we saw in \S\ref{sec:scadyn} and in particular from Fig.~\ref{fig:wwpscale} that we do not have the standard inflation slow roll perturbative expansion parameter. Explicitly, we do not have small $\vp/V$ in the freezing or thawing regions unless $w\lesssim-0.995$; more generally, $-M_P\vp/V\gtrsim 5\sqrt{1+w}$ today. We examine here whether we can substitute a physics based hierarchy of dynamics parameters. \subsection{Slope parameters \label{sec:slope}} The Klein-Gordon equation can be rewritten dimensionlessly as \beq \phi''+(2-q)\phi'=-\vp/H^2\equiv \eta_1, \eeq where as before a prime denotes a derivative with respect to $\ln a$, and $q=-a\ddot a/\dot a^2$ is the deceleration parameter. In terms of the $w-w'$ dynamics equation, \beqa w'&=&-3(1-w^2)-\sqrt{2(1-w^2)} \,\vp/(HV^{1/2})\\ &\equiv& -3(1-w^2)+\sqrt{2(1-w^2)} \,\eta_2. \eeqa So the parameters $\eta_1$, $\eta_2$ are called out by the physics. If they are small, one could solve the equations perturbatively. Note that $\eta_2=\eta_1(V/H^2)^{-1/2}$ so we always have $\eta_2>\eta_1$. We violate the lower bound of the freezing region, $w'<3w(1+w)$, when \beq \eta_2^2<\frac{9}{2}\frac{\epsilon}{2-\epsilon}, \eeq where $\eps=1+w$ is the tilt parameter. The analogous condition such that $w'<0$ is $\eta_2^2<(9/2)\eps(2-\eps)$, and such that field is decelerating ($\ddot\phi<0$) is $\eta_2^2<18\eps/(2-\eps)$. Neither $\eta_1$ nor $\eta_2$ are particularly small unless $\eps=1+w$ is. For example $\eta_2\gtrsim 1.5\sqrt{\eps}$, giving $\eta_2>0.1$ for $w>-0.995$. Even $\eta_1>0.1$ for $w>-0.94$. The hierarchy among the parameters is fixed for the region of interest: $\eps<\eta_1<\eta_2$, so there is no phase space classification in this respect, as there is for large field, small field, and hybrid models in inflation \cite{kinney}. However that hierarchy involved the second derivative of the potential, so it is worth a brief look at that quantity. \subsection{Tracking parameter \label{sec:track}} The tracking parameter is defined to be \beq \Gamma\equiv\frac{V\vpp}{\vp^2}. \eeq This is not generally a small parameter either. Indeed, models whose energy density tracks \cite{tracker} the evolution of the dominant energy component fulfill the conditions \beq \Gamma>1\qquad;\qquad \frac{d\ln(\Gamma-1)}{d\ln a}\ll 1. \eeq Within the class of tracking models (so now a particular subset of scalar field cosmologies), at high redshifts within the matter dominated epoch the field obeys \beq \Gamma=1-\frac{w}{2(1+w)}. \label{eq:gammaw} \eeq (Note the equation of state deviation $1+w$ at high redshift may not be small.) In \S\ref{sec:brane} we found that $H^\alpha$ models (including braneworlds) follow freezing trajectories. This is not surprising because they are basically trackers. A component starting with constant $w$ at early times is equivalent there to a modification $H^\alpha$ with $\alpha=2(1+w)$. The tracking parameter $\Gamma=(2+\alpha)/(2\alpha)$ and it initially acts like an inverse power law potential $V\sim\phi^{-n}$ with $n=2\alpha/(2-\alpha)$. To relate the dynamics of the time variation to the tracking condition, we invert Eq.~(\ref{eq:gammaw}) to write \beqa w&=&-2(\Gamma-1)/[1+2(\Gamma-1)] \\ w'&=&\frac{dw}{d\ln a}=w(1+w)\frac{d\ln(\Gamma-1)}{d\ln a}. \eeqa However, this is of limited use since we are unlikely to be able to probe $w'$ in the high redshift, $z\gg1$, regime where tracking might hold. Strong acceleration today, with $w\lesssim-0.7$, requires the breakdown of tracking. However, the analogy to $H^\alpha$ models presents an important insight into why the general freezing region is bounded below by $w'=3w(1+w)$. At early times the contribution to the Friedmann expansion equation by the dark energy is small and $\delta H^2\sim a^{-3(1+w)}\sim H^{2(1+w)}$. That is, the effective $\alpha\approx 2(1+w)$. One can generalize this to $\alpha=2\langle 1+w\rangle$ when time variation of the equation of state becomes relevant, where angle brackets denote an averaging over $\ln a$. At late times, the dark energy density dominates the expansion, $\delta H^2\sim H^2$, as it approaches $w=-1$ (freezes). For any epoch we can define an instantaneous value of $\alpha$. Equation~(\ref{eq:wpalpha}) then gives the relation for $w'$. As freezing models approach $w=-1$, Eq.~(\ref{eq:wpalpha}) indicates they should do so along $w'=3w(1+w)$. Furthermore, since the bracketed term is less than one, then this trajectory represents a general lower bound to the freezing region.\footnote{One caveat involves dark energy models that possess an internal cosmological constant, i.e.\ nonzero minimum to the potential, or otherwise act as the sum of two or more components. These cannot be represented as $H^\alpha$ models and the freezing bound does not apply.} \subsection{Dynamics, Mass, and Spatial Inhomogeneities} \label{sec:inhom} Dynamical models must also possess spatial inhomogeneities in the field at some level. The equation for these is given by perturbation of the Klein-Gordon equation (\ref{eq:kg}), \beq \delta\ddot\phi+3H\delta\dot\phi+(k^2+\vpp)\delta\phi=-\dot h\dot\phi/2, \label{eq:pert} \eeq where $k$ is the wavenumber and $h$ is the trace of the metric perturbation \cite{ma}. Just as matter density perturbations are damped on scales below the Jeans length related to the sound speed in the background medium, so the spatial inhomogeneities in the scalar field will be absent on length scales less than that corresponding to the effective mass $\sqrt{\vpp}$. Using eq.\ (\ref{eq:wp}) and $\vpp=\dot \vp/\dot\phi$ we can calculate the critical mass scale (also see \cite{caldwellvpp}), with \beqa \vpp/H^2 &=& (2+3w+2q)\frac{w'}{1+w}+ \frac{1}{4}\left(\frac{w'}{1+w}\right)^2 \nonumber \\ &\,& -\frac{1}{2}\frac{w''}{1+w}+\frac{3}{4}(1-w)(5+3w+2q). \eeqa Note $w'/(1+w)$ and $w''/(1+w)$ are well behaved and generally nonzero as $w\to-1$. This shows that the goal of exploring the temporal and spatial dynamics of dark energy runs into double jeopardy. If the time variation is weak, $|w'/(1+w)|\ll1$, then the effective mass $m=\sqrt{\vpp}\lesssim H$. (The scale $H$ today corresponds to $10^{-33}$ eV; dark energy would be a very light scalar field). This means the Compton wavelength of the scalar field perturbations is larger than the horizon, and so spatial inhomogeneities are also difficult to detect. Thus, for $|w'|<1+w$, i.e.\ between the thawing and freezing regions there is a ``dead zone'' of phase space, where we can detect neither time variation nor spatial inhomogeneity. For appreciable time variation, $|w'|>1+w$, one can have $m>H$ and so the possibility of subhorizon clustering. However for models within the freezing or thawing regions, one is restricted to $m\lesssim 2H$ so this could only occur on the largest scales (largest angles or lowest multipoles). Note that as the field approaches $w=-1$, the mass stays nonzero (except it vanishes along the upper boundary of thawing and along the null line). However, the amplitude of spatial perturbations vanishes as can be seen from Eq.~(\ref{eq:pert}) with $\dot\phi=0$. These properties make scalar field inhomogeneity an extremely begrudging probe of the nature of dark energy, much less friendly than the dynamics. \section{Conclusion} \label{sec.concl} We have deepened and elaborated the understanding of the role that the dark energy dynamics, through the $w-w'$ phase plane, can play in leading our understanding of the nature of dark energy. This includes the foundations of the null line, coasting line, constant pressure line, and phantom line dividing the phase plane into distinct, physical regions. We also elucidate the uppermost and lowermost boundaries of the thawing and freezing regions. The physical structure has been extended beyond canonical scalar fields, including specific instances of modified gravity scenarios such as scalar-tensor, braneworld, and $H^\alpha$ models, and barotropic and polytropic generalizations of the Friedmann equation. We outlined similarities and differences with the scalar field case, showing that many act as freezing fields, and that we should be able to clearly distinguish certain models that do not possess a deSitter future. Mocker models, implementing a unification of dark matter and dark energy, were shown to have difficulties purely from dynamical considerations, in addition to their problems in structure formation. Dark energy is demonstrated to be generically not amenable to a slow roll description -- a major difference from early universe inflation -- as one of its ``Goldilocks'' conundra. This makes the dark energy problem in some sense even more challenging than the early universe. However, it also opens the possibility that if some physical bound can be placed on the flatness of the potential, e.g.\ due to quantum corrections, then this implies a barrier around the cosmological constant $\Lambda$ model. This would offer hope, possibly accessible to next generation experiments, that dark energy could definitely be distinguished from $\Lambda$, if it is not $\Lambda$. That would be exciting! The dynamics of the dark energy, in the form of the equation of state ratio $w$ and its time variation $w'$, provides powerful insight into the new physics behind cosmic acceleration. Spatial inhomogeneities in the dark energy are seen to be much weaker and less forthcoming, unless one entertains direct couplings. While we are not guaranteed to zero in on the physics -- there is a dead zone of minimal dynamics and possibly a ``confusion'' zone near the cosmological constant -- any highly precise and accurate result would be an enormous success in enlightening us on the dark universe. $\,$ \\ \section*{Acknowledgments} I benefited greatly from numerous discussions with my collaborator Robert Caldwell. I also thank Carlo Baccigalupi, Robert Scherrer, and participants in the workshop Cosmological Frontiers in Fundamental Physics at Perimeter Institute (which I thank for hospitality), the workshop Dark Energy from Fundamentals (DarkFun 3, hosted by the SNAP Collaboration), and the LBNL particle theory group. This work has been supported in part by the Director, Office of Science, Department of Energy under grant DE-AC02-05CH11231.
Title: Weak-Lensing Detection at z~1.3: Measurement of the Two Lynx Clusters with Advanced Camera for Surveys
Abstract: (Abridged) We present a HST/ACS weak-lensing study of RX J0849+4452 and RX J0848+4453, the two most distant (at z=1.26 and z=1.27, respectively) clusters yet measured with weak-lensing. The two clusters are separated by ~4' from each other and appear to form a supercluster in the Lynx field. Using our deep ACS F775W and F850LP imaging, we detected weak-lensing signals around both clusters at ~4 sigma levels. The mass distribution indicated by the reconstruction map is in good spatial agreement with the cluster galaxies. From the SIS fitting, we determined that RX J0849+4452 and RX J0848+4453 have similar projected masses of ~2.0x10^14 solar mass and ~2.1x10^14 solar mass, respectively, within a 0.5 Mpc (~60") aperture radius.
https://export.arxiv.org/pdf/astro-ph/0601334
\title{WEAK-LENSING DETECTION AT $z\sim1.3$: MEASUREMENT OF THE TWO LYNX CLUSTERS WITH ADVANCED CAMERA FOR SURVEYS} \author{M.J. JEE\altaffilmark{1}, R.L. WHITE\altaffilmark{2}, H.C. FORD\altaffilmark{1}, G.D. ILLINGWORTH\altaffilmark{3} J.P. BLAKESLEE\altaffilmark{4}, B. HOLDEN\altaffilmark{3}, AND S. MEI\altaffilmark{1}} \altaffiltext{1}{Department of Physics and Astronomy, Johns Hopkins University, 3400 North Charles Street, Baltimore, MD 21218.} \altaffiltext{2}{Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218.} \altaffiltext{3}{University of California Observatories/Lick Observatory, University of California, Santa Cruz, CA 95064.} \altaffiltext{4}{Department of Physics and Astronomy, Washington State University, WA 99164.} \keywords{gravitational lensing --- dark matter --- cosmology: observations --- X-rays: galaxies: clusters --- galaxies: clusters: individual (\objectname{RX J0849+4452},~\objectname{RX J0848+4453}) --- galaxies: high-redshift} \section{INTRODUCTION} It has become clear that massive clusters are not extremely rare at high redshifts ($z>0.8$) and the presence of these large collapsed structures when the age of the Universe is less than half its present value is no longer in conflict with our current understanding of the structure formation, especially in a $\Lambda$-dominated flat cosmology. Pursuit of galaxy clusters to higher and higher redshift is important in the extension of the evolutionary sequences to earlier epochs, when the effect of the different cosmological frameworks becomes more discriminating. A great deal of observational efforts have been made in the last decade in enlarging the sample of high-redshift clusters. X-ray surveys have provided an efficient method of cluster identification and probe of cluster properties because a hot intracluster medium (ICM) within the cluster generates strong diffuse X-ray emission and is believed to be in quasi-equilibrium with gravity. However, it is questionable how well the clusters selected by their X-ray excess can provide the unbiased representation of the typical large scale structure at the cluster redshift. If the degree of the virialization decreases significantly with redshift and is strongly correlated with X-ray temperature, the cosmological dimming $\sim (1+z)^{-4}$ can bias our selection progressively towards higher and higher mass, relaxed structures. Among other important approaches to detect high-redshift clusters is a red-cluster-sequence (RCS) survey using the distinctive spectral feature in cluster ellipticals. This so-called 4000\AA~break feature is well-captured by a careful combination of two passbands, and Gladders \& Yee (2005) recently reported 67 candidate clusters at a photometric redshift of $0.9 < z < 1.4$ from the $\sim10$\% subregion of the total $\sim100~\mbox{deg}^2$ RCS survey field. A related method but giving a higher contrast of cluster members with respect to the background sources is to use deep near-infrared (NIR) imaging (e.g., Stanford et al. 1997) for the selection of cluster candidates. High-redshift clusters identified in these color selection methods are expected to serve as less biased samples encompassing the lower mass regime at high redshifts. In the current paper, we study two $z\sim1.3$ clusters, namely RX J0849+4452 and RX J0848+4453 (hereafter Lynx-E and Lynx-W for brevity), using the deep F775W and F850LP (hereafter $i_{775}$ and $z_{850}$, respectively) images obtained with the Advanced Camera for Surveys (ACS) on the $Hubble$ $Space$ $Telescope$ ($HST$). Interestingly, although these two clusters are separated by only $\sim4\arcmin$ from each other, they were discovered by different methods. Standford et al. (1997) discovered Lynx-W in a NIR survey as an overdense region of the $J-K > 1.9$ galaxies and spectroscopically confirmed 8 cluster members. They also analyzed the archival ROSAT-PSPC observation of the region and found diffuse X-ray emission near the confirmed cluster galaxies. However, they could not rule out the possibility that the X-ray flux might be coming from the foreground point sources because of the PSPC PSF is too broad to identify such objects. The subsequent study of the field using the $Chandra$ observations showed that, although the previous ROSAT-PSPC observation is severely contaminated by the X-ray point sources adjacent to the cluster, the cluster is still responsible for some diffuse X-ray emission. Both the X-ray temperature and luminosity of the cluster appear to be low ($T_X\sim1.6$~keV and $L_{bol}\sim0.69\times10^{44} ~ \mbox{ergs}~\mbox{s}^{-1}$; Stanford et al. 2001). Lynx-E was, on the other hand, first discovered in the ROSAT Deep Cluster Survey (RDCS) as a cluster candidate and follow-up near-infrared imaging showed an excess of red ($1.8<J-K<2.1$) galaxies around the peak of the X-ray emission (Rosati et al. 1999). They also showed that five galaxies around the X-ray centroid have redshifts that are consistent with the cluster redshift at $z=1.26$ using the Keck spectroscopic observations. From the $Chandra$ data analysis, Stanford et al. (2001) determined the cluster temperature and luminosity to be $T_X=5.8_{-1.7}^{+2.8}$ keV and $L_{bol}=3.3_{-0.5}^{+0.9}\times10^{44} \mbox{ergs}~\mbox{s}^{-1}$, respectively. The rather large difference in the X-ray properties of these two clusters may be viewed as representing the characteristics of the sample obtained from different survey methods. Lynx-E, the X-ray selected cluster, has much higher X-ray temperature and luminosity than Lynx-W, the NIR-selected cluster. If the stronger X-ray emission means higher dynamical maturity, the more compact distribution of the Lynx-E galaxies provides an alternate support of this hypothesis. For dynamically relaxed systems the observed X-ray properties can be easily translated into the mass properties under the assumption of hydrostatic equilibrium. However, as we probe into the higher and higher-redshift regime, it is natural to expect that there will be more frequent occasions when the equilibrium assumption loses its validity in deriving the mass properties of the system. In addition, at $z>1$ we expect to have a growing list of low-mass clusters that are also X-ray dark because of the evasively low-temperature, as well as the substantial cosmological dimming. Therefore, it is plausible to suspect that these two $z\sim1.3$ clusters (especially Lynx-W, the poorer X-ray system) might lie on a border where the X-ray observations alone start to become insufficient to infer the mass properties. Weak-lensing provides an alternative approach to deriving the mass of a gravitationally bound system without relying on assumptions about the dynamical state. This can help us to probe the properties of the high-redshift clusters in lower mass regimes, where the X-ray measurements alone may not provide useful physical quantities. In our particular case, weak-lensing is an important tool to test how the masses of the two Lynx clusters at $z\sim1.3$ compare with their X-ray measurements. Especially for Lynx-W, weak-lensing seems to be the unique route for probing the cluster mass, considering the poor and amorphous X-ray emission. Another interesting question is whether the low X-ray temperature of Lynx-W arises simply from a low mass or from a yet poor thermalization of the ICM. However, the detection of weak-lensing signal at $z\sim1.3$ is difficult and much more so if the lens is not very massive. In our previous investigation of the two $z\sim0.83$ high-redshift clusters (Jee et al. 2005a, hereafter Paper I; Jee et al. 2005b, hereafter Paper II), we were able to detect clear lensing signals. They revealed the complicated dark matter substructure of the clusters in great detail. The effective source plane (defined by the effective mean redshift of the background galaxies) in Paper I and II is at $z_{eff}\sim1.3$, corresponding to the redshift of the lenses targeted in the current paper! Therefore, the number density of background galaxies decreases substantially compared to our $z\sim0.8$ studies and, in addition, the higher fraction of non-background population in our source sample inevitably dilutes the resulting lensing signal quite severely. Furthermore, the accurate removal of instrumental artifacts becomes more critical as stronger signals come from more distant, and thus fainter and smaller galaxies. They are more severely affected by the point-spread-function (PSF). Nevertheless, our analyses of RDCS 1252-2927 at $z=1.24$ (Lombardi et al. 2005; Jee et al. in preparation) demonstrate that weak-lensing can still be applied to clusters even at these redshifts and reveals the cluster mass distribution with high significance. Returning to the X-ray properties, the low-energy quantum efficiency (QE) degradation of the $Chandra$ instrument can cause noticeable biases in cluster temperature measurements. Although there have been many suggestions regarding this issue, it was not until recently that a convergent prescription to remedy the situation has become available from the $Chandra$ X-ray Center\footnote{see http://cxc.harvard.edu/ciao3.0/threads/apply\_acisabs/ or http://cxc.harvard.edu/ciao3.2/releasenotes/}. Because we suspect that the previous X-ray analyses of the Lynx clusters suffered from the relatively insufficient understanding of this problem, we have also re-analyzed the archival $Chandra$ data to enable a fairer comparison between the weak-lensing and X-ray measurements. Throughout the paper, we assume a $\Lambda$CDM cosmology favored by the Wilkinson Microwave Anisotropy Probe (WMAP), where $\Omega_M$, $\Omega_{\Lambda}$, and $H_0$ are 0.27, 0.73, and 71 $\mbox{km}~\mbox{s}^{-1}~\mbox{Mpc}^{-1}$, respectively. All the quoted uncertainties are at the 1 $\sigma$ ($\sim68$\%) level. \section{OBSERVATIONS} \subsection{ACS Observation \label{subsection_acsobservation}} Deep ACS/WFC imaging of the Lynx clusters were carried out as part of ACS Guaranteed Time Observation (GTO) during 2004 March in three contiguous pointings, which cover a strip of $\sim9\arcmin\times3\arcmin$ region. A slight overlap ($\sim30\arcsec$) was made between the pointings and the strip is oriented in such a way that the two cluster centers are approximately located near the overlap region. Each pointing was observed in $i_{775}$ and $z_{850}$ passbands with 3 and 5 orbits of integration, respectively. We used the ACS GTO pipeline (``APSIS"; Blakeslee et al. 2003) to remove cosmic rays, correct geometric distortion via drizzle algorithm (Fruchter and Hook 2002), and register different exposures. Apsis meets all the requirements of weak lensing analysis (Paper I and II), offering a precise ($\sim0.015$ pixels) image registration via the ``match'' program (Richmond 2002) after correcting for geometric distortion (Meurer et al. 2003). In Figure~\ref{fig_lynx_illustration} we present the pseudo-color image of the entire ACS field with the blow-ups of the two Lynx clusters. Lynx-E is well-portrayed by the somewhat compact distribution of the cluster red sequence around the brightest cluster galaxies (BCGs). It appears that the cluster has a strongly lensed blue giant arc $\sim4.5\arcsec$ south of the BCGs. The spectroscopic redshift of this arc candidate has not yet been determined. The red sequence of Lynx-W looks somewhat scattered and there seem to be no distinct BCGs characterizing the cluster center though the excess of the early-type galaxies in the region clearly defines the cluster locus. The detection image was created by combining the two passband images using inverse variance weighting. Objects are detected through the SExtractor program (Bertin \& Arnouts 1996) by searching for at least five connected pixels above 1.5 times the sky rms. The field contains several bright stars whose diffraction spikes not only induce a false detection, but also contaminate the neighboring objects. We manually selected and removed these objects. The catalog contains a total of 8737 galaxies. \subsection{Chandra Observation} We retrieved the $Chandra$ observation of the Lynx field from the Chandra X-ray Center. The field was observed with the Advanced CCD Imaging Spectrometer I-array (ACIS-I) in the faint mode at a focal temperature of -120 K. The observation consists of two exposures: $\sim65$ ks and $\sim125$ ks integrations on 2000 May 3 and 4, respectively. The raw X-ray events were processed with the $Chandra$ Interactive Analysis of Observations (CIAO) software version 3.2 and the Calibration Database (CALDB) version 3.1, which provide the correction for time-dependent gain variation and the low-energy quantum efficiency degradation without requiring any external guidance. We identified and flagged hot pixels and afterglow events using the $acis\_build\_badpix$, $acis\_classify\_hotpix$ and $acis\_find\_hotpix$ scripts while selecting only the standard $ASCA$ events (0,2,3,4, and 6). Figure~\ref{fig_xrayoverimage} shows the adaptively smoothed $Chandra$ X-ray contours of the Lynx field overlaid on the ACS image. This adaptive smoothing is performed using the CIAO CSMOOTH program with a minimum significance of 3 $\sigma$ and the contours are spaced in square-root scale. Because of the low counts from the two high-redshift clusters, the 3 $\sigma$ significance condition can only be met with rather large smoothing kernels. Therefore, the round appearance of the contours should not be misinterpreted as indicating the relaxed status of the systems. When the contours are reproduced with a smaller, constant kernel smoothing, Lynx-W looks much more irregular than Lynx-E. The X-ray centroids of Lynx-E and W are in good spatial agreement with those of cluster optical lights. The foreground cluster RXJ 0849+4456 (Holden et al. 2001) at z=0.57 appears to be also strong in X-ray emission, but is located outside the ACS pointings ($\sim5\arcmin$ and $\sim3\arcmin$ apart from Lynx-E and W, respectively). The multi-wavelength analysis of this cluster is presented by Holden et al. (2001) and they found that the cluster can be further resolved into two groups at z=0.57 and 0.54. Our subsequent X-ray analyses are confined to the two high-redshift clusters at $\bar{z}=1.265$ present within the current ACS pointings. \section{ACS DATA ANALYSIS} As in our previous investigations (Paper I and II), we measure galaxy shapes and model the point-spread-function of the observation using shapelets (Bernstein \& Jarvis 2002; Refregier 2003). Readers are referred to Paper I and II for detailed description of the ellipticity measurements. \subsection{Cluster Luminosity} Our current spectroscopic catalog of the ACS Lynx field (B. Holden et al. in preparation) contains 150 objects and 32 of them belong to either of the two high-redshift clusters ($1.24<z<1.28$);12 galaxies are at $z>1.31$ and the rest of them (106 objects) are foreground objects. We supplemented the cluster member galaxy catalog with the cluster red sequence (Mei et al. 2005) using $i_{775}-z_{850}$ colors. In order to minimize the systematics from internal gradients and the different PSF sizes (the PSF of $z_{850}$ is $\sim10$\% broader than that of $i_{775}$), the galaxies are deconvolved with the CLEAN (H\"{o}gbom et al. 1974) algorithm. After an effective radius $R_e$ is determined for each galaxy, we measured the object colors within a circular aperture defined by $R_e$. When the estimated $R_e$ was less than three pixels, we used a three pixel aperture instead (the median $R_e$ is $\sim5$~pixels). At $z\sim1.265$, the 4000\AA~break is shifted slightly blue-ward of the effective wavelength of the $z_{850}$ filter. Therefore, this filter combination is less than ideal, but the red sequence is still visible down to $z_{850}\sim24$ in the $i_{775}-z_{850}$ versus $z_{850}$ plot (Figure~\ref{fig_cm}). We visually examined each candidate and discarded the objects that do not seem to have early-type morphology, or whose redshifts (if known) are inconsistent with the cluster redshifts. The final cluster member catalog contains 68 objects. The rest-frame $B$ band at the cluster redshift is approximately redshifted to the ACS $z_{850}$ band and we derive the following photometric transformation from the synthetic photometry with the Spectral Energy Distribution (SED) templates of Kinney et al. (1996). \begin{equation} B_{rest} = z_{850} - (0.70 \pm 0.02) (i_{775}-z_{850}) + (1.08 \pm 0.01) - DM \label{photran}, \end{equation} \noindent where DM is the distance modulus of 44.75 at $\bar{z}=1.265$. From the above selection of the cluster galaxies, we estimate that Lynx-E and W encloses $L_B\sim1.5\times10^{12}$ and $\sim0.8\times10^{12} L_{B\sun}$, respectively within 0.5 Mpc ($\sim60\arcsec$) radius. Of course, these values correspond to the lower limits because we neglected the contribution from the blue galaxies (except for the several spectroscopically confirmed ones), as well as the less luminous population ($z_{850}>24$). However, we do not attempt to determine the correction factors in the current paper because the number of galaxies in both of our spectroscopic and red sequence samples is insufficient to support our statistical derivation. \subsection{PSF Correction} ACS/WFC has a time- and position-dependent PSF (Paper I) and the ability to properly model the PSF pattern in the observed cluster field is critical in subsequent galaxy ellipticity analysis. In paper I and II, we demonstrated that the PSF of WFC sampled from the 47 Tucanae field can be used to describe the PSF pattern of the cluster images where only a limited number of stars are available, but can be used as a diagnosis of the model accuracy. We selected the stars in the Lynx field via a typical magnitude versus half-light radius plot (Figure~\ref{fig_starselect}). Figure~\ref{fig_starfield}a show the WFC PSF pattern in the $i_{775}$ image of the Lynx field, which is similar to the ones in our previous cluster weak lensing studies. The PSFs are elongated in the lower-left to upper-right direction. An analogous pattern is also observed in the $z_{850}$ band. However, the wings of the $z_{850}$ are stretched approximately parallel to the row of the CCD (telescope V2 axis) and the feature becomes observable when the wings of the PSFs are more heavily weighed (Heymans et al. 2005). In our calibration of the ACS (Sirianni et al. 2005), we also observed an opposite pattern (i.e., with an ellipticity nearly perpendicular to Figure~\ref{fig_starfield}a), it seems that this PSF pattern is more frequently observed, at least in our GTO surveys of $\sim15$ clusters. Because the focus offsets of different HST visits are likely to vary, one may desire to find the closest PSF template for every individual exposure and perform PSF corrections one by one. However, we find that in our GTO cluster observations the PSF patterns in different exposures do not vary considerably. Therefore, we chose a single PSF template for each filter and created a PSF map for the entire $3\times 1$ mosaic image by placing the template PSFs on each pointing. In order to minimize the model-data discrepancy due to the slight focus variation, we fine-tuned our model for each exposure by shearing the PSF by an amount $\delta \eta$, which can be expressed in shapelet notation as \begin{equation} b_{pq}^{\prime} = \mbox{\bf{S}}_{\delta \eta} b_{pq}, \end{equation} \noindent where $b_{pq}$ is the shapelet componet of the PSF and the evaluation of matrix elements of the shear operator $\mbox{\bf{S}} _ {\delta\eta}$ can be found in Bernstein \& Jarvis (2002). Figure~\ref{fig_starfield}b displays the residual ellipticities of the same stars in the $i_{775}$ when the PSF is circularized with rounding kernels (Fischer \& Tyson 1997; Kaiser 2000; Bernstein \& Jarvis 2002). The dramatic reduction of the PSF anisotropy is also distinct when the ellipticity components ($e_+$ and $e_{\times}$) before and after the corrections are compared (Figure~\ref{fig_star_anisotropy}). This rounding kernel test verifies that our PSF models describe the PSF pattern of the cluster observation very precisely. Although one can continue with this rounding kernel method and make a subsequent measurement of the galaxy shape in this ``rounded'' images (e.g., Fischer \& Tyson 1997), we prefer to remove the PSF effect through straightforward deconvolution in $shapelets$ because the latter gives more satisfactory results for very faint galaxies (Paper I; Hirata \& Seljak 2003). Besides, the $z_{850}$ PSF is rather complicated because of the ellipticity variation between core and wing metioned above, and this PSF effect can be more efficiently corrected by the deconvolution. \subsection{Mass Reconstruction \label{section_mass_reconstruction}} In order to maximize the weak-lensing signal, it is important to select the source population in such a way that the source sample contains the minimal contamination from cluster and foreground galaxies. Because only two passband images of the Lynx field are available, direct determination of reliable photometric redshift for an individual galaxy is impossible. Therefore, we chose to select the background galaxies based on their ($i_{755}-z_{850}$) colors and $z_{850}$ magnitudes. The redshift distribution of this sample can be indirectly inferred when we apply the same selection criteria to other deep multi-band HST observations such as the Ultra Deep Field (UDF; Beckwidth et al. 2003) project, for which reliable photometric redshift information is obtainable down to the limiting magnitude of our cluster observation (D. Coe et al., in preparation). We selected the $24<z_{850}<28.5$ galaxies whose $i_{775}-z_{850}$ colors are bluer than those of the cluster redsequence ($i_{775}-z_{850}\lesssim0.7)$ as ``optimal'' background population by examining the resulting tangential shears around the two $\bar{z}=1.265$ clusters. This selection yields a total of 6742 galaxies ($\sim204 \mbox{arcmin}^{-2}$). Assuming that the cosmic variance between the Lynx and UDF is not large, we estimate that approximately 60 per cent of the selection is behind the Lynx clusters. Our final ellipticity catalog was created by combining the $i_{775}$ and $z_{850}$ bandpass ellipticities. Of course, there is a subtlety in this procedure because an object can have intrinsically different shapes and thus ellipticities in different passbands. We adopted the methodology presented by Bernstein and Jarvis (2002) to optimally combine the galaxy ellipticities. In our previous weak-lensing analyses (Paper I and II), we found that this scheme indeed reduced the mass reconstruction scatters compared to the case when only single passband images were used; the improvement increases as fainter galaxies are included. As a consistency check, we compared the shapes and lensing signals from the two passband images, and confirmed that the results are statistically consistent. We show the distortion and mass reconstruction of the Lynx field from this combined shape catalog in Figure~\ref{fig_whisker}. Although the systematic alignments of source galaxies around the cluster centers are subtle in the whisker plot (left panel), the resulting mass reconstruction (right panel) clearly shows the dark matter concentration associated with the cluster galaxies. The mass map is generated using the maximum likelihood algorithm and is smoothed with a FWHM$\sim40\arcsec$ Gaussian kernel. We verify that other methods (e.g., Seitz \& Schneider 1995; Lombardi \& Bertin 1999) also produce virtually identical results. The two mass clumps are in good spatial agreement with both the cluster light and X-ray emission. Within a radius of $1\arcmin$, both clumps are found to be significant, above the $4 \sigma$ level (determined from bootstrap resampling). Figure~\ref{fig_high_resolution} shows the high-resolution (smoothed with a FWHM$\sim20\arcsec$ kernel) version of the mass maps overlaid on the ACS images. The clump associated with Lynx-E is offset $\sim10\arcsec$ from the BCGs and the Lynx-W clump seems to lie on the western edge of the cluster galaxy distribution. In Paper I and II, we have reported significant mass-galaxy offsets for two clusters at $z\sim0.83$ and discussed the possibility that those offsets may signal the merging substructures. Although it is tempting to interpret the mass-galaxy offsets in the current study as also implying the similar merging of the two Lynx clusters, our investigation of the mass centroid distribution using the bootstrap resampling shows that the significance is only marginal (i.e., the $r\sim10\arcsec$ circle roughly encloses $\sim70$\% of the centroid distribution). It is encouraging to observe that the foreground cluster at $z\sim0.54$ affects the distortion of source galaxies and reveals itself in the weak lensing mass reconstruction (Figure~\ref{fig_whisker}) though most of its galaxies are outside our ACS field (see Figure~\ref{fig_xrayoverimage} for the location of the X-ray emission from the foreground cluster). As shown by this foreground cluster and its manifestation in the mass map, light coming from background galaxies is perturbed by all the objects lying in their paths to the observer. Considering the high-redshifts ($\bar{z}=1.265$) of the Lynx clusters, the likelihood of such interlopers is high. In addition, if the masses of the two high-redshift clusters are not very large, even a moderately massive foreground object can generate a similar lensing signal because it has higher lensing efficiency for a fixed source plane (unless the source plane is located at substantially higher than $z\sim1.3$). In an attempt to separate this lower-redshift contribution from our weak lensing mass map presented in Figure~\ref{fig_whisker}b, we created an alternate source sample by selecting the brighter ($22<z_{850}<25$) galaxies. This time we did not exclude the galaxies whose $i_{775}-z_{850}$ colors correspond to that of the cluster red-sequence because they also serve as well-defined source plane at $z\sim1.3$ and their shapes should be perturbed by any lower-redshift mass clumps. We present this second version of the mass reconstruction in Figure~\ref{fig_mass_fore}. It is remarkable to observe that in this version the two high-redshift clusters disappear whereas many of the assumed foreground features (including the cluster at $z=0.54$) still remain. The comparison of this second mass reconstruction with the previous result also indicates that some of the foreground mass clumps might affect the shape of the contours of the high-redshift clusters at large radii; the mass clump of Lynx-E seems to have a neighboring foreground clump at its southwestern edge, and the southern edge of the Lynx-W clump also slightly touches the foreground structure (Figure~\ref{fig_high_resolution}) (However, far fewer galaxies were used for this second version of mass reconstruction and thus the position of these structures have much less significance). This apparent substructure in projection may bias our measurements of the total mass. We discuss this issue in \textsection\ref{section_mass_estimate}. \subsection{Redshift Distribution of Source Galaxies} As detailed in Paper I, the redshift distribution of the source galaxies of the Lynx field was inferred from the photometric redshift catalog of the UDF. We also used the two photometric catalogs created from the Great Observatories Origins Deep Survey (GOODS; Giavalisco et al. 2004) and the degraded UDF in order to estimate the contamination of the cluster members in the source sample for $z_{850}<26$ and $z_{850}>26$, respectively. Figure~\ref{fig_zdist} shows the magnitude distribution of the source galaxies (top panel) with the estimated mean redshift (bottom panel) for each magnitude bin. It appears that the number density excess due to the cluster galaxy contamination is not significant throughout the entire magnitude range. However, we must remember that the sample contains substantial contamination of foreground galaxies, which dilute the lensing signal. We measure the mean redshift in terms of the following: \begin{equation} \beta_{l} = \left < \mbox{max} ( 0, \frac{D_{ls}} {D_s}) \right > \label{eqn_beta}, \end{equation} \noindent where $D_s$, $D_l$, and $D_{ls}$ are the angular diameter distance from the observer to the source, from the observer to the lens and from the lens to the source, respectively. We obtain $<\beta>=0.155$ for the entire source galaxies. The value corresponds to a single source plane at $z_{eff}\simeq1.635$ and the critical surface mass density ($\Sigma_c = c^2 (4 \pi G D_l \beta)^{-1}$) has the physical unit of $\sim6180 M_{\sun}/\mbox{pc}^2$ at the redshift of the lens $\bar{z}=1.265$. \subsection{Weak-lensing Mass Estimation \label{section_mass_estimate}} A first guess of the mass can be obtained by fitting the SIS model to the observed tangential shears around the clusters. We chose the origin of the tangential shears as the centroids of the mass clumps in Figure~\ref{fig_whisker}. The neighboring foreground structures at $z\simeq0.54$ as well as the proximity of the field boundary restrict us to the use of the tangential shears at radii no greater than $\sim80\arcsec$. In addition, we discarded the measurements at $r<30\arcsec$ in order to minimize the possible substructure artifact and the contamination of the lensing signal from the cluster members. Although this precaution leaves us with only a small fraction of the total measurements, the lensing signal is clearly detected for both clusters at the $\sim3\sigma$ level in the tangential shear plots (Figure~\ref{fig_tan_shear}). It is plausible that the severly decreased shears at $r<30\arcsec$ for Lynx-E might be in part caused by the aforementioned contamination from the cluster members. We verified that the lensing signal disappeared when the background galaxies were rotated by 45$\degr$ (null test). Note that the uncertainties in Figure~\ref{fig_tan_shear} reflect only the statistical errors set by the finite number of background galaxies. In Paper II, we demonstrated that the large scale structures lying in front of and behind the high-redshift cluster MS 1054-0321 ($z\simeq0.83$) were dominant source of errors in the mass determination, responsible for approximately 15\% of the total cluster mass. This fractional uncertainty increases substantially with cluster redshifts because the lensing by the foreground cosmic structures become more efficient than the lensing by clusters whose redshifts approach those of source galaxies. However, for the current clusters, we expect that the large statistical errors still overwhelm the cosmic shear effects. When we repeat the analysis of Paper II for the current clusters, we estimate that the uncertainties of the Einstein Radius for the SIS fit marginally increases from $\sigma_{er}=0\arcsec.75$ and $0\arcsec.77$ to $\sigma_{er}=0\arcsec.81$ and $0\arcsec.83$ for Lynx-E and W, respectively. The Einstein radius of $\theta_E=2\arcsec.45\pm0\arcsec.81$ (with respect to the effective source plane at $z_{eff}\simeq1.635$) for Lynx-E corresponds to a mass of $M(r)=(4.0\pm1.3)\times10^{14} (r/\mbox{Mpc})~M_{\sun}$ and a velocity dispersion of $740_{-134}^{+113}\mbox{km}\mbox{s}^{-1}$. Similar values of $M(r)=(4.2\pm1.4)\times10^{14} (r/\mbox{Mpc})~M_{\sun}$ and $\sigma_{SIS}=762_{-133}^{+113}\mbox{km}\mbox{s}^{-1}$ are obtained for Lynx-W as implied by its comparable Einstein Radius $\theta_E=2\arcsec.60\pm0\arcsec.83$. As mentioned in \textsection\ref{subsection_acsobservation}, we note that there is a strongly lensed arc candidate at $r\simeq4.5 \arcsec$ for Lynx-E, which can provide a useful consistency check. In general, Einstein radii depend on source redshifts, and the relation steepens if a lens is at a high redshift. If the Einstein radius of the arc is assumed to be $\theta_E=4.5\arcsec$, this implies that the redshift of the object should lie at $1.8<z<3.2$ in our adopted cosmology (the uncertainty reflects only the errors of the Einstein radius from the SIS fit result). Because we have only $i_{775}$ and $z_{850}$ band images, the photometric redshift estimation of this arc candidate is unstable. Nevertheless, if we use the HDFN prior and truncate it below $z=1.2$, the color ($i_{775}-z_{850}=0.098$) of the object is consistent with the SED of the starburst galaxy at $1.7<z<3.7$. Alternatively, we can also estimate the cluster mass based on the two parameter-free methods, namely the aperture mass densitometry and the rescaled mass reconstruction. Although these two parameter-free approaches need some feedbacks from the above SIS fitting result to lift the mass-sheet degeneracy, in general they provide more robust methodology. They are less affected by the cluster substructure or the deviation from the assumed radial profile. However, one drawback of this approach is that the measurement is more severely influenced by the cosmic shear effect than in the case of the SIS fitting because the aperture mass densitometry uses less amount of information (i.e., decreased tangential shears in outer range). With the $r=80-90\arcsec$ region as a control annulus for both clusters, we computed the cluster mass profiles from these two parameter-free methods (Figure~\ref{fig_mass_summary}); from the SIS fit results, we determine the mean mass density in the annulus to be $\bar{\kappa}=0.014\pm0.004$ and $0.015\pm0.005$ for Lynx-E and W, respectively. As observed in Paper I and II, the mass estimation obtained from the rescaled mass reconstruction (dotted) is in good agreement with the aperture mass densitometry (open circle). We also note that both methods gives masses consistent with the SIS fit results. Because we used the SIS fit results above to lift the mass-sheet degeneracy, it is useful to examine how the result change when an NFW profile is assumed, instead. Unfortunately, the low lensing signal in the limited range does not allow us to constrain the two free parameters of the NFW profile simultaneously; the two parameters trade off with each other without significantly altering the quality of the fit. Freezing the concentration parameter to $c=4$, nevertheless, yields $r_s=180\pm37$ ($187\pm34$) kpc for Lynx-E (W), predicting the mean mass density of $\bar{\kappa}=0.015\pm0.020$ ($0.016\pm0.021$) in the control annulus. Different choices for the concentration parameter $c$ do not change these results substantially (for instance, the choice of $c=6$ gives $\bar{\kappa}\simeq0.012$ for Lynx-E). In \textsection\ref{section_mass_reconstruction} we demonstrated that both clusters might have neighboring foreground mass clumps in projection. Therefore, it is worthwhile to assess how much these foreground structures affect our mass estimation. Because the redshift information of the foreground masses are not available, we cannot subtract their contribution directly from our mass map. Instead, we attempted to minimize their effects by replacing the mass density of the region that is occupied by the foreground mass clumps with the azimuthal average from the rest. Of course, we do not expect that this scheme yields cluster masses that are completely free from foreground contamination, since the azimuthal averages taken at other regions might be biased. However, this method is still an important test because a significant difference in resulting mass estimation must be detected if the foreground contamination is indeed severe. The mass-sheet lifted mass map is convenient for this type of analysis. We replaced the southwestern region ($\sim220\degr<\theta<\sim260\degr$; the angle is measured from the north axis counterclockwise) of the Lynx-E clump and the southern region ($\sim130\degr<\theta<\sim195\degr$) of the Lynx-W clump with the azimuthal averages taken at different angles. The solid lines in Figure~\ref{fig_mass_summary} represent the mass profiles obtained from this measurement. For both clusters, this new measurements give slightly lower values, but the change is only marginal. We estimate that both Lynx-E and W have a similar mass of $(2.0\pm0.5) \times 10^{14} M_{\sun}$ within 0.5 Mpc ($\sim60\arcsec$) aperture radius from this approach. The uncertainties here are estimated from 5000 bootstrap resampling of the source galaxies and we do not include the cosmic shear effects because it is non-trivial to estimate the effect for this rescaled mass map approach. We adopt the conventional definition of the virial radius, where the enclosed mean density within the sphere becomes 200 times the critical density $\rho_c(z)=3H(z)^2/8\pi G$ at the redshift of the cluster. Although the factor 200 above is most meaningful in the mass-dominated flat universe, we retain this definition so as to enable a consistent comparison with the values of other clusters found in the literature. The assumption of the spherical symmetry (SIS) allows us to estimate $r_{200}\simeq0.75$~Mpc and $M_{200}\simeq2.0\times 10^{14} M_{\sun}$ for both Lynx clusters. These virial properties are much smaller than the clusters at $z\sim0.83$ studied in Paper I and II. We reported that CL 0152-1357 has a virial radius of $r_{200}\sim1.1$~Mpc and a virial mass of $M_{200}\sim4.5\times10^{14}~M_{\sun}$ in Paper I. For MS 1054-0321, Paper II quoted $r_{200}\simeq1.5$~Mpc and $M_{200}\simeq1.1\times10^{15}~M_{\sun}$. If we assume that the two Lynx clusters are approaching each other perpendicular to the line of sight at a free-fall speed, our order-of-magnitude estimation predicts that the two Lynx cluster will merge into a single cluster whose virial mass exceeds $\sim 4.0\times 10^{14} M_{\sun}$ in a time scale of $t\sim2$~Gyrs (or at $z\sim0.8$). \section{$CHANDRA$ X-RAY ANALYSIS} \subsection{Cluster Temperature and Luminosity \label{section_temperature}} The X-ray spectra of Lynx-E and Lynx-W were extracted from the circular regions ($\bar{r}\sim36\arcsec$) positioned at their approximate X-ray centroids after the point sources (Stern et al. 2002) are removed. The redistribution matrix file (RMF) and the area response file (ARF) were created using the CIAO tool version 3.2 with the calibration database (CALDB) version 3.1, which properly accounts for the time-dependent low-energy QE degradation, as well as charge transfer inefficiency (CTI). The photon statistics is somewhat poor mainly because the clusters are at a high-redshift ($\bar{z}=1.265$) and thus the differential surface brightness dimming is severe $\sim (1+z)^4$. Especially, as implied by its low temperature ($T<2$ keV), the Poissonian scatter of the Lynx-W is worse. Therefore, we constructed the spectra for both clusters with a minimum count of 40 per spectral bin. We think that this choice makes the spectral fitting stable without diluting the overall shape of photon distribution too much. Because it is impossible to constrain the iron abundance given the statistics, we fixed the metallicity at 0.36 $Z_{\sun}$. This assumes that both Lynx clusters possess similar metallicity to RDCS 1252.9-2927 at $z=1.24$ (Rosati et al. 2004). However, as noted by Stanford et al. (2001), we observed only minor changes even when different values were tried. The Galactic hydrogen column density was also fixed at $\mbox{n}_H=2.0\times10^{20}\mbox{cm}^{-2}$ (Dickey \& Lockman 1990). We used the $\chi^2$ minimization modified by Gehrels (1986;CHI-GEHRELS), who extended the conventional $\chi^2$ statistics so that it can handle the deviation of the Poissonian from the Gaussian at the low-count limit. Figure~\ref{fig_spec} shows the best-fit MEKAL plasma spectra (Kaastra \& Mewe 1993; Liedahl, Osterheld, \& Goldstein 1995) for both clusters. We obtain $T=3.8_{-0.7}^{+1.3}$~keV for Lynx-E with a reduced $\chi^2$ of 0.79 (19 degrees of freedom). Lynx-W is determined to have $T=1.7_{-0.4}^{+0.7}$~keV with a reduced $\chi^2$ of 1.16 (7 degrees of freedom). The observed fluxes are estimated to be $F (0.4-7\mbox{keV})=1.5_{-0.2}^{+0.3}\times10^{-14}\mbox{ergs~cm}^{-2}~\mbox{s}^{-1}$ and $F (0.4-4\mbox{keV})=7.2_{-0.5}^{+1.4}\times10^{-15}\mbox{ergs}~\mbox{cm}^{-2}~\mbox{s}^{-1}$, which can be transformed into the rest-frame (also, apertured-corrected to $\sim0.5$~Mpc) bolometric (0.01 - 40 keV) luminosity of $L_X=(2.1\pm0.5)\times10^{44}$ and $(1.5\pm0.8)\times10^{44}~\mbox{ergs}~\mbox{s}^{-1}$ for Lynx-E and Lynx-W, respectively (note that the shallow surface brightness profile of Lynx-W requires a rather large aperture correction factor). \subsection{X-ray Surface Brightness Profile and Mass Determination \label{section_sb}} The azimuthally averaged radial profiles were created from the exposure-corrected $Chandra$ image. In Figure~\ref{fig_betafit} we display these radial profiles with the best-fit isothermal beta models for both clusters. As is indicated by their X-ray image and cluster galaxy distribution, Lynx-E has a higher concentration ($\beta=0.71\pm0.12$ and $r_c=13.2\pm3.2$) of the ICM than Lynx-W ($\beta=0.42\pm0.07$ and $r_c=4.9\arcsec\pm2.8\arcsec$). Together with the cluster temperatures determined in \textsection\ref{section_temperature}, these structural parameters can be converted to the cluster mass under the assumption of hydrostatic equilibrium. In general, many authors report cluster masses within a spherical volume rather than a cylindrical volume spanning from the observer to the source plane, which however is the preferred and natural choice in weak-lensing measurements. This different geometry is often a source of confusion and subtlety in mass comparison between both approaches. Therefore, in this paper we present our X-ray mass estimates in a cylindrical volume in order to ensure more straightforward comparison with the weak-lensing result using the following equation (Paper II): \begin{equation} M_{ap}(r)= 1.78 \times 10^{14} \beta \left ( \frac{T}{\mbox{keV}} \right ) \left ( \frac{r}{\mbox{Mpc}} \right ) \frac{r/r_c}{\sqrt{1+(r/r_c)^2}} M_{\sun} \label{eqn_xray_mass_2d} \end{equation} \noindent For Lynx-E we obtain $M(r\leq0.5~\mbox{Mpc})=2.3_{-0.4}^{+0.8}\times10^{14} M_{\sun}$, which is in good agreement with our weak-lensing measurement. On the other hand, the X-ray mass of Lynx-W ($M(r\leq0.5~\mbox{Mpc})=6.3_{-1.5}^{+2.6}\times10^{13} M_{\sun}$) is much lower than the weak lensing estimation. We will discuss a few possible scenarios for this discrepancy in \textsection\ref{summary}. \section{COMPARISON WITH OTHER STUDIES} The first attempt to estimate the mass of Lynx-W was made by Stanford et al. (1997) using the X-ray luminosity from the ROSAT-PSPC observation and the velocity dispersion obtained from the Keck spectroscopy of 8 galaxies. They converted the luminosity $L_X\sim1.5\times10^{44} \mbox{ergs}~\mbox{s}^{-1}$ to $M(r<2.3~ \mbox{Mpc})\sim7.8\times10^{14} M_{\sun}$ assuming $\beta=0.8$. A similar value of $M(r<2.3 \mbox{Mpc})=5.4_{-2.3}^{+3.1}\times10^{14} M_{\sun}$ was estimated from the velocity dispersion of $\sigma=700\pm180 \mbox{km}~\mbox{s}^{-1}$ (note that they adopted $h_{100}=0.65$ and $q_0=0.1$). Although both masses are consistent with each other, their X-ray luminosity measurement seems to have suffered a severe contamination from the neighboring point sources, which are now identified in the $Chandra$ observation. In their presentation of the $Chandra$ analysis, Stanford et al. (2001) did not attempt to estimate the mass of Lynx-W because of the large uncertainty of the temperature measurement, as well as the apparent asymmetry of the X-ray emission. Our predicted velocity dispersion of $\sigma_{SIS}=762_{-133}^{+113}\mbox{km}\mbox{s}^{-1}$ from the SIS fit result is consistent with their most recent determination of the velocity dispersion $\sigma=650\pm170~\mbox{km}~\mbox{s}^{-1}$ from the spectroscopic redshifts of the 9 member galaxies. Lynx-W was also selected as one of the 28 X-ray clusters for the study of the X-ray scaling relation at high redshifts by Ettori et al. (2004). From the re-analysis of the $Chandra$ data, they obtained $\beta=0.97\pm0.43$, $r_c=163\pm70$~kpc, and $T_X=2.9\pm0.8$~keV, which predicts a projected mass of $M(r\leq 0.5~\mbox{Mpc})= 3.0\pm1.5\times10^{14} M_{\sun}$ (eqn.~\ref{eqn_xray_mass_2d}). This mass is consistent with our weak-lensing estimation ($2.0\pm0.5) \times 10^{14} M_{\sun}$, but much higher than the value from our re-analysis of the same $Chandra$ data ($6.3_{-1.5}^{+2.6}\times10^{13} M_{\sun}$). In general, many detailed steps in the $Chandra$ X-ray analysis such as the QE correction, background modeling, flare removal, spectral aperture, etc. affect the final result, and much more if the source is faint. Therefore, it is difficult, if not impossible, to trace the exact causes of the differences. Nevertheless, we note that there is an important difference in the calibration of the low-energy quantum efficiency correction between the results. Ettori et al. (2004) used the ACISABS correction method (Chartas and Getman 2002) to account for the low-energy QE degradation, which is however now officially disapproved by the $Chandra$ $Data$ $Center$. We also demonstrate in Paper II that the use of this ACISABS model causes a difference of $\sim1$~keV in the temperature determination of MS1054-0321. We suspect that the effect should be more important in Lynx-W because of its low temperature and luminosity. Stanford et al. (2001) obtained an X-ray temperature of $5.8_{-1.7}^{+2.8}$ keV for Lynx-E. Combined with their determination of $\beta=0.61\pm0.12$ and $r_c=11\arcsec.14\pm3\arcsec.41$, this gives a projected mass of $M(r\leq0.5\mbox{Mpc})=3.1_{-0.9}^{+2.4} \times 10^{14} M_{\sun}$ (eqn.~\ref{eqn_xray_mass_2d}), which is slightly higher than our X-ray re-analysis of the same $Chandra$ data by $\sim35$\% though the error bars from both results marginally overlap. Vihklinin et al. (2002) included Lynx-E in their sample of the 22 distant clusters to study the evolution the X-ray scaling relation. With the early understanding of the low-energy QE problem of the $Chandra$, they obtained $T_X=4.7\pm1.0$~keV, $r_c=167$~kpc, and $\beta=0.85\pm0.33$. Using the ACISABS correction. Ettori et al. (2004) reported $T_X=5.2_{-1.1}^{+1.6}$~keV, $r_c=128\pm40$~kpc, and $\beta=0.77\pm0.19$. The results from these two papers are statistically consistent with, but slightly higher than our values ($T=3.8_{-0.7}^{+1.3}$keV, $r_c=111\pm27$, and $\beta=0.71\pm0.12$), which predicts the lowest projected mass of $M(r\leq0.5\mbox{Mpc})=2.3_{-0.4}^{+0.8}\times10^{14} M_{\sun}$. As already mentioned in the discussion of the Lynx-W temperature above, we suspect that the difference in temperatures mainly stems from the different correction methods of the low-energy QE degradation. Although our understanding of the $Chandra$ instrument still evolves and this may neccesitate some updates to our results, it is encouraging to note that this X-ray mass is closest to our independent lensing determination of the cluster mass of $M(r\leq0.5\mbox{Mpc})=(2.0\pm0.6)\times10^{14} M_{\sun}$ from the SIS fit result. Our spectroscopic catalog currently provides the redshifts of 11 member galaxies within a $r=80\arcsec$ radius (B. Holden in prep). Based on Tukey's biweight estimator, we obtain a velocity dispersion of $720\pm140 \mbox{km}~\mbox{s}^{-1}$ (without assuming a Gaussian distribution). This direct measurement agrees with the predicted velocity dispersion of $740_{-134}^{+113}\mbox{km}\mbox{s}^{-1}$ from the lensing analysis (\textsection\ref{section_mass_estimate}). In addition, the cluster temperature $T_X=3.8_{-0.7}^{+1.3}$keV with $\beta=0.71$ is translated into $\sigma_v=662_{-64}^{+106} \mbox{km}\mbox{s}^{-1}$ (from $\beta=\mu m_p \sigma_v^2/kT_X$), in good agreement with both results. \section{DISCUSSION AND CONCLUSIONS \label{summary}} We have presented a weak-lensing analysis of the two Lynx clusters at $\bar{z}=1.265$ using the deep ACS $i_{775}$ and $z_{850}$ images. Our mass reconstruction clearly detects the dark matter clumps associated with the two high-redshift clusters and other intervening objects within the ACS field, including the known foreground cluster at $z=0.57$. In order to verify the significance of the cluster detection and to separate the high-redshift signal from the low-redshift contributions, we performed a weak-lensing tomography by selecting an alternate lower-redshift source plane. This second mass reconstruction does not show the mass clumps around the high-redshift clusters, while maintaining most of the other structures seen in the first mass map. This experiment strongly confirms that the weak-lensing signals observed in the first mass reconstruction are real and come from the high-redshift Lynx clusters. Interestingly, both clusters are found to have similar weak-lensing masses of $\sim 2.0\times 10^{14} M_{\sun}$ within 0.5 Mpc ($\sim60\arcsec$) aperture radius despite their discrepant X-ray properties. Our re-analysis of the Chandra archival data with the use of the latest calibration of the low-energy QE degradation shows that Lynx-E and W have temperatures of $T=3.8_{-0.7}^{+1.3}$ and $1.7_{-0.4}^{+0.7}$~keV, respectively. Combined with the X-ray surface brightness profile measurements, the X-ray temperature of Lynx-E gives a mass estimate in good agreement with the weak-lensing result. On the other hand, the X-ray mass of Lynx-W is much smaller than the weak-lensing estimation nearly by a factor of three. According to our experiment in \textsection\ref{section_mass_estimate}, it is unlikely that any foreground contamination or cosmic shear effect in weak-lensing measurement causes this large discrepancy. Apart from a simplistic, but valid possibility that Lynx-W might have a filamentary structure extended along the line of sight, yielding a substantial, projected mass but with yet only low-temperature thermal emission, we can also consider the self-similarity breaking (e.g., Ponman et al. 1999; Tozzi \& Norman 2001; Rosati, Stefano, \& Norman et al. 2002) typically observed for low-temperature X-ray systems. There have been quite a few suggestions that a non-gravitational heating (thus extra entropy) might prevent the ICM from further collapsing at the cluster core. The effect is supposed to be more pronounced in colder systems whose virial temperature is comparable to the temperature created by this non-gravitational heating, leading to shallower gas profiles than those of high-temperature systems (e.g., Balogh et al. 1999; Tozzi \& Norman 2001). Interestingly, our determination of the surface brightness profile of Lynx-W is much shallower ($\beta=0.42\pm0.07$) than that of Lynx-E ($\beta=0.71\pm0.12$) (however, Ettori et al. (2004) obtained $\beta=0.97\pm0.43$ for Lynx-W). The relatively loose distribution of the cluster galaxies in Lynx-W without any apparent BCG defining the cluster center leads us to consider another possibility that the system might be dynamically young and the ICM has not fully thermalized within the potential well. If we imagine that the ICM is not primordial, but has been ejected from the cluster galaxies at some recent epoch, it is plausible to expect that the X-ray temperature of the ICM might yet under-represent the depth of the cluster potential well. Tozzi et al. (2003) investigated the iron abundance in the ICM at $0.3<z<1.3$ and argued that the result was consistent with no evolution of the mean iron abundance out to $z\simeq1.2$. If we assume that, as they suggested, Type Ia SNe are the dominant sources of this iron enrichment and have already injected their metals into the ICM by $z\sim1.2$, a significant fraction of clusters at $z\gtrsim 1.2$ may possess dynamically young ICM. Recently, Nakata et al. (2005) reported with a photometric redshift technique the discovery of seven other cluster candidates around these two Lynx clusters possibly forming a $z\sim1.3$ supercluster. Although further evidence is needed that the individual clumps are dynamically bound, the clear enhancement of the red galaxies consistent with the color at the redshift of the two known Lynx clusters is worthy of our attention. If they are indeed found to be forming groups/clusters at $z\sim1.3$, but missed by X-ray observations because of their low X-ray contrast, the detailed studies of these young high-redshift structures will provide a critical benchmark in testing our understanding of the structure formation as well as the individual galaxy evolution in the context of different environments. Deep two band ($i_{775}$ and $z_{850}$ $HST$/ACS imaging of the five out of the seven group/cluster candidates of Nakata el al. (2005) are scheduled in $HST$ Cycle 14 (Prop. 10574, PI. Mei). Studies similar to the current investigation will not only test whether there exist dark matter clumps around the candidate galaxies, but also quantify the environments for the investigation of the cluster galaxy color/morphology evolution. ACS was developed under NASA contract NAS5-32865, and this research was supported by NASA grant NAG5-7697. We are grateful for an equipment grant from Sun Microsystems, Inc. Some of the data presented herein were obtained at the W.M. Keck Observatory, which is operated as a scientific partnership among the California Institute of Technology, the University of California and the National Aeronautics and Space Administration. The Observatory was made possible by the generous financial support of the W.M. Keck Foundation. The authors wish to recognize and acknowledge the very significant cultural role and reverence that the summit of Mauna Kea has always had within the indigenous Hawaiian community. We are most fortunate to have the opportunity to conduct observations from this mountain. \clearpage
Title: Radial velocity measurements of B stars in the Scorpius-Centaurus association
Abstract: We derive single-epoch radial velocities for a sample of 56 B-type stars members of the subgroups Upper Scorpius, Upper Centaurus Lupus and Lower Centaurus Crux of the nearby Sco-Cen OB association. The radial velocity measurements were obtained by means of high-resolution echelle spectra via analysis of individual lines. The internal accuracy obtained in the measurements is estimated to be typically 2-3 km/s, but depends on the projected rotational velocity of the target. Radial velocity measurements taken for 2-3 epochs for the targets HD120307, HD142990 and HD139365 are variable and confirm that they are spectroscopic binaries, as previously identified in the literature. Spectral lines from two stellar components are resolved in the observed spectra of target stars HD133242, HD133955 and HD143018, identifying them as spectroscopic binaries.
https://export.arxiv.org/pdf/astro-ph/0601643
\title{Radial velocity measurements of B stars in the Scorpius-Centaurus association} \author{E.\,Jilinski\inst{1,2}, S. Daflon\inst{1}, K.\,Cunha \inst{1} and R.\,de la Reza \inst{1} } \offprints{E.\,Jilinski, Observat\'orio Nacional/MCT, Rua Gal. Jose Cristino 77, S\~ao Cristov\~ao, Rio de Janeiro, Brazil. e-mail: jilinski@on.br} \institute{Observat\'orio Nacional/MCT, Rio de Janeiro, Brazil \and Main Astronomical Observatory, Pulkovo, St. Petersburg, Russia} \date{} \authorrunning{E.\,Jilinski et al.} \titlerunning{Radial velocities in the Sco-Cen association} \abstract{We derive single-epoch radial velocities for a sample of 56 B-type stars members of the subgroups Upper Scorpius, Upper Centaurus Lupus and Lower Centaurus Crux of the nearby Sco-Cen OB association. The radial velocity measurements were obtained by means of high-resolution echelle spectra via analysis of individual lines. The internal accuracy obtained in the measurements is estimated to be typically 2-3 km\,s$^{-1}$, but depends on the projected rotational velocity of the target. Radial velocity measurements taken for 2-3 epochs for the targets HD120307, HD142990 and HD139365 are variable and confirm that they are spectroscopic binaries, as previously identified in the literature. Spectral lines from two stellar components are resolved in the observed spectra of target stars HD133242, HD133955 and HD143018, identifying them as spectroscopic binaries. \keywords{stars: early-type - stars: binaries: spectroscopic - stars: kinematics - stars: radial velocities in open clusters and associations: individual: Scorpius-Centaurus association. } } \section{Introduction} The Scorpius-Centaurus association is the nearest association of young OB stars to the Sun. Blaauw (1960, 1964) divided this association into three stellar subgroups: Upper Scorpius (US), Upper Centaurus Lupus (UCL) and Lower Centaurus Crux (LCC). LCC and UCL have roughly similar ages of about $16-20$ Myr, while US is younger with an estimated age of $\sim$ 5 Myr (Mamajek et al.2002; Sartori et al. 2003). This complex OB association of unbound stars is of great interest because, as recently shown, it is related to the origins of nearby moving groups of low mass post-T Tauri stars with ages around 10 Myr: the $\beta$ Pictoris Moving Group, the TW Hydra association, and the $\eta$ and $\epsilon$ Chamaleonis groups (Mamajek et al. 2000; Ortega et al. 2000, 2004; Jilinski et al. 2005). In addition, the Scorpius-Centaurus association also appears to be the source of a large bubble of hot gas in which the Sun is plunged. All these structures are believed to have been possibly triggered by supernova explosions taking place in UCL and LCC during the last ~13 Myr (Ma\'iz-Apell\'aniz 2001). The technique adopted for investigating the origins of the $\beta$ Pictoris Moving Group, for example, consists of tracing back the 3-D stellar orbits of the members of these moving groups until their main first orbits confinement was found, as well as the past mean positions of LCC and UCL. This enabled, investigator not only to determine the dynamical age of this moving group, but also to investigate properties of their birth clouds (Ortega et al. 2002, 2004). It is also possible to find the past positions of the possible supernovae that triggered the formation of these groups by tracing back the orbit of a runaway OB star, which could have been the result of a supernova explosion in LCC or UCL (see, for example, Hoogerwerf et al. 2001 and Vlemmings et al. 2004). While the past evolution of these moving groups of low mass stars appears to be a relatively simple problem (as the dynamical ages are not so old), the dynamical evolution of the older and more numerous subgroups LCC and UCL appears to be more difficult. There is the possibility of the presence of several generations of hot stars during the mainstream of the OB association evolution (Garmany 1994). Substructure in LCC and UCL was found by de Bruijne (1999), based on Hipparcos data. The formation of the younger US subgroup could have been triggered by UCL some 6-8 Myr ago (Preibisch et al. 2001). All these studies require reliable radial velocities in order to calculate space velocities. In this paper we present single-epoch radial velocity (RV) measurements for 56 B-type stars members of LCC, UCL and US subgroups, to contribute to studies of their dynamics so as to unravel their origins. \section{Observations and reduction} A sample of 56 B-type stars from the Scorpius-Centaurus association was observed during observing runs in May 16-20 and July 7, 2002, with the 1.52m telescope equipped with the FEROS echelle spectrograph (Kaufer et al. 2000; resolving power R=48,000, wavelength coverage between 3900 and 9200\AA.) with a CCD detector at the European Southern Observatory (ESO) \footnote {Observations obtained under the ON/ESO agreement}. The target stars were selected from the list in Humphreys \& McElroy (1984) and from the comprehensive study of OB associations based on Hipparcos observations by de Zeeuw et al. (1999). The observed targets are listed in Table~1. From this sample, according to de Zeeuw et al. (1999), 15 targets are confirmed members of the LCC, while 15 stars are members of the UCL and 11 stars are from the US subgroup. For the remaining 15 stars in our sample, membership to any of these subgroups was not certain. The spectra were reduced with the MIDAS reduction package and consisted of the following standard steps: CCD bias correction, flat-fielding, extraction, wavelength calibration, correction of barycentric velocity, as well as spectrum rectification and normalization. The one-dimensional spectra were then treated by tasks in the NOAO/IRAF data package. The signal-to noise ratio obtained in the observed spectra was typically larger than 100 and typical exposure times varied between 300 seconds for the brightest stars (V $\sim$ 3) and 1200 seconds for stars with V $\sim$ 5. Typical spectra are shown in Figure~1 . The top panel corresponds to the target star HD~122980 and the bottom panel to HD~112092. Both stars have sharp lines with projected rotational velocities ($v \sin i$) less than ~40 km\,s$^{-1}$. The spectral region displayed shows identifications of several lines that were used in the RV determinations. The FEROS bench spectrograph and set up have proven to have high spectral stability for RV measurements as concluded from a study of radial-velocity standard stars: a r.m.s. of 21 m\,s$^{-1}$ has been obtained for a data set of 130 individual measurements (Kaufer et al. 2000). \section{Analysis and discussion} The cross-correlation technique, which is used for precise RV determinations in later type stars, when applied to the hotter OB stars can be problematic as early type star spectra show few absorption lines. These lines are in many cases, intrinsically broad (up to a few hundreds km\,s$^{-1}$) due to stellar rotation. In addition, there is also the possibility of line variability affecting their line profiles (Steenbrugge et al. 2003). Therefore, the cross-correlation peak that defines the value of radial velocity can be very broad and contain important sub-structures caused by blending of spectral lines that appear to have different widths. In addition to having high $v \sin i$ values many OB stars are binary and it is not straightforward to apply the cross-correlation method and to identify them as double-lined binaries; in order to obtain the orbital solution a long set of observations is needed. Detailed cross-correlation technique analyses applied to determinations of radial velocities of early-type stars has been presented in a number of recent publications (see, for example, Verschueren et al. 1997; Verschueren et al. 1999a; Griffin et al. 2000). Griffin et al. (2000), in particular, discuss in detail the difficulties in obtaining accurate RV measurements from cross-correlation in early-type stars spectra. In this study, having high-resolution observations covering a large spectral range, radial velocity values for the target stars were obtained from measurements of the positions of individual spectral lines of He I, C II, N II, O II, Mg II, Si II and Si III, relative to their rest wavelengths. Radial-velocity standard stars were not observed. (The adopted linelists can be found in Daflon et al. 2001, 2003.) We inspected and identified all unblended lines visible in the spectral range between 3798 \AA\ (H\,{\sc i}) and 7065 \AA\ (He\,{\sc i}) in each target star: the number of measurable lines varied between 10 and 74, depending on the star spectral type, rotation velocity, possible multiplicity, but also on the signal-to-noise of the obtained spectra. (We note, however, that for the double lined binary HD 133242, it was possible to measure positions only for 4 lines in component A and 6 lines in component B.) Mean radial velocities using all measurable lines (${\rm RV}$) and respective dispersions were calculated for the individual target stars. In Table~1 we assemble our RV results as well as results from the literature. In the two first columns of this table we list the HD numbers of the observed stars with the respective spectral types; in the columns 3 and 4 we list the heliocentric Julian Date (HJD) and the measured radial velocities, plus the number of measured lines in brackets. In the other columns we list results from the literature: columns 5 and 6 list the RV$_{\rm GCRV}$ and associated error or quality, from the General Catalogue of Radial Velocities (GCRV; quality flags A to E, or I for insufficient data); column 7 presents the projected rotational velocity from Brown \& Verschueren (1997) and, when not available in this source, the $v\,sin i$ was taken from the compilation of Glebocki \& Stawikowski (2000); in column 9 we list, when available, literature references where information about duplicity can be found for the stars. The stars are separated according to the different subgroups in the Scorpius-Centaurus association, following the membership probabilities P(m) listed by de Zeeuw et al. (1999). The internal precision of our RV determinations can be represented by the scatter obtained from the RV measurements line-by-line, which is listed in column 4 of Table 1. These are typically smaller than $\sim$2.0 ${\rm km\,s}^{-1}$ for stars with estimated $v\,sin i$ smaller than 100 ${\rm km\,s}^{-1}$. We note, however, that when the target $v\,sin i$ are large, the uncertainties in the derived RVs can be significantly larger due to uncertainties in defining the line center. This can be seen in Figure~2, where we show the obtained line-to-line scatter versus the target projected rotational velocity (as taken from Brown \& Verschueren 1997 and Glebocki \& Stawikowski 2000). In order to evaluate possible systematic effects that line selection could have on the RV results, we selected a homogeneous set of 28 spectral lines of H\,{\sc i}, He\,{\sc i}, Si\,{\sc iii} and Mg\,{\sc ii}, that could be measured in most of the studied spectra, and recalculated the mean radial velocity values for all possible stars. A comparison of the mean radial velocities ${\rm RV}$ with ${\rm RV}_{\rm 28lines}$ (obtained using only the selected 28 lines) indicates that there are non-significant systematic differences between the two determinations ${\rm RV} - \, {\rm RV}_{\rm 28lines}= -1.0 {\rm km\,s}^{-1}$, with $\sigma = 3.1 {\rm km\,s}^{-1}$. \subsection{Membership} Table~1 lists the target stars according to their membership the 3 subgroups as assigned by de Zeeuw et al. (1999). Most of the stars in the Lower Centaurus Crux subgroup are flagged as binaries in the literature, except for HD103079, HD106490 and HD108483. For these 3 stars we measured radial velocities of $19.3 {\rm km\,s}^{-1}$, $15.3 {\rm km\,s}^{-1}$ and $12.8 {\rm km\,s}^{-1}$, respectively, with RV$_{mean}$=15.8 $\pm$3.2 ${\rm km\,s}^{-1}$. This mean value is in general agreement with the mean radial velocity calculated by de Zeeuw et al. (1999) for LCC, which is of $12 {\rm km\,s}^{-1}$. For the subgroup UCL, our sample has 2 non-binary stars (HD121790 and HD128345) and RV$_{mean}$=$9.6 {\rm km\,s}^{-1}$ which is $\sim 5 {\rm km\,s}^{-1}$ higher than the de Zeeuw et al. (1999) mean value of $4.9 {\rm km\,s}^{-1}$. For the Upper Scorpius subgroup all the sample stars have been flagged as binaries in the literature. For the five target stars that had not been identified as members of any of the three subgroups in the Sco-Cen association (listed as "others" in Table~1) and for which we have no information on duplicity, we can attempt to discuss their membership status based on the comparison of the radial velocities measured here and in the literature. We find that the measured radial velocities for HD109026 (RV=4.0 ${\rm km\,s}^{-1}$) and HD110335 (RV= 4.8 ${\rm km\,s}^{-1}$) are consistent with the mean radial velocity for UCL of 4.9 ${\rm km\,s}^{-1}$s. For the target star HD109026 we have RV$_{GCRV}$= 2.5 ${\rm km\,s}^{-1}$, therefore it could be considered initially as having constant RV (within the uncertainties) and possibly a member of the Upper Centaurus Lupus subgroup. For HD110335, we find a larger discrepancy between our measurement (RV=4.8 ${\rm km\,s}^{-1}$) and the RV value in the GCRV (RV$_{GCRV}$=12.5 ${\rm km\,s}^{-1}$). The RVs, however, are marginally consistent given the expected uncertainty brackets that affect the 2 determinations. If this is really the case, HD110335 could be also considered as a possible member of the UCL subgroup. In addition, the target star HD120640 (RV$_{mean}$=-1.8 ${\rm km\,s}^{-1}$ from this study and RV$_{GCRV}$=-4.7 ${\rm km\,s}^{-1}$) can be assumed here to have constant radial velocity. These measurements are consistent with the mean RV value of -4.6 ${\rm km\,s}^{-1}$ listed by de Zeeuw et al. (1999) for this subgroup. The two other stars in our sample of 'others' (HD115846 and HD105937) for which we derived RV= -21.8 ${\rm km\,s}^{-1}$ and RV=22.7 ${\rm km\,s}^{-1}$, respectively, have values in the GCRV of RV$_{GCRV}$=3.0 ${\rm km\,s}^{-1}$ and RV$_{GCRV}$=15.0 ${\rm km\,s}^{-1}$. We found no information in the literature about these stars being confirmed binary stars, but the variation in RV for HD115846 exceeds the expected uncertainties: this target probably has a non-constant RV, which prevents further considerations about it belonging to any of the Sco-Cen subgroups. HD105937 has an RV only marginally constant within the uncertainties, but its mean RV is not compatible with any of the subgroups. \subsection{Duplicity} Results from a search for duplicity information for the targets stars (column 9; Table~1) indicate that a large number of stars in our sample are flagged as binaries in the literature. For most of these targets we have only one single-epoch RV measurement and our results alone cannot be used to infer duplicity. However, the RVs derived in this study can be added to RV databases and contribute to long term studies of their orbits. Only a small number of stars had not been previously flagged as binaries in the different studies in the literature. For this subsample of 10 stars, considered a priori as RV constants, it is possible to compare our RV values with the averaged radial velocities assembled in the GCRV. This comparison is shown in Figure~3. Our RV determinations compare favorably with the RVs from the GCRV with a scatter of the order of the estimated uncertainties. ${\rm RV}_{\rm GCRV} - {\rm RV} = 0.7$ and $\sigma = 4.9 {\rm km\,s}^{-1}$. (This was calculated excluding one discrepant star, HD115846, which could be a binary system.) Taking into account the mean precision of RV determinations from the GCRV as $\pm 5 {\rm km\,s}^{-1}$, the external precision of our RV determinations may be evaluated as approximately $\pm 5 {\rm km\,s}^{-1}$. For those targets with more than one epoch RV measurement in our study, a subsample showed radial velocity variations larger than the expected uncertainties: HD120307, HD142990 and HD139365. Since these had been previously identified as SBs in the literature (Levato et al. 1987 and Batten et al. 1989), our results confirm their duplicity. Four other stars with multiple epoch observations in this study showed a constant RV within the uncertainties: HD116087, HD130807, HD132200 and HD120640. For 3 targets in our sample (HD133242, HD133955 and HD143018) we were able to separate and identify lines of two stellar components, classifying them as double lined spectroscopic binaries. Their combined spectra showing spectral lines from two stars are shown in Fig~4. Two of these stars (HD143018 and HD133955) were previously identified in Batten et al. (1999) as a spectroscopic binaries. \begin{acknowledgements} We thank the anonymous referee for suggestions that significantly improved the paper. E.G.J. thanks FAPERJ and MCT Brazil for financial support. \end{acknowledgements} {} \newpage \begin{table} \begin{center} \begin{tabular}{llcrcccc} \multicolumn{8}{c}{Table 1. Radial velocities of observed Sco-Cen OB stars}\\ \hline \hline \,\,\, HD &\,\,\,\,Sp& HJD &RV\,[N]\,\,\,\,\,\,\,\,& RV$_{\rm GCRV}$\,\,\, & eRV* & V$\sin$ i & Duplicity \\ & & & (km\,s$^{-1}$)\,\,\,\,\,\, & (km\,s$^{-1}$) & (km\,s$^{-1}$) & (km\,s$^{-1}$) & \\ \hline \multicolumn{8}{c}{Lower Centaurus Crux} \\ \,\,\,98718&B5Vn & 52411.0285 & 26.8$\pm$2.7 [19] & 9.4 & C & 340$^a$ & ST \\ 103079 & B4V & 52411.0425 & 19.3$\pm$1.0 [58] & 20.6 & B & 47$^b$ & \\ 105382 & B6IIIe & 52413.0429 & 15.8$\pm$1.1 [50] & 16.4 & B & 75$^b$ & ST \\ 106490 & B2IV & 52413.0617 & 15.3$\pm$1.5 [35] & 22.0 & B & 135$^b$ & \\ 106983 & B2.5V & 52411.0787 & 12.1$\pm$0.8 [50] & 15.8 & A & 65$^b$ & VDH \\ 108257 & B3Vn & 52411.0875 & 19.9$\pm$2.7 [19] & 5.0 & C & 298$^b$ & VDH \\ 108483 & B2V & 52411.1008 & 12.8$\pm$1.4 [29] & 8.0 & C & 169$^b$ & \\ 109668 & B2IV-V & 52411.1127 & $-$0.1$\pm$1.4 [38] & 13.0 & C & 114$^b$ & ST \\ 110879 & B2.5V & 52411.1177 & 56.9$\pm$2.5 [33] & 42.0 & D & 139$^b$ & VDH \\ 110956 & B3V & 52411.1229 & 15.6$\pm$0.7 [69] & 16.4 & B & 26$^b$ & VDH \\ 112091 & B5Vne & 52411.1449 & 15.9$\pm$1.9 [19] & 13.0 & C & 242$^b$ &VDH\\ 112092 & B2IV-V & 52411.1389 & 14.4$\pm$0.6 [72] & 13.9 & A & 34$^b$ & VDH \\ 113703 & B5V & 52415.1623 & 1.2$\pm$1.4 [29] & 6.0 & C & 140$^b$ & VDH,ST \\ 113791 & B1.5V & 52411.1615 & 58.5$\pm$0.8 [72] & 14.3 & C & 15$^b$ & SB8 \\ 116087 & B3V & 52411.2216 & 12.3$\pm$1.6 [21] & 6.0 & C & 233$^b$ & VDH,ST \\ & & 52414.1034 & 9.4$\pm$3.9 [10] & & & & \\ \multicolumn{8}{c}{Upper Centaurus Lupus} \\ 120307 & B2IV & 52413.1160 & 25.4$\pm$0.9 [54] & 9.1 & & 65$^b$ & SB8,L87\\ & & 52413.1562 & 7.4$\pm$0.8 [39] & & & & \\ & & 52414.1143 & 11.0$\pm$0.7 [31] & & & & \\ 121743 & B2IV & 52414.1453 & 9.6$\pm$0.8 [57] & 5.3 & 1.4 & 79$^b$ & L87\\ 121790 & B2IV-V & 52414.1523 & 9.2$\pm$1.4 [35] & 4.8 & B & 124$^b$ & \\ 122980 & B2V & 52414.1594 & 10.5$\pm$0.6 [74] & 9.6 & 2.8 & 15$^b$ & L87 \\ 128345 & B5V & 52414.1828 & 9.9$\pm$1.9 [25] & 8.0 & D & 186$^b$ & \\ 129056 & B1.5III& 52414.1909 & 18.3$\pm$0.8 [72] & 5.4 & 0.6 & 16$^b$ & VDH\\ 130807 & B5IV & 52414.1957 & 7.1$\pm$0.6 [67] & 7.3 & A & 27$^b$ & VDH,ST\\ & & 52414.2150 & 7.2$\pm$0.5 [56] & & & & \\ 132200 & B2IV & 52414.2097 & 4.6$\pm$0.6 [70] & 8.0 & 0.9 & 32$^b$ &VDH,L87,ST\\ & & 52414.2266 & 4.9$\pm$0.6 [55] & & & & \\ 133242A\,& B5IV & 52414.2599 &$-$52.5$\pm$0.7 [4] & 4.5 & C & 140$^a$ & VDH\\ 133242B\,& & & 81.5$\pm$2.1 [6] & & & & \\ 133955A\,& B3V & 52414.2696 & 61.2$\pm$1.4 [10] & 9.8 & B & 135$^b$ & SB8\\ 133955B\,& & &$-$31.6$\pm$1.0 [10] & & & & \\ 134687 & B3IV & 52414.2777 & 23.7$\pm$0.8 [74] & 13.5 & D & 13$^b$ & SB8 \\ 136504 & B2IV-V & 52414.2904 & $-$5.7$\pm$1.1 [60] & 7.9 & C & 41$^b$ & SB8,ST \\ 137432 & B4Vp & 52414.2976 & 6.3$\pm$1.0 [41] & $-$0.8 & E & 77$^b$ & VDH,SB8 \\ 139365 & B2.5V & 52414.1535 & 33.3$\pm$2.3 [23] &$-$14.0 & E & 134$^b$ & SB8 \\ & & 52414.3140 & $-$5.8$\pm$1.1 [25] & & & & \\ 140008 & B5V & 52414.3210 & 2.6$\pm$0.8 [56] & 3.9 & B & 11$^b$ & SB8,ST \\ \hline \end{tabular} \end{center} \end{table} \begin{table} \begin{center} \begin{tabular}{llcrcccc} \multicolumn{8}{c}{Table 1. Continued} \\ \hline \hline \,\,\, HD &\,\,\,\,Sp& HJD &RV\,[N]\,\,\,\,\,\,\,\,& RV$_{\rm GCRV}$\,\,\, & eRV* & V $\sin$ i & Duplicity \\ & & & (km\,s$^{-1}$)\,\,\,\,\,\, & (km\,s$^{-1}$) & (km\,s$^{-1}$) & (km\,s$^{-1}$) & \\ \hline \multicolumn{8}{c}{Upper Ssorpius} \\ 142669 & B2IV-V & 52415.2944 & 2.5$\pm$1.0 [41] & 3.3 & E & 98$^b$ & SB8 \\ 142883 & B3V & 52415.3002 &$-$54.3$\pm$0.5 [70] &$-$27.5 & E & 14$^b$ & SB8 \\ 142990 & B5V & 52481.0065 & $-$5.6$\pm$3.5 [33] &$-$12.1 & 3.4 & 178$^b$ & L87 \\ & & 52481.1270 &$-$10.9$\pm$2.9 [34] & & & & \\ 143018A\,& B1V+ & 52414.3559 & 113.8$\pm$2.8 [25] &$-$11.7 & D & 100$^b$ & SB8 \\ 143018B\,& & &$-$173.6$\pm$3.2 [21]& & & & \\ 144217 & B0.5V & 52414.3646 & 9.1$\pm$1.3 [43] & $-$1.0 & & 91$^b$ &SB8,VDH,L87,ST\\ 144470 & B1V & 52415.2873 & $-$0.6$\pm$1.1 [39] & $-$4.4 & 3.0 & 100$^b$ & L87 \\ 147165 & B1III & 52414.3498 &$-$25.6$\pm$1.2 [50] & 2.5 & & 56$^b$ &SB8,L87,ST\\ 147888 & B3/B4V & 52481.1481 & $-$3.8$\pm$2.3 [27] & $-$6.8 & 2.9 & 175$^a$ & L87 \\ 147932 & B5V & 52481.2006 & $-$2.8$\pm$2.7 [23] &$-$11.0 & 2.4 & 186$^a$ & L87 \\ 148184 & B2Vne & 52481.1797 & $-$4.7$\pm$2.1 [34] &$-$19.0 & & 148$^b$ &SB8,L87\\ 149438 & B0V & 52414.3446 & 1.6$\pm$0.8 [43] & 1.7 & 0.8 & 10$^b$ & L87 \\ \multicolumn{8}{c}{Other} \\ 104841 & B2IV & 52411.0594 & $-$8.9$\pm$0.5 [73] & 16.1 & I & 25$^b$ & SB8 \\ 105435 & B2IVne & 52411.0725 & 3.8$\pm$2.8 [15] & 11.0 & C & 298$^b$ & VDH\\ 105937 & B3V & 52413.0558 & 22.7$\pm$1.5 [33] & 15.0 & C & 129$^b$ & \\ 109026 & B5V & 52411.1069 & 4.0$\pm$1.6 [31] & 2.5 & D & 188$^b$ & \\ 110335 & B6IVe & 52481.0996 & 4.8$\pm$1.8 [23] & 12.5 & B & 250$^a$ & \\ 111123 & B0.5IV & 52411.1357 & 9.8$\pm$0.7 [53] & 10.3 & A & 40$^b$ & VDH\\ 115846 & B3IV & 52481.97 &$-$21.8$\pm$1.6 [27] & 3.0 & 4.0 & 168$^b$ & \\ 116072 & B2.5Vn & 52411.1714 & 18.8$\pm$2.5 [21] & 3.0 & C & 233$^b$ & VDH\\ 118716 & B1III & 52413.1107 & 14.0$\pm$1.2 [32] & 3.0 & B & 114$^b$ & VDH\\ 120640 & B2Vp & 52413.1209 & $-$2.1$\pm$0.8 [63] & $-$4.7 & 0.8 & 21$^b$ & \\ & & 52414.1219 & $-$1.5$\pm$0.6 [49] & & & & \\ 126341 & B2IV & 52414.1712 &$-$21.1$\pm$1.0 [73] &$-$21.5 & B & 15$^b$ & VDH \\ 132058 & B2III & 52414.2050 & 0.1$\pm$1.0 [48] & 0.2 & 0.9 & 92$^b$ & L87 \\ 132955 & B3V & 52414.2424 & 5.1$\pm$0.5 [70] & 3.7 & 2.1 & 8$^b$ & VDH \\ 144218 & B2V & 52414.3716 & 0.6$\pm$0.8 [57] & $-$5.6 & 0.8 & 56$^b$ &VDH,L87,ST,SB8\\ 151985 & B2IV & 52414.3352 & 1.9$\pm$0.7 [56] & 1.3 & 0.7 & 52$^b$ & L87,SB8 \\ \hline \end{tabular} \end{center} \begin{list}{}{} \item * A: errors $\le$ 2.5 ${\rm km\,s}^{-1}$; B: 2.5 $<$ errors $\le$ 5.0 ${\rm km\,s}^{-1}$; C: 5.0 $<$ errors $\le$ 10.0 ${\rm km\,s}^{-1}$; D: errors $\ge$ 10 ${\rm km\,s}^{-1}$; E: too uncertain (from Table 3 of Barbier-Brossat \& Figon \cite{bbf00}) \item a: from Glebocki \& Stawikowski (\cite{gle00}) and b: from Brown \& Verschueren (\cite{ver97}) \item SB8 - Eighth Orbital Elements of Spectroscopic Binaries (Batten et al. \cite{bat89}) \item L87 - Levato et al. (1987) \item VDH - Visual Double Stars in Hipparcos (Dommanget \& Nys \cite{dom00}) \item ST - Shatsky \& Tokovinin (2002) \end{list} \end{table}
Title: Quantifying the Luminosity Evolution in Gamma-ray Bursts
Abstract: We estimate the luminosity evolution and formation rate for over 900 GRBs by using redshift and luminosity data calculated by Band, Norris, $&$ Bonnell (2004) via the lag-luminosity correlation. By applying maximum likelihood techniques, we are able to infer the true distribution of the parent GRB population's luminosity function and density distributions in a way that accounts for detector selection effects. We find that after accounting for data truncation, there still exists a significant correlation between the average luminosity and redshift, indicating that distant GRBs are on average more luminous than nearby counterparts. This is consistent with previous studies showing strong source evolution and also recent observations of under luminous nearby GRBs. We find no evidence for beaming angle evolution in the current sample of GRBs with known redshift, suggesting that this increase in luminosity can not be due to an evolution of the collimation of gamma-ray emission. The resulting luminosity function is well fit with a single power law of index $L'^{-1.5}$, which is intermediate between the values predicted by the power-law and Gaussian structured jet models. We also find that the GRB comoving rate density rises steeply with a broad peak between $1<z<2$ followed by a steady decline above $z> 3$. This rate density qualitatively matches the current estimates of the cosmic star formation rate, favoring a short lived massive star progenitor model, or a binary model with a short delay between the formation of the compact object and the eventual merger.
https://export.arxiv.org/pdf/astro-ph/0601146
\title{Quantifying the Luminosity Evolution in Gamma-ray Bursts} \author{Dan Kocevski \altaffilmark{1} and Edison Liang\altaffilmark{}} \altaffiltext{1}{Physics Department, University of California, Berkeley, Berkeley, Ca 94709 } \altaffiltext{2}{Department of Physics and Astronomy, Rice University, Houston, Tx 77005 } \email{kocevski@rice.edu} \email{liang@spacibm.rice.edu} \keywords{gamma rays: bursts---gamma rays: theory} \section{Introduction} \label{sec:Introduction_ch3} There are currently roughly three dozen gamma-ray burst events (GRBs) for which we have independently measured redshifts. Most of these redshift determinations come from either identification of absorption lines in the afterglow spectra, attributed to the gas in the host galaxy, or from observations of emission lines from the host galaxy. The combination of these techniques has resulted in a small but growing GRB sample with redshifts ranging from 0.0085 to 4.5 and a distribution peaking around $z \sim 1$. From this small sample, it is already abundantly clear that the isotropic equivalent energy $E_{iso}$ released in the prompt GRB phase is not a standard candle. The total radiated energy taken at face value (i.e. when not correcting for a beaming factor $d\Omega$) clearly spans several orders of magnitude, ranging from $10^{47}$ for the closest event, GRB 980425 at $z=0.0085$ \citep{Kulkarni98}, to $10^{54}$ for GRB 990123 at $z=1.6004$ (Kulkarni et al. 1999). Recently \citet{Sazonov04} and \citet{Soderberg04} have reported on gamma-ray observations of a nearby underluminous GRB occurring at redshift of $z=0.106$. These new findings have added to the speculation that there is either a substantial under luminous population of GRBs which cannot be seen at large distances and/or that nearby events ($z <$ 0.15) are underluminous compared to distant counterparts, pointing to the evolution of the average energy emitted by a GRB with time. A measure to the extend to which luminosity evolution exists in the GRB population, along with their true luminosity function and density distribution, may yield important clues regarding the nature of gamma-ray bursts and how they're progenitors have evolved with time. Although the physics of the underlying GRB engine is hidden from direct observation and is yet uncertain, the total GRB energy budget is most likely linked to the mass and/or rotational energy of the GRB progenitor. Understanding of how this energy budget has changed with time may offer constraints on progenitor properties and may ultimately point to the physics leading to their explosions. Since GRB progenitors are most likely linked to compact objects (supermassive rotating star, black hole or neutron star mergers) understanding how the GRB luminosity function evolves with time may give insight to the host environment in the early universe, namely the star formation rate or initial stellar mass functions at high redshifts. Any attempt at quantifying the evolution of intrinsic source properties of parent populations must account for Malquist type biases. Detection thresholds prevent events below a certain flux from being observed, resulting in the detection of only bright objects at large distances. Combined with the fact that bright events are typically rare, it is very easy in astronomy to incorrectly conclude that the distant universe is filled with extremely bright rare objects. Any attempt at measuring the correlation between luminosity and redshift without properly accounting for selection effects will grossly overestimate the correlation strength between the two variables. Flux limited samples are a classic problem in astronomy, which manifested prominently in early quasar studies. Fortunately, straight forward methods have been devised to account for such effects based on maximum likelihood techniques. These methods allow for the correct estimation of the correlation strength between a truncated data set as well as an estimate on the underlying parent population. The "catch" of such techniques is that the overall normalization of the resulting parent distributions cannot be determined, although their functional forms are constructed in such a way to account for the data truncations. These techniques also have to limitation of requiring a large sample sizes and more importantly, an extremely good understanding of the survey's detection thresholds (i.e. the flux cutoff for magnitude limited samples). The use of the current sample of GRBs with known redshift is limited by both of these restrictions. The current size of a little over two dozen bursts does not lend itself well to producing statistically robust results, especially in the high and low redshift regimes for which only a handful of events have been detected. Furthermore, the sample is an accumulation of observations from several different spacecraft, all of varying detector thresholds. It would seem that these limitations could only be overcome by the accumulation of a larger data set with consistent detector thresholds which is expected to come from the Swift spacecraft and the upcoming GLAST mission. Fortunately, several authors have announced empirical Cepheid like correlations linking intrinsic burst properties, such as luminosity \citep{norris00} and the total radiated energy \citep{amati02,ghirlanda04a} to other GRB observable. These correlations may allow for the determination of burst redshifts directly from the gamma-ray data, which has the advantage of being relatively insensitive to extinction and observable at far greater distances than afterglow line measurements. The first of these correlations was reported by \citet{norris00}. Using 6 BATSE detected bursts with known redshift, they found an anti-correlation between the {\it source} frame lag between the 25-50 keV and 100-300 keV emission and the absolute luminosity of the GRB. More recently, \citep{ghirlanda04a} reported an empirical correlation between the collimation correction total energy $E_{\gamma}$ radiated by the burst and the rest frame energy at which most of the prompt radiation is emitted $E_{pk}$. Using these relationships, it is now possible to estimate "pseudo" redshifts for a much larger number of GRBs detected by the BATSE instrument which perviously lacked any information as to their distance. More importantly, the BATSE detector threshold is relatively well understood for the entire sample, making the resulting pseudo redshift data excellent for statistical analysis. In this paper, we examine the issue of luminosity and density evolution by using a sample of over 900 BATSE GRBs for which the luminosity and redshift where recently estimated by \citet{band04} through the use of the lag-luminosity correlation. We limit our analysis to the lag-luminosity correlation primarily due to the lack of jet opening angle $\theta_{j}$ information that is required for the use of the $E_{\gamma}$-$E_{pk}$ relation. This relationship requires knowledge of $\theta_{j}$ in order to determine the collimation factor, which is only known for bursts with measured jet break times and hence cannot be used with the BATSE sample in consideration for this paper. We found that the more general, and much broader, correlation between intrinsic $E_{pk}$ and $E_{iso}$ reported by \citet{amati02} did not provide redshift constraints for a majority of the bursts in our sample. This is consistent with recent observations by \citet{nakar} and \citet{bandpreece} who also found large fractions of their BATSE samples to be inconsistent with the Amati correlation. Therefore we limit the current analysis to distances estimated through the use of the lag-luminosity relationship. To our sample, we apply statistical techniques developed by \citet{Lynden-Bell71} and \citet{efron92} and first applied to GRB analysis by \citet{lloyd02} to measure the underlying luminosity and density distribution in a way that properly accounts for the detection thresholds of the BATSE instrument. We find a strong (11.63 $\sigma$) correlation between luminosity and redshift that can be parameterized as $L(z) = (1+z)^{1.7 \pm 0.3}$. The resulting cumulative luminosity function $N(L')$ is well fit by double power law separated by a break energy of about $10^{52}$ ergs s$^{-1}$, with the differential luminosity function $dN/dL'$ exhibiting a power law shape of $L^{-1.5}$ below this luminosity. We show that the GRB comoving rate density increases roughly as $\rho_{\gamma}(z) \propto (1+z)^{2.5}$ to a redshift of $z \approx 1$ followed by a flattening and eventual decline above $z>3$. This rate density is in qualitatively agreement with recent photometric estimates of the cosmic star formation rate (SFR), as would be expected from massive short lived progenitors. In $\S 2$, we describe the data set that we use in our study. In $\S 3$ we discuss the statistical methods applied to this data to estimate the GRB luminosity function and comoving rate density as well as to test for any correlation between luminosity and redshift. In $\S 4$ we present the resulting demographic distribution functions of this analysis followed in $\S 5$ by a discussion of the implications of the shape and evolution of the luminosity function and comoving rate density on various jet profile. We show that there is no evidence for beaming angle evolution in the current sample of GRBs with known redshift, suggesting that the variation of the observed luminosity with redshift can not be due to an evolution of the collimation of gamma-ray emission. We conclude by examining how the similarity between the SFR and the GRB comoving rate density tentatively favors short lived progenitor models. \section{Data} \label{sec:data_ch3} For this analysis we utilize data for 1438 BATSE detected GRBs presented in \citet{band04}, hereafter BNB04. This sample includes peak photon flux $f_{pk}$ in the 50-300 keV band on a 256 ms timescale, the burst duration $T_{90}$, and measured lags and their uncertainties for each burst. From these lag measurements, the authors infer each burst's luminosity and redshift by use of the lag-luminosity correlation, allowing also for an estimation of the intrinsic $E_{pk}$ and $E_{iso}$ for each burst. Of these 1438 bursts, 1218 have positive lags making them suitable for this analysis. This data is shown in Figure \ref{fig:data} with an imposed flux cut set at 0.5 photon cm$^{-2}$ s$^{-1}$, leaving a total of 985 bursts. The lags measurements used in this sample where made using a cross-correlation analysis similar to that previously employed by \citet{band93b} and \citet{norris00}. The cross correlation method has been widely used in x-ray and gamma-ray astronomy, and is well suited for timing analysis between two signals. In this application, the normalized discrete cross correlation function is given by \begin{equation} \label{eq:CCF} CCF(\tau)=\sum_{i}^{N-1}\frac{f_{i}(t)*g_{i}(t-t')}{\sigma_{f}\sigma_{g}} \end{equation} where $t'$ is commonly referred to as the lag between $f(t)$ and $g(t)$ and $\sigma_{f}=\langle f(t)^{2}\rangle^{1/2}$. By maximizing the CCF function (i.e. by maximizing the area of the product of the two functions) as a function of $t'$, an estimate of the temporal offset of the two signals can be made. If $g(t)$ leads $f(t)$ by $t_{0}$ (i.e. $f(t)=g(t+t_{0})$) than the CCF curve peaks at $t'=t_{0}$. In BNB04, the authors utilize 64ms count data gathered by BATSE's Large Area Detectors (LADs) which provide discriminator rates with 64 ms resolution from 2.048 s before the burst to several minutes after the trigger \citep{fish94}. The discriminator rates are gathered in four broad energy channels covering approximately 25-50, 50-100, 100-300, and 300 to about 1800 keV allowing for excellent count statistics since the photons are collected over a wide energy band. BNB04 measure the temporal offset or lag between channel 3 (100-300 keV) and channel 1 (25-50 keV) light curves to produce the CCF31 lags listen in their sample. The shifting of the GRB spectra out of (or into) the observers frame, otherwise known as the k-correction, was accounted for in the analysis performed in BNB04. They perform spectral fits for most of the bursts in their sample and for those which cannot yeild a fit, a "Band" spectral model with average parameters is assumed for the spectra. The effects of time dilation and k-correction are then used to obtain the source frame lag and also applied to the energy flux to obtain a bolometric luminosity. As was the case in the original \citet{norris00} paper, the CCF method used in BNB04 can result in lag measurements which are less than the 64ms time resolution of the BATSE instrument. In these cases, the associated errors of these values tend to be quite large, reducing the significance of their associated luminosity and redshift values. These errors are taken into consideration in the maximum likelihood techniques performed in our analysis. Therefore, bursts with extremely short lags (and hence high luminosity's) are weighted accordingly. A plot of the lag-luminosity plane for the events under consideration along with the errors in the lag measurements are shown in Figure \ref{fig:laglumplane}. \subsection{Estimating Redshifts} \label{sec:redshifts} Using these lag measurements, BNB04 utilize the lag-luminosity correlation to estimate the luminosity of each event. This empirical correlations was reported by \citet{norris00} who used the CCF method to measure the lag between BATSE's channel 3 and channel 1 energy light curves for 6 GRBs with independently measured redshift. They concluded that there was an anti-correlation between the {\it source} delay in the low and high energy emission and the absolute luminosity of the GRB showing that high luminosity events exhibited very small intrinsic (source frame) lag, whereas fainter bursts exhibited the largest time delay. This empirical correlation can be expressed as \begin{equation} \label{eq:laglum} L = 2.51\times10^{51} (\Delta t'/0.1)^{-1.15} \end{equation} where $\Delta t'$ is the source frame lag related to the observed lag $\Delta t'_{obs}$ by a time dilation factor of $(1+z)^{-1}$. The fact that the lag-luminosity correlation relates two source frame quantities (i.e. luminosity and intrinsic lag) would make it seem that knowledge of the redshift is needed \emph{a priori}. As it turns out this is not the case. A simple numerical iteration routine can be used to solve for the redshift of a GRB which lacks any information as to its distance. This is done by first making an initial guess for $z$ (say $z \sim 1$) to obtain the lag in the comoving frame $\Delta t' = \Delta t'_{obs}/(1+z)$. This in turn gives us an initial value for the luminosity through the use of the lag-luminosity relation. This luminosity is then used in combination with the burst's energy flux to obtain a value for the luminosity distance $D_{L}$ through the standard relation \begin{equation} \label{eq:DL1} D_{L} = \sqrt{\frac{L/d\Omega}{f_{256}}} \end{equation} where $f_{256}$ is the peak flux in the 256 ms timescale and $d\Omega$ is the beaming factor. This distance is then compared to the $D_{L}$ that can be calculated directly from the guessed redshift $z$ by assuming standard cosmological parameters ($H_{o} = 65$ km s$^{-1}$, $\Omega_{m} = 0.3$, $\Omega_{\Lambda} = 0.7$) and using the expression \begin{equation} \label{eq:DL2} D_{L} = (1+z)\frac{c}{H_{0}}\int_{0}^{z}\frac{dz}{\sqrt{\Omega_{m}(1+z)^{3}+\Omega_{\Lambda}}} \end{equation} The value for $z$ is then varied until the luminosity distances obtained from the two separate methods converge to within some predetermined precision. We note that it has been suggested by \citet{salmonson01} and \citet{norris02} that the lag-luminosity relationship should be a broken power law in order to accommodate GRB 980425. This burst was associated with SN 1998bw and when using the distance to the supernova, the GRB appears under luminous compared to the other bursts that fall on the lag-luminosity correlation. In their analysis, BNB04 note that this break has been suggested to fit a single point, which may or may not be associated with the SNe event and hence decide to use a single power law of -1.15. The physical origin of the lag-luminosity correlation is not immediately clear. Fundamentally, this observed lag is due to the evolution of the GRB spectra to lower energies, so a relationship between the rate of spectral decay and luminosity is expected \citet{kocevski03a}. This implies that the mechanisms resulting in the "cooling" of the GRB spectra is intimately related to the total energy budget of a GRB or its collimation factor. Other purposed theories attempt to explain the lag-luminosity correlation as being due to the effect of the viewing angle of the GRB jet \citep{krm02, Ioka01}, and or kinematic effects \citep{salmonson00}. In any case, the use of this correlation is similar to methods used to calibrate Type Ia supernova luminosities based on the empirical correlation between their peak magnitude and rate of light curve decay (Phillips 1999). The lack of a clear physical interpretation of these correlations does not immediately preclude their use in determining, or refining, luminosity estimates. \section{Analysis} \label{sec:analysis_ch3} The luminosity and redshift data calculated by BNB04 gives us an enormous sample from which to investigate the evolution of the GRB luminosity function. As with any cosmological source, it is important and revealing to understand of how the average luminosity and density has evolved with cosmic time. Attempting to do so by simply measuring the correlation coefficient between the flux truncated luminosity and redshift data in the BNB04 sample without properly accounting for the detector selection effects would grossly overestimate the correlation strength. This is true whenever an estimate of correlation is made between two variables that suffer from data truncations, with the resulting correlation coefficient representing the truncation itself and not the underlying relation. There have been several methods developed in astronomy to account for such selection effects, based largely on maximum likelyhood techniques (see Petrosian 1992 for a review). In our analysis, we use a nonparametric statistical technique originally proposed by \citet{Lynden-Bell71} for applications in flux limited quasar studies. This so called C-Method has been used successfully to reconstruct underlying parent distributions for quasars and GRBs samples by \citet{Maloney99}, and \citet{lloyd02} respectively. The parent luminosity and redshift distributions which the method estimates allows for the construction of a GRB luminosity function, a measure of the number of bursts per unit luminosity, and an estimate on the comoving rate density, a measure of the number of bursts per unit comoving volume and time. The C-Method has two important limitations, or stipulations, to its use. First, the truncation limit below which no observations can be made must be well known. This is not a problem in our case, since the detector threshold of the BATSE instrument is well understood and BNB04 quantify the truncation limit of their sample. Secondly, the parent luminosity and redshift distributions can only be estimated in a bivariate manner if the two variables are uncorrelated. This is a limitation of all nonparametric techniques which rely on the assumption of stochastic independence. Therefore, it is necessary to first determine the degree of correlation between the two variables, in our case luminosity and $Z = 1+z$, and then produce an uncorrelated data set through the transformation $L \rightarrow L' = L/g(z)$, where $g(z)$ parameterizes the luminosity evolution. Using this uncorrelated data set, it is then possible to apply the C-Method to estimate the underlying parent luminosity and redshift distributions. To estimate the degree of correlation we use a simple test of independence for truncated data put forth by \citet{efron92} which is based in part on Lynden-Bell's C-Method. Below we describe the details of both Lynden-Bell's C-Method and the Efron $\&$ Petrosian independence test and how they are applied in our analysis. \subsection{Test of Independence} \label{sec:efron} If the variables $x$ and $y$ in a data set are independent, then the rank $R_{i}$ of $x_{i}$ within that set should be distributed uniformly between 1 and $N$ with an expected mean $E=(1/2)(N+1)$ and variance $V=(1/12)(N^{2}-1)$. It is common practice to normalize the rank $R_{i}$ such that for independent variables $R_{i}$ has a mean of 0 and a variance of 1 by defining the statistic $T_{i}=(R_{i}-E)/V$. A specialized version of the Kendell $\tau$ statistic can be constructed to produce a single parameter whose value directly rejects or accepts the hypothesis of independence. This quantity is commonly defined as \begin{equation} \label{eq:tau0} \tau = \frac{\Sigma_{i} (R_{i}-E)}{\sqrt{\sum_{i}V}} \end{equation} Based on this definition, a $\tau$ of 1 indicates a 1 $\sigma$ correlation whereas a $\tau$ of 0 signifies a completely random data set. See \citet{efron92} for a more detailed (and elucidating) proof of the applicability of normalized rank statistics. The modified version of this rank statistic proposed by \citet{efron92} to test the independence of truncated data is based on a simple concept. Instead of measuring the ranks $R_{i}$ for the entire set of observables, rather deal with data subsets which can be constructed to be independent of the truncation limit suffered by the entire sample. This is done by creating "associated sets" which include all objects that could have been observed given a certain limiting luminosity. We can define an associated set as \begin{equation} \label{eq:Ji} J_{i} \equiv \{j:L_{j} > L_{i}, L_{lim,j} < L_{i}\} \end{equation} In other words, for each burst $i$ there can be constructed a data subset that includes all events within the range $L_{i} < L < \infty$ and $0 < z < z_{max}(L_{i})$. The boundaries of an associated set for a given burst $i$ are shown as dotted lines in Figure \ref{fig:associatedsets}. In this scenario, we expect the rank $R_{i}$ of $z_{i}$ within the associated set \begin{equation} \label{eq:Ri} R_{i} \equiv \{j\in J_{i} : z_{j} \leq z_{i}\} \end{equation} to be uniformly distributed between 1 and $N_{j}$, where $N_{j}$ is the number of points in the associated set $J_{i}$. Using these new ranks, we can again construct the mean and variance, except that now we replace $N$ with $N_{j}$ such that $E=(1/2)(N_{j}+1)$ and $V=(1/12)(N_{j}^{2}-1)$. The specialized version of Kendell's $\tau$ statistic is now given by \begin{equation} \label{eq:tau} \tau = \frac{\Sigma_{i} (R_{i}-E_{i})}{\sqrt{\sum_{i}V_{i}}} \end{equation} where the mean and variance are calculated separately for each associated set and summed accordingly to produce a single value for $\tau$. This parameter represents the degree of correlation for the entire sample with proper accounting for the data truncation. With this statistic in place, it is a simple matter to find the parametrization that best describes the luminosity evolution. This is accomplished by first choosing a functional form for the luminosity evolution, which in our case we choose a simple power law dependence $g(z) = (1+z)^{\alpha}$. We can then make the transformation $L \rightarrow L' = L/g(z)$ and vary $\alpha$ until $\tau \rightarrow 0$. An example of how well these methods are able to estimate underlying correlations in truncated data is shown in Figure \ref{Fig:fakedata}. Here we have plotted a distribution of fake luminosity and redshift data with some known power law dependence $L \propto (1+z)^{p}$ which is subjected to a flux cut $L_{lim} \propto (1+z)^{q}$ represented by the red dashed line. The crosses show the observable data whereas the dots represent the data that would otherwise be undetectable. The long dashed line is the best fit to the truncated data without any knowledge of the flux cut whereas the dash dot line is the reconstructed correlation when taking into account the flux threshold. This method fails when the undetected data points become significantly larger than the observable data set, with the exact boundary at which this occurs depending on the difference in the power law indices between the underlying correlation and the flux threshold. Since these quantities cannot be known a priori, it is explicitly assumed that a large data sample contains a sufficient amount of events above the flux threshold for the method to work. A histogram of the difference between the known correlation index and the reconstructed index $(p-q)$ for multiple such simulations is shown in Figure \ref{Fig:alphadist}. The error, or difference between the known $p$ and the measured $q$ is peaked about zero with a fwhm which roughly matches that error estimates that correspond to the 1 $\sigma$ range for this parameter given by the condition $|\tau|<1$. \subsection{Determination of Distribution Functions} \label{sec:distributions} Once a parametric form that removes the the correlation between $L$ and $z$ is known, it is possible to use nonparametric maximum likelyhood techniques to estimate the underlying parent luminosity and redshift distributions. This luminosity distribution $\Phi(L_{i})$ represents the cumulative GRB luminosity function with the redshift distribution $\sigma(z_{i})$ representing the GRB density evolution. \citet{Petrosian92} has shown that many, if not most, of the familiar nonparametric methods used in astronomy to produce $\Phi(L_{i})$ and $\sigma(z_{i})$ reduce fundamentally to Lynden-Bell's C-Method. Consider the area, or number of events, in the box produced by the associated set shown in Figure \ref{fig:associatedsets}. If $N_{1}$ represents the number of points with $L \geq L_{1}$, then let $dN_{1}$ represent the number of points in the infinitesimal column between $L_{1}$ and $L_{1}+dL_{1}$. The general premise behind the C-Method is that if the two variables $(L,z)$ are stochastically independent, then the ratio between $N_{1}$ and $dN_{1}$ should equal the ratio between $d\Phi$ and the true cumulative distribution function $\Phi(L_{1})$ \begin{equation} \label{eq:dnn} \frac{dN_{1}}{N_{1}} = \frac{d\Phi}{\Phi_{1}} \end{equation} which can then be integrated to find $\Phi(L)$. In the case of discrete data points, this integration becomes a summation, yielding the solution \begin{equation} \label{eq:lyndenbell1} \Phi(L_{i}) = \prod_{k=2}^{j}\left(1+\frac{1}{N_{j}}\right) \end{equation} where $N_{j}$, is the number of bursts in the box defined by $0<z<z_{max}(L_{j})$ and $L_{j}<L<\infty$. The value $N_{j}$ is the same as Lynden-Bell's $C^{-}_{j}$ in that it does not count the $L_{i}$ object that is used to form the associated set. Similarly, we can construct the underlying cumulative redshift distribution function $\sigma(z_{i})$ by reversing the definition of the associated set such that $M_{j}$ represents the number of bursts in the box $0<z<z_{i}$ and $L_{min}(z_{i})<L<\infty$. Then \begin{equation} \label{eq:lyndenbell2} \sigma(z_{i}) = \prod_{k=2}^{j}\left(1+\frac{1}{M_{j}}\right) \end{equation} As mentioned in $\S$ 1, there are several important limitations to the C-method. First, the overall normalization of $\Phi(L_{i})$ and $\sigma(z_{i})$ is arbitrary, so information regarding the absolute numbers and densities cannot be obtained. Despite this, the shape of the bivariate distribution is constructed in such a way that it accounts for the data truncations. Due to this limitation, all distributions presented in this paper will have arbitrary normalizations. Secondly, it is clear from Equation \ref{eq:lyndenbell1} and \ref{eq:lyndenbell2} that the cumulative distribution function is not defined when either $N_{j}$ or $M_{j}$ are zero. This limitation restricts the use of the C-method to samples with a data size sufficiently large to ensure that all associated sets greater than $j=2$ contain a nonzero number of points. \section{Results} \label{sec:results_ch3} \subsection{Luminosity Evolution} \label{sec:lumevolution} We apply the test of independence outlines in the $\S$ 3.1 to the entire BNB04 GRB sample to test for luminosity evolution. For this analysis we use the flux threshold suggested by BNB04 of $f_{min}$ = 0.5 photons cm$^{-2}$ s$^{-1}$, decreasing the sample size to 985 bursts. Applying this method, we find evidence for a strong 11.63 $\sigma$ correlation between luminosity and redshift. This evolution is well parameterized by a power law of the form $g(z) = (1+z)^{\alpha}$, with an optimal value for the power law index (i.e when $\tau(\alpha)$=0 given the transformation $L \rightarrow L' = L/g(z)$) of $\alpha$=1.7 $\pm$ 0.3. The error estimates on $\alpha$ correspond to the 1 $\sigma$ range for this parameter given by the condition $|\tau|<1$. A plot of $\tau(\alpha)$ vs. $\alpha$ with the corresponding 1 $\sigma$ levels are shown in Figure \ref{fig:taualpha}. These findings indicate that the average luminosity (modulo a beaming factor $d\Omega$) of GRBs in the universe has evolved with time. Because of the lack of beaming information, it may also be possible that the luminosity is remaining constant while the beaming factor $d\Omega$ is actually evolving. As will be discussed in $\S$ 5, there is no observational evidence to suggest that this is the case. It should also be noted that $\tau(\alpha)$ appears to be strongly affected by the choice of the flux threshold assumed for the sample. Plotted in Figure \ref{fig:alphacut} is the optimal value for $\alpha$ vs. $f_{min}$. Not surprisingly, if we assume no flux threshold (i.e. $f_{min}=0$), $\tau$ approaches the overestimated value received from the standard Kendell $\tau$ statistic. $\alpha$ similarly approached the value obtained by simply performing a power-law fit to the truncated data. $\alpha$ decreases steeply with increasing $f_{min}$, never reaching a stable plateau as one would hope would happen as the $f_{min}$ approaches the $\emph{true}$ threshold of the detector. This underscores the importance of having a good understanding the thresholds of the detector used to collect the sample. BNB04 make a strong case for a threshold of $f_{min}$=0.5 photons cm$^{-2}$ s$^{-1}$ based on where they see a strong drop off of detected events in the $L-Z$ plane (see their Figures \ref{fig:alphacut} $\&$ \ref{fig:alphacut}) and we adopt this value for all analysis presented in this paper. \subsection{Luminosity Function} \label{sec:lumfunc} The deduced parametric form describing the luminosity evolution allows us to use the C-method on the uncorrelated parameters $L'$ and $Z$ to obtain the cumulative luminosity function $\Phi(L')$. Shown in Figure \ref{fig:cumluminosity} is the cumulative $\Phi(L')$ distribution plotted as $\Phi(>L')$ as a function of $L'$ for all 985 bursts. Because the luminosity evolution has been explicitly removed, this distribution represents the luminosity function in the present epoch. Fitting a double power law to the curve yields $\Phi(>L') \propto L'^{-0.623}$ and $\Phi(>L') \propto L'^{-1.966}$ for the low and high luminosity ranges respectively, separated by a break at a luminosity of roughly $\sim 10^{52}$. These slopes are very similar to those reported by \citet{lloyd02} who found a GRB cumulative luminosity function with power law slopes of $k_{1}=-0.51$ and $k_{2}=-2.33$ below and above a break at about $L'=5.9\times10^{51}$. These values can also be compared to the luminosity functions found by \citet{Maloney99} who employ the C-method to account for selection effects in quasar samples. They find that the quasar luminosity function exhibits a double power law form with indices of $k_{1}=-1.16$ and $k_{2}=-3.59$. Next, we differentiate the cumulative luminosity function with a 3-point Lagrangian interpolation to find the differential luminosity function $d\Phi/dL'$, or what is commonly referred to as simply the luminosity function $\psi(L')$. This function represents the total number of bursts with luminosity between $L'$ and $L'+dL'$. A plot of the $\psi(L')$ vs. $L'$ is shown in Figure \ref{fig:lumfunction}. The function falls roughly as $\psi(L') \propto L'^{-1.5}$ below the break energy of $\sim 10^{52}$ with a sharp decline for higher $L'$. This power law index is identical to the slope found by \citet{lloyd02} who found $L'^{-1.5}$ and similar to the index found by \citet{Schaefer01} who found $L^{-1.7 \pm 0.1}$ from $(L,z)$ data estimated from a combined use of the lag-luminosity function and variability-luminosity function, although the latter did not account for any selection biases in their data set. This value is also similar to results of several studies that used the measured flux distribution with an assumed density distribution $\rho(z)$, such as \citep{Schmidt01} who uses the star formation rate to infer a $\rho(z)$ and finds $\psi(L') \propto L^{-1.4}$. The shape of the GRB luminosity function has important implications to jet model theories which predict specific power law indices for various jet structures. A comparison between theorized shapes and our deduced values will be discussed in more detail in $\S$\ref{sec:discussion_ch3}. \subsection{Density Evolution} \label{sec:densityevolution} Using the alternative definition of the associated set, we can construct the cumulative density distribution $\sigma(z)=\int_{0}^{z}\rho(z)(dV/dz)dz$, or the total number of GRBs per comoving volume, up to a given redshift. The cumulative distribution is shown in Figure \ref{fig:cumdensityvolume} plotted as $\sigma(>z)$ as a function of $z$. The distribution of GRBs appears to increase smoothly with $z$, without a pronounced break at any distance, but with a flattening at high redshift indicating a drop off of events between $5\leq z \leq 10$. To get a better look at this density evolution, we can plot the cumulative density distribution $\sigma(z)$ as a function of comoving volume $V(z)$ as seen in Figure \ref{fig:cumdendist} If the density of GRBs per comoving volume $V(z)$ is constant, i.e. $\rho(z)=\rho_{0}$, then it should follow that $\sigma(z) \propto V(z)$. We can test for evolution by fitting $\sigma(z)$ vs $V(z)$ to a simple power law $\sigma(z) \propto V(z)^{\beta}$ and looking for deviations from the constant density case. An index of $\beta \neq 1$ indicates the presence of density evolution, with $\beta > 1$ and $\beta < 1$ signifying an increasing and decreasing population respectively. Using the definition of $V(z)$ in a flat universe of \begin{equation} \label{eq:volume} V(z) = \frac{4\pi}{3}\left[\frac{c}{H_{0}}\int_{0}^z\frac{dz}{\sqrt{\Omega_{m}(1+z)^{3}+\Omega_{\Lambda}}}\right]^{3} \end{equation} we find that the cumulative density distribution increase with $z$ roughly as $\sigma(z) \propto V^{1.25}$ at low redshifts before falling off at higher redshifts. From these results we can deduce that the GRB density has undergone complicated evolution, increasing as $\rho \sim V^{0.25}$ before peaking between $z \sim 1-2$ and then decreasing. To obtain a more quantitative look at the shape of the comoving rate density $\rho(z)$, we again use a 3-point Lagrangian interpolation routine on $\sigma(z)$ to find the differential cumulative distribution $d\sigma/dZ$. We can then convert this differential distribution into a comoving rate density through the relation: \begin{equation} \label{eq:rho} \rho(Z)=\frac{d\sigma}{dZ}(1+z)\left(\frac{dV}{dZ}\right)^{-1} \end{equation} In Figure \ref{fig:comovingratedensity} we show the resulting comoving rate density plotted as a function of $z$. It can be seen that the GRBs density function increases out to a redshift between $1\leq z \leq$ then flattens before beginning to show signs of a turn over at a redshift of $z > 3$. This is in contrast to previous estimates of the comoving rate density by \citet{Schaefer01}, \citet{lloyd02}, and \citep{Yonetoku04} all of who find a flattening of the GRB population with no apparent turn over out to a redshift of $z \sim 10$. It is also in contrast to results reported by \citet{Murakami} who also used the lag-luminosity correlation to estimate the GRB formation rate. There the authors find the GRB formation rate increases steadily out to a redshift of at least 4, but it should be noted that this work did not take into account the detector selection effects discussed above so a direct comparison may not be appropriate. As opposed to these previous findings, the turn over observed in our data quantitatively matches the global behavior of the star formation rate of the universe which has been observed to peak between $1 \leq z \leq 2$ followed by a steady decline \citep{madau96,Steidel99}. A more detailed comparison between the GRB comoving rate density and the supernova and star formations rates will be continued in $\S 5$. \section{Discussion} \label{sec:discussion_ch3} We find an 11.63 $\sigma$ correlation between the luminosity and redshift data deduced from the lag-luminosity correlation, strongly suggesting an evolution of the average luminosities of GRBs. We show that this correlation can be parameterized as a power law as $L(z) = (1+z)^{1.7 \pm 0.3}$. This value agrees extremely well with the results presented in \citet{lloyd02} who found a power law index of $\alpha = 1.4$ after performing a similar analysis on $(L,z)$ data estimated using the variability-luminosity correlation. These results imply that the average energy emitted per unit time per unit solid angle by GRBs was much higher in the distant past compared to relatively recent events. This is consistent with previous studies showing strong source evolution and also recent observations of under luminous nearby GRBs. Due to our lack of knowledge regarding the beaming angle of the bursts in our sample, it is also possible that the increase in the apparent luminosity is due to an increasing collimation at higher redshifts. As we will discuss in more detail below, we find no evidence for beaming angle evolution in the current sample of GRBs with known redshift and jet opening angle, suggesting that this increase in luminosity can not be due simply to an evolution of the collimation of the gamma-ray emission. \subsection{Comparison to Other Objects} \label{sec:comparisons} Such a steep luminosity evolution is not uncommon in other astrophysical objects that show evolution with redshift. \citet{Maloney99} perform a similar analysis using the statistical techniques described in this paper on a combination of several quasar samples and find that the quasar luminosity function evolves as $L(z) = (1+z)^{2.58}$ up to a redshift of at least 2. There is evidence that this evolution may then become constant up to a redshift of at least 3 \citep{Boyle93}. We find no such break in the luminosity evolution of GRBs, which in our case can be adequately fit by a single power law between at least $0 < z < 10$. The authors also find a density evolution of $\sigma(z) \propto V^{1.19}$ similar to the power law of $\sigma(z) \propto V^{1.25}$ that we find in GRBs at low redshifts. A more detailed look at their comoving rate density estimate shows that the quasar density rises as $\rho \sim (1+z)^{2.5}$ before peaking at $z \approx 2$ and then declining rapidly as roughly $\rho(z) ~ (1+z)^{-5}$ for $z > 2.0$. This is qualitatively similar to the trend we deduce from the GRB sample, which rises as $\rho \sim (1+z)^{2.4}$ to a $z \approx 1$ although the proceeding decline is much more shallow as $\rho \sim (1+z)^{-0.6}$ and extends to at least a redshift of $z \approx 6$ before dropping off sharply. There is also evidence for significant evolution in the luminosities of star forming galaxies, which is perhaps a more relevant comparison to GRBs because of their suggested association with active star forming regions \citep{Djorg98}. Hopkins (2004) used a compilation of recent star formation rate density measurements as a function of redshift to constrain the evolving luminosity function of star-forming galaxies. He finds that the preferred evolution in a standard cosmology is given by $L(z) = (1+z)^{2.70 \pm 0.60}$ out to a redshift as high a $z \approx 6$. At the same time he finds evidence for a very shallow density evolution given by $\rho(z) \sim (1+z)^{0.15 \pm 0.60}$, markedly different from steep density evolution $\rho(z) \sim (1+z)^{2.5}$ that we estimate for GRBs at low redshift. This would indicate that GRB luminosities have evolved at a slower rate, but that their density in the past rises much more steeply compared to the number of star forming galaxies. It could also mean that the number of GRBs per star forming galaxy has evolved rapidly with cosmic time. Perhaps more interesting is the comparison between the GRB comoving rate and luminosity densities with the global star formation rate history of the universe. Because GRBs suffer little extinction and are potentially detectable out to redshifts of $z \approx 10$, they could offer a unique tracer to the SFR history. They would allow for a more complete sampling of dust enshrouded star forming regions that may be missed in traditional SFR estimates based on the UV "drop-out" technique that is currently employed to identify Lyman break galaxies. The shape of the SFR at low redshifts $z<1$ is relatively well understood, showing an order of magnitude increase from $0\leq z \leq 1$ \citep{madau96,fall96}. These early estimates suggest that the star formation activity peaks around $z \sim 1$ followed by a rapid decline at higher redshifts. However, further observations of hundreds of Lyman break galaxies at redshifts of $z \sim 3$ and 4 have shown that the SFR may remain constant after reaching a maximum around $1 \leq z \leq 2$ \citep{Steidel99}. Recent deep surveys with the Subaru \citep{Iwata03} and Hubble Space Telescopes \citep{Bouwens03} out to $z \sim 5$ and 6 show evidence for a mild evolution of the SFR at redshifts $z>3$, with measurements based on photometric redshifts showing a constant SFR out to $z \approx 6$ \citep{Fontana03}. These recent SFR estimates qualitatively match the deduced GRB comoving rate density shown in Figure \ref{fig:comovingratedensity}. At low redshifts $z < 1$, the GRB rate density increases as $\rho(z) \sim (1+z)^{2.5}$ roughly matching the rise in the SFR over the same range, with a peak somewhere between $1\leq z\leq 2$. The following flattening and decline between $2\leq z\leq 6$ in the GRB $\rho(z)$ matches the global properties of the SFR estimated from the recent deep surveys. Of course the comparisons between the GRB comoving rate density and the SFR are simply phenomenological, since we have as of yet no way of connecting the amount of star formation for a given amount of GRBs. Ultimately this conversion factor depends on knowledge of the GRB progenitor and the initial stellar mass function (IMF) and how it changes with redshift. In the case of the collapsar model \citep{woosley00, macfadyen99}, the rate of GRBs produced for a given SFR would increase sharply with redshift, as is the case for all core collapse events, due primarily to the redshift dependence on the IMF. However, the connection between the GRB $\rho(z)$ and the SFR would be straightforward since the progenitors would consist of massive stars with short lifetimes making them direct indicators of the SFR at that redshift. If the mass range of the progenitors and the redshift dependence of the IMF is know (or assumed) then it would be possible to calculate a constant that directly relates $\rho(z)$ to the SFR. The case is more complicated for binary merger models since there would be a substantial delay between the formation of the progenitor star and the final merger event that produces the GRB. The distribution in the delay times is not well known for SNe events, much less GRBs, but it is expected to be large enough to dissociate the GRB $\rho(z)$ and the active SFR at a given redshift. The peak we observed in our deduced values for $\rho(z)$ matching the peak of the current SFR estimates hardly seems coincidental, tentatively favoring the core collapse models. It is interesting then to compare our demographic results to that of various types of supernovae. There is overwhelming observational evidence and theoretical discussion suggesting a GRBs-SNe connection, including observations of supernovae bumps in afterglow lightcurves \citep{stanek03,hjorth} and a deduced collimation corrected energy that is narrowly clustered around the typical SNe energy of $10^{51}$ ergs \citep{Frail01,bloom03b}. Although the intrinsic luminosity of type Ia SNe are \emph{a priori} \emph{assumed} to be constant with redshift (hence no luminosity evolution), we can still compare the formation rates of SNe Ia/b/c to GRBs, although the b/c events are obviously of more relevance to GRB models. Unfortunately, very little SNe data is available for the high $z$ universe with only 7 SNe at $z>1.25$ of the 42 SNe detected in the redshift range of $0.2 \leq z \leq 1.6$ by the Advanced Camera for Surveys (ACS) on the Hubble Space Telescope \citep{riess04}. Data on core collapse supernova accounts for only 17 events of this sample, going out to a maximum range of $z\approx0.7$. \citet{dahlen04} use this data to estimate the core collapse SNe (CC SNe) rate between $0.3 \leq z \leq 0.7$ and find a steep (about a factor of $\sim 7$) increase in the SNe $\rho(z\approx0.7)$ compared to the local rate presented by \citet{capp99}. Shown in Figure \ref{fig:comovingratedensity} are their data points for CC SNe plotted over the GRB comoving density, both rates normalized to 1 at $z=0.7$. The two data points, although limited, do agree with the rise of the GRB $\rho(z)$. A direct SNe-GRB comparison at higher redshifts will have to wait until the launch of the SNAP spacecraft which is predicted to find thousands of supernovae, including a significant number at high redshift. \subsection{The Nature of the Luminosity Evolution} \label{sec:lumevolution2} The observed luminosity evolution that we observed in the $(L,z)$ data leads to the conclusion that the GRB progenitor population has most likely evolved in such a way as to create more energetic or more narrowly beamed bursts in the distant past. Speculations on the nature of this evolution are dependant on the progenitor model and how the properties of their population are affected by the conditions of the early universe. In the case of highly rotating massive stars (i.e. the collapsar model), the overall progenitor mass and/or rotation rate could be the determining factor. There is ample evidence suggesting that the so called population III stars were much more massive than their present day counterparts. This is suggested by recent work showing that the stellar initial mass function (IMF) has evolved with time, having a much higher value in the distant past. This higher IMF is due to various factors, although it primarily is due to the lower metallicity in the early universe. The amount of material lost to stellar ejecta has also been shown to be dependant on the stellar metallicity, causing these early stars to retain more of their mass until their eventual collapse. Although the relationship between progenitor mass and emitted energy and/or beaming angle is not straight forward, there are reasons to think that this increase in average mass could result in an increase in the total energy budget available to a burst. \citet{macfadyen99} show that under the right conditions, the collapsar model could produce more energetic bursts with increasing stellar mass, up to some limit dictated by the energy needed by the GRB jet to punch through the stellar envelope. Proponents of black hole accretion disk models have also shown that the rate of accretion onto the central engine of the GRB increases dramatically as a function of the progenitor mass, increasing the overall energy available to the burst. Unfortunately, a simple increase in the overall energy budget cannot by itself explain the deduced luminosity evolution. \citet{Frail01} and \citet{bloom03} have recently shown observational evidence suggesting that the collimation corrected GRB energies $E_{\gamma}$ are actually narrowly cluster around the $10^{51}$ ergs typical of SNe explosions. They come to this conclusion by correcting the observed prompt isotropic equivalent energy release $E_{iso}$ of several GRBs with known redshift by a factor of $1 - \cos \theta_{j}$, where $\theta_{j}$ is the canonical jet opening angle. These angles are derived from broadband breaks observed in the afterglow light curves attributed to the slowing of the GRB jet to the point where the relativistic beaming angle of the radiation $1/\Gamma$ becomes greater then $\theta_{j}$. \citet{bloom03}, using a larger sample of bursts, show that the corrected energies cluster around $1.33\times10^{51}$ ergs with a variance of 0.35 dex, or a factor of 2.2. \citet{guetta03} has reported a similar result when correcting for the isotropic luminosity $L_{iso}$, although not as narrow as the $E_{\gamma}$ results. If the collimation corrected energy and luminosity are indeed invariant with redshift, it directly implies that the brightening of the apparent isotropic equivalent luminosity is actually due largely to an increase in the beaming factor as a function of redshift and not an increase in the overall energy of the burst. There are physical arguments that can be made in the case of the collapsar model that would suggest that more massive progenitor stars could indeed produce more collimated jet outflows. Plotted in Figure \ref{fig:fluencevsredshift} are the $E_{\gamma}$ estimates from \citet{Friedman05} for a little over two dozen GRBs. Furthermore, plotted in Figure \ref{fig:beamingangle} is the canonical jet opening angle for the same two dozen GRBs. By apply a standard Kendall rank order $\tau$ statistic we can measure the degree of correlation in these two samples in a nonparametric way (i.e. without assuming an underlying correlation type). We find a correlation strength of $\tau = 0.093$ between $E_{\gamma}$ and redshift and $\tau = 0.163606$ between $\theta_{j}$ and redshift, where a $tau$ of 1 signifies a significant correlation. Therefore, there is no deduced redshift dependency that would suggest any evolution of the jet opening angle or $E_{\gamma}$ with redshift in the pre-Swift data set. This lack of redshift dependency stands to be tested in the Swift era as more GRBs with measured jet break times are observed over a broader redshift range, but if it is confirmed then it would imply that the evolution of some jet property other than the collimation factor must be responsible for the brightening of GRBs with redshift. Speculations on the nature of this evolution are dependant on the jet model used to explain the emission. The simplest model assumes a uniform energy distribution per solid angle $\epsilon(\theta)$ across the jet with a sharp drop beyond $\theta_{j}$. In this scenario, the observed distribution in GRB energies is directly due to the diversity of jet opening angles, as is the observed values of the jet break time $t_{j}$. The lack of any concrete evidence for an evolution of $\theta_{j}$ with redshift combined with the observation that the collimation corrected $E_{\gamma}$ and $L_{\gamma}$ are very narrowly clustered, create difficulty for the uniform jet model to explain any kind of evolution in luminosity. One of the observed conditions above would have to be broken in order to accommodate any such evolution with this model. In a structured jet model, the GRB jets are identical having a quasi-universal shape with a fixed opening angle and a nonuniform energy distribution per solid angle. The diversity in the observed jet break time and isotropic energies would then be a result of varying viewing angles away from the jet axis $\theta_{v}$. Furthermore, an observer viewing the GRB at a small $\theta_{v}$ would see an extremely powerful burst, with the observed luminosity declining as some function of increasing $\theta_{v}$. A jet structure with a functional form of $\epsilon(\theta)^{-2}$ is required to reproduce the observed $t_{j} \propto E_{iso}$, i.e the \citet{Frail01} and \citet{bloom03} results. If the requirement of a narrow $E_{\gamma}$ and $L_{\gamma}$ distribution is broken, then any power law structure $\theta^{k}$ could still produce the observed steepening in the afterglow light curve. In this case, the luminosity evolution would manifest itself not as an narrowing of $\theta_{j}$ but rather as an overall increase in the normalization of $\epsilon(\theta_{j})$. Another possibility would be an evolution of the morphology of $\epsilon(theta)$ as a function of redshift. If the power law index $\epsilon(\theta_{j}) \sim \theta_{j}$ has evolved with time, or if $\epsilon(\theta_{j})$ has evolved from a non-power-law shape (e.g. a Gaussian profile), such that $\epsilon(\theta_{j})$ varies more slowly with viewing angle, then there would be a markedly different luminosity distribution at high redshift. A third, rather implausible, explanation is a preferentially small viewing angle $\theta_{v}$ at higher redshift, although there is no physical reason to think that this is at all possible. Therefore, it would seem that evidence of luminosity evolution in the presence of the observation that $E_{\gamma}$ and $L_{\gamma}$ are narrowly distributed and the lack of any evidence of an evolution of $\theta_{j}$ with redshift, favors a quasi-universal jet model over a uniform jet model. This is primarily due to the inability of the uniform jet model to explain any kind of luminosity evolution with redshift without a parallel evolution in the jet opening angle, something that is not currently observed. \subsection{Luminosity Functions and Jet Model Discrimination} \label{sec:jetmodels} Because the energy distribution $\eta(\theta_{j})$ of the structured jet model is well defined, it can make specific predictions regarding the GRB luminosity function $\phi(L)$. In the case of power-law structured jets $\epsilon(\theta_{j}) \propto \theta_{j}^{-k}$, resulting in a predicted luminosity function with a slope of $\gamma = 1 + 2/k$ \citep{zhang02}. The "canonical" $k=2$ model would yield a luminosity function $\propto L^{-2}$ \citep{rossi02}, whereas the quasi-universal gaussian structured jet model predicts $\propto L^{-1}$ \citep{lloyd04}. Although the uniform jet model also exhibits a well defined $\epsilon(\theta_{j})$, it cannot make any firm predictions about the shape of the GRB luminosity function due to the random variation $\theta_{j}$. The luminosity function deduced from our analysis is well fit by a single power low with an index of $\propto L'^{-1.5}$ for luminosities below about 10$^{52}$, with a sharp decline for higher luminosities. This value is intermediate between the expected value for the $k=2$ power law and gaussian structure models. An index of $L'^{-1.5}$ actually predicts a power law model with $k=4$ which is much steeper than the $k=2$ shape needed to keep $E_{iso}\theta^{2} \sim$ constant. It is also possible that a simple gaussian or power law profile for $\eta(\theta)$ is simply an oversimplification. It has been pointed out that by \citet{lamb03} that $\eta(\theta)$ would have to fall off steeper than $\theta^{-2}$ at large angles if the quasi-uniform jet model is to explain the relative numbers of x-ray flashes to GRBs. This may explain why so many studies have found a sharp decline above some break energy possibly indicating a different value for $k$ at low and high luminosities, i.e large and small opening angles respectively. In any case, it would not be unfeasible to think that the jet morphology is more complicated than a simple power or gaussian profile, as is suggested by our results. \section{Conclusions} \label{conclusion:ch3} In this work we perform demographic studies on a large sample of luminosity and redshift data found through the use of the lag-luminosity correlation. By applying maximum likelihood techniques, we are able to obtain an estimate of the luminosity evolution, luminosity function, and density distributions in a way that accounts for detector selection effects. We find that there exists a strong (11.63 $\sigma$) correlation between luminosity and redshift that can be parameterized as $L(z) = (1+z)^{2.58}$. The resulting cumulative $\Phi(L')$ and differential $d\Phi/dL'$ luminosity functions are well fit by double power laws separated by a break energy of about $10^{52}$ ergs s$^{-1}$, with $d\Phi/dL'$ exhibiting a power law shape of $L^{-1.5}$ below this luminosity. This value does not immediately discriminate against any proposed structured jet models, but it may indicate that a more complicated jet profile is need to explain the observed luminosity function of GRBs. The GRB comoving rate density is found to increase as $\rho_{\gamma}(z) \propto (1+z)^{2.5}$ to a redshift of $z \approx 1$ followed by flattening and eventual decline above $z>6$. Although, the conversion between $\rho_{\gamma}(z)$ and an estimate of the SFR cannot be quantitatively made due to the uncertainty regarding the GRB progenitors and their initial stellar mass functions, it can be said that $\rho_{\gamma}(z)$ does qualitatively follow recent photometric estimates of the SFR, as would be expected from massive short lived progenitors. We stress that these conclusions are based on the validity of the lag-luminosity correlation, which still stands to be confirmed as new redshift data becomes available. A full confirmation, and most probably further calibration, of this distance indicator will have to wait for addition detections with the Swift spacecraft, which should have the collecting area necessary to obtain high signal-to-noise energy dependant light curves from which to measure statistically significant lags. Even with the slew of directly measured luminosity and redshift data expected to come from the Swift mission, empirical distance indicators may still play an important roll in expanding the available GRB data set through the use of archival BATSE and BepooSAX data for statistical studies such as the analysis performed in this work. \bigskip \section*{Figure Captions} \bigskip {\bf Fig. 1.} - The luminosity and redshift data used in our analysis as deduced from the lag-luminosity correlation. The dashed line represents an imposed 0.5 photons cm$^{-2}$ s$^{-1}\ldots$ cut to the original 1438 bursts analyzed by BNB04, producing the sharp cutoff in the data. This leaves a total of 985 bursts with a median redshift of 1.64. \bigskip {\bf Fig. 2.} - A plot of the lag-luminosity plane for the events under consideration. The error in the lag and flux measurements are used to determine the uncertainty in the luminosity values which in effect is used as a weight in the maximum likelihood technique that estimates the correlation strength between $L$ and $z$. \bigskip {\bf Fig. 3.} - A representation of the associated sets used in the Lynden-Bell technique. For each data point with $(L_{i},z_{i})$, the solid line represents the minimum luminosity or maximum redshift that the burst could have had and still have been observed. Employing the Lynden-Bell technique with associated sets defined as $N_{j}$ ($L_{j} < L < \infty$, $0 < z < z_{max}(L_{i})$) produces the cumulative distribution for the vertical axis, whereas the $M_{j}$ associated set produces the distribution for the data represented on the horizontal axis. \bigskip {\bf Fig. 4.} - Generated luminosity and redshift data used to test the ability of the Efron $\&$ Petrosian method to estimate the correlation strength and functional dependence for data of a given flux cut. \bigskip {\bf Fig. 5.} - A histogram of the difference between the known correlation index and the reconstructed index $(p-q)$ for a large set of generated $(L,z)$ data with arbitrarily imposed flux thresholds. The error in the method is tightly centered around $p-q=0$. \bigskip {\bf Fig. 6.} - The correlation statistic $\tau$ is plotted as a function of power law index $g_{\alpha}(z)=(1+z)^{\alpha}$ parameterizing the luminosity evolution. The solid line represents the $\alpha$ index that minimizes the correlation between $L'$ and $z$. The dotted lines show the 1 $\sigma$ \emph{statistical} error in the $\alpha$ parameter. \bigskip {\bf Fig. 7.} - The correlation parameter $\alpha$ is plotted as a function of the flux cut applied to the BNB04 sample. The optimal $\alpha$ is highly dependent on the choice of the cut value, showing the importance of a good understanding of the detector flux threshold. \bigskip {\bf Fig. 8.} - The luminosity and redshift data used in our analysis as deduced from the lag-luminosity correlation. The dashed line represents an imposed 0.5 photons cm$^{-2}$ s$^{-1}\ldots$ cut to the original 1438 bursts analyzed by BNB04, producing the sharp cutoff in the data. This leaves a total of 985 bursts with a median redshift of 1.64. \bigskip {\bf Fig. 9.} - The present epoch GRB luminosity function $\psi(L') = d\Phi/dL'$, representing the number of events between the luminosity $L'$ and $L'+dL'$. The function falls as roughly $\propto L'^{-1.5}$ for luminosities below the break energy of $10^{52}$ ergs s$^{-1}$. \bigskip {\bf Fig. 10.} - The cumulative density distribution $\sigma(z)=\int_{0}^{z}\rho(z)(dV/dz)dz$, representing the total number of events up to a given redshift. The flattening at high redshift indicates a drop off of events around $z \sim 5-10$. \bigskip {\bf Fig. 11.} - The cumulative density function $\sigma(z)$ plotted as a function of comoving volume. The dashed line represents the increase in the number of sources if the GRB density were constant throughout the history of the universe. The GRB density has increased as $\rho \sim V^{0.25}$ before peaking between $z \sim 1-2$ and then decreasing. \bigskip {\bf Fig. 12.} - The comoving rate density $\rho(z)$ as a function of redshift. The rate density of sources can be seen to follow the evolution deduced from Figure \ref{fig:cumdensityvolume}, increasing to a redshift of 1-2 then flattening before decreasing at higher redshifts. The circles represent high redshift cc rates from \citet{dahlen04} whereas the square point is the local rate found \citet{capp99}. The increase in the cc event rate with redshift qualitatively matches the overall increase in the GRB comoving rate density. \bigskip {\bf Fig. 13.} - Plot of the collimation corrected total emitted energy of 23 GRBs with known redshift and beaming angles. No significant correlation can be seen with redshift. \bigskip {\bf Fig. 14.} - The beaming angles $\theta_{j}$ of 23 GRBs with known redshift. The lack of a significant correlation with redshift is quite evident. \bigskip \section*{Figures}
Title: Low radiative efficiency accretion at work in active galactic nuclei: the nuclear spectral energy distribution of NGC4565
Abstract: We derive the spectral energy distribution (SED) of the nucleus of the Seyfert galaxy NGC4565. Despite its classification as a Seyfert2, the nuclear source is substantially unabsorbed. The absorption we find from Chandra data (N_H=2.5 X 10^21 cm^-2) is consistent with that produced by material in the galactic disk of the host galaxy. HST images show a nuclear unresolved source in all of the available observations, from the near-IR H band to the optical U band. The SED is completely different from that of Seyfert galaxies and QSO, as it appears basically ``flat'' in the IR-optical region, with a small drop-off in the U-band. The location of the object in diagnostic planes for low luminosity AGNs excludes a jet origin for the optical nucleus, and its extremely low Eddington ratio L_o/L_Edd indicates that the radiation we observe is most likely produced in a radiatively inefficient accretion flow (RIAF). This would make NGC4565 the first AGN in which an ADAF-like process is identified in the optical. We find that the relatively high [OIII] flux observed from the ground cannot be all produced in the nucleus. Therefore, an extended NLR must exist in this object. This may be interpreted in the framework of two different scenarios: i) the radiation from ADAFs is sufficient to give rise to high ionization emission-line regions through photoionization, or ii) the nuclear source has recently ``turned-off'', switching from a high-efficiency accretion regime to the present low-efficiency state.
https://export.arxiv.org/pdf/astro-ph/0601629
command. \newcommand{\myemail}{chiab@stsci.edu} \slugcomment{Submitted to ApJ} \shorttitle{Low efficiency accretion in NGC~4565} \shortauthors{Chiaberge et al.} \begin{document} \title{Low radiative efficiency accretion at work in active galactic nuclei: the nuclear spectral energy distribution of NGC~4565} \author{M. Chiaberge\altaffilmark{1}} \affil{Space Telescope Science Institute, 3700 San Martin Dr., Baltimore, MD 21218} \email{chiab@stsci.edu} \author{R. Gilli\altaffilmark{2}} \affil{INAF - Osservatorio astronomico di Bologna, Via Ranzani 1, 40127 Bologna, Italy} \author{F. D. Macchetto\altaffilmark{3}} \affil{Space Telescope Science Institute, 3700 San Martin Dr., Baltimore, MD 21218} \and \author{William B. Sparks} \affil{Space Telescope Science Institute, 3700 San Martin Dr., Baltimore, MD 21218} \altaffiltext{1}{On leave from INAF, Istituto di Radioastronomia, via P. Gobetti 101, I-40129 Bologna} \altaffiltext{2}{Visiting Programmer, Space Telescope Science Institute} \altaffiltext{3}{On assignment from ESA} \keywords{galaxies: active --- galaxies: nuclei --- accretion, accretion disks --- galaxies: individual (NGC~4565)} \section{Introduction} Low luminosity active galactic nuclei (LLAGN) are believed to be powered by accretion of matter onto the central supermassive black hole, similarly to powerful AGN. In a large fraction of LLAGN, the central black hole is as massive as in powerful distant quasars ($M_{BH} \sim 10^{8}- 10^{9} M_\sun$), thus their very low nuclear luminosity implies that accretion occurs with very low radiative efficiency \citep[or at very low rates; see e.g.][]{ho04coz,papllagn}. If so, the physics of the accretion process may be different from the ``standard'' optically thick, geometrically thin accretion disks. Starting from the ``ion-supported tori'' of \citet{rees82}, a number of theoretical models have been developed to describe such radiatively inefficient accretion flows \citep[RIAF, e.g. advection-dominated accretion flows, ADAF, advection-dominated inflow-outflow solutions, ADIOS, convection-dominated accretion flows, CDAF][]{narayanyi94,quataert99,abramowicz02}. But because of the very low radiation they emit at all wavelengths, these objects (if they at all exist) are very difficult to be observed. Recently, the AGN nature of optical nuclear components seen in HST images of a sample of LLAGN have been unambiguously established. \citet{Maoz05} have shown that among a sample of 17 LLAGN, 15 of them show variability over a timescale of a few months, which demonstrate their non--stellar origin. However, it is still unclear whether the radiation is from a jet or from the accretion flow. LLAGN have also been found to lie on the so-called ``fundamental plane of black hole activity'' \citep{merloni03,falcke04}, which attempts to unify the emission from all sources around black holes, over a large range of masses and luminosities, from Galactic sources to powerful quasars. But the origin of such a ``fundamental plane'' and its relationship with the origin of the radiation is still a matter of debate \cite[e.g][]{bregman05,koerding06}. RIAF models have been applied to several sources belonging to different classes, such as low-luminosity radio galaxies, ``normal'' ellipticals, the Galactic center Sagittarius~A \citep[e.g.][]{quataert99,dimatteo00,dimatteo03}. However, for most of these objects, the models cannot be properly constrained, mostly because the nuclear radiation is swamped by other processes. This is particularly problematic in the optical band, which appears to be crucial to fix the models, where the stellar emission from the host galaxy is substantial. It is indeed in the IR-to-UV region that different accretion disk models are expected to show the largest difference in spectral shape. RIAFs should lack both the "big blue-bump" and the IR (reprocessed) bump, which instead characterize optically thick, geometrically thin accretion disk emission and the surrounding heated dust. For example, in low luminosity radio galaxies non-thermal emission from the jet dominates the optical nuclear radiation \citep{pap1}, while the Galactic center is not visible in the optical because it is hidden by a large amount of dust. Therefore, neither of the above mentioned classes of objects appear to be suitable laboratories to test models of low-efficiency accretion through the analysis of their overall SED. Recently, among LLAGN which show very low Eddington ratios $L_{bol}/L_{Edd} << 10^{-3}$, \citet{papllagn} have found that a class of LLAGN, mainly composed by LINERs and low-luminosity Seyfert~1 galaxies, show faint optical unresolved nuclei in HST images that may be interpreted as direct radiation from a very low efficiency accretion flow. In fact, when the radio-optical properties of LLAGN are considered, Seyfert, LINERs and low luminosity radio galaxies separate into different regions of diagnostic planes, according to the properties of their nuclei. If this interpretation is correct, we now have a powerful tool to identify the nature of the nuclear radiation (i.e. jet-dominated or accretion-dominated). The best possibility of detecting radiation from an ADAF-like process would then be to study unobscured Seyferts of lowest luminosity, as well as a sub-class of LINERs. In all other objects other radiation processes dominate. In this paper we further test the picture outlined in \citet{papllagn} by studying the nuclear spectral energy distribution of a galaxy, \object{NGC~4565}, that seems to be a perfect candidate for hosting a RIAF around the central supermassive black hole. The object is part of the ``Palomar sample'' of LLAGN \citep{ho97}, and it is included in both the \citet{merloni03} and \citet{falcke04} samples that were used to define the ``fundamental plane of black hole activity''. It is worth mentioning that NGC~4565 does not show any significant peculiarity in that plane. NGC~4565 is a nearby (d=9.7 Mpc) LLAGN classified as a Seyfert~1.9 because of the possible presence of a faint, relatively broad (FWHM = 1750 km s$^{-1}$) H$\alpha$ line. However, as \citet{ho97broad} have pointed out, the detection of a broad component is highly uncertain. As we show in the following, although it is a Type 2 Seyfert, this object is only moderately absorbed, and the nuclear radiation is visible in the optical spectral region. NGC~4565 may thus represent the first clear example of low-luminosity accretion onto a supermassive black hole in the optical band. In Section~\ref{obs} we describe the {\it Chandra} and {\it HST} observations, the data analysis procedures and flux measurements, in Section~\ref{discussion} we present the results, we derive the spectral energy distribution and we discuss its interpretation. In Section~\ref{conclusions} we give a summary of our findings and we draw conclusions. \section{Observations and data analysis} \label{obs} We use X-ray data taken with \facility{Chandra} satellite, and IR through optical \facility{HST} images. In the following we describe the data and the analysis procedures. \subsection{{\it Chandra} data} A 60 ksec ACIS-S observation of NGC~4565 (performed in 2003, PI D. Wang) is publicly available in the {\it Chandra} archive. We retrieve the {\it Chandra} data and analyze them using standard CIAO 3.2.2 procedures, applying the latest calibration files in the CALDB 3.1.0 database. The X-ray image reveals a wealth of pointlike sources, many of which located along the NGC~4565 disk. The two brightest sources correspond to an off-nuclear source at $\sim 50$ arcsec from the nucleus, and to the nucleus itself \citep[see also the XMM image in][]{foschini02}. To avoid contamination from faint nearby X-ray sources, the 0.4-7 keV nuclear spectrum is extracted in a circular aperture of 6 pixel radius ($\sim 3$ arcsec, corresponding to an encircled energy fraction of $>97\%$ at 1.5 keV). The background is evaluated in a large annulus around the nucleus. Faint X-ray sources are not masked out from the background region, since their presence has a negligible impact on our results (the total background flux including faint sources is less than $1\%$ of the nuclear flux). Given the moderate nuclear count rate (0.036 counts/sec), X-ray photon pileup is under control ($\lesssim 4\%$). Spectral analysis is carried out with XSPEC v11.3.1, with the column density of our Galaxy fixed to $1.3\:10^{20}$ cm$^{-2}$ \citep{dickey90}. The spectrum is re-binned to have at least 20 photons per bin to allow use of the $\chi^2$ statistics, errors are quoted at the $90\%$ c.l. for one interesting parameter. We find that an absorbed power law model provides a very good description of the data ($\chi^2/dof=51/81$), the best fit photon index and column density being $\Gamma=1.91^{+0.22}_{-0.19}$ and $N_H=2.5\pm0.6\;10^{21}$ cm$^{-2}$, respectively. The observed source fluxes in the 0.5-2 keV (soft) and 2-10 keV (hard) bands are $9.0\;10^{-14}$ erg cm$^{-2}$ s$^{-1}$ and $2.1\;10^{-13}$ erg cm$^{-2}$ s$^{-1}$, respectively. When corrected for absorption, these correspond to intrinsic nuclear luminosities of $1.9\;10^{39}$ erg s$^{-1}$ and $2.5\;10^{39}$ erg s$^{-1}$ in the soft and hard band, respectively. We note that the derived X-ray spectral parameters, fluxes and luminosities are in good agreement with those measured in a 14 ksec XMM-$Newton$ observation performed in 2001 \citep{cappi06}. \subsection{{\it HST} data} \begin{deluxetable}{l r c c c} \tabletypesize{\scriptsize} \tablewidth{0pt} \tablecaption{Nuclear fluxes from HST observations \label{fluxes}} \tablehead{\colhead{Instrument/Filter} & \colhead{T$_{exp}$} & \colhead{Program ID} & \colhead{Wavelength} & \colhead{$F_\lambda$} \\ \colhead{} & \colhead{[s]} & \colhead{} & \colhead{[\AA]} & \colhead{ }\\} \startdata ACS-HRC/F330W & 1200 & 9379 & 3367 & 8.1 \\ WFPC2/F450W & 600 & 6092 & 4575 & 18 \\ WFPC2/F555W & 320 & 6685 & 5468 & 21 \\ WFPC2/F814W & 480 & 6092 & 8023 & 11 \\ NICMOS/F160W & 384 & 7331 & 16074 & 5.9 \\ \enddata \tablecomments{Fluxes (in units of $10^{-17}$ erg cm$^{-2}$s$^{-1}$ \AA$^{-1}$) have been corrected for local extinction using $N_H=2.5\times 10^{21} cm^{-2}$ and standard $A_V/N_H = 5 \times 10^{-22}$ ratio.} \end{deluxetable} HST data are available in the MAST archive at STScI from the near IR to the optical U-band. Images were taken as part of different programs, with the following instruments and filters: NICMOS (F160W), WFPC2 (F814W, F555W, F450W), ACS/HRC (F330W). These filters approximate the H,I,V,B and U bands in the HST system. The images are processed with the standard on-the-fly reprocessing calibration pipeline \citep[see][]{acshandbook}. The optical images show the bulge of the galaxy partially covered by a prominent dust lane or disk seen almost edge-on (Fig. \ref{galaxy}) The inclination of the ``disk'' is such that the central region of the bulge is not covered by a large amount of dust, and a faint nuclear compact source (to which we refer as the {\it nucleus}) is visible in all images. \subsubsection{Nuclear photometry} In the U-band, the emission from the bulge stars is low, and the nucleus is the by far the brightest source in the field of view of the ACS/HRC (Fig. \ref{profiles}). Photometry of the nucleus is thus straightforward in the U-band, also thanks to the higher resolution, smaller projected pixel-size of the HRC. On the other hand, in the IR (NICMOS) and optical I,V and B band WFPC2 images (the target is always located in one of the WF cameras) the contrast with the underlying stellar background is low, thus the measurement of the nuclear flux is more problematic. For the photometry of nuclear unresolved sources superimposed to the stellar emission of the host galaxy, we undertake two different approaches, as described in the following. 1) Aperture photometry with the {\it IRAF } task {\it radprof}, measuring the background close to the unresolved nucleus, at a distance of $\sim 0.4^{\prime\prime}$ from the center of the point source, and setting the aperture radius at the same distance. Note that the ``background'' here is the stellar emission of the galaxy in the vicinity of the nucleus. Therefore, this approach works well for nuclei in elliptical galaxies with flat radial brightness profiles, i.e. Nuker-law ``core'' galaxies \citep{faber97} which have a flat ($\gamma < 0.3$, $\Sigma \propto r^{-\gamma}$, where $\Sigma$ is the surface brightness) slope in the inner region \citep[see also the discussion in][]{barbara}. Clearly, this is because in this case the ``background'' measured at a distance of $\sim 0.4^{\prime\prime}$ is a good estimate of the stellar emission at the center of the nucleus. On the other hand, for both ``power-law'' ellipticals and spirals bulges, the profile in the innermost regions (i.e. in the central 1-2 arcsec) is significantly steeper ($\gamma \sim 0.8$). In this latter case, the measurement and even the identification of faint nuclei is more difficult, because a ``peaked'' brightness profile of the bulge may hamper the detection of the central emission from the AGN. Furthermore, for the IR images, which have a lower angular resolution, the background cannot be measured close enough to the center of the nucleus, and thus may be significantly underestimated. We find that this would lead to overestimate the nuclear flux by a factor as large as $\sim 5$. Thus, while we used this method to measure the nuclear flux in the F330W image \citep[aperture correction was taken into account, following the prescriptions given in][]{sirianniacs}, we had to adopt a different strategy for the WFPC2 and NICMOS images. 2) An alternative approach consists in deriving the radial brightness profile of the galaxy, and measure any nuclear excess. Multiple component models are often used to reproduce the central regions of galaxies and measure the flux of nuclear sources \citep[see e.g. the discussion in][]{quillen01}. But since here we are not interested in modeling the galaxy bulge on large scales, we only derive the radial brightness profile in the central $\sim 2$ arcsec. Then we produce a model galaxy with the same slope as observed in the region $R > 0.2$ arcsec and we assume that the profile can be extrapolated to the center of the bulge, all the way to $R=0$. As shown in Fig. \ref{profiles} (solid line), the effect of the finite resolution of HST ($\sim 0.1$ arcsec) produces a flattening of the observed profile, at a distance of $\sim 0.15$ arcsec. The observed profile, obtained by fitting ellipses to the galaxy image using the {\it IRAF} task {\it ellipse}, shows a significant excess (filled circles). We produce synthetic PSF's using the {\it TinyTim} software \citep{krist}, which, for WFCPC2, produces an accurate representation of the central region of the PSF, thus appropriate for our purpose. We align the synthetic PSF's with the position of the nucleus, we multiply the PSF image by an appropriate constant $K$ and we subtract the two images. We change the value of $K$ until the profile of the nuclear regions do not produce a ``hole'' at the center of the galaxy. The obtained profile is shown in Fig. \ref{profiles} as the empty circles. To convert the flux of the nucleus from counts to physical units, we multiply the count rate ($CR = K/t_{\rm exp}$) by the keyword {\it PHOTFLAM} in the image header (for NICMOS $CR=K$). In Fig.~\ref{profiles} we also show the radial profiles obtained by subtracting PSFs that are 20\% brighter and 20\% fainter than our reference value. It is not straightforward to estimate the error on the flux measurements obtained using method 2, because the main uncertainty is the assumption that the radial profile of the galaxy can be extrapolated all the way to the center, at R=0. After subtracting synthetic PSFs with different total counts and comparing the resulting profiles with our model profile, we prefer to adopt a rather conservative value of $\sim 20\%$ for the error of the IR and optical fluxes. With future observations, which should be taken using a dithering strategy aimed at improving the PSF sampling, the error could be significantly reduced. The statistical error on the F330W flux (obtained with method 1) is 7\%. A summary of the photometry is given in Table~\ref{fluxes}. The F450W and F555W filter pass bands include relatively strong emission lines (mainly [OIII]5007 and H$\beta$). However, since the pass bands are $\sim 1000$\AA~~ wide, the observed flux is likely to be dominated by continuum emission (see also Sect. \ref{discussion}). Since the near-IR and the U-band images were not taken simultaneously to the optical data, the SED may be affected by variability. We checked for variability of the nuclear source in the optical (F814W and F450W), for which two sets of observations with the same filters, taken at a distance of $\sim 1$ year, are available. The nuclear fluxes are consistent within the errors, thus no variability is found between the 2 observations. However, our estimated $20\%$ error on the optical fluxes does not allow us to exclude variations of smaller amplitude, as observed in other LLAGN \citep{Maoz05}. \section{Results and discussion} \label{discussion} \subsection{The nuclear SED of NGC~4565} The absorption corrected nuclear spectral energy distribution is shown in Fig.~\ref{sed}. The HST data are de-reddened using $N_H = 2.5 \times 10^{21}$ cm$^{-2}$, which converts to $A_V = 1.25$, assuming Galactic gas-to-dust ratio. Although in AGN the gas-to-dust ratio may differ from the local value, we believe that this choice is justified in the case of NGC~4565. As it is clear from the large field of view image of the galaxy (Fig.~\ref{galaxy}), this is a spiral seen almost edge on. Therefore, it is reasonable to assume that a significant amount of dust and gas in the disk of the galaxy project on our line-of-sight to the nucleus. Assuming a circular geometry, from the observed ellipticity of the disk in the image we find that the orientation of the disk is likely not to exceed $\sim 10^\circ$. For comparison, we can check the absorption we find in our Galaxy for $10^{\circ}$ Galactic Latitude. We obtain $A_V \sim 1.0 - 1.5$, where the lower value is found for Galactic longitude $\sim 180^\circ$, the higher value is for $\sim 0^\circ$ (from NED). These values may actually increase substantially if we observe the same Galactic Latitude from the Galaxy center. This simple check shows that the absorption we measure in the X-rays is compatible with that provided by galactic dust in the disk. This supports our hypothesis that the moderate absorption observed to the nucleus of NGC~4565 is not produced locally, in the vicinity of the nucleus. In this case, Galactic dust-to-gas ratio may be used to convert $N_H$ derived from the X-rays to optical $A_V$. The nuclear SED appears basically flat ($\alpha \sim 1$, $F_\nu \propto \nu^{-\alpha}$) from the 1.6$\mu$m to 4500 \AA, with possibly a small peak between 5000 and 4000 \AA~~ and a small drop-off in the U-band. This peak may be real, or due to a possible contamination from emission lines (mainly [OIII] and H$\beta$) that fall in the F555W and F450W filters pass bands. Unfortunately, since neither images with narrow-band filter nor nuclear spectra are available to date, a certain ambiguity persists. However, all other filters are free from strong lines, thus the intrinsic SED cannot be dramatically different from what we show here. Whatever the nature of such a small peak, it is clear that neither a significant UV bump nor IR thermal emission from hot dust, which are characteristic of AGNs, are visible in NGC~4565. Furthermore, note that the luminosity in the X-ray is not higher than in the optical, even after absorption has been taken into account (see Fig.~\ref{xspec}). For comparison, in Fig.~\ref{sed} we show the nuclear SED of two specific objects, together with the average SED of radio-quiet QSO as taken from \citet{elvis94}. The two objects are a Seyfert~1 galaxy (\object{NGC~3516}) and of a Seyfert~2 (\object{Fairall~49}), for which the absorption, estimated from X-ray observations, is $N_H = 1.4 \times 10^{22}$ \citep{iwasawa04}. These two Seyferts have similar bolometric luminosity, but they are both clearly more powerful than NGC~4565, by $\sim 3$ orders of magnitude. The Type~1 object clearly shows a concave spectrum, which is interpreted as the signature of the presence of the blue bump in the UV and of dust heated by the central AGN in the near-IR. The Type~2 galaxy, instead, is very bright in the IR, while the flux is dramatically reduced for higher (optical) frequencies, as a result of absorption.\footnote{It might seem surprising to detect the nucleus in Fairall~49, which is a Seyfert~2 absorbed by a large amount of $N_H$ in the X-rays. In fact, if converted to $A_V$ using standard Galactic dust-to-gas ratio, this would correspond to 7 mag extinction in V and 11.5 mag in the F330W filter. Two possible explanations have been proposed: i) part of the optical nuclear flux in Fairall~49, if not all, might not be radiation from the accretion disk seen directly (through a moderate amount of dust). Instead, the nucleus might be in part (or completely) obscured in the optical-to-UV band, and the bulk of the observed emission may be scattered light; ii) the properties of the circumnuclear absorber are different from the Galactic dust and this would result in a non-standard $A_V/N_H$ ratio. This has been discussed by various authors \citep[e.g.][]{granato97,maiolino01}.} Not all Seyferts show such a clear behavior, but these objects serve as good examples to be compared with the peculiar SED of NGC~4565. \subsection{A low radiative efficiency accretion disk} How do we interpret the nuclear emission in NGC~4565? As pointed out in the introduction, the source is included in the sample of both \citet{merloni03} and \citet{falcke04}. The object does not appear to show any peculiarity, and its location along the ``fundamental plane of black hole activity'' does not provide us with information on the emission process. Therefore, in order to answer the above question, we use diagnostics introduced by \citet{papllagn}. As already mentioned in Section~1, we found that the radio to optical nuclear luminosity ratio (i.e. the ``nuclear radio-loudness'') $L_r/L_o$ for LLAGNs gives us important information on the nature of the source. In particular, we can infer whether we are observing synchrotron emission in both bands (if $\log (L_r/L_o)\sim 3$), or in the optical we have some kind of excess radiation which, in the case of unabsorbed Seyferts is most likely interpreted as radiation from the accretion process. Furthermore, when the optical luminosity to Eddington luminosity ratio $L_o/L_{Edd}$ is plotted against the nuclear radio loudness (Fig. \ref{r_edd}) different classes of LLAGNs nicely separate into three different regions of the diagram. Seyfert nuclei with relatively high accretion efficiency objects occupy the top-left part of the diagram ($L_o/L_{Edd} \sim 10^{-2}-10^{-3}$ and $\log(L_r/L_o) \sim 1$); LINERs separate into two subclasses, which we named according to their nuclear radio-optical ratio as ``radio--quiet'' LINERs (bottom-left side, $L_o/L_{Edd} < 10^{-4}$ and $\log(L_r/L_o) \sim 1$), and ``radio--loud'' LINERS (bottom-right); radio galaxies (bottom-right part of the diagram) have the same Eddington ratio as for radio--quiet LINERs, but a much higher $\log(L_r/L_o)$. Note that in the plane of Fig. \ref{r_edd}, the objects in which we observe an extra-component in the optical, in excess of synchrotron emission, are those that lie on the left side. Therefore, those are the objects in which emission from the accretion process can be detected. In fact, in perfect agreement with the shape of its SED, NGC~3516 lies in top-left panel of the plane, where relatively high efficiency accretion disks are. On the other hand, we did not mark the location of the Seyfert~2 galaxy Fairall~49 on the plot of Fig.~\ref{r_edd}, since the nuclear emission is most likely to be affected by significant obscuration. However, we point out that assuming that the IR flux can be used as instead of the optical, the radio to IR ratio would be similar to the Seyfert 1s in the plot (at $\log L_r/L_o = 0.88$). Let us explore how this applies to NGC~4565. First of all, we calculate the ratio between the nuclear radio flux and the optical flux. \citet{neil05} measured a radio core flux of 3.2 mJy, which implies that $\log(L_r/L_o) = 1.6$. Therefore, NGC~4565 has an excess in the optical of at least 2 dex with respect to the expected synchrotron emission (i.e. the optical counterpart of the radio core should be $> 2$ dex fainter than the measured optical flux, unless the radio-to optical spectral index has unreasonable values for synchrotron radiation). Thus it is reasonable to interpret the optical nucleus as radiation from the accretion process. In this case, assuming a central black hole mass of $2.8 \times 10^{7} M_\sun$, as derived from the M-$\sigma$ relation of \citet{tremaine02}, and using $\sigma$ value from the LEDA database\footnote{http://leda.univ-lyon1.fr/}, the resulting Eddington ratio $L_o/L_{Edd}$ is extremely low, $2 \times 10^{-6}$. It is important to note that in the case of ``typical'' Seyferts, which show a blue bump, a significant bolometric correction is needed (however, this should not exceed a factor of $\sim 15$). For NGC~4565 a big blue bump is clearly not present, thus our value of $L_o$ is a good estimate of $L_{bol}$. In the diagnostic diagram of Fig.~\ref{r_edd}, NGC~4565 lies in the lower-right quadrant, among ``radio-quiet'' LINERs \citep[the other two Seyferts in the same region of the plot are M~81 and NGC~4639, as discussed in][]{papllagn}. In order to reconcile NGC~4565 with other Seyferts, which are confined in the top-left quadrant, the central black hole mass would have to be at least a factor of $\sim 100$ lower. This would substantially violate the $M_{BH} - \sigma$ relation. We conclude that the nucleus of NGC~4565 is a very low-efficiency accretion object and that we are observing the accretion process directly in the optical. This is extremely important since models of advection-dominated accretion flows are particularly sensitive to the optical-UV spectral region. For example, as shown by \citet{quataert99} the presence of winds in the disk can dramatically change the shape of the observed SED in the range of frequencies between $\nu \sim 10^{13}$ Hz and $\nu \sim 10^{15}$ Hz. However, a detailed comparison with a set of models is beyond the scope of this paper. RIAF models are very sensitive to many different physical parameters, therefore detailed modeling is useful only if the SED is well determined, from the radio-mm to the X-ray band and, if possible, when simultaneous data are available. A similar study of the nuclear emission has been performed by \citet{moran99} for NGC~4395, ``the least luminous Seyfert~1''. In that case, the nuclear luminosity is even lower than in NGC~4565, but the central black hole mass in NGC~4395 is dramatically lower. A recent estimate based on reverberation mapping gives a value of $M_{BH} = 3.6 \times 10^{5} M_\sun$ \citep{peterson05}. Such a low black hole mass implies $L_{bol}/L_{Edd} \sim 10^{-3}$ or higher if, as \citet{moran99} point out, intrinsic nuclear absorption is present. This seems in fact to be the case since \citep{moran05} obtain $N_H \sim 10^{22}$ cm$^{-2}$ analyzing {\it Chandra} data, although most of the absorption might be produced by ionized gas. Therefore, although it is clear that even if NGC~4395 displays peculiar characteristics, its Eddington ratio is not different from the average of low luminosity Seyferts in the Palomar and CfA samples \citep{papllagn,ho04coz}. Instead, NGC~4565 has completely different physical properties, as appears from its extremely low value of the Eddington ratio, and it is a perfect candidate for hosting a radiative inefficient accretion process. \subsection{Where is the narrow line region in NGC~4565?} One further implication of the observations we present here is worth mentioning. The flux of the [OIII]5007 emission line ($F_{[OIII]} = 1.5 \times 10^{-14}$ erg s$^{-1}$ cm$^{-2}$), as measured from the ground with a 2'' beam size \citep{ho97}, and the diagnostic line ratios are typical of Seyfert galaxies. Even considering the interesting (although only slightly different) classification scheme for LLAGN proposed by \citet{kewley06} based on SDSS data, NGC~4565 still falls in the region occupied by Seyferts. However, if we assume that all of the [OIII] flux is produced in the unresolved nucleus, this would result in a count rate higher by factor of 5 and 20 than we measure in the nucleus in the F555W and F450W, respectively. This implies that the narrow emission line region (NLR) must be extended. It is particularly interesting to investigate the properties of the NLR relatively to the nuclear properties, since it is sometimes assumed that radiative inefficient accretion cannot provide a sufficient photon field to ionize the surrounding medium and create the NLR. If this is true, we can speculate that NGC~4565 may have recently transitioned from a relatively high-efficiency accretion state (as in ``normal'' Seyferts) to a very low-efficiency accretion process. This might also reconcile its classification as Seyfert with the fact that its nucleus is located among LINERs in the plane of Fig.~\ref{r_edd}. The spectral classification is in fact based on the large-scale properties of the emission-line gas, that may still be powered by a higher radiation field (possibly having also a different spectral shape), because of light travel time effects. However, the equivalent width of the [OIII] line, as measured with respect to the nuclear continuum emission, is $EW_{[OIII]} \sim 100$\AA. This value is in the range normally spanned by Type~1 AGN \citep[see e.g.][]{kinney91,pap4,marziani03}. Therefore, this may indicate that the continuum emission from RIAFs is sufficient to account for the observed [OIII] flux, and Seyferts' NLR can be powered by radiatively inefficient accretion flows. However, in the scenario in which the accretion disk has changed its state, $EW_{[OIII]}$ may assume a low value if only a fraction of the NLR gas is still highly ionized under the effect of the ``past'' high--efficiency state of the accretion. Clearly, only high spatial-resolution images with narrow band filters, and nuclear spectra, can provide further information to understand the recent history and ionizing mechanism of both the nucleus and NLR of NGC~4565. \section{Summary and conclusions} \label{conclusions} We have derived the spectral energy distribution of a peculiar low-luminosity Seyfert~2 galaxy which, despite its spectral classification, basically shows no evidence for local nuclear absorption. The SED is peculiar, as it is almost flat in a $\log \nu - log (\nu F\nu)$ representation, with no sign of both a UV bump and thermally reprocessed IR emission. The very low luminosity of the source associated with a relatively high central black hole mass imply an extremely small value of the Eddington ratio ($L_o/L_{Edd} \sim 10^{-6}$). This, together with the position occupied by this object on diagnostic planes for low luminosity AGN, represents clear evidence for a low radiative efficiency accretion process at work in the innermost regions of NGC~4565. The direct detection of optical emission from such radiative inefficient processes is particularly important for providing constraints to ADAF models or similar. NGC~4565 is therefore a perfect candidate for studying RIAFs, and more observations aimed at achieving a complete coverage of the SED, from the radio-mm to the X-ray bands, should be taken in order to test the models. As part of a ``search for RIAFs in LLAGN'', it would also be extremely useful to derive the SED of objects that are located in the same region of the diagnostic planes as NGC~4565. The fact that the [OIII] emission line flux is substantial in this object implies that an extended narrow line region, similar to other Seyfert galaxies, is still present in NGC~4565. A possible intriguing scenario is that the active nucleus has recently ``turned-off'', switching from a high efficiency, standard, accretion disk, to a radiative inefficient accretion process. However, since the EW of the [OIII]5007 emission line is rather small, with the present data we cannot rule out that the amount of ionizing photons from the RIAF is sufficient to produce the observed [OIII] flux. \acknowledgments We thank the anonymous referee for her/his comments that greatly improved the paper. We acknowledge Dave Axon, Alessandro Capetti and Alice Quillen for useful comments. RG acknowledges support from the STScI Visitor Program. This research has made use of the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. {\it Facilities:} \facility{HST}, \facility{Chandra}.
Title: Plane waves in metric-affine gravity
Abstract: We describe plane-fronted waves in the Yang-Mills type quadratic metric-affine theory of gravity. The torsion and the nonmetricity are both nontrivial, and they do not belong to the triplet ansatz.
https://export.arxiv.org/pdf/gr-qc/0601074
\title{Plane waves in metric-affine gravity} \author{Yuri N.~Obukhov\footnote{On leave from: Dept. of Theoret. Physics, Moscow State University, 117234 Moscow, Russia}} \address{Institute for Theoretical Physics, University of Cologne, 50923 K\"oln, Germany} \bigskip \noindent PACS: 04.50.+h, 04.30.-w, 04.20.Jb. \section{Introduction} Although Einstein's general relativity theory is satisfactorily supported by experimental tests on a macroscopic level, the gravitational interaction on a microscopic scale is not well understood. The gravitational gauge models provide an alternative description of gravitational physics in the microworld \cite{PR}. A variety of models arise within the framework of the gauge approach to gravity (Poincar\'e, teleparallel, metric-affine, supergravity, to mention but a few), and their corresponding {\it kinematic} schemes are well established at present. However, the {\it dynamic} aspects of the gauge gravity models have been rather poorly studied up to now. This includes the choice of the basic Lagrangian of the theory, as well as the detailed analysis of possible physical effects. The derivation of a new exact solutions for these models may bring new insight to the understanding of gravitational physics on small scales. The plane-fronted gravitational waves represent an important class of exact solutions which generalize the basic properties of electromagnetic waves in flat spacetime to the case of curved spacetime geometry. The relevant investigation of the gravitational waves in general relativity has a long and rich history, see, e.g., \cite{peres,pen1,pen2,griff,vdz,exact}. The discussion of the possible generalizations of such solutions revealed the exact wave solutions in Poincar\'e gauge gravity \cite{adam,chen,sippel,vadim,singh,babu}, in teleparallel gravity \cite{tele}, in generalized Einstein theories \cite{gurses,lovelock}, in supergravity \cite{sg1,sg2,sg3,sg4,sg5}, as well as, more recently, in superstring theories \cite{gimon,ark1,ark2,str1,str2,str3,str4}. Some attention has also been paid to the higher-dimensional generalizations of the gravitational wave solutions \cite{coley1,coley2,hervik,ndim}. It was demonstrated \cite{dirk1,dirk2,king,vas1,vas2,vas3} that gravitational wave solutions are also admitted in the metric-affine theory of gravity (MAG) with the propagating torsion and nonmetricity fields. The latter results are, however, restricted either to the case of torsion waves only, or to the triplet class of solutions with a specific ansatz for torsion and nonmetricity \cite{eff,tri} and for a special form of the Lagrangian. The aim of this paper is to describe the plane gravitational waves for the general Yang-Mills type quadratic MAG Lagrangian with nontrivial torsion and nonmetricity configurations that do not belong to the triplet ansatz. The motivation is twofold. On the one hand, the systematic study of the space of solutions represents a significant aspect of the development of any field-theoretic model. On the other hand, the wave phenomena as such are of fundamental importance, and the construction and comparison of the wave solutions in different models may clarify the physical contents of and the relations between the microscopic and macroscopic gravitational theories (in particular, general relativity, Poincar\'e gauge gravity and MAG). The metric-affine spacetime is described by the metric $g_{\alpha\beta}$, the coframe 1-forms $\vartheta^{\alpha}$, and the linear connection 1-forms $\Gamma_{\beta}{}^{\alpha}$. These are interpreted as generalized gauge potentials, while the corresponding field strengths are the nonmetricity 1-form $Q_{\alpha\beta}=-Dg_{\alpha\beta}$ and the 2-forms of torsion $T^{\alpha}=D\vartheta^{\alpha}$ and curvature $R_{\beta}{}^{\alpha}=d\Gamma_{\beta}{}^{\alpha} + \Gamma_{\gamma}{}^{\alpha} \wedge\Gamma_{\beta}{}^{\gamma}$. The metric-affine geometry reduces to a purely Riemannian one as soon as torsion and nonmetricity both vanish. The teleparallel geometry arises when the curvature is trivial, $R_{\beta} {}^{\alpha} = 0$, whereas a vanishing nonmetricity $Q_{\alpha\beta}=0$ yields the Riemann-Cartan geometry of spacetime. It is well known that for every metric $g_{\alpha\beta}$ there exists a unique torsion-free and metric-compatible connection represented by the Christoffel symbols. We will denote this Riemannian connection by $\widetilde{\Gamma}_{\beta} {}^{\alpha}$, and hereafter the tilde will denote purely Riemannian geometrical objects and covariant differentials constructed from them. Our general notations and conventions for the basic geometric objects, the holonomic and anholonomic indices, the choice of the metric signature are that of \cite{PR}. The plan of the paper is as follows. In the next Sec.~\ref{emwave}, we recall the definition of the ordinary electromagnetic wave. This is used then in Sec.~\ref{ansatz} for the description of the corresponding ansatz for a gravitational plane wave in MAG. The properties of the resulting curvature, torsion and nonmetricity are discussed in Sec.~\ref{NTC}. Finally, in Sec.~\ref{eqs} we demonstrate that the proposed ansatz provides the exact solution for the general quadratic MAG model. The conclusions are outlined in Sec~\ref{conclusion}. \section{Electromagnetic plane waves}\label{emwave} An electromagnetic plane wave is described by a 1-form $u$ which satisfies \begin{eqnarray} d\,{}^\ast du &=& 0,\label{ddu}\\ k\wedge{}^\ast du = 0,\qquad k\wedge du &=& 0.\label{kdu} \end{eqnarray} The propagation 1-form $k$ is null (i.e., $k\wedge{}^\ast k =0$) and geodetic, $k\wedge{}^\ast dk =0$, and it corresponds to a congruence with zero shear, expansion and rotation. This is typical for the plane wave, and the equations (\ref{kdu}) represent the so-called radiation conditions imposed on the electromagnetic field. With the electromagnetic potential $A = u$, the electromagnetic field strength $F = du$ satisfies the vacuum Maxwell equation (\ref{ddu}). We do not specify the spacetime metric (and hence the Hodge operator ${}^\ast$) as the flat Minkowski one. It will be convenient not to fix the spacetime geometry at this stage. Such an electromagnetic wave construction underlies the derivation of the corresponding gravitational wave solutions as described in \cite{pleb,orr,pod1,pod2,pod3}, for example. In our study, we will use the similar constructions by extending the Riemannian results to their non-Riemannian counterparts. \section{Wave ansatz for metric, coframe and affine connection}\label{ansatz} Let us denote the local spacetime coordinates as $x^i = \{\sigma, \rho, z^2, z^3\}$. The upper case Latin indices, $A,B,\dots = 0, 1$, will label the first 2 spacetime dimensions which are relevant to a $pp$-wave. In particular, $x^A = \{\sigma, \rho\}$ are the wave coordinates with the wave fronts described by the surfaces of constant $\sigma$, and $\rho$ is an affine parameter along the wave vector of the null geodesic. The lower case Latin indices, $a,b,\dots = 2,3$, refer to the remaining spatial coordinates: $x^a = \{z^2, z^3\}$. The Greek indices, $\alpha, \beta, \dots = 0,\dots, 3$, label the local anholonomic (co)frame components. We denote separate frame components by a circumflex over the corresponding index in order to distinguish them from coordinate components. The can now formulate the wave ansatz for the MAG gravitational potentials $g_{\alpha\beta}$, $\vartheta^{\alpha}$, and $\Gamma_{\beta}{}^{\alpha}$ as follows. We choose the half-null metric \begin{equation} g_{\alpha\beta} = \left(\begin{array}{cc}g_{AB}&0\\ 0&g_{ab}\end{array} \right),\qquad g_{AB} = \left(\begin{array}{cc}0&1\\ 1&0\end{array}\right), \quad g_{ab} = \delta_{ab}.\label{met} \end{equation} The components of the coframe 1-form are given by \begin{equation}\label{cof} \vartheta^{\widehat{0}} = -\,d\sigma,\qquad \vartheta^{\widehat{1}} = {\frac 12}\,H(\sigma, z^a)\,d\sigma + d\rho,\qquad \vartheta^a = dz^a,\quad a=2,3. \end{equation} Finally, the ansatz for the affine connection reads: \begin{equation}\label{conn} \Gamma^{\alpha\beta} = k^{[\alpha}\varphi^{\beta]}\,k + k^\alpha k^\beta\,u. \end{equation} The corresponding dual frame basis (such that $e_\alpha\rfloor\vartheta^\beta = \delta_\alpha^\beta$) reads: \begin{equation} e_{\widehat{0}} = -\,\partial_\sigma + H\,\partial_\rho,\qquad e_{\widehat{1}} = \partial_\rho,\qquad e_a = \partial_a.\label{frame} \end{equation} We now make a crucial assumption that the 1-forms $u$ and $k$ above fulfill the radiation conditions (\ref{ddu}), (\ref{kdu}). Moreover, using the local coordinate system adapted for a plane wave and the half-null nature of the coframe (\ref{cof}), we can put $k = \vartheta^{\widehat{0}}$ without loss of generality. As usual, we introduce the components as $k_\alpha = e_\alpha\rfloor k$. Finally, the components of $\varphi_\alpha$ are determined by the function $H$ as follows: \begin{equation} \varphi_{\widehat{0}} = 0,\qquad \varphi_{\widehat{1}} = 0,\qquad \varphi_a = \partial_a H. \end{equation} Although this choice looks to be rather ad hoc, it is actually well motivated by the corresponding Riemannian solution, cf. \cite{ndim}. Indeed, the ansatz (\ref{met}) and (\ref{cof}) for the metric and the coframe is exactly the same as in the purely Riemannian case, whereas the connection (\ref{conn}) minimally extends the Christoffel connection (given in Appendix of \cite{ndim}) via the term proportional to the non-Riemannian parameter $u$. In the next section we demonstrate that both the torsion and the nonmetricity are determined by this 1-form. It is worthwhile to note that the wave 1-form is closed, $dk = 0$, whereas the wave covector is covariantly constant, \begin{equation} D k_\alpha = -\,\Gamma_\alpha{}^\beta\,k_\beta = 0,\qquad D k^\alpha = \Gamma_\beta{}^\alpha\,k^\beta = 0.\label{Dk} \end{equation} Here we used the fact that this covector is null, $k^\alpha k_\alpha =0$ and orthogonal to $\varphi_\alpha$, i.e. $k^\alpha\varphi_\alpha =0$. \section{Nonmetricity, torsion and curvature}\label{NTC} Given the above ansatz for the MAG potentials -- metric (\ref{met}), coframe (\ref{cof}) and linear connection (\ref{conn}) -- it is straightforward to find the corresponding gauge field strengths. The nonmetricity, torsion and curvature read, explicitly: \begin{eqnarray} Q_{\alpha\beta} &=& 2k_\alpha k_\beta\,u,\label{nonm}\\ T^\alpha &=& k^\alpha\,u\wedge k,\label{tor}\\ R^{\alpha\beta} &=& 2\gamma^{[\alpha}k^{\beta]}\wedge k + k^\alpha k^\beta\,du.\label{curv} \end{eqnarray} Here the covector-valued 1-form is defined by $\gamma_\alpha = - {\frac 12} \,\underline{d}\varphi_\alpha$, where the differential $\underline{d}$ is taken with respect to the $z^a$ coordinates only. This 1-form has the obvious properties: $k^\alpha\gamma_\alpha = 0$, $\vartheta^\alpha\wedge \gamma_\alpha = 0$ and $e_A\rfloor\gamma_\alpha = 0$, $A = 0,1$. When $u = 0$, we find zero torsion and nonmetricity. In this sense, we may consider the nontrivial $u$ to represent the true post-Riemannian geometric structures which we are primarily interested in. Let us compute the irreducible parts of the curvature 2-form. It is well known \cite{PR} that the curvature for the general linear connection can be decomposed into 11 irreducible parts. Following \cite{PR}, we first decompose the curvature 2-form into the skew-symmetric and symmetric forms, \begin{equation} W^{\alpha\beta} = R^{[\alpha\beta]} = 2\gamma^{[\alpha}k^{\beta]}\wedge k, \qquad Z^{\alpha\beta} = R^{(\alpha\beta)} = k^\alpha k^\beta\,du.\label{WZ} \end{equation} Then we have to calculate the interior products with the frame and exterior products with the coframe. In tensor language, this corresponds to computing the various contractions of the curvature tensor. All the contractions of the symmetric curvature are trivial. Namely, $Z = Z_\alpha{}^\alpha = 0$ in view of the nullity of the wave vector ($k_\alpha k^\alpha = 0$), whereas \begin{equation} e_\alpha\rfloor Z^{\alpha\beta} = -\,k^\beta\,{}^\ast(k\wedge{}^\ast du) =0, \qquad \vartheta_\alpha\wedge Z^{\alpha\beta} = k^\beta\,k\wedge du =0, \end{equation} due to the fact that $k_\alpha\vartheta^\alpha = k$ and using the radiation conditions (\ref{kdu}). As a result, all the irreducible parts of the symmetric curvature form are zero except for the first piece: \begin{equation} Z_{\alpha\beta} = {}^{(1)}\!Z_{\alpha\beta}.\label{Z1} \end{equation} In tensor language this means that all the contractions of the symmetric part of the curvature tensor are trivial, i.e., this tensor is totally trace-free and dual trace-free. The symmetric part of the curvature has no Riemannian counterpart, this is a totally post-Riemannian object. For the skew-symmetric curvature we find straightforwardly: \begin{equation} e_\alpha\rfloor W^{\alpha\beta} = (e_\alpha\rfloor\gamma^\alpha)\,k^\beta\,k, \qquad \vartheta_\alpha\wedge W^{\alpha\beta} = 0. \end{equation} Accordingly, if we demand that the zero-form $e_\alpha\rfloor\gamma^\alpha$ vanishes, then all the contractions of the skew-symmetric curvature are also trivial. This condition imposes the partial differential equation on the unknown function $H$: \begin{equation}\label{ddH} e_\alpha\rfloor\gamma^\alpha = -{\frac 12}\,\partial_a\partial^a H = 0. \end{equation} Provided that $H(\sigma, z^a)$ is a solution of the Laplace equation (\ref{ddH}), we ultimately find that all the irreducible parts of the skew-symmetric curvature form are zero except for the first piece \begin{equation} W_{\alpha\beta} = {}^{(1)}\!W_{\alpha\beta}.\label{W1} \end{equation} This is again the pure tensor part which is totally trace-free and dual trace-free, in complete analogy with the symmetric curvature. The 2-form ${}^{(1)}\!W_{\alpha\beta}$ is a direct non-Riemannian generalization of the Weyl tensor. \section{Field equations: quadratic MAG model}\label{eqs} Let us consider the general Yang-Mills type (curvature quadratic) Lagrangian for the MAG model which was studied recently in the literature (see, for example, \cite{dirk1,dirk2,king,vas1,vas2,vas3}): \begin{eqnarray} V_{\rm MAG}&=& -\,{\frac{1}{2}}\,R^{\alpha\beta} \wedge{}^*\!\left(\sum_{I=1}^{6}w_{I}\,^{(I)}W_{\alpha\beta} + w_7\,\vartheta_\alpha\wedge(e_\gamma\rfloor {}^{(5)}W^\gamma{}_{\beta} ) \nonumber\right.\\ && +\,\left.\sum_{I=1}^{5}{z}_{I}\,^{(I)}Z_{\alpha\beta} + z_6 \,\vartheta_\gamma\wedge (e_\alpha\rfloor ^{(2)}Z^\gamma{}_{\beta}) +\sum_{I=7}^{9}z_I\,\vartheta_\alpha\wedge(e_\gamma\rfloor ^{(I-4)}Z^\gamma{}_{\beta} )\right)\label{QMA}\,. \end{eqnarray} The 16 dimensionless coupling constants $w_1, \ldots w_7$, $z_1, \ldots z_9$ describe the contributions of all possible quadratic invariants which can be constructed from the components of the curvature in a general MAG theory \cite{remark}. The vacuum gravitational field equations of the MAG theory read \cite{PR}: \begin{eqnarray} DH_{\alpha}- E_{\alpha}&=& 0,\label{first}\\ DH^{\alpha}{}_{\beta}-E^{\alpha}{}_{\beta}&=& 0.\label{second} \end{eqnarray} The gravitational gauge field momenta are introduced by partial differentiation, \begin{equation} H_\alpha = - \frac{\partial V}{\partial T^\alpha}\,,\quad H^\alpha{}_\beta= -\frac{ \partial V}{\partial R_\alpha{}^\beta}\,,\quad M^{\alpha\beta} = - 2\frac{\partial V}{\partial Q_{\alpha\beta}}\,,\label{3} \end{equation} whereas the canonical gauge field currents of the gravitational energy--momentum and of the hypermomentum, respectively, are defined as the following expressions, linear in the Lagrangian and in the gauge field momenta: \begin{eqnarray} E_{\alpha} & := & \frac{\partial V}{\partial\vartheta^\alpha} =e_{\alpha}\rfloor V + (e_{\alpha}\rfloor T^{\beta}) \wedge H_{\beta} + (e_{\alpha}\rfloor R_{\beta}{}^{\gamma})\wedge H^{\beta}{}_{\gamma} + {1\over 2}(e_{\alpha}\rfloor Q_{\beta\gamma}) M^{\beta\gamma}\,,\\ E^{\alpha}{}_{\beta} & := &\frac{\partial V}{\partial\Gamma_\alpha{}^\beta}= - \vartheta^{\alpha}\wedge H_{\beta} - g_{\beta\gamma}M^{\alpha\gamma}\,. \end{eqnarray} For the purely curvature quadratic Lagrangian (\ref{QMA}), we obviously have $H_\alpha =0$ and $M^{\alpha\beta} = 0$. Hence $E^{\alpha}{}_{\beta} = 0$, and as a result the field equations (\ref{first}) and (\ref{second}) reduce to \begin{eqnarray} E_{\alpha} = e_{\alpha}\rfloor V + (e_{\alpha}\rfloor R_{\beta}{}^{\gamma})\wedge H^{\beta}{}_{\gamma} &=& 0,\label{first2}\\ D H^\alpha{}_\beta &=& 0.\label{second2} \end{eqnarray} For the above gravitational wave ansatz, we have verified that all the irreducible parts of the curvature are trivial except for the pure tensor pieces (\ref{Z1}) and (\ref{W1}). Accordingly, the direct computation of the gravitational hypermomentum 3-form then yields \begin{equation} H^\alpha{}_\beta = {}^\ast\!\left(w_1\,{}^{(1)}W^\alpha{}_\beta + z_1\,{}^{(1)}Z^\alpha{}_\beta\right).\label{Hab} \end{equation} Let us now demonstrate that the gravitational wave ansatz above provides an exact solution for the the gravitational field equations (\ref{first2}) and (\ref{second2}). The following two facts are crucial for this. The first one is the property of the curvature 2-form (\ref{curv}) \begin{equation}\label{kR} k_\alpha R^\alpha{}_\beta = 0,\qquad k^\beta R^\alpha{}_\beta = 0, \end{equation} which is obviously satisfied due to the null nature of the wave vector $k$ and its orthogonality to the 1-form $\gamma_\alpha$. The second fact concerns the well-known double duality property for the first irreducible part of the curvature: \begin{equation} {}^{\ast (1)}W_{\alpha\beta}= {\frac 1 2}\eta_{\alpha\beta\mu\nu} \,{}^{(1)}W^{\mu\nu},\label{ddW1} \end{equation} We begin with the first equation (\ref{first2}). The property (\ref{kR}) obviously yields for the gravitational wave configuration $V_{\rm MAG}= -\,{\frac{1}{2}}\,R^{\alpha\beta}\wedge H_{\alpha\beta} = 0$ as well as the contraction $(e_{\alpha}\rfloor R_{\beta}{}^{\gamma})\wedge H^{\beta}{}_{\gamma} = 0$. Hence, our ansatz solves the first equation. Finally, we turn to the second equation (\ref{second2}). Substituting (\ref{Hab}), we notice that the second term vanishes, \begin{equation} z_1\,D\,{}^\ast\!\left({}^{(1)}Z^\alpha{}_\beta\right) = z_1\,k^\alpha k_\beta\,d\,{}^\ast du = 0, \end{equation} due to the property (\ref{Dk}) and the radiation conditions (\ref{ddu}). As a result, the second equation (\ref{second2}) reduces to $w_1\,D\,{}^\ast\!\left({}^{(1)}W^\alpha{}_\beta\right) = 0$. Since ${}^{(1)}W^\alpha{}_\beta = g^{\alpha\gamma}\,{}^{(1)}W_{\gamma\beta}$, we have $D\,{}^\ast\!\left({}^{(1)}W^\alpha{}_\beta\right) = g^{\alpha\gamma} \,D\,{}^\ast\!\left({}^{(1)}W_{\gamma\beta}\right) + Q^{\alpha\gamma}\wedge {}^\ast\!\left({}^{(1)}W_{\gamma\beta}\right)$. Using the explicit form the nonmetricity (\ref{nonm}), we prove that the last term vanishes. Thus we find, in the end, \begin{equation} w_1\,D\,{}^\ast\!\left({}^{(1)}W_{\alpha\beta}\right) = 0. \end{equation} Now we can use the double duality identity (\ref{ddW1}) and obtain \begin{equation}\label{dW} {\frac {w_1} 2}\,\eta_{\alpha\beta\mu\nu}\,D\,{}^{(1)}W^{\mu\nu} = 0, \end{equation} where we used the fact that the covariant derivative of the Levi-Civita tensor $D\eta_{\alpha\beta\mu\nu} = -\,2Q\,\eta_{\alpha\beta\mu\nu} = 0$ vanishes due to the absence of the Weyl covector, $Q = {\frac 14} Q^\alpha{}_\alpha = 0$. In order to demonstrate that our ansatz solves (\ref{dW}), we can use the MAG Bianchi identity which reads $DR_\alpha{}^\beta = 0$. We find $D\, (g_{\alpha\gamma}\,R^{\gamma\beta}) = g_{\alpha\gamma}D R^{\gamma\beta} - Q_{\alpha\gamma}\wedge R^{\gamma\beta}$. Because of (\ref{nonm}) and (\ref{curv}), the last term vanishes for the gravitational wave configuration. Thus, in view of the Bianchi identity, the gravitational wave curvature satisfies $D R^{\alpha\beta} = 0$. It now remains to use the explicit formulas (\ref{curv}), (\ref{WZ}), (\ref{Z1}) and (\ref{W1}) to verify that \begin{equation} D{}^{(1)}W^{\alpha\beta} = 0, \end{equation} since the covariant derivative of the symmetric curvature vanishes identically $D{}^{(1)}Z^{\alpha\beta} = k^\alpha k^\beta\,ddu \equiv 0$. Consequently, (\ref{dW}) is satisfied by the gravitational wave ansatz, and this completes the proof that such a configuration is, indeed, an exact solution of the MAG field equations. \section{Discussion and conclusion}\label{conclusion} The extension of the Riemannian geometry of Einstein's general relativity to the post-Riemannian structures of the metric-affine gravity can be motivated by a number of reasons. Among them, we mention the problem of quantization (see the discussion of the renormalizable MAG models in \cite{lee1,lee2}), the theory of defects in the continuous media with microstructure (for a overview, see \cite{frank} and \cite{PR}), the physics of hadrons in terms of extended structures (see \cite{nee1,nee2,nee3} and more details and references in \cite{PR}), the study of the early universe (in particular, relating the post-Riemannian structures to the dark matter problem, see \cite{dirk3,dirk4,dirk5}). Finally, one can show that the MAG models may arise as the effective theories in the context of the dilaton-axion-metric low-energy limit of the string theory (see, e.g., \cite{dil1,dil2,dil3,dil4}). The study of the exact solutions of the MAG field equations is important for understanding and development of the physical aspects mentioned above. In this paper, we have derived a new plane wave solution of the general Yang-Mills type (curvature quadratic) metric-affine theory of gravity. This extends the previous study of the waves in the Yang-Mills type models of the Poincar\'e gauge gravity \cite{adam,chen,sippel,vadim,singh,babu}. As compared to the other exact wave solutions available in the literature \cite{dirk1,dirk2,king,vas1,vas2,vas3}, the new configuration has the following characteristic properties: (i) the spacetime metric is not a flat Minkowski one but the metric of the Riemannian gravitational plane wave determined by the single harmonic function $H(\sigma, z^a)$, (ii) there are not only torsion waves present but the nonmetricity has a nontrivial wave behavior as well, (iii) the post-Riemannian sector of the torsion and nonmetricity does not belong to the triplet ansatz. It is worthwhile to note that the triplet ansatz might be considered as a useful tool which helps to avoid a possible problem of the well-posedeness of the field equations by reducing them to the effective Einstein-Maxwell system of equations. However, this ansatz is applicable only to a quite narrow class of the MAG models, namely to those Lagrangians (\ref{QMA}) where the only nontrivial coupling constant $z_4\neq 0$ is allowed. Our results apply to the general case with all the 16 nontrivial coupling constants $w_1, \ldots w_7$, $z_1, \ldots z_9$. The well-posedness of the general MAG model was never studied in the literature, and this question clearly represents an open potentially interesting and important problem within the metric-affine approach to gravity. The results obtained have a number of interesting mathematical and physical applications. To begin with, the curvature quadratic Lagrangian is potentially important for a quantized theory of gravity. Furthermore, the long-distance character of the wave solutions makes them a convenient tool for the tests of the additional properties of matter besides the mass (energy-momentum), namely, the hypermomentum which includes the spin and the dilaton/shear charges. This is of particular interest for the study of the elastic media with defects, and for the physics of hadrons (see the references quoted above). It is worthwhile to stress that the new solutions are obtained as the direct generalization of the general-relativistic wave solutions. When the 1-form $u$ vanishes, the post-Riemannian geometric quantities disappear. On the other hand, the Riemannian (metric) sector of the solution has the same form as in general relativity, and thus all the earlier mathematical and physical analyses \cite{peres,pen1,pen2,griff,vdz,exact} are directly applicable to our case. In physical terms this means that the usual general relativistic detectors (with mass as the only gravitational charge) will not distinguish between the gravitational waves of the Einstein theory and the new MAG waves. This is in complete agreement with the correspondence principle which underlies the dynamical structure of the metric-affine gravity: MAG is not supposed to replace the general relativity theory in the well established macroscopic domain, but rather to extend the latter in the microscopic domain by taking into account the additional physical properties of matter (such as the spin, dilation and hypermomentum currents). The above conclusion is based on the fact that the equations of motion in MAG for the test particles without the spin, dilaton and proper hypermomentum exactly coincide with the equations of motion of the test massive matter in general relativity \cite{nee}. Furthermore, it is worthwhile to recall that within the Poincar\'e gauge gravity the equations of motion of matter are also known to coincide with the general-relativistic equations of motion for the bodies with the trivial average spin value \cite{yass}. The corresponding generalization is expected to be valid in MAG for the macroscopic bodies with vanishing average spin, dilaton and proper hypermomentum, although the detailed relevant analysis is still missing in the literature. The new solution gives a natural generalization of the definitions of a gravitational plane-fronted wave. In accordance with \cite{orr}, a gravitational wave is defined by the existence of a null, geodetic, shear-, twist, and expansion-free vector field $k$ and the Weyl 2-form subject to the algebraic conditions \begin{equation} {}^{(1)}W^{\alpha\beta}\,k_\beta = 0,\qquad {}^{(1)}W^{[\alpha\beta}\,k^{\gamma]} =0.\label{wavedef1} \end{equation} According to \cite{sippel}, a gravitational wave is defined by a covariantly constant null field and a quadratic algebraic condition on the components of the curvature tensor \begin{equation} R_{\mu\nu\alpha}{}^\beta\,R_{\rho\sigma\beta}{}^\alpha = 0.\label{wavedef2} \end{equation} As we can immediately verify, the non-Riemannian curvature (as well as its Riemannian constituent) satisfies both (\ref{wavedef1}) and (\ref{wavedef2}). In \cite{king,vas1,vas2,vas3} the notion of the pseudo-instanton solutions of the MAG field equations was introduced. The latter are described by a metric-compatible linear connection for which only one of the eleven irreducible parts of the curvature is nontrivial. Our gravitational wave solution provides a minimal generalization of the pseudo-instanton, in the sense that the nonmetricity does not vanish and that the curvature has {\it two} purely tensor irreducible parts (\ref{Z1}) and (\ref{W1}). The applications mentioned above refer to the fundamental MAG theory. However, we recall that MAG also arises as an effective theory within the framework of the dilaton-axion-metric low energy limit of the string models. Accordingly, one can use our solution as a technical tool to construct the exact wave configurations in the string motivated models where the plane waves play essential role \cite{gimon,ark1,ark2,str1,str2,str3,str4}. This construction will be described in detail elsewhere. \bigskip {\bf Acknowledgment} This work was supported by the Deutsche Forschungsgemeinschaft (Bonn) with the grant HE~528/20-1. I thank Friedrich Hehl for the reading of the manuscript and for the discussion of the results obtained.
Title: Large Scale CO Observations of a Far-Infrared Loop in Pegasus; Detection of a Large Number of Very Small Molecular Clouds Possibly Formed via Shocks
Abstract: We have carried out large scale 12CO and 13CO observations with a mm/sub-mm telescope NANTEN toward a loop-like structure in far infrared whose angular extent is about 20x20 degrees around (l, b) ~ (109, -45) in Pegasus. The 12CO distribution is found to consist of 78 small clumpy clouds whose masses range from 0.04 Mo to 11 Mo. About 83% of the 12CO clouds have very small masses less than 1.0 Mo. 13CO emission shown in the 19 of the 78 12CO clouds was detected in the region where the column density of H2 derived from 12CO is greater than 5x10(20) cm(-2), corresponding to Av of ~ 1 mag, which takes into account that of HI. We find no indication of star formation in these clouds in IRAS and 2MASS Point Source Catalogs. The very low mass clouds, M < 1 Mo, identified are unusual in the sense that they have very weak 12CO peak temperature of 0.5 K to 2.7 K and that they aggregate in a region of a few pc with no main massive clouds of ~ 100 Mo. A comparison with a theoretical work on molecular cloud formation (Koyama & Inutsuka 2002) suggests that the very low-mass clouds may have been formed in the shocked layer through the thermal instability. The star HD886 (B2IV) may be the source of the mechanical luminosity via stellar winds to create shocks, forming the loop-like structure where the very low-mass clouds are embedded.
https://export.arxiv.org/pdf/astro-ph/0601315
\title{Large Scale CO Observations of a Far-Infrared Loop in Pegasus; Detection of a Large Number of Very Small Molecular Clouds Possibly Formed via Shocks} \author{H.\ Yamamoto\altaffilmark{1}, A.\ Kawamura\altaffilmark{1}, K.\ Tachihara\altaffilmark{2}, N.\ Mizuno\altaffilmark{1}, T.\ Onishi\altaffilmark{1} and Y.\ Fukui\altaffilmark{1}} \altaffiltext{1}{Department of Astrophysics, Nagoya University, Chikusa-ku, Nagoya, Japan 464-8602; hiro@a.phys.nagoya-u.ac.jp} \altaffiltext{2}{Graduate School of Science and Technology, Kobe University, 1-1 Rokko-dai, Nada-ku, Kobe, Japan 657-8501} \keywords{ISM: clouds --- ISM: individual(High Latitude Clouds) --- radio lines: ISM --- stars: formation --- stars: winds} \section{INTRODUCTION} High Galactic latitude molecular clouds (hereafter HLCs) are typically located at $\mid$b$\mid$ $\gtrsim$ 20$\degr$--30$\degr$. Since the Gaussiun scale height of CO is estimated to be $\sim$ 100 pc in the inner Galactic disk (e.g., Magnani et al. 2000), HLCs are likely located very close to the Sun, within a few hundred pc or less. Their proximity to the Sun and the low possibility of overlapping with other objects along the line of sight enable us to study them with a high spatial resolution and to compare CO data unambiguously with the data at other wavelengths. HLCs have lower molecular densities compared with dark clouds where the optical obscuration is significant. Therefore, HLCs are often called as translucent clouds (e.g., van Dishoeck \& Black 1988) and most of the known HLCs are not the sites of active star formation, although a few of them are known to be associated with T Tauri stars (e.g., Magnani et al. 1995; Pound 1996; Hearty et al. 1999). Given the very small distances of HLCs, it is a challenging task for observers to make a complete survey for HLCs over a significant portion of the whole sky. $^{12}$CO ($J$ = 1--0) emission has been used to search for HLCs because the line emission in the mm band is strongest among the thermally or sub-thermally excited spectral lines of interstellar molecular species. It is however difficult to cover an area as large as tens of square degrees subtended by some of the HLCs because of the general weakness of the $^{12}$CO emission, typically $\sim$ a few K (e.g., Magnani et al. 1996), with existing mm-wave telescopes in a reasonable time scale. HLCs have been therefore searched for by employing various large-scale datasets at other wavelengths including the optical obscuration (Magnani et al. 1985; Keto \& Myers 1986), the infrared radiation (Reach et al. 1994), and the far-infrared excess over H{\small \,I} (=FIR excess)(Blitz et al. 1990; Onishi et al. 2001). On the other hand, unbiased surveys in CO at high Galactic latitudes have been performed at very coarse grid separations of 1$\degr$ resulting in a small sampling factor of a few \% (Hartmann et al. 1998; Magnani et al. 2000). Most recently, Onishi et al. (2001) discovered 32 HLCs or HLC complexes. This search was made based on the FIR excess, demonstrating the correlation among FIR excess clouds with CO clouds is a useful indicator of CO HLCs. Previous CO observations of individual HLCs at higher angular resolutions show that HLCs exhibit often loop-like or shell-like distributions having filamentary features with widths of several arc min or less (Hartmann et al. 1998; Magnani et al. 2000; Bhatt 2000), and in addition that HLCs often compose a group, whose angular extent is $\sim$ 10 degrees or larger. In order to better understand the structure of HLCs and to pursue the evolution of HLC complexes, CO observations covering tens of square degrees at a high angular resolution are therefore crucial. The past observations of such complexes of HLCs are limited to a few regions including Polaris flare (Heithausen \& Thaddeus 1990), Ursa Major (Pound \& Goodman 1997) and the HLC complex toward MBM 53, 54, and 55 (Yamamoto et al. 2003). Pound \& Goodman (1997) showed an arc-like structure of the molecular cloud system and suggested that the origin of such structures could be some explosive events. Most recently, Yamamoto et al. (2003) carried out extensive observations of the molecular cloud complex including MBM 53, 54, and 55 and suggest that the HLCs may be significantly affected by past explosive events based on the arc-like morphologies of molecular hydrogen (see also Gir et al. 1994). The region of MBM 53, 54, and 55 is of particular interest among the three, because it is associated with a large H{\small \,I} cloud of $\sim$ 590 $M_{\sun}$ at a latitude of $-$35 degrees and because there is a newly discovered HLC of 330$M_{\sun}$, HLCG92$-$35, which is significantly H{\small \,I} rich with a mass ratio $M$(H$_{2}$)/$M$(H{\small \,I}) of $\sim$ 1, among the known HLCs (Yamamoto et al. 2003). This cloud was in fact missed in the previous surveys based on optical extinction (Magnani et al. 1985). Subsequent to these observations we became aware of that the region is also very rich in interstellar matter as shown by the 100$\mu$m dust features (Kiss et al. 2004). There is a loop-like structure shown at 100 $\mu$m around ($l$, $b$) $\sim$ (109$\degr$, $-$45$\degr$). Toward the center of the loop, an early type star HD886(B2IV) is located and may play a role in creating the loop. Its proper motion is large at a velocity of a few km s$^{-1}$, suggesting that the stellar winds of the star might have continued to interact with the surrounding neutral matter over a few tens of pc in $\sim$ a few Myr. Magnani et al. (1985) and Onishi et al. (2001) yet observed only a small part of this region. In order to reveal the large scale CO distribution of the region, we have carried out observations toward ($l$, $b$) $\sim$ (109$\degr$, $-$45$\degr$) by $^{12}$CO ($J$ = 1--0) and $^{13}$CO ($J$ = 1--0) with NANTEN 4-meter millimeter/sub-mm telescope of Nagoya University at Las Campanas, Chile. We shall adopt the distance of 100 pc from the sun to the loop-like structure which is equal to the distance of the B2 star in the center of the loop, and is also a typical value for the HLCs. \section{OBSERVATIONS} $^{12}$CO ($J$ = 1--0) and $^{13}$CO ($J$ = 1--0) observations were made with the 4-meter telescope, NANTEN, of Nagoya University at Las Campanas Observatory of Carnegie Institutions of Washington, Chile. The front-end was an SIS receiver cooled down to 4 K with a closed-cycle helium gas refrigerator (Ogawa et al. 1990). The backend was an acousto-optical spectrometer with 2048 channels, and the total bandwidth was 40 MHz. The frequency resolution was 35 kHz, corresponding to a velocity resolution of $\sim$ 0.1 km s$^{-1}$. A typical system noise temperature was $\sim$ 200 K (SSB) at 115.271 GHz and $\sim$ 150 K (SSB) at 110.201 GHz. The half-power beam width was about 2$\farcm$6, corresponding to 0.076 pc at a distance of 100 pc. The pointing accuracy was better than 20$\arcsec$, as established by radio observations of Jupiter, Venus, and the edge of the Sun in addition to optical observations of stars with a CCD camera attached to the telescope. The observed region in $^{12}$CO was $\sim$ 240 square degrees toward the whole area of the loop-like structure centered at around ($l$, $b$) $\sim$ (109$\degr$, $-$45$\degr$) shown in a 100 $\mu$m map by Schlegel et al. (1998). First, the $^{12}$CO observations were made at a grid spacing of 8$\arcmin$$\times$cos($b$) and 8$\arcmin$ in Galactic longitude and latitude, respectively. Then, the regions where the $^{12}$CO emission is significantly detected were observed at a grid spacing of 4$\arcmin$$\times$cos($b$) and 4$\arcmin$ in Galactic longitude and latitude, respectively. The $^{13}$CO observations were made in and around the whole area where the peak temperature of $^{12}$CO emission is higher than 2.0 K at a grid spacing of 2$\arcmin$$\times$cos($b$) and 2$\arcmin$ in Galactic longitude and latitude, respectively. The periods of $^{12}$CO observations were several sessions between 2002 May and November and those of $^{13}$CO were those between 2003 April and August. All the observations were made by frequency switching whose interval is 20 MHz, corresponding to $\sim$ 50 km s$^{-1}$. The integration times per point of $^{12}$CO and $^{13}$CO observations were typically $\sim$ 30 s and $\sim$ 75 s, respectively, resulting in typical rms noise temperatures per channel of $\sim$ 0.35 K and $\sim$ 0.15 K in the radiation temperature, $T_{\rm R}^*$, respectively. In reducing the spectral data, we subtracted forth-order polynomials for the emission-free parts in order to ensure a flat spectral baseline. Total numbers of observed points of $^{12}$CO and $^{13}$CO are 16890 and 3100, respectively. We employed a room-temperature blackbody radiator and the sky emission for the intensity calibration. An absolute intensity calibration and the overall check of the whole system were made by observing Orion KL [$\alpha$(1950) = 5$^{\rm h}$32$^{\rm m}$47.$^{\rm s}$0, $\delta$(1950) = $-$5$\degr$24$\arcmin$21$\arcsec$] every 2 hours. We assumed the $T_{\rm R}^*$ of Orion KL to be 65 K for $^{12}$CO and 10 K for $^{13}$CO. \section{RESULTS} \subsection{$^{12}$CO Observation} \subsubsection{Distribution and Past Detection of $^{12}$CO Clouds} Figure 1 shows the distribution of the velocity-integrated intensity map of $^{12}$CO emission. We defined a $^{12}$CO cloud as a collection of more than two contiguous observed positions whose integrated intensity exceeds 0.77 K km s$^{-1}$ (5$\sigma$). Based on the definition, we identified 78 molecular clouds in this region. Molecular clouds are concentrated from ($l$, $b$) $\sim$ (107$\degr$, $-$37$\degr$) to (116$\degr$, $-$45$\degr$) and around ($l$, $b$) $\sim$ (114$\degr$, $-$52$\degr$). Most of the molecular clouds are very small, having size of $\lesssim$ 1$\degr$. Figure 2 shows the distribution of the CO superposed on the SFD 100 $\mu$m (Schlegel et al. 1998), which was derived from a composite of the COBE/DIRBE and IRAS/ISSA maps, with the foreground zodiacal light and confirmed point sources removed. CO clouds are distributed along the infrared loop whose diameter is $\sim$ 25 pc. We detected little CO emission within the loop-like structure, while toward some of the local peaks of SFD 100 $\mu$m there is no CO emission. Figure 3 shows the peak radial velocity distribution derived from the present $^{12}$CO data set. The velocity in Figure 3 is derived by a single gaussiun fitting from all CO spectra. The velocity range of the molecular clouds is from $-$18.3 km s$^{-1}$ to 0.3 km s$^{-1}$ and there is no systematic large scale velocity gradients. Some of the molecular clouds have already been known by previous observations. Molecular clouds toward ($l$, $b$) $\sim$ (110$\fdg$18, $-$41$\fdg$23) and (117$\fdg$36, $-$52$\fdg$28) are identified by Magnani et al. (1985) and named as MBM 1 and MBM 2, respectively. DIR117$-$44 and DIR105$-$38 identified by Reach et al. (1998) are also identified in CO toward ($l$, $b$) $\sim$ (116$\fdg$5, $-$44$\fdg$0) and (105$\fdg$0, $-$38$\fdg$0) by Onishi et al. (2001). Magnani et al. (1986) detected CO emission at ($l$, $b$) $\sim$ (112$\degr$, $-$40$\degr$). Magnani et al. (2000) also covered this region even though they made observations on a locally Cartesian grid with 1$\degr$(true angle) spacing in longitude and latitude for a beam size of 8$\farcm$8, they detected CO emission at eight positions of ($l$, $b$) $\sim$ (103$\fdg$2, $-$38$\fdg$0), (103$\fdg$2, $-$39$\fdg$0), (104$\fdg$4, $-$39$\fdg$0), (106$\fdg$8, $-$37$\fdg$0), (108$\fdg$0, $-$52$\fdg$0), (109$\fdg$5, $-$51$\fdg$0), (110$\fdg$4, $-$41$\fdg$0), and (111$\fdg$0, $-$50$\fdg$0) in the present region while they missed the present small molecular clouds whose sizes are less than several arc min in Figure 1 due to the coarse grid spacing. \subsubsection{Physical Properties of $^{12}$CO Molecular Clouds} Seventy-eight $^{12}$CO molecular clouds are identified in the present region. For each molecular cloud, $\Delta V$ derived from single Gaussian fitting was from 0.5 to 3.7 km s$^{-1}$, and the radial velocity, $V_{\rm LSR}$, ranges from $-$15.7 to $-$0.1 km s$^{-1}$. The maximum brightness temperature, $T_{\rm R}^*$($^{12}$CO) ranges from 0.5 to 5.7 K. The radius of a cloud, $R$, which is defined as the radius of an equivalent circle having the same area, i.e., and $R$(pc)$=\sqrt{A/\pi}$ where A is the total cloud surface area within the 5$\sigma$--contour level, ranges from 0.07 to 0.79 pc. The peak column density of molecular hydrogen, $N$(H$_{2}$), in each cloud derived by assuming a conversion factor of 1.0$\times$10$^{20}$ cm$^{-2}$/(K km s$^{-1}$) (Magnani et al. 2000) ranges from 8.0$\times$10$^{19}$ to 1.7$\times$10$^{21}$ cm$^{-2}$ with the present detection limit, 7.7$\times$10$^{19}$ cm$^{-2}$, corresponding to mass detection limit of 0.014 $M_{\sun}$. We estimate the molecular mass, $M$($^{12}$CO), by using the following formula \begin{equation} M(^{12}{\rm CO}) = \mu m_{{\rm H}} \Sigma[D^2 \Omega N({\rm H}_{2})], \end{equation} where $\mu$ is the mean molecular weight, assumed to be 2.8 by taking into account a relative helium abundance of 25\% in mass, $m_{{\rm H}}$ is the mass of the atomic hydrogen, $D$ is the distance from the Sun to the molecular clouds, and $\Omega$ is the solid angle subtended by a unit grid spacing of (4$\arcmin$)$\times$(4$\arcmin$$\times$cos($b$)). $M$($^{12}$CO) ranges from $\sim$ 0.04 to $\sim$ 11 $M_{\sun}$ and the total mass of molecular clouds is $\sim$ 64 $M_{\sun}$. These physical properties are listed in Table 1 and the histograms of $T_{\rm R}^*$($^{12}$CO), $\Delta V$, log($R$), and log($N$(H$_{2}$)) of these clouds are shown in Figure 4. Histograms in Figure 4 are divided into three different categories, \textit{Usual Cloud} (hereafter UC) whose mass is greater than 1 $M_{\sun}$, \textit{Small Cloud} (hereafter SC) whose mass is between 0.1 and 1 $M_{\sun}$, and \textit{Very Small Cloud} (hereafter VSC) whose mass is less than 0.1 $M_{\sun}$. It is remarkable that there are a number of molecular clouds having mass less than 1 $M_{\sun}$ and that the fractions of SC and VSC are 43/78 $\sim$ 55\% and 22/78 $\sim$ 28\% in the present region, respectively. In addition, the sizes of SC and VSC are equal to or less than 0.1 pc. We also note that the peak temperatures of SC and VSC are typically in a range from 0.5 K to 2.7 K, well below that of UC in the same region. \subsection{The Detection and Physical Properties of the $^{13}$CO Molecular Clouds} Figure 5 shows the distribution of the velocity-integrated intensity map of the $^{13}$CO emission superposed on the $^{12}$CO distribution. The total area of the $^{13}$CO observations is $\sim$ 29 square degrees toward 38 of the 78 $^{12}$CO clouds. We observed all of 13 UCs, 24 of 43 SCs, and 3 of 22 VSCs. We detected $^{13}$CO emission at 11 of the 13 UCs, 8 of the 24 SCs, and none of the 3 VSCs, indicating a trend that the $^{13}$CO intensity increases with $^{12}$CO cloud mass. A $^{13}$CO cloud is defined in the same way as for a $^{12}$CO cloud except for the lowest integrated intensity level, 0.3 K km s$^{-1}$ (3$\sigma$). Based on the definition, we identified 33 $^{13}$CO clouds. For the 33 $^{13}$CO molecular clouds, $\Delta$$V$ derived from single Gaussian fitting is $\sim$ 1.5 km s$^{-1}$ and $V_{\rm LSR}$ of them ranges from $-$13.1 to $-$1.9 km s$^{-1}$. Other physical properties, the maximum brightness temperature, $T_{\rm R}^*$($^{13}$CO), and $R$ range from 0.3 to 2.3 K and from 0.04 to 0.21 pc, respectively. The physical parameters including the molecular column density and mass (hereafter $M_{\rm LTE}$) are derived on the assumption of local thermodynamic equilibrium (LTE). To derive the column density of molecular hydrogen, the optical depth of $^{13}$CO is estimated by using the following equations, \begin{equation} \tau(^{13}{\rm CO})={\rm ln}\left[1-\frac{T_{\rm R}^{*}(^{13}{\rm CO})}{5.29}\left\{\frac{1}{{\rm exp}(5.29/T_{\rm ex})-1}-0.164\right\}^{-1}\right], \end{equation} where $T_{\rm ex}$ is the excitation temperature of the $J$ = 1--0 transition of CO in K and was derived from \begin{equation} T_{\rm ex}=\frac{5.53}{{\rm ln}\left\{1+5.53/\left[T_{\rm R}^{*}(^{12}{\rm CO})+0.819\right]\right\}}. \end{equation} $T_{\rm ex}$ was estimated to be 9.4 K from our $^{12}$CO data. The $^{13}$CO column density, $N$($^{13}$CO), is estimated by \begin{equation} N(^{13}{\rm CO})=2.42\times10^{14}\nonumber\times\frac{\tau(^{13}{\rm CO})T_{\rm ex}({\rm K})\Delta V({\rm km \hspace{0.1cm} s^{-1}})}{1-{\rm exp}[-5.29/T_{\rm ex}({\rm K})]} ({\rm cm^{-2}}). \end{equation} The ratio of $N$(H$_{2}$)/$N$($^{13}$CO) was assumed to be 7$\times$10$^{5}$ (Dickman 1978). The $M_{\rm LTE}$ of a cloud from $N$(H$_{2}$) is derived by the same way as $^{12}$CO (see equation (1)). The column density and $M_{\rm LTE}$ range from 2.3$\times$10$^{20}$ to 1.7$\times$10$^{21}$ cm$^{-2}$ and 0.03 to 1.41 $M_{\sun}$, respectively, where the detection limit in the column density is 2.0$\times$10$^{20}$ cm$^{-2}$, coressponding to mass limit of 0.009 $M_{\sun}$, smaller than that of $^{12}$CO because the observations of $^{13}$CO were made by higher grid sampling and lower rms noise fluctuations than those of $^{12}$CO, respectively. Figure 6 shows the histograms of each physical property. The virial mass, $M_{\rm vir}$, of a cloud was derived by using the following equation, assuming isothermal, spherical, and uniform density distribution with no external magnetic pressure: \begin{equation} M_{\rm vir} = 209 \times R \times \Delta V_{\rm comp}^{2}, \end{equation} where $R$ and $\Delta V_{\rm comp}$ are the radius (pc) and line width (km s$^{-1}$) of the composite profile obtained by averaging all the spectra within a cloud, respectively (for details of the line width of composite profiles, see Yonekura et al. 1997; Kawamura et al. 1998). From this equation, $M_{\rm vir}$ is estimated to be in a range from 4.7 to 197 $M_{\sun}$. These physical properties are also listed in Table 2. \section{CORRELATIONS AMONG THE CLOUD PHYSICAL PARAMETERS} \subsection{Mass Spectrum and Size Linewdith Relation} Figure 7a and 7b show the mass spectrum of the present $^{12}$CO and $^{13}$CO clouds. The spectra have been fitted by the maximum-likelihood method (Crawford et al. 1970), and it is found that they are well fitted by a single power law as follows; $dN/dM$ $\propto$ $M^{-1.53\pm0.13}$ for the $^{12}$CO clouds and $dN/dM$ $\propto$ $M^{-1.36\pm0.10}$ for the $^{13}$CO clouds. These values of the spectral indices seem to be similar to those for the higher mass range (e.g., Yonekura et al. 1997). Figure 8 shows a plot of size, $R$, versus line width, $\Delta V$, of the $^{13}$CO clouds in this region and for a comparison with other HLCs, MBM 53, 54, and 55 complex (Yamamoto et al. 2003). We can make fitting as follows by using a least-squares fitting, log($\Delta V$) = (0.22$\pm$0.43) $\times$ log($R$) + (0.37$\pm$0.52) (c.c.=0.23) for the present region and log($\Delta V$) = (0.43$\pm$0.32) $\times$ log($R$) + (0.53$\pm$0.28) (c.c.=0.37) for MBM 53, 54, and 55 complex. The low correlation coefficient (c.c.) indicates that there is no significant correlation between $R$ and $\Delta V$ because of a small range of $R$. Here we do not show the same relationship for $^{12}$CO, because the non-circular shape of the $^{12}$CO clouds may not be appropriate to derive reliable $R$. \subsection{$M_{\rm LTE}$ vs. $M_{\rm vir}$} Figure 9 shows a plot of $M_{\rm LTE}$ versus $M_{\rm vir}$. The present $^{13}$CO clouds are located far above the equilibrium line where $M_{\rm LTE}$ is equal to $M_{\rm vir}$, indicating that the $^{13}$CO clouds are not in the virial equilibrium. This indicates that none of the molecular clouds are gravitationally bound. These parameters can be fitted by using a least-squares fitting as follows, log($M_{\rm vir}$) = (0.91$\pm$0.30) $\times$ log($M_{\rm LTE}$) + (2.23$\pm$0.29) (c.c.=0.66) for present molecular clouds and log($M_{\rm vir}$) = (0.77$\pm$0.13) $\times$ log($M_{\rm LTE}$) + (0.16$\pm$0.08) (c.c.=0.74) for MBM53, 54, and 55 complex. As mentioned in Yamamoto et al. (2003), the present molecular clouds also tend to be more virialized as the mass increases. For the Gemini and Auriga, and Cepheus-Cassiopeia regions, the indices of $M_{\rm LTE}$ for $M_{\rm vir}$ of $^{13}$CO clouds are estimated to be 0.72$\pm$0.03 and 0.62$\pm$0.03 for the cloud mass range of $M_{\rm LTE}$ $<$ 10$^{4}$ $M_{\sun}$ and 10$^{2}$ $M_{\sun}$ $<$ $M_{\rm LTE}$ $<$ 10$^{5}$ $M_{\sun}$, respectively (Kawamura et al. 1998; Yonekura et al. 1997). Although the mass ranges of MBM 53, 54, and 55 complex and of this region are 10$^{-1}$ $M_{\sun}$ $<$ $M_{\rm LTE}$ $<$ 10$^{2}$ $M_{\sun}$ and 10$^{-2}$ $M_{\sun}$ $<$ $M_{\rm LTE}$ $<$ 1 $M_{\sun}$, respectively, difference in the power-law indices among these regions is small and a tendency that the SCs have large ratios of $M_{\rm vir}$/$M_{\rm LTE}$ is commonly seen. \section{COMPARISON WITH OTHER WAVELENGTH DATA} \subsection{No Sign of Star Formation} In order to look for sings of star formation associated with the present molecular clouds, we searched the IRAS point source catalog for candidates of protostellar objects satisfying the following criteria: (1) point sources having a data quality flag better than 2 in 4 bands, (2) flux ratios at 12, 25, 60 $\mu$m satisfying both log($F_{12}$/$F_{25}$) $<$ $-$0.3 and log($F_{25}$/$F_{60}$) $<$ 0, and not identified as galaxies or planetary nebulae and stars. We find that there are no cold IRAS point sources satisfying these criteria in the present molecular clouds. We also find that there are no IRAS point sources having a spectrum like a T-Tauri type star or no YSOs identified from Point Source Catalog of Two-Micron All-Sky Survey in this region. Here we select the 2MASS sources whose signal to noise ratio of valid measurements in all bands are greater than 10 and extract the sources which have the spectra like T Tauri stars in ($J$$-$$H$)--($H$$-$$K$) color--color diagram (e.g., Meyer et al. 1997) . These results suggest that the present molecular clouds are not the site of recent star-formation, or that the region is not remnants of past star formation. Such a low level of star formation is similar to the other HLCs including MBM 53, 54, and 55 complexes. \subsection{Comparison with HI} Figure 10 shows the integrated intensity map of H{\small \,I} taken from a Leiden-Dwingeloo H{\small \,I} survey (Hartmann \& Burton 1997) superposed on the integrated intensity of CO. The integrated velocity range is from $-$16 to 0 km s$^{-1}$, corresponding to the velocity range of the $^{12}$CO emission. Because an angular resolution of $\sim$ 30$\arcmin$ is coarser than that of the present CO observations by a factor of $\sim$ 10, we discuss here only the overall comparison between CO and H{\small \,I} distributions. The H{\small \,I} distribution is loop-like and the molecular clouds are distributed nearly along the H{\small \,I} loop. Figure 11 shows the position-velocity diagram of H{\small \,I} integrated from $-$40$\fdg$5 to $-$39$\fdg$5 and $-$52$\fdg$5 to $-$51$\fdg$5 in Galactic latitude, respectively. The hole like structures can be seen in H{\small \,I}, suggesting that these H{\small \,I} clouds are expanding. These expanding structures in Figures 11(a) and 11(b) correspond to the Galactic northern and southern HI shells shown as the thick dashed (semi) ellipses in Figure 10, respectively, while the expanding motion is not seen in CO (See Figure 3). These two expanding shells are also identified by an infrared radiation (Kiss et al. 2004). From Figure 10, an H{\small \,I} cloud around ($l$, $b$) $\sim$ (109$\degr$, $-$52$\degr$) seems to be located on the left side of the southern expanding shell and the molecular clouds are distributed nearby. It is difficult to distinguish with which H{\small \,I} shell the molecular clouds located around ($l$, $b$) $\sim$ (109$\degr$, $-$52$\degr$) are associated because the H{\small \,I} velocities of the two shells are similar with each other. The shape of $^{12}$CO and SFD100 $\mu$m radiation around ($l$, $b$) $\sim$ (109$\degr$, $-$52$\degr$) in Figure 2 is also similar to the left side of the expanding shell. If this is true, there may be two expanding structures in the present region. HD886 (109$\fdg$43, $-$46$\fdg$68) is located near the center of the northern expanding shell, indicating that HD886 may be affecting the northern expanding shell. The parallax of HD886 has been measured to be 9.79$\pm$0.81 mas (Perryman et al. 1997), corresponding to a distance from the Sun of 102$^{+9}_{-8}$ pc. The proper motion has also been measured to be $\mu\alpha*$=0$\farcs$0047 yr$^{-1}$, $\mu\delta$=$-$0$\farcs$0082 yr$^{-1}$ by Perryman et al. (1997). From this proper motion, the velocity of HD886 in the L-B map is estimated to be $\sim$ 4.3$\times$10$^{-6}$ pc yr$^{-1}$ at $\sim$ 100 pc (see Figure 10). We use typical values of the stellar wind for B2(IV) star on $dM$/$dt$=10$^{-9}$ $M_{\sun}$ yr$^{-1}$ and $V_{\infty}$ = 1000 km s$^{-1}$ (e.g., Snow 1982). From these parameters, the energy injected to the northern shell is estimated to be $\sim$ 10$^{47}$ ergs in a few $\times$10$^{6}$ yr. The expanding energy of the northern shell is estimated to be $\sim$ 10$^{48}$ ergs from the atomic and molecular hydrogen, using that the masses of amomic and molecular hydrogen associated with the northern shell are $\sim$ 400 and 42 $M_{\sun}$, respectively, the expanding velocity is $\sim$ 7 km s$^{-1}$ which is estimated from Figure 11 and the equation of $E_{\rm exp}$=1/2$M$$V_{\rm exp}^{2}$. Since the expanding energy of the atomic and molecular hydrogen is comparable to the energy from HD886, additional source of energy other than HD886 such as photo evaporation is needed to explain the expanding energy because the energy conversion efficiency of the stellar wind is $\lesssim$ 10\%. The energy of the southern shell is estimated to be $\sim$ 10$^{47}$ ergs, by using that the masses of atomic and molecular hydrogen are $\sim$ 1000 and 16 $M_{\sun}$, respectively, and that the expanding velocity is $\sim$ 9 km s$^{-1}$. We could not find possible candidates of the energy source for the expanding feature in the literature (SIMBAD) and there are no counterparts in optical or X-ray wavelength. Although we could not identify the possible candidates, we cannot exclude a possibility that these objects may have escaped from the region in a few $\times$ 10$^{6-7}$ yr after forming these structures. \section{DISCUSSION} \subsection{Physical States of the Small Clouds} The present observations have revealed numerous molecular clouds having very small mass of less than 1 $M_{\sun}$. It is of considerable interest to pursue the physical states of these SCs from view points of cloud physics and chemistry as well as of the origin of molecular clouds. We shall hereafter focus on the low mass $^{12}$CO clouds whose mass is less than 1 $M_{\sun}$. The total number of such clouds is 65 among 78. The $^{13}$CO emission has been searched for toward 24 of the 43 $^{12}$CO low-mass clouds whose mass is in a range of 0.1--1.0 $M_{\sun}$ and has been detected from 8 of them. Figure 12 shows correlations of the molecular column density, estimated from $^{12}$CO and $^{13}$CO, of the clouds where both $^{12}$CO and $^{13}$CO emission are detected. It is seen that almost all of the $^{12}$CO clouds having molecular column density greater than 5$\times$10$^{20}$ cm$^{-2}$, corresponding to the visual extinction of 0.55 mag if we use the relationship of $N(\rm H_2)$=9.4$\times$10$^{20}$$\times$$A$v cm$^{-2}$ (Bohlin et al. 1978; Hayakawa et al. 1999), show significant $^{13}$CO emission. We note that there are 22 $^{12}$CO clouds whose mass is less than 0.1 $M_{\sun}$; for those it is doubtful that the $^{13}$CO emission is so significant as those whose mass is a range of 0.1--1.0 $M_{\sun}$ although only 3 of them were searched for the $^{13}$CO emission in the present study. We note that we ignore the possible contribution of atomic hydrogen in the above relationship. If we take into account the distribution of atomic hydrogen as $N$(H{\small \,I}), the visual extinction would increase by 0.2 $\sim$ 0.3 mag. This may be explained as that the $^{13}$CO emitting regions become significant when $A$v becomes larger than $\sim$ 1 mag, marginally enough to shield the ultraviolet radiation to protect $^{13}$CO molecules (e.g., Warin et al. 1996), although $^{13}$CO molecules may be also affected from the ultraviolet radiation because the $N(\rm H_2)$ derived from $^{13}$CO is lower than that derived from $^{12}$CO in most of molecular clouds. The peak intensity ratio of $^{12}$CO and $^{13}$CO is around 5, much smaller than the terrestrial abundance ratio of 89, indicating that the $^{12}$CO emission is optically thick in the clouds where the both emissions are detected. The maximum $^{12}$CO peak temperature of the present brightest $^{12}$CO emission is 6 K, and this suggests that the excitation temperature is consistent with the kinetic temperature of $\sim$ 10 K, typical to the local dark clouds. The low mass clouds show lower $^{12}$CO peak temperatures down to 1 K, significantly less than the brightest peak intensities. It is not clear if this is due to the lower excitation temperatures or due to smaller filling factors significantly less than 1. In order to clarify this point we need observations of the present low mass clouds at much higher angular resolutions. \subsection{Origin of the Very Small Clouds} Figure 13 shows the histograms of mass and sizes of $^{12}$CO clouds. In the present region, we have detected a large number of molecular clouds of mass less than 0.1 $M_{\sun}$ and sizes less than 0.1 pc not detected so far in the other regions. We may ask why molecular clouds of mass less than 0.1 $M_{\sun}$ and size less than 0.1 pc such as the present clouds have not been detected so far. The main reason for this is perhaps the paucity of high-resolution observational studies of nearby molecular clouds. Most of the observations of the local high latitude clouds were made at lower resolutions of 10 arc-min or at a coarse grid spacing of 1 degree, both of which are unable to detect and resolve the present low mass clouds. This suggests that the low mass clouds similar to the present ones may not be uncommon in the interstellar space and warrant more extensive searches for them in the other parts of the sky. It is interesting to compare the physical parameters of the present VSCs with theoretical studies. The typical size of these HLCs, $\sim$ 0.1 pc, is significantly less than the Jeans length of 1.3 pc -- 7 pc for molecular gas with $T$=10 K and $n$(H$_{2}$)=10--100 cm$^{-3}$. If we take temperature higher than 10 K, the length becomes even larger. Figure 14 shows the radial distribution of mass surface density which is derived by dividing the mass in circular annulus of a radius by the area of the circular annulus. Mass surface density in the present region is fairly flat in radius because there is no massive molecular cloud in the center of the VSCs. On the other hand, mass surface density in the other regions has a gradient for radius, indicating that there are small molecular clouds around a massive molecular cloud whose mass is several dozen $M_{\sun}$ or greater (e.g., Sakamoto 2002, Sakamoto \& Sunada 2003). In these regions, the gravity of the massive molecular cloud may contribure to the formation of these small clouds by increasing the pressure in the surroundings. These suggest that mechanisms other than gravitational instability might contribute to the formation of present VSCs. A theory of molcular cloud formation is discussed by Koyama \& Inutsuka (2002). According to them, molecular clouds smaller than the Jeans-length can be formed in the shocked layer through the thermal instability. We shall present some considerations by comparing the observational results with their theoretical results. Present VSCs are likely to be affected by HD886 in the last few $\times$10$^{6}$ yr (for details, see section 5.2). Although the mechanical luminosity from HD886 injected to the loop-like structure during a few $\times$10$^{6}$ yr is low to explain the expanding of the interstellar matter, it is a possibility that the stellar wind of HD886 is the source of shock. Koyama \& Inutsuka (2002) assumed shock velocity of 26 km s$^{-1}$ and density of 0.6 cm$^{-3}$ as an initial pre-shock condition and find that the region of density greater than 100 cm$^{-3}$ grows in size to $\sim$ 0.2 pc $\times$ 0.1 pc in 1.06$\times$10$^{6}$ yr and the internal structure consists of some filaments. The velocity dispersion of CO derived by Koyama \& Inutsuka (2002) is a few km s$^{-1}$. The size of the smallest molecular clouds and the velocity dispersion of CO are comparable to those derived from Koyama \& Inutsuka (2002). We cannot resolve the internal structure of the VSCs because the present VSCs are detected with only a few points for each. The typical column density derived by Koyama \& Inutsuka (2002) is 2$\times$10$^{20}$ cm$^{-2}$, while the column density of the VSCs is estimated to be $\sim$ 1.6$\times$10$^{20}$ cm$^{-2}$ (see Figure 4). The surface filling factor of the region of density greater than 100 cm$^{-3}$ in Koyama \& Inutsuka (2002) is roughly estimated to be 30--40\%. In this surface filling factor the column density estimated from the observations is consistent with that derived from Koyama \& Inutsuka (2002) and the low temperature of the VSCs is consistent with the result that their peak temperature is lower than that of the typical local dark clouds. These results indicate a potential that the present VSCs are formed in the shock compressed layer through thermal instability. In order to compare the observational results with the theoretical simulation on internal temperature and density structure in more details, the observations of higher resolutions are needed. \section{CONCLUSIONS} We have made a large-scale survey of high Galactic latitude molecular clouds in the $J$ = 1--0 lines of $^{12}$CO and $^{13}$CO toward a large scale structure located around ($l$, $b$) $\sim$ (109$\degr$, $-$45$\degr$) with NANTEN. This survey spatially resolved the distribution of molecular gas associated with the large scale structure. The main conclusions of the present study are summarized as follows: \begin{enumerate} \item The $^{12}$CO observation covered the entire large loop-like structure. The loop-like structure consits of very small clumpy clouds. The $^{12}$CO clouds are concentrated on the north to north-west of the loop-like structure and toward the south of that. We identified 78 $^{12}$CO clouds in the observed region. The total mass is estimated to be $\sim$ 64 $M_{\sun}$ if we assume the conversion factor from CO intensity to $N$(H$_2$) as 1.0$\times$10$^{20}$ cm$^{-2}$/(K km s$^{-1}$). \item We performed $^{13}$CO observations in and around the whole area where the peak temperature of $^{12}$CO is more than 2.0 K. We identified 33 $^{13}$CO clouds and derived physical properties under the assumption of LTE. \item The mass spectra are well fitted by a power law, $dN/dM$ $\propto$ $M^{-1.53\pm0.13}$ for the $^{12}$CO clouds and $dN/dM$ $\propto$ $M^{-1.36\pm0.10}$ for the $^{13}$CO clouds. These spectral indices are similar to those derived in the other regions. \item The size and the line width relation of $^{13}$CO clouds is fitted by a least-squares method, log($\Delta V$) = (0.22$\pm$0.43) $\times$ log($R$) + (0.37$\pm$0.52) (c.c.=0.23), but the correlation is not good. \item Present $^{13}$CO clouds are far from the virial equilibrium, indicating that $^{13}$CO clouds are not gravitationally bound. $M_{\rm vir}$ and $M_{\rm LTE}$ relation can be fitted by a least-squares method as log($M_{\rm vir}$) = (0.91$\pm$0.30) $\times$ log($M_{\rm LTE}$) + (2.23$\pm$0.29) (c.c.=0.66). This index is slightly different from the indices in the other regions although the tendency that molecular clouds are more vilialized as the mass increases is consistent with the other regions. \item There is no sign of star formation from the comparison of IRAS point sources and Point Source Catalog of Two-Micron All-Sky survey in the present region. This suggests that molecular clouds in this region are not the site of present star formation or the remnants of past star formation. \item There may be two expanding shells in the present region as inferred from H{\small \,I} although we cannot identify them from CO. The total mechanical luminosity of HD886 during the last few $\times$ 10$^{6}$ yr is comparable to the expanding energy of the northern expanding H{\small \,I} shell. This indicates that some additional source of energy other than HD886 is needed to explain the expanding energy. \item $^{13}$CO emission is significantly detected in the $^{12}$CO clouds having molecular column density greater than 5$\times$10$^{20}$ cm$^{-2}$. This may be explained as that the $^{13}$CO emitting regions become significant when $A$v becomes larger than $\sim$ 1 mag, marginally enough to shield the ultraviolet radiation to protect $^{13}$CO molecules. \item There is a possibility that very small clouds have been formed in the shoked layer through the thermal instability. The stellar wind of HD886 may be the source to creat shocks, forming the loop-like structure where the very small clouds are embedded. \end{enumerate} \acknowledgments We greatly appreciate the hospitality of all staff members of the Las Campanas Observatory of the Carnegie Institution of Washington. The NANTEN telescope is operated based on a mutual agreement between Nagoya University and the Carnegie Institution of Washington. We also acknowledge that the operation of NANTEN can be realized by contributions from many Japanese public donators and companies. Three of the authors (N.M., T.O., and Y.F.) acknowledge financial support from the scientist exchange program under bilateral agreement between JSPS (Japan Society for the Promotion of Science) and CONICYT (the Chilean National Commission for Scientific and Technological Research). This research has made use of the SIMBAD astronomical database operated by CDS, Strasbourg, France. This publication makes use of data products from the Two Micron All Sky Survey, which is a joint project of the University of Massachusetts and the Infrared Processing and Analysis Center/California Institute of Technology, funded by the National Aeronautics and Space Administration and the National Science Foundation. This research has made use of the IRAS point sources from the NASA/IPAC Infrared Science Archive, which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \catcode`@=\active \def@{\phantom{0}} \begin{deluxetable}{ccccccccc} \tablewidth{0pt} \tablecaption{Physical Properties of $^{12}$CO Clouds \label{tbl-1}} \tablehead{ \colhead{No.} & \colhead{$l$} & \colhead{$b$} & \colhead{$T_{\rm R}^{\ast}$} & \colhead{$\Delta V$} & \colhead{$V_{\rm LSR}$} & \colhead{$R$} & \colhead{$N$(H$_2$)} & \colhead{$M_{\rm CO}$} \\ \colhead{(1)} & \colhead{(2)} & \colhead{(3)} & \colhead{(4)} & \colhead{(5)} & \colhead{(6)} & \colhead{(7)} & \colhead{(8)} & \colhead{(9)} } \startdata @1 & @95.52 & $-$50.00 & 3.2 & 1.3 & $-$10.7 & 0.31 & @4.7 & @1.07 \\ @2 & @96.11 & $-$53.13 & 1.1 & 1.3 & @$-$9.4 & 0.09 & @0.8 & @0.07 \\ @3 & 101.43 & $-$41.53 & 1.5 & 1.5 & @$-$9.5 & 0.17 & @1.9 & @0.27 \\ @4 & 101.91 & $-$41.20 & 0.7 & 0.5 & @$-$9.5 & 0.08 & @0.9 & @0.04 \\ @5 & 102.09 & $-$41.20 & 1.9 & 1.4 & @$-$8.9 & 0.10 & @3.4 & @0.12 \\ @6 & 102.19 & $-$41.07 & 0.9 & 1.1 & @$-$8.9 & 0.08 & @1.0 & @0.04 \\ @7 & 102.80 & $-$40.27 & 2.5 & 1.6 & @$-$8.4 & 0.16 & @4.5 & @0.30 \\ @8 & 103.32 & $-$40.33 & 1.1 & 1.3 & @$-$8.3 & 0.11 & @1.6 & @0.11 \\ @9 & 103.47 & $-$40.60 & 1.0 & 1.2 & @$-$7.3 & 0.08 & @1.0 & @0.04 \\ 10 & 103.74 & $-$39.33 & 5.2 & 2.9 & $-$10.1 & 0.59 & 12.3 & @7.02 \\ 11 & 104.20 & $-$38.93 & 0.8 & 1.0 & $-$15.7 & 0.10 & @2.0 & @0.09 \\ 12 & 104.48 & $-$38.53 & 1.4 & 3.5 & @$-$8.6 & 0.16 & @4.5 & @0.44 \\ 13 & 104.81 & $-$38.80 & 1.4 & 1.2 & @$-$5.0 & 0.10 & @1.5 & @0.08 \\ 14 & 105.07 & $-$38.80 & 1.1 & 1.4 & @$-$5.7 & 0.08 & @1.8 & @0.06 \\ 15 & 105.08 & $-$52.27 & 1.0 & 1.2 & @$-$0.1 & 0.07 & @1.2 & @0.04 \\ 16 & 105.10 & $-$38.07 & 2.5 & 2.4 & $-$12.7 & 0.26 & @5.7 & @1.19 \\ 17 & 105.22 & $-$53.20 & 1.2 & 1.3 & @$-$3.7 & 0.09 & @1.1 & @0.05 \\ 18 & 105.40 & $-$48.13 & 1.6 & 1.8 & @$-$9.0 & 0.12 & @3.0 & @0.20 \\ 19 & 105.77 & $-$38.40 & 5.6 & 1.9 & @$-$5.0 & 0.47 & 10.5 & @5.60 \\ 20 & 105.80 & $-$39.60 & 1.7 & 1.8 & @$-$9.8 & 0.24 & @3.1 & @0.70 \\ 21 & 105.82 & $-$54.00 & 1.1 & 1.4 & @$-$4.8 & 0.07 & @1.3 & @0.04 \\ 22 & 106.76 & $-$36.53 & 2.7 & 3.0 & @$-$9.8 & 0.53 & @8.3 & @4.29 \\ 23 & 106.97 & $-$37.87 & 1.7 & 1.7 & @$-$1.2 & 0.12 & @3.4 & @0.17 \\ 24 & 107.32 & $-$37.60 & 1.2 & 1.5 & @$-$2.3 & 0.12 & @1.6 & @0.11 \\ 25 & 107.38 & $-$52.00 & 1.0 & 2.6 & @$-$7.5 & 0.12 & @2.1 & @0.13 \\ 26 & 107.87 & $-$53.93 & 0.8 & 1.6 & @$-$4.9 & 0.11 & @1.8 & @0.11 \\ 27 & 108.21 & $-$53.73 & 2.2 & 1.4 & @$-$5.1 & 0.16 & @2.9 & @0.25 \\ 28 & 108.33 & $-$53.13 & 2.2 & 1.2 & @$-$0.2 & 0.07 & @3.5 & @0.08 \\ 29 & 108.78 & $-$52.60 & 5.7 & 2.4 & @$-$6.6 & 0.79 & 16.6 & 11.13 \\ 30 & 109.00 & $-$52.13 & 3.3 & 1.9 & @$-$1.5 & 0.31 & @5.1 & @1.43 \\ 31 & 109.00 & $-$50.07 & 2.5 & 1.6 & @$-$4.7 & 0.13 & @5.1 & @0.26 \\ 32 & 109.11 & $-$50.60 & 0.9 & 2.2 & @$-$5.9 & 0.09 & @1.6 & @0.06 \\ 33 & 109.17 & $-$37.60 & 4.0 & 1.5 & @$-$4.4 & 0.36 & @5.6 & @1.83 \\ 34 & 109.55 & $-$52.87 & 1.3 & 2.0 & @$-$7.1 & 0.14 & @2.5 & @0.16 \\ 35 & 109.68 & $-$38.27 & 4.6 & 1.4 & @$-$8.1 & 0.17 & @5.4 & @0.52 \\ 36 & 109.76 & $-$38.00 & 2.7 & 2.0 & @$-$4.9 & 0.21 & @4.7 & @0.63 \\ 37 & 109.78 & $-$53.33 & 2.8 & 1.5 & @$-$6.4 & 0.18 & @2.6 & @0.30 \\ 38 & 109.84 & $-$50.73 & 4.0 & 2.6 & @$-$7.8 & 0.18 & @5.0 & @0.40 \\ 39 & 109.85 & $-$38.60 & 0.8 & 1.8 & @$-$8.1 & 0.12 & @2.1 & @0.13 \\ 40 & 110.05 & $-$50.40 & 2.1 & 3.1 & @$-$7.4 & 0.10 & @3.6 & @0.14 \\ 41 & 110.06 & $-$41.27 & 4.4 & 2.9 & @$-$6.1 & 0.47 & 10.2 & @4.21 \\ 42 & 110.32 & $-$48.93 & 2.9 & 0.9 & @$-$4.2 & 0.12 & @2.6 & @0.17 \\ 43 & 110.94 & $-$41.00 & 1.3 & 3.4 & @$-$6.3 & 0.17 & @5.7 & @0.67 \\ 44 & 110.98 & $-$39.13 & 1.2 & 2.2 & @$-$9.6 & 0.12 & @2.9 & @0.20 \\ 45 & 111.08 & $-$50.13 & 1.6 & 1.0 & @$-$7.2 & 0.13 & @1.7 & @0.11 \\ 46 & 112.04 & $-$39.87 & 3.4 & 2.3 & @$-$5.4 & 0.26 & @6.4 & @1.11 \\ 47 & 112.40 & $-$40.07 & 1.3 & 2.6 & @$-$7.2 & 0.10 & @3.6 & @0.19 \\ 48 & 112.55 & $-$39.60 & 1.2 & 1.9 & @$-$3.6 & 0.16 & @3.0 & @0.29 \\ 49 & 112.64 & $-$50.07 & 0.7 & 1.5 & @$-$6.5 & 0.14 & @1.7 & @0.14 \\ 50 & 112.72 & $-$41.13 & 1.2 & 1.9 & @$-$6.2 & 0.11 & @2.4 & @0.13 \\ 51 & 112.72 & $-$41.13 & 1.2 & 1.9 & @$-$6.2 & 0.11 & @2.4 & @0.13 \\ 52 & 112.72 & $-$39.67 & 1.2 & 2.6 & @$-$5.3 & 0.10 & @1.9 & @0.09 \\ 53 & 112.84 & $-$40.27 & 1.3 & 1.7 & @$-$5.1 & 0.08 & @2.9 & @0.11 \\ 54 & 113.15 & $-$39.47 & 1.5 & 1.2 & @$-$4.7 & 0.16 & @2.5 & @0.23 \\ 55 & 113.23 & $-$52.07 & 4.4 & 1.3 & @$-$7.3 & 0.30 & @5.7 & @1.32 \\ 56 & 113.42 & $-$42.33 & 3.4 & 2.4 & $-$10.5 & 0.23 & @5.9 & @0.89 \\ 57 & 113.47 & $-$39.20 & 0.5 & 1.9 & @$-$5.7 & 0.10 & @1.4 & @0.08 \\ 58 & 113.56 & $-$49.93 & 1.3 & 0.9 & @$-$7.5 & 0.12 & @1.1 & @0.09 \\ 59 & 113.94 & $-$51.67 & 1.4 & 1.6 & @$-$9.7 & 0.07 & @3.2 & @0.07 \\ 60 & 114.38 & $-$51.73 & 3.9 & 1.6 & @$-$9.7 & 0.17 & @6.7 & @0.55 \\ 61 & 114.49 & $-$50.80 & 1.5 & 1.8 & @$-$7.8 & 0.09 & @3.2 & @0.12 \\ 62 & 114.52 & $-$41.53 & 2.2 & 1.1 & @$-$7.8 & 0.15 & @2.1 & @0.20 \\ 63 & 114.83 & $-$51.07 & 0.8 & 1.8 & @$-$8.1 & 0.10 & @1.4 & @0.07 \\ 64 & 114.90 & $-$41.73 & 1.5 & 1.2 & @$-$8.3 & 0.08 & @1.1 & @0.04 \\ 65 & 115.10 & $-$43.80 & 1.6 & 1.0 & @$-$2.1 & 0.11 & @1.5 & @0.09 \\ 66 & 115.16 & $-$45.33 & 1.2 & 1.6 & @$-$8.4 & 0.17 & @1.6 & @0.25 \\ 67 & 115.24 & $-$43.40 & 1.8 & 2.0 & @$-$3.5 & 0.23 & @2.7 & @0.57 \\ 68 & 115.63 & $-$43.60 & 0.8 & 1.0 & @$-$3.3 & 0.08 & @1.6 & @0.06 \\ 69 & 115.74 & $-$46.20 & 1.5 & 1.0 & @$-$6.8 & 0.13 & @2.3 & @0.18 \\ 70 & 116.20 & $-$43.73 & 3.2 & 1.5 & @$-$2.6 & 0.16 & @3.6 & @0.31 \\ 71 & 116.33 & $-$44.80 & 4.2 & 2.3 & @$-$3.9 & 0.59 & 10.0 & @7.13 \\ 72 & 116.45 & $-$50.53 & 2.3 & 1.1 & @$-$7.3 & 0.17 & @3.5 & @0.35 \\ 73 & 116.64 & $-$52.33 & 0.9 & 1.5 & @$-$7.6 & 0.07 & @2.1 & @0.05 \\ 74 & 116.86 & $-$43.87 & 0.8 & 3.7 & @$-$4.7 & 0.18 & @2.9 & @0.41 \\ 75 & 117.01 & $-$50.73 & 2.7 & 0.8 & @$-$7.5 & 0.10 & @2.0 & @0.10 \\ 76 & 117.11 & $-$44.33 & 2.3 & 1.1 & @$-$2.6 & 0.14 & @5.1 & @0.25 \\ 77 & 118.12 & $-$52.13 & 1.1 & 1.1 & @$-$6.7 & 0.12 & @1.8 & @0.12 \\ 78 & 118.23 & $-$52.67 & 4.2 & 2.0 & @$-$7.8 & 0.43 & @8.8 & @3.12 \\ \enddata \tablecomments{Col. (1) : Cloud number, Col. (2)--(3) : Cloud peak ($l,b$) position in degree, Col. (4) : Peak temperature in K, Col. (5) : Line width of the composite spectrum in km s$^{-1}$, Col. (6) : Peak velocity of the composite spectrum in km s$^{-1}$, Col. (7) : Radius of the molecular cloud in pc, Col. (8) : Column density of peak position in 10$^{20}$ cm$^{-2}$, Col. (9) : Mass of the molecular cloud in $M_{\sun}$. Col. (4) to (6) are derived by using a single Gaussiun fitting. } \end{deluxetable} \clearpage \begin{deluxetable}{ccccccccccc} \tablewidth{0pc} \tablecaption{Physical Properties of $^{13}$CO Clouds} \tablehead { \\ \colhead{No.} & \colhead{$l$} & \colhead{$b$} & \colhead{$T_{\rm R}^{\ast}$} & \colhead{$\Delta V$} & \colhead{$V_{\rm LSR}$} & \colhead{$R$} & \colhead{$\tau(^{13}\rm CO$)} & \colhead{$N(\rm H_2)$} & \colhead{$M_{\rm LTE}$} & \colhead{$M_{\rm vir}$} \\ \colhead{(1)} & \colhead{(2)} & \colhead{(3)} & \colhead{(4)} & \colhead{(5)} & \colhead{(6)} & \colhead{(7)} & \colhead{(8)} & \colhead{(9)} & \colhead{(10)} & \colhead{(11)} } \startdata 16a & 105.11 & $-$38.03 & 0.7 & 1.7 & $-$11.4 & 0.06 & 0.11 & @4.0 & 0.08 & @34.6 \\ 16b & 105.12 & $-$37.80 & 0.7 & 1.0 & $-$13.1 & 0.07 & 0.12 & @3.6 & 0.09 & @13.5 \\ 19a & 105.54 & $-$38.63 & 0.8 & 1.1 & @$-$5.5 & 0.09 & 0.14 & @2.5 & 0.17 & @23.2 \\ 19b & 105.77 & $-$38.37 & 2.3 & 1.7 & @$-$4.0 & 0.21 & 0.46 & 16.6 & 1.41 & 129.8 \\ 20@ & 103.70 & $-$39.33 & 1.2 & 1.5 & @$-$9.7 & 0.13 & 0.22 & 10.0 & 0.56 & @57.9 \\ 22@ & 106.93 & $-$36.40 & 0.7 & 2.1 & @$-$9.4 & 0.05 & 0.13 & @3.7 & 0.06 & @45.2 \\ 26@ & 107.87 & $-$53.93 & 0.3 & 1.8 & @$-$7.7 & 0.07 & 0.13 & @2.3 & 0.09 & @46.4 \\ 27@ & 108.10 & $-$53.77 & 0.8 & 3.1 & @$-$6.2 & 0.09 & 0.13 & @3.7 & 0.21 & 176.1 \\ 29a & 107.82 & $-$51.70 & 1.2 & 1.6 & @$-$5.0 & 0.08 & 0.21 & @7.0 & 0.21 & @40.7 \\ 29b & 108.78 & $-$52.63 & 1.4 & 2.0 & @$-$4.6 & 0.14 & 0.26 & 11.1 & 0.87 & 120.6 \\ 29c & 108.84 & $-$52.03 & 1.2 & 0.7 & @$-$6.2 & 0.11 & 0.21 & @5.9 & 0.29 & @10.6 \\ 29d & 108.89 & $-$52.17 & 1.3 & 1.1 & @$-$6.9 & 0.08 & 0.23 & @3.8 & 0.13 & @18.4 \\ 29e & 109.00 & $-$52.40 & 1.1 & 2.1 & @$-$7.8 & 0.12 & 0.20 & @6.0 & 0.41 & 110.6 \\ 29f & 109.00 & $-$52.13 & 0.9 & 1.6 & @$-$1.9 & 0.04 & 0.15 & @5.8 & 0.05 & @22.5 \\ 29g & 109.16 & $-$51.87 & 1.0 & 1.7 & @$-$5.8 & 0.10 & 0.17 & @2.3 & 0.25 & @57.6 \\ 29h & 109.53 & $-$51.17 & 0.8 & 0.9 & @$-$6.5 & 0.05 & 0.15 & @2.9 & 0.04 & @@8.7 \\ 29i & 109.64 & $-$51.57 & 0.3 & 2.4 & @$-$7.0 & 0.05 & 0.12 & @3.2 & 0.04 & @59.7 \\ 29j & 109.80 & $-$51.37 & 0.4 & 1.5 & @$-$6.7 & 0.05 & 0.14 & @2.5 & 0.04 & @22.3 \\ 33a & 109.17 & $-$37.60 & 0.8 & 1.0 & @$-$4.4 & 0.05 & 0.13 & @2.4 & 0.05 & @10.7 \\ 33b & 109.21 & $-$37.80 & 0.9 & 0.9 & @$-$4.1 & 0.06 & 0.16 & @4.7 & 0.08 & @10.6 \\ 33c & 109.34 & $-$37.80 & 0.7 & 0.7 & @$-$4.5 & 0.05 & 0.13 & @3.2 & 0.05 & @@4.7 \\ 35@ & 109.72 & $-$38.27 & 0.9 & 0.7 & @$-$8.1 & 0.05 & 0.16 & @3.2 & 0.05 & @@5.1 \\ 38@ & 109.90 & $-$50.73 & 1.1 & 1.8 & @$-$7.7 & 0.11 & 0.19 & 10.5 & 0.48 & @75.3 \\ 41a & 110.11 & $-$41.27 & 1.0 & 2.9 & @$-$5.6 & 0.11 & 0.23 & 10.1 & 0.41 & 197.4 \\ 41b & 110.19 & $-$41.07 & 0.6 & 2.0 & @$-$7.6 & 0.04 & 0.10 & @2.3 & 0.03 & @33.1 \\ 43@ & 111.12 & $-$41.00 & 0.7 & 1.1 & @$-$5.6 & 0.04 & 0.12 & @4.3 & 0.03 & @@9.4 \\ 46@ & 112.08 & $-$39.87 & 0.8 & 0.9 & @$-$5.1 & 0.06 & 0.14 & @3.7 & 0.07 & @@9.3 \\ 55@ & 113.17 & $-$52.03 & 1.2 & 0.7 & @$-$7.2 & 0.11 & 0.22 & @4.3 & 0.24 & @10.6 \\ 56@ & 113.17 & $-$42.60 & 0.8 & 0.9 & $-$10.9 & 0.04 & 0.14 & @3.7 & 0.04 & @@6.5 \\ 60@ & 114.32 & $-$51.70 & 0.8 & 1.2 & @$-$9.7 & 0.08 & 0.15 & @4.8 & 0.19 & @24.9 \\ 71a & 115.49 & $-$44.40 & 0.9 & 1.0 & @$-$3.5 & 0.10 & 0.16 & @3.7 & 0.20 & @22.2 \\ 71b & 116.21 & $-$44.97 & 1.2 & 1.6 & @$-$3.9 & 0.18 & 0.22 & @9.1 & 1.04 & @93.9 \\ 78@ & 118.19 & $-$52.70 & 1.5 & 1.2 & @$-$8.1 & 0.12 & 0.28 & @9.6 & 0.53 & @36.7 \\ \enddata \begin{flushleft} {\footnotesize Note---Col. (1) : Cloud number of $^{13}$CO taken from that of $^{12}$CO cloud with which the $^{13}$CO cloud is associated. If plural $^{13}$CO clouds are associated with one $^{12}$CO cloud, a sequential alphabet is added, Col. (2)--(3) : Cloud peak ($l$, $b$) position in degree, Col. (4) : Peak temperature in K, Col. (5) : Line width of the composite spectrum in km s$^{-1}$, Col. (6) : Peak velocity of the composite spectrum in km s$^{-1}$, Col. (7) : Radius of the molecular cloud in pc, Col. (8) : Optical depth of $^{13}$CO, Col. (9) : Column density of peak position in 10$^{20}$ cm$^{-2}$, Col. (10) : Mass of the molecular cloud assuming the LTE in $M_{\sun}$, Col. (11) : Virial mass of the molecular cloud in $M_{\sun}$. Col. (4) to (6) are derived by using a single Gaussiun fitting.} \end{flushleft} \end{deluxetable}
Title: On the variation of the fine-structure constant: Very high resolution spectrum of QSO HE 0515-4414
Abstract: We present a detailed analysis of a very high resolution (R\approx 112,000) spectrum of the quasar HE 0515-4414 obtained using the High Accuracy Radial velocity Planet Searcher (HARPS) mounted on the ESO 3.6 m telescope at the La Silla observatory. The HARPS spectrum, of very high wavelength calibration accuracy (better than 1 m\AA), is used to search for possible systematic inaccuracies in the wavelength calibration of the UV Echelle Spectrograph (UVES) mounted on the ESO Very Large Telescope (VLT). We have carried out cross-correlation analysis between the Th-Ar lamp spectra obtained with HARPS and UVES. The shift between the two spectra has a dispersion around zero of \sigma\simeq 1 m\AA. This is well within the wavelength calibration accuracy of UVES (i.e \sigma\simeq 4 m\AA). We show that the uncertainties in the wavelength calibration induce an error of about, \Delta\alpha/\alpha\le 10^{-6}, in the determination of the variation of the fine-structure constant. Thus, the results of non-evolving \Delta\alpha/\alpha reported in the literature based on UVES/VLT data should not be heavily influenced by problems related to wavelength calibration uncertainties. Our higher resolution spectrum of the z_{abs}=1.1508 damped Lyman-\alpha system toward HE 0515-4414 reveals more components compared to the UVES spectrum. Using the Voigt profile decomposition that simultaneously fits the high resolution HARPS data and the higher signal-to-noise ratio UVES data, we obtain, \Delta\alpha/\alpha=(0.05\pm0.24)x10^{-5} at z_{abs}=1.1508. This result is consistent with the earlier measurement for this system using the UVES spectrum alone.
https://export.arxiv.org/pdf/astro-ph/0601194
\title{On the variation of the fine-structure constant: Very high resolution spectrum of QSO HE {0515$-$4414} \author{Hum Chand\inst{1}, Raghunathan Srianand\inst{1}, Patrick Petitjean\inst{2,3},\\ Bastien Aracil\inst{2,4}, Ralf Quast\inst{5}, Dieter Reimers\inst{5} } \thanks{ Based on observations collected at the European Southern Observatory (ESO), under Programe ID No. 072.A-0244 with HARPS on the 3.6~m telescope operated at the La Silla Observatory and Programe ID 066.A-0212 with UVES/VLT at the Paranal observatory.} } \titlerunning{Variation of the fine-structure constant} \authorrunning{H. Chand et al.} \author{Hum Chand\inst{1}, Raghunathan Srianand\inst{1}, Patrick Petitjean\inst{2,3},\\ Bastien Aracil\inst{2,4}, Ralf Quast\inst{5}, Dieter Reimers\inst{5} } \offprints{H. Chand \\~\email{hcverma@iucaa.ernet.in}} \institute{$^1$IUCAA, Post Bag 4, Ganeshkhind, Pune 411 007, India\\ $^2$Institut d'Astrophysique de Paris, UMR7095 CNRS, Universite Pierre \& Marie Curie, 98 bis boulevard Arago, 75014 Paris.\\ $^3$LERMA, Observatoire de Paris, 61 Rue de l'Observatoire, F-75014 Paris, France\\ $^4$Department of Astronomy, University of Massachusetts, 710 North Pleasant Street, Amherst, MA 01003-9305, USA\\ $^5$Hamburger Sternwarte, Universitat Hamburg, Gojenbergsweg 112, D-21209 Hamburg, Germany\\ } \date{Received date/ Accepted date} \abstract {} {We present a detailed analysis of a very high resolution (R~$\approx 112,000$) spectrum of the quasar HE {0515$-$4414} obtained using the High Accuracy Radial velocity Planet Searcher (HARPS) mounted on the ESO 3.6~m telescope at the La Silla observatory. The main aim is to use HARPS spectrum of very high wavelength calibration accuracy (better than 1~m\AA), to constrain the variation of $\alpha=e^2/\hbar c$ and investigate any possible systematic inaccuracies in the wavelength calibration of the UV Echelle Spectrograph (UVES) mounted on the ESO Very Large Telescope (VLT).} {A cross-correlation analysis between the Th-Ar lamp spectra obtained with HARPS and UVES is carried out to detect any possible shift between the two spectra. Absolute wavelength calibration accuracies, and how that translate to the uncertainties in \dela are computed using Gaussian fits for both lamp spectra. The value of \dela at \zabs~=~1.1508 is obtained using Many Multiplet method, and simultaneous Voigt profile fits of HARPS and UVES spectra.} {We find the shift between the HARPS and UVES spectra has mean around zero with a dispersion of $\sigma\simeq 1$ m\AA. This is shown to be well within the wavelength calibration accuracy of UVES (i.e $\sigma\simeq 4$ m\AA). We show that the uncertainties in the wavelength calibration induce an error of about, \dela $~\le 10^{-6}$, in the determination of the variation of the fine-structure constant. Thus, the results of non-evolving \dela reported in the literature based on UVES/VLT data should not be heavily influenced by problems related to wavelength calibration uncertainties. Our higher resolution spectrum of the \zabs~=~1.1508 Damped Lyman-$\alpha$ system toward HE {0515$-$4414} reveals more components compared to the UVES spectrum. Using only \feii lines of \zabs~=~1.1508 system, we obtain \dela~=~${ (0.05\pm0.24)\times10^{-5}}$. This result is consistent with the earlier measurement for this system using the UVES spectrum alone.} {} \keywords{ {\em Quasars:} absorption lines -- {\em cosmology:} observations } \section{Introduction} \label{sect:Int} Some of the modern theories of fundamental physics, such as SUSY, GUT and Super-string theory, allow possible space and time variations of the fundamental constants, thus motivating an experimental search for such a variation (Uzan 2003 and 2004 for a detail review on the subject). Murphy et al. (2003), applying the Many Multiplet method (MM method) to 143 complex metal line systems, claimed a non-zero variation of the fine-structure constant, $\alpha=e^2/\hbar c$: $\langle\Delta\alpha/\alpha\rangle = (-0.57\pm0.11)\times10^{-5}$ for $0.2\le z \le 3.5$, where $\Delta\alpha/\alpha=(\alpha_{z}-\alpha_{0})/\alpha_{0}$, with $\alpha_{0}$ being the present value and $\alpha_{z}$ its value at redshift $z$. This result, if true, would have very important implications to our understanding of fundamental physics and has therefore motivated new activities in the field. Search for the possible time-variation of $\alpha$ using alkali doublets has started long ago (Bahcall et al. 1967). The alkali-doublet method is a clean method for constraining the variation in $\alpha$ using spectral lines because it uses transitions from the same species (Wolfe et al. 1976; Levshakov 1994; Potekhin et al. 1994; Cowie \& Songaila, 1995; Varshalovich et al. 1996; Varshalovich et al. 2000; Murphy et al. 2001a; Martinez et al. 2003; Chand et al. 2005). The tightest constraint obtained using this method till date is \dela = (0.15$\pm$0.44) $\times10^{-5}$ at $z \sim 2$ (Chand et al. 2005). \par Studies based on heavy element molecular absorption lines seen in the radio/mm wavelength range are more sensitive than that based on optical/UV absorption lines. They usually provide constraints on the variation of a combination of the fine-structure constant, the proton g-factor ($G_{p}$) and the electron-to-proton mass ratio ($\mu$). Murphy et al. (2001b) have obtained \dela = $(-0.10\pm0.22)\times10^{-5}$ at z = 0.2467 and \dela = $(-0.08\pm0.27)\times 10^{-5}$ at z = 0.6847, assuming a constant proton g-factor ($G_{p}$). It has been pointed out that OH lines are very useful in simultaneously constraining various fundamental constants (Chengalur \& Kanekar 2003; Kanekar et al. 2004; Darling 2003, 2004). These studies have provided \dela = $(0.6\pm1.0)\times 10^{-5}$ for an absorption system at \zabs = 0.247 toward PKS~1413+135. Such studies have not been performed yet at higher redshift (i.e $z$~$\ge1$) due to the lack of molecular absorption systems. \par Constraints on the variations of $\alpha$ are also obtained from terrestrial measurements. The most stringent constrain has been obtained from the analysis of the Oklo phenomenon. Fujii et al. (2000) find that $\Delta\alpha/\alpha = (-0.8\pm1.0)\times10^{-8}$ over a period of about 2 billion years (or $z\simeq0.45$). Laboratory experiments also give very stringent constraints on the local variation of $\alpha$. Marion et al. (2003) have obtained, $\Delta\alpha/\alpha\Delta t = (-0.4\pm16)\times10^{-16}\,{\rm yr}^{-1}$, by comparing the hyperfine transition in $^{87}$Rb and $^{133}$Cs over a period of 4 years assuming no variation in the magnetic moments. Fischer et al. (2004) have obtained, $\Delta\alpha/\alpha\Delta t = (-0.9\pm2.9)\times10^{-16}\,{\rm yr}^{-1}$, by comparing the absolute $1S-2S$ transition of atomic hydrogen to the ground state of Cesium. A linear extrapolation gives a constraint of $-1.3\times10^{-6}\le$ \dela$\le1.9\times10^{-6}$ at $z = 1$ for the most favored cosmology ($\Omega_m = 0.27$, $\Omega_{\Lambda} =0.73$ and $ h=0.71$). \par Clearly all the experimental results summarized above are consistent with no variation of $\alpha$. However, these results do not directly conflict with the positive detection by Murphy et al. (2003) either because of the insufficient sensitivity of the method (as in the case of alkali doublets) or because of the different redshift coverage (as in the case of radio and terrestrial measurements). However, recent attempts using the MM method (or its modified version) applied to very high quality UVES spectra have resulted in null detections. The analysis of Fe~{\sc ii} multiplets and Mg~{\sc ii} doublets in a homogeneous sample of 23 systems has yielded a stringent constraint, \dela = $(-0.06\pm0.06)\times10^{-5}$ (Chand et al. 2004; Srianand et al. 2004). Modified MM method analysis of \zabs = 1.1508 toward HE 0515$-$4414 that avoids possible complications due to isotopic abundances has resulted in \dela = $(0.01 \pm 0.17)\times 10^{-5}$ (Quast et al. 2004). Levshakov et al. (2005b) have re-analysis this system using the single ion differential alpha measurement method as described in Levshakov et al. (2005a), and obtained \dela = $(-0.007\pm0.084)\times10^{-5}$. Clearly all studies based on VLT-UVES data are in contradiction with the conclusions of Murphy et al. (2003). \par A first possible concern about these studies is the real accuracy and robustness of the various calibration procedures. A second possible source of uncertainty comes from the multi-component Voigt-profile decomposition. It is very important to check how sensitive the derived constraints are to the profile decomposition. This can be done by performing the analysis on data of higher resolution than typical UVES (or HIRES) spectra. The best way to investigate all this is to compare data taken by UVES (or HIRES) with data on the same object taken with another completely independent, well controlled, and higher spectral resolution instrument. The advent of HARPS mounted on the ESO 3.6~m telescope makes this possible. Unfortunately this is only possible on the brightest quasar in the southern sky, HE~0515$-$4414. \par This forms the basic motivations of this work. We report the analysis of the \zabs = 1.15 DLA system toward QSO HE 0515$-$4414 (De la Varga et al. 2000, Quast et al. 2004, 2005) using very high resolution (R$\sim112,000$) spectra obtained with HARPS mounted on the ESO 3.6~m telescope. The organization of the paper is as follows. The HARPS observations of HE 0515$-$4414 are described in Section 2. Calibration accuracy and comparison with the UVES observations are discussed in Section 3. In Section 4 we present the joint analysis of the HARPS and UVES spectra. Results are summarized and discussed in Section 5. \section{Observations} \label{sect:Obs} The spectrum of HE 0515$-$4414 used in this work was obtained with the High Accuracy Radial velocity Planet Searcher (HARPS) mounted on the ESO 3.6~m telescope at the La Silla observatory. HARPS is a fiber-fed spectrograph and is therefore less affected by any fluctuation in the seeing conditions (Mosser et al. 2004). It is installed in the Coud\'e room of the 3.6~m telescope building and is enclosed in a box in which vacuum and constant temperature are maintained. The instrument has been specifically designed to guarantee stability and high-accuracy wavelength calibration. \par The observations were carried over four nights in classical fiber spectroscopy mode, with one fiber on the target and the other on the sky. The CCD was read in normal low readout mode without binning. The echelle order extraction from the raw data frame is done using the HARPS reduction pipeline. The error spectrum is computed by modeling the photon noise with a Poisson distribution and CCD readout noise with a Gaussian distribution. The calibrated spectrum is converted to vacuum wavelengths according to Edl\'en (1966) and the heliocentric velocity correction is done manually using the dedicated MIDAS (ESO-Munich Image Data Analysis Software) procedure. Special attention was given while merging the orders. While combining overlapping regions, higher weights were assigned to the wavelength ranges toward the center of the order compared to the one at the edges. The resulting 1-D spectrum covers the wavelength range from 3800 to 6900 \AA, with a gap between 5300 to 5330 \AA~ caused by the transition between the two CCDs used in HARPS. In total, we obtained 14 individual exposures, each of duration between 1 and 1.5 hour. Combination of individual exposures is performed using a sliding window and weighting the signal by the errors in each pixel. The final error spectrum was obtained by adding quadratically in each pixel the extracted errors and the rms of the 14 individual measurements. The final combined spectrum has a S/N ratio of about 30 to 40 per pixel of size $\sim$0.015 \AA~ and a spectral resolution of R$\approx$ 112,000.\par To make quantitative comparisons, as will be discussed in the next section, we have also used the UVES spectrum of this QSO. The details of the UVES observation and data reduction can be found in Quast et al. (2004). However we have used our procedures for air-to-vacuum wavelength conversion, heliocentric velocity correction and for the addition of individual exposures as in the case of the HARPS spectrum. \section{Accuracy of wavelength calibration} \label{sect:comp_cal} In this Section we investigate (i) the cross-correlation between the Th-Ar lamp spectra obtained with HARPS and UVES, (ii) the absolute wavelength calibration accuracies of HARPS and UVES and (iii) how the uncertainties in the wavelength calibration translate into uncertainties in \dela measurements in the case of HARPS and UVES. \subsection{Cross-correlation of UVES and HARPS Th-Ar spectra} \label{crossCorr:subsec} To estimate how well the UVES and HARPS wavelength scales agree, one can in principle use the narrow heavy element absorption lines seen in the spectra of the QSO. However not only the number of such lines is small but also, due to differences in the resolutions and S/N ratios, spurious shifts can be introduced in the analysis. In order to avoid this, we perform a cross-correlation analysis between the Th-Ar lamp spectra obtained with UVES and HARPS. We have 4 and 14 Th-Ar lamp exposures respectively for UVES and HARPS observations in the setting that covers the wavelength range where Fe~{\sc ii} and Mg~{\sc ii} absorption lines from the \zabs~=~1.1508 absorption system are seen. We have combined all the extracted Th-Ar exposures after subtracting a smooth continuum corresponding to the background light. \par The cross-correlation analysis was performed on groups of five consecutive unblended emission lines that are clearly seen in both the UVES and HARPS spectra. For this, both spectra were re-sampled to an uniform wavelength scale using cubic spline and the pixel-by-pixel cross correlation was performed by shifting the UVES spectrum with respect to the HARPS spectrum. The results of the cross-correlation at places where absorption lines at \zabs~=~1.1508 are redshifted are shown in Fig.~\ref{crossc.fig}. All the curves shown in this figure have their peak at zero pixel shift with a typical pixel size of 15 m\AA. In order to derive sub-pixel accuracy in the cross-correlation, we have fitted a Gaussian to the cross correlation curves as is shown by dotted lines (Fig.~\ref{crossc.fig}) and derive its centroid accurately. The corresponding values are given in each panel. The relative shifts between the two spectra are less than 1 m\AA~ except in one case where it is 1.7~m\AA. We note that the quadratic refinement technique (instead of a Gaussian fitting) also gives similar results. To derive the global trend of the relative shift, we have extended our cross-correlation analysis, to the entire wavelength range. The result of the analysis is shown in Fig.~\ref{crossHARPSUV.fig}. The shifts are obtained in the same way as in Fig.~\ref{crossc.fig}. The average of the mean relative shifts over the entire wavelength range is 0.01 m\AA~ with an rms deviation of 1.09 m\AA. In what follows we investigate the absolute wavelength calibration accuracies of the two instruments. \subsection{Testing absolute wavelength calibration error of UVES and HARPS} \label{abscal:subsec} To test the absolute wavelength calibration accuracy we compare the central wavelength of strong un-blended emission lines in the extracted Th-Ar lamp spectrum with the wavelengths tabulated in Cuyper et al. (1998). We model the emission lines by a single Gaussian function. The best-fit line-centroid along with other parameters of the models and errors are determined by a \chisq minimization procedure. In many cases we find it difficult to fit the lines with reduced $\chi^{2}\approx 1$. In such cases we have scaled the flux errors by square root of the reduced \chisq and re-run the fitting procedure. In this way, we have avoided any underestimation of the errors on the best fit parameters, assuming that the actual errors on the flux of the Th-Ar lamp spectrum was somehow underestimated.\par The difference between the best-fit line centroid, in the extracted lamp spectra and the wavelength quoted by Cuyper et al. (1998) is plotted in Fig.~\ref{dlam.fig}. The wavelength range shown in this figure is the one covered by the main \feii and \mgii lines of the \zabs~=~1.1508 system. We find the rms of the deviation ($\Delta\lambda$ in Fig.~\ref{dlam.fig}) around zero to be, respectively, 0.87~m\AA~ and 4.08~m\AA~ for the HARPS and UVES lamp spectra. This clearly demonstrates that the shifts between the HARPS and the UVES lamp spectra measured from the cross-correlation analysis (i.e $\le1$ m\AA) are well within the wavelength calibration accuracy of UVES.\par In addition, we have used the best-fit FWHM of the Gaussian fit of the lamp lines to derive the spectral resolution ($R=\lambda/FWHM$) of the spectrum. The resolution measurements are shown in Fig.~\ref{reso.fig}. The mean resolution and standard deviation for HARPS and UVES are found to be $R= 112,200$ and $\sigma=8,400$; $R= 55,100$ and $\sigma=7,600$ respectively.\par \subsection{Effect of calibration error on $\Delta\alpha/\alpha$ measurement} \label{calDal:subsec} Next we investigate how the scatter in wavelength calibration ($\Delta\lambda$) translates into a scatter in \dela. We follow the method used by Murphy et al. (2003) for this purpose. We randomly choose 3 emission lines in the lamp spectrum, with a rest wavelength close to each of the observed wavelengths of the \feii and \mgii lines used in the analysis of the variation of $\alpha$. There are two \mgii lines, $\lambda$2796 and $\lambda$2803, and five \feii lines, $\lambda$2344, $\lambda$2374, $\lambda$2382, $\lambda$2586, and $\lambda$2600. Thus we have 21 ($7\times3$) lines per realization. By choosing 3 lines, we mimic 3 distinct components in the actual absorption system. We assume that the measured shift in the emission line centroid away from the actual value is caused by the variation in $\alpha$. To estimate this variation, we use the analytic fitting function given by Dzuba et al. (2002), \begin{equation} w = w_o + q x. \end{equation} Here, $w_o$ and $w$ are, respectively, the vacuum wave number (in units of cm$^{-1}$) measured in the laboratory and the modified wave number due to a change in $\alpha$; $x=(\Delta\alpha/\alpha+1)^2-1$ and $q$ is the sensitivity coefficient. At each chosen lamp emission line we assign the $q$ value of the neighboring metal absorption transition. All the lamp emission lines in each realization are fitted simultaneously with Gaussians, for one fixed value of \dela. Here, the \dela value is used to modify the rest wavelength of the emission lines using the $q$ coefficients given by Dzuba et al. (2002) for the corresponding metal lines. This procedure is repeated for a range of \dela, from $-2.0\times10^{-5}$ to $2.0\times 10^{-5}$ in steps of $0.02\times10^{-5}$ to achieve \chisq as a function of \dela. The \chisq versus \dela curve is used to extract the best fitted \dela (with error-bars) in a similar way as is used in the absorption system (discussed in the next Section). The measured spurious \dela for 100 random realizations are plotted in Fig.~\ref{dalpha.fig} both for HARPS (left-hand side middle panel) and UVES (left-hand side lower panel) lamp spectra. In the top panel we give the results for similar analysis of UVES spectrum considering 6 Fe~{\sc ii} lines (i.e including Fe~{\sc ii}$\lambda$1608 instead of Mg~{\sc ii} doublet) alone. \par We notice that the measured values of \dela obtained in this experiment have a Gaussian-shape distribution with $\sigma$~of~$0.02\times 10^{-5}$ for HARPS and $\sigma \simeq 0.1\times10^{-5}$ for UVES. As the system under consideration is known to have much more than 3 components, the above quoted values are conservative errors due to uncertainties in the wavelength calibration. Murphy et al. (2003) have also carried out such analysis for HIRES Th-Ar lamp spectra. Their weighted mean from the sample of 128 sets of Th-Ar lines resulted in $\langle\Delta\alpha/\alpha\rangle_{ThAr}=(0.4\pm0.8)\times10^{-7}$. If one assumes a Gaussian distribution for the individual values, then the central limits theorem implies that the typical $\sigma$ from one set of Th-Ar lines in the case of HIRES should be around $0.09\times10^{-5}$ ($\equiv0.8\times10^{-7}\times\sqrt{128}$), which is similar to our value for UVES Th-Ar lamp spectra (i.e $\sigma = 0.1\times10^{-5}$).\par \subsection{Effect of using different Th-Ar line tables on wavelength calibration} Th-Ar reference wavelengths are taken from the compilations of Palmer et al. (1983) for Thorium lines and Norl\'en et al. (1973) for Argon lines. The line lists built from these compilations and commonly used for echelle spectroscopy calibration are available on the web-pages of the European Southern Observatory (ESO\footnote{http://www.eso.org/instruments/uves/tools/tharatlas.html}) and the National Optical Astronomy Observatory (NOAO\footnote{http://www.noao.edu/kpno/specatlas/thar/thar.html}). The two tables differ slightly, because the ESO Th-Ar line table is not accurate up to 4 decimal places as is the case with NOAO Th-Ar line table. For the extraction of UVES lamp spectra we have used the Th-Ar line table provided by NOAO. To investigate whether the use of ESO table could induce systematic shifts in \dela, we have also extracted the same UVES Th-Ar lamp spectrum using the Th-Ar line table provided by ESO. We fit a Gaussian function to the un-blended Th-Ar line as described in sub-section~\ref{abscal:subsec} and get the deviation, $\delta\lambda_{fit}$, of the best-fit centroid with respect to the corresponding value in the NOAO Th-Ar table. The deviation ($\delta\lambda_{fit}$) is plotted in Fig.~\ref{tab_diff.fig} as a function of the difference in the wavelengths tabulated by ESO and NOAO, $\Delta\lambda_{tab}$. If the wavelength uncertainties caused by the inaccurate wavelengths listed in ESO Th-Ar table for some of the Th-Ar lines are larger than the errors allowed by the dispersion solution, then we expect a correlation between $\delta\lambda_{fit}$ and $\Delta\lambda_{tab}$. The lack of such a correlation and the larger scatter of $\delta\lambda_{fit}$ compared to $\Delta\lambda_{tab}$ in the figure, show that the effect of inaccurate rest-wavelengths of a few lines in the ESO line list is negligible.\par To complement this, we perform the cross-correlation between the lamp spectra calibrated using the two wavelength tables. The cross-correlation is performed in a similar way as described in sub-section~\ref{crossCorr:subsec}. Here we have shifted the UVES lamp spectrum calibrated using the ESO Th-Ar table over the same lamp spectrum calibrated using the NOAO Th-Ar line table. The result of the cross-correlation is shown in the upper panel of Fig.~\ref{crossNoaoEso.fig}. From the figure it can be seen that the relative shift is not completely random. However the relative shift is most of the time less than 2m\AA~and even 1m\AA, which is well within the UVES calibration accuracy.\par We also repeat the exercise to derive how these wavelength calibration uncertainties translate into \dela as described in detail in sub-section~\ref{calDal:subsec} for the case when one uses for calibration the ESO Th-Ar line table (Fig.~\ref{dalpha.fig} for UVES lamp uses NOAO table). The result is shown in the lower left-hand side panel of the Fig.~\ref{crossNoaoEso.fig} for 100 realizations. The histogram shown in the lower right-hand side panel shows that the fiducial \dela is distributed like a Gaussian. As a result, we can conclude that the \dela measurements in the literature (Chand et al. 2004 \& 2005, Quast et al. 2004) using the ESO Th-Ar line table, should not be significantly affected by this possible systematic effect.\par \section{Analysis} \label{sect:Ana} In this section we present the results on the measurement of \dela using the HARPS and UVES spectra. The details of the analysis used here, validation of the procedure using simulated spectra and the error budget from $\chi^2$ analysis can be found in Chand et al. (2004, 2005). Here, we mainly concentrate on (i) comparing the methods used by Chand et al. (2004, 2005) to derive \dela with that used by Quast et al. (2004) and (ii) understanding the effect of the decomposition of the absorption profiles into multiple narrow Voigt-profile. \par \subsection{Re-analysis of the red sub-system in the UVES data} In the analysis of Chand et al (2004, 2005) \dela is not explicitly used as fitting parameter. Instead \chisq versus \dela curve is used to get the best fitted value of \dela. However, Quast et al. (2004) use the Voigt profile analysis keeping \dela also as a fitting parameter in addition to $N$, $b$ and $z$. Chand et al. (2005), using analytic calculations, have shown that both the approaches should give the same result. Here we check this by re-analysing the absorption lines of the \zabs~=~1.1508 system toward HE {0515$-$4414} using \chisq versus \dela curve.\par The absorption lines of this system is spread over about 730 \kms (Quast et al. 2004). We have divided the whole system in two well detached blue and red sub-systems. The blue sub-system covers the velocity range $-$570 to $-$100 \kms and the red sub-system covers the velocity range $-$20 to $+$110 \kms with respect to \zabs=1.1508. Our best fit Voigt-profiles to the blue and red sub-system using the UVES spectrum, is shown respectively in the left and right-hand side panels of Fig.~\ref{uves.fig}. The vertical dotted lines are best fitted velocity components obtained in this study and the long dashed vertical lines mark the velocity components of the Quast et al. (2004). Apart form the component around $\sim90$~\kms, we find almost perfect matching between the components obtained with two different fitting codes. The variation of $\chi^2$ as a function of \dela using this initial fit (Fig.~\ref{uves.fig}) is shown in the left-hand side panel of Fig.~\ref{res.fig}. The scatter seen in the \chisq curve is mainly due to low column density of many components in blue sub-system (see the discussion in Chand et al. 2004). The position of the minimum in the \chisq curve remains uncertain till either we smooth the curve or fit some smoothing polynomial to it. Therefore we have fitted a polynomial function of 4th order minimizing the rms deviation. The best fit of the \chisq curve is shown by the solid line (left-hand side panel of Fig.~\ref{res.fig}). Its minimum gives \dela~=~${(0.10\pm0.22)\times10^{-5}}$, using $\chi^2_{min}+1$ statistics. The derived position of the minimum does not depart significantly when we use a 2nd or 3rd order polynomial fit to the $\chi^2$ data points. Our best fitted value, \dela = ${(0.10\pm0.22)\times10^{-5}}$, is very much consistent with that obtained by Quast et al. (2004) (\dela~=~${[0.01\pm0.17]\times10^{-5}}$). The best fitted column densities and Doppler parameters in individual components also agree well (see Fig.~\ref{compfit}). \begin{table*} {\large \caption{Results of the Voigt profile fit of Fe~{\sc ii} lines at \zabs = 1.1508 toward HE {0515$-$4414}.} \begin{tabular}{rrrcr} \hline\hline \\ {C.N} & \multicolumn{1}{c}{\zabs} & \multicolumn{1}{c}{b} & \multicolumn{1}{c}{log[N(Fe~{\sc ii})]} & \multicolumn{1}{c}{V$^{a}$} \\ & & \multicolumn{1}{c}{(\kms)} & \multicolumn{1}{c}{(cm$^{-2}$)}& \multicolumn{1}{c}{(\kms)} \\ \\ \hline\hline \\ 1 &$1.146938\pm0.00000^{\dagger}$ &$ 1.70\pm0.22$ &$ 11.38\pm 0.14$ &$ -538.79\pm 00.00$ \\ 2 &$1.146969\pm0.000098$ &$ 2.34\pm0.25$ &$ 12.30\pm 0.03$ &$ -534.46\pm 13.71$ \\ 3 &$1.147008\pm0.00000^{\dagger}$ &$ 4.47\pm0.75$ &$ 11.90\pm 0.06$ &$ -529.02\pm 00.00$ \\ 4 &$1.147117\pm0.001030$ &$ 7.45\pm1.01$ &$ 12.01\pm 0.04$ &$ -513.80\pm 143.7$ \\ 5 &$1.147169\pm0.000410$ &$ 4.25\pm0.88$ &$ 11.58\pm 0.09$ &$ -506.54\pm 57.27$ \\ 6 &$1.147249\pm0.000106$ &$ 4.63\pm0.22$ &$ 11.92\pm 0.04$ &$ -495.37\pm 14.83$ \\ 7 &$1.147312\pm0.00000^{\dagger}$ &$ 4.90\pm0.45$ &$ 11.23\pm 0.17$ &$ -486.57\pm 00.00$ \\ 8 &$1.147416\pm0.000096$ &$ 4.70\pm0.19$ &$ 11.93\pm 0.04$ &$ -472.05\pm 13.33$ \\ 9 &$1.147587\pm0.000255$ &$ 4.49\pm0.67$ &$ 11.24\pm 0.15$ &$ -448.18\pm 35.65$ \\ 10 &$1.147809\pm0.000113$ &$ 3.47\pm0.22$ &$ 11.91\pm 0.04$ &$ -417.19\pm 15.84$ \\ 11 &$1.147911\pm0.000133$ &$ 3.39\pm0.25$ &$ 11.81\pm 0.04$ &$ -402.96\pm 18.58$ \\ 12 &$1.147980\pm0.000215$ &$ 3.75\pm0.57$ &$ 12.12\pm 0.10$ &$ -393.33\pm 30.04$ \\ 13 &$1.148101\pm0.000543$ &$ 4.99\pm1.12$ &$ 11.75\pm 0.07$ &$ -376.44\pm 75.84$ \\ 14 &$1.148501\pm0.000218$ &$ 7.52\pm0.44$ &$ 11.56\pm 0.10$ &$ -320.62\pm 30.47$ \\ 15 &$1.148783\pm0.000287$ &$ 2.97\pm0.57$ &$ 11.09\pm 0.18$ &$ -281.27\pm 40.03$ \\ 16 &$1.149088\pm0.000096$ &$ 2.11\pm0.21$ &$ 12.44\pm 0.03$ &$ -238.72\pm 13.32$ \\ 17 &$1.149112\pm0.000057$ &$ 6.46\pm0.11$ &$ 12.52\pm 0.03$ &$ -235.38\pm 07.97$ \\ 18 &$1.149489\pm0.000398$ &$ 4.30\pm0.43$ &$ 12.03\pm 0.03$ &$ -182.79\pm 55.50$ \\ 19 &$1.149547\pm0.000470$ &$ 5.50\pm0.53$ &$ 12.20\pm 0.02$ &$ -174.70\pm 65.55$ \\ 20 &$1.149817\pm0.000061$ &$ 4.14\pm0.12$ &$ 12.08\pm 0.03$ &$ -137.05\pm 08.56$ \\ 21 &$1.149915\pm0.000108$ &$ 5.12\pm0.21$ &$ 12.01\pm 0.03$ &$ -123.38\pm 15.02$ \\ 22 &$1.150548\pm0.00000^{\dagger}$ &$ 0.26\pm0.0^{\ddagger}$&$ 11.21\pm 0.16$ &$ -35.13\pm 00.00$ \\ 23 &$1.150659\pm0.00000^{\dagger}$ &$ 17.85\pm0.0^{\ddagger}$&$ 12.21\pm 0.06$ &$ -19.65\pm 00.00$ \\ 24 &$1.150688\pm0.000107$ &$ 2.98\pm0.18$ &$ 12.58\pm 0.02$ &$ -15.61\pm 14.94$ \\ 25 &$1.150747\pm0.00000^{\dagger}$ &$ 4.62\pm0.78$ &$ 12.47\pm 0.32$ &$ -7.39\pm 00.00$ \\ 26 &$1.150792\pm0.000102$ &$ 1.95\pm0.18$ &$ 13.26\pm 0.03$ &$ -1.11\pm 14.20$ \\ 27 &$1.150819\pm0.00000^{\dagger}$ &$ 8.16\pm1.67$ &$ 13.46\pm 0.07$ &$ 2.65\pm 00.00$ \\ 28 &$1.150864\pm0.000126$ &$ 1.07\pm0.25$ &$ 12.88\pm 0.05$ &$ 8.92\pm 17.53$ \\ 29 &$1.150903\pm0.00000^{\dagger}$ &$ 3.65\pm2.03$ &$ 12.58\pm 0.37$ &$ 14.36\pm 00.00$ \\ 30 &$1.150962\pm0.000190$ &$ 4.21\pm0.21$ &$ 13.47\pm 0.03$ &$ 22.58\pm 26.50$ \\ 31 &$1.151063\pm0.000207$ &$ 6.68\pm0.39$ &$ 13.09\pm 0.02$ &$ 36.66\pm 28.89$ \\ 32 &$1.151113\pm0.00000^{\dagger}$ &$ 6.00\pm1.61$ &$ 12.34\pm 0.11$ &$ 43.63\pm 00.00$ \\ 33 &$1.151152\pm0.00000^{\dagger}$ &$ 3.37\pm0.83$ &$ 12.25\pm 0.08$ &$ 49.06\pm 00.00$ \\ 34 &$1.151218\pm0.000235$ &$ 7.13\pm0.38$ &$ 13.29\pm 0.02$ &$ 58.26\pm 32.68$ \\ 35 &$1.151314\pm0.000158$ &$ 6.21\pm0.17$ &$ 13.56\pm 0.02$ &$ 71.64\pm 22.06$ \\ 36 &$1.151406\pm0.00000^{\dagger}$ &$ 15.40\pm0.0^{\ddagger}$&$ 12.72\pm 0.02$ &$ 84.46\pm 00.00$ \\ \\ \hline \multicolumn{5}{l}{`$^{a}$' relative velocity with respect to $z_{\rm abs}=1.1508$.}\\ \multicolumn{5}{l}{`$^{\dagger}$' The redshift ($z$) of these components are kept fixed.}\\ \multicolumn{5}{l}{`$^{\ddagger}$' The Doppler parameter, $b$, of these components are kept fixed.}\\ \label{model.tab} \end{tabular} } \end{table*} The larger errors in the measured quantities in the present study is mainly due to higher values of the error assigned to the flux in individual pixels. Thus the analysis presented here clearly shows that the analysis used by us in Chand et al (2004, 2005) produces consistent results.\par In addition we have also performed the analysis of UVES spectra by excluding the weaker \feii lines from the blue sub-system and heavily saturated strong \feiistr lines from the red-subsystem, (see discussion in Chand et al. 2004). In this case the \chisq curve is found relatively less fluctuating as compare to the left-hand side panel of Fig.~\ref{res.fig}, and has resulted in \dela = $(0.00\pm0.26)\times 10^{-5}$. \subsection{\dela from the HARPS data} The decomposition of the absorption profiles in sub-components is expected to be better defined from the HARPS spectrum because of its superior spectral resolution. In Fig.~\ref{comp_stru.fig} we compare the profiles of the \feii lines in the red sub-system as observed with HARPS and UVES. The best multi-component Voigt-profiles fit using the UVES spectrum alone is over plotted. To fit the HARPS data we need additional components, as is apparent in the region around $-$20 to $+$30 \kms where consistent differences are seen for all profiles between the HARPS spectrum and the fit to the UVES data alone. However, the UVES spectrum has the advantage of having higher S/N. Thus, in our analysis we fitted simultaneously both HARPS and UVES data using the same component structure and the appropriate instrumental functions. We initially fitted the HARPS data and used the derived parameters to fit the UVES data. The process was repeated until the residuals along the profiles are symmetrically distributed around zero and the best-fit parameters from these two data sets are consistent with one another within measurement uncertainties. In this exercise we have not included the line \feiia (covered only in the UVES spectrum) so that our derived component structure is not artificially bias towards \dela$=0$. \par Our best-fit Voigt-profile components that simultaneously fit the HARPS and UVES spectra are shown in Fig.~\ref{fit_blueboth.fig},\ref{fit_redboth.fig} respectively for the blue and red sub-systems. The best-fit parameters are listed in Table.~\ref{model.tab}. The component identification number (C.N), redshift ($z$), velocity dispersion ($b$), and Fe~{\sc ii} column density ($N$), for each component are listed respectively in columns 1, 2, 3 and 4. The last column of the table lists the relative velocity of the components with respect to \zabs=1.1508. We find that the blue and red sub-system (Fig.~\ref{fit_blueboth.fig},\ref{fit_redboth.fig}) require respectively 3 and 6 extra components compared to the minimum number required to fit the UVES spectrum alone with $\chi^2=1$.\par We evaluate the best-fit \dela value using the high resolution HARPS spectrum for the five main Fe~{\sc ii} lines and the UVES spectrum for Fe~{\sc ii}$\lambda1608$ considering both the blue and red sub-systems simultaneously. Here it should be noted that the \feiia is crucial for \dela measurement due to its opposite sensitivity for \dela (negative $q$ coefficient) compared to the other main \feii lines. However as its observed wavelength range ($\approx 3460$\AA) is not covered by the HARPS spectral coverage (3800 - 6900\AA), we have to use it from the UVES spectrum for constraining the \dela value. The $\chi^2$ versus \dela curve is shown in the right-hand side panel of Fig.~\ref{res.fig}. The scatter seen in the \chisq curve is mainly due to the low S/N ratio and low column density of many components as can be seen from Table~\ref{model.tab} (see the discussion in Chand et al. 2004). The continuous curve gives the 4th order polynomial fit to the $\chi^2$ data points using rms minimisation. Its minimum gives \dela~=~${(0.05\pm0.24)\times10^{-5}}$, using $\chi^2_{min}+1$ statistics. This result is consistent with the Quast et al. (2004) measurement (\dela~=~${[0.01\pm0.17]\times10^{-5}}$) based on the UVES spectrum and lesser number of components. Thus in this particular case lack of information on the additional components in the UVES spectrum does not seem to affect the final result. \section{Result and discussion} \label{sect:result} In this paper, we present a very high resolution (R~=~112,000) spectrum of QSO HE 0515$-$4414 obtained using HARPS. We have used the high wavelength calibration accuracy and high spectral resolution capabilities of HARPS to address the following issues.\par We compare the lamp spectra obtained with UVES and HARPS. Using cross-correlation analysis we show that any possible relative shift between the two spectra are within 2~m\AA. Using Gaussian fits to unblended lamp emission lines, we find that the absolute wavelength calibration of HARPS is very robust with rms deviation of 0.87~m\AA~with respect to the wavelengths tabulated in Cuyper et al. (1998). This is about a factor of 4 better than that of UVES ($\sigma=4.08$ m\AA,~see Fig.~\ref{dlam.fig}). Thus the small shifts noted between the HARPS and UVES lamp spectra are well within the typical wavelength calibration accuracy of UVES. We have derived the error on \dela measurements due to the calibration accuracy alone. For UVES and HARPS spectra this is found to be respectively $\sigma=0.96\times10^{-6}$ and $\sigma=0.19\times10^{-6}$ for a typical system with three well detached components. The value obtained for the UVES spectrum is also consistent with that of HIRES (Murphy et al. 2003). \par This shows that HARPS is the ideal instrument for this kind of measurement. Unfortunately it is mounted on the 3.6~m telescope at La~Silla and only HE~0515$-$4414 is bright enough to be observed in a reasonable amount of time. This shows as well that the UVES spectra reduced (or calibrated) with the UVES pipeline and used in the literature to constrain \dela (Srianand et al. 2004 and Chand et al. 2004, Quast et al. 2004, Chand et al. 2005) do not suffer from major systematic error in the wavelength calibration. \par We have obtained the accurate multi-component structure using the higher resolution data (R~$\approx 112,000$ for HARPS compared to $\approx 55,000$ for UVES). The best fit to the profiles obtained by fitting simultaneously the HARPS data (of higher resolution) and the UVES data (of better S/N ratio) require additional components as compared to the fit using the UVES data alone (Quast et al. 2004). Using this new sub-component decomposition and both HARPS and UVES data, we find \dela~$=(0.05\pm0.24)\times10^{-5}$. This is consistent with the results derived by Quast et al. (2004) from the UVES data alone. Indeed, we have in addition re-analyzed the UVES data which was used in Quast et al. (2004) (without using the component structure from HARPS data), to estimate the effect of different independent algorithms used to obtain error spectra, to combine the data, to fit the continuum and to fit the absorption lines. We find that the best-fit parameters as well as the \dela measurement (\dela~=~${[0.10\pm0.22]\times10^{-5}}$), obtained by our independent analysis, are consistent with that of Quast et al. (2004) (\dela~=~${[0.01\pm0.17]\times10^{-5}}$).\par We note that the precision on the \dela measurement obtained using the HARPS spectrum, which is of high resolution and low S/N ratio, is similar to that obtained from the UVES spectrum, which is of lower resolution and higher S/N ratio. Therefore, the improvement in the wavelength calibration accuracy by an order of magnitude using HARPS will be effective to improve the constrain on \dela only if high S/N ratio can also be obtained. This could be possible if an instrument such as HARPS can be mounted on bigger telescopes. \section*{Acknowledgments} HC thanks CSIR, INDIA for the grant award No. 9/545(18)/2KI/EMR-I. RS and PPJ gratefully acknowledge support from the Indo-French Centre for the Promotion of Advanced Research (Centre Franco-Indien pour la Promotion de la Recherche Avanc\'ee) under contract No. 3004-3. PPJ also thanks IUCAA (Pune, India) for hospitality during the time part of this work was completed. RQ has been supported by the DFG under Re353/48.
Title: Low and intermediate mass star yields.II: The evolution of nitrogen abundances
Abstract: We analyze the impact on the Galactic nitrogen abundances of using a new set of low and intermediate mass star yields. These yields have a significant yield of primary nitrogen from intermediate mass stars. We use these yields as an input to a Galactic Chemical Evolution model and study the nitrogen abundances in the halo and in the disc, and compare them with models obtained using other yield sets and with a large amount of observational data. We find that, using these new yields, our model adequately reproduce the observed trends. In particular, these yields solve the historical problem of the evolution of nitrogen, giving the right level of relative abundance N/O by the production of a primary component in intermediate mass stars. Moreover, using different evolutionary rates in each radial region of the Galaxy, we may explain the observed N dispersion.
https://export.arxiv.org/pdf/astro-ph/0601326
\title{Low and intermediate mass star yields.\\ II: The evolution of nitrogen abundances} \author{Marta Gavil\'{a}n\inst{1}, Mercedes Moll\'{a} \inst{2} and James F. Buell\inst{3}} \offprints{Marta Gavil\'{a}n} \institute{Departamento de F\'{\i}sica Te\'{o}rica, Universidad Aut\'onoma de Madrid, 28049 Cantoblanco, Spain\\ \email{marta.gavilan@uam.es}\\ \and Departamento de Investigaci\'{o}n B\'{a}sica, C.I.E.M.A.T., Avda. Complutense 22, 28040 Madrid, Spain \\ \email{mercedes.molla@ciemat.es}\\ \and Department of Mathematics and Physics, Alfred State College, Alfred, NY 14802, USA \\ \email{BuellJF@alfredstate.edu}\\ } \date{Received ; accepted } \titlerunning{Low and intermediate mass star yields} \authorrunning{Gavil\'{a}n, Moll\'{a} \& Buell } \abstract {} {We analyze the impact on the Galactic nitrogen abundances of using a new set of low and intermediate mass star yields. These yields have a significant yield of primary nitrogen from intermediate mass stars.} {We use these yields as an input to a Galactic Chemical Evolution model and study the nitrogen abundances in the halo and in the disc, and compare them with models obtained using other yield sets and with a large amount of observational data.} { We find that, using these new yields, our model adequately reproduce the observed trends. In particular, these yields solve the historical problem of the evolution of nitrogen, giving the right level of relative abundance N/O by the production of a primary component in intermediate mass stars. Moreover, using different evolutionary rates in each radial region of the Galaxy, we may explain the observed N dispersion.} {} \keywords{ stars: -- galaxies: abundances -- galaxies: evolution-- galaxies: spirals} \section{Introduction} Most elements are created in the interiors of stars by nucleosynthesis processes \citep[see][for a review]{wall97}, starting with hydrogen and progressing toward heavy elements. These processes are called {\sl primary production}. Some elements, however, can be formed from nuclei heavier than hydrogen originally present in the star. They are called {\sl secondary}. This is the case of nitrogen, that can be created during the CNO cycle using seeds of original carbon and/or oxygen. From a theoretical point of view, it has been considered that massive stars produce secondary nitrogen \citep{pei87}, while low and intermediate mass (LIM) stars have mechanisms, like the third dredge-up and the Hot Bottom Burning processes, to produce both, primary and secondary nitrogen \citep{edm78,all79}. The third dredge-up event is a consequence of the thermal pulses in the star, and transport C and He to the outer layers. The Hot Bottom Burning occurs when the CNO cycle takes place at the base of the convective envelope. Observationally, there are several open questions about the primary or secondary character of nitrogen that up to now remain unsolved. When N and O data are represented as log(N/O) {\sl vs} log(O/H), including the galactic stars, H{\sc ii} regions for the Milky Way Galaxy (MWG), external galaxies \citep{gar95,gar99,vze98,izo99} and the high redshift data \citep[][and references therein]{pet02,pro02,cen03}, a clear positive slope appears for abundances larger than $\rm 12+log(O/H)=7.8-8$ dex which indicates a secondary behavior, but the plot shows a flat slope for low metallicities that can only be explained with a primary component of nitrogen. Taking into account that this flat slope occurs for low abundances, the first idea proposed, shared by some authors \citep{pag79,dia86,dah95}, is that observations would be reproduced if the nitrogen ejected by massive stars would be primary, while intermediate mass stars might have both primary and secondary components. Thus, some authors have tried to look for mechanisms that explain how massive stars could produce primary nitrogen. This is the case of \citet{mey02} that have recently proposed rotation as a possible source of primary nitrogen, since low metallicity stars show a bigger rotation than high metallicity ones. \cite{chi03-1} have used these yields in their chemical evolution models, concluding that they are only a lower limit for the primary nitrogen production since the Hot Bottom Burning is not considered in their calculation. In fact, \cite{chi05} find that an extra-- production of N in low metallicity massive stars by a large factor, between 40 and 200 along the mass range, is necessary to explain the data of very metal-poor halo stars since these yields do not produce a sufficient amount of primary N. Moreover, if the production of primary nitrogen would proceed from massive stars, the left side of the (N/O) {\sl vs} (O/H) plot should not show any scatter. Although some authors claim to observe \citep{izo99,pil03} this no-scatter, recent observations from low metallicity objects \citep{pet02,pro02,cen03,isr04,spi05} do show a clear dispersion. \citet{ser83} already claimed that a secondary production by intermediate mass stars must exist and suggested that the zero slope may be explained by two factors: 1) a delay in the ejection of N to the ISM due to the different mean-lifetimes of stars and 2) the gas infall effects. The advantage of taking a delay into account is that the great data scatter can be explained by considering different evolutionary states for each galaxy and, therefore, this possibility has been supported by a large number of authors \citep{vila93,pil92,pil93,vze98b,hen00}. These last ones also include gas flows --infall or/and outflow--, and low efficiency for the star formation rate (the equivalent mechanism to produce a delay) in the low evolved regions, in order to reproduce the flat slope in the (N/O) vs (O/H) plot. They conclude that the secondary production of nitrogen should dominate in high metallicity environments while the primary one should act at low metallicities. Some new yields for LIM stars have been given in \citet[][hereinafter Paper I]{gav05}, where they were adequately evaluated and calibrated by using them in a Galaxy chemical evolution model. It was shown that the results about C and O abundances adequately reproduce the Galactic and Solar Neighborhood data. The purpose of this work is to analyze the impact of these stellar yields on the nitrogen abundances. In particular, we check, using the same prescriptions of Paper I, if the contribution to N given by these yields for LIM stars is sufficient to justify the amount of primary nitrogen the observations point out. We describe the yields in Section 2, analyzing in particular the primary and secondary components of the nitrogen production. In section 3 we describe briefly the chemical evolution model. Section 4 is devoted to the results, and the conclusions are presented in section 5. \section{Low and intermediate mass yields: The secondary and primary components of nitrogen} \label{prim-t-sec} The aim of this work is the study of the nitrogen behavior, using the same set of yields as in Paper I, that we call BU yields. For comparison purposes we also take the LIM stars yields from \cite{hoe97} and \cite{mar01} that we call VG and MA, respectively. The complete table of yields BU was already given in Paper I for five metallicities: -0.2, -0.1, 0.0, +0.1 and +0.2, expressed as $\rm log(Z/Z_{\odot})$, where solar abundances are taken from \cite{gre98} \footnote{The use of these solar abundances implies that $Z_{\odot}=0.02$. Recently, \cite{asp05} have obtained lower abundances which lead to a value $Z_{\odot}=0.012$. However, these new determinations are still questioned by some authors \citep{bah05,dra05,ant05} because they do not fit the helioseismological constraints.}. We summarize the behavior of the carbon and nitrogen yields for LIM stars, as shown in Fig.~\ref{yields}. In panel a) we see that $^{12}C$ yield is extremely small for stars with mass lower than 2 $\rm M_{\odot}$, since they do not experience third dredge up events. However, stars begin to suffer these kinds of events for smaller masses at lower metallicity. In other words, in the low mass range, the metallicity and the $^{12}C$ yield are anti-correlated. When the stars have enough mass to undergo Hot Bottom Burning (HBB), the $^{12}C$ yield drops abruptly because of the conversion of carbon into nitrogen. The $^{14}N$ yield presents a local maximum in the mass range from 3.5 to 5 M$_{\odot}$, depending on the metallicity, then decreases before beginning to increase again as a function of stellar mass. The largest amount of nitrogen is produced by stars of intermediate mass because HBB and the 2$^{\rm nd}$ dredge-up occur only in stars with $M>3.5-5 M_{\odot}$. As the HBB increases the luminosity and the mass-loss rate, stars that suffer this process have shorter TP-AGB lifetimes. The local maximum occurs in the transition between stars with HBB and those without. The increase at higher masses is due to the shortened time between third dredge-up events. The yields at the lowest masses are due to the 1$^{st}$ dredge-up. The most important difference among the used yields resides in the contribution of primary and secondary components of nitrogen by LIM stars. In Fig.~\ref{prim_total} we represent the fraction of primary $^{14}N$ for the three used sets, as labeled, as a function of mass (M $\leq$ 8 $M_{\odot}$, except for MA for which M $\leq 5$ $M_{\odot}$). In panel a) we show the results for solar abundances. All of them show a similar behavior with a maximum for masses around 3.5-4 $M_{\odot}$. We must clear some points about the components of N. The only difference between primary and secondary nitrogen is the origin of the carbon atom producing it. Although the idea is conceptually clear, it is not so simple to separately compute both components. Thus, although BU and MA give the two components separately for each model, VG do not. These authors, however, show their yields in each phase of stellar evolution. If we consider that all the nitrogen created in the AGB phase is primary, about $ \sim 90$ \% of the N ejected by LIM stars will be primary. This is sometimes assumed when these yields are used. This hypothesis, that we call {\it AGB technique} leads to a primary N component excessively large and is not totally adequate. Let us return to the definition: secondary N proceeds from the burning of original $^{12}C$. If a fraction of the original carbon is burned in the pre-AGB phase, it produces secondary N. Sometimes, this gives a negative $^{12}C$ yield. But, not all the initial carbon is consumed before the AGB phase. If we take, as an example, a star of 4$M_{\odot}$ of solar abundance, that is with $X(^{12}C)=0.28\times 10^{-2}$, it has an initial $^{12}C$ abundance $4 \msun X(^{12}C) = 1.12 \times 10^{-2} \msun$ The pre-AGB phase carbon yield is $yC12_{pre} = 0.300\times 10^{-4}$, so the mass of this element present in the star before the AGB begins is: \begin {equation} M(^{12}C)=yC12_{pre}M_{ini}+M_{end} X(^{12}C) \end{equation} where $M_{end}$ is the mass of the star at the end of this first phase: 3.95 $\msun$. Therefore, there is a mass $\sim 1.118\times 10^{-2} M_{\odot}$ of $^{12}C$, from which $M_{end}X(^{12}C) \sim 1.102\times 10^{-2}M_{\odot}$ corresponds to original carbon. This implies that a quantity of the original carbon is still available to form nitrogen in the following phases. Thus, a fraction of the total nitrogen produced in the AGB phase (given by the addition of the two values given by VG in their tables denoted AGB yields and final AGB yields) may be secondary. In order to calculate this component from the total AGB yields we use the fraction, called $r$, between the secondary to the total nitrogen yield, $r=^{14}N_{S}/^{14}N$. Taking into account that the secondary N proceeds from the existing carbon used as a seed, we assume that $r$ is equal to the ratio between the old carbon and the new plus old carbon: \begin{equation} r= \frac{^{14}N_{S}}{^{14}N} = \frac{^{12}C_{old}}{^{12}C_{old}+^{12}C_{new}} \end{equation} This method (hereinafter called $r$ method), may only be applied to stars which suffer the HBB and produce primary $^{14}N$, that is, those for which the core mass before the HBB is larger than $M_{HBB}=0.8 M_{\odot}$, usually stars with $M> 3.5-4 M_{\odot}$. Otherwise, the nitrogen yield is all secondary. The results of VG shown in Fig.~\ref{prim_total} proceed from this calculation. We have then computed the integrated yields of $^{14}N$ produced by LIM stars that we present in Fig.~\ref{yields_integrados} as a function of metallicity Z. In panel a), we represent the BU results as solid circles to which we have performed a least-squares fit shown by the solid (red) line. This integrated yield for $^{14}N$, equivalent to the yield produced by a single stellar population, is located between the two other sets in this panel, with a similar dependence on Z that VG \footnote{The integrated yields for VG are slightly different than those obtained by \cite{hen00} due to the Initial Mass Function (IMF) used by us, from \cite{fer92}} but with lower absolute values. More significant however, is the dependence on metallicity of the ratio of primary to total integrated yields, $^{14}N_{P}/^{14}N$, shown in panel b). This ratio increases for decreasing metallicity for all sets, as expected, although for VG the integrated yield is quite different if we consider AGB technique than $r$ method. This metallicity effect can be easily explained: low metallicity stars have smaller radii and take longer to reach super-winds, so they have more time to experience more third dredge-up events than solar metallicity stars. As a consequence, they have more fresh $^{12}{\rm C}$ in their envelopes and they can make more primary nitrogen by the HBB process. On the other hand, due to the lower amount of original carbon, they produce less secondary nitrogen. For VG yields the ratio is almost constant at a value of 15\% when the $r$ technique, represented by the short-dashed (blue) line, is used, for metallicities greater then 0.01, although it also increases for metallicities lower than this value. While it is $ \sim 90$ \% when the $AGB$ technique, represented by a dot-short-dashed (blue) line, is used, showing a smaller variation with Z. It is interesting that the integrated yield for solar abundance in the BU case \footnote{For $Z< 0.0126$ we use the same yields set, that is, this one from $Z=0.0126$. It is always possible to extrapolate the trend obtained for the other Z sets, that we show with the straight line joining the points corresponding to $Z=0.0126 $ and $Z=0.0159$} is around 20\%, very similar to the value computed by \citet{all79} two decades ago on the basis of the observations available at that time. All these considerations indicate that the primary nitrogen appears at a different time scale in the ISM depending on the value of Z. The first primary N will be ejected when stars of 8 (5 for MA) $M_{\odot}$ die, while Z is still low. \section{The chemical evolution model and its calibration} \subsection{Description} The model used in this work is the Multiphase Chemical Evolution Model described in \citet{fer92,fer94}, in the version presented in \citet{mol05} and in Paper I. For LIM stars, we use the same yields BU than in these two last papers, and for comparison purposes those from MA and VG. For massive stars we have chosen \cite{por98} and \citet[][hereinafter PCB and WW, respectively]{woo95}. We have run different models computed with different combination of yields: BU + WW, VG + WW and MA + PCB, which we distinguish as BU, VG and MA, respectively. The nitrogen study is usually done by comparing its behavior relative to iron and oxygen, so it is very important to have a careful calibration for these two elements. Oxygen calibration was done in Paper I. The SNIa are the main manufacturer of iron. The yields for type Ia supernova (SNIa) explosions are taken from \cite{iwa99} and \cite{bra86}. The evolution of this element in the model is quasi-independent of the normal stars yields. However, since iron is mainly produced by SNIa, even if its yield is very well known, its abundance is very dependent on the method to compute the rate of these explosions. For this purpose we analyzed the results obtained with different possibilities in order to eliminate, if possible, uncertainties in the iron abundance evolution. This point is relatively important because the Age-metallicity relation and the G-dwarf metallicity distribution are usually used as calibration methods for chemical evolutions models. In our case, furthermore, we compared our results with observed stellar nitrogen abundances, most of which are given as [N/Fe], and so, we checked that the Iron evolution is adequately reproduced by our models before this comparison can be made We used three methods to compute the SNIa rates as given by the following authors: the classical one \citep{mat86,fer93}, the one given by \cite{tor89}, and other, more recent, described in \cite{ruiz00}, hereinafter named MAT, TOR and RL, respectively. The first authors estimate the SN rates by using only the Initial Mass Function. The method, well described in depth in both cited works, is summarized as follows: a proportion of the stellar masses in a given range [$M_{min}$-- M$_{max}$] will be in binary systems and a fraction of them will develop type Ia supernova. Based on this idea, a mass function for the secondary stars is computed from the original one. Finally the SNIa rate depends on the number of secondary stars that died in each time step, which implies that the time scale for the iron appears in the ISM is controlled by the mean lifetimes of these secondary stars. Actually, this time scale does not depend only on the secondary mean lifetimes, since there are other processes that also participate in the conversion of a binary system into a SNIa explosion. It is necessary to take into account the effects of the distances between both stellar components, the orbital velocities and other parameters to finally obtain the time taken for the system to explode since the moment of its formation. \cite{tor89} performed these calculations for several combinations of possible candidates of binary system or SNIa scenarios (Double Degenerate, Single Degenerate, etc...), providing the supernova rate as a function of time normalized for a binary system of 1 $\rm M_{\sun}$. All the physical processes and assumptions are included in their calculations, so we only need to include the selected functions in our code and multiply them by the number of binary systems, avoiding the need of computing the secondary and primary initial mass function as defined in the previously described method. A similar technique has also been performed more recently by \cite{ruiz00}. A numerical table has been provided to us by Ruiz-Lapuente ( private communication) with the time evolution of the supernova rates for a single stellar population, computed under updated assumptions about different scenarios and probabilities of occurrence. We have computed the supernova rates using the three methods, thus producing three models MAT, TOR and RL. These different techniques affect mostly the iron abundances, the other elemental abundances being equal for all of them. Therefore, we will compare the three type of SN rate calculations by using only BU yields in the analysis of the Iron abundance evolution, as well as in the calibration of the model (next section). We will compare the three set of LIM stars yields when N be studied, using only the RL technique and only for the comparison of the relative abundance [N/Fe] we will show the nine possible combinations of models. The main disparity among the three techniques described above resides in the different evolution of the SN rate in time. As we see in Fig.~\ref{tasas}, MAT is the technique that presents highest values of SNIa/SNII at any time, reaching the maximum at 2.5 Gyr. RL has a maximum between 2 and 5 Gyr, with values approximately 1/2 or 1/3 of those given by MAT. Nevertheless it still is within the error bar given by observations \citep{capp99,capp04,mann05}. Note that this value has been reduced for the most recent determinations compared with the oldest ones. TOR model is the only one with low values. Even if it presents a maximum before the first Gyr, this will not be seen in the results because its value is very small. From the first Gyr, SNIa/SNII has positive slope and it almost reaches the observed value at the present time. \footnotesize \begin{flushleft} \begin{table*} \begin{tabular}{lccccccc} \hline \noalign{\smallskip} Reference & Fe & C & N & O & R & Age \\ \noalign{\smallskip} \hline \noalign{\smallskip} \cite{ake04}(AKE) & X & X & --- & X & --- & --- \\ \cite{bar88} & X & --- & --- & X & --- & --- \\ \cite{bar89} & X & --- & --- & X & --- & --- \\ \cite{barry88}(BA) & X & --- & --- & -- & --- & X \\ \cite{boe99} & X & --- & --- & X & --- & --- \\ \cite{car87}(CARB) & X & X & X & --- & --- & --- \\ \cite{car98}(CARR) & X & --- & --- & --- & --- & X \\ \cite{car00-2} & X & X & X & X & --- & --- \\ \cite{cav97} & X & --- & --- & X & --- & --- \\ \cite{che00} & X & --- & --- & X & --- & X \\ \cite{cle81} & X & X & X & X & --- & --- \\ \cite{daf04}(DAF) & --- & X & X & X & X & --- \\ \cite{dep02} & X & X & X & X & --- & --- \\ \cite{ecu04} & X & --- & X & --- & --- & --- \\ \cite{edv93}(EDV) & X & --- & X & X & X & X \\ \cite{gus99} & --- & X & --- & --- & X & X \\ \cite{fri90} & X & X & --- & --- & --- & --- \\ \cite{gra00} & X & X & X & X & --- & --- \\ \cite{gum98} & --- & X & X & X & X & --- \\ \cite{isr98,isr01} & X & --- & --- & X & --- & --- \\ \cite{isr04}(ISR) & X & --- & X & X & --- & --- \\ \cite{lai85} & X & X & X & --- & --- & --- \\ \cite{mel01}, & X & X & --- & --- & --- & --- \\ \cite{mel02} & & & & & & \\ \cite{mis00} & X & --- & --- & X & --- & --- \\ \cite{nis02,nis02b} & X & --- & --- & X & --- & --- \\ \cite{red03}(RED) & X & X & X & --- & --- & X \\ \cite{roc00,roc00b}(RO) & X & --- & --- & --- & --- & X \\ \cite{rol00}, & X & X & X & X & X & --- \\ \cite{sma97}, & & & & & & \\ \cite{sma01} & & & & & & \\ \cite{shi02} & X & X & X & --- & --- & --- \\ \cite{smi01} & X & --- & --- & X & --- & --- \\ \cite{spi05}(SPI) & X & X & X & X & --- & --- \\ \cite{tom84}, & X & X & X & X & --- & --- \\ \cite{tom86,tom95}, & & & & & & \\ \cite{twa80}(TW) & X & --- & --- & --- & --- & X \\ \cite{wes00} & X & X & X & X & --- & --- \\ \hline \noalign{\smallskip} \end{tabular} \caption{References for CNO stellar abundances used for the comparison with model results.} \label{authors} \end{table*} \end{flushleft} \normalsize \subsection{Calibration of the model: Iron evolution in the Solar Vicinity} The results for iron abundance obtained with these three methods are shown in Fig.~\ref{AMR}, the Age-Metallicity Relation (AMR) and in Fig.~\ref{G_dwarf}, the G-Dwarf distribution, for the Solar Vicinity. For this comparison we have shown only BU yields, keeping in mind that the set of yields will have only small effects on this relation. Nevertheless, this model uses WW yields for massive stars and these authors claimed that this could produce too much iron, and advised in \cite{tim95} to divide the iron ejections at least by two. In order to calculate how much that WW iron excess will be, we have calculated four different models for BU yields and RL technique, where massive stars iron production is divided by 1, 1.5, 2 and 3. Results are represented in panel a) where it is clearly shown that a factor of 2 is a good compromise that we will use in panel b). In this last panel we present the Age-Metallicity relation for the three SNIa cases, where all of them are in reasonable agreement with data, given their large dispersion. Although there are very little differences between models, it can be seen that the iron appears later and takes a little more time to reach high values when MAT and RL techniques are used, than for the TOR SNIa method, but all of them reach the solar abundance. Regarding the G-dwarf distribution, represented in Fig.~\ref{G_dwarf}, differences appear mainly between TOR and the others because it provides a narrower distribution than the others. The three models are able to reproduce the low metallicity tail without showing any G-dwarf problem. In Fig.~\ref{feo} we show the relation between iron and oxygen. As before, in panel a) the BU + RL model is presented varying the massive stars iron ejection. In this case the differences are clearer than in the AMR case. We chose the model Fe/2 that we will use for the rest of the paper. In panel b) we plot the model results using BU yields with the three SNIa techniques. As in the previous case, the LIM stars yields do not change the results because oxygen is ejected by massive stars and iron is mainly produced by SNIa events. The three models have very similar behavior. \section{Results analysis: The nitrogen abundances} We devote this section to analyzing the results obtained with our models for nitrogen abundances and comparing them with the available observational data. We divide these results in four parts: a) the evolution of Nitrogen over time for the Solar Vicinity (assumed located at a galactocentric distance of 8 kpc), b) the radial distributions of elements in the disc, c) the relation of log(N/O) with the oxygen evolution and d) the relation of [N/Fe] with iron. \subsection{Time evolution of nitrogen} Fig.~\ref{abunt} shows the evolution of nitrogen with the solar abundance values --large filled symbols-- taken from \cite{gre98} -- circle--, \cite{hol01} --square-- and \cite{asp05}, --cross--, by assuming an age of 4.5 Gyr for the sun. For the interstellar medium abundances at 13.2 Gyr, large empty symbols, we use the abundances given by \cite{mey97,mey98}, --circle--, and \cite{pei99} --square--. The small open circles are abundances for objects with given stellar ages, at a radial galactocentric distance between 7.5 and 9.5 kpc. Models BU, VG and MA are represented by the (red) solid, the (blue) short-dashed and the (green) long-dashed lines, respectively. Both models VG, following the two possible techniques to calculate the proportion of primary nitrogen, techniques $r$ and $AGB$ described in Section~ \ref{prim-t-sec}, yield results indistinguishable for times larger than 1 Gyr, so we represent only the results for the first one. In panel a) MA and VG models give a greater value than BU since at the lowest metallicity their nitrogen yield are higher than the corresponding one from BU (see Fig.\ref{prim_total}). Then, once an abundance higher than $\sim$0.004 is reached, it continues increasing smoothly until the present time, reproducing both, solar and ISM, abundances. The shape shown by the three models are similar and all of them reproduce both the solar value and the ISM value. The same kind of information can also be extracted from the relative abundances represented in panel b). In panel b), we show the time evolution of log(N/O). Since there is a good agreement in fitting the abundance of oxygen for all models, \citep[see][]{gav05}, the differences in this plot must be only due to the nitrogen production. The disagreement between the different models is important for times shorter than 1.5 Gyr, when intermediate mass and massive stars are the main contributors and the distinct primary/secondary ratios effects are evident there. Model MA has a strong increase in the first Gyr due the primary component and then it flattens up. Model BU has a ratio of primary nitrogen larger than MA for all Z except for the lowest one, what produces a smoother evolution, and, finally, N remains below MA. The resulting final N/O ratios are similar in both models and in agreement with observations. The behavior mostly primary of model VG, when all AGB nitrogen is considered as such, implies a very strong increase of the abundance at the earliest times. After that, both methods give a smooth slope, reaching an absolute value around -0.5 dex higher than the observed one. The good behavior of BU yields is also evident in panel c) where log(N/C) is shown. MA model presents a maximum at the first Gyr, the decrease is due to the higher amount of carbon ejected in that model (see Paper I); so the absolute value at the present time is only marginally reached. The shape for VG model is similar than the BU one but with a nitrogen excess. All models seem to fit the solar and ISM data but the model BU is the best one in reproducing the stellar data, and, more importantly, it is the best one at fitting all of the data at the same time. \subsection {Nitrogen abundance in the Galactic disc} Now, we will explore the radial distributions of nitrogen over the galactic disc, shown in Fig.~\ref{gran}. Data correspond to H{\sc ii} regions from references as labeled in the figure and to stars from references in Table~\ref{authors}. The radial distribution is more or less reproduced within the errors by all models. Actually, the shape of the radial distribution is well fitted in all cases, independently of the absolute values, since this results as an effect of the ratio infall/SFR along the galactocentric radius, produced by the scenario of our MWG model, and therefore is rather independent of the yields used. However, the slope of the radial distribution at the two ends of the disc, in the center and in the outer regions, is a matter of discussion. Thus, \cite{vil96} claim that the gradients are not as steep in these regions as in the rest of the galactic disc. The same occurs in the inner disc where the most recent data from \cite{sma01} show that the distribution flattens. Our models have been tuned to fit these two sets of data, and due to that, the resulting overall gradient is smaller than that obtained by other authors. As can be seen in Fig.~\ref{gran} MA and VG models produce a gradient flatter than observed. In Fig.~\ref{gran}b) and c), the radial distributions log(N/O) and log(N/C) are plotted, as they are considered important for the study of different yields. In panel b) the radial distribution of log(N/O) showed by data presents a clear slope, although there is some data that shows a flatter distribution in the outer regions. A steep radial distribution for N/O is expected because oxygen is produced by massive stars. If nitrogen was ejected by massive stars, its secondary character would cause it to enter the ISM after the oxygen. Instead, if it would be ejected by intermediate stars, the time needed for their evolution would be larger. Thus, in both cases, the nitrogen appears in the ISM after the oxygen does. Once again, the large dispersion of the data prevents a clear selection of {\sl the best model}, however the BU model seems most adequate to reproduce the H{\sc ii} regions data from \cite{vil96} and \citet{fic91}. The MA model shows a radial gradient flatter than indicated by observations, with higher absolute values compared to the mean values of data, mostly in the outer disc, and VG models, as before, show higher values. The same arguments are also valid when panel c) is analyzed. In this case a slight negative gradient is shown for log(N/C). The small amount of data and the great dispersion prevent the selection of any model as better than the others, although it seems clear that MA remains below most of them as corresponds to the large production of C, and that VG lies in the upper side of the data. It is apparent that BU model shows a better behavior compared to the data. Once again we stress the importance of using adequate yields to reproduce the whole set of data at the same time. Yield BU seems to be in the adequate range of production of N, C and O, since the model appears in the zone occupied by of data in the three panels. It is necessary to remember that open dots represent stellar abundances. We have tried to select only those corresponding to young stars, but we do not know the age of the complete set of stars with available data. In this case we have preferred to use the available abundances; thus, it is possible that some data does not correspond to young enough stars. \subsection {Nitrogen {\sl vs} Iron} In this case, as the iron evolution may have also an influence over the model results, we have presented the relation between nitrogen and iron, Fig.~\ref{nfe}, with a different panel, a), b) and c), for each set of yield, MA, BU and VG, respectively. The three possible methods to compute the SNIa, RL, TOR, and MAT are shown with solid (red), short-dashed (blue) and long-dashed (green) lines, respectively, in each panel. The first thing we observe is that the effects of the different SNIa techniques are almost indistinguishable. Therefore, the main features of each model at those metallicities are due to yields. In other words, we may analyze the behavior of the nitrogen corresponding to each yield set disregarding the accuracy in the SNIa calculations. \label{nfe} All results agree in the sense that the first nitrogen to be ejected is secondary, as due to the massive stars, so the initial slope is positive and large, --although this behavior is not shown in the figure because it occurs when oxygen abundances is lower that 5 dex-- but they differ in when the slope begins to change. When the N ejected by LIM stars appear, there is a strong increase due to the change from a secondary to a primary behavior. In MA yields LIM stars eject less primary nitrogen, and later, since it is ejected as secondary for stars up to 5 $M_{\odot}$. Then the main contributors to primary N are the stars with masses between $2 M_{\odot}$ and $3 M_{\odot}$. For this reason the slope does not change until it reaches $\rm [Fe/H] = -1.5$, the moment in which these stars begin to die. When BU yields are used, the trend changes earlier in the evolution due to the contribution of the primary nitrogen ejected by stars in the range 4--8 $M_{\odot}$. As their life is so brief, the ejection occurs at $\rm [Fe/H] = -4$. From then, the slope is close to zero: the signature of primary nitrogen. The case of VG shows a behavior similar to BU. We would like to remark that the data dispersion is so great that all the models lie in the data area, regardless of their big discrepancies, although the region of the metal-rich objects ($[Fe/H] > -1.5$) is particularly well fitted in panel b) by Model BU. It is necessary to use the very low metallicity data to clarify which model works better. In fact, the most recent observations from \cite{isr04,spi05} show a slope flatter than before, (even with a negative slope) which is a behavior more consistent with our model BU than with the model obtained with MA yields. This last model might be considered acceptable when the available low metallicity data were only those from \cite{car87}, but when using the new determinations of N abundances for this kind of objects, the conclusion is that BU better reproduces the generic trend of data. It is also necessary to take into account that most of the metal-poor objects do not belong to the disc but to the halo. In this way, we represent the halo model results for the zone that infall over the disc at galactocentric distance equal to 8 Kpc, at the right panels b), d) and f) of Fig.\ref{nfe}. We see that the trend shown by the recent observations from \cite{isr04,spi05} is more compatible with BU and VG than MA. \subsection {Nitrogen {\sl vs} oxygen} Finally, we show in Fig.~\ref{no} the classical and well known graphic of the relative abundance of N {\sl vs} O as $\log{(N/O)}$ {\sl vs} $12 +\log{(O/H)}$. We show the final results for the Solar Neighborhood of the computed models. Model BU reproduces well the expected behavior of N when the whole data is taken into account. Not only the level of N is so adequate but also the shape is smoother than the one shown by the other two models. The observed trend at low metallicity can be well reproduced with BU yields because they have the appropriate primary to secondary ratio, and the adequate integrated nitrogen yields. MA yields have also a primary nitrogen component, but the integrated nitrogen yield has a metallicity dependence in the opposite way as BU for the lowest Z, so the trend shown by data can not be well reproduced. Both VG models have the right shape and are almost the same for $12 +\log{(O/H)} \ge 8$. The problem is that the integrated yields are high. It would be necessary to change the input parameters as the infall rate or the efficiencies to form stars in order to fit the solar abundances. In that case, probably, other data will not be reproduced. Only the BU model shows simultaneously the good shape and adequate absolute abundances. For comparison purposes, we have also shown the resulting model using VG yields but assuming that the Nitrogen is completely secondary. If the flat behavior were caused by massive stars, the data dispersion would be very small. A problem arises when the metal-poor objects \citep{isr04,spi05} are included in the figure, as can be seen in Fig.~\ref{no}. Some values follow the described trend over the flat line, but there exist some lower abundances which are around $log(N/O)\sim -2$. This behavior is not compatible with a primary component proceeding only from massive stars. We show in Fig.\ref{no_bis} the evolution given by BU model for four different radial regions of the Galaxy: two inner ($\sim$ 2 and 4 kpc) more evolved regions --(magenta) dotted and (blue) short-dashed lines--, the solar vicinity ($\sim$ 8 kpc) as in the previous figure, the (red) solid line, and an outer one ($\sim$ 18k pc), --the (green) long-dashed line--, where the evolution takes place slowly. In panel a) we show the results for the halo zones and in panel b) for the disk regions. In both panels we have included the stellar data for the MWG. The numbers on the graph indicate the evolutionary time, in million years, that corresponds to that point of the line. This is necessary because the $12+log(O/H)$ value is not the same for each radius at the same value of time. The halo regions have similar evolutions independent of their distance from the center of the Galaxy. All of them reproduce well the recent data from \cite{isr04,spi05} obtained for $5.5 < 12+log(O/H) < 8$, and the disk regions fit the stellar data obtained for $12+log(O/H) > 8$ of the disk. Their evolutionary tracks, however, are very different, as corresponds to their distinct input parameters (infall rates, initial gas masses, efficiencies to form stars...) which are translated into very different star formation histories. Thus, the dispersion of the MWG data can be well explained on the basis of a primary production of nitrogen from LIM stars, higher for the lowest metallicities, and with different star formation efficiencies in the different regions. We represent the same results for the disk regions in Fig.\ref{no_hii} compared with data referring to Galactic HII regions, taken from the same authors than those of Fig.~\ref{no}, but without limiting the possible galactocentric distance. Other galaxy data \citep{gar95,gar99,vze98} and \cite{izo99} are also shown. The large open triangle is the recent estimate obtained from \cite{izo05} for the lowest metallicity known galaxy. We have also added the DLA objects data from \cite{pet02,pro02,cen03} as solid points. We want to remark that the disk regions evolve in good agreement with all of them, showing the inner regions a steeper evolution while the outer one shows a very flat evolution with a high and constant value $log(N/O)\sim -1.2$ dex, similar to the behavior of dwarf galaxies. These results suggest that the observed dispersion in this kind of plot, when other galaxies data (such as dwarf or DLA galaxies) are included, might be reproduced if different star formation histories have occurred in different galaxies. This argument has already been invoked by other authors, in particular by \cite{hen00} and \cite{pra03}. It was even demonstrated by \cite{pil03}, who analyzed data for different radial regions in spiral galaxies and showed the changes of the evolutionary track in the plane N/O {\sl vs} O/H for each one of them. It is evident that this kind of behavior may be represented by our models and that the new yields may reproduce better the whole set of data. In fact, these yields have already been used in a grid of chemical evolution models for a large number of theoretical galaxies \cite{mol05}. A discussion about the resulting N/O abundances and its possible dispersion for different objects is done in \cite{mol05b}. This figure and the behavior of (C/O) vs O/H, shown in Paper I, are the main clues to consider the present yields the most adequate to represent the evolution of galaxies. The production of carbon by LIM stars is sufficient to obtain an increase in C/O without the need to invoke mass loss by massive star winds, and the N/O behavior may be well reproduced with different star formation efficiencies due to the adequate level of the primary component produced by LIM stars and to the right dependence of this component with Z. \section{Conclusions} Our conclusion can be summarized as follows: \begin{enumerate} \item The primary component of nitrogen, necessary to explain the trend of N/O with O/H, may be mostly produced by LIM stars, and adequately fits all the data, including the observed dispersion. In this way the integrated yield produced by LIM stars must be directly proportional to Z, while the ratio $N_{P}/N_{tot}$ must increase for Z decreasing. A primary component, larger for lowest metallicities, has an important effect on explaining low abundance range data. \item The dependence of the N yield on stellar mass would have a maximum around 5-6 $M_{\odot}$, while the primary component shows other around 3.5-5 $M_{\odot}$. This constraints the time where these contributions have important effects on the evolution. \item The high dispersion on N/O data for low and high metallicity galactic regions may be explained with these yields as we have demonstrated with the radial regions of the disc models which have different star formation efficiencies. Our findings go in the same address than \cite{hen00} and \cite{pra03} using VG yields, but we claim that it is easier to reproduce the whole data set when BU yields are used. Our model for MWG with small efficiencies to form stars is consistent with data. \item These results seem suggest that models with differences in the star formation histories for different types of galaxies, such as those calculated in \cite{mol05}, might produce final abundances with high dispersion, in agreement with the observed one when dwarfs galaxies or DLA galaxies are included in a plot N/O-O/H, such as we will show in \cite{mol05b}. \item As \cite{chi03-2}, we also support that the halo and the disc have different evolutions. The set of stellar data [C/Fe], [N/Fe] and [N/C] may be divided into two trends. The first one is well reproduced by our disc models, while the second one is well fitted by our halo results. \item Due to the primary N component of BU yields, and since the intermediate stars have short lifetimes, it is possible to produce high [N/Fe] abundances even at low metallicities which are in perfect agreement with the recent halo stars data obtained by \cite{isr04,cay04,spi05}. \item We claim that the fit of the whole set of data with only one model is not an easy task. We may reproduce the observed trend with BU yields combined with yields from WW. In summary, our model BU reproduce reasonably well the whole CNO data set. \end{enumerate} \begin{acknowledgements} This work has been partially supported by the Spanish PNAYA project AYA2004--8260-C03-03. We acknowledge Pilar Ruiz-Lapuente for her personal contribution in the SNIa rates data. We also thanks Jose Manuel V\'{\i}lchez for his valuable suggestions and the referees, Leonid S. Pilyugin and Angeles I. D\'{\i}az their comments that have improved this paper. \end{acknowledgements} \bibliographystyle{aa} \bibliography{bibliografia}
Title: Late Light Curves of Normally-Luminous Type Ia Supernovae
Abstract: The use of Type Ia supernovae as cosmological tools has reinforced the need to better understand these objects and their light curves. The light curves of Type Ia supernovae are powered by the nuclear decay of $^{56}Ni \to ^{56}Co \to ^{56}Fe$. The late time light curves can provide insight into the behavior of the decay products and their effect of the shape of the curves. We present the optical light curves of six "normal" Type Ia supernovae, obtained at late times with template image subtraction, and the fits of these light curves to supernova energy deposition models.
https://export.arxiv.org/pdf/astro-ph/0601088
\runauthor{Lair,et al.} \begin{frontmatter} \title{Late Light Curves of Normally-Luminous Type Ia Supernovae} \author[Clemson University]{Jessica C. Lair} \author[Clemson University]{Mark D. Leising} \author[Steward Observatory]{Peter A. Milne} \author[Steward Observatory]{G. Grant Williams} \address[Clemson University]{Department of Physics and Astronomy, Clemson University, Clemson, SC 29634} \address[Steward Observatory]{Steward Observatory, University of Arizona, Tucson, AZ 85721} \begin{keyword} supernovae \end{keyword} \end{frontmatter} \section{Introduction} Type Ia Supernovae (SNe Ia) are thought to be the thermonuclear explosion of a white dwarf \citep[see][and references therein]{2000A&ARv..10..179L}. The light curves of SNe Ia are powered by deposition in the SN ejecta of the $\gamma$-ray and positron products of the $^{56}Ni\rightarrow ^{56}Co\rightarrow ^{56}Fe$ decay \citep*{1969ApJ...157..623C}. The extreme brightness and seemingly uniform light curves of SNe Ia make them good candidates for use as standard candle distance indicators. In more recent years, it has been shown that Type Ia supernovae do not have uniform light curve magnitudes, shape or spectra. The light curves can, however, be normalized to account for this inhomogeneity, thus allowing these objects to be used at standard candle distance indicators \citep[e.g.][]{1993ApJ...413L.105P,1996ApJ...473...88R}. Between 100-200 days after the explosion the ejecta become transparent to the $\gamma$-rays and the light curve is powered by the deposition of the positron kinetic energy into the ejecta. The escape of a fraction of these positrons from the ejecta has been suggested as a possible source of the Galactic 511 keV annihilation radiation \citep*{1999ApJS..124..503M}. There are currently two methods of modeling the late emission of SNe Ia. One is radiation transport with complete and instantaneous trapping of the positrons. \cite{1980PhDT.........1A} showed, by comparing a model to the late time spectra of SN 1972E, that the ejecta will cool leading to an increased fraction of the emission coming out in the infrared, the so named ``infrared catastrophe" (IRC). Other studies of radiation transport \citep[e.g.][]{1996ssr..conf..211F,2004A&A...428..555S}, have reproduced the IRC, but they also predict the abrupt fall off of the optical light curves as the emission shifts into the NIR and ultimately into the IR, which is not seen in observed light curves. The other method consists of positron energy deposition modeling without radiation transport. In this type of modeling \citep[e.g.][]{1980ApJ...237L..81C,1997A&A...328..203C,1998ApJ...500..360R,1999ApJS..124..503M} , optical band light curves are used as tracers of the bolometric luminosity and fit to model energy deposition curves. The results show model curves with varying degrees of positron escape fitting the light curves. One weakness in this model fitting technique is in using the optical bands as tracers of bolometric. \citet{2001ApJ...559.1019M} constructed bolometric curves using BVRI bands and showed those curves roughly fitting the positron escape energy deposition curves. \section{BVRI Photometry using Template Subtraction} We preformed aperture photometry on six ``normal" SNe Ia at late epochs, SN 2000E, SN 2000ce, SN 2000cx, SN 2001C, SN 2001bg, SN 2001dp. Some of these SNe were located in very complicated regions in their host galaxies. For this reason, we chose to do template image subtraction, on all but SN 2000cx, before preforming the aperture photometry. All data reduction, image subtraction and aperture photometry was performed using the Image Reduction and Analysis Facility (IRAF) software \footnote{IRAF is distributed by the National Optical Astronomy Observatory. http://iraf.noao.edu}. The combined light curves can be seen in Figures \ref{radmodel} \& \ref{posmodel}, where they are normalized to be zero magnitude at 200days past explosion assuming an 18d rise time to peak light. The data set for SN 2000E includes photometry from \cite{2003ApJ...595..779V}, and the data set for SN 2000cx includes data from \cite{2001PASP..113.1178L}, \cite{2002PhDT........10J}, \cite{2003PASP..115..277C}, and \cite{2004A&A...428..555S}, where the data from our observations are plotted as the filled symbols. \section{Light Curve Decline Rates} The decline rates, the slope of the light curve, between 200-500 days were calculated for these light curves and the averages are shown in Figure \ref{slopes}, where the solid line is average for the six SNe. The calculated averages for B,V,R,\& I bands were 1.43 (0.07), 1.46 (0.04), 1.36(0.04), 0.95 (0.06), respectively, in magnitudes per day. The shaded bar represents the average decline rate for 16 normal/super-luminous SNe Ia from \cite{2001ApJ...559.1019M} with a $1\sigma$ error bar. In the R-band, there is a second average, represented by the dot-dashed line, which is the average decline rate leaving out SN 2000ce and SN 2001C. This was done only to show the agreement with the Milne et al. averages. As shown in Figure \ref{slopes}, the B,V,\& R bands have decline rates of $\sim 1.4$ mag/day but the I-band has a much shallower slope of 0.95 mag/day. This is in agreement with the decline rates of SN 2000cx as shown by \cite{2004A&A...428..555S}. These results suggest that a slower I-band decline rate is a general feature of the late light curves of normal/super-luminous SNe Ia, and is possibly suggesting a shift in the late emission to longer wavelengths. A major result of \cite{2004A&A...428..555S} was the constant late time emission seen in the NIR curves of SN 2000cx, which supports the idea that the emission is moving into the NIR and eventually into the IR resulting in an IRC. Our results from the analysis of these SNe reinforce the need for more observations of SNe Ia in the NIR in an attempt to reproduce what was seen in SN 2000cx and also in SN 1998bu \citep{2004A&A...426..547S} \section{Discussion} Figure \ref{radmodel} shows the combined light curves of the six SNe plotted on the radiation transport models of \cite{2004A&A...428..555S}. The V-band model light curve has been normalized to be zero magnitude at 200d along with the data. The B, R, \& I band model light curves have been adjusted so that the colors of the model are preserved. The dotted curve is the model including photoionization and the dot-dashed curve is the model without photoionization. Figure \ref{posmodel} shows the combined light curves of the six SNe plotted on the positron energy deposition curves of \cite{2001ApJ...559.1019M}. The model curves have been normalized to be zero magnitude at 200d. The solid curve is the energy deposition with the positron kinetic energy trapped and deposited into the ejecta and the dashed curve is the energy deposition curve with a radially combed magnetic field allowing a fraction of the positrons to escape the ejecta without depositing their kinetic energy. As can be seen in these figures, the shape of the B, V, \& R band light curves could be explained by either the color evolution in the radiation transport model or the escape of positrons from the ejecta, while the I-band has a slower decline rate than both models. This suggests that a model combining radiation transport with positron transport would be preferred to attempt to explain the late light curves of SNe Ia. One thing is clear from the light curves; these SNe show very little deviation from each other in a given band, implying that within this class of normally-luminous SNe Ia there is only one answer for the question of positron escape from SNe Ia ejecta.
Title: SINFONI's take on Star Formation, Molecular Gas, and Black Hole Masses in AGN
Abstract: We present some preliminary (half-way) results on our adaptive optics spectroscopic survey of AGN at spatial scales down to 0.085arcsec. Most of the data were obtained with SINFONI which provides integral field capability at a spectral resolution of R~4000. The themes on which we focus in this contribution are: star formation around the AGN, the properties of the molecular gas and its relation to the torus, and the mass of the black hole.
https://export.arxiv.org/pdf/astro-ph/0601417
\title*{SINFONI's take on Star Formation, Molecular Gas, and Black Hole Masses in AGN} \titlerunning{Star Formation, Molecular Gas, \& Black Holes Masses in AGN} \author{R.~Davies, R.~Genzel, L.~Tacconi, F.~M\"uller~Sanchez, J.~Thomas, \and S.~Friedrich} \authorrunning{R. Davies et al.} \institute{Max Planck Institut f\"ur extraterrestrische Physik, Postfach 1312, 85741, Garching, Germany} \section{The AGN Sample} \label{dav:sec:sample} The primary criteria for selecting AGN were that (1) the nucleus should be bright enough for adaptive optics correction, (2) the galaxy should be close enough that small spatial scales can be resolved, and (3) the galaxies should be ``well known'' so that complementary data can be found in the literature. These criteria were not applied strictly, since some targets were also of particular interest for other reasons. The resulting sample of 9 AGN is listed in Table~\ref{dav:tab:sample}. The observations of these are now completed, and while the data for some objects has been fully analysed, others are still in a preliminary stage. Additional AGN will likely be added once the Laser Guide Star Facility is commissioned. \begin{table} \begin{centering} \caption{AGN sample} \label{dav:tab:sample} % \begin{tabular}{llrlrll} \hline\noalign{\smallskip} Target & Classification & Dist. & \ \ \ & \multicolumn{3}{c}{Observations} \\ & & (Mpc) && Date & \ \ \ & Instrument\\ \noalign{\smallskip}\hline\noalign{\smallskip} Mkn 231$^1$ & ULIRG / Sy 1 / QSO & 170 && May '02 && Keck / NIRC2 \\ NGC 7469$^2$ & Sy 1 & 66 && Nov '02 && Keck / NIRSPAO \\ IRAS 05189-2524 \ \ & ULIRG / Sy 1 & 170 && Dec '02 && VLT / NACO \\ Circinus$^3$ & Sy 2 & 4 && Jul '04 && VLT / SINFONI \\ NGC 3227$^4$ & Sy 1 & 17 && Dec '04 && VLT / SINFONI \\ NGC 3783 & Sy 1 & 42 && Mar '05 && VLT / SINFONI \\ NGC 2992 & Sy 1 & 33 && Mar '05 && VLT / SINFONI \\ NGC 1068 & Sy 2 & 14 && Oct '05 && VLT / SINFONI \\ NGC 1097 & LINER / Sy 1 & 18 && Oct '05 && VLT / SINFONI \\ \noalign{\smallskip}\hline \end{tabular} \end{centering} $^1$ Davies et al. 2004a \cite{dav:dav04a}; $^2$ Davies et al. 2004b \cite{dav:dav04b}; $^3$ M\"uller Sanchez et al. 2006 \cite{dav:mul06}; $^4$ Davies et al. 2006 \cite{dav:dav06}; \end{table} One immediate result, which has a bearing on the classifications in the table, is the frequent detection of broad Br$\gamma$ -- i.e. with FWHM at least 1000\,km\,s$^{-1}$. An example of this is given in Fig.~\ref{dav:fig:n2992}. In only 3 galaxies was no broad Br$\gamma$ detected: Circinus, NGC\,1068, and NGC\,1097 (in which even the narrow Br$\gamma$ is so weak that it is almost lost in the stellar absorption features). \section{Star Formation} \label{dav:sec:starform} The topics we address here are the spatial scales on which stars exist around the AGN, the age and star formation history of these stars, and their contribution to the bolometric luminosity with respect to that of the AGN itself. The stellar K-band (or equivalently H-band) continuum can be distinguished from the non-stellar continuum via the depth of stellar absorption features such as the CO bandheads, because for any ensemble of stars the intrinsic depth will not vary much once late-type stars appear (see Davies et al. 2006 \cite{dav:dav06} for a more detailed discussion of this). Doing so immediately allows one to assess the physical size scale of the stellar population close to the AGN (see Fig.~\ref{dav:fig:plot3}). In addition it permits a lower limit to be put on the bolometric luminosity originating in stars. This is because, while a stellar population which is still forming stars will have $L_{\rm bol}/L_{\rm K} \sim 50$ (or even higher if it is very young), even an old passively evolving population has $L_{\rm bol}/L_{\rm K} \sim 20$. In most cases we are able to apply tighter constraints than this by considering other diagnostics. For example, from the morphology and kinematics one can estimate the fractions of the narrow Br$\gamma$ flux that are associated with stars and with the AGN's narrow line region. Similarly, it is often possible to estimate the fractions of the radio continuum associated with the AGN and stars: the former will be unresolved and have very high brightness temperatures (see Condon et al. 1991 \cite{dav:con91}). The ratio of either of these to the stellar K-band continuum can provide strong constraints on the star formation time scales and hence the bolometric luminosity from stars close around the AGN. Our preliminary results are: \begin{itemize} \item In all 9 cases we have resolved a stellar population around the AGN on the scales we have achieved (0.08--0.3$''$); and the stellar luminosity increases as one approaches the AGN. \item In the 5 cases we have analysed in detail so far (Mkn\,231, NGC\,7469, IRAS\,05189-2524, Circinus, NGC\,3227), the stellar population is young: the range of ages we find is 40--120\,Myr \item The (young) stellar luminosity is comparable to that of the AGN on scales of 1\,kpc (Mkn\,231, IRAS\,05189-2524); is 10--50\% of the AGN on scales of 50--100\,pc (NGC\,7469, NGC\,3227); and is a few percent of the AGN on scales of 10--20\,pc (Circinus). \end{itemize} \section{Molecular Gas} \label{dav:sec:torus} The H$_2$ morphologies traced by the 1-0\,S(1) line show a much greater diversity than the stellar distributions, as typified in Fig.~\ref{dav:fig:plot3}. This might be expected since it is known that distribution of gas is strongly influenced by dynamical resonances and outflows. However, when analysing the morphologies on $\sim$10\,pc scales, one needs to remember that the 1-0\,S(1) line traces only hot (typically 1000--2000\,K) gas, and hence the very local environment will have an important impact on the observed luminosity distribution: for example, is there a particularly massive star cluster nearby or has there been a recent supernova? With this caveat in mind, our preliminary results are: \begin{itemize} \item the 1-0\,S(1) emission is stronger closer to the AGN (with the exception of NGC\,1068) indicating the gas distribution is also concentrated towards the nucleus on scales of 10--50\,pc. \item the kinematics show ordered rotation (again excepting NGC\,1068) but also remarkably high velocity dispersion -- in the range $\sigma = 70$--140\,km\,s$^{-1}$, giving $V_{\rm rot}/\sigma \sim 1$. This means that the gas must be rather turbulent, most likely due to heating from the AGN and/or star formation, and as a result is probably geometrically thick. \item Given the size scales on which models predict the molecular torus around AGN should exist (10--100\,pc, e.g. most recently Schartmann et al. 2005 \cite{dav:sch05}), and the fact that the torus must have a large enough scale height to collimate ionisation cones, it is reasonable to propose that the gas we have seen in these data is associated with the torus. \end{itemize} \section{Black Hole Masses} \label{dav:sec:bhmass} Since it was first discovered, the relation between the mass of the supermassive black hole $M_{\rm BH}$ and the velocity dispersion $\sigma_*$ of the surrounding spheroid has become a cornerstone of galaxy evolution and black hole growth in the cosmological context. However, almost without exception the `reliable' black hole masses (typically based on stellar kinematics and resolving the black hole's radius of influence) have been derived only for nearby bulge dominated E/S0 quiescent galaxies (see the review by Ferrarese \& Ford 2005 \cite{dav:fer05}). While extremely challenging, it is therefore crucial to determine stellar dynamical black hole masses in AGN -- not only to verify that the $M_{\rm BH} - \sigma_*$ relation holds for galaxies which are by definition active, but to assess its scatter for these galaxies, and to provide a comparison to reverberation masses which might then allow one to constrain the geometry of the broad line region. The high spatial resolution and integral field capability of SINFONI provide an ideal combination to do this, and we have successfully derived $M_{\rm BH}$ in NGC\,3227 from stellar kinematics -- the first time for a Seyfert~1 -- using Schwarzschild orbit superposition techniques. Details of the specific code, which is based on that used by the Nuker team, are given in Thomas et al. (2004) \cite{dav:tho04}. While the inclination and mass-to-light ratios are often uncertain parameters, for NGC\,3227 they are relatively well constrained. Nevertheless, we have explored the range of values which the modelling would permit and find it to be consistent with those expected, giving us confidence that the results are physically meaningful and reasonably robust. The resulting range of permissible black hole masses is $M_{\rm BH} = 5\times10^6$--$2\times10^7$\,$M_\odot$. The range is a result of the degeneracy between the black hole mass and the `effective' mass-to-light ratio of the stellar population, which includes the contribution of the gas mass. If the gas is significantly less concentrated than the stars, then the higher $M_{\rm BH}$ is possible; on the other hand if the gas is strongly centrally concentrated in a similar way to the stars, then $M_{\rm BH}$ must be correspondingly lower. That the mass we find is within a factor of 2--3 of the masses found by other methods suggests that all are satisfactory to this level of accuracy. However, the fact that the mass is also likely to be a factor of a few below that implied by the $M_{\rm BH} - \sigma_*$ relation, while in contrast the stellar dynamical mass of Cen\,A (Silge et al. 2005, \cite{dav:sil05}) is a factor of several greater, may indicate that for AGN the scatter around this relation could be very considerable. \printindex
Title: Star formation in the nearby universe: the ultraviolet and infrared points of view
Abstract: This work presents the main ultraviolet (UV) and far-infrared (FIR) properties of two samples of nearby galaxies selected from the GALEX ($\lambda = 2315$\AA, hereafter NUV) and IRAS ($\lambda = 60\mu$m) surveys respectively. They are built in order to get detection at both wavelengths for most of the galaxies. Star formation rate (SFR) estimators based on the UV and FIR emissions are compared. Systematic differences are found between the SFR estimators for individual galaxies based on the NUV fluxes corrected for dust attenuation and on the total IR luminosity. A combined estimator based on NUV and IR luminosities seems to be the best proxy over the whole range of values of SFR. Although both samples present similar average values of the birthrate parameter b, their star-formation-related properties are substantially different: NUV-selected galaxies tend to show larger values of $b$ for lower masses, SFRs and dust attenuations, supporting previous scenarios for the star formation history (SFH). Conversely, about 20% of the FIR-selected galaxies show high values of $b$, SFR and NUV attenuation. These galaxies, most of them being LIRGs and ULIRGs, break down the downsizing picture for the SFH, however their relative contribution per unit volume is small in the local Universe. Finally, the cosmic SFR density of the local Universe is estimated in a consistent way from the NUV and IR luminosities.
https://export.arxiv.org/pdf/astro-ph/0601235
\title{Star formation in the nearby universe: the ultraviolet and infrared points of view} \author{J. Iglesias-P\'{a}ramo\altaffilmark{1,10}, V. Buat\altaffilmark{1}, T. T. Takeuchi\altaffilmark{1}, K. Xu\altaffilmark{2}, S. Boissier\altaffilmark{3}, A. Boselli\altaffilmark{1}, D. Burgarella\altaffilmark{1}, B. F. Madore\altaffilmark{3}, A. Gil de Paz\altaffilmark{3}, L. Bianchi\altaffilmark{4}, T. A. Barlow\altaffilmark{2}, Y.-I. Byun\altaffilmark{5}, J. Donas\altaffilmark{1}, K. Forster\altaffilmark{2}, P.G. Friedman\altaffilmark{2}, T. M. Heckman\altaffilmark{6}, P. N. Jelinski\altaffilmark{7}, Y.-W. Lee\altaffilmark{5}, R. F. Malina\altaffilmark{1}, D. C. Martin\altaffilmark{2}, B. Milliard\altaffilmark{1}, P. F. Morrissey\altaffilmark{2}, S. G. Neff\altaffilmark{8}, R. M. Rich\altaffilmark{9}, D. Schiminovich\altaffilmark{2}, M. Seibert\altaffilmark{2}, O. H. W. Siegmund\altaffilmark{7}, T. Small\altaffilmark{2}, A. S. Szalay\altaffilmark{6}, B. Y. Welsh\altaffilmark{7} } \and \author{T. K. Wyder\altaffilmark{2}} \altaffiltext{1}{ Laboratoire d'Astrophysique de Marseille, 13376 Marseille, FRANCE} \altaffiltext{2}{ Space Astrophysics Laboratory, Mail Stop 405-47, California Institute of Technology, 1200 East California Boulevard, Pasadena, CA 91125} \altaffiltext{3}{ Observatories of the Carnegie Institution of Washington, 813 Santa Barbara Street, Pasadena, CA 91101} \altaffiltext{4}{ Center for Astrophysical Sciences, Johns Hopkins University, 3400 North Charles Street, Baltimore, MD 21218} \altaffiltext{5}{ Center for Space Astrophysics, Yonsei University, Seoul 120-749, Korea} \altaffiltext{6}{ Department of Physics and Astronomy, Johns Hopkins University, Homewood Campus, Baltimore, MD 21218} \altaffiltext{7}{ Space Sciences Laboratory, University of California at Berkeley, 601 Campbell Hall, Berkeley, CA 94720} \altaffiltext{8}{ Laboratory for Astronomy and Solar Physics, NASA Goddard Space Flight Center, Greenbelt, MD 20771} \altaffiltext{9}{ Department of Physics and Astronomy, University of California, Los Angeles, CA 90095} \altaffiltext{10}{ Instituto de Astrof\'{\i}sica de Andaluc\'{\i}a, 18008 Granada, SPAIN} \keywords{surveys: GALEX --- ultraviolet: galaxies --- infrared: galaxies} \section{Introduction} What is the best way to measure the SFR of galaxies on large scales and at different redshifts? The possibility of estimating the SFR of a galaxy directly from the luminosity at a single wavelength would be a major advantage for anyone wanting to compute the SFR per unit volume at a given redshift. This quantity could be derived directly from the luminosity function (LF) at this wavelength and at this redshift. Under these conditions, large area surveys at single wavelengths might suffice. The recent SFR of a galaxy is often measured from the light emitted by the young stars: given their short lifetimes their luminosity is directly proportional to the rate at which they are currently forming. The UV and FIR luminosities of star forming galaxies are both closely related to the recent star formation: most of the UV photons are originally emitted by stars younger than $\sim 10^{8}$~yr, but many of these photons are re-processed by the dust present in galaxies and re-emitted at FIR wavelengths. Strictly speaking, neither of these fluxes can be used alone to estimate the SFR independently of the other one (e.g. Buat \& Xu 1996; Hirashita et al. 2003; Iglesias-P\'{a}ramo et al. 2004). Because of the previous lack of data at both wavelengths, attempts have been made using only the rest frame UV (Lilly et al. 1996; Madau et al. 1996; Steidel et al. 1999; and more recently Schiminovich et al. 2005 with GALEX data), or just FIR data (Rowan-Robinson et al. 1997; Chary \& Elbaz 2001), but only a few authors have compared both (Flores et al. 1999, Cardiel et al. 2003). The SFR estimator based on the UV luminosity suffers from attenuation by dust and it has to be corrected in order to properly trace the SFR: for UV-selected samples of galaxies the attenuation can reach more than 1~mag (e.g. Iglesias-P\'{a}ramo et al. 2004; Buat et al. 2005). On the other hand the FIR emission is not free of problems because the dust can also be heated by old stars, and can be a non-negligible correction for many star forming galaxies (Lonsdale \& Helou 1987; Sauvage \& Thuan 1992). Neither of these two indicators taken alone is an accurate estimator of the SFR except perhaps for starburst galaxies where (a) the dust attenuation was found to follow a tight relation with the slope of the spectrum at UV wavelengths (Meurer et al. 1999), thus allowing one to estimate the dust attenuation with only information on UV fluxes, and (b) the contribution to the dust emission coming from old stars can be neglected (Sauvage \& Thuan 1992). In the most general case, the best estimator of the SFR should contain combined information of the luminosities at both wavelength ranges (Hirashita et al. 2003). The UV and FIR fluxes are thus complementary for tracing star formation and it is well known that the FIR/UV ratio is a proper indicator of the dust attenuation (Buat et al. 1999; Witt \& Gordon 2000; Panuzzo et al. 2003). Although other indicators of the recent SFR of galaxies have been extensively used in the literature, a detailed discussion of their quality as SFR tracers will not be discussed in this paper. The GALEX mission (Martin et al. 2005a) is imaging the high-Galactic latitude sky at two UV wavelengths ($\lambda=1530$\AA, FUV; $\lambda=2315$\AA, NUV) and is providing the astronomical community with unprecedented data (both in quantity and quality). The UV data combined with existing FIR datasets (from the IRAS, ISO or Spitzer missions) now allow us to carry out detailed studies of the UV and FIR properties of galaxies, with special emphasis on the derivation of the dust attenuation and star formation activity in star forming galaxies. With this purpose in mind we selected two samples of galaxies: one from the GALEX-All Imaging Sky (AIS) and the other from the IRAS PSCz (redshifts, infrared and optical photometry, and additional information for 18,351 IRAS sources, mostly selected from the Point Source Catalog) and FSC (Faint Source Catalog) for which UV and FIR fluxes are available. With these datasets in hand we undertaken a study of the properties related to their emission at these wavelengths. Both samples were extracted from the same region of the sky ($\sim 600$~degrees$^{2}$, constrained by the status of the GALEX survey when this work was initiated). From the GALEX catalog we built a complete sample of galaxies down to $AB_{NUV}=16$~mag\footnote{AB magnitudes are defined as: $AB_{\nu} = -2.5 \log f_{\nu} -48.6$, where $f_{\nu}$ is the monochromatic flux density expressed in erg~s$^{-1}$~cm$^{-2}$~Hz$^{-1}$.} and cross-correlated it with the IRAS database (FSC), allowing non-detections at 60$\mu$m (fluxes lower than 0.2 Jy). The FIR-selected sample was built from the IRAS catalog (PSCz) with a limiting flux at 60$\mu$m of 0.6~Jy as the only constraint. The resulting list of PSCz sources was cross-correlated with the GALEX database, allowing again non-detections in the NUV. Both the NUV and the 60$\mu$m limits used to build the samples were chosen to allow for a very small number of non-detections and to sample the galaxy population over a large range of values of the dust attenuation. Besides an analysis of the SFR, for these NUV and FIR-selected samples of galaxies (chosen with well-defined selection criteria) can also be used to place important constraints on models designed to predict the statistical properties of galaxy populations. The attenuation of star light by interstellar dust, and its emission in the far infrared are usually computed very crudely in models of galactic evolution. Dust attenuation is usually deduced in such approaches from other quantities such as the mass gas and the metallicity (e.g. Guiderdonni \& Rocca-Volmerange 1987; Devriendt \& Guiderdoni 2000; Balland et al. 2003). The properties of large samples of galaxies observed both in the UV and FIR with clear selection criteria, such as the ones presented in this paper, provide an important statistical constraint for the calibration of the treatment of dust in such models. The first paper in this series (Buat et al. 2005) based on these samples was mainly devoted to the dust attenuation properties. In the present work we discuss various aspects relating to the NUV and 60$\mu$m emission of our sample galaxies, including systematic differences in the 60$\mu$m and NUV luminosities and we will focus on their star formation related properties. This paper is organized as follows: the samples are presented in section~2. The relation between NUV and 60$\mu$m luminosities is discussed in section~3. Section~4 is devoted to the derivation of the SFR and to a comparison of various estimators as well as to a discussion on the star formation activity related properties of the samples. The derivation of the local cosmic SFR density by different methods is discussed in section~5. The main conclusions are presented in section~6. Throughout this paper we adopt the following cosmological parameter set: $(h, \Omega_0, \lambda_0)=(0.7,0.3,0.7)$, where $h\equiv H_0/100\; (\mbox{km\,s}^{-1}\mbox{Mpc}^{-1})$. \section{Observational dataset} \subsection{NUV-selected sample \label{uvsel_sam}} In order to build the NUV-selected sample, we used 833 fields of the GALEX All sky Imaging Survey (AIS) each having an exposure time equal to or larger than 50s. We used only the central 0.5~deg radius circles in each field to ensure a uniform image quality: the resulting sky coverage is 615 deg $^2$. Within this area we selected all the galaxies of the GALEX AIS survey with $AB_{NUV} \leq 16$~mag. This bright limit was chosen in order to ensure IRAS detections of all the galaxies with attenuation larger than $\sim 0.3$~mag (for a limit of 0.2 Jy at 60 $\mu$m from the IRAS FSC using the calibration of Buat et al. 2005) and highly significant upper limits for the less attenuated galaxies. On the one hand, the moderate angular resolution of GALEX (FWHM $\sim 6$~arcsecs) does not allow for a secure discrimination of stars from small galaxies. On the other hand the GALEX pipeline can induce some shredding of larger galaxies. This results in multiple detections which (cumulatively) correspond to a single galaxy because of the misidentification of star forming regions as if they were individual galaxies. The main consequence of this is the underestimation of the fluxes of large galaxies. Corollary catalogs were thus required in order to perform a reliable selection of galaxies. Our starting point was the catalog of NUV detections produced by the GALEX pipeline\footnote{version 0.2.0, September 2003, with the correction to the NUV and FUV magnitudes reported in Buat et al. 2005}, which made use of the Sextractor code (Bertin \& Arnouts 1996). We made the assumption that all the galaxies brighter than $AB_{NUV}=$16~mag, even if they are shredded, should contain at least one Sextractor detection brighter than $AB_{NUV}=18$~mag. Then we associated an object from databases of well known stars and galaxies (SIMBAD, 2MASS, LEDA) with each of the Sextractor detections brighter than $AB_{NUV}=18$~mag. The problem of shredding was mostly resolved by using LEDA. As this database contains the optical diameters of the galaxies, NUV detections corresponding to shredded galaxies can be associated with their parent galaxies provided they are located within the aperture defined by the optical diameters and the position angle listed in LEDA. For the detections of galaxies not shredded we used SIMBAD and 2MASS in order to classify them as stars or galaxies. Finally, objects not associated with any known source were classified as ``dubious''. In order to test the quality of our selection method, we cross-correlated our final catalog with the SDSS DR1, which spatially overlap one fifth of our sample. The spectral and photometric information of the SDSS together with its higher angular resolution made possible an optimal classification of all the objects detected in the field: all the objects present in both GALEX and SDSS catalogs were found to be properly classified. Dubious objects were found to be noise detections or ghosts generated mainly near the edges of the GALEX frames. Thus we ended up with a catalog of {\em bona fide} galaxies or fragments of galaxies (i.e. belonging to the large galaxies shredded by Sextractor) brighter than $AB_{NUV}=18$~mag. The next step was to obtain aperture photometry of these objects in order to select out only the galaxies brighter than $AB_{NUV}=16$~mag. Photometry of our sample galaxies was performed in the GALEX NUV and FUV bands. Since the selection criterion for our sample galaxies was imposed in the NUV band, we also took the NUV images as our reference for the total photometry. We performed aperture photometry using a set of elliptical apertures, the total photometry corresponding to the aperture where convergence of the growth curve is achieved. Once we determined the total NUV flux, the photometry in the FUV band was obtained by performing background subtracted aperture photometry within the same elliptical aperture where convergence of the NUV growth curve was achieved. This way, the NUV$-$FUV colors are consistent. Some galaxies were contaminated by the flux of nearby bright stars or galaxies. The contaminating sources were then masked in the NUV and FUV frames in order to obtain proper NUV and FUV fluxes for our galaxies. Table~\ref{error} shows the typical uncertainties of the NUV and FUV magnitudes. NUV and FUV magnitudes were corrected for Galactic extinction using the Schlegel et al. (1998) dust map and the Cardelli et al. (1989) extinction curve. In the end, a total of 95 non-stellar objects brighter than $AB_{NUV}=16$~mag were found. One of them, classified as a QSO, was excluded of the sample. FIR fluxes at 60$\mu$m were obtained from the IRAS Faint Source Catalog (FSC, Moshir et al. 2000) for 68 of our 95 galaxies. We discarded all these sources for which the cirrus parameter (as listed by the FSC) is larger than 2 because it can result in uncertain fluxes. The Scan Processing and Integration Facility (SCANPI) was used to obtain the FIR fluxes for the remaining 27 objects. Three of these galaxies (UGC~11866, UGCA~438 and UGC~12613) were not detected at 60$\mu$m. We adopted a conservative upper limit of 0.2Jy at 60$\mu$m (as given in the FSC) for these galaxies. Four galaxies of the sample were not covered by the IRAS survey. \subsection{FIR-selected sample \label{firsel_sam}} The FIR-selected sample was extracted from the IRAS PSCz (Saunders et al. 2000) over the 509~$\deg^{2}$ in common with the GALEX AIS fields having exposure times larger than 90~sec. In order to keep only good quality FIR data we discarded those galaxies for which the probability of a correct optical identification of the FIR-selected galaxies was lower than 50\%, as listed in the PSCz. As for the NUV-selected sample, galaxies for which the cirrus parameter (as listed in the PSCz) was larger than 2 were discarded. A total of 163 galaxies were selected; all but two of them (Q00443+1038 and Q23367-0448) have published radial velocities. As galaxies were selected from the PSCz, the imposed limiting flux at 60$\mu$m was 0.6~Jy. This limit, combined with the one estimated for the GALEX AIS at NUV ($AB_{NUV} \sim 20.5$~mag, Morrissey et al. 2005), results in detections at NUV for all the galaxies with dust attenuation as large as $\sim 4.4$~mag (Buat et al. 2005). Indeed, only two galaxies (Q00443+1038 and Q00544+0347) were not detected in the NUV frames and a total of 23 were not detected in the FUV frames. The NUV and FUV photometry of the FIR-selected galaxies was performed using the same technique as for the NUV-selected galaxies. \subsection{Completeness and bias of the samples} Before drawing conclusions about the properties of the samples we have to check on just how representative they are. Because of the reduced statistics of the samples, if the sampled volumes are not large enough it could be that some luminosity ranges are oversampled or undersampled with respect to reference samples defined over larger volumes of the local Universe. We check the representative nature of our samples by building LFs and comparing them to the standard ones at $z=0$, constructed from larger samples of galaxies (NUV LF of Wyder et al. (2005) and 60$\mu$m LF of Takeuchi et al. (2003)). We calculate both NUV and 60$\mu$m LFs of our sample by the Lynden-Bell method (Lynden-Bell 1971), implemented to obtain the normalization using the formulation of Takeuchi et al. (2000). The calculation of the uncertainty is based on a bootstrap resampling method (Takeuchi et al.~(2000)). We note that the Lynden-Bell method is robust against density inhomogeneities, and hence we can trust the LF so determined (see Takeuchi et al.~2000 for details). The results are shown in Figure~\ref{nlf}. The error bars correspond to 1$\sigma$ uncertainties. The agreement between the LFs of our samples and the corresponding LFs from larger samples is very good, so we are confident that in spite of their small size our samples are representative of flux-limited NUV and FIR samples in the local Universe. We also compare the volumes from which the samples were extracted. Figure~\ref{histo_velo_all}a shows the redshift distributions for both samples. The median values are 0.013 and 0.027 for the NUV and FIR-selected samples respectively. At a first glance, this means that the FIR selection samples a volume 8 times larger than the corresponding NUV selection. However, we must recall that the redshift distribution of a flux limited sample is strongly dependent on the shape of the LF, and as shown in Figure~\ref{nlf}, the NUV and 60$\mu$m LFs are very different. In Figure~\ref{histo_velo_all}b we show the theoretical redshift distributions for NUV and FIR-selected samples with the same limiting fluxes as our two samples\footnote{Details on the calculation are given in Appendix~\ref{depth}.}. As can be seen, the theoretical median values of the redshift for both samples are consistent with the ones obtained from the observational data. A limiting magnitude of $AB_{NUV} = 18$~mag is required to obtain similar median values of the redshift distributions of both samples, as we show in Figure~\ref{histo_velo_all}c. And in this case most of the galaxies detected in NUV will not have any counterpart in FIR. This behavior of the redshift distribution can be understood intuitively. Indeed, the flux density selection procedure omits intrinsically low-luminosity objects from the sample, whereas bright objects are hardly affected by the flux selection. To the degree that the LF is well reconstructed from the flux/magnitude-limited samples, these samples can be said to be representative, with respect to the luminosity and/or flux density, and it is indeed the case for the present work. \section{Relation between $L_{60}$ and $L_{NUV}$ \label{tsu_vero}} The physical link between the FIR and UV luminosities of galaxies is rather complex. On the one hand, both are related to the light of young stars, so one expects a correspondence between the two. On the other hand, the FIR emission is due to the absorption of the UV light thereby leading to an anti-correlation. Since our samples were built to avoid upper limits -- i.e. most of the galaxies selected in NUV (or at 60 $\mu$m) are also detected at 60 $\mu$m (or in NUV) -- we are able to discuss statistically the intrinsic relation between both the two wavelengths and outline the specifics of NUV versus FIR selection effects. In Figure~\ref{l60_lnuv} we plot $L_{NUV}$ versus $L_{60}$ for both samples \footnote{Throughout the paper the NUV and 60 $\mu$m luminosities $L_{NUV}$ and $L_{60}$ will be calculated as $\nu L_{\nu}$ expressed in solar units. The adopted value for the bolometric solar luminosity is $L_{\odot} = 3.83 \times 10^{33}$~erg~s$^{-1}$.}. The two samples exhibit very different behaviors: the NUV-selected galaxies show a good correlation between both luminosities, with a correlation coefficient (in log units) of $r \simeq 0.8$. On the contrary the dispersion is very large for FIR-selected galaxies and the correlation coefficient is low: $r \simeq 0.3$. NUV-selected galaxies appear intrinsically less luminous at 60$\mu$m than FIR-selected ones. This is also true for the sum of both luminosities, $L_{tot} = L_{NUV} + L_{60}$, which is supposed to be a crude estimate of the bolometric luminosity of galaxies related to recent star formation (e.g. Martin et al. 2005b). Although the luminosity distribution within each sample is the combined effect of the LFs and selection criteria, this result is confirmed by other studies and from the comparison of the 60$\mu$m and NUV LFs themselves (e.g. Martin et al. 2005b; Buat et al. 2005). Both distributions flatten at higher luminosities, reflecting a general increase of the dust attenuation already pointed out in the literature by several authors (Wang \& Heckman 1996; Buat et al. 1999; Sullivan et al. 2001; Vijh et al. 2003). One could argue that the difference in luminosity between the two samples is a consequence of bias in the sampling. We show in Figure~\ref{l60_lnuv} the lines corresponding to our lower (upper) limit of the NUV attenuation above (below) which the NUV (FIR) selected sample is complete. Thus, the fact that only very few low-luminosity and low-attenuation FIR-selected galaxies are detected must be taken as real. Low-luminosity, high-attenuation galaxies should have been detected by our FIR survey if they were present. For the same reason, very luminous galaxies should have been detected in the NUV survey if they existed. The good correlation found between $L_{NUV}$ and $L_{60}$ for the NUV-selected galaxies has to be related to their rather low dust attenuation: in these galaxies both $L_{NUV}$ and $L_{60}$ represent a significant part of the total luminosity of the galaxies. This result holds for intrinsically faint galaxies ($L_{tot} \leq 2 \times 10^{9} L_{\odot}$). The very loose correlation found for FIR-selected galaxies may also be explained by the effect of the dust attenuation. With a mean dust attenuation larger than 2~mag, the NUV luminosity becomes a poor tracer of their total luminosity whereas the 60$\mu$m luminosity is not very different from the bolometric emission of the young stars. Some fluctuations in the percentage of NUV photons escaping the galaxies can induce large variations in the NUV observed luminosity on an absolute scale without any strong physical difference on the scale of the total luminosity of the galaxies. We make a final comment on the so-called ``UV luminous galaxies'' (UVLGs, defined as those with $L_{FUV} \geq 2 \times 10^{10}$~L$_{\odot}$ in Heckman et al. 2005). We found 3 ULVGs in our NUV-selected sample and a total of 8 (including the previous 3) in the FIR-selected sample. All but one of these galaxies are more luminous at 60$\mu$m than in the NUV (in fact most of them are LIRGs), and their attenuation is typically larger than 1~mag. This means that these galaxies are not only UV luminous but also very luminous from a bolometric point of view. \section{Selection effects on observational quantities and physical properties of galaxies} The main aim of this section is to show the effect of the selection criteria of samples of galaxies on observational and physical quantities. We will now show that the selection criteria of a sample of galaxies play an important role in defining the nature of the galaxies selected and thus, in their averaged properties. Accordingly we warn against the unqualified comparison of results obtained from samples of galaxies selected on the basis of different criteria. In order to reduce the uncertainties associated with the FIR and NUV fluxes we impose further constraints on our galaxy sample: \begin{itemize} \item Ellipticals, S0s as well as AGNs (Seyferts and QSOs) were excluded since the origin of their 60$\mu$m and NUV fluxes is clearly not associated to recent star formation. The necessdary classification information is available for most of the NUV-selected galaxies; but this turned out not to be the case for the FIR-selected galaxies, so contamination of the sample by ellipticals and/or AGN among these galaxies cannot be totally excluded. Galaxies with extraneous radio sources (from NVSS and/or FIRST) within the IRAS beam were also excluded since part of the FIR flux of these galaxies could be due to contaminating background objects. \item Multiple galaxies, not resolved by the IRAS beam but clearly resolved into various components in the GALEX frames were excluded, since a one-to-one 60$\mu$m-NUV association is not possible for them. \end{itemize} After applying these criteria we ended up with 59 and 116 galaxies from the original NUV and FIR-selected samples, respectively. Hereafter we will use these restricted subsamples for our subsequent analysis of the star formation related properties, although we will keep the terminology FIR and NUV-selected samples to refer to the restricted subsamples. Given that all the galaxies were extracted from the same region of the sky, some of them belong to both subsamples. Their basic properties are listed in Table~2: (1) Identifier of the galaxy; (2) ``Y'' (``N'') indicates whether the galaxies is included (or not) in the NUV-selected sample; (3) ``Y'' (``N'') indicates whether the galaxies is included (or not) in the FIR-selected sample; (4) R.A.(J2000 equinox) of the source ; (5) Declination (J2000 equinox); (6) Radial velocity in km~s$^{-1}$, obtained from NED or LEDA; (7) Distance to the source in Mpc, corrected for the Local Group Infall to Virgo and $H_{0} = 70$~km~s$^{-1}$~Mpc$^{-1}$; (8) Morphological type, from NED or LEDA; (9) IRAS identifier: ``F'' for FSC; ``Q'', ``O'' or ``R'' for PSCz; ``SCANPI'' for absence in both catalogs. Table~\ref{photo} gives some useful photometric data for the galaxies in the restricted subsamples: (1) Optical identifier; (2) NUV magnitude corrected for Galactic extinction; (3) FUV magnitude corrected for Galactic extinction; (4) Flux density at 60$\mu$m in Jy; (5) Flux density at 100$\mu$m in Jy; (6) $H$ magnitude from 2MASS Extended Source Catalog. For galaxies with no detection by 2MASS we adopt the limiting value of $H = 13.9$~mag (3~mJy) as given by Jarrett et al. (2000); (7) NUV attenuation in mag, derived as in Buat et al. (2005); (8) FUV attenuation derived as indicated in Buat et al. (2005). For some galaxies Eqs.~\ref{attenuv_eq} and \ref{attefuv_eq} gave negative values of the NUV and FUV attenuations which is, of course, unphysical. In fact, this is an artifact of the polynomial fitting used to derive a functional form for the attenuation in Buat et al. (2005). Throughout this paper they will be considered as zero. Given that the FIR fluxes were extracted from different catalogs (PSCz for the FIR-selected sample and FSC/SCANPI for the NUV-selected sample), for those galaxies belonging to both samples we list the FIR entries corresponding to the PSCz catalog. For those galaxies not present in the FSC and not detected by SCANPI at 60$\mu$m, we list $f_{60} \leq 0.2$Jy, which is the nominal limiting flux of the FSC. No estimate of an upper limit at 100 $\mu$m is given for these galaxies. For galaxies with no detection in 2MASS we adopt the limiting value of $H = 13.9$~mag (3~mJy) as published by Jarrett et al. (2000). In Table~\ref{sfr_tab} we list some star-formation properties which will be used in the forthcoming discussion: (1) Identifier of the galaxy; (2) $SFR_{NUV}$ from Eq.~\ref{sfrnuv_eq}; (3) $SFR_{FUV}$ from Eq.~\ref{sfrfuv_eq}; (4) $SFR_{dust}$ from Eq.~\ref{sfrfir_eq}; (5) $SFR_{tot}(NUV)$ from Eq.~\ref{sfrtot_eq}; (6) $SFR_{tot}(FUV)$ from Eq.~\ref{sfrtot_eq} but modified by using $SFR^{0}_{FUV}$ instead of $SFR^{0}_{NUV}$; (7) $\left< SFR \right>$ averaged over the galaxy's lifetime, estimated as indicated in Appendix~\ref{apendice_b}. \subsection{SFR derivations} This section is devoted to a detailed comparison of the recent SFR as seen in the FIR and NUV-selected samples. Although other estimators of the recent SFR can be found in the literature (see Kennicutt 1998 for an interesting review on several methods to derive the SFR), we focus on only two of them, those using the NUV and FIR fluxes. Our aim is to compare commonly used recipes to derive SFR from the UV and FIR luminosity of the galaxies. Therefore we will make very classical calculations, as described below. For consistency we re-derive the calibrations in a homogeneous way, adapted to the GALEX bands: the formulae are found to be very similar to those of Kennicutt (1998). The underlying physical justification for deriving the SFR of a galaxy from the UV luminosity is the following: most of the UV photons emerging from a galaxy originate in the atmospheres of stars younger than $\sim 10^{8}$~yr. Thus, the SFR is proportional to the UV luminosity emitted by the young stars under the assumption that the SRF is approximately constant over this timescale. This is reasonable given that Salim et al. (2005) and Burgarella et al. (2005) found that the intensity of the youngest burst in large samples of nearby galaxies contributes typically less than 5\% to the total. However, the presence of dust absorbs a part of the UV light escaping from galaxies and breaks down the proportionality between the SFR and the observed UV luminosity. As star-forming galaxies may present a large variety of relative geometries between stars and dust, the scattering of the stellar photons through the interstellar medium may introduce a fraction of them in the line of sight before they escape the galaxy. Thus, the effect of the dust differs from a pure extinction but is a complex combination of absorption and scattering. Following Gordon et al. (1997) we will use the term `dust attenuation' for this global process at work in galaxies. The most commonly accepted method to estimate the dust attenuation at UV wavelengths is to use the ratio of FIR-to-UV fluxes (Buat \& Xu 1996; Meurer et al. 1999; Gordon et al. 2000). Several analytical expressions are already available in the literature for different UV wavelengths (Panuzzo et al. 2003; Kong et al. 2004; Buat et al. 2005). All these expressions are fairly consistent except at high values of the dust attenuation, where some dispersion appears (e.g. Meurer et al. 1999, Kong et al. 2004, Buat et al. 2005 at $\lambda \sim 1500$\AA.). In this work we use the prescription of Buat et al. (2005) to obtain the corrected NUV and FUV luminosities: \begin{equation} A_{NUV} = -0.0495 x^{3} + 0.4718 x^{2} + 0.8998 x + 0.2269 \label{attenuv_eq} \end{equation} where $x = \log L_{IR}/L_{NUV}$ and \begin{equation} A_{FUV} = -0.0333 y^{3} + 0.3522 y^{2} + 1.1960 y + 0.4967 \label{attefuv_eq} \end{equation} where $y = \log L_{IR}/L_{FUV}$. Once the observed NUV and FUV luminosities have been corrected for dust attenuation, the SFRs can be derived using the following expressions\footnote{This formula has been derived from Starburst99 (Leitherer et al. 1999) and assuming solar metallicity, and a Salpeter IMF from 0.1 to 100~$M_{\odot}$.}: \begin{equation} \log SFR_{NUV} (M_{\odot}~yr^{-1}) = \log L_{NUV,corr} (L_{\odot}) - 9.33 \label{sfrnuv_eq} \end{equation} \begin{equation} \log SFR_{FUV} (M_{\odot}~yr^{-1}) = \log L_{FUV,corr} (L_{\odot}) - 9.51 \label{sfrfuv_eq} \end{equation} In Figure~\ref{sfrnuvsfrnuvfuv} we show the ratio of $SFR_{NUV}/SFR_{FUV}$ as a function of $L_{tot}~(= L_{NUV} + L_{60})$, which traces the bolometric luminosity related to recent star formation and has the advantage of being a purely observational quantity. As this figure shows, both quantities are equivalent with a dispersion of about 20\%. Since our sample is NUV-selected, hereafter we will use NUV as our reference wavelength for star formation related properties. The luminosity at IR wavelengths provides a different avenue to the derivation of the SFR. Dust absorbs photons at UV wavelengths and re-emits most of them at IR wavelengths ($8 - 1000\mu$m). Under the hypothesis that all the UV photons are absorbed by dust, the IR luminosity would be a direct tracer of the SFR of a galaxy. One source of uncertainty is the difficulty in estimating the total IR luminosity from the FIR flux at only one or two wavelengths. In this paper we use the prescription of Dale et al. (2001) and derive $L_{IR}$ by using $f_{60}$ and $f_{100}$. For the galaxies for which only $f_{60}$ is available we use the mean value of $f_{60}/f_{100}$ estimated using the galaxies detected at both wavelengths. If we assume the same scenario as for Eq.~\ref{sfrnuv_eq}, the SFR can be expressed as: \begin{equation} \log SFR_{dust} (M_{\odot}~yr^{-1}) = \log L_{IR} (L_{\odot}) - 9.75 \label{sfrfir_eq} \end{equation} However, Eq.~\ref{sfrfir_eq} is a good approximation only for the most extreme starbursts, since many of the FIR-selected galaxies are, in fact, detected at UV wavelengths. A further limitation of this method concerns the fraction of the total IR luminosity heated by old stars (the cirrus component, hereafter represented by $\eta$), which should be removed before applying Eq.~\ref{sfrfir_eq}. This quantity is known to depend on the morphological type of galaxies (Sauvage \& Thuan 1992), but a precise estimate for individual galaxies is subject to large uncertainties (Bell~2003). The SFRs estimated from these methods are often compared in the literature for individual objects or for large samples of galaxies. In order to see whether they are consistent with each other we show here a comparison of the two using the galaxies of our two samples. Figure~\ref{sfrnuvdustb} shows the ratio $SFR_{NUV}/SFR_{dust}$ as a function of $L_{tot}$; each sample shows a different behavior. For the NUV-selected sample (blue, filled circles), $SFR_{NUV}$ is always larger than $SFR_{dust}$ (and the ratio can be as high as 3) but the discrepancy is lowered as $L_{tot}$ (and $A_{NUV}$) increase. This result is expected since we have seen in Section~3 that low luminous galaxies are brighter in the NUV than at 60$\mu$m. This affirms that $SFR_{dust}$ cannot give a proper estimation of the SFR for these galaxies. The FIR-selected galaxies extend the trend found for the NUV-selected sample to higher luminosities. For values of $L_{tot} \geq 3 \times 10^{10}$ (and for higher values of the dust attenuation), where no NUV-selected galaxies are present, $SFR_{dust}$ systematically exceeds $SFR_{NUV}$ by a factor of $\sim 2$. One reason that could play a role in this inconsistency between the two estimators is that the dust attenuation is not properly estimated for very dusty galaxies. In any case it does not make sense to use the corrected UV luminosity to measure the SFR for these IR bright galaxies. In fact, Charmandaris et al. (2004) have reported decoupled IR and UV emissions for some dusty galaxies, which could be at the basis of the discrepancy found between $SFR_{NUV}$ and $SFR_{dust}$ found in this work for galaxies with large attenuation. The conclusion of this analysis seems to be that $SFR_{NUV}$ is a good tracer of the SFR for low values of the attenuation, and in the opposite extreme $SFR_{dust}$ must be used for very heavily attenuated galaxies. There is no obvious way to delimit these two different regimes, or to chose which and which of the two indicators should be used in the intermediate cases. And so we warn users about any undiscriminated comparison of $SFR_{NUV}$ and $SFR_{dust}$ for samples of galaxies selected with different criteria. An alternative tracer of the SFR containing information from NUV and IR luminosities has already been discussed in the literature (Hirashita et al. 2003; Iglesias-P\'{a}ramo et al. 2004, Bell 2003): \begin{equation} SFR_{tot} = SFR^{0}_{NUV} + (1 - \eta) \times SFR_{dust} \label{sfrtot_eq} \end{equation} where $\eta$ accounts for the IR cirrus emission and $SFR^{0}_{NUV}$ is obtained following Eq.~\ref{sfrnuv_eq} but using $L_{NUV,obs}$ (that is the observed NUV luminosity) instead of $L_{NUV,corr}$. This estimator has the advantages of being free of the model dependence of the attenuation correction, and it contains information of the observed NUV and the IR luminosities. One limitation of this estimator, $\eta$, is the adopted value of the IR cirrus contribution. Hirashita et al. (2003) and Iglesias-P\'{a}ramo et al. (2004) reported a value of $\eta \sim 0.4$ for normal disk galaxies. Accurate values of $\eta$ for individual galaxies are not easily obtained and instead, averaged values are often used. However, this parameter is strongly dependent on the sample of galaxies under study and cannot be easily generalized. Whereas an average value of $\eta \sim 0.4$ seems to apply for normal disk galaxies, a value of $\eta \sim 0$ seems to better represent the properties of starbursts (Hirashita et al. 2003). Bell (2003) also proposed a cirrus correction for a compilation of galaxies from the literature with FUV, optical, IR and radio luminosities. He found $\eta \sim 0.32 \pm 0.16$ for galaxies with $L_{IR} \leq 10^{11}$~L$_{\odot}$ and $\eta \sim 0.09 \pm 0.05$ for galaxies with $L_{IR} > 10^{11}$~L$_{\odot}$. For our NUV-selected sample (similar to the normal star forming galaxies of Hirashita et al.) a value of $\eta \sim 0.2$ gives similar values for $SFR_{NUV}$ and $SFR_{tot}(NUV)$. Although our NUV-selected sample must contain galaxies more active than that of Hirashita et al. (since their selection is based on optical fluxes rather than on UV fluxes), this result gives an idea of the uncertainties related to the determination of $\eta$. For practical issues, throughout this paper we will use the value of $\eta$ of Bell (2003) -- not far from that of Hirashita et al. (2003) -- when computing $SFR_{tot}$, but keeping in mind that the uncertainties reported by this author are of the order of 50\%. Another limitation of $SFR_{tot}$ is that it depends on the wavelength at which we measure the UV flux. In order to illustrate this point we show in Figure~\ref{ldustnuvldustfuv} the ratio of $SFR_{tot}(NUV)/SFR_{tot}(FUV)$ as a function of $SFR_{tot}(NUV)$ for both samples. As can be seen, for the NUV-selected galaxies $SFR_{tot}(NUV)$ is systematically larger than $SFR_{tot}(FUV)$ by about $20\%$. This discordance for the NUV-selected galaxies is due to the fact that the UV attenuation is not grey: $A_{NUV} \leq A_{FUV}$ for most galaxies (see Buat et al. (2005) and Table~\ref{photo}), and since we showed in Figure~\ref{sfrnuvsfrnuvfuv} that $SFR_{NUV} \approx SFR_{FUV}$, it is obvious that $SFR^{0}_{NUV} \geq SFR^{0}_{FUV}$. On the contrary, for the brightest FIR-selected galaxies the agreement between $SFR_{tot}(NUV)$ and $SFR_{tot}(FUV)$ is good since for these galaxies $SFR_{tot}$ is dominated by $SFR_{dust}$. We conclude that $SFR_{tot}$ is stable to within 20\% for whatever UV wavelength at which we measure the UV flux. We compare now $SFR_{tot}(NUV)$ to the classical estimators $SFR_{NUV}$ and $SFR_{dust}$, in order to set their domain of applicability. Figure~\ref{sfrtotsfrnuv}a shows the comparison between $SFR_{NUV}$ and $SFR_{tot}$. At low values of the SFR both quantities are almost identical for the NUV-selected galaxies. This is expected since for these galaxies both $A_{NUV}$ and $L_{IR}$ are almost negligible and $SFR_{NUV} \approx SFR_{NUV}^{0}$. As the SFR grows, we note an increase of $SFR_{NUV}$ with respect to $SFR_{tot}$, but always within $\sim 15\%$. This increase could be due to the choice for the cirrus correction and/or to the fact that $A_{NUV}$ does not exactly corresponds to the dust emission (since factors other than absorption do play a role in the attenuation, like for example the relative geometry between stars and dust). Finally, the NUV-selected galaxies with the largest values of SFR show a decrease of $SFR_{NUV}$ with respect to $SFR_{tot}$. We stress that these galaxies have $L_{IR} > 10^{11}$~L$_{\odot}$ and so their cirrus correction is different from for the rest. All in all we find that for the NUV-selected galaxies, basically those with $SFR_{NUV} \leq 15$~$M_{\odot}$~yr$^{-1}$, $SFR_{NUV}$ and $SFR_{tot}$ are equivalent to within $\sim 15\%$. The FIR-selected galaxies show a different behavior. Whereas those with $L_{IR} < 10^{11}$~L$_{\odot}$ show an $\sim 15\%$ excess of $SFR_{NUV}$ with respect to $SFR_{tot}$, similar to the NUV-selected galaxies, for those with $L_{IR} > 10^{11}$~L$_{\odot}$, $SFR_{NUV}$ is well below $SFR_{tot}$. This is easily understood as a consequence of the already mentioned discrepancy between $SFR_{dust}$ and $SFR_{NUV}$ for galaxies dominated by their IR emission. In Figure~\ref{sfrtotsfrnuv}b we compare $SFR_{dust}$ and $SFR_{tot}$. The NUV-selected galaxies follow a very dispersed trend with $SFR_{dust}/SFR_{tot}$ increasing with $SFR_{tot}$. This behavior is due to the fact that $SFR_{dust}$ lacks the UV contribution which is dominant in these galaxies. The FIR-selected galaxies obey two different trends: for galaxies with $SFR_{tot} \leq 15$~$M_{\odot}$~yr$^{-1}$ the ratio $\log SFR_{dust}/SFR_{tot} \sim 0$, although with a dispersion of $\sim 0.2$~dex. This large dispersion is due to the contribution of the NUV luminosity to $SFR_{tot}$, which is important for the less attenuated galaxies. On the contrary, at large values of $SFR_{tot}$ the average value of $\log SFR_{dust}/SFR_{tot} \sim 0.04$~dex with a very small dispersion. This is a consequence of the fact that most of these galaxies have $L_{IR} > 10^{11}$ and are dominated by their IR emission, so the difference between $SFR_{dust}$ and $SFR_{tot}$ corresponds basically to the cirrus correction applied to $SFR_{tot}$, which is minimal. Bell (2003) proposed a calibration of the SFR similar to the one described in Eq.~\ref{sfrtot_eq} but using FUV as the reference UV wavelength. His method is based on the relation he found between $L_{IR}/L_{FUV}$ and $L_{IR}$ ($L_{IR}/L_{FUV} \sim \sqrt{L_{IR}/10^9}$) for a compilation of galaxies from the literature with FUV, optical, IR and radio luminosities. One can see in Figure~\ref{lfuvtirltir} that our NUV-selected galaxies follow well the Bell's relation whereas it is not the case for the FIR-selected sample. The galaxy sample used by Bell is therefore closer to a UV selection than to an IR one. Again we emphasize the importance of the selection biases in deriving SFRs. The overall conclusion emerging from this study is that $SFR_{tot}$ seems to be a proper estimator of the SFR of galaxies whatever their dust content is, since it avoids the main problems of the clasical estimators $SFR_{NUV}$ and $SFR_{dust}$ and is consistent with them within their respective domains of applicability to within $\sim 15\%$. We again warn against indiscriminate comparisons of the SFR of galaxies estimated from these classical estimators since the results could be strongly affected by selection biases as we have illustrated in this section. The combined uncertainty of $SFR_{tot}$ due to the choice of the UV wavelength at which we measure the UV flux and to the cirrus contribution to the IR luminosity is $\lesssim 55\%$. Throughout this paper, we adopt $SFR_{tot}(NUV)$ as our proxy to trace the recent SFR. \subsection{Star formation history} The determination of the SFR of a galaxy gives information about the total number of young stars that are being formed. But this does not necessarily mean that the light coming from this galaxy is dominated by these young stars given that most galaxies are composed of a mixture of various stellar populations of different ages. This parameter is of major importance in understanding the SFH of the Universe. Recent results based on large amounts of SDSS data suggest that the higher the mass of a galaxy, the earlier its stars were formed (Heavens et al. 2004), thus supporting the so-called ``downsizing'' explanation for the SFH of galaxies already proposed by several authors (e.g. Cowie et al. 1996; Brinchmann \& Ellis 2000; Boselli et al. 2001). We devote this section to the study of the SFH of the galaxies in our samples. A quantitative estimation of the SFH of a galaxy requires information relating the relative contribution of young and old stars. The birthrate parameter (hereafter $b$) has been proposed as a quantitative estimator of the recent SFH of a galaxy (Scalo 1986). It is defined as the ratio between the current and the past-averaged SFR: \begin{equation} b = \frac{SFR}{\left< SFR \right>} \label{b_eq} \end{equation} Since $b$ depends on the overall SFH of the galaxy, an accurate estimation from observational quantities is complex and involves several approximations. A detailed derivation of $b$ following the prescriptions of Boselli et al. (2001) can be found in Appendix~\ref{apendice_b}. As explained in Section~4.1, the NUV luminosity is sensitive to the SFR over a timescale of $\sim 10^{8}$~yr, and thus $b$ is not sensitive to shorter-timescale variations in the SFH. However, this is not a serious problem since Burgarella et al. (2005) have shown that less than 20\% of the galaxies in either sample have bursts younger than $10^{8}$~yr. Figure~\ref{histob} shows the distributions of $b$ for both samples of galaxies. The median values of both distributions are similar: 0.50 and 0.58 for the NUV and FIR-selected galaxies respectively. In Figure~\ref{bvssfrtot}a we show the relation between $SFR_{tot}$ and $b$ for both samples. In the range of overlap between the two samples (approximately $0.5 \leq \log SFR_{tot} \leq 1.5$) the values of $b$ are consistent, but beyond this region two different trends are seen: the NUV-selected galaxies show no trend of b with the SFR, whereas the FIR-selected galaxies show an increase of $b$ for high SFR. This bimodal behavior of $b$ is also seen in Figure~\ref{bvssfrtot}b, where $b$ is plotted as a function of the attenuation. Again, for the NUV-selected galaxies $b$ at lower values of the attenuation, although this trend is very dispersed. The opposite holds for the FIR-selected galaxies, with galaxies with high $b$ being the most attenuated. Thus, the picture emerging from this study is that galaxies dominated by young stellar populations fall into two categories: those showing low SFRs and low attenuation, which naturally appear in UV surveys, and those with high SFRs and large attenuation, mainly detected in FIR surveys. \subsection{The link between the $H$ luminosity and star formation properties of galaxies} The baryonic mass of galaxies is a key parameter in understanding their evolution. It has been proposed as the parameter which governs the SFH, rather than the morphological type, for example (Boselli et al. 2001). In addition, it is often used to derive some properties like the dust attenuation in semi-empirical models of formation and evolution of galaxies. For this reason we devote this section to a discussion of the effects of the sample selection on the relation between the mass and the star formation related properties of galaxies. As explained in the previous section, we will use the $H$-band luminosity as a tracer of the galaxy mass. First we show in Figure~\ref{lhsfrtot} $SFR_{tot}$ as a function of the $H$-band luminosity. Both samples show a positive relation between these two quantities, which means that more massive galaxies are also currently forming more young stars. This result is expected since we are comparing two extensive quantities. However, whereas the relation followed by the NUV-selected galaxies shows a small dispersion, the FIR-selected galaxies exhibit a more dispersed relation, especially at the most massive end. At high galaxian masses the range in SFR spans almost two orders of magnitude, which is not seen in the NUV-selected sample. In Figure~\ref{hvsanuv} we show the dust attenuation as a function of the $H$-band luminosity. Two different trends are seen in this plot. The NUV-selected galaxies show a fairly good correlation between the two quantities, with the dispersion increasing towards high $H$-band luminosities. On the contrary, the FIR-selected galaxies span an interval of almost 5~mag in dust attenuation and no correlation at all is shown with the galaxian mass. While the trend followed by the NUV-selected galaxies could be interpreted as a result of the mass -- metallicity relation reported for samples of spiral and irregular galaxies (Garnet \& Shields 1987; Zaritsky 1993) in the sense that more metallic galaxies contain more dust, there is no simple explanation for the lack of any trend shown by the FIR-selected galaxies. Finally, we show in Figure~\ref{lh_b} the $b$ parameter as a function of the $H$-band luminosity. The NUV-selected galaxies follow the classical trend that low-luminosity galaxies have larger values of $b$ (e.g. Boselli et al. 2001). Some of the FIR-selected galaxies also follow this trend, although about 20\% of them that show large masses and large values of $b$. As shown in Figure~\ref{hvsanuv}, these galaxies are among the most attenuated of the FIR-selected sample. Overall, our galaxies are shifted towards higher values of $b$ with respect to the sample of galaxies of Boselli et al. (2001). These authors adopt a slightly different IMF than we do ($M_{up} = 80$~M$_{\odot}$ against 100~M$_{\odot}$) and different evolutionary synthesis codes. Nevertheless, the large shift in $b$ found between the samples can probably not be explained by these differences alone. The correction for dust attenuation could also partially explain the shift in $b$, since Boselli et al. assume average values of 0.20~mag for Sds and later types, and 0.60~mag for types earlier than Sd. For our NUV and FIR-selected samples, the values of the dust attenuation estimated from the FIR/UV flux ratio of each of the individual galaxies show higher averaged values for the two categories of morphologies than those of Boselli et al., which would imply higher SFRs. However, this effect is diluted by the fact that for many objects in their sample, Boselli et al. estimate the SFR as the mean value of $SFR_{\rm H \alpha}$ and $SFR_{UV}$. A further factor that could be responsible for the shift in $b$ is the different selection effect of the sample: the sample of Boselli et al. (2001) is drawn from the nearby clusters Virgo, Cancer, Coma and A1367 and from the Coma-A1367 supercluster. Although not a unique selection criterion was applied, this sample can be defined as an optically selected sample of galaxies with a normal H{\sc i} content. Thus, in their sample there is a non-negligible fraction of Sa-Sab, bulge-dominated galaxies, which tend to lower the average value of $b$ (see their Figure~2). Since our selections are based on NUV and FIR fluxes, we argue that we are surely avoiding these kind of objects. Anyway, one important point is that the NUV-selected galaxies follow the same relation between mass and $b$ as the optically selected ones (disregarding the absolute calibration of both quantities) and that a fraction of the FIR-selected galaxies do not follow this trend. We have seen that the relation of the star-formation-related properties with the mass of galaxies strongly depends on the selection procedure of the sample: whereas for NUV-selected galaxies low-luminosity galaxies are also low mass, show low attenuation and have high values of $b$. A selection based on the FIR fluxes yields a different result: a population having high attenuation, high mass and strong star-formation activity appears. This population is absent in the NUV-selected sample. Since these galaxies present very high values of the attenuation (most of them are LIRGs and/or ULIRGs), their UV (and optical) fluxes are strongly dimmed and for this reason they are often excluded from flux limited surveys. However, even if these galaxies show such extreme properties, they do not put into question the downsizing picture for the SFH of galaxies since their contribution to the local cosmic SFR density is very low (see Takeuchi et al. 2005). \section{The local cosmic SFR density from different estimators} We saw in Section~4.1 that a proper estimation of the SFR is not possible with information restricted only to either NUV or FIR fluxes. However, big surveys usually provide large amounts of data only at single wavelengths, thus an estimation of the density of SFR over cosmological volumes has to be carried out under these constraints. The accuracy of the cosmic density of SFR has already been discussed by Hirashita et al. (2003) using FOCA UV data. In this section we make a similar analysis using the new GALEX data. The usual way to estimate the average SFR density is to construct the monochromatic LF and then to weight the corresponding contribution to the SFR at a given luminosity with the probability of finding a galaxy with this luminosity. \begin{equation} \rho_{\lambda} = \kappa_{\lambda} \int L_{\lambda} \phi(L_{\lambda}) dL_{\lambda} \label{sfrdensity} \end{equation} where $\rho_{\lambda}$ is the SFR density estimated from the flux at a given $\lambda$, $\kappa_{\lambda}$ is the conversion factor between $SFR_{\lambda}$ and $L_{\lambda}$, and $\phi(L_{\lambda})$ is the LF. A simple calculation using Eqs.~\ref{sfrnuv_eq}, \ref{sfrdensity} and the NUV LF of Wyder et al. (2005) yields a value of $\rho_{NUV} = 0.009^{+0.007}_{-0.004}$~M$_{\odot}$~yr$^{-1}$~Mpc$^{-3}$ for the local cosmic SFR density. Unfortunately, there is not a simple relation linking the observed NUV luminosity and the attenuation (see Figure~3 in Buat et al. 2005), so we adopt the median attenuation of our NUV-selected sample, which is $A_{NUV} = 0.78$~mag. After correcting for this median attenuation we obtain $\rho_{NUV,corr} = 0.018^{+0.013}_{-0.008}$~M$_{\odot}$~yr$^{-1}$~Mpc$^{-3}$. Takeuchi et al. (2005) report a value of the local cosmic star formation density derived from $L_{IR}$ of $\rho_{IR} = 0.013^{+0.008}_{-0.005}$~M$_{\odot}$~yr$^{-1}$~Mpc$^{-3}$. However, we recall that this quantity does not account for the fraction of UV escaping photons or for the cirrus IR heating. For our FIR-selected sample, the median contribution of the NUV escaping luminosity to $SFR_{tot}$ is 17\% which leads to $\rho'_{IR} = 0.016^{+0.010}_{-0.006}$~yr$^{-1}$~Mpc$^{-3}$. After correcting for the cirrus contribution assuming $\eta = 0.32$ we obtain $\rho_{IR,corr} = 0.011^{+0.007}_{-0.004}$~M$_{\odot}$~yr$^{-1}$~Mpc$^{-3}$ which is well below $\rho_{NUV,corr}$, although still within the 1-$\sigma$ uncertainty. Although we have previously accepted that a cirrus correction of $\eta = 0.32$ is valid for galaxies with $L_{IR} \leq 10^{11}$~L$_{\odot}$ and $\eta = 0.09$ for brighter galaxies, we argue that adopting just $\eta = 0.32$ for all the galaxies is not a bad approximation for this calculation since it can be seen in Takeuchi et al. (2005) that the contribution of $L_{IR} \phi(L_{IR})$ to the total $\int L_{IR} \phi(L_{IR}) dL_{IR}$ of galaxies with $L_{IR} > 10^{11}$~L$_{\odot}$ is very low and can hardly affect our calculations. Proceeding in an analogous way as in Section~4.1, we can add both contributions to get the total cosmic SFR, and we get $\rho_{tot} = \rho_{NUV} + (1 - \eta) \times \rho_{IR}~(= 0.009 + 0.009) = 0.018$~M$_{\odot}$~yr$^{-1}$~Mpc$^{-3}$, which is in good agreement with $\rho_{NUV,corr}$, which means that correcting $\rho_{NUV}$ with a median attenuation is a good approximation. We also stress that the $\rho_{tot}$ is almost equally shared between the NUV and the IR contribution. The discrepancy found with $\rho_{IR}$ alone may be due to the average corrections applied. In order to obtain a better agreement between $\rho_{NUV,corr}$, $\rho_{tot}$ and $\rho_{IR,corr}$ two points should be studied in more detail: \begin{itemize} \item A detailed knowledge of the cirrus component is required since this contribution is probably multivalued for a given value of $L_{IR}$. Although the morphological type seems to drive this parameter (Sauvage \& Thuan 1992), it could also be dependent on $b$ since this parameter also measures the relative weight of the young and old stellar populations. A more detailed study of a large samples of galaxies, covering a wide range of galaxian properties, could shed light on its fractional contribution to the total cosmic $L_{IR}$ density. \item The bivariate LF $\phi(L_{NUV},L_{IR})$ appears to be the best way to estimate the fraction of UV photons escaping from the galaxy, required to correct $\rho_{IR}$. It is also required since for large values of $L_{NUV}$, the attenuation can take on multiple values, and thus an average value, as the one used in this work, might be not the most appropriate. \end{itemize} Under these conditions, the cosmic SFR densities would be expressed as: \begin{equation} \rho_{NUV,corr} = \int\int \kappa_{NUV} \times L_{NUV} \times 10^{A_{NUV}(L_{NUV},L_{IR})/2.5} \times \phi(L_{NUV},L_{IR}) dL_{NUV} dL_{IR} \end{equation} and \begin{equation} \rho_{IR,corr} = \int\int \kappa_{IR} \times \left[1 - \eta(L_{IR},L_{NUV})\right] \times L_{IR} \times (1 + (\kappa_{NUV}/\kappa_{IR}) \times (L_{NUV}/L_{IR})) \times \phi(L_{NUV},L_{IR}) dL_{NUV} dL_{IR} \end{equation} or if we use the approximation of Eq.~\ref{sfrtot_eq} \begin{equation} \rho_{tot} = \int\int \left( \kappa_{IR} \times \left[1 - \eta(L_{IR},L_{NUV})\right] \times L_{IR} + \kappa_{NUV} \times L_{NUV} \right) \times \phi(L_{NUV},L_{IR}) dL_{NUV} dL_{IR} \end{equation} \section{Conclusions} We have performed a detailed study of the star formation properties of two samples of galaxies selected on the basis of their NUV and FIR fluxes, which were found to be representative of the nearby Universe when compared to samples drawn from larger volumes. The main conclusions of this work are: \begin{enumerate} \item $L_{NUV}$ and $L_{60}$ are tightly correlated for the NUV-selected galaxies. The opposite holds for FIR-selected galaxies, which span a large range of $L_{60}$ for a given value of $L_{NUV}$ and show larger values of attenuation. Intrinsically bright galaxies are more luminous at FIR than at NUV wavelengths, including the UV luminous galaxies (UVLGs), and they show moderate to high attenuation. \item The SFR deduced from the NUV fluxes, corrected for the dust attenuation ($SFR_{NUV}$), are not found to be consistent with those calculated using the total dust emission ($SFR_{dust}$). Whereas $SFR_{NUV}$ is larger than $SFR_{dust}$ for galaxies with low attenuation ($A_{NUV} \lesssim 1$~mag) the inverse is found for bright, but highly extinguished galaxies, mostly selected in IR: $SFR_{NUV}$ is likely to underestimate the actual SFR in these galaxies by a factor $\sim 2$. A combined estimator based on UV and IR luminosities with a cirrus correction depending on the IR luminosity seems to be the best proxy over the whole range of values of SFR. As a practical recipe we found that $SFR_{tot}$ and $SFR_{NUV}$ yield similar results for $SFR_{tot} \lesssim 15$~M$_{\odot}$~yr$^{-1}$, whereas $SFR_{tot}$ and $SFR_{dust}$ are almost equivalent for $SFR_{tot} \gtrsim 15$~M$_{\odot}$~yr$^{-1}$. \item NUV-selected galaxies follow a trend whereby low-mass galaxies show lower SFRs, low attenuation and higher values of $b$, indicating the existence of a dominant young stellar population. On the contrary, about 20\% of the FIR-selected sample shows high attenuation, high SFRs and also large values of $b$, most of them being LIRGs and/or ULIRGs. In spite of their discordant properties, these galaxies are not sufficiently abundant in the local Universe to question the downsizing picture for the SFH seen at $z = 0$ from optical surveys. \item The cosmic SFR densities of the local Universe, estimated from the NUV and IR luminosities, are consistent to within 1-$\sigma$, although the difference between the two values is large, when average corrections for the attenuation, UV escaping photons and IR cirrus component are applied. The sum of the individual contributions is quite consistent with the value obtained from the NUV luminosities corrected for average attenuation. A better knowledge of the cirrus contribution to $L_{IR}$ and of the bivariate LF is required in order to better understand the large differences found between the monochromatic estimators of the local SFR density. \end{enumerate} \acknowledgments GALEX is a NASA Small Explorer, launched in 2003 April. We gratefully acknowledge NASA's support for construction, operation, and science analysis for the GALEX mission, developed in cooperation with the Centre National d'Etudes Spatiales of France and the Korean Ministry of Science and Technology. This publication makes use of data products from the Two Micron All Sky Survey, which is a joint project of the University of Massachusetts and the Infrared Processing and Analysis Center/California Institute of Technology, funded by the National Aeronautics and Space Administration and the National Science Foundation. This research has made use of the NASA/IPAC Extragalactic Database (NED) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. The Lyon Extragalactic Database (LEDA) is available at http://leda.univ-lyon1.fr/. TTT has been financially supported by the Japan Society for the Promotion of Science. \appendix \section{Relation of the Mean Redshift of Galaxies and the LF: Representativity of a Survey Depth\label{depth}} We show the strong dependence of the mean redshift on the shape of the LF. This means that the mean distance of galaxies does not represent the depth of a survey, but rather reflects an intrinsic property of the sample. Since we treat a local sample of galaxies, we first approximate their distance by the classical Hubble's law: \begin{eqnarray} cz \simeq \frac{r}{H_0} \;, \end{eqnarray} where $c$ is the velocity of light, and $r$ represents the distance. We define ${\cal N}$ to be surface density of galaxies on the sky, and denote the LF as an explicit function of the characteristic luminosity, $L_*$ as $\phi(L/L_*)$. Then, ${\cal N}$ is written as (Peebles 1993): \begin{eqnarray} {\cal N} = \int \int \phi \left[ \frac{L(r)}{L_*} \right] \frac{dL(r)}{L_*} r^2 dr \;. \end{eqnarray} Using the detected flux density $S$, this can be expressed as \begin{eqnarray} {\cal N} &=& \int \int \phi \left[ \frac{4\pi}{L_*} \left( \frac{c}{H_0}\right)^2 z^2 S \right] \frac{4\pi}{L_*} \left( \frac{c}{H_0}\right)^2 z^2 dS \left( \frac{c}{H_0}\right)^3 z^2 dz \nonumber \\ &=& \int \int \phi \left[ \frac{4\pi}{L_*} \left( \frac{c}{H_0}\right)^2 z^2 S \right] \frac{4\pi}{L_*} \left( \frac{c}{H_0}\right)^5 z^4 dS dz \;. \end{eqnarray} We observe the distribution function of the sources with a {\sl fixed} flux density $S$ as \begin{eqnarray}\label{eq:dif_nc} \frac{d^2{\cal N}}{dzdS} = \phi \left( \alpha z^2 S \right) \alpha \left( \frac{c}{H_0}\right)^3 z^4 \;, \end{eqnarray} where \begin{eqnarray} \alpha \equiv \frac{4\pi}{L_*}\left(\frac{c}{H_0}\right)^2\;, \end{eqnarray} The mean redshift of a flux-limited survey (limiting flux density $S$), $\langle z \rangle_{>S}$, is then defined as\footnote{Note that this is a different quantity defined by Equation~(5.134) of Peebles (1993).} \begin{eqnarray}\label{eq:meanz} \langle z \rangle_{>S} \equiv \frac{\displaystyle \int_S^\infty \int_0^\infty \frac{d^2{\cal N}}{dzdS'} zdzdS'}{\displaystyle \int_S^\infty \int_0^\infty \frac{d^2{\cal N}}{dzdS'} dzdS'} \;. \end{eqnarray} The numerator is obtained as \begin{eqnarray}\label{eq:numer} \int_S^\infty \int \frac{d^2{\cal N}}{dzdS'} zdzdS' &=& \frac{1}{2\alpha^2} \left(\frac{c}{H_0}\right)^3 \left[ \int_0^{\infty} \phi (x) x^2 dx\right] \int_S^\infty {S'}^{-3} dS'\nonumber \\ &=& \frac{1}{2\alpha^2} \left(\frac{c}{H_0}\right)^3 \left[ \int_0^{\infty} \phi (x) x^2 dx\right] \frac{S^{-2}}{2}\;. \end{eqnarray} Here, $x \equiv \alpha z^2 S'$, a luminosity normalized by the characteristic luminosity of the LF, $L_*$ and expressed in terms of flux density $S'$. Since the luminosity must be positive, the lower bound on the integration with respect to $x$ is 0, and upper bound is large, effectively taken to be $+\infty$. Similarly the denominator becomes \begin{eqnarray}\label{eq:denom} \int_S^\infty \int_0^\infty \frac{d^2{\cal N}}{dzdS'} dzdS' &=& \frac{1}{2\alpha^{3/2}} \left(\frac{c}{H_0}\right)^3 \left[ \int_0^\infty \phi (x) x^{3/2} dx\right] \int_S^\infty {S'}^{-5/2} dS' \nonumber \\ &=& \frac{1}{2\alpha^2} \left(\frac{c}{H_0}\right)^3 \left[ \int_0^\infty \phi (x) x^2 dx\right] \frac{2S^{-3/2}}{3}\;. \end{eqnarray} Both the numerator and the denominator of this part is the moment of the LF. This means that this is a function of its shape. Combining Equations~(\ref{eq:meanz}), (\ref{eq:numer}), and (\ref{eq:denom}), we have \begin{eqnarray}\label{eq:meanz2} \langle z \rangle_{>S} &=& 3\alpha^{-1/2} \frac{\displaystyle \int_0^\infty \phi (x) x^2 dx}{ \displaystyle \int_0^\infty \phi (x) x^{3/2} dx} S^{-1/2} \nonumber \\ &=& 3 \left(\frac{L_*}{4\pi}\right)^{1/2}\left(\frac{H_0}{c}\right) \frac{\displaystyle \int_0^\infty \phi (x) x^2 dx}{ \displaystyle \int_0^\infty \phi (x) x^{3/2} dx} S^{-1/2} \;. \end{eqnarray} Let us carefully examine Equation~(\ref{eq:meanz2}). First, the dependence of the mean redshift on the limiting flux density $S$ is a power of $-1/2$. Also, it has the same order of dependence on $L_*$. In contrast, the integral part of Equation~(\ref{eq:meanz2}) has an important meaning. Since this part depends on the second-order moment, the tail of the LF affects the value very strongly. As we mentioned, we integrate over the (normalized) luminosity up to a certain very large value, and this part is a ratio between the moments of order $3/2$ and 2. Hence the contribution from a large value of $x$ controls the value. Consequently, the mean redshift $\langle z \rangle_{>S}$ is very sensitive to the LF shape. This aspect is clearly seen in the comparison of the expected redshift distributions in NUV and $60\mu$m calculated from the LFs and the limiting flux density or magnitude, because the shapes of the LFs at these wavelengths are very different (Buat \& Burgarella 1998; Takeuchi et al. 2005). We show the comparison in Figures~\ref{histo_velo_all}b and ~\ref{histo_velo_all}c. In these figures we fix the limiting flux density at $60\mu$m to be 0.6~Jy, while we change the limiting magnitude from $AB_{NUV}=16.0$~mag (the actual value we adopt in this work) to 18.0~mag. The medians for the two wavelengths are very different when we adopt a limiting value of $AB_{NUV}=16.0$~mag. Only when we adopt a limiting value of $AB_{NUV}=18.0$~mag, which corresponds to a very sensitive survey, the median values approach for the NUV and FIR-selected samples. \section{Estimating the birthrate parameter \label{apendice_b}} Boselli et al. (2001) give a detailed recipe to estimate $\left< SFR \right>$, based only on observable quantities and/or adopted values for the parameters (see Gavazzi et al. 1996). Here we follow their prescriptions, using the $H$-band luminosity to estimate the past averaged SFR and the same parameters as Boselli et al. (2001): \begin{equation} b = \frac{SFR \times t_{0} \times (1 - R)}{L_{H} \times (M_{tot}/L_{H}) \times DM_{cont}} \end{equation} where $SFR$ in this work is averaged over $10^{8}$~yr, $t_{0}$ is the age of the disk (assumed to be equal to 13~Gyr), $R$ is the fraction of gas re-injected by stars through stellar winds into the interstellar medium during their lifetime (taken to be equal to 0.3 for a Salpeter IMF), $L_{H}$ is the $H$-band luminosity estimated as $\log L_{H} = 11.36 - 0.5 \times H + 2 \times \log D$ (in solar units) where $D$ is the distance to the source (in Mpc), $M_{tot}$ is the dynamical mass at the B-band 25~mag~arcsec$^{-2}$ isophotal radius, $M_{tot}/L_{H}$ is taken to be constant and equal to 4.6 (in solar units) and $DM_{cont}$ is the dark matter contribution to the $M_{tot}/L_{H}$ ratio at the optical radius, assumed to be equal to 0.5. \clearpage \newpage \begin{table} \footnotesize \begin{center} \caption{Typical uncertainties of the NUV and FUV magnitudes as a function of the magnitude.\label{error}} \begin{tabular}{ccc} \tableline\tableline AB Magnitude & $\sigma$(NUV) & $\sigma$(FUV) \\ interval & (mag) & (mag) \\ \tableline $\leq 16$ & 0.01 & 0.02 \\ $16 ~ ... ~ 18$ & 0.03 & 0.05 \\ $18 ~ ... ~ 20$ & 0.09 & 0.15 \\ $20 ~ ... ~ 22$ & 0.26 & 0.40 \\ \tableline \end{tabular} \end{center} \end{table} \begin{table} \footnotesize \begin{center} \caption{Basic properties of the sample galaxies: (1) Name; (2) Flag indicating membership to the NUV-selected subsample; (3) Flag indicating membership to the FIR-selected subsample; (4) R.A. (J2000); (5) Dec. (J2000); (6) Heliocentric velocity; (7) Distance derived from the velocity corrected for the Local Group infall onto Virgo and $H_{0} = 70~km~s^{-1}~Mpc^{-1}$; (8) Morphological type from NED; (9) IRAS identificator: F for FSC origin; Q,O,R for PSCz origin; SCANPI for absence in both catalogs.\label{pro}} \begin{tabular}{lcccccrcl} \tableline\tableline Name & UVsel & FIRsel & R.A. & Dec. & vel. & Dist & Type & IRAS Id.\\ & & & \multicolumn{2}{c}{(J2000.0)} & (km~sec$^{-1}$) & (Mpc) & & \\ \tableline MRK~544 & Y & N & 0 4 48.70 & $-$ 1 29 54.60 & 7110 & 101.39 & S? & F00022-0146 \\ NGC~10 & Y & Y & 0 8 34.56 & $-$33 51 27.25 & 6811 & 94.62 & Sbc & Q00060-3408 \\ NGC~35 & Y & Y & 0 11 10.46 & $-$12 1 14.74 & 5964 & 83.95 & Sb & Q00086-1217 \\ NGC~47 & Y & Y & 0 14 30.42 & $-$ 7 10 6.28 & 5700 & 80.54 & Sbc & Q00119-0726 \\ NGC~101 & Y & N & 0 23 54.72 & $-$32 32 9.06 & 3383 & 45.92 & Sc & F00214-3248 \\ \tableline \end{tabular} \end{center} Note: The distances to UGCA~438 and UGC~12613 were taken from Karachentsev et al. (2002) and Hoessel et al. (1990) respectively. \end{table} \begin{table} \footnotesize \begin{center} \caption{Photometric properties of the sample galaxies: (1) Optical identifier; (2) NUV magnitude corrected for Galactic extinction; (3) FUV magnitude corrected for Galactic extinction; (4) Flux density at 60$\mu$m in Jy; (5) Flux density at 100$\mu$m in Jy; (6) $H$ magnitude from 2MASS Extended Source Caztalog. For galaxies with no detection at 2MASS we adopt the limiting value of $H = 13.9$~mag (3~mJy) as given by Jarrett et al. (2000); (7) NUV attenuation in mag, derived as indicated in Buat et al. (2005); (8) FUV attenuation derived as indicated in Buat et al. (2005). For some galaxies Eqs.~\ref{attenuv_eq} and \ref{attefuv_eq} gave negative values of the NUV and FUV attenuations which is senseless. In fact, this is an artifact of the polynomial fitting used to derive a functional form for the attenuation in Buat et al. (2005). Throughout this paper they will be considered as zero.\label{photo}} \begin{tabular}{lccccccc} \tableline\tableline Name & $AB_{NUV}$ & $AB_{FUV}$ & $f_{60}$ & $f_{100}$ & $H$ & $A_{NUV}$ & $A_{FUV}$ \\ & \multicolumn{2}{c}{(mag)} & \multicolumn{2}{c}{(Jy)} & (mag) & \multicolumn{2}{c}{(mag)} \\ \tableline MRK~544 & 15.75 & 15.97 & 0.49 & 1.11 & 12.50 & 0.71 & 0.96 \\ NGC~10 & 15.48 & 15.98 & 0.66 & 2.97 & 9.50 & 1.59 & 2.10 \\ NGC~35 & 15.55 & 15.89 & 1.31 & 2.34 & 11.91 & 1.29 & 1.67 \\ NGC~47 & 15.46 & 15.97 & 0.85 & 2.57 & 10.25 & 1.02 & 1.49 \\ NGC~101 & 14.78 & 14.98 & 0.55 & 1.75 & 10.74 & 0.50 & 0.72 \\ \tableline \end{tabular} \end{center} \end{table} \begin{table} \footnotesize \begin{center} \caption{Star formation related properties of the sample galaxies: (1) Identifier of the galaxy; (2) $SFR_{NUV}$ from Eq.~\ref{sfrnuv_eq}; (3) $SFR_{FUV}$ from Eq.~\ref{sfrfuv_eq}; (4) $SFR_{dust}$ from Eq.~\ref{sfrfir_eq}; (5) $SFR_{tot}(NUV)$ from Eq.~\ref{sfrtot_eq}; (6) $SFR_{tot}(FUV)$ from Eq.~\ref{sfrtot_eq} but using $SFR^{0}_{FUV}$ instead of $SFR^{0}_{NUV}$; (7) $\left< SFR \right>$ averaged over the galaxy's lifetime, estimated as indicated in Appendix~\ref{apendice_b}.\label{sfr_tab}} \begin{tabular}{lcccccc} \tableline\tableline Name & $\log SFR_{NUV}$ & $\log SFR_{FUV}$ & $\log SFR_{dust}$ & $\log SFR_{tot}(NUV)$ & $\log SFR_{tot}(FUV)$ & $\log \left< SFR \right>$ \\ & \multicolumn{6}{c}{($M_{\odot}$~yr$^{-1}$)} \\ \tableline MRK~544 & 0.83 & 0.85 & 0.60 & 0.80 & 0.75 & 0.81 \\ NGC~10 & 1.23 & 1.24 & 1.07 & 1.09 & 1.02 & 1.95 \\ NGC~35 & 0.98 & 1.00 & 0.78 & 0.84 & 0.79 & 0.88 \\ NGC~47 & 0.87 & 0.86 & 0.78 & 0.85 & 0.77 & 1.51 \\ NGC~101 & 0.45 & 0.46 & 0.13 & 0.43 & 0.38 & 0.83 \\ \tableline \end{tabular} \end{center} \end{table} \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage
Title: Bubbles in Planetary Nebulae and Clusters of Galaxies: Jet Bending
Abstract: We study the bending of jets in binary stellar systems. A compact companion accretes mass from the slow wind of the mass-losing primary star, forms an accretion disk, and blows two opposite jets. These fast jets are bent by the slow wind. Disregarding the orbital motion, we find the dependence of the bending angle on the properties of the slow wind and the jets. Bending of jets is observed in planetary nebulae which are thought to be the descendants of interacting binary stars. For example, in some of these planetary nebulae the two bubbles (lobes) which are inflated by the two opposite jets, are displaced to the same side of the symmetry axis of the nebula. Similar displacements are observed in bubble pairs in the center of some clusters and groups of galaxies. We compare the bending of jets in binary stellar systems with that in clusters of galaxies.
https://export.arxiv.org/pdf/astro-ph/0601032
\title{BUBBLES IN PLANETARY NEBULAE AND CLUSTERS OF GALAXIES: JET BENDING} \author{Noam Soker and Gili Bisker} \affil{Department of Physics, Technion$-$Israel Institute of Technology, Haifa 32000 Israel; soker@physics.technion.ac.il.} \keywords{stars: mass loss --- binaries: close --- planetary nebulae: general --— intergalactic medium --— ISM: jets and outflows --- galaxies: clusters: general} \section{INTRODUCTION} The nebular gas in planetary nebulae (PNs) originates in the envelope of asymptotic giant branch (AGB) stars that are the descendants of intermediate mass stars (initial masses $\sim 1-8 M_\odot$). Such stars rotate very slowly, and their mass loss is expected to be spherical. Indeed, AGB stellar winds usually consist of a more or less spherically symmetric outflow at rates of $\sim 10^{-7}-10^{-5} M_\odot \yr ^{-1}$. Most PNs, though, possess a global axisymmetrical structure rather than a spherical structure in their inner region, indicating a non-spherical shaping process. Among the several PN shaping models (Balick \& Frank 2002), one of the most successful is the jet-shaping model. If the jets are not well collimated they are termed collimated fast wind (CFW). The presence of jets in PNs was deduced from observations more than 20 years ago (e.g., Feibelman 1985). Gieseking et al. (1985) found collimated outflow in the PN NGC 2392, and noted the similarity of these jets with that of young stellar objects, and speculated that such outflows exist in many similar PNs. On the theoretical side, Morris (1987) suggested that two jets blown by an accreting companion (the secondary star) can form bipolar nebulae. This model is strongly supported by the similarity of bipolar PNs to many bipolar symbiotic nebulae which are known to be shaped by jets (e.g., Schwarz et al. 1989; Corradi \& Schwarz 1995). Soker (1990) proposed that the two fast low-ionization emission blobs (FLIERs or {\it ansae}) along the symmetry axis of many elliptical PNs are formed by jets blown during the last phase of the AGB or the post-AGB phase of the PN progenitor. The high quality HST images led Sahai \& Trauger (1998) to suggest that in many PNs the non-spherical structures are formed solely by jets. Projecting from similar astronomical objects, the formation of massive jets, to distinguish from magnetized low density pulsar jets, require the presence of accretion disks. The only source of angular momentum sufficient to form accretion disks in evolved stars is the orbital angular momentum of a stellar (or in some cases substellar) companion. The disk can be formed around the progenitor during the late post-AGB phase, when it is already small (Soker \& Livio 1994), or, more likely, around a stellar companion accreting mass, forming an accretion disk, and blowing two jets. The past seven years have seen further consolidation of the bipolar jet-shaping model in binary systems, addressed both in observations (e.g., Parthasarathy et al. 2000; Sahai \& Nyman 2000; Miranda et al. 2001a,b; Corradi et al. 2001; Guerrero et al. 2001; Vinkovic et al. 2004; Huggins et al. 2004; Pena et al.\ 2004; Balick \& Hajian 2004; Arrieta et al. 2005; Oppenheimer et al. 2005; Sahai et al. 2005) and in theory (e.g., Soker 2002, 2005; Lee \& Sahai 2003, 2004; Livio \& Soker 2001; Garcia-Arredondo\& Frank 2004; Velazquez et al. 2004; Riera et al. 2005). Many of the PNs in the observations listed above posses point symmetric morphology, i.e., several symmetry axes rotate with respect to each other through a common origin, indicating precessing jets. The most likely explanation for precession is an accretion disk in the presence of a companion. Soker \& Rappaport (2000) further discussed the jet shaping process and have shown that the statistical distribution of bipolar PNs can be accounted for in the binary model. Further support for the formation of jets in binary systems comes from X-ray observations hinting at jets in a PN (Kastner et al. 2003) similar to X-ray jets in symbiotic systems (Kellogg et al. 2001; Galloway \& Sokoloski 2004). Garcia-Arredondo \& Frank (2004) were the first to conduct 3D numerical simulations of the interaction of jets launched by a secondary star with the slow primary wind. Their high quality results strengthen the general stellar-binary jets model, and in particular the conjecture (Soker \& Rappaport 2000) that a narrow waist can be formed by jets. It should be stressed that not all PNs are shaped by jets, but bubble pairs are formed by jets. X-ray images of active galactic nuclei in clusters of galaxies indeed show that double-jets, observed in the radio band, can form a bubble pair with a narrow waist between them, similar to narrow waists in PNs with no need for enhanced equatorial mass loss rate, although enhanced equatorial mass loss rate might occur in many PNs. The subject of the similarity between some morphological structures in clusters of galaxies, as revealed via X-ray observations, and in PNs, as revealed in the visible band, was studied in a series of four papers. \newline {\it Paper 1} ( Soker 2003b). In that paper ( see also Soker 2003a, and section 5 in Soker 2004c) the similarity in morphological structures was discussed \footnote{The similar morphologies are compared in the appendix of the astro-ph version of the present paper.}. This similarity is not trivial. Two opposite jets are observed in many young stellar objects (YSOs), however, bubbles pairs similar to those in PNs and in clusters of galaxies are not usually observed around YSOs. \newline {\it Paper-2 } (Soker 2004a). It was found that to inflate fat, more or less spherical, bubbles the opening angle of the jets should be large; the half opening angle measured from the symmetry axis of each jet should typically be $\alpha \gtrsim 40 ^\circ$, or the jets might precess. \newline {\it Paper-3} (Soker 2004b). Paper 3 studies the stability of off-center low-density fat bubbles in clusters of galaxies and in PNs to the Rayleigh-Taylor instability. \newline {\it Paper 4} (Pizzolato \& Soker 2005). Pizzolato \& Soker examined the point symmetric structure of the bubble pair in the cluster MS 0735.6+7421 (McNamara et al. 2005) and compared it to the point symmetric structure of PNs. Point symmetric PNe are thought to be shaped by stellar binary interactions; namely, the presence of a companion to the PN’s progenitor star is required. Pizzolato \& Soker (2005) suggested that similar point-symmetric structures in the X-ray deficient cavities of galaxy clusters might be associated with the presence of massive binary black holes. In this paper, the fifth in the series, we examine the bending of the two jets, and the subsequent bending of the two bubbles inflated by the jets, to the same side of their original symmetry axis (jets' axis). Such displacement relative to the symmetry axis of bubbles touching the center is seen, for example, in the Perseus cluster of galaxies (Fabian et al. 2000; Dunn et al. 2006), and in the PN NGC 3587 (Guerrero et al. 2003). Dunn et al. (2006) discuss the departure of the two bubbles from their alignment along a cluster center and explain this departure by the two opposite bubbles detaching from the precessing jets at different times. We consider this displacement to result from the ram pressure of the intra cluster medium (ICM). Displacement of bubbles at a distance from the center are seen in the PN NGC 6886 (Terzian \& Hajian 2000), and the group of galaxies HCG 62 (Vrtilek et al. 2002). We focus on PNs and related binary stellar objects, e.g., the massive binary stellar system $\eta$ Carinae ($\S 2.1$). The departure of PNs and related binary systems from axisymmetry has been previously studied (Soker \& Hadar 2002 and references therein). Our goal here is to derive a simple expression for the bending of jets in binary stellar systems ($\S 2.2$). This expression is not a substitute for future numerical simulations. The results for typical binary systems ($\S 2.2$) can account for some morphological structures in PNs and related systems. Readers interested in only using the relations and the results, can skip $\S 2.1$ and go directly to $\S 2.2$. In $\S 3$ we compare the situation with jet bending in cooling flow clusters, and $\S 4$ is a summary. \section{BENDING IN A BINARY STELLAR SYSTEM} \subsection{Assumptions and Equations} When a compact secondary star accretes from the AGB (or post-AGB) stellar wind only part of the AGB wind is accreted, and the rest expands outward and forms the medium that the jets expand into. In addition, when the jet is still close to the binary system, the AGB wind hits the jet on its side, causing the jet to deflect (Soker \& Rappaport 2000). Like precession, this can have large effects on the descendant PN morphology. However, while precession leads to point-symmetric nebula, the deflection of the two oppositely ejected jets is to the same side, leading the two opposite lobes to be bent to the same side; this is the {\it bent} departure from axisymmetry according to the classification of Soker \& Hadar (2002). The bending interaction can clear the way to radiation, possibly ionizing radiation, from the central binary system to more strongly illuminate the same side in both lobes (bubbles). Due to the orbital motion, this structure forms a revolving light source. Livio \& Soker (2001) suggested such a revolving ionizing source model to explain the positional shift of the bright knots in the inner nebular lobes of the M2-9 nebula (Doyle et al. 2000). Soker \& Rappaport (2000) derived a simple expression for the bending angle of a narrow jet. In this section we relax some of the assumptions made by Soker \& Rappaport and derive a more accurate expression for the bending angle, while still keeping the expression simple. The goal is to derive a simple approximate relation that will give the jet's bending angle upon specifying the jet's parameters and slow wind parameters. The bending interaction is drawn schematically by Soker \& Rappaport (2000) and Livio \& Soker (2001), and it is shown in Figure \ref{draw1}; 3D images of numerical simulations are presented by Garcia-Arredondo \& Frank (2004). The slow wind has a spherical mass loss rate of $\dot M_s$ and a relative speed to the primary star of $v_s$. A small fraction of this wind is accreted by the secondary star, forms an accretion disk that blows two jets, with a mass loss rate of $\dot M_j$ into the two jets together, and with a speed of $v_j$ perpendicular to the equatorial plane relative to the secondary star. Although the jets can have a large opening angle and, in many cases, are likely to have a large opening angle, in the present study we assume a narrow jet with a half opening angle $\alpha \ll 1$, and also assume that the jet is bent as one entity (sound crossing time across the jet is very short). The density per unit length along the jet axis is \begin{equation} m_j = \frac {\dot M_j}{2 v_j} \label{mj} \end{equation} (recall that $\dot M_j$ is the mass loss rate into the two jets together). We move to a frame of reference attached to the secondary star in its orbital motion, with a velocity relative to the primary of ${\bf {v}}_{\rm orb}=v_r \hat r + v_\theta \hat \theta$, where $v_\theta \simeq r \dot \theta$, $\theta$ is the relative angle of the two stars in the equatorial plane, and $r$ is the projected distance from the primary to the jet on the equatorial plane; $v_r<0$ when the two stars approach each other. We consider a narrow jet's segment at a height $z$ above (or below) the equatorial plane. The slow wind segment that hit this segment left the primary at an angle $\beta$ to the equatorial plane (see fig. \ref{draw1}) \begin{equation} \sin \beta = \frac{z} {(z^2+r^2)^{1/2}}. \label{stheta} \end{equation} The slow wind that hits the jet at a high $z$ above the equatorial plane has a relative velocity to the jet of \begin{equation} v_{rel}= [v_\theta^2+(v_s \cos \beta-v_r)^2+(v_s \sin \beta)^2]^{1/2}. \label{vrel} \end{equation} We consider a fast jet $v_j \gg v_s$ that initially expands perpendicularly to the orbital plane, but is then bent by the ram pressure of the slow wind and acquires a velocity parallel to the equatorial plane $v_p$. The ram pressure exerted by the slow wind on the jet in a direction parallel to the equatorial plane is \begin{equation} P_{ram}= \rho \left\{ [v_\theta^2+(v_s \cos \beta-v_r)^2]^{1/2}-v_p \right\}^2, \label{pram1} \end{equation} where the density of the slow wind \begin{equation} \rho = \frac {\dot M_s}{4 \pi v_s (r^2+z^2)}. \label{rho} \end{equation} The equation for accelerating the jet in a direction parallel to the equatorial plane (perpendicular to the initial direction of the jet), under the assumption of a fast jet, $v_p \ll v_j$, reads \begin{equation} \frac{dv_p}{dt}= \frac{P_{ram} 2 z \tan \alpha}{m_j} \label{dvpdt} \end{equation} Under the assumption of a fast jet, $z=v_j t$ and $dt=dz/v_j$. We also scale velocities by the slow wind speed \begin{equation} u_r \equiv \frac{v_r}{v_s}; \quad u_\theta \equiv \frac{v_\theta}{v_s}; \quad u_p \equiv \frac{v_p}{v_s}; \quad u_j \equiv \frac{v_j}{v_s}. \label{defvs} \end{equation} The equation of motion reads \begin{equation} \frac{du_p}{dz}= A \left\{ \left[ u_\theta^2+\left( \frac{r}{\sqrt{r^2+z^2}}-u_r \right)^2 \right]^{1/2} -u_p \right\}^2 \frac{z}{r^2+z^2}, \label{dvpdz} \end{equation} where \begin{equation} A= \frac{\tan \alpha}{\pi} \frac{\dot M_s}{\dot M_j} \label{adef} \end{equation} The meaning of the different terms in equation (\ref{dvpdz}) are as follows. (1) The factor $A$ is proportional to the ratio of colliding masses. Bending efficiency increases with $A$. (2) The terms $u_\theta$ and $u_r$ result from the orbital motion of the secondary star, which blows the jets, relative to the slow wind. (3) The term $r/(r^2+z^2)^{1/2}$ results from the ram pressure of the slow wind on the jet. The slow wind moves at a velocity $v_s$; but since velocity was scaled by $v_s$, a factor of unity multiplies this term. (4) The numerator in the last term is due to the increase in the jet cross section, and it increases the bending efficiency as the jet expands. (5) The denominator in the last term is the decrease in the slow wind density, and it makes bending less efficient as distance from the primary star grows. \subsection{Results for Impulsive Jets} We consider a case in which the jets are blown by a secondary star that is less massive than the primary star. companion. The slow wind is blown by the primary star, residing close to the center of mass of the binary system. The formulation derived above is applicable to continuously blown jets, or jets blown impulsively. However, for the bubbles in PNs, or similar object, to be significantly displaced by the mechanism discussed in $\S 2.1$, the jet should be blown during a short time compare to the orbital period. (Significant displacement from axisymmetry for continuously blown jets can be acquired if the binary system has a large eccentricity; see references in Soker \& Hadar 2002.). In many PNs, the jets' ejection (PN jets refer to the jets blown by the PN progenitor) can take place over a short time period (e.g., Meaburn 2006), which we take to be shorter than the orbital period. For example, the orbital period can be 5-50 years (orbital separation of $\sim 3-20 \AU$), and the ejection event a few years, as in symbiotic-nova outbursts on an accreting WD companion. The mass accretion rate from the primary stellar wind, $\dot M_2$, by a companion of mass $M_2$ at an orbital separation $r_0$ is \begin{equation} \frac{\dot M_2}{\dot M_s} \simeq 0.05 \left( \frac {M_2}{0.6 M_\odot} \right)^{2} \left( \frac {v_{rel}}{15 \km \s^{-1}} \right)^{-4} \left( \frac {r_0}{10 \AU} \right)^{-2}. \end{equation} If in impulsive jets' ejection $\dot M_j \sim 0.2 \dot M_2$, then for the above mass accretion rate $A \simeq 5.6 \alpha/10^\circ$. In short eruption events, like disk instability or nova-like outbursts on an accreting WD, it might be that $\dot M_j > 0.2 \dot M_2$, and $r_0$ span a range of $\sim 1-30 \AU$. Therefore, we consider $A$ to be be in the range $A \sim 0.1-100$. The jet is bent, according to equation (\ref{dvpdz}), and $u_p$, the velocity component parallel to the equatorial plane and perpendicular to initial velocity of the jet reaches an asymptotic velocity of \begin{equation} u_{pa} = \left[ u_\theta^2+\left( \frac{r}{\sqrt{r^2+z^2}}-u_r \right)^2 \right]^{1/2}. \label{upa} \end{equation} where $u_{pa}$ is in unit of the slow wind speed $v_s$. The asymptotic (final) velocity $u_p$ due to the orbital tangential velocity $u_\theta$ does not depend on the factor $A$ or the jet speed $v_j$ (or $u_j=v_j/v_s$). This is approximately true for the radial orbital component $u_r$ as well, meaning that the initial jet velocity component along the secondary stellar orbital motion is quite efficiently reduced to zero. The departure from axisymmetry due to the orbital motion of the star blowing the jet will be small. Therefore, in imposing a noticeable large-scale departure from axisymmetry, where the two jets are bent to the same side, the bending due to the slow wind outflow from the primary star must be considered. This bending is less efficient because the slow wind velocity is not perpendicular to the jet velocity after the jet leaves the equatorial plane, as seen by the decreasing of the term $r/(r^2+z^2)^{1/2}$. Ignoring the orbital motion, equation (\ref{dvpdz}) reads \begin{equation} \frac{du_p}{dz}= A \left[ \frac{r}{(r^2+z^2)^{1/2}} -u_p \right]^2 \frac{z}{r^2+z^2}, \label{dup} \end{equation} This equation is supplemented by another equation for the jet propagation along the direction perpendicular to the equatorial plane. For a fast jet, $v_j \gg v_s$, this reads, \begin{equation} \frac{dr}{dz}= \frac{u_p}{u_j}. \label{dr} \end{equation} Figure \ref{upf1} presents the numerical solutions of the last two coupled equations for initial jet's speed $u_j\equiv v_j/v_s=6$ and for three values of $A$ as function of the distance from the equatorial plane $z$ in units of the orbital separation $r_0$. The velocity $u_p$ is plotted in the upper panel, in the middle panel the projection of the jet distance on the equatorial plane $r$ is drawn (in units of $r_0$), while the lower panel presents the acceleration $du_p/dz$. In Figure \ref{upaf}, we show the asymptotic velocity $u_{pa}$ as a function of $A$ for $u_j=6$ (the thick line). Changing the initial jet's speed $u_j$ does not change the solution for $u_p$, while the quantity $r-r_0$ is proportional to $u_j^{-1}$, because the bending angle is given by $\tan \phi =u_p/u_j$, so that for faster jets the bending angle decreases; the dependence is $\phi \propto v_j^{-1}$. This can be understood as follows. As the jet speed $v_j$ increases, the time of accelerating the jet by the slow wind's ram pressure along a distance $dz$ decreases as $v_j^{-1}$; however, the density in the jet decreases as $v_j^{-1}$ as well. Hence the total change in $v_p$ (or $u_p$) along a distance $dz$ does not depend on $v_j$ under our assumptions, in particular the assumption $v_j$ (or $u_j$) is constant and is not influenced by the interaction with the slow wind. As seen from Figure \ref{upf1}, for a fast jet (and for typical values used here) $u_p$ almost reaches its terminal speed while $r \simeq r_0$, where $r_0$ is the initial orbital separation. Note that most of the bending occurs for $r$ not much larger than $r_0$. For $r=r_0$ equation (\ref{dup}) reads then \begin{equation} \frac{du_p}{dz}= A \left[ \frac{r_0}{(r_0^2+z^2)^{1/2}} -u_p \right]^2 \frac{z}{r_0^2+z^2}, \label{dup0} \end{equation} The solution near the origin, when $u_p \ll 1$ is \begin{equation} u_{p0} (r \sim r_0) \simeq \frac{A}{2} \frac{z^2}{r_0^2+z^2}. \label{up0} \end{equation} The asymptotic velocity is reached when the numerical value inside the square brackets in equation (\ref{dup0}) is very small, or \begin{equation} u_{pac} (r \sim r_0) \simeq \frac{r_0}{(r_0^2+z^2)^{1/2}}. \label{upaa} \end{equation} The change of behavior between the solution near the jets' origin and the asymptotic solution takes place when $u_{p0} \sim u_{pa}$, which by equations (\ref{up0}) and (\ref{upaa}) is \begin{equation} u_{pac} \sim \left( 1+ \frac{1}{A^2} \right)^{1/2} -\frac{1}{A}. \label{upas} \end{equation} This very crude expression for the asymptotic transverse velocity is drawn by a thin line on Figure \ref{upaf}. As the jet leaves the launching accretion disk, it is very dense and no bending occurs, namely, $du_p/dz$ is very small. At large distances from the jet's origin, the angle $\delta$ is small and bending is no longer efficient. The bending is most efficient at some intermediate value of $z$, after the density of the jets decreases as they expand, but before the angle $\delta$ decreases much. Practically, this intermediate value of $z$ is quite close to the jet's origin, $z \la r_0$, as is seen in the lower panel of Figure \ref{upf1}. If the jet pair in a binary system is known to be blown in a time period much shorter than the orbital period (an impulsive jet pair), and a bending is observed, the bending angle can be used with the thick line in Figure \ref{upaf} to find the constant $A$ given in equation (\ref{adef}). Thus the relation between the three quantities: the primary stellar mass loss rate $\dot M_s$, the secondary stellar mass loss rate to the two jet $\dot M_j$, and the half opening angle of the jets $\alpha$, can be found. \section{BENDING IN CLUSTERS} \label{clben} Many jet pairs blown by radio galaxies are observed to be bent as a result of the relative motion of the galaxy and the ICM (e.g., Bliton et al. 1998). Radio jets which are strongly bent are called narrow-angle tailed (NAT) radio galaxies, while those with slightly bent jets are called wide-angle tailed (WAT) radio galaxies. Many of the WAT radio galaxies are dominant galaxies in clusters, like cD galaxies (Owen \& Rudnick 1976; Burns et al. 1979). A bulk motion of the IC, e.g., as a result of cluster merger, can efficiently bend radio jets (Bliton et al. 1998). A bulk ICM motion relative to the central cD galaxy can exist as a result of merging with a sub-cluster (group of galaxies), as found in several cases (Dupke \& Bregman 2005, 2006; Fujita et al. 2006). The bending process of jets by the ICM was extensively studied (e.g. Balsara \& Norman 1992); the calculations are not repeated here. Basically, the bending of jets in clusters is characterized by the curvature radius $R_c$ of the bent jet. Because the ambient density changes slowly with distance from the cluster center, unlike the case in PNs, a constant ambient density is assumed in the region where most of the jet's bending occurs. An approximate expression for the radius of curvature is (Sarazin et al. 1995) \begin{equation} R_{\rm {curv}} \sim 2 \left( \frac {L_{2j}}{3 \times 10^{43} \erg \s^{-1}} \right) \left( \frac {R_j}{1 \kpc} \right)^{-1} \left( \frac {v_j}{0.1 c} \right)^{-1} \left( \frac {v_a}{300 \km \s^{-1}} \right)^{-2} \left( \frac {n_e}{0.1 \cm^{-3}} \right)^{-1} \kpc \label{rcur} \end{equation} where $n_e$ is the ambient electron density, $R_j$ is the jet's radius where most bending takes place, the total mechanical (kinetic) power of the two jets, $L_{2j}$, was scaled according to Birzan et al. (2004), and the relative ambient to jet speed $v_a$ according to Malumuth (1992), with a $90^\circ$ angle between the initial jet velocity and the ambient flow. To achieve a noticeable bending in cD clusters we require $R_{\rm curv} \la 10 \kpc$. This implies that even if the relative velocity of the cD galaxy to the ambient medium component perpendicular to the jet axis is $\sim 100 \km \s^{-1}$ we get the required bending, as observed in some jets, or bubbles, blown by cD galaxies. This required bending will occur also for a narrower jet or a faster jet with $v_j$ close to the speed of light $c$, Our primary interest is to compare the bending process in binary stars to that of jets blown by the dominant cD galaxies at the centers of cooling flow clusters (or galaxies). Common to bending of jets in binary progenitors of PNs (referred to as bending in PNs) and bending of jets at the centers of clusters of galaxies is that the jets are bent by the ram pressure due to the relative motion of the medium the jets expand to and interact with, both in PNs and clusters. This is unlike point-symmetric structures which result from precession; in both PNs and clusters precession is due to the accretion disk that launches the jets and not due to the ambient medium. There are some differences between bending in clusters and bending in binary systems, such as PNs' progenitors. \begin{enumerate} \item {\it Relative velocity.} The bending considered in PNs is due to the outflow velocity of the slow wind blown by the primary star. This implies that the angle between the jets velocity and the ambient medium velocity, $\delta$ in Figure \ref{draw1}, decreases very fast. It decreases even if bending does not occur. In clusters the velocity is due to the motion of the cD galaxy relative to the ICM. The angle decreases only because of the bending, and it decreases slowly. \item {\it Densities.} The jet's density decreases as the jet moves outward, both in clusters and in stellar binary systems. However, in PNs the density of the bending slow wind (see Fig. \ref{draw1}) decreases as well, $\rho_s \propto (r^2+z^2)$ (denominator of last term in eq. \ref{dvpdz}), while near the center of clusters the ICM density profile is much shallower and decreases slowly with increasing distance. \item {\it Asymmetry in clusters.} The bending process in PNs is the same for the two opposite jets, but this is not necessarily the case in clusters of galaxies. If the jets are not blown perpendicular to the relative velocity between the ICM and the galaxy, then the jet expanding to the same direction the galaxy moves to will feel a larger ram pressure opposing its expansion velocity, and it will be slowed down more efficiently. More important, as this jet is bent, the angle of the relative velocity between the ICM and the galaxy to the jet's axis will increase to $90^\circ$ before decreasing. In the opposite jet this angle decreases continuously. Therefore, the jet expanding against the ICM motion will be bent more than the other jet. Asymmetry between the two jets in clusters can be also caused by the presence of asymmetric strong magnetic fields in the ICM (Soker 1997), and/or density inhomogeneities such as clouds (Sarazin et al. 1995). \item {\it Late stages of the bending process. } After reaching their asymptotic bending angle $\phi$, jets in binary stellar systems will not bend any more. If a jet inflates a bubble, it will move outward radially along the streaming slow wind material. In clusters the situation is different because of the flow structure mentioned in points (1) and (2) above and because the low density bubbles buoyant outward. The result is that although the radio jets of cD galaxies are not bent much, after they become subsonic the bending is very efficient (Eilek et al. 1984; Odea \& Owen 1986), and the asymmetry between the two sides can substantially increase (Burns et al. 1986). \item{\it The effective bending location.} From the differences in points 1,2, and 4 it turns out that in stellar binary systems most of the bending occurs close to the jets' origin (lower panel of fig. \ref{upf1}). In clusters the bending becomes more efficient as the jet expands and slows down. In particular, if the jet inflate bubbles, they move slowly, have very low density, and large cross section. Thus, in clusters the departure from axisymmetry will be most noticeable in bubbles. \end{enumerate} Despite the differences listed above, there are some striking morphological similarities of bubbles displaced from the symmetry axis in PNs and clusters; two cases are mentioned in $\S 1$ (see appendix of the astro-ph version of the paper). \section{SUMMARY} In recent years the jet shaping model for many PNs and similar objects, like the massive binary star $\eta$ Carinae, acquired considerable acceptance. It should be stressed that not all PNs were shaped by jets, and not all morphological structure in PNs were formed by jets. Bubble pairs, though, are most likely inflated by double jets, and the jets are probably blown by a stellar secondary star. The secondary star accretes mass from the primary's slow wind, forms an accretion disk and blow two jets, either continuously or impulsively, that is, during a time shorter than the orbital period. In some PNs, the line joining the centers of the two bubbles in a pair does not pass through the center of the nebula, meaning that the bubbles are displaced such that the nebular structure departs from axisymmetry. The explanation is that the two jets that inflate the bubbles were bent to the same side by the ram pressure of the slow wind (Fig. \ref{draw1}). We therefore set the goal of deriving a simple and approximate relation between the bending angle of the jets and the properties of secondary stellar jets and the primary slow wind. For fast jets, $v_j \gg v_s$, the important factor is the quantity $A$ defined in equation (\ref{adef}). The relation between the jet's asymptotic transverse speed $v_{pa}$ (see fig. \ref{draw1}) and $A$ is presented by the thick line in Figure \ref{upaf}, and a very crude approximation is given in equation (\ref{upas}) ($v_{pa}$ is in units of the slow wind speed $v_s$). If the jets are impulsive, then the bending will be easier to observe; otherwise it is averaged over different directions as the binary system rotate. If $A$ is not too small, and the orientation of the nebula is such that the bending is not along the line of sight, then observations may reveal the two jets or the bubbles (lobes) inflated by the jets to be displaced to the same side of the symmetry axis. Examples of such PNs are listed and classified by Soker \& Hadar (2002). In some clusters, X-ray-deficient bubble (cavity) pairs that were inflated by jets blown by the central cD galaxy, show displacement from axisymmetry similar to visible-deficient bubble (lobe) pairs observed in PNs (see appendix of the astro-ph version of this paper). We therefore set a second goal of comparing the bending process of jets in these two groups of objects. Two factors of the bending process are common to these two classes of objects: 1) the bending results from the ram pressure perpendicular to the jet axis, and 2) the ram pressure is exerted by the same external medium the jets expand to and interact with. However, there are some significant differences listed in \S\ref{clben}. (1) Because the bending in binary systems results from the slow wind blown by the primary star, the ambient density decreases faster with distance than the ambient density of the ICM in the centers of clusters. (2) Also, the angle $\delta$ (see Fig. \ref{draw1}) between the jet velocity and ambient slow wind velocity in binary systems decreases with distance along the jet axis, even when bending does not occur. In clusters the relative velocity is due to the bulk ICM motion and changes only because of bending. (3) In binary systems the two opposite jets are likely to be blown perpendicular to the orbital plane, thus they will be bent in the same way. In clusters, the jets' axis need not be perpendicular relative to the bulk motion of the ICM relative to the central black hole that blows the jets, so the jet facing the ICM flow will be bent more efficiently. In binary systems such an asymmetry between the two jets can occur if the jets (more specifically the accretion disk that launches the jets) precess. (4) In clusters, after the bubbles (cavities; lobes) are inflated, they buoy outward. They are more susceptible than the jets to the ram pressure, and departure from axisymmetry may substantially increase. This process does not exist in binary systems because the circumbinary ambient matter is not in hydrostatic equilibrium, but rather the ambient matter expands at a high Mach number. (5) In binary stars most of the bending occurs when the jets are at a distance $z \la r_0$, where $r_0$ is the orbital separation. In clusters the bending becomes more efficient at larger and larger distances. We hope that the study presented in this paper will motivate researchers to pay more attention to the departure from axisymmetry of bubble (cavity; lobe) pairs in both clusters of galaxies and PNs. \acknowledgments \acknowledgments This research was supported in part by the Asher Fund for Space Research at the Technion.
Title: Thermal Structure and Radius Evolution of Irradiated Gas Giant Planets
Abstract: We consider the thermal structure and radii of strongly irradiated gas giant planets over a range in mass and irradiating flux. The cooling rate of the planet is sensitive to the surface boundary condition, which depends on the detailed manner in which starlight is absorbed and energy redistributed by fluid motion. We parametrize these effects by imposing an isothermal boundary condition $T \equiv T_{\rm deep}$ below the photosphere, and then constrain $T_{\rm deep}$ from the observed masses and radii. We compute the dependence of luminosity and core temperature on mass, $T_{\rm deep}$ and core entropy, finding that simple scalings apply over most of the relevant parameter space. These scalings yield analytic cooling models which exhibit power-law behavior in the observable age range $0.1-10 {\rm Gyr}$, and are confirmed by time-dependent cooling calculations. We compare our model to the radii of observed transiting planets, and derive constraints on $T_{\rm deep}$. Only HD 209458 has a sufficiently accurate radius measurement that $T_{\rm deep}$ is tightly constrained; the lower error bar on the radii for other planets is consistent with no irradiation. More accurate radius and age measurements will allow for a determination of the correlation of $T_{\rm deep}$ with the equilibrium temperature, informing us about both the greenhouse effect and day-night asymmetries.
https://export.arxiv.org/pdf/astro-ph/0601317
\title{Thermal Structure and Radius Evolution of Irradiated Gas Giant Planets} \author{Phil Arras and Lars Bildsten} \affil{ Kavli Institute for Theoretical Physics \\ Kohn Hall, University of California, \\ Santa Barbara, CA 93106; arras@kitp.ucsb.edu, bildsten@kitp.ucsb.edu} \keywords{planetary systems--planets and satellites:general} \section{Introduction} \begin{deluxetable*}{lccccccr} \tabletypesize{\small} \tablewidth{0pt} \tablecaption{Transiting Extrasolar Planets \label{tab1} } \tablehead{ \colhead{object} & \colhead{$a$(au)} & \colhead{$M_{\rm p}$($M_{\rm J}$)} & \colhead{$R_{\rm p}$($R_{\rm J}$)} & \colhead{ $T_{\rm eq}[K]$ \tablenotemark{a} } & \colhead{ $T_{\rm deep}[K]$\tablenotemark{b}} & \colhead{Age (Gyr)} & \colhead{Reference} } \startdata OGLE-TR-132 & 0.031 & $1.19 \pm 0.13$ & $1.13 \pm 0.08$ & 2100& $\leq 2200$ & 0--1.4 & 1 \\ OGLE-TR-56 & 0.023 & $1.24 \pm 0.13$ & $1.25 \pm 0.08$ & 2100 & $ 1000-3100$ & $3 \pm 1$ & 2,3,12 \\ HD 209458 & 0.046 & $0.69 \pm 0.05$ & $\rm 1.31^{+0.05}_{-0.05}$ & 1500 & 2200-2800 & 4--7 & 4,5\\ OGLE-TR-10 & 0.042 & $0.63 \pm 0.14$ & $1.14 \pm 0.09$ & 1500 & $\leq 2600$ & -- & 6,12 \\ OGLE-TR-113 & 0.023 & $1.35 \pm 0.22$ & $\rm 1.08^{+0.07}_{-0.05}$ & 1300 & $\leq 2100$ & -- & 7 \\ TrES-1 & 0.039 & $0.73 \pm 0.04$ & $\rm 1.08^{+0.05}_{-0.05}$ & 1200 & $\leq 1000$ & $2.5 \pm 1.5$ & 5,8 \\ OGLE-TR-111 & 0.047 & $0.52 \pm 0.13$ & $\rm 0.97 \pm 0.06$ &1000 & $\leq 1200$ & -- & 9,12 \\ HD 149026 \tablenotemark{c} & 0.042 & $0.36\pm 0.04$ & $0.725 \pm 0.05$ & 1700 & & $2.0 \pm 0.8$ & 10 \\ HD 189733 & 0.031 & $1.15\pm0.04$ & $1.26\pm 0.03$ & 1200 & $\leq 3200$ & & 11 \\ \enddata \tablenotetext{a}{Here $T_{\rm eq} \equiv T_{\rm *}(R_*/2a)^{1/2}$. See the discussion following eq.\ (\ref{eq:Teq}).} \tablenotetext{b}{Allowed range of $T_{\rm deep}$ given range of mass, radius and age. If no age range given in the literature, we (arbitrarily) give the maximum value of $T_{\rm deep}$ for an age less than 10Gyr. However, given an accurate age range, the figures in \S \ \ref{sec:applications} can be used to obtain stronger constraints than given here.} \tablenotetext{c}{HD 149026's small radius clearly indicates a large core size or heavy element abundance. The present paper does not include heavy element cores, so we do not discuss HD 149026 further.} \tablerefs{(1) \cite{2004A&A...424L..31M}, (2) \cite{2004ApJ...609.1071T}, (3) \cite{2003ApJ...596.1327S}, (4) \cite{2002ApJ...569..451C}, (5) \cite{2005ApJ...621.1072L}, (6) \cite{2005ApJ...624..372K}, (7) \cite{2004A&A...421L..13B}, (8) \cite{2004ApJ...616L.167S}, (9) \cite{2004A&A...426L..15P}, (10) \cite{2005ApJ...633..465S},(11)\cite{2005astro.ph.10119B}, (12) \cite{2006astro.ph..1024S} } \end{deluxetable*} Following the discovery of the planet orbiting 51 Peg \citep{1995Natur.378..355M,1995AAS...187.7004M}, more than 160 planets have been found around nearby stars using precision Doppler spectroscopy. \footnote{For up to date catalogs, see http://exoplanets.org/ and http://obswww.unige.ch/~udry/planet/planet.html.} Theories of planet formation now have the demanding task of explaining the existence of gas giants with semi-major axes one hundred times smaller than Jupiter, others with order unity orbital eccentricities, a detailed spectrum of (minimum) planet masses, and metallicity correlations with the parent star. The discovery of transiting planets in the last five years (see Table \ref{tab1}) challenges not only theories for the origin of short-period gas giants, but also their structure and thermal evolution, spectrum, and interior fluid dynamics. Measurements of planetary mass, radius, and (stellar) age test cooling models which predict radius as a function of mass and age. The atmospheres of two planets have been directly observed. For HD 209458b, absorption lines (due to stellar photons passing through planet's atmosphere) have been found \citep{2002ApJ...568..377C, 2003Natur.422..143V, 2004ApJ...604L..69V}, and the first detections of photons emitted by planets outside our solar system have been made for the thermal emission from HD 209458b \citep{2005Natur.434..740D} and TrES-1 \citep{2005ApJ...626..523C}. These observations directly constrain the atmospheric structure, temperature profile and chemical composition near the photosphere. Evolution of the short orbital period transiting exoplanets is significantly different than for Jupiter and Saturn due to proximity of the parent star \citep{1996ApJ...459L..35G}. Irradiation increases the photospheric temperature by nearly an order of magnitude relative to an isolated planet. Irradiation also decreases the cooling rate, and hence the rate of shrinkage, by altering the surface boundary condition \citep{2000ApJ...534L..97B}. This is immediately apparent in Table \ref{tab1} as many transiting extra-solar giant planets (EGP's) have radii significantly larger than Jupiter. As short period planets are expected to be tidally synchronized \citep{1996ApJ...459L..35G,1997ApJ...481..926M}, the strong day-night temperature contrast will drive winds to transport heat from the day to the night side \citep{2002A&A...385..166S}. Hence the atmospheric temperature profile depends on a combination of detailed radiative transfer calculations for absorption of starlight, and hydrodynamics to model day-night winds and dissipation of wind kinetic energy. Lastly, tides raised on the planet by the parent star may significantly affect its thermal evolution \citep{2001ApJ...548..466B,2002A&A...385..166S}. The free energy available by synchronizing the planet's spin or circularizing the orbit are comparable or larger than the thermal energy. Hence {\it if} the heat can be deposited sufficiently deep in the planet in less than a cooling timescale, the cooling can be slowed, or even reversed. However, it is uncertain if tides can deposit heat deep in the planet \citep{1997ApJ...484..866L, 2004ApJ...610..477O, 2004astro.ph..7628W} . Evolutionary models show that the cooling rate is quite sensitive to the uncertain surface boundary condition \citep{2002A&A...385..156G}. This boundary condition has been implemented using various approximations. Full radiative transfer calculations \citep{2001ApJ...556..885B, 2003ApJ...594.1011H} of {\it static} atmospheres include the stellar irradiation self-consistently, and determine the temperature structure for a given cooling flux from the deep interior. These calculations compute (rather than assume) the albedo, and determine the temperature rise due to absorption of starlight (the greenhouse effect). Such detailed radiative transfer solutions have been incorporated as boundary conditions for some evolutionary calculations \citep{2003A&A...402..701B, 2003ApJ...594..545B}. However, as day-night and equator-pole winds are not included, assumptions must be made about how the stellar flux is deposited over the surface of the planet (only day-side versus evenly over the entire surface, etc.) which directly affect the temperature profile. Other evolutionary calculations (e.g. Bodenheimer et al. 2003) solve the radiation diffusion equation and set the temperature at (infrared) optical depth $2/3$ to be the equilibrium temperature, ignoring additional temperature increase due to absorption of starlight. Lastly, a number of groups \citep{2002A&A...385..166S, 2003ApJ...587L.117C, 2005ApJ...618..512B, 2005ApJ...629L..45C, 2005A&A...436..719I} are beginning to model the day-night winds on tidally locked, short orbital period planets, and the role of clouds \cite{2003ApJ...589..615F} . As we stress here, the crucial parameter for the cooling rate is the temperature at the radiative-convective boundary, which is orders of magnitude deeper in pressure than the photosphere. The plan of the paper is as follows. The uncertain surface boundary condition is discussed in \S \ \ref{sec:surfacebc}, motivating the surface isotherm used in our models. Details of cooling models and microphysical input are described in \S \ \ref{sec:numerics}. In \S \ \ref{sec:luminosity} we compute the dependence of the luminosity on planet mass, core entropy and irradiation. An analytic solution for the temperature profile in the radiative zone, and the position of the radiative-convective boundary are derived in \S \ \ref{sec:transition}. These results are collected together in \S \ \ref{sec:cooling} to derive an analytic cooling model which exhibits simple power-law dependence on time. The radii of irradiated gas giant planets are discussed in \S \ \ref{sec:radiusev}, and the analytic formula for the radius given in eq.\ (\ref{eq:Rtfit}). We apply our models to the observed transiting planets and give constraints on the temperature of the deep surface isotherm in \S \ \ref{sec:applications}. Our main conclusions are summarized in \S \ \ref{sec:conclusions}. \section{ Surface Boundary Condition } \label{sec:surfacebc} The surface boundary condition we adopt is to set the temperature $T \equiv T_{\rm deep}$ at a sufficiently large optical depth that the stellar light is fully absorbed, and the radiation diffusion approximation is valid. This choice of surface boundary condition has also recently been advocated by \citet{2005A&A...436..719I}, based on the results of time-dependent radiative models for the atmosphere of HD 209458b. We motivate our choice with a simple toy problem, and then discuss its relation to detailed radiative transfer solutions for the atmosphere. The atmosphere is heated by absorption of starlight, and possibly dissipation of day-night winds and tidal flows. Let there be an energy deposition rate $\varepsilon$ per unit volume in a radiative region of thickness $\Delta z$. Choose boundary conditions $T=0$ at the top (for simplicity) and outward flux $F=0$ at the base of the heated layer. The latter choice is required in steady state so that the temperature deeper in the atmosphere not increase in time. The flux generated in the layer, which exits the planet, is $F = \varepsilon \Delta z$, and the temperature of the deep atmosphere is $T_{\rm deep} \sim (\tau F/\sigma)^{1/4}$, where $\tau=\kappa \rho \Delta z$ is the optical depth, $\rho$ is the density and $\kappa$ is the opacity. Hence an atmosphere subject to intense heating is expected to develop a deep isothermal region below the heated layer, the temperature determined primarily by the energy flux and depth of the layer, through $\tau$. This estimate of the deep isotherm temperature is similar to that found for absorption of starlight for the proper choice of $\tau$ \citep{2003ApJ...594.1011H}. We now discuss the temperature profile for static atmospheres in more detail. In the absence of external irradiation, the photosphere of a planet will cool to a temperature $T_{\rm cool} \sim (F_{\rm cool}/\sigma)^{1/4} \sim 100\ {\rm K}$ in a few Gyr's, where $F_{\rm cool}$ is the flux from the deep interior. A characteristic temperature at small optical depth for an irradiated planet can be defined by balancing absorbed and emitted energy flux. For a star with mass $M_*$, radius $R_*$ and effective temperature $T_*$ a distance $a=(GM_*P_{\rm orb}^2/4\pi^2)^{1/3}$ away, this ``equilibrium'' temperature is \be && T_{\rm eq} \equiv T_* (R_*/2a)^{1/2} \nonumber \\ & \simeq & 1400\ {\rm K} \left(\frac{{\rm 3\ day}}{P_{\rm orb}}\right)^{1/3} \left(\frac{T_*}{6000\ {\rm K}}\right) \left(\frac{R_*}{R_\odot}\right)^{1/2} \left(\frac{M_\odot}{M_*}\right)^{1/6}, \label{eq:Teq} \ee an order of magnitude larger than for an isolated planet. Hence the surface boundary condition is drastically altered from the isolated case. In general, the irradiated boundary condition will cause the planet to cool slower \citep{2000ApJ...534L..97B}, as we discuss in detail. As significant horizontal temperature variation is expected above the photosphere, $T_{\rm eq}$ is an average temperature which gives the correct outgoing flux. Eq.\ (\ref{eq:Teq}) assumes zero reflection of the stellar photons, and should be multiplied by $(1-A)^{1/4}$ for nonzero Bond albedo $A$. The (optical) incoming stellar photons not scattered back out of the planet are absorbed at the starlight's photosphere, typically at a pressure $\la 10^6\ {\rm dyne\ cm^{-2}}$. Radiative balance implies an outgoing (infrared) flux $F \sim (T_{\rm eq}/T_{\rm cool})^4 F_{\rm cool} \sim 10^4 F_{\rm cool}$ generated by thermal emission. This large flux may lead to a significant increase in temperature above $T_{\rm eq}$ (the greenhouse effect, e.g. Hubeny et al. 2003). This situation continues to a depth at which the starlight is fully absorbed, at which point the temperature profile becomes isothermal. Hence, a semi-infinite atmosphere subject to external irradiation, and with no internal flux deep in the atmosphere, becomes isothermal at large optical depth. We label the temperature of this deep isotherm $\Tdeep$. Now including the internal cooling flux $F_{\rm cool}$, the temperature will again rise toward the interior, the gradient eventually becoming large enough for convection to occur. Since the cooling luminosity is generated in deep layers with sufficiently large optical depth that the stellar light is fully absorbed, the radiation diffusion approximation is valid there. Furthermore, we will show in \S \ \ref{sec:transition} that the temperature profile becomes isothermal within a pressure scale height of the radiative-convective boundary. Hence the problem of determining the cooling luminosity is insensitive to many of the details of the absorption of starlight. The only input needed from the full radiation transfer problem near the photosphere is the temperature of the deep isotherm, $T_{\rm deep}$. \footnote{ We expect that the degree to which this layer is isothermal depends on the number of pressure scale heights separating the optical photosphere from the radiative-convective boundary. Larger irradiation and lower core entropy should make this layer more nearly isothermal.} For tidally locked planets, the day side will be significantly hotter than the night side in static atmospheres with negligible day-night winds. A more uniform temperature distribution results if winds can carry heat from the day to the night side without suffering radiative losses (e.g. Iro et al. 2005). We will show that the cooling luminosity is determined in deep layers with thermal time $t_{\rm th} \ga 10^3\ {\rm yr}$. While significant day-night temperature asymmetries may exist near the optical photosphere, winds moving at even a tiny fraction, $\sim 10^{-5}$, of the sound speed could deposit heat on the night side in less than a thermal time at the depths where the cooling luminosity is determined. Hence we have a strong expectation of a near-spherically symmetric, isothermal temperature profile deep in the radiative layer. \section{ Numerical Models for the Interior } \label{sec:numerics} In the deep interior where the diffusion approximation is valid, we solve the mechanical and thermal structure equations \citep{1959flme.book.....L} \be \frac{dm}{dr} & = & 4\pi r^2 \rho, \\ \frac{dP}{dr} & = & - \frac{Gm\rho}{r^2}, \\ \frac{dT}{dr} & = & \frac{dP}{dr} \frac{T}{P} \nabla, \\ \frac{dl}{dr} & = & \frac{dm}{dr} \left( \varepsilon - T \frac{\partial S}{\partial t} \right) \label{eq:entropy}, \ee for the interior mass $m$, pressure $P$, temperature $T$, and outward luminosity $l$, as a function of radius $r$. Here $S$ is the entropy per gram, and $\nabla=d\ln T/d\ln P$ is the logarithmic temperature gradient. The energy generation $\varepsilon$ is set to zero throughout this paper, as we study passively cooling planets. The subscript ``cool'' on the luminosity will be assumed for the rest of the paper. As the eddy turnover time is much shorter than the cooling time and convection is quite efficient, entropy is very nearly constant in space in the convection zone, but decreases in time due to cooling. Hence we treat $\partial S/\partial t$ as a constant in the convection zone. For numerical convenience, we use this same value of $\partial S/\partial t$ in the surface radiative zone. A negligible luminosity is generated there however, so this error does not affect our results. While the entropy equation (\ref{eq:entropy}) is valid on timescales longer than an eddy turnover time ($\sim {\rm yrs}$) in the convective core, the assumption of nearly spatially constant luminosity is only valid in the radiative envelope on timescales longer than the thermal time ($\sim 10^3\ {\rm yr}$) there. As this is much shorter than the global cooling time, we expect our numerical cooling models to be as accurate as a relaxation (Henyey-type) code. The equation of state (EOS) from \cite{1995ApJS...99..713S} (SCVH) is used with a mixture of $70\%$ hydrogen and $30\%$ helium, ignoring metals, and using the tables which smooth over the plasma phase transition. There are been several improvements to SCVH \citep{1999astro.ph..9168S, 2004ApJ...608.1039F} using recent laser shock- compression data and including the effects of helium phase separation, which change the radii at the few percent level. We use the solar composition ``condensed'' phase opacities from \citet{2001ApJ...556..357A}, which includes the effects of grains in the equation of state, but ignores their opacity, as is appropriate if the grains have condensed out. Mixing length theory is used to calculated $\nabla$ in convective regions, and radiative diffusion in radiative regions. The mixing length is set equal to the pressure scale height. Two boundary conditions are needed at the surface. First, the surface temperature is set to $\Tdeep$, the temperature of the deep isotherm discussed in \S \ \ref{sec:surfacebc}. The second boundary condition is that we specify the surface to be at the (arbitrarily chosen) pressure $P=10^4\ {\rm dyne\ cm^{-2}}$. As the surface layer is isothermal, the contribution to the radius from near-surface layers is larger than for the radiative zero temperature profile, hence care is needed when comparing the radii computed here with previous work. As the radius is somewhat dependent on the problem at hand (optical photosphere versus infrared photosphere, corrections due to geometry in a transit, etc.) we have made this arbitrary choice of the surface for simplicity. The change in radius between pressures $P_1$ and $P_2$ is $\Delta R=\int_{P_1}^{P_2} d\ln P (k_bT/\mu m_p g) \simeq (k_bT/\mu m_p g) \ln(P_2/P_1)$. For example, the radius must be decreased by $\Delta R=-0.022R_J$ for an outer boundary condition $P=10^6\ {\rm dyne\ cm^{-2}}$ for $T=1000\ {\rm K}$ and mean molecular weight $\mu=2.43$ ($70\%$ molecular hydrogen and $30\%$ neutral helium). We do not include a solid core in the present calculations. We make a single model of a planet as follows. Planet mass $M$, core entropy $S$, and surface temperature $\Tdeep$ are treated as fixed parameters. Assuming values for the planet's radius $R$, cooling luminosity $L=l(R)$, $\partial S/\partial t$, and central pressure $P_c$, we integrate outward from the center and inward from the surface. The four parameters are adjusted to make the integration variables ($m$, $P$, $T$, and $l$) continuous at a fitting radius. Given the subroutine to solve for a single model, evolving the planet in time is trivial. As we specify the core entropy $S$, and have solved for $\partial S/\partial t$, we compute the time it takes to cool from one entropy to the next. \section{ Impact of irradiation on heat loss } \label{sec:luminosity} We now show the dependence of the cooling luminosity on the depth of the radiative-convective boundary, emphasizing the role of the opacity deep in the planet. Many of the luminosity dependences can be understood with purely local arguments, without the need to build a global planet model. Hence, qualitative statements can be made about cooling of EGP's under irradiation just given EOS and opacity tables. We make comparisons between the local arguments and the global numerical calculations as well. The convective core is capable of transporting enormous luminosities through fluid motion. Hence it is the large thermal resistance of the outer radiative envelope that determines the cooling flux. For an opacity which increases inward from the surface, this resistance is largest at the base of the radiative layer, hence it is the radiative-convective boundary that determines the cooling flux. This boundary is moved to higher pressures by irradiation \citep{1996ApJ...459L..35G}. The outward flux carried by radiative diffusion is \be F & = & - \frac{16\sigma T^3}{3\kappa \rho} \frac{dT}{dr}, \label{eq:flux} \ee where $\kappa$ is the Rosseland mean opacity. The maximum flux which can be carried by radiative diffusion is found using the adiabatic temperature gradient $dT/dr|_{\rm ad}=(\nabla_{\rm ad}T/P)(-Gm\rho/r^2)$, where $\nabla_{\rm ad}=\partial \ln T/\partial \ln P |_S$ ($=2/7$ for an ideal gas with five degrees of freedom) is the adiabatic temperature gradient. Multiplying by $4\pi r^2$, the maximum luminosity per unit mass which can be carried by radiative diffusion at a local temperature $T$, pressure $P$, opacity $\kappa(T,P)$, and enclosed mass $m \simeq M$ is \be \frac{L}{M} & = & \frac{64\pi G}{3} \frac{\sigma T^4}{\kappa P} \nabla_{\rm ad}. \label{eq:Lmax} \ee Choosing an entropy $S$, the right hand side of eq.\ (\ref{eq:Lmax}) can be evaluated along an adiabat out from the center, yielding the cooling flux for a specified temperature $T_{\rm rcb}$ at the radiative-convective boundary. Eq.\ (\ref{eq:Lmax}) shows that the luminosity per unit mass depends solely on the entropy and irradiating flux, and that the luminosity is proportional to the planet's mass. \footnote{It is commonly stated that the luminosity {\it decreases} with increasing mass. This is true at fixed core temperature, rather than fixed entropy. The derivatives can be related by $\partial \ln L/\partial \ln M|_{T_c}=\partial \ln L/\partial \ln M|_S + \partial \ln L/\partial S|_M \partial S/\partial \ln M|_{T_c}$. At fixed core temperature, entropy increases for decreasing mass (see Figure \ref{fig:Tcore_vs_M}).} Figure \ref{fig:T_vs_P} shows the run of temperature versus pressure from numerical models for a range of $M$, $S$ and $\Tdeep$. Choosing $S$ and $\Tdeep$, the temperature profile must follow the adiabat deep in the planet and the isotherm near the surface. An even stronger statement can be made, however. The temperature profile over the entire planet from the center to the top of the deep isotherm depends only on $S$ and $\Tdeep$, and is independent of $M$ (since eq.\ (\ref{eq:Tprofile}) and (\ref{eq:Pdeep}) depend only on $F/g \propto L/M$). Next, for a given irradiation flux (fixed $\Tdeep$), the radiative-convective transition burrows deeper into the planet with time (decreasing $S$). Increasing the irradiation flux at fixed $S$ also moves the radiative-convective region deeper into the planet. Temperature changes in response to small changes in flux in optically thick regions occur on the thermal time, estimated from eq.\ (\ref{eq:entropy}) to be \be t_{\rm th} & = & \frac{PC_pT}{gF} \simeq 10^4\ {\rm yr} \left(\frac{P}{10^8\ {\rm dyne\ cm^{-2}}} \right) \left( \frac{T}{10^3\ {\rm K}} \right) \nonumber \\ && \left( \frac{10^3\ {\rm cm\ s^{-2}}}{g} \right) \left( \frac{10^4\ {\rm erg\ cm^{-2}\ s^{-1}}}{F} \right). \ee Here we have used typical numbers from Figure \ref{fig:T_vs_P} for the radiative-convective boundary. Note that this estimate is much longer than the adjustment time near the optical photosphere ($\sim {\rm days}$, e.g. Iro et al. 2005), as the cooling flux is $\sim 10^4$ times smaller than the stellar flux, and the heat content increases $\propto TP$. As the thermal time at the radiative-convective boundary is so much longer than the horizontal sound travel time ($\sim {\rm days}$), we expect the day-night temperature asymmetry to be small there. Figures \ref{fig:LoverM_vs_T} and \ref{fig:LoverM_vs_T_int} show the local calculation of $L/M$ evaluated along adiabats, and the global calculation of $L/M$ versus $\Tdeep$, respectively. The x-axis in Figure \ref{fig:LoverM_vs_T} is the local temperature, which should be interpreted as $T_{\rm rcb}$, the temperature of the radiative-convective boundary. Care must be taken in Figure \ref{fig:LoverM_vs_T} in regions where $L(T)$ increases inward. As we show in \S \ \ref{sec:cooling}, $L(T)$ must {\it decrease} inward in order for convection to begin. Hence, if the chosen isotherm intersects a region of positive slope, such as the bump near $T=2000-2500\ {\rm K}$, the convection zone actually begins at a deeper point at which the slope $L(T)$ is again negative. Such regions correspond to the flat parts of the curves in Figure \ref{fig:LoverM_vs_T_int}. The result is that the luminosity generally decreases with irradiation temperature, or is roughly constant, but should not increase. This is the origin of the result found by previous investigators \citep{2000ApJ...534L..97B} that irradiated planets cool slower. Comparison of Figures \ref{fig:LoverM_vs_T} and \ref{fig:LoverM_vs_T_int} show rough agreement in regions where $L(T)$ is decreasing, the main discrepancies due to the ratio $T_{\rm rcb}/\Tdeep$ not being precisely a constant (see Figure \ref{fig:T_vs_P}). Figure \ref{fig:LoverM_vs_S_num} shows luminosity versus core entropy for the numerical models. If $\Tdeep$ is constant during the evolution, Figure \ref{fig:LoverM_vs_S_num} shows the change in luminosity as the planet cools. Comparison of lines with different $\Tdeep$ clearly shows the monotonic decrease in luminosity as the irradiation temperature is increased. Aside from models with large mass ($M=3.2M_J$) and irradiation temperature ($\Tdeep=3500\ {\rm K}$) at entropies so low ($S<8k_b/m_p$) as to be unreachable in a Hubble time, the luminosity is proportional to the mass and the curves overlie each other. \section{ Radiative-convective boundary } \label{sec:transition} We now derive a simplified analytic model for the temperature profile at the transition from the surface radiative zone to the core convection zone. We relate $T_{\rm deep}$ to $T_{\rm rcb}$, the radiative-convective boundary temperature where the cooling luminosity is determined. The scalings derived here are used in \S \ \ref{sec:cooling} to derive the scalings of the cooling luminosity. We assume constant gravity $g$, ideal gas pressure $P=\rho k_b T/\mu m_p$ and power-law opacity \footnote{ Significant features in the opacity may be treated as broken power-laws.} $\kappa\equiv\kappa_0 \rho^a T^b\equiv\kappa_1 P^a T^{b-a}$. In the radiative zone, \be F & = & \frac{16\sigma T^3 g}{3\kappa}\frac{dT}{dP} = \frac{a+1}{4+a-b} \frac{16\sigma g}{3\kappa_1}\frac{dT^{4+a-b}}{dP^{a+1}}. \ee When integrating this equation, it's essential to retain the constant of integration. Defining the temperature gradient for a radiative zero solution $\nabla_\infty=(a+1)/(4+a-b)$, we find \be T^{4+a-b} & =& {\rm constant} + \nabla_\infty^{-1} \left(\frac{3\kappa_1 F}{16\sigma g} \right) P^{a+1}. \ee At small pressure, $T \simeq \Tdeep$, so we write the temperature profile as \be T & =& \Tdeep \left[ 1 + \left(P/P_{\rm deep} \right)^{a+1} \right]^{1/(4+a-b)}, \label{eq:Tprofile} \ee which becomes isothermal below the pressure \be P_{\rm deep} & = & \left( \nabla_\infty \frac{16 \sigma g \Tdeep^{4+a-b}}{3\kappa_1 F} \right)^{1/(a+1)}. \label{eq:Pdeep} \ee The logarithmic temperature gradient is then \be \nabla & = & \nabla_\infty \frac{ \left(P/P_{\rm deep} \right)^{a+1} }{1 + \left(P/P_{\rm deep} \right)^{a+1} }, \label{eq:nabla} \ee which decreases sharply over a pressure scale height. A plot of $\nabla$ versus $P$ is shown in Figure \ref{fig:del_vs_P_allard_cond} for several values of $\Tdeep$. The upper envelope of the curves is set by the adiabatic gradient in the convection zone. Increasing $T_{\rm deep}$ moves the boundary inward along the adiabat, $P_{\rm deep} \propto \Tdeep^{1/\nabla_{\rm ad}}$, aside from regions where the opacity changes irregularly. Eq.\ (\ref{eq:nabla}) agrees well with the numerical integrations in regions where the opacity is smooth. To solve for the transition from radiative to convective zone, we set $\nabla=\nabla_{\rm ad}$. As $a+1>0$, one must have the inequality $\nabla_\infty \geq \nabla_{\rm ad}$ for a convection zone to exist. We find the temperature and pressure at the boundary are \be T_{\rm rcb} & = & \Tdeep \left( \frac{\nabla_\infty}{\nabla_\infty-\nabla_{\rm ad}} \right)^{1/(4+a-b)} \label{eq:Ttr} \nonumber \\ P_{\rm rcb} & = & P_{\rm deep} \left( \frac{\nabla_{\rm ad}}{\nabla_\infty-\nabla_{\rm ad}} \right)^{1/(a+1)}, \ee so they differ by a factor of order unity from $\Tdeep$ and $P_{\rm deep}$ unless $|\nabla_{\rm ad}-\nabla_\infty|<<\nabla_\infty$. The decrease of $\nabla_{\rm ad}$ at large pressures seen in Figure \ref{fig:del_vs_P_allard_cond} makes the ratio $T_{\rm rcb}/\Tdeep$ closer to unity for large $T_{\rm deep}$ and $P$, as seen in Figure \ref{fig:T_vs_P}. Also note that at fixed $S$ and $\Tdeep$, $T_{\rm rcb}$ is largely independent of mass. \section{ Analytic cooling model } \label{sec:cooling} In \S \ \ref{sec:luminosity} we found that the luminosity scales with core entropy and irradiating flux over much of the relevant parameter space. Here we show that the core temperature also scales simply with mass and entropy when sufficiently degenerate. As a consequence, we derive an analytic model in which entropy has a simple power-law time dependence at late times. We compare the power-law model against numerical time integrations. The scaling of luminosity with $\Tdeep$ and $S$ can be found by substituting $P(T,S)$ into eq.\ (\ref{eq:Lmax}). Using the thermodynamic relation \be \frac{dS}{C_p} & = & \frac{dT}{T} - \nabla_{\rm ad} \frac{dP}{P}, \label{eq:thermo} \ee and expanding about a reference point $T_{\rm ref}$, $P_{\rm ref}$ and $S_{\rm ref}$, the adiabat is \be P & \simeq & P_{\rm ref} \left( \frac{T}{T_{\rm ref}} \right)^{1/\nabla_{\rm ad}} \exp\left(-\frac{\Delta S}{C_p\nabla_{\rm ad}} \right), \label{eq:adiabat} \ee where $\Delta S=S-S_{\rm ref}$, and we approximate $\nabla_{\rm ad}$ and $C_p$ as constants. For an ideal gas, $C_p \nabla_{\rm ad}=k_b/\mu m_p$, but particle interactions and molecular dissociation reduce $C_p \nabla_{\rm ad}$ below the ideal value. Inserting eq.\ (\ref{eq:adiabat}), (\ref{eq:Ttr}), and the power-law form of the opacity into eq.\ (\ref{eq:Lmax}), we find \be L & \simeq & L_{\rm ref} \left( \frac{T_{\rm deep}}{T_{\rm ref}} \right)^{-\alpha} \exp\left[ \beta \frac{(S-S_{\rm ref})}{k_b/m_p} \right] \label{eq:Lpl} \ee where the exponents are (Figures \ref{fig:LoverM_vs_T_int} and \ref{fig:LoverM_vs_S_num}) \be \alpha & \simeq & (4+a-b)\left(\frac{\nabla_\infty}{\nabla_{\rm ad}} - 1\right) \simeq 0.0-10.0, \nonumber \\ \beta & \simeq & (a+1) \frac{k_b/m_p}{C_p \nabla_{\rm ad}} \simeq 2.5-3.5. \ee Examination of the exponent $\alpha$ shows that irradiation slows the cooling, i.e. luminosity decreases as $T_{\rm deep}$ increases. The condition $\nabla_\infty > \nabla_{\rm ad}$ is required for a core convection zone to exist, hence $\alpha \geq 0$. Evaluation of $\alpha$ depends on the detailed density and temperature dependence of the opacity, which can be found in Figure \ref{fig:LoverM_vs_T}. Features of note are the positive slope near $2000-3000\ {\rm K}$ at which point $\alpha$ become small, and also the steep decrease for $T_{\rm rcb} \geq 3000\ {\rm K}$. The exponent $\beta$ can be estimated for an ideal gas (solar mixture, molecular hydrogen and neutral helium) and density independent opacity to be $\beta \simeq \mu \sim 2.4$. This ideal limit is expected for small $T_{\rm deep}$ and hence low density. As $T_{\rm deep}$ is increased, molecular interactions make the gas less ideal, reducing the value of $C_p \nabla_{\rm ad}$ and increasing $\beta$. This qualitative trend may be seen in Figure \ref{fig:LoverM_vs_S_num}. The core temperature increases during the initial contraction phase when the core is non-degenerate. A maximum is reached when $k_bT_c \simeq E_F$, the Fermi energy, and subsequently $T_c$ decreases as entropy decreases. In this degenerate phase, the core temperature depends mainly on mass and entropy, with only a weak dependence on irradiation. Figure \ref{fig:Tcore_vs_M} shows core temperature versus mass for four adiabats and a range of irradiation temperatures. The dependence on $\Tdeep$ gives only a slight broadening of each adiabat. The dependence on mass is quite simple when sufficiently degenerate. Figure \ref{fig:Tc_vs_S} shows the dependence of core temperature on entropy for a range of masses and irradiation temperatures, showing a simple exponential dependence at low entropy. For the degenerate phase we write $T_c$ in the form (see Figure \ref{fig:Tcore_vs_M}) \be T_c(M,S) & = & T_{c,\rm ref}\left( \frac{M}{M_{\rm ref}}\right)^\gamma \exp\left[ \delta \frac{(S-S_{\rm ref})}{k_b/m_p}\right]. \label{eq:Tcpl} \ee Using hydrostatic balance $P \propto M^2/R^4$, and parameterizing $R \propto M^\lambda$, we estimate the exponents to be \be \gamma & \simeq & \nabla_{\rm ad}(2-4\lambda)\simeq 0.6-0.7 \nonumber \\ \delta & \simeq & k_b/C_pm_p\simeq 0.5. \ee Next we solve for the change in core entropy with time for the analytic model. We treat $T_{\rm deep}$ and $M$ as constants during the evolution. The entropy equation integrated over the convective core gives \be \frac{\partial S}{\partial t} & = & - \frac{L/M}{fT_c}, \label{eq:intentropy} \ee where $f=\int (dm/M)(T/T_c)\simeq 0.6-0.7$ and we have treated $\partial S/\partial t$ as constant in space. Plugging eq.\ (\ref{eq:Tcpl}) and (\ref{eq:Lpl}) into eq.\ (\ref{eq:intentropy}), we find the following solution for the entropy with time \be \exp\left[\frac{S-S_{\rm ref}}{k_b/m_p} \right] & = & \left( 1 + \frac{t}{t_{\rm S}} \right)^{-1/(\beta-\delta)}, \label{eq:Svst} \ee where the characteristic cooling time is \be && t_{\rm S}(M,T_{\rm deep}) \nonumber \\ & = & \left( \frac{f}{\beta - \delta } \right) \left( \frac{ k_b T_c/m_p}{ L/M } \right)_{\rm ref} \left( \frac{T_{\rm deep}}{T_{\rm deep,ref}} \right)^\alpha \left( \frac{M}{M_{\rm ref}} \right)^\gamma. \label{eq:tS} \ee This solution has a number of notable features: \begin{list}{} \item (1) The fiducial evolution time, $Mk_b T_c/m_p L$, is just the {\it initial} time to radiate away the core's thermal energy. \item (2) At late times, $t \geq t_S$, the solution is a power-law in time. As the initial cooling time is very short, the power-law occurs for most of the planet's lifetime. \item (3) The exponent of the power-law involves the change of luminosity and core temperature with respect to entropy. \item (4) Evolution timescale is slowed for large irradiation or large planet mass. The dependence on planet mass comes purely from the dependence of $T_c$ on planet mass (at fixed entropy). The dependence on irradiation is primarily through the luminosity. \end{list} Figure \ref{fig:S_vs_time} shows entropy versus time for a range of mass and irradiation. Note the large spread in $S$ at a given age. For low irradiation, $S \simeq 6-8k_b/m_p$ in the age range $1-10\ {\rm Gyr}$, while $S$ can be as high as $\simeq 10k_b/m_p$ for $\Tdeep=3500\ {\rm K}$. While the range of $\Tdeep$ shown here gives fairly good power-laws, we note that at $\Tdeep<500\ {\rm K}$ there is a break occurring at $\simeq 1\ {\rm Gyr}$, due to the increase in luminosity seen in Figure \ref{fig:LoverM_vs_S_num} below $S=8k_b/m_p$. While we have used the SCVH EOS and Allard et.al.(2001) opacities for numerical estimates, the analytic solution makes it particularly clear which quantities need by evaluated for a given opacity table and EOS. The luminosity is sensitive only to the local conditions at the radiative-convective boundary, while the core temperature involves building static (i.e. not time-dependent) models. \section{ Radius Evolution } \label{sec:radiusev} Planets are initially nondegenerate in their core, and undergo rapid contraction until the core is degenerate. If $T_{\rm eff}$ is constant during the contraction, the energy equation $d/dt(-3GM^2/7R)=-4\pi R^2 \sigma T_{\rm eff}^4$ is solved to find the change in radius with time (see, e.g. Bildsten et al. 1997) \be R(t) & \simeq & 8\ R_J\ \left( \frac{M}{M_J} \right)^{2/3} \left( \frac{300\ {\rm K}}{T_{\rm eff}} \right)^{4/3} \left( \frac{1\ {\rm Myr}}{t} \right)^{1/3}. \ee The core temperature $T_c \simeq GM\mu m_p/Rk_b \propto t^{1/3}$ is increasing during the non-degenerate phase, and reaches a maximum when $k_bT \simeq E_F$. For an ideal gas, $k_bT/E_F$ is a function only of entropy, so that the maximum temperature would occur at the same entropy for all planet masses. Coulomb interactions suppress the value of $\nabla_{\rm ad}$ below $2/5$, so that if $T_c \propto M^{4/3}$ and $P_c \propto M^{10/3}$, the entropy at maximum temperature will increase a bit with mass. This can be seen in Figure \ref{fig:Tc_vs_S}, as the two higher masses have maxima at higher entropy (off the plot) than the lowest mass. Once degeneracy sets in, the radius is primarily determined in the degenerate core of the planet, although as irradiation is increased the contribution from the outer envelope becomes larger due to the increased scale height. This is clarified by writing the radius as an integral over pressure \be r(P) & = & \int^{P_c}_P d\ln P \left( \frac{P}{\rho g} \right) . \ee In the degenerate core, $P/\rho \propto E_F$ while in the nondegenerate envelope $P/\rho \propto k_bT$. The contribution from the core is larger when $E_F \gg k_b T$ unless the number of pressure scale heights in the envelope is much larger than the core. The equation of state in the core for $M<M_J$ is complicated by strong Coulomb interactions. For illustrative purposes, an approximate equation of state including the leading order contributions from Coulomb interactions as well as ideal ion pressure is \be P & = & n_e \left( \frac{2}{5} E_F - \frac{3}{10} \frac{Z^2e^2}{a_i} \right) + \frac{\rho k_b T}{\mu_i m_p}, \label{eq:simpleeos} \ee where $a_i=(4\pi \rho/\mu_i m_p)^{1/3}$ is the mean ion spacing and $\mu_i m_p$ is the mean ion mass. The energy scales relevant for the core are $E_F=(\hbar^2/2m_e)(3\pi^2 \rho/\mu_e m_p)^{2/3} \simeq 26\ {\rm eV}(\rho\mu_e^{-1}{\rm g\ cm^{-3}})^{2/3}$, $k_b T \simeq 0.9\ {\rm eV} (T/10^4\ {\rm K})$, and $E_{\rm coul}=Z^2e^2/a_i \simeq 20\ {\rm eV} Z^2 (\rho\mu_i{\rm g\ cm^{-3}})^{1/3}$ is the Coulomb interaction energy between a nucleus and uniform electron cloud. Ignoring the ion pressure term, the density at zero pressure is $\rho_{\rm zp}\simeq 0.2 \mu_e Z^2 {\rm g\ cm^{-3}}$. As the central density in Jupiter mass objects is near $\rho_{\rm zp}$, Coulomb interactions (and further the tendency to form bound states) stiffen the EOS, and are important in determining the radius. Figure \ref{fig:R_vs_M_irr} shows mass versus radius for a range of core entropy and irradiation. The effects of irradiation are seen to be most severe at low mass and low entropy, since $\Tdeep$ is becoming a significant fraction of the core temperature. At $M \simeq M_J/2$ and low entropy, the range of irradiation temperatures shown here can change the radius by as much as $50\%$. Radii for fully adiabatic planets (not shown here) agree well with the $\Tdeep=500\ {\rm K}$ lines. Figure \ref{fig:R_vs_S} shows radius versus entropy for a range of masses and irradiation temperature. At late times in the evolution when the entropy is small, the radius is converging to some constant value which depends on {\it both} $M$ and $\Tdeep$. If the planet were allowed to cool under a constant irradiation field indefinitely, it would approach an isothermal state \citep{1977Icar...30..305H} at $T=\Tdeep$ with a radius \footnote{In principle, $R_0$ can be calculated by integrating the structure equations for a given EOS. In practice, such low temperatures and high densities are not covered by the SCVH EOS. In this paper, we compute the isothermal radius by fitting evolutionary curves of radius versus entropy, defining $R_0$ by extrapolating to the small entropy limit.} $R=R_0$. Although in practice planets will never reach this isothermal state, it is the {\it minimum} radius to which the planet is evolving. Furthermore, it is the {\it deviation} around the isothermal radius, $\delta R=R-R_0$ which is changing with age. As we now show, $\delta R$ has a particularly simple behavior with time over the entire observable range $\delta R \leq R$. To motivate the following numerical calculations, we first discuss the change in radius for a fluid element in mass shell $m$ as the entropy is changed. The radius of a mass shell in the convection zone can be written \be r^3(m,S) & = & \frac{3}{4\pi} \int_0^m \frac{dm'}{\rho(m',S)}, \ee hence for fixed interior mass the change in radius with respect to entropy is \be \frac{\partial r}{\partial S} & = & - \frac{1}{4\pi r^2} \int_0^m \frac{dm'}{\rho(m',S)} \frac{\partial \rho(m',S)}{\partial S} \rfloor_{m'}. \ee Given an equation of state $\rho(P,S)$, and switching to radius as the integration variable, we find \be \frac{\partial r}{\partial S} & = & - \frac{1}{r^2} \int_0^r {r'}^2 dr' \left( \frac{1}{C_p} \frac{\partial \ln \rho}{\partial \ln T}\rfloor_{P} + \Gamma_1^{-1} \frac{\partial \ln P}{\partial S}\rfloor_{m'} \right) \label{eq:drds} \ee where eq.\ (\ref{eq:thermo}) has been used. The second term in eq.\ (\ref{eq:drds}) mainly corresponds to a uniform shift in pressure in the core, due to the radius changing. Near the surface this term must go to zero since pressure is proportional to external mass, which is fixed. Consequently, the first term is most important. From eq.\ (\ref{eq:simpleeos}), the volume expansion term is $\partial \ln \rho /\partial \ln T|_P \propto k_bT/E_F$, with a significant correction due to Coulomb interactions which acts to increase the expansion since the electron pressure is effectively lowered. Hence the change in radius in the core is proportional \footnote{ Eq.\ (\ref{eq:simpleeos}) has ignored contributions to the Coulomb correction which depend on temperature, and do not scale linearly with temperature. Using the EOS in \citet{2000PhRvE..62.8554P}, we find the contribution of these terms to the volume expansion seems to be somewhat smaller than the ideal ion pressure. } to $T_c$. As $T_c$ depends exponentially on the entropy (eq.\ [\ref{eq:Tcpl}]), the contribution to the radius from the degenerate core depends exponentially on entropy. In the nondegenerate envelope, $\partial \ln \rho /\partial \ln T|_P \simeq -1$. Plugging this result into eq.\ (\ref{eq:drds}) implies that the change in radius due to the nondegenerate envelope scales linearly with entropy. As a consequence, it is less important than the exponential dependence from the core. A suite of evolutionary calculations has been done for $M/M_J=0.32,1.0,3.2$ and $\Tdeep[K]=500, 1000, ..., 3500$ starting from high entropy and evolved to ages greater than $15\ {\rm Gyr}$. Given the run of $R(S)$, we fit a function \be R(S) & = & R_0+\delta R_0\exp(\eta m_p S/k_b) \label{eq:Rfit} \ee to determine the isothermal radius $R_0$, coefficient $\delta R_0$, and exponent $\eta$. The coefficients $R_0$, $\delta R_0$ and $\eta$ depend on $M$ and $\Tdeep$. The small entropy points were more heavily weighted to force the fit to agree there. The weighting was adjusted until the fit agreed for as large a region in $S$ as possible (for the plots here we used weighting $\propto R^{10}$.) A comparison of the fit against the data for one example is given in Figure \ref{fig:goodfitexample}. The agreement is good at small entropies, and gets worse for large entropy as degeneracy is lifted. We find good agreement between $\eta \sim 0.5-0.7$ and $\delta$ from eq.\ (\ref{eq:Tcpl}), as expected if $\delta R \propto T_c$. The deviation of the radius about the isothermal value is plotted for all runs over the age range $0.1-10\ {\rm Gyr}$ in Figure \ref{fig:dR_vs_S}. Recall that $R_0$ is different for each line. Note that each line is approximately a power-law, even to $\delta R/R_0 \simeq 1$, where the degenerate approximation breaks down. Hence the fitting formula often works better than naively expected. \begin{deluxetable}{rrrrr} \tablecolumns{5} \tablewidth{0pc} \tablecaption{Parameters for the Fitting Function $R(t)$ in eq.\ (\ref{eq:Rtfit}). } \tablehead{ \colhead{$M/M_J$} & \colhead{$\Tdeep [K]$} & \colhead{$\eta/(\beta-\delta)$} & \colhead{$R_0/R_J$} & \colhead{$\delta R_1/R_J$} } \startdata 0.316 & 500 & 0.31 & 0.836 & 0.308 \\ 0.316 & 1000 & 0.25 & 0.881 & 0.330 \\ 0.316 & 1500 & 0.23 & 0.905 & 0.359 \\ 0.316 & 2000 & 0.23 & 0.952 & 0.361 \\ 0.316 & 2500 & 0.32 & 1.05 & 0.401 \\ 0.316 & 3000 & 0.43 & 1.15 & 0.867 \\ 0.316 & 3500 & 0.50 & 1.30 & 2.49 \\ 1.00 & 500 & 0.16 & 0.825 & 0.320 \\ 1.00 & 1000 & 0.16 & 0.894 & 0.273 \\ 1.00 & 1500 & 0.15 & 0.902 & 0.280 \\ 1.00 & 2000 & 0.16 & 0.930 & 0.266 \\ 1.00 & 2500 & 0.24 & 0.992 & 0.264 \\ 1.00 & 3000 & 0.25 & 0.993 & 0.439 \\ 1.00 & 3500 & 0.35 & 1.09 & 0.727 \\ 3.16 & 500 & 0.20 & 0.915 & 0.243 \\ 3.16 & 1000 & 0.16 & 0.934 & 0.238 \\ 3.16 & 1500 & 0.17 & 0.947 & 0.227 \\ 3.16 & 2000 & 0.18 & 0.962 & 0.217 \\ 3.16 & 2500 & 0.26 & 0.994 & 0.239 \\ 3.16 & 3000 & 0.30 & 1.02 & 0.379 \\ 3.16 & 3500 & 0.41 & 1.10 & 0.673 \\ \enddata \label{tab:fits} \end{deluxetable} We now combine the power-law cooling result in eq.\ (\ref{eq:Svst}) and (\ref{eq:tS}) with the fit for the radius in eq.\ (\ref{eq:Rfit}) to find \be R(t) & = & R_0 + \delta R_0 \exp(\eta m_pS_{\rm ref}/k_b) \left( 1 + \frac{t}{t_{\rm S}} \right)^{-\eta/(\beta-\delta)}. \ee At late times $t \gg t_S$, the deviation in radius from the isothermal value is a power-law in time. In order to provide useful fits to our evolutionary tracks, we parametrize this late time power-law as \be R(t) & = & R_0 + \delta R_1 \left( \frac{{\rm 1\ Gyr}}{t} \right)^{\eta/(\beta-\delta)}, \label{eq:Rtfit} \ee where $R_0$ is again the isothermal radius and $\delta R_1=\delta R_0 \exp(\eta m_pS_{\rm ref}/k_b) (t_{\rm S}/{\rm 1\ Gyr})^{\eta/(\beta-\delta)}$ is the deviation at an age of $1\ {\rm Gyr}$. We fit tracks of $R(t)$ to find the coefficients $R_0$, $\delta R_1$ and $\eta/(\beta-\delta)$ in the same way as the fits for $R(S)$ in eq.\ (\ref{eq:Rfit}). The coefficients are given in Table \ref{tab:fits}. Comparison between the numerical evolutionary tracks for $R(t)$ and the analytic fit in eq.\ (\ref{eq:Rtfit}) are given in Figure \ref{fig:R_vs_age}. The agreement is generally very good. Approximate values and scalings of the coefficients in Table \ref{tab:fits} can be understood as follows. The expected power-law index $\eta/(\beta-\delta) \simeq 0.6/3.0 = 0.20$ agrees well with the temperature range $\Tdeep=1000-2000\ {\rm K}$ where $L$ is independent of $\Tdeep$. At large irradiation, Figure \ref{fig:dR_vs_S} shows $\eta$ increases and Figure \ref{fig:LoverM_vs_S_num} shows that $\beta$ decreases, explaining the increase in $\eta/(\beta-\delta)$. Since $\delta R_1 \propto t_S^{\eta/(\beta-\delta)} \propto \Tdeep^{\alpha \eta/(\beta-\delta)}$, regions of constant (decreasing) slope in Figure \ref{fig:LoverM_vs_T_int} correspond to $\delta R_1$ being constant (increasing). The magnitude of $\delta R_1$ can be estimated from Figure \ref{fig:dR_vs_S} and eq.\ (\ref{eq:tS}). Interestingly, $R_0$ can be somewhat bigger for $M=0.32M_J$ than for the higher masses. While a larger radius is expected for strong irradiation, we caution the reader about interpretation of the exact values for $R_0$. It would be interesting to compare the values obtained by fitting tracks with actual calculations of isothermal planets given a sufficiently accurate low temperature EOS. Given measurements of planetary mass, radius and age, $T_{\rm deep}$ can be constrained. Figure \ref{fig:Tdeep_vs_age} shows the value of $T_{\rm deep}$ required to explain a planet of a given mass and radius, as represented by different lines, as a function of age. The lines slope up to the right since the cooling must be slower (higher $T_{\rm deep}$) to reach the same radius at larger age. As the lines are not horizontal or vertical, there is significant degeneracy between $T_{\rm deep}$ and age. Large radii in the age range $1-10\ {\rm Gyr}$ can only be explained by large irradiation temperatures for the mass range $0.32-3.2 M_J$. For each mass and radius, there is a minimum age which is set by the unirradiated planet, resulting in a steep slope down to the left. We shall use Figure \ref{fig:Tdeep_vs_age} in \S \ \ref{sec:applications} to constrain $T_{\rm deep}$ for the observed transiting planets. \section{Applications to Transiting Planets} \label{sec:applications} We now compare our theory to the observed masses and radii of the transiting planets (Table \ref{tab1}). Figure \ref{fig:R_vs_M_data} shows radius versus mass for the observed transiting planets. The points with errorbars are the data. The three different hatched regions show $T_{\rm deep}=500, 2500, 3000\ {\rm K}$ from bottom to top. The change is gradual from $T_{\rm deep}=500$ to $2500$, and then accelerates for higher temperatures (see Figure \ref{fig:LoverM_vs_T_int}). Within each hatched region, a spread of ages from $1$ (top) to $10\ {\rm Gyr}$ (bottom) is shown. The radius of HD 149026 is so small as to be well outside the plot. It clearly has a large abundance of heavy elements. The radii of the other eight planets can be broadly explained with solar composition, ages in the range $1-10\ {\rm Gyr}$, and temperatures deep in the atmosphere $T_{\rm deep} \leq 3000\ {\rm K}$. The largest radii requiring the most irradiation are HD 209458, HD 189733 and OGLE-TR-56. There are significant uncertainties in fitting stars on the main sequence to find stellar ages. Hence there is motivation to understand how a range of ages affects the range of observed radii. Figure \ref{fig:dR_vs_S} shows deviation from the isothermal radius by factors $1.1-2$ in the age range $0.1-10\ {\rm Gyr}$. The length of each track gives an idea of the uncertainty in radius due to an uncertainty in age. Using the fitting formula in eq.\ (\ref{eq:Rtfit}), the fractional difference in radius between ages $t_1$ and $t_2$ is $\simeq [\eta/(\beta-\delta)](\delta R_1/R_0)\ln(t_2/t_1)$. For the strongly irradiated case, if we choose characteristic values $\eta/(\beta-\delta)=0.5$, $\delta R_1/R_0=0.5$, and a factor of two error in age $t_2=2t_1$, the fractional error in radius is $17\%$. Hence, the age dependence for strongly irradiated planets is important because (i) the decrease in time is steep, and (ii) strong irradiation increases the size of $\delta R_1$ relative to $\delta R_0$. Next, the parameter $T_{\rm deep}$ is crucial for the cooling rate, but is not directly measurable. Here we constrain $T_{\rm deep}$ using measured mass, radius and age. We then compare $T_{\rm deep}$ to the equilibrium temperature. We interpolate over the ${\rm age}-T_{\rm deep}$ tracks in Figure \ref{fig:Tdeep_vs_age} for the mass and radii appropriate for each planet (except HD 149026, which we do not discuss). Since the uncertainty in $T_{\rm deep}$ due to the error bar in planet mass is smaller than that due to the error bar in radius, we fix the planet mass at the central value and only vary the radius. For those planets with an age range quoted in the literature, we show the age range in the plot, and derive the range of $T_{\rm deep}$ consistent with the age range. These values are listed in Table \ref{tab1}. For those planets with no age determination, we find the maximum value of $T_{\rm deep}$ consistent with an age less than $10\ {\rm Gyr}$. The constraints on $T_{\rm deep}$ for all planets except HD 149026 are shown in Figures \ref{fig:HD209458TRES1}, \ref{fig:OGLETR56OGLETR10}, \ref{fig:OGLETR111OGLETR113} and \ref{fig:OGLETR132HD189733}, and summarized in Table \ref{tab1}. The $T_{\rm deep}$ of HD 209458 is best constrained due to the small error bar on mass and radius, as well as detailed fitting of the parent star to find the age. From Figure \ref{fig:HD209458TRES1} we find $T_{\rm deep}=2200-2800\ {\rm K}$; HD 209458b is not consistent with an un-irradiated planet. OGLE-TR-56 has a weak lower limit on $T_{\rm deep}$ which is far less than the equilibrium temperature. All other planets with age constraints have only upper limits, set by the upper limit on the radius, since the lower limit on the radius is consistent with no irradiation. Plots are provided for those planets with no age constraints at the present time. The equilibrium temperature $T_{\rm eq} \equiv T_{\ast}(R_\ast/2a)^{1/2}$ is measurable, but plays no part in our model. On the other hand, the temperature of the deep isotherm is not measurable, but is crucial for the cooling rate. From radiative transfer models, we expect these two temperatures to be roughly proportional, the exact ratio determined by the size of the greenhouse effect (\S \ref{sec:surfacebc}). Hence, they should be strongly correlated. Figure \ref{fig:TeqvsTdeep} shows a plot of $T_{\rm eq}$ versus $T_{\rm deep}$. The large error bars on $T_{\rm deep}$, due to large error bars on the radius, prevent one from drawing robust conclusions. It is in principle possible for a correlation (sloping up to the right) to exist given the current error bars, however, it is not required. Tighter constraints on $T_{\rm deep}$ require the following: \begin{itemize} \item Significantly smaller error bars on the radii of OGLE-TR-56 and OGLE-TR-132. \item An age estimate is needed for HD 189733. If it is found to have an age $\geq 1\ {\rm Gyr}$, $T_{\rm deep}$ will be well constrained with a value much larger than $T_{\rm eq}$, similar to HD 209458b. \item Age estimates are needed for OGLE-TR-10, OGLE-TR-111, and OGLE-TR-113. However, given the present error bar on radii of OGLE-TR-10 and OGLE-TR-113, $T_{\rm deep}$ will be constrained only at the factor of two level. OGLE-TR-111 is an interesting case, as it must be older than $\sim 3-4\ {\rm Gyr}$ to be consistent with our model. \end{itemize} We encourage efforts in these directions. \section{conclusions} \label{sec:conclusions} We have presented calculations of cooling and radius evolution for strongly irradiated planets. Novel aspects of this model are the following: \begin{itemize} \item We argue that the generic outcome of strong surface heating, whether it be due to absorption of stellar flux or dissipation of winds and tidal flow, is that a deep isothermal region exists above the radiative-convective boundary. The thermal time in this layer is sufficiently long that the temperature profile is approximately spherically symmetric, irrespective of the size of the asymmetry near the photosphere. We assign this region the temperature $T_{\rm deep}$ and treat it as a boundary condition for the cooling models. \item We show that the cooling flux is determined at the radiative-convective boundary, which is much deeper than the photosphere. Scalings of the flux with core entropy, $T_{\rm deep}$, and mass are computed. \item These scalings allow us to derive an analytic model for the cooling, which shows power-law decrease over a large range of parameter space. The part of the radius which changes in time (the deviation from the isothermal planet) is also a power-law. An analytic formula for radius evolution is given in eq.\ (\ref{eq:Rtfit}), with coefficients in Table \ref{tab:fits}. \item While we have used the SCVH EOS and Allard et.al.(2001) opacities for numerical estimates, the analytic solution makes it particularly clear which quantities need to be evaluated for a given opacity table and EOS. The luminosity is sensitive only to the local conditions at the radiative-convective boundary, while the core temperature involves building static (i.e. not time-dependent) models. \end{itemize} We have compared our theory to observed masses and radii for the transiting planets in Table \ref{tab1} (except for HD 149026, which clearly has a large abundance of heavy elements). Our findings are as follows: \begin{itemize} \item Figure \ref{fig:R_vs_M_data} shows mass versus radius for eight transiting planets, compared to our model. The radii can be broadly explained with solar composition, ages in the range $1-10\ {\rm Gyr}$, and temperatures deep in the atmosphere $T_{\rm deep} \leq 3000\ {\rm K}$. The largest radii requiring the most irradiation to explain are HD 209458, HD 189733 and OGLE-TR-56. \item Figures \ref{fig:HD209458TRES1}, \ref{fig:OGLETR56OGLETR10}, \ref{fig:OGLETR111OGLETR113} and \ref{fig:OGLETR132HD189733} show constraints on $T_{\rm deep}$ using measured masses, radii, and ages (when available), and their uncertainties. We find that only HD 209458b is well constrained, with $T_{\rm deep}=2200-2800\ {\rm K}$. OGLE-TR-56 has a weak lower limit, and the other six planets have only upper limits, due to the large measurement uncertainty in the radius, or lack of an age determination. These constraints are summarized in Table \ref{tab1}. \item The equilibrium temperature $T_{\rm eq}$ is measurable, but plays no part in our model. The deep isothermal temperature $T_{\rm deep}$ is not measurable, but is crucial for the cooling rate. Radiative transfer calculations find these two temperatures should be strongly correlated. Figure \ref{fig:TeqvsTdeep} shows $T_{\rm eq}$ versus $T_{\rm deep}$. As only upper limits on $T_{\rm deep}$ are available for all but HD 209458b and OGLE-TR-56, it is difficult to draw conclusions at the present time. It is in principle possible for a correlation to exist given the current error bars, however, it is not required. \end{itemize} We hope that our models have illuminated the need for more accurate ages and radii. Once those are in hand, our calculations will provide a measurement of $T_{\rm deep}$ of adequate accuracy to compare to $T_{\rm eq}$, thus constraining greenhouse physics and day-night transport. \vspace{4cm} \acknowledgements This project arose out of a lunchtime conversation with Adam Burrows discussing cooling models for gas giant planets. We thank Tristan Guillot and France Allard for helpful advice on opacities. We also thank Omer Blaes, Shane Davis, Eric Pfahl and Evan Scannapieco for useful discussions. We would also like to thank the referee for constructive comments which improved the presentation of this paper. Phil Arras was supported by the NSF Astronomy and Astrophysics Postdoctoral Fellowship, and the Kavli Institute for Theoretical Physics during this project. This work was supported by the National Science Foundation under grants PHY99-07949 and AST02-05956.
Title: The flat synchrotron spectra of partially self-absorbed jets revisited
Abstract: Flat radio spectra with large brightness temperatures at the core of AGN and X-ray binaries are usually interpreted as the partially self-absorbed bases of jet flows emitting synchrotron radiation. Here we extend previous models of jets propagating at large angles to our line of sight to self-consistently include the effects of energy losses of the relativistic electrons due to the synchrotron process itself and the adiabatic expansion of the jet flow. We also take into account energy gains through self-absorption. Two model classes are presented. The ballistic jet flows, with the jet material travelling along straight trajectories, and adiabatic jets. Despite the energy losses, both scenarios can result in flat emission spectra, however, the adiabatic jets require a specific geometry. No re-acceleration process along the jet is needed for the electrons. We apply the models to observational data of the X-ray binary Cygnus X-1. Both models can be made consistent with the observations. The resulting ballistic jet is extremely narrow with a jet opening angle of only 5". Its energy transport rate is small compared to the time-averaged jet power and therefore suggests the presence of non-radiating protons in the jet flow. The adiabatic jets require a strong departure from energy equipartition between the magnetic field and the relativistic electrons. These models also imply a jet power two orders of magnitude higher than the Eddington limiting luminosity of a 10 solar mass black hole. The models put strong constraints on the physical conditions in the jet flows on scales well below achievable resolution limits.
https://export.arxiv.org/pdf/astro-ph/0601103
\title[Flat spectra of self-absorbed jets]{The flat synchrotron spectra of partially self-absorbed jets revisited} \author[C.R. Kaiser]{C. R. Kaiser\thanks{crk@soton.ac.uk}\\ School of Physics \& Astronomy, University of Southampton, Southampton SO17 1BJ } \begin{keywords} radiation mechanisms: non-thermal -- radio continuum: general -- methods: analytical -- galaxies: active -- stars: individual: Cygnus X-1 -- stars: outflows \end{keywords} \section{Introduction} The centres or cores of many AGN show a flat radio spectrum in the sense that for the flux density as a function of frequency $\nu$ we observe $F_{\nu} \propto \nu^{\alpha}$ with $\alpha \sim 0$. The high surface brightness temperature associated with these spectra suggests a synchrotron origin of the emission. Observations with high spatial resolution reveals that the flat spectrum arises in the base of jet flows which continue to much larger scales \citep[for a review see][]{tc91}. Similar flat or inverted ($\alpha > 0$) radio spectra are also observed in X-ray binaries in the low-hard state \citep[e.g.][]{rf01}. If optically thin, the flat synchrotron spectrum would imply a power-law energy distribution of the radiating relativistic electrons with a slope of unity. Such a distribution is very unlikely to arise for the usually assumed mechanism for the acceleration of the electrons at shock fronts \citep[e.g.][]{ab78}. A magnetised plasma containing very energetic electrons with a power-law energy distribution will produce a power-law spectrum at high frequencies. The slope of the spectrum, typically $\alpha < 0$, is determined by the slope of the energy distribution. However, below a critical frequency the radiating electrons will re-absorb some of the photons. In this self-absorbed, optically thick regime the spectrum has a power-law slope of $5/2$, independent of the slope of the electron energy distribution \citep[e.g.][]{rl79}. The spectrum of a uniform, self-absorbed synchrotron source therefore shows a pronounced peak. \citet{bk79} pointed out that in a jet the plasma conditions are changing along the flow and therefore the peaks of the self-absorbed spectra of different parts of the jet can occur at different frequencies. If the plasma conditions change such that the spectra peak at the same level, then the overall spectrum, observed with a spatial resolution insufficient to resolve the individual parts of the jet, will be flat. Their model has become the standard tool for interpreting observations of flat radio spectra from jetted sources. In the \citet{bk79} model the jet is assumed to have a conical geometry, i.e. the velocity with which the jet is expanding sideways, is constant. The bulk velocity of the jet material along the jet axis is also assumed to be constant. The magnetic field is assumed to be directed perpendicular to the jet axis and `frozen' into the jet plasma. Adiabatic losses of the electrons are mentioned by the authors, but are assumed to be replenished by an unknown, continuous re-acceleration process along the entire jet. The same assumption is made for radiative energy losses associated with the emission of synchrotron radiation. The subsequent model of \citet{am80} includes a simplified treatment of energy losses of the electrons due to adiabatic expansion and radiative processes. It also allows for more confined jets, i.e. the jets are not necessarily conical. With these assumptions, the model is unable to produce a flat emission spectrum. A similar model was developed by \citet{hj88}. They consider adiabatic, but not radiative, energy losses of the electrons. The jet geometry is again conical, but they also investigate a more confined jet. The model can predict flat spectra, but \citet{hj88} point out that these may only arise under special circumstances, particularly in the case of confined jets. The model of \citet{gm98} includes a detailed treatment of the energy losses of the relativistic electrons, but it concentrates only on the optically thin part of the spectrum of jets propagating close to the line of sight for which numerical solutions are presented. The perhaps most comprehensive study of jet emission models is that of \citet{sr82} which includes the effects of energy losses on the electron population, but neglects the effects of self-absorption on the electron energy spectrum. In this paper we extend the previous models by including adiabatic and radiative energy losses and gains (due to absorption) for the electrons as well as investigating various possibilities for the evolution of the magnetic field. We consider two distinct cases: The ballistic and the adiabatic jet models. In the ballistic case the jet material follows straight trajectories and does not behave like a fluid, because individual fluid elements do not interact with each other. In many ways this model is similar to the \citet{bk79} model, but we show that because of self-absorption effects we do not need to invoke a re-acceleration process to achieve flat emission spectra. in the adiabatic jet model the relativistic electrons suffer from adiabatic energy losses as well as radiative losses. Again we show that the models can produce flat spectra without re-acceleration of the electrons, but only for a very specific jet geometry. The emphasis of our treatment is on the construction of analytical models and so we concentrate on jets propagating at large angles to the line of sight, i.e. the viewing angle is larger than the inverse of the Lorentz factor of the jet flow. In Section \ref{conical} we briefly discuss the basic properties of our jets in terms of their geometry, the evolution of the magnetic field and that of the relativistic electrons. We present the first fully analytical solution of the equations governing partially self-absorbed synchrotron emission from a jet in Section \ref{radiation}. Section \ref{simple} summarises the model results for the case without radiative energy losses as studied in many previous models. In Section \ref{cutoff} we develop the formalism for including radiative energy losses in the model and the resulting spectra are discussed in Section \ref{losses}. We apply the model to the data obtained for Cygnus X-1 in Section \ref{obs} and derive the properties of this jet. Finally, we summarise our conclusions in Section \ref{conc}. \section{The model} In this Section we derive the emission properties of partially self-absorbed jets neglecting radiative energy losses of the relativistic electrons. Note that we are concentrating on jets at comparatively large angles to our line of sight, $\vartheta$. As we will point out further down this greatly simplifies the determination of the optical depth of the jet material. \subsection{The basic jet properties} \label{conical} \subsubsection{Jet geometry and velocity} We take the $x$-axis as the centre of a jet that is rotationally symmetric about this axis. The geometrical shape of the jet is then given by a one-dimensional function $r(x)$ defining the jet radius with respect to the $x$-axis. Analogous to previous work we parameterize this function as $r(x)= r_0 \left(x / x_0 \right)^{a_1} = r_0 l^{a_1}$, where $x_0$ is an arbitrary position along the $x$-axis defining the dimensionless coordinate $l$ and $r_0$ is a constant scaling factor. The value of the exponent $a_1$ depends on the details of the confinement of the jet. Confinement by external pressure is the simplest mechanism \citep{br74}, but can lead to problems with the collimation of the jet \citep{bbr84}. Confinement by magnetic fields has also been suggested by various authors, but it is unlikely that magnetic fields alone, without additional gas pressure, can collimate the jet on large scales \citep{mb95}. For our purposes here we do not need to specify the details of the jet confinement and we will assume that $0\le a_1 \le1$. In principle one could also envisage highly overpressured jets with an accelerating expansion rate, i.e. $a_1 > 1$. However, the pressure in such jets would fall very rapidly and they would quickly evolve to a situation where $a_1 \le 1$. The extreme case of a highly overpressured jet is that of a jet flow expanding into a (near) vacuum. In such a ballistic jet, as opposed to the adiabatic, confined jet discussed above, the jet freely expands in the direction perpendicular to the jet axis. In this process random, `thermal' energy of the jet material is converted to ordered, kinetic energy associated with the sideways expansion. The random energy of the electrons giving rise to the synchrotron emission is reduced by this adiabatic expansion. However, in Section \ref{cyg} we apply the ballistic jet model to the observations of the jet in Cygnus X-1. There we will find a very small opening angle for the ballistic jet implying very small adiabatic expansion losses. Therefore we can assume that in the limiting case of a ballistic jet studied here the relativistic electrons do not suffer energy losses other than those associated with radiation processes. We assume in this paper that the velocity of the jet material along the jet axis, $v_{\rm j}$, is constant. While this is justified in the case of the ballistic jet, the confined, adiabatic jets can be accelerated, for example, by a pressure gradient in the confining medium. In the model of \cite{br74} the Lorentz factor of the bulk velocity is proportional to $p_{\rm x}^{-1/4}$, where $p_{\rm x}$ is the pressure of the external medium. As long as the external pressure gradient is shallow, the Lorentz factor of the jet flow will be only a very weak function of the position along the jet axis. Similar arguments hold for a magnetically confined jet. The constant bulk velocity of the jet also implies that a given volume element $\Delta V$ travelling with the jet flow will only expand sideways according to $\Delta V \propto r^2$. A constant jet velocity also simplifies the model greatly as we can ignore the effects of varying length contraction along the jet axis \citep{gk04}. \subsubsection{Magnetic field} The strength of the magnetic field changes during the sideways expansion of the jet material. In general, we parameterize the evolution of the magnetic field as $B = B_0 l^{-a_2}$. For flux freezing of the magnetic field and using flux conservation we have that the field component parallel to the jet axis, $B_{\parallel}$, is proportional to $r^{-2}$. Also, the magnetic field components perpendicular to the jet axis, $B_{\perp}$, are proportional to $r^{-1}$. For an initially mixed field, $B_{\perp}$ will always become the dominant component and so $a_2 = a_1$. The perpendicular magnetic field may also contribute to the confinement of the jet. For completeness we also consider a purely parallel configuration of the magnetic field with $a_2 = 2a_1$. Finally, if the magnetic field is constantly tangled by turbulent motions in the jet material on scales smaller than the jet radius, then it can remain isotropic and it behaves like a relativistic fluid with $B = B_0 l^{-4 a_1 / 3}$ and $a_2 = 4 a_1 /3$ \citep[e.g.][]{hb00}. This is analogous to the behaviour of the magnetic field in an isotropic expansion \citep[e.g.][]{ml94}, but is clearly incompatible with flux freezing. The last case of a permanently isotropic field cannot be realised in the ballistic jet as it would require that the jet material behaves like a fluid. \subsubsection{Relativistic electrons} In order to produce synchrotron emission the jets must contain a population of relativistic electrons. We assume that the latter has a power-law energy distribution of the form \begin{equation} N(E)\,{\rm d}E = \kappa E^{-p} \, {\rm d}E, \end{equation} where $E$ is the electron energy, $E=\gamma m_{\rm e} c^2$, and $\kappa$ is a scaling independent of $E$. $\gamma$ is the Lorentz factor associated with the relativistic motion of the electrons. In this section we do not impose a high-energy cut-off on the energy distribution and we neglect radiative energy losses. Even so the energy distribution of the electrons changes as the jet expands. We represent the evolution of the electron distribution by setting $\kappa = \kappa _0 l^{-a_3}$. For a given volume of jet material $\Delta V$ particle conservation demands that \begin{equation} \Delta V \kappa \gamma ^{-p} \, {\rm d}\gamma = \Delta V_0 \kappa _0 \gamma _0^{-p} \, {\rm d}\gamma_0, \label{dist} \end{equation} where all quantities with subscript `0' refer to their values at $x=x_0$. Therefore for the ballistic jet we have $a_3 = 2 a_1 = 2$. For the adiabatic jet we need to include energy losses due to the jet expansion. Since most of the electrons are highly relativistic we have \citep[e.g.][]{ml94} \begin{equation} \frac{\partial \gamma}{\partial t} = - \frac{1}{3} \gamma \frac{\partial \ln \left( \Delta V \right)}{\partial t}, \label{adiabat} \end{equation} which has the solution \begin{equation} \gamma = \gamma _0 \left( \frac{\Delta V}{\Delta V_0} \right)^{-1/3}, \label{adsol} \end{equation} and it follows that \begin{equation} \frac{\partial \gamma _0}{\partial \gamma} = \left( \frac{\Delta V}{\Delta V_0} \right)^{1/3}. \end{equation} Re-arranging equation (\ref{dist}) and substituting yields \begin{equation} \kappa \gamma ^{-p} \, {\rm d} \gamma = \frac{\Delta V_0}{\Delta V} \kappa _0 \left( \frac{\Delta V}{\Delta V_0} \right)^{-p/3} \gamma^{-p} \left( \frac{\Delta V}{\Delta V_0} \right)^{1/3} \, {\rm d} \gamma. \end{equation} Collecting terms and remembering that $\Delta V \propto r(x)^2$ we find $a_3 = (4+2p) a_1/ 3$. \subsubsection{Individual models} On the basis of the discussion above we formulate five individual models distinguished by the magnetic field behaviour. The ballistic jet models, B1 and B2, as well as the adiabatic models, A1 and A2, correspond to a perpendicular and parallel field structure, respectively. The adiabatic model A3 represents the case of an isotropic magnetic field in the jet. The relevant coefficients describing the jet geometry and the behaviour of the magnetic field and relativistic particles are summarised in Table \ref{expo}. \begin{table} \begin{tabular}{llccccc} & & $a_1$ & $a_2$ & $a_3$ & $a_4$ & $ a_5$\\ \hline ballistic & B1 & 1 & 1 & 2 & $-4-p$ & $\frac{-5}{4+p}$\\[1.5ex] & B2 & 1 & 2 & 2 & $-6-2p$ & $\frac{-3}{2p+3}$\\[1.5ex] adiabatic & A1 & $a$ & $a$ & $\frac{\left(4 + 2p \right) a}{3}$ & $\frac{-a \left(8 + 7p \right)}{3}$ & $\frac{-3 \left( 3a +2 \right)}{a \left(8 +7p \right)}$\\[1.5ex] & A2 & $a$ & $2a$ & $\frac{(4+2p)a}{3}$ & $\frac{-2 a\left( 7 +5p \right)}{3}$ & $\frac{-3 \left(2a +1 \right)}{a \left(7+5p \right)}$\\[1.5ex] & A3 & $a$ & $\frac{4 a}{3}$ & $\frac{(4+2p)a}{3}$ & $\frac{-2a \left(5+4p \right)}{3}$ & $ \frac{-5a - 3}{a \left(5 +4p \right)}$\\ \hline \end{tabular} \caption{Exponents of the model parameters used in this paper. See text for details.\label{expo} } \end{table} \subsection{Partially self-absorbed synchrotron emission from a jet} \label{radiation} From the expressions for $r$, $B$ and $\kappa$ defined in the previous Section, we can now build a model for the emission from the jet. For this purpose we split the jet into small segments of length ${\rm d} x$ along the $x$-axis. We assume that the segments move along the jet axis at a constant velocity $v_{\rm j} = \beta _{\rm j} c$ corresponding to a Lorentz factor $\gamma _{\rm j}$. The jet axis is at an angle $\vartheta$ to the line of sight of the observer and so the Doppler factor for an approaching (`$-$') or receding (`$+$') jet is $\delta _{\mp} = \left[ \gamma _{\rm j} \left( 1 \mp \beta _{\rm j} \cos \vartheta \right)\right]^{-1}$. The observable monochromatic intensity of one such segment taking into account absorption is \begin{equation} I_{\nu} = \delta _{\mp}^3 \frac{J_{\nu}}{4 \pi \chi_{\nu}} \left( 1 - e^{-\chi_{\nu} r(x)} \right), \end{equation} where $J_{\nu}$ is the emissivity per unit volume and $\chi_{\nu}$ is the absorption coefficient. For convenience in the development of the model the frequency $\nu$ is measured in the restframe of the jet material. It is related to the observing frequency by $\nu_{\rm ob} = \delta_{\mp} \nu$. Here and in the following we assume that the average path of a photon through the jet has the length $r(x)$. An exact calculation of the radiative transfer of photons through various jet elements would have to take into account relativistic aberration effects. It is therefore complex and impossible in an analytical model. The assumption of an average path length $r$ will not introduce a large error as long as the angle to the observer's line of sight is large. The jet segment has a surface area of $2 \pi r(x)\,{\rm d} x$ and so the observable flux density of the segment is given by \begin{equation} {\rm d} F_{\nu} = \delta _{\mp}^3 \frac{r(x) J_{\nu}}{2 D^2 \chi_{\nu}} \left( 1 - e^{-\chi_{\nu} r(x)} \right) \, {\rm d} x, \label{pflux} \end{equation} where $D$ is the distance of the jet from the observer. Substituting the dimensionless variable $l=x/x_0$, we can express $J_{\nu}$ and $\chi_{\nu}$ in SI units as \citep{ml94} \begin{eqnarray} J_{\nu} & = & J_0 \nu^{\left(1-p \right)/2} l^{-a_3-a_2 \left( p+1 \right)/2}\nonumber\\ \chi_{\nu} & = & \chi _0 \nu^{\left(-p-4 \right)/2} l^{-a_3-a_2 \left(p+2 \right)/2}, \label{jk} \end{eqnarray} with \begin{eqnarray} J_0 & = & 2.3 \times 10^{-25} \left( 1.3 \times 10^{37} \right)^{\left(p-1 \right)/2} c_1(p) \nonumber \\ & & B_0^{\left( p + 1\right)/2} \kappa _0 \,{\rm W\,m^{-3}\,Hz^{-1}}\nonumber\\ \chi_0 & = & 3.4 \times 10^{-9} \left( 3.5 \times 10^{18} \right)^p c_2(p) B_0^{\left( p +2 \right)/2} \kappa _0 \, {\rm m^{-1}}, \label{jk0} \end{eqnarray} and the constants $c_1(p)$ and $c_2(p)$ given by equations 18.49 and 18.74 in \citet{ml94}. Substituting into equation (\ref{pflux}) and integration gives the total flux density of the jet as \begin{equation} F_{\nu} = \delta _{\mp}^2 \frac{x_0 r_0 J_0}{2 D^2 \chi_0} \nu^{5/2} \int _{l_{\rm min}}^{l_{\rm max}} l^{a_1+a_2/2} \left[ 1 - e^{-\tau} \right] \, {\rm d}l, \label{lflux} \end{equation} where $l_{\rm min}$ and $l_{\rm max}$ are the physical limits of the jet flow along the $x$-axis and the optical depth of the jet material is given by \begin{equation} \tau (l)= \chi_{\nu} r(x) = \chi _0 r_0 \nu^{\left(-p-4\right)/2} l^{a_1-a_3-a_2 \left(p+2 \right)/2}. \label{depth} \end{equation} The reduction in the number of Doppler factors arises from our assumption of a steady state of the jet flow. In principle further relativistic corrections must be applied in the case of mixed optically thin and thick emission, but these corrections are small \citep{tc91} and we neglect them here for simplicity. It is convenient to recast equation (\ref{lflux}) with the help of equation (\ref{depth}) as an integration over optical depth, \begin{equation} F_{\nu} = \delta _{\mp}^2 \frac{x_0 r_0 J_0}{a_4 D^2 \chi _0} \nu^{5/2} \tau _0^{-a_5} \int _{\tau _{\rm max}}^{\tau _{\rm min}} \tau^{a_5-1} \left( 1- e^{-\tau} \right) \, {\rm d} \tau, \label{flux} \end{equation} where \begin{equation} a_4 = 2a_1 -2a_3-\left(p+2\right) a_2 \end{equation} and \begin{equation} a_5 = \frac{2a_1+a_2+2}{a_4}. \end{equation} The coefficients $a_4$ and $a_5$ are listed for the ballistic and adiabatic jets in Table \ref{expo}. $\tau_0$ is given by setting $l=1$ in equation (\ref{depth}) while $\tau_{\rm max} = \tau (l_{\rm min})$ and $\tau_{\rm min} = \tau (l_{\rm max})$, which reflects the fact that the optical depth is always greatest in the innermost regions of the jet. The solution of equation (\ref{flux}) is given by \begin{equation} F_{\nu} = \delta _{\mp}^2 \frac{x_0 r_0 J_0}{a_4 D^2 \chi_0} \nu^{5/2} \tau_0^{-a_5} \left[ \Gamma \left(a_5, \tau \right)+\frac{1}{a_5} \tau^{a_5} \right]_{\tau_{\rm max}}^{\tau_{\rm min}}. \label{sol} \end{equation} Here, the incomplete $\Gamma$-function is defined as \begin{equation} \Gamma \left( a , z \right) = \int_z^{\infty} t^{a-1} e^{-t} \, {\rm d} t. \end{equation} \subsection{Spectra in the absence of radiative energy losses and without a high-energy cut-off} \label{simple} We can immediately recover the well-known solutions for an entirely optically thin ($\tau _{\rm min} \ll 1$ and $\tau _{\rm max} \ll 1$) and an entirely optically thick ($\tau_{\rm min} \gg 1$ and $\tau_{\rm max} \gg 1$) jet. We note that for all choices of $a_1$ discussed above and for physical reasonable values for the power-law exponent $2\le p \le 3$ we find $a_5 < 0$ (see Table \ref{expo}). The incomplete $\Gamma$-function has the series representation \citep{gr00} \begin{equation} \Gamma \left(a, z \right) = \Gamma \left( a \right) - \sum_{n=0}^{\infty} \frac{\left( -1 \right)^n z^{a+n}}{n! \left( a+n \right)}. \label{series} \end{equation} For $\tau \ll 1$ we ignore all terms beyond $n=1$ and thus obtain \begin{equation} \left[ \Gamma \left( a_5, \tau \right)+\frac{1}{a_5} \tau^{a_5} \right]_{\tau_{\rm max}}^{\tau_{\rm min}} \sim \frac{1}{1+a_5} \left( \tau _{\rm min}^{1+a_5} - \tau _{\rm max}^{1+a_5} \right). \end{equation} Because of equation (\ref{depth}) we have $\tau_0 \propto \tau_{\rm min} \propto \tau _{\rm max} \propto \nu^{\left(-4-p\right)/2}$, and from equation (\ref{sol}) it then follows that $F_{\nu} \propto \nu^{\left(1-p \right)/2}$ as expected. For large optical depths we can use the limit for the incomplete $\Gamma$-function, $\lim_{z\rightarrow \infty} \Gamma \left(a, z \right) =0$ \citep{gr00}. Thus \begin{equation} \left[ \Gamma \left( a_5, \tau \right)+\frac{1}{a_5} \tau^{a_5} \right]_{\tau_{\rm max}}^{\tau_{\rm min}} \sim \frac{1}{a_5} \left( \tau_{\rm min}^{a_5} - \tau_{\rm max}^{a_5} \right), \label{largetau} \end{equation} and from the proportionality of the optical depths terms it then follows that $F_{\nu} \propto \nu^{5/2}$, again as expected. The final special case is that of a spatially very extended jet or `long' jet. If the physical dimensions of the long jet, $l_{\rm min}$ and $l_{\rm max}$, are such that $\tau_{\rm max} \rightarrow \infty$ and $\tau_{\rm min} \rightarrow 0$, then for $-1 < a_5 < 0$ we have \begin{equation} \left[ \Gamma \left( a_5, \tau \right)+\frac{1}{a_5} \tau^{a_5} \right]_{\tau_{\rm max}}^{\tau_{\rm min}} \sim \Gamma \left( a_5 \right), \end{equation} which implies \begin{equation} F_{\nu} \propto \nu^{\left[5 +\left( p+4 \right) a_5\right]/2}. \label{longslope} \end{equation} Using the results summarised in Table \ref{expo}, we recover the result of \citet{bk79} that the ballistic jet with the magnetic field perpendicular to the jet axis (model B1) has a flat spectrum, i.e. $F_{\nu}$ is independent of $\nu$, if it is extended and $-1 < a_5 < 0$. The ballistic jet with a parallel magnetic field, model B2, can never produce a flat spectrum for positive $p$ because equation (\ref{longslope}) predicts $F_{\nu} \propto \nu^ {\left(7p +3 \right) / \left[2 \left( 2p + 3\right) \right]}$. For the adiabatic jet models we can substitute the expressions for $a_5$ and find that a flat spectrum is predicted if the exponent $a$ for the geometrical shape of the jet, $r(x) \propto x^a$, is given by \begin{eqnarray} {\rm Model\ A1:} && a = \frac{3p +12}{13p +2}\nonumber\\ {\rm Model\ A2:} && a = \frac{3p+12}{19p+11}\\ {\rm Model\ A3:} && a = \frac{3p+12}{15p+5} \nonumber. \label{flatrel} \end{eqnarray} Figure \ref{flat} plots the relation for model A3. It is interesting that for all adiabatic jets geometrical shapes described by exponents $a$ in the range $1/3 \ltappeq a \ltappeq 2/3$ are required for flat spectra. Note that for $a_5 < -1$ the $n=1$ term in the series in equation (\ref{series}) dominates for $\tau_{\rm min} \rightarrow 0$. In that case we have for the long jet \begin{equation} \left[ \Gamma \left( a_5, \tau \right)+\frac{1}{a_5} \tau^{a_5} \right]_{\tau_{\rm max}}^{\tau_{\rm min}} \sim \frac{1}{1+a_5} \tau _{\rm min}^{1+a_5}, \end{equation} similar to the entirely optically thin jet. The spectrum of the long jet is then also optically thin, i.e. $F_{\nu} \propto \nu^{\left(1-p \right)/2}$, and a flat spectrum is not possible for physically reasonable values of the exponent $p$. \begin{table} \begin{tabular}{lc} Model parameter & Value\\ \hline $x_0$ & $47$\,AU\\ $r_0$ & $8.9 \times 10^7$\,m\\[1ex] $p$ & $2.5$\\ $B_0$ & $2.4$\,mT\\[1ex] $\kappa _0$ & $3.3\times10^{-7}$\,J$^{1.5}$\,m$^{-3}$\\ $D$ & $2$\,kpc\\[1ex] $l_{\rm min}$ & $4.3\times 10^{-5}$\\ $l_{\rm max}$ & $200$\\[1ex] $\gamma_{\rm max} \left( t_{\rm min} \right)$ & $10^6$\\ $v_{\rm j}$ & $0.97\,c$\\ $\vartheta$ & $40^{\circ}$\\ \hline \end{tabular} \caption{Parameters used to illustrate the spectral properties of the jet models. These model parameters are also used to explain the observational data of Cygnus X-1 in Section \ref{cyg} when using a ballistic jet model with a magnetic field perpendicular to the jet axis, model B1. \label{modpara}} \end{table} Unless the jet is exceedingly short, there will always be a range of frequencies for which $\tau_{\rm min} \rightarrow 0$ and $\tau_{\rm max} \rightarrow \infty$ and the long jet scenario applies. An example is shown in Figure \ref{tauspec} where we plot the integrand in equation (\ref{lflux}) and the optical depth of the jet material as a function of $l$ for a single frequency for the ballistic jet. The contribution to the overall flux of the jet peaks close to $\tau=1$. For the long jet scenario to apply the jet must be long enough so that substantial emission from either side of the peak contributes to the overall flux. Below the frequency range of the long jet the spectrum will follow the optically thick case, $F_{\nu} \propto \nu^{5/2}$, and above this range the optically thin case applies with $F_{\nu} \propto \nu^{(1-p)/2}$. The solid line in Figure \ref{illustration2} illustrates this generic overall shape of the jet spectrum for the ballistic jet model B1. For this Figure and the following we have used the model parameters summarised in Table \ref{modpara}. The observations of the partially self-absorbed jet of Cyg X-1 are well explained by the model for these parameters (see Sections \ref{losses} and \ref{cyg}). For comparison, the solid line in Figure \ref{adillu2} shows the spectrum of the adiabatic jet model A3 with $a=0.46$ for the same set of parameters. The value of $a$ was chosen according to equation (\ref{longslope}) to allow for a flat spectrum at intermediate frequencies. It is interesting to note that the frequencies for which the optical depth of the jet material is unity at $l_{\rm min}$ and $l_{\rm max}$ are located well within the optically thick and thin regimes respectively. The transition from the, in this case, flat spectrum of the long jet occurs closer to $\tau \left( l_{\rm max} \right) \sim 10^{-4}$ and $\tau \left( l_{\rm min} \right) \sim 100$. Obviously, for sufficiently short jets the frequency range over which the long jet case applies may vanish altogether. \section{Including energy losses of the electrons} In the previous Section we did not consider the effect of radiative energy losses of the relativistic electrons on the predicted spectra. The adiabatic jet models include the effect of adiabatic energy losses on the overall energy distribution of the relativistic electrons. However, because we did not impose a high-energy cut-off to this distribution, we did not have to consider the effect of adiabatic losses on such a cut-off. In this Section we introduce a high-energy cut-off at $\gamma _{\rm max}$ and include the effect of adiabatic energy losses on this cut-off. \subsection{Evolution of the high-energy cut-off} \label{cutoff} \subsubsection{Adiabatic and synchrotron losses} Other than adiabatic energy losses, the radiative losses due to synchrotron radiation modify the energy distribution of the relativistic electrons away from a simple power-law, unless $p=2$ \citep[e.g.][]{nk62}. In the following we will make the simplifying assumption that the energy losses only shift the sharp high-energy cut-off to lower energies while not altering the power-law shape or exponent of the power-law distribution. This approximation does not introduce a large error as the deviation from the original power-law is significant only near the cut-off. Also, the expressions for the synchrotron emissivity and absorption coefficient given in equations (\ref{jk}) and (\ref{jk0}) are strictly valid only for power-law energy distribution extending from $\gamma =1$ to $\gamma_{\rm max} \rightarrow \infty$. However, the expressions involved in the derivation of $J_{\nu}$ and $\chi_{\nu}$ decay sufficiently quickly for $\gamma \gtappeq 10$ that the results for finite $\gamma _{\rm max}$ do not deviate greatly from those presented in the previous Section for $\gamma_{\rm max} \rightarrow \infty$ \citep{rl79,ml94}. In the optically thin regime the evolution of the Lorentz factor of a given relativistic electron in the rest frame of the jet material is described by \citep[e.g.][]{ml94} \begin{equation} \dot{\gamma} = -\frac{4}{3} \frac{\sigma _{\rm T} u_0}{m_{\rm e} c} \left( \frac{t}{t_0} \right)^{-2 a_2} \gamma^2 - \frac{2a_1}{3t} \gamma, \label{evol} \end{equation} where the first term on the right describes the energy losses due to synchrotron radiation and the second term reproduces equation (\ref{adiabat}) for the adiabatic losses where we substituted for $\Delta V$. Because of our assumption of a constant bulk velocity for the jet material along the $x$-axis, $v_{\rm j}$, we can express the dimensionless coordinate $l=x/x_0$ also as a time variable, i.e. $l=\gamma _{\rm j}v_{\rm j}t/x_0$. Because of time dilation, we have to include $\gamma_{\rm j}$. Thus $t=l x_0 / \left( v_{\rm j} \gamma _{\rm j} \right)$ and $t_0=x_0/ \left( v_{\rm j} \gamma _{\rm j} \right)$. $\sigma _{\rm T}$ is the Thomson cross section and $u_0=B_0^2/\left(2 \mu_0 \right)$ is the energy density of the magnetic field at $x_0$. The solution of equation (\ref{evol}) is found as \begin{equation} \gamma \left( t \right) = \frac{\gamma \left( t_{\rm min}\right) t^{-2 a_1 /3}}{t_{\rm min}^{-2 a_1/3} + \frac{4 \sigma _{\rm T} u_0}{3 a_6 m_{\rm e} c} t_0^{2a_2} \gamma \left( t_{\rm min}\right) \left( t^{a_6}-t_{\rm min}^{a_6} \right)}, \label{gamadiabat} \end{equation} with $a_6 = 1-2a_2-2a_1/3$ and $t_{\rm min} = l_{\rm min} x_0 / \left( v_{\rm j} \gamma _{\rm j} \right)$. Electrons which were injected into the jet at $t_{\rm min}$ or, equivalently, $l_{\rm min}$ with a Lorentz factor $\gamma \left( t_{\rm min}\right)$, have a Lorentz factor $\gamma \left( t \right)$ at $t$ or, equivalently, $l$. For the ballistic jet the adiabatic, second term on the right of equation (\ref{evol}) vanishes and we have instead \begin{equation} \gamma \left( t \right) = \frac{\gamma \left( t_{\rm min}\right)}{1+\frac{4 \sigma _{\rm T} u_0}{3 \left(1-2a_2 \right) m_{\rm e} c} t_0^{2a_2} \gamma \left( t_{\rm min} \right)\left( t^{1-2a_2}-t_{\rm min}^{1-2a_2} \right)}. \label{gamrad} \end{equation} Note that the exponents $a_6$ and $1-2a_2$ are usually negative. This implies in the case of the ballistic jet models (no adiabatic losses) that the Lorentz factors of electrons do not necessarily decrease forever, but converge to a finite value for $t\gg t_{\rm min}$. The somewhat surprising result simply reflects the fact that the synchrotron losses rapidly decline in the decreasing magnetic field of the expanding jet. For the adiabatic case the adiabatic losses continue at all times and so $\gamma \propto t^{-2a_1 /3} \propto l^{-2 a_1 / 3}$ for $t\gg t_{\rm min}$. For optically thin conditions the evolution of the high-energy cut-off $\gamma _{\rm max}$ also obeys equations (\ref{gamadiabat}) and (\ref{gamrad}). Below we will refer to the high-energy cut-off in the optically thin regime as $\gamma _{\rm thin}$. However, for large parts of the spectrum the jet is optically thick. Electrons with a given Lorentz factor $\gamma$ emit most of their radiation at the critical frequency $\nu \sim \nu _{\rm g} \gamma^2$, where the gyro-frequency is defined as $\nu_{\rm g} = e B / \left(2 \pi m_{\rm e}\right)$. An electron emitting at a critical frequency for which the jet is optically thick gains energy through synchrotron self-absorption. Ideally we would include an energy gain term for the self-absorption effect into equation (\ref{evol}) and then derive the electron evolution as before. While this approach leads to analytic solutions when only considering the systematic energy gain of electrons of a single energy \citep{mr67}, it is not applicable in most cases because the stochastic energy gain for power-law energy distributions is comparable to the systematic term. In this case, only numerical solutions are possible, because the stochastic term depends on the entire energy distribution \citep{rm69}. The full numerical treatment of synchrotron losses and gains in the optically thick regime is beyond the scope of this paper. However, electrons radiating mainly at frequencies for which the jet is optically thick, do on average not lose or gain energy due to radiative effects, even if they are relatively close to the surface of the jet \citep{rm69}. Thus the high-energy cut-off in the optically thick regime, $\gamma _{\rm thick}$, is given by the requirement that $\tau \left( \gamma _{\rm thick} \right) \sim 1$. The electrons at this cut-off emit mainly at a frequency $\nu_{\rm thick} = \nu_{\rm g} \gamma _{\rm thick}^2$ and so we find from equation (\ref{depth}) \begin{equation} \gamma _{\rm thick} = \left[ \frac{4 \pi m_{\rm e}}{3 e B_0} \left( \chi_0 r_0 \right)^{2 / \left(p+4 \right)} l^{a_7} \right]^{1/2}, \label{gamthick} \end{equation} with \begin{equation} a_7 = \left(a_1 -a_3 - a_2 \frac{p+2}{2}\right) \frac{2}{p+4}+a_2. \end{equation} For the ballistic jet with perpendicular magnetic field, model B1, $a_7 =0$ and so $\gamma _{\rm thick}$ is constant along the entire length of the jet. In other words, electrons with Lorentz factors equal or below $\gamma _{\rm thick}$ never loose their energy to radiation unless they are very close to the jet surface. The existence of a constant high-energy cut-off is required for a flat spectrum from the jet. While \citet{bk79} invoked an unknown re-acceleration mechanism to ensure $\gamma _{\rm max} ={\rm constant}$, we have shown here that such a process is unnecessary because of the energy gains associated with synchrotron self-absorption. For a parallel magnetic field in the ballistic jet, model B2, we have $a_7= 2/ \left(p+4 \right)$. The Lorentz factor of relativistic electrons for which the jet is optically thick is {\em increasing\/} for increasing $l$ in this model. Therefore, if $l_{\rm thick}$ is the position along the jet where $\gamma _{\rm thin} = \gamma _{\rm thick}$, we have $\gamma_{\rm max} =\gamma_{\rm thick}$ at this position and $\gamma _{\rm max} = {\rm constant}$ for all $l>l_{\rm thick}$. In the adiabatic cases we find \begin{eqnarray} {\rm Model\ A1:} && a_7 = a \frac{4 \left( 1 - p \right)}{3 \left( p + 4 \right)}\nonumber\\ {\rm Model\ A2:} && a_7 = a \frac{2 \left( 5 - 2p \right)}{3 \left( p + 4 \right)}\\ {\rm Model\ A3:} && a_7 = a \frac{2 \left( 3 -2 p \right)}{3 \left(p+4 \right)}.\nonumber \end{eqnarray} Even for optically thick conditions the electron energy distribution does not deviate greatly from the original power-law with a high-energy cut-off $\gamma _{\rm max}$ for exponents $2\le p \le 3$ \citep{rm69}. Thus, for each position $l$ along the jet we can now determine $\gamma_{\rm max}$ and thereby the entire electron energy distribution. For small $l$ the cut-off is given by $\gamma _{\rm max} =\gamma _{\rm thin}$. Further down the jet $\gamma _{\rm thin}$ will first become equal to and then fall below $\gamma _{\rm thick}$ and for the ballistic jet models $\gamma _{\rm max} = \gamma _{\rm thick}$ afterwards. In the adiabatic jet models two competing effects can diminish $\gamma _{\rm max}$ further after passing through the point at which $\gamma _{\rm thick} = \gamma _{\rm thin}$. In most cases equation (\ref{gamthick}) implies a further reduction of $\gamma _{\rm thick}$ for increasing $l$. This means that $\gamma _{\rm max}$ would also decrease. At the same time, adiabatic losses lead to $\gamma _{\rm max} = \gamma _{\rm thick} \left( l / l_{\rm thick} \right) ^{-2 a_1 / 3}$. It is straightforward to show that for all adiabatic models the adiabatic losses of $\gamma _{\rm max}$ are the dominant effect. By again making the assumption that all electrons only emit at their critical frequency, we can now define for a given frequency $\nu$ a maximum distance $l'_{\rm max}$ along the jet axis where the jet material is still contributing to the overall emission, \begin{equation} l'_{\rm max} = \left(\frac{2 \pi m_{\rm e} \nu}{e B_0 \gamma _{\rm max}^2} \right)^{-1/a_2}. \label{ldmax} \end{equation} Clearly, as long as $l'_{\rm max}$ is larger than the physical extent of the jet, $l_{\rm max}$, the jet spectrum is not affected by energy losses of the relativistic electrons at frequency $\nu$ and we can use the results of the previous Section. For $l'_{\rm max} < l_{\rm max}$ we have to take into account energy losses of the electrons by using $l'_{\rm max}$ instead of $l_{\rm max}$ in the calculation of $\tau_{\rm min}$. \subsubsection{Losses due to Compton scattering} In the optical thick parts of the jet Compton scattering of the synchrotron photons off the relativistic electrons may become important. We do not include energy losses of the relativistic electrons due to Compton scattering in the jet in our calculations as these would require a full treatment of radiative transfer. However, it is obviously necessary to test whether these losses are important when applying the model to observational data and so we give the necessary expressions below. The relevant limit for the energy density of the synchrotron photon field is most conveniently expressed in terms of the brightness temperature \citep[e.g.][]{rl79}, \begin{equation} T_{\rm b} = \frac{c^2 I_{\nu}}{2 \delta_{\mp}^3 k_{\rm B} \nu^2}, \end{equation} where $k_{\rm B}$ is the Boltzman constant and $\nu$ is the emitted frequency rather than the observing frequency. For $T_{\rm b} \ltappeq 10^{12}$\,K Compton losses are not important compared to the energy losses due to synchrotron radiation. The maximum brightness temperature for a given frequency is reached at the position along the jet where the optical depth of the jet material roughly equals unity for photons of this frequency. Hence in our model we have \begin{equation} T_{\rm b, max} = \frac{c^2 J_0}{8 \pi k_{\rm B} \chi _0} \left( \chi _0 r_0 \right)^{1/ \left( p+4 \right)} \left( 1 - e^{-1} \right) l^{ \left( a_1 + a_2 - a_3 \right) / \left( p + 4 \right)}. \label{bright} \end{equation} The maximum brightness temperature is either constant or only a weak function of $l$ in all our models. Also, using the dependencies of $J_0$ and $\chi _0$ given in equation (\ref{jk0}) we find that $T_{\rm b, max}$ depends only weakly on the other model parameters $B_0$, $\kappa _0$ and $r_0$. In some jets relativistic induced Compton scattering may be more important than direct Compton scattering discussed above. Induced Compton scattering causes significant energy losses for the relativistic electrons if \citep{sk94} \begin{equation} \frac{k_{\rm B} T_{\rm b}}{m_{\rm e} c^2} \tau _{\rm T} \ge 1, \label{crit} \end{equation} where $\tau_{\rm T}$ is the Thomson depth of the jet material, \begin{equation} \tau_{\rm T} = n_{\rm e} \sigma_{\rm T} r. \end{equation} Here, $n_{\rm e}$ is the number density of electrons, $\sigma _{\rm T}$ is the Thomson cross-section and we have again assumed that the average path length a photon travels through the jet material is equal to the jet radius, $r$. For our power-law energy distribution of the electrons an upper limit for the electron density is given by $n_{\rm e} \le \kappa \left( m_{\rm e} c^2 \right)^{1-p}$. The Thomson depth of the jet materials in our models is then limited by \begin{equation} \tau_{\rm T} \le \left(m _{\rm e} c^2 \right)^{1-p} \sigma _{\rm T} \kappa _0 r_0 l^{a_1-a_3}. \label{thom} \end{equation} Equations (\ref{bright}) and (\ref{thom}) can be used to ensure that the models are applicable to a given observational data set, i.e. that they do not suffer from Compton losses which are not included in the models. \subsection{Spectra with energy losses and a high-energy cut-off} \label{losses} \subsubsection{Iterative construction of model spectra} For a given set of model parameters we can construct a model spectrum. In practice this will involve the determination of $\tau_{\rm min}$ and $\tau _{\rm max}$ from equation (\ref{depth}) to be substituted into equation (\ref{sol}). For a given frequency $\nu$, we set $l= l_{\rm min}$ and calculate $\tau_{\rm max}$. Determining $\tau _{\rm min}$ is more involved as it requires the calculation of $l'_{\rm max}$. This calculation involves an implicit equation and so cannot be done analytically. Here we describe one possible iterative procedure for the determination of $l'_{\rm max}$. The first step is to choose a trial distance $l$ such that $l_{\rm min} \le l \le l_{\rm max}$. The strength of the magnetic field in the jet material at $l$ is $B \left( l \right) = B_0 l^{-a_2}$. The maximum frequency at which jet material located at $l$ is still contributing to the emission is given by $\nu _{\rm max} \left( l \right) = \nu_{\rm g} \left( l \right) \gamma _{\rm max} \left( l \right)$. For the next iteration we need to compare $\nu_{\rm max} \left( l \right)$ with $\nu$. Therefore we must next derive $\gamma _{\rm max} \left( l \right)$, the high-energy cut-off of the electron energy distribution at $l$. From equations (\ref{gamadiabat}, adiabatic jet models) or (\ref{gamrad}, ballistic jet models) we can determine $\gamma _{\rm thin} \left( l \right)$. We calculate $\gamma _{\rm thick}$ from equation (\ref{gamthick}). If $\gamma_{\rm thin} \left( l \right) > \gamma _{\rm thick}$, then $\gamma _{\rm max} \left( l \right)= \gamma _{\rm thin} \left( l \right)$. Otherwise, for the ballistic jet models $\gamma _{\rm max} \left( l \right) = \gamma _{\rm thick}$. For the adiabatic jet models $\gamma _{\rm max} \left( l \right) = \gamma _{\rm thick} \left( l / l_{\rm thick} \right)^{-2a_1 /3}$. The required distance $l_{\rm thick}$ must be found from the implicit equation resulting from setting $\gamma \left( t \right) = \gamma _{\rm thick}$ in equation (\ref{gamadiabat}). Finally, if $\nu_{\rm max} \left( l \right) > \nu$, then the trial distance in the next iteration should be larger than the current one. In the case of $\nu_{\rm max} \left( l \right) < \nu$, the trial distance should be decreased. The iterations can be stopped when $\nu_{\rm max} \left( l \right) \sim \nu$ within the required accuracy and at that point we can set $l'_{\rm max} =l$ and then proceed to calculate $\tau_{\rm min}$. \subsubsection{Example spectra} The model parameters in Table \ref{modpara} were chosen to explain the observations of the jet in Cygnus X-1 with the ballistic jet model B1 including radiative energy losses (see Section \ref{cyg}). This does not imply that energy losses will always be important in all jets and at all frequencies. The spectrum of the ballistic and adiabatic jets with energy losses of the electrons are shown in Figures \ref{illustration2} for model B1 and \ref{adillu2} for model A3. When energy losses of the electrons are taken into account, then we cannot in general expect that $\tau_{\rm min} \ll 1$ for a given frequency. Therefore the spectrum of the jet will not necessarily be that of the long jet described by equation (\ref{longslope}). In fact, in most cases $\tau_{\rm min}$ will considerably exceed unity. If $\tau_{\rm max} \rightarrow \infty$, then equation (\ref{largetau}) applies and we find \begin{equation} F_{\nu} \sim \delta_{\mp}^2 \frac{x_0 r_0 J_0}{a_4 a_5 D^2 \chi_0} \nu^{5/2} \left( \frac{\tau_{\rm min}}{\tau_0} \right)^{a_5}. \label{reduced} \end{equation} For the ballistic jet models we have argued in the previous Section that $\gamma _{\rm max} ={\rm constant}$ for $l \ge l_{\rm thick}$. For rapid radiative energy losses of the electrons at $l$ in the range $l_{\rm min} < l < l_{\rm thick}$, the distance $l_{\rm thick}$ itself will not depend strongly on the observing frequency. From equation (\ref{ldmax}) we then find $l'_{\rm max} \propto \nu ^{-1/a_2}$ and substituting into equation (\ref{depth}) we get \begin{equation} \tau_{\rm min} \propto \nu^{\left( a_3 -a_1-a_2 \right) / a_2}. \label{taumin} \end{equation} Note that the exponent does not depend on the slope of the electron energy distribution, $p$, nor on the geometrical shape of the jet described by $a$ (see Table \ref{expo}). Finally, from equation (\ref{reduced}) we obtain the slope of the spectrum as \begin{equation} F_{\nu} \propto \nu^{\left( 4a_2-2a_1-2 \right) / \left( 2 a_2 \right)}. \end{equation} For the ballistic jet with perpendicular magnetic field, model B1, we have $F_{\nu} ={\rm constant}$ as in the case without energy losses of the electrons, which is confirmed by Figure \ref{illustration2}. For a parallel field structure, model B2, the spectrum would follow $F_{\nu} \propto \nu$. Model B2 is incompatible with a flat spectrum. The slope in the optically thin part of the spectrum in Figure \ref{illustration2} is steeper compared to the case of no energy losses because of the effect of the high-energy cut-off in the electron energy spectrum. In the case of the adiabatic jet models $\gamma _{\rm max} \propto l^{-2 a_1 /3}$ and so $l'_{\rm max} \propto \nu^{-3 / \left( 4 a_1 + 3 a_2 \right)}$. Again substituting into equation (\ref{depth}) yields \begin{equation} \tau_{\rm min} \propto \nu^{- \left(7 a_1 + 3 a_2 \right) / \left( 4 a_1 + 3 a_2 \right)}, \end{equation} where we also used $a_3 = \left( 4 + 2p \right) a_1 /3$ as appropriate for the adiabatic jet models. Again the exponent of this expression does not depend on $p$ or $a_1$. The shape of the spectrum is now \begin{equation} F_{\nu} \propto \nu^{\left( 7 a_1 + 6 a_2 -3 \right) / \left(4 a_1 +3 a_2 \right)}. \end{equation} The exponent of the power-law spectrum predicted by model A3 for $a_1 = 0.46$ is then 1.06 which is confirmed by the slope of the spectrum in Figure \ref{adillu2} below about $10^{12}$\,Hz. The emission at high frequencies comes from the innermost parts of the jet close to $l_{\rm min}$. Radiative losses had not enough time there to completely change the electron energy distribution. This explains the peak in the spectrum. At the highest frequencies only optically thin parts of the jet contribute to the overall emission and lead to a negative power-law similar to the case without energy losses. Note however that the slope of this power-law is somewhat steeper due to the decreasing high-energy cut-off. The adiabatic jet models are all consistent with a flat spectrum provided the shape of the jet described by $a_1$ takes a suitable value (Model A1: $a_1 = 3/13$, A2: $a_1 = 3/19$ and A3: $a_1 = 1/5$). In all three adiabatic models the jet needs to be strongly confined, i.e. $a_1 < 1/4$, to achieve a flat emission spectrum. The slope of the spectrum is quite sensitive to the value of $a_1$. For example, a change of $a_1$ in model A3 from $1/5$ to $1/3$ results in a change of the power law exponent of the spectrum from zero to 0.75. For the following discussion we note that all the spectral slopes calculated above, with the only exception of the purely optically thin case, are all independent of the exponent $p$ of the power-law describing the energy distribution of the electrons. Therefore we can readily apply the model to observational data even if the optically thin part of the spectrum is not observed and thus we do not know the value of $p$. \section{Application to observations} \label{obs} The jet emission models depend on a number of parameters. Some of these parameters can be constrained by applying general considerations and others may be inferred from applying the model predictions to observational data with a view to determining the physical conditions within the jet. Here we discuss all of the relevant parameters in turn and demonstrate below how observations of the jet in Cygnus X-1 may be used to infer the properties of this object. The scale height $x_0$ can always be chosen arbitrarily to provide a convenient location along the jet axis at which to define the exact values of other quantities. In many cases we will be mainly interested in that part of the jet spectrum which is strongly affected by absorption. As we have seen in the previous Section, we then do not need to know the exponent of the power-law energy distribution of the electrons, $p$, as it does not influence the slope of the predicted spectrum. However, estimates for other quantities derived from the model, for example the strength of the magnetic field, depend weakly on $p$. The distance of the jet, $D$, the bulk velocity of the jet material, $v_{\rm j}$, and the viewing angle of the jet axis to our line of sight, $\vartheta$, cannot normally be determined by the model itself and need to be measured by other means. The parameters describing the geometrical shape and size of the jet, $r_0$, $l_{\rm min}$ and $l_{\rm max}$, could in principle be determined from observations. However, $l_{\rm min}$ is probably too small to be resolvable even with a large improvement on current resolution limits. Currently only upper limits exist for the radius of jets in Galactic X-ray binaries (Miller-Jones et al., in preparation) while for AGN jets $r_0$ is sometimes resolved \citep[e.g.][]{jbl99}. The maximum extent of a jet at a given observing frequency is sometimes measured and an example is provided by the observations of the jet in Cygnus X-1 of \citet{ssf01} which we use in the following Section. It should be borne in mind that at one observing frequency we can always only measure $l'_{\rm max}$ given by equation (\ref{ldmax}) rather than the physical extent of the jet flow $l_{\rm max}$. However, $l_{\rm max}$ only determines the low frequency cut-off of the spectrum, but is not important for the model otherwise. In the case of the adiabatic jet models resolved observations can, in principle, also determine the shape of the jet as described by the parameter $a$. However, in practice it is easier to infer the value of $a$ from the slope of the observed self-absorbed spectrum as this is a strong function of $a$. Finally, the normalization of the electron energy distribution, $\kappa _0$, and the strength of the magnetic field, $B_0$, cannot be determined directly from observations, but must be inferred from the model. We can reduce the number of free parameters by assuming that the energy densities of the magnetic field and of the relativistic electrons are initially in equipartition. In this case, \begin{equation} u_0 = \frac{B_0^2}{2 \mu _0} = \int_{E_{\rm min}}^{E_{\rm max}} \kappa _0 E^{1-p} \, {\rm d}E. \end{equation} For an energy distribution extending over all physically meaningful Lorentz factors ($1 \le \gamma \le \infty$) we then have \begin{equation} \kappa _0 \sim \left( p-2 \right) \frac{B_0^2}{2 \mu _0} \left( m_{\rm e} c^2 \right)^{p-2}. \end{equation} With these considerations we can now apply the model to observations and determine relevant parameters for the observed jet. \subsection{Application to Cygnus X-1} \label{cyg} In the following we apply the model to the jet observed in the X-ray binary Cygnus X-1. \citet{ssf01} report a resolved jet extending to about 15\,mas from the position of the X-ray binary system at an observing frequency of 8.4\,GHz. We set the bulk velocity of the jet material to 0.97\,$c$ and the viewing angle to our line of sight to $\vartheta = 40^{\circ}$ \citep{ssf01}. For a distance of 2\,kpc \citep{gzp99} the observed, projected jet length then corresponds to a real jet length of roughly 47\,AU. For convenience we set $x_0$ equal to this value and so $l'_{\rm max} = 1$ at 8.4\,GHz. Only one jet is observed and it is therefore reasonable to assume that this is the approaching jet. Note that the bulk velocity of the jet and the viewing angle imply $\delta _- \sim 1$ and so $\nu \sim \nu_{\rm ob}$. The total flux density at the same frequency is 13\,mJy. In a map at 15\,GHz from an earlier observing epoch, the jet may also be marginally extended along the same axis \citep{ssg98,ssf01}. The extension is 2\,mas or less which corresponds to $l'_{\rm max} \left( 15\,{\rm GHz} \right) \le 0.13$. In the following we will mainly concentrate on the observational data at 8.4\,GHz. There are no simultaneous observations at any frequency other than 8.4\,GHz and so we cannot be certain what the spectral slope of the jet emission was. However, during the low/hard X-ray state the source usually shows a flat spectrum extending up to at least 220\,GHz \citep{fpdtb99}. We assume here that the spectral slope at the time the radio jet at 8.4\,GHz was observed was zero. The extent of the flat spectrum to high frequencies is also not known. However, for GX 339-4 the flat jet spectrum is observed to near-IR wavelengths \citep{cf02} while for XTE J1118+480 it may extend to the near-UV \citep{hmh00}. For the purpose of illustrating the model, we assume here that the flat spectrum extends to the near-IR of wavelengths of about 1\,$\mu$m. The flat spectrum always arises from regions in the jet which are at least partially self-absorbed. Thus the exact value of $p$ is not very important and we set $p=2.5$. \subsubsection{Ballistic jet models} We have seen above that the ballistic jet with a parallel magnetic field configuration, model B2, is inconsistent with a flat spectrum. In this Section we therefore concentrate on model B1, a ballistic jet with a magnetic field perpendicular to the jet axis. With the assumptions made above equation (\ref{reduced}) reduces to \begin{equation} F_{\nu} \sim 5.9 \times 10^{-7} \delta _{\mp}^2 x_0 r_0 D^{-2}B_0^{-1/2}\nu^{5/2} {l'_{\rm max}}^{5/2} \, {\rm mJy} \end{equation} for the ballistic jet model B1. Here and in the following, all quantities are measured in SI units unless indicated otherwise. Substituting the measurements discussed above we get an expression for $r_0$ as a function of $B_0$. A second equation relating the same quantities can be found from substituting $\gamma _{\rm thick}$ from equation (\ref{gamthick}) for $\gamma_{\rm max}$ in equation (\ref{ldmax}) resulting in \begin{equation} l'_{\rm max} = 2.0 \times 10^{11} \nu^{-1} B_0^{9/13} \kappa _0^{4/13} r_0^{4/13} = 1, \end{equation} where we again made use of the observed quantities. For initial equipartition we can eliminate $\kappa _0$ and solve for $B_0$. We can then calculate all other model parameters and they are summarised in Table \ref{modpara}. The model spectrum is plotted as the dashed line in Figure \ref{illustration2}. The maximum brightness temperature of the jet emission does not depend on $l$ for model B1 and from equation (\ref{bright}) we find $T_{\rm b, max} =1.1\times10^{10}$\,K. We can also calculate at what distance $l$ along the jet the jet material is dense enough so that relativistic induced Compton scattering becomes important. From equation (\ref{crit}) we find that this loss process would only play a role for $l < 1.5\times10^{-7}$ which is well inside $l_{\rm min}$. We therefore conclude that our model can be applied to the jet of Cygnus X-1. The jet radius $r_0$ is very small. The ballistic jet model is conical and therefore we can define a jet opening angle as $\theta = 2 r_0 / x_0 = 5"$ which is much smaller than the observational upper limit of $2^{\circ}$ \citep{ssf01}. It is of course possible to assume a larger radius for the jet by dropping the assumption of equipartition. However, setting the jet radius equal to the observational upper limit would imply that the jet material is out of equipartition by several orders of magnitude. The physical limits of the jet flow, $l_{\rm min}$ and $l_{\rm max}$, can be derived from equation (\ref{depth}) by setting $\tau \left( l_{\rm min} \right) = \tau \left( l_{\rm max} \right) = 1$ for those frequencies at which the flat spectrum is required to break to the optically thick or thin regime, respectively. In our example of Cygnus X-1 we used a lower break of just under 100\,MHz and an upper break of $3\times 10^{14}$\,Hz corresponding to a wavelength of 1\,$\mu$m. The resulting limit $l_{\rm min}=4.3\times10^{-5}$, associated with the break to optical thin conditions, corresponds to a distance of about 6000 Schwarzschild radii from the central black hole with a mass of 10\,M$_{\odot}$ \citep{gzp99}. Since for the ballistic jet model B1 $l_{\rm min}$ decreases linearly with the break frequency, the lower physical limit decreases to about 500 Schwarzschild radii if the flat spectrum extends to the near-UV. Clearly the determination of the high frequency break of the flat part of the spectrum can put interesting constraints on the distance from the central black holes at which jets become ballistic. Similarly the low frequency break constrains the overall extent of the ballistic jet flow. The observable extent of the jet flow depends linearly on the observing frequency. Since $l'_{\rm max} \left(8.4\,{\rm GHz} \right) = 1$, we would expect that $l'_{\rm max} \left( 15\,{\rm GHz} \right) \sim 0.6$. This prediction significantly exceeds the tentative extension of the Cygnus X-1 jet of $l'_{\rm max} \left( 15\,{\rm GHz} \right) \le 0.13$ reported in \citet{ssg98}. However, the observations at 15\,GHz were not simultaneous with those at 8.4\,GHz. The strength of the magnetic field, $B_0$, at $x_0$ implies a field strength of around 50\,T at the distance $l_{\rm min}$. If $l_{\rm min}$ is reduced because the flat spectrum extends to the near-UV, then the magnetic field in the jet has a strength of 500\,T about 500 Schwarzschild radii away from the black hole. Again this demonstrates that our model can provide useful constraints on the conditions at the very base of the observed jets despite them not being spatially resolved. The strength of the magnetic field, $B_0$, and the constant in the expression for the density of relativistic electrons, $\kappa _0$, can be used to estimate the power of the jet. We find that the energy transport rate associated with the magnetic field and the relativistic particles alone is $5.2\times10^{25}$\,W. A further $1.1 \times 10^{25}$\,W is added by the kinetic energy of the electrons. Not surprisingly these numbers are comparable to the estimates using the \citet{bk79} model \citep{fpdtb99}. If there is a cold proton for every relativistic electron in the jet, then its energy transport rate in terms of kinetic energy is $1.1\times 10^{28}$\,W. This power is about one order of magnitude below the time-averaged energy transport rate of the Cygnus X-1 jet, recently estimated from the observed interaction of the jet with the surrounding ISM \citep{gfk05}. The flux density of the flat jet spectrum does not vary significantly between available observations. Unless the jet power varies considerably over timescales longer than the timespan since the first available radio observations of Cygnus X-1, then our estimate strongly suggests the presence of a proton-electron plasma in the jet. \subsubsection{Adiabatic jet models} All three adiabatic jet models are consistent with a partially flat emission spectrum. In fact, if we choose the geometrical parameter $a_1$ appropriate for a flat spectrum for each model, then the differences between the adiabatic models become small as the differences in the behaviour of the magnetic field as a function of $l$ between them is compensated for by the different degrees of confinement of the jet. Therefore and to simplify the discussion below, we focus on the adiabatic model A3 with an isotropic magnetic field. None of the conclusions change greatly for models A1 and A2. For the adiabatic jet model A3 our assumptions with equation (\ref{reduced}) lead to \begin{equation} F_{\nu} \sim 8.8 \times 10^{-7} \delta _{\mp}^2 x_0 r_0 D^{-2}B_0^{-1/2}\nu^{5/2} {l'_{\rm max}}^{5/3} \, {\rm mJy}, \end{equation} where we have set $a=1/5$ to allow for a flat section in the spectrum. The calculation of $\gamma _{\rm max}$ now requires $l_{\rm thick}$, which must be determined from the implicit equation (\ref{gamadiabat}) for a given magnetic field strength, $B_0$. Figure \ref{lmaxest} demonstrates that the allowed minimum for $l'_{\rm max}$ exceeds unity for the assumption of initial equipartition. Since observations require $l'_{\rm max} =1$ at $8.4$\,GHz, the adiabatic models are incompatible with observations unless we drop the requirement of equipartition. We now introduce a reduction factor $f$ such that the initial energy density of the relativistic electrons is a fraction $f$ of the initial energy density of the magnetic field. An example of the results for $l'_{\rm max}$ for $f=10^{-6}$ is shown in Figure \ref{lmaxest}. There are now two possible solutions for the strength of the magnetic field. However, we also require that the lower size limit of the jet, $l_{\rm min}$, accommodates a break of the flat spectrum to the optically thin regime in the near-IR. This lower limit cannot lie inside the last stable orbit of the central black hole. Thus we obtain another constraint on the solution because $l_{\rm min} \ge 3 R_{\rm S}$, where $R_{\rm S} = 3 \times 10^4$\,m is the Schwarzschild radius of a 10\,M$_{\odot}$ black hole. The additional constraint is also plotted in Figure \ref{lmaxest}. Only one of the two possible solutions for $f=10^{-6}$ is consistent with this constraint. It is also interesting to note that the solution requiring initial equipartition is also inconsistent with a physical meaningful value of $l_{\rm min}$. Using the remaining solution for $f=10^{-6}$ as an example, we can compute all remaining model parameters which are summarised in Table \ref{failedtab}. Figure \ref{failedspec} shows the resulting spectrum. It is flat over a wide range of frequencies with the required flux level. The dip at around $10^{14}$\,Hz is a result of the diminishing optical depths of the jet to radiation emitted by electrons with the limiting Lorentz factor $\gamma _{\rm max} = \gamma_{\rm thick} \left( l / l_{\rm thick} \right)^{-2 a_1/3}$. For higher frequencies those parts of the jet close to $l_{\rm min}$ which still contain electrons with Lorentz factors in excess of $\gamma _{\rm thick}$ contribute to the emission and cause the peak. The position of the peak is located at the frequency with optical depth $\tau_{\rm max} \sim 1$ at $l_{\rm min}$. Clearly the spectrum is not flat from radio to near-IR frequencies because of the emission peak. A flat spectrum extending over the entire radio to near-IR range could be achieved by moving the peak to higher frequencies. However, an appropriate adjustment of the model parameters would also tighten the constraints on the reduction factor $f$ and thereby on $B_0$. \begin{table} \begin{tabular}{lc} Model parameter & Value\\ \hline $x_0$ & $47$\,AU\\ $r_0$ & $1.3 \times 10^9$\,m\\[1ex] $p$ & $2.5$\\ $B_0$ & $1.2$\,T\\[1ex] $\kappa _0$ & $7.6\times10^{-8}$\,J$^{1.5}$\,m$^{-3}$\\ $D$ & $2$\,kpc\\[1ex] $l_{\rm min}$ & $2.0\times 10^{-8}$\\ $l_{\rm max}$ & $200$\\[1ex] $\gamma_{\rm max} \left( t_{\rm min} \right)$ & $10^6$\\ $v_{\rm j}$ & $0.97\,c$\\ $\vartheta$ & $40^{\circ}$\\ \hline \end{tabular} \caption{Parameters for the adiabatic jet model A3 without initial equipartition. The model spectrum is shown in Figure \ref{failedspec}.\label{failedtab}} \end{table} The maximum brightness temperature of the adiabatic model used here arises at $l_{\rm min}$ and is with $7.3\times 10^9$\,K well below the limit for efficient Compton scattering. The distance at which relativistic induced Compton scattering would become important in the adiabatic jet is $l=1.7\times 10^{-16}$. Again we are justified to neglect energy losses due to Compton scattering. The jet radius, $r_0$, for the adiabatic jet is larger than for the ballistic jet. Formally, we cannot define a jet opening angle for adiabatic jets discussed here, because their shape is not conical. However, if a conical shape was assumed, then the opening angle inferred from the radius at $x_0$ would be 1.3'. In our example, the lower limit of the physical extent of the jet, $l_{\rm min}$, corresponds to 4.7\,$R_{\rm S}$. While this is close to the theoretical limit and gravitational redshift would affect the spectrum, a further decreased value of the reduction factor would increase this limit at the expense of requiring an even stronger magnetic field. For adiabatic jets the constraints on the nature of the jet flow extend to even closer distances from the central black hole than in the ballistic case. For the adiabatic jet model A3 the observable extent of the jet flow is proportional to $\nu^{-3 / \left( 8 a \right)}$. In our example this relation implies that $l'_{\rm max} \left( 15\,{\rm GHz} \right) \sim 0.3$, which is still larger than the observed value of \citet{ssg98}, but closer than the prediction of the ballistic model. However, as mentioned above, the observations are not simultaneous and that may explain the discrepancy in both cases. The strength of the magnetic field of 1.2\,T is high at $x_0$. However, because of the much more collimated geometry of the jet, the field strength only increases to 140\,T at $l_{\rm min}$. Nevertheless, the energy transport rate of the jet due to the magnetic field is very large with $10^{34}$\,W determined at $l_{\rm min}$. Due to the small value for the reduction factor $f$, the contribution of the relativistic electrons to any energetic considerations is negligible, even if the kinetic energy of possibly associated protons is taken into account. The derived jet power exceeds by far all previous estimates and is inconsistent with the time-averaged jet power \citep{gfk05}, unless the jet flow is suppressed for long periods on very long timescales. Note also, that this jet power corresponds to one hundred times the Eddington limiting luminosity of a 10\,M$_{\odot}$ black hole. The large jet power is caused by the significant reduction in the radiative efficiency of the synchrotron process well away from equipartition conditions. \section{Conclusions} \label{conc} We construct a model for the synchrotron emission of partially self-absorbed jets. The model does not invoke a re-acceleration process for the relativistic electrons. All electrons are accelerated only once at the lower physical limit of the jet, $l_{\rm min}$. It is not necessary to postulate an unknown re-acceleration mechanism as was suggested in previous work \citep{bk79}, because synchrotron self-absorption counteracts excessive energy losses. Two classes of models are considered. The ballistic jets have a conical geometry and their contents do not suffer adiabatic energy losses. This situation may arise when jets are initially highly overpressured with respect to their environments. They expand unimpeded and random thermal energy is converted into bulk kinetic energy, but not dissipated to any external medium. At the end of this very rapid expansion the mean free path of the jet material exceeds the physical dimensions of the jet itself and follows ballistic trajectories. The other class of models considered here are adiabatic jets confined by either the pressure of the gas surrounding them, surface magnetic fields or both. Their geometrical shape is dictated by the details of the confinement mechanism which is not the subject of this paper. The jet material dissipates energy to the external gas during its adiabatic expansion. Both classes of models can predict flat emission spectra if energy losses of individual electrons are neglected. The ballistic jet with a magnetic field perpendicular to the jet axis produces a flat spectrum without further assumptions. The adiabatic jet models require a specific jet geometry to allow for flat emission spectra. Both model classes can predict flat emission spectra, even when taking energy losses of the electrons and the magnetic field into account. Synchrotron self-absorption prevents radiative energy losses below a critical Lorentz factor $\gamma _{\rm thick}$. In the ballistic case $\gamma _{\rm thick}$ is constant along the jet flow and because adiabatic energy losses are absent, the energy distribution of the relativistic electrons remains stationary. The emission properties of the jet in this case are essentially identical to the model of \citet{bk79}. For the adiabatic jet the with a perpendicular magnetic field electrons with Lorentz factors at and below $\gamma _{\rm thick}$ continue to lose energy because of the sideways expansion of the jet. However, for a geometrical shape given by $r \propto x^{1/5}$ the resulting spectrum is again flat. Other configurations of the magnetic field can also lead to flat emission spectra. The spectral slope is very sensitive to the jet shape. For example, slightly changing the jet shape to $r \propto x^{1/4}$ results in a spectrum with $F_{\nu} \propto \nu^{0.38}$. We show an application of the model to observations of a resolved jet in the X-ray binary Cygnus X-1 \citep{ssf01}. As input we use the flux density of the jet in the flat part of the spectrum, the physical extent of the resolved jet and an assumed break of the flat spectrum to optically thin conditions in the near-IR. Both model classes can be made consistent with the observational constraints. However, in doing so the adiabatic jet models require a significant departure from energy equipartition between the magnetic field and the relativistic particles. The associated reduced radiative efficiency of the jet plasma implies extremely high energy transport rates for the jet of around $10^{34}$\,W. This jet power exceeds the Eddington limiting luminosity of a 10\,M$_{\odot}$ black hole by two orders of magnitude. The ballistic jet is consistent with current observations and requires energy transport rates well below the time-averaged jet power \citep{gfk05}. This result holds even if the kinetic energy of one non-radiating proton per relativistic electron is taken into account. Unless Cygnus X-1 ceases to produce a jet for long periods of time, then the ballistic model requires a proton-electron jet plasma to explain the large accumulated energy in the region where the jet interacts with the surrounding gas. The jet opening angle of 5" required by the ballistic model is very small. The velocity perpendicular to the jet axis established in the jet acceleration region must be very small for this model to work. If the ballistic jet results from a rapid initial expansion of a highly overpressured jet, then the requirements are even more severe. This result places stringent limits on the acceleration process. Both model classes can probe the conditions in jets on scales unresolvable with current telescopes. The most important measurement in this respect is the high frequency cut-off of the flat spectrum which is formed closest to the jet acceleration region. In order to resolve the remaining uncertainties with the model, the most helpful measurements would be to resolve the jet along its axis in quasi-simultaneous observations at two different frequencies. As mentioned above, even a factor 2 between the observing frequencies employed would help to constrain the geometrical shape of the jet and thereby help to decide which of the model classes is compatible with observations. Obviously from the work presented here we cannot rule out the possibility of continued re-acceleration of the radiating electrons in the jet. However, such an energetisation process is not necessary to produce flat spectra from self-absorbed, synchrotron emitting jets. \section*{Acknowledgments} It is a pleasure to thank E. Gallo, J. Miller-Jones and R.P. Fender for helpful discussions. I also thank the referee A. Marscher for many helpful comments that improved the paper. \def\newblock{\hskip .11em plus .33em minus .07em} \bibliography{crk} \bibliographystyle{mn2e}
Title: Planets and Asteroids in the gamma Cephei System
Abstract: The binary star system gamma Cephei is unusual in that it harbours a stable giant planet around the larger star at a distance only about a tenth of that of the stellar separation. Numerical simulations are carried out into the stability of test particles in the system. This provides possible locations for additional planets and asteroids. To this end, the region interior to the planet is investigated in detail and found to permit structured belts of particles. The region between the planet and the secondary star however shows almost no stability. The existence of an Edgeworth-Kuiper belt analogue is found to be a possibility beyond 65 au from the barycentre of the system, although it shows almost no structural features. Finally, the region around the secondary star is studied for the first time. Here, a zone of stability is seen out to 1.5 au for a range of inclinations. In addition, a ten Jupiter-mass planet is shown to remain stable about this smaller star, with the habitability and observational properties of such an object being discussed.
https://export.arxiv.org/pdf/astro-ph/0601609
\label{start} \begin{keywords} stars: individual: $\gamma$ Cephei (HR 8974) -- planetary systems -- binaries: general -- methods: \textit{N}-body simulations \end{keywords} \section{Introduction} $\gamma$ Cephei is one of the closest separation binary systems that contains a planet. The primary and secondary stars are separated by 18.5 au. The planet (designated $\gamma$ Cep Ab) has a minimum mass of 1.7 $\MJ$ and follows an eccentric orbit about the larger star alone. The observational parameters of $\gamma$ Cephei, as determined by \citet{Ha03}, are summarised in~Table~\ref{tab:orbels}. Although the data are reasonably reliable, there is some uncertainty regarding the masses of the components. The inclination to the line of sight of a spectroscopic binary is usually indeterminate. This means that the mass of the primary is generally reckoned from spectral type. Then, the mass function (e.g., Smart 1977) permits a minimum mass to be assigned to the secondary, albeit crudely (Griffin, Carquillat \& Ginestet 2002). For $\gamma$ Cephei, the mass of the primary is $\approx 1.57 M_{\odot}$ from photospheric modelling \citep{Fu04}. The most recent determination from stellar evolution models is $\approx 1.7 M_{\odot}$~\citep{Af05}. The mass of the secondary is given by \citet{Dv03} as $\approx 0.4 M_{\odot}$. However, assuming the original value of the primary's mass of 1.57 $M_{\odot}$, a minimum mass can be derived as $\approx 0.34$ $M_{\odot}$ from the velocity amplitude fitted by \citet{Ha03}. There have been a number of studies of the dynamics of the $\gamma$ Cephei system to date. Foremost is the numerical investigation of \citet{Dv03}, which does use an earlier, and slightly different, set of orbital parameters (see Table \ref{tab:altorbels}). They used Burlisch-Stoer and Fast Liapunov Indicator methods to show that the $\gamma$ Cephei system is stable over Myr time-scales. They carried out test particle integrations as a guide to the possible existence of further planets. The results show a stable inner region between 0.5 and 0.8 au and then an additional stable zone at low inclination around 1 au. \citet{Dv03} noted that this coincides with the 3:1 mean motion resonance. This resonance is stable in the $\gamma$ Cephei system, but is unstable in the Solar system. \citet{Dv03} conclude that Earth-mass planets (up to 90 $\ME$) can exist in this region which is fortunately just on the edge of the habitable zone \footnote{The habitable zone has boundaries which depend on assumptions regarding stellar luminosity and effective temperature, as well as the climate model adopted for the hypothetical habitable planet.}. An extension to this work by \citet{Dv04} showed that the planet's eccentricity is less important than that of the two stars for the stability of a second planet. \citet{Ha05} also carried out numerical studies of the system's stability using a Burlisch-Stoer integrator, for a range of possible configurations of the planet's and binary's semimajor axes, eccentricities and inclinations, confirming and extending the results of \citet{Dv03}. Here, we use both numerical simulations and analytic calculations to study zones of stability as possible locations of additional planetary companions to either star and asteroid or Edgeworth-Kuiper belt analogues. Edgeworth-Kuiper belts are of particular interest in binary systems, as they may be being observed (indirectly) in exosystems such as $\tau$ Ceti and $\eta$ Corvi (Greaves et al. 2004; Wyatt et al. 2005). The results fall into three categories: possible planetary companions and asteroid belts centred on the primary star (\S 2), planetary companions around the secondary star (\S 3) and finally possible Edgeworth-Kuiper belt about both stars (\S 4). \begin{table*} \centerline{ \scriptsize \begin{tabular}{|l|r@{ $\pm$ }l|r@{ $\pm$ }l|r@{ $\pm$ }l|} \hline Name &\multicolumn{2}{|c|}{$\gamma$ Cep A} & \multicolumn{2}{|c|}{$\gamma$ Cep B} & \multicolumn{2}{|c|}{$\gamma$ Cep Ab } \\ \hline Class &\multicolumn{2}{|c|}{K1IV sub-giant star} & \multicolumn{2}{|c|}{M dwarf star } & \multicolumn{2}{c}{Planet}\\ Mass & 1.59 & 0.12 $M_{\odot}$ & \multicolumn{2}{|c|}{0.4 $M_{\odot}$} & 1.7 & 0.4 $M_{J}$\\ Period (days) &\multicolumn{2}{|c|}{--} & 20750.6579 & 1568.6 & 905.574 & 3.08\\ Semimajor axis (au) &\multicolumn{2}{|c|}{--} & 18.5 & 1.1 & 2.13 & 0.05\\ Eccentricity &\multicolumn{2}{|c|}{--} & 0.361 & 0.023 & 0.12 & 0.05\\ Longitude of periastron ($^{\circ}$) &\multicolumn{2}{|c|}{--} & 158.76 & 1.2 & 49.6 & 25.6\\ Time of periastron passage (JD) &\multicolumn{2}{|c|}{--} & 2448429.03 & 27.0 & 2453121.925 & 66.9\\ \hline \end{tabular}} \large \caption{Best fit orbital parameters for the $\gamma$ Cephei system from \citet{Ha03}. Mass of star B is from \citet{Dv03}.} \label{tab:orbels} \end{table*} \begin{table} \centerline{ \scriptsize \begin{tabular}{|l|c|c|c|} \hline Name & $\gamma$ Cep A & $\gamma$ Cep B & $\gamma$ Cep Ab\\ \hline Class & K1IV sub-giant star & M dwarf star & Planet\\ Mass & 1.6$M_{\odot}$ & 0.4$M_{\odot}$ & 1.76$M_J$\\ Period (days) & -- & 25567.5 & 902.2 \\ Semimajor axis (au) & -- & 21.36 & 2.15\\ Eccentricity & -- & 0.44 & 0.209\\ \hline \end{tabular}} \large \caption{Orbital parameters for the $\gamma$ Cephei system used by \citet{Dv03}.} \label{tab:altorbels} \end{table} \section{Planets and Asteroid Belts around the Primary} \subsection{Algorithm} \label{sec:algorithm} For all the simulations, we use the parameters of the $\gamma$ Cephei system given in Table~\ref{tab:orbels}. Here and elsewhere in the paper, the equations of motion are integrated using a conservative Burlisch-Stoer method provided in the {\tt MERCURY} software package \citep{Ch99}. Although not as fast as sympletic methods, this was chosen because of its ability to provide close encounter data and handle highly eccentric objects. The two stars and planet are simulated as point masses for gravitational interactions. Any additional objects are taken as massless test particles to decrease integration times. Test particles are removed from the simulations when they collide with the primary star, or pass an ejection distance of the order of several hundred au. A collision with the primary means that the test particle has a position that lies within the body of the primary, as judged by its stellar radius of $0.02$ au (Hatzes et al. 2003). Close encounters are allowed to occur, and are defined as taking place whenever a test particle enters within one Hill radius $\RH$ of the secondary star or the planet, defined as \begin{equation} \RH = a \left( {\frac{m}{3M}} \right)^{1/3} \end{equation} where $m$ is the mass of the secondary or planet and $M$ is the mass of the primary. This works out as $\approx 8.1$ au for the secondary and $\approx 0.15$ au for the planet. To maintain accuracy, the variation in the system's total energy and angular momentum is monitored throughout each simulation. Using this to constrain the initial timestep to 1 day and the tolerance in the Burlisch-Stoer algorithm \citep{Pr99} to $10^{-12}$ leads to an overall fractional change in the system's energy $\Delta E/E$ of about $10^{-8}$ over a 100 Myr period. This can therefore be considered the maximum time the system can be accurately followed. All the simulations presented here are typically 1 Myr in time-scale, for which $\Delta E/E \approx 10^{-11}$ or better. It is straightforward to show that the $\gamma$ Cephei system has long term stability. A 100 Myr integration shows no major variation in the orbits. Regular short period variations do occur, for example, a slight oscillation of the planet's semimajor axis over time-scales equal to both its orbital period and the binary's orbital period. An additional secular variation is seen over a period of about 5500 years, evident in the eccentricity and longitude of the planet only. This secular period is in good agreement with the results of quadrupole theory (see eq. (36) of \cite{Fo00}). \subsection{Test Particles Interior to the Orbit of the Planet} \label{sec:tpip} Figure~\ref{fig:inc_illus} shows two possible configurations of test particles in the system considered here. In the first, the test particles are inclined to the common plane of the binary and planet. In the second, both the test particles and planet are coplanar, yet inclined relative to the plane of the binary. The binary is always assumed to be viewed edge-on (that is, it lies in a plane perpendicular to the plane of the sky). If the planet is at an inclination $i_{\rm Ab}$ relative to this, its mass must be increased accordingly by dividing by $\sin i$ where $i = 90^\circ - i_{\rm Ab}$. With the planet and binary coplanar, we begin by investigating the stability of test particles in the region interior to the known planet. A grid of particles with semimajor axis from 0.5 to 1.85 au and inclination from 0$^{\circ}$ to 50$^{\circ}$ with resolution 0.05 au and 5$^{\circ}$ respectively is integrated for 1 Myr. Thirty-six particles are started at each grid point with varying initial longitudes of pericentre ($\omega = 0^\circ, 60^\circ, \dots 300^\circ)$ and longitudes of ascending node ($\Omega = 0^\circ, 60^\circ, \dots 300^\circ$). The orbits are initially circular. The stability is then determined by computing the mean survival time $\ts$ in Myrs at each grid point averaged over the 36 test particles. The results are shown in Figure~\ref{fig:stabmapinner} and can be compared to figure 2 of \citet{Dv03}. Note that our stability index is slightly different to the criteria used by \citet{Dv03}, who removed test particles after they become orbit crossing. This may miss the occasional test particle that is stable, for example, if it lies in a Trojan-like orbit. We only remove test particles if they collide with the central star or are ejected from the system. The map shows test particles with semimajor axes less than $\approx 1.4$ au are stable. However, there is a strip of instability between roughly 0.8 and 1.0 au, creating an island of stability at low inclinations between 1.0 and 1.3 au. Some of the structure in the map can be clearly related to the positions of the mean motion resonances (MMRs) with the planet (indicated in Fig.~\ref{fig:stabmapinner}). The 4:1, 3:1, 5:2 and 2:1 resonances seem to mark transitions from stability to instability. For example, the 5:2 MMR divides the island at $\approx 1.15$ au. The lack of effect of some of the higher order MMRs, such as the 5:1 case, is probably due to their comparative weakness and narrowness. The lack of stability beyond $\approx 1.4$ au is readily explained. The gravitational reach of the planet as a multiple of the Hill radius (Jones, Underwood \& Sleep 2005) places the limit on stability at 1.31 au, a good match with the results here. Note that in calculating this limit the maximum eccentricity of the planet obtained during the simulation has been used. Many of the test particles in the high inclination region of Figure~\ref{fig:stabmapinner} show evidence of Kozai cycles, as illustrated by the orbits in Figure~\ref{fig:orbitsb}. The \citet{Ko62} instability is well-known from studies of high inclination comets and asteroids in the Solar system. It sets in at inclinations greater than a critical value of $i_{\rm crit} = {\rm asin} \sqrt{0.4} \approx 39.23^\circ$. During Kozai cycles, the eccentricity and inclination vary so as to maintain approximate constancy of the integral of motion $I_{\rm K} = \sqrt{1-e_{\rm tp}^2} \cos i_{\rm tp}$, where $e_{\rm tp}$ is the test particle's eccentricity. As the semimajor axis increases, the amplitude of eccentricity and inclination librations becomes larger, thus accounting for the increased instability evident in this region of Figure~\ref{fig:stabmapinner}. As already seen, the MMRs with the planet are important in shaping the regions of stability. However, one major feature unexplained by this is the instability strip between roughly 0.8 and 1.0 au. Although at zero inclination the edges are marked by the 4:1 and 3:1 MMRs, the instability strip shows a pronounced evolution with inclination which is suggestive of a secular resonance instead. The classical Laplace-Lagrange linear theory \citep{Mu00}, although derived for low eccentricity and inclination regimes around a dominant central mass, can be applied to give a first approximation of the locations of these resonances. A secular resonance for a test particle occurs when its precession rate has exactly the same magnitude as an eigenfrequency of the system. The eigenfrequencies are easily calculable for the three body system made up of the two stars and planet and are the eigenvalues of the 2x2 matrices $\mathbf{A}$ and $\mathbf{B}$ respectively, which have components \begin{eqnarray} A_{jj} & = & + n_j \frac{1}{4} \frac{m_k}{M+m_j} \alpha \bar{\alpha} b^{(1)}_{3/2}(\alpha),\nonumber\\ A_{jk} & = & - n_j \frac{1}{4} \frac{m_k}{M+m_j} \alpha \bar{\alpha} b^{(2)}_{3/2}(\alpha),\nonumber\\ B_{jj} & = & - A_{jj},\\ B_{jk} & = & - A_{jk} \frac{ b^{(1)}_{3/2}(\alpha)}{ b^{(2)}_{3/2}(\alpha)}.\nonumber \end{eqnarray} Here, $n_j$ is the mean motion of object $j$ (1 represents the planet and 2 represents the secondary), $m_j$ and $a_j$ are the mass and semimajor axis of object $j$, $M$ is the mass of the primary (the central object), $\alpha = a_1/a_2$ and $b^{(1)}_{3/2}(\alpha)$ and $b^{(2)}_{3/2}(\alpha)$ are Laplace coefficients. Using the values for the masses and semimajor axes given in Table \ref{tab:orbels} along with $n_1 = 145.2^\circ \yr^{-1}$, $n_2 =6.337 ^\circ \yr^{-1}$ and $\alpha=0.1151$ gives the Laplace coefficients as $b^{(1)}_{3/2}(\alpha)=0.3542^\circ \yr^{-1}$ and $b^{(2)}_{3/2}(\alpha)=0.05089^\circ \yr^{-1}$, employing the {\tt MATHEMATICA} routines of \citet{Mu00}. Calculating the matrices and solving for the eigenfrequencies gives \begin{eqnarray} g_1 &=&0.04338^\circ \yr^{-1},\nonumber\\ g_2 &=& 0.00005211^\circ\yr^{-1},\\ f &=& -0.04343^\circ\yr^{-1},\nonumber \end{eqnarray} where $g_1$ and $g_2$ are the eigenvalues of $\mathbf{A}$ and $f$ is the degenerate eigenvalue of $\mathbf{B}$. Note that the $g_1$ and $f$ eigenfrequencies have about the same magnitude, whilst $g_2$ is almost zero due to the large mass ratio between the planet and secondary star. The precession rate of the test particle is given by \begin{equation} A_{\rm tp} = n \frac{1}{4} \left(\frac{m_1}{M}\alpha_1\bar{\alpha}_1b^{(1)}_{3/2}(\alpha_1) + \frac{m_2}{M}\alpha_2\bar{\alpha}_2b^{(1)}_{3/2}(\alpha_2) \right) \end{equation} where $n$ is the particle's mean motion. For the region interior to the planet $\alpha_j = \bar{\alpha}_j= a/a_j$, where $a$ is the test particles semimajor axis. Plotting $A_{\rm tp}$ as a function of $a$ from 0.5 to 1.85 au shows a resonant location where the $g_1$ and $f$ eigenfrequencies intersect the curve at $\approx 0.8$ au. This supports the idea that the location of the inner edge of the instability strip on the map coincides with a secular resonance. Comparing our results with those of \citet{Dv03}, it is easy to see that the broad trends are similar, despite slightly differing orbital elements. The main stable regions are slightly closer to the star in \citet{Dv03}. This may be due to a higher eccentricity of both the planet and the secondary, which means that they approach the central star more closely, reducing separations with the test particles. \citet{Dv03} find that the 3:1 MMR is stable, in contrast to asteroids in the Solar System in the same resonance with Jupiter. Here, we find that the resonance is unstable, with a particle following a fairly steady evolution until switching to a mode where its eccentricity is rapidly driven to unity on a time-scale of 10 kyrs, as shown in Figure~\ref{fig:orbitsa}. The case where the planet is also inclined ($i_{Ab} \neq 0$) has not been previously studied. This is a more likely configuration for a system that has formed in a common protoplanetary disc. Here, we investigate this case by using the same grid of test particles as before, but with the planet sharing the same inclination as the test particles and with $\Omega_{Ab} = 0$. This means that, to reproduce the same radial velocity dataset, the mass of the planet must be increased by dividing by $\sin i$, where $i = 90^\circ - i_{Ab}$. The results are displayed in Figure~\ref{fig:stabmapincl}. As compared to the earlier case of Figure~\ref{fig:stabmapinner}, the unstable region has expanded, especially at high inclinations. This is partly caused by the change in the extent of the gravitational reach of the planet, as shown by the dotted curve in Figure~\ref{fig:stabmapincl}. Although this does scale with the increasing planetary mass, the sharp change at $40^\circ$ inclination is due to the planet becoming subject to Kozai cycles. The large increase in eccentricity here means that the planet's periastron is much closer to the star, and hence its gravitational influence is larger. At $50^\circ$, the planet is unstable, colliding with the central star after $\approx 0.5$ Myr. At first sight, it may seem that the rest of the increased instability evident in Figure~\ref{fig:stabmapincl} is caused by the increased mass of the planet. However, experiments show that this is not so. For example, an integration of the $30^\circ$ case with the planet inclined but at minimum mass ($1.7 \MJ$) shows very little difference to Figure~\ref{fig:stabmapincl}. This suggests that the cause lies in the relative inclination of test particle and planet to the plane of the binary. In the case of Figure~\ref{fig:stabmapinner}, the amplitude of libration of $i_{\rm tp}$ and $e_{\rm tp}$ of test particles is generally modest for all cases below $i_{\rm crit}$, the critical value for the Kozai instability. For test particles in Figure~\ref{fig:stabmapincl}, the amplitude is no longer small and becomes larger with increasing inclination, rapidly driving particles into the regime where the Kozai instability is effective. This is understandable as in the former case, the forces due to the masses in the system are always directed towards the plane of the binary, whereas in the latter case this is not true. As the inclination increases, the misalignment between the forces due to the stars and the force due to the planet increases, and so the amplitude of libration increases. \subsection{Test Particles Interior to the Orbit of the Binary} \label{sec:tpib} Holman \& Wiegert (1999) have already studied the stability of test particles in binary systems. These may orbit either a single star or both stars. Here, we study test particles around the primary star. For this case, Holman \& Wiegert (1999) introduced the notion of a critical semimajor axis $a_{\rm crit}$, which is the largest circular orbit around the primary for which a ring of test particles survives for at least $10^4$ binary periods ($\approx 600$ kyrs in the case of $\gamma$ Cephei). Using their eq~(1), we find $a_{\rm crit} = 4.0 \pm 0.6$ au. In other words, our expectation is that all test particles starting out at semimajor axes greater than 4.0 au will be rapidly swept out. The existence of the comparatively large and eccentric planet $\gamma$ Cephei Ab will cause further de-stabilization. To investigate this, simulations of test particles in the region from 0.5 to 20.0 au are carried out. They have initially zero eccentricity and are set up on a grid with resolution 0.5 au in starting semimajor axis. The range of inclinations is restricted from $0^\circ$ to $30^\circ$ for the prograde case, and from $150^\circ$ to $180^\circ$ for the retrograde case, both in steps of $10^\circ$. Seventy-two particles are started at each grid point with varying initial longitudes of pericentre ($\omega = 0^\circ, 30^\circ, \dots 330^\circ)$ and longitudes of ascending node ($\Omega = 0^\circ, 60^\circ, \dots 300^\circ$). The orbits of the test particles are followed for 1 Myr, and any close encounters are recorded. Although this is a limited range of inclinations, it is expected that those not investigated are largely unstable. This receives confirmation from exploration integrations in the case of $70^\circ$ inclination, for which very few test particles survive throughout the entire region. Figures~\ref{fig:inner} and \ref{fig:innerAb} show the ultimate fates of the test particles within 5 au. They differ in that the planet is in the plane of the binary and test particles are inclined in Figure~\ref{fig:inner}, whilst both planet and test particles are similarly inclined in Figure~\ref{fig:innerAb}. Four panels showing the results of selected simulations are displayed in each case, corresponding to prograde with $i_{\rm tp} =0^\circ$ and $30^\circ$, retrograde with $i_{\rm tp} = 180^\circ$ and $150^\circ$. In the inner regions, test particles close to the planet are swept out on a precession time-scale ($\approx 5.5$ kyrs). Within $\approx 3$ au, prograde test particles are either ejected or collide with the central star after a close encounter with the planet. This is evident from the particles coloured orange and red in the upper panels of both Figures~\ref{fig:inner} and \ref{fig:innerAb}. The retrograde case is different. The lower panels of Figure~\ref{fig:inner} show swathes of stable particles coloured either brown or green, according to whether they reside within the Hill sphere of the secondary or not. Note that the secondary's periastron is at $11.8$ au, so particles out to $3.7$ au are within its Hill sphere at some point. In the case $i_{\rm tp} = 180^\circ$, the retrograde stability zone extends out to $\approx 7$ au. As the inclination decreases ($i_{\rm tp} \rightarrow 150^\circ$), the stability zone shrinks to the annuli between $0.5$ to $1.0$ au and $3.0$ to $5.0$ au. The lower panels of Figure~\ref{fig:innerAb} show the case when both test particles and planet are retrograde. In the case $i_{\rm tp} = 180^\circ$, the large stability region seen previously has almost completely disappeared. There are only a few remaining test particles that survive the 1 Myr integration. The similarity of the two right-hand panels of this figure shows that the inner region's evolution is almost entirely controlled by the planet. \begin{table*} \caption{\label{table:pevtable} The number of test particles interior to the orbit of the binary at each starting semimajor axis and inclination that survive for 1 Myrs. There are 72 test particles initially for all cases except $i=0^\circ$ and $i=180^\circ$, which have 12. } \centerline{ \scriptsize \begin{tabular}[width=0.9\textwidth]{c||c|c|c|c|c|c|c|c|c|c|c|c|c|c|c||c|c|c|c|c} \hline \null &\multicolumn{20}{c}{Starting Semimajor Axis [in au]}\\ Inclination & 0.5 & 1.0 & 1.5 & 2.0 & 2.5&3.0 & 3.5 & 4.0 & 4.5 & 5.0 & 5.5 & 6.0 & 6.5 & 7.0 & 7.5 & 12.0 & 12.5 & 13.0 & 13.5 & 14.0 \\ \hline $0^\circ$ & 12 & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - \\ $10^\circ$ & 72 & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & 2 \\ $20^\circ$ & 72 & - & - & - & - & - & - & - & - & - & - & - & - & - & - & 1 & - & 2 & - & - \\ $30^\circ$ & 72 &72 & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - \\ $150^\circ$ & 72 &67 & - & - & - &23 &72 &70 &67 &72 &60 &11 &14 & - & - & - & - & - & - & - \\ $160^\circ$ & 72 &72 &72 & - & - &54 &72 &72 &72 &72 &62 &39 &25 & 3 & 1 & - & - & - & - & - \\ $170^\circ$ & 72 &72 &72 & - &24 &72 &72 &72 &72 &72 &68 &49 &15 & 4 & 1 & - & - & - & - & - \\ $180^\circ$ & 12 &12 &12 & - &11 &12 &12 &12 &12 &12 &12 &12 & 8 & 4 & 1 & - & - & - & - & - \\ \hline $0^\circ$ & 12 & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - \\ $10^\circ$ & 72 & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - \\ $20^\circ$ & 60 &17 & - & - & - & - & 1 & 2 & - & - & - & - & - & - & - & 1 & - & 1 & 1 & - \\ $30^\circ$ & 35 & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - & - \\ $150^\circ$ & 37 & - & - & - & - & - & - & - & 5 & - & 1 & - & - & - & - & - & - & - & - & - \\ $160^\circ$ & 69 & - & - & - & - & - & 2 & 1 &10 &13 &15 &10 & 4 & 1 & - & - & - & - & - & - \\ $170^\circ$ & 72 & - & - & - & - & - &25 &16 & 4 & 9 &11 &11 & - & - & - & - & - & - & - & - \\ $180^\circ$ & 12 &12 & - & - & - & - & 6 & 5 & 4 & 2 & 2 & - & - & - & - & - & - & - & - & - \\\hline \end{tabular} } \end{table*} The numbers of test particles surviving after 1 Myr at each semimajor axis are given in Table~\ref{table:pevtable}. Shown in Figure~\ref{fig:survivtimes} are the mean survival times plotted against semimajor axis for a variety of inclinations. The mean survival time $\ts$ can be computed by averaging over the results for the differing longitudes of ascending node and pericentre at fixed semimajor axis and inclination. The averaging is over 12 test particles for $i_{\rm tp} = 0^\circ$ and $180^\circ$ and 72 for all other cases. Test particles that survive to the end of the integrations are included with a survival time of 1 Myr, so that the computed mean survival time is a lower limit in these cases. Figure~\ref{fig:survivtimes} shows that there is a region of enhanced stability with $\ts \approx 100$ kyrs -- for all the studied inclinations -- centred at $\approx 3.5$ au, just beyond the gravitational reach of the planet but within the critical semimajor axis and just within the region affected by the secondary. Most test particles here (and also beyond this region) suffer a close encounter with the secondary before ejection or collision with the primary. The effect of the secondary star becomes increasingly important with increasing semimajor axis and $\ts$ falls to $\approx$ 1 kyr. There is, for the prograde cases, a region of enhanced stability around $12$ au for some inclinations, clearly visible in Figure~\ref{fig:survivtimes}. This is due to a few particles surviving here for the full length of the integration. As stated in the introduction, the mass of star B is uncertain. By altering this parameter and rerunning some of the simulations the importance of the uncertainty can be seen. For the coplanar configuration of test particles described in this section, changing the mass of star B by $\sim 20 \%$ results in almost no difference in global statistical properties, such as average survival times. This would indicate that the system's dynamics are not significantly affected by the uncertainty in the secondaries mass. \section{Planets and Asteroid Belts around the Secondary} \label{sec:b} Here, we investigate whether planets could exist around star B. This possibility has not been investigated before for this system. The critical semimajor axis for stability, as defined by Holman \& Wiegert (1999), is $1.9 \pm 1.0$ au. To investigate further, test particles are set up from $0.5$ to $2.5$ au in steps of $0.05$ au for inclinations $0^\circ$ to $30^\circ$ and $150^\circ$ to $180^\circ$ in steps of $10^\circ$. The particles are on initially circular orbits again, and spaced in longitude of perihelion and ascending node by $60^\circ$ as before. The results of two of the prograde cases are shown in Figure~\ref{fig:starB}. Prograde test particles are not stable beyond $1.5$ au. In the case of $10^\circ$, $20^\circ$ and $30^\circ$ inclinations, the unstable test particles generally survive at least ten times longer than those in the $0^\circ$ case. The unstable test particles within about $2$ au all collide with the primary, whilst those beyond this point have a range of fates. Holman \& Wiegert's critical semimajor axis does not match up as well as in Section~\ref{sec:tpib}, but still agrees within the rather large uncertainty. In the retrograde cases, almost all the test particles are stable, with the exception of 7 particles at $2.45$ au in the $150^\circ$ case. Integrating more distant particles shows that the retrograde stability reaches out to about $3.5$ au. The stability of the test particles can be further investigated by plotting the evolution of their eccentricity and semimajor axis, as shown in Figure~\ref{fig:starBae}. For all the simulations, there is not much change in the inclination of the test particles. The variation in eccentricity and semimajor axis of stable particles increases as the initial distance from star B increases, and is similar for all the prograde cases. However, the variation is much larger for the (more stable) retrograde cases. When $i_{\rm tp} = 150^\circ$, the variations are no longer smoothly increasing and show abrupt jumps, indicating that this case may not be long term stable. As test particles can remain stable around the secondary, this raises the question of whether a massive planet could also persist here. So, it is interesting to consider whether this could be detectable in the radial velocity curve of the primary star. To investigate this, the simple case of a $10$ Jupiter-mass planet in a coplanar, circular orbit with initial semimajor axis $0.5$ au and initial longitude $0^\circ$ is integrated. The orbital elements of the secondary are adjusted so that the centre of mass of it and its putative planet orbits the primary with the elements shown in Table 1 for star B alone. As expected from the test particle results, the planet remains stable for the full Myr length of the integration and shows very little variation in its orbit. However, the 1 Myr integrations here may overestimate the extent of the stability zone for planets, as instabilities can appear even after 100 Myr of apparent stability (Jones, Sleep \& Chambers 2001). To calculate the radial velocity curve, the system is assumed to be at zero inclination relative to the line of sight ($i = 90^\circ$). Figure~\ref{fig:rv} shows the radial velocity curve of the primary, together with the residuals with respect to the case with no extra planet. There is no detectable signal with the period of the extra planet. The small variations shown, which have the period of the binary, amount to 25 ms$^{-1}$ over 1000 yr timespans, would be undetectable. \section{Edgeworth-Kuiper Belt Analogues} \label{sec:kb} The final region remaining to be studied is that exterior to the binary. This may host particles analogous to the Edgeworth-Kuiper belt in our own Solar system. There is observational evidence for the existence of extrasolar Edgeworth-Kuiper belts from infrared imaging of dusty discs around other stars (e.g., Wyatt 2003, Greaves et al. 2004), so the longevity of such a structure around $\gamma$ Cephei is worth investigation. Numerical integrations of test particles around binary systems suggest that they will remain stable in this region (e.g., Harrington 1977). In addition to the criterion in Section~\ref{sec:tpib}, Holman \& Wiegert (1999) also provide an empirical rule-of-thumb to determine prograde test particle stability exterior to the binary. Here, the critical semimajor axis $a_{\rm crit}$ is that beyond which almost all test particles survive for at least $10^4$ binary periods. Using eq. (3) of their paper, we find that $a_{\rm crit} = 64 \pm 2$ au. However, Holman \& Wiegert (1999) caution that there can sometimes be islands of instability beyond the critical radius associated with $n$:1 MMRs. There is an older criterion due to Harrington (1977), who also considers the case of retrograde test particles. Harrington's equation suggests that prograde test particles are stable for $a \gta 56$ au and retrograde stable for $a \gta 45$ au. By stability, Harrington means that the particles show bounded motion with no secular or large periodic variations in their elements over his -- by nowadays standards -- very small integration timespans. Using the same method as in Sections~\ref{sec:tpip} and \ref{sec:tpib}, test particles are set up with inclinations $0$ to $30^\circ$ and $150^\circ$ to $180^\circ$ in the (barycentric) region from $20$ to $150$ au in steps of 5 au. The mean survival times are shown in Figure~\ref{fig:outer}. The prograde test particles exhibit a sharp cut-off, with those beyond $65$ au being stable, independent of the starting inclination. This figure also shows that the retrograde test particles survive to the end of the integration for starting semimajor axes beyond 40 au for $180^\circ$ inclination, 45 au for $170^\circ$ and $160^\circ$ inclinations and 60 au for $150^\circ$ inclination. The range of initial conditions for which retrograde particles survive is larger than for prograde (e.g., Harrington 1977). For all inclinations, unstable particles from about $20$ to $40$ au are removed quickly and generally are ejected after a close encounter with the secondary. Unstable particles from about $40$ to $65$ au are less rapidly removed and tend to collide with star A or be ejected from the system, without experiencing a prior close encounter. We see that both Harrington's (1977) and Holman \& Wiegert's (1999) stability criteria seem to give a reasonable description of the results of our simulations. The average maximum eccentricity and change in semimajor axis is shown in Figure~\ref{fig:baryboth} as a function of initial semimajor axis for all eight inclinations. All the test particles show a small secular variation in their orbital elements. The Edgeworth-Kuiper belt in our Solar system shows some structure due to locations of MMRs with the giant planets (e.g., Luu \& Jewitt 2002). However, the results in Figure~\ref{fig:outer} do not indicate any such resonant features. This is most likely due to the coarse semimajor axis grid employed. To investigate this, the coplanar case was re-run, but now with a spacing in semimajor axis of $1$ au. The mean survival times and fates of individual particles for this simulation are shown in Figure~\ref{fig:14}. The first location where all test particles survive is now $64$ au, agreeing exactly with Holman \& Wiegert's value of $a_{\rm crit}$. There is then an unstable band from $67$ to $70$ au that seems to match up to the 7:1 MMR. The plot gives no evidence for any other resonant effects beyond this location. This is understandable since the locations of the other major resonances are outside the region of Hill stability (for example, the 3:2 resonance that is important in our own Solar system), leaving only those of very high order to affect stable test particles. The distinct difference in fates of test particles is obvious in this plot, with the blue coloured points, indicating close encounters with star B, not occurring beyond about $45$ au. At this point, the test particles are yellow or grey, indicating either ejection or collision with star A without undergoing any close encounter. \section{Conclusions} In this paper, we have carried out a suite of test particle integrations for the $\gamma$ Cephei system, as a guide to possible locations for additional planets. For test particles in the plane of the binary and planet, there are three zones of stability for 1 Myr timespans at least. These are [1] the region interior to the planet, which is stable within the bands $1.2$ and $1.3$ au, $1.05$ to $1.15$ au, and $0.75$ to at least $0.5$ au, [2] the region around star B from $1.5$ au into at least $0.5$ au and [3] the region around both stars extending out from about $65$ au. These can be used to constrain possible locations of additional planetary companions. The region interior to the planet has a complicated structure. Low order resonances, such as the 4:1, 3:1 and 5:2, mark transitions between stable and unstable regimes. There is a secular resonance located at $0.8$ au from the primary that also plays a role in the dynamics of the test particles. The results for this region match up well with the previous work of Dvorak et al. (2003) despite the differences in the parameters of the system and the methods used. This implies that the slight improvements in the observations of the system have not significantly altered its dynamical characteristics. One difference though is that, unlike Dvorak et al. (2003), we find that test particles close to the 3:1 resonance are unstable, just as for the asteroid belt in the Solar System. In agreement with Dvorak et al. (2003), we find that small planets interior to $\gamma$ Cephei Ab are long-lived. Another as yet unconsidered possibility seen from the results here is the existence of an asteroid belt interior to the giant planet. However, the stability seen for test particles in this region disappears when $\gamma$ Cephei Ab is inclined in the same plane as the test particles. Now, the test particles are rapidly driven to inclinations which are subject to the Kozai instability. The region interior to the binary and exterior to the planet is unpromising. It seems unlikely that any prograde planet or asteroid can remain stable between $\gamma$ Cephei Ab and the binary, although retrograde test particles in this region are more long-lived. The region around the secondary is stable out to $1.5$ au for test particles. We have shown that planets up to 10 Jupiter-masses in circular orbits at $0.5$ au can survive for at least 1 Myr. This seems to be a promising place for additional planets to reside, in a similar manner to satellites about a planet around a single star. In the final region, that exterior to the binary, test particles are also stable. Although planets would be able to reside out here, the existence of an Edgeworth-Kuiper belt structure is also a possibility. Once again the retrograde particles are more stable than those that are prograde. Whilst in every region studied this is true, such objects are perhaps unlikely, if the origin of the system was a common disk. If particles are captured from elsewhere, then this may become a possibility. Retrograde asteroids and comets certainly exist in the Solar system. Since the survival of additional planets in the system has been shown to be a possibility, it is interesting to consider their habitability. There are a number of estimates of the habitable zone around star A in the literature. For example, Dvorak et al. (2003) place it at $1.0$ to $2.2$ au, in which case habitable planets could exist at the very edge of the zone. However, Haghighipour (2005) places the habitable zone at $3.1$ to $3.8$ au, while Jones et al. (2005) place it at $2.07$ to $4.17$ au. These different locations reflect differences in the criteria for the location of the habitable zone or the assumed stellar luminosity and effective temperature. The zone $2.07$ to $4.17$ au is unstable for all except retrograde test particles. So, this would permit the possibility of a habitable planet only if retrograde. The detection of any such small planet is challenging, given the small (of the order of a few ms$^{-1}$) radial velocity signatures that they cause. In addition, a retrograde planet appears in radial velocity curves merely as a prograde one with a $180^\circ$ phase difference. Although retrograde objects -- especially planets -- are thought to be unlikely, it has been shown that giant planets in binary systems can end up with retrograde orbits after close encounters within a planetary system (Marzari et al. 2005). A more interesting possibility is that of a habitable planet around star B, since it has been shown that planets can survive here. Theories of planetary formation do not preclude this possibility (e.g., Armitage, Clarke \& Palla 2003). The habitable zone for an M dwarf extends out to about 0.3 au (Kasting, Whitmire \& Reynolds 1993), which is within the stable zone found here. The large primary star nearby might also act as a `shield' from comets and asteroids similar to Jupiter for the Earth. As seen in the results in Sections~\ref{sec:b} and \ref{sec:kb}, very few test particles collide with the secondary star once the region interior to the binary has been rapidly cleared. Edgeworth-Kuiper belt objects that are perturbed into the inner regions of the system are also likely to suffer encounters with the primary, thus leaving the secondary and its environs largely unscathed. The detection of such a habitable planet around star B has, unfortunately, been shown to be virtually impossible from the radial velocity signature of star A alone. Should the secondary be resolved this would change, and any companion giant planet would be easily detectable. The region around star B, none the less, remains as a promising place for the existence of a habitable planet in this system. \section*{Acknowledgments} PEV acknowledges financial support from the Particle Physics and Astronomy Research Council. We thank the referee for his helpful report. \label{lastpage}
Title: Stable Models of superacceleration
Abstract: We discuss an instability in a large class of models where dark energy is coupled to matter. In these models the mass of the scalar field is much larger than the expansion rate of the Universe. We find models in which this instability is absent, and show that these models generically predict an apparent equation of state for dark energy smaller than -1, i.e., superacceleration. These models have no acausal behavior or ghosts.
https://export.arxiv.org/pdf/astro-ph/0601517
\bibliographystyle{prsty} \preprint{UCI-TR-2006-1} \title{Stable Models of Super-acceleration} \author{Manoj\ Kaplinghat and Arvind\ Rajaraman} \affiliation{Department of Physics and Astronomy\\ University of California, Irvine, California 92697, USA} \date{\today} \pacs{98.70.Vc} \section{Motivation\label{sec:motivation}} Observations of distant Type Ia supernovae \cite{riess98, perlmutter99} and the cosmic microwave background \cite{spergel03} together strongly prefer an accelerated expansion of the universe in the recent past. In the standard cosmological model this is accommodated by introducing ``dark energy'', a component which has a significantly negative pressure causing the expansion of the universe to accelerate. In the standard cosmological model, dark energy is completely decoupled from the rest of the matter in the universe except for its gravitational effects. It is interesting to consider more general models in which the dark matter and dark energy have a coupling. Such models could have new nontrivial signatures in cosmology and structure formation. One simple class of such models is a model in which the vacuum energy density depends on the matter density. We shall consider a class of these models in which the dark energy responds to changes in the matter density on a time scale shorter than the expansion time scale. For example, one can consider models with scalar field dark energy coupled to matter (\eg \cite{casas91,anderson97,amendola99,bean01,comelli03,farrar03,chimento03}), in which the mass of the scalar field is much larger than the expansion rate (for example, the MaVaN scenario \cite{fardon03}). As we show below, these models generically suffer from an instability which we label AZK-instability. The AZK-instability was pointed out in the context of mass-varying neutrinos (MaVaN) \cite{afshordi05}. A similar effect was identified in the context of unified dark energy models \cite{beca05}. This instability can also occur in models of dark energy coupled to matter, such as the MaVaN scenario \cite{fardon03}, the Chameleon dark energy scenario \cite{brax04} and the Cardassian expansion scenario \cite{freese02}. Not all models in the above scenarios are necessarily unstable (for example, \cite{fardon06,koivisto05,takahashi06}). This will become clear when we discuss the instability. In this paper, we will construct a large class of models in which this instability is avoided. We find that these models generically predict an apparent equation of state (pressure over energy density) $\wde$ which is less than -1 (such a phase is labeled super-acceleration~\cite{kaplinghat03b}). That is, a model of interacting dark energy can be incorrectly interpreted as a theory with super-acceleration if the interactions are not taken into account. For example, the coupling of dark energy to matter could be such that the total matter density decreases more slowly than $1/a^3$ (where $a$ is the scale factor of the universe). When we interpret observations in such a universe with a canonical matter density term (that decreases with expansion as $1/a^3$) and dark energy, we would infer an equation of state for dark energy more negative than it truly is \cite{huey04,das05}. There is no physical reason why this inferred equation of state cannot be below -1. This is particularly interesting because current data seem to favor a dark energy density which is almost constant or even increasing with time \cite{caldwell99,schuecker02,tonry03,knop03,choudhury03,alam03,melchiorri03,majerotto04,astier06,schaefer05}. and exciting results can be expected in the future \cite{weller01,frieman02,linder03,kratochvil04}. SNIa observations currently favor a phase of super-acceleration. Future SNIa and CMB observations have the potential to detect super-acceleration \cite{kaplinghat03b}. No other combination has been shown to robustly detect the signature of super-acceleration, although combining SNIa and baryon oscillation \cite{astier06} or weak lensing data set seem promising. Note that a measurement of just the average equation of state \cite{saini03} is not sufficient for this purpose \cite{maor01}. This was made explicit recently \cite{csaki05} using a simple single scalar field model. Scalar field models with canonical kinetic terms always produce $\wde > -1$. Effective models with the opposite sign kinetic term \cite{caldwell99,schulz01} imply $\wde < -1$ but are unstable \cite{carroll03} unless more than one scalar field \cite{feng04,guo04,hu04,wei05,urena05} or quantum effects \cite{onemli04} are considered. Models with higher derivative terms or scalar-tensor theories can give rise to an apparent $\wde < -1$ \cite{boisseau00}, but are constrained \cite{carroll04,vikman04,abramo05}. Interpreting an alternative gravity theory in the context of 4-d GR can also lead to super-acceleration \cite{mcinnes01,sahni02,pietroni02,elizalde04,nojiri05,martin06}. Some Cardassian models may have $\wde <-1$ \cite{wang03,koivisto04,freese05} while still satisfying the dominant energy condition. Another possible way to get super-acceleration with no instabilities is to appeal to photon-axion mixing (conversion of photons to axions) in a universe dominated by a cosmological constant (or quintessence) \cite{csaki04}. In our models, the superacceleration arises due to interactions of dark energy and matter. Our models therefore provide super-acceleration with none of the attendant problems that plague most of the above models. Furthermore, the interactions are generic; we do not need to fine-tune couplings in order to avoid theoretical pitfalls or observational constraints. We therefore believe that considering interactions of dark energy is the best way to generate models of superacceleration. \section{AZK-instability\label{sec:instability}} In this section we will consider a general class of models in which the dark energy density is coupled to the non-relativistic matter density. For an example of how this could occur, suppose that non-relativistic matter particles are coupled to a scalar field. Thus the local density of the matter particles can influence the vacuum expectation value (vev) of the scalar field. The change in the potential of the scalar then affects the dark energy, thus coupling matter and dark energy. In this class of models, the matter fields will be taken to have a matter density $\rhom$. They are coupled to a scalar field $\chi$ (dark energy) through Yukawa like couplings. We take the potential to be \bea E&=&\int d^3x\ V(\chi,\rhom)\,,\\ &=&\int d^3x\ \left[ V_0(\chi)+m\rhom+\lambda g(\chi)\rhom \right]\,.\label{eq:defineV}\eea We will assume that \( m_\chi^2=V''_0(\chi_0)+\lambda g''(\chi_0)\rhom\), the mass-squared of the scalar field about its vev $\chi_0$, is very large so that the $\chi$ field always sits at the minimum of its effective potential. This is the central assumption of our paper. The mass will certainly have to be larger than the expansion rate of the universe to be consistent with this assumption. We will also assume that the mass is large enough to satisfy the constraints imposed by experiments that probe the strength of a fifth force. In the absence of the last term, this is the potential energy of two decoupled fluids. The first term corresponds to a cosmological constant term (since we have assumed that the field $\chi$ is always at the minimum). The second term is the energy density of a dark matter fluid with density $\rhom$ and particle mass $m$. The last term couples these two fluids, and leads to interesting effects. In particular $\chi_0$, the value of the scalar field at its minimum is now found by solving the equation \bea V'_0(\chi_0)+\lambda g'(\chi_0)\rhom =0 \,,\label{eq:potmin} \eea where $V_0'$ and $g'$ are derivatives of $V_0$ and $g$ with respect to $\chi$. Thus $\chi_0$ is now a function of $\rhom$. We can make the dependence of $\chi_0$ on $\rhom$ explicit in the following way. Consider small deviations in $\rhom$. The vev of the scalar field shifts to account for this change in $\rhom$. Taking a further derivative, we find \bea (V''_0(\chi_0)+\lambda g''(\chi_0)\rhom){\partial\chi_0\over\partial\rhom}+\lambda g'(\chi_0) =0 \,.\label{eq:dchi0dn}\eea This explicitly shows how $\chi_0$ varies as $\rhom$ varies. In writing Eq.~\ref{eq:defineV}, we neglected the kinetic term in comparison to the potential. This is necessary if the scalar field is to behave as dark energy and, as we now show, consistent with our assumption of a large mass for the scalar field. Note that $\dot{\chi}=\dot{\rhom}\partial \chi_0/\partial \rhom$. Working out this expression, we find that $\dot{\chi}^2/V$ for $\chi=\chi_0$ is given by $(V_0^{\prime 2}/V m_\chi^2)(\dot{\rhom}/m_\chi\rhom)^2$. Lets look at changes to the scalar field potential around $\chi=\chi_0$. Unless there are strong fine-tunings and cancellations, we will have $V_0'\pert{\chi} < V$ and $m_\chi^2(\pert{\chi})^2/2 < V$, which together imply that $2V_0^{\prime 2}/V m_\chi^2 < 1$. Hence the natural expectation is that $\dot{\chi}^2/V \sim H^2/m_\chi^2$. For large enough $m_\chi$, the kinetic term is negligible. We now show that there is an instability in this system. We start with a configuration where the dark matter is evenly distributed, and the $\chi$ field is at its minimum $\chi_0$ everywhere. Now consider small fluctuations in the matter density $\d \rhom$ which preserve $\int_\tau d^3x\ \d \rhom=0$, i.e., the total number in volume $\tau$. The integral is over some region $\tau$, much smaller than the Hubble volume, over which the fluctuations are coherent. Such a fluctuation leads to a change in the total energy. The energy change proportional to $\d \rhom$ vanishes because of Eq.~\ref{eq:potmin} and the condition that $\int_\tau d^3x\ \d \rhom=0$. The energy change to next order is \bea \pert{E}= {1\over 2}\int d^3x (\pert{\rhom})^2\left({\partial\chi_0\over\partial\rhom}\right)\left( m_\chi^2 {\partial\chi_0\over\partial\rhom} +2\lambda g'(\chi_0)\right) \\ = -{1\over 2}\int d^3x (\pert{\rhom})^2\lambda^2{\left[g'(\chi_0)\right]^2 \over m_\chi^2 }~~~\eea Therefore the leading correction to the energy is always negative, implying that the configuration is unstable to the growth of these fluctuations. We dub this the AZK-instability. This instability was first noted in the context of the MaVaN scenario \cite{afshordi05}. We have neglected gravity and the expansion of the universe in the above analysis. We neglected gravity because the relevant length scales are much smaller than the Jeans length; the instability occurs on all scales and hence the effect is most severe on microscopic scales. The analysis above was thus for a region $\tau$ much smaller than that where gravity would be important. We neglected the expansion of the universe because the relevant time scales are much smaller than the age of the universe. In addition our setup started with a smooth distribution of matter. For this one must go to scales smaller than the free-streaming scale of dark matter particles. For example, the comoving free-streaming scale of a typical neutralino dark matter particle is of the order of parsec. We do not study this system on larger cosmologically relevant scales. It is, however, unlikely that the system will still able to drive the accelerated expansion of the universe since the generic AZK instability is intimately related to the adiabatic sound speed of the fluid \cite{afshordi05}. The result above assumes that the scalar field is much heavier than the expansion rate of the universe. This constraint is easy to satisfy and the large mass makes the model more robust to radiative corrections (for example, see \cite{fardon06}). Secondly, the calculation is only valid for modes which have a wavelength much larger than $1/m_\chi$; for shorter wavelengths, we cannot assume that the scalar field relaxes to the minimum quickly enough. \section{Avoiding the AZK-instability\label{sec:avoiding-instability}} To avoid this instability, we look at more general couplings. Consider now a model where the total energy is \bea E=\int d^3x\ \left[ V_0(\chi)+m \rhom+\lambda g(\chi)\rhom^n \right] \,,\eea and we choose $\lambda > 0$ without loss of generality. Again we assume that the scalar field tracks the minimum of the potential and hence we have, \bea V'_0(\chi_0)+\lambda g'(\chi_0)\rhom^n=0 \,, \\ (V''_0(\chi_0)+\lambda g''(\chi_0)\rhom^n)\left({\partial\chi_0\over\partial\rhom}\right)+\lambda g'(\chi_0)n\rhom^{n-1}=0\,.\eea Following our earlier calculation, we find \bea \pert{E}= {1\over 2}\int d^3x \left({\pert{\rhom} \over \rhom}\right)^2 \left(-{\left[n \lambda g'(\chi_0) \rhom^n \right]^2 \over m_\chi^2}\right. \nonumber \\ \left. +\lambda n(n-1)g(\chi_0)\rhom^n { \over }\right)\,,\eea Therefore, the instability is avoided if \bea -n^2\lambda^2\rhom^n {\left[g'(\chi_0)\right]^2 \over m_\chi^2} +n(n-1)\lambda g(\chi_0) > 0 \,.\label{eq:condition-on-n} \eea We note that the first term is always negative and gets large with $\rhom$ unless $g'(\chi_0)$ decreases fast enough. Looking at the second term we note that any value of $0 < n\leq 1$ is unstable independent of the form of $g(\chi)$ except for the requirement that $g(\chi_0) > 0$ which is required anyway for the potential to be bounded from below. A robust way to avoid the instability is to choose $n<0$, which makes the second term positive. This is, of course, not sufficient to guarantee the inequality in Eq.~\ref{eq:condition-on-n}. We need the magnitude of the second term to be larger than that of the first. This is easy to arrange. We again look at changes to the potential as we vary $\chi$ about $\chi_0$. If the potential is not fine-tuned to give rise to cancellations between terms in the Taylor expansion, then $n\lambda g' \rhom^n \pert{\chi} < V$ and also $m_\chi^2(\pert{\chi})^2/2 < V$. Putting these two expressions together yields $2n^2\lambda^2 (g')^2 \rhom^{2n}/m_\chi^2 < V \sim \lambda g \rhom^n$. Hence we see that it is natural, if $n<0$, for the inequality in Eq.~\ref{eq:condition-on-n} to be satisfied. It is also possible to avoid the instability by choosing $n>1$. However, this region of model space will be heavily constrained by observations. In situations where the matter density gets large, i.e., in collapsed structures, the last term in the potential dominates. It would make the dark energy density in galaxies large, change structure formation and clustering properties of dark matter halos. Therefore, these kinds of models would be tightly constrained. In order for these models to be viable, $\lambda$ would have to be small and the model would essentially be the same as that with two decoupled fluids. Thus the requirement of AZK-stability and observational constraints naturally lead us to consider models where $n<0$. We now look at observational consequences of such a coupling. \section{AZK-stability and Super-acceleration\label{sec:super-acceleration}} The coupling term above with $n<0$ introduces a very interesting effect: this model has super-acceleration. That is, observations will seem to show a phase with dark energy equation of state less than -1. To see this, we first note that the observational quantity that is important is the pressure. We will fit to the observations a model with matter that scales with the expansion as $1/a^{3}$, and dark energy with some equation of state $w_{\rm DE}$. Note that adding or removing a component of energy density that scales as $1/a^3$ does not change the pressure of the fluid. Hence very generally $P_{\rm tot}=P_{\rm DE}$. $P_{\rm tot}$ is defined by the equation $\dot{V}=-3H(V+P_{tot})$ from which we find $P_{tot}=-V_0(\chi_0)+\lambda g(\chi_0)\rhom^n(n-1)$. We set the equation of state $\wde \equiv P_{\rm tot}/(V-m\rhom)$ and find, \bea \wde =-1+{n\lambda g(\chi_0)\rhom^n\over V_0(\chi_0) +\lambda g(\chi)\rhom^n}\,.\eea Now since $n<0$, the second term is actually negative, and we have $w_{DE}<-1$ i.e. super-acceleration. We emphasize that this super-acceleration is {\it not} accompanied by any of the problems normally associated with theories with equation of state less than -1. There is no acausal behavior, and there are no ghosts. This is because the super-acceleration in our model results from an interaction which is ignored in the fitting of theory to observations. If we fit our observations using a canonical matter density term and dark energy, then the interaction has the effect of making the the effective equation of state for dark energy more negative. \section{Sound speed\label{sec:sound-speed}} Here we present an alternative derivation of the instability in terms of the sound speed of the combined fluid. A negative sound speed squared would signal instability. On length scales much larger than $m_\chi^{-1}$, the evolution of the system is adiabatic and hence the sound speed is \begin{equation}\label{eq:define-c} c_a^2 = {\dot{P}_{\rm tot} \over \dot{V}}\,. \end{equation} The adiabatic sound speed in this theory can then be expressed as \begin{eqnarray} c_a^2 & = & { \rhom \partial \wde / \partial \rhom + \wde(1+\wde) \over 1+\wde+m\rhom/(V-m\rhom)} \,.\label{eq:c-wde}\\ &=&{\rhom \over M} \left[ {\partial^2 V(\chi_0,\rhom) \over \partial \rhom^2} - m_\chi^2\left({\partial \chi_0 \over \partial \rhom}\right)^2 \right]\,,\label{eq:c-V}\\ & = & { \rhom \partial \wdem / \partial \rhom + \wdem(1+\wdem) \over 1+\wdem} \,,\label{eq:c-wdem} \end{eqnarray} where $\wdem \equiv P_{\rm tot}/V$ is the equation of state of the total fluid. For a universe with an accelerating expansion $\wdem < -1/3$. For a wide class of models with $\wdem<0$ and either the $\rhom \wdem'$ term sub-dominant or negative, we have $c_a^2 < 0$ and the system is unstable. This is just the AZK-instability. Lets now look in more detail at Eq. \ref{eq:c-wde}. First, consider the case where $\wde>-1$: the denominator is positive and if the $\wde'$ term is sub-dominant or negative, then AZK-instability sets in. It is clear that this instability may not be present in models with $\wde < -1$. We also note that this instability will likely set in well before the current epoch because at early times $\rhom/(V-\rhom) \gg 1$. For this case where $\wde(1+\wde) > 0$, the sign and magnitude of the $\rhom\wde'$ term is important. In particular, the requirement that the $\rhom\wde'$ term is sub-dominant may not be trivial to obtain \cite{linder06}. While the above derivation shows us how the instability arises, it does not provide us with an intuitive understanding of what happens to the matter. In order to better understand that we look at the Boltzmann equation for the matter coupled to a scalar field. The scalar field gives the matter a mass term that can vary spatially and temporally. Following AZK \cite{afshordi05}, we write down the Boltzmann equation for matter neglecting gravity and hence only valid on small scales. These are the scales of interest since we have assumed $m_\chi \gg H$. We write down the first order perturbations to this equation and expand the perturbations in plane wave modes. Denoting the effective mass of the matter particle by $M(\chi)$ we find, \bea \omega \pert{f}({\vec p},{\vec k}) &-& (\gamma M)^{-1}{\vec p}\cdot{\vec k}\pert{f}({\vec p},{\vec k}) \\ &+& \gamma^{-1}\pert{M}({\vec k}){\vec k}\cdot\nabla_{\vec p}f({\vec p})=0 \,.\label{eq:boltzmann} \eea We then find the perturbation to the matter density $\pert{\rhom}({\vec k})$ using the above equation. In the limit that matter is non-relativistic, the resulting equation has a simple form. We find that the variation in effective mass of the particle is given by $\pert{M}({\vec k}) = (M / \rhom) c_s^2 \pert{\rhom}({\vec k})$ where we have defined $c_s=\omega/k$, the sound speed of matter. The above equation is valid for perturbations $\pert{M}$ on all scales at which our assumptions hold. As pointed out in \cite{afshordi05}, there is no scale in the equation for $c_s^2$ because we are studying scales where it is correct to assume that the scalar field adjusts to changes in the matter density, and gravity is unimportant. We now turn to the fluid description and write $M=V(\chi_0,\rhom)/\partial \rhom$. Using Eq.~\ref{eq:dchi0dn} for $d\chi_0/d\rhom$, one may then obtain perturbations in $M$ as $\pert{M} = (M / \rhom) c_a^2 \pert{\rhom}$ where $c_a^2$ is given by Eq.~\ref{eq:c-V}. In the framework of a scalar degree of freedom coupled to matter, both descriptions must be valid and hence we find that $c_s^2=c_a^2$. The instability may therefore be analyzed in terms of $c_a^2$. All of our analyses in earlier sections go through if we work with $c_a^2$ and we conclude that models with super-acceleration provide a generic way to avoid the AZK instability. \section{Conclusions\label{sec:conclusions}} In this paper, we have explored the possibility that dark energy may interact with matter. Such a hypothesis is natural if the explanation for dark energy requires extra scalar degrees of freedom. Unfortunately, as we have shown here, these models suffer from a generic instability when the mass of the scalar field is very large. We have verified that this instability is also present in scalar-tensor theories where the scalar plays the role of dark energy, and also in models with multiple scalar fields. We then looked for models where this instability could be avoided, and found a large class of such models. Most interestingly, we found that in these models, the {\em apparent} equation of state of the dark energy density is generically smaller than -1. This super-acceleration is a result of the fact that we fit observations with models that have non-interacting matter and dark energy fluids. There is a theoretical prejudice against models of $\wde<-1$ due to their apparent theoretical problems. The observational data certainly do not disfavor $\wde< -1$. Indeed a large region of the parameter space allowed by SNIa observations corresponds to a constant $\wde < -1$. Here we have shown that stable models with $\wde< -1$ may be constructed without encountering ghosts or acausal behavior. These models are no more fine-tuned than quintessence models. Thus theoretical bias against $\wde <-1$ should be treated with circumspection, and not be given any weight when interpreting observational data.
Title: The ultra-cool white dwarf companion of PSR J0751+1807
Abstract: We present optical and near-infrared observations with Keck of the binary millisecond pulsar PSR J0751+1807. We detect a faint, red object - with R=25.08+-0.07, B-R=2.5+-0.3, and R-I=0.90+-0.10 - at the celestial position of the pulsar and argue that it is the white dwarf companion of the pulsar. The colours are the reddest among all known white dwarfs, and indicate a very low temperature, Teff~4000 K. This implies that the white dwarf cannot have the relatively thick hydrogen envelope that is expected on evolutionary grounds. Our observations pose two puzzles. First, while the atmosphere was expected to be pure hydrogen, the colours are inconsistent with this composition. Second, given the low temperature, irradiation by the pulsar should be important, but we see no evidence for it. We discuss possible solutions to these puzzles.
https://export.arxiv.org/pdf/astro-ph/0601205
\title{The ultra-cool white dwarf companion of PSR~J0751+1807} \author{C. G. Bassa\inst{1} \and M. H. van Kerkwijk\inst{2} \and S. R. Kulkarni\inst{3}} \institute{Astronomical Institute, Utrecht University, PO Box 80\,000, 3508 TA Utrecht, The Netherlands\\ \email{c.g.bassa@astro.uu.nl} \and Department of Astronomy and Astrophysics, University of Toronto, 60 Saint George Street, Toronto, ON M5S 3H8, Canada \and Palomar Observatory, California Institute of Technology 105-24, Pasadena, CA 91125, USA% } \offprints{C.G. Bassa} \date{Received / Accepted} \abstract{We present optical and near-infrared observations with Keck of the binary millisecond pulsar PSR~J0751+1807. We detect a faint, red object -- with $R=25.08\pm0.07$, $B-R=2.5\pm0.3$, and $R-I=0.90\pm0.10$ -- at the celestial position of the pulsar and argue that it is the white dwarf companion of the pulsar. The colours are the reddest among all known white dwarfs, and indicate a very low temperature, $T_\mathrm{eff}\approx4000$\,K. This implies that the white dwarf cannot have the relatively thick hydrogen envelope that is expected on evolutionary grounds. Our observations pose two puzzles. First, while the atmosphere was expected to be pure hydrogen, the colours are inconsistent with this composition. Second, given the low temperature, irradiation by the pulsar should be important, but we see no evidence for it. We discuss possible solutions to these puzzles. \keywords{Pulsars: individual (\object{PSR~J0751+1807}) -- binaries: close -- stars: neutron -- white dwarfs} } \section{Introduction}\label{sec:intro} Among the pulsars in binaries, the largest group, the low-mass binary pulsars, has low-mass white-dwarf companions. Before the companions became white dwarfs, their progenitors filled their Roche lobe and mass was transferred to the neutron stars, thereby spinning them up and decreasing their magnetic fields. Considerations of the end of this stage, where the white dwarf progenitor's envelope becomes too tenuous to be supported further, allow one to make predictions for relations between the orbital period and white dwarf mass, and orbital period and eccentricity (for a review, e.g., \citealt{pk94,sta04}). Furthermore, after the cessation of mass transfer, two clocks will start ticking at the same time: the neutron star, now visible as a millisecond pulsar, will spin down, while the secondary will contract to a white dwarf and start to cool. Consequently, the spin-down age of the pulsar should equal the cooling age of the white dwarf. From optical observations of white-dwarf companions to millisecond pulsars one can estimate the white-dwarf cooling age and compare it with the pulsar spin-down age. Initial attempts to do this \citep{hp98a,hp98b,sdb00} revealed a dichotomy in the cooling properties of white dwarfs in the sense that some white dwarf companions to older pulsars have cooled less than those of younger pulsars. In particular, the companions of PSR~J0437$-$4715 \citep{dbv93,sbb+01} and PSR~B1855+09 \citep{kbkk00,rt91} have temperatures of about 4000--5000\,K, with characteristic pulsar ages of 5\,Gyr. This is in contrast to the companion of PSR~J1012+5307 \citep{llfn95,kbk96,cgk98}, which has a higher temperature (8600\,K), while it orbits an older pulsar (8.9\,Gyr). A likely cause for this dichotomy is the difference in the thickness of the envelope of hydrogen surrounding the helium core of the white dwarf \citep{ashp96}. After the cessation of mass transfer, the white dwarfs have relatively thick ($\sim\!10^{-2}$\,M$_\odot$) hydrogen envelopes which are able to sustain residual hydrogen shell-burning, keeping the white dwarf hot and thereby slowing the cooling \citep{dsbh98}. The shell burning, however, can become unstable and lead to thermal flashes which can reduce the mass of the envelope. White dwarfs with such reduced, relatively thin ($\la10^{-3}$\,M$_\odot$) hydrogen envelopes cannot burn hydrogen and, as a result, cool faster. The transition between thick and thin hydrogen envelopes was predicted to lie near 0.18--0.20\,M$_\odot$ (where heavier white dwarfs have thin envelopes; \citealt{ashp96,seg00,asb01}). Until recently, PSR~J1012+5307, with an orbital period $P_\mathrm{b}=0.60$\,d, was the only system for which a thick hydrogen envelope was required to match the two timescales. Given the relation between the white dwarf mass and the orbital period \citep{jrl87,rpj+95,ts99}, companions in similar or closer orbits should have similar or lower mass, and thus have thick hydrogen envelopes as well. This was confirmed by the recent discovery of two new, nearby, binary millisecond pulsars with orbital periods similar to that of PSR~J1012+5307; PSR~J1909$-$3744 (1.53\,d, \citealt{jhb+05}) and PSR~J1738+0333 (0.354\,d, Jacoby et al., in prep.; see \citealt{kbjj05} for preliminary results). For both, the temperatures and characteristic ages are similar to those of PSR J1012+5307, and thus one is led to the same need for a thick hydrogen envelope. These discoveries, combined with the thin envelopes inferred for PSR~J0034$-$0534 (1.59\,d) and binaries with longer periods, suggest that the transition occurs at a mass that corresponds to an orbital period just over 1.5\,d (\citealt{kbjj05}). All systems with shorter orbital periods should have thick hydrogen envelopes. The two known millisecond pulsars with white dwarf companions that have shorter orbital periods than PSR~J1012+5307 but do not have optical counterparts, are PSR~J0613$-$0200, with a 1.20\,d period, and PSR~J0751+1807, which has the shortest orbital period of all binary millisecond pulsars with $M_\mathrm{c}>0.1$\,M$_\odot$ companions, 0.26\,d \citep{lzc95}. The latter system is of particular interest because the companion mass has been determined from pulse timing ($M_\mathrm{WD}=0.19\pm0.03$\,M$_\odot$ at 95\% confidence; \citealt{nss+05}), so that one does not have to rely on the theoretical period-mass relationship. Intriguingly, for PSR J0751+1807, optical observations from \citet{lcf+96} set a limit to the temperature of 9000\,K, which is only marginally consistent with it having a thick hydrogen envelope. Based on this, \citet{esa01}, suggested the hydrogen envelope may have been partially lost due to irradiation by the pulsar. The faintness of the companion to PSR~J0751+1807 aroused our curiosity and motivated us to obtain deep observations to test the theoretical ideas discussed above. We describe our observations in Sect.~\ref{sec:observations}, and use these to determine the temperature, radius and cooling history in Sect.~\ref{sec:tandr}. In Sect.~\ref{sec:irradiation}, we investigate irradiation by the pulsar, finding a surprising lack of evidence for any heating. We discuss our results in Sect.~\ref{sec:discussion}. \section{Observations and data reduction}\label{sec:observations} The PSR~J0751+1807 field was observed with the 10~meter Keck I and II telescopes on Hawaii on five occasions. On December 11, 1996 the Low Resolution Imaging Spectrometer (LRIS, \citealt{occ+95}) was used to obtain $B$ and $R$-band images, while the Echellette Spectrograph and Imager (ESI, \citealt{sbe+02}) was used on December 21, 2003 to obtain deeper $B$ and $R$-band, as well as $I$-band images. The $R$-band filter used that night was the non-standard ``Ellis $R$'' filter. The observing conditions during the 1996 night were mediocre, with 0\farcs8--1\farcs1 seeing and some cirrus appearing at the end of the night. The conditions were photometric during the 2003 night, and the seeing was good, 0\farcs6--0\farcs8. The third and fourth visit were with LRIS again, now at Keck I, on January 7 and 8, 2005. The red arm of the detector was used to obtain $R$-band images. The seeing on the first night in 2005 was rather bad, about 1\farcs5 and improved to about 1\farcs0 on the second night. The conditions on these nights were not photometric. Finally, a series of 36 dithered exposures, each consisting of 5 co-added 10\,s integrations, were taken through the $K_\mathrm{s}$ filter with the Near Infrared Camera (NIRC; \citealt{ms94}) on January 26, 2005. The conditions were photometric with 0\farcs6 seeing. Standard stars (\citealt{lan92,ste00}) were observed in 1996 and 2003, while a 2MASS star \citep{csd+03} in the vicinity of PSR~J0751+1807 was observed to calibrate the NIRC data. A log of the observations is given in Table~\ref{tab:obs}. The images were reduced using the Munich Image Data Analysis System (MIDAS). The $BRI$ images were bias-subtracted and flat-fielded using dome flats. The longer exposures in each filter were aligned using integer pixel offsets, and co-added to create average images. The near-infrared images were corrected for dark current using dark frames with identical exposure times and number of co-adds as those used for the science frames. Next, a flatfield frame was created by median combining the science frames. After division by this flatfield, the science frames were registered using integer pixel offsets and averaged. \begin{table} \begin{minipage}[t]{\columnwidth} \centering \caption[]{Observation log.}\label{tab:obs} \renewcommand{\footnoterule}{} \begin{tabular}{l@{\hspace{0.5cm}} c@{\hspace{0.5cm}} c@{\hspace{0.5cm}} c@{\hspace{0.5cm}} c@{\hspace{0.5cm}} } \hline Field & Time (UT) & Filter & $t_\mathrm{int}$ (s) & $\sec z$ \\ \hline \multicolumn{5}{l}{December 11, 1996, LRIS} \\[0.2ex] SA\,95 & 08:23--08:25 & $R$ & $2+10$ & 1.07 \\ & 08:27--08:29 & $B$ & $2+10$ & 1.07 \\[0.1ex] SA\,95 & 09:28--09:31 & $B$ & $2+10$ & 1.07 \\ & 09:33--09:35 & $R$ & $2+10$ & 1.08 \\[0.1ex] PSR~J0751+1807 & 09:45 & $R$ & $10$ & 1.39 \\ & 09:47--09:59 & $R$ & $2\times300$ & 1.36 \\ & 10:01 & $R$ & $600$ & 1.31 \\ & 10:13 & $B$ & $600$ & 1.26 \\[0.8ex] \multicolumn{5}{l}{December 21, 2003, ESI} \\[0.2ex] PSR~J0751+1807 & 10:06--10:27 & $R$ & $3\times360$ & 1.14 \\ & 10:29--10:57 & $I$ & $6\times240$ & 1.08 \\ & 11:00--11:33 & $B$ & $3\times600$ & 1.04 \\[0.1ex] NGC\,2419 & 11:40 & $B$ & $10+30$ & 1.06 \\ & 11:44 & $R$ & $10+30$ & 1.06 \\ & 11:47 & $I$ & $10+30$ & 1.06 \\[0.8ex] \multicolumn{5}{l}{January 7, 2005, LRIS} \\[0.2ex] PSR~J0751+1807 & 11:54--12:53 & $R$ & $5\times600$ & 1.05 \\[0.8ex] \multicolumn{5}{l}{January 8, 2005, LRIS} \\[0.2ex] PSR~J0751+1807 & 11:42--12:51 & $R$ & $6\times600$ & 1.05 \\[0.8ex] \multicolumn{5}{l}{January 26, 2005, NIRC} \\[0.2ex] PSR~J0751+1807 & 08:06--08:56 & $K_\mathrm{s}$ & $36\times50$ & 1.07 \\ 2MASS star\footnote{2MASS\,J07510621+1807253} & 08:59 & $K_\mathrm{s}$ & $0.4$ & 1.02\\ \hline \end{tabular} \end{minipage} \end{table} \subsection{Astrometry} For the astrometric calibration, we selected 14 stars from the second version of the USNO CCD Astrograph catalogue (UCAC2; \citealt{zuz+04}) that overlapped with the 10\,s $R$-band LRIS image of December 1996. Of these, 11 were not saturated and appeared stellar and unblended. The centroids of these objects were measured and corrected for geometric distortion using the bi-cubic function determined by J.~Cohen (1997, priv.\ comm.)\footnote{http://alamoana.keck.hawaii.edu/inst/lris/coordinates.html}. We fitted for zero-point position, plate scale and position angle. The inferred uncertainty in the single-star measurement of these 11 stars is $0\farcs057$ and $0\farcs083$ in right ascension and declination, respectively, and is consistent with expectations for the UCAC measurements of approximately $0\farcs020$ for stars of 14th magnitude and $0\farcs070$ for stars 2 magnitudes fainter. This solution was transferred to the 10\,min $R$-band LRIS image using 91 stars that were present on both images and were stellar, unsaturated and not blended. Again the zero-point position, plate scale and position angle were left free in the fit and the final residuals were $0\farcs016$ and $0\farcs019$ in right ascension and declination. The UCAC is on the International Celestial Reference System (ICRS) to $\la\!0\farcs01$, and hence the final systematic uncertainty with which our coordinates are on the ICRS is dominated by our first step, and is $\sim\!0\farcs03$ in each coordinate. Our images, with the position of PSR~J0751+1807 \citep{nss+05} indicated, are shown in Figure~\ref{fig:fc}. On the 10\,min LRIS $R$-band images from 1996 and 2005, we find a faint object, hereafter star X, at the position of the pulsar. It is also, though marginally, present in the two 5\,min $R$-band images from 1996, but not detected in the 10\,min $B$-band LRIS image of that observing run. Star X is clearly present in the 2003 ESI $R$ and $I$-band images, and marginally in the $B$-band image. It is not detected in the near-infrared observations (Fig.~\ref{fig:fc}). Positions for star X and other objects inferred using the astrometry of the 10\,min LRIS $R$-band image are listed in Table~\ref{tab:phot}. The pulsar position at the time of the 1996 LRIS observation, using the \citet{nss+05} position and proper motion, is $\alpha_\mathrm{J2000}=07^\mathrm{h}51^\mathrm{m}09\fs1574(1)$, $\delta_\mathrm{J2000}=+18\degr07\arcmin38\farcs624(10)$. We find that star X is offset from the pulsar position by $-0\farcs01\pm0\farcs06$ in right ascension and $0\farcs04\pm0\farcs06$ in declination, well within the $1\sigma$ uncertainties (including those on the pulsar position). Given the low density of about 47 stars per square arcminute and the excellent astrometry, the probability of a chance coincidence in the 95\% confidence error circle, which has a radius of $0\farcs24$, is only 0.1--0.2\%. Since, as we will see, it is hard to envisage how the companion could be fainter than the object detected, we are confident that star X is the companion of PSR~J0751+1807. \subsection{Photometry}\label{sec:photometry} The DAOPHOT II package (\citealt{ste87}), running inside MIDAS, was used for the photometry on the averaged images. We followed the recommendations of \citet{ste87}: instrumental magnitudes were obtained through point spread function (PSF) fitting and aperture photometry on brighter stars was used to determine aperture corrections. For the calibration of the optical images, instrumental magnitudes of the standard stars, determined using aperture photometry, were compared against the values of \citet{ste00}. We used the standard Keck extinction coefficients of 0.17, 0.11 and 0.07\,mag per airmass for $B$, $R$ and $I$, respectively. Colour terms were not required for the LRIS $B$ and $R$ bands, but were significant for the ESI bands: $0.107 (B-R)$ for $B$, $0.083 (B-R)$ for $R$, and $-0.004 (R-I)$ for $I$, i.e., the ESI $B$, $R$ are redder than the standard bands, while ESI $I$ is slightly bluer. The root-mean-square residuals of the ESI calibrations are about 0.05\,mag in $B$, and 0.03\,mag in $R$ and $I$, while those of the LRIS calibration are 0.08\,mag in $B$ and 0.05\,mag in $R$; we adopt these as the uncertainty in the zero-points. The near-infrared observations were calibrated through aperture photometry with 1\farcs5 (10\,pix) apertures using the 2MASS star, fitting for a zero-point only, as the difference in airmass between the science and calibration images is small. We adopt an uncertainty in the $K\mathrm{s}$ zero-point of 0.1\,mag. Calibrated ESI magnitudes for star X and selected other stars in the field are listed in Table~\ref{tab:phot}. Star X is barely above the detection limit of the ESI $B$-band observations, hence the large error. It is not detected in the LRIS $B$-band and the NIRC $K_\mathrm{s}$-band observations, and, scaling from the magnitude of a star with a signal-to-noise ratio of about 10 and 6, we estimate the 3$\sigma$ detection limits at $B=26.8$ and $K_\mathrm{s}=21.3$, respectively. The former is consistent with the ESI detection. None of the stars in Table~\ref{tab:phot} are covered by the small $38\arcsec\times38\arcsec$ field-of-view of NIRC, hence we do not have near-infrared magnitudes for these. The 1996 LRIS $R$-band magnitude is $25.13\pm0.11$, which is consistent with the ESI measurement. Since the conditions during the 1996 LRIS observations may not have been photometric, however, this may be a coincidence. To check for variability, we tied the instrumental LRIS $R$ band magnitudes directly to the ESI $R$ and $I$ ones, using 38 stars that both images had in common and that had magnitude uncertainties below 0.1 mag. As expected given the non-standard ``Ellis $R$'' filter on ESI, we required a large colour term, $-0.302(R_\mathrm{inst}-I_\mathrm{inst})$, but with this the fit was adequate, with root-mean-square residuals of 0.14 mag. Compared to the fit, the ESI minus LRIS difference in $R$-band magnitude is insignificant, $-0.03\pm0.13$\,mag. Similarly, comparing instrumental $R$-band magnitudes from 2005 January 7 with those taken 2005 January 8 and 1996 December 11, fitting for an offset only, results in magnitude differences of $0.03\pm0.07$ and $-0.16\pm0.12$\,mag, respectively. Thus, no large variations in brightness are seen; we will see in Sect.~\ref{sec:irradiation} that this is somewhat surprising. \begin{table} \caption[]{LRIS Astrometry and ESI photometry of the companion of PSR~J0751+1807 and stars in the field. The nomenclature of the stars is according to Fig.~\ref{fig:fc}. The uncertainties listed in parentheses are instrumental, i.e., they do not include the zero-point uncertainty in the astrometric tie (about $0\farcs03$ in each coordinate) or of photometric calibration (0.05 mag in $B$ and 0.03 mag in both $R$ and $I$).} \label{tab:phot} \begin{tabular} {l@{\hspace{0.15cm}} l@{\hspace{0.15cm}} l@{\hspace{0.15cm}} l@{\hspace{0.15cm}} l@{\hspace{0.15cm}} l@{\hspace{0.15cm}} } \hline \hline ID & \multicolumn{1}{c}{$\alpha_\mathrm{2000}$} & \multicolumn{1}{c}{$\delta_\mathrm{2000}$} & \multicolumn{1}{c}{$B$\phantom{0}} & \multicolumn{1}{c}{$R$\phantom{0}} & \multicolumn{1}{c}{$I$\phantom{0}} \\ & $\phantom{00}^\mathrm{h}\phantom{00}^\mathrm{m}\phantom{00}^\mathrm{s}$ &$\phantom{00}\degr\phantom{00}\arcmin\phantom{00}\arcsec$ & & & \\ \hline X & 07 51 09.158(4) & 18 07 38.66(6) & 27.56(25) & 25.08(7) & 24.18(7) \\[0.2em] A & 07 51 09.933(1) & 18 07 05.97(1) & 21.73(1) & 19.30(1) & 18.31(1) \\ B & 07 51 10.844(1) & 18 07 52.91(1) & 22.80(1) & 21.03(1) & 20.32(1) \\ C & 07 51 10.891(1) & 18 07 35.69(1) & 24.30(2) & 21.81(1) & 20.63(1) \\ D & 07 51 10.739(1) & 18 07 32.79(1) & 24.28(6) & 22.50(5) & 21.99(6) \\ E & 07 51 08.519(1) & 18 07 59.89(2) & 24.56(7) & 22.87(5) & 22.38(8) \\ F & 07 51 08.859(2) & 18 07 08.83(3) & 24.94(4) & 24.00(5) & 23.29(4) \\ G & 07 51 08.908(4) & 18 07 35.71(5) & 25.65(8) & 24.51(5) & 23.85(6) \\ H & 07 51 10.691(3) & 18 07 24.69(6) & 25.69(7) & 24.94(9) & 24.34(8) \\ \hline \end{tabular} \end{table} \section{Temperature, radius, and cooling history}\label{sec:tandr} We use our observations of star X, the companion of PSR~J0751+1807, to constrain its temperature, radius, and atmospheric constituents, and discuss our result that the white dwarf does not have the expected thick hydrogen envelope. \subsection{Colours, temperature, and atmospheric composition} We first use the colours of star X to constrain its temperature. The red colours are largely intrinsic, as the maximum reddening towards PSR~J0751+1807 ($l=202.73$, $b=21.09$) is small, $E_{B-V}=0.05\pm0.01$ \citep{sfd98}. This value is consistent with the low value found for the interstellar absorption $N_\mathrm{H}\sim4\times10^{20}$\,cm$^{-2}$, as estimated from {\it ROSAT} X-ray observations of PSR~J0751+1807 by \citet{btl+96}. For comparison, the relation by \citet{ps95} predicts an $N_\mathrm{H}\approx3\times10^{20}$\,cm$^{-2}$ for the above reddening. Given the distance of $\sim\!0.6\,$kpc \citep{nss+05}, we expect most of the reddening to be in the foreground to the pulsar. Hence, the dereddened colours are $(B-R)_0=2.40\pm0.27$ and $(R-I)_0=0.86\pm0.10$. In Fig.~\ref{fig:tcd}a, we compare the intrinsic colours of star X with those of other white-dwarf companions of millisecond pulsars, other white dwarfs, and models. We find that the colours of star X are the reddest for any known millisecond pulsar companion or white dwarf. The pulsar companion that comes closest is that of PSR~J0437$-$4715 ($B-R=2.12\pm0.06$, $R-I=0.56\pm0.02$ [\citealt{dbv93}] and negligible extinction\footnote{As inferred from the dust maps of \cite{sfd98}; \cite{dbv93} estimate $E_{B-V}=0.07$ from the work of \cite{knu79}.}); the most similar white dwarf is WD~0346+246 ($B-R=2.2\pm0.1$, $R-I=0.76\pm0.08$, \citealt{osh+01}). Thus, star X is likely as cool or even cooler than the $T_{\rm eff}\simeq3700\,$K inferred for those two sources (PSR~J0437$-$4715: \citealt{dbv93}; Hansen 2002, priv.\ comm.; WD~0346+246: \citealt{osh+01,ber01}). Also shown in Fig.~\ref{fig:tcd}a are colours expected from model atmospheres of \citet{sarb01} and of Hansen (2004, priv.\ comm.), which are specifically tailored to the low-mass, helium-core companions of millisecond pulsars, as well as those for updated low-gravity ($\log g=7$), pure hydrogen atmosphere models\footnote{For updated versions of the \citet{bwb95} models, see http://www.astro.umontreal.ca/\~{}bergeron/CoolingModels/} of \citet{bwb95}. One sees that the colours of the companion of PSR~J0437$-$4715, as well as those of the hotter companions of PSR~J1012+5307 and J0218+4232, are consistent with these models. For star X, however, the colours are not consistent, as the models never venture redwards of $R-I\approx0.7$ and $B-R\approx2.0$. The change in direction of the tracks is seen in all models for hydrogen-rich, metal-free atmospheres; it reflects a change in the dominant source of opacity, from bound-free absorption of H$^-$ at higher temperatures to collision-induced absorption of H$_2$ at lower ones (\citealt{lcs91,sbl+94,han98}). The latter process is highly non-grey, and leads to absorption predominantly longward of the $R$-band. As a result, the $R-I$ colour becomes bluer with decreasing temperatures, while $B-R$ remains roughly constant. Could star X have a different composition? Due to the high gravity of white dwarfs, metals settle out of the atmosphere. However, some white dwarfs have atmospheres dominated not by hydrogen, but by helium. For the latter, the opacity sources are all fairly grey, and hence the colours continue to redden with decreasing temperatures. Indeed, the colours of star X are consistent with the predictions of the updated $\log g=7$ pure helium models after \citet{bwb95} at $T_\mathrm{eff}\simeq4200$\,K (Fig.~\ref{fig:tcd}a). From an evolutionary perspective, however, a pure helium atmosphere is not expected. Low-mass white dwarfs such as the companions to millisecond pulsars are all formed from low-mass stars whose evolution was truncated by mass transfer well before helium ignition (for recent models, see \citealt{ts99,ndm04}). As a result, they should have helium cores surrounded by relatively thick, 0.01 to 1\% of the mass, hydrogen envelopes (\citealt{dsbh98,asb01}). Indeed, among the low-mass white-dwarf companions to pulsars \citep{kbjj05} as well as among low-mass white dwarfs in general \citep{blr01}, only hydrogen-dominated atmospheres have been observed. In principle, at low temperatures, the hydrogen envelope might become mixed in with the helium core. Even if fully mixed, however, the remaining amounts of hydrogen would strongly influence the spectrum. Indeed, the effects of collision-induced absorption {\em increase} with increasing helium abundance up to $N({\rm He})/N({\rm H})\simeq10^5$ (\citealt{bl02}). From Fig.~\ref{fig:tcd}a, it is clear that the predictions for hydrogen-dominated atmospheres are also a somewhat poor match to the colours of the cooler normal white dwarfs with hydrogen in their atmospheres (as inferred from absorption at H$\alpha$, \citealt{blr01}; filled circles in the figure). For most, this appears to be due to missing blue opacity in the models (see \citealt{blr01} for a detailed study); the visual through infrared fluxes are reproduced well by the models, and show unambiguously that collision-induced absorption by H$_2$ is important. Indeed, the absorption is evident in the optical colours of some objects, in particular LHS~3250 (shown in Fig.~\ref{fig:tcd}) and SDSS~J133739.40+000142.8 (\citealt{bl02} and references therein). For our purposes, however, the case of the ultra-cool white dwarf WD~0346+246 is most relevant. For this source, the colours cannot be reproduced with either pure hydrogen or helium, but require a mixed atmosphere, dominated by helium (with fractional hydrogen abundances ranging from $10^{-9}$ to $10^{-1}$, depending on assumptions about the contribution of other opacity sources; \citealt{osh+01,ber01}, though recent work puts these abundances in to doubt, P.\ Bergeron 2005, priv.\ comm.). For all cases, the temperature is around 3700\,K. The similarity in the colours of WD~0346+246 and star~X would suggest that star~X has a similar, maybe slightly lower, temperature. From the above, we find that we cannot determine the temperature of the companion of PSR~J0751+1807 with certainty, since we do not know its composition. Most likely, however, it is somewhere between the temperature inferred for WD~0346+246 and that indicated by the (pure helium) models, i.e.\ in the range of, say 3500--4300\,K. A more stringent test could be provided by the near-infrared observations, as the $R-K$ colour (which is similar to $R-K_\mathrm{s}$) differs for different predictions. At a temperature of 4000\,K the $\log g=7$ \citet{bwb95} models predict $R-K$ colours of 2.7 and 1.6 for pure helium and pure hydrogen atmospheres, respectively. For the same temperature, $R-K=1.6$ is predicted by the 0.196\,M$_\odot$ model by \citet{sarb01}. Finally, for WD~0346+246, with presumably a mixed hydrogen/helium atmosphere, \citet{osh+01} observed $R-K=-0.7$. Unfortunately, our near-infrared observations only limit the colour to $R-K<3.8$, which does not constrain any of these predictions. \subsection{Brightness, distance and radius} So far, we have only discussed the colours and temperature. We now turn to the absolute magnitude and radius. In Fig.~\ref{fig:cmd}b, we show $M_R$ as a function of $R-I$. For star~X, we computed the absolute $R$-band magnitude $M_R$ using the parallax of $\pi=1.6\pm0.8$\,mas as measured through radio timing \citep{nss+05}. The resulting distance of $0.6^{+0.6}_{-0.2}$\,kpc is consistent with that estimated from the dispersion measure which predicts $1.1\pm0.2$\,kpc, using a dispersion measure of $30.2489\pm0.003$\,pc\,cm$^{-3}$ \citep{nss+05} and the recent model of the Galactic electron distribution of \citet{cl02}. Correcting for the reddening, this implies $M_R=15.97^{+0.88}_{-1.51}$. Given the similarities in the above absolute magnitude of star~X and that of WD~0346+246 ($M_R=16.1\pm0.3$; \citealt{hsh+99,osh+01}), and assuming similar temperature, one finds that the radius of star~X should be comparable to the $R=0.010$\,R$_\odot$ for WD~0346+246 \citep{ber01}. However, the large uncertainty in the parallax of PSR~J0751+1807 allows radii between 0.007--0.021\,R$_\odot$. For the white-dwarf mass of $\sim\!0.19$\,M$_\odot$ inferred from pulse timing \citep{nss+05}, this is consistent the $\sim\!0.022$\,R$_\odot$ expected from the 0.196\,M$_\odot$ model by \citep{sarb01}. As can be seen in Fig.~\ref{fig:cmd}, the absolute magnitude is also consistent with the predicted values from the $\log g=7$ pure helium model by \citet{bwb95}. At a temperature of $T_\mathrm{eff}=4250$\,K, this model has a radius of 0.020\,R$_\odot$ and a mass of 0.15\,M$_\odot$, somewhat smaller than the observed 0.19\,M$_\odot$. To correct for the small difference in mass, we computed white dwarf radii for the observed temperature and mass of the companion and used these to scale the absolute magnitudes of the pure helium track in Fig.~\ref{fig:cmd}. At 0.19\,M$_\odot$ and $T_\mathrm{eff}=4000$\,K, the \citet{pab00} helium core white dwarf mass-radius relation predicts 0.021\,R$_\odot$. This is very similar to the radius predicted by the \citet{bwb95} $\log g=7$ pure helium models, and as such, the absolute magnitudes are comparable. We conclude that, with in the large uncertainties on the parallax distance, the absolute magnitude and radius that we derive for the companion of PSR~J0751+1807 are consistent with the predictions for a pure helium atmosphere. We note that of the models presented in Fig.~\ref{fig:cmd}, those of \citet{bwb95} have been extensively tested to explain the population of nearby white dwarfs \citep{blr01,ber01,bl02} and use a very detailed description of the white dwarf atmosphere combined with the latest opacities (P.\,Bergeron 2005, priv.\ comm.). This is not the case for the models of Serenelli et al.\ and Hansen, and thus we should be careful in using their models quantitatively. Indeed, as can be seen from Fig.~\ref{fig:cmd}, their models do not reproduce the observations of cool white dwarfs well. For instance, for the companion of PSR~J0437$-$4715, which has a well-determined mass of $0.236\pm0.017$\,M$_\odot$ and distance of $139\pm3\,$pc \citep{sbb+01}, the models of \cite{sarb01}, while consistent with the observed $B-R$ and $R-I$ well, do not reproduce $R-I$ and $M_R$ simultaneously. In contrast, the $0.2$\,M$_\odot$ model of Hansen (2004, priv.\ comm.) does pass through the $R-I$, $M_R$ point, but cannot reproduce both colours. It may be that both problems reflect uncertainties in the model atmospheres used by Hansen and \citet{sarb01}. It would be worthwhile to couple the evolutionary models of these authors with the updated, very detailed atmospheric model of \citet{bwb95}. \subsection{Cooling history and nature of the envelope}\label{sec:cooling} Despite the uncertainty in the models and in the composition of the atmosphere, our observations show that the companion of PSR~J0751+1807 has cooled much more than expected if the amount of hydrogen was thick enough for significant residual nuclear burning (Sect.~\ref{sec:intro}). Indeed, the temperature is as expected if no residual hydrogen burning occurred. For instance, at the characteristic age of the pulsar, $\tau=7.1$\,Gyr \citep{nss+05}, the 0.196\,M$_\odot$ of \citet{sarb01}, which has a thin envelope, predicts a temperature of about 3200\,K, which is roughly consistent with what is observed. With a pure helium atmosphere, a slightly colder temperature, of $\sim\!2500$\,K, is expected, though this is a less secure estimate due to uncertainties in the opacities \citep{hp98a} The presence of a thin (or no) hydrogen envelope is not expected, however, since thick envelopes are inferred for other optically identified companions in short-period systems (see Sect.~\ref{sec:intro}). What could be wrong with this expectation? It was based on two theoretical ideas: (i) that below a certain critical mass, no shell flashes occur and hydrogen layers will be thick; and (ii) that the companion mass monotonously increases with increasing orbital period. These assumptions appeared to be confirmed by the available data: for PSR~J0751+1807, with a period of 0.26\,d, the companion mass of 0.16--0.21\,M$_\odot$ (95\% conf.; \citealt{nss+05}) is similar to what is found for two other short-period systems with companions for which thick hydrogen envelopes are inferred, and less than the masses for longer period systems with thin-envelope companions. Specifically, PSR~J1012+5307 (0.60\,d, 0.12--0.20\,M$_\odot$) and PSR~J1909$-$3744 (1.53\,d, 0.19--0.22\,M$_\odot$) have thick envelopes while PSR~J0437$-$4715 (5.74\,d, 0.20--0.27\,M$_\odot$) and PSR~B1855+09 (12.33\,d, 0.24--0.29\,M$_\odot$) have thin envelopes (see Fig.~\ref{fig:pbm2} and \citealt{kbjj05} and reference therein). Thus, while the uncertainties do not exclude that the companion of PSR~J0751+1807 is so massive that it its envelope was diminished by shell flashes, the existing data make it unlikely. Two explanations for a thin envelope remain. First, there may be differences in metallicity among the progenitors of pulsar companions. \citet{sarb02} studied the evolution of low-mass pulsar companions with sub-solar metallicity and found that, since the thermonuclear flashes are induced by the reactions of the CNO-cycle, the threshold mass between thin and thick hydrogen envelopes increases with decreasing metallicity of the white dwarf progenitor. Thus, it may be that the companion of PSR~J0751+1807 had a sufficiently higher metallicity that it was above the threshold for shell flashes, while companions in other short-period systems had lower metallicity and hence were below the threshold, despite having higher masses. The next possibility is that the white dwarf was indeed formed with a thick envelope, which was subsequently removed by an action other than shell flashes. Based on the upper limit on the temperature of \citet{lcf+96}, \citet{esa01} already argued that the pulsar companion could not have the thick hydrogen envelope, and they proposed a scenario where part of the envelope was removed by pulsar irradiation. \citeauthor{esa01} found that irradiation driven mass-loss could remove as much as 0.01\,M$_\odot$ from the thick hydrogen envelope (mostly while the companion is contracting following the cessation of mass transfer). A possible problem with the above suggestions, is that none predict the removal of the entire hydrogen envelope, while the observed colours seem most consistent with a pure helium or at least helium-dominated atmosphere. \section{Irradiation by the pulsar?}\label{sec:irradiation} Above, we have treated the companion as if it were an isolated object rather than member of a binary system. Might the presence of a relatively energetic pulsar influence our observations? The observed pulsar period and period derivative imply a spin-down luminosity $L_\mathrm{SD}=(2 \pi)^2 I \dot{P}/P^3=7.5\times10^{33}\,I_{45}$\,ergs\,s$^{-1}$ \citep{lzc95,nss+05}, where $I=10^{45}\,I_{45}{\rm\,g\,cm^2}$ is the pulsar moment of inertia. For a $2.1$\,M$_\odot$ pulsar and a $0.19$\,M$_\odot$ companion, the orbital separation is $a=2.3$\,R$_\odot$, and, consequently, the irradiative flux of the pulsar wind incident on the companion is $f_{\rm irr}=2.1\times10^{10}\,I_{45}$\,erg\,s$^{-1}$\,cm$^{-2}$. This is about twice the flux of the companion itself, $f_{\rm th}=\sigma T_\mathrm{eff}^4=1.06\times10^{10}$\,erg\,s$^{-1}$\,cm$^{-2}$ for $T_\mathrm{eff}=3700$\,K. Therefore, the presence of the pulsar and its irradiation may be important. Given the irradiation, one would expect the side of the companion facing the pulsar to be brighter than the side facing away from it. Thus, from Earth, the companion should appear faintest at phase 0.25 and brightest at phase 0.75 (using the convention that at phase 0, the pulsar is at the ascending node). This is indeed seen in other pulsar binaries, with the black widow pulsar PSR~B1957+20 perhaps the most spectacular example \citep{vpasc+88,fgld88}. For star X, assuming a fraction $\eta$ of the incident flux is absorbed and reradiated as optical flux, the flux from the bright side of the companion should be a factor $1+\frac{2}{3}\eta f_{\rm irr}/f_{\rm th}$ brighter (here, the factor $\frac{2}{3}$ reflects projection effects). Observationally, the inferred values of $\eta$ range from 0.1 to 0.6 (\citealt{ok03}, and references therein), and thus one expects a maximum change in bolometric flux by a factor 1.13 to 1.8. For the $R$-band flux, the range is 1.2 to 2.2 (assuming it scales like a black-body spectrum, $\propto\!T^6$ around 3700\,K). We confirmed this using a detailed light-curve synthesis model (described briefly in \citealt{sklk99}). For star~X, no effect is seen. Using the PSR~J0751+1807 ephemeris from \citet{nss+05}, we find that during the ESI $R$-band observations the orbital phase ranged from 0.22 to 0.25, while the 1996 LRIS $R$-band images were taken at phases 0.86--0.90, and the 2005 LRIS images at phases 0.01--0.14 on January 7, and 0.77--0.93 on January 8. Thus, these observations span the orbital phases necessary to test for any modulation in brightness. Indeed, using the inclination inferred from timing, $i=66^{+4}_{-7}\,$deg \citep{nss+05}, we find that during the ESI observations only 4 to 5\% of the irradiated part of the companion surface was in view, while during the 1996 LRIS observations is was 78\% to 85\%. As a consequence, we expect to see nearly the maximum change in brightness. Nevertheless, in Sect.~\ref{sec:photometry}, we found no significant variation, $R_{\rm LRIS}-R_{\rm ESI}=0.03\pm0.13$; thus, to $\sim\!99$\% confidence, the variation is smaller than 0.3\,mag, which implies $\eta<0.15$. The lack of observed modulation could be taken to indicate that the irradiation is not very effective, e.g., because the albedo is large (i.e., $\eta$ is small), the pulsar emission is non-isotropic, or the spin-down luminosity is overestimated. We believe these options are not very likely (for a discussion in a slightly different context, see \citealt{ok03}), which leads us to consider the only alternative, that one of the assumptions underlying the above estimates is wrong. In particular, we assumed implicitly that irradiated flux is reprocessed and re-emitted instantaneously, i.e., transfer of flux inside and around the companion are assumed to have negligible effect. For the companions of black-widow pulsars, this is reasonable, since for these relatively large objects, tides will have ensured synchronous rotation. Any flux transfer would thus have to be due to winds and/or convection, which plausibly happens on a timescale long compared to the thermal time of the layer in which the pulsar flux is reprocessed. The companion of PSR~J0751+1807, however, is well within its Roche-lobe, and tidal dissipation should be negligible. We can estimate its current rotation period from its prior evolution, following the reasoning used by \cite{kk95} for the companion of PSR B0655+64. Briefly, during mass transfer, the companion filled its Roche-lobe and tides ensured the system was synchronised and circularised. Once mass transfer ceased and the companion started to contract to a white dwarf, however, the tides became inefficient, and the rotational evolution of the companion was determined by conservation of angular momentum. For our estimates, we split the total moment of inertia of the progenitor into two parts, one from the core, $I_\mathrm{core}=k_{\rm core}^2M_\mathrm{core} R_\mathrm{core}^2$ and one from the envelope, $I_\mathrm{env}=k_{\rm env}^2M_\mathrm{env} R_\mathrm{L}^2$; here $k$ is the radius of gyration and $R_{\rm L}$ is the radius of the Roche lobe. After contraction of the envelope, one is left with a white dwarf with $I_\mathrm{WD}=k_{\rm WD}^2 M_\mathrm{WD} R_\mathrm{WD}^2$. If we now assume that $I_\mathrm{core} \simeq I_\mathrm{WD}$ and ignore differences in radius of gyration, conservation of angular momentum yields $\Omega_\mathrm{rot} / \Omega_\mathrm{orb} \simeq 1+ M_\mathrm{env} R_\mathrm{L}^2 / M_\mathrm{WD} R_\mathrm{WD}^2$. In reality, likely the envelope will be more centrally concentrated than the white dwarf, i.e., $k_\mathrm{env}<k_{\rm WD}$, and tidal dissipation will be important in the initial stages of the contraction. This will reduce the spin-up. On the other hand, the hot core of the progenitor will be larger than the white dwarf, i.e., $I_{\rm core}>I_{\rm WD}$. In any case, it follows that unless the envelope mass is very small, the white dwarf should be significantly spun up. Model predictions for the envelope mass of helium-core white dwarfs differ. The $0.196$\,M$_\odot$ model by \citet{sarb01}, has an envelope mass of $6.7\times10^{-3}$\,M$_\odot$ (as given in \citealt{asb01}), whereas a model of similar mass ($M_\mathrm{WD}=0.195$\,M$_\odot$) by \citet{dsbh98} has one of $3.1\times10^{-2}$\,M$_\odot$. Using these values, taking $M_\mathrm{WD}=0.19$\,M$_\odot$, $R_\mathrm{WD}=0.021$\,R$_\odot$ and $R_\mathrm{L}=0.48$\,R$_\odot$, and ignoring differences in $k$, we find current rotation periods a factor 18--85 faster than the orbital period, or 20 to 5 minutes. Given that thick envelopes seem inconsistent with the low observed temperature (Sect.~\ref{sec:tandr}), the slower end of the range seems more likely. To estimate the timescale on which the pulsar flux is reprocessed, we assume that the incident particles are predominantly highly energetic, and that they penetrate to, roughly, one Thompson optical depth. This corresponds to a column depth of $N=1.5\times10^{24}{\rm\,cm^{-2}}$, for which the thermal timescale $t\simeq NkT/\sigma T_\mathrm{eff}^4\simeq1\,$min, where the numerical estimate is for $T=T_\mathrm{eff}=3700\,$K. This is shorter than the rotation periods estimated above, suggesting that rotation may not be too important. On the other hand, our estimate is very rough. For instance, at one Thompson depth, the opacity at optical wavelengths is much smaller than unity for the cool temperatures under consideration \citep{sbl+94}. Thus, the material likely radiates less efficiently than a black body, which would make the thermal timescale longer. Furthermore, the irradiation will change the temperature and ionisation structure of the atmosphere, further complicating matters. (Indeed, could this be the underlying cause for the fact that the colours deviate so strongly from those expected for a pure hydrogen atmosphere?) Finally, it might induce strong winds which equalise the temperature on both hemispheres (as is the case for Jupiter). \section{Conclusions}\label{sec:discussion} We have optically identified the white dwarf companion of the binary millisecond pulsar PSR~J0751+1807. We find that the companion has the reddest colours of all known millisecond pulsar companions and white dwarfs. These colours indicate that the companion has a very low (ultra-cool) temperature of $T_\mathrm{eff}\sim\!3500-4300$\,K. Furthermore, the colours suggest that the white dwarf has a pure helium atmosphere, or a helium atmosphere with some hydrogen mixed in, as invoked for the field white dwarf WD~0346+246 which has similar colours \citep{osh+01,ber01}. Our observations are inconsistent with evolutionary models, from which one would expect a pure hydrogen atmosphere. Indeed, as for other short-period systems, the hydrogen envelope is expected to be thick enough to sustain significant residual hydrogen burning, leading to temperatures far in excess of those observed. It may be that the mass of the envelope was reduced due to shell flashes or irradiation by the pulsar, as was proposed by \citet{esa01}. However, we see no evidence for irradiation, despite the fact that the pulsar spin-down flux inpinging on the white dwarf is roughly double the observed thermal flux. Clues to what happens might be found from more detailed studies of the spectral energy distribution, or more accurate phase-resolved photometry. Finally, a deeper observation at infrared wavelengths would allow one to distinguish between the different atmosphere compositions for the companion: for a pure helium atmosphere, black-body like colours are expected, while if any hydrogen is present, the infrared flux would be strongly depressed (as is seen for WD 0346+246). With adaptive optics instruments, such observations should be feasible. \begin{acknowledgements} We thank Norbert Zacharias for providing preliminary UCAC2 data. We also would like to thank the referee, Pierre Bergeron, for his useful suggestions and for pointing out the existence of his updated models. The observations for this paper were taken at the W. M. Keck Observatory, which is operated by the California Association for Research in Astronomy, a scientific partnership among the California Institute of Technology, the University of California, and the National Aeronautics and Space Administration. It was made possible by the generous financial support of the W. M. Keck Foundation. MIDAS is developed and maintained by the European Southern Observatory. This research made use of the SIMBAD and ADS data bases and of data products from the Two Micron All Sky Survey, which is a joint project of the University of Massachusetts and the Infrared Processing and Analysis Center/California Institute of Technology, funded by the National Aeronautics and Space Administration and the National Science Foundation. We acknowledge support from NWO (C. G. B.), NSERC (M. H. v. K.), and from NASA and NSF (S. R. K.). \end{acknowledgements} \bibliographystyle{aa}
Title: The Bolocam 1.1 mm Lockman Hole Galaxy Survey: SHARC II 350 micron Photometry and Implications for Spectral Models, Dust Temperatures, and Redshift Estimation
Abstract: We present 350 micron photometry of all 17 galaxy candidates in the Lockman Hole detected in a 1.1 mm Bolocam survey. Several of the galaxies were previously detected at 850 microns, at 1.2 mm, in the infrared by Spitzer, and in the radio. Nine of the Bolocam galaxy candidates were detected at 350 microns and two new candidates were serendipitously detected at 350 microns (bringing the total in the literature detected in this way to three). Five of the galaxies have published spectroscopic redshifts, enabling investigation of the implied temperature ranges and a comparison of photometric redshift techniques. Lambda = 350 microns lies near the spectral energy distribution peak for z = 2.5 thermally emitting galaxies. Thus, luminosities can be measured without extrapolating to the peak from detection wavelengths of lambda > 850 microns. Characteristically, the galaxy luminosities lie in the range 1.0 - 1.2 x 10^13 L_solar, with dust temperatures in the range of 40 K to 70 K, depending on the choice of spectral index and wavelength of unit optical depth. The implied dust masses are 3 - 5 x 10^8 M_solar. We find that the far-infrared to radio relation for star-forming ULIRGs systematically overpredicts the radio luminosities and overestimates redshifts on the order of Delta z ~ 1, whereas redshifts based on either on submillimeter data alone or the 1.6 micron stellar bump and PAH features are more accurate.
https://export.arxiv.org/pdf/astro-ph/0601582
command. \shorttitle{Bolocam 1.1 mm Lockman Hole Galaxy Survey} \shortauthors{Laurent et~al.} \newcommand{\deriv}[2]{\frac{d{#1}}{d{#2}}} \newcommand{\D}{\,d} \newcommand{\Avg}[1]{\left\langle #1 \right\rangle} \newcommand{\vect}[1]{\ensuremath{\mathbf{#1}}} \begin{document} \title{The Bolocam 1.1 mm Lockman Hole Galaxy Survey: SHARC II 350 $\mu$m Photometry and Implications for Spectral Models, Dust Temperatures, and Redshift Estimation} \author{ G.~T. Laurent\altaffilmark{1,2}, J. Glenn\altaffilmark{1}, E. Egami\altaffilmark{3}, G.~H. Rieke\altaffilmark{3}, R.~J. Ivison\altaffilmark{4}, M.~S. Yun\altaffilmark{5}, J.~E. Aguirre\altaffilmark{6,1}, P.~R. Maloney\altaffilmark{1}, \& D.~Haig\altaffilmark{7}} \altaffiltext{1}{Center for Astrophysics and Space Astronomy \& Department of Astrophysical and Planetary Sciences, University of Colorado, 593 UCB, Boulder, CO 80309-0593} \altaffiltext{2}{glaurent@colorado.edu} \altaffiltext{3}{Steward Observatory, University of Arizona, 933 North Cherry Avenue, Tucson, AZ 85721} \altaffiltext{4}{UK Astronomy Technology Centre, Royal Observatory, Blackford Hill, Edinburgh EH9} \altaffiltext{5}{Department of Astronomy, University of Massachusetts, Amherst, MA 01002} \altaffiltext{6}{Jansky Fellow, National Radio Astronomy Observatory} \altaffiltext{7}{Physics and Astronomy, Cardiff University, 5 The Parade, P.O. Box 913, Cardiff CF24 3YB, Wales, UK} \keywords{galaxies: high-redshift --- galaxies: starburst --- submillimeter} \section{Introduction} Surveys at submillimeter and millimeter wavelengths have detected hundreds of galaxy candidates by their thermal dust emission. The galaxies (hereafter referred to as submillimeter galaxies) characteristically have redshifts $z > 1$ and inferred luminosities of $L \sim 10^{13}$ L$_\odot$ and star formation rates of $10^3$ M$_\odot$ per year (assuming dust heating by young stars). Such enormous luminosities and star formation rates, or accretion rates in the case of super-massive black hole growth, make submillimeter galaxies strong candidates for the progenitors of massive galaxies at the current epoch. Clearly, it is crucial to characterize the spectral energy distributions (SEDs) where their emission peaks ($\lambda$ = 350 $\mu$m for 40 K dust at a redshift of $z = 2.5$), measure their redshifts and luminosity functions, determine their power sources, and integrate them into theories of galaxy formation. Although submillimeter galaxy SEDs peak at a few hundred microns for all but the highest redshifts, $\lambda \ge$ 850 $\mu$m surveys have been most successful at detecting galaxies because of the lower atmospheric noise and greater transmission, and less stringent telescope surface requirements. Most of the detections have been low signal-to-noise ratio (just over thresholds of 3-4 $\sigma$), necessitating multiwavelength confirmation. Furthermore, the SEDs have been extrapolated shortward from 850-1200 $\mu$m over the peak, or between 850-1200 $\mu$m and the far-infrared, to estimate dust temperatures, luminosities, and star formation rates. Clearly, 350 $\mu$m photometry can confirm galaxy candidates and sample the SEDs near their peaks for more precise inferences of physical parameters. Similarly, because of the difficulty in obtaining spectroscopic redshifts of large numbers of highly obscured galaxies, various photometric redshift estimation techniques have emerged, notably based on the far-infrared to radio luminosity relation in ULIRGs \citep{carilli99,yun02} and the stellar continuum bump in the infrared for {\it Spitzer}-detected galaxies \citep{egami04,sawicki02}. However, despite the difficulty, candidate spectroscopic redshifts have been obtained for $\sim$ 73 galaxies \citep{chapman05}. Thus, with well-determined dust-emission SEDs, including 350 $\mu$m, photometric techniques can be compared to spectroscopic redshifts. In this paper, we present 350 $\mu$m photometry of all 17 submillimeter galaxy candidates from the Bolocam Lockman Hole survey \citep{laurent05}, some with previous 850 $\mu$m and 1200 $\mu$m detections. Two new galaxy candidates were serendipitously detected at 350 $\mu$m, bringing the total number of 350 $\mu$m-discovered galaxies to three \citep{khan05}. We combine these data with infrared and radio data to derive improved luminosities, explore the range of implied dust temperatures and spectral indices, and compare photometric redshift techniques. This comparison is timely with the imminent launch of the {\it Herschel Space Observatory} (scheduled for August 2007), which will detect thousands of galaxies at far-infrared and submillimeter wavelengths, but for which spectroscopic redshifts will be attainable for only a small fraction. Throughout the paper, a cosmology of H$_0 = 70$ km s$^{-1}$ Mpc$^{-1}$, $\Omega_M = 0.3$, and $\Omega_\Lambda$ = 0.7 is assumed. \section{The 350 $\mu$m SHARC II Galaxy Survey} \label{section:sharc} Observations at multiple submillimeter wavelengths are vital both to confirm the Bolocam sources (as 6 false detections are expected from Monte-Carlo simulations) and to make photometric redshift and temperature estimates. The 350 $\mu$m SHARC-II observations combined with the Bolocam 1.1 mm galaxy survey provides a flux density ratio that is strongly dependent on redshift for a given temperature. This is because the rest wavelength corresponding to the observed wavelength of 350 $\mu$m with SHARC-II is near the peak of the grey-body spectrum (for a $z \sim 2$ galaxy at 40 K), and Bolocam's 1.1 mm observations climb the steep $\nu^{2+\beta}$ ($\beta$ $\approx$ 1.5) modified Rayleigh-Jeans side of the spectral energy distribution. Follow-up observations of each of the Bolocam Lockman Hole galaxy candidates \citep{laurent05} were taken with the Submillimeter High Angular Resolution Camera (SHARC II) at the Caltech Submillimeter Observatory. The observations were taken on three observing runs: 2004 March-April, 2005 January, and 2005 February. The brightest Bolocam sources (1 and 2) were observed over 8 hours of total integration time during the 2004 March-April run, although most of the run was lost due to poor weather. Bolocam sources 1 and 16 were observed over 18 hours of integration time during the 2005 January run, again with much of the run lost due to poor weather. The 2005 February run was characterized by much better weather, and all the Bolocam sources except for 1, 5, 8 and 16 were observed over 35 hours of integration time\footnote{The primary weather measurement correlated with the SHARC II mapping speed is the CSO 225 GHz heterodyne, narrowband, ``tipper tau'' monitor, which measures the zenith atmospheric attenuation. The 2004 March-April, 2005 January, and 2005 February Lockman Hole observations yielded $\tau_{225\mathrm{GHz}}$ ranges and 75th percentiles of $\tau_{225\mathrm{GHz}}=0.046-0.087$, $\tau_{75\%}=0.076$, $\tau_{225\mathrm{GHz}}=0.044-0.120$, $\tau_{75\%}=0.093$ and $\tau_{225\mathrm{GHz}}=0.030-0.074$, $\tau_{75\%}=0.047$, respectively.}. When combined with the observations of Bolocam sources 5 and 8 by \cite{kovacs05}, the entire Bolocam sample was observed over these observing runs. Observations with SHARC II were taken in the point source observing mode, with a Lissajous (parametric sinusoidal curve) scan pattern using the SWEEP command of telescope. The Lissajous pattern was scanned in altitude and azimuth, with amplitudes of 30$\arcsec$ and 20$\arcsec$, respectively. When combined with the 2.6$\arcmin$ $\times$ 1.0$\arcmin$ SHARC II field-of-view and 9$\arcsec$ FWHM instrument beam size, this resulted in a uniform coverage region of 95$\arcsec$ $\times$ 18$\arcsec$, with a border of additional coverage (60$\arcsec$ $\times$ 40$\arcsec$) outside this region. Each observation had a fixed length of 10 minutes to ensure uniform coverage even on individual scans. Integration times and the resulting depths of each of the SHARC II fields are listed in Table \ref{table:sharcdetections}. \setlength{\tabcolsep}{1mm} \begin{deluxetable}{ccccccccccc} \tabletypesize{\scriptsize} \tablecaption{SHARC II Photometry and New Galaxy Candidates} \tablewidth{0pt} \tablehead{ \colhead{} & \colhead{} & \colhead{} & \colhead{SHARC II/Bolocam} & \colhead{Bolocam 2 $\sigma$} & \colhead{SHARC II} & \colhead{SHARC II} & \colhead{} & \colhead{} & \colhead{} \\ \colhead{Bolocam} & \colhead{SHARC II} & \colhead{SHARC II} & \colhead{Offset} & \colhead{Error Circle} & \colhead{R.A} & \colhead{Dec} & \colhead{S/N} & \colhead{S$_\nu$} & \colhead{$\sigma$} \\ \colhead{Source} & \colhead{Source} & \colhead{\# 10 min. Scans} & \colhead{(")} & \colhead{(")} & \colhead{(J2000.0)} & \colhead{(J2000.0)} & \colhead{} & \colhead{(mJy)} & \colhead{(mJy)} \\ } \startdata 1&7 &30,16&13.2 &21&10:52:57.1&57:21:01&3.7&38.0&14.0\\ 2&8 &17,23&15.8 &21&10:51:18.6&57:16:36&3.5&20.9&7.9\\ 3&10 &20&9.1 &21&10:52:13.0&57:15:46&3.2&14.0&5.6\\ "&11&20&17.3 &21&10:52:14.0&57:16:02&3.1&15.1&6.2\\ 4&21&20&(18.3) &22&10:52:04.8&57:18:39&&$\le$ 15.4&\\ 5&6 (Kov\'{a}cs 4)&&15.0&22&10:52:30.9&57:22:06&5.9&40.4&8.6\\ 6&1 & 5&12.0 &22&10:51:14.1&57:14:21&6.8&63.6&18.4\\ "&9 & 5&18.0 &22&10:51:17.8&57:14:20&3.3&27.6&10.8\\ 7&20&15&(16.7) &22&10:51:28.6&57:30:50&&$\le$ 10.7&\\ 8&5 (Kov\'{a}cs 5)&&16.3&23&10:52:38.8&57:24:38&6.2&40.5&8.1\\ 9&18&14&(15.4) &23&10:53:05.0&57:15:08&&$\le$ 19.6&\\ 10&16&36&(22.4) &23&10:51:31.0&57:23:35&&$\le$ 25.1&\\ 11&19&13&(19.9) &23&10:52:49.0&57:13:01&&$\le$ 18.9&\\ 12&14&18&(1.5) &23&10:51:15.5&57:15:22&&$\le$ 20.5&\\ 13&17&13&(20.5) &23&10:52:34.9&57:18:13&&$\le$ 23.7&\\ 14&2 &16&20.5 &24&10:52:01.7&57:24:43&5.9&24.1&7.3\\ 15&15&15&(24.0) &24&10:51:47.9&57:28:57&&$\le$ 20.0&\\ 16&13 &93&1.6 &25&10:52:27.3&57:25:13&3.0&44.0&18.3\\ 17&12&16&6.0 &26&10:52:00.6&57:24:21&3.1&15.5&6.3\\ \hline (9)&3 &14&- &23&10:53:08.3&57:15:01&4.8&28.4&9.2\\ (16)&4 &93&- &25&10:52:32.3&57:24:48&3.8&37.0&13.4\\ \enddata \tablecomments{SHARC II detections and 3 $\sigma$ upper limits at each of the Bolocam sources, in order of descending brightness at 1.1 mm. The Bolocam sources in parentheses correspond to SHARC II detections well outside of the Bolocam 2 $\sigma$ positional error circle, and are therefore believed not to be associated with the Bolocam source. Two SHARC II detections from \cite{kovacs05} are also included. The SHARC II source numbers are listed in order of 350 $\mu$m S/N. } \label{table:sharcdetections} \end{deluxetable} \setlength{\tabcolsep}{2mm} The reduction of the raw SHARC II data was accomplished with the use of the "deep" cleaning utility of the Comprehensive Reduction Utility for SHARC II (CRUSH\footnote{http://www.submm.caltech.edu/$\sim$sharc/crush/index.htm}). Observations of pointlike galaxies, quasars, protostellar sources, H$_\mathrm{II}$ regions, and evolved stars were used to construct pointing models for each of the observing runs. Observations of the pointing sources were taken with a scan strategy identical to that of the science fields. A subset of the pointing sources were used for flux density calibration, with reference 350 $\mu$m flux densities obtained from the SHARC II website\footnote{http://www.submm.caltech.edu/$\sim$sharc/}. Source extraction was performed on the CRUSH-cleaned maps, with each map (corresponding to a single Bolocam candidate) consisting of all of the individual scans co-added together. The algorithm was begun by doing a cut on the uniform coverage region, defined as the set of pixels for which the coverage is $\ge$ 60\% of the maximum per-pixel coverage. The uniform coverage region is a contiguous region in the center of each map. Next, an RMS in sensitivity units (the flux density of each pixel times the square root of the integration time for that pixel in units of mJy s$^{1/2}$) was computed in the uniform coverage region. This RMS is valid for the entire uniform coverage region since variations in coverage have been accounted for by the $t_i^{1/2}$ coverage normalization, where $t_i$ is the total integration time for pixel $i$. All pixels with coverage-normalized flux densities exceeding 3 $\sigma$ (``hot pixels'') were flagged as potential sources. Then hot pixels were grouped into multi-pixel sources by making the maximal group of adjacent hot pixels, including those within $\sqrt{2}$ pixels (i.e., diagonally adjacent). The peak flux density, right ascension and declination of the source candidates were computed by centroiding two-dimensional Gaussians on the groups. The uncertainty in the flux density of each source is given by the pixel-to-pixel RMS at the centroid location of the source. \section{Positional Uncertainties} \label{section:positionalerrors} The large beam sizes of submillimeter and millimeter wave instruments (31$\arcsec$, 14$\arcsec$, 11$\arcsec$, and 9$\arcsec$ FWHM for Bolocam, SCUBA, MAMBO, and SHARC II, respectively) makes it difficult to identify likely optical and radio counterparts to the galaxy candidates. Despite the large beam sizes, however, individual sources can be centroided to much higher precision than the quoted beam size. To help constrain this issue of source matching between the various surveys, a positional error circle was estimated for each of the submillimeter and millimeter band detections. For the Bolocam detections, Monte Carlo simulations were performed by injecting sources into the timestream and running the reduction pipeline and source extraction algorithm. This simulation was repeated for a range of source flux densities. The resulting centroiding error as a function of flux density (5.4 - 9.1$\arcsec$) was added in quadrature with the RMS telescope pointing error (9.1$\arcsec$) to yield a range of 2 $\sigma$ positional error circles of 21 - 26$\arcsec$. A similar approach was used to estimate the centroiding errors for SCUBA \citep{scott02} and MAMBO \citep{greve04}, yielding 2.2 - 10.4$\arcsec$ and 1.2 - 4.3$\arcsec$ respectively. When added in quadrature to the quoted pointing errors (4$\arcsec$ and 3$\arcsec$, respectively), this yields 2 $\sigma$ positional errors of 9.2 - 22$\arcsec$ and 6.5 - 13$\arcsec$, respectively. As the centroiding error as a function of flux density for the SHARC II observations was not available, values of 3.0 - 4.0$\arcsec$ were empirically determined from the SHARC II centroiding fits of the Bolocam sources, which, when added in quadrature with the pointing error of 3.8$\arcsec$, yields 2 $\sigma$ positional errors of 9.8 - 11$\arcsec$. These error circles were used to correlate the sources between the different surveys to find coincident detections. \section{Results} \subsection{SHARC II 350 $\mu$m Detections} Postage stamp images of each of the SHARC II fields are shown in Figure \ref{figure:postage}. Each image has been cropped to 60$\arcsec$ $\times$ 60$\arcsec$, centered on the Bolocam source positions (dotted circle). The SHARC II source candidate list is presented in Table \ref{table:sharcdetections}, where the sources are listed in order of Bolocam source number. Seven Bolocam galaxy candidates were detected by SHARC II at $> 3 \sigma$ (Bolocam 1, 2, 3, 6, 14, 16, 17). Two of these sources (Bolocam 3, 6) were found to have two SHARC II counterparts. Two Bolocam candidates (5 and 8) were observed by \cite{kovacs05} and are also included in the list. 3 $\sigma$ upper limits are given for each of the Bolocam fields with no positive detections. Note that the flux density uncertainties in the last column of Table \ref{table:sharcdetections} include uncertainties of $\sim$ 20\% due to calibration error (as determined by the dispersion of the calibration source flux densities). The correlation between the two source lists based upon the positional error circles and detection offsets is lower than expected, as only 2 of the 7 Bolocam sources with a single SHARC II counterpart have errors within 1 $\sigma$ (5 are expected from a normal distribution). This may be due to underestimating the Bolocam pointing error (cf.\ \S\ \ref{section:positionalerrors} and Laurent et al.\ 2005). An additional two sources were detected in the survey (in the fields of Bolocam 9 and 16); they are not associated with the Bolocam sources because their locations are well outside of the Bolocam positional error circles (see \S\ \ref{section:positionalerrors}). While 0.8 false detections are expected from Gaussian statistics in the SHARC II survey (given the 60\% uniform coverage cut and the 9\arcsec\ SHARC II beam), these serendipitous sources nevertheless may be real. SHARC II sources 3 and 4 were detected at high significance at 350 $\mu$m and lie in regions of positive flux density in the Bolocam map, with S/N ratios of 2.9 and 2.1, respectively. \subsection{SHARC II / Bolocam Correspondence} \label{section:summary} In addition to the Bolocam 1.1 mm and SHARC II 350 $\mu$m detections, existing multiwavelength coverage (submillimeter, radio, infrared, optical, and X-ray) of the Lockman Hole was used to identify likely counterparts and characterize coincident sources. A detailed description of each of these surveys is found in Appendix \ref{section:coverage}. A comprehensive summary of the counterparts to the Bolocam detections (including the coverage by each survey) is listed in Table \ref{table:detections}. Given the large size of the Bolocam beam, identifying likely counterparts requires a certain amount of judgement. Detailed maps of the sources can be found in Figure \ref{fig:circles}. Additional notes on individual objects are discussed below. Bolocam.LE.1100.1 -- We conclude that Bolocam source 1 is likely to be a submillimeter galaxy given the coincident Bolocam, MAMBO, and SHARC II detections. In addition, highly plausible radio, {\it Spitzer}, and faint optical counterparts exist. Bolocam.LE.1100.2 -- The SHARC II detection falls 16$\arcsec$ to the ENE, but contains both radio sources in its error circle. The two 20 cm radio sources \citep{yun05,biggs06} are unrelated due to their separation. We treat each of the radio sources separately when fitting photometric redshifts. The coincident northern radio source with bright optical counterparts has a low photometric redshift (see next section) and is likely to be a low-redshift galaxy as the SDSS survey concludes. Bolocam and SHARC II may also be detecting the \cite{biggs06} southern radio source, which has a very faint optical counterpart. Bolocam.LE.1100.3 -- The position of this 6.0 mJy Bolocam source is at the edge of the good coverage region of the SCUBA survey. All of the optical counterparts to the radio sources are relatively bright (22 - 24 magnitude), although curiously, the SDSS catalog classifies the optical counterpart to the northeast radio source as a star. (Given the radio and 350 $\mu$m SHARC II counterparts, we conclude that the SDSS classification may be incorrect.) Given that both of the SHARC II sources have radio, {\it Spitzer}, and optical counterparts, each (or both) are likely candidates as submillimeter galaxies and each could contribute to the flux density of the Bolocam source. Note that when estimating photometric redshifts where source confusion may be present (for this Bolocam source and elsewhere), no attempt was made to partition the Bolocam flux density among multiple submillimeter sources (due to the large uncertainties in position). Bolocam.LE.1100.5 -- Four radio detections fall within the Bolocam positional error circle: one within the SCUBA and SHARC II error circles and just outside the edge of the MAMBO error circle, one on the edge of the MAMBO error circle, and the other two near the edge of the Bolocam positional error circle. The southwest radio source has 5-color SDSS photometry and is classified as a galaxy (extended). The fact that three of the radio sources lie outside both the SCUBA and SHARC II error circles makes them unlikely to be the correct counterpart of the submillimeter detections. We therefore choose the northeast radio source to be the more likely counterpart (which is confirmed by the fact that Chapman et al.\ 2005 were able to obtain a spectroscopic reshift for this submillimeter galaxy at this radio position, as discussed in Appendix \ref{section:previousredshift}). \noindent A MAMBO detection located just outside the Bolocam error circle was not detected by SHARC II or SCUBA (3 $\sigma$ upper limit of 35.4 mJy), but has a \cite{chapman05} spectroscopic redshift of 1.956. Due to the large size of the Bolocam beam, the Bolocam flux densities and positions may be influenced by source confusion. Bolocam.LE.1100.6 -- Two SHARC II counterparts fall within the Bolocam positional error circle, each with radio counterparts (with the eastern source containing two radio counterparts). The SDSS survey classifies both the radio source associated with the western SHARC II source and the eastern radio source associated with the eastern SHARC II source as galaxies. Each of the three radio sources may be contributing to the Bolocam flux density due to source confusion. We treat each of the radio sources separately when fitting photometric redshifts. Bolocam.LE.1100.8 -- Given the fact that the southern radio source (of the pair of two radio sources oriented N-S) has Bolocam, SCUBA, MAMBO, SHARC II, and {\it Spitzer} detections, along with a spectroscopic redshift, we conclude that this Bolocam source is real. Nevertheless, the northern radio source cannot be ruled out as a galaxy also contributing to the submillimeter fluxes. Bolocam.LE.1100.14 -- This 4.4 mJy Bolocam detection is likely influenced by source confusion, given three closely spaced submillimeter sources (SCUBA sources 1, 4, and 8, with the latter two lying near, but outside of the Bolocam positional error circle). The location of the southern radio source relative to the Bolocam position is greater than the 16$\arcsec$ radius used for the {\it Spitzer} counterparts catalog, and thus no {\it Spitzer} data is available. This set of coincident sources is the most likely counterpart to the Bolocam source. Just outside of the Bolocam error circle lies another 850 $\mu$m SCUBA detection (LE850.4) to the northeast, with a coincident \cite{ivison02} 20 cm radio source as well as published 3.6, 4.5, 5.8, and 8.0 $\mu$m {\it Spitzer} counterparts. The SCUBA source that coincides with Bolocam source 14 (LH850.1) is also the brightest SCUBA source and has been the target of many published multi-wavelength studies. In addition to the extensive radio, infrared, optical and X-ray surveys discussed in Appendix \ref{section:coverage}, a faint (K $\simeq$ 23.5) galaxy counterpart was positively identified \citep{lutz01} at the radio position. The source was found to be extended (20-30 kpc), clumpy (on subarcsecond scales) and very red ({\it I - K} $>$ 6.2). Bolocam.LE.1100.16 -- We conclude that this Bolocam source is in fact a submillimeter galaxy, given the large number of multiwavelength detections and a radio source with a confirmed spectroscopic redshift. Bolocam.LE.1100.17 -- This 4.0 mJy Bolocam detection is likely influenced by source confusion, given two nearby submillimeter sources. The 850 $\mu$m SCUBA, 1.1 mm MAMBO, and 850 $\mu$m SHARC II coincident detections to the northeast are the likely counterparts to Bolocam source 14 and are discussed in detail in Bolocam.LE.1100.14. The SDSS catalog curiously classifies the northwest radio source as a star. We conclude that both the southeast and northwest radio sources associated with the Bolocam source may be submillimeter galaxies, given the large number of multiwavelength detections. While a confirmed spectroscopic redshift exists near the southeast radio source, self-consistent photometric redshifts and multiple optical counterparts at the radio position suggest that the spectroscopic redshift may not correspond to the radio / submillimeter sources (see \S\ \ref{subsection:bolocam17}). \subsection{SHARC II Non-Detections} From extensive Monte-Carlo simulations of the Bolocam data set \citep{laurent05}, 6 false detections (Poisson distributed) are expected. This represents a large fraction (6/17) of the overall source catalog and is a consequence of the relatively low 3 $\sigma$ detection threshold used in the source detection algorithm. Eight of the Bolocam sources (4, 7, 9, 10, 11, 12, 13, and 15) were found to show no secure counterparts at 350 $\mu$m, although two of the sources (Bolocam 9, 12) exhibit flux densities just below the 3 $\sigma$ detection threshold (with a coincident radio detection for Bolocam source 9). Here we describe each of the SHARC II non-detections of the Bolocam sources in detail. Bolocam.LE.1100.4 -- A single radio counterpart \citep{biggs06} lies near the edge of the MAMBO positional error circle, with an SDSS classification of the optical counterpart as a galaxy. While well within the Bolocam positional error circle, the location of the radio source relative to the Bolocam position is greater than the 16$\arcsec$ radius used for the {\it Spitzer} counterparts catalog, and thus no {\it Spitzer} data is available. Given the coincident Bolocam and Mambo detections, along with a lack of SHARC II and SCUBA detections, this source could possibly be a very high redshift galaxy ($z > 4$), such that the SED falls below the 3 $\sigma$ detection threshold of the SCUBA 850 $\mu$m survey. Bolocam.LE.1100.7 -- The lack of multiwavelength observations makes it difficult to determine whether this Bolocam detection is real (and associated with the coincident radio detection). The lack of {\it Spitzer} and SHARC II counterparts to the radio source, however, leads us to believe that the Bolocam source may be a spurious detection. Bolocam.LE.1100.9 -- It is interesting to point out that the SHARC II upper limit in the Bolocam error circle is just below the 3 $\sigma$ detection flux density threshold and coincides with the radio position. The lack of multiwavelength observations makes it difficult to determine whether this Bolocam detection is real. The presence of {\it Spitzer} counterparts and a possible dim SHARC II detection, however, leads us to believe that the Bolocam source may be real. Bolocam.LE.1100.10 -- Given the lack of counterparts, we conclude that there is little evidence to suggest that this detection represents a submillimeter galaxy and is likely a spurious detection. Bolocam.LE.1100.11 -- The lack of more multiwavelength data makes it difficult to determine whether this Bolocam detection is real. While lacking a SHARC II detection, the radio source with an SDSS classification as a galaxy (extended object) leads us to believe that the Bolocam source may be real. Bolocam.LE.1100.12 -- A portion of the Bolocam error circle lies outside the deep \cite{ivison05} optical Subaru R-band field. Similar to Bolocam source 9, we point out that the SHARC II upper limit in the Bolocam error circle is just below the 3 $\sigma$ detection flux density threshold. The lack of more multiwavelength data makes it difficult to confirm whether this Bolocam detection is real. Bolocam.LE.1100.13 -- This Bolocam source lacks 850 $\mu$m SCUBA detections (although a SCUBA source is located just outside of the Bolocam positional error circle). We suggest that there is little evidence that the Bolocam detection represents a submillimeter galaxy and is likely a spurious detection. Bolocam.LE.1100.15 -- The lack of MAMBO and SHARC II counterparts makes it difficult to confirm the Bolocam detection. Nevertheless, the SDSS classification of the position coincident with the radio source as a galaxy (extended) leads us to believe that the Bolocam source may be real. \subsection{Submillimeter Spectral Energy Distributions} \label{subsection:correlations_between_spectra} The submillimeter spectral energy distributions (SEDs) of the coincident SHARC / Bolocam detections is shown in Figure \ref{figure:sed_all}. Five of the 17 Bolocam galaxy candidates (5, 8, 14, 16, 17) have spectroscopic redshifts from Chapman et al. (2005, see Appendix \ref{section:previousredshift}). In order to properly compare the SEDs, it is necessary to shift each of the SEDs to a common redshift. Thus, each observed SED was brought to a redshift of 2.0 (the mean redshift of the five Bolocam galaxies) using the spectroscopic redshifts. The composite SED of these five Bolocam galaxies can be seen in Figure \ref{figure:revertzoom}. In addition to redshifting the SEDs to align their rest wavelengths, a cosmological dimming term was applied by assuming a flat ($\Omega_k=0$), $\Omega_\Lambda$ = 0.7 cosmology. Finally, to account for variations in their intrinsic brightnesses, we normalize the flux densities of these five Bolocam galaxies by tying together their SEDs at the observed Bolocam wavelength of 1.1 mm. To account for the spread of the redshifted wavelengths of the 1.1 mm Bolocam observations, the flux densities were normalized to the \cite{laurent05} model based on the observations cited in the \cite{blain02} paper. The model assumes a single dust temperature of 40 K ($\beta=1.6$) and is overplotted as a solid line in Figure \ref{figure:revertzoom}. Note that only the Bolocam observations are constrained to pass through this model. Upon inspection, we find that at least four of the five Bolocam galaxies with spectroscopic redshifts (5, 8, 14, 16) exhibit very similar SEDs in the submillimeter and millimeter wavelengths. They are modeled adequately by the 40 K composite SED based on nearby dusty IRAS galaxies, high redshift submillimeter galaxies, gravitationally lensed high-redshift galaxies, and high redshift AGN. The Bolocam galaxy 17, however, appears to peak at a much higher wavelength and lower flux density than the others. We believe that there is enough source confusion to question whether the \cite{chapman05} redshift for this galaxy ($z$ = 0.689) corresponds to the SED shown (see \S\ \ref{subsection:bolocam17}). If the spectroscopic redshift is valid, the SED is modeled much better by a T = 20 K ($\beta=1.0$) grey-body dust spectrum. \section{Redshifts} \label{section:photometric_redshifts} \subsection{Introduction} With the multiwavelength photometry of the Bolocam sources, we fit photometric redshifts using various models based on different portions of the SED. Photometric redshifts based on the far-IR-to-radio correlation were derived using the models of \cite{carilli99} and \cite{yun02}. The shape of the submillimeter and millimeter part of the spectrum was also fit without the radio points, assuming a blackbody emission spectrum modified by a dust emissivity term \citep{wiklind03,laurent05}. A brief description of each of the models, fitting methods, and the redshift results are discussed in the next section. \subsection{Redshift Techniques} \label{section:redshift_techniques} This section attempts to briefly describe each of the five photometric redshift techniques used in this paper and the results of the fits when applied to the Bolocam galaxy candidates. The following section will compare the relative merits of each of the photometric redshift fitting techniques and discuss the results of redshift distributions. 1) FIR-Radio Spectral Index -- \cite{carilli99} used the semianalytic, linear relationships derived by Condon (1992) between the massive star formation rate and the radio synchrotron luminosity and far-IR dust emission from active star-forming galaxies to show that the spectral index between these two frequencies, $\alpha^{350}_{1.4}$, is a well behaved function of redshift: \begin{eqnarray} \alpha^{350}_{1.4} = -0.24 - [0.42 \times (\alpha_{\mathrm{radio}} - \alpha_{\mathrm{submm}}) \times \mathrm{log} (1+z)], \end{eqnarray} where we adopt the standard value in Condon (1992) of -0.8 for $\alpha_{\mathrm{radio}}$, and a value of +3.2 for $\alpha_{\mathrm{submm}}$ (an average of the spectral indices between 270 and 850 GHz for M82 and Arp 220. The relation is believed to be a result of relativistic electrons accelerated in supernova remnants (producing synchrotron radiation) and dust heated by the interstellar radiation field (with a thermal peak of $\sim 380 \mu$m for a galaxy with $z = 2$ and T = 40 K). Photometric redshifts determined using only the Bolocam and radio flux densities are listed in Table \ref{table:redshifts_yunandcarilli}. Redshift results from Bolocam sources with multiple radio counterparts are listed using the higher S/N detection in the case of coincident detections by independent surveys or are listed together in the case of multiple counterparts detected by a single group. The error bars listed in Table \ref{table:redshifts_yunandcarilli} (and elsewhere throughout this paper) were obtained from Monte-Carlo simulations of the fits and represent statistical errors due to measurement uncertainty in the flux densities. The flux densities at each observed wavelength were varied about their mean value assuming a Gaussian distribution of flux errors. Each Monte-Carlo SED was then fit to the photometric redshift models with a standard, least-squares minimization fitting routine. Each simulation was repeated 1000 times, with the error bars quoted being the minimum-length 1 $\sigma$ confidence intervals from the resulting histogram of redshifts. It should be noted that these confidence intervals represent only the statistical goodness of fit and that uncertainties in the templates themselves are expected to dominate the photometric redshift errors. 2) Entire FIR-Radio SED -- \cite{yun02} utilized the entire Far-IR to radio spectral energy distribution to estimate photometric redshifts and SFRs. The redshift template is based upon the theoretical models of thermal dust emission, thermal bremsstrahlung (free-free) emission, and nonthermal synchrotron emission for dusty starburst galaxies. Photometric redshift fits of the five Bolocam galaxy candidates (5, 8, 14, 16, 17) with spectroscopic redshifts \citep{chapman05} are shown in Figure \ref{figure:photoz}, with best-fit redshifts (and errors) also listed in Table \ref{table:redshifts_yunandcarilli}. The solid lines in the figure represent the best fit spectrum to the submillimeter, millimeter, and radio point shown. The dotted line represents a second fit using the \cite{yun02} model, this time fixing the spectroscopic redshift and normalizing (varying only the SFR) to the submillimeter points. \begin{deluxetable}{ccccccccccc} \tabletypesize{\tiny} \tablecaption{Photometric Redshifts} \tablewidth{0pt} \tablehead{ \colhead{Bolocam} & \colhead{N$_{\mathrm{Radio}}$} & \colhead{N$_{\mathrm{Submm}}$} & \colhead{N$_{\mathrm{{\it Spitzer}}}$} & \colhead{Carilli \& Yun} & \colhead{Yun \& Carilli} & \colhead{Wiklind} & \colhead{Laurent} & \colhead{MRK231} & \colhead{ARP220} & \colhead{Chapman}\\ \colhead{1.1 mm} & \colhead{} & \colhead{} & \colhead{} & \colhead{1999} & \colhead{2002} & \colhead{2003} & \colhead{et al.\ 2005} & \colhead{} & \colhead{} & \colhead{et al.\ 2005}\\ \colhead{Number} & \colhead{} & \colhead{} & \colhead{} & \colhead{z$_{\mathrm{phot}}$} & \colhead{z$_{\mathrm{phot}}$} & \colhead{z$_{\mathrm{phot}}$} & \colhead{z$_{\mathrm{phot}}$} & \colhead{z$_{\mathrm{phot}}$} & \colhead{z$_{\mathrm{phot}}$} & \colhead{z$_{\mathrm{spec}}$} } \startdata 1 &2(S) &3 &3 &4.6$^{+0.3}_{-0.3}$ &4.1$^{+0.3}_{-0.3}$ &3.2$^{+0.4}_{-0.5}$ &3.2$^{+0.3}_{-0.4}$ &15$^{+1.4}_{-1.5}$ &3.0$^{+0.2}_{-0.3}$ &\\ " &1(N) &" &3 &4.6$^{+0.3}_{-0.3}$ &4.1$^{+0.3}_{-0.3}$ &" &" &1.8$^{+1.0}_{-0.8}$ &2.0$^{+0.7}_{-0.7}$ &\\ 2 &2,1(N,S)&2 &0 &2.2$^{+0.2}_{-0.2}$, 3.9$^{+0.3}_{-0.3}$ &0.6$^{+0.1}_{-0.1}$, 2.9$^{+0.7}_{-0.8}$ &4.2$^{+0.6}_{-0.8}$ &3.9$^{+0.4}_{-0.4}$ & & &\\ 3 &2(NE) &2 &5 &2.6$^{+0.2}_{-0.2}$ &0.7$^{+0.1}_{-0.1}$ &4.7$^{+0.8}_{-0.7}$ &4.2$^{+0.4}_{-0.5}$ &0.2$^{+0.2}_{-0.2}$ &3.1$^{+0.2}_{-0.2}$ &\\ " &2(SW) &2 &3 &3.7$^{+0.3}_{-0.3}$ &1.6$^{+0.3}_{-0.4}$ &4.9$^{+0.8}_{-0.7}$ &4.3$^{+0.5}_{-0.4}$ &10$^{+0.9}_{-0.8}$ &2.3$^{+0.7}_{-0.2}$ &\\ 4 &1 &2 &0 &5.7$^{+0.7}_{-0.8}$ &5.1$^{+0.7}_{-0.9}$ & & & & &\\ 5 &3(E) &4 &5 &4.3$^{+0.4}_{-0.4}$ &3.4$^{+0.2}_{-0.3}$ &2.5$^{+0.5}_{-0.4}$ &2.6$^{+0.4}_{-0.5}$ &1.2$^{+0.2}_{-0.2}$ &2.6$^{+0.2}_{-0.2}$ &2.611\\ 6 &2,2(E,W)&2(E) &0 &4.2$^{+0.5}_{-0.4}$, 3.1$^{+0.3}_{-0.3}$ &3.5$^{+0.4}_{-0.5}$, 2.0$^{+0.3}_{-0.3}$ &3.2$^{+0.7}_{-0.7}$ &3.2$^{+0.5}_{-0.6}$ & & &\\ " &2 &2(W) &0 &1.6$^{+0.2}_{-0.2}$ &0.8$^{+0.1}_{-0.1}$ &1.9$^{+0.4}_{-0.5}$ &2.0$^{+0.5}_{-0.6}$ & & &\\ 7 &1 &1 &0 &4.2$^{+0.6}_{-0.8}$ &3.6$^{+0.6}_{-1.0}$ & & & & &\\ 8 &2(N) &4 &4 &5.1$^{+0.5}_{-0.6}$ &4.7$^{+0.3}_{-0.3}$ &3.1$^{+0.4}_{-0.6}$ &3.1$^{+0.3}_{-0.5}$ &0.6$^{+0.3}_{-0.3}$ &2.4$^{+0.6}_{-0.2}$ &\\ " &1(S) &" &4 &5.1$^{+0.6}_{-0.6}$ &4.7$^{+0.4}_{-0.4}$ &" &" &0.7$^{+0.4}_{-0.1}$ &3.1$^{+0.2}_{-0.6}$ &3.036\\ 9 &2 &1 &5 &3.9$^{+0.4}_{-0.4}$ &3.2$^{+0.5}_{-0.5}$ & & &1.3$^{+0.2}_{-0.2}$ &1.8$^{+0.2}_{-0.5}$ &\\ 10 &0 &1 &0 & & & & & & &\\ 11 &1 &1 &2 &3.2$^{+0.4}_{-0.5}$ &2.5$^{+0.4}_{-0.6}$ & & &0.4$^{+0.2}_{-0.2}$ &0.8$^{+0.3}_{-0.2}$ &\\ 12 &0 &1 &0 & & & & & & &\\ 13 &2,1(NE,SE)&1 &0 &3.9$^{+0.4}_{-0.4}$, 4.8$^{+0.6}_{-0.6}$ &3.2$^{+0.5}_{-0.4}$, 4.4$^{+0.8}_{-0.8}$ & & & & &\\ " &1(E) &" &4 &5.0$^{+0.6}_{-0.7}$ &4.6$^{+0.8}_{-1.0}$ & & &0.7$^{+0.3}_{-0.2}$ &2.4$^{+0.3}_{-0.5}$ &\\ 14 &4 &5 &5 &3.7$^{+0.4}_{-0.3}$ &3.1$^{+0.2}_{-0.2}$ &3.2$^{+0.3}_{-0.4}$ &3.2$^{+0.3}_{-0.3}$ &7.3$^{+0.6}_{-0.2}$ &3.0$^{+0.2}_{-0.3}$ &2.148\\ 15 &1 &1 &5 &4.1$^{+0.5}_{-0.4}$ &3.5$^{+0.6}_{-0.5}$ & & &0.6$^{+0.2}_{-0.2}$ &1.0$^{+0.2}_{-0.2}$ &\\ 16 &2 &4 &5 &4.1$^{+0.5}_{-0.5}$ &3.2$^{+0.2}_{-0.2}$ &1.9$^{+0.3}_{-0.5}$ &2.0$^{+0.4}_{-0.4}$ &0.4$^{+0.2}_{-0.2}$ &2.3$^{+0.3}_{-0.2}$ &2.142\\ 17 &3(SE) &4 &10 &3.8$^{+0.5}_{-0.4}$ &2.8$^{+0.3}_{-0.3}$ &3.4$^{+0.5}_{-0.7}$ &3.3$^{+0.4}_{-0.4}$ &3.4$^{+0.3}_{-0.3}$ &3.1$^{+0.2}_{-0.2}$ &0.689\\ " &1(NW) &" &5 &5.3$^{+0.7}_{-0.8}$ &4.1$^{+0.4}_{-0.5}$ &" &" &0.3$^{+0.2}_{-0.2}$ &3.0$^{+0.2}_{-0.6}$ &\\ \enddata \tablecomments{Best fit photometric redshifts of the Bolocam galaxy candidates using the models of \cite{carilli99}, \cite{yun02}, \cite{wiklind03}, and \cite{laurent05}, and cool and warm ULIRGs Arp 220 and MRK 231. N$_{\mathrm{Radio}}$, N$_{\mathrm{Submm}}$, and N$_{\mathrm{{\it Spitzer}}}$ are the number of coincident radio, submillimeter and {\it Spitzer} infrared points, respectively.} \label{table:redshifts_yunandcarilli} \end{deluxetable} 3) Modified Blackbody -- \cite{wiklind03} found that observations of local ULIRGs exhibit a remarkably low dispersion in the far-IR to millimeter wavelengths ($\lambda > 50 \mu$m), independent of whether the power source of the thermal emission is due to AGN or intense star formation. \cite{wiklind03} fit a simple blackbody emission spectrum (modified by a dust emissivity term) to sample of 37 local ULIRGs from \cite{klaas01}: \begin{eqnarray} \label{equation:greybody} f_\nu \propto \epsilon_\nu B_\nu(T) \propto [1-\exp{(-\tau_\nu)}] B_\nu(T), \end{eqnarray} where $B_\nu(T)$ is the Planck function evaluated at dust temperature, $T$, and frequency, $\nu$, and $\tau_\nu$ is the optical depth of the dust: \begin{eqnarray} \nonumber \tau_\nu = \left(\frac{\nu}{\nu_0}\right)^\beta. \end{eqnarray} \cite{wiklind03} made no assumption about the Wien side of the spectrum, as only the submillimeter ($\ge$ 450 $\mu$m) and millimeter points were fit. Using the best-fit parameters from \cite{wiklind03}: $\beta = 1.8$, $\nu_0 = 1.2 \times 10^{12}$ Hz (250 $\mu$m), and $T_d = 68$ K, we fit photometric redshifts to the SHARC II 350 $\mu$m, SCUBA 450 and 850 $\mu$m, Bolocam 1.1 mm, and MAMBO 1.2 mm flux densities of the galaxies detected in our Bolocam survey. The two parameter fit (redshift and overall flux density normalization) yields redshifts for 9 of the 17 bolocam galaxies with $\ge$ 2 submillimeter/millimeter points. Seven of the Bolocam galaxies (Bolocam 7, 9, 10, 11, 12, 13, 15) have no counterpart in the submillimeter/millimeter and one (Bolocam 4) has detections only at 1.1 and 1.2 mm, which is an insufficient wavelength spread in order to properly constrain the galaxy redshift using this two parameter model. Redshift results for each of the 9 Bolocam galaxies are listed in Table \ref{table:redshifts_yunandcarilli}. The best fit models to the five galaxies with \cite{chapman05} redshifts (Bolocam 5, 8, 14, 16, 17) are shown in Figure \ref{figure:wiklind}. Similar to the method of \cite{wiklind03}, \cite{laurent05} created a composite SED of nearby dusty {\it IRAS} galaxies, high-redshift submillimeter galaxies, gravitationally lensed high-redshift galaxies, and high-redshift AGNs \citep[][and references therein]{blain02}, and found fit parameters of $T$ = 40 K, $\nu_0$ = 3700 GHz, and $\beta$ = 1.6 for Equation \ref{equation:greybody}. Redshift results for each of the 9 Bolocam galaxies are listed in Table \ref{table:redshifts_yunandcarilli}. The best fit models to the five galaxies with \cite{chapman05} redshifts (Bolocam 5, 8, 14, 16, 17) are shown in Figure \ref{figure:wiklind}. 4) Near IR Stellar Bump -- \cite{egami04} used the multiband imaging capabilities of the IRAC and MIPS IR cameras of the {\it Spitzer Space Telescope} to observe 38 VLA radio sources in the Lockman Hole. They classified the resulting IR SEDs into two types: those showing a clear near-IR stellar continuum hump at a rest wavelength of 1.6 $\mu$m (due to the minimum opacity of the H$^-$ ion at 1.6 $\mu$m from photo-detachment and free-free transitions, which results in a local maximum in the the SEDs of cool stars, Sawicki 2002), and those with a featureless power-law continuum (from AGN). We fit the {\it Spitzer} IR counterparts of the Bolocam galaxies with each of these spectra, using the \cite{egami04} models of a cool ULIRG Arp 220 \citep[from][]{silva98} and a warm (dominated by an AGN) ULIRG Mrk 231. Only two fit parameters were used: the redshift, and an overall normalization. Photometric redshifts were fit for each of the 12 Bolocam galaxies (Bolocam 1, 3, 5, 8, 9, 10, 11, 13, 14, 15, 16, 17) with $\ge$ 2 IR {\it Spitzer} points. Three Bolocam galaxies (Bolocam 2, 6, 12) were outside of the field surveyed by \cite{egami04}. Two galaxies (Bolocam 4, 7) have {\it Spitzer} counterparts, but due to the high density of {\it Spitzer} sources in the field, they could not be uniquely associated with the Bolocam sources (because of a lack of another coincident detection in the submillimeter and/or radio). The best-fit photometric redshifts for both Arp 220 and Mrk 231 (fitting only the {\it Spitzer} near- and mid-infrared points) are shown in Figure \ref{figure:spitzer}, with the resulting redshifts also listed in Table \ref{table:redshifts_yunandcarilli}. \subsection{Comparison of Photometric Redshift Techniques} \label{subsection:discussion} Comparing the results of each of the photometric redshift techniques with the spectroscopic redshifts of \cite{chapman05} yields widely varying degrees of agreement. Histograms of redshift errors for each of the photometric redshift models \citep[other than][in which coincident radio detections were treated separately]{carilli99} are shown in Figure \ref{figure:histogram}. The histogram from fitting models of Arp 220 and Mrk 231 to the {\it Spitzer} IRAC and MIPS observations are the fits that result in the lowest residual $\chi^2$ (Arp 220 for Bolocam 5, 8, 14, and 16, and Mrk 231 for Bolocam 17). The \cite{yun02} model \citep[as well as][]{carilli99} yields systematically high photometric redshifts compared to the spectroscopic redshifts by \cite{chapman05}. The comparison of model SEDs for photometric (solid line) and spectroscopic (dotted line) redshifts in Figure \ref{figure:photoz} suggests that SED data points at the extreme ranges of wavelength coverage strongly influence the model fit and that the systematic tendency to derive a high redshift is primarily driven by the lower than expected radio continuum flux density. This is supported by the fact that fits to only the submillimeter and millimeter-wave points yield much more accurate photometric redshifts (see below). This is perhaps not surprising, given recent evidence \citep{chapman05} suggesting a large degree of dispersion in the radio-to-far-IR correlation at higher redshift. Nevertheless, it is unlikely that the \cite{yun02} model template, which is derived from the ensemble average of 23 infrared luminous galaxies in the local universe, is systematically biased by radio bright objects because $\ge$98\% of all FIR-selected galaxies follow the well known and tight radio-FIR correlation, independent of FIR luminosity \citep{yun01}. Aside from Bolocam source 17 whose spectroscopic redshift by \cite{chapman05} appears suspect (see \S\ \ref{subsection:bolocam17}), these comparisons suggest that the observed radio continuum in Bolocam galaxies is 2-5 times fainter than predicted by the synchrotron flux densities (which dominate thermal brehmsstrahlung by a factor of $\sim$ 13 at 20 cm) from the low-redshift ULIRGs for which the local FIR-radio correlation was derived (see further discussions in \ref{subsection:radio}). In contrast to the far-IR-to-radio correlation photometric redshift techniques, both the \cite{wiklind03} and \cite{laurent05} modified blackbody curves correctly estimate the redshifts for three of the five Bolocam galaxies with spectroscopic redshifts (within the uncertainties of the photometric redshift techniques). The strength of the submillimeter / millimeter-only photometric redshift technique is twofold. First, while abandoning the radio points limits the number of points (as few as two, in some cases) to which we may fit a model, we ensure that the physics that dominates the region of the spectrum to which we are fitting is directly relevant to star formation-heated dust emission. Synchrotron radiation, by contrast, is dependent upon high energy electrons streaming through interstellar magnetic fields, whose properties may vary as a function of environment (e.g.\ inverse-Compton losses for cosmic rays with higher CMB energy densities at high redshift -- see \S\ \ref{subsection:radio}). Indeed, these galaxies are typically at least an order of magnitude more luminous than the low-redshift, infrared-luminous galaxies from which the FIR-radio correlation was derived. Second, having the Bolocam 1.1 mm flux densities on the Rayleigh-Jeans side of the spectrum and the SHARC II 350 $\mu$m flux densities near the peak of the SED makes the 350 $\mu$m/1.1 mm flux density ratio a strong function of redshift. This can be seen in Figure \ref{figure:fluxratio}, which shows the flux density ratios between various wavebands based on the \cite{laurent05} model SED. The importance of the SHARC II 350 $\mu$m waveband is apparent. For intermediate to high ($z < 5$) redshifts, the SHARC II flux density drops rapidly as a power law ($\sim \nu^{-1.7}$ due to the hotter components of dust) with redshift on the Wien side of the spectrum, while the millimeter-wave climbs up the steep Rayleigh-Jeans portion of the SED. Photometric redshifts using the 450 or 850 $\mu$m wavebands are less sensitive than the 350 $\mu$m / 1.1 mm wavebands. Extending this analysis shows the discriminatory power of the BLAST and {\it Herschel} space telescope 250 $\mu$m bands in conjunction with a millimeter waveband, although the far-IR waveband begins to probe a range of hotter dust temperatures. (The flux density ratio of these wavebands may not be well correlated, as discussed in \S\ \ref{subsection:irspectrum}.) BLAST \citep[Balloon-borne Large-Aperture Submillimeter Telescope,][]{devlin01} is a balloon-based instrument which incorporates a 2-meter primary mirror and is equipped with large-format bolometer cameras operating at 250, 350, and 500 $\mu$m which, when complete, will provide the first sensitive large-area (0.5-40 deg$^2$) submillimeter surveys at these wavelengths. The bolometer arrays are prototypes of the Spectral and Photometric Imaging Receiver (SPIRE) focal plane cameras for the {\it Herschel} satellite \citep{griffin01}, which will further investigate the formation and evolution of AGNs and star formation in high redshift submillimeter galaxies. It is important to note that while both the \cite{wiklind03} and \cite{laurent05} modified blackbody curves both produce reasonably accurate photometric redshifts (and produce nearly identical SEDs on the Rayleigh-Jeans portion of the spectrum), they make substantially different assumptions about the dust properties: $T_{\mathrm{dust}}$ and $\beta$ are 68 K and 1.8 for the \cite{wiklind03} model SED and 40 K and 1.6 for the \cite{laurent05} model SED. This points to the degeneracy of the dust temperature and the grain emissivity index. While the shapes of the submillimeter SEDs are reasonably modeled by either dust model and thus predict photometric redshifts with some accuracy, essentially no information about the dust temperatures can be inferred. In fact, representing the dust SED with two (or more) components produces similar $\chi^2$ values (and thus similar redshifts), with both temperatures {\it lower} than that of the single dust temperature model \citep{wiklind03}. The photometric redshifts determined by fitting the {\it Spitzer} infrared observations yield redshifts that are equivalent to both the \cite{laurent05} and \cite{wiklind03} model SEDs. This confirms the conclusions of \cite{egami04}, in which starburst-dominated galaxies ("cold") show remarkably similar SEDs in the infrared. The resulting photometric redshifts are highly sensitive to the 1.6 $\mu$m continuum hump (and PAH features), with a sharp minimum in $\chi^2$. (The ARP 220 fits may be biased towards particularly good fits as 4 of the 5 galaxies with spectroscopic redshifts lie between 2 $\lesssim z \lesssim$ 3, which is optimal for the 8 $\mu$m PAH feature to be shifted into the observed 20 $\mu$m IRAC waveband.) AGN-dominated ("warm") galaxies also show very similar SEDs, but lacking a strong continuum feature in the infrared, are subject to larger redshift fitting uncertainties; a brighter, higher redshift galaxy is characterized by a similar shape in the infrared portion of the SED as a cooler, low-redshift galaxy, with the Wien side of the spectrum well-modeled with a power-law \citep{blain99}. \subsection{Bolocam Source 17: Spectroscopic Misidentification?} \label{subsection:bolocam17} We bring special attention to photometric redshift analysis of Bolocam source 17 (corresponding to SCUBA source 8), as the \cite{laurent05}, \cite{wiklind03}, and {\it Spitzer} IR models are consistent in overpredicting the redshift of this galaxy ($z_{\mathrm{spec}}$ = 0.689) by $\ge 4 \sigma$ ($z_{\mathrm{phot}}$ = 3.3$^{0.4}_{0.5}$, 3.4$^{0.6}_{0.6}$, and 3.4$^{0.3}_{0.3}$, respectively). We point out that the large offset between the spectroscopic and photometric redshifts is possibly the result of source confusion, as two radio sources (both with {\it Spitzer} and optical counterparts) fall near the center of the Bolocam, SCUBA, MAMBO, and SHARC II error circles, within 4$\arcsec$ of each other. (Egami et al.\ 2004 refer to the northwest and southeast radio sources as LE850.8a and LE850.8b, respectively.) The northwest radio source is believed by \cite{lehmann01} to be the counterpart to the ROSAT X-ray emission, who find a redshift of 0.974 using optical Keck spectroscopy. Using XMM-Newton observations, however, \cite{ivison02} conclude that the X-ray source corresponds to the southeast radio source. Indeed, the linear fit of a combination of ARP 220 and MRK 231 ULIRG models to the {\it Spitzer} points coincident with the southeast radio source yields a 100\% warm (AGN dominated) component. It is near this radio position that \cite{chapman05} find a spectroscopic redshift of 0.689. In fact, the X-ray emission observed with both instruments appears to fall between these two radio sources. Optical R-band images from both \cite{yun05}, and \cite{ivison05} show multiple optical counterparts at the southeast radio source position. Furthermore, the spectroscopic position quoted by \cite{chapman05} appears to coincide with an optical source $\sim$ 2$\arcsec$ to the south of the southeast radio source, a source detected with four-color SDSS photometry (in addition to Yun et al.\ 2005 and Ivison et al.\ 2005 R-band photometry) and cataloged as a low-redshift galaxy. We conclude that it is possible that the submillimeter detections may either be suffering from source confusion from two or more galaxies, or that the \cite{chapman05} redshift corresponds to a source other than that of the southeast radio detection. If the latter is true, then the consistent redshifts predicted by the Laurent et al./Wiklind (2005, 2003) and {\it Spitzer} IR models may further point to the accuracy of these photometric redshift techniques. \section{Discussion} \subsection{IR Spectrum} \label{subsection:irspectrum} SEDs over the entire IR-radio spectral range of the five Bolocam galaxies with spectroscopic redshifts are shown in Figure \ref{figure:revert}. These spectra have the same redshift, cosmological dimming, and normalization corrections as in Figure \ref{figure:revertzoom}. While four of the five galaxies have closely correlated spectra in the submillimeter region of the spectrum, the infrared spectra ({\it Spitzer} 3.6, 4.5, 5.8, and 8.0 $\mu$m IRAC and 24 $\mu$m MIPS observations) exhibit a large degree of dispersion. This dispersion may be the result of several things: 1) Because the spectra have been normalized to a T = 40 K ($\beta=1.6$) spectrum at their 1.1 mm Bolocam flux densities to account for intrinsic brightness variation between the galaxies, the normalization will result in an artificial reduction in the submillimeter flux density dispersions. This effect is not likely to dominate, as the flux density normalization has a $\sim$ 10\% effect on the flux densities of the galaxies. 2) The {\it Spitzer} detection associated with Bolocam source 14 lies systematically low compared to the other three galaxies well-modeled by a T=40 K dust spectrum. While these {\it Spitzer} observations \citep{egami04} are from a different data set than the remaining {\it Spitzer} observations (this work), it is unlikely that their flux densities are systematically uncertain by nearly an order of magnitude. 3) The 40 K dust temperature model of the submillimeter portion of the spectrum for Bolocam sources 5, 8, 14, and 16 assume a single dust temperature for each of the four sources. The temperature fit parameter is somewhat degenerate with other fit parameters, including the critical frequency, $\nu_0$, where the optical depth of the dust is unity. Thus, if Bolocam source 14 has a lower characteristic dust temperature, then the infrared portion of the 40 K model will significantly overestimate the infrared flux density. \cite{chapman05} estimates the temperature of Bolocam source 14 to be 33 K from two photometric points (850 $\mu$m SCUBA and 1.4 GHz VLA radio observations) and the dust SED templates of \cite{dale02}. This temperature uncertainty likely dominates our uncertainty in matching the infrared flux densities. 4) In addition to heating by the ultraviolet and optical flux density from young stars associated with ongoing star formation, the thermal dust emission responsible for the bright submillimeter flux densities may be contributed to by an energetic AGN. While not dominating the total bolometric output from the galaxy, they may have a non-negligible (20\%) contribution \citep{alexander05}. If this is the case, then the shape of the infrared continuum may be vary according to the relative contribution of star formation rates for these galaxies. 5) Another possible explanation for the larger dispersion in {\it Spitzer} infrared flux density as compared to our single dust temperature model may be from the fact that the infrared flux densities trace separate epochs of star formation within the galaxy. It is plausible that the dust heated by the ultraviolet and optical flux density from from current star formation is not well correlated to the current infrared flux density of older stars (from previous star formation). Furthermore, models of UV to millimeter emission of star clusters embedded in optically thick giant molecular clouds (GMCs) suggest that the near-infrared to far-infrared portion of starburst galaxy SEDs vary considerably with age of the starburst \citep{efstathiou00}. \subsection{Radio Spectrum / FIR-Radio Correlation} \label{subsection:radio} The composite SED (including the radio points) of the 5 Bolocam sources with spectroscopic redshifts is shown in Figure~\ref{figure:revert}. Two interesting observations can be made about the radio continuum emission associated with these galaxies: (1) the radio continuum is lower (or the submillimeter continuum is higher) on average than the well established radio-FIR correlation for the local universe by Yun, Reddy, \& Condon (2001, cf.\ \S\ \ref{subsection:discussion}); and (2) like the infrared flux densities, the 6 and 20 cm VLA radio flux densities for the four galaxies with closely correlated spectra in the submillimeter region of the spectrum show a large degree of dispersion (factor of 5 in 20 cm flux density). This scatter is much larger than the quoted uncertainties of the VLA radio observations (which constrain the 20 cm radio flux density of each galaxy to better than 20\%) This degree of dispersion is also larger than the factor of 3 scatter in the radio-FIR correlation seen among the FIR selected galaxies in the local universe. In fact, while the spectroscopic redshifts are similar for the four galaxies (2.1 $< z <$ 3.0), varying the 20 cm flux density over the observed range causes the best fit photometric redshifts of \cite{yun02} to vary from $z = 2.7$ to 4.8. This dispersion undoubtably contributes to the large errors of the photometric redshifts discussed in \S\ \ref{section:photometric_redshifts}. Deep radio continuum imaging using the VLA is a technically challenging task, and the disparate 20 cm flux densities by $\sim$ 50\% reported for Bolocam sources 2 and 6 (cf.\ \S\ \ref{section:summary}) exemplifies the difficulty of the photometry at radio wavelengths. Most systematic noises in interferometry tend to suppress the brightness of astronomical sources, and part of the lower radio continuum flux density might be related to the imaging and photometry problems. It is important to note that due to the low S/N ratios at which the sources have been detected in the submillimeter wavebands ($\le 4 \sigma$ for the vast majority of SHARC, SCUBA, Bolocam, and MAMBO detections), flux bias \citep{laurent05} plays a major role in overestimating the flux densities at these wavelengths. While this factor indeed results in a systematic shift of the entire submillimeter portion of the spectrum to higher flux densities, it is unlikely that the magnitude of this effect ($\sim$ 20\% for the Bolocam flux densities) could fully account for the lower than expected (from the FIR-radio correlation) radio continuum flux densities. Calibration errors may also contribute to a systematic overestimate of the submillimeter flux densities. Finally, as two or more independent radio sources are found within the Bolocam error circle in 8 out of 17 cases in Figure 2, source confusion or source blending may also contribute to the apparently lower radio continuum flux if only one radio source is identified as the counterpart. The lower radio continuum flux density and the larger scatter may also reflect an actual breakdown in the radio-FIR correlation. Inverse-Compton losses for the high energy cosmic rays responsible for the synchrotron radiation is thought to be significant at $z > 2$, and a possible breakdown in the radio-FIR correlation has been considered previously (see Condon 1992, Carilli \& Yun 2000). This effect is demonstrated in Figure \ref{figure:spitzer}, in which models of local ULIRGs Arp 220 and Mrk 231 systematically overestimate the radio flux densities of these submillimeter galaxies. Higher quality data on a larger sample of high redshift systems are needed to examine the importance of inverse-Compton loss and the possible breakdown in the radio-FIR correlation. X-ray heating of the circum-nuclear gas and dust is an important source of luminosity in the far-IR if a luminous AGN is present (Maloney, Hollenbach, \& Tielens 1996). Radio-quiet AGN FIR emission could reduce the 1.4 GHz flux density with respect to the FIR heating. Alexander et al.\ (2005, and references therein) make the case using X-ray detections and spectral indices that many, perhaps most, submillimeter galaxies have AGN but that they are not bolometrically important. However, the statistics of X-ray detected submillimeter galaxies for which hard/soft ratios can be measured is not large and it cannot be ruled out that many submillimeter galaxies are Compton thick (N$_\mathrm{H}$ $>$ 1.5 $\times$ 10$^{24}$ cm$^{-2}$). We conclude that the generally low radio flux densities in our sample could be due to small number statistics, source confusion, or generally depressed radio emission, perhaps due to quenching of high energy cosmic rays, although radio-quiet, Compton-thick AGN contributions to the dust heating cannot be ruled out. There are new AGN versus star formation spectral diagnostics emerging (Ivison et al., Egami et al.), and it is possible that to definitively settle the issue may ultimately require ALMA, Constellation-X, and interferometric FIR spectral line diagnostic capability. \subsection{Inferred Luminosities} To obtain the intrinsic bolometric luminosities of the five Bolocam galaxies with spectroscopic redshifts, the spectra of \S\ \ref{subsection:correlations_between_spectra} (based on the Laurent et al.\ 2005 model) were integrated for each galaxy in their respective rest frames. The resulting bolometric luminosities are listed in Figure \ref{figure:revertzoom}. The four galaxies well-modeled by a 40 K dust spectrum have luminosities ranging from $L = (1.0-1.2) \times 10^{13}$ $\mathrm{L}_\odot$. The lower redshift galaxy (Bolocam source 17) has an inferred luminosity two orders of magnitude lower ($L = 1.3 \times 10^{11}$ $\mathrm{L}_\odot$). If the spectroscopic redshift of 0.689 does not apply to this galaxy and it instead lies at $z$ = 3.4 (the photometric redshift predicted by the Laurent et al. 2005 / Wiklind 2003 and {\it Spitzer} IR models), then its luminosity of $L = 8.2 \times 10^{12}$ $\mathrm{L}_\odot$ agrees well with the others. \subsection{Stellar and Dust Masses Implied from the Integrated Submillimeter Luminosities} Three hundred and fifty micron observations combined with far-infrared and millimeter-wavelength photometry enables accurate measurements of luminosity for galaxies near $z$ = 2 because no interpolation across the peak of the SED is required. Characteristically, integration of the SEDs of the galaxies in our sample from far-infrared to millimeter-wavelengths yields luminosities of $\sim$ 1 $\times$ 10$^{13}$ L$_\odot$. Assuming: 1) that the luminosity derives from star formation (young stars, which may overestimate stellar masses due to contribution from intermediate mass giants), 2) a characteristic (Salpeter) form of the initial mass function (IMF) from \cite{chabrier03}, and 3) all of the optical and ultraviolet radiation is reprocessed to long wavelengths by dust, enables the stellar mass content of the galaxies to be approximately estimated. We adopt the $M^{3.5}$ luminosity function of \cite{demircan91}. Because the luminosity function is so steep, the derived mass depends strongly on the assumed upper mass limit of integration. The lower limit of integration is not well constrained by the data, although masses less than 0.3 M$_\odot$ are not likely to dominate the mass because the IMF flattens considerably at low mass. Lower mass limits of 0.7 and 1.0 M$_\odot$ could be relevant because: 1) \cite{dwek98} argued that a Salpeter IMF for $z$ $>$ 1 galaxies cannot extend much lower than this without producing too many low mass stars that would be present today and 2) \cite{chabrier03} suggests (with caution) that the high-z IMF could cut off $>$ 1 M$_\odot$ based on multiple circumstantial lines of evidence. Varying $m_l$ from 0.3 to 1.0 M$_\odot$ and limiting $m_u$ to $\le$ 50 M$_\odot$ yields a mininum stellar mass of 10$^{10}$ M$_\odot$ and a maximum of a few $\times$ 10$^{11}$ M$_\odot$, consistent with the stellar mass content of large elliptical galaxies, as previously pointed out by many authors \citep[e.g.,][]{smail02, lilly99}. This range must be considered an upper limit because AGN could be responsible for some of the dust heating. For all of the galaxies with secure 350 $\mu$m detections, especially those with \cite{chapman05} spectroscopic redshifts, it is clear that the Bolocam 1.1 mm observations lie on the optically thin Rayleigh-Jeans side of the SED, and therefore enable dust mass estimates. The flux density, $S_\nu$, of a galaxy at an observed frequency, $\nu$, is related to the dust mass, M, by \begin{eqnarray} \nonumber S_\nu=B_{\nu'}(T) \frac{\kappa_\nu M (1+z)}{D_L^2}, \end{eqnarray} where $D_L$ is the luminosity distance to redshift z, $B_{\nu'}(T)$ is the Plank function evaluated at the emitted frequency, $\nu'$, and $\kappa_\nu$ is the dust opacity. Using the range of observed Bolocam flux densities (4.0 - 6.8 mJy), assuming a redshift of 2.1, and applying the dust cross section of $\kappa_\nu$ = 12.4 cm$^2$/g \citep{ossenkopf94} most relevant for high mass star formation (high gas density and thin ice mantle model), leads to dust masses of 3 - 5 $\times$ 10$^8$ M$_\odot$ (2 - 3 $\times$ 10$^8$ M$_\odot$) for a dust temperature of 40 K (50 K). Blindly applying a dust-to-gas mass ratio of 1/100 implies gas masses of 3 - 5 $\times$ 10$^{10}$ M$_\odot$. These gas masses are comparable to those from a recent sample of 8 submillimeter galaxies of \cite{genzel04} and \cite{neri03} which yield median molecular gas masses (from CO emission) of 2.2 $\times$ 10$^{10}$ M$_\odot$ and 2.8 $\times$ 10$^{10}$ M$_\odot$, with median dynamical masses of 1.1 $\times$ 10$^{11}$ M$_\odot$ and 6.2 $\times$ 10$^{10}$ M$_\odot$ (assuming the most probable inclination angle of sin $i = 2/\pi$), respectively. These gas mass estimates are uncertain to at least a factor of a few due to 1) the Bolocam flux density bias, which causes the measured flux densities to be overestimated by 10 - 30\%, 2) our assumed values of $\kappa$ and T, which may vary by a factor of a few and $\pm$ 20 K, respectively, and 3) increasing our assumed redshift of 2.1 (the mean of the 5 Chapman spectroscopic redshifts of the Bolocam galaxies) to $z$ = 2.4 {\it increases} our calculated gas mass by 30\%. Nevertheless, taking these factor into account still imply that a considerable fraction of the mass could already by in stars and substantial gas remains for star formation. This major epoch of galaxy formation at approximately $z$ = 2 is consistent with the conclusions of \cite{fontana04} from spectral fitting of a sample of 500 elliptical galaxies at 0.2 $\le$ $z$ $\le$ 2.5 that approximately 35\% of elliptical galaxy stellar mass was assembled by $z$ = 2 and approximately 80\% by $z$ = 1. \section{Conclusions} We have obtained 350 $\mu$m SHARC II observations toward galaxy candidates from the Bolocam Lockman Hole survey. The Lockman Hole has rich, deep, multiwavelength observations enabling detailed studies of galaxies. The 350 $\mu$m photometry is near the peaks of the SEDs of galaxies with characteristic temperatures of $\sim$ 50 K and redshifts of $z$ $\sim$ 2 to 3. They therefore enable measurements of luminosities and estimates of temperatures and photometric redshifts without interpolating over the peak of the FIR thermal SEDs. Seven galaxies detected at 1.1 mm with Bolocam were detected at 350 $\mu$m, two of which have two 350 $\mu$m counterparts; these were combined with two 350 $\mu$m detections from the survey of \cite{kovacs05}, bringing the total number of Bolocam galaxies detected with SHARC II to nine. Two additional galaxies not associated with the Bolocam sources were also detected. The SHARC II detections range in significance from 3.0 $\sigma$ to 6.8 $\sigma$, with flux densities ranging from 14 mJy to 64 mJy. We combined our observations with 850 $\mu$m and 1.2 mm photometry from the literature to fit the submillimeter/millimeter-wave spectra to thermal dust models. We found that two models with significantly different dust temperatures (40 K and 68 K) and spectral indices $\beta$ (1.6 and 1.8, respectively) yielded similar quality fits owing to the degeneracy in T and $\beta$, rendering them indistinguishable without better SED sampling. However, there is little consequence of the degeneracy to the derived luminosities, photometric redshifts, and dust masses within the statistical uncertainties. Five of the galaxies have spectroscopic redshifts in the literature, with four ranging from $z$ = 2.1 to 3.0 and one at $z$ = 0.689. The four high-z galaxies have luminosities of (1.0 - 1.2) $\times$ 10$^{13}$ L$_\odot$, while the $z$ = 0.689 galaxy is best fit by a 20 K, $\beta$ = 1.0, spectrum with a much lower luminosity: 1.3 $\times$ 10$^{11}$ L$_\odot$. (Given the source confusion in the optical and radio, along with consistent photometric redshifts, we suggest that the $z$ = 0.689 spectroscopic redshift of Bolocam source 17 may be a misidentification.) The characteristic dust masses for the four high-$z$ spectroscopic galaxies are 4 $\times$ 10$^8$ M$_\odot$, implying gas masses of 4 $\times$ 10$^{10}$ M$_\odot$. The dominant uncertainties in this estimation are the dust opacity and the gas-to-dust conversion factor, which make the estimation uncertain to a factor of a few. Assuming a Salpeter IMF and that the submillimeter emission derives completely from star formation yields stellar masses of 10$^{10}$ to a few times 10$^{11}$ M$_\odot$, broadly consistent with the stellar content of modern-day elliptical galaxies. The photometric redshifts of the full sample of seven galaxies span the range of $z$ = 2.0 to $z$ = 4.3, with statistical uncertainties of $\Delta z$ = 0.3 to 0.6 (1 $\sigma$). Photometric redshifts utilizing composite radio/FIR spectra representative of local star-forming ULIRGs yields systematically higher redshifts, on the order of $\Delta z$ = 1. For the four galaxies with optical spectroscopic redshifts the anomolously high redshifts arise from systematically low 1.4 GHz observed flux densities. The discrepancy could arise from small number statistics, inverse-Compton losses of high energy cosmic rays off the CMB, heating by radio-quiet AGN, or suppressed synchrotron emission from supernova remnants in the unusually luminous galaxies. For comparison, photometric redshifts derived using only Spitzer MIPS and IRAC data points yielded slightly more precise and accurate redshifts than the submillimeter/millimeter-wave data alone, with discriminatory power between heating by AGN and star formation (albeit with limited bolometric luminosity constraints). \acknowledgments We thank Attila Kov\'{a}cs for providing us with SHARC II detections of Bolocam sources 5 and 8. We also acknowledge the support of the CSO director and staff, the support of Kathy Deniston, and helpful conversations with Steven Eales. This work was supported in part by NSF grants AST-0098737, AST-9980846, and AST-0206158 and PPARC grants PPA/Y/S/2000/00101 and PPA/G/O/2002/00015. G.\ T.\ L.\ acknowledges NASA for GSRP Fellowship NGT5-50384, D.\ J.\ Haig and D.\ Dowell for their assistance during the SHARC II observing runs, and the entire Bolocam instrument team.
Title: Enhanced Mass-to-Light Ratios in UCDs through Tidal Interaction with the Centre of the Host Galaxy
Abstract: A recent study of ultra-compact dwarf galaxies (UCDs) in the Virgo cluster revealed that some of them show faint envelopes and have measured mass-to-light ratios of 5 and larger, which can not be explained by simple population synthesis models. It is believed that this proves that some of the UCDs must possess a dark matter halo and may therefore be stripped nuclei of dwarf ellipticals rather than merged star cluster complexes. Using an efficient N-body method we investigate if a close passage of a UCD through the central region of the host galaxy is able to enhance the measured mass-to-light ratio by tidal forces leaving the satellite slightly out of virial equilibrium and thereby leading to an overestimation of its virial mass. We find this to be possible and discuss the general problem of measuring dynamical masses for objects that are probably interacting with their hosts.
https://export.arxiv.org/pdf/astro-ph/0601330
\label{firstpage} \title[Enhanced $M/L$-ratios in UCDs]{Enhanced Mass-to-Light Ratios in UCDs through Tidal Interaction with the Centre of the Host Galaxy} \author[M. Fellhauer and P. Kroupa] {M. Fellhauer$^{1,2,3}$ \thanks{madf@ast.cam.ac.uk} and P. Kroupa$^{1,2}$ \thanks {pavel@astro.uni-bonn.de} \\ $^{1}$ Argelander Institute for Astronomy, University Bonn, Auf dem H\"{u}gel 71, 53121 Bonn\\ $^{2}$ The Rhine Stellar-Dynamical Network \\ $^{3}$ Institute of Astronomy, University of Cambridge, Madingley Road, Cambridge CB3 0HA} \pagerange{\pageref{firstpage}--\pageref{lastpage}} \pubyear{2005} \begin{keywords} galaxies: dwarfs -- galaxies: interactions -- galaxies: kinematics and dynamics -- methods: N-body simulations \end{keywords} \section{Introduction} \label{sec:intro} \citet{has05} investigated ultra-compact dwarf galaxies (UCDs) around M87, the central galaxy of the Virgo cluster. By measuring the surface brightness profiles and assessing the projected velocity dispersion of the objects they concluded that some UCDs of their sample have mass-to-light ratios of the order $5$--$9$. Furthermore, they find that some of the UCDs show faint envelopes. This supports the notion that these UCDs may be stripped nuclei of dwarf ellipticals. UCDs were first discovered by \citet*{hil98,hil99} during a study of globular clusters and dwarf galaxies around the central galaxy in the Fornax cluster. These objects are compact with effective radii of about $15$--$25$~pc and a hundred to several hundred pc in extension, and they are massive with masses of a few $10^{6}$ up to a few $10^{7}$~M$_{\odot}$. There are several theories about the origin of UCDs: (I) They could be the most luminous end of the distribution function of very massive globular clusters \citep{hil99,mie02,dir03}. (II) They could be the remnants of stripped dwarf ellipticals \citep*{bek01,bek03,dep05}. In this 'threshing' scenario a nucleated dwarf elliptical looses its envelope and most of its dark matter content due to tidal interaction with the host galaxy, such that only the 'naked' nucleus remains. (III) They could be amalgamated young massive star clusters formed in a star cluster complex during the star-burst caused by the tidal perturbation and possible disruption of a gas-rich galaxy \citep{fel02,fel05}. This scenario is well-established theoretically and was first proposed by \citet{kro98}. It must have been profusively active during the early hierarchical structure formation epoch when gas-rich substructures merged to the present-day major galaxies. \begin{table*} \centering \begin{minipage}{10.5cm} \caption{Table of our initial model parameters. The columns denote the initial mass of our model ($M_{\rm pl} = M_{\rm ini}$), the scale length (Plummer radius, $R_{\rm pl}$; analytical \& measured as described in Sect.~\ref{sec:res}), the characteristic crossing time ($T_{\rm cr}$), the central projected (line-of-sight) velocity dispersion ($\sigma_{0,p}$; analytical \& measured as described in Sect.~\ref{sec:res}), and the scaling factor ($A$) to compute the virial mass (see Eq.~\ref{eq:virplum}).} \label{tab:para} \begin{tabular}[t!]{rrrrrrr} \hline Mass [M$_{\odot}$] & $R_{\rm pl}$ [pc] & measured & $T_{\rm cr}$ [Myr] & $\sigma_{0,p}$ [km\,s$^{-1}$] & measured & $A$ \\ \hline \hline $10^{7}$ & $25$ & $23.7$ & $3.69$ & $15.92$ & $16.20$ & $1608$ \\ $10^{7}$ & $50$ & $46.6$ & $10.43$ & $11.25$ & $11.71$ & $1565$ \\ $10^{7}$ & $100$ & $94.6$ & $29.50$ & $7.96$ & $8.43$ & $1487$ \\ $10^{7}$ & $250$ & $221.0$ &$116.61$ & $5.03$ & $6.30$ & $1140$ \\ \hline $10^{8}$ & $25$ & $25.2$ & $1.17$ & $50.33$ & $50.2$ & $1573$ \\ $10^{8}$ & $50$ & $47.4$ & $3.30$ & $35.59$ & $36.8$ & $1558$ \\ $10^{8}$ & $100$ & $92.8$ & $9.33$ & $25.16$ & $26.9$ & $1489$ \\ $10^{8}$ & $250$ & $213.0$ & $36.88$ & $15.92$ & $19.91$ & $1184$ \\ \hline \end{tabular} \end{minipage} \end{table*} It is still under debate whether the UCDs, which fill the gap between globular clusters and dwarf galaxies, follow the fundamental plane (effective radius / velocity dispersion -- total luminosity) relation for globular clusters or for dwarf galaxies \citep*{has05,evs05}. While, in the $M_{V}$--$\sigma_{0}$-plane \citep[see e.g.][their Fig.~7]{has05}, they lie closer to the globular clusters, their relation seems to rather follow the one of dwarf galaxies. Furthermore they occupy exactly the space between dwarf ellipticals and their nuclei. This seems to point to formation theory (II). But the surface brightness profiles of the bright UCDs in Fornax are much more extended (i.e.\ larger effective radii) compared with nuclei of dEs \citep{dep05}. On the other hand, \citet{mar04} found an intermediate age object (W3, age $300$--$500$~Myr) in the merger remnant galaxy NGC~7252. The mass ($M = 8 \cdot 10^{7}$~M$_{\odot}$), size ($r_{\rm eff} = 17.5$~pc, and velocity dispersion of $45$~km\,s$^{-1}$) strongly suggest this object to be a UCD rather than a globular cluster. The age of this object, which corresponds to the time elapsed since the major interaction, unambiguously shows that this object can not be a stripped nucleus of a dwarf elliptical. A dwarf elliptical can not be stripped in a time interval of only $500$~Myr. \citet{fel05} showed that W3 could be the merger object of a massive star cluster complex which was formed during the interaction. The subsequent merging of star clusters forms an object with properties similar to those of W3. The evolution of the simulated super cluster shows that it transforms into a UCD such as those found in Fornax \citep{hil99,phi01}, Abell~1689 \citep{mie04}, and Virgo \citep{has05,evs05}. Moreover, \citet{fel05} showed that, due to its high mass, the object was able to retain an envelope of bound stars which initially were expelled from the individual clusters during the merger process. Thus, an envelope around a UCD is not a proof of its cosmological origin as a dwarf elliptical. Still the puzzle of high mass-to-light ratios of some of the UCDs in Virgo remains. These mass-to-light ratios (5--9) are found for the UCDs lying closest to the host galaxy. This suggests that deviations from virial equilibrium may play a role. The reasoning is that dwarf galaxies are on radial rather than circular orbits, if the UCDs formed as star-cluster complexes during the merging of major gas-rich substructures. This allows the satellites to pass close to the galactic centre. Tidal forces in the central region of the host galaxy are rather strong. Hence, the dwarf object looses stars and subsequently departs from virial equilibrium. Some of the lost stars (stars which are no longer bound to the object) do not immediately leave the object or its vicinity but disperse slowly along the orbit. Thus, a line-of-sight velocity dispersion measurement may be contaminated by these unbound stars which inflate the velocity dispersion. Furthermore, the gravitational shock of the central passage leads to an expansion of the dwarf galaxy which is later reversed again. But still this expansion through tidal heating may result in a measurable increase of the core radius once the dwarf object is again outside the host galaxy. Both of these effects may lead to an overestimation of the measured dynamical mass resulting in a higher mass-to-light ratio. These effects are not new to the astronomical community and are studied intensively by various authors (e.g. \citep{kro97} for dwarf spheroidals, \citep{may01,may02} for dwarf discs and dwarf spheroidals or \citep{bek03} for dwarf ellipticals). With this paper we want to extend these studies onto massive (compared to globular clusters) and compact (compared to other dwarf galaxies) objects like UCDs. It is definitely clear that these effects must be very strong if an object comes close to the galactic centre. But on the other hand these very close passages may be highly unlikely. By means of numerical simulations we intend to find out which sets of orbits allow an enhanced mass-to-light ratio, i.e.\ how close the UCD has to approach to the centre of its host galaxy, for which we choose the parameters of M87 to account for the UCDs with high mass-to-light ratios found around the central galaxy of Virgo. \section{Setup} \label{sec:setup} We model the parent galaxy as an analytical potential because on one single passage dynamical friction for a dwarf galaxy of mass $M \leq 10^{8}$~M$_{\odot}$ affects the orbit to at most $2$--$3$~per cent, estimated using Chandrasekhar's formula \citep{cha43} as described in \citet{por02}. We also look for only one central passage because unbound stars disperse along the orbit and are lost at subsequent passages. Some of them might still be around the object after the second or third central passage but definitely not for dozens of orbits. As a model for the analytical potential we choose the parameters for M87, the central galaxy in the Virgo cluster, consisting of a NFW-profile for the dark halo, a Hernquist profile (H) for the visible matter (stars) and a central super-massive black hole (BH) \citep*{mcl99,ves03,dim03}. The density profile of the host galaxy is \begin{eqnarray} \label{eq:pot} \rho_{\rm tot}(r) & = & \rho_{\rm NFW}(r) + \rho_{\rm H}(r) \nonumber \\ & = & \frac{\rho_{0,{\rm NFW}} \ r_{\rm s,NFW}} {r \left( 1 + \frac{r}{r_{\rm s,NFW}} \right)^{2}} + \frac{M_{\rm H} \ r_{\rm s,H}} {2 \pi \ r \left( r_{\rm s,H} + r \right)^{3}}, \end{eqnarray} with the following parameters, \begin{eqnarray} \label{eq:param} \rho_{0,{\rm NFW}} & = & 3.17 \cdot 10^{-4} \ {\rm M}_{\odot}/{\rm pc}^{3}, \\ r_{\rm s,NFW} & = & 560 \ {\rm kpc}, \\ M_{\rm H} & = & 8.1 \cdot 10^{11} \ {\rm M}_{\odot}, \\ r_{\rm s,H} & = & 5.1 \ {\rm kpc}, \\ M_{\rm BH} & = & 3 \cdot 10^{9} \ {\rm M}_{\odot}. \end{eqnarray} The above parameters denote from top to bottom the characteristic density and the scale-length of the NFW-profile, the total mass and the scale-length of the Hernquist-profile and finally the mass of the central super-massive black hole. We model the UCD as a Plummer-sphere \citep{plu11} in the numerical realisation described by \citet*{aar74}, \begin{eqnarray} \label{eq:plummer} \rho_{\rm pl}(r) & = & \frac{3M_{\rm pl}}{4\pi R_{\rm pl}^{3}} \left( 1 + \frac{r^{2}}{R_{\rm pl}^{2}}\right)^{-5/2}, \end{eqnarray} with varying scale-lengths, $R_{\rm pl}$, and initial masses, $M_{\rm pl}$, because \citet{dep05} found a Plummer-profile to fit four out of five UCDs in the Fornax cluster quite well. Furthermore, the Plummer model is analytically simple and the Plummer radius is not only the scale-length of the model but also its half-light radius (the projected radius from within which half of the light of the object is emitted). The masses of UCDs range between several million M$_{\odot}$ up to several tens of millions. So we choose for our models $10^{7}$ and $10^{8}$~M$_{\odot}$ as initial masses ($M_{\rm pl} = M_{\rm ini}$). We vary the initial scale-length of our models to be $R_{\rm pl} = 25$, $50$, $100$ and $250$~pc to determine the influence of the concentration of the objects. The cut-off radius ($R_{\rm lim}$; the radius where we truncate the Plummer distribution of our object) of all our models was kept constant at $500$~pc which is larger than the initial tidal radius. A detailed list of our model parameters can be found in Table~\ref{tab:para}. The Plummer spheres are modelled using $10^{6}$ particles. The object is set-up and integrated in isolation until equilibrium, as measured by the constancy of the $90$~\% Lagrangian radius, is reached \citep{kro97}. The UCD model is then placed at a distance of $10$~kpc to the centre of the host galaxy with no radial velocity. This means that the effect of the central passage is the most harmful possible because it is the slowest possible. The faster the passage would be (if the satellite was to start further out) the less tidal influence the central passage would have. No UCD is found closer than $10$~kpc from the centre of its host and therefore it acts as a minimum apogalacticon, which has the strongest tidal effect possible. To vary the minimum distance (perigalacticon) we give our models different tangential velocities which are listed in Tab.~\ref{tab:vel}. We assume the problem to be spherically symmetric, so we are able to place the trajectory of our model in the $x$-$y$-plane of our simulation area without restricting the problem. \begin{table} \centering \caption{List of minimum distances and the corresponding tangential velocities at the start of the simulation.} \label{tab:vel} \begin{tabular}[t!]{rr} \hline $D_{\rm min}$ [pc] & $v_{\rm tan}$ [km\,s$^{-1}$] \\ \hline \hline 0 & 0.0 \\ 50 & 6.0 \\ 100 & 10.9 \\ 150 & 15.9 \\ 250 & 25.4 \\ 500 & 48.1 \\ 1000 & 89.3 \\ 1500 & 126.0 \\ 2000 & 158.9 \\ \hline \end{tabular} \end{table} We use the particle-mesh code {\sc Superbox} to carry out the simulations. {\sc Superbox} has high-resolution sub-grids which stay focused on the core of the dwarf object while it is moving through the host galaxy. The resolution of the innermost grid containing the core is $3$~pc. For a detailed description of the code see \citet{fel00}. \section{Results} \label{sec:res} We carried out a parameter survey of 76 simulations covering different dwarf galaxy objects and different orbits. The parameter range is shown in Tables~\ref{tab:para} and~\ref{tab:vel}. For each parameter set one simulation over two central passages is carried out. For the determination of our results we look at the satellite when it reaches apogalacticon again after the first central passage. \subsection{The analytical method} \label{sec:theo} The surface density profile of the satellite is fitted by a Plummer profile of the form \begin{eqnarray} \label{eq:plum-surf} \Sigma(r) & = & \Sigma_{0} \left( 1 + \frac{r^{2}} {R_{\rm pl}^{2}}\right)^{-2}, \end{eqnarray} by applying a non-linear least-squares Marquardt-Levenberg algorithm. From this procedure we take the fitted Plummer radius for our determination of the virial mass. It can be shown that the Plummer radius exactly coincides with the half-light (projected half-mass) radius of a Plummer sphere. Furthermore we determine the line-of-sight velocity dispersion profile. In cases where it is possible, i.e.\ the profile is not too contaminated by unbound stars, we again fit the Plummer profile for the line-of-sight velocity dispersion, \begin{eqnarray} \label{eq:plum-sig} \sigma_{\rm p}(r) & = & \sigma_{\rm 0,p} \left( 1 + \frac{r^{2}} {R_{\rm pl}^{2}}\right)^{-1/4}. \end{eqnarray} The projected velocity dispersion, as well as the surface density, is measured along all three Cartesian coordinates in logarithmically spaced, concentric rings centred on the object. For the measurement all stars within a certain distance $R_{\rm max}$ in front and behind the centre of the object are taken into account. The procedure is done along all three Cartesian axes and we take the arithmetic mean value (i.e.\ a mean profile), because we do not know which orientation our objects have with respect to the observer. Therefore effects may be very strong measuring along the trajectory of the object but almost not visible in the perpendicular direction. Any random direction is likely to yield an intermediate result. For the surface density we always fit a Plummer profile even if the object is completely destroyed and does not follow a Plummer distribution at all. But in most of our models the remaining object is still fairly well represented by a Plummer profile even if it is on the way to complete destruction. In cases where we are not able to fit the velocity dispersion profile with a Plummer profile (see e.g.\ first panel of Fig.~\ref{fig:rmax}) we take an average of all measured line-of-sight velocity dispersion values within the innermost $10$~pc in projected radius to determine the central line-of-sight velocity dispersion. We now determine the virial mass taking the formula from \citet{has05} which is based on the theoretical work of \citet{kin66}, \begin{eqnarray} \label{eq:vir} M_{\rm vir} & = & \frac{9}{2 \pi G} \ \frac{\nu}{\alpha p} \ R_{\rm c} \ \sigma_{\rm 0,p}^{2}, \end{eqnarray} where $R_{\rm c}$ is the core radius and $\sigma_{\rm 0,p}$ is the central value of the projected velocity dispersion. The parameters $\nu$, $\alpha$, and $p$ are dependent on the kind of King model one wants to fit. Because we are using Plummer spheres we accumulate these parameters together with the constants into one single parameter $A$, which we determine for the isolated Plummer sphere before we start the simulation, \begin{eqnarray} \label{eq:virplum} M_{\rm vir} & = & A \ R_{\rm pl} \ \sigma_{\rm 0,p}^{2}, \end{eqnarray} where the values of the scaling factor $A$ can be found in Table~\ref{tab:para}. As one can see these values are lower for higher masses and increase for larger scale-lengths. Nevertheless we use the same $A$ determined for the isolated model for all final models stemming from this initial model, even if the mass-loss is significant. This may lead to a slight underestimation of the final virial mass. On the other hand {\sc Superbox} calculates the number of bound particles (energy below zero) at each time-step. Therefore we know the real bound mass $M_{\rm b}$. With these two masses we can determine a virial-mass-to-bound-mass ratio ($M_{\rm vir} / M_{\rm b}$ or short M/M) which can then be multiplied by a 'normal' mass-to-light ratio for a population of stars with the determined age and metallicity and without dark matter to obtain the dynamically measured $M/L$-ratio. This M/M-ratio is plotted in Fig.~\ref{fig:all}, when the satellites have reached their apogalactica again. At this point the satellites are almost back to virial equilibrium again if they are not on the way to complete dissolution. The results show clearly that only objects which are not compact and/or have a low mass can be influenced enough by one central passage to enhance the M/M to account for the high mass-to-light ratios found in UCDs. But these satellites will also not survive the next central passage or have not survived the first one at all. The best representation of real UCDs is the model with an initial mass of $10^{7}$~M$_{\odot}$ and an initial Plummer radius of $25$~pc. The results for this model are shown as tri-pointed stars in the first panel of Fig.~\ref{fig:all} and the lowest line in the first panel of Fig.~\ref{fig:result}. The results show clearly that either the satellite gets completely dissolved if $D_{\rm min}$ is closer than $50$--$100$~pc or shows no deviation in M/M at all (Fig.~\ref{fig:result}). In Fig.~\ref{fig:result} we distinguish the models according to their survival of the passage and if they show enhanced M/M-ratios. The panels show clearly that only dissolved or almost dissolved objects show enhanced M/M-ratios. All surviving objects show no deviation from virial equilibrium. \subsection{The 'observational' method} \label{sec:obs} The results in Sect.~\ref{sec:theo} are based on the exact knowledge of the theoretical central line-of-sight velocity dispersion. Observers on the other hand do not have this information. The usual way to determine the velocity dispersion of a marginally resolved object is to place a slit on the object and obtain a spectrum. From this spectral information one chooses one or two spectral lines and fits a template spectrum convolved with the instrumental line-width function and a Gaussian velocity distribution. This Gaussian velocity distribution measures the line-of-sight velocity dispersion of the whole object. Based on assumed theoretical models this value is then corrected to the central line-of-sight velocity dispersion. In the case of the simple Plummer model one has to multiply the result for the whole object by a factor of 1.25. Thus, the central velocity dispersion is \begin{eqnarray} \label{eq:gauss1} \sigma_{0,p}^{\rm pl} & = & \sqrt{\frac{3 \pi G M_{\rm pl}} {64 R_{\rm pl}}}, \end{eqnarray} while the projected velocity dispersion integrated over the whole Plummer sphere is \begin{eqnarray} \label{eq:gauss2} \sigma_{\rm obs,p}^{\rm pl} & = & \frac{1}{M_{\rm pl}} \ \int_{0}^{\infty} 2 \pi r' \Sigma(r') \sigma_{p}(r') {\rm d}r', \nonumber \\ & = & \sqrt{\frac{3 \pi G M_{\rm pl}} {100 R_{\rm pl}}}, \end{eqnarray} so that \begin{eqnarray} \label{eq:gauss3} \frac{\sigma_{0,p}^{\rm pl}} {\sigma_{\rm obs,p}^{\rm pl}} & = & 1.25. \end{eqnarray} $\sigma_{0,p}^{\rm pl}$ denotes the central line-of-sight velocity dispersion and $\sigma_{\rm obs,p}^{\rm pl}$ the weighted line-of-sight velocity dispersion integrated over the whole object. As one can see in Fig.~\ref{fig:rmax} it can be a crucial point how strongly a measurement of the line-of-sight velocity dispersion is contaminated by unbound stars. The observational values can highly overestimate the real central value. Unbound stars have a different velocity distribution than the bound ones. They are either travelling in front or behind the object (seen along the trajectory; this is shown in Fig.~\ref{fig:contour}) and are faster or slower. But the velocity distribution of these stars which are still located in the vicinity of the object and stem from one central passage, peak on either side of the Gaussian distribution of the bound stars, leading to an effective broadening of the measured velocity distribution. Fitting just a single Gaussian will therefore lead to an enhanced measured velocity dispersion and an overestimation of the central velocity dispersion value of the satellite. This is demonstrated in Fig.~\ref{fig:gauss} where the velocity distribution of all stars is shown together with the best fitting single Gaussian and the fitting curve of a triple Gaussian which takes the unbound stars into account. The measured values for the central line-of-sight velocity dispersion are shown in Table~\ref{tab:sig}. \begin{table} \centering \caption{Central velocity dispersion along the x-axis derived by different methods. Shown are the values of the same three simulations plotted in Figs.~\ref{fig:rmax} to~\ref{fig:gauss} now labelled 'dissolved', 'enhanced' and 'massive', respectively. The rows show the different values of the central velocity dispersion. First row is the direct measurement of the central dispersion, 'single' denotes the central dispersion derived from the fit of a single Gaussian to the data and 'triple (c)' denotes the value of a fit using three Gaussians where the central one (c) is used to compute the velocity dispersion of the bound object.} \label{tab:sig} \begin{tabular}[t!]{r|rrr} \hline [km/s] & dissolved & enhanced & massive \\ \hline \hline true actual & $10.6 \pm 3.1$ & $9.49 \pm 0.05$ & $35.7 \pm 0.2$ \\ single & $44.6 \pm 0.3$ & $15.90 \pm 0.30$ & $38.4 \pm 0.1$ \\ triple (c) & $13.3 \pm 0.1$ & $12.35 \pm 0.04$ & $35.3 \pm 0.3$ \\ \hline \end{tabular} \end{table} Clearly this effect is strongest in the plane of the orbit and is not visible perpendicular to it. This can be seen in Fig.~\ref{fig:rmax} where the symbol for the $z-$axis-value in all three panels shows the same value as the direct measurement. Actually for strongly disturbed systems (i.e.\ high mass-loss) the values in the $z$-direction are below the line because we do not account for the change of the constant $A$ which should increase for lower masses. But in any random orientation of the object with respect to the observer the effect of the enhanced M/M-ratio should at least be partly visible. Summing up our results we state the following: Satellites which are out of virial equilibrium, i.e.\ in the state of dissolution show enhanced virial masses and therefore the real mass content is overestimated. Figure~\ref{fig:obs} shows in the first panel the M/M-ratio plotted against the ratio of final to initial bound mass of the object. The dividing dashed line separates objects which have lost more than $50$~per cent of their mass during the central passage and are already dissolved or are not likely to survive the next passage, from objects which are stable for several more passages through the centre. While on the left side the derived M/M-ratios (using the observational method) can climb up to very high values, the stable objects show only slight enhancements of the ratio if at all. But in the second panel one already sees that even if the object itself is in virial equilibrium ($M_{\rm vir}/M_{\rm b} \approx 1.0$) there are satellites which show a broadening of the velocity distribution leading to an enhanced mass-ratio if measured the observational way. The third panel finally shows an enlargement of this area and one finds M/M-ratios overestimated by up to a factor of five. Surviving objects which show an enhancement in their M/M-ratio, if measured with this observational method, are already marked in Fig.~\ref{fig:result} with a small 'e'. As one can see there is a range of critical distances ($D>100$~pc and $D<1$~kpc) to the centre of the host galaxy where an UCD like the ones found in Virgo could show an overestimated mass-to-light ratio. \subsection{Observability} \label{sec:observ} In the previous section we claimed that with the 'observational' method the mass-to-light ratios of UCDs could be overestimated, because the velocity distribution is not Gaussian any more but 'contaminated' by unbound stars around the object. In this section we show that there is almost no chance for an observer to detect this 'non-Gaussianity' of the velocity distribution. When measuring a spectrum of a distant object, observers have to deal with two major shortcomings. First there is the intrinsic line-width produced mainly by the instrument. State-of-the-art instruments like UVES or Flames can reduce this line-width down to about $2$~km\,s$^{-1}$. But the observations of the UCDs in Virgo were made with an instrumental line-width of about $25$~km\,s$^{-1}$. The second effect an observer has to take into account is noise in the spectrum. Taking the velocity distribution from the enhanced M/M-ratio simulation we fold it with a Gaussian of the width $\sigma_{\rm i}$ to mimic the instrumental line-width and determine at which line-width the 'features' of our distribution are washed out. This happens at an instrumental line-width of $\sigma_{\rm i}=7.5$~km\,s$^{-1}$ (as shown in Fig.~\ref{fig:dist}). Then we take the best state-of-the-art line-width of $\sigma_{\rm i}=2$~km\,s$^{-1}$ and add random white noise to the distribution until again the 'features' are almost invisible again. This happens already at a signal-to-noise ratio of $20$ (see Fig.~\ref{fig:dist} lower left panel). In the final panel of Fig.~\ref{fig:dist} we fold our distribution with the line-width of the observations of the Virgo-UCDs and add noise to mimic the same S/N-ratio as in the observations. There is no deviation from Gaussiantity visible any more. \section{Discussion \& Conclusion} \label{sec:disc} We have shown with our models that dwarf satellites around a giant elliptical galaxy like M87 can have an enhanced mass-to-light ratio due to close passages to the centre of the host galaxy. While very close passages lead to the destruction of the satellite there are orbits which allow for enough 'damage' to the satellite to enhance the mass-ratio (measured the same way an observer would do) without completely destroying the object. The amount of destruction is larger if the satellite is less massive and less concentrated. But the loss of about $20$~per cent of the initial mass is enough to have the object surviving several more close passages and to mimic an enhanced mass-to-light ratio. For models comparable to the UCDs in Virgo and Fornax the range of possible minimum distances during central passages is about $100$ to $1000$~pc. All passages closer to the centre lead to complete destruction and all passages further away show no measurable effect at all, except for a mass-loss of the order of a few per cent. We therefore conclude that the enhanced M/L-ratios measured for the Virgo UCDs by \citet{has05} may not be due to dark matter. A more general result of our study is the discrepancy between the derived virial masses if one has access to the correct properties of the satellites compared to the virial masses derived the way an observer would measure. While very massive objects which are almost unaffected by tidal forces show the same results within the uncertainties one has to be careful if objects are less massive and are surrounded by a cloud of tidally stripped stars. These stars are either faster or slower in the mean but their mean values are not too different to the bulk velocity of the bound stars to clearly disentangle the 'populations' (populations in the sense of within, in front or behind the object). Especially if the broadening of a spectral line is estimated by fitting the template line folded with a Gaussian for the instrumental line-width and a single Gaussian for the velocity distribution can the deduced central velocity dispersion be too high thus leading to a mass-to-light ratio that is too large by up to a factor of ten. Effects like this have to be given serious consideration when measuring velocity dispersions of faint and distant objects. \\ \noindent {\bf Acknowledgements:} \noindent MF thankfully announces financial support through DFG-grant KR1635/5-1 and PPARC. We also want to thank M. Hilker and T. Richtler for useful comments regarding how to mimic observations. \label{lastpage}
Title: Upper limit on the ultra-high-energy photon flux from AGASA and Yakutsk data
Abstract: We present the interpretation of the muon and scintillation signals of ultra-high-energy air showers observed by AGASA and Yakutsk extensive air shower array experiments. We consider case-by-case ten highest energy events with known muon content and conclude that at the 95% confidence level (C.L.) none of them was induced by a primary photon. Taking into account statistical fluctuations and differences in the energy estimation of proton and photon primaries, we derive an upper limit of 36% at 95% C.L. on the fraction of primary photons in the cosmic-ray flux above 10^20 eV. This result disfavors the Z-burst and superheavy dark-matter solutions to the GZK-cutoff problem.
https://export.arxiv.org/pdf/astro-ph/0601449
\title{Upper limit on the ultra-high-energy photon flux from AGASA and Yakutsk data} \author{G.I.~Rubtsov$^1$, L.G.~Dedenko$^{2,3}$, G.F.~Fedorova$^3$, E.Yu.~Fedunin$^3$, A.V.~Glushkov$^4$, D.S.~Gorbunov$^1$, I.T.~Makarov$^4$, M.I.~Pravdin$^4$, T.M.~Roganova$^3$, I.E. Sleptsov$^4$ and S.V.~Troitsky$^1$} \affiliation{$^1$Institute for Nuclear Research of the Russian Academy of Sciences, Moscow 117312, Russia} \affiliation{$^2$Faculty of Physics, M.V.~Lomonosov Moscow State University, Moscow 119992, Russia} \affiliation{$^3$D.V.~Skobeltsin Institute of Nuclear Physics, M.V.~Lomonosov Moscow State University, Moscow 119992, Russia} \affiliation{$^4$Yu.G.~Shafer Institute of Cosmophysical Research and Aeronomy, Yakutsk 677980, Russia} \date{January 13, 2006} \pacs{98.70.Sa, 96.40.De, 96.40.Pq} \section{Introduction} \label{sec:intro} One of the most intriguing puzzles in astroparticle physics is the observation of air showers initiated by particles with energies beyond the cutoff predicted by Greisen and by Zatsepin and Kuzmin~\cite{gzk}. Compared to lower energies, the energy losses of protons increase sharply at $\approx 5\times 10^{19}$~eV since pion production on cosmic microwave background photons reduces the proton mean free path by more than two orders of magnitude. This effect is even stronger for heavier nuclei, while photons are absorbed due to pair production on the radio background with the mean free path of a few Mpc. Thus, the cosmic-ray (CR) energy spectrum should dramatically steepen at $\approx 7\times 10^{19}$~eV for any homogeneous distribution of CR sources. Despite the contradictions in the shape of the spectrum, the existence of air showers with energies in excess of $10^{20}$~eV is firmly established by several independent experiments using different techniques (Volcano Ranch~\cite{exp}, Fly's Eye~\cite{Bird}, Yakutsk~\cite{YakutskExperiment}, AGASA~\cite{agasares}, HiRes~\cite{HiRes} and Pierre Auger~\cite{Auger} experiments). Some explanations for these showers, like the $Z$-burst or top-down models, predict a significant fraction of photons above typically $8\times 10^{19}$~eV (for reviews see, e.g., Refs.~\cite{reviews}). Indications for the presence of neutral particles at lower energies were found in Refs.~\cite{neutral}. Thus, the determination of the photon fraction in the CR flux is of crucial importance, and the aim of this work is to derive a stringent limit on this fraction in the integral CR flux above $10^{20}$~eV. To this end, we compare the reported information on signals measured by scintillation and by muon detectors for observed showers with those expected by air shower simulations. We focus on the surface detector signal density at 600 meters $S(600)$ (known as charged particle density) and the muon density at 1000 m, $\rho_{\mu}(1000)$, which are used in experiments as primary energy and primary mass estimators, respectively. We study individual events of AGASA~\cite{AGASA_Eest} and of the Yakutsk extensive air shower array (Yakutsk in what follows)~\cite{YakutskExperiment} with \textit{reconstructed} energies above $8\times 10^{19}$~eV and measured muon content. We reject the hypothesis that any of showers considered was initiated by a photon primary at the 95\% confidence level (C.L.). We then derive as our main result an upper limit of 36\% (at 95\%~C.L.) on the fraction $\epsilon_\gamma$ of primary photons with \textit{original} energies above $10^{20}$~eV (the difference between original and reconstructed energies is discussed in Sec.~\ref{sec:data}). The rest of the paper is organized as follows. In Sec.~\ref{sec:data} we discuss the experimental data set which we use for our study. In Sec.~\ref{sec:simulations}, the details of the simulation of the artificial shower libraries and comparison of the simulated and real data are given. This section contains the description of our method and the main results. We discuss how robust these results are with respect to changes in assumptions, to analysis procedure, and to variations in the experimental data, in Sec.~\ref{sec:robustness}. In Sec.~\ref{sec:comparison}, we discuss the differences between our approach and previous studies, which allowed us to put a significantly more stringent limit on the gamma-ray fraction. Our conclusions are briefly summarized in Sec.~\ref{sec:conclusions}. \section{Experimental data} \label{sec:data} AGASA was operating from 1990 to 2003 and consisted of 111 surface scintillation detectors (covering an area of about $100$~km$^2$) and 27 muon detectors. The areas of the AGASA muon detectors varied between 2.8 and 20~m$^2$. The detectors consisted of 14--20 proportional counters aligned under a shield of either 30~cm of iron or 1~m of concrete and were placed below or close to scintillation detectors. The threshold energy was 0.5~GeV$/\cos\theta_\mu$ for muons with zenith angle $\theta_\mu$~\cite{AGASAmu}. During 14 years of operation, AGASA had observed 11 events with reported energies above $10^{20}$~eV and zenith angles $\theta<45^\circ$~\cite{agasares,AGASAarrDir}. Among them, six events had $\rho_{\mu}(1000)$ determined~\cite{AGASAmu}. Yakutsk is observing CRs of highest energies since 1973, with detectors in various configurations. With {$\theta< 60^\circ$}, it has observed three events above $10^{20}$~eV, all with measured muon content. Before 1978, only one muon detector with the area of 8~m$^2$ and threshold energy $0.7$~GeV$/\cos\theta_\mu$ was in operation. Later, it has been replaced by six detectors with areas up to 36~m$^2$ and the threshold energy of $1.0$~GeV$/\cos\theta_\mu$~\cite{Yakutsk_mu}. In our study, we combine the AGASA and Yakutsk datasets, motivated by the following. First, both datasets are obtained from surface array experiments operated with similar plastic scintillation detectors. Second, the energy estimation procedures of the two experiments are compatible, within the reported systematic errors at $\sim 10^{20}$~eV, if differences in the observational conditions are taken into account~\cite{SakakiThesis}. Finally, the values of the CR flux at $10^{20}$~eV reported by the two experiments are consistent within their $1\sigma$ errors. The shower energy estimated by an experiment (hereafter denoted as $E_{\rm est}$) is in general different from the true primary energy (denoted as $E_0$) because of natural shower fluctuations, etc. Moreover, the energy estimation algorithms used by surface-array experiments normally assume that the primary is a proton. While the estimated energy for nuclei depends only weakly on their mass number, the difference between photons and hadrons is significant. For photons, the effects of geomagnetic field~\cite{GMF} result in directional dependence of the energy reconstruction. Thus, the event energy reported by the experiment should be treated with care when we allow the primary to be a photon. In this study we include events with $E_{\rm est} \ge 8\times 10^{19}$~eV because of possible energy underestimation for photon-induced showers; these events contribute to the final limit, derived for $E_0>10^{20}$~eV, with different weights. For AGASA, we use the events given in Ref.~\cite{agasares} that pass the ``cut B'' defined in Ref.~\cite{AGASAmu}, that is having at least one \footnote{We thank K.~Shinozaki for bringing a misprint in Ref.~\cite{AGASAmu} to our attention: ``more than one'' was written there.} muon detector hit between 800~m and 1600~m from the shower axis. The $\rho_\mu(1000)$ of the individual events can be read off from Fig.~2 of Ref.~\cite{AGASAmu}. Yakutsk muon detectors have larger area and are more sensitive both to weak signals far from the core and to strong signals for which AGASA detectors might become saturated. This allowed the Yakutsk collaboration to relax the cuts, as compared to AGASA, and to obtain reliable values of $\rho_\mu(1000)$ using detectors between 400~m and 2000~m from the shower axis~\cite{Knurenko,Yakutsk-muon-new}. Providing these cuts, six AGASA and four Yakutsk events entered the dataset in our study (see Table~\ref{events} for the event details). \begin{table*} \caption{\label{events} Description of the individual events used in this work. Columns: (1), event number; (2), experiment; (3), date of the event detection (in the format dd.mm.yyyy); (4), the reported energy assuming a hadronic primary (in units of $10^{20}$~eV); (5), the zenith angle (in degrees); (6) the azimuth angle (in degrees, $\phi =0$ corresponds to a particle coming from the South, $\phi =90^\circ$ -- from the West); (7) number of muon detectors used to reconstruct muon density; (8) muon density at 1000~m from the shower axis (in units of m$^{-2}$); (9), probability that this event was initiated by a photon with $E>10^{20}$~eV; (10), probability that this event was initiated by a non-photon with $E>10^{20}$~eV, assuming correct energy determination. The sum $p_1^{(i)} +p_2^{(i)}$ gives the weight of this event in the final limit on $\epsilon _\gamma $. The probability that the primary had the energy $E<10^{20}$~eV is $1-p_1^{(i)}-p_2^{(i)}$.} \begin{center} \begin{ruledtabular} \begin{tabular}{cccdddcddd} $i$ & Experiment&Date & \multicolumn{1}{c}{$E_{\rm obs}$} & \multicolumn{1}{c}{$\theta$} & \multicolumn{1}{c}{$\phi$}& \multicolumn{1}{c}{$n_{\rm det}$}& \multicolumn{1}{c}{$\rho_{\mu}^{(i)}(1000)$}& \multicolumn{1}{c}{$p_1^{(i)} $}& \multicolumn{1}{c}{$p_2^{(i)}$}\\ (1)& (2)&(3)& \multicolumn{1}{c}{(4)}& \multicolumn{1}{c}{(5)}& \multicolumn{1}{c}{(6)}& \multicolumn{1}{c}{(7)}& \multicolumn{1}{c}{(8)}& \multicolumn{1}{c}{(9)}& \multicolumn{1}{c}{(10)}\\ \hline 1 &AGASA & 10.05.2001 &2.46 & 36.5& 79.2 & 3& 8.9 &0.000 & 1.000\\ 2 &AGASA & 03.12.1993 &2.13 & 22.9& 55.5 & 1& 10.7 &0.001 & 0.998\\ 3 &AGASA & 11.01.1996 &1.44 & 14.2& 27.5 &$>1$&8.7 &0.013 & 0.921\\ 4 &AGASA & 06.07.1994 &1.34 & 35.1&234.9 & 1& 5.9 &0.003 & 0.887\\ 5 &AGASA & 22.10.1996 &1.05 & 33.7&291.6&$>1$&12.6 &0.000 & 0.581\\ 6 &AGASA & 22.09.1999 &1.04 & 35.6&100.0 &$>1$&9.3 &0.000 & 0.565\\ 7 &Yakutsk& 18.02.2004 &1.60 & 47.7&180.8 & 5& 19.6 &0.000 & 0.876\\ 8 &Yakutsk& 07.05.1989 &1.50 & 58.7&230.6 & 5& 11.8 &0.000 & 0.868\\ 9 &Yakutsk& 21.12.1977 &1.10 & 46.1&346.8 & 1& 8.0 &0.000 & 0.645\\ 10&Yakutsk& 02.05.1992 &0.85 & 55.7&163.0 & 5& 4.7 &0.000 & 0.303\\ \end{tabular} \end{ruledtabular} \end{center} \end{table*} \section{Simulations and results} \label{sec:simulations} In order to interpret the data, for each of the ten events, we generated a shower library containing 1000 showers induced by primary photons \footnote{For the illustration in Fig.~\ref{fig:event3}, 500 proton-induced showers were simulated and processed in a similar way.}. Thrown energies $E_0$ of the simulated showers were randomly selected (see below the discussion of the initial spectra) between $5\times 10^{19}$~eV and $5\times 10^{20}$~eV to take into account possible deviations of $E_{\rm est}$ from $E_0$. The arrival directions of the simulated showers were the same as those of the corresponding real events. The simulations were performed with CORSIKA~v6.204~\cite{Heck:1998vt}, choosing QGSJET~01c~\cite{Kalmykov:1997te} as high-energy and FLUKA~2003.1b~\cite{fluka} as low-energy hadronic interaction model. Electromagnetic showering was implemented with EGS4~\cite{Nelson:1985ec} incorporated into CORSIKA. Possible interactions of the primary photons with the geomagnetic field were simulated with the PRESHOWER option of CORSIKA~\cite{Homola:2003ru}. As discussed in Sec.~\ref{sec:robustness:models}, this choice of the interaction models results in a conservative limit on gamma-ray primaries. As suggested in Ref.~\cite{Thin}, all simulations were performed with thinning level $10^{-5}$, maximal weight $10^6$ for electrons and photons, and $10^4$ for hadrons. For each simulated shower, we determined $S(600)$ and $\rho_{\mu}(1000)$. For the calculation of $S(600)$, we used the detector response functions from Refs.~\cite{Sakaki,YakutskGEANT}. For a given arrival direction, there is one-to-one correspondence between $S(600)$ and the quantity called estimated energy, $E_{\rm est}$. The relation is determined by the standard analysis procedure of the two experiments~\cite{AGASA_Eest,Yakutsk_Eest}. This allows us to select simulated showers compatible with the observed ones by the signal density. The quantity $S(600)$ is reconstructed not precisely. In terms of estimated energy, for AGASA events, the reconstructed energies are are distributed with a Gaussian in $\log \left(E_{\rm est}/\bar E_{\rm rec}\right)$; the standard deviation of $E_{\rm est}$ is $\sigma\approx 25\%$ \cite{SakakiThesis}. For Yakutsk events, the corresponding $\sigma$ has been determined event-by-event and is typically 30--45\% \cite{PravdinICRC2005}. To each simulated shower, we assigned a weight $w_1$ proportional to this Gaussian probability distribution in $\log E_{\rm est}$ centered at the observed energy $\bar E_{\rm rec}=E_{\rm obs}$. Additionally, each simulated shower was weighted with $w_2$ to reproduce the thrown energy spectrum $\propto E_0^{-2}$ which is typically predicted by non-acceleration scenarios (see Sec.~\ref{sec:robustness:spectrum} for a discussion of the variations of the spectral index). For each of the ten observed events, we obtained a distribution of muon densities $\rho _\mu (1000)$ representing photon-induced showers compatible with the observed ones by $S(600)$ and arrival directions. To this end, we calculated $\rho _\mu (1000)$ for each simulated shower by making use of the same muon lateral distribution function as used in the analysis of real data~\cite{AGASAmu,Yakutsk_mu}. To take into account possible experimental errors in the determination of the muon density, we replaced each simulated $\rho _\mu (1000)$ by a distribution representing possible statistical errors (50\% and 25\% Gaussian for AGASA cut B~\cite{AGASAmu50} and Yakutsk~\cite{Yakutsk-muon-new}, respectively). The distribution of the simulated muon densities is the sum of these Gaussians weighted by $w_1w_2$. A typical distribution of simulated $\rho _\mu (1000)$ is given in Fig.~\ref{fig:event3}, for gamma- and proton-induced simulated showers compatible with the event 3 by $S(600)$ and the arrival direction. We will see below that this particular event has the largest probability of gamma interpretation among all ten events in the data set; still the proton interpretation looks perfect for it. This is the case for all events except event 7, which has too high $\rho _\mu (1000)$ for a proton; possible nature of its primary particle will be discussed elsewhere. To estimate the allowed fraction $\epsilon_\gamma $ of primary photons among CRs with $E_0>10^{20}$~eV, we compare, for each observed event, two possibilities: (i)~that it was initiated by a photon primary with $E_0>10^{20}$~eV and (ii)~that it was initiated by any other primary with $E_0>10^{20}$~eV for which the experimental energy estimation works properly. Let us consider the $i$th observed event. Denote by $M$ the weighted number of showers contributed to the $\rho _\mu (1000)$ distribution for the simulated photon-induced showers compatible with the $i$th event by arrival direction and $S(600)$ (throughout this paragraph, the weighted number is the sum of corresponding weights, that is $M$ is the sum of weights of all 1000 showers simulated for the $i$th event). Some of the simulated showers contributed to the part of the distribution for which $\rho _\mu (1000)>\rho _\mu ^{(i)}(1000)$, where $\rho _\mu ^{(i)}(1000)$ is the observed value for this event. The weighted number of these showers is $M'$. Some part $l$ of this $M'$ corresponds to showers with $E_0>10^{20}$~eV, the rest ($M'-l$) to $E_0<10^{20}$~eV. The probability $p_1^{(i)}$ of case (i) is $p_1^{(i)}=l/M$, while the probability that the event is consistent with a photon of $E_0<10^{20}$~eV is $p_1^{\prime(i)}=(M'-l)/M$. Moreover, the probability that the event is described by any other primary is $1-p_1^{(i)}-p_1^{\prime(i)}=1-M^\prime/M$. We assume that the experimental energy estimation works well for non-photon primaries and determine the fraction $\xi$ of events with $E>10^{20}$~eV simply from the Gaussian $\log(E_{\rm est})$ distribution, so the probability of the case (ii) is $p_2^{(i)}=\xi(1-M^\prime/M)$. The values of $p_{1,2}^{(i)}$ are presented in Table~\ref{events}. Note that $p_1^{(i)}+p_2^{(i)}<1$ because of a non-zero probability that a simulated shower is initiated by a primary with $E_0<10^{20}$~eV. This happens especially for events with reported energies close to $10^{20}$~eV and reduces considerably the effective number of events contributing to the limit on $\epsilon_\gamma$: since we are interested in the limit for $E_0>10^{20}$~eV only, each event contributes to the result with the weight $(p_1^{(i)}+p_2^{(i)})$. Inspection of Table~\ref{events} demonstrates that the total effective number of events with $E_0>10^{20}$~eV (the sum of $p_1^{(i)}$ and $p_2^{(i)}$ over all ten events) is 7.67. If the $i$th primary particle was a photon with $E_0>10^{20}$~eV with the probability $p_1^{(i)}$ and a non-photon with $E_0>10^{20}$~eV with the probability $p_2^{(i)}$, one can easily calculate the probability ${\cal P}(n_1,n_2)$ to have $n_1$ photons and $n_2$ non-photons in the set of $N=10$ observed events ($0\le n_1+n_2 \le N$, the rest $N-n_1-n_2$ events have $E_0<10^{20}$~eV). From the set of $N$ events, one should take all possible non-overlapping subsets of $n_1$ and $n_2$ events and sum up probabilities of these realisations (since $p_{1,2}^{(i)}\ne p_{1,2}^{(j)}$, these probabilities are different for different realisations with the same $n_1$ and $n_2$). Now, suppose that the fraction of the primary photons at $E_0>10^{20}$~eV is $\epsilon _\gamma $. Then, the probability to have $n_1$ photons and $n_2$ non-photons at $E_0>10^{20}$~eV is $ \epsilon _\gamma ^{n_1} \left(1-\epsilon _\gamma \right)^{n_2} $, and the probability that the observed muon densities were obtained with a given $\epsilon _\gamma $ is $$ {\cal P}(\epsilon _\gamma )=\sum\limits_{n_1,n_2=0}^N {\epsilon_\gamma} ^{n_1} \left(1-\epsilon_\gamma \right)^{n_2} {\cal P}(n_1,n_2) \, $$ (cf.\ Ref.~\cite{Homola} for a particular case $n_1+n_2=N$; note that the combinatorial factor is included in the definition of ${\cal P}(n_1,n_2)$). The cases $n_1+n_2<N$ reflect the possibility that some of the $N$ events correspond to primaries with $E_0<10^{20}$~eV. In our case, the probability ${\cal P}(\epsilon_\gamma)$ is a monotonically decreasing function of $\epsilon_\gamma$. Thus the upper limit on $\epsilon_\gamma$ at the confidence level $\alpha'$ is obtained by solving the equation ${\cal P}(\epsilon_\gamma )=1-\alpha'$. For our dataset, the 95\%~C.L.\ upper limit on the photon fraction is $\epsilon_\gamma<0.33$. The limit on $\epsilon_\gamma$ is rather weak compared to the individual values of $p_1^{(i)}$ because of the small number of observed events. However, some of the photon-induced showers may escape from our study because they may not pass the muon measurement quality cuts or their estimated energy is below $8\times 10^{19}$~eV. Possible reasons for an underestimation of the energy may be either the LPM effect~\cite{LPM} or substantial attenuation of gamma-induced showers at large zenith angles. To estimate the fraction of these ``lost'' events, we have simulated 1000 gamma-induced showers for each experiment with arrival directions distributed according to the experimental acceptance. We find that the fraction of the ``lost'' events is $\sim 3.5\%$ for AGASA and $\sim 15\%$ for Yakutsk. The account of these fractions, weighted with the relative exposures of both experiments, results in the final upper limit, $$ \epsilon_\gamma <36\% ~~(95\%~{\rm C.L.}). $$ In Fig.~\ref{fig:limits}, we present our limit on $\epsilon_\gamma $ (AY) together with previously published limits on the same quantity. Also, typical theoretical predictions are shown for the superheavy dark-matter, topological-defect and $Z$-burst models. Our limit on $\epsilon_\gamma $ is currently the strongest one at $E_0>10^{20}$~eV. It disfavours some of the theoretical models such as the $Z$-burst and superheavy dark-matter scenarios. \section{Robustness of the results} \label{sec:robustness} In this section, we discuss systematic uncertainties of our limit that are related to the air shower simulations, to the data interpretation and to selection cuts. \subsection{Systematic uncertainty in the $S(600)$ and energy determination} \label{sec:robustness:energy} The systematic uncertainty in the absolute energy determination is 18\% and 30\% for AGASA~\cite{AGASA_Eest} and Yakutsk~\cite{YakutskExperiment}, respectively. These systematic errors originate from two quite different sources: (a)~the measurement of $S(600)$ and (b)~the relation between $S(600)$ and primary energy. The probabilities $p_1^{(i)}$ that a particular event may allow for a gamma-ray interpretation are not at all sensitive to the $S(600)$-to-energy conversion because we select simulated events by $S(600)$ and not by energy. These probabilities may be affected by relative systematics in determination of $\rho _\mu (1000)$ and $S(600)$. On the other hand, in the calculation of $p_2^{(i)}$ we assumed that the experimental energy determination is correct for non-photon primaries; the values of $p_2^{(i)}$ and the effective number of events contributing to the limit on $\epsilon _\gamma $ at $E_0>10^{20}$~eV would change if the energies are systematically shifted. In our case (all $p_1^{(i)}\approx 0$), the reported value of $\epsilon _\gamma $ would be applicable to the shifted energy range in that case. Thus, the 95\%~C.L. conclusion that none of the ten events considered here was initiated by a photon is robust with respect to any changes in the $S(600)$-to-energy conversion. As for the limit on $\epsilon _\gamma $ we report, instead of $E_0>10^{20}$~eV, it would be applicable to a different energy range if all experimental energies are systematically shifted. One should note that theoretical predictions, e.g. the curves shown in Fig.~\ref{fig:limits}, would also change because they are normalised to the observed AGASA spectrum. \subsection{Interaction models and simulation codes} \label{sec:robustness:models} Our simulations were performed entirely in the CORSIKA framework, and any change in the interaction models or simulation codes, which affects either $S(600)$ or $\rho _\mu (1000)$, may affect our limit. We have studied the model dependence of our results by comparing different low- and high-energy hadronic interaction models (GHEISHA~\cite{GHEISHA} versus FLUKA, SIBYLL 2.1~\cite{SIBYLL} versus QGSJET). Our result is quite stable with respect to these changes. In all cases, individual values of $p_1^{(i)}$ are always close to zero, thus the limit on $\epsilon_\gamma$ is not affected. The change of the low energy model does not at all affect the reported values. In use of SIBYLL compared with QGSJET, $\rho_\mu(1000)$ is $\sim 20\%$ smaller for photon- induced showers. While $S(600)$ is almost unchanged, events in our dataset are better explained by showers initiated by heavier nuclei and the probability of photon-induced showers is even smaller. A similar effect is expected for the coming interaction model QGSJET~II~\cite{QGSJET-II}. We also performed simulations with the help of the hybrid code~\cite{hybrid} which reproduced the CORSIKA results with high accuracy. Another popular simulation code, AIRES~\cite{Aires}, differs from CORSIKA mainly in the low-energy hadronic interaction model (which is fixed in AIRES to be the Hillas splitting algorithm), hence we hope that simulations with AIRES would not significantly affect our results. Comparison with AIRES will be presented elsewhere. The values presented here were obtained for the standard parameterization of the photo-nuclear cross section given by the Particle Data Group~\cite{PDG} (implemented as default in CORSIKA). The muon content of gamma-induced showers is in principle sensitive to the extrapolation of the photonuclear cross section to high energies. The hybrid code~\cite{hybrid} allows for easy variations of the cross section; we checked that the results are stable for various reasonable extrapolations, in agreement with Ref.~\cite{RissePgamma}. \subsection{Primary energy spectrum} \label{sec:robustness:spectrum} For our limit, we used the primary photon spectrum $E_0^{-\alpha }$ for $\alpha =2$. While the individual probabilities $p_{1,2}^{(i)}$ are not affected by the change of the spectral index $\alpha$ because the simulated events are selected by $S(600)$ anyway, the value of $\alpha $ changes the fraction of ``lost'' photons and, correspondingly, the final limit on $\epsilon _\gamma $. Variations of $1\le \alpha \le 3$ result in the photon fraction limits between 36\% and 37\% (95\%~C.L.). \subsection{Width of the $\rho _\mu $ distribution} \label{sec:robustness:width} Clearly, the rare probabilities of high values of $\rho _\mu (1000)$ in the tail of the distribution for primary photons depend on the width of this distribution. The following sources contribute to this width: \begin{itemize} \item variations of the primary energy compatible with the observed $S(600)$ (larger energy correspond to larger muon number and $\rho _\mu (1000)$); \item physical shower-to-shower fluctuations in muon density for a given energy (dominated by fluctuations in the first few interactions, including preshowering in the geomagnetic field); \item artificial fluctuations in $S(600)$ and $\rho _\mu (1000)$ due to thinning; \item experimental errors in $\rho _\mu (1000)$ determination. \end{itemize} While the first two sources are physical and are fully controlled by the simulation code, the variations of the last two may affect the results. \subsubsection{Artificial fluctuations due to thinning} \label{sec:robustness:thinning} It has been noted in Ref.~\cite{Badagnani} that the fluctuations in $\rho _\mu (1000)$ due to thinning may affect strongly the precision of the composition studies. For the thinning parameters we use, the relative size of these fluctuations is~\cite{our-thinning} $\sim 10\%$ for $\rho _\mu (1000)$ and $\sim 5\%$ for $S(600)$. Thus with more precise simulations, the distributions of muon densities should become more narrow, which would reduce the probability of the gamma-ray interpretation of each of the studied events even further. \subsubsection{Experimental errors in $\rho _\mu (1000)$ determination} \label{sec:robustness:muon-errors} The distributions of $\rho _\mu (1000)$ we use accounted for the error in experimental determination of this quantity. The size of the errors was taken from the original experimental publications~\cite{Yakutsk-muon-new,AGASAmu50}. In principle, this error depends on the event quality and on the muon number itself, which is lower for simulated gamma-induced showers than for the observed ones. However, e.g.\ Ref.~\cite{AGASAmu} states that for the AGASA cut A (two or more muon detectors), the error is 40\%, lower than 50\% we use~\cite{AGASAmu50}. Note that Ref.~\cite{AGASAmu} discusses muon densities as low as 0.04~m$^{-2}$ and even 0~m$^{-2}$, much lower than $\sim 1$~m$^{-2}$ typical for our simulated gamma-induced events. Still, we tested the stability of our limit by taking artificially high values of experimental errors in muon density: 100\% for AGASA and 50\% for Yakutsk. The limit on $\epsilon _\gamma $ changes to 37\% (95\%~C.L.) in that case. \subsection{Data selection cuts} \label{sec:robustness:cuts} Since all events in the data set are unlikely to be initiated by primary photons (all $p_1^{(i)}\approx 0$), the limit on $\epsilon _\gamma $ is determined by statistics only and is affected if the number of events is changed. Here, we discuss possible variations of the data set corresponding to more stringent quality cuts which reduce the event number and weaken the limit. \subsubsection{Zenith angle} \label{sec:robustness:ZA} All Yakutsk events in the data set have zenith angles $45^\circ<\theta <60^\circ$, so the cut $\theta<45^\circ$ imposed by AGASA reduces the sample to six AGASA events which results in the limit $\epsilon _\gamma <50\%$ (95\%~C.L.). One should note however that AGASA muon detectors are not sensitive to inclined showers, which is not the case for Yakutsk. \subsubsection{Core inside array} \label{sec:robustness:event7} Another cut imposed on the AGASA published dataset is the location of the core inside array. The event number 7 does not satisfy this criterion; its exclusion from the data set results in $\epsilon _\gamma <40\%$ (95\%~C.L.). \subsubsection{More than one muon detector} Reconstruction of the muon density at 1000~m from a single muon detector reading requires extrapolation of the lateral distribution function with an averaged slope. Though it is well-studied, the data points corresponding to events with a single muon detector hit might be considered less reliable than those with two or more hits. With the account of the events with two or more hits only, we are left with seven events (four AGASA and three Yakutsk) which weakens the 95\%~C.L.\ limit to $\epsilon _\gamma <48\%$. \section{Comparison with other studies} \label{sec:comparison} Some of the previous studies used the AGASA~\cite{AGASAmu,Homola} and Yakutsk~\cite{Knurenko} muon data to limit the gamma-ray primaries at high energies. Our results differ from the previous ones not only because we join the data sets of the two experiments. Two major distinctive features of our approach allowed us to put the stringent limit: \begin{itemize} \item both $\rho _\mu (1000)$ and $S(600)$ were tracked for simulated showers within framework of a {\em single } simulation code (CORSIKA in our case); \item each event was studied individually, without averaging over arrival directions. \end{itemize} In Refs.~\cite{AGASAmu,Knurenko}, no conclusion was derived about $\epsilon _\gamma $ at $E>10^{20}$~eV, and the data points corresponding to highest-energy events were found to be quite close to the gamma-ray domain. To our opinion, the main source of this effect is averaging over arrival directions which introduced additional fluctuations for gamma-ray primaries due to direction-dependent preshowering (see Fig.~\ref{fig:2events} for an illustration). In Ref.~\cite{Homola} which discussed the same six AGASA events, all simulated showers for an event with the observed energy $E_{\rm obs}$ had energies $1.2 E_{\rm obs}$ (up to the energy reconstruction uncertainty of 25\%). This conversion had been obtained as the average over $\theta<36^\circ$ in Ref.~\cite{AGASAmu} using AIRES simulation code~\cite{Aires}. That is, not only the average results were applied to individual showers, but effectively muon densities were simulated with CORSIKA while energies -- with AIRES, though the two codes result in a systematically different relations between energy and $S(600)$. Artificially high energies resulted in higher, closer to observed, muon densities for simulated photonic showers. In our event-by event simulations with CORSIKA, the energies of gamma-ray primaries whose $S(600)$ were compatible to observed values, were not higher by a factor 1.2, but in fact even lower than $E_{\rm obs}$ for some of the events: besides the difference in simulation codes, this is partially due to non-uniform distribution of the highest-energy AGASA events on the celestial sphere~\cite{AGASAarrDir,NorthSouth} which makes the usage of averaged energies poorly motivated. The impact of two other sources of difference between our approach and that of Ref.~\cite{Homola} is less important for the final result: (i)~Ref.~\cite{Homola} does not account for the ``lost'' photons and (ii)~the detector error is applied in our study to the simulated events while in Ref.~\cite{Homola} -- to the observed ones. The difference with Ref.~\cite{Homola} is illustrated in Fig.~\ref{fig:DifferentEnergies}, where $\rho _\mu (1000)$ is plotted versus $E_0$ for simulated gamma-induced showers with the arrival direction of the event \#1. For simulated events compatible with the real event by $S(600)$, the average point is shown together with one sigma error bars. Horizontal error bars correspond to variations in $E_0$ compatible with $S(600)$. Vertical error bars include variations in simulated $\rho _\mu (1000)$ and 50\% detector error. The point corresponding to simulated showers with $E_0=1.2E_{\rm obs}$ from Ref.~\cite{Homola} has a larger $\rho _\mu (1000)$. Horizontal error bars correspond to the energy reconstruction accuracy. Vertical error bars include variations in simulated $\rho _\mu (1000)$ reported in Ref.~\cite{Homola} and 40\% detector error applied to the observed value, added in quadrature. We see that the main source of the disagreement is in the values of $E_0$ which push, for the case of Ref.~\cite{Homola}, the simulated muon densities closer to the observed one. \section{Conclusions} \label{sec:conclusions} To summarize, we have studied the possibility that the highest-energy events observed by the AGASA and Yakutsk experiments were initiated by primary photons. Comparing the observed and simulated muon content of these showers, we reject this possibility for each of the ten events at $E>8\cdot 10^{19}$~eV at least at the 95\% C.L. An important ingredient in our study is the careful tracking of differences between the original and reconstructed energies. This allows us to put an upper bound of 36\% at 95\% C.L.\ on the fraction $\epsilon_\gamma $ of primary photons with original energies $E_0>10^{20}$~eV, assuming an isotropic photon flux and $E_0^{-2}$ spectrum. This limit is the strongest one up to date. It strongly disfavors the $Z$-burst and constrains severely superheavy dark-matter models. The method that we have used is quite general and may be applied at other energies and to other observables. We are indebted to M.~Kachelrie\ss, K.~Shinozaki and M.~Teshima for numerous helpful discussions and collaboration at initial stages of this work. We thank L.~Bezroukov, V.~Bugaev, R.~Engel, D.~Heck, A.~Ringwald, M.~Risse, V.~Rubakov, D.~Semikoz and P.~Tinyakov for helpful discussions and comments on the manuscript. This study was performed within the INTAS project 03-51-5112. We acknowledge also support by fellowships of the Russian Science Support Foundation and of the Dynasty foundation (D.G.\ and S.T.), by the grants NS-2184.2003.2 (D.G., G.R.\ and S.T.), NS-1782.2003.2, RFFI 03-02-16290 (L.D., G.F., E.F.\ and T.R.), NS 748.2003.2, RFFI 03-02-17160, RFFI 05-02-17857, FASI 02.452.12.7045 (A.G., I.M., M.P.\ and I.S.). G.R.\ and S.T.\ thank the Max-Plank-Institut f\"ur Physik (Munchen), where a significant part of this work was done, for warm hospitality. Computing facilities of the Department of Theoretical Physics, Institute for Nuclear Research (Moscow), were used to perform the simulations of air showers.
Title: Sersic Properties of Disc Galaxy Mergers
Abstract: Sersic parameters characterising the density profiles of remnants formed in collision-less disc galaxy mergers are obtained; no bulge is included in our simulations. For the luminous component we find that the Sersic index is n ~ (1.5,5.3) with <n> ~ 3 +/- 1 and an effective radius of R_e ~ (1.6,12.9) kpc with <R_e> ~ 5 +/- 3 kpc. A strong correlation of n with the central projected density I_0 is found [n ~ I_0^(-0.14)] which is consistent with observations. No positive linear correlation between the size (R_e) and structure (n) of our remnants is found; we do not advocate the existence of this. The photometric plane (PHP) of the luminous component [n ~ R_e^(0.05) I_0^(0.15)] agrees well, within the uncertainties and the assumption of a constant mass-to-light ratio, with those observationally determined for ellipticals. We found that the surface defined by Sersic parameters (n, R_e, mu_0) in log-space is not a true plane, but a pseudo-plane with a small curvature at low values of n owed to intrinsic properties of the Sersic model. The dark haloes of the remnants have a 3-dimensional Sersic index of <n> ~ 4 +/- 0.5 that are smaller than the ones obtained for dark haloes in LCDM cosmologies <n> ~ 6 +/- 1. A tight dark Sersic ``plane'' (DSP) is also defined by the parameters of the remnants haloes with n \~ r_e^(0.07) rho_0^(0.10). We conclude that collision-less merger remnants of pure disc galaxies have Sersic properties and correlations consistent with those of observed in early-type galaxies and local remnants. It seems that a ``primordial'' bulge in spirals is not a necessary condition to form bona fide ellipticals on grounds of the Sersic properties of remnants.
https://export.arxiv.org/pdf/astro-ph/0601412
\date{Accepted ------. Received ------; in original form ------} \pagerange{\pageref{firstpage}--\pageref{lastpage}} \pubyear{2004} \label{firstpage} \begin{keywords} galaxies: kinematics and dynamics -- galaxies: formation -- galaxies: fundamental parameters -- galaxies: interactions -- galaxies: elliptical -- methods: $N$-body simulations. \end{keywords} \section{Introduction} Hierarchical galaxy formation theory (e.g. Cole~et~al.~2000, De~Lucia~et~al.~2005, Bower~et~al.~2005) considers that early-type galaxies have an accretion/merger origin, as was originally suggested by Toomre (1977). Observational (e.g. Schweizer~1998, Struck~2005, Rothberg \& Joseph~2006, Kaviraj~et~al.~2006) and theoretical (e.g. Naab \& Burkert 2003, Meza~et~al.~2005, Naab~et~al.~2005) evidence supports this picture although several topics remain unsolved (e.g. Peebles~2002, Tantalo \& Chiosi~2004). Early-type galaxies show several correlations among their colours, luminosities, velocity dispersions, effective radii and surface brightness (e.g. Baum~1990, Faber~\&~Jackson~1976, Kormendy~1977, Djorgovski~\& Davis~1987, Dressler~et~al.~1987, Bernardi~et~al.~2003). These correlations provide constraints to any theory of formation and evolution of these galaxies. Furthermore, their properties are linked with the distribution of luminous and dark matter, that would be important when comparing with models of formation of elliptical galaxies. Observational studies [e.g. Caon, Capaccioli \& D'Onofrio 1993 (CCD93), Graham \& Colles~1997, Binggeli \& Jerjen 1998, D'Onofrio~2001 (D01), Trujillo~et~al.~2004] have found that the surface brightness density profiles of early-type galaxies are better described by a S\'ersic (1968) $R^{1/n}$--profile than the classical de~Vaucouleurs (1948) $R^{1/4}$--profile. The index $n$ is directly related with the cur\-va\-ture and ``concentration'' of the light profile (Trujillo, Graham \& Caon~2001). Several observational relationships have been found between the index $n$ and, for example, the total luminosity ($L$), effective radius ($R_{\rm e}$) and central velocity dispersion (e.g. CCD93, Prugniel \& Simien~1997, Graham \& Guzm\'an~2003). Also, it has been found a linear relation among $\log n$, $\log R_{\rm e}$ and $\mu_0$ (central brightness) termed the Photometric Plane (PHP) for early-type galaxies [e.g. Khosroshahi~et~al.~2000 (K00), Graham~2002], analogous to the Fundamental Plane (Djorgovski \& Davis 1987, Dressler~et~al.~1987). Recently, Rothberg~\&~Joseph (2004, RJ04) have found that nearby merger remnants have a peak in the $n$-distribution at $n\approx 2$ with most values in the range of $1<n<6$, and in some cases it is found that $n > 8$. On other hand, theoretical studies of S\'ersic properties of merger remnants have appeared recently. For example, G\'onzalez-Garc\'{\i}a~\&~Balcells (2005, GGB) and Naab~\&~Trujillo~(2005, NT) find in collision-less simulations that bulge-less progenitors lead to ranges of $n\! \in\! (2.4,3.2)$ and $(1.2,3.1)$, respectively; when a single S\'ersic function is used to fit the entire remnant. For progenitors with a bulge component they obtain about the same range of S\'ersic index, $n\! \in \!(3,8)$. Since ``bona fide'' ellipticals have values $n\gta 4$, they reach the conclusion that collision-less merger remnants of pure disc galaxies do not lead to concentrations, indicated by $n$, similar to those found in intermediate or giant elliptical galaxies (e.g. Graham~et~al.~1996). The above findings suggest that a primordial bulge in spirals is a necessary condition to form bona fide ellipticals in the hierarchical merging scenario. However, we show below, collision-less mergers of pure discs can cover the range of observed values of the shape parameter $n$, and can reproduce adequately other observational correlations. S\'ersic model in a de-projected form has been recently used to represent the dark matter distribution in $\Lambda$CDM haloes (Navarro~et~al.~2004, Merritt~et~al.~2005, Graham~et~al.~2005, Prada~et~al.~2005), in order to have a better estimation of the inner asymptotic logarithmic derivative. A mean value of a 3D S\'ersic index $\approx 6$, with a scatter of $\approx 1$, has been found in these works. So it is of interest to determine the three-dimensional S\'ersic parameters that characterise our remnants. In this work, we study the structural properties of remnants as provided by fitting a S\'ersic profile to their luminous and dark mass distribution. The paper has been organised as follows: in $\S$\ref{sec:model} we present a summary of the properties of our progenitors, some details of the simulations performed, as well as some basic characteristics of S\'ersic profile; both projected and deprojected. In $\S$\ref{sec:results} we present distributions and correlations, in two and three-dimensions, found among the different S\'ersic parameters for our remnants, and compare them with observations. S\'ersic properties of the dark haloes of the remnants are determined, some correlations presented, and compared with those obtained in cosmological simulations. Some final comments are given in $\S$\ref{sec:discussion} and a summary of our conclusions. \section{Simulations and S\'ersic Functions}\label{sec:model} \subsection{Galaxy models} The galaxy models used in this work have been already described in Aceves \& Vel\'azquez (2005) and follow the method outlined by Shen, Mo \& Shu (2002) to obtain the global properties of the discs, once the haloes properties are known. Our numerical galaxies do \emph{not} include a bulge-like component. The dark haloes follow a modified NFW (Navarro, Frenk \& White~1997) model with an exponential cutoff. The discs have a typical exponential density profile, and satisfy the Tully-Fisher relation at redshift $z\!=\! 1$; roughly a look-back time of $8\,$Gyr in a $\Lambda$CDM cosmology with Hubble parameter $h=0.7$. Only discs satisfying the Efstathiou, Lake \& Negroponte~(1982) stability criterion were used. In this work, an additional simulation to those reported in Table~1 of Aceves \&~Vel\'azquez (2005) has been done. This is a merger from the resulting remnants of $M01$ and $M05$, label $MM$. All simulations were carried out using a parallel version of {\sc gadget}-1.1 code, a tree base code (Springel, Yoshida \& White 2001), and evolved for $\approx \! 8\,$Gyr with conservation of energy better than $0.25$ percent. \subsection{Density Profiles} We fit only S\'ersic profiles to our merger remnants; no bulge-disc decomposition is attempted since progenitors lack any bulge component. The S\'ersic surface luminous-mass density profile is given by \begin{equation} \Sigma(R) = \Sigma_0 \, {\rm e}^{ -b (R/R_{\rm e})^{1/n}}\;, \label{eq:sersic} \end{equation} where $R$ is the projected spherical radius, $R_{\rm e}$ is the effective radius, $n$ the index of the profile, $b=b(n)\approx 2n - 0.324$ (Ciotti \& Bertin 1999) and $\Sigma_0$ the central surface density. Index $n$ is associated with the curvature and the concentration of the profile (Trujillo, Graham \& Caon~2001); $n\!=\!1$ corresponds to an exponential profile while the classical de Vaucouleurs (1948) profile is obtained for $n=4$. The accumulated projected luminous mass, $M_{\rm L}(R)$, is given by \begin{equation} M_{\rm L}(R)= \int_0^R \Sigma(R) {\rm d}(\pi R^2) = \frac{2\pi n \gamma(\alpha,x)}{b^{2n}}\, \Sigma_0 R^2_{\rm e} \,, \end{equation} where $\alpha \equiv 2n$, $x \equiv b(R/R_{\rm e})^{1/n}$, and $\gamma (\alpha,x)$ is the incomplete gamma function. The total projected luminosity mass is given by \begin{equation} M_{\rm L} = \frac{2\pi n}{b^{2n}} \Gamma(2n) \, \Sigma_{\rm 0} R^2_{\rm e}\, , \label{eq:Lmass} \end{equation} being $\Gamma(\alpha)$ the complete gamma function. A summary of S\'ersic projected profile properties is given by Graham \& Driver (2005). When comparing our simulations with observations we assume a constant mass-to-light ratio, so that $\Sigma \!\propto \! I$; where $I$ refers to the surface brightness. The three-dimensional (3D) S\'ersic profile is \begin{equation} \rho(r) = \rho_{\rm 0} \, {\rm e}^{ -d (r/r_{\rm e})^{1/n}}\;, \label{eq:3Dsersic} \end{equation} where $r$ is the spatial radius, $d\approx 3n-1/3+0.005/n^2$ (Graham~et~al.~2005) such that $r_{\rm e}$ is the half-mass spatial radius. The total mass is determined from \begin{equation} M_{\rm t} = \frac{4\pi n}{d^{3n}} \Gamma(3n) \, \rho_0 r^3_{\rm e} \,. \label{eq:3Dmass} \end{equation} S\'ersic parameters for the luminous component were computed along 400 different random line-of-sights. To each projection a circularly averaged density profile $\Sigma(R)$ was determined, and a S\'ersic profile (\ref{eq:sersic}) fitted by $\chi^2$--minimisation using the Levenberg-Marquardt method (Press~et~al.~1992) to obtain $\{n,R_{\rm e},\Sigma_0\}$. S\'ersic parameters for dark haloes are obtained by a similar procedure, but using equation (\ref{eq:3Dsersic}). \subsubsection{Fitting Range} The fitting set of parameters depend on the methodology used to obtain them. In particular, there have been indications that these parameters depend on both the covered range of surface brightness range (e.g. Capaccioli, Caon \& D'Onofrio 1992) and the spatial radial interval for fitting (e.g. Kelson~et~al.~2000). Also, the determination of fitted parameters degrades when the inner parts of a galaxy are not well considered. For example, the index $n$ tends more to be a representation of the outer slope of the profile than of the curvature of the luminosity distribution (Graham~et~al.~1996). The treatment and quality of data has also an effect on the fitted parameters. For example, CCD93 obtain higher values of $n$ for NGC~4406, NGC~4552 and NGC~1399 (14.9, 13.9, 16.8) in comparison with D01 ($6.5$, $7.2$, $6.1$). We have considered two radial intervals for our fits in order to asses their effect on the S\'ersic parameters. The first radial interval, $I_1$, is taken from our numerical resolution value $\xi_{\rm i}=100\,$pc to the outer radius $\eta_{95}$, which encloses 95 percent of the projected luminous mass and is determined directly from the simulations; thus, $I_1=[\xi_{\rm i},\eta_{95}]$. The second one, $I_2$, uses another inner point at $\xi_{\rm f}=10\xi_{\rm i}$,\footnote{ \footnotesize For reference, in a $\Lambda$CDM cosmology with $h=0.7$ we have that $1''=464\,{\rm pc}$ at the distance of the Coma cluster ($z\!=\!0.023$), $977\,$pc at $z\!=\!0.05$, and $4.5\,$kpc at $z\!=\!0.3\,$.} and outer point at $\eta_{70}$; this enclosing 70 percent of the luminous mass. For each line-of-sight used, two uniform random numbers $\xi \in [\xi_{\rm i},\xi_{\rm f}]$ and $\eta \in [\eta_{70},\eta_{95}]$ are generated that in turn define $I_2=[\xi,\eta]$. In the Appendix we discuss some effects the radial range of a fit has on the parameters estimated using synthetic models. \section{Results}\label{sec:results} In this section we present the results of the fittings done, both ``luminous'' and dark, to the merger remnants, as well as several relationships among them based in observational studies. Table~\ref{tab:global} lists different global physical properties of our remnants obtained directly from the $N$-body simulations. Column (2) is the total half-mass radius $R_{\rm h}$, (3) the virial radius $R_{\rm v}$, (4) the virial velocity $V_{\rm v}$, (5) the total luminous mass ${\rm M}_{\rm lum}$ and (6) the total bounded mass ${\rm M}_{\rm tot}$, and column (7) is the virial ratio at the end of the simulation. The last column (8) provides the ratio of the total mass of the secondary to the primary galaxy in the simulations. The merger labelled as $MM$ corresponds to the simulation where the resulting remnants of $M01$ and $M05$ were merged together in a parabolic encounter. \begin{table} \begin{minipage}{80mm}% \caption{Physical properties of remnants}\label{tab:global} \begin{tabular}{lcrrrrll} \hline {\sc id}& $R_{\rm h}$ & $R_{\rm v}$ & $V_{\rm v}$ & $ \frac{{\rm M}_{\rm lum}}{10^{10}}$ & $ \frac{ {\rm M}_{\rm tot}}{10^{11} }$ & $\frac{2T}{|W|}$ & $\frac{2}{1}$\\ & [kpc] & [kpc] & [km/s] & [M$_\odot$] & [M$_\odot$] & & \\ \hline $M01$ & 66.9 & 156.1 & 213.0 & 10.00 & 16.60 & 0.99 & 0.32\\ $M02$ & 29.8 & 71.4 & 108.5 & 0.60 & 1.95 & 0.99 & 0.46 \\ $M03$ & 24.6 & 56.4 & 99.0 & 0.54 & 1.29 & 0.99 & 0.53\\ $M04$ & 41.6 & 96.2 & 132.8 & 1.55 & 3.98 & 0.99 & 0.74\\ $M05$ & 22.2 & 48.6 & 100.3 & 0.83 & 1.15 & 1.00 & 0.93\\ $M06$ & 27.3 & 63.4 & 96.9 & 0.81 & 1.41 & 0.99 & 0.87\\ $M07$ & 24.0 & 55.3 & 105.8 & 1.02 & 1.45 & 0.99 & 0.51\\ $M08$ & 33.8 & 80.5 & 92.5 & 0.37 & 1.62 & 0.98 & 0.97\\ $M09$ & 28.7 & 66.1 & 103.5 & 1.41 & 1.66 & 0.99 & 0.98\\ $M10$ & 33.3 & 74.9 & 110.8 & 1.66 & 2.19 & 0.99 & 0.70\\ $M11$ & 32.4 & 76.6 & 178.2 & 4.47 & 5.62 & 1.00 & 0.14\\ $M12$ & 32.1 & 74.9 & 147.0 & 2.39 & 3.72 & 1.01 &0.18\\ $MM$ & 68.2 & 163.1 & 216.7 & 10.72 & 17.78 & 1.02 & 0.07 \\ \hline \end{tabular} \end{minipage} \end{table} \begin{table} \centering \begin{minipage}{140mm} \caption{Mean parameters using radial range $I_1$}\label{tab:fits1} \begin{tabular}{lcrccc} \hline {\sc id}& $n$ & $R_{\rm e}$ & $-\mu_0$ & $-M_{T}$ & {\sc rms} \\ & & [kpc] & [${\rm M}_\odot/{\rm kpc}^2$] & [M$_\odot$] & \\ \hline $M01$ & $ 4.3 \pm 0.4$ & $ 9.2 \pm 1.9$ & $28.5 \pm 0.5$ & $27.6$ & $0.10$ \\ $M02$ & $ 2.1 \pm 0.1$ & $ 2.7 \pm 0.2$ & $23.3 \pm 0.3$ & $24.4$ & $0.10$ \\ $M03$ & $ 1.9 \pm 0.1$ & $ 2.5 \pm 0.0$ & $23.0 \pm 0.1$ & $24.3$ & $0.12$ \\ $M04$ & $ 2.8 \pm 0.1$ & $ 3.9 \pm 0.2$ & $25.1 \pm 0.2$ & $25.5$ & $0.14$ \\ $M05$ & $ 3.9 \pm 0.2$ & $ 1.7 \pm 0.1$ & $28.4 \pm 0.4$ & $24.8$ & $0.10$ \\ $M06$ & $ 2.5 \pm 0.1$ & $ 4.2 \pm 0.3$ & $23.6 \pm 0.4$ & $24.7$ & $0.12$ \\ $M07$ & $ 3.1 \pm 0.2$ & $ 2.6 \pm 0.3$ & $26.1 \pm 0.5$ & $25.0$ & $0.10$ \\ $M08$ & $ 2.6 \pm 0.1$ & $ 2.1 \pm 0.2$ & $24.5 \pm 0.3$ & $23.9$ & $0.12$ \\ $M09$ & $ 2.7 \pm 0.2$ & $ 8.1 \pm 0.8$ & $23.2 \pm 0.6$ & $25.3$ & $0.20$ \\ $M10$ & $ 3.2 \pm 0.2$ & $ 6.4 \pm 0.7$ & $24.8 \pm 0.6$ & $25.5$ & $0.14$ \\ $M11$ & $ 1.6 \pm 0.1$ & $ 9.3 \pm 1.4$ & $22.0 \pm 0.5$ & $26.6$ & $0.08$ \\ $M12$ & $ 3.2 \pm 0.2$ & $ 4.3 \pm 0.6$ & $26.0 \pm 0.6$ & $25.9$ & $0.19$ \\ $MM$ & $ 2.4 \pm 0.1$ & $ 9.1 \pm 0.7$ & $24.5 \pm 0.1$ & $27.6$ & $0.19$ \\ \hline \end{tabular} \end{minipage} \end{table} \begin{table} \centering \begin{minipage}{140mm} \caption{Mean parameters using random radial range $I_2$}\label{tab:fits2} \begin{tabular}{lcrccc} \hline {\sc id}& $n$ & $R_{\rm e}$ & $-\mu_0$ & $-M_{T}$ & {\sc rms} \\ & & [kpc] & [${\rm M}_\odot/{\rm kpc}^2$] & [M$_\odot$] & \\ \hline $M01$ & $ 5.8 \pm 1.3$ & $13.2 \pm 5.5$ & $30.9 \pm 2.0$ & $27.8$ & $0.05$ \\ $M02$ & $ 2.1 \pm 0.3$ & $ 2.4 \pm 0.3$ & $23.7 \pm 0.8$ & $24.4$ & $0.05$ \\ $M03$ & $ 2.2 \pm 0.4$ & $ 2.3 \pm 0.2$ & $23.9 \pm 0.8$ & $24.3$ & $0.05$ \\ $M04$ & $ 3.3 \pm 1.0$ & $ 3.1 \pm 0.5$ & $26.5 \pm 2.1$ & $25.4$ & $0.07$ \\ $M05$ & $ 3.2 \pm 1.1$ & $ 2.1 \pm 0.6$ & $26.6 \pm 2.1$ & $24.8$ & $0.04$ \\ $M06$ & $ 2.7 \pm 0.4$ & $ 3.7 \pm 0.3$ & $24.3 \pm 0.9$ & $24.7$ & $0.06$ \\ $M07$ & $ 2.6 \pm 0.7$ & $ 3.0 \pm 1.0$ & $24.8 \pm 1.4$ & $25.0$ & $0.06$ \\ $M08$ & $ 2.1 \pm 0.3$ & $ 2.3 \pm 0.5$ & $23.3 \pm 0.8$ & $23.9$ & $0.04$ \\ $M09$ & $ 2.9 \pm 0.5$ & $ 6.9 \pm 0.9$ & $23.8 \pm 1.2$ & $25.3$ & $0.10$ \\ $M10$ & $ 3.5 \pm 0.6$ & $ 5.8 \pm 0.7$ & $25.7 \pm 1.4$ & $25.5$ & $0.06$ \\ $M11$ & $ 1.6 \pm 0.3$ & $ 9.6 \pm 1.9$ & $22.0 \pm 0.8$ & $26.6$ & $0.06$ \\ $M12$ & $ 3.2 \pm 1.4$ & $ 6.3 \pm 4.0$ & $25.5 \pm 2.4$ & $26.0$ & $0.16$ \\ $MM$ & $ 3.2 \pm 0.6$ & $ 9.2 \pm 1.5$ & $26.3 \pm 1.0$ & $27.6$ & $0.07$ \\ \hline \end{tabular} \end{minipage} \end{table} Tables~\ref{tab:fits1} and~\ref{tab:fits2} summarise the mean values of the fitted S\'ersic parameters $\{n,R_{\rm e},\mu_0\}$ ($\mu_0\!=\!-2.5\log \Sigma_0$), the total ``magnitude'' ($M_T\!\equiv\! -2.5\log M_{\rm L}$) and the {\sc rms} of the fit, for the different projections for both radial intervals $I_1$ and $I_2$; respectively. Here $M_{\rm L}$ is determined from the fitted values using equation (\ref{eq:Lmass}). Standard deviations are listed for the S\'ersic parameters. The values of $M_{\rm L}$ determined from the fits agree very well with $M_{\rm lum}$. \subsection{Luminous Distributions} \subsubsection{Shape parameter}\label{ssec:n} Figure~\ref{fig:Nglxs}~({\it top}) shows the frequency distribution of $n$ for a set of observational data in optical wave bands (D01, La~Barbera~et~al.~2005) and in the near-infrared ($K$) band [La~Barbera~et~al~2005, Ravikumar~et~al.~2005 (R05)]. A total of 169 galaxies in the optical and 156 in the $K$ band were used here. The frequency distribution of 41 merger remnants observed in the $K$-band by Rothberg \& Joseph (2004) are also indicated as a shaded histogram. The mean and standard deviations of these data sets are indicated, as well as their median. In Figure~\ref{fig:Nglxs}~({\it bottom}) we show the distribution of $n$ for our merger remnants using the radial fitting intervals $I_1$ and $I_2$. The frequency distribution for our $N$-body remnants peak at a value $n\approx 3$ in both cases; although using $I_2$ it shows a somewhat broader distribution. For $I_1$ it is found that $n\in (1.5,5.3)$ and for $I_2$ that $n\! \in \! (1.4,9.5)$. These values are in good agreement with those found in intermediate mass ellipticals (e.g. Graham~\&~Guzm\'an~2003, de Jong~et~al.~2004, Trujillo, Burkert \& Bell~2004, Ellis~et~al.~2005), some brightest cluster galaxies (e.g. Graham~et~al.~1996), dwarf ellipticals (e.g. Binggeli \& Jerjen~1998, Young \& Currie 2001), and the local merger remnants of RJ04. Our results using $I_1$, and \emph{no} bulge, are consistent with the values found by NT and GGB for their models with a bulge in the progenitors. Furthermore, using interval $I_2$ lead to some values $n \! \approx \! 9$. This does not seem to be due to the methodology in the computation of the surface density profiles. NT construct artificial images analogous to the observational procedure while GGB fit ellipses to isodensity contours, both considering a wide range in the radial fitting range, and obtaining similar ranges for $n$. It is likely that differences in the way models of the progenitors are set up be probably one of the reasons behind the differences with our results; see $\S$\ref{sec:discussion}. \subsubsection{Effective Radius} Figure~\ref{fig:Nre} ({\it top}) shows the observed frequency distribution of effective radii $R_{\rm e}$ for the data considered in $\S$\ref{ssec:n} and that corresponding to our remnants ({\it bottom}). For the fitting radial range $I_1$ we obtain $R_{\rm e}\! \in \!(1.6,12.9)\,$kpc, and for $I_2$ we have $R_{\rm e}\! \in \! (1.6,34.5)\,$kpc. The average value of the observational data is about $4\,$kpc and for our remnants is about $5\,$kpc. It can be noticed that $R_{\rm e}$ shows a larger dispersion of values than the index $n$ depending on the fitting interval. This was also noticed by Binggeli \& Jerjen (1998). Our remnants have a lower bound of $R_{\rm e}\!\approx \!1.5\,$kpc, while the observational data considered here can reach smaller values $R_{\rm e} \! \gta \! 0.5\,$kpc. We are not able to reproduce the small values of $R_{\rm e}$ mainly because in our sample of initial conditions no pairs of small progenitors were included. On other hand, values of $R_{\rm e}\gta 10\,$kpc can be reproduced by our more massive remnants ($M01$ and $MM$); see Tables~\ref{tab:fits1} and~\ref{tab:fits2}. A unique comparison with the distribution of$R_{\rm e}$ values found by NT and GGB is not possible, since their models can be scaled to arbitrary physical units; a thing that is not possible here due to the way our disc galaxy progenitors were built up. Nonetheless, if we use the range of dimensionless values found by NT ($1\! < \! R_{\rm e} \! < \! 1.7$) for systems classified as pure ``bulges'', and use a length unit of $3.5\,$kpc (i.e., the radial scale-length of the Milky Way) to transform their results to physical units, we find that both results are consistent. Also, we obtain qualitatively the same behaviour as the one shown in their Figure~18 where a sharp cut at the lower-end of the distribution, as well as an extended tail at larger values. Considering the observational values of $R_{\rm e}$ and those in our $N$-body remnants, we can establish with confidence that the simulations can reproduce quite well the observed range of values. Even some large values of $R_{\rm e}$ found in giant ellipticals (e.g., Graham~et~al.~1996) are reproduced. \subsection{Luminous Correlations} Several works (e.g. CCD93, D01, R05) have found a series of correlations among S\'ersic parameters in early-type galaxies. We now turn to study some of these and compare them with the properties of our numerical remnants. Firstly, we consider two-dimensional correlations, and then turn to consider the so called Photometric Plane (PHP) [e.g. K00]. \subsubsection{Two Dimensional Correlations}\label{sec:2Dcorrel} In the work of Caon~et~al.~(1993) it was stated that a linear positive correlation between $n$ and $R_{\rm e}$ exists for early-type galaxies; they find that $n\propto R_{\rm e}^{0.52}$ for early-type galaxies in Virgo. A similar conclusion was reached by D'Onofrio, Capaccioli \& Caon (1994) analysing galaxies in Fornax. Combining the data of both works one finds $n\propto R_{\rm e}^{0.50}$ with a Pearson's linear correlation coefficient $r=0.72$. The statement of CCD93 that structure (as indicated by $n$) of an elliptical depends on its size $R_{\rm e}$ has been supported by the analysis of Trujillo, Graham \& Caon (2001). Figure~\ref{fig:NRcorrel} shows index $n$ against $R_{\rm e}$ for some of the data considered here, as well as the values obtained for our disc galaxy merger remnants. A linear least-square fit to the data of D01 leads to $n\propto R_{\rm e}^{0.37}$, and for RJ04 mergers $n\propto R_{\rm e}^{0.26}$; with linear correlation coefficients $r=0.73$ and $0.39$, respectively. A similar fit to our remnants yields $n\propto R_{\rm e}^{0.22}$ with $r=0.39$. However, the observational data plotted in Figure~\ref{fig:NRcorrel} shows a large scatter around the assumed linear correlation; a fact already noticed by other authors (e.g. Trujillo~et~al.~2001). These fluctuations are quantified by considering the {\sl coefficient of determination} ($r^2$) that measures the proportion of the variance of one variable that is predictable from the other (e.g. Ryan~1997). As indicated in Figure~\ref{fig:NRcorrel}, the coefficients of determination are rather small, and the {\sc rms} of the fits are large, so its is not clear that a {\it true} linear correlation exists between $\log n$ and $\log R_{\rm e}$. We notice also that the $N$-body remnant $M05$, being the smallest one, has the second largest $n$ value in our simulations. These results lead us to state that there is {\it no} linear positive correlation between the ``structure'' and size of an elliptical. It seems that the values of $n$ and $R_{\rm e}$ are restricted by some physical mechanism to a finite region of the S\'ersic parameter space; an option also indicated by Trujillo~et~al.~(2001). On other hand, a stronger observational correlation has been found between $n$ and the central brightness $\mu_0$ in ellipticals (e.g. K00, Graham \& Guzm\'an~2003) and it appears to extend to dwarf ellipticals (e.g. Binggeli \& Jerjen 1998, R05). In Figure~\ref{fig:NMu0correl} we plot these quantities for the observational data of D01, R05, RJ04, and for comparison our $N$-body remnants. A linear fit $\log n$--$\mu_0$ to these data leads to $n\propto I_0^{0.17}$ for D01, $n\propto I_0^{0.14}$ for both RJ04 and the ellipticals in R05, and $n\propto I_0^{0.14}$ for our merger remnants. All fits have $r^2\gta 0.9$ and an {\sc rms}~$\lta 0.07$, that lead us to conclude that $\log n$--$\mu_0$ is a true linear correlation, at least for the range of $n$ values considered. As shown, the numerical remnants presented here are able to reproduce very well the $\log n$--$\mu_0$ correlation and its tightness. We recall that we have assumed a constant mass-to-light ratio to convert $\Sigma_0$ to $I_0$ in order to compare with observations. Hence, it appears that at least in the range of masses of our remnants (Table~\ref{tab:global}), there is no need to assume a dependence of the mass-to-light ratio dependence with mass (or luminosity) to reproduce the observations. \subsubsection{The Photometric ``Plane''} Several authors [e.g. K00, Graham~2002, Khos\-ro\-sha\-hi~et~al.~2004 (K04), La~Barbera~et~al.~2005, R05] have found that S\'ersic parameters $\{n, R_{\rm e}, \mu_0\}$ of early-type galaxies define a plane in log-space of the form \begin{equation} \log n = a \log R_{\rm e} + b \, \mu_0 + c\,, \label{eq:php} \end{equation} which is termed the ``photometric plane'' (PHP). Some authors instead of $\mu_0$ use the mean effective brightness $\langle \mu \rangle_{\rm e}$ (e.g. Graham~2002, La~Barbera~et~al.~2005). The different 2D correlations of $\S$\ref{sec:2Dcorrel} can be considered then projections of the PHP. We have computed, by a linear-square fit procedure, the coefficients of this PHP for our remnants under the assumption of a constant mass-to-light ratio. Since we find that using $\langle \mu \rangle_{\rm e}$ leads to about twice the {\sc rms} in $\log n$ than using $\mu_0$, we restrict ourselves to an expression of the form (\ref{eq:php}). \begin{table} \centering \begin{minipage}{140mm} \caption{Photometric Plane coefficients}\label{tab:php} \begin{tabular}{lccrc} \hline {\sc id}& $a$ & $-b$ & $c$ & {\sc rms}$_n$ \\ \hline K00 & $0.17 \pm 0.03$ & $0.069\pm 0.007$ & $1.2\pm 0.1$ & 0.04 \\ D01 & $0.09 \pm 0.02$ & $0.058\pm 0.004$ & $1.3\pm 0.1$ & 0.06 \\ K04 & $0.21 \pm 0.09$ & $0.074\pm 0.013$ & $1.7\pm 0.3$ & 0.13 \\ R05E & $0.15 \pm 0.02$ & $0.066\pm 0.003$ & $1.1\pm 0.0$ & 0.04 \\ R05dE & $0.16 \pm 0.04$ & $0.082\pm 0.004$ & $1.6\pm 0.1$ & 0.07 \\ RJ04 & $0.11 \pm 0.04$ & $0.054\pm 0.004$ & $1.0\pm 0.0$ & 0.06 \\ M($I_1$) & $0.05 \pm 0.00$ & $0.057\pm 0.001$ & $-1.0\pm 0.0$ & 0.05 \\ M($I_2$) & $0.05 \pm 0.01$ & $0.057\pm 0.001$ & $-1.0\pm 0.0$ & 0.05 \\ \hline \end{tabular} \end{minipage} \end{table} In Table~\ref{tab:php} we list the values of the coefficients of (\ref{eq:php}) found in the works of K00, K04, RJ04, the elliptical and dwarf ellipticals of R05, and those we obtain from the data of D01. Also shown are the coefficients found for our merger remnants for both fitting radial intervals $I_1$ and $I_2$; M($I_1$) and M($I_2$), respectively. In Figure~\ref{fig:PHPrem} we plot the PHP from these data, using for illustrative purposes the values from the data of D01 to define the abscissa axis. The remnants' coefficient $b$ in (\ref{eq:php}), associated with $\mu_0$, is rather consistent with the observed ones, aside of those found in dEs (R05). The coefficient $a$, associated with $R_{\rm e}$, is less well reproduced. This is not surprising taking into account the large dispersion in the $n$--$R_{\rm e}$ relation (see Figure~\ref{fig:NRcorrel}). As several authors have pointed out (e.g. K00, K04, R05) a slight curvature towards small values of $n$ is observed, a feature that tends to be reproduced here by the effect of merger $M11$ that has $n\! \approx \! 1.5$. The {\sc rms} is similar for both the observational data and our simulations. The best overall agreement is obtained with the data of D01. We consider that our $a$ and $b$ values are rather consistent with the whole set of values listed in Table~\ref{tab:php}, and that the numerical remnants are able to reproduce the PHP. In $\S$\ref{sec:discussion} we argue that the PHP is not really a plane, but a ``pseudo-plane'' with a small curvature at low values of $n$ due to the intrinsic properties of S\'ersic model. \subsection{Dark Haloes}\label{sec:dark} Dark haloes in cosmological simulations are started to being described by a 3D S\'ersic function (Merritt~et~al.~2005, Prada~et~al.~2005, Graham~et~al.~2005), of the form \begin{equation} \rho(r) = \rho_0 \exp[ -d_n (r/r_{\rm e})^{1/n} ] \label{eq:3dsersic} \end{equation} with $r$ being the spatial radius, $r_{\rm e}$ a 3D ``effective radius'', and $d_n \approx 3n-1/3+0.005/n^2$ (Graham~et~al.~2005). It has been found that (\ref{eq:3dsersic}) provides a better fit to dark haloes than the typical NFW or M99 (Moore~et~al.~1999) density profiles. S\'ersic indices of about $6$, with a scatter of $\approx 1$, are found for the cosmological dark haloes. \begin{table} \centering \begin{minipage}{140mm} \caption{Dark haloes 3D S\'ersic fits}\label{tab:darkies} \begin{tabular}{lccrcc} \hline {\sc id}& $n$ & $r_{\rm e}$ & $\log \rho_0$ & & {\sc rms} \\ & & [kpc] & [M$_\odot/{\rm kpc}^3$] & & \\ \hline $M01$ & 3.3 & 86.7 & 9.08 & & 0.05 \\ $M02$ & 3.5 & 28.4 & 9.81 & & 0.03\\ $M03$ & 4.5 & 26.5 & 10.90 & & 0.04\\ $M04$ & 4.0 & 42.1 & 10.17 & & 0.08\\ $M05$ & 4.3 & 20.4 & 10.93 & & 0.06\\ $M06$ & 4.4 & 25.2 & 10.88 & & 0.05\\ $M07$ & 3.9 & 23.0 & 10.32 & & 0.05\\ $M08$ & 3.8 & 28.9 & 9.99 & & 0.05\\ $M09$ & 4.5 & 29.0 & 10.92 & & 0.07\\ $M10$ & 4.0 & 31.9 & 10.24 & & 0.12\\ $M11$ & 3.7 & 41.3 & 10.03 & & 0.06\\ $M12$ & 3.3 & 39.5 & 9.38 & & 0.05\\ $MM$ & 3.0 & 82.6 & 8.78 & & 0.04\\ \hline \end{tabular} \end{minipage} \end{table} We have fitted 3D S\'ersic profiles (\ref{eq:3dsersic}) to the dark haloes of our remnants. The radial range of the fit was from the convergence radius $r_{\rm c}$ (e.g. Power~et~al.~2003) to the dynamical virial radius of the remnant (Table~\ref{tab:global}). However, instead of using the orbital period at the $r_{200}$ radius to determine $r_{\rm c}$ as done in cosmological simulations, we used the orbital period at the virial radius. In Figure~\ref{fig:darkprofiles} we display the fitted 3D S\'ersic profiles, along with their residuals. In Table~\ref{tab:darkies} the values of the 3D S\'ersic parameters are listed for each remnant, as well as the corresponding {\sc rms}. It can be seen that the 3D S\'ersic profile (\ref{eq:3dsersic}) is a very good representation of the density distribution up to the virial radius of each remnant. This is in concordance with the behaviour of the 3D S\'ersic profile for characterising cosmological haloes. The haloes of the $N$-body remnants have a mean 3D S\'ersic index $\langle n \rangle\! = \! 3.9 \pm 0.5$; the uncertainty being the standard deviation. A value that is lower than that found for cosmological haloes. However, it is consistent with the mean value of the dark haloes of the progenitors $\langle n \rangle\! = \! 3.7\pm 0.3$; as expected from the pre\-ser\-vation of the cuspyness of dark haloes in mergers (e.g. Boylan-Kolchin \& Ma~2004, Aceves~\&~Vel\'azquez~2006, Kazantzidis, Zentner \& Kravtsov~2006). Differences in results are expected since the outer radius of the fits are not the same; we use a dynamical virial radius while for cosmological haloes the fits are done up to $r_{200}$ or further out (e.g. Prada~et~al.~2005). It should be noticed that S\'ersic fits (\ref{eq:3dsersic}) by Graham~et~al.~(2005) have values for $r_{\rm e}$ (see their Table~1) in some cases larger than their cosmological virial radius; these last ones listed in Table~1 of Diemand, Moore \& Stadel~(2004). For example, their haloes G02 and G03 have $r_{\rm e}\!=\! 391.4$~kpc and 405.6~kpc, respectively, while their virial radii are $337\,$kpc and $299\,$kpc; halo B09 shows even a more larger discrepancy. Unfortunately, they do not provide their numerical half-mass radii to make a direct comparison with the values $r_{\rm e}$ they obtained. Also, Merritt~et~al.~(2005) and Prada~et~al.~(2005) do not provide the fitted values $r_{\rm e}$ and the numerical half-mass radii. This makes uncertain any comparison of our results with these works. Figure~\ref{fig:darkies} shows different relations among the 3D S\'ersic parameters for the haloes of our remnants: $n$--$\rho_0$ and $\rho_0$--$r_{\rm e}$, and in analogy to the PHP we have constructed a 3D dark S\'ersic plane (DSP). We find, assuming a log-linear correlation, that $n\propto \rho_0^{0.08}$ and $\rho_0\propto r_{\rm e}^{-3.10}$ with coefficients of determination $0.96$ and $0.70$, respectively. This indicates that $n$--$\rho_0$ can be considered with confidence a true log-linear positive correlation, as its was in the corresponding 2D case, but we do not deem on the same level $\rho_0$--$r_{\rm e}$. The DSP found for the remnants is \begin{equation} \log n \! = \! (0.07 \pm 0.02) \log r_{\rm e} - (0.04 \pm 0.00) \bar{\mu}_0 - (0.48 \pm 0.07)\; \label{eq:dsp} \end{equation} where $\bar{\mu}_0\!=\!-2.5 \log \rho_0$. This turns out to be a very tight correlation, with a coefficient of determination of $0.98$ and {\sc rms} of $0.007$, for the range of haloes masses considered in our simulations (see Table~\ref{tab:global}). \section{Final Remarks and Conclusions}\label{sec:discussion} The results found here, as well as those of NT and GGB, show that the merger scenario is capable of reproducing the S\'ersic properties of observed elliptical galaxies. This work shows, however, that the presence of a ``primordial'' bulge in the progenitors is not necessary to satisfy, for example, the observed values of the shape parameter $n$; as was suggested by NT and GGB. It is likely that the different results with NT and GGB have their origin on the initial properties of the progenitors; in particular, their dark matter distribution. We have used cuspy (NFW-type) dark haloes in contrast to those used by NT (pseudo-isothermal) and GGB (Lowered Evans) that have a constant density core. It is probable that the higher concentration of dark matter used here, affected the distribution of luminous matter in a way to increase the index $n$ which is correlated with the luminous concentration (Trujillo~et~al.~2001). Other initial conditions of the progenitors and of the encounters such as energy and angular momentum, both intrinsic and orbital, may have played a role in the final concentration of luminous matter of the remnants, as indicated by the index $n$. A systematic study of the way different dynamical elements determine the S\'ersic index is, however, beyond the scope of the present work. We have shown that haloes of remnants define a tight dark S\'ersic plane (DSP) analogous to that of the luminous matter and with less dispersion. No indication of curvature is present, at difference to what is noted in the PHP at low values of $n$. We argue here that this curvature is real and is related to an intrinsic property of a S\'ersic profile. Consider the expression for the total luminous matter associated with a S\'ersic profile (\ref{eq:Lmass}). This can be written in log-space as \begin{equation} \log L_{\rm T} = \log n - 0.4\, \mu_0 + 2 \log R_{\rm e} + \log f_2(n) \label{eq:plane2} \end{equation} with the ``form factor'' $f_2(n)= 2 \pi \Gamma(2n)/b^{2n}$. A given set of galaxies with equal $L_{\rm T}$ and different S\'ersic parameters would define an exact log-plane, except for the presence of the $\log f_2(n)$ term. In the 3D S\'ersic function (\ref{eq:3dsersic}) the analogous form factor is $f_3(n) = 4 \pi \Gamma(3n)/d^{3n}$. These non-constant terms introduce a systematic change in a PHP-like expression. The importance of the form factor is larger for values $n\lta 1$ and smaller for $n \gta 1$; both $f_2(n)$ and $f_3(n)$ are shown in Figure~\ref{fig:Nfactor}. Thus, the form factor of the S\'ersic model determines the curvature observed in the PHP. This explains why no curvature is found by La~Barbera~et~al.~(2005) whose galaxies show $n \gta 2$, but this can be seen in dwarf ellipticals with several values of $n\lta 1$ (K04). Also, the DSP does not show such curvature since $n \gta 3$ (see Figure~\ref{fig:darkies}). Furthermore, the dispersion about these ``planes'' is determined by the luminosity or dark mass range of the galaxy sample. It remains to study the central phase-space densities of the remnants, to see if they are consistent with the estimates for ellipticals (e.g. Carlberg~1986), and to analyse their kinematical properties with those observed in elliptical galaxies. We plan to study these topics in a future work. In summary, our main conclusions are as follows: \begin{enumerate} \item Collision-less mergers of \emph{pure} disc galaxies yield values and distributions of S\'ersic parameters consistent with those observed for bona fide ellipticals. The existence of a bulge in merging spirals does not appear to be a necessary condition on grounds of S\'ersic properties of the remnants. \item The suggested positive log-linear correlation between the size ($R_{\rm e}$) and structure ($n$) in ellipticals is not supported. However, the strong $\log~n$--$\mu_0$ linear correlation found in observational studies is supported by our merger simulations. On other hand, the PHP is fairly well reproduced. For these results a constant mass-to-light ratio is assumed. \item The final dark haloes of remnants show values of $n\approx 4$ lower than those found in cosmological simulations $n\approx 6$. The difference may be attributable to the non-equivalence outer radius, where the dynamical virial radius was used in our case to carry out the fitting by a S\'ersic profile. Haloes define a tight Dark S\'ersic Plane (DSP) in three dimensions, with no indication of curvature at the level of the smaller $n$ obtained. \item The curvature observed in the PHP at low values of $n$ is an intrinsic manifestation of the properties of S\'ersic model, due to the presence of a non-constant term dependent of $n$. \end{enumerate} \section*{Acknowledgments} This research was funded by CONACyT-M\'exico project 37506-E. We appreciate the kindness of Chazhiyat Ravikumar for providing us observational data used in this work. \section*{Appendix} We briefly discuss here the effect of the radial interval on the fitting process of a S\'ersic profile to a mass distribution. To do this we generate exact S\'ersic profiles with $R_{\rm e}\!=\! 1$, $L_{\rm T}\!=\!1$, and $n=2,4,8$. Also, random fractional errors $\le \{1,10,20\}$\% are introduced. The radial fitting interval is chosen as follows. A random inner point $\xi$ is selected from the interval $[0.03,0.5]R_{\rm e}$ while the outer radius, $\eta$, is randomly generated from the interval $[R_{70},R_{95}]$; where $R_{70}$ and $R_{95}$ correspond to the radii containing 70\% and 95\% percent of the projected mass. These two points define our radial fitting interval $[\xi,\eta]$. To corroborate the importance the importance of the underlying mass distribution we use also a Hernquist (1990) mass model with an without errors. \subsubsection*{S\'ersic Distribution} Table~\ref{tab:app1} lists the mean values of $\{n,R_{\rm e},\mu_0\}$ obtained from fitting 1000 Monte Carlo experiments for three S\'ersic models with errors as indicated above. Each line lists, in order of the ascending error introduced to the theoretical S\'ersic profile, the parameters recovered from the fit inside the random interval $[\xi,\eta]$. The standard deviation for each quantity is provided. These results show that the determination of S\'ersic parameters is very robust, for errors $\lta 10$\%, against the size of the fitting region. As the error in the ideal S\'ersic distribution increases the dispersion grows. This is more clearly appreciated for index $n$. In the limit of zero error, even for a random radial fitting interval, the model parameters are recovered exactly. For this case, we conclude that the radial fitting range does not has an important effect on S\'ersic fitted parameters. \begin{table} \centering \begin{minipage}{8cm} % \caption{S\'ersic fits.}\label{tab:app1} \begin{tabular}{cccc} \hline $n_{\rm{true}}$ & $n$ & $R_{\rm e}$ & $-\mu_0$ \\ \hline 2 & $2.000\pm 0.015$ & $1.000 \pm 0.003$ & $\hphantom{0}0.956\pm \hphantom{0}0.030$ \\ & $2.014\pm 0.162$ & $1.002 \pm 0.031$ & $\hphantom{0}0.981 \pm \hphantom{0}0.319$ \\ & $2.061\pm 0.418$ & $1.009 \pm 0.082$ & $\hphantom{0}1.067 \pm \hphantom{0}0.819$ \\ \hline 4 & $4.001\pm 0.037$ & $1.000 \pm 0.004$ & $\hphantom{0}4.940 \pm \hphantom{0}0.080$ \\ & $4.041\pm 0.408$ & $1.001 \pm 0.043$ & $\hphantom{0}5.026 \pm \hphantom{0}0.876$ \\ & $4.222\pm 1.749$ & $1.010 \pm 0.139$ & $\hphantom{0}5.407 \pm \hphantom{0}3.719$ \\ \hline 8 & $8.002\pm 0.087$ & $0.999 \pm 0.007$ & $13.261 \pm \hphantom{0}0.196$ \\ & $8.119\pm 1.042$ & $0.998 \pm 0.071$ & $13.521 \pm \hphantom{0}2.316$ \\ & $8.663\pm 4.776$ & $1.006 \pm 0.251$ & $14.719 \pm 10.407$ \\ \hline \end{tabular} \end{minipage} \end{table} \subsubsection*{Hernquist Distribution} We now consider that the case where the underlying mass distribution follows a Hernquist model. Here, $R_{\rm hl}$ denotes its theoretical projected half-light (mass) radius. It is found that in the fitting interval $[0.03,2.79]R_{\rm hl}$ , the underlying Hernquist's profile is fitted by a S\'ersic profile with index $n=2.6$ and $R_{\rm e}=0.82$ in agreement with NT. For a radial fitting interval of $[0.03,14.5]R_{\rm hl}$ we find that $n\!=\!3.67$ and $R_{\rm e}\!=\!1.10$. This indicates that the process of fitting a S\'ersic profile is far more sensitive when the underlying mass distribution does not follows a S\'ersic one. The above was already noted by Boylan-Kolchin, Ma \& Quataert~(2005), where a systematic change in S\'ersic parameters was found when trying to fit a Hernquist profile. In Figure~\ref{fig:app1} ({\it left}) we reproduce this systematic effect, for the S\'ersic index $n$, $R_{\rm e}$, and the mean effective surface brightness $\langle I_{\rm e}\rangle$. If a random error $\le 10$\% is introduced the trend is preserved but a large dispersion results; especially as the inner radius of radial interval of the fit is increased. Figure~\ref{fig:app2} shows the distribution of fitted values with no error ({\it solid line}) and with a random error $\le 10$\% ({\it dashed line}) for an underlying Hernquist profile where the radial interval was obtained from $\xi \in [0.06,0.50]R_{\rm hl}$ and $\eta\in [\eta_{70},\eta_{95}]$. The mean and standard deviations of the distribution are indicated. For comparison, the histogram of values corresponding to a S\'ersic model with $n=4$ with a random error $\le 10$\% is also shown ({\it dotted line}); see Table~\ref{tab:app1}, second line in the entry for $n=4$. From the above results, it follows that when the underlying mass distribution is {\it not} of a S\'ersic type, the fitted values have a rather large dispersion even in the presence of no error. In particular, higher values of the index $n$ are obtained for different radial ranges of the fit. In order to have a confident estimate of $n$, and other parameters, one has to sample rather deep inside and outside the luminous (mass) distribution; from about $0.1$ to $6\,R_{\rm hl}$. In practise, for example, sampling very near the centre of a galaxy may pose problems due to resolution effects. This is a particular problem for observations of galaxies at different redshifts, using the same angular resolution but representing different physical scales, and can lead to uncertain S\'ersic parameters. \bsp \label{lastpage}
Title: The effect of environment on the UV colour-magnitude relation of early-type galaxies
Abstract: We use \textit{GALEX} (Galaxy Evolution Explorer) near-UV (NUV) photometry of a sample of early-type galaxies selected in \textit{SDSS} (Sloan Digital Sky Survey) to study the UV color-magnitude relation (CMR). $NUV-r$ color is an excellent tracer of even small amounts ($\sim 1$% mass fraction) of recent ($\la 1$ Gyr) star formation and so the $NUV-r$ CMR allows us to study the effect of environment on the recent star formation history. We analyze a volume-limited sample of 839 visually-inspected early-type galaxies in the redshift range $0.05 < z < 0.10$ brighter than $M_{r}$ of -21.5 with any possible emission-line or radio-selected AGN removed to avoid contamination. We find that contamination by AGN candidates and late-type interlopers highly bias any study of recent star formation in early-type galaxies and that, after removing those, our lower limit to the fraction of massive early-type galaxies showing signs of recent star formation is roughly $30 \pm 3%$ This suggests that residual star formation is common even amongst the present day early-type galaxy population. We find that the fraction of UV-bright early-type galaxies is 25% higher in low-density environments. However, the density effect is clear only in the lowest density bin. The blue galaxy fraction for the subsample of the brightest early-type galaxies however shows a very strong density dependence, in the sense that the blue galaxy fraction is lower in a higher density region.
https://export.arxiv.org/pdf/astro-ph/0601036
\title{The Effect of Environment on the UV Color-Magnitude Relation of Early-type Galaxies} \author{ K. Schawinski,\altaffilmark{1} S. Kaviraj,\altaffilmark{1} S. Khochfar,\altaffilmark{1} S.-J. Yoon,\altaffilmark{2,1} S. K. Yi,\altaffilmark{2,1,10} J.-M. Deharveng,\altaffilmark{3} A. Boselli,\altaffilmark{3} T. Barlow,\altaffilmark{4} T. Conrow,\altaffilmark{4} K. Forster,\altaffilmark{4} P. G. Friedman,\altaffilmark{4} D. C. Martin,\altaffilmark{4} P. Morrissey,\altaffilmark{4} S. Neff,\altaffilmark{5} D. Schiminovich,\altaffilmark{6} M. Seibert,\altaffilmark{4} T. Small,\altaffilmark{4} T.Wyder,\altaffilmark{4} L. Bianchi,\altaffilmark{7} J. Donas,\altaffilmark{3} T. Heckman,\altaffilmark{7} Y.-W. Lee,\altaffilmark{2} B. Madore,\altaffilmark{8} B. Milliard,\altaffilmark{3} R. M. Rich\altaffilmark{9}\& A. Szalay\altaffilmark{7} } \altaffiltext{1}{Department of Physics, University of Oxford, Oxford OX1 3RH, UK} \altaffiltext{2}{Center for Space Astrophysics, Yonsei University, Seoul 120-749, Korea} \altaffiltext{3}{Laboratoire d'Astrophysique de Marseille, 13376 Marseille Cedex 12, France} \altaffiltext{4}{California Institute of Technology, MC 405-47, Pasadena, CA 91125} \altaffiltext{5}{Laboratory for Astronomy and Solar Physics, NASA Goddard Space Flight Center, Greenbelt, MD 20771} \altaffiltext{6}{Department of Astronomy, Columbia University, MC 5246, New York, NY 10027} \altaffiltext{7}{Department of Physics and Astronomy, Johns Hopkins University, Baltimore, MD 21218} \altaffiltext{8}{IPAC,770 S. Wilson Ave., Pasadena, CA 91125} \altaffiltext{9}{Department of Physics and Astronomy, University of California, Los Angeles, CA 90095} \altaffiltext{10}{Send offprint request to yi@yonsei.ac.kr} \keywords{galaxies: elliptical and lenticular, cD -- galaxies: evolution -- galaxies: formation -- galaxies: fundamental parameters} \section{Introduction} There is observational evidence pointing to a very simple evolutionary model for early-type galaxies. This model of \textit{Monolithic Collapse} was first proposed by \citet*{1962ApJ...136..748E} to explain the origin of the Milky Way halo. According to this model, the Milky Way halo formed through the rapid collapse of a cloud of gas very early on in the history of the universe, forming all of its stars in an initial burst followed by a passive evolution of the stellar population. A similar model is often invoked as the simplest explanation for the old and seemingly homogeneous stellar populations seen in early-type galaxies \citep{1975MNRAS.173..671L}. The apparently universal relationship between galaxy color and luminosity in early-type galaxies was first studied in detail by \citet*{1977ApJ...216..214V}, even though the relation had been observed before \citep{1959PASP...71..106B, 1961ApJS....5..233D, 1968AJ.....73.1008M}. This \textit{Color-Magnitude Relation} (CMR) is often used as a tool for understanding the formation and evolution of early-type galaxies. A seminal investigation on the optical CMR was undertaken by \citet*{1992MNRAS.254..589B} on the elliptical galaxies in the Virgo and Coma clusters. Their study revealed a remarkably small intrinsic scatter around the mean relation. In the context of the monolithic paradigm, they interpreted the small scatter as the result of a small age dispersion amongst galaxies of the same age and the slope as a result of a mass-metallicity relation \citep{1997A&A...320...41K}. Further, they concluded that massive early-type galaxies did not have any major episodes of star formation at redshifts z $< 2$. More massive galaxies are likely to be in deeper potential wells and are therefore more able to retain metals ejected from supernovae from the initial generations of young stars at high redshift, leading to the observed mass-metallicity relation \citep{1974MNRAS.169..229L}. In addition to this, the observed levels of $\alpha$-enhancement \citep{1992ApJ...398...69W,1993MNRAS.265..553C, 1994MNRAS.270..743C, 1994MNRAS.270..523C, 1997A&A...320...41K, 2000AJ....120..165T} in many giant ellipticals imply that the initial formation starburst was of a very short duration of less than 1 Gyr \citep{1993ApJ...405..538B, 1999MNRAS.302..537T}. Later studies have found that there is no significant evolution in the optical CMR out to z=1 and further \citep{1997ApJ...483..582E, 1998ApJ...501..571G,1998ApJ...492..461S, 2000ApJ...541...95V, 2003ApJ...596L.143B, 2005ApJ...635..243F}. All this adds up to a picture of massive early-type galaxies forming in an initial, intense starburst at high redshift followed by a relatively-passive evolution. However, we know since the simulations of \citet*{1972ApJ...178..623T} that the product of a spiral-spiral merger can be an elliptical galaxy \citep{1983MNRAS.205.1009N, 1988ApJ...331..699B, 1992ApJ...400..460H, 2003ApJ...597..893N}. An alternative approach to understanding early-type galaxies takes into account dynamical interactions and mergers. In the \textit{Hierarchical Merger} paradigm, small galaxies form first and later assemble into larger objects \citep{1978MNRAS.183..341W}. The advent of semi-analytical models (SAMs) in the 1990s has greatly enhanced our understanding of galaxy evolution in such a hierarchical universe. \citet{1996MNRAS.283L.117K} find that in the Canada-France Redshift Survey, only 1/3 of elliptical and lenticular galaxies at redshift z=1 were fully assembled and showed colors expected of old passively evolving systems. There is of course older evidence for a strong dependence of the population of early-type galaxies on density and redshift. \citet{1980ApJ...236..351D} found that approximately 80\% of galaxies in a sample of 55 clusters were of early-type morphology, a much higher fraction than in the field suggesting that the denser cluster environment does affect galaxy evolution. When \citet{1984ApJ...285..426B} looked at higher redshift clusters, they found that the fraction of blue, spiral galaxies in cluster environments increased with redshift. Later studies confirmed that this trend was not a selection effect \citep{1997ApJ...490..577D, 2000ApJ...541...95V}. This evolution is accompanied with an increase in merger rates \citep{1998ApJ...497..188C, 1999ApJ...520L..95V, 2001ApJ...561..517K}. In a purely monolithic collapse model, the star formation history of early-type galaxies is almost trivial, as they are composed of uniformly old stars. As soon as we allow for any sort of hierarchical merging, the story becomes much more complex. Rather than being uniform, the star formation histories become highly degenerate, as disparate stellar populations from progenitor galaxies are mixed together. Beyond this simple addition, the merging history of the galaxy and its progenitors adds further complication as entirely new populations are created during interactions and mergers. Thus, assigning a single age to the stellar population of an early-type galaxy is misleading - there is no single age. The typically-derived luminosity-weighted ages are in this sense nontrivial to interpret. We know now that the combined effects of age, dust, metallicity and - potentially - a multitude of progenitors are highly degenerate. \citet*{1992MNRAS.254..589B} took the apparent uniformity and low intrinsic scatter as a very strong constraint on the evolution of the Virgo \& Coma Early-type population. While monolithic evolution is the simplest possible explanation of these observations, however, it does not necessarily exclude other interpretations. \citet{2005MNRAS.360...60K} have argued using merger models that the optical early-type CMR is useful for constraining evolution models \textit{only} if we believe \textit{a priori} in a monolithic model. The effect of progenitor bias - the fact that a progressively larger fraction of the progenitor set of present-day ellipticals is contained in late-type star-forming galaxies at higher redshift - means that we are \textit{not} probing the entire star formation history of early-type galaxies, but rather a progressively more biased subset. Besides, the level of star formation predicted by SAMs incorporating AGN and supernova feedback is very low; on the order of a few percent by stellar mass. Optical filters, including $U$-band, are not sufficiently sensitive to detect such a low-level star-forming activity. This is why we must turn to the UV. The \textit{Galaxy Evolution Explorer (GALEX)} \citep{2005ApJ...619L...1M} near-UV filter is capable of detecting even a small ($\sim 1\%$ mass fraction) young stellar population and so ideal for tracing the recent star formation history of early-type galaxies. The UV color-magnitude relation allows us to identify the last important episode of star formation in galaxies. \citet{2005ApJ...619L.111Y} have already shown using \textit{GALEX} information that a significant fraction of massive early-type galaxies at low redshift exhibit levels of star formation undetectable in the optical but visible in the UV. Our paper presents the results of our search for the effect of environment on the recent star formation. We assume a standard $\Lambda$CDM cosmology with $(\Omega_{M}, \Omega_{\Lambda}) =(0.3, 0.7)$ and a Hubble constant of $H_{0}=70\, {\rm km\,s^{-1}\,Mpc^{-1}}$ \citep{2003ApJS..148..175S}. \section{Sample Selection \label{sample_selection}} The \textit{GALEX} \textit{Medium Imaging Survey} (MIS) is a wide-area survery with limiting magnitudes of 22.6 AB in the far-UV ($FUV; 1344-1786 A $) and 23.0 AB in the near-UV filter ($NUV; 1771-2831 A $) (\cite{2005ApJ...619L...7M}) with substantial overlap with the \textit{Sloan Digital Sky Survey} DR3 (\cite{2002AJ....123..485S}, \cite{2000AJ....120.1579Y}, \cite{2005AJ....129.1755A}) . We define a sample of early-type galaxies within \textit{SDSS} and then cross-match it to \textit{GALEX} detections. We use the \textit{GALEX} Internal Release 1.1 MIS data. \subsection{Early-type Galaxy Selection in \textit{SDSS} \label{criteria}} A fundamental problem in the study of early-type galaxies is that there are no fixed criteria for their classification. In terms of the Hubble Sequence, everything equal to or earlier than a lenticular is an early-type galaxy, but even this innocent definition is highly subjective, varies between different observers, and strongly depends on the image quality used to evaluate it. There is danger in classifying galaxies as early-types using the properties that are based on the presumption that early-type galaxies are old, red, dead, uniform and dustless, e.g, colors or spectral features. By making such supposition, any sub-population of early-type galaxies departing from this set of prejudices is liable to be rejected. Such a sample is then biased against \textit{precisely those} early-type galaxies that can tell us the most about galaxy evolution. In order to create an unbiased, volume-limited sample, we match \textit{GALEX} NUV detections to a catalog of early-type galaxies identified in the \textit{SDSS}. The paramount effort of \cite{2003AJ....125.1817B} (hereafter B03) has already generated such a catalog of $\sim 9000$ galaxies. They were selected on a number of \textit{SDSS} pipeline parameters. Such a catalog is of no doubt extremely useful to study the overall properties of galaxies in a statistical sense but less than perfect to our investigation which is searching for ``abnormality'' of early-type galaxies. For instance, B03 uses the Principal Component Analysis Technique which is biased strongly against star-forming ellipticals (e.g. \citet{2004ApJ...601L.127F}). Second, the sample generated this way is bound to be contaminated by late-type interlopers despite the effort of cleaning the sample in various ways (see B03 for details). In a visual inspection of a sample of bright ($M_{r} < -22$) early-types from the B03 catalog, we found up to 30\% contamination. These were not only Sa galaxies with small or faint spiral arms, but also edge-on disk and a number of face-on spirals. Such late-types are generally actively star-forming and should be removed from our sample. Besides, it is difficult to estimate the rate of false rejections (that is, early-types falsely rejected) if we use a catalog generated by a different group a priori. Some of these false rejections may be due to the \textit{spectral} part of the B03 criteria. \citet{2005ApJ...619L.107R} find the same contamination problem when they employed a method similar to B03. To avoid these problems, we define a simple set of \textit{morphology-driven} criteria with no assumptions at all on color or spectral energy distribution (SED). We define early-type galaxies to be those bulge-dominated galaxies that lack clearly visible spiral arms. We use these criteria to create an \textit{inclusive} rather than \textit{exclusive} sample to avoid rejecting too many genuine early-types. In order to select early-type galaxies over late-types, we consider the surface-brightness profiles in three bands and select those which have a very high likelihood of being a de Vaucouleurs profile rather than an exponential profile. To do this, we use the \textit{fracDev} parameter, which is the weight of the deVaucouleurs profile of the linear combination which best fits the image in each band. We select galaxies in DR3 with: \begin{enumerate} \item \textit{SED Quality}: The spectrum is of good quality ($S/N> 10$). \item \textit{fracDev\textunderscore g} $> 0.95$ We use the $g$ profile as it is sensitive to blue disk and arm stellar populations to ensure that Spiral galaxies are rejected. \item \textit{fracDev\textunderscore r} $> 0.95$ The $r$ band traces bulge populations and so will select bulges that follow an $r^{1/4}$ profile. \item \textit{fracDev\textunderscore i} $> 0.95$ The $i$ band strengthens the constraint derived from the $r$ band profile. \end{enumerate} For relatively bright galaxies ($r < 16.8$), this method is very reliable. The number of galaxies accepted that do not appear to be early-types upon visual inspection is on the order of $\sim 15\%$. Similarly, the number of galaxies that appear to be early-types amongst those which are rejected due to low values of $fracDev$ is $\sim 10\%$, which gives us confidence that we are not excluding a significant part of the early-type population. This level of contamination nevertheless requires a careful visual inspection, which is performed after the matching process. \subsection{Matching to \textit{GALEX}-MIS} The initial selection of early-type galaxies in \textit{SDSS} DR3 yields a total of 89248 galaxies without any constraints on luminosity or redshift. The detections in the \textit{GALEX} MIS survey are then cross-matched to this catalog. All early-type galaxies within each \textit{GALEX} field of view (FOV) are flagged and retained. We then perform a simple proximity search algorithm to find all those \textit{GALEX} detections that are within the $4 \arcsec$ angular resolution limit of \textit{GALEX} of each \textit{SDSS} early-type. All unique matches are flagged and kept together with all galaxies within \textit{GALEX} fields that are not detected. \subsection{Visual Inspection of Galaxy Morphology} The most dangerous contaminant when constructing a sample of supposed early-type galaxies are Sa galaxies. We set the difference between S0 (which we keep) and Sa (which we reject) to be \textit{the presence of distinct spiral arms}. This can be challenging when the galaxies in question are at higher redshift or faint. In order to quantify how well we can distinguish Sa galaxies based on SDSS images alone, we compare these to the \textit{COMBO-17} S11 field, which overlaps with \textit{SDSS} DR3 and has a number of galaxies at $ 0.10 \la z \la 0.13$. This image is significantly deeper (24 000 sec) and has better seeing ($\sim 0.7\arcsec$) than \textit{SDSS} images, so they allow us to identify morphology with much higher accuracy. We selected the brightest galaxies in the S11 field, ranging in $R$-band magnitude from 16.56 to 17.31. From this experiment, we conclude that the reliability of visual inspection is dependent first on redshift and seeing and second on apparent magnitude. In order to set a reliable redshift and magnitude limit, we limit our sample to $z < 0.1$ and $r < 16.8$. \subsection{The Volume-limited Sample} In order to create an unbiased sample, we need to take into account a number of factors. At $z < 0.05$, \textit{SDSS} spectroscopy begins to be incomplete for bright galaxies, so $z = 0.05$ is our lower limit \citep{2002AJ....123..485S, 2002AJ....124.1810S}. The \textit{GALEX} MIS limiting magnitude in NUV is 23.0 AB \citep{2005ApJ...619L...7M}, but many fields have longer exposure and some have been visited multiple times and co-added, giving us no uniform NUV magnitude limit. This is a problem since if we wished to probe the reddest early-types out to $NUV-r \sim 7.5$, we could only probe the most massive galaxies within a small redshift slice. In order to maximise the range in absolute magnitude to a reasonable part of the high end of the luminosity function, we must leave the reddest galaxies incomplete in some fields. We nevertheless retain them as non-detections. If we choose $r < 16.8$ as an apparent magnitude limit out to which visual inspection can be done reliably, the $NUV = 23.0$ hard limit guarantees us completeness to $NUV-r = 6.2$ which corresponds roughly to the top of the red sequence which was introduced by Yi et al. (2005, Figure 3). However, the fact that many images go up to a magnitude deeper than $NUV = 23.0$ means we can still probe the red end of the UV color-magnitude relation. Colors redder than $NUV-r \sim 6.5$ cannot be produced by an old stellar population of any age on its own; these galaxies must contain dust to achieve such red colors. Since we are primarily interested in studying those early-type galaxies that show signs of recent star formation, this is a safe limit. In addition to this, there is a significant fraction of \textit{SDSS} galaxies in the \textit{GALEX} field of view that are not matched to \textit{any} \textit{GALEX} counterpart (Figure \ref{fig1}). Even when matching to a sample of spectroscopic galaxies of \textit{all} morphologies in \textit{SDSS} roughly $10\%$ do not have \textit{GALEX} detections (Figure \ref{fig1}). Thus we must assume that they are either too faint in the UV, too dusty, or a combination of both and so can be assumed to be red on the UV color-magnitude diagram for the purpose of deriving the fractions of UV-bright galaxies. Some of these non-detections might also be due to mechanical problems in astrometry near the edge of the detector. Nevertheless, by making a number of assumptions on these non-detections, we can still derive some information from them. At $z = 0.1$, $r = 16.8$ is equivalent to an absolute magnitude limit of $M_{r} = -21.5$, so the limits on our sample are $z =[0.05, 0.10]$ and the color-magnitude relation can be probed out to $M_{r} = -21.5$. For comparison, $M_* = -20.83$ for all morphologies in an \textit{SDSS} sample \citep{2001AJ....121.2358B}. We then perform a visual inspection of all matched galaxies in our sample and place them into one of three categories: \begin{enumerate} \item Elliptical galaxies (847) \item Lenticular galaxy (112) \item Other (126) \end{enumerate} The ``Other'' category includes all galaxies rejected for either non early-type morphology or for the presence of nearby, bright blue stars which might contaminate the UV flux. The apparently low number of lenticular galaxies is due to the fact that we were very stringent about giving out the label ``lenticular''. If there was any doubt between elliptical and S0, we placed it in the elliptical category. In a study of 146 early-type galaxies of the Coma cluster, \citet{1994ApJ...433..553J} find that the separation of early-type galaxies into elliptical and lenticular is very difficult and that many face-on lenticulars have been misclassified as elliptical galaxies. In Section \ref{ba}, we discuss the relationship between recent star formation and axis ratio, where this effect becomes important. \subsection{Discussion of Random $\&$ Systematic Errors} The random errors in the $NUV-r$ color are dominated by the errors in the NUV. The mean 1-$\sigma$ error is 0.17 magnitudes, which is much smaller than the overall scatter of the observed colors (Figures \ref{fig3}, \ref{fig6}). The \textit{GALEX} photometry are taken from Internal Release 1.1, which is known to underestimate the errors. The errors are recalculated for our analysis following the instruction given iin the GALEX WEB site. Virtually all of our galaxies are unresolved in \textit{GALEX} NUV due to the large size of the NUV point spread function (4$\arcsec$ FHWM). Due to this large difference between the optical and UV resolutions, we do not attempt to use matched apertures. Since we use total fluxes, we do not expect color gradients to affect $NUV-r$ colors. \subsection{AGN Contamination $\&$ Removal} The other major problem is the presence of AGN. In the local universe, AGN hosts are preferentially massive elliptical galaxies. A strong AGN can easily produce a UV flux similar to that of a small mass fraction of young stars. In order to minimize the contamination from the galaxies whose UV fluxes are possibly dominated by an AGN non-thermal spectrum rather than a thermal stellar spectrum we apply two methods. First, we perform a BPT analysis \citep{1981PASP...93....5B} wherein galaxies are classified using a number of emission line ratios into either quiescent, star-forming or AGN. We employ a method similar to the one devised by Kauffmann et al. (2003). The line ratios used are [NII]/H$\alpha$ and [OIII]/H$\beta$. A full description of our method can be found in Kaviraj et al. (2005, in preparation). Classification using such a BPT diagram is only reliable when all four emission lines have sufficient $S/N$. The $S/N$ cut we employ in this study is $S/N > 3$ for all four lines. We reject all galaxies consequently classified as Seyfert, LINER or transition objects and only retain those which are quiescent or star-forming. It is interesting to note that most objects classified as star-forming were in fact galaxies rejected by the visual inspection as late-types (see Figure \ref{fig2}). Most of our early-type galaxies do not appear in \ref{fig2} because they do not show emission lines with $S/N>3$. For a discussion of how the AGN were identified, see \citet{2003MNRAS.346.1055K}. This process removed 11\% of our volume-limited sample after visual inspection. In order to ensure that we have as few AGN as possible left in our sample, we checked if any strong radio sources were left. The VLA FIRST survey \citep{1995ApJ...450..559B} covers about 80 \% of our galaxies at 1.4 GHz with $5\arcsec$ resolution. We removed all strong radio sources with a luminosity $L_{\rm 1.4 Ghz} > 10^{23} {\rm W\,Hz^{-1}}$. This cutoff was chosen as it is often assumed that below this luminosity, AGN activity and star formation are degenerate, whereas above it, most sources are AGN. We cross-checked this value to be consistent with the radio luminosities of our BPT-selected AGN. We only identify 8 further sources, which gives us confidence in the reliability of our BPT diagnostics. In total, this leaves us with a sample of 839 early-type galaxies to analyze. A catalog of these 839 galaxies can be made available upon request. We now construct the UV color-magnitude relation using this sample. In Figure \ref{fig3}, we show the entire sample of \textit{GALEX}-\textit{SDSS} matches with their classification into early-types, rejected late-types and AGN candidates. \section{Method \label{method}} In this section we describe our methods for classifying environment and for how we separate recent star formation (RSF) galaxies from ``UV-upturn'' galaxies. \subsection{Defining a Parameter for Local Environment \label{env_section}} We wish to define \textit{a quantitative way for measuring environment} that makes as much use of the information we are given as possible. Two-dimensional projected number densities would offer some information, but without redshift information, they can easily be rendered meaningless for anything but the most nearby clusters (e.g. Coma) due to fore- and background contaminants. It is possible to apply statistical methods to correct for this, but since \textit{SDSS} spectroscopy is available to us for all our galaxies and their surroundings, we can make use of spectroscopic redshifts to determine proximity. The high redshift accuracy of \textit{SDSS} spectroscopy ($\sigma_{z} = 1.7 \times 10^{-4} \pm 2 \times 10^{-5}$ for our sample , corresponding to $\sim 0.5 Mpc$ in our redshift range) allows us to compute the number density of neighboring galaxies \citep{2002AJ....124.1810S}. The SDSS spectroscopic completeness limit of $r = 17.77$ imposes a cut-off in absolute magnitude of $M_{r} = -20.55$ at our upper redshift limit of $z=0.1$. This allows us to sample the luminosity function to about $M_*$, which for an \textit{SDSS} sample is $M_* = -20.83$ \citep{2001AJ....121.2358B}. Any method that relies on number density has to deal with the fact that dense clusters give rise to peculiar velocities that can translate to shifts of up to several hundred ${\rm km\,s^{-1}}$, which can correspond to shifts of up to $\sim 10$ Mpc. \cite{2004ApJ...601L..29H} for example use a cylindrical volume elongated along the z-axis to $16h^{-1}$ Mpc to deal with this. Thus, their method corrects as well as is possible for dense environments. However, such a fixed- volume method does not take into account the density-dependence of peculiar velocity. We therefore attempt to correct for this by employing an adaptive volume: for each galaxy, we initially count all neighbors within a certain radius $\sigma$, ignoring the fact that the distance along redshift may be distorted. This number is capped at 10. We use this number $n$ as a guide to adaptively change the extent of our redshift search radius. We define the scale factor $c_{z}$ as follows: \begin{equation} c_{z} = 1+0.2n \label{cz} \end{equation} The scale factor is used to scale the value of $\sigma$ along the redshift axis by up to a factor 3 for the highest density environments to compensate for the ``finger of god'' effect. This is only a zeroth order approximation however and modelling will be needed to devise a more reliable method for scaling to a realistic volume. We then employ a Gaussian distribution to give more weight to closer neighbors and use $c_{z}$ to increase the extent of this Gaussian along the z-axis. We define the \textit{Adaptive Gaussian Environment Parameter} $\rho_{g}$ as the sum over all neighbors within the ellipse defined by \begin{equation} {(\frac{r_{a}}{3 \sigma})^2 + (\frac{r_{z}}{3 c_{z} \sigma})^2 \leq 1}, \end{equation} that is, we search out to $3 \sigma$: \begin{equation} \rho_{g}(\sigma) = \frac{1}{\sqrt{2 \pi}\sigma } \exp \left[ -\frac{1}{2} \left(\frac{r_{a}^2}{\sigma^2} + \frac{r_{z}^2}{c_{z}^2 \sigma^2} \right) \right] \label{rho_g1} \end{equation} where $r_{a}$ is the angular distance in Mpc to each surrounding galaxy, $r_{z}$ is the distance along the line-of-sight in Mpc to each surrounding galaxy, and $\sigma$ is an arbitrary dospersion parameter. This weighting scheme is biased towards nearby galaxies and so is a more realistic measure than a raw number density or overdensity. When measuring this parameter for a particular galaxy, the galaxy itself is \textit{not} counted towards the total. For this project, we adopt a fiducial value of $\sigma$ of 2.0 Mpc. The choice of $\sigma = 2$\,Mpc is somewhat arbitrary. We chose it so that the scale-length of our measure was focused approximately on the scale of large groups and small clusters. Perturbing $\sigma$ does not change our results within 1-$\sigma$. But what does this parameter actually measure? It is blind as to whether the structure around it is gravitationally bound or in equilibrium, so it is not a way of separating clusters from the field. Rather, it is a measure of the number and proximity of galaxies around a point in space, a more sophisticated number density. Despite that, it is useful to have physical sense on the values of $\rho_g$. First of all, the spatial distribution of our galaxies is mapped in Figure \ref{map1}\footnote{Created using POV-ray (http://www.povray.org)}. The birghter regions are denser. For comparison, we compute $\rho_g$ for the bright cluster galaxies in the C4 Catalog \citep{2005AJ....130..968M} within our redshift range. Indeed, most of them have high values of $\rho_g$ (Figure \ref{C4_density}). Typically, the central galaxy of a typical cluster with 10 $L_*$ galaxies randomly distributed within a $r \sim 3$\,Mpc sphere would have $\rho_g \sim 1$. A galaxy at the edge of the same cluster however would have $\rho_g \sim 0.5$. All galaxies in a group with three $L_*$ galaxies within a $r \sim 1$\,Mpc sphere would have similar values of $\rho_g$ to the cluster outskirts. Typical field galaxies would have $\rho_g \la 0.2$. We divide our final sample into three numerically equal environment bins, which we arbitrarily label ``low'', ``medium'' and ``high'' density (see Table \ref{environment_bins}). The (1) low, (2) medium, and (3) high density roughly correspond to (1) fields, (2) groups and cluster outskirts and (3) cluster centers, respectively. We also tested whether a mass weight could improve our measure. We tested a weight of the form: \begin{equation} \rho_{g}(\sigma) = \frac{f(mass)}{\sqrt{2 \pi}\sigma }\, \exp \left[ -\frac{1}{2} \left(\frac{r_{a}^2}{\sigma^2} + \frac{r_{z}^2}{c_{z}^2 \sigma^2} \right) \right] \label{rho_g2} \end{equation} where we chose $f$ to be a linear function of absoulte $r$-band magnitude such that a galaxy at the lower limit of $M_{r} = -20.55$ counted as 1 and the most massive neighbors of $M_{r} \sim -23$ counted three times as much. This made no difference within error to our result, so we do not use such a mass weight to avoid introducing unnecessary complication, so we adopt $f(mass) = 1$. \begin{table}[h] \caption{Environment Bins \label{environment_bins}} \begin{center} \begin{tabular}{c|ccc} \hline Bin & $\rho_{g}(\sigma = 2.0 Mpc)$ & Label\\ \hline \hline $0 - \frac{1}{3}$ & 0.00 $<$ $\rho_{g}$ $\le$ 0.21 & Low density\\ $\frac{1}{3} - \frac{2}{3}$ & 0.21 $<$ $\rho_{g}$ $\le$ 0.58 & Medium density\\ $\frac{2}{3} - 1$ & 0.58 $<$ $\rho_{g}$ $\le$ 4.68 & High density\\ \end{tabular} \end{center} \tablecomments{These bins are derived by splitting our sample of 839 galaxies into three euqal-number bins. The values of $\rho_{g}$ represend the boundaries between them.} \end{table} \subsection{Recent Star Formation and the UV-Upturn Phenomenon \label{upturn}} Many early-type galaxies exhibit the UV-upturn phenomenon (\cite{1979ApJ...228...95C}, \cite{1988ApJ...328..440B}) characterized by unusually strong UV flux rising with decreasing wavelength in the range ($1000-2500 A$). The UV-upturn phenomenon is thought to be due to the presence of low-mass, core helium-burning horizontal branch (HB) and evolved HB stars (\cite{1997ApJ...486..201Y}). We therefore face the problem that the moderate UV flux that we see in many of our early-type galaxies may in fact be due to such an old stellar population, or, even more difficult to resolve, due to both old and young stars. There is however a limit to how much NUV flux an early-type galaxy can produce via UV-upturn. This limit can be explored using both theoretical and observational methods. Ideally, we wish to combine both to derive a conservative limit beyond which we can be certain to probe recent star formation only. However, the UV upturn theory is still debated and thus observational evidence should take precedence. The IUE satellite conducted a survey of UV spectra of nearby elliptical galaxies \citep{1988ApJ...328..440B}. Among the strongest known nearby UV-upturn galaxies is NGC 4552, which has an $NUV-r$ color of 5.4 mag. We therefore choose $NUV-r = 5.4$ as a conservative lower boundary in color. At $NUV-r < 5.4$, all galaxies are considered to have experienced a recent episode of star formation, although part of their UV flux may come from a UV-upturn. Above this limit, a galaxy might either (or both) be forming stars or (and) exhibiting UV-upturn - we cannot distinguish the two using \textit{GALEX} NUV alone. Considering that the IUE SEDs were obtained from the UV-bright central regions of galaxies, our $NUV-r=5.4$ cut is conservative and puts some fraction of star-forming galaxies into the quiescent galaxy bins. \subsection{Comparison Between the Optical and UV-CMR \label{comparison}} In Figure \ref{fig6}, we plot the optical $u-r$ and $g-r$ color-magnitude relations on the same scale as the $NUV-r$. We label galaxies not classified as AGN by the BPT diagram above $NUV-r= 5.4$ as Quiescent (QST) and those bluer as Recent Star Formation (RSF). We do not include a slope in this cut-off, although one might suggest a slope based on the red-sequence slope for example as found by Yi et al. (2005, Figure 3), because any slope over our magnitude range would likely be very small and complex (albeit not impossible) to explain theoretically. We can see that the $g-r$ is completely insensitive. It cannot be used to detect recent star formation in early-type galaxies. Even $u-r$ color does not break this degeneracy. While the scatter of the UV-bright RSF galaxies is slightly greater, the bulk of them are indistinguishable from quiescent ones. In order to properly study recent star formation in early-type galaxies the UV information is essential. In total, $30\% \pm 3$ of our 839 early-type galaxies with $M_{r} < -21.5$ are classified as RSF using this scheme. This RSF galaxy fraction is probably a lower limit, first because of our conservative UV-upturn criterion and because we do not correct our UV data for internal extinction. \section{The Effect of Environment on Early-type Galaxies \label{env_dep}} In this Section, we investigate two related questions: does the UV color-magnitude relation depend on environment? And does the $fraction$ of early-type galaxies showing signs of recent star formation depend on environment? \subsection{The Color-Magnitude Relation \& Environment} It is well known that more massive early-type galaxies reside in denser environments \citep{1980ApJ...236..351D, 1984ApJ...281...95P} even though the slope and zero-point of their color-magnitude relations do not appear to depend on environment. \citet{2004ApJ...601L..29H} find that (1) the color-magnitude relation for their sample of 55,158 early-type galaxies in \textit{SDSS} does not depend on environment and that (2) the most luminous galaxies reside preferrentially in the most high-density environments. (2) is not surprising as the most massive ellipticals are known to reside at the centers of clusters \citep{1983ApJ...274..491B}. In their analysis, \citet{2003AJ....125.1882B} also find little dependence of the color-sigma relation on environment. Further, \citet{2005AJ....129...61B} suggest that the color-magnitude relation is entirely a consequence of the fact that both the luminosities and colors are correlated with sigma, a proxy to mass; that the color-sigma relation is in fact the more fundamental relation. In Figure \ref{color_cmr}, we show the UV color-magnitude relation for the three equal-number environment bins defined in Table \ref{environment_bins}. From this, we can see that there are two obvious differences between the low, medium and high density color-magnitude relations. As expected, the higher-density CMR extends to more massive galaxies. However, the low-density CMR extends to bluer colors than the high-density one. This is observational evidence for a change in the range of color of the UV-CMR with environment. We test the statistical significance of both of these environmental differences. \subsection{The Dependence of $NUV-r$ Color on Environment \label{color_dep}} The first quantity we consider is $NUV-r$ color. In Figure \ref{fig8}, we show how $NUV-r$ color varies with $\rho_{g}$. The range in $NUV-r$ remains more or less constant over the entire range of $\rho_{g}$, but the distribution itself varies with $\rho_{g}$. We can make this variation more apparent by plotting the cumulative color distribution of the three environment bins in Figure \ref{fig8}. In this plot, we not only show the cumulative distribution itself, but also a Monte Carlo re-simulation of the color distribution. In order to assess to what extent the difference between the environment bins is, we regenerate the distribution by randomly changing the color by the error and recomputing the distribution. The ``medium'' and ``high'' density curves are statistically indistinguishable. On the other hand, the low density bin (the yellow line in Figure \ref{fig8}) diverges from the other two at blue colors. We test the significance of this difference using both Kolmogorov-Smirnov (KS) and Kuiper test. The test significances are the probability that one of the distribution is drawn from a different parent distribution. The results are shown in Tables \ref{color_KS} and \ref{color_kuiper}. \begin{table}[!ht] \caption{KS-Test of NUV-r color dependence on environment \label{color_KS}} \begin{center} \begin{tabular}{c|ccc} \hline bin & low & medium & high \\ \hline \hline low & - & 89.220\% & 99.239\% \\ medium & 89.220\% & - & 25.130\% \\ high & 99.239\% & 25.130\% & - \\ \end{tabular} \end{center} \tablecomments{Table of Kolmogorov-Smirnov test significance comparing the distribution of $NUV-r$ color in the three environment bins.} \end{table} \begin{table}[!ht] \caption{Kuiper Test of NUV-r color dependence on environment \label{color_kuiper}} \begin{center} \begin{tabular}{c|ccc} \hline bin & low & medium & high \\ \hline \hline low & - & 83.428\% & 99.657\% \\ medium & 83.428\% & - & 55.753\% \\ high & 99.657\% & 55.753\% & - \\ \end{tabular} \end{center} \tablecomments{Table of Kuiper test significance comparing the distribution of $NUV-r$ color in the three environment bins.} \end{table} \subsection{The Dependence of Mass on Environment \label{mass_env}} \begin{table}[!ht] \caption{KS-Test of r-band absolute magnitude dependence on environment \label{mr_KS}} \begin{center} \begin{tabular}{c|ccc} \hline bin & low & medium & high \\ \hline \hline low & - & 99.139\% & 90.816\% \\ medium & 99.139\% & - & 50.728\% \\ high & 90.816\% & 50.728\% & - \\ \end{tabular} \end{center} \tablecomments{Table of Kolmogorov-Smirnov test significance comparing the distribution of $M_{r}$ in the three environment bins.} \end{table} Figure \ref{fig9} shows how $M_{r}$ varies with $\rho_{g}$. If we then plot the cumulative $M_{r}$ distribution for the three environment bins (Figure \ref{fig9}), we see a clear dependence of absolute magnitude on environment. Even in our volume-limited sample with a narrow baseline in luminosity there is a clear trend for brighter galaxies to be in higher-density environments. In Table \ref{mr_KS}, we give the KS-test significance for the differences between the $M_{r}$ distributions in each bin. \section{The Dependence of Recent Star Formation Activity on Environment} The fact that the distribution of $NUV-r$ color of massive early-type galaxies changes between low- and high-density environments may suggest that the recent star formation history of those galaxies is different. In order to quantify this, we use the criterion for recent star formation outlined in Section \ref{upturn}. It is not possible to directly convert NUV flux into an actual star formation rate, chiefly due to our inability to quantify dust extinction, to which the near-UV is extremely sensitive. We therefore merely classify our galaxies as RSF and QST and calculate the RSF galaxy fractions for subsamples in different environments in an attempt to find general trends. We calculate the recent star-forming fraction of early-types by dividing the number of galaxies bluer than $NUV-r = 5.4$ (RSF) by the total number of galaxies in this bin - that is, both those bluer and redder than $NUV-r = 5.4$ as well as those not detected by \textit{GALEX} but classified as early-type galaxies during the visual inspection. We include these non-detections as QSTs on the assumption that they are red galaxies further reddened by dust beyond the MIS detection limit. It is an intriguing possibility that at least some of these galaxies are dusty because they are actually forming stars, but we cannot make this distinction using \textit{GALEX}. In total, $30\% \pm 3$ of our 839 early-type galaxies with $M_{r} < -21.5$ are classified as RSF. The ellipticals, the bulk of our sample, have an RSF fraction of $29\% \pm 3$, while the lenticulars have an RSF fraction of $39\% \pm 5$. The division into ellipticals and lenticulars is based on visual inspection. We mentioned in \S 2.4 that our visual classification was generous to ellipticals. Hence, if a half of our ellipticals were in truth lenticulars, and if 39\% were the true RSF galaxy fraction for lenticulars, the RSF fraction for true ellipticals would be as low as 20\%. \subsection{RSF and Axis Ratio \label{ba}} The RSF galaxy fraction of those galaxies identified as lenticulars in the visual inspection is higher than that of the ellipticals. While there is no natural way to distinguish ellipticals and lenticulars \citep{1994ApJ...433..553J}, we can look at the change in UV properties with axis ratio. This still suffers from the fact that orientation can obscure true axis ratio. In Figure \ref{fig10}, we show the distribution of $NUV-r$ color with $r$-band axis ratio together with the RSF percentage as a function of $b/a$. Even amongst the roundest elliptical galaxies such as E0/1, there still is a significant fraction of star-forming galaxies. The RSF percentage appears to have a weak dependence on $b/a$ rising upto $\sim 50$\% for the most flattened galaxies (which corresponds to the 39\% we find for the visually identified lenticulars) but the trend is statistically insignificant. All this should be viewed in light of the bias against visually identifying face-on lenticular galaxies \citep{1994ApJ...433..553J}; it is likely that a fraction of the round early-types are such mis-classified objects and that the RSF fraction for genuine, round ellipticals is lower. \subsection{The Dependence of the RSF Galaxy Fraction on Environment} We now divide our sample into the three equal-number environment bins (see Table \ref{environment_bins}) to see whether the RSF fraction depends on environment. As expected from the results in Section \ref{color_dep}, the low-density environment bin shows a pronounced enhancement of the fraction of galaxies showing signs of recent star formation (see Figure \ref{fig11}). The medium- and high-density bins are consistent with having the same fraction. In order to constrain this enhancement further, we then divide our sample into 5 equal-number bins to see at what values of $\rho_{g}$ this increase lies and in particular whether there is any change at very low or high values. The red 5-bin curve in Figure \ref{fig11} shows that there is no change at high density and that the enhancement of the fraction of RSF galaxies begins at values of $\rho_{g} \sim 0.4$. This corresponds roughly to one $M_*$ galaxy per cubic Megaparsec, a loose definition of the ``field''. Thus, the enhancement of star formation in our sample is primarily due to the galaxies in the field. Our environment parameter on the other hand only probes neighbors down to $M_{r} \sim -20.5$, so these galaxies may well merely lack large neighbors - that they may simply be the dominant galaxy in a small group. Large surveys of galaxy star formation rates show a strong dependence on environment. Studies in both \textit{2dF} (\cite{2002MNRAS.334..673L}) and \textit{SDSS}. (\cite{2003ApJ...584..210G}) find that above a certain ``break'' local density, star formation rapidly declines. This ``break'' or ``characteristic'' local galactic density is given as $\sim 1 h^{2} {\rm Mpc}^{2}$, so the enhancement we see is similar. However, they also find a continuing decrease in star formation rate with increasing galactic density, which we do \textit{not} see. A direct comparison to our result is not possible however, since we cannot trace actual star formation rates, but rather only the fraction of galaxies showing signs of $recent$ star formation. \subsection{Breaking the Mass-Environment Degeneracy \label{mass_deg}} In Section \ref{mass_env}, we have shown the well-known fact that more massive galaxies prefer higher-density environments. It is also well-known that smaller galaxies tend to have higher star formation rates and are bluer, i.e. that the color-magnitude relation has a slope. This raises the possibility that the dependence of the RSF fraction for our entire volume-limited sample is nothing but an effect of mass. In order to test whether this is the case, we have to break the mass-environment degeneracy. Similarly to the environment bins, we divide our sample into three equal-number absolute magnitude bins. Together with the three environment bins, this gives us three curves of RSF percentage as a function of environment like Figure \ref{fig12}. These nine sub-samples are indicated by the dashed lines on Figure \ref{fig9}. The resulting curves are shown in Figure \ref{fig12}. From this, it is apparent that the effect of environment that we are seeing is not due to a stellar mass effect, as all three curves follow almost the same trend of a high RSF fraction at low density and a low RSF fraction at high density. Intriguingly, the high mass bin ($-23.82 < M_{r} \leq -22.13$) departs from the others at high density, though this remains just above a 1-$\sigma$ result. It should also be noted that the strongest density dependence is found among the brightest galaxies. \section{Summary} We have used the UV color-magnitude relation of low redshift, massive early-type galaxies to study their recent star formation history. Our sample is volume-limited, ranging in redshift from z=0.05 to 0.1 and is limited in absolute magnitude to $M_{r} < -21.5$. Our sample is highly unlikely to be contaminated by any significant number of late-type galaxies, as all our galaxies have been visually inspected. In order to classify galaxies by their environment, we have devised a method for measuring environment that takes the proximity, and not just the number density of neighboring galaxies, into account (see Section \ref{env_section}). This method can easily be modified to different samples within SDSS and can take into account a larger part of the luminosity function if restricted to lower redshift limits than $z=0.1$. Our measure works very well for the brightest cluster galaxies in the C4 cluster catalog and in addition also performs for field galaxies. In our sample of 839 early-type galaxies with $M_{r} < -21.5$, the recent star formation (RSF) galaxy fraction is $30 \pm 2$\%. Our ellipticals, the bulk of our sample, have an RSF fraction of $29\% \pm 3$, while the lenticulars show $39\% \pm 5$. This implies that \textit{residual star formation is common amongst the present day early-type galaxy population}. Our estimates are very likely lower limits on the true fractions, as our criteria for RSF are conservative in the consideration of internal extinction and the UV contribution from the old populations. The UV color-magnitude relation differs from the optical color-magnitude relation \citep{1992MNRAS.254..589B, 2004ApJ...601L..29H} in that it does vary more clearly with environment. The recent star formation history of early-type galaxies also varies with environment. It is well-known that more massive galaxies reside in higher-density environments (Figure \ref{fig9}), but we show for the first time that UV-bright early-type galaxies preferentially reside in low density environments. The RSF fraction is a function of environment and drops by 25\% from field to group but then puzzlingly remains relatively constant at higher densities, even when split into luminosity bins (Figure \ref{fig12}). Interestingly, the most massive galaxies ($-23.82 < M_{r} \leq -22.13$) show the strongest dependence on environment and alone exhibit a further drop in RSF fraction from medium to high density. One possible way to understand the drop in the RSF fraction between low and medium density is in the context of ram pressure stripping. Galaxies moving fast in the deep gravitational potential of a galaxy cluster are bound to lose most of their gas during their orbital motion \citep{1972ApJ...176....1G}. The density dependence of gas content in galaxies has long been established empirically as well \citep{1981ApJ...247..383G}. Our RSF fraction-density relation is in the right direction. The gas goes into the ICM and so could potentially explain the star formation we do see in high density environments. Another noteworthy observation is the fact that those early-type galaxies which have been identified as AGN - by emission lines and/or radio - are significantly bluer than those who are not (see Figure \ref{fig3}). We have removed these AGN from our sample since we cannot disentangle the UV flux from the AGN from that of a possible young stellar population. However, the blueness of the AGN colors is intriguing - are we really just seeing the AGN itself, or is this from the star formation triggered by the jets and outflows from the AGN \citep{1998A&A...331L...1S}? In the latter case, the RSF fraction would increase further from our estimates, and the AGN regulations might present a possible physical mechanism responsible for the star formation that we observe. In fact, \citet{1995ApJ...452..549H}, using IUE observations of nearby Type 1 and 2 Seyferts, suggest that at most 20\% of the UV continuum emission seen in them can originate from the nucleus itself. This means that the vast majority of our AGN candidate (removed) would qualify as RSF galaxies. It is important to note that we only deal with \textit{fractions} of star-forming galaxies and not actual star-formation rates. Thus, our RSF fractions simply give us an indication of how likely an early-type galaxy with certain properties and in a certain environment is to have experienced recent star formation. Neither environment, luminosity nor axis ratio seems to be the primary physical quantity that regulates recent star formation in early-type galaxies. The relative insensitivity to environment in any environment denser than the field is also surprising and warrants further study. The observational trends presented here give us new constraints for theoretical models of galaxy evolution. \acknowledgements Special thanks are given to M. Bernardi who kindly supplied her early-type galaxy catalog which provided us with a great insight on our catalog generation. We warmly thank C. Wolf for making the \textit{COMBO-17} S11 field image available to us. We would also like to thank E. Gawiser, L. Miller, S. Rawlings, J. Silk, R. Davies, I. Jorgensen, M. Sarzi, J. Magorrian, S. Salim, M. Urry and K. Kotera for helpful comments and discussions. GALEX (Galaxy Evolution Explorer) is a NASA Small Explorer, launched in April 2003. We gratefully acknowledge NASA's support for construction, operation, and science analysis for the GALEX mission, developed in cooperation with the Centre National d'Etudes Spatiales of France and the Korean Ministry of Science and Technology. This was supported by Yonsei University Research Fund of 2005 (S.K.Yi). \bibliographystyle{astroads} \bibliography{bibliography}
Title: Constraints on SN Ia progenitor time delays from high-z SNe and the star formation history
Abstract: We re-assess the question of a systematic time delay between the formation of the progenitor and its explosion in a type Ia supernova (SN Ia) using the Hubble Higher-z Supernova Search sample (Strolger et al. 2004). While the previous analysis indicated a significant time delay, with a most likely value of 3.4 Gyr, effectively ruling out all previously proposed progenitor models, our analysis shows that the time-delay estimate is dominated by systematic errors, in particular due to uncertainties in the star-formation history. We find that none of the popular progenitor models under consideration can be ruled out with any significant degree of confidence. The inferred time delay is mainly determined by the peak in the assumed star-formation history. We show that, even with a much larger Supernova sample, the time delay distribution cannot be reliably reconstructed without better constraints on the star-formation history.
https://export.arxiv.org/pdf/astro-ph/0601454
\date{Accepted 2006 March 01. Received 2006 February 28; in original form 2006 January 19} \label{firstpage} \begin{keywords} supernovae: general --- cosmology: observational. \end{keywords} \section{Introduction} Type Ia supernovae have been used extensively as standard distance indicators and have provided the best evidence to date for an acceleration of the Universe \citep{rie98, per99, rie04}. Future missions, e.g. \emph{GAIA} and \emph{SNAP}, will greatly increase the number of detected SNe Ia and significantly reduce the statistical errors in the determination of cosmological parameters. However, the nature of the progenitors of type Ia supernovae is still unknown and the empirically calibrated \emph{Phillips relation} \citep{phi93} is not fully understood physically. Several progenitor scenarios are under discussion, but there is no consensus due to uncertainties in the evolutionary processes \citep{HKN96, HKN99, LV97, lan00, HP04} and the explosion mechanism \citep{HN00, roe05, gam05}. One of the signatures of the various scenarios is the distribution of time delays between the formation of the progenitor systems and their explosion, which could give rise to a significant difference between the redshift dependence of the supernova rate (SNR) and the star-formation history (SFH). \citet{str04}, hereafter S04, aimed to detect this difference by studying the distribution of 25 high-z SNe Ia in the \emph{Hubble} Higher-z Supernova Search sample \citep{rie04}. Their approach was to infer the mean time delay of the distribution using a Bayesian analysis, which assumed different parametrized time-delay distributions and adopted the star-formation history (SFH) from \citet{gia04}, hereafter G04. They concluded that mean time delays shorter than $\sim 2$~Gyr ought to be excluded with a 95 per cent confidence level, ruling out essentially all progenitor scenarios currently under discussion. In a recent re-assessment of the constraints, \citet{str04erratum} obtained a 95 per cent lower limit ranging from 0.2 to 1.6~Gyr for different time-delay distributions. In the corrected best-fitting model, the 95 per cent confidence interval ranged from 1 to 4.4~Gyr with the most likely value at 3.4~Gyr. Unlike core collapse SNe (SNe II, Ib/c) that originate from massive progenitors with relatively short main-sequence (MS) lifetimes ($\sim 3-20$~Myr), SNe Ia are believed to be thermonuclear explosions of white dwarf stars (WDs) whose progenitors have MS lifetimes ranging from $\sim 30$~Myr to several billion years. This implies a minimum time delay for SNe Ia of the order of $\sim 30$~Myr. Most of the SN Ia progenitor scenarios that have been proposed involve mass transfer on to a CO WD in a binary system, either through the expansion and Roche lobe overflow of an evolved companion (\emph{single degenerate [SD] scenarios}) or through the slow release of gravitational waves, orbital shrinking, Roche lobe overflow and merging of a compact double WD system (\emph{double-degenerate [DD] scenarios}). Both scenarios have associated time-delay distributions that have been estimated with binary population synthesis codes (BPS), where the properties of binary systems are followed from their birth up to the explosion stage through the many different evolutionary paths. Independently of the particular treatment of the binary interactions, the resulting time-delay distributions differ considerably since their characteristic time-scales have different origins. The SD scenario is controlled by the process of mass accretion, which has to occur at just the correct critical rate in order to allow the growth of the mass of the companion WD up to the Chandrasekhar limit \citep{nom91}. The dominant evolutionary path seems to occur via the accretion of matter on to a CO WD from a slightly evolved MS star, the CO WD + MS -- SD scenario \citep{vH92, rap94, LV97, lan00, HP04}. In this channel, the accretion rate is determined mainly by the mass of the donor star, which must lie in a narrow range in order to satisfy the required accretion-rate constraints. As a consequence, the distribution of MS lifetimes and the time-delay distribution of the channel are relatively narrow, peaking at $\sim 670$~Myr and rapidly becoming negligible after $\sim 1.5$~Gyr. Although recent simulations have suggested that other evolutionary paths within the SD framework are of minor importance \citep{HP04}, it is quite possible that their contribution has been underestimated. This is particularly important for the CO WD + RG -- SD scenario, where a red-giant (RG) star accretes matter on to a CO WD star \citep{HKN96}; in this channel the time-delay distribution extends up to several Gyr. The DD scenario \citep{IT84, web84}, in contrast, is controlled by the time that it takes for the binary system to coalesce, which depends roughly on the fourth power of the separation of the double-degenerate system \citep{sha83}. As a result, the time-delay distribution can be described by a low time-delay cutoff ($\sim 30-100$~Myr) and an approximately power-law decline up to the age of the Universe. The lower time-delay cutoff can be explained by the time required to form the most massive degenerate systems with the shortest MS lifetimes, whereas the power-law tail can be explained by the power-law relation between coalescence time and separation of the double-degenerate systems. However, the expected accretion rates in the DD scenario are a problem: they are so high that present calculations suggests that this leads to accretion-induced collapse (AIC) and the formation of a compact object rather than a thermonuclear explosion \citep{NI85, SN85, SN98, TWT94, nom91}. Therefore, the currently generally most favoured progenitor scenario is the SD scenario. Because the seemingly dominant evolutionary path of this channel would need to be discarded if the mean time-delay were found to be higher than 2~Gyr, it is important to confirm the significance of the S04 results. In this work we have studied the SN Ia time-delay distribution using the sample of S04 and the same basic analysis, but introducing alternative SFHs found in the literature, avoiding binning effects as much as possible and using a Goodness of Fit (GoF) test that is generally recommended for small samples. We discuss the data and analysis in Section \ref{sec:analysis}, show the results and Monte Carlo simulations in Sections \ref{sec:results} and \ref{sec:Monte Carlo} and discuss their significance in Sections \ref{sec:discussion} and \ref{sec:conclusions}. Throughout his paper, we adopted a value for the Hubble constant of $H_{\rm 0} = 70$ km s$^{-1}$ Mpc$^{-1}$, present ratios of matter, curvature and dark energy density over the critical density of $\Omega_M = 0.3$, $\Omega_K = 0$ and $\Omega_{\Lambda} = 0.7$, respectively, and a `dark energy' pressure over density ratio ('equation of state') of $w =-1$. \section{Data and Analysis} \label{sec:analysis} The analysis is based on the \emph{Hubble} Higher-z Supernova Search sample \citep{rie04, dah04, str04}, which contains 25 SNe Ia found in the GOODS field (13 in the Hubble Deep Field North, HDFN, and 12 in the Chandra Deep Field South, CDFS) in the redshift range 0.21 to 1.55. The SNe were discovered in four difference images that were produced by observing both fields five times in intervals of approximately one month, comparable to the typical duration of the main SN Ia light curve peak. To infer the underlying time-delay distribution of SNe Ia, S04 compared the observed redshift distribution in the sample with a parametrized predicted distribution, derived from the G04 SFH convolved with three alternative time-delay distributions. Each distribution was parametrized by its mean time delay, which was recovered using a Bayesian analysis. Among these, the distribution that best fit the data was a `narrow Gaussian', which after being corrected was centred on 3.4~Gyr with a FWHM of $\sim 1.5$~Gyr. The 95 per cent confidence interval for the mean time delay ranged from 1.0 to 4.4~Gyr. The alternatives `wide Gaussian' and e--folding distributions had a mean time delay above 0.2 and 1.6~Gyr, respectively, with more than 95 per cent confidence. Only the shapes of the distributions are compared, i.e. the analysis is scale--free, and the associated efficiencies of SNe per unit formed mass are calculated later by normalising the models to the SN numbers and are not used to constrain the models. This means that the sample must ideally span a redshift range that includes both the rising and declining parts of the SNR, i.e. where the SNR is not approximately linear. A recent study \citep[see][Fig.~6--8]{BT05} could not fully exploit information on the SN redshift distribution because their sample did not reach to a sufficiently high redshift ($z>1$), as the authors indicate in the text. A similar situation is found in the work of \citet{GYM04}. Thus, because the \emph{Hubble} Higher-z Supernova Search sample is the deepest SN sample available, it is the most suitable one for constraining the time-delay distribution of SNe Ia. However, the formal errors quoted in S04 reflect only the limited size of the sample and not other systematic uncertainties, such as those associated with the SFH. In the following Sections \ref{sec:SNR} to \ref{sec:tc} we introduce the formalism that gives the SNR, the number of detected SNe per unit redshift and the control times used in the derivations. In Section \ref{sec:timedist} we discuss alternative time-delay distributions and in Section \ref{sec:SFH} alternative SFHs. The Bayesian analysis is described in Section \ref{sec:Bayes} and further modifications concerning binning effects and the GoF test are discussed in Section \ref{sec:newanalysis}. \subsection{The SN Ia rate} \label{sec:SNR} The rate of SNe Ia per unit time per unit co-moving volume ($SNR_{\rm Ia}$) is given by the star-formation rate per unit time per unit co-moving volume ($SFR$) convolved with the normalised distribution of explosions per unit time of the progenitor channel (the time-delay distribution, $\phi$), and multiplied by the number of SNe per unit formed mass (the efficiency, $\nu$). We assume that neither $\nu$ nor $\phi$ evolve with redshift: \begin{align} SNR_{\rm Ia}(z) = \nu \int_{\rm t(z_R)}^{t} SFR(t')\ \phi(t - t', \tau) \ dt' \label{eq:SNR}, \end{align} where $t = t(z)$, $\tau$ is some characteristic time-scale defined in Section ~\ref{sec:timedist} and $z_R$ is the redshift associated with the time when the first stars formed, approximately the epoch of reionisation. We assumed $z_R =10$, as in S04. \subsection{Distribution of detected SNe} \label{sec:nIa} The number of detected SNe Ia per unit redshift interval ($n_{\rm Ia}$) is given by the multiplication of the rate of SNe Ia per unit time per unit co-moving volume ($SNR_{\rm Ia}$), a time dilation factor, $(1+z)^{-1}$, the control time of the survey ($t_{\rm c}$) and the volume per unit redshift being surveyed, $\displaystyle{ \frac{dV}{dz d\omega} \Delta \omega}$: \begin{align} n_{\rm Ia}(z) = \frac{SNR_{\rm Ia}(z)}{1+z} \ t_{\rm c}(z) \ \frac{\mathrm{d}V(z)}{\mathrm{d}z \mathrm{d}\omega} \ \Delta\omega \label{eq:nIa}, \end{align} where in our cosmology the volume derivative formula simplifies to: \begin{align} \frac{\mathrm{d}V}{\mathrm{d}z \mathrm{d}\omega} = d_C^2 \frac{\mathrm{d}(d_C)}{\mathrm{d}z}\text{, where} \end{align} \begin{align} d_C = c H_0^{-1} \int_0^z \mathrm{d}u \left[ (1+u)^3 \Omega_M + \Omega_\Lambda \right] ^{-1/2} \label{eq:dV1}, \end{align} and hence, \begin{align} \frac{\mathrm{d}V(z)}{\mathrm{d}z \mathrm{d}\omega} = c H_0^{-1} \left[ (1+z)^3 \Omega_M + \Omega_\Lambda \right] ^{-1/2} d_C^2(z) \label{eq:dV2}. \end{align} \subsection{The control time} \label{sec:tc} The control time can be understood as the total observing time multiplied by the probability of detecting a SN at a given redshift. We used the same values as S04, that were calculated taking into account the expected extinction, spectra, light curve shapes and peak magnitude dispersion of SNe Ia, the way each field was revisited, and the efficiency of the detection algorithm (but see Section \ref{sec:binning}). The control times are defined by \begin{align} t_{\rm c}(z) = \iiint P(t | M_\lambda, A_\lambda, z) \ P(M_\lambda) \ P(A_\lambda) \ dM_\lambda \ \mathrm{d}A_\lambda \ \mathrm{d}t, \end{align} where $P(t | M_\lambda, A_\lambda, z)$ is the probability of detecting a new SN at time t, given its rest-frame luminosity and its host galaxy extinction and redshift. It depends on the assumed spectra through K--corrections, the sensitivity of the survey and the efficiency of the detection algorithm. $P(M_\lambda)$ is the probability of having a given SN rest-frame luminosity. It was estimated based on the characteristic relation between peak luminosity and light curve shape of SNe, and the observed dispersion of SN Ia peak luminosities. $P(A_\lambda)$ is the probability of having a given host galaxy extinction at the given rest-frame wavelength. It was assumed to be proportional to $e^{- A_\lambda}$. For more details see the original discussion in S04. \subsection{Time delay distributions} \label{sec:timedist} All the time delay distributions were parametrized by their mean time delays, $\tau$. S04 used an exponential distribution and two families of Gaussian distributions whose width scale with the mean time delay. The e-folding distributions are given by: \begin{align} \phi(t,\tau) = \frac{e^{-t/\tau}}{\tau}. \end{align} The two alternative Gaussian distributions are grouped into the families of `narrow' ($\sigma_{\tau} = 0.2 \tau$) and `wide' ($\sigma_{\tau} = 0.5 \tau$) distributions, of the form: \begin{align} \phi(t,\tau) = \frac{1}{\sqrt{2 \pi \sigma_{\tau}^2}} e^{ -\tfrac{(t - \tau)^2}{2 \sigma_{\tau}^2} }. \end{align} The previous distributions are defined only for positive values. However, negative time delays must be allowed in order to avoid statistical bias in a small sample and to get confidence intervals that do not artificially discard short time delays. Moreover, a preference for negative time delays would signal SFHs that peak too late in time. Thus, we considered a fourth time delay distribution, a Gaussian distribution with fixed width ($\sigma = 0.5$~Gyr) that allows either for positive or negative time delays: \begin{align} \phi(t,\tau) = \frac{1}{\sqrt{2 \pi \sigma^2} } e^{ -\tfrac{ (t-\tau)^2 }{ 2 \sigma^2 }}. \end{align} We also added a log-normal distribution, which is associated with processes where the source of uncertainty has multiplicative effects rather than additive ones, as is the case for Gaussian distributions. The best-fitting models to the theoretical time delays were in most cases log-normal distributions, whose width $\sigma$, in units of $\rm{log}(t)$, was kept fixed and determined by the best-fitting model of the theoretical time delays: \begin{align} \phi(t,\tau) = \frac{1} {\sqrt{2 \pi \rm{ln}(\sigma)^2}} e^{- \tfrac{\rm{ln}(t/\tau)^2} {2\ \rm{ln}(\sigma)^2} } \frac{1}{t}. \end{align} The theoretical time delay distributions from \citet{HP04} were also examined with a GoF test. They were produced assuming either the CO WD + MS -- SD scenario or the DD scenario with different binary evolution parameters. In Fig.~\ref{fig:loglog-TheoryDelay} the theoretical time delay distributions of the SD and DD scenario together with two time delay distributions with different mean time delays are plotted, including the best-fitting model from S04. \subsection{The Star-Formation History (SFH)} \label{sec:SFH} Because we consider alternative prescriptions for the SFH that have incomplete redshift information, further complications arise. The ideal redshift coverage of the SFH should range from zero to the redshift of the first star formation, farther than the highest-redshift object presently known in the Universe. This is a consequence of the redshift range of the detected SNe and the long time delays that have to be considered. If high-redshift SNe only exploded after long time delays, their progenitors would need to form at redshifts up to $\sim 30$, as Fig.~\ref{fig:SNeHisto} shows. If the determination of the SFH did not cover the required redshift range, we used as an approximation either a power-law extrapolation in time or a scaled version of the optical--UV derivation. We found that the method is not very sensitive to this approximation when the position of the peak of the SFH is well constrained, since it is the difference between the peaks of the SFH and the SNR that mainly constrains the best-fitting models. The alternative prescriptions of the SFH we have used are the following: \begin{itemize} \item The SFH by \citet{gia04}, G04. We have used continuous approximations for the extinction corrected and not corrected models, inferred from deep optical--UV observations of galaxies in the GOODS field. Both versions differ by a factor of $\sim 3$ in the redshift range of interest. The continuous approximations are the ones used in S04, which peak at $z \sim 2.7$ in the extinction corrected model (M1) and at $z \sim 1.8$ in the model that is not corrected for extinction (M2). \item The best-fitting model of \citet{CE01}, hereafter CE01. It was derived from the integrated cosmic infrared background (CIRB) and covers the redshift range from 0 to 4.5. At $\rm{z > 4.5}$ we tried a power-law extrapolation in time or a scaled version of G04 at $\rm{z > 3}$. Mainly because the peak of the SFH occurs very late in time, at $z \sim 0.8$, we found that the inferred time delays are not very sensitive to this approximation. However, a constant SFH model is within the error bars at high--z. \item The SFH from \citet{hea04}, hereafter H04, inferred from the `fossil record' of stellar populations in the \emph{Sloan Digital Sky Survey} (SDSS). We interpolated a Spline function to the binned SFH, which peaks at $z \sim 0.4$, in order to obtain a smooth approximation. It is not usually recommended to approximate data in this way, but we think that our approximation preserves the general differences between this SFH and the alternative prescriptions. We also tried a scaled version of G04 at $\rm{z > 3}$, or a constant star-formation history for all times, since this SFH is very flat at $z \gtrsim 1$ and peaks very late in time with respect to the detected SNe. Both alternatives gave very similar results to the fit to the original binned data. \end{itemize} Additionally, we considered one of the most recent determinations of the SFH using \emph{Spitzer} data, presented by \citep{PG05}. However, both its limited redshift coverage and its dependence on the assumed galaxy luminosity function makes the high-redshift extrapolation ambiguous. For this reason, we did not try a continuous approximation of this SFH in the analysis, although it must be considered a reliable result. The variance between its different versions only demonstrates the persisting uncertainties in our knowledge of the SFH. In Fig.~\ref{fig:loglog-SFH} we show the four continuous approximations of the SFH and the binned SFH from PG05. The continuous approximations are based on the extinction corrected and not corrected SFH from G04, the best-fitting model from CE01 with a power-law extrapolation in time at high redshift and a continuous approximation of the best-fitting model from H04, which is constant at high redshift. It is apparent that there is a range of SFHs in the literature that do not agree and, importantly, peak at very different times. Thus, it is important to understand the systematic errors associated with this uncertainty. For a recent estimation of the uncertainties on the SFH see also Fig.~2 and 4 from \citet{HB06}. \subsection{The Bayesian probability} \label{sec:Bayes} Using Bayes theorem with a uniform prior, the probability of a mean time delay $\tau$ with a time-delay distribution $\phi(t, \tau)$ and a SFH in the form of $SFR(t')$, given the set of SN redshifts, $\{ z_i \}$, is proportional to: \begin{align} P( SFR(t'), \phi(t, \tau), \tau \vert \{z_i\}) \varpropto P( \{z_i\} \vert SFR(t'), \phi(t, \tau), \tau). \end{align} Thus, it is proportional to the probability of the particular SN redshift distribution: \begin{align} P( \{z_i\} \vert SFR(t'), \phi(t, \tau), \tau) \varpropto \prod_{i=1}^{25} n_{\rm Ia}(z, \tau), \end{align} where $n_{\rm Ia}(z,\tau)$ has been normalised for every $\tau$ because the analysis is scale free. It depends on $\tau$ through the time-delay distributions of Section \ref{sec:timedist} and equations \ref{eq:SNR} and \ref{eq:nIa}. Hence, for a given combination of SFH, time-delay distribution and mean time delay, the predicted number of SNe per unit redshift can be expressed as a probability distribution in redshift. Subsequently, the probability of the set of SNe can be calculated for every $\tau$. \subsection{New analysis} \label{sec:newanalysis} The main differences between the S04 calculations and this work are a result of considering a range of alternative SFHs. Further differences are as follows: \subsubsection{Redshift binning effects} \label{sec:binning} The main advantage of using the observed redshift distribution of SNe Ia instead of the corresponding SNR \citep{GYM04} is that the analysis can be done in a way that avoids binning and the subsequent loss of information. Moreover, we have found that the analysis is very sensitive to the way the volume derivative is calculated in equation \ref{eq:nIa}. Because the SN sample is small, binning the data is not recommended, and all the calculations should be done continuously. Binning can introduce a relative overestimation of the volume derivative at low redshift, effectively pushing the most probable time delays to higher values. As a result, the lower limit of the Bayesian analysis can be overestimated by more than $\sim 1$~Gyr and the peak of the Bayesian probability distribution by $\sim 500$~Myr, according to our calculations. This is consistent with the corrected results in \citet{str04erratum}. However, in order to get a continuous version of the control times we have used an interpolation of the values calculated in S04 at redshift intervals of $0.2$. Recalculating the control times with more redshift resolution would be a better approach, but we have not tried it in this analysis. \subsubsection{Goodness of fit test (GoF)} A maximum likelihood analysis must be accompanied by a GoF test to check that the parametrized model has an appropriate form to start with, and only then can the confidence intervals be trusted at all. Accordingly, a $\chi^2$ test was used to check consistency between the predicted and observed redshift histograms of SNe Ia in S04. However, the $\chi^2$ test is not reliable when the number of elements per bin is not greater than five in 80 per cent of the bins \citep{WJ03}. Instead, we have used a Kolmogorov--Smirnov test (KS test) as our goodness of fit test, which is the recommended test to use when the sample size is small and because the analysis is done continuously. Furthermore, selecting the best-fitting models with the KS test can be used as an alternative parameter estimator. \subsubsection{Confidence intervals} \label{sec:ConfInt} In S04 the confidence intervals were obtained starting from the mode (the maximum) of the Bayesian probabilities, partly because in the original results the probabilities at low time delays were negligible. However, because in some cases the probability distributions are relatively flat, or not negligible at zero time delay, the definition of the confidence intervals becomes important. In our calculations, taking 95 per cent confidence intervals around the maximum vs. from the median can make a difference of typically $\sim 500$~Myr. One way to avoid this problem it to use the parameter region that is not rejected by the GoF test with a certain confidence level. In this approach, we obtain 95 and 68 per cent confidence intervals that are unambiguously defined. \subsubsection{Photometric redshifts} Because the spectra of SNe are characterised by many blended lines broadened by high velocity dispersion, SN redshifts are determined from their host galaxies. Of the 25 SNe Ia in the sample, six have only photometric redshifts, three of them in each field. We found that the photometric redshift of SN 2003al, $0.91 \pm 0.2$, has a better estimate in the public COMBO--17 catalogue \citep{wol04} of $0.82 \pm 0.04$. Additionally, in \citet{str05}, the photometric redshift of SN 2003lu, $0.11 \pm ^{0.13}_{0.11}$, has a better estimate of $0.14 \pm 0.01$. \section{Results} \label{sec:results} \subsection{Varying the SFH} \label{sec:resultsSFH} If alternative SFHs are allowed, the Bayesian probabilities associated with a given time-delay distribution have a wide range of preferred time delays. The Bayesian probabilities and KS test associated rejection probabilities show a preference for values ranging from very long ($\sim 4$~Gyr) to very short, and even negative ($\sim -3$~Gyr) if the SFH peaks very late in time (see Fig.~\ref{fig:probs}). The relation between the peak of the SFH and the peak of the Bayesian probability distribution is, to zeroth order, such that later peaked SFHs give shorter time delays. Interestingly, inspection of Fig.~\ref{fig:probs} (upper-left panel) shows two types of maxima in the Bayesian probabilities: one whose position decreases with later peaked SFHs and another that is fixed at approximately $\sim 3.5$~Gyr, even for different SFHs. The first peak approximately reflects the time difference between the peaks of the SFH and the SNR. The second peak reflects the relative absence of SNe at high--z. Because no SNe were detected between the epoch of reionisation and $z \sim 1.5$, or between $t \sim 0.5$~Gyr and $t \sim 4$~Gyr, the Bayesian analysis marginally favours models that do not produce SNe in the first $\sim 3.5$~Gyr after the assumed epoch of reionisation. The upper plot of Fig.~\ref{fig:SNeHisto} is illustrative of this effect. With the current data, it is the first peak which is statistically dominant for all the SFHs, but this may change with deeper and wider SN surveys in the future. \subsection{Kolmogorov--Smirnov test} With the KS test we find best-fitting mean time delays and confidence intervals that are free from the problems explained in Section \ref{sec:ConfInt}. The rejection probabilities for the `fixed width Gaussian' and e-folding time-delay distributions are shown in Fig.~\ref{fig:probs}. We found that all the combinations of SFH and time-delay distributions had an associated parameter region that is accepted by the KS test, which validates the use of the Bayesian analysis. Additionally, the parameter estimation seems robust in the sense that it gives results that are consistent with what the Bayesian analysis shows. Moreover, the addition of this test shows that the negative time-delay peak for the H04 SFH is favoured over the long time-delay peak (see Fig.~\ref{fig:probs}), which may be an indication that this SFH is not compatible with the SN data. \subsection{Confidence intervals} The confidence intervals were defined as the parameter region that cannot be rejected with a certain confidence level based on the KS test. We found that only the extinction corrected SFH from \citet{gia04} has a 95 per cent confidence lower limit greater than zero, i.e. around $1$~Gyr. All the alternative SFHs did not result in a lower limit for the time delays greater than zero. A summary of the confidence intervals obtained with the Gaussian and e-folding time delay distributions is shown in Fig.~\ref{fig:summaries}. \subsection{Varying the time-delay distribution} The non-rejection regions of the five time-delay distributions tested in this work can be grouped into three families of results: one family associated with the `fixed width Gaussian' (that allows for negative time delays) and `narrow Gaussian' distributions, another with the `wide Gaussian' and log-normal distributions and one associated with the e-folding distribution. As a general rule, the narrower the test time-delay distribution, the narrower the associated Bayesian probabilities. However, it is in the long time-delay region where the changes are more noticeable, as can be seen in Fig.~\ref{fig:summaries}. This is because the abrupt transition that occurs in the SFH at the epoch of reionisation is reflected in a less smooth SNR when narrower time-delay distributions are assumed. Hence, the wider the time-delay distribution, the less pronounced the second peak in the Bayesian probabilities (see Section ~\ref{sec:resultsSFH}) and the longer the time delays allowed. \subsection{Theoretical time-delay distributions} We performed KS tests of the theoretical time-delay distribution varying the BPS parameters and the assumed SFHs. As a result, if the extinction corrected SFH from G04 is assumed, the CO WD + MS -- SD scenario alone has a 3 per cent probability of not being rejected and the best-fitting models are obtained for a DD scenario which has a very high mass transfer efficiency. Conversely, if any of the theoretical models are assumed to be true, the best-fitting models are, in almost all the combinations, obtained with the SFH from \citet{CE01}. In Table \ref{tab:summary_models} we show the non-rejection probabilities for the different combinations of SFH, theoretical scenario and BPS parameters tested in this work. The BPS parameters are: $\alpha_{\rm CE}$, the common--envelope ejection efficiency efficiency as in \citet{HP04}; $\alpha_{\rm RLOF}$, the Roche lobe overflow mass transfer efficiency, and $Z$, the metallicity. \begin{table}\centering \caption{Summary of the KS non-rejection probabilities (per cent) for different combinations of SFH, theoretical time-delay distribution and BPS parameters. Unless stated otherwise, the standard parameters are $\alpha_{\rm CE} = 1.0$~, $\alpha_{\rm RLOF} = 1.0$~ and $Z = 0.02$~.} \begin{tabular}{@{}l ccc c ccc@{}} \cmidrule{1-8} & \multicolumn{3}{c}{SD scenario -- $\alpha_{\rm{CE}}:$} & & \multicolumn{3}{c}{DD scenario -- $\alpha_{\rm{CE}}:$} \\ \cmidrule{2-8} SFH & 0.5 & 0.75 & 1.0 & & 0.5 & 0.75 & 1.0\\ \cmidrule{1-8} G04 (M1) & 3.1 & 3.0 & 3.1 & & 5.7 & 11.4 & 25.9 \\ G04 (M2) & 12.2 & 11.7 & 12.4 & & 15.8 & 29.1 & 51.6 \\ CE01 & 49.8 & 47.8 & 50.8 & & 48.1 & 69.2 & 85.4 \\ H04 & 28.1 & 28.1 & 28.0 & & 25.7 & 24.2 & 21 \\ \cmidrule{1-8} & \multicolumn{3}{c}{DD scenario -- $Z:$} & & \multicolumn{3}{c}{DD scenario -- $\alpha_{\rm{RLOF}}:$} \\ \cmidrule{2-8} SFH & 0.001 & 0.004 & 0.02 & & 0.5 & 0.75 & 1.0 \\ \cmidrule{1-8} G04 (M1) & 11.0 & 13.2 & 25.9 & & 25.9 & 47.4 & 56.5 \\ G04 (M2) & 30.1 & 31.2 & 51.6 & & 51.6 & 82.6 & 78.5 \\ CE01 & 70.2 & 71.2 & 85.4 & & 85.4 & 80.7 & 62.5 \\ H04 & 25.2 & 23.1 & 21.0 & & 21.0 & 18.1 & 17.3 \\ \cmidrule{1-8} \end{tabular} \label{tab:summary_models} \end{table} \subsection{Constraints on the SFH assuming the theoretical models} In addition to the previous constraints on the time delay distribution we followed the approach of \citet{GYM04} to constrain the SFH. The method assumes a particular time delay distribution and a SFH that consists of a broken power--law smoothly joined at the transition redshift $z_0$, proportional to $\sim (1+z)^\alpha$ at high $z$ and to $\sim (1+z)^\beta$ at low $z$, i.e.: \begin{align} SFR(z) \propto \left\{ \left( \frac{1+z_0}{1+z} \right)^{5 \alpha} + \left ( \frac{1+z_0}{1+z} \right) ^{5 \beta} \right\} ^{-1/5}. \label{eq:SFRGY} \end{align} For more details see \citet{GYM04}. We assumed the theoretical time delay distributions and performed a KS test for different combinations of power--law indices and transition redshift, $z_0$. We marginalised the probabilities on $\alpha$ because it was found that the dependency on this parameter is weak, which is expected because the SN sample contains very few objects at high redshift. The resulting probability distributions for different values of $z_0$ and $\beta$, assuming either the SD or DD scenario is shown in Fig.~\ref{fig:KS_z0beta}. Virtually all measurements of the SFH suggest $SFR(z) \propto (1+z)^\beta$, with $\beta$ in the range $\sim$~2 to 4 \citep{wil02, PG05}. For $\beta \gtrsim 2$ we find that the peak of the SFH is most likely between $z \sim 0.7$ and $1.2$. However, any rejection at 90 per cent confidence is only possible for a peak location at $z \gtrsim 2$. The flatter the recent decline in the SFH (lower $\beta$), the less clearly constrained is the peak of the SFH. \section{Monte Carlo simulations} \label{sec:Monte Carlo} In this section we examine whether our approach is subject to systematic biases due to the method itself and the choice of the SFH. To assess the robustness of the method, we performed Monte Carlo simulations for each progenitor scenario separately, where we took large sets of simulated SNe (typically 10,000) to minimize statistical errors. Because the DD scenario can extend to times comparable to the age of the universe, its time delay distribution cannot be sampled completely; we therefore expect that the recovered values will always be biased towards shorter mean time delays. We found the following: (1) such an expected bias is indeed present for the DD scenario, but is always below $0.5$~Gyr; (2) for large sample sizes, the smallest statistical errors are obtained using the e-folding time delay distribution, independent of the theoretical scenario assumed. To quantitatively estimate the systematic errors associated with the choice of SFH, we proceeded in the following way: (1) in order to minimize the statistical errors and study the systematics, we performed Monte Carlo simulations with 10,000 mock SNe drawn from the theoretical scenarios, using the e-folding time-delay distribution in the analysis; (2) we produce a mock sample of SNe with a SFH that differs from the one used in the Bayesian analysis; (3) a comparison between the different recovered mean time delays then gives an estimate of the systematic error due to the choice of SFH (see Fig.~\ref{fig:Monte Carlo} and Fig.~\ref{fig:Monte Carlo2}). The result is consistent with what can be concluded from Section~\ref{sec:results}, i.e. that the bias on the mean time delay can be of the order of $\sim 2$~Gyr, even ignoring the H04 SFH. We then tested the robustness of the algorithm with small samples. In order to do this, we performed Monte Carlo simulations with 10,000 different sets of 25 mock SNe drawn from the theoretical models. We repeated the Bayesian and GoF analysis and found the following: (1) when the theoretical time delays are relatively short and the SFH peaks early in time, or when the time delays are long and the SFH peaks late in time, the Bayesian probabilities tend to have very positive or very negative skewness and, consequently, the mode of the Bayesian probabilities either overestimates or underestimates the theoretical mean time delays; (2) the distribution of non-rejection probabilities using the KS test is flat, meaning that there is no significant bias towards short or long time delays. Hence, we conclude that the mode of the Bayesian probability is not a good estimator for the mean time delays, while the KS test confidence intervals are robust even with small samples, general results consistent with the discussion in \citet{pre92}. \section{Discussion} \label{sec:discussion} \subsection{Large scale structure} Large-scale structure (LSS) effects are important in small pencil beam surveys when studying the time delay of SNe Ia, even when the SFH and the SNR have been measured in the same field. Usually, the SFR is measured as an `instantaneous' observed rate at a particular location in space and time, where for this study a prediction of the SNR will be needed for comparison. However, this prediction ought to be based on the SFR in the same position in space, not only at the time when the SNe are seen to explode, but at earlier times as well. Since the latter figure can not be measured directly, we have to obtain the SFR for earlier times by looking at a higher redshift, i.e. at a different location in space. Cosmic variance and the dependence of star formation on local environment will lead to a SFR measurement that is somewhat different from the past SFR of the target location in the same field. This SFR variance will appear whether the SFR is measured from the same fields or not, since the locations in space where the SFR and SNR are measured will be unrelated. A 1~Gyr time delay already translates into proper distance differences of $\sim 500$~Mpc at $z=1$. So, if we do not know the individual star-formation histories of galaxies in the supernova field, we are best advised to use our best knowledge of the cosmic SFH, while LSS still leaves an imprint on the observed SNR history. We investigate the large scale structure in the CDFS where the spectroscopic redshift survey \emph{VIMOS VLT Deep Survey} \citep[][VVDS]{LF04} overlaps almost precisely with the GOODS field and is complete to $M_{\rm{V}} < -19.5$ at $z<1$. In Fig.~\ref{fig:vvds_histo} we show a histogram of galaxy redshifts with $M_{\rm{V}}<-19.5$ in bin sizes chosen to contain a constant co-moving volume of $(25\ \rm{Mpc})^3$. A non-evolving and homogeneous galaxy distribution should appear flat in this representation. Well-known over-densities or wall-like structures are clearly apparent, especially in the redshift range from 0.5 to 0.75. Two wall-like structures near $z \approx 0.67$ and $z \approx 0.74$ are conspicuous in the distribution and have been observed in both x-ray and optical surveys \citep[see e.g.][for independent confirmation]{G03, wol04}. Also shown is a histogram of SNe Ia with secure (spectroscopic) redshift determinations. SNe with photo-z's only have been omitted due to the large error in redshift which makes any association with structures in the galaxy distribution difficult. The average bin in the redshift range from 0 to 1 contains 7.8 galaxies, whereas the average SNe-weighted bin contains 16.7 galaxies (2.14 times the normal average), a possible indication that the inferred SNR is affected by LSS in this field and redshift range. In order to understand how these numbers depend on the choice of our bins, we have repeated the calculations for bin widths of $(\Delta V)^{1/3} = (20,21.. 25)$~Mpc and found ratios of 2.52, 2.38, 2.15, 2.22, 2.00 and 2.14, respectively. Interestingly, the ratios that result using only SNe Ia-weighted bins are even higher: 2.87, 2.68, 2.88, 3.29, 2.55 and 2.97. The imprint of LSS in the SNR distribution could be corrected in the analysis of time delays. If the underlying mass density on a given direction $\hat n$ were of the form $\rho(\hat n,z)=\rho_0(z) (1 + \delta(\hat n,z))$, the over-densities could be corrected by multiplying the \emph{control times} by the same factor, $1 + \delta(\hat n, z)$. However, we do not correct for density variations, because we lack a good determination of the LSS with consistent quality in both fields. \subsection{SN rates and efficiencies} In the process of matching the observed number of SNe with a model, it is possible to explicitly calculate the SN production efficiencies and the supernova rate (SNR). Not only does the shape of the SNR history constrain the progenitor models via the delay times, but also the SN efficiency must be compatible with the model. As a first consistency test, we compare the directly measured SNRs with the parametrized estimates. In Fig.~\ref{fig:SNR} we show two versions using the results of the parameter estimation with the G04 (M1) and CE01 SFHs. Clearly, the extinction corrected SFH from G04 with the `narrow Gaussian' distribution is consistent with long time delays ($\sim 3.5$~Gyr), whereas the SFH from CE01 matches best with an e-folding distribution of shorter time delays ($\sim 1.5$~Gyr). The second alternative is only marginally favoured by the KS test, so both possibilities seem equally plausible. It is important to mention that in \citet{str05} a deeper SN search in the smaller Ultra Deep Field and its parallel fields (UDF/P) resulted in the detection of four additional SNe with $z<1.4$. The lack of high-redshift SNe is one of the predictions of the S04 best-fitting model and hence it supports the result. However, the authors also considered the SFH from CE01, but without any time delay, and concluded that it is not possible to rule out this SFH with more than 50 per cent confidence. The SN efficiencies required to explain the observed SNR pose a problem for the theoretical SD scenario, which produces too few SNe. This is true for all SFHs and combination of parameters, and amounts to a shortfall of $4\times$ to $10\times$ depending on the assumed SFH. The DD scenario can reproduce the required efficiencies for some combinations of BPS parameters. However, changes in the little constrained binary fraction and mass ratio distributions, or possibly in the initial mass function (IMF), may solve this problem in the future. One additional challenge is that the time-delay distribution or the supernova efficiencies may evolve with time: if the delay distribution is made up of several components from different SN production channels, their relative contribution could change with time; the supernova efficiencies themselves may evolve with time, especially considering that the accretion processes could depend on environmental factors such as the metallicity \citep{kob98}. The relative absence of SNe at high redshift could be a result of a metallicity effect, rather than a reflection of the time delays. If this is truly the case, our error bars on the time delays would be largely underestimated. The use of compatible star formation and chemical enrichment histories will be required to tackle this problem. \subsection{\emph{Spitzer} SFH} We have shown that it is crucial to determine the SFH that is to be used in the analysis. The recent determination of the SFH from \citet{PG05}, using three different extrapolations of the galaxy luminosity function, was shown in Fig.~\ref{fig:loglog-SFH}. This SFH was obtained by combining infrared \emph{Spitzer} observations with optical--UV data of galaxies in the GOODS field. The resemblance between this SFH and the CE01 best-fitting model justifies the use of the latter. Interestingly, the recent near--infrared and sub--millimetric determination of the SFH from \citet{WCB05} peaks at $z \sim 1$, similarly to what is obtained in CE01. These new results emphasize the persistent uncertainties in our knowledge of the SFH and the problem of extinction corrections in the optical. \section{Conclusions} \label{sec:conclusions} We have found that systematic errors associated with the use of alternative star-formation histories are comparable if not larger than the statistical errors reported in \citet{str04}. The position of the peak of the SFH was found to be the crucial parameter for the recovered time delays: Later peaked SFHs result in lower time delays and vice versa. Furthermore, the confidence intervals for the time delays depend on the functional form of the delay distributions assumed in the analysis. The use of wider time-delay distributions, in particular the e-folding model, gives considerably longer upper limits for the time delays. For the data set under investigation, we found that the KS test is better suited to obtain confidence intervals than a Bayesian analysis. The KS test confidence intervals are unambiguously defined, and we have confirmed their validity using Monte Carlo simulations. In contrast, the skewness of the Bayesian probabilities and the small sample size result in somewhat arbitrary confidence intervals. A KS test using the shape of the theoretical time-delay distributions shows that the extinction corrected model from G04 is incompatible with the CO WD + MS -- SD scenario, although other SFHs are compatible. The DD scenario cannot be rejected at 95 per cent confidence with any combination of SFH and time-delay distribution. Starting from theoretical time-delay distributions, they consistently favour the SFH from CE01. If we wish to constrain the time-delay distribution and possibly discard progenitor scenarios from the redshift distribution of supernovae, it is of foremost importance to determine the cosmic star-formation history more accurately. Otherwise, the uncertainty in the SFH will continue to limit the interpretation of SN data sets of any conceivable size. Secondly, we would need a better understanding of our progenitor models: if we ignore factors affecting efficiencies, their cosmic evolution will make the delay time distributions evolve and produce a situation where it is hard to disentangle these different effects without qualitatively different observations. Only after having solved these issues would deeper and wider surveys help to constrain the delay times and progenitor models by decreasing statistical noise and reducing the influence of environmental cosmic variance on the supernova samples. Different observations have already produced some evidence for shorter time delays: Recent work by \citet{man05} showed that the most efficient host galaxies for SNe Ia production among all galaxies are irregulars, and \citet{dv05} showed that among elliptical galaxies it is in particular the radio-loud galaxies, which are also believed to be associated with recent star formation. Also, \citep{gar05} have shown that SNe Ia of normal luminosity occur particularly in those elliptical galaxies with quite substantial star-formation rates. Also, SNe Ia in galaxy clusters seem to indicate time delays that are shorter than 2~Gyr \citep{MGY04}. Finally, it is clear that, if the SNe Ia phenomenon is composed of several production channels, all conclusions we have drawn apply to the dominant channel. A generally less common channel with different characteristics could again be the dominant channel in a subset of galaxies with older-age or higher-metallicity stellar populations. It is already established that under-luminous 91bg-type SNe Ia preferably occur in non-star-forming hosts, such as ellipticals. It remains to be explored theoretically, whether the CO WD + RG -- SD scenario could be related specifically to old populations. We should also consider the possibility that we may not even have found the dominant progenitor channel for normal SNe Ia \citep{ham03, tou05}. Even if the SFH peaks at redshift $\sim 1$ and the recovered time delays are consequently low, the associated SNR (see Fig.~\ref{fig:SNR}, lower panel) tends to be over-estimated in the highest-z bin. This effect could be interpreted as the signature of a long time delay component that does not contribute to the total SNR when the universe is too young. However, because the highest-z bin only contains two SNe, this does not lead to low non-rejection probabilities. A study of bimodal time delay distributions could only be done with this method if the uncertainties in the SFH were significantly reduced and the metallicity cutoff on the efficiencies was properly quantified. If different channels produce SNe Ia from progenitors of distinctly different age or metallicity, then an increase of the supernova sample could greatly help the identification of the various plausible progenitors. However, such a data set would most be beneficial if it is complemented with host galaxy characterisation \citep[see][]{vdb05} and spectra with better signal \citep[see][]{ben05}. \section*{Acknowledgments} We are indebted to Louis-G.Strolger for fruitful discussions and providing us with the control times required in the analysis. We also thank Guillaume Blanc, Ranga-Ram Chary, Ben Panther, Alan Heavens and Pablo P\'erez-Gonz\'alez for discussions related to the SFH and Klaus Meisenheimer and an anonymous referee for comments that significantly improved the manuscript. F.F. was supported by a Fundaci\'on Andes -- PPARC Gemini studentship. C.W. was supported by a PPARC Advanced Fellowship. This work was in part supported by a Royal Society UK-China Joint Project Grant (Ph.P and Z.H.), the Chinese National Science Foundation under Grant Nos. 10521001 and 10433030 (Z.H.) and a European Research \& Training Network on Type Ia Supernovae (HPRN-CT-20002-00303).
Title: General stability criterion of inviscid parallel flow
Abstract: A more restrictively general stability criterion of two-dimensional inviscid parallel flow is obtained analytically. First, a sufficient criterion for stability is found as either $-\mu_1<\frac{U''}{U-U_s}<0$ or $0<\frac{U''}{U-U_s}$ in the flow, where $U_s$ is the velocity at inflection point, $\mu_1$ is the eigenvalue of Poincar\'{e}'s problem. Second, this criterion is generalized to barotropic geophysical flows in $\beta$ plane. Based on the criteria, the flows are are divided into different categories of stable flows, which may simplify the further investigations. And the connections between present criteria and Arnol'd's nonlinear criteria are discussed. These results extend the former criteria obtained by Rayleigh, Tollmien and Fj{\o}rtoft and would intrigue future research on the mechanism of hydrodynamic instability.
https://export.arxiv.org/pdf/physics/0601043
\preprint{APS/123-QED} \title{General stability criterion of two-dimensional inviscid parallel flow } \author{Liang Sun} \email{sunl@ustc.edu.cn; sunl@ustc.edu} \affiliation{Dept. of Modern Mechanics, and School of Earth and Space Sciences, \\ University of Science and Technology of China, Hefei, 230026, China.} \date{\today} \pacs{47.20.-k, 47.20.Cq, 47.20.Ft, 47.15.Ki } The stability due to shear in the flow is one of the fundamental and the most attracting problems in many fields, such as fluid dynamics, astrophysical fluid dynamics, oceanography, meteorology et al. The shear instability has been intensively investigated, which is to the greatly helpful understanding of other instability mechanisms in complex shear flows. For the inviscid parallel flow with horizontal velocity profile of $U(y)$, the general way is to investigate the growth of linear disturbances by means of normal mode expansion, which leads to the famous Rayleigh's equation \cite{Rayleigh1880}. Using this equation, Rayleigh \cite{Rayleigh1880} first proved a necessary criterion for instability, i.e., Inflection Point Theorem. Then, Fj{\o}rtoft \cite{Fjortoft1950} found a stronger necessary criterion for instability. These criteria are well known and have been applied to understanding the mechanism of hydrodynamic instability \cite{Drazin1981,Huerre1998,CriminaleBook2003}. Unfortunately, both criteria are only necessary criteria for instability, except for some special cases of the symmetrical or monotone velocity profiles. Tollmien \cite{Tollmien1935} gave a heuristic result that the criteria are also sufficiency for instability in these special cases. The stable criteria also provide a way to categorize the velocity profiles of the flows. According to Rayleigh's criterion, the flows are stable if $U''(y)\neq 0$, where $U''(y)$ denotes $d^2U/dy^2$. And according to Fj{\o}rtoft's criterion, there is another kind of stable flows if $U''(U-U_s)>0$, where $U_s$ is the velocity at the inflection point $U''_s=0$. Then if $U''(U-U_s)<0$, can the flow still be stable? Is there another kind of stable flows besides the above flows? To answer these questions, a more restrictive criterion is needed. And the criterion itself is important for both theoretic researches and real applications. The aim of this letter is to obtain such a stability criterion. and other instabilities may be understood via the investigation here. For this purpose, Rayleigh's equation for an inviscid parallel flow is employed \cite{Rayleigh1880,Drazin1981,Huerre1998,SchmidBook2000,CriminaleBook2003}. For a parallel flow with mean velocity $U(y)$, the streamfunction of the disturbance expands as a series of waves (normal modes) with real wavenumber $k$ and complex frequency $\omega=\omega_r+i\omega_i$, where $\omega_i$ denotes the grow rate of the waves. The flow is unstable if and only if $\omega_i>0$. We study the stability of the disturbances by investigating the growth rate of the waves, this method is known as normal mode method. The amplitude of waves, namely $\phi$, satisfies \begin{equation} (\phi''-k^2 \phi)-\frac{U''}{(U-c)}\phi=0, \label{Eq:stable_parallelflow_RayleighEq} \end{equation} where $c=\omega/k=c_r+ic_i$ is the complex phase speed. The real part of complex phase speed $c_r=\omega_r/k$ is the wave phase speed. In fact, Rayleigh's equation is the vorticity equation of the disturbance \cite{Drazin1981,Huerre1998}. This equation is to be solved subject to homogeneous boundary conditions \begin{equation} \phi=0 \,\, at\,\, y=a,b. \label{Eq:stable_parallelflow_RayleighBc} \end{equation} There are three main categories of boundaries: (i) enclosed channels with both $a$ and $b$ being finite, (ii) boundary layer with either $a$ or $b$ being infinite, and (iii) free shear flows with both $a$ and $b$ being infinite. It is obvious that the criterion for stability is $\omega_i=0$ ($c_i=0$), for that the complex conjugate quantities $\phi^*$ and $c^*$ are also a physical solution of Eq.(\ref{Eq:stable_parallelflow_RayleighEq}) and Eq.(\ref{Eq:stable_parallelflow_RayleighBc}). Multiplying Eq.(\ref{Eq:stable_parallelflow_RayleighEq}) by the complex conjugate $\phi^{*}$ and integrating over the domain $a\leq y \leq b$, we get the following equations \begin{equation} \displaystyle\int_{a}^{b} [(\|\phi'\|^2+k^2\|\phi\|^2)+\frac{U''(U-c_r)}{\|U-c\|^2}\|\phi\|^2]\, dy=0% \label{Eq:stable_parallelflow_Rayleigh_Int_Rea} \end{equation} and \begin{equation} \displaystyle c_i\int_{a}^{b} \frac{U''}{\|U-c\|^2}\|\phi\|^2\,dy=0. \label{Eq:stable_parallelflow_Rayleigh_Int_Img} \end{equation} Rayleigh used only Eq.(\ref{Eq:stable_parallelflow_Rayleigh_Int_Img}) to prove his theorem. Fj\o rtoft noted that Eq.(\ref{Eq:stable_parallelflow_Rayleigh_Int_Rea}) should also be satisfied, then he obtained his necessary criterion. To find a more sufficient criterion, we shall investigate the conditions for $c_i=0$. Unlike the former investigations, we consider this problem in a totally different way: if the velocity profile is stable ($c_i=0$), then the hypothesis $c_i\neq0$ should result in contradictions in some cases. Following this, some more restrictive criteria can be obtained. To find a stronger criterion, we need to estimate the ratio of $\int_{a}^{b} \|\phi'\|^2 dy$ to $\int_{a}^{b} \|\phi\|^2 dy$. This is known as Poincar\'{e}'s problem: \begin{equation} \int_{a}^{b}\|\phi'\|^2 dy=\mu\int_{a}^{b}\|\phi\|^2 dy, \label{Eq:stable_parallelflow_Poincare} \end{equation} where the eigenvalue $\mu$ is positive definition for any $\phi \neq 0$. The smallest eigenvalue value, namely $\mu_1$, can be estimated as $\mu_1>(\frac{\pi}{b-a})^2$, like Tollmien \cite{Tollmien1935} did. Then using Poincar\'{e}'s relation Eq.(\ref{Eq:stable_parallelflow_Poincare}), a new stability criterion may be found: the flow is stable if $-\mu_1<\frac{U''}{U-U_s}<0$ everywhere. To get this criterion, we introduce an auxiliary function $f(y)=\frac{U''}{U-U_s}$, where $f(y)$ is finite at the inflection point. We will prove the criterion by two steps. At first, we prove proposition 1: if the velocity profile is subject to $-\mu_1<f(y)<0$, then $c_r\neq U_s$. \iffalse Proof: Otherwise, $f(y)$ \begin{equation} -\mu_1<\frac{U''}{U-U_s}=\frac{U''(U-U_s)}{(U-U_s)^2}\leq\frac{U''(U-U_s)}{(U-U_s)^2+c_i^2}, \end{equation} \fi Proof: Since $-\mu_1<f(y)<0$, then \begin{equation} -\mu_1<\frac{U''}{U-U_s}=\frac{U''(U-U_s)}{(U-U_s)^2}\leq\frac{U''(U-U_s)}{(U-U_s)^2+c_i^2}. \label{Eq:stable_parallelflow_Rayleigh_inequ} \end{equation} Substitution of $c_r=U_s$ and Eq.(\ref{Eq:stable_parallelflow_Rayleigh_inequ}) into Eq.(\ref{Eq:stable_parallelflow_Rayleigh_Int_Rea}) results in \begin{equation} \displaystyle\int_a^b [\|\phi'\|^2+k^2\|\phi\|^2+\frac{U''(U-U_s)}{\|U-c\|^2}\|\phi\|^2]\, dy > 0. \end{equation} This contradicts Eq.(\ref{Eq:stable_parallelflow_Rayleigh_Int_Rea}). So proposition 1 is proved. Then, we prove proposition 2: if $-\mu_1<f(y)<0$ and $c_r\neq U_s$, there must be $c_i^2=0$. Proof: If $c_i^2\neq0$, then multiplying Eq.(\ref{Eq:stable_parallelflow_Rayleigh_Int_Img}) by $(c_r-U_t)/c_i$, where the arbitrary real constant $U_t$ does not depend on $y$, and adding the result to Eq.(\ref{Eq:stable_parallelflow_Rayleigh_Int_Rea}), it satisfies \begin{equation} \displaystyle\int_a^b [(\|\phi'\|^2+k^2\|\phi\|^2)+\frac{U''(U-U_t)}{\|U-c\|^2}\|\phi\|^2]\, dy=0. \label{Eq:stable_parallelflow_Sun_Int} \end{equation} But the above Eq.(\ref{Eq:stable_parallelflow_Sun_Int}) can not hold for some special $U_t$. For example, let $U_t=2c_r-U_s$, then there is $(U-U_s)(U-U_t)<\|U-c\|^2$, and \begin{equation} \frac{U''(U-U_t)}{\|U-c\|^2}= f(y)\frac{(U-U_s)(U-U_t)}{\|U-c\|^2}>-\mu_1. \label{Eq:stable_parallelflow_Sun_Ust} \end{equation} This yields \begin{equation} \int_a^b \{\|\phi'\|^2+[k^2+\frac{U''(U-U_t)}{\|U-c\|^2}]\|\phi\|^2\} dy>0, \end{equation} which also contradicts Eq.(\ref{Eq:stable_parallelflow_Sun_Int}). So proposition 2 is also proved. Using 'proposition 1: if $-\mu_1<f(y)<0$ then $c_r\neq U_s$' and 'proposition 2: if $-\mu_1<f(y)<0$ and $c_r\neq U_s$ then $c_i = 0$', we find a stability criterion. If the velocity profile satisfies $-\mu_1<\frac{U''}{U-U_s}<0$ everywhere in the flow, it is stable. Moreover, the above proof is still valid for $0<f(y)$, which is equivalent to Fj\o rtoft's criterion. Thus we have the following theorem. Theorem 1: If the velocity profile satisfies either $-\mu_1<\frac{U''}{U-U_s}<0$ or $0<\frac{U''}{U-U_s}$, the flow is stable. This criterion is more restrictive than Fj\o rtoft's criterion. As known from Fj\o rtoft's criterion, the necessary condition for instability is that the base vorticity $\xi=-U'$ has a local maximal in the profile. Noting that $U''/(U-U_s)\approx \xi_s''/\xi_s$ near the inflection point, where $\xi_s$ is the vorticity at inflection point, it means that the base vorticity $\xi$ must be convex enough near the local maximum for instability, i.e., the vorticity should be concentrated somewhere in the flow for instability. A simple example can be given by following Tollmien's way \cite{Tollmien1935}. As shown in Fig.\ref{Fig:vorticity_profile}, there are three vorticity profiles within the interval $-1\leq y\leq 1$, which have local maximal at $y=0$. Profile 2 ($U=-2\sin(\pi y/2)/\pi$) is neutrally stable, while profile 1 ($U=-\sin(y)$) and profile 3 ($U=-\sin(2y)/2$) are stable and unstable, respectively. Moreover, the stabile criterion for the parallel inviscid flows can be applied to the barotropic geophysical flows in $\beta$ plane, like Kuo did \cite{KuoHL1949}. This is a generalized stable criterion, we state it as a new theorem. Theorem 2: The flow is stable, if the velocity profile satisfies either $-\mu_1<\frac{U''-\beta}{U-U_s}<0$ or $0<\frac{U''-\beta}{U-U_s}$ in the flow, where $U_s$ is the velocity at the point $U''=\beta$. The criteria proved above may shed light on the investigation of vortex dynamics. Both Theorem 1 and Fig.\ref{Fig:vorticity_profile} show that it is the vorticity profile rather than the velocity profile that dominates the stability of the flow. This means that the distribution of vorticity dominates the shear instability in parallel inviscid flow, which is essential to understanding the role of vorticity in fluid. So an unstable flow might be controlled just by adjusting the vorticity distribution according to present results. This is an very fascinating problem, but can not be discussed in detail here. To show the power of the criteria obtained above, we consider the stability of velocity profile $U=\tanh(\alpha y)$ within the interval $-1\leq y\leq 1$, where $\alpha$ is a constant. This velocity profile is an classical model of mixing layer, and has been investigated by many researchers (see \cite{Huerre1998,SchmidBook2000,CriminaleBook2003} and references therein). Since $U''(U-U_s)=-2\alpha^2\tanh^2(\alpha y)/\cosh^2(\alpha y) <0$ for $-1\leq y\leq 1$, it might be unstable for any $\alpha$ according to both Rayleigh's and Fj\o rtoft's criteria. But it can be derived from Theorem 1 that the flow is stable for $\alpha^2<\pi^2/8 $. For example, we choose $\alpha_1=1.1$ and $\alpha_2=1.3$ for velocity profiles $U_1(y)$ and $U_2(y)$. The growth rate of the profiles can be obtained by Chebyshev spectral collocation method \cite{SchmidBook2000} with 100 collocation points, as shown in Fig.\ref{Fig:Growth}. It is obvious that $c_i=0$ for $U_1$ and $c_i>0$ for $U_2$, which agrees well with the criteria obtained above. This is also a counterexample that Fj\o rtoft's criterion is not sufficient for instability. So this new criterion for stability is more useful in real applications. The present stable criteria give a affirmative answer to the questions at the beginning, i.e., there are some stable flows if $U''(U-U_s)<0$. Based on the former criteria, the velocity profiles can be categorized as follows: (\romannumeral1) without inflection point (Reyleigh's criterion), (\romannumeral2) $f(y)>0$ (Fj{\o}rtoft's criterion), and (\romannumeral3) $\mu_1<f(y)<0$ (present criterion). Then the flow might be unstable only for $f(y)<\mu_1$ and $f(y)$ changing sign within the interval. However, if $f(y)$ changes sign somewhere within the interval $[a,b]$, then the flow is stable. For that $f(y)$ changing sign implies $U'''_s=0$ but $U''''_s\neq0$, so $U''$ does not change sign near the inflection point. Thus $c_i$ must vanish in Eq.(\ref{Eq:stable_parallelflow_Rayleigh_Int_Img}), i.e., the flow is stable for $f(y)$ changing sign within the interval. In this way, the flow might be unstable only for $f(y)<\mu_1$ somewhere, which will intrigue further studies on this problem. In fact, there are still stable flows if $\mu_1<f(y)$ is violated. Recall the proof of theorem 1, it is found that the following Rayleigh's quotient $I(f)$ plays a key role in determination the stability of the flows. \begin{equation} I(f)=\min_{\phi} \frac{\int_{a}^{b} [\,\|\phi'\|^2+f(y)\|\phi\|^2\,]\, dy}{\int_{a}^{b} \|\phi\|^2} \label{Eq:stable_paralleflow_sun_Energy} \end{equation} Noting that the proof of theorem 1 is still valid in the case of $I(f)>0$. We have such result: the flows are stable if $I(f)>0$. Though this criterion is more restrictive than that in theorem 1, it is inconvenient for the real applications due to unknown value of Rayleigh's quotient $I(f)$. Theorem 1 is more convenient for the real applications in different research fields. The idea of categorization the velocity profiles of the flows may simplify the investigation of stability problem. It can be seen from Rayleigh's equation Eq.(\ref{Eq:stable_parallelflow_RayleighEq}) that the stability of profile $U(y)$ is not only Galilean invariant, but also independent from the the magnitude of $U(y)$ due to linearity. So the stability of $U(y)$ is the same as that of $AU(y)+B$, where $A$ and $B$ are any arbitrary nonzero real numbers. As the value of $U''(U-U_s)$ in Fj\o rtoft's criterion is only Galilean invariant but not magnitude free, it satisfies only part of the Rayleigh's equation's properties. On the other hand the value of $U''/(U-U_s)$ satisfies both conditions, this is the reason why the criteria in both Arnol'd's theorems and present theorems are the functions of $U''/(U-U_s)$. Since the stability of inviscid parallel flow depends only on the velocity profile's geometry shape, namely $f(y)$, and the magnitude of the velocity profile can be free, then the instability of inviscid parallel flow could be called "geometry shape instability" of the velocity profile. As the above investigation shows that the inviscid shear instability is only associated with the geometry of velocity profile. The concept of "geometry shape instability" would be help in further investigations. This distinguishes from the viscous instability, which is also associated with the magnitude of the velocity profile. As mentioned above, we have investigated the stability of the flows via Rayleigh's equation, while Arnol'd \cite{Arnold1969} considered the hydrodynamic stability in a totally different way. He studied the conservation law of the inviscid flow via Euler's equations and found two nonlinear stability conditions by means of variational principle. So what is the relationship between the linear criteria and the nonlinear ones? It is very interesting that the linear stability criteria match Arnol'd's nonlinear stability theorems very well. Applying Arnol'd's First Stability Theorem to parallel flow, the stable criterion is $0<C_1<(U-U_s)/U''<C_2<\infty$ everywhere in the flow, where $C_1$ and $C_2$ are constants. This corresponds to Fj\o rtoft's criterion for linear stability, and is well known \cite{Drazin1981,Dowling1995}. Here we find that Theorem 1 proved above corresponds to Arnol'd's Second Stability Theorem, i.e., the stable criterion is $0<C_1<-(U-U_s)/U''<C_2<\infty$ everywhere in the flow. Given $C_1=1/\mu_1$, Arnol'd's Second Stability Theorem is equivalent to Theorem 1. Moreover, the proofs here are similar to Arnol'd's variational principle method. For the arbitrary real number $U_t$, which is like a Lagrange multiplier in variational principle method, plays a key role in the proofs. So that the above Theorem 1 is similar to Arnol'd's theorems. Unfortunately, Arnol'd's nonlinear stability theorems, though quite useful in the geophysical flows \cite{Dowling1995}, are seldom known by the scientists in other fields. The main reason is that the proofs of Arnol'd's theorems are very advanced and complex in mathematics for most general scientists in different fields to understand. Although Dowling \cite{Dowling1995} suggested that Arnol'd's idea need to be added to the general fluid-dynamics curriculum, his suggestion has not been followed even 10 years later. Compare with Arnol'd's theorems, the theorems proved here are equivalent in some sense but much simpler and easier to understand, therefore it is more convenient to use our new results in applications. In summary, the general stability criteria are obtained for inviscid parallel flow. These results, which are equivalent to Arnol'd's nonlinear theorems, extend the former theorems proved by Rayleigh, Tollmien and Fj\o rtoft. Based on the criteria, the velocity profiles are divided into different categories, which may simplify the further investigations. In general, these criteria would intrigue future research on the mechanism of hydrodynamic instability and to understand the mechanism of turbulence. And it also sheds light on the flow control and investigation of the vortex dynamics. The author thanks Prof. Sun D-J at USTC, Dr. Yue P-T at UBC (Canada) and two anonymous referees for their useful comments. This work was original from author's dream of understanding the mechanism of instability in the year 2000, when the author was a graduated student and learned the course of hydrodynamic stability by Prof. Yin X-Y at USTC. \iffalse \bibliography{MSH1} \fi
Title: Stellar Multiplicity and the IMF: Most Stars Are Single
Abstract: In this short communication I compare recent findings suggesting a low binary star fraction for late type stars with knowledge concerning the forms of the stellar initial and present day mass functions for masses down to the hydrogen burning limit. This comparison indicates that most stellar systems formed in the galaxy are likely single and not binary as has been often asserted. Indeed, in the current epoch two-thirds of all main sequence stellar systems in the Galactic disk are composed of single stars. Some implications of this realization for understanding the star and planet formation process are briefly mentioned.
https://export.arxiv.org/pdf/astro-ph/0601375
. \begin{document} \title{Stellar Multiplicity and the IMF: Most Stars Are Single} \author{Charles J. Lada\altaffilmark{1}} \altaffiltext{1}{Harvard-Smithsonian Center for Astrophysics, 60 Garden Street, Cambridge, MA 02138, USA; clada@cfa.harvard.edu} \keywords{stars: binary, formation} \section{Introduction} \label{sec:introduction} Ever since Mitchell (1767) pointed out that the observed frequency of visual double stars was too high to be due to random chance, the study of binary stars has occupied an important place in astrophysics. William Herschel (1802) discovered and cataloged hundreds of visual pairs and produced the first observations of a rudimentary binary orbit. In doing so he established that the double stars were indeed physical pairs and that Newtonian physics operated nicely in the distant sidereal universe. By the beginning of the twentieth century tens of thousands of binary stars were known and cataloged (e.g., Burnham 1906). By the middle to late twentieth century the first systematic attempts to establish the binary frequency of main sequence F and G stars suggested that a very high fraction (70 - 80\%) of all such stellar systems consist of binary or multiple stars (Heintz 1969; Abt \& Levy 1976; Abt 1983). The most comprehensive and complete study of the multiplicity of G stars was performed by Duquennoy \& Mayor (1991) who argued that two-thirds of all such stellar systems are multiple. It has often been assumed but never clearly demonstrated that similar statistics applied to stars of all spectral types. This assumption has led to the commonly held opinion that most all stars form in binary or multiple systems with the Sun (and its system of planets) being atypical as a single star. But how robust is the assumption that the binary statistics for G stars is representative of all stars? Over the last decade two important developments have occurred in stellar research which directly bear on this question. First, the functional form of the stellar initial mass function (IMF) has been better constrained by observations of both field stars (e.g., Kroupa, 2002) and young embedded clusters (e.g., Muench et al. 2002). The IMF has been found to peak broadly between 0.1 - 0.5 \msun, indicating that most stars formed in the Galactic disk are M stars. Second, surveys for binary stars have suggested that the binary star frequency may be a function of spectral type (e.g., Fischer \& Marcy 1992). In particular, there have been a number of attempts to ascertain the binary frequency of M type stars and even for L and T dwarfs, objectss near and below the hydrogen burning limit. These studies suggest that the binary frequency declines from the G star value, being only around 30\% for M stars (e.g., Leinert et al. 1997; Reid \& Gizis 1997; Delfosse et al. 2004; Siegler et al. 2005) and as much as a factor of 2 lower for L and T dwarfs (e.g., Gizis et al. 2003). I argue in this communication that these two facts together suggest that most stellar systems in the Galaxy consist of single rather than binary or multiple stars. \section{The Single Star Fraction and Spectral Type} \label{sec:observations} In this section I use data compiled from the literature to examine the single star fraction as a function of stellar spectral type, in particular for the range spanning G to M stars. I consider the single star fraction (SSF) to be the fraction of stellar systems without a {\it stellar} companion, that is, primary stars without a companion whose mass exceeds 0.08 \msun. Figure 1 displays the single star fraction as a function of spectral type for G and later type stars. This plot suggests that the SSF is significantly greater for M stars than for G stars. Indeed the SSF for M stars appears to be at least 70\%. It is difficult to evaluate the significance of this difference at face value given that the differing binary surveys suffer from differing biases and varying degrees of incompleteness. The systematic differences that can arise between the surveys mostly derive from varying sensitivities to primary/secondary separations and mass ratios. Below I attempt to evaluate the results from the surveys used to construct Figure 1. In their seminal study, Duquennoy \& Mayor (1991) obtained a spectroscopic survey of a distance-limited complete sample of F7-G9 stars in the Northern Hemisphere and within 22 pc of the Sun. They examined radial velocities obtained for these stars over a 13 year period. They combined their detections of spectroscopic binaries with known visual binaries and common proper motion pairs to examine 164 primaries for evidence of multiplicity. They derive multiplicity ratios of 57:38:4:1 for single:double:triple:quadruple systems, respectively. They considered all the various detection biases to estimate the incompleteness of their study and concluded that there was a slight bias against detecting low mass companions, this resulted in a 14\% upward correction to the multiplicity fraction such that 57\% of systems were estimated to be multiple for a primary/companion mass ratio, q $>$ 0.1. They further extrapolated this incompleteness correction to include substellar secondaries and estimated a multiplicity fraction of 2/3 and a single star fraction of 1/3 for their sample. However, in recent years sensitive and precise radial velocity surveys of 1330 single FGKM stars have indicated a paucity of substellar companions within 5 AU of the primary stars (Marcy \& Butler 2000; Marcy et al. 2005). In addition coronographic imaging surveys have found a similar dearth of substellar companions around GK and M stars over separations between 75 and 300 AU (McCarthy \& Zuckerman 2004). The existence of this so-called ``brown dwarf desert'' indicates that Duquennoy \& Mayor may have overestimated the multiplicity fraction of G stars and the true value is likely 57\% or even somewhat smaller. For the purposes of this paper I adopt 57\% as the multiplicity fraction of G type stars and thus 43\% for the SSF. The first extensive examination of the multiplicity of M stars was performed by Fischer \& Marcy (1992) who studied radial velocity, speckle and visual binary data for a sample of stars within 20 pc. The full range of separations, $a$ $<$ 10$^4$ AU, was examined, similar to the G star study. These authors pointed out that M star surveys suffer less from the effects of incompleteness than G star surveys because the M star sample is on the whole a factor of 2 closer in distance and M star primaries are sufficiently faint to enable detection of very faint companions more readily. They derived a SSF of 58\% which is higher than the G star value. Reid \& Gizis (1997) determined the SSF for a volume complete sample of 79 M2-M4.5 primary stars within 8 pc of the Sun and derived a SSF of 70 $\pm$ 12\% for this sample. The range of binary separations they were able to probe was 0.1 - 10$^4$ AU. A similar volume complete search for M dwarf binaries within 5 pc of the Sun was performed by Leinert et al. (1997) who reported a SSF of 74 $\pm$ 19\%. However, their sample of 29 stars is smaller than the Reid \& Gizis (1997) and Fischer \& Marcy (1997) samples accounting for the larger uncertainty. More recently Delfosse et al. (2004) presented statistics for a much larger sample of 100 M dwarfs which they estimated was 100\% complete for stellar mass companions over the entire separation range and out to 9 pc from the Sun. Delfosse et al. (2004) derive a multiple star fraction of 26 $\pm$ 3 \% which corresponds to a SSF of 74 $\pm$ 6\%. This may represent the most accurate determination for the M star SSF yet made. I note here that even if one considers substellar companions this estimate for the SSF will not likely alter significantly since as mentioned earlier, surveys have revealed a dearth of substellar companions to G, K {\it and} M stars (Marcy \& Butler 2000; McCarthy and Zuckerman 2004). Surveys for multiplicity among very late M stars and even L and T dwarfs have also been recently reported. These studies typically explore more limited separation ranges and somewhat smaller samples of stars. The multiplicity fractions they find are however all lower than that reported for the earlier type M stars. For example, Siegler et al. (2005) examined a magnitude-limited survey of 36 M6 - M 7.5 stars and derived a binary fraction of 9 $\pm$ 4\% corresponding to a SSF of 91 $\pm$ 5\%. However this sample is not volume limited and may be incomplete. Thus the inferred SSF is likely an upper limit. Despite this limitation Siegler et al. were able to conclude that wide (a$>$20 AU) binaries are very rare among these stars. Although not considered for inclusion in Figure 1 because of the large fraction of brown dwarfs in their samples, surveys by Gizis et al. (2003) and Bouy et al. (2003) find similarly small binary fractions for ultra low mass objects. For example, Gizis et al. examined 82 nearby late M and L dwarfs and derived a (incompleteness corrected) binary fraction of 15 $\pm$ 5\% (corresponding to a SSF of 85 $\pm$ 14\%) for separations, $a > $ 1.6 AU. Estimating the possible contribution of companions at smaller separations they suggest a binary star fraction (BSF) of 15 $\leq$ BSF $\leq$ 25 \% corresponding to 75 $\leq$ SSF $\leq$ 85 \% for these objects near and just below the hydrogen burning limit. Bouy et al. (2003) examined the binary statistics for a sample of 134 late M and L field dwarfs and estimated a binary fraction for a separation range of about 2 - 140 AU of only 10\% corresponding to a SSF of 90\% for these objects. They also noted a dearth of companions with wide (i.e., $a >$ 15 AU) separations. Although these surveys of very low mass and substellar objects suffer from some degree of incompleteness it is quite unlikely that sensible corrections for such effects would decrease the estimated single star fraction to a value similar to that of G stars or even typical M stars. The observations discussed above lead to the conclusion that the single star fraction is a function of spectral type and increases from about 43\% for G stars to $\sim$ 85\% for brown dwarfs. The most secure estimate for M stars appears to be about 74\% based on the complete volume-limited sample of Delfosse et al. (2004) for M stars with stellar companions. \section{M Stars and the IMF} \label{sec:results} The stellar IMF is one of the most fundamental distribution functions in astrophysics. A great deal of effort has been expended in determining its form since the first attempt to measure its shape by Salpeter (1954). He found that the IMF is a power-law which decreases with stellar mass for field stars with masses in the range between 1-10 \msun. More recent determinations of the IMF for field stars and young embedded clusters have expanded the mass range covered by Salpeter. These studies have found the IMF to break from a single power-law shape near 0.5 \msun\ and to have a broad peak between $\sim$ 0.1 - 0.5 \msun. On either side of this peak the IMF falls off rapidly (e.g., Miller \& Scalo 1979; Kroupa 2002; Muench et al. 2002; Chabrier 2003; Luhman et al. 2006). The broad peak of the IMF encompasses the M stars and indicates that these stars are the most numerous objects created in the star formation process. This is illustrated in Figure 2 which shows the cumulative fraction of all stars above the hydrogen burning limit given by the IMF. Two different IMFs are plotted which span the range of modern day determinations of this function. One is the log-normal field star IMF derived by Miller \& Scalo (1979) and the other represents a determination of the IMF for the embedded Trapezium cluster in Orion in which the IMF is characterized by a series of broken power-laws (Muench et al. 2002). This latter IMF is very similar to that determined for the field by Kroupa (2002) but is more sensitive to substellar masses (not plotted). The vertical dashed line shows the boundary for the M star population. The fraction of all stars {\it above the hydrogen burning limit (HBL)} that are M stars is 73\% for the Muench et al. IMF and 78\% for the Miller-Scalo IMF. (It is important to note here that these two IMFs are essentially primary star IMFs, that is, IMFs that do not include companion star masses.) This analysis indicates that roughly 3/4 of all stars formed are M stars. The IMF represents the frequency distribution of stars at birth and differs from the present day mass function (PDMF) which represents the frequency distribution of all stars currently living within the Galactic disk. Stellar evolution has significantly depleted the high mass end of the PDMF relative to the IMF. Therefore, the fraction M stars in the PDMF is somewhat higher than the fraction in the IMF. Indeed, for the PDMF derived by Miller \& Scalo (1979) we find from Figure 2 that 84\% of all stars in the Galactic disk are M stars. \vskip 0.2in \section{The Total Single Star Fraction} \label{sec:SSF} To estimate the total fraction of single stars, I assume that all stars earlier than M are characterized by the single star fraction for G stars determined by Duquennoy \& Mayor (1991), that is, $SSF_{<M} =$ 43\%. The single star fraction for M-type stars (i.e., $SSF_{M}$) is assumed to be that (74\%) determined by Delfosse et al. (2004) for a complete, volume limited sample. The total SSF is then simply given by: \small $$ {\rm SSF(total)} = SSF_{<M} \times ETF + SSF_{M} \times MTF $$ \normalsize \noindent Here $MTF$ is the M-type fraction, that is, the fraction of all stars that are M-type stars and $ETF = 1 - MTF$ is the early-type fraction, that is the fraction of all stars that have spectral types earlier than M. To determine the SSF for all stars produced at any one time by the star formation process I adopt the Muench et al. and Miller-Scalo IMFs, specifically, MTF = 0.73 and 0.78, respectively. The total SSF is found to be 66\% and 67\% for these two IMFs, respectively. Therefore, single stars must ultimately account for as many as two-thirds of all stellar systems that formed at any one time in the Galaxy. Similarly, if we consider the MTF (0.84) for the Miller-Scalo PDMF we find the total SSF to be 69\%. Thus, {\it two thirds of all (main sequence) primary stars currently residing in the Galactic disk are single stars}. \section{Discussion and Conclusions} \label{sec:discussion} The primary result of this paper is the recognition that most stellar systems in the Galaxy consist of single rather than binary stars. This fact has important consequences for star and planet formation theory. For example, contrary to the current accepted paradigm that most, if not all, stars form in binary or multiple systems (e.g., Larson 1972, 2001; Mathieu 1994), this result could indicate that the theoretical frameworks developed to explain the formation of single, sunlike stars (e.g., Shu, Adams \& Lizano 1987) have wide applicability. Indeed, when appropriately modified for a cluster-forming environment (e.g., Myers 1998; Shu, Li \& Allen 2004), they may even describe most star forming events in the Galaxy. On the other hand, most stars could still initially form in binary or multiple systems provided that most such systems promptly disintegrate via dynamical interactions or decay in an early, perhaps even protostellar, stage of evolution (e.g., Kroupa 1995; Sterzik \& Durisen 1998, Reipurth 2000). The current paradigm that most, if not all stars, form in binaries was strengthened by early multiplicity surveys of pre-main sequence (PMS) stars. In particular, surveys of the PMS population of the Taurus cloud indicated a binary fraction that was twice that of field G stars (Ghez et al. 1993; Leinert et al. 1993; Reipurth \& Zinnecker 1993). However, most field stars are now known to have formed in embedded clusters, environments quite different than represented by the Taurus PMS population (e.g., Lada \& Lada 2003). Binary surveys of both young embedded and Galactic clusters have revealed binary fractions indistinguishable from that of the field (e.g., Petr et al. 1998; Duch\^ene, Bouvier \& Simon 1999; Patience \& Duch\^ene 2001). The most simple and straightforward hypothesis to explain these two facts and the finding of a high SSF in this paper is that the most common outcome of the star formation process is a single rather than multiple star. Observations of dust emission and extinction of molecular cloud cores have found that the shape of the primordial or dense core mass function is very similar to that of the stellar IMF except that the core mass function is offset to higher mass by a factor of 2-3 (e.g., Stanke et al. 2005, Alves, Lombardi \& Lada 2005). These observations indicate that a 1-to-1 mapping of core mass to stellar mass, modified by a more or less constant star formation efficiency of 30-50\%, is possible, if not likely. This idea is consistent with single star systems being most often produced once the cores undergo collapse. The fact that stellar multiplicity is a function of stellar mass, however, may provide important clues to the nature of the physical process of star formation. For example, Durisen, Sterzik \& Pickett (2001) have shown that if individual protostellar cores can further fragment and produce small N clusters, the dynamical decay of these clusters into binary and single stars can in certain circumstances produce a binary star fraction that declines with decreasing primary mass, similar to what is observed. However, to be consistent with the SSF derived here and to simultaneously produce reasonable binary component separations, such models would require N $\geq$ 5, within a region $\sim$ 300 AU in size (Sterzik \& Durisen 1998). This would correspond to a stellar surface density ($\sim$ 7.5 $\times$ $10^5$ stars pc$^{-2}$) about two orders of magnitude higher than the peak density (7.2 $\times$ 10$^3$ stars pc$^{-2}$) measured for the rich Trapezium cluster (Lada et al. 2004). Such ultra-dense protostellar groups have not yet been identified, but could be revealed with high resolution infrared imaging surveys of deeply embedded candidates. A related possibility, proposed by Kroupa (1995) and collaborators, posits that all stars are formed in binaries in modestly dense embedded clusters. Dynamical interactions between these systems can disrupt some binaries and modify the separations of others. These models can produce the observed dependance of binary frequency with mass, but at the expense of a SSF (50\%) that is too low to be consistent with that derived here. These models could be made consistent with the high Galactic SSF by assuming more compact configurations for the birth clusters, however it is unclear whether the required higher cluster densities would remain consistent with observed values. Another possibility is that binary star formation is related to the initial angular momentum content of the primordial cores. In this case the initial angular momentum of a protostellar core would be expected to be a function of core mass, with low mass cores being endowed with considerably less angular momentum than high mass cores. A systematic molecular-line survey of cores of varying mass within a molecular cloud could test this idea. A related possibility is that turbulence may play a role in the propensity for a core to fragment. For example, Shu, Li \& Allen (2004) posit that the break in the stellar IMF at 0.5 \msun\ is a result of the transition from turbulent to thermal support of the envelopes of dense pre-collapse cloud cores. The more massive the core, the more turbulence is required to insure its support. Ammonia observations of dense cores in fact do suggest that massive cores are more turbulent than low mass cores (Jijina, Myers \& Adams 1999). Perhaps increased cloud turbulence in the more massive dense cores can also promote, in some fashion, more efficient core fragmentation and a higher incidence of binary star formation. In this context it would be interesting to know if the trend of increasing stellar multiplicity with stellar mass continues to the more massive A, B and O stars, as has been suggested in some studies (e.g., Preibisch, Weigelt, \& Zinnecker 2001, Shatsky \& Tokovinin 2002). Finally I note that the large fraction of single star systems in the field is consistent with the idea that most stars could harbor planetary systems unperturbed by binary companions and thus extra-solar planetary systems that are characterized by architectures and stabilities similar to that of the solar system could be quite common around M stars, provided planetary systems can form around M stars in the first place. \vskip -0.1in \acknowledgments I am indebted to August Muench for constructing the cumulative IMFs presented in Figure 2 and many useful discussions. I thank David Latham and Bo Reipurth for their careful reading of the paper and detailed suggestions and Kevin Luhman, Geoff Marcy, Frank Shu and Pavel Kroupa for useful comments which improved the paper.
Title: A Chandra View of Dark Matter in Early-Type Galaxies
Abstract: We present a Chandra study of mass profiles in 7 elliptical galaxies, of which 3 have galaxy-scale and 4 group-scale halos demarcated at 1E13Msun. These represent the best available data for nearby objects with comparable X-ray luminosities. We measure ~flat mass-to-light (M/L) profiles within an optical half-light radius (Reff), rising by an order of magnitude at ~10Reff, which confirms the presence of dark matter (DM). The data indicate hydrostatic equilibrium, which is also supported by agreement with studies of stellar kinematics in elliptical galaxies. The data are well-fitted by a model comprising an NFW DM profile and a baryonic component following the optical light. The distribution of DM halo concentration parameters (c) versus Mvir agrees with LCDM predictions and our observations of bright groups. Concentrations are slightly higher than expected, which is most likely a selection effect. Omitting the stellar mass drastically increases c, possibly explaining large concentrations found by some past observers. The stellar M/LK agree with population synthesis models, assuming a Kroupa IMF. Allowing adiabatic compression (AC) of the DM halo by baryons made M/L more discrepant, casting some doubt on AC. Our best-fitting models imply total baryon fractions \~0.04--0.09, consistent with models of galaxy formation incorporating strong feedback. The groups exhibit positive temperature gradients, consistent with the "Universal" profiles found in other groups and clusters, whereas the galaxies have negative gradients, suggesting a change in the evolutionary history of the systems around Mvir=1E13 Msun.
https://export.arxiv.org/pdf/astro-ph/0601301
\title{A Chandra View of Dark Matter in Early-Type Galaxies} \author {\href{mailto:phumphre@uci.edu}{Philip J. Humphrey}\altaffilmark{1}, David A. Buote\altaffilmark{1}, Fabio Gastaldello\altaffilmark{1}, Luca Zappacosta\altaffilmark{1}, James S. Bullock\altaffilmark{1}, Fabrizio Brighenti\altaffilmark{2,3} and William G.~Mathews\altaffilmark{3}} \altaffiltext{1}{Department of Physics and Astronomy, University of California at Irvine, 4129 Frederick Reines Hall, Irvine, CA 92697-4575} \altaffiltext{2}{Dipartimento di Astronomia, Universit\`{a} di Bologna, Via Ranzani 1, Bologna 40127, Italy} \altaffiltext{3}{University of California Observatories, Lick Observatory, University of California at Santa Cruz, Santa Cruz, CA 95064} \keywords{Xrays: galaxies--- galaxies: elliptical and lenticular, cD--- galaxies: halos--- galaxies: ISM--- dark matter} \section{Introduction} The nature and distribution of dark matter (DM) in the Universe is one of the fundamental problems facing modern physics. Cold DM lies at the heart of our current (\lcdm) cosmological paradigm, which predicts substantial DM halos for objects at all mass-scales from galaxies to clusters. Although \lcdm\ has been remarkably successful at explaining large-scale features \citep[\eg][]{spergel03a,perlmutter99a}, observations of galaxies have been more problematical for the theory. Dissipationless dark matter simulations find that dark matter halos are well characterized by a ``Universal'' mass density profile \citep[][hereafter NFW]{navarro97} over a wide range of Virial masses (\mvir) \citep[e.g.][]{bullock01a}. Low mass halos tend to form first in hierarchical cosmologies and are consequently more tightly concentrated than their later forming, high mass counterparts. This tendency produces a predicted correlation between the DM halo concentration parameter (c, which is ratio between Virial radius, \rvir, and the characteristic scale of the density profile) and \mvir \citep{navarro97}. However, since mass and formation epoch are not perfectly correlated, we expect a significant scatter at fixed Virial mass \citep{jing00a,bullock01a,wechsler02a}. The tight link between halo formation epoch and concentration implies that the precise relation between c and \mvir\ is sensitive to the underlying Cosmological parameters, including \sigmaeight\ and the dark energy equation of state \citep{kuhlen05a}, making an observational test of this relation potentially a very powerful tool for cosmology. The mass profiles of galaxies also may provide valuable clues as to the way in which galaxies form in DM halos. In particular, as baryons cool and collapse into stars, the associated increase in the central mass density should in turn modify the shape of the DM halo. This process is typically modelled assuming adiabatic contraction (AC) of the DM particle orbits \citep[\eg][]{blumenthal86a,gnedin04a}. If the galaxy halo subsequently evolves by major mergers, simulations are unclear as to whether these features would persist \citep[\eg][]{gnedin04a} or whether the merging process may destroy this imprint of star formation, or even mix the DM and baryons sufficiently to produce a {\em total} gravitating mass profile more akin to NFW \citep{loeb03a,elzant04a}. Observational tests of the predictions of \lcdm\ have proven controversial. In clusters of galaxies there is overwhelming evidence for DM, and an increasing body of work verifying the predictions of \lcdm. In particular recent, high-quality \chandra\ and \xmm\ observations have revealed mass profiles in remarkable agreement with the Universal profile from deep in the core to a large fraction of \rvir\ \citep[\eg][]{lewis03a,zappacosta06a,vikhlinin05b}, and a distribution of c {\em versus} \mvir\ in good agreement with \lcdm\ \citep{pointecouteau05a}. In galaxies, however, the picture is much less clear. Rotation curve analysis of low surface brightness (LSB) disk galaxies has suggested significantly less cuspy density profiles than expected \citep[\eg][]{swaters00a}. Although this discrepancy led to a serious discussion of modifications to the standard paradigm \citep[\eg][]{hogan00a,spergel00a,zentner02a,kaplinghat05a,cembranos05a}, recent results, taking account of observational bias and the 3-dimensional geometry of the DM halos, have done much to resolve the discrepancy \citep[\eg][]{swaters03a,simon05a}. However, some significant discrepancies remain, not least of which is that the DM halos of these galaxies appear less concentrated than expected \citep[\eg][]{gonzalez00a,kassin06a}. A possible explanation is that LSB galaxies are preferentially found in low-concentration halos \citep{bullock01a,bailin05a,wechsler05a}, making additional constraints at the galaxy scale extremely important. % In many respects, kinematical mass measurements are far more challenging for early-type than spiral galaxies. As essentially pressure-supported systems little is known {\em a priori} about the velocity anisotropy tensor of the stars in elliptical galaxies, which is problematical for the determination of the mass from stellar motions. Nonetheless, stellar kinematical measurements have widely been used as a means to measure the gravitating matter within $\sim$the optical half-light radius (\reff) of elliptical galaxies \citep[\eg][]{binney90a,vandermarel91a,gerhard01a}. These studies tend to find relatively flat mass-to-light (M/L) ratios within \reff, implying that most of the matter within this radius is baryonic. Consideration of the tilt in the fundamental plane can also lead to the same conclusion \citep{borriello03a}. In contrast, \citet{padmanabhan04a} pointed out that dynamical M/L ratios within \reff\ are much larger than predicted by realistic stellar population synthesis models for stars alone, allowing \gtsim 50\% of the mass within \reff\ to be dark. % Attempts to extend kinematical studies of elliptical galaxies to larger radii, where DM should be dominant, have proven controversial. In particular \citet{romanowsky03a} argued against the existence of DM in a small sample of elliptical galaxies, based on planetary nebulae dynamics within $\sim$5\reff. We note that this sample was heavily biased towards very X-ray faint objects, which might hint at low-mass halos since they have not held onto their hot gas. In any case \citet{dekel05a} pointed out that their conclusions were very sensitive to the uncertainty in the velocity anisotropy tensor, for plausible values of which the data were consistent with substantial DM halos. In fact globular cluster dynamics in one of these systems, NGC\thin 3379, does imply a significant amount of DM \citep{pierce06a,bergond06a}. As more kinematical studies of early-type galaxies at large radii are appearing, it is becoming clear that at least some elliptical galaxies host considerable DM halos \citep[\eg][]{statler99a,romanowsky05a}. There persist some questions, however, as to the extent to which all galaxies have DM halos consistent with \lcdm. In particular \citet{napolitano05a} argued that a substantial number of early-type galaxy halos appear less concentrated than expected. Gravitational lensing provides further evidence that, at least some, early-type galaxies possess substantial DM halos \citep[\eg][]{kochanek95a,fischer00a,rusin02a}. Since weak lensing of galaxies only provides useful mass constraints in a statistical sense, the relatively rare instances of strong lensing are required to study DM in individual systems. Nonetheless it has been possible in a few cases to decompose the mass into stellar and DM components, albeit with strong assumptions or additional observational constraints \citep[\eg][]{rusin03a,treu04a}. X-ray observations of the hot gas in early-type galaxies provide a complementary means to infer the mass-profiles {\em via} techniques similar to those used in studying clusters. Since the X-ray emission from early-type galaxies is typically not very bright, prior to the advent of \chandra\ and \xmm\ this was limited by the relatively sparse information on the radial temperature and density profiles of the hot gas which could be determined by prior generations of satellites. Notwithstanding this limitation, large M/L ratios (consistent with substantial DM) were inferred for a number of X-ray bright galaxies, albeit with strong assumptions concerning the temperature and density profiles \citep[\eg][]{forman85a,loewenstein99b}. Using a novel technique which relied, instead, on the ellipticity of the X-ray halo, \citet{buote94} were able robustly to detect DM in the isolated elliptical NGC\thin 720 \citep[see also][]{buote96a,buote98d,buote02b}. Detailed measurements of the radial mass distribution were, however, largely restricted to a few massive systems, which may be entwined with a group halo \citep[\eg][]{irwin96,brighenti97a}. Nevertheless \citet{brighenti97a} were able to decompose the mass profiles of two systems, NGC\thin 4472 and NGC\thin 4649, into stellar and DM components. \citet{sato00a} investigated the \mvir-c relation using \asca\ for a sample of objects ranging from massive clusters to $\sim$3 elliptical galaxies. The limited spatial resolution of \asca\ necessitated some assumptions about the density profiles and, crucially, the authors neglected any stellar mass component in their fits. This omission may explain the very steep \mvir-c relation (with c$_{200}$\gtsim 30 for the galaxies) found by these authors, in conflict with \lcdm\ \citep{mamon05a}. Although mass profiles of early-type galaxies are beginning to appear which exploit the improved sensitivity and resolution of \chandra\ and \xmm, many of the most interesting constraints on DM are still restricted to massive systems, which may be at the centres of groups. For example, \citet{fukazawa06a} reported \chandra\ and \xmm\ M/L profiles for $\sim$50 galaxies and groups, confirming $\sim$flat profiles within \reff\ which rise at larger radii. However, the constraints at large radii were dominated by the massive (group-scale) objects so the implications for the DM content of normal galaxies are unclear. Furthermore, the authors included a substantial number of highly disturbed systems, in which hydrostatic equilibrium may be questioned, and failed to account for the unresolved sources which dominate the emission in the lowest-\lx\ objects in their sample\footnote{Although the authors account for unresolved sources when measuring the gas temperature, they do not account for it when computing the gas density, where its effect is more pronounced}. Recently, however, detailed \chandra\ and \xmm\ mass profiles have begun to appear for isolated early-type galaxies, also confirming the presence of massive DM halos \citep[\eg][]{osullivan04b,khosroshahi04a}. This paper is part of a series \citep[see also][]{gastaldello06a,zappacosta06a,buote06b,buote06a} using high-quality \chandra\ and \xmm\ data to investigate the mass profiles of galaxies, groups and clusters. This provides an unprecedented opportunity to place definitive constraints upon the \mvir-c relation over $\sim$2 orders of magnitude in \mvir. In this paper, we focus on the temperature, density and mass profiles of seven galaxies and poor groups chosen from the \chandra\ archive. In order to compare to theory we perform spherically-averaged analysis, leaving a discussion of the ellipticities of the X-ray halos to a future paper. In \S~\ref{sect_targets} we discuss the target selection. The data-reduction is described in \S~\ref{sect_reduction} and the X-ray morphology is addressed in \S~\ref{sect_imaging}. We discuss the spectral analysis in \S~\ref{sect_spectra}, the mass analysis in \S~\ref{sect_mass}, the systematic uncertainties in our analysis in \S~\ref{sect_systematics} and reach our conclusions in \S~\ref{sect_discussion}. The three systems for which we find \mvir$<10^{13}$\msun\ are optically isolated and so we refer to them as ``galaxies'', and the other systems in our sample as groups. We discuss this in more detail in \S~\ref{discussion_groups}. In this paper, all error-bars quoted represent 90\% confidence limits, unless otherwise stated, and we computed Virial quantities assuming a ``critical overdensity'' factor for the DM halos of $\rho_{\rm halo}/\rho_{\rm crit} = 103$ (where $\rho_{\rm halo}$ is the mean density of a sphere of mass \mvir\ and radius \rvir). \section{Target Selection} \label{sect_targets} We chose, for this initial study, to focus on objects observed with \chandra. \chandra\ data are particularly valuable for the study of galaxies since the unprecedented spatial resolution makes it possible to resolve the temperature and density profiles deep into the galaxy core, allowing us to disentangle the stellar and dark matter, and resolve them into discrete components. We initially chose a set of potential target systems from detections listed in the X-ray catalogue of \citet{osullivan01a} which have non-grating ACIS data in the \chandra\ archive. To eliminate bright groups and cluster cDs in the sample, we excluded galaxies with \lx\gtsim $10^{43}$\ergps. In order to perform the required spatially-resolved spectroscopy, we required at least $\sim$5000 hot gas photons. The potential targets were processed and the 0.1--10.0~keV image examined for evidence of large-scale disturbances (\S~\ref{sect_imaging}). We included some systems with low-amplitude asymmetries which should not strongly disturb hydrostatic equilibrium (we discuss this in more detail in \S~\ref{sect_asymmetry}). Preliminary analysis was conducted to estimate the Virial mass of the object (\S~\ref{sect_mass}). Since we aimed to focus on lower-mass objects, systems for which a fit using a simple NFW profile yielded \mvir\gtsim $10^{13}$\msun were discounted. Massive objects of this type are the focus of another study \citep{gastaldello06a}. The most promising candidates for study found {\em via} this method were chosen for detailed analysis. The properties of the 7 objects in our sample and the \chandra\ exposures are shown in Table~\ref{table_obs}. Our selection criteria naturally bias the sample towards X-ray bright galaxies. One might expect that galaxies sitting in deep potential wells are more likely to retain hot gas than those with little dark matter, and so our results may be biased somewhat towards those galaxies with substantial dark halos \citep[in contrast to the opposite bias in the analysis of][]{romanowsky03a}. As we are selecting objects which are not heavily disturbed, we are also biased towards galaxies which have not recently undergone a major merger. For the purposes of this paper, however, we do not require statistical completeness, and we will discuss how to take account of these selection effects in \citet{buote06b}. \begin{deluxetable*}{lllllllrrr} \tablecaption{The galaxy sample\label{table_sample}} \tabletypesize{\scriptsize} \tablehead{ \colhead{Galaxy} & \colhead{Type} & \colhead{\lb} & \colhead{\lk} & \colhead{Dist} & \colhead{Scale} & \colhead{\reff} & \colhead{ObsID} & \colhead{Date} & \colhead{Exposure} \\ \colhead{} & \colhead{} & \colhead{($10^{10}$\lsun)} &\colhead{($10^{11}$\lsun)} & \colhead{(Mpc)} & \colhead{(\arcsec\ kpc$^{-1}$)} & \colhead{(kpc)} & \colhead{} & \colhead{(dd/mm/yy)} & \colhead{(ks)} } \startdata NGC\thin 720 & E5 & 3.1 & 1.7 & 25.7 & 8.1 & 3.1 & 492 & 12/10/00 & 17 \\ NGC\thin 1407 & E0 & 6.4 & 3.1 & 26.8 & 7.8 & 4.4 & 791 & 16/08/00 & 38 \\ NGC\thin 4125 & E6 pec Liner & 4.7 & 1.8 & 22.2 & 9.4 & 3.3 & 2071& 09/09/01 & 63 \\ NGC\thin 4261 & E2-3 Liner Sy3& 4.4 & 2.2 & 29.3 & 7.1 & 3.4 & 834 & 06/05/00 & 34 \\ NGC\thin 4472 & E2/S0(2) Sy2 & 7.5 & 3.2 & 15.1 & 14 & 4.0 & 321 & 12/06/00 & 34 \\ NGC\thin 4649 & E2 & 5.1 & 2.5 & 15.6 & 13 & 3.2 & 785 & 20/04/00 & 21 \\ NGC\thin 6482 & E Liner & 10.9 & 3.2 & 58.8 & 3.6 & 3.4 & 3218& 20/05/02 & 18 \enddata \tablecomments{The galaxies in the sample. Distances were obtained from \citet{tonry01}, corrected for the the new Cepheid zero-point \citep{jensen03}, except for NGC\thin 6482, for which we adopted the kinematical distance modulus from \leda. \lb\ was obtained from \leda, corrected to our distance. \ks-band luminosities (\lk) and effective radii (\reff) were obtained from \twomass. We assumed ${\rm M_{B\odot}=5.48}$ and ${\rm M_{K\odot}=3.41}$ \citep[\eg][]{maraston98a}. We also list the image scale (Scale), which is the number of arc seconds corresponding to 1 kpc. We list the observation ID (ObsID) and total exposure times, after having eliminated flaring intervals.}\label{table_obs} \end{deluxetable*} \section{Data reduction} \label{sect_reduction} For data reduction, we used the \ciao\ 3.2.2 and \heasoft\ 5.3 software suites, in conjunction with \chandra\ calibration database (\caldb) version 3.1.0. Spectral-fitting was conducted with \xspec\ 11.3.1w. In order to ensure the most up-to-date calibration, all data were reprocessed from the ``level 1'' events files, following the standard \chandra\ data-reduction threads\footnote{\href{http://cxc.harvard.edu/ciao/threads/index.html}{http://cxc.harvard.edu/ciao/threads/index.html}}. We applied the standard correction to take account of the time-dependent gain-drift as implemented in the standard \ciao\ tools. To identify periods of enhanced background (``flaring''), which seriously degrades the signal-to-noise (S/N) and complicates background subtraction \citep{markevitch02} we accumulated background lightcurves for each exposure from low surface-brightness regions of the active chips. We excluded obvious diffuse emission and data in the vicinity of any detected point-sources (see below). Periods of flaring were identified by eye and excised. Small amounts of residual flaring not removed by this procedure can be important in low surface-brightness regions at large radii, but this was taken into account in our treatment of the background (\S~\ref{sect_bkd}). The final exposure times are listed in Table~\ref{table_obs}. Point source detection was performed using the \ciao\ tool {\tt wavdetect} \citep{freeman02}. Point sources were identified in full-resolution images of the \acis\ focal-plane, containing all active chips (except the S4 chip, which suffers from serious ``streaking'', which can lead to false detections). To maximise the likelihood of identifying sources with peculiarly hard or soft spectra, images were created in three energy bands, 0.1--10.0~keV, 0.1--3.0~keV and 3.0--10.0~keV. Sources were detected separately in each image. In order to minimize spurious detections at node or chip boundaries we supplied the detection algorithm with exposure-maps generated at energies 1.7~keV, 1.0~keV and 7~keV respectively (although the precise energies chosen made little difference to the results). The detection algorithm searched for structure over pixel-scales of 1, 2, 4, 8 and 16 pixels, and the detection threshold was set to ensure $\sim$0.1 spurious detections per image. The source-lists obtained within each energy-band were combined and duplicated sources removed, and the final list was checked by visual inspection of the images. The data in the vicinity of any detected point source were removed so as not to contaminate the diffuse emission. As discussed in \citet[][see also \citealt{kim03a}]{humphrey04a} a significant fraction of faint X-ray binary sources will not have been detected by this procedure, and so we include an additional component to account for it in our spectral fitting (\S~\ref{sect_spectra}). For each galaxy, we extracted spectra in a number of concentric annuli, centred on the nominal X-ray centroid. We determined the centroid iteratively by placing a 0.5\arcmin\ radius aperture at the nominal galaxy position (obtained from \ned) and computing the X-ray centroid within it. The aperture was moved to the newly-computed centroid, and the procedure repeated until the computed position converged. Typically the X-ray centroid agreed with that from \ned. The widths of the annuli were chosen so as to contain approximately the same number of background-subtracted photons and ensure there were sufficient photons in each to perform useful spectral-fitting. The data in the vicinity of any detected point-sources were excluded, as were the data from the vicinity of chip gaps, where the instrumental response may be uncertain. We extracted products from all active chips, excluding the S4, since it suffers from considerable ``streaking'' noise. Appropriate count-weighted spectral response matrices were generated for each annulus using the standard \ciao\ tasks {\bf mkwarf} and {\bf mkacisrmf}. \subsection{Background estimation} \label{sect_bkd} One of the chief difficulties in performing spectral-fitting of diffuse emission is the proper treatment of the background. A set of standard blank-field ``template'' files are available for \chandra\ as part of the \caldb. We found, however, that the background template files are not sufficiently accurate to use in the very low surface brightness regions at large radii, which are crucial to determine interesting global mass constraints. The background comprises cosmic, instrumental and non X-ray (particle) components. The cosmic component is known to vary from field to field, while the non X-ray background exhibits long-term secular variability. To mitigate the latter effect, several authors have adopted the practice of renormalizing the background template to ensure good agreement with their data at high energies (\gtsim 10~keV). Such a procedure, however, also renormalizes the (uncorrelated) cosmic X-ray background and instrumental line features, which can lead to serious over- or under-subtraction. Given these reservations we chose to use an alternative background estimation procedure. Our method involved modelling the background, somewhat akin to the approach of \citet{buote04c}. All of the targets were centred on the \acis-S3 chip, which is back-illuminated (BI). To obtain constraints on the background, we extracted spectra from a $\sim$2\arcmin\ region centred on the S1 chip, which is also BI, and from an annulus centred at the galaxy centroid and with an inner and outer radii typically $\sim$2.5\arcmin\ and 3.3\arcmin. We excluded data from the vicinity of any point-sources found by the source detection algorithm. Although the diffuse emission from each galaxy typically had a very low surface-brightness on the S1 CCD, we found that using two regions in this way with different contributions of source emission enabled the background components to be most cleanly disentangled from the source. The \acis\ focal plane also consists of front-illuminated (FI) chips, which have significantly different (and lower) background. To obtain an estimate of the background for these chips, we extracted spectra from the entirety of each chip, excluding detected sources and data towards the edge of the chips where the exposure-map may be uncertain. In order to constrain the model, we fitted all spectra simultaneously, without background subtraction, using \xspec. Our model consisted of a single \apec\ plasma (to take account of the diffuse emission from the galaxy; the ``source''), plus background components. These comprised a power law with $\Gamma=1.41$ (to account for the hard X-ray background), two \apec\ models with solar abundances and kT$=$ 0.2 and 0.07~keV (to account for the soft X-ray background) and, to model the instrumental and particle contributions, a broken power law model and two Gaussian lines with energies 1.7 and 2.1~keV and negligible intrinsic widths. We have found that this model can be used to parameterize adequately the template background spectra. In general, the instrumental contributions of the FI chips were very similar in shape. Therefore, the background components of all the FI chips were tied, assuming the normalization scaled with the spectral extraction area. For the BI chips, there was some evidence that the S1 chip background can be somewhat larger at energies \gtsim 5~keV (although this is variable). In order to disentangle the source and background components, given the general lack of photons in these spectra, we tied the abundances and temperatures of the ``source'' \apec\ components between the extraction regions, but allowed the normalizations to be free. Where there was a significant improvement in the fit-statistic if this assumption was relaxed, we allowed the abundances or temperatures to fit freely. Notwithstanding, this assumption should not significantly affect our results. This model was able to fit all of the data well. In our subsequent spectral analysis, we did not background-subtract the data using the standard templates, but took into account the background by using appropriately scaled versions of the models fitted to each CCD, which were added according to the overlap between the source region and the CCD. We found that the standard background templates fared much worse than these modelled background estimates when the data were from regions of low surface-brightness. We discuss the impact of the background treatment on our results in \S~\ref{sect_systematics_bkd}. \section{X-ray images} \label{sect_imaging} The X-ray image of each galaxy was examined to identify any obvious surface-brightness disturbances or asymmetries which would be indicative of clear deviations from hydrostatic equilibrium. We note that low-level X-ray asymmetries, such as the ``fingers of emission'' identified by \citet{randall03} in the adaptively-smoothed images of NGC\thin 4649, probably do not merit concern\footnote{Although the authors suggested these may arise from bulk convective flow, the spectra do not agree with simulations of such.}, as, provided care is taken to avoid seriously disturbed emission regions, reliable mass profiles can be inferred even in mildly disturbed systems \citep{buote95a}. In Fig~\ref{fig_images} we show the 0.1--10.0~keV \acis-S3 images of each of the systems. These images were first processed to remove point-sources, using the \ciao\ tool {\em dmfilth}, which replaces photons in the vicinity of each point-source with a locally-estimated background. NGC\thin 4261 contains an AGN which appears as a bright central X-ray source and there is evidence of a small, low surface-brightness jet \citep{zezas05a}. We have also removed these sources from the image. The images were flat-fielded with the 1.7~keV monochromatic exposure-map (although this analysis is insensitive to the choice of energy), and then smoothed by convolution with a 5\arcsec\ gaussian, to make large-scale structure more apparent. Due to the low surface-brightness nature of the emission at large radii, it is difficult to appreciate X-ray emission outside $\sim$a few arc minutes in many of the images. However, detailed spectral analysis and azimuthally-averaged surface brightness analysis reveals substantial hot gas extending beyond the edge of the S3 chip in each system. None of the objects show very obvious disturbances in their X-ray emission on the \acis-S3 chip \citep[such as those found in NGC\thin 4636:][]{jones02a}. Some low-amplitude features are evident such as the faint jet in NGC\thin 4261 (which is not visible in the above images), a possible north-south asymmetry in NGC\thin 1407 and some asymmetry, in particular an off-axis X-ray enhancement, in NGC\thin 4125. Based on adaptively-smoothed \xmm\ images, \citet{croston05a} argued that the X-ray emission in NGC\thin 4261 is anti-correlated with the galaxy radio lobes. By inspection of the \xmm\ images, this actually appears to be a very low-amplitude effect. It is not obvious in the \chandra\ images, although the X-ray isophotes do align somewhat perpendicularly to the jet. In any case, this does not appear to have significantly disturbed hydrostatic equilibrium, since there is excellent agreement between our inferred mass profile and a model comprising stellar plus DM components (\S~\ref{sect_mass}), which would be an extraordinary coincidence if hydrostatic equilibrium had been strongly disturbed. The limited field-of-view makes it difficult to assess asymmetries and disturbances on the other chips. NGC\thin 4472 is known, however, to exhibit a disturbance outside $\sim$6\arcmin\ \citep{irwin96}, but mass analysis inside this radius should be reliable. We assess the impact of all these features in \S~\ref{sect_asymmetry}. \section{Spectral Analysis} \label{sect_spectra} Spectral-fitting was carried out in the energy-band 0.5--7.0~keV, to avoid calibration uncertainties at lower energies and to minimize the instrumental background, which dominates at high energies. The spectra were rebinned to ensure a S/N ratio of at least 3 and a minimum of 20 photons per bin (to validate $\chi^2$ fitting). We fitted data from all annuli simultaneously using \xspec. To model the hot gas we adopted a {\bf vapec} component, plus a bremsstrahlung component for all annuli within the twenty-fifth magnitude isophote (\dtwentyfive) of each galaxy, taken from the Third Reference Catalog of Bright Galaxies \citep[RC3:][]{devaucouleurs91}, to account for undetected point-sources \citep[this model gives a good fit to the composite spectrum of the detected sources in nearby galaxies:][]{irwin03a}. We used a slightly modified form of the existing \xspec\ {\bf vapec} implementation so that \zfe\ is determined directly, but for the remaining elements the abundance ratios (in solar units) were directly determined with respect to Fe. This was useful since, in general, the data did not enable us to determine any abundance {\em ratio} gradients and so we tied the abundance ratios between all annuli. Where abundances or abundance ratios could not be constrained, they were fixed at the Solar value. We adopted the solar photospheric abundances standard of \citet{asplund04a}. We refer the interested reader to \citet{humphrey05a} for a detailed discussion of this choice and how to convert our results to older abundance standards. In the interests of physically reasonable results, we constrained all abundances and abundance ratios to the range 0.0--5.0~times solar. The absorbing column density (\nh) was fixed at the Galactic value \citep{dickey90}; the effect of varying \nh\ is discussed in \S~\ref{sect_systematics_spectra}. For NGC\thin 4261, our innermost annulus contained substantial contamination from the central AGN. However, this was sufficiently absorbed that the thermal emission from the gas can be clearly disentangled from it. To account for the AGN emission, we fitted a highly absorbed (\nh${\rm =10^{+6}_{-4}\times 10^{22} cm^{-2}}$) power law component ($\Gamma=1.4\pm0.8$). We discuss the impact of including this annulus on our fits in \S~\ref{sect_asymmetry}. To account for projection effects, we used the {\bf projct} model implemented in \xspec. This model, unfortunately, does not take into account the emission from gas outside the outermost shell, which is also projected into the line-of-sight. To take account of this effect, we assumed that the emission outside this shell has the same spectral shape as the emission in that shell and a density profile well-described by a $\beta$-model \citep[\eg][]{buote00c}. We included an extra spectral component to our fits of each annulus to account for projected emission from this gas. To estimate the parameters of the $\beta$-profile, we we fitted the galaxy surface brightness, using dedicated software, in the 0.1--3.0~keV band. Although a single $\beta$-model did not always match the fine detail of the surface brightness profiles, it adequately parameterized the data for our purposes (our results are not expected to be strongly dependent upon the parameters of this fit). We obtained good fits to the spectra of each galaxy with this model. The best-fitting abundances were in excellent agreement with those of other early-type galaxies \citep{humphrey05a}, and are shown in Table~\ref{table_abundances}. We note that \citet{randall05a} found \zsi/\zfe$\simeq$1.7 for NGC\thin 4649 (adjusting to our abundances standard) when fitting the data from single, large aperture, which they argued points to substantial enrichment from SN~II, in stark contrast to the predominantly SN~Ia enrichment we found in such galaxies \citep{humphrey05a}. From our analysis, however, \zsi/\zfe$\simeq$1, which is more consistent with our results for other systems. The discrepancy appears to be related to the ``Fe bias'' \citep[where \zfe\ is systematically underestimated if one assumes multi-temperature gas is isothermal:][]{buote00c} which has suppressed their large aperture \zfe\ in comparison to their spatially-resolved results (which agree better with our measurement). Error-bars were computed {\em via} the Monte-Carlo technique which we have extensively used in past analyses \citep[\eg][]{buote03a}. We simulated spectra from the best-fit models, which were then fitted exactly analogously to the real data. We performed 25 simulations, which were sufficient to assess the distribution of the fit parameters about the best-fit values; the standard deviation of this distribution corresponds to the 1-$\sigma$ confidence region.Assuming that we have found the global minimum, and the fit statistic follows a $\chi^2$ distribution this is statistically equivalent to searching the parameter space for changes in the fit statistic. Temperature and density profiles are discussed below (\S~\ref{sect_temp_profiles} and \S~\ref{sect_mass_results}) \begin{deluxetable*}{lrrrrrrrr} \tabletypesize{\scriptsize} \tablecaption{Emission-weighted average abundances\label{table_abundances}} \tablehead{ \colhead{Galaxy} & \colhead{$\chi^2$/dof} & \colhead{\zfe} & \colhead{\zo/\zfe} & \colhead{\zne/\zfe} & \colhead{\zmg/\zfe} & \colhead{\zsi/\zfe} & \colhead{\zs/\zfe} & \colhead{\zni/\zfe} } \startdata NGC\thin 720$^1$ & 383.4/357& $0.80^{+0.45}_{-0.24}$ & 0.30$\pm 0.28$ & 0.68$\pm0.67$ & 1.26$\pm0.35$ & \ldots& \ldots & \ldots \\ NGC\thin 1407$^1$ & 222/221& 2.1$^{+1.1}_{-0.9}$$^\dagger$& 0.37$^{+0.21}_{-0.25}$ & \ldots & 1.10$\pm 0.23$ & 1.21$^{+0.31}_{-0.27}$ & 2.2$\pm1.1$ & 3.3$^{+1.7}_{-1.3}$\\ NGC\thin 4125 & 327/307 & 0.55$^{+0.22}_{-0.13}$ & 0.29$^{+0.13}_{-0.09}$ & 0.62$\pm$0.14 & 0.33$\pm0.20$ & \ldots & \ldots & \ldots \\ NGC\thin 4261 & 307/319 & 1.72$\pm0.50^\dagger$ & $<$0.23 & 0.36$^{+0.79}_{-0.36}$ & 0.83$\pm$0.23& 1.2$\pm$0.4& \ldots & 1.8$^{+2.3}_{-1.8}$ \\ NGC\thin 4649 & 563/491 & 2.32$^{+0.87}_{-0.37}$ & $<0.15$ & \ldots & 0.97$\pm$0.13 & 1.02$\pm$0.13 & \ldots & 1.42$^{+0.85}_{-0.73}$ \\ NGC\thin 4472$^1$& 785/740 & $1.4^{+1.7}_{-0.4}$$^\dagger$& 0.51$\pm$0.12 & 0.95 $\pm0.44$ & 1.02$\pm$0.11 & 1.25$\pm$0.11 & 2.36$\pm$0.33 & 3.28$\pm0.61$\\ NGC\thin 6482 & 256/262 & $>$2.5 & 0.34$\pm$0.20 & \ldots & 1.15$\pm$0.18 & 1.3$\pm$0.3 & \ldots & 3.2$^{+1.5}_{-1.2}$ \enddata \tablecomments{The best-fitting globally-averaged emission-weighted abundances and abundance ratios for each galaxy, shown along with the quality of fit. Statistical errors represent the 90\% confidence region. Where we were able to constrain an abundance gradient, we estimated an emission-weighted \zfe, extrapolated over a large aperture \citep[see][]{humphrey05a}; those affected galaxies are marked ($^\dagger$). $^1$---results taken from \citet{humphrey05a}. Where parameters could not be constrained, they were fixed at the Solar value, and listed as ``\ldots''.} \end{deluxetable*} \section{Mass modelling} \label{sect_mass} \subsection{Assumed potential method} \label{sect_potential} We adopted two complementary approaches in order to determine the mass profiles of the galaxies in the sample. The first method, discussed here, was found to be less sensitive to the assumptions of the modelling and therefore was adopted as our default. We discuss our alternative approach in \S~\ref{sect_xmass}. Starting with a parameterised model for the temperature (T) and gravitating mass ($M_{grav}$) profiles, the equation of hydrostatic equilibrium can be solved for \rhog\ thus: \begin{equation} \ln \left( \frac{\rho_g}{\rho_{g0}} \right) = - \ln \left( \frac{T}{T_0} \right) - G \mu m_p \int^R_{R_0} \frac{M_{grav}(<R)}{kT R^2} dR \label{eqn_hydrostatic_rho} \end{equation} where R is the radius from the centre of the gravitational potential, \rhog\ is the gas density, $\rho_{g0}$ and $T_0$ are density and temperature at some ``reference'' radius $R_0$, k is Boltzmann's constant, G is the universal gravitational constant, $m_p$ is the atomic mass unit and $\mu$ is the mean atomic weight of the gas. In our fitting we explicitly ignored the contribution of the gas to the gravitating mass, but we subsequently verified this contributed \ltsim 1\% of the total gravitating matter within 100~kpc, justifying this assumption. We developed software to fit \rhog\ and temperature profiles simultaneously using this procedure. For speed we assumed that the density and temperature data-points were each evaluated at a single point, the radius of which was given by: \begin{equation} \bar{R_i} = \left( 0.5*(Rin_i^{1.5}+Rout_i^{1.5}) \right)^{2/3} \label{eqn_radius} \end{equation} where $Rin_i$ and $Rout_i$ were the inner and outer radius of the bin \citep[see][]{lewis03a}. \subsection{Temperature profiles} \label{sect_temp_profiles} There were considerable differences in the temperature profiles from object to object (Fig~\ref{fig_temp}), so that we were not able to adopt a ``universal'' profile for all of the systems. {\em A priori} we do not expect any particular form for the temperature profile and so we determined appropriate functional forms for our temperature models empirically. Based on experience, the following ``toolbox'' of models provided adequate flexibility to ensure at least one model can describe the temperature profiles reasonably well \citep[see][]{buote06a}: \begin{eqnarray} T & = & T_0 + T_1 \left[1+x^{-\epsilon}\right]^{-1} \label{eqn_trise2}\\ T & = & \left[T_0 + T_1 x^{p_1}\right]e^{-x^{p_e}} + T_2 x^{p_2} \left[1 - e^{-x^{p_e}}\right] \label{eqn_pow2expcut2} \\ T & = & \frac{A}{A+B}\left[ T_0+T_1\left( \frac{x_1}{1+x_1}\right)^{p_1} \right] + \nonumber\\ & & \frac{B}{A+B}\left[ T_2+T_3(1+x_2)^{-p_2} \right] \label{eqn_twophase2} \end{eqnarray} where $x=(r/r_c)$, $x_1=(r/r_{c1})$, $x_2=(r/r_{c2})$, $A=(1+r/r_{t1})^{-3\beta_1}$ and $B=\epsilon (1+r/r_{t2})^{-3\beta_2}$. $T_0$, $T_1$, $T_2$, ${T_3}$, $r_c$, $r_{c1}$, $r_{c2}$, $r_{t1}$, $r_{t2}$, $p_1$, ${p_2}$, $p_e$ and $\epsilon$ are parameters of the fit. For NGC\thin 4261, we ignored temperature data-points from 15--25~kpc, which were poorly-determined and seemed erroneous. We experimented with fitting the projected (rather than deprojected) spectra, and found no evidence of any features (in either temperature or density) around this range of radii, strongly implying that they arise solely due to deprojection noise. The temperature profiles and best-fit models are shown in Fig~\ref{fig_temp}. Our deprojected temperature profiles generally agree with those appearing in the literature for these objects. (Although most of these are projected profiles, typically deprojection does not strongly alter the overall character of the temperature profile.) \citet{osullivan03a} reported \rosat\ profiles for all of the galaxies which, although substantially less well-constrained, agree well with our results. Likewise our NGC\thin 4472 temperature profile agrees well with the (less well-constrained) \rosat\ profile of \citet{irwin96}. Our profile for NGC\thin 4649 is in reasonable agreement with the projected \xmm\ measurements of \citet{randall05a}, and likewise our measured profile of NGC\thin 6482 agrees with the deprojected results of \citet{khosroshahi04a}. Our temperature profiles for NGC\thin 1407, NGC\thin 720 and NGC\thin 4472 were also in agreement with those we reported in \citet{humphrey05a}. \subsection{Mass-fitting results} \label{sect_mass_results} \begin{deluxetable*}{llll} \tablecaption{Quality of the mass fits\label{table_chisq}} \tabletypesize{\scriptsize} \tablehead{ \colhead{Galaxy} & \colhead{NFW} & \colhead{NFW+stars} & \colhead{AC NFW+stars} } \startdata NGC\thin 720 & 1.9/9& 1.0/8& 0.9/8\\ NGC\thin 1407 & 26.7/9& 20.7/8& 20.7/8\\ NGC\thin 4125 & 23.4/11 & 9.5/10 & 10.8/10 \\ NGC\thin 4261 & 22.6/12 & 14.0/11 & 14.0/11 \\ NGC\thin 4472 & 35.2/20 & 34.9/20 & 35.2/20 \\ NGC\thin 4649 & 30.2/7 & 11.0/6 & 11.5/6 \\ NGC\thin 6482 & 0.5/5 & 2.4/4 & 1.7/4 \enddata \tablecomments{The $\chi^2$/dof of the fits to the density and temperature profiles used to infer the mass, for the three basic mass-models adopted. For the NFW+stars and AC NFW+stars models, we constrain \fbaryons\ to Eq~\ref{eqn_baryons}.} \end{deluxetable*} \begin{deluxetable*}{lllll|rrrr} \tablecaption{Best-fitting NFW+stars results\label{table_results}} \tabletypesize{\scriptsize} \tablehead{ \colhead{Galaxy} & \multicolumn{4}{l}{\fbaryons=Eq~\ref{eqn_baryons}} & \multicolumn{4}{l}{0.032$\leq$\fbaryons$\leq$0.16}\\ \colhead{} & \colhead{\mvir ($10^{12}$\msun)} & \colhead{\rvir (kpc)} & \colhead{c} & \colhead{\fbaryons} &\colhead{\mvir ($10^{12}$\msun)} & \colhead{\rvir (kpc)} & \colhead{c} & \colhead{\fbaryons} } \startdata NGC720 &$6.6^{+2.4}_{-3.0}$ &$480^{+50}_{-90}$ &$18.^{+30.}_{-8.}$ &$0.044^{+0.037}_{-0.003}$ &$6.6^{+6.0}_{-4.3}$ &$480\pm 120$ &$18.^{+49.}_{-10.}$ &$0.044^{+0.095}_{-0.012}$ \\ NGC1407 &$16.\pm 6.$ &$650^{+80}_{-100}$ &$18.^{+11.}_{-7.}$ &$0.065^{+0.041}_{-0.001}$ &$21.\pm 15.$ &$720^{+140}_{-200}$ &$15.^{+16.}_{-6.}$ &$0.032^{+0.130}_{-0.001}$ \\ NGC4125 &$6.2^{+0.8}_{-2.3}$ &$470^{+20}_{-70}$ &$10.^{+5.}_{-2.}$ &$0.039^{+0.035}_{-0.001}$ &$7.2^{+1.4}_{-4.9}$ &$500^{+30}_{-160}$ &$9.3^{+11.}_{-2.1}$ &$0.032^{+0.13}_{-0.001}$ \\ NGC4261 &$67.^{+41.}_{-15.}$ &$1040^{+200}_{-90}$ &$3.7\pm 1.7$ &$0.14^{+0.01}_{-0.03}$ &$57.^{+260}_{-15.}$ &$990^{+760}_{-100}$ &$4.0\pm 2.0$ &$0.16^{+0.00}_{-0.13}$ \\ NGC4472 &$33.^{+6.}_{-10.}$ &$820^{+50}_{-100}$ &$13.^{+4.}_{-2.}$ &$0.084^{+0.037}_{-0.001}$ &$63.^{+17.}_{-44.}$ &$1020^{+90}_{-300}$ &$10.0^{+7.}_{-2.}$ &$0.032^{+0.13}_{-0.00}$ \\ NGC4649 &$35.^{+7.}_{-13.}$ &$840^{+60}_{-120}$ &$21.^{+6.}_{-3.}$ &$0.086^{+0.037}_{-0.001}$ &$93.^{+26.}_{-73.}$ &$1200^{+100}_{-500}$ &$15.^{+11.}_{-3.}$ &$0.032^{+0.12}_{-0.00}$ \\ NGC6482 &$7.1^{+4.4}_{-1.7}$ &$500^{+90}_{-40}$ &$18.^{+13.}_{-8.}$ &$0.075^{+0.013}_{-0.032}$ &$3.6^{+5.5}_{-1.5}$ &$390^{+140}_{-70}$ &$38.^{+76.}_{-24.}$ &$0.16^{+0.00}_{-0.10}$ \enddata \tablecomments{The best-fitting results for the NFW+stars model. All error-bars shown correspond to 90\% confidence regions. The fit results for the AC NFW+stars model are very similar, and are shown in Fig~\ref{fig_confidence}. Results are shown for the fits using the two different constraints on \fbaryons\ we adopted (see text).} \end{deluxetable*} We tested three different mass-models against the data. In order to investigate the suggestion that historically large c values found based on X-ray analysis were an artefact of the omission of the stellar mass, as well as to investigate the scenario of \citet{loeb03a}, we first tested a model comprising a single NFW profile. Although stellar kinematical results would seem to rule out the \citeauthor{loeb03a} picture, our analysis of more massive systems \citep{gastaldello06a} does suggest that the stellar mass may not be uniformly required in all systems. In order to take into account the stellar mass, we fitted a model comprising an NFW DM component, plus a \citet[][hereafter H90]{hernquist90} mass component, the \reff\ of which being fixed to that measured in the \ks-band (Table~\ref{table_obs}). The H90 model is, in projection, a good approximation to the familiar de Vaucouleurs profile of elliptical galaxies. To test whether the DM halos retained any evidence of their response to baryonic condensation, we further adopted an H90 component, plus an NFW component modified by the adiabatic contraction model of \citet{gnedin04a}\footnote{Available publicly from \href{http://www.astronomy.ohio-state.edu/$\sim$ognedin/contra/}{http://www.astronomy.ohio}-state.edu/$\sim$ognedin/contra/}. Hereafter, we refer to these three models as, respectively, NFW, NFW+stars and AC NFW+stars. Our computed \mvir\ for each system included both dark and stellar mass. For NGC\thin 4472 and NGC\thin 4649, which lie in Virgo, there is the possibility that the DM halo may have experienced some tidal truncation at a radius $<$\rvir. Our measured Virial quantities relate to the original halo prior to truncation. We also experimented with replacing the NFW component with the less cuspy \citet[][hereafter N04]{navarro04a} model, which gives an improved fit to DM halos in high-resolution N-body simulations. However, since the \mvir-c relation was calibrated using the NFW model we treat this choice as a systematic effect and it is discussed in \S~\ref{sect_n04}. For NGC\thin 4261 we ignored a deviant \rhog\ data-point at $\sim$11~kpc, in addition to the excluded temperature data-points discussed above. % In Fig~\ref{fig_density} we show the density profiles (along with the best-fitting AC NFW+stars model, which is described below). In Fig~\ref{fig_confidence}(a) we show the best-fitting 1-$\sigma$ contours of c {\em versus} \mvir\ for the NFW model fitted to each galaxy. The fits were typically, but not uniformly, poor (Table~\ref{table_chisq}). We found very large ($\gg$20) values for c, completely inconsistent with the expectation of N-body simulations. Good constraints on the global halo properties typically require interesting density and temperature constraints over as large a radial range as possible. In our case, the absence of data outside $\sim$50--100~kpc ($\sim$0.1--0.2\rvir) therefore makes the inner data-points critical in determining the profile of the halo. Unfortunately, since the scale radius of a galaxy-size DM halo is $\sim$10--30~kpc, there is some degeneracy between the DM and stellar mass components at small radii. As we discuss in \S~\ref{sect_mass_to_light} there are considerable uncertainties in estimating a reliable mass-to-light (M/L) ratio from the characteristics of the stellar population. We found that the results are extremely sensitive to the stellar M/L adopted; we found that varying this ratio by as little as 20\% could cause \mvir\ variations of $\sim$50--100\% \citep[see][]{humphrey05b}. It was therefore necessary to allow the stellar mass to be determined as a parameter of the fit. This, unfortunately, made it very difficult to constrain \mvir\ or c, unless additional constraints were applied to the fit. One way to achieve this is to constrain the fit to lie on the mean \mvir-c relation predicted from N-body simulations \citep[\eg][]{bullock01a}. Although this would prevent our measuring \mvir\ and c independently, it would enable us to determine whether the data are consistent with the mean relation. However, this relation was determined for an unbiased sample of DM halos, whereas our selection criteria (\S~\ref{sect_targets}) should bias us towards systems which have not recently had a merger (implying earlier-forming, hence more concentrated, objects). Furthermore, individual halos are not expected to lie exactly on the mean \mvir-c relation, but be scattered about it. Nevertheless, we experimented with applying this constraint. The data for each galaxy were consistent with this model, but we found \mvir\ was generally poorly constrained, and extremely sensitive to any scatter we introduced about the mean \mvir-c relation. A far more useful way to constrain the fit was to restrict the total baryon fraction (\fbaryons) in the system. Such a constraint is useful since we found that, for a given system, \fbaryons\ determined from our fits was strongly anti-correlated with the measured \mvir. To estimate \fbaryons, we computed the gas mass by extrapolating our \rhog\ model from the centre of the innermost radial bin to the Virial radius. The contribution of stars to the total baryon fraction was derived from the stellar mass found by our fits. We crudely took into account the fact that not all of the stellar mass within \rvir\ is necessarily contained in the central galaxy by scaling this mass by the ratio of the total B-band light of all putative ``group'' members listed in the catalogue of \citet[][hereafter G93]{garcia93} to that of the central galaxy. This is likely to overestimate slightly the stellar mass content, since it assumes the same stellar M/L ratio for all low-mass companions/ group members, whereas some fraction of these are likely to have substantial young stellar populations, with lower M/L ratios. We discuss the impact of this assumption in \S~\ref{sect_systematics_baryon_fraction}. G93 lists NGC\thin 4649 as belonging to the NGC\thin 4472 ``group'', whereas they both have distinct X-ray halos, indicating they are, in fact, distinct systems. As a zeroth order approximation, we therefore assumed that the total B-band luminosity was divided between the two ``subgroups'' in proportion to the central galaxy's B-band luminosity. In practice, between $\sim$25\% (for NGC\thin 4472) and 84\% (for NGC\thin 4125) of the B-band light of the system resides in the central galaxy. NGC\thin 6482 was not listed in G93, but as it is known to be relatively isolated \citep{khosroshahi04a}, we assumed that $\sim$80\% of its mass is in the central galaxy, consistent with the other relatively isolated systems. Based on hydrodynamical simulations incorporating gas cooling and supernovae feedback, \citet{kay03a} predicted \fbaryons\ as a function of Virial temperature for systems with \mvir\gtsim a few $\times 10^{12}$\msun. Fitting their data by eye, converting from Virial temperature to mass, and assuming a Universal baryon fraction of 0.16, we estimate \begin{eqnarray} f_{b} = & 0.062 \log_{10}(M_{14})+0.13 & (M_{14}<1.02) \label{eqn_baryons} \\ f_{b} = & 0.14 & (M_{14}>1.02)\nonumber \end{eqnarray} where ${M_{14}=M_{vir}/10^{14} M_\odot}$, with an approximate scatter of $\pm$0.02. Adopting this constraint (including the allowed range of scatter) and fitting the NFW+stars model resulted in significant improvements in the fit quality over the simple NFW model (Table~\ref{table_chisq}). We show the resulting c-{\em versus}-\mvir\ contours in Fig~\ref{fig_confidence}(b), and summarise our results in Table~\ref{table_results}. Clearly adding the stellar mass component allows the DM halos to be substantially less concentrated, since less DM is required in the centre of the halo. These results were in much better agreement with the results of N-body simulations than those obtained with the NFW model. There is, however, a slight trend towards more concentrated halos than \lcdm. Since the results of simulations can be sensitive to the rather uncertain process of feedback, we additionally adopted, as a somewhat less restrictive constraint on \fbaryons, 0.03$<$\fbaryons$<0.16$, the lower limit being $\sim$the lowest values found in \citeauthor{kay03a}'s simulations. The shallower potential of very low-mass halos makes it more difficult for them to hold onto their hot gas, and so our lower limit on \fbaryons\ may be an overestimate if the Virial mass is small. However, imposing a lower limit on \fbaryons\ in the fitting algorithm actually works to exclude the most massive solutions, which we would expect to be closer to baryonic closure \citep{mathews05a}. It is conceivable that some more massive (\mvir\gtsim $10^{13}$\msun) systems are rare examples of ``dark'' groups which have unusually low baryon fractions, a possibility we return to in \S~\ref{sect_ngc1407}. Notwithstanding, the results are shown in Figs~\ref{fig_confidence}(c), which are qualitatively similar to those obtained applying the more restrictive constraint on \fbaryons, although with larger uncertainty. In Fig~\ref{fig_mass_to_light} we show the gravitating mass to K-band light (\mgrav/\lk) ratio profile implied by our best-fit models for each system. In each case we found that the X-ray emission was considerably more extended than the optical light. We also show data-points estimated from ``parameterized profile'' mass modelling (\S~\ref{sect_xmass}), which tend to agree reasonably well; the slight systematic differences between the profiles are an artefact of the assumptions used to derive the data-points and we discuss this in detail in \S~\ref{sect_xmass}. Clearly \mgrav/\lk\ increases very slowly with radius within \reff, rising very steeply outside this range. This arises naturally from the very different shapes of the stellar and DM halos, and is similar to M/L profiles seen from stellar kinematics and the results of \citet{brighenti97a} for NGC\thin 4472 and NGC\thin 4649. By \rvir, \mgrav/\lk\ reaches as high as $\sim$20--40 \msun/\lsun\ for the galaxy-scale systems or $\sim$100-200 \msun/\lsun\ for the group-like objects. We stress that this only includes the light of the central galaxy which, for the group-like systems may be a little as $\sim$25\% of the total luminosity. \subsection{Parameterized profile mass modelling} \label{sect_xmass} We briefly discuss here an alternative technique to determine the mass profiles of X-ray bright objects which we have extensively employed in our previous studies, as well as the companion papers to this present work \citep[\eg][]{lewis03a,buote06a,gastaldello06a,zappacosta06a}. This technique, which we here dub the ``parameterized profile'' method involves parameterizing independently the temperature and density profiles of the system with simple, empirical models. These functions were then inserted into the equation of hydrostatic equilibrium, which we solved for the mass enclosed within any given radius. The temperature profiles were parameterized with the models discussed in \S~\ref{sect_temp_profiles}, and to fit \rhog\ we adopted, where appropriate a, $\beta$-model, a ``double-$\beta$'' model or a ``cusped-$\beta$'' model, defined, respectively, as: \begin{eqnarray} \rho_g & = & \rho_{g0} \left[ 1+ (r/r_c)^2 \right]^{-3\beta /2} \label{eqn_beta}\\ \rho_g & = & \sqrt{\rho_{g0} \left[ 1+ (r/r_c)^2 \right]^{-3\beta} + \rho_{g1} \left[ 1+ (r/r_{c2})^2 \right]^{-3\beta_2}} \label{eqn_twobeta} \\ \rho_g & = & \rho_{g0} 2^{3\beta/2-\epsilon/2}(r/r_c)^{-\epsilon}\left[ 1+ (r/r_c)^2 \right]^{-3\beta /2+\epsilon/2} \label{eqn_cusp} \end{eqnarray} Where the parameters $\rho_{g0}$, $\rho_{g1}$, $r_c$, $r_{c2}$, $\beta$, $\beta_2$ and $\epsilon$ are determined by the fit. Fitting these models to the simulated temperature and density profiles discussed in \S~\ref{sect_spectra} (which were used therein to estimate the error-bars on kT and \rhog\ in each data-bin) allowed us to estimate the scatter in the mass data-points arising from statistical noise, and hence the error-bars. % For a full discussion of this technique, we refer the interested reader to \citet{buote06a}, who demonstrate the good agreement typically found between this method and the assumed potential modelling of \S~\ref{sect_potential}, when fitting high-quality data. However, the mass data-points, especially at the innermost and outermost radii, are rather sensitive to the parameterized models adopted to fit the temperature and, especially, \rhog. The systematic uncertainty introduced by the choice of \rhog\ model can be considerably larger than the statistical error. For our purposes the absence of data at very large radii, which are vital to constrain the curvature of the mass model, exacerbated by the uncertainty introduced at small radii due to the uncertain stellar mass-to-light ratio, magnified the impact of these systematic effects. Notwithstanding these reservations, it is still interesting to compare the results obtained {\em via} both mass-fitting methods. We show in Fig~\ref{fig_mass} the mass data-points computed using parameterized potential modelling, along with the best-fitting mass models found in \S~\ref{sect_potential}. Clearly there is good overall agreement between the two methods although there are some systematic differences, which reflect the systematics inherent in our choice of parameterized model for \rhog. \subsection{Stellar mass-to-Light ratios} \label{sect_mass_to_light} \begin{deluxetable*}{llll|rr} \tablecolumns{5} \tablecaption{Stellar mass-to-light ratios\label{table_mass_to_light}} \tabletypesize{\scriptsize} \tablehead{ \colhead{Galaxy} & \colhead{\lk/\lb} & \multicolumn{2}{c}{Fitted \mstars/\lk\ (\msun/\lsun)} & \multicolumn{2}{c}{Pop.\ synthesis \mstars/\lk\ (\msun/\lsun)}\\ \colhead{} & \colhead{} & \colhead{NFW+stars} & \colhead{AC NFW+stars} & \colhead{Salpeter IMF} & \colhead{Kroupa IMF} } \startdata NGC\thin720 & 5.5 & 0.77$^{+0.52}_{-0.71}$& 0.54$^{+0.42}_{-0.48}$& 0.54$\pm$0.11& 0.35$\pm$0.07\\ NGC\thin 1407 & 4.8 & 0.52$^{+0.25}_{-0.32}$& 0.35$\pm$0.25& 1.6$\pm$0.2& 1.1$\pm$0.1\\ NGC\thin 4125 & 3.8 & 0.72$\pm$0.11& 0.53$\pm$0.11& 1.7$\pm$0.5& 1.1$\pm$0.4\\ NGC\thin 4261 & 5.0 & 1.2$\pm$0.1& 1.0$\pm$0.1& 1.9$\pm$0.1& 1.3$\pm$0.1\\ NGC\thin 4472 & 4.3 & 0.51$^{+0.24}_{-0.31}$& 0.36$\pm$0.1& 1.3$\pm$0.3& 0.83$\pm$0.15\\ NGC\thin 4649 & 4.9 & 0.90$\pm$0.13& 0.65$\pm$0.12& 1.7$\pm$0.2& 1.1$\pm$0.1\\ NGC\thin 6482 & 2.9 & $0.73^{+0.18}_{-0.27}$& 0.52$^{+0.19}_{-0.23}$& 1.58$\pm$0.02& 1.05$\pm$0.02\\ \enddata \tablecomments{K-band stellar mass-to-light ratios measured from our fits to the data using both the NFW+stars and the AC NFW+stars models. Since AC tends to increase the cuspiness of the DM profiles, \mstars/\lk\ is substantially lower for the AC NFW+stars models. We also show the predicted \mstars/\lk\ values derived from simple stellar population synthesis, assuming either the Salpeter or \citet{kroupa01a} IMF.} \end{deluxetable*} It is interesting to compare the stellar M/L ratios (\mstars/L) determined by our fitting to the expectations of stellar population synthesis models. In order to ensure that the optical light traces the stellar mass as closely as possible, we opted to perform this comparison in the K-band. Table~\ref{table_mass_to_light} shows \mstars/\lk\ determined from our models using eq~\ref{eqn_baryons} to constrain \fbaryons. Since AC tends to increase the cuspiness of the DM profile we found a significantly lower mass-to-light ratio for the AC NFW+stars model than for the NFW+stars model. To compare our measured \mstars/\lk\ to single burst stellar population synthesis predictions, we first estimated a mean emission-weighted stellar age and metallicity for each galaxy, as outlined in Appendix~\ref{sect_stars}. We linearly interpolated synthetic \mstars/\lk\ values based on the stellar population models of \citet{maraston98a} from updated model-grids made available by the author\footnote{\href{http://www-astro.physics.ox.ac.uk/~maraston/Claudia's_Stellar_Population_Models.html}{http://www-astro.physics.ox.ac.uk/$\sim$maraston/Claudia's}-\\\_Stellar\_Population\_Models.html}. For typical early-type galaxies, K-band and \twomass\ \ks-band magnitudes should differ by $<$0.1~magnitudes \citep{carpenter01a}, so we were able to compare directly the synthetic K-band M/L ratios with our measured \mstars/\lk\ ratios. The predicted \mstars/\lk\ ratios are shown in Table~\ref{table_mass_to_light} for different assumptions about the stellar IMF, which is poorly-known in early-type galaxies. In this case we show predicted \mstars/\lk\ assuming a standard Salpeter IMF, and for the IMF of \citet{kroupa01a}. It is immediately clear that these ratios are very sensitive to this choice; \mstars/\lk\ is typically $\sim$50--60\% higher if the Salpeter IMF is used. Our measured \mstars/\lk\ for the NFW+stars models are typically $\sim$20\% lower than the synthetic M/L ratios, assuming the Kroupa IMF. Using the AC NFW+stars models, the discrepancy is $\sim$40\%. Assuming a Salpeter IMF, the discrepancies for both models are considerably larger. This would seem to rule out the Salpeter IMF, in agreement with the conclusions of \citet{padmanabhan04a}. The best-fitting \mvir\ and c are sensitive to \mstars/\lk. If we fix \mstars/\lk\ to the synthetic value, this essentially pushes all the galaxies, except NGC\thin 4261 (for which the measured and synthetic values are in excellent agreement) and NGC\thin 720 in the direction of the high-\mvir\ range of their confidence contours shown in Fig~\ref{fig_confidence}. For NGC\thin 720, \mvir\ is lowered and c increased. The fits are then typically much worse ($\Delta \chi^2 \sim$7--35), and the loci in the \mvir-c plane slightly more discrepant with simulations. There are a considerable number of systematic uncertainties in the computation of the synthetic M/L ratios, not the least of which is the very uncertain IMF, which could probably account for the modest discrepancy with our NFW+stars results (see \S~\ref{discussion_mass_to_light}). In the case of NGC\thin 720, the rather young age inferred for the stellar population ($\sim$3~Gyr) leads to a significantly lower synthetic \mstars/\lk\ than measured. Fitting template models to spatially-resolved spectra of this system, \citet{rembold05a} found evidence of a significant age gradient, which falls from $\sim$12~Gyr in the centre to $\sim$3~Gyr by 1~kpc. This may, therefore, represent a system in which a relatively small fraction of the stellar component, produced in a modest, recent star-formation event (``frosting'') dominates the optical line emission. In this case, the synthetic \mstars/\lk\ may be underestimated. We return to this issue in \S~\ref{discussion_mass_to_light}. \section{Systematic errors} \label{sect_systematics} \begin{deluxetable*}{lllllllllll} \tablecaption{Systematic error budget} \tabletypesize{\scriptsize} \tablehead{ \colhead{Galaxy} & \colhead{Best-fit} & \colhead{$\Delta$stat} & \colhead{$\Delta$N04} & \colhead{$\Delta$stars} & \colhead{$\Delta$bkd} & \colhead{$\Delta$asym} & \colhead{$\Delta$temp} & \colhead{$\Delta$spectra} & \colhead{$\Delta$dist} & \colhead{$\Delta$baryons} } \startdata \multicolumn{11}{c}{\mvir/$10^{12}$\msun}\\ \hline NGC720 &$6.6$ &$^{+2.4}_{-3.0}$ &$+0.06$ &$^{+0.5}_{-0.3}$ &$-1.7$ &$-0.2$ &$-0.9$ &$^{+0.7}_{-0.2}$ &$^{+0.9}_{-0.6}$ &$-0.01$ \\ NGC1407 &$16$ &$\pm6$ &$-0.1$ &$^{+1}_{-0.5}$ &$-3$ &$+3$ &$^{+1}_{-3}$ &$^{+1}_{-3}$ &$^{+3}_{-4}$ &$-0.3$ \\ NGC4125 &$6.2$ &$^{+0.8}_{-2.3}$ &$-0.2$ &$^{+1.7}_{-2.1}$ &$^{+0.09}_{-0.2}$ &$^{+0.2}_{-0.03}$ &$-1.0$ &$^{+0.5}_{-2.0}$ &$^{+0.7}_{-0.6}$ &$-0.3$ \\ NGC4261 &$67$ &$^{+41}_{-15}$ &$+2$ &$^{+15}_{-24}$ &$+9$ &$^{+5}_{-0.5}$ &$^{+0.08}_{-0.6}$ &$^{+13}_{-0.04}$ &$^{+4}_{-7}$ &$-3$ \\ NGC4472 &$33$ &$^{+6}_{-10}$ &$+0.5$ &$^{+6}_{-0.05}$ &$+2$ &$+12$ &$-4$ &$^{+8}_{-1}$ &$\pm 3$ &$-5$ \\ NGC4649 &$35$ &$^{+7}_{-13}$ &$-8$ &$^{+16}_{-3}$ &$+2$ &$-1$ &$^{+63}_{-2}$ &$^{+6}_{-2}$ &$\pm 3$ &$-19$ \\ NGC6482 &$7.1$ &$^{+4.4}_{-1.7}$ &$-0.8$ &$^{+1.9}_{-0.3}$ &$-0.1$ &$+0.8$ &$^{+0.1}_{-3.6}$ &$^{+0.3}_{-0.8}$ &$^{+0.9}_{-0.8}$ &$-0.2$ \\ \hline\multicolumn{11}{c}{c}\\ \hline NGC720 &$18$ &$^{+30}_{-8}$ &$+0.1$ &$\pm2$ &$+6$ &$+0.2$ &$+2$ &$^{+2}_{-1}$ &$\pm 2$ &$+0.03$ \\ NGC1407 &$18$ &$^{+11}_{-7}$ &$-0.8$ &$^{+2}_{-3}$ &$^{+4}_{-3}$ &$-4$ &$^{+4}_{-1}$ &$^{+6}_{-3}$ &$^{+5}_{-3}$ &$+0.02$ \\ NGC4125 &$10$ &$^{+5}_{-2}$ &$-1$ &$^{+5}_{-4}$ &$^{+0.9}_{-0.1}$ &$^{+0.05}_{-1.0}$ &$+2$ &$^{+2.}_{-0.7}$ &$^{+2}_{-1}$ &$+0.3$ \\ NGC4261 &$3.7$ &$\pm 1.7$ &$-1.4$ &$^{+3.6}_{-1.2}$ &$^{+0.10}_{-0.7}$ &$-0.3$ &$^{+0.08}_{-0.03}$ &$-1.5$ &$^{+0.8}_{-0.1}$ &$+0.08$ \\ NGC4472 &$13$ &$^{+4}_{-2}$ &$-2$ &$^{+0.4}_{-2}$ &$-0.7$ &$-5$ &$+1$ &$^{+0.2}_{-3.}$ &$\pm 2.$ &$+1.0$ \\ NGC4649 &$21$ &$^{+6}_{-3}$ &$-2$ &$^{+3}_{-6}$ &$-0.9$ &$+2$ &$^{+3}_{-4}$ &$^{+0.4}_{-3}$ &$\pm 3$ &$+7$ \\ NGC6482 &$18$ &$^{+13}_{-8}$ &$+1$ &$^{+2}_{-6}$ &$+3$ &$+6$ &$^{+51.}_{-0.4}$ &$+2$ &$\pm 3$ &$-0.7$ \\ \hline \multicolumn{11}{c}{\mstars/\lk (NFW+stars)}\\ \hline NGC720 &$0.77$ &$^{+0.52}_{-0.71}$ &$+0.17$ &$^{+0.28}_{-0.18}$ &$-0.15$ &$+0.07$ &$-0.04$ &$\pm 0.10$ &$\pm 0.16$ &$-0.001$ \\ NGC1407 &$0.52$ &$^{+0.25}_{-0.32}$ &$+0.06$ &$^{+0.71}_{-0.16}$ &$^{+0.12}_{-0.04}$ &$+0.14$ &$-0.19$ &$^{+0.09}_{-0.18}$ &$^{+0.10}_{-0.08}$ &$+0.001$ \\ NGC4125 &$0.72$ &$\pm 0.11$ &$+0.04$ &$^{+0.62}_{-0.20}$ &$^{+0.01}_{-0.03}$ &$+0.06$ &$-0.04$ &$+0.05$ &$\pm 0.15$ &$-0.003$ \\ NGC4261 &$1.2$ &$\pm 0.1$ &$+0.05$ &$^{+0.7}_{-0.8}$ &$^{+0.008}_{-0.001}$ &$^{+0.04}_{-0.05}$ &$^{+0.0009}_{-0.006}$ &$^{+0.09}_{-0.010}$ &$\pm 0.2$ &$-0.006$ \\ NGC4472 &$0.51$ &$^{+0.24}_{-0.31}$ &$+0.06$ &$^{+0.45}_{-0.06}$ &$+0.05$ &$+0.20$ &$-0.01$ &$^{+0.13}_{-0.02}$ &$^{+0.11}_{-0.09}$ &$-0.01$ \\ NGC4649 &$0.91$ &$\pm 0.13$ &$+47$ &$^{+0.80}_{-0.19}$ &$+0.02$ &$-0.06$ &$-0.16$ &$^{+0.05}_{-0.009}$ &$\pm 0.18$ &$-0.03$ \\ NGC6482 &$0.73$ &$^{+0.18}_{-0.27}$ &$+0.13$ &$^{+0.50}_{-0.04}$ &$-0.06$ &$-0.15$ &$-0.51$ &$-0.05$ &$\pm 0.15$ &$-0.01$ \\ \hline \multicolumn{11}{c}{\mstars/\lk (AC NFW+stars)}\\ \hline NGC720 &$0.54$ &$^{+0.42}_{-0.48}$ &$-0.12$ &$\pm 0.16$ &$-0.14$ &$+0.06$ &$-0.05$ &$^{+0.03}_{-0.10}$ &$^{+0.12}_{-0.09}$ &$-0.003$ \\ NGC1407 &$0.35$ &$\pm 0.25$ &$-0.06$ &$^{+0.41}_{-0.08}$ &$^{+0.12}_{-0.05}$ &$+0.12$ &$-0.14$ &$^{+0.07}_{-0.13}$ &$^{+0.07}_{-0.05}$ &$-0.001$ \\ NGC4125 &$0.53$ &$\pm 0.11$ &$-0.04$ &$^{+0.46}_{-0.14}$ &$^{+0.01}_{-0.03}$ &$+0.06$ &$-0.04$ &$+0.04$ &$^{+0.11}_{-0.09}$ &$-0.002$ \\ NGC4261 &$1.0$ &$\pm 0.1$ &$+0.02$ &$\pm0.6$ &$^{+0.02}_{-0.004}$ &$^{+0.03}_{-0.04}$ &$-0.006$ &$^{+0.1}_{-0.003}$ &$\pm 0.2$ &$-0.009$ \\ NGC4472 &$0.36$ &$\pm0.1$ &$-0.04$ &$^{+0.27}_{-0.03}$ &$+0.04$ &$+0.17$ &$-0.02$ &$^{+0.09}_{-0.01}$ &$^{+0.08}_{-0.06}$ &$-0.010$ \\ NGC4649 &$0.65$ &$\pm 0.12$ &$-0.05$ &$^{+0.57}_{-0.12}$ &$+0.02$ &$-0.06$ &$-0.13$ &$^{+0.06}_{-0.008}$ &$\pm 0.13$ &$-0.05$ \\ NGC6482 &$0.52$ &$^{+0.19}_{-0.23}$ &$-0.07$ &$^{+0.37}_{-0.03}$ &$-0.05$ &$-0.13$ &$^{+0.005}_{-0.03}$ &$-0.04$ &$^{+0.11}_{-0.09}$ &$-0.01$ \\ \enddata \tablecomments{The estimated error-budget for each of the galaxies. Excepting the statistical error, these values estimate a likely upper bound on the sensitivity of the (best fit) value of each parameter to various data-analysis choices, and should {\em not} be added in quadrature with the statistical errors. The systematic uncertainties on \mvir\ and c are estimated for the NFW+stars model. In addition to the best-fit values, we show the 90\% confidence interval for each parameter ($\Delta$stat). We also show estimated upper-limits on the systematics likely to arise by making various changes to our default analysis choices. This includes adopting the N04 DM model ($\Delta$N04), varying the shape of the stellar mass component ($\Delta$stars), varying the background ($\Delta$bkd), excluding data in the vicinity of asymmetries ($\Delta$asym), adopting alternative temperature models ($\Delta$temp), changing spectral analysis choices ($\Delta$spectra), varying the distance ($\Delta$dist) or assuming that all of the stellar baryons are in the central galaxy ($\Delta$baryons).} \label{table_syserr} \end{deluxetable*} In this section we address the sensitivity of our results to various data-analysis choices which were made. An estimated upper limit on the sensitivity of our results to these choices is shown in Table~\ref{table_syserr}. These numbers reflect the sensitivity in the best-fit parameter to each potential source of systematic error, and we stress they should {\em not} be added in quadrature with the statistical errors. We outline in detail below how each of these systematics were estimated. Those readers uninterested in the technical details of our analysis may like to proceed directly to \S~\ref{sect_discussion}. \subsection{DM profile shape} \label{sect_n04} As discussed above, we experimented with replacing the NFW model by the revised N04 model, which is less cuspy. We caution that the \mvir-c relation was derived assuming NFW. We fixed the $\alpha$ parameter for this model to 0.17, the mean value determined from simulations since the inner slope of the DM halo is degenerate to some degree with the stellar mass. The quality of the fits (N04+stars, AC N04+stars) were typically similar to those using the simple NFW profile. There were some slight systematic differences in the inferred \mvir\ and c as compared to NFW. It is interesting to note that this model, which is less cuspy than NFW, gave slightly larger \mstars/\lk, although not sufficiently to bring our measurements completely into agreement with the synthetic estimates. For the adiabatically-compressed N04 model, \mstars/\lk\ did not increase, but this is unsurprising since the stellar component significantly modifies the shape of the inner DM halo in this model. We note that the predicted typical inner slope for \lcdm\ halos is still under debate. If, instead of the N04 profile we had adopted the cuspier profile of \citet{diemand05a}, then the resultant \mstars/\lk would have been even smaller, and in worse agreement with stellar population models. \subsection{Shape of the stellar potential} \label{sect_syserr_potential} To account for the stellar component, we adopted an H90 model, the effective radius of the model being fixed to that determined by \twomass. However, it is not entirely clear that the H90 model is an adequate description of the stellar mass. There are some deviations between H90 and the de Vaucouleurs model fitted as the {\em de facto} standard to the optical light profiles of elliptical galaxies, particularly in the critical central regions. Furthermore, the K-band light profiles of elliptical galaxies may, in fact, be better described by the \sersic\ profile \citep[\eg][]{brown03a}. To investigate the sensitivity of our results to the H90 assumption, therefore, we experimented with adopting a \sersic\ stellar mass potential \citep[\eg][]{prugniel97a}. To determine the two parameters of this model (the \sersic\ index and half-light radius) we obtained the \ks-band \twomass\ images of each galaxy from \ned, and fitted the surface brightness profiles using dedicated software. A \sersic\ model fitted the \ks-band light profile of each galaxy in the radial range 5\arcsec--3\arcmin\ reasonably well. The fitted profiles tended to be slightly more centrally peaked than H90, which resulted primarily in slightly {\em lower} inferred \mstars/\lk\ ratios when adopted as mass models. We also experimented with replacing the H90 model with a de Vaucouleurs model \citep{mellier87a}, and adopting \reff\ values from \citet{pahre99a}. Elliptical galaxies exhibit radial colour gradients, which may reflect gradients in the metallicity or age of the stellar population (see discussion in \S~\ref{discussion_mass_to_light}). These may therefore imply a radial gradient in the stellar M/L ratio. It is beyond the scope of this present work to take such a gradient into account. However, we investigated the sensitivity of our results to the precise shape of the optical light profile we adopted by experimenting with replacing the K-band \reff\ for each galaxy with the (typically larger) B-band value listed in RC3. For NGC\thin 6482, for which \reff\ is not listed in RC3, we simply increased \reff\ by 50\%. \subsection{Background subtraction} \label{sect_systematics_bkd} One of the major potential sources of systematic uncertainty in measuring the mass profiles of galaxies is the background subtraction technique. This is especially important in the low surface-brightness regime at large radii in our galaxies. In order to estimate the likely magnitude of uncertainty arising from our modelling, when initially fitting the background components (\S~\ref{sect_bkd}) we artificially adjusted the slope of the instrumental background components, which dominate at high energy, to the limits of their 90\% confidence regions, refitting the other components and then refitted all the spectra with these revised background models. \subsection{X-ray asymmetries} \label{sect_asymmetry} We note that there are some low-level asymmetries in the X-ray surface brightness profiles (\S~\ref{sect_imaging}). In order to assess the potential impact of these features, we experimented with excluding or including the features. In particular, we tried excluding data from the vicinity of the jet and AGN in NGC\thin 4261. We also excluded data from an off-axis X-ray asymmetry in NGC\thin 4125 and excluded data for NGC\thin 4472 outside 6\arcmin, where \citet{irwin96} pointed out that the X-ray data become asymmetric. These choices did not dramatically affect our results, indicating that these features did not indicate a significant violation of hydrostatic equilibrium, at least in an azimuthally-averaged sense. To gain an insight into possible asymmetries in other sources, we tried re-extracting all our spectra, and re-deriving the mass profiles, from suitably-oriented semi-annuli (thereby excluding one half of the emission from each system). \subsection{Temperature profile} In principle multiple temperature profiles may be able to fit the same data adequately well but give rise to slightly different global halo parameters. In particular our constraints upon \fbaryons, the computation of which requires the extrapolation of the density (and hence temperature) profiles to large radius, may make \mvir\ and c sensitive to this effect. To test this, we experimented with cycling through each of our adopted temperature profiles (eq~\ref{eqn_trise2}--\ref{eqn_twophase2}). Provided the fits were of comparable quality to our preferred choice, the impact on the best-fit parameters reflect the systematic uncertainty in this choice. Furthermore, we also experimented with excluding the central bin from the temperature profiles of NGC\thin 1407 and NGC\thin 4649, which may indicate a central disturbance (although there is no obvious X-ray morphological disturbance in this region). These choices did not strongly affect our results. \subsection{Spectral-fitting choices} \label{sect_systematics_spectra} A variety of choices are made in the spectral-fitting, each of which can affect, to some degree, the inferred \rhog\ and temperature in each radial bin. A thorough discussion of these effects is given in \citet{humphrey05a}. {\em Column density.} In order to take account of possible local deviations in the line-of-sight \nh\ from the value of \citet{dickey90}, we experimented with allowing \nh\ to vary by $\pm$25\%. {\em Bandwidth.} To estimate the impact of the bandwidth on our fits, we experimented with fitting the data in the energy ranges 0.7--7.0~keV, 0.5--2.0~keV and 0.4--7.0~keV, in addition to our preferred choice 0.5--7.0~keV. {\em Plasma code.} There are some uncertainties in the correct modelling of the individual emission lines, in particular those of Fe. This can systematically lead to differences in the inferred temperature and density depending on choice of plasma code. We therefore experimented with replacing the APEC model with the MEKAL plasma model. {\em Unresolved source component.} We included a 7.3~keV bremsstrahlung component to account for unresolved point sources within \dtwentyfive. This model is generally a good fit to the resolved point sources in early-type galaxies, but is an empirical result which may not be appropriate to model all unresolved sources in all early-type galaxies. We therefore tested the sensitivity of our results to this model, by replacing the bremsstrahlung component with a simple power law (with $\Gamma=$1.5) or varying the temperature of the component by $\pm$25\% % \subsection{Distance uncertainty} The estimated distance to the object enters into our mass determination (Eq~\ref{eqn_hydrostatic_rho}) primarily through the impact on the radial scale. To assess its impact on our fitting, we experimented by varying the distance by $\pm$20\%. \subsection{Stellar baryon fraction} \label{sect_systematics_baryon_fraction} In our analysis, we restricted \fbaryons\ to enable interesting constraints on \mvir\ and c. For the stellar contribution to the baryon fraction, we assumed that mass is divided among group members following the B-band light, which is not formally correct since \mstars/L ratios are very sensitive to the age of the stellar population. To estimate how much impact this makes to our fits, we experimented with assuming that all the stellar mass is in the central galaxy, which should place an upper limit on the uncertainty arising from this choice. \section{Discussion} \label{sect_discussion} \subsection{Hydrostatic equilibrium} Our fit results provide strong evidence that the gas is in hydrostatic equilibrium in these systems. Despite highly nontrivial temperature and density profiles, we were able to recover smooth mass profiles in remarkably good agreement with expectation for these systems, using two complementary techniques. If the gas is significantly out of hydrostatic equilibrium, this would represent a remarkable ``conspiracy'' between the density and temperature profiles. It is unsurprising that the gas is close to hydrostatic equilibrium in these systems, since we took care to choose objects with relaxed X-ray morphology. Based on N-body/ hydrodynamical analysis, X-ray measurements are expected to give reliable constraints on the DM in systems without obvious substructure \citep{buote95a}. Further support for hydrostatic equilibrium is provided by the general agreement between our measured \mstars/\lk\ ratios and those predicted by SSP models, coupled with the agreement between the measured \mvir-c relation and that expected. Similarly a comparison between our results and masses determined from stellar dynamics provides even more evidence that our measured mass profiles are reliable. Dynamically-determined \mgrav/\lb\ within the B-band \reff\ are typically found to be $\sim$4--10 \citep{gerhard01a,trujillo04a}. We found \mgrav/\lb\ within the B-band \reff\ (taken from RC3 or \citealt{faber89}) for our systems ranged from $\sim$3 to $\sim$8, in good agreement with this result. For NGC\thin 4649, outside \reff\ there is excellent agreement between our measured \mgrav/L profile and that obtained from globular cluster kinematics, although at small radii the X-ray data lie $\sim$30\% lower (K.\ Gebhardt et al, in preparation). \citet{vandermarel91a} constructed stellar kinematical models for 5 galaxies in our sample (NGC\thin 720, NGC\thin 1407 and NGC\thin 4261, NGC\thin 4472 and NGC\thin 4649), under the assumption of a constant M/L profile. Strictly speaking a direct comparison cannot be made between their \mgrav/\lb\ measurements and our results since our data indicate this assumption is incorrect. However, if we simply assume that these M/L ratios represent those integrated out to \reff, the X-ray inferred masses vary from $\sim$40\% lower to $\sim$10\% higher than those from kinematics. \citet{kronawitter00a} report \mgrav/\lb$\sim$8$\pm1.5$ for NGC\thin 4472 within $\sim$50\arcsec, at which radius our X-ray determined value is $\sim$50\% lower. The discrepancies between the X-ray and dynamical masses are only modest (the X-ray mass being on average $\sim$20\% lower), indicating that the data must be close to hydrostatic equilibrium. Turbulence is expected to contribute only $\sim$10\% pressure support in clusters, which are believed to be more turbulent than galaxies, \citep{rasia06a}. Therefore, on a case-by-case basis, the observed differences are most likely a manifestation of the mass-anisotropy degeneracy \citep[\eg][]{dekel05a}. \subsection{Mass profiles} We obtained detailed mass profiles for 3 galaxies and 4 group-scale systems, out to $\sim$10\reff. The data clearly show M/L profiles which are $\sim$flat within \reff\ and rise considerably outside this range. This confirms the presence of substantial DM in at least some early-type galaxies and indicates that a stellar mass component dominates within $\sim$\reff. This is consistent with studies of stellar kinematics and similar to the mass decomposition analysis of \citet{brighenti97a}. The data are well-fitted by a model comprising a stellar mass (H90) component and an NFW DM profile. Omitting the stellar mass component led to systematically poorer fits, smaller \mvir\ and vastly larger c ($\gg$20), which are inconsistent with the predictions of \lcdm. This effect is easy to understand--- if we add a compact stellar mass component to an (extended) NFW profile, we increase the mass in the core which, by definition, makes the halo more concentrated. However, it is not entirely clear whether this effect, pointed out by \citet{mamon05a}, can completely account for the significantly steeper \mvir-c relation found by \citet{sato00a}. Based on our analysis of group-scale halos \citep{gastaldello06a} we found that the inclusion of the stellar mass component does not have a strong effect on c in most systems with \mvir \gtsim 2$\times 10^{13}$\msun, provided the data are fitted to a sufficiently large fraction of \rvir. The data did not allow us to distinguish statistically between the simple NFW+stars model and scenarios in which the DM halo experiences adiabatic compression due to star formation (however, see \S~\ref{discussion_mass_to_light}), or the NFW profile was replaced with the alternative N04 profile. Comparing our inferred \mvir\ and c to the predictions of \lcdm\ we find general agreement. There is some evidence, however, that the concentrations are systematically higher than one would expect, although the error-bars are typically large. Such a trend is also seen in our analysis of groups \citep{gastaldello06a}. Whilst the slope of the \mvir-c relation therefore implied by our data is difficult to explain by varying the cosmological parameters within reasonable limits \citep{buote06b}, we suspect that the discrepancy can be resolved by taking into account the selection function of our galaxies. Our data were not selected in a statistically complete manner and, by choosing objects with relatively relaxed X-ray morphologies we are probably selecting objects which have not had a recent major merger. This systematically biases us towards early-forming, hence higher concentration halos. In fact, it is striking that all three {\em de facto} galaxies in our sample are relatively isolated systems (\S~\ref{discussion_groups}). Such systems preferentially might be expected to occupy high-c halos \citep{zentner05a}, which does appear to be the case for 2 out of 3 of the galaxies. We will return to these issues in detail in \citet{buote06b}. \subsection{Galaxies, Groups and Fossil Groups} \label{discussion_groups} All three of the lowest-mass systems in our sample are very isolated optically. NGC\thin 6482 matches the isolation criteria adopted to identify so-called ``fossil groups'' \citep{khosroshahi04a}. NGC\thin 4125 and NGC\thin 720 are both listed as ``groups'' in G93, but closer inspection actually reveals they are also very isolated. Excepting the central galaxy, only one of the putative members of the NGC\thin 720 ``group'' listed in the G93 catalogue \citep[which omits the dwarf galaxy population studied by][]{dressler86a}, actually lies within the projected \rvir\ (but outside 0.75$\times$\rvir) and it is 2.4 magnitudes fainter in B than the central galaxy. \citeauthor{dressler86a} remarked upon the optical isolation of this galaxy. Of the two putative companion galaxies to NGC\thin 4125 given in G93 which lie within the projected \rvir\ (but outside 0.67$\times$\rvir), both are much fainter (by 2.3 and 3.9 magnitudes, respectively) in B than the central galaxy. In contrast, the four remaining systems in our sample are much less optically isolated. \citet{schindler99a} show the clear over-density of early-type galaxies around NGC\thin 4649 and NGC\thin 4472, and almost 60 group members are associated with these systems by G93. \citet{gould93a} identified at least 10 members of the NGC\thin 1407 group, from the dynamics of which he inferred a mass broadly consistent with our measured \mvir\ (\S~\ref{sect_ngc1407}). \citet{helsdon03a} report 57 galaxies associated with the NGC\thin 4261 group within $\sim$1~Mpc projected radius, which is consistent with our measured \rvir. Rather than an isolated galaxy \citet{khosroshahi04a} identify NGC\thin 6482 as a ``fossil group''. Fossil groups are group-sized X-ray halos centred on essentially a single elliptical galaxy \citep{ponman94,vikhlinin99a,jones03}. The typical interpretation of these objects is groups in which all of the \lstar\ members have merged. Confusingly, using almost the same selection criteria, \citet{osullivan04b} classify the galaxy NGC\thin 4555 as an ``isolated elliptical galaxy'' and posit a very different formation scenario. This object appears to be more massive than NGC\thin 6482; the authors found \mgrav $\sim 3\times 10^{12}$\msun\ within 60~kpc which, assuming an NFW profile with c$=$15 would imply \mvir $\sim 2\times 10^{13}$\msun. Nonetheless, both of these systems have more in common (both optically and in the X-ray) with each other, and the other isolated ellipticals in our sample, than, for example, the massive (\mvir\gtsim$10^{14}$\msun), hotter (kT$\sim$2~keV) fossil groups considered by \citet{vikhlinin99a}. We suspect that the distinction made between ``isolated elliptical'' and ``fossil group'' for these two systems is largely semantic, and consider NGC\thin 6482 more properly an isolated galaxy, too. The clear division in the galaxy content of our sample clearly lends itself to the nomenclature ``galaxies'' for the three lowest-mass systems, and ``groups'' for the others. Strikingly, this separation between galaxies and groups also appears consistent with a difference in temperature profiles (\S~\ref{discussion_temp}). That this distinction appears commensurate with \mvir$\sim 10^{13}$\msun\ is suggestive that this mass-scale may be a useful yard-stick with which to compare to other systems. The error-bars on our mass estimates are sufficiently large that the 90\% confidence regions of several of the objects (notably NGC\thin 720, NGC\thin 6482 and NGC\thin 1407) actually straddle $10^{13}$\msun. However, it is clear that {\em on average}, the systems with \mvir \ltsim $10^{13}$\msun\ are galaxies. We note that the \mvir\ adopted here is that {\em before} any tidal truncation which is almost certainly occurring as NGC\thin 4472 and NGC\thin 4649 merge with Virgo (their untruncated \rvir\ would stretch much of the distance to M\thin 87). \mvir\ does not exactly correlate with formation epoch, so that lower-mass halos may still be in the process of forming (hence contain multiple galaxies of similar magnitude), and more massive halos may contain single, dominant ellipticals (fossil groups). Nonetheless, classifying halos primarily on the basis of \mvir\ provides a straightforward way to locate them in the formation hierarchy. Traditionally, galaxy-like and group-like systems are distinguished on the basis of local over-densities of galaxies. However, placing optically-identified groups into a cosmological context requires a firm understanding not only of the formation of DM halos but also how galaxies populate them, which is much less well-understood \citep[\eg][]{kravtsov04a}. This problem is compounded by the difficulties faced by optical group-finding algorithms in identifying very poor groups \citep[\eg][]{gerke05a}. Not only can a significant fraction of putative groups be chance superpositions of galaxies, particularly along filaments, but adjacent groups can be merged, such as happened for NGC\thin 4649 and NGC\thin 4472 in G93. If there are only a few identified members, small-number statistics and the treatment of interlopers can affect their interpretation \citep[\eg][]{gould93a}. To complicate matters further, some authors refer to {\em any} over-density of galaxies as a group, even a Milky Way-sized galaxy and its dwarf satellites \citep[\eg][]{tully05a}. \subsection{Stellar Mass-to-Light Ratios} \label{discussion_mass_to_light} Comparing our measured stellar M/L ratios to the predictions of simple stellar population (SSP) models, we found reasonable agreement provided one assumes a \citet{kroupa01a} IMF. There is modest disagreement, even when the less-cuspy N04 DM model was adopted. Considering the uncertainties in the SSP modelling (discussed below), however, we believe this discrepancy is not significant. If we allowed the DM profile to be modified by adiabatic compression, we obtained substantially smaller \mstars/\lk\ values from our data, (since it increases the cuspiness of the halo) which are more discrepant with the SSP models. This result casts doubt on AC being as significant an effect as currently modelled. However, the data alone did not allow us statistically to distinguish between the NFW+stars and AC NFW+stars models. Nonetheless, this result is joining a growing body of literature which similarly calls into question whether AC operates as predicted \citep{zappacosta06a,kassin06a,sand04a}. There are a number of major uncertainties in the computation of the stellar mass-to-light ratios from the SSP models. Specifically, the results are very dependent upon the assumed IMF, which is not confidently known in early-type galaxies. Furthermore there is some evidence that early-type galaxies frequently contain multiple stellar populations of different ages, including a significant young population \citep[\eg][]{rembold05a,nolan06a}. Depending on the mass fraction of the young component, this may substantially reduce \mstars/\lk\ in the galaxy, hence possibly reconciling the data and the AC NFW+stars model. A small amount of star formation may also give rise to a population of stars which can dominate the light in the galaxy core, giving rise to significantly lower synthetic \mstars/\lk\ than measured. This may be the case in NGC\thin 720 (see \S~\ref{sect_mass_to_light}). More problematically, there are known to be significant abundance, or possible age, gradients in the stellar populations of early-type galaxies \citep[\eg][]{trager00a,kobayashi99a,rembold05a}, which would translate into stellar \mstars/\lk\ gradients. Our simple modelling did not allow us to account for such an effect {\em per se}. Although we suspect that such gradients will primarily lead to a \mstars/\lk\ value which reflects an average for the galaxy, \mstars/\lk\ does depend to some extent upon the shape of the assumed stellar potential. Properly taking account of this effect is beyond the scope of this present work, but may bring the synthetic M/L ratios and our results into better agreement. Clearly this is only one of a number of other systematic effects which may also reconcile the slight discrepancy (Table~\ref{table_syserr}). \subsection{Baryon fractions} An interesting result from our analysis is that these systems, despite having masses \gtsim 5$\times 10^{12}$\msun, do not appear in general to be baryonically closed. To some extent this trend was enforced by applying Eq~\ref{eqn_baryons} to constrain the data. However, the excellent fits we obtained by this method, in conjunction with the good agreement between the measured \mvir-c relation and the predictions of \lcdm\ and, crucially, our measurements at the group scale (which do not employ this restriction: \citealt{gastaldello06a}), indicate that the inferred \fbaryons ($\sim$0.04--0.09; Table~\ref{table_results}) are accurate. Furthermore, if we relaxed this constraint and instead restricted \fbaryons\ to a finite range, we also found that the data tended to favour modest values of \fbaryons. In particular, for any given system, the measured \mvir\ and \fbaryons\ were strongly anti-correlated, so that our upper \mvir\ constraint is in part imposed by the {\em lower} limit we place on \fbaryons. Given the shapes of the \mvir-c contours (Fig~\ref{fig_confidence}), it is clear that good agreement with the \mvir-c relation predicted from simulations tends, therefore, to require rather modest values of \fbaryons. This would suggest that strong feedback plays an important role in the evolution of these objects. \subsection{Temperature profiles} \label{discussion_temp} By inspection of the temperature profiles (Fig~\ref{fig_temp}) it is immediately clear that, for all of the galaxy-scale systems in our sample the temperature profiles have negative gradients. In contrast the group-scale objects have positive temperature gradients, similar to observations of other X-ray bright groups and clusters \citep{gastaldello06a,vikhlinin05b,piffaretti05a}. This radical difference in the temperature profiles seems consistent with our division of galaxies and groups at \mvir $\sim 10^{13}$\msun. The origin of this distinct demarcation between objects around $10^{13}$\msun\ is unclear, however. Negative temperature gradients are expected for isolated galaxies containing relatively cool gas, such as that arising from stellar mass-loss. In the deep stellar potential well, compressive heating of the gradually inflowing gas can dominate over radiative cooling to produce a negative temperature slope. In contrast, if hotter ($\sim$1--2~keV) baryons are allowed to flow in, radiative cooling dominates to produce a positive temperature gradient \citep{mathews03a}. It is by no means clear, however, why the hot baryons appear to be present only in the systems with \mvir\gtsim $10^{13}$\msun. One possibility is the local environment; all of the galaxy-scale objects are rather isolated, whereas the groups NGC\thin 4472 and NGC\thin 4649, in particular, are found in a relatively dense cluster environment, which could provide a reservoir of hot baryons. However, such an explanation cannot easily account for the positive temperature gradient in NGC\thin 1407, which is comparatively isolated, or the isolated system NGC\thin 4555, which appears only slightly more massive than our galaxies. It is possible that selection effects may have played some role in the bimodal temperature profile behaviour, since both NGC\thin 4125 and NGC\thin 6482 are classified in \ned\ as LINERS, and NGC\thin 720 has a dominant young stellar population (Appendix~\ref{sect_stars}). However, none of these systems show strong X-ray morphology disturbances in the core, which might indicate a substantial energy input from star-formation or AGN activity. In any case, the cooling time in the core of NGC\thin 720 is only $\sim$200~Myr, substantially less than the implied time since the last major burst of star formation, and so the negative temperature gradient cannot simply be related to energy injection during a starburst. Furthermore, at least two of the group-scale systems also harbour AGN and do not show obvious negative temperature gradients in the core. Another example of an object we believe to be a galaxy (rather than a group) which exhibits a negative temperature gradient is the S0 NGC\thin 1332 \citep{humphrey04b}. A possible counter-example to this trend might by the ``isolated elliptical galaxy'' NGC\thin 4555, which exhibits a temperature profile akin to the groups in our sample \citep{osullivan04b}. However, as we discuss in \S~\ref{discussion_groups}, this probably has comparable \mvir\ to the groups. Another intriguing feature of two of the group scale objects is a central temperature peak, similar to a feature we found in the cluster A\thin 644 \citep{buote05a}. In that system, we found a significant offset between the X-ray centroid and the emission peak in an otherwise fairly relaxed object. We suggested that both of these features may be related to the cD ``sloshing'' in the potential well of the cluster, which is relaxing following disturbance by, for example, a merger. We do not find obvious evidence of a similar offset in either NGC\thin 1407 or NGC\thin 4649. However, these groups may be in a comparably more relaxed (evolved) state than A\thin 644. Alternatively, the central peaks may be related to past AGN activity heating the gas in the core of the galaxies, from which the system has had time to relax dynamically but not cool completely. \subsection{Is NGC\thin 1407 a ``dark group''?} \label{sect_ngc1407} Based on the group member dispersion velocity \citet{gould93a} suggested that NGC\thin 1407 may lie in a massive (\gtsim a few $\times 10^{13}$\msun) DM halo. Although such a conclusion was strongly dependent on the association of the galaxy NGC\thin 1400, which exhibits a large peculiar velocity, with the group, we can now confirm the presence of a substantial DM halo around this system. Both the temperature profile and our best-fit mass are similar to the bright X-ray group NGC\thin 5044 \citep{buote06a}, and yet it is almost 2 orders of magnitude fainter in \lx. NGC\thin 5044 appears to be close to baryonic closure \citep{mathews05a}, and so has likely retained most of its large gaseous halo. On the other hand NGC\thin 1407 is not baryonically closed (we estimate \fbaryons$\simeq$0.06) and so the loss of much of its hot gas envelope easily explains its lower \lx/\lb. Since the masses of the two systems are not considerably different, this points to substantial variation in the evolutionary history of these two groups. In particular, feedback may have operated more efficiently in evacuating the gas from NGC\thin 1407. \citeauthor{gould93a}'s preferred mass estimate ($\sim10^{14}$\msun) would imply a remarkably high M/L ratio for the system (\mvir/\lb$\sim$900\msun/\lsun), making NGC\thin 1407 a bona fide ``dark group''. The existence of such an object would provide a valuable insight into the process of star formation in DM halos, as it would imply star formation was somehow inhibited in that system. This mass estimate is, however, considerably larger than our preferred value $\sim 1.5 \times 10^{13}$\msun, which implies a more modest M/L ratio (\mvir/\lb$\sim$140\msun/\lsun). To some extent, though, our constraint on \fbaryons, which was necessary to obtain interesting \mvir\ constraints, has probably enforced this behaviour. Such a restriction may not be valid in a system with an unusual star-formation history and so we experimented with freeing \fbaryons. To enable \mvir\ to be constrained, we restricted c to lie on the best-fit \mvir-c relation found by \citet{bullock01a}. The best-fitting mass, \mvir$=(9.7^{+17.8}_{-6.2})\times 10^{13}$\msun, was in good agreement with \citet{gould93a}'s values, but implies a baryon fraction only of $\sim$0.003. Since this fit was statistically indistinguishable from the preferred model, we cannot determine which mass estimate is more likely. \section{Summary} Using \chandra\ we have obtained detailed mass profiles centred on 7 elliptical galaxies, of which 3 were found to have {\em de facto} galaxy-scale halos, with \mvir $< 10^{13}$\msun, and 4 had group-scale ($10^{13}$\msun$<$\mvir$< 10^{14}$\msun) halos. These represent the best available data for nearby objects with comparable \lx. In summary: \begin{enumerate} \item The M/L ratio profiles were $\sim$flat within \reff\ and rose sharply outside this region, indicating substantial DM in all 7 systems. \item The data were well-described by a two component model, comprising an NFW potential for the DM and a H90 stellar mass model. We were not able statistically to distinguish between this scenario and one in which the DM profile was modified by ``adiabatic compression'' due to baryonic infall. Similarly, we could not distinguish between the NFW and the revised N04 DM halo profiles. \item The distribution of the galaxies in the \mvir-c plane was in broad agreement with the predictions of \lcdm, although with a slight trend toward more concentrated halos, in good agreement with our modelling of X-ray bright groups and poor clusters \citep{gastaldello06a}. This probably represents a galaxy selection bias to earlier-forming systems, and we will discuss how we might account for it in \citet{buote06b}. Allowing AC to modify the shape of the DM halo did not appreciably affect the \mvir-c relation. \item Omitting the stellar mass component resulted in systematically poorer fits, smaller \mvir\ and unphysically large c, confirming the conclusions of \citet{mamon05a}. This may explain very large values of c found by some previous X-ray observers \citep[\eg][]{sato00a,khosroshahi04a}. \item For the NFW+stars model, \mstars/\lk\ was found to be in approximate agreement with the predictions of simple stellar population synthesis models, assuming a \citet{kroupa01a} IMF. The AC NFW+stars models have significantly lower \mstars/\lk\ which seems to cast doubt on the AC scenario, although this conclusion is sensitive to the considerable uncertainties in the theoretical modelling. \item Despite having \mvir \gtsim 5$\times 10^{12}$\msun, typically \fbaryons $\sim$0.04--0.09 for each galaxy, implying that feedback has played an important role in the evolution of these systems. \item The temperature profiles of the galaxy-scale systems all exhibited negative radial gradients, whereas the group-scale objects exhibited positive gradients, similar to the ``Universal'' temperature profiles being found in other X-ray bright groups and clusters. This implies a strict line of demarcation between systems at \mvir $\sim 10^{13}$\msun. \item In two of the groups, we found central temperature peaks, similar to that found in the cluster A\thin 644 \citep{buote05a}, but no obvious central disturbances in X-ray morphology. This may relate to past AGN activity, following which the heated gas in the core of the galaxy has relaxed but not cooled. \item We confirm the suggestion of \citet{gould93a} that the elliptical galaxy NGC\thin 1407 lies at the centre of a massive DM halo, possibly making it a ``dark group'' with an unusually large M/L. Our best-fitting \mvir\ is considerably lower than that of \citeauthor{gould93a}, implying M/L more consistent with normal groups. Nonetheless, if we relax the assumptions of our modelling very large masses (\mvir$\sim 10^{14}$\msun) are allowed. \end{enumerate} \begin{acknowledgements} We would like to thank Oleg Gnedin for making available his adiabatic compression code. We would also like to thank Karl Gebhardt for communicating with us results from his paper in preparation. We thank Louisa Nolan for interesting discussions on the stellar populations of galaxies. This research has made use of data obtained from the High Energy Astrophysics Science Archive Research Center (HEASARC), provided by NASA's Goddard Space Flight Center. This research has also made use of the NASA/IPAC Extragalactic Database (\ned) which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with NASA. In addition, this work also made use of the HyperLEDA database (\href{http://leda.univ-lyon1.fr}{http://leda.univ-lyon1.fr}). Support for this work was provided by NASA under grant NNG04GE76G issued through the Office of Space Sciences Long-Term Space Astrophysics Program. \end{acknowledgements} \appendix \section{Stellar population parameters} \label{sect_stars} \begin{deluxetable*}{lllllll} \tablecaption{Stellar population parameters\label{table_ssp}} \tabletypesize{\footnotesize} \tablehead{ \colhead{Galaxy} & \colhead{indices} & \colhead{ref.} & \colhead{${\rm [\alpha/Fe]}$} & \colhead{age} & \colhead{${\rm [Z/H]_0}$} & \colhead{$\rm <[Z/H]>$} \\ \colhead{} & \colhead{} & \colhead{} & \colhead{} & \colhead{(Gyr)} & \colhead{} & \colhead{}} \startdata NGC\thin720$^\dagger$ & H$\beta$, Mgb, Fe5270, Fe5335 & 2 & 0.37$\pm0.05$ & 2.9$^{+1.3}_{-0.3}$ & 0.65$\pm$0.13& 0.48$\pm$0.18 \\ NGC\thin 1407$^\dagger$ & H$\beta$, Mgb, Fe5270, Fe5335 & 1 & 0.33$\pm 0.02$ & 12$\pm2$ & 0.35$\pm0.06$ & 0.08$\pm0.06$\ddag \\ NGC\thin 4125 & H$\beta$, Mgb, Fe5270, Fe5335 & 3 & 0.33$\pm 0.16$ & 13$\pm8$ & 0.16$\pm$0.25 & -0.11$\pm$0.25\ddag \\ NGC\thin 4261 & H$\beta$, Mgb, Fe5270, Fe5335 & 2 & 0.25$\pm 0.02$ & 15$\pm1$ & 0.30$\pm0.03$ & -0.03$\pm$0.10 \\ NGC\thin 4472$^\dagger$ & H$\beta$, Mgb, Fe5270, Fe5335 & 2 & 0.25$\pm0.03$ & 9$\pm2$ & 0.36$\pm0.05$ & 0.17$\pm$0.12 \\ NGC\thin 4649 & H$\beta$, Mgb, Fe5335 & 2 & 0.25$\pm0.02$ & 13$\pm2$ & 0.41$\pm0.04$ & 0.23$\pm$0.15 \\ NGC\thin 6482 & Mgb, Fe5270, Fe5335 & 3 & 0.30$\pm0.15$ & 12 & 0.28$\pm0.15$ & 0.06$\pm0.15$\ddag \\ \enddata \tablecomments{Stellar population parameters determined from Lick index fitting. The indices used in fitting are listed (indices), as is the reference (ref) whence they were taken. Those mean stellar abundances (${\rm <[Z/H]>}$) marked \ddag\ were estimated from the central abundance (${\rm [Z/H]_0}$) adopting the mean abundance gradient ${\rm [Z/H]_0}-{\rm <[Z/H]>}=0.27$\citep[see][]{humphrey05a}. Where no error-bar is given, the parameter was frozen. Table references: 1--- \citet{beuing02a}, 2--- \citet{trager00a}, 3--- \citet{trager98a} Results for galaxies marked $^\dagger$ were taken from \citet{humphrey05a} } \end{deluxetable*} The mass-to-light ratio of a stellar population is dependent upon both the age and the metal abundance ([Z/H]) of the stars. To estimate these quantities we searched the literature to obtain Lick/IDS indices for each galaxy, which we fitted with the simple stellar population (SSP) models of \citet{thomas03a}, using the technique outlined in \citet{humphrey05a}. Briefly, we constructed a model by linearly interpolating the SSP models as a function of stellar age, metallicity and $\alpha$-element to Fe ratio, which was then fitted {\em via} a $\chi^2$ minimization technique to those indices shown in Table~\ref{table_ssp}. \citet{trager00a} provided indices measured in two apertures, which enabled us to take account of any abundance gradients, as outlined in \citet{humphrey05a}. Where only a central Lick index was available, we estimated the total emission-weighted abundance by correcting the central metallicity by -0.27 dex. We did not attempt to take account of possible age gradients. The results, including the Lick indices adopted and the reference whence the indices were obtained, are shown in Table~\ref{table_ssp}. This method implicitly assumes that all the stars were created in a single burst of star formation, which may be over-simplistic if there are, in fact, multiple bursts of star formation in early-type galaxies \citep[\eg][]{rembold05a,nolan06a}. We note that we were not able to obtain acceptable solutions for NGC\thin 6482 if we used the H$\beta$ index, which is the most sensitive age indicator. This galaxy is classified in \ned\ as a LINER and is rather blue for an old stellar population (${\rm M_B-M_K=3.4\pm0.2}$, whereas a 12~Gyr, solar abundance population is expected to have ${\rm M_B-M_K\simeq3.9}$: \citealt{maraston98a}). Both of these factors might suggest the presence of a significant young population of stars \citep[although see][]{cidfernandes04a}. It is beyond the scope of this present work, however, to attempt to take account of this effect. \bibliographystyle{apj_hyper} \bibliography{paper_bibliography.bib}
Title: PBH and DM from cosmic necklaces
Abstract: Cosmic strings in the brane Universe have recently gained a great interest. I think the most interesting story is that future cosmological observations distinguish them from the conventional cosmic strings. If the strings are the higher-dimensional objects that can (at least initially) move along the compactified space, and finally settle down to (quasi-)degenerated vacua in the compactified space, then kinks should appear on the strings, which interpolate between the degenerated vacua. These kinks look like ``beads'' on the strings, which means that the strings turn into necklaces. Moreover, in the case that the compact manifold is not simply connected, the string loop that winds around a non-trivial circle is stable due to the topological reason. Since the existence of degenerated vacua and a non-trivial circle is the common feature of the brane models, it is important to study cosmological constraints on the cosmic necklaces and their stable winding states in the brane Universe.
https://export.arxiv.org/pdf/hep-ph/0601014
\title{% PBH and DM from cosmic necklaces } \author{% Tomohiro Matsuda\footnote{E-mail:matsuda@sit.ac.jp} } \address{% $^1$Theoretical Physics Group, Saitama Institute of Technology, Saitama 369-0293, Japan } \abstract{Cosmic strings in the brane Universe have recently gained a great interest. I think the most interesting story is that future cosmological observations distinguish them from the conventional cosmic strings. If the strings are the higher-dimensional objects that can (at least initially) move along the compactified space, and finally settle down to (quasi-)degenerated vacua in the compactified space, then kinks should appear on the strings, which interpolate between the degenerated vacua. These kinks look like ``beads'' on the strings, which means that the strings turn into necklaces. Moreover, in the case that the compact manifold is not simply connected, the string loop that winds around a non-trivial circle is stable due to the topological reason. Since the existence of degenerated vacua and a non-trivial circle is the common feature of the brane models, it is important to study cosmological constraints on the cosmic necklaces and their stable winding states in the brane Universe.} \section{Introduction} In this talk we will explain the cosmological consequences of the production of Dark Matter(DM) and Primordial Black Hole(PBH) from the loops of the cosmic necklaces. To begin with, I think it is fair to explain why necklaces\cite{vilenkin_book} are produced in brane models, since in many papers it is discussed that ``only strings are produced in the brane Universe''\cite{tutorial}. Of course I think this claim is not wrong, however somewhat misleading for non-specialists. To explain what is misleading in the ``standard scenario'', we have a figure in Fig.\ref{fig:fig0}. In general, the distance between branes may appear in the four-dimensional effective action as a Higgs field of the effective gauge dynamics. At least in this case, it is natural to consider the cosmological defects coming from the spatial variation of the Higgs field, which corresponds to the ``deformation'' of the branes\cite{matsuda_deformation}. Is the spatial variation of the Higgs field unnatural in the brane Universe? The answer is, of course, no. One should therefore consider at least two different kinds of defects in brane models: One is induced by the brane creation that is due to the spatial variation of the tachyon condensation, while the other is induced by the brane deformation that is due to the spatial variation of the brane distance. Along the line of the above arguments, it is possible to construct Q-ball's counterpart in brane models\cite{matsuda_Q-balls}, which can be distinguished from the conventional Q-balls by their decay process. We therefore have an expectation that strings can be distinguished from the conventional ones, if one properly considers their characteristic features. Now let us discuss about the validity of the conventional Kibble mechanism. Of course the Kibble mechanism is an excellent idea that explains the nature of the cosmological defect formation. However, if there is oscillation of the brane distance that may be induced by the brane inflation or by a later phase transition that changes the brane distance, the four-dimensional counterpart of the brane distance (i.e. the Higgs field) oscillates in the effective action. In the four-dimensional counterpart, defect production induced by such oscillation is already discussed by many authors, including the production of the sphaleron domain walls which otherwise cannot be produced in the Universe\cite{matsuda_deformation}. The defect production induced by such oscillation may or may not be explained by the Kibble mechanism, however it should be fair to distinguish it from the ``conventional'' Kibble mechanism. Let us summarize the above discussion about the defect production in the brane Universe. Actually, it is possible to produce all kinds of defects in the brane Universe, however it is impossible to produce defects other than the strings {\bf simply as the result of the brane creation that is induced by the conventional Kibble mechanism}. One should therefore be careful about the assumption that is made in the manuscript, which may or may not be explicit. The necklaces are produced as the hybrid of the brane creation and the brane deformation. It should be noted that the stable loops of the necklaces that we will discuss in this talk may appear in the four-dimensional gauge dynamics, irrespective of the existence of the branes\cite{050906x}. The stabilization of the necklace loops is first discussed in Ref.\cite{matsuda_necklace} for brane models and in Ref.\cite{050906x} for necklaces embedded in four-dimensional gauge dynamics. In order to produce necklaces in the brane Universe, the motion in the compactified direction is important. I know that in the ``standard scenario'' it is sometimes discussed that the position of the strings are fixed by the potential that is induced by the supersymmetry breaking, and the position is a homogeneous parameter of the Universe because all the decay products (Typically, they are F, D, and $(p,q)$ strings) lie (at least initially) along the same plane of the original hypersurface on which the tachyon condensation took place. However, in this case one may hit upon the idea that the potential for a string cannot be identical to all the other kinds of the strings. One may therefore obtain many kinds of strings that may move independently along different hypersurfaces, with exponentially small intersection ratios. Moreover, I think it is not reasonable(but may be possible) to assume that the string motion is utterly restricted by the potential even in the most energetic epoch just after inflation. Please remember that in general the moving (inflating) brane carries kinetic energy, and the brane annihilation should be an energetic process, although one may admit that there could be exceptional scenarios. I therefore think that the decay products should have kinetic energy, which is enough to climb up the potential hill at least just after brane inflation. \section{PBH and DM from necklaces} The scenario of the PBH formation from strings is initiated by Hawking\cite{Hawking}, who utilizes the huge kinetic energy of the shrinking loops. However, the probability of finding loops that can shrink into their small Schwarzschild radius is very rare due to their random shape and motion, which weakens the obtained bound for the string tension up to $G\mu < 10^{-6}$. In our previous paper hep-ph/0509061\cite{0509061}, we have just extended Hawking's idea to the networks that include monopoles attached to the strings. It should be emphasized that both in the above scenarios, the kinetic energy of the shrinking object plays crucial role. Now let us discuss about our new idea for producing PBHs. The most obvious discrepancy is that in our new scenario we discard the benefit of the kinetic energy. We consider stable relics that are produced from necklace loops, whose mass is large enough to turn into black holes, even after they have dissipated their kinetic energy during the loop oscillation. The stability of the loops is due to their windings around the compactified space. Of course the production of PBHs is delayed compared to the Hawking's scenario, however the obtained bound is much stronger than the original scenario due to the high ($\sim 1$) production ratio. This new mechanism of PBH production is first advocated in hep-ph/0509062\cite{050906x}. Let us explain how one can count the winding number of a necklace loop. Please see Fig.\ref{fig:count}. We introduce $\chi(t)$, which is the step length between each random walk that corresponds to the right or the left movers in the compactified direction. Since the left and the right movers can annihilate on the necklaces, the actual distance between ``beads'' becomes much larger than $\chi(t)$. We therefore introduce another parameter $d(t)$, which is the typical length between the remaining ``beads''. Although the annihilation could be efficient, the simple statistical argument shows that the typical number of the beads that remain after annihilation is about $n^{1/2}$, if the initial number of the random walk is given by $n$. If the strings are in the scaling epoch, the typical length of the loops is $l(t)\sim \alpha t$ when they are disconnected from the string networks. Then one can obtain the typical mass of the stable relics, $M_{coils}\sim n(t)^{1/2}m \sim [l(t)/\chi(t)]^{1/2}m$, where $m$ is the typical mass of the beads. Here we should note that disregarding the annihilation, $\chi(t)\propto t^{-1}$ is already obtained in Ref.\cite{vilenkin-necklace}, which means that $d(t)$ should evolve as $d(t)\sim t^{-1}$ at least during the short periods between each annihilation. Of course $d(t)$ is discontinuous at each annihilation, however the underlying parameter $\chi(t)$ is continuous and depends on time as $\chi(t) \sim t^{-1}$. Using the above ideas, we can calculate the typical mass of the stable relics that are produced from necklace loops. The calculation of the PBH density is straightforward. We have obtained the result \begin{equation} G\mu < 10^{-21} \times \left[\frac{p}{10^{-2}}\right]^{4/5} \left[\frac{\gamma}{10^{-2}}\right]^{1/5} \left[\frac{t_n}{M_p/\mu}\right]^{3/5} \left[\frac{d(t_n)}{M_p/\mu}\right]^{3/5} \left[\frac{m}{10^{16} GeV}\right]^{-6/5}, \end{equation} where we have assumed $\alpha \sim \gamma G\mu$, and $t_n$ is the time when necklaces are produced. $p$ is the reconnection ratio that should be $\sim 1$ for the conventional cosmic strings, but $p \ll 1$ is possible in our case. The string loops are produced at any time, and the typical mass of the stable relics depends on the time when they are produced, because the typical length scale of the string network increases with time both in the friction-dominated and in the scaling epoch. Therefore, the relics that are produced in an earlier epoch may be too light to turn into black holes. The ``light'' relics are the ``monopoles'' if the cosmic strings are D-branes. However, the ``magnetic charge'' of the ``monopoles'' may or may not be identical to the conventional magnetic charge of the electromagnetism. Therefore, they are the candidate of DM, and possibly the origin of the troublesome monopole problem only if they carry the conventional magnetic charge. In our paper hep-ph/0509064\cite{050906x}, we have examined if the DM relics can put significant bound on the tension of the cosmic strings. We have obtained the result for $m \sim M_{GUT} \sim 10^{16}GeV$, \begin{equation} \label{finalresult} G\mu< 10^{-23}\times \left[\frac{p}{10^{-2}}\right]^{9/10} \left[\frac{1}{\beta_s}\right]^{9/10} \left[\frac{10^{-3}}{r}\right]^{9/10} \end{equation} where $r$ is the mass ratio between the string part and the beads on the necklaces, which becomes a constant in the scaling epoch\cite{050906x}. The difficulty in lowering the typical energy scale is discussed in Ref.\cite{matsuda_baryo} for baryogenesis and Ref.\cite{matsuda_CMB} for the mechanism to generate density perturbations.
Title: Stochastic Gravitational Wave Production After Inflation
Abstract: In many models of inflation, the period of accelerated expansion ends with preheating, a highly non-thermal phase of evolution during which the inflaton pumps energy into a specific set of momentum modes of field(s) to which it is coupled. This necessarily induces large, transient density inhomogeneities which can source a significant spectrum of gravitational waves. In this paper, we consider the generic properties of gravitational waves produced during preheating, perform detailed calculations of the spectrum for several specific inflationary models, and identify problems that require further study. In particular, we argue that if these gravitational waves exist they will necessarily fall within the frequency range that is feasible for direct detection experiments -- from laboratory through to solar system scales. We extract the gravitational wave spectrum from numerical simulations of preheating after $\lambda \phi^4$ and $m_{\phi}^2 \phi^2$ inflation, and find that they lead to a gravitational wave amplitude of around $\Omega_{gw}h^2\sim 10^{-10}$. This is considerably higher than the amplitude of the primordial gravitational waves produced during inflation. However, the typical wavelength of these gravitational waves is considerably shorter than LIGO scales, although in extreme cases they may be visible at scales accessible to the proposed BBO mission. We survey possible experimental approaches to detecting any gravitational wave background generated during preheating.
https://export.arxiv.org/pdf/astro-ph/0601617
\title{Stochastic Gravitational Wave Production After Inflation} \author{Richard Easther} \author{Eugene A. Lim} \affiliation{Department of Physics, Yale University, New Haven CT 06520, USA \\ Email: {\tt richard.easther@yale.edu} {\tt eugene.lim@yale.edu}} \section{Introduction} In recent years, the Cosmic Microwave Background (CMB) has been our primary window into the primordial universe. Scale invariant density perturbations, sourced by quantum fluctuations during inflation and processed by the photon-baryon plasma, are by far the most satisfying explanation for the observed temperature anisotropies in the CMB. Inflation also predicts the existence of a scale invariant spectrum of primordial gravitational waves, sourced by the same quantum fluctuations that underlie the scalar perturbations. Gravitational waves are only weakly coupled to matter fields, and move freely through the universe from the moment they are produced. They are thus the deepest probe of the early universe of which we currently know \cite{Starobinsky:1979ty,Rubakov:1982df,Fabbri:1983us,Abbott:1984fp,Allen:1987bk,Turner:1993vb,Boyle:2005se}. In the short term, the best hope for observing primordial gravitational waves is via their contribution to the B-mode of the CMB polarization. However, foreground weak lensing puts a fundamental limit on a B-mode signal sourced by gravitational waves \cite{Knox:2002pe,Hirata:2003ka,Seljak:2003pn}. Consequently, attention has recently focussed upon direct detection experiments \cite{Boyle:2005se,Ungarelli:2005qb,Crowder:2005nr,Smith:2005mm} which, while enormously challenging, may ultimately be more sensitive to a stochastic gravitational wave background than the CMB B-mode. These experiments are sensitive to frequencies far higher than those that contribute to the CMB, since the physical sizes of detectors necessarily range from laboratory scales through to solar system scales. However, as the inflationary spectrum is almost scale-free, the amplitude at short scales is not dramatically different from that seen at CMB scales, at least for simple models of inflation. The primordial spectrum is not alone: gravitational waves are produced whenever there are large, time-dependent inhomogeneities in the matter distribution. These more pedestrian gravitational waves are generated ``classically'', much like radiation produced by the oscillation of an electron, whereas the inflationary tensor perturbations are sourced by quantum fluctuations in the spacetime background. During the evolution of the universe, there are several hypothetical mechanisms which would generate large, local inhomogeneities. These include first order phase transitions \cite{Kamionkowski:1993fg}, brane-world models \cite{Sahni:2001qp,Sami:2004xk}, networks of cosmic strings \cite{Vilenkin:1981bx}, or inhomogenous neutrino diffusion \cite{Dolgov:2001nv}. Here, we investigate a further possibility: gravitational waves produced during preheating \cite{Felder:2000hq,Garcia-Bellido:1997wm,Greene:1997ge,Greene:1997fu,Traschen:1990sw,Kofman:1994rk,Greene:1998nh,Giudice:1999fb,Khlebnikov:1997di,Garcia-Bellido:1998gm,Kofman:1997yn,Bassett:1998wg,Easther:1999ws,Parry:1998pn,Finelli:1998bu,Liddle:1999hq,Bassett:2005xm,Suyama:2004mz,Tsujikawa:2002nf,Finelli:2001db,Tilley:2000jh,Henriques:2003ga}, a period of non-thermal evolution following the end of inflation. While the detailed dynamics of resonance and preheating is a complicated, nonlinear problem, the basic picture is very simple. For a given combination of fields, parametric resonance occurs when some subset of Fourier modes have exponentially growing solutions, driven by the oscillating inflaton field. The resonant modes are quickly pumped up to a large amplitude $\Phi^2 \approx \langle \chi^2 \rangle$. Resonance ends when nonlinearities render further growth kinematically expensive, leaving the universe far from thermal equilibrium. Thermalization occurs as the excited modes dissipate their energy over the entire spectrum via self-interaction. The rapid rise in the mode amplitude can be associated with an exponentially growing occupation number (at least for bosonic species). If one transforms into position-space, the highly pumped modes correspond to large, time dependent inhomogeneities, ensuring the matter distribution has a non-trivial quadrupole moment, sourcing the production of gravitational radiation. This topic was first addressed by Khlebnikov and Tkachev \cite{Khlebnikov:1997di}, and has not been widely discussed within the experimental community. We believe that the time is ripe for revisiting this question. Since \cite{Khlebnikov:1997di} was written, technological approaches to gravitational wave detection have advanced considerably. Moreover, there have been significant advances in the theoretical understanding of preheating in more complicated inflationary models. Finally, numerical simulations benefit from the gains in computational power over this interval. It will turn out that for ``typical'' cosmological parameters, gravitational radiation sourced by preheating has a peak frequency in the MHz band. Coincidentally, a new generation of detectors has been proposed which is tuned to gravitational waves in this range \cite{Ballantini:2005am}, although their strain sensitivity would need improve over current values by around $10^6$ in order to detect the spectra we compute here. This is clearly ambitious, but probably not unreasonable when compared to the 25 year interval anticipated before BBO [Big Bang Observer] will be in a position to detect the inflationary gravitational wave spectrum \cite{Phinney}. Since preheating occurs in many (although not all) inflationary models, any gravitational wave signal associated with preheating would provide a new and currently unexplored window into inflationary physics. It is worth emphasizing that the physical principles that underlie the calculations in this paper are well established and do not rely on exotic physics -- in particular the mechanism that governs the generation of the gravitational waves is simply the usual quadrupole related emission.\footnote{The possibility that preheating dynamics are directly affected by ``back-reaction'' from metric perturbations has also been discussed \cite{Finelli:1998bu,Parry:1998pn,Easther:1999ws,Bassett:1998wg,Finelli:2001db,Suyama:2004mz} but we do not address this effect in this work.} Finally, this signal carries information about the epoch at the end of inflation, opening a new window into the early universe. With the stakes this high any possible observational opportunity must be carefully explored. Unlike the primordial spectrum, gravitational waves induced by preheating (or any subsequent phase transition) will not be scale-free; indeed their spectrum is indicative of the complicated processes that generate them. This is both a boon and a possible pitfall. A scale dependent spectrum necessarily contains more information than a scale-free spectrum, so the detection and subsequent mapping of a gravitational wave background induced by preheating would yield a rich trove of information. However, gravitational wave detectors are subject to physical limitations, since they all ultimately measure deformations in (an array of) physical objects induced by passing gravitational waves. Thus, even in principle, it is hard to imagine the direct detection of gravitational waves below atomic scales, or beyond solar system scales. To be sure, this is a large range but it is very much shorter than any relevant cosmological scale. Fortunately, by a happy numerical coincidence, any gravitational wave spectrum generated during preheating will peak at scale between a few meters and millions of kilometers -- which overlaps with the ``golden window'' open to direct detection experiments. In Section (\ref{sect:preheating}), we briefly review preheating and discuss previous work. Following that, in Section (\ref{sect:lengthscales}) we demonstrate that preheating gravitational waves from inflation scales ranging from TeV to the GUT scale will peak around $1$ Hz to $10^8$ Hz. In Section (\ref{sect:production}) we discuss our methodology for numerically computing the gravitational wave spectrum. In Section (\ref{sect:results}) we present our results for $\lambda \phi^4$ and $m^2 \phi^2$ inflation, confirming and extending the results of \cite{Khlebnikov:1997di}. We note that while the inflationary dynamics of the two models are very similar, their resonant behaviors diverge considerably. Finally, we discuss the theoretical and observational implications of our results and lay out future plans to further improve the computational methodology in Section (\ref{sect:conclusions}). \section{Preheating} \label{sect:preheating} A key problem facing any inflationary model is to ensure that inflation ends. This issue is highlighted in Guth's foundational paper, describing what is now known as old inflation, which is driven by a meta-stable false vacuum does not successfully terminate \cite{Guth:1980zm}. In new or chaotic (slow-roll) inflation \cite{Linde:1981mu,Albrecht:1982wi,Linde:1983gd}, it was thought that inflation was followed by a period of {\em reheating\/}, where energy slowly bleeds from the inflaton field as it oscillates about the minimum of its potential \cite{Albrecht:1982mp}. Generating the correct perturbation spectrum typically requires that inflaton self-coupling is extremely weak, and this small coupling must be protected from loop corrections. Consequently, the coupling between the inflaton and other particles is necessarily tiny ($\lesssim 10^{-6}$ for $\lambda \phi^4$), ensuring that reheating proceeds slowly. Preheating provides a vastly more efficient mechanism for extracting energy from the inflaton field, and it proceeds non-perturbatively and non-thermally via a process known as parametric resonance \cite{Traschen:1990sw,Kofman:1994rk}. This is akin to stimulated emission in a laser: during preheating, individual momentum modes of fields coupled to the inflaton (or the inflaton itself, in some cases) have exponentially growing amplitudes. Consider the following action \begin{eqnarray} S&=&\int dx^4 \sqrt{-g}\left[\frac{\mpl ^2 R}{16\pi }-\frac{1}{2}(\partial \phi)^2 - V(\phi)- \right. \nonumber \\ &&\left.\frac{1}{2}(\partial \chi)^2-\frac{1}{2}g^2\phi^2\chi^2\right]. \label{eqn:action} \end{eqnarray} As usual the Hubble parameter, $H$ and scale factor, $a$, are related by $H = \dot{a}/a$ and during inflation the dynamics are described by% \begin{eqnarray} H^2 &=& \frac{8 \pi}{3 \mpl^2} \left[ \frac{\dot{\phi}^2}{2} + \frac{\dot{\chi}^2}{2} + {\cal V}(\phi,\chi) \right] \, , \\ \dot{H} &=& -\frac{8 \pi}{ \mpl^2} \left[ \frac{\dot{\phi}^2}{2} + \frac{\dot{\chi}^2}{2} \right] \, , \\ \ddot{\phi} &+& 3 H \dot{\phi} + \frac{\partial {\cal V}}{\partial \phi} =0 \, ,\\ \ddot{\chi} &+& 3 H \dot{\chi} + \frac{\partial {\cal V}}{\partial \chi} =0 \, , \end{eqnarray} where ${\cal V}=V(\phi)+1/2g^2\phi^2\chi^2$ is a shorthand for the potential terms in (\ref{eqn:action}). For simplicity we consider only couplings to scalar fields although fermionic preheating has also been investigated \cite{Greene:1998nh,Giudice:1999fb}. In this action, $\phi$ is the inflaton and the inflationary dynamics are fixed by the potential $V(\phi)$, which is assumed to possess a minimum. At the end of inflation, the potential energy of the field is quickly converted into kinetic energy, and the field oscillates with frequency $m_{\phi} = \sqrt{d^2V(\phi)/d\phi^2}$ evaluated at the minimum of $V(\phi)$. The solution for $\phi$ is then approximately \begin{equation} \phi(t)=\Phi(t)\sin m_{\phi} t. \end{equation} Meanwhile, the equation of motion for the $\chi$ field after expanding it in Fourier modes is \cite{Kofman:1994rk} \begin{equation} \chi_k''+3H\chi_k'+(A(k)-2q\cos 2z)\chi_k=0 \label{eqn:eom}. \end{equation} In the limit where we can ignore the expansion of the universe this is the Mathieu equation which is simply a harmonic oscillator with a periodic forcing function. Here we have made the identification $A(k)=k^2/(m_{\phi}^2 a^2)+2q$, rescaled the time to $z=m_{\phi}t$ and used a prime to denote differentiation with respect to $z$. The crucial \emph{resonance parameter} is \begin{equation} q=g^2\Phi^2/(4m_\phi^2) \, . \label{eqn:resonanceparameter} \end{equation} The Mathieu equation possesses both oscillatory and exponential solutions; for each individual mode $k$ one can compute $A(k)$ and $q$ to determine whether or not it goes into resonance \cite{Abramowitz}. Roughly speaking, for broad resonance where a large number of modes are excited we need $q>1$ \cite{Kofman:1997yn}. As first described by \cite{Traschen:1990sw,Kofman:1994rk}, some $\chi_k$ will have exponentially growing solutions for realistic parameter values. A full treatment of \emph{parameteric resonance} in an expanding universe is complicated, but we can make several generic statements. Firstly, preheating is very efficient and proceeds much more rapidly than reheating, which relies on tree level couplings between the inflaton and other matter fields (which, at minimum are provided by gravitational interactions). Parametric resonance typically lasts less than a Hubble time, and in some models will complete in a few oscillations of the inflaton. This is because the resonant modes are rapidly pumped up to an amplitude $\langle \chi \rangle^2 \approx \Phi^2$, cutting off resonance as it becomes kinematically expensive. Once preheating ends, the pumped-up modes dissipate their energy via self-interaction with other modes, thermalizing the universe. Preheating leads to an initially non-thermal distribution of energy in the $\chi_k$ states. At high frequencies, corresponding to Fourier modes that are much shorter than the size of the post-inflationary Hubble horizon, the effective mass of the $\chi_k$ state is much larger than the function amplitude and resonance does not occur. Meanwhile, at low frequencies, causality ensures that modes longer than the Hubble horizon are unlikely to be in resonance.\footnote{It is actually not impossible to have resonance at very small $k$, but it does not occur in generic models of preheating. For a summary see \cite{Tsujikawa:2002nf}.} Consequently, we expect the spectrum to be narrow and centered around a wavelength which is dependent on the energy scale at the end of inflation. For the models we study in detail here, the gravitational wave spectrum induced by preheating peaks at scales $1\sim 2$ orders of magnitude shorter than the energy scale at the end of inflation. This comoving scale can be converted into a physical scale in the present universe once the post-inflationary behavior of the scale factor $a(t)$ is specified. In Section (\ref{sect:lengthscales}) we show that the peak wavelength has a physical wavelength that scales as $H_e$ \begin{equation} l_0 \propto \frac{1}{\sqrt{H_e}} \propto \frac{1}{ V(\phi_e)^{1/4}} . \label{eqn:positionscaling} \end{equation} where the $e$ subscript denotes the value of $\phi$ and $H$ at the end of inflation. Lowering the inflation scale reduces the reheating scale, and reddens the gravitational wave spectrum. If the longest possible modes are excited in GUT scale inflation, the signal peaks around $10^{7} \sim 10^{8}$ Hz, although the excited modes are generally slightly shorter. As we will see below, this is a very challenging frequency range for any direct detection experiment. Reducing the inflation scale pushes the signature towards more easily observable frequencies. Given that parametric resonance naturally cuts off at both small and large scales, the spectrum of any gravitational waves will cover a fixed range of wavelengths. As the power is thus restricted to a relatively narrow band, the total gravitational radiation remains safely below the bound from big bang nucleosynthesis \cite{Maggiore:1999vm}. The bottom line is that for a short moment the universe is highly inhomogenous, providing a fertile ground for the generation of gravitational waves. Needless to say, preheating is a highly non-linear process and analytical estimates can only take us so far. Fortunately, given an action such as (\ref{eqn:action}), this is a problem that can be solved numerically; we simply derive the equations of motion and evolve them numerically on an expanding lattice. This mimics the growth of the universe, but ignores the back-reaction of metric perturbations on the field evolution -- an assumption that is self-consistent, as while $\delta \rho/\rho$ can be large, the metric perturbations are typically small. We use a modified version of the publicly available package {\sc LatticeEasy} \cite{Felder:2000hq} for the numerical computations. To our knowledge, the generation of gravitational waves by preheating has been thoroughly examined only once before, by Khlebnikov and Tkachev \cite{Khlebnikov:1997di}. Their work is the starting point for this paper: we elaborate and expand upon their treatment, considering a broader range of models, and taking recent developments in detector technology into account. In particular, preheating can occur at a very broad range of scales, for example via hybrid inflation \cite{Linde:1991km,Garcia-Bellido:1997wm,Garcia-Bellido:1998gm}. If the scale gets low enough, the peak wavelength can be close to the scales probed by next generation observatories such as BBO \cite{Phinney}. In Section (\ref{sect:results}), we reproduce numerical results of \cite{Khlebnikov:1997di} for the $\lambda \phi^4$ model, and present new results for the $m^2 \phi^2$ model. While the inflationary behavior of these two models is similar, their resonance structure is very different, providing a useful crosscheck on the generality of our analytic estimates. In future work we will extend these numerical calculations to hybrid inflation and fermionic preheating. \section{Peak wavelength and amplitude} \label{sect:lengthscales} We begin our detailed analysis with a general discussion of the different parameters that determine the amplitude and wavelength of any gravitational waves generated during preheating. Consider the ``usual'' gravitational wave power spectrum generated by quantum fluctuations of the background, \begin{equation} \Omega_{gw,inf}(k)h^2 = \Omega_{r}h^2\frac{32}{9}\left(\frac{V_e}{\mpl^4}\right) \left(\frac{g_0}{g_*}\right)^{1/3}, \label{eqn:quantumgrav} \end{equation} where $V^{1/4}_{inf}$ is the energy scale of inflation and $\Omega_r h^2\approx 4 \times 10^{-5}$ is the total density of radiation today. The effective number of degrees of freedom in the radiation at matter-radiation equality and today are given by $g_*$ and $g_0$ respectively. This form of the power spectrum is slightly non-standard as tensor modes are usually expressed via $P_h=8\pi H^2/\mpl^2$. There are two salient features to this spectrum. The first is that it is (almost) scale-invariant, since each mode is frozen out at an approximately constant energy scale, namely the inflation scale. The second is that the power is minute; for the most optimistic scenario where inflation occurs around the GUT scale, $\Omega_{gw,inf}(k) h^2 < 10^{-14}$. The current upper limit on the scale of inflation from WMAP observations for single-field inflation models is $V^{1/4}<3.3 \times 10^{16}$ GeV, corresponding to $\Omega_{gw,inf}h^2<2\times 10^{-15}$ \cite{Peiris:2003ff,Smith:2005mm} The gravitational waves which we are considering in this paper are not directly sourced by quantum fluctuations; instead they are generated by the classical motion of particles during preheating. As is well-known, accelerated motion generates a quadrupole moment, leading to the generation of gravitational radiation. During preheating at the end of inflation, large inhomogeneities in the matter fields are generated by the selective pumping of modes in parametric resonance. These large inhomogeneities, as first shown in \cite{Khlebnikov:1997di} for the $\lambda \phi^4$ model, are sufficiently large to produce gravitational waves with amplitudes many orders of magnitude larger than those produced by the quantum fluctuations. An analytical estimate \cite{Khlebnikov:1997di}, for an inflaton with an effective oscillation frequency $\bar{m}$, coupled to a massless scalar field $\chi$ with a $g^2\phi^2\chi^2$ term, yields the peak amplitude at the resonance mode $k \sim H_r$ \begin{equation} \Omega_{gw}(k\sim H_e)h^2\approx \Omega_{r}h^2\frac{\bar{m}^2}{g^2 \mpl^2}\left(\frac{g_0}{g_*}\right)^{1/3}. \label{eqn:simpleamplitude} \end{equation} In other words, the amplitude probes the oscillation scale, in contrast to the primordial spectrum which probes the inflation scale. If we plug in the usual field values for chaotic inflation $\bar{m}=m_{\phi}$ at the end of inflation $\phi\approx \mpl$ such that $V_e=m_{\phi}^2\phi^2/2 = m_{\phi}^2\mpl^2 /2 $, we see from equation (\ref{eqn:quantumgrav}) that the amplitude of the gravitational waves generated by preheating is $1/g^2$ larger than the inflationary spectrum. This is a significant boost, as we expect $g^2 \lesssim 10^{-6}$. Note that in models such as hybrid inflation \footnote{We note that in hybrid inflation, preheating amplification of the perturbations is achieved through a combination of parametric resonance and tachyionic instabilities. We thank Gary Felder for pointing this out to us.} one can decouple the oscillation frequency $\bar{m}$ from the inflaton mass $m$, and this simple relationship has to be revisited \cite{Garcia-Bellido:1997wm,Garcia-Bellido:1998gm}. However, only a finite range of modes excited during preheating. If the power was generated at scales corresponding to, say, atomic distances today, then our hope of detecting any gravitational waves induced by preheating would be dashed. On causal grounds, we expect that resonant modes have a wavelength roughly equal to or less than the Hubble length at the end of inflation, $1/H_e$: \begin{equation} H_e \sim \frac{ \sqrt{V_e}}{\mpl} \, , \end{equation} where $V_e$ is the inflationary potential. After inflation, the universe reheats to a temperature $\Trh$. During the subsequent radiation dominated phase, the Hubble parameter scales as $H = H_*(a_*/a_e)^2$ until matter-radiation equality at $T_*$. Meanwhile, the scale factor evolves as $a_* = a_0 (g_0/g_*)^{1/2} (T_0/T_*)$ from matter-radiation equality until today when $a_0\equiv 1$. Thus for a physical length $l$ \cite{Khlebnikov:1997di} and \emph{physical} wavevector $k$ we have \begin{eqnarray} l&=&\frac{1}{k} \frac{g_*^{1/2}}{g_0^{1/3}}\left(\frac{8\pi^3}{90}\right)^{-1/4}\frac{\sqrt{H_e M_p}}{T_0} \nonumber \\ & \approx& 0.5\frac{\sqrt{M_p H_e}}{k} ~\mathrm{cm} \end{eqnarray} % or \begin{equation} f=6 \times 10^{10} \frac{k}{\sqrt{M_p H_e}} ~\mathrm{Hz} \label{eqn:frequency} \end{equation} where we have used $g_0/g_* = 1/100$ in the second line. Plugging in the lowest excitable frequency $k = H_e$, where we more or less expect peak gravitational waves production to occur, we obtain the scaling relation (\ref{eqn:positionscaling}), as claimed earlier. The inverse scaling is particularly important: it means that the pertinent wavelengths are longer for smaller inflationary energy scales. If we assume instantaneous reheating after inflation for GUT scale inflation $H_e \approx 10^{13}$ GeV, $l\approx 1-10$ meters, and $f\approx 10^{7}$ to $10^{8}$ Hz. Lowering the inflationary scale reduces power in the primordial gravitational wave spectrum making it harder to detect, as quantified by equation (\ref{eqn:quantumgrav}). However, this also reddens the peak power of any preheating generated gravitational waves, making them easier for us to observe. This follows because the strain sensitivity $\tilde{h}_f$ of a detector scales as $\Omega_{gw}/f^3$ \cite{Maggiore:1999vm}, i.e. for the same value of $\Omega_{gw}$ we have to build a more sensitive detector if the frequencies are higher. In addition, if inflation occurs at a lower scale, then the gravitational wave energy density will be diluted less by expansion following preheating, again increasing our chance of observing them. On the other hand, if the gravitational waves are generated at a lower scale the off-diagonal terms of $T_{\mu\nu}$ are smaller for fixed $\delta \rho / \rho$, and will be less efficient sources of gravitational radiation. On the basis of the limited calculations performed in this paper, we see that the last two effects roughly cancel and $\Omega_{gw}$ does not depend strongly on $H_e$. However, further work will be needed before we can safely say that this is true of all models which undergo preheating. In this naive analysis, a few subtle points have been glossed over. The peak resonance modes are usually not exactly at the Hubble scale; instead they are frequently $1\sim 2$ orders of magnitude smaller \cite{Greene:1997fu,Kofman:1997yn,Greene:1997ge}. This has the effect of pushing the observable modes to a bluer band. On the other hand, preheating does not always start immediately after inflation ends; peak particle production occurs when the amplitude of the field perturbations $\delta \phi/\phi$ grows to order unity, which need not happen quickly. In the models looked at here, the Hubble parameter during peak gravitational wave production is about $1\sim 2$ orders of magnitude smaller than $H_e$, shifting the observable modes to a redder band. It is also worth noting that the gravitational waves are generated causally within the Hubble volume, and thus the phases of the individual modes are uncorrelated -- unlike the primordial spectrum. This is a generic feature of all causally generated perturbation spectra, and is a powerful discriminant \cite{Dodelson:2003ip}. Unfortunately, direct detection experiments cannot dinstinguish the coherence (or lack of) of the gravitational waves as their signal is an integral over some time interval greater than the frequency scale. To do this, one must find a processed \emph{imprint} on a fixed time-slice. While there is an upper bound on the inflationary energy scale from the contribution of tensor modes to the CMB, the lower bound is very weak. At minimum, the post-inflationary universe must be hot enough to permit baryogenesis and nucleosynthesis. We conservatively assume that the former occurs via electroweak scale processes, so we can easily have $V_e^{1/4}$ as low as the TeV scale. Nuclear reactions necessarily take place at MeV scales and ensuring successful nucleosynthesis provides an absolute lower limit on the reheating temperature. This corresponds to gravitational wave peak wavelength scales ranging from laboratory scales through to solar system scales. \section{Gravitational Wave Production} \label{sect:production} Equation (\ref{eqn:simpleamplitude}) suggests that the preheating induced gravitational wave spectrum is larger than its primordial counterpart. To obtain an actual power spectrum, the highly nonlinear physics of preheating forces us to turn to numerical methods. We use {\sc LatticeEasy\/} \cite{Felder:2000hq} to simulate the evolution of the early universe, solving the equations of motion for a set of interacting scalar fields in a flat Friedmann-Robertson-Walker (FRW) Universe. The fields become highly inhomogeneous, which is important for the generation of gravitational waves. This does not immediately make the {\em metric\/} perturbations large \cite{Ishibashi:2005sj}. Consequently, we can solve the nonlinear field evolution numerically while assuming a rigid spacetime background, and then extract the spectrum of gravitational radiation produced during preheating. We extended {\sc LatticeEasy\/} to compute the gravitational wave spectrum generated during preheating. We follow the approach of \cite{Khlebnikov:1997di} (see also \cite{Kamionkowski:1993fg}), reproducing their results for the quartic $\lambda \phi^4 /4 + g^2\phi^2\chi^2/2$ model. In addition, we compute the gravitational wave spectrum for the quadratic inflation model $m^2\phi^2/2 +g^2\phi^2\chi^2/2$. We leave the simulation of other models such as the negative coupling $-g^2\phi^2\chi^2/2$ \cite{Greene:1997ge} or hybrid inflation models to future work. We now sketch the approach we use to compute the spectrum. We begin by considering the energy radiated in gravity waves in a frequency interval $d\omega$ and a solid angle $d\Omega$, given by \begin{equation} \frac{dE}{d\Omega}=2G\Lambda_{ij,lm}\omega^2 T^{ij*}(\vec{\mathbf{k}},\omega)T^{lm}(\vec{\mathbf{k}},\omega) d\omega \label{eqgrav} \end{equation} where $T^{ij}$ is the stress tensor describing the source matter fields. Here $i$,$j$ run over the spatial indices and the projection tensor is given by \cite{WeinbergBook} \begin{eqnarray} \Lambda_{ij,lm}(\hat{k})=\delta_{ij}\delta_{lm}-2\hat{k}_j\hat{k}_m\delta_{il}+\frac{1}{2}\hat{k}_i\hat{k}_j\hat{k}_l\hat{k}_m \nonumber \\ -\frac{1}{2}\delta_{ij}\delta_{lm}+\frac{1}{2}\delta_{ij}\hat{k}_l\hat{k}_m+\frac{1}{2}\delta_{jl}\hat{k}_i\hat{k}_m \end{eqnarray} with unit vector $\hat{k}\equiv \vec{\mathbf{k}}/\omega$. Strictly speaking, this formula is only valid for linearized gravity in Minkowski space. A more accurate calculation will involve solving the equations of motion for linearized gravity on a curved background. To see why the use of this formula is justified, consider the gravitational wave energy emitted by a three dimensional box of \emph{conformally flat} spacetime with physical size $l\times l\times l$. The energy density in this box is \begin{equation} d\rho(\omega)=8\pi G l^{-3} \Lambda_{ij,lm}\omega^2 T^{ij*}(|\mathbf{\vec{k}}|,\omega)T^{lm}(|\mathbf{\vec{k}}|,\omega) d\omega \label{eqn:gravdensity} \end{equation} where we have assumed that the spectrum is isotropic.\footnote{Isotropy allows us to choose any direction for the projection vector $\Lambda_{ij,lm}$: we chose for simplicity $\hat{k}=(1,0,0)$. We checked that this assumption is robust by showing that the final simulation results are not sensitive to different choices of direction.} From causal arguments alone, only modes of wavelengths equal to or shorter than $1/H$ will be generated, imposing a natural cut-off at long scales. Thus, provided we choose $l\geq 1/H$ we will effectively capture the essential physics. Depending on how efficient preheating is in a particular model, the entire phase can last for several Hubble times. However, the gravitational waves are produced near the end of preheating, as the inhomogeneities in the fields become large. We therefore expect the gravitational wave power to be generated in a short burst, and numerical simulations confirm this suspicion. Thus it is a good approximation to assume that the gravitational wave source is localized in the box \cite{Kosowsky:1991ua}. In our simulations, we begin our computations at the end of inflation, near the beginning of the parameteric resonance phase. We end our simulations when the fields are stabilized and parameteric resonance ends. We subdivide the the spacetime into discrete 4-D boxes of spatial sizes $L^3$ and time interval $\tau=L$, where $L$ and $\tau$ are the conformal length and time respectively. Our physical box size thus scales roughly as $a^3$. The choice of $\tau=L$ is purely operational, allowing us to fix our Fourier variables to be the conformal frequency and conformal wavevector such that $|\mathbf{\vec{k}}_{conf}|=\omega_{conf}$. One can in principal decompose the conformal time differently, but that would unnecessarily complicate matters. We fix $L$ so that during the period of gravitational wave production $aL \geq 1/H$ and the box is larger than the effective Hubble horizon. We assume that each ``box'', labeled $\alpha$, is a localized source, and compute the total gravitational wave density produced for each box $\rho_{gw}^{(\alpha)}$ using equation (\ref{eqn:gravdensity}). We then sum them up, diluting them appropriately as follows \begin{equation} \frac{d\rho_{gw}(a_{e})}{d\ln \omega}=\sum_{\alpha} \frac{d\rho_{gw}^{(\alpha)}(a_{\alpha})}{d\ln \omega}\left(\frac{a_{e}}{a_{\alpha}}\right)^{4} \label{eqn:finalpower} \end{equation} where $a_{\alpha}$ is the scale factor taken at the middle of the box in conformal time and $a_{e}$ is the scale factor at the end of inflation. Meanwhile, for each box $\alpha$ \begin{equation} \frac{d\rho_{gw}(a_{\alpha})}{d\ln \omega}=8\pi G \omega^3 l_{a_{\alpha}}^{-3}\Lambda_{ij,lm}T^{ij*}T^{lm} \end{equation} where $l_{a_{\alpha}}^3$ is the physical size of the box at time $a_{\alpha}$ and $\omega$ is the physical frequency. Finally, putting everything together, the total density of gravitational waves today is given by \begin{equation} \Omega_{gw}h^2=\Omega_r h^2 \frac{d\rho_{gw}(a_{e})}{d\ln \omega}\left(\frac{g_0}{g_*}\right)^{1/3} \label{eqn:finalpowertoday} \end{equation} We should mention that in equation (\ref{eqn:finalpower}) and hence equation (\ref{eqn:finalpowertoday}), we have implicitly assumed that the universe is radiation dominated at the end of preheating, which is not true for certain chaotic models. \section{Numerical Results} \label{sect:results} In this section, we give numerical results for the gravitational wave spectrum produced during resonance in two different models: $\lambda \phi^4$ and $m^2\phi^2$. While the inflationary dynamics of these two systems are very similar, there is considerable divergence in the resonance structure between the models, making this a useful generalization of \cite{Khlebnikov:1997di}. \subsection{Quartic Inflation ($\lambda \phi^4$)} To test our code, we reproduce Khlebnikov and Tkachev's results \cite{Khlebnikov:1997di} for $\lambda \phi^4$ with a $\phi^2\chi^2$ term \begin{equation} {\cal V}(\phi,\chi)=\frac{\lambda}{4}\phi^4 + \frac{1}{2}g^2\phi^2 \chi^2. \end{equation} From the perspective of preheating, this model is atypical \cite{Greene:1997fu} as it possess only a weak resonance band. Even so, we still see significant production of gravity waves. Following \cite{Khlebnikov:1997di}, we set $\lambda=10^{-14}$ and $g^2/\lambda=120$, corresponding to a resonance parameter $q\approx 120$ from equation (\ref{eqn:resonanceparameter}). In this model, inflation ends around the GUT scale, where $\phi_0\approx \mpl$, or $H_{end}\approx 10^{12}$ GeV. We begin our simulation on a $256^3$ size lattice from that time and run it until preheating ends around $H\approx 10^{7}$ GeV. Parameteric resonance peaks around $H_{peak}\approx 10^{8}$ GeV, and the size of the box is chosen to ensure that its physical size at this time $l \approx 1/H_{peak}$. With a $\lambda \phi^4$ potential, the background spacetime scales like a radiation dominated universe during parametric resonance. Using (\ref{eqn:frequency}), the present frequency associated with the Hubble parameter during preheating is $10^{6}$Hz. From figure (\ref{fig:quarticstdresult}), we see that the peak frequency is actually $10^{7}\sim 10^{8}$ Hz, suggesting that the peak resonance modes are about two orders of magnitude smaller than the Hubble wavelength. The amplitude of the gravitational waves peaks at around $\Omega_{gw}\approx 10^{-9}$, consistent with equation (\ref{eqn:simpleamplitude}). From the plot, we see that even at the lower end of the relevant frequency range which is easier to detect, $\Omega_{gw} \sim10^{-11}$. This is 3 orders of magnitude larger than the primordial spectrum. Beyond that at lower frequencies, we expect the spectrum to undergo a steep $k^3$ decline. This $k^3$ superhorizon tail is a common property for a spectrum which is causally generated inside the Hubble horizon \cite{Liddle:1999hq}. In cases where the inflationary scale and thus the intrinsic stochastic background is very low (so it does not mask the signal) this $k^3$ tail might be easier to detect than the peak wavelengths, given the physical limitations on realistic detectors. \subsection{Quadratic Inflation ($m_{\phi}^2\phi^2$)} We now turn to the $m_{\phi}^2\phi^2$ model \cite{Linde:1983gd} with the same interaction term as before \begin{equation} {\cal V}(\phi,\chi)=\frac{1}{2}m_{\phi}^2\phi^2 + \frac{1}{2}g^2\phi^2 \chi^2. \end{equation} The amplitude of the CMB temperature anisotropy requires $m_{\phi}\approx 10^{13}$ GeV. At the end of inflation $\phi \approx \mpl$. Choosing $g^2=2.5 \times 10^{-7}$, gives a resonance parameter of $q\approx 2.5\times 10^{5}$, via (\ref{eqn:resonanceparameter}). Again we begin our simulation after inflation ends at $H_{end}\approx 10^{13}$ GeV through the peak preheating phase at $H_{peak}\approx 10^{11}$ GeV, until the end of preheating. Since $m_{\phi}$ is non-zero, the universe evolves as if it was matter dominated\footnote{The equation of state $w=\langle p \rangle / \langle \rho \rangle$ fluctuates rapidly between $1$ and $-1$ around a center value of $0$.} with the scale factor growing 30-fold. At the end of preheating, we assume that the universe reheats normally and enters a radiation dominated phase. This is in principle not a valid assumption, as some numerical results have shown that it is difficult for quadratic inflation to reheat to radiation domination without a trilinear coupling \cite{Podolsky:2005bw}. In this paper, we are using it as a toy model to illustrate preheating for a inflation model with a different mass scale. We simulated this model on a $256^3$ lattice, with the results as shown in figure (\ref{fig:chaoticstdresults}). Using (\ref{eqn:simpleamplitude}), we expect $\Omega_{gw}h^2\approx 10^{-10}$ at peak which matches the result of our detailed calculation. Although inflation ends at $H_{end}\approx 10^{13}$ GeV, peak resonance occurs at $H_{peak}\approx 10^{11}$ GeV, which sets the comoving size of our lattice. Figure (\ref{fig:chaoticstdresults}) shows that the peak frequency is actually $10^{8}\sim 10^{9}$ Hz, a couple of orders of magnitude smaller than the Hubble parameter during reheating. Finally, we present the results for a model with a lower mass, $m_{\phi}=10^{12}$ GeV in figure (\ref{fig:chaoticstdresults2}). This model is ruled out by CMB data, but it demonstrates the way in which the gravitational wave spectrum generated by preheating depends on the inflationary scale. In this model $H\approx 10^{12}$ GeV, and as expected from equation (\ref{eqn:positionscaling}), the peak location is reddened by a factor of $\sqrt{10}$. The observable power remains comparable to the previous model. The emitted power is reduced, but since the overall expansion of the universe is reduced by the lower reheating temperature, the values of $\Omega_{gw}h^2$ today is roughly fixed. It is tempting to conjecture that the cancellation between these two effects will be seen in other preheating induced gravitational wave spectra, and that the $\Omega_{gw} h^2 \sim 10^{-10}$ seen here will prove to be a generic value \cite{Garcia-Bellido:1997wm,inpreparation}. \section{Summary and Future Prospects} \label{sect:conclusions} We have carefully investigated the production of gravitational waves during preheating, reproducing the work of Khlebnikov and Tkachev \cite{Khlebnikov:1997di} for the $\lambda \phi^4$ inflation model and extending it to the $m_{\phi}^2\phi^2$ case. For both models we show numerically that preheating is a sizable source of gravitational waves with frequencies of around $10^{6}\sim 10^{8}$ Hz, and peak power of $\Omega_{gw}h^2\approx 10^{-9} \sim 10^{-11}$. We present simple scaling arguments to predict the overall properties of the spectrum for a broader class of inflationary models. We see that the spectrum of gravitational waves induced by preheating peaks at a scale proportional to $1/\sqrt{H}$, where $H$ is the Hubble parameter during preheating, and generally somewhat smaller than the scale of inflation. Thus, lowering the inflationary scale reddens the spectrum and makes it easier to observe. This is in contrast to the primordial inflationary spectrum, which is roughly scale invariant and becomes harder to observe as inflationary scale is lowered. We now ask what we can learn about inflation if we detect a spectrum of gravitational waves generated during preheating. Its two most basic features, the peak frequency and the amplitude, represent the reheat and the oscillation scales respectively. As the reheat scale is lower than the inflation scale, its detection would impose a lower bound on the inflationary scale. Its usefulness as a probe of inflation is amplified if we have a separate probe of the scale of inflation, say from the CMB B-mode observations. The oscillation scale is harder to interpret, as it is often highly model dependent. For single scalar field inflation, such as the models considered here, knowledge of the reheat scale would constrain the coupling constant $g^2$. More optimistically, the \emph{structure} of the spectrum encodes information about resonance and preheating, so if we can predict the structure accurately we potentially probe the detailed mechanics of preheating. Such an endeavour will require more careful computations and simulations than we present in this paper. Further progress on this problem can be made in two ways \cite{inpreparation}. The first is to further refine the code to accommadate a larger class of models, particularly hybrid inflation models which have an essentially arbitrary inflationary scale. Alternatively, as alluded to in Section (\ref{sect:production}), a more sophisticated theoretical calculation would be to directly solve the evolution equations for the off-diagonal parts of the perturbed Einstein tensor, which are sourced by $T_{ij}$. This would avoid any ambiguity concerning the use of a formula that is only strictly applicable in flat space, and it would avoid the need to run the code for a finite number of ``boxes'', since we would only need to take a Fourier transform at the end of the computation. Investigating these gravitational waves is timely, since there is currently considerable interest in the direct detection of gravitational waves. At the moment, several proposals based on different technologies are being actively pursued. At the solar-system scale, the space-based interferometer LISA \cite{LISA} will probe frequencies from $10^{-2}$ Hz, which is probably too small for the gravitational waves we are considering. The proposed BBO \cite{Phinney} and also the Deci-hertz Interferometer Gravitational Wave observatory (DECIGO) \cite{Seto:2001qf} missions are sensitive to frequencies on the order of $1$ Hz, and would probe gravitational waves arising from preheating after TeV scale inflation. The array of terrestrial interferometers also probes frequencies corresponding to preheating following low-scale inflation. These experiments include GEO600 \cite{GEO600}, LIGO \cite{LIGO}, TAMA \cite{TAMA} and VIRGO \cite{VIRGO}. These are sensitive to scales between $100\sim 1000$ Hz and may be able to probe a stochastic background in the interesting range $\Omega_{gw}h^2 \sim 10^{-10}$, if they are correlated. At even higher frequencies in the KHz range, we have a slew of resonant bars detectors \cite{ALLEGRO,AURIGA,EXPLORER,Blair:1997as}. Once correlated, these resonant bars have a potential to reach $\Omega_{gw}h^2\approx 10^{-5}$ \cite{Maggiore:1999vm}, which puts the signals we see here out of their reach. However, theoretical studies suggest that correlating hollow spherical detectors may eventually allow us to reach $\Omega_{gw}h^2\approx 10^{-9}$ \cite{Coccia:1997gy}. At even higher frequencies from $10^{3} \sim 10^{5}$ Hz, there has been a proposal to build a superconducting resonant cavity detector called the Microwave Apparatus for Gravitational Wave Observation (MAGO) \cite{Pegoraro:1977uv,Reece:1984gv,Ballantini:2005am}. Although the strain sensitivity $\tilde{h}_f \approx 10^{-21} \textrm{Hz}^{-1/2}$ for the prototype is expected to be comparable to resonant bar detectors at $4\times 10^{3}$ Hz, the large $f^3$ suppression from the relation $\Omega_{gw} h^2 \propto \tilde {h}_f^2 f^3$ means that at these frequencies we can only reach $\Omega_{gw} h^2\approx {\cal{O}}(1)$. An improvement of $5\sim 6$ orders of magnitude in the strain sensitivity is needed to reach $\Omega_{gw}h^2\approx 10^{-10}$, a possibility which may be achieved by further refinements to the prototype and/or construction of an array of such detectors \cite{Ballantini:2005am}. By way of comparison, we note that the best hope for observing a primordial gravitational wave background is currently provided by BBO, which has a lead time of at least 20-25 years. In this context hoping for a large extrapolation of detector technologies at high frequencies is perhaps not excessively optimistic. In Figure~\ref{fig:sensitivityplot} we sketch the sensitivities of the leading interferometric detectors along with the expected stochastic background produced during preheating for the models we discuss in detail here. Finally, if one was to ever make a concerted attempt to detect a gravitational wave spectrum associated with preheating, one would need to be understand other potential sources that could supply a stochastic background of gravitational waves. This includes any first order phase transition in the early universe (such as the electroweak scale), or decays from cosmic strings. In addition, the presence of a rising component in the spectrum illustrates the dangers of using a (locally) positive spectral index of any detected stochastic gravitational wave background to rule out inflation in favour of alternative cosmogenesis ideas such as ekpyrosis \cite{Khoury:2001wf,Boyle:2003km} or pre-big bang scenarios \cite{Gasperini:1992em}. The potential for gravitational waves to provide a clean probe of inflation has rightfully drawn considerable attention, and strongly motivates attempts to detect the primordial gravitational spectrum. However, cosmic evolution is seldom tidy and gravitational waves are produced as long as large inhomogeneities are present. Preheating is a mechanism which will generate large inhomogeneities, and will necessarily be accompanied by the generation of a stochastic background of gravitational waves. The challenge now is to better determine their properties, and to assess possible strategies for their detection. \section*{ Acknowledgments} We thank Latham Boyle, Gary Felder, Gianluca Gemme, Tom Giblin, Will Kinney, Hiranya Peiris, Geraldine Servant, Igor Tkachev, and David Wands for a number of useful discussions. We are particularly indebted to Gary Felder and Igor Tkachev for their work on {\sc LatticeEasy\/}. This work is supported in part by the United States Department of Energy, grant DE-FG02-92ER-40704.
Title: Helicity-Rotation-Gravity Coupling for Gravitational Waves
Abstract: The consequences of spin-rotation-gravity coupling are worked out for linear gravitational waves. The coupling of helicity of the wave with the rotation of a gravitational-wave antenna is investigated and the resulting modifications in the Doppler effect and aberration are pointed out for incident high-frequency gravitational radiation. Extending these results to the case of a gravitomagnetic field via the gravitational Larmor theorem, the rotation of linear polarization of gravitational radiation propagating in the field of a rotating mass is studied. It is shown that in this case the linear polarization state rotates by twice the Skrotskii angle as a consequence of the spin-2 character of linear gravitational waves.
https://export.arxiv.org/pdf/gr-qc/0601054
\title{Helicity-Rotation-Gravity Coupling for Gravitational Waves} \author{Jairzinho Ramos} \affiliation{Physics Department, Drexel University, Philadelphia, Pennsylvania 19104, USA} \author{Bahram Mashhoon} \affiliation{Department of Physics and Astronomy, University of Missouri-Columbia, Columbia, Missouri 65211, USA} \pacs{04.20.Cv} \section{Introduction} In a recent paper on the purely gravitational spin-rotation coupling, Shen \cite{Shen} has treated the coupling of graviton spin to the gravitomagnetic field. In this way, spin-gravity coupling has been extended to include gravitational waves. The subject of spin-rotation-gravity coupling for a particle of spin $s$ has been reviewed in Refs. \cite{Mash} and \cite{Ryder} and discussions of more recent advances are contained in Refs. [4-12]; however, these treatments have ignored the $s=2$ case. Shen's field-theoretical approach is based on a weak-field approximation scheme that emphasizes the self-interaction of the nonlinear gravitational field \cite{Shen}. The purpose of the present paper is to investigate the consequences of the helicity-gravitomagnetic field coupling for weak gravitational waves. To provide a comprehensive treatment, we begin with the analysis of the propagation of free gravitational waves in a Minkowski spacetime background from the standpoint of a uniformly rotating observer. We then generalize our helicity-rotation coupling results to the propagation of gravitational waves in the field of a rotating astronomical mass via the gravitational Larmor theorem. The plan of this paper is as follows. In Section II, we study the reception of a free gravitational wave by a rotating observer in Minkowski spacetime. Section III deals with the modification of Doppler effect and aberration for gravitational waves caused by the helicity-rotation coupling. The results are extended to the gravitational case in Section IV using the gravitational Larmor theorem. It is then possible to study the influence of the gravitomagnetic field of a rotating source on the propagation of high-frequency gravitational radiation. The rotation of the linear polarization state of gravitational radiation propagating in the field of a rotating source is directly calculated in Section V using the eikonal approximation. Section VI contains a discussion of our results. In this paper, we choose units such that $c=1$, moreover, the signature of the metric is $+2$ in our convention. \renewcommand{\theequation}{2.\arabic{equation}} \setcounter{equation}{0} \section*{II. HELICITY-ROTATION COUPLING} Imagine a class of uniformly rotating observers ${\cal O}'$ in a global inertial frame of reference. For the sake of concreteness, we choose Cartesian coordinates such that the observers rotate with a frequency $\Omega$ about the $z$ axis, each on a circle of radius $\rho$, $0 \leq \rho < 1/\Omega$, parallel to the $(x,y)$ plane of the inertial system. The local orthonormal tetrad frame of each observer is given in the $(t,x,y,z)$ system by \begin{eqnarray} \Lambda^{\mu}_{(0)}&=&\gamma(1,-v \sin\phi,v \cos\phi,0), \label{b1} \\ \Lambda^{\mu}_{(1)}&=&(0,\cos\phi,\sin\phi,0), \label{b2} \\ \Lambda^{\mu}_{(2)}&=&\gamma(v,-\sin\phi,\cos\phi,0), \label{b3} \\ \Lambda^{\mu}_{(3)}&=&(0,0,0,1), \label{b4} \end{eqnarray} where $\phi=\Omega t=\gamma\Omega\tau$, $v=\Omega\rho$, $\gamma$ is the Lorentz factor corresponding to $v$ and $\tau$ is the proper time such that $\tau=0$ at $t=0$. Regarding the motion of each observer, we note that $\Lambda^{\mu}_{(1)}$ and $\Lambda^{\mu}_{(2)}$ indicate the radial and tangential directions, respectively, in cylindrical coordinates. Let us first consider an incident plane monochromatic gravitational wave of frequency $\omega$ and wave vector ${\bf k}=\omega(0,0,1)$, so that each observer rotates about the direction of wave propagation. The gravitational potential of the incident radiation is given by the symmetric tensor $h_{\mu\nu}$, which represents a small perturbation of the background Minkowski metric $\eta_{\mu\nu}$. Therefore, only terms that are linear in $h_{\mu\nu}$ will be considered throughout. In the transverse-traceless gauge, $h_{0\mu}=0$ and the potential for circularly polarized gravitational radiation is given by the matrix $(h_{ij})={\rm Re}(P_{\pm})$, where \begin{equation} P_{\pm}=(\epsilon_{\oplus} \pm i\epsilon_{\otimes})\hat{h}({\bf k})e^{i\omega(z-t)}. \label{b5} \end{equation} The upper (lower) sign corresponds to positive (negative) helicity and the two independent linear polarization states are denoted by \begin{equation} \epsilon_{\oplus}=\left( \begin{array}{ccc} 1 & 0 & 0 \\ 0 & -1 & 0 \\ 0 & 0 & 0 \end{array}\right), \,\,\,\,\,\,\,\,\,\,\,\ \epsilon_{\otimes}=\left( \begin{array}{ccc} 0 & 1 & 0 \\ 1 & 0 & 0 \\ 0 & 0 & 0 \end{array}\right). \label{b6} \end{equation} All of the operations involving wave functions are linear; therefore, only the real part of the relevant quantities will be of any physical significance. Consider the measurement of spacetime curvature by the static observers ${\cal O}$; the Riemann tensor can be expressed as a $6 \times 6$ matrix $({\cal R}_{AB})$, where the indices $A$ and $B$ range over the set $\{01,02,03,23,31,12\}$. The gravitational field is given by the measured components of the Riemann tensor and for a Ricci-flat field, \begin{equation} {\cal R}=\left( \begin{array}{cc} E & H \\ H & -E \end{array}\right), \label{b7} \end{equation} where $E$ and $H$ are $3 \times 3$ symmetric and traceless matrices corresponding to the electric and magnetic components of the curvature tensor. It turns out that for all of the circularly polarized gravitational waves considered in this section, the gauge-invariant Riemann tensor is of the form of Eq. (\ref{b7}) with $H=\pm i E$, which, taking due account of the proper definitions of curvature-based gravitoelectric and gravitomagnetic fields \cite{Mash3}, corresponds exactly to the analogous electromagnetic case. It is therefore sufficient to focus attention only on the electric part of the Riemann tensor. For the static observers ${\cal O}$, the field of circularly polarized radiation is thus given by $E=\frac{1}{2}\omega^{2}P_{\pm}$. The gravitational field as determined by the rotating observers ${\cal O}'$ is given by $R_{\mu\nu\rho\sigma}\Lambda^{\mu}_{(\alpha)}\Lambda^{\nu}_{(\beta)}\Lambda^{\rho}_{(\gamma)}\Lambda^{\sigma}_{(\delta)}$. These may be expressed via the real part of ${\cal R}'={\cal L}{\cal R}{\cal L}^{\dagger}$, where ${\cal L}$ is a real $6 \times 6$ matrix that can be determined from Eqs. (\ref{b1})-(\ref{b4}). We find that \begin{equation} {\cal L}=\left( \begin{array}{cc} {\cal A} & {\cal B} \\ -{\cal B} & {\cal A}, \end{array}\right), \label{b8} \end{equation} where ${\cal A}$ and ${\cal B}$ are given by \begin{equation} {\cal A}=\left( \begin{array}{ccc} \gamma\cos\phi & \gamma\sin\phi & 0 \\ -\sin\phi & \cos\phi & 0 \\ 0 & 0 & \gamma \end{array}\right), {\cal B}=v\gamma\left( \begin{array}{ccc} 0 & 0 & -1 \\ 0 & 0 & 0 \\ \cos\phi & \sin\phi & 0 \end{array}\right). \label{b9} \end{equation} It is then possible to express ${\cal R}'$ as \begin{equation} {\cal R}'=\left( \begin{array}{cc} C_{\pm} & \pm iC_{\pm} \\ \pm iC_{\pm} & -C_{\pm} \end{array}\right), \label{b10} \end{equation} where $C_{\pm}$ is given by \begin{equation} C_{\pm}=\frac{1}{2}\omega^{2}\left( \begin{array}{ccc} \gamma^{2} & \pm i\gamma & \pm iv\gamma^{2} \\ \pm i\gamma & -1 & -v\gamma \\ \pm iv\gamma^{2} & -v\gamma & -v^{2}\gamma^{2} \end{array}\right)\hat{h}({\bf k})e^{i\omega z-i(\omega \mp 2\Omega) t}. \label{b11} \end{equation} Thus the frequency measured by the rotating observers via the temporal dependence of Eq. (\ref{b11}) is \begin{equation} \omega'=\gamma(\omega \mp 2\Omega), \label{b12} \end{equation} which clearly exhibits the contribution of helicity-rotation coupling \cite{Mansouri}. This expression is the exact spin-2 analog of the electromagnetic result \cite{Mash1,Neutze} that has been observationally verified for $\omega\gg\Omega$. Let us note that a simple application of the Doppler formula would lead to the transverse Doppler frequency $\omega'_{D}=\gamma\omega$; however, the Doppler formula must be modified by taking into account the helicity-rotation coupling as in Eq. (\ref{b12}). For a packet of free gravitational waves propagating along the axis of rotation of the observer, the frequency of each Fourier component would be affected as in Eq. (\ref{b12}). A complete discussion of the physical implications of Eq. (\ref{b12}) will not be given here, since such treatment would be entirely analogous to the electromagnetic case that has been discussed in detail in Ref. \cite{Mash}. Let us now consider the extension of helicity-rotation coupling to the case of {\it oblique} incidence. Expressing the incident plane wave in terms of spherical waves whose dependence upon time $t$ and the azimuthal coordinate $\varphi$ is of the form ${\rm exp}(-i\omega t+im\varphi)$, and taking into account the fact that a transformation to the frame of the rotating observer involves $(r,\vartheta, \varphi) \rightarrow (r,\vartheta, \varphi')$ in terms of spherical polar coordinates such that $\varphi=\varphi'+ \Omega t$, we find that \begin{equation} \omega'=\gamma(\omega-m\Omega), \,\,\,\,\,\,\,\ m=0,\pm 1, \pm 2,\ldots . \label{b13} \end{equation} Here $m$ is the multipole parameter such that $m\hbar$ is the total (orbital plus spin) angular momentum of the radiation field along the direction of rotation of the observer. When this direction coincides with the direction of wave propagation, only the spin contributes to $m$ in Eq. (\ref{b13}) and hence we recover Eq. (\ref{b12}) for radiation of definite helicity $(m=\pm 2)$. It proves interesting to consider the other special case where the contribution of the orbital angular momentum of the obliquely-incident radiation field vanishes, namely, the reception of the radiation by a rotating observer ${\cal O}'_{0}$ at ${\bf x}=0$. In this case, the noninertial observer is at rest at the origin of spatial coordinates, but refers its measurements to axes rotating with frequency $\Omega$. The tetrad frame for ${\cal O}'_{0}$ is given by Eqs. (\ref{b1})-(\ref{b4}) with $v=0$ and $\gamma=1$. Imagine therefore an incident plane circularly-polarized gravitational wave with wave vector ${\bf k}=\omega{\bf \hat{k}}$, where ${\bf \hat{k}}=(0,-\sin\theta, \cos\theta)$. According to the static inertial observers, the natural orthonormal triad for the wave is $({\bf \hat{x}},{\bf \hat{n}},{\bf \hat{k}})$, where ${\bf \hat{n}}={\bf \hat{k}} \times {\bf \hat{x}}=(0,\cos\theta,\sin\theta)$. It is straightforward to express the potential of the wave in the transverse-traceless gauge as $(\tilde{h}_{ij})={\rm Re}(\tilde{P}_{\pm})$, where \begin{equation} \tilde{P}_{\pm}=(\tilde{\epsilon}_{\oplus} \pm i\tilde{\epsilon}_{\otimes})\hat{h}({\bf k})e^{i({\bf k}\, .\, {\bf x}-\omega t)} \label{b14} \end{equation} and \begin{eqnarray} \tilde{\epsilon}_{\oplus}&=&\left( \begin{array}{ccc} 1 & 0 & 0 \\ 0 & -\cos^{2}\theta & -\sin\theta\cos\theta \\ 0 & -\sin\theta\cos\theta & -\sin^{2}\theta \end{array}\right), \nonumber \\ \tilde{\epsilon}_{\otimes}&=&\left( \begin{array}{ccc} 0 & \cos\theta & \sin\theta \\ \cos\theta & 0 & 0 \\ \sin\theta & 0 & 0 \end{array}\right). \label{b15} \end{eqnarray} The gravitational field measured by the static inertial observers, $\tilde{{\cal R}}$, has the standard form and its electric part is given by $\frac{1}{2}\omega^{2} \tilde{P}_{\pm}$. Similarly, the field according to the special rotating observer ${\cal O}'_{0}$ is $\tilde{{\cal R}}'$ with electric components given by $\frac{1}{2}\omega^{2}\tilde{P}'_{\pm}$. To determine $\tilde{P}'_{\pm}$, let \begin{equation} R_{\Omega}=\left( \begin{array}{ccc} \cos\phi & \sin\phi & 0 \\ -\sin\phi & \cos\phi & 0 \\ 0 & 0 & 1 \end{array}\right) \label{b16} \end{equation} be the rotation matrix that relates the spatial frame $({\bf \hat{x}}',{\bf \hat{y}}',{\bf \hat{z}}')$ of the rotating observer to that of the static inertial observers. We find that \begin{equation} \tilde{P}'_{\pm}=R_{\Omega}\tilde{P}_{\pm}R^{\dagger}_{\Omega}. \label{b17} \end{equation} It then follows after some algebra that $\tilde{P}'_{\pm}$ is given by \begin{equation} \tilde{P}'_{\pm}=\hat{h}({\bf k})\sum^{2}_{m=-2}\mu_{m}{\cal M}^{(m)}e^{-i(\omega-m\Omega)t}, \label{b18} \end{equation} so that in this case the measured frequencies are $\omega'_{0}=\omega-m\Omega$, where $m=0,\pm 1,\pm 2$. Here ${\cal M}^{(m)}$ are $3 \times 3$ symmetric and traceless matrices given by ${\cal M}^{(0)}={\rm diag}(\frac{1}{2},\frac{1}{2},-1)$, \begin{equation} {\cal M}^{(\pm 1)}=\left( \begin{array}{ccc} 0 & 0 & \pm i \\ 0 & 0 & -1 \\ \pm i & -1 & 0 \end{array}\right),\,\ {\cal M}^{(\pm 2)}=\left( \begin{array}{ccc} 1 & \pm i & 0 \\ \pm i & -1 & 0 \\ 0 & 0 & 0 \end{array}\right); \label{b19} \end{equation} moreover, the coefficients $\mu_{m}$ can be expressed as \begin{eqnarray} \mu_{0}&=&\sin^{2}\theta, \,\,\ \mu_{1}=\pm a_{\pm}\sin\theta, \,\,\ \mu_{2}=(a_{\pm})^{2}, ~~~~~~~~ \label{b20} \\ \mu_{-1}&=&\mp a_{\mp}\sin\theta, \,\,\ \mu_{-2}=(a_{\mp})^{2}, \label{b21} \end{eqnarray} where \begin{equation} a_{\pm}=\frac{1}{2}(1\pm \cos\theta). \label{b22} \end{equation} It follows from these results that ${\cal O}'_{0}$ can express its measurements as a Fourier sum of frequencies $\omega'_{0}=\omega-m\Omega$ with amplitudes given in Eqs. (\ref{b18})-(\ref{b22}). It is reasonable to assume that the weight $W_{m}$, or intensity, assigned to each Fourier component is given by the sum of the squares of the absolute magnitudes of the elements of the corresponding matrix. Thus, $W_{m}=|\hat{h}({\bf k})|^{2}w_{m}$, where \begin{eqnarray} w_{0}&=&\frac{3}{2}\sin^{4}\theta, \,\, w_{1}=4(a_{\pm})^{2}\sin^{2}\theta, \,\, w_{2}=4(a_{\pm})^{4}, ~~~~~~~~~~ \label{b23} \\ w_{-1}&=&4(a_{\mp})^{2}\sin^{2}\theta, \,\,\,\,\,\,\,\,\,\,\,\ w_{-2}=4(a_{\mp})^{4}, \label{b24} \end{eqnarray} and one can show that \begin{equation} \sum_{m}w_{m}=4. \label{b25} \end{equation} Let us now define the {\it relative} weight of each frequency $(\omega-m\Omega)$ in the Fourier sum to be \begin{equation} \wp_{m}=\frac{W_{m}}{\sum_{m} W_{m}}=\frac{1}{4}w_{m}; \label{b26} \end{equation} therefore, the average frequency measured by the observer ${\cal O}'_{0}$ can be computed and the result is \begin{equation} \langle \omega'_{0} \rangle = \sum_{m}(\omega-m\Omega)\wp_{m}=\omega \mp 2\Omega \cos\theta. \label{b27} \end{equation} These considerations can be given a proper physical interpretation based upon the eigenstates of a particle with spin $2\hbar$. That is, according to the representation theory of the rotation group, the eigenstates of the particle with respect to the coordinate system $({\bf \hat{x}},{\bf \hat{n}},{\bf \hat{k}})$ can be transformed to the $({\bf \hat{x}},{\bf \hat{y}},{\bf \hat{z}})$ system using the matrix $({\cal D}^{(j)}_{mm'})$ for $j=2$ \cite{Edmonds}. Each element of this matrix is given, up to a phase factor, by $d^{(j)}_{mm'}(\theta)$ with $j=2$. This latter matrix can be expressed \cite{Edmonds} as \begin{equation} \left( \begin{array}{ccccc} a^{2}_{+} & -b_{+} & b_{0} & -b_{-} & a^{2}_{-} \\ b_{+} & c_{+} & c_{0} & c_{-} & -b_{-} \\ b_{0} & -c_{0} & a_{0} & c_{0} & b_{0} \\ b_{-} & c_{-} & -c_{0} & c_{+} & -b_{+} \\ a^{2}_{-} & b_{-} & b_{0} & b_{+} & a^{2}_{+} \end{array}\right), \label{b28} \end{equation} where $a_{+}$ and $a_{-}$ are given by Eq. (\ref{b22}) and \begin{equation} a_{0}=\frac{1}{4}(1+3\cos2\theta), \,\,\ b_{0}=\frac{\sqrt{6}}{4}\sin^{2}\theta, \,\,\ b_{\pm}=-a_{\pm}\sin\theta, ~~~~~~~~~~~ \label{b29} \end{equation} \begin{equation} c_{0}=\frac{\sqrt{6}}{4}\sin2\theta, \,\,\,\,\,\,\,\,\ c_{\pm}=\frac{1}{2}(\cos\theta \pm \cos2\theta). \label{b30} \end{equation} Consider an incident graviton with definite helicity $\pm 2\hbar$ as in Eq. (\ref{b14}). The probability amplitude that the graviton has an angular momentum $m\hbar$ along the $z$ axis, corresponding to a frequency $\omega'_{0}=\omega-m\Omega$ as measured by ${\cal O}'_{0}$, is given up to a phase factor by $\zeta_{m}$, where \begin{equation} (\zeta_{m})=\left( \begin{array}{c} a^{2}_{\pm} \\ \pm b_{\pm} \\ b_{0} \\ \pm b_{\mp} \\ a^{2}_{\mp} \end{array}\right). \label{b31} \end{equation} Here the upper (lower) sign refers to an initial incident state of positive (negative) helicity. Equation (\ref{b31}) is obtained from the first and last columns of the matrix (\ref{b28}), since the state of the particle in the $({\bf \hat{x}},{\bf \hat{y}},{\bf \hat{z}})$ system is obtained from the application of the matrix (\ref{b28}) on the helicity states of the incident graviton. It follows that the probability that an incident graviton of helicity $\pm 2\hbar$ has spin $m\hbar$, $m=0,\pm 1,\pm 2$, along the direction of rotation of the observer is given by $|\zeta_{m}|^{2}$, where \begin{equation} \wp_{m}=|\zeta_{m}|^{2} \label{b32} \end{equation} based upon the comparison of Eqs. (\ref{b26}) and (\ref{b31}). Thus $\wp_{m}$ is indeed the probability that observer ${\cal O}'_{0}$ would measure frequency $\omega - m\Omega$ and hence the average measured frequency is in fact $\langle \omega'_{0} \rangle$ given by Eq. (\ref{b27}). Introducing the helicity vector ${\bf \hat{H}}=\pm {\bf \hat{k}}$, the average frequency measured by the observer can be written as $\langle \omega'_{0} \rangle=\omega-2{\bf \hat{H}}\, .\, {\bf \Omega}$, which is another expression of helicity-rotation coupling. In view of Eq. (\ref{b12}), we may interpret the expression for $\langle \omega'_{0} \rangle$ as follows: the rotation frequency of the observer ${\bf \Omega}=\Omega {\bf \hat{z}}$ may be decomposed into a component of magnitude $\Omega \cos\theta$ parallel to the wave vector ${\bf k}$ and a component of magnitude $\Omega \sin\theta$ perpendicular to ${\bf k}$. On the average, the latter component does not contribute to the measured frequency; hence $\langle \omega'_{0} \rangle=\omega \mp 2(\Omega\cos\theta)$ in agreement with Eq. (\ref{b12}). The average frequency measured by the observer is expected to be the same as the result that would be obtained in the JWKB regime in accordance with the quasi-classical approximation. In the case under consideration, this corresponds to the high-frequency regime for gravitational waves; that is, waves with $\omega\gg\Omega$. To see how this comes about explicitly, we follow the approach that has been developed for the electromagnetic case \cite{Mash1} and adapt it to the gravitational case under consideration here. It is first necessary to define a triad $({\bf \hat{\alpha}},{\bf \hat{\beta}},{\bf \hat{k}})$ that can provide a more natural polarization basis for the rotating observer ${\cal O}'_{0}$ such that ${\bf \hat{\alpha}}$ and ${\bf \hat{\beta}}$ remain ``fixed'' in the rotating frame as much as possible. These unit vectors are defined by \begin{eqnarray} {\bf \hat{\alpha}}&=&{\bf \hat{x}}\cos\Phi + {\bf \hat{n}}\sin\Phi, \label{b33} \\ {\bf \hat{\beta}}&=&-{\bf \hat{x}}\sin\Phi + {\bf \hat{n}}\cos\Phi, \label{b34} \end{eqnarray} where $\Phi$ is given by \begin{equation} \sin\Phi=\frac{1}{D}\cos\theta\sin\Omega t, \,\,\,\,\ \cos\Phi=\frac{1}{D}\cos\Omega t \label{b35} \end{equation} and $D>0$ can be obtained from \begin{equation} D^{2}=\cos^{2}\theta + \sin^{2}\theta \cos^{2}\Omega t. \label{b36} \end{equation} We note that $\Phi$ reduces to $\Omega t$ for $\theta=0$ and to $-\Omega t$ for $\theta=\pi$, while $\Phi=0$ for $\theta=\pi/2$. The orthonormal triad $({\bf \hat{\alpha}},{\bf \hat{\beta}},{\bf \hat{k}})$ is related to $({\bf \hat{x}},{\bf \hat{y}},{\bf \hat{z}})$ by a rotation; in fact, the transformation that takes $({\bf \hat{x}},{\bf \hat{y}},{\bf \hat{z}})$ to $({\bf \hat{\alpha}},{\bf \hat{\beta}},{\bf \hat{k}})$ is given by \begin{equation} T=\left( \begin{array}{ccc} \cos\Phi & \cos\theta\sin\Phi & \sin\theta\sin\Phi \\ -\sin\Phi & \cos\theta\cos\Phi & \sin\theta\cos\Phi \\ 0 & -\sin\theta & \cos\theta \end{array}\right). \label{b37} \end{equation} We wish to find the new polarization basis for linearly polarized gravitational waves based on the new triad; this can be achieved by a similarity transformation of the basis given in Eq. (\ref{b6}), and the result is \begin{equation} \xi=T^{\dagger}\epsilon_{\oplus}T, \,\,\,\,\,\,\ \nu=T^{\dagger}\epsilon_{\otimes}T. \label{b38} \end{equation} It is possible to connect this new basis with Eq. (\ref{b15}) as follows \begin{eqnarray} \xi&=&\tilde{\epsilon}_{\oplus} \cos2\Phi + \tilde{\epsilon}_{\otimes} \sin2\Phi, \label{b39} \\ \nu&=&-\tilde{\epsilon}_{\oplus} \sin2\Phi + \tilde{\epsilon}_{\otimes} \cos2\Phi. \label{b40} \end{eqnarray} The rotation by $2\Phi$ takes a simple form for the circular polarization basis, namely, \begin{equation} \xi \pm i \nu = (\tilde{\epsilon}_{\oplus} \pm i\tilde{\epsilon}_{\otimes})e^{\mp 2i\Phi}. \label{b41} \end{equation} Using this relation in Eq. (\ref{b14}), we have \begin{equation} \tilde{P}_{\pm} = (\xi \pm i\nu)\hat{h}({\bf k})e^{-i\omega t \pm 2i\Phi}, \label{b42} \end{equation} so that from Eq. (\ref{b17}), \begin{equation} \tilde{P}'_{\pm}=(\xi' \pm i\nu')\hat{h}({\bf k})e^{-i\omega t \pm 2i\Phi}, \label{b43} \end{equation} where the new polarization basis involving \begin{equation} \xi'=R_{\Omega}\xi R^{\dagger}_{\Omega}, \,\,\,\,\ \nu'=R_{\Omega}\nu R^{\dagger}_{\Omega}, \label{b44} \end{equation} is naturally adapted to the rotating observer in the sense that the temporal dependence in Eq. (\ref{b43}) has been transferred to the phase of the wave as much as possible. Let us note that $\xi'$ and $\nu'$ are obtained from $\epsilon_{\oplus}$ and $\epsilon_{\otimes}$, respectively, by a unitary transformation involving the orthogonal matrix $R_{\Omega}T^{\dagger}$. According to Eq. (\ref{b43}), observer ${\cal O}'_{0}$ receives a circularly polarized wave with a phase $-\omega t \pm 2\Phi$. The frequency of the wave is defined to be the negative gradient of the phase with respect to time; hence, $\omega'_{0}=\omega \mp 2\partial \Phi / \partial t$. It follows from Eq. (\ref{b36}) that \begin{equation} \frac{\partial \Phi}{\partial t}=\frac{\Omega \cos\theta}{D^{2}}. \label{b45} \end{equation} In practice, the frequency determination would necessitate the reception of at least a few oscillations of the wave. During such a period of time $t$, assuming that the observations begin at $t=0$, $\epsilon'=\Omega t \ll 1$ due to the fact that $\Omega \ll \omega$; moreover, \begin{equation} D^{-2}=1+\epsilon'^{2}\sin^{2}\theta + O(\epsilon'^{4}) \label{b46} \end{equation} by Eq. (\ref{b36}). Thus $\omega'_{0}=\omega \mp 2\Omega\cos\theta$ in the high-frequency regime. This result is equivalent to the intuitive expectation that from the standpoint of the rotating observer, the spin of the radiation field should precess in the opposite sense. This circumstance can be restated in terms of the rotation of the state of linear polarization of the gravitational wave as explained in Section IV. Henceforward, we limit our considerations to the high-frequency regime $(\omega \gg \Omega)$. It is then possible to generalize the main result that we have obtained for the fixed rotating observers ${\cal O}'_{0}$ to the case of arbitrary rotating observers and thereby determine the modifications in the Doppler effect and aberration that are brought about by the helicity-rotation coupling. This is done in the next section. \renewcommand{\theequation}{3.\arabic{equation}} \setcounter{equation}{0} \section*{III. MODIFIED DOPPLER AND ABERRATION FORMULAS FOR GRAVITATIONAL WAVES} In a background global inertial frame, we first consider the class of noninertial observers that are at rest in the inertial frame but refer their measurements to rotating axes as discussed in Section II. Each of these observers carries a tetrad frame of the form \begin{eqnarray} {\lambda}'^{\mu}_{(0)}&=&(1,0,0,0), \,\,\ {\lambda}'^{\mu}_{(1)}=(0,\cos\phi,\sin\phi,0), \label{c1} \\ {\lambda}'^{\mu}_{(2)}&=&(0,-\sin\phi,\cos\phi,0), \,\,\ {\lambda}'^{\mu}_{(3)}=(0,0,0,1), ~~~~~~~ \label{c2} \end{eqnarray} where $\phi=\Omega t$ as before. We are interested in the determination of frequency and wave vector of incident high-frequency $(\omega\gg\Omega)$ gravitational waves. To this end, we focus attention on the noninertial observer ${\cal O}'_{0}$ at the origin of spatial coordinates and imagine the class of observers that are at the rest in the rotating system of ${\cal O}'_{0}$, each with a tetrad of the form of Eqs. (\ref{b1})-(\ref{b4}). The gravitational field measured by this class of rotating observers is given by $\tilde{{\cal R}}'={\cal L}\tilde{{\cal R}}{\cal L}^{\dagger}$. A complete Fourier analysis of $\tilde{{\cal R}}'$ in time and space is needed to determine the frequency and wave-vector content of the incident radiation according to ${\cal O}'_{0}$. In the high-frequency regime $(\omega \gg \Omega)$, the measurements of ${\cal O}'_{0}$ can be restricted in space to the cylindrical domain of radius $\ll \Omega^{-1}$. The analysis of frequency determination for $\omega\gg\Omega$ has been given in Section II and a corresponding analysis of the wave-vector determination will not be carried out here, since it is entirely analogous to the electromagnetic case presented in detail in Ref. \cite{Hauck}. It follows from an analysis similar to the one given in Section 2 of Ref. \cite{Hauck} that ${\bf k}'_{0}={\bf k}_{0}$. Thus we conclude that to lowest order \begin{equation} \omega'_{0}=\omega - s{\bf \hat{H}}\, .\, {\bf \Omega} , \,\,\,\ {\bf k}'_{0}={\bf k} \label{c3} \end{equation} for ${\cal O}'_{0}$ in the high-frequency regime. The dispersion relation for ${\cal O}'_{0}$ is then $\omega'_{0}=k'_{0}\mp s{\bf \hat{k}}'_{0}\, . \, {\bf \Omega}$, where $k'_{0}=|{\bf k}'_{0}|$. Here $s=1$ for electromagnetic waves and $s=2$ for gravitational waves. Indeed, these results hold for each member of the rotating class of observers that are at rest and carry the tetrad frame $\lambda'^{\mu}_{(\alpha)}$ given by Eqs. (\ref{c1})-(\ref{c2}). The generalization of Eq. (\ref{c3}) to the case of rotating observers that are not at rest in the background global inertial frame can be simply obtained from the observation that at each event along the circular path of an observer ${\cal O}'$, its tetrad $\Lambda^{\mu}_{(\alpha)}$ is related to the tetrad $\lambda'^{\mu}_{(\alpha)}$ of the static rotating observer at that event by a Lorentz boost \begin{eqnarray} \Lambda^{\mu}_{(0)}=\gamma[\lambda'^{\mu}_{(0)}+v\lambda'^{\mu}_{(2)}], \,\,\ \Lambda^{\mu}_{(1)}=\lambda'^{\mu}_{(1)}, \label{c4} \\ \Lambda^{\mu}_{(2)}=\gamma[\lambda'^{\mu}_{(2)}+v\lambda'^{\mu}_{(0)}], \,\,\ \Lambda^{\mu}_{(3)}=\lambda'^{\mu}_{(3)}. \label{c5} \end{eqnarray} It follows from Lorentz invariance that $(\omega',{\bf k}')$ for ${\cal O}'$ are related to $(\omega'_{0},{\bf k}'_{0})$ by the standard Doppler and aberration formulas; that is \begin{eqnarray} \omega'&=&\gamma(\omega'_{0}-{\bf v}\, .\, {\bf k}'_{0}), \label{c6} \\ {\bf k}'&=&{\bf k}'_{0}+\frac{(\gamma-1)}{v^{2}}({\bf v}\, .\, {\bf k}'_{0}){\bf v}-\gamma\omega'_{0}{\bf v}. \label{c7} \end{eqnarray} Substituting Eq. (\ref{c3}) in Eqs. (\ref{c6})-(\ref{c7}), we obtain the modified Doppler and aberration formulas for gravitational waves in the high-frequency regime \begin{eqnarray} \omega'&=&\gamma[(\omega-s{\bf \hat{H}}\, .\, {\bf \Omega})-{\bf v}\, .\, {\bf k}], \label{c8} \\ {\bf k}'&=&{\bf k}+\frac{(\gamma-1)}{v^{2}}({\bf v}\, .\, {\bf k}){\bf v}-\gamma(\omega-s{\bf \hat{H}}\, .\, {\bf \Omega}){\bf v}. ~~~~~~~ \label{c9} \end{eqnarray} Eq. (\ref{c8}) may be interpreted in terms of Eq. (\ref{b13}) in the eikonal approximation, namely, \begin{equation} \omega'=\gamma(\omega-{\bf j}\, .\, {\bf \Omega}), \,\,\,\ {\bf j}={\bf r}\times {\bf k}+s{\bf \hat{H}}, \label{c10} \end{equation} where $\hbar{\bf j}$ is the total angular momentum of the graviton $(s=2)$ or the photon $(s=1)$. The results of this section for $s=2$ are expected to be of importance in experiments involving the reception of gravitational waves by antennas rotating with frequency $\Omega\ll\omega$. Clearly, current large-scale Earth-fixed gravitational-wave antennas rotate with the frequency of rotation of the Earth. \renewcommand{\theequation}{4.\arabic{equation}} \setcounter{equation}{0} \section*{IV. HELICITY-GRAVITY COUPLING} It is possible to employ Einstein's principle of equivalence in order to extend the helicity-rotation coupling discussed in previous sections to the propagation of gravitational waves in a curved spacetime background. For this purpose, we consider a solution of the linearized gravitational field equations. We assume that this background field is due to isolated gravitating sources that move slowly compared to the speed of light in vacuum. It is possible to describe such a background field in terms of gravitoelectromagnetism (``GEM'') in close analogy with Maxwell's electrodynamics. GEM has been thoroughly reviewed in Ref. \cite{Mash3} and references cited therein. In this approach to GEM, the metric of the curved spacetime background is of the form \begin{equation} -(1-2f)dt^{2}-4({\bf S}\, .\, d{\bf x})dt+(1+2f)\delta_{ij}dx^{i}dx^{j}, \label{d1} \end{equation} where in our convention $f$ and ${\bf S}$ are respectively the gravitoelectric and gravitomagnetic potentials subject to a (``Lorentz'') gauge condition \begin{equation} \frac{\partial f}{\partial t}+{\bf \nabla}\, .\, (\frac{1}{2}{\bf S})=0. \label{d2} \end{equation} These potentials are connected to sources via the linearized gravitational field equations. For the treatment in this section and in conformity with our general linear approach, we assume that in the presence of an incident gravitational wave the deviation from the Minkowski spacetime is a linear superposition of the perturbations due to the isolated sources and the incident wave. The GEM fields are defined by \begin{equation} {\bf F}=-{\bf \nabla}f-\frac{1}{2}\frac{\partial {\bf S}}{\partial t}, \,\,\,\ {\bf B}={\bf \nabla}\times {\bf S}, \label{d3} \end{equation} in our special convention \cite{Mash3} that has been designed to provide the closest possible connection with the standard formulas of classical electrodynamics. Eqs. (\ref{d2}) and (\ref{d3}) together with the linearized gravitational field equations lead to the Maxwell equations for the GEM fields \cite{Mash3}. It follows from a detailed discussion of Einstein's principle of equivalence within the context of GEM that the gravitoelectric field is in effect locally equivalent to a translationally accelerated system, while the gravitomagnetic field is in effect locally equivalent to a rotating system \cite{Mash3}. Traditionally, Einstein's heuristic principle of equivalence refers to the accelerated ``elevator'' in relation with the gravitoelectric field of the source; however, the rotation of the elevator is in general necessary as well to take due account of the corresponding gravitomagnetic field. In keeping with the electromagnetic analogy, this application of Einstein's principle of equivalence is the content of the gravitational Larmor theorem \cite{Larmor, Mas}. It turns out that a spinning particle at rest in the exterior field of a rotating mass precesses with frequency ${\bf B}$, which is equivalent to what would be observed from a local frame of reference rotating with frequency ${\bf \Omega}=-{\bf B}$. Following this general line of thought, we may conclude from the results of Section III that for observers at rest in the exterior GEM background, the local dispersion relation \begin{equation} \omega=k \pm s{\bf \hat{k}}\, .\, {\bf B} \label{d4} \end{equation} is approximately valid for high-frequency incident gravitational ($s=2$) or electromagnetic ($s=1$) waves. Eq. (\ref{d4}) follows from Eq. (\ref{c3}) with ${\bf \Omega} \rightarrow-{\bf B}$ in accordance with the gravitational Larmor theorem. Recognizing that the {\it local} dispersion relation (\ref{d4}) ignores the usual global GEM effects, such as the bending of the incident beam of radiation, one may nevertheless employ Eq. (\ref{d4}) globally in order to uncover the specific consequences of helicity-gravity coupling. The results may then be superposed on the standard GEM effects in line with our general linear perturbative approach. An interesting consequence of the coupling of helicity with the gravitomagnetic field is the rotation of the state of linear polarization of a gravitational wave that propagates in the field of a rotating mass. To illustrate this effect, we assume that the background field is {\it stationary} and is due to a rotating astronomical source. The GEM fields have been determined in this case \cite{Tey}.The curl of the gravitomagnetic field ${\bf B}$ vanishes in the exterior of the source; hence, \begin{equation} {\bf B}=-{\bf \nabla}Q, \label{d5} \end{equation} where $Q$ is the gravitomagnetic scalar potential. Far from the source \begin{eqnarray} f &\sim& \frac{GM}{r}, \,\,\,\ {\bf S} \sim \frac{G {\bf J}\times {\bf r}}{r^{3}},\label{d6}\\ Q &\sim& \frac{G {\bf J}\, .\, {\bf r}}{r^{3}}, \,\,\,\,\ {\bf B} \sim \frac{GJ}{r^{3}}[3({\bf \hat{J}}\, .\, {\bf \hat{r})}{\bf \hat{r}}-{\bf \hat{J}}], \label{d7} \end{eqnarray} where $M$ is the mass and ${\bf J}=J{\bf \hat{z}}$ is the angular momentum of the source. Consider a linearly polarized gravitational wave starting at $z=z_{0}$ far from the source and propagating outward along its rotation axis. Let $\Pi^{\mu\nu}_{1}$ and $\Pi^{\mu\nu}_{2}$ be the linear polarization tensors for the wave. Assuming that at $z=z_{0}$ the state of the wave is given by the real part of $\psi^{\mu\nu}=\hat{\psi}\Pi^{\mu\nu}_{1}{\rm exp}(-i\omega t)$, where $\hat{\psi}, |\hat{\psi}| \ll 1$, is a constant amplitude, then for any $z$ \begin{equation} \psi^{\mu\nu}=\frac{1}{2}\hat{\psi}[(\Pi^{\mu\nu}_{1}+i\Pi^{\mu\nu}_{2})e^{i{\cal S}_{+}-i\omega t}+(\Pi^{\mu\nu}_{1}-i\Pi^{\mu\nu}_{2})e^{i{\cal S}_{-}-i\omega t}]. \label{d8} \end{equation} Here ${\cal S}$ is given by \begin{equation} {\cal S}_{\pm}=\int^{z}_{z_{0}}k_{\pm}dz, \label{d9} \end{equation} where $k_{+}$ and $k_{-}$ are the wave numbers of the positive and negative-helicity components of the gravitational wave. Expressing ${\cal S}_{\pm}$ as \begin{equation} {\cal S}_{+}={\cal S}_{0}-\Delta, \,\,\,\ {\cal S}_{-}={\cal S}_{0}+\Delta, \label{d10} \end{equation} we find that $\psi^{\mu\nu}$ can be written as \begin{equation} \psi^{\mu\nu}=\hat{\psi}(\Pi^{\mu\nu}_{1}\cos\Delta + \Pi^{\mu\nu}_{2}\sin\Delta)e^{i{\cal S}_{0}-i\omega t}. \label{d11} \end{equation} Inspection of this equation reveals that as the wave propagates, the linear polarization state rotates by an angle $\Delta$ given by \begin{equation} \Delta=\frac{1}{2}\int^{z}_{z_{0}}(k_{-}-k_{+})dz. \label{d12} \end{equation} To compute this angle, we write Eq. (\ref{d4}) for the positive and negative helicity components of the wave, \begin{equation} \omega=k_{+}+sB_{z}, \,\,\,\,\ \omega=k_{-}-sB_{z}, \label{d13} \end{equation} so that we find \begin{equation} \Delta=s\int^{z}_{z_{0}}B_{z}dz=s[Q(z_{0})-Q(z)]. \label{d14} \end{equation} Using Eq. (\ref{d7}), we finally have \begin{equation} \Delta=sGJ(\frac{1}{z^{2}_{0}}-\frac{1}{z^{2}}). \label{d15} \end{equation} For electromagnetic radiation $(s=1)$, this gravitomagnetic rotation of the plane of polarization was first studied by Skrotskii \cite{Skr}; detailed discussions and references are contained in Ref. \cite{Kop} and references cited therein. The angle of rotation for gravitational radiation $(s=2)$ is twice the Skrotskii angle. Let us note that $\Delta$ vanishes for waves propagating from $-\infty$ to $+\infty$ along the $z$ axis; this is an instance of a general result discussed in Appendix A, where the consequences of Eq. (\ref{d4}) are worked out in a more general context. The angle of rotation of the state of linear polarization $\Delta$ is independent of the frequency (or wavelength) of the radiation; therefore, Eq. (\ref{d15}) is valid in the limit of vanishing wavelength, namely, the JWKB (or eikonal) limit. In this limit, gravitational waves propagate along a null geodesic; in this case, a full treatment is contained in the next section. \renewcommand{\theequation}{5.\arabic{equation}} \setcounter{equation}{0} \section*{V. ROTATION OF LINEAR POLARIZATION} The purpose of this section is to compute (within the eikonal approximation scheme) the rate of rotation of the state of linear polarization of high-frequency gravitational waves propagating in the exterior background field of a rotating astronomical source. Consider the propagation of gravitational radiation on a background spacetime such that the waves cause a small perturbation. Let $\bar{g}_{\mu\nu}(x)$ be the metric tensor of the background field in a given coordinate system and $g_{\mu\nu}=\bar{g}_{\mu\nu}+h_{\mu\nu}$ be the spacetime metric tensor. Under an infinitesimal coordinate transformation $x'^{\mu}=x^{\mu}-\epsilon^{\mu}(x)$, \begin{equation} h'_{\mu\nu}(x)=h_{\mu\nu}(x)+\epsilon_{\mu | \nu}+\epsilon_{\nu | \mu}, \label{e1} \end{equation} where the vertical bar denotes covariant differentation with respect to $\bar{g}_{\mu\nu}$. We also raise and lower indices, etc., with $\bar{g}_{\mu\nu}$. Thus the perturbation is determined up to a gauge transformation given by Eq. (\ref{e1}). Introducing the trace-reversed potential \begin{equation} \psi_{\mu\nu}=h_{\mu\nu}-\frac{1}{2}\bar{g}_{\mu\nu}\bar{g}^{\rho\sigma}h_{\rho\sigma}, \label{e2} \end{equation} we find that under a gauge transformation \begin{equation} \psi'_{\mu\nu}=\psi_{\mu\nu}+\epsilon_{\mu | \nu}+\epsilon_{\nu | \mu}-\bar{g}_{\mu\nu}\, \epsilon^{\sigma}_{\,\ | \sigma}. \label{e3} \end{equation} It is convenient to impose the transverse gauge condition \begin{equation} \psi^{\mu\nu}_{\,\,\,\,\ | \nu}=0, \label{e4} \end{equation} which does not fix the gauge completely, however. It turns out that any solution of \begin{equation} \epsilon^{\mu \,\,\ \nu}_{\,\ | \nu}+\bar{R}^{\mu}_{\,\ \sigma}\epsilon^{\sigma}=0 \label{e5} \end{equation} in Eq. (\ref{e3}) leads to $\psi'^{\mu\nu}_{\,\,\,\,\,\,\ | \nu}=0$ if Eq. (\ref{e4}) is assumed. In the transverse gauge, the gravitational field equations imply that $\psi_{\mu\nu}$ satisfies the wave equation \cite{Eisenhart} \begin{equation} \psi^{\,\,\,\,\,\,\,\,\,\,\ \sigma}_{\mu\nu | \sigma}+2\bar{R}_{\mu\rho\nu\sigma}\psi^{\rho\sigma}=0, \label{e6} \end{equation} where $\bar{g}_{\mu\nu}$ is assumed to be Ricci-flat in the spacetime region under consideration here. To describe the propagation of the wave function $\psi_{\mu\nu}$ in the eikonal approximation, we seek a solution of Eq. (\ref{e6}) in the form \begin{equation} \psi_{\mu\nu}={\rm Re}\{\tilde{\chi}_{\mu\nu}(x;\epsilon)e^{i \epsilon^{-1}\sigma(x)}\}, \label{e7} \end{equation} where $\epsilon$, $0<\epsilon \ll 1$, is directly proportional to the wavelength of the radiation. In the eikonal approximation, $\tilde{\chi}_{\mu\nu}(x;\epsilon)$ is expressed as an asymptotic series in powers of $\epsilon$ \begin{equation} \tilde{\chi}_{\mu\nu}(x;\epsilon)=\chi_{\mu\nu}(x)+\epsilon \rho_{\mu\nu}(x)+\epsilon^{2}\kappa_{\mu\nu}(x)+.... \label{e8} \end{equation} Let $k_{\mu}=\partial \sigma(x) / \partial x^{\mu}$ be the propagation vector of the wave; then, the substitution of equations (\ref{e7}) and (\ref{e8}) in the gauge condition (\ref{e4}) and propagation equation (\ref{e6}) results in series that contain powers of $1/ \epsilon$ in addition to powers of $\epsilon$. It follows from Eq. (\ref{e4}) that there is only one such term involving $1/ \epsilon$ and in the eikonal limit $(\epsilon \rightarrow 0)$, the coefficient of this term must vanish; therefore, \begin{equation} \chi_{\mu\nu}k^{\nu}=0. \label{e9} \end{equation} Moreover, the propagation equation (\ref{e6}) involves $1/ \epsilon^{2}$ and $1/ \epsilon$ terms and the coefficients of these terms must also vanish in the eikonal limit, hence we have respectively \begin{equation} k_{\mu}k^{\mu}=0, \label{e10} \end{equation} and \begin{equation} i(2\chi_{\mu\nu | \sigma}k^{\sigma}+\chi_{\mu\nu}k^{\sigma}_{\,\ | \sigma})-k^{\sigma}k_{\sigma}\rho_{\mu\nu}=0. \label{e11} \end{equation} Let us first note that equation (\ref{e10}) implies that the radiation follows a null geodesic in the eikonal limit. It follows from $k_{\mu}=\partial \sigma(x) / \partial x^{\mu}$ that $k_{\mu | \nu}=k_{\nu | \mu}$. Thus, taking covariant derivative of Eq. (\ref{e10}), we get that $k_{\mu | \nu}k^{\mu}=0$; hence, \begin{equation} k_{\nu | \mu}k^{\mu}=0. \label{e12} \end{equation} The geodesic equation follows from Eq. (\ref{e12}) and $k^{\mu}=dx^{\mu}/ d \lambda$, where $\lambda$ is an affine parameter along the path. Equations (\ref{e9}) and (\ref{e11}) describe the propagation of the wave amplitude $\chi_{\mu\nu}$ along the null geodesic, since it follows from Eq. (\ref{e11}) that \begin{equation} 2\chi_{\mu\nu | \sigma}k^{\sigma}+\chi_{\mu\nu}k^{\sigma}_{\,\ | \sigma}=0. \label{e13} \end{equation} An immediate consequence of this relation is that $\Sigma^{0}=\chi^{*}_{\mu\nu}\chi^{\mu\nu}$ satisfies the conservation law \begin{equation} (\Sigma^{0}k^{\sigma})_{ | \sigma}=0, \label{e14} \end{equation} which can be interpreted as the conservation of the ``graviton'' number along the null geodesic congruence. These results have been based on terms involving $1/ \epsilon$ and $1/ \epsilon^{2}$; taking account of the other terms in the eikonal series, i.e., those involving $\epsilon^{n}, n=0,1,2,...$, would simply specify the manner in which the general wave amplitude $\tilde{\chi}_{\mu\nu}(x;\epsilon)$ varies along the null geodesic. This eikonal (or JWKB) treatment of gravitational radiation has been previously considered in Ref. \cite{Isaacson}. A critical assessment of the eikonal approximation scheme is contained in Ref. \cite{Mash4}. In the eikonal approximation scheme, the curves $x^{\mu}=x^{\mu}(\lambda)$ that have $k^{\mu}=d x^{\mu}/ d \lambda$ as tangent vectors are null geodesics orthogonal to the surfaces of constant phase $\sigma$. Imagine a bundle of such null rays in a congruence characterized by the propagation vector $k^{\mu}$. Let us define a null tetrad system $(k^{\mu},l^{\mu}, n^{\mu}, n^{* \mu})$ such that $k^{\mu}l_{\mu}=-1$ and $n^{\mu}n^{*}_{\mu}=1$ are the only nonvanishing scalar products among the four null vectors. Starting from an observer's orthonormal tetrad frame $\lambda^{\mu}_{(\alpha)}$, the null frame is constructed as follows: \begin{eqnarray} k^{\mu}=\frac{a}{\sqrt{2}}[\lambda^{\mu}_{(0)}+\lambda^{\mu}_{(3)}] , \,\ l^{\mu}=\frac{1}{a\sqrt{2}}[\lambda^{\mu}_{(0)}-\lambda^{\mu}_{(3)}] ~~~~~~~~~ \label{e15} \end{eqnarray} are real, while $n^{\mu}$ and its complex conjugate $n^{* \mu}$ are complex, since $n^{\mu}$ is defined by \begin{equation} n^{\mu}=\frac{1}{\sqrt{2}}[\lambda^{\mu}_{(1)}+i\lambda^{\mu}_{(2)}]. \label{e16} \end{equation} Here $a=-\sqrt{2}\, k_{\mu}\lambda^{\mu}_{(0)}$ is a nonzero constant. The tetrad system is assumed to be parallel propagated along the congruence. The symmetric and transverse tensor $\chi_{\mu\nu}$ can be locally expressed in terms of the parallel-propagated null tetrad as \begin{equation} \chi_{\mu\nu}=\Phi_{+}n_{\mu}n_{\nu}+\Phi_{-}n^{*}_{\mu}n^{*}_{\nu}+k_{\mu}\Gamma_{\nu}+k_{\nu}\Gamma_{\mu}. \label{e17} \end{equation} It turns out that in this expansion $n_{\mu}n_{\nu}$ corresponds to a positive helicity wave and $n^{*}_{\mu}n^{*}_{\nu}$ corresponds to a negative helicity wave; thus, $\chi_{\mu\nu}$ consists in general of a positive helicity part with amplitude $\Phi_{+}$, a negative helicity part with amplitude $\Phi_{-}$ and a gauge part involving a transverse vector $\Gamma_{\mu}$ such $k^{\mu}\Gamma_{\mu}=0$ (see the appendix of Ref. \cite{Mash5}). We note that $\langle \psi^{\mu\nu}\psi_{\mu\nu}\rangle=\Sigma^{0}/2$ and \begin{equation} \Sigma^{0}=\chi^{\mu\nu}\chi^{*}_{\mu\nu}=|\Phi_{+}|^{2}+|\Phi_{-}|^{2}, \label{e18} \end{equation} so that the intensity of the wave depends only on its {\it irreducible} part \begin{equation} \hat{\psi}_{\mu\nu}=\Phi_{+}n_{\mu}n_{\nu}+\Phi_{-}n^{*}_{\mu}n^{*}_{\nu}. \label{e19} \end{equation} Using this irreducible part of $\chi_{\mu\nu}$, we find from Eq. (\ref{e13}) that \begin{equation} \frac{d \Phi_{+}}{d\lambda}+\hat{\theta}\, \Phi_{+}=0, \,\,\ \frac{d \Phi_{-}}{d\lambda}+\hat{\theta}\, \Phi_{-}=0, \label{e20} \end{equation} where $\hat{\theta}=(1/2)k^{\sigma}_{\,\ | \sigma}$ is the {\it expansion} of the null congruence $k^{\mu}$. Let $A$ be the area of the cross section of a bundle of null rays in the congruence; then, \begin{equation} \frac{d A}{d\lambda}=2\hat{\theta} A. \label{e21} \end{equation} It follows that $|\Phi_{+}|^{2} A$ and $|\Phi_{-}|^{2} A$ are separately conserved along the trajectory; that is, the number of positive or negative helicity null rays (``gravitons'') is independently conserved along the congruence. This approach to the polarization of gravitational waves in the eikonal limit demonstrates that the parallel transport of polarization tensors $n_{\mu}n_{\nu}$ and $n^{*}_{\mu}n^{*}_{\nu}$ for positive and negative helicities, respectively, is a natural interpretation of Eq. (\ref{e13}). This circumstance is a spin-2 analog of the spin-1 electromagnetic case, where the irreducible part of the wave amplitude corresponding to the vector potential is of the form $\phi_{+}n_{\mu}+\phi_{-}n^{*}_{\mu}$. Furthermore, it is possible to define Stokes parameters $\Sigma^{\alpha}$ for gravitational radiation such that $\eta_{\alpha \beta}\Sigma^{\alpha}\Sigma^{\beta}=0$ and \begin{equation} \Sigma^{1}=\Phi_{+}\Phi^{*}_{-}+\Phi_{-}\Phi^{*}_{+}, \,\,\ \Sigma^{2}=-i(\Phi_{+}\Phi^{*}_{-}-\Phi_{-}\Phi^{*}_{+}), \label{a22} \end{equation} \begin{equation} \Sigma^{3}=|\Phi_{+}|^{2}-|\Phi_{-}|^{2}, \label{e23} \end{equation} along the lines developed in Ref. \cite{Mash5}. It is important to discuss the uniqueness of the representation (\ref{e17}). At any given event, the observer tetrad frame $\lambda^{\mu}_{(\alpha)}$ is unique up to a Lorentz transformation. We are interested in a subgroup of the Lorentz group that preserves $k^{\mu}$; in fact, this is the little group of $k^{\mu}$ that is isomorphic to the Euclidean group in the plane. Therefore, under the action of the little group, $k'^{\mu}=k^{\mu}$, \begin{eqnarray} l^{' \mu}&=&l^{\mu}+b\, n^{\mu}+b^{*}n^{* \mu}+|b|^{2}k^{\mu}, \label{e24} \\ n^{' \mu}&=&e^{-i\Theta}(n^{\mu}+b^{*}k^{\mu}). \label{e25} \end{eqnarray} Here $\Theta$ is real and corresponds to the constant angle of rotation in the $(\lambda^{\mu}_{(1)},\lambda^{\mu}_{(2)})$ plane. For $\Theta=0$, Eqs. (\ref{e24}) and (\ref{e25}) with a complex constant $b$ correspond to the Abelian subgroup of the little group under which \begin{equation} \hat{\psi}'_{\mu\nu}=\hat{\psi}_{\mu\nu}+k_{\mu}G_{\nu}+k_{\nu}G_{\mu}, \label{e26} \end{equation} where $k^{\mu}G_{\mu}=0$ and $G_{\mu}$ is given by \begin{equation} G_{\mu}=\Phi_{+}b^{*}n_{\mu}+\Phi_{-}b\, n^{*}_{\mu}+\frac{1}{2}(\Phi_{+}b^{* 2}+\Phi_{-}b^{2})k_{\mu}. \label{e27} \end{equation} Equation (\ref{e26}) amounts to a gauge transformation of $\hat{\psi}_{\mu\nu}$; therefore, under this gauge subgroup the Stokes parameters for the radiation field remain invariant. On the other hand, with $b=0$ and under rotation of angle $\Theta$, $\hat{\psi}_{\mu\nu}$ remains invariant if \begin{equation} \Phi'_{+}=e^{2i\Theta}\Phi_{+}, \,\,\ \Phi'_{-}=e^{-2i\Theta}\Phi_{-} \label{e28} \end{equation} which demonstrates that $\Phi_{+}$ ($\Phi_{-}$) is the amplitude of the graviton in the positive (negative) helicity state. Moreover, the Stokes parameters undergo a rotation as well, since a rotation of the angle $\Theta$ in the ($\lambda^{\mu}_{(1)},\lambda^{\mu}_{(2)}$) plane induces a rotation of angle $-4\Theta$ in the ($\Sigma^{1},\Sigma^{2}$) plane \cite{Mash5}. It proves useful to define real linear polarization tensors $\Pi^{\mu\nu}_{1}$ and $\Pi^{\mu\nu}_{2}$ such that \begin{equation} n^{\mu}n^{\nu}=\frac{1}{\sqrt{2}}(\Pi^{\mu\nu}_{1}+i\Pi^{\mu\nu}_{2}). \label{e29} \end{equation} Here $\Pi_{1}$ and $\Pi_{2}$ are independent polarization states such that $\hat{\psi}^{\mu\nu}$ can be written as \begin{equation} \hat{\psi}^{\mu\nu}=L_{1}\Pi^{\mu\nu}_{1}+L_{2}\Pi^{\mu\nu}_{2}, \label{e30} \end{equation} where $L_{1}$ and $L_{2}$ are the linear polarization amplitudes. Under a constant rotation of angle $\Theta$ in a plane perpendicular to the spatial direction of propagation of the wave, \begin{eqnarray} \Pi'^{\mu\nu}_{1}&=&\Pi^{\mu\nu}_{1}\cos2\Theta + \Pi^{\mu\nu}_{2}\sin2\Theta, \label{e31} \\ \Pi'^{\mu\nu}_{2}&=&-\Pi^{\mu\nu}_{1}\sin2\Theta + \Pi^{\mu\nu}_{2}\cos2\Theta, \label{e32} \end{eqnarray} so that one linear polarization state turns into another under a rotation of $\Theta=\pi/4$. In general, $\Theta=\pi/(2s)$ is the angle ``between'' the linear polarization states $\Pi_{1}$ and $\Pi_{2}$. The constant parameter $a$ in the definition of $k^{\mu}$ and $l^{\mu}$ in Eq. (\ref{e15}) is related to the choice of the affine parameter $\lambda$, which is defined up to a linear transformation $\lambda \rightarrow \lambda'={\rm constant}+\lambda/A_{0}$ where $A_{0}\neq 0$ is a constant. Under this affine transformation $\lambda^{\mu}_{(\alpha)}$ is unchanged, but $a \rightarrow aA_{0}$ and hence $k^{\mu} \rightarrow A_{0}k^{\mu}$, $l^{\mu}\rightarrow A^{-1}_{0}l^{\mu}$ and $n^{\mu}$ is unchanged. A null rotation of the null tetrad is defined to be a 4-parameter group that consists of the combined action of the little group together with an affine transformation. It is the most general transformation that leaves the {\it spatial direction} of wave propagation vector $k^{\mu}$ invariant. It is important to note that under a rotation with angle $\Theta$, the linear polarization states of a massless spin-1 field would rotate by $\Theta$, while that of a spin-2 field would rotate by $2\Theta$. It has been shown in Ref. \cite{Kop} that in a general gravitomagnetic field ${\bf B}$, the plane of linear polarization of electromagnetic radiation rotates by an angle \begin{equation} \alpha_{Skrotskii}=\int {\bf B}\, .\, d{\bf x}, \label{e33} \end{equation} where the integral is evaluated along the spatial path of the null ray; see Section VII of Ref. \cite{Kop} for a detailed derivation. The Skrotskii effect \cite{Skr} is the gravitomagnetic analog of the Faraday effect. Therefore, the state of linear polarization of gravitational radiation would rotate by an angle $2\alpha_{Skrotskii}$ in a general gravitomagnetic field. Thus, we recover, in the eikonal limit, the result of Section IV. \section*{VI. DISCUSSION} It is useful to provide estimates of the spin-rotation-gravity coupling effects presented in this work. For Earth-based gravitational-wave antennas that rotate with the Earth, the effective rotation frequency is therefore about $10^{-5} {\rm Hz}$ in Eqs. (\ref{c8})-(\ref{c9}), so that the incident gravitational waves that would be relevant in this case satisfy the high-frequency condition, namely, $\omega\gg\Omega$. Ignoring the helicity-rotation coupling would introduce a small systematic Doppler bias of magnitude $2\Omega/ \omega$. Let us next consider, as in Section IV, a linearly polarized gravitational wave that propagates outward to infinity starting from the north pole of a rotating astronomical system of radius $R_{0}$. We assume that $\omega \gg B_{0}$, where $B_{0}=2GJ/R^{3}_{0}$ is the gravitomagnetic (Larmor) frequency of the system. The rotation of the linear polarization state of the wave is in the same sense as the rotation of the source and the net angle of rotation is given by Eq. (\ref{d15}), namely, $\Delta=2GJ/R^{2}_{0}$. For a homogeneous sphere of mass $M$ rotating with frequency $\Omega$, $\Delta=0.8GM\Omega$. This amounts to $\Delta \approx 0.025 \, {\rm rad}$ (or about $1.5^{\circ})$ for a millisecond pulsar of mass $M \approx M_{\odot}$ and rotational period $\approx 10^{-3} \, {\rm s}$. For the Earth, however, the corresponding result would be negligibly small, that is, $\Delta \approx 10^{-15} \, {\rm rad}$. \renewcommand{\theequation}{A\arabic{equation}} \setcounter{equation}{0} \section*{APPENDIX A} The purpose of this appendix is to work out {\it to first order in the helicity-gravity coupling} the solution of Hamilton's equations for the dispersion relation (\ref{d4}). The Hamiltonian for the ray motion is given in this case by \begin{equation} {\cal H}({\bf r},{\bf k})=k \pm s{\bf \hat{k}}\, . \, {\bf B}({\bf r}). \label{a1} \end{equation} Hamilton's equations are \begin{eqnarray} \frac{d{\bf r}}{dt}&=&{\bf v}_{g}, \label{a2} \\ \frac{d{\bf k}}{dt}&=&-{\bf \nabla}[\pm s{\bf \hat{k}}\, . \, {\bf B}({\bf r})], \label{a3} \end{eqnarray} where \begin{equation} {\bf v}_{g}=\frac{\partial \omega}{\partial {\bf k}}=\frac{1}{\omega}({\bf k} \pm s{\bf B}) \label{a4} \end{equation} is the group velocity of the rays to first order in the coupling to the gravitomagnetic field. The background is stationary; therefore, for any ray that is a solution of equations (\ref{a2}) and (\ref{a3}), $\omega={\cal H}({\bf r},{\bf k})$ is a constant of the motion. Let $({\bf r}_{+},{\bf k}_{+})$ denote the solution of the equations of motion for a positive-helicity ray and $({\bf r}_{-},{\bf k}_{-})$ denote the corresponding solution for a negative-helicity ray. Working to first order in the helicity-gravity coupling, we let \begin{eqnarray} {\bf k}_{+}&=&{\bf k}_{0}+{\bf \kappa}(t), \,\,\,\,\ {\bf k}_{-}={\bf k}_{0}-{\bf \kappa}(t), \label{a5} \\ {\bf r}_{+}&=&{\bf r}_{0}+{\bf q}(t), \,\,\,\,\ {\bf r}_{-}={\bf r}_{0}-{\bf q}(t), \label{a6} \end{eqnarray} where ${\bf k}_{0}$ is constant, $\omega=|{\bf k}_{0}|$ and \begin{equation} \frac{d{\bf r}_{0}}{dt}={\bf \hat{k}}_{0} \label{a7} \end{equation} is the equation of motion of the unperturbed ray in the absence of spin-gravity coupling. Substituting Eqs. (\ref{a5})-(\ref{a7}) in the equations of motion (\ref{a2}) - (\ref{a3}), we find \begin{eqnarray} \frac{d{\bf q}}{dt}&=&\frac{1}{\omega}({\bf \kappa}+s{\bf B}), \label{a8} \\ \frac{d{\bf \kappa}}{dt}&=&-s{\bf \nabla}({\bf \hat{k}}_{0}\, . \, {\bf B}). \label{a9} \end{eqnarray} Using the fact that ${\bf B}=-{\bf \nabla}Q$, we can write ${\bf \nabla}({\bf \hat{k}}_{0}\, . \, {\bf B})=({\bf \hat{k}}_{0}\, . \, {\bf \nabla}){\bf B}$, which can be expressed as $d{\bf B}/dt$ via Eq. (\ref{a7}) in our approximation scheme. Hence it follows from Eq. (\ref{a9}) that ${\bf \kappa}+s{\bf B}$ is a constant of the motion. This implies that the right-hand side of Eq. (\ref{a8}) is a constant as well; therefore, \begin{equation} {\bf \kappa} + s{\bf B}=\omega{\bf V}_{0}, \label{a10} \end{equation} where ${\bf V}_{0}=d{\bf q}/dt$ is a constant and ${\bf B}$ is evaluated along the average (unpolarized) ray. The requirement that $\omega$ be the same for both rays can be implemented using Eqs. (\ref{a1}) and (\ref{a5}) and the result is \begin{equation} {\bf \hat{k}}_{0}\,\ . \, ({\bf \kappa}+s{\bf B})=0, \label{a11} \end{equation} so that ${\bf v}_{g}\, . \, {\bf \hat{k}}_{0}=1$. This completes the solution of the equations of motion. It follows from ${\bf q}(t)={\bf q}(0)+{\bf V}_{0}t$ and Eq. (\ref{a6}) that the positive and negative helicity rays diverge away from the path of the average (unpolarized) radiation, since ${\bf \hat{k}}_{0}\, . \, {\bf V}_{0}=0$. The integration constant ${\bf V}_{0}$ must be determined from the boundary conditions. Consider, for instance, the emission of rays of radiation that originate at $z=z_{0}$ on the axis of rotation of an astronomical mass. Let the initial direction of propagation of the radiation be normal to the $z$ axis. Then, it follows from these initial data that ${\bf V}_{0}=2sG{\bf J}/(\omega z^{3}_{0})$. Thus there will be a differential deflection of the radiation such that the positive and negative helicity rays separate at a constant rate about the average direction of propagation ${\bf \hat{k}}_{0}$. The total gravitomagnetic splitting angle between the rays would be $2V_{0}=4sGJ/(\omega z^{3}_{0})$, which amounts to a few degrees for gravitational waves of frequency $10^{3}\, {\rm Hz}$ grazing the north pole of a millisecond pulsar. We emphasize that the polarization-dependent differential deflection of rays will occur, in our treatment, only for radiation that originates near a rotating source. The gravitomagnetic field ${\bf B}$ rapidly falls off to zero far from the source; therefore, there will be no differential deflection of rays in scattering situations. This conclusion agrees in the electromagnetic case $(s=1)$ with the results of recent investigations \cite{Gua}. One can use the results of this appendix to estimate the polarization-dependent time delay in the arrival of rays originating near a rotating astronomical source \cite{Mas}. Moreover, one can generalize the treatment of the rotation of linear polarization given in Section IV. In fact, it follows from Eqs. (\ref{a5}), (\ref{a10}) and (\ref{a11}) that \begin{equation} \frac{1}{2}({\bf k}_{-}-{\bf k}_{+})\, . \, {\bf \hat{k}}_{0}=s{\bf \hat{k}}_{0}\, . \, {\bf B}, \label{a12} \end{equation} where ${\bf \hat{k}}_{0}\, . \, {\bf B}=-dQ/dt$ along the unperturbed ray. Therefore, we find, following a method analogous to that used in Section IV, that the angle of rotation of linear polarization is \begin{equation} \Delta = s[Q({\bf r}_{i})-Q({\bf r}_{f})] \label{a13} \end{equation} for a linearly polarized ray of radiation traveling from ${\bf r}_{i}$ to ${\bf r}_{f}$. Thus for rays that are incident from infinity and travel to infinity, the net angle of rotation of linear polarization vanishes. Finally, we should mention that one can think of the effects discussed here in terms of the helicity dependence of the index of refraction of an effective medium for the propagation of rays. From the definition \begin{equation} |{\bf k}_{\pm}|=\omega n_{\pm}, \label{a14} \end{equation} we find that \begin{equation} n_{\pm}=n_{0}\mp \frac{s}{\omega}{\bf \hat{k}}\, . \, {\bf B}, \label{a15} \end{equation} where $n_{0}=1$, since we have neglected here the polarization-independent bending of rays given by $n_{0} \approx 1+2f$.
Title: Avoiding Dark Energy with 1/R Modifications of Gravity
Abstract: Scalar quintessence seems epicyclic because one can choose the potential to reproduce any cosmology (I review the construction) and because the properties of this scalar seem to raise more questions than they answer. This is why there has been so much recent interest in modified gravity. I review the powerful theorem of Ostrogradski which demonstrates that the only potentially stable, local modification of general relativity is to make the Lagrangian an arbitrary function of the Ricci scalar. Such a theory can certainly reproduce the current phase of cosmic acceleration without Dark Energy. However, this explanation again seems epicyclic in that one can construct a function of the Ricci scalar to support any cosmology (I give the technique). Models of this form are also liable to problems in the way they couple to matter, both in terms of matter's impact upon them and in terms of the long range gravitational force they predict. Because of these problems my own preference for avoiding Dark Energy is to bypass Ostrogradski's theorem by considering the fully nonlocal effective action built up by quantum gravitational processes during the epoch of primordial inflation.
https://export.arxiv.org/pdf/astro-ph/0601672
\title*{Avoiding Dark Energy with 1/R Modifications of Gravity} \titlerunning{1/R Modifications of Gravity} \author{R. P. Woodard} \institute{Department of Physics, University of Florida, Gainesville, FL 32611-8440, USA \texttt{woodard@phys.ufl.edu}} \section{Introduction} \label{sec:1} The case for alternate gravity is easily made. The best that can be done from observing cosmic motions is to infer the metric $g_{\mu\nu}$ in some coordinate system. From this one can reconstruct the Einstein tensor and then ask whether or not general relativity predicts it in terms of the observed sources of stress-energy, \begin{equation} \Bigl(R_{\mu\nu} - \frac12 g_{\mu\nu} R\Bigr)_{\rm rec} = 8 \pi G \Bigl( T_{\mu\nu}\Bigr)_{\rm obs} \; ? \end{equation} One way of explaining any disagreement is by positing the existence of an unobserved, ``dark'' component of the stress-energy tensor, \begin{equation} \Bigl(T_{\mu\nu}\Bigr)_{\rm dark} \equiv \frac1{8\pi G} \Bigl(R_{\mu\nu} - \frac12 g_{\mu\nu} R\Bigr)_{\rm rec} - \Bigl(T_{\mu\nu}\Bigr)_{\rm obs} \; . \end{equation} This always works, but recent observations make it seem epicyclic. The theory of nucleosynthesis implies that no more than about 4\% of the energy density currently required to make general relativity agree with all observations can consist of any material with which we are presently familiar \cite{BBN} --- and only a fraction of this 4\% is observed. Just to make general relativity agree with the observed motions of galaxies and galactic clusters we must posit that {\it six times} the mass of ordinary matter comes in the form of nonbaryonic, cold dark matter \cite{CDM}. Although there are some plausible candidates for what this might be, no Earth-bound laboratory has yet succeeded in detecting it. I belong to the minority of physicists who feel that this factor of six already strains credulity. Easing that strain is what led Milgrom to propose MOND \cite{MM}, which can be viewed as a phenomenological modification of gravity in the regime of very small accelerations. There is an impressive amount of observational data in favor of this modification \cite{SM} --- although see \cite{GSKVK}. Bekenstein has recently constructed a fully relativistic field theory \cite{JDB} which reproduces MOND, and a preliminary analysis of the resulting cosmology works better than many experts thought possible \cite{SMFB}. However, the worst problem for conventional gravity comes on the largest scales. To make general relativity agree with the Hubble plots of distant Type Ia supernovae \cite{SNCP,SNST,SNLS}, with the power spectrum of anisotropies in the cosmic microwave background \cite{CMB} and with large scale structure surveys \cite{LSS}, one must accept an additional component of ``dark energy'' that is about {\it eighteen times} larger than that of ordinary matter. This would mean that 96\% of the current universe's energy exists in forms which have so far only been detected gravitationally! Even people who believe passionately in dark matter (and hence accept the factor of six) find this factor of $6 \!+\! 18 \!=\! 24$ difficult to swallow. That is why there has been so much recent interest in modifying gravity to make it predict observed cosmic phenomena without the need for dark energy, and sometimes even without the need for dark matter. I want to stress that the issue is one of plausibility. There is no problem inventing field theories which give the required amount of dark energy. The simplest way of doing it is with a minimally coupled scalar \cite{CW,PR}, \begin{equation} \mathcal{L} = -\frac12 \partial_{\mu} \varphi \partial_{\nu} g^{\mu\nu} \sqrt{-g} - V(\varphi) \sqrt{-g} \; . \label{quint} \end{equation} The usual procedure is to begin with a scalar potential $V(\varphi)$ and work out the cosmology, but it is easy to start with whatever cosmological evolution is desired and {\it construct} the potential which would support it. I will go through the construction here, both to make the point and so that it can be used later. On the largest scales the geometry of the universe can be described in terms of a single function of time known as the scale factor $a(t)$, \begin{equation} ds^2 = -dt^2 + a^2(t) d\vec{x} \cdot d\vec{x} \; . \end{equation} The logarithmic time derivative of this quantity gives the Hubble parameter, \begin{equation} H(t) \equiv \frac{\dot{a}}{a} \; . \end{equation} If we specialize to a solution $\varphi_0(t)$ of the scalar field equations which depends only upon time, the two nontrivial Einstein equations are, \begin{eqnarray} 3 H^2 & = & 8 \pi G \Bigl(\frac12 \dot{\varphi}_0^2 + V(\varphi_0)\Bigr) \; , \label{E1} \\ -2 \dot{H} - 3 H^2 & = & 8 \pi G \Bigl(\frac12 \dot{\varphi}_0^2 - V(\varphi_0)\Bigr) \; . \label{E2} \end{eqnarray} Let us assume $a(t)$ is known as an explicit function of time, and construct $\varphi_0(t)$ and $V(\varphi)$. By adding (\ref{E1}) and (\ref{E2}) we obtain, \begin{equation} -2 \dot{H} = 8 \pi G \dot{\varphi}_0^2 \; . \label{twoeqns} \end{equation} The weak energy condition implies $\dot{H}(t) \leq 0$ so we can take the square root and integrate to solve for $\varphi_0(t)$, \begin{equation} \varphi_0(t) = \varphi_I \pm \int_{t_I}^t dt' \sqrt{\frac{-2 \dot{H}(t')}{ 8 \pi G}} \; . \label{phi} \end{equation} One can choose $\varphi_I$ and the sign freely. Because the integrand in (\ref{phi}) is always positive, the function $\varphi_0(t)$ is monotonic. This means we can invert to solve for time as a function of $\varphi_0$. Let us call the inverse function $T(\varphi)$, \begin{equation} \psi = \varphi_0\Bigl(T(\psi)\Bigr) \; . \label{inv} \end{equation} By subtracting (\ref{E2}) from (\ref{E1}) we obtain a relation for the scalar potential as a function of time, \begin{equation} V = \frac1{8\pi G} \Bigl( \dot{H}(t) + 3 H^2(t)\Bigr) \; . \end{equation} The potential is determined as a function of the scalar by substituting the inverse function (\ref{inv}), \begin{equation} V(\varphi) = \frac1{8\pi G} \Biggl\{ \dot{H}\Bigl(T(\varphi)\Bigr) + 3 H^2\Bigl(T(\varphi)\Bigr) \Biggr\} \; . \end{equation} This construction gives a scalar which supports any evolution $a(t)$ (with $\dot{H}(t) < 0$) all by itself. Should you wish to include some other, known component of the stress-energy, simply add the energy density and pressure of this component to the Einstein equations, \begin{eqnarray} 3 H^2 & = & 8 \pi G \Bigl(\frac12 \dot{\varphi}_0^2 + V(\varphi_0) + \rho_{\rm known}\Bigr) \; , \\ -2 \dot{H} - 3 H^2 & = & 8 \pi G \Bigl(\frac12 \dot{\varphi}_0^2 - V(\varphi_0) + p_{\rm known}\Bigr) \; . \end{eqnarray} Provided $\rho_{\rm known}$ and $p_{\rm known}$ are known functions of either time or the scale factor, the construction goes through as before.\footnote{ This construction seems to be due to Ratra and Peebles \cite{PR}. Recent examples of its use include \cite{TW2,SRSS,NO0}.} Using this method one can devise a new field $\varphi(x)$ which will support {\it any} cosmology with $\dot{H}(t) < 0$. However, the introduction of such a ``quintessence'' field raises a number of questions: \begin{enumerate} \item{Where does $\varphi$ reside in fundamental theory?} \item{Why can't $\varphi$ couple to fields other than the metric? And if it does couple to other fields, why haven't we detected its influence in Earth-bound laboratories?} \item{Why did $\varphi$ come to dominate the stress-energy of the universe so recently in cosmological time?} \item{Why is the $\varphi$ field so homogeneous?} \end{enumerate} When a phenomenological fix raises more questions than it answers people are naturally drawn to investigate other fixes. One possibility is that general relativity is not the correct theory of gravity on cosmological scales. In this talk I shall review gravitational Lagrangians of the form, \begin{equation} \mathcal{L} = \frac1{16 \pi G} \Bigl(R + \Delta R[g]\Bigr) \sqrt{-g} \; , \label{ansatz} \end{equation} where $\Delta R[g]$ is some local scalar constructed from the curvature tensor and possibly its covariant derivatives. Examples of such scalars are, \begin{equation} \frac1{\mu^2} R^{\alpha\beta} R_{\alpha\beta} \qquad , \qquad \frac1{\mu^4} g^{\mu\nu} R_{,\mu} R_{,\nu} \qquad , \qquad \mu^2 \sin\Bigl(\frac1{\mu^4} R^{\alpha\beta\rho\sigma} R_{\alpha\beta\rho\sigma}\Bigr) \; . \end{equation} I begin by reviewing a powerful no-go theorem which pervades and constrains fundamental theory so completely that most people assume its consequence without thinking. This is the theorem of Ostrogradski \cite{MO}, who essentially showed why Newton was right to suppose that the laws of physics involve no more than two time derivatives of the fundamental dynamical variables. The key consequence for our purposes is that the only viable form for the functional $\Delta R[g]$ in (\ref{ansatz}) is an algebraic function of the undifferentiated Ricci scalar, \begin{equation} \Delta R[g] = f(R) \; . \end{equation} I review the Ostrogradski result in section 2, and hopefully immunize you against some common misconceptions about it in section 3. In section 4 I explain why $f(R)$ theories do not contradict Ostrogradski's result. I also demonstrate that, in the absence of matter, $f(R)$ theories are equivalent to ordinary gravity, with $f(R) = 0$, plus a minimally coupled scalar of the form (\ref{quint}). Then I use the construction given above to show how one can choose $f(R)$ to enforce an arbitrary cosmology. This establishes that an $f(R)$ can be found to support any desired cosmology. In section 5 I discuss problems associated with the particular choice function $f(R) = -\frac{\mu^4}{R}$. Section 6 presents conclusions. \section{The Theorem of Ostrogradski} \label{sec:2} Ostrogradski's result is that there is a linear instability in the Hamiltonians associated with Lagrangians which depend upon more than one time derivative in such a way that the dependence cannot be eliminated by partial integration \cite{MO}. The result is so general that I can simplify the discussion by presenting it in the context of a single, one dimensional point particle whose position as a function of time is $q(t)$. First I will review the way the Hamiltonian is constructed for the usual case in which the Lagrangian involves no higher than first time derivatives. Then I present Ostrogradski's construction for the case in which the Lagrangian involves second time derivatives. And the section closes with the generalization to $N$ time derivatives. In the usual case of $L = L(q,\dot{q})$, the Euler-Lagrange equation is, \begin{equation} \frac{\partial L}{\partial q} - \frac{d}{dt} \frac{\partial L}{\partial \dot{q}} = 0 \; . \label{ELE1} \end{equation} The assumption that $\frac{\partial L}{\partial \dot{q}}$ depends upon $\dot{q}$ is known as {\it nondegeneracy}. If the Lagrangian is nondegenerate we can write (\ref{ELE1}) in the form Newton assumed so long ago for the laws of physics, \begin{equation} \ddot{q} = \mathcal{F}(q,\dot{q}) \qquad \Longrightarrow \qquad q(t) = \mathcal{Q}(t,q_0,\dot{q}_0) \; . \label{newt} \end{equation} From this form it is apparent that solutions depend upon two pieces of initial value data: $q_0 = q(0)$ and $\dot{q}_0 = \dot{q}(0)$. The fact that solutions require two pieces of initial value data means that there must be two canonical coordinates, $Q$ and $P$. They are traditionally taken to be, \begin{equation} Q \equiv q \qquad {\rm and} \qquad P \equiv \frac{\partial L}{\partial \dot{q}} \; . \label{ctrans} \end{equation} The assumption of nondegeneracy is that we can invert the phase space transformation (\ref{ctrans}) to solve for $\dot{q}$ in terms of $Q$ and $P$. That is, there exists a function $v(Q,P)$ such that, \begin{equation} \frac{\partial L}{\partial \dot{q}} \Biggl\vert_{q = Q \atop \dot{q} = v} = P \; . \label{invct} \end{equation} The canonical Hamiltonian is obtained by Legendre transforming on $\dot{q}$, \begin{eqnarray} H(Q,P) & \equiv & P \dot{q} - L \; , \\ & = & P v(Q,P) - L\Bigl(Q,v(Q,P)\Bigr) \; . \end{eqnarray} It is easy to check that the canonical evolution equations reproduce the inverse phase space transformation (\ref{invct}) and the Euler-Lagrange equation (\ref{ELE1}), \begin{eqnarray} \dot{Q} & \equiv & \frac{\partial H}{\partial P} = v + P \frac{\partial v}{ \partial P} - \frac{\partial L}{\partial \dot{q}} \frac{\partial v}{\partial P} = v \; , \\ \dot{P} & \equiv & -\frac{\partial H}{\partial Q} = -P \frac{\partial v}{ \partial Q} + \frac{\partial L}{\partial q} + \frac{\partial L}{\partial \dot{q}} \frac{\partial v}{\partial P} = \frac{\partial L}{\partial q} \; . \end{eqnarray} This is what we mean by the statement, ``the Hamiltonian generates time evolution.'' When the Lagrangian has no explicit time dependence, $H$ is also the associated conserved quantity. Hence it is ``the'' energy by anyone's definition, of course up to canonical transformation. Now consider a system whose Lagrangian $L(q,\dot{q},\ddot{q})$ depends nonde\-gen\-er\-ate\-ly upon $\ddot{q}$. The Euler-Lagrange equation is, \begin{equation} \frac{\partial L}{\partial q} - \frac{d}{dt} \frac{\partial L}{\partial \dot{q}} + \frac{d^2}{dt^2} \frac{\partial L}{\partial \ddot{q}} = 0 \; . \label{ELE2} \end{equation} Non-degeneracy implies that $\frac{\partial L}{\partial \ddot{q}}$ depends upon $\ddot{q}$, in which case we can cast (\ref{ELE2}) in a form radically different from Newton's, \begin{equation} q^{(4)} = \mathcal{F}(q,\dot{q},\ddot{q},q^{(3)}) \qquad \Longrightarrow \qquad q(t) = \mathcal{Q}(t,q_0,\dot{q}_0,\ddot{q}_0,q^{(3)}_0) \; . \end{equation} Because solutions now depend upon four pieces of initial value data there must be four canonical coordinates. Ostrogradski's choices for these are, \begin{eqnarray} Q_1 \equiv q \qquad & , & \qquad P_1 \equiv \frac{\partial L}{\partial \dot{q}} - \frac{d}{dt} \frac{\partial L}{\partial \ddot{q}} \; , \label{ct1} \\ Q_2 \equiv \dot{q} \qquad & , & \qquad P_2 \equiv \frac{\partial L}{\partial \ddot{q}} \; . \label{ct2} \end{eqnarray} The assumption of nondegeneracy is that we can invert the phase space transformation (\ref{ct1}-\ref{ct2}) to solve for $\ddot{q}$ in terms of $Q_1$, $Q_2$ and $P_2$. That is, there exists a function $a(Q_1,Q_2,P_2)$ such that, \begin{equation} \frac{\partial L}{\partial \ddot{q}} \Biggl\vert_{{q = Q_1 \atop \dot{q} = Q_2} \atop \ddot{q} = a} = P_2 \; . \label{invct2} \end{equation} Note that one only needs the function $a(Q_1,Q_2,P_2)$ to depend upon {\it three} canonical coordinates --- and not all four --- because $L(q,\dot{q},\ddot{q})$ only depends upon three configuration space coordinates. This simple fact has great consequence. Ostrogradski's Hamiltonian is obtained by Legendre transforming, just as in the first derivative case, but now on $\dot{q} = q^{(1)}$ and $\ddot{q} = q^{(2)}$, \begin{eqnarray} \lefteqn{H(Q_1,Q_2,P_1,P_2) \equiv \sum_{i=1}^2 P_i q^{(i)} - L \; , } \\ & & = P_1 Q_2 + P_2 a(Q_1,Q_2,P_2) - L\Bigl(Q_1,Q_2,a(Q_1,Q_2,P_2)\Bigr) \; . \label{Host} \end{eqnarray} The time evolution equations are just those suggested by the notation, \begin{equation} \dot{Q_i} \equiv \frac{\partial H}{\partial P_i} \qquad {\rm and} \qquad \dot{P}_i \equiv - \frac{\partial H}{\partial Q_i} \; . \end{equation} Let's check that they generate time evolution. The evolution equation for $Q_1$, \begin{equation} \dot{Q}_1 = \frac{\partial H}{\partial P_1} = Q_2 \; , \end{equation} reproduces the phase space transformation $\dot{q} = Q_2$ in (\ref{ct2}). The evolution equation for $Q_2$, \begin{equation} \dot{Q}_2 = \frac{\partial H}{\partial P_2} = a + P_2 \frac{\partial a}{ \partial P_2} - \frac{\partial L}{\partial \ddot{q}} \frac{\partial a}{\partial P_2} = a \; , \end{equation} reproduces (\ref{invct2}). The evolution equation for $P_2$, \begin{equation} \dot{P}_2 = -\frac{\partial H}{\partial Q_2} = -P_1 - P_2 \frac{\partial a}{ \partial Q_2} + \frac{\partial L}{\partial \dot{q}} + \frac{\partial L}{ \partial \ddot{q}} \frac{\partial a}{\partial Q_2} = -P_1 + \frac{\partial L}{ \partial \dot{q}} \; , \end{equation} reproduces the phase space transformation $P_1 = \frac{\partial L}{\partial \dot{q}} - \frac{d}{dt} \frac{\partial L}{\partial \ddot{q}}$ (\ref{ct1}). And the evolution equation for $P_1$, \begin{equation} \dot{P}_1 = -\frac{\partial H}{\partial Q_1} = -P_2 \frac{\partial a}{\partial Q_1} + \frac{\partial L}{\partial q} + \frac{\partial L}{\partial \ddot{q}} \frac{\partial a}{\partial Q_1} = \frac{\partial L}{\partial q} \; , \end{equation} reproduces the Euler-Lagrange equation (\ref{ELE2}). So Ostrogradski's system really does generate time evolution. When the Lagrangian contains no explicit dependence upon time it is also the conserved Noether current. By anyone's definition, it is therefore ``the'' energy, again up to canonical transformation. There is one, overwhelmingly bad thing about Ostrogradski's Hamiltonian (\ref{Host}): it is {\it linear} in the canonical momentum $P_1$. This means that no system of this form can be stable. In fact, there is not even any barrier to decay. Note also the power and generality of the result. It applies to {\it every} Lagrangian $L(q,\dot{q},\ddot{q})$ which depends nondegenerately upon $\ddot{q}$, independent of the details. The only assumption is nondegeneracy, and that simply means one cannot eliminate $\ddot{q}$ by partial integration. This is why Newton was right to assume the laws of physics take the form (\ref{newt}) when expressed in terms of fundamental dynamical variables. Adding more higher derivatives just makes the situation worse. Consider a Lagrangian $L\left(q,\dot{q},\dots,q^{(N)}\right)$ which depends upon the first $N$ derivatives of $q(t)$. If this Lagrangian depends nondegenerately upon $q^{(N)}$ then the Euler-Lagrange equation, \begin{equation} \sum_{i=0}^N \left(-{d \over dt}\right)^i {\partial L \over \partial q^{(i)}} = 0 \; , \label{ELEN} \end{equation} contains $q^{(2N)}$. Hence the canonical phase space must have $2N$ coordinates. Ostrogradski's choices for them are, \begin{equation} Q_i \equiv q^{(i-1)} \qquad {\rm and} \qquad P_i \equiv \sum_{j=i}^N \Bigl(- \frac{d}{dt}\Bigr)^{j-i} \frac{\partial L}{\partial q^{(j)}} \; . \end{equation} Non-degeneracy means we can solve for $q^{(N)}$ in terms of $P_N$ and the $Q_i$'s. That is, there exists a function $\mathcal{A}(Q_1,\ldots,Q_N,P_N)$ such that, \begin{equation} \frac{\partial L}{\partial q^{(N)}} \Biggl\vert_{q^{(i-1)} = Q_i \atop q^{(N)} = \mathcal{A}} = P_N \; . \label{nondeg} \end{equation} For general $N$ Ostrogradski's Hamiltonian takes the form, \begin{eqnarray} H & \equiv & \sum_{i=1}^N P_i q^{(i)} - L \; , \\ & = & P_1 Q_2 + P_2 Q_3 + \cdots + P_{N-1} Q_N + P_N \mathcal{A} - L\Bigl(Q_1,\ldots,Q_N,\mathcal{A}\Bigr) \; . \label{HN} \end{eqnarray} It is simple to check that the evolution equations, \begin{equation} \dot{Q}_i \equiv \frac{\partial H}{\partial P_i} \qquad {\rm and} \qquad \dot{P}_i \equiv -\frac{\partial H}{\partial Q_i} \; , \end{equation} again reproduce the canonical transformations and the Euler-Lagrange equation. So (\ref{HN}) generates time evolution. Similarly, it is Noether current for the case where the Lagrangian contains no explicit time dependence. So there is little alternative to regarding (\ref{HN}) as ``the'' energy, again up to canonical transformation. One can see from (\ref{HN}) that the Hamiltonian is linear in $P_1, P_2, \ldots P_{N-1}$. Only with respect to $P_N$ might it be bounded from below. Hence the Hamiltonian is necessarily unstable over half the classical phase space for large $N$! \section{Common Misconceptions} \label{sec:3} The no-go theorem I have just reviewed ought to come as no surprise. It explains why Newton was right to expect that physical laws take the form of second order differential equations when expressed in terms of fundamental dynamical variables.\footnote{The caveat is there because one can always get higher order equations by solving for some of the fundamental variables.} Every fundamental system we have discovered since Newton's day has had this form. The bizarre, dubious thing would be if Newton had blundered upon a tiny subset of possible physical laws, and all our probing over the course of the next three centuries had never revealed the vastly richer possibilities. However --- {\it deep sigh} --- particle theorists don't like being told something is impossible, and a definitive no-go theorem such as that of Ostrogradski provokes them to tortuous flights of evasion. I ought to know, I get called upon to referee the resulting papers often enough! No one has so far found a way around Ostrogradski's theorem. I won't attempt to prove that no one ever will, but let me use this section to run through some of the misconceptions which have been in back of attempted evasions. To fix ideas it will be convenient to consider a higher derivative generalization of the harmonic oscillator, \begin{equation} \mathcal{L} = -\frac{g m}{2 \omega^2} \ddot{q}^2 + \frac{m}2 \dot{q}^2 - \frac{m\omega^2}2 q^2 \; . \label{HDO} \end{equation} Here $m$ is the particle mass, $\omega$ is a frequency and $g$ is a small positive pure number we can think of as a coupling constant. The Euler-Lagrange equation, \begin{equation} -m \Bigl( \frac{g}{\omega^2} q^{(4)} + \ddot{q} + \omega^2 q\Bigr) = 0 \; , \label{HDE} \end{equation} has the general solution, \begin{equation} q(t) = A_+ \cos(k_+ t) + B_+ \sin(k_+ t) + A_- \cos(k_- t) + B_- \sin(k_- t) \; . \label{gensol} \end{equation} Here the two frequencies are, \begin{equation} k_{\pm} \equiv \omega \sqrt{ \frac{1 \mp \sqrt{1 \!-\! 4 g}}{2 g} } \; , \end{equation} and the initial value constants are, \begin{eqnarray} A_+ = \frac{k_-^2 q_0 \!+\! \ddot{q}_0}{k_-^2 \!-\! k_+^2} \qquad & , & \qquad B_+ = \frac{k_-^2 \dot{q}_0 \!+\! q^{(3)}_0}{k_+ (k_-^2 \!-\! k_+^2)} \; , \\ A_- = \frac{k_+^2 q_0 \!+\! \ddot{q}_0}{k_+^2 \!-\! k_-^2} \qquad & , & \qquad B_- = \frac{k_+^2 \dot{q}_0 \!+\! q^{(3)}_0}{k_- (k_+^2 \!-\! k_-^2)} \; . \end{eqnarray} The conjugate momenta are, \begin{eqnarray} P_1 = m \dot{q} + \frac{g m}{\omega^2} q^{(3)} \qquad & \Leftrightarrow & \qquad q^{(3)} = \frac{\omega^2 P_1 \!-\! m \omega^2 Q_2}{g m} \; , \\ P_2 = - \frac{g m}{\omega^2} \ddot{q} \qquad & \Leftrightarrow & \qquad \ddot{q} = -\frac{\omega^2 P_2}{g m} \; . \end{eqnarray} The Hamiltonian can be expressed in terms of canonical variables, configuration space variables or initial value constants, \begin{eqnarray} H & = & P_1 Q_2 - \frac{\omega^2}{2 g m} P_2^2 - \frac{m}2 Q_2^2 + \frac{m \omega^2}2 Q_1^2 \; , \label{H1} \\ & = & \frac{g m}{\omega^2} \dot{q} q^{(3)} - \frac{g m}{2 \omega^2} \ddot{q}^2 + \frac{m}2 \dot{q}^2 + \frac{m \omega^2}2 q^2 \; , \label{H2} \\ & = & \frac{m}2 \sqrt{1 \!-\! 4 g} \, k_+^2 (A_+^2 \!+\! B_+^2) - \frac{m}2 \sqrt{1 \!-\! 4 g} \, k_-^2 (A_-^2 \!+\! B_-^2) \; . \label{H3} \end{eqnarray} The last form makes it clear that the ``$+$'' modes carry positive energy whereas the ``$-$'' modes carry negative energy. \subsection{Nature of the Instability} \label{sub:3.1} It's important to understand both how the Ostrogradskian instability manifests and what is physically wrong with a theory which shows this instability. Because the Ostrogradskian Hamiltonian (\ref{HN}) is not bounded below with respect to more than one of its conjugate momenta, one sees that the problem is not reaching arbitrarily negative energies by setting the dynamical variable to some {\it constant value}. Rather it is reaching arbitrarily negative energies by making the dynamical variable have a certain {\it time dependence}. People sometimes mistakenly believe they have found a higher derivative system which is stable when all they have checked is that the Hamiltonian is bounded from below for constant field configurations. For example, from expression (\ref{H2}) we see that our higher derivative oscillator energy is bounded below by zero for $q(t) = {\rm const}$! Negative energies are achieved by making $\ddot{q}$ large and/or making $q^{(3)}$ large while keeping $\dot{q} \!+\! g q^{(3)}/\omega^2$ fixed. Another crucial point is that the same dynamical variable typically carries both positive and negative energy degrees of freedom in a higher derivative theory. For our higher derivative oscillator this is apparent from expression (\ref{gensol}) which shows that $q(t)$ involves both the positive energy degrees of freedom, $A_+$ and $B_+$, and the negative energy ones, $A_-$ and $B_-$. And note from expression (\ref{H3}) that I really mean positive and negative {\rm energy}, not just positive and negative frequency, which is the usual case in a lower derivative theory. People sometimes imagine that the energy of a higher derivative theory decays with time. That is not true. Provided one is dealing with a complete system, and provided there is no external time dependence, the energy of a higher derivative system is conserved, just as it would be under those conditions for a lower derivative theory. This conservation is apparent for our higher derivative oscillator from expression (\ref{H3}). The physical problem with nondegenerate higher derivative theories is not that their energies decay to lower and lower values. The problem is rather that certain sectors of the theory become arbitrarily highly excited when one is dealing with an interacting, continuum field theory which has nondegenerate higher derivatives. To understand this I must digress to remind you of some familiar facts about the Hydrogen atom. If you consider Hydrogen in isolation, there is an infinite tower of stationary states. However, if you allow the Hydrogen atom to interact with electromagnetism only the ground state is stationary; all the excited states decay through the emission of a photon. Why is this so? It certainly is {\it not} because ``the system wants to lower its energy.'' The energy of the full system is constant, the binding energy released by the decaying atom being compensated by the energy of the recoil photon. Yet the decay always takes place, and rather quickly. The reason is that decay is terrifically favored by entropy. If we prepare the Hydrogen atom in an excited state, with no photons present, there is {\it one} way for the atom to remain excited, whereas there are an {\it infinite} number of ways for it to decay because the recoil photon could go off in any direction. Now consider an interacting, continuum field theory which possesses the Ostrogradskian instability. In particular consider its likely particle spectrum about some ``empty'' solution in which the field is constant. Because the Hamiltonian is linear in all but one of the conjugate momenta we can increase or decrease the energy by moving different directions in phase space. Hence there must be both positive energy and negative energy particles --- just as there are in our higher derivative oscillator. Just as in that point particle model, the same continuum field must carry the creation and annihilation operators of {\it both} the positive and the negative energy particles. If the theory is interacting at all --- that is, if its Lagrangian contains a higher than quadratic power of the field --- then there will be interactions between positive and negative energy particles. Depending upon the interaction, the empty state can decay into some collection of positive and negative energy particles. The details don't really matter, all that matters is the counting: there is {\it one} way for the system to stay empty versus a continuous {\it infinity} of ways for it to decay. This infinity is even worse than for the Hydrogen atom because it includes not only all the directions that recoil particles of fixed energies could go but also the fact that the various energies can be arbitrarily large in magnitude provided they sum to zero. Because of that last freedom the decay is instantaneous. And the system doesn't just decay once! It is even {\it more} entropicly favored for there to be two decays, and better yet for three, etc. You can see that such a system instantly evaporates into a maelstrom of positive and negative energy particles. Some of my mathematically minded colleagues would say it isn't even defined. I prefer to simply observe that no theory of this kind can describe the universe we experience in which all particles have positive energy and empty space remains empty. Note that we only reach this conclusion if the higher derivative theory possesses both interactions and continuum particles. Our point particle oscillator has no interactions, so its negative energy degree of freedom is harmless. Of course it is also completely unobservable! However, it is conceivable we could couple this higher derivative oscillator to a discrete system without engendering an instability. The feature that drives the instability when continuum particles are present is the vast entropy of phase space. Without that it becomes an open question whether or not there is anything wrong with a higher derivative theory. Of course we live in a continuum universe, and any degree of freedom we can observe must be interacting, so these are very safe assumptions. However, people sometimes delude themselves that there is no problem with continuum, interacting higher derivative models of the universe on the basis of studying higher derivative systems which could never describe the universe because they either lack interactions or else continuum particles. In this sub-section we have learned: \begin{enumerate} \item{The Ostrogradskian instability does not drive the dynamical variable to a special, constant value but rather to a special kind of time dependence.} \item{A dynamical variable which experiences the Ostrogradskian instability will carry both positive and negative energy creation and annihilation operators.} \item{If the system interacts then the ``empty'' state can decay into a collection of positive and negative energy excitations.} \item{If the system is a continuum field theory the vast entropy at infinite momentum will make the decay instantaneous.} \end{enumerate} \subsection{Perturbation Theory} \label{sub:3.2} People sometimes mistakenly believe that the Ostrogradskian instability is avoided if higher derivatives are segregated to appear only in interaction terms. This is not correct if one considers the theory on a fundamental level. One can see from the construction of section {\ref{sec:2} that the fact of Ostrogradski's Hamiltonian being unbounded below depends only upon nondegeneracy, irrespective of how one organizes any approximation technique. However, there is a way of imposing constraints to make the theory agree with its perturbative development. If this is done then there are no more higher derivative degrees of freedom, however, one typically loses unitarity, causality and Lorentz invariance on the nonperturbative level. I constructed the higher derivative oscillator (\ref{HDO}) so that its higher derivatives vanish when $g \!=\! 0$. If we solve the Euler-Lagrange equation (\ref{HDE}) exactly, without employing perturbation theory, there are four linearly independent solutions (\ref{gensol}) corresponding to a positive energy oscillator of frequency $k_+$ and a negative energy oscillator of frequency $k_-$. However, we might instead regard the parameter $g$ as a coupling constant and solve the equations perturbatively. This means substituting the ansatz, \begin{equation} q_{\rm pert}(t) = \sum_{n=0}^{\infty} g^n x_n(t) \; , \label{pertan} \end{equation} into the Euler-Lagrange equation (\ref{HDE}) and segregating terms according to powers of $g$. The resulting system of equations is, \begin{eqnarray} \ddot{x}_0 + \omega^2 x_0 & = & 0 \; , \label{E0} \\ \ddot{x}_1 + \omega^2 x_1 & = & -\frac1{\omega^2} x^{(4)}_0 \; , \\ \ddot{x}_2 + \omega^2 x_2 & = & -\frac1{\omega^2} x^{(4)}_1 \; , \end{eqnarray} and so on. Because the zeroth order equation involves only second derivatives, its solution depends upon only two pieces of initial value data, \begin{equation} x_0(t) = q_0 \cos(\omega t) + \frac{\dot{q}_0}{\omega} \sin(\omega t) \; . \end{equation} The first correction is, \begin{equation} x_1(t) = -\frac{\omega t}2 q_0 \sin(\omega t) + \frac{t}2 \dot{q}_0 \cos(\omega t) - \frac1{2 \omega} \dot{q_0} \sin(\omega t) \; , \end{equation} and it is easy to see that the sum of all corrections gives, \begin{equation} q_{\rm pert}(t) = q_0 \cos(k_+ t) + \frac{\dot{q}_0}{k_+} \sin(k_+ t) \; . \label{pertsol} \end{equation} What is the relation of the perturbative solution (\ref{pertsol}) to the general one (\ref{gensol})? The perturbative solution is what results if we change the theory by imposing the constraints, \begin{eqnarray} \ddot{q}(t) = - k_+^2 q(t) \qquad & \Longleftrightarrow & \qquad P_2 = \frac{m}2 \Bigl(1 \!-\! \sqrt{1 \!-\! 4g}\Bigr) Q_1 \; , \label{C1} \\ q^{(3)}(t) = - k_+^2 \dot{q}(t) \qquad & \Longleftrightarrow & \qquad P_1 = \frac{m}2 \Bigl(1 \!+\! \sqrt{1 \!-\! 4g}\Bigr) Q_2 \; . \label{C2} \end{eqnarray} Under these constraints the Hamiltonian becomes, \begin{equation} H_{\rm pert} = \sqrt{1 \!-\! 4g} \Bigl( \frac{m}2 Q_2^2 + \frac{m k_+^2}2 Q_1^2 \Bigr) \; , \end{equation} which is indeed that of a single harmonic oscillator. From the full theory, perturbation theory has retained only the solution whose frequency is well behaved for $g \rightarrow 0$, \begin{equation} k_+ = \omega \Bigl(1 + \frac12 g + \frac78 g^2 + O(g^3) \Bigr) \; . \label{lowk} \end{equation} It has discarded the solution whose frequency blows up as $g \rightarrow 0$, \begin{equation} k_- = \frac{\omega}{\sqrt{g}} \Bigl(1 -\frac12 g - \frac58 g^2 + O(g^3)\Bigr) \; . \label{highk} \end{equation} So what's wrong with this? In fact there is nothing wrong with the procedure for our model. If the constraints (\ref{C1}-\ref{C2}) are imposed at one instant, they remain valid for all times as a consequence of the full equation of motion. However, that is only because our model is free of interactions. Recall that this same feature means the positive and negative energy degrees of freedom exist in isolation of one another, and there is no decay to arbitrarily high excitation as there would be for an interacting, continuum field theory. When interactions are present it is more involved but still possible to impose constraints which change the theory so that only the lower derivative, perturbative solutions remain. The procedure was first worked out by Ja\'en, Llosa and Molina \cite{JLM}, and later, independently, by Eliezer and me \cite{EW}. To understand its critical defect suppose we change the ``interaction'' of our higher derivative oscillator from a quadratic term to a cubic one, \begin{equation} -\frac{g m}{2 \omega^4} \, \ddot{q}^2 \longrightarrow -\frac{g m}{6 \ell \omega^4} \, \ddot{q}^3 \; . \end{equation} Here $\ell$ is some constant with the dimensions of a length. As with the quadratic interaction, the new equation of motion is fourth order, \begin{equation} -m \Biggl[ \frac{d^2}{dt^2} \Bigl(\frac{g \ddot{q}^2}{2 \ell \omega^4}\Bigr) + \ddot{q} + \omega^2 q \Biggr] = 0 \; , \end{equation} Its general solution depends upon four pieces of initial value data. However, by isolating the highest derivative term of the free theory, \begin{equation} \ddot{q} =- \omega^2 q -\frac{d^2}{dt^2} \Bigl(\frac{g \ddot{q}^2}{2 \ell \omega^4} \Bigr) \; , \label{rewrite} \end{equation} and then iteratively substituting (\ref{rewrite}), we can delay the appearance of higher derivatives on the right hand side to any desired order in the coupling constant $g$. For example, two iterations frees the right hand side of higher derivatives up to order $g^2$, \begin{eqnarray} \ddot{q} & = & -\omega^2 q -\frac{d^2}{dt^2} \Biggl\{ \frac{g}{2 \ell \omega^4} \Biggl[ -\omega^2 q - \frac{d^2}{dt^2} \Bigl( \frac{g \ddot{q}^2}{2 \ell \omega^4}\Bigr)\Biggr]^2\Biggr\} \; , \\ & = & -\omega^2 q + \frac{g}{\ell} \Bigl( \omega^2 q^2 \!-\! \dot{q}^2\Bigr) + \frac{g^2}{2 \ell^2 \omega^4} \, q \frac{d^2}{dt^2} \Bigl( \ddot{q}^2\Bigr) \nonumber \\ & & \hspace{.5cm} - \frac{g^2}{2 \ell^2 \omega^6} \frac{d^2}{dt^2} \Bigl[ q \frac{d^2}{dt^2} \Bigl( \ddot{q}^2 \Bigr)\Bigr] - \frac{g^3}{8 \ell^3 \omega^{12}} \frac{d^2}{ dt^2} \Biggl[ \frac{d^2}{dt^2} \Bigl( \ddot{q}^2\Bigr) \Biggr]^2 \; . \end{eqnarray} This obviously becomes complicated fast! However, the lower derivative terms at order $g^2$ are simple enough to give if I don't worry about the higher derivative remainder, \begin{equation} \ddot{q} = -\omega^2 q + \frac{g}{\ell} \Bigl( \omega^2 q^2 \!-\! \dot{q}^2 \Bigr) + \frac{g^2}{\ell^2} \Bigl( -6 \omega^2 q^3 \!+\! 14 q \dot{q}^2\Bigr) + O(g^3) \; . \end{equation} If we carry this out to infinite order, {\it and drop the infinite derivative remainder}, the result is an equation of the traditional form, \begin{equation} \ddot{q} = f(q,\dot{q}) \; . \end{equation} The canonical version of this equation gives the first of the desired constraints. The second is obtained from the canonical version of its time derivative. The constrained system we have just described is consistent on the perturbative level, but not beyond. It does not follow from the original, exact equation. That would be no problem if we could define physics using perturbation theory, but we cannot. Perturbation theory does not converge for any known interacting, continuum field theory in $3\!+\!1$ dimensions! The fact that the constraints are not consistent beyond perturbation theory means there is a nonperturbative amplitude for the system to decay to the arbitrarily high excitation in the manner described in sub-section \ref{sub:3.1}. The fact that the constraints treat time derivatives differently than space derivatives also typically leads to a loss of causality and Lorentz invariance beyond perturbation theory. A final comment concerns the limit of small coupling constant, i.e., $g \rightarrow 0$. One can see from the frequencies (\ref{lowk}-\ref{highk}) of our higher derivative oscillator that the negative energy frequency diverges for $g \rightarrow 0$. Disingenuous purveyors of higher derivative models sometimes appeal to people's experience with {\it positive energy} modes by arguing that, ``the $k_-$ mode approaches infinite frequency for small coupling so it must drop out.'' That is false! The argument is quite correct for an infinite frequency {\it positive} energy mode in a stable theory. In that case exciting the mode costs an infinite amount of energy which would have to be drawn from de-exciting finite frequency modes. However, a {\it negative} energy mode doesn't decouple as its frequency diverges. Rather it couples {\it more strongly} because taking its frequency to infinity opens up more and more ways to balance its negative energy by exciting finite frequency, positive energy modes. \subsection{Quantization} \label{sub:3.3} People sometimes imagine that quantization might stabilize a system against the Ostrogradskian instability the same way that it does for the Hydrogen atom coupled to electromagnetism. This is a failure to understand correspondence limits. Conclusions drawn from classical physics survive quantization unless they depend upon the system either being completely excluded from some region of the canonical phase space or else inhabiting only a small region of it. For example, the classical instability of the Hydrogen atom (when coupled to electromagnetism) derives from the fact that the purely Hydrogenic part of the energy, \begin{equation} E_{\rm Hyd} = \frac{\Vert \vec{p}\Vert^2}{2m} - \frac{e^2}{\Vert \vec{x}\Vert} \; . \end{equation} can be made arbitrarily negative by placing the electron close to the nucleus at fixed momentum. Because this instability depends upon the system being in a very small region of the canonical phase space, one might doubt that it survives quantization, and explicit computation shows that it does not. In contrast, the Ostrogradskian instability derives from the fact that $P_1 Q_2$ can be made arbitrarily negative by taking $P_1$ either very negative, for positive $Q_2$, or else very positive, for negative $Q_2$. {\it This covers essentially half the classical phase space!} Further, the variables $Q_2$ and $P_1$ commute with one another in Ostrogradskian quantum mechanics. So there is no reason to expect that the Ostrogradskian instability is unaffected by quantization. \subsection{Unitarity vs. Instability} \label{sub:3.4} Particle physicists who quantize higher derivative theories don't typically recognize a problem with the stability. They maintain that the problem with higher derivatives is a breakdown of unitarity. In this sub-section I will again have recourse to the higher derivative oscillator (\ref{HDO}) to explain the connection between the two apparently unrelated problems. Let us find the ``empty'' state wavefunction, $\Omega(Q_1,Q_2)$ that has the minimum excitation in both the positive and negative energy degrees of freedom. The procedure for doing this is simple: first identify the positive and negative energy lowering operators $\alpha_{\pm}$ and then solve the equations, \begin{equation} \alpha_+ \vert \Omega \rangle = 0 = \alpha_- \vert \Omega \rangle \; . \label{wavef} \end{equation} We can recognize the raising and lowering operators by simply expressing the general solution (\ref{gensol}) in terms of exponentials, \begin{eqnarray} \lefteqn{q(t) = \frac12 (A_+ \!+\! i B_+) e^{-ik_+ t} + \frac12 (A_+ \!-\! i B_+) e^{ik_+ t} } \nonumber \\ & & \hspace{2cm} + \frac12 (A_- \!+\! i B_-) e^{-ik_-t } + \frac12 (A_- \!-\! i B_-) e^{ik_- t} \; . \end{eqnarray} Recall that the $k_+$ mode carries positive energy, so its lowering operator must be proportional to the $e^{-ik_+ t}$ term, \begin{eqnarray} \alpha_+ & \sim & A_+ + i B_+ \; , \\ & \sim & \frac{m k_+}2 \Bigl(1 \!+\! \sqrt{1 \!-\! 4g}\Bigr) Q_1 + i P_1 - k_+ P_2 - \frac{i m}2 \Bigl(1 \!-\! \sqrt{1 \!-\! 4g}\Bigr) Q_2 \; . \end{eqnarray} The $k_-$ mode carries negative energy, so its lowering operator must be proportional to the $e^{+i k_- t}$ term, \begin{eqnarray} \alpha_- & \sim & A_- - i B_- \; , \\ & \sim & \frac{m k_-}2 \Bigl(1 \!-\! \sqrt{1 \!-\! 4g}\Bigr) Q_1 - i P_1 - k_- P_2 + \frac{i m}2 \Bigl(1 \!+\! \sqrt{1 \!-\! 4g}\Bigr) Q_2 \; . \end{eqnarray} Writing $P_i = -i \frac{\partial}{\partial Q_i}$ we see that the unique solution to (\ref{wavef}) has the form, \begin{equation} \Omega(Q_1,Q_2) = N \exp\Biggl[-\frac{m \sqrt{1 \!-\! 4g}}{2 (k_+ \!+\! k_-)} \Bigl(k_+ k_- Q_1^2 + Q_2^2\Bigr) - i \sqrt{g} m Q_1 Q_2\Biggr] \; . \label{true} \end{equation} The empty wave function (\ref{true}) is obviously normalizable, so it gives a state of the quantum system. We can build a complete set of normalized stationary states by acting arbitrary numbers of $+$ and $-$ raising operators on it, \begin{equation} \vert N_+ , N_-\rangle \equiv \frac{(\alpha_+^{\dagger})^N_+}{\sqrt{N_+ !}} \frac{(\alpha_-^{\dagger})^N_-}{\sqrt{N_- !}} \vert \Omega \rangle \; . \end{equation} On this space of states the Hamiltonian operator is unbounded below, just as in the classical theory, \begin{equation} H \vert N_+ , N_- \rangle = \Bigl(N_+ k_+ - N_- k_-\Bigr) \vert N_+ , N_- \rangle \; . \end{equation} This is the correct way to quantize a higher derivative theory. One evidence of this fact is that classical negative energy states correspond to quantum negative energy states as well. Particle physicists don't quantize higher derivative theories as we just have. What they do instead is to regard the negative energy lowering operator as a positive energy raising operator. So they define a ``ground state'' $\vert \overline{\Omega} \rangle$ which obeys the equations, \begin{equation} \alpha_+ \vert \overline{\Omega} \rangle = 0 = \alpha_-^{\dagger} \vert \overline{\Omega} \rangle \; . \label{falsewave} \end{equation} The unique wave function which solves these equations is, \begin{equation} \overline{\Omega}(Q_1,Q_2) = N \exp\Biggl[-\frac{m \sqrt{1 \!-\! 4g}}{2 (k_- \!-\! k_+)} \Bigl(k_+ k_- Q_1^2 - Q_2^2\Bigr) + i \sqrt{g} m Q_1 Q_2 \Biggr] \; . \label{wrong} \end{equation} This wave function is {\it not} normalizable, so it doesn't correspond to a state of the quantum system. At this stage we should properly call a halt to the analysis because we aren't doing quantum mechanics anymore. The Schrodinger equation $H \psi(Q) = E \psi(Q)$ is just a second order differential equation. It has two linearly independent solutions {\it for every} energy $E$: positive, negative, real, imaginary, quaternionic --- it doesn't matter. The thing that puts the ``quantum'' in quantum mechanics is requiring that the solution be normalizable. Many peculiar things can happen if we abandon allow normalizability \cite{RPW1,TW0}. However, my particle theory colleagues ignore this little problem and define a completely formal ``space of states'' based upon $\vert\overline{\Omega} \rangle$, \begin{equation} \vert \overline{N_+ , N_-}\rangle \equiv \frac{(\alpha_+^{\dagger})^{N_+}}{ \sqrt{N_+ !}} \frac{(\alpha_-)^{N_-}}{\sqrt{N_- !}} \vert \overline{\Omega} \rangle \; . \end{equation} None of these wavefunctions is any more normalizable than $\overline{\Omega}( Q_1,Q_2)$, so not a one of them corresponds to a state of the quantum system. However, they are all positive energy eigenfunctions, \begin{equation} H \vert \overline{N_+ , N_-} \rangle = \Bigl(N_+ k_+ + N_- k_-\Bigr) \vert \overline{N_+ , N_-} \rangle \; . \end{equation} My particle physics colleagues typically say they {\it define} $\vert \overline{\Omega}\rangle$ to have unit norm. Because they have not changed the commutation relations, \begin{equation} [\alpha_+,\alpha_+^{\dagger}] = 1 = [\alpha_-,\alpha_-^{\dagger}] \; , \end{equation} the norm of any state with odd $N_-$ is negative! The lowest of these is, \begin{equation} \langle \overline{0,1} \vert \overline{0,1}\rangle = \langle \overline{\Omega} \vert \alpha_-^{\dagger} \alpha_- \vert \overline{ \Omega} \rangle = - \langle \overline{\Omega} \vert \overline{\Omega} \rangle \; . \end{equation} As I pointed out above, the reason this has happened is that we aren't doing quantum mechanics any more. We ought to use the normalizable, but indefinite energy eigenstates. What particle physicists do instead is to reason that because the probabilistic interpretation of quantum mechanics requires norms to be positive, the negative norm states must be excised from the space of states. At this stage good particle physicists note that that the resulting model fails to conserve probability \cite{KS}. Just as the correctly-quantized, indefinite-energy theory allows processes which mix positive and negative energy particles, so too the indefinite-norm theory allows processes which mix positive and negative norm particles. It only conserves probability on the space of ``states'' which includes both kinds of norms. If we excise the negative norm states then probability is no longer conserved. So good particle physicists reach the correct conclusion --- that nondegenerate higher derivative theories can't describe our universe --- by a somewhat illegitimate line of reasoning. But who cares? They got the right answer! Of course {\it bad} particle physicists regard the breakdown of unitarity as a challenge for inspired tinkering to avoid the problem. Favorite ploys are the Lee-Wick reformulation of quantum field theory \cite{LW} and nonperturbative resummations. The analysis also typically involves the false notion that high frequency ghosts decouple, which I debunked at the end of sub-section \ref{sub:3.2}. When the final effort is written up and presented to the world, some long-suffering higher derivative expert gets called away from his research to puzzle out what was done and explain why it isn't correct. {\it Sigh}. The problem is so much clearer in its negative energy incarnation! I could list many examples at this point, but I will confine myself to citing a full-blown paper debunking one of them \cite{TW1}. It is also appropriate to note that Hawking and Hertog have previously called attention to the mistake of quantizing higher derivative theories using nonnormalizable wave functions \cite{HH}. \subsection{Constraints} \label{sub:3.5} The only way anyone has ever found to avoid the Ostrogradskian instability on a nonperturbative level is by violating the single assumption needed to make Ostrogradski's construction: nondegeneracy. Higher derivative theories for which the definition of the highest conjugate momentum (\ref{nondeg}) cannot be inverted to solve for the highest derivative can sometimes be stable. An interesting example of this kind is the rigid, relativistic particle studied by Plyushchay \cite{MSP,DZ}. Degeneracy is of great importance because {\it all theories which possess continuous symmetries are degenerate,} irrespective of whether or not they possess higher derivatives. A familiar example is the relativistic point particle, whose dynamical variable is $X^{\mu}(\tau)$ and whose Lagrangian is, \begin{equation} L = -m \sqrt{-\eta_{\mu\nu} \dot{X}^{\mu} \dot{X}^{\nu}} \; . \end{equation} The conjugate momentum is, \begin{equation} P_{\mu} \equiv \frac{m \dot{X}_{\mu}}{\sqrt{-\dot{X}^2}} \; . \end{equation} Because the right hand side of this equation is homogeneous of degree zero one can not solve for $\dot{X}^{\mu}$. The associated continuous symmetry is invariance under reparameterizations $\tau \rightarrow \tau'(\tau)$, \begin{equation} X^{\mu}(\tau) \longrightarrow X^{\prime \mu}(\tau) \equiv X^{\mu}\Bigl({\tau' }^{-1}(\tau)\Bigr) \; . \end{equation} The cure for symmetry-induced degeneracy is simply to fix the symmetry by imposing gauge conditions. Then the gauge-fixed Lagrangian should no longer be degenerate in terms of the remaining variables. For example, we might parameterize so that $\tau = X^0(\tau)$, in which case the gauge-fixed particle Lagrangian is, \begin{equation} L_{\rm GF} = -m \sqrt{1 - \dot{\vec{X}} \cdot \dot{\vec{X}} } \; . \end{equation} In this gauge the relation for the momenta is simple to invert, \begin{equation} P_i \equiv \frac{m \dot{X}_i}{\sqrt{1 - \dot{\vec{X}} \cdot \dot{\vec{X}} }} \qquad \Longleftrightarrow \qquad \dot{X}^i = \frac{P^i}{\sqrt{m^2 + \vec{P} \cdot \vec{P}}} \; . \end{equation} When a continuous symmetry is used to eliminate a dynamical variable, the equation of motion of this variable typically becomes a {\it constraint}. For symmetries enforced by means of a compensating field --- such as local Lorentz invariance is with the antisymmetric components of the vierbein \cite{RPW2} --- the associated constraints are tautologies of the form $0 = 0$. Sometimes the constraints are nontrivial, but implied by the equations of motion. An example of this kind is the relativistic particle in our synchronous gauge. The equation of the gauge-fixed zero-component just tells us the Hamiltonian is conserved, \begin{equation} \frac{d}{d\tau} \Biggl( \frac{m \dot{X}_0}{\sqrt{-\eta_{\mu\nu} \dot{X}^{\mu} \dot{X}^{\nu}}} \Biggr) = 0 \longrightarrow \frac{d}{dt} \Bigl( \sqrt{m^2 + \vec{p} \cdot \vec{p} } \Bigr) = 0 \; . \end{equation} And sometimes the constraints give nontrivial relations between the canonical variables that generate residual, time-independent symmetries. In this case another degree of freedom can be removed (``gauge fixing counts twice,'' as van Nieuwenhuizen puts it). An example of this kind of constraint is Gauss' Law in temporal gauge electrodynamics. Were it not for constraints of this last type, the analysis of a higher derivative theory with a gauge symmetry would be straightforward. One would simply fix the gauge and then check whether or not the gauge-fixed Lagrangian depends nondegenerately upon higher time derivatives. If it did, the conclusion would be that the theory suffers the Ostrogradskian instability. However, when constraints of the third type are present one must check whether or not they affect the instability. This is highly model dependent but a very simple rule seems to be generally applicable: {\it if the number of gauge constraints is less than the number of unstable directions in the canonical phase space then there is no chance for avoiding the problem}. Because the number of constraints for any symmetry is fixed, whereas the number of unstable directions increases with the number of higher derivatives, one consequence is that gauge constraints can at best avoid instability for some fixed number of higher derivatives. For example, the constraints of the second derivative model of Plyushchay are sufficient to stabilize the system \cite{MSP,DZ}, but one would expect it to become unstable if third derivatives were added. People sometimes make the mistake of believing that the Ostrogradskian instability can be avoided with just a single, global constraint on the Hamiltonian. For example, Boulware, Horowitz and Strominger \cite{BHS} showed the energy is zero for any asymptotically flat solution of the higher derivative field equations derived from the Lagrangian, \begin{equation} \mathcal{L} = \alpha R^2 \sqrt{-g} + \beta R^{\mu\nu} R_{\mu \nu} \sqrt{-g} \; . \end{equation} As I explained in sub-section \ref{sub:3.1}, the nature of the Ostrogradskian instability is not that the energy decays but rather that the system evaporates to a very highly excited state of compensating, positive and negative energy degrees of freedom. As long as $\beta \neq 0$, there are six independent, higher derivative momenta at each space point, whereas there are only four local constants --- or five if $\alpha$ and $\beta$ are such as to give local conformal invariance. Hence there are two (or one) unconstrained instabilities per space point. There are an infinite number of space points, so the addition of a single, global constraint does not change anything. I should point out that Boulware, Horowitz and Strominger were aware of this, cf. their discussion of the dipole instability. The case of $\beta = 0$ is special, and significant for the next section. If $\alpha$ has the right sign that model has long been known to have positive energy \cite{AAS0,AS}. This result in no way contradicts the previous analysis. When $\beta = 0$ the terms which carry second derivatives are contracted in such way that only a single component of the metric carries higher derivatives. So now the counting is {\it one} unstable direction per space point versus four local constraints. Hence the constraints can win, and they do if $\alpha$ has the right sign. \subsection{Nonlocality} \label{sub:3.6} I would like to close this section by commenting on the implications of Ostrogradski's theorem for fully nonlocal theories. In addition to nonlocal quantum field theories \cite{KW0,CHY,JJ} this is relevant to string field theory \cite{GJ1,GJ2,GJ3}, to noncommutative geometry \cite{NC,HPR}, to regularization techniques \cite{EMKW,KW1,KW2} and even to theories of cosmology \cite{TW2,SW1,BMS}. The issue in each case is whether or not we can think of the fully nonlocal theory as the limit of a sequence of ever higher derivative theories. When such a representation is possible the nonlocal theory must inherit the Ostrogradskian instability. The higher derivative representation is certainly valid for string field theory because, otherwise, there would be cuts and poles that would interfere with perturbative unitarity. So string field theory suffers from the Ostrogradskian instability \cite{EW}. The same is true for theories where the nonlocality is of limited extent in time \cite{RPW3}, although not everyone agrees \cite{JL,RPW4}. However, when the nonlocality involves inverse differential operators there need be no problem \cite{EW,SW1}. Indeed, the effective action of any quantum field theory is nonlocal in this way \cite{BGVZ,BM}! Nor is there necessarily any problem when the nonlocality arises in the form of algebraic functions of local actions \cite{BNW}. \section{$\Delta R[g] = f(R)$ Theories} \label{sec:4} From the lengthy argumentation of the previous two sections one might conclude that the only potentially stable, local modification of gravity is a cosmological constant, $\Delta R[g] = - 2\Lambda$. However, a close analysis of sub-section \ref{sub:3.5} reveals that it is also possible to consider algebraic functions of the Ricci scalar. In this section I first explain why such theories can avoid the Ostrogradskian instability. I then demonstrate that they are equivalent to general relativity with a minimally coupled scalar, provided we ignore matter. Finally, I exploit this equivalence, with the construction described in the Introduction, to show how $f(R)$ can be chosen to enforce any evolution $a(t)$. \subsection{Why They Can Be Stable} \label{sub:4.1} The alert reader will have noted that the $R + R^2$ model \cite{AAS0,AS} avoids the Ostrogradskian instability. It does this by violating Ostrogradski's assumption of nondegeneracy: the tensor indices of the second derivative terms in the Ricci scalar are contracted together so that only a single component of the metric carries higher derivatives. This component does acquire a new, higher derivative degree of freedom, and the energy of this degree of freedom is indeed opposite to that of the corresponding lower derivative degree of freedom, just as required by Ostrogradski's analysis. However, that lower derivative degree of freedom is the {\it Newtonian potential}. It carries negative energy, but it is also completely fixed in terms of the other metric and matter fields by the $g_{00}$ constraint. So the only instability associated with it is gravitational collapse. Its higher derivative counterpart has positive energy, at least on the kinetic level; it can still have a bad potential, and the model is indeed only stable for one sign of the $R^2$ term. None of these features depended especially upon the higher derivative term being $R^2$. Any function for the Ricci scalar would work as well. Note that we cannot allow derivatives of the Ricci scalar, because Ostrogradski's theorem says the next higher derivative degree of freedom would carry negative energy and there would be no additional constraints to protect it. We also cannot permit more general contractions of the Riemann tensor because then other components of the metric would carry higher derivatives. These components are positive energy in general relativity, so their higher derivative counterparts would be negative, and there would again be no additional constraints to protect the theory against instability. \subsection{Equivalent Scalar Representation} \label{sub:4.2} The general Lagrangian we wish to consider takes the form, \begin{equation} \mathcal{L} = \frac1{16 \pi G} \Bigl( R + f(R)\Bigr) \sqrt{-g} \; . \end{equation} If we ignore the coupling to matter the modified gravitational field equation consists of the vanishing of the following tensor, \begin{equation} \frac{16 \pi G}{\sqrt{-g}} \frac{\delta S}{\delta g^{\mu\nu}} = [1 \!+\! f'(R)] R_{\mu\nu} - \frac12 [R \!+\! f(R)] g_{\mu\nu} + g_{\mu\nu} [f'(R)]^{;\rho}_{ ~\rho} - [f'(R)]_{;\mu\nu} \; . \label{MGR} \end{equation} There is an old procedure for reformulating this as general relativity with a minimally coupled scalar. I don't know whom to credit, but I will give the construction. The first step is to define an ``equivalent'' theory with an auxiliary field $\phi$ which is defined by the relation. \begin{equation} \phi \equiv 1 + f'(R) \qquad \Longleftrightarrow \qquad R = \mathcal{R}(\phi) \; . \end{equation} Inverting the relation determines the Ricci scalar as an algebraic function of $\phi$. We can then define an auxiliary potential for $\phi$ by Legendre transformation, \begin{equation} U(\phi) \equiv \Bigl(\phi \!-\! 1\Bigr) \mathcal{R}(\phi) - f\Bigl(\mathcal{R}(\phi)\Bigr) \qquad \Longrightarrow \qquad U'(\phi) = \mathcal{R}(\phi) \; . \end{equation} Now consider the equivalent scalar-tensor theory whose Lagrangian is, \begin{equation} \mathcal{L}_{\rm E} \equiv \frac1{16 \pi G} \Bigl(\phi R - U(\phi)\Bigr) \sqrt{-g} \; . \end{equation} Its field equations are, \begin{eqnarray} \frac{16 \pi G}{\sqrt{-g}} \frac{\delta S_{\rm E}}{\delta \phi} & = & R - U'(\phi) = 0 \; , \label{E11} \\ \frac{16 \pi G}{\sqrt{-g}} \frac{\delta S_{\rm E}}{\delta g^{\mu\nu}} & = & \phi R_{\mu\nu} - \frac12 \Bigl(\phi R \!-\! U(\phi)\Bigr) g_{\mu\nu} + g_{\mu\nu} \phi^{;\rho}_{~\rho} - \phi_{\mu\nu} = 0 \; . \label{E12} \end{eqnarray} The scalar equation (\ref{E11}) implies $\phi \!=\! 1 \!+\! f'(R)$, whereupon the tensor equations (\ref{E12}) reproduce the original modified gravity equations (\ref{MGR}). The final step is to define a new metric $\widetilde{g}_{\mu\nu}$ and a new scalar $\varphi$ by the change of variables, \begin{eqnarray} \widetilde{g}_{\mu\nu} \equiv \phi \, g_{\mu\nu} \qquad & \Longleftrightarrow & \qquad g_{\mu\nu} = \exp\Bigl[-\sqrt{\frac{4\pi G}3} \, \varphi\Bigr] \, \widetilde{g}_{\mu\nu} \; , \label{gtrans} \\ \varphi \equiv \sqrt{\frac3{4\pi G}} \, \ln(\phi) \qquad & \Longleftrightarrow & \qquad \phi = \exp\Bigl[\sqrt{\frac{4\pi G}3} \, \varphi\Bigr] \; . \end{eqnarray} In terms of these variables the equivalent Lagrangian takes the form, \begin{equation} \mathcal{L}_E = \frac1{16 \pi G} \widetilde{R} \sqrt{-\widetilde{g}} -\frac12 \partial_{\mu} \varphi \partial_{\nu} \varphi \, \widetilde{g}^{\mu\nu} \sqrt{-\widetilde{g}} - V(\varphi) \sqrt{-\widetilde{g}} \; , \label{finalL} \end{equation} where the scalar potential is, \begin{equation} V(\varphi) \equiv \frac1{16 \pi G} U\Biggl(\exp\Bigl[ \sqrt{\frac{4 \pi G}3} \, \varphi\Bigr]\Biggr) \exp\Bigl[-\sqrt{\frac{16 \pi G}3} \, \varphi\Bigr] \; . \end{equation} This is general relativity with a minimally coupled scalar, as claimed. \subsection{Reconstructing $f(R)$ from Cosmology} \label{sub:4.3} I want to show how to choose $f(R)$ to support an arbitrary $a(t)$.\footnote{ For a somewhat different construction which achieves the same end, see \cite{NO0,NO00}.} Recall from the Introduction that one can choose the potential of a quintessence model such as (\ref{finalL}) to support any homogeneous and isotropic cosmology for its metric $\widetilde{g}_{\mu\nu}$. However, we cannot immediately exploit this construction because it is the metric $g_{\mu\nu}$ which is assumed known, not $\widetilde{g}_{\mu\nu}$. We must explain how to infer the one from the other without knowing $f(R)$. Because the relation (\ref{gtrans}) between $g_{\mu\nu}$ and $\widetilde{g}_{ \mu\nu}$ is a conformal transformation, it makes sense to work in a coordinate system in which each metric is conformal to flat space. This is accomplished by changing from co-moving time $t$ to conformal time $\eta$ though the relation, $d\eta = dt/a(t)$, \begin{equation} ds^2 = -dt^2 + a^2(t) d\vec{x} \cdot d\vec{x} = a^2 \Bigl(-d\eta^2 + d\vec{x} \cdot d\vec{x}\Bigr) \; . \end{equation} The $\widetilde{g}_{\mu\nu}$ element takes the same form in conformal coordinates, but note that its different scale factor implies a different co-moving time, \begin{equation} d\widetilde{s}^2 = \widetilde{a}^2 \Bigl(-d\eta^2 + d\vec{x} \cdot d\vec{x}\Bigr) = -d\widetilde{t}^{~2} + \widetilde{a}^2(\widetilde{t}\,) d\vec{x} \cdot d\vec{x} \; . \end{equation} From relation (\ref{gtrans}) we infer, \begin{equation} a(t) = \widetilde{a}(\widetilde{t}\,) \exp\Bigl[-\sqrt{\frac{\pi G}3} \, \varphi_0(\widetilde{t}\,) \Bigr] \; . \label{keyrel} \end{equation} We denote differentiation with respect to $\eta$ by a prime, and one should note the relation between derivatives with respect to the various times, \begin{equation} \frac{\partial}{\partial \eta} = a \frac{\partial}{\partial t} = \widetilde{a} \frac{\partial}{\partial \widetilde{t}} \; . \end{equation} Differentiating the logarithm of (\ref{keyrel}) with respect to $\eta$ and using the relation (\ref{twoeqns}) between $\widetilde{a}$ and $\varphi_0$ gives, \begin{equation} \frac{a'}{a} = \frac{\widetilde{a}'}{\widetilde{a}} -\sqrt{\frac{\pi G}3} \, \varphi_0' = \frac{\widetilde{a}'}{\widetilde{a}} -\sqrt{-\frac1{12} \widetilde{a}'} \; . \end{equation} This is a nonlinear but first order differential equation for the variable $\widetilde{a}$ in terms of the known function, $a(t(\eta))$. At the worst it can be solved numerically. Once we have $\widetilde{a}$ the potential $V(\varphi)$ can be constructed using the procedure explained in the Introduction. We then compute the auxiliary potential, \begin{equation} U(\phi) = 16 \pi G \phi^2 V\Bigl( \sqrt{\frac3{4\pi G}} \, \ln(\phi)\Bigr) \; . \end{equation} The auxiliary field can be expressed in terms of the Ricci scalar from the algebraic relation, \begin{equation} U'(\phi) = R \qquad \Longleftrightarrow \qquad \phi = \Phi(R) \; . \end{equation} And we finally recover the function $f(R)$ by Legrendre transformation, \begin{equation} f(R) = \Bigl(\Phi(R) \!-\! 1\Bigr) R - U\Bigl(\Phi(R)\Bigr) \; . \end{equation} \section{Problems with $f(R) = -\frac{\mu^4}{R}$} \label{sec:5} In view of the construction of sub-section \ref{sub:4.3} it is not surprising but rather {\it inevitable} that an $f(R)$ can be found to support late time acceleration, or indeed, any other evolution. However, the method is not guaranteed to produce a simple model, so the discovery that $f(R) = -\mu^4/R$ works is quite noteworthy \cite{CDTT,CCT}.\footnote{Although extensions involving $R^{\mu\nu} R_{\mu\nu}$ and $R^{\rho\sigma\mu\nu} R_{\rho\sigma\mu\nu}$ have also been studied \cite{CDDETT}, they must be ruled out on account of the Ostrogradskian instability.} It may also be significant that models of this type seem to follow from fundamental theory \cite{NO1}. To derive acceleration in this model consider its field equations, \begin{equation} \Bigl(1 \!+\! \frac{\mu^4}{R^2}\Bigr) R_{\mu\nu} - \frac12 \Bigl(1 \!-\! \frac{\mu^4}{R^2}\Bigr) R g_{\mu\nu} + \Bigl(g_{\mu\nu} \square - D_{\mu} D_{\nu}\Bigr) \frac{\mu^4}{R^2} = 8 \pi G T_{\mu\nu} \; . \label{theeqn} \end{equation} Setting $T_{\mu\nu} \!=\! 0$ and searching for constant Ricci scalar solutions gives, \begin{equation} \Bigl(1 \!+\! \frac{\mu^4}{R^2}\Bigr) R_{\mu\nu} - \frac12 \Bigl(1 \!-\! \frac{\mu^4}{R^2}\Bigr) R g_{\mu\nu} = 0 \qquad \Longleftrightarrow \qquad R_{\mu\nu} = \pm \frac{\sqrt{3}}4 \mu^2 g_{\mu\nu} \; . \end{equation} The plus sign corresponds to acceleration. In addition to proposing the model, Carroll, Duvvuri, Trodden and Turner \cite{CDTT} also showed that it suffers from a very weak tachyonic instability in the absence of matter. Because the only new higher derivative degree of freedom resides in the Ricci scalar, we may as well derive an equation for it alone from the trace of (\ref{theeqn}), \begin{equation} -R + \frac{3\mu^4}{R} + \square \Bigl(\frac{3\mu^4}{R^2}\Bigr) = 0 \; . \end{equation} Now perturb about the accelerated solution, \begin{equation} R = +\sqrt{3} \mu^2 + \delta R \quad \Longrightarrow \quad -2 \delta R -\frac{2}{\sqrt{3} \mu^2} \square \delta R + O(\delta R^2) = 0 \; . \end{equation} By comparing the linearized equation for $\delta R$ with that of a positive mass-squared scalar, \begin{equation} (\square - m^2) \varphi = 0 \; , \label{comp} \end{equation} we see that $\delta R$ behaves like a tachyon with $m^2 = -\sqrt{3} \mu^2$. However, because explaining the current phase of acceleration requires $\mu \sim 10^{-33}~{\rm eV}$, the resulting instability is not very serious. I should note that the existence of a tachyonic instability in no way contradicts the Ostrogradskian analysis that this model's higher derivative degree of freedom carries positive kinetic energy. \subsection{Inside Matter} \label{sub:5.1} Dolgov and Kawasaki \cite{DK} showed that a radically different result emerges when this model is considered inside a static distribution of matter, \begin{equation} T_{\mu\nu} = \rho \delta_{\mu}^0 \delta_{\nu}^0 \qquad {\rm with} \qquad 8 \pi G \rho \equiv M^2 \gg \mu^2 \; . \label{Dolgov} \end{equation} In that case the trace of (\ref{theeqn}) gives, \begin{equation} -R + \frac{3 \mu^4}{R} + \square \Bigl(\frac{3 \mu^4}{R^2} \Bigr) = - M^2 \; . \end{equation} As might be expected, the static Ricci scalar solution in this case is dominated by $M$ rather than $\mu$, \begin{equation} R_0 = \frac12 \Bigl(M^2 \!+\! \sqrt{M^4 \!+\! 12 \mu^4}\Bigr) \simeq M^2 \; . \end{equation} Perturbing about this solution gives, \begin{equation} R = R_0 + \delta R \quad \Longrightarrow \quad - \delta R -\frac{3\mu^4}{ R_0^2} \delta R -\frac{6 \mu^4}{R_0^3} \square \delta R + O(\delta R^2) = 0\; . \end{equation} Comparing with the reference scalar (\ref{comp}) now reveals an enormous tachyonic mass, \begin{equation} m^2 = -\frac{R_0}2 -\frac{R_0^3}{6 \mu^4} \simeq -\frac{M^6}{6 \mu^4} \; ! \end{equation} Plugging in the numbers for the density of water ($\rho \sim 10^3~{\rm kg/m}^3$) gives $M \sim 10^{-18}~{\rm eV}$, implying a tachyonic mass of magnitude $\vert m\vert \sim 10^{12}~{\rm eV} = 10^3~{\rm GeV}$! As disastrous as this problem might seem, Dick \cite{RD} and Nojiri and Odintsov \cite{NO2} have shown that it can be avoided by changing the model slightly, \begin{equation} f(R) = -\frac{\mu^4}{R} + \frac{\alpha}{2 \mu^2} R^2 \quad \Longrightarrow \quad -R + \frac{3 \mu^4}{R} + 3 \square \Bigl( \frac{\mu^4}{R^2} + \frac{\alpha}{\mu^2} R \Bigr) = 0 \; . \label{Ext} \end{equation} Because an $R^2$ term has global conformal invariance, it makes no contribution to the trace for constant $R$. Hence the cosmological solution of $R = + \sqrt{3} \mu^2$ is not affected, nor is the static solution inside the matter distribution (\ref{Dolgov}). However, the equation for linearized perturbations inside matter changes to, \begin{equation} -\delta R - \frac{3 \mu^4}{R_0^2} \delta R + 3 \Bigl(-\frac{2 \mu^4}{R_0^3} \!+\! \frac{\alpha}{\mu^2} \Bigr) \square \delta R = 0 \; . \end{equation} The instability of Dolgov and Kawasaki was driven by the smallness of $2\mu^4/ R_0^3$. By simply taking $\alpha$ positive and of order one the tachyon becomes a positive mass-squared particle of $m^2 \sim \mu^2/\alpha$. \subsection{Outside Matter} \label{sub:5.2} Marc Soussa and I analyzed force of gravity outside a matter distribution \cite{SW2}. Although our analysis was for the original $f(R)= -\mu^4/R$ model, there would be only slight differences for the extended model (\ref{Ext}). So our result seems to foreclose this possibility, but see \cite{NO3}. The tachyonic instability could be studied using the perturbed Ricci scalar, but the gravitational force requires use of the metric. We perturbed about the de Sitter solution with Hubble constant $H = \mu/(48)^{\frac14}$ in co-moving coordinates, \begin{equation} ds^2 = -(1 \!-\! h_{00}) dt^2 + 2 a(t) h_{0i} dt dx^i + a^2(t) (\delta_{ij} \!+\! h_{ij}) dx^i dx^j \quad {\rm with} \quad a(t) = e^{H t} \; . \end{equation} In the gauge, \begin{equation} h_{\mu\nu}^{~~,\nu} - \frac12 h_{\mu} + 3 h_{\mu}^{~\nu} [\ln(a)]_{,\nu} = 0 \; , \end{equation} with $h \equiv -h_{00} \!+\! h_{ii}$, the perturbed Ricci scalar takes the form, \begin{equation} \delta R = -\frac12 \partial^2 h + 2 H \partial_0 h \; . \label{deltaR} \end{equation} Our strategy was first to solve the de Sitter invariant equation for the perturbed Ricci scalar, then reconstruct the gauge-fixed metric. We assumed a matter density of the form, \begin{equation} \rho(t,\vec{x}) = \frac{3 M}{4 \pi R_g^3} \theta\Bigl(R_g - a(t) \vert \vec{x} \vert \Bigr) \; . \end{equation} The exterior field equation has a simple expression in terms of the coordinate $y \equiv a(t) H \vert \vec{x}\vert$, \begin{equation} \Biggl[\Bigl(1 \!-\! y^2\Bigr) \frac{d^2}{dy^2} + \frac2{y} \Bigl(1 \!-\! 2 y^2 \Bigr) \frac{d}{dy} + 12 \Biggr] \delta R = 0 \; . \end{equation} The solution takes the form, \begin{equation} \delta R = \beta_1 f_0(y) + \beta_2 f_{-1}(y) \; , \label{genform} \end{equation} where $f_0$ and $f_{-1}$ are hypergeometric functions whose series expansions are, \begin{eqnarray} f_0(y) & = & 1 - 2 y^2 + \frac15 y^4 + \ldots \; , \\ f_{-1}(y) & = & \frac1{y} \Bigl( 1 - 7 y^2 + \frac{14}3 y^4 + \ldots \Bigr) \; . \end{eqnarray} We only need the behavior for small $y$ because $y \!=\! 1$ is the Hubble radius! Matching to the source at $y = H R_g$ determines the combination coefficients to be, \begin{equation} \beta_1 \simeq \frac{3 G M}{R_g^3} \qquad , \qquad \beta_2 \simeq -12 G M H^3 \; . \end{equation} This last step might seem bogus because we needed to regard the mass density as a small perturbation on the cosmological energy density $\mu^4$, whereas the opposite would be the case for galaxies or clusters of galaxies. However, this will only make changes of order one in the $\beta_i$'s. In particular, the asymptotic solution must still take the form (\ref{genform}). The next step is solving for the trace of the perturbed metric. It turns out that relation (\ref{deltaR}) can also be expressed very simply using the variable $y$, \begin{equation} \Biggl[ \Bigl(y^2 \!-\! 1\Bigr) \frac{d}{dy} + \frac1{y} \Bigl(5 y^2 \!-\!2 \Bigr) \Biggr] h'(y) = \frac{2}{H^2} \delta R \; . \end{equation} We only need to solve for the derivative of $h$ because that is what gives the gravitational force in the geodesic equation. The solution is, \begin{equation} h'(y) = -\frac{2 G M}{H^2 R_g^3} y + O(y^3) \; . \end{equation} This should be compared to the general relativistic prediction, \begin{equation} h'_{\rm GR}(y) = -\frac{4 G M H}{y^2} + O(1) \qquad \Longrightarrow \qquad \frac{h'}{h'_{\rm GR}} = \frac12 \Bigl(\frac{\Vert \vec{x}\Vert}{R_g}\Bigr)^3 \; . \end{equation} One consequence is that the force between the Milky Way and Andromeda galaxies would be about a million times larger than predicted by general relativity! \section{Conclusions} \label{sec:6} The potential of a quintessence scalar can be chosen to support any cosmology, but the epicyclic nature of this construction suggests we consider modifications of gravity. Ostrogradski's theorem \cite{MO} limits local modifications of gravity to just algebraic functions of the Ricci scalar. Models of this form can give a late phase of cosmic acceleration such as we are currently experiencing. However, they can be tuned to give anything else as well. They seem every bit as epicyclic as scalar quintessence. Further, the $f(r) = -\mu^4/R$ model is problematic, both inside and outside matter sources.\footnote{Observations also rule out the somewhat different version of this model that results from regarding the connection and the metric as independent, fundamental variables in the Palatini formalism \cite{AEMM}.} An interesting and largely overlooked possibility for modifying gravity is the fully nonlocal effective action that results from quantum gravitational corrections. In weak field perturbation theory it has long been known that the most cosmologically significant one loop corrections are not of the $R^2$ form usually studied but rather of the form $R \ln(\square) R$ \cite{EMV}. More potentially interesting is the possibility of very strong infrared effects from the epoch of primordial inflation \cite{TW3,RBM}. It can be shown that quantum gravitational corrections to the inflationary expansion rate grow with time like powers of $\ln(a)$. Although suppressed by very small coupling constants, the exponential growth in $a(t)$ during inflation must eventually cause the effect to become nonperturbatively strong \cite{TW4,TW5}. Similar secular growth occurs as well for minimally coupled scalar field theories \cite{OW1,OW2}, in which context Starobinski\u{\i} has developed a technique for summing the leading powers of $\ln(a)$ at each loop order \cite{AAS,SY}. If Starobinski\u{\i}'s technique can be generalized to quantum gravity \cite{RPW5,TW6} it might result in a nonlocal effective gravity theory for late time cosmology in which a large, bare cosmological constant is almost completely screened by a nonperturbative quantum gravitational effect. In such a formalism the current phase of acceleration might result from a very slight mismatch between the bare cosmological constant and the quantum effect which screens it. It is even conceivable that one could reproduce the phenomenological successes of MOND \cite{MM,SM} with such a nonlocal metric theory, although it would have to unstable against decay into galaxy-scale gravitational waves \cite{SW3}. \vskip 1cm \centerline{Acknowledgements} It is a pleasure to acknowledge conversations and correspondence on this subject with S. Deser, A.D. Dolgov, D.A. Eliezer, S. Odintsov, M.E. Soussa, A. Strominger and M. Trodden. This work was partially supported by NSF grant PHY-244714 and by the Institute for Fundamental Theory at the University of Florida. \printindex
Title: Separation of dwarf and giant stars with ROTSE-IIId
Abstract: 136 stars which were known to be the members of open cluster NGC 752 were observed at R band with ROTSE-IIId telescope located at the Turkish National Observatory (TUG) site. The data had been evaluated together with BV and 2MASS photometric data. A new practical method for separating dwarf and giant was described and applied. Evaluating the colour magnitude--diagrams with Padova isochrones revealed metallicity similar to the Sun and an age of 1.41 Gyr for the open cluster NGC 752.
https://export.arxiv.org/pdf/astro-ph/0601681
\lhead[\thepage]{A.N. Bilir et al.: Separation of dwarf and giant stars with ROTSE-IIId} \rhead[Astron. Nachr./AN~{\bf XXX} (200X) X]{\thepage} \headnote{Astron. Nachr./AN {\bf 32X} (200X) X, XXX--XXX} \title{Separation of dwarf and giant stars with ROTSE--IIId} \author{S. Bilir \and T. G\"uver \and M. Aslan } \institute{Istanbul University Science Faculty, Department of Astronomy and Space Sciences, 34119, University-Istanbul, Turkey} \date{} \abstract{136 stars which were known to be the members of open cluster NGC 752 were observed at $R$ band with ROTSE--IIId telescope located at the Turkish National Observatory (TUG) site. The data had been evaluated together with BV and 2MASS photometric data. A new practical method for separating dwarf and giant was described and applied. Evaluating the colour magnitude--diagrams with Padova isochrones revealed metallicity similar to the Sun and an age of 1.41 Gyr for the open cluster NGC 752. \keywords{Galaxy: open cluster and associations, stars: colour-magnitude diagrams, stars: giants} } \correspondence{sbilir@istanbul.edu.tr} \section{Introduction} One of the problems of the Galactic astronomy is the estimation of Galactic model parameters of giant stars in our Galaxy. In many studies, the Galactic model parameters are estimated without any discrimination between dwarfs and giants, whereas some researchers estimated model parameters only for certain star categories (e.g. Pritchet 1983, Bahcall \& Soneira 1984, Buser \& Kaeser 1985 and Mendez \& Altena 1996). A very recent work is an example for this where the Galactic model parameters were estimated using only giants (Cabrera-Lavers, Garzon \& Hammersley 2005). The separation of field dwarfs and field giants plays an important role for such kinds of works. The most efficient classical methods of identifying dwarf and giant stars utilize spectroscopy. By inspecting spectral line profiles, one has to estimate surface gravity to discriminate between higher and lower pressure stellar atmospheres to be sure for the identification. This is, however time consuming and tiring. A rather easier procedure is to separate dwarfs and evolved stars (subgiants or giants) such as to obtain a luminosity function consistent with the local luminosity function of nearby stars due to Gliese \& Jahreiss (1991) and Jahreiss \& Wielen (1997). The procedure of this separation is based on the fact that the local luminosity functions obtained for many fields indicates a systematic excess of star counts relative to the luminosity function of nearby stars for the fainter segment, i.e. $M(V)\geq5^{m}.5$, and a deficit for brighter segment, $M(V)<5^{m}.5$. (in $RGU$ system $M(G)\geq6^{m}$ and $M(G)<6^{m}$, respectively). The works of Karaali (1992); Ak, Karaali \& Buser (1998); Karata\c{s}, Bilir \& Karaali (2000); Karaali et al. (2000); Karata\c{s}, Karaali \& Buser (2001); Karaali, Bilir \& Buser (2004); Bilir, Karaali \& Buser (2004) and Karata\c{s} et al. (2004) can be given as examples for application of this procedure. Recently, a new method were suggested by Bilir et al. (2006) for separating the field dwarfs and field giants. This new method is based on the comparison of the Two Micron All Sky Survey (2MASS, hereafter) $J$, $H$, $K_{s}$ with the $V$ magnitudes down to the limiting magnitude of $V=16$. In this work we extend application of this method to the open cluster NGC 752 observed by a robotic telescope ROTSE-IIId (Akerlof et al. 2003). This paper is organized as follows. In Section 2 the BV, 2MASS and ROTSE data are presented. In Section 3 the method is applied to ROTSE and 2MASS data, and the separation of dwarf and giant stars is tested. In Section 4 colour-magnitude diagrams (CMDs) of NGC 752 is compared to the Padova isochrones. Finally, the conclusion is given in Section 5. \begin{figure} \center \resizebox{5cm}{7cm}{\includegraphics*{fig01.eps}} \caption {$V \times (B-V)$ CMD of 136 probable member stars in NGC 752.} \end {figure} \begin{figure*} \center \resizebox{17cm}{5.51cm}{\includegraphics*{fig02.eps}} \caption {V and 2MASS magnitudes of 136 stars in NGC 752. (a) $J_{0} \times V_{0}$, (b) $H_{0} \times V_{0}$, and (c) $Ks_{0} \times V_{0}$. The solid lines in diagrams were drawn according to eqs. (1), (2), and (3).} \end {figure*} \section{Observations} \subsection{The BV and 2MASS Data} NGC 752=C0154+374 ($\alpha=01^{h}57^{m}41^{s}$, $\delta=+37^{o}47^{'}06^{''}$; $l=137^{o}.13$, $b=-23^{o}.25$; J2000) has been subject of many studies, because it is the nearest intermediate-age cluster, with 427 pc (Dzervitis \& Paupers 1993) distance from the Sun. It is usually considered as metal-deficient with respect to the Sun, $[Fe/H]=-0.15 \pm$ 0.05 dex, slightly reddened $E(B-V)=0.035 \pm 0.005$, with distance module $(m-M)=8.25\pm$0.10 (Daniel et al. 1994). Accurate proper motion and radial velocity measurements show that there are 136 probable member stars of the open cluster (Daniel et al. 1994). $V$ magnitudes and $(B-V)$ colour indices used in this study were taken from Daniel et al. (1994) and they were given in Table 1. The $V\times(B-V)$ CMD in Fig. 1 shows that stars with $V<10$ and $(B-V)>0.80$ are giants. Recently the 2MASS, including the Point-Source Catalogue and Atlas, has produced huge amounts of data to be explored in the coming years (Skrutskie et al. 1997). The photometric system comprises Johnson's $J$ (1.25 $\mu$m) and $H$ (1.65 $\mu$m) bands with the addition of $K_{s}$ (2.17 $\mu$m), slightly bluer than Johnson's K. The 2MASS sky coverage, homogeneity and depth will certainly make this set of filters a photometric standard reference for the future. 2MASS data of the 136 probable member stars in NGC 752 are obtained by Vizier\footnote{http://vizier.u-strasbg.fr/viz-bin/Vizier?-source=2MASS} in CDS and they are given in Table 1. We used the equations of Fiorucci \& Munari (2003) for the determination of the total absorptions for the bands $V$, $J$, $H$ and $K_{s}$, i.e. $A(V)=3.1E(B-V)$, $A(J)=0.887E(B-V)$, $A(H)=0.565E(B-V)$ and $A(K_{s})=0.382E(B-V)$. Thus the de-reddened magnitudes were obtained as follows: $V_{0}=V-A(V)$, $J_{0}=J-A(J)$, $H_{0}=H-A(H)$ and $(K_{s})_{0}=K_{s}-A(K_{s})$. The subscript ``0'' indicates de-reddened magnitude. $J_{0} \times V_{0}$, $H_{0} \times V_{0}$ and $(K_{s})_{0} \times V_{0}$ diagrams for the cluster stars are given in Fig. 2. The solid lines represent the equations in Bilir et al. (2006), i.e. \begin{equation} J_{0} = 0.957V_{0} - 1.079, \end{equation} \begin{equation} H_{0} = 0.931V_{0} - 1.240, \end{equation} \begin{equation} (K_{s})_{0} = 0.927V_{0} - 1.292. \end{equation} 14 stars below the lines are the giants in Fig. 1, whereas 122 stars above the lines are the dwarfs of the same cluster. The distribution of different star categories at different sides of the lines in three figures were presented here to confirm separation of dwarfs and giants. \subsection{ROTSE data} The Robotic Optical Transient Experiment (ROTSE-III) consists of four 0.45m worldwide robotic, automated telescopes situated at different locations on Earth. They are designed for fast ($\sim$ 6 sec) responses to Gamma-Ray Burst (GRB) triggers from satellites such as Swift. Each ROTSE telescope has a $1.85 \times 1.85$ deg$^{2}$ field of view, and uses a Marconi 2048 $\times$ 2048 back illuminated thinned CCD. These telescopes operate without filters, and have wide passband which peaks around 550 nm (Akerlof et al. 2003). In this work, we present optical observations of NGC 752 performed by ROTSE-IIId, telescope located at Turkish National Observatory (TUG) site, Bak$\i$rl$\i$tepe, Turkey. The observations took place between MJD 53637 (September 2005) and MJD 53649 (October 2005). A total of about 217 CCD frames were analyzed. After determining the instrumental magnitudes (Bertin \& Arnouts, 1996), they were reduced to ROTSE magnitudes via comparing all the field stars with the USNO A2.0 $R$-band catalog. All the processes were done in an automated mode. \begin{figure} \center \resizebox{6.51cm}{6.22cm}{\includegraphics*{fig03.eps}} \caption {$R_{0} \times V_{0}$ magnitudes of 136 stars in NGC 752.} \end {figure} \begin{figure*} \center \resizebox{17cm}{5.51cm}{\includegraphics*{fig04.eps}} \caption {$R_{0}$ and 2MASS magnitudes of 136 stars in NGC 752. (a) $J_{0} \times R_{0}$, (b) $H_{0} \times R_{0}$, and (c) $Ks_{0} \times R_{0}$. The solid lines in diagrams were drawn according to eqs. (5), (6), and (7).} \end {figure*} $R$-band magnitudes of the 136 stars are given in Table 1. ROTSE magnitudes are also de-reddened in order to homogenize the data. The total absorption for the $R$ band could be determined by the equation of Fiorucci \& Munari (2003), i.e. $A(R)=2.613E(B-V)$. Thus the de-reddened magnitude in $R$ becomes $R_{0}=R-A(R)$. \section{Application of the method to ROTSE-IIId data} We used the following relation between the $V_{0}$ magnitude and the ROTSE-IIId magnitude $R_{0}$ for 136 stars of the open cluster NGC 752 (Fig. 3) in order to apply the method to ROTSE-IIId data: \begin{equation} V_{0}=(1.019\pm0.006)R_{0}+0.092\pm0.070~~(\sigma=0.10). \end{equation} Thus, substituting the value of $V_{0}$ in (4) into (1), (2), and (3) we obtain the relations between 2MASS and ROTSE-IIId magnitudes, i.e. \begin{equation} J_{0} = 0.975R_{0} - 0.991, \end{equation} \begin{equation} H_{0} = 0.949R_{0} - 1.154, \end{equation} \begin{equation} (K_{s})_{0}= 0.945R_{0} - 1.207. \end{equation} The diagrams $J_{0} \times R_{0}$, $H_{0} \times R_{0}$ and $(K_{s})_{0} \times R_{0}$ for the NGC 752 cluster stars and the line corresponding to the eqs. (5), (6), and (7) are given in Fig. 4. One can see that dwarfs and giants lie at opposite sides of the line in these figures, especially the relation between $(K_{s})_{0} \times R_{0}$ (Fig. 4c) is the most successful in separating of the two different star categories. \section{Age estimation for the open cluster NGC 752 via two photometries} We estimated the age of the open cluster NGC 752 by the means of BV and 2MASS data to show the advantage. The absolute magnitudes of the stars were determined by the corresponding apparent magnitudes and the distance module of the cluster. The Padova isochrones were taken from Girardi et al. (2002) and Bonatto, Bica \& Girardi (2004) for BV\footnote {http://pleiadi.pd.astro.it/isoc\_photsys.02/isoc\_photsys.02.html} and 2MASS\footnote{ http://pleiadi.pd.astro.it/isoc\_photsys.01/isoc\_2mass/index.html} photometry, respectively. The isochrone sets were computed with updated opacities, and equations of state, and a moderate amount of convective overshoot. The basic isochrone set presented in Girardi et al. (2002) covers a very wide range of initial masses (from 0.15 to $\sim 100 M_{\odot}$), metallicities, and photometric systems, being well suited for studies of clusters of all ages. The isochrones mentioned above were fitted to the CMDs in Figs. 5-7 for two sets of chemical compositions, i.e. Z=0.019, Y=0.273 (panel a) and Z=0.008, Y=0.250 (panel b). The isochrones in Fig. 5a fits to the main-sequence and turn-off segments of the $M_{V} \times (B-V)_{0}$ diagram and reveal an age of $t=1.26$ Gyr, whereas in Fig. 5b, the isochrones could be fitted only to the giant branch of the same CMD, resulting a larger age, i.e. $t=1.78$ Gyr. The isochrones could not be fitted to all segments of the $M_{J} \times (J-H)_{0}$ diagram in Fig. 6 either. In Fig. 6a the fit is better to the main-sequence and turn-off segments, however it is only to the giant branch in Fig. 6b. The best fit is accomplished with the isochrone of age $t=1.41$ Gyr to the $M_{J} \times (J-K_{s})_{0}$ in Fig. 7a. In fact, the isochrone fits to all segments, main-sequence, turn-off and giant branch, for a metallicity close to the solar one which is expected (Daniel et al. 1994). Thus, the comparison of the six diagrams in Figs. 5-7 reveals that the CMD $M_{J} \times (J-K_{s})_{0}$ is the best one which fits for the age estimation. \begin{figure} \center \resizebox{8cm}{6.92cm}{\includegraphics*{fig05.eps}} \caption {$M_{V} \times (B-V)_{0}$ CMDs of NGC 752. (a) Z=0.019 and 1.12, 1.26 and 1.41 Gyr and (b) Z=0.008 and 1.58, 1.78 and 2.00 Gyr of Padova isochrones.} \end {figure} \begin{figure} \center \resizebox{8cm}{6.92cm}{\includegraphics*{fig06.eps}} \caption {$M_{J} \times (J-H)_{0}$ CMDs of NGC 752. (a) Z=0.019 and 1.00, 1.12 and 1.26 Gyr and (b) Z=0.008 and 1.41, 1.59 and 1.78 Gyr of Padova isochrones.} \end {figure} \begin{figure} \center \resizebox{8cm}{6.92cm}{\includegraphics*{fig07.eps}} \caption {$M_{J} \times (J-Ks)_{0}$ CMDs of NGC 752. (a) Z=0.019 and 1.26, 1.41 and 1.59 Gyr and (b) Z=0.008 and 1.59, 1.78 and 1.99 Gyr of Padova isochrones.} \end {figure} \section{Conclusion} In this study, we have reduced the relations between the 2MASS and $V$ magnitudes by which field dwarfs and giants can be separated, to the USNO A2.0 $R$-band magnitudes. The $R_{0}$ magnitudes of 136 stars in open cluster NGC 752 were transferred to the $V_{0}$ magnitudes, and the relations between $J_{0}$, $H_{0}$, $(K_{s})_{0}$ and $R_{0}$ were derived by means of the relations between $J_{0}$, $H_{0}$, $(K_{s})_{0}$ and $V_{0}$ given by Bilir et al. (2006). Dwarf and giant stars identified by the CMD of open cluster NGC 752 lie at different sides of the line representing the relation between the 2MASS and $R_{0}$ magnitudes, in the $J_{0} \times R_{0}$, $H_{0} \times R_{0}$ and $(K_{s})_{0} \times R_{0}$ diagrams. The best one is the last diagram, i.e. $(K_{s})_{0} \times R_{0}$. Thus, dwarf-giant separation could be carried out also in the ROTSE-IIId data. Proven to be successful, this practical method can provide good contributions to the studies of Galactic model parameters in which separation of dwarfs and giants were needed. A set of Padova isochrones were fitted to the CMDs of the open cluster NGC 752 using BV and 2MASS photometric data. It turned out that the isochrone with chemical composition Z=0.019 and Y=0.273 which reveals an age of 1.41 Gyr for the open cluster NGC 752 could be fitted to all segments, i.e. main-sequence, turn-off and giant branch, of the $M_{J} \times (J-K_{s})_{0}$ two-colour diagram. This result is very close to the age 1.24$\pm$0.20 Gyr which Salaris, Weiss \& Percival (2004) calculated from the morphology of 71 open clusters in our Galaxy. The metal-abundance of the cluster given by Daniel et al. (1994), i.e. $[Fe/H]=-0.15\pm0.05$ dex, is a strong confirmation for our result. \begin{acknowledgements} We thank international ROTSE collaboration and TUG for the optical facilities (Project number: TUG-ROTSE.05.14). This research has made use of the SIMBAD database, operated at CDS, Strasbourg, France. This publication makes use of data products from the Two Micron All Sky Survey, which is a joint project of the University of Massachusetts and the Infrared Processing and Analysis Center/California Institute of Technology, funded by the National Aeronautics and Space Administration and the National Science Foundation. We would also like to thank Dr. Salih Karaali for helpful comments and suggestions, Dr. Zeki Eker for reading the whole manuscript and correction and Dr. Tansel Ak for various helps. This work was supported by the Research Fund of the University of Istanbul. Project number: BYP 914. \end{acknowledgements} \begin{table*} {\scriptsize \center \caption{BV, ROTSE and 2MASS magnitudes and its errors of 136 probable member stars of open cluster NGC 752. ID name is the same as Daniel et al. (1994) and the coordinates are for the epoch 2000.} \begin{tabular}{rcccccccccccccc} \hline ID & $\alpha$ & $\delta$ & \multicolumn{2} {c} {V} & \multicolumn{2} {c} {B-V}& \multicolumn{2} {c} {R} & \multicolumn{2} {c} {J}& \multicolumn{2} {c} {H}& \multicolumn{2} {c} {$K_{s}$}\\ \hline 143 & 01 54 31.04 & +37 29 31.6 & 11.140 & & 0.470 & & 10.970 & 0.010 & 10.343 & 0.019 & 10.184 & 0.024 & 10.151 & 0.020 \\ 215 & 01 54 56.71 & +37 58 28.9 & 11.400 & & 0.430 & & 11.200 & 0.012 & 10.615 & 0.021 & 10.416 & 0.017 & 10.378 & 0.018 \\ 222 & 01 54 59.65 & +37 28 59.8 & 11.651 & 0.009 & 0.489 & 0.007 & 11.365 & 0.011 & 10.607 & 0.021 & 10.377 & 0.018 & 10.293 & 0.021 \\ 237 & 01 55 02.85 & +38 16 38.1 & 12.418 & 0.000 & 0.519 & 0.000 & 12.110 & 0.017 & 11.364 & 0.021 & 11.105 & 0.017 & 11.094 & 0.020 \\ 245 & 01 55 07.13 & +37 32 36.9 & 14.190 & & 0.660 & & 13.830 & 0.026 & 12.838 & 0.022 & 12.464 & 0.021 & 12.397 & 0.021 \\ 259 & 01 55 12.62 & +37 50 14.4 & 9.496 & 0.013 & 0.954 & 0.006 & 9.092 & 0.016 & 7.818 & 0.023 & 7.370 & 0.023 & 7.228 & 0.016 \\ 264 & 01 55 15.29 & +37 50 31.2 & 9.569 & 0.005 & 0.996 & 0.003 & 9.090 & 0.017 & 7.795 & 0.023 & 7.317 & 0.016 & 7.197 & 0.018 \\ 300 & 01 55 26.18 & +38 08 22.0 & 13.610 & & 0.700 & & 13.200 & 0.019 & 12.182 & 0.018 & 11.798 & 0.021 & 11.743 & 0.021 \\ 305 & 01 55 27.68 & +37 34 04.5 & 10.152 & 0.003 & 0.411 & 0.001 & 9.940 & 0.011 & 9.293 & 0.018 & 9.103 & 0.015 & 9.062 & 0.016 \\ 308 & 01 55 27.67 & +37 59 55.2 & 9.285 & 0.011 & 0.966 & 0.009 & 8.902 & 0.020 & 7.618 & 0.021 & 7.163 & 0.021 & 7.039 & 0.020 \\ 313 & 01 55 29.29 & +37 50 26.2 & 9.968 & 0.013 & 0.470 & 0.001 & 9.718 & 0.014 & 9.001 & 0.023 & 8.792 & 0.018 & 8.696 & 0.017 \\ 350 & 01 55 39.37 & +37 52 52.4 & 8.922 & 0.002 & 1.011 & 0.006 & 8.439 & 0.024 & 7.151 & 0.020 & 6.674 & 0.016 & 6.547 & 0.016 \\ 356 & 01 55 42.39 & +37 37 54.3 & 9.161 & 0.003 & 1.010 & 0.001 & 8.687 & 0.020 & 7.395 & 0.019 & 6.900 & 0.020 & 6.800 & 0.018 \\ 361 & 01 55 44.75 & +37 54 42.5 & 13.899 & 0.000 & 0.724 & 0.000 & 13.497 & 0.025 & 12.438 & 0.023 & 12.054 & 0.023 & 11.988 & 0.019 \\ 363 & 01 55 44.93 & +38 08 21.4 & 10.420 & & 0.360 & & 10.189 & 0.013 & 9.619 & 0.021 & 9.464 & 0.021 & 9.410 & 0.020 \\ 372 & 01 55 47.40 & +37 42 26.4 & 9.890 & 0.005 & 0.464 & 0.000 & 9.656 & 0.011 & 8.971 & 0.029 & 8.741 & 0.027 & 8.717 & 0.018 \\ 391 & 01 55 53.54 & +37 49 26.6 & 13.890 & & 0.730 & & 13.501 & 0.024 & 12.457 & 0.023 & 12.072 & 0.021 & 11.985 & 0.023 \\ 397 & 01 55 55.53 & +37 28 33.2 & 9.831 & 0.000 & 0.477 & 0.005 & 9.585 & 0.011 & 8.839 & 0.021 & 8.636 & 0.020 & 8.576 & 0.016 \\ 413 & 01 55 59.44 & +37 40 48.5 & 12.303 & 0.016 & 0.517 & 0.022 & 12.035 & 0.013 & 11.181 & 0.021 & 10.924 & 0.017 & 10.842 & 0.016 \\ 429 & 01 56 02.95 & +37 36 32.7 & 14.270 & & 0.850 & & 13.791 & 0.033 & 12.741 & 0.023 & 12.338 & 0.022 & 12.225 & 0.021 \\ 435 & 01 56 03.69 & +37 59 22.4 & 11.467 & 0.017 & 0.447 & 0.010 & 11.223 & 0.013 & 10.572 & 0.020 & 10.384 & 0.018 & 10.324 & 0.018 \\ 455 & 01 56 08.96 & +37 39 52.6 & 10.512 & 0.013 & 0.395 & 0.003 & 10.298 & 0.020 & 9.657 & 0.021 & 9.496 & 0.017 & 9.442 & 0.014 \\ 461 & 01 56 10.30 & +37 44 59.9 & 10.054 & 0.002 & 0.384 & 0.006 & 9.752 & 0.041 & 9.239 & 0.021 & 9.115 & 0.018 & 9.015 & 0.019 \\ 465 & 01 56 11.11 & +37 45 11.1 & 11.229 & 0.012 & 0.412 & 0.007 & 10.896 & 0.045 & 10.419 & 0.021 & 10.235 & 0.019 & 10.194 & 0.018 \\ 472 & 01 56 12.88 & +38 01 43.2 & 11.060 & & 0.460 & & 10.813 & 0.140 & 10.127 & 0.020 & 9.897 & 0.018 & 9.866 & 0.020 \\ 475 & 01 56 13.70 & +37 15 56.9 & 12.847 & 0.000 & 0.624 & 0.000 & 12.496 & 0.016 & 11.778 & 0.022 & 11.514 & 0.026 & 11.435 & 0.022 \\ 477 & 01 56 13.96 & +37 47 04.7 & 10.572 & 0.006 & 0.364 & 0.006 & 10.375 & 0.015 & 9.778 & 0.020 & 9.627 & 0.019 & 9.589 & 0.021 \\ 479 & 01 56 14.28 & +37 58 14.2 & 10.938 & 0.011 & 0.442 & 0.005 & 10.696 & 0.014 & 10.013 & 0.021 & 9.828 & 0.018 & 9.777 & 0.017 \\ 486 & 01 56 15.51 & +37 38 41.5 & 10.074 & 0.000 & 0.493 & 0.006 & 9.824 & 0.012 & 9.115 & 0.026 & 8.918 & 0.017 & 8.853 & 0.016 \\ 505 & 01 56 18.63 & +37 37 39.3 & 10.778 & 0.005 & 0.399 & 0.008 & 10.543 & 0.014 & 9.903 & 0.021 & 9.719 & 0.017 & 9.669 & 0.016 \\ 506 & 01 56 18.90 & +37 58 00.4 & 8.971 & 0.018 & 1.003 & 0.011 & 8.456 & 0.044 & 7.174 & 0.018 & 6.723 & 0.024 & 6.606 & 0.023 \\ 512 & 01 56 21.65 & +37 36 08.2 & 9.375 & 0.022 & 1.029 & 0.004 & 8.896 & 0.018 & 7.558 & 0.020 & 7.045 & 0.020 & 6.915 & 0.018 \\ 517 & 01 56 22.57 & +37 39 17.8 & 14.230 & & 0.830 & & 13.833 & 0.030 & 12.681 & 0.021 & 12.265 & 0.024 & 12.163 & 0.022 \\ 520 & 01 56 23.10 & +37 38 03.0 & 12.850 & & 0.570 & & 12.530 & 0.016 & 11.684 & 0.021 & 11.390 & 0.023 & 11.342 & 0.022 \\ 542 & 01 56 29.45 & +37 55 14.7 & 14.350 & & 0.490 & & 14.045 & 0.032 & 13.214 & 0.024 & 12.878 & 0.025 & 12.828 & 0.026 \\ 552 & 01 56 32.05 & +37 34 22.2 & 12.921 & 0.016 & 0.581 & 0.036 & 12.425 & 0.025 & 11.696 & 0.021 & 11.417 & 0.023 & 11.351 & 0.020 \\ 555 & 01 56 32.96 & +37 56 46.4 & 11.786 & 0.001 & 0.482 & 0.006 & 11.514 & 0.014 & 10.779 & 0.020 & 10.519 & 0.015 & 10.489 & 0.018 \\ 563 & 01 56 34.46 & +38 08 49.4 & 13.680 & & 0.910 & & 13.216 & 0.020 & 12.020 & 0.019 & 11.565 & 0.017 & 11.483 & 0.022 \\ 575 & 01 56 36.88 & +37 45 12.7 & 13.840 & & 0.760 & & 13.477 & 0.029 & 12.384 & 0.020 & 12.030 & 0.019 & 11.944 & 0.018 \\ 580 & 01 56 39.22 & +37 51 41.1 & 10.398 & 0.007 & 0.373 & 0.005 & 10.194 & 0.015 & 9.630 & 0.020 & 9.458 & 0.017 & 9.416 & 0.018 \\ 619 & 01 56 47.60 & +37 24 30.4 & 10.280 & 0.022 & 0.415 & 0.005 & 10.054 & 0.013 & 9.412 & 0.020 & 9.226 & 0.018 & 9.173 & 0.019 \\ 622 & 01 56 48.61 & +37 29 11.2 & 10.503 & 0.001 & 0.391 & 0.007 & 10.321 & 0.013 & 9.714 & 0.020 & 9.584 & 0.021 & 9.533 & 0.020 \\ 626 & 01 56 49.77 & +38 01 21.7 & 9.158 & 0.016 & 0.439 & 0.006 & 8.902 & 0.032 & 8.202 & 0.026 & 8.105 & 0.057 & 8.005 & 0.024 \\ 630 & 01 56 50.44 & +38 01 58.1 & 8.961 & 0.011 & 0.811 & 0.020 & 8.661 & 0.059 & 7.379 & 0.021 & 6.936 & 0.020 & 6.824 & 0.023 \\ 641 & 01 56 53.06 & +37 52 09.3 & 10.270 & 0.004 & 0.434 & 0.004 & 10.036 & 0.014 & 9.371 & 0.020 & 9.164 & 0.015 & 9.122 & 0.018 \\ 648 & 01 56 54.33 & +37 23 51.9 & 12.108 & 0.000 & 0.585 & 0.000 & 11.782 & 0.014 & 10.952 & 0.021 & 10.693 & 0.019 & 10.652 & 0.018 \\ 653 & 01 56 55.38 & +38 04 45.8 & 12.410 & & 0.540 & & 12.088 & 0.013 & 11.262 & 0.019 & 10.986 & 0.017 & 10.929 & 0.020 \\ 654 & 01 56 55.77 & +37 47 59.3 & 11.196 & 0.000 & 0.396 & 0.014 & 11.022 & 0.015 & 10.385 & 0.020 & 10.216 & 0.017 & 10.155 & 0.018 \\ 655 & 01 56 56.15 & +38 08 16.2 & 13.040 & & 0.720 & & 12.665 & 0.017 & 11.706 & 0.021 & 11.370 & 0.019 & 11.294 & 0.018 \\ 659 & 01 56 56.35 & +37 39 51.3 & 10.116 & 0.003 & 0.424 & 0.001 & 9.898 & 0.014 & 9.220 & 0.018 & 9.039 & 0.017 & 8.959 & 0.018 \\ 667 & 01 56 57.59 & +37 23 20.5 & 10.925 & 0.021 & 0.377 & 0.004 & 10.720 & 0.170 & 10.131 & 0.021 & 9.990 & 0.021 & 9.925 & 0.018 \\ 682 & 01 57 02.51 & +37 53 07.7 & 11.255 & 0.012 & 0.447 & 0.002 & 11.007 & 0.014 & 10.332 & 0.019 & 10.205 & 0.023 & 10.178 & 0.024 \\ 684 & 01 57 02.81 & +38 14 03.5 & 12.480 & & 0.500 & & 12.153 & 0.015 & 11.415 & 0.021 & 11.180 & 0.019 & 11.123 & 0.020 \\ 687 & 01 57 03.12 & +38 08 02.6 & 8.927 & 0.002 & 1.026 & 0.004 & 8.424 & 0.024 & 7.136 & 0.024 & 6.663 & 0.015 & 6.537 & 0.017 \\ 689 & 01 57 03.19 & +37 55 44.5 & 11.788 & 0.005 & 0.455 & 0.009 & 11.529 & 0.015 & 10.838 & 0.019 & 10.641 & 0.019 & 10.582 & 0.020 \\ 694 & 01 57 03.64 & +38 05 11.7 & 11.779 & 0.011 & 0.448 & 0.007 & 11.520 & 0.013 & 10.720 & 0.021 & 10.478 & 0.019 & 10.439 & 0.019 \\ 699 & 01 57 04.89 & +38 07 33.1 & 13.001 & 0.024 & 0.627 & 0.048 & 12.647 & 0.016 & 11.652 & 0.021 & 11.322 & 0.018 & 11.222 & 0.018 \\ 701 & 01 57 05.47 & +37 50 42.8 & 13.060 & & 0.690 & & 12.700 & 0.017 & 11.684 & 0.019 & 11.339 & 0.021 & 11.241 & 0.020 \\ 720 & 01 57 10.50 & +38 02 06.6 & 12.367 & 0.020 & 0.494 & 0.034 & 12.064 & 0.013 & 11.322 & 0.020 & 11.066 & 0.019 & 11.019 & 0.018 \\ 722 & 01 57 10.35 & +37 25 55.3 & 13.500 & & 0.910 & & 13.221 & 0.021 & 12.178 & 0.019 & 11.813 & 0.019 & 11.724 & 0.018 \\ 723 & 01 57 10.53 & +37 27 26.6 & 13.770 & & 0.870 & & 13.339 & 0.033 & 12.030 & 0.019 & 11.557 & 0.021 & 11.489 & 0.020 \\ 728 & 01 57 12.13 & +37 59 24.8 & 9.420 & 0.005 & 0.471 & 0.004 & 9.169 & 0.011 & 8.485 & 0.020 & 8.296 & 0.016 & 8.246 & 0.024 \\ 731 & 01 57 12.17 & +37 56 04.7 & 11.956 & 0.000 & 0.595 & 0.000 & 11.608 & 0.013 & 10.722 & 0.019 & 10.426 & 0.019 & 10.335 & 0.020 \\ 745 & 01 57 14.27 & +37 46 51.0 & 9.874 & 0.011 & 0.400 & 0.007 & 9.659 & 0.016 & 9.032 & 0.021 & 8.878 & 9.995 & 8.790 & 0.018 \\ 756 & 01 57 17.11 & +37 26 08.8 & 10.207 & 0.012 & 0.428 & 0.004 & 9.971 & 0.013 & 9.302 & 0.018 & 9.122 & 0.017 & 9.047 & 0.020 \\ 768 & 01 57 19.42 & +37 59 23.5 & 12.072 & 0.011 & 0.490 & 0.000 & 11.811 & 0.014 & 11.080 & 0.020 & 10.874 & 0.018 & 10.830 & 0.019 \\ 772 & 01 57 20.73 & +37 51 43.1 & 10.188 & 0.009 & 0.476 & 0.003 & 9.923 & 0.012 & 9.213 & 0.019 & 9.022 & 0.019 & 8.937 & 0.018 \\ 783 & 01 57 22.29 & +37 36 23.2 & 12.240 & & 0.610 & & 11.946 & 0.015 & 11.054 & 0.018 & 10.771 & 0.019 & 10.717 & 0.024 \\ 786 & 01 57 22.97 & +37 38 21.8 & 13.170 & & 0.730 & & 12.746 & 0.016 & 11.703 & 0.018 & 11.327 & 0.017 & 11.261 & 0.020 \\ 790 & 01 57 23.81 & +37 52 11.9 & 12.267 & 0.012 & 0.529 & 0.018 & 11.938 & 0.015 & 11.117 & 0.018 & 10.860 & 0.019 & 10.756 & 0.018 \\ \end{tabular} } \end{table*} \begin{table*} {\scriptsize \begin{tabular}{rcccccccccccccc} \hline ID & $\alpha$ & $\delta$ & \multicolumn{2} {c} {V} & \multicolumn{2} {c} {B-V}& \multicolumn{2} {c} {R} & \multicolumn{2} {c} {J}& \multicolumn{2} {c} {H}& \multicolumn{2} {c} {$K_{s}$}\\ \hline 791 & 01 57 24.01 & +38 06 10.4 & 12.684 & 0.017 & 0.543 & 0.017 & 12.370 & 0.015 & 11.566 & 0.020 & 11.306 & 0.019 & 11.261 & 0.021 \\ 798 & 01 57 26.01 & +37 43 19.7 & 10.454 & 0.005 & 0.419 & 0.004 & 10.226 & 0.013 & 9.565 & 0.018 & 9.401 & 0.019 & 9.322 & 0.020 \\ 799 & 01 57 26.17 & +37 39 20.3 & 11.304 & 0.004 & 0.423 & 0.004 & 11.073 & 0.013 & 10.419 & 0.018 & 10.225 & 0.015 & 10.184 & 0.020 \\ 806 & 01 57 27.47 & +37 35 10.4 & 10.756 & 0.000 & 0.385 & 0.000 & 10.555 & 0.015 & 9.948 & 0.018 & 9.817 & 0.017 & 9.748 & 0.018 \\ 814 & 01 57 28.26 & +37 24 02.6 & 10.219 & 0.011 & 0.367 & 0.007 & 10.015 & 0.014 & 9.427 & 0.021 & 9.287 & 0.019 & 9.209 & 0.020 \\ 823 & 01 57 30.93 & +37 54 57.9 & 10.273 & 0.003 & 0.417 & 0.002 & 10.037 & 0.014 & 9.379 & 0.018 & 9.225 & 0.019 & 9.165 & 0.020 \\ 824 & 01 57 31.86 & +37 53 40.6 & 11.629 & 0.011 & 0.440 & 0.005 & 11.375 & 0.015 & 10.720 & 0.019 & 10.533 & 0.019 & 10.452 & 0.018 \\ 828 & 01 57 32.58 & +37 42 05.8 & 13.943 & 0.000 & 0.681 & 0.000 & 13.537 & 0.024 & 12.452 & 0.031 & 12.095 & 0.035 & 12.051 & 0.020 \\ 847 & 01 57 35.91 & +37 58 23.1 & 14.200 & & 1.030 & & 13.704 & 0.025 & 12.385 & 0.021 & 11.891 & 0.023 & 11.769 & 0.018 \\ 849 & 01 57 36.23 & +37 45 10.0 & 9.917 & 0.010 & 0.424 & 0.005 & 9.682 & 0.013 & 9.049 & 0.024 & 8.860 & 0.023 & 8.773 & 0.019 \\ 857 & 01 57 37.69 & +37 49 00.7 & 10.028 & 0.011 & 0.470 & 0.002 & 9.780 & 0.013 & 9.072 & 0.021 & 8.841 & 0.020 & 8.774 & 0.019 \\ 858 & 01 57 37.62 & +37 39 37.8 & 8.958 & 0.003 & 1.085 & 0.004 & 8.416 & 0.030 & 7.070 & 0.021 & 6.555 & 0.018 & 6.412 & 0.018 \\ 859 & 01 57 37.77 & +37 49 50.4 & 13.200 & & 0.670 & & 12.842 & 0.016 & 11.735 & 0.021 & 11.364 & 0.023 & 11.244 & 0.018 \\ 864 & 01 57 38.78 & +38 08 30.3 & 12.885 & 0.013 & 0.578 & 0.017 & 12.553 & 0.017 & 11.711 & 0.019 & 11.424 & 0.019 & 11.374 & 0.016 \\ 867 & 01 57 38.97 & +37 46 12.2 & 9.041 & 0.007 & 1.004 & 0.006 & 8.556 & 0.019 & 7.263 & 0.020 & 6.813 & 0.024 & 6.669 & 0.031 \\ 868 & 01 57 39.46 & +37 52 25.8 & 10.465 & 0.014 & 0.381 & 0.005 & 10.262 & 0.014 & 9.676 & 0.021 & 9.485 & 0.021 & 9.418 & 0.017 \\ 888 & 01 57 43.97 & +37 51 42.1 & 10.447 & 0.007 & 0.427 & 0.003 & 10.208 & 0.012 & 9.557 & 0.021 & 9.371 & 0.021 & 9.279 & 0.017 \\ 889 & 01 57 44.46 & +38 11 06.8 & 12.802 & 0.015 & 0.551 & 0.018 & 12.475 & 0.014 & 11.654 & 0.018 & 11.395 & 0.019 & 11.323 & 0.016 \\ 890 & 01 57 44.74 & +37 59 18.4 & 10.087 & 0.011 & 0.453 & 0.006 & 9.837 & 0.014 & 9.172 & 0.021 & 8.985 & 0.021 & 8.901 & 0.018 \\ 897 & 01 57 46.07 & +38 04 28.4 & 10.506 & 0.037 & 0.407 & 0.003 & 10.382 & 0.170 & 9.653 & 0.018 & 9.481 & 0.019 & 9.407 & 0.014 \\ 901 & 01 57 47.15 & +37 47 30.3 & 10.981 & 0.006 & 0.388 & 0.007 & 10.763 & 0.014 & 10.126 & 0.021 & 9.985 & 0.019 & 9.897 & 0.018 \\ 917 & 01 57 51.42 & +37 39 52.4 & 14.310 & & 0.520 & & 13.590 & 0.026 & 12.596 & 0.019 & 12.203 & 0.019 & 12.089 & 0.020 \\ 921 & 01 57 52.00 & +37 27 46.0 & 12.644 & 0.011 & 0.553 & 0.007 & 12.349 & 0.013 & 11.507 & 0.018 & 11.275 & 0.017 & 11.226 & 0.021 \\ 935 & 01 57 55.20 & +37 52 46.0 & 11.620 & & 0.450 & & 11.417 & 0.014 & 10.750 & 0.021 & 10.553 & 0.021 & 10.520 & 0.018 \\ 937 & 01 57 54.97 & +37 20 26.6 & 10.980 & & 0.380 & & 10.762 & 0.014 & 10.161 & 0.018 & 10.015 & 0.017 & 9.971 & 0.019 \\ 941 & 01 57 56.45 & +37 50 01.0 & 10.706 & 0.006 & 0.424 & 0.001 & 10.462 & 0.014 & 9.807 & 0.021 & 9.640 & 0.023 & 9.556 & 0.019 \\ 950 & 01 57 57.79 & +37 48 22.3 & 11.467 & 0.020 & 0.480 & 0.007 & 11.160 & 0.077 & 10.393 & 0.020 & 10.225 & 0.021 & 10.163 & 0.019 \\ 952 & 01 57 58.24 & +37 26 06.4 & 12.650 & & 0.540 & & 12.356 & 0.014 & 11.556 & 0.019 & 11.282 & 0.019 & 11.268 & 0.019 \\ 953 & 01 57 58.85 & +37 41 26.8 & 12.352 & 0.038 & 0.600 & 0.008 & 12.012 & 0.014 & 11.144 & 0.018 & 10.870 & 0.019 & 10.797 & 0.020 \\ 955 & 01 57 59.37 & +37 54 53.8 & 9.968 & 0.010 & 0.445 & 0.001 & 9.730 & 0.013 & 9.084 & 0.021 & 8.885 & 0.023 & 8.804 & 0.018 \\ 964 & 01 58 02.79 & +38 02 30.4 & 12.912 & 0.016 & 0.582 & 0.014 & 12.568 & 0.015 & 11.668 & 0.019 & 11.385 & 0.019 & 11.273 & 0.018 \\ 983 & 01 58 06.31 & +37 38 06.6 & 13.110 & & 0.610 & & 12.709 & 0.017 & 11.756 & 0.019 & 11.401 & 0.019 & 11.314 & 0.020 \\ 988 & 01 58 07.71 & +37 39 57.0 & 10.934 & 0.008 & 0.373 & 0.004 & 10.714 & 0.016 & 10.144 & 0.019 & 9.987 & 0.019 & 9.953 & 0.017 \\ 993 & 01 58 09.26 & +37 28 35.5 & 13.590 & 0.000 & 0.708 & 0.000 & 13.226 & 0.018 & 12.271 & 0.021 & 11.923 & 0.021 & 11.855 & 0.021 \\ 999 & 01 58 10.66 & +37 24 05.9 & 13.550 & & 0.670 & & 13.189 & 0.019 & 12.236 & 0.021 & 11.899 & 0.023 & 11.819 & 0.019 \\ 1000 & 01 58 11.43 & +37 39 33.4 & 11.410 & 0.012 & 0.419 & 0.006 & 11.175 & 0.014 & 10.568 & 0.021 & 10.379 & 0.019 & 10.332 & 0.018 \\ 1003 & 01 58 12.27 & +37 32 38.2 & 11.193 & 0.003 & 0.472 & 0.003 & 10.933 & 0.014 & 10.251 & 0.021 & 10.013 & 0.018 & 9.942 & 0.017 \\ 1007 & 01 58 13.40 & +38 11 41.5 & 13.018 & 0.025 & 0.556 & 0.042 & 12.641 & 0.016 & 11.742 & 0.020 & 11.467 & 0.017 & 11.375 & 0.016 \\ 1008 & 01 58 12.71 & +37 34 40.4 & 10.959 & 0.005 & 0.371 & 0.004 & 10.743 & 0.015 & 10.188 & 0.021 & 10.009 & 0.018 & 9.964 & 0.019 \\ 1012 & 01 58 12.94 & +37 15 20.2 & 12.417 & 0.015 & 0.539 & 0.000 & 12.113 & 0.016 & 11.336 & 0.020 & 11.083 & 0.023 & 11.047 & 0.019 \\ 1017 & 01 58 15.33 & +37 33 19.6 & 13.260 & & 0.660 & & 12.920 & 0.017 & 12.014 & 0.021 & 11.673 & 0.019 & 11.602 & 0.019 \\ 1023 & 01 58 16.88 & +37 38 15.9 & 11.250 & 0.015 & 0.424 & 0.010 & 11.005 & 0.013 & 10.407 & 0.021 & 10.203 & 0.019 & 10.159 & 0.019 \\ 1026 & 01 58 19.00 & +38 32 14.0 & 11.089 & 0.016 & 0.377 & 0.017 & 10.851 & 0.010 & 10.303 & 0.020 & 10.134 & 0.023 & 10.079 & 0.017 \\ 1027 & 01 58 18.42 & +38 06 54.0 & 12.597 & 0.020 & 0.541 & 0.036 & 12.282 & 0.022 & 11.497 & 0.020 & 11.237 & 0.017 & 11.184 & 0.018 \\ 1083 & 01 58 27.61 & +37 35 22.2 & 11.920 & 0.011 & 0.484 & 0.007 & 11.676 & 0.014 & 10.977 & 0.021 & 10.721 & 0.019 & 10.656 & 0.018 \\ 1089 & 01 58 29.84 & +37 51 37.4 & 9.303 & 0.007 & 0.963 & 0.008 & 8.836 & 0.021 & 7.621 & 0.026 & 7.166 & 0.026 & 7.040 & 0.021 \\ 1107 & 01 58 34.42 & +37 40 15.1 & 13.660 & & 0.660 & & 13.273 & 0.020 & 12.299 & 0.022 & 11.923 & 0.019 & 11.884 & 0.021 \\ 1117 & 01 58 36.91 & +37 45 10.6 & 9.598 & 0.002 & 0.406 & 0.002 & 9.370 & 0.012 & 8.747 & 0.027 & 8.559 & 0.026 & 8.502 & 0.016 \\ 1123 & 01 58 38.12 & +37 32 15.7 & 11.507 & 0.000 & 0.417 & 0.007 & 11.278 & 0.014 & 10.647 & 0.020 & 10.467 & 0.019 & 10.395 & 0.019 \\ 1129 & 01 58 40.07 & +37 38 05.1 & 11.910 & 0.011 & 0.481 & 0.007 & 11.629 & 0.014 & 10.911 & 0.021 & 10.691 & 0.019 & 10.599 & 0.018 \\ 1151 & 01 58 47.96 & +38 26 08.2 & 10.063 & 0.013 & 0.444 & 0.014 & 9.788 & 0.014 & 9.158 & 0.021 & 8.947 & 0.018 & 8.912 & 0.019 \\ 1161 & 01 58 50.00 & +37 59 46.6 & 14.600 & & 0.680 & & 14.060 & 0.040 & 12.715 & 0.019 & 12.215 & 0.019 & 12.108 & 0.020 \\ 1165 & 01 58 50.44 & +37 20 52.0 & 10.462 & 0.000 & 0.390 & 0.000 & 10.239 & 0.013 & 9.643 & 0.020 & 9.485 & 0.023 & 9.436 & 0.020 \\ 1172 & 01 58 52.93 & +37 48 57.0 & 9.060 & 0.003 & 1.036 & 0.006 & 8.529 & 0.036 & 7.270 & 0.019 & 6.795 & 0.023 & 6.643 & 0.018 \\ 1178 & 01 58 53.94 & +37 34 42.6 & 13.390 & & 0.830 & & 13.028 & 0.015 & 11.711 & 0.021 & 11.268 & 0.019 & 11.197 & 0.022 \\ 1196 & 01 58 57.32 & +37 39 40.9 & 13.810 & & 0.910 & & 13.339 & 0.019 & 12.150 & 0.021 & 11.703 & 0.017 & 11.611 & 0.022 \\ 1204 & 01 58 59.87 & +38 01 18.9 & 11.680 & & 0.460 & & 11.446 & 0.014 & 10.796 & 0.021 & 10.632 & 0.023 & 10.577 & 0.019 \\ 1263 & 01 59 14.82 & +38 00 55.2 & 9.006 & 0.011 & 1.019 & 0.007 & 8.483 & 0.025 & 7.199 & 0.019 & 6.767 & 0.018 & 6.648 & 0.016 \\ 1270 & 01 59 18.04 & +37 49 49.4 & 14.090 & & 0.700 & & 13.457 & 0.060 & 12.552 & 0.019 & 12.159 & 0.021 & 12.069 & 0.022 \\ 1284 & 01 59 19.91 & +37 23 23.1 & 12.893 & 0.000 & 0.677 & 0.000 & 12.476 & 0.013 & 11.472 & 0.021 & 11.140 & 0.021 & 11.052 & 0.020 \\ 1296 & 01 59 26.08 & +37 40 39.9 & 14.750 & & 0.530 & & 14.111 & 0.035 & 13.123 & 0.023 & 12.767 & 0.022 & 12.676 & 0.027 \\ 1304 & 01 59 29.63 & +38 16 04.3 & 11.341 & 0.008 & 0.412 & 0.011 & 11.076 & 0.014 & 10.485 & 0.022 & 10.293 & 0.022 & 10.251 & 0.020 \\ 1365 & 01 59 47.27 & +37 49 53.8 & 13.290 & & 0.710 & & 12.935 & 0.018 & 12.005 & 0.022 & 11.696 & 0.021 & 11.669 & 0.019 \\ 1407 & 01 59 56.83 & +37 58 10.5 & 12.949 & 0.000 & 0.552 & 0.000 & 12.630 & 0.016 & 11.782 & 0.027 & 11.575 & 0.034 & 11.454 & 0.023 \\ 1474 & 02 00 21.98 & +38 02 41.0 & 10.698 & 0.009 & 0.355 & 0.016 & 10.441 & 0.013 & 9.894 & 0.027 & 9.730 & 0.030 & 9.692 & 0.022 \\ 1602 & 02 01 05.97 & +37 42 23.6 & 9.961 & 0.011 & 0.462 & 0.014 & 9.671 & 0.013 & 8.978 & 0.020 & 8.786 & 0.032 & 8.733 & 0.023 \\ \hline \end{tabular} } \end{table*}
Title: Constraints on UED KK-neutrino dark matter from magnetic dipole moments
Abstract: Generically, universal extra dimension (UED) extensions of the standard model predict the stability of the lightest Kaluza-Klein (KK) particle and hence provide a dark matter candidate. For UED scenarios with one extra dimension, we model-independently determine the size of the induced dimension-five magnetic dipole moment of the KK-neutrino, $\nu^{(1)}$. We show that current observational bounds on the interactions of dipole dark matter place constraints on UED models with KK-neutrino dark matter.
https://export.arxiv.org/pdf/hep-ph/0601161
\preprint{ \hfill \begin{minipage}[t]{3in} \begin{flushright} \vspace{0.0in} OUTP-0605P \end{flushright} \end{minipage} } \title{Constraints on UED KK-neutrino dark matter from magnetic dipole moments} \author{Thomas Flacke} \email{t.flacke1@physics.ox.ac.uk} \affiliation{Rudolf Peierls Centre for Theoretical Physics, University of Oxford, 1 Keble Road, Oxford OX1 3NP, United Kingdom} \author{David W.~Maybury} \email{d.maybury1@physics.ox.ac.uk} \affiliation{Rudolf Peierls Centre for Theoretical Physics, University of Oxford, 1 Keble Road, Oxford OX1 3NP, United Kingdom} \date{January 24, 2006} \pacs{12.60.-i, 14.60.St, 95.30.Cq, 95.35.+d} \keywords{extra dimensions, dark matter, dipole moments} \section{Introduction} While the standard model proves an excellent framework for fundamental interactions at energy scales up to at least the sub TeV range, it nevertheless leaves a number of fundamental problems. Theoretically, one of the most outstanding puzzles centers on the origin and mechanism of electroweak symmetry breaking, and the quantum mechanical stability of the hierarchy generated between the electroweak scale and the Planck scale. In addition, recent astrophysical observations \cite{DarkMatter} concord with $0.094 < \Omega_{CDM}h^2 < 0.129$, indicating the presence of cold non-baryonic dark matter as the principle form of matter in the Universe, of which the standard model provides no explanation. The most popular candidate for dark matter assumes a non-standard model, stable, electrically neutral, and weakly interacting particle -- the WIMP hypothesis. Clearly, from both a theoretical and phenomenological perspective, the standard model requires extension. In this letter we wish to explore the consequences of the universal extra dimension (UED) \cite{Appelquist:2000nn} extension of the standard model. In the UED scenario, all standard model particles can freely propagate in the bulk of one or more extra dimensions and thus each standard model particle is associated with a Kaluza-Klein (KK) tower of states. Each state in the KK tower has the same spin as its standard model counterpart. An important consequence of UED models concerns the existence of a conserved discrete symmetry, KK-parity, which guarantees the stability of the lightest KK particle (LKP) and thus provides a dark matter candidate. Suitable thermal relic dark matter candidates that have been studied extensively \cite{Servant:2002aq} include the first KK-excitations of the hypercharge boson, the photon, and the neutrino \ie $B^{(1)}$, $\gamma^{(1)}$, or $\nu^{(1)}$. The tree-level mass spectrum of the KK-excitations of UED models reveals a nearly degenerate spectrum. As an example, a UED model with one extra-dimension compactified on an $S_1/Z_2$ orbifold of radius $R$, leads to the tree level mass relation \be\label{mass} m^{(n)} = \sqrt{(n/R)^2 + (m^{(0)})^2} \ee for the n-th KK mode, where $m^{(0)}$ constitutes the zero-mode mass (\ie the standard model particle value). Quantum corrections typically dominate over zero mode level contributions and therefore the resulting mass spectrum depends crucially on radiative effects. In general, a moderately split UED mass spectrum \cite{Cheng:2002ej} develops. In the minimal UED model (MUED) \cite{Cheng:2002ej}, one-loop calculations suggest that the LKP is well approximated by the KK hypercharge boson, $B^{(1)}$, and numerous studies have examined the thermal production and prospects of direct and indirect detection of $B^{(1)}$ and $\gamma^{(1)}$ LKP dark matter \cite{Servant:2002aq,KKDMcollection}. However, the non-renormalizability of UED models imply the existence of an ultraviolet cut-off, typically of the order of a few tens of TeV, at which point the model requires UV completion. As such, UED models must be regarded as an effective theory. The presence of incalculable boundary terms arising from the UV complete theory can potentially change the mass spectrum, resulting in different LKP candidates. Recent studies of the relic density of $B^{(1)}$ LKP dark matter with the full MUED spectrum \cite{coannmat,coannkribs} reveal substantial observational tension with constraints from electroweak precision data \cite{Flacke:2005hb}. Thus, non-minimal models with brane-localized terms appear as a likely alternative if UED models are to provide a successful phenomenology. Furthermore, model independent studies \cite{Servant:2002aq} show that $B^{(1)}$, $\gamma^{(1)}$, and $\nu^{(1)}$ can all be thermally produced with abundances sufficient to provide the dark matter. Constraints on minimal UED models from limits on weak neutral current nucleon-$\nu^{(1)}$ elastic scattering in direct searches, together with thermal dark matter production mechanisms, disfavour $\nu^{(1)}$ dark matter \cite{Servant:2002hb}. However, given the need for possible new non-minimal interactions, we consider further consequences of KK-neutrino dark matter model-independently. While there exists compelling evidence for dark matter in the form of WIMPS, there also exists strong constraints on possible electromagnetic interactions of dark matter, even in the limit of complete charge neutrality. A neutral Dirac fermion can posses both a permanent magnetic dipole moment, $\mu$, and a permanent electric dipole moment, $d$, arising from the dimension-five operator, \be \mathcal{L}_{D} = \frac{i}{2} \bar f \sigma_{\mu\nu}\left(\mu +\gamma_5 d \right)f F^{\mu\nu}. \ee While the presence of the magnetic dipole moment does not violate any discrete symmetries, the electric dipole moment requires the violation of parity and CP. Severe constraints exist on the dipole moments of $\sim 1$ TeV dark matter WIMPS \cite{Sigurdson:2004zp}. (We are aware that the authors of \cite{Sigurdson:2004zp} are currently revising their estimates and, as a result, the strength of the constraints in \cite{Sigurdson:2004zp} may change substantially \cite{private}.) Thus, if a model predicts Dirac fermionic dark matter, it is important to determine the strength of the induced dipole moment. Since the KK-neutrino of UED models is a Dirac fermion from the 4-dimensional perspective, UED models that assume KK-neutrino dark matter are constrained by the strength of the induced dipole operator. In this letter we derive model independent bounds on KK-neutrino dark matter by examining the induced dipole moment and comparing the predictions with the current observational bound. Section II provides a brief review on the properties of KK-fermions in UED models along with a discussion on dipole moments relevant to the calculation presented in section III. Finally, in section IV we present our conclusions. \section{KK-neutrino LKP and induced dipole moments} In UED models, standard model fields become identified with the zero modes of KK towers of states once the extra dimensions are integrated out of the theory. For concreteness, we will restrict our discussion to five dimensional UED models compactified on an $S_1/Z_2$ orbifold. Five dimensional theories do not posses a chirality condition since $\gamma_5$ becomes part of the five dimensional Clifford Algebra. Thus, in order to arrive at a chiral theory at the zero mode level, we require one five dimensional \emph{Dirac} spinor for every \emph{Weyl} spinor of the standard model. By use of the orbifold boundary conditions half the number of states project out of the spectrum leaving a zero mode level chiral theory. In the case of the lepton doublet, the decomposition of the corresponding five dimensional Dirac spinor reads (\cf \eg \cite{Appelquist:2000nn}), \be \label{fermiondecomp} \begin{split} \hat{\mathcal{L}}(x^{\mu},y) = \frac{1}{\sqrt{\pi R}} P_L\mathcal{L}(x^\mu) + \sqrt{\frac{2}{\pi R}}\sum_{n=1}\left[P_L\mathcal{L}(x^\mu)\cos(\frac{ny}{R})\right.&\\ \left.+P_R\mathcal{L}(x^\mu)\sin(\frac{ny}{R})\right]&\\ \end{split} \ee where $\hat{\mathcal{L}}$ denotes the five dimensional Dirac spinor, $\mathcal{L}$ denotes a four dimensional \emph{4 component} spinor with $P_{L,R} = (1\pm\gamma_5)/2$, and $\hat{\mathcal{L}}$ is associated with the (4D left-handed) lepton doublet via the identification $P_L\mathcal{L}(x^\mu)\leftrightarrow (\nu_L,e_L)$. We see from eq(\ref{fermiondecomp}) that $\hat{\mathcal{L}}$ contains a purely left-handed zero mode while all higher KK modes contain both chiralities. Therefore, the non-zero mode KK-level fermions appear as Dirac spinors from the 4-dimensional perspective. Specifically, the KK-neutrino charged under the weak $SU(2)$ appears with both chiralities. Translational invariance in the fifth direction implies (once the extra dimension becomes integrated out) that a UED model compactified solely on $S_1$ conserves KK-excitation-number in every vertex. The orbifold reduces the conserved KK-number to a discrete $Z_2$ symmetry, called KK-parity. The conservation of KK-parity in every vertex implies the stability of the lightest KK-particle. Since the fermions receive mass through Yukawa couplings after electroweak symmetry breaking and through the KK-level expansion itself, the theory, in general, requires a unitary transformation to connect mass and gauge eigenstates. For example, in the lepton sector at the $j$th KK-level we have, \be \left(\begin{array}{c} \mathcal{E}^j \\ \mathcal{L}^j \end{array} \right) = \left(\begin{array}{cc}-\gamma_5\cos\alpha_j & \sin\alpha_j\\ \gamma_5\sin\alpha_j & \cos\alpha_j \end{array}\right)\left(\begin{array}{c} \mathcal{E}^{\prime j} \\ \mathcal{L}^{\prime j} \end{array} \right) \ee where $\mathcal{E}^j$ and $\mathcal{L}^j$ denote the $j^{\mbox{th}}$ KK mode of the lepton $SU(2)$ singlet and doublet respectively, and where \be \tan (2\alpha_j) = \frac{m_l^{(0)}}{j/R}. \ee As the lepton masses are small compared to $j/R$, we ignore the effects of the mixing matrix for the remainder of this letter. Furthermore, we ignore the effects of lepton flavour violation and neutrino mixing. A Dirac fermion can posses a dipole moment, derived from the transition amplitude (\cf \eg \cite{neutrinobook}), \be T = -i\epsilon^\mu q^\nu \bar f (p^\prime) \sigma_{\mu\nu} \left(F_2 + G_2 \gamma_5\right) f(p) \ee where $q=p^\prime -p$. The magnetic moment is defined by $\mu = F_2(0)$ while the electric dipole is defined by $ d = G_2(0)$. At low energies compared to the mass of the particle, the photon does not distinguish between $\mu$ or $d$ provided that one ignores other time-reversal violating observables. We will focus on the limits established on $\mu$ throughout. On dimensional grounds, we naively expect the induced magnetic dipole moment to scale as, \be \mu \lesssim e \frac{M_{\nu^{(1)}}}{R^{-2}} \simeq \frac{e}{M_{\nu^{(1)}}} = 1.022\times 10^{-6}\mu_B\left(\frac{\mathrm{TeV}}{M_{\nu^{(1)}}}\right) \ee where $M_{\nu^{(1)}}$ denotes the mass of KK-neutrino, $\nu^{(1)}$, and $R$ indicates the radius of compactification. \section{Calculation} The KK-neutrino develops a magnetic dipole moment through the diagrams tabulated in figure \ref{Figgraphs}. In general, the entire Kaluza-Klein tower of states participate, however we estimate the leading order effect by considering only the first level KK excitations. Furthermore, we restrict our calculation to KK-number conserving graphs since KK-number violation proceeds with volume suppression. For simplicity, we ignore flavour violation in the lepton sector and we assume that the lightest KK-neutrino is $\nu_e^{(1)}$. Relaxing these assumptions will not significantly alter our conclusions. As we make no assumptions on the exact UED spectrum, we consider the mass of the KK-$W$ ($W^{(1)}$) and the KK-electron ($e^{(1)}$) as free parameters. In our numerical calculations we do not consider a KK-electron/KK-neutrino mass difference in excess of 5\% as any substantial splitting will lead to unacceptably large contributions to the $T$ parameter. The relevant UED Feynman rules are listed in \cite{Buras:2002ej} and we calculate in the Feynman-'t Hooft gauge. In the limit of exact $M_{\nu^{(1)}}$-$M_{e^{(1)}}$ degeneracy and where the effects of Yukawa couplings are ignored, we arrive at the semi-analytic result, \be \label{dipole_result} \begin{split} \mu = & \frac{eg^2}{(4\pi)^2}\frac{1}{M_{\nu^{(1)}}} \times\\ &\left\{\frac{3}{2}\ln (\epsilon)+r+\frac{7}{2}+\frac{1}{2r}-\frac{5}{2}(r-1)\ln \left(\frac{r}{r-1}\right)\right.\\ &\left. -(r-1)^2\ln \left(\frac{r}{r-1}\right) + \mathcal{O}(\sqrt{\epsilon})\right\} \end{split} \ee with the approximation $\epsilon\equiv M_{W^{(0)}}^2/M_{\nu^{(1)}}^2\ll 1$ and $r\equiv M_{W^{(1)}}^2/M_{\nu^{(1)}}^2\simeq 1$. Numerically, we find agreement with our semi-analytical result as seen in figure \ref{plot_1}. In figure \ref{plot_1} we display the result of the dipole moment as a function of the KK-neutrino mass. The upper curve illustrates the magnetic dipole moment with exact degeneracy $M_{e^{(1)}}$-$M_{\nu^{(1)}}$ while holding the KK-$W$ mass at its tree level value. The middle curve plots the effect of a KK-electron 5\% heavier than the KK-neutrino. This has an $\mathcal{O}(1)$ effect on the dipole moment. The predicted value of the dipole moment exceeds the current upper bound for a KK-neutrino mass $M_{\nu^{(1)}}\sim 1 \hspace{1mm}\mathrm{TeV}$ by more than five orders of magnitude \cite{Sigurdson:2004zp}. The lower curve displays the effect of maintaining $M_{\nu^{(1)}}$-$M_{e^{(1)}}$ degeneracy while varying the KK-$W$ mass. Allowing the mass difference, $M_{W^{(1)}}$-$M_{\nu^{(1)}}$, to vary by up to 5\% has at most an $\mathcal{O}(10)$ effect. We should note that the calculation presented above determines only the radiatively induced part of the KK-neutrino magnetic dipole moment. The presence of boundary terms or effects arising from the UV complete theory may also contribute a non-renormalizable dimension-five dipole operator which, a priori, may be of the same order as the radiative part itself. \section{Conclusion} UED models have attracted attention as a possible extension to the standard model. A particular appealing feature of the model class centers on the existence of plausible dark matter candidates as the result of KK-parity conservation. While the minimal UED model suggests $B^{(1)}$ dark matter \cite{Cheng:2002ej}, detailed studies of the relic abundance in the minimal UED model \cite{coannmat,coannkribs, Matsumoto:2005uh} in combination with electroweak precision constraints \cite{Flacke:2005hb} show strong observational tension, and thus provide motivation for new possible UED model building avenues. The need for non-minimality has been reported \cite{Hewett:2004py}. Extensions of the MUED scenario by incalculable boundary terms arising from the UV completion of the model can lead to a different LKP and therefore different possible dark matter candidates. We have taken a model independent approach, following \cite{Servant:2002aq}, and examined the consequences of UED KK-neutrino dark matter. As the KK-neutrino is a Dirac fermion, UED models predict an induced KK-neutrino dipole moment. We find that the induced dipole moment, typically $\mu \lesssim 10^{-7} \mu_B$, strongly conflicts -- by over five orders of magnitude -- with the current observational bounds stated in \cite{Sigurdson:2004zp} for TeV scale dipole dark matter. We reiterate that the constraints provided by \cite{Sigurdson:2004zp} are currently under revision, and the strength of the stated bounds are expected to change \cite{private}. The constraints on magnetic dipole moments given in \cite{Sigurdson:2004zp} would provide the strongest limits on KK-neutrino dark matter. Even in the absence of the strong limits provided by \cite{Sigurdson:2004zp}, the bounds on dipole moments remain an important constraint on future model building. Not only will new, non-minimal models that predict KK-neutrino LKP need to circumvent the constaints provided by \cite{Servant:2002hb}, but will also have to evade the constraints provided by radiatively induced magnetic dipole moments, which are generically at least as large as the current experimental bounds. While we have restricted our discussion to five-dimensional models compactified on $S_1/Z_2$, we expect that the qualitative features carry over to UED models with multiple extra dimensions. We have also assumed the absence of any fine-tuning between the radiatively induced magnetic dipole moment and possible non-renormalizable dimension-five dipole operators arising from the UV complete theory or boundary terms. Our results indicate that observational limits on dipole dark matter can place significant constraints on UED scenarios where the KK-neutrino is the LKP dark matter candidate. \section{Acknowledgments} We would like to thank J.~March-Russell, G.~Starkman, B.~A.~Campbell, and K.~Sigurdson for useful discussions. DM wishes to acknowledge the support of the Natural Science and Engineering Research Council of Canada and the Canada-United Kingdom Millennium Research Fellowship. The work of TF is supported by ``Evangelisches Studienwerk Villigst e.V." and PPARC Grant No. PPA/S/S/2002/03540A. This work was also supported by the ``Quest for Unification" network, MRTN 2004-503369.
Title: Detection and Fundamental Applications of Individual First Galaxies
Abstract: First galaxies formed within halos of mass M=E7.5-E9 Msun at z=30-40 in the standard cold dark matter (CDM) universe may each display an extended hydrogen 21-cm absorption halo against the cosmic microwave background with a brightness temperature decrement of del T=-(100-150)mK at a radius 0.3 < r < 3.0 comoving Mpc, corresponding to an angular size of 10-100 arcseconds. A 21-cm tomographic survey in the redshift shell z=30-40 (at 35-45MHz), which could be carried out by the next generation of radio telescopes, is expected to be able to detect millions of first galaxies and may prove exceedingly profitable in enabling (at least) four fundamental applications for cosmology and galaxy formation. First, it may yield direct information on star formation physics in first galaxies. Second, it could provide a unique and sensitive probe of small-scale power in the standard cosmological model hence physics of dark matter and inflation. Third, it would allow for an independent, perhaps "cleaner" characterization of interesting features on large scales in the power spectrum such as the baryonic oscillations. Finally, possibly the most secure, each 21-cm absorption halo is expected to be highly spherical and faithfully follow the Hubble flow. By applying the Alcock-Paczynski test to a significant sample of first galaxies, one may be able to determine the dark energy equation of state with an accuracy likely only limited by the accuracy with which the matter density can be determined independently.
https://export.arxiv.org/pdf/astro-ph/0601010
\title{Detection and Fundamental Applications of Individual First Galaxies} \author{Renyue Cen\altaffilmark{1}} \altaffiltext{1} {Princeton University Observatory, Princeton University, Princeton, NJ 08544; cen@astro.princeton.edu} \received{\date} \accepted{ } \keywords{galaxies - radio - intergalactic medium - cosmology: theory} \section{Introduction} It is of wide interest to detect and understand the first generation of galaxies, expected to form in the redshift range $z=30-50$ in the standard CDM universe (Spergel \etal 2003). Extensive literatures on 21-cm properties of neutral hydrogen in the dark ages and during cosmological reionization have long focused on large-scale fluctuations of the intergalactic neutral hydrogen and global spectral features (e.g., Hogan \& Rees 1979; Scott \& Rees 1990). In this {\it Letter} we point out a unique feature possessed by the first {\it individual} galaxies of mass $10^{7.5}-10^9\msun$ formed at $z=30-40$ --- a large hydrogen 21-cm absorption halo against the cosmic microwave background (CMB). Each 21-cm absorption halo has a size $10^{''}-100^{''}$ with a brightness temperature decrement of $\delta T=-(100-150)$~mK at $35-45$MHz, which could serve as a visible proxy for each galaxy that otherwise may be undetectable. The next generation of radio telescopes, such as LOFAR, may be able to detect such a signal. A range of fundamental applications is potentially possible with a redshift (i.e., 21-cm tomographic) survey of the first galaxies in the redshift shell $z=30-40$, which may hold the promise to revolutionize the field of cosmology and shed illuminating light on dark matter, dark energy and inflation physics. Throughout, a standard (Wilkinson Microwave Anisotropy Probe) WMAP-normalized CDM model is used (unless indicated otherwise): $\Omega_M=0.31$, $\Lambda=0.69$, $\Omega_b=0.048$, $H_0=69$km/s/Mpc, $n_s=0.99$ and $\sigma_8=0.90$. \section{Large 21-cm Absorption Halos of First Galaxies} A first-generation galaxy is expected to emit UV and X-ray radiation, each carving out an H II region of size (ignoring recombination): $r_{\hbox{HII}} \sim 43 ({M_h\over 10^7\msun})^{1/3} ({c_*\over 0.1})^{1/3} ({f_{esc}\over 0.1})^{1/3} ({N_p\over 8\times 10^4})^{1/3}\kpc$ comoving, where $M_h$ is the halo mass, $c_*$ the star formation efficiency, $f_{esc}$ the ionizing photon escape fraction into the intergalactic medium (IGM), and $N_p$ the number of hydrogen ionizing photons produced by each baryon formed into stars, ($\sim 10^{4.5-5}$ for a massive metal-free Population III IMF; Bromm, Kudritzki, \& Loeb 2001). Hard X-ray photons ($\ge 1$keV) produced escape deep into the IGM with a distance of $\sim 500-1000$ comoving megaparsecs, building an X-ray background. Sandwiched between small H II regions and the X-ray sea sits a quite large \lya\ scattering region (Loeb \& Rybicki 1999), resulting in a four-layer structure, as depicted in Figure 1. The IGM in the vicinity % of a galaxy interacts with ionizing UV and soft X-ray as well as near \lya\ photons emanating from the galaxy, which can be computed. The most important and relevant physical processes are (1) the interaction between neutral gas and near \lya\ photons emitted by the central host galaxy, which couples the spin temperature of the IGM to its kinetic temperature (Wouthuysen 1952; Field 1958), and (2) the interaction between neutral gas and ionizing UV and soft X-ray photons emanating from the host galaxy, which provides a heating source for the otherwise cold IGM up to some small radius. In the absence of heating the kinetic temperature $T_k$ of the IGM would be equal to $T_{\hbox{IGM}} \approx 18 \left({1+z\over 31}\right)^2~\hbox{K}$ (for the redshift range of interest here) since beginning of decoupling with CMB at $z\sim 200$ (Peebles 1993). We perform spherically the transfer of UV and soft X-ray radiation from the host galaxy outward to compute (1) the development of the HII region around the galaxy as a function of time, (2) the evolution of the temperature of the surrounding IGM, subject to UV and soft X-ray heating by the host galaxy, as a function of radius and time, and (3) the \lya\ coupling coefficient $y_\alpha$ as a function of radius and time. We assume a uniform IGM density equal to the mean gas density of the universe (i.e., $\Delta=0$ in equation 1 below). Two scenarios of metal-free star formation in first galaxies are considered: (1) all stars have a single mass of $200\msun$ (VMS), advocated by Oh \etal (2001) and Qian \& Wasserburg (2002), and (2) an IMF has the Salpeter slope of $2.35$ with a lower cutoff of $25\msun$ and an upper cutoff of $120\msun$ (SAL), close to what favored by Umeda \& Nomoto (2003), Tumlinson, Venkatesan, \& Shull (2004) and Tan \& McKee (2004). We adopt the library of stellar spectra and ages from Schaerer (2002); we use a black-body radiation spectrum for each star of chosen mass with an effective surface temperature from Table 3 of Schaerer (2002) but slightly adjusted so as to produce the correct ratio of the number of photons above helium II Lyman limit to the number of photons above hydrogen Lyman limit, averaged over the lifetime of each star (see Table 4 of Schaerer 2002). We use an escape fraction for Lyman limit photons $f_{esc}$ from the host galaxy and self-consistently a frequency dependent escape fraction for other UV and soft X-ray photons assuming that they are subject to the same absorbing column in the galaxy. All escaped photons (emerging from the virial radius) are then subject to the (time-dependent) combined absorption of H I, He I and He II in the IGM, self-consistently computed. Since the \lya\ scattering region is mostly neutral with a residual ionized fraction of $2\times 10^{-4}$ left from recombination (Peebles 1993), we assume that $\eta=14\%$ of X-ray energy is used to heat the gas (Shull \& Van Steenberg 1985) with the heating rate per hydrogen atom at radius $r$ computed with the following formula: ${dE\over dt} = \int_0^\infty {\eta L_\nu\over 4\pi h\nu r^2}(\sigma_\nu(HI)(h\nu-h\nu_{H}) + \xi\sigma_\nu(HeI)(h\nu-h\nu_{HeI}) +\xi\sigma_\nu(HeII)(h\nu-h\nu_{HeII})) e^{-\tau_\nu} d\nu$, where $L_\nu$ is the luminosity per unit frequency of the galaxy; $\nu_{HI}$, $\nu_{HeI}$ and $\nu_{HeII}$ are ionization potentials of H, He I and He II, respectively; $\sigma_\nu(HI)$ $\sigma_\nu(HeI)$ and $\sigma_\nu(HeII)$ are photo-ionization cross-sections of H, He I and He II, respectively; $\xi$ is ratio of helium number density to hydrogen number density; $\tau_\nu$ is the optical depth from the galaxy to radius $r$ at frequency $\nu$. The spin temperature of neutral hydrogen (Field 1958, 1959) is then given by $T_{\hbox{s}}={T_{\hbox{cmb}}+y_\alpha T_k + y_c T_k\over 1+y_\alpha+y_c}$, where $y_c\equiv {C_{10}\over A_{10}} {T_*\over T_k}$ is the collisional coupling coefficient with the collisional de-excitation rate $C_{10}={4\over 4} \kappa (1-0) n_H$, $\kappa (1-0)$ taken from Zygelman (2005), $n_{\rm H}$ is the mean hydrogen density and $T_*=0.0682$~K is the hydrogen hyperfine energy splitting. In the expression for the \lya\ coupling coefficient $y_\alpha = {P_{10}T_*\over A_{10}T_k}$, $A_{10}=2.87\times 10^{-15}$~s$^{-1}$ is spontaneous emission coefficient of the 21-cm line, the indirect de-excitation rate $P_{10}$ of the hyperfine structure levels is related to the total \lya\ scattering rate $P_\alpha$ by $P_{10}=4P_\alpha/27$ (Field 1958). Here $P_\alpha=\int F_\nu \sigma(\nu) d\nu$ with $F_\nu$ being the \lya\ photon flux (in units of cm$^{-2}$~s$^{-1}$) and $\sigma(\nu)=\sigma_\alpha \phi(\nu) = {3\over 8\pi}\lambda_\alpha^2 A_\alpha \phi(\nu)$ being the cross section for \lya\ scattering (MMR), where $\lambda_\alpha=1.216\times 10^{-5}$~cm is the wavelength of the \lya\ line, $A_\alpha=6.25\times 10^8$~s$^{-1}$ is the spontaneous Einstein coefficient for \lya\ line and $\phi(\nu)$ is the normalized \lya\ line (Voigt) profile with $\int \phi(\nu) d\nu=1$. Most of the \lya\ scattering is accomplished by UV photons slightly on the blue side of the \lya\ that redshift into \lya\ resonant line due to the Hubble expansion (note that $\Delta\nu/\nu\sim 10^{-3}$ due to Hubble expansion at $r\sim 1$Mpc comoving), not the intrinsic \lya\ line photons that escape from the host galaxy and redshift to the damping wing (Madau, Meiksin, \& Rees 1997; MMR). Additional physical processes that were not treated previously, including higher-order Lyman lines that result in cascade in two-photon emission, fine structure of \lya\ resonance and spin-flip scattering, introduce corrections of order unity to $P_\alpha$ (CM; Hirata 2005; Chuzhoy \& Shapiro 2005) but all these corrections terms are insignificant for our case, and we only apply the relatively large correction term $S_c$ ($\sim 1.5$) as shown in Figure 4 of CM due to a spectral shape change near \lya\ . The observed brightness temperature increment/decrement against the CMB is \begin{eqnarray} \delta T = 41 (1+\Delta)x_H ({T_{\hbox{s}}-T_{\hbox{cmb}}\over T_{\hbox{s}}})({\Omega_b h^2\over 0.02})({0.15\over \Omega_Mh^2})^{1/2} ({1+z\over 31})^{1/2}~\hbox{mK}, \end{eqnarray} \noindent where $\Delta$ is gas overdensity relative to the mean, $x_H$ neutral hydrogen fraction, $T_{\hbox{cmb}}=2.73(1+z)$~K CMB temperature and other symbols have their usual meanings. Figure 2 shows the profile of $\delta T$ for four cases of halo masses with each choice of IMF. Let us examine each of the four regions (sketched in Figure 1) with respect to 21-cm observations. Inside the virial radius (the red circle in Figure 1) the gas is overdense with $\delta\ge 100$ and a positive large-amplitude emission signal may result, if a significant amount of neutral hydrogen gas exists within. However, the size of this regions falls below $0.1^"$ and its signal is unlikely to be detectable in the foreseeable future. The H II region (inside the blue circle in Figure 1) is ionized hence $\delta T=0$. In the region outside the \lya\ scattering region (exterior to the green region in Figure 1) the spin temperature of the IGM has been progressively attracted to the temperature of the CMB with gradually weakening coupling to the gas kinetic temperature by atomic collisions, producing a small but non-negligible 21-cm absorption signal at the redshift of interest ($z\sim 30-40$) (Loeb \& Zaldarriaga 2004). It is the \lya\ scattering region that is of most interest here. In the inner part of the \lya\ scattering region (shown in magenta in Figure 1) the IGM is significantly heated by UV and soft X-ray photons to exceed the CMB temperature, while its spin temperature is very strongly coupled to its kinetic temperature by \lya\ scattering. As a result, the inner radial region at $0.01-0.04$~Mpc/h comoving for minihalos and $0.04-0.4$~Mpc/h comoving for large halos displays an emission signal against CMB with an amplitude of $\delta T\sim 30$~mK (a shark dorsal fin-like feature in Figure 2). Going outward (shown in green in Figure 1), the soft X-ray heating abates (because the cumulative optical depth to these photons increases quickly) but the \lya\ scattering remains strong, up to a distance of about $10$~Mpc/h comoving. Consequently, a strong 21-cm absorption signal against the CMB with an amplitude of $\delta T= -(100-150)$~mK at $\sim 35-45$MHz (for $z=30-40$) on a scale of $0.3-3$~Mpc/h comoving, corresponding to an angular scale of $10^{''}-100^{''}$, is produced for large halos. This is the 21-cm absorption halo --- a unique and strong feature for the large first galaxies. We note that the absorption signal cast by minihalos (the two top sets of thin curves in each panel in Figure 2) is relatively weak due to a combination of low mass and low star formation efficiency. We will therefore focus on large halos for practical observability purposes. Besides stars, no other soft X-ray source in the galaxy is assumed to concurrently exist. We shall examine the validity of this assumption in detail. The density of the interstellar medium plays an important role for some of the potentially relevant processes considered here and it is assumed to be $n(z) = n_0 (1+z)^3$, where local interstellar density $n_0=1$~cm$^{-3}$. This assumption should hold in hierarchical structure formation model for the following reasons. The mean gas density scales as $(1+z)^3$ and halos at low and high redshift in cosmological simulations show similarities when density and length are measured in their respective comoving units (e.g., Navarro, Frenk, \& White 1997; Del Popolo 2001). The spin parameters (i.e., angular momentum distribution) of both high and low redshift halos have very similar distributions peaking at a nearly identical value $\lambda\sim 0.05$ (Peebles 1969; White 1984; Barnes \& Efstathiou 1987; Ueda \etal 1994; Steinmetz \& Bartelmann 1995; Cole \& Lacey 1996; Bullock \etal 2001). Thus, cooling gas in galaxies at low and high redshift should collapse by a similar factor before the structure becomes dynamically stable (e.g., rotation support sets in), resulting in interstellar densities scaling as $(1+z)^3$. Direct simulations (Abel \etal 2002; Bromm \etal 2002) suggest a gas density of $10^3-10^4$cm$^{-3}$ by the end of the initial free fall for minihalos at $z\sim 20$, verifying this simple analysis. We will estimate each of several possible types of soft X-ray emission sources in turn. First, let us estimate soft X-ray emission from supernova remnants. Assuming the standard cooling curve (Sutherland \& Dopita 1993) we find that at $z=30$ a supernova blastwave with an initial explosion energy of $5\times 10^{52}$~erg (for a star of mass $200\msun$; e.g., Heger \& Woosley 2002) would enter its rapid cooling phase at a temperature of $2.7\times 10^7$~K. This implies that the energy emitted at $\sim 100-300$~eV from the cooling shell is about $7\%$. A $200\msun$ mass would release $2.5\times 10^{54}$~erg total energy due to nuclear burning, out of which $0.3\%$ is released in photons at $100-300$eV for our adopted radiation spectrum. Therefore, the ratio of total photon energy from the supernova remnant to that from the star is $0.4-0.5$. Thus, for the VMS IMF, stellar soft X-ray appears to dominate over that from its supernova remnant. The soft X-ray contribution from supernova remnant cooling increases relatively compared to that from the star itself with decreasing stellar mass and we estimate that the overall contribution from the two components may become comparable for the SAL IMF case, averaged over time. Second, we will examine X-rays produced from cooling of supernova-accelerated relativistic electrons by CMB photons via inverse Compton (IC) process (e.g., Oh 2001). For adiabatic shocks, as is appropriate in our case, the IC spectral energy distribution has a two-power-law form: $L_\nu\propto$~constant at $E<E_{break}$ and $L_\nu\propto \nu^{-1}$ at $E>E_{break}$. The break energy is $E_{break}=70$~keV, independent of redshift with the assumed scaling of the interstellar medium density with redshift. Then the ratio of energy from IC to that from stars is found to be $10^{-4}-10^{-3}$ in the $100-300$eV band, depending on the exact upper energy cutoff (assuming $10\%$ of supernova explosion energy is utilized to accelerate relativistic electrons in shocks). Clearly, contribution to soft X-rays from IC process is unimportant. Third, X-ray binaries during the relatively short lifetime of massive stars may be rare, for top-heavy IMFs of concern here. We can make an estimate based on the calculation by Rappaport, Podsiadlowski \& Pfahl (2005), who give an ultra luminous X-ray binary formation rate of $3\times 10^{-5}$ per supernova. It is clear that even if each X-ray binary is able to release as much energy as in a supernova explosion itself and all in the soft X-ray band, the resulting contribution will be less than a fraction of a percent of that from stars. Fourth, stellar mass black holes (BH) of $\sim 10-100\msun$ may be produced in significant numbers with a top-heavy IMF as well as a central galactic BH. It seems that stellar BH accretion is likely significantly suppressed and small due to feedback effect from stars on surrounding gas (e.g., Mori, Umemura, \& Ferrara 2004; Alvarez, Bromm, \& Shapiro 2005). A concomitant contribution of soft X-rays from central BH accretion in the lifetime of a $200\msun$ star is approximately $(M_{BH}/M_*)\times (1/0.007)) \times (t_*/t_E) \times (f_{BH,SX}/f_{*,SX}) =0.6 f_{BH,SX} (M_{BH}/M_*/0.003)$, after inserting soft X-ray emission fraction for a $200\msun$ stellar spectrum of $f_{*,SX}=0.003$, stellar lifetime $t_*=2.2\times 10^6$~yrs and Eddington time $t_E=4.4\times 10^8$~yrs, where $f_{BH,SX}$ is the energy fraction released by the BH accretion in the soft X-ray band ($100-300$eV). Thus, if the ratio of black hole mass to (bulge) stellar mass follows the local Magorrian (Magorrian \etal 1998) relation, then, unless most of the accretion energy is released in the soft X-ray band, contribution from central BH accretion to the soft X-ray band is relatively small. Fifth, thermal bremsstrahlung emission from gravitational shock heated gas is likely negligible due to a low gas temperature ($T\sim 10^4$K). Finally, soft X-rays from massive first stars themselves are thought to be produced by stellar winds and quite uncertain. Recent work suggests that the winds hence soft X-ray emission from metal-free stars are expected to be insignificant (e.g., Krticka \& Kubat 2005). In summary, soft X-rays from neglected, possible sources other than that from the stellar photospheres would, at most, make a modest correction to what is adopted in our calculation. To ascertain our conclusion, we compute a case with the amplitude of soft X-ray intensity at $h\nu \ge 100$~eV artificially raised by a factor of $10$ and do not find any significant effect that would qualitatively change our results (Figure 2). The reason is that the IGM quickly becomes optically thick to a few $100$eV soft X-ray photons at $\sim 1$~Mpc comoving. Therefore, our results should be quite robust. Since we are concerned with gas of relatively low temperature $\sim 20$K, heating by a cumulative (hard) X-ray background may become relevant at some redshift. We estimate when this may happen in the CDM model. While an X-ray background may be generated by a variety of processes, black hole accretion at the centers of galaxies are thought to be the most dominant (e.g., Ricotti \& Ostriker 2003; Kuhlen \& Madau 2005), estimated as follows. Let us suppose the energy extraction efficiency from black accretion is $\alpha$ and a fraction $f_x$ of the released energy is in the form of hard X-rays. Then, the X-rays may collectively heat up the IGM temperature at most (ignoring Compton cooling and assuming all X-ray photons in the background are consumed by the IGM) by an increment \begin{equation} \Delta T_{\hbox{xray}} = 1.1({f_{\hbox{coll}}\over 10^{-6}}) ({c_*\over 0.1})({\alpha\over 0.1}) ({M_{BH}/M_*\over 0.003}) ({f_x\over 0.029})~\hbox{K} \end{equation} \noindent (assuming 14\% of X-ray energy is used to heat the IGM; Shull \& Van Steenberg 1985), where $f_{\hbox{coll}}$ is the fraction of matter that has collapsed to halos where stars have formed. Under the reasonable assumption that the parameters have their adopted fiducial values (for $\alpha$ see Yu \& Tremaine 2002, $f_x$ see Elvis \etal 1994, for $c_*$ see Gnedin 2000), it becomes evident that, for the very first galaxies formed in the universe that comprise a collapsed mass fraction less than $10^{-6}$, heating of the IGM by an X-ray background radiation field may be small. Figure 3 shows cumulative halo mass functions at redshift $z=(20,30,40)$, based on Press-Schechter (1974) formalism (using $\delta_c=1.67$ and the standard model of Spergel \etal 2003), which should be accurate for the exponentially falling regime of interest here (Sheth \& Tormen 1999; Jenkins \etal 2001). Figure 3 suggests that at redshift $z\ge 30$ heating of the IGM by an X-ray background is small. Neglected contributions to the X-ray background from other sources (X-ray binaries, supernova remnants, etc) (e.g., Oh 2001, Cen 2003) will likely just add a modest numerical correction factor for equation (2) and the net effect, if any, would push the epoch for significant heating by an X-ray background to a slightly higher redshift (note that structure formation is exponentially increasing with decreasing redshift at $z=30-40$). But even a factor of a few upward correction to equation (2) would still leave the temperature of IGM $z=30$ relatively unaffected by an X-ray background. Another critical issue is whether heating of IGM by \lya\ photons is important. In a recent accurate calculation based on Fokker-Planck approximation, Chen \& Miralda-Escud\'e (2004; CM) show that the heating rate by \lya\ photons is much lower than previous estimates (MMR). We recast their important result (equation 17 of CM and using Figure 3 of CM), the heating rate per hydrogen atom and per Hubble time, $\beta$, in the following way: \begin{equation} \beta\equiv {\Gamma_c\over H n_{H} k_{B}} = 0.08 ({P_\alpha\over P_{th}})~\hbox{K} \end{equation} \noindent at $z=30$, where $k_{\rm B}$ is the Boltzmann constant, $H$ is the Hubble constant and $P_{th}=2.4\times 10^{-11}({1+z\over 31})$~s$^{-1}$ is the thermalization rate for \lya\ scattering at $z=30$, above which \lya\ scattering brings down the spin temperature to the gas kinetic temperature (MMR). The Hubble time is $1.4\times 10^8$~yrs at $z=30$. So, over the duration of a stellar lifetime $6\times 10^6$~yrs (of the least massive stars in our model, $25\msun$), the gas will be heated up by $0.003$~K at ${P_\alpha\over P_{\rm th}}=1$. For the regime of interest where we see the strong 21-cm absorption signal (Figure 2) we find ${P_\alpha\over P_{\rm th}}=1-10$. Thus, heating of surrounding IGM by \lya\ photons emanating from the host galaxy can be safely neglected. In addition, since we are concerned with early times when the universe is far from being ionized and the number of \lya\ photons per hydrogen atom is significantly less than unity, indicating that heating by the background \lya\ photons can also be safely neglected (CM). Furthermore, heating rate by high order Lyman series photons is still lower and thus negligible (Pritchard \& Furlanetto 2005). \section{Fundamental Applications} We have demonstrated a unique feature of first galaxies. A large 21-cm survey of the first galaxies will be invaluable. Demanding that each of the physical quantities be resolved by a factor of $10$ would translate to the following requirements: an angular resolution of $\sim 1^{''}$, a spectral resolution of $\sim 4$kHz ($\Delta\nu\sim 40$kHz across a radius of $1$Mpc/h at $z=30$ due to the Hubble expansion) and a sensitivity of $\sim 10$~mK at $35-45$~MHz. Among the next generation of radio telescopes currently under construction/consideration, LOFAR (http://www.lofar.org) appears to be best positioned to be able to carry out such a survey, at least for some of the brighter and larger galaxies. LOFAR is currently designed to reach a frequency as low as $10$MHz with an angular resolution of $3^{''}-4^{''}$ at $35-45$MHz, a sensitivity of $10$~mK and a spectral resolution (i.e., processing capability) of $1$~kHz (e.g., Rottgering 2003), while MWA (http://web.haystack.mit.edu/arrays/MWA/index.html), PaST (http://web.phys.cmu.edu/$\sim$past/index.html) and SKA (http://www.skatelescope.org) appear to be placed out of the $35-45$MHz range, as they stand now. Figure 4 shows the density of galaxies versus the maximum absorption cross section (i.e., in the plane of the large circle centered on the galaxy perpendicular to the line of sight) with $\delta T<-100$~mK. Here we point out four fundamental and potentially ground-breaking applications regarding cosmology and galaxy formation, if a 21-cm tomographic survey of galaxies in the redshift shell $z=30-40$ is carried out, which may be able to detect millions of galaxies. First, a characteristic sharp fall-off at $5-10$ square arcseconds and a characteristic peak of the number of 21-cm absorption halos is expected, as seen in Figure 4 due to a lower star formation efficiency in (and low mass of) minihalos, as suggested by available simulations (Abel \etal 2002). This should yield direct information on physics of cooling and star formation in first galaxies, which may be unobtainable otherwise by any other means in the foreseeable future. Note that the left cutoff and the peak density are functions of $c_*$ and $g({\rm IMF})$ (noting the differences between SAL and VMS cases in Figure 4), where $g({\rm IMF})$ denotes dependence on the properties of IMF such as stellar lifetime and spectrum. A full parameter space exploration will be given in a separate paper with more detailed treatments. We expect that $c_*$ and $g({\rm IMF})$ may be determined separately, when jointly analyzed with the density of absorption halos in the context of the standard CDM model. Second, we see that the density of strong 21-cm absorption halos depends strongly on $n_s$, as testified by the large difference (a factor of $\sim 50$) between solid and dotted curves in Figure 4. One may then obtain a constraint on $n_s$, which is made possible because the effect due to difference in IMF may be isolated out, as discussed above, thanks to the features in the density of absorption halos (e.g., peak location and sharp fall-off at the low end). Let us estimate a possible accuracy of such measurements. At $n\sim 10^{-6}$h$^3$Mpc$^{-3}$ one would find 0.1 million galaxies in the redshift shell between $z=28$ and $z=32$, giving a relative fraction (Poisson) error of $0.3\%$. By comparing the solid and dotted curves in Figure 4, we find that a constraint on $n_s$ with $\Delta n_s=0.01$ ($\sim 3\sigma$) may be achieved. This may have the potential to discriminate between inflationary theories (e.g., Liddle \& Lyth 1992; Peiris \etal 2003). In addition, the constraint placed on the temperature (or mass) of dark matter particles or running of the spectral index may be still tighter, because a significant, finite dark matter temperature or a running index tends to suppress small-scale power exponentially thus amplify the effects. The high-sensitivity constraint on small-scale power is afforded by the physical fact that we are dealing with rare $\ge 5-6\sigma$ peaks in the matter distribution. Third, clustering of galaxies may be computed using such a survey containing potentially hundred of thousands to millions of galaxies in a comoving volume of size $\sim 100$Gpc$^3$ (for the redshift shell $z=28-32$). Both the survey volume and the number of observable galaxies within are large enough to allow for accurate determinations of the correlations of first galaxies, particularly on large scales. It may then provide an independent, perhaps ``cleaner" characterization of interesting features in the power spectrum such as the baryonic oscillations, with the advantage that they are not subject to subsequent complex physical processes, including cosmological reionization, gravitational shock heating of the IGM and complex interplay between galaxies and IGM, which in turn might introduce poorly understood biases in galaxy formation. A comparison between clustering of first galaxies and local galaxies (e.g., Eisenstein \etal 2005) will provide another, high-leverage means to gauge gravitational growth and other involved processes between $z=30$ to $z=0$. Finally, the scale ($\sim 1$Mpc comoving) of the 21-cm absorption halo signals is much greater than the nonlinear scale and virial radius (both $\sim 1$kpc comoving). Thus, the IGM region in the 21-cm absorption halo is expected to closely follow the Hubble flow. Since near \lya\ photons (between \lya\ and Ly$\beta$) are not subject to absorption by hydrogen (and helium) atoms whose distribution might be complex, they escape into the IGM in a spherical fashion. Additionally, since the dependence on $\Delta$ is linear (see equation 2), density inhomogeneities are likely to average out (to zero-th order) and results do not depend sensitively on uncertain linear density fluctuations in the IGM. Although there might exist ``pores" in the domain of 21-cm absorption halo due to fluctuations in local IGM temperatures which may be caused by local shock heating due to formation of minihalos, the overall effect is likely negligible, because the mass fraction contained in all halos down to a mass as small as $M_h=10^{5}\msun$ is about $10^{-4}$ at $z=30$. Furthermore, at $n=10^{-6}$h$^3$Mpc$^{-3}$ the mean separation between the galaxies is $100$~Mpc/h, much larger than the size of \lya\ scattering regions of size $\sim 1$Mpc/h, so overlapping of the latter should be very rare (taking into account the known fact that they are strongly clustered in the standard cosmological model with {\it gaussian} random fluctuations; Mo \& White 1996). For these reasons, each 21-cm absorption halo is expected to be highly spherical in real space. Therefore, 21-cm absorption halos are ideal targets to apply the Alcock-Paczy\'nski (1979) test. Accurate measurements of angular size $\Delta\theta$ and radial depth $\Delta v$ for a sample of galaxies would yield a sample of $d_A(z) H(z)$, where $d_A(z)$ and $H(z)$ are the angular diameter distance and Hubble constant, respectively, both of which are, in general, functions of $\Omega_M$, $w$($\equiv p/\rho$) and $k$, with $w$ describing the equation of state for dark energy and $k$ being the curvature of the universe. As an example, let us assume that $\Omega_M$($\approx 0.3$) has been fixed exactly by independent observations and $k=0$ and that $w\approx -1$. Then one obtains $|d ln[d_A(z) H(z)]/dw|=0.45$ at $z=30$ (Huterer \& Turner 2001). Let us suppose a relative measurement error on each individual $d_A(z) H(z)$ is $20\%$, then with ten thousand galaxies, one could obtain a highly accurate constraint on $w$ with $\Delta w = 20\%/0.45/\sqrt{10000} \sim 0.004$. Likely, the accuracy of $w$ determined by this method may eventually be limited by the accuracy with which $\Omega_M$ (and $k$) can be determined by independent observations, due to the degenerate nature. We stress that this method is valid for each individual first galaxy and unaffected by uncertainties, for example, in the precise abundance of such galaxies. In post-survey analyzes one faces the practical issue of extracting the wanted signals from the raw data, whose amplitude is expected to be dominated by foreground radio sources, including galactic synchrotron radiation, galactic and extragalactic free-free emission, and extragalactic point sources (e.g., Di Matteo, Ciardi, \& Miniati 2004). While seemingly daunting, it has already been shown that signals of the amplitude proposed here may be recovered with relatively high fidelity, when one takes into account the expected, potent differences in the spectral and angular properties between the 21-cm signal and foreground contaminants (e.g., Zaldarriaga, Furlanetto, Hernquist 2004; Santos, Cooray, \& Knox 2005; Wang \etal 2005). Since the 21-cm absorption halos are expected to be rather regular and simple, one might be able to significantly enhance the signal by using additional techniques, such as matched filter algorithm, in combination with foreground ``cleaning" methods. Finally, the amount of data in such a high spatial and frequency resolution 3-dimensional survey will be many orders of magnitude larger than that of WMAP. Computational challenges for analyzing it will be of paramount concern and most likely demand new and innovative approaches. \section{Conclusions} It is shown that a first galaxy hosted by a halo of mass $M=10^{7.5}-10^9\msun$ at $z=30-40$ possesses a large 21-cm absorption halo against the CMB with a brightness temperature decrement $\delta T=-(100-150)$~mK and an angular size of $10^{''}-100^{''}$. A 21-cm tomographic survey of galaxies in the redshift shell at $z=30-40$ may detect millions of galaxies and may yield critical information on cosmology and galaxy formation. A successful observation may need an angular resolution of $\le 1^{''}$, a spectral resolution of $\le 4$kHz, and a sensitivity of $\le 10$~mK at $35-45$~MHz. LOFAR appears poised to be able to execute this unprecedented task, at least for the high end of the distribution. At least four fundamental applications may be launched with such a survey, which could potentially revolutionize cosmological study and perhaps the field of astro-particle physics. First, it may provide unprecedented constraint on star formation physics in first galaxies, for there is a proprietary sharp feature related to the threshold halo mass for efficient atomic cooling. Second, it may provide a unique and sensitive probe of the small-scale power in the cosmological model hence physics of dark matter and inflation, by being able to, for example, constrain $n_s$ to an accuracy of $\Delta n_s=0.01$ at a high confidence level. Constraints on the nature of dark matter particles, i.e., mass or temperature, or running of index could be still tighter. Third, clustering of galaxies that may be computed with such a survey will provide an independent set of characterizations of potentially interesting features on large scales in the power spectrum including the baryonic oscillations, which may be compared to local measurements (Eisenstein \etal 2005) to shed light on gravitational growth and other involved processes from $z=30$ to $z=0$. Finally, the 21-cm absorption halos are expected to be highly spherical and trace the Hubble flow faithfully, and thus are ideal systems for an application of the Alcock-Paczy\'nski test. Exceedingly accurate determinations of key cosmological parameters, in particular, the equation of state of the dark energy, may be finally realized. As an example, it does not seem excessively difficult to determine $w$ to an accuracy of $\Delta w\sim 0.01$, if $\Omega_M$ has been determined to a high accuracy by different means. If achieved, it may have profound ramifications pertaining dark energy and fundamental particle physics (e.g., Upadhye, Ishak, \& Steinhardt 2005). If a null detection of the proposed signal is found, as it might turn out, implications may be as profound. It might be indicative of some heating and/or reionizing sources in the early universe ($z=30-200$) that precede or are largely unrelated to structure formation, possibly due to yet unknown properties of dark matter particles or dark energy. Alternatively, star formation and/or BH accretion in first galaxies may be markedly different from our current expectations. \acknowledgments I thank Dr. Daniel Schaerer for helpful information on Pop III stars. This research is supported in part by grants AST-0206299, AST-0407176 and NAG5-13381.
Title: Radial velocities of population II binary stars. II
Abstract: Here we publish the second list of radial velocities for 91 Hipparcos stars, mostly high transverse velocity binaries without previous radial velocity measurements. The measurements of radial velocities are done with a CORAVEL-type radial velocity spectrometer with an accuracy better than 1 km/s. We also present the information on eight new radial velocity variables - HD 29696, HD 117466AB, BD +28 4035AB, BD +30 2129A, BD +39 1828AB, BD +69 230A, BD +82 565A and TYC 2267-1300-1 - found from our measurements. Two stars (HD 27961AB and HD 75632AB) are suspected as possible radial velocity variables.
https://export.arxiv.org/pdf/astro-ph/0601212
\begin{center} \vbox{\scriptsize \tabcolsep 2pt \begin{tabular}{rlcrc|rlcrc} \multicolumn{10}{c}{\parbox{120mm}{\baselineskip=8.5pt {\smallbf Table 3.}{ \small Individual radial velocities.}}}\\ \noalign{\vskip1mm} \tablerule No. & ~~Star name & HJD & ${V_r}$~~ & ${\varepsilon}1 $ & No. & ~~Star name & HJD & ${V_r}$~~ & ${\varepsilon}1 $ \\ & & +2400000 & km/s & km/s & & & +2400000 & km/s & km/s \\ \tablerule \noalign{\vskip1mm} 1. & HD\,225220\,AB & 51794.481 & --24.3 & 0.6 & ~~34. & & 52994.830 & --6.9 & 0.6 \\ & & 52950.633 & --24.4 & 0.6 & ~~35. & HIP\,17876 & 52992.880 & --10.3 & 0.8 \\ 2. & HIP\,375 & 51783.569 & --9.8 & 0.7 & ~~36. & HD\,26735\,A & 51893.445 & 64.3 & 0.7 \\ & & 51794.486 & --8.6 & 0.7 & & & 51893.453 & 64.8 & 0.8 \\ & & 52950.672 & --9.4 & 0.6 & & & 52991.865 & 63.0 & 0.7 \\ 3. & HDE\,236325 & 51784.493 & 40.7 & 0.6 & ~~37. & HD\,26735\,B & 51893.441 & 62.8 & 0.8 \\ & & 51795.494 & 40.5 & 0.6 & & & 51893.449 & 63.6 & 0.8 \\ 4. & BD\,+38 35\,AB & 51786.503 & --9.1 & 0.8 & & & 52991.879 & 64.4 & 0.6 \\ & & 51794.507 & --9.7 & 0.8 & & & 52994.842 & 63.6 & 0.6 \\ 5. & HDE\,232169 & 51784.502 & --3.6 & 0.7 & ~~38. & HD\,27961\,AB & 52949.913 & 36.3 & 0.7 \\ & & 51795.501 & --4.3 & 0.7 & ~~39. & BD\,+03 609 & 52990.846 & 53.4 & 0.7 \\ 6. & BD\,+23 80\,A & 51786.522 & --17.1 & 0.7 & & & 52995.791 & 52.5 & 0.7 \\ & & 51794.525 & --17.1 & 0.7 & ~~40. & HIP\,21089\,AB & 51893.465 & 32.0 & 1.0 \\ 7. & BD\,+23 80\,B & 51786.526 & --17.7 & 1.2 & & & 52990.859 & 32.4 & 0.7 \\ & & 51794.531 & --17.1 & 1.5 & & & 52995.799 & 32.5 & 0.7 \\ 8. & HDE\,236523\,AB & 51784.537 & --9.7 & 0.8 & ~~41. & HIP\,21842 & 52990.901 & 4.4 & 0.8 \\ & & 51795.508 & --7.7 & 0.8 & & & 52995.805 & 4.7 & 0.7 \\ & & 52950.689 & --10.7 & 0.6 & & & 52995.813 & 5.4 & 0.7 \\ 9. & HD\,6094\,AB & 52950.700 & 8.7 & 0.7 & ~~42. & BD\,+44 1142\,AB & 51892.484 & 0.2 & 0.8 \\ 10. & HD\,6448\,A & 51784.552 & --25.3 & 0.6 & & & 52949.965 & --1.3 & 0.6 \\ 11. & HD\,6448\,B & 51784.559 & --25.6 & 0.8 & ~~43. & LP\,360-6\,AB & 51892.500 & --1.5 & 1.0 \\ 12. & BD\,+83 28\,A & 51794.546 & --18.2 & 0.7 & & & 52940.027 & --1.4 & 0.7 \\ 13. & BD\,+14 232\,AB & 51786.563 & 25.1 & 0.7 & & & 52940.031 & --2.5 & 0.8 \\ & & 51794.593 & 24.5 & 0.7 & & & 52950.018 & --4.1 & 0.7 \\ 14. & BD\,+75 65\,A & 51786.569 & --22.8 & 0.7 & ~~44. & HIP\,25413 & 52990.944 & 24.1 & 0.9 \\ & & 51794.572 & --22.4 & 0.7 & & & 53000.810 & 23.9 & 0.8 \\ 15. & HIP\,8607 & 51795.541 & 25.0 & 0.8 & ~~45. & HDE\,244359\,AB & 52991.897 & 98.8 & 0.7 \\ & & 52953.828 & 23.9 & 0.8 & & & 53001.834 & 99.4 & 0.7 \\ & & 52959.766 & 23.3 & 0.7 & ~~46. & HD\,36195\,AC & 52991.918 & --21.7 & 0.6 \\ 16. & HD\,14106 & 51795.554 & 51.8 & 0.7 & ~~47. & WDS\,05329+5208\,B & 52991.926 & --22.0 & 0.8 \\ 17. & HD\,14202 & 51784.580 & --10.9 & 0.7 & ~~48. & HD\,248330 & 52991.946 & 24.9 & 0.6 \\ & & 51795.563 & --11.7 & 0.7 & ~~49. & HD\,39448 & 52991.968 & 45.6 & 0.6 \\ 18. & HIP\,10774\,B & 51784.575 & --10.4 & 0.7 & ~~50. & HD\,40412 & 52991.981 & --46.5 & 0.7 \\ & & 51795.567 & --10.4 & 0.7 & ~~51. & HIP\,120002 & 52991.992 & --46.5 & 0.7 \\ 19. & HD\,14511\,AB & 51784.587 & 13.1 & 0.7 & ~~52. & BD\,+77 250 & 51795.611 & --52.6 & 0.8 \\ & & 51786.582 & 12.2 & 0.7 & & & 51894.547 & --53.5 & 0.8 \\ & & 51795.574 & 12.5 & 0.7 & ~~53. & HD\,49622\,AB & 51892.477 & 10.5 & 0.9 \\ 20. & BD\,+00 494\,AB & 52990.792 & 31.8 & 0.7 & & & 52940.039 & 10.8 & 0.7 \\ 21. & BD\,+39 692\,AB & 52991.736 & --18.2 & 0.6 & & & 52950.026 & 8.6 & 0.6 \\ & & 52994.801 & --18.3 & 0.6 & ~~54. & BD\,+71 380\,A & 51894.535 & 25.8 & 0.8 \\ 22. & BD\,+01 549\,AB & 52990.807 & 53.1 & 0.7 & ~~55. & BD\,+65 559\,AB & 51894.563 & 52.3 & 0.9 \\ & & 52994.769 & 52.5 & 0.6 & ~~56. & HD\,56244\,AB & 52990.994 & --36.8 & 0.6 \\ 23. & BD\,+00 523\,AB & 52990.816 & --4.0 & 0.7 & & & 52994.004 & --37.1 & 0.6 \\ & & 52994.777 & --3.7 & 0.7 & ~~57. & BD\,+28 1365 & 52991.009 & 38.9 & 0.8 \\ 24. & BD\,+00 549\,A & 51893.340 & 88.1 & 0.9 & & & 53001.981 & 37.6 & 0.7 \\ 25. & BD\,+15 452\,AB & 52953.840 & 25.3 & 0.7 & ~~58. & BD\,+39 1967\,AB & 52991.027 & 33.7 & 1.0 \\ & & 52958.846 & 25.0 & 0.6 & & & 53001.993 & 37.5 & 0.8 \\ 26. & HD\,20289\,AB & 52991.744 & 16.8 & 0.6 & ~~59. & HD\,60820\,AB & 51892.590 & 63.0 & 0.6 \\ & & 52994.794 & 17.4 & 0.6 & ~~60. & BD\,+25 1788\,AB & 52991.047 & 58.4 & 0.7 \\ 27. & HD\,20369\,AB & 52990.830 & --65.1 & 0.6 & & & 53002.002 & 59.6 & 0.7 \\ & & 52994.785 & --65.7 & 0.7 & ~~61. & HD\,71185 & 51894.578 & --65.2 & 0.7 \\ 28. & BD\,+12 472\,AB & 52991.755 & --14.9 & 0.8 & ~~62. & BD\,+40 2062\,AB & 52992.028 & 44.2 & 0.9 \\ & & 52994.808 & --14.4 & 0.7 & & & 53002.009 & 44.7 & 0.9 \\ 29. & HIP\,16332\,AB & 51795.589 & --55.2 & 0.8 & ~~63. & HD\,73889 & 52993.021 & 56.0 & 0.6 \\ 30. & HDE\,232816 & 51786.599 & 4.3 & 0.7 & ~~64. & HD\,74861\,AB & 52951.025 & --25.5 & 0.6 \\ & & 51795.515 & 4.0 & 0.7 & & & 52952.030 & --25.9 & 0.6 \\ 31. & BD\,+16 492\,AB & 52992.824 & 23.5 & 0.6 & & & 52953.027 & --23.6 & 0.7 \\ & & 52994.822 & 23.0 & 0.6 & & & 52953.030 & --24.4 & 0.7 \\ 32. & HD\,23439\,A & 52949.936 & 50.5 & 0.6 & ~~65. & BD\,+46 1436 & 52991.056 & 58.6 & 0.7 \\ & & 52958.865 & 50.4 & 0.7 & & & 52996.065 & 58.8 & 0.7 \\ 33. & HD\,23439\,C & 52949.952 & --0.5 & 0.7 & ~~66. & HD\,75632\,AB & 52985.016 & 48.0 & 0.7 \\ & & 52958.899 & --1.6 & 0.7 & & & 52987.017 & 45.8 & 0.6 \\ 34. & HD\,23865\,AB & 52992.836 & --8.0 & 1.0 & ~~67. & BD\,+49 2588 & 51784.328 & 24.9 & 0.7 \\ & & 52992.841 & --9.7 & 0.9 & & & 51785.297 & 26.3 & 0.7 \\ \end{tabular} } \end{center} \newpage \begin{center} \vbox{\scriptsize \tabcolsep 2pt \begin{tabular}{rlcrc|rlcrc} \multicolumn{10}{c}{\parbox{125mm}{\baselineskip=8.5pt {\smallbf Table 3.}{ \small \hskip 1.5mm Continued}}}\\ \noalign{\vskip1mm} \tablerule No. & ~~Star name & HJD & ${V_r}$~~ & ${\varepsilon}1 $ & No. & ~~Star name & HJD & ${V_r}$~~ & ${\varepsilon}1 $ \\ & & +2400000 & km/s & km/s & & & +2400000 & km/s & km/s \\ \tablerule \noalign{\vskip1mm} 68. & BD\,+68 986\,A & 51784.338 & 1.5 & 0.8 & ~~81. & & 51022.427 & 16.6 & 0.7 \\ & & 51794.337 & 2.0 & 0.7 & & & 51784.436 & 17.3 & 0.7 \\ 69. & BD\,+75 641 & 51016.382 & --118.8 & 0.7 & & & 51795.384 & 18.0 & 0.7 \\ & & 51018.403 & --119.4 & 0.7 & ~~82. & BD\,+14 4697\,AB & 51779.467 & --57.0 & 0.7 \\ & & 51018.423 & --119.3 & 0.7 & & & 51786.446 & --55.6 & 0.7 \\ & & 51022.408 & --119.0 & 0.7 & ~~83. & HDE\,239892\,AB & 51021.457 & --19.0 & 1.1 \\ & & 51785.285 & --117.5 & 0.7 & & & 51021.467 & --22.2 & 1.0 \\ & & 51795.370 & --117.3 & 0.7 & & & 51022.448 & --18.6 & 1.0 \\ 70. & HD\,177349 & 51780.375 & --1.9 & 0.8 & & & 51784.447 & --21.3 & 1.3 \\ & & 51785.339 & --0.9 & 0.7 & & & 51795.447 & --20.5 & 0.9 \\ 71. & HIP\,93600 & 51780.383 & --30.6 & 0.9 & ~~84. & HDE\,235919\,AB & 51784.459 & --30.4 & 0.7 \\ & & 51785.343 & --31.4 & 0.7 & & & 51795.458 & --30.3 & 0.7 \\ 72. & BD\,+49 3113 & 51784.352 & --8.9 & 0.7 & ~~85. & HDE\,240021 & 51021.481 & --1.9 & 1.1 \\ & & 51785.310 & --8.8 & 0.7 & & & 51021.496 & --3.7 & 1.4 \\ 73. & HDE\,239292\,AB & 51784.364 & --83.3 & 0.7 & & & 51022.464 & --2.2 & 1.1 \\ & & 51794.363 & --83.3 & 0.7 & & & 51784.471 & --5.2 & 1.0 \\ 74. & HDE\,356668\,AB & 51780.396 & 10.8 & 0.9 & & & 51795.468 & --3.0 & 1.2 \\ & & 51785.376 & 10.8 & 0.7 & ~~86. & HDE\,236100 & 51779.512 & --38.8 & 0.7 \\ 75. & HDE\,340730 & 51784.387 & --10.5 & 0.8 & & & 51783.420 & --39.7 & 0.7 \\ & & 51785.391 & --10.8 & 0.7 & ~~87. & BD\,+21 4948 & 51783.461 & 3.3 & 0.7 \\ 76. & GSC\,2161-755 & 51785.403 & --35.5 & 0.9 & & & 51786.453 & 4.4 & 0.8 \\ 77. & HDE\,235300 & 51784.377 & --110.9 & 0.7 & ~~88. & PPM\,11774 & 51794.464 & --33.3 & 0.7 \\ & & 51794.374 & --111.5 & 0.7 & ~~89. & HDE\,236223 & 51021.515 & 1.0 & 0.7 \\ 78. & BD\,+13 4614\,AB & 51780.406 & --24.1 & 0.8 & & & 51022.479 & 1.7 & 0.7 \\ & & 51786.381 & --25.1 & 0.8 & & & 51784.482 & 2.3 & 0.7 \\ 79. & BD\,+25 4462\,A & 51780.449 & --21.4 & 1.0 & & & 51795.482 & 1.6 & 0.7 \\ & & 51786.389 & --20.7 & 1.1 & ~~90. & HDE\,223523\,AB & 52950.644 & --90.5 & 0.6 \\ 80. & G\,212-24\,AB & 51780.481 & --101.7 & 0.9 & & & 52953.684 & --91.7 & 0.7 \\ & & 51786.415 & --102.0 & 0.8 & ~91. & BD\,+43 4596\,AB & 51783.555 & 9.2 & 0.8 \\ 81. & HDE\,235634 & 51010.522 & 17.5 & 0.7 & & & 51786.461 & 9.1 & 0.7 \\ & & 51016.403 & 17.0 & 0.8 & & & 51794.447 & 9.9 & 0.7 \\ \tablerule \end{tabular} } \end{center}
Title: Large dust grains in the inner region of circumstellar disks
Abstract: CONTEXT: Simple geometrical ring models account well for near-infrared interferometric observations of dusty disks surrounding pre-main sequence stars of intermediate mass. Such models demonstrate that the dust distribution in these disks has an inner hole and puffed-up inner edge consistent with theoretical expectations. AIMS: In this paper, we reanalyze the available interferometric observations of six intermediate mass pre-main sequence stars (CQ Tau, VV Ser, MWC 480, MWC 758, V1295 Aql and AB Aur) in the framework of a more detailed physical model of the inner region of the dusty disk. Our aim is to verify whether the model will allow us to constrain the disk and dust properties. METHODS: Observed visibilities from the literature are compared with theoretical visibilities from our model. With the assumption that silicates are the most refractory dust species, our model computes self-consistently the shape and emission of the inner edge of the dusty disk (and hence its visibilities for given interferometer con gurations). The only free parameters in our model are the inner disk orientation and the size of the dust grains. RESULTS: In all objects with the exception of AB Aur, our self-consistent models reproduce both the interferometric results and the near-infrared spectral energy distribution. In four cases, grains larger than 1.2 micron, and possibly much larger are either required by or consistent with the observations. The inclination of the inner disk is found to be always larger than 30 deg, and in at least two objects much larger.
https://export.arxiv.org/pdf/astro-ph/0601438
\title{Large dust grains in the inner region of circumstellar disks} \author{ Andrea Isella \inst{1,2}, Leonardo Testi \inst{1} and Antonella Natta \inst{1} } \institute{ Osservatorio Astrofisico di Arcetri, INAF, Largo E.Fermi 5, I-50125 Firenze, Italy \and Dipartimento di Fisica, Universit\'a di Milano, Via Celoria 16, 20133 Milano, Italy } \offprints{isella@arcetri.astro.it} \date{Received ...; accepted ...} \authorrunning{ISELLA, TESTI \& NATTA} \titlerunning{Large grains in circumstellar disks} \abstract { Simple geometrical ring models account well for near-infrared interferometric observations of dusty disks surrounding pre-main sequence stars of intermediate mass. Such models demonstrate that the dust distribution in these disks has an inner hole and puffed-up inner edge consistent with theoretical expectations. } { In this paper, we reanalyze the available interferometric observations of six intermediate mass pre-main sequence stars (CQ Tau, VV Ser, MWC 480, MWC 758, V1295 Aql and AB Aur) in the framework of a more detailed physical model of the inner region of the dusty disk. Our aim is to verify whether the model will allow us to constrain the disk and dust properties.} { Observed visibilities from the literature are compared with theoretical visibilities from our model. With the assumption that silicates are the most refractory dust species, our model computes self-consistently the shape and emission of the inner edge of the dusty disk (and hence its visibilities for given interferometer configurations). The only free parameters in our model are the inner disk orientation and the size of the dust grains.} { In all objects with the exception of AB Aur, our self-consistent models reproduce both the interferometric results and the near-infrared spectral energy distribution. In four cases, grains larger than $\sim$1.2\um, and possibly much larger are either required by or consistent with the observations. The inclination of the inner disk is found to be always larger than $\sim 30^{\circ}$, and in at least two objects much larger.} \keywords{} \section {Introduction} Understanding the properties and evolution of the dust grains contained in proto-planetary disks around pre-main sequence stars is important because they are the seeds from which planets may form. We have now strong evidence that grains in disks are very different from the grains in the diffuse interstellar medium and in the molecular clouds from which disks form, as reviewed, e.g., by Natta et al. (2006). In many objects, observations with millimeter interferometers have provided strong evidence that the grains in the outer and cooler regions of the disk (further than 50AU from the star) have been hugely processed, and have grown from sub-micron sizes to millimeter and centimeter ones. Closer to the star, however, in the regions were planets are more likely to form, observational evidence has been confined to grains close to the disk surface. For these, which however account for a tiny fraction of the total dust mass, emission in the silicate features has shown a correlation between the shape of the feature and its strength that is interpreted as due to growth of the grains from size $a\sim0.1\mu$m to $a\sim1\mu$m (van Boekel et al. 2003, 2004; Meeus et al. 2003). In this inner disk, the properties of the grains in the disk midplane are still unknown. In the last few years, due to the new long baseline near-infrared interferometers, many important steps forward in the study of the internal regions of circumstellar disks have occurred. The available near-infrared interferometric observations of T Tauri (TTS) and Herbig Ae (HAe) stars (Eisner et al. 2003, 2004; Millan-Gabet et al. 2001; Tuthill et al. 2001; Monnier et al. 2005) confirm the idea that the inner disk properties are controlled by the dust evaporation process which produce a ``puffed-up'' inner rim at the dust destruction radius (Natta et al. 2001; Dullemond, Dominick \& Natta 2001, hereafter DDN01). In these models, the location and shape of the rim depends on the properties of grains located not on the disk surface but on its midplane. Isella \& Natta (2005, hereafter IN05) have recently proposed models of the ``puffed-up''inner rim which include a self-consistent description of the grain evaporation and its dependence on the gas density. IN05 have explored a large range of grain properties, and discussed how the location of the rim depends on grain properties. In this paper, we will use the IN05 models to analyze the existing interferometric data of the best observed HAe stars to explore, in practice, the constraints on grain properties provided by this technique and their uncertainties. As a byproduct of the modeling process, one obtains also the orientation of the inner disk (i.e. its inclination with respect to the line of sight and its position angle); this can be compared with the orientation of the outer disk, obtained from millimeter observations of the molecular gas and dust emission and/or scattered light in the optical. The paper is organized as follows. In \S2 we describe the available interferometric observations of the target stars. The IN05 model for the inner rim is briefly summarized in \S3 and used to fit the observations of the individual objects in \S4. A comparison of the results with previous analysis of the same data is presented in \S5. Our results are discussed in \S6. Conclusions follow in \S7. \section{Target stars and observations} Our sample is composed of six HAe stars (AB Aur, CQ Tau, VV Ser, MWC 480, MWC 758 and V1295 Aql), for which near-infrared interferometric observations exist in the literature. Table \ref{tab.sources} summarizes the physical properties of the target stars. All the stars are classified as young stellar objects with masses ranging from 1.5 to 4.3 solar mass and a spectral type between A0/B9 and A8/F2. CQ Tau and VV Ser belong to the family of UXORs and are characterized by large and irregular variability. We use visibility measurements of the target stars from the literature, obtained with interferometric observations carried out with PTI (Palomar Testbed Interferometer) in K band ($\lambda_0=2.2\mu$m, $\Delta\lambda = 0.4\mu$m) described in Eisner et al. (2004). For AB Aur and V1295Aql, IOTA observations are also available (Millan-Gabet et al. 2001) for the K'($\lambda_0=2.16\mu$m, $\Delta\lambda = 0.32\mu$m) and H ($\lambda_0=1.65\mu$m, $\Delta\lambda = 0.30\mu$m) bands. \begin{table*} \centering \begin{minipage}[t]{12.5cm} \caption{Stellar parameters.} \label{tab.sources} \begin{tabular}{c c c c l l l l} % \hline\hline Source & Alternate Name & Spectral Type & d & T & L & M & Av \\ ~ & ~ & ~ & (pc) & (K) & (\Lsun) & (\Msun) & ~ \\ \hline AB Aur & HD 31293 & A0pe & 144 & 9772 & 47 & 2.4 & 0.5 \\ MWC 480 & HD 31648 & A2/3ep+sh & 140 & 8700 & 25 & 2.2 & 0.25 \\ MWC 758 & HD 36112 & A5IVe & 230 & 8128 & 22 & 2.0 & 0.22 \\ CQ Tau & HD 36910 & A8 V/F2 IVea & 100 & 8000 & 5 & 1.5 & 1.00 \\ VV Ser & HBC 282 & B9/A0 Vevp & 260 & 10200 & 49 & 3.0 & 3.6 \\ V1295 Aql & HD 190073 & B9/A0 Vp+sh & 290 & 8912 & 83 & 4.3 & 0.19 \\ \hline \end{tabular} \\\\ Stellar parameters are from Hillenbrand et al. (1992), van den Ancker et al. (1998), Chiang et al. (2001), Strizys et al. (1996), Mannings et al. (2000) and references therein. \end{minipage} \end{table*} \section{Model description} \label{sec:model} We use a model based on the assumption that the near-infrared emission of HAe stars originates in the ``puffed up'' inner rim which forms in the circumstellar disk at the dust evaporation radius (Natta et al 2001; DDN01). In IN05 we revised the concept of the ``puffed up''inner rim, introducing the dependence of the dust evaporation temperature on the local gas density, following the dust model of Pollack et al. (1994). The main result is that the surface of the rim presents a curved shape (see Fig.1), whose features are summarized in the following. \subsection{The dust evaporation and the ``puffed-up'' inner rim} \label{sec:mod_1} Following the suggestion of Natta et al. (2001), we assume that the dust component of a circumstellar accreting disk is internally truncated by the dust evaporation process, forming an inner hole of radius $R_{evp}$ inside of which only gas can survive. If the radiation absorption due to this inner gas is negligible (as is often the case; see, e.g., Muzerolle et al. 2004), dust evaporation occurs where the equilibrium temperature $T_d$ of grains embedded in the unattenuated stellar radiation field, equals their evaporation temperature $T_{evp}$. In IN05 we used an analytical solution of the radiation transfer problem (Calvet et al. 1991, 1992) to calculate the grain temperature inside the disk and we showed that evaporation occurs at a distance from the star that can be expressed as: \begin{equation} \label{eq:Revp} R_{evp}[AU] = 0.034 \cdot \left( \frac{1500}{T_{evp}} \right)^2 \sqrt{ \frac{L_{\star}}{L_\odot} \left( 2+\frac{1}{\epsilon} \right) }, \end{equation} where $L_{\star}$ is the stellar luminosity and $\epsilon$ is the ratio of the Planck mean opacity at $T_{evp}$ to that at the stellar effective temperature $T_\star$, $\epsilon=\kappa_P(T_{evp})/\kappa_P(T_{\star})$. The quantity $\epsilon$ measures the cooling efficiency of the grains; it depends on the wavelength dependence of the absorption efficiency of the grains and varies with grain composition and size. If the dust in the proto-planetary disk is composed of different types of grains, the location and structure of the inner rim depends on the properties of the grains with the highest evaporation temperature. In the dust model proposed by Pollack et al. (\cite{PH94}), the most refractory grains are silicates for which $T_{evp}$ is given by the relation \begin{equation} \label{eq:Tevp} T_{evp}(r,z) = 2000 \cdot [\rho_g(r,z)]^{0.0195}, \end{equation} valid for the gas density $\rho_g$ in the range between $10^{-18}g\, cm^{-3}$ and $10^{-5}g\, cm^{-3}$. In the following analysis, we will therefore assume that the inner disk dust is made of silicates, with optical properties given by Weingartner \& Draine (2001); thus, $\epsilon$ is uniquely defined by the grain radius $a$, and we will use $a$, rather than $\epsilon$, as a model parameter. Assuming that the proto-planetary disk is in hydrostatic equilibrium in the gravitational field of the central star and that it is isothermal in the vertical direction $z$, the gas density $\rho_g(r,z)$ has its maximum value on the midplane and decreases with $z$ as \begin{equation} \label{eq:rhog} \rho_g(r,z)=\rho_{g,0}(r) \exp(-z^2/2h(r)^2) , \end{equation} where $h$ is the pressure scale height of the disk. The midplane density can be expressed as a power-law of $r$ $\rho_{g,0}(r) = \rho_{g,0}(r_0) (r_0/r)^{\gamma}$, with $\gamma$ of the order of 2--3 (see, e.g., Chiang \& Goldreich 1997). The decrease of $\rho_g$ with $z$, combined with Eq.\ref{eq:Tevp}, implies that the silicate evaporation temperature varies from, i.e., $\sim1500K$ on the midplane (assuming a typical gas density of $\sim10^{-7}$g/cm$^3$) to $\sim 1000K$ at $z/h=6.4$ and $\sim 800K$ at $z/h=8$. Since $T_{evp}$ decreases with $z$ , it is immediately clear from Eq.\ref{eq:Revp} that the distance from the star at which dust evaporates increases with $z$, describing a curved surface as shown in Fig.\ref{fig:disk}. The dependence of $T_{evp}$ on the gas density is an important factor when computing the shape of the rim in the vertical direction, where the gas density varies by many orders of magnitude while the distance from the star is practically unchanged. In the radial direction, we expect relatively small variations of $\rho_{g,0}$, for any reasonable value of the disk mass, so that the distance of the rim from the star, measured in the midplane, is practically independent of the gas density. The emission of the rim is computed assuming that it originates from the surface characterized by an effective temperature $T_{eff}=T(\tau_d=2/3)$, where $\tau_d$ is the optical depth for the emitted radiation. The $T_{eff}$ surface, therefore, defines the observed location and shape of the rim. In IN05, we discussed how the $T_{evp}$ and the $T_{eff}$ surfaces behave for small and large silicate grains, and showed that the $T_{eff}$ surface moves closer to the star for increasing grain size until a critical value, which for silicates is about 1.2 \um. Larger grains produce rims with $T_{eff}$ surfaces practically independent of $a$. Therefore, for a fixed stellar luminosity, silicates with $a\sim 1.2$\um\ give the minimum value of the distance of the rim from the star. Conversely, if the measured rim distance is equal to this minimum value, one can derive from it only a lower limit ($\sim 1.2$\um) to the grain size. The rim emission peaks at near-infrared wavelengths. At $\lambda \simless 5-7$\um, one can assume that the observed flux is the sum of the stellar + rim emission, with only negligible contribution from the outer disk (see, e.g., DDN01). We model the stellar photospheric flux using standard Kurucz model atmospheres. \subsection{Visibility model} Due to the limited coverage of the $u-v$ plane of the existing near-infrared interferometers, it is not possible at present to recover full images from the available data, and one has to resort to the analysis of the visibilities on given interferometric baselines. Starting from the synthetic images of the inner rim (see IN05), we compute the predicted visibility values using a Fast Fourier Transform recipe. For face-on inclination, due to the circular symmetry of the rim image, the visibility depends only on the length of the baseline $B$. For inclination greater than zero, the image of the rim has an ``elliptical'' shape: the minor axis decreases with increasing inclination and the upper half of the rim becomes brighter than the lower part. For baselines oriented along the direction of the minor axis of the rim image, the visibility decreases more slowly than for those oriented along the major axis. For all other orientations of the baseline, the visibility will have values intermediate between these two (see Fig.\ref{fig:vis}). Moreover, due to the Earth rotation during the observation, the baseline corresponding to a fixed telescope pair moves in the u-v plane describing an ellipse. Along this ellipse, each point is related to the hour angle $HA$ of the target object in the sky. In the next section we use the $V^2$-$B$ plot and $V^2$-$HA$ plot, to show how the models fit the observations. The visibility model takes into account the emission of the central star, modeled as a uniform disk of radius $R_{\star}$. If $F_{\star}$ and $F_{d}$ are respectively the stellar and the inner rim flux at the wavelength of the observation, the total visibility is given by the relation: \begin{equation} V^2 = \left( \frac{F_\star V_\star + F_d V_d}{F_\star + F_d} \right) ^2, \end{equation} where $V_\star$ and $V_d$ are the visibility values of the star and of the disk. Note that for the PTI configuration (baselines between 84m and 100m) $V_\star$ is in all cases very closed to 1. \section {Comparison with the observations} \label{sec:fit} \subsection{Model parameters} Once the stellar and dust properties are fixed, the model-predicted visibilities depend on the dust grain radius $a$, which completely defines the rim structure, and two parameters (observational parameters in the following) that describe the orientation of the disk, namely the inclination $\iota$ and position angle PA. For each star, we firstly compute the predicted rim structure varying the size of the grains from very small to very large values. As discussed in \S3, we assume that silicates are the most refractory component; we take the optical properties of astronomical silicates defined by Weingartner \& Draine (\cite{WD01}). Other disk parameters (i.e., mass and density radial profile) can be neglected in this analysis. We fix $\rho_{g,0}(R_{rim})\sim 10^{-7}$ g/cm$^3$ which gives a total disk mass of about 0.1\Msun, for a fiducial value of $\gamma=2.5$ and an outer disk radius of the disk of 200AU. Once the structure of the ``puffed up'' inner rim is calculated, the predicted visibility depends on the orientation of the disk in the sky, defined by the inclination $\iota$ of the midplane of the disk with respect to the line of sight and its position angle $PA$, measured from north to east and relative to the major axis of the projected image of the disk on the sky. The inclination is defined so that $\iota=0^{\circ}$ identifies a face-on disk while $\iota=90^{\circ}$ corresponds to an edge-on disk. For inclinations higher than 80$^\circ$ the rim emission is likely absorbed by the outer regions of the disk and the IN05 model can not be applied. In practice, we compute visibility models for each object varying $a$, $\iota$ and $PA$ independently. We then select the best models calculating the reduced $\chi^2$ between the visibility data and the theoretical values calculated at the same points in the u-v plane. The observed near-infrared fluxes, and the IOTA data when available, are then ``a posteriori'' used to check the quality of the fit and, when possible, to reduce the degeneracy due to the small number of visibility points. For some stars, the existing data do not constrain the parameters, but still define a range, outside of which the fit to the data is very poor. Tab.\ref{tab:fit} shows in column 2 the best values of the astronomical silicate radius $a$ and, in column 3, the corresponding values of the radius of the inner rim $\Rrim$. The two observational parameters $i$ and $PA$ are given in columns 4 and 5, respectively. Note that the free parameters are in boldface; $\Rrim$ is a derived quantity. \begin{table*} \centering \begin{minipage}[t]{17.5cm} \caption{Best fitting model parameters. } \label{tab:fit} \begin{tabular}{l|l l l l | l l l | l l } \hline\hline ~ & \multicolumn{4}{c}{IN05 model} & \multicolumn{3}{c}{Eisner et al. (2004)} & \multicolumn{2}{c}{outer disk} \\ Source &{\boldmath $a$ } & {$\Rrim$} & {\boldmath $\iota$} & {\bf PA}& $\Rrim$ & $\iota$ & PA & $\iota$ & PA \\ ~ &($\mu$m) & (AU) & (deg) &(deg) & (AU) & (deg) &(deg) & (deg) & (deg) \\ \hline MWC 758 &{\boldmath $\geq 1.2$} & 0.32 & {\bf 40} & {\bf 145} & 0.21 & $36^{+3}_{-2}$ & $127^{+4}_{-3}$ & 46 & $ 116^{+6 \, (a)}_{-5}$\\ VV Ser & {\boldmath $\geq 1.2$} & 0.54 & {\boldmath$50-70$} & {\boldmath$60-120$}& 0.47 & $42^{+6}_{-2}$ & $166^{+17}_{-6}$& $72\pm5$ & $13\pm5 ^{(b)}$ \\ CQ Tau & {\boldmath $0.3 - \geq 1.2$} & $0.16 - 0.25$ & {\boldmath $40 - 55$} & {\boldmath$145- 190$} & 0.23& $48^{+3}_{-4}$ & $ 106^{+4}_{-5}$ &$63^{+10}_{-15}$ & $2\pm13^{(c)}$ \\ V1295 Aql & {\boldmath $0.3 - \geq1.2$} & $0.7 - 1.2$ & {\boldmath$40 - 65$} & &0.55 & $23^{+15}_{-23}$ & & ~ &\\ MWC 480 & {\boldmath $0.2 - 0.3$} & $0.53 - 0.63$ & {\boldmath$30 - 65$} & ~& 0.23 & $28^{+2}_{-1}$ & $145^{+9}_{-6}$ & $ 20 - 40$ &$147 - 180^{(a,d)}$ \\ AB Aur & \multicolumn{3}{c}{impossible to fit} && 0.25 & $8^{+7}_{-8}$ & &$15 - 35$ &$50 - 110 ^{(e,f,g,h)}$ \\ \hline \end{tabular} \\\\ From column 2 to 5 are reported the best fit parameters for the ``puffed-up'' inner rim, obtained with the IN05 model: the grain radius $a$, the radius of the inner rim $\Rrim$, the inclination $\iota$ and the position angle $PA$. The free parameters of the model are presented in bold face. Columns 6,7 and 8 show the values of the radius of the inner rim, the inclination and the position angle, obtained by Eisner et al. (2004). Finally, the last two columns show the available estimates of inclination and position angle for the external region of the disk: $(a)$ Mannings et al. (1997); $(b)$ Pontoppidan et al. (2006); $(c)$ Testi et al. (2001, 2003); $(d)$ Simon et al. (2000); $(e)$ Fukagawa et al.(2004); $(f)$ Grady et al. (1999); $(g)$ Corder et al. 2005; $(h)$ Pi\'etu et al. (2005). \end{minipage} \end{table*} \subsection{MWC 758} \label{sec:MWC758} PTI visibilities are fitted by a family of models, with parameters varying between the two extreme cases shown in Fig.\ref{fig:MWC758_vis2}. In one case, the disk has small grains of radius $a=0.17$ $\mu$m, $\iota= 48^{\circ}$ and $PA=134^{\circ}$; in the other, big grains with $a \geq 1.2 \mu$m, $\iota =40^{\circ}$ and $PA= 145^{\circ}$. Models with $a$ values within this range will fit the observed visibilities equally well, provided that we vary $\iota$ and $PA$ in an appropriately way. However, if we consider also the constraints set by the SED at near-infrared wavelengths, we find that only models with big grains fit it reasonably well (see the right panel of Fig.\ref{fig:MWC758_vis2}). The best-fitting model $(\chi^2_r=2.0)$ has then $a \geq 1.2$\um, inner rim radius is $\Rrim=0.32$AU, rim effective temperature (at $z$=0) is 1460K. The near-infrared flux, $L_{NIR}$, integrated between 2$\mu$m and 7$\mu$m is $25\%$ of the total stellar luminosity, similar to the observed value. Once we fix $a$, the formal uncertainties on $\iota$, estimated from the surface where the reduced $\chi^2$ equals $\chi^2_{min}+1$, are quite small, $\pm 3^{\circ}$. More realistic uncertainties are of the order of $10^{\circ}$ for both $\iota$ and $PA$. Note that in MWC 758 the PTI visibilities define quite well the orientation of the disk, even when the SED is not used to constrain the grain size. In particular, the inclination cannot be lower than about $30^\circ$. \subsection{VV Ser} The results for the star VV Ser are shown in Fig.\ref{fig:VVSer_HA}. As for MWC 758, the interferometric observations allow different sets of model parameters. Namely, we obtain similar values of the reduced $\chi^2$ $(\sim1.2)$ for all grain sizes $a\simgreat 0.4$\um. Over this range of $a$, inclination and position angle vary in the intervals $45^\circ - 80^\circ$ and $60^\circ - 120^\circ$, respectively, with lower inclinations for larger grains. The correlation between $\iota$ and $PA$ is very strong, and the uncertainties in these two parameters remain very large even for fixed $a$. Although the fit is never very good, the VV Ser SED is better accounted for by large grains (see the right panel of Fig.\ref{fig:VVSer_HA}) and in Table 2 we show the best values of the parameters for $a\geq 1.2$\um. The rim effective temperature is 1400 K, the near-infrared excess is 21\% of \Lstar. \subsection{CQ Tau} The limited number of visibility points does not allow us to constrain all the parameters of the disk. Fig.\ref{fig:CQTau_HA} shows two models, with the same level of confidence $(\chi^2_r \sim 1)$; the two models have similar orientations (inclinations of 52$^{\circ}$ and 46$^{\circ}$ with position angles of 168$^{\circ}$ and 164$^{\circ}$ respectively) but very different grain sizes ($a=0.3\mu$m and $a \geq 1.2 \mu$m) and radii of the inner rim (0.25AU and 0.16AU, respectively). For $a \geq 1.2 \mu$m, the effective temperature of the rim (at z=0) is $T_{eff}=1480K$, while $T_{eff}=1050$ for $a=0.3\mu$m. The SEDs of the two models are compatible with the observed fluxes, with $L_{NIR}=13\%-18\%$ \Lstar. All the models with $a$ within this range, and similar $\iota$ and $PA$, have similar $\chi^2$ values. Outside this range, models give a much poorer fit to the data. As for MWC 758, the visibility data constrain well the orientation of the disk on the sky. The inclination, in particular, has to be quite large, $40^{\circ} \simless \iota \simless 55^{\circ}$. \subsection{V1295 Aql} \label{sec:V1295Aql} The PTI observations of V1295 Aql are characterized by a very small number of visibility points and the disk parameters are hardly constrained. Even adding the IOTA data does not help due to the big errors that affect these observations. Note also that PTI and IOTA observations are performed at different wavelengths, K and H respectively. As for CQ Tau, we show the models for the two extreme sets of parameters that give an equally good fit (Fig.\ref{fig:V1295_HA}) with $\chi^2_r \sim 1$. All the intermediate combinations of $a$, $\iota$ and PA can explain the observations as well. In order to fit the values of visibility, an inclination ranging between $40^\circ$ and $65^\circ$ is required. More face-on systems can not in general reproduce the visibility spread in the IOTA data and the $V^2-HA$ behaviour of the PTI points. The grain radius varies from $a\geq1.2\mu$m ($\Rrim=0.7$AU, $T_{eff}=1400K$) to $a=0.3\mu$m ($\Rrim=1.2$AU, $T_{eff}=970K$) while the position angle can not be defined at all. The right panel of Fig.\ref{fig:V1295_HA} shows the comparison between the observed and the predicted SED. More inclined disks can in general reproduce better the photometric measurements around 1.5$\mu$m. The near-infrared emission of disk with an inclination of less than 50$^\circ$ is peaked at about $4\mu$m and can not reproduce the infrared excess between 1$\mu$m and 2$\mu$m; in both models, the integrated near-infrared flux is $L_{NIR}\sim 20\%$ \Lstar. \subsection{MWC 480} Also in this case, due to the narrow range of available baselines, the PTI visibilities of MWC 480 are consistent with different sets of parameters at the same level of confidence $(\chi^2_r\sim2)$. Fig.\ref{fig:MWC480_vis2} shows the two extreme disk configurations characterized by quite similar parameters for the dust grain size ($a=0.2\mu$m$-0.3\mu$m; $\Rrim=0.63$AU-0.53AU and $T_{eff}\simeq 1250K$), but very different values of the inclination ($\iota=35^{\circ}$ and $\iota=60^{\circ}$) and the position angle ($\psi=60^{\circ}$ and $\psi=168^{\circ}$). Several intermediate configurations reproduce the observed data as well. The degeneracy can not be removed even using the SED (Fig.\ref{fig:MWC480_vis2}) which is similar in the two models ($L_{NIR}/L_{\star} = 0.18\%-0.14\%$) and it is only roughly consistent with the photometric values. \subsection{AB Aur} \label{sec:fitABAur} AB Aur is the only HAe star that cannot be fitted with the IN05 models. The PTI visibilities require a face-on rim, consistent with the inclination derived from large-scale images in scattered light (Grady et al.~1999; Fukagawa et al.~2004) and at millimeter wavelengths (Corder et al.~2005; Pi\'etu et al.~2005). For these inclinations, the $V^2$ data imply a very small inner radius, about two times smaller that the smallest $R_{rim}$ obtained using the IN05 model (see Fig.\ref{fig:ABAur_vis2}). If, to put the discrepancy in a more physical context, we take $T_{evp}$ as a free parameter, we find good agreement with the PTI data for $T_{evp}\sim 2800K$ (dashed line), a value by far too high not only for silicates but also for any other type of grains (e.g., Pollack et al.~1994). The situation becames even less clear if we consider also the IOTA observations (squared points), since they seem to indicate the presence of a more inclined disk with an inner radius between 0.26AU and 0.51AU. No additional information can be obtained from the analysis of the spectral energy distribution, since all the inner rim models with an effective temperature between 1500K and 2500K are compatible with the photometric data. We will come back to AB Aur in \S 6. \section {Comparison with previous analysis} Fits to the same interferometric data analyzed in \S\ref{sec:fit} have been obtained by Eisner et al. (2004, hereafter E04) assuming a toroidal shape for the inner ``puffed-up'' rim, based on the simplified DDN01 model. In these fits, the free parameters are the location of the rim $\Rrim$ and the two observational parameters, $\iota$ and $PA$. The E04 results are shown in Table 2. We note that for three objects (MWC 758, VV Ser and CQ Tau) the E04 inclinations are in agreement within the errors with the values obtained with the IN05 model, while for the other two (V1295 Aql and MWC 480) the E04 $\iota$ estimates are consistent with the lowest value of the range derived in this paper. The largest differences are in the derived values of $R_{rim}$: the IN05 inner radii are always larger than E04 results, with a maximum difference of a factor $\sim 3$ if we consider our maximum $\Rrim$ in the MWC 480 system. While for CQ Tau the two values are almost the same, for all the other stars the difference is a factor 1.5 and 2. This discrepancy is mainly due to the difference between IN05 and E04 models. In particular, in IN05, the curved shape of the emitting surface is self-consistently calculated, allowing a more correct determination of the dependence of the rim emission on the inclination of the disk. Moreover, the IN05 model takes into account the effect of the radiation transport within the disk, even if in an approximate way (see Appendix A in IN05). This supplementary heating is neglected by E04, who calculate the dust temperature taking into account only the direct stellar radiation. The ratio between the two values of $\Rrim$ is given by the relation \begin{equation} \label{eq:R/R} \frac{ \Rrim }{ \hat{R}_{rim} } = \sqrt{\hat{\epsilon} \left(2 + 1/\epsilon \right) }, \end{equation} where the $\hat\epsilon$, $\hat{R}_{rim}$ and $\hat{T}_{evp}^2$ are the values used by E04 and the inner radius $\hat{R}_{rim}$ is given by the relation \begin{equation} \label{eq:R04} \hat{R}_{rim} = \frac{1}{\hat{T}_{evp}^2} \sqrt{ \frac{L_{\star}}{4\pi\sigma} {\frac{1}{\hat{\epsilon} } } }. \end{equation} Assuming the same value of the dust emissivity ($\epsilon=\hat{\epsilon}$), the ratio $\Rrim/\hat{R}_{in}$ is $\sim1$ for $\epsilon<<1$. The difference increases for larger $\epsilon$ and is maximum when $\epsilon$ and $\hat\epsilon$ are very different. Finally, there also differences due to the fact that E04 assumes $T_{evp}$ (or, equivalently, $\Rrim$) as a free parameter, while in the IN05 model $T_{evp}$ is self-consistently determined starting from the choice of the type of grains and the gas density in the disk (see Eq.\ref{eq:Tevp}). For all our target stars, the resulting values of $T_{evp}$ vary between 1370K and 1460K, and are in some cases significantly different from those given by E04. \section{Discussion} The results presented in \S\ref{sec:fit} show that, with the exception of AB Aur, the IN05 self-consistent models of the ``puffed-up'' inner rim can explain the available observations, both visibilities and SEDs, of HAe stars. They can be used to derive information about the properties of the dust present in the innermost region of the circumstellar disk, the location of the inner rim and the orientation of the disk on the sky, using a minimum number of assumptions. \subsection{Presence of large grains} \label{sec:large} As shown in \S\ref{sec:fit}, the IN05 models reproduce the interferometric data under the assumption that the most refractory dust in the inner disk is made of silicates, with properties typical of astronomical silicates (Weingartner \& Draine \cite{WD01}). In four cases, grain sizes larger than $\sim 1.2$\um\ are either required by or consistent with the observations. Only in one case are the data better fitted with $a \sim 0.2-0.3\mu$m. Grains in the rim are thus larger, and often much larger, than grains in the interstellar medium ($a=0.01-0.1\mu$m, Weingartner \& Draine \cite{WD01}), confirming that grain growth has taken place in the innermost disk regions (van Boekel et al. 2004). Even if the predicted near-infrared excess agrees well with the photometric observations, some interesting differences exist between the theoretical and the observed spectral energy distributions. With the exception of CQTau, the predicted SEDs always peak at a wavelength slightly longer than found in the observations: the flux at short wavelengths (between 1.5\um\ and 2.2\um) is thus generally underestimated while the flux between 2.2\um\ and 7\um\ is overestimated. This may be due to the fact that in our models the SED is computed assuming that each point on the surface of the rim emits as a black body at the local effective temperature. This approximation is energetically correct but may not reproduce the exact wavelength dependence of the emitted radiation (see Appendix in IN05). The rim models fail only in the case of AB Aur, where silicates, of whatever size, produce rims that are too distant from the star to be consistent with the observations. As shown in \S 4.7, PTI and IOTA data give somewhat contradictory results, and more interferometric data are clearly required. However, unless further observations drastically change the present picture, the discrepancy between the rim model predictions and the data is highly significant, and some of the basic underlying assumptions need to be changed. It is possible that in AB Aur grains more refractory than silicates dominate the dust population in the inner disk; however, the $T_{evp}$ required ($\sim 2800$K) is too high for any dust species likely present in disks (e.g., Pollack et al. 1994). It is more likely that gas in the dust-depleted inner region absorbs a significant fraction of the stellar radiation, shielding the dust grains which are therefore cooler than in our models. This requires a high gas density in the inner disk, as expected if the accretion rate is high; in general, the accretion rates of HAe stars (including AB Aur) are low enough to ensure that the gaseous disk remains optically thin (Muzerolle et al.~2004). However, our knowledge of the accretion properties and gas disks of HAe stars is very poor, and should be investigated further. AB Aur may be more than just an oddity. The presence of optically thick gas inside the inner rim has been proposed to explain the near-infrared interferometric observations of some very bright Herbig Be stars, for which the visibility data suggest inner disk radii many times too small to be consistent with the ``puffed-up'' rim models (Malbet et al. 2005, Monnier et al. 2005). If this is the case, AB Aur could be the low-luminosity tail of the same phenomenon, which, given its small distance and large brightness, could be used to understand a whole class of objects. \subsection{Inclination and position angle} Inclination and $PA$ are well constrained by the existing data only in one case (MWC~758). We want to stress, however, that values of the parameters outside the ranges given in Table 2 do not fit the data at all. In particular, there are no objects consistent with face-on disks, or in general with a centro-symmetric brightness distribution. This rules out models, such as those of Vinkovic et al. (2005), where most of the near-infrared flux is contributed by a spherically symmetric shell around the star, rather than by a circumstellar disk. Two stars (CQ Tau and VV Ser) belong to the group of UXOR variables, which are interpreted as objects with disks seen close to edge-on (Grinin et al. 2001, Natta \& Whitney 2001, Dullemond et al. 2003). We derive for them large inclinations, in agreement with this interpretation. For some of our targets, there are in the literature estimates of the orientation of the outer disk on the plane of the sky obtained with millimeter interferometers (Testi el al. 2001, 2003; Manning \& Sargent 1997; Pi\'etu et al.~2005; Corder et al.~2005). These determinations refer to the outer disk, i.e., to spatial scales of 50--100 AU at least. The comparison with the values derived in the near-infrared for the inner disk (on scales of less than 1 AU) can provide information on possible distortions in the disk, e.g. variations of the inclination with radius. For the three disks with millimeter data (MWC 758, CQ Tau and MWC 480), there is agreement (within the uncertainties) between the inclination obtained by infrared and millimeter observations However, it is certainly premature to exclude the existence of disk distortions, given the large uncertainties that affect both the millimeter and the near-infrared estimates. More accurate interferometric observations in the two wavelength ranges and self-consistent models of the disk at all physical scales are required. An interesting case is that of VV Ser, whose disk has recently been imaged as a shadow seen against the background emission in the 11.3 PAH feature (Pontoppidan et al. 2006); These authors derive an inclination (of the outer disk) of about 70$^{\circ}$ and a position angle of $13^\circ \pm 5^\circ$. While the inclination is consistent with the upper limit of the range we obtain for the inner disk, the position angle is off by almost $90^\circ$. This discrepancy is intriguing, and deserves further investigation. \subsection{Improving the model constrains} \label{sec:Const} Near-infrared interferometric observations of disks around pre-main sequence stars are still few and sparse. It is clear from our analysis that even in the most favorable cases more visibility data at different baselines are necessary to narrow the range of possible disk inclinations and grain properties. Given the huge demands of telescope time that interferometric observations require, it is useful to make use of model predictions in preparing the observations and in choosing the baseline configurations that can constrain the disk structure. CQ Tau represents an example of how the IN05 model can be used in this context. To better constrain the inner rim structure, one will need observations with baselines longer than 130m (available with the VLT interferometer), for which the predicted values of the squared visibility parameters are very different for different disk models (see Fig.\ref{fig:CQTau_HA}). On the other hand, observations with baseline shorter than 60m could better constrain the inner radius of MWC 480, since a degeneracy in the models is present at longer baselines. V1295 Aql represents a still different case, in which the degeneracy in the values of the predicted squared visibility can be removed observing at distant hour angles for the same baseline configurations, in order to determine the visibility variations at the same baseline, due to the inclination of the disk. \section{Summary and conclusions} In this paper we have analyzed the near-infrared interferometric observations of the six best observed HAe stars using the rim models developed by Isella \& Natta (2005). Our aim was to explore the potential of near-infrared interferometry to constrain the properties of the grains in the inner disks of these stars. The basic assumptions of the IN05 rim models are that the inner disk structure is controlled by the evaporation of dust in the unattenuated stellar radiation field, as expected if the gaseous disks have low optical depth, and that the most refractory grains are silicates. The IN05 self-consistent models for the ``puffed-up'' inner rim reproduce both the interferometric observations and the near-infrared spectral energy distribution of all the objects we have studied, with the exception of AB Aur, which we have briefly discussed. For the five stars where we are able to obtain a good fit to the data, we can estimate the grain sizes in the rim, i.e., in the midplane of the inner disk. We find that in four cases grains larger than $\sim 1.2$\um\ are either required by or consistent with the data. Only in one case do we find that the existing data require $a\sim 0.2-0.3$\um. Note that this value of $a=1.2$\um\ is a lower limit to the grain size: grains can be much larger, since the rim location and shape do not change significantly if the grains grow further. As a result of the model-fitting, one derives also the inclination and position angle of the disk on the plane of the sky. We find that, in general, these parameters are not well constrained by the existing data. However, in all cases we can fit, inclinations lower than $30^{\circ}$ are not consistent with the observations and the surface brightness distribution can not be circularly symmetric. This rules out a spherical envelope as the dominant source of the near-infrared emission. For some objects, estimates of the inclination of the outer disk have been obtained from millimeter interferometric observations; within the uncertainties, they agree with the values obtained for the inner disk. Our analysis shows that near-infrared interferometry is a very powerful tool for understanding the properties of the inner disks, in particular when combined with physical models of these regions. However, at present the existing data are for many objects still too sparse in their coverage of the u-v plane to allow an accurate determination of the disk parameters. We expect that this will be improved in the future. In this context, since near-infrared interferometry is and will remain a very time demanding technique, we stress the importance of using physical models of the inner region of the disk in planning future observations. \begin{acknowledgements} We are indebted to Josh Eisner and Rafael Millan-Gabet for providing us with the PTI and IOTA data. The authors acknowledge partial support for this project by MIUR PRIN grant 2003/027003-001. \end{acknowledgements}
Title: Evolution of the Color-Magnitude Relation in High-Redshift Clusters: Blue Early-Type Galaxies and Red Pairs in RDCS J0910+5422
Abstract: The color-magnitude relation has been determined for the RDCS J0910+5422 cluster of galaxies at redshift z = 1.106. Cluster members were selected from HST ACS images, combined with ground--based near--IR imaging and optical spectroscopy. The observed early--type color--magnitude relation (CMR) in (i_775 -z_850) versus z_850 shows intrinsic scatters in color of 0.042 +/- 0.010 mag and 0.044 +/- 0.020 mag for ellipticals and S0s, respectively. From the scatter about the CMR, a mean luminosity--weighted age t > 3.3 Gyr (z > 3) is derived for the elliptical galaxies. Strikingly, the S0 galaxies in RDCS J0910+5422 are systematically bluer in (i_775 - z_850) by 0.07 +/- 0.02 mag, with respect to the ellipticals. The ellipticity distribution as a function of color indicates that the face-on S0s in this particular cluster have likely been classified as elliptical. Thus, if anything, the offset in color between the elliptical and S0 populations may be even more significant. The color offset between S0 and E corresponds to an age difference of ~1 Gyr, for a single-burst solar metallicity model. A solar metallicity model with an exponential decay in star formation will reproduce the offset for an age of 3.5 Gyr, i.e. the S0s have evolved gradually from star forming progenitors. The early--type population in this cluster appears to be still forming. The blue early-type disk galaxies in RDCS J0910+5422 likely represent the direct progenitors of the more evolved S0s that follow the same red sequence as ellipticals in other clusters. Thirteen red galaxy pairs are observed and the galaxies associated in pairs constitute ~40% of the CMR galaxies in this cluster.
https://export.arxiv.org/pdf/astro-ph/0601327
command. \newcommand{\vdag}{(v)^\dagger} \newcommand{\myemail}{smei@pha.jhu.edu} \def\etal{et~al.} \def\hst{{\it HST}} \def\vi{\ifmmode(V{-}I)\else$(V{-}I)$\fi} \def\viz{\ifmmode(V{-}I)_0\else$(V{-}I)_0$\fi} \def\gz{\ifmmode(g_{475}{-}z_{850})\else$(g_{475}{-}z_{850})$\fi} \def\gzz{\ifmmode(g_{475}{-}z_{850})_0\else$(g_{475}{-}z_{850})_0$\fi} \newcommand\lta{\mathrel{\rlap{\lower 3pt\hbox{$\mathchar"218$}} \raise 2.0pt\hbox{$\mathchar"13C$}}} \newcommand\gta{\mathrel{\rlap{\lower 3pt\hbox{$\mathchar"218$}} \raise 2.0pt\hbox{$\mathchar"13E$}}} \def\mbari{\ifmmode\overline{m}_I\else$\overline{m}_I$\fi} \def\mbarz{\ifmmode\overline{m}_z\else$\overline{m}_z$\fi} \def\mbar{\ifmmode\overline{m}\else$\overline{m}$\fi} \def\Mbar{\ifmmode\overline{M}\else$\overline{M}$\fi} \def\Mbarz{\ifmmode\overline{M_z}\else$\overline{M}_z$\fi} \shorttitle{CL0910} \shortauthors{Mei et al.} \tolerance=100000000 \begin{document} \title{Evolution of the Color-Magnitude Relation in High-Redshift Clusters: Blue Early-Type Galaxies and Red Pairs in RDCS~J0910+5422} \author{S. Mei\altaffilmark{1}, J. P. Blakeslee\altaffilmark{1}, S. A. Stanford\altaffilmark{2,3}, B. P.~Holden \altaffilmark{4}, P. Rosati\altaffilmark{6}, V. Strazzullo\altaffilmark{20,6}, N.~Homeier\altaffilmark{1}, M. Postman\altaffilmark{1,5}, M. Franx\altaffilmark{12}, A. Rettura\altaffilmark{6}, H. Ford\altaffilmark{1}, G. D. Illingworth \altaffilmark{4}, S. Ettori\altaffilmark{19}, R.J.~Bouwens\altaffilmark{4}, R. Demarco\altaffilmark{1}, A.R. Martel\altaffilmark{1}, M. Clampin\altaffilmark{5}, G.F. Hartig\altaffilmark{5}, P. Eisenhardt\altaffilmark{7}, D.R.~Ardila\altaffilmark{1}, F. Bartko\altaffilmark{8}, N. Ben\'{\i}tez\altaffilmark{18}, L.D. Bradley\altaffilmark{1}, T.J. Broadhurst\altaffilmark{9}, R.A. Brown\altaffilmark{5}, C.J. Burrows\altaffilmark{5}, E.S. Cheng\altaffilmark{10}, N.J.G. Cross\altaffilmark{17}, P.D. Feldman\altaffilmark{1}, D.A. Golimowski\altaffilmark{1}, T. Goto\altaffilmark{1}, C. Gronwall\altaffilmark{13}, L. Infante\altaffilmark{14} R.A. Kimble\altaffilmark{11}, J.E. Krist\altaffilmark{5}, M.P. Lesser\altaffilmark{15}, F. Menanteau\altaffilmark{1}, G.R. Meurer\altaffilmark{1}, G.K. Miley\altaffilmark{12}, V. Motta\altaffilmark{14}, M. Sirianni\altaffilmark{5}, W.B. Sparks\altaffilmark{5}, H.D. Tran\altaffilmark{16}, Z.I.~Tsvetanov\altaffilmark{1}, R.L. White\altaffilmark{5}, \& W. Zheng\altaffilmark{1}} \altaffiltext{1}{Dept.\ of Physics \&Astronomy, Johns Hopkins University, Baltimore, MD 21218; smei@pha.jhu.edu} \altaffiltext{2}{Department of Physics, University of California, Davis, CA 94516} \altaffiltext{3}{Institute of Geophysics and Planetary Physics, Lawrence Livermore National Lab, Livermore, CA 94551} \altaffiltext{4}{Lick Observatory, University of California, Santa Cruz, CA 95064} \altaffiltext{5}{Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218} \altaffiltext{6}{European Southern Observatory, Karl-Schwarzschild-Str. 2, D-85748 Garching, Germany} \altaffiltext{7}{Jet Propulsion Laboratory, CalTech, 4800 Oak Grove Drive, Pasadena, CA 91125} \altaffiltext{8}{Bartko Science \& Technology, 14520 Akron Street, Brighton, CO 80602.} \altaffiltext{9}{School of Physics and Astronomy, Tel Aviv University, Tel Aviv 69978, Israel} \altaffiltext{10}{Conceptual Analytics, LLC, 8209 Woburn Abbey Road, Glenn Dale, MD 20769.} \altaffiltext{11}{NASA Goddard Space Flight Center, Code 681, Greenbelt, MD 20771.} \altaffiltext{12}{Leiden Observatory, Postbus 9513, 2300 RA Leiden, Netherlands.} \altaffiltext{13}{Dept.\ of Astronomy \& Astrophysics, Penn State University, University Park, PA 16802.} \altaffiltext{14}{Dept.\ de Astronom\'{\i}a y Astrof\'{\i}sica, Pontificia Universidad Cat\'{\o}lica, Casilla 306, Santiago 22, Chile.} \altaffiltext{15}{Steward Observatory, University of Arizona, Tucson, AZ 85721.} \altaffiltext{16}{W. M. Keck Observatory, 65-1120 Mamalahoa Hwy., Kamuela, HI 96743} \altaffiltext{17}{Royal Observatory Edinburgh, Blackford Hill, Edinburgh, EH9 3HJ, UK} \altaffiltext{18}{Instituto de Astrof\'\i sica de Andaluc\'\i a (CSIC), Camino Bajo de Hu\'etor 50, Granada 18008, Spain } \altaffiltext{19}{INAF - Osservatorio Astronomico, via Ranzani 1, 40127 Bologna, Italy } \altaffiltext{20}{Dipartimento di Scienze Fisiche, Universit\`a Federico II, I-80126 Napoli,Italy} \keywords{galaxies: clusters: individual (RDCS~J0910+5422) -- galaxies: elliptical and lenticular --- galaxies: evolution} \section{Introduction} The Advanced Camera for Surveys (ACS; Ford et al. 2002), by virtue of its high spatial resolution and sensitivity, allows us to study galaxy clusters in great detail up to redshifts of unity and beyond. At these redshifts, galaxy clusters are still assembling and galaxies are evolving towards the populations that we observe today. Recent results from our ACS Intermediate Redshift Cluster Survey (Blakeslee et al. 2003a; Lidman et al. 2004; Demarco et al. 2005; Goto et al. 2005; Holden et al. 2005a; Holden et al. 2005b; Homeier et al. 2005; Postman et al. 2005) have shown that galaxy clusters at redshift around unity show many similarities with local clusters, in terms of galaxy populations and their distribution, but also significant differences in galaxy morphology, ellipticity, and mass--luminosity ratios. The strongest evolution observed in the early--type population is a deficit of a S0 population in this sample when compared to lower redshift samples (Postman et al. 2005). This would give evidence that the formation of the S0 population is still under way in clusters at redshift unity. One of the most striking similarities is that the tight relation between early-type galaxy colors and luminosities that applies locally (the color--magnitude relation; CMR) is already in place at redshifts as high as $z \sim 1.3$ (e.g. Stanford et al.\ 1997; Mullis et al.\ 2005). The CMR in local samples of galaxy clusters presents universal properties, in terms of scatter and zero point (Bower et al. 1992; van Dokkum et al. 1998, Hogg et al. 2004; L\'opez--Cruz et al. 2004; Bell et al. 2004; Bernardi et al. 2005; McIntosh et al. 2005) that evolve back in time in agreement with passively evolving models (Ellis et al.\ 1997; Stanford, Eisenhardt, \& Dickinson 1998; van Dokkum et al. 2000, 2001; Blakeslee et al. 2003a; Holden et al. 2004; De Lucia et al. 2004; Blakeslee et al. 2005). ACS enables accurate measurement of the scatter around the CMR, with enough precision to seriously constrain galaxy formation age, which is impossible to obtain from ground--based data (see for example Holden et al. 2004). The measurement of the CMR scatter of the first cluster in our ACS cluster survey, RXJ1252.9-292, permitted us to constrain the mean luminosity--weighted age for the ellipticals to be $> 2.6$~Gyr ($z > 2.7$) (Blakeslee et al. 2003a), based on simple modeling. In this paper, we extend the results obtained in Blakeslee et al. (2003a) to RXJ~0910+5422. RXJ~0910+5422 is part of the ACS cluster survey (guaranteed time observation, GTO, program \#9919), that includes eight clusters in the redshift range at $0.8 < z < 1.3$, selected in the X--ray, optical and near--IR (Ford et al.\ 2004). RXJ~0910+5422 was selected from the ROSAT Deep Cluster Survey (Rosati et al. 1998) and confirmed with near-IR and spectroscopic observations by Stanford et al. (2002). Extensive followup spectroscopy at the Keck Observatory has been carried out in a magnitude limited sample reaching $K_s = 20.0$~mag in the central $3$~arcmin (Stanford et al.\ 2005; in preparation). The mean redshift of the cluster was measured to be $z=1.106$ (Stanford et al. 2002). In this paper, we combine ACS imaging with ground--based spectroscopy and near--IR imaging to constrain galaxy ages and formation histories from the study of their color--magnitude relation. We discuss the properties of the elliptical (E) and lenticular (S0) populations separately in the light of simple galaxy formation scenarios. \section{Observations} RXJ~0910+5422 was observed in March 2004 with the ACS WFC (Wide Field Camera) in the F775W ($i_{775}$) and F850LP ($z_{850}$) bandpasses, with total exposure times of 6840~s and 11440~s, respectively. The ACS WFC scale is 0.05\arcsec/pixel, and its field of view is $210\arcsec \times 204\arcsec$. The APSIS pipeline (Blakeslee et al. 2003b), with a {\it Lanczos3} interpolation kernel, was used for processing the images. The ACS photometric zero--points (AB system) are 25.654~mag and 24.862~mag in $i_{775}$ and $z_{850}$, respectively (Sirianni et al. 2005). A Galactic reddening of $E(B-V)=0.019$ towards RXJ~0910+5422 was adopted (Schlegel et al. 1998), with $A_{i775}=0.039$ and $A_{z850}=0.029$ (Sirianni et al. 2005). The ACS WFC field covers an area that at the redshift of this cluster, $z=1.106$, corresponds to $\approx 1 Mpc^2$ in the WMAP cosmology (Spergel et al. (2003): $\Omega_m =0.27$, $\Omega_{\Lambda} =0.73$, $h=0.71$, adopted as our standard cosmology hereafter). Fig.~\ref{cluster} shows the ACS color image with X--ray contours from Chandra ACIS (Advanced CCD Imaging Spectrometer) data that have been adaptively smoothed (Stanford et al. 2002). Near-IR $JK_s$ and optical $i$-band images were obtained at Palomar Observatory as described in detail by Stanford et al. (2002). Optical spectroscopy of galaxies in RXJ~0910+5422 was obtained using the Low Resolution Imaging Spectrometer (LRIS; Oke et al. 1995) on the Keck 1 and 2 telescopes (Stanford et al. 2005; in preparation). Our typical errors in redshift correspond to errors in velocity between 100 and 300~km/s. Objects for spectroscopy were chosen initially from the catalog of objects with $K_s < 20.0$~mag (Vega magnitudes) within the IR imaging area; outside of this area objects were chosen with $i > 21$~mag from the $i$-band image to fill out masks. Our final sample included 66\% of the objects with $K_s < 20.0$~mag. Spectra were obtained using the 400 lines mm$^{-1}$ grating for all runs except for the initial two discovery masks as reported in Stanford et al.\ (2002). Nine more masks were observed using LRIS during four runs between January 2001 and February 2003. Usually each mask was observed in a series of four 1800~s exposures, with small spatial offsets along the long axis of the slits. On average, the seeing was 0.9~$\arcsec$. The blue side data were generally not used since the rest frame wavelengths probed at $z = 1.1$ fall far to the blue of the spectral features of interest for galaxies in the cluster. In total, 149 redshifts were obtained. The slit mask data were separated into slitlet spectra and then reduced using standard long-slit techniques. A fringe frame was constructed for each exposure from neighboring exposures, each offset from the previous by 3$\arcsec$, in an observing sequence for each mask, and then subtracted from each exposure to greatly reduce fringing in the red. The exposures for each slitlet were reduced separately and then co-added. One-dimensional spectra were extracted for each targeted object, as well as the occasional serendipitous source. Wavelength calibration of the 1-D spectra was obtained from arc lamp exposures taken immediately after the object exposures. A relative flux calibration was obtained from long-slit observations of the standard stars HZ44, G191B2B, and Feige 67 (Massey \& Gronwall 1990). \section{Object selection and photometry} SExtractor (Bertin \& Arnouts 1996) was used to find objects in the $i_{775}$ and $z_{850}$ images and measure their magnitudes. Threshold and deblending settings were used as in Ben\'{\i}tez et al. (2004). Although we have extensive spectroscopy, the ACS imaging reaches considerably deeper along the cluster luminosity function. Thus, we have chosen to use colors ($i_{775}$ - $z_{850}$) and $(J-K_s)$ to isolate a set of probable cluster members. In Fig.~\ref{colage}, the ($i_{775}$ - $z_{850}$) and $(J-K_s)$ colors are shown as a function of galaxy age, using BC03 stellar population models, redshifted to $z{=}1.106$. Early--type cluster members would have ages of at least 0.5~Gyr, corresponding to $(i_{775} - z_{850}) >$0.8~mag and $(J-K_s) > 1.45$~mag. At first, we give a larger color margin and select as potential cluster members all morphologically-classified early-type galaxies with $0.5 <(i_{775} - z_{850}) < 1.2$~mag and $(J-K_s) > 1.45$~mag, down to $z_{850} = 24$~mag (the limiting magnitude of Postman et al. (2005) morphological classification, that included all clusters in our sample at redshift unity). Our results in this paper are based on this morphological classification and a detailed discussion of the uncertainties in this classification can be found in that work. This selected sample includes 38 galaxies within the ACS field. Our final colors were measured within galaxy effective radii ($R_e$), to avoid biases due to galaxy internal gradients, following the approach in Blakeslee et al. (2003a) and van Dokkum et al. (1998, 2000). $R_e$ values were derived with the program GALFIT (Peng et al. 2002), constraining the {\it Sersic} index $n \le 4$ (as in Blakeslee et al. 2003a). To remove differential blurring effects (the PSF is $\sim10\%$ broader in the $z_{850}$ band) each galaxy image in both $i_{775}$ and $z_{850}$ was deconvolved using the CLEAN algorithm (H{\"o}gbom et al. 1974). The $(i_{775} - z_{850})$ colors were measured on the deconvolved images within a circular aperture of radius equal to $R_e$, or 3 pixels, whichever is larger. Our median $R_e$ is $\approx$~5.5~pixels ($\approx$~13kpc at z=1.106). Our final results do not change (within the uncertainties) if the effective radii are calculated via a two component (Sersic bulge + exponential disk) surface brightness decomposition technique using GIM2D (Marleau \& Simard 1998; Rettura et al., in preparation), that permits us to better fit the galaxy light profile. The photometric uncertainties due to flat fielding, PSF variations, and the pixel-to-pixel correlation for ACS (Sirianni et al. 2005) were estimated by measuring the standard deviation of photometry in the background for circular apertures in the range of the measured effective radii. These photometric errors were added in quadrature to the Poisson uncertainties in the measured fluxes for each object. The derived errors in the colors are between 0.01 and 0.03~mag down to $z_{850} = 24$~mag. SExtractor MAG\_AUTO were used for the $z_{850}$ magnitude in the color--magnitude relation; these are fairly robust, though may systematically miss a small fraction of the light (Ben\'{\i}tez et al.\ 2004). We finally color--selected 34 early--type (E, S0 and S0/a) galaxies with $0.8 <(i_{775} - z_{850}) < 1.1$~mag within 2$\arcmin$ from the cluster center, taken as the center of the X--ray emission (Stanford et al.\ 2002). Images of the color--selected galaxies are shown in Fig~\ref{epostage}, Fig~\ref{sopostage} and Fig~\ref{sapostage}. Moreover, there are late--type galaxies with luminosities that are similar to the red--sequence bright early--type galaxies (Fig.~\ref{spipostage}). Of the 34 color--selected galaxies, 15 are spectroscopically confirmed cluster members, one (S0/a, with magnitude $z_{850} = 24.2$~mag) is a confirmed non-member, and the others were not targeted for spectroscopy. The selection in $(i_{775} - z_{850})$ at $z=1.1$ therefore appears to be robust: only one of the 16 selected galaxies with measured redshifts is a non-member. We expect few of the other 18 to be interloper field galaxies. \section{Color-Magnitude Relation} The color--magnitude relation for the final color--selected objects is shown in Fig~\ref{cmd}. Red dots are ellipticals, orange squares and stars are S0 and S0/a galaxies, respectively. Smaller black symbols represent early--type galaxies that do not lie on the red sequence. Small triangles are late-type galaxies. Boxes are plotted around confirmed cluster members. Confirmed interlopers are circled in the figure. Surprisingly, the two brightest cluster members are not ellipticals, but S0. The brightest of these two galaxies lies $\approx$~700~kpc ( $\approx$1.2$\arcmin$) from the cluster center, and the other bright S0 at $\approx$~300~kpc ($\approx$~0.6$\arcmin$). Moreover, there are late--type galaxies with luminosities that are similar to the red--sequence bright early--type galaxies. Two of them lie on the red--sequence and are confirmed cluster members, at $\approx$~80~kpc from the cluster center (see also below in the discussion of the color and morphology distribution as a function of distance from the cluster center). We fitted the following linear color--magnitude relation to various subsamples of the galaxies: \begin{equation} i_{775} - z_{850} = c_0 + Slope (z_{850} - 23) \end{equation} The solid line in Fig~\ref{cmd} is the fit to the color--magnitude relation for the ellipticals, the black dotted line is the fit to the CMR for the S0s, and the dashed--dotted line the fit to the full sample, within 2$\arcmin$ from the cluster center (see discussion below). The dashed line is the fit to the full sample of early--type galaxies in RXJ1252.9-292 from Blakeslee et al. (2003a), scaled to this redshift with BC03 evolved stellar population models, with solar metallicity and a formation age of 2.6~Gyr (since Blakeslee et al. 2003a obtains elliptical mean ages ~$>$~2.6~Gyr). The long--dashed vertical line is the magnitude limit of the morphological classification $z_{850} = 24$~mag. The results for different morphological samples are given in Table~\ref{results}. A robust linear fit based on Bisquare weights (Tukey's biweight; Press et al. (1992)) has been used to fit the color--magnitude relation. Uncertainties on the parameters were estimated by bootstrapping on 10,000 simulations. The scatter around the fit was estimated from a biweight scale estimator (Beers, Flynn \& Gebhardt 1990), that is insensitive to outliers, in the same set of bootstrap simulations. The internal color scatter ($\sigma_{int}$) was measured in two ways: 1) to the scatter around the fit, we have subtracted in quadrature the average uncertainty due to the galaxy color error; and 2) we have calculated the internal scatter for which the $\chi^2$ of the fit would be unity. Both methods give us internal scatters consistent to within a few 0.001~mag. All galaxies in this sample lie within three sigma from the fit. The X--ray distribution appears to be very symmetric, and largely confined within 1\farcm5 from the cluster center. We calculated the CMR zero point and scatter within 1$\arcmin$ (which corresponds to a scale of $\approx$~0.5~Mpc at this redshift), and within 1\farcm5 ($\approx$0.7~Mpc), and within 2$\arcmin$ ($\approx$~1~Mpc, the scale used for the analysis of RXJ1252.9-292). According to the results in Table~\ref{results}, the internal color scatter increases when adding populations between 1$\arcmin$ and 2$\arcmin$, especially for the S0 and S0/a populations, as also observed in local samples (e.g. van Dokkum et al. 1998), with only a small increase in sample size. We will therefore focus on the results obtained for color--selected galaxies within 1$\arcmin$ from the cluster center (where 90\% of the color-selected galaxies lie). The slope of the elliptical CMR ($-0.033 \pm 0.015$) is slightly steeper than the observed slope in RXJ1252.9-292 ($-0.020 \pm 0.009$), and in Coma when the latter are shifted to the observed colors at $z \sim 1.1$, using non-evolving BC03 stellar population models, but still consistent within the uncertainties. We do not find a flatter (with respect to Coma) slope as in Stanford et al. (2002). However, the S0 sample shows a much shallower slope ($0.005 \pm 0.023$) than the ellipticals, resulting in a much flatter slope for ellipticals and S0s together ($-0.024 \pm 0.020$). This can explain why a shallower slope was found in that work, in which elliptical and S0s were not separated. The spectroscopically-selected elliptical plus S0 slope ($-0.010 \pm 0.034$) is also flattened by the S0 population, while the spectroscopically-selected ellipticals have a slope ($-0.021 \pm 0.046$) similar to that of RXJ1252.9-292 and Coma. All the difference in slope are however within the uncertainties and are statistically insignificant. Using Bruzual \& Charlot (2003; BC03) stellar population models, as in Blakeslee et al. (2003a), we derive a constraint on the age of the stellar populations in the galaxies from galaxy colors and the scatter of the CMR (van Dokkum et al. 2001; Blakeslee et al. 2003a). Two simple models have been considered and our conclusions will depend on the chosen models. The first model is a {\it single burst} model, in which galaxies form in single bursts at random times $t_f$, between the age of the cluster and the recombination epoch. The second is a model with {\it constant star formation} in a range of time between $t_1$ and $t_2$, randomly chosen to be between the age of the cluster and the recombination epoch. Colors for 10,000 galaxies were simulated with their scatter around the CMR to be dependent on the burst age. In Fig~\ref{scatterage}, we show the simulated scatters as a function of burst age, with solar, half solar and twice solar metallicity model. We will assume solar metallicity in what follows. From the scatter ($\sigma_{int}=0.042\pm 0.010$) in the colors of the galaxies classified as ellipticals, we obtain ages $>2.1$~Gyr ($z >2$), with a mean luminosity--weighted age $\overline t=3.31$~Gyr ($z_f \approx$~3.1), assuming the random single burst model. From the constant star formation model, we obtain ages $>1.6$~Gyr ($z >1.7$), with a mean luminosity--weighted age $\overline t=3.26$~Gyr ($z_f \approx$~3). This agrees with the conclusion (e.g. Blakeslee et al. 2003a; Holden et al. 2004; Lidman et al. 2004; De Lucia et al. 2005) that the elliptical population in clusters of galaxies formed at $z_f >$~3, and has evolved mainly passively until $z = 1.1$. In the (U-B)--rest frame (using BC03 stellar population models with solar metallicity and age equal to 4~Gyr), a scatter in $(i_{775}-z_{850})$ of $0.042\pm 0.010$ corresponds to $0.050 +/- 0.011$. As pointed out in van Dokkum (2000) and Blakeslee et al. (2003a), CMR scatters vary little with redshift. The Blakeslee et al. (2003a) scatter for the elliptical CMR in RDCS~1252--2927 ($0.024 \pm 0.008$), correspond to a scatter of $0.042 \pm 0.014$ in the (U-B)--rest frame, indistinguishable within the uncertainties from our result. The scatter in the CMR for galaxies classified as S0 ($\sigma_{int}=0.044\pm 0.020$) is comparable to the one in the E CMR, but the galaxy colors are bluer and are not compatible with a population that is as old as the ellipticals. All the S0s lie below the elliptical color--magnitude relation. In fact, between the elliptical and the S0 CMR fits there is a zero point difference of $0.07 \pm 0.02$~mag, with the S0s being bluer than the ellipticals. One of the three S0/a galaxies has a color that is 0.07~mag redder than the elliptical CMR, and another has a color that is $\approx$~0.15 bluer than the CMR relation for Es. The inclusion of the S0/a galaxies does not significantly change the fitted CMR for the S0s. When we consider all galaxies within 2$\arcmin$ from the cluster center, the S0 and total early--type slopes are similar, while the color offset in the CMR is still present ($0.05 \pm 0.02$~mag). The S0 population of this cluster has a very peculiar CMR with respect to the average cluster of galaxies. In fact, for several other studies the CMR of the S0 population has a similar zero point and on average a larger scatter with respect to the elliptical population (van Dokkum et al. 1998; Blakeslee et al. 2003a; Holden et al. 2004; De Lucia et al. 2004), quite different from our results. We will discuss this peculiar behavior in detail in the rest of the paper, including an examination of orientation effects on the classification In Fig.~\ref{ircolors}, the near-IR (Vega magnitudes) and $(i_{775} - z_{850})$ (AB magnitudes) colors are shown compared with single burst stellar population model predictions from BC03. The S0 colors are consistent with young ($<$~2~Gyr) solar metallicity, or older ($<$~3.5~Gyr), half solar metallicity populations. If the difference in E and S0 mean colors is mainly due to metallicity, then even if the two populations were formed at the same epoch, ellipticals must have been able to retain more metals than the S0s, i.e., they were more massive at a given luminosity (given the observed mass--metallicity relation for early-type galaxies, e.g., Tremonti et al. 2004, Bernardi et al. 2005, and references therein). This would imply higher mass--to--light ratios for the ellipticals with respect to the S0s. However, the lack of strong evolution in the slope and scatter of the CMR from the present out to $z{\sim}1$ suggests that the CMR is mainly the result of a metallicity--mass (i.e. metallicity--magnitude) relation (e.g., Kodama \& Arimoto 1997; Kauffman \& Charlot 1998 Vazdekis et al. 2001; Bernardi et al. 2005). So, at a given magnitude we do not expect large metallicity variations. If the offset is due to a different star formation history, a model with solar metallicity and with an exponential decay of the star formation will reproduce the offset at a galaxy mean age of $\approx$~3.5~Gyr. This age is consistent with the small scatter observed in the S0 CMR. We would then be observing galaxies that followed different star formation: single burst and passive evolution for the ellipticals and exponentially decaying star formation for the S0s. An exponential decay in the star formation is observed in field spiral samples (Rowan--Robinson 2001). If this is the case, our S0 population might be the evolved product of an old spiral population that was already in place in this cluster when the ellipticals formed and then gradually lost available gas for star formation. If the E vs S0 color difference is mainly due to a difference in age, for a solar metallicity and a single burst BC03 template with age 4~Gyr, the color difference corresponds to an age difference of $\sim$1~Gyr. For clusters of galaxies at z~$>$~1, the cluster members on the red sequence are only a part of all the progenitors of present--day early--type galaxies. Some of today's galaxy progenitors would have been bluer than the red sequence at these redshifts (van Dokkum \& Franx 2001). In the S0 population of this cluster, we may be seeing the transitional progenitor population that in $\sim$1~Gyr will evolve onto the same red sequence as now occupied by the ellipticals. Either of the latter two scenarios would be consistent with the Postman et al. (2005) observed deficit of the S0 population of our ACS cluster sample, when compared to lower redshift samples, implying that part of the S0 population is still forming in clusters at redshifts around unity. \section{Galaxy shape properties} Since the galaxies classified as S0 in RXJ~0910+5422 are found to be systematically bluer (with respect to the red sequence) than the S0 populations observed in previous studies, we wish to examine further the properties of these galaxies in terms of their shapes and light distributions, and how they compare to the elliptical and spiral samples in this and other clusters. The shape parameters that we will consider are Concentration and Asymmetry (Abraham et al. 1996; Conselice et al. 2004), Sersic index $n$, and galaxy axial ratios. \subsection{Asymmetry and Concentration} In Fig~\ref{cas} (left), we compare the Asymmetry A and the Concentration C for ellipticals, S0s, and spirals with $(i_{775} - z_{850})$ colors between 0.5 and 1.2~mag. The Asymmetry parameter is obtained by subtracting a 180--degree rotated image from each original galaxy image, summing the residuals and including a correction for the background. The Concentration parameter is defined as in Abraham et al. (1996) as the sum of the galaxy flux within an aperture $r_{0.3}$ divided the total flux. $r_{0.3}$ is calculated using the SExtractor fit to the galaxies at 1.5~$\sigma$ above the background. The obtained semi--major and semi--minor axes from this fit were multiplied by 0.3 to derive the $r_{0.3}$ aperture (see also Homeier et al. 2005b). Early--type and late--type galaxies lie on different regions in this A vs C plane. All our red sequence S0s have $A < 0.3$ and $C > 0.3$. All but one have $A < 0.2$ and $C > 0.3$. This is the same locus in the A--C plane that is occupied by most early--type galaxies in Abraham et al. (1996) and in our Postman et al. (2005) low--redshift sample (Fig~\ref{cas}, right). This last sample includes 5 strong lensing clusters observed as part of our ACS GTO program [Zw1455+2232 ($z{=}0.258$), MS1008-1224 ($z{=}0.301$), MS1358+6245, CL0016+1654 ($z{=}0.54$), and MS J0454-0300 ($z{=}0.55$)]. Visual and automated classification for this cluster in the $i_{775}$--band for all galaxies with $i_{775}< 22.5$ was performed by Postman et al. (2005). We conclude that the S0 population presents statistical parameters typical of an early--type population (low asymmetry and high compactness). \subsection{Sersic Indices} Fig~\ref{justsersic} plots Sersic index $n$ as a function of galaxy effective radius $R_e$, both from our GALFIT modeling, for ellipticals, S0s and spirals with $(i_{775} - z_{850})$ colors between 0.5 and 1.3~mag. Red sequence ($(i_{775} - z_{850})$ color between 0.8 and 1.1~mag) galaxies are shown by large diamonds. Most spirals have $n<2$ and most early--types have $n>2$, as expected. However, the $n$ values do not permit us to discriminate between S0s and ellipticals in a unique way, unless they are combined with goodness--of--fit information for the Sersic model (e.g. the {\it Bumpiness} parameter introduced by Blakeslee et al. 2005) \section{Axial Ratios} \subsection{Axial Ratio Distribution} In Fig~\ref{ab} we compare the apparent axial ratio from SExtractor versus effective radii for elliptical and S0s with $(i_{775} - z_{850})$ colors between 0.5 and 1.3~mag. The axial ratios have been verified by using ELLIPROF (the isophotal fitting software that is used for Surface Brightness Fluctuations analysis in Tonry et al. 1997 and Mei et al. 2005) on each galaxy image, after the cleaning procedure. As above, red sequence galaxies are shown by large diamonds. The red sequence ellipticals and the bluer S0s have different axial ratio distributions, with all red sequence S0s showing axial ratios $\frac{b}{a} \lesssim 0.7$, and nearly all red sequence ellipticals with $\frac{b}{a} > 0.7$. Assuming axisymmetric disks (oblate ellipsoids) viewed with random orientation, and with a Gaussian distributed intrinsic axial ratio (with mean equal to $0.3 \pm 0.1$ (extreme thin disk), $0.5 \pm 0.1$ (early--type galaxy) and $0.75 \pm 0.1$ (elliptical) (Jorgensen \& Franx 1994)), one would expect $\gtrsim\,$40\%, $\gtrsim\,$60\%, and $\gtrsim\,$90\%, respectively, of the S0s to have axial ratios above 0.7. Just 9\% (1 out of 11) of the red sequence S0s are observed to have an axial ratio this large (or 22\% for the full S0 sample in this cluster field) (Fig~\ref{ab2}; top). The random probability that the S0 axial ratios would show such a low fraction with $\frac{b}{a} > 0.7$ is less than 1\%. This is a very simple model, but it points out a lack of round S0s, indicating either that there is some orientation bias in the classification, or that this class of objects is intrinsically prolate in shape. Jorgensen \& Franx (1994) found a similar deficit of round S0s in the center of the Coma cluster. They concluded that part of the face--on S0s were classified as elliptical galaxies. Fabricant et al. (2000) also found a deficit of round S0s in the cluster CL~1358+62, at z=0.33. Their analysis shows that ellipticities and bulge--to--total--light--ratio do not allow us to distinguish elliptical from S0 galaxies. The other two ACS GTO clusters at $z{\,>\,}1$ (RXJ1252.9-292 and RX~J0848+4452) do not show a similar lack of round S0s, as 90\% of the red sequence ellipticals (out of $\approx\,$70) and 47\% of the S0s (out of $\approx\,$35) have $\frac{b}{a} > 0.7$ (Fig~\ref{ab2}; bottom). % This bias is also not observed in other clusters of our ACS Intermediate Redshift Cluster Survey (see Fig~5 from Postman et al. 2005), for which more than 40\% of the S0 galaxies have axial ratios above 0.7. The observed peculiarity of the RXJ~0910+5422 S0 axial ratio distribution might call into question our result above that the S0s have a significant color offset with respect to the ellipticals. For instance, if there is a bias in our color measurement procedure which causes elongated objects to have colors that are too blue, then the color offset found above may be artificial. Such a color bias might occur if the high inclination angles bias our $R_e$ measurements to higher values, and if the S0s become progressively bluer at larger radii. We first examine this possibility, then proceed to discuss resolutions to the peculiarity of the S0 axial ratio distribution, with the aim to establish if a misclassification of face--on S0s as ellipticals would bias the measurement of the offset between the elliptical and S0 CMR zero points. \subsubsection{Internal color gradients} It is conceivable that our $(i_{775} - z_{850})$ colors could be biased by aperture effects in the nearly edge--on S0s, for which the (possibly) bluer outer disks might contribute more to the galaxy colors than in the rounder ellipticals population. If this effect were severe enough, it might mimic the offset in color of the S0s and ellipticals found above. We test this possibility here. S0 and elliptical internal color profiles are shown in Fig~\ref{gradso} and Fig~\ref{grade}. The gradients have been calculated with aperture photometry on the same images used to calculate our $(i_{775} - z_{850})$ colors. Circles are $(i_{775} - z_{850})$ colors at different radii, the cross the $(i_{775} - z_{850})$ color at the effective radius, used for the CMR. The S0 colors profiles do not show strong gradients. In particular, the $(i_{775} - z_{850})$ colors calculated at the effective radii are not systematically bluer than colors determined at smaller radii. Most of S0 galaxies have flat profiles; one has a blue inward gradient (ACS ID 1621; $z_{850}=23.72$~mag, $(i_{775} - z_{850})=1$~mag). In two galaxies (ACS ID 1393 and 3177) the colors in the central 0.15~$\arcsec^2$ are redder than the color at the effective radius. When compared with elliptical gradients, on average S0 colors do not appear biased towards higher effective radii and bluer colors than the ellipticals. We note that three elliptical galaxies show blue inward gradients (ACS ID 1753, 1519, 3323). \subsubsection{Orientation or intrinsic shape: Axial ratios vs $(i_{775} - z_{850})$ colors} Orientation biases are known to occur in the classification of ellipticals and S0s in local galaxy samples. For instance, Rix \& White (1990, 1992) showed, based on both isophotal and dynamical modeling, that a large fraction of ellipticals contain a disk component with at least $\sim\,$20\% of the light, but which is hidden due to projection effects. Jorgensen \& Franx (1994) found a strong deficit of round S0s in a sample of 171 galaxies in the central square degree of the nearby Coma cluster, and concluded that inclination angle played a large part in the classification of Es and S0s. Michard (1994) proposed that, except for the bright boxy ellipticals without rotational support, early-type galaxies comprise a single class of oblate rotators with orientation being the main criterion for classification as either E or S0. On the other hand, van den Bergh (1994) explained the predominance of flattened S0s by invoking two distinct subpopulations: bright disky objects intermediate between ellipticals and spirals, and a fainter population of prolate objects. We now address the question of whether the galaxies classified as S0 in RXJ~0910+5422 are preferentially flattened in shape because of an orientation bias in the classifications or intrinsically prolate shapes. If it is an orientation bias, then this could mean either that (1) face-on S0s have been misclassified as ellipticals because their disks are not apparent, or (2) that edge-on spirals tend to be called S0s because the spiral structure is obscured. Either would result in a predominance of flattened S0s. However, in the former case, the misclassification of face-on S0s as ellipticals would tend to blur any color separation between the two classes, while in the latter case, a color offset might be introduced between the two classes because of contamination by bluer spirals. Because we do observe a color offset between galaxies classified as E and S0, with the S0s being bluer, it is possible to look for the ``missing'' population of face-on blue galaxies by examining ellipticity versus color. If a population of round blue objects is found, we can then determine the nature of the classification bias, and whether it biases our color offset measurement. Histograms of the axial ratio distributions for ellipticals and S0s are shown in the top panel of Fig.~\ref{ab2}. We find that 60\% of the early-type red sequence galaxies have $\frac{b}{a} > 0.7$, but 95\% of these low-ellipticity galaxies are classified as Es. However, if we split galaxies instead by color, using $(i_{775} - z_{850}) < 0.99$~mag as the separation point, then we find that 54\% of all galaxies bluer than this separation have $\frac{b}{a} > 0.7$, while 43\% of the red sequence galaxies bluer than this [$0.8 < (i_{775} - z_{850}) < 0.99$~mag] have $\frac{b}{a} > 0.7$ (Fig.~\ref{ab3}). Thus, the deficit of round S0 galaxies (which are also significantly bluer than the mean of the E class) is not found when the early-type galaxies are split based purely on color. This suggests that some of the bluer round galaxies are the face-on counterparts of those classified as S0. Fig~\ref{blue} shows axial ratios vs $(i_{775}- z_{850})$ color residuals with respect to the total early-type galaxy CMR relation, for all galaxies in the RXJ~0910+5422 red sequence. Galaxy types are coded with different symbols; using the color residuals in this way takes out the effect of the magnitude dependence of the colors. There are five round blue ellipticals, i.e., with $\frac{b}{a} > 0.7$ and on the blue side of the early-type CMR. If these are the face-on counterparts of the blue S0s, then the S0 axial ratio distribution becomes much more in line with expectations for a randomly oriented disk population (35\% are rounder than $\frac{b}{a}{\,=\,}0.7$). Further, we note that four other ellipticals were classified E/S0 in Postman et al.\ (2005); if these are also taken as face-on S0s, then the axial ratio distribution comes in very close agreement with expectations. We conclude that the face-on S0s in RXJ~0910+5422 are classified as ellipticals, just as in local early-type galaxy samples. From our axial ratio simulations and statistics from our z$\approx$1 ACS GT0 sample, the number of blue S0 galaxies might be between 25 and 40\% higher than estimated in section~5.3.1. Moreover, the inclusion of a few blue S0s in the elliptical red--sequence will have the effect of increasing very slightly the observed CMR scatter for the ellipticals. The uncertainty in the Postman et al. (2005) S0 fraction in RXJ~0910+5422 is significantly larger than 40\% and hence a systematic change by this amount for one cluster does not alter any of the results or conclusions in that paper. Throughout the paper, we continue to use elliptical and S0 classifications from Postman et al. (2005), but keeping in mind the presence of this possible projection effect. \section{Color trends, velocity dispersion and merging activity} Stanford et al. (2002) have analyzed the X--ray and near-IR properties of this cluster. This system appears to be fairly relaxed based on its regular X--ray profile; however, they find indications that the cluster is in an early phase of formation. In fact, the Chandra ACIS data show evidence for temperature structure, possibly due to an infalling group or mass streaming along a filament. The soft component of the X--ray emission (0.5--2~keV) dominates the X--ray center of the cluster, while to the south there is a harder component (2--6~keV) (see Fig. 6 from Stanford et al. 2002). The cluster does not have a central BCG or cD galaxy, and the X--ray emission center does not correspond to an optical grouping of galaxies; rather a number of luminous confirmed cluster members are linearly distributed, at least as projected on the sky (as shown in Fig~\ref{cluster}). We do not observe any strong trend of the galaxy color $(i_{775} - z_{850})$ with distance from the cluster X--ray center (Fig.~\ref{colors}). Surprisingly, though not statistically significant, the bluer galaxies are concentrated toward the cluster center, instead of the outskirts, as in the other ACS intermediate redshift cluster sample (Demarco et al. 2005; Goto et al. 2005; Homeier et al. 2005; Postman et al. 2005). This tendency might support the hypothesis that we are observing sub-groups of galaxies in an edge--on sheet, e.g a group of bluer disk galaxies on a redder, older population of ellipticals. There are 10 confirmed elliptical and 5 confirmed S0 members in the center of the cluster (R~$<$~500~kpc). The average S0 redshift is 1.102~$\pm$~0.002 and the average elliptical redshift is 1.105~$\pm$~0.007 (the given uncertainties are standard deviations around the mean). This indicates small relative velocities (the two redshifts are indistinguishable given the errors) between the classes and is true regardless of the possible classification bias discussed above. Unfortunately, our spectroscopic sample does not permit us to track the cluster central structure in detail. The average relative velocity between the confirmed E and S0s of $\approx$~500~km/s (that also corresponds to the cluster velocity dispersion; see below) is fairly small. If merging of two distinct groups of galaxies is happening along the line of sight, we expect much higher velocity dispersions and/or relative velocities between the infalling S0s and the ellipticals. Stanford et al. (2002) also suggest that active galaxy-galaxy merging should be observed, based on the X--ray temperature structure. To investigate any on--going dynamical activity, we calculated the cluster velocity dispersion, and the merger rate. From the 25 spectroscopically confirmed members (all galaxy types included) the line--of--sight rest--frame velocity dispersion is $\sigma = 675 \pm 190$~km/s, using the software ROSTAT from Beers, Flynn, \& Gebhardt (1990). The available Chandra data give X-ray temperatures ranging from $kT = 7.2^{+ 2.2}_{-1.4}$~keV (Stanford et al.\ 2002) to $kT = 6.6^{+ 1.7}_{-1.3}$~keV (Ettori et al.\ 2004). Wu et al. (1999; see also Rosati et al. 2002) gives the relationship between $kT$ and $\sigma$ for relaxed clusters, which predicts that the velocity dispersion corresponding to the measured X-ray temperature should be $\approx$~1000~km/s, considerably higher than we have found. Again, in the case of merging groups (along the line of sight) we would also have expected a higher velocity dispersion. \section{Red galaxy pairs} The quality of the ACS data allows us to discern merging activity among the cluster galaxies. If we assume that galaxy pairs with projected separations less than $20 h^{-1}_{70}$~kpc are physically associated, we observe 13 associated early--type galaxies, nine galaxies of which lie on a filamentary structure about $\approx$~100~kpc from the cluster center (Fig.~\ref{cluster}). As noted above, RXJ~0910+5422 lacks any cD galaxy near the center of the X--ray emission (see also Fig.~\ref{cluster}), but rather has a filamentary group of galaxies around the X--ray center. The nine early-type interacting galaxies within this filamentary structure (at radius of $\sim100 h^{-1}_{70}$~kpc from the X--ray center) include 3 unique pairs (yellow arrows in Fig.~\ref{cluster}), plus a galaxy triplet (the three components marked with red arrows in Fig.~\ref{cluster}). Each of the 3 pairs consists of a bright elliptical with a smaller companion (all closer than $10 h^{-1}_{70}$~kpc), while the triplet is a large E with two smaller S0s (also closer than $10 h^{-1}_{70}$~kpc; one of these two S0s is ACS ID 1621, the S0 with an inward blue gradient). One of the pairs (the middle pair in the figure) and the two nearest galaxies in the triplet, have essentially zero relative velocity and thus are likely merger candidates. Also two of the ellipticals with blue inward gradients lie on the central filaments, and they are both small satellites of a larger galaxy. The other two pairs, which do not lie on the filamentary structure, are at ~$250h^{-1}_{70}$~kpc (two E with similar size, both with weak O~II emission (Stanford et al. 2002) and ~$300 h^{-1}_{70}$~kpc (one S0/a and one E of similar size) from the X--ray center, and have relative velocities of ~10,000~km/s and ~2000~km/s, respectively. The presence of the low-velocity pairs is consistent with the low velocity dispersion, and provides evidence for the on--going hierarchical growth of the cluster (e.g., van Dokkum et al. 1999). The pairing of the bright ellipticals with smaller S0s also argues against the view that we are observing a blue S0-dominated group infalling into a red cluster elliptical population, but rather complex stellar population evolution within a filamentary structure. The red--sequence S0/a confirmed members also lie within this filamentary structure. The observations of a significant number of red galaxy pairs in a cluster at z~$\sim$~1 is interesting in the context of the recent findings by van Dokkum (2005) of red galaxy interactions in $\approx$~70\% of 86 early--type galaxies in a selected sample of nearby red galaxies from the MUSYC (Multiwavelength Survey by Yale--Chile; Gawiser et al. 2005) and the NOAO Deep Wide--Field Survey. This work concluded that most of the ellipticals in local samples were assembled by red galaxy--galaxy mergers, denominated {\it dry} mergers because they would involve gas--poor early--type galaxies. At higher redshift, Tran et al. (2005b) confirmed red galaxy mergers first observed by van Dokkum (1999) in MS~1054-03 at z=0.83. Tran et al. selected mergers as associated pairs with projected separation less than $10 h^{-1}_{70}$~kpc and relative line--of--sight velocities less than 165~km/s. As in RXJ~0910+5422, the red early--type galaxies involved in these mergers are among the brightest cluster members. Their results suggest that most early--type galaxies grew from passive red galaxy--galaxy mergers. In our sample we observe a triplet and three red galaxy pairs with projected distances less than $10 h^{-1}_{70}$~kpc. Of those, the triplet and one galaxy pair show zero relative velocity. Of the two other pairs, composed of a bright and a fainter companion, redshifts are not available for the faint companions. \section{Cluster luminosity function} To obtain a deeper understanding of the RXJ~0910+5422 galaxy population, we constructed galaxy luminosity functions in the following way. We start with the original Sextractor catalog (described in Section~3). All objects with magnitudes brighter than $i_{F775W} = 21.1$~mag were considered as foreground objects. Nine of these bright objects are confirmed non-members. The remaining seven are objects that do not belong to the red sequence and whose sizes and luminosities are much larger than those of the confirmed members, in particular those of the bright red sequence galaxies; therefore they are very unlikely to be at the cluster redshift. The contribution to the luminosity function from both foreground and background field galaxies (hereinafter, the field) has been estimated from the galaxy counts in a reference field. The control region is taken from the GOODS-S (Great Observatories Origins Deep Survey--South; Giavalisco et al. 2004) ACS field, observed in the same filter as the cluster field. Point--like objects were eliminated in a consistent way in the cluster and in the control field, by identification of the stellar locus in the diagnostic plot of the SExtractor parameters MAG\_AUTO vs FLUX\_RADIUS (the selected objects have FWHM equal to the PSF in the image). Cluster and control field luminosity functions were normalized to the cluster area. Both cluster and field counts were binned with a bin size of 0.5~mag. For each bin, the field counts are subtracted from the cluster counts, taking into account the extensive spectroscopic sample (more than 60\% of the objects used for the LF determination brighter than M$^{*}$ have measured redshifts). Known interlopers were excluded from the analysis. The uncertainties in the cluster counts after subtraction of the field contribution are calculated by adding in quadrature Poissonian uncertainties. The luminosity functions are shown in Fig~\ref{lumfun}. The filled circles with errors are the total background--corrected cluster luminosity function. We do not include errors from cosmic variance due to the choice of the background control region. The red sequence elliptical and spheroidal (S0 and S0/a) luminosity functions are shown respectively in blue and green. The red histogram is the luminosity function of all early--type galaxies with color $0.8 <(i_{775} - z_{850}) < 1.1$~mag, excluding confirmed non--members. The histogram of red sequence galaxies in RXJ1252.9-292 is shown as the dashed red line. The background contribution is very small for the early-type sample. The red arrows show the histogram values after background subtraction. The rest-frame $B$~magnitudes are shown on the top of the plot, calculated from colors obtained from the BC03 stellar population model and templates (Sbc, Scd) from Coleman, Wu \& Weedman (1980). The solid black line is a Schechter function fit to the total cluster luminosity function. It is obtained by calculating the C (Cash 1979) statistic (a maximum likelihood statistic to fit data with Poissonian errors) on a grid in the M$^{*}$ -- $\alpha$ plane for each combination of M$^{*}$ and $\alpha$: first the normalization ($\phi^{*}$) is calculated in order to reproduce the observed number of galaxies in the observed magnitude range, then the C statistic is computed as $C=-2\Sigma_{i_{bin}} n_{i} \ln m_{i} - m_{i} - \ln n_{i}!$, where $n_{i}$ is the observed number of galaxies in the $i\,$th bin and $m_{i}$ is the number of galaxies predicted in that bin by the Schechter function with parameters M$^{*}$, $\alpha$, and $\phi^{*}$. The combination M$^{*}$,$\alpha$ which minimizes the C statistic is taken as the best-fit. If the C statistic is defined as above, the 1- 2- and 3-$\sigma$ confidence levels for M$^{*}$ and $\alpha$ can be estimated from $\Delta C = 2.3, 6.17,11.8$. We obtain M$^{*}=22.6^{+0.6}_{-0.7}$~mag and $\alpha=-0.75\pm0.4$. Most of the faint--end population is composed of S0 and S0/a galaxies. The two brightest galaxies in the red sequence are S0. With respect to RXJ1252.9-292, a bright population of red sequence ellipticals is missing in RXJ~0910+5422. However, the large Poissonian errors on the bright end of the cluster population prevent us from definitively excluding the hypothesis that the two clusters could be drawn from the same parent population. Similarly, small number statistics do not permit us to study the luminosity functions of the different red and blue faint populations in this cluster. \section{Discussion and Conclusions} In this paper we have studied the color--magnitude relations of galaxies in RXJ\,0910+5422 to constrain their ages and formation histories. Our results show that the color--magnitude relation for the elliptical galaxies is consistent both in slope and scatter with that of RXJ\,1252.9-292 (Blakeslee et al. 2003a, Lidman et al. 2004) and recent results from Holden et al. (2004) and De Lucia et al. (2004), confirming that elliptical galaxies in galaxy clusters show a universal color--magnitude relation consistent with an old passively evolving population even at $z \sim 1$. From the color--magnitude relation of the ellipticals, we derive a mean luminosity--weighted age $\overline t > 3.3$~Gyr ($z_f > $~3). We find that the S0s in RXJ~0910+5422 define a color--magnitude sequence with a scatter similar to that found for the ellipticals, but shifted bluer by $0.07 \pm 0.02$~mag. This is peculiar with respect to previous cluster studies, which more typically found that the S0s followed the same CMR as the ellipticals, but with somewhat larger scatter (Bower et al. 1992; Ellis et al.\ 1997; Stanford et al.\ 1997; L\'opez--Cruz et al. 2004; Stanford, Eisenhardt, \& Dickinson 1998; van Dokkum et al. 2000, 2001; Blakeslee et al. 2003a; Holden et al. 2004). Only one earlier study, van Dokkum et al. (1998) found a significantly bluer S0 population. We examine this population of blue S0s in some detail, noting that there is a strong predominance of flattened systems with axial ratios $\frac{b}{a} > 0.7$, and conclude that the face-on members of the population have likely been classified as ellipticals. If so, the color offset between the two classes would become even more significant, and the true CMR scatter for the ellipticals would be slightly lower than we have estimated. This peculiarity is not observed in other clusters of our ACS Intermediate Redshift Cluster Survey, and its amplitude is smaller than the uncertainties adopted in Postman et al. (2005). If the observed color difference between the ellipticals and S0s is mainly due to metallicity at the same age, this would imply that the redder ellipticals were able to retain more metals than the S0s, i.e., they are more massive. However, current data suggest that the CMR is mainly the result of a metallicity--mass (i.e. metallicity--magnitude) relation (e.g., Kodama \& Arimoto 1997; Kauffman \& Charlot 1998 Vazdekis et al. 2001; Bernardi et al. 2005). This implies that we do not expect large metallicity variations at a given magnitude. If, instead, the offset is mainly due to age, then the implied age difference would be $\sim$1~Gyr for single-burst solar metallicity BC03 models. It could also result from different star formation histories, with the S0s experiencing a more extended period of star formation. A model with solar metallicity and with an exponential decay of the star formation reproduces the offset at a galaxy mean age of $\approx$~3.5~Gyr. The blue S0s may comprise a group infalling from the field onto a more evolved red cluster population, or they may be a transitional cluster population not yet evolved all the way onto the elliptical red sequence (van Dokkum \& Franx 2001). Assuming passive evolution, they will reach this red sequence after about 1~Gyr. High fractions of faint blue late type galaxies were observed in substructures infalling in a main cluster (e.g. Abraham et al. 1996; Tran et al. 2005a), and proposed as the progenitors of faint S0s in clusters. The view of this cluster as a structure still in formation is supported by X--ray observations of the cluster temperature structure (Stanford et al. 2002), the lack of a cD galaxy, and its filamentary structure that suggests merging of substructures. However, we derive a small cluster velocity dispersion, unusual for merging substructures. Moreover, the blue S0s in this sample span the same luminosity range of the bright ellipticals, are distributed towards the center of the cluster, and some of the faintest ones are physically associated with brighter ellipticals belonging to the central filamentary structure. These elements would argue against the bluer S0s being a young group merging with an existing red cluster population, and support the hypothesis that we are observing a transitional blue S0 population in a cluster core that is still evolving onto the elliptical red sequence. This result is also consistent with the deficit of S0s observed in our ACS cluster sample, when compared to lower redshift samples, that implies that the S0 population is not yet in place but still forming in clusters at redshifts around unity (Postman et al. 2005). Interestingly, we also observe in this cluster potential progenitors for bright S0 galaxies: four bright spirals (spectroscopically confirmed cluster members) with $z_{850}$ brighter than 22.5~mag and $(i_{775}-z_{850})$ between 0.5 and 1.3~mag. Red galaxy pairs are also observed. A triplet and three red galaxy pairs have projected distances less than $10 h^{-1}_{70}$~kpc, and of those, the triplet and one pair show zero relative velocity. This would be the evidence of red galaxy mergers at z$\sim$~1. van Dokkum (2005) and Tran et al. (2005b) have observed mergers of red galaxies in a nearby elliptical sample and in MS 1054-03 at z=0.83, respectively. They suggested a scenario in which most of the early--type galaxies were formed from passive red galaxy--galaxy mergers, called {\it dry} mergers, because they involve gas--poor early--type galaxies. Future papers will analyze the ages and masses of the cluster members using our optical spectroscopy along with newly obtained Spitzer IRAC imaging. A larger sample would be needed to draw firmer conclusions about the formation of S0s. \begin{acknowledgements} ACS was developed under NASA contract NAS 5-32865, and this research has been supported by NASA grant NAG5-7697 and by an equipment grant from Sun Microsystems, Inc. The {Space Telescope Science Institute} is operated by AURA Inc., under NASA contract NAS5-26555. We are grateful to K.~Anderson, J.~McCann, S.~Busching, A.~Framarini, S.~Barkhouser, and T.~Allen for their invaluable contributions to the ACS project at JHU. We thank W. J. McCann for the use of the FITSCUT routine for our color images. SM thanks Tadayuki Kodama for useful discussions. \end{acknowledgements} \newpage \clearpage \begin{small} \begin{table*} \begin{center} \caption{Color--Magnitude Relations \label{results}} \vspace{0.25cm} \begin{tabular}{cccccccccccc} \tableline \tableline\\ Sample &$N$&$c_0$&$Slope$ & $\sigma_{int}$\\ &&(mag)&&(mag)&\\ \tableline \tableline\\ E+S0+S0/a$^1$& 31& 0.99 $\pm$ 0.01& -0.030 $\pm$ 0.020& 0.060 $\pm$ 0.008\\ E+S0$^{1a}$& 14& 1.00 $\pm$ 0.02& -0.010 $\pm$ 0.033& 0.054 $\pm$ 0.009\\ E$^1$& 19& 1.02 $\pm$ 0.01& -0.033 $\pm$ 0.015& 0.042 $\pm$ 0.011\\ E$^{1a}$& 10& 1.02 $\pm$ 0.04& -0.020 $\pm$ 0.044& 0.047 $\pm$ 0.022\\ S0$^1$& 9& 0.95 $\pm$ 0.02& 0.005 $\pm$ 0.023& 0.044 $\pm$ 0.02\\ S0+S0/a$^{1}$& 12& 0.95 $\pm$ 0.02& -0.007 $\pm$ 0.027& 0.057 $\pm$ 0.015\\ \tableline \\ E+S0+S0/a$^2$& 32& 0.99 $\pm$ 0.01& -0.032 $\pm$ 0.019& 0.060 $\pm$ 0.008\\ E+S0$^{2a}$& 15& 0.99 $\pm$ 0.02& -0.021 $\pm$ 0.034& 0.054 $\pm$ 0.009\\ E$^2$& 19& 1.02 $\pm$ 0.01& -0.033 $\pm$ 0.015& 0.042 $\pm$ 0.011\\ E$^{2a}$& 10& 1.02 $\pm$ 0.04& -0.020 $\pm$ 0.044& 0.047 $\pm$ 0.022\\ S0$^2$& 10& 0.95 $\pm$ 0.02& -0.012 $\pm$ 0.036& 0.051 $\pm$ 0.018\\ S0+S0/a$^{2}$& 13& 0.96 $\pm$ 0.02& -0.015 $\pm$ 0.033& 0.065 $\pm$ 0.015\\ \tableline \\ E+S0+S0/a$^3$& 34& 0.99 $\pm$ 0.01& -0.036 $\pm$ 0.018& 0.059 $\pm$ 0.008\\ E+S0$^{3a}$& 15& 0.99 $\pm$ 0.02& -0.022 $\pm$ 0.035& 0.054 $\pm$ 0.009\\ E$^3$& 20& 1.01 $\pm$ 0.01& -0.032 $\pm$ 0.015& 0.044 $\pm$ 0.010\\ E$^{3a}$& 10& 1.02 $\pm$ 0.04& -0.021 $\pm$ 0.046& 0.047 $\pm$ 0.022\\ S0$^3$& 11& 0.96 $\pm$ 0.02& -0.022 $\pm$ 0.038& 0.053 $\pm$ 0.015\\ S0+S0/a$^{3}$& 14& 0.96 $\pm$ 0.02& -0.024 $\pm$ 0.034& 0.065 $\pm$ 0.013\\ \tableline \tableline \end{tabular} \end{center} $1$: within 1~$\arcmin$ \\ $2$: within 1.5~$\arcmin$ \\ $3$: within 2~$\arcmin$ \\ $a$: only confirmed members \end{table*} \end{small}
Title: First scattered light images of debris disks around HD 53143 and HD 139664
Abstract: We present the first scattered light images of debris disks around a K star (HD 53143) and an F star (HD 139664) using the coronagraphic mode of the Advanced Camera for Surveys (ACS) aboard the Hubble Space Telescope (HST). With ages 0.3 - 1 Gyr, these are among the oldest optically detected debris disks. HD 53143, viewed ~45 degrees from edge-on, does not show radial variation in disk structure and has width >55 AU. HD 139664 is seen close to edge-on and has belt-like morphology with a dust peak 83 AU from the star and a distinct outer boundary at 109 AU. We discuss evidence for significant diversity in the radial architecture of debris disks that appears unconnected to stellar spectral type or age. HD 139664 and possibly the solar system belong in a category of narrow belts 20-30 AU wide. HD 53143 represents a class of wide disk architecture with characteristic width >50 AU.
https://export.arxiv.org/pdf/astro-ph/0601488
\title{First scattered light images of debris disks around HD 53143 and HD 139664} \author{Paul Kalas\altaffilmark{1}, James R. Graham\altaffilmark{1}, Mark C. Clampin\altaffilmark{2}, \& Michael P. Fitzgerald\altaffilmark{1}} \affil{} \altaffiltext{1}{Astronomy Department, University of California, Berkeley, CA 94720} \altaffiltext{2}{Goddard Space Flight Center, Greenbelt, MD 20771} \keywords{stars: individual(\objectname{HD 53143, HD 139664}) - circumstellar matter} \section{Introduction} The configuration of our solar system is perhaps the most significant starting point for our understanding of planet formation. Therefore a fundamental question is whether or not the architecture of our solar system is common relative to other planetary systems. One point of comparison is the structure of our Kuiper Belt relative to other systems, which are typically seen as debris disks in scattered light or thermal emission. In scattered light, some debris disks, such as $\beta$ Pic and AU Mic, have central holes, but are radially extended to hundreds of AU radii \citep{smith84, kalas04}. Other debris disks, such as HR 4796A and Fomalhaut consist of relatively narrow rings with sharp inner and outer boundaries \citep{schneider99, kalas05a}. However, a narrow-belt architecture has not previously been detected in scattered light among stars similar in spectral type and age to the Sun. HD 53143 (K1V) and HD 139664 (F5V) are two stars $\sim$18 parsec from the Sun known to have circumstellar dust due to excess thermal emission at far-infrared wavelengths (Aumann 1985; Stencel \& Backman 1991; Table 1). Various indicators place the age of HD 53143 at $1.0 \pm 0.2$ Gyr \citep{decin00, song00, nord04}, whereas HD 139664 may be a younger system with age $0.3^{+0.7}_{-0.2}$ Gyr \citep{lach99, montes01, mallik03, nord04}. For these two stars the infrared excess corresponds to a dust mass 3 - 10 times smaller than that of $\sim$10 Myr-old systems such as AU Mic and $\beta$ Pic \citep{kalas04}. Direct imaging of debris disks with masses this small is observationally challenging, but it is now feasible using the optical coronagraph in ACS. \section{Observations \& Data Analysis} We utilized the HST ACS High Resolution Camera (HRC) with a 1.8$\arcsec$ diameter occulting spot to artificially eclipse each star (Table 1; Fig. 1). Five F stars were observed in consecutive orbits, as were five K stars, in order to minimize differences in the point spread function (PSF) due to telescope thermal variations. Each PSF was subtracted using the four other stars in each set of observations. The relative intensity scaling and registration between images was iteratively adjusted until the residual image showed a mean radial profile equal to zero intensity. After the excess nebulosity was detected around HD 53143 and HD 139664, we determined that no surface brightness asymmetries were detected between each side of each disk. To improve the signal-to-noise, we mirror averaged the data (Figs. 2 \& 3). Mirror averaging splits the image into two halves along the axis that is perpendicular to the disk midplane and bisects the disk. One side is transposed onto the other side and the data are then averaged. Asymmetries due to scattering phase function effects will be coadded \citep{kalas96}. In effect mirror averaging doubles the integration time spent on the circumstellar disk given that the broad features between each side are symmetric. As a test, we also subtracted the two disk halves from each other and confirmed that the assumption of symmetry is valid. \section{Results} The two disks have different morphologies due to different inclinations and intrinsically different architectures. To quantify the viewing geometries and structural properties of the disks, we produce a series of simulated scattered-light disks that explore the parameters of inclination to the line of sight, inner and outer disk radius, and the radial and vertical variation of dust number density \citep[Fig. 2;][]{kalas96}. We reinsert the simulated disks across each star in a direction orthogonal to the observed midplanes and select those models that most closely resemble the properties of the observed disks. Table 1 summarizes our findings. The shape of the midplane surface brightness distribution differs significantly for each system (Fig. 3). The HD 53143 midplane surface brightness decreases monotonically with projected radius, approximately as $r^{-3}$, where $r$ is the projected radius. In the simulated disk, the radial number density distribution decreases as $q^{-1}$, where $q$ is radius in the disk cylindrical coordinate system. Our solar system's Zodiacal dust complex has a comparable dependence of grain number density as a function of radius ($q^{-1.34}$; Kelsall et al. 1998), controlled mainly by the force of Poynting-Robertson (PR) drag that causes small grains to spiral into the Sun \citep{burns79}. Therefore the HD 53143 disk may simply represent a population of unseen parent bodies that collisionally replenish dust that is redistributed radially by PR drag. Our simulations show that the outer radius of the observed disk is a sensitivity-limited value at approximately 6$\arcsec$ radius (110 AU). The inner radius is also sensitivity-limited to 3$\arcsec$ radius (55 AU). Therefore, the debris disk around HD 53143 is at least 55 AU wide. Material surrounding HD 139664, on the other hand, is confined to a narrow belt, as indicated by a turnover in the midplane surface brightness profile between 4.5$\arcsec$ and 5.5$\arcsec$ (79 - 96 AU; Fig. 3). Our disk simulations show that the peak in the dust distribution occurs at 83 AU, decreasing as $q^{-2.5}$ from 83 - 109 AU, and with a sharp outer truncation at 109 AU (Figs. 2 \& 3). We tested model disks that have outer radii $>$109 AU and found that these disks would have been detectable in our data as far as 10$\arcsec$ radius (175 AU; Fig. 3). Given an absence of significant gas, the belt-like nature of HD 139664 is most likely a structure related to planet formation. \citet{kenyon04} find that a dust belt with peak surface brightness at $\sim$80 AU radius forms within a planetesimal disk at age 400 Myr. The appearance of this belt signals the recent formation of a $\geq$1000 km planet that gravitationally stirs planetesimals in its viscinity. \citet{liou99} and \citet{moro02}, on the other hand, simulate the concentrations of dust in trans-Neptunian space that arise due to trapping in mean motion resonances. A natural explanation for the belt-like morphology of HD 139664 is that the $\sim$83 AU peak in the dust distribution corresponds to either an interior or exterior mean motion resonance created by a companion to HD 139664. Large grains may dominate the belt's peak at 83 AU, with smaller grains passing quickly through the resonance regions due to radiation forces. \section{Discussion} Taking a census of eight debris disks resolved in scattered light and the predicted dust distribution in our Kuiper Belt, we observe two basic architectures that are not correlated with stellar mass and luminosity, but must depend on other environmental factors (Table 2). Debris systems are either narrow belts or wide disks. Both types have central dust depletions, but the distinguishing characteristic is the presence or absence of a distinct outer edge. HR 4796A, Fomalhaut, HD 139664 and the Sun are examples of narrow-belt systems. The belt systems appear to have radial widths ranging between 20 and 30 AU, and the inner edges may begin as close as 25 AU (Sun), or as far as 133 AU (Fomalhaut). HD 32297, $\beta$ Pic, HD 107146, HD 53143 and AU Mic are examples of disks with sensitivity-limited outer edges that imply disk widths $>$50 AU. The F, G, and K stars are interesting because {\it a priori} we might expect to find planetary systems similar to our own, yet we discover significant diversity in the outer regions that correspond to Neptune and our Kuiper Belt. With age $\sim$1 Gyr, HD 53143 is among the oldest known extrasolar debris disks, yet its wide-disk architecture resembles that of the $\sim$10 Myr old systems of $\beta$ Pic and AU Mic. The lingering dust mass (Table 1) throughout the system could signal the absence of giant planets that otherwise sweep clear the parent bodies (comets and asteroids) responsible for the producing the dust disk. Yet the presence of a dust disk out to at least 110 AU radius shows that the primordial circumstellar disk probably contained the prerequisite mass of gas and dust to form giant planets. By contrast, the younger system HD 139664 has already developed a narrow-belt architecture. Narrow-belt architectures for the underlying population of planetesimals may originate from early stochastic dynamical events, such as a close stellar flyby, that strip disk mass and dynamically heat the surviving disk \citep{ida00, adams01}. Theoretical simulations show that a reduction in disk mass, combined with dynamical heating, produces a less stable planetary system that is more likely to eject giant planets from their formation site to much larger radii, as has been proposed for the origin of Neptune \citep{thommes99,tsiganis05}. Therefore, planetesimal belts not only evolve into a narrow structures because of external stochastic events, but they may be found at large distances from the central star due to the subsequent outward migration of interior planets. The collisionally replenished dust population will spread away from any narrow belt of planetesimals. If the scattered light appearance continues to manifest a narrow structure, then both the inner and outer edges are probably maintained by other gravitational perturbers such as stellar or sub-stellar companions. However, only HR 4796 has a known stellar companion that may truncate the outer radius of the dust belt \citep{aug99}. If there is no confinement mechanism for the outer radius, then the architecture will manifest as a wide-disk. For example, $\beta$ Pic and AU Mic may have narrow belts of planetesimals \citep{aug01, strubbe05}, but the observed dust disk widths extend to hundreds of AU. Though the predicted structure of our Kuiper Belt places the solar system among the narrow disk architectures, it is conceivable that the dust component extends to greater radii \citep{trujillo01}. Thus the Sun's classification as a narrow-disk system is tentative. \section{Summary} We present the first optical scattered light images of debris disks surrounding relatively old main sequence F and K stars. Material around HD 139664 is concentrated at 83 AU radius, with a distinct outer edge at 109 AU, and a depleted, but not empty, region at $<$83 AU radius. Dust surrounding HD 53143 has a monotonic $q^{-1}$ variation in grain number density, and the disk edges from 55 AU to 110 AU are sensitivity-limited values. The different radial widths appear consistent with a more general grouping of debris disks into either narrow or wide architectures. These two categories are probably an oversimplification of significant diversity in the formation and evolution of debris disks. Future observations should test for common traits among these stars, such as stellar multiplicity and the existence of planets. \acknowledgements {\bf Acknowledgements:} Based on observations with the NASA/ESA Hubble Space Telescope obtained at the Space Telescope Science Institute (STScI), which is operated by the Association of Universities for Research in Astronomy. Support for Proposal number GO-9475 was provided by NASA through a grant from STScI under NASA contract NAS5-26555. \clearpage \clearpage \clearpage \begin{deluxetable}{lllll} \tabletypesize{\scriptsize} \tablecaption{Star (rows 1-7) and disk (rows 8-15) properties \label{tbl-1}} \tablewidth{0pt} \tablehead{ \colhead{} & \colhead{HD 53143} & \colhead{HD139664} } \startdata Age (Gyr) &1.0$\pm$0.2 &$0.3^{+0.7}_{-0.2}$\\ Spectral Type &K1V & F5IV-V \\ Mass (M$_\odot$) &0.8 &1.3\\ T$_{eff}$ (K) &5224 & 6653\\ Luminosity (L$_\odot$) &0.7 &3.3\\ Distance (pc) &18.4 & 17.5 \\ $m_V$ (mag) &6.30 & 4.64 \\ &&\\ Peak disk surf. bright. (mag/arcsec$^{2}$) &22.0$\pm$0.3 &20.5$\pm$0.3\\ Disk position angle (degrees) &147$\pm$2 &77 $\pm$ 0.5\\ Inclination (degrees) &40 - 50 &85 - 90\\ Disk number density gradients ($q^{\alpha}$) &-1 &+3.0 \& -2.5 \\ Inner dust depletion (AU) & $<55$ &83 \\ Maximum outer radius (AU) & $>110$ &109 \\ Optical depth from IRAS data\tablenotemark{a} &$2.5\times10^{-4}$ & $0.9\times10^{-4}$ \\ Optical depth from HST data\tablenotemark{b} &$>1.6\times10^{-5}$ &$1.0\times10^{-5}$ \\ Total Dust Mass (g)\tablenotemark{c} &$>7.1\times10^{23}$& $5.2\times10^{23}$ \\ \enddata \tablenotetext{a}{The fractional dust luminosity \citep{zuck04}. } \tablenotetext{b}{Derived from the model disks. Since the disks are optically thin, we sum the cumulative light from the model disk, $m_d$, and quote optical depth as $10^{(m_d - m_V) / -2.5)}$. We assume albedo=1 and the optical depth will scale inversely with the assumed albedo. } \tablenotetext{c}{Dust mass follows from the HST optical depth and assumes a uniform particle radius of 30 $\mu$m, density 2.5 g cm$^{-3}$ and albedo=1.0. } \end{deluxetable} \clearpage \begin{deluxetable}{lllllll} \tabletypesize{\scriptsize} \tablecaption{Debris disk architectures from scattered light properties\tablenotemark{a} \label{tbl-1}} \tablewidth{0pt} \tablehead{ \colhead{Name} & \colhead{$r_{in}$ (AU)} & \colhead{$r_{out}$ (AU)} & \colhead{Width (AU)} & \colhead{d (pc)} & \colhead{SpT} & \colhead{References} } \startdata $\beta$ Pic &$\sim$90 &$>$1835 & \bf{$>$1745} & 19.3 & A5V & 1, 2\\ HD 32297 &$<$40 &$>$1680 & \bf{$>$1640} & 112 & A0 & 3, 4\\ AU Mic &$\sim$12 &$>$210 & \bf{$>$198} & 9.9 & M1Ve & 5, 6\\ HD 53143 &$<$55 &$>$110 & \bf{$>$55} & 18.4 & K2V & This work.\\ HD 107146 &$\sim$130 &$>$185 & \bf{$>$55} & 28.4 & G2V & 7\\ &&\\ HD 139664 & 83 & 109 & \bf{26} & 18.5 &F5V & This work.\\ Fomalhaut & 133 & 158 & \bf{25} & 7.7 & A3V & 8\\ HR 4796A & 60 & 80 & \bf{20} & 67.1 & A0V & 9\\ Sun\tablenotemark{b} & 25 & 50 & \bf{25} & -- & G2V & 10, 11 \enddata \tablenotetext{a}{We do not include systems younger than 10 Myr that are likely to possess significant primordial circumstellar gas. Column 2 gives the inner disk radius corresponding to the approximate peak of dust number density. Column 3 gives the outer radius. } \tablenotetext{b}{From simulations of relatively large grains trapped in resonances with Neptune. } \tablecomments{REFERENCES: (1) \citet{pantin97}; (2) \citet{larwood01}; (3) \citet{ssh05}; (4) \citet{kalas05b}; (5) \citet{krist05}; (6) \citet{kalas04}; (7) \citet{ardila04}; (8) \citet{kalas05a}; (9) \citet{schneider99}; (10) \citet{liou99}; (11) \citet{moro02} } \end{deluxetable}
Title: Morpho-kinematic modeling of gaseous nebulae with SHAPE
Abstract: We present a powerful new tool to analyse and disentangle the 3-D geometry and kinematic structure of gaseous nebulae. The method consists in combining commercially available digital animation software to simulate the 3-D structure and expansion pattern of the nebula with a dedicated, purpose built rendering software that produces the final images and long slit spectra which are compared to the real data. We show results for the complex planetary nebulae NGC 6369 and Abell 30 based on long slit spectra obtained at the San Pedro Martir observatory.
https://export.arxiv.org/pdf/astro-ph/0601585
\section{Introduction} \label{sec:introduction} In recent years, the discovery of a variety of complex structures in planetary nebulae has opened many questions regarding the origin and evolution of these objects (e.g. L\'opez 2000). Deviations from simple expanding shells can include collimated outflows, poly-polar and point-symmetric structures, rings or disks. These observations have lead to a wealth of theoretical research into the effects of stellar magnetic fields, rapid rotation and binarity of the central stars, and their evolutionary path from spherically symmetric to bipolar mass-loss (e.g. Balick \& Frank, 2002, and references therein). In the absence of spherical symmetry, the tilt of the nebula with respect to the line of sight and the location and position angle of the slit on the nebula can often result in complicated position-velocity (P-V) diagrams that can be difficult to interpret. The correct interpretation of the nebular 3-D geometry and kinematic structure of PNe is key to the understanding of the dynamics ruling their origin and evolution. Modeling of line emission intensity maps have been used to obtain density distributions over the face of the nebula in order to assess 3-D structures, assuming pure photoionization from the central star (e.g. Monteiro et al. 2004; Morisset, Stasi\'nska \& Pe\~na, 2005) but without incorporating kinematic information or assuming simple velocity laws (e.g. Ragazzoni et al. 2001). In this paper we present a new interactive 3-D modeling tool called \shape which combines the versatility of commercial 3-D modeling software with a rendering module specifically developed for application in astrophysical research. The application of this method yields a 3-D emissivity and velocity distribution for the object. Furthermore, different velocity laws can be applied to different sections of the nebula to reproduce the complex velocity patterns often observed in PNe. We exemplify the power of this new method with models of the particularly complex planetary nebulae Abell 30 and NGC 6369. \section{Problem Definition} The problem we attack with \shape is to characterize the current 3-D morphology and velocity field of a nebula based on imagery and spectral kinematic information. Detailed knowledge of this information leads to a better understanding of the physical structure and dynamical evolution of a gaseous nebula. The projected image on the sky of an extended nebula provides bidimensional spatial information of its structure. On the other hand, the velocity field provides information on the radial component of the velocity vector along the line of sight and conveys limited but useful information on its depth or third spatial dimension. However, an unambiguous solution of the complete 3-D structure at least requires full knowledge of the velocity field. This situation is usually not given, although topological and symmetry information apparent in the images and spectra may help resolve ambiguities. The simplest case occurs if the velocity of a volume element is constant over most of the expansion time. In complex objects, this type of velocity distribution can be expected if the nebula has evolved from a relatively short mass-loss event and is now moving ballistically (e.g. Zijlstra et al. 2001) or from a continuous interaction of a wind with small scale structures (Steffen \& L\'opez, 2004). In these cases, after a sufficiently long time, the velocity pattern becomes proportional to the distance from the center (a hubble-like velocity law). The expansion of such a nebula is self-similar, i.e. the global shape is conserved over time. Under such conditions, the velocity vector is proportional to the position vector of every material element in the nebula. The shape of the nebula along the line of sight is mapped linearly into the corresponding component of the velocity vector. Hence, the Doppler-shift, which is equivalent to the velocity component along the line of sight, is a map of the structure that is lost in a direct image, i.e. the long-slit spectrum allows a view of the nebula from a direction perpendicular to the line of sight. This situation can clearly be appreciated in the case of axisymmetric bipolar nebulae, where the line profiles also show a bipolar structure that represents the depth or third dimension of the 2-D image on the sky. In many planetary nebulae an expansion velocity proportional to distance from central star seems to be a reasonable approximation at least for the brightest regions (e.g. Wilson 1950, Weedman 1968). However, more complex velocity structures can be expected when one or more mass loss events arise over a significant timescale compared with the age of the nebula. Sabbadin et al. (2000) have used the assumption of a radial velocity field proportional to the distance from the central star to reconstruct the 3-D structure of a number of nebulae with a "tomographic" method. This tomography works well as long as there are no significant deviations from the hubble-like velocity law. \section{The 3D modeling system} In this section we describe our new version of the code \shape as a tool to find the 3-D structure and kinematics of gaseous nebulae. Originally \shape was based on a description of structure and kinematics using parametric geometrical equations on a regular 3-D grid (Steffen et al. 1996). This code was adapted with a simple graphical interface by Harman et al. (2004). It has been applied to a variety of objects from individual knots in planetary nebulae (L\'opez, Steffen \& Meaburn, 1997) to moderately complex structures in active galactic nuclei (Steffen et al. 1996). In the present upgraded version we have devised a completely different approach based on particle systems, rather than a regular grid, in combination with a commercial 3-D animation package, which we describe next. As our 3-D modeling software we use {\em Autodesk 3DStudio Max 7} ({\em 3DStudio Max} is a trademark of {\em Autodesk Media and Entertainment}, see the website {\em www.discreet.com} for detailed software information). We apply the available tools of this software to create a particle and velocity distribution in space and time in order to model an object. In particular we use the {\it ParticleFlow} particle system to generate particle distributions which are then exported and rendered in \shape. \shape renders images and spectral information from the kinematics of the model particle distribution. Key parameters such as orientation of the object on the sky, location and width of the slit, seeing values and spectral and spatial resolution are handled interactively in the graphical interface of \shape. The general modeling process is as follows. With the inspection of available observations one obtains an initial rough idea of the structure and topology of the object, which is then reproduced in the modeling software. For this purpose one produces a distribution of particles in space with its corresponding velocities. The particles may be distributed over a topologically complex surface or throughout a volume. The resulting emissivity is integrated along the line of sight. The modeling software allows a very complex object to be built and since all the features may be variable in time, the time evolution of the object may also be explored. As a guide during modeling, a spectal preview feature has been developed. For a limited number of particles it allows a rough version of the P-V diagrams to be visualized during modeling in the viewport of {\em 3DStudio Max}. The particle data are then exported into an external file. This file is read by the core code of \shape, which produces the rendered image and corresponding P-V diagram. In contrast to the real-time feature described above, the core code of \shape is not limited by the number of particles. It has been tested with up to one million particles. The rendered images are then compared with the observations. To improve a model, changes may be introduced at any point of the procedure until, after a number of iterations, a satisfactory approximation to the observations are obtained. Alternative solutions to an ambiguous dataset will not be automatically found, but may be sought for separately. The processing of the data, including the viewing direction, the emissivity, internal structure, size and other parameters of individual particles is controled via a dedicated graphical interface. This interface has been programmed as a module of the modeling software {\em 3DStudio Max} and is fully integrated in its interface. An important analytic feature of \shape is the possibility to apply different colors to sections of a complex object. This allows a clear distinction of them in the P-V diagrams, which helps considerably in the interpretation of the observed spectra. By combining models of various emission lines, color images are obtained to be compared with similar images from observations (see Figure \ref{fig:images}). A red-blue coloring mode allows a clear distinction between red and blue-shifted regions in the model image. Sequences of varying parameters like slit position and width, as well as orientation of the object allow a systematic search for the best parameters that match the observations or the production of image sequences for animations which help considerably to visualize the 3-D structure of the object. The P-V diagrams of objects that have been modeled as evolving in time can also be visualized as a time sequence in animation form. The most important current limitation of \shape is that only optically thin nebulae can be modeled. The code does not perform any physical radiation transport or line emission calculation. What it does is to directly assign a relative emissivity distribution, which is sufficient for its main purpose, the characterization of the structure and kinematics of a nebula. Moreover, \shape can also be used to produce complex density distributions and kinematics from which photoionization models can be calculated using codes like NEBU\_3D (Morisset et al, 2005). The current apparent drawbacks can hence be largely overcome by combining \shape with codes which calculate radiation transport. We aim at this next step in the near future. \subsection{The rendering code} The core code of \shape renders images and P-V diagrams from the position and velocity data provided with {\em 3DStudio Max}. Each particle is read from a file and its emission mapped to the image and spectrum. The intensity and color of the emission from a particle may depend on the position in the object or on the substructure of which it is part. The particles may also have a finite size larger than single image pixels, as a well as an internal emissivity structure. At this time this internal emissivity distribution may be either constant, an exponential or a gaussian fall-off. This is useful mostly for the initial modeling stages, before the convolution with the seeing parameters are applied, because convolution eliminates all internal particle structure. Particle sizes smaller than the seeing together with a sufficiently large number of particles ensure adequate sampling of the object structure. The image blurring due to seeing is modeled by convolution with a gaussian point-spread function and including emission from within one FWHM of the seeing disk outside the spectrograph slit. The contribution of this emission decreases with distance from the edge of the slit. The instrumental resolution is included in the convolution of the raw image and P-V diagram with gaussian kernals of FWHM corresponding to the spatial and velocity resolution of the instruments. A Fast-Fourier-Transform algorithm is applied to calculate the convolutions with gaussian kernels. After convolution, the P-V diagrams can be used to obtain one-dimensional line profiles adding all or sections of the emission along the spacial domain, which are often useful in work with low spatial resolution on the nebulae. The core code of \shape has been programmed updating the Fortran code of the original version. In the following sections we present two example models to reproduce observations of objects with different degrees of complexity. An additional good example for an application of \shape on the complex structure of NGC~6302 can be found in Meaburn et al. (2005). \section{Observations and models} \subsection{NGC~6369} Our model of NGC~6369 is based on a long-slit spectrum obtained with MES (Meaburn et al. 2003) on the 2.1 m telescope at the San Pedro Martir Observatory. The spectral resolution is 10 km s$^{-1}$ and the seeing 1.5 arcsec. The image has been obtained from the Hubble Heritage Team, NASA, STScI (Figure \ref{fig:images}). The spectrograph slit was located in the east-west direction, which is coincident with the main axis of the object (as indicated in the model image in Figure \ref{fig:images}). The slit width corresponds to 1.5 arcsec on the sky. The observed spectra (Figure \ref{fig:spectra}, left) show that the central "barrel" and the "lobes" are connected and probably conform a single topological closed surface. Therefore we have started the modeling process with an initially spherical shape and deformed it such that it matches our initial estimate of the 3-D shape according to the image and spectrum. We then applied a brightness map to the surface which approximates the emissivity distribution from the H$\alpha$ and NII emission lines. The structure and emission distribution is then adjusted interactively until a satisfactory match was found. Figure \ref{fig:modvec} shows the 3-D structure of NGC~6369 depicted as a central ellipsoid with two opposite, off-axis, protruding lobes. We have modeled the nebula's expansion with two different velocity laws. Based on the observation that the highest and lowest projected velocities in the spectra are similar for the barrel and the lobes, we considered a constant expansion velocity directed perpendicularly to the surface. This corresponds roughly to an energy driven bubble (in this case three different bubbles, the barrel and the two lobes). In this case we did not find any shape that would reasonably resemble the image and spectra simultaneously. The second model assumes that the velocity is proportional to distance and the direction of motion is radial from the center of the nebula. The resulting surface model with representative velocity vectors is shown in Figure \ref{fig:modvec}. For the H$\alpha$ emission we assume that emission comes from a region somewhat inside this surface and for NII slightly outside, with some overlapping around the surface that is shown (as seen in the observed spectra and images). The rendered image and spectrum are shown in Figures (\ref{fig:images}) and (\ref{fig:spectra}), respectively. The particle number density is proportional to the brightness on the 3-D surface (see Fig.\ref{fig:modvec}), which is not accurately represented in the image due to lighting effects of this visualization, which emphasizes the topology of the 3-D structure. Our modeling indicates, that the line of sight is close to the ellipsoid's symmetry axis (tilt $\approx 15^\circ$) and the axis of the lobes is tilted approximately $40^\circ\pm 10^\circ$ with respect to that axis. The ratio between the height and the diameter of the barrel is not well constrained by the current modeling and observations. We estimate that this ratio is of the order $3/2$. The model images and P-V diagrams have spatial and spectral resolution corresponding to those in our observations, which are 1.5 arcsec and 10 km s$^{-1}$ using a slitwidth of 1.5 arcsec in the simulation. We find that the model with the velocity law proportional to distance produces an acceptable fit to the observed image and spectrum. This suggests that this flow is in a relaxed momentum driven state; though the lobes appear to move slightly slower than expected in this case (Figure \ref{fig:spectra}). This might be an indication for a short-lived collimated outflow, which is not acting on the lobes anymore, and/or a stronger interaction of the lobes with the ambient medium due to their lower densities. Monteiro et al. (2004) proposed a ``diabolo''-type structure for NGC~6369 based on the analysis of spectral imaging data of a number of ions. The diabolo model in this case necessarily implies a narrow waist in the equatorial plane with its symmetry axis close to the line of sight. With this orientation such a narrow torus-like waist is expected to have very low velocity along the line of sight, and in the H$\alpha$ long\-slit spectrum it should appear as a narrow feature near the systemic velocity. The observations presented in Figure \ref{fig:spectra} do not show evidence for a narrow low-velocity waist. However, bright emission regions near the systemic velocity in the line profiles are apparent. In our model this is due to both, an intrinsic equatorial density enhancement as seen in the image (Figure \ref{fig:images}) as well as a long tangential line of sight through the spheroidal main nebula. Thus, our model does not require a waist to reproduce this feature. NGC~6886 and NGC~6565 are two remarkably similar nebulae to NGC~6369, both in terms of morphology and line-profile structure. Turatto et al. (2002) have modeled the structure of NGC~6565 with their tomographic method obtaining similar results to ours. \subsection{Abell 30} Abell~30 is a hydrogen deficient planetary nebula with a spherical [OIII] shell and complex knotty and filamentary structures in the central region. The inner region contains unusual cometary knots with large velocity spikes in the P-V diagrams. Meaburn \& L\'opez (1996) describe its kinematics as "dramatic". In this section we present a model produced with \shape as a further illustration of its flexibility to handle complex kinematic structures which may be very different from a hubble-like velocity law. Observations of Abell~30 have been adopted from figures 1 and 2a in Meaburn \& L\'opez (1996) which are reproduced here in Figure \ref{fig:abell30_obs}. The observations may be compared with our model model image and P-V diagram for one slit position as shown in Figure (\ref{fig:abell30}). The outer shell was modeled as a sphere with some variations in brightness and a constant radial expansion. A random distribution of knots and filaments was used for the inner regions. For the cometary knots, particles where emitted from discrete points which then interacted with a central spherically symmetric wind, accelerating the particles outwards and producing the high-speed features in the P-V diagrams. Support for the existence of such a wind in Abell~30 comes from observations of X-rays (Chu, Chang \& Conway, 1997). For the model we adapted the parameters similar to those of the observations with seeing between 1 and 2 arcsec and a width of the slit of 1.9 arcsec. For the model we used 1.5 and 1.9 arcsec respectively. Since the object is basically spherically symmetric, except for the random distribution of brightness variations of the outer shell and the distribution of knots in the central region, the specific orientation of the model is not a fundamentally important parameter. At this time no attempt was made to match individual small-scale features with those in Abell~30. Still, the spectrum in Figure \ref{fig:abell30} matches the observed one very well. This model shows how different methods can be used to combine substructures which have a variety of complexity and kinematic signatures. \section{Discussion and Conclusions} In this paper we have presented the upgraded version of \shape, with a novel approach to the determination of the three-dimensional structure and kinematics of gaseous nebula. \shape combines the capabilities of commercial 3-D modeling software with a purpose-built rendering software and graphical control interface for its application to astrophysical nebulae. \shape produces images and P-V diagrams for highly complex nebulae. The results can be compared directly with observations and help to understand structure, kinematics and orientation of the objects. The kinematics as observed from long-slit spectra often provides sufficient information about the 3-D structure and topology of an object for the modeling with \shape to provide a self-consistent 3-D structure. In addition the \shape models can be applied directly as input density distributions for photo-ionization codes. In \shape one can define and combine different velocity laws that fit complex structures. Solutions are expected to be most reliable if the object shows evidence for a significant degree of symmetry, then the full 3-D structure and kinematics can be deduced unambiguously. In other cases the object may be devided into regions or subsystens which allows them to be solved separately. We are currently building a catalogue of of synthetic emission line profiles with \shape that should be a useful reference to interpret long-slit observations of PNe with diverse morphologies and orientations. In this paper we showed applications of \shape to the planetary nebulae Abell~30 and NGC~6369. As a result of our application of \shape to new observations of the planetary nebula NGC~6369, we propose that its basic structure is that of a ellipsoidal or barrel-shaped main nebula with bipolar protrusions at a large angle to the symmetry-axis of the main nebula. Abell~30 has been modeled combining different velocity patterns to fit the expanding shell and the inner high-velocity knots. \acknowledgements We acknowledge support from DGAPA-UNAM projects IN111803 and IN112103 as well as CONACYT projects 37214 and 43121.
Title: Two-body problem with the cosmological constant and observational constraints
Abstract: We discuss the influence of the cosmological constant on the gravitational equations of motion of bodies with arbitrary masses and eventually solve the two-body problem. Observational constraints are derived from measurements of the periastron advance in stellar systems, in particular binary pulsars and the solar system. Up to now, Earth and Mars data give the best constraint, Lambda < 10^{-36} km^{-2}; bounds from binary pulsars are potentially competitive with limits from interplanetary measurements. If properly accounting for the gravito-magnetic effect, this upper limit on $\Lambda$ could greatly improve in the near future thanks to new data from planned or already operating space-missions.
https://export.arxiv.org/pdf/astro-ph/0601612
\title{Two-body problem with the cosmological constant and observational constraints} \author{Philippe Jetzer} \email{jetzer@physik.unizh.ch} \author{Mauro Sereno} \email{sereno@physik.unizh.ch} \affiliation{Institut f\"{u}r Theoretische Physik, Universit\"{a}t Z\"{u}rich, Winterthurerstrasse 190, CH-8057 Z\"{u}rich , Switzerland} \date{November 21, 2005} \pacs{04.25.Nx,04.80.Cc,95.10.Ce,95.30.Sf,95.36.+x,96.30.-t,97.60.Gb,98.80.Es} \keywords{cosmological constant, pulsars, solar system} \section{Introduction} Aged nearly one century, Einstein's cosmological constant $\Lambda$ still keeps unchanged its cool role to solve problems. $\Lambda$, despite being just one number, was able to respond to very different needs of the scientific community, from theoretical prejudices about universe being static (which provided the original motivation for introducing $\Lambda$ in 1917) to observational hints that the universe is dominated by unclustered energy density exerting negative pressure, as required by data of exquisite quality which became available in the last couple of decades. Although it is apparently plagued by some theoretical problems about its size and the coincidence that just in the current phase of the universe the energy contribution from $\Lambda$ is of the same order of that from non-relativistic matter, the cosmological constant still provides the most economical and simplest explanation for all the cosmological observations \cite{pad05}. The interpretation of the cosmological constant is a very fascinating and traditional topic. $\Lambda$ might be connected to the vacuum density, as suggested by various authors (see \cite{pe+ra03} for an historical account), and could offer the greatest contribution from cosmology to fundamental physics. The big interest in the cosmological constant has recursively raised attempts in putting observational bounds on its absolute value from completely different phenomena. $\Lambda$, supposed to be $\sim 10^{-46}\mathrm{km}^{-2}$ from observational cosmology analyses, is obviously of relevance on cosmological scales but it could play some role also in local problems. Up to now, no convincing methods for constraining $\Lambda$ in a laboratory have been proposed \cite{je+st05}, but interesting results have been obtained considering planetary motions in the solar system \cite{isl83,wri98,ker+al03}. The effects of $\Lambda$ become stronger for diluted mass conglomeration but they get enhanced also through various mechanisms \cite{now+al02,ba+no05}. As an example, conditions for the virial equilibrium can be affected by $\Lambda$ for highly flattened objects \cite{now+al02}. On the scale of the Local Volume, a cosmological constant could have observable consequences by producing lower velocity dispersion around the Hubble flow \cite{tee+al05}. Up to now, local physical consequences of the existence of a cosmological constant were investigated studying the motion of test bodies in the gravitational field of a very large mass. This one-body problem can be properly considered in the framework of the spherically symmetric Schwarzschild vacuum solution with a cosmological constant, also known as Schwarzschild-de Sitter or Kottler space-time. The rotation of the central source can also be accounted for using the so-called Kerr-de Sitter space-time \cite{ker+al03}. Here, we carry out an analysis of the gravitational $N$-body problem with arbitrary masses in the weak field limit with a cosmological constant. This study is motivated by the more and more central role of binary pulsars, from the discovery of the pulsar PSR B1913+16 in 1974 \cite{hu+ta75}, in testing gravitational and relativistic effects. The gravitational two-body equations of motion for arbitrary masses were first derived in absence of spin by Einstein, Infeld and Hoffmann (EIH) \cite{ein+al38}. The problem was later addressed in more general cases, subsequently accounting for spins and quadrupole moments \cite[and reference therein]{ba+oc75}. Here, we take the further step to consider a cosmological constant. The paper is as follows. In section~\ref{sec:field}, we discuss the gravitational weak field limit in presence of a cosmological constant and introduce the relevant approximations. Section~\ref{sec:EIHeq} presents the generalization of the EIH equations of motion, whereas section~\ref{sec:twobo} is devoted to the study of the two-body problem. In section~\ref{sec:obser}, we review how measurements of precession of pericentre in stellar system can constrain $\Lambda$. In particular, we consider binary pulsars and the solar system. Section~\ref{sec:concl} contains some final considerations. \section{Field equations with a cosmological constant in post-Newtonian approximation} \label{sec:field} Einstein's equations with the cosmological constant are \begin{equation} R_{\mu \nu} - \Lambda g_{\mu \nu} = \frac{8 \pi G}{c^4} S_{\mu \nu} , \end{equation} where $G$ is the gravitational constant, $c$ the vacuum speed of light and \begin{equation} S_{\mu \nu} \equiv T_{\mu \nu} - \frac{1}{2} g_{\mu \nu} T^{\lambda}_{\lambda}, \end{equation} with $ T_{\mu \nu}$ being the energy-momentum tensor. The weak field expansion can start by introducing a nearly Lorentzian system for weak, quasi-stationary fields, in which \begin{equation} g_{\mu \nu} = \eta_{\mu \nu} + h_{\mu \nu}, \ \ \ |h_{\mu \nu}| \ll 1. \end{equation} Actually, the Minkowski metric $\eta_{\mu \nu}$ is not a vacuum solution of the field equations with a cosmological constant, but for $ | \Lambda | \ll 1$ an approximate solution in a finite region can still be found by an expansion around $\eta_{\mu \nu}$. In the post-Newtonian (pN) approximation, metric components can be expanded in powers of \begin{equation} \varepsilon \sim \left( \frac{G M}{c^2 R} \right)^{1/2} \sim \frac{v}{c} \sim \frac{p}{\rho}, \end{equation} where $M$, $R$, $v$, $p$ and $\rho$ represent typical values for the mass, length, velocity, pressure and energy density of the system, respectively. In what follows, $^{(n)}g_{\mu \nu}$ and $^{(n)}T_{\mu \nu}$ will denote terms of order $\varepsilon^n$ and $\varepsilon^n (M/R)$, respectively. To perform a proper treatment in presence of a $\Lambda$-term we have to consider the suitable approximation order for $\Lambda$. We assume that the size of the contributions due to the cosmological constant is at most comparable to the post-Newtonian terms, i.e, ${\cal{O}}(\Lambda g_{00}) \gs {\cal{O}} (G\ ^{(2)} T^{00}/c^2)$. This condition can be rewritten as \begin{equation} \label{fie1a} \Lambda \ls \frac{R_\mathrm{g}^2}{R^4} , \end{equation} where $R_\mathrm{g} \equiv G M/c^2$ is the gravitational radius. Eq.~(\ref{fie1a}) is easily satisfied by gravitational bound systems with $M \sim M_\odot$ and $ R\sim 1-10^2~\mathrm{AU}$ if $\Lambda \ls 10^{-33}~\mathrm{km}$, a value well above the estimated one from cosmological constraints and also greater than the limits we will derive in section \ref{sec:obser} considering stellar systems. Hereafter, we will put $c=1$. With such an approximation order, we can use classical results within the standard pN gauge. Following \citet{str04}, the approximate field equations read \begin{eqnarray} \Delta^{(2)} g_{00} & = & -8 \pi G \ ^{(0)} T^{00}, \\ \Delta^{(4)} g_{00} & = & \ ^{(2)}g_{ij}\ ^{(2)}g_{00,ij}+ ^{(2)}g_{ij,j}\ ^{(2)} g_{00,i} - \frac{1}{2}\ ^{(2)} g_{00,i}\ ^{(2)} g_{00,i} - \frac{1}{2}\ ^{(2)} g_{00,i}\ ^{(2)} g_{jj,i} \\ & -& 8 \pi G \left(\ ^{(2)} T^{00}-2^{\ (2)}g_{00}\ ^{(0)}T^{00} +^{\ (2)}T^{ii} \right) +2 \Lambda , \nonumber \\ \Delta^{(3)} g_{oi} & = & - \frac{1}{2}\ ^{(2)}g_{jj,0i}+ ^{\ (2)}g_{ij,0j} + 16 \pi G ^{\ (1)} T^{i0} , \\ \Delta^{(2)} g_{ij} & = & - 8 \pi G \delta_{ij} \ ^{(0)} T^{00} . \end{eqnarray} The components of the metric can be expressed in terms of potentials. Let $\phi_\mathrm{N}$ be the Newtonian potential, \begin{equation} \phi_\mathrm{N} = -G \int \frac{^{(0)} T^{00} (t,\bfx^{'})}{|\bfx-\bfx^{'}|}d^3 x^{i} . \end{equation} According to our approximation order, the cosmological constant appears only in the equation for $^{(4)} g_{00}$. This can be re-arranged to give \begin{equation} \Delta^{(4)} ( g_{00} +2 \phi_\mathrm{N}^2 )= -8 \pi G \left(\ ^{(2)} T^{00} + ^{(2)} T^{ii} \right) +2 \Lambda \end{equation} Together with the classical pN potential $\psi$, \begin{equation} \psi = -G \int \frac{d^3 x^{'}}{|\bfx-\bfx^{'}|} \left( \ ^{(2)}T^{00}+ ^{(2)} T^{ii} \right) , \end{equation} we introduce $\phi_\Lambda$, solution of \begin{equation} \Delta \phi_\Lambda = -\Lambda . \end{equation} In presence of a cosmological constant, there is an upper limit on the maximum distance within which the Newtonian limit holds and boundary conditions must then be chosen at a finite range \citep{now01}. When these boundary conditions are chosen on a sphere whose origin coincides with the origin of the coordinate system, $\phi_\Lambda$ can be expressed as \begin{equation} \phi_\Lambda = -\frac{1}{6} \Lambda |\bfx|^2 , \end{equation} where we have neglected correction terms which appear because of boundary conditions. Due to a positive cosmological constant, the origin of the coordinate system has a distinguished dynamical role with a radial force directed away from it \cite{adl+al65}. Since the choice of the origin is arbitrary, any point in the space will experience repulsion from any other point. Finally, introducing the standard pN potentials, \begin{eqnarray} \xi_i & = & - 4 G \int \frac{d^3 x^{'}}{|\bfx-\bfx^{'}|}\ ^{(1)} T^{i0} (t,\bfx^{'}), \\ \chi & = & - \frac{G}{2} \int |\bfx-\bfx^{'}| ^{(0)} T^{00} d^3 x^{'} (t,\bfx^{'}), \end{eqnarray} the metric components read \begin{eqnarray} ^{(2)}g_{00} & = & -2 \phi_\mathrm{N} , \label{metr1} \\ ^{(4)}g_{00} & = & -2 ( \phi_\mathrm{N}^2 +\psi + \phi_\Lambda ), \label{metr2} \\ ^{(2)}g_{ij} & = & -2 \delta_{ij} \phi_\mathrm{N} , \label{metr3} \\ ^{(3)}g_{0i} & = & \xi_i +\chi_{,i0} . \label{metr4} \end{eqnarray} For a point-like mass at the centre of the coordinate system, the above expressions reduce to the weak field limit at large radii of the Kottler space-time. \subsection{Equations of motion for a test particle} The motion of a particle in an external gravitational field can be described by the Lagrangian \begin{equation} {\cal L}= 1- \sqrt{ - g_{\mu \nu} \left( \frac{d x^{\mu}}{dt} \right) \left( \frac{d x^{\nu}}{dt} \right)}. \end{equation} Using the metric components in equations~(\ref{metr1}-\ref{metr4}), we get \begin{equation} {\cal {L}} \simeq \frac{1}{2} v^2+ \frac{1}{8} v^4 -\phi_\mathrm{N} -\frac{1}{2} \phi_\mathrm{N}^2 -\psi -\phi_\Lambda -\frac{3}{2}\phi_\mathrm{N} v^2 +v^i \left( \xi_i + \frac{\partial \chi}{\partial t \partial x^i} \right) . \end{equation} The corresponding Euler-Lagrange equations of motion in a 3-dimensional notation read, \begin{equation} \frac{d {\bf v}}{dt} \simeq -\nabla \left( \phi_\mathrm{N} +2\phi_\mathrm{N}^2+\psi \right) - \frac{\partial \bfxi}{\partial t} - \frac{\partial^2 }{\partial t^2} \nabla \chi + \bfv {\times} (\nabla {\times} \bfxi) + 3 \bfv \frac{\partial \phi_\mathrm{N}}{\partial t} +4 \bfv (\bfv {\cdot} \nabla)\phi_\mathrm{N}-\bfv^ 2\nabla \phi_\mathrm{N} +\frac{\Lambda}{3} \bfx . \end{equation} The above expression reduces to equation (20) in \cite{ker+al03} when neglecting pN corrections. \section{The Einstein-Infeld-Hoffmann equations} \label{sec:EIHeq} Since the contribution from the cosmological constant is of higher-order, it does not couple with other corrections. The Lagrangian of an $N$-body system of point-like particles can be written as \begin{equation} {\cal L} \simeq {\cal L}_{(\Lambda =0)} + \delta{\cal L}_\Lambda, \end{equation} where ${\cal L}_{\Lambda =0}$ is the total Lagrangian in absence of $\Lambda$. The Lagrangian ${\cal L}_a $ of particle $a$ in the field of other particles is \begin{equation} {\cal L}_a \simeq {\cal L}_{a(\Lambda =0)} +\frac{\Lambda}{6} \bfx_a^2, \end{equation} where ${\cal L}_{a(\Lambda =0)}$ is given in equation (5.94) in \cite{str04} The total Lagrangian reads \begin{equation} {\cal L} \simeq {\cal L}_{(\Lambda =0)} + \sum_a \frac{\Lambda}{6} m_a \bfx_a^2, \end{equation} with ${\cal L}_{(\Lambda =0)}$ given in \cite[equation~(5.95)]{str04}. The corresponding Euler-Lagrange equations are the Einstein-Infeld-Hoffmann equations corrected for a $\Lambda$ term, \begin{equation} \dot{\bfv}_a = - G \sum_{b \neq a} \left( \frac{\bfx_{ab}}{r_{ab}} \right)+ \delta \mathbf{F}_{\mathrm{pN} (\Lambda=0)} + \frac{\Lambda}{3} \bfx_a \end{equation} where $\mathbf{F}_{\mathrm{pN} (\Lambda=0)}$ is the post-Newtonian perturbing function \cite[equation~(5.96)]{str04}. \section{The two-body problem} \label{sec:twobo} The total Lagrangian for two particles can be written as \begin{equation} {\cal L} \simeq \frac{1}{2}m_\mathrm{a} v_\mathrm{a}^2 +G \frac{m_\mathrm{a} m_\mathrm{b}}{x}+ \frac{1}{2}m_\mathrm{b} v_\mathrm{b}^2 + \delta {\cal L}_{\mathrm{pN} (\Lambda=0)} + \delta {\cal L}_{\Lambda} \end{equation} where $\bfx \equiv \bfx_\mathrm{a} - \bfx_\mathrm{b} $ is the separation vector and $\delta {\cal L}_{\mathrm{pN} (\Lambda=0)}$ and $\delta {\cal L}_{\Lambda}$ are the pN and $\Lambda$-contributions, respectively. It is \cite{str04} \begin{equation} \delta {\cal L}_{\mathrm{pN} (\Lambda=0)} = \frac{1}{8} (m_\mathrm{a} v_\mathrm{a}^4 + m_\mathrm{b} v_\mathrm{b}^4) + G \frac{m_\mathrm{a} m_\mathrm{b}}{2 r} \left[ 3 (v_\mathrm{a}^2+v_\mathrm{b}^2) - 7 \bfv_\mathrm{a} \dot \bfv_\mathrm{b} - (\bfv_\mathrm{a} \dot {\bf n})(\bfv_\mathrm{b} \dot {\bf n}) \right] -\frac{G^2}{2}\frac{m_\mathrm{a} m_\mathrm{b} (m_\mathrm{a}+m_\mathrm{b})}{x^2} \end{equation} with ${\bf n} \equiv {\bf x}/x$ and \begin{equation} \delta {\cal L}_{\Lambda} = \frac{\Lambda}{6} (m_\mathrm{a} x_\mathrm{a}^2 + m_\mathrm{b} x_\mathrm{b}^2) . \end{equation} Due to cosmological constant, the energy of the system is modified by a contribution $-\delta {\cal L}_{\Lambda}$. The pN and $\Lambda$ corrections are additive and can be treated separately. We are interested in examining the effect of a non vanishing $\Lambda$ term. Let us consider the centre of mass and relative motions. Introducing ${\bf X} \equiv \left( m_\mathrm{a} \bfx_\mathrm{a} +m_\mathrm{b} \bfx_\mathrm{b} \right)/M $, with $M \equiv m_\mathrm{a}+m_\mathrm{b}$, the Lagrangian can be re-written as \begin{equation} {\cal L} \simeq \frac{1}{2} M V^2 + \frac{\Lambda}{6} M X^2 +\frac{1}{2} \mu v^2 + \frac{\Lambda}{6} \mu x^2 + G \frac{M \mu}{x}, \end{equation} with $\mu \equiv m_\mathrm{a} m_\mathrm{b}/M$. Due to cosmological constant, the centre of mass of the system is subject to an effective repulsive force given by $\Lambda {\bf X}/3$ per unit mass. The equations for the relative motion are those of a test particle in a Schwarzschild-de~Sitter space-time with a source mass equal to the total mass of the two-body system. Since the perturbation due to $\Lambda$ is radial, the orbital angular momentum is conserved and the orbit is planar. The main effect of $\Lambda$ on the orbital motion is a precession of the pericentre \cite[and references therein]{kr+wh03,ker+al03}. Following the analysis of the Rung-Lenz vector in \cite{ker+al03} and restoring the $c$ factors, we get for the contribution to the precession angular velocity due to $\Lambda$, \begin{equation} \dot{\omega}_\Lambda = \frac{\Lambda c^2 P_\mathrm{b}}{4 \pi} \sqrt{1-e^2}, \end{equation} where $e$ is the eccentricity and $P_\mathrm{b}$ the Keplerian period of the unperturbed orbit. This contribution should be considered together with the post-Newtonian periastron advance, $\dot{\omega}_{\mathrm{pN}} = 3 (2 \pi/P_\mathrm{b})^{5/3} (G M/c^3)^{2/3} (1-e^2)^{-1} $. The ratio between these two contributions can be written as, \begin{equation} \frac{ \dot{\omega}_\Lambda } {\dot{\omega}_{\mathrm{pN}}} = \frac{\bar{R}}{R_\mathrm{g}} \frac{\rho_\Lambda }{ \rho } = \frac{1}{6}\frac{\bar{R}^4}{R_\mathrm{g}^2} \Lambda , \end{equation} where $\bar{R} = a (1-e^2)^{3/8}$ is a typical orbital radius with $a$ the semi-major radius of the unperturbed orbit, $\rho \equiv M/(4 \pi \bar{R}^3/3)$ is a typical density of the system and $\rho_\Lambda \equiv c^2 \Lambda /8 \pi G $ is the energy density associated to the cosmological constant. The effect of $\Lambda$ can be significant for very wide systems with a very small mass. \section{Observational constraints} \label{sec:obser} In this section, we derive observational limits on $\Lambda$ from orbital precession shifts in stellar systems and in the solar system. \subsection{Interplanetary measures} \label{sec:obserA} Precessions of the perihelia of the solar system planets have provided the most sensitive local tests for a cosmological constant so far \cite{isl83,wri98,ker+al03}. Estimates of the anomalous perihelion advance were recently determined for Mercury, Earth and Mars \cite{pit05a,pit05b}. Such ephemerides were constructed integrating the equation of motion for all planets, the Sun, the Moon and largest asteroids and including rotations of the Earth and of the Moon, perturbations from the solar quadrupole mass moment and asteroid ring in the ecliptic plane. Extra-corrections to the known general relativistic predictions can be interpreted in terms of a cosmological constant effect. We considered the 1-$\sigma$ upper bounds. Results are listed in Table~\ref{tab:plan}. Best constraints come from Earth and Mars observations, with $\Lambda \ls 10^{-36}\mathrm{km}^{-2}$. Major sources of systematic errors come from uncertainties about solar oblateness and from the gravito-magnetic contribution to secular advance of perihelion but their effect could be in principle accounted for \citep{ior05}. In particular, the general relativistic Lense-Thirring secular precession of perihelia is compatible with the determined extra-precessions \citep{ior05}. The accuracy in determining the planetary orbital motions will further improve with data from space-missions like BepiColombo, Messenger and Venus express. By considering a post-Newtonian dynamics inclusive of gravito-magnetic terms, the resulting residual extra-precessions should be reduced by several orders of magnitude, greatly improving the upper bound on $\Lambda$. The orbital motion of laser-ranged satellites around the Earth has been also considered to confirm general relativistic predictions. Observations of the rates of change of the nodal longitude of the LAGEOS satellites allowed to probe the Lense-Thirring effect with an accuracy of $\sim10\%$, i.e. about half a milliarcsecond per year \cite{ci+pa04}. Other proposed missions, such as the LARES/WEBER-SAT satellite \cite{ciu04}, should further increase this experimental precision. In general, since effects of $\Lambda$ become significant only for very dilute systems, even very accurate measurements of orbital elements of Earth's satellites can not help in constraining the cosmological constant. For a satellite with a typical orbital semi-major axis of about 12,000~km, in order to get a bound on $\Lambda$ as accurate as those inferred from Earth and Mars perihelion shifts (i.e. $\Lambda \ls 10^{-36}~\mathrm{km}^{-2}$), changes in orbital elements should be measured with a today unattainable precision of a few tens of picoseconds of arc per year, about six order of magnitude better than today accuracy. \begin{table} \caption{\label{tab:plan} Limits on the cosmological constant due to extra-precession of the inner planets of the solar system.} \begin{ruledtabular} \begin{tabular}{lrrr} Name & $\delta\dot{\omega}\footnotemark[1]$ (arcsec/year)& $\dot{\omega}_\Lambda$ ($\deg$/year)& $\Lambda_\mathrm{lim}~(\mathrm{km}^{-2}) $ \\ \hline Mercury & $-0.36(50)\times 10^{-4}$ & $9.61{\times} 10^{25} \Lambda /(1~\mathrm{km}^{-2})$ & $4 {\times} 10^{-35}$ \\ Venus & $0.53(30)\times 10^{-2}$ & $2.51{\times} 10^{26}\Lambda /(1~\mathrm{km}^{-2})$ & $9 {\times} 10^{-33}$ \\ Earth & $-0.2(4)\times 10^{-5}$ & $4.08{\times} 10^{26}\Lambda /(1~\mathrm{km}^{-2})$ & $1{\times} 10^{-36}$ \\ Mars & $0.1(5)\times 10^{-5}$ & $7.64{\times} 10^{26} \Lambda /(1~\mathrm{km}^{-2})$& $ 2{\times} 10^{-36}$ \\ \end{tabular} \end{ruledtabular} \footnotetext[1]{From \cite{pit05a}.} \end{table} \subsection{Binary pulsars} \begin{table} \caption{\label{tab:1} Binary pulsars with known post-Keplerian parameter $\dot{\omega}$ and corresponding limits on the cosmological constant. The identification of the companion is often uncertain. We refer to the original papers for a complete discussion.} \begin{ruledtabular} \begin{tabular}{lllllcc} PSR Name & $P_\mathrm{b}$ (days) & $e$ & $\dot{\omega}$ ($\deg$/year)& $\dot{\omega}_\Lambda$ ($\deg$/year)& $ \Lambda_\mathrm{lim}~(\mathrm{km}^{-2})$ & ref.\\ \hline \multicolumn{7}{c}{Double neutron star binaries} \\ \hline J1518+4904 & 8.634000485 & 0.2494849 & 0.0111(2) & $9.335 {\times} 10^{24}\Lambda /(1~\mathrm{km}^{-2})$ & $2 {\times} 10^{-29}$ & \cite{nic+al96} \\ B1534+12 & 0.2736767 & 0.420737299153 & 1.755805(3) & $2.772 {\times} 10^{23}\Lambda /(1~\mathrm{km}^{-2})$ & $1 {\times} 10^{-29}$ & \cite{wil05} \\ B1913+16 & 0.323 & 0.617 & 4.226595(5) & $2.838{\times} 10^{23} \Lambda /(1~\mathrm{km}^{-2})$ & $2 {\times} 10^{-29}$ & \cite{wil05} \\ J1756-2251 & 0.319633898 & 0.180567 & 2.585(2) & $3.510 {\times} 10^{23} \Lambda /(1~\mathrm{km}^{-2})$ & $6 {\times} 10^{-27}$ & \cite{fau+al05} \\ J1811-1736 & 18.779168 & 0.82802 & 0.009(2) & $1.176{\times} 10^{25} \Lambda /(1~\mathrm{km}^{-2})$ & $2 {\times} 10^{-28}$ & \cite{lyn+al00} \\ J1829+2456 & 1.176028 & 0.13914 & 0.28(1) & $1.300 {\times} 10^{24} \Lambda /(1~\mathrm{km}^{-2})$ & $ 8 {\times} 10^{-27}$ & \cite{cha+al04} \\ B2127+11C & 0.68141 & 0.335282052 & 4.457(12) & $7.168{\times} 10^{23}\Lambda /(1~\mathrm{km}^{-2})$ & $2{\times} 10^{-26}$ & \cite{wil05} \\ B2303+46 & 12.34 & 0.65837 & 0.01019(13) & $1.037 {\times} 10^{25} \Lambda /(1~\mathrm{km}^{-2}) $ & $1 {\times} 10^{-29}$ & \cite{th+ch99} \\ \hline \multicolumn{7}{c}{Neutron star/white dwarf binaries} \\ \hline J0621+1002 & 8.3186813 & 0.00245744 & 0.0116(8) & $9.288 {\times} 10^{24} \Lambda /(1~\mathrm{km}^{-2})$ & $9 {\times} 10^{-29}$ & \cite{spl+al02} \\ J1141-6545 & 0.171876 & 0.1976509587 & 5.3084(9) & $1.881 {\times} 10^{23}\Lambda /(1~\mathrm{km}^{-2}) $ & $5 {\times} 10^{-27}$ & \cite{wil05} \\ J1713+0747 & 67.82512987 & 0.0000749406 & 0.0006(4)\footnotemark[1] & $7.573{\times} 10^{25} \Lambda /(1~\mathrm{km}^{-2}) $ & $8 {\times} 10^{-30}$ & \cite{spl+al05} \\ B1802-07 & 2.617 & 0.212 & 0.0578(16) & $2.856{\times} 10^{24}\Lambda /(1~\mathrm{km}^{-2}) $ & $ 6 {\times} 10^{-28}$ & \cite{th+ch99} \\ J1906+0746 & 0.085303(2) & 0.165993045(8) & 7.57(3) & $9.392{\times}10^{22} \Lambda /(1~\mathrm{km}^{-2})$ & $ 3 {\times} 10^{-25}$ & \cite{lor+al05} \\ \hline \multicolumn{7}{c}{Double pulsars} \\ \hline J0737-3039 & 0.087779 & 0.102251563 & 16.90(1) & $9.750{\times} 10^{22} \Lambda /(1~\mathrm{km}^{-2}) $ & $1 {\times} 10^{-25}$ & \cite{wil05} \\ \hline \multicolumn{7}{c}{Unknown companion} \\ \hline B1820-11 & 357.7622(3) & 0.79462(1) & 0.01\footnotemark[1] & $2.425 {\times} 10^{26} \Lambda /(1~\mathrm{km}^{-2}) $ & $4 {\times} 10^{-29}$ & \cite{ly+mc89} \\ \end{tabular} \end{ruledtabular} \footnotetext[1]{Upper limit} \end{table} Binary pulsars have been providing unique possibilities of probing gravitational theories. Relativistic corrections to the binary equations of motion can be parameterized in terms of post-Keplerian parameters \cite{wil93}. As seen before, the advance of periastron of the orbit, $\dot{\omega}$, depends on the total mass of the system and on the cosmological constant. In principle, because Keplerian orbital parameters such as the eccentricity $e$ and the orbital period $P_\mathrm{b}$ can be separately measured, the measurement of $\dot{\omega}$ together with any two other post-Keplerian parameters would provide three constraints on the two unknown masses and on the cosmological constant. As a matter of fact for real systems, the effect of $\Lambda$ is much smaller than $\dot{\omega}_{\mathrm{pN}}$, so that only upper bounds on the cosmological constant can be obtained by considering the uncertainty on the observed periastron shift. We considered binary systems with measured periastron shift, see Table~\ref{tab:1}. The effect of $\Lambda$ is maximum for B1820-11 and PSR J1713+0747. Despite of the low accuracy in the measurement of $\dot{\omega}$, PSR J1713+0747 provides the best constraint on the cosmological constant, $\Lambda \ls 8 {\times}10^{-30}\mathrm{km}^{-2}$. Uncertainties as low as $\delta\dot\omega \gs 10^{-6}$ have been achieved for very well observed systems, such as B1913+16 and B1534+12. Such an accuracy for B1820-11 would allow to push the bound on $\Lambda$ down to $10^{-33}\mathrm{km}^{-2}$. Better constraints could be obtained by determining post-Keplerian parameters in very wide binary pulsars. We examined systems with known period and eccentricity as reported in \cite{lor05}. The binary pulsar having the most favourable orbital properties for better constraining $\Lambda$ is the low eccentricity B0820+02, located in the Galactic disk, with $\dot{\omega}_\Lambda \sim 1.4{\times}10^{27}\Lambda / (1~\mathrm{km}^{-2})\deg/\mathrm{days}$. For binary pulsars J0407+1607, B1259-63, J1638-4715 and J2016+1948, the advance of periastron due to the cosmological constant is between 7 and $9{\times}10^{26}\Lambda/ (1~\mathrm{km}^{-2}) \deg/\mathrm{days}$. All of these shifts are of similar value or better than the Mars one. A determination of $\dot{\omega}$ for B0820+02 with the accuracy obtained for B1913+16, i.e. $\delta \dot{\omega}\gs~10^{-6} \deg/\mathrm{days}$ would allow to push the upper bound down to $ 10^{-34}-10^{-33}\mathrm{km}^{-2}$. \section{Conclusions} \label{sec:concl} We considered the $N$-body equations of motion in presence of a cosmological constant. The impact of $\Lambda$ on the two-body system was explicitly derived. Due to the anti-gravity effect of the cosmological constant, the barycentre of the system drifts away. The relative motion is like that of a test particle in a Schwarzschild-de Sitter space-time with a source mass equal to the total mass of the two-body system. The main effect of $\Lambda$ is the precession of the pericentre on the orbital motion. We determined observational limits on the cosmological constant from measured periastron shifts. With respect to previous similar analyses performed in the past on solar system planets, our estimate was based on a recent determination of the planetary ephemerides properly accounting for the quadrupole moment of the Sun and for major asteroids. The best constraint comes from Mars and Earth, $\Lambda \ls 1-2\times 10^{-36}\mathrm{km}^{-2}$. Due to the experimental accuracy, observational limits on $\Lambda$ from binary pulsars are still not competitive with results from interplanetary measurements in the solar system. Accurate pericentre advance measurements in wide systems with orbital periods $\gs 600~\mathrm{days}$ could give an upper bound of $\Lambda \ls 10^{-34}-10^{-33}\mathrm{km}^{-2}$, if determined with the accuracy performed for B1913+16, i.e. $\delta\dot\omega \gs 10^{-6}\deg/\mathrm{years}$. For some binary pulsars, observations with an accuracy comparable to that achieved in the solar system could allow to get an upper limit on $\Lambda$ as precise as one obtains from Mars data. The bound on $\Lambda$ from Earth or Mars perihelion shift is nearly $\sim 10^{10}$ times weaker than the determination from observational cosmology, $\Lambda \sim 10^{-46}\mathrm{km}^{-2}$, but it still gets some relevance. The cosmological constant might be the non perturbative trace of some quantum gravity aspect in the low energy limit \cite{pad05}. $\Lambda$ is usually related to the vacuum energy density, whose properties depends on the scale at which it is probed \cite{pad05}. So that, in our opinion, it is still interesting to probe $\Lambda$ on a scale of order of astronomical unit. Measurements of periastron shift should be much better in the next years. New data from space-missions should get a very high accuracy and might probe spin effects on the orbital motion \cite{oco04,ior05}. A proper consideration of the gravito-magnetic effect in these analyses plays a central role to improve the limit on $\Lambda$ by several orders of magnitude. {\it Note added}. After submission of this work, L.~Iorio \cite{ior05b} presented an analysis of solar system data similar to our results in section~\ref{sec:obserA}. \begin{acknowledgments} The authors thank N. Straumann for stimulating discussions. M.S. is supported by the Swiss National Science Foundation and by the Tomalla Foundation. \end{acknowledgments}
Title: The Vertical Structure of the Outer Milky Way HI Disk
Abstract: We examine the outer Galactic HI disk for deviations from the b=0 plane by constructing maps of disk surface density, mean height, and thickness. We find that the Galactic warp is well described by a vertical offset plus two Fourier modes of frequency 1 and 2, all of which grow with Galactocentric radius. Adding the m=2 mode accounts for the large asymmetry between the northern and southern warps. We use a Morlet wavelet transform to investigate the spatial and frequency localization of higher frequency modes; these modes are often referred to as "scalloping." We find that the m=10 and 15 scalloping modes are well above the noise, but localized; this suggests that the scalloping does not pervade the whole disk, but only local regions.
https://export.arxiv.org/pdf/astro-ph/0601697
\title{The Vertical Structure of the Outer Milky Way HI Disk} \author{E.S. Levine\altaffilmark{1}, Leo Blitz\altaffilmark{1}, and Carl Heiles\altaffilmark{1} } \affil{Department of Astronomy, University of California at Berkeley, Mail Code 3411, Berkeley, CA 94720 USA} \email{elevine@astron.berkeley.edu} \keywords{Galaxy: disk --- Galaxy: structure --- Galaxy: kinematics and dynamics --- ISM: structure --- radio lines: general} \section{Introduction} Although the topography of the gas disk of the Milky Way has previously been mapped \citep[and many others]{W1957, HJK1982,BT1986}, its complete Fourier structure has never been quantitatively described, though \citet{BM1998} have approximated its three lowest frequency terms. Which spatial oscillation frequencies are most important? Is the scalloping a local or global effect? This paper constitutes an in depth analysis of the shape of the outer HI disk, as well as the first quantitative analysis of the scalloping. The large-scale warp in the gas disk of the Milky Way has been known since 1957 \citep{B1957,K1957,W1957,KHC1957}. The warp has a large amplitude, rising to a height greater than 4 kpc at a Galactocentric radius of 25 kpc in the northern data. It is also asymmetric; in the south the gas falls about 1 kpc below the plane before rising back to it. By observing other galaxies, it is possible to develop a general understanding of how warps behave. \citet{B1991} found that at minimum half of all galaxies are warped, and that galaxies with smaller dark matter halo core radii are less likely to be warped. \citet{B1990} claimed that a warp's line of nodes starts out straight, and at a transition begins to advance in the direction of rotation, with some exceptions. Another survey has shown that warps are common in galaxies with HI disks that are extended compared to their optical components, and are often asymmetric \citep{GSK2002}. In the Milky Way, many components other than HI also participate in the warp. A partial list includes dust \citep{FBD1994}, CO \citep{WBB1990}, solar neighborhood stars \citep{D1998}, and IRAS point sources \citep{DS1989}. Many efforts have been directed toward understanding the warp on a theoretical basis. Bending modes have long been suspected as the mechanism creating and maintaining the warp. Early work studying the evolution of bending mode oscillations was hampered by a lack of knowledge regarding galactic halos, but showed that the shape of the density falloff near the edge of the disk plays an important role in the stability of bending modes \citep{HT1969,T1983}. The distribution of matter in the halo controls the ability of the disk to sustain long-lived bending wave warps, so studying the properties of the warp will tell us about the shape of the halo \citep{B1978,S1984}. \citet{SC1988} argued that bending mode oscillations are plausible when the halo and the self-gravity of the disk are taken into account, but \citet{BJD1998} showed that in this situation the warp will wind up within a few dynamical times. Previous work has largely been concerned with the existence and behavior of $m=1$ warps only, though \citet{S1995} demonstrated the stability of $m=0$ modes in an axisymmetric halo. Several other mechanisms have been suggested as possibilities for creating and maintaining a warp \citep{KG2000}. Gravitational interaction with satellites such as the Magellanic Clouds is a promising candidate \citep{B1957,W1998,WB2005}, but there is a longstanding debate of whether tidal effects are strong enough to produce the observed effect \citep{K1957}. Indeed, even in galaxies with companions similar to the LMC, tidal amplification may not be strong enough to account for the size of the warp \citep{GKD2002}. Accretion of matter onto the halo is another plausible cause \citep{JB1999}, or matter can accrete directly onto onto the disk and torque the gas orbits \citep{LBB2002,S2004}. Intergalactic magnetic fields can act on a slightly ionized gas disk to produce a warp \citep{BFS1990}, or the intergalactic medium can excite a warp through a wind \citep{KW1959}. Throughout this paper, we will refer to shorter wavelength (5--25 kpc) oscillations in azimuth as scalloping. \citet{GKW1960} first noticed a ``waviness'' in the gas layer of the inner Galaxy; observational evidence for scalloping turns out to be prominent in the outer Galaxy \citep{HJK1982,KBH1982}. \citet{F1983} investigated the possibility that the Milky Way scalloping results from the Kelvin-Helmholz instability, and \citet{S1995} suggested that the scalloping should only be present in the outer parts of Galactic disks. \citet{SF1986} find azimuthal corrugations in HI and other components along the spiral arms after removing the warp. In an N-Body simulation, \citet{EE1997} found evidence for spiral corrugations due to gravitational interaction with a satellite galaxy. In this paper we will only investigate oscillation in Galactocentric azimuth, and not in radius. In \S 2 we describe the method of transforming the Heliocentric data cube into Galactocentric coordinates, and present maps of the surface density, average height off the $b=0\degr$ plane, and vertical thickness of the outer Galaxy. While these maps are constructed using new data, they are not significantly different from previous work. In \S 3 we perform global and local analyses on these maps to better understand the warp and the scalloping. \section{Method}\label{sec:proc} \subsection{Data Processing} We use the 21 cm Leiden/Argentine/Bonn (LAB) data \citep{LAB,HB1997,BALMPK2005,ABLMP2000} to conduct a quantitative study of the warp and scalloping. The LAB survey is a combination of the LDS data set \citep{HB1997} with Southern sky observations from the IAR \citep{ABLMP2000}; however, much attention has been paid toward ensuring a uniform data set. The data are corrected for stray radiation. The combined survey maps the entire sky within $-450\le v_r\le 400$ km s$^{-1}$ with a resolution of 1.3 km s$^{-1}$; this velocity range includes all of the gas in the Galaxy in circular rotation. We used the Hanning smoothed data, which have a velocity resolution of 1.9 km s$^{-1}$, and used only data with $|b|\le30\degr$. The signal of the warp in the north can be weakly traced beyond this elevation limit \citep{B1985}, but the vast majority of the warp signal is included within our $b$ range. The LAB survey contains a large number of angularly small emission features, especially at high latitudes, which are not contiguous with the gas in the disk, and may not even be in circular rotation. Features like these are particularly troublesome at larger Galactic latitudes since they can contain enough gas to contaminate our calculations, especially at large Galactocentric radii. These objects are removed with a median filter so we can focus on the gas in the disk. Points with brightness temperature $T_b\le 0$ are temporarily filled with the value $0.01$ K for this filtering only. We then cycle through the LAB data cube and calculate the median of each point and its 12 nearest neighbors in $\ell$ and $b$ at the same line-of-sight velocity $v_r$ (a two dimensional diamond shaped filter); call this median $T_m(\ell,b,v_r)$. Any point with $T_b>10~T_m$ is replaced with 10 $T_m$. Two examples of objects caught by this filter are M31 and NGC 6822. The survey grid in $\ell,b,$ and $v_r$ is not convenient for analyzing Galactic properties. Ideally, our measurements would be equally spaced in the Galactocentric cylindrical coordinates $R,\phi,$ and $z$; we must interpolate a Galactocentric grid from the LSR-centered data. The Galactic azimuth $\phi$ is defined such that it converges with galactic longitude $\ell$ at large $R$. To convert from observed $\ell,b,$ and $v_r$ to $R,\phi$, and $z$ we use the following mapping functions: \begin{eqnarray}\label{eqn:transform} \ell&=&\sin^{-1}\left[\frac{R}{r'}\sin\phi\right]\nonumber\\ b&=&\tan^{-1}\frac{z}{r'}\nonumber\\ v_r&=&\sin\ell\cos b\left[\frac{R_0}{R}\Theta(R)-\Theta_0\right]\nonumber\\ &&+v_\Pi(R)\cos\phi\left(1-\frac{R_0^2}{R^2}\sin^2\ell\right)^{1/2}\cos b. \end{eqnarray} Here, $\Theta(R)$ is the Galactic rotation curve, which we assume to be 220 km s$^{-1}$ everywhere \citep{BB1993}. $\Theta_0$ and $R_0$ are 220 km s$^{-1}$ and 8.5 kpc, respectively. $\bf{r}$ is the vector connecting the Sun's location to the point under investigation; $\bf{r'}$ is the projection of this vector onto the plane of the disk (with magnitude $r'$). Equation \ref{eqn:transform} is the transformation that results from an assumption of elliptical gas orbits with major axis along the $\phi=90\degr, 270\degr$ line. To derive this, we assume the gas moves on the orbit: \begin{eqnarray}\label{eqn:ellipse} v_\phi&=&220~\mathrm{km~s^{-1}}\nonumber\\ v_R&=&v_\Pi(R)\cos\phi. \end{eqnarray} $v_R$ is the magnitude of the velocity in the Galactocentric radial direction. $v_\Pi$ is a parameterization of the ellipticity of the orbit, which is free to vary with Galactocentric radius; we discuss our method of calculating $v_\Pi(R)$ in the Appendix. At all points, $v_R/v_\phi<0.1$. These equations are simply the epicyclic approximation for an orbit with epicyclic frequency 1 and the angle of ellipse orientation fixed. Although the gas orbits in the Galaxy are not likely to correspond to the fixed ellipse orientation we describe, this configuration minimizes the correction to $v_r$ for gas far from $\ell=0\degr$ or $180\degr$ (see \citet{V1999} for more detail on elliptical gas orbits). Without a correction for radial motion of the gas, there is a large asymmetry between the surface densities at Galactic longitudes on either side of $\ell =0\degr$ and $\ell =180\degr$ \citep{HJK1982}. This must be taken into account, or features in these two regions such as the surface density will appear discontinuous and distorted. Assuming an outward velocity for the Local Standard of Rest (LSR) will correct this discontinuity to some degree \citep{K1962,KW1965}. However, it seems that the best fit for the motion relative to the LSR changes with radius, implying that the effect is global, rather than local \citep{BS1991}. To reduce the magnitude of the discontinuity using gas orbits, one can use a Galactocentric radial velocity roughly of the form $\cos\phi$ (or $\cos\ell$) \citep{KT1994}. We exclude all points that lie within $345\degr\le\ell\le15\degr$ or $165\degr\le\ell\le195\degr$. Points in these two wedges have velocities along the line of sight that are too small with respect to their random velocities to establish reliable distances. All points in this region are set to $T_b = 0$. Using (\ref{eqn:transform}) we construct a Galactocentric grid $T_b(R,\phi,z)$ by trilinear-interpolating from the grid $T_b(\ell,b,v_r)$. We do this by calculating the coordinates of an $(R,\phi,z)$ point in $(\ell,b,v_r)$ space, and interpolating from $T_b(\ell,b,v_r)$. The resolution of the Galactocentric grid is set by the spacing of the LSR centered grid, but in this paper we are not interested in small-scale disk structure. A grid of 100 points in 10 kpc $\le R \le$ 30 kpc, 350 points in $-\pi\le\phi\le \pi$, and 141 points in -20 kpc$\le z \le$ 20 kpc gives us sufficient resolution to answer the questions we are interested in. Undersampled grids fail to utilize all of the information in the data; our choice of grid spacing is both is an undersampling and an oversampling of the information in $T_b(\ell,b,v_r)$ depending on the position in the disk. Consider two points near $R=10$ kpc and $\ell = 15\degr$, where we have poor Galactocentric resolution in $\phi$. The LAB survey has $\Delta \ell = 0.5\degr$; at this location this corresponds to $\Delta\phi\approx0.9\degr$. The spacing in our Galactocentric grid is larger than $1\degr$, thus we are undersampled everywhere in the $\phi$ dimension. In \S \ref{sec:lomb} we will perform an azimuthal frequency analysis of each ring; none of the frequencies we examine approach the Nyquist frequency of the data. Near the midplane at $R=30$ kpc and $\ell =15\degr$, lines of constant $b$ are separated by around 300 pc, and near the top of the grid they are separated by around 400 pc; there is some oversampling in the $z$ dimension by no more than a factor of 2. The most severe case of oversampling is in the $R$ dimension, where along $\ell = 15\degr$ we have only 14 velocity resolution elements in our $R$ range for the 100 grid points. On the other hand, at $\ell=90\degr$, there are more than 60 resolution elements in our range. We have chosen a grid spacing in $R$ that oversamples to a varying degree depending on location in the disk. From $T_b(R,\phi,z)$, we can recover $\rho(R,\phi,z)$ using the method outlined in \citet{K1968}. We assume $T_s$, the spin temperature, is 155 K everywhere. This is slightly higher than the maximum brightness temperature found in the LAB survey in the region we are concerned with. Choosing a larger number to force the optically thin limit ($T_b\ll T_s$) makes a difference only in a small number of areas in the inner radii of our grid. The vast majority of the points are optically thin with any reasonable $T_s$, and are not affected by this choice. The transformation to a density grid depends on $|dv_r/dr|$. Using elliptical orbits makes calculating this quantity slightly more difficult than with a flat rotation curve. Although the full derivative can be written analytically, we just calculate it numerically. Points with $T_b<0$ are set to $\rho =0$. \subsection{Surface Density and Mean Height maps} The grid $\rho(R,\phi,z)$ contains information about the density of HI in the Galaxy, minus whatever has been removed by the median filter and the excluded regions. Previous studies have proceeded by calculating a mean height $\bar{z}(R,\phi)$ for the gas. However, the Galaxy is a complicated place that contains a variety of HI structures in addition to the disk. In particular, there are many extended clouds located near the disk as well as spurs that split off from the disk. None of these will have been removed by the median filter, which acts only on comparatively small areas of the sky. \citet{V1999} developed an alternative method to remove some of these features; he masked out a map of high velocity cloud complexes. Since we are only interested in the shape of the disk itself, these additional components must be filtered out before our calculation of the mean height. We perform a dispersion filter that operates as follows. \begin{enumerate} \item Calculate the total mass surface density\begin{equation}M(R,\phi)=\sum_{i=1}^{N=141} \rho(R,\phi,z_i)\Delta z\end{equation} where $\Delta z$ is the $z$ bin size. \item Calculate \begin{equation}\bar{z}(R,\phi)=\frac{\sum_{i=1}^{N=141} z_i \rho(R,\phi,z_i) \Delta z}{M(R,\phi)}.\end{equation} \item Calculate the second moment\begin{equation}d^2(R,\phi)=\frac{\sum_{i=1}^{N=141}[z_i-\bar{z}(R,\phi)]^2 \rho(R,\phi,z_i) \Delta z}{M(R,\phi)}.\end{equation} \item Run through each point in $(R,\phi,z)$ space. For any point that does not lie within $2d$ of $\bar{z}$, set $\rho = 0$. Call the grid that results from this dispersion filter $\rho_d(R,\phi,z)$. \end{enumerate} For step 4, we experimented with several different cutoffs (in the range $1-3d$); our results do not depend strongly on which cutoff we choose. We construct the Galactic disk surface density and mean height maps from $\rho_d(R,\phi,z)$. For example, we can sum the density over the $z$ dimension to construct the surface density associated with the disk (Figure \ref{fig:sigma}): \begin{equation} \Sigma(R,\phi)= \sum_{i=b}^t \rho_d(R,\phi,z_i) \Delta z \end{equation} The indices $t$ and $b$ represent the top and bottom $z_i$ that have not been zeroed out by the dispersion filter. The resulting figure clearly demonstrates the falloff of the disk surface density with radius. The contour lines are nicely continuous across the $\ell=0\degr$ and $\ell=180\degr$ lines because of our use of elliptical orbits; see the Appendix for a version of this figure without correcting for these orbits. The jagged nature of the contours with $R\la 18$ kpc is due in part to spiral arms \citep{LBH2006}. There is a region near $R\approx 27$ kpc and $\phi\approx90\degr$ with a smaller surface density than other regions at the same radius; this region has somewhat unusual features in all of our maps. Also notice the excluded regions near the Sun-Galactic center line; these gaps will appear in all of our plots. We will be looking closely at the mass weighted mean height of the gas disk (Figure \ref{fig:height}), \begin{equation} h(R,\phi)=\frac{1}{\Sigma(R,\phi)}\sum_{i=b}^{t} z_i \rho_d(R,\phi,z_i) \Delta z. \end{equation} This height is calculated with respect to the Galactic midplane defined by $b=0\degr$. The Galactic warp is the most immediately evident feature in this map; the gas in the northern hemisphere peaks at $h\approx5$ kpc, while the southern gas descends only to $h\approx-1.5$ kpc, consistent with previous maps of the Galaxy. At least three vertical oscillations of magnitude $\approx 1$ kpc can be seen in the south from $\phi\approx-120\degr$ to $\phi\approx-20\degr$; these have previously been called the ``scalloping''. The approximate extent of the scalloping is marked with an arc connecting two ``S'' labels. The region of low surface density noted in the discussion of Figure \ref{fig:sigma} ($R=27$ kpc, $\phi=90\degr$) has a height that seems anomalous when compared to surrounding gas (it is marked with an ``X''). Several features are elongated along lines of constant $\ell$ indicating some level of contamination by turbulent velocities and/or local gas. In the past these have been dubbed ``fingers of God'' because they all point back to the Sun. In \S \ref{sec:warp} we will need a measure of the uncertainty in the mean height of each point in the map. We define this error using the sum of squared residuals, as is usual for least squares fits: \begin{equation} e^2(R,\phi)=\frac{\sum_{i=b}^{N=t}(z_i-h(R,\phi))^2 \rho_d(R,\phi,z_i) \Delta z}{(t-b+1)\Sigma(R,\phi)}, \end{equation} where $t-b+1$ is the number of points in the calculation of the mean and surface density. This is only an approximation because it does not account for any of the uncertainty introduced in $T_b(\ell,b,v_r)$, $v_\Pi(R)$, the interpolation to $\rho(R,\phi,z)$, or the dispersion filter. Typically, $e(R)/R\approx0.01$. Because of effects like turbulence, anomalous velocities, and spiral arms, the features in the mean height map may not correspond to the actual shape of the Galaxy. The severity of this effect can be seen by looking at a contour map of $\mathrm{d}r/\mathrm{d}v_r$ (see Figure \ref{fig:uncert}). While similar to previously published plots \citep{BT1986}, this figure also includes the correction for elliptical gas orbits. Given a turbulent velocity of magnitude $v_t$, features with coherent scales less than $v_t \mathrm{d}r/\mathrm{d}v_r$ could potentially be false signals. Furthermore, even real features will be blurred out or even incorrectly positioned over the same length scale. A typical turbulent velocity is around 8 km s$^{-1}$. Small regions can differ from the flat rotation curve by 20-30 km s$^{-1}$ \citep{BB1993}. We do not expect that these distortions will significantly effect our Fourier analysis. The steep contours near $\ell=0\degr$ and $\ell=180\degr$ demonstrate the need for the excluded regions, because small irregularities in velocity there result in large changes in distance. \subsection{Thickness map} We also construct a measure of the thickness of the disk. In contrast to the previous section, we keep all of the complicated features of the HI in our calculation of the thickness. We do this because these features do contain information regarding pressure and gravitational force, although they are a nuisance when calculating the mean height. In particular, our method for finding the thickness of the disk relies heavily on the tails of the vertical density distribution. Thus, in this section, we will work with $\rho$ instead of $\rho_d$. Following \citet{HJK1982}, we define the first and third quartile points as $z_{j1}$ and $z_{j3}$ as the smallest and largest indices, respectively, that satisfy: \begin{eqnarray} \sum_{i=0}^{j1}\rho(R,\phi,z_i)\Delta z&\ge&M(R,\phi)/4\nonumber\\ \sum_{i=j3}^{141}\rho(R,\phi,z_i)\Delta z&\ge&M(R,\phi)/4. \end{eqnarray} We then define the half thickness (a factor of 2 smaller than \citet{HJK1982} to ease comparison with recent work): $T_h(R,\phi)=(z_{j3}-z_{j1})/2.$ Note that this implies that $T_h$ is quantized by $\Delta z/2$ (about 140 pc). This can lead to inaccuracies in $T_h$ in places where the thickness is small, i.e. $R\approx10$ kpc. $T_h(R,\phi)$ is shown in Figure \ref{fig:disp}. Some points near the sun actually have calculated half-thicknesses of zero; this is due to the comparatively small thickness of the disk in that region combined with our limited LAB survey range $|b|\le30\degr$ and poor grid resolution in $z$. The flaring of the disk with radius is immediately evident, as $T_h$ increases in magnitude by a factor of about 8 between $R\approx10$ and 30 kpc; this flaring was first seen in \citet{LK1963}. Asymmetry between the northern and southern halves of the disk is also prominent; the southern half of the Galaxy has a lower average thickness than the northern. This asymmetry was also evident in \citet{HJK1982} and \citet{BT1986}. The low surface density region around $R\approx27$ kpc has a very large thickness; the gas in the region has clearly been disturbed and dispersed by some mechanism. There are regions of increased thickness close to $\ell=15\degr$ and $345\degr$ near $R\approx 10$ kpc probably caused by local gas. Figure \ref{fig:rfuncs} plots the surface density and half-thickness of the HI layer averaged over $\phi$ as a function of $R$. This figure confirms our visual impressions from the surface density and dispersion maps by showing the falloff in the surface density and the flaring of the thickness with radius. The contamination by local gas described in the previous paragraph as well as the poor $z$ resolution problem mentioned at the beginning of the section account for the rise in the average of $T_h$ with decreasing $R$ near $R\approx 10$ kpc; there is no evidence the thickness of the disk actually behaves this way. A gradient-expansion least-squares routine \citep{M1963} gives the best fit for the surface density beyond 14 kpc (where the exponential falloff begins) as: \begin{equation} \Sigma(R)=4.5\times\exp[-(R-14~\mathrm{kpc})/ 4.3~ \mathrm{kpc}] ~\mathrm{M}_\odot~ \mathrm{pc}^{-2}. \end{equation} \section{Analysis} \subsection{Global Structure} Using the maps we have constructed, we conduct a quantitative investigation of the disk vertical structure. Since we are studying a disk, we use a method that is independent of rotation in $\phi$ and treats $\phi =0$ and $\phi = 2\pi$ as the same point. Furthermore, the data are unevenly sampled due to the excluded regions within $15\degr$ of the Sun-Galactic center line. We complete the analysis without extrapolating $h(R,\phi)$ in these regions, to avoid introducing any artifacts into the signal. Throughout this paper, we refer to different frequency oscillations in the disk. These frequencies will always be labeled by the number of oscillations they will complete in a full $2\pi$; thus the $m=1$ mode has a one maximum and one minimum in the disk. \subsubsection{Galactic Warp}\label{sec:warp} The Galactic warp is the most prominent feature in Figure \ref{fig:height}. A Lomb periodogram analysis of each radius ring (see \S \ref{sec:lomb}) reveals that the power in each of the $0,1,$ and 2 modes is consistently larger than that in any other mode for $R\ga 20$ kpc. At some radii $m=3$ is the next strongest mode, and at others it is $m=4$. Accordingly, we characterize the warp by an offset in the $z$ direction, plus two Fourier modes with frequency 1 and 2. We fit each ring with the function: \begin{equation}\label{eqn:fit} W(\phi)=W_0 + W_1\sin(\phi -\phi_1)+W_2\sin(2\phi-\phi_2). \end{equation} Each of the three amplitudes $W_i$ and two phases $\phi_i$ in this fit is a function of radius, because we fit each radius ring independently. We use the gradient-expansion fitting algorithm to perform this fit, weighting each point by the inverse of the squared estimate of the uncertainty in the mean height, $e^2(R,\phi)$. The results of this fit for the rings at $R=16,22,$ and $28$ kpc are shown in Figures \ref{fig:warp16},\ref{fig:warp22}, and \ref{fig:warp28}. Error bars in these plots represent $e(R,\phi)$. Each of these plots is a good fit; the offset and the two Fourier components are both necessary and sufficient to describe the large-scale structure of the disk. For illustrative purposes, we will follow the ring at $R=28$ kpc through each step of the analysis. Following the evolution of the fit parameters at different places in the disk will tell us how the warp changes with radius. The three amplitude parameters each increase monotonically, with the $m=0$ mode possibly reaching an asymptotic value near the far end of our radius range (Figure \ref{fig:warpamp}). At $R\approx11$ kpc, the $m=1$ mode dominates the shape of the warp; the other two modes do not become important until $R\approx 15$ kpc. This plot implies that the $m=1$ mode has power even at the edge of our grid, thus we cannot establish the onset of the warp. We do a linear least-squares fit on the growth of each warp parameter using the function \begin{equation}\label{eqn:warpfit} W_n=k_0+k_1\left(R-R_k\right)+k_2\left(R-R_k\right)^2 \end{equation} where only points at $R_k$ and beyond are weighted in the fit. $R_k$ is arbitrarily chosen to be near where each of the three modes starts growing. The value of $k_0$ for each fit is strongly correlated to the choice of $R_k$. \begin{deluxetable}{rrrrr} \tablecaption{\label{tab:warpfit} Parameters resulting from a linear least-squares fit to the warp} \tablehead{ \colhead{$m$}&\colhead{$R_k$ (kpc)} &\colhead{$k_0$ (pc)}&\colhead{$k_1$ (pc kpc$^{-1}$)}&\colhead{$k_2$ (pc kpc$^{-2}$)}} \startdata 0&15&-66&150&-0.47\\ 1&10&9&197&-3.1\\ 2&15&-70&171&-5.3 \enddata \end{deluxetable} \citet{BM1998} (hereafter BM) discuss an approximation to the warp that is of similar form to our fit. They also fit the first three modes, but they fix the line of zeros to lie along the Sun-Galactic center line. In our fit, this would be equivalent to setting $\phi_1$ and $\phi_2$ to zero. Also, BM fix $W_0$ and $W_2$ to be the same. Note that our warp data are adjusted for elliptical gas orbits, and are filtered according to the procedure described in \S\ref{sec:proc}, whereas BM's data are not. Fig.~\ref{fig:binney} compares the mode amplitudes calculated from the data, the fit from BM, and our fit. BM overestimate the strength of the warp starting at around $R\approx22$ kpc for the $m=2$ mode, 24 kpc for the $m=0$, and 27 kpc for the $m=1$. However, the BM fit matches our data fairly well for the radii where the warp is growing most rapidly. The line of maxima for the $m=1$ mode and one of the two lines of maxima for the $m=2$ mode are marked on the contour map of the warp fit (Figure \ref{fig:phase}). Since $\phi_1$ and $\phi_2$ are actually the line of zeros, the lines of maxima are shifted $90\degr$ and $45\degr$ from these values in our fit, respectively. The parameter $\phi_2$ is not well determined at small radii in our fit because the amplitude of the $m=2$ Fourier component is very small in that region. For this reason, we do not plot $\phi_2$ in the region where $W_2$ is less than 150 pc. There is little evidence for precession in the lines of maxima for the two modes, and the line of maxima for the $m=1$ mode is roughly aligned with one of the lines of maxima of the $m=2$ mode; for each radius their difference in $\phi$ is less than $12\degr$. \citet{B1988} has examined how the two lines of zeros for the mean height change in $\phi$ as a function of radius; they appear to stay roughly aligned with the Sun-Galactic center line as $R$ changes from $R_0$ to 26 kpc. Fig.~\ref{fig:phase} shows that the line of zeros for the sum of our three warp modes falls within the excluded region, consistent with the earlier work. The three component fit shown in Figure \ref{fig:phase} does a good job of reproducing the large-scale features in the mean height map. This is not surprising, given that the $m=0,1,$ and 2 modes are the strongest in the Lomb periodogram; differences between the two plots are due to power in high frequency modes. We will now examine the differences between the mean height map and the warp fit. \subsubsection{Scalloping}\label{sec:lomb} From the fit to the warp, we can determine a function for the scalloping: $s(R,\phi)=h(R,\phi)-W(R,\phi)$. We continue to assume no information regarding the shape of the disk in the excluded regions. We can now look for higher $m$ modes of oscillation that may be present in $s(R,\phi)$. Such a signal would be present if the gas were coherently moving up and down as a function of $\phi$. The Lomb periodogram is a useful numerical method for detecting periodic signals in unevenly spaced data \citep{NumRec}. It provides the same amplitude we would determine by using a linear least squares fit, even if the signal is not sinusoidal, while allowing for a straightforward error analysis. The data are weighted equally per point, which is necessary to deal with unevenly sampled data. To ease comparison with the warp component amplitudes in Fig.~\ref{fig:warpamp}, we use an unnormalized Lomb periodogram and take the square root of the power to get the amplitude. We use a Monte Carlo algorithm to determine the noise level for the Lomb periodogram. For each ring, the null hypothesis is Gaussian white noise with the same dispersion as the data in $s(R)$. We construct $10^3$ sets of noise for each ring, run the Lomb periodogram on each set, and record the highest peak in the Lomb amplitude. We then determine the distribution of peaks in these amplitudes, and define the 95\% confidence interval as the amplitude just larger than 95\% of the noise amplitude peaks. We use the same technique to determine the 99\% confidence level. To conclude that a signal is real and not caused by noise, the signal strength must cross these confidence levels. We found nearly identical noise levels by scrambling the order of the data instead of using Gaussian white noise, indicating that these calculations are robust. The Lomb periodogram of $s(R,\phi)$ for the $R=28$ kpc ring is shown in Figure \ref{fig:lombring28}. 95\% and 99\% confidence levels are marked as thresholds in amplitude. At this radius, there is significant strength in modes 4-6, 10, and 15. These modes are markedly weaker than the warp components at the same radii, which have amplitudes of 1-2 kpc. In the case of unevenly spaced or missing data points, simple sine waves are not eigenmodes of the system. One way to see this is to take a Lomb periodogram of a pure sine wave on our $\phi$ grid with missing data; a small amount of the power will leak into other frequencies. Thus subtracting out the warp from $h(R,\phi)$ influences the amplitudes of the higher order modes from the Lomb periodogram because individual $m$ modes are not independent. We argue that the subtraction is nonetheless acceptable because the warp and the scalloping appear to be due to physically distinct phenomena; studying them is much easier once they have been separated. Removing the strongest modes will also result in a more accurate power spectrum of the weaker frequencies, since we eliminate the power leakage from the stronger to the weaker modes. Figure \ref{fig:lombfreqs} shows how the amplitudes in several different $m$ modes evolve with $R$. Modes 4-6, 10, and 15 appear to increase in strength at outer radii, while mode 3 is strongest in the intermediate radii in our map. Other modes have no detectable strength because they do not cross the 95\% confidence level at any radii. Also notice how the confidence thresholds rise with radius because of the increase in the dispersion of $s(R,\phi)$. Another consequence of using the Lomb periodogram analysis on unevenly spaced data is that it is possible to be fooled by a false signal due to ringing from interference between different $m$ modes. It is difficult to protect against this, but if two strong modes were interfering with each other to produce a third signal, we would expect at least two of the signals to increase in strength at the same radius. The strongest modes in Figure \ref{fig:lombfreqs} become significant at different radii and have visually different evolution with radius; we conclude that these modes are real. \subsection{Local Structure}\label{sec:wave} The Lomb periodogram cannot be used to study the local structure of the disk since it cannot determine where in azimuth each mode is strong. Imagine a situation where, like a falling stone creating ripples on the surface of a pond, something passes through the HI disk and excites a local series of vertical oscillations. The oscillation will add power to some frequencies in the Lomb periodogram, but this effect may be dwarfed by oscillations elsewhere in the disk. We wish to detect these sorts of perturbations and learn where in the disk they are prominent. One way to draw out this type of structure is to use a wavelet analysis. Wavelets are ideal for our purposes because they are localized in both position and frequency space; in other words, they will show where in the disk a given mode of oscillation is dominant. In wavelet theory, it is beneficial to choose a mother wavelet that is similar in shape to the signals you are searching for. Since we are looking for sinusoidal perturbations, we use the normalized Morlet wavelet with $\omega_0=6$; this is simply a sine wave times a Gaussian envelope. In the same way that a Fourier transform breaks down the frequency structure of a signal using sines and cosines, a wavelet transform breaks down a signal in terms of a set of these Morlet wavelets centered at different spatial positions and with a range of frequencies. However, in a continuous wavelet transform like the one we perform, the different wavelet functions are not orthogonal. To avoid having to interpolate or zero out the excluded regions, we will examine the northern and southern halves of the Galaxy separately. As in \S \ref{sec:lomb}, we will work with the mean height function once the warp has been subtracted, $s(R,\phi)$. Points in frequency and position space that would involve the excluded regions are defined as being inside the ``cone of influence''; points in this region are subject to edge effects, and are therefore discarded. The points affected by this are not just those in the excluded regions but also those adjacent to these regions, within some range set by the wavelength (and thus the frequency) of the particular mode \citep{TC1998}. Thus, lower frequency modes will have a larger portion of the disk fall inside the cone of influence, and have to discard a larger range of points. It is also important to have some analytic measure of which peaks in the filtered power spectrum are significant. Significance levels are discussed in detail in \citet{TC1998}. We construct a combination of parameters that have a $\chi^2$ distribution, and count as significant those that cross the 95\% confidence level threshold. Again, we use Gaussian white noise to model the randomness in the height function. The wavelet power spectrum, $\overline{W_n}^2$, for the northern half of the $R=28$ kpc ring is plotted in Figure \ref{fig:specring}; the southern half is shown in Figure \ref{fig:specring2}. The mathematical details of the wavelet transform are summarized in the Appendix. This ring has several regions with significant power. These regions have a large width in both dimensions because wavelets do not have precise resolution in either position or frequency space. Much like the uncertainty principle, the cost of using a technique that gives both position and frequency information is mediocre resolution in both. The power around $\phi\approx90\degr$ comes from the sharp dip at that azimuth; as the figure shows, this causes ringing for a large range of frequencies. The power at $-45\degr\ga\phi\ga-90\degr$ is due to the ``scalloping'' in that region. The filtered wavelet power spectra for several different bands are plotted in Figure \ref{fig:waveplot}. The band that is labeled $m=7$ is actually the filtered sum of the power spectra that satisfy $6.5\le m \le 7.5$, and so on. Note that since we calculate the power spectra for the northern and southern halves independently, the significance contours are different for the two hemispheres, in addition to being a function of $R$. In practice, this occurs because the variance of $s(\phi)$ for the northern half of the Galaxy is larger than that for the southern, and the strength required of the power spectrum to cross the 95\% confidence level is directly proportional to the variance. \section{Discussion}\label{sec:disc} We grouped the $m=0,1$ and 2 modes together in the warp because of their similar magnitudes in a Lomb periodogram. A close look at their dependence with $R$ (Fig.~\ref{fig:warpamp}) demonstrates that there are additional similarities. Both the 0 and the 2 mode are near zero until about 15 kpc from the Galactic Center. The 0 mode grows linearly from this point outwards while the 2 mode declines a small amount, and then grows approximately linearly by about 1.2 kpc. In comparison, the $m=1$ mode starts out fairly large ($\approx 300$ pc), declines slightly and then grows by about an order of magnitude. Though the $m=2$ mode does break off from this pattern at larger radii, all three components grow essentially monotonically, approximately linearly, and with similar slopes over the range $15\la R\la 22$ kpc. This may be a clue that their origins involve the same physics, and helps to justify our classification of higher frequency modes as scalloping. Studying the radial dependence of the warp inside the solar circle seems a fruitful way to learn more about these three components. The filtered wavelet power spectrum maps (Fig.~\ref{fig:waveplot}) are a representation of the scalloping in the outer Galaxy. Due to frequency-position uncertainty relations, it is impossible to establish precisely the oscillation frequency of any local disturbance. This uncertainty manifests itself in the wavelet transform by perturbations that are only somewhat localized in frequency and position space, and therefore have some width in both. These maps demonstrate that the $m\approx10$ scalloping found near $\ell\approx 310\degr$ and $R\approx 25$ kpc mentioned in previous work \citep{HJK1982,KBH1982} is real. The strong power around $m=10$ in the wavelet transform is accompanied by a large amplitude in the Lomb periodogram for $m=10$ at the corresponding radii. Wavelet transforms are also subject to the same problems that bedevil traditional Fourier transform approaches. For example, there is a large amount of power around $R\approx30$ kpc and $\phi\approx90\degr$ in $8\le m\le15$. This power is most likely not a result of scalloping in all of these different frequencies over some range of $\phi$; instead, it is probably due to ringing. As in a Fourier decomposition, sharp changes in height will cause ringing in all frequencies of a wavelet transform. Indeed, the height map (Fig.~\ref{fig:height}) does have an abrupt drop near $\phi\approx90\degr$ which could cause this ringing; we mark this region with an `X'. The same feature can be seen after the warp has been subtracted in the ring at 28 kpc plotted in Figure \ref{fig:specring}. We use the local and global analyses in conjunction to determine whether there is any scalloping mode that is present over a full $2\pi$ ring in the outer Galaxy. This is important because it will determine whether the mechanism that causes scalloping operates on a global or local scale. For this task, both techniques are necessary because even a pure $m=10$ oscillation will have some width in frequency space when put through a wavelet transform. However, this same oscillation will have a sharp peak in the Lomb periodogram. Therefore, in order to state that some mode exists over a full $2\pi$, we require both significant strength in the Lomb periodogram and significant power over a large range of $\phi$ in the wavelet transform. This combination will conclusively determine whether the scalloping is a global or local phenomenon. For modes with $m\le6$ this technique is not useful. These modes are clearly important because their Lomb periodograms show significant power for $3\le m\le 6$. Unfortunately, these modes have large enough wavelengths that the excluded regions always interfere with the wavelet transform regardless of where in the disk we look. A method of reducing the size of the excluded regions would alleviate this problem. However, this is likely to be difficult because eliminating the excluded regions would require a detailed knowledge of both $v_R$ and the distribution of gas velocities due to turbulence in the disk. Portions of these region are optically thick, which will make density determination impossible. With these constraints, the Lomb periodogram leads us to conclude that the modes $m=10$ and 15 are the most fruitful places to look for scalloping that travels a full circle around the disk. Other modes do have interesting features, but their smaller Lomb amplitudes imply that we could be fooled by the imperfect frequency resolution of the wavelet power spectrum. For example, the modes $7\le m\le9$ have significant wavelet power for large portions of the northern half of the galaxy, but none of these modes has significant signal for the corresponding radius range in the Lomb periodogram. For the $m=10$ mode the only large region with significant power in the south is the one around $\ell\approx310\degr$ that we noted previously. The power in this perturbation has fallen below the 95\% confidence level by $\ell\approx270\degr$, implying that this scalloping is a local effect. The same appears to be true for the $m=15$ oscillations, although this mode does have a region of high significance near $R\approx30$ kpc over a large range of $\phi$. However, the northern part of this signal is the region where we believe ringing to play an important role. It is therefore possible that the $m=15$ mode does have significant wavelet power over the full $2\pi$ near $R\approx 30$ kpc, but the evidence is not conclusive because the northern part of the wavelet power is most likely not due to scalloping. With the exception of the $m=15$ mode, no frequency we examine with the wavelet transform carries significant power around an entire ring along with a correspondingly significant Lomb periodogram strength. For this reason, we conclude that scalloping generally appears to be a local phenomenon. The modes $3\le m \le 6$ remain a possible exception, since we were not able to study them with a wavelet transform. It remains unclear what mechanism acts as the energy source for the scalloping. One possible cause is a massive object passing through the disk that excites local vertical oscillations in the HI gas. Azimuthally traveling wavefronts can be created by the magnetic field threaded through the disk; if the field is primarily azimuthal in nature, vertical oscillations will have larger phase and group velocities in the azimuthal direction than in the radial direction. This could lead to scalloping such as that seen towards $\ell\approx 310\degr$. \section{Conclusions} We fit the global shape of the warp on our grid of concentric rings using a vertical offset and two sinusoidal modes. Outside of $R\approx20$ kpc, each of these three modes has more power than any of the higher frequencies we look at. The amplitude increases with radius over our entire radius range for the 1 mode, and starting from around 15 kpc for the 0 and 2 modes. The growth of the 0 and 2 modes results in asymmetry in the warp; this growth begins near where the stellar disk ends. The line of maxima of the $m=1$ mode is essentially coincident with one of the lines of maxima of the $m=2$ mode. There is little evidence for precession or winding of these two modes. A global analysis with the Lomb periodogram shows that each $m$ mode evolves differently with radius. The most interesting include $m=3-6,10$, and 15, each of which start at small radii below the 95\% significance level and then cross it further out in the disk. An analysis combining the global Lomb periodogram and the local wavelet transform shows that none of the modes with $7\le m\le15$ have strength over a full ring of the disk. Using a wavelet transform, we show that the scalloping observed by previous authors near $\ell\sim 310\degr$ is real. We therefore conclude that the scalloping is a local effect. Lower frequency modes proved impossible to study with wavelets due to the presence of the excluded regions. \acknowledgments Wavelet software was provided by C.~Torrence and G.~Compo and is available at URL: http://paos.colorado.edu/research/wavelets/ \citep{TC1998}. Many thanks to Peter Kalberla for providing a copy of the LAB data set. Thanks to Eugene Chiang for discussions of dynamics. ESL and LB are supported by NSF grant AST 02-28963. CH is supported by NSF grant AST 04-06987. \appendix \section{Measuring radial motion of the gas} This appendix describes our method for finding a functional form for $v_\Pi(R)$, the magnitude of the elliptical corrections to circular rotation. A naive surface density plot using only circular rotation has a large degree of asymmetry across the lines $\ell=0\degr$ and $180\degr$ \citep{KBH1982}. Figure \ref{fig:novlsr} shows the calculated surface density without any correction for elliptical gas orbits, but with the filters we describe in \S\ref{sec:proc}. Observational experience tells us that the surface density should not have discontinuities in cardinal directions; the elliptical orbits we describe in \S \ref{sec:proc} provide the strongest corrections in the directions where the $\cos \phi$ term in (\ref{eqn:transform}) is large, and smaller corrections elsewhere. The algorithm is based on matching the surface density of the HI disk on either side of the excluded region centered at $\ell = 180\degr$. We draw confidence from the fact that this fit also does a good job matching the contours around $l=0\degr$ even though they are not included in the fit. In order to calculate a surface density, we must first know $v_\Pi(R)$ because it enters into the derivative $|dv_r/dr|$. Thus, determining $v_\Pi(R)$ from anything connected to $\rho$ is a circular problem. We must assume some form of $v_\Pi(R)$ and check to see if it results in surface densities that are well matched. We choose the functional form: \begin{equation} v_\Pi(R)=\alpha\frac{(R-R_0)}{R_0}+\beta\frac{(R-R_0)^2}{R_0^2}, \end{equation} and do not include a zeroth order term to ensure that $v_\Pi(R)$ passes through zero for the solar circle. \citet{RS1980} showed that the LSR does not have a radial velocity with respect to the Galactic center. Our algorithm follows these steps: \begin{enumerate} \item Construct $v_\Pi(R)$ for some combination of $\alpha$ and $\beta$ \item Interpolate from the LAB survey to find $T_b(R,\ell,z)$ for a grid in $R_i$ and $z_j$ over the survey range $\left|b\right| \le 30\degr$ and $155\degr \le \ell \le 165\degr$ or $195\degr\le \ell \le205\degr$ using $v_\Pi(R)$ \item Sum over $z$ to find the surface density $\Sigma(R,\ell)$ \item Average over $\ell$ for the two subsets of $\ell$ to find $\Sigma_{165}(R)$ and $\Sigma_{195}(R)$ \item Calculate a modified $\chi^2$ statistic to determine how well the surface densities are matched \end{enumerate} The modified $\chi^2$ is defined as \begin{equation} \chi^2=\sum_{R_i} \left[\frac{\Sigma_{165}(R_i)-\Sigma_{195}(R_i)}{\Sigma_{165}(R_i)<\Sigma_{195}(R_i)}\right]^2 \end{equation} where the $<$ operator returns the smaller of its two operands. Using this algorithm we search for the values of $\alpha$ and $\beta$ that minimize the modified $\chi^2$. We find good matches for $\alpha=8.67$ and $\beta=-1.08$. We also tried a weighting that only fit points beyond the perturbations of the spiral arms, i.e. $R>2R_0$. Although this did change $v_\Pi(R)$ by 25\% or so, it had no qualitative effect on our other results. \section{Wavelets} The continuous wavelet transform for a discrete series of points $z_n$ is given by \begin{equation}\label{eqn:wavetrans} W_n(s)=\sum^{N-1}_{n'=0} z_{n'} \psi^*\left[ \frac{(n'-n)\Delta\phi}{s}\right] \end{equation} where $N$ is the number of points in the series, $n$ is a position index, $\Delta\phi$ is the spacing of the points in $\phi$ space, and $\psi^*$ is the complex conjugate of the normalized Morlet wavelet \citep{TC1998}: \begin{equation} \psi\left(\eta\right) =\left(\frac{\Delta\phi}{s}\right)^{1/2}\pi^{-1/4}e^{i\omega_0\eta}e^{-\eta^2/2}. \end{equation} The scale of the transform is $s$; for the Morlet wavelet this is simply related to the wavelength by $\lambda=1.03s$. Evenly spaced points are necessary in order to use this transform, but since we are treating the two halves of the Galaxy separately, $s(R,\phi)$ is split into two halves, each with equally spaced points. We perform this transform for the dense set of frequencies given by \begin{equation} s_j=s_02^{j\Delta j}, ~~~j=0,1,\ldots,J. \end{equation} Here, $s_0$ is the smallest scale that can be sampled, $2\Delta\phi$. $\Delta j$ is a measure of how densely we sample in scale space; because computation time is not large, we choose relatively dense sampling throughout: $\Delta j = 0.0125$. From (\ref{eqn:wavetrans}), we construct the wavelet power spectrum, $\left|W_n(s)\right|^2$. We calculate the wavelet power spectrum for the dense set of frequencies, and then filter over a range of scales to find the power in a frequency band. The filtered power spectrum is given by \begin{equation} \overline{W_n}^2=\frac{\Delta j\Delta \phi}{C_\delta}\sum_{j=j_1}^{j_2}\frac{\left|W_n(s_j)\right|^2}{s_j} \end{equation} where $C_\delta$ is a reconstruction factor that depends on the choice of wavelet. For the Morlet wavelet with $\omega_0 =6$, $C_\delta=0.776$.
Title: Recovery of the global magnetic field configuration of 78 Virginis from Stokes IQUV line profiles
Abstract: The surface magnetic field configuration of the Ap star HD 118022 (78 Vir) has been reconstructed in the framework of the magnetic charge distribution (MCD) method from the analysis of Stokes $IQUV$ spectra obtained using the MuSiCoS spectropolarimeter at Pic du Midi Observatory. Magnetically-sensitive Fe~{\sc ii} lines were primarily employed in the analysis, supposing that iron is evenly distributed over the stellar surface. We show that the Stokes $IQUV$ profile shapes and variations of 78 Vir can be approximately fit assuming a global magnetic field configuration described by a slightly decentered, inclined magnetic dipole of polar surface intensity approximately 3.3~kG. The derived inclinations of the stellar rotational axis to the line of sight $i=24\pm 5\degr$ as well as to the magnetic dipole axis $\beta=124\pm5\degr$ are in good agreement with previous estimations by other authors, whereas the sky-projected position angle\thanks{$\Omega$ increases clockwise from the axis to the North Celestial Pole and relates to the azimuth angle $\Theta$ specified by Landolfi at al.~(\cite{Landolfi+93}) as $\Omega=360\degr-\Theta$.} of the stellar rotation axis $\Omega\sim110\degr$ is reported here for the first time. In addition, several lines of Cr~{\sc ii} and Ti~{\sc ii} were studied, yielding evidence for non-uniform surface distributions of these elements, and magnetic field results similar to those derived from Fe.
https://export.arxiv.org/pdf/astro-ph/0601677
\title{Recovery of the global magnetic field configuration of\\ 78 Virginis from Stokes $IQUV$ line profiles} \author{V.R. Khalack \inst{1, 2} and G.A. Wade \inst{3}} \offprints{V. Khalack \\ \email{khalakv@umoncton.ca}} \institute{ D\'{e}partement de physique et d'astronomie, Universit\'{e} de Moncton, Moncton, N.-B., E1A 3E9, Canada \and Main Astronomical Observatory, 27 Zabolotnoho Str., 03680, Kyiv, Ukraine \and Department of Physics, Royal Military College of Canada, PO Box 17000 stn 'FORCES', Kingston, Ontario, Canada K7K 4B4 } \date{Received {\it date will be inserted by the editor}\ Accepted {\it date will be inserted by the editor} } \abstract{The surface magnetic field configuration of the Ap star HD~118022 (78 Vir) has been reconstructed in the framework of the magnetic charge distribution (MCD) method from the analysis of Stokes $IQUV$ spectra obtained using the MuSiCoS spectropolarimeter at Pic du Midi Observatory. Magnetically-sensitive Fe~{\sc ii} lines were primarily employed in the analysis, supposing that iron is evenly distributed over the stellar surface. We show that the Stokes $IQUV$ profile shapes and variations of 78 Vir can be approximately fit assuming a global magnetic field configuration described by a slightly decentered, inclined magnetic dipole of polar surface intensity approximately 3.3~kG. The derived inclinations of the stellar rotational axis to the line of sight $i=24\pm 5\degr$ as well as to the magnetic dipole axis $\beta=124\pm5\degr$ are in good agreement with previous estimations by other authors, whereas the sky-projected position angle\thanks{ $\Omega$ increases clockwise from the axis to the North Celestial Pole and relates to the azimuth angle $\Theta$ specified by Landolfi at al.~(\cite{Landolfi+93}) as $\Omega=360\degr-\Theta$.} of the stellar rotation axis $\Omega\sim110\degr$ is reported here for the first time. In addition, several lines of Cr~{\sc ii} and Ti~{\sc ii} were studied, yielding evidence for non-uniform surface distributions of these elements, and magnetic field results similar to those derived from Fe. \keywords{stars: chemically peculiar -- stars: magnetic fields -- line: polarisation -- stars: individual: HD~118022, 78~Virginis}} \titlerunning{Recovery of the 78 Vir global magnetic field} \authorrunning{V.R. Khalack and G.A. Wade} \section{Introduction \label{intro}} 78 Virginis (HD~118022) is a bright Ap star, and the first star other than the sun in which magnetic field was discovered (Babcock~\cite{Babcock47}). The longitudinal magnetic field of 78 Vir, as first observed by Babcock, is variable, and constantly negative. 78 Vir is also marginally variable in broadband light, in radial velocity, and shows variations of its line profiles (Preston~\cite{Preston69}). As suggested by Preston (\cite{Preston69}), the observed properties of 78 Vir can be explained in the framework of the oblique rotator model (Stibbs~\cite{Stibbs50}), in which the star rotates with a period of approximately 3.7 days and has a rotation axis forming a small angle with respect to the line of sight. The first comprehensive modelling of the surface magnetic field structure of 78 Vir was performed by Borra (\cite{Borra80}). Borra obtained high-resolution measurements of circular polarisation across the Fe\,{\sc ii} $\lambda$4520.2\AA\, line, which he attempted to reproduce using nine different magnetic field configuration models. As argued by Borra (\cite{Borra80}), the profiles were well reproduced with an inclined dipole geometry that contains a moderate quadrupolar component. Nevertheless, he also pointed out that a decentered ($a$=0.2) dipole model provides {essentially} the same fit quality. Borra also concluded that the 78 Vir presents to an observer a remarkably uniform magnetic field over most of its visible disk, at most phases. It is unfortunate that due to the small inclination of the rotational axis (about $25\degr$) a significant part of the stellar surface remains hidden - a part which might have a less uniform field geometry. The analysis of broadband linear polarization (BBLP) measurements of several Ap stars (including 78 Vir) by Leroy~et~al.~(\cite{Leroy+96}) suggests that the magnetic fields of many Ap stars exhibit departures from the standard oblique rotator model assuming a pure dipole field geometry. According to those authors, significant discrepancies between the BBLP observations and the ``canonical model'' results {(Landolfi~et~al. \cite{Landolfi+93})} required the assumption of local departures from a dipolar field. In particular, Leroy~et~al. (\cite{Leroy+96}) showed that the BBLP variations of most stars could be reproduced assuming a dipole magnetic field geometry with slightly expanded field lines over some parts of the magnetic equator. For 78 Vir, the observed BBLP variations do not show especially strong departures from the dipolar case, although according to Leroy~et~al.~(\cite{Leroy+96}) a modified dipolar model provides a somewhat better fit to the BBLP curves for this star. In order to better constrain the magnetic field configuration of 78 Vir, in this paper we undertake a more detailed modelling of the surface magnetic field structure based on the analysis of high resolution line profiles in all 4 Stokes parameters. In Sec.~\ref{obs} we discuss the properties of obtained Stokes $IQUV$ spectra, and in Sec.~\ref{fund} we summarise the fundamental characteristics of 78 Vir. In Sec.~\ref{mod} we describe the features of the magnetic field modelling framework, and derive the global magnetic field configuration of the star, comparing observed and computed Stokes profiles for various spectral features. In Sec.~\ref{discuss} we discuss the derived global magnetic field characteristics and their agreement with earlier studies. \section{Observations} \label{obs} Spectropolarimetric observations of 78 Vir were obtained in 1997 February, 1998 February and 1999 January using the 2m T\'elescope Bernard Lyot at Observatoire du Pic du Midi (Wade~et~al.~\cite{Wade+00a}). The MuSiCoS cross-dispersed \'echelle spectrograph (Baudrand \& B\"{o}hm~\cite{BB92}) and dedicated polarimeter module (Donati et al.~\cite{Donati+99}) were employed for the observations. The MuSiCoS spectrograph is a table-top instrument, which allows the acquisition of a stellar spectrum in a given polarization state (Stokes $V$, $Q$ or $U$) throughout the spectral range from 4500 to 6600~\AA\ with a resolving power of about 35 000, in a single exposure. The spectrograph is fed by a double optical fibre directly from the Cassegrain-mounted polarimeter. The optical characteristics of the polarization analyser, as well as the spectropolarimeter observing procedures, are described in detail by Donati et al.~(\cite{Donati+99}). Observing details specific to the acquisition, reduction and analysis of the 78 Vir spectra are provided by Wade~et~al.~(\cite{Wade+00a}). The journal of spectropolarimetric observations is reported by Wade~et~al.~(\cite{Wade+00a}). The 52 Stokes $V$, $Q$ and $U$ spectra, obtained on 18 different nights, cover the whole rotational period of 78 Vir approximately uniformly, and provide an average S/N of about 370. \section{Fundamental parameters of 78 Vir} \label{fund} 78 Vir was classified by Cowley et al.~(\cite{C2J2}) as A1pCrSrEu. Adelman~(\cite{Adelman73a, Adelman73b}) performed both a line identification and abundance analysis of this star. The distance to 78 Vir $d=48\pm15$ pcs and its radius $R=1.77\pm0.68R_{\rm \sun}$ are derived by Monier~(\cite{Monier92}) using the Infrared Flux Method. On the other hand, taking into account the mean angular diameter $\theta_{\rm a}$= 0.343~{milliarcsec} obtained by Monier~(\cite{Monier92}) for this star and its distance $d=56.2\pm2.5$~pc derived from the Hipparcos parallax (ESA~\cite{ESA97}), we find a somewhat larger value for the stellar radius $R=2.06\pm0.17~R_{\rm \sun}$. Using the Hipparcos visual magnitude and parallax we have found the absolute visual magnitude of 78 Vir, $M_{\rm v}= 1.18\pm0.10$. Taking into account the bolometric correction (Flower~\cite{Flower96}), which corresponds to Monier's (1992) $9200\pm 290$~K effective temperature and the bolometric zeropoint correction BC$_{\rm V}$=-0.07, which is obtained assuming $M_{bol}^{\sun}$=4.74 for the Sun (Bessell~et~al.~\cite{Bessell+98}), we can find the absolute luminosity for 78 Vir $L_{\rm \star}=(27.3\pm2.5)L_{\rm \sun}$. For 78 Vir Monier~(\cite{Monier92}) also estimates the integrated flux $f_{\star}=(2.7\pm 0.27)\times 10^{-7} erg\; sm^{-2} s^{-1}$. Taking into account the distance to the star, this provides the luminosity $L_{\rm \star}=(26.5\pm2.8)L_{\rm \sun}$. Finally, employing the stellar radius and the effective temperature (see Table~\ref{tab2}) we obtain a third estimate of the absolute luminosity $L_{\rm \star}=27.4\pm4.5L_{\rm \sun}$. All of these values are in good mutual agreement. The luminosity $L_{\rm \star}=(27.3\pm2.5)~L_{\rm \sun}$ together with the effective temperature (see Table~\ref{tab2}) allow us to find the stellar mass $M_{\rm \star}=2.18\pm0.06M_{\rm \sun}$ and age $3.0^{+0.7}_{-1.3}\times10^{8}$ years (see Fig.~\ref{HRdiaram}) by spline interpolation (Sandwell~\cite{Sandwell87}) in the model evolutionary tracks of Schaller~et~al.~(\cite{Schaller+92}) for metallicity $Z$=0.02. The derived age suggests that 78 Vir has completed approximately 37\% of its main sequence life. The mass and radius provide $\log g = 4.16\pm 0.07$. This value is marginally inconsistent with that of the Monier~(\cite{Monier92}), but it is in good agreement with the value $\log g =4.20$ reported by King~et~al.~(\cite{King+03}). A recent investigation of the Ursa Major stream characteristics performed by King~et~al.~(\cite{King+03}) derived $Z=0.016-0.02$ and age $(5\pm1)\times10^{8}$ years for the stream. They found that 78 Vir is a "certain member" of the stream from photometric data, while the kinematic characteristics of the star argue for its "probable non-membership". The derived age for 78 Vir is marginally inconsistent with the UMa stream age. The published metallicity of 78 Vir is $\log(Fe/H)_{\star}-\log(Fe/H)_{\sun}=^{-0.13}_{+1.58}$ (Cayrel~de~Strobel~et~al.~\cite{Cayrel+97}) and it is not clear if this star truly has a solar metallicity. For this reason the derived uncertainties of the mass and age may be somewhat larger than indicated here. \begin{table}[t] \caption[]{Fundamental parameters of 78 Vir.} \begin{tabular}{lcc} \hline Parameter & Value & Reference\\ \hline Spectral type& A1pCrSrEu&Cowley et al.~(\cite{C2J2})\\ Age & $3.0^{+0.7}_{-1.3}\times10^{8}$y.& this work\\ Distance & 56.2$\pm$2.5 pcs& Hipparcos (ESA~\cite{ESA97})\\ Period & 3$.^{d}$7220$\pm0.^{d}$003 & Preston~(\cite{Preston69}) \\ $T_{\rm eff}$&9200$\pm$290K &Monier~(\cite{Monier92}) \\ $\log{g}$ &4.50$\pm$ 0.25 & ibidem \\ $M_{\rm v}$ & 1.18$\pm$0.10 &this work\\ $L_{\rm \star}$& 27.3$\pm$2.5 $L_{\sun}$&this work\\ $R_{\rm \star}$& 2.06$\pm$0.17$R_{\sun}$& this work\\ $M_{\rm \star}$& 2.18$\pm$0.06$M_{\sun}$& this work \\ $V_{\rm e}\sin{i}$& 12$\pm$1 km s$^{-1}$& this work \\ $i$ & 25$\degr\pm$5$\degr$ &Leroy~et~al.~(\cite{Leroy+96}) \\ $\beta$ & 120$\degr\pm$5$\degr$&ibidem \\ \hline \end{tabular} \label{tab2} \end{table} Babcock~(\cite{Babcock47}) observed that the longitudinal magnetic field of 78 Vir was variable, but always negative. Analysing Babcock's data along with his own measurements, Preston~(\cite{Preston69}) determined the periodic character of the magnetic field variability and derived the ephemeris: \begin{equation}\label{ephemeris} {\rm JD}\ {\rm (magnetic\;\; maximum)}= 2 434 816.9 + 3.^{d}7220\cdot {\rm E}. \end{equation} \noindent Several authors have confirmed this period on the basis of further magnetic field measurements (Wolff~\&~Wolff~\cite{W+W71}; Wolff~\&~Bonsack~\cite{W+B72}; Wolff~\cite{Wolff78}; Borra~\cite{Borra80}; Borra~\& Landstreet~\cite{B+L80}; Landstreet~\cite{Landstreet82}; {Wade~et~al.~\cite{Wade+00b}), while Landstreet~(private communication) has estimated the accuracy of the period determination to be $\pm0.^{d}003$. Preston~(\cite{Preston69}) also found that the crossover effect is very pronounced and is present throughout most of the magnetic cycle. This feature is particularly remarkable at phase 0.85 (Wolff~\&~Bonsack~\cite{W+B72}, see also Wade et al.~\cite{Wade+00a}). Applying the oblique rotator model with the distorted dipole approximation, % Leroy~(private communication) found that they must adopt a period 3.7218 days in order to fit the observed BBLP variations. As discussed in Sect.~\ref{intro}, their model, using local modifications of the axisymmetric magnetic field imposed near the magnetic equator, differs only mildly from a dipole configuration. The derived rotational axis inclination and dipole obliquity $i=25\degr$ and $\beta=120\degr$ provide a good match to the longitudinal field and linear polarisation variations. Nevertheless, the period inferred is somewhat shorter than that obtained by other authors (Eq.~\ref{ephemeris}), and Wade~et~al.~(\cite{Wade+00b}) reported that this period is only barely consistent with the available longitudinal field measurements. 78 Vir also shows photometric variability in the visible and the infrared with the same period (Eq.~\ref{ephemeris}). Combining the visual flux variability in the $uvby$ bands, obtained by Wolff~\&~Wolff~(\cite{W+W71}), with their own new observations, Catalano~\&~Leone~(\cite{C+L94}) have tried to improve the period determination. Their analysis results in the new ephemeris \begin{equation}\label{ephemeris-new} {\rm JD}\ {\rm (y\; min.)}=2 434 816 + (3.^{d}722084 \pm 0.^{d}000042) {\rm E}, % \end{equation} \noindent which was recently confirmed by Leone~\& Catanzaro~(\cite{L+C01}). These authors have shown with the help of Hipparcos photometry that the light variations of 78 Vir are not purely sinusoidal. Taking into account all archival magnetic field observations as well as their own data, Leone~\&~Catanzaro~(\cite{L+C01}) have shown that the longitudinal field measurements are in phase only when the 3.722084 day period is adopted. Moreover, by adopting the 3.7218 day period suggested by Leroy~(private communication), they find a 0.12-cycle phase shift between the Hipparcos light curves and the other light curves. Monier~(\cite{Monier92}) has derived for 78 Vir from the simulation of the energy distribution in the spectral range from 1200\AA\, to 22000\AA\, an effective temperature $T_{\rm eff}=9200\pm290$~K, a surface gravity $\log{g}=4.5$ and a photospheric metallicity $[M/H]=10$. In addition, using a model atmosphere with these values of $\log{g}$ and $[M/H]$, he has found {that the effective temperature $T_{\rm eff}=9300$K provides the best description of the spectral energy distribution in the region from 1200\AA\, to 8000\AA\, at rotational phase 0.0, while $T_{\rm eff}=9200$K seems to be the best effective temperature, at phase 0.5. These temperatures are somewhat lower than most previous estimates, ranging from 9700K to 10700K (Mihalas~\&~Henshaw~\cite{M+H66}; Wolff~\cite{Wolff67}; Jugaku~\&~Sargent~\cite{J+S68}), probably because the models used by these authors were unblanketed and calculated for solar abundances. The fundamental parameters of 78 Vir are summarised in Table~\ref{tab2}. \section{Modelling of the Stokes $IQUV$ spectra} \label{mod} According to the discussion of Sect.~\ref{fund}, an ATLAS9 model (Kurucz~\cite{Kurucz94}) with parameters $T_{\rm eff}=9250$K, $\log{g}=4.5$, [M/H]=0 and microturbulent velocity $v_{\rm t}$=0 km s$^{-1}$ % successfully approximates the stellar atmosphere of 78 Vir. We describe the structure of the surface magnetic field of 78 Vir (assumed to be a rigidly rotating and spherically symmetric star) using the {\em magnetic charge distribution} (MCD) method (Gerth~et~al.~\cite{gerth+}; Khalack~et~al.~\cite{khalack+}). Originally, the MCD method considered a system of spatially separated {\it point field sources} with {\it virtual magnetic charges} in the stellar interior. {In order to provide zero magnetic flux through the stellar surface the sum of ``magnetic charges'' should be kept to zero.} These sources produce a magnetic field whose potential at each point on the stellar surface is specified by the superposition of potentials of individual sources (Gerth~et~al.~\cite{gerth+}; Gerth~\&~Glagolevskij~\cite{gerth+01, gerth+04}). Since the number of sources is usually more than one, we actually operate with a system of several magnetic dipoles located in the stellar interior. In order to minimize the number of free model parameters, the most convenient way is to consider a system of two sources, which mathematically formulate a magnetic dipole (Khalack~et~al.~\cite{khalack+}). When the dipole is centered on the stellar centre and the dipole size is much smaller than the stellar radius (Khalack~\cite{khalack02}), the MCD model transforms to the conventional model of a symmetric inclined magnetic rotator (Stibbs~\cite{Stibbs50}). Otherwise, we deal with a more complex magnetic field configuration.% The mathematical verification of the MCD model as well as the procedure of specification of the angle $\beta$ between the rotational and magnetic dipole axes, the coordinates and the field strengths of the positive and negative magnetic poles on the basis of the derived free model parameters is described in detail by Khalack~et~al.~(\cite{khalack+03}). \begin{table}[t] \caption{Here the individual columns specify the ion, the wavelength, the quantum number $J$ and the Land\'e factor for lower and upper atomic levels, the observed Stokes $I$ profile depth and the line scaling factor (LSF) (see Sect.~\ref{integral}), derived from the comparison of simulated equivalent width of the Stokes $Q$ and $U$ Zeeman profiles with the BBLP data (Leroy~\cite{Leroy95}). Asterisks marks the LSF calculated from Stokes $Q$ and $U$ profiles simulated for a field structure that is derived from only Stokes $I$ and $V$ profile variability (see Table.~\ref{fe2sm}).} \begin{tabular}{lccccccl} \hline Ion& $\lambda$, \AA &$J_{\rm lo}$&$g_{\rm lo}$&$J_{\rm up}$&$g_{\rm up}$&$I_{\rm obs}$&LSF\\ \hline Fe\,{\sc ii} & 4620.52& 3.5 & 1.21 & 3.5 & 1.40 & 0.33&0.035\\ Fe\,{\sc ii} & 4635.32& 2.5 & 1.20 & 3.5 & 1.13 &0.32&0.036$^*$\\ Fe\,{\sc i} & 4635.85& 1.0 & 1.49 & 2.0 & 1.89 & & \\ Fe\,{\sc ii} & 4923.93& 2.5 & 2.00 & 1.5 & 2.40 &0.58&0.079\\ Ti\,{\sc i} & 5017.95& 4.0 & 1.50 & 5.0 & 1.18 & & \\ Cr\,{\sc i} & 5018.15& 2.0 & 1.16 & 3.0 & 1.09 & & \\ Fe\,{\sc ii} & 5018.44& 2.5 & 2.00 & 2.5 & 1.87 &0.60&0.102\\ Cr\,{\sc ii} & 5018.84& 3.5 & 1.24 & 3.5 & 1.42 & & \\ Fe\,{\sc ii} & 5100.61& 4.5 & 1.54 & 4.5 & 1.34 & & \\ Fe\,{\sc ii} & 5100.66& 4.5 & 1.31 & 3.5 & 1.40 & & \\ Fe\,{\sc ii} & 5100.73& 4.5 & 1.54 & 5.5 & 1.36 & & \\ Fe\,{\sc ii} & 5100.85& 1.5 & 0.80 & 1.5 & 0.72 &0.47&0.083$^*$\\ Fe\,{\sc i} & 5168.90& 3.0 & 1.50 & 3.0 & 1.75 & & \\ Fe\,{\sc ii} & 5169.03& 2.5 & 2.00 & 3.5 & 1.70 &0.59&0.178\\ Fe\,{\sc i} & 5169.30& 4.0 & 1.26 & 3.0 & 1.32 & & \\ Fe\,{\sc ii} & 5197.48& 2.5 & 1.20 & 1.5 & 0.72 & & \\ Fe\,{\sc ii} & 5197.58& 2.5 & 0.57 & 1.5 & 0.44 &0.39&0.073\\ Fe\,{\sc ii} & 5362.74& 3.5 & 1.15 & 4.5 & 1.24 & & \\ Fe\,{\sc ii} & 5362.87& 4.5 & 1.15 & 3.5 & 1.40 &0.45&0.112$^*$\\ Cr\,{\sc i} & 5362.96& 3.0 & 1.33 & 3.0 & 1.06 & & \\ Fe\,{\sc ii} & 5362.97& 3.5 & 1.44 & 4.5 & 1.34 & &\\ Cr\,{\sc ii} & 5363.88& 3.5 & 1.71 & 3.5 & 1.39 & & \\ Fe\,{\sc ii} & 6247.35& 1.5 & 0.40 & 2.5 & 0.62 & & \\ Fe\,{\sc ii} & 6247.56& 2.5 & 1.33 & 1.5 & 1.72 &0.36&0.036$^*$\\ Fe\,{\sc ii} & 6432.68& 2.5 & 2.00 & 2.5 & 1.65 &0.24&0.026\\ Fe\,{\sc ii} & 6516.08& 2.5 & 2.00 & 3.5 & 1.58 &0.25&0.032\\ \hline \end{tabular} \label{tab3} \end{table} \subsection{Procedure \label{proc}} From previous analyses of the Stokes $IQUV$ spectra of 78 Vir, it has been found by Wade~et~al.~(\cite{Wade+00a}) that the {three} strong Fe\,{\sc ii} lines $\lambda$4923.93\AA, $\lambda$5018.44\AA\, {and $\lambda$5169.03\AA\,} show the strongest Zeeman signatures of any lines in the optical spectra of CP stars. The real magnetic field structure of 78 Vir is expected to differ only marginally from a centered magnetic dipole (based on previous modeling efforts: Borra~\cite{Borra80}).% The most attention is therefore paid to the variability of the Stokes $I$ and $V$ profiles, which contain the most information about the global surface magnetic field configuration. The Stokes $I$ variation provides information about the surface distribution of the chemical abundance, as well as the stellar radial velocity $V_{\rm r}$ and projected rotational velocity $V_{\rm e}\sin{i}$. The Stokes $V$ profiles are sensitive to the global field configuration, in particular the lower-order multipolar components. The Stokes $Q$ and $U$ profiles also provide some constraint on the global field morphology, but are most sensitive to the smaller-scale structure of the field. Unfortunately, the available Stokes $Q$ and $U$ profiles have low relative S/N ratio (Wade~et~al.~\cite{Wade+00a}), and in this study they are used primarily to check the results of the Stokes $I$ and $V$ profile simulations and to determine the sky-projected position angle of the rotational axis, $\Omega$. According to Khalack et al.~(\cite{khalack+, khalack+03}) this angle ($0\leq\Omega<360\degr$) is counted clockwise from the rotation axis to the North Celestial Pole in the plane of the sky, and is related to the azimuth angle $\Theta$ specified by Landolfi et al.~(\cite{Landolfi+93}) by $\Omega=360\degr-\Theta$. The opposite counting of the position angle in the MCD model is compensated for by the negative sign in definition of the $B_{\rm y}$-component of the local magnetic field. Hence a right-handed reference frame is applicable. We have examined the spectra for additional lines with prominent polarisation signatures, in order to compile a list of lines that are especially appropriate for this task. To perform the line identification we use the VALD-2 resources (Kupka~et~al.~\cite{Kupka+99}; Ryabchikova~et~al.~\cite{Ryab+99}). The adopted list contains Fe\,{\sc ii}, Fe\,{\sc i}, Cr\,{\sc ii}, Cr\,{\sc i}, Ti\,{\sc ii}, Ti\,{\sc i} and Mg\,{\sc ii} lines, but for the present study we will concentrate primarily on the strongest Fe\,{\sc ii} lines. The main reason for this is that Fe is presumably almost uniformly distributed over the stellar surface (see Sect.~\ref{res}) and in this way we exclude from the fitting procedure a model of the surface abundance distribution. Table~\ref{tab3} presents the final list of Fe\,{\sc ii} lines used in the simulation procedure, together with lines of some other elements that are responsible for blends. Some of the Fe\,{\sc ii} lines listed in Table~\ref{tab3} have no detectable features in the Stokes $Q$ and $U$ spectra. Nevertheless, they are comparatively strong lines (in the Stokes $I$ spectra) with clear variability in the Stokes $V$ spectra, and are analysed without the linear polarization data in order to check our final results. All line profiles are simulated using the {\sc Zeeman2} polarised spectrum synthesis code (Landstreet~\cite{Landstreet88}; Wade~et~al.~\cite{Wade+01}). The code has been modified to include a magnetic field described within the framework of the MCD method (Khalack~et~al.~\cite{khalack+03}), and to allow for an automatic minimization of the model parameters using the {\it downhill simplex method} (Press~et~al.~\cite{press+}). The relatively poor efficiency of the downhill simplex method, requiring a large number of function evaluations, is a well-known problem. Repeating the minimization routine 3$\div$4 times in the vicinity of a supposed minimum in the parameter space allows us to check if the method converges to a global minimum. In our case, this technique requires a comparatively long computational time due to the large amount of analysed observational data and the large number of free parameters. \begin{table*}[th] \parbox[t]{\textwidth}{ \caption[]{Results of Fe\,{\sc ii} lines simulation for $T_{\rm eff}=9250$K, $\log{g}=4.5$ and $v_{\rm t}$=0 km~s$^{-1}$. % The first column indicates the free parameters, while the other columns specify the parameter values for the given line profile. The last column provides the averaged estimation errors for each parameter. The first 6 rows show the fit quality for the each analysed Stokes profile (Eq.~\ref{chi2}), weighted and unweighted (Eqs.~\ref{chi2wa}-\ref{chi2o}) $\chi^2$-function. The following 12 rows show the best fit values of the free model parameters, while the final 9 rows provide the characteristics of the magnetic dipole poles, which are derived from the free parameters. } \label{fe2sm} \vspace{0.in} \begin{tabular}{l|cccccccccccc} \hline\hline Line, \AA & 4620 & 4635 & 4923 & 5018 & 5100 & 5169 & 5197 & 5362 & 6247 & 6432 & 6516 & $\sigma_{\rm er}$\\ \hline $\chi^2_{I}$ & 2.92& 10.91& 8.96& 6.97& 7.40& 13.10& 16.06& 5.54& 2.27& 8.83& 5.94& \\ 4$\chi^2_{V}$& 6.14& 7.86& 9.14& 8.69& 6.97& 9.09& 8.71& 6.64& 5.00& 12.24& 7.39& \\ 6$\chi^2_{Q}$& 7.75& - & 7.73& 5.76& - & 8.23& 6.35& - & - & 12.15& 9.30& \\ 6$\chi^2_{U}$& 7.42& - & 10.04& 6.82& - & 10.79& 6.80& - & - & 9.99& 6.19& \\ $\chi^2_w$ & 6.06& 9.39& 8.97& 7.06& 7.19& 10.30& 9.45& 6.09& 3.64& 10.81& 7.20& \\ $\chi^2$ & 1.75& 6.44& 3.55& 2.81& 4.57& 4.64& 5.11& 3.60& 1.76& 3.90& 2.59& \\ \hline Qr, kG & 180 & 254 & 303 & 202 & 230 & 221 & 151 & 192 & 398 & 173 & 289 & 50\\ $a_{1}, 10^{-3}$&8.4& 6.7 & 4.3 & 6.7 & 5.3 & 7.3 & 12.6 & 9.7 & 5.4 & 8.9 & 6.4 & 1.2\\ $\lambda_{1}$& 13\dr& 24\dr& 34\dr& 41\dr& 33\dr& 15\dr& -4\dr& 19\dr& 69\dr& 23\dr& 20\dr& 5\dr\\ $\delta_{1}$ &-45\dr&-46\dr&-52\dr&-51\dr&-60\dr&-30\dr&-40\dr&-44\dr&-39\dr&-45\dr&-39\dr& 8\dr\\ $a_{2}, 10^{-3}$&3.1& 3.7 & 3.5 & 5.2 & 4.3 & 2.5 & 1.6 & 3.7 & 5.1 & 3.8 & 2.7 & 0.3\\ $\lambda_{2}$&118\dr&125\dr&144\dr&135\dr&145\dr& 85\dr&111\dr&101\dr&113\dr& 96\dr&111\dr& 10\dr\\ $\delta_{2}$ &-16\dr& -9\dr& -9\dr&-14\dr&-12\dr&-11\dr&-11\dr&-26\dr&-11\dr&-27\dr&-11\dr& 10\dr\\ $\Omega$ &109\dr& - &109\dr&109\dr& - &109\dr&102\dr& - & - &164\dr&128\dr& 17\dr\\ $i$ & 22\dr& 24\dr& 22\dr& 23\dr& 25\dr& 21\dr& 29\dr& 27\dr& 26\dr& 29\dr& 24\dr& 5\dr\\ $\log(Fe/N_{\rm tot})$& -3.36& -2.82& -3.26& -3.24& -3.20& -3.12& -3.25& -2.86& -3.17& -2.97& -3.50& 0.22\\ $V_r$ & -8.78& -8.42& -7.82& -7.51& -8.69& -7.23& -7.93& -8.25& -7.99& -6.57& -6.74& 1.18\\ $V_{\rm e}\sin{i}$& 10.9 & 11.1 & 11.9 & 12.0 & 11.5 & 12.6& 11.9 & 13.0 & 11.9 & 13.7 & 11.5 & 0.9\\ \hline $\beta$ &125\dr&124\dr&122\dr&120\dr&123\dr&118\dr&129\dr&126\dr&123\dr&126\dr&122\dr& 5\dr\\ $a_{0}, 10^{-3}$&4.5& 3.9 & 2.9 & 6.0 & 3.4 & 4.3 & 6.5 & 5.8 & 4.8 & 5.6 & 3.6 & 1.6 \\ $a, 10^{-3}$ & 4.5 & 3.8 & 2.6 & 0.7 & 3.4 & 3.4 & 6.2 & 4.4 & 2.1 & 3.8 & 3.4 & 1.6\\ $B_{p}$, kG & 3.25 & 3.95 & 3.47 & 3.18 & 3.13 & 3.02 & 4.00 & 3.45 & 3.42 & 2.69 & 3.92 & 0.7\\ $\lambda_{p}$& -9\dr&-10\dr& -6\dr& -7\dr&-10\dr& -8\dr&-11\dr& -8\dr&-11\dr& -8\dr& -8\dr& 5\dr\\ $\delta_{p}$ &-35\dr&-33\dr&-30\dr&-30\dr&-33\dr&-28\dr&-37\dr&-36\dr&-34\dr&-36\dr&-32\dr& 12\dr\\ $B_{n}$, kG & -3.18& -3.82& -3.46& -3.16& -3.12& -2.95& -3.85& -3.36& -3.41& -2.62& -3.86& 0.7\\ $\lambda_{n}$&170\dr&170\dr&174\dr&173\dr&170\dr&172\dr&168\dr&171\dr&169\dr&171\dr&172\dr& 5\dr\\ $\delta_{n}$ & 35\dr& 33\dr& 30\dr& 30\dr& 33\dr& 28\dr& 37\dr& 35\dr& 33\dr& 36\dr& 32\dr& 12\dr\\ \hline\hline \end{tabular} } \end{table*} The free model parameters employed in the line profile simulation can be divided into two groups: those that describe the magnetic field structure, and those that describe the stellar geometry and atmospheric characteristics. The first group includes the parameters $Q_{\rm r}$, $a_{\rm 1}$ and $a_{\rm 2}$ (which specify the modulus of ``magnetic charges" and their distance from the center of the star, expressed in the units of stellar radius), and the parameters $\lambda_{\rm 1}, \delta_{\rm 1}$ and $\lambda_{\rm 2}, \delta_{\rm 2}$ (which specify the spherical coordinates of the ``charges" in the rotational reference frame of the star). Meanwhile, the second group includes the parameters $\Omega$ (the position angle), $i$ (the rotational axis inclination with respect to the line of sight), $V_{\rm r}$ (the heliocentric radial velocity), $V_{\rm e}\sin{i}$ (the projected rotational velocity), and $\log(N_{\rm x}/N_{\rm tot})$ (the abundance(s) of the element(s) forming the line to be modeled). Instead of $V_{\rm e}\sin{i}$ we can obviously use only $V_{\rm e}$ as a free parameter, but $V_{\rm e}\sin{i}$ is preferable here, because it allows us to compare our estimate of this parameter directly with the results of other authors. The minimum number of free model parameters is 11 for the MCD model if we do not take into account the linear polarization data, and 12 if we do. Sometimes, to obtain a good fit to the line profiles we need to include in the simulation lines of other chemical elements (see Table~\ref{tab3}). In this case the number of free model parameters grows with the number of chemical elements under investigation, but usually does not exceed 15 parameters. Initially we specify arbitrarily the values of the free model parameters in the range of their possible values and simulate the Stokes $IQUV$ spectra at the phases of the observations. We then compare the simulated profiles with the observed spectra, and calculate the reduced $\chi^\mathrm{2}$, which we adopt as a measure of the fit quality. The expression for the $\chi^\mathrm{2}$ function, which reflects the agreement between (for example) the simulated Stokes $I$ ($I_{\rm \lambda_j}(\varphi_{\rm i})$) and observed Stokes $I$ ($I^{obs}_{\rm i,\lambda_j}$) spectra is given by: \begin{equation} \label{chi2} \chi^{2}_{I} = \displaystyle \frac{1}{N_{I}} \sum^{N_{I}}_{i=1} \displaystyle \frac{1}{N_{i}} \sum^{N_{i}}_{j=1} \left( \displaystyle \frac{I^{obs}_{i,\lambda_j}-I_{\lambda_j}(\varphi_{i})} {\sigma[I^{obs}_{i,\lambda_j}]}\right)^{2}, \end{equation} \noindent where $\sigma[I^{obs}_{\rm i,\lambda_j}]$ corresponds to the measurement errors, $\varphi_{\rm i}$ is the rotational phase of the observation, $N_{\rm I}$ specifies the number of spectra, while $N_{\rm i}$ is the number of pixels in each analysed line profile. In fact, although the simulated spectra are calculated with approximately the same resolving power as the observed spectra, this does not provide a direct coincidence of wavelengths in the simulated and observed spectra. Therefore, during the $\chi^\mathrm{2}$ function evaluation the simulated spectral intensity at the exact observed wavelength is calculated using a linear interpolation. The scale of variability and the measurement errors are different for each of the four Stokes $IQUV$ spectra (Wade~et~al.~\cite{Wade+00a}). In order to balance (from a statistical point of view) the information flow from each of the four Stokes spectra, {the respective $\chi^\mathrm{2}$ contributions are weighted. The weights are derived from the comparison of the best fit values of $\chi^2_{\rm V}, \chi^2_{\rm Q}, \chi^2_{\rm U}$ with $\chi^2_{\rm I}$ (when the ordinary $\chi^{2}$ function is minimized) and provide approximately the same weighted value of these functions for the majority of analysed lines (see Table~\ref{fe2sm}). The weighted $\chi^\mathrm{2}_{\rm w}$ functions is specified} in the following way: \begin{equation} \label{chi2wa} \chi^{2}_{w}=\displaystyle \frac{1}{4}[\chi^{2}_{I}+ 4\chi^{2}_{V}+6(\chi^{2}_{Q}+\chi^{2}_{U})], \end{equation} \noindent or \begin{equation} \label{chi2wIV} \chi^{2}_{w}=\displaystyle \frac{1}{2}[\chi^{2}_{I}+4\chi^{2}_{V}], \end{equation} \noindent if we work only with the Stokes $I$ and $V$ spectra. The value of the $\chi^{2}_{\rm w}$ function is reduced throughout the minimization procedure, varying the values of all free model parameters. When it reaches its global minimum, we can also calculate the ordinary $\chi^{2}$ function: \begin{equation} \label{chi2o} \chi^{2}=\displaystyle \frac{1}{4}[\chi^{2}_{I}+ \chi^{2}_{V}+\chi^{2}_{Q}+\chi^{2}_{U}]. \end{equation} \noindent We begin by simulating the strong Fe\,{\sc ii} $\lambda$4923.927~\AA, $\lambda$5018.44 and $\lambda$5169.03 lines, for which the minimization process has been repeated for 13$\div$15 times starting from different locations in the free model parameter space in order to avoid local minima. In the case of 78 Vir, the downhill simplex method converges to four global minima, which are caused by the decentered dipole model symmetry in the MCD method. Two global minima correspond to the parameter sets $i, \beta, \Omega$ and $i, \beta, 180\degr+\Omega$ which differ only by the angle $\Omega$ and provide exactly the same value of the $\chi^{2}_{\rm w}$-function. Models with the opposite direction of stellar rotation ($180\degr-i, 180\degr-\beta, \Omega$ or $180\degr-i, 180\degr-\beta, 180\degr+\Omega$) do not result in minima. The other configurations ($i, 180\degr-\beta, \Omega$ and $i, 180\degr-\beta, 180\degr+\Omega$) correspond to models in which the positive magnetic pole faces the observer. These configurations are not applicable to the case of 78 Vir, where we always see the negative magnetic pole, while the positive pole is partially visible just for phases close to $\varphi=0.0$ (Babcock~\cite{Babcock47}; Borra~\cite{Borra80}). The other two global minima have parameter sets which differ from the previous ones by the location of the ``magnetic charges". The two ``magnetic charges" can be located in two different stellar hemispheres (separated by the equatorial plane), or in a single hemisphere. If they are not significantly shifted from the stellar centre, they can form the same dipole axis and will produce almost the same surface magnetic field configuration. The differences between the two models can be revealed by the sensitivity of $\chi^{2}_{\rm w}$ function on the parameter values, but the minima will still exist. Preference is given to the deepest global minimum, obtained for the model with the two ``magnetic charges" located in the same stellar hemisphere. For the other lines the minimization process was performed only 3$\div$4 times (although if we tested the contribution of blends to the analysed profiles, we ran the minimization routine several more times). Supposing that all analysed lines contain the signatures of the same magnetic field structure, we chose the initial locations in parameter space not far from the parameter set obtained from the strong line analysis. In order to evaluate the fit errors (and therefore the uncertainties on the derived free parameters), we calculate deviations of the simulated profiles produced as a result of small variations of each of the free parameters, thus introducing a small shift along one axis in the $\chi^\mathrm{2}$ hyper-space from the point of the function minimum value. Using this procedure, and taking into account the uncertainties of the observational data and the obtained minimal value of $\chi^\mathrm{2}$-function, we can estimate the errors of the best-fit parameters. \subsection{Results} \label{res} Our initial assumption that iron is uniformly distributed over the surface of 78 Vir appears to be valid, given that the analysed Fe\,{\sc ii} lines reveal no significant variability of the Stokes $I$ profiles with rotational phase. Each Fe\,{\sc ii} line (or group of lines) shown in Table~\ref{tab2} was analyzed independently of the others, using a stellar atmosphere model with $T_{\rm eff}=9250$K, $\log{g}=4.5$, $v_{\rm t}$=0 km s$^{-1}$. For each best-fit simulation the derived free model parameters are given in Table~\ref{fe2sm}. During the simulation of some Fe\,{\sc ii} lines the Stokes $Q$ and $U$ profiles were not taken into account, and consequently the angle $\Omega$ is not determined for those lines. The other data in Table~\ref{fe2sm} are derived from the free model parameters (Khalack~et~al.~\cite{khalack+03}). Here $\beta$ defines the angle between the magnetic dipole axis and the stellar rotation axis (the angle exists if these two axes cross each other). The distance of the magnetic dipole center from the center of the star and one-half of the magnetic dipole size are represented by variables $a_{\rm 0}$ and $a$, respectively. The variables ($B_{\rm p}, \lambda_{\rm p}, \delta_{\rm p}$) and ($B_{\rm n}, \lambda_{\rm n}, \delta_{\rm n}$) specify the location of the positive and negative magnetic poles at the stellar surface and the respective strength of the magnetic field. \subsubsection{The strong lines \label{strong}} The Fe\,{\sc ii} $\lambda$4923.927, $\lambda$5018.44 and $\lambda$5169.03 lines are the most prominent Fe lines in the spectrum of 78 Vir, and show the clearest Stokes $Q$ and $U$ signatures. The first profile is formed essentially by the single line Fe\,{\sc ii} $\lambda$4923.927 (see Table~\ref{tab3}) and appears to contain no important blends. The agreement between the observed and simulated data for Fe\,{\sc ii} $\lambda$4923.927 is similar to that shown in Fig.~\ref{fe5018IVQU} (for Fe~{\sc ii}~$\lambda 5018$). Independent Stokes~$I$ spectra (each corresponding to a slightly different phase) were obtained with each of the Stokes $V$, $Q$ and $U$ spectra, and although only one set of Stokes $I$ profiles are presented in the figure, all were taken into account during the calculation of the $\chi_{\rm I}^2$-function (Eq.~\ref{chi2}). The best-fit values of the Stokes $IVQU$ $\chi^2$-functions for the Fe\,{\sc ii} $\lambda$4923.927 line are given in the third column of Table~\ref{fe2sm}. The second line profile is composed mainly of the Fe\,{\sc ii} $\lambda$5018.44\AA\, line, but is also contaminated by contributions from the Ti\,{\sc i} $\lambda$5017.95, Cr\,{\sc i} $\lambda$5018.15 and Cr\,{\sc ii} $\lambda$5018.84 lines. It seems that the Cr\,{\sc ii} $\lambda$5018.84\AA\, is responsible for the blend in the red wing of the Stokes~$I$ and $V$ profiles (Fig.~\ref{fe5018IVQU}). Supposing a uniform Cr distribution, this blend is well fit for a Cr abundance of $\log Cr/N_{\rm tot}$=-3.38 dex. The best fit $\chi^2$ values for the Stokes $IVQU$ profiles of Fe\,{\sc ii} $\lambda$5018.44 are given in the fourth column of Table~\ref{fe2sm}.% In the case of Fe\,{\sc ii} $\lambda$5169.033, the profile is blended by the weaker Fe\,{\sc i} $\lambda$5168.898 and $\lambda$5169.296 lines (see Table~\ref{tab3}). According to Kochukhov et al.~(\cite{Kochukhov+04}) $\log gf$=-0.786 given in the GRIFON list for the Fe\,{\sc i} $\lambda$5169.296\AA\, line is too high to match the solar spectrum, and they recommend to use a decreased $\log gf$=-2.15. This value is employed here for the respective profile simulation and provides much better agreement with the observed data. The behaviour of the best fit Stokes~$IVQU$ profiles with stellar rotational phase for this particular line is very similar to that shown in Fig.~\ref{fe5018IVQU} for Fe\,{\sc ii} $\lambda$5018.44. The sixth column of Table~\ref{fe2sm} presents the best fit parameters and the model characteristics for Fe\,{\sc ii} $\lambda$5169.033. The concordance of the observed and simulated Stokes~$I$ and $V$ profile variations as a function of stellar rotation is extremely good for these lines. On the other hand, the fit to the Stokes $Q$ and $U$ profiles is only approximate. The simulated profiles show generally the same intensity, qualitative structure and variability as the observations. However, even with these relatively noisy data, it is clear that the model does not reproduce the observations within the errors at some phases (for example, see phases 0.1174, 0.5339 and 0.9851 in Fig. 2). In order to verify the assumption of a uniform surface distribution of iron in the atmosphere of 78 Vir, we have calculated the difference in radial velocity between the observed and simulated Stokes $I$ profiles for the strong Fe\,{\sc ii} lines $\lambda$4923.927, $\lambda$5018.44 and $\lambda$5169.033\AA\, (see Fig.~\ref{VelocityDiff}). The mean (averaged for all the available observational phases) radial velocities obtained from the simulation are taken into account in the calculation of the velocities of the respective simulated profiles. Fig.~\ref{VelocityDiff} shows that the derived differences in $V_{\rm r}$ almost coincide for all three lines. A moderate disagreement may exist only for the data obtained in the vicinity of rotational phase $\varphi$=0.5, when the negative magnetic pole is most visible. The variation of radial velocity, which shows a reasonably coherent variation with phase from about -1 to +1~km/s, may reflect a mildly non-uniform distribution of Fe, unmodelled structure in the magnetic field, or other unaccounted-for physical processes in the stellar atmosphere. \subsubsection{The moderate strength lines} From the list of the Fe\,{\sc ii} lines selected for analysis (see Table~\ref{tab3}) the $\lambda$4620.52, $\lambda$5197.58, $\lambda$6432.68 and $\lambda$6516.08 lines are weaker than the $\lambda$4923.927, $\lambda$5018.44 and $\lambda$5169.03 lines. Due to the weaker polarised signal, a few spectra with comparatively large observational errors have been excluded from the simulation (phases 0.5757, 0.5847 and 0.5914). The line Fe\,{\sc ii} $\lambda$4620.521 is primarily responsible for the formation of the observed profile. No blends were taken into account during the simulation in this case. % The Fe\,{\sc ii} $\lambda$5197.577 line is blended by the weak Fe\,{\sc ii} $\lambda$5197.48 line, but provides the main contribution to the observed profile. The behaviour of the best fit Stokes~$IVQU$ profiles with stellar rotational phase for these lines is very similar to that shown in Fig.~\ref{fe6516IVQU} for Fe\,{\sc ii} $\lambda$4620.521\AA. The first and seventh columns of Table~\ref{fe2sm} % present the best fit parameters and the model characteristics for Fe\,{\sc ii} lines $\lambda$4620.521\AA\, and $\lambda$5197.577\AA, respectively. The Fe\,{\sc ii} $\lambda$6432.68\AA\, line is also essentially unblended. The agreement between the observed and simulated Stokes~$IVQU$ data for this particular line is very similar to that presented by Fig.~\ref{fe6516IVQU}. The Fe\,{\sc ii} $\lambda$6516.08\AA\, line is the primary contributor to the formation of the corresponding observed Stokes~$IVQU$ profiles. The agreement of the best-fit simulation with the observed data for Fe\,{\sc ii} $\lambda$6516.08\AA\, line is shown at Fig.~\ref{fe6516IVQU}. The tenth and eleventh columns at the Tables~\ref{fe2sm} contain the best fit parameters and model characteristics for the Fe\,{\sc ii} $\lambda$6432.68\AA\, and $\lambda$6516.08\AA\, lines, respectively. The best fit of the simulated Stokes~$IVQU$ profiles for the Fe\,{\sc ii} $\lambda$6516.08\AA\, line is statistically better than that obtained for Fe\,{\sc ii} $\lambda$6432.68\AA. \subsubsection{Separate analysis of % the Stokes $I$ and $V$ spectra} The other Fe\,{\sc ii} lines $\lambda$4635.316\AA, the $\lambda$5100\AA-group, $\lambda$5362.87\AA\, and $\lambda$6247.56\AA\, show no variability in the Stokes~$Q$ and $U$ spectra and hence only the Stokes~$I$ and $V$ spectra are taken into account during the simulation. These lines are selected for the analysis in order to check the resulting magnetic field structure of 78 Vir on the basis of a more thorough line list. The Fe\,{\sc ii} $\lambda$4635.316 line is the main contributor to the formation of the corresponding observed Stokes~$I$ and $V$ profiles, although it is blended by the comparatively weak Fe\,{\sc i} $\lambda$4635.846\AA\, line. The agreement between the observed and simulated data for this particular line is similar to that shown in Fig.~\ref{fe5100IV}. The superposition of the Fe\,{\sc ii} $\lambda$5100.607, $\lambda$5100.664, $\lambda$5100.727 and $\lambda$5100.852 lines is responsible for the formation of Stokes~$I$ and $V$ profiles at $\sim 5100.7$~\AA\ (see Table~\ref{tab3}). Fig.~\ref{fe5100IV} shows the simulated Stokes~$I$ and $V$ profiles, which fit well the observed profiles (see also the respective $\chi^2$ value in the fifth column of Table~\ref{fe2sm}). In the region of the Fe\,{\sc ii} $\lambda$5362.87 line the observed Stokes~$I$ and $V$ profiles are formed mainly by this line and by a weak contribution from the Fe\,{\sc ii} $\lambda$5362.74, $\lambda$5362.98, Cr\,{\sc i} $\lambda$5362.87 and Cr\,{\sc ii} $\lambda$5363.88 lines. The Fe\,{\sc ii} $\lambda$6247.557 line also provides the main contribution to the observed Stokes~$I$ and $V$ profiles and is blended by the Fe\,{\sc ii} $\lambda$6247.35 line. The best fit simulated data show almost the same fit quality as is shown in Fig.~\ref{fe5100IV}. Simulation of the Stokes~$I$ and $V$ profiles in the regions of these lines results in almost the same configuration of the magnetic field structure as obtained from the analysis of the stronger Fe\,{\sc ii} lines, for which all four Stokes parameters were taken into account during the simulation. \subsection{Integral magnetic field characteristics} \label{integral} In order to check the agreement of the derived magnetic field model with other available magnetic field data for 78 Vir, we calculate the intensity-weighted, averaged (over the visible stellar disk) longitudinal magnetic field $B_{\rm l}$ and the normalized equivalent widths of the Stokes~$Q$ and $U$ profiles for all the analysed phases. The normalization procedure is performed in accordance with the method described by Wade~et~al. (\cite{Wade+00b}) over the passband of the analysed line profiles. All the available longitudinal magnetic field measurements for 78 Vir are plotted in Fig.~\ref{Blon1} of Leone~\&~Catanzaro~(\cite{L+C01}). This figure shows the good agreement of the $B_{\rm l}$ measurements obtained by Wade~et~al.~(\cite{Wade+00b}) using the Least-Squares Deconvolution (LSD) technique (Donati~et~al.~\cite{Donati+97}) with the majority of other observational data. Therefore, in this paper we just compare our results with the LSD longitudinal field data. As demonstrated by Wade~et~al. (\cite{Wade+00b}), the 78 Vir Stokes~$Q$ and $U$ LSD equivalent widths are proportional to the BBLP measured at that phase (Leroy~\cite{Leroy95}) with a line scaling factor. This factor might be different for the analysed lines (see the last column in Table~~\ref{tab3}), but is the same for the both Stokes~$Q$ and $U$ equivalent widths for a particular spectral line. Neither the $B_{\rm l}$ measurements nor the BBLP data are included directly into the model minimization procedure. However, given that the model successfully reproduces the Stokes profiles with which these quantities are fundamentally related, we should expect that the model magnetic field configuration is capable of reproducing them. Using the same stellar atmosphere model, the longitudinal field calculated from the model fits to the strongest Fe\,{\sc ii} line $\lambda$5018.44\AA\, and to the weak Fe\,{\sc ii} line $\lambda$6432.68\AA\, show a good agreement with the LSD $B_{\rm l}$ data (see Fig.~\ref{Blon1}). The longitudinal fields calculated from the simulation results for the other Fe\,{\sc ii} lines appear to be shifted downward in longitudinal field intensity relative to the LSD $B_l$ variation (see Fig.~\ref{Blon1}). \begin{table*}[th] \parbox[t]{3.5in}{ \center{\caption[]{The same as at Table~\ref{fe2sm}, but for Cr\,{\sc ii} and Ti\,{\sc ii} lines} \label{crti}} \vspace{0.in} \begin{tabular}{l|ccccccc|ccc} \hline\hline Line & Cr\,{\sc ii}& Cr\,{\sc ii}& Cr\,{\sc ii}& Cr\,{\sc ii}& Cr\,{\sc ii}&Cr\,{\sc ii}& & Ti\,{\sc ii}& Ti\,{\sc ii}& \\ \AA & 4592 & 4634 & 5237 & 5310 & 5407 & 5421 & $\sigma_{\rm er}$ & 5188 & 5336 & $\sigma_{\rm er}$ \\ \hline $\chi^2_{\rm I}$ & 14.82& 13.12& 10.51& 6.47& 6.20& 9.48& & 21.53& 12.45& \\ $4\chi^2_{\rm V}$& 8.83& 6.39& 12.70& 8.64& 11.11& 11.50& & 24.28& 9.30& \\ $6\chi^2_{\rm Q}$& 5.62& 6.18& - & 8.18& 8.55& 11.76& & 8.77& 5.79& \\ $6\chi^2_{\rm U}$& 6.48& 5.88& - & 6.63& 9.17& 12.33& & 7.45& 6.45& \\ $\chi^2_{\rm w}$ & 8.95& 7.89& 11.60& 7.49& 8.76& 11.27& & 15.51& 8.50& \\ $\chi^2$ & 4.76& 4.18& 6.84& 2.77& 2.98& 4.09& & 7.58& 4.20& \\ \hline Qr, kG & 175 & 186 & 185 & 182 & 161 & 171 & 30 & 287 & 271 & 40 \\ $a_{\rm 1}, 10^{-3}$&10.2& 9.5 & 9.7 & 8.9 & 9.6 & 8.6 & 1.5 & 3.7 & 3.8 & 1.4 \\ $\lambda_{\rm 1}$& 19\dr& 15\dr& 16\dr& 25\dr& 17\dr& 21\dr& 5\dr& 28\dr& 31\dr&17\dr\\ $\delta_{\rm 1}$ &-41\dr&-42\dr&-45\dr&-49\dr&-40\dr&-46\dr& 5\dr&-39\dr&-41\dr&19\dr\\ $a_{\rm 2}, 10^{-3}$&4.4& 4.1 & 3.8 & 3.9 & 3.6 & 3.7 & 0.3 & 2.8 & 3.5 & 1.0 \\ $\lambda_{\rm 2}$&103\dr&120\dr&111\dr&115\dr&109\dr&120\dr& 8\dr&113\dr&113\dr&19\dr\\ $\delta_{\rm 2}$ &-17\dr&-18\dr&-17\dr&-12\dr& -2\dr&-19\dr&10\dr& -8\dr&-14\dr&12\dr\\ $\Omega$ &109\dr&117\dr& - &123\dr&115\dr&109\dr&14\dr& 97\dr& 90\dr&19\dr\\ $i$ & 26\dr& 26\dr& 26\dr& 26\dr& 22\dr& 27\dr& 5\dr& 25\dr& 18\dr& 5\dr\\ $\log(El/N_{\rm tot})$ &-3.90 & -4.02& -4.04& -3.91& -4.18& -4.17& 0.19& -5.48& -5.68& 0.22\\ $V_{\rm r}$ & -8.25& -8.02& -8.80& -8.41& -8.05& -8.25& 1.23& -5.61& -7.91& 1.26\\ $V_{\rm e}\sin{i}$&11.7 & 10.8 & 11.8 & 11.5 & 10.9 & 12.2 & 1.1 & 13.4 & 13.3 & 1.5 \\ \hline $\beta$ &123\dr&119\dr&126\dr&131\dr&127\dr&124\dr& 5\dr&117\dr&112\dr& 5\dr\\ $a_{\rm 0}, 10^{-3}$&6.1& 5.2 & 5.5 & 5.2 & 5.1 & 4.9 & 1.4 & 2.5 & 2.9 & 1.7 \\ $a, 10^{-3}$ & 5.0 & 5.1 & 5.0 & 4.6 & 5.1 & 4.4 & 1.4 & 2.1 & 2.2 & 1.7 \\ $B_{\rm p}$, kG & 3.52 & 3.86 & 3.71 & 3.36 & 3.32 & 3.05 & 0.4 & 2.46 & 2.41 & 0.5 \\ $\lambda_{\rm p}$&-11\dr&-10\dr&-11\dr& -8\dr& -8\dr& -7\dr& 8\dr&-18\dr&-23\dr& 8\dr\\ $\delta_{\rm p}$ &-33\dr&-29\dr&-36\dr&-41\dr&-37\dr&-34\dr&10\dr&-27\dr&-22\dr&11\dr\\ $B_{\rm n}$, kG & -3.44& -3.78& -3.62& -3.29& -3.24& -2.99& 0.4 & -2.45& -2.41& 0.5 \\ $\lambda_{\rm n}$&169\dr&170\dr&169\dr&172\dr&171\dr&173\dr& 8\dr&161\dr&156\dr& 8\dr\\ $\delta_{\rm n}$ & 33\dr& 29\dr& 35\dr& 40\dr& 37\dr& 34\dr&10\dr& 27\dr& 22\dr&11\dr\\ \hline\hline \end{tabular} } \end{table*} The calculated Stokes~$Q$ and $U$ equivalent widths for all the analysed line profiles are multiplied by the respective line scaling factor in order to fit them to the LSD Stokes~$Q$ and $U$ data. For each analysed line a scaling factor is determined with the help of the Least Squares method. The respective results are given in Table~\ref{tab3}, taking into account that the LSD equivalent widths of Stokes~$Q$ and $U$ profiles have a line scaling factor of 0.1 (Wade~et~al.~\cite{Wade+00b}) from comparison with the BBLP data (Leroy~\cite{Leroy95}). Fig.~\ref{Wade-4923} presents the Stokes $Q$ and $U$ equivalent widths variability with phase, derived from the best fit simulations of Fe\,{\sc ii} $\lambda$4923.927 (with $\Omega=109\degr$) and Fe\,{\sc ii} $\lambda$6516.08 (with $\Omega=128\degr$). They are scaled by a factor of 0.08 (for $\lambda 4923$) and 0.03 (for $\lambda 6516$) and plotted over the Stokes $Q$ and $U$ equivalent widths derived from LSD profiles. It is remarkable that the simulated polarimetric data for Fe\,{\sc ii} $\lambda$4923.927\AA\, and $\lambda$5018.44\AA\, lines have almost the same scaling factor as the LSD Stokes $Q$ and $U$ equivalent widths. Theoretically, the line scaling factor depends on the Land\'e factor and on the degree of line saturation (Wade~et~al.~\cite{Wade+00a}). Table~\ref{tab3} shows that in general the lines with a high mean Land\'e factor have also a comparatively high line scaling factor. Nevertheless, this dependence is not clearly pronounced by the sample of derived data. This is because, for some lines, this effect is reduced by blend contamination of the analysed Stokes $Q$ and $U$ profiles (for example the Fe\,{\sc ii} $\lambda$5100 line group, $\lambda$5197.03 and $\lambda$5362.87 lines) or by weak Stokes $Q$ and $U$ profile variability, which results in a comparatively low precision of the calculated equivalent widths. The LSD net linear polarisation (in the same way as the BBLP) reflect the contributions of many spectral lines with significant polarization. These lines belong to a number of chemical elements, which apparently have non-uniform and usually quite different abundance distributions. Such a composition of polarised features results in LSD (or BBLP) data that may differ substantially from the Stokes $Q$ and $U$ equivalent widths derived for a particular Fe\,{\sc ii} line (see Fig.~\ref{Wade-4923}). This is illustrated by the rather large differences between the two simulated curves in Fig.~\ref{Wade-4923}. Moreover, lines of elements with non-uniform abundance distributions can provide substantially different values of the sky-projected position angle of the stellar rotation axis (see Sect.~\ref{other}), and can exhibit totally different net linear polarisation variations (e.g. 180$\degr$ out of phase with those shown here). The spectral resolution and the observational errors of the available spectra do not allow us to map confidently the detailed distribution of chemical abundances nor the local patterns of the surface magnetic field. The uncertainties in the details of the abundance distributions and the magnetic field structure limit the precision of $\Omega$ as well as the quality of simulated Stokes $Q$ and $U$ equivalent widths. Given these limitations, we would characterise the agreement of the observed and simulated variations shown in Fig.~\ref{Wade-4923} as very acceptable. \subsection{Lines of other elements \label{other}} 78 Vir is a chromium-rich star (Cowley~et~al.~\cite{C2J2}) and has some strong Cr\,{\sc ii} lines, which are located in the observed spectral range. The most prominent of them, Cr\,{\sc ii} $\lambda$4592.049, $\lambda$4634.07, $\lambda$5237.329, $\lambda$5310.687, $\lambda$5407.604 and $\lambda$5420.922\AA\,, have been selected for simulation. Atomic parameters of spectral lines were extracted from the VALD database (Kupka~et~al.~\cite{Kupka+99}) and from Raassen~\&~Uylings~(\cite{R+U98}) ({\it ftp://ftp.wins.uva.nl/pub/orth}) in the case of Cr\,{\sc ii} lines. These lines are not significantly blended by lines of other elements and show variability of the Stokes $IV$ parameters with phase. Some show marginally-detected Stokes $QU$ signatures as well. We analyse also the two strong Ti\,{\sc ii} $\lambda$5188.68 and $\lambda$5336.771 lines. These Ti lines are not significantly blended (although we take into account a small contribution of the V\,{\sc i} $\lambda$5188.885 blend to the profile formed by Ti\,{\sc ii} $\lambda$5188.68). Table~\ref{crti} presents the results of the best fit simulations for the Cr and Ti lines. % The best fit quality is on average similar to that obtained for the Fe\,{\sc ii} lines. The Stokes $I$ profiles for these particular lines vary weakly with rotational phase, suggesting the presence of non-uniform distributions of Cr\,{\sc ii} and Ti\,{\sc ii} in the atmosphere of 78 Vir. To take into account the non-uniform abundance distributions the stellar surface is divided into 30 areas, each characterised by an independent local abundance of the analysed chemical element. The local abundances are included as free parameters in the simulation procedure and the model operates in this case with 41 parameters. The best fit simulations of the aforementioned lines tentatively suggest that chromium is enhanced in the vicinity of the negative magnetic pole, while titanium is underabundant in this region (see Fig.~\ref{Abun}). The improvement in the fit to Cr~{\sc ii} $\lambda 5310$ resulting from a non-uniform abundance distribution is highly significant (reduction of $\chi^2_I$ by over 40\%), although the improvement for Ti~{\sc ii} $\lambda 5336$ is much less so (just 10\%). The averaged chromium and titanium abundances are given in Table~\ref{crti}. It is remarkable that the best fit simulations of Cr\,{\sc ii} lines provide almost the same global magnetic field structure and the sky-projected position angle of the stellar rotation axis as the Fe lines (see Tables~\ref{fe2sm},~\ref{crti}). Meanwhile, the Ti\,{\sc ii} simulation results in somewhat lower values of $\Omega$ and $\beta$. These differences may results from the non-uniform Ti\,{\sc ii} distribution in the stellar atmosphere. \section{Discussion} \label{discuss} The abundance of Fe\,{\sc ii} derived in this study is $\log{Fe/N_{\rm tot}}=-3.16\pm 0.20$ (see Table~\ref{fe2sm}). From analysis of spectra of 78 Vir in the regions 3850-3870~\AA\ and 3868-4650~\AA\ Adelman~(\cite{Adelman73b}) obtained an Fe\,{\sc ii} abundance of $\log{Fe/N_{\rm H}}=-2.89$dex. The differences in the iron abundances {may be due to the inclusion/exclusion of polarized transfer, microturbulence, effects of Balmer line wings, etc.}, as well as from the different effective temperatures that have been employed for the abundance analysis (for 78 Vir Adelman~(\cite{Adelman73a}) applied $T_{\rm eff}=9950$K, significantly higher than the temperatures employed in our model. % For the 11 different Fe features studied here, we observe no systematic difference in strength between weak and strong lines. Therefore, the results of this study are consistent with the absence of important Fe stratification in the atmosphere of 78~Vir. According to our results the global magnetic field structure of 78 Vir is well-described by a slightly decentered magnetic dipole. Usually the starting point of a simulation (in the free parameter hyperspace) is chosen to correspond to a central magnetic dipole, but for all the analysed lines the final model results in a slightly decentered magnetic dipole. This fact is in good qualitative agreement with the results previously obtained by Borra (\cite{Borra80}). In that paper Borra considered the classical decentered dipole model, where the magnetic dipole size is insignificant in comparison with the stellar radius, and obtained $a_{\rm 0}$=0.2. Our model provides a smaller $a_{\rm 0}=0.006\pm 0.002$ due to the non-zero dipole size ($2a=0.008\pm 0.003$), that allows us to take into account the non-symmetrical (with respect to the stellar center) magnetic field configuration of the star. The magnetic field intensity and location of the positive and negative magnetic poles in the stellar rotational reference frame are: \begin{equation}\label{poles} \begin{array}{lll} B_{p}\!=\!3.4\pm 0.7~{\rm kG},&\!\lambda_{p}\!=\!-9^\circ\!\pm 5^\circ,& \!\delta_{p}\!=\!-33^\circ\!\pm 12^\circ; \\ B_{n}\!=\! -3.3\pm 0.7~{\rm kG},&\!\lambda_{n}\!=\!171^\circ\!\pm 5^\circ,& \!\delta_{n}\!=\!33^\circ\!\pm 12^\circ. \end{array} \end{equation} \noindent Due to the shift of the dipole from the stellar centre, the positive magnetic pole has a slightly stronger field intensity than the negative pole, but not all the analysed lines are equally sensitive to this difference. Nevertheless, the final model is hardly distinguishable from a symmetric central dipole model, which provides a slightly poorer agreement between the simulated and observed data. For Fe\,{\sc ii} lines % the centered dipole model results in a $\chi^{2}$-function which exceeds by 3\%$\div$5\% the best fit result from the decentered dipole model. This fact can be related to the quality of the data and to the model sensitivity. The dipole offset could be a result of the particular geometry of 78 Vir, because we never completely see the positive magnetic pole of the star. The modeled global magnetic field configuration of 78 Vir is illustrated in Fig.~\ref{structure} according to the best fit results obtained for Fe\,{\sc ii} $\lambda$5018\AA. Five of the Fe\,{\sc ii} lines analysed using the entire Stokes vector result in a plane-of-sky orientation of the rotational axis of $\Omega=110\degr\pm17\degr$, while the last two lines (Fe\,{\sc ii} $\lambda$6432\AA\, and $\lambda$6516\AA), with comparatively weaker variability of the observed Stokes $Q$ and $U$ profiles, provide higher values of this angle. Fig.~\ref{XOmega} shows the dependence of the $\chi^{2}_{\rm Q}$ and $\chi^{2}_{\rm U}$ functions on the angle $\Omega$ for the three strong Fe\,{\sc ii} lines (see Subsec.~\ref{strong}). Both functions reach their minima at $\Omega=109\degr$ and $\Omega=289\degr$. The second value describes a model which is indistinguishable from the configuration specified in Table~\ref{fe2sm} due to the definition of the Stokes Q and U parameters. As discussed in Sect. 3, 78 Vir is a probable member of the Ursa Major stream. It approaches us with a mean radial velocity $V_{\rm r}=-8.1\pm1.0$ km~s$^{-1}$ (Wade~et~al.~2000a). This velocity varies by about $\pm2$ km~s$^{-1}$ with the phase of stellar rotation (Preston~\cite{Preston69}, this work). The possible variability of the radial velocity is not taken into account in the simulation procedure. Respectively, our estimations of the radial velocity for different Fe\,{\sc ii} lines are distributed around the aforementioned value within the range $\pm1.5$ km~s$^{-1}$. Nevertheless, the $V_{\rm r}$ variability through the rotational cycle has been determined from the strong Fe\,{\sc ii} lines $\lambda$4923.927\AA\, $\lambda$5018.44\AA\, and $\lambda$5169.033\AA\, (see Subsec.~\ref{strong}) and for the other Fe\,{\sc ii} lines as well. We show that the mean radial velocity derived from the simulation varies from line to line (see Table~\ref{fe2sm}). The other Fe\,{\sc ii} lines provide a similar $\Delta V_{\rm r}$ variability. The precision of the mean radial velocity determination depends on the observational errors and on the quality of the description of the surface magnetic field structure. For the blended profiles it also depends on the accuracy of oscillator strengths. The analysed lines show weak, coherent variations of $\Delta V_{\rm r}$ with the period of stellar rotation (see Fig.~\ref{VelocityDiff}). This suggests a possible moderately non-uniform distribution of Fe\,{\sc ii}. Nevertheless, the non-uniform iron distribution will not lead to a significantly different surface magnetic field structure. The simulation of Cr\,{\sc ii} lines with the assumption of a non-uniform chromium distribution resulted in almost the same field structure that we obtained from the simulation of Fe\,{\sc ii} lines. The other best fit parameters are the rotational axis inclination $i=24\degr\pm5\degr$ and the dipole axis obliquity $\beta=124\degr\pm5\degr$, which are in a good agreement with the estimates of Leroy~et~al.~(\cite{Leroy+96}). The derived $V_{\rm e}\sin{i}=12\pm1~{\rm km~s^{-1}}$ provides $V_{\rm e}=29\pm 4~{\rm km~s^{-1}}$ and is consistent with that of Preston~(\cite{Preston71}). The derived surface magnetic field variability interval ranges from 2.1~kG to 3.2~kG and covers the value $B_s$=2.9~kG estimated by Preston~(\cite{Preston71}). Besides, as Fig.~\ref{Blon1} shows, the longitudinal magnetic field variation obtained from the two Fe\,{\sc ii} lines simulation is also in good agreement with the LSD $B_{\rm l}$ data (Wade~et~al.~\cite{Wade+00b}). These facts justify the applicability of the slightly decentered magnetic dipole model with the aforementioned values of free parameters for the global magnetic field structure description at 78 Vir. Unfortunately in the case of 78 Vir the intensity of the Stokes $Q$ and $U$ profiles is similar to the observational errors. The data are unsuitable for performing an analysis of the small-scale structure of the magnetic field, using a more sophisticated technique such as Magnetic Doppler Imaging (MDI; Kochukhov~\&~Piskunov~\cite{K+P02}). However, it appears, given the relatively good agreement between the intensity, structure and variability of the observed and simulated Stokes $QU$ profiles, that the real magnetic field of 78 Vir does not depart strongly from the configuration derived here. At the same time, we have noted that the Stokes $QU$ profiles are not, at some phases, fit to within their errors. Therefore these data, notwithstanding their relatively low S/N, already suggest the limitations of the MCD model framework. A similar analysis of the Stokes $IV$ profile variability was performed by Kochukhov et al.~(\cite{Kochukhov+02}) for $\alpha^2$~CVn. Those authors applied MDI using ``multipolar regularization'', analysing the Stokes $IV$ line profiles and obtained a good agreement between observed and computed profiles for Fe\,{\sc ii}, Cr\,{\sc ii} and Si\,{\sc ii} lines. The quality of the Stokes $I$ and $V$ profile fits obtained in this paper is only marginally poorer than that obtained by Kochukhov et al.~(\cite{Kochukhov+02}), although remarkably the magnetic field model framework employed here is much simpler. New spectro-polarimetric observations of 78 Vir with higher spectral resolution and signal-to-noise ratio (especially for the Stokes $Q$ and $U$ spectra) are required for further refinement of the global magnetic field structure. With such data, a model with additional parameters (such as MDI) could be applied, in order to study the local magnetic field topology and the detailed relationship between the magnetic field and the abundance distributions. \begin{acknowledgements} The authors are grateful to Prof. John Landstreet and to the referee, Dr. O. Kochukhov, for their valuable remarks and advice that have led to the improvement of this paper. The authors acknowledge grant support from the Natural Sciences and Engineering Research Council of Canada, and the Department of National Defence of Canada (DND-ARP). \end{acknowledgements}
Title: Massive and Red Objects predicted by a semianalytical model of galaxy formation
Abstract: We study whether hierarchical galaxy formation in a concordance $\Lambda$CDM universe can produce enough massive and red galaxies compared to the observations. We implement a semi-analytical model in which the central black holes gain their mass during major mergers of galaxies and the energy feedback from active galaxy nuclei (AGN) suppresses the gas cooling in their host halos. The energy feedback from AGN acts effectively only in massive galaxies when supermassive black holes have been formed in the central bulges. Compared with previous models without black hole formation, our model predicts more massive and luminous galaxies at high redshift, agreeing with the observations of K20 up to $z\sim 3$. Also the predicted stellar mass density from massive galaxies agrees with the observations of GDDS. Because of the energy feedback from AGN, the formation of new stars is stopped in massive galaxies with the termination of gas cooling and these galaxies soon become red with color $R-K>$5 (Vega magnitude), comparable to the Extremely Red Objects (EROs) observed at redshift $z\sim$1-2. Still the predicted number density of very EROs is lower than observed at $z\sim 2$, and it may be related to inadequate descriptions of dust extinction, star formation history and AGN feedback in those luminous galaxies.
https://export.arxiv.org/pdf/astro-ph/0601685
\title{Massive and Red Objects predicted by a semianalytical model of galaxy formation} \author{X. Kang$^{1,2}$, Y. P. Jing$^{1}$, J. Silk$^{2}$} \affil{$^1$ Shanghai Astronomical Observatory, Nandan Road 80, Shanghai, China} \affil{$^2$ Astrophysics, University of Oxford, Denys Wilkinson Building, Keble Road, Oxford OX1 3RH, UK} \affil{e-mail: kangx@astro.ox.ac.uk} \keywords{galaxies: formation---galaxies: evolution---galaxies: luminosity function,mass function} \section{Introduction} There are many recent observations of high-redshift galaxies that probe the star formation history of the Universe. The finding of many massive galaxies, especially massive Extreme Red Objects (EROs), at high redshift is particularly interesting. These observations show that some EROs are passive ellipticals, and were already in place at redshift z$\sim 2$. It is usually argued that in a Cold Dark Matter (CDM) universe, structures form via a hierarchical formation process in which small galaxies form first at early times, and massive galaxies form later through the continuous mergers of the smaller systems. With representative semi-analytical models (SAMs; Kauffmann et al. 1999, Somerville \& Primack 1999, Cole et al. 2000), it was found that in the concordance $\Lambda$CDM universe, it is difficult to produce enough massive and red galaxies that look like those observed(e.g. Cimatti et al. 2002a, Glazebrook et al. 2004). On the other hand, the existence of the observed massive galaxies at high redshift is not necessarily in conflict with the concordance $\Lambda$CDM model, because the conversion of just ten percent of baryons in dark matter halos of mass $M >10^{13}M_{\odot}$ to stars is sufficient to produce the number of observed massive galaxies (Somerville 2004a). Many authors have studied the formation of these massive, red objects using SAMs or Smoothed Particle Hydrodynamics (SPH) simulations. It was shown that the SAMs (Kauffmann et al. 1999, Somerville \& Primack 1999, Cole et al. 2000) cannot produce enough massive/red objects at redshift $z>1$ (e.g. Firth et al 2002, Somerville et al. 2004b, Daddi et al. 2005). The SPH simulations (e.g. Nagamine et al. 2004, 2005) have succeeded in producing massive and red galaxies at high redshift, but at the cost of introducing more uncertainties. First, it is unknown if these SPH simulations can produce the local galaxy luminosity function. It seems that these simulations produce too many bright galaxies at $z=0$ (Nagamine et al. 2004). Secondly, Nagamine et al. (2005) used a high dust extinction for the entire galaxy population, but the observations show that some EROs are passive ellipticals with little dust extinction (Cimatti et al. 2002b). The main reason that the SAMs fail to produce enough massive and luminous galaxies at high redshift is that the gas cooling and star formation in early massive halos is over-suppressed. In previous SAMs, the gas cooling in massive halos is switched off in order not to produce more luminous central galaxies than observed at redshift $z=0$. The suppression of gas cooling is also motivated by the X-ray observations that massive cooling flows are not observed in groups and clusters (e.g. Peterson et al. 2003). But as the consequence, the gas cooling may be over-suppressed at high redshift if a simplified prescription is used for the cooling cutoff. For example, in the Munich group model and also in Kang et al. (2005), the gas cooling is shut off by hand in halos with the virial velocity greater than $350km/s$. Since the halo mass is much lower at high redshift than at the present for a given virial velocity, the gas cooling is suppressed in this model for halos with the virial mass greater than 2.5$\times 10^{12}M_{\odot}$ at z $=$ 3. This artificial cooling switch-off seems to be the main reason that these models do not produce as many massive galaxies as observed. In this paper, we implement a new model in which the energy from AGN is used to suppress the cooling of hot gas in halos. Following Kauffmann \& Haehnelt (2000) we use a simple model wherein black holes gain most of their mass during major mergers. Our implementation of the feedback from AGN is very similar to that used recently by Croton et al. (2006) and Bower et al. (2005), and resembles a combination of their models. In our model, the total energy from the AGN is proportional to the Eddington luminosity of the central black hole and the efficiency of reheating the gas is proportional to a power of the virial velocity of the galaxy. Then the energy compensates for the radiative energy of the cooling gas, and the actual cooling rate is determined by the ratio between the two energies. The cooling is totally suppressed if the energy from AGN is larger than the energy radiated by the cooling gas. Compared with the previous model used by Kang et al. (2005) with an artificial cut-off of the gas cooling in the halos with the virial velocity larger than $350km/s$, the gas cooling and AGN feedback in the new model are treated in a more self-consistent way. The $M_{\rm bh}$-$\sigma$ relation of black hole mass $M_{\rm bh}$ and the bulge velocity dispersion $\sigma$ implies that massive black holes are present only in massive spheroids. In our present model, the energy feedback from AGN indeed is efficient in galaxies with a massive spheroid. We also require that the star formation rate in quiescent disks is reduced at high redshift as motivated by the observed evolution of cosmological cold gas content with redshift (Keres et al. 2005); thus the gas-rich mergers result in earlier formation of supermassive black holes in massive central bulges. Once the energy feedback is enough to suppress the gas cooling, the termination of new star formation will soon make the galaxies red. We will compare the model prediction of the number density of luminous galaxies with the K20 survey, and find that good agreement holds up to z$\sim$3, beyond which there is little observational data. Compared with previous SAMs, our present model can also produce some very red ($R-K>5$, magnitudes are given in the Vega system unless otherwise stated) passive ellipticals which are observed by the Great Observatories Origins Deep Survey (GOODS) at z$\sim 1-2$. We arrange our paper as follows. In section 2, we briefly introduce our new model with AGN feedback and compare our model predictions with the local galaxy population. In section 3, we give the model predictions and compare them with the observations at high redshift. Finally, we discuss our results and conclude our work in section 4. \section{Model} The SAM that we use here was described in detail by Kang et al. (2005) who studied the local galaxy population. The merger tree is constructed based on a high-resolution N-body simulation (Jing \& Suto 2002) of 512$^{3}$ particles in a box of 100$h^{-1}{\rm Mpc}$. The cosmological parameters adopted there are $\Omega_{m} = 0.3$, $\Omega_{\Lambda} = 0.7$, $h=0.7$, $\sigma_{8} = 0.9$. Here we still use this simulation, but the SAM model is modified in two ways. 1. We adopt a star formation efficiency $\alpha \sim (1+z)^{-1}$ in a quiescent disk that was shown to give a better match with the evolution of cosmological cold gas content with redshift (Kauffmann \& Haehnelt 2000, P$\acute{\rm e}$roux et al. 2003, Keres et al. 2005). In the recent model of Durham group (Baugh et al. 2005, Bower et al. 2005), they adopt a constant star formation timescale for the disk. The star formation timescale used in our model is the dynamical time of the disk which scales with redshift as $(1+z)^{-1.5}$. So the star formation rate ($\dot{M_{\ast}}=\alpha M_{cold}/t_{dyn}$) of our model differs from that of the Durham model only slightly. Note that the relatively lower star formation rate in quiescent disks leaves more cold gas which helps to produce massive black holes during galaxy mergers at high redshift. 2. We include a model for the growth of black holes and for the energy feedback from AGN to suppress the gas cooling. As the $M_{\rm bh}$-$\sigma$ relation indicates that the central black holes grow with the growth of the spheroid components, it is plausible that the black holes get their mass through major mergers. But it is far from clear about the exact way that the black holes accrete the surrounding material. Here following Kauffmann \& Haehnelt (2000), we use a simple parameterised form to describe the cold gas accreted by the black hole during a major merger, \begin{equation} \Delta M_{bh} = F_{acc} \frac {M_{cold}} {1+(280km/s/V_{vir})^{2}} \end{equation} where $M_{cold}$ is the total cold gas in merging galaxies, and $V_{vir}$ is the virial velocity of the post-merger host halo. We normalize the parameter $F_{acc}$ by best matching the observed $M_{bulge}-M_{bh}$ relation at z=0 (H\"aring \& Rix 2004). During the gas accretion by black holes, part of the gravitational energy will be converted into radiations which in turn will heat the surrounding cold gas. But it is again unclear in a quantitative way about how much the radiation is produced and how efficiently the cold gas is re-heated. Croton et al. (2006) use a simple phenomenological model to describe the accretion rate which depends on the hot gas fraction and circular velocity of the halo, but the efficiency of heating the gas by AGN are the same in all halos of different mass. Sijacki \& Springel (2006) have shown that heating efficiency from a AGN bubble is lower in low mass halos. Here we simply assume that the energy from the central AGN is proportional to the Eddington luminosity $L_{edn}$ and the heating efficiency is proportional to a power of the virial velocity of the host halo. Thus the heating rate ejected into the gas is taken as, \begin{equation} L_{reheat}=F_{0}(V_{vir}/V_{\star})^{n}L_{edn}\,. \end{equation} If we denote the cooling rate in a halo of gas temperature $T$ by $\dot{M}_{0,cool}$ in the case of no AGN feedback, then the cooling rate $\dot{M}_{cool}$ in the presence of AGN feedback is: \begin{equation} \frac {\dot{M}_{cool}} {\dot{M}_{0,cool}} = 1 - \frac {L_{reheat}} {\frac {3} {4}\dot{M}_{0,cool}V_{vir}^{2}}. \end{equation} If the heating rate from AGN $L_{reheat}/\frac{3}{4}V_{vir}^2$ is larger than the radiative cooling rate ${\dot{M}_{0,cool}}$, the gas cooling is totally suppressed. We normalize the parameters $F_{0}$, $V_{\star}$ to get a good match to the galaxy luminosity function at z=0. In our model we obtain $F_{0}=2\times 10^{-5}$ and $V_{\star}=200km/s$ and $n=4$. In Fig.~\ref{fig:Bh_Bulge} we plot the relation between the bulge mass and the black hole mass. The data points show for the model galaxies and the solid line the best fit to the observations by H$\ddot{\rm a}$ring \& Rix (2004). Here $F_{acc}$ is taken to be $0.01$. It is seen that a simple model of black hole growth with a free parameter can reproduce the observed $M_{bulge}-M_{bh}$ relation. After the black hole mass is normalized, we then tune the parameters in equation 2 to get good fits to the local galaxy luminosity functions. In Fig.~\ref{fig:LF_z0} we show the luminosity function at B$_{j}$ and K bands. The upper panel shows a comparison with the 2dFGRS at B$_{j}$ band. The solid circles show the observational data of 2dFGRS, and the thick solid histogram associated with Poisson errors is our model prediction. The lower panel shows the comparison at K band where the circles are from Cole et al. (2001) and squares are the observations by Huang et al. (2003). We find that the new model can produce the local galaxy luminosity functions at blue and near-IR bands which are respectively sensitive to the current star formation rate and the total stellar mass in the galaxies. It has been shown (Croton et al. 2006, Bower et al. 2005) that without an effective energy feedback, the predicted luminosity functions at the bright end are too flat with many more luminous galaxies predicted than observed. Note that here our model predictions at high luminosity ends are still slightly higher than observed. This might point to the fact that a more detailed model is needed for AGN heating in massive halos which we will address in future work. \section{Results at high redshift} As discussed in Section 1, the gas cooling in our new model is not suppressed artificially but by heating due to the energy injected from AGN in the galaxy center. So compared to previous SAMs without AGN, the gas cooling and star formation can continues until a massive spheroid forms at the galaxy center. It is expected that this model can produce more massive and luminous galaxies at high redshift. In Fig.~\ref{fig:K20_LF} we show the predicted rest-frame K band luminosity function at z$\sim 1.5$. The squares with error bars are the observational results from K20 (Pozzetti et al. 2003). The solid circles are the predictions by the new model and the triangles show the results predicted by Kang et al. (2005) where they adopted a artificial shut off of gas cooling in galaxies with $V_{vir}>350km/s$. We also re-plot the results of K band luminosity function at z=0 by the solid line, taken from from lower panel of Fig.\ref{fig:LF_z0}. It is clearly seen from the plot that the new model produces more massive galaxies and the agreement with the observations is very good. Also note that the good agreement holds for faint galaxies as well, whereas it was reported previously that SAM models produce more faint galaxies than observed (Pozzetti et al. 2003). Another test, firstly proposed by Kauffmann \& Charlot (1998), is the evolution of the surface number density of galaxies at a fixed limiting magnitude, which also widely used to constrain the models. There are plenty of data from GOODS that are already publicly available (Giavalisco et al. 2004). In Fig.~\ref{fig:GOODS_num} we show the predicted redshift surface number density of galaxies with K$<20$. The square points show the results of K20 and triangles are the data from GOODS. The new model predictions are shown as the solid line, and the dashed line shows the prediction by the model of Kang et al. (2005). Here we find that compared with Somerville et al. (2004b) who predicted much fewer luminous galaxies at $z>1.5$, the agreement between our model and the observations holds much better up to z$\sim 3$. Here we also show how dust extinction will change the result. The dotted line is the new model with the simple dust extinction model of Calzetti et al. (2000) with $E(B-V)=0.1$. Clearly dust extinction has no significant effect on the predicted number of galaxies in the observed-frame K band up to z=3. Though the predicted numbers of luminous galaxies agree with the observations, it would be interesting to check the predicted color distributions. The color is dependent on the star formation history and on the dust extinction. At high redshift the galaxy mergers are very frequent and the dust extinction is significant in these starburst galaxies, but no reliable model of dust extinction is available for such galaxies. Observations show that at z$\sim 1-2$ the EROs have contributions both from passive ellipticals with little dust and from dust-enshrouded starburst galaxies (Cimatti et al. 2002b, Cimatti et al. 2003, Yan \& Thompson 2003, Yan et al. 2004, Moustakas et al. 2004). Because there are significant uncertainties in the dust extinction modelling for the starburst galaxies, we think that the predicted number density of passive ellipticals should set a more meaningful constraint on the galaxy formation model. Here we take a simple model of dust extinction. We classify the galaxies with starbursts produced during the major mergers in the past 0.1Gyr as young starburst galaxies and those otherwise as passive galaxies. We then use the Calzetti et al. (2000) reddening law to model the dust extinction effect on the galaxy color. The amount of dust in passive and young starburst galaxies is difficult to assess, and here we simply assume a small reddening $E(B-V)=0.05$ for the passive galaxies. The dust extent in young starburst galaxy is expected to be high. Observations of EROs show that some extremely red galaxies have heavy dust extinction with $E(B-V)=0.4$. But the average extinction should be lower. Here we assume a Gaussian distribution of $E(B-V)$ with a mean of 0.1 and a dispersion of 0.05 for the young starburst galaxies. Our main motivation is to see if a simple dust reddening model can produce the main features of the observed color distribution. In Fig.~\ref{fig:GOODS_color} we show the observed $R-K$ (both in the AB magnitude system, $(R-K)_{AB} \simeq (R-K)_{Vega}-1.65$) color distribution with a comparison with the data which are from the GOODS Southern field in an area of 160 arcmin$^{2}$ (Somerville et al. 2004b). The upper panel shows the GOODS data, which is from Figure 2 of Somerville et al. (2004b). The model galaxies are selected using the magnitude cut and are normalized to the same area of 160 arcmin$^{2}$. The total number of galaxies selected in our model is 1595 which is $6\%$ higher than the GOODS data points used here. The lower panel shows the model predictions. In each panel we also show the evolution track of single burst stellar populations with solar metallicity, the Salpeter IMF, and the ages (at $z=0$) of 13.35 and 11.7 Gyrs (i.e. $z_{f}=26, 2.6$) based on the model of BC03 (Bruzual \& Charlot 2003). From the figure, our model can reproduce the main features of the observed galaxies: 1) many extremely red galaxies ($R-K>4$) at $z>1$; 2) the bimodal color distribution, red passive and young starburst galaxies at $z>1.5$. Still there are some discrepancies. The predicted numbers of blue galaxies are too prominent at z$<1.5$ and this might be due to the inadequate treatment of star formation rate, stellar initial mass function, or the dust extinction model. Also the predicted number of extremely red galaxies with $(R-K)_{AB}>3.35$ at $z \sim 2$ is still lower than observed. In our model there are enough luminous galaxies but insufficient number of very red galaxies, which means that the star formation (at $\sim 2$) in the current model are still high. There are two possible reasons for this discrepancy. First the star formation is not strong enough in the past in our model, as we do not include any star formation during minor mergers which are also frequent at early times. Second the energy from central AGN is not high enough to suppress the hot gas cooling. Observations have shown that there are already massive black holes ($\sim 10^{9}M_{\odot}$) at $z \sim 6$ (Fan et al. 2001), so the growth of black holes in massive galaxies might be much quicker at early time than in our model in which the fraction of cold gas accreted by black hole is constant with time. We will address this in a forthcoming paper (Kang et al. 2006). Glazebrook et al. (2004) used the Gemini Deep Deep Survey (GDDS) to obtain the stellar mass distribution from $z \simeq$ 0.7 to 2. The evolution of stellar mass density does place important constraints on the formation model of massive spheroids. But due to the uncertainties in fitting the multi-broad band colors of high redshift galaxies including those of the IMF and dust extinction, the constraints are weak. In Fig.~\ref{fig:stellar_GDDS}, we show the stellar mass density of galaxies with stellar mass above certain limits. The lines show the predicted stellar density in galaxies with stellar mass in the range indicated in the plot. Black lines are for this model and blue lines are from the model of Kang et al. (2005) where they used an artificial cut of gas cooling in the halos with $V_{vir} > 350 km/s$. We can still see a good match between the model and the data. Although it seems that the stellar mass density with $M_{\star}> 10^{10.46}M_{\odot}$ is higher than the data points, it agrees with the integral of the star formation rate (see figure 4 of Glazebrook et al. 2004). Note that galaxies with $M_{\star}>10^{11}M_{\odot}$ are in the sharply declining tail of the mass function, therefore a small uncertainty in the estimated stellar mass can introduce a very large uncertainty in the number density. The hexagon in the plot shows the stellar mass density of massive galaxies with $M_{\star}>10^{11}M_{\odot}$ recently obtained by van Dokkum et al. (2006) making use of the deep multi-wavelength GOODS, FIRES and MUSYC surveys. It is seen from the black dashed lines that our model prediction is slight lower than the data by a factor of 2. At high redshift the cosmic variance is so large in the observed catalogs (about $60\%$, Somerville et al. 2004c) that the discrepancy might not be serious. \section{Discussion} Here we have implemented a new semi-analytical model in which the energy from AGN suppresses hot gas cooling in massive halos. The growth of black holes and bulges, and the gas cooling, are determined in a self-consistent way. In our description, the AGN feedback becomes efficient in massive galaxies after a massive black hole is formed in the galaxy center. The AGN feedback model has drawn much recent attentions. The main motivation is that in massive groups and clusters cooling flows are not observed. There should be some physical process to reheat the cooling region, and the energy from AGN has been proposed as an effective source (e.g. B$\ddot{\rm o}$hringer et al. 2002, Begelmen et al. 2002, Sijacki \& Springel 2006). At the same time, the AGN feedback models have also been incorporated into the SAMs recently and it has been shown that AGN feedback can produce a break of the luminosity function at the bright end and produce the color-magnitude relation observed in SDSS (Croton et al. 2006, Bower et al. 2005). Our model of AGN feedback is very similar to theirs in spirit, but the detailed prescription is different. In this paper we use this model to address some issues about the number distribution and color distribution of galaxies at high redshift. We compare the model predictions with the K20 and GOODS surveys. Our conclusions are as follows. \begin{itemize} \item The predicted number distribution of $K<20$ galaxies matches well with that of the GOODS and K20 galaxies up to a redshift of z $\sim$ 3; \item The predicted color distribution is similar to that observed in the surveys and many extremely red galaxies ($R-K_{AB}>4$) are produced, which has not been seen in previous models (Somerville et al. 2004b). At $z > 1.5$ the galaxy population already displays a bimodal color distribution; \item The predicted stellar mass density can marginally agree with the GDDS observation even with the uncertainties in the IMFs; \end{itemize} These results demonstrate that it is not difficult to produce massive and red galaxies at z $\sim$ 1-2 in the concordance CDM universe. The stellar mass in galaxy centers continues to grow until the energy from central AGN is high enough to suppress the gas cooling. In our model the black holes acquire most of their mass during major mergers, so the AGN energy feedback is expected to be effective after the last major merger which led to massive bulge formation at galactic centers. In our model we can produce some of those passive ellipticals at z$\sim 1-2$ with extremely red colors $(R-K)_{AB}>4$. Many observations have shown that the star formation rate was higher in massive galaxies at high redshift and these support the "downsizing" formation scenario (Cowie et al. 1996). It is often argued that hierarchical galaxy formation cannot reproduce the downsizing formation process. But recent works (de Lucia et al. 2005, Bower et al. 2005, Scannapieco et al. 2005) have shown that models with AGN feedback in the hierarchical universe can reproduce the downsizing process in which the massive galaxies forms earlier. In this paper, we also find that the predicted luminous and massive galaxies are increased to the degree that is in agreement with the observations, though the predicted number of red galaxies may still be fewer than observed. Once more observations are available on the dust extinction in these galaxies, the number density and evolution of red passive ellipticals will put more stringent constraints on the galaxy formation models. It is also possible that a new ingredient is needed, such as the star formation induced by AGN feedback prior to disruption of the cold gas supply (Silk 2005), in order to make bulge formation more efficient and to account for the chemical evolution of massive early-type galaxies. \acknowledgments We thank Mashiro Nagashima for kindly providing the GOODS data, and Manfred Georg Kitzbichler for the binned data of GOODS and K20. Xi Kang acknowledge support from the Royal Society China Royal Fellowship Fellowship scheme. This work is supported in part by NSFC(No. 10373012, 10533030) and by Shanghai Key Projects in Basic research (04jc14079, 05xd14019). \newpage \clearpage \clearpage \clearpage \clearpage \clearpage \clearpage
Title: Lensed Quasar Hosts
Abstract: Gravitational lensing assists in the detection of quasar hosts by amplifying and distorting the host light away from the unresolved quasar core images. We present the results of HST observations of 30 quasar hosts at redshifts 1 < z < 4.5. The hosts are small in size (r_e <~ 6 kpc), and span a range of morphologies consistent with early-types (though smaller in mass) to disky/late-type. The ratio of the black hole mass (MBH, from the virial technique) to the bulge mass (M_bulge, from the stellar luminosity) at 1<z<1.7 is broadly consistent with the local value; while MBH/M_bulge at z>1.7 is a factor of 3--6 higher than the local value. But, depending on the stellar content the ratio may decline at z>4 (if E/S0-like), flatten off to 6--10 times the local value (if Sbc-like), or continue to rise (if Im-like). We infer that galaxy bulge masses must have grown by a factor of 3--6 over the redshift range 3>z>1, and then changed little since z~1. This suggests that the peak epoch of galaxy formation for massive galaxies is above z~1. We also estimate the duty cycle of luminous AGNs at z>1 to be ~1%, or 10^7 yrs, with sizable scatter.
https://export.arxiv.org/pdf/astro-ph/0601391
\begin{frontmatter} \title{Lensed Quasar Hosts\thanksref{label1}} \thanks[label1]{Observations presented in this paper were obtained using the Hubble Space Telescope, operated by the Space Telescope Science Institute under contract to NASA.} \author[STScI]{Chien Y. Peng}, \author[Steward]{Chris D. Impey}, \author[MPIA]{Hans-Walter Rix}, \author[CfA]{Emilio E. Falco} \author[Rutgers] {Charles R. Keeton}, \author[OSU]{Chris S. Kochanek}, \author[CfA]{Joseph Leh\'ar}, \& \author[CfA]{Brian A. McLeod} \address[STScI]{Space Telescope Science Institute, 3700 San Martin Drive, Baltimore, MD 21218} \address[Steward]{Steward Observatory, Univ. of Arizona, 933 N. Cherry Ave., Tucson, AZ 85721} \address[MPIA]{Max-Planck-Institut f\"{u}r Astronomie, K\"onigstuhl 17, Heidelberg, D-69117, Germany} \address[CfA]{Harvard-Smithsonian Center for Astroph., 60 Garden St., Cambridge, MA 02138} \address[Rutgers]{Department of Physics \& Astronomy, Rutgers University, 136 Frelinghuysen Road, Piscataway, NJ 08854} \address[OSU]{Department of Astronomy, The Ohio State University, 4055 McPherson Lab, 140 West 18th Avenue, Columbus, OH 43210} \begin{keyword} quasar \sep host galaxy \sep gravitational lensing \sep evolution \sep supermassive black hole \end{keyword} \end{frontmatter} \section{Introduction} Elsewhere in these proceedings the reader can find summaries of previous work on quasar hosts. We concentrate here on the benefits and challenges of using gravitational lensing as a technique for measuring the host galaxy, especially in regimes where traditional direct imaging methods are difficult -- at high redshift, or when the host is either sub-luminous or compact. The context for this work is the coevolution of galaxies and supermassive black holes. In the local universe, a tight relation is observed between bulge mass (\mbulge) and black hole mass (\mbh) measured with stellar kinematics \citep{Geb00, Fer00}. We now extend the relation to $z \gtrsim 1$ using the virial technique to estimate $M_{BH}$ (e.g. Kaspi et al., 2000; Vestergaard \& Peterson, 2006), and the stellar luminosity to estimate \mbulge\ \citep{Kor95,Mag98}. The existence, slope, and scatter of a \mbh/\mbulge\ relation at high redshift can be used to analyze the relative growth rates of galaxies and their (presumably ubiquitous) central engines. The CfA-Arizona Space Telescope Lens Survey (CASTLES) is a project to image all known, multiply-imaged, quasars in a homogeneous set of optical and near infrared passbands using the {\it Hubble Space Telescope} ({\it HST}). With the number of lens systems now near 100, data are in hand for 80 targets. The observations are shallow, 1-2 orbits per filter, but the excellent surface brightness sensitivity of the {\it HST} leads to a host detection in most cases. The overall CASTLES project is described by \citet{Fal01}; early results in the lensed host search are given by \citet{Rix01} in the same conference proceedings; and detections and models of the hosts of several individual quasars have been published \citep{Imp98,Koc00,Kee00}. Gravitational lensing is a large and growing field of astrophysics so only the rudiments can be given here; a number of excellent book and reviews are available for the full formalism and diverse applications \citep{Bla92,Sch92,Cla02}. About one in 500 quasars has a sight-line passing close enough to the central potential of a massive galaxy for multiple image formation. It has taken surveys of tens of thousands of radio and optically selected quasars to yield the sample of $\sim$100 objects (see the CASTLES web site\footnote{http://www.cfa.harvard.edu/castles} and that of Li\'ege group\footnote{http://vela.astro.ulg.ac.be/themes/extragal/gravlens}). Even though adaptive optics techniques from the ground are improving, stable point spread functions, obtained with {\it HST}, are still essential for reliable modeling. In gravitational lensing, the AGN is magnified into multiple images, but remains unresolved, whereas the extended light from the host galaxy maps into arcs or Einstein rings (ER). A lens model is needed to extract the full information content of the lensed host light. In principle, a typical ER having a radius $\sim$ 1 arcsec means that {\it HST} imaging potentially yields 50-100 resolution elements in the host galaxy in the deep images from the survey. \begin {figure} \hskip 1.2cm \vskip -0.1in \vbox{ \hbox{ \hskip +1.0in \psfig{file=pg1115.ps,height=3.5truein,angle=0} } } \vskip -0.1in \caption {An example of the lens modeling technique. (a) Original NIC2 image of PG~1115+080 ($z_{\mbox QSO} = 1.72$). (b) The host galaxy Einstein ring, after removing the best fit lensing galaxy and quasar point sources. (c) The best fitting residuals. (d) The parametric model of the host galaxy in the source plane.} \label{fig:pg1115} \end {figure} \section{Modeling} The image modeling (Fig.~\ref{fig:pg1115}) uses a custom-built program called LENSFIT (Peng et al. 2006, in prep.), which is based on a methodology that has been well-tested with the GALFIT algorithm \citep{Pen02}. The model for the light profiles of the host and foreground (lens) galaxy uses a S\'ersic model with a concentration index $n$ that is often used to quantify the gross morphology of galaxies (e.g. $n=1$ for late-type, while $n=4$ for early-type). Both the quasar point source and the host galaxy light profile are propagated through the lens model to produce the image distortion, and multiple images. External shear is included to model the tidal influence due to neighbors. All the parameters are simultaneously varied to reduce the $\chi^2$ on a pixel-by-pixel basis. The models are often very robust in well resolved systems ($\theta \ge 1$ arcsec), due to the spatial separation between the host and the lens, and their different shapes. \begin {figure} \hskip 1.2cm \vskip -0.1in \vbox{ \hbox{ \hskip +1.0in \psfig{file=radio.ps,height=3.3truein,angle=0} } } \vskip -0.1in \caption {The (deprojected) host galaxy luminosity with redshift. Open points: radio loud quasar hosts. Solid points: radio quiet or unknown (most likely radio quiet). The stellar evolution tracks correspond to $z_{form} = 5$, all normalized to $L^*$ by $z=0$. The E/S0 and Sb/c tracks have an initial burst of stars, followed by star formation rates that produce colors of E/S0 and Sb/c type galaxies by $z=0$. The ``No Evolution'' model corresponds to a redshifted spectrum of a $z=0$ E/S0 galaxy. } \label{fig:radio} \end {figure} \section{Results} Here we present a summary of our findings, which are detailed along with a description of our analysis techniques elsewhere (Peng et al. 2006, in prep.). We select on lensing geometry size ($\theta \gtrsim 0.7$ arcsec), which does not a priori bias the intrinsic AGN luminosity selection, or the host luminosities. Therefore, we expect the sample of the AGNs to be randomly drawn from the AGN luminosity function, where the lower limits are determined by various lensing search programs. The heterogeneity of different surveys, however, will not affect our primary conclusions about the relationships between \mbh\ and \mbulge, since both quantities are measured in the {\it same} objects. \subsection {General Properties} Figure \ref{fig:radio} shows the $H$-band host galaxy luminosities (lensing distortion removed) versus redshift. Overall, the host luminosities from CASTLES appear to agree well with non-lensing studies \citep{Kuk01,Rid01}. The host luminosities range from 1 to 20 times $L^*_V$ today, while the AGNs are 0 to 3 magnitudes brighter than the host in restframe $B$ to $V$ band. Despite their brightnesses, the host galaxies appear to be fairly small in size (typical $r_e \lesssim 6$ kpc) for their central AGNs. As we shall see later, when coupled with information about their \mbh, both the host luminosities and sizes lead to the conclusion that the bulges may be undermassive compared to present-day normal galaxies. Lastly, the S\'ersic index values suggest that while a number of quasar hosts at $z\gtrsim 1.5$ have steep central concentrations consistent with the presence of a bulge, many (30\% -- 50\%) also have low S\'ersic values ($n \le 2$) more analogous to later-type galaxies today. Even those galaxies with high S\'ersic indices may not qualify as bona fide -- fully formed and passively evolving -- ellipticals, given their small sizes and black hole masses. {\it Radio-loud (RLQ) vs. Radio-quiet (RQQ) hosts}\ \ \ \ \ Quantifying differences between RLQ and RQQ hosts has historically been controversial. Figure~\ref{fig:radio} shows that there is not a clear difference in the host luminosities between RLQ (open) and RQQ (closed) AGNs in the lensing sample. The diverse selection criteria from disjointed surveys are, however, hard to quantify. It is worth to keep in mind, however, that in a study by \citet{Kuk01} RLQs were drawn from the rare and extreme radio-loud sources which may require atypically large BHs to produce. Consequently, the finding of luminous hosts in those RLQs may reflect a correlation between \mbh\ and \mbulge\ at high redshifts. The issue of radio correlation with host properties remains unsettled. {\it AGN Duty Cycle}\ \ \ \ \ We can estimate the duty cycle of nuclear activity for each object with a host detection. A rough estimate of the duty cycle is: $D \sim\Phi_Q(L_{QSO},z)/\Phi_G(L_{Gal},z)$, where $\Phi_Q$ and $\Phi_Q$ are the luminosity functions of quasars and galaxies, respectively, appropriate to a given redshift. At $z\gtrsim 1$, the median duty cycle is 1\%, or $10^7$ years, with a sizable scatter. \subsection {Black Hole vs. Bulge Evolution} Based on quasar and host luminosities we can study the \mbh\ vs. bulge properties at $z\gtrsim 1$ (see Peng et al. 2005 for details), where \mbh's are obtained using the virial technique \citep {Kas00,Ves06}. {\it $1.7 \lesssim z\lesssim 4.5$}\ \ \ \ \ Fig.~\ref{fig:highz} shows the \mbh\ vs. the restframe $R$-band bulge luminosity (\lr) for the host galaxies at $1.7 \lesssim z \lesssim 4.5$. Determining \lr\ requires $K$-corrections, computed using an Sbc SED; the dashed lines shows the small systematic effect of using an E/S0 (right) or Im (left) SED. It is clear that a correlation between \mbh\ and bulge luminosity was already present at a lookback time of 10--12 Gyr. Remarkably, the high-redshift hosts appear to lie on the {\em same} relation as $z=0$ normal galaxies, implying that the high-$z$ hosts are {\em undermassive} in comparison. To explain why, Fig.\ 3b shows the host luminosities after we account for passive evolution of $dM_R/dz = -0.8$ mag; specifically, we hold $M_{BH}$ fixed and shift the color points in Fig.\ 3a to the left. Now the $z\gtrsim 1.7$ hosts are displaced from the local host relation (solid line) by a factor of 3--6 in luminosity, which translates into a mass deficit of a factor of 3--6 in the quasar hosts compared to local galaxies with the same \mbh. {\it $1\lesssim z \lesssim 1.7$}\ \ \ \ \ In contrast, by $z\approx 1$, Fig.~\ref{fig:lowz} shows that the mass deficit of the hosts is mostly reduced (to within a factor of unity in mass), after accounting solely for passive evolution. Thus, massive bulges that correspond to luminous E/S0 galaxies today may have been nearly assembled by $z\approx 1$. By requiring that the hosts evolve onto the local \mbh\ vs. \lr\ of Figs.~\ref{fig:highz} and \ref{fig:lowz}, we can illustrate a growth in the \mbh/\mbulge\ ratio with redshift relative to today, shown in Figure~\ref{fig:growth}. We find that the ratio of \mbh/\mbulge\ increases roughly to a factor of 10 higher than today (assuming an Sbc-type SED) out at $z=4$. This conclusion depends somewhat on the assumption of the SED and evolutionary history: the ratio at earlier times would be higher than shown for a bluer SED than Sb/c, or for a faster fading rate than passive evolution. \begin {figure} \hskip 1.2cm \vskip -0.1in \vbox{ \hbox{ \hskip +0.4in \psfig{file=bh-bulge-highz.ps,height=4.5truein,angle=-90} } } \vskip -0.1in \caption {The relationship of the black hole mass, \mbh, vs. bulge absolute luminosity (\lr, bottom axis; \mr, top axis), at low $z$ (solid round points) and $z\gtrsim 1.7$ (open points). Solid lines: fit to $z\approx0$ solid points. All open points assume a modern-day Sbc-type SED for $K$-correction and their average is represented by dotted lines. Dashed lines illustrate assumptions of bluer(Im, left)/redder(E, right) SEDs. Open circle: gravitationally lensed quasar hosts. Open triangles: \citet{Rid01}. Open squares: \citet{Kuk01}. Vertical line in points: a possible lower limit in \mbh\ due to AGNs being broad absorption line QSOs. Criss-crossed points: potential problem with lens identification, host detection, radio-loud quasar, or narrow line AGN. Panel ({\it a}): The observational data. Panel ({\it b}): The same data in {\it a}, but the open points are shifted horizontally by assuming that the hosts evolve {\it passively} with $z_f=5$ by $d$\mr/$dz$ = $-0.8$ mag. See also \citet{Pen05} for details. } \label{fig:highz} \end {figure} \begin {figure} \hskip 1.2cm \vskip -0.1in \vbox{ \hbox{ \hskip +0.4in \psfig{file=bh-bulge-lowz.ps,height=4.5truein,angle=-90} } } \vskip -0.1in \caption {The same diagram as Figure \ref{fig:highz}, except for redshift of $1 \lesssim z \lesssim 1.7$ quasar hosts. The vertical line in the square points indicates that the AGN either has a strong narrow Mg~{\sc ii} line component or is strongly absorbed in the wings, causing a potentially low \mbh\ estimate. The dotted line is displaced from the solid line, representing the local \mbh-\lr\ relation, by $-0.5$ (Fig. {\it a}) and $+0.5$ (Fig. {\it b}) magnitude. Note the very slight bias between the non-lensed and lensed datasets, which might be explained by the difference in the median redshift of $\left<z\right>_{\mbox{med}} = 1.45$ for the lens sample and $\left<z\right>_{\mbox{med}} = 0.94$ for the non-lenses. However, three of the non-lensed data points may also have a lower limit on the \mbh\ estimate, as noted above.} \label {fig:lowz} \end {figure} \begin {figure} \hskip 1.2cm \vskip -0.1in \vbox{ \hbox{ \hskip +0.12in \psfig{file=growth1.ps,height=2.4truein,angle=0} \hskip +0.16in \psfig{file=growth2.ps,height=2.4truein,angle=0} } } \vskip -0.1in \caption {The growth of the \mbh/\mstar\ ratio as a function of ({\it a}) redshift and ({\it b}) age of the universe in Gyrs. Circles are gravitational lens data points, while triangles are from direct imaging of hosts using {\it HST} NICMOS $H$-band \citep{Rid01,Kuk01}. Point styles are the same as Fig.~\ref{fig:highz}. The \mbh/\mstar\ ratio appears to rise quickly beyond $z\approx 1$ and may slow, and perhaps flatten, to a factor of $6-10\times$ local value by $z\approx 3$. A fading rate of $dM_R/dz=-0.8$ is assumed here (passive evolution since $z_{form} = 5$). } \label{fig:growth} \end {figure} \section{Conclusions} Detailed modeling of 30 well-observed systems from a total sample of 80 lensed quasars has provided new insights into the properties of host galaxies at $1 < z < 4.5$. About half have S\'ersic model fits indicative of early type galaxies. However, combined with their small sizes of $r_e < 6$ kpc, luminosities, and \mbh, it appears that luminous, fully-formed, ellipticals are in a minority as hosts of luminous quasars at $z\gtrsim 2$. No difference is seen between the luminosities of radio-loud and radio-quiet quasars in the sample, with a caveat on sample selection. Even at $z \gtrsim 2$, the host galaxies follow nearly the same relationship between \mbh\ and luminosity as at low redshifts, but the bulges must gain in mass by a factor of 3-6 between $1\lesssim z \lesssim 4.5$. However, by $z\approx1$, the mass deficit is mostly gone. Thus massive bulges at $z\approx 1$ may be consistent with being passively evolving, or may still grow by at most a factor of 1. Our estimate of the AGN duty cycle is $\approx 1\%$, or $10^7$ years. Ongoing work includes using color information to constrain the host star formation histories, obtaining sub-millimeter data to measure star formation rate, and characterizing detailed host morphology with deeper {\it HST} imaging.
Title: Constraints on the coupled quintessence from cosmic microwave background anisotropy and matter power spectrum
Abstract: We discuss the evolution of linear perturbations in a quintessence model in which the scalar field is non-minimally coupled to cold dark matter. We consider the effects of this coupling on both cosmic microwave background temperature anisotropies and matter perturbations. Due to the modification of the scale of cold dark matter as $\rho_{c} = \rho_{c}^{(0)} a^{-3 + \xi}$, we can shift the turnover in the matter power spectrum even without changing the present energy densities of matter and radiation. This can be used to constrain the strength of the coupling. We find that the phenomenology of this model is consistent with current observations up to the coupling power $n_{c} \leq 0.01$ while adopting the current parameters measured by WMAP. Upcoming cosmic microwave background observations continuing to focus on resolving the higher peaks may put strong constraints on the strength of the coupling.
https://export.arxiv.org/pdf/astro-ph/0601333
\baselineskip=16pt \begin{titlepage} \rightline{astro-ph/0601333} \rightline{January 2006} \begin{center} \vspace{0.5cm} \large {\bf Constraints on the coupled quintessence from cosmic microwave background anisotropy and matter power spectrum} \vspace*{5mm} \normalsize {\bf Seokcheon Lee$^{\,1}$}, {\bf Guo-Chin Liu,$^{\,2}$ and {\bf Kin-Wang Ng$^{\,1,2}$}} \smallskip \medskip $^1${\it Institute of Physics,\\ Academia Sinica, Taipei, Taiwan 11529, R.O.C.} $^2${\it Institute of Astronomy and Astrophysics,\\ Academia Sinica, Taipei, Taiwan 11529, R.O.C.} \smallskip \end{center} \vskip0.6in \centerline{\large\bf Abstract} We discuss the evolution of linear perturbations in a quintessence model in which the scalar field is non-minimally coupled to cold dark matter. We consider the effects of this coupling on both cosmic microwave background temperature anisotropies and matter perturbations. Due to the modification of the scale of cold dark matter as $\rho_{c} = \rho_{c}^{(0)} a^{-3 + \xi}$, we can shift the turnover in the matter power spectrum even without changing the present energy densities of matter and radiation. This can be used to constrain the strength of the coupling. We find that the phenomenology of this model is consistent with current observations up to the coupling power $n_{c} \leq 0.01$ while adopting the current parameters measured by WMAP. Upcoming cosmic microwave background observations continuing to focus on resolving the higher peaks may put strong constraints on the strength of the coupling. \vspace*{2mm} \end{titlepage} \section{Introduction} \setcounter{equation}{0} Analysis of the Hubble diagram of high redshift Type Ia supernovae (SNe Ia) has discovered that the expansion of the Universe is currently accelerating \cite{SCP}. In addition, combining measurements of the acoustic peaks in the angular power spectrum of the cosmic microwave background (CMB) anisotropy which indicate the flatness of the Universe \cite{CMB} and the matter power spectrum of large scale structure (LSS) which is inferred from galaxy redshift surveys like the Sloan Digital Sky Survey (SDSS) \cite{SDSS} and the $2$-degree Field Galaxy Redshift Survey ($2$dFGRS) \cite{2dFGRS} has confirmed that a component with negative pressure (dark energy) should be added to the matter component to make up the critical density today. The cosmological constant and/or a quintessence field are the most commonly accepted candidates for dark energy. The latter is a dynamical scalar field leading to a time dependent equation of state, $\omega_{\phi}$. Also, this scalar field has a fluctuating, inhomogeneous component in order to conserve the equivalence principle corresponding to the response of the new component to the inhomogeneities in the surrounding cosmological fluid \cite{CDS}. Several new observational effects produced by the existence of quintessence are imprinted in the CMB anisotropies and the matter power spectrum when compared to models with the cosmological constant. The locations of the acoustic peaks in the CMB angular power spectrum are shifted due to their dependence on the amount of dark energy today and at last scattering as well as $\omega_{\phi}$ \cite{Doran, SLee}. Usually the Universe is dominated by the quintessence at late times ($z \sim {\cal O}(1)$), when the gravitational potential associated with the density perturbations is changed due to a time dependent $\omega_{\phi}$. This enhances the CMB anisotropies at large angular scales by the integrated Sachs-Wolfe effect (ISW) \cite{ISW}. Thus the amplitudes of both CMB angular and matter power spectra decrease at large scale for the fixed Cosmic Background Explorer Satellite (COBE) normalization compared to those in the cold dark matter model with a cosmological constant ($\Lambda$CDM)~\cite{Dodelson1}. The possibility that a scalar field at early cosmological times follows an attractor-type solution and tracks the evolution of the visible matter-energy density has been explored \cite{Ratra}. This may help alleviate the severe fine-tuning associated with the cosmological constant problem. However, this still cannot explain the reason why the dark energy and the dark matter have comparable energy densities at present. Recently models considering the coupling of quintessence to dark matter have been investigated as a possible solution for this late time coincidence problem \cite{coupQ}. However, several of these models using a simple coupling can be ruled out by observational constraints \cite{coupQ1}. These non-minimally coupled quintessence models have several different observational effects compared to the minimally coupled models. One of the most important effects is a different scaling of the cold dark matter (CDM) compared to that of CDM of the minimally coupled case. Since CDM scales as $\rho_{c} = \rho_{c}^{(0)} a^{-3 + \xi}$ where $\xi < 0$, there will be more CDM energy density at early epoch compared to the case with $\xi = 0$ ({\it i.e.} $n _c =0$). As the coupling is increased, the locations of the acoustic peaks are also shifted to smaller scales. However, the amplitudes of the odd-number peaks decrease due to the decrease of the baryon density at early time and the increasing ISW effect as the coupling scaled with the COBE normalization. The amplitudes of the even-number peaks increase due to the increase of the CDM energy density. Even though the ratios of the height of the first peak to those of higher peaks between different couplings are quite similar to one another, these are quite different from the $\Lambda$CDM model \cite{KMT}. The location of the turnover in the matter power spectrum corresponds to the scale that entered the Hubble radius when the universe became matter-dominated ($a_{eq}$). Thus the coupling might shift the turnover scale. This paper is organized as follows. In the next section we show the basic equations of linear perturbations of the coupled quintessence model. In Sec. 3, we derive the formal entropy perturbation due to multiple fluids. We also consider the possibility of isocurvature perturbations of the quintessence. We check the coupling effects on CMB and matter power spectrum in Sec. 4. The effect of coupling on the metric perturbation is considered in Sec. 5. Our conclusion is in the last section. \section{Linear perturbations} \setcounter{equation}{0} We will consider the perturbation effect of the quintessence model which is coupled to the CDM. First we start from the metric in the conformal Newtonian (longitudinal) gauge, which is restricted only for the scalar mode of the metric perturbations \cite{Perturb}. The line element is given by \be ds^2 = a^2(\eta) \Biggl[ -\Bigl(1 + 2 \Psi(\eta, \vec{x}) \Bigr) d\eta^2 + \Bigl(1 - 2 \Phi(\eta,\vec{x}) \Bigr) dx^i dx_i \Biggr], \label{CNG} \ee where $\eta$ is the conformal time, $\Psi$ is the amplitude of perturbation in the lapse function, and $\Phi$ is the amplitude of perturbation of a unit spatial volume. We will consider only a flat universe case. If the coupling is derived by a Brans-Dicke Lagrangian, the radiation is decoupled from the dark energy \cite{Amendola0}. Due to the strong constraint on the coupling to the baryons from the local gravity experiments such as radar time-delay measurements, we will assume that the baryons are decoupled from quintessence \cite{Damour1}. Also from the violation of the weak equivalence principle, even though it is in a still way that is locally unobservable, we can have the species-dependent couplings \cite{Damour2}. So we will make an assumption that the scalar field $\phi$ is coupled only to CDM by means of a general function $\exp[B_{c}(\phi)]$ and there is no coupling to the baryons or the radiation \cite{Coupled}. We can write the general equation including this interaction as \be S = - \int d^4x \sqrt{-g} \Biggl\{ \frac{\bar{M}^2}{2} [\partial^{\mu} \phi \partial_{\mu} \phi - R] + V(\phi) - {\cal L}_{c} - {\cal L}_{r} - {\cal L}_{b} \Biggr\}, \label{lagrangian} \ee where $\bar{M} = 1/ \sqrt{8\pi G}$ is the reduced Planck mass and ${\cal L}_{i}$s denote respectively CDM, radiation, and baryons. If we regard the matter as a gas of pointlike particles with masses $m_{c}$ and paths $x_{c}^{\nu}(t)$, then we can write ${\cal L}_{c}$ as \be {\cal L}_{c} = - \frac{m_{c}}{\sqrt{-g}} \delta(\vec{x} - \vec{x}_{c}(t)) (- g_{\mu\nu} \dot{x}_{c}^{\mu} \dot{x}_{c}^{\nu})^{1/2}, \label{Lm} \ee where $m_{c} = e^{B_{c}(\phi)} m^{*}_{c}$ ($m_{c}^{*}$ being a bare mass of the CDM) ~\cite{Peebles,LOP}. Each fluid element has an energy-momentum tensor $T^{\mu}_{(\beta) \nu}$ where $\beta$ includes all of the species. The total energy-momentum tensor is covariantly conserved, however the energy-momentum transfer between CDM and quintessence is written as \ba \sum_{\beta} \nabla_{\mu} T^{\mu}_{(\beta) \nu} &=& 0, \label{dTbeta} \\ \nabla_{\mu} T^{\mu}_{(\gamma) \nu} &=& 0, \label{dTgamma} \\ \nabla_{\mu} T^{\mu}_{(d) \nu} &=& Q_{(d) \nu}, \label{dTd} \ea where $\gamma$ denotes baryons or radiation, $d$ denotes CDM or quintessence, and $Q_{(d) \nu}$ is the energy-momentum transfer vector which shows the energy transfer to the $d$-fluid \cite{Kodama}. This transfer vector is constrained as \be \sum_{d} Q_{(d) \nu} = 0. \label{sumQ} \ee From the above action (\ref{lagrangian}) we have the following equation which will give the constraint equation for the interaction between the quintessence and the CDM: \be \bar{M}^2 \Box \phi - \frac{\partial V}{\partial \phi} - \frac{\partial B_{c}}{\partial \phi} {\cal L}_{c} = 0. \label{boxphi} \ee From this we can find the unperturbed part of the field equation, \be \phi'' + 2 {\cal H} \phi' + \frac{a^2}{\bar{M}^2} \frac{\partial V(\phi)}{\partial \phi} = - \frac{a^2}{\bar{M}^2} \frac{\partial B_{c}(\phi)}{\partial \phi} \rho_{c}, \label{phieq} \ee where the prime means $d/d\eta$, ${\cal H} = a'/a$, and $\rho_{c}$ means the energy density of CDM. As we mentioned before the energy-momentum of each species may not be conserved due to the scalar field coupling even though the total energy momentum does conserve. If we use this fact, then we can rewrite Eqs.~(\ref{dTbeta}), (\ref{dTgamma}), and (\ref{dTd}) as \ba \rho_{\tot}' &=& - 3 {\cal H} (\rho_{\tot} + p_{\tot}), \label{rhotot'} \\ \rho_{\gamma}' &=& - 3 {\cal H} (\rho_{\gamma} + p_{\gamma}), \label{rhogamma'} \\ \rho_{c}' &=& - 3 {\cal H} (\rho_{c} + p_{c}) + B_{c,\phi} \phi' \rho_{c} \equiv -3 {\cal H}(\rho_{c} + p_{c})(1 - {\cal B}_{c}), \label{rhoc'} \\ \rho_{\phi}' &=& -3 {\cal H}(\rho_{\phi} + p_{\phi}) - B_{c,\phi} \phi' \rho_{c} \equiv -3 {\cal H}(\rho_{\phi} + p_{\phi})(1 - {\cal B}_{\phi}), \label{rhophi'} \ea where $B_{c,\phi} = \partial B_{c}/ \partial \phi$. Henceafter we will adopt the potential and the coupling as given in Ref.~\cite{LOP}: \ba V(\phi) &=& V_{0} \exp \Bigl( \frac{\lambda\phi ^2 }{2} \Bigr), \label{V} \\ \exp [B_{c}(\phi)] &=& \Biggl(\frac{b_c+V(\phi)/V_0}{1+b_c}\Biggr)^{n_c}, ~~~ {\rm with}~~ b_c+1>0, \label{BF} \ea where $V_0$, $\lambda$, $b_c$, and $n_c$ are constant parameters. Here we will consider only $b_{c} =0$ case. \begin{center} \end{center} From these we have the evolution of the background quantities as shown in the first panel of Fig.~\ref{fig:Qnxi}. The parameters we use in this figure for the present energy density contrasts of the quintessence and the CDM are $\Omega_{\phi}^{(0)} = 0.76$ and $\Omega_{c}^{(0)} = 0.20$, respectively. To be compatible with observational data, the energy density of the quintessence must be subdominant during the big-bang nucleosynthesis (BBN) $\Omega_{\phi}^{\rm{BBN}}(a \sim 10^{-10}) < 0.2$ at $T \sim 1\,{\rm MeV}$~\cite{Ratra}, and the energy density at present must be comparable to the preferred range for dark energy, $0.60 \leq \Omega_{\phi}^{(0)}(a=0) \leq 0.85$ ~\cite{CMB}. The stronger bound on the energy density of the quintessence during BBN, $\Omega_{\phi}^{\rm{BBN}}(a \sim 10^{-10}) < 0.045$~\cite{Bean}, is also well satisfied in this model. We note that the equation of state (EOS) of the quintessence $\omega_{\phi} \equiv p_\phi/\rho_\phi \approx \omega_r = 1/3$ during the radiation dominated era as in other tracker solutions. Instead of falling from the tracking value of $1/3$ towards $-1$, as in the uncoupled case, $\omega_{\phi}$ increases towards higher values ($\leq +1$), before ultimately dropping to $-1$ at later times. This is due to the fact that when the matter density increases, the effect on the effective potential $V_{eff}(\phi)= V(\phi) + \rho_{c}$ given in Eq. (\ref{phieq}) drives the faster evolution of $\phi$. From Eq.~(\ref{rhoc'}), we find that the scaling of the CDM energy density differs from the usual $a^{-3}$ due to the coupling : \ba \rho_{c}(a) &=& \rho_{c}^{(0)} a^{-3} \Bigl( \exp[B_{c}(\phi(x)) - B_{c}(\phi(0))] \Bigr) \equiv \rho_{c}^{(0)} a^{-3 + \xi}, \label{rhoca} \\ \xi \cdot x &=& B_{c}(\phi(x)) - B_{c}(\phi(0)), \label{xi} \ea where we set the present scale factor as one ($a^{(0)} =1$), $x = \ln a$, and $\xi$ is the deviation of the CDM redshift as a result of the coupling to a scalar field. This is shown in the second panel of Fig.~\ref{fig:Qnxi}. As the coupling is increased, the magnitude of $\xi$ is increased. The $\xi$ depends also on the scale factor. Consequently, the location of the turnover in the matter power spectrum will be shifted due to the coupling as well (see below). We can find the time-component of the source term in the unperturbed background ($\bar{Q}_{(d) 0}$) from Eqs.~(\ref{dTd}), (\ref{sumQ}), (\ref{rhoc'}), and (\ref{rhophi'}) which is given by \be \bar{Q}_{(c) 0} = - \bar{Q}_{(\phi) 0} = - B_{c,\bar{\phi}} \bar{\phi}' \bar{\rho}_{c}. \label{Q0} \ee Due to the conservation of the total energy-momentum (\ref{dTbeta}), we can find the constraint equation of ${\cal B}_{c}$ and ${\cal B}_{\phi}$, \be (\rho_{c} + p_{c}) {\cal B}_{c} + (\rho_{\phi} + p_{\phi}) {\cal B}_{\phi} = 0. \label{bconstrain} \ee % To consider the perturbation we can decompose the scalar field as \be \phi(\eta, \vec{x}) = \bar{\phi}(\eta) + \delta \phi (\eta, \vec{x}) \label{pphi}, \ee where $\bar{\phi}$ is the unperturbed part and $\delta \phi$ is the perturbed part of the scalar field. We will express the perturbed parts of each quantities by means of Fourier expansions. So the above perturbed scalar field will be expressed as \be \delta \phi (\eta, \vec{x}) = \sum_{k} \delta \phi_{k}(\eta) e^{i\vec{k} \cdot \vec{x}}, \label{Fourier1} \ee where \be \delta \phi_{k}(\eta) = \frac{1}{V} \int \delta \phi(\eta, \vec{x}) e^{-i \vec{k} \cdot \vec{x}} d^3{\vec x}. \label{Fourier2} \ee Therefore, the energy-momentum tensor of the scalar field can be decomposed into the unperturbed part and the perturbed one. The background energy-momentum tensors are \ba ^{(0)}T^{0}_{(\phi)0} &=& - \Bigl( \frac{\bar{M}^2}{2 a^2} \bar{\phi'}^{2} + V(\bar{\phi}) \Bigr) \equiv - \bar{\rho}_{\phi}, \label{0rhophi} \\ ^{(0)}T^{i}_{(\phi)j} &=& \Bigl(\frac{\bar{M}^2}{2 a^2} \bar{\phi'}^{2} - V(\bar{\phi})\Bigr) \delta^{i}_{j} \equiv \bar{p}_{\phi} \delta^{i}_{j} \equiv \omega_{\phi} \bar{\rho}_{\phi} \delta^{i}_{j}, \label{0pphi} \ea and the first-order perturbed parts are \ba \delta T^{0}_{(\phi)0} &=& \frac{1}{a^2} \Bigl( \bar{M}^2 \bar{\phi'}^{2} \Psi - \bar{M}^2 \bar{\phi}' \delta \phi' - a^2 \frac{\partial V(\bar{\phi})}{\partial \bar{\phi}} \delta \phi \Bigr), \label{deltarhophi} \\ \delta T^{i}_{(\phi)j} &=& \frac{1}{a^2} \Bigl( - \bar{M}^2 \bar{\phi'}^{2} \Psi + \bar{M}^2 \bar{\phi}' \delta \phi' - a^2 \frac{\partial V(\bar{\phi})}{\partial \bar{\phi}} \delta \phi \Bigr) \delta^{i}_{j}, \label{deltapphi} \\ \delta T^{0}_{(\phi)i} &=& -\frac{\bar{M}^2}{a^2} \bar{\phi}' \partial_{i} (\delta \phi). \label{delta0iphi} \ea We can repeat the similar consideration for the other components. We will regard the CDM as a perfect fluid of energy density $\rho_{c}$ and pressure $p_{c}$. To the linear order in the perturbations the energy momentum tensors are given by \ba T^{0}_{(c) 0} &=& -\Biggl( \bar{\rho}_{c} + \frac{\partial B_{c}(\bar{\phi})} {\partial \bar{\phi}} \delta \phi \bar{\rho}_{c} + \delta \rho_{c} \Biggr) \equiv -\Biggl( \bar{\rho}_{c} + \delta \bar{\rho}_{c} \Biggr), \label{rhoc} \\ T^{0}_{(c) i} &=& ( \bar{\rho}_{c} + \bar{p}_{c}) {\it v}_{(c) i} = -T^{i}_{(c) 0}, \label{T0ic} \\ T^{i}_{(c) j} &=& \Biggl( \bar{p}_{c} + \frac{\partial B_{c}(\bar{\phi})} {\partial \bar{\phi}} \delta \phi \bar{p}_{c} + \delta p_{c} \Biggr) \delta^i_j + \Sigma^i_{(c) j} \equiv \Biggl( \bar{p}_{c} + \delta \bar{p}_{c} \Biggr) \delta^i_j + \Sigma^i_{(c) j}, \label{pc} \ea where ${\it v}^{i} = dx^{i}/d\eta$ and $\Sigma^i_{j}$ is an anisotropic shear perturbation. Here we define the perturbed part of the CDM as $\delta \bar{\rho}_{c} = B_{c,\bar{\phi}} \delta \phi \bar{\rho}_{c} + \delta \rho_{c}$, where the first term of the right hand side is due to the coupling of the scalar field to t the CDM~\footnote{We can show this as follows: $\rho_{c} = e^{B_c(\phi)} \rho_{c}^{*} = ( e^{B_c(\bar{\phi})} + B_c(\bar{\phi})_{,\bar{\phi}} \delta \phi e^{B_c(\bar{\phi})}) (\bar{\rho}_{c}^{*} + \delta \rho_{c}^{*}) \sim \bar{\rho}_{c} + B_c(\bar{\phi})_{,\bar{\phi}} \delta \phi \bar{\rho}_{c} + \delta \rho_{c}$, where we means $B_c(\bar{\phi})_{,\bar{\phi}} = \partial B_c(\bar{\phi})/ \partial \bar{\phi}$ and $\rho_{c}^{*}$ is the bare energy density of the CDM.}. For the baryons and the radiation we have the following equations, \ba T^{0}_{(\gamma) 0} &=& -\Bigl( \bar{\rho}_{\gamma} + \delta \rho_{\gamma} \Bigr), \label{rhogamma} \\ T^{0}_{(\gamma) i} &=& \Biggl( \bar{\rho}_{\gamma} + \bar{p}_{\gamma} \Biggr) {\it v}_{(\gamma)i} = - T^{i}_{(\gamma) 0}, \label{T0igamma} \\ T^{i}_{(\gamma) j} &=& \Bigl( \bar{p}_{\gamma} + \delta p_{\gamma} \Bigr) \delta^i_j + \Sigma^i_{(\gamma) j}. \label{pgamma} \ea For a flat Friedmann-Robertson-Walker universe we can collect the unperturbed equations for each species: \ba 3 {\cal H}^2 &=& \frac{a^2}{\bar{M}^2} \Bigl( \bar{\rho}_{r} + \bar{\rho}_{b} + \bar{\rho}_{c} + \bar{\rho}_{\phi} \Bigr) \equiv \frac{a^2}{\bar{M}^2} \sum_{\beta} \bar{\rho}_{\beta} \equiv \frac{a^2}{\bar{M}^2} \bar{\rho}_{cr}, \label{H} \\ {\cal H}' &=& - \frac{a^2}{6 \bar{M}^2} \sum_{\beta} (1 + 3 \omega_{\beta}) \bar{\rho}_{\beta} = - \sum_{\beta} \frac{(1 + 3 \omega_{\beta})}{2} \bar{\Omega}_{\beta} {\cal H}^2, \label{H'} \\ \bar{\rho}_{\gamma}' &=& - 3 {\cal H} (\bar{\rho}_{\gamma} + \bar{p}_{\gamma}) , \label{barrhogamma'} \\ \bar{\rho}_{c}' &=& - 3 {\cal H} (\bar{\rho}_{c} + \bar{p}_{c}) + B_{c,\bar{\phi}} \bar{\phi}' \bar{\rho}_{c} \equiv -3 {\cal H}(\bar{\rho}_{c} + \bar{p}_{c})(1 - \bar{{\cal B}}_{c}), \label{barrhoc'} \\ \bar{\rho}_{\phi}' &=& -3 {\cal H}(\bar{\rho}_{\phi} + \bar{p}_{\phi}) - B_{c,\bar{\phi}} \bar{\phi}' \bar{\rho}_{c} \equiv -3 {\cal H}(\bar{\rho}_{\phi} + \bar{p}_{\phi})(1 - \bar{{\cal B}_{\phi}}). \label{barrhophi'}\ea If we include terms up to the first order of the energy transfer vector (\ref{dTd}), then we can write $Q_{(d) \nu}$ as \ba Q_{(d) 0} &=& - \bar{Q}_{(d)} (1 + \Psi) - \delta Q_{(d)}, \label{Q0p} \\ Q_{(d) i} &=& \Bigl( f_{(d)} + \bar{Q}_{(d)} v_{(d) i} \Bigr)_{,i}. \label{Qip} \ea where $\delta Q_{(d)}$ and $f_{(d)}$ is the energy and momentum transfer of the CDM or quintessence, respectively. These terms should be included in the coupled quintessence models when we use the conformal Newtonian gauge \cite{Kodama}. If we missed this term, then we would have $2 \Psi B_{c,\bar{\phi}}$ instead of $3 \Psi B_{c,\bar{\phi}}$ in the last term of Eq. ~(\ref{deltaphi}). With using Eqs.~(\ref{Q0p}) and (\ref{Qip}), we can find the perturbed part of the energy momentum conservation equations in $k$-space. The perturbed equations for the scalar field and the fluids which can be obtained from Eqs.~(\ref{dTgamma}) and (\ref{dTd}) are given by \ba \delta \phi '' &+& k^2 \delta \phi + 2 {\cal H} \delta \phi' + 2 \Psi \frac{a^2}{\bar{M}^2} V_{,\bar{\phi}} + \frac{a^2}{\bar{M}^2} V_{,\bar{\phi}\bar{\phi}} \delta \phi - (\Psi' + 3 \Phi') \bar{\phi}' \nonumber \\ &=& - \frac{a^2}{\bar{M}^2} \bar{\rho}_{c} \Biggl( B_{c,\bar{\phi}\bar{\phi}} \delta \phi + B_{c,\bar{\phi}}^2 \delta \phi + B_{c,\bar{\phi}} \delta_{c} + 3 \Psi B_{c,\bar{\phi}} \Biggr) \nonumber \\ &=& - \frac{a^2}{\bar{M}^2} \bar{\rho}_{c} \Biggl( B_{c,\bar{\phi}\bar{\phi}} \delta \phi + B_{c,\bar{\phi}} \bar{\delta}_{c} + 3 \Psi B_{c,\bar{\phi}} \Biggr) , \label{deltaphi} \\ \frac{\delta \rho_{c}'}{\bar{\rho}_{c}} &=& -(1 + \omega_{c}) (\theta_{c} - 3 \Phi') - 3 {\cal H} \Bigl(\frac{\delta p_{c}}{\delta \rho_{c}} + 1 \Bigr) \delta_{c} + B_{c,\bar{\phi}} \bar{\phi}' \Bigl( \delta_{c} + \Psi \Bigr), \label{deltarhoc} \\ \frac{\delta \bar{\rho}_{c}'}{\bar{\rho}_{c}} &=& -(1 + \omega_{c}) (\theta_{c} - 3 \Phi') - 3 {\cal H} \Bigl(\frac{\delta \bar{p}_{c}}{\delta \bar{\rho}_{c}} + 1 \Bigr) \bar{\delta}_{c} + B_{c,\bar{\phi} \bar{\phi}} \bar{\phi}' \delta \phi \nonumber \\ &+& B_{c,\bar{\phi}} \Bigl( \bar{\phi}' \bar{\delta}_{c} + \delta \phi' + \bar{\phi}' \Psi \Bigr), \label{deltabarrhoc} \\ \delta_{c}' &=& -(1 + \omega_{c})(\theta_{c} - 3 \Phi') - 3 {\cal H} \Bigl(\frac{\delta p_{c}}{\delta \rho_{c}} - \omega_{c} \Bigr) \delta_{c} + B_{c,\bar{\phi}} \bar{\phi}' \Psi, \label{deltac} \\ \bar{\delta}_{c}' &=& -(1 + \omega_{c})(\theta_{c} - 3 \Phi') - 3 {\cal H} \Bigl(\frac{\delta p_{c}}{\delta \rho_{c}} - \omega_{c} \Bigr) \bar{\delta}_{c} + B_{c,\bar{\phi} \bar{\phi}} \bar{\phi}' \delta \phi + B_{c,\bar{\phi}} \delta \phi' \nonumber \\ &+& B_{c,\bar{\phi}} \bar{\phi}' \Psi, \label{bardeltac} \\ \theta_{c}' &=& - {\cal H} (1 -3 \omega_{c}) \theta_{c} - \frac{\omega_{c}'}{1 + \omega_{c}} \theta_{c} + \frac{\delta p_{c}/\delta \rho_{c}}{ 1 + \omega_{c}} k^2 \delta_{c} - k^2 \sigma_{c} + k^2 \Psi \nonumber \\ &+& \fr{\omega_{c}}{(1+\omega_{c})} k^2 B_{c,\bar{\phi}} \delta \phi - \fr{\omega_{c}}{(1+\omega_{c})} B_{c,\bar{\phi}} \bar{\phi}' \theta_{c} \label{thetac} \\ &=& - {\cal H} (1 -3 \omega_{c}) \theta_{c} - \frac{\omega_{c}'}{1 + \omega_{c}} \theta_{c} + \frac{\delta \bar{p}_{c}/\delta \bar{\rho}_{c}}{ 1 + \omega_{c}} k^2 \bar{\delta}_{c} - k^2 \sigma_{c} + k^2 \Psi \nonumber \\ &-& \fr{\omega_{c}}{(1+\omega_{c})} B_{c,\bar{\phi}} \bar{\phi}' \theta_{c}, \label{barthetac} \ea where we define $\delta_{c} = \delta \rho_{c}/ \bar{\rho}_{c}$, $\bar{\delta}_{c} = \delta \bar{\rho}_{c}/ \bar{\rho}_{c}$, $\theta_c=i \vec{k} \cdot \vec{v}$, $\omega_c$ denotes the equation of state parameter (EOS) of the CDM, and $\sigma_c$ is related to the CDM anisotropic stress perturbation $\Pi_{c}$, by $\sigma_c = 2 \Pi_{c} \, \bar{p_{c}}/3(\bar{\rho}_c + \bar{p}_c)$. We express both $\delta_{c}'$ and $\bar{\delta}_{c}'$ explicitly in order to show that the equations of the energy density perturbation can be different with different definitions. However, it is the coupled energy density which is measured in observations. As such, we will use $\bar{\delta}_{\beta}$ as the energy density contrast of each species. % Every term containing $B_{c}(\phi)$ in the above equations comes from the coupling. If we drop all these terms, then they are obviously identical to the expressions Ref.~\cite{MaB}. If we have more than one ideal gas components, then we have entropy perturbation ($\delta S$) which can be expressed as \be \delta p_{\tot} = \sum_{\beta} \Bigl[ (\partial p_{\beta}/\partial \rho_{\beta})|_{S} \, \delta \rho_{\tot} + (\partial p_{\beta}/ \partial S_{\beta})|_{\rho} \, \delta S_{\tot} \Bigr] \equiv c_{\tot}^2 \delta \rho_{\tot} + p_{\tot} \Gamma_{\tot} , \label{deltaS} \ee where $c_{\tot}^2$ is the overall adiabatic sound speed squared and $\Gamma_{\tot}$ is the total entropy perturbation. We will consider this more carefully in the following section. We can write the perturbed equations for the CDM from the above generic perturbation equations (\ref{bardeltac}) and (\ref{barthetac}). However due to the coupling between the baryons and the photons we also need to consider the Thomson scattering term in the baryon-photon fluid. After including this we have the following equations. \ba \delta_{b}' &=& -\theta_{b} + 3 \Phi', \label{deltab} \\ \theta_{b}' &=& -{\cal H} \theta_{b} + c_{s}^2 k^2 \delta_{b} + k^2 \Psi + \frac{4 \bar{\rho}_{\gamma}}{3 \bar{\rho}_{b}} a n_{e} \sigma_{T} (\theta_{\gamma} - \theta_{b}), \label{thetab} \\ \delta_{c}' &=& -\theta_{c} + 3 \Phi' + B_{c,\bar{\phi}} \bar{\phi}' \Psi, \label{deltac2} \\ \bar{\delta}_{c}' &=& -\theta_{c} + 3 \Phi' + B_{c,\bar{\phi} \bar{\phi}} \bar{\phi}' \delta \phi + B_{c,\bar{\phi}} \delta \phi' + B_{c,\bar{\phi}} \bar{\phi}' \Psi, \label{bardeltac2} \\ \theta_{c}' &=& -{\cal H} \theta_{c} + k^2 \Psi, \label{thetac2} \\ \delta_{r}' &=& - \frac{4}{3} \theta_{r} + 4 \Phi' , \label{deltar} \\ \theta_{r}' &=& k^2 \Bigl( \frac{1}{4} \delta_{r} - \sigma_{r} \Bigr) + k^2 \Psi + a n_{e} \sigma_{T} (\theta_{b} - \theta_{\gamma}), \label{thetar} \ea where $n_{e}$ is the electron number density, $\sigma_{T}$ is the cross section for the Thomson scattering. \section{Entropy perturbation} \setcounter{equation}{0} Let us start from the definition of the total energy density perturbation and that of the total pressure perturbation, \ba \delta \bar{\rho}_{\rm{tot}} &=& \sum_{\beta} \delta \bar{\rho}_{\beta}, \label{deltarhotot} \\ \delta \bar{p}_{\rm{tot}} &=& \sum_{\beta} \delta \bar{p}_{\beta} . \label{deltaptot} \ea For a given $p_{\rm{tot}}(\rho,S)$, the pressure fluctuation can be expressed as \be \delta p_{\tot} \equiv c_{\tot}^2 \delta \rho_{\tot} + p_{\tot} \Gamma_{\rm{int}} + p_{\tot} \Gamma_{\rel}, \label{deltaptot2} \ee where $S$ is an entropy, $c_{\tot}^2$ is the overall adiabatic sound speed squared, $\Gamma_{\rm{int}}$ and $\Gamma_{\rel}$ are the intrinsic and the relative entropy perturbations respectively. We have \ba p_{\tot} \Gamma_{\rm{int}} &=& \sum_{\beta} p_{\beta} \Gamma_{\beta} , \label{Gammaint} \\ p_{\tot} \Gamma_{\rel} &=& \sum_{\beta} (c_{\beta}^2 - c_{\tot}^2) \delta \rho_{\beta}, \label{Gammarel} \\ c_{\tot}^2 &=& \frac{\sum_{\beta} c_{\beta}^2\rho_{\beta}'}{\rho_{\tot}'}, \label{ctot} \ea where $\Gamma_{\rm{int}}$ is the sum of the intrinsic entropy perturbation of each fluid and $\Gamma_{\rel}$ arises from the relative evolution between fluids with different sound speeds. As we mentioned before the energy momentum of each species may not be conserved due to the scalar field coupling even though the total energy momentum does conserve. We can rewrite the equation for the total adiabatic sound speed (\ref{ctot}) as \be c_{\tot}^2 = \sum_{\beta} c_{\beta}^2 (1 -{\cal B}_{\beta}) \frac{\rho_{\beta} + p_{\beta}}{\rho_{\tot} + p_{\tot}}. \label{ctot2} \ee Now we can rewrite the relative entropy perturbation as \be p_{\tot} \Gamma_{\rel} = \frac{1}{2} \sum_{\beta,\alpha} \frac{(\rho_{\beta} + p_{\beta})(\rho_{\alpha} + p_{\alpha})} {\rho_{\tot} + p_{\tot}}(c_{\beta}^2 - c_{\alpha}^2) S_{\beta \alpha} + \sum_{\beta} {\cal B}_{\beta} c_{\beta}^2 (\rho_{\beta} + p_{\beta}) \Delta, \label{Gammarel2} \ee where \ba S_{\beta \alpha} &=& \Delta_{\beta} - \Delta_{\alpha}, \label{S} \\ \Delta_{\beta} &=& \frac{\delta \rho_{\beta}}{\rho_{\beta} + p_{\beta}}, \label{Deltabeta} \\ \Delta &=& \frac{\delta \rho_{\tot}}{\rho_{\tot} + p_{\tot}} , \label{Delta} \ea where $S_{\beta \alpha}$ is the entropy perturbation \cite{Kodama}. Due to the $\Delta$-term the relative entropy perturbation is non-vanishing even without the non-adiabatic perturbation. This is improper, so we redefine the new quantities as \ba \hat{\Delta}_{\beta} &=& \frac{\delta \rho_{\beta}}{(1 - {\cal B}_{\beta})(\rho_{\beta} + p_{\beta})}, \label{hatDeltabeta} \\ \hat{S}_{\beta \alpha} &=& \hat{\Delta}_{\beta} - \hat{\Delta}_{\alpha}. \label{S2} \ea With these quantities we can rewrite Eq.~(\ref{Gammarel2}) as \be p_{\tot} \Gamma_{\rel} = \frac{1}{2} \sum_{\beta,\alpha} \frac{(1 - {\cal B}_{\beta})(1 - {\cal B}_{\alpha})(\rho_{\beta} + p_{\beta})(\rho_{\alpha} + p_{\alpha})} {\rho_{\tot} + p_{\tot}}(c_{\beta}^2 - c_{\alpha}^2) \hat{S}_{\beta \alpha}. \label{Gammarel3} \ee \begin{center} \end{center} \subsection{Isocurvature condition} Due to the out of thermal equilibrium nature of the quintessence, we need to check the isocurvature evolution of the scalar field perturbation \cite{Isocur}. We can analytically show this in the tracking region. First of all the adiabatic sound speed squared $c_{\phi}^2$ of the quintessence can be represented as \be c_{\phi}^2 = \frac{\bar{p}_{\phi}'}{\bar{\rho_{\phi}}'} = 1 + \frac{2 \bar{\phi}' V_{,\bar{\phi}}}{3 {\cal H} (\bar{\rho}_{\phi} + \bar{p}_{\phi})( 1 - {\cal B}_{\phi})} = \omega_{\phi} - \fr{\omega_{\phi}'}{3 {\cal H} ( 1 + \omega_{\phi})(1 - {\cal B}_{\phi})}, \label{cphi} \ee where we have used Eq.~(\ref{rhophi'}). From this equation we can find the second derivative of the potential, \be \frac{a^2}{\bar{M}^2} V_{,\bar{\phi} \bar{\phi}} = \frac{3}{2} {\cal H} c_{\phi}^{2'}(1 - {\cal B}_{\phi}) + \frac{3}{2} {\cal H}^2(c_{\phi}^2 - 1)(1 - {\cal B}_{\phi}) \Bigl[ \frac{{\cal H}'}{{\cal H}^2} - 1 - \frac{3}{2}(c_{\phi}^2 + 1)(1 - {\cal B}_{\phi}) \Bigr] - \frac{3}{2} {\cal H}(c_{\phi}^2 - 1) {\cal B}_{\phi}'. \label{Vdouble} \ee The relation between ${\cal H}^2$ and ${\cal H}'$ can be found from Eq.~(\ref{H'}) : \ba \fr{{\cal H}'}{{\cal H}^2} &=& - \fr{1}{2} ( 1 + 3 \omega_{\tot}), \label{HH} \\ \omega_{\tot} &=& \sum_{\beta} \omega_{\beta} \Omega_{\beta}, \label{omegatot} \ea where $\omega_{\tot}$ is the weighted EOS. As long as we use the background as shown in the pervious model, the tracking region is well established during the radiation dominated era~\cite{LOP}. During this era we can find the following relation, \be c_{\phi}^2 = \frac{\bar{p}_{\phi}'}{\bar{\rho}_{\phi}'} = \omega_{\phi} = \omega_{r}. \label{trackingcphi} \ee This is shown in the first panel of Fig.~\ref{fig:cpBn}. As the coupling is increased we can have the non-monotonic behavior of $c_{\phi}^2$ as shown in the $n_{c} = 10^{-2}$ case. In addition, we can rewrite the coupling term as \be {\cal B}_{\phi} = - \frac{ B_{c,\bar{\phi}} \bar{\phi}' \bar{\rho}_{c}}{3 {\cal H} (\bar{\rho}_{\phi} + \bar{p}_{\phi})} = - \frac{{\cal H}}{\bar{\phi}'} B_{c,\bar{\phi}} \bar{\Omega}_{c} . \label{calBphi} \ee As shown in the second panel of Fig.~\ref{fig:cpBn}, this term is negligible during the radiation dominated epoch. The coupling drives a faster evolution of $\phi$ when matter energy dominates the Universe and the magnitude of ${\cal B}_{\phi}$ depends on the energy density of the CDM. With these facts we can have the approximate expression of Eq.~(\ref{Vdouble}), \be \frac{a^2}{\bar{M}^2} V_{,\bar{\phi} \bar{\phi}} \simeq \frac{3}{2} {\cal H}^2(c_{\phi}^2 - 1) \Bigl[ \frac{{\cal H}'}{{\cal H}^2} - \frac{3}{2}(c_{\phi}^2 + 1) \Bigr] = - \frac{3}{4} {\cal H}^2(c_{\phi}^2 - 1) \Bigl[ 3 \omega_{\tot} + 3 c_{\phi}^2 + 4 \Bigr], \label{Vdouble2} \ee where we have used Eq.~(\ref{HH}) in the second equality. The isocurvature mode in the radiation dominated epoch can be obtained from Eq.~(\ref{deltaphi}) and using the fact that $a \propto \eta$ during the radiation dominated era. To check this we can put $\Phi = 0$ and then we have \be \delta \phi'' + \frac{4}{(3 \omega_{r} + 1) \eta} \delta \phi' + \Bigl[ k^2 - (c_{\phi}^2 - 1) \frac{3}{(3 \omega_{r} + 1)^2 \eta^2} (3 \omega_{r} + 3 c_{\phi}^2 + 4 ) \Bigr] \delta \phi = 0. \label{deltaphi2} \ee If we use the tracking solution (\ref{trackingcphi}), then we can rewrite the above equation as \be \delta \phi'' + \frac{2}{\eta} \delta \phi' + \Bigl[ k^2 + \frac{3}{\eta^2} \Bigr] \delta \phi = 0. \label{deltaphi3} \ee The solutions of this equation are Bessel functions, \be \delta \phi(\eta) = {\rm const} ~ \eta^{-1/2} ~ J_{\pm |i \sqrt{11}/2|} (k \eta) . \label{soldeltaphi1} \ee Thus both solutions decay in time. At the superhorizon scale the $k$-dependent term can be neglected and we obtain the power-law solutions, $\delta \phi \propto \eta^{\nu}$, where the power index $\nu = (-1 \pm i \sqrt{11})/2$. Therefore, any initial nonzero isocurvature fluctuation of the quintessence is damped to zero with time. We will use only the adiabatic perturbations in the following section. \begin{center} \end{center} \section{Effects of coupling} \setcounter{equation}{0} The non-minimally coupled quintessence models have been investigated as a possible solution for the late time coincidence problem \cite{coupQ}. The coupling gives rise to the additional mass and source terms of the evolution equations for CDM and scalar field perturbations. This also affects the perturbation of radiation indirectly through the background bulk ${\cal H}$ and the metric perturbations \cite{Bean2}. The value of the energy density contrast of the CDM ($\Omega_{c}$) is increased in the past when the coupling is increased. \subsection{CMB} \setcounter{equation}{0} The temperature anisotropy measured in a given direction of the sky can be expanded in spherical harmonics as \be \Theta \equiv \frac{\Delta T}{T}(\hat{n})=\sum_{\ell m}a_{\ell m}Y_{\ell m} (\hat{n}), \label{Theta} \ee where $\Theta$ is the temperature brightness function that is the fractional perturbation of the temperature of the photons $T = T_{0}(1 + \Theta)$, $\hat{n}$ is the direction of the photon momentum, and $a_{\ell m}$ are the multipoles. We can also expand the brightness function as a Legendre polynomial, $P_{\ell}$ : \be \Theta = \sum_{\ell} (-i)^{\ell} \, (2 \ell + 1) \, \Theta_{\ell} \, P_{\ell} . \label{Theta2} \ee From the inflationary scenario, the multipoles are Gaussian random variables which satisfy \be \langle a^*_{\ell m} \, a_{\ell' m'}\rangle = C_{\ell} \, \delta_{\ell \ell'} \, \delta_{m m'}. \label{Cl} \ee The angular power spectrum ($C_{\ell}$) contains all the information about the statistical properties of CMB and is defined as \cite{Hu} \be \frac{2 \ell + 1}{4 \pi} C_{\ell} = \frac{1}{2 \pi^2} \int d \eta \frac{dk}{k} \frac{k^3 |\Theta_{\ell}(k,\eta)|^2}{2 \ell +1}. \label{spectrum} \ee In the standard recombination model the acoustic oscillations will be frozen into the CMB. A generalization of the free-streaming equation in a flat universe gives the resulting anisotropies : \be \fr{\Theta_{\ell}(\eta)}{2 \ell + 1} = [\Theta_0 + \Psi](\eta_{ls}) j_{\ell}(k(\eta - \eta_{ls})) + \Theta_{1}(\eta_{ls}) \fr{1}{k} \fr{d}{d \eta} j_{\ell}(k(\eta - \eta_{ls})) + \int_{\eta_{\,ls}}^{\eta} (\Psi' - \Phi') j_{\ell} (k(\eta - \tilde{\eta})) d \tilde{\eta} \label{Spectrum2} \ee where $j_{\ell}$ is the spherical Bessel function and $\eta_{\,ls}$ is the conformal time at last scattering. Photon density perturbation is related to the temperature perturbation in the matter rest frame : \be \delta_{r} = 4 \Theta_0 + 4 {\cal H} \fr{\theta_{\tot}}{k^2}, \label{deltarm} \ee and the gravitational potentials will be given in Eq. (\ref{deltaG00}) - (\ref{deltaGij}) in the following section. We use the fact that in the absence of anisotropic stress, the two scalar potentials $\Psi$ and $\Phi$ defined in the conformal Newtonian gauge (\ref{CNG}) are equal and they coincide with the usual gravitational potential in the Newtonian limit. Now, we investigate the effects of non-minimal coupling of a scalar field to the CDM on the CMB power spectrum. Firstly, the Newtonian potential at late times changes more rapidly as the coupling increases, as shown in Eq.~(\ref{deltaG00}). This leads to an enhanced ISW effect as indicated in the last term of Eq.~(\ref{Spectrum2}). Thus we have a relatively larger $C_{\ell}$ at large scales ({\it i.e.} small $\ell$). Thus, if the CMB power spectrum normalized by COBE, then we will have smaller quadrupole \cite{COBE}. This is shown in the first panel of Fig.~\ref{fig:cl}. One thing that should be emphasized is that we use different parameters for the $\Lambda$CDM and the coupled quintessence models to match the amplitude of the first CMB anisotropy peak. The parameter used for the quintessence model is indicated in Fig.~\ref{fig:cl} ({\it i.e.} $\Omega_{\phi}^{(0)} = 0.76$, $\Omega_{m}^{(0)} = 0.191$, $\Omega_{b}^{(0)} = 0.049$, and $h = 0.7$, where $h$ is the present Hubble parameter in the unit of $100 {\rm km s^{-1} Mpc^{-1}}$). However, these parameters are well inside the $1 \, \sigma$ region given by the WMAP data. We use the WMAP parameters for the $\Lambda$CDM model ({\it i.e.} $\Omega_{\phi}^{(0)} = 0.73$, $\Omega_{m}^{(0)} = 0.23$, $\Omega_{b}^{(0)} = 0.04$, and $h = 0.72$). In both models we use the same spectral index $n_{s} =1$. The heights of the acoustic peaks at small scales ({\it i.e} large $\ell$) can be affected by the following two factors. One is the fact that the scaling of the CDM energy density deviates from that of the baryon energy density. Therefore for the given CDM and baryon energy densities today, the energy density contrast of baryons at decoupling ($\Omega_{b}^{(ls)}$) is getting lower as the coupling is being increased. This suppresses the amplitude of compressional (odd number) peaks while enhancing rarefaction (even number) peaks. The other is that for models normalized by COBE, which approximately fixes the spectrum at $\ell \simeq 10$, the angular amplitude at small scales is suppressed in the coupled quintessence. This is shown in the second panel of Fig.~\ref{fig:cl}. The third peak in this model is smaller than that in the $\Lambda$CDM model. The WMAP data do not show the value of the third peak but quote a compilation of other experiments \cite{Wang}. The ratio of the amplitude between the second and the third peaks is $1.03 \pm 0.02$. In the $\Lambda$CDM model this value is $0.986$ and in our model these values are $1.08$ and $1.11$ for without and with the coupling equal to $n_c=0.01$ respectively. We also show the $n_{c} = 0.02$ case in this figure. With the same parameters this case can be ruled out from the current data. In all cases, we have used the CMBFAST code~\cite{CMBFAST} with the modified Boltzmann equations to compute the CMB power spectrum. Secondly, for $\ell>200$ we can see that the locations of the acoustic peaks are slightly shifted to smaller scales ({\it i.e.} larger $\ell$). This can be explained as follows. The locations of peaks and troughs can be parametrized as \be \ell_{m} = \ell_{A} (m - \varphi_m) = \ell_{A} (m - \bar{\varphi} - \delta \varphi_{m} ), \label{lm} \ee where $\ell_{A}$ is the acoustic scale dependent on the geometry of the Universe, $\bar{\varphi}$ is the overall peak shift, and $\delta \varphi_{m}$ is the relative shift of the $m$-th peak relative to the first one~\cite{Doran}. The overall peak shift, $\bar{\varphi}$ is given by \be \bar{\varphi} \simeq 0.267 \Bigl(\fr{r_{ls}}{0.3} \Bigr)^{0.1}, \label{varphi} \ee where $r_{ls} = \rho_r^{(ls)} / \rho_m^{(ls)}$ is the ratio of the energy densities of radiation to matter at last scattering. The shift is due to both driving effects from the decay of the gravitational potential and contributions from the Doppler shift of the oscillating fluid. The acoustic scale $\ell_{A}$ depends on both the sound horizon $s_{ls}$ at decoupling and the angular diameter distance $D$ to the last scattering surface: \be \ell_{A} = \pi \fr{D}{s_{ls}} = \pi \fr{\eta_{0} - \eta_{\,ls}}{\bar{c}_{s} \eta_{\, ls}} , \label{la} \ee where $\bar{c}_{s}$ is the average sound speed before last scattering : \be \bar{c}_{s} = {\eta}_{\,ls}^{-1} \int_{0}^{\eta_{\,ls}} c_{s} d \eta \hspace{0.2in} {\rm with} \hspace{0.2in} c_s^{-2} = 3 + \fr{9}{4} \fr{\rho_{b}}{\rho_{r}} . \label{cs} \ee Also from the Hubble parameter, \be \Biggl(\frac{d a}{d \eta}\Biggr)^2 = H_{0}^2 \Biggl\{ \Omega_{r}^{(0)} + \Omega_{b}^{(0)} a + \Omega_{m}^{(0)} a^{1+ \xi} + \fr{\rho_{\phi}}{\rho_{cr}^{(0)}} a^{4} \Biggr\}, \label{da1} \ee where $H_{0}$ is the present value of the Hubble parameter, we can find the angular diameter distance, \be D=\eta_{0} - \eta_{ls} = H_{0}^{-1} \int_{a_{ls}}^{a^{(0)}} \fr{da}{\sqrt{\Omega_{r}^{(0)} + \Omega_{b}^{(0)} a + \Omega_{m}^{(0)} a^{1+ \xi} + \rho_{\phi} / \rho_{cr}^{(0)} a^4}}. \label{da2} \ee The sound horizon is not affected by the coupling and the effect of coupling on the angular diameter distance is also quite small in our model. However, the overall shift and the relative shift are affected by the coupling. As the coupling is increased, the value of $r_{ls}$ is decreased. Hence, the locations of the peaks are shifted to the right. However, this shift is quite small. We show this in the second panel of Fig.~\ref{fig:cl}. The heights and the locations of the acoustic peaks in various models are summarized in Table~\ref{tab:1}. There is a significant difference for the heights of the peaks of the second and the third peaks between the models. Thus upcoming observations continuing to focus on resolving the higher peaks may constrain the strength of the coupling. \begin{table}[htb] \begin{center} \caption{Summary of the heights and the locations of the considered models. The heights of peaks, $A_{i}$ have been scaled by a factor of $10^{10}$. $A_{i:j}$ is the ratio of the $i$-th peak height to the $j$-th. The locations of peaks, $\ell_{i}$. \label{tab:1}} \vskip .3cm \begin{tabular}{|c|c|c|c|c|c|c|c|c|} \hline Model & $A_{1}$ & $A_{2}$ & $A_{3}$ & $\ell_{1}$ & $\ell_{2}$ & $\ell_{3}$ & $A_{1:2}$ & $A_{2:3}$ \\ \hline $\Lambda$CDM & $7.26$ & $3.25$ & $3.30$ & $220$ & $536$ & $812$ & $2.23$& $0.986$ \\ \hline $n_{c} = 0$ & $7.04$ & $3.23$ & $3.00$ & $217$ & $532$ & $811$ & $2.18$& $1.08$ \\ \hline $n_{c} = 10^{-2}$ & $7.02$ & $3.36$ & $3.03$ & $218$ & $535$ & $818$ & $2.09$& $1.11$ \\ \hline \end{tabular} \end{center} \end{table} \subsection{Matter power spectrum} \begin{center} \end{center} The effects of quintessence on structure formation in several other models have been investigated \cite{WS, DSW}. Structure as a function of physical scale size is usually described in terms of a power spectrum : \be P(k) = \langle|\delta_{k}|^2\rangle = A k^{n_{s}} T^{2}(k), \label{Pk} \ee where $A$ is the COBE normalization, $n_s$ is a power index, and $T(k)$ is the transfer function. The coupling of quintessence to the CDM can change the shape of matter power spectrum because the location of the turnover corresponds to the scale that entered the Hubble radius when the Universe became matter-dominated. This shift on the scale of matter and radiation equality is indicated in the second panel of Fig.~\ref{fig:Qnxi}: \be a_{eq} \simeq \fr{\rho_{r}^{(0)}}{\rho_{c}^{(0)}} \exp [B_{c}(\phi_{0}) - B_{c}(\phi_{eq})], \label{aeq} \ee where $\rho_{r}^{(0)}$ and $\rho_{c}^{(0)}$ are the present values of the energy densities of radiation and CDM respectively, and the approximation comes from the fact that the present energy density of CDM is bigger than that of baryons ($\rho_{c}^{(0)} > \rho_{b}^{(0)}$). Increasing the coupling shifts the epoch of matter-radiation equality further from the present, thereby moving the turnover in the power spectrum to smaller scale. If we define $k_{eq}$ as the wavenumber of the mode which enters the horizon at radiation-matter equality, then we will obtain \be k_{eq} = \fr{2 \pi}{\eta_{eq}}. \label{keq} \ee However, from the previous subsection, we notice that the value of $\eta_{eq}$ remains unchanged for different couplings and this degeneracy is indicated in Fig.~\ref{fig:mp}. We have used different parameters for the $\Lambda$CDM and quintessence models; the matter power spectra look different between the models. There is a slight suppression in the quintessence models. Note that a bias factor could resolve the discrepancy and perhaps a parameter fitting may also help this. However, the detailed parameter fitting is beyond the scope of this paper. We can write the equation of the matter fluctuation in the synchronous gauge during the matter dominated epoch : \be \bar{\delta}_{c}'' + {\cal H} \bar{\delta}_{c}' - \fr{3}{2} {\cal H}^2 \fr{( \delta \bar{\rho}_{c} + \delta \bar{p}_{c} )}{\bar{\rho}_{cr}} - \fr{(a F(\phi))'}{a} = 0, \label{deltacsyn} \ee where the coupling term $F(\phi)$ is given by \be F(\phi) = B_{c,\bar{\phi} \bar{\phi}} \bar{\phi}' \delta \phi + B_{c,\bar{\phi}} \delta \phi' = n_c \lambda (\bar{\phi} \delta \phi)'. \label{coupc} \ee This equation can be rewritten by the structure growth exponent $f$, which is defined as \be f(a) = \fr{d \ln \bar{\delta}_{c}(a)}{d x}. \label{fa} \ee Hence, we have the following equation for $f(a)$ which is identical to Eq. (\ref{deltacsyn}): \be \fr{d f}{d x} + f^2 + \Bigl( 2 + \fr{d \ln ({\cal H} / a )}{d x} \Bigr) f - \fr{3}{2} \Omega_{c} = \fr{d \Bigl[a F(\phi) \Bigr]}{d x} \fr{1}{a {\cal H} \bar{\delta}_{c}}, \label{faeq} \ee If we use Eq.~(\ref{rhoca}), then we can find that $a \propto \eta^{2/(1-\xi)}$ and ${\cal H} = 2/(1- \xi) \eta^{-1}$ for a matter dominated universe. As we can see from the above Eq.~(\ref{faeq}) that the coupling term is negligible because $\phi$ varies much slower than ${\cal H}$. Thus we can ignore the last term in Eq.~(\ref{deltacsyn}). From this we can rewrite the above equation in a matter dominated era : \be \bar{\delta}_{c}'' + \fr{2}{(1-\xi) \eta} \bar{\delta}_{c}' - \fr{6}{(1-\xi)^2 \eta^2} \bar{\delta}_{c} = 0. \label{deltacsyn2} \ee This equation has two solutions, \be \bar{\delta}_{c}^{\pm} = c_{\pm} \eta^{\nu_{\pm}}, \label{deltacsol} \ee where $c_{\pm}$ are arbitrary constants and \be \nu_{\pm} = \fr{-(1+\xi) \pm \sqrt{24 + (1 + \xi)^2}}{2(1 - \xi)}. \label{nupm} \ee $\bar{\delta}_{c}^{+}$ indicates a growing mode, which is only relevant today because $|\bar{\delta}_{c}^{+}|$ is small at early time and we can ignore the decaying mode $\bar{\delta}_{c}^{-}$. If we remove the coupling effect in this solution, then we can recover the well known solution $\delta_{c}^{+} = \eta^{2/3}$. This effect is shown in Fig.~\ref{fig:mp}. As the coupling is increased, we have little more matter at early times and this increases the height of the matter power spectrum. Again this effect is tiny and we can hardly see the difference between various couplings. \section{Metric perturbation} \setcounter{equation}{0} In addition to these, the perturbed equations of the metric can be obtained from the Einstein equations: \ba k^2 \Phi + 3 {\cal H} \Bigl( \Phi' + {\cal H} \Psi \Bigr) &=& \frac{3}{2} \frac{{\cal H}^2}{\bar{\rho}_{tot}} \delta T^0_{(tot)0}, \label{deltaG00} \\ \Phi'' + {\cal H} ( \Psi' + 2 \Phi') + \Bigl(2{\cal H}' + {\cal H}^2 \Bigr) \Psi + \frac{k^2}{3} ( \Phi - \Psi) &=& \frac{1}{2} \frac{{\cal H}^2}{\bar{\rho}_{tot} } \delta T^i_{(tot)i}, \label{deltaGii} \\ k^2 \Bigl( \Phi' + {\cal H} \Psi \Bigr) &=& \frac{3}{2} {\cal H}^2 \sum_{\beta} (1 + \omega_{\beta}) \bar{\Omega}_{\beta} \theta_{\beta} , \label{deltaG0i} \\ k^2 ( \Phi - \Psi) &=& \frac{9}{2} {\cal H}^2 \sum_{\beta} (1 + \omega_{\beta}) \bar{\Omega}_{\beta} \sigma_{\beta} , \label{deltaGij} \ea where we have used the unperturbed equations~(\ref{H}) and (\ref{H'}) . From Eqs.~(\ref{deltaG00}) and (\ref{deltaGii}), we can find the metric perturbation equation, \ba \Phi'' + {\cal H}(\Psi' + 5 \Phi') + 2 ( {\cal H}' + 2{\cal H}^2) \Psi + \frac{k^2}{3}(4\Phi - \Psi) \nonumber \\ = -\frac{a^2}{2\bar{M}^2} [ (1 - c_{tot}^2) \delta \rho_{tot} - p_{tot} \Gamma_{int} - p_{tot} \Gamma_{rel} ]. \label{metricper} \ea If we use Eq.~(\ref{deltaG00}), this equation can be rewritten as \ba \Phi'' + {\cal H} \Bigl[\Psi' + (2 + 3 c_{tot}^2) \Phi' \Bigr] + \Bigl[ 2 {\cal H}' + {\cal H}^2(1 + 3c_{tot}^2) \Bigr] \Psi + \frac{k^2}{3} \Bigl[ (1 + 3 c_{tot}^2) \Phi - \Psi \Bigr] \nonumber \\ = \frac{a^2}{2\bar{M}^2} \Bigl( p_{tot} \Gamma_{int} + p_{tot} \Gamma_{rel} \Bigr) \equiv \frac{a^2}{2\bar{M}^2} p_{tot} \Gamma_{tot}, \label{metricper2} \ea where we define $\Gamma_{tot} = \Gamma_{int} + \Gamma_{rel}$ in the last equality. The last term in this equation comes from the coupling of the scalar field. So even if we start from the adiabatic condition ( $p_{tot} \Gamma_{tot} =0$), we can analytically solve the above equation for the specific case. Let us first put $\Psi = \Phi$ (no anisotropic stress) and $k^2 c_{tot}^2 \Phi = 0$ (consider the superhorizon scale), then Eq.~(\ref{metricper2}) is simplified as \be \Phi'' + 3 {\cal H} (1 + c_{tot}^2) \Phi' + \Bigl[ 2 {\cal H}' + {\cal H}^2(1 + 3c_{tot}^2) \Bigr] \Phi = \frac{a^2}{2\bar{M}^2} p_{tot} \Gamma_{tot}. \label{metricper3} \ee If you use the background equations, this equation can be rewritten as \be \Phi'' - \Bigl( \ln[\rho_{tot} + p_{tot}] \Bigr)' \Phi' + \Bigl( \ln[\rho_{tot}/(\rho_{tot} + p_{tot})] \Bigr)' {\cal H} \Phi = \frac{3 {\cal H}^2}{2} \frac{p_{tot}}{\rho_{tot}} \Gamma_{tot}. \label{metricper4} \ee If we consider the adiabatic condition ($p_{tot} \Gamma_{tot} = 0$), then the above equation is identical to the equation in Bardeen's article~\cite{Bardeen}.As such, the metric perturbation can be rewritten as the curvature perturbation~\cite{Lyth}, \be \zeta = \frac{({\cal H}^{-1} \Phi' + \Phi)}{(1 + \omega_{tot})} + \frac{3}{2} \Phi. \label{zeta} \ee With this we can express Eq.~(\ref{metricper3}) in the adiabatic case as \be \zeta' = 0. \label{etaprime} \ee This result looks the same to the minimally coupled case and there seems to have no difference from the non-minimally coupled case. Nevertheless, the coupling information is absorbed in both $c_{tot}^2$ and the perturbation equation of each species. \section{Conclusions} \setcounter{equation}{0} We have analyzed the linear perturbations of the cosmological models for the scalar field with its self-interaction potential, $V(\phi) = V_{0} \exp(\lambda \phi^2 /2)$, and its coupling to the CDM, $\exp[B_{c}(\phi)] = \exp(n_{c} \lambda \phi^2 /2)$. The evolution of the non-perturbed background scalar field $\bar{\phi}$ occurred in the tracking regime throughout the radiation dominated epoch. The full analysis of the non-perturbed and perturbed equation of each species in the conformal Newtonian gauge has been done including the proper energy-momentum transfer vector due to the coupling, $Q_{(d) \nu}$. The Boltzmann equations have been modified as to account for the coupling between a scalar field and the CDM. We have seen that the energy-momentum of each species may not be conserved as a result of the coupling of the scalar field to the CDM. Thus we have redefined the concepts of entropy perturbations. We have shown that the isocurvature perturbation of the quintessence has been damped to zero with time in the tracking regime during the radiation epoch. Thus we have constrained our considerations on the adiabatic perturbation. We have considered the CMB anisotropy spectrum and the matter power spectrum for the non-minimally coupled models. Additional mass and source terms in the Boltzmann equations induced by the coupling give the rapid changes of the Newtonian potential $\Phi$ and enhance the ISW effect in the CMB power spectrum. The modification of the evolution of the CDM, $\rho_{c} = \rho_{c}^{(0)} a^{-3 + \xi}$, changes the energy density contrast of the CDM at early epoch. We have adopted the current cosmological parameters measured by WMAP within $1 \sigma$ level. With the COBE normalization and the WMAP data we have found the constraint of the coupling $n_{c} \leq 0.01$. The locations and the heights of the CMB anisotropy peaks have been changed due to the coupling. Especially, there is a significant difference for the heights of the second and the third peaks among the models. Thus upcoming observations continuing to focus on resolving the higher peaks may constrain the strength of the coupling. The suppression of the amplitudes of the matter power spectra could be lifted by a bias factor. However, a detailed fitting is beyond the scope of this paper. The turnover scale of the matter power spectrum may be used to constrain the strength of the coupling $n_{c}$. Finally, we have investigated the metric perturbations including the coupling between the scalar field and the CDM. There is no difference to the curvature perturbation $\zeta$ for the different couplings in the adiabatic case. However, the effects of the coupling have been absorbed in the Boltzmann equations already. \section{Acknowledgements} \setcounter{equation}{0} This work was supported in part by the National Science Council, Taiwan, ROC under the Grant NSC94-2112-M-001-024 (K.W.N.).
Title: Stellar and Gas properties of High HI Mass-to-Light Ratio Galaxies in the Local Universe
Abstract: We present a multi-wavelength study (BVRI band photometry and HI line interferometry) of nine late-type galaxies selected from the HIPASS Bright Galaxy Catalog on the basis of apparently high HI mass-to-light ratios (3 M_sun/L_sun < M_HI/L_B < 27 M_sun/L_sun). We found that most of the original estimates for M_HI/L_B based on available photographic magnitudes in the literature were too high, and conclude that genuine high HI mass-to-light ratio (>5 M_sun/L_sun) galaxies are rare in the Local Universe. Extreme high M_HI/L_B galaxies like ESO215-G?009 appear to have formed only the minimum number of stars necessary to maintain the stability of their HI disks, and could possibly be used to constrain galaxy formation models. They may to have been forming stars at a low, constant rate over their lifetimes. The best examples all have highly extended HI disks, are spatially isolated, and have normal baryonic content for their total masses but are deficent in stars. This suggests that high M_HI/L_B galaxies are not lacking the baryons to create stars, but are underluminous as they lack either the internal or external stimulation for more extensive star formation.
https://export.arxiv.org/pdf/astro-ph/0601321
\title{Stellar and Gas properties of High \hi{} Mass-to-Light Ratio Galaxies in the Local Universe} \author{Bradley E. Warren\altaffilmark{1} and Helmut Jerjen} \affil{Research School of Astronomy and Astrophysics, Australian National University, Mount Stromlo Observatory, Cotter Road, Weston ACT 2611, Australia} \email{bewarren@mso.anu.edu.au, jerjen@mso.anu.edu.au\\} \and \author{B\"arbel S. Koribalski} \affil{Australia Telescope National Facility, CSIRO, PO Box 76, Epping NSW 1710, Australia} \email{Baerbel.Koribalski@csiro.au} \altaffiltext{1}{Affiliated with the Australia Telescope National Facility, CSIRO.} \keywords{galaxies: irregular --- galaxies: dwarf --- galaxies: evolution --- galaxies: photometry --- galaxies: ISM --- galaxies: kinematics and dynamics --- galaxies: individual (\esoq{}, ESO\,572-G009, ESO\,428-G033, \mcg{}, ESO\,473-G024, IC\,4212, ESO\,348-G009, ESO\,121-G020, ESO\,505-G007, \atg{})} \section{Introduction} \label{sec:intro} The number of low mass dark matter halos predicted by models of a CDM dominated Universe far exceeds the quantity of observed dwarf galaxies, typically by several orders of magnitude \citep*[see][]{kau93,moo99,kly99}. In this context we consider a galaxy to be a dark matter halo that contains baryons. Consequently the slope of dark matter mass functions generally rises much more steeply than observed galaxy luminosity functions (\citealp*{tre02,hil03,bla03}; but see also \citealp{bla04}). While there are some physical processes that could help narrow this discrepancy \citep*{kly99,sha04}, the theoretical low mass halo frequency is not reduced enough to reconcile them with observations. So we are left with the conclusions that either the current most favored cosmological models significantly over-estimate the number of low mass dark matter halos present in the Local Universe, or the observations have failed to find the vast majority of low mass galaxies to date. If the latter were true, then it is important to look at why dwarf galaxies could be missed and how we could detect them. Two reasons why galaxies might not have been found yet in optical surveys are that they could exhibit low stellar densities ($<$ 1\Msun{}\,pc$^{-2}$) or most of their baryons are in invisible form. In the extreme case, they do not contain any baryonic matter at all and are in fact ``empty'' dark matter halos. These possibilities could be dark matter halos in which the star formation from accreted gas has been halted, suppressed, never began, or there were simply no baryons to form stars to begin with. Such objects lacking stars might be considered as being ``dark galaxies,'' which could be easily missed in surveys biased towards optical wavelengths \citep{dis76}. However, dark galaxies have yet to be found in the blind \hi{} surveys such as HIPASS \citep{kor04,doy05}. If they do exist, large numbers of low mass dark galaxies could naturally steepen the mass functions from observations, which are mostly derived from optical or near-infrared galaxy luminosity functions without consideration of other baryonic matter. Recent large scale observations of neutral hydrogen gas (\hi{}) are now allowing mass functions to be derived based on non-stellar properties. \citet{zwa03} produced one of the most extensive \hi{} mass functions based on a catalogue of the 1000 \hi{}-brightest galaxies in the Southern hemisphere and found a similar low mass end slope to other observational studies, again in contradiction with $\Lambda$CDM models. Galaxies that have failed to convert most of their primordial gas into stars, and thus retained a high proportion of \hi{}, may well provide a partial solution. While not entirely ``dark'' these galaxies are hard to detect optically, but may be detectable through 21cm line observations. The \hi{} mass-to-light ratio (\mlr{}, which compares the \hi{} mass to the {\em B} band luminosity) of these objects could be significantly higher than the typical ratios measured for late-type galaxies, so that they would be in a lower mass bin in a luminosity function than they would be for a baryonic mass function. If high \mlr{} galaxies existed in significant numbers then they could help correct the discrepancy in two ways, by including more previously unknown galaxies, and by shifting known galaxies to higher mass bins than they would be placed in with purely optical results. An example of an extreme \hi{} mass-to-light ratio object, which could be described as a ``dim'' galaxy, is the nearby dwarf irregular \esoq{} with \mlr{} = $22 \pm 4$\mls{} \citep*[][ hereafter \pI{}]{war04}. This faint low surface brightness dwarf irregular was found to be spatially isolated (1.7~Mpc from the nearest neighbor), with a low current star formation rate ($\la 2.5 \times 10^{-3}$\Msun{}\,yr$^{-1}$). It has an extended regularly rotating \hi{} disk, which can be traced out to over six times the Holmberg radius of the optical galaxy, making it one of the most extended \hi{} envelopes relative to the optical extent. \pI{} included an analysis of the \hi{} gas surface density of \esoq{}. The azimuthally averaged surface density at all radii was below the critical gas surface density needed for large scale star formation as defined by the \citet{too64} stability criteria \citep{ken89,mar01}. It was proposed in \citet{ver02} that a large fraction of low mass halos may form \citeauthor{too64} stable gas disks and become ``dark'' galaxies, possibly 95\% of objects with halo masses of $\la 10^{10}$\Msun{}. If so \esoq{} would be just the tip of the iceberg. If we can find more galaxies similar to \esoq{}, where the gas density after gravitational collapse is too low for efficient star formation, it may go some way to explaining the discrepancy between the dark matter halo mass function and the observed galaxy luminosity function. To continue our study of the stellar and gas properties of galaxies with high \mlr{} we have selected a sample of nine galaxies from the BGC in the approximate range 3\mls{} $<$ \mlr{} $<$ 27\mls{}. In this paper, \S~\ref{sec:sample} looks at what was previously known about the sample galaxies. \S~\ref{sec:obs} summarizes our 21cm and optical observations. \S\S~\ref{sec:radio} and \ref{sec:optp} present the results of the \hi{} line observations and optical photometry, respectively. \S~\ref{sec:dis-indi} compares the properties of individual galaxies. \S~\ref{sec:dis} contains the discussion of these results and the possible implications, while \S~\ref{sec:conc} gives our conclusions. \section{Galaxy Selection} \label{sec:sample} \begin{deluxetable}{lccccccccc} \tabletypesize{\scriptsize} \tablecaption{Summary of Previously Measured Galaxy Properties. \label{tab:prop}} \tablewidth{0pt} \tablehead{\colhead{Name} & \colhead{Center} & \colhead{Galactic} & \multicolumn{3}{c}{Bright Galaxy Catalog} & \colhead{LEDA} & \colhead{SFD98} & \multicolumn{2}{c}{BGC + LEDA} \\ \colhead{HIPASS Name} & \colhead{$\alpha$(J2000.0)} & \colhead{$l$} & \colhead{\vsys{}} & \colhead{$D$} & \colhead{\FHI{}} & \colhead{\mB{}} & \colhead{\AB{}} & \colhead{\MB{}} & \colhead{\mlr{}} \\ & \colhead{$\delta$(J2000.0)} & \colhead{$b$} & \colhead{(\kkms{})} & \colhead{(Mpc)} & \colhead{(\jjks{})} & \colhead{(mag)} & \colhead{(mag)} & \colhead{(mag)} & \colhead{(\mmls{})} \\ \colhead{(1)} & \colhead{(2)} & \colhead{(3)} & \colhead{(4)} & \colhead{(5)} & \colhead{(6)} & \colhead{(7)} & \colhead{(8)} & \colhead{(9)} & \colhead{(10)}} \startdata \mcg{} & $00^{\rm h}\,19^{\rm m}\,11^{\rm s}$ & 62\fdg6 & $669 \pm 3$ & 9.5 & $16.0 \pm 2.5$ & $15.8 \pm 0.2$ & $0.08 \pm 0.01$ & $-14.1 \pm 0.2$ & $4.8 \pm 1.6$ \\ HIPASS\,J0019--22 & --22\degr\,40\arcmin\,14\arcsec{} & --81\fdg4 \\ \\ ESO\,473-G024 & $00^{\rm h}\,31^{\rm m}\,23^{\rm s}$ & 75\fdg7 & $540 \pm 4$ & 7.6 & $7.2 \pm 1.8$ & $16.2 \pm 0.2$ & $0.08 \pm 0.01$ & $-13.3 \pm 0.2$ & $3.0 \pm 1.3$ \\ HIPASS\,J0031--22 & --22\degr\,46\arcmin\,02\arcsec{} & --83\fdg7 \\ \\ ESO\,121-G020 & $06^{\rm h}\,15^{\rm m}\,53^{\rm s}$ & 266\fdg5 & $577 \pm 5$ & 4.1 & $14.1 \pm 2.9$ & $17.0 \pm 0.4$ & $0.17 \pm 0.03$ & $-11.3 \pm 0.4$ & $11 \pm 6$ \\ HIPASS\,J0615--57 & --57\degr\,43\arcmin\,24\arcsec{} & --27\fdg3 \\ \\ ESO\,428-G033 & $07^{\rm h}\,25^{\rm m}\,49^{\rm s}$ & 244\fdg2 & $1727 \pm 4$ & 19.5 & $12.8 \pm 2.7$ & $17.9 \pm 0.5$ & $1.10 \pm 0.18$ & $-14.7 \pm 0.5$ & $10 \pm 7$ \\ HIPASS\,J0725--30B & --30\degr\,55\arcmin\,05\arcsec{} & --6\fdg9 \\ \\ \esoq{} & $10^{\rm h}\,57^{\rm m}\,30^{\rm s}$ & 284\fdg1 & $598 \pm 2$ & 4.2 & $104.4 \pm 11.5$ & $16.4 \pm 0.4$ & $0.95 \pm 0.15$ & $-12.6 \pm 0.4$ & $24 \pm 12$ \\ HIPASS\,J1057--48 & --48\degr\,11\arcmin\,02\arcsec{} & 10\fdg5 \\ \\ ESO\,572-G009 & $11^{\rm h}\,53^{\rm m}\,23^{\rm s}$ & 284\fdg1 & $1745 \pm 3$ & 20.4 & $8.5 \pm 1.9$ & $17.4 \pm 0.2$ & $0.16 \pm 0.03$ & $-14.3 \pm 0.2$ & $10 \pm 4$ \\ HIPASS\,J1153--18 & --18\degr\,09\arcmin\,59\arcsec{} & 42\fdg6 \\ \\ ESO\,505-G007 & $12^{\rm h}\,03^{\rm m}\,30^{\rm s}$ & 289\fdg5 & $1785 \pm 4$ & 20.8 & $20.5 \pm 3.2$ & $17.7 \pm 0.2$ & $0.36 \pm 0.06$ & $-14.2 \pm 0.2$ & $27 \pm 9$ \\ HIPASS\,J1203--25 & --25\degr\,28\arcmin\,22\arcsec{} & 36\fdg1 \\ \\ IC\,4212 & $13^{\rm h}\,12^{\rm m}\,09^{\rm s}$ & 312\fdg0 & $1484 \pm 2$ & 18.1 & $47.5 \pm 4.6$ & $16.4 \pm 1.1$ & $0.19 \pm 0.03$ & $-15.0 \pm 1.1$ & $23 ^{ +45}_{ -16}$ \\ HIPASS\,J1311--06 & --06\degr\,58\arcmin\,31\arcsec{} & 55\fdg5 \\ \\ ESO\,348-G009 & $23^{\rm h}\,49^{\rm m}\,23^{\rm s}$ & 349\fdg8 & $648 \pm 4$ & 8.4 & $13.4 \pm 2.2$ & $16.7 \pm 0.7$ & $0.06 \pm 0.01$ & $-13.0 \pm 0.7$ & $9 \pm 7$ \\ HIPASS\,J2349--37 & --37\degr\,46\arcmin\,23\arcsec{} & --73\fdg2 \\ \enddata \end{deluxetable} The HIPASS Bright Galaxy Catalog \citep[][ hereafter BGC]{kor04} lists the 1000 \hi{}-brightest extragalactic sources (by \hi{} peak flux density) in the Southern hemisphere ($v_{\rm sys} < 8\,000$\kms{}). Photometric and structural parameters for the BGC's {\it optical} counterparts were obtained in 2002 from the Lyon-Meudon Extragalactic Database \citep[LEDA,][ and references therein, now moved to HyperLEDA]{pat97} to study the statistical properties and various scaling relations of these galaxies. First estimates of \hi{} mass-to-light ratios were obtained for 789 BGC galaxies that had mean apparent {\em B} band photographic magnitudes from LEDA, using the equation: \begin{equation} \frac{{\cal M}_{\rm HI}}{L_{\rm B}} = 1.5 \times 10^{-7} F_{\rm HI}~ 10^{0.4(m_{\rm B}-A_{\rm B})}~~\frac{{\cal M}_{\sun}}{L_{\sun,{\rm B}}} , \label{eqn:mlr} \end{equation} where \MHI{} is the \hi{} mass in solar units, \LB{} is the {\em B} band luminosity in solar units, \FHI{} is the integrated \hi{} flux density in \jjks{}, \mB{} is the apparent {\em B} magnitude, and \AB{} is the {\em B} band Galactic extinction. Extinction correction from the host galaxy is not included for reasons explained in \S~\ref{sec:dis-extinct}. Fig.~\ref{fig:mlmbt} shows the log(\mlr{}) distribution for these galaxies as a function of their absolute {\em B} magnitude: \begin{equation} M_{\rm B,0} = m_{\rm B} - A_{\rm B} - 5\log(D) - 25 ~~{\rm mag} , \label{eqn:abmag} \end{equation} where the galaxy distances, $D$ (Mpc), were calculated from the Local Group velocities given in the BGC. Throughout this paper we adopt H$_0$ = 75\kms{}\,Mpc$^{-1}$. The relation between these two quantities seems to suggest that many of the low luminosity galaxies listed in the BGC have high \mlr{}, up to 27\mls{}, well above typical values for late-type galaxies of less than 1\mls{} \citep[median \mlr{} of 0.78\mls{} in][ for type Sm/Im galaxies]{rob94}. The nine filled circles mark the positions of the galaxies subject to our detailed follow up observations, including \esoq{} (\pI{}). The galaxies were chosen for various reasons, mostly because of a high estimated \mlr{}, but also for reasons of unusual morphology, and after initial ATCA follow up observations showed some unexpected results (see \S~\ref{sec:structure}). Previously measured properties of those nine galaxies are summarised in Table~\ref{tab:prop}. The columns are as follows: (1) commonly used galaxy name and HIPASS source name; (2) J2000.0 Right Ascension and Declination as given in the RC3 \citep{dev91}; (3) Galactic longitude and latitude; (4) \hi{} systemic velocity as given in the BGC; (5) galaxy distance derived from the velocity relative to the barycentre of the Local Group as given in the BGC; (6) total integrated \hi{} flux density as given in the BGC; (7) apparent {\em B} band photographic magnitude as listed in LEDA; (8) \citet*[ hereafter SFD98]{sch98} Galactic dust extinction in the {\em B} band; (9) absolute {\em B} band magnitude calculated as in eqn.~\ref{eqn:abmag} using the LEDA magnitude and SFD98 extinction; (10) preliminary estimate of \hi{} mass-to-light ratio calculated using eqn.~\ref{eqn:mlr} from the LEDA, BGC and SFD98 data. \section{Observations} \label{sec:obs} Each galaxy was observed in two different wavelength regimes. Optical CCD photometry was obtained with the Australian National University (ANU) 2.3-meter Telescope at the Siding Spring Observatory. \hi{} (21cm) line data were obtained with the Australia Telescope Compact Array (ATCA). \begin{deluxetable}{lcccccccc} \tabletypesize{\scriptsize} \tablecaption{Summary of Observations for each Galaxy. \label{tab:obs}} \tablewidth{0pt} \tablehead{ \colhead{Name} & \multicolumn{3}{c}{Optical} & \colhead{\phantom{0000}} & \multicolumn{4}{c}{Radio} \\ \colhead{} & \colhead{Band} & \colhead{Exposure Time} & \colhead{Seeing} & & \colhead{Arrays} & \colhead{Time On Source} & \colhead{Central Freq.} & \colhead{Phase Cal.} \\ \colhead{} & \colhead{} & \colhead{(seconds)} & \colhead{(arcsec)} & & \colhead{} & \colhead{(hours)} & \colhead{(MHz)} & \colhead{} \\ \colhead{(1)} & \colhead{(2)} & \colhead{(3)} & \colhead{(4)} & & \colhead{(5)} & \colhead{(6)} & \colhead{(7)} & \colhead{(8)} } \startdata \mcg{} & {\em B} & 3000 ($10\times300$) & 2\farcs2 & & H75B & $\sim1.5$ & 1417 & PKS\,0023--263 \\ & {\em V} & 2400 ($8\times300$) & 2\farcs0 & & H168B & $\sim8.6$ & 1417 & PKS\,0023--263 \\ & {\em R} & 1800 ($6\times300$) & 1\farcs9 \\ & {\em I} & 1800 ($6\times300$) & 2\farcs2 \\ \\ ESO\,473-G024 & {\em B} & 3000 ($10\times300$) & 2\farcs5 & & H75B & $\sim1.5$ & 1417 & PKS\,0023--263 \\ & {\em V} & 2400 ($8\times300$) & 3\farcs3 & & H168B & $\sim1.0$ & 1417 & PKS\,0023--263 \\ & {\em R} & 1800 ($6\times300$) & 3\farcs3 \\ & {\em I} & 1800 ($6\times300$) & 2\farcs0 \\ \\ ESO\,121-G020 & {\em B} & 3000 ($10\times300$) & 2\farcs1 & & 750D & $\sim10.5$ & 1417 & PKS\,0407--658 \\ & & & & & 1.5B & $\sim10.9$ & 1417 & PKS\,0407-658 \\ & {\em R} & 1800 ($6\times300$) & 1\farcs9 & & EW352 & $\sim2.4$ & 1416 & PKS\,0537--441 \\ \\ ESO\,428-G033 & {\em B} & 3000 ($5\times600$) & 2\farcs2 & & 750D & $\sim10.6$ & 1412 & PKS\,0614--349 \\ & {\em V} & 2400 ($4\times600$) & 2\farcs0 & & 1.5B & $\sim10.3$ & 1412 & PKS\,0614--349 \\ & {\em R} & 1800 ($3\times600$) & 1\farcs9 & & EW352 & $\sim1.1$ & 1414 & PKS\,0614--349 \\ & {\em I} & 1800 ($3\times600$) & 3\farcs0 & & EW367B & $\sim0.8$ & 1414 & PKS\,0614--349 \\ \\ \esoq{} & {\em B} & 3000 ($10\times300$) & 1\farcs9 & & EW352 & $\sim11.6$ & 1417 & PKS\,1215--457 \\ & {\em V} & 2400 ($8\times300$) & 1\farcs9 & & 750A & $\sim10.6$ & 1417 & PKS\,1215--457 \\ & {\em R} & 1800 ($6\times300$) & 1\farcs8 & & 6A & $\sim11.5$ & 1417 & PKS\,1215--457 \\ & {\em I} & 1800 ($6\times300$) & 2\farcs0 \\ \\ ESO\,572-G009 & {\em B} & 1800 ($3\times600$) & 1\farcs9 & & H75B & $\sim1.0$ & 1412 & PKS\,1127--145 \\ & {\em V} & 2400 ($8\times300$) & 2\farcs7 \\ & {\em R} & 1800 ($6\times300$) & 2\farcs2 \\ \\ ESO\,505-G007 & {\em B} & 3000 ($10\times300$) & 2\farcs0 & & H75B & $\sim1.0$ & 1412 & PKS\,1127--145 \\ & {\em V} & 1800 ($6\times300$) & 2\farcs2 & & H168B & $\sim1.9$ & 1412 & PKS\,1127--145 \\ & {\em R} & 1800 ($6\times300$) & 2\farcs2 & & H75B & $\sim9.2$ & 1412 & PKS\,1151--348 \\ \\ IC\,4212 & {\em B} & 3000 ($5\times600$) & 2\farcs9 & & H75B & $\sim8.8$ & 1413 & PKS\,1308--220 \\ & {\em V} & 2400 ($4\times600$) & 2\farcs3 \\ & {\em R} & 1800 ($3\times600$) & 1\farcs9 \\ \\ ESO\,348-G009 & {\em B} & 3000 ($10\times300$) & 1\farcs7 & & 750D & $\sim10.6$ & 1417 & PKS\,0008--421 \\ & {\em V} & 2400 ($8\times300$) & 1\farcs4 & & 1.5B & $\sim9.8$ & 1417 & PKS\,0008--421 \\ & {\em R} & 1800 ($6\times300$) & 1\farcs5 & & EW352 & $\sim1.7$ & 1417 & PKS\,0008--421 \\ & {\em I} & 1800 ($6\times300$) & 1\farcs4 \\ \enddata \end{deluxetable} \subsection{Radio Observations} \label{sec:obs-rad} ATCA \hi{} line observations of the selected galaxies were carried out between June 2002 and June 2003. The galaxies ESO\,121-G020, ESO\,428-G033, and ESO\,348-G009 ($\delta < -30$\degr{}) were observed for $2 \times \sim12$ hours in different East-West arrays, while \esoq{} was observed for $3 \times \sim$12 hours. For the other five galaxies ($\delta > -30$\degr{}) we used the compact hybrid arrays that include antennas on the Northern spur, resulting in a rather large synthesized beams. The galaxies \mcg{}, ESO\,505-G007, and IC\,4212 were observed for $\sim$10 hours, while ESO\,572-G009 and ESO\,473-G024 were only observed in snapshot mode ($\sim1 - 2$ hours taken over a 12 hour period). \hi{} snapshot observations that were initially taken for the other galaxies (except \esoq{} and IC\,4212) were added to the other observations. Details of the \hi{} observations for each galaxy are given in Table~\ref{tab:obs}. The columns are as follows: (5) ATCA configurations used; (6) approximate time on source for each array; (7) central observing frequency; and (8) phase calibrator. We used a bandwidth of 8 MHz, divided into 512 channels, resulting in a channel width of 3.3\kms{}. The velocity resolution of the \hi{} data is $\sim$4\kms{}. The primary calibrator for all observations was PKS\,1934--638. Data reduction and analysis were performed with the {\sc MIRIAD} package using standard procedures, with further analysis using {\sc AIPS}, {\sc GIPSY}, and {\sc KARMA}. Channels affected by Galactic \hi{} emission were discarded where appropriate. After continuum subtraction, the \hi{} data were Fourier-transformed using ``natural'' weighting and a channel width of 4\kms{}. The data were cleaned and restored with the synthesized beam (the size of which is given in Table~\ref{tab:radio} for each galaxy). Primary beam correction was applied. \hi{} distributions (0th moment) were obtained for all galaxies using cutoffs between 3 and 4$\sigma$ and are shown in Fig.~\ref{fig:himap}, while the corresponding \hi{} spectra are shown in Fig.~\ref{fig:hispectra}. For ESO\,121-G020, ESO\,428-G033, \esoq{}, and ESO\,348-G009 mean velocity fields and dispersion maps were also produced as they had sufficient resolution. \subsection{Optical Photometry} \label{sec:obs-opt} {\em BVRI} band CCD images were obtained at the 2.3m telescope as a series of 300\,s or 600\,s exposures during observing runs between April 2002 and February 2004 using the Nasmyth Imager (SITe $1124 \times 1024$ thinned CCD). The imager has a circular field of view with a diameter of 6\farcm62 and a pixel size of 0\farcs59. Table~\ref{tab:obs} gives a summary of the observations taken for each galaxy in each band. The columns are as follows: (2) broad band (Cousins) filters used; (3) total observing time in each of the optical bands including the number of individual exposures; and (4) atmospheric seeing in the final optical images. Most observations were taken at low airmass. Twilight sky flat fields in all bands and bias images were obtained at the same time. On each photometric night several \citet{lan92} standard stars were taken together with shallow 120\,s {\em BVRI} images of the galaxy fields to perform the photometric calibration of the deeper images. Data reduction, photometric calibration, and analysis were carried out within IRAF using standard procedures. After overscan subtraction, bias subtraction, and flatfielding, individual sets of images were registered and the sky level was subtracted. The images for each band were then combined into a single image (to increase signal-to-noise, remove cosmic rays, etc.) and the photometric calibration applied. Fig.~\ref{fig:optimage} shows the resulting master images in the {\em B} band for all nine galaxies. Foreground stars were removed by replacing them with the surrounding sky so that only the galaxy remained. Special care was taken to restore the light distribution under any stars superimposed onto the galaxies, e.g. using the mirror image from across the galaxies center. For more details of this technique see \citet{jer03}. To illustrate the final result Fig.~\ref{fig:starsub} shows the {\em B} band images of the galaxies after cleaning. \section{Radio Properties} \label{sec:radio} ATCA \hi{} follow-up observations are needed to obtain accurate positions of the targeted HIPASS BGC sources and to reliably identify their optical counterparts. For those galaxies where we have high angular resolution \hi{} observations (ESO\,121-G020, ESO\,428-G033, \esoq{}, and ESO\,348-G009), we also analyse their \hi{} structure and kinematics, including the galaxy rotation curve. \subsection{\hi{} Structure} \label{sec:structure} \begin{deluxetable}{lcccccc} \tabletypesize{\scriptsize} \tablecaption{ATCA \hi{} Results. \label{tab:radio}} \tablewidth{0pt} \tablehead{ \colhead{Name} & \colhead{Beam (\hi{})} & \colhead{\Speak{}} & \colhead{\FHI{}} & \colhead{\vsys{}} & \colhead{\whalf{}} & \colhead{\wxx{}} \\ \colhead{} & \colhead{(arcsec)} & \colhead{(Jy)} & \colhead{(\jjks{})} & \colhead{(\kkms{})} & \colhead{(\kkms{})} & \colhead{(\kkms{})} \\ \colhead{(1)} & \colhead{(2)} & \colhead{(3)} & \colhead{(4)} & \colhead{(5)} & \colhead{(6)} & \colhead{(7)} } \startdata \mcg{} & $193\times129$ & $0.232 \pm 0.009$ & $16.2 \pm 0.6$ & $670 \pm 2$ & $112 \pm 2$ & $126 \pm 2$ \\ \\ ESO\,473-G024 & $359\times205$ & $0.17 \pm 0.02$ & $5.7 \pm 0.9$ & $542 \pm 3$ & $37 \pm 2$ & $50 \pm 3$ \\ \\ ESO\,121-G020 & $32\times30$ & $0.204 \pm 0.006$ & $9.1 \pm 0.3$ & $583 \pm 2$ & $47 \pm 2$ & $61 \pm 4$ \\ ~~\atg{} & $32\times30$ & $0.075 \pm 0.006$ & $2.7 \pm 0.2$ & $554 \pm 4$ & $36 \pm 3$ & $56 \pm 8$ \\ \\ ESO\,428-G033 & $45\times30$ & $0.179 \pm 0.005$ & $14.5 \pm 0.3$ & $1728 \pm 2$ & $94 \pm 2$ & $110 \pm 2$ \\ \\ \esoq{} & $38\times35$ & $2.128 \pm 0.005$ & $122 \pm 4$ & $597 \pm 1$ & $64 \pm 2$ & $90 \pm 4$ \\ \\ ESO\,572-G009 & $477\times327$ & $0.25 \pm 0.03$ & $7.2 \pm 1.3$ & $1740 \pm 4$ & $36 \pm 2$ & $49 \pm 2$ \\ \\ ESO\,505-G007 & $413\times304$ & $0.347 \pm 0.009$ & $21 \pm 3$ & $1776 \pm 2$ & $69 \pm 3$ & $88 \pm 5$ \\ \\ IC\,4212 & $444\times357$ & $0.342 \pm 0.009$ & $46.0 \pm 1.0$ & $1476 \pm 1$ & $158 \pm 2$ & $172 \pm 2$ \\ \\ ESO\,348-G009 & $44\times27$ & $0.165 \pm 0.005$ & $13.1 \pm 0.3$ & $648 \pm 1$ & $86 \pm 2$ & $100 \pm 3$ \\ \enddata \end{deluxetable} The \hi{} distributions of all observed galaxies are shown in Fig.~\ref{fig:himap} overlaid onto second generation Digitized Sky Survey (DSS\,{\sc II}) {\em R} band images. The synthesized beam sizes for the individual observations differ significantly due to the use of either East-West or northern hybrid arrays. The ATCA \hi{} spectra of the observed galaxies are shown in Fig.~\ref{fig:hispectra}, with the BGC spectra plotted for comparison. The ATCA \hi{} observations of the galaxy ESO\,121-G020 reveal a previously uncatalogued galaxy at a projected distance of only 3\arcmin{} (see Fig.~\ref{fig:himap}{\em c}). The two galaxies are well resolved spatially, and their systemic velocities differ by only $29 \pm 6$\kms{}, less than the 50\% velocity width of either galaxy. In the following we refer to this companion as \atg{} according to its \hi{} centre position. The \hi{} spectra of the two galaxies as well as the global spectrum of the galaxy pair is shown in Fig.~\ref{fig:121spect} (Fig.~\ref{fig:hispectra}{\em c} shows the global spectrum only). \atg{} has no previous optical measurements and we included it in our optical follow-up observations (see \S~\ref{sec:optp}). Because \atg{} contributes $\sim$20\% to the \hi{} flux density of HIPASS J0615--57, the \hi{} mass to light ratio for ESO\,121-G020 decreases slightly. The galaxy ESO\,505-G007 likewise has a neighbor, the catalogued galaxy ESO\,505-G008, which lies at a projected distance of $\sim7$\arcmin{}. However, the two galaxies are not entirely spatially or spectrally resolved in our \hi{} observations. ESO\,505-G007 appears to have a low inclination in the optical images and its \hi{} line is relatively narrow. In contrast, the galaxy ESO\,505-G008 is seen close to edge-on and shows a broad \hi{} line, as seen in Fig.~\ref{fig:hispectra}{\em g}. By fitting two point sources to the low-resolution \hi{} distribution we determine an \hi{} flux density of $21 \pm 3$\jks{} for ESO\,505-G007, in excellent agreement with the BGC value for HIPASS\,J1203--25 despite the confusion, and $8 \pm 3$\jks{} for ESO\,505-G008. The ATCA \hi{} maps of the other seven galaxies match the optical counterparts identified in the BGC. For \mcg{} (Fig.~\ref{fig:himap}{\em a}) the \hi{} distribution is extended North-South and aligned with the stellar distribution (see \S~\ref{sec:optp}), although little other structure is distinguishable with the large beam. For IC\,4212 (Fig.~\ref{fig:himap}{\em h}) the source appears extended compared to the large beam, and deconvolution with imfit in {\sc MIRIAD} indicates that the \hi{} extends $\sim$200\arcsec{}. We find that for all galaxies, except \esoq{} (see \pI{}), the measured ATCA \hi{} flux densities are in agreement with the BGC values (including the combined ESO\,121-G020/\atg{} system). Our results for the ATCA \hi{} observations of the nine galaxies are listed in Table~\ref{tab:radio}. The columns are as follows: (1) galaxy name (with \atg{} included); (2) size of the synthesized beam; (3) \hi{} peak flux density; (4) integrated \hi{} flux density; (5) \hi{} systemic velocity from the \hi{} line; (6) velocity width of the \hi{} line at 50\% of the peak flux density; and (7) velocity width of the \hi{} line at 20\% of the peak flux density. \subsection{\hi{} Gas Dynamics} \label{sec:velo} \begin{deluxetable}{lccccc} \tabletypesize{\scriptsize} \tablecaption{Rotation curve fit for galaxies with high-resolution ATCA \hi{} data. \label{tab:rot}} \tablewidth{0pt} \tablehead{ \colhead{Name} & \colhead{\vsys{}} & \colhead{{\it PA}} & \colhead{{\em i}} & \colhead{\vmax{}} & \colhead{$r_{\rm max}$} \\ \colhead{} & \colhead{(\kkms{})} & \colhead{(degrees)} & \colhead{(degrees)} & \colhead{(\kkms{})} & \colhead{(arcsec)} \\ \colhead{(1)} & \colhead{(2)} & \colhead{(3)} & \colhead{(4)} & \colhead{(5)} & \colhead{(6)}} \startdata ESO\,121-G020 & $584.5 \pm 1.0$ & $262 \pm 2$ & $78 \pm 5$ & $21 \pm 2$ & $80 \pm 10$ \\ \\ ESO\,428-G033 & $1717 \pm 3$ & $295 \pm 5$ & $11 \pm 5$ & $200^{+160}_{-60}$ & $130 \pm 5$ \\ \\ \esoq{} & $597 \pm 1$ & $119 \pm 2$ & $36 \pm 10$ & $51 \pm 8$ & $370 \pm 20$ \\ \\ ESO\,348-G009 & $646 \pm 2$ & $245 \pm 3$ & $80 \pm 5$ & $50 \pm 5$ & $160 \pm 15$ \\ \enddata \end{deluxetable} The \hi{} velocity fields (1st moment) of the galaxies ESO\,121-G020/\atg{}, ESO\,428-G033, \esoq{} (\pI{}), and ESO\,348-G009 are shown in Fig.~\ref{fig:hivfield}. All galaxies show clear signs of rotation. The velocity field of the galaxy ESO\,428-G033 appears similar to that of \esoq{}, displaying fairly regular, undisturbed gas dynamics. The extreme velocity contours on both the approaching and receding sides close, suggesting that either the rotation curve turns down at large radii or the galaxy is warped. The position angle ($PA$) of ESO\,428-G033, as measured from its \hi{} velocity field, is aligned with the bright optical emission, although the nature of the stellar feature is unclear (see \S~\ref{sec:dis-compd}). Likewise, the position angle of the galaxy ESO\,348-G009 aligns well with the optical disk (see \S\S~\ref{sec:optp} and \ref{sec:dis-comp} for more on the optical details of the galaxies). The analysis of the rotation curve for \esoq{} using {\sc rocur} in {\sc AIPS} was presented in \pI{}. For the other three galaxies we instead used the equivalent procedure {\sc rotcur} in {\sc GIPSY} \citep[both use the tilted ring algorithm described by][]{beg89}. We used the standard procedure of narrowing the free parameters (centre position, \vsys{}, position angle, and inclination) down one at a time until the best fit to all parameters was obtained and a rotation curve could be produced. All three fits were done with 10\arcsec{} rings (the \esoq{} fit used 12\arcsec{} rings). After an initial fit which included both sides of the galaxy, the fit was done individually for the approaching and receding sides of the galaxy to check for any asymmetry. Of the three new fits, the only galaxy which had significant differences between the two sides was ESO\,428-G033, where the inclination fit was lower on the approaching side, which may have been a result of the very low inclination (this was included in the uncertainties). The results of fitting rotation curves to the \hi{} velocity fields of these four galaxies is listed in Table~\ref{tab:rot}. The columns are as follows: (1) galaxy name; (2) systemic velocity; (3) position angle of the galaxy's receding side; (4) inclination angle; (5) maximum rotation velocity; and (6) maximum radius. The final rotation curves are shown in Fig.~\ref{fig:hirotcur}. The very low inclination of ESO\,428-G033 produced a high uncertainty in the rotation velocity values as shown by the curves plotted. In the case of ESO\,348-G009 the curve appears to still be rising at the last points of the rotation curve, suggesting that the \hi{} is not tracing the galaxy out to the radius at which the maximum rotation velocity is reached (so \vmax{} should be considered a lower limit in this case). Similarly for ESO\,121-G020, where we may just be reaching the point where the curve flattens out. In the other two galaxies we appear to have reached the flat part of the curve. \section{Optical Properties} \label{sec:optp} \begin{deluxetable}{lccccccc} \tabletypesize{\scriptsize} \tablecaption{2.3m Telescope Optical Results. \label{tab:opt}} \tablewidth{0pt} \tablehead{ \colhead{Name} & \colhead{Band} & \colhead{$m_{T}$\tablenotemark{a}} & \colhead{$\mu_{0}$\tablenotemark{a}} & \colhead{$\langle \mu \rangle _{\rm eff}$\tablenotemark{a}} & \colhead{$r_{\rm eff}$} & \colhead{$r_{\rm H,0}$} & \colhead{\AG{}} \\ \colhead{} & \colhead{} & \colhead{(mag)} & \colhead{(mag arcsec$^{-2}$)} & \colhead{(mag arcsec$^{-2}$)} & \colhead{(arcsec)} & \colhead{(arcsec)} & \colhead{(mag)} \\ \colhead{(1)} & \colhead{(2)} & \colhead{(3)} & \colhead{(4)} & \colhead{(5)} & \colhead{(6)} & \colhead{(7)} & \colhead{(8)}} \startdata \mcg{} & {\em B} & $15.32 \pm 0.06$ & $22.34 \pm 0.02$ & $23.77 \pm 0.04$ & $19.6 \pm 0.7$ & $51.0 \pm 2.0$ & $0.08 \pm 0.01$ \\ & {\em V} & $14.91 \pm 0.05$ & $21.89 \pm 0.01$ & $22.98 \pm 0.04$ & $16.4 \pm 0.6$ & -- & $0.06 \pm 0.01$ \\ & {\em R} & $14.40 \pm 0.05$ & $21.60 \pm 0.01$ & $22.68 \pm 0.05$ & $18.1 \pm 0.7$ & -- & $0.05 \pm 0.01$ \\ & {\em I} & $13.60 \pm 0.06$ & $21.17 \pm 0.01$ & $22.76 \pm 0.05$ & $27.1 \pm 1.0$ & -- & $0.04 \pm 0.01$ \\ \\ ESO\,473-G024 & {\em B} & $16.38 \pm 0.06$ & $24.63 \pm 0.04$ & $25.30 \pm 0.04$ & $24.3 \pm 0.8$ & $32.4 \pm 0.8$ & $0.08 \pm 0.01$ \\ & {\em V} & $15.44 \pm 0.07$ & $23.73 \pm 0.03$ & $24.46 \pm 0.03$ & $25.4 \pm 0.7$ & -- & $0.06 \pm 0.01$ \\ & {\em R} & $15.19 \pm 0.03$ & $23.51 \pm 0.03$ & $24.18 \pm 0.02$ & $25.1 \pm 0.4$ & -- & $0.05 \pm 0.01$ \\ & {\em I} & $14.76 \pm 0.07$ & $23.42 \pm 0.08$ & $24.35 \pm 0.02$ & $33.0 \pm 0.9$ & -- & $0.04 \pm 0.01$ \\ \\ ESO\,121-G020 & {\em B} & $15.27 \pm 0.05$ & $23.36 \pm 0.03$ & $23.95 \pm 0.02$ & $21.7 \pm 0.4$ & $47.0 \pm 2.0$ & $0.17 \pm 0.03$ \\ & {\em R} & $14.62 \pm 0.04$ & $22.71 \pm 0.03$ & $23.29 \pm 0.02$ & $21.6 \pm 0.5$ & -- & $0.11 \pm 0.02$ \\ ~~\atg{} & {\em B} & $17.01 \pm 0.06$ & $22.81 \pm 0.04$ & $23.34 \pm 0.02$ & $7.4 \pm 0.3$ & $18.6 \pm 0.6$ & $0.17 \pm 0.03$ \\ & {\em R} & $16.36 \pm 0.06$ & $22.07 \pm 0.02$ & $22.72 \pm 0.02$ & $7.5 \pm 0.2$ & -- & $0.11 \pm 0.02$ \\ \\ ESO\,428-G033 & {\em B} & $16.90 \pm 0.10$ & $23.33 \pm 0.11$ & $24.69 \pm 0.02$ & $14.4 \pm 0.5$ & $37.2 \pm 1.2$ & $1.10 \pm 0.18$ \\ & {\em V} & $16.13 \pm 0.10$ & $23.12 \pm 0.02$ & $24.23 \pm 0.05$ & $16.7 \pm 0.8$ & -- & $0.85 \pm 0.14$ \\ & {\em R} & $15.61 \pm 0.08$ & $22.56 \pm 0.02$ & $23.48 \pm 0.05$ & $15.0 \pm 0.9$ & -- & $0.68 \pm 0.11$ \\ & {\em I} & $15.04 \pm 0.09$ & $22.01 \pm 0.03$ & $22.95 \pm 0.04$ & $15.2 \pm 0.7$ & -- & $0.50 \pm 0.08$ \\ \\ \esoq{} & {\em B} & $16.13 \pm 0.07$ & $24.97 \pm 0.03$ & $25.48 \pm 0.02$ & $29.7 \pm 0.6$ & $57.6 \pm 0.6$ & $0.95 \pm 0.15$ \\ & {\em V} & $14.89 \pm 0.06$ & $23.65 \pm 0.03$ & $24.14 \pm 0.02$ & $28.3 \pm 0.7$ & -- & $0.73 \pm 0.12$ \\ & {\em R} & $14.38 \pm 0.05$ & $23.16 \pm 0.02$ & $23.64 \pm 0.03$ & $28.4 \pm 0.5$ & -- & $0.59 \pm 0.09$ \\ & {\em I} & $13.76 \pm 0.06$ & $22.91 \pm 0.04$ & $23.40 \pm 0.03$ & $33.9 \pm 0.8$ & -- & $0.43 \pm 0.07$ \\ \\ ESO\,572-G009 & {\em B} & $16.79 \pm 0.05$ & $24.96 \pm 0.05$ & $26.03 \pm 0.02$ & $28.1 \pm 0.7$ & $30.6 \pm 1.2$ & $0.16 \pm 0.03$ \\ & {\em V} & $15.65 \pm 0.07$ & $23.69 \pm 0.02$ & $25.12 \pm 0.02$ & $31.1 \pm 0.8$ & -- & $0.12 \pm 0.02$ \\ & {\em R} & $15.40 \pm 0.04$ & $23.42 \pm 0.03$ & $24.75 \pm 0.01$ & $29.6 \pm 0.5$ & -- & $0.10 \pm 0.02$ \\ \\ ESO\,505-G007 & {\em B} & $14.20 \pm 0.06$ & $23.99 \pm 0.04$ & $24.00 \pm 0.01$ & $36.4 \pm 0.7$ & $82.8 \pm 1.2$ & $0.36 \pm 0.06$ \\ & {\em V} & $14.48 \pm 0.05$ & $23.89 \pm 0.04$ & $23.96 \pm 0.01$ & $31.3 \pm 0.6$ & -- & $0.28 \pm 0.04$ \\ & {\em R} & $13.97 \pm 0.04$ & $23.46 \pm 0.04$ & $23.66 \pm 0.02$ & $34.6 \pm 0.5$ & -- & $0.22 \pm 0.04$ \\ \\ IC\,4212 & {\em B} & $14.11 \pm 0.04$ & $22.38 \pm 0.02$ & $24.42 \pm 0.02$ & $46.0 \pm 1.0$ & $98.4 \pm 0.6$ & $0.19 \pm 0.03$ \\ & {\em V} & $13.69 \pm 0.05$ & $22.00 \pm 0.03$ & $23.98 \pm 0.01$ & $45.7 \pm 0.9$ & -- & $0.14 \pm 0.02$ \\ & {\em R} & $13.29 \pm 0.06$ & $21.53 \pm 0.02$ & $23.53 \pm 0.02$ & $44.5 \pm 0.9$ & -- & $0.12 \pm 0.02$ \\ \\ ESO\,348-G009 & {\em B} & $14.81 \pm 0.07$ & $23.78 \pm 0.04$ & $24.79 \pm 0.03$ & $39.5 \pm 0.9$ & $72.0 \pm 0.6$ & $0.06 \pm 0.01$ \\ & {\em V} & $14.68 \pm 0.06$ & $23.42 \pm 0.03$ & $24.40 \pm 0.03$ & $35.1 \pm 1.0$ & -- & $0.04 \pm 0.01$ \\ & {\em R} & $14.48 \pm 0.05$ & $23.13 \pm 0.03$ & $24.03 \pm 0.03$ & $32.4 \pm 0.8$ & -- & $0.04 \pm 0.01$ \\ & {\em I} & $13.75 \pm 0.10$ & $22.75 \pm 0.03$ & $23.72 \pm 0.02$ & $39.3 \pm 1.1$ & -- & $0.03 \pm 0.01$ \\ \enddata \tablenotetext{a}{Correction for Galactic extinction not applied.} \end{deluxetable} The observed {\em B} images taken on the 2.3m Telescope of all nine galaxies are in Fig.~\ref{fig:optimage}, while Fig.~\ref{fig:starsub} shows the same images after star subtraction. Star contamination is obviously a problem for \esoq{} and especially ESO\,428-G033, both of which are close to the Galactic Plane, so special care was taken with these images. Some of the galaxies (notably \mcg{}, \esoq{}, ESO\,505-G007 and ESO\,348-G009) were affected by moderately bright foreground stars sitting on top of the galaxy that had to be removed with care. As mentioned in \S~\ref{sec:radio}, a companion galaxy to ESO\,121-G020 was found in the \hi{} imaging. The optical image for ESO\,121-G020 also includes \atg{}, and the photometry for that galaxy is included with the other results here. For each galaxy a growth curve was measured on the star-subtracted images from the luminosity weighted center in 2 pixel ($\sim$1\farcs2) circular aperture rings to obtain the total intensity and a surface brightness profile. Our results for the optical {\em BVRI} photometry taken on the 2.3m Telescope are in Table~\ref{tab:opt}. No Galactic extinction correction was applied to the apparent magnitude and surface brightness. The columns are as follows: (1) galaxy name; (2) broad band (Cousins) filter used; (3) total apparent magnitude; (4) central surface brightness; (5) effective surface brightness (i.e. the average surface brightness out to the half light radius); (6) half light (effective) radius; (7) radius out to $\mu = 26.6$~mag arcsec$^{-2}$ (i.e. the Holmberg radius in the {\em B} band, extinction corrected); and (8) Galactic extinction correction from SFD98. Surface brightness profiles for the nine initially selected galaxies in each of the observed bands are in Fig.~\ref{fig:sbpro}, while the profile of \atg{} is in Fig.~\ref{fig:sbpatca}. All profiles have been corrected for Galactic extinction, but no attempt to correct for inclination was made as it is difficult to calculate given the morphology of many galaxies, and as the correction for edge on galaxies (such as ESO\,348-G009) would be unrealistic without correction for the thickness of the stellar disk. The profiles for several galaxies show some of the underlying morphology. \mcg{} (Fig.~\ref{fig:sbpro}{\em a}) has two distinct components, an inner bright bulge region which is prominent in the optical image, and a surrounding low surface brightness disk (almost 4~mag fainter in surface brightness). IC\,4212 (Fig.~\ref{fig:sbpro}{\em h}) appears to have a small central bulge, and the effect of the spiral arms are evident in the bumps of the profile. ESO\,572-G009 (Fig.~\ref{fig:sbpro}{\em f}) also seems to have a small central bulge, which is seen in the image (Fig.~\ref{fig:starsub}{\em f}). ESO\,348-G009 and ESO\,428-G033 (Figs.~\ref{fig:sbpro}{\em d} and {\em i}) exhibit pure exponential disks, while ESO\,473-G024 (Fig.~\ref{fig:sbpro}{\em b}) is similar but with a flatter central region. The various light profiles of ESO\,505-G007 (Fig.~\ref{fig:sbpro}{\em g}) are flat in the central region as the luminosity weighted centre of this galaxy is a faint region between many denser sites that are probably star formation regions. The optical structure of this galaxy (see Fig.~\ref{fig:starsub}{\em g}) possibly reflects a highly disturbed galaxy with increased star formation due to interaction with the neighboring galaxy ESO\,505-G008 (the source which contaminates its \hi{} spectrum). The new galaxy \atg{} notably has a much steeper exponential profile than its companion ESO\,121-G020, as well as a higher central surface brightness. This denser structure may be the result of recent star formation (there is a bright region in the centre of the galaxy which could be a star formation site), and may have lead to it previously being missed in optical surveys if it were mistaken for a star \citep[it is falsely identified as a star in the USNO star catalogue,][]{mon03}. For only two galaxies, ESO\,473-G024 and \esoq{}, did we find {\em B} band apparent magnitudes that are in agreement with the values quoted in LEDA (even within their sometimes large error bars). In the case of all seven other galaxies our value of \mB{} was brighter than that given in LEDA, sometimes by several magnitudes. The effect this has on the \hi{} mass-to-light ratios and why this occurred so often in our sample will be discussed in the following sections (\S\S~\ref{sec:dis-comp} and \ref{sec:dis-slide}). \section{Discussion of Individual Galaxies} \label{sec:dis-indi} \subsection{Comparing Optical and \hi{} Properties} \label{sec:dis-comp} Now that we have more accurate optical and \hi{} measurements we can recalculate many of the physical properties of the galaxies, including \mlr{}. Table~\ref{tab:summary} summarises new parameters from our observations. The columns are as follows: (1) galaxy name; (2) {\em B} band absolute magnitude; (3) {\em B} band luminosity; (4) \hi{} mass; (5) \hi{} mass-to-{\em B} band luminosity ratio (\mlr{}); (6) total dynamical mass; (7) \hi{} mass-to-total mass ratio; and (8) total mass-to-{\em B} band luminosity ratio. The latter three values (\Mtot{}, \mtmr{}, and \tmlr{}) are only given for the four galaxies where we fit rotation curves. \begin{deluxetable}{lccccccc} \tabletypesize{\scriptsize} \tablecaption{Summary of Derived Galaxy Properties from ATCA and 2.3m Data. \label{tab:summary}} \tablewidth{0pt} \tablehead{ \colhead{Name} & \colhead{\MB{}} & \colhead{\LB{}} & \colhead{\MHI{}} & \colhead{\mlr{}} & \colhead{\Mtot{}} & \colhead{\mtmr{}} & \colhead{\tmlr{}} \\ \colhead{} & \colhead{(mag)} & \colhead{($\times 10^{7}$\Lsun{})} & \colhead{($\times 10^{7}$\Msun{})} & \colhead{(\mmls{})} & \colhead{($\times 10^{9}$\Msun{})} & \colhead{} & \colhead{(\mmls{})} \\ \colhead{(1)} & \colhead{(2)} & \colhead{(3)} & \colhead{(4)} & \colhead{(5)} & \colhead{(6)} & \colhead{(7)} & \colhead{(8)}} \startdata \mcg{} & $-14.65 \pm 0.06$ & $11.3 \pm 0.6$ & $34.5 \pm 1.3$ & $3.0 \pm 0.3$ \\ \\ ESO\,473-G024 & $-13.10 \pm 0.06$ & $2.7 \pm 0.1$ & $7.8 \pm 1.2$ & $2.8 \pm 0.6$ \\ \\ ESO\,121-G020 & $-12.97 \pm 0.06$ & $2.39 \pm 0.13$ & $3.61 \pm 0.12$ & $1.50 \pm 0.13$ & $0.16 \pm 0.05$ & $0.23 \pm 0.11$ & $7 \pm 4$ \\ ~~\atg{} & $-11.23 \pm 0.07$ & $0.48 \pm 0.03$ & $1.07 \pm 0.08$ & $2.2 \pm 0.3$ \\ \\ ESO\,428-G033 & $-16.7 \pm 0.2$ & $28 \pm 5$ & $130 \pm 3$ & $4.5 \pm 0.9$ & $110^{+190}_{-70}$ & $0.012^{+0.020}_{-0.008}$ & $\sim380$ \\ \\ \esoq{} & $-12.9 \pm 0.2$ & $2.3 \pm 0.4$ & $50.8 \pm 1.7$ & $22 \pm 4$ & $4.5 \pm 1.6$ & $0.11 \pm 0.04$ & $200 \pm 110$ \\ \\ ESO\,572-G009 & $-14.92 \pm 0.06$ & $14.5 \pm 0.8$ & $71 \pm 13$ & $4.8 \pm 1.1$ \\ \\ ESO\,505-G007 & $-17.75 \pm 0.08$ & $196 \pm 14$ & $210 \pm 30$ & $1.1 \pm 0.2$ \\ \\ IC\,4212 & $-17.36 \pm 0.05$ & $137 \pm 6$ & $356 \pm 8$ & $2.55 \pm 0.17$ \\ \\ ESO\,348-G009 & $-14.87 \pm 0.07$ & $13.8 \pm 0.9$ & $21.8 \pm 0.5$ & $1.56 \pm 0.16$ & $3.8 \pm 1.1$ & $0.06 \pm 0.02$ & $27 \pm 10$ \\ \enddata \end{deluxetable} \subsubsection{\mcg{}} Our {\em B} band apparent magnitude for this galaxy is $\sim0.5$~mag brighter than that expressed in LEDA, which was based on the APM plate scan results of \citet{mad90}. The galaxy has a distinctive optical morphology (Fig.~\ref{fig:starsub}{\em a}). A shallow CCD image shows only a small, bright, centrally concentrated circular region like a BCD. However, deeper observations reveal a second, low surface brightness disk component extending well beyond this central bulge, which is clearly visible in our surface brightness profile for this galaxy (Fig.~\ref{fig:sbpro}{\em a}). Difficulty in measuring the full extent of this faint disk could account for the disagreement between our measurement and the previous result of \citet{mad90}. The \hi{} flux density result on the other hand is in excellent agreement with the BGC and the observations of \citet{fou90}, suggesting we have recovered almost all the Parkes flux density. The result of our combined \hi{} and optical values is that the \hi{} mass-to-light ratio drops moderately to $3.0 \pm 0.3$\mls{}, still higher than ``typical'' late-type galaxies but not unusually so. Despite the large beam for the \mcg{} observations, we can clearly see the \hi{} envelope is extended in a North-South direction. This is in the same direction as the outer low surface brightness optical disk, and the two correlate quite well in both shape and extent. \subsubsection{ESO\,473-G024} Of the nine sample galaxies, ESO\,473-G024 had the best agreement with the {\em B} band apparent magnitude quoted by LEDA, also agreeing with the results in \citet{lau89} and \citet{mad90}. Our \hi{} flux density is in agreement with both the BGC and \citet{fou90}. Consequently, \mlr{} remained at a moderate ratio of $2.8 \pm 0.6$\mls{}, one of the higher ratios of this sample after accurate measurements. It is a well studied dwarf irregular from the Sculptor group, and the \mlr{} has been noted before \citep{ski03a,ski03b}. In the optical the galaxy is extended in the North-South direction, and has a number of potential star formation regions (see Fig.~\ref{fig:starsub}{\em b}), some of which were studied by \citet[][ see also \S~\ref{sec:dis-implicat}]{ski03a,ski03b}. \subsubsection{ESO\,121-G020 and \atg{}} The unexpected discovery of the companion galaxy \atg{} to the South East of ESO\,121-G020 must affect the \mlr{} previously calculated for this galaxy, as the BGC \hi{} flux density measurement for HIPASS\,J0615--57 includes the \hi{} emission from both galaxies. But, more importantly, our optical {\em B} magnitude determined here is $\sim1.7$~mag brighter than the value in LEDA, although it is closer to the \citet{lau89} total magnitude ($15.85 \pm 0.09$~mag). The new ratio for the combined system is $1.62 \pm 0.18$\mls{}, and for the individual galaxies it is $1.50 \pm 0.13$\mls{} for ESO\,121-G020 and $2.2 \pm 0.3$\mls{} for \atg{}. Down to our sensitivity limits there is neither an \hi{} nor a stellar bridge between the two galaxies detected. \subsubsection{ESO\,428-G033} \label{sec:dis-compd} At only 6\fdg9 from the Galactic Plane, ESO\,428-G033 suffers from relatively high Galactic extinction \citep[$1.10 \pm 0.18$~mag in {\em B},][]{sch98} and foreground star contamination (see Fig.~\ref{fig:optimage}{\em d}). Our {\em B} band magnitude of $16.90 \pm 0.10$~mag is $\sim1$~mag brighter than the LEDA value (which has a high uncertainty of 0.5~mag), but agrees with the \citet{lau89} measurement ($16.83 \pm 0.09$~mag). Despite this, with our new data the galaxy still has a high \hi{} mass-to-light ratio of $4.5 \pm 0.9$\mls{}, which is in good agreement with the results of \citet{kra92}. The optical morphology of the galaxy is difficult to determine due to the high foreground star obscuration. We can only discern an elongated stellar feature (see Fig.~\ref{fig:starsub}{\em d}), which could be a galaxy disk. Alternatively, this could be a central bar, while the fainter spiral arms remain undetected. The \hi{} distribution of ESO\,428-G033 is nearly circular, indicating a disk that is close to face on, which favors the optical feature being a bar rather than a disk. The position angle of velocity field is aligned with the optical emission. \subsubsection{\esoq{}} This low surface brightness dwarf irregular galaxy was discussed in detail in \pI{}, where we confirmed that it did have the unusually high \mlr{} as initially suggested by the combination of the BGC results and the magnitudes listed in LEDA. It is included here for comparison with the other galaxies and we refer the reader to the previous work for more details on it. After obtaining new estimates of \mB{} and \FHI{} for all our sample galaxies, \esoq{} remains as the stand out galaxy with \mlr{} = $22 \pm 4$\mls{}. To our knowledge it has one of the highest \hi{} mass-to-light ratios that has been confirmed by accurate measurement to date for any galaxy system, being approximately double the ratio of the best example in the literature DDO\,154 \citep{car89}, and about four times the ratio of our next highest sample objects (ESO\,572-G009 and ESO\,428-G033). Like DDO\,154 and another known high \mlr{} galaxy NGC\,3741 \citep{beg05}, \esoq{} has a highly extended \hi{} envelope, over six times the optical Holmberg radius. \citet{beg05} also point out that all three galaxies are isolated and have low tidal indexes from nearby galaxies. The nearest neighbor we can identify to \esoq{} is approximately 1.7 Mpc away in the Centaurus A Group. \subsubsection{ESO\,572-G009} Although our \hi{} flux density was in good agreement with the BGC and \citet{fou90}, our apparent {\em B} magnitude is about half a magnitude brighter than that given in LEDA and that of \citet{lau89}. Despite this correction ESO\,572-G009 remains one of the few galaxies in our sample to retain a high \mlr{} at $4.8 \pm 1.1$\mls{}. Morphologically it is a faint low surface brightness galaxy and appears to have two stellar components, a brighter cigar shaped central region (likely to be a central bar) surrounded by a fainter disk that extends only a short distance (see Fig.~\ref{fig:starsub}{\em f}). \subsubsection{ESO\,505-G007} A previous measurement of the integrated \hi{} flux density taken on the Effelsberg Radio Telescope by \citet[ $19.3 \pm 2.2$\jks]{ric87} agrees with both the BGC's and our ATCA results. However, literature \mB{} measurements vary wildly \citep[both $17.64 \pm 0.09$~mag and $15.95 \pm 0.09$~mag from][ dependent on the isophotal level it is measured to]{lau89}. Our measurement is $\sim$3.5~mag brighter than that listed in LEDA (corresponding to a 25 times more luminous). The severe underestimate of the {\em B} band apparent magnitude quoted by LEDA for this galaxy has a dramatic effect on the \hi{} mass-to-light ratio. From the highest ratio of the 789 BGC galaxies with {\em B} magnitudes in LEDA, \mlr{} for ESO\,505-G007 has dropped down to a more typical ratio of $1.18 \pm 0.12$\mls{}, the lowest of our sample galaxies. We note that optically ESO\,505-G007 has an unusual irregular morphology (see Fig.~\ref{fig:starsub}{\em g}), with several large clumpy structures and some ragged spiral arm-like features, which could suggest recent disruption and star formation triggered by interaction with ESO\,505-G008. \subsubsection{IC\,4212} IC\,4212 is an unusual galaxy to have in our sample since its optical morphology is that of a face on spiral rather than the dwarf irregular we might expect for the magnitude given in LEDA. It has two bright, loosely wound arms, several fainter arms and a small bright central bar (see Fig.~\ref{fig:starsub}{\em h}). The uncertainty in the apparent magnitude given in LEDA of $\pm 1.1$~mag is quite large. Our optical measurements find that IC\,4212 is a much brighter galaxy than suggested by the \mB{} value LEDA lists (over two magnitudes). This means that the \mlr{} is not the extreme value initally suggested (the third highest ratio of the 789 BGC galaxies). However, IC\,4212 retains at a ratio of $2.55 \pm 0.17$\mls{}, which relatively high for a galaxy with such a distinct spiral structure, typical values for late type spirals being less than 1\mls{} \citep{rob94}. We measure a deconvolved \hi{} diameter of $\sim$200\arcsec{}, suggesting it may extend significantly beyond the optical disk. \subsubsection{ESO\,348-G009} The apparent {\em B} magnitude we measured for ESO\,348-G009 was again brighter than the value listed in LEDA, while the \hi{} flux density is consistent with the BGC result. This means we end up with an \hi{} mass-to-light ratio of $1.56 \pm 0.16$\mls{}, down from the preliminary value of $9 \pm 7$\mls{}. The optical image shows an edge on disk galaxy extending East-West, with some clumpy structures visible along the length of the disk (see Fig.~\ref{fig:starsub}{\em i}). The position angle of the \hi{} velocity field is aligned closely with the stellar disk. \subsection{Distance Uncertainties} \label{sec:dis-dist} As was discussed in \pI{}, the use of distances calculated from Local Group velocities for galaxies in the Local Universe can be problematic as the peculiar velocities in nearby groups are potentially of similar order to the redshifts themselves, and the local Hubble flow can differ from the cosmological expansion. Evidence from studies on the Sculptor group \citep{jer98} suggests that the local velocity-distance relationship is much steeper in the direction of this group than for galaxies further out due to the probable gravitational influence of the Local Group. If we use the Hubble constant of \citet[ H$_0$ = 119\kms\,Mpc$^{-1}$]{jer98} for the three Sculptor Group members in our sample \citep[membership confirmed using][]{cot97} then the Local Group velocity distances to these galaxies would be less than listed in Table~\ref{tab:prop}, with 6.0 Mpc for \mcg{}, 4.8 Mpc for ESO\,473-G024, and 5.3 Mpc for ESO\,348-G009. This would put all three on the far side of the Sculptor group \citep{jer98}, and would mean that the distance dependent quantities in Table~\ref{tab:summary} (everything except \mlr{}) would require adjustment. \section{Discussion} \label{sec:dis} \subsection{The revised \hi{} mass-to-light ratios} \label{sec:dis-slide} The plot in Fig.~\ref{fig:slidenine} shows \mlr{} versus \MB{} (as in Fig.~\ref{fig:mlmbt}) with the new positions of the nine target galaxies resulting from our observations. The lines connect our results (large points with error bars) to the initial estimates (open circles). While the new measurements resulted in lower \mlr{} values for all selected galaxies, the decrease is particularly significant for galaxies which had preliminary values of \mlr{} $> 5$\mls{}. While disappointing, our result is not too surprising given that we selected galaxies from the BGC with initially the most extreme \mlr{} values and large uncertainties in their optical magnitudes, so we preferentially selected galaxies with underestimated \mB{}. For many of our target galaxies optical magnitudes exist that agree with our results \citep[e.g.][]{lau89}, but the mean magnitudes available from LEDA were generally highly underestimated. All but one of the selected galaxies, \esoq{}, now have revised \hi{} mass-to-light ratios in the range $\sim$1-5\mls{}. Such revisions are by no means uncommon among claims of high \mlr{} galaxies as we discussed in \S~6.2 of \pI{}, and as seen in \citet{vzee97} and \citet{chu02}. The faint luminosity of these galaxies and the difficulty in getting high quality data in both the optical and radio regime make examples of high \mlr{} galaxies difficult to find. Only a few other galaxies with confirmed high \mlr{} are in the literature, most notably DDO\,154 \citep[9.4\mls{}][]{car89,hof93}, UGCA\,292 \citep[7.0\mls{}][]{you03}, and NGC\,3741 \citep[5.8\mls{}][]{beg05}. While we must be careful of small number statistics, the trend of all the findings of current studies strongly suggest that there do not appear to be large numbers of ``dim'' galaxies like \esoq{} in the local Universe. Therefore high \mlr{} galaxies cannot account for much of the discrepancy between observations and theoretical predictions of low mass galaxy numbers. But there are other possible ways that galaxies could be missed observationally, and several suggestions how they might be detected. Some methods have been proposed for finding ``empty'' dark matter halos, such as the suggestion to use the Milky Way halo microlensing statistics to look for dark matter satellite influence \citep{wid98}, or to analyse the gravitational lensing of quasars to determine the dark matter sub-halos of the lensing object \citep{moo99,dal02}. However, the existence of halos without baryons is still highly speculative. True ``dark'' galaxies in the form of isolated, rotating, extragalactic \hi{} clouds have so far proven elusive \citep{rya02,kor04,doy05}. Some isolated \hi{} sources have been found in HIPASS \citep{kil00,ryd01,ryd04} and were interpreted as high velocity clouds or tidal debris by the respective authors. A recent claim of a ``dark galaxy'' close to the one-armed spiral NGC\,4254, in the outskirts of the Virgo Cluster \citep{min05} also appears to be tidal debris \citep{bek05}; see also \citep{oos05}. \citet{tay05} have discussed theoretically that an isolated \hi{} cloud which formed without a stellar component is likely to be unstable to star formation, and therefore would not remain dark. \subsection{The importance of dust extinction} \label{sec:dis-extinct} As well as the accuracy of the {\em B} band photometry and \hi{} flux density it is also important to discuss the one contributor to \mlr{} that is beyond the scope of our observations, the dust extinction due to both our Galaxy (see Table~\ref{tab:opt}) and the host galaxy (``internal'' extinction). As we noted in \pI{} Galactic extinction is particularly important for \esoq{} due to its sky position only 10\fdg5 from the Galactic Plane. ESO\,428-G033 at $b$ = -6\fdg9 is the only other galaxy that is subject to similar Galactic extinction (\AB{} = $1.10 \pm 0.18$~mag, SFD98). The uncertainties in the SFD98 Galactic extinction grow proportionally to the value. This means that it only contributes a significant fraction to the total error in \mlr{} for \esoq{} and ESO\,428-G033. Close to the Galactic Plane the dust distribution can be patchy. Fig.~\ref{fig:schlegel} shows the SFD98 dust extinction maps in terms of \AB{} for the regions around ESO\,428-G033 and \esoq{}, with the last \hi{} contour from Fig.~\ref{fig:himap} superimposed for reference. Both maps show relatively low variation in the amount of extinction over the field around the galaxy, less than the uncertainty in \AB{} in both cases. This suggests that the SFD98 value at the position of both galaxies is an accurate representation of the true Galactic extinction. ESO\,428-G033 was studied by \citet{kra92} as part of an investigation of galaxies in a region of the Galactic Plane with reduced dust extinction. Their results are in agreement with our measurements, and in general they found that the properties of their sample objects were typical of other samples of nearby galaxies despite the relatively high extinction. It is worth noting that even if the Galactic extinction was much higher at the position of \esoq{}, say by 0.5~mag, the galaxy would still have a very high \mlr{} of $14 \pm 4$\mls{}. While we can at least get a relatively accurate estimate of Galactic extinction, dust extinction internal to the host galaxy itself is much harder to quantify. Dust extinction in late-type galaxies is poorly understood but thought to be lower than in early-type spirals due to low metalicity and ineffective dust accretion processes \citep{dwe98,hir99}. We would expect a disk galaxy to have significant extinction when viewed close to edge on, like ESO\,348-G009 ($i = 80$\degr{}$\pm 5$\degr{}, see Table~\ref{tab:rot}), due to the geometry of the dust distribution. We discussed \esoq{}'s possible internal extinction in \pI{}, concluding that it is most likely low for this close to face on galaxy. The galaxy ESO\,121-G020 may have a high inclination angle of $78$\degr{}$\pm 5$\degr{} (Table~\ref{tab:rot}), in contrast to the optical dimensions, and a moderate amount of internal extinction. The optical emission of the galaxy ESO\,428-G033 is highly obscured by foreground stars. Due to the current inability to estimate internal extinction we have not accounted for it in our \mlr{} calculations. \subsection{Physical Characteristics of High \mlr{} Galaxies Implications for Their Existence} \label{sec:dis-implicat} Several common elements between the galaxies with the most extreme \mlr{}'s are becoming more evident. The three best examples, \esoq{} ($22 \pm 4$\mls{}, \pI{}), DDO\,154 \citep[9.4\mls{},][]{car89,hof93}, and NGC\,3741 \citep[5.8\mls{},][]{beg05}, all have \hi{} envelopes that are 5 to 8 times the optical Holmberg radius, as IC\,4212 may also have. This may be because the gas is at a low density and is in a stable state, as was seen for \esoq{} in \pI{}. All these galaxies also have low tidal indexes \citep{kar04}, indicating {\em they are isolated in space} and have little external stimulation to form stars. Despite having very low stellar content for their dynamical masses, the baryonic masses of the galaxies are always of the order of $\sim10\%$ of the total dynamical mass, which is a typical fraction seen in galaxies from $L_{*}$ to the dwarf regime \citep{beg05}. This suggests {\em that high \mlr{} galaxies are not lacking the baryons to create stars, but are underluminous as they lack either the internal or external stimulation for further star formation}. \citet{ski03b} obtained spectra of \hii{} regions within five Sculptor group galaxies including two in our sample, ESO\,473-G024 and ESO\,348-G009. For ESO\,473-G024, for which we found a moderately high \mlr{}, they were able to produce oxygen and nitrogen abundances. The oxygen abundance indicated a low metalicity which is typical of other late type galaxies. Normally dwarf galaxies with a similar low metalicity have low nitrogen to oxygen ratios in a narrow range around an average of $\log{N/O} \simeq -1.6$ \citep{izo99}. It is thought this is because these galaxies are undergoing their first burst of star formation, and that the nitrogen from this burst has not yet had time to dissipate into the ISM \citep[coming from type {\sc ii} supernovae of intermediate mass stars, while oxygen comes from higher mass stars, see][]{ski03b}. However, \citet{ski03b} found that the N/O ratio for ESO\,473-G024 was relatively high. Most importantly for our study, they compared these results to a study of the high \mlr{} galaxy DDO\,154 by \citet{ken01} which shows the same trend for N/O. In both galaxies nitrogen from any past star formation events has had time to disperse and there is no current burst of star formation to reduce the N/O ratio, neither galaxy having a particularly high current star formation rate \citep[][ respectively]{ski03a,ken01} and like \esoq{} they may be considered quiescent galaxies. In fact, their current SFR and luminosity \citep[which can be used as a rough estimate of the average past star formation rate,][]{tin80} are similar to what was found for \esoq{} in \pI{}. This suggests that {\em high \hi{} mass-to-light ratio galaxies may have been forming stars at a low, constant rate over their lifetimes.} In order to understand further the significance of this result we would need to expand this study to spectra of \hii{} regions in other high \mlr{} galaxies (especially \esoq{}) and look at other ratios which may indicate the timing of star formation events. Metalicity may also be an important element in determining the fraction of baryons which remain in gas form \citep{tay05}. A close look at the plots of \mlr{} versus \MB{} (Figs.~\ref{fig:mlmbt} and \ref{fig:slidenine}) shows that there is potentially an upper envelope to a galaxy's \hi{} mass-to-light ratio at a given luminosity. Low luminosity galaxies appear to be able to have a higher portion of their detectable baryons in the form of neutral hydrogen than galaxies around $L_{*}$, where the baryonic mass is dominated by stars even for the most gas rich galaxies. What this suggests is that {\em there is a minimum quantity of stars a galaxy will form that goes as a function of initial baryonic mass}. Support for this idea can also be found in the theoretical work of \citet{tay05}. Whether or not a galaxy forms more than this minimum is likely to be a function of such factors as environment and the galaxy's dark matter properties. \citet{tay05} developed models to determine whether a neutral gas disk without stars (a ``dark galaxy'') could remain dynamically stable or if some gas will collapse and form stars. They found that without an internal radiation field the majority of the gas in the disk will become gravothermally unstable, even for galaxies with very low baryonic masses (down to $5\times10^{6}$\Msun{}). They also found that the fraction of unstable gas decreases as the baryonic mass decreases. This may provide an explanation to why we see the slope in the upper envelope for \mlr{}, the lower mass galaxies only having to convert a much smaller fraction of their baryons to stars in order to become stable. Galaxies such as \esoq{} and DDO\,154 are close to our upper envelope, and may define the extreme cases of galaxies which have formed only the minimum number of star required in order to remain stable and have not experienced any other events which may trigger star formation. In this way they might be used to distinguish between various models for galaxy collapse by defining the minimum star formation required for stability. We will further explore this possibility, and other properties that vary with \mlr{}, with a larger sample of 37 late type dwarf galaxies in an upcoming paper. \section{Conclusions} \label{sec:conc} We obtained accurate optical CCD apparent magnitudes and \hi{} flux densities for nine late type dwarf galaxies and recalculated their \hi{} mass-to-light ratios. The new \mlr{} values are: \begin{itemize} \item $22 \pm 4$\mls{} for \esoq{}, \item $4.8 \pm 1.1$\mls{} for ESO\,572-G009, \item $4.5 \pm 0.9$\mls{} for ESO\,428-G033, \item $3.0 \pm 0.3$\mls{} for \mcg{}, \item $2.8 \pm 0.6$\mls{} for ESO\,473-G024, \item $2.6 \pm 0.2$\mls{} for IC\,4212, \item $1.6 \pm 0.2$\mls{} for ESO\,348-G009, \item $1.5 \pm 0.1$\mls{} for ESO\,121-G020, and \item $1.2 \pm 0.1$\mls{} for ESO\,505-G007. \end{itemize} Many of these \hi{} mass-to-light ratios are significantly below the initial estimates, due to inaccurate magnitude estimates in the literature. This strongly emphasises the importance of having accurate observations in both the \hi{} line and the optical. Based on the new \hi{} mass-to-light ratio distribution we conclude that genuine ``dim'' galaxies with high ratios (\mlr{}$>$5\mls{}) are rare in the local Universe. A previously uncatalogued companion galaxy to ESO\,121-G020 was found at a projected distance of 3\arcmin{}. \atg{} has an \hi{} mass of $\sim10^{7}$\Msun{} and \mlr{} of $2.2 \pm 0.3$\mls{}. This was the only such companion detected, and is well within the beam of the Multibeam instrument used by the HIPASS survey. Despite our low resolution \hi{} observations we were able to separate the galaxy ESO\,505-G007 from its close neighbor ESO\,505-G008 and determined \hi{} flux densities of $21 \pm 3$ and $8 \pm 3$\jks{}, respectively. The best examples of high \mlr{} dwarf galaxies in the literature all have highly extended \hi{} disks, are spatially isolated and have normal baryonic content for their dynamical masses. The galaxies are not lacking the baryons to create stars, but are underluminous as they lack either the internal or external stimulation for further star formation. Future examination of element abundances within star formation sites of high \mlr{} galaxies may reveal important clues about their star formation history. Recent observations \citep{ski03b,ken01} support the idea that high \mlr{} galaxies may have been forming stars at a low, constant rate over their lifetimes as proposed in \pI{}. There may be a minimum quantity of stars a galaxy will form that depends on the initial baryonic mass, which is supported by the theoretical work of \citet{tay05}. If this is true then maybe high \hi{} mass-to-light ratio galaxies have over their lifetimes only formed the minimum number of stars necessary to maintain the stability of their \hi{} gas disk. \section*{Acknowledgments} We are grateful for the assistance of Ken Freeman and Lister Staveley-Smith in this project, especially for their assistance with observations. We would like to thank Erwin de Blok for his help with various aspects of the \hi{} data reduction and interpretation. We would also like to thank Marilena Salvo and Gayandhi de Silva for their observing assistance. Our thanks also go to the anonymous referee for their useful comments, especially regarding the \hi{} spectra. The 2.3-meter Telescope is run by the Australian National University as part of Research School of Astronomy and Astrophysics. The Australia Telescope Compact Array and the Parkes Radio Telescope are part of the Australia Telescope that is funded by the Commonwealth of Australia for operation as a National Facility managed by CSIRO. This research has made use of the NASA/IPAC Extragalactic Database (NED), which is operated by the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration. The Digitized Sky Survey (DSS) was produced at the Space Telescope Science Institute under U.S. Government grant NAG W-2166, based on photographic data obtained using the UK Schmidt Telescope. \clearpage
Title: The late time evolution of Gamma-Ray Bursts: ending hyperaccretion and producing flares
Abstract: We consider the properties of a hyperaccretion model for gamma-ray bursts (GRBs) at the late time when the mass supply rate is expected to decrease with time. We point out that the region in the vicinity of the accretor and the accretor itself can play an important role in determining the rate and time behavior of the accretion and ultimately the energy output. Motivated by numerical simulations and theoretical results, we conjecture that the energy release can be repeatedly stopped and then restarted by the magnetic flux accumulated around the accretor. We propose that the episode or episodes when the accretion resumes correspond to X-ray flares discovered recently in a number of GRBs.
https://export.arxiv.org/pdf/astro-ph/0601272
\title[] {The late time evolution of Gamma-Ray Bursts: ending hyperaccretion and producing flares} \author[] {\parbox[]{6.in} {Daniel Proga$^1$ and Bing Zhang$^1$ \\ \footnotesize $^1$ Department of Physics, University of Nevada, Las Vegas, NV 89154, USA, e-mail: dproga@physics.unlv.edu; bzhang@physics.unlv.edu} \date{Accepted . Received ; in original form }} \label{firstpage} \begin{keywords} accretion, accretion discs -- gamma rays: bursts -- methods: numerical -- MHD \end{keywords} \section{Introduction} Gamma-Ray Bursts (GRB) are generally believed to be powered by hyperaccretion onto a compact, stellar mass object. The total amount of the available fuel is considered to be the key factor determining the burst duration. Within merger scenarios for short-duration GRBs, a neutron star (NS) is accreted onto another NS or onto a stellar mass black hole (BH; e.g, Paczy\'nski 1986, 1991; Eichler et al. 1989; Narayan et al. 1992; Fryer et al. 1999). Within the collapsar model for long-duration GRBs, up to 20 $\MSUN$ of a stellar envelope collapses onto the star's core which is a NS or a BH (e.g., Woosley 1993; Paczy\'nski 1998; MacFadyen \& Woosley 1999; Popham, Woosley, \& Fryer 1999; Proga et al. 2003). For short- and long-duration events, the accretion rate, $\MDOT_a$ must be of order of 1~$\MSUN~{\rm s^{-1}}$, yielding a duration of less than a few seconds for the former and a duration as long as tens to hundreds of seconds for the latter. These duration estimates are made under the assumption that all the available fuel is accreted during the GRB activity at a time-averaged constant rate. Recent GRB observations obtained with {\it Swift} motivate us to review the above assumption and some other aspects of GRB models. In particular, early X-ray afterglow lightcurves of nearly half of the long-duration GRBs show X-ray flares (Burrows et al. 2005; Romano et al. 2006; Falcone et al. 2006). X-ray flares are also found to follow the short-duration GRB 050724 (Barthelmy et al. 2005) whose host galaxy is early-type, which is consistent with the merger origin. The flares generally rise and fall rapidly, with typical rising and falling time scales much shorter than the epoch when the flare occurs. This time behavior strongly supports the ``internal'' origin of the flares (Burrows et al. 2005; Zhang et al. 2006; Fan \& Wei 2005), in contrast to the ``external'' origin of the power-law decay afterglows. The internal model not only offers a natural interpretation of the rapid rise and decay behavior of the flares, but also demands a very small energy budget (Zhang et al. 2006). Within this picture, the data require a {\it restart} of the GRB central engine (i.e., a restart of accretion). Fragmentations in the collapsing star (King et al. 2005) or in the outer parts of the accretion disc (Perna et al. 2006) have been suggested to be responsible for the observed episodic flaring behavior. These two flare models appeal to one of the basic ingredients of an accretion powered engine -- the mass accretion rate -- and conjecture that the episodic energy output is driven by changes in the mass supply and subsequently accretion rate. In this picture, the inner part of the accreting system {\em passively} responds to changes in the accretion flow at larger radii. Here, we point out that the region in the vicinity of the accretor and the accretor itself can play an important role of determining the rate and time behavior of the accretion and the energy output. In particular, we conjecture that the energy release can be repeatedly stopped and then restarted, provided the mass supply rate decreases with time even if the decrease is smooth. For both merger and collapsar GRB models, a decrease of the mass supply rate is expected, especially in the late phase of activity, because the mass density decreases with increasing radius. In our model, we appeal to the fact that, as mass is being accreted onto a BH, the magnetic flux is accumulating in the vicinity of the BH. Eventually, this magnetic flux must become dynamically important and affect the inner accretion flow, unless the magnetic field is very rapidly diffused. In the remaining part of the paper we list and discuss theoretical arguments and results from a variety of numerical magnetohydrodynamic (MHD) simulations of accretion flows that support our model. We also provide analytic estimates to show that our model can quantitatively account for the observed features of the flares. \section[]{Magnetic model for GRBs and their flares} \subsection{Insights from numerical models} Generally, our model for the flares is based on the results from the numerical simulation of an MHD collapsar model for GRBs carried out by Proga et al. (2003) and the results from a number of simulations of radiatively inefficient accretion flows (RIAFs) onto a BH (Proga \& Begelman 2003, PB03 hereafter; Igumenshchev, Narayan, \&Abramowicz 2003, INA03 hereafter). Proga et al. (2003) performed time-dependent axisymmetric MHD simulations of the collapsar model. These MHD simulations included a realistic equation of state, neutrino cooling, photodisintegration of helium, and resistive heating. The progenitor was assumed to be spherically symmetric but with spherical symmetry broken by the introduction of a small, latitude-dependent angular momentum and a weak split-monopole magnetic field. The main conclusion from the simulations is that, within the collapsar model, MHD effects alone are able to launch, accelerate and sustain a strong polar outflow. The MHD outflow provides favorable initial conditions for the subsequent production of a baryon-poor fireball (e.g., Fuller, Pruet \& Abazajian 2000; Beloborodov 2003; Vlahakis \& K$\ddot{\rm o}$nigl 2003; M\'{e}sz\'{a}ros 2002), or a magnetically dominated ``cold fireball'' (Lyutikov \& Blandford 2002), though the specific toroidal magnetic field geometry Proga et al. derived differs from some of these models (e.g., Vlahakis \& K$\ddot{\rm o}$nigl 2003; Lyutikov \& Blandford 2002). The latest Swift UV-Optical Telescope (UVOT) observations indicate that the early reverse shock emission is generally suppressed ( Roming et al. 2005), which is consistent with the suggestion that at least some GRBs are Poynting-flux-dominated outflows (Zhang \& Kobayashi 2005). To study the extended GRB activity, one would like to follow the collapse of the entire star. However, such studies are beyond current computer and model limits. Therefore, we explore instead the implications of the published simulations and consider the physics of the collapsing star to infer the properties and physical conditions in the vicinity of a BH during the late phase of evolution, i.e., when a significant fraction of the total available mass is accreted. The long time evolution of axisymmetric MHD accretion flows was studied by PB03 who explored simulations very similar to those performed by Proga et al. (2003) but with much simpler micro physics (i.e., an adiabatic equation of state, no neutrino cooling or photodisintegration of helium). Proga et al. (2003) found that despite the more sophisticated micro physics of the MHD collapsar simulations the flow cooling is dominated by advection not neutrino cooling. As a result, the early phase of the time evolution, and the dynamics of the innermost flow, are very similar in both the RIAF simulations and the collapsar simulations. In particular, after an initial transient behavior, the flow settles into a complex convolution of several distinct, time-dependent flow components including an accretion torus, its corona and outflow, and an inflow and outflow in the polar funnel (see the left panel in Fig. 1 for a schematic picture of such a flow). The accretion through the torus is facilitated by the magnetorotational instability (MRI, e.g., Balbus \& Halwey 1991) which also dominates the overall dynamics of the inner flow. In the remaining part of the paper, we will assume that the late evolution of the MHD collapsar simulations is similar to the late evolution of the RIAFs simulations. This assumption is justifiable because the flows in the collapsar and RIAFs simulations are similar during the early phase of the evolution (i.e., their dymanics and cooling are dominated by MRI and advection, respectively) The late evolution of RIAFs shows that the torus accretion can be interrupted for a short time by a strong poloidal magnetic field in the vicinity of a BH. This result is the main motivation for this paper, as it shows that the extended GRB activity may be a result of an accretion flow modulated by the ``magnetic-barrier'' and gravity. Because this barrier halts the accretion flow intermittently (see Figs.~6 \& 8 in PB03), the accretion rate is episodic (see Fig.3 of PB03). This potentially gives a natural mechanism for flaring variability in the magnetic-origin models of GRBs as we first mentioned in Fan, Zhang \& Proga (2005; see the middle panel of Fig. 1 here, for a cartoon picture of the accretion halted by the magnetic-barrier.) The importance of accumulating of the magnetic flux has been explored and observed by others in various astrophysical contexts (e.g., Bisnovatyi-Kogan \& Ruzmaikin 1974, 1976; Narayan, Igumenshchev \& Abramowicz 2003; INA03). In particular, INA03 carried out a three-dimensional (3D) MHD simulation (their model B) to late model times. They found that the magnetic flux accumulates, initially near the BH and then farther out, and the field becomes dynamically dominant. At late times, mass is able to accrete only via narrow streams, in a highly nonaxisymmetric manner (see also Narayan et al. 2003). The main difference between PB03's and INA03's results is the extent and duration of the magnetic dominance. In PB03, the magnetic dominance is a {\em transient} whereas in INA03 is a {\em persistent} state. The reason for this difference is the treatment of the magnetic field: for the initial conditions, PB03 used the split-monopole magnetic field and any changes in the magnetic flux near the BH during the evolution are due to the chaotic, small-scale fields generated in the disc. The detailed analysis show that the disc properties in PB03's simulations are determined by MRI. In particular, MRI is responsible for the complex field structure and for the disc toroidal field being one or even two orders of magnitude higher than the poloidal field (see figs. 9 and 10 in PB03 and fig. 3 in Proga et al. 2003). On the other hand, in their model B, INA03 set up a poloidal field configuration in the injected gas in such a way that the portion of the material that accretes always carries in the same sign of the vertical component of the magnetic field. The simulations carried out by PB03 and INA03 differ also in the assumed geometry (axisymmetric versus fully 3D). INA03 and PB03 do not explore all cases including the case where the external or initial field has zero net flux or the field with the poloidal component changing sign on length scales much smaller than the size of the mass reservoir \footnote{ In the case where the initial or external flux has zero-net flux, a large scale coherent field might in some circumstances be generated by MRI (e.g., Livio, Pringle, \& King 2003). If so the central magnetic flux could vary with time but still be dynamical signifacant for some periods of time.}. Additionally, these simulations also do not give definitive answers to the problems for which they were designed. Nevertheless, they give interesting insights into the general problem of MHD accretion flows. In particular, they suggest that magnetic fields can provide an important parameter determining the time scale for the accretion; i.e., it can be significantly longer than the local dynamical time scale. This can have important implications for the observed X-flares in GRBs, as we argue here, and for X-ray spectral states for BH binaries as discussed by Spruit \& Uzdensky (2005, SU05 hereafter). In fact, the work by SU05 describes very well the general physics and theory of magnetic flux accumulated by an accretion flow. Therefore we now turn our attention to some theoretical aspects of the problem as presented by SU05. \subsection{Theory of the magnetic barrier and accretion flow} SU05 considered a new mechanism of efficient inward transport of the large-scale magnetic field through a turbulent accretion disc. The key element of the mechanism is concentration of the external field into patches of field comparable in strength to the MRI turbulence in the disc. They focused on how to increase the magnetic flux at the center in the context of BH binaries. In particular, they argue that the capture of external magnetic flux by accretion disc and its subsequent compression in the inner regions of the disc may explain both changes in the radiation spectrum and jet activity in those objects. However, their model and physical arguments are generic and applicable to our problem. One can expect that as the strength of the magnetic field increases at the center, the field may eventually suppress MRI turbulence and reduce the mass accretion rate and the power in the outflow. This should be the case especially for GRBs because the mass inflow rate at the late time is most likely much lower than at the early time. The disc may become a Magnetically-Dominated Accretion Flow (MDAF) as proposed by Meier (2005) or the fields in the polar funnel can expand toward the equator and reconnect as in PB03's simulations. In the latter, the torus is pushed outward by the magnetic field. At this time, the gas starts to pile up outside the barrier; eventually it can become unstable to interchange instabilities at the barrier outer edge as suggested by SU05 or the gas in the torus can squash the magnetic field (compare Fig.~5 and 6 in PB05 or the middle and right panel in Fig.1 here). When interchange instabilities operate, magnetic flux from the bundle mixes outward into the disc while the disc material enters the barrier. In the accretion disk context, interchange instabilities have been studied by a few authors (e.g., Spruit et al. 1995; Lubow \& Spruit 1995; Stehle 1996; Stehle \& Spruit 2001; Li \& Narayan 2004). These studies showed that the onset of small-scale modes typical of interchanges (as in Rayleigh-Taylor instabilities) takes place only at rather large field strengths, due to a stabilizing effect of the Keplerian shear. The interchange instability operates at moderate field strengths, but only at low shear rates (less than Keplerian). However for most of the time, we expect high shear rates in a torus because a low shear torus quickly becomes Keplerian due to MRI (e.g., PB03 and Proga et al. 2003). We note that SU05 interpreted INA03 accretion through the barrier, in the form of blobs and streams as a product of interchange instabilities. SU05 also suggested that the field strength at which these instabilities become effective is most usefully expressed in terms of the degree of support against gravity provided by the magnetic stress $B_R B_Z$. According to SU05, the instabilities become effective when the radial magnetic force, $F_m\sim 2 B_R B_Z/4\pi$, is of the order of a few percent of the gravitational force, $F_g=GM\Sigma/R^2$, where $M$ is the central mass, $R$ is the radius, and $\Sigma$ is the surface density. For $B_R \approx B_Z$, there is a range in field strengths between the value at which MRI turbulence is suppressed and the value where dynamical instability of the barrier itself sets in, where no known instability operates (Stehle \& Spruit 2001). In this range, the disk material cannot mix or penetrate the magnetic field accumulated at the center (e.g., the middle panel of Fig. 1). Instead, mass builds up outside a region with such field strengths until the magnetic field at the center is compressed enough for instability to set in. Thus, both numerical work and theoretical models of magnetized accretion flows show that the inner most part of the flow and accretor can respond {\em actively} to changes of the accretion flow at larger radii. In particular, the inner most accretion flow can be halted for a very long time as shown by INA03 or it can be repeatedly halted and reactivated as shown in PB03. \subsection{Analytic estimates} We finish this section with order-of-magnitude estimates of a few key features of our X-ray flare model. We start by estimating the strength and flux of magnetic field required to support the gas. The gas of the surface density, $\Sigma$ can be supported against gravity by the magnetic tension if $F_g\sim F_m$. The surface density can be estimated from $\Sigma=\MDOT /2\pi R \epsilon v_{ff}~{\rm g~cm^{-2}}$, where $\epsilon v_{ff}$ is the flow radial velocity assumed to be a fraction $\epsilon$ of the free fall velocity, $v_{ff}$. Assuming $B_r \approx B_z=B$, the force balance yields the field strength $B \sim 2\times 10^{16}~\epsilon_{-3}^{-1/2} r^{-5/4} \MDOT_1^{1/2} M_3^{-1} $~G, where $\epsilon_{-3} \equiv 10^{3} \epsilon$, $r \equiv R/R_S=R/(2GM_{BH}/c^2)$, $\MDOT_1=\MDOT/1~\MSUN~{\rm s^{-1}}$, and $M_3=M/3 \MSUN$. We estimate the magnetic flux as $\Phi\sim\pi r^2 R_S^2 B(r)= 5\times10^{28}~\epsilon_{-3}^{-1/2}r^{3/4}\MDOT_ 1^{1/2} M_3~{\rm cm^2~G}$ from which we obtain an estimate to the magnetospheric radius $r_m \approx 60~\epsilon_{-3}^{2/3} \MDOT_1^{-2/3} M_3^{-4/3} \Phi_{30}^{4/3}$, where $\Phi_{30}\equiv \Phi/(10^{30}~{\rm cm^2 G})$. Substituting the expression for $B$ into the expression for the surface density, one finds that a given magnetic flux can support the gas with the surface density of $\Sigma_B=5\times10^{19}~\Phi_{30}^2 M_3^{-3}r^{-2}~{\rm g~cm^{-2}}$. To stop accretion with the hyper rate of $1\MSUN~{\rm s^{-1}}$ onto a 3$\MSUN$ black hole at r=3 (i.e., for $r_m$ to be 3), the magnetic flux of order $\Phi_{30} \sim 0.11$ is required. We now assume that such a magnetic flux is accumulated during hyperaccretion and that it does not change with time. Under these assumptions, $r_m=300$ for the mass supply rate of $10^{-3}~\MSUNYR$ representative of the late time evolution . This relatively large radius demonstrates one of our key points that the innermost part an accreting system can actively respond, via magnetic fields, to changes in the inflow at large radii. To estimate the conditions needed to restart accretion, the accretion energetics and related time scales, we ask what is the mass of a disc with $\Sigma$ high enough to reduce $r_m$ from 300 to 3 or so. To answer this question, we adopt Popham et al.' (1999) model of neutrino-dominated discs. Popham et al. assumed that neutrino cooling produces a thin disc (Shakura \& Sunyaev 1973) for accretion rates require to power GRBs. Using the disc solution for the density and height (eqs. 5.4 and 5.5 in Popham et al. 1999), we can express the disc surface density as $\Sigma_\alpha=1.8\times10^{19}~\alpha^{-1.2} M_3^{-0.8}\MDOT_1r^{-1.25}$~g, where $\alpha$ is the dimensionless parameter scaling the stress tensor and the gas pressure (Shakura \& Sunyaev 1973). Equating $\Sigma_B$ with $\Sigma_\alpha$, one can estimate the mass accretion rate of an $\alpha$ disc and compute $M_D$ by integrating $\Sigma_\alpha$ over radius. For $\Phi_{30}=0.11$ and $\alpha=10^{-2}$ the accretion rate through the $\alpha$ disc is $0.03~\MSUN~s^{-1}$ and $M_D$ for $r$ between 3 and 300 is 0.32~$\MSUN$. This mass accretion rate is more than one order of magnitude lower than the rate of $\sim 1~\MSUN~s^{-1}$ typical for the early time evolution. Thus, our estimates are consistent with the fact that the X-ray flare luminosity is at least one or two orders of magnitude lower the prompt gamma-ray emission (see section 3). If this disc mass is a result of slow mass accumulation during the late evolutionary stage, then it will take about 400 s to rebuild the disc for the mass supply rate of $10^{-3}~\MSUN~s^{-1}$ and 12 s to accrete all this mass at the disc accretion rate of $0.03~\MSUN~s^{-1}$. The latter is a lower estimate for the flare duration because, for simplicity, we assumed a relatively high, {\em constant} disc accretion rate. It is very likely that the rate changes with time as the shape of the light curve during the flares indicates. In our model, the mass supply rate controls the epochs when the flares happen: the disc is rebuilt on the time scale which increases with time because the mass supply slowdowns. Additionally, the flare duration is coupled to the epoch through the mass of the rebuilt disc. Thus our model is capable of accounting for the observed duration - time scale correlation. \section{Discussion and conclusions} The detailed analysis of the X-ray flares revealed that they generally have lower luminosities (by at least one or two orders of magnitude) than the prompt gamma-ray emission. Additionally, the total energy of the flare is also typically smaller than that of the prompt emission, although in some cases both could be comparable (e.g. for GRB 050502B, Falcone et al. 2006). Moreover multiple flares are observed in some GRBs and the durations of these flares seem to be positively correlated with the epochs when the flares happen, i.e. the later the epoch, the longer the duration (O'Brien et al. 2005; Falcone et al. 2006; Barthelmy et al. 2005). The flare analysis also showed that the later the epoch the lower the flare luminosity. The above qualitative properties of the flares provide important constraints on models of them. Perna et al.'s (2006) disc fragmentation model promises to account for the duration - time scale correlation and the duration - peak luminosity anticorrelation. However, the physical process or processes causing fragmentation are uncertain. It is also uncertain that the conditions for the disc fragmentation are met in GRB progenitors. This seems to be the case especially for the collapsar model as a relatively high rotation of the progenitor is required. We also note that magnetic fields can suppress or even prevent disc fragmentation (e.g., Banerjee \& Pudritz 2006). Here, we propose that the X-ray flares in GRBs are consequences of the fact that during the late time evolution of a hyperaccretion system the mass supply rate should decrease with time while the magnetic flux accumulating around a BH should increase. In particular, we point out that the flux accumulated during the main GRB event can change the dynamics of the inner accretion flow. We argue that the accumulated flux is capable of halting intermittently the accretion flow. In our model, the episode or episodes when the accretion resumes correspond to X-ray flares. A comparison of our analytic estimates from Section 2.3 with the observed X-ray flare characteristics, shows that our model is not only physically based but also can both qualitatively and quantitatively account for some aspects of the flares -- such as the peak times. In general, our model fits under the general label of the magnetic jet model for GRBs as we appeal to the magnetic effects to play the key role not only during the main event but also during the late evolution. The importance of the magnetic effects for the X-ray flares can be argued based on energy budget of the accretion model (Fan et al. 2005). The X-ray flares discovered in GRBs are relatively new and unexpected phenomena. They give a strong incentive to apply the existing models of hyperaccretion systems to circumstances where the mass supply is reduced. Studies of this kind should reveal whether one needs to introduce additional physics in order to explain the flares. If so one should explore the effects of this on the early evolution of GRBs and check whether they are consistent with GRBs observations. Our X-ray flare model has the advantage that it is essentially the same as the MHD collapsar model for GRBs, with only one justifiable change in a key physical property of the collapsar model: a decrease of the mass supply rate with time. \section*{Acknowledgments} We thank D. Meier, D. Uzdensky, and a referee for useful comments. This work is supported by NASA under grants NNG05GB68G (DP) and NNG05GB67G (BZ). \bsp
Title: Silicate Emission in the Spitzer IRS spectrum of FSC 10214+4724
Abstract: We present the first MIR spectrum of the z=2.2856 ultraluminous, infrared galaxy FSC 10214+4724, obtained with the Infrared Spectrograph onboard the Spitzer Space Telescope. The spectrum spans a rest wavelength range of 2.3-11.5 microns, covering a number of key diagnostic emission and absorption features. The most prominent feature in the IRS spectrum is the silicate emission at rest-frame 10 microns. We also detect an unresolved emission line at a rest wavelength of 7.65 microns which we identify with [NeVI], and a slightly resolved feature at 5.6 microns identified as a blend of [Mg VII] and [Mg V]. There are no strong PAH emission features in the FSC 10214+4724 spectrum. We place a limit of 0.1 micron on the equivalent width of 6.2 micron PAH emission but see no evidence of a corresponding 7.7 micron feature. Semi-empirical fits to the spectral energy distribution suggest about 45% of the bolometric luminosity arises from cold 50 K dust, half arises from warm (190 K) dust, and the remainder, 5%, originates from hot (640 K) dust. The hot dust is required to fit the blue end of the steep MIR spectrum. The combination of a red continuum, strong silicate emission, little or no PAH emission, and no silicate absorption, makes FSC 10214+4724 unlike most other ULIRGs or AGN observed thus far with IRS. These apparently contradictory properties may be explained by an AGN which is highly magnified by the lens, masking a (dominant) overlying starburst with unusually weak PAH emission.
https://export.arxiv.org/pdf/astro-ph/0601061
\title{Silicate Emission in the {\it Spitzer}$^1$\ IRS$^2$\ spectrum of FSC 10214+4724} \altaffiltext{1}{based on observations obtained with the {\it Spitzer Space Telescope}, which is operated by JPL, California Institute of Technology for the National Aeronautics and Space Administration} \altaffiltext{2}{The IRS is a collaborative venture between Cornell University and Ball Aerospace Corporation that was funded by NASA through JPL.} \author{H. I. Teplitz\altaffilmark{3}, L. Armus\altaffilmark{3}, B.T. Soifer\altaffilmark{3}, V. Charmandaris\altaffilmark{4,5,6}, J. A. Marshall\altaffilmark{4}, H. Spoon\altaffilmark{4}, C. Lawrence\altaffilmark{7}, L. Hao\altaffilmark{4}, S. Higdon\altaffilmark{4}, Y. Wu\altaffilmark{4}, M. Lacy\altaffilmark{3}, P. R. Eisenhardt\altaffilmark{7}, T. Herter\altaffilmark{4}, J.R. Houck\altaffilmark{4} } \altaffiltext{3}{Spitzer Science Center, MS 220-6, Caltech, Pasadena, CA 91125. hit@ipac.caltech.edu} \altaffiltext{4}{Astronomy Department, Cornell University, Ithaca, NY 14853} \altaffiltext{5}{Chercheur Associ\'e, Observatoire de Paris, F-75014, Paris, France} \altaffiltext{6}{University of Crete, Dept. of Physics, GR-71003 Heraklion, Greece} \altaffiltext{7}{Jet Propulsion Laboratory, California Institute of Technology, 4800 Oak Grove Drive, Pasadena, CA 91109} \keywords{ cosmology: observations --- galaxies: evolution --- galaxies: high-redshift --- galaxies: individual (FSC 10214+4724) } \section{Introduction} FSC 10214+4724, at a redshift of $z=2.2856$\ \citep{Rowan-Robinson 1991}, was initially thought to be the most luminous object in the Universe, but was later revealed to be gravitationally lensed by a foreground galaxy \citep{Broadhurst 1995, Graham 1995, Serjeant 1995, Eisenhardt 1996}. The lensing model of \cite{Eisenhardt 1996}\ suggests that the central (optical/UV) source is magnified by a factor of $\sim 100$, and that the lensing arc is an image of the central $\sim 0^{\prime \prime}.005$\ (40 pc) of the source at B-band rest wavelength. They further conclude that the bolometric luminosity of the source is produced over a larger region (240 pc), implying a bolometric magnification factor of 30 and an intrinsic luminosity of $\sim 2 \times 10^{13}$\ \Lsun\ , placing it in the class of ultraluminous infrared galaxies (ULIRGs). The large magnification makes FSC 10214+4724 a unique subject for the study of high redshift ULIRGs, offering greater effective sensitivity and spatial resolution than possible in observations of unlensed sources. A central question in understanding FSC 10214+4724 is the relative contribution of starburst and AGN components to its luminosity. The rest-frame UV-optical spectrum appears to be similar to that of Seyfert 2 galaxies \citep{Elston 1994}. Strong UV polarization shows that much of the UV continuum results from scattered light, and broad lines in the polarization spectrum identify the presence of an AGN \citep{Lawrence 1993, Goodrich 1996}. The bulk of the luminosity, however, is emitted in the infrared \citep[as much as 99\%,][]{Rowan-Robinson 1991}. CO observations point to a large reservoir of molecular gas \citep{Solomon 1992, Scoville 1995}. Sub-mm and millimeter detections \citep{Downes 1992, Rowan-Robinson 1993,Benford 1999} suggest substantial emission from cold dust, usually associated with extended star formation. Recent {\it Chandra}\ observations show the object to have weak X-ray emission, consistent with vigorous star formation or a Compton-thick AGN \citep{Alexander 2005}. Lens models suggest differential amplification, so that the central (AGN-dominated) region is more highly magnified than the surrounding starburst \citep[e.g.,][]{Eisenhardt 1996, Lacy 1998}. The sensitivity and large wavelength coverage of the {\it Spitzer}\ Infrared Spectrograph \citep[IRS][]{Houck 2004}\ makes it possible to explore the dust emission and absorption features in the rest-frame mid-infrared spectra of dusty galaxies at low and high redshift, and thus identify the power sources which may be hidden in the UV and optical. In this paper, we present the first MIR spectrum of FSC 10214+4724. We describe the observations and data reduction in Section 2, present the spectrum and a dust emission model fit to the spectral energy distribution in Section 3, and discuss the implications in Section 4. Throughout, we assume a $\Lambda$-dominated flat universe, with $H_0=70$\ km s$^{-1}$\ Mpc$^{-1}$, $\Omega_{\Lambda}=0.7, \Omega_{m}=0.3$. \section{Observations and Data Reduction} Spectra of FSC 10214+4724 were obtained with the IRS on 19 April 2004. The data were taken in the first order of the low resolution, short wavelength module (SL-1; 7.5-14.2 \mic) and in both orders of the low resolution, long wavelength module (LL-1 and LL-2, 14.2-21.8 and 20.6-38 \mic, respectively). The spectral resolution varies from 60 to 120 across each order. Individual ramp durations were 60 seconds in SL-1 and 120 in LL-1 and LL-2. Spectra were taken in the standard ``staring'' mode, with four exposures at each of two positions, separated by one third of the slit length. A total of 8 individual spectra were taken in each sub-slit, for total on-source integration times of 480, 960, and 960 seconds. Spectra were reduced using the S11 pipeline at the {\it Spitzer}\ Science Center, which includes ramp fitting, dark sky subtraction, droop correction, linearity correction, and wavelength calibration. One-dimensional spectra were extracted from the un-flatfielded two-dimensional spectra using the SMART data reduction package \citep{Higdon 2004}. The extractions are then flux calibrated with the IRS spectrum of a star ($\alpha$\ Lac), extracted in an identical manner using SMART. The data have been sky-subtracted by differencing the two nod positions along the slit, before spectral extraction. As a final step, we have scaled the SL-1 and LL-2 1D spectra by 3\% and 8\%, respectively, to match the LL-1 spectrum in the overlap region and produce a single low-resolution spectrum from $7.5-38\mu$m (observed frame). Mid-IR photometry of FSC10214+4724 at 3.6, 4.5, 5.8 and 8.0 $\mu$m was obtained on 20 May 2004 using the Infrared Array Camera \citep[IRAC;][]{Fazio 2004}. The galaxy was observed in full array mode with a cyclic 5 point dither pattern resulting in a total on source integration time of 1 minute per filter. The source was unresolved and its Full-Width at Half Maximum (FWHM) varied between 1.5\arcsec and 2\arcsec. We performed photometry on the final mosaics produced by the SSC pipeline (S11.0.2) using an aperture of 3.6 arcseconds radius following the method described in the IRAC data Handbook. The resulting flux densities are accurate to $<$5\%. No attempt was made to correct the IRAC photometry for the contribution of the lensing galaxy which is unresolved from the source. \section{Results} We plot the IRS spectrum of FSC 10214+4724 in Figure \ref{fig: spectrum}. The most prominent feature is a steep rise at the red end ($\sim 8-10$\ \mic\ rest wavelength). Extending the spectral energy distribution (SED) with the IRAS 60 and 100 \mic\ photometry \citep{Moshir 1990} makes clear that there is not simply a steeply rising continuum, but rather is a broad emission feature on top of a somewhat less steep continuum. We identify this feature as silicate emission, centered at $\sim 10$\ \mic. It is difficult to measure an accurate equivalent width for the feature, because the IRS wavelength coverage does not extend far enough to the red. Three weak emission lines are also present. We identify the narrow emission line at rest wavelength 7.65 \mic\ as the [Ne VI] fine structure line. This line has a rest-frame equivalent width (EW) of 0.02 \mic, roughly comparable to that observed in low redshift Seyferts \citep{Sturm 1999, Lutz 2000}. The emission feature observed at $\sim 5.5$\ \mic\ in the rest-frame appears to be marginally resolved. We identify it as a blend of [Mg VII] and [Mg V] at 5.503 and 5.610 \mic, respectively. These lines have been seen in nearby AGN \citep{Sturm 2002}\ and have ionization potentials of about 186 and 105 eV, similar to the 158 eV ionization potential of [Ne VI]. While there is also a rotational transition of H$_{2}$ at $5.51\mu$m (the S(7) line) which is often quite strong in ULIRGs, this feature is always accompanied by much stronger emission from the other, lower transition, H$_{2}$ rotational lines, which are not present. The rest-frame equivalent width of the [Mg VII] and [Mg V] lines combined is $\sim 0.03\pm 0.01$\ \mic. While uncertain, this EW appears to be a factor of 3-5 higher than in local AGN \citep{Sturm 2002}. The EW of the Ne and Mg lines is a factor of a few higher than the strongest of the narrow, high-excitation emission lines (C IV, He II, [Ne IV]) seen in the rest-frame UV spectrum of FSC 10214+4724 \citep{Rowan-Robinson 1991, Goodrich 1996}. There is a marginal detection of broad emission at 6.2 \mic\ rest-frame, corresponding to the wavelength of polycyclic aromatic hydrocarbon (PAH) emission. Although the feature is clearly visible in Figure \ref{fig: spectrum}, examination of the individual extractions shows that it is more prominent in LL-2 than in LL-1. The observed wavelength (20.3 \mic) places the feature near the noisy blue end of LL-1, but at a wavelength that is usually regarded as reliable. We take the measurement, $EW \sim 0.1$\ \mic\ in the rest-frame, as an upper limit. We also note that the corresponding 7.7 and 8.6 \mic\ PAH emission features are not seen, despite being redshifted into a clean part of the spectrum and the fact that they are usually 1.5--2 times stronger than the 6.2 \mic\ feature in starburst-dominated low redshift ULIRGs. The low ratio of 7.7 to 6.2 \mic\ PAH equivalent width, while highly unusual, is not impossible under certain conditions \citep{Kessler-Silacci 2005}. The 11.3 \mic\ PAH feature might also be expected, but that line falls near the red end of LL-1 where the noise precludes a meaningful limit. Nonetheless, the lack of other PAH lines indicates that the 6.2 \mic\ feature should be treated with caution. \section{Discussion} The shape of the SED of FSC 10214+4724 is dominated by dust at a variety of temperatures. Substantial cold dust must be present, given the strong emission at long wavelengths \citep[$> 40$\ \mic, rest-frame; e.g., ][]{Rowan-Robinson 1993}. At the same time, warmer dust will be required to explain the 5-10 \mic\ continuum. Dust warmer than 100 K can produce the silicate emission \citep[e.g.,][]{Li 2001}, but a hot dust component (several hundred K) is required to explain the $\sim 5$\ \mic\ continuum. The rest-frame near-infrared (NIR) will have a contribution from both star light and hot dust. We have fit the SED with a multi-component model which includes three graphite and silicate dust grain components and a 3500K blackbody stellar component \citep{Marshall 2006}. The three dust components in the model are not meant to represent three distinct physical structures, but rather they are indicative of the average temperature ranges within the source. The model was fit to the SED from observed frame NIR to millimeter wavelengths (see Figure \ref{fig: SED fit}. NIR data included the photometry of \cite{Soifer 1991}\ and the IRAC data described in Section 2. Longer wavelength data included IRAS photometry at 60 and 100 \mic, the 350 \mic\ detection of \cite{Benford 1999}, and the sub-mm and mm data of \cite{Rowan-Robinson 1993}\ and \cite{Downes 1992}. A minimum of three dust components are required to fit the FSC 10214+4724 SED, one at $\sim 50$\ K (cold), one at $\sim 190$\ K (warm), and one at $\sim 640$\ K (hot). Their relative contributions to the bolometric luminosity are given in Table \ref{tab: SED fit} . Note that hot dust contribution is an upper limit, because it is dominated by the IRAC data and no correction has been made for contamination by the foreground lensing galaxy. The stellar emission is a negligible contribution to the bolometric luminosity, but is needed to fit the shortest wavelength data. component. The cold dust component ($51 \pm 6$\ K) is in good agreement with the estimate of \cite{Benford 1999}, 55 K. Each component contains a distribution of grains with different equilibrium temperatures depending on their size and composition. Additionally, each component contains emission from dust at different radial distances, and therefore temperatures, from the illuminating source. We define the characteristic temperature of a component to be the temperature of the most luminous grain size at the distance from the source contributing the majority of the luminosity. This luminosity dominating distance corresponds to a $\tau(UV) \sim 0.5$, where approximately half of the UV-source photons have been absorbed. Each component therefore contains dust above and below the characteristic temperature. With this definition, the characteristic temperature roughly corresponds to the expected peak in the dust modified (grey-body) Planck function. The observed SED of FSC 10214+4724 appears consistent with low redshift, AGN-dominated ULIRGs. Such sources can have cold dust components which account for up to 40\%\ of the bolometric luminosity, due to both AGN-heated cold dust far from the nucleus and the presence of a small underlying starburst \citep{Armus 2004, Armus 2006}. \cite{Rowan-Robinson 2000} also estimates that the AGN contributes most of the luminosity of FSC 10214+4724. In addition, FSC 10214+4724 shows silicate emission, similar to other AGN observed with the IRS, and a hot (graphite) dust continuum. However, the hot dust contribution is quite small compared to many AGN-dominated ULIRGs or QSOs. The only obvious fine-structure lines are [Ne VI], [Mg VII] and [Mg V], high-ionization species not observed in starburst galaxies. There is little if any PAH emission. The upper limit to the $6.2\mu$m EW is approximately a factor of $5-10$ lower than most pure starburst galaxies \citep{Brandl 2005}\ and ULIRGs dominated by star formation \citep{Armus 2004, Armus 2006}. However, differential magnification is likely to be enhancing the central AGN, reducing the PAH EW and making the silicate emission more obvious in the IRS spectrum. \cite{Eisenhardt 1996}\ estimate a magnification factor of $\sim 100$\ for the central 40 pc, but only a factor of $\sim 30$\ out to at least 240 pc. The intrinsic contribution of the starburst to the bolometric luminosity could be larger than that inferred by our current model. We assume that the warm and hot dust heated by the AGN are magnified by an additional factor of 3.3, but that both the cold dust heated by the AGN and any dust heated by starburst are not. The ratio of cold to warm dust in local AGN- and starburst-dominated ULIRGs is approximately 2:3 and 3:1, respectively (Armus et al. 2006, in preparation). Taking these assumptions together, we estimate the intrinsic AGN contribution to the bolometric luminosity of FSC 10214+4724 to be $\sim 35$\% after correction for the differential magnification. If the differential magnification factor is correct, this AGN contribution is probably an upper limit because AGN-dominated ULIRGs often have measurable starbursts. If the starburst contributes 65\%\ of the intrinsic luminosity, it is surprising that little or no PAH emission is seen. Taking our limit of 0.1 \mic\ $>EW_{\mbox{rest}}$\ and a differential magnification factor of 3.5 would place the limit of the EW for PAH emission within the range of other star-forming ULIRGs\citep[Armus et al. 2005, in prep;][]{Brandl 2005}. However, the uncertainty in our PAH measurement leaves this possibility unconfirmed; a lower signal to noise ratio would have made our limit higher. Furthermore, the lack of evidence for comparably strong PAH emission at 7.7 \mic\ makes it likely that the EW at 6.2 \mic\ is overestimated. Nonetheless, the limit demonstrates that substantial star formation can be present in FSC 10214+4724 and not visible in the (differentially magnified) IRS spectrum. In figure \ref{fig: SED compare}, we compare the rest frame MIR spectrum of FSC 10214+4724 to the IRS spectra of three other sources. The intrinsic spectrum of FSC 10214+4724 is redder than the observed spectrum, given the differential magnification, so the differences seen in the figure would be greater if corrected for lensing. The comparison sources are an AGN-dominated ULIRG (FSC 15307+3252), a silicate-emitting QSO (PG 1351+640) and a pure starburst galaxy (NGC 7714). PG 1351+640 has similar silicate emission, but has a much stronger hot dust continuum at $\sim 5$\ \mic\ \citep{Hao 2005}. In fact, the cold dust emission in FSC 10214+4724 gives it a much steeper mid-to-far infrared slope than most AGN with silicate emission. The 30 \mic\ to 6 \mic\ flux density ratio of 46 in FSC 10214+4724, is higher than the reddest of the quasars in \cite{Hao 2005}\ or \cite{Siebenmorgen 2005}, with a ratio of 11, and the reddest AGN in \cite{Weedman 2005}, NGC 1275, which has a ratio of 30. The ULIRG FSC 15307+325 has a similar MIR slope to FSC 10214+4724 in the 2-10 \mic\ region, but has silicate absorption rather than emission, and a much weaker cold dust continuum at longer wavelengths. The continuum of NGC7714 is very red with a 30 \mic\ to 6 \mic\ flux density ratio of 120, and a strong stellar contribution at the short wavelengths. In many AGN-dominated ULIRGs, the MIR spectrum is even bluer, with a strong continuum in the $\sim 5$\ \mic\ region \citep[e.g., ][]{Laurent 2000}. ULIRGs with steeper mid-infrared continua tend also to have silicate absorption, not emission. We see no evidence in the spectrum of FSC 10214+4724 for an underlying silicate absorption feature at 9.7 \mic. Furthermore, the profile of the silicate emission is consistent with that seen in most other AGN observed with the IRS \citep{Hao 2005, Siebenmorgen 2005, Weedman 2005}; although see \cite{Sturm 2005}\ for a counter example and discussion of the factors influencing the profile shape. While some absorption may be filled in with emission, the absorption and emission profiles are not necessarily identical. A coincidence of opacity, temperature, and grain mixture in the emitting and absorbing regions would be required. The AGN contribution to the bolometric luminosity in FSC 10214+4724 is apparently substantial but does not dominate. Nonetheless, the PAH emission expected for a starburst is weak or absent. An unusual dust geometry would be required for an AGN alone to explain the SED, given the amount of cold dust. The other available evidence points to a standard AGN configuration. The shape of the spectrum between $2-12\mu$m is qualitatively similar to dusty torus models \citep{Pier and Krolik 1992}\ with inner radius to height ratios of $a/h\sim 0.3$, and opening angles of $\sim 50$\deg -- consistent with a high-luminosity equivalent of Seyfert 2 galaxy, like NGC 1068. The similarlity to NGC 1068 hass previously been noted\citep[e.g.,][]{Barvainis 1995}. The presence of silicate emission in the IRS spectrum rules out models where the torus is seen edge-on, or where the AGN is completely obscured by foreground (cold) dust. \acknowledgements This work is based in part on observations made with the {\it Spitzer Space Telescope}, which is operated by the Jet Propulsion Laboratory, California Institute of Technology under NASA contract 1407. Support for this work was provided by NASA through an award issued by JPL/Caltech. \clearpage \begin{deluxetable}{lll} \tablecaption{SED fit parameters \label{tab: SED fit}} \tablehead{ \colhead{Dust Component} & \colhead{Temperature (K) } & \colhead{$L/L_{tot}$\ (\%)} } \startdata Hot & $638\pm20$ & 5.6 \\ Warm & $191\pm 2$ & 51.3 \\ Cold & $51\pm 6$ & 43.3 \enddata \end{deluxetable} \clearpage \clearpage \clearpage
Title: NLTE spectral analysis of GW Vir pulsators
Abstract: GW Vir variables are the pulsating members in the spectroscopic class of PG 1159 stars. In order to understand the characteristic differences between pulsating and non-pulsating PG 1159 stars, we analyse FUSE spectra of eleven objects, of which six are pulsating, by means of state-of-the-art NLTE model atmospheres. The numerous metal lines in the FUV spectra of these stars allow a precise determination of the photospheric parameters. We present here preliminary results of our analysis.
https://export.arxiv.org/pdf/astro-ph/0601566
\title{NLTE Spectral Analysis of GW~Vir Pulsators} \author{E. Reiff,$^1$ D. Jahn,$^1$ T. Rauch,$^{1}$ K. Werner,$^1$ and J. W. Kruk$^2$} \affil{$^1$Institut f\"ur Astronomie und Astrophysik, Universit\"at T\"ubingen, Germany \\ $^2$Department of Physics and Astronomy, The Johns Hopkins University, Baltimore, USA} \section{Introduction} \label{sec:introduction} GW~Vir variables \index{GW~Vir variables} belong to the spectroscopic class of the PG~1159 stars \index{PG~1159 stars} (Wesemael, Green \& Liebert 1985), which is named after the prototype \index{PG~1159$-$035} PG~1159$-$035. These objects are strongly hydrogen-deficient post-AGB stars \index{post-AGB stars} which pass through the hottest stage of stellar evolution. Their effective temperatures range between 75~000 and 200~000\,K, surface gravities vary from $\log g = 5.5-8.0\,\,[\mathrm{cm\,s}^{-2}]$. The so-called born-again scenario (a late thermal pulse which transferred these objects back to the AGB followed by a second post-AGB evolution, Iben et al\@. 1983) is mainly accepted as an explanation for the H-deficiency and can reproduce well the observed abundances. PG~1159 stars have spectra which are dominated by lines of \ion{He}{ii}, \ion{C}{iv}, and \ion{O}{vi} (Werner et al. 2004), their atmospheres show a typical surface composition of He:C:O = 33:50:17 by mass. Beside these main constituents there are several lines of trace elements, such as neon, nitrogen, silicon, sulfur, phosphorus, and fluorine (Reiff et al. 2005, Werner et al. 2005). Presently 37 PG~1159 stars are known, eleven of them proved to be pulsators. The pulsating members of the PG~1159 class are referred as GW~Vir variables. They are non-radial g-mode pulsators with periods from 300\,s up to 1000\,s, in some cases exceeding even 2000\,s (Nagel \& Werner 2004). The favored excitation mechanism for the pulsations is the $\kappa$-mechanism associated with cyclic ionization of carbon and oxygen (Quirion, Fontaine \& Brassard 2004). In the $\log T_{\mathrm{eff}} - \log g$ diagram the GW~Vir variables are located among the PG~1159 stars in the so-called GW~Vir instability strip. Spectral analyses of pulsating and non-pulsating PG~1159 stars were used by Dreizler \& Heber (1998) to define empirically the edges of this instability strip. But it is still puzzling that also non-pulsating PG~1159 stars are located within the instability strip. In our analysis we try to find more characteristic properties to distinguish between pulsating and non-pulsating members of this class. \section{Observations and First Results} \label{sec:observations_results} For our analysis we selected pulsating and non-pulsating PG1159 stars for which high resolution (R $\approx$ 20~000) FUV spectra obtained with the Far Ultraviolet Spectroscopic Explorer (FUSE) are available. The resulting sample comprises eleven objects. The FUSE spectra are processed within the standard Calfuse pipeline process. A log of all observations used for this analysis is listed in Table \ref{tab:log}. Besides the FUV spectra we also used spectra obtained with STIS, GHRS and IUE as well as optical spectra. The model atmospheres and synthetic line profiles are computed with the T\"ubingen Model Atmosphere Package (Werner et al\@. 2003, Rauch \& Deetjen 2003). The line-blanketed NLTE model atmospheres are in radiative and in hydrostatic equilibrium. Besides the main constituents of the atmospheres of PG~1159 stars, helium, carbon, and oxygen, our model atmospheres also contain neon and nitrogen. For the abundances of these elements we use atmospheric parameters taken from the literature which are summarized in Table \ref{tab:parameters}. For neon an abundance of 2\% mass fraction was assumed for all models, according to Werner \& Rauch (1994) and Werner et al\@. (2004). Although the abundances in the literature were mostly determined in analyses of optical spectra the synthetic spectra can also fit the FUV spectra well, which confirms the literature values for abundances in most cases. In Fig.\,\ref{fig:pg1159} we display the FUSE spectrum of PG\,1159$-$035 together with our synthetic spectrum. As lines of sulfur and silicon were also identified in several objects we included those elements in the synthetic spectra, too. Both were treated with line formation calculations without back-reaction on the atmospheric structure. We assumed solar abundances for both elements. \begin{table}[t] \caption{Log of the FUSE observations used for this analysis.} \label{tab:log} \footnotesize \begin{center} \begin{tabular}{l l r c} \noalign{\smallskip} \tableline \noalign{\smallskip} Object & Observation ID &\multicolumn{1}{c}{$t_\mathrm{exp}$} & Aperture \\ \noalign{\smallskip} \tableline \noalign{\smallskip} RX\,J2117.1+3412\index{RX\,J2117.1+3412} & P1320501 & 8232\,s & LWRS\\ PG\,1144+005\index{PG\,1144+005} & P1320201 & 6859\,s & LWRS\\ PG\,1520+525\index{PG\,1520+525} & P1320101 & 3648\,s & LWRS\\ PG\,1159$-$035\index{PG\,1159-035} & Q1090101 & 6321\,s & LWRS\\ K\,1$-$16\index{K\,1-16} & M1031010 &11271\,s & HIRS\\ HS\,2324+3944 \index{HS\,2324+3944} & P1320601 & 4004\,s & LWRS\\ Abell 78\index{Abell 78} & B1100101 & 9972\,s & LWRS\\ & B1100102 & 7894\,s & LWRS\\ NGC 7094\index{NGC 7094} & P1043701 &23183\,s & LWRS\\ Abell 43\index{Abell 43} & B0520202 &12150\,s & LWRS\\ PG\,1424+535\index{PG\,1424+535} & P1320301 &11132\,s & LWRS\\ PG\,1707+427\index{PG\,1707+427} & P1320401 &14599\,s & LWRS\\ \noalign{\smallskip} \tableline \end{tabular} \end{center} \end{table} Silicon is detectable in at least three objects, which are PG~1159$-$035, and the two cooler stars PG~1424+535 and PG~1707+427. Models with a solar Si abundance can fit the doublets at 1122/1128\,\AA\ and 1393/1402\,\AA\ well. In all spectra sulfur lines are detected, but our preliminary fits also suggest abundances less than solar. In Fig. \ref{fig:SiS_lines} we display part of the FUSE spectrum of PG~1424+535 with a preliminary fit of the sulfur and silicon lines, both with solar abundances. In former analyses by Dreizler \& Heber (1998) it was suggested that the nitrogen abundance is a characteristic difference between pulsating and non-pulsating PG~1159 stars, as nitrogen was detected in all GW~Vir pulsators with a rather high abundance of 1\% by mass, while in stable PG~1159 stars no nitrogen could be detected, except for PG~1144+005 (which is considered outside the instability strip). In order to confirm previously determined N abundances we tried to fit the N resonance doublet at 1238/1242\,\AA. For this purpose we also analysed the STIS spectrum of PG~1159$-$035, which has a high resolution (0.1\,\AA) and high S/N. In this spectrum the interstellar component of the resonance doublet is clearly separated from the photospheric component. This allows to determine the N abundance much more precisely than before and it seems to turn out that the N abundance is also significantly lower, about 0.1\% by mass, than suggested by Dreizler \& Heber (1998). Fig. \ref{fig:N_comparison} shows the N resonance doublets of three objects, the pulsators PG~1159$-$035 and PG~1707+427 and the non-pulsator PG~1424+535. While the comparison of PG~1707+427 and PG~1424+535 seems to confirm the characteristic difference in the N abundance, the new fit to the photospheric components in the STIS spectrum of PG~1159$-$035 shows that the N abundance is only 0.1\% by mass, but still higher than in the non-pulsator PG~1424+535. Further analyses are necessary to confirm these preliminary results. \begin{table} \caption{Summary of the atmospheric parameters of our program stars taken from the literature. All objects are of spectral type PG1159 except for Abell\,78, which is a [WC]-PG1159 transition object. The last column indicates whether the star is pulsating or not.} \label{tab:parameters} \footnotesize \begin{center} \begin{tabular}{l r c r c r r r c} \noalign{\smallskip} \tableline \noalign{\smallskip} Object & $T_{\mathrm{eff}}$ & $\log g$ &\multicolumn{1}{c}{H} & He &\multicolumn{1}{c}{C} & \multicolumn{1}{c}{N} & \multicolumn{1}{c}{O} & Puls.\\ \noalign{\smallskip} \cline{4-8} \noalign{\smallskip} & $[$kK$]$ & (cgs) &\multicolumn{5}{c}{(mass fractions)} &\\ \noalign{\smallskip} \tableline \noalign{\smallskip} RX\,J2117.1+3412 & 170 & 6.0 & & 38.0 & 56.0 & & 6.0 & $\times$\\ PG\,1144+005 & 150 & 6.5 & & 39.0 & 58.0 & 1.5 & 1.6 & \\ PG\,1520+525 & 150 & 7.5 & & 44.0 & 39.0 & & 17.0 & \\ PG\,1159$-$035 & 140 & 7.0 & & 33.0 & 49.0 & 1.0 & 17.0 & $\times$\\ K 1$-$16 & 140 & 6.4 & & 33.0 & 50.0 & & 17.0 & $\times$\\ HS\,2324+3944 & 130 & 6.2 & 21.0 & 41.0 & 37.0 & & 1.0 & $\times$\\ Abell 78 & 110 & 5.5 & & 33.0 & 50.0 & 2.0 & 15.0 & \\ NGC 7094 & 110 & 5.7 & 36.0 & 43.0 & 21.0 & & & \\ Abell 43 & 110 & 5.7 & 36.0 & 43.0 & 21.0 & & & $\times$ \\ PG\,1424+535 & 110 & 7.0 & & 50.0 & 44.0 & & 6.0 & \\ PG\,1707+427 & 85 & 7.5 & & 43.0 & 38.5 & 1.5 & 17.0 & $\times$\\ \noalign{\smallskip} \tableline \end{tabular} \end{center} \end{table} \acknowledgements{This research is supported by the DFG under grant WE\,1312/30-1 (E.R.), by the DLR under grant 50\,OR\,0201 (T.R.) and the FUSE project, funded by NASA contract NAS5-32985 (J.W.K.).} {}
Title: A high accuracy computed water line list
Abstract: A computed list of H$_{2}$$^{16}$O infra-red transition frequencies and intensities is presented. The list, BT2, was produced using a discrete variable representation two-step approach for solving the rotation-vibration nuclear motions. It is the most complete water line list in existence, comprising over 500 million transitions (65% more than any other list) and it is also the most accurate (over 90% of all known experimental energy levels are within 0.3 cm$^{-1}$ of the BT2 values). Its accuracy has been confirmed by extensive testing against astronomical and laboratory data. The line list has been used to identify individual water lines in a variety of objects including: comets, sunspots, a brown dwarf and the nova-like object V838 Mon. Comparison of the observed intensities with those generated by BT2 enables physical values to be derived for these objects. The line list can also be used to provide an opacity for models of the atmospheres of M-dwarf stars and assign previously unknown water lines in laboratory spectra.
https://export.arxiv.org/pdf/astro-ph/0601236
\date{Accepted XXXX. Received XXXX; in original form XXXX} \pagerange{\pageref{firstpage}--\pageref{lastpage}} \pubyear{2005} \label{firstpage} \begin{keywords} water, line list, BT2, molecular spectra \end{keywords} \section{Introduction} Water is the most abundant molecule in the universe after H$_{2}$ and CO. It is present in many astrophysical environments including the atmospheres of: M dwarfs (Allard et al.1994), brown dwarfs (Allard et al. 1996), K and M giants and supergiants (Jennings and Sada 1998; Ryde et al. 2002; Tsuji 2001) and oxygen-rich AGB stars (Barlow et al. 1996). It occurs in: sun-spots (Wallace et al. 1992; Polyansky et al. 1997), nova outflows (Banerjee et al. 2005), Mira variables (Hinkle and Barnes 1979), T Tauri eruptive variables (Shiba et al. 1993), dark molecular clouds (Gensheimer et al. 1996), young stellar objects (Carr et al. 2004), comets (Mumma et al. 1996; Dello Russo et al. 2000), the ISM (Cernicharo 1994), masers (Cheung et al. 1969; Gonz\'{a}lez-Alphonso et al. 1995) and planetary atmospheres. An accurate water line list is thus essential for interpreting spectra from all of these sources and in modelling stellar atmospheres at temperatures up to 4,000 K. The importance of water has given rise to many laboratory investigations of its spectrum. Ludwig 1971; Camy-Peyret et al. 1977; Bernath 1996 all investigated hot water line positions. However, techincal problems and the huge number of transitions (many of which appear blended) mean that only in the region of 80,000 (out of a total of more than a billion) transitions are known experimentally and there are few hot water lines for which intensities have been determined. The spectrum of water, which extends over a wide wavelength range from millimetre to near ultra-violet, is due to quantised changes in the rotation-vibration energy of the atomic nuclei moving in the electronic potential well. Essentially, the water molecule is only able to absorb or emit in its ground electronic state as the energy of the first stable excited electronic state is above the dissociation energy. In practice, some emissions do occur from short-lived excited electronic states. H$_{2}$O electronic transitions from diffuse interstellar clouds provide an example (Smith et al. 1981), but these transitions, which occur in the ultra-violet, can be disregarded in almost all other situations. Water is a triatomic asymmetric top molecule. Its rotation-vibration spectrum is more complicated than those of most other triatomic molecules. In common with all non-linear triatomics, H$_{2}$O has six degrees of internal freedom (three of rotation and three of vibration). However, the lightness of the hydrogen atoms means that the rotation constants are large, and this gives rise to an open spectrum that extends over a wide frequency range. Moreover, the `floppy' nature of the molecule means that the movement of the hydrogen atoms is generally anharmonic and consequently transitions involving changes of more than one vibrational quanta often occur. Also, since many of the vibrational frequencies are nearly resonant with other frequencies, it is common for vibrational bands to overlap and for states to interact in ways that cannot easily be predicted by perturbation theory but are amenable to a variational approach. The importance of the H$_{2}$O molecule in astronomy and the complexity of its spectrum have created a great deal of interest in the possibility of generating the spectrum synthetically. Previous synthetic line lists include: MT (Miller et al. 1994), VT1 (Viti et al. 1997) VT2 (Viti Ph.D.thesis 1997), PS, otherwise called AMES, (Partridge \& Schwenke 1997) and SCAN (J{\o}rgensen et al. 2001). All previous synthetic line lists have suffered from a number of problems that are discussed below. The most successful attempts have employed similar variational nuclear motion procedures to that used in producing the BT2 water line list. However, for reasons detailed in the next section, none of the earlier lists is considered to be satisfactory. We have addressed these problems and consequently the BT2 line list is an accurate tool for astronomers working in a variety of fields. \section{Background to the calculations} \subsection{Variational Techniques} Variational techniques represent the best approach to solving the the nuclear motion problem (Tennyson 1992), and they are examined in detail in Ba\v{c}i\'{c} and Light (1989). Here we use a discrete variable representation (DVR). In a DVR, the wavefunctions are defined by a complete set of weighted, orthogonal grid points, each wavefunction having a different set of weightings. This method is capable of generating accurate solutions, the accuracy being determined by the the number and appropriateness of the points. It is efficient for a large number of situations. The DVR approach has the advantage that the potential matrix elements are diagonal, and hence are easily evaluated. Even more important is the fact that the dipoles can be reduced to a similar form (Tennyson et al. 2004). The current work is not unique in employing a DVR approach to solve the nuclear motion problem for water (see for example, Viti et al. 1997). However, improved physics, in the form of a highly accurate potential energy surface (PES), the methodology embodied in the DVR3D program suite and increases in computational power have made it possible to produce a line list that is more complete and more accurate than any previous list, even those employing similar methodolodies. The DVR3D approach generates four rotation-vibration symmetry blocks for the H$_{2}$O molecule, which we label ee, eo, oe and oo. The first e/o term is the vibrational basis symmetry `q' and the second, e/o is the standard quantum number `p', the rotational parity. These symmetry blocks are not the same as the C$_{2v}$(M) symmetry blocks $\Gamma_{rv}$: A$_1$, A$_2$, B$_1$, B$_2$, but are related to them (see standard texts on molecular spectroscopy, such as Bunker \& Jensen 2005). The nuclear permutation operation in which the two identical protons comprising the hydrogen nuclei are interchanged gives rise two two dicrete states of the molecule: ortho (O) and para (P).The nuclear spins may couple symmetrically or antisymmetrically. The antisymmetric coupling that gives rise to the O form of the H$_{2}$O molecule is triply degenerate, whilst the symmetric coupling, that gives rise to the P form of the molecule is undegenerate. For any J, apart from J=0, there are two ortho and two para symmetry blocks (in the case of J=0 there is one O and one P block) There are two possible arrangements, depending on whether J is odd or even and these are detailed in Table 1. \begin{table} \begin{center} \caption{Symmetry Blocks} \begin{tabular}{cccccccccc} \hline { }& \multicolumn{4}{c}{\bf J even}&&\multicolumn{4}{c}{\bf J odd}\\ \cline{2-5}\cline{7-10} \bf q & e & e & o & o && e & e & o & o\\ \bf p & e & o & e & o && e & o & e & o\\ \bf O/P & P & O & O & P && O & P & P & O\\ \bf Code & 1 & 3 & 4 & 2 && 3 & 1 & 2 & 4\\ \hline \end{tabular} \end{center} {\footnotesize q is the vibrational basis symmetry and p is the rotational parity. These are labelled as either symmetric (e), or antisymmetric (o) states. O/P = Ortho/Para. The `Code' is the notation for symmetry used in the Levels File (Table 2).} \end{table} The difference in degeneracies of the O and P states impacts on the partition function of the molecule as well as on line intensity. Also, the fact that O-P and P-O transitions are forbidden, has spectroscopic consequences. \subsection{PES and the energy levels} The energies of the quantised rotation-vibration states are the eigenvalue solutions that satisfy the Shr\"{o}dinger equation for the oxygen and two hydrogen nuclei moving within the electronic potenial. However, the electronic potential within which the charged nuclei are moving is itself a function of the actual internuclear geometry. Therefore, in order to solve for the nuclear motion, it is necessary to have an accurate model of how the potential varies with the nuclear geometry. The problem is rendered tractable by adopting the Born-Oppenheimer approximation which separates the nuclear and electronic motions and has as its basis the fact that, due to their lightness, the electrons may be considered to react immediately to any changes in the nuclear geometry. The most accurate PESs are computed using an \emph{ab initio} starting point, with the resulting surface being empirically adjusted to improve the agreement between the computed energies and experimental data (Partridge and Schwenke 1997). We used the potential energy surface fit B of Shirin et al. (2003), which is based on a highly accurate \emph{ab initio} surface with adjustments for electronic relativistic and adiabatic (also known as the Born-Oppenheimer diagonal correction) effects and fitted to the available experimental data. At the time of writing, this surface is the most accurate available and it is the single most important factor affecting the accuracy of our results. Non-adiabatic corrections to the Born-Oppenheimer approximation are also important in the case of water (Schwenke 2003). However, a full theoretical non-adiabatic adjustement to the PES has been shown to produce no significant benefits compared to simplified approaches (Tennyson et al., 2002). These authors examine two such approaches: using separate reduced masses for the vibrational and rotational motions (this systematically over-corrects bending motions and under-corrects stretches), and a simplified version of the full correction which includes only terms that scale with the kinetic energy terms in $\theta$ and r. The latter approach is preferred and the ajustment is effected through the DVR3D program rather than changes to the PES. \subsection{Transition intensities} The intensities of the allowed transitions between rotation-vibration states are determined by the dipole transition moments for these pairs of states. They are: \begin{equation} \label{Eq:A} \quad\qquad \qquad \qquad \qquad \langle \psi^{\prime}|\overline\mu|\psi^{\prime\prime}\rangle \end{equation} where $\psi^{\prime}$ and $\psi^{\prime\prime}$ are the wavefunctions of the two states,and $\overline{\mu}$ is the electronic dipole moment vector. Therefore, in order to calculate intensities, a dipole moment surface (DMS) is required. The parameter computed by the DVR3D program suite is the Einstein A coefficient for, A$_{if}$, for each transition. This is the coefficient of spontaneous emission between the upper and lower states. It is related to the dipole transition moment for the pair of states and to J for the upper state. A$_{if}$, is independent of temperature, relates to a single molecule, has units of s$^{-1}$ and is given by: \begin{equation} \label{Eq:B} \qquad A_{if} = \frac{64\pi^{4}}{3c^{3}h}\nu^{3}g_{i}(2J^{\prime\prime}+1)|\langle \psi^{\prime} | \overline{\mu} | \psi^{\prime\prime} \rangle|^2 \end{equation} where prime and double prime represent the upper, i, and lower, f, states respectively. The quantity usually derived from observation, is the line intensity, I, which has units of cm/molecule. I is temperature-dependent and in emission is related to A$_if$ by the expression: \begin{equation} \label{Eq:C} I = \frac{C(2J^{\prime}+1)}{Q_{vr}(T) \nu^{2} g_i}\exp\left(\frac{-hcE^{\prime\prime}}{kT}\right)\left[1-\exp\left(\frac{-hc\nu}{kT}\right)\right]A_{if} \end{equation} where $\nu$ is the frequency in cm$^{-1}$, E$^{\prime\prime}$ is the energy of the lower ro-vibrational level in cm$^{-1}$. $Q_{vr}$(T) is the ro-vibrational partition function and is dimensionless, and $g_i$ is the nuclear spin degeneracy and caries only one subscript since transitions between different nuclear spin states are not allowed (Miani and Tennyson 2004). Boltzmann's constant, k, has units of JK$^{-1}$ and the constant C has the value (8$\pi$c)$^{-1}$ = 1.3271x10$^{-12}$ s cm$^{-1}$. Since the BT2 line list includes A$_{if}$ for each transition, the above equation enables the line intensities to be computed at any given temperature. Unlike the PES, where the most accurate are \emph{ab initio} surfaces that have been fitted to the available experimental data, in the case of DMS, limitations on the accuracy of experimental line strengths means that at the current time the most accurate surfaces are purely \emph{ab initio} (Lynas-Gray et al. 1995). Two such DMS were tested in the intensity part of our calculations. Our initial BT1 line list was computed using the same PES as BT2 and a preliminary Lynas-Gray et al. (in preparation) DMS. The line positions in BT1 are identical to those in BT2 (the same PES having being used for both). However, we observed that the Einstein A coefficients of the weaker lines generated using the Lynas-Gray et al. DMS were often too large when compared with experiment. This part of the computation was therefore repeated using the DMS of Schwenke and Partridge (2000). The results, which are contained in BT2, show much better agreement between the computed strengths of weak lines and experiment. The Einstein A coefficients of the stronger lines in BT1 and BT2 generally agree to within two percent and both agree reasonably well with experimental values. In addition to being superior to the Lynas-Gray et al. surface that we tested, Schwenke and Partridge's DMS represents a major improvement on the earlier DMS of Partridge and Schwenke (1997) and from our analysis, as well from the same authors, is the most accurate in existence. \section{Calculating the line list} The BT1 and BT2 line lists were computed using the DVR3D suite of programs (Tennyson et al. 2004) on three Sun 5 Microsystem V880 mainframe computers: Enigma and Ra which are clustered using high speed interconnects, each having 8 processors and 32 Gb of RAM and 432 Gb of disk storage and PSE, which has 24 processors and 96 Gb of RAM and 1,296 Gb of disk storage at UCL's Hiperspace computing centre. The final, `DIPOLE' stage of the suite was amenable to parallelisation with little time penalty. Other parts of the program were run on single processors, which avoided coding problems and was more efficient in computer time. The total number of processor hours employed in generating the BT1 line list (including the preliminary convergence testing which is discussed below) was 55,000 hrs and a further 10,000 hrs. were used in repeating the `DIPOLE' runs to generate BT2 and in testing the outputs. DVR3D calculates the bound rotation-vibration energy levels, the wavefunctions on a grid in three dimensional space, and the dipole transition strengths of the allowed transitions. The final part of the DVR3D suite, `SPECTRA' is able to compute temperature-dependent spectra over any selected frequency range, convolved with either the natural line width or some other selected profile, such as that given by the resolving power of a particular spectrometer. DVR3D uses an exact (within the Born-Oppenheimer approximation) kinetic energy operator. The program uses a discrete variable notation (DVR) with two radial and one angular co-ordinate for the nuclear motion problem (vibrational and rotational motions being treated spearately). In establishing the working parmeters for our calculations, we aimed to provide a line list that would be complete and accurate at the temperatures of late series K-series stars (up to 4,000 K) and at wavelengths down to 0.8 $\mu$m. Preliminary calculations showed that in order to achieve this it would be necessary to include all states lying at energies up to 30,000 cm$^{-1}$ relative to the ground state of the system. Previous workers (Miller et al. 1994 and Partridge and Schwenke 1997) selected lower energy cut-offs. Our cut-off for the total angular momentum was J=50. We calculate that the highest value of J which has ro-vibrational energies of less than 30,000 cm$^-1$ is 58. However, we also estimate that by terminating our calculation at J=50 we omit less than 500 levels out of a total of more than 505 million and none of the omitted levels has an energy less than 23,490 cm$^{-1}$, with the majority being at energies above 28,000 cm$^{-1}$. Consequently, even at a temperature of 4,000 K these missing levels contribute less than 0.02\% of the total partition function. Moreover by omitting Js above 50 we saved in the region of 8,000 processor hours. In order to generate accurate eigenvalues, it is essential that all calculations are fully converged within the limitations imposed by computing power and time. The lack of convergence in earlier lists has already been noted (Polyansky et al. 1997). DVR3D has the option of using Jacobi or Radau co-ordinates. The latter was selected as being more appropriate for water. Our Radau grid is defined in terms of two radial co-ordinates, r$_{1}$ and r$_{2}$, each with 28 points, and one angular co-ordinate, $\theta$ having 44 points. These numbers are determined by convergence testing at J$=$20. The eigenvalue solutions are particularly sensitive to the number of radial points, r$_n$. However, the computation time rises very rapidly as this number is increased, so the selection of r$_n$ represents a compromise and is the principal factor impacting on convergence. The angular grid points are arrived at using associated Legendre polynomials for the underlying basis sets, whilst the radial grid points are set up using Morse oscillator-like underlying basis sets, which are defined in terms of parameters, r$_{e}$, $\omega_{e}$ and D$_{e}$. These must be entered into `DVR3DRJZ', the first module of the DVR suite. The parameters have physical counterparts which are respectively: the equilibrium bond length, harmonic frequency and dissociation energy of the water molecule. However, the parameters in the Morse oscillator function differ from the physical values and must be determined by empirical testing. Although in principle, DVR3D is not strictly a variational method, in practice it is found that the Variational Principle does apply; this fact is employed in obtaining values for r$_{e}$, $\omega_{e}$ and D$_{e}$. The method is to alter the three parameters in a systematic manner until DVR3DRJZ generates a set of pure vibrational states (the J=0, para states), the sum of whose energies is a minimum (251 eigenvalues were employed). We conducted this part of the process manually. The investigation was complicated by the existence of local minima, which are not global minima for the three variables. The set of parameters giving this result was: r$_{e}$ = 2.05, $\omega_{e}$ = 0.008 and D$_{e}$ = 0.20 all in atomic units (the equivalent dissociation energy is 43,895 cm$^{-1}$). A good choice of these parameters (which are a function of the system and of the particular energy range that is of interest) is fundamental to the accuracy of the ensuing calculations. The eigenvalues are particluarly sensitive to r$_{e}$ and $\omega_{e}$ and it was observed that differences of as little as 0.05 in the first of these two parameters and 0.002 in the second could affect the energy levels by in the region of 0.01cm$^{-1}$ for states with energies in the region of 10,000 cm$^{-1}$. Larger deviations in these basis functions from their determined optimum values produced correspondingly greater errors and in the case of states with energies over 20,000 cm$^{-1}$, a bad choice of parameters could easily result errors in the individual levels in excess of 20 cm$^{-1}$. Consequently considerable time was spent in determining the values of the Morse oscillator-like basis set. Two other inputs are required by the vibrational module, DVR3DRJZ. These are: the maximum size of the intemediate Hamiltonian, which we chose after testing as 2500 and the number of eigenvectors to be saved for use in the rotational module, `ROTLEV3B', the optimum value for which was found by testing to be 700. ROTLEV3B also requies one variable to be determined; this is IBASS, the size of the Hamiltonian in the rotation module. IBASS varies with J. Its value was established by convergence testing at J=20 as being $530\times(J+1-p)$, where p is the rotational parity and has the value 0 for even parity states and 1 for odd parity states. Hamiltonians of this size are expensive in terms of computing time at high Js. Nevertheless, it was easy to demonstrate that lower values of IBASS produce results that are not converged. This is significant, as earlier workers using similar DVR techniques have used lower IBASS values. Viti et al. 1997, for example, used $200\times(J+1-p)$, and PS used an even lower effective number, particularly at high J. In selecting the various parameters referred to above, we regularly tested for convergence. We estimate that our choice of r$_{n}$=28 accounted for approximately half of our total convergence error, which we estimated as being less than 0.01 cm$^{-1}$ at 10,000 cm$^{-1}$ and in the region of 0.02 cm$^{-1}$ at 20,000 cm$^{-1}$. \section{Results} The BT2 water line list is available electronically in compressed form at: \emph {ftp://cdsarc.u-strasbg.fr/cats/VI/119} The data are in two parts. The first, the `` Levels File'' is a list of 221,097 energy levels, ordered by J and symmetry block. About 25,000 of these energy levels have been labelled with the appropriate angular momentum (J,$K_a,K_c$) and vibrational ($\nu_1$,$\nu_2$,$\nu_3$) quantum numbers. An extract from the Levels File with an explanation of the contents of each of the 11 columns in the file is given in Table 2. The second part of BT2 is the ``Transitions File''. This has 505,806,202 entries. Each transition references upper and lower energy levels in the Levels File and gives the Einstein $A_{if}$ coefficient for the transition. An extract from the Transitions File is given in Table 3. In uncompressed form the BT2 Transitions File is 12.6 Gb of data. Therefore, in order to facilitate use of the list, the transitions have been ordered by frequency and separated into 16 smaller files, each representing a specific frequency range. In addition to the files containing the actual line list, the Strasbourg site contains a Fortran program, spectra-BT2.f90, that will enable users to generate emission or absorption spectra from BT2 by specifying various parameters including: temperature, frequency range, cut-off intensity and line width. There is also a facility to generate spectra with full ro-vibrational assignments if required. The method of using spectra-BT2.f90 is detailed in a `readme-spectra' file and there are also examples of a job file and an output file. \begin{table} \caption{Extract from the BT2 Levels File} \begin{tabular}{|cccrccccccc|} \hline A & B & C & D & E & F & G & H & I & J & K \\[0.5ex] \hline 2284 & 2 & 2 & 5 & 3885.718672 & 0 & 0 & 1 & 2 & 2 & 1 \\ 2285 & 2 & 2 & 6 & 4777.145956 & 0 & 3 & 0 & 2 & 1 & 1 \\ 2286 & 2 & 2 & 7 & 5332.258064 & 1 & 1 & 0 & 2 & 1 & 1 \\ 2287 & 2 & 2 & 8 & 5472.371851 & 0 & 1 & 1 & 2 & 2 & 1 \\ 2288 & 2 & 2 & 9 & 6254.694085 & 0 & 4 & 0 & 2 & 1 & 1 \\ 2289 & 2 & 2 &10 & 6876.917089 & 1 & 2 & 0 & 2 & 1 & 1 \\ 2290 & 2 & 2 &11 & 7027.396535 & 0 & 2 & 1 & 2 & 2 & 1 \\ 2291 & 2 & 2 &12 & 7293.201639 & 2 & 0 & 0 & 2 & 1 & 1 \\ 2292 & 2 & 2 &13 & 7376.617020 & 1 & 0 & 1 & 2 & 2 & 1 \\ 2293 & 2 & 2 &14 & 7536.864373 & 0 & 0 & 2 & 2 & 1 & 1 \\ \hline \end{tabular} {\footnotesize A: Row in file, B: J, C: Symmetry (1-4: see Table 1), D: Row in block, E: $\nu$ in cm$^{-1}$ F, G, H: $\nu_1$, $\nu_2$, $\nu_3$. I, J, K: J, K$_a$,K$_c$. } \end{table} \begin{table} \begin{center} \caption{Extract from BT2 Transitions File} \begin{tabular}{|ccc|} \hline A & B & C \\[0.5ex] \hline 1000 & 239 & 9.671E+01 \\ 1001 & 239 & 1.874E+00 \\ 1002 & 239 & 4.894E-03 \\ 1003 & 239 & 1.140E-04 \\ 1004 & 239 & 1.707E-02 \\ 1005 & 239 & 8.473E-08 \\ 1006 & 239 & 6.535E-04 \\ 1007 & 239 & 7.157E+00 \\ 1008 & 239 & 6.403E-06 \\ 1009 & 239 & 9.861E-05 \\ \hline \end{tabular} \\ {\footnotesize A, B: Row numbers in the Levels File (upper and lower levels are not identified as the program tests for these). C: A$_{if}$ (s$^{-1}$).} \end{center} \end{table} \begin{table} \begin{center} \caption{Comparison of BT2 and PS (Partridge \& Schwenke 1997) with 14,889 experimentally-determined energy levels} \begin{tabular}{|ccc|} \hline Within & BT2 & PS \\ cm$^{-1}$ & \% & \% \\ \hline 0.1 & 48.7 & 59.2 \\ 0.3 & 91.4 & 85.6 \\ 1.0 & 99.2 & 92.6 \\ 3.0 & 99.9 & 96.5 \\ 5.0 & 100 & 97.0 \\ 10.0& 100 & 98.1 \\ \hline \end{tabular} \\ \end{center} \end{table} \subsection{Comparing the BT2 and Partridge \& Schwenke line lists} Several water line lists are in regular use by astronomers and the the most accurate list previously is that of Partridge and Schwenke (PS). Table 4 compares BT2 and PS energy levels with known experimental values (Tennyson et al. 2001). It will be seen from this table that although the PS list is more accurate than BT2 in the cases where agreement with experiment is better than 0.1 cm$^{-1}$, this is not the case generally. Specifically, based on a sample of 14,889 levels, whilst 99.9\% of the BT lines are within 3.0 cm$^{-1}$ of experiment 3.5\% of the PS lines are outside this range. Other line lists (MT, VT2, SCAN) perform significantly worse than this. Examining deviations from experiment by energy is even more revealing, for it is seen that PS is increasingly unreliable above 10,000 cm$^{-1}$, which is the region with the greatest number of transitions. \begin{table} \begin{center} \caption{Distribution of levels in the BT2 and PS (Partridge \& Schwenke 1997) disageeing with experiment by more than 2 cm$^{-1}$ - by frequency} \begin{tabular}{rccc} \hline Level Energy & Number & BT2 & PS \\ cm$^{-1}$ & in range & No. & No. \\ \hline 20,000 - 26,300 & 575 & 9 & 334 \\ 15,000 - 20,000 & 2,813 & 10 & 105 \\ 10,000 - 15,000 & 6,323 & 8 & 58 \\ 7,000 - 10,000 & 3,263 & 3 & 9 \\ $<$ 7,000 & 1,914 & 0 & 0 \\ \hline Total & 14,889 & 30 & 506 \\ \hline \end{tabular} \\ \end{center} \end{table} \subsection{Labelling the levels} As with observed water lines, the production of a synthetic line list prompts the question of how to assign quantum numbers to the transitions. The BT2 line list contains over 505 million transitions, but the DVR3D suite only provides data on J and the symmetry block of each energy level. Since each individual line is a transition between two energy states, the problem reduces to one of labelling the 221,097 energy levels, but this is still a large task. Many papers have been devoted to assigning quantum numbers to energy levels that have been deduced from experimental line frequencies (e.g. Zobov et al. 2000). So far, less than 15,000 experimentally-determined levels have been labelled with their 3 rotational and 3 vibrational quantum numbers, even though much effort has been expended on the task. 25,870 levels in the BT2 list have been labelled using methodologies detailed below. BT2 labels 270 of the 416 levels which have J=0 and energies below 30,000 cm$^{-1}$. However, the proportion labelled is less at higher Js. The process of labelling BT2 started by identifying particular energies with experimental levels that have already been determined. In addition, we have labelled many of the energy levels that are unknown experimentally, using several different methods. A large number of vibrational states for J=0 have been labelled by visual inspection of the nodal structure of the wavefunction as described in Mussa and Tennyson (1998). This method has the disadvantage that observations can only easily be made on two-dimensional sections of the three-dimensional wavefunction and the procedure can be misleading. In addition, Mussa and Tennyson oberved that at energies in the region of our 30,000 cm$^{-1}$, a high proportion of the wavefunctions are irregular with no identifyable nodal structure. . A second method involved examining the $A_{if}$ coefficients for pure rotational transitions between an unlabelled vibrational state and a known vibrational state; the strongest transitions being those where standard selection rules are obeyed. We found this method (see Tolchenov et al 2005) to be useful for J$<$10 A third method that was found to be useful in labelling higher J states involved the use of an algorithm to identify sets of levels within the same parity block having the same K$_{a}$, $\nu_{1}$, $\nu_{2}$, $\nu_{3}$ quantum numbers, but different values of J and K$_{c}$. The method was originally developed for labelling the energy levels of the HCN and HNC isomers (Barber et al. 2000). However, because of the density of the energy levels in water, an extra term was introduced when labelling the water levels. The algorithm used in this case was: \begin{equation} \label{Eq;6} E_{J_{n}}\cong 4E_{J_{n-1}} -6E_{J_{n-2}} +4E_{J_{n-3}} -E_{J_{n-4}} \end{equation} However, when resonance between levels caused the behaviour pattern to be erratic, or no E$_{J_{n-4}}$ value existed, the original, simpler algorithm was used: \begin{equation} \label{Eq;7} E_{J_{n}}\cong 3E_{J_{n-1}} -3E_{J_{n-2}} +E_{J_{n-3}} \end{equation} where E$_{J_{n}}$ is the energy of the state in the same symmetry block having J = n and the same set of K$_{a}$, $\nu_{1}$, $\nu_{2}$, $\nu_{3}$ quantum numbers. The results of the labelling exercise are included in the BT2 Level File. This means that when synthetic spectra are generated many of the transitions are fully labelled. This feature is useful when generating synthetic spectra for astronomical or laboratory applications as is discussed in the next section. \subsection{Completeness} If we compare the partition function, Q(T) for water computed at a particular temperature using the BT and PS line lists with the most accurately known value at this temperature, it is possible to estimate the the completeness of the line lists and the amounts of opacity that are missing in spectra generated by the two line lists at the selected temperature. A calculation of the partition function of water at 3,000K using the 221,097 energy levels in BT2 yields a value that is 99.9915 percent of the Vidler \& Tennyson (2000) value, which indicates that levels missing from BT2 only contribute about 85 parts in a million to the partition function of water at this temperature, the reason being that there is a diminishingly small probability of states above 30,000 cm$^{-1}$ being occupied at this temperature. For comparison, the PS line list, which has 28,000 cm$^{-1}$ cut-off gives a partition function at 3,000 K that is only 99.493$\%$ of the Vidler \& Tennyson value. Although the exclusion of levels above 30,000 cm$^{-1}$ does not materially affect the completeness of the BT2 list, it does affect absorption at shorter wavelengths. If we consider a photon of of wavelength 1 $\mu$m (energy 10,000 cm$^{-1}$). This is able to be absorbed by a water molecule in a particular rotation-vibration eigenstate provided that there is another eigenstate exactly 10,000 cm$^{-1}$ above this lower state into which the molecule may be excited. It follows that since the BT2 line list has an upper cut-off of 30,000 cm$^{-1}$, none of the energy levels in the list above 20,000 cm$^{-1}$ are capable of being excited by a 1$\mu$m photon, since there is no corresponding upper level. If we examine the extent to which Q(3,000K) computed from BT2, but excluding all levels above 20,000 cm$^{-1}$, falls short of the Vidler \& Tennyson value we will have an indication (this is an upper limit as it takes no account of blending effects) of the opacity that has been excluded by adopting a 30,000 cm$^{-1}$ cut-off. Performing the calculation gives a shortfall of 0.83$\%$. In the case of PS, only energy levels below 18,000 cm$^{-1}$ are able to absorb a 1$\mu$m photon, and computing Q(3,000K) using only PS states up to 18,000 cm$^{-1}$ shows that these comprise only 98.37$\%$ of the Vidler value. Hence it will be seen that the opacity deficit at 1$\mu$m is in the region of twice as great in the PS list as in BT and the ratio increases at shorter wavelengths. \section{Sample Applications} Although BT2 shows good agreement with experimentally known lines this is not a sufficient test of its accuracy as the PES used to generate BT2 was fitted to the known experimental data (this is also true of the PS line list). The most effective way of checking the accuracy of a line list is to test its ability to predict or identify previously unknown lines in astronomical or laboratory spectra. BT2 has been used successfully as outlined below. It should be noted here that some of the earlier spectroscopic applications used BT1 as they predate BT2. However, the line positions of the two lists are identical, and the intensities of the strong lines are similar for the two lists. \subsection{Astronomical Spectra} \subsubsection{$\epsilon$ Indi Ba} A synthetic spectrum generated at 1,500K using BT1 was able to reproduce the previously unknown absorption feature observed at 1.554 $\mu$m in the spectrum of the early T dwarf $\epsilon$ Indi Ba, discovered by Volk et al. (2003), as being a blend of six water lines with no individual line contributing more than 25$\%$ of intensity (Smith et al. 2003). \subsubsection{Comet 153P/Ikeya-Zhang (2002 C1)} BT1 was used to compute the frequencies and Einstein A coefficients of the 64 transitions (up to J=7) that make up each of the 7 hot bands of water detected in Comet 153P/Ikeya-Zhang. Dello Russo et al. (2004) applied these data in determining the rotational temperature of the comet on three dates. \subsubsection{Temperatures of comet forming regions in early solar nebula} The hot-band transitions identified by BT1 in Dello Russo et al. (2004) were classified into ortho and para using the symmetry information contained within the BT2 energy file (see column C Table 1). Transitions between different nuclear spin parities can be ignored (Miani and Tennyson 2004). Dello Russo et al. (2005) were able to deduce the primordial O/P water composition of three comets: C/1999 H1, C/1999 S4 and C/2001 A2 and hence the temperatures of the different regions of the early solar nebula in which the comets were formed. The normal O/P ratio is 3:1, but since the lowest ortho level lies 23.8 cm$^{-1}$ ($\sim$ 34 K) above the ground state (which is a para state), the O/P ratio is sub-normal at temperatures below $\sim$ 50 K. A comparison of the Einstein A coefficients in BT2 with those actually used in BT1 shows that Dello Russo et al.'s results would have been the same had the later line list been used. \subsubsection{Detection of water lines in nova-like object V838 Mon} Synthetic spectra generated by BT2 were used to identify absorption features observed in the 1.73 to 1.75 $\mu$m region of the spectrum of the nova-like object, V838 Monocerotis on five separate dates as being due to blended water lines. Quantum numbers were assigned to the 17 strong transitions that comprise the 5 absorption features (Banerjee et al. 2005). Sixteen of the lines were found to be in the (0 0 0)-(0 1 1) band. This is the first time that individual water lines have been identified in a nova-like outflow region. In addition, BT2 was used to compute the theoretical intensities of the five absorption features as a function of temperature and column density (assuming LTE). The results indicated that the water features were arising from a cool $\sim$ 750--900 K region around V838 Mon that was cooling at a rate of $\sim$ 100 K per year. Column densities computed for the five dates also showed a reduction with time. \subsubsection{Sunspot spectra} Following their assignment of high temperature laboratory water lines in which extensive use was made of BT2 (see `Laboratory Spectra' below), Coheur et al. (2005) revisited the sunspot absorpton spectrum in the 9.89--12.95 $\mu$m region of Wallace et al. (1995) which had been partially assigned by Polyansky et al. 1997 and Zobov et al. 1999. Coheur et al. used their high temperature laboratory assignments to identify a substantial number of previously unassigned sunspot lines. \subsection{Laboratory Spectra} Hot water spectra have been analysed in the laboratory over the last thirty years. Methods generally involve high resolution Fourier transform spectrometry of vapour which may be at an elevated temperature, such as in an oxy-acetylene flame (Camy-Pyret et al. 1977, Coheur et al. 2005). The BT2 line list has already been used on a number of occasions to analyse spectra generated at both high and low temperatures at various wavelengths and this has has added considerably to the existing database of experimentally known energy levels and transitions. Coheur et al. (2005) working with experimental spectra generated at 3,000 K in the 5--20 $\mu$m region labelled about 600 previously unidentified levels using the BT2 line list. The identification of these levels was an important factor in their subsequent assignment of 8,500 of the 10,100 lines that they observed in the 0.385-1.081 $\mu$m region. Most of the states that were labelled by Coheur et al. were either high J states having low bending modes or else lower J states with higher bending modes. These states had defied earlier analysis because the previous line lists used in earlier work, were unable to accurately predict the energies of states above 15,000 cm$^{-1}$ or to treat accurately the high bending mode states, even those below 15,000 cm$^{-1}$. They note that whilst the agreement between BT2 and observation is generally within 0.1 cm$^{-1}$ for the energy levels that they observe for low J states, the disageement can be as great as 0.8 cm$^{-1}$ for some of the very high J states. Nevertheless, they comment that since for a given vibrational and K$_a$ state the diffence between the BT2 list and experiment increases smoothly with J, they are able to use BT2 to predict the positions of unknown higher J levels with an accuracy of 0.02 cm$^{-1}$. This is considerably less than the experimentally determined line widths (full widths at half maximum) which were in the range 0.05--0.10 cm$^{-1}$ due to broadening at 1 atm. and T $=$ 3000 K. In a follow-up paper analysing the 2--5$\mu$m region Zobov et al., 2006 use BT2 to label approximately 700 previously unidentified ro-vibrational energy levels for water, observed in laboratory torch spectra. Tolchenov et al. 2005 analysed long-pathlength room temperature spectra. They use three separate line lists in their work. However, only the BT2 list is found to be reliable over all transition frequencies and in the case of lines having frequncies above 16,000 cm$^{-1}$ it is the only one that they use. In addition, Tolchenov et al. encounter difficulties with the labelling adopted in previous studies, finding, for example, that in some cases different states have been labelled with the same quantum numbers. They therefore undertake a systemmatic re-labelling exercise using the BT2 energies. These labels are incorporated into the BT2 list for states with J $\le$ 9. Dupr\'{e} et al.(2005) found BT2 similarly reliable for predictions of long-pathlength room temperature spectra in the near ultra-violet. They observed 62 R-brach transitions in the 8$\nu$ polyad and were able to determine 36 energy levels, previously unknown experimentally. \section{Conclusion} We present a new synthetic water line list which gives the energies of 221,097 states with cut-offs of J=50 and E=30,000 cm$^{-1}$. 25,870 of the lower energy levels have been labelled with a full set of three rotational and three vibrational quantum numbers. BT2 lists 505,806,202 trasitions. It has been extensively tested against experimental observations and also compared with other lists. It has been shown to be the most complete and accurate water line list in existence. We make our results freely available in electronic form via \emph{ftp://cdsarc.u-strasbg.fr/cats/VI/119}, in the hope that BT2 line list will be a vauable tool for astronomers in both spectroscopy and atmospheric modelling applications. One of the problems facing the modellers of the atmospheres of cool stars and brown dwarfs is the disageement between observation and model. Nevertheless, considerable progress has been made in recent years in such areas as: convection, molecular abundances, non-LTE effects and dust (Hauschildt et al. 1999; Tsuji 2002; Alexander et al. 2003). At the same time, there have been advances in computing the opacity effects of the various species that are included in the models. At the temperature of late M-dwarfs (2,500K), water is by far the most important contributor to stellar atmospheric opacity, typically contributing over 60$\%$ of all opacity in the infra-red. It is also an extremely important contributor to the opacity of L and T brown dwarfs, although at these lower temperatures (down to 900K), methane and ammonia play an increased role. As part of our on-going work in applying BT2 we plan to compare observed M-dwarf spectra with synthetic spectra produced with the Phoenix model (Hauschildt, Allard \& Baron 1999) using several water line lists, including BT2. This is an extension of work already conducted on modelling oxygen rich cool stars. Jones et al. (2005) compare the CO 2-0 bands in the 2.297-2.210 $\mu$m region of several M dwarfs and an L dwarf with opacity calculations using the PS and BT line lists and conclude that for this particular wavelength range `While the Partridge-Schwenke line list is a reasonable spectroscopic match [for the BT line list] at 2,000K, by 4,000K it is missing around 25 $\%$ of the water vapour opacity.' \section{Acknowledgments} R.J.Barber has been sponsored in this work by PPARC.\\ Calculations were performed at the UCL Hiperspace computing centre and we thank its manager, Callum Wright, for his assistance. We also thank Svetlana Voronina for performing consistency checks on our energy level labels. \label{lastpage}
Title: Self-Gravitating Phase Transitions: Point Particles, Black Holes and Strings
Abstract: We compute the quantum string entropy S_s(m,j) of the microscopic string states of mass m and spin j in two physically relevant backgrounds: Kerr (rotating) black holes and de Sitter (dS) space-time. We find a new formula for the quantum gravitational entropy S_{sem} (M, J), as a function of the usual Bekenstein-Hawking entropy S_{sem}^(0)(M, J). We compute the quantum string emission by a black hole in de Sitter space-time (bhdS). In all these cases: (i) strings with the highest spin, and (ii) in dS space-time, (iii) quantum rotating black holes, (iv) quantum dS regime, (v) late bhdS evaporation, we find a new gravitational phase transition with a common distinctive universal feature: A square root branch point singularity in any space-time dimensions. This is the same behavior as for the thermal self-gravitating gas of point particles (de Vega-Sanchez transition), thus describing a new universality class.
https://export.arxiv.org/pdf/astro-ph/0601601
\centerline{\bf SELF-GRAVITATING PHASE TRANSITIONS:} \centerline{\bf POINT PARTICLES, BLACK HOLES AND STRINGS} \begin{center}{\bf Norma G. Sanchez} \footnote{Norma.Sanchez@obspm.fr} \end{center} \begin{center}{Observatoire de Paris, LERMA, Laboratoire Associ\'e au CNRS UMR 8112, 61, Avenue de l'Observatoire, 75014 Paris, France.} \end{center} \centerline{\bf Abstract} We compute the quantum string entropy $S_s(m,j)$ of the microscopic string states of mass $m$ and spin $j$ in two physically relevant backgrounds: Kerr (rotating) black holes and de Sitter (dS) space-time. We find a {\bf new} formula for the quantum gravitational entropy $S_{sem} (M, J)$, as a function of the usual Bekenstein-Hawking entropy $S_{sem}^{(0)}(M, J)$. We compute the quantum string emission by a black hole in de Sitter space-time (bhdS). In {\bf all} these cases: (i) strings with the highest spin, and (ii) in dS space-time, (iii) quantum rotating black holes, (iv) quantum dS regime, (v) late bhdS evaporation, we find a {\bf new} gravitational phase transition with a common distinctive {\bf universal} feature: A square root {\bf branch point} singularity in any space-time dimensions. This is the same behavior as for the thermal self-gravitating gas of point particles (de Vega-Sanchez transition), thus describing a new {\bf universality class}. Nous calculons l'entropie $S_s(m, j)$ des \'etats microscopiques de masse $m$ et spin $j$ des cordes quantiques dans deux espaces-temps physiquement relevants: Le trou noir en rotation (de Kerr) et l'espace-temps de de Sitter (dS). Nous trouvons une {\bf nouvelle} formule pour l'entropie gravitationnelle quantique $S_{sem} (M, J)$ comme fonction de l'entropie de Bekenstein-Hawking $S_{sem}^{(0)}(M)$. Nous calculons l' \'emission quantique des cordes par un trou noir avec constante cosmologique (bhdS). Dans {\bf tous} ces cas: (i) cordes avec le spin maximal, et (ii) dans l'espace-time de dS, (iii) trous noirs de grand moment angulaire, (iv) regime quantique de dS, (v) derni\`ere \'etape d'\'evaporation bhdS, nous trouvons une {\bf nouvelle} transition de phase avec la m\^eme caract\'eristique distinctive {\bf universelle} : un {\bf point de ramification} racine carr\'ee \`a la transition, pour toute dimension de l'espace-temps. C'est le m\^eme comportement que pour le gaz auto gravitant de particules ponctuelles (transition de de Vega-Sanchez), d\'efinissant ainsi une nouvelle {\bf classe d'universalit\'e}. \\ {\bf Key words}: gravitational phase transitions, rotating black holes, quantum strings, de Sitter, self-gravitating gaz \\ A theory of quantum gravity, such as string theory, should account for an unified and consistent description of both black holes and elementary particles, and the physics of the early universe as well. \\ A central object in string theory is $\rho_s (m,j)$, the microscopic string density of states of mass $m$ and spin $j$. We find $\rho_s (m,j)$ by solving the non-linear and quantum string {\bf dynamics} in the curved space-times considered, refs. \cite {1}-\cite{5}. We compute the quantum string entropy from the microscopic density of states $\rho_s (m,j)$ in two physically relevant space-times: the rotating (Kerr) black hole and the de Sitter space-time, refs \cite{6},\cite{7}. Combination of the two effects at the transition: black hole and cosmological constant are also treated for a black hole in de Sitter space-time (bhdS), ref. \cite{7}. \\ The Hawking temperature $T_{sem}$ and the string (Hagedorn) temperature $T_s$ are the same concept in different (semiclassical and quantum) gravity regimes, ref. \cite{2}. Similarly, it holds for the Bekenstein-Hawking entropy and the string entropy, ref. \cite{2}. $T_s$ is the precise quantum dual of $T_{sem}=\hbar c /(2\pi k_B L_{c\ell})$, $ L_{c\ell}$ being the classical radius: $2R_S/(D-3)$, ($R_S$ is the black hole radius), or $c/H$ for Sitter space-time, $H$ being the Hubble constant, (in any $D$ space-time dimensions). For $L_{c\ell}\gg \ell_{Planck}$, ie. for low curvature or low $H\ll c/\ell_{Planck}$, (semiclassical gravity regime), the Bekenstein-Hawking entropy $S_{sem}^{(0)}$ is the leading term of the whole entropy $S_{sem}$, {\bf but} for high curvature or very small radius near $\ell_{Planck}$, (quantum gravity regime), a {\bf new} phase transition operates and the whole entropy $S_{sem}$ is drastically {\bf different} from the Bekenstein-Hawking entropy $S_{sem}^{(0)}$, refs \cite{6},\cite{7}. \\ The phase transitions we found for quantum strings in Minkowski, Kerr and de Sitter backgrounds, as well as for the quantum regimes of rotating black holes and de Sitter space, are {\bf all} of the same nature: a {\bf square root} branch point singularity, (the value of the critical temperature being different in each case). This phase transition does not occur in Anti de Sitter space-time (AdS): in the AdS background, the density of mass states is finite, the negative cosmological constant "pushes" the critical temperature to infinity and there is no singularity at all, refs \cite{1}, \cite{2}. \\ The origin of this string or gravitational {\bf square root} phase transition is gravity at finite temperature or in the presence of a repulsive force (angular momentum, positive cosmological constant). This behavior is similar to that found for the selfgravitating gas of point particles (de Vega-Sanchez transition), refs.\cite{8}, \cite{9}. In our context here, the temperature is not an external parameter, but {\bf intrinsic} to the theory, determined by the mass of the system (strings, or space-time backgrounds). Furthemore, our treatment here is not statistical mechanical , but {\bf microscopic} or dynamical. \\ Here, we focalizes on the phase transition we found for strings and the quantum black hole and de Sitter regimes. More results and implications on these new issues can be found in refs. \cite{6},\cite{7}. Black hole evaporation ends as quantum string decay into elementary particles (most massless), ie pure (non mixed) quantum states refs \cite {2},\cite{6}, \cite{7}.\\ {\bf Quantum String Entropy in the Kerr background}: The string entropy $S_s(m,j)$ in a Kerr black hole background is given by $\rho_s(m,j)= e^{\frac{S_s(m,j)}{k_B}} $, where the microscopic string density of states of mass $m$ and spin $j$, $\rho_s(m, j)$, has been found in ref. \cite{6} and can be expressed as $$ \rho_s(m, j) \sim \Big(\frac{S_s^{(0)}}{k_B}\Big)^{-a} ~ e^{\Big(\frac{S_s^{(0)}}{k_B}\Big)} ~F(S_s^{(0)}, j)~~, $$ $$ F(S_s^{(0)}, j) = \Delta_s^{-a-1} e^{\frac{S_s^{(0)}}{k_B} \frac{(1 - \Delta_s) ^2}{2 \Delta_s}} \cosh^{-2} \Bigg( \frac{S_s^{(0)}}{4 k_B} \frac{(1 - \Delta_s^2)}{\Delta_s} \Bigg) $$ ~~~~~~~~~~$\Delta_s = \sqrt{1 - \frac{4 j }{m^2 \alpha^{'} c}}=\sqrt{1 - \frac{j}{\hbar} \Big( \frac{k_B~b}{S_{s}^{(0)}} \Big)^2}~~,~~~4j\leq m^2 \alpha^{'} c$~. $S_s^{(0)}$ is the zero order string entropy for $j=0$: $S_s^{(0)} = \frac{1}{2} \frac{mc^2}{t_s}$ ~~~(closed~ strings), ~~~~ $S_s^{(0)} = \frac{mc^2}{t_s}$~~~(open~ strings), and $t_s$ is the string Hagedorn temperature, ~~~~$t_s = \frac{m_s c^2}{k_B b}$, ~~~being ~~$b = 2 \pi \sqrt{ \frac{D- 2 }{6}}$, $a = D$ (closed strings),~~ $a = (D-1)/2$ (open strings). $D$ is the number of space-time dimensions. $m_s$ is the fundamental string mass: $m_s = \sqrt{\frac{\hbar}{\alpha'c}}\equiv~\frac{\ell_s}{\alpha'}$. $\alpha'$ is the fundamental string constant ($\alpha'^{-1}$ is a mass linear density), $\ell_{s}$ is the fundamental string length. $\Delta_s$ can be expressed as : $$ \Delta_s= \sqrt{1 - \frac{j}{\hbar} ~\Big(\frac{m_s}{m} \Big)^2} = \sqrt{1 - \frac{j}{\hbar}~ \Big(\frac {t_s}{T} \Big)^2} ~~,~~ T = \frac{m c^2}{k_B b}. $$ $F({m}, j)$ takes into account the effect of the angular modes $j$, being $ F({m},j=0) = 1$. Therefore, the string entropy $S_s(m,j)$ in the Kerr background is given by $$ S_s(m,j) = S_s^{(0)} - a ~k_B~\ln \Big (~\frac{S_s^{(0)}}{k_B}~ \Big)~+~k_B~\ln F(S_s^{(0)}, j) $$ The last new term $k_B\ln F(S_s^{(0)},j)$ is enterely due to the spin $j \neq 0$. Interestingly enough, this formula expresses the string entropy $S_s(m,j)$ for mass $m$ and spin $j$ in terms of the string entropy $S_s^{(0)}$ for $j=0$. For $j=0$, we recover the usual expression. The logarithmic terms have a negative sign, the effect of the spin is to reduce the entropy. $S_s(m,j)$ is maximal for $j=0$ (ie, for $\Delta_s = 1$).\\ For $j < (m/m_s)^2$, and $m \gg m_{s}$, that is for low $j$ and very excited string states, $S_s^{(0)}(m,j)$ is the leading term, but for high $j$, that is $j \rightarrow m^2 \alpha^{'} c$, ie $\Delta_s \rightarrow 0$, the situation is {\it very different} as we see it below. Moreover, $S_s(m,j)$ allows us to write the full expression for the gravitational Kerr black hole entropy $S_{sem}$, as a function of the Bekenstein-Hawking entropy $S_{sem}^{(0)}$.\\ {\bf Extremal String States. A New Transition}: In ref. \cite {6} we considered a new kind of string states, the states in which $j$ reachs its {\bf maximal} value, that is $j = m^2 \alpha^{'} c$, we call these states {\bf "extremal string states"}. In this case, $\Delta_s = 0$. For $\Delta_s \rightarrow 0$, the entropy $S_s(m,j)$ behaves as $$ S_s(m,j)_{extremal} = -(a + 1)~k_B~ \ln \frac {\Delta_s}{2}~ +~ k_B~\ln 2~ +~ \Delta_s ~\Big(~\frac{3} {4}~S_s^{(0)}~-~ak_B \Big)~+ ~ O (\Delta_s^{2}) $$ In the extreme limit $(j/\hbar) \rightarrow (m/m_s)^2 $:~~$\Delta_{s ~extremal}= \sqrt{2}\sqrt{1 - \Big(\frac{j}{\hbar}\Big)^{1/2} ~\frac{m_s}{m}}$\\ and $S_s(m,j)_{extremal}$ is dominated by $$ S_s(m,j)_{extremal} = -(a + 1)~k_B~ \ln ~ \sqrt{1 - \Big(\frac{j}{\hbar}\Big)^{1/2} ~\frac{t_s}{T}}~~+~O(1) $$ This shows a {\bf phase transition} takes place at $T \rightarrow \sqrt{(j/\hbar)}~t_s$, we call it {\bf extremal} transition. This is {\bf not} the usual (Hagedorn/Carlitz) string phase transition occuring for $m \rightarrow \infty $, $T \rightarrow t_s$; such transition is also present for $j\neq 0$ since $\rho (m,j)$ has the same $m \rightarrow \infty $ behaviour as $\rho(m)$. The extremal transition we found here is a {\bf gravitational} like phase transition: the square root {\bf branch point} behaviour near the transition is analogous to that found in the thermal self-gravitating gas of (non-relativistic) particles (by both mean field and Monte Carlo methods), refs. \cite {8},\cite{9}. And this is also the same behaviour found for the microscopic density of states and entropy of strings in de Sitter background, refs. \cite{2}, \cite{7}.\\ As pointed out in ref.\cite {2}, this string behaviour is {\bf universal}: the logarithmic singularity in the entropy (or pole singularity in the specific heat) holds in any number of dimensions, and is due to the gravitational interaction in the presence of temperature, similar to the Jeans's like instability at finite temperature but with a more complex structure. A particular new aspect here is that the transition shows up at high angular momentum, (while in the thermal gravitational gaz or for strings in de Sitter space, angular momentum was not considered, (although it could be taken into account)). Since $j\neq 0$, the extremal transition occurs at a temperature $ T_{sj}~=~ \sqrt{j/\hbar}~T_s $, {\bf higher} than the string temperature $T_{s}$. That is, the angular momentum which acts in the sense of the string tension, appears in the transition as an "effective string tension" $(\alpha^{'}_j~)^{-1}$: a smaller $\alpha^{'}_j~ \equiv ~ \sqrt{\hbar/j}~ \alpha^{'}$ ~ (and thus a {\bf higher} tension).\\ {\bf Quantum String Entropy in de Sitter Background}: The entropy of quantum strings in de Sitter background is defined by $ \rho_{s}(m, H) = e^{\frac{S_{s}(m,H)}{k_B}}$, where $\rho_{s}(m, H)$ is the microscopic string density states of mass $m$ in de Sitter space time, and $H$ is the Hubble constant. In terms of the zero order string entropy in flat space time ($H=0$) $S_s^{(0)}$, the string density of mass states $\rho_s(m, H)$ for both open and closed strings can be expressed as :$$\rho_s (m, H) =\Big( \frac{S_s^{(0)}}{k_B} \sqrt{f(x)} \Big)^{-a}~ e^{\Big(\frac{S_s^{(0)}}{k_B}\sqrt{f(x)}\Big)}~F(x) ~~~~,~~~~ F(x) = \frac{1}{\Delta_s \sqrt{f(x)}},$$ ~~~~ $$f(x)= \frac{2}{1+\Delta_s}~~, ~~a=\frac{(D-1)}{2} ~~(open),~~ a=D ~~(closed) , $$ $x$~ being the dimensionless variable~~$x(m, H)\equiv \frac{1}{2}\Big(\frac{m}{M_s}\Big)= \frac{m_s}{b M_s}\frac{S_s^{(0)}}{k_B}$~~ and ~~$\Delta_s$~~ is given by: $$\Delta_s~~=~~\sqrt{1-4x^2}~~=~~\sqrt{1 -\Big(\frac{m}{M_s}\Big)^2}$$ $M_s$ is the highest string mass in de Sitter space time ~\cite{1},\cite{2},\cite{3},\cite{5}$$ M_s = \frac{L_{c\ell}}{\alpha'} = \frac{c}{H ~ \alpha'};~~~~\Bigg(\frac{m_s}{M_s} \Bigg) = \frac{\ell_{s}}{L_{c\ell}}= \frac{H}{c}\ell_{s}$$ $L_{c\ell}= c/H$ being the classical de Sitter radius. Furthermore, $M_s$ defines the quantum string de Sitter length $L_s$ and the de Sitter string temperature $T_s$: $$ L_s = \frac{\hbar }{M_s ~c}= \frac{\ell_s^2}{L_{c\ell}}=\frac{\hbar\alpha'}{c^2}H~~,~~ T_s = \frac{1}{2\pi k_B} ~ M_s~c^2 = \frac{\hbar c}{2 \pi k_B} ~\frac{1}{L_s}= \frac{1}{2 \pi k_B} ~\frac{c^3}{H \alpha'} $$ $T_s$ appears to be a true critical temperature as we will see below. \\ For small $x$, (small $H m\alpha '/c$), $f(x)$ can be naturally expressed as a power expansion in $x$. In particular, for $H=0$, we have $x=0$ and $f(x)=1$, and we recover the flat space time string solution: $ \rho_s(m, H=0) \simeq \Big( \frac{S_s^{(0)}}{k_B} \Big)^{-a} ~e^{\big( \frac{S_s^{(0)}}{k_B} \big)}$. From $\rho_s(m,H)$, we can read the full string entropy in de Sitter space : $$ S_s(m,H) = \hat {S_s}^{(0)}(m,H) -a~k_B~\ln ~\big(\frac{\hat {S_s}^{(0)}(m, H)}{k_B}\big) + k_B ~\ln F(m,H) $$ $$\hat {S_s}^{(0)}(m,H)\equiv S_s^{(0)}\sqrt{f(x)} $$ The mass domain is $ 0 \leq m \leq M_{s}$, ie. $ 0 \leq x \leq 1/2 $, (which implies $ 0\leq \Delta_s \leq 1$, ie. $1 \leq f(x)\leq 2 $). All terms in the entropy $S_s(m,H)$ except the first one have negative sign. For $\Delta_s \neq 0$, (ie. $ m \neq M_{s}$), the entropy $S_s(m,H)$ of string states in de Sitter space is smaller than the string entropy for $H=0$. The effect of the Hubble constant is to reduce the entropy.\\ For low masses $m \ll M_{s}$, the entropy is a series expansion in $(H m\alpha' /c \ll 1)$, like a low H expansion around the flat $H=0$ solution, $S_s^{(0)}$ being its leading term. But for high masses $m \rightarrow M_s$, that is $(H m \alpha^{'}/c) \rightarrow 1$, (i.e. $\Delta_s \rightarrow 0$), the situation is {\bf very different} as we see it below. Moreover, $S_s(m,H)$ allows us to write the whole expression for the semiclassical de Sitter entropy $S_{sem}(H)$, as a function of the (Bekenstein-Hawking) de Sitter entropy $S_{sem}^{(0)}(H)$.\\ {\bf The String de Sitter Phase Transition}: For high masses $m \sim M_s$, (or in terms of temperature $T \sim T_s$), the entropy behaves as: $$ S_s(T,H)_{T \sim T_s} = k_B \ln \sqrt{1 - \frac{T}{T_s}}~-k_B\ln~2 ~+~k_B \frac{ b}{\sqrt{2}}~(\frac{T_s}{t_s})~-~ a k_B \ln~(\frac{T_s}{t_s}) $$ where ~ $T = m c^2/2 \pi k_B $. We see that a phase transition takes place at $m = M_s$, ie, $T = T_s$. This is a {\bf gravitational} like phase transition: the square root {\bf branch point} behaviour near the transition is analogous to the thermal self-gravitating gas phase transition of point particles, refs. \cite {8},\cite{9}. This is also the same behaviour of the microscopic density of states and entropy of strings with the spin modes included ref. \cite{6}. \\ The transition occurs at the string de Sitter temperature $T_{s}$ {\bf higher} than the (flat space) string temperature $t_{s}$: $(T_s/t_s)= (b M_s /2\pi m_s) = (b L_{c\ell}/2\pi \ell_s)$. This is so since in de Sitter background, the flat space-time string mass $m_s$, (or Hagedorn temperature $t_s$) is the scale mass, (or scale temperature), in the {\bf low} $Hm$ regime. For high masses, the critical string mass, (temperature), in de Sitter background is $T_s$, instead of $m_s$, $(t_s)$. In de Sitter space, $H$ "pushes" the string temperature beyond its flat space Hagedorn value $t_s$ . That is, $H$, which acts in the sense of the string tension, does appear in the transition temperature as an "effective string tension" $(\alpha^{'}_H) ^{-1} $ : a smaller $\alpha^{'}_H = (\hbar/c)(~H \alpha'/c)^2$, (and thus a {\bf higher} tension). The effect of $H$ in the transition is similar to the effect of angular momentum $j$. \\ Furthemore, for high masses ($m \sim M_s$), the {\bf partition function} $\ln Z$ of a {\bf gas of strings} in de Sitter space-time, behaves as $$ (\ln Z)_{T \sim T_s}\sim V_{D-1}\left(\frac{k_B T_{sem}}{\hbar c}\right)^{D-1}~~ \sqrt{1 - \frac{T_{sem}}{T_s}} $$ where $T_s$ is the string de Sitter temperature and $ T_{sem}$ is the semiclassical (Gibbons-Hawking) de Sitter temperature. $\ln Z$ shows a singular behavior for $T_{sem} \rightarrow T_{s}$ which is general for {\bf any} space-time dimensions $D$. Again, this is a {\bf square root branch point} at $T_{sem}= T_s$. That is, a {\bf phase transition} takes place for $T_{sem} \rightarrow T_s$, which implies $M_{c\ell}\rightarrow m_{s}$, $L_{c\ell}\rightarrow \ell_{s}$.\\ Furthemore, the high mass behaviour of $\ln Z$ implies that $T_{sem}$ has to be bounded by $T_s$, $(T_{sem} < T_s)$, which means: $L_{c\ell} > \ell_s, ~~i.e., ~~H < {c}/{\ell_s}$. In the de Sitter string phase transition $T_{sem} \rightarrow T_{s}$, $H$ reachs a maximum value sustained by the string tension $\alpha'^{-1}$ (and the fundamental constants $\hbar$, $c$ as well): $$H_s = c ~\sqrt{\frac{c}{\alpha'\hbar}}, ~~~~ i.e.,~~\Lambda_s = \frac{1}{2 {\ell_s}^2}(D-1)(D-2)$$ The highly excited $m \rightarrow M_s$ string gas in de Sitter space undergoes a phase transition at high temperature $T_{sem} \rightarrow T_{s}$, into a condensate stringy state. This means that the background itself becames a string state. \\ These results also allow to consider the string regimes of a black hole in a de Sitter (or asymptotically) de Sitter background. This allow to study the effects of the cosmological constant $\Lambda$ on the quantum string emission by black holes, and the string bounds on the semiclassical (Gibbons-Hawking) {\bf black hole-de Sitter} (bhdS) temperature~ $T_{sem~bhdS}$, ref. \cite{7}. We computed the quantum string emission cross section $\sigma_{string}$ by a black hole in de Sitter (or asymptotically de Sitter) space-time (bhdS), ref. \cite{7}. For $T_{sem~bhdS}\ll T_{s}$, (early evaporation stage), it shows the QFT Hawking emission with temperature $T_{sem~bhdS}$, (semiclassical regime). For $T_{sem~bhdS}\rightarrow T_{s}$, $\sigma_{string}$~exhibits a phase transition into a string de Sitter state of size $L_s = \ell_s^2/L_{c\ell}$, ($\ell_s= \sqrt{\hbar \alpha'/c}$), and string de Sitter temperature $T_s$. For high masses ($m \sim M_s$), we find for the $\sigma_{string}$ leading behavior: $$ \sigma_{string} ~~(T \sim T_s) \sim V_{D-1}~\Gamma_A~\left(\frac{k_B T_{sem~bhdS}}{\hbar c}\right)^{D-1} \sqrt{1~-~\frac{T_{sem~bhdS}}{T_s}} $$ Instead of featuring a single pole singularity in the temperature (Carlitz transition), it features a square root {\bf branch point} (de Vega-Sanchez transition, refs.\cite{8},\cite{9}). {\bf New} bounds on the black hole radius $r_g$~ emerge in the bhdS string regime: it can become $r_g = L_s/2$, or it can reach a more quantum value, $r_g = 0.365 ~\ell_s$, ref. \cite{7}.\\ {\bf Semi-classical and quantum (string) black hole regimes}: Our analysis of the string canonical partition function, the black hole quantum string emission and the string bounds on the black hole, shows that a semiclassical black hole (BH) with size $L_{c\ell}$, mass $M$, temperature $ T_{sem}$, density of states $ \rho_{sem}$ and entropy $S_{sem}$, namely $(BH)_{sem}= (L_{c\ell}, M, T_{sem}, \rho_{sem}, S_{sem})$, evolves through evaporation into a quantum string state of size $L_{s}$, mass $m$, temperature $T_{s}$, density of states $\rho_{s}$ and entropy $S_{s}$, namely $(BH)_{s}$ = $(L_{s}, m , T_{s}, \rho_{s}, S_{s})$. The quantities in the set $(BH)_{sem}$ are precisely the semiclassical expressions of the respective ones in the set $(BH)_{s}$. In the quantum string regime, the black hole size $L_{c\ell}$ becomes the string size $L_{s}$, the Hawking temperature $T_{sem}$ becomes the string temperature $T_s$, the black hole entropy $S_{sem}$ becomes the string entropy $S_s$. The sets $(BH)_{s}$ and $(BH)_{sem}$ are the same quantities but in different (quantum and semiclassical/classical) domains. That is, $(BH)_{s}$ is the quantum dual of the semiclassical set $(BH)_{sem}$ in the precise sense of the wave-particle (de Broglie) duality. This is the usual classical/quantum duality but in the gravity domain, which is {\bf universal}, not linked to any symmetry or isommetry nor to the number or the kind of space-time dimensions. From the semiclassical and quantum (string) black hole regimes $(BH)_{sem}$ and $(BH)_{s}$, we can write the full gravity entropy $S_{sem}(M, J)$ for the {\bf Kerr black hole} such that it becomes the string entropy $S_s(m,j)$ in the quantum string regime, namely: $$ S_{sem}(M, J) = S_{sem}^{(0)}(M, J) - a ~k_B~\ln \Big (~\frac{S_{sem}^{(0)}(J, M)}{k_B}~ \Big)~+~k_B~\ln~ F(S_{sem}^{(0)}, J) $$ where $S_{sem}^{(0)}(M, J)$ is the Kerr black hole Bekenstein-Hawking entropy and $F(S_{sem}^{(0)}, J)$ is given by $$ F = \Delta^{-1} \Big(\frac{1 + \Delta}{2\Delta}\Big)^{a}~e^{\Big(\frac{1 -\Delta}{1 +\Delta}\Big) \frac{S_{sem}^{(0)}(M, J)}{\Delta k_B}} \cosh^{-2} \Bigg(\frac{(1 - \Delta)}{\Delta}\frac{S_{sem}^{(0)}(M, J)}{2 k_B}\Bigg) $$ For $J=0:~~~~~~F = 1 ~~~~, ~~~~ S_{sem}^{(0)}(M,J=0)~\equiv~ S_{sem}^{(0)} ~=~ 4 \pi k_{B} \Big(\frac{M}{m_{Pl}}\Big)^2$ $S_{sem}^{(0)}$ being the Schwarzschild black hole Bekenstein-Hawking entropy. $\Delta$ is given by : $$ \Delta = \sqrt{1 - \Big(\frac{J}{\hbar}\Big)^2 \Big(\frac{4 \pi k_B~}{S_{sem}^{(0)}} \Big)^2}=~ \sqrt{1 - \Big(\frac{J}{\hbar}\Big)^2 \Big(\frac{m_{Pl}}{M} \Big)^4}~=~ \sqrt{1 - \Big(\frac{J}{\hbar}\Big)^2 \Big(\frac{T_{sem}}{T} \Big)^2} $$ where $T = M c^2 / 8\pi k_B$ and $T_{sem}$ is the Schwarzshild Hawking temperature. $m_{Pl}$ is the Planck mass. Notice the {\bf new} term $k_B\ln F(S_{sem}^{(0)},J)$ in $S_{sem}(M, J)$ enterely due to $J \neq 0$. For $\Delta \neq0$, the effect of the angular momentum is to reduce the entropy. $S_{sem}(M, J)$ is maximal for $J=0$ (ie, for $\Delta = 1$); for $ J=0$ we have: $S_{sem}(M)~ = ~S_{sem}^{(0)}~-~ a ~k_B~\ln S_{sem}^{(0)}$.\\ $S_{sem}(M,J)$ provides the full Kerr black hole entropy as a function of the Kerr Bekenstein-Hawking entropy $S_{sem}^{(0)}(M,J)$. Interestingly enough, the full Kerr black hole entropy $S_{sem}(M, J)$ can be also written enterely in terms of the Schwarschild Bekenstein-Hawking entropy $S_{sem}^{(0)}$, this is done in ref. \cite{6}. For $M\gg m_{Pl}$ and $J < GM^2/c$, $S_{sem}^{(0)}$ is the leading term of this expression, {\bf but} for high angular momentum, (nearly extremal or extremal case $J= GM^2/c$), a gravitational {\bf phase transition} operates and the whole entropy $S_{sem}$ is drastically {\bf different} from the Bekenstein-Hawking entropy $S_{sem}^{(0)}$, as we precisely see it below.\\ {\bf A New feature. The Extremal Black Hole Phase Transition:} When $J$ reachs its {\bf maximal} value, that is $J = M^2 G/c^2$, then $\Delta = 0$ and the term $S_{sem}^{(0)}(M,J)$ is minimal : $S_{sem}^{(0)} (M,J)_{extremal} ~ = ~\frac{1}{2}~S_{sem}^{(0)}$. But for $\Delta \rightarrow 0$, the last term in the expression for $S_{sem}(M,J)$ substracts the first one, the poles in $\Delta$ cancel out, yielding: $$ S_{sem}(M,J)_{extremal} = -(a + 1)~k_B~ \ln \frac {\Delta}{2}~ +~ k_B~\ln 2~ +~ \Delta ~\Big(~\frac{3} {4}~S_{sem}^{(0)}~-~ak_B~\Big)~+ ~ O(\Delta^{2}) $$ In the extreme limit $(J/\hbar) \rightarrow (M/m_{Pl})^2 $ :~~ $\Delta_{extremal} = ~ 2 ~\sqrt{1 - \Big(\frac{J}{\hbar}\Big)^{1/2}~ \frac {m_{Pl}}{m}}$\\and $S_{sem}(M,J)_{extremal}$ is dominated by $$ S_{sem}(M,J)_{extremal} = -(a + 1)~k_B~ \ln ~\sqrt{1 - \Big(\frac{J}{\hbar}\Big)^{1/2}~\frac{t_{Pl}}{T}}~+~O(1) $$ This shows that a {\bf phase transition} takes place at $T \rightarrow \sqrt{(J/\hbar)}~t_{Pl} $, equivalently , at $T \rightarrow (J/\hbar)~T_{sem} $, we call it {\bf extremal transition}. $t_{Pl}$ is the Planck temperature. The characteristic features of this gravitational transition can be discussed on the lines of the extremal string transition we analysed for the extremal string states. Our discussion on the string case translates into the respective black hole quantities but we do not extend on these new features and implications here.\\ {\bf Semi-classical and quantum (string) de Sitter regimes}: From the microscopic string density of mass states $\rho_s(m, H)$, we have shown that for $m\rightarrow M_s$,~ i.e. $T\rightarrow T_s$, the string undergoes a phase transition into a semiclassical phase with mass $M_{cl}$ and temperature $T_{sem}$. Conversely, from the string canonical partition function in de Sitter space and from the quantum string emission by a black hole in de Sitter space, we have shown that for $T_{sem}\rightarrow T_s$, the semiclassical (Q.F.T) regime with Hawking-Gibbons temperature $T_{sem}$ undergoes a phase transition into a string phase at the string de Sitter temperature $T_{s}$. This means that in the quantum string regime, the semiclassical mass density of states $\rho_{sem}$ and entropy $S_{sem}$ become the string density of states $\rho_s$ and string entropy $S_{s}$ respectively. Namely, a semiclassical de Sitter state, $(dS)_{sem}= (L_{c\ell}, M_{c\ell}, T_{sem}, \rho_{sem}, S_{sem})$, undergoes a phase transition into a quantum string state $(dS)_{s}$ = $(L_{s}, m , T_{s}, \rho_{s}, S_{s})$. The sets $(dS)_{s}$ and $(dS)_{sem}$ are the same quantities but in different (quantum and semiclassical/classical) regimes. This is the usual classical/quantum duality but in the gravity domain, which is {\bf universal}, not linked to any symmetry or isommetry nor to the number or the kind of dimensions. From the semiclassical and quantum de Sitter regimes $(dS)_{sem}$ and $(dS)_{s}$, we can write the full de Sitter entropy $S_{sem}(H)$, with quantum corrections included, such that it becomes the de Sitter string entropy $S_s(m,H)$ in the string regime: the full de Sitter entropy $S_{sem}(H)$ is given by $$ S_{sem}~(H) = \hat{S}_{sem}^{(0)}~(H) -a~k_B~\ln ~(\frac{\hat{S}_{sem}^{(0)}~(H)}{k_B})+ k_B \ln~F(H)$$ $$\hat{S}_{sem}^{(0)}~(H)\equiv S_{sem}^{(0)}~(H) \sqrt{f(X)}~~~,~~~F(H) = \Delta \sqrt{f(X)}~~,~~ a=D~~,~~ f(X)= \frac{2}{1+\Delta},$$ $$\Delta ~\equiv~\sqrt{1-4X^2}~=~ \sqrt{1 -\Big(\frac{\pi k_B}{S_{sem}^{(0)}(H)}\Big)^2}, ~~X(H)\equiv \frac{\pi k_B}{2 S_{sem}^{(0)}(H)}= \frac {M_{sem}}{M_{cl}}= \Big(\frac{m_{Pl}}{M_{cl}}\Big)^2. $$ $S_{sem}^{(0)}(H)$ is the usual Bekenstein-Hawking entropy of de Sitter space. $M_{cl}= c^3/GH $ is the de Sitter mass scale, $M_{sem}= m_{pl}/M_{cl}$ is the semiclassical mass. \\ In this expression, the mass domain is $m_{pl} \leq M_{c\ell} \leq \infty$, that is, $0\leq X \leq 1/2$. (ie. $0\leq \Delta \leq 1$). The same formula but with $\hat{X}(H) = S_{sem}^{(0)}(H)/2\pi k_B$, instead of $X(H)$, describes $S_{sem}(H)$ in the mass domain $0 \leq M_{cl} \leq m_{pl}$. This provides the whole de Sitter entropy $S_{sem}(H)$ as a function of the Bekenstein-Hawking entropy $S_{sem}^{(0)}(H)$.\\ The limit $X \rightarrow 0$ corresponds to $M_{cl}\gg m_{Pl}$, that is $L_{c\ell}\gg \ell_{Pl}$ , (low $H \ll c/\ell_{Pl}$ or low curvature regime). In this limit, $\Delta \rightarrow 1$, $f(X)\rightarrow 1$ and $S_{sem}^{(0)}(H)$ is the leading term of $S_{sem}(H)$, with its logarithmic correction: $$ S_{sem}(H) = S_{sem}^{(0)}(H) -ak_B~\ln \Big(\frac{S_{sem}^{(0)}(H)}{k_B}\Big) $$ But for {\bf high} Hubble constant, $H \sim c/\ell_{Pl}$, (ie. $M_{cl}\sim m_{Pl}$), (that is {\bf high} curvature or quantum gravity regime), $S_{sem}^{(0)}(H)$ is sub-dominant, a gravitational {\bf phase transition} operates and the whole entropy $S_{sem}(H)$ is drastically {\bf different} from the Bekenstein-Hawking entropy $S_{sem}^{(0)}(H)$, as we precisely see below.\\ {\bf The de Sitter Gravitational Phase Transition} : For $\Delta \rightarrow 0$, that is for $ M_{c\ell}\rightarrow M_{sem}$, the full de Sitter entropy $S_{sem}(H)$ behaves as:$$ S_{sem}(H)_{\Delta \sim 0}~~ =~~ k_B~ \ln \Delta ~+~O(1)$$ In this limit, the Bekenstein-Hawking entropy $S_{sem}^{(0)}(H)$ is sub-leading, O(1), (~$S_{sem}^{(0)}(H)_{\Delta = 0} = \pi k_B$~). In terms of the mass, or temperature:$$ \Delta~ =~ \sqrt{1 - \Big(\frac{m_{Pl}}{M_{cl}} \Big)^4}~=~ \sqrt{1 - \Big(\frac{T_{sem}}{T} \Big)^2} ~~,~~~~T = \frac{1}{2\pi k_B}M_{cl} c^2$$ $T_{sem}$ is the semiclassical (Gibbons-Hawking) de Sitter temperature . In the limit $ M_{c\ell}\rightarrow M_{sem}$, which implies $ M_{cl} \rightarrow m_{Pl} $, $S_{sem}(H)$ is dominated by $$ S_{sem}(H)_{\Delta \rightarrow 0} = -~k_B~ \ln ~\sqrt{1 - \frac{T_{sem}}{T}}~~+~O(1) $$ This shows that a {\bf phase transition} takes place at $T \rightarrow T_{sem}$. This implies that the transition occurs for $M_{cl} \rightarrow m_{Pl}$, ie $T \rightarrow t_{Pl}$, (that is for high $H \rightarrow c/\ell_{Pl}$). This is a {\bf gravitational} like transition, similar to the de Sitter string transition we analysed above: the signature of this transition is the square root {\bf branch point} behavior at the critical mass (temperature) {\bf analogous} to the thermal self-gravitating gas behavior of point particles, refs \cite {8}, \cite{9}, and to the string gas in de Sitter space, ref.\cite {7}. This is also the same behaviour as that found for the entropy of the Kerr black hole in the high angular momentum $ J\rightarrow M^2G/c$ regime, (extremal transition), ref.\cite{6}. This is {\bf universal}, in any number of space-time dimensions.