diff --git "a/density of states/1.json" "b/density of states/1.json" new file mode 100644--- /dev/null +++ "b/density of states/1.json" @@ -0,0 +1 @@ +[ { "title": "0704.3218v2.Density_of_bulk_trap_states_in_organic_semiconductor_crystals__discrete_levels_induced_by_oxygen_in_rubrene.pdf", "content": "arXiv:0704.3218v2 [cond-mat.mtrl-sci] 19 Jun 2007Density of bulk trap states in organic semiconductor crysta ls:\ndiscrete levels induced by oxygen in rubrene\nC. Krellner,∗S. Haas, C. Goldmann,†K. P. Pernstich, D. J. Gundlach,‡and B. Batlogg§\nLaboratory for Solid State Physics, ETH Zurich, CH-8093 Zur ich, Switzerland\n(Dated: October 26, 2018)\nAbstract\nThe density of trap states in the bandgap of semiconducting o rganic single crystals has been\nmeasured quantitatively and with high energy resolution by means of the experimental method of\ntemperature-dependent space-charge-limited-current sp ectroscopy (TDSCLC). This spectroscopy\nhas been applied to study bulk rubrene single crystals, whic h are shown by this technique to be\nof high chemical and structural quality. A density of deep tr ap states as low as ∼1015cm−3is\nmeasured in the purest crystals, and the exponentially vary ing shallow trap density near the band\nedge could be identified (one decade in the density of states p er∼25meV). Furthermore, we have\ninduced and spectroscopically identified an oxygen-relate d sharp hole bulk trap state at 0.27 eV\nabove the valence band.\n1I. INTRODUCTION\nThe performance of organic thin-film transistors (OTFTs) is stead ily improving, and\nthe charge carrier mobility, as a key figure of merit, has reached va lues comparable to\nthat of hydrogenated amorphous silicon.1,2,3,4,5As with most semiconductors, the electrical\nperformance is determined to a high degree by modifications of the id eal crystal, such as\nintentional doping and other chemical or structural effects which create electrically active\nstates in the band gap. For a thorough understanding of the intrin sic capabilities and\nlimitations of organic semiconductors, it is now highly desirable to quan titatively study\nsuch in-gap states, their density of states (DOS) spectrum, the ir origin, and their stability.\nField-effect transistors (FETs) on organic single crystals are well s uited for determining\nthe surface properties, with mobilities near 20cm2/V s routinely achieved (for a review see\nGershenson et al.6). Bulk properties, however, cannot be determined with this surfa ce-\nsensitive method and thus one has to use alternative techniques, s uch as the time-of-flight\nmethod as shown in the pioneering work by Karl and co-workers,7thermally stimulated\ncurrent, or space-charge-limited-current (SCLC) measuremen ts. For instance, SCLC mea-\nsurements with coplanar electrodes were used by Lang et al.8to detect a metastable trap\nstate in pentacene single crystals, where the SCLC changes over s everal orders of magnitude,\nindicating trap filling.\nAdditional experimental techniques are used to detect defect st ates in organic semicon-\nductors: Recently, oxygen-related states at the surface of na turally oxidized rubrene single\ncrystals were detected by photoluminescene measurements.9Kelvin-probe force microscopy\nnot only offers a direct imaging of the potential across the channel of an OTFT, but addi-\ntionally allows one to extract the DOS spectrum at the semiconducto r/insulator interface.10\nInthisstudy we go beyond thebasicconcept of SCLC andusethe sp ectroscopic character\nof temperature-dependent SCLC spectroscopy (TDSCLC) as de scribed in the theory by\nSchaueret al.11to derive the bulk DOS spectrum in ultrapure rubrene single crystals . We\nfound that these crystals of high structural and chemical quality have a broad distribution of\ndeep states with a low total density of trap states, in addition to a s teep band tail. We also\nuse the spectroscopic technique to identify the energetic position of oxygen-induced bulk\ntrap states in rubrene single crystals.\n2II. METHOD\nA central aspect of TDSCLC is to exploit the spectroscopic charac ter inherent in the\ntemperature dependence of the SCLC due to the energy window as sociated with the Fermi-\nDirac statistics.11It is assumed that the SCLC is dominated by the charge carriers tha t are\nthermally excited from a localized trap into delocalized band states. T herefore the valence\nband edge is seen as separating localized and delocalized states, and it is chosen as the\nreference point for the energy scale ( Ev= 0), with positive energy toward the midgap. The\nbasic equations are Ohm’s law in the form j=eµ0nf(x)F(x) and the Poisson equation\ndF/dx=−ens(x)/(ǫǫ0). HereF(x) is the electric field strength along the direction xof\ncurrent flow, jis the current density, ǫǫ0is the product of the dielectric constant and electric\npermittivity (3.5 for rubrene), µ0is the microscopic band mobility, nf(x) is the density of\nfree carriers, ns(x) is the total density of carriers (free and trapped), which is given by\nthe convolution of the density of trap states h(E) in the energy gap with the Fermi-Dirac\nfunction f(E,EF,T), i.e.,ns=/integraltext\nEh(E)f(E,EF,T)dE. The shape of the current-voltage\ncharacteristic j(U) reflects the increment of the space charge with respect to the s hift of the\nFermi energy and thus mirrors the energy dependence of the DOS ,\ndns\ndEF=1\nkBTǫǫ0\neL2(2m−1)\nm2(1+C) (1)\nwith\nC=B(2m−1)+B2(3m−2)+d[ln(1+B)]/dlnU\n1+B(m−1). (2)\nHereListhethickness ofthecrystal withelectrodesonoppositefaces, Utheappliedvoltage,\nkBthe Boltzmann constant, m=dlnj/dlnUthe logarithmic slope of the j(U) curve, and B\ncontains higher-order derivatives of j(U),B=−[dm/dlnU]/[m(m−1)(2m−1)]. The right\nhand side of Eq. (1) can be calculated from the current-voltage ch aracteristic measured at\nonly one temperature. For a complete reconstruction of the DOS, however, it is necessary\nto relate a given voltage Uto the energy of states which are being filled at this value of U.\nA first starting point is to extract an activation energy EAfrom the Arrhenius plot of the\nTDSCLC data for a given U, i.e.EA=−dlnj/d(kBT)−1. Because h(E) in general is not\nsymmetric around EF,12the energy EDof the incremental change of space charge is slightly\nshifted from EA, typically by a fraction of kBT13\nED=EA+(3−4m)n\n(2m−1)(m−1)mkBT. (3)\n3Heren=−d(EA/kBT)/dlnUis the derivative of the activation energy with respect to the\napplied voltage. To extract the DOS from the shape of the j(U) curves, it is necessary to\ndeconvolute Eq. (1) with respect to df/dE F, using a high accuracy deconvolution method\nbased on spline functions.14\nIII. EXPERIMENTAL DETAILS\nTherubrenecrystalsweregrownbyphysicalvaportransport15underastreamofultrapure\nArgon 6.0. The starting material (Aldrich purum) was sublimed three times in vacuum.\nConsiderable effort was made to avoid contamination with any organic substances, e.g.,\nby using glass tubes cleaned in acids. Typical crystals are platelike, w here the direction\nperpendicular to the surface corresponds to the long axis of the o rthorhombic unit cell.16\nWe note that these high-quality rubrene single crystals show in-plan e field-effect mobilities\nat the surface of up to 10 cm2/V s.20\nAn optimized sample preparation method was found in a slightly adapte d “flip-crystal”\napproach,21benefiting from the minimized sample handling. The thin crystals (pref erred\nthickness <2µm) are placed on glass substrates with 20nm Au on 5nm Cr electrode s tripes,\nwhere they stick by electrostatic adhesion. A 20nm gold top electro de is then evaporated\nonto the crystal, which is slightly cooled during the evaporation ( Tmask=−8◦C) to minimize\nthermal damage. The overlap oftheelectrodes results inatypical measurement crosssection\nofA∼2·10−5cm2. Finally, electrical connections to a sample holder were made with silve r\nepoxy and 25- µm-thick gold wires. The thickness of the crystals was measured by a tomic\nforce microscopy, as optical inspection turned out to be unreliable for the ultrathin crystals.\nThe electrical measurements were performed in a closed-cycle cry ostat in inert helium\natmosphere and in darkness, covering a maximal temperature ran ge of 30–350K. By the use\nof helium exchange gas, a constant and reliable sample temperature , measured at the back\nsideofthesampleholder, couldbeadjustedwitharesistiveheaterc oil(drivenbyaLakeshore\n331 controller). A Keithley 6517A electrometer was used for the ele ctrical measurements.\nBy a proper shielding, leakage currents well below 1pA at 100V were a chieved. To inject\nholes from the bottom contact, a negative voltage was applied at th e top contact; V >0\nresults in hole injection from the top contact. The superiority of lam inated vs evaporated\ncontacts was reported previously.22,23\n4The typical measurement procedure for TDSCLC was as follows. Fir st, initial tests at\nroom temperature were performed in order to check the reprodu cibility of repeated mea-\nsurements of the j(U) characteristic. Thereafter the sample was cooled to typically 100 K\nat a rate of 2K/min. Prior to the measurement of the current-volt age characteristic (at\neach chosen temperature), an initial delay of 20min ensured therm al equilibrium. The\nj(U) curves were measured by a stepwise increase of the applied voltag e and measuring the\nquasiequilibrium current. A current measurement delay of 10s turn ed out to be sufficiently\nlong compared to the settling time of the system. Typically 50 points w ere measured per\nvoltage decade. The maximal voltage was limited such that the curre nt density would not\nexceed 0.1A/cm2in order to avoid crystal damage.\nUsuallyj(U) was measured every 10K between 100 and 200K. At lower tempera tures\nit is difficult to get accurate j(U) curves, because the current increases extremely rapidly\nwith voltage. At higher temperatures, the broadening of the Ferm i statistics has an adverse\ninfluence on the spectroscopic character of TDSCLC. In particula r, shallow states can only\nbe reliably measured at low temperature.\nThe issue of possible long-term charge trapping as result of a j(U) sweep needs careful\nattention during the course of a measurement. In most crystals, subsequent sweeps at the\nsame low temperature yield identical curves. This indicates a detrap ping of the charge on\na time scale faster than that of a sweep ( ∼10 min). For a few samples, however, it was\nnecessary to warm the crystal to room temperature after ever y sweep in order to restore the\ninitial condition. Subsequent j(U) sweeps at the same low temperature then yield the same\nresults. If charge trapping or sample deterioration influence the m easurement, the current\nat a given fixed voltage will not be thermally activated (as opposed to the Arrhenius plot in\nFig. 2). Therefore, the measurement itself is a valuable self-consis tency test.\n5/s48/s46/s49 /s49 /s49/s48/s49/s48/s45/s49/s48/s49/s48/s45/s56/s49/s48/s45/s54/s49/s48/s45/s52/s49/s48/s45/s50\n/s48/s46/s48/s49 /s48/s46/s49 /s49/s49/s48/s45/s54/s49/s48/s45/s52/s49/s48/s45/s50\n/s84 /s32/s61/s32/s51/s48/s48/s32/s75/s82/s117/s98/s114/s101/s110/s101\n/s115/s105/s110/s103/s108/s101/s32/s99/s114/s121/s115/s116/s97/s108\n/s84 /s32/s61/s32/s50/s48/s48/s32/s75/s84 /s32/s61/s32/s49/s49/s48/s32/s75\n/s32/s32/s106/s32/s91/s65/s47/s99/s109/s50\n/s93\n/s85 /s32/s91/s86/s93\n/s32/s32\n/s32/s32\nFIG. 1: Space-charge-limited-current (SCLC) density vs ap plied voltage at different temperatures\nfor a rubrene single crystal (Ru65-2, L= 0.6µm). The temperature step is 10 K. The inset\nshowsj(U) at 300 K. The straight line indicates the Ohmic behavior of t hermally generated charge\ncarriers.\nIV. EXPERIMENTAL RESULTS\nA. DOS in rubrene single crystals\nIn Fig. 1, the j(U) curves at different temperatures are shown, measured perpen dicular\nto the molecule layers in a rubrene single crystal (Ru65-2, L= 0.6µm). The “Ohmic”\nregion at low voltages is indicated by a straight line (see inset of Fig. 1) . The onset of SCLC\nat∼0.1V indicates that holes are effectively injected from the laminated go ld contact. At\nlower temperatures the number of thermally excited carriers decr eases exponentially and the\nOhmic region disappears below the sensitivity of the measurement se tup (∼10−14A).\nIn Fig. 2 the Arrhenius plots of the current density are shown for t he same data set as\nin Fig. 1. For clarity we show only data points for selected voltages. T he slope yields the\nactivation energy EA(U) for each applied voltage Uand we note the high quality of these\nplots. The resulting EA(U) is given in the inset and is in agreement with recent FET data24.\nThe effective Fermi energy is moved from ∼0.45eV at the lowest voltage to ∼0.1eV at the\nhighest injection voltage. The distinct change of slope near 0.27eV is worth noting, as it\nobviously reflects amarked increase ofthetrapdensity. Wenotet hat, amongotherevidence,\nthe smooth variation of EAat lowUindicates space-charge-limited-transport rather than\n6/s53 /s54 /s55 /s56 /s57/s49/s48/s45/s56/s49/s48/s45/s54/s49/s48/s45/s52/s49/s48/s45/s50/s49/s48/s48\n/s49 /s49/s48/s48/s46/s49/s48/s46/s50/s48/s46/s51/s48/s46/s52\n/s48/s46/s50/s54/s48/s46/s52/s50/s48/s46/s55/s50/s49/s46/s54/s82/s117/s35/s54/s53/s45/s50\n/s32/s32/s32/s106/s32/s91/s65/s47/s99/s109/s50\n/s93\n/s49/s48/s48/s48/s47 /s84 /s32/s91/s75/s45/s49\n/s93/s85 /s91/s86/s93\n/s51/s46/s49/s69\n/s65/s32/s40 /s85 /s41\n/s32/s85 /s32/s91/s86/s93/s32/s69\n/s65/s32/s91/s101/s86/s93\nFIG. 2: Current density for a fixed applied voltage Uvs the inverse temperature for the data\nplotted in Fig. 1. The straight lines are the Arrhenius fits us ed to determine the activation energy\nEA(U). The inset shows the resulting EA(U).\ntransport limited by contacts.25\nThe analysis described above is used to extract the density of trap states from these\nTDSCLC measurements. In order to calculate higher-order deriva tives of the experimental\nj(U) andEA(U) curves according to Eq.(1), a smoothing spline fit was applied to th e\nmeasured data, keeping the fit within 1% of the raw data. The result ing density of trap\nstates after the deconvolution is shown in Fig. 3. The edge of the va lence band is used as\nthe reference level and the positive energy axis points to the cent er of the band gap. Three\nmain features can be discerned. (1) An exponential increase of th e DOS toward the band is\nobserved for all rubrene single crystals. However, the characte ristic energy kBTtover which\nthe DOS is reduced by a factor of evaries from sample to sample. The sample Ru52-3\nhas a broad exponential distribution with kBTt= 210meV and the highest density of trap\nstates. Broad tail states with kBTt= 180meV were recently reported for pentacene single\ncrystals26. The sample Ru65-1 with the lowest deep trap density of order 1015cm−1eV−1in\nthe energy range from 0.45 to 0.1eV has a characteristic exponent ial distribution parameter\nofkBTt= 180meV. (2) Of particular interest in this sample are the shallow tra p states\nbelow∼0.1eV with a steeper slope ( kBTt= 11meV), reminiscent of band tail states. Due\nto the large increase of the trap density the quasi Fermi level is pin ned close to the band\nedge, and the amount of injected charge cannot fill these shallow t ail states. In this energy\n7/s48/s46/s48 /s48/s46/s49 /s48/s46/s50 /s48/s46/s51 /s48/s46/s52 /s48/s46/s53/s49/s48/s49/s53/s49/s48/s49/s54/s49/s48/s49/s55/s49/s48/s49/s56\n/s82/s117/s35/s53/s50/s45/s51\n/s82/s117/s35/s54/s53/s45/s49/s82/s117/s35/s54/s53/s45/s50/s82/s117/s35/s53/s50/s45/s50\n/s32/s32/s68/s79/s83 /s32/s91/s99/s109/s45/s51\n/s101/s86/s45/s49\n/s93\n/s69 /s32/s91/s101/s86/s93\nFIG. 3: (Color online) Density of states above the valence ba nd for four rubrene single crystals.\nrange the charge transport is still activated ( EA∼0.05eV). (3) The fine structure of the\nDOS indicates small features for all crystals, due to discrete trap levels in the band gap.\nThe observed fine structure was confirmed using the energy depe ndence of the statistical\nshift.11\nIncomparison, data of organicthin films show a very similar qualitative DOSspectrum,10\nexhibiting tail states and exponentially increasing deep states, albe it at much higher con-\ncentrations than in the purest rubrene single crystals. Again, the FET geometry in Ref. 10\nmight have emphasized defects near the interface, because FET d evices made from the same\ntype of high quality rubrene single crystals show approximately thre e orders of magnitude\nhigher interface trap density than the bulk trap density presente d here.27\nIt is remarkable that the representative selection of rubrene cry stals vary in their DOS,\nalthough grown under basically identical conditions.28This variation may originate from\nindividual micro-conditions during and after crystal growth, i.e., diff erent actual growth\ntemperature, growth rate, (thermo)mechanical strain, and diff erent atmospheric conditions\nin the device fabrication process. On the other hand, measuremen ts of different cross sec-\ntions on the samecrystal result in virtually identical DOS spectra, which is a convincing\nverification of the data evaluation.\n8/s48/s46/s48 /s48/s46/s49 /s48/s46/s50 /s48/s46/s51 /s48/s46/s52 /s48/s46/s53/s49/s48/s49/s53/s49/s48/s49/s55/s49/s48/s49/s57/s49/s48/s49/s53/s49/s48/s49/s55/s49/s48/s49/s57/s48/s46/s48 /s48/s46/s49 /s48/s46/s50 /s48/s46/s51 /s48/s46/s52 /s48/s46/s53\n/s82/s117/s35/s55/s49/s45/s52\n/s110/s111/s116/s32/s105/s108/s108/s117/s109/s105/s110/s97/s116/s101/s100/s82/s117/s35/s54/s53/s45/s50\n/s105/s108/s108/s117/s109/s105/s110/s97/s116/s101/s100\n/s32/s82/s117/s98/s114/s101/s110/s101/s32\n/s32/s82/s117/s98/s114/s101/s110/s101/s32/s43/s49\n/s79\n/s50/s32/s43/s32/s79\n/s51\n/s32/s68/s79/s83 /s32/s91/s99/s109/s45/s51\n/s101/s86/s45/s49\n/s93\n/s69 /s32/s91/s101/s86/s93/s32\n/s32/s82/s117/s98/s114/s101/s110/s101/s32\n/s32/s82/s117/s98/s114/s101/s110/s101/s32/s43/s49\n/s79\n/s50\n/s32/s32\nFIG. 4: (Color online) Density of trap states in rubrene sing le crystals before and after exposure\nto1O2. The oxygen-induced defect acts as a hole trap at an energy of 0.27eV above the valence\nband edge. Sample Ru65-2 was illuminated in an oxygen atmosp here to form the1O2directly at\nthe rubrene crystal surface. Sample Ru71-4 was exposed to ox ygen excited by uv light.\nB. Oxygen-induced bulk trap states\nIn order to understand the role of bulk traps in organic single cryst als and to demonstrate\nthe power of TDSCLC spectroscopy, we investigated the influence of oxygen on the trap\ndensity of rubrene single crystals. It is well known that the reactio n of rubrene with singlet\noxygen1O2formstheendoperoxide.29Rubreneitselfcanactasthesensitizer whenthetriplet\nstate of rubrene is populated by an intersystem crossing from the singlet state excited with\nvisible light. The energy of this triplet state at 1.2 eV above the groun d state is transferred\nto the molecular oxygen resulting in1O2which reacts with the rubrene molecule.\nAfter measuring the full DOS in the as-grown crystals (open symbo ls in Fig. 4), we\nilluminated the sample Ru65-2 with visible light under oxygen atmospher e for four hours.\n9For comparison the sample Ru71-4 was directly exposed for four ho urs to1O2by exciting\nmolecular oxygen with uv light in the vicinity of the sample; the sample its elf was held\nin the dark. The exposure of rubrene to1O2results in a large peak in the density of\ntrap states at 0.27 eV above the valence band (filled symbols in Fig. 4) . The amount of\noxygen-induced trap states in sample Ru65-2 is estimated to be Nox\nt≈2·1017cm−3, which\ncorresponds to ∼100 ppm. Because the sample was illuminated in an oxygen atmosphere\nwithout removing from the cryostat we can exclude origins other th an oxygen for this hole\ntrap. We note that the energetic position of the identified oxygen t rap state is in agreement\nwith recently published photoluminescence measurements on oxydiz ed rubrene.9In sample\nRu71-4the oxygen-induced trapstates arewithin themeasureme nt error at thesame energy,\nbut the density of the trap states is slightly higher than for the illumin ated sample ( Nox\nt≈\n3·1017cm−3). In addition a shoulder appears in the energetic distribution on the side closer\nto the valence band. This difference may be due to the formation of O 3under uv light which\nmight also react with the rubrene single crystal.\nThe trap level at E= 0.27eV already exists in the pristine rubrene single crystals, with\nmuch lower concentrations. This is due to the device fabrication pro cess; the samples are\nhandled in room air under microscope illumination, which results in a similar reaction as\nfor the illuminated sample Ru65-2. The oxygen-related trap level wa s also observed in other\nrubrene samples with various concentrations, probably as a result of a different exposure\ntime to air during handling.\nV. CONCLUSIONS\nIn conclusion, we have implemented temperature-dependent spac e-charge-limited-current\nspectroscopy and demonstrate it to be a powerful tool to quant itatively measure the density\nof bulk trap states with high energy resolution. Applied to high-qualit y rubrene crystals\nthis method reveals the existence of states within ∼0.1 eV of the band edge, reminiscent of\nband tails, and a smooth distribution of states deeper in the gap. Dis crete peaks are also\nobserved, and through a controlled exposure of the crystals to a ctivated oxygen, a distinct\nand stable trap level at 0.27eV has been created. Applying this appr oach to other organic\nsemiconductors will be very helpful in the quest to identify intrinsic a nd extrinsic factors\nthat dominate charge transport in organic semiconductors.\n10Acknowledgments\nThe authors thank H.-P. Staub and K. Mattenberger for technica l assistance with crystal\ngrowth and the low-temperature measurements and D. Oberhoff f or helpful discussions and\nsupport with MATLAB.\n∗Present address: Max Planck Institute for Chemical Physics of Solids, D-01187 Dresden, Ger-\nmany\n†Present address: Philips Research Laboratories Aachen, D- 52066 Aachen, Germany\n‡Present address: National Institute for Standards and Tech nology (NIST), Gaithersburg, MD,\nUSA\n§Electronic address: batlogg@solid.phys.ethz.ch\n1G. Horowitz, Adv. Mater. 10, 365 (1998).\n2D. J. Gundlach, Y. Y. Lin, T. N. Jackson, S. F. Nelson, and D. G. Schlom, Electron Device\nLetters, IEEE 18, 87 (1997).\n3C. D. Dimitrakopoulos and P. R. L. Malenfant, Adv. Mater. 14, 99 (2002).\n4T. W. Kelley, D. V. Muyres, P. F. Baude, T. P. Smith, and T. D. Jo nes, inMat. Res. Soc.\nSymp. Proc. Vol. 771 (2003), p. L6.5.1.\n5H. Klauk, M. Halik, U. Zschieschang, G. Schmid, W. Radlik, an d W. J. Weber, J. Appl. Phys.\n92, 5259 (2002).\n6M. E. Gershenson, V. Podzorov, and A. F. Morpurgo, Rev. Mod. P hys.78, 973 (2006).\n7W. Warta and N. Karl, Phys. Rev. B 32, 1172 (1985).\n8D. V. Lang, X. Chi, T. Siegrist, A. M. Sergent, and A. P. Ramire z, Phys. Rev. Lett. 93, 076601\n(2004).\n9O. Mitrofanov, D. V. Lang, C. Kloc, J. M. Wikberg, T. Siegrist , W.-Y. So, M. A. Sergent, and\nA. P. Ramirez, Phys. Rev. Lett. 97, 166601 (2006).\n10O. Tal, Y. Rosenwaks, Y. Preezant, N. Tessler, C. K. Chan, and A. Kahn, Phys. Rev. Lett. 95,\n256405 (2005).\n11F. Schauer, S. Neˆ sp˚ urek, and H. Valeri´ an, J. Appl. Phys. 80, 880 (1996).\n12V. I. Arkhipov, P. Heremans, E. V. Emelianova, and G. J. Adria enssens, Appl. Phys. Lett 79,\n114154 (2001).\n13F. Schauer, R. Novotny, and S. Neˆ sp˚ urek, J. Appl. Phys. 81, 1244 (1997).\n14M. Deutsch and I. Beniaminy, Rev. Sci. Instrum. 53, 90 (1982).\n15R. A. Laudise, C. Kloc, P. G. Simpkins, and T. Siegrist, J. Cry st. Growth 187, 449 (1998).\n16The direction perpendicular to the surface is parallel to th e crystallographic a-17,18orc-axis19,\ndepending on the space group setting.\n17O. D. Jurchescu, A. Meetsma, and T. T. M. Palstra, Acta Crysta llogr. B62, 330 (2006).\n18B. D. Chapman, A. Checco, R. Pindak, T. Siegrist, and C. Kloc, J. Cryst. Growth 290, 479\n(2006).\n19V. C. Sundar, J. Zaumseil, V. Podzorov, E. Menard, R. L. Wille tt, T. Someya, M. E. Gershen-\nson, and J. A. Rogers, Science 303, 1644 (2004).\n20C. Goldmann, S. Haas, C. Krellner, K. P. Pernstich, D. J. Gund lach, and B. Batlogg, J. Appl.\nPhys.96, 2080 (2004).\n21J. Takeya, C. Goldmann, S. Haas, K. P. Pernstich, B. Ketterer , and B. Batlogg, J. Appl. Phys.\n94, 5800 (2003).\n22J. Zaumseil, K. W. Baldwin, and J. A. Rogers, J. Appl. Phys. 93, 6117 (2003).\n23R. W. I. de Boer, M. Jochemsen, T. M. Klapwijk, A. F. Morpurgo, J. Niemax, A. K. Tripathi,\nand J. Pflaum, J. Appl. Phys. 95, 1196 (2004).\n24V. Y. Butko, J. C. Lashley, and A. P. Ramirez, Phys. Rev. B 72, 081312R (2005).\n25S. Neˆ sp˚ urek, O. Zmeˆ skal, and F. Schauer, Phys. Status Sol idi A85, 619 (1984).\n26D. V. Lang, X. Chi, T. Siegrist, A. M. Sergent, and A. P. Ramire z, Phys. Rev. Lett. 93, 086802\n(2004).\n27C. Goldmann, C. Krellner, K. P. Pernstich, S. Haas, D. J. Gund lach, and B. Batlogg, J. Appl.\nPhys.99, 034507 (2006).\n28The samples Ru65-1 and Ru65-2 are from a different growth batch than the samples Ru52-2\nand Ru52-3, but with identical growth parameters.\n29T. Takahashi, Y. Harada, N. Sato, K. Seki, H. Inokuchi, and S. Fujisawa, Bull. Chem. Soc. Jpn.\n52, 380 (1979).\n12" }, { "title": "0705.1738v1.Equation_of_state_of_isospin_asymmetric_nuclear_matter_in_relativistic_mean_field_models_with_chiral_limits.pdf", "content": "arXiv:0705.1738v1 [nucl-th] 12 May 2007Equation of state of isospin-asymmetric nuclear matter in\nrelativistic mean-field models with chiral limits\nWei-Zhou Jiang1,2, Bao-An Li1, and Lie-Wen Chen1,3\n1Department of Physics, Texas A&M University-Commerce, Com merce, TX 75429, USA\n2Institute of Applied Physics, Chinese Academy of Sciences, Shanghai 201800, China\n3Institute of Theoretical Physics, Shanghai Jiao Tong Unive rsity, Shanghai 200240, China\nAbstract\nUsing in-medium hadron properties according to the Brown-Rho sca ling due to the chiral\nsymmetryrestorationat highdensities and consideringnaturalnes softhe couplingconstants,\nwe have newly constructed several relativistic mean-field Lagrang ianswith chiral limits. The\nmodelparametersareadjustedsuchthat thesymmetricpartof theresultingequationofstate\natsupra-normaldensitiesisconsistentwiththatrequiredbythec ollectiveflowdatafromhigh\nenergy heavy-ion reactions, while the resulting density dependenc e of the symmetry energy\nat sub-saturation densities agrees with that extracted from the recent isospin diffusion data\nfrom intermediate energy heavy-ion reactions. The resulting equa tions of state have the\nspecial feature of being soft at intermediate densities but stiff at h igh densities naturally.\nWith these constrained equations of state, it is found that the rad ius of a 1.4M⊙canonical\nneutron star is in the range of 11.9 km ≤R≤13.1 km, and the maximum neutron star mass\nis around 2.0 M⊙close to the recent observations.\nKeywords: Equation of state, nuclear matter, symmetry energy , relativistic mean-field\nmodels, chiral limits. PACS numbers: 21.65.+f, 26.60.+c, 11.30.Rd\n1 Introduction\nThe Equation of State (EOS) of isospin asymmetric nuclear ma tter plays a crucial role in many\nimportant issues in astrophysics, see, e.g., Refs. [1, 2, 3] . It is also important for understanding\nboth the structure of exotic nuclei and the reaction dynamic s of heavy-ion collisions, see, e.g.,\nRef. [4]. Within the parabolic approximation, the energy pe r nucleon in isospin asymmetric\nnuclear matter can be written as E/A=e(ρ)+Esym(ρ)δ2wheree(ρ) is the EOS of symmetric\nnuclearmatter, the Esym(ρ)isthesymmetryenergyand δ= (ρn−ρp)/ρistheisospinasymmetry.\nBoth thee(ρ) and theEsym(ρ) are important in astrophysics although maybe for different i ssues.\n1For instance, themaximum mass of neutronstars is mainly det ermined by theEOSof symmetric\nnuclear matter e(ρ) while the radii and cooling mechanisms of neutron stars are determined\ninstead mainly by the symmetry energy Esym[1, 5]. The nuclear physics community has been\ntrying to constrain the EOS of symmetric nuclear matter usin g terrestrial nuclear experiments\nfor more than three decades, see, e.g.[6], for a review. On th e other hand, a similarly systematic\nandsophisticated studyonthedensitydependenceofthesym metryenergy Esymusingheavy-ion\nreactions only started about ten years ago stimulated mostl y by the progress and availability of\nradioactive beams[7]. Compared to our current knowledge ab out the EOS of symmetric nuclear\nmatter, the symmetry energy Esymis still poorly known especially at supra-normal densities [8]\ngiven the recent progress in constraining it at densities le ss than about 1.2 ρ0using the isospin\ndiffusion data from heavy-ion reactions [9, 10, 11].\nThe aim of this work is to investigate the EOS of isospin asymm etric nuclear matter within\nthe relativistic mean-field (RMF) model with in-medium hadr on properties governed by the BR\nscaling. From the point of view of hadronic field theories, th e symmetry energy is governed\nby the isovector meson exchange. Studying in-medium proper ties of isovector mesons is thus\nof critical importance for understanding the density depen dence of the symmetry energy. We\nfirst construct model Lagrangians respecting the chiral sym metry restoration at high densities.\nThe model parameters are adjusted such that the symmetric pa rt of the resulting EOS at\nsupra-normal densities is consistent with that required by the collective flow data from high\nenergy heavy-ion reactions [6], while the resulting densit y dependence of the symmetry energy\nat sub-saturation densities agrees with that extracted fro m the recent isospin diffusion data\nfrom intermediate energy heavy-ion reactions [9, 10, 11]. T he constrained EOS is then used to\ninvestigate several global properties of neutron stars.\n2 Relativistic mean-field models with chiral limits\nIn-medium properties of the isovector meson ρcan be studied through the special symmetry\nbreaking and restoration. The local isospin symmetry in the Yang-Mills field theory, where the\nρmeson may be introduced as a gauge boson of the strong interac tion, can serve as a possible\ncandidatetostudythein-mediumpropertiesof ρmeson. However, since πNinteractionsactually\ndominate the strong interaction in hadron phase, it was rath er difficult to understand how the\nin-medium properties of the massive ρmeson could be consistent with the restoration of local\n2isospin symmetry. On the other hand, within the microscopic theory for the strong interaction,\nnamely, the QCD which is a color SU(3) gauge theory, the chira l symmetry is approximately\nconserved. The spontaneous chiral symmetry breaking and it s restoration can be manifested\nin effective QCD models. Based on the latter, Brown and Rho (BR) proposed the in-medium\nscaling law [12] implying that hadron masses and meson coupl ing constants in the Welacka\nmodel [13] approach zero at the chiral limit. The scaling was treated in the hadronic phase\nbefore the chiral symmetry restoration.\nAs an effective QCD field theory, the hidden local symmetry theo ry has been developed\nto include the ρmeson in addition to the pion in the framework of the chiral pe rturbative\ntheory by Harada and Yamawaki [14, 15] and it is shown that the ρmeson becomes massless\nat the chiral limit. This supports the mass dropping senario of the BR scaling. There are\nalso experimental indication for the mass dropping, i.e., t he dielectron mass spectra observed\nat the CERN SPS [16, 17], the ωmeson mass shift measured at the KEK [18] and the ELSA-\nBonn [19], as well as the analysis of the STAR data [20, 21]. Ho wever, data from the NA60\nCollaboration for the dimuon spectrum[22] seem to favor the explanation of ρmeson broadening\nbased on a many-body approach [23]. So far, the controversy i s still unsettled [21]. The chiral\nsymmetry and its spontaneous breaking are closely related t o the mass acquisition and dropping\nof hadrons. Since the chiral symmetry is a characteristic of the strong interaction within the\nQCD, it is favorable to include in the RMF models effects of the c hiral symmetry through\nthe BR scaling law. However, this does not mean that the contr ibution of the many-body\ncorrelations [24] is excluded. Actually, the contribution of the many-body correlations can be\nincluded phenomenologically into the RMF models to reprodu ce the saturation properties of\nnuclear matter.\nThe in-medium ρmeson plays an important role in modifying the density depen dence of the\nsymmetry energy. For most RMF models that the ρmeson mass is not modified by the medium,\nthe symmetry energy is almost linear in density. The introdu ction of the isoscalar-isovector\ncoupling in RMF models can soften the symmetry energy at high densities [3]. Meanwhile,\nit reproduces the neutron-skin thickness in208Pb as that given by the non-relativistic models\n(about 0.22fm) [25], consistent with the available data [26 ]. In well-fitted RMF models that\ngive a value of incompressibility κ= 230 MeV, a large coefficient of non-linear self-interacting\nωmeson term is required [25] and thus the naturalness breaks d own. Moreover, the isoscalar-\n3isovector coupling in the RMF models increases the effective ρmeson mass with density, which\nleads the model to be far away from the chiral limit.\nThe Walecka model with the density-dependent parameters is the simplest version to incor-\nporate the effects of chiral symmetry. The Lagrangian is writt en as\nL=ψ[iγµ∂µ−M∗+g∗\nσσ−g∗\nωγµωµ−g∗\nργµτ3bµ\n0]ψ+1\n2(∂µσ∂µσ−m∗2\nσσ2)\n−1\n4FµνFµν+1\n2m∗2\nωωµωµ−1\n4BµνBµν+1\n2m∗2\nρb0µbµ\n0 (1)\nwhereψ,σ,ω, andb0are the fields of the nucleon, scalar, vector, and isovector- vector mesons,\nwith their in-medium scaled masses M∗,m∗\nσ,m∗\nω, andm∗\nρ, respectively. The meson coupling\nconstants and hadron masses with asterisks denote the densi ty dependence, given by the BR\nscaling. The energy density and pressure read, respectivel y,\nE=1\n2C2\nωρ2+1\n2C2\nρρ2δ2+1\n2˜C2\nσ(m∗\nN−M∗)2+/summationdisplay\ni=p,n2\n(2π)3/integraldisplaykFi\n0d3k E∗, (2)\np=1\n2C2\nωρ2+1\n2C2\nρρ2δ2−1\n2˜C2\nσ(m∗\nN−M∗)2−Σ0ρ+1\n3/summationdisplay\ni=p,n2\n(2π)3/integraldisplaykFi\n0d3kk2\nE∗(3)\nwhereCω=g∗\nω/m∗\nω,Cρ=g∗\nρ/m∗\nρ,˜Cσ=m∗\nσ/g∗\nσ,E∗=/radicalBig\nk2+m∗\nN2withm∗\nN=M∗−g∗\nσσthe\neffective mass of nucleon, and kFis the Fermi momentum. The incompressibility of symmetric\nmatter can be expressed explicitly as [27]\nκ= 9ρ∂µ\n∂ρ= 9ρ(C2\nω+2Cωρ∂Cω\n∂ρ+∂EF\n∂ρ−∂Σ0\n∂ρ) (4)\nwhere the chemical potential is given by µ=∂E/∂ρand the Fermi energy is EF=/radicalBig\nk2\nF+m∗\nN2.\nThe rearrangement term is essential for the thermodynamic c onsistency to derive the pressure\nin (3) and its expectation value in the mean field Σ 0is given by\nΣ0=−ρ2Cω∂Cω\n∂ρ−ρ2δ2Cρ∂Cρ\n∂ρ−˜Cσ∂˜Cσ\n∂ρ(m∗\nN−M∗)2−ρs∂M∗\n∂ρ. (5)\nThe density dependence of parameters is usually described b y the scaling functions that are the\nratios of thein-mediumparameters tothoseinthefreespace . Thechoice of scaling functionsand\ntheir coefficients are constrained by thesaturation propert iesof nuclear matter andexperimental\ndataaboutthein-mediummass droppingofvector mesons. Mor eover, wealsouseasaconstraint\nthepressurewithin the density range 2 −4.6ρ0extracted from measurements of nuclear collective\nflows in heavy-ion collisions [6]. The scaling function may t ake the form [28]:\nΦ(ρ) =1\n1+yρ/ρ0(6)\n4withy= 0.28 for the vector meson mass, giving Φ( ρ0) = 0.78 found in QCD sum rules [29].\nRecently, in Ref. [30] where the memory effect in dimuon yield w as studied by considering the\nmass dropping of ρmeson, the authors cited a scaling function [31]\nΦ(ρ) = 1−yρ/ρ0 (7)\nwithy= 0.15. Data extracted from the γ−Areaction by the TAP collaboration indicate a value\nofy≈0.13 [32]. In addition, data from the KEK photon-induced nucle ar reaction indicate that\ntheωmeson mass dropping is about the same order of magnitude at th e saturation density [33].\nSong firstly built effective models based on the BR scaling usin g the scaling functions (6)\nfor both hadron masses and vector coupling constants [28]. A reasonable incompressibility for\nnuclear matter was obtained by introducing the non-linear s elf-interacting meson terms with\ncoefficients satisfying the hypothesis of naturalness that is originated from the chiral symmetry\nand QCD scaling [34, 35, 36, 37]. In Ref. [38], along the line o f [28], the BR scaling function\n(6) was considered only for hadron masses with a small value o fyin the RMF models to study\nnuclear matter properties. In a more recent work [39], the BR scaling function (6) was taken\nfor the scalar and vector meson coupling constants with resp ective values of y, and the scaling\nfunction (7) was taken for hadron masses in the RMF models wit hout the self-interacting meson\ninteractions. Thoughthemodelsbuilt intheseworkscan giv e rather gooddescriptions of nuclear\nsaturation properties, the pressures calculated in the den sity region of ρ= 2−4.6ρ0, however,\nare still far away from that extracted from measurements of n uclear collective flows in heavy-ion\ncollisions [6]. People have already tried to improve the sit uation by including the non-linear\nmeson self-interacting terms. Unfortunately, it is not sat isfactory because even unreasonably\nlarge coefficients of non-linear meson terms that break down t he hypothesis of naturalness can\nnot reduce the pressure lower enough to pass the experimenta l pressure-density region.\nThe effective nucleon mass is not dominated by the BR scaling (a t least at the normal\nnuclear matter density) since it is usually around 0 .65M∗at normal density while the mass\ndropping given by the BR scaling is less than 15% ( y≤0.15 in (7)). This implies that the\nscalar meson coupling constant that plays a crucial role in t he effective nucleon mass can adopt\na different density-dependent scaling from that of the vector meson. In particular, the high\npressure predicted by various models at high densities shou ld be lowered, and this requires the\nscalar coupling constant to decrease more slowly. Furtherm ore, if the scaling function (7) for\ntheωmeson mass is preferred by experiments, one may take the same scaling function (7) for\n5Table 1: Parameter sets fitted at the saturation density ρ0= 0.16fm−3. The vacuum hadron\nmasses are M= 938MeV, mσ= 500MeV, mω= 783MeV and mρ= 770MeV except for mσ=\n600MeV for the parameter set S3. The coupling constants give n here are those at zero density.\nFor parameter sets SL3 and S3, the non-linear σself-interacting coefficients are introduced (see\ntext). The parameter set SL1∗has two more parameters yρ= 0.654 andyω= 0.0365. The\nsymmetry energy is fitted to 31.6MeV at ρ= 0.16fm−3for all models. The critical density ρc\nfor the chiral symmetry restoration is given by the value yin (7) for zero hadron mass.\ngσgωgρy x κ (MeV)M∗/M M∗\nn/M ρ c/ρ0\nSL1 8.6388 10.4634 3.7875 0.126 0.234 230.0 1.0 0.679 7.94\nSL1∗9.7414 12.5535 5.8644 0.126 0.238 230.0 1.0 0.600 7.94\nSL2 6.1664 10.9682 3.9866 0.11 0.381 219.5 0.89 0.763 9.09\nSL3 9.8627 12.4928 3.6128 0.126 - 250.0 1.0 0.620 7.94\nS3[28] 5.3210 15.3134 3.6035 0.28 - 250.0 0.78 0.617 -\nthe coupling constant of ωmeson to avoid the infinity of pressure at the chiral limit whe re the\nscalar density vanishes. In this way, the effect of density dep endence from the vector meson part\nis cancelled out since the energy density (3) only relies on t he ratiosCωandCρ. Generally, we\nmay take the scaling functions for the coupling constants of vector mesons as\nΦρ(ρ) =1−yρ/ρ0\n1+yρρ/ρ0,Φω(ρ) =1−yρ/ρ0\n1+yωρ/ρ0. (8)\nFor hadron masses (including nucleons, if they have), (7) is taken with the same value of yused\nin (8). For the σmeson coupling constant, the same form as (6) is taken but wit h a coefficient\ndenoted by x:\nΦσ(ρ) =1\n1+xρ/ρ0. (9)\n3 Results and discussions\nWe first adjust the parameters to reproduce the saturation pr operties including the binding\nenergy per nucleon E/A−M=−16 MeV, the zero pressure, the incompressibility κand the\neffective nucleon mass m∗\nNat saturation density ρ0= 0.16fm−3. The resulting parameter sets\nSL1 and SL2 and the corresponding saturation properties are tabulated in Table 1. In SL1 the\nnucleon mass scaling is not considered and the effective nucle on mass is just m∗\nN=M−g∗\nσσ. In\nSL2, the nucleon mass scaling is included, and a much larger e ffective nucleon mass at normal\ndensity is obtained. There are just two coefficients xandyused in the scaling functions in SL1\nand SL2. Without the inclusion of the non-linear meson self- interacting terms, the saturation\n6110102103\n2 4 6 8FsuGoldSL1\nSL1*SL2SL3 S3\n ρB/ρ0 p (MeV/fm3)\nFigure 1: The pressure as a function of density for different mo dels. The shaded region is given\nby experimental error bars[6].\npropertiesat ρ0= 0.16fm−3(seeTable1) andthepressurewithinthedensity region ρ= 2−4.6ρ0\nare both nicely reproduced with the SL1 and SL2 (see Fig. 1). I t is worth mentioning that the\nvector potential which is quadratic in density for the const antCωis known to result in a higher\npressure above the experimental region. The softening of th e pressure here is attributed to the\ncontribution of rearrangement terms. For the SL1, only the r earrangement term from the σ\nmeson survives.\nThe parameter set SL3 does not consider the scaling of scalar meson coupling constant but\nincludes in the Lagrangian the non-linear σself-interacting terms U(σ) =g2σ3/3+g3σ4/4 with\ncoefficients g2= 23.095 andg3=−29.678. Though these large coefficients are needed to fit the\nsaturation properties, they seem to be inconsistent with th e hypothesis of naturalness [35, 36].\nIn S3, we introduce a g2=−1.096 to obtain the given κin Table 1. This is a little different\n7from the original one [28]. As a comparison, we can see in Fig. 1 that the SL3 and S3 parameter\nsets are not consistent with the pressure constrained by the collective flow data.\nWith (3), the symmetry energy in the RMF models can be derived as\nEsym=1\n2∂2(E/ρ)\n∂δ2=1\n2C2\nρρ+k2\nF\n6EF. (10)\nIt consists of contribution from the ρmeson (potential part) and nucleons (kinetic part). In\nSL1 and SL2, we adopt the same scaling function for coupling c onstants of ρandωmesons:\nΦρ(ρ) = Φω(ρ) = 1−yρ/ρ0. In this way, the ratio Cρis just a constant that does not rely on\nthe density. An almost linear dependence of the symmetry ene rgy on the density is expected\nfrom Eq.(10). The symmetry energy as a function of density is shown in Fig.2. In Refs. [10, 11],\nthe symmetry energy extracted from the isospin diffusion data is parameterized as Esym=\n31.6(ρ/ρ0)γwith 0.69≤γ≤1.05 . All the parameter sets based on the BR scaling discussed\nabove lead to the symmetry energies between those parameter ized withγ= 0.69 and 1.05 in the\nwhole density region. In Ref. [11], the authors pointed out t hat with the in-medium nucleon-\nnucleon cross sections, a symmetry energy of Esym(ρ) = 31.6(ρ/ρ0)0.69forρ<1.2ρ0was found\nmostacceptable comparedtotheisospindiffusiondata. Howev er, thesymmetryenergy athigher\ndensities is not constrained at all. It is thus interesting t o examine predictions within the RMF\nmodels with the chiral limit at high densities.\nOur strategy is to first fit the symmetry energies constrained at low densities by modifying\nthe coefficient yρin (8), and then predict the symmetry energy at high densitie s. The modified\nsymmetry energy is shown in Fig.2 with the dashed blue curve d enoted as SL1∗. Considering\nthe nucleon effective mass at normal density is a little large i n SL1, we introduce a coefficient\nyωin (8) to reduce it to the value of 0 .6M. The new parameter set is called SL1∗and also listed\nin Table 1. The different parameterizations in SL1 and SL1∗lead to significantly difference in\npressure at high densities as shown in Fig.1. However, the su ppression of the symmetry energy\nin SL1∗at high densities is dominated by the density dependence of t he ratioCρin (10). This\ncan be seen by comparing results shown in Table 1 and Fig.2 tha t the density dependence of\nthe symmetry energy does not change much by the different nucle on effective masses of various\nparameter sets. In the density-dependent RMF model [24], th e dropping of ratios like the Cωor\nCρis induced by the many-body correlation contributions. Tho ugh the mechanism is different\nfrom the present study, the correlation effect beyond the mean -field approximation may play\nsome role in obtaining the coefficients yρandyω.\n80100200300\n0 2 4 6MDI(x=0)\nMDI(x=-1)\nSL2\nSL3\nSL1*\nSL1\n ρB/ρ0 Esym (MeV) 020\n0 1ρ/ρ0Esym\nFigure 2: The symmetry energy as a function of density for diffe rent models. The MDI(x=0)\nand MDI(x=-1) results are taken from [10].\nThe symmetry energy at high densities from the MDI interacti on with x=0 is quite close\nto that given by the SL1∗. This is not surprising since the SL1∗is rendered to have the same\nsymmetry energies as the MDI(x=0) at low densities. On the ot her hand, this justifies the\nconsistency of the symmetry energy given by both models at hi gh densities. One can freely\nadjustyρto fit the symmetry energies given by the MDI(x=-1) model at lo w densities and then\nexamine the behavior at high densities. However, the Cρwill increase with density by doing\nso, which is contrary to the empirical cases [24, 25]. Theref ore, based on the BR scaling the\nsymmetry energy at high densities is expected to be between t hose given by the SL1 and the\nSL1∗.\nWe now turn to the astrophysical implications of the EOS’s co nstrained above. Generally\nspeaking, these EOS’s have the special features of being sof t at moderate densities and stiff at\n9012\n10 12 14SL3SL2\nSL1 SL1*\nSL1′R1.4\n11.9≤R1.4≤13.1\n R (km)M / MO⋅\nFigure 3: The neutron star mass versus the radius for various models. The parameter set SL1′\nis just the same as the SL1 but with yρ= 0.526.\nhigh densities (see Fig. 1). It is thus interesting to compar e predictions using these EOS’s with\ntherecent astrophysical observations. Recent studies on t he millisecond pulsarPSR J0751+1807\nsuggest that it has a mass of 2 .1±0.2(+0.4\n−0.5) with 1σ(2σ) confidence[40]. Since the maximum\nmass of neutron star is more sensitive to the EOS at high densi ties, this requires a rather stiff\nEOS at least at high densities. In Fig.3, we plot the mass-rad ius correlation of neutron stars\nobtained from solving the standard TOV equation. In the mode ls with chiral limits, hadron\nmasses approach zero at certain critical baryon densities ( see Table 1). The maximum central\nenergy density in a neutron star is that calculated at the max imum baryon density. With\nSL1, the maximum neutron star mass is 2.04 M⊙with a radius of R=9.89km. With SL2, these\ntwo observables are 2.17 M⊙and R=10.13km, respectively. As the nucleon effective mass at\nnormal density is reduced by introducing the parameter yω, the maximum neutron star mass\n10also drops. This is owing to that the introduction of yωsoftens the EOS at high densities and\nthe energy density at the critical density is thus lowered. W ith SL1∗, the nucleon effective\nmass isM∗\nn= 0.6Matρ0, and the maximum neutron star mass is 1 .7M⊙. For a moderate\nvalue ofM∗\nn= 0.65M, we can obtain a moderately larger neutron star mass 1 .94M⊙. Recent\nmeasurements on the neutron star EXO 07482-676 gave a mass of M= 2.10±0.28M⊙and a\nradius of 13 .8±1.8km [41]. The author indicated that the existence of such a ma ssive neutron\nstar rules out all the soft EOS’s of neutron-star matter. Amo ng the models studied here,\nonly the parameter set SL3 can give values close to the measur ement for the EXO 07482-676.\nHowever, the parameter set SL3 gives a pressure at ρ≤4.6ρ0that is too strong compared to\nthat constrained by the collective flow data in heavy-ion col lisions [6]. For the other parameter\nsets, the radii of maximum mass neutron stars are just around 10km, which is below that given\nin [41], though the maximum mass can be within the error bars o f the measurement (except for\nSL1∗).\nCompared to the EOS obtained using the FSUGold parameter set [25] the EOS’s in the\npresent study are stiffer at high densities although they have the same incompressibility κ= 230\nMeV at normal density (see Fig.1). The EOS’s obtained here th us also result in larger maximum\nmasses of neutron stars. In Ref. [5], the authors predicted a radius span of 11.5km > and the surface potential may be written as: \n \n() ( ) ( ) () )r(n nn n rer,n e 0 0 0 total δµµΦ Φ∂∂++≈ (3) \n \nWhen we substract the average value of the surface potential measured at high \nelectron density (positive back -gate voltage, Eq. (3)) and equal hole density (negative \nback-gate voltage) from the surface potential recorder close to the Dirac point (Eq. \n(2)), we derive: \n)r(nn )n( ))r(n(2)r,n( )r,n( e)r,0( e0 0 total 0 totaltotal δµδµΦ ΦΦ∂∂− =+−− \n \n )r(nn )n( \nn ))r(n( 0δµ δµ⋅⎟\n⎠⎞⎜\n⎝⎛\n∂∂−∂∂≈ \n \n \n )r(nn ))r(n( δδµ⋅∂∂≈ (4) \n \n \nThe density variations in the graphe ne sheet can then be obtained from: \n ()()()\n()\nn )r(n 2r,n er,n er,0 e\n)r(n0 total 0 totaltotal\n∂∂+−−\n≈δµΦ ΦΦ\nδ (5) \n \n \nB. Direct pick-up: \n \nThe small size of the graphene flake and the finite distance between the single \nelectron transistor (SET) and th e graphene flake result in a directly measured potential \ncontribution from the highly doped Si-back gate. The inset to Fig. 5 illustrates that some electrical field lines originating from the back-gate reach th e SET. The parasitic \ncharge associated with these fringing fi elds is also detected by the SET. \n The detected amount of charge depends both on the distance of the SET to the \ngraphene edge as well as on the vertical di stance between tip and flake. Fig. 5 shows \nthe dependence of the direct pickup measur ed as a function of height above a gold \nelectrode. The gold electrode serves as a re ference since it entirely screens the back-\ngate potential underneath and only the fringi ng fields are detected . The distance from \nthe edge is chosen to be equal to the one in the real measurement on top of the \ngraphene flake. As expected, once the SET comes close to the surface the dir ect pick-up of the \nfringing fields drops and appr oaches zero. For typical he ights and distances to the \nsample edge we derive a contribution from the fringing fields of approximately 0.15 \nmeV 10\n-10 cm2. \n \n \n \n \nFig. 5: The direct pick-up cont ribution to the measured inverse compressibility as a function of \nthe distance between the tip and the sample. The ti p is placed above a gold electrode at a \nfixed distance from the border of the electrode typical for the configuration in the experiment \non a graphene flake. The inset schematically illu strates the origin of this direct pick-up. \n \n \n " }, { "title": "0705.3003v1.Negative_Energy_in_Superposition_and_Entangled_States.pdf", "content": "arXiv:0705.3003v1 [quant-ph] 21 May 2007Negative Energy in Superposition and Entangled States\nL.H. Ford∗\nInstitute of Cosmology\nDepartment of Physics and Astronomy\nTufts University, Medford, MA 02155\nThomas A. Roman†\nDepartment of Mathematical Sciences\nCentral Connecticut State University\nNew Britain, CT 06050\nAbstract\nWe examine the maximum negative energy density which can be a ttained in various quantum\nstatesofamasslessscalarfield. Weconsiderstatesinwhich eitheroneortwomodesareexcited, and\nshow that the energy density can be given in terms of a small nu mber of parameters. We calculate\nthese parameters for several examples of superposition sta tes for one mode, and entangled states\nfor two modes, and find the maximum magnitude of the negative e nergy density in these states.\nWe consider several states which have been, or potentially w ill be, generated in quantum optics\nexperiments.\nPACS numbers: 04.62.+v,03.65.Ud,42.50.Pq\n∗Email: ford@cosmos.phy.tufts.edu\n†Email: roman@ccsu.edu\n1I. INTRODUCTION\nIt has been proven, beginning in the early 1960’s [1], that there alway s exist states\nwith negative energy density in quantum field theory. Some specific e xamples include the\nCasimir effect [2] and squeezed states [3], both of which have been ex perimentally realized.\n(Although the energy density itself is far too small to be directly mea sured.) Negative\nenergy is also required for black hole evaporation, and hence for th e consistency of the\nlaws of black hole physics with those of thermodynamics. On the othe r hand, unrestricted\namounts of negative energy could produce bizarre effects, for ex ample, violations of the\nsecond law of thermodynamics [4, 5]. However, the same laws of quan tum field theory which\nallow the existence of negative energy also appear to severely rest rict its magnitude and\nduration in such a way as to prevent gross large-scale effects. The se bounds are known\nas quantum inequalities, and quite a large body of work now exists on t he subject. For\nsome recent reviews of quantum inequalities, see Refs. [6, 7, 8]. Qu antum inequality bounds\nhave been proven, for example, for the minimally coupled scalar, elec tromagnetic, and Dirac\nfields. It should be pointed out that the potential macroscopic pro blems arise not because\nof the existence of negative energy per se, but from the arbitrar y separation of negative and\npositive energy. It is this behavior which the quantum inequalities pro hibit. Many possible\nconfigurations of separated negative and positive energy can eas ily be ruled out, and known\npermitted examples involve the subtle intertwining of the two [9]. Whet her the currently\nknown examples are representative of the general case is unknow n. Hence, the study of\nfurther examples could prove useful.\nSince the negative energy densities in these states aretoo small to be directly measurable,\nexperiments in quantum optics may offer the best possibilities for indir ect detection of\nnegative energy. (However, see also Refs. [10, 11].) A first link betw een quantum optics and\nthe work on quantum inequalities has been forged in a recent paper b y Marecki [12]. For\nsqueezed states, he proved quantum inequality-type bounds on t he magnitude and duration\nof the squeezing.\nQuantum optics has seen enormous experimental and theoretical advances in the last\ntwenty years. This marriage of optics with quantum field theory has resulted in experiments\nwhich were formerly purely “gedanken” becoming those which are no w routinely performed\nin the laboratory. Highly non-classical states, such as Schr¨ oding er “cat states” and squeezed\n2states, have been produced and play a part in everything from qua ntum computers to\nnoise reduction in laser interferometer gravitational wave detect ors. The “cat states” of the\nelectromagnetic field are superpositions of coherent states and h ave been created experi-\nmentally [13, 14]. The experiments which have been done so far have p roduced mesoscopic\nsuperpositions, in which the mean photon number is of order 10. This is somewhat short of\na true Schr¨ odinger cat state, which would be a superposition of tw o or more classical config-\nurations, that is, coherent states with very large occupation num bers. More recently, there\nhave been proposals for methods of creating superpositions of sq ueezed vacuum states [15].\nAn interesting question arises: can one start with two quantum sta tes which do not\ninvolve negative energy and by superposing them obtainnegative en ergy? The answer is yes;\nthe classic standard example being the vacuum + two-particle state (for a nice discussion\nsee Ref. [16]). Is this true for the superposition of other states a s well? More generally,\nwhat effects does superposition have on negative energy? Could on e also go the other way,\ni.e., start with two states involving negative energy and by superpos ing them diminish or\neradicate the negative energy? In this paper, we will address such questions for several\nclasses of states. In Sect. II, we develop some formalism for para meterizing the maximum\nmagnitude of negative energy that can occur for states of a minima lly coupled scalar field\nin Minkowski spacetime with either one or two modes excited. We give s everal examples\nof superpositions for a single mode in Sect. III, including superposit ions of two coherent\nstates, two squeezed vacuum states, and a coherent state with a squeezed vacuum state. In\nSect. IV, we move to the two-mode case. This allows us to consider e xamples of entangled\nstates involving either squeezed vacua or coherent states for th e two modes. A summary of\nour conclusions is presented in Section V.\nII. ENERGY DENSITY WITH ONE OR TWO MODES ARE EXCITED\nIn this paper, we will consider a massless scalar field in flat spacetime, for which the\nstress tensor operator is\nTµν=ϕ,µϕ,ν−1\n2gµνϕ,σϕ,σ. (1)\nThe normal-ordered energy density operator is\n:T00:=1\n2[: ˙ϕ2: + : (∇ϕ)2:], (2)\n3where\nϕ=/summationdisplay\nk(akfk+ak†fk∗), (3)\nwith thefk(x,t) being the mode functions.\nA. Two Modes Excited\nWe wish to consider the case where all modes except for two are in th e vacuum state. For\nthe first mode, let f1,a, anda†be the mode function, annihilation operator, and creation\noperator, respectively, and let f2,b, andb†be the corresponding quantities for the second\nmode. The expectation value of the energy density in an arbitrary q uantum state can be\nexpressed as\nρ=/an}b∇acketle{t:T00:/an}b∇acket∇i}ht= Re/braceleftbigg\n/an}b∇acketle{ta†a/an}b∇acket∇i}ht(|˙f1|2+|∇f1|2)+/an}b∇acketle{ta2/an}b∇acket∇i}ht[˙f2\n1+(∇f1)2]+/an}b∇acketle{tb†b/an}b∇acket∇i}ht(|˙f2|2+|∇f2|2)\n+/an}b∇acketle{tb2/an}b∇acket∇i}ht[˙f2\n2+(∇f2)2]+2/an}b∇acketle{ta†b/an}b∇acket∇i}ht(˙f∗\n1˙f2+∇f1∗·∇f2)+2/an}b∇acketle{tab/an}b∇acket∇i}ht(˙f1˙f2+∇f1·∇f2)/bracerightbigg\n.(4)\nLet\nn1=/an}b∇acketle{ta†a/an}b∇acket∇i}ht, n2=/an}b∇acketle{tb†b/an}b∇acket∇i}ht, R1eiγ1=/an}b∇acketle{ta2/an}b∇acket∇i}ht, R2eiγ2=/an}b∇acketle{tb2/an}b∇acket∇i}ht, R3eiγ3=/an}b∇acketle{ta†b/an}b∇acket∇i}ht, R4eiγ4=/an}b∇acketle{tab/an}b∇acket∇i}ht.\n(5)\nAll of the information needed to give the two-mode energy density, Eq. (4), at a given\nquantum state is encoded in the above set of six amplitudes and four phases.\nIn the case of a traveling waves, we may take the mode functions to be\nfj=i/radicalBig\n2ωjVei(kj·x−ωjt), (6)\nwhereωj=|kj|, forj= 1,2 andVis a normalization volume. In this case, the mean energy\ndensity may be expressed as\nρ=1\nV/braceleftbigg\nn1ω1+n2ω2+R1ω1cos[2(k1·x−ω1t)+γ1]+R2ω2cos[2(k2·x−ω2t)+γ2]\n+R3√ω1ω2(1+ˆk1·ˆk2) cos[(k2−k1)·x−(ω2−ω1)t+γ3]\n+R4√ω1ω2(1+ˆk1·ˆk2) cos[(k2+k1)·x−(ω2+ω1)t+γ4]/bracerightbigg\n. (7)\nWe will also consider the case of a standing wave which depends upon o nly one space\ncoordinate, in which case the mode functions can be taken to be\nfj=1/radicalBig\nωjVsin(ωjx)e−iωjt. (8)\n4The energy density now becomes\nρ=1\nV/braceleftbigg\nn1ω1+n2ω2+R1ω1cos(2ω1x) cos(2ω1t−γ1)+R2ω2cos(2ω2x) cos(2ω2t−γ1)\n+ 2R3√ω1ω2cos[(ω2−ω1)x] cos[(ω2−ω1)t−γ3]\n+ 2R4√ω1ω2cos[(ω1+ω2)x] cos[(ω1+ω2)t−γ4]/bracerightbigg\n. (9)\nB. One Mode Excited\nA useful special case is when only one mode is excited. In this case, w e may setn1=n,\nR1=R,γ1=γ, andR2=R3=R4=γ2=γ3=γ4= 0. In this case, we need only the\nthree real numbers n,R, andγto determine the energy density in a given state. For the\ncase of a traveling wave, we have\nρ=ω\nV{n+Rcos([2(k·x−ωt)+γ]} (10)\nWe can see from Eq. (10) that the minimum value of ρis\nρmin=−ω\nV(R−n), (11)\nand hence we can have negative energy density only if R > n. In the case of a standing\nwave, Eq. (9) becomes\nρ=ω\nV[n+Rcos(2ωx) cos(2ωt−γ)]. (12)\nAgain, the minimum value of ρis given by Eq. (11).\nIII. SUPERPOSITIONS FOR ONE MODE\nIn this section, we examine some explicit examples of superpositions in volving a single\nmode. In each case, we need only calculate the quantity R−nto determine the maximum\nmagnitude of the negative energy.\nA. Superposition of Two Coherent States\nFirst we consider a superposition of coherent states. Coherent s tates are eigenstates of\nthe annihilation operator, that is\na|α/an}b∇acket∇i}ht=α|α/an}b∇acket∇i}ht. (13)\n5Let\nψ/an}b∇acket∇i}ht=N[|α/an}b∇acket∇i}ht+η|β/an}b∇acket∇i}ht], (14)\nwhere|α/an}b∇acket∇i}htand|β/an}b∇acket∇i}htare two different coherent states for the same mode, ηis a complex\nnumber, and Nis a normalization factor (see, for example, Sec. 7.6 of Ref. [17]). W e also\nassume that the states are normalized so that /an}b∇acketle{tα|α/an}b∇acket∇i}ht=/an}b∇acketle{tβ|β/an}b∇acket∇i}ht= 1. As a result we have that\n/an}b∇acketle{tψ|ψ/an}b∇acket∇i}ht= 1 =N2[1+|η|2+η/an}b∇acketle{tα|β/an}b∇acket∇i}ht+η∗/an}b∇acketle{tβ|α/an}b∇acket∇i}ht]. (15)\nThe coherent states are not orthonormal; their overlap integral is given by (see for example,\nEq.(3.6.24) of Ref. [18]):\n/an}b∇acketle{tα|β/an}b∇acket∇i}ht=e−1\n2(|α|2+|β|2−2α∗β). (16)\nTherefore the square of the normalization factor is\nN2= [1+|η|2+2e−1\n2(|α|2+|β|2)Re(ηeα∗β)]−1. (17)\nThe mean number of particles is found to be\nn=N2/bracketleftBig\n|α|2+|ηβ|2+2e−1\n2(|α|2+|β|2)Re(ηα∗βeα∗β)/bracketrightBig\n, (18)\nand\n/an}b∇acketle{ta2/an}b∇acket∇i}ht=N2/bracketleftBig\nα2+|η|2β2+e−1\n2(|α|2+|β|2)(ηβ2eα∗β+η∗α2eαβ∗)/bracketrightBig\n. (19)\nLet\nα=|α|eiδ1, β=|β|eiδ2,andη=|η|eiδ. (20)\nThen the quantities n,R, andγare functions of six real parameters, the magnitudes and\nphases ofα,β, andη. However, one finds that only γdepends upon all six. The magnitudes\nnandRdepend only upon the difference δ2−δ1, and are hence functions of five parameters.\nWe are primarily interested in the quantity R−n, which measures the maximum magnitude\nof the negative energy density. Hence set δ1= 0 and write\nF(|α|,|β|,|η|,δ2,δ) =R−n, (21)\nand letGbe a five-dimensional vector given by\nG=/parenleftBigg∂F\n∂|α|,∂F\n∂|β|,∂F\n∂|η|,∂F\n∂δ2,∂F\n∂δ/parenrightBigg\n. (22)\n6One may use Eqs. (21) and (22) as the basis of a numerical algorithm to search for points\nof maximum Fand hence maximally negative energy density. Start at a random poin t in\nthe five-dimensional parameter space, and compute FandG. IfF >0, then this choice of\nparameters is a quantum state with negative energy density. The c omponents of Gindicate\nthe direction in which Fis increasing most rapidly. One then moves along this direction\nuntil a local maximum of Fis located. A preliminary, non-exhaustive, search located two\nsuch local maxima, at ( |α|,|β|,|η|,δ2,δ)≈(0.8,0.8,1,3.14,0) and at ( |α|,|β|,|η|,δ2,δ)≈\n(0,1.61,1,0,0). (One can trivially generate a third maximum by interchange of αandβ\nin the latter case.) The first example corresponds to α=−β= 0.8 and the second to a\nsuperposition of a coherent state and the vacuum. Interestingly , the maximum magnitude\nof the negative energy density is about the same in both examples, w ithF=R−n≈0.278,\nand henceρmin≈ −0.278ω/V. We do not have an explanation as to why these two choices\ngive the same value of R−n. The mean particle number is n≈0.36 in the first example and\nn≈1.0 in the second. This example illustrates that a superposition of two c oherent states\ncan produce negative energy density, and the maximum negative en ergy density arises for\nmean particle number of order one.\nB. Superposed Squeezed Vacuum States\n1. A Single-Mode Squeezed Vacuum State\nWe begin with a review of the features of the expectation value of th e energy density in\na single squeezed vacuum state. Our state is given by:\n|ψ/an}b∇acket∇i}ht=|ξ/an}b∇acket∇i}ht,withξ=reiδ, (23)\nwhereris the squeeze parameter and δis a phase parameter. The squeeze operator S(ξ) is\ngiven by\nS(ξ) =e1\n2[ξa2−ξ∗(a†)2]. (24)\nThis operator is unitary since\nS†(ξ) =S(−ξ) =S−1(ξ). (25)\n7Thesingle-modesqueezed state |ξ/an}b∇acket∇i}htisproducedbythesqueeze operatoractingonthevacuum\nstate\n|ξ/an}b∇acket∇i}ht=S(ξ)|0/an}b∇acket∇i}ht. (26)\nThe state |ξ/an}b∇acket∇i}htcan be written, after some work (see Eq. (3.7.5) of Ref. [18]), in te rms of the\neven Fock states as\n|ξ/an}b∇acket∇i}ht=√\nsechr∞/summationdisplay\nn=0/radicalBig\n(2n)!\nn!/bracketleftbigg\n−1\n2eiδtanhr/bracketrightbiggn\n|2n/an}b∇acket∇i}ht. (27)\nWe also have that\nS†(ξ)aS(ξ) =acoshr−a†eiδsinhr,\nS†(ξ)a†S(ξ) =a†coshr−ae−iδsinhr. (28)\nIn the state |ξ/an}b∇acket∇i}htwe have the expectation values\nn=/an}b∇acketle{ta†a/an}b∇acket∇i}ht=/an}b∇acketle{t0|S†(ξ)a†aS(ξ)|0/an}b∇acket∇i}ht=/an}b∇acketle{t0|S†(ξ)a†S(ξ)S†(ξ)aS(ξ)|0/an}b∇acket∇i}ht= sinh2r,\n/an}b∇acketle{ta2/an}b∇acket∇i}ht=/an}b∇acketle{t0|S†(ξ)a2S(ξ)|0/an}b∇acket∇i}ht=/an}b∇acketle{t0|S†(ξ)aS(ξ)S†(ξ)aS(ξ)|0/an}b∇acket∇i}ht=−eiδsinhrcoshr,(29)\nwhere we have made use of Eqs. (28). Thus R= sinhrcoshrand\nR−n= sinhr(coshr−sinhr) (30)\nattains its maximum value of 0 .5 asr→ ∞.\n2. Superposition of Squeezed Vacuum States\nWe now calculate the energy density in a superposition of two single-m ode squeezed\nvacuum states of the form\n|ψ/an}b∇acket∇i}ht=N[|ξ/an}b∇acket∇i}ht+η|−ξ/an}b∇acket∇i}ht], (31)\nwhere, for simplicity, we will choose\nξ=r, δ= 0, (32)\nand set\nη=|η|eiθ. (33)\n8In this state we have\nn=/an}b∇acketle{ta†a/an}b∇acket∇i}ht=N2[/an}b∇acketle{tξ|a†a|ξ/an}b∇acket∇i}ht+|η|2/an}b∇acketle{t−ξ|a†a|−ξ/an}b∇acket∇i}ht+η/an}b∇acketle{tξ|a†a|−ξ/an}b∇acket∇i}ht+η∗/an}b∇acketle{t−ξ|a†a|ξ/an}b∇acket∇i}ht]\n=N2/bracketleftbigg\nsinh2r(1+|η|2)−2|η|cosθsechrtanh2r\n(1+tanh2r)3/2/bracketrightbigg\n, (34)\nwhere we have made use of Eq. (A4) in the Appendix. A similar calculatio n, using Eq. (A5),\nyields\n/an}b∇acketle{ta2/an}b∇acket∇i}ht=N2/bracketleftbigg\n(|η|2−1)sinhrcoshr+2i|η|sinθsechrtanhr\n(1+tanh2r)3/2/bracketrightbigg\n, (35)\nand\nR=|/an}b∇acketle{ta2/an}b∇acket∇i}ht|=N2/bracketleftbigg\n(|η|2−1)2sinh2rcosh2r+4|η|2sin2θsech2rtanh2r\n(1+tanh2r)3/bracketrightbigg1\n2.(36)\nThe normalization of our state is given by\n/an}b∇acketle{tψ|ψ/an}b∇acket∇i}ht= 1 =N2[1+|η|2+η/an}b∇acketle{tξ|−ξ/an}b∇acket∇i}ht+η∗/an}b∇acketle{t−ξ|ξ/an}b∇acket∇i}ht]. (37)\nThe square of the normalization factor is then\nN2=/bracketleftbigg\n1+|η|2+2|η|cosθ/radicalBig\nsech(2r)/bracketrightbigg−1\n, (38)\nwhere we have used Eq. (A8) of the Appendix.\nFrom Eqs. (34) and (36), we can compute the quantity R−n, which gives the maximum\nmagnitude of the negative energy density, as function of θandr. For fixed θ, one typically\nfinds thatR−nattains a maximum value for some value of r, usually of order one. A\ntypical case of θ= 0 is illustrated in Fig. 1. The case η= 0 is just the single squeezed\nvacuum state discussed in Sect. IIIB1 . This case gives the maximum negative energy\ndensity,R−n= 0.5, for large r. All other values of η, corresponding to superposed\nsqueezed vacua, give somewhat smaller amounts of negative energ y density, and attain their\nmaximum negative energy density at finite values of r.\nC. Superposition of Coherent and Squeezed Vacuum States\nIn this subsection, we consider states of the form\n|ψ/an}b∇acket∇i}ht=N[|ξ/an}b∇acket∇i}ht+η|α/an}b∇acket∇i}ht], (39)\n90.5 11.5 2\n-0.4-0.20.20.4\nrR - n θ = 0\nη = 0\nη = 0.25\nη = 0.5\nFIG. 1: The quantity R−nfor two superposed squeezed vacua, Eq. (31), is plotted for t he case\nθ= 0 for various values of η. The case η= 0 is the single squeezed vacuum, and gives more\nnegative energy density than do any of the superpositions. F or non-zero η, there is a maximum\nvalue for R−nat a finite value of r.\nwhere|ξ/an}b∇acket∇i}htis a squeezed vacuum state, and |α/an}b∇acket∇i}htis a coherent state. We may use Eq. (A10) to\nfind\nN2=/braceleftbigg\n1+|η|2+2√\nsechre−1\n2|α|2Re/bracketleftbigg\nηexp/parenleftbigg\n−1\n2e−iδα2tanhr/parenrightbigg/bracketrightbigg/bracerightbigg−1\n.(40)\nSimilarly, we find\nn=/an}b∇acketle{ta†a/an}b∇acket∇i}ht=N2/braceleftbigg\nsinh2r+|ηα|2\n−2e−1\n2|α|2√\nsechrtanhrRe/bracketleftbigg\nηe−iδα2exp/parenleftbigg\n−1\n2e−iδα2tanhr/parenrightbigg/bracketrightbigg/bracerightbigg\n(41)\nand\n/an}b∇acketle{ta2/an}b∇acket∇i}ht=N2/braceleftbigg\n−sinhrcoshreiδ+|η|2α2+ηα2√\nsechre−1\n2|α|2exp/parenleftbigg\n−1\n2e−iδα2tanhr/parenrightbigg\n+η∗√\nsechre−1\n2|α|2[(α∗)2eiδtanhr−1]eiδtanhrexp[−1\n2eiδ(α∗)2tanhr]/bracerightbigg\n,(42)\nusing Eqs. (A11) and (A13).\nLet us consider the case where δ= 0,η= 1, andαis real. In this case, R−nis plotted\nin Fig. 2 for various values of α. Here the maximum negative energy density, R−n≈0.23\n100.20.40.60.8 1\n-0.4-0.20.2\nα = 1α = 0\nα = 0.8 α= 0.6α = 0.4R - n\nr\nFIG. 2: Here R−nis plotted for the superposition of a coherent and a squeezed vacuum state,\nEq. (39) with η= 1, as a function of rfor various values of α. Note that for α/ne}ationslash= 0,R−ncan\nbe positive, corresponding to negative energy, for both sma ller and larger values of r, but has an\nintermediate region where there is no negative energy.\nis attained for large r. Note that this state has less negative energy than does the sque ezed\nvacuum by itself. [See Eq. (30).] For αnon-zero, we find that R−ninitially decreases,\nreaches a minimum value, and then increases again. For the case α= 0.6, for example, there\nis negative energy for r <0.2 and again for r >0.65, but not for intermediate values of r.\nNote thatr= 0 is a superposition of the vacuum and a coherent state, a special case of the\nstate treated in Sect. IIIA.\nA limit of special interest is when α= 0 and we have the superposition of the vacuum\nwith a squeezed vacuum state. In this case,\nN2=/bracketleftBig\n1+|η|2+2Re(η)√\nsechr/bracketrightBig−1, (43)\nn=/an}b∇acketle{ta†a/an}b∇acket∇i}ht=N2sinh2r, (44)\nand\n/an}b∇acketle{ta2/an}b∇acket∇i}ht=−N2sinhrcoshreiδ/bracketleftBig\n1+η∗(sechr)5\n2/bracketrightBig\n. (45)\nTheα= 0 curve in Fig. 2 is this limit for η= 1. Ifηis real and negative, η=−|η|, we then\n111 2 3 4\n-2.5-2-1.5-1-0.5R - n\nr\nFIG. 3: Here R−nis plotted as a functionof rforα= 0andη=−1, asuperpositionof thevacuum\nand a squeezed vacuum. The maximum negative energy occurs wh enr≈2, where R−n≈0.3.\nhave\nR=|/an}b∇acketle{ta2/an}b∇acket∇i}ht|=N2sinhrcoshr/vextendsingle/vextendsingle/vextendsingle1−|η|(sechr)5\n2/vextendsingle/vextendsingle/vextendsingle, (46)\nand\nR−n=sinhr/bracketleftBig\ncoshr/vextendsingle/vextendsingle/vextendsingle1−|η|(sechr)5\n2/vextendsingle/vextendsingle/vextendsingle−sinhr/bracketrightBig\n1+|η|2−2|η|√\nsechr. (47)\nIn the case that η=−1 the right-hand side of Eq. (47) is plotted as a function of rin Fig. 3.\nHere we find the maximum negative energy, R−n≈0.3 atr≈2. This is slightly less\nnegative energy than can be found in a single squeezed vacuum stat e. Note that as r→0,\nthis state becomes |2/an}b∇acket∇i}ht, a two-particle state with positive energy density everywhere. Th is is\nthe reason that the behavior in Fig. 3 differs from the α= 0 curve in Fig. 2. In the latter\ncase,η= 1, and there is negative energy for all values of r.\nIV. TWO-MODE ENTANGLED STATES\nIn this section, we will consider several examples of entangled stat es involving two modes.\n12A. An Entangled Squeezed State - the Barnett-Radmore State\nOur first example of a two-mode entangled squeezed state was des cribed by Barnett and\nRadmore [18] and is defined by\n|ψ/an}b∇acket∇i}ht=SAB|0/an}b∇acket∇i}ht, (48)\nwhere|0/an}b∇acket∇i}htis the vacuum state for both modes, and\nSAB= e(ξ∗ab−ξa†b†)(49)\nis a two-mode squeeze operator. If one were to expand the state |ψ/an}b∇acket∇i}htin terms of number\neigenstates, the expansion would contain states with an even tota l number of particles, with\nhalf of these particles in each mode. One has the following identities [18 ]\nSAB(−ξ)aSAB(ξ) =acoshr−b†eiδsinhr\nSAB(−ξ)a†SAB(ξ) =a†coshr−be−iδsinhr\nSAB(−ξ)bSAB(ξ) =bcoshr−a†eiδsinhr\nSAB(−ξ)b†SAB(ξ) =b†coshr−ae−iδsinhr. (50)\nNote that\nS†\nAB(ξ) =S−1\nAB(ξ) =SAB(−ξ). (51)\nOne may use these relations to show that\nn1=n2= sinh2r, R 1=R2=R3= 0, R4= sinhrcoshr,andγ4=δ+π.(52)\nThe minimum energy density in this state is\nρmin(BR) =−sinhr\nV[2√ω1ω2coshr−(ω1+ω2)sinhr]. (53)\nThis is never more negative than the minimum energy density that wou ld be obtained if\nthe two modes were individually in squeezed vacuum states. The latte r energy density is\nρmin(2SQ) =−(ω1+ω2)(R−n)/V, whereR−nis given by Eq. (30). Thus we can write\nρmin(BR)−ρmin(2SQ) =sinhrcoshr\nV(√ω1−√ω2)2, (54)\nwhich is always non-negative and approached zero only when the two modes have nearly the\nsame frequency.\n13B. An Second Entangled Squeezed State - the Zhang State\nIn this subsection, we will consider a second possibility for an entang led two-mode\nsqueezed state, which was discussed by Zhang [15]. This state is defi ned by\n|ψ/an}b∇acket∇i}ht=N/parenleftBig\n|¯ξ/an}b∇acket∇i}hta|¯η/an}b∇acket∇i}htb+eiθ|ξ/an}b∇acket∇i}hta|η/an}b∇acket∇i}htb/parenrightBig\n, (55)\nwhere|¯ξ/an}b∇acket∇i}htaand|ξ/an}b∇acket∇i}htaare single-mode squeezed vacuum states for mode a, and|¯η/an}b∇acket∇i}htband|η/an}b∇acket∇i}htbare\nsuch states for mode b. In general, ξ,¯ξ,η, and ¯ηcan be four arbitrary complex parameters.\nHowever, we will restrict our attention to the case where they are real and satisfy\nξ=η=−¯ξ=−¯η. (56)\nIn this case,\nN=/braceleftBig\n2[1+Re(eiθ/an}b∇acketle{t−ξ|ξ/an}b∇acket∇i}hta/an}b∇acketle{t−η|η/an}b∇acket∇i}htb)]/bracerightBig−1\n2= [2(1+cos θsech2r)]−1\n2, (57)\nwhere we have used Eq. (A8) for each of the two modes. Similarly, we find\nn1=n2= 2N2sinh2r/bracketleftBigg\n1−cosθ\n(cosh2r)3\n2/bracketrightBigg\n, (58)\nand\nR1=R2=N2|sinθ|tanh2r√\ncosh2r. (59)\nHere we have use Eqs. (A4) and (A5), as well as the identity sinh2r+cosh2r= cosh(2r).\nIn addition, we find R3=R4= 0 andγ1=γ2=−π/2.\nIn this case, the energy density, Eq. (7), becomes\nρ=1\nV/parenleftbigg\nn1(ω1+ω2)+R1{ω1cos[2(k1·x−ω1t)+γ1]+ω2cos[2(k2·x−ω2t)+γ1]}/parenrightbigg\n(60)\nWe can always choose the spatial position xand timetso as to make both cosine functions\nequal to −1, in which case we achieve the minimum allowed energy density in this sta te of\nρmin=−ω1+ω2\nV(R1−n1). (61)\nFrom Eqs. (57), (58) and (59), we find R1−n1as a function of θandr. In general,\nthe behavior of this entangled state is similar to that of the superpo sed squeezed vacua\nillustrated in Fig. 1. However, there is one limit of particular interest, which is when r≪1\n140.01\r 0.02\r 0.03\r 0.04\r\n-1\r-0.75\r-0.5\r-0.25\r0.25\r0.5\r0.75\r1\r θ = 0.99 π\nθ = 0.95 π\nθ = 0.99 πr\nFIG. 4: The quantities R1−n1(solid lines) and n1(dashed lines) are plotted for two values of θ\nas functions of rfor the entangled squeezed state defined in Eqs. (55) and (56) . In the limit that θ\nis close to π, one can have appreciable negative energy at small values of r. We see that the peak\nnegative energy, R1−n1≈0.25 occurs at r≈0.007 forθ= 0.99πand atr≈0.035 forθ= 0.95π,\nwhereas the mean particle number is about the same for both ca ses,n1≈0.2.\nand 0<|π−θ| ≪1. (Note that if θ=π, thenR1= 0, and there is no negative energy.) If\nwe take the limit r≪1, for fixed θ/ne}ationslash=π, then we find the asymptotic forms\nn1∼1−cosθ\n1+cosθr2, (62)\nand\nR1−n1∼|sinθ|\n1+cosθr. (63)\nIn the case that 0 <|π−θ| ≪1, the coefficient in the expression for n1can be large, so\nwe can have an unusually large particle number in relation to the value o fr. The quantities\nR1−n1andn1are plotted in Fig. 4 for two values of θclose toπ. In this case, we can have\na reasonable amount of negative energy at very small values of the squeeze parameter, r.\n15C. Entangled Coherent States\nIn this subsection, we consider a state of the same form as that in E q. (55), but involving\nentangled coherent states for two modes, which was also discusse d by Zhang [15]. Let\n|ψ/an}b∇acket∇i}ht=N/parenleftBig\n|α/an}b∇acket∇i}hta|β/an}b∇acket∇i}htb+eiθ|α′/an}b∇acket∇i}hta|β′/an}b∇acket∇i}htb/parenrightBig\n, (64)\nwhere|α/an}b∇acket∇i}hta,etcare single-mode coherent states. We will restrict our attention to the case\nwhere the magnitudes of the four complex coherent state parame ters are all equal, and\nα′=−αandβ′=−β. Thus\n|α|=|β|=|α′|=|β′|=σ, (65)\nand\nδ1−δ′\n1=±π, δ 2−δ′\n2=±π. (66)\nHereδ1,δ′\n1,δ2,δ′\n2are the phases of α,α′,β,β′, respectively. In this case, we find\nN=/bracketleftBig\n2(1+cosθe−4σ2)/bracketrightBig−1\n2, (67)\nand\nn1=n2= 2σ2N2(1−cosθe−2σ2),\nR1=R2= 2σ2N2(1+cosθe−2σ2),\nR3=σ2N2(1−cosθe−4σ2),\nR4=σ2, (68)\nas well asγ1= 2δ′\n1,γ2= 2δ′\n2,γ3=δ2−δ1, andγ4=δ1+δ2. Letφ1=k1·x−ω1tand\nφ2=k2·x−ω2t. We then set\nφ1+δ1=φ2+δ2=π\n2, (69)\nwhich can always be done by a suitable choice of xandt. The energy density for a two-mode\ntraveling wave state, Eq. (7) now becomes\nρ=1\nV/bracketleftBig\nn1ω1+n2ω2−R1ω1−R2ω2+(R3−R4)√ω1ω2(1+ˆk1·ˆk2)/bracketrightBig\n.(70)\nIf the two modes are close in wavenumber, so that ω1≈ω2=ωandˆk1≈ˆk2, and we set\nθ= 0 then\nρ=−4ω\nVf(σ), (71)\n160.5 11.5 20.050.10.150.2f (σ) \nσ\nFIG. 5: The function f(σ), given by Eq. (72), is plotted. Its maximum, at σ≈0.7, describes the\ncase of maximal negative energy density for the entangled co herent state defined in Eqs. (64), (65)\nand (66).\nwhere\nf(σ) =σ2e−2σ2(1+e−2σ2)\n1+e−4σ2. (72)\nThe function f(σ) is plotted in Fig. 5, where we see that it attains a maximum value of\nabout 0.22 atσ≈0.7. This corresponds to a negative energy density of ρ≈ −0.88ω/V,\nwhich is about three times as negative as the maximally negative energ y density found in\nthe superposed coherent states discussed in Sect. IIIA.\nV. SUMMARY\nIn this paper we have developed a formalism for parameterizing the e nergy density in\nstates of a massless scalar field in which either one or two modes are e xcited. We found\nexplicit expressions for the energy density for the cases of trave ling waves and of standing\nwaves in one spatial direction. In all cases, the maximum negative en ergy density which can\nbe achieved in a given state can be expressed in terms of our parame ters.\nWe next applied this approach to find the maximum negative energy de nsity in several\nstates, including some states which are of current interest in quan tum optics. For the case\n17of a single mode, we considered three possible superposition states : (1) two coherent states,\n(2) two squeezed vacuum states, and (3) a coherent state and a squeezed vacuum state.\nThe superposition of two coherent states can be described as a Sc hr¨ odinger “cat state” in\nthe sense that it would be a superposition of two classical configura tions in the limit of\nlarge coherent state parameter. Here we find that the maximal ne gative energy density\nis achieved with mean photon numbers slightly less than one. This is an e xample where a\nquantumsuperpositionstatehasnegativeenergydensity, event hougheachcomponentofthe\nsuperposition would have positive energy density by itself. Intheca se of thesuperposition of\ntwo squeezed vacuum states, one finds the opposite effect. Altho ugh here the superposition\nstate does has negative energy density, it is somewhat less negativ e than in the case of a\nsingle squeezed vacuum state. Furthermore, the most negative e nergy density now occurs\nfor small mean photon number, as opposed to large number in the ca se of a single squeezed\nvacuum state. In the case of a superposition of a coherent state and a squeezed vacuum\nstate, we find that for fixed coherent state parameter, there is negative energy density for\nsmall squeeze parameter, and again for larger values, but there is an intermediate range\nwhere the energy density is always positive.\nWe next examined some two-mode states involving entanglement bet ween the two modes,\nincludingtwoexamplesofentangledsqueezedvacuumstates. Thefi rstexample, theBarnett-\nRadmore state [18], exhibits somewhat less negative energy density than would be found if\neach mode were separately in a squeezed vacuum state. In the sec ond example, the Zhang\nstate [15], we find results similar to those in the superposition of sque ezed vacuum states.\nThere is negative energy in the Zhang state, but only for small mean particle numbers.\nFinally, we examined a two-mode entangled coherent state, which als o exhibits negative\nenergy for small mean particle number. It is also similar to the case of a superposition\nof coherent states of a single mode, but the entangled state has s omewhat more negative\nenergy density.\nOneofthemotivationsforthisinvestigation istodrawlinks between t heoretical studies of\nviolations of the weak energy condition, and experimental work in qu antum optics. We hope\nthat this line of work will lead to further experimental studies of sub vacuum phenomena.\n18Acknowledgments\nWe would like to thank Piotr Marecki for useful discussions. This wor k was supported in\npart by the National Science Foundation under Grant PHY-055575 4 to LHF.\nAPPENDIX A\nIn this appendix, we will calculate some of the matrix elements of oper ators such as a†a\nanda2which are needed to find the energy density in the states treated in this paper. We\nbegin with matrix elements between squeezed vacuum states. The d iagonal matrix elements\n/an}b∇acketle{tξ|a†a|ξ/an}b∇acket∇i}htand/an}b∇acketle{tξ|a2|ξ/an}b∇acket∇i}htare given by Eq. (29). We need off-diagonal matrix elements of the\nform/an}b∇acketle{t−ξ|a†a|ξ/an}b∇acket∇i}htand/an}b∇acketle{t−ξ|a2|ξ/an}b∇acket∇i}ht, whereξis real. If we set δ= 0, so that ξ=r, then Eq. (27)\nbecomes\n|ξ/an}b∇acket∇i}ht=√\nsechr∞/summationdisplay\nn=0/radicalBig\n(2n)!\nn!/parenleftbigg\n−1\n2tanhr/parenrightbiggn\n|2n/an}b∇acket∇i}ht. (A1)\nThis leads to the result\n/an}b∇acketle{t−ξ|a†a|ξ/an}b∇acket∇i}ht=/an}b∇acketle{tξ|a†a|−ξ/an}b∇acket∇i}ht= 2sechr∞/summationdisplay\nn=1(2n)!\nn!(n−1)!(−1)n/parenleftbigg1\n2tanhr/parenrightbigg2n\n.(A2)\nUse the fact that\n∞/summationdisplay\nn=1(2n)!\nn!(n−1)!(−1)n/parenleftbigg1\n2x/parenrightbigg(2n−2)\n=−2\n(1+x2)3/2, (A3)\nto find\n/an}b∇acketle{t−ξ|a†a|ξ/an}b∇acket∇i}ht=/an}b∇acketle{tξ|a†a|−ξ/an}b∇acket∇i}ht=−sechrtanh2r\n(1+tanh2r)3/2. (A4)\nSimilarly, we may use Eq. (A1) to show that\n/an}b∇acketle{t−ξ|a2|ξ/an}b∇acket∇i}ht=−/an}b∇acketle{tξ|a2|−ξ/an}b∇acket∇i}ht= sechr∞/summationdisplay\nn=1(2n)!\nn!(n−1)!(−1)n/parenleftbigg1\n2tanhr/parenrightbigg(2n−1)\n=−sechrtanhr\n(1+tanhr)3/2.\n(A5)\nBecause these matrix elements are real, we have that\n/an}b∇acketle{t−ξ|(a†)2|ξ/an}b∇acket∇i}ht=/an}b∇acketle{tξ|(a†)2|−ξ/an}b∇acket∇i}ht=/an}b∇acketle{t−ξ|a2|ξ/an}b∇acket∇i}ht. (A6)\nIn the present case, the squeeze operator is\nS(ξ) =S(r) = e1\n2r[a2−(a†)2]. (A7)\n19From this relation, we see that\n/an}b∇acketle{t−ξ|ξ/an}b∇acket∇i}ht=/an}b∇acketle{tξ|−ξ/an}b∇acket∇i}ht=/an}b∇acketle{t0|S2(r)|0/an}b∇acket∇i}ht=/an}b∇acketle{t0|S(2r)|0/an}b∇acket∇i}ht=/radicalBig\nsech(2r). (A8)\nNext we derive the matrix elements involving both a coherent state a nd a squeezed\nvacuum state that are needed in Sect. IIIC. A coherent state ma y be represented in terms\nof number eigenstates as [18]\n|α/an}b∇acket∇i}ht= e−1\n2|α|2∞/summationdisplay\nℓ=0αℓ\n√\nℓ!|ℓ/an}b∇acket∇i}ht. (A9)\nThis may be combined with Eq. (27) to show that\n/an}b∇acketle{tξ|α/an}b∇acket∇i}ht=/an}b∇acketle{tα|ξ/an}b∇acket∇i}ht∗= e−1\n2|α|2√\nsechr∞/summationdisplay\nn=0α2n\nn![−1\n2e−iδtanhr]n\n= e−1\n2|α|2√\nsechrexp/parenleftbigg\n−1\n2e−iδα2tanhr/parenrightbigg\n, (A10)\nand\n/an}b∇acketle{tξ|a†a|α/an}b∇acket∇i}ht=/an}b∇acketle{tα|a†a|ξ/an}b∇acket∇i}ht∗=−e−1\n2|α|2√\nsechre−iδα2tanhrexp/parenleftbigg\n−1\n2e−iδα2tanhr/parenrightbigg\n.(A11)\nNote that\n/an}b∇acketle{tξ|a2|α/an}b∇acket∇i}ht=/an}b∇acketle{tα|(a†)2|ξ/an}b∇acket∇i}ht∗=α2/an}b∇acketle{tξ|α/an}b∇acket∇i}ht. (A12)\nFinally, we show that\n/an}b∇acketle{tξ|(a†)2|α/an}b∇acket∇i}ht=/an}b∇acketle{tα|a2|ξ/an}b∇acket∇i}ht∗= e−1\n2|α|2√\nsechr∞/summationdisplay\nn=12n(2n−1)\nn!α2n−2/parenleftbigg\n−1\n2e−iδtanhr/parenrightbiggn\n= e−1\n2|α|2√\nsechrd2\ndα2∞/summationdisplay\nn=0α2n\nn!/parenleftbigg\n−1\n2e−iδtanhr/parenrightbiggn\n= e−1\n2|α|2√\nsechrd2\ndα2exp/parenleftbigg\n−1\n2e−iδα2tanhr/parenrightbigg\n= e−1\n2|α|2√\nsechr/parenleftBig\ne−iδα2tanhr−1/parenrightBig\ne−iδtanhr\n×exp/parenleftbigg\n−1\n2e−iδα2tanhr/parenrightbigg\n. (A13)\n[1] H. Epstein, V. Glaser, and A. Jaffe, Nuovo Cim. 36, 1016 (1965).\n[2] H.B.G. Casimir, Proc. K. Ned. Akad. Wet. B51, 793 (1948); L.S. Brown and G.J. Maclay,\nPhys. Rev. 184, 1272 (1969).\n20[3] R. E. Slusher, L. W. Hollberg, B. Yurke, J. C. Mertz, and J. F. Valley, Phys. Rev. Lett. 55,\n2409, (1985).\n[4] L. H. Ford, Proc. Roy. Soc. Lond. A364, 227 (1978).\n[5] L. H. Ford, Phys. Rev. D43, 3972 (1991).\n[6] L.H. Ford, ”Spacetime in Semiclassical Gravity”, in 100 Years of Relativity - Space-time\nStructure: Einstein and Beyond , edited by A. Ashtekar, (World Scientific, Singapore, 2006) ,\ngr-qc/0504096.\n[7] T.A. Roman, “Some Thoughts on Energy Conditions and Worm holes”, in Proceedings of the\nTenth Marcel Grossmann Meeting on General Relativity , edited by S.P. Bergliaffa and M.\nNovello, (World Scientific, Singapore, 2006), gr-qc/04090 90.\n[8] C.J. Fewster, “Energy inequalities in quantum field theo ry”, inXIVth International Congress\non Mathematical Physics , edited by J.C. Zambrini (World Scientific, Singapore, 2005 ), see\nupdated version, math-ph/0501073.\n[9] A. Borde, L.H. Ford, and T.A. Roman, Phys. Rev. D65, 084002 (2002), gr-qc/0109061.\n[10] L.H. Ford, P.G. Grove, and A.C. Ottewill, Phys. Rev. D 46, 4566 (1992).\n[11] P.C.W. Davies, A.C. Ottewill, Phys. Rev. D65, 104014 (2002), gr-qc/0203003.\n[12] P. Marecki, Phys. Rev. A 66, 053801 (2002), quant-ph/0203027.\n[13] A. Auffeves, P. Maioli, T. Meunier, S. Gleyzes, G. Nogues, M. Brune, J. M. Raimond, and S.\nHaroche, Phys. Rev. Lett. 91, 230405 (2003).\n[14] P. K. Pathak and G. S. Agarwal, Phys. Rev. A 71, 043823 (2005).\n[15] Zhi-Ming Zhang, Generating superposition and entanglement of squeezed vac uum states ,\nquant-ph/0604128.\n[16] M.J. Pfenning and L.H. Ford, Quantum Inequality Restrictions on Negative Energy Densiti es\nin Curved Spacetimes Doctoral Dissertation, (Dept. of Physics and Astronomy, Tu fts Univer-\nsity), gr-qc/9805037, Sec. 2.1.1.\n[17] C.C. Gerry and P.L. Knight, Introductory Quantum Optics , (Cambridge University Press,\nCambridge, 2005).\n[18] S.M. Barnett and P.M. Radmore, Methods in Theoretical Quantum Optics , (Oxford University\nPress, New York, 1997), Chap. 3.\n21" }, { "title": "0705.3574v1.The_geometry_of_density_states__positive_maps_and_tomograms.pdf", "content": "arXiv:0705.3574v1 [quant-ph] 24 May 2007The geometry of density states, positive maps\nand tomograms\nV.I. Man’ko,1G. Marmo,2E.C.G. Sudarshan3and F. Zaccaria2\n1P.N. Lebedev Physical Institute, Leninskii Prospect, 53, N oscow 119991 Russia\n2Dipartimento di Scienze Fisiche, Universit` a “Federico II ” di Napoli and Istituto\nNazionale di Fisica Nucleare, Sezione di Napoli, Complesso Universitario di Monte\nSant Angelo, Via Cintia, I-80126 Napoli, Italy\n3Physics Department, Center for Particle Physics, Universi ty of Texas, 78712\nAustin, Texas, USA\nEmailsmanko@sci.lebedev.ru marmo@na.infn.it\nsudarshan@physics.utexas.edu zaccaria@na.infn.it\nAbstract. The positive and not completely positive maps of density mat rices,\nwhich are contractive maps, are discussed as elements of a se migroup. A new kind\nof positive map (the purification map), which is nonlinear ma p, is introduced.\nThe density matrices are considered as vectors, linear maps among matrices are\nrepresented by superoperators given in the form of higher di mensional matrices.\nProbability representation of spin states (spin tomograph y) is reviewed and U(N)-\ntomogram of spin states is presented. Properties of the tomo grams as probability\ndistribution functions are studied. Notion of tomographic purity of spin states is\nintroduced. Entanglement and separability of density matr ices are expressed in\nterms of properties of the tomographic joint probability di stributions of random\nspin projections which depend also on unitary group paramet ers. A new positivity\ncriterion for hermitian matrices is formulated. An entangl ement criterion is given\nin terms of a function depending on unitary group parameters and semigroup of\npositive map parameters. The function is constructed as sum of moduli of U(N)-\ntomographic symbols of the hermitian matrix obtained after action on the density\nmatrix of composite system by a positive but not completely p ositive map of the\nsubsystem density matrix. Some two-qubit and two-qutritt s tates are considered as\nexamples of entangled states. The connection with the star- product quantisation is\ndiscussed. The structure of the set of density matrices and t heir relation to unitary\ngroupandLie algebra oftheunitarygroup are studied.Nonli near quantumevolution\nof state vector obtained by means of applying purification ru le of density matrices\nevolving via dynamical maps is considered. Some connection of positive maps and\nentanglement with random matrices is discussed and used.\nKeywords: unitary group, entanglement, adjoint representation, tom ogram, oper-\nator symbol, random matrix.\n1. Introduction\nThestates in quantum mechanics are associated with vectors in Hilbert\nspace [1] (it is better to say with rays) in the case of pure sta tes.\nFor mixed state, one associates the state with density matri x [2, 3].\nc/circlecopyrt2018Kluwer Academic Publishers. Printed in the Netherlands.\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.12 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\nIn classical mechanics (statistical mechanics), the state s are associ-\nated with joint probability distributions in phase space. T here is an\nessential difference in the concept of states in classical and quantum\nmechanics. This difference is clearly pointed out by the pheno menon of\nentanglement. Thenotion of entanglement [4] is related to t he quantum\nsuperposition rule of the states of subsystems for a given mu ltipartite\nsystem. For pure states, the notion of entanglement and sepa rability\ncan be given as follows.\nIf the wave function [5] of a state of a bipartite system is rep resented\nas the product of two wave functions depending on coordinate s of the\nsubsystems, the state is simply separable; otherwise, the s tate is sim-\nply entangled. An intrinsic approach to the entanglement me asure was\nsuggested in [6]. The measure was introduced as the distance between\nthe system density matrix and the tensor product of the assoc iated\nstates. For the subsystems, the association being realized via partial\ntraces. There are several other different characteristics an d measures\nof entanglement considered by several authors [7–13]. For e xample,\nthere are measures related to entropy (see, [14–24]). Also l inear entropy\nof entanglement was used in [25–27], “concurrences” in [28, 29] and\n“covariance entanglement measure” in [30]. Each of the enta nglement\nmeasures describes some degree of correlation between the s ubsystems’\nproperties.\nThe notion of entanglement is not an absolute notion for a giv en\nsystem but depends on the decomposition into subsystems. Th e same\nquantum state can be considered as entangled, if one kind of d ivision\nof the system into subsystems is given, or as completely dise ntangled,\nif another decomposition of the system into subsystems is co nsidered.\nFor instance, the state of two continuous quadratures can be entan-\ngled in Cartesian coordinates and disentangled in polar coo rdinates.\nCoordinates are considered as measurable observables labe lling the\nsubsystems of the given system. The choice of different subsys tems\nmathematically implies the existence of two different sets of the sub-\nsystems’ characteristics (we focus on bipartite case). We m ay consider\nthe Hilbert space of states H(1,2) orH(1′,2′). The Hilbert space for\nthe total system is, of course, the same but the index (1 ,2) means\nthat there are two sets of operators P1andP2, which select subsystem\nstates 1 and 2. The index (1′,2′) means that there are two other sets\nof operators P′\n1andP′\n2, which select subsystem states 1′and 2′.The\noperatorsP1,2andP′\n1′,2′have specific properties. They are represented\nas tensor products of operators acting in the space of states of the\nsubsystem 1 (or 2) and unit operators acting in the subsystem 2 (or\n1). In other words, we consider the space H, which can be treated as\nthe tensor product of spaces H(1) andH(2) orH(1′) andH(2′). In\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.2The geometry of density states, positive maps and tomograms 3\nthe subsystems 1 and 2, there are basis vectors |n1/angbracketrightand|m2/angbracketright, and in\nthe subsystems 1′and 2′there are basis vectors |n′\n1/angbracketrightand|m′\n2/angbracketright.The\nvectors|n1/angbracketright |m2/angbracketrightand|n′\n1/angbracketright |m′\n2/angbracketrightform the sets of basis vectors in\nthe composite Hilbert space, respectively. These two sets a re related\nby means of unitary transformation. An example of such a comp osite\nsystem is a bipartite spin system.\nIf one has spin- j1[the space H(1)] and spin- j2[the space H(2)]\nsystems, the combined system can be treated as having basis\n|j1m1/angbracketright |j2m2/angbracketright.\nAnotherbasisinthecomposite-system-state spacecanbeco nsidered\nintheform |jm/angbracketright, wherejisoneofthenumbers |j1−j2|,|j1−j2|+1,...,\nj1+j2andm=m1+m2. The basis |jm/angbracketrightis related to the basis\n|j1m1/angbracketright |j2m2/angbracketrightby means of the unitary transform given by Clebsch–\nGordon coefficients C(j1m1j2m2|jm). From the viewpoint of the given\ndefinition, the states |jm/angbracketrightare entangled states in the original basis.\nAnother example is the separation of the hydrogen atom in ter ms of\nparabolic coordinates used while discussing the Stark effect .\nThe spin states can be described by means of the tomographic\nmap [31–33]. For bipartite spin systems, the states were des cribed\nby the tomographic probabilities in [34, 35]. Some properti es of the\ntomographic spin description were studied in [36]. In the to mographic\napproach, the problems of the quantum state entanglement ca n be\ncast into the form of some relations among the probability di stribution\nfunctions. On the other hand, to have a clear picture of entan glement,\none needs a mathematical formulation of the properties of th e density\nmatrix of the composite system, a description of the linear s pace of\nthe composite system states. Since a density matrix is hermi tian, the\nspace of states may be embedded as a subset of the Lie algebra o f\nthe unitary group, carrying the adjoint representation of U(n2), where\nn2= (2j+ 1)2is the dimension of the spin states of two spinning\nparticles. Thus one may try to characterize the entanglemen t phenom-\nena by using various structures present in the space of the ad joint\nrepresentation of the U(n2) group.\nThe aim of this paper is to give a review of different aspects of\ndensity matrices and positive maps and connect entanglemen t prob-\nlems with the properties of tomographic probability distri butions and\ndiscuss the properties of the convex set of positive states f or composite\nsystembytakingintoaccount thesubsystemstructures.Weu sed[6]the\nHilbert–Schmidt distance to calculate the measure of entan glement as\nthe distance between a given state and the tensor productof t he partial\ntraces of the density matrix of the given state. In [37] anoth er measure\nof entanglement as a characteristic of subsystem correlati ons was intro-\nduced. This measure is determined via the covariance matrix of some\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.34 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\nobservables. A review of different approaches to the entangle ment no-\ntion and entanglement measures is given in [38] where the app roach to\ndescribe entanglement and separability of composite syste ms is based,\ne.g., on entropy methods.\nDue to a variety of approaches to the entanglement problem, o ne\nneeds to understand better what in reality the word “entangl ement”\ndescribes. Is it a synonym of the word “correlation” between two sub-\nsystems or does it have to capture some specific correlations attributed\ncompletely and only to the quantum domain?\nThe paper is organized as follows.\nIn section 2 we discuss division of composite systems onto su bsys-\ntems and relation of the density matrix to adjoint represent ation of\nunitary group in generic terms of vector representation of m atrices;\nwe study also completely positive maps of density matrices. In sec-\ntion 3 we consider a vector representation of probability di stribution\nfunctions and notion of distance between the probability di stributions\nand density matrices. In section 4 we present definition of se parable\nquantum state of a composite system and criterion of separab ility. In\nsection 5 the entanglement is considered in terms of operato r symbols.\nParticular tomographic probability representation of qua ntum states\nand tomographic symbols are reviewed in section 6. Symbols o f mul-\ntipartite states are studied in section 7. In section 8 spin t omography\nis reviewed. An example of qubit state is done in section 9. Th e uni-\ntary spin tomogram is introduced in section 10 while in secti on 11\ndynamical map and corresponding quantum evolution equatio ns are\ndiscussed as well as examples of concrete positive maps. Con clusions\nand perspectives are presented in section 12.\n2. Composite system\nIn this section, we review the meaning and notion of composit e system\nin terms of additional structures on the linear space of stat e for the\ncomposite system.\n2.1.Difference of states and observables\nIn quantum mechanics, there are two basic aspects, which are associ-\nated with linear operators acting in a Hilbert space. The firs t one is\nrelated to the concept of quantum state and the second one, to the\nconcept of observable. These two concepts of state and obser vable are\npaired via a map with values in probability measures on the re al line.\nOften states are described by Hermitian nonnegative, trace -class, ma-\ntrices. The observables are described by Hermitian operato rs. Though\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.4The geometry of density states, positive maps and tomograms 5\nbothstates and observables are identified with Hermitian ob jects, there\nis an essential difference between the corresponding objects . The ob-\nservables have an additional product structure. Thus we may consider\nthe product of two linear operators corresponding to observ ables.\nFor the states, the notion of product is redundant. The produ ct of\ntwo states is not a state. For states, one keeps only the linea r structure\nof vector space. For finite n-dimensional system, the Hermitian states\nand the Hermitian observables may be mapped into the Lie alge bra\nof the unitary group U(n). But the states correspond to nonnegative\nHermitian operators. Observables can be associated with bo th types\nof operators, including nonnegative and nonpositive ones. The space\nof states is therefore a convex-linear space which, in princ iple, is not\nequipped with a product structure. Due to this, transformat ions in the\nlinearspaceofstates neednotpreserveanyproductstructu re.Intheset\nofobservables,onehastobeconcernedwithwhatishappenin gwiththe\nproduct of operators when some transformations are perform ed. State\nvectors can be transformed into other state vectors. Densit y operators\nalso can be transformed. We will consider linear transforma tions of the\ndensityoperators. Thedensity operator hasnonnegative ei genvalues. In\nany representation, diagonal elements of density matrix ha ve physical\nmeaning of probability distribution function.\nDensity operator can be decomposed as a sum of eigenprojecto rs\nwith coefficients equal to its eigenvalues. Each one of the pro jectors\ndefines a pure state. There exists a basis in which every eigen projector\nof rank one is represented by a diagonal matrix of rank one wit h only\none matrix element equal to one and all other matrix elements equal\nto zero. Other density matrices with similar properties bel ong to the\norbit of the unitary group on the starting eigenprojector. D ependingon\nthe number of distinct nonzero values determines the class o f the orbit.\nSince density matrices of higher rank belong to an appropria te orbit\nof a convex sum of the different diagonal eigenprojectors (in s pecial\nbasis), we may say that generic density matrices belong to th e orbits of\nthe unitary group acting on the diagonal density matrices wh ich belong\nto the Cartan subalgebra of the Lie algebra of the unitary gro up. Any\nconvex sum of density matrices can be treated as a mean value o f a\nrandom density matrix. The positive coefficients of the conve x sum can\nbe interpreted as a probability distribution function whic h makes the\naveragingprovidingthefinalvalueoftheconvex sum.Theset ofdensity\nmatrices may be identified with the union of the orbits of the u nitary\ngroup acting on diagonal density matrices considered as ele ments of\nthe Cartan subalgebra.\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.56 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\n2.2.Matrices as vectors, density operators and\nsuperoperators\nWhen matrices represent states it may be convenient to ident ify them\nwithvectors. Inthiscase, adensitymatrixcanbeconsidere dasavector\nwith additional properties of its components. If the identi fications are\ndone elegantly, we can see the real Hilbert space of density m atrices in\nterms of vectors with real components. In this case, linear t ransforms\nof the matrix can be interpreted as matrices called superope rators. It\nmeans that density matrices–vectors undergoing real linea r transfor-\nmations are acted on by the matrices representing the action of the\nsuperoperators of the linear map. This construction can be c ontinued.\nThus we can get a chain of vector spaces of higher and higher di men-\nsions. Let us first introduce some extra constructions of the map of\na matrix onto a vector. Given a rectangular matrix Mwith elements\nMid, wherei= 1,2,...,nandd= 1,2,...,m, one can consider the\nmatrix as a vector /vectorMwithN=nmcomponents constructed by the\nfollowing rule:\nM1=M11,M2=M12,Mm=M1m,\n(1)\nMm+1=M21,...MN=Mnm.\nThus we construct the map M→/vectorM=ˆt/vectorMMM.\nWe have introduced thelinear operator ˆt/vectorMMwhich maps the matrix\nMonto avector /vectorM.Now weintroducetheinverseoperator ˆ p/vectorMMwhich\nmapsagivencolumnvectorinthespacewithdimension N=mnontoa\nrectangular matrix. Thismeansthat given avector /vectorM=M1,...,MN,\nwe relabel its components by introducing two indices i= 1,...,nand\nd= 1,...,m. The relabeling is accomplished according to (1). Then\nwe collect the relabeled components into a matrix table. Eve ntually we\nget the map\nˆp/vectorMM/vectorM=M. (2)\nThe composition of these two maps\nˆt/vectorMMˆp/vectorMM/vectorM=1·/vectorM (3)\nacts as the unit operator in the linear space of vectors.\nGiven an×nmatrix the map considered can also be applied. The\nmatrix can be treated as an n2-dimensional vector and, vice versa, the\nvector of dimension n2may be mappedby this procedureonto the n×n\nmatrix.\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.6The geometry of density states, positive maps and tomograms 7\nLet us consider a linear operator acting on the vector /vectorMand related\nto a linear transform of the matrix M. First, we study the correspon-\ndence of the linear transform of the form\nM→gM=Ml\ng (4)\nto the transform of the vector\n/vectorM →/vectorMl\ng=Ll\ng/vectorM. (5)\nOne can show that the n2×n2matrixLl\ngis determined by the tensor\nproduct of the n×nmatrixgandn×nunit matrix, i.e.,\nLl\ng=g⊗1. (6)\nAnalogously, the linear transform of the matrix Mof the form\nM→Mg=Mr\ng (7)\ninduces the linear transform of the vector /vectorMof the form\n/vectorM →/vectorMr\ng=ˆt/vectorMMMr\ng=Lr\ng/vectorM, (8)\nwhere then2×n2matrixLr\ngreads\nLr\ng= 1⊗gtr. (9)\nSimilarity transformation of the matrix Mof the form\nM→gMg−1(10)\ninduces the corresponding linear transform of the vector /vectorMof the form\n/vectorM →/vectorMs=Ls\ng/vectorM, (11)\nwhere then2×n2matrixLs\ngreads\nLs\ng=g⊗(g−1)tr. (12)\nStarting with vectors, one may ask how to identify on them a pr oduct\nstructure which would make ˆ p/vectorMNinto an algebra homomorphism. An\nassociative algebraic structure on the vector space may be d efined by\nimitating the procedure one uses to define star-products on t he space\nof functions on phase space. One can define the associative pr oduct of\ntwoN-vectors/vectorM1and/vectorM2using the rule\n/vectorM=/vectorM1⋆/vectorM2, (13)\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.78 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\nwhere\n/vectorMk=N/summationdisplay\nl,s=1Kk\nls(/vectorM1)l(/vectorM2)s. (14)\nIf one applies a linear transform to the vectors /vectorM1,/vectorM2,/vectorMof the form\n/vectorM1→/vectorM′\n1=L/vectorM1,/vectorM2→/vectorM′\n2=L/vectorM2,/vectorM →/vectorM′=L/vectorM,\nand requires the invariance of the star-product kernel, one finds\n/vectorM′\n1⋆/vectorM′\n2=/vectorM′,ifL=G⊗G−1tr, G∈GL(n).\nThe kernel Kk\nls(structure constants) which determines the associative\nstar-product satisfies a quadratic equation. Thus if one wan ts to make\nthe correspondence of the vector star-product to the standa rd matrix\nproduct (row by column), the matrix Mmust be constructed appro-\npriately. For example, if the vector star-product is commut ative, the\nmatrixMcorrespondingtothe N-vector/vectorMcanbechosenasadiagonal\nN×Nmatrix.Thisconsiderationshowsthatthemapofmatrices on the\nvectors provides the star-product of the vectors (defining t he structure\nconstantsorthekernelofthestar-product)and,conversel y, ifonestarts\nwith vectors and uses matrices with the standard multiplica tion rule,\nit will be the map to be determined by the structure constants (or by\nthe kernel of the vector star-product).\nThe constructed space of matrices associated with vectors e nables\noneto enlarge thedimensionality of thegroup acting in the l inear space\nof matrices in comparison with the standard one, i.e., we may relax the\nrequirement of invariance of the product structure. In gene ral, given\nan×nmatrixMthe left action, the right action, and the similarity\ntransformation of the matrix are related to the complex grou pGL(n).\nOn the other hand, the linear transformations in the linear s pace ofn2-\nvectors/vectorMobtained by using the introduced map are determined by\nthematrices belongingto thegroup GL(n2). Thereare transformations\non the vectors which cannot be simply represented on matrices. If\nM→Φ(M) is a linear homogeneous function of the matrix M, we\nmay represent it by\nΦab=Baa′,bb′Ma′b′.\nUnder rather clear conditions, Baa′,bb′can be expressed in terms of its\nnonnormalized left and right eigenvectors:\nBaa′,bb′=/summationdisplay\nνxaa′(ν)y†\nbb′(ν),\nbeing an index for eigenvalues, which corresponds to\nΦ(M) =xMy†=n2/summationdisplay\nν=1x(ν)My†(ν).\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.8The geometry of density states, positive maps and tomograms 9\nTherearepossiblelineartransformsonthematrices andcor respond-\ning linear transforms on the induced vector space which dono t give rise\nto a group structure but possess only the structure of algebr a. One can\ndescribe the map of n×nmatricesM(source space) onto vectors /vectorM\n(target space) usingspecificbasisinthespaceofthematric es. Thebasis\nis given by the matrices Ejk(j,k= 1,2,...,n) with all matrix elements\nequal to zero except the element in the jth row and kth column which\nis equal to unity. One has the obvious property\nMjk= Tr(MEjk). (15)\nInourprocedure,thebasismatrix Ejkismappedontothebasiscolumn-\nvector/vectorEjk, which has all components equal to zero except the unity\ncomponent related to the position in the matrix determined b y the\nnumbersjandk. Then one has\n/vectorM=n/summationdisplay\nj,k=1Tr(MEjk)/vectorEjk. (16)\nFor example, for similarity transformation of the finite mat rixM, one\nhas\n/vectorMs\ng=N/summationdisplay\nj,k=1Tr/parenleftig\ngMg−1Ejk/parenrightig/vectorEjk. (17)\nNow we will define the notion of ‘composite’ vector which corr e-\nsponds to dividing a quantum system into subsystems.\nWe will use the following terminology.\nIn general, the given linear space of dimensionality N=mnhas\na structure of a bipartite system, if the space is equipped wi th the\noperator ˆp/vectorMMand the matrix M(obtained by means of the map) has\nmatrix elements in factorizable form\nMid→xiyd. (18)\nThisM=x⊗ycorresponds to the special case of nonentangled states.\nOtherwise, one needs\nM=/summationdisplay\nνx(ν)⊗y(ν).\nIn fact, to consider in detail the entanglement phenomenon, in the\nbipartite system of spin-1/2, one has to introduce a hierarc hy of three\nlinear spaces. The first space of pure spin states is the two-d imensional\nlinear space of complex vectors\n|/vector x/angbracketright=/parenleftbiggx1\nx2/parenrightbigg\n. (19)\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.910 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\nIn this space, the scalar product is defined as follows:\n/angbracketleft/vector x|/vector y/angbracketright=x∗\n1y1+x∗\n2y2. (20)\nSo it is a two-dimensional Hilbert space. We do not equip this space\nwith a vector star-product structure. In the primary linear space, one\nintroduces linear operators ˆMwhich are described by 2 ×2 matrices\nM. Due to the map discussed in the previous section, the matric es\nare represented by 4-vectors /vectorMbelonging to the second complex 4-\ndimensional space. The star-product of the vectors /vectorMdetermined by\nthe kernel Kk\nlsis defined in such a manner in order to correspond to\nthe standard rule of multiplication of the matrices.\nIn addition to the star-product structure, we introduce the scalar\nproduct of the vectors /vectorM1and/vectorM2, in view of the definition\n/angbracketleft/vectorM1|/vectorM2/angbracketright= Tr(M†\n1M2), (21)\nwhich is the trace formula for the scalar product of matrices .\nThis means introducing the real metric gαβin the standard notation\nfor scalar product\n/angbracketleft/vectorM1|/vectorM2/angbracketright=4/summationdisplay\nα,β=1(M1)∗\nαgαβ(M2)β, (22)\nwhere the matrix gαβis of the form\ngαβ=\n1 0 0 0\n0 0 1 0\n0 1 0 0\n0 0 0 1\n, gαjgjβ=δαβ. (23)\nThe scalar product is invariant under the action of the group of non-\nsingular 4 ×4 matricesℓ, which satisfy the condition\ng=ℓ†gℓ. (24)\nThe product of matrices ℓsatisfies the same condition since g2= 1.\nThus, the space of operators ˆMin the primary two-dimensional\nspace of spin states is mapped onto the linear space which is e quipped\nwith a scalar product (metric Hilbert space structure) and a n associa-\ntive star-product (kernel satisfying the quadratic associ ativity equa-\ntion). In the linear space of the 4-vectors /vectorM, we introduce linear\noperators (superoperators), which can be associated with t he algebra\nof 4×4 complex matrices.\nLet usnow focuson densitymatrices. Thismeans thatourmatr ixM\nis considered as a density matrix ρwhich describes a quantum state.\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.10The geometry of density states, positive maps and tomograms 11\nWe consider here the action of the unitary transformation U(n) of\nthe density matrices and corresponding transformations on the vector\nspace. If one has the structure of a bipartite system, we also consider\nthe action of local gauge transformation both in the “source space” of\ndensitymatrices andinthe“target space”of thecorrespond ingvectors.\nThen×ndensity matrix ρhas matrix elements\nρik=ρ∗\nki,Trρ= 1,/angbracketleftψ|ρ|ψ/angbracketright ≥0. (25)\nSince the density matrix is hermitian, it can always be ident ified as an\nelement of the convex subset of the linear space associated w ith the Lie\nalgebra ofU(n) group, on which the group U(n) acts with the adjoint\nrepresentation\nρ→ρU=UρU†. (26)\nThe system is said to be bipartite if the space of representat ion is\nequipped with an additional structure. This means that for\nn2=n1·n2,\nwhere, for simplicity, n1=n2=n, one can make first the map of\nn×nmatrixρonton2-dimensional vector /vector ρaccording to the previous\nprocedure, i.e., one equips the space by an operator ˆt/vector ρρ. Given this\nvector one makes a relabeling of the vector /vector ρcomponents according to\nthe rule\n/vector ρ→ρid,ke, i,k= 1,2,...,n 1, d,e= 1,2,...,n 2,(27)\ni.e., obtaining again the quadratic matrix\nρq= ˆpρq/vector ρ/vector ρ. (28)\nThe unitary transform (26) of the density matrix induces a li near\ntransform of the vector /vector ρof the form\n/vector ρ→/vector ρU= (U⊗U∗)/vector ρ. (29)\nThere exist linear transforms (called positive maps) of the density ma-\ntrix, which preserve its trace, hermicity, and positivity. In some cases,\nthey have the following form introduced in [39]\nρ0→ρU=LUρ0=/summationdisplay\nkpkUkρ0U†\nk,/summationdisplay\nkpk= 1,(30)\nwhereUkare unitary matrices and pkare positive numbers.\nIf the initial density matrix is diagonal, i.e., it belongs t o the Cartan\nsubalgebra of the Lie algebra of the unitary group, the diago nal ele-\nmentsoftheobtainedmatrixgivea“smoother”probabilityd istribution\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.1112 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\nthan the initial one. A generic transformation preserving p reviously\nstated properties may be given in the form (see [39, 40])\nρ0→ρV=LVρ0=/summationdisplay\nkVkρ0V†\nk,/summationdisplay\nkV†\nkVk= 1.(31)\nFor example, if Vk(k= 1,2,...,N) are taken as square roots of or-\nthogonal projectors onto complete set of Nstate, the map provides\nthe map of the density matrix ρ0onto diagonal density matrix ρ0d\nwhich has the same diagonal elements as ρ0has. In this case, the\nmatricesVkhave the only nonzero matrix element which is equal to\none. Such a map may be called “decoherence map” because it rem oves\nall nondiagonal elements of the density matrix ρ0killing any phase\nrelations. In quantum information terminology, one uses al so the name\n“phase damping channel.” More general map may be given if one takes\nVkasNgeneric diagonal density matrices, in which eigenvalues ar e\nobtained by Ncircular permutations from the initial one. Due to this\nmap, one has a new matrix with the same diagonal matrix elemen ts\nbut with changed nondiagonal elements. The purity of this ma trix is\nsmaller then the purity of the initial one. This means that th e map\nis contractive. All matrices with the same diagonal element s up to\npermutations belong to a given orbit of the unitary group.\nFor a large number of terms with randomly chosen matrices Vkin\nthesumin(31), theabovemapgives themost stochasticdensi ty matrix\nρ0→ρs=L1ρ0= (n)−11.\nIts four-dimensional matrix L1for the qubit case has four matrix el-\nements different from zero. These matrix elements are equal to one.\nThey have the labels L11,L14,L41,L44. The map with two nonzero\nmatrix elements L41=L44= 0provides pure-statedensity matrix from\nanyρ0. The transform (30) is the partial case of the transform (31) . We\ndiscuss the transforms separately since they are used in the literature\nin the presented form.\nOne can see that the constructed map of density matrices onto\nvectors provides the corresponding transforms of the vecto rs, i.e.,\n/vector ρ0→/vector ρU=/summationdisplay\nkpk(Uk⊗U∗\nk)/vector ρ0 (32)\nand\n/vector ρ0→/vector ρV=/summationdisplay\nk(Vk⊗V∗\nk)/vector ρ0. (33)\nIt is obvious that the set of linear transforms of vectors, wh ich preserve\ntheir properties of being image of density matrices, is esse ntially larger\nthan the standard unitary transform of the density matrices .\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.12The geometry of density states, positive maps and tomograms 13\nFormulae (32) and (33) mean that the positive map superopera tors\nacting on the density matrix in the vector representation ar e described\nbyn2×n2matrices\nLU=/summationdisplay\nkpk(Uk⊗U∗\nk) (34)\nand\nLV=/summationdisplay\nkVk⊗V∗\nk, (35)\nrespectively.\nThe positive map is called “noncompletely positive” if\nL=/summationdisplay\nkVk⊗V∗\nk−/summationdisplay\nsvs⊗v∗\ns,/summationdisplay\nkV†\nkVk−/summationdisplay\nsv†\nsvs= 1.\nThis map is related to a possible “nonphysical” evolution of a subsys-\ntem.\nFormula(34) can beconsideredinthecontext of randommatri x rep-\nresentation. In fact, the matrix LUcan be interpreted as the weighted\nmean value of the random matrix Uk⊗U∗\nk. The dependence of matrix\nelements and positive numbers pkon indexkmeans that we have a\nprobabilitydistributionfunction pkandaveragingoftherandommatrix\nUk⊗U∗\nkby means of the distribution function. So the matrix LUreads\nLU=/angbracketleftU⊗U∗/angbracketright. (36)\nLet us consider an example of a 2 ×2 unitary matrix. We can consider\na matrix of the SU(2) group of the form\nu=/parenleftbiggα β\n−β∗α∗/parenrightbigg\n,|α|2+|β|2= 1. (37)\nThe 4×4 matrix LUtakes the form\nLU=\nℓ m m∗1−ℓ\n−n s −q n\n−n∗−q∗s∗n∗\n1−ℓ−m−m∗ℓ\n. (38)\nThe matrix elements of the matrix LUare the means\nm=/angbracketleftαβ∗/angbracketright,\nℓ=/angbracketleftαα∗/angbracketright,\nn=/angbracketleftαβ/angbracketright, (39)\ns=/angbracketleftα2/angbracketright,\nq=/angbracketleftβ2/angbracketright.\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.1314 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\nThe moduli of these matrix elements are smaller than unity.\nThe determinant of the matrix LUreads\ndetLU= (1−2ℓ)/parenleftig\n|q|2−|s|2/parenrightig\n+4Re/bracketleftig\nq∗m∗n+mns∗/bracketrightig\n.(40)\nIf one represents the matrix LUin block form\nLU=/parenleftbiggA B\nC D/parenrightbigg\n, (41)\nthen\nA=/parenleftbiggℓ m\n−n s/parenrightbigg\n, B=/parenleftbiggm∗1−ℓ\n−q n/parenrightbigg\n, (42)\nand\nD=σ2A∗σ2, C=−σ2B∗σ2, (43)\nwhereσ2is the Pauli matrix.\nOne can check that the product of two different matrices LUcan\nbe cast in the same form. This means that the matrices LUform a\n9-parameter compact semigroup. It means that the product of two\nmatrices from the set (semigroup) belongs to the same set. It means\nthat composition is inner like the one for groups. There is a u nity\nelement in the semigroup, however, there exist elements whi ch have no\ninverse. In our case, these elements are described, e.g., by the matrices\nwith zero determinant. Also the elements, which are matrice s with\nnonzero determinants, have no inverse elements in this set, since the\nmap corresponding to the inverse of these matrices is not pos itive one.\nFor example, in the case ℓ= 1/2 andm= 0, one has the matrices\nA=/parenleftbigg1/2 0\n−n s/parenrightbigg\n, B=/parenleftbigg0 1/2\n−q n/parenrightbigg\n. (44)\nThe determinant of the matrix LUin this case is equal to zero. All the\nmatrices LUhave the eigenvector\n/vector ρ0=\n1/2\n0\n0\n1/2\n, (45)\ni.e.,\nLU/vector ρ0=/vector ρ0. (46)\nThis eigenvector corresponds to the density matrix\nρ1=/parenleftbigg1/2 0\n0 1/2/parenrightbigg\n, (47)\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.14The geometry of density states, positive maps and tomograms 15\nwhich is obviously invariant of the positive map.\nFor random matrix, one has correlations of the random matrix ele-\nments, e.g., /angbracketleftαα∗/angbracketright /negationslash=/angbracketleftα/angbracketright/angbracketleftα∗/angbracketright.\nThe matrix Lp\nLp=\n1 0 0 0\n0 0 1 0\n0 1 0 0\n0 0 0 1\n(48)\nmaps the vector\n/vector ρin=\nρ11\nρ12\nρ21\nρ22\n(49)\nonto the vector\n/vector ρt=\nρ11\nρ21\nρ12\nρ22\n. (50)\nThis means that the positive map (48) connects the positive d ensity\nmatrix with its transpose (or complex conjugate). This map c an be\npresented as the connection of the matrix ρwith its transpose of the\nform\nρ→ρtr=ρ∗=1\n2/parenleftig\nρ+σ1ρσ1−σ2ρσ2+σ3ρσ3/parenrightig\n.\nThere is no unitary transform connecting these matrices.\nThere is noncompletely positive map in the N-dimensional case,\nwhich is given by the generalized formula (for some ε)\nρ→ρs=−ερ+1+ε\nN1N.\nIn quantum information terminology, it is called “depolari zing chan-\nnel.”\nFor the qubit case, matrix form of this map reads\nL=\n1−ε\n20 01+ε\n2\n0−ε0 0\n0 0−ε0\n1+ε\n20 01−ε\n2\n. (51)\nThusweconstructedthematrixrepresentationofthepositi vemapof\ndensityoperators ofthespin-1/2system. Thisparticulars et of matrices\nrealize the representation of the semigroup of real numbers −1≤ε≤1.\nIf one considers the product ε1ε2=ε3, the result ε3belongs to the\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.1516 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\nsemigroup. Only two elements 1 and −1 of the semigroup have the\ninverse. These two elements form the finite subgroup of the se migroup.\nThe semigroup itself without element ε= 0 can be embedded into the\ngroup of real numbers with natural multiplication rule. Eac h matrix\nLhas an inverse element in this group but all the parameters of the\ninverse elements ηlive out of the segment −1,1. The group of the real\nnumbersiscommutative. Thematricesofthenonunitaryrepr esentation\nof this group commute too. It means that we have nonunitary re ducible\nrepresentation of the semigroup which is also commutative. To con-\nstruct this representation, one needs to use the map of matri ces on the\nvectors discussed in the previous section. Formulae (31) an d (35) can\nbe interpreted also in the context of the random matrix repre sentation,\nbut we use the uniform distribution for averaging in this cas e. So one\nhas equality (35) in the form\nLV=/angbracketleftV⊗V∗/angbracketright (52)\nand the equality\n/angbracketleftV†V/angbracketright= 1, (53)\nwhich provides constraints for the random matrices Vused.\nUsing the random matrix formalism, the positive (but not com -\npletely positive) maps can be presented in the form\nL=/angbracketleftV⊗V∗/angbracketright−/angbracketleftv⊗v∗/angbracketright,/angbracketleftV†V/angbracketright−/angbracketleftv†v/angbracketright= 1.\nOne can characterize the action of positive map on a density m atrixρ\nby the parameter\nκ=Tr(Lρ)2\nTrρ2=µLρ\nµρ≤1.\nAs a remark we note that in [39] the positive maps (30) and (31) were\nused to describe the non-Hamiltonian evolution of quantum s tates for\nopen systems.\nWe haveto pointout that, in general, such evolution isnot de scribed\nby first-order-in-time differential equation. As in the previ ous case, if\nthere are additional structures for the matrix in the form\nρid,ke→xiydzkte, (54)\nwhich means associating with the initial linear space two ex tra lin-\near spaces where xi,zkare considered as vector components in the\nn1-dimensional linear space and yd,teare vector components in n2-\ndimensional vector space, we see that one has bipartite stru cture of\nthe initial space of state [bipartite structure of the space of adjoint\nrepresentations of the group U(n)].\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.16The geometry of density states, positive maps and tomograms 17\nUsually the adjoint representation of any group is defined pe r se\nwithout any reference to possible substructures. Here we in troduce\nthe space with extra structure. In addition to being the spac e of the\nadjointrepresentation of thegroup U(n), ithasthestructureofabipar-\ntite system. The generalization to multipartite ( N-partite) structure is\nstraightforward. One needs only the representation of posi tive integer\nn2in the form\nn2=N/productdisplay\nk=1n2\nk. (55)\nIf one considers the more general map given by superoperator (35)\nrewritten in the form\nLV=/angbracketleftV⊗V∗/angbracketright,/angbracketleftV†V/angbracketright= 1,\nthe number of parameters determining the matrix LVcan be easily\nevaluated. For example, for n= 2,\nV=/parenleftbigga b\nc d/parenrightbigg\n, V∗=/parenleftbigga∗b∗\nc∗d∗/parenrightbigg\n,\nwhere the matrix elements are complex numbers, the normaliz ation\ncondition provides four constraints for the real and imagin ary parts of\nthe matrix elements of the following matrix:\nLV=\n/angbracketleft|a|2/angbracketright /angbracketleftab∗/angbracketright /angbracketleftba∗/angbracketright /angbracketleftbb∗/angbracketright\n/angbracketleftac∗/angbracketright /angbracketleftad∗/angbracketright /angbracketleftbc∗/angbracketright /angbracketleftbd∗/angbracketright\n/angbracketleftca∗/angbracketright /angbracketleftcb∗/angbracketright /angbracketleftda∗/angbracketright /angbracketleftdb∗/angbracketright\n/angbracketleftcc∗/angbracketright /angbracketleftcd∗/angbracketright /angbracketleftdc∗/angbracketright /angbracketleftdd∗/angbracketright\n,\nnamely,\n/angbracketleft|a|2/angbracketright+/angbracketleft|c|2/angbracketright= 1,/angbracketleft|b|2/angbracketright+/angbracketleft|d|2/angbracketright= 1,/angbracketlefta∗b/angbracketright+/angbracketleftc∗d/angbracketright= 0.\nDuetothestructureofthematrix LV,therearesixcomplexparameters\n/angbracketleftab∗/angbracketright,/angbracketleftac∗/angbracketright,/angbracketleftad∗/angbracketright,/angbracketleftbc∗/angbracketright,/angbracketleftbd∗/angbracketright,/angbracketleftcd∗/angbracketright\nor 12 real parameters.\nThe geometrical picture of the positive map can be clarified i f one\nconsiders the transform of the positive density matrix into another\ndensity matrix as the transform of an ellipsoid into another ellipsoid. A\ngeneric positive transform means a generic transform of the ellipsoid,\nwhich changes its orientation, values of semiaxis, and posi tion in the\nspace. But the transform does not change the ellipsoidal sur face into a\nhyperboloidalor paraboloidal surface. For purestates, th epositive den-\nsity matrix defines the quadratic form which is maximally deg enerated.\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.1718 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\nIn this sense, the “ellipsoid” includes all its degenerate f orms corre-\nsponding to the density matrix of rank less than n(inn-dimensional\ncase). The number of parameters defining the map /angbracketleftV⊗V∗/angbracketrightin the\nn-dimensional case is equal to n2(n2−1).\nThe linear space of Hermitian matrices is also equipped with the\ncommutator structure defining the Lie algebra of the group U(n). The\nkernel that defines this structure (Lie product structure) i s determined\nby the kernel that determines the star-product.\n3. Distributions as vectors\nIn quantum mechanics, one needs the concept of distance betw een the\nquantum states. In this section, we consider the notion of di stance\nbetween the quantum states in terms of vectors. First, let us discuss\nthe notion of distance between conventional probability di stributions.\nThis notion is well known in the classical probability theor y.\nGiven the probability distribution P(k),k= 1,2,...N, one can\nintroduce the vector /vectorPin the form of a column with components P1=\nP(1), P2=P(2),..., P N=P(N).The vector satisfies the condition\nN/summationdisplay\nk=1Pk= 1. (56)\nThis set of vectors does not form a linear space but only a conv ex sub-\nset. Nevertheless, in this set one can introduce a distance b etween two\ndistributions by using the one suggested by the vector space structure\nof the ambient space:\nD2=/parenleftig/vectorP1−/vectorP2/parenrightig2=/summationdisplay\nkP1kP1k+/summationdisplay\nkP2kP2k−2/summationdisplay\nkP1kP2k.(57)\nOf course, one may use other identifications of distribution s with\nvectors.\nSince allP(k)≥0, one can introduce Pk=/radicalbig\nP(k) as components\nof the vector /vectorP. The/vectorPcan be thought of as a column with nonnegative\ncomponents. Then the distance between the two distribution s takes the\nform\nD2=/parenleftig/vectorP1−/vectorP2/parenrightig2= 2−2/summationdisplay\nk/radicalig\nP1(k)P2(k). (58)\nThe two different definitions (56) and (57) can be used as distan ces\nbetween the distributions.\nLet us discuss now the notion of distance between the quantum\nstates determined by density matrices. In the density-matr ix space (in\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.18The geometry of density states, positive maps and tomograms 19\nthe set of linear space of the adjoint U(n) representation), one can\nintroduce distances analogously. The first case is\nTr(ρ1−ρ2)2=D2(59)\nand the second case is\nTr(√ρ1−√ρ2)2=D2. (60)\nIn fact, the distances introduced can be written naturally a s norms of\nvectors associated to density matrices\nD2=|/vector ρ1−/vector ρ2|2(61)\nand\nD2=/parenleftig/vector(√ρ1)−/vector(√ρ2)/parenrightig2, (62)\nrespectively.\nIn the above expressions, we use scalar product of vectors /vector ρ1and/vector ρ2\nas well as scalar products of vectors/vector(√ρ1) and/vector(√ρ2), respectively.\nBoth definitions immediately follow by identification of eit her matri-\ncesρ1andρ2with vectors according tothe map of theprevious sections\nor matrices√ρ1and√ρ2with vectors. Since the density matrices ρ1\nandρ2have nonnegative eigenvalues, the matrices√ρ1and√ρ2are\ndefined without ambiguity. This means that the vectors/vector(√ρ1) and\n/vector(√ρ2) are also defined without ambiguity. It is obvious that using this\nconstruction and introducing linear map of positive vector s/vector√ρ, one in-\nduces nonlinear map of density matrices. Other analogous fu nctions, in\naddition to square root function, can be used to create other nonlinear\npositive maps.\n4. Separable systems and separability criterion\nAccording to the definition, the system density matrix is cal led sep-\narable (for composite system) but not simply separable, if t here is\ndecomposition of the form\nρAB=/summationdisplay\nkpk/parenleftig\nρ(k)\nA⊗ρ(k)\nB/parenrightig\n,/summationdisplay\nkpk= 1,1≥pk≥0.(63)\nThis is Hilbert’s problem of biquadrates. Is a positive biqu adratic the\npositive sum of products of positive quadratics? In this for mula, one\nmay use also sum over two different indices. Using another labe lling in\nsuch sum over two different indices, this sum can be always repr esented\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.1920 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\nas the sum over only one index. The formula does not demand ort hogo-\nnality of the density operators ρ(k)\nAandρ(k)\nBfor different k. Since every\ndensity matrix is a convex sum of pure density matrices, one c ould\ndemand that ρ(k)\nAandρ(k)\nBbe pure. This formula can be interpreted in\nthe context of random matrix representation reading\nρAB=/angbracketleftρA⊗ρB/angbracketright, (64)\nwhereρAandρBare considered as random density matrices of the\nsubsystems AandB, respectively. One can use the clarified structure\nofthedensity matrix set as theunionoforbits obtained by ac tion of the\nunitary group on projectors of rank one with matrix form cont aining\nonly one nonzero matrix element. Then the separable density matrix\nof bipartite composite system can be always written as the su m of\nn1n2tensor products (or corresponding mean tensor product), i. e., in\n(64) the factors are state projectors. Each of tensor produc ts contains\nrandom unitary matrices of local transforms of the fixed loca l projector\nfor one subsystem and for the second subsystem. It means that an\narbitrary projector of rank one of a subsystem can be always p resented\nin the product form ρ(k)\nA=u(k)\nAρAu(k)†\nA, whereu(k)\nAis a unitary local\ntransform and ρAis a fixed projector.\nThere are several criteria for the system to be separable. We suggest\nin the next sections a new approach to the problem of separabi lity\nand entanglement based on the tomographic probability desc ription of\nquantum states. The states which cannot be represented in th e form\n(63) by definition are called entangled states [38]. Thus the states are\nentangled if in formula (63) at least one coefficient (or more) piis\nnegative which means that the positive ones can take values g reater\nthan unity.\nLet us discuss the condition for the system state to be separa ble.\nAccording to the partial transpose criterion [41], the syst em is sepa-\nrable if the partial transpose of the matrix ρAB(63) gives a positive\ndensity matrix. This condition is necessary but not sufficien t. Let us\ndiscuss this condition within the framework of positive-ma p matrix\nrepresentation. For example, for a spin-1/2 bipartite syst em, we have\nshown that the map of a density matrix onto its transpose belo ngs\nto the matrix semigroup of matrices L. One should point out that\nthis map cannot be obtained by means of averaging with all pos itive\nprobability distributions pk. On the other hand, it is obvious that the\ngeneric criterion, which contains the Peres criterion as a p artial case,\ncan be formulated as follows.\nLet us map the density matrix ρABof a bipartite system onto vector\n/vector ρAB. Let the vector /vector ρABbe acted upon by an arbitrary matrix, which\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.20The geometry of density states, positive maps and tomograms 21\nrepresents the positive maps in subsystems AandB. Thus we get a\nnew vector\n/vector ρ(p)\nAB=/parenleftig\nLA⊗LB/parenrightig\n/vector ρAB. (65)\nLet us construct the density matrix ρ(p)\nABusing the inverse map of the\nvectors onto matrices. If the initial density matrix is sepa rable, the new\ndensity matrix ρ(p)\nABmust be positive (and separable).\nIn the case of the bipartite spin-1/2 system, by choosing LA= 1 and\nwithLBbeing the matrix coinciding with the matrix gαβ, we obtain\nthe Peres criterion as a partial case of the criterion of sepa rability\nformulated above. Thus, our criterion means that the separa ble matrix\nkeeps positivity under the action of the tensor product of tw o semi-\ngroups. In the case of the bipartite spin-1/2 system, the 16 ×16 matrix\nof the semigroup tensor product of positive contractive map s (52) is\ndetermined by 24 parameters. Among these parameters, one ca n have\nsome correlations.\nLet us discuss the positive map (52) which is determined by th e\nsemigroup for the n-dimensional system. It can be realized also as\nfollows.\nThen×nHermitian genericmatrix ρcan bemappedontoessentially\nrealn2-vector/vector ρby the map described above. The complex vector /vector ρis\nmapped onto the real vector /vector ρrby multiplying by the unitary matrix\nS, i.e.,\n/vector ρr=S/vector ρ, /vector ρ =S−1/vector ρr. (66)\nThe matrix Sis composed from nunity blocks and the blocks\nS(jk)\nb=1√\n2/parenleftbigg1 1\n−i i/parenrightbigg\n, (67)\nwherejcorresponds to a column and kcorresponds to a row in the\nmatrixρ.\nFor example, in the case n= 2, one has the vector /vector ρrof the form\n/vector ρr=\nρ11√\n2Reρ12√\n2Imρ12\nρ22\n. (68)\nOne has the equalities\n/vector ρ2\nr=/vector ρ2= Trρ2. (69)\nThe semigroup preserves the trace of the density matrix. Als o the\ndiscrete transforms, which are described by the matrix with diagonal\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.2122 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\nmatrix blocks of the form\nD=\n1 0 0 0\n0 1 0 0\n0 0−1 0\n0 0 0 1\n, (70)\npreserve positivity of the density matrix.\nFor the spin case, the semigroup contains 12 parameters.\nThus, the direct product of the semigroup (52) and the discre te\ngroup of the transform Ddefines positive map preserving positivity\nof the density operator. One can include also all the matrice s which\ncorrespond to other not completely positive maps. The consi dered rep-\nresentation contains onlyreal vectors andtheirreal posit ivetransforms.\nThis means that one can construct representation of semigro up of\npositive maps by real matrices.\n5. Symbols, star-product and entanglement\nIn this section, we describe how entangled states and separa ble states\ncan be studied using properties of symbols and density opera tors of\ndifferent kinds, e.g., from the viewpoint of the Wigner functi on or\ntomogram. The general scheme of constructing the operator s ymbols is\nas follows [36].\nGiven a Hilbert space Hand an operator ˆAacting on this space,\nlet us suppose that we have a set of operators ˆU(x) acting transitively\nonHparametrized by n-dimensional vectors x= (x1,x2,...,x n). We\nconstruct the c-number function fˆA(x) (we call it the symbol of the\noperator ˆA) using the definition\nfˆA(x) = Tr/bracketleftigˆAˆU(x)/bracketrightig\n. (71)\nLet us suppose that relation (71) has an inverse, i.e., there exists a set\nof operators ˆD(x) acting on the Hilbert space such that\nˆA=/integraldisplay\nfˆA(x)ˆD(x)dx,TrˆA=/integraldisplay\nfˆA(x)TrˆD(x)dx.(72)\nOne needs a measure in xto define the integral in above formulae.\nThen, we will consider relations (71) and (72) as relations d etermin-\ning the invertible map from the operator ˆAonto the function fˆA(x).\nMultiplying both sides of Eq. (2) by the operator ˆU(x′) and taking the\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.22The geometry of density states, positive maps and tomograms 23\ntrace, one can satisfy the consistency condition for the ope ratorsˆU(x′)\nandˆD(x)\nTr/bracketleftigˆU(x′)ˆD(x)/bracketrightig\n=δ/parenleftbigx′−x/parenrightbig. (73)\nThe consistency condition (73) follows from the relation\nfˆA(x) =/integraldisplay\nK(x,x′)fˆA(x′)dx′. (74)\nThe kernel in (74) is equal to the standard Dirac delta-funct ion, if the\nset of functions fˆA(x) is a complete set.\nIn fact, we could consider relations of the form\nˆA→fˆA(x) (75)\nand\nfˆA(x)→ˆA. (76)\nThe most important property of the map is the existence of the asso-\nciative product (star-product) of the functions.\nWe introduce the product (star-product) of two functions fˆA(x) and\nfˆB(x) corresponding to two operators ˆAandˆBby the relationships\nfˆAˆB(x) =fˆA(x)∗fˆB(x) := Tr/bracketleftigˆAˆBˆU(x)/bracketrightig\n. (77)\nSince the standard product of operators on a Hilbert space is an asso-\nciative product, i.e., ˆA(ˆBˆC) = (ˆAˆB)ˆC, it is obvious that formula (77)\ndefines an associative product for the functions fˆA(x), i.e.,\nfˆA(x)∗/parenleftig\nfˆB(x)∗fˆC(x)/parenrightig\n=/parenleftig\nfˆA(x)∗fˆB(x)/parenrightig\n∗fˆC(x).(78)\nUsingformulae(71) and (72), onecan writedown acompositio n rule\nfor two symbols fˆA(x) andfˆB(x), which determines the star-product\nof these symbols. The composition rule is described by the fo rmula\nfˆA(x)∗fˆB(x) =/integraldisplay\nfˆA(x′′)fˆB(x′)K(x′′,x′,x)dx′dx′′.(79)\nThe kernel in the integral of (79) is determined by the trace o f the\nproduct of the basic operators, which we use to construct the map\nK(x′′,x′,x) = Tr/bracketleftigˆD(x′′)ˆD(x′)ˆU(x)/bracketrightig\n. (80)\nThe kernel function satisfies the composition property K∗K=K.\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.2324 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\n6. Tomographic representation\nIn this section, we will consider an example of the probabili ty repre-\nsentation of quantum mechanics [42]. In the probability rep resentation\nof quantum mechanics, the state is described by a family of pr obabil-\nities [43–45]. According to the general scheme, one can intr oduce for\nthe operator ˆAthe function fˆA(x), where\nx= (x1,x2,x3)≡(X,µ,ν),\nwhich we denote here as wˆA(X,µ,ν) depending on the position Xand\nthe parameters µandνof the reference frame\nwˆA(X,µ,ν) = Tr/bracketleftigˆAˆU(x)/bracketrightig\n. (81)\nWe call the function wˆA(X,µ,ν) the tomographic symbol of the oper-\natorˆA. The operator ˆU(x) is given by\nˆU(x)≡ˆU(X,µ,ν) = exp/parenleftbiggiλ\n2(ˆqˆp+ ˆpˆq)/parenrightbigg\nexp/parenleftbiggiθ\n2/parenleftig\nˆq2+ ˆp2/parenrightig/parenrightbigg\n|X/angbracketright/angbracketleftX|\n×exp/parenleftbigg\n−iθ\n2/parenleftig\nˆq2+ ˆp2/parenrightig/parenrightbigg\nexp/parenleftbigg\n−iλ\n2(ˆqˆp+ ˆpˆq)/parenrightbigg\n=ˆUµν|X/angbracketright/angbracketleftX|ˆU†\nµν. (82)\nThe tomographic symbol is the homogeneous version of the Moy al\nphase-space density. The angle θand parameter λin terms of the\nreference phase-space frame parameters are given by\nµ=eλcosθ, ν =e−λsinθ,\nthat is, ˆqand ˆpare position and momentum operators\nˆq|X/angbracketright=X|X/angbracketright (83)\nand|X/angbracketright/angbracketleftX|is the projection density. One has the canonical transform\nof quadratures\nˆX=ˆUµνˆqˆU†\nµν=µˆq+νˆp,\nˆP=ˆUµνˆpˆU†\nµν=1+/radicalbig\n1−4µ2ν2\n2µˆp−1−/radicalbig\n1−4µ2ν2\n2νˆq.\nUsing the approach of [46] one obtains the relationship\nˆU(X,µ,ν) =δ(X−µˆq−νˆp).\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.24The geometry of density states, positive maps and tomograms 25\nIn the case we are considering, the inverse transform determ ining the\noperator in terms of the tomogram [see Eq. (72)] will be of the form\nˆA=/integraldisplay\nwˆA(X,µ,ν)ˆD(X,µ,ν)dXdµdν, (84)\nwhere\nˆD(x)≡ˆD(X,µ,ν) =1\n2πexp(iX−iνˆp−iµˆq).(85)\nThe trace of the above operator, which provides the kernel de ter-\nmining the trace of an arbitrary operator in the tomographic represen-\ntation, reads\nTrˆD(x) =eiXδ(µ)δ(ν).\nThe function wˆA(X,µ,ν) satisfies the relation\nwˆA(λX,λµ,λν ) =1\n|λ|wˆA(X,µ,ν). (86)\nThis means that the tomographic symbols of operators are hom oge-\nneous functions of three variables.\nIf one takes two operators ˆA1andˆA2, which are expressed through\nthe corresponding functions by the formulas\nˆA1=/integraldisplay\nwˆA1(X′,µ′,ν′)ˆD(X′,µ′,ν′)dX′dµ′dν′,\n(87)\nˆA2=/integraldisplay\nwˆA2(X′′,µ′′,ν′′)ˆD(X′′,µ′′,ν′′)dX′′dµ′′dν′′,\nandˆAdenotes theproductof ˆA1andˆA2,then thefunction wˆA(X,µ,ν),\nwhichcorrespondsto ˆA,isthestar-productofthefunctions wˆA1(X,µ,ν)\nandwˆA2(X,µ,ν). Thus this product\nwˆA(X,µ,ν) =wˆA1(X,µ,ν)∗wˆA2(X,µ,ν)\nreads\nwˆA(X,µ,ν) =/integraldisplay\nwˆA1(x′′)wˆA2(x′)K(x′′,x′,x)dx′′dx′,(88)\nwith kernel given by\nK(x′′,x′,x) = Tr/bracketleftigˆD(X′′,µ′′,ν′′)ˆD(X′,µ′,ν′)ˆU(X,µ,ν)/bracketrightig\n.(89)\nThe explicit form of the kernel reads\nK(X1,µ1,ν1,X2,µ2,ν2,Xµ,ν)\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.2526 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\n=δ/parenleftig\nµ(ν1+ν2)−ν(µ1+µ2)/parenrightig\n4π2exp/parenleftbiggi\n2/braceleftig\n(ν1µ2−ν2µ1)+2X1+2X2\n−/bracketleftigg\n1−/radicalbig\n1−4µ2ν2\nν(ν1+ν2)+1+/radicalbig\n1−4ν2µ2\nµ(µ1+µ2)/bracketrightigg\nX/bracerightigg/parenrightigg\n.\n(90)\n7. Multipartite systems\nLet us assume that for multimode ( N-mode) system one has\nˆU(/vector y) =N/productdisplay\nk=1⊗ˆU/parenleftig\n/vector x(k)/parenrightig\n, (91)\nˆD(/vector y) =N/productdisplay\nk=1⊗ˆD/parenleftig\n/vector x(k)/parenrightig\n, (92)\nwhere\n/vector y=/parenleftig\nx(1)\n1,x(1)\n2,...,x(1)\nm,x(2)\n1,x(2)\n2,...,x(N)\nm/parenrightig\n. (93)\nThis means that the symbol of the density operator of the comp osite\nsystem reads\nfρ(/vector y) = Tr/bracketleftig\nˆρN/productdisplay\nk=1⊗ˆU(/vector x(k))/bracketrightig\n. (94)\nThe inverse transform reads\nˆρ=/integraldisplay\nd/vector yfρ(/vector y)N/productdisplay\nk=1⊗ˆD(/vector x(k)), d/vector y =N/productdisplay\nk=1m/productdisplay\ns=1dx(k)\ns.(95)\nNow we formulate the properties of the symbols in the case of\nentangled and separable states, respectively.\nGiven a composite m-partite system with density operator ˆ ρ.\nIf the nonnegative operator can be presented in the form of a “ prob-\nabilistic sum”\nˆρ=/summationdisplay\n/vector zP(/vector z)ˆρ(a1)\n/vector z⊗ˆρ(a2)\n/vector z⊗···⊗ˆρ(am)\n/vector z, (96)\nwith positive probability distribution function P(/vector z), where the com-\nponents of /vector zcan be either discrete or continuous, we call the state a\n“separable state.” Without loss of generality, all factors in the tensor\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.26The geometry of density states, positive maps and tomograms 27\nproducts can be considered as projectors of rank one. This me ans that\nthe symbol of the state can be presented in the form\nfρ(/vector y) =/summationdisplay\n/vector zP(/vector z)m/productdisplay\nk=1f(ak)\nρ(/vector xk,/vector z). (97)\nAnalogous formula can be written for the tomogram of separab le state.\nWe point out that in the multipartite case one can introduce r an-\ndom symbols and represent the symbol of separable density ma trix\nof composite system as mean value of pointwise products of sy mbols of\nsubsystem density operators. As in the bipartite case, one c an use sum\nover different indices but this sum can be always reduced to the sum\nover only one indexcommon forall the subsystems.Itis impor tant that\nfor separable state its symbol always can be represented as t he sum\ncontaining number of summants which is equal to dimensional ity of\ncomposite system. Each term in the sum is equal to mean value o f ran-\ndom projector. The random projector is constructed as the pr oduct of\ndiagonal projectorsofrankoneineach subsystemconsidere dinrandom\nlocal basis obtained by means of random unitary local transf orms.\n8. Spin tomography\nBelow we concentrate on bipartite spin systems.\nThe tomographic probability (spin tomogram) completely de ter-\nmines the density matrix of a spin state. It has been introduc ed in\n[31, 32, 36]. The tomographic probability for the spin- jstate is defined\nvia the density matrix by the formula\n/angbracketleftjm|D†(g)ρD(g)|jm/angbracketright=W(j)(m,/vector0), m=−j,−j+1,...,j,\n(98)\nwhereD(g) is the matrix of SU(2)-group representation depending on\nthe group element gdetermined by three Euler angles. It is useful to\ngeneralize the construction of spin tomogram.\nOne can introduce unitary spin tomograms w(m,u) by replacing in\nabove formula (98) the matrix D(g) by generic unitary matrix u. For\nthe case of higher spins j= 1,3/2,2,..., then×nprojector matrix\nρ1=\n1 0···0\n0 0···0\n· · ··· ·\n· · ··· ·\n0 0···0\n, n= 2j+1 (99)\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.2728 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\nhas the unitary spin tomogram denoted as\nw1(j,u) =|u11|2, w1(j−1,u) =|u12|2, ... w 1(−j,u) =|u1n|2.\n(100)\nOther projectors\nρk=\n0 0··· ···0\n· · ··· · ·\n0···1···0\n· · · ··· ·\n0 0· ···0\n, (101)\nin which unity is located in kth column, have the tomogram wk(m,u)\nof the form\nwk(j,u) =|uk1|2, wk(j−1,u) =|uk2|2, ... w k(−j,u) =|ukn|2.\n(102)\nIn connection with the decomposition of any density matrix i n theform\nρ=/summationdisplay\njkρjkEjk, (103)\nthe unitary spin tomogram can be presented in form of the deco mpo-\nsition\nwρ(m,u) =/summationdisplay\njkρjkwjk(m,u), (104)\nwherewjk(m,u) are basic unitary spin symbols of transition operators\nEjkof the form\nwjk(m,u) =/angbracketleftjm|u†Ejku|jm/angbracketright. (105)\nIf one uses a map\nρ→ρ′, (106)\nthe unitary spin tomogram is transformed as\nwρ(m,u)→w′\nρ(m,u) =/summationdisplay\njkρ′\njkwjk(m,u). (107)\nIf the transform (106) is a linear one\nρjk→ρ′\njk=Ljk,psρps, (108)\nthe transform reads\nw′\nρ(m,u) =/summationdisplay\npsρpsw′\nps(m,u). (109)\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.28The geometry of density states, positive maps and tomograms 29\nHere\nw′\nps(m,u) =/summationdisplay\njkLjk,pswjk(m,u) (110)\nis the linear transform of the basic tomographic symbols of t he opera-\ntorsEjk.\nLet us now discuss some properties of usual spin tomograms.\nThe set of the tomogram values for each /vector0 is an overcomplete set.\nWe need only a finite number of independent locations which wi ll give\ninformationonthedensitymatrixofthespinstate. Duetoth estructure\nof the formula, there are only two Euler angles involved. The y are\ncombined into the unit vector\n/vector0 = (cosφsinϑ,sinφsinϑ,cosϑ). (111)\nThis is the Hopf map from S3toS2.\nThe physical meaning of the probability W(m,/vector0) is the following.\nIt is the probability to find, in the state with the density mat rixρ,\nthe spin projection on direction /vector0 equal tom. For a bipartite system,\nthe spin tomogram is defined as follows:\nW(m1m2/vector01/vector02) =/angbracketleftj1m1j2m2|D†(g1)D†(g2)ρD(g1)D(g2)|j1m1j2m2/angbracketright.\n(112)\nItcompletely determinesthedensitymatrix ρ.Ithasthemeaningofthe\njoint probability distribution for spin j1andj2projections m1andm2\non directions /vector01and/vector02.Since the map ρ⇋Wis linear and invertible,\nthedefinitionof separablesystemcan berewritteninthefol lowingform\nfor the decomposition of the joint probability into a sum of p roducts\n(of factorized probabilities):\nW(m1m2/vector01/vector02) =/summationdisplay\nkpkW(k)(m1/vector01)˜W(k)(m2/vector02).(113)\nThis form can be considered to formulate the criterion of sep arability\nof the two-spin state.\nOne can present this formula in the form\nW(m1m2/vector01/vector02) =/angbracketleftW(m1/vector01)˜W(m2/vector02)/angbracketright, (114)\nwherewe interpretthepositive numbers pkas probability distributions.\nThus separability means the possibility to represent joint probability\ndistribution in the form of average product of two random pro bability\ndistributions.\nThe state is separable iff the tomogram can be written in the fo rm\n(113) with/summationtext\nkpk= 1,pk≥0.Itseems that wesimplyusethedefinition\nbut, in fact, we cast the problem of separability into the for m of the\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.2930 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\nproperty of the positive joint probability distribution of two random\nvariables. This is an area of probability theory and one can u se the\nresults and theorems on joint probability distributions. I f one does not\nuse any theorem, one has to study the solvability of relation (113)\nconsidered as the equation for unknown probability distrib utionpkand\nunknown probability functions W(k)(m1/vector01) andW(k)(m2/vector02).\n9. Example of spin-1/2 bipartite system\nFor the spin-1/2 state, the generic density matrix can be pre sented in\nthe form\nρ=1\n2(1+/vector σ·/vector n), /vector n= (n1,n2,n3), (115)\nwhere/vector σare Pauli matrices and /vector n2≤1,with the vector /vector nfor a pure\nstate being the unit vector. This decomposition means that w e use as\nbasis in 4-dimensional vector space the vectors correspond ing to the\nPauli matrices and the unit matrix, i.e.,\n/vector σ1=\n0\n1\n1\n0\n, /vector σ 2=\n0\n−i\ni\n0\n, /vector σ 3=\n1\n0\n0\n−1\n,/vector1 =\n1\n0\n0\n1\n.\n(116)\nThe density matrix vector\n/vector ρ=\nρ11\nρ12\nρ21\nρ22\n(117)\nis decomposed in terms of the basis vectors\n/vector ρ=1\n2/parenleftig\n/vector1+n1/vector σ1+n2/vector σ2+n3/vector σ3/parenrightig\n. (118)\nThis means that the spin tomogram of the spin-1/2 state can be given\nin the form\nW/parenleftbigg1\n2,/vector0/parenrightbigg\n=1\n2+/vector n·/vector0\n2, W/parenleftbigg\n−1\n2,/vector0/parenrightbigg\n=1\n2−/vector n·/vector0\n2.(119)\nWe can consider tomograms of specific spin state. If the state is pure\nstate with density matrix\nρ+=/parenleftbigg1 0\n0 0/parenrightbigg\n, (120)\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.30The geometry of density states, positive maps and tomograms 31\nthe spin tomogram W(m,/vector0), where\nm=±1\n2,/vector0 = (sinθcosϕ,sinθsinϕ,cosθ)\nhas the values\nW+/parenleftbigg1\n2,/vector0/parenrightbigg\n= cos2θ\n2, W+/parenleftbigg\n−1\n2,/vector0/parenrightbigg\n= sin2θ\n2.(121)\nThe tomogram of the pure state\nρ−=/parenleftbigg0 0\n0 1/parenrightbigg\n, (122)\nhas the values\nW−/parenleftbigg1\n2,/vector0/parenrightbigg\n= sin2θ\n2= cos2π−θ\n2,\n(123)\nW−/parenleftbigg\n−1\n2,/vector0/parenrightbigg\n= cos2θ\n2= sin2π−θ\n2.\nThe spin tomogram of the diagonal density matrix\nρd=/parenleftbiggρ110\n0ρ22/parenrightbigg\n(124)\nequals\nWd(m,/vector0) =ρ11W+(m,/vector0+)+ρ22W−(m,/vector0−). (125)\nThe generic density matrix which has eigenvalues ρ11andρ22can be\npresented in the form\nu0ρdu†\n0, (126)\nwhere the unitary matrix u0has columns containing components of\nnormalized eigenvectors of the density matrix ρ.\nThis means that the tomogram of the state with the matrix ρreads\nWρ(m,/vector0) =/angbracketleftm|u†u0ρdu†\n0u|m/angbracketright. (127)\nThe elements of the group can be combined\nu†u0= ˜u. (128)\nThus the tomogram becomes\nWρ(m,/vector0) =Wd(m,/vector0′), (129)\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.3132 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\nwhere the angle /vector0′corresponds to the Euler angle calculated from the\nproduct of two unitary matrices u†\n0u.\nOnecanusethepropertyofnumbers ρ11andρ22tointerpretformula\n(125) as averaging\nWd(m,/vector0) =/angbracketleftW(m,/vector0′)/angbracketright, (130)\nwhere one interprets two functions W+(m,/vector0) andW−(m,/vector0′) as the\nrealization of “random” probability distribution functio nW±(m,/vector0).\nThen one has\nWρ(m,/vector0) =/angbracketleftW(m,/vector0′)/angbracketright. (131)\nThe new vector /vector0′has the parameter θ′obtained from the initial pa-\nrameterθby action of the unitary matrix on the initial unitary matrix\nu.\nInserting these probability values into relation (113) for each value\nofk, we get the relationships\nW/parenleftbigg1\n2,1\n2,/vector01,/vector02/parenrightbigg\n=1\n4+1\n2/parenleftigg/summationdisplay\nkpk/vector nk/parenrightigg\n·/vector01+1\n2/parenleftigg/summationdisplay\nkpk/vector n∗\nk/parenrightigg\n·/vector02\n+/summationdisplay\nkpk/parenleftig\n/vector nk·/vector01/parenrightig/parenleftig\n/vector n∗\nk·/vector02/parenrightig\n, (132)\nW/parenleftbigg1\n2,−1\n2,/vector01,/vector02/parenrightbigg\n=1\n4+1\n2/parenleftigg/summationdisplay\nkpk/vector nk/parenrightigg\n·/vector01−1\n2/parenleftigg/summationdisplay\nkpk/vector n∗\nk/parenrightigg\n·/vector02\n−/summationdisplay\nkpk/parenleftig\n/vector nk·/vector01/parenrightig/parenleftig\n/vector n∗\nk·/vector02/parenrightig\n, (133)\nW/parenleftbigg\n−1\n2,1\n2,/vector01,/vector02/parenrightbigg\n=1\n4−1\n2/parenleftigg/summationdisplay\nkpk/vector nk/parenrightigg\n·/vector01+1\n2/parenleftigg/summationdisplay\nkpk/vector n∗\nk/parenrightigg\n·/vector02\n−/summationdisplay\nkpk/parenleftig\n/vector nk·/vector01/parenrightig/parenleftig\n/vector n∗\nk·/vector02/parenrightig\n. (134)\nOne has the normalization property\n1/2/summationdisplay\nm1,m2=−1/2W(m1m2/vector01/vector02) = 1. (135)\nOne easily gets\nW/parenleftbigg1\n2,1\n2,/vector01,/vector02/parenrightbigg\n+W/parenleftbigg1\n2,−1\n2,/vector01,/vector02/parenrightbigg\n=1\n2+/parenleftigg/summationdisplay\nkpk/vector nk/parenrightigg\n·/vector01.(136)\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.32The geometry of density states, positive maps and tomograms 33\nThis means that the derivative in /vector01on the left-hand side gives\n∂\n∂/vector01/bracketleftbigg\nW/parenleftbigg1\n2,1\n2,/vector01,/vector02/parenrightbigg\n+W/parenleftbigg1\n2,−1\n2,/vector01,/vector02/parenrightbigg/bracketrightbigg\n=/parenleftigg/summationdisplay\nkpk/vector nk/parenrightigg\n.(137)\nAnalogously\n∂\n∂/vector02/bracketleftbigg\nW/parenleftbigg1\n2,1\n2,/vector01,/vector02/parenrightbigg\n+W/parenleftbigg\n−1\n2,1\n2,/vector01,/vector02/parenrightbigg/bracketrightbigg\n=/parenleftigg/summationdisplay\nkpk/vector n(⋆)\nk/parenrightigg\n.(138)\nTaking the sum of (133) and (134)) one sees that\n1\n2∂\n∂/vector0i∂\n∂/vector0j/bracketleftbigg\nW/parenleftbigg1\n2,−1\n2,/vector01,/vector02/parenrightbigg\n+W/parenleftbigg\n−1\n2,1\n2,/vector01,/vector02/parenrightbigg/bracketrightbigg\n=−/summationdisplay\nkpk(nk)i(n(⋆)\nk)j. (139)\nSince we look for the solution where pk≥0, we can introduce\n/vectorNk=√pk/vector nk,/vectorN(⋆)\nk=√pk/vector n(⋆)\nk. (140)\nThis means that the derivative in (139) can be presented as a t ensor\n−Tij=/summationdisplay\nk(Nk)i(N(⋆)\nk)j. (141)\nOne has/summationdisplay\nkpk/vector nk=/summationdisplay\nk√pk/vectorNk, (142)\n/summationdisplay\nkpk/vector n⋆\nk=/summationdisplay\nk√pk/vectorN(⋆)\nk. (143)\nThe conditions of solvability of the obtained equations is a criterion for\nseparability or entanglement of a bipartite quantum spin st ate. Using\nthe arguments on the representation of the tomogram (tomogr aphic\nsymbol) as sum of random basic projector symbols we get that f or two\nqubits the separable state has the tomogram with following p roperties.\nAll four values of joint probability distribution function are equal to\nmean values of product of two cosine of two different angles squ ared,\nproduct of sine of two different angles squared and product of s ine and\ncosine squared, respectively. The entangled matrix does no t provide\nsuch structure.\nAs an example, we consider the Werner state. For the Werner st ate\n(see, e.g., [47]) with the density matrix\nρAB=\n1+p\n40 0p\n2\n01−p\n40 0\n0 01−p\n40\np\n20 01+p\n4\n,\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.3334 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\n(144)\nρA=ρB=1\n2/parenleftbigg1 0\n0 1/parenrightbigg\n,\nonecan reconstruct theknownresultsthat for p<1/3 thestate is sepa-\nrable and for p>1/3 the state is entangled, since in the decomposition\nof the density operator in the form (113) the state\nρ0=1\n4\n1 0 0 0\n0 1 0 0\n0 0 1 0\n0 0 0 1\n(145)\nhas the weight p0= (1−3p)/4.\nForp>1/3, the coefficient pobecomes negative.\nThere is some extension of the presented consideration.\nLet us consider the state with the density matrix (nonnegati ve and\nHermitian)\nρ=\nR110 0R12\n0ρ11ρ120\n0ρ21ρ220\nR210 0R22\n,Trρ= 1. (146)\nUsing the procedure of mapping the matrix onto vector /vector ρand applying\nto the vector the nonlocal linear transform corresponding t o the Peres\npartial transposeandmakingtheinversemapofthetransfor medvector\nonto the matrix, we obtain\nρm=\nR110 0ρ12\n0ρ11R120\n0R21ρ220\nρ210 0R22\n. (147)\nIn the case of separable matrix ρ, the matrix ρmis a nonnegative\nmatrix. Calculating the eigenvalues of ρmand applying the condition\nof their positivity, we get\nR11R22≥ |ρ12|2, ρ 11ρ22≥ |R12|2. (148)\nViolation of these inequalities gives a signal that ρis entangled. For\nWerner state (144), Eq. (148) means\n1+p>0,1−p>2p, (149)\nwhich recovers the condition of separability p<1/3 mentioned above.\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.34The geometry of density states, positive maps and tomograms 35\nThe joint probability distribution (112) of separable stat e is positive\nafter making the local and nonlocal (partial transpose-lik e) transforms\nconnected with positive map semigroup. But for entangled st ate, func-\ntion (112) can take negative values after making this map in t he func-\ntion and replacing on the right-hand side of this equality th e product of\ntwo matrices D(g) by generic unitary transform u. This is a criterion\nof entanglement in terms of unitary spin tomogram of the stat e of\nmultiparticle system.\nA simpler and more transparent case is the generalized Werne r\nmodel with density matrix\nρ=1\n4(1+µ1σ1⊗τ1+µ2σ2⊗τ2+µ3σ3⊗τ3).(150)\nHere the density matrix is expressed in terms of tensor produ cts of two\nsets of Pauli matrices σkandτk(k= 1,2,3), which are chosen in the\nstandard form.\nIts eigenvalues are\n1−µ1−µ2−µ3,1+µ1+µ2−µ3,1+µ1−µ2+µ3,1−µ1+µ2+µ3.\nThese eigenvalues are related to the vertices of a regular te trahedron.\nThe partially time-reversed density matrix is\n/tildewideρ=1\n4(1−µ1σ1⊗τ1−µ2σ2⊗τ2−µ3σ3⊗τ3),(151)\nwhich may be viewed as\n/tildewideρ=L(1)⊗L(2)ρwithL(1)ρ(1)=ρ(1), L(2)ρ(2)= 1−ρ(2).\n(152)\nThe eigenvalues of this are\n1+µ1+µ2+µ3,1+µ1−µ2−µ3,1−µ1+µ2−µ3,1−µ1−µ2+µ3.\nTheseform an inverted tetrahedron and they have the common d omain\nwhich is a regular octahedron. The unitary spin tomograms ca n be\nwritten down by inspection and we may verify that all the rela tions\nrequired by the separability criterion (see the next sectio n for details)\naresatisfied byany pointinsidetheoctahedron for ρandforL(1)⊗L(2)ρ\nbuttherelations connected withpositivity condition expr essedinterms\nof positivity of unitary spin tomogram fail when it lies outs ide.\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.3536 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\n10. Tomogram of the group U(n)\nIn this section we discuss in more detail the separability cr iterion using\nintroduced notion of unitary spin tomogram.\nIn order to formulate a criterion of separability for a bipar tite spin\nsystem with spin j1andj2, we introduce the tomogram w(/vectorl,/vector m,g(n))\nfor the group U(n), where\nn=n1n2, n 1= 2j1+1, n 2= 2j2+1,\nandg(n)are parameters of the group element. Vectors /vectorland/vector mlabel a\nbasis|/vectorl,/vector m/angbracketrightof the fundamental representation of the group U(n). For\nexample, since this representation is irreducible, being r educed to the\nrepresentation of the U(n1)⊗U(n2) subgroup of the group U(n), the\nbasis can be chosen as the product of basis vectors:\n|j1,m1/angbracketright |j2,m2/angbracketright=|j1,j2,m1,m2/angbracketright. (153)\nDue to the irreducibility of this representation of the grou pU(n) and\nits subgroup, there exists a unitary transform u/vectorl/vector m\nj1j2m1m2|/vectorl,/vector m/angbracketrightsuch\nthat\n|j1,j2,m1,m2/angbracketright=/summationdisplay\n/vectorl/vector mu/vectorl/vector m\nj1j2m1m2|/vectorl,/vector m/angbracketright, (154)\n|/vectorl/vector m/angbracketright=/summationdisplay\nm1m2(u−1)/vectorl/vector m\nj1j2m1m2|jl,j2,m1,m2/angbracketright.(155)\nOne can define the U(n) tomogram for a Hermitian nonnegative n×n\ndensity matrix ρ, which belongs to the Lie algebra of the group U(n),\nby a generic formula\nw(/vectorl,/vector m,g(n)) =/angbracketleft/vectorl,/vector m|U†(g(n))ρU(g(n))|/vectorl,/vector m/angbracketright.(156)\nFormula (156) defines the tomogram in the basis |/vectorl,/vector m/angbracketrightfor arbitrary\nirreducible representation of the unitary group. But below we focus\nonly on tomograms connected with spins.\nLet us define the U(n) tomogram using the basis |j1,j2,m1,m2/angbracketright\nnamely for fundamental representation, i.e.,\nw(j1,j2)(m1,m2,g(n))\n=/angbracketleftj1,j2,m1,m2|U†(g(n))ρU(g(n))|j1,j2,m1,m2/angbracketright.(157)\nThis unitary spin tomogram becomes the spin-tomogram [34] f or the\ng(n)∈U(2)⊗U(2) subgroup of the group U(n). The properties of this\ntomogram follow fromits definition as the joint probability distribution\nof two random spin projections m1,m2depending on g(n)parameters.\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.36The geometry of density states, positive maps and tomograms 37\nOne has the normalization condition\n/summationdisplay\nm1,m2w(j1,j2)(m1,m2,g(n)) = 1. (158)\nAlso all the probabilities are nonnegative, i.e.,\nw(j1,j2)(ml,m2,g(n))≥0. (159)\nDue to this, one has\n/summationdisplay\nm1,m2|w(j1,j2)(ml,m2,g(n))|= 1. (160)\nFor the spin-tomogram,\ng(n)→/parenleftig/vectorO1,/vectorO2/parenrightig\n(161)\nand\nw(j1,j2)(ml,m2,g(n))→w(m1,m2,/vectorO1,/vectorO2). (162)\nThe separability and entanglement condition discussed in t he previ-\nous section for a bipartite spin-tomogram can be considered also from\nthe viewpoint of the properties of a U(n) tomogram. If the two-spin\nn×ndensitymatrix ρisseparable,itremainsseparableundertheaction\nofthegenericpositivemapofthesubsystemdensitymatrice s. Thismap\ncan be described as follows.\nLetρbemappedontovector /vector ρwithn2components.Thecomponents\nare simply ordered rows of the matrix ρ, i.e.,\n/vector ρ=/parenleftig\nρ11,ρ12,...,ρ1n,ρ21,ρ22,...,ρnn,/parenrightig\n. (163)\nLet then2×n2matrixLbe taken in the form\nL=/summationdisplay\nspsL(j1)\ns⊗L(j2)\ns, p s≥0,/summationdisplay\nsps= 1,(164)\nwhere then1×n1matrixL(j1)\nsand then2×n2matrixL(j2)\nsdescribe the\npositive maps of density matrices of spin- j1and spin-j2subsystems,\nrespectively. We map vector /vector ρonto vector /vector ρL\n/vector ρL=L/vector ρ (165)\nand construct the n×nmatrixρL, which corresponds to the vector /vector ρL.\nThen we consider the U(n) tomogram of the matrix ρL, i.e.,\nw(j1,j2)\nL(ml,m2,g(n))\n=/angbracketleftj1,j2,m1,m2|U†(g(n))ρLU(g(n))|j1,j2,ml,m2/angbracketright.(166)\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.3738 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\nUsing this tomogram we introduce the function\nF(g(n),L) =/summationdisplay\nm1,m2/vextendsingle/vextendsingle/vextendsinglew(j1,j2)\nL(m1,m2,g(n))/vextendsingle/vextendsingle/vextendsingle. (167)\nFor separable states, this function does not depend on the U(n)-group\nparameterg(n)and positive-map matrix elements of the matrix L.\nFor the normalized density matrix ρof the bipartite spin-system,\nthis function reads\nF(g(n),L) = 1. (168)\nFor entangled states, this function depends on g(n)andLand is not\nequalto unity. Thispropertycan bechosen asanecessary and sufficient\ncondition for separability of bipartite spin-states. We in troduce also\ntomographic purity parameter µkofkth order by the formula\nµk(g(n),L) =/summationdisplay\nm1m2/vextendsingle/vextendsingle/vextendsinglew(j1,j2)\nL(m1,m2,g(h))/vextendsingle/vextendsingle/vextendsinglek.\nFor identity semigroup element Land specific g(n)\n0unitary transform\ndiagonalizing the density matrix, the tomographic purity µ2is identical\nto purity parameter of the state ρ. The parameters for k= 2,3,...,\ncorrespond to Tr ρk+1.\nIn fact, the formulated approach can be extended to multipar tite\nsystems too. The generalization is as follows.\nGivenNspin-systems with spins j1,j2,...,jN, let us consider the\ngroupU(n) with\nn=N/productdisplay\nk=1nk, n k= 2jk+1. (169)\nLet us introduce the basis\n|/vector m/angbracketright=N/productdisplay\nk=1|jkmk/angbracketright (170)\ninthelinear spaceofthefundamental representation ofthe groupU(n).\nWe define now the U(n) tomogram of a state with the n×nmatrixρ:\nwρ(/vector m,g(n)) =/angbracketleft/vector m|U†(g(n))ρU(g(n))|/vector m/angbracketright. (171)\nFor a positive Hermitian matrix ρwith Trρ= 1, we formulate the\ncriterion of separability as follows.\nLet the map matrix Lbe of the form\nL=/summationdisplay\nsps/parenleftigN/productdisplay\nk=1⊗L(k)\ns/parenrightig\n, p s≥0,/summationdisplay\nsps= 1,(172)\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.38The geometry of density states, positive maps and tomograms 39\nwhereL(k)\nsis the positive-map matrix of the density matrix of the\nkth spin subsystem. We construct the matrix ρLas in the case of the\nbipartite system using the matrix L. The function\nF(g(n),L) =/summationdisplay\n/vector m|wρL(/vector m,g(n))| ≥1 (173)\nis equal to unity for separable state and depends on the matri xLand\nU(n)-parameters g(n)for entangled states.\nThis criterion can beapplied also in the case of continuous v ariables,\ne.g., for Gaussian states of photons. Function (173) can pro vide the\nmeasure of entanglement. Thus one can use the maximum value ( or a\nmean value) of this function as a characteristic of entangle ment. In the\nprevious section, we considered the generalized Werner sta tes. Using\nthe above criterion, one can get the domain of values of the pa rameters\nof the states for which one has separability or entanglement . In fact,\nthe separability criterion is related to the following posi tivity criterion\nof finite or infinite (trace class) matrix A. The matrix Ais positive iff\nthe sum of moduli of diagonal matrix elements of the matrix UAU†\nis equal to a positive trace of the matrix Afor an arbitrary unitary\nmatrixU.\n11. Dynamical map and purification\nIn this section, we consider the connection of positive maps with pu-\nrification procedure. In fact, formula\nρ→ρ′=/summationdisplay\nkpkUkρU†\nk, (174)\nwhereUkare unitary operators, can be considered in the form\nρ→ρ′=/summationdisplay\nkpkρk, p k≥0,/summationdisplay\nkpk= 1.(175)\nHere the density operators ρkread\nρk=UkρU†\nk, (176)\nand the maps which are not sufficiently general keep the most de gener-\nate density matrix fixed.This formis theform of probabilist ic addition.\nThis mixture of density operators can be purified with the hel p of a\nfiducial rank one projector P0\nρ′→ρ′′=N\n/summationdisplay\nkj√pkpjρkP0ρj/radicalbigTrρkP0ρjP0\n, (177)\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.3940 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\nwhereNis a normalization constant\nN−1= Tr\n/summationdisplay\nkj√pkpjρkP0ρj/radicalbigTrρkP0ρjP0\n. (178)\nThe normalization is unnecessary if all ρkare mutually orthogonal. We\ncall this map a purification map. It maps the density matrix of mixed\nstate on the density matrix of pure state.\nThe map (174) could be interpreted as the evolution in time of the\ninitialmatrix ρ0consideringunitaryoperators Uk(t)dependingontime.\nThus one has\nρ0→ρ(t) =/summationdisplay\nkpkUk(t)ρ0U†\nk(t). (179)\nIn this case, the purification procedure provides the dynami cal map of\na pure state\n|ψ0/angbracketright/angbracketleftψ0|→|ψ(t)/angbracketright/angbracketleftψ(t)|, (180)\nwhere|ψ(t)/angbracketrightobeys a nonlinear equation and, in the general case, this\nmap does not definea one parameter group of transformations n ot even\nlocally.\nFor some specific cases, the evolution (179) can be described by a\nsemigroup.Thedensitymatrix(179)obeysthenafirst-order differential\nequation in time for this case [27–29].\nMore specifically, the reason why there is no differential equa tion in\ntime for the generic case is due to the absence of the property\nρij(t2) =/summationdisplay\nmnKmn\nij(t2,t1)ρmn(t1), (181)\nwhere the kernel of evolution operator satisfies\nKmn\nij(t3,t2)Kpq\nmn(t2,t1) =Kpg\nij(t3,t1). (182)\nThus, via a purification procedure and a dynamical map applie d to\na density matrix we get a pure state (nonlinear dynamical map ). This\nmapcanbeusedinnonlinearmodelsofquantumevolution.Man ylinear\npositive maps both completely positive and not completely p ositive are\ncontractive. We define a positive map Las “contractive” or “dilating”\nifTr(Lρ)2≤Tr(ρ)2orTr(Lρ)2≥Tr(ρ)2, respectively. Thismeans, for\nexample, that purity parameter µ= Trρ2after performing the positive\nmap generically becomes smaller. There are maps for which th e purity\nparameter is preserved, for example,\nρ→ρtr, ρ→ −ρ+2\nN1. (183)\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.40The geometry of density states, positive maps and tomograms 41\nThese linear maps include also unitary transform\nρ→uρu†. (184)\nThere are maps which provide dilation. For qubit system, mat rix\nL=\n1 0 0 1\n0 0 0 0\n0 0 0 0\n0 0 0 0\n. (185)\nacting on arbitrary vector /vector ρ0correspondingto a density matrix ρ0gives\nthe pure state\nρf=/parenleftbigg1 0\n0 0/parenrightbigg\n. (186)\nThe matrices Lεthe inverse matrices exist for ε/negationslash= 0. But these inverse\nmatricesdonotprovidepositivetracepreservingmaps.Sin cethepurifi-\ncation procedure provides a positive map, which increases t he purity\nparameter, the composition of linear map with the purificati on map\nprovides the possibility to recover the initial density mat rixρwhich\nwas the object of action of a positive linear map. It means tha t the\npurification map ˆLpcan give\nˆLP0(Lρ) = 1ρ (187)\nfor any density matrix ρbut the choice of fiducial projector depends\nonρ(the initial condition).\nThus one has also for completely positive maps\nρ→ρ′=/summationdisplay\nkρ′\nk, ρ k=VkρV†\nk,/summationdisplay\nkV†\nkVk= 1.(188)\nMaking polar decomposition\nρk=√ρ0kUk, U kU†\nk= 1, ρ 0k≥0\nand introducing the positive numbers pk= Trρ0k, we construct the\nmap\nρ′→ρ′′=\n\n/summationdisplay\nkj√pkpj˜ρkP0˜ρj+ ˜ρjP0˜ρk/radicalbigT˜ρkP0˜ρjP0\n\n,/summationdisplay\nkpk= 1, pk˜ρk=ρk.\n(189)\nThematrix ρ′′isamatrixofrankoneforanyrankonefiducialprojector\nP0. The projector P0is restricted to be not orthogonal to the generic\nmatrixρk. TakingNorthogonal projectors P(s)\n0(s= 1,2,...,N) and\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.4142 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\nobtainingNprojectorsρ′′\ns, one can combine them in order to get the\ninitial matrix ρ. It means that one can take convex sum of the N\npure states ρ′′\nsto recover the initial mixed state ρ. Another way to\nmake the state with higher purity was demonstrated usingthe modified\npurification procedure in [48]. For qubit state, one has\nρ=p1ρ1+p2ρ2+κ√p1p2ρ1P0ρ2+ρ2P0ρ1√Trρ1P0ρ2P0, p1+p2= 1,(190)\nwhere the decoherence parameter 0 ≤κ≤1 is used. If κ∼1, we\nincrease purity.\nLet us discuss the map (188) using its matrix form, i.e.,\nραβ→ρ′\nαβ=/summationdisplay\nijLαβ,ijρij. (191)\nThe matrix Lαβ,ijis expressed in terms of the matrices Vkas\nLαβ,ij=/summationdisplay\nk(Vk)αi(V∗\nk)βj. (192)\nOne can construct another positive map [49]\nρ→ρ′=/summationdisplay\nkrkTr(Rkρ), (193)\nwhererkare density matrices and Rkare positive operators satisfying\nthe normalization condition\n/summationdisplay\nkRk= 1. (194)\nThe matrix corresponding to this map (called entanglement b reaking\nmap [50]) reads\nLb\nαβ,ij=/summationdisplay\nk(rk)αβ(R∗\nk)ij. (195)\nThe entanglement breaking map is contractive positive map. There\nexist some special cases of completely positive maps. For ex ample,\nρ→ −ερ+1+ε\nNρ (196)\ndiffers from the depolarizing map by replacing the unity opera tor by\nthe density operator. Another map reads\nρ→1−diagρ\nN. (197)\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.42The geometry of density states, positive maps and tomograms 43\nThe decoherence map (phase damping map) of the kind\nρij→/braceleftbiggρij, i=j\nλρij, i/negationslash=j,(198)\nwhere|λ|<1 provides contractive map with uniform change of off-\ndiagonal matrix elements of the density matrix.\nLet us discuss the property of tomogram of bipartite system w ith\ndensity matrix ρ12. If the density matrix is separable, than the depo-\nlarizing map of the second subsystem provides the following density\nmatrix\nρ12→ρε=−ερ12+1+ε\nN2ρ(1)⊗12, (199)\nwhere\nρ(1)= Tr2(ρ12) (200)\nand 12is theN2-dimensional unity matrix. Then one has the property\nof unitary spin tomogram\nwε(m1,m2,g(n)) =−εw12(m1,m2,g(n))+1+ε\nN2w(m1,m2,g(n)),(201)\nwhereg(n)is matrix of U/parenleftig\n(2j1+1)(2j2+1)/parenrightig\nunitary transform;\nwε(m1,m2,g(n)) is the tomogram of transformed density matrix of bi-\npartite system;\nw(m1,m2,g(n)) is the unitary spin tomogram of tensor product of par-\ntial traceρ(1)over the second subsystem’s coordinates of the density\nmatrixρ12and unity operator 1 2;\nw12(m1,m2,g(n))istheunitaryspintomogram ofthestatewithdensity\nmatrixρ12.\nThe criterion of separability means\nj1/summationdisplay\nm1=−j1j2/summationdisplay\nm2=−j2/vextendsingle/vextendsingle/vextendsingle/vextendsingle1+ε\n2j2+1w(m1,m2,g(n))−εw12(m1,m2,g(n))/vextendsingle/vextendsingle/vextendsingle/vextendsingle= 1\n(202)\nfor arbitrary g(n)andε.\nFor Werner states ρW, the tomogram of transformed state (in this\ncase, it means that p→ −εp) is related to the initial-state tomogram\nwW\nwε(m1,m2,g(n)) =−εw12(m1,m2,g(n))+1+ε\n4.(203)\nThe criterion of separability yields\n1/2/summationdisplay\nm1,m2=−1/2/vextendsingle/vextendsingle/vextendsingle/vextendsingle1+ε\n4−εwW(m1,m2,g(n))/vextendsingle/vextendsingle/vextendsingle/vextendsingle= 1.(204)\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.4344 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\nEquality (204) takes place for arbitrary g(n)andεonly for |p| ≤1/3.\nForp>1/3, the above sum depends on g(n)andεand it is larger than\none.\nIt is obvious if one calculates the tomogram using the elemen t of the\nunitary group of the form\ng(n)\n0=\n0 0 0 1\n0 1 0 0\n0 0 1 0\n1 0 0 0\n. (205)\nAt this point, the sum (204) reads\n1/2/summationdisplay\nm1,m2=−1/2/vextendsingle/vextendsingle/vextendsingle/vextendsingle1+ε\n4−εwW(m1,m2,g(n))/vextendsingle/vextendsingle/vextendsingle/vextendsingle= 3/vextendsingle/vextendsingle/vextendsingle/vextendsingle1+εp\n4/vextendsingle/vextendsingle/vextendsingle/vextendsingle+/vextendsingle/vextendsingle/vextendsingle/vextendsingle1−3pε\n4/vextendsingle/vextendsingle/vextendsingle/vextendsingle.\n(206)\nOne can see that this sum equals to one independently on the va lue\nof parameter |ε| ≤1 only for values |p| ≤1/3.Forp= 1, the maxi-\nmum value of the sum equals 2 = (1 + 3 ε)/2 (ε= 1). This value can\ncharacterize the degree of entanglement of Werner state.\nWe have introduced positive nonlinear map of density matrix which\nis purification map. The purification map can be combined with con-\ntractive maps discussed.Thetomographic-probability dis tributions un-\nder discussion can be completely described by their charact eristic func-\ntions. This means that the relation of tomogram property to e ntan-\nglement can be formulated in terms of the properties of chara cteristic\nfunctions.\nOne can also check the criterion using example of two-qutrit e pure\nentangled state with wave function\n|ψ/angbracketright=1√\n31/summationdisplay\nm=−1|um/angbracketright |vm/angbracketright. (207)\nThe sum defining the criterion of separability for specific U(9) trans-\nformg(n)\n0which is diagonalizing the hermitian matrix Lε|ψ/angbracketright/angbracketleftψ|\nreads\nF(ε,g(n)\n0) = 8/vextendsingle/vextendsingle/vextendsingle/vextendsingle1+ε\n9/vextendsingle/vextendsingle/vextendsingle/vextendsingle+/vextendsingle/vextendsingle/vextendsingle/vextendsingle1−8ε\n9/vextendsingle/vextendsingle/vextendsingle/vextendsingle. (208)\nFor 1/2> ε >1/8, this sum is larger than one, that means that the\nstate is entangled. For ε= 1/2, the function has maximum and it is\nequal to 5/3.\nThe entanglement of the considered state can be detected usi ng\npartial transposition criterion too.\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.44The geometry of density states, positive maps and tomograms 45\nFor the case of pure entangled state of two-qutrite system wi th the\nwave function\n|Φ/angbracketright=1√\n2/parenleftig\n|u1/angbracketright |v1/angbracketright+|u0/angbracketright |v0/angbracketright/parenrightig\n, (209)\ninwhichthestateswithspinprojections m=−1donotparticipate, the\npartial transpose criterion does also detect entanglement . Our criterion\nyields forspecific U(9) transform g(n)\n0, which diagonalizes thehermitian\nmatrixLε|Φ/angbracketright/angbracketleftΦ|the following expression for the function F(ε,g(n)\n0),\nwhich reads\nF(ε,g(n)\n0) = 5|1+ε|\n6+|1−5ε|\n6. (210)\nThe function takes maximum value for ε= 1/2 that equals to 3/2.\nThis value is smaller than 5/3 of the previous case. It corres ponds\nto our intuition that the superposition of three product sta tes of two\nqutrite system is more entangled than the superposition of o nly two\nsuch product states.\nThe criterion can be extended to multipartite spin system.\nWe have to apply for n-partite system the transform of the density\nmatrixρof the form\nL/vector ε=L(1)\nε1⊗L(2)\nε2⊗...⊗L(n)\nεn, (211)\nwhere the transform L(k)\nεkacts as depolarizing map on the kth subsys-\ntem. If the state is separable\nρ=/summationdisplay\nkpkρ(1)\nk⊗ρ(2)\nk⊗...⊗ρ(n)\nk,/summationdisplay\nkpk= 1, pk≥0,(212)\neach of the terms ρ(j)\nk(j= 1,2,...,n) in the tensor product is replaced\nby the term\nρ(j)\nk→ −εjρ(j)\nk+1+εj\nNj1j. (213)\nThis means that the transformed density matrix reads\nρ→L/vector ερ=/summationdisplay\nkpk\nn/productdisplay\nj=1⊗/parenleftigg\n−ερ(j)\nk+1+εj\nNj1j/parenrightigg\n.(214)\nThe unitary spin tomogram of the transformed density matrix takes\nthe form (/vector ε=ε1,ε2,...,εn)\nw/vector ε(m1,m2,...,m n,g(N)) =/summationdisplay\nkpkw(k)\npr(m1,m2,...,m n,g(N),/vector ε),\n(215)\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.4546 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\nwhereN=/producttextn\ns=1(2js+ 1) and element g(N)is the unitary matrix in\nN-dimensional space. The tomogram w(k)\npr(m1,m2,...,m n,g(N),/vector ε) is\nthe joint probability distribution of spin projections ms=−js,−js+\n1,...,js, which dependson theunitarytransform g(N)in thestate with\ndensity matrix\nρk=n/productdisplay\ns=1⊗/parenleftbigg\n−εsρ(s)\nk+1+εs\nNs1s/parenrightbigg\n. (216)\nFor the elements\ng(N)\npr=n/productdisplay\ns=1⊗us(2js+1),\nwhereus(2js+1) is unitary matrix, the tomogram (215) takes the form\nof sum of the products\nw/vector ε(m1,m2,...,m n,g(N)\npr) =/summationdisplay\nkpkn/productdisplay\ns=1wk/parenleftig\nms,us(2js+1),εs/parenrightig\n,(217)\nwithwk(ms,us(2js+ 1),εs) being the unitary spin tomograms of the\nsthspinsubsystem with transformeddensity matrix Lεsρ(s)\nk. If oneuses\nas the matrix us(2js+ 1), the matrix of unitary irreducible represen-\ntation of the SU(2) group, the tomogram wkdepends only on the two\nparameters defining the point on the sphere S2.\nFor a separable state of the multipartite system, one has\n/summationdisplay\nm1,...,mn/vextendsingle/vextendsingle/vextendsinglew/vector ε(m1,m2,...,m n,g(N))/vextendsingle/vextendsingle/vextendsingle= 1 (218)\nfor all elements g(N)and all parameters /vector ε.\nFor entangled state, there can be some values of parameters /vector εand\ngroup elements g(N)for which the sum is larger than one.\n12. Conclusions\nWe summarize the results of the paper.\nThenotionofentangledstates (firstdiscussedbySchr¨ odin ger[4, 51])\nhas attracted a lot of efforts to find a criterion and quantitati ve charac-\nteristics of entanglement. A criterion based on partial tra nspose trans-\nform of subsystem density matrix (complex conjugation of th e subsys-\ntem density matrix or its time reverse) provides the necessa ry and\nsufficient condition of separability of the system of two qubi ts and\nqubit-qutritt system [52]. The phase-space representatio n of the quan-\ntum states and time reverse transform (change of the signs of the\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.46The geometry of density states, positive maps and tomograms 47\nsubsystem momenta) of the Wigner function in the case of Gaus sian\nstate was applied to study the separability and entanglemen t of pho-\nton states in [13]. Recently it was pointed out that the tomog raphic\napproach of reconstructing the Wigner function of quantum s tate [43–\n45] can be developed to consider the positive probability di stribution\n(tomogram) as an alternative to density matrix (or wave func tion)\nbecause the complete set of tomograms contains the complete informa-\ntion on the quantum state [42]. This representation (called probability\nrepresentation) was constructed also for spin states inclu ding a bi-\npartite system of two spins. Up to now the problem of entangle ment\nwas not discussed in the tomographic representation. Some r emarks\non tomograms and entanglement of photon states in the proces s of\nRaman scattering were done in [53]. The tomographic approac h has\nthe advantage of dealing with positive probabilities and on e deals with\nstandard probability distributions which are positive and normalized.\nWe studied the properties of separable and entangled state o f mul-\ntipartite system using the tomographic probability distri butions. The\npositive and completely positive maps of density matrices [ 39, 54] in-\nduce specific properties of the tomograms. The properties of the posi-\ntive maps were studied in [55]. We formulated necessary and s ufficient\nconditions of separability and entanglement of multiparti te systems in\nterms of properties of the quantum tomogram. Since the tomog rams\nwere shown [36] to be related to the star-product quantizati on proce-\ndure [56], we discuss entanglement and separability proper ties in terms\nof generic operator symbols. The tomographic symbols of gen eric spin\noperators were studied in [36]. Then we focused on propertie s of entan-\nglement and separability of a bipartite system using spin to mograms\n(SU(2)-tomograms) and tomograms of the U(N)-group.\nThe idea of the approach suggested can be summarized as follo wing.\nThe positive but not completely positive linear maps of a sub system\ndensity matrix do preserve the positivity of separable dens ity matrix\nof the composite system. These maps contain also maps which d o not\npreservethepositivity oftheinitial densitymatrixof ane ntangled state\nfor the composite system. It means that the set of all linear p ositive\nmaps of the subsystem density matrix (this set is semigroup) creates\nfrom the initial entangled positive density matrix of compo site system\na set of hermitian matrices including the matrices with nega tive eigen-\nvalues. To detect the entanglement we use the tomographic sy mbols\nof the obtained hermitian matrices.The tomographic symbol s of state\ndensity matrices (state tomograms)are standard probabili ties.In view\nof this the tomographic symbols of the obtained hermitian ma trices\ncorresponding to initial separable state preserve all the p roperties of\nthe probability representation including positivity and n ormalization.\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.4748 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\nBut in case of entangled state the tomographic symbols of the obtained\nhermitian matrices can take negative values.The different be haviour\nof tomograms of separable and entangled states of composite systems\nunderaction ofthesemigroupofpositivemapsofthesubsyst emdensity\nmatrix provides the tomographic criterion of the separabil ity.\nTo conclude, we point out the main result of the work.\nWefoundthecriterionofseparabilitywhichisgivenbyequa tion(173).\nThe criterion is valid for multiparticle spin system. The cr iterion can\nbe called “tomographic criterion” of separability. The tom ographic cri-\nterion can be considered also for symplectic tomograms of mu ltimode\nphoton states. The condition of separability is sufficient be cause there\nalways exists a unitary group element by means of which any he rmitian\nmatrix can be diagonalized.It means that tomographic symbo l of non-\npositivehermitianmatrixhasnonpositivevaluesforsomeu nitarygroup\nparameters. The suggested criterion is connected with prop erties of\nthe constructed function (173) which for given density matr ix depends\non unitary group parameters gand the parameters of positive map\nsemigroupL. For separable density matrix the dependence on unitary\ngroup parameters and the semigroup parameters disappears a nd the\nfunction becomes constant equal to unity. For entangled sta tes the\nfunction differs from unity and depends on both group and semig roup\nparameters. The suggested criterion can be considered as so me com-\nplementary test of separability together with other criter ia available in\nthe literature (see, for example, [38, 52]). We point out tha t suggested\ncriterion differs from available usual ones by the kind of the n ecessary\nnumerical calculations. To apply this criterion one needs t o calculate\nthe sum of moduli of diagonal matrix elements of product of th ree\nmatrices. One of the matrices is hermitian and two others are unitary\nones. This proceduredoes not need the calculation of the eig envalues of\na matrix. The structure of positive (including not complete ly positive)\nmap semigroup with elements Lneeds extra investigation (see, for\nexample, [57]). We found also a test of entanglement based on the\nproperty of unitary spin tomogram.\nThe discussed purification map can be applied to find new quant um\nevolution equations in addition to known ones [58–61]. The a pplication\nof different forms of positive maps [55, 62] and supermatrix re presen-\ntation of the maps [63, 64] are useful for better understandi ng of the\ncomputations. Entanglement phenomena can be considered us ing sym-\nbols of density matrix of different kinds, e.g., particular qu asidistribu-\ntions[65] aswell astomographicsymbols[36]. Thedifference ofsymbols\nof entangled and separable density operators for different sc hemes of\nthe star-product quantization needs further investigatio ns as well as\ntest of entanglement of some generalizations of Werner stat e [66, 40]\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.48The geometry of density states, positive maps and tomograms 49\nin multipartite case. A relation of tomographic approach to different\npositive maps [67] should be investigated. The tomographic symbols\nare analytic in group parameters. This can be used to find extr ema of\ntomograms which give information on degree of entanglement .\nAcknowledgments\nV. I. M. and E. C. G. S. thank Dipartimento di Scienze Fisiche, Uni-\nversit´ a “Federico II” di Napoli and Istitito Nazionale di F isica Nu-\ncleare, Sezione di Napoli for kind hospitality. V. I. M. is gr ateful to\nthe Russian Foundation for Basic Research for partial suppo rt under\nProject No. 01-02-17745.\nReferences\n[1] P.A.M. Dirac 1958“The Principles ofQuantumMechanics” (Oxford:Pergamon)\n[2] L. D. Landau 1927 Z. Phys 45430\n[3] J. von Neumann 1932 “Mathematische Grudlagen der Quante nmechanik”\n(Berlin: Springer); Nov. 1927 G¨ ottingenische Nachrichten 11S245\n[4] E. Schr¨ odinger 1935 Naturwissenschaften 23807; 823; 844\n[5] E. Schr¨ odinger 1926 Ann. d. Phys. Lpz79489\n[6] V. I. Man’ko, G. Marmo, E. C. G. Sudarshan and F. Zaccaria 1 999J. Russ.\nLaser Res. 20421; 2002 J. Phys. A: Math. Gen. 357173\n[7] M. Horodecki, P. Horodecki and R. Horodecki 1996 Phys. Lett. A2231\n[8] S. Hill and W. K. Wootters 1997 Phys. Rev. Lett. 785022\nW. K. Wootters 1998 Phys. Rev. Lett. 802245\n[9] K. Zyczkowski, P. Horodecki, A. Sanpera and M. Lewenstei n 1998Phys. Rev. A\n58883\n[10] S. Popescu and D. Rohrlich 1997 Phys. Rev. A56R3319\n[11] S. Abe and A. K. Rajagopal 2002 PhysicaA289157\n[12] C. H. Bennett, D. P. Di Vincenzo, J. A. Smolin and W. L. Woo tters 1996 Phys.\nRev.A543824\n[13] R. Simon 2002 Phys. Rev. Lett. 842726\n[14] S. M. Barnett and S. J. D. Phoenix 1989 Phys. Rev. A402404; 1991 ibid 44\n535\n[15] A. Mann, B. C. Sanders and W. J. Munro 1995 Phys. Rev. A51989\n[16] C. H. Bennet, H. J. Bernstein, S. Popescu and B. Schumach er 1996Phys. Rev.\nA532046\n[17] V. Vedral and M. B. Plenio 1998 Phys. Rev. A571619\n[18] B.-G. Englert, M. L¨ offler, O. Benson, B. Varcoe, M. Weidi nger and H. Walter\n1998Fortschr. Phys. 46897\n[19] M. Horodecki, P. Horodecki and R. Horodecki 2000 Phys. Rev. Lett. 842014\n[20] S. Parker, S. Bose and M. B. Plenio 2000 Phys. Rev. A61032305\n[21] C. H. Bennet, S. Popescu, D. Rohrlich, J. A. Smolin and A. V. Thapliyal 2001\nPhys. Rev. A63012307\n[22] K. Pi¸ atek and W. Leo´ nski 2001 J. Phys. A: Math. Gen. 344951\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.4950 V.I. Man’ko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria\n[23] K. Audenaert, J. Esert, E. Jan´ e, M. B. Plenio, S. Virman i and R. B. De Moor\n2001Phys. Rev. Lett. 87217902\n[24] K. Furuya, M. C. Nemes and G. Q. Pellegrino 1998 Phys. Rev. Lett. 805524\n[25] K.ˇZyczkowski, P. Horofecki, A. Savpera and M. Lewenstein 1998 Phys. Rev. A\n58883\n[26] W. J. Munro, D. F. V. James, A. G. White and P. G. Kwait 2001 Phys. Rev.\nA64030302\n[27] G. Vidal and R. F. Werner 2002 Phys. Rev. A65032314\n[28] F. Coffman, J. Kundu and W. K. Wootters 2000 Phys. Rev. A61052306\n[29] P. Badziag, P. Deuar, M. Horodecki, P. Horodecki and R. H orodecki 2002 J.\nMod. Opt. 491289\n[30] M. A. Andreata, A. V. Dodonov and V. V. Dodonov 2002 J. Russ. Laser Res.\n23531\n[31] V. V. Dodonov and V. I. Man’ko 1997 Phys. Lett. A229335\n[32] Olga Man’ko and V. I. Man’ko 1997 JETP85430\n[33] A. B. Klimov, O. V. Man’ko, V. I. Man’ko, Yu. F. Smirnov an d V. N. Tolstoy\n2002J. Phys. A: Math. Gen. 356101\n[34] V. A. Andreev and V. I. Man’ko 1998 JETP87239\n[35] V. I. Man’ko and S. S. Safonov 1998 Yad. Fiz. 61658\n[36] O. V. Man’ko, V. I. Man’ko and G. Marmo 2000 Phys Scr. 62446; 2002 J.\nPhys. A: Math. Gen. 35699\n[37] V. V. Dodonov, A. S. M. De Castro and S. S. Misrahi 2002 Phys. Lett. A296\n73\nA. S. M. De Castro and V. V. Dodonov 2003 J. Russ. Laser Res. 2393; 2003 J.\nOpt. B: Quantum Semiclass. Opt. 5S593\n[38] Special Issue on Entanglement 2002 J. Math. Phys. 43No. 9\n[39] E. C. G. Sudarshan, P. M. Mathews and J. Rau 1961 Phys. Rev. 121920\n[40] E. C. G. Sudarshan and A. Shaji 2003 “Structure and param etrization of\nstochastic maps of density matrix” ArXiv quant-ph/0205051 v2; 2003 J. Phys.\nA: Math. Gen. 36(in press)\n[41] A. Peres 1996 Phys. Rev. Lett. 771413\n[42] S. Mancini, V. I. Man’ko and P. Tombesi 1996 Phys. Lett. A2131; 1997Found.\nPhys.27801\n[43] J. Bertrand and P. Bertrand 1987 Found. Phys. 17397\n[44] K. Vogel and H. Risken 1989 Phys. Rev. A402847\n[45] S. Mancini, V. I. Man’ko and P. Tombesi 1995 Quantum Semiclass. Opt. 7615\nG. M. D’Ariano, S. Mancini, V. I. Man’ko and P. Tombesi 1996 Quantum\nSemiclass. Opt. 81017\n[46] M. A. Man’ko, V. I. Man’ko and R. V. Mendes 2001 J. Phys. A: Math. Gen.\n248321\n[47] V. I. Man’ko, G. Marmo, E. C. G. Sudarshan and F. Zaccaria 2003 “Entangle-\nment in probability representation of quantum states and to mographic criterion\nof separability” J. Opt. B: Quantum Semiclass. Opt. (in press); 2003 J. Russ.\nLaser Res. 24507\n[48] V. I. Man’ko, G. Marmo, E. C. G. Sudarshan and F. Zaccaria 2000 “Inner\ncomposition law of pure-spin states” in “Spin-Statistics C onnection and Com-\nmutation Relations” R. C. Hilborn and G. M. Tino (eds.) AIP Conference\nProceedings 54592\n[49] A. S. Holevo 1999 Russ. Math. Surveys 531295\n[50] P. W. Shor 2002 J. Math. Phys. 434334\n[51] E. Schr¨ odinger 1935 Proc. Cambridge Philos. Soc. 31555\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.50The geometry of density states, positive maps and tomograms 51\n[52] P. B. Slater 2003 J. Opt. B: Quantum Semiclass. Opt. 5S691\n[53] S. V. Kuznetsov, O. V. Man’ko and N. V. Tcherniega 2003 J. Opt. B: Quantum\nSemiclass. Opt. 6S503\n[54] W. F. Stinespring 1955 Proc. Amer. Math. Soc. 6211\n[55] L. C. Woronowicz 1976 Rep. Math. Phys. 10165\n[56] F. Bayen, M. Flato, C. Fronsdal, A. Lichnerovicz and D. S ternheimer 1975 Lett.\nMath. Phys. 1521\n[57] C. King 2002 “The capacity of the quantumdepolarizing c hannel” ArXivquant-\nph/0204172 v2\n[58] J. E. Moyal 1949 Proc. Cambridge Philos. Soc. 4599\n[59] A. Kossakovski 1972 Rep. Math. Phys. 3247\n[60] G. Lindblad 1976 Comm. Math. Phys. 48119\n[61] V. Gorini, A. Kossakovski and E. C. G. Sudarshan 1978 Rep. Math. Phys. 18\n149\n[62] K. Kraus 1973 Ann. Phys. NY64311\n[63] M. D. Choi 1975 Lin. Alg. Appl. 10285; 1970 Canadian J. Math. 24520; 1976\nIllinois J. Math. 48119\n[64] T. F. Havel 2003 J. Math. Phys. 44534\n[65] E. C. G. Sudarshan 1963 Phys. Rev. Lett. 10277\nC. L. Mehta and E. C. G. Sudarshan 1965 Phys. Rev. B138274\n[66] R. Werner 1989 Phys. Rev. A404277\n[67] A. Kossakowski 2003 “A class of linear positive maps in m atrix algebras” ArXiv\nquant-ph/0307132 v1\nMMSZ-Bregenz.tex; 10/11/2018; 12:33; p.51MMSZ-Bregenz.tex; 10/11/2018; 12:33; p.52" }, { "title": "0706.3116v1.Series_expansion_for_the_density_of_states_of_the_Ising_and_Potts_models.pdf", "content": "SERIES EXPANSION FOR THE DENSITY OF STATES OF THE\nISING AND POTTS MODELS\nDANIEL ANDR \u0013EN\n1.Introduction\nThe Lentz-Ising model of ferromagnetism has been thoroughly studied since its\nconception in the 1920's[Len20]. It was solved in the 1-dimensional case by Ising\nhimself in 1925[Isi25] and in the 2-dimensional case without an external \feld by\nOnsager in 1944[Ons44]. For an introduction to the model, see Cipra [Cip87].\nThe partition function on a graph Gonnvertices and medges is de\fned as\n(1) Z(G;x;y) =X\ni;jaijxiyj:\nHereaijcounts the number of induced subgraphs of Gwith (n\u0000j)=2 vertices and\n(m\u0000i)=2 edges in the boundary. We refer to the index ias the energy and the\nindexjas the magnetization.\nThe traditional partition function studied in statistical physics is then obtained\nby evaluating it at a certain point\n(2) Z\u0000\nG;eK;eH\u0001\nwhereK=\u0000J=kBT,H=\u0000h=kBTandJandhare parameters describing the\ninteraction through edges and with an external magnetic \feld respectively, Tis the\ntemperature and kBthe Boltzmann constant.\nThe main goal in the study of the Ising model on a graph G, or some family\nof graphs, is usually to study the model in the vicinity of a critical temperature ,\ndenotedTc, where the model undergoes a phase transition and there determine the\nbehaviour of various critical properties.\n2.Definitions and notation\nLetG= (V;E) be a graph with vertex set VwithjVj=nvertices and edge set\nEwithjEj=medges. Let a state \u001bbe a function from the vertices Vto the set\nf\u00061gand let \n be the set of all states. We can then de\fne the energy of the graph\nG= (V;E) in state\u001bto be\n(3) E(G;\u001b) =X\nuv2E\u001b(u)\u001b(v)\nthemagnetization to be\n(4) M(G;\u001b) =X\nv2V\u001b(v)\nand formulate a generating function that counts them all as\n(5) Z(G;x;y) =X\n\u001b2\nxE(G;\u001b)yM(G;\u001b)=X\nijaijxiyj\nwhere the last equality de\fnes the coe\u000ecients aij=aij(G). We often drop Gwhen\nwe can deduce the graph from the context.\nWe will also need these de\fnitions later:\n1arXiv:0706.3116v1 [cond-mat.str-el] 21 Jun 20072 DANIEL ANDR \u0013EN\nDe\fnition 1 (T-join) .A T-join (T;A) in a graph G= (V;E) is a subset T\u0012V\nof vertices and a subset A\u0012Eof edges such that each vertex in Tis incident with\nan odd number of edges in Aand each vertex in VnTis incident with an even\nnumber of edges from A.\nObserve that the cardinality of Thas to be even since we can not have a subgraph\nwith an odd number of vertices of odd degree.\nDe\fnition 2 (Cut) .A cut [S;\u0016S] in a graph G= (V;E) is a subset of edges, induced\nby a partition S[\u0016S=V, that have one endpoint in Sand the other in \u0016S. Let\nj[S;\u0016S]jbe the number of edges in the cut.\nDe\fnition 3 (Locally vertex transitive) .We say that a sequence of graphs fGig1\ni=1\nare locally vertex transitive if for each Rthere exists an Nsuch that all balls of\nradiusR, in the graph metric, around each vertex in each graph in the subsequence\nfGig1\ni=Nare isomorphic.\n3.Series expansion\nWe can try to make our problem simpler by setting y= 1 and get\n(6) Z(x;1) =Z(x) =X\nijaijxi1j=X\niaixi\nwhere once again the last equality is the de\fnition of the aicoe\u000ecients. What do\nthese coe\u000ecients count? Since the state \u001bpartitions the vertex set Vin two parts\nand theE(G;\u001b) counts edges with one vertex in one part and the other vertex in\nthe other part as negative and the other edges as positive we get that aiis twice\nthe number of cuts of sizem\u0000i\n2(we count each cut twice since we can interchange\nthe partitions). This has a natural reformulation using even subgraphs, namely:\nTheorem 4 (van der Waarden) .Letaibe the number of cuts of sizem\u0000i\n2and let\nbibe the number of even subgraphs with iedges, then:\n(7)X\niaieiK= 2 coshmKX\nibitanhiK\nProof. The \frst sum in (7) is the moment generating function for the sequence ai.\nThekthmoment of aican be written as\n(8) \u0016k=X\niaiik=X\n\u001b2\n X\nuv2E\u001b(u)\u001b(v)!k\nWe now expand the multinomial\u0000P\nuv2E\u001b(u)\u001b(v)\u0001kwhere each term can be seen\nas an choice of kout ofmedges (not necessarily distinct). Now observe that if we\nhave chosen an even number of edges incident with a vertex, vsay, we will have an\neven number of \u001b(v)'s in the product so they contribute +1. If we have chosen an\nodd number of vertices we can \fnd a smallest (in some arbitrarily order) such odd\nvertexvand we see that if we change the state \u001bto the state \u001b0with\u001b(v) =\u0000\u001b0(v)\nand all other values equal, we will get a bijection between states witch contribute\n+1 and\u00001 and with at least one vertex of odd degree. Our conclusion is that\nwe only count the choices where we have an even degree at each vertex. We will\nhowever count subgraphs where we have the opportunity to choose each edge a\nmultiple number of times. If we reduce the multiple edges modulo 2 we get a\nsimple subgraph of even degree. The \\surviving\" edges are the ones that where\nchosen an odd number of times so an even number of those \\odd\" edges have to be\nincident at each vertex.SERIES EXPANSION FOR THE DENSITY OF STATES OF THE ISING AND POTTS MODELS 3\nIf we now change our view and instead of adding up the kmoments, change the\norder of summation, and add up along the index of the number of \\surviving\" odd\nedgesiwe get a simple connection between the simple subgraphs and subgraphs\nwith multiple edges. We can construct an even multiedge subgraph by \frst select\na simple subgraph with even degree at each vertex and then multiply each edge an\nodd number of times and then select a number of edges not in the even subgraph\nto multiply an even number of times. So if we \frst choose an even subgraph with i\nedges and multiply each edge an odd number of times we get the generating func-\ntionbisinhiK, and then choose a number of edges outside the even subgraph and\nmultiply these an even number of times, we get the generating function coshm\u0000iK,\nand we end up with an multiedge subgraph with an even number of edges incident\nto each vertex. If we now sum over all iwe get\n(9)X\nibisinhjKcoshm\u0000iK= coshmKX\nibitanhiK\nand we see each of these graphs twice and therefor this gives (7). \u0003\nTo formulate the full two variable connection we need T-joins instead of even\ndegree subgraphs and also consider the size of the sets in the vertex partition\ninduced by the state. Also note that in this theorem we have slightly changed the\nmeaning of aijandbijto make the proof using standard graph theoretic notation.\nThe following theorem can be found in e.g. [Big77]:\nTheorem 5. LetG=G(V;E)be a graph, aijthe number of cuts [S;\u0016S]with\nj[S;\u0016S]j=iandjSj=j. Letbijbe the number of T-joins (T;A)withjAj=iand\njTj=j. Then\nX\nijbijxiyj= 2\u0000jVjX\nijaij(1\u0000x)i(1 +x)jEj\u0000i(1\u0000y)j(1 +y)jVj\u0000j\nProof. Fix a subset T\u0012Vof vertices and a subset A\u0012Eof edges from the graph\nG=G(V;E). LetS\u0012Vbe another subset of vertices and [ S;\u0016S] be the cut de\fned\nby the edges from Sto\u0016S=VnS. Let the weight of the vertices in T\\Sbe\u0000y,\nthe weight of the vertices in T\\\u0016Sbey, the weight of the edges from Athat lies in\nthe cut [S;\u0016S] be\u0000xand the rest of the edges from Ahave weight x. Let the total\nweight of (T;A) with respect to the cut [ S;\u0016S] be the product of the weights of the\nedges and vertices in ( T;A). We say that the weight is positive if the coe\u000ecient in\nfront ofxjAjyjTjis positive and negative otherwise. By magnitude we denote the\nweight without the sign.\n(T;A) can fail to be a T-join in basically three ways. First the cardinality of\nTcan be odd, secondly there can exist a smallest vertex (in an arbitrary order of\nthe vertices) vthat is incident with an odd number of edges from Aand does not\nbelong toT, and \fnally there can exist a smallest vertex vthat is incident with an\neven number of edges from Aand belongs to T.\nIn the \frst case we have two cuts [ S;\u0016S] and [ \u0016S;S] in which the magnitude of the\nweight will be the same but with opposite sign.\nIn the two latter cases we have a bijection between cuts with v2Sandv =2S\n(we simply move the vertex vbetweenSand \u0016S) that once again give the same\nmagnitude and di\u000berent signs of the weight. If we sum over all cuts the total\ncontribution of such a choice of ( T;A) will cancel.\nIf (T;A) indeed is a T-join the weight will always be positive since we either have\nan even number of vertices in T\\Sand an even number of edges crossing the cut\nor an odd number of vertices in T\\Sand an odd number of vertices crossing the\ncut. All in all we end up with an even number of minus signs and thus a positive\nweight.4 DANIEL ANDR \u0013EN\nIf we now sum over all choices ( T;A) andSwe will count each T-join 2jVjtimes.\nIf we rearrange our summation (i.e. we \frst choose a cut and then go through all\nchoices ofTandA) we get the theorem. \u0003\nTo get Theorem 4 we have to shift the indices and substitute xfore\u00002Kandy\nfor 0 and \fnally double all values since we count all cuts twice in (7).\n(10) 2emKX\ni;jaije\u00002iK0j= [since 00= 1] =X\ni2aie(m\u00002i)K=\n2emK2\u0000mX\ni;jbij(1\u0000e\u00002K)i(1 +e\u00002K)m\u0000i1n=\n2\u0012eK+e\u0000K\n2\u0013mX\nibi\u00121\u0000e\u00002K\n1 +e\u00002K\u0013i\n=\n2 coshmKX\nibitanhiK\nwherebidenotes the number of T-joins with iedges. Now, since aicounts the\nnumber of cuts with iedges in Theorem 5, 2 aie(m\u00002i)Kcorresponds to the aieiKin\nTheorem 4 via reindexing and the fact that each cut is counted twice in Theorem\n4.\nSince we have a symmetry between T-joins and cuts we have the following corol-\nlary:\nCorollary 6. With the same notation as in theorem 5 we have\nX\ni;jaijxiyj= 2\u0000jEjX\ni;jbij(1\u0000x)i(1 +x)jVj\u0000i(1\u0000y)j(1 +y)jEj\u0000j\nProof. If we choose a T-join instead of a cut the weight of ( T;A) will always be\npositive if and only if ( T;A) is a cut. In other cases the contributions once again\ncancel out. \u0003\n3.1.The thermodynamic limit. In Physics we are interested in the so called\nthermodynamic limit of a sequence of graphs fGjg1\nj=1. This is de\fned as\n(11) f(x) = lim\nj!11\njV(Gj)jlogZ(Gj;x;1)\nwhen it exists. An example of such a family is fCn\u0002Cng1\nn=3. If the graph family\nis locally vertex transitive its easy to see that the number of connected T-joins\nof \fxed size will grow proportionally to jV(G)j. From that follows that the total\nnumber of T-joins of a \fxed size will grow as a polynomial with degree equal to the\nmaximal number of connected components in the T-joins of that size and thus will\nthe thermodynamic limit exist.\nIf we change notation so that our index set instead is the number of vertices n\nin our graph sequence and use bj(n) to denote the number of T-joins with jedges\nwe see that the thermodynamic limit is\n(12) lim\nn!11\nnlog 2 coshmxX\njbj(n) tanhjx= log 2 +dcoshx+X\njbjtanhjx\nwhich de\fnes a new set of bj:s that happens to be rational numbers and d=m=n,\nthe average degree.SERIES EXPANSION FOR THE DENSITY OF STATES OF THE ISING AND POTTS MODELS 5\n3.2.Taylor series. It can be of interest to plot these functions, or at least some\napproximation of them. Since we have a phase transition in all interesting cases its\nhard to \fnd one function that works for the entire interval. In this section we will\ninstead develop two Taylor approximations, one for the high-temperature case and\none for the low-temperature case. The \frst one is simple, take\n(13)u(K) =X\niaiKi= log 2 +dlog coshK+ lim\nn!11\nnlogX\nibi(n) tanhiK\nand Taylor expand around K= 0. If you want the function K(u) you can invert\nu(K) as a Taylor series. Its often better to plot a Pad\u0013 e-approximation of u(K).\nThe second one needs a little more work:\n(14)d\ndKu(K) =d\ndK\"\nlog 2 +dlog coshK+ lim\nn!11\nnlogX\nibi(n) tanhiK#\n=\n\u0014x= tanhK\nd\ndK= (1\u0000x2)d\ndx\u0015\n=dx+ (1\u0000x2)d\ndxlim\nn!11\nnlogX\nibi(n)xi=\ndx+ (1\u0000x2)d\ndxlim\nn!11\nnlog1\n2nX\niam\u00002i(n)(1\u0000x)i(1 +x)m\u0000i=\ndx+ (1\u0000x2)d\ndxlim\nn!1\"\n\u0000log 2 +dlog(1 +x) +1\nnlog exp \nnX\niai\u00121\u0000x\n1 +x\u0013i!#\n=\ndx+d(1\u0000x2)\n1 +x+ (1\u0000x2)d\ndxX\nai\u00121\u0000x\n1 +x\u0013i\n=d+ (1\u0000x2)d\ndxX\nai\u00121\u0000x\n1 +x\u0013i\n=\nd+2(x\u00001)\n(x+ 1)X\nai\u00121\u0000x\n1 +x\u0013i\u00001\n=d\u00002X\nai\u00121\u0000x\n1 +x\u0013i\n=\nd\u00002X\nai\u00121\u0000tanhK\n1 + tanhK\u0013i\nBy not doing the last substitution x= tanhKits easy to invert this function and\nplot it in the low temperature (high energy) portion of the scale.\n4.Two other series expansions\nFrom the basic thermodynamic limit one can construct two other sequences that\nare of a more combinatorial \ravour. These are sequences with integer coe\u000ecients.\nThis is because they, instead of weighting the counts, make an explicit order in\nwhich you have to choose things and thus avoid dividing with large factorials. The\ncorrespondence with the thermodynamic limit series are:\n(15)X\ni\u00150bi(n)xi= exp0\n@nX\ni\u00151bixi1\nA=Y\ni\u00151\u00121\n1\u0000xi\u0013n\fi\n=Y\ni\u00151\u00121\n1\u0000b\fixi\u0013n\nThe coe\u000ecients are connected trough these simple triangular linear equations:\nibi=X\ntjit\ft (16)\nibi=X\nrs=irb\fs\nr (17)\ni\fi=X\nrst=i\u0016(r)tb\fs\nt (18)6 DANIEL ANDR \u0013EN\n4.1.What they are counting. The new series can be seen as placing connected\nsubgraphs at each vertex in some order. First we place all subgraphs (including the\nempty one) rooted at vertex one, then the subgraphs rooted at vertex two and so\non in such a way that we never form a new connected component. In this way we\navoid symmetries and we get a integer sequence. We can still calculate, in principle,\nthe di\u000berent \fiandb\fifor locally vertex transitive graphs.\n5.Plots\nWe shall now compare these Taylor expansions of the series with some sampled\ndata to compare how well behaved the series are around the critical energy. As we\nshall see, the series are rather far from what can be expected to be the truth.\n0.2 0.4 0.6 0.8 10.10.20.30.40.5\nFigure 1. K(u) for the three-dimensional simple cubic lattice to-\ngether with sampled data for the cubes of linear order 32, 64 and\n128.\n5.1.Simple cubic lattice. We start of with the simple cubic lattice in three,\nfour and \fve dimensions. As can be expected, the longest series expansion is for\nthe three-dimensional simple cubic lattice. The four- and \fve-dimensional simple\ncubic lattices have much shorter high- and low-temperature expansions. Figure\n1 to 3 shows K(u) curves for the three lattices respectively. The pictures are of\ndiagonal Pad\u0013 e-approximants of the Taylor expansions. The Pad\u0013 e-approximant for\nthe function of the three-dimensional simple cubic lattice is fairly accurate to about\nK\u0019:22. For the four- and \fve-dimensional lattices the accuracy is a lot lower as\ncan be seen from the pictures.\n6.Guessing the radius of convergence\nThe interesting part is of course to try to \fnd the radius of convergence for\nthe series of the function u(K) since that would give the critical temperature Kc.\nA fairly good guess for Kcfor the simple cubic lattice is that it is approximately\n0:2216546 (see e.g. [HRL+]). A simple observation is that the radius of convergence\nmust be smaller than any 1 =b\fi1=isince otherwise the in\fnite product (15) will notSERIES EXPANSION FOR THE DENSITY OF STATES OF THE ISING AND POTTS MODELS 7\n0.2 0.4 0.6 0.8 10.10.20.30.40.5\nFigure 2. K(u) for the four-dimensional simple cubic lattice to-\ngether with sampled data for the cube of linear order 4, 6, 8, 12,\n16, 32, 48 and 64.\n0.2 0.4 0.6 0.8 10.10.20.30.40.5\nFigure 3. K(u) for the \fve-dimensional simple cubic lattice to-\ngether with sampled data for the cube of linear order 4, 6, 8, 12,\n16 and 32.\nconverge. In Table 1 the row labelled min is the minimum over all known values of\nb\fifor the simple cubic lattice. In trying to extrapolate this value we have used a\npolynomial of degree 3 and a min square approximation to the data set (1 =i;1=b\fi1=i)\nand looked at the constant term in the resulting polynomial. This gives the result in\nthe row labelled f(0). Since the coe\u000ecients is some form of combinatorial quantity8 DANIEL ANDR \u0013EN\nb\fibib\u000bi+b\u000bi\u0000a+\nia\u0000\ni\nmin 0.291989 0.291989 0.618531 0.616299 0.618531 0.616307\nf(0) 0.227727 0.227839 0.54278 0.543852 0.542846 0.539196\n\u000bii\f\r0.221418 0.221451 0.523726 0.520324 0.523747 0.542904\n\u000bi\r 0.258385 0.258429 0.578131 0.565115 0.578137 0.56453\nTable 1. Di\u000berent ways to try to guess the radius of convergence\nfor the simple cubic lattice.\nit is not unthinkable that they grow exponentially. We have thus also tried to\n\ft the data ( i;b\fi) to the functions \u000bii\f\rand\u000bi\rand calculated 1 =\u000bthat gives\nus the guessed radius of convergence. This is the last two lines of Table 1. The\nsecond column is the same thing for the coe\u000ecients of bi. When we come to the low\ntemperature part we get into some trouble since the coe\u000ecients of the sequences\naiandb\u000bihave both positive and negative signs. If we try to use the same analysis\nas for the (all positive) biandb\fi, we get very erratic numbers, so instead we split\nthe sequences in a positive and a negative part and do the analysis separately. As\nobserved by others, the low temperature series seems to have a complex root that\nis closer to the origin than the physically important real root. Fortunately does\nthe extrapolation with \u000bii\f\rfor thebiandb\fiseries give a decent idea of what the\ncritical temperature may be.\nReferences\n[Big77] Norman Biggs. Interaction models . Cambridge University Press, Cambridge, 1977. Course\ngiven at Royal Holloway College, University of London, October{December 1976, London\nMathematical Society Lecture Note Series, No. 30.\n[Cip87] Barry A. Cipra. An introduction to the Ising model. Amer. Math. Monthly , 94(10):937{\n959, 1987.\n[HRL+] Roland H aggkvist, Andres Rosengren, Per H\u0017 akan Lundow, Klas Markstr om, Daniel\nAndr\u0013 en, and Petras Kundrotas. On the Ising model for the simple cubic lattice. Manu-\nscript.\n[Isi25] Ernst Ising. Beitrag zur Theorie des Ferromagnetismus. Z.Physik , 31:253{258, 1925.\n[Len20] Wilhelm Lenz. Beitrag zum Verst andnis der magnetishen Erscheinungen in festen\nK orpern. Z. Physik , 21:613{615, 1920.\n[Ons44] Lars Onsager. Crystal statistics. I. A two-dimensional model with an order-disorder tran-\nsition. Phys. Rev. (2) , 65:117{149, 1944.SERIES EXPANSION FOR THE DENSITY OF STATES OF THE ISING AND POTTS MODELS 9\nAppendix A.Tables\nWe have collected the various series we have found in this appendix. Some of\nthese are old and thus not especially long and some are from newer calculations\nand longer.\nn \u000bn b\u000bn an\n6 1 1 1\n10 3 3 3\n12 -4 -4 -7/2\n14 15 15 15\n16 -33 -33 -33\n18 104 104 313/3\n20 -282 -285 -561/2\n22 849 849 849\n24 -2460 -2470 -9847/4\n26 7485 7485 7485\n28 -22542 -22647 -45069/2\n30 69392 69384 346966/5\n32 -213738 -214299 -427509/2\n34 666750 666750 666750\n36 -2086785 -2092121 -12520405/6\n38 6583341 6583341 6583341\n40 -20852223 -20892996 -83409453/4\n42 66425750 66424630 464980286/7\n44 -212410377 -212770353 -424819905/2\n46 682202205 682202205 682202205\n48 -2198562644 -2201602421 -17588511087/8\n50 7110521070 7110521022 35552605353/5\n52 -23065955826 -23093964696 -46131904167/2\n54 75045653088 75045278168 675410878105/9\n56 -244806881325 -245063348553 -979227570369/4\n58 800606679471 800606679471 800606679471\n60 -2624325216574 -2626724535242 -13121625909861/5\n62 8621219166681 8621219166681 8621219166681\n64-28379404026078 -28402366460136 -113517616531821/4\nTable 2. Simple cubic lattice10 DANIEL ANDR \u0013ENn \fnb\fn bn\n4 3 3 3\n6 22 22 22\n8 186 183 375/2\n10 1980 1980 1980\n12 24032 23793 24044\n14 319170 319170 319170\n16 4514664 4497993 18059031/4\n18 67003462 66999920 201010408/3\n20 1032455736 1030496478 5162283633/5\n22 16397040750 16397040750 16397040750\n24 266958785298 266673642443 266958797382\n26 4437596650548 4437596650548 4437596650548\n28 75078511535604 75027576950427 525549581866326/7\n30 1289656872697576 1289654284203514 6448284363491202/5\n32 22447149807206352 22437033558570606 179577198475709847/8\n34 395251648062268272 395251648062268272 395251648062268272\n36 7031220729573330428 7028971745154518717 21093662188820520521/3\n38 126225408651399082182 126225408651399082182 126225408651399082182\n40 2284608766597864770492 2284077801219364370517 4569217533196761997785/2\n42 41655898158709803301884 41655887320814523000436 291591287110968623857940/7\n44 764611114761740442269316 764476683289070060492337 8410722262379235048686604/11\n4614120314204713719766888210 14120314204713719766888210 14120314204713719766888210\nTable 3. Simple cubic latticeSERIES EXPANSION FOR THE DENSITY OF STATES OF THE ISING AND POTTS MODELS 11\nn a4\nnb4\nna5\nnb5\nn\n4 0 6 0 10\n6 0 76 0 180\n8 1 1371 0 5025\n10 0 30152 1 178696\n18 0 5\n20 28 -11/2\n22 -64 45\n24 127/3 -95\n26 228 50\n28 -834 5\n30 1116 1471/3\n32 5911/4 -1685\n34 -10404 1885\n36 21460 -775/2\n38 -1956 5445\n40 -595179/5 -112951/4\n42 1076092/3 49545\n44 -344316 -62795/2\n46 -1132588 71442\n48 10842287/2\n50 -9187444\n52 -5820150\n54 73867260\n56-1294335811/7\n58 95069292\n60 2609680726/3\n62 -3217644924\nTable 4. Simple cubic lattice in dimension 4 ( x4\nn) and 5 (x5\nn)12 DANIEL ANDR \u0013EN\nn an bn\n4 0 6\n6 2 44\n7 0 36\n8 0 384\n9 0 688\n10 6 4572\n11 6 11148\n12 -14 66158\n13 0 190662\n14 30 1051668\n15 60 3471452\n16 -108 17917704\n17 -144 65438160\n18 1118/3 971627006/3\n19 498 1265584728\n20 -714 30625029636/5\n21 -2366 25078631014\n22 3270 -3816568476\n23 7704\n24 -8106\n25 -27372\n26 19842\n27 114342\n28 -68892\n29 -377376\n30 643952/5\n31 1431726\n32 -137718\n33 -5365756\n34 -33330\n35 18644574\n36 -970922/3\n37 -69012330\n38 -32516754\n39 249820316\n40 162829320\n41 -869879742\n42-5660822830/7\n43 3155460756\nTable 5. Simple cubic lattice, 3-states Potts model" }, { "title": "0707.3099v3.Existence_of_a_Density_Functional_for_an_Intrinsic_State.pdf", "content": "arXiv:0707.3099v3 [nucl-th] 19 Aug 2008Existence of a Density Functional for an Intrinsic State\nB. G. Giraud\nbertrand.giraud@cea.fr, Institut de Physique Th´ eorique ,\nDSM, CE Saclay, F-91191 Gif/Yvette, France\nB. K. Jennings\njennings@triumf.ca, TRIUMF, Vancouver BC, V6T2A3, Canada\nB. R. Barrett\nbbarrett@physics.arizona.edu, Department of Physics,\nUniversity of Arizona, Tucson, AZ 85721, USA\n(Dated: November 1, 2018)\nA generalization of the Hohenberg-Kohn theorem for finite sy stems proves the existence of a\ndensity functional (DF) for a symmetry violating intrinsic state, out of which a physical state with\ngood quantum numbers can be projected.\nI. INTRODUCTION\nDensity functional [1] theory (DFT) was initially defined for ground s tates. These have good quantum numbers.\nEverynuclearphysicistknowsthat, forinstance, thegroundsta teof20Ne isa0+andthat its densityis, thus, isotropic,\nnot an ellipsoid. Every molecular physicist knows that, for instance, the ground state of the ammonia molecule is a\ngood parity state, not just the pyramid described by the Born-Op penheimer approximation. In particular, the nuclear\nDF must generate spherical solutions for the some thousand 0+nuclear ground states, whether nuclei are intrinsically\ndeformed or not. The same need for isotropic solutions extends to the non-local generalization of the DFT [2] - [3].\nBut the theory of rotational bands and/or parity vibrations, whe ther in nuclear, atomic or molecular physics, most\noften relates ground states to wave packets, often named “intr insic states”, which are symmetry breaking, namely do\nnot transform in an irreducible representation (irrep) of the symm etry group Sof the Hamiltonian. Therefore, one\nmay raise the question of DFT for intrinsic states rather than eigen states.\nGiven the physical Hamiltonian Hwith its symmetry group S,calculations providing a “non S-irrep” state as a\nsolution for a minimum energy cannot be labelled as the result of “the” DF. Such a state, labelled intrinsic, is actually\njust a convenient wave packet, to be subsequently projected on to good quantum numbers to account for physical\nlevels. Such intrinsic calculations should rather exhibit a special Hamilt onian, which might be called an intrinsic\nHamiltonian, distinct from the physical one, if such calculations are t o be legitimized. Or they should be interpreted\nas one variety of the Hartree-Fock, Hartree-Bogoliubov, etc. v ariational approaches. This is implicit or even explicit\nin calculations with an energy density functional, implying non-localities, see for instance [4] - [7]. Energy d ensity and\nparticle density are different concepts.\nIt turns out that the particle density which has been used for the f oundation of DF theory mainly concerns eigen-\nstates of the physical Hamiltonian, in principle at least, while the ener gy density, used for Skyrme force calculations\nin nuclear physics for instance, mainly provides intrinsic states. This note presents a particleDF theory for intrinsic\nstates, not for eigenstates of the Hamiltonian. We show how the ph ysical Hamiltonian can be reconciled with the\nproper definition of a DF for an intrinsic state and how the resulting in trinsic state can be accepted as a useful wave\npacket, out of which states with good quantum numbers can be pro jected. In particular, in the case of molecules,\nour approach will consider both the electrons and the nuclei. Our int rinsic state can take into account both kinds of\ndegrees of freedom. Section II describes a functional out of whic h a variational principle derives for an intrinsic state,\nand out of which a DF for the intrinsic density is obtained. Section III gives an example of variational equations to\nbe solved in practice. Section IV rewrites the formalism into a slightly s impler form. Section V contains a discussion\nof our result and suggests an ansatz for intrinsic Hamiltonians.\nII. BASIC FORMALISM\nFor a first argument, dealing with one kind of identical particles only, letHbe their physical Hamiltonian and\nφ,∝angbracketleftφ|φ∝angbracketright= 1,be a trial wave packet, most often not transforming under an irre p of the symmetry group SofH.\nFor instance, for fermions, φmay be an arbitrary Slater determinant, but we let φbe also a more general wave\nfunction, including some amount of correlations. States ψ∝Pφwith good quantum numbers can then be projected\nout ofφby a projector P,a fixed operator. In the following, we shall systematically use the pr operties,P2=Pand2\n[P,H] = 0.It may happen that ∝angbracketleftφ|P|φ∝angbracketrightvanishes, but such cases usually make a domain of zero measure in th e usual\nvariational domains, where φevolves. In any case, since His an operator bounded from below, the functional of φ,\n∝angbracketleftφ|PH|φ∝angbracketright/∝angbracketleftφ|P|φ∝angbracketright,is bounded from below. Embed now the system in an external, local fie ld,U=/summationtext\niu(ri).The\nlocal, real potential uis taken bounded from below, but is otherwise arbitrary. In particu lar, it may usually have\nnone of the symmetries of H.Then, given u,the following functional of φ,\nF[φ] =∝angbracketleftφ|PH|φ∝angbracketright\n∝angbracketleftφ|P|φ∝angbracketright+∝angbracketleftφ|U|φ∝angbracketright, (1)\nis bounded from below. To find the lowest energy with the quantum nu mbers specified by Pone can use a constrained\nsearch [8], in which one first considers only states that show a given d ensity profile τ(r),then one lets τvary,\nInfφ/bracketleftbigg∝angbracketleftφ|PH|φ∝angbracketright\n∝angbracketleftφ|P|φ∝angbracketright+∝angbracketleftφ|U|φ∝angbracketright/bracketrightbigg\n= Infτ/bracketleftbigg/parenleftbigg\nInfφ→τ∝angbracketleftφ|PH|φ∝angbracketright\n∝angbracketleftφ|P|φ∝angbracketright/parenrightbigg\n+/integraldisplay\ndrτ(r)u(r)/bracketrightbigg\n. (2)\nThe process goes in two steps, namely, i) a minimization within a given pa rticle density profile, τ(r)≡ ∝angbracketleftφ|c†\nrcr|φ∝angbracketright,for\nNparticles, with c†\nrandcrthe usual creation and annihilation operators at position r,then, ii) a minimization with\nrespect to the profile. The inner minimization clearly defines a DF, F[τ]≡Infφ→τ(∝angbracketleftφ|PH|φ∝angbracketright/∝angbracketleftφ|P|φ∝angbracketright).\nActually, it is more general [9] - [13] to use many-body density matric esBinN-body space, meaning mixed as well\nas pure states, and yielding a density τ(r) in one-body space,\nInfτ/bracketleftbigg\nInfB→τ/parenleftbiggTrBPH\nTrBP+TrBU/parenrightbigg/bracketrightbigg\n, (3)\nbut we shall use wave-functions in the following, namely B=|φ∝angbracketright∝angbracketleftφ|,for obvious pedagogical reasons. We shall assume\nthat thisInf φactuallydefinesanabsoluteminimum, Min φ,reachedatsomesolutionΦ ofthecorrespondingvariational\nprinciple. Moreover, we shall assume, temporarily at least, that th e solution Φ is unique. Uniqueness is not obvious,\nhowever, if only because many φ’s can give the same P|φ∝angbracketright,and, whenuvanishes, this variational principle, Eq. (2),\nreduces to the well-known “variation after projection” [14] metho d for Hartree-Fock calculations for instance.\nAnyhow,τanduare clearly conjugate in a functional Legendre transform, with δF/δτ=−u.Finally, ifρ(r)\ndenotes the profile of Φ when u→0,then the lowest energy with good quantum numbers is nothing but F[ρ].\nThe minimization, with respect to τ,of the functional, F[τ],provides simultaneously the density of the intrinsic\n(unprojected!) state and the projected energy. Notice, inciden tally, thatF[τ] depends on the choice of the variational\nset of trial functions φwhere the “inner minimization” is performed. Furthermore, it obviou sly depends on P.\nAmore generalargumentis possible, with morethan one kind ofident ical particles. Trialstatescan be, forinstance,\nproducts of determinants, one for each kind of fermions, and per manents, one for each kind of bosons. Consider for\ninstance the ammonia molecule, with i) its active electrons, ii) its three protons and iii) its nitrogen ion. It is trivial\nto include a center-of-mass trap into Hto factorize into a spherical wave packet the center of mass motio n of this\nself-bound system and avoid translational degeneracy problems. The complete Hamiltonian H, trial states φand\ndensity operators Bdepend on and describe simultaneously the electron, proton and nit rogen ion coordinates and\nmomenta. A DF in just the electronic density space, however, resu lts from the definition,\nF[τ] = Inf B→τTrBPH\nTrBP, (4)\nwherePprojectsgood quantumnumbers forthe wholesystem and traces aretakenoverall degreesoffreedom, while τ\nis set as only an electronic density. Interactions between heavy de grees of freedom, between heavy and electronic ones,\nand between electrons, are taken into account by the trace in the numerator. No Born-Oppenheimer approximation\nis needed for this “global” definition. For the sake of simplicity, howev er, we return in the following to the case of one\nkind of particles only. Most considerations which follow have obvious g eneralization for multicomponent systems.\nIII. VARIATIONAL EQUATIONS\nLetδφbe an infinitesimal variation of the trial function in its allowed domain. T hen, at first order, one obtains,\nδF=∝angbracketleftδφ|PH|φ∝angbracketright\n∝angbracketleftφ|P|φ∝angbracketright+∝angbracketleftδφ|U|φ∝angbracketright−∝angbracketleftδφ|P|φ∝angbracketright∝angbracketleftφ|PH|φ∝angbracketright\n(∝angbracketleftφ|P|φ∝angbracketright)2+\n∝angbracketleftφ|PH|δφ∝angbracketright\n∝angbracketleftφ|P|φ∝angbracketright+∝angbracketleftφ|U|δφ∝angbracketright−∝angbracketleftφ|P|δφ∝angbracketright∝angbracketleftφ|PH|φ∝angbracketright\n(∝angbracketleftφ|P|φ∝angbracketright)2. (5)3\nIf one defines the “gradient operator”,\nG=PH\n∝angbracketleftφ|P|φ∝angbracketright+U−P∝angbracketleftφ|PH|φ∝angbracketright\n(∝angbracketleftφ|P|φ∝angbracketright)2, (6)\nthen, obviously, δF=∝angbracketleftδφ|G|φ∝angbracketright+∝angbracketleftφ|G|δφ∝angbracketright.Note, incidentally, that Gis Hermitian.\nAt the minimum position Φ, the variation δFvanishes for any δφ.Replaceδφbyiδφto see that the difference,\n−∝angbracketleftδφ|G|Φ∝angbracketright+∝angbracketleftΦ|G|δφ∝angbracketright,vanishes as well. Then, trivially, at Φ ,both∝angbracketleftδφ|G|Φ∝angbracketrightand∝angbracketleftΦ|G|δφ∝angbracketrightvanish simultaneously,\n∝angbracketleftδφ|G|Φ∝angbracketright=∝angbracketleftΦ|G|δφ∝angbracketright= 0,∀δφ. (7)\nIn the special case of Slater determinants, let |ph∝angbracketright ≡c†\npch|Φ∝angbracketrightdenote any particle-hole state built upon |Φ∝angbracketrightas the\n“reference vacuum” for quasi-particles. Here c†andcare the familiar fermionic creation and annihilation operators,\nrespectively. Then the particle-hole matrix elements of Gvanish,\n∝angbracketleftph|G|Φ∝angbracketright= 0,∀ph. (8)\nAs long as a solution of this stationarity condition, Eq. (8), is not rea ched, the matrix elements, ∝angbracketleftph|G|φ∝angbracketright,define\nthe direction of the gradient of Fin the hyperplane tangent to the manifold of Slater determinants. A gradient\ndescent algorithm, |δφ∝angbracketright=−η/summationtext\nph|ph∝angbracketright∝angbracketleftph|G|φ∝angbracketright,whereηis a small step parameter, then leads to the solution. Notice,\nhowever, that the phrepresentation is covariant with φ.Thephbasis has to be recalculated at each step. Being state\ndependent, Gmust also be recalculated at each step.\nIV. SIMILAR THEORY, WITH A LAGRANGE MULTIPLIER\nThe slightly complicated gradient operator, Eq. (6), leads to a varia tional condition, Eq. (7), which combines\nthe matrix elements of three operators, namely PH, PandU.Define the number λ=∝angbracketleftΦ|PH|Φ∝angbracketright/∝angbracketleftΦ|P|Φ∝angbracketrightas a yet\nunknown Lagrange multiplier; it can be considered as an arbitrary pa rameter and shall be adjusted self-consistently\nlater, when Φ is reached. Then Eq. (7) also reads,\n∝angbracketleftδφ|(PH−λP+∝angbracketleftΦ|P|Φ∝angbracketrightU)|Φ∝angbracketright= 0. (9)\nIfφwere completely unrestricted, namely if δφwere completely general, this equation, Eq. (9), would mean that Φ is\nan eigenstate of the operator G.Since intrinsic states are understood to belong to restricted sets of states, the result\nΦ is only an approximate eigenstate of G.\nTo avoid the cumbersome coefficient, ∝angbracketleftΦ|P|Φ∝angbracketright,which multiplies U,it is convenient to define the auxiliary operator,\nH=PH−λP+W, (10)\nwhereW=/summationtext\niw(ri) is, likeU,an arbitrary, local, real, external field, bounded from below. It is o bvious that Gand\nHdefine a common solution Φ if uandware suitably proportional to each other, w=∝angbracketleftΦ|P|Φ∝angbracketrightu.In the following,\nhowever, we set Hab initio. It is an operator bounded from below. We are interested in its “almo st ground state”\nΞ and assume that this state is unique. A connection between a solut ion Ξ in this section and a solution Φ in the\nprevious section can easily be tested later.\nDefine again a constrained search for the lowest energy,\nInfφ∝angbracketleftφ|H|φ∝angbracketright= Infτ(Infφ→τ∝angbracketleftφ|H|φ∝angbracketright) = Inf τ/parenleftbigg\nFλ[τ]+/integraldisplay\ndrτ(r)w(r)/parenrightbigg\n, (11)\nwhere theλ-dependent DF, Fλ,is defined as,\nFλ[τ]≡Infφ→τ∝angbracketleftφ|(PH−λP)|φ∝angbracketright. (12)\nIt is again convenient, for pedagogy at least, to assume that this I nfφinduces an absolute minimum, reached at a\nposition Ξ in the variational space. The same assumption states tha t, givenλ,the absolute minimum of Fλ[τ] is\nFλ[σ] =∝angbracketleftΞ|(PH−λP)|Ξ∝angbracketright, (13)\nwhereσis the density of Ξ .LetEdenote this energy, E(λ)≡ ∝angbracketleftΞ|(PH−λP)|Ξ∝angbracketright.A simple manipulation then gives,\ndE\ndλ=−∝angbracketleftΞ|P|Ξ∝angbracketright. (14)\nA Legendre transform, using λand∝angbracketleftΞ|P|Ξ∝angbracketrightas conjugate variables, is thus available to return the matrix elemen t\n∝angbracketleftΞ|PH|Ξ∝angbracketrightas a function of the matrix element ∝angbracketleftΞ|P|Ξ∝angbracketright.Then one has just to locate the minimum of their ratio.\nNote again that the theory depends on the variational space wher eφevolves. But, in any case, one obtains\nsimultaneously the density of Ξ,the best intrinsic state, and the energy of its projected state P|Ξ∝angbracketright.4\nV. SUMMARY, DISCUSSION AND CONCLUSION\nIf only because of the need for spin densities [15] in the description of polarizable systems, the problem of symmetry\nconservation, or restoration, in DF theory has already received m uch attention in atomic and molecular physics [16]\n- [19]. It has been revisited here, in the spirit of the projected Hart ree-Fock method with variation afterprojection\n[14]: a variational principle for the density of an intrinsic state, witho ut symmetry, optimizes the energy of a state\nwith good quantum numbers. The idea was already introduced in the c ontext of particle number projection [20]. We\nhave shown in Secs. II and IV that our approach allows generalizatio ns of the Hohenberg-Kohn existence theorem.\nIt can be stressed that the present approach is concerned with t he density of an intrinsic state, not that of an\neigenstate. This is a major difference with all the other DF theories t hat we are aware of. Note, in particular, how\nour functional differs from a functional of a symmetrized [16] dens ity .\nWe showed in Sec. IV that a way to define the intrinsic Hamiltonian amou nts to a linear combination, H=\n−λP+PH,of the projector Pon the desired quantum numbers, and the laboratory Hamiltonian mu ltiplied by that\nsameP.Here, a subtle question must be raised, that of the nature of the in trinsic state. The more flexible the trial\nfunctions for this state, the better the projected state and th e lower the projected energy. However, full flexibility\ncontradicts simplicity, and, moreover, uniqueness of the intrinsic s tate; many different packets |φ∝angbracketrightcan give the same\nP|φ∝angbracketright.Symmetry projection brings correlations which, therefore, mustbe absent from the intrinsic state. This is why\nvariational domains for intrinsic states must necessarily be much narrower than the full Hilbert space.\nIn practice, fortunately, intrinsic states are confined to non-line ar, curved [21] manifolds, such as coherent states,\nSlater determinants, etc., which do not make linear subspaces. The intrinsic state, therefore, is not an exact eigenstate\nofH.It just minimizes a related quantity, the projected energy. It mus t be concluded that DF theory for an intrinsic\nstate necessarily depends on two factors, namely, i) obviously the quantum numbers to be projected out, but also ii)\nthe variational space retained for this intrinsic state.\nAcknowledgements : It is a pleasure for B.R.B. and B.G.G. to thank the TRIUMF Laborator y, Vancouver, B.C.,\nCanada, for its hospitality, where part of this work was done. The N atural Science and Engineering Research Council\nof Canada is thanked for financial support. TRIUMF receives fede ral funding via a contribution agreement through\nthe National Research Council of Canada. B.R.B. also thanks Servic e de Physique Th´ eorique, Saclay, France, and\nthe Gesellschaft f¨ ur Schwerionenforschung mbh Darmstadt, Ge rmany, for their hospitality, where parts of this work\nwere carried out, and acknowledges partial support from the Alex ander von Humboldt Stiftung and NSF Grant\nPHY0555396. Thecontributionoftwoanonymousrefereesinmakin gthispaperclearerisalsogratefullyacknowledged.\n[1] P. Hohenberg and W. Kohn, Phys. Rev. 136, B864 (1964)\n[2] T.L. Gilbert, Phys. Rev. B 12, 2111 (1975)\n[3] R. M. Dreizler and E. K. U. Gross, Density Functional Theory , Springer, Berlin/Heidelberg (1990); see also the referen ces\nin their review\n[4] J. Dobaczewski, H. Flocard and J. Treiner, Nucl. Phys. A 422, 103 (1984)\n[5] J. Meyer, J. Bartel, M. Brack, P. Quentin and S. Aicher, Ph ys. Lett. B 172, 122 (1986)\n[6] T. Duguet, Phys. Rev. C 67, 044311 (2003); T. Duguet and P. Bonche, Phys. Rev. C 67, 054308 (2003)\n[7] G.F. Bertsch, B. Sabbey and M. Uusn¨ akki, Phys. Rev. C 71, 054311 (2005)\n[8] M. Levy, Proc. Natl. Acad. Sci. USA 766062 (1979)\n[9] E.H. Lieb, Int. J. Quant. Chem. 24, 243 (1983)\n[10] H. Englisch and R. Englisch, Phys. Stat. Solidi B123, 711 (1984); B124, 373 (1984)\n[11]´A. Nagy and M. Levy, Phys. Rev. A 63, 052502 (2001)\n[12] J.P. Perdew and S. Kurth, in A Primer in Density Functional Theory , C. Fiolhais, F. Nogueira and M. Marques, eds.,\nLecture Notes in Physics, 620, Springer, Berlin (2003); see also the references in their r eview\n[13] R. van Leeuwen, Advances Quant. Chem. 43, 25 (2003)\n[14] R. Dreizler, P. Federman, B.G. Giraud and E. Osnes, Nucl . Phys.A 113145 (1968); S. Das Gupta, J.C. Hocquenghem\nand B.G. Giraud, Nucl. Phys. A 168625 (1971)\n[15] O. Gunnarson and B. J. Lundquist, Phys. Rev. B 13, 4274 (1976)\n[16] A. G¨ orling, Phys. Rev. A 47, 2783 (1993)\n[17] A. G¨ orling, Phys. Rev. A 59, 3359 (1999)\n[18] M. Levy and ´A. Nagy, Phys. Rev. Lett. 83, 4361 (1999)\n[19] A. G¨ orling, Phys. Rev. Lett. 85, 4229 (2000)\n[20] M.V. Stoitsov, J. Dobaczewski. R. Kirchner, W. Nazarew icz and J. Terazaki, Phys. Rev. C 76014308 (2007); J.A. Sheik\nand P. Ring, Nucl. Phys. A 66571 (2000)\n[21] B.G. Giraud and D.J. Rowe, J. Physique Lett. 40, L177-L180 (1979); B.G. Giraud and D.J. Rowe, Nucl. Phys. A 330,\n352 (1979)" }, { "title": "0707.3759v1.The_space_of_density_states_in_geometrical_quantum_mechanics.pdf", "content": "arXiv:0707.3759v1 [quant-ph] 25 Jul 2007THE SPACE OF DENSITY STATES IN GEOMETRICAL\nQUANTUM MECHANICS\nJES´US CLEMENTE-GALLARDO AND GIUSEPPE MARMO\nAbstract. We present a geometrical description of the space of density states\nof a quantum system of finite dimension. After presenting a br ief summary\nof the geometrical formulation of Quantum Mechanics, we pro ceed to describe\nthe space of density states D(H) from a geometrical perspective identifying\nthe stratification associated to the natural GL(H)–action on D(H) and some\nof its properties. We apply this construction to the cases of quantum systems\nof two and three levels.\nKeywords :Density states, projective space, geometric quantum mechanics\nPACS: 03.65.-w, 03.65.Ta\n1.Introduction\nA comparison of the frameworks underlying classical and quantum m echanics\nshows that the two descriptions have several common mathematic al structures.\nHowever, a striking difference emerges: the classical setting is geo metrical and non-\nlinear while the quantum is algebraic and linear. The emphasis on the und erlying\nlinearity in quantum mechanics is usually attributed to the description of the inter-\nferencephenomena[10]. Therefore,thecarrierspaceofquantu msystemsisrequired\nto be a Hilbert space Hfrom the beginning. The Hermitian structure is required\nto describe the probabilistic interpretation of Quantum Mechanics. However, it\nis exactly this probabilistic interpretation which forces on us the iden tification of\nphysical states not with the Hilbert space but rather with the spac e of rays, i.e.\nthe complex projective space of H, sayRH. Of course RHis a genuine nonlinear\nmanifold and on it the Hermitian structure gives rise to a K¨ ahler stru cture.\nThe appearance of this manifold in the quantum setting calls for a geo metrical\nformulation of Quantum Mechanics. It is clear that in the manifold view point we\nhave to give up the usual “superposition of states” and the notion of operators,\neigenvectors and eigenstates as usually presented. Nevertheles s, due to their phys-\nical relevance and interpretation we must be able to recover these “attributes” for\nquantum systems also at the manifold level. The overall formulation m ust allow\nfor nonlinear transformations and therefore only tensorial obje cts should be iden-\ntified with physically relevant quantities. To fully exploit the geometric al picture,\none prefers to work with real differential manifolds, i.e. one replace s the complex\nvector space Hwith its realification HR. The Hermitian structure then splits into a\ncomplex structure, a symplectic structure and a Riemannian struc ture (compatible\namongthem to define a K¨ ahlerstructure) Hermitian operatorsar etransformed into\nfunctions by replacing them by their expectation values. These fun ctions project\nontoRH. With the help of the Poissontensor associated with the symplectic s truc-\nture it is possible to give rise to a flow by integrating the Hamiltonian vec tor field\n12 JES ´US CLEMENTE-GALLARDO AND GIUSEPPE MARMO\nassociated with the expectation value function corresponding to a given Hermitian\noperator [1, 2, 3, 4, 5, 6, 7, 8, 9, 15, 23].\nThe symplectic structure appearing in Quantum Mechanics makes als o possible\nto consider it as a “classical field theory” associated with a Lagrang ian description\nwith relativity group the Galilei group [21], since we deal with non-rela tivistic\nQuantum Mechanics. Thus, classical and quantum descriptions hav e in common a\nsymplectic structure. However, this is the only quantum feature t hat has a direct\nclassical analogue. Some characteristic features like the quantum uncertainties\nand state vector reduction in a measurement process are strictly related to the\nadditional complex structure, available in Quantum Mechanics but no t present in\nClassicalMechanics [20]. This additional structure lies at the heart o f the difference\nbetween the mathematical structures underlying the two theorie s, much more than\nthe linear structure.\nGoing back to the manifold view point introduced in Quantum Mechanics , i.e.\nthe identification of RHas the true manifold of physical states, we have to recover\nthe notion of superposition of physical states [19] . This has been d one and creates\na deep relation with Pancharatnam connection, Bargmann invariant s and geomet-\nric phases [22]. In addition we have to recover the notion of “eigenve ctor” and\n“eigenvalue”. As a matter of fact by considering the expectation v alue function\nassociated with any operator we find that their critical points will co rrespond to\neigenstates and their values at those critical points correspond t o the eigenvalues.\nIf we consider the expectation value-functions of generic operat ors we get com-\nplex valued functions on RHand they provide us with a C∗–algebra, thus paving\nthe way for the geometrization of the C∗–algebraic approach to quantum theories.\nIn this latter approach usually the space of states is identified with t he space of\nnormalized positive linear functionals on the C∗–algebra describing the quantum\nsystem. As a provisional geometrization of this approach we shall c onstruct these\nspaces with the help of the momentum map associated with the symple ctic action\nof the unitary group on the K¨ ahler manifold RH.\n2.A brief exposition of geometrical quantum mechanics\nThe aim ofthis section is to present a brief summary ofthe set of the geometrical\ntools which characterize the description of Quantum Mechanics [13 , 14, 18].\n2.1.The states. The first step consists in replacing the usual complex vector\nspace structure of the Hilbert space Hof a quantum system by the corresponding\nrealification of the vector space. We shall denote as HRthe resulting vector space.\nIn this realification process the complex structure on Hwill be represented by a\ntensorJonHR.\nThe natural identification is then provided by\nψR+iψI=ψ∈ H /ma√sto→(ψR,ψI)∈ HR.\nUnder this transformation, the Hermitian product becomes, for ψ1,ψ2∈ H\n/an}bracketle{t(ψ1\nR,ψ1\nI),(ψ2\nR,ψ2\nI)/an}bracketri}ht= (/an}bracketle{tψ1\nR,ψ2\nR/an}bracketri}ht+/an}bracketle{tψ1\nI,ψ2\nI/an}bracketri}ht)+i(/an}bracketle{tψ1\nR,ψ2\nI/an}bracketri}ht−/an}bracketle{tψ1\nI,ψ2\nR/an}bracketri}ht).\nTo consider HRjust as a real differential manifold, the algebraic structures avail-\nable onHmust be converted into tensor fields on HR. To this end we have to\nintroduce the tangent bundle THRand its dual the cotangent bundle T∗HR. TheTHE SPACE OF DENSITY STATES IN GEOMETRICAL QUANTUM MECHANIC S 3\nlinear structure available in HRis encoded in the vector field ∆\n∆ :HR→THRψ/ma√sto→(ψ,ψ)\nWe can consider the Hermitian structure on HRas an Hermitian tensor on THR.\nWith every vector we can associate a vector field\nXψ:φ→(φ,ψ)\nTherefore, the Hermitian tensor, denoted in the same way as the s calar product\nwill be\n/an}bracketle{tXψ1,Xψ2/an}bracketri}ht=/an}bracketle{tψ1,ψ2/an}bracketri}ht\nThe scalar product above is written as /an}bracketle{tψ1,ψ2/an}bracketri}ht=g(Xψ1,Xψ2) +iω(Xψ1,Xψ2),\nwheregis now a symmetric tensor and ωa skew-symmetric one. It is also possible\nto write them as a pull-back by means of the dilation vector field ∆ as:\n(∆∗(g+iω))(ψ,φ) =/an}bracketle{tψ,φ/an}bracketri}htH\nThe properties of the Hermitian product ensure that:\n•the symmetric tensor is positive definite and non-degenerate, and hence\ndefines a Riemannian structure on the real vector space.\n•the skew-symmetrictensorisalsonondegenerate, andisclosedwit h respect\nto the natural differential structure of the vector space. Henc e, the tensor\nis a symplectic form.\nAs the inner product is sesquilinear, it satisfies\n/an}bracketle{tψ1,iψ2/an}bracketri}ht=i/an}bracketle{tψ1,ψ2/an}bracketri}ht,/an}bracketle{tiψ1,ψ2/an}bracketri}ht=−i/an}bracketle{tψ1,ψ2/an}bracketri}ht.\nThis implies\ng(Xψ1,Xψ2) =ω(JXψ1,Xψ2).\nWe also have that J2=−I, and hence that the triple ( J,g,ω) defines a K¨ ahler\nstructure. This implies, among other things, that the tensor Jgenerates both\nfinite and infinitesimal transformations which are orthogonal and s ymplectic.\nLinear transformations are converted into (1 ,1)–tensor fields by setting A→TA\nwhere\nTA:THR→THR(ψ,φ)/ma√sto→(ψ,Aφ).\nThe association A→TAis an associative algebra isomorphism. It is possible to\nrecover the Lie algebra of vector fields by setting XA=TA(∆). Complex linear\ntransformations will be represented by (1 ,1)–tensor fields commuting with J.\nFor finite dimensional Hilbert spaces it may be convenient to introduc e adapted\ncoordinates on HandHR. Fixing an orthonormal basis {|ek/an}bracketri}ht}of the Hilbert space\nallows us to identify this product with the canonical Hermitian produc t ofCn:\n/an}bracketle{tψ1,ψ2/an}bracketri}ht=/summationdisplay\nk/an}bracketle{tψ1,ek/an}bracketri}ht/an}bracketle{tek,ψ2/an}bracketri}ht\nThe group of unitary transformations on Hbecomes identified with the group\nU(n,C), its Lie algebra u(H) withu(n,C) and so on.\nThe choice of the basis also allows us to introduce coordinates for th e realified\nstructure:\n/an}bracketle{tek,ψ/an}bracketri}ht= (qk+ipk)(ψ),\nand write the geometrical objects introduced above as:\nJ=∂pk⊗dqk−∂qk⊗dpkg=dqk⊗dqk+dpk⊗dpkω=dqk∧dpk4 JES ´US CLEMENTE-GALLARDO AND GIUSEPPE MARMO\nIf we combine them in complex coordinates we can write the Hermitian s tructure\nin a simple way zn=qn+ipn:\nh=/summationdisplay\nkd¯zk⊗dzk\nIn an analogous way we can consider a contravariant version of the se tensors. It\nis also possible to build it by using the isomorphism THR↔T∗HRassociated to\nthe Riemannian tensor g. The result in both cases is a K¨ ahler structure for the\ndual vector space H∗\nRwith the dual complex structure J∗, a Riemannian tensor G\nand a (symplectic) Poisson tensor Ω: The coordinate expressions w ith respect to\nthe natural base are:\n•the Riemannian structure G=/summationtextn\nk=1/parenleftBig\n∂\n∂qk⊗∂\n∂qk+∂\n∂pk⊗∂\n∂pk/parenrightBig\n,\n•the Poisson tensor Ω =/summationtextn\nk=1/parenleftBig\n∂\n∂qk∧∂\n∂pk/parenrightBig\n•while the complex structure has the form\nJ=n/summationdisplay\nk=1/parenleftbigg∂\n∂pk⊗dqk−∂\n∂qk⊗dpk/parenrightbigg\n2.2.The observables. The space of observables (i.e. of self-adjoint operators\nacting on H) may be identified with the dual u∗(H) of the real Lie algebra u(H),\naccording to the pairing between the unitary Lie algebra and its dual given by\nA(T) =i\n2TrAT\nUnder the previous isomorphism, u∗(H) becomes a Lie algebra with product\ndefined by\ni[A,B] = [A,B]−= (AB−BA)\nWe can also transfer the Jordan product:\n[A,B]+= 2A◦B=AB+BA\nBoth structures are compatible. As a result, u∗(H) becomes a Jordan-Lie algebra\n(see [11, 16]).\nWe can also define a suitable scalar product, given by:\n/an}bracketle{tA,B/an}bracketri}ht=1\n2TrAB\nwhichturnsthespaceintoarealHilbertspace. Thisscalarproduct istherestriction\nof the one on gl(H) defined as /an}bracketle{tM,N/an}bracketri}ht=1\n2TrM†N.\nBesides this scalar product is compatible with the Lie-Jordan struct ure in the\nfollowing sense:\n/an}bracketle{t[A,ξ],B/an}bracketri}htu∗(H)=/an}bracketle{tA,[ξ,B]/an}bracketri}htu∗(H)/an}bracketle{t[A,ξ]+,B/an}bracketri}htu∗(H)=/an}bracketle{tA,[ξ,B]+/an}bracketri}htu∗(H)\nThese algebraic structures may be given a tensorial translation in t erms of the\nassociation A/ma√sto→TA. However we can also associate complex valued functions with\nlinear operators A∈gl(H) by means of the scalar product\ngl(H)∋A/ma√sto→fA=1\n2/an}bracketle{tψ,Aψ/an}bracketri}htH.\nIn more intrinsic terms we may write\nfA=1\n2(g(∆,XA)+iω(∆,XA)).THE SPACE OF DENSITY STATES IN GEOMETRICAL QUANTUM MECHANIC S 5\nHermitian operators give rise thus to quadratic real valued functio ns.\nThe association of operators with quadratic functions allows also to recover the\nalgebraic structures on u(H) andu∗(H) by means of approapriate (0 ,2)–tensors on\nHR. By using the contravariant form of the Hermitian tensor G+iΩ given by:\nG+iΩ = 4∂\n∂zk⊗∂\n∂¯zk=∂\n∂qk⊗∂\n∂qk+∂\n∂pk⊗∂\n∂pk+i∂\n∂qk∧∂\n∂pk,\nit is possible to define a bracket\n{f,h}H={f,h}g+i{f,h}ω\nIn particular, for quadratic real valued functions we have\n{fA,fB}g=fAB+BA= 2fA◦B{fA,fB}ω=−ifAB−BA\nThe imaginary part, i.e. {·,·}ω, defines a Poisson bracket on the space of func-\ntions. Both brackets allow us to define a tensorial version of the Lie -Jordan algebra\nof the set of operators.\nFor Hermitian operators we recover previously defined vector field s:\ngradfA=/tildewideA; HamfA=/tildewideriA\nwhere the vector fields associated with operators, we recall, are d efined by:\n/tildewideA:HR→THRψ/ma√sto→(ψ,Aψ)\n/tildewideriA:HR→THRψ/ma√sto→(ψ,JAψ)\nWe can alsoconsider the algebraicstructure associatedto the full bracket{·,·}H,\nas we associated above the Jordan product and the commutator o f operators to the\nbrackets {·,·}gand{·,·}ωrespectively. It is simple to see that it corresponds to the\nassociative product of the set of operators, i.e.\n{fA,fB}H={fA,fB}g+i{fA,fB}ω=fAB+BA+ifAB−BA= 2fAB\nThis particular bilinear product on quadratic functions may be writte n also as\na star product\n{fA,fB}H= 2fAB=/an}bracketle{tdfA,dfB/an}bracketri}htH∗=fA⋆fB\nThe set of quadratic functions endowed with such a structure tur ns out to be a\nC∗–algebra.\nWe see then that we can reconstruct all the information of the alge bra of oper-\nators starting only with real-valued functions defined on HR. We have thus\nProposition 1. The Hamiltonian vector field Xf(defined as Xf=ˆΩ(df)) is a\nKilling vector field for the Riemannian tensor Gif and only if fis a quadratic\nfunction associated with an Hermitian operator A, i.e. there exists A=A†such\nthatf=fA.\nFinally, we can consider the problem of how to recover the eigenvalue s and\neigenvectors of the operators at the level of the functions of HR. It is simple to see\nthat\n•eigenvectors correspond to the critical points of functions fA, i.e.\ndfA(ψ∗) = 0 iffψ∗is an eigenvector of A6 JES ´US CLEMENTE-GALLARDO AND GIUSEPPE MARMO\n•the corresponding eigenvalue is recovered by the value\nfA(ψ∗)\n/an}bracketle{tψ∗,ψ∗/an}bracketri}ht\nThus we can conclude that the K¨ ahler manifold ( HR,J,ω,g) contains all the\ninformation of the usual formulation of Quantum Mechanics on a com plex Hilbert\nspace.\nUp to now we have concentrated our attention on states and obse rvables. If we\nconsider observables as generators of transformations, i.e. we c onsider the Hamil-\ntonian flow associated to the corresponding functions, the invaria nce of the tensor\nGimplies that the evolution is actually unitary. It is, therefore, natur al, to consider\nthe action of the unitary group on the realification of the complex ve ctor space.\n3.The momentum map: geometrical structures on g∗\nThe unitary action of U(H) onHinduces a symplectic action on the symplectic\nmanifold ( HR,ω). By using the association\nF:HR×u(H)→R(ψ,A)/ma√sto→1\n2/an}bracketle{tψ,Aψ/an}bracketri}ht=fiA(ψ),\nwe find, with FA=fiA:HR→R, that\n{F(A),F(B)}ω=iF([A,B]).\nThus if we fix ψ, we have a mapping F(ψ) :u(H)→R. Thus with any element\nψ∈ Hwe have an element in u∗(H). Hence it defines a momentum map\nµ:H →u∗(H),\nwhich provides us with a symplectic realization of the natural Poisson manifold\nstructure available in u∗(H). We can write the momentum map from HRtou∗(H)\nas\nµ(ψ) =|ψ/an}bracketri}ht/an}bracketle{tψ|\nIf we make the convention that the dual u∗(H) of the (real) Lie algebra u(H)\nis identified with Hermitian operators by means of a scalar product, t he product\npairing between Hermitian operators A∈u∗(H) and the anti-Hermitian element\nT∈u(H) will be given by\nA(T) =i\n2Tr(AT)\nIf we denote the linear function on u∗(H) associated with the element iA∈u(H)\nbyˆA, we have\nµ∗(ˆA) =fA\nThe pullback of linear functions on u∗(H) is given by the quadratic functions on\nHRassociated with the corresponding Hermitian operators.\nIt is possible to show that the contravariant tensor fields on HRassociated with\nthe Hermitian structure are µ–related with a complex tensor on u∗(H):\nµ∗(G+iΩ) =R+iΛ;\nwhere the two new tensors Rand Λ are defined by\nR(ξ)(ˆA,ˆB) =/an}bracketle{tξ,[A,B]+/an}bracketri}htu∗=1\n2Tr(ξ(AB+BA))THE SPACE OF DENSITY STATES IN GEOMETRICAL QUANTUM MECHANIC S 7\nand\nΛ(ξ)(ˆA,ˆB) =/an}bracketle{tξ,[A,B]−/an}bracketri}htu∗=1\n2iTr(ξ(AB−BA))\nClearly,\nG(µ∗ˆA,µ∗ˆB)+iΩ(µ∗ˆA,µ∗ˆB) =µ∗(R(ˆA,ˆB)+iΛ(ˆA,ˆB)).\nAs we know that u∗(H) is foliated by symplectic manifolds, we wish to consider\nmore closely the map from HRto the minimal symplectic orbit on u∗(H).\n4.The complex projective space\nAs we have already remarked, the association of states of the qua ntum system\nwith vectors in the Hilbert space needs further qualifications becau se of the prob-\nabilistic interpretation required in Quantum Mechanics. More specific ally, states\nshould be identified with rays in the Hilbert space, i.e. equivalence class es of vec-\ntors, orbits of non-null vectors under the action of C0=C−{0}. The equivalence\nclass of the vector ψ∈ Hwill be denoted then as\n[ψ] ={λψ,λ∈C0}\nAs the infinitesimal generatorsofthe realand imaginarypartsoft he actionof C0\nonHRaregivenby the dilationvectorfield ∆ and the vectorfield J(∆) respectively,\nit is clear that we have to undertake the projection of the relevant tensors on HR\nto the complex projective space or ray space RH.\nWithout entering in too many details, we find that we have to modify Gand Ω\nby a conformal factor to turn them into projectable tensors. Sp ecifically we have\n˜G=g(∆,∆)G,˜Λ =g(∆,∆)Λ.\nThese tensors are projectable onto non-degenerate contrava riant tensors on RH\nand givesrise to aLie-Jordanalgebrastructureon the spaceofre alvalued functions\nwhose Hamiltonian vector fields are also Killing vector fields for the pro jection˜G.\nAs a matter of fact a theorem by Wigner allows us to state that thes e functions\nare necessarily projections of expectation values of Hermitian ope rators\neA(ψ) =/an}bracketle{tψ,Aψ/an}bracketri}ht\n/an}bracketle{tψ,ψ/an}bracketri}ht.\nThe action of the unitary group may also be projected and gives rise to a sym-\nplectic action on RH. The momentum map from HRprojects onto the momentum\nmap from RHbecause it is equivariant with respect to the action of ∆ HonHR\nand the action of ∆ u∗onu∗(H).\nFromµ∗(ˆA) =fAwe find\n∆Hµ∗(ˆA) = 2µ∗(∆u∗ˆA) = 2µ∗(ˆA)\nThe momentum map for the projected action may be written in the fo rm\nµ([ψ]) =|ψ/an}bracketri}ht/an}bracketle{tψ|\n/an}bracketle{tψ,ψ/an}bracketri}ht=ρψ.\nThis map identifies RHwith the Hermitian operatorsin u∗(H) which are of rank\none and are projectors, i.e.\nρψρψ=ρψ,Trρψ= 1.\nAs the ray space is a principal bundle with base manifold RHand structure\ngroupC0, we may look for a connection one form.8 JES ´US CLEMENTE-GALLARDO AND GIUSEPPE MARMO\nThe connection one-form θis given by\nθ(ψ) =/an}bracketle{tψ,dψ/an}bracketri}ht\n/an}bracketle{tψ,ψ/an}bracketri}ht,\nwith associated curvature form ω=dθ, because the structure group is Abelian.\nThis curvature form coincides with the symplectic structure on RHarising from\nthe projection of ˜Λ (conformally related to Λ).\nIt is also possible to write the Hermitian tensor which coincides with the Her-\nmitian tensor on RHwhen evaluated on horizontal vector fields. We thus have\n/an}bracketle{tdψ,dψ/an}bracketri}ht\n/an}bracketle{tψ,ψ/an}bracketri}ht−/an}bracketle{tψ,dψ/an}bracketri}ht/an}bracketle{tdψ,ψ/an}bracketri}ht\n/an}bracketle{tψ,ψ/an}bracketri}ht2.\nIt is not difficult to see that both ∆ and J(∆) are annihilated by this tensor.\nThe embedding of RHintou∗(H) by means of the momentum map allows us\nto consider convex combinations of the image µ(RH)⊂u∗(H). The convex com-\nbinations will generate the space of density states, i.e. normalized p ositive linear\nfunctionals on the Lie-Jordan algebra of observables. This convex body inherits\nsome structures from those existing on u∗(H), which are particularly important\nand useful when we are interested in describing evolutions of state s which are not\nunitary. In particular they inherits a Poisson structure and a Jord an structure. In\nthe next section we shall study more closely the space of density st ates.\n5.The space of density states\nAs we have already remarked the space of density states is the spa ce of positive\nnormalized linear functionals on the real linear space of observables . A theorem by\nGleason[12] assertsthat these functionals may be represented b y suitable operators\nwhen the trace is used as a bilinear pairing. By using this theorem we ca n start by\nconsidering states directly as appropriate operators.\nLet us introduce first the space of all non-negatively defined oper ators, i.e. the\nspace of all those ρ∈gl(H) which can be written in the form\nρ=T†T T∈gl(H).\nWe shall denote by PHthis space of operators, which is a convex cone in u∗(H).\nBy imposing the condition Tr ρ= 1 we select in PHthe convex body of density\nstates which we denote by D(H). We shall also consider non-negative Hermitian\noperatorsanddensitystatesofrank k, anddenoteas Pk(H)andDk(H)respectively\nthe corresponding spaces.\nThe complex projective space is in one-to-one correspondence wit hD1(H). In-\ndeed, any state in D(H) can be written as a convex combination of distinct states\nρ=p1ρ1+(1−p1)ρ2, with 0≤p1≤1. We shall call extremal states those which\ncan not be written in this form (i.e. as convex combination of two ρ1andρ2). The\nextremal states are thus given by D1(H).\nAsΛandRarenotinvertiblein u∗(H), itisconvenienttousethepairingbetween\nu(H) andu∗(H) defined by the trace, to introduce two tensor fields on u∗(H). We\nset then\n˜J,R:Tu∗(H)→Tu∗(H)THE SPACE OF DENSITY STATES IN GEOMETRICAL QUANTUM MECHANIC S 9\ndefined as\n˜Jξ(XA) = (ξ,[A,ξ]−) = Λ(ξ)(dˆA)\nRξ(XA) = (ξ,[A,ξ]+) =R(ξ)(dˆA)\nThe image of ˜Jis the Hamiltonian involutive distribution associated with linear\nHamiltonian functions, and we shall denote it as DΛ. The image of Ris also a\ndistribution, which we shall denote as DR, but in this case it is not involutive.\nIt is possible to see that combining DRandDΛwe can define two distributions\nD0=DR∩DAandD1=DR+DΛwhich are indeed involutive.\nWe noticethat the tensors ˜JandRcommute, i.e. ˜J◦R=R◦˜J. Morespecifically\nwe have\n˜J(ξ)◦R(ξ)(XA) =R(ξ)◦˜J(ξ)(XA) = [A,ξ2]−.\nAs a result, we find that the distribution D0becomes:\nD0(ξ) ={[A,ξ2];A∈u∗(H)}.\nOnRH,D0coincides with DΛ.\nThe distribution D1is involutive and the leaves are related to orbits of the\nfollowingGL(H)–action:\nGL(H)×u∗(H)→u∗(H) (T,ξ)/ma√sto→TξT†.\nWe obtain some interesting results [13, 14]:\n(1) The Hermitian operators ρandρ′belong to the same GL–orbit if and only\nif they have the same number K+of positive eigenvalues and the same\nnumberK−of negative eigenvalues (counted with multiplicities).\n(2) AnyGL–orbitintersectingthe positive cone PHlies entirely in PH; sothat\nPHis stratified by the GL–orbits. These GL–orbits in PHare determined\nby the rank of the operator, i.e. they are exactly Pk(H).\nWhen we restrict to the space of density states by imposing the con dition Trρ=\n1, thisGL–action will not preserve the states. It is however possible to defin e a\nnew action that maps D(H) into itself by setting\nGL(H)×D(H)→ D(H) (T,ρ)/ma√sto→TρT†\nTr(TρT†).\nThis action does preserve the rank of ρand then the following proposition holds\ntrue:\nProposition 2. The decomposition of the convex body of density states D(H)into\norbits of the GL(H)–actionρ/ma√sto→TρT†\nTr(TρT†)is exactly the stratification\nD(H) =n/uniondisplay\nk=1Dk(H),\ninto states of a given rank.\nTheboundaryofthe convexbody ofdensitystatesconsistsofst atesofranklower\nthann, i.e.∂D(H) =/uniontextn−1\nk=1Dk(H), and each stratum is a smooth submanifold in\nu∗(H). However, the boundary ∂D(H) is not smooth (for n >2). We have the\nfollowing theorem:10 JES ´US CLEMENTE-GALLARDO AND GIUSEPPE MARMO\nTheorem 1. Every smooth curve γ:R→u∗(H)through the convex body of density\nstates is tangent, at every point, to the stratum to which it b elongs, i.e.\nγ(t)∈ Dk(H)⇒Tγ(t)∈Tγ(t)Dk(H).\nOne may gain further insight on the “location” of the boundary by us ing the\nnotion of “face”\nDefinition 1. A non-empty closed convex subset K0of a closed convex set Kis\ncalled a faceofKif any closed segment in Kwith an interior point in K0lies\nentirely inK0.\nThus, for any ρ∈ D(H) we may consider the decomposition H= Imρ+ Kerρ\ninto the kernel and the image of ρ. We have:\nProposition 3. The face of D(H)throughρ∈ Dk(H)consists of states Awhich\nare “projectable” with respect to the projection defined by Kerρ, i.e.KerA⊂Kerρ.\nThe face through ρis then equivalent to D(Imρ).\nThe inner product defined by the trace allows to define a probability t ransition\nfunction\np(ρ1,ρ2) = Trρ1ρ2,\nwhenρ1andρ2belong to the boundary ∂eD(H), the space of extremal states.\nThis function satisfies\n0≤p(ρ1,ρ2)≤1p(ρ1,ρ2) =p(ρ2,ρ1).\nMoreover,p(ρ1,ρ2) = 1 if and only if ρ1=ρ2.\nIt is not difficult to show that the Hamiltonian vector fields which leave Rin-\nvariant will preserve also the probability transition functions. This r esult is related\nto a theorem by Wigner and may be used to recover the space of den sity states\nstarting with a Poisson space carrying a compatible probability trans ition function.\nFurther details and a full treatement of Poisson spaces with a tran sition probability\nfunction have been considered by Landsman (see [17]).\n6.Two examples: gu(2)andgu(3)\n6.1.States of a two level system. Weshallconsiderinsomedetailtwoexamples.\nThe first one is the two level system with carrier space H=C2. We consider u(2)\nandu∗(2) and make a specific choice of basis\nσ0=/parenleftbigg1 0\n0 1/parenrightbigg\nσ1=/parenleftbigg0i\n−i0/parenrightbigg\nσ2=/parenleftbigg0 1\n1 0/parenrightbigg\nσ3=/parenleftbigg1 0\n0−1/parenrightbigg\nWe recall that\nσ1σ2=iσ3σ2σ3=iσ1σ3σ1=iσ2,\nalong with\nσ2σ1=−iσ3σ3σ2=−iσ1σ1σ3=−iσ2;\nwhich may be obtained considering the conjugate-transpose of an y product.\nWe define coordinate functions by writing\ny0(A) =1\n2Tr(σ0A), ya(A) =1\n2Tr(σaA).THE SPACE OF DENSITY STATES IN GEOMETRICAL QUANTUM MECHANIC S 11\nIn these coordinates the corresponding Poisson brackets for th e canonical Lie-\nPoisson structure on the dual of the Lie algebra read:\n{y0,ya}= 0{ya,yb}= 2ǫabcyc.\nThe expression of the Poisson tensor thus becomes:\nΛ = 2/parenleftbigg\ny1∂\n∂y2∧∂\n∂y3+y2∂\n∂y3∧∂\n∂y1+y3∂\n∂y1∧∂\n∂y2/parenrightbigg\nIt is also possible to construct the Riemann-Jordan tensor in the fo rm:\nR=∂\n∂y0⊗s/parenleftbigg\ny1∂\n∂y1+y2∂\n∂y2+y3∂\n∂y3/parenrightbigg\n+\ny0/parenleftbigg∂\n∂y0⊗∂\n∂y0+∂\n∂y1⊗∂\n∂y1+∂\n∂y2⊗∂\n∂y2+∂\n∂y3⊗∂\n∂y3/parenrightbigg\nwhere⊗smeans the symmetrized tensor product.\n6.2.Distributions associated with ΛandR.It is easy to see that the Hamil-\ntonian distribution is generated by\nH1=y3∂\n∂y2−y2∂\n∂y3, H2=y1∂\n∂y3−y3∂\n∂y1, H3=y2∂\n∂y1−y1∂\n∂y2,\nwhile the distribution associated with the Riemann-Jordan tensor is\nX0=ya∂\n∂ya+y0∂\n∂y0Xa=ya∂\n∂y0+y0∂\n∂ya\nIt is clear that X0is central and {Xa}are boosts of a four dimensional Lorentz\ngroup, therefore their commutator will provide us with the Lie algeb ra of the rota-\ntion group:\n[Xa,Xb] =ya∂\n∂yb−yb∂\n∂ya.\nThe intersection of the distribution associated with Rand the Hamiltonian dis-\ntribution associated will indeed be generated by the Hamiltonian vect or fields and\nis involutive with leaves which are symplectic two dimensional spheres. The dis-\ntribution generated by the union of the two distributions is the full L orentz group\ncentrally extended with the dilations. As the Lorentz group admits a s a covering\nSL(2,C)the central extension is isomorphic to GL(2,C). This is a generalproperty\nholding true in any dimension (see [14]).\nWe find that\nLemma 1. The rank of Λis zero ify2\n1+y2\n2+y2\n3= 0and the rank is equal to 2 if\ny2\n1+y2\n2+y2\n3>0.\nThe situation is richer with R:\nLemma 2. The rank of Ris\n•zero ify2\n0+y2\n1+y2\n2+y2\n3= 0\n•two ify0= 0andy2\n1+y2\n2+y2\n3>0.\n•three fory2\n0=y2\n1+y2\n2+y2\n3\n•four ify2\n0/ne}ationslash=y2\n1+y2\n2+y2\n312 JES ´US CLEMENTE-GALLARDO AND GIUSEPPE MARMO\n6.3.Density states. As we have already seen in the previous sections the set\nof states is identified with a subset of u∗(H) satisfying a positivity condition and\na normalization condition. In the specific situation we are considering , a generic\nHermitian matrix A=y0σ0+yaσawill define a state if\nTrA= 1 0≤y0+y3≤1,0≤y0−y3≤1,detA≥0.\nExplicitly we have\ny0=1\n2,/parenleftbigg1\n2+y3/parenrightbigg/parenleftbigg1\n2−y3/parenrightbigg\n−((y1)2+(y2)2)≥0,\nor\n(y3)2+(y2)2+(y1)2≤1\n4.\nThus in our parametrization states are determined by points in R4on the hy-\nperplaney0=1\n2, and on this three dimensional space are identified by the points\nin the ball of radius1\n2. When referring to states we replace Awithρand write:\nρ=/parenleftbigg1\n2+y3y2+iy1\ny2−iy11\n2−y3/parenrightbigg\n. (1)\nThe pure states corresponding to the vector ( z1,z2)∈C2with unit norm z1¯z1+\nz2¯z2= 1 has a density state\nρ=/parenleftbigg¯z1\n¯z2/parenrightbigg\n⊗(z1,z2) =/parenleftbiggz1¯z1¯z1z2\n¯z2z1z2¯z2/parenrightbigg\n.\nWithin the previous parametrization we find\ny3=1\n2(z1¯z1−z2¯z2), y1= Im(¯z1z2), y2= Re(¯z1z2),\nand for these points the inequality is saturated thus implying that th ey lie on the\nsurface of the ball of radius1\n2. We shall denote the set of density states by D. This\nset is the convex hull of the sphere of pure states. For any ρ∈ Dthere exist pure\nstatesρ1andρ2and a positive number psuch thatρ=pρ1+(1−p)ρ2.\nThese states, points on the surface, are in one-to-one corresp ondence with the\nunit rays in C2and the map is given by the momentum map associated with the\nsymplectic action of U(2) onRH ∼CP1. The ball of the density states is foliated\nby symplectic leaves associated with the coadjoint action of U(2), which coincide\nalso with the orbits of the SU(2) group.\nThe analysis of these orbits may also be done by considering the orbit s passing\nthrough diagonal matrices, in other terms\nρ=S/parenleftbigga0\n0b/parenrightbigg\nS†a+b= 1a≥0, b≥0.\nWe visualize the situation with the help of the following diagram: the seg ment\nconnecting (1\n2,1\n2) with (1,0) parametrizes the family of two dimensional spheres.\nThe point (1\n2,1\n2) coincides with the center of the ball and (1 ,0) belongs to the\noutmost sphere of pure states.THE SPACE OF DENSITY STATES IN GEOMETRICAL QUANTUM MECHANIC S 13\n(1,0)(0,1)\n(0.5,0.5)\nPSfrag replacementsab\nWhat we have described is usually known as the Bloch sphere represe ntation of\none qubit. The decomposition of a density states ρ, a point in the ball, as a convex\nsum of two pure states ρ1=|ψ1/angbracketright/angbracketleftψ1|\n/angbracketleftψ1,ψ1/angbracketrightandρ2=|ψ2/angbracketright/angbracketleftψ2|\n/angbracketleftψ2,ψ2/angbracketright, is given geometrically by\ndrawing a straight line through ρ: the states ρ1andρ2are the intersections of the\nline with the sphere. Evidently this decomposition may be done in a two p arameter\nfamily of ways.\nPSfrag replacements\ny1y2y3\nρρ1\nρ2ρ′\n1\nρ′\n2\n6.4.States of a three level system. NowH=C3. The states are normalized\npositive3 ×3matricesinside u∗(3). Wefirstconsiderthegeometricaltensorsdefined\nby means of the momentum map construction. We choose a basis for u(3) given by\nthe Gell-Mann matrices\nλ1=\n0 1 0\n1 0 0\n0 0 0\nλ2=\n0−i0\ni0 0\n0 0 0\nλ3=\n1 0 0\n0−1 0\n0 0 0\n\nλ4=\n0 0 1\n0 0 0\n1 0 0\nλ5=\n0 0−i\n0 0 0\ni0 0\nλ6=\n0 0 0\n0 0 1\n0 1 0\n14 JES ´US CLEMENTE-GALLARDO AND GIUSEPPE MARMO\nλ7=\n0 0 0\n0 0−i\n0i0\nλ8=1√\n3\n1 0 0\n0 1 0\n0 0−2\nλ0=/radicalbigg\n2\n3\n1 0 0\n0 1 0\n0 0 1\n\nThey satisfy the scalar product relation\nTrλµλν= 2δµν\nTheir commutation and anti-commutation relations are written in ter ms of the\nantisymmetric structure constants and symmetric d–symbols dµνρ. We find\n[λµ,λν] = 2iCµνρλρ[λµ,λρ]+= 2/radicalbigg\n2\n3λ0δµν+2dµνρλρ.\nThe numerical values turn out to be\nC123= 1, C 458=C678=√\n3\n2\nC147=−C156=C246=C257=C345=−C367=1\n2\nThe values of these symbols show the different embeddings of SU(2) intoSU(3)⊂\nU(3). For the other coefficients we have\ndjj0=−d0jj=−dj0j=/radicalbigg\n2\n3j= 1,···,8\n−d888=d8jj=djj8=dj8j=1√\n3j= 1,2,3\nd8jj=djj8=dj8j=−1\n2√\n3j= 4,5,6,7\nd3jj=djj3=dj3j=1\n2j= 4,5d3jj=djj3=dj3j=−1\n2j= 6,7\nd146=d157=d164=d175=−d247=d256=d265=−d274=1\n2\nd416=−d427=d461=−d472=d517=d526=d562=d571=1\n2\nd614=d625=d641=d652=d715=−d724=d751=−d742=1\n2\nThe indices appearing in the non-null structure constants are iden tifying the\ncorresponding λ–matrices whose pairwise commutators define SU(2)–subgroups.\nIt is now possible to introduce coordinate functions\nyµ(A) =1\n2TrλµA.\nIn these coordinates, a generic Hermitian matrix Acan be written as\nA=y0λ0+yrλr\nThe vector ( y0,/vector y)∈R9plays a similar role to the one we saw on U(2). Under\nconjugation with S∈SU(3), any matrix Acan be written as\nA=S\na0 0\n0b0\n0 0c\nS†.THE SPACE OF DENSITY STATES IN GEOMETRICAL QUANTUM MECHANIC S 15\nThe scalar product induced on vectors on R8will be invariant under the action\nofSO(8). It is now possible to write the Poisson tensor\nΛ = 2Cµνρyρ∂\n∂yµ∧∂\n∂yν\nand the Riemann-Jordan tensor\nR=∂\n∂y0⊗syµ∂\n∂yµ+y0∂\n∂yr⊗∂\n∂yr+dµνρyµ∂\n∂yν⊗s∂\n∂yρ.\nNow the analysis of the various distributions is more cumbersome, ho wever it is\neasy to identify a few elements:\nR(dy0) =yµ∂\n∂yµ,\nwhich is the dilation vector field on R9; whileR(dyr) =yr∂\n∂y0+y0∂\n∂yr+dµνryµ∂\n∂yν,\nwhere it is possible to identify a boost structure plus a correction du e to the d–\nsymbols. InanycasetheunionoftheHamiltoniandistributionandthe Riemannian-\nJordan distribution generates GL(3,C).\nThe set of states will again be identified as the subset of the Hermitia n matrices\nwhich are normalized and satisfy the positivity condition. If we set\nρ=\na¯h g\nh b¯f\n¯g f c\na,b,c∈Rf,g,h∈C.\nThe conditions for ρto be a state are:\n•a+b+c= 1\n•a≥0,b≥0,c≥0.\n• |f|2≤bc,|g|2≤ca,|h|2≤ab.\n•detρ=abc+2Re(fgh)−(a|f|2+b|g|2+c|h|2)≥0\nThese matrices form a convex set of R8. The trace condition allows to identify\nthis subset as a subset of the vector space of R8corresponding to the dual space of\nthe Lie algebra of SU(3).\nExtremal states are in one-to-one correspondence with the minim al symplectic\norbit of the unitary group according to the coadjoint action and co rresponds to\nCP2, the complex projective space of C3.\nPure states, rank one projectors, are given by vectors ( z1,z2,z3)∈C3with the\nnormalization condition z1¯z1+z2¯z2+z3¯z3= 1 as\n\nz1¯z1¯z1z2¯z1z3\n¯z2z1z2¯z2¯z2z3\n¯z3z1z3¯z2¯z3z3\n\nPrevious inequalities are saturated by these states.\nThese extremal states may be written in terms of the λ–matrices\nρψ=|ψ/an}bracketri}ht/an}bracketle{tψ|\n/an}bracketle{tψ,ψ/an}bracketri}ht=1\n3(I+√\n3naλa),\nwithnana= 1 andn⋆n=n, the star product being\n(a⋆b)l=√\n3dljkajbk16 JES ´US CLEMENTE-GALLARDO AND GIUSEPPE MARMO\nBy using the “radial-angular” parametrization of states, say\nρ=s\na0 0\n0b0\n0 0c\ns†a≥0,b≥0,c≥0, s∈SU(3),\nwe may study the structure of this union of symplectic orbits by con sidering the\nfamily of diagonal matrices with the positivity condition (elements of a positive\nWeyl chamber in the Abelian Cartan subalgebra). The hyperplane Tr ρ= 1 iden-\ntifies a triangle with the intersection with positive axes ( Oa,Ob,Oc ); i.e. in the\npositive octant.\nPSfrag replacements\nabc\n0\nEach internal point of the triangle correspondsto a 6–dimensional symplectic orbit,\nout of which we may consider convex combinations. Due to the action ofSU(3)\ncontaining the action of the discrete Weyl group, the symplectic or bits are actually\nparametrized by the following smaller triangle.THE SPACE OF DENSITY STATES IN GEOMETRICAL QUANTUM MECHANIC S 17\na=1c=0b=1a=0b=0c=1\nPSfrag replacements\na\nb\nc\n0\nWhena=b=c=1\n3we have the “maximally mixed state” which play a crucial\nrole when we consider composite systems and entangled states (th e orbit passing\nthrough this point degenerates to a zero dimensional orbit). On th e boundary\nof the bigger triangle the rank of ρis either 1 or 2. However the orbits passing\nthrough these points are diffeomorphic to CP2. For a generic point, the orbits are\ndiffeomorphic to SU(3)/U(1)×U(1). It appears quite clearly that the set of states\nis a stratified manifold characterized by the rank of the state. We s hall not indulge\nfurther on the geometrical analysis and refer to the literature fo r further details\nand applications.\nReferences\n[1] M.C. Abbati, R. Cirelli, P. Lanzavecchia and A. Mani´ a, Pure states of general quantum\nmechanical systems as K¨ ahler bundle , Nuovo Cimento B 83, pp 43–60, 1984\n[2] J.S. Anandan A Geometric approach to Quantum Mechanics Found.Phys. 21, pp 1265-1284,\n1991\n[3] A. Ashtekar and T.A. Schilling, Geometrical formulation of Quantum Mechanics , onEin-\nstein’s path , pp 23-65, New York: Springer, 1999\n[4] A.Benvegn´ u, N. Sansonetto and M. Spera Remarks on geometric quantum mechanics\nJ.Geom.Phys. 51, pp 229-243, 2004\n[5] A. Bloch, An infinite-dimensional Hamiltonian system on a projective Hilbert space , Trans.\nAMS302, pp 787–796, 1987\n[6] D. Brody and L.P. Hughston, Geometric quantum mechanics , J. Geom. Phys. 38, pp 19–53,\n2001\n[7] R. Cirelli and P. Lanzavecchia, Hamiltonian vector fields in Quantum Mechanics , Nuovo\nCimento B, 79, pp 271–283, 1984\n[8] R. Cirelli, A. Mani´ a and L. Pizzocchero, J.Math.Phys. 3 1, pp 2891-2903, 1990 (part I and II)\n[9] D. Cruscinski and A. Jamiolkowski Geometric phases in classical and Quantum Me-\nchanics Birkhauser,Boston,2004\n[10] P.A.M. Dirac, The Principles of Quantum Mechanics , Clarendon Press, Oxford, 2nd edition,\n1936.\n[11] G.G. Emch, Foundations of 20th century Physics , North Holland, Amsterdam, 198418 JES ´US CLEMENTE-GALLARDO AND GIUSEPPE MARMO\n[12] A.M.Gleason Measures on the closed subspaces of a Hilbert space J.Math.Mech 6, pp 885-893,\n1957\n[13] J. Grabowski, M. Ku´ s and G. Marmo, Geometry of quantum systems: density states and\nentanglement , J. Phys. A:Math. Gen 38, pp 10217-10244, 2005\n[14] J. Grabowski, M. Ku´ s and G. Marmo, Symmetry, group actions and entanglement , Open\nsys. & Information dyn.13, pp 343-362, 2006\n[15] A. Heslot, Quantum mechanics as a classical theory , Phys Rev D 31, pp 1341–1348, 1985\n[16] N.P. Landsman, Mathematical topic between Classical and Quantum Mechanic s,\nSpringer-Verlag, 1998\n[17] N.P. Landsman, Poisson spaces with a transition probability , Rev Math Phys 9, pp 29-57,\n1997\n[18] V. I. Manko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria, The geometry of density states ,\nRep. Math. Phys 55, pp 405-422, 2005\n[19] V. I. Manko, G. Marmo, E.C.G. Sudarshan and F. Zaccaria, Interference and entanglement:\nan intrinsic approach , J. Phys A: Math Gen 35, pp 7137-7157, 2002\n[20] G.Marmo, G. Scolarici, A.Simoni, F. Ventriglia, The Quantum-Classical Transition:The Fate\nof the Complex Structure Int. J.Mod.Geom.Meth.Phys. 2, pp 127-145, 2005\n[21] G.Marmo and G.Vilasi, Symplectic Structures and Quantum Mechanics Mod.Phys.Letters\nB10,pp 545, 1996\n[22] N. Mukunda, Arvind, E. Ercolessi, G. Marmo, G. Morandi, R. Simon, Bargmann invariants,\nnull phase curves and a theory of geometric phase , Phys Rev A 67, 042114, 2003\n[23] D.J. Rowe, A. Ryman and G. Rosensteel, Many body quantum mechanics as a symplectic\ndynamical system , Phys Rev A 22, pp 2362-2372, 1980\nBIFI-Universidad de Zaragoza, Corona de Arag ´on 42, 50009 Zaragoza-SPAIN\nDipartamento di Scienze Fisiche, Universit ´a Federico II and INFN- Sezione Napoli,\nVia Cintia I-80126 Napoli-ITALY" }, { "title": "0707.4428v2.Only_n_Qubit_Greenberger_Horne_Zeilinger_States_are_Undetermined_by_their_Reduced_Density_Matrices.pdf", "content": "arXiv:0707.4428v2 [quant-ph] 12 Feb 2008Onlyn-Qubit Greenberger-Horne-Zeilinger States are Undetermi ned by their\nReduced Density Matrices\nScott N. Walck∗and David W. Lyons†\nLebanon Valley College, Annville, PA 17003\n(Dated: December 4, 2007)\nThe generalized n-qubit Greenberger-Horne-Zeilinger (GHZ) states and thei r local unitary equiv-\nalents are the only states of nqubits that are not uniquely determined among pure states by their\nreduced density matrices of n−1 qubits. Thus, among pure states, the generalized GHZ state s are\nthe only ones containing information at the n-party level. We point out a connection between local\nunitary stabilizer subgroups and the property of being dete rmined by reduced density matrices.\nPACS numbers: 03.67.Mn,03.65.Ta,03.65.Ud\nQuantifying and characterizing multi-party quantum\nentanglement is a fundamental problem in the field of\nquantum information. Roughly speaking, one expects\nthe states that are “most entangled”to be the most valu-\nable resourcesfor carryingout quantum information pro-\ncessing tasks such as quantum communication and quan-\ntum teleportation, and to give the most striking philo-\nsophical implications in terms of the rejection of local\nhidden variable theories [1].\nAlthough no single definition of “most entangled”\nseems possible, since we know that multi-party entangle-\nment occurs in many types that admit at best a partial\norder [2], it is still worthwhile to consider properties that\ncarry some of the spirit of “most entangled.”\nOne such property is the failure of a state to be de-\ntermined by its reduced density matrices. As reduced\ndensity matrices contain correlation information pertain-\ning to fewer than the full number of parties in the sys-\ntem, states exhibiting entanglement involving allparties\nmust possess information beyond that contained in their\nreduced density matrices. Linden, Popescu, and Woot-\nters put forward this suggestion in [3, 4] and proved the\nsurprising result that almost all n-party pure states are\ndetermined by their reduced density matrices. In other\nwords, the set of n-party pure states undetermined by\ntheir reduced density matrices is a set of measure zero.\nIn [5], Di´ osi gave a constructive method that succeeds in\nalmost all casesfor determining a 3-qubit pure state from\nits reduced density matrices. Nevertheless, the question\nof precisely which states are determined by their reduced\ndensity matrices remained open.\nIn this Letter, we show that the only n-qubit states\nthat are undetermined among pure states by their\nreduced density matrices are the generalized n-qubit\nGreenberger-Horne-Zeilinger (GHZ) states,\nα|00···0/an}bracketri}ht+β|11···1/an}bracketri}ht, α,β /ne}ationslash= 0\nand their local unitary (LU) equivalents. This means\nthat, among pure states, the generalized GHZ states are\nthe only ones containinginformation at the n-partylevel.\nFor the case n= 3, this result was reported previously in[3]. Part of our argument employs the methods of [5] in\nan essential way.\nLetDnbe the set of n-qubit density matrices. If ρ∈\nDnis ann-qubit density matrix, and j∈ {1,...,n}is a\nqubit label, wemayforman( n−1)-qubitreduceddensity\nmatrixρ(j)= trjρby taking the partial trace of ρover\nqubitj. Let\nPTr :Dn→Dn\nn−1\nbe the map ρ/ma√sto→(ρ(1),...,ρ (n)) that associates to ρits\nn-tuple of (n−1)-qubit reduced density matrices. The\nmap PTr is neither injective (one-to-one) nor surjective\n(onto). Its failure to be surjective means that there are\nn-tuples of (n−1)-qubit density matrices that cannot be\nproduced from any n-qubit density matrix by the partial\ntrace. The question of whether a collection of ( n−1)-\nqubit reduced density matrices could have come from an\nn-qubit density matrix by the partial trace is the subject\nof recent and ongoing investigations [6, 7]. The failure\nof PTr to be injective means that multiple n-qubit states\ncan have the same reduced density matrices. States ρ1/ne}ationslash=\nρ2with PTr(ρ1) = PTr(ρ2) require more information for\ntheir determination than is contained in their ( n−1)-\nqubit reduced density matrices.\nLet\nPn={ρ∈Dn|ρ2=ρ}\nbe the set of pure n-qubit states. If we are interested\nprimarily in pure states, we can restrict the partial trace\nmap to pure n-qubit states.\nptr = PTr |Pn:Pn→Dn\nn−1\nGiven a pure state ψwith|ψ/an}bracketri}ht/an}bracketle{tψ| ∈Pn, the set\nptr−1(ptr(ψ)) contains all pure states with the same re-\nduced density matrices as ψ. (We abbreviate ptr( ψ) =\nptr(|ψ/an}bracketri}ht/an}bracketle{tψ|).)\nWedefineastate ψtobedetermined among pure states\nif ptr−1(ptr(ψ)) contains only |ψ/an}bracketri}ht/an}bracketle{tψ|, andundetermined\namong pure states if ptr−1(ptr(ψ)) contains more than2\none state. Similarly, we define a state ρ∈Dnto bede-\ntermined among arbitrary states if PTr−1(PTr(ρ)) con-\ntains onlyρ, andundetermined among arbitrary states if\nPTr−1(PTr(ρ)) contains more than one state.\nThe surprising result of Linden and Wootters [4] is\nthat almostall n-qubit pure statesaredeterminedamong\narbitrary states.\nNevertheless, there are pure states that are undeter-\nmined among pure states (and consequently undeter-\nmined among arbitrary states). For example, consider\nthe one-parameter family of n-qubit states\n|η/an}bracketri}ht=1√\n2|00···0/an}bracketri}ht+η√\n2|11···1/an}bracketri}ht,\nwhereηisacomplexnumberwithmagnitudeone. If η1/ne}ationslash=\nη2, then|η1/an}bracketri}htand|η2/an}bracketri}htare different states with different\ndensity matrices |η1/an}bracketri}ht/an}bracketle{tη1| /ne}ationslash=|η2/an}bracketri}ht/an}bracketle{tη2|, yet they share the\nsame reduced density matrices, that is, ptr( |η1/an}bracketri}ht/an}bracketle{tη1|) =\nptr(|η2/an}bracketri}ht/an}bracketle{tη2|).\nWe see that almost all pure n-qubit states are deter-\nmined among pure states, yet n-qubit GHZ states are\nundetermined among pure states. The question then be-\ncomes, precisely which states ψare undetermined among\npure states?\nMain Result. Ann-qubit state ψis undetermined\namong pure states if and only if ψis LU-equivalent to\na generalized n-qubit GHZ state.\nProof.Letψbe ann-qubit pure state.\nSuppose that ψis LU-equivalent to a generalized n-\nqubit GHZ state, so we have\nU|ψ/an}bracketri}ht=α|00···0/an}bracketri}ht+β|11···1/an}bracketri}ht,\nwhereUis a local unitary transformation. Define\nU|ψ′/an}bracketri}ht=α|00···0/an}bracketri}ht −β|11···1/an}bracketri}ht. Then |ψ/an}bracketri}ht/an}bracketle{tψ| /ne}ationslash=\n|ψ′/an}bracketri}ht/an}bracketle{tψ′|, sinceαβ/ne}ationslash= 0, but ptr( ψ) = ptr(ψ′). Hence,\nψis undetermined among pure states.\nConversely, suppose that |ψ/an}bracketri}htis undetermined among\npure states. Then there is an n-qubit state vector |ψ′/an}bracketri}ht /ne}ationslash=\neiα|ψ/an}bracketri}htthat hasthe samereduceddensitymatricesas |ψ/an}bracketri}ht.\nClaim: If |ψ/an}bracketri}htand|ψ′/an}bracketri}hthave the same reduced density\nmatrices, then for each qubit j∈ {1,...,n}, there is\na one-qubit local unitary transformation Ljsuch that\n|ψ′/an}bracketri}ht=Lj|ψ/an}bracketri}ht.\nTo prove this, let j∈ {1,...,n}be a qubit label. Let\nρjdenote the one-qubit reduced density matrix of |ψ/an}bracketri}htfor\nqubitj. We write ρjas a spectral decomposition,\nρj=1/summationdisplay\nij=0pij\nj|ij/an}bracketri}ht/an}bracketle{tij|,\nfor some orthonormal basis |ij/an}bracketri}ht, wherep0\njandp1\njare the\neigenvalues of ρj. Ifp0\nj/ne}ationslash=p1\nj, then the orthonormal basis\n|ij/an}bracketri}htis uniquely determined up to a phase. If p0\nj=p1\nj,\nthen any one-qubit orthonormal basis can be used.The (n−1)-qubit reduced density matrix\nρ(j)= trj|ψ/an}bracketri}ht/an}bracketle{tψ|,\nobtained by taking the partial trace of |ψ/an}bracketri}ht/an}bracketle{tψ|over qubit\nj, has the same nonzero eigenvalues as ρj,\nρ(j)=1/summationdisplay\nij=0pij\nj|ij;(j)/an}bracketri}ht/an}bracketle{tij;(j)|.\nIfp0\nj/ne}ationslash=p1\nj, then the (n−1)-qubit eigenvectors |0;(j)/an}bracketri}htand\n|1;(j)/an}bracketri}htare unique up to a phase. If p0\nj=p1\nj, then the\neigenvectorsof ρ(j)with eigenvalue p0\nj=p1\nj= 1/2 consti-\ntute a two-dimensional subspace of the 2n−1-dimensional\nvector space of ( n−1)-qubit vectors, and any orthonor-\nmal pair of vectors in this subspace may be chosen as a\nbasis.\nWe choose the one-qubit orthonormal basis |ij/an}bracketri}htand\nthe (n−1)-qubit orthonormal basis |ij;(j)/an}bracketri}htso that|ψ/an}bracketri}ht\ncan be written\n|ψ/an}bracketri}ht=/radicalBig\np0\nj|0/an}bracketri}ht⊗j|0;(j)/an}bracketri}ht+/radicalBig\np1\nj|1/an}bracketri}ht⊗j|1;(j)/an}bracketri}ht,\nwhere⊗jis the tensor product that inserts a one-qubit\nket just before the jth factor in the ( n−1)-qubit ket\n|ij;(j)/an}bracketri}ht.\nNow|ψ′/an}bracketri}htcan be regarded as the state of a bipartite\nsystem composed of qubit jand all qubits but j, and has\na Schmidt decomposition with respect to those subsys-\ntems,\n|ψ′/an}bracketri}ht=/radicalBig\nq0\nj|0′/an}bracketri}ht⊗j|0′;(j)/an}bracketri}ht+/radicalBig\nq1\nj|1′/an}bracketri}ht⊗j|1′;(j)/an}bracketri}ht,\nwhere|0′/an}bracketri}ht,|1′/an}bracketri}htare orthonormal one-qubit vectors and\n|0′;(j)/an}bracketri}ht,|1′;(j)/an}bracketri}htare orthonormal ( n−1)-qubit vectors.\nTaking the partial trace over qubit j, we have\ntrj|ψ′/an}bracketri}ht/an}bracketle{tψ′|=q0\nj|0′;(j)/an}bracketri}ht/an}bracketle{t0′;(j)|+q1\nj|1′;(j)/an}bracketri}ht/an}bracketle{t1′;(j)|.\nSince this must be equal to ρ(j), it must have the eigen-\nvalues ofρ(j),q0\nj=p0\njandq1\nj=p1\nj. We consider two\ncases, depending on whether ρ(j)has distinct eigenvalues\nor not. Let us treat first the case of distinct eigenvalues,\np0\nj/ne}ationslash=p1\nj. In this case, the eigenvector |0′;(j)/an}bracketri}htcan be\noff by at most a phase from the eigenvector |0;(j)/an}bracketri}ht, and\nsimilarly for |1′;(j)/an}bracketri}ht. The same argument applied to the\none-qubit reduced density matrix ρjshows that |0′/an}bracketri}htcan\nbe off by at most a phase from |0/an}bracketri}ht, and similarly for |1′/an}bracketri}ht.\nIn this case, then, we can write\n|ψ′/an}bracketri}ht=/radicalBig\np0\nj(Lj|0/an}bracketri}ht)⊗j|0;(j)/an}bracketri}ht+/radicalBig\np1\nj(Lj|1/an}bracketri}ht)⊗j|1;(j)/an}bracketri}ht,\n(1)\nwithLja 2×2 diagonal unitary matrix.\nLet us treat next the case of repeated eigenvalues,\np0\nj=p1\nj. In this case, the eigenvectors |0′;(j)/an}bracketri}htand\n|1′;(j)/an}bracketri}htmustmerelyspanthesametwo-dimensionalcom-\nplex space that is spanned by |0;(j)/an}bracketri}htand|1;(j)/an}bracketri}ht. In this3\ncase, the primed eigenvectors must be related to the un-\nprimed eigenvectors by a two-dimensional unitary trans-\nformation,\n|0′;(j)/an}bracketri}ht=u00|0;(j)/an}bracketri}ht+u01|1;(j)/an}bracketri}ht\n|1′;(j)/an}bracketri}ht=u10|0;(j)/an}bracketri}ht+u11|1;(j)/an}bracketri}ht\nwith\n/bracketleftbiggu00u01\nu10u11/bracketrightbigg\n∈U(2).\nThe same argument applied to the one-qubit reduced\ndensity matrix ρjshows that there must be some 2 ×2\nunitary matrix vlmwith\n|0′/an}bracketri}ht=v00|0/an}bracketri}ht+v01|1/an}bracketri}ht\n|1′/an}bracketri}ht=v10|0/an}bracketri}ht+v11|1/an}bracketri}ht.\nIn this case, then, we can write equation (1) with Ljthe\n2×2unitary matrixequaltothe product ofthe transpose\nofvlmwithulm. (We have abused notation by using the\nsymbolLjto represent both the 2 ×2 unitary matrix,\nandalsothelocalunitarytransformationon n-qubitstate\nvectors\nI⊗···⊗I⊗Lj⊗I⊗···⊗I,\nwith the 2 ×2 matrixLjin thejth slot of this tensor\nproduct, and one-qubit (2 ×2) identity operators in all\nother slots.) This completes the proof of the Claim.\nFor each pair of qubit labels j,k, we have\n|ψ/an}bracketri}ht=L−1\nkLj|ψ/an}bracketri}ht.\nNext, spectrally decompose each Ljwith unitary matri-\ncesUjso that\nDj=UjLjU−1\nj\nare diagonal. We have\nD−1\nkDjU1···Un|ψ/an}bracketri}ht=U1···UnU−1\nkD−1\nkUkU−1\njDjUj|ψ/an}bracketri}ht\n=U1···Un|ψ/an}bracketri}ht.\nUsing the multi-index I= (i1i2···in), where each ijis\nzero or one, and the basis\n|I/an}bracketri}ht=|i1i2···in/an}bracketri}ht=|i1/an}bracketri}ht⊗|i2/an}bracketri}ht⊗···⊗|in/an}bracketri}ht,\nexpand\nU1···Un|ψ/an}bracketri}ht=/summationdisplay\nIcI|I/an}bracketri}ht\nand write\nDj=eiαj/bracketleftbiggeiβj0\n0e−iβj/bracketrightbigg\n,whereeiβj/ne}ationslash=e−iβj, since|ψ′/an}bracketri}htis a different state from\n|ψ/an}bracketri}ht. Now we have\ncI=cIexp{i[αj−αk+(−1)ijβj−(−1)ikβk]}\nfor all multi-indices Iand allj,k∈ {1,...,n}. We see\nthat for each multi-index I, eithercI= 0 or\nexp{i[αj−αk+(−1)ijβj−(−1)ikβk]}= 1\nfor allj,k∈ {1,...,n}. LetJbe a multi-index with\ncJ/ne}ationslash= 0. IfIis any multi-index that agrees with Jin at\nleast one entry (say the jth qubit entry), and disagrees\nwithJin at least one entry (say the kth qubit entry),\nthencI= 0. We conclude that cJandcJ, whereJis the\nmulti-index consisting of the complements of each of the\nnbits in multi-index J, are the only nonzero coefficients\ninU1···Un|ψ/an}bracketri}ht. Consequently, |ψ/an}bracketri}htis LU-equivalent to a\ngeneralized n-qubit GHZ state.\nMuch ofthis argumentcarries overto the more general\nsituation of a system of nparties in which party jhas\ndimensiondj(partyjis a qubit if and only if dj= 2). In\nparticular, the Claim carries over. Suppose that |ψ/an}bracketri}htand\n|ψ′/an}bracketri}htare states of a system of nparties in which party j\nhas dimension dj. If|ψ/an}bracketri}htand|ψ′/an}bracketri}hthave the same reduced\ndensitymatrices,thenforeachparty j∈ {1,...,n},there\nis a one-party local unitary transformation Ljsuch that\n|ψ′/an}bracketri}ht=Lj|ψ/an}bracketri}ht. The second part of the argument is com-\nplicated by the possibility of repeated eigenvalues in the\ntransformations Dj. We leave this as a question for fu-\nture work.\nIt is worthwhile to point out the significance of stabi-\nlizer subgroups of the local unitary group in this work.\nThe local unitary group for n-qubit density matrices is\nthe groupG=SU(2)n, consisting of a special unitary\ntransformationoneachqubit. Each(pureormixed)state\nρhas a stabilizer subgroup Iρconsisting of elements of G\nthat leaveρfixed under the action gρg−1forg∈G. We\nhave seen that a state that is undetermined among pure\nstates has a special type of enlarged stabilizer subgroup.\nThere are two alternative formulations of the main re-\nsult that may suggest promising avenues for the classi-\nfication of entanglement types. One can precisely char-\nacterize the pure n-qubit states that are undetermined\namong pure states in terms of the stabilizer subalgebra\nof the state, that is the Lie algebra of the stabilizer sub-\ngroup of the state. It is shown in [8] that the generalized\nn-qubit GHZ state has (for n≥3) stabilizer subalgebra\nKρ=\n\nn/summationdisplay\nj=1itjZj/vextendsingle/vextendsingle/vextendsingle/vextendsingle/vextendsingle/vextendsinglen/summationdisplay\nj=1tj= 0\n\n,\nwhereZjisthe Paulimatrix σzapplied toqubit j. States\nofnqubits undetermined amongpure states areprecisely\nthose which are LU-equivalent to states with this subal-\ngebra.4\nA second alternative formulation of the main result is\nin terms of the dimension of the stabilizer subgroup. For\nn= 3andn≥5,ann-qubitstateisundeterminedamong\npure states if and only if it is not a product state and its\nstabilizer subgroup has dimension n−1 [9].\nWe have shown that all n-qubit states other than gen-\neralizedn-qubit GHZ states and their LU-equivalents are\ncompletelydeterminedbytheirreduceddensitymatrices.\nIs it necessaryto specify allofthe reduced density matri-\nces? Which states are undetermined by specifying only\nn−1 (rather than all n) of the (n−1)-qubit reduced\ndensity matrices? Is it a larger set than the generalized\nGHZs? The answer is yes, and stabilizers can help us\nunderstand this. For example, the state\n|χ/an}bracketri}ht=1√\n3(|0000/an}bracketri}ht+|0001/an}bracketri}ht+|1111/an}bracketri}ht)\nis undetermined by its 3-qubit reduced density matrices\nobtained by taking the partial trace over qubit 1, the\npartialtraceoverqubit 2, andthepartialtraceoverqubit\n3. It is not LU-equivalent to a generalized 4-qubit GHZ\nstate (it has a different stabilizer subalgebra structure,\nand stabilizer subalgebra structure is an LU-invariant).\nNote thatZ1|χ/an}bracketri}ht=Z2|χ/an}bracketri}ht=Z3|χ/an}bracketri}ht /ne}ationslash=eiα|χ/an}bracketri}ht. The state\nZ1|χ/an}bracketri}hthas the same 3-qubit reduced density matrices as\n|χ/an}bracketri}htwhen taking the partial trace over qubit 1, 2, or 3.\nThese two states have a different 3-qubit reduced density\nmatrix when taking the partial trace over qubit 4.\nA pure state’s LU-equivalence class is often considered\nto contain all of the information about the entanglement\nofthestate. Aninterestingquestionis: Arethere n-qubit\npure states with entanglement information that is not\ncontained in their reduced density matrices? We might\ninterpret this question as equivalent to the question: Are\ntheren-qubit pure states for which the LU-equivalence\nclass of the state is undetermined by its reduced densitymatrices? The answer to this question is no. Every n-\nqubit pure state can be determined (among pure states)\nup to a local unitary transformation by its reduced den-\nsity matrices. This can be seen directly from the Claim\nat the beginning of our proof. Two pure states that have\nthe same reduced density matrices must be LU equiva-\nlent.\nWe have not answered the question of which n-qubit\npure states are undetermined among arbitrary states by\ntheir reduced density matrices. The set of n-qubit states\nundetermined among arbitrary states must contain the\ngeneralized GHZ states, bit it could be strictly larger\nthan the set that is undetermined among pure states.\nThis remains an open question.\nThe authors thank the National Science Foundation\nfor their support of this work through NSF Award No.\nPHY-0555506.\n∗Electronic address: walck@lvc.edu\n†Electronic address: lyons@lvc.edu\n[1] D. Bouwmeester, J.-W. Pan, M. Daniell, H. Weinfurter,\nand A. Zeilinger, Phys. Rev. Lett. 82, 1345 (1999).\n[2] W.D¨ ur, G. Vidal, andJ.I.Cirac, Phys.Rev.A 62, 062314\n(2000), arXiv:quant-ph/0005115.\n[3] N. Linden, S. Popescu, and W. K. Wootters, Phys. Rev.\nLett.89, 207901 (2002).\n[4] N. Linden and W. K. Wootters, Phys. Rev. Lett. 89,\n277906 (2002).\n[5] L. Di´ osi, Phys. Rev. A 70, 010302(R) (2004).\n[6] Y.-J. Han, Y.-S. Zhang, and G.-C. Guo, Phys. Rev. A 70,\n042309 (2004).\n[7] J.-M. Cai, Z.-W. Zhou, S. Zhang, and G.-C. Guo, Phys.\nRev. A75, 052324 (2007).\n[8] S. N. Walck and D. W. Lyons, Phys. Rev. A 76, 022303\n(2007), arXiv:0706.1785 [quant-ph].\n[9] D. W. Lyons, S. N. Walck, and S. A. Blanda,\narXiv:0709.1105 [quant-ph]." }, { "title": "0709.1973v1.Density_of_states_of_helium_droplets.pdf", "content": "arXiv:0709.1973v1 [physics.atm-clus] 13 Sep 2007Density of states of helium droplets\nKlavs Hansen1\nDepartment of Physics, G¨ oteborg University, SE-412 96 G¨ o teborg, Sweden\nMichael D. Johnson, Vitaly V. Kresin\nDepartment of Physics and Astronomy,\nUniversity of Southern California, Los Angeles, Californi a 90089-0484,\nUSA\nNovember 1, 2018\nAbstract\nAccurate analytical expressions for the state densities of liquid4He droplets are\nderived, incorporatingthe ripplonand phonondegrees of fr eedom. Themicrocanon-\nical temperature and the ripplon angular momentum level den sity are also evalu-\nated. The approach is based on inversions and systematic exp ansions of canonical\nthermodynamic properties.\n1 Introduction\nImportant dynamical processes in finite systems such as nuclei, po lyatomic molecules,\nnanoclusters, atomic clouds, droplets frequently turn out to be s tatistical in nature:\nevaporation/fragmentation, radiation, emission of electrons, eq uilibration between inter-\nnal degrees of freedom or between host and solvent molecules. Wh en such a system\nis thermally isolated, e.g. when flying in a beam or suspended in a trap, t he proper\nstatistical-mechanics treatment is that of the microcanonical ens emble where the energy\nEis fixed and not the temperature of an external heat bath. The de nsity of states, or\nlevel density, ρ(E)dErepresents the number of quantum states between energy Eand\nE+dE. For separable degrees of freedom this is number of normal mode c ombination\nsuch that their energies add up to a total internal energy lying in th is interval. The func-\ntion plays a crucial role in the thermal description of microcanonical systems. For low\nexcitation energies ρ(E) can be represented by a sum of delta functions, corresponding\nto excitations of a only a few of the individual modes, but for even mo derate excitation\n1email: klavs@physics.gu.se\nphone:+46 (0)31 772 3432\nFAX: +46 (0)31 772 3496\n1energies the density of these delta functions becomes so large tha t is is well described as\na continuous function of energy. In this situation it is most convenie nt to use a density\nsmoothed over the discreteness of the energy levels. In addition t o energy systems de-\nscribed with the microcanonical ensemble have a conserved total a ngular momentum, so\nthe correspondingly resolved density of states, ρ(E,J), is often of relevance.\nFree liquid helium nanodroplets [1, 2] represent an interesting syste m for a statistical\ntreatment. One reason is that helium is the only element which cannot be described in\nterms of classical dynamics for any internal degrees of freedom u nder the experimental\nconditions used to study the droplets. This makes the system inter esting in its own right.\nFor example, ”magic number” maxima in the size distributions of small4He clusters have\nbeen shown to correlate with the ability of the cluster to accommoda te elementary ex-\ncitation modes [3]. A second reason is the use of the droplets as micro -cryostats used\nto investigate other clusters and molecules. Evaporative cooling ge nerates internal ener-\ngies corresponding to temperatures of ≈0.4 K and is used to thermalize impurities to\nthis otherwise unreachable temperature for gas phase molecule an d cluster beams. Under-\nstanding these processes requires accurate density-of-state s expressions for the elementary\nexcitation spectrum.\nThe calculation of level densities requires that the excitation spect rum is known. At\nlowtemperatures, therelevantnormalmodesof4HeNclusterswithintheliquiddropmodel\nareripplonswhich arequantized capillarysurfacewaves, andphono nswhicharequantized\nbulk compression waves. For large droplets these modes are separ able to a good approxi-\nmation [4], a fact that greatly facilitates a statistical analysis of the excitation spectrum.\nFor a spherical droplet, both ripplon and phonon modes possess we ll-defined eigenvalue\nspectra characterized by angular momentum for ripplons, and ang ular momentum and\nmode index for phonons. A calculation of the total density of state s requires enumeration\nof all possible normal mode combinations, with individual energies and angular momenta\nadding up to a given total EandJ.\nA leading-order evaluation of ρ(E) for ripplons was carried out by Brink and Stringari\n[5]. Subsequently, Lehmann [6] presented a comprehensive discuss ion of the densities of\nstates for ripplons and phonons computed by direct numerical cou nting, and showed that\nthe resultant plots of the logarithm of the level densities could be we ll parameterized by\npolynomial fits. These fits were then used to calculate other therm odynamic functions\nand to analyze droplet cooling with angular momentum conservation c onstraints [7, 8].\nIn this paper, we show that accurate density of states functions can be obtained by\nanalytic evaluation. This is appealing in its own right, as the calculations take advantage\nof several elegant and generally useful tools from the literature. In addition, having ana-\nlytic expressions for various types of elementary excitations prov ides a systematic method\nfor treating situations where several types of normal modes are excited simultaneously, or\nwhen the spectrum of elementary excitations is modified.\nThe plan of the paper is as follows. In Section 2 we calculate the ripplon density of\nstates as a function of energy, Section 3 considers its angular mom entum dependence, Sec-\ntion 4 is devoted to phonon excitations, Section 5 to the angular mom entum of phonons,\nand Section 6 to the total ρ(E) function. Section 7 comments on the similarity between\n2the spectra considered here and those of multielectron bubbles in b ulk liquid helium, and\npresents a summary.\n2 Ripplon density of states\nAs mentioned above, ripplons are quantized waves on the droplet su rface. For a spherical\nliquid drop, the elementary excitation spectrum is given by [9]\nεℓ= ¯hω0/radicalBig\nℓ(ℓ−1)(ℓ+2). (1)\nHereℓ≥2 is the angular momentum quantum number of the wave and\nω0=/radicalbiggσt\nDR3=/radicalBigg\n4πσt\n3maN, (2)\nwhereσtis the coefficient of surface tension, Dthe mass density, Rthe droplet radius, ma\nthe atomic mass, and Nthe number of atoms in the droplet. If the parameters of bulk\nliquid helium are used, we have ¯ hω0≈/parenleftBig\n3.8/√\nN/parenrightBig\nK in temperature units [6]. Below, the\nripplon energy will be expressed in dimensionless units scaled to the qu antity ¯hω0. Each\nmode has a degeneracy of (2 ℓ+1).\nCanonical approximation\nAfirstapproximationtothelevel densitycanbederivedinthecanon icalensemble picture,\nwhere it is assumed that the system possesses a definite temperat ureT, and the system’s\ninternal energy is associated with its most probable value. The ener gy density of states\nof a finite system is then given by [5, 10, 11, 12]\nρ(E) =eS\n/radicalBig\n−2π(∂E/∂β), (3)\nwhereβ≡(kBT)−1and\nS=βE+lnZ (4)\nis the entropy; Zis the canonical partition function. In the following we will use units\nwherekB= 1. The square root appearing in the equation involves the heat cap acity\nand appears because the canonical entropy includes an approxima tely gaussian integral\nover the thermally populated states with a width given by the heat ca pacity and the\ntemperature, see, e.g., Ref. [12].\nSince the ripplon elementary excitations are bosons we have\nlnZ=−ℓmax/summationdisplay\nℓ=2(2ℓ+1)ln/parenleftBig\n1−e−βεℓ/parenrightBig\n. (5)\n3The canonical thermal energy of the ripplon ensemble is\nE=−∂(lnZ)/∂β. (6)\nTo leading order, we can replace the sum in Eq. (5) by an integral fro m zero to\ninfinity, and approximate the energy eigenvalues (1) by εℓ≈ℓ3/2. The integral then\nstraightforwardly evaluates to\nlnZ= Γ/parenleftbigg7\n3/parenrightbigg\nζ/parenleftbigg7\n3/parenrightbigg\nβ−4/3= 1.685β−4/3, (7)\nand from Eq. (6) the (dimensionless) energy is\nE= 2.247β−7/3. (8)\nAssembling everything into Eq.(3) and expressing the answer in term s of the energy,\nwe find\nρrip(E)≈0.311E−5/7exp(2.476E4/7), (9)\nwhich is the same answer as in Ref. [5].\nMicrocanonical ensemble\nThe above calculation can be improved in two places. One obvious refin ement is to\nevaluate the sum (5) with greater care and to use more precise eige nvalues. A deeper\nconceptual question is how to compute thermodynamic quantities f or a finite isolated\nsystem for which the total internal energy is a conserved quantit y and not an expectation\nvalueandtheuseofa”temperature”mustbecarefullydefined. At horoughdiscussion was\ngiven by Andersen et al. in Ref. [12] with the conclusion that the conv enient canonical\nformalism may be retained, but the canonical expression for the en ergy (6) must be\ncorrected as follows:\nE=−∂(lnZ)/∂β−β−1. (10)\nHereEis the fixed excitation energy of the system and βis understood as the ”micro-\ncanonical temperature” defined as\nβ≡∂[lnρ(E)]/∂E. (11)\nThe procedure taken is as follows. First, the sum in Eq. (5) is calculat ed using the\nfirst three terms of the Euler-Maclaurin summation formula [13]. With the form of the\nspectrum given in Eq.(1), this formula becomes\n−lnZ=∞/integraldisplay\n2(2ℓ+1)ln/parenleftBig\n1−e−βεℓ/parenrightBig\ndℓ+5\n2ln/parenleftBig\n1−e−βε2/parenrightBig\n(12)\n−1\n12d\ndℓ/bracketleftBig\n(2ℓ+1)ln/parenleftBig\n1−e−βεℓ/parenrightBig/bracketrightBig/vextendsingle/vextendsingle/vextendsingle/vextendsingle/vextendsingle\nℓ=2+...\n4Theupper limit, ℓmax, hasbeenset toinfinityasbefore. Theactualvalueisontheorder of\nℓmax≈2πR/λmin≈2πR/(2d),whereλisthewavelength and distheinteratomicdistance\n[4]. In the liquid drop approximation ( R=N1/3d/2) one then has ℓmax≈πN1/3/2. In\nview of Eqs. (1, 2) this yields a size-independent ripplon Debye tempe rature of εmax≈7.5\nK. Using this value to estimate the error in ln Z, the leftover terms are found to be on the\norder of ( βεmax/4−7ℓmax/6)exp(−βεmax).ForT= 1Kthis is a relative contribution to\nln(Z) of less than 10−2/N1/3which will be ignored.\nA tedious calculation of Eq.(12) involving expansions of exponentials in powers of β\nresults in [14]\nlnZ= 1.685β−4/3+0.639β−2/3−349\n96+7\n3ln(2√\n2β)+... (13)\nThe first term coincides with Eq. (7), and the rest are finite-size an d spectral correc-\ntions. Note that all the numerical coefficients derive from explicit ex pressions involving\nspecial functions. The expansion Eq.(13) has been checked agains t a numerical sum. The\ncomparison is shown in Fig. 1 for Helmholtz’ free energy, F=−Tln(Z). Already at tem-\nperatures where Tis equal to the lowest excitation energy ε2= ¯hω0√\n6, the free energy\nis well represented by the above expression. At higher energies th e agreement improves\nmonotonically.\nKnowing the partition function, we can now use Eq. (10) to determin e the relation\nbetween the microcanonical energy and temperature:\nE= 2.247β−7/3+0.426β−5/3−10\n3β−1. (14)\nAgain, the first term reproduces Eq. (8).\nIn order to proceed with the calculation of the entropy, the heat c apacity and the\nlevel density in Eq.(3), we need to invert the relation (14) which expr essesE(β) to get\nβ(E). This is done by the iterative method of successive approximations . The result is\nan expansion for β−1in powers of E−2/7,\nβ−1≡0.7069E3/7−0.07239E1/7+0.7212E−1/7+..., (15)\nwhere the coefficients are calculated from those entering Eq.(14). Now the prefactor and\nthe exponent in Eq.(3) can be evaluated, using Eqs. (4),(13),(14), andβ(E), finally\nyielding\nρrip(E) = 0.205E−12/7exp/parenleftBig\n2.476E4/7+0.507E2/7/parenrightBig\n. (16)\nLet us emphasize again that all the numerical coefficients encode an alytical expressions.\nEq.(16), which is the main result of this section, may be compared with the canonical\napproximation (9), an exact numerical count carried out with the h elp of the Beyer-\nSwinehart algorithm [15], and the form written down in Ref. [6] as an e mpirical fit to\nthe numerical count in the interval E=50-2500. Fig. 2 shows such a comparison, and\ndemonstrates that the analytical expression gives an excellent re presentation of the exact\nresult [16].\n53 Ripplon angular momentum density\nThe next step is to generalize the ripplon state density to a function which is not only\nenergy- but also angular momentum-resolved. This problem has bee n comprehensively\nstudied in nuclear physics [10, 11]. One way of visualizing the net angula r momentum of\na large distribution of excitations with varying ( ℓ,ℓz) is as the result of random angular\nmomentum coupling, in which case the central limit theorem applies and one expects to\nfind a normal distribution. Indeed, the above references show th atρ(E,J) is essentially\na product of ρ(E) and a Gaussian factor:\nρ(E,J) =ρ(E)2J+1\n2(2π)1/2σ3e−J(J+1)\n2σ2. (17)\nIt is permissible here to replace J(J+1) by ( J+1\n2)2.\nIt is still necessary to establish the variance σ2. An elegant way to do this to leading\norder by means of an extended grand canonical distribution is desc ribed in Bethe’s review\n[10], where the method is applied to a system of non-interacting ferm ions in a spheri-\ncal potential box. Here we follow the same procedure for a system of bosonic ripplon\nexcitations.\nThe idea is first to evaluate the projection Mof the net angular momentum /vectorJof the\ndroplet in terms of the contributions of individual normal modes at t emperature T. The\nfact that Mis a conserved quantity is accounted for by a separate Lagrange m ultiplier,\nor ”chemical potential” γ, such that\nγ=−∂S/∂M (18)\n(Sis the entropy). We can calculate Mdirectly by summing over all modes:\nM=∞/summationdisplay\nℓ=2ℓ/summationdisplay\nm=−ℓm\neβεℓ−γm−1≈∞/integraldisplay\n0dℓℓ/integraldisplay\n−ℓm·dm\neβℓ3/2−γm−1(19)\n(in reduced energy units). Expanding the integrand to first order inγ[10], we find\nM=20\n27Γ/parenleftbigg5\n3/parenrightbigg\nζ/parenleftbigg5\n3/parenrightbigg\nγβ−4/3, (20)\nand Eq. (8) allows us to express the result in terms of the droplet en ergy. To leading\norder we have:\nγ= 1.776ME−8/7. (21)\nNow we can use Eq. (18) with Eq. (21) to obtain the entropy variatio n:\nS(E,M) =S(E,0)−M2/(2σ2) (22)\nwith /parenleftBig\n2σ2/parenrightBig−1= 0.888E−8/7. (23)\n6The second term in Eq. (22) leads to a normal distribution in M. The distribution in\nJcan be shown to have the same variance [10, 11]. Therefore Eqs. (1 7) and (23) define\nρrip(E,J).\nThenumerical evaluationoftherotationaldensity ofstatesinRef . [6]ledtoessentially\nthesameformofthestatedensity function, withthefactorcorr esponding to(2 σ2)−1fitted\nas 0.868E−8/7+0.964E−13/7, which affirms the analytical result (23): the factors deviate\nby less than 2% for E=100-2500.\nA shorter estimate of the variance is illustrative. The number of qua nta in one mode\n(ℓ,m) is on the order of T/εℓfor levels up to εℓ≃Tand zero for higher quantum energies.\nThe total number of excited quanta is then\nn≈T2/3/summationdisplay\nℓ=2(2ℓ+1)T/ℓ3/2≈4T4/3, (24)\nwhere the sum was approximated by an integral and Tis written in terms of the ω0\nunit. With the energy-temperature relation (14) we get the leading order value for en-\nergy per quantum ∝angbracketlefte∝angbracketright=E/n= 2.247T/4,and from this an average of ∝angbracketleftℓ∝angbracketright=∝angbracketlefte∝angbracketright2/3=\nT2/3(2.247/4)2/3.The standard deviation σofℓis then, according to the ’random walk’\nargument used above, σ=√n∝angbracketleftℓ∝angbracketright. Inserting the calculated ∝angbracketleftℓ∝angbracketrightand expressing the result\nin terms of the total energy, one has\n/parenleftBig\n2σ2/parenrightBig−1= 2−1/3(2.247)−4/21E−8/7= 0.68E−8/7, (25)\nin surprisingly sensible agreement with the above result.\nOne may seek to describe the angular momentum distribution in the lan guage of\na rotational energy and a moment of inertia I, associating [11] the exponential in Eq.\n(17) with a Boltzmann factor involving β¯h2J(J+ 1)/(2I), i.e.,I= ¯h2βσ2. Using the\ncanonical-ensemble results, Eqs.(8) and (23), we can express the ”ripplon moment of\ninertia” in terms of the ripplon excitation energy (in reduced units):\nI= 0.797E5/7. (26)\n4 Phonon density of states\nSurface ripplons are the lowest-temperature droplet excitations ; bulk phonons appear\nnext. These are compression sound waves which arise as solutions o f the wave equation\nwithin the volume of the spherical drop. As such, their energies are given by\nεn,ℓ= ¯hukn,ℓ (27)\nwhereuis the speed of sound and the wave number kn,ℓis determined by the boundary\ncondition at the surface. If the Dirichlet boundary condition is adop ted [4, 6], then kn,ℓ=\nan,ℓ/R, wherean,ℓis thenth root of the jℓspherical Bessel function. For a free surface, a\nmore appropriate boundary condition is the Neumann one, in which ca sekn.ℓ=a′\nn,ℓ/R,\n7witha′\nn,ℓthe root of the Bessel function derivative, j′\nℓ. The energy scale is set by the\nlongest wave length, i.e., k∼π/R, so we can express phonon energies in units of\n˜ε= ¯huπ/R, (28)\nwhich works out to ˜ ε=/parenleftBig\n25.5N−1/3/parenrightBig\nK in temperature units if the speed of sound in bulk\n4He is used [6]. The leading-order behavior of the phonon state densit y can be determined\nin a straightforward way by invoking the standard expression for t he Debye heat capacity\n(per unit volume) of bulk phonons:\nCbulk=2π2\n15k4\nB\n¯h3u3T3. (29)\nMultiplying this by 4 πR3/3 and using the fact that (in the canonical framework)\nC=∂E/∂Tand (kBT)−1=∂S/∂E, we can use integrations to express Sin terms of E.\nThen, from Eq.(3) we find that to first order, ln ρph(E)≈S(E) = 3.41E3/4. This matches\nthe leading term of the fit to a direct numerical count in Ref. [6] which is 3.331E3/4.\nThe Debye temperature for phonons in liquid4He is≈25 K [21], corresponding to a\ntotal phonon thermal energy (from Eq.31) of ≈1000NK. We can therefore use the low\ntemperature approximation throughout.\nTheprospectofrefiningthecalculationbyanalyticallyevaluatingast atisticalsumover\nthe precise spectrum (27) may seem bleak, as the Bessel function roots which ”contribute\nin an essential manner are just the zeros for which the usual form ulae (like McMahon’s\nexpansion) are bad approximations” [17]. However, rescue comes f rom an elegant math-\nematical framework known as the Weyl expansion [17, 18, 19]. It pr ovides a systematic\nexpression for the smoothed density of eigenmodes in a finite cavity in terms of volume,\nsurface, and curvature terms. As described in the above refere nces, this is a very general\ntheory, valid for both scalar and vector wave equations, and applic able to a wide variety\nof physical phenomena.\nRef. [20] applied this formalism to the specific heat of metal nanopar ticles. The finite-\nsize correction to the specific heat (29) derived there is immediately usable for our droplet\nproblem:\nC=Cbulk+\n(−)9ζ(3)\n4πk3\nB\n¯h2u2T2\nR+1\n6k2\nB\n¯huT\nR2. (30)\nThe + sign applies to the Neumann and the - sign to the Dirichlet bounda ry conditions.\nAlthough we focus on the Neumann condition, the Dirichlet case will be included for\ncompleteness.\nWe now follow almost the same sequence as in the bulk limit described abo ve: Eq.\n(30) is multiplied by the droplet volume and integrated once to obtain ( withEandTin\nunits of ˜ε)\nE(T) =2π6\n45T4+\n(−)ζ(3)π2T3+π2\n9T2, (31)\n8and a second time to obtain S(T) asS=/integraltextT\n0(C/T′)dT′. The first function is inverted by\niteration to yield\nT(E) = 0.391E1/4−\n(+)0.069−0.006E−1/4. (32)\n[Calculating T(E) instead of β(E) is more convenient in this case.] Eq. (3) is then\nused to obtain the density of states. The calculation is done to the fi rst three orders\ninE, in correspondence to the three terms in the heat capacity expan sion (30). The\nmicrocanonical correction (10) in the present case turns out to c ontribute only in the\nnext order of smallness. The result of the calculation is as follows:\nρph(E) =AE−5/8exp/parenleftbigg\n3.409E3/4+\n(−)0.908E1/2+0.482E1/4/parenrightbigg\n(33)\nOnce again, the + sign is for the Neumann boundary condition on the p honon wave at the\ndroplet surface and (-) for the Dirichlet condition. Using the bulk ca nonical heat capacity\nin Eq.(30) gives a pre-exponential factor of A= 0.26.\nFig.3 compares the exact Beyer-Swinehart count for the phonon s pectrum with the\nfull Eq.(33) and with the level density based on the bulk Debye heat c apacity, Eq. (29),\ni.e., where only the first term in the exponent is present. Fig.(4) show s a more detailed\ncomparison of Eq.(33) and the exact-count phonon level density. We find good agreement\nbetween analytical expression and the exact computation, althou gh not as good as for the\nripplon case.\nTheestimate oftheprefactor AinEq. (33)cannot beexpected tobecorrectbecause it\ndoesnotincludehigher-orderexpansiontermsintheexponenttha twouldyieldcorrections\nof the same order. A comparison with the numerical count suggest s a correction in the\nform of a factor exp( −0.62E0.2). Although this correction is larger than the error found\nfor the ripplon level density, it is nevertheless still relatively small. An effective value of\nA≈0.05 can be used for energies below 400.\n5 Phonon angular momentum density\nA computation of the angular momentum resolved phonon level dens ity suffers from the\ndifficulties with expressing the lowest Bessel function eigenvalues wit h a simple functional\nform. In contrast to ρph(E) there is, to our knowledge, no solution in the literature\nfor this problem. As will be clear from the results presented in sectio n 6 below, the\ncontribution to the level density from the phonons is minor compare d to that of the\nripplons, and the required precision in the calculation of the angular s pecified phonon\ncontribution is therefore correspondingly smaller. In this section w e will make an order\nof magnitude estimate, based on the leading order term of McMahon ’s expansion of the\nroots of the Bessel functions [13]. For the Neumann boundary con dition the roots are\n(n+ℓ/2−3/4)π≈(n+ℓ/2)π. With the phonon energy scale used, Eq.(28), the quantum\nenergies are thus n+ℓ/2. When states with energies up to Tare averaged, the linear\ndependence of the quantum energy on ℓgives an average value of ∝angbracketleftℓ∝angbracketright ∼T. Since also the\nn-dependence is linear the constant of proportionality is on the orde r of unity. The total\n9number of states below energy Tis on the order of T3. Combining these estimates give,\nusing the same type of ’random walk’ estimate as Eq. (25) for the rip plons, that\n(2σ2\nph)−1∼1\nT3·T2=/parenleftBigg2π6\n45/parenrightBigg5/4\nE−5/4≈100E−5/4. (34)\nThe ratio of the σ’s for the phonons and ripplons (here denoted σrip) with the leading\norder terms in the caloric curves Eqs.(14,31) and the proper energ y scaling is\nσph\nσrip∼0.002(T[K])7/6N1/6. (35)\nThis is small compared to unity up to extremely large droplet sizes. Th e conclusion\nthat the width of the phonon angular momentum distribution can be ig nored holds very\nwell, even if the estimate of the width should be incorrect by as much a s an order of\nmagnitude.\n6 Combined level density\nA helium droplet may have both ripplon and phonon oscillations excited a t the same time\n(and, at higher temperatures, rotons as well [1]). The coupling bet ween these normal\nmodes is weak at bulk liquid surfaces [22] and in large droplets [4], thus their energy\ncontents may remain independently defined for some length of time, and the individual\nstate densities will then come from Eqs. (16) and (33). The questio n of equilibration\ndynamics of excitations in superfluid droplets and the relevant time s cales is a very in-\nteresting one, and has not yet been addressed in detail. Below, we d iscuss an estimate\nof state densities in circumstances when the ripplon and phonon exc itations do achieve\nstatistical equilibrium.\nIn principle, the level density of combined excitations can be calculat ed by direct\nsummation, as described in Section 2. This would be a very involved pro cedure, because\nthe ripplon and phonon quantum energies have different dispersion r elations and scale\ndifferently with size. Alternatively, one can calculate the level densit y as a convolution.\nAlso in this task does one benefit from formulating the general prob lem in terms of the\nmicrocanonical temperature. The convolution to be performed is\nρ(E) =/integraldisplayE\n0ρrip(E−ε)ρph(ε)dε. (36)\nFor not extremely large droplets the largest part of the excitation energy resides in the\nripplons. Indeed, the ratio between the energies of the ripplon and phonon subsystems is,\ncanonically:\nEph\nErip≈6.8×10−3N1/3(T[K])5/3. (37)\n(Temperature expressed in Kelvins.) It is clear that for temperatu res under 1 K (i.e.,\nthose which lie safely below the Debye cut-off values specified above a nd below the onset\n10ofrotonmodes)anddropletsofuptoseveraltensofthousands ofatomsinsize, thephonon\nenergy contents is a fraction of the ripplon energy. Under these c onditions one can treat\nthe ripplon degrees of freedom as a heat bath and calculate the pho non contribution with\nan expansion of the integrand of Eq. (36) around some small phono n energy. We will use\nthe simplest choice of zero phonon energy, and to increase the pre cision we expand the\nlogarithm of the level density. Thus\nρ(E) =/integraldisplayE\n0ρrip(E)exp/bracketleftBigg\n−εdln[ρrip(E)]\ndE+1\n2ε2d2ln[ρrip(E)]\ndE2−.../bracketrightBigg\nρph(ε)dε.(38)\nThe upper limit of the integral can be replaced by infinity without serio us loss of precision\nbecause the integrand peaks well below this value. We recognize the first derivative in\nthe exponential as the microcanonical temperature 1 /Tof the ripplon system at energy\nE, see Eq. (11), and therefore have\nρ(E) =ρrip(E)/integraldisplay∞\n0e−ε/Tρph(ε)exp/bracketleftBigg1\n2ε2d2ln[ρrip(E)]\ndE2+.../bracketrightBigg\ndε. (39)\nThe second exponential in the integrand can be expanded, with the integral of the first\nterm yielding the phonon canonical partition function at T,Zph(T):\nρ(E) =ρrip(E)/braceleftBigg\nZph(T)+/integraldisplay∞\n0e−ε/Tρph(ε)/bracketleftBigg1\n2ε2d2ln[ρrip(E)]\ndE2+.../bracketrightBigg\ndε/bracerightBigg\n.(40)\nTo leading order and ignoring the difference between the canonical a nd microcanonical\ntemperatures, this simplifies to\nρ(E) =ρrip(E)Zph(T)/braceleftBigg\n1−Cph\n2Crip−E2\nph\n2CripT2+.../bracerightBigg\n. (41)\nHence the ratio of the term which is second order in εto the zero order term in Eq.(40)\nis approximately\nCph\n2Crip+E2\nph\n2CripT2= 6×10−3(T[K])5/3N1/3+4×10−6(T[K])14/3N4/3.(42)\nFor not excessively large or warm droplets we can leave out the corr ection terms and thus\nhave\nρ(E) =ρrip(E)Zph(T), (43)\nwhereZph(T) as stated above is the phonon canonical partition function at the micro-\ncanonical ripplon temperature corresponding to the ripplon energ yE.\nThe exponential part of the phonon canonical partition function c an be calculated,\ne.g., by integration of the standard relation Eq.(6) with the caloric cu rve in Eq.(31). This\nprocedure does not determine the integration constant which tra nslates into a multiplica-\ntive constant on the total level density, Eq. (43). This constant ,c, is approximately the\n11product of the pre-exponential from Eq.(33) and the prefactor that appears in Eq.(3) (i.e.,\nthe value given by a saddlepoint expansion of the phonon level densit y in the calculation\nof the canonical particion function). The result is\nc≈/radicalBig\n2πT2CphAE−5/8, (44)\nwhereCphisagainthephononheatcapacity. Theleadingorderexpressions fo rthephonon\nparameters Cph(T),Eph(T) give, taking into account the different scaling of energies for\nphonons and ripplons, the total level density\nρ(E) =ρrip(E)·0.526N1/6exp/parenleftBig\n0.04713N−1/2T3+0.01317N1/3T2+0.1634N1/6T/parenrightBig\n(45)\nwith the equation given in ripplon energy units and T=T(E) given by Eq.(15). The\nconstant of 0 .05 for the phonon level density pre-exponential, mentioned at the end of\nSec. 4, was used here also.\nThis result is compared with numerical convolutions for N= 103in Fig.5. The\nnumerical convolution was also calculated for N= 104with a similar result.\nOne remark about Eqs.(43,45) is in place: These equations should only be used for\ncalculations of microcanonical properties. For the calculation of ca nonical properties one\nshould use the product partition function, Zripp,ph=ZripZph. A naive application of\nEq.(45) in a calculation of the partition function of the combined ripplo n-phonon system\nwill give a divergent result at all temperatures. The origin of this dive rgence is the\nbreakdown at high excitation energies of the approximations leading to the equation.\n7 Conclusions\nWe have presented an analytical evaluation of the statistical dens ity of states functions\nof the elementary excitations (surface ripplons and volume phonon s) of isolated liquid-\ndrop helium nanoclusters. These functions are expressed in terms of microcanonically\nconserved quantities: energy and angular momentum. The obtaine d formulas accurately\nmatch numerically computed curves as the energy level densities va ry over∼150��300\norders of magnitude.\nOther interesting helium systems to which the results may be applicab le include\nmicron-sized superfluid fog [23] and multielectron bubbles in liquid heliu m. The lat-\nter are spherical voids inside bulk He, with a thin shell of electrons linin g the inner wall\n(see, e.g., Refs. [24, 25] and references therein). They can unde rgo small-amplitude shape\noscillations, i.e., ripplons, whose frequency under zero external ap plied pressure has the\nformω2\nℓ∝(ℓ2−1)(ℓ−2) , which for large ℓapproaches the same form as the droplet\nripplon dispersion, Eq. (1). This implies that the statistical mechanic s of these bub-\nbles should be similar to that of nanodroplets. One distinction is that t he bubble are\nsubmerged into a bulk helium thermal bath, therefore for them the canonical ensemble\ntreatment is rigorously correct and not just a convenient approx imation.\nFinally, it should be pointed out that the results obtained in the prese nt paper have a\nuniversal form and are expressed in terms of dimensionless scaled e nergies, therefore they\nare generally applicable to the statistics of droplets of various subs tances besides helium.\n128 Acknowledgments\nThis work was supported by the Swedish National Research Council (VR), the U.S. Na-\ntional Science Foundation under Grant No. PHY-0245102, and a Lic k fellowship to M.J.\n13References\n[1] J.P. Toennies and A.F. Vilesov, Angew. Chem. Intern. Ed. 43(2004) 2622\n[2] F.Stienkemeier and K.K.Lehmann, J. Phys. B 39(2006) R127\n[3] R. Guardiola, O. Kornilov, J. Navarro, andJ.P. Toennies, J. Chem. Phys. 124(2006)\n084307\n[4] A. Tamura, Phys. Rev. B 53(1996) 14475\n[5] D.M.Brink and S.Stringari, Z. Phys. D 15(1990) 257\n[6] K.K.Lehmann, J. Chem. Phys. 119(2003) 3336\n[7] K.K.Lehmann, J. Chem. Phys. 120(2003) 513\n[8] K.K.Lehmann and A.M.Doktor, Phys. Rev. Lett. 92(2004) 173401\n[9] L.D.Landau and E.M.Lifshitz, Fluid Mechanics , 2nd ed. (Butterworth-Heinemann,\nOxford, 1987), §62\n[10] H.A.Bethe, Rev. Mod. Phys. 9(1937) 69, §53\n[11] T.Ericson, Adv. Phys. 9(1960) 425\n[12] J.U.Andersen, E.Bonderup, and K.Hansen, J. Chem. Phys. 114(2001) 6518\n[13]Handbook of Mathematical Functions , ed. by M.Abramowitz and I.A.Stegun (Dover,\nNew York, 1972)\n[14] K.Hansen et al., to be published\n[15] T.Beyer and D.F.Swinehart, Comm. Assoc. Comput. Machines 16(1973) 379\n[16] To resolve the numerical comparison, we empoyed two more digit s on the leading\norder term in the exponential than written out in the text.\n[17] H. P. Baltes and E. R. Hilf, Spectra of Finite Systems (Bibliographisches Institut,\nMannheim, 1976)\n[18] H. P. Baltes and E. R. Hilf, Comput. Phys. Commun. 4(1972) 208\n[19] M.BrackandR.K.Bhaduri, SemiclassicalPhysics (Addison-Wesley, NewYork1997)\n[20] H. P. Baltes and E. R. Hilf, Solid State Commun. 12(1973) 369\n[21] A.D.B. Woods and R.A. Cowley, Rep. Prog. Phys. 36(1973) 1135\n[22] M.W. Reynolds, I.D. Setija, and G.V. Shlyapnikov, Phys. Rev. B 46(2001) 575\n14[23] H. Kim, K. Seo, B. Tabbert, and G. A. Williams, Europhys. Lett. 58(2002) 395\n[24] M. M. Salomaa and G. A. Williams, Phys. Rev. Lett. 47(1981) 1730\n[25] J. Tempere, I. F. Silvera, and J. T. Devreese, Phys. Rev. Lett. 87(2001) 275301\n151 10 100 10 -1 10 010 110 210 310 410 5\nε2\n -F [h ω0]\nT [h ω0]\nFigure1: The (negative of) the ripplon free energies, calculated wit h Eq.(13) (dotted line)\nand the summation in Eq.(5) (full line) which is exact apart from settin g the upper limit\nto infinity. The temperature corresponding to the energy of the lo west excitation, ε2, is\nindicated.\n160 2000 4000 6000 8000 10000 -1.0 -0.5 0.0 0.5 1.0 1.5 2.0 \n relative error \nE [h ω0]\nFigure 2: Comparison of the ripplon level densities calculated accord ing to Eq.(16) (full\nline) and Ref.[6] (dashed for E <2500, dotted line for E >2500). The fit in Ref.[6] was\nlimited to energies between 50 and 2500, in the reduced units used he re and is calculated\nas the derivative of the numerical fit to the integrated level densit y. The expressions have\nbeen divided by the exact Beyer-Swinehart result, causing the osc illatory behavior at low\nenergy, and the curves plotted are the logarithms of these ratios . The curve of Ref.[5]\n(not shown) is around 3.\n170 200 400 600 800 1000 1200 0100 200 300 400 500 600 700 800 \nExcitation energy [ ε]\n ln ρ(E)\n~\nFigure 3: Phonon level densities calculated according to the exact B eyer-Swinehart count\n(open circles), Eq.(33) (full line), and the level density derived fro m the bulk Debye heat\ncapacity, i.e. corresponding to Eq.(33) but including only the first te rm in the exponential\n(dashed line).\n180 50 100 150 200 250 300 350 400 -0.20 -0.15 -0.10 -0.05 0.00 0.05 0.10 0.15 0.20 \n ln( ρ) / ln( ρBS ) -1 \nE [ ε]\nFigure 4: A comparison of Eq.(33) and the exact-count phonon leve l density, showing\nessentially the relative difference in the entropy of the phonon syst em in the two calcula-\ntions.\n190 200 400 600 800 1000 020 40 60 80 100 120 \n \nT = 1K \n(if ripplons \nalone excited) ln( ρ(E) [(h ω0)-1 ]) \nE [h ω0]\nFigure 5: The convoluted level densities for phonons and ripplons fo r droplet size 103.\nThe numerical convolution is the full line, and the approximate result in Eq.(45) the,\nhardly discernible, dashed line. The level densities for ripplons alone ( dotted line) is\ngiven for reference. The arrow indicates the energy content of t he ripplon excitations at\na temperature of 1 K.\n20" }, { "title": "0711.2258v1.Density_operators_and_selective_measurements.pdf", "content": "arXiv:0711.2258v1 [math-ph] 14 Nov 2007Density operators and selective measurements.\nW/suppress lodzimierz M. Tulczyjew\nValle San Benedetto, 2\n62030 Monte Cavallo, Italy\nAssociated with\nDivision of Mathematical Methods in Physics\nUniversity of Warsaw\nHo˙ za 74, 00-682 Warszawa\nand\nIstituto Nazionale di Fisica Nucleare,\nSezione di Napoli\nComplesso Universitario di Monte Sant’Angelo\nVia Cinthia, 80126 Napoli, Italy\ntulczy@libero.it\n1. Introduction.\nIt is widely believed that statistical interpretation of quantum mech anics requires that density op-\nerators representing quantum states be normalized. We present a description of selective measurements\nin terms of density operators. The description is inspired by Schwing er’s Algebra of Microscopic Mea-\nsurements [1], (see also [2]). Density operators used are not norma lized. We do not know applications\nof density operators requiring normalization.\n2. Beams of particles.\nThe physical space is an affine space Mof dimension 3 modelled on a vector space V. There is\nan Euclidean metric tensor\ng:V→V∗. (1)\nWe consider beams of particles of mass mand constant energy Ein the direction of a unit vector\nz∈V. The internal states of the particles are elements of a unitary vec tor spaceUof dimension r\nover the field Cof complex numbers. The elements of Uarekets|u/an}bracketri}htand elements of the dual space\nU∗arebras/an}bracketle{ta|. The unitary structure establishes an antilinear isomorphism of UwithU∗assigning\nto each ket |u/an}bracketri}hta unique bra /an}bracketle{tu†|. The number\n/an}bracketle{tu1†|u2/an}bracketri}ht (2)\nis thescalar product of vectors |u1/an}bracketri}htand|u2/an}bracketri}ht\nWe introduce a sequence of points\nx0,x1,x2,...,x n (3)\nsatifying inequalities\n/an}bracketle{tg(z),xi−xi−1/an}bracketri}ht>0, (4)\nand a corresponding sequence of planes\nXi={x∈M;/an}bracketle{tg(z),x−xi/an}bracketri}ht= 0}. (5)\nIn the immediate neighbourhood of each plane Xithe beam is not subject to external interaction and\nis represented by a plane wave\n|ψi(x)/an}bracketri}ht=|Ai/an}bracketri}htexp (ik/an}bracketle{tg(z),x−x0/an}bracketri}ht+δi),|Ai/an}bracketri}ht ∈U, k =√\n2mE\n¯h(6)\nwith time dependence separated. The probability flux through a unit surface element of the plane Xi\nis expressed by\n¯hk\nm/an}bracketle{tAi†|Ai/an}bracketri}ht. (7)\n1Between the plane Xi−1and the plane Xithe beam passes through a selective device. Its action\non the wave function is described as the action of a linear transition o perator\nMi:U→U:|ψi−1(x)/an}bracketri}ht /ma√sto→ |ψi(x)/an}bracketri}ht=Mi|ψi−1(x)/an}bracketri}ht. (8)\nIf\nM1,M2,M3,...,M n (9)\nis the sequence of transition operators and\n|ψ0(x)/an}bracketri}ht=|A0/an}bracketri}htexp (ik/an}bracketle{tg(z),x−x0/an}bracketri}ht) (10)\nis the initial state, then\n|ψi(x)/an}bracketri}ht=|Ai/an}bracketri}htexp (ik/an}bracketle{tg(z),x−x0/an}bracketri}ht+δi)\n=Mi···M2M1|A0/an}bracketri}htexp (ik/an}bracketle{tg(z),x−x0/an}bracketri}ht)(11)\nand\n/an}bracketle{tψi†(x)|ψi(x)/an}bracketri}ht=/an}bracketle{tA0†|M1†M2†···Mi†Mi···M2M1|A0/an}bracketri}ht (12)\nThe flux of particles through unit surface element of Xiis given by\n¯hk\nm/an}bracketle{tA0†|M1†M2†···Mi†Mi···M2M1|A0/an}bracketri}ht. (13)\n3. Mixed states and density operators.\nThe expression (13) can be presented in the form\n¯hk\nmtr/parenleftbig\nMi···M2M1|A0/an}bracketri}ht/an}bracketle{tA0†|M1†M2†···Mi†/parenrightbig\n. (14)\nIn this new expression the pure initial state is represented by the d ensity operator |A0/an}bracketri}ht/an}bracketle{tA0†|and can\nbe replaced by a mixed state represented by a positive Hermitian den sity operator T. The expression\n¯hk\nmtr/parenleftbig\nMi···M2M1TM1†M2†···Mi†/parenrightbig\n(15)\nis the result.\n4. Selective measurements.\nWe set the density operator Tin the expression (15) equal to\nT=m\n¯hkrI, (16)\nwhereIis the identity operator. The operator Tis normalized in the sense that\n¯hk\nmtrT= 1 (17)\nalthough this normalization is of no importance. The expression\nPi=¯hk\nmtr/parenleftbig\nMi···M2M1TM1†M2†···Mi†/parenrightbig\n=1\nrtr/parenleftbig\nMi···M2M1M1†M2†···Mi†/parenrightbig\n(18)\nrepresents the probability of detecting a particle crossing a unit su rface element of the plane Xiin unit\ntime. The state of the initial beam emitted by a source at X0is totally mixed.\n2We want to describe the following experimental arrangement. The in itial beam emitted at X0\nundergoes a preliminary selection by a sequence of devices represe nted by the sequence of operators\nM1,M2,...,M j. (19)\nThe beam undergoes further selection passing through a sequenc e of devices represented by operators\nN1=Mj+1,N2=Mj+2,...,N n−j=Mn. (20)\nThe particles are detected at Xnby a non selective detector. After the preliminary selection the sta te\nof the beam is represented by the density operator\nM=MjMj−1···M1TM1†···Mj−1†Mj†=m\n¯hkrMjMj−1···M1M1†···Mj−1†Mj†(21)\nwith the probability of non selective detection\nPin=¯hk\nmtr/parenleftbig\nMjMj−1···M1TM1†···Mj−1†Mj†/parenrightbig\n=¯hk\nmtrM (22)\nThe beam arrives at Xnin a state represented by the density operator\nNn−jNn−j−1···N1MjMj−1···M1TM1†···Mj−1†Mj†N1†···Nn−j−1†Nn−j†(23)\nIt is detected with the probability\nPout=¯hk\nmtr/parenleftbig\nNn−jNn−j−1···N1MjMj−1···M1TM1†···Mj−1†Mj†N1†···Nn−j−1†Nn−j†/parenrightbig\n=¯hk\nmtr/parenleftbig\nN1†···Nn−j−1†Nn−j†Nn−jNn−j−1···N1MjMj−1···M1TM1†···Mj−1†Mj†/parenrightbig\n=¯hk\nmtr(NM )(24)\nwith\nN=N1†···Nn−j−1†Nn−j†Nn−jNn−j−1···N1. (25)\nThe density operator Ncharacterizes the selective detector. The probability Poutis measured at Xjby\nthe selective detector. This measurement is performed on the mixe d state represented by the operator\nM. The relative probability\nPout/Pin= tr(NM )/trM (26)\nshould be considered the result of the selectve measurement desc ribed. Arbitrary normalization can\nbe imposed on M. Normalization of Nwould distort the result of the measurement.\n5. An example.\nIn addition to the metric tensor\ng:V→V∗(27)\nwe introduce in the model space Vof the physical space Man orientation odefined as an equivalence\nclass of bases.\nWe analyse the internal states of a beam of particles of spin 1 /2. States of particles are represented\nby wave functions with values in an unitary space of complex dimension 2. The set of hermitian traceless\noperators in Uis a real vector space Sof dimension 3.\nThe trace tr ( ab) of a product is a non negative real number and the mapping\nS×S→R: (a,b)/ma√sto→tr (ab) (28)\n3is bilinear and symmetric. The spectrum of an operator a∈Sis a pair {α,−α}of real numbers and\nthe spectrum of the operator aais the set {α2,α2}. It follows that tr ( aa) = 0 if and only if a= 0. In\nconclusion we have a Euclidean scalar product\n(|) :S×S→R: (a,b)/ma√sto→(a|b) =1\n2tr (ab). (29)\nWe introduce the Pauli morphism\nσ:V→S. (30)\nThis morphism is an isometry such that the operator\n1\n2σ(w) :U→U (31)\nassociated with each unit vector w∈Vis thespin operator in the direction w. Its spectrum is the\nset{1/2,−1/2}and its eigenvectors represent states of the particle with spin 1 /2 and−1/2 in the\ndirection of w.\nWe introduce a number of operators in the space U:\n1) The projection operator\nK(w) =1\n2(I+σ(w)) (32)\nassociated with a unit vector w∈V. This operator projects onto the space of eigenstates of the\nspin operator 1 /2σ(w) corresponding to the eigenvalue 1 /2.\n2) Aphase shift operator\nD(δ) = exp(iδ)I. (33)\n3) Anattenuation operator\nR(ρ) = exp(−ρ/2)I. (34)\n4) A unitary unimodular operator\nG:U→U. (35)\nThis operator represents a rotation\nE:V→V (36)\nin the sense that\nGσ(w)G−1=σ(Ew). (37)\n5) The operator\nQ(w) =1\n2(I+σ(w)) (38)\nassociated with a vector w∈Vof norm/bardblw/bardbl /ne}ationslash= 1. This operator is not a projection operator.\nConsider a beam undergo a preliminary selection by devices represen ted byM1=K(w) and\nM2=D(δ). The vector wis orthogonal to the direction of the beam and the first of the devic es is\na Stern-Gerlach filter. It is accompanied by an unavoidable phase sh ift. The state prepared by these\ndevices is a pure state represented by the density operator\nM=M2M1TM1†M2†=m\n2¯hkK(w). (39)\nThe selective detector is composed of devices represented by ope ratorsN1=R(ρ),N2=D(δ′), and\nN3=K(w′). The attenuation R(ρ) my be due to the beam passing through a potential barrier. The\ndensity operator\nN=N1†N2†N3†N3N2N1= exp(−ρ)K(w′) (40)\n4represents the selective detector. The result of the selective me asurement is the relative probability\nPout/Pin= tr(NM )/trM=1\n2(1 + (w′|w)) (41)\nsince\ntrM= 1 (42)\nand\ntr(NM ) = tr (exp( −ρ)K(w′)K(w))\n=1\n4tr (exp(−ρ)(I+σ(w′))(I+σ(w)))\n=1\n4tr (exp(−ρ)(I+σ(w′) +σ(w) +σ(w′)σ(w)))\n=1\n2(1 + (w′|w))(43)\n6. References.\n[1] J. Schwinger, The Algebra of Microscopic Measurements , Proc. Natl. Acad. Sc. US, 45(1959)\n[2] F. A. Kaempffer, Concepts in Quantum Mechanics , Academic Press, New York and London (1965)\n5" }, { "title": "0711.3748v1.On_the_effect_of_weak_disorder_on_the_density_of_states_in_graphene.pdf", "content": "arXiv:0711.3748v1 [cond-mat.mes-hall] 23 Nov 2007On the effect of weak disorder on the density of states in graph ene\nBal´ azs D´ ora∗\nMax-Planck-Institut f¨ ur Physik Komplexer Systeme, N¨ oth nitzer Str. 38, 01187 Dresden, Germany\nKlaus Ziegler\nInstitut f¨ ur Physik, Universit¨ at Augsburg, D-86135 Augs burg, Germany\nPeter Thalmeier\nMax-Planck-Institut f¨ ur Chemische Physik fester Stoffe, 01 187 Dresden, Germany\n(Dated: October 31, 2018)\nThe effect of weak potential and bond disorder on the density o f states of graphene is studied. By\ncomparing the self-consistent non-crossing approximatio n on the honeycomb lattice with perturba-\ntion theory on the Dirac fermions, we conclude, that the line ar density of states of pure graphene\nchanges to a non-universal power-law, whose exponent depen ds on the strength of disorder like\n1-4g/√\n3πt2, withgthe variance of the Gaussian disorder, tthe hopping integral. This can result\nin a significant suppression of the exponent of the density of states in the weak-disorder limit. We\nargue, that even a non-linear density of states can result in a conductivity being proportional to the\nnumber of charge carriers, in accordance with experimental findings.\nPACS numbers: 81.05.Uw,71.10.-w,72.15.-v\nI. INTRODUCTION\nGraphene is a single sheet of carbon atoms with a honeycomb lattice, exhibiting interesting transport\nproperties1,2,3,4,5. These are ultimately connected to the low-energy quasiparticles o f graphene, i.e. two-dimensional\nDirac fermions. Its conductivity depends linearly on the carrier den sity, and reaches a universal value in the limit of\nvanishingcarrierdensity1,2. Theformerhas been explainedby the presenceofchargedimpurit ies, while the latter does\nnot allow a charged disorder6,7. Moreover, in the presence of a magnetic field, the half-integer qu antum Hall-effect is\nexplained in terms of the unusual Landau quantization and by the ex istence of zero energy Landau level2,3.\nThe density of states in pure graphene is linear around the particle- hole symmetric filling (called the Dirac point),\nand vanishes at the Dirac point. This is a common feature both in the la ttice description and in the continuum.\nIn addition, the lattice model also shows a logarithmic singularity at th e hopping energy, which is absent in the\ncontinuum or Dirac description.\nWhen disorder is present, the emerging picture is blurred. Field-the oretical approaches to related models (quasi-\nparticles in a d-wave superconductor) predict a power-law vanishin g with non-universal8or universal9exponent or a\ndiverging8density of states, depending on the type and strength of disorde r. Away from the Dirac point a power-law\nwith positive non-universal exponent is also supported by numerica l diagonalization of finite size systems10,11. At\nand near the Dirac point the behavior of the density of states is less clear. Some approaches favour a finite density of\nstates at the Dirac point12, whereas others predict a vanishing DOS8,9or an infinite DOS13. There is some agreement\nthat away from the Dirac point and for weak disorder the DOS behav es like a power law with positive exponent\nρ(E)∼ρ0|E|γ(γ >0). (1)\nThe purpose of the present paper is to investigate, how a non-univ ersal (therefore disorder dependent) power-law\nexponent (found numerically in Refs. 10,11) can emerge for weak dis order (compared to the bandwidth), and what\nits physical consequences are. We determine the exponent based on the comparison of the self-consistent non-crossing\napproximation on the honeycomb lattice and of the perturbative tr eatment of the Dirac Hamiltonian. The exponent\ndecreases linearly with disorder. Then, using this generally non-linea r density of states, we evaluate the conductivity\naway from the Dirac point by using the Einstein relation. We show, tha t based on the specific form of the diffusion\ncoefficient, this can result in a conductivity, depending linearly on the carrier concentration1, and in a mobility,\ndecreasing with increasing disorder. These are in accord with recen t experiment on K adsorbed graphene14. By\nvarying the K doping time, the impurity strength was controlled. The conductivity away from the Dirac point still\ndepends linearly on the charge carrier concentration, but its slope , the mobility decreases steadily with doping time.\nOur results apply to other systems with Dirac fermions such as the o rganic conductor15α-(BEDT-TTF) 2I3.2\nII. HONEYCOMB DISPERSION\nWe start with the Hamiltonian describing quasiparticles on the honeyc omb lattice, given by16,17:\nH0=h1σ1+h2σ2, (2)\nwhereσj’s are the Pauli matrices, representing the two sublattices. Here,\nh1=−t3/summationdisplay\nj=1cos(ajk), h 2=−t3/summationdisplay\nj=1sin(ajk), (3)\nwitha1=a(−√\n3/2,1/2),a2=a(0,−1) anda3=a(√\n3/2,1/2) pointing towards nearest neighbours on the hon-\neycomb lattice, athe lattice constant, tthe hopping integral. The resulting honeycomb dispersion is given by\n±/radicalbig\nh2\n1+h2\n2, which vanishes at six points in the Brillouin zone. To take scattering in to account, we consider the\nmutual coexistence of both Gaussian potential (on-site) disorde r (with matrix element Vo,r, satisfying ∝angbracketleftVo,r∝angbracketright= 0 and\nvariance ∝angbracketleftVo,rVo,r′∝angbracketrightg0=goδrr′) and bond disorder in only one direction (in addition to the uniform hop ping with\nmatrix element Vb,r, satisfying ∝angbracketleftVb,r∝angbracketright= 0 and variance ∝angbracketleftVb,rVb,r′∝angbracketright=gbδrr′), which is thought to describe reliably the\nmore complicated case of disorder on all bonds18. In graphene, ripples can represent the main source of disorder, and\nare approximated by random nearest-neighbour hopping rates, w hile potential disorder might only be relevant close\nto the Dirac point19. The corresponding term in the Hamiltonian is\nV=Vo,rσ0+Vb,rσ1, (4)\nwhich results in H=H0+V.\nFIG. 1: (Color online) A small fragment of the honeycomb latt ice is shown. The thick red lines denote the uni-directional bond\ndisorder, on-site disorder acts on the lattice points.\nWithout magnetic field, the self-energy for the Green’s function, w hich takes all non-crossing diagrams to every\norder into account (non-crossing approximation, NCA), can be fo und self-consistently from20\nΣ(iωn) =1\n1−(go+gb)G2\n[(go+gb)−(go−gb)2G2]G−G, (5)\nwhereiωnis the fermionic Matsubara frequency and\nG=G0[iωn−Σ(iωn)]. (6)\nHere,G0is the unperturbed local Green’s function on the honeycomb lattice given by\nG0(z) =Ac\n(2π)2/integraldisplayzd2k\nz2−t2[4cos(√\n3kx/2)cos(3ky/2)+2cos(√\n3kx)+3], (7)\nwhereAc= 3√\n3a2/2 is the area of the unit cell, and the integral runs over the hexagon al Brillouin zone with corners\ngiven by the condition h2\n1+h2\n2= 0. This can further be brought to a closed form using the results o f Ref. 21. On the\nother hand, in the continuum representation, using the Dirac Hamilt onian, the above Green’s function simplifies to\nG0(z) =2Acσ0\n(2π)2/integraldisplayd2k\nz+v(kxσ1+kyσ2)=−Aczσ0\n2πv2ln/parenleftbigg\n1−λ2\nz2/parenrightbigg\n, (8)3\nandv= 3ta/2. The cutoff λcan be found by requiringthe number ofstates in the Brillouin zoneto be preservedin the\nDirac case as well. This leads to λ=/radicalbig\nπ√\n3t. Another possible choice relies on the comparison of the low frequen cy\nparts of the Green’s function in the lattice and in the continuum limit, w hich reveals the presence of ln( ω/3t) terms.\nThis leads to λ= 3t, which coincides with the real bandwidth on the lattice. We are going t o use this form in the\nfollowing. The difference of the variances becomes important when c alculating the 2nd order correction (in variance)\nto the self-energy. It is clear from Eq. (5), that the same self-en ergy is found for pure potential or unidirectional bond\ndisorder. The effect of their coexistence is the strongest, when t hey possess the same variance. From this, the density\nof states follows as\nρ(ω) =−1\nπImG(ω+iǫ) (9)\nwithǫ→0+. Without disorder, we have the linear density of states ρ(ω≪t) =Ac|ω|/2πv2. At zero frequency, in\nthe limit of weak disorder, the self-energy is obtained as\nΣ(0) =−iλexp/parenleftbigg\n−πv2\nAc(go+gb)/parenrightbigg\n, (10)\nwhich translates into a residual density of states as\nρ(0) =λ\nπ(g0+gb)exp/parenleftbigg\n−πv2\nAc(go+gb)/parenrightbigg\n. (11)\nFrom this expression, weak disorder is defined by the condition go+gb≪t2. The exponential term indicates the\nhighly non-perturbative nature of density of states at the Dirac p oint: all orders of perturbation expansion vanish\nidentically at ω= 0.\nFrom Eq. (5) it is evident, that the interference of the mutual coe xistence of both on-site and bond disorder should\nbe the most pronounced when go=gb. The frequency dependence of the density of states on the hone ycomb lattice\ncan be obtained by the numerical solution of the self-consistency e quation, Eq. (5), and is shown in Fig. 2. For small\nfrequency and disorder, there is hardly any difference between pu re on-site or unidirectional bond disorder and their\ncoexistence. However, at higher energies and disorder strength , they start to deviate from each other. At ω=t,\nthe weak logarithmic divergence is washed out with increasing disorde r strength. Such features are absent from the\nDirac description, which concentrates on the low energy excitation s. Interestingly, for weak disorder, the residual\nDOS remains suppressed as suggested by Eq. (11), but the initial s lope in frequency changes. In order to determine,\nwhether the exponent or its coefficient or both change with disorde r, we perform a perturbation expansion in disorder\nstrength using the Dirac Hamiltonian to quantify the resulting densit y of states, and compare it to the numerical\nsolution of the self-consistent non-crossing approximation using t he honeycomb dispersion.\nIII. POWER-LAW EXPONENT\nThe expansion of the one-particle Green’s function in disorder at z=E+iǫleads to\nG(z) =G0+G0VG0+G0VG0VG0+..., (12)\nwhereV=Vo,rσ0+Vb,rσ1describes both Gaussian potential and unidirectional bond disorde r,G= (z−H0−V)−1\nandG0= (z−H0)−1, andHo=v(kxσ1+kyσ2) is the Dirac Hamiltonian. After averaging over disorder, we get\n∝angbracketleftGrr∝angbracketright=G0;rr+g/summationdisplay\nr′G0;rr′G0;r′r′G0;r′r+... (13)\nwithg=go+gb. The Green’s function G0is translational invariant and reads from Eq. (8) at real frequenc ies as\nG0;rr=Ac|E|\n2πv2/bracketleftbigg\nln/parenleftbiggλ2\nE2−1/parenrightbigg\n−iπ/bracketrightbigg\n(14)\nThis implies\n∝angbracketleftGrr∝angbracketright=G0;rr+g(G2\n0)rrG0;rr+o(g2) =G0;rr[1+g(G2\n0)rr]+o(g2) (15)4\n0 0.5 1 1.5 200.050.10.150.20.250.30.350.4\nPSfrag replacements\nω/ttρ(ω)\n00.511.522.533.5400.020.040.060.080.10.120.14\nPSfrag replacements\n(go+gb)/t2tρ(0)\nFIG. 2: (Color online) The density of states is shown in the le ft panel for pure on-site ( gb= 0) or unidirectional bond ( go= 0)\ndisorder (solid line) for ( go+gb)/t2= 0.1, 0.4, 0.8 and 1.6 with decreasing DOS at ω=t. The red dashed line represents\nthe coexisting bond and unidirectional bond disorder with g0=gb. The black dashed-dotted line denotes the free case with a\nlinear density of states at low energies, exhibiting a logar ithmic divergence at ω=tin the pure limit. The right panel shows\nthe residual density of states for on-site or unidirectiona l bond disorder (blue solid line) and their coexistence with g0=gb(red\ndashed line). The black dashed-dotted line denotes the appr oximate expression, Eq. (11), for weak disorder. For go+gb≤0.4t2,\nthe residual density of states is negligible.\n−0.5 −0.25 0.25 0 0.500.020.040.060.080.10.12\nPSfrag replacements\nω/ttρ(ω)\ngoorgb go=gb\n00.1 0.2 0.3 0.4 0.50.550.60.650.70.750.80.850.90.951\nPSfrag replacements\n(go+gb)/t2exponent ( γ)\nFIG. 3: (Color online) The low energy density of states is sho wn in the left panel for ( go+gb)/t2= 0.004, 0.1, 0.2, 0.3, 0.4 and\n0.5 from bottom to top, for pure on-site ( gb= 0) or unidirectional bond ( go= 0)disorder at positive energies (blue solid line),\nand for their coexistence at negative energies at go=gb(red dashed line). The black dashed-dotted line denotes the power-law\nfit asρ(ω) =ρ0+2ρ1(ω/t)γ. The green vertical dotted line separates the two parts. The right panel visualizes the exponents as\na function of the variance of the disorder, go+gb, for on-site or bond disorder (blue solid line) and their coe xistence ( go=gb).\nThe black dashed-dotted line denotes the result of perturba tion theory: γ= 1−4(go+gb)/√\n3πt2. As is seen, the agreement\nis excellent in the limit of weak disorder. Note, that gdenotes the variance of the disorder.\nMoreover, we have with Eq. (14)\n(G2\n0)rr=Ac\n(2πv2)2/integraldisplayd2k\n[z+v(kxσ1+kyσ2)]2=−∂G0;rr\n∂z=Ac\n2πv2/bracketleftbigg\nln(1−λ\nz2)−2λ2\nz2−λ2/bracketrightbigg\nσ0≈Ac\n2πv2/bracketleftbigg\n−2ln/parenleftbigg|E|\nλ/parenrightbigg\n−iπ/bracketrightbigg\n(16)5\nforλ≫ |E| ≫ǫ. Therefore, we obtain\n∝angbracketleftGrr∝angbracketright=G0;rr/bracketleftbigg\n1−Acg\nπv2ln/parenleftbigg|E|\nλ/parenrightbigg\n−igAc\n2v2/bracketrightbigg\n+o(g2). (17)\nFrom this, the density of states follows as\nρ(E) =−1\nπIm∝angbracketleftGrr∝angbracketright=AcE\n2πv2/bracketleftbigg\n1−2gAc\nπv2ln/parenleftbiggE\nλ/parenrightbigg/bracketrightbigg\n. (18)\nIf we further assume that (I) the density of states as a function ofEsatisfies a power law (inspired by Refs. 9,10,11)\nand (II) the disorder strength gis small, we can formally consider Eq. (18) as the lowest order expans ion in disorder,\nand sum it up to a scaling form as\nρ(E) =Acλ\n2πv2/parenleftbiggE\nλ/parenrightbigg1−(2gAc/πv2)\n. (19)\nHence, thissuggeststhatthelineardensityofstatesofpuregra phenechangesintoanon-universalpower-lawdepending\non the strength of the disorder as γ= 1−(4g/π√\n3t2). Note, that the exponent does not depend on the ambiguous\ncutoffλ. We mention that Eq. (18) might suggest other closed forms than E q. (19). By using the renormalization\ngroupproceduretoselectthemostdivergentdiagramsatagiveno rderg(similarlytoparquetsummationintheKondo\nproblem), one can sum it up as a geometrical series19,22. However, the resulting expression contains a singularity\naround|Σ(0)|(Eq. (10), playing the role of the Kondo temperature here), and is valid at high energies compared\nto|Σ(0)|as Eq. (18). To avoid such problems, we use a different scaling funct ion, suggested by the results of Refs.\n9,10,11.\nWe compare this expression to the numerical solution of the self-co nsistent non-crossing approximation in Fig. 3.\nTo extract the exponent, we fit the data with ρ(E) =ρ0+2ρ1(|E|/t)γ, and extract ρ0,1and the exponent γ. As can be\nseen in the left panel, the power-law fits are excellent in an extended frequency window up to t/4. This suggests that\nthis effect should also be observable experimentally as well. The obtain ed value of ρ0is negligibly small, as follows\nfrom Eq. (11). From the fits, we deduce the exponent and its coeffi cient, which is shown in the right panel. It agrees\nwell with the result of perturbation theory, Eq. (19) in the limit of we ak disorder. The suppression of the exponent is\nsignificant, and can be as big as 30-35% around ( go+gb)/t2∼0.4. Similar phenomenon has been observed for Dirac\nfermions on a square lattice in the presence of random hopping10,11, where disordered systems were studied by exact\ndiagonalization. The decreasing exponent with disorder agrees with our results.\nIV. CONDUCTIVITY FOR NON-LINEAR DENSITY OF STATES\nNow we turn to the discussion of the conductivity in graphene. A pos sible starting point is the Einstein relation23,\nwhich states for the conductivity\nσ=e2ρD, (20)\nwhereρis the density of states and Dis the diffusion coefficient, both at the Fermi energy EF. Assuming a general\npower-law density of states, as found above, we have ρ(E) =ρ1(E/λ)γ. In the weak-disorder limit and away from the\nDirac point E= 0, we can safely neglect any tiny residual value. Moreover, the diff usion coefficient in this case is of\nthe form D=D1E, which is validated from the Boltzmann approach in the presence of c harged impurities6. On the\nother hand, at the Dirac point there is a exponentially small density o f states and a finite non-zero diffusion coefficient\nD∝g/ρin the presence of uncorrelated bond disorder24such that the conductivity is of order 1, in units of e2/h. In\nthe following, however, we will concentrate on the regime away from the Dirac point. Putting these results together,\nwe find\nσ=e2ρ1D1Eγ+1\nF\nλγ. (21)\nThis can be simplified further by noticing that the total number of ch arge carriers, participating in electric transport,\ncan be expressed as\nn=/integraldisplayEF\n0ρ(E)dE=ρ1Eγ+1\nF\n(γ+1)λγ. (22)6\nBy inserting this back to Eq. (21), we can read off the conductivity a s\nσ=e2D1(γ+1)n. (23)\nFrom this we can draw several conclusions. First, it predicts that a way from the Dirac point, where our approach\npredicts a general power-law density of states, the conductivity varies linearly with the density of charge carriers, in\nagreement with experiments1. Second, the mobility of the carriers, which is the coefficient of the nlinear term in the\nconductivity, behaves as\nµ=e/parenleftbigg\n2−4g\nπ√\n3t2/parenrightbigg\nD1, (24)\nwhere we used our approximate expression for the exponent in the density of states, Eq. (19). This means, that\nwith increasing disorder ( g), the mobility decreases steadily, in agreement with recent experim ents on K adsorbed\ngraphene14. There, the graphene sample was doped by K, representing a sour ce of charged impurities. Nevertheless,\nthese centers also distort the local electronic environment, and a ct as bond and potential disorder as well. The\nobserved conductivity varied linearly with the carrier concentratio nn, similarly to Eq. (23). Moreover, the mobility\n(the slope of the nlinear term) decreased steadily with the doping time (and hence the im purity concentration),\nwhich, in our picture, corresponds to a reduction of the exponent γas well as the mobility, Eq. (24).\nTo study the properties close to the Dirac point we have to go beyon d the perturbative regime. Then we realize\nthat the density of states does not vanish at E= 0. As an approximation we add a small contribution near the Dirac\npoint\nρ(E) =ρ0δη(E)+ρ1/parenleftbiggE\nλ/parenrightbiggγ\n,\nin form of a soft Dirac Delta function\nδη(E) =1\nπη\nE2+η2(η >0).\nThis implies a particle density nwhich does not vanish at the Dirac point:\nn(EF) =/integraldisplayEF\n0ρ(E)dE≈ρ0+ρ1\n(γ+1)λγEγ+1\nF. (25)\nMoreover, the diffusion coefficient does neither diverge nor vanish a t the Dirac point18,24such that we can assume\nD(E) =D0δη(E)+D1E .\nFrom the Einstein relation we get the conductivity which provides an in terpolation between a behavior linear in n\naway from the Dirac point and a minimal conductivity at the Dirac point :\nσ∼e2\nh/braceleftBigg\nD0ρ0δ2\nη(EF) for EF∼0\nD1ρ1Eγ+1\nF/λγ∼(1+γ)nforEF≫0. (26)\nThis, together with Eq. (25), implies for EF≫0 the same behavior as in Eq. (23) with the mobility of Eq. (24).\nThe value of the minimal conductivity can be adjusted by choosing th e parameter ηproperly. Therefore Eq. (26)\nprovides us with a qualitative understanding of the conductivity in gr aphene for arbitrary carrier density.\nV. CONCLUSIONS\nWe have studied the effect of weak on-site and bond disorder on the density of states and conductivity of graphene.\nBy using the honeycomb dispersion, we determine the self-energy d ue to disorder in the self-consistent non-crossing\napproximation. The density of states at the Dirac point is filled in for a rbitrarily weak disorder. We investigate\nthe possibility of observing non-linear density of states away from t he Dirac point, motivated by numerical studies\non disordered Dirac fermionic systems. By comparing the results of non-crossing approximation on the honeycomb\nlattice to perturbation theory in the Dirac case, we conclude, that a disorder dependent exponent can account for\nthe evaluated density of states. The exponent decreases linearly with the variance for weak impurities. Then, by\nusing the obtained power-law DOS, we evaluate the conductivity awa y from the Dirac point through the Einstein\nrelation. We find, that this causes the conductivity to depend linear ly on the carries concentration by assuming that\nthe diffusion coefficient is linear in energy6, and the mobility decreases steadily with increasing disorder. These can\nalso be relevant for other systems with Dirac fermions15.7\nAcknowledgments\nWe acknowledge enlighting discussions with A. V´ anyolos. This work wa s supported by the Hungarian Scientific\nResearch Fund under grant number OTKA TS049881 and in part by t he Swedish Research Council.\n∗Electronic address: dora@pks.mpg.de\n1K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, M. I. Kats nelson, I. V. Grigorieva, S. V. Dubonos, and A. A. Firsov,\nNature438, 197 (2005).\n2A. K. Geim and K. S. Novoselov, Nature Materials 6, 183 (2007).\n3V. P. Gusynin and S. G. Sharapov, Phys. Rev. B 73, 245411 (2006).\n4E. McCann, K. Kechedzhi, V. I. Fal’ko, H. Suzuura, T. Ando, an d B. L. Altshuler, Phys. Rev. Lett. 97, 146805 (2006).\n5V. V. Cheianov and V. I. Fal’ko, Phys. Rev. Lett. 97, 226801 (2006).\n6K. Nomura and A. H. MacDonald, Phys. Rev. Lett. 96, 256602 (2006).\n7A. H. C. Neto, F. Guinea, N. M. R. Peres, K. S. Novoselov, and A. K. Geim, arXiv:cond-mat/0709.1163v1.\n8A. W. W. Ludwig, M. P. A. Fisher, R. Shankar, and G. Grinstein, Phys. Rev. B 50, 7526 (1994).\n9A. A. Nersesyan, A. M. Tsvelik, and F. Wenger, Phys. Rev. Lett .72, 2628 (1994).\n10Y. Morita and Y. Hatsugai, Phys. Rev. Lett. 79, 3728 (1997).\n11S. Ryu and Y. Hatsugai, Phys. Rev. B 65, 033301 (2001).\n12K. Ziegler, M. H. Hettler, and P. J. Hirschfeld, Phys. Rev. Le tt.77, 3013 (1996).\n13W. A. Atkinson, P. J. Hirschfeld, A. H. MacDonald, and K. Zieg ler, Phys. Rev. Lett. 85, 3926 (2000).\n14J. H. Chen, C. Jang, M. S. Fuhrer, E. D. Williams, and M. Ishiga mi, arXiv:0708.2408.\n15N. Tajima, S. Sugawara, M. Tamura, R. Kato, Y. Nishio, and K. K ajita, Europhys. Lett. 80, 47002 (2007).\n16G. W. Semenoff, Phys. Rev. Lett. 53, 2449 (1984).\n17N. M. R. Peres, F. Guinea, and A. H. Castro Neto, Phys. Rev. B 73, 125411 (2006).\n18K. Ziegler, arXiv:cond-mat/0703628.\n19P. M. Ostrovsky, I. V. Gornyi, and A. D. Mirlin, Phys. Rev. B 74, 235443 (2006).\n20A. V´ anyolos, B. D´ ora, K. Maki, and A. Virosztek, New J. Phys .9, 216 (2007).\n21T. Horiguchi, J. Math. Phys. 13, 1411 (1972).\n22I. L. Aleiner and K. B. Efetov, Phys. Rev. Lett. p. 236801 (200 6).\n23A. A. Abrikosov, Fundamentals of the Theory of Metals (North-Holland, Amsterdam, 1998).\n24K. Ziegler, Phys. Rev. Lett. 97, 266802 (2006)." }, { "title": "0712.1194v1.The_Hartree_Fock_ground_state_of_the_three_dimensional_electron_gas.pdf", "content": "arXiv:0712.1194v1 [cond-mat.str-el] 7 Dec 2007The Hartree-Fock ground state of the three-dimensional ele ctron gas\nShiwei Zhang1and D. M. Ceperley2\n1Department of Physics, College of William and Mary, William sburg, VA 23187, USA\n2NCSA and Department of Physics, University of Illinois at Ur bana-Champaign, Urbana, IL 61801, USA\nIn 1962, Overhauser showed that within Hartree-Fock (HF) th e electron gas is unstable to a spin\ndensity wave (SDW) instability. Determining the true HF gro und state has remained a challenge.\nUsing numerical calculations for finite systems and analyti c techniques, we study the HF ground\nstate of the 3D electron gas. At high density, we find broken sp in symmetry states with a nearly\nconstant charge density. Unlike previously discussed spin wave states, the observed wave vector of\nthe SDW is smaller than 2 kF. The broken-symmetry state originates from pairing instab ilities at\nthe Fermi surface, a model for which is proposed.\nPACS numbers: 71.10.Ca,71.10.-w,71.15.-m,75.30.Fv\nThe three-dimensional electron gas is one of the basic\nmodels of many-body physics, and has been investigated\nfor over 70 years [1, 2, 3, 4, 5, 6, 7]. As the simplest\nmodel system representing an itinerant metal, it consists\nof interacting electrons in a uniform neutralizing charge\nbackground, described by the Hamiltonian:\nH=−¯h2\n2m/summationdisplay\ni∇2\ni+1\n2/summationdisplay\ni/negationslash=je2\n|ri−rj|+constant ,(1)\nwhere the sums are over particle indices. Its properties\nare routinely used in density functional theory, e.g., in\nlocal density approximations, as a reference state in cal-\nculations of electronic structure of real materials [6].\nThe simplest approach to an interacting many-fermion\nsystem suchasthe electrongas(jellium) is the mean-field\nHartree-Fock (HF) method, which finds the Slater deter-\nminant wave function minimizing the variational energy.\nIn unpolarized jellium, the “conventional” solution is a\nparamagnetic state with spin symmetry, the restricted\nHF (rHF) solution in quantum chemistry.\nThe rHF solution, however,is notthe exactHF ground\nstate of jellium. In 1962, Overhauser [3] proved that the\nrHF solution is unstable with respect to spin and charge\nfluctuations at any density. The global minimum en-\nergy state within HF is a spontaneously broken symme-\ntry state. The properties of this global ground state have\nremained unknown [7]. This is surprising, given the fun-\ndamentalimportanceofboth theelectrongasandthe HF\napproach. The correlation energy of the homogeneous\nelectron gas is a commonly used fundamental concept,\nbut its definition is in terms of the HF energy of the\nelectron gas.\nIn this paper, we numerically find the HF ground state\nfor finite systems. Our motivation, aside from solving\nthis mathematical puzzle, was to understand the mech-\nanism for the broken symmetry state. Further, it was\nhoped that the solution would suggest candidate ground\nstates, for jellium or for other systems, that can then\nbe explored by accurate many-body approaches such as\nquantum Monte Carlo [5, 8]. We focus on high andmedium densities. An analytic approach is used to aug-\nment and extend the results to the thermodynamic limit.\nWe find that a different pairing instability characterizes\nthe high-density ground state.\nWe consider N(N↑=N↓=N/2) electrons in a cu-\nbic supercell of volume Ω = L3. The density is spec-\nified by the average distance between electrons: rs≡\n(3Ω/4πN)1/3/aB. We write a Slater determinant as\n|Φ∝an}b∇acket∇i}ht=|φ↑\n1,φ↑\n2,···,φ↑\nN↑∝an}b∇acket∇i}ht⊗|φ↓\n1,φ↓\n2,···,φ↓\nN↓∝an}b∇acket∇i}ht,(2)\nwith|φσ\nj∝an}b∇acket∇i}ht=/summationtext\nkcσ\nj(k)|k∝an}b∇acket∇i}htwhere|k∝an}b∇acket∇i}htis a plane-wave basis\nfunction. In the rHF solution, any |k∝an}b∇acket∇i}htwithk≡ |k| ≤kF\nis fully occupied, while all others are empty. Our basis\ncontains all plane-waves with k < kcut.\nTo find the global ground state, i.e., the unrestricted\nHF (uHF) solution, we use an iterative projection\n|Φ(m+1)∝an}b∇acket∇i}ht=e−τHHF(Φ(m))|Φ(m)∝an}b∇acket∇i}ht, (3)\nwhereHHF(Φ(m)) is the HF Hamiltonian, i.e., the mean-\nfield approximation of Eq. ( 1). The wave function re-\nmains a single Slater determinant in the projection [8].\nIfτis sufficiently small, the energy will decrease in each\nstep and the projection will converge to a HF solution as\nm→ ∞. To ensure that the solution is not a local mini-\nmum, we often start from multiple randominitial states\n|Φ(0)∝an}b∇acket∇i}htand verify that the same final state is reached.\nThe smallness of the energy scale relevant to the sym-\nmetry breaking at high density presents a difficulty;\nfinite-size effects can easily be larger than differences in\nenergy of different phases. As a consequence, the stable\nstructures vary wildly with N. There are subtle com-\nmensuration effects in both r-space (Wigner crystal) and\nk-space (spin waves). For example, in open-shell systems\nthere is always a uHF solution, since a broken symmetry\nstate can be formed in a partially filled shell to lower the\nexchange energy, with no cost to the kinetic or Hartree\nenergy (see below). In closed-shell systems, on the other\nhand, there seems to be a critical value of rc\ns(N), below\nwhich no uHF state exists for a given value of Nunder\nperiodic boundary condition (PBC).2\nTo breakthe shell structure, we impose twisted bound-\nary conditions [9] on the orbitals (and hence on the wave-\nfunction): φ(r+Lˆα) =ei2πθαφ(r) where α=x,y,z.\nThis applies a shift of kθ= 2π/vectorθ/Lto the plane-wave ba-\nsis./vectorθ= 0 corresponds to PBC, the Γ-point for solids.\nFor generic twist angles /vectorθ, the rHF solution is non-\ndegenerate.\nLetuswritetheHamiltonianinsecondquantizedform,\nomitting an overall constant:\nˆH=¯h2\n2m/summationdisplay\nσ,kk2c†\nk,σck,σ+1\n2Ω/summationdisplay\nΛ′4πe2\nQ2ˆV(Λ),(4)\nwherec†\nk,σandck,σare creation and annihilation op-\nerators. In the second term, Λ denotes the variables\n{k,k′,Q,σ,σ′},Qis a reciprocal lattice vector, the′on\nthe summation indicates Q∝ne}ationslash= 0, and\nˆV(Λ)≡c†\nk−Q,σc†\nk′+Q,σ′ck′,σ′ck,σ. (5)\nThe HF Hamiltonian ˆHHF(Φ(m)) needed in Eq. ( 3) is\nthe same as ˆH, but with ˆV(Λ) replaced by the linearized\nformˆVHF(Λ) = ˆv(Λ)−∝an}b∇acketle{tˆv(Λ)∝an}b∇acket∇i}ht/2, where\nˆv(Λ)≡2/bracketleftbig\n∝an}b∇acketle{t/summationtext\nk′′c†\nk′′,σ′ck′′−Q,σ′∝an}b∇acket∇i}htδk′,k+Q(6)\n−∝an}b∇acketle{tc†\nk′+Q,σ′ck+Q,σ′∝an}b∇acket∇i}htδσ,σ′/bracketrightbig\nc†\nk,σck′,σ.\nThe expectation ∝an}b∇acketle{t..∝an}b∇acket∇i}htis with respect to |Φ(m)∝an}b∇acket∇i}ht. The vari-\national energy is Ev(Φ)≡ ∝an}b∇acketle{tΦ|ˆH|Φ∝an}b∇acket∇i}ht/∝an}b∇acketle{tΦ|Φ∝an}b∇acket∇i}ht. We also\ncompute a “growth estimator” of the energy, Eg≡\n−ln[∝an}b∇acketle{tΦ(m+1)|Φ(m+1)∝an}b∇acket∇i}ht/∝an}b∇acketle{tΦ(m)|Φ(m)∝an}b∇acket∇i}ht]/2τ, in the projec-\ntion. At convergence, Ev(Φ(m)) =Ev(Φ(m+1)) =Eg,\nwhich means that |Φ(m)∝an}b∇acket∇i}htis an eigenfunction of ˆHHF.\nHence, the projection gives a true solution of the HF\nHamiltonian, not just a Slater determinant with a varia-\ntional energy lower than the rHF value.\nCalculations were carried out with a set of random /vectorθ\nand the results averaged and errors estimated. For each\nN, a fixed set of kθ-points were used at different values\nofrsto correlate the runs. At larger rs, less statistical\naccuracy is needed and fewer kθ-points were used. Typi-\ncally the plane-wave basis cutoff was set to kcut∼2-3kF,\ni.e., a kinetic energy cutoff of 5-10 EF. The resulting ba-\nsis set error is negligible for all but the largest rs(= 7).\nFast Fourier transform (FFT) techniques were used to\nspeed up each step in the projection [8], and the orbitals\nre-orthonormalized as necessary.\nThe top panel of Fig. 1shows the energy difference δE\nbetweentheuHFgroundstateandtherHFstate. As rsis\nreduced, the magnitude of δEdecreases before becoming\nnearly flat at high densities ( rs<3) for each value of N.\nThe energy lowering remains finite across rs, showing\nthat the broken symmetry uHF solution exists for all\ndensities, consistent with Overhauser’s proof [3]. At low\nrs,δEis roughly −10−4Ry in the large-sized systems.1 1.5 2 2.5 3 3.5 4-0.001-0.0008-0.0006-0.0004-0.00020δE\nN=54\nN=64\nN=66\nN=128\nN=246\nN=512\nN=528 (a)\n1 2 3 4 5 6 7\nrs (a.u.)-0.04-0.03-0.02-0.0100.010.02δE\n-0.002-0.00100.0010.002K\nVH\nVex\n(b)\nFIG. 1: (color online) Energy differences (in Ry) between the\nHF ground state and the rHF state. (a) The energy lowering\nper electron, δE= (E−ErHF)/N, vs.rsfor different values of\nN [10]. (b) The three components of δE, kinetic ( K), Hartree\n(VH), and exchange ( Vex), forN= 54 (solid lines) and N=\n66 (dashed lines). The inset is an enlargement of the low\nrs. Error bars are estimated from the results for various kθ-\npoints.\nCalculations for larger systems are needed to clarify its\nbehavior at small rs. The energy differences are very\nsmall compared to the rHF energy which, at N→ ∞, is\nErHF/N= (2.21/r2\ns−0.916/rs)Ry: the relativeenergy\nreduction vanishes as rs→0. From the bottom panel,\nwe see that the lower total energy of the uHF ground\nstate is achieved by reducing the exchange energy at the\ncost of increased kinetic energy (exciting electrons into\n|k∝an}b∇acket∇i}htstates with k > k F). The Hartree energy remains\nunchanged up to rs∼3.\nWe find that the momentum distribution in the uHF\nsolutionis spin-independent, i.e., n↑(k) =n↓(k) is anun-\nbroken symmetry in the broken symmetry ground state.\nThis is illustrated in Fig. 2forrs= 3, but holds in all\nour calculations when fully converged, at all rsandN.\nFigure2also shows that, as rsdecreases, the occu-\npancy of k > k Fstates becomes less pronounced, and\nsignificant modifications to the Fermi sphere become in-\ncreasinglyconfinedtotheimmediatevicinityoftheFermi\nsurface (FS). At small rs, such modifications tend to be\nprincipally single pairing states. An example is seen at\nrs= 1 inN= 54: two plane-wave vectors are involved,\nsuch that a pair of rHF orbitals |φ↑∝an}b∇acket∇i}ht=|φ↓∝an}b∇acket∇i}ht=|k∝an}b∇acket∇i}htbecome\n|φ/arrowbothv∝an}b∇acket∇i}ht=ck|k∝an}b∇acket∇i}ht±ck′|k′∝an}b∇acket∇i}ht (7)\nwherek≤kFandk′> kF, and|ck|2+|ck′|2= 1. Such a\npairing state by itself forms a linear spin-density wave3\n0.5 1 1.5 2 2.5 3\nk/kF00.20.40.60.81n(k)\nrs=1\nrs=2\nrs=3\nrs=3, dn\nrs=4\nrs=70.9 1 1.1 1.2\nk/kF-10-4010-4n(k) - nrHF(k)\nN=246\nN=54\nFIG. 2: (color online) Momentum distribution n(k) at dif-\nferentrsvs.k/kF. The main graph has N= 54, with\n/vectorθ= (−0.368,0.172,−0.364). The arrow indicates kF. The\ninset shows [ n(k)−nrHF(k)] atrs= 1 for N= 246, with\n/vectorθ= (−0.494,−0.425,0.144). Note the primary spike at kF,\nand the paired smaller spikes with one on each side of kF.\n(SDW), with constant charge density and unchanged\nHartree energy. As the inset in Fig. 2shows, additional\npairing states form in the uHF solution involving wave\nvectors further from the FS. Although their amplitudes\nbecome very small as rsis reduced, these are important\nto the true uHF ground state, as we discuss later.\nReal-space properties are examined in Fig. 3. The\nchargedensityis ρ(r) =n↑(r)+n↓(r)andthespindensity\nisσ(r) =n↑(r)−n↓(r). We measure their Fourier trans-\nforms, e.g., Sρ(q) =|/integraltext\nρ(r)eiq·rdr|2/N. Atrs= 7, the\nN= 54 system is an antiferromagnetic bcc Wigner crys-\ntal. Asrsdecreases, electrons become less localized and\nless particle-like. Fluctuation in the charge density be-\ncomes much smaller, as indicated by the rapid reduction\ninSρ(q). This is consistent with the vanishing Hartree\nenergy in Fig. 1. Although Sσ(q) also decreases with\nrs, it is much larger and spin symmetry remains bro-\nken. Atrs= 4, the SDW no longer has a bcc structure,\nand its symmetry between x,y, andzis broken. At\nsmallrs, the electrons are highly delocalized and wave-\nlike. Charge variations are effectively compensated for\nby spatial “double occupancy” of ↑and↓electrons, as in\nthe pairing state discussed in Fig. 2.\nWe measure the characteristic wave vector of the spin\norchargedensitywaveby q=|q|, whereqisthepeakpo-\nsition of S(q). A wave vector of qσ∼2kFseems to have\nalways been assumed in previous investigations of the\nSDW states [3, 7]. However, we find the maximum spin\nordering is at smaller wave vectors, as shown in Fig. 3\nforN= 54. Consistent results are seen for larger N, e.g.\natrs= 2,qσ/kF= 1.2(2), 1.4(2), 1.5(3), and 1 .5(2) for\nN= 66, 128, 246, and 528, respectively.\nWe next prove analytically that an SDW instability\nwhose wave vector is qσ<2kFindeed exists. Let us con-10-510-310-2100S(q)spin\ncharge\n0 1 2 3 4 567\nrs (a.u.)123q/kF5 10 15 20\n5101520\nFIG. 3: (color online) Left: Peak values and locations of the\nFourier transforms of spin and charge densities for differen t\nvalues of rsinN= 54. Errors are estimated from the values\nat different kθ-points. Note the logarithmic scale in S(q).\nRight: Contour plot of the spin density σ(x,y,z) for a slice\nparallel tothe x-yplane. The systemhas N= 512and rs= 1.\nTheq-vector [2 ¯6¯3] has the leading Sσ(q) value, followed by\n[33¯3], and then [4 ¯15] with Sσfour times smaller.\nsider a system of large but finite N, withkθ= 0. From\nthe rHF reference state, we create a broken symmetry\nstate with two pairing orbitals as in Eq. ( 7), using two\npoints at the FS, k(a highest occupied state) and k′(a\nlowest unoccupied state), as illustrated in Fig. 4. The\nenergy cost consists of kinetic and exchange terms [7]:\n∆K∼2|ck′|2¯h2\nmkF∆k (8)\n∆Vex∼ |ck′|2e2\nπ∆kln2kF\n∆k, (9)\nwherekF= 1/(αrs), withα= (4/9π)1/3. Our choice\nofkandk′gives: ∆ k=|k′| − |k| ∼(1/2lF)(2π/L) =\n1/(2l2\nFαrs), where lFis defined by kF≡lF(2π/L), i.e.,\nlF= (3/8π)1/3N1/3. For fixed rs, the exchange term\ndominates if Nis sufficiently large. Choosing ckandck′\nto be real and of O(1), we can write Eq. ( 9) as:\n∆Vkk′\nex∼e2\nπαrslnlF\nl2\nF. (10)\nWe now createa “satellite”pairingstate in the vicinity\nofkandk′,i.e., with k′−k=s′−sand|s−k| ∼ O(2π/L).\nWe choose the excitation amplitude to be of the particu-\nlar form: cs′∼1/(lnlF)1+δ(δ >0), and thereby cs∼1.\nUsing Eq. ( 9) and noting that the difference in the mag-\nnitude of the wave vectors, ∆ s≡ |s′| − |s|, is less than\nthe size of circles in Fig. 4, we obtain an upper bound to\nthe energy cost for creating the {s,s′}-pairing state\n∆Vss′\nex∼2e2\nπαrs1\nlF(lnlF)1+2δ. (11)\nThedecreaseinexchangeenergybecauseof“constructive\ninterference” between the two parallel pairs is:\n∆Vks\nex∼ −2ckck′cscs′4πe2\nL31\n|k−s|24\nk\nk's\ns'FS\nFS\nFIG. 4: (color online) Cartoon of pairing state with qσ<\n2kF. The primary pairing state {k,k′}is at the FS. The\nsatellite pairing state {s,s′}can be in the shaded areas,\nwithinO(2π/L) ofkandk′. Dashed arrow lines illustrate\nOverhauser[3, 7] pairing at 2 kF.\n∼ −e2\n2π2αrs1\nlF(lnlF)1+δ. (12)\nFor sufficiently large lF, i.e., a large enough system size,\n|∆Vks\nex|can always be made larger than the energy costs\nin Eqs. ( 10) and (11). Hence this is an SDW state with\nlower energy than the rHF state.\nThewavevectoroftheconstructedSDWis qσ=k′−k.\nAs Fig.4shows,qσdoes not need to be 2 kF. The angle\nbetween k′andqσ,θ, can range from 0 ( qσ= 2kF) to\nπ/2 (qσ= 0). As θincreases, more {k,k′}pairing states\nbecome available on the FS, while the number of possi-\nble satellite pairs, i.e., the volume of the shaded areas,\ndecreases. The optimal choice would be in between. In\nfact, as a crude estimate, the number of {k,k′}pairs is\n∝2πk2\nFsinθ, and the number of {s,s′}pairs for each is\n∝(π−θ). Maximizing their product gives qσ∼1.52kF,\nwhich is consistent with our data.\nIn previous approaches [3, 7], pairing is constructed\nfrom orbitals directly across the FS. The energy lowering\nis driven by interference between such pairs (dashed ar-\nrows in Fig. 4), which in our model belong to primary\npairing states. Our approach differs by including the\nsatellite pairing states. The interference between the pri-\nmary and satellite states is what makes a general qσpos-\nsible. This model is supported by the exact numerical\ndata in Fig. 2. Clearly, the true ground state goes be-\nyond this model: the energy will be further lowered by\nhaving more {s,s′}pairs and multiple {k,k′}states, etc.\nWe have shown that the uHF states at intermediate\nand high densities have nearly constant charge density.\nThey arewave-like,and arisefrom pairingbetween states\non the FS separated by distance qσ. The momentum\ndistribution is spin-independent. Its deviation from the\nFermi sea is increasingly confined to the vicinity of the\nFS asrs→0. The SDW wave vector is qσ∼1.5(2)kF,\nand its structures are determined primarily by the short-\nrange exchange potential [11]. In the rHF solution, the\nsizeofthe exchangeholeis rx≡2.34rs[6]. AlinearSDWwith a wave length (i.e., characteristic like-spin separa-\ntion) ofrxhas wave vector q= 1.40kF. Asrs→0, the\nnumerical results are sensitive to the detailed topology\nof the FS in the finite-size systems. The outcome of the\nSDWstructurecanvarygreatly,asitisadelicatebalance\nto optimize pairing among a small number of plane-wave\nstates at the FS that can participate. Our results indi-\ncate that, at high densities, the HF ground state tends\nto further break spatial symmetry and favor one or two\ndimensions. As illustrated in the right panel of Fig. 3,\nfor example, multiple ’pockets’ can coalesce into locally\norevenglobally(e.g., stripe-like)connectedstructures,in\ncontrastwiththeWignercrystalstatewhereeachpocket,\ncorresponding to one electron, is fully localized.\nTo conclude, we have determined the true HF ground\nstate for finite electron gas. Combining numerical and\nanalytic results, we have described the origin and char-\nacteristics of the broken symmetry state at high density,\nand the novel pairing mechanism that drives it.\nThis work was supported by NSF (DMR-0535529 and\nDMR-0404853) and ARO (48752PH). We are grateful\nto H. Krakauer for help with the plane-wave machinery.\nWe acknowledgeuseful discussions with H. Krakauerand\nR. M. Martin.\n[1]E. P. Wigner, Phys. Rev. 46, 1002, (1934); Trans. Fara-\nday Soc. 34678 (1938).\n[2]F. Bloch, Z. Phys. 57, 549 (1929).\n[3]A. W. Overhauser, Phys. Rev. Lett. 3, 414 (1959); Phys.\nRev.1281427 (1962).\n[4]J. R. Trail et. al.Phys. Rev. B 68, 045107 (2003).\n[5]D. Ceperley, Phys. Rev. B 18, 3126 (1978); D. M. Ceper-\nley and B. J. Alder, Phys. Rev. Lett. 45, 566 (1980);\nF. H. Zong et. al., Phys. Rev. E 66, 036703 (2002).\n[6]R. M. Martin, “Electronic Structure: basic theory and\npractical methods”, Cambridge University Press, 2004.\n[7]G. F. Giuliani and G. Vignale, “Quantum Theory of the\nElecton Liquid”, Cambridge University Press, 2005.\n[8]S. Zhang and H. Krakauer, Phys. Rev. Lett. 90, 136401\n(2003); M. Suewattana et. al., Phys. Rev. B 75, 245123\n(2007).\n[9]C. Linet. al., Phys. Rev. E 64, 016702 (2001).\n[10]Systems with up to N= 66 were projected to conver-\ngence following this approach. Some larger systems were\nquenched from a “frozen core” state, where Nfc(< N)\nelectrons are frozen in the rHF state while the remaining\nN−Nfcelectrons are active in the projection. Nfcis then\ngradually reduced in subsequent projections. This proce-\ndure can sometimes get “stuck” in a local minimum. It is\nthus possible that the energy gain in the larger systems\nis underestimated and is a lower bound.\n[11]The structure which maximizes the near-neighbor dis-\ntance (NND) between like-spin electrons is the close-\npacking fcc, which would lead to an NaCl-like SDW, but\nwith a characteristic NND of 2 .28rs, less than rx." }, { "title": "0801.0852v1.Statistics_of_local_density_of_states_in_the_Falicov_Kimball_model_with_local_disorder.pdf", "content": "arXiv:0801.0852v1 [cond-mat.str-el] 6 Jan 2008Statistics of local density of states in\nthe Falicov-Kimball model with local disorder\nMinh-Tien Tran\nAsia Pacific Center for Theoretical Physics, Pohang, Republ ic of Korea, and\nInstitute of Physics and Electronics, Vietnamese Academy o f Science and Technology, Hanoi, Vietnam.\nStatistics of the local density of states in the two-dimensi onal Falicov-Kimball model with local\ndisorder is studied by employing the statistical dynamical mean-field theory. Within the theory\nthe local density of states and its distributions are calcul ated through stochastic self-consistent\nequations. The most probable value of the local density of st ates is used to monitor the metal-\ninsulator transition driven by correlation and disorder. N onvanishing of the most probable value\nof the local density of states at the Fermi energy indicates t he existence of extended states in the\ntwo-dimensional disordered interacting system. It is also found that the most probable value of the\nlocal density of states exhibits a discontinuity when the sy stem crosses from extended states to the\nAnderson localization. A phase diagram is also presented.\nPACS numbers: 71.27.+a, 71.23.An, 71.30.+h, 71.10.Fd\nI. INTRODUCTION\nElectron interaction and disorder strongly influence\nthe properties of materials. In particular, the motion\nof charge carrier particles can be suppressed by Coulomb\ninteraction and disorder, and the suppression leads to\na metal-insulator transition (MIT). In the pure system\nwithout disorder the MIT can occur and is purely driven\nby electron correlations.1The transition is commonly re-\nferred to the Mott-Hubbard MIT. In the presence of dis-\norder, the MIT can occur even without electron interac-\ntions. The state of system changes from extended phase\nto the Anderson localization due to coherent backscat-\ntering from randomly distributed impurities.2The Mott-\nHubbard MIT is characterized by opening a gap of the\ndensity of states (DOS) at the Fermi energy, while the\nAnderson localization is a gapless insulator. At the An-\nderson localization the spectra change from continuous\nto dense discrete points. It is characterized by vanishing\nof the most probable value of the local density of states\n(LDOS).2,3,4The most probable value of the LDOS dis-\ncriminates between the metal and insulator phases. The\nimportance of the distribution ofthe LDOS in the proper\ndescription of the Anderson localization was stressed by\nAnderson.2,3,4The very distribution of the LDOS deter-\nmines the Anderson localization but not average quan-\ntities of the LDOS. From the distribution of the LDOS\none can detect both the vanishing of the most probable\nvalue of LDOS as well as the gap opening of the total\nDOS, thus the distribution of the LDOS is a valuable\ntool for determining the MIT driven by both disorder\nand correlation.\nIn recent years, a generalized Curie-Weiss mean-field\ntheory, the dynamical mean-field theory (DMFT) was\ndeveloped.5The DMFT essentially captures local tem-\nporal fluctuations. It has been widely applied to study\ncorrelated electron systems. The DMFT describes the\nMott-Hubbard MIT well.5However, it works with the\narithmetic average of the LDOS, and cannot determinethe Anderson localization in disordered systems. A sta-\ntistical variant of the DMFT, which is usually referred\nto the statistical DMFT, has been introduced to study\nsystems with both disorder and interaction.6,7It can\nbe viewed as the DMFT formulated in real space with\ngeneral inhomogeneous solutions.8Within the statistical\nDMFT, the self energy is a local function of frequency,\nbut it also depends on the site index. In the presence of\ndiagonaldisorder, the self energyis also diagonalrandom\nvariables, and gives additional dynamical random contri-\nbutions to disorder. It generates a set of self-consistent\nstochastic equations. The statistical DMFT essentially\ndeals with the LDOS, hence it is capable of studying\nthe Anderson-Mott-Hubbard MIT. In parallel with the\nstatistical DMFT, a typical medium theory (TMT) was\nalso introduced to study the Anderson localization.9The\nTMT is based on the DMFT too. However, instead of\nthe arithmetic average DOS, it works with the geomet-\nric average DOS. The geometric average DOS is incor-\nporated into the self-consistent stochastic DMFT equa-\ntions, which result into self-consistent equations of the\nDMFT fashion. The TMT is essentially a mean-field\ntheory of both disorder and correlation, while in the sta-\ntistical DMFT disorder is treated exactly, and only the\ncorrelation effects are treated in a mean-field manner.\nThe TMT was employed to study the Anderson-Mott-\nHubbard MIT in correlated electron systems with local\ndisorder.10,11\nIn the presence of disorder the LDOS forms a stochas-\ntic ensemble. The stochastic ensemble of the LDOS\nmusthavecharacteristicswhichdiscriminatebetween the\nmetallic and insulator phases, and one has to search the\ncharacteristics for determining the MIT. Examples for\nthe characteristics are the most probable value of the\nLDOS in the original Anderson theory of localization2,3,4\northegeometricaverageoftheLDOSintheTMT.9How-\never, nearby the critical point of the MIT, quantum fluc-\ntuations may induce outliers of the statistical description\nof the stochastic ensemble of the LDOS. For a random\nsample an estimator is called robust if it is insensitive to2\noutliers.12The most probable value, arithmetic average,\nmedian are the examples of robust estimator. The geo-\nmetric average is not a robust estimator because it does\nnot fulfil the linear property of the robust estimator.12\nFrom the point ofview ofrobust statistics the most prob-\nable value is a better estimator of a random sample than\nthe geometric average. Moreover, in general, the geo-\nmetric average DOS is not the most probable value of\nthe LDOS, hence it does not truly represent a typical\nDOS, although it is closer to the most probable value of\nthe LDOS than the arithmetic averageof the LDOS. The\ngeometric average DOS is sensitive to small values of the\nLDOS at individual sites, even when these values do not\nrepresentthemostprobablevalueoftheLDOS.Adecline\nof the geometric average DOS with increasing the disor-\nder strength does not necessarily imply the approach to\nthe Anderson localization.13Nevertheless, when the geo-\nmetric average is embedded into the self-consistent cycle\nof the DMFT, the whole method, i.e the TMT, can de-\nscribe the Anderson localization.9,10,11\nThe interplay between disorder and correlation in the\nMIT theory is a long standing problem. In the clean\nor noninteracting limits the MIT was well studied.5,14\nHowever, the correlation effects in disordered systems\nstill remain unclear. In particular, despite experiments\nfound evidences of the metallic behavior in the two di-\nmensional electron systems, in theory it is not clear how\nelectron correlations induce metallic phase in low dimen-\nsional disordered systems.15In the present paper we con-\nsider the interplay of disorder and short range interac-\ntion in the MIT. Usually, the short range interaction\nis modelled by the Hubbard interaction.16Here we take\nan alternative point of view. Rather than try to study\nthe Hubbard model we take a simpler model, the spin-\nless Falicov-Kimball model (FKM).17The relation of the\nFKM to the Hubbard model is analogous to the relation\nbetween the Ising and Heisenberg models of magnetism.\nThe FKM describes itinerant electrons interacting via a\nrepulsive contact potential with localized electrons (or\nions). It can also be viewed as a simplified Hubbard\nmodel where electrons with down spin are frozen and do\nnot hop. Certainly, within the TMT the phase diagram\nof the Anderson-Mott-Hubbard MIT in the FKM and\nthat in the Hubbard model share common features.10,11\nMoreover, the FKM exhibits a rich phase diagram. In\nhomogeneous phase it exhibits the Mott-Hubbard type\nof MIT, although the model does not describe the Fermi\nliquid picture. At low temperature different phases with\nlong-range order may exist depending on the doping and\ninteraction strength.18,19,20The FKM can also be incor-\nporated into different models to study various aspects\nof electron correlations, for instance the charge ordered\nferromagnetism in manganites.21,22,23In the disordered\nFKM we can study different realizations of the inter-\nplay of disorder and electron correlations. With local\ndisorder the FKM exhibits the Anderson-Mott-Hubbard\nMIT.11It will be studied in the present paper by em-\nploying the statistical DMFT. After solving of the self-consistent equations of the statistical DMFT we obtain\nthe distributions of the LDOS. We determine the most\nprobable value of the LDOS and use it to monitor the\nAnderson localization. We find extended states which\noccur in the region from weak to intermediate strengths\nof interaction and disorder. For intermediate values of\ndisorder or interaction there is a reentrance of the An-\nderson localization. At strong disorder the system is a\ngapless insulator, while at strong interaction the system\nis a Mott-Hubbard insulator. We also find at the cross-\ning point from extended states to localization the most\nprobable value of the LDOS exhibits a discontinuity.\nThe plan of the present paper is as follows. In Sec. II\nwe present the statistical DMFT through its application\nto the FKM with local disorder. Numerical results are\npresented in Sec. III. In Sec. IV conclusion and remarks\nare presented.\nII. STATISTICAL DYNAMICAL MEAN-FIELD\nTHEORY\nIn this section we describe the statistical DMFT\nthrough its application to the FKM with local disorder.\nThe Hamiltonian of the system reads\nH=/summationdisplay\ntijc†\nicj+/summationdisplay\niεic†\nici−µ/summationdisplay\nic†\nici\n+Ef/summationdisplay\nif†\nifi+U/summationdisplay\nic†\nicif†\nifi, (1)\nwherec†\ni(ci) andf†\ni(fi) is the creation (annihilation)\noperator of itinerant and localized electrons at site i, re-\nspectively. tijis the hopping integral of itinerant elec-\ntrons between site iandj. In the following we take into\naccount only nearest neighbor hopping, i.e., tij=−tfor\nnearest neighborsites, and tij= 0 otherwise. We will use\ntas the energy unit. Uis the local interaction of itiner-\nantandlocalizedelectrons. µisthechemicalpotentialfor\nitinerant electrons. It controls the electron density. Ef\nis the energy level of localized electrons. It also serves as\nthe chemical potential of localized electrons and controls\nthe density oflocalized electrons. In the following we will\nconsider only the symmetric half filling case. It turns out\nthat it is equivalent to µ=U/2 andEf=−U/2.18εi\nare independent random variables. They represent lo-\ncal disorder in the model. We will consider the random\nvariables with uniform distribution\nP(εi) =1\nWΘ/parenleftbiggW\n2−|εi|/parenrightbigg\n, (2)\nwhere Θ( x) is the step function, and Wrepresents the\ndisorder strength. When the disorder is absent, model\n(1) is the pure FKM. In homogeneous phase which oc-\ncurs at high temperature it exhibits a Mott-Hubbard\nMIT.18,19,20When the interactionisabsent( U= 0), itin-\nerant and localized electrons are decoupled, and the itin-\nerant electron part of model (1) represents the Anderson3\nmodel, and it would exhibit the Anderson localization.2\nThus when both disorder and interaction are present,\nmodel (1) would exhibit the complex Anderson-Mott-\nHubbard MIT transition at high temperature.11The\nFKM may be realized by loading two kinds of fermion\natomswith light and heavymassesin optical lattice. The\nlight atoms play the role of itinerant electrons, while the\nheavyatoms are kept immobile as the localized electrons.\nThe main idea of the statistical DMFT is to formulate\nthe DMFT of the system with a realization of disorder in\nrealspace.6Whenthedisorderisrealized,thelocalGreen\nfunction is not homogeneous anymore. In this case we\ncan adopt the inhomogeneous DMFT8for treating the\ninteraction part. Within the approach the disorder is\ntreated exactly, while the effects of electron interaction\nresult into the self energy which is self-consistently calcu-\nlated by the DMFT equations. However, the self energy\ndepends on the site index. The electron Green function\ncan be written in real space\nG(ω) = [G−1\n0(ω)−Σ(ω)]−1, (3)\nwhereΣ ij(ω)istheselfenergyoftheelectronGreenfunc-\ntionGij(ω) =/angbracketleft/angbracketleftci|c†\nj/angbracketright/angbracketrightω.G0(ω) is the noninteracting\nGreen function. For the FKM with a realization of dis-\norderG0(ω) = [ωδij−tij−εi+µδij]−1.εiis random\nvariables realized accordingly to the probability distribu-\ntion (2). Equation (3) is just the Dyson equation written\ninthematrixform. Within thestatisticalDMFT,theself\nenergy is approximated by a local function of frequency.\nHowever, this local function can vary from site to site,\ni.e.,\nΣij(ω) =δijΣi(ω). (4)\nThe approximation is strictly local. In infinite dimen-\nsions the self energy is purely local. For finite dimensions\nthe approximation neglects nonlocal correlations. The\nnonlocal correlations can be systematically incorporated\nby cluster extensions of the DMFT.24The site depen-\ndence of the self energy is generated via random vari-\nablesεi. As a consequence the self energy Σ i(ω) is also\nstochastic variables. With this feature the effective mean\nfield and the local Green function also are local stochas-\ntic variables. The Dyson equation (3) shows that the\nself energy gives additional local random contributions\nto disorder of the system. These contributions are due\nto both interaction and the interplay between interac-\ntion and disorder. However, in difference to the random\nvariables εithe contributions are dynamical. They take\ninto account temporal local quantum fluctuations gen-\nerated by interaction and disorder. They also broaden\nthe random energy levels generated by disorder. The self\nenergy Σ i(ω) is determined from an effective single site.\nOnce the effective single site is solved the self energy is\ncalculated by the Dyson equation\nΣi(ω) =G−1\ni(ω)−G−1\ni(ω), (5)\nwhereGi(ω) is the bare Green function of the effective\nsingle site and represents the effective mean field actingon sitei.Gi(ω) is the electron Green function of the\neffectivesinglesite. Theself-consistentconditionrequires\nthat the Green function Gi(ω) of the effective single site\nmust concisewith the local Green function of the original\nlattice. i.e.,\nGi(ω) =Gii(ω). (6)\nIn the Appendix we show the exact derivation of the self\nconsistentequationforinhomogeneoussystemsininfinite\ndimensions. Equations (3)-(6) form the self-consistent\nsystemofequationsforthelatticeGreenfunctionandthe\nselfenergy. They areprincipalequationsofthe statistical\nDMFT. Since the local Green function is stochastic vari-\nables, the self-consistent equations are naturally stochas-\ntic too. In Eq. (6) the right hand side is a functional of\nstochastic variables Gi(ω), thus the self-consistent con-\ndition (6) generates a stochastic chain of Gi(ω) via the\niteration process. The LDOS is defined as usually\nρi(ω) =−1\nπImGii(ω+iη),\nwhereη= 0+is an infinitely small positive number.\nFor the FKM the effective single site problem has the\nfollowing action\nSi[c†\ni,ci] =−/integraldisplayβ\n0dτdτ′c†\ni(τ)G−1\ni(τ−τ′)ci(τ′)\n+/integraldisplayβ\n0dτU(c†\nicif†\nifi)(τ)+βEff†\nifi,(7)\nwhereβ= 1/Tis the inverse of temperature. The parti-\ntion function corresponding to the action is\nZi= Trfi/integraldisplay\nDc†\niDcie−Si[c†\ni,ci]. (8)\nThispartitionfunctioncanbecalculatedexactly, because\nthetraceoverthelocalizedelectronsisindependentofthe\ndynamics of itinerant electrons. We obtain20\nZi= 2exp/bracketleftBig/summationdisplay\nnln/parenleftBigG−1\ni(iωn)\niωn/parenrightBig\neiωnη/bracketrightBig\n+\n2exp/bracketleftBig\n−βEf+/summationdisplay\nnln/parenleftBigG−1\ni(iωn)−U\niωn/parenrightBig\neiωnη/bracketrightBig\n,(9)\nwhereωn= (2n+1)πTis the Matsubara frequency. The\nGreen function can directly be calculated from the par-\ntition function. Without difficulty one obtains\nGi(iωn) =W0i\nG−1\ni(iωn)+W1i\nG−1\ni(iωn)−U,(10)\nwhereW1i=f(/tildewideEi),W0i= 1−W1i. Here f(x) =\n1/(exp(βx)+1) is the Fermi-Dirac distribution function,\nand\n/tildewideEi=Ef+T/summationdisplay\nnln/parenleftBig1\n1−UGi(iωn)/parenrightBig\neiωnη.(11)4\nNote that the weight factors W0i,W1iare not simply a\nnumber. They are functionals of the local Green func-\ntion. One can show that W1i=/angbracketleftf†\nifi/angbracketrightis the density\nof localized electrons at site i. So far we have obtained\nthe complete solution of the effective single site. This to-\ngether with the statistical DMFT equations (3)-(6) fully\ndeterminethedynamicsofitinerantelectronswithafixed\ndisorder realization. Within the statistical DMFT, the\ndisorder is treated exactly, while the correlation effects\nare taken into account through the mean field contribu-\ntions of the DMFT. Once the self-consistent equations\nof the statistical DMFT are solved we obtain the LDOS.\nWith many different realizations of disorder a data en-\nsembleofthe LDOS isobtained. The sizeofthe ensemble\ndepends on the lattice size and the number of disorder\nrealizations. From the ensemble of the LDOS we can de-\ntermine its probability distributions as well as the most\nprobable value of the LDOS. The most probable value of\nthe LDOS is used to monitor the MIT driven by disorder\nand correlation.\nThe present approach proposes a statistical local de-\nscription of the MIT in disordered interacting systems.\nThe statistical aspect is a proper description since it\nworks directly with the ensemble of LDOS and use\nthe most probable value of the LDOS to monitor the\nMIT.2,3,4The local treatment in the spirit of DMFT ne-\nglects nonlocal correlations. This weakness may be seri-\nous in low dimensional systems. However, for the two di-\nmensional FKM nonlocal correlations give nonsignificant\ncontributions to the DOS in the homogeneous phase.25\nThe DMFT calculations for the two-dimensional FKM\nalso show reasonable results.26Nevertheless, the statis-\ntical DMFT can be considered as a simplest approach\nincorporating both disorder and correlation to the MIT.\nIII. NUMERICAL RESULTS\nIn this section we present the numerical results of the\nstatistical DMFT equations. In general, the matrix in-\nversion in Eq. (3) can be performed only for a finite size\nlattice. We consider a two dimensional square lattice\nwith the linear size L. Thus, we perform numerical cal-\nculations for the lattice of size Land finite number Nd\nof disorder realizations. In particular numerical calcu-\nlations were performed for L= 12 and Nd= 100. We\ntake temperature T= 1 which is high enough for homo-\ngeneous phase and avoiding any long-range order. Cer-\ntainly, the homogeneous phase is insensitive to temper-\nature, however at low temperature it is unstable against\nthe long-range ordered phase.20The Mott-Hubbard like\nMIT in the FKM occursonly for the homogeneousphase.\nStrictly speaking, the Anderson localization is well de-\nfined only at zero temperature. Here we study the in-\nterplay between the Anderson localization and the Mott-\nHubbard MIT by investigating the single particle Green\nfunction at finite temperature. However, in the absence\nof long range orders the single particle Green function/s48/s46/s48/s48/s48/s46/s48/s53/s48/s46/s49/s48/s48/s46/s49/s53\n/s48/s53/s49/s48/s49/s53\n/s48/s46/s48/s48/s48/s46/s48/s53/s48/s46/s49/s48\n/s48/s53/s49/s48\n/s45/s49/s53 /s45/s49/s48 /s45/s53 /s48 /s53 /s49/s48 /s49/s53/s48/s46/s48/s48/s48/s46/s48/s51/s48/s46/s48/s54\n/s48/s53/s49/s48/s49/s53/s97/s40 /s41/s85/s61/s50\n/s87/s61/s52\n/s32/s124/s100\n/s97/s40 /s41/s47/s100 /s124/s32/s32\n/s97/s40 /s41\n/s32/s85/s61/s56\n/s87/s61/s52\n/s32/s124/s100\n/s97/s40 /s41/s47/s100 /s124\n/s85/s61/s56\n/s87/s61/s49/s48/s32\n/s97/s40 /s41\n/s32\n/s32/s124/s100\n/s97/s40 /s41/s47/s100 /s124\nFIG. 1: The total DOS ρa(ω) and its derivative |dρa(ω)/dω|\nfor various interactions and disorders.\nis insensitive to temperature. Hence, we can consider\nthe single particle Green function at finite temperature\nas it would be at zero temperature. The symmetric half\nfilling case with fixed µ=U/2 and localized electron\ndensitynf= 1/2 is considered. The energy level Efof\nlocalized electrons is determined accordingly by condi-\ntionnf=/summationtext\ni/angbracketleftf†\nifi/angbracketright/L2for each disorder realization. We\nsolvethe statisticalDMFT equationsinrealfrequencyby\niterations. The small positive number η= 0.01 is used.\nWithout disorder the FKM exhibits the Mott-Hubbard\nMIT with critical value Uc≈4. Thus when system is\nclean the system state is metallic for U < U c, and is\ninsulator for U > U c.\nA. Total DOS and band edge\nFirst we determine the band edge of the system. Usu-\nally, the band edge is determined from vanishing condi-\ntion of the total DOS. The total DOS is defined as the\narithmetic average of the LDOS\nρa(ω) =1\nNdL2/summationdisplay\ndisorder/summationdisplay\niρi(ω). (12)\nHowever, strictly speaking, in the numerical calcula-\ntions the total DOS never vanishes since η= 0.01 was\nused. In Fig. 1 we plot the total DOS and its deriva-\ntive|dρa(ω)/dω|. One can see that at the band edge the\ntotal DOS sharply changes and its derivative exhibits a5\n/s48/s49/s48/s50/s48/s51/s48\n/s48/s53/s49/s48\n/s48/s53/s49/s48\n/s48/s46/s48 /s48/s46/s49 /s48/s46/s50/s48/s53/s49/s48/s49/s53\n/s48/s46/s48 /s48/s46/s49 /s48/s46/s50 /s48/s46/s51/s48/s49/s48/s50/s48/s51/s48/s48/s53/s48/s49/s48/s48/s49/s53/s48\n/s32/s32\n/s87/s61/s50\n/s32/s32\n/s87/s61/s52/s32/s80/s40\n/s105/s41\n/s32/s87/s61/s54\n/s32 /s32/s87/s61/s56\n/s32\n/s105/s32/s87/s61/s49/s48\n/s32/s32\n/s32/s32\n/s105/s87/s61/s49/s50\nFIG. 2: (Color online) Probability distribution P(ρi) of the\nLDOSρiat the Fermi energy for various disorder Win the\nweak interaction case ( U= 2).\npronounced peak. We use the position of the peak to\ndetermine the band edge. The value of the total DOS\nat the band edge, ρbe, also serves as the cutoff of the\nDOS. Below this value ρbethe DOS approximately van-\nishes. For strong interaction and weak disorder, the total\nDOS opens a gap at the Fermi energy. It is similar to\nthe Mott-Hubbard insulator in the clean case. For such\ncases we also use the peaks of |dρa(ω)/dω|to determine\nthe band gap. For strong interaction and strong disor-\nder, the gap opened at the Fermi energy closes. As we\nwill see later, the system is still localized, but gapless. It\nis a crossover from the Mott-Hubbard to the Anderson\ninsulator by disorder.\nB. Anderson-Mott-Hubbard MIT\nThe probability distribution of the LDOS is con-\nstructed from statistical data of the LDOS which is ob-\ntained after solving the statistical DMFT equations. In\nFig. 2 we present the histogram of the probability dis-\ntribution of the LDOS at the Fermi energy for U= 2\nand various disorders. This value of interaction ( U= 2)\ncorresponds to the metallic phase in the clean limit.\nFor weak disorders the probability distribution has a\nmonomodal structure with a shaped peak at its most\nprobable value. In this regime the most probable value\nof the LDOS has a finite value, thus the system is in\nan extended state. This is an unambiguous evidence of/s48 /s53 /s49/s48 /s49/s53/s48/s46/s48/s48/s48/s46/s48/s52/s48/s46/s48/s56/s48/s46/s49/s50/s48/s46/s49/s54\n/s97\n/s103\n/s109/s112/s118\n/s32/s32\n/s87\nFIG. 3: The most probable value ρmpv, arithmetic average ρa\nand geometric average ρgof the LDOS at the Fermi energy\nvia disorder strength Win the weak interaction case ( U= 2).\nthe existence of extended states in the two-dimensional\ndisordered interacting system. As the disorder strength\nincreases the peak broadens, and then a second peak is\ndeveloped. Thus the probability distribution has a bi-\nmodal structure. The high value mode is nearly fixed\nalmost independently on the disorder strength. The low\nvalue mode moves towardsto zerovalue. Actually, as the\ndisorder strength increases the low value mode becomes\nthe most probable value of the LDOS, and it approxi-\nmately vanishes at some value of disorder. The vanishing\nofthe most probablevalueofthe LDOS is detected in the\nsense that its value is below the density cutoff ρbe. The\nvanishing of the most probable value of the LDOS mani-\nfeststhe Andersonlocalization. Thusthe systemexhibits\na MIT from extended state to the Anderson localization\nas the disorder strength increases. Perhaps, the bimodal\nstructure of the probability distribution of the LDOS is\nduetospecialfeaturesoftheFKM.Thelowvaluemodeis\ndueto the Andersonlocalizationwhen disorderincreases.\nThe high value mode reflectsthe non-Fermi liquid behav-\nior of the FKM in the weak interaction regime. For weak\ninteractions the chemical potential remains pinned at the\neffectivelevelof flocalizedelectrons.27In the presenceof\ndisorder most of the LDOS still persist with the pinning\nproperty, thus most of the LDOS at the chemical poten-\ntial keep the same value. The bimodal structure may be\nabsent in the systems where the Fermi liquid properties\nare maintained.\nIn principle, we can determine the most probablevalue\nas the value at which the probability distribution reaches\nits global maximum. However, the probability distribu-\ntion is constructed by histogram, its most probable value\nis sensitive to the width of histogram bars. To avoid the\ninaccuracy, we determine the most probable value by the\nhalf sample mode algorithm.12This algorithm is a fast\nroutine for locating the most probable value of a finite\nstatistical sample. The half sample mode algorithm is\nbased on finding the smallest interval that contains half\nnumber of the sample points. The most probable value6\n/s48/s52/s48/s56/s48/s49/s50/s48\n/s48/s49/s48/s50/s48/s51/s48\n/s48/s49/s48/s50/s48/s51/s48/s52/s48/s53/s48\n/s48/s46/s48/s48 /s48/s46/s48/s53/s48/s50/s48/s52/s48/s54/s48/s56/s48\n/s48/s46/s48/s48 /s48/s46/s48/s53 /s48/s46/s49/s48/s48/s50/s48/s52/s48/s54/s48/s56/s48/s49/s48/s48/s48/s49/s48/s48/s50/s48/s48/s51/s48/s48/s52/s48/s48/s53/s48/s48/s32\n/s87/s61/s52\n/s32/s32/s32\n/s32/s87/s61/s53\n/s32/s32/s80/s40\n/s105/s41/s87/s61/s54\n/s32/s32 /s32/s87/s61/s49/s48\n/s32\n/s105/s32/s87/s61/s49/s49\n/s32\n/s105/s32 /s32/s87/s61/s49/s50\nFIG. 4: (Color online) Probability distribution P(ρi) of the\nLDOSρiat the Fermi energy for various disorder Win the\nintermediate interaction case ( U= 6).\nmust lie in the obtained half sample. Repeat this half\nsample procedure until obtain the half sample with two\nor three sample points. Then one can easily locate the\nmost probable value of the sample. In Fig. 3 we plot the\nmost probable value as well as the arithmetic and geo-\nmetric averageof the LDOS at the Fermi energy for com-\nparison. It shows as the disorder strength increases the\nmost probable value shifts from the higher value mode to\nlowervalueone. Afteracriticalvalueofdisorderstrength\n(Wc≈11.1) the most probable value approximately van-\nishes, thus the system changes to the Anderson localized\nphase. At the crossing point the most probable value ex-\nhibits a discontinuity. The most probable value of the\nLDOS is not an order parameter of MIT, since it does\nnot associate with any symmetry breaking, but the dis-\ncontinuity of the most probable value of the LDOS may\nbe considered as a sign of the first order phase transi-\ntion. Figure 3 also shows that both the arithmetic and\ngeometric average of the LDOS never coincide with the\nmostprobablevalueoftheLDOS.Thearithmeticandge-\nometric averages of the LDOS monotonously decrease as\nthe disorder strength increases. However, the decreases\ndo not necessarily imply the approach to the Anderson\nlocalization.13\nIn Fig. 4 we present the histogram for the probability\ndistribution of the LDOS at the Fermi energy for various\ndisorders and U= 6. This value of interaction ( U= 6)\ncorresponds to the insulator phase in the clean limit. In\ndifference to the weak interaction case, at weak disorders\nthe probabilitydistribution ofthe LDOS at the Fermi en-ergy has almost a delta function like structure. It means\nthat almost all LDOS at the Fermi energy vanish. The\ntotal DOS opens a gap at the Fermi energy. It is anal-\nogous to the Mott-Hubbard insulator in the clean limit.\nAs the disorder strength increases, the delta peak broad-\nens, and the probability distribution of the LDOS shows\na long tail. However, the most probable value of the\nLDOS still approximately vanishes, while the arithmetic\nand geometric average are finite. The total DOS now\ncloses the opened gap at the Fermi energy. The system\nstate is still localized, however gapless. It is analogous to\nthe Anderson localized phase in the noninteracting limit.\nWe still refer it to the Anderson localization, although\nthe physics nature may be different. In general, at finite\ninteraction and finite disorder there are no precise def-\ninitions of the Mott-Hubbard and Anderson insulating\nphases.10These phases rigorously exist only in the clean\nor noninteracting limits. In disordered interacting sys-\ntems, both phases are characterized by vanishing of the\nmost probable value of the LDOS at the Fermi energy.\nHowever, the total DOS of the Mott-Hubbard insulator\nopens a gap at the Fermi energy, while the one of the\nAnderson localization is gapless. The scenario of clos-\ning the opened gap by disorder in the strong interaction\ncase can be understood from the atomic limit.28In the\nclean atomic limit each site has two energy levels, one\nsingle occupancy and one double occupancy, separated\nby interaction strength U. At the half filling, each site\nis occupied either by one itinerant electron or by one\nlocalized electron. The double occupancy levels remain\nempty, thus the charge gap is equal to U. When disorder\nis added, each of these two energy levels is shifted by ran-\ndomly fluctuating site energy −W/2< εi< W/2. For\nW < U the situation remains unchanged, thus there is\nstill a charge gap for electron excitations. When W > U,\nthe double occupancy levels at some sites may be shifted\nlower than the single occupancy levels. Thus the double\noccupancy levels of a fraction of sites are either occu-\npied or empty. As a result the charge gap is closed. The\nphase may be interpreted as a mixture of Anderson and\nMott-Hubbard insulators.\nIn contrast to the weak interaction case, where the\nprobability distribution of the LDOS has the bimodal\nstructure, in the intermediate and strong interaction\ncases the probability distribution of the LDOS keeps its\nmonomodal structure. As the disorder strength increases\nfurther, the most probable value of the LDOS first in-\ncreases from zero value and then decreases back to zero\nvalue. In the first stage the system changes from the lo-\ncalized state to extended state, while in the second stage\nthe system changes back from the extended state to lo-\ncalized state. It is a reentry effect of the Anderson local-\nization. In Fig. 5 we plot the most probable value, the\narithmetic and geometric average of the LDOS at the\nFermi energy as a function of the disorder strength for\nU= 6. In difference to the weak interaction case, as the\ndisorder strength increases, the most probable value and\nthe averages of the LDOS increase from zero value, reach7\n/s50 /s52 /s54 /s56 /s49/s48 /s49/s50 /s49/s52/s48/s46/s48/s48/s48/s46/s48/s49/s48/s46/s48/s50/s48/s46/s48/s51/s48/s46/s48/s52\n/s97\n/s103\n/s109/s112/s118\n/s32/s32\n/s87\nFIG. 5: The most probable value ρmpv, arithmetic average ρa\nandgeometric average ρgof theLDOSat theFermi energyvia\ndisorder strength Win the strong interaction case ( U= 6).\ntheirmaximum, andthendecrease. Italsoshowsthatthe\ngeometric average of the LDOS approximately vanishes\nonly in the Mott-Hubbard insulator phase. However in\nthis phase the arithmetic average and the most proba-\nble value of the LDOS approximately vanish too. In the\nAnderson localized phase only the most probable value\nof the LDOS vanishes. In the region of intermediate val-\nues of the disorder strength, the most probable value of\nthe LDOS is finite. This evidence unambiguously indi-\ncates the existence of extended states for intermediate\ndisorders. It also shows that disorder can drive the sys-\ntem from the insulating state to extended one. However,\nthe insulating state should be the gapless localized state.\nDisorder cannot drive a Mott-Hubbard insulator directly\nto extended state. One may speculate the MIT scenario\nof the intermediate interaction case as a screening of dis-\norder which leads to close the gap at the Fermi energy,\nand then the standard scenario of the Anderson localiza-\ntion as in the weak interaction case. At the transition\npoint the most probable value also shows a discontinuity\nlike in the weak interaction case. We can use the dis-\ncontinuity to detect the crossing point from extended to\nlocalized states.\nIn Fig. 6 we plot the phase diagram. It clearly dis-\ntinguishes three phase regions. Extended phase is those\nstates that the most probable value of the LDOS at the\nFermi energy is finite. The insulator phase is charac-\nterized by the vanishing of the the most probable value\nof the LDOS at the Fermi energy. This phase is sep-\narated into the Anderson localization where the total\nDOS is gapless and the Mott-Hubbard insulator where\nthe total DOS opens a gap at the Fermi energy. The\nAnderson localization occurs for strong disorder, while\nthe Mott-Hubbard insulator occurs for strong correla-\ntion. The extended phase appears only for weak and\nintermediate disorder and correlation. For a weak dis-\norder, as the interaction increases, the system changes\nfrom the extended phase to the Anderson localization,\nand finally to the Mott-Hubbard insulator. For an in-/s48 /s49 /s50 /s51 /s52 /s53 /s54 /s55 /s56/s48/s50/s52/s54/s56/s49/s48/s49/s50/s49/s52\n/s32/s32/s87\n/s85/s101/s120/s116/s101/s110/s100/s101/s100/s32/s112/s104/s97/s115/s101\n/s77/s111/s116/s116/s45/s72/s117/s98/s98/s97/s114/s100\n/s32/s32/s32/s32/s32/s32/s32/s112/s104/s97/s115/s101/s65/s110/s100/s101/s114/s115/s111/s110/s32/s108/s111/s99/s97/s108/s105/s122/s97/s116/s105/s111/s110\nFIG. 6: Phase diagram for the Mott-Hubbard-AndersonMIT.\nThe dotted line separates the phase region with finite gap in\nthe total DOS.\ntermediate value of disorder, as the interaction increases,\nthe system changesfrom the Anderson localizationto the\nextended phase, and then back again to the Anderson lo-\ncalization. This is a reentry effect of the Anderson local-\nization. For a weak interaction as the disorder strength\nincreases the system changes from extended to localized\nphase. For intermediate and strong interactions there is\na crossover from the Mott-Hubbard insulator to the An-\nderson insulator by closing the gap at the Fermi energy\nby disorder. For a fixed intermediate interaction there is\nalso the reentry effect of the Anderson localization when\nthe disorder strength is varied. The phase diagram is\nqualitatively analogous to the one calculated within the\nTMT.11Although the geometric mean alone does not in-\ndicate the Anderson localization, its fully embedding in\nthe self consistent cycle of the DMFT may describe the\nAnderson localization.9,10,11However, within the statis-\ntical DMFT there is a discontinuity of the most proba-\nble value of the LDOS at the phase boundary, whereas\nwithin the TMT the geometric average is continuous at\nthe phase boundary. The two limiting cases W= 0 and\nU= 0 are special. In the clean limit W= 0, there is\nthe Mott-Hubbard MIT, although the metallic phase is\nnot a Fermi liquid. The noninteracting limit U= 0 is\ncontroversial in two dimension. Our result agrees well\nwith the real space renormalization group calculations.29\nHowever,scalingtheorydid notfind anytruemetallicbe-\nhaviorin the system.30There is only a crossoverfrom ex-\nponentially to logarithmicallylocalized states. Certainly,\nthe present phase diagram is constructed from statistics\nof the LDOS, and the actual transport properties still\nremain unclear. Nevertheless, it was demonstrated that\nthe two dimensional disordered Hubbard model can have\na delocalizing effect.31,32,33,348\n/s48/s46/s48/s55/s48/s48/s46/s48/s55/s53\n/s53/s48 /s49/s48/s48 /s49/s53/s48 /s50/s48/s48/s48/s46/s48/s48/s48/s48/s46/s48/s48/s53/s48/s46/s48/s49/s48\n/s32/s32/s97\n/s32/s87/s61/s49/s48/s46/s53\n/s32/s87/s61/s49/s49/s46/s53\n/s32\n/s32/s78\n/s100/s32\n/s109/s112/s118\nFIG. 7: The statistical average of the most probable value\n(mpv) and the arithmetic average (a) of the LDOS at the\nFermi energy via the number of disorder realizations. The\nerror bars are their standard deviations. ( U= 2,L= 12,\nNb= 5).\nC. Finite size effects\nThe results of the previous subsections basically per-\nmit finite size effects. There are two sources of the finite\nsize effects. One is the finite size Lof the lattice, and\nthe second is the finite number Ndof disorder realiza-\ntions. First, we study the finite size effects of Nd. We\nfix the lattice size L= 12, and consider different num-\nbers of disorder realizations. For each Ndwe generate\nNbdifferent bins of size L2Ndfor disorder realizations.\nFor each bin we calculate the most probable value as well\nas the arithmetic average of the LDOS at the Fermi en-\nergy. Then we calculate their statistical average ραand\nstandard deviation σα, i.e.,\nρα=1\nNbNb/summationdisplay\nn=1ρ(n)\nα,\nσ2\nα=1\nNb−1Nb/summationdisplay\nn=1/parenleftbig\nρ(n)\nα−ρα/parenrightbig2,\nwhereαdenotes the most probable value (mpv) or the\narithmetic average (a). ρ(n)\nαis the most probable value\nor the arithmetic average of the LDOS, obtained from\nthe nth bin calculations. In Fig. 7 we plot the statistical\naverage and the standard deviation of the most proba-\nL810121416\nNd2401501108060\nN1568015000158401568015360\nTABLE I: The lattice size L, the number of disorder realiza-\ntionsNd, and the total number N, which are used for study\nof the finite size effects./s48/s46/s48/s54/s53/s48/s46/s48/s55/s48/s48/s46/s48/s55/s53\n/s56 /s49/s48 /s49/s50 /s49/s52 /s49/s54/s48/s46/s48/s48/s48/s48/s46/s48/s48/s53/s48/s46/s48/s49/s48/s32/s87/s61/s49/s48/s46/s53\n/s32/s87/s61/s49/s49/s46/s53\n/s32/s32/s97\n/s32\n/s32/s76/s32\n/s109/s112/s118\nFIG. 8: The most probable value ρmpvand the arithmetic\naverageρaof the LDOS at the Fermi energy via the lattice\nsizeL. The number of disorder realizations for each lattice\nsize is given in Table I. ( U= 2).\nble value or the arithmetic average of the LDOS at the\nFermi energy for various NdandNb= 5. We choose\ntwo values of disorder nearby the transition point. One\nis in the extended phase, and the other is in Anderson\nlocalized phase. Figure 7 shows the arithmetic average\nof the LDOS (the total DOS) has little fluctuations even\nforNd= 50. It almost independent on Ndalready from\nNd= 50. The most probable value of the LDOS fluctu-\nates in the extended phase more strongly than in the lo-\ncalized phase. Certainly, in the localized phase the most\nprobablevalueoftheLDOSapproximatelyvanishes, that\nthe fluctuations of the vanishing value is negligible when\nNdvaries. The most probable value of the LDOS seems\nto be reasonable from Nd= 100. Thus at Nd= 100 the\nfinite size effects of Ndare small and do not significantly\nchange the results.\nNext, we study the finite size effects of the lattice size.\nThe DMFT calculations for clean systems show the fi-\nnite site effects are small and controllable.8In disordered\nsystems, one can notice that the finite size effects of the\nensemble of the LDOS depend mostly on the size of the\nensemble, i.e., on the number N=L2Nd.Nis the total\nnumber of LDOS which are obtained in numerical cal-\nculations. Therefore to study the finite size effects of\nLalone, when the lattice size Lis varied, we have to\nchangeNdaccordingly, that the total number Nkeeps\nmore or less the same value. In Table I we present sev-\neral values of L, and corresponding values of Ndthat the\ntotal number Nis around 15000. We use the parame-\nter values in Table I for study of the finite size effects of\nL. For all cases the whole bin of disorder realizations is\nkept more or less the same. We also choose two values\nof disorder nearby the transition point. One is in the\nextended phase, and the other is in the localized phase.\nThe most probable value and the arithmetic average of\nthe LDOS at the Fermi energy via the lattice size are9\nplotted in Fig. 8. It shows that the arithmetic average\nof the LDOS is almost independent on the lattice size,\nat least from L= 8. The most probable value of the\nLDOS also slightly fluctuates as the lattice size varies.\nThese fluctuations mainly are due to the statistical fluc-\ntuations of finite size of disorder realizations. As we al-\nready showed in Fig. 7, the statistical fluctuations of the\nmost probable value of the LDOS in the extended phase\nis larger than in the localized phase. This feature is con-\nsistent with Fig. 8, where the most probable value of the\nLDOS fluctuates in the extended phase stronger than in\nthe localized phase. Both results show that the finite size\neffects in our study are small and do not change signifi-\ncantly the picture of MIT. It is interesting to note that\nthe DMFT can be performed for finite size lattices, and\nthe obtained results are not significantly different from\nthe ones of the thermodynamical limit.8\nIV. CONCLUSIONS\nIn this paper we have studied the Mott-Hubbard-\nAnderson MIT in the two-dimensional FKM with local\ndisorder by the statistical DMFT. Within the statisti-\ncal DMFT the correlation effects are resulted into addi-\ntional dynamical local random variables, which are self-\nconsistently determined from the local single site dynam-\nics. The probability distribution and the most probable\nvalue of the LDOS are calculated. The localized phase\nis detected by vanishing condition of the most probable\nvalue of the LDOS at the Fermi energy. The scenario\nof the MIT in the system depends on the interaction.\nFor weak interactions, which correspond to the metallic\nphase in the clean limit, the system changes from ex-\ntended to localized states as the disorder strength in-\ncreases. For intermediate interactions, which correspond\nto the insulating phase in the clean limit, as the disor-\nder strength increases the system crosses from the Mott-\nHubbard to the Anderson insulator, and then it transits\nto extended states and goes back again to the Anderson\nlocalized phase. Thus there is a reentrance of the An-\nderson localization. At the crossing point from extended\nto localized states the most probable value of the LDOS\nexhibits a discontinuity. For strong interactions only lo-\ncalized states exist. There is only a crossover from the\nMott-Hubbard to the Anderson insulator by closing the\nopened gap at the Fermi energy by disorder. The results\nalsoconfirmthedelocalizingeffect inthetwodimensional\ndisordered interacting system. However, the phase dia-\ngramwasdetermined only by statistics ofthe LDOS, and\nthe actual transport properties of the system still remain\nunclear. We leave the problem for further study.\nAcknowledgments\nThe author would like to thank the Asia Pacific Cen-\nter forTheoreticalPhysicsfor the hospitality. He alsoac-knowledgesuseful discussionswithHanyonChoiandJae-\njun Yu. The author is grateful to thank the Max Planck\nInstitute for the Physics of Complex Systems at Dresden\nfor sharing computer facilities where the numerical cal-\nculations were performed. This work was supported by\nthe Asia Pacific Center for Theoretical Physics, and by\nthe Vietnam National Program on Basic Research.\nAPPENDIX A: SELF CONSISTENT EQUATION\nOF THE INHOMOGENEOUS DMFT IN\nINFINITE DIMENSIONS\nIn this Appendix we present the derivation of the self\nconsistent equation of the inhomogeneous DMFT in infi-\nnite dimensions. Formally, we can derive the self consis-\ntentequationforinhomogeneoussystemsinthesameway\nas for homogeneous systems.5Certainly, the derivation\nfor homogeneous systems is based on the cavity method\nand the Hilbert transform.5Since the cavity method for\nhomogeneoussystems is also formulated in real space, we\ncan follow it closely. First a disorder realization is fixed,\nand then all fermions are traced out except for a single\nsitel. In the infinite dimensions we obtain the Green\nfunction which represents the effective mean field5\nG−1\nl(iωn) =iωn+µ−εl−∆l(iωn),(A1)\n∆l(iωn) =/summationdisplay\nijtliG(l)\nij(iωn)tjl, (A2)\nwhereG(l)\nij(iωn) is the Green function of the model with\nsitelremoved. ∆ l(iωn) can be considered as a hybridiza-\ntion function. The cavity Green function can be ex-\npressed through the original lattice Green function5\nG(l)\nij(iωn) =Gij(iωn)−Gil(iωn)Glj(iωn)\nGll(iωn).(A3)\nInserting Eq. (A3) into Eq. (A2) one obtains\n∆l(iωn) =/bracketleftbig\nt·G(iωn)·t/bracketrightbig\nll\n−/bracketleftbig\nt·G(iωn)/bracketrightbig\nll/bracketleftbig\nG(iωn)·t/bracketrightbig\nll\nGll(iωn),(A4)\nwheretisthehoppingmatrix. ThelatticeGreenfunction\ncan be rewritten as\nG(iωn) =/bracketleftbig\nξ(iωn)−t−Σ(iωn)/bracketrightbig−1,(A5)\nwhereξij(iωn) = (iωn+µ−εi)δij. In infinite dimensions\nthe self energy is purely local,5hence the self energy ma-\ntrixΣ(iωn) is diagonal. For homogeneous systems the\nterms in the right hand side of Eq. (A4) are calculated\nby using the Fourier and Hilbert transforms.5For inho-\nmogeneous systems the Fourier and Hilbert transforms\nare replaced by the matrix multiplication and inversion\nin real space. One can notice that\nt·G·t=t·G·/bracketleftbig\nξ−Σ−G−1/bracketrightbig10\n=t·G·/parenleftbig\nξ−Σ/parenrightbig\n−t, (A6)\nt·G=/bracketleftbig\nξ−Σ−G−1/bracketrightbig\n·G\n=/parenleftbig\nξ−Σ/parenrightbig\n·G−1. (A7)\nUsing relations (A6)-(A7), from Eq. (A4) we obtain\n∆l(iωn) =−/bracketleftbig\nξ(iωn)−Σ(iωn)/bracketrightbig\nll\n+/bracketleftBig/parenleftbig\nξ(iωn)−Σ(iωn)/parenrightbig\n·G(iωn)·/parenleftbig\nξ(iωn)−Σ(iωn)/parenrightbig/bracketrightBig\nll\n−/bracketleftbig/parenleftbig\nξ(iωn)−Σ(iωn)/parenrightbig\n·G(iωn)−1/bracketrightbig\nll/bracketleftbig\nG(iωn)·/parenleftbig\nξ(iωn)−Σ(iωn)/parenrightbig\n−1/bracketrightbig\nll/Gll(iωn).(A8)\nHere we have used tll= 0. Since both matrices ξ(iωn)\nandΣ(iωn) are diagonal, from Eqs. (A1) and (A8) we\nobtain\nG−1\nl(iωn) = Σl(iωn)+1/Gll(iωn).(A9)Equation(A9)isjusttheselfconsistentequations(5)-(6).\nIt is exact in infinite dimensions for any inhomogeneous\nsystem. The above derivation of the self consistent equa-\ntion can be considered as a generalization of the homo-\ngeneous one. In infinite dimensions, the sum in Eq. (A2)\nruns over infinitely many neighbouring sites, that the hy-\nbridization function is replaced by its average value. As\na consequence, all spatial fluctuations ofthe environment\nsurrounding a given site are suppressed, thus the Ander-\nson localization is prohibited.7At infinite dimensions the\nAnderson localization is absent in any disordered inter-\nacting system.\n1N. F. Mott, Metal-Insulator Transition , 2nd ed. (Taylor\nand Francis, London, 1990).\n2P. W. Anderson, Phys. Rev. 109, 1492 (1958).\n3P. W. Anderson, Rev. Mod. Phys. 50, 191 (1978).\n4R. Abou-Chacra, P. W. Anderson, and D. J. Thouless, J.\nPhys. C 6 1734 (1973).\n5A. Georges, G. Kotliar, W. Krauth, and M. J. Rozenberg,\nRev. Mod. Phys. 68, 13 (1996).\n6V. Dobrosavljevic and G. Kotliar, Phys. Rev. Lett. 78,\n3943 (1997).\n7V. Dobrosavljevic and G. Kotliar, Phil. Trans. R. Soc.\nLond. A 356, 57 (1998).\n8Minh-Tien Tran, Phys. Rev. B 73, 205110 (2006).\n9V. Dobrosavljevic, A. A. Pastor, and B. K. Nikolic, Euro-\nphys. Lett. 62, 76 (2003).\n10K. Byczuk, W. Hofstetter, and D. Vollhardt, Phys. Rev.\nLett.94, 056404 (2005).\n11K. Byczuk, Phys. Rev. B 71, 205105 (2005).\n12D. R. Bickel and R. F. Fruhwirth, Comp. Stat. & Data\nAnalysis 50, 3500 (2006).\n13Y. Song, W. A. Atkinson, and R. Wortis, Phys. Rev. B 76,\n045105 (2007).\n14P. A. Lee and T. V. Ramakrishnan, Rev. Mod. Phys. 57,\n287 (1985).\n15E. Abrahams, S. V. Kravchenko, and M. P. Sarachik, Rev.\nMod. Phys. 73, 251 (2001).\n16J. Hubbard, Proc. Roy. Soc. (London) A 276, 238 (1963).\n17L. M. Falicov and J. C. Kimball, Phys. Rev. Lett. 22997\n(1969).\n18T. Kennedy, Rev. Math. Phys. 6, 901 (1994).19J. K. Freericks, E. H. Lieb, and D. Ueltschi, Phys. Rev.\nLett.88, 106401 (2002).\n20J. K. Freericks and V. Zlatic, Rev. Mod. Phys. 75, 1333\n(2003).\n21Tran Minh-Tien, Phys. Rev. B. 67, 144404 (2003).\n22Van-Nham Phan and Minh-Tien Tran, Phys. Rev. B 72,\n214418 (2005).\n23T. V. Ramakrishnan, H. R. Krishnamurthy, S. R. Hassan,\nand G. Venketeswara Pai, Phys. Rev. Lett. 92, 157203\n(2004).\n24T. Maier, M. Jarrell, T. Pruschke, and M. H. Hettler, Rev.\nMod. Phys. 77, 1027 (2005).\n25M. H. Hettler, M. Mukherjee, M. Jarrell, and H. R. Krish-\nnamurthy, Phys. Rev. B 61, 12739 (2000).\n26J. K. Freericks, Phys. Rev. B 48, 14797 (1993).\n27Q. Si, G. Kotliar, and A. Georges, Phys. Rev. B 46, 1261\n(1992).\n28M. C. O. Aguiar, V. Dobrosavljevic, E. Abrahams, and\nG. Kotliar, Phys. Rev. B 73, 115117 (2006).\n29P. A. Lee, Phys. Rev. Lett. 42, 1492 (1979).\n30E. Abrahams, P. W. Anderson, D. C. Licciardello, and\nT. V. Ramakrishnan, Phys. Rev. Lett. 42, 673 (1979).\n31P. J. H. Denteneer, R. T. Scalettar, and N. Trivedi, Phys.\nRev. Lett. 83, 4610 (1999).\n32I. F. Herbut, Phys. Rev. B 63, 113102 (2001).\n33B. Srinivasan, G. Benenti, and D. L. Shepelyansky, Phys.\nRev. B67, 205112 (2003).\n34D. Heidarian and N. Trivedi, Phys. Rev. Lett. 93, 126401\n(2004)." }, { "title": "0801.3131v1.Density_induced_suppression_of_the_alpha_particle_condensate_in_nuclear_matter_and_the_structure_of_alpha_cluster_states_in_nuclei.pdf", "content": "arXiv:0801.3131v1 [nucl-th] 21 Jan 2008Density-induced suppression of the α-particle condensate in nuclear matter and the\nstructure of αcluster states in nuclei\nY. Funaki1, H. Horiuchi2, G. R¨ opke3, P. Schuck4,5, A. Tohsaki2, T. Yamada6\n1Nishina Center for Accelerator-Based Science, The Institu te of Physical\nand Chemical Research (RIKEN), Wako, Saitama 351-0198, Jap an\n2Research Center for Nuclear Physics (RCNP), Osaka Universi ty, Osaka 567-0047, Japan\n3Institut f¨ ur Physik, Universit¨ at Rostock, D-18051 Rosto ck, Germany\n4Institut de Physique Nucl´ eaire, 91406 Orsay Cedex, France\n5Universit´ e Paris-Sud, F-91406 Orsay-C´ edex, France and\n6Laboratory of Physics, Kanto Gakuin University, Yokohama 2 36-8501, Japan\nAt low densities, with decreasing temperatures, in symmetr ic nuclear matter α-particles are\nformed, which eventually give raise to a quantum condensate with four-nucleon α-like correlations\n(quartetting). Starting with a model of α-matter, where undistorted αparticles interact via an\neffective interaction such as the Ali-Bodmer potential, the suppression of the condensate fraction at\nzero temperature with increasing density is considered. Us ing a Jastrow-Feenberg approach, it is\nfound that the condensate fraction vanishes near saturatio n density. Additionally, the modification\nof the internal state of the αparticle due to medium effects will further reduce the conden sate.\nIn finite systems, an enhancement of the Sstate wave function of the c.o.m. orbital of αparticle\nmotion is consideredas thecorrespondence tothecondensat e. Wavefunctions havebeenconstructed\nfor self-conjugate 4 nnuclei which describe the condensate state, but are fully an tisymmetrized on\nthe nucleonic level. These condensate-like cluster wave fu nctions have been successfully applied to\ndescribe properties of low-density states near the nαthreshold. Comparison with OCM calculations\nin12C and16O shows strong enhancement of the occupation of the S-state c.o.m. orbital of the\nα-particles. This enhancement is decreasing if the baryon de nsity increases, similar to the density-\ninduced suppression of the condensate fraction in αmatter. The ground states of12C and16O show\nno enhancement at all, thus a quartetting condensate cannot be formed at saturation densities.\nI. INTRODUCTION\nThe properties of nuclear matter at very low densities\nand low temperatures are dominated by the formation\nof clusters, in particular αparticles. As a well-known\nconcept,αmatter has been introduced where symmetric\nnuclear matter is described by a system of αparticles,\nweakly interacting via effective α-αpotentials fitted\nto the scattering phase shifts, such as the Ali-Bodmer\ninteraction potential [1, 2, 3].\nThis concept becomes less valid with increasing den-\nsity. First, at finite temperatures other correlations and\nalso single nucleon states appear so that we have a mix-\nture of different constituents, described in chemical equi-\nlibrium by a mass action law. Secondly, at higher densi-\nties the internal fermionic structure of the αparticles be-\ncomes of relevance so that the four-nucleon bound state\nwill be modified by medium effects. A consistent ap-\nproach can be given by quantum statistical methods [4].\nUsingthermodynamic Greenfunctions, the effects ofself-\nenergyand Pauli blockingare included so that the bound\nstates are dissolved when the density exceeds a critical\nvalue. For αparticles this critical density, which is also\ndependent on temperature, is about ρ0/5, withρ0= 0.17\nfm−3as the saturation density [5].\nAn important phenomenon is the formation of a quan-\ntum condensate with strong four nucleon correlations at\nlow temperatures [6]. At low densities where αparticles\nare well defined weakly interacting constituents of sym-\nmetric nuclear matter, we have Bose-Einstein condensa-tion ofαparticles. With increasing density, quartetting\noccurs with medium-modified αparticles and disappears\nat a density of about ρ0/3. Note that quartet condensa-\ntion has recently also been considered in the context of\ncold atom physics [7].\nThe Bose-Einstein condensation for ideal quantum\ngases is a well-known phenomenon. The occupation of\nsingle-particle states is given by the Bose distribution\nfunction. Below a critical temperature Tc, to obey nor-\nmalization, the state of lowest energy is macroscopically\noccupied. This macroscopically enhanced coherent occu-\npation of the lowest quantum state is denoted as quan-\ntum condensate. As well known, the fraction of bosons\nfound in the condensate results for the ideal Bose gas as\nncond/n= 1−(T/Tc)3/2.\nHowever, this simple picture is no longer valid, if inter-\naction is taken into account. For a recent determination\nofTcin the interacting case, see Ref. [8]. Here, we want\nto concentrate on interaction effects at zero temperature.\nIn general, the condensate fraction is given by the prop-\nerties of the density matrix which contains a part which\nfactorizes. According to Penrose and Onsager [9], the\nquantum condensate in a homogeneousinteracting boson\nsystem at zero temperature is given by the off-diagonal\nlong-rangeorderinthe densitymatrix. Thenon-diagonal\ndensity matrix in coordinate representation can be fac-\ntorized so that in the limit |r−r′| → ∞follows\nlim\n|r−r′|→∞ρ(r,r′) =ψ∗\n0(r)ψ0(r′)+γ(r−r′).(1)\nThe last contribution γ(r) disappears at large distances,2\nwhereas the first contribution determines the condensate\nfraction in infinite matter as\nn0=∝an}bracketle{tΨ|a†\n0a0|Ψ∝an}bracketri}ht\n∝an}bracketle{tΨ|Ψ∝an}bracketri}ht. (2)\nExploratory calculation of the condensate fraction of α\nmatterwillbegiveninthefollowingSec.II.Incontrastto\nRef. [6, 8] where the transition temperature Tcfor quar-\ntetting was considered, we considerhere the zerotemper-\naturecaseand analyzethe groundstatewavefunction. It\nwill be shown that due to the interaction, the condensate\nfraction is suppressed with increasing density.\nAn important question is whether such properties of\ninfinite nuclear matter are of relevance for finite nuclei.\nAs well known, e.g., pairing obtained in nuclear matter\nwithin the BCS approach is also clearly seen in finite nu-\nclei. Nuclei with densities near the saturation density are\nwell described by the quasiparticle picture which leads to\nthe shell model for finite nuclei. At low densities, a fully\ndevelopedαcluster structure similar to αmatter is ex-\npected. Cluster structures in finite nuclei have been well\nestablished. A density functional approach is able to in-\nclude correlations and to bridge between infinite matter\nand finite nuclei.\nAn interesting aspect of finite nuclei is the enhance-\nment of the occupation of single α-particle states sim-\nilar to Bose-Einstein condensation in α-particle matter\nor condensation of bosonic atoms in traps. Recently,\ngas-like states have been investigated in self-conjugate\n4nnuclei [10], and a special ansatz for the wave function\n(THSR ansatz), which is similar to the condensate state\nin infinite matter, has been shown to be appropriate in\ndescribing low-density isomers. In particular,8Be and\nthe Hoyle state of12C are well described with this THSR\nwave function. Investigation of states near the four α\nthreshold in16O is in progress [11, 12]. Predictions for\n20Ne have been given in [13].\nIn Sect. III, we will explain how the suppression of\nthe condensate fraction, calculated for infinite nuclear\nmatter, is also seen in the low-density isomers of self-\nconjugate 4 nnuclei, in particular for n= 3 (12C). First\nresults forn= 4 (16O) are also given. General conclu-\nsions are drawn in Sec. IV.\nII. SUPPRESSION OF CONDENSATE\nFRACTION IN αMATTER AT ZERO\nTEMPERATURE\nThetheoryofPenroseandOnsager[9]wasfirstapplied\nto a system with hard core repulsion. Depending on the\nfilling factor, the suppression of the condensate was cal-\nculated. In particular, for liquid4He with a filling factor\nof 28% at normal conditions, the condensate fraction is\nreduced to ≈8% in good agreement with experimental\nobservations. To give an estimation for αmatter, with\nan “excluded volume” of about 20 fm3[14], such a fillingfactor of 28 % would arise at ≈ρ0/3 so that a substan-\ntial reduction of the condensate fraction already below\nsaturation densities is expected for αmatter.\nWithinamoresystematicapproach,wefollowthework\nofClarketal. [15]. Wecalculatethereductionofthecon-\ndensate fraction as function of the baryon density within\nperturbation theory. A uniform Bose gas of αparticles,\ninteracting via the potential Vα(r), is considered, disre-\ngardingany change of the internal structure of the αpar-\nticles at increasing density. In particular, the dissolution\nof theαparticle as a four-nucleon bound state because\nof the Pauli blocking is not taken into account.\nThe simplest form of a trial wave function incorpo-\nrating the strong spatial correlations implied by the\ninteraction potential is the familiar Jastrow choice,\nψ(r1,...,rA) =/producttext\nin(ri,m−ri,n)2]. The wave function\nχ(s,t) of the relative motion of 3 αclusters is obtained\nbysolvingtheenergyeigenvalueproblemofthefullthree-\nbody equation of motion; ∝an}bracketle{tφ3\nα|(H−E)|A{χ(s,t)φ3\nα}∝an}bracketri}ht=\n0, whereHis the microscopic Hamiltonian consisting of\nthe kinetic energy, effective two-nucleon potential, and\nthe Coulomb potential between protons. The difference\nbetween the works by Uegaki et al. and Kamimura et\nal. lies in the adopted effective two-nucleon force, besides\nthe differing techniques of solution.\nBoth calculations by Uegaki et al. and Kamimura et\nal. reproduced reasonably well the observed binding en-\nergy and r.m.s. radius of the ground 0+\n1state which is\nthe state with normal density, while they both predicted\na very large r.m.s. radius for the second 0+\n2state which\nis larger than the r.m.s. radius of the ground 0+\n1state\nby about 1 fm, i.e. by over 30%. The observed 0+\n2state\nlies slightly above the 3 αbreakup threshold and the en-\nergies of the calculated 0+\n2state reproduced reasonably\nwell the observed value although the value by Uegaki et\nal. is slightly higher than the 3 αbreakup threshold by\nabout 1 MeV. The second 0+state of12C is well known\nas the key state for the synthesis of12C in stars (Hoyle\nstate) and also as one of the typical mysterious 0+states\nin light nuclei which are very difficult to understand from\nthe point of view of the shell model [23].\nAlternatively, the 0+\n2state with dilute density can be\ndescribed by a gas-likestructure of 3 α-particles which in-\nteract weakly among one another, predominantly in rel-\nativeSwaves. The S-wave dominancy in the 0+\n2state\nstructure had been already suggested by Horiuchi on the\nbasisofthe3 αOCM(orthogonalityconditionmodel)cal-\nculation [24]. It should be mentioned that not only thebinding energy, but also other properties of the 0+\n2state\nsuchaselectronscatteringformfactorsarewelldescribed\nwithin the calculations given in Refs. [21, 22, 24].\nRecently, based on the investigations of the possibility\nofα-particle condensation in low-density nuclear mat-\nter [6], the present authors proposed a conjecture that\nnear thenαthreshold in self-conjugate 4 nnuclei there\nexist excited states of dilute density which are composed\nof a weekly interacting gas of self-bound αparticles and\nwhich can be considered as an nαcondensed state [10].\nThis conjecture was backed by examining the structure\nof12Cand16Ousinganew α-cluster wavefunction ofthe\nα-cluster condensate type. The new α-cluster wave func-\ntion, which will be denoted as THSR wave function, ac-\ntually succeeded to place a level of dilute density (about\none third of saturation density) in each system of12C\nand16O in the vicinity of the 3 respectively 4 αbreakup\nthreshold, without using any adjustable parameter. In\nthe case of12C, this success of the new α-cluster wave\nfunction mayseemrathernatural, asweexplained above.\nThemicroscopic3 αclustermodelshadpredictedthatthe\n0+\n2in the vicinity of the 3 αbreakup threshold has a gas-\nlike structure of 3 α-particles which interact weakly with\neach other predominantly in relative Swaves. Having\nput forward that Hoyle like states in 4 nself-conjugate\nnuclei may be a general and common phenomenon is the\nmerit of the work in [10].\nThe THSR wave function of the α-cluster condensate\ntype used in Ref. [10] represents a condensation of α-\nclusters in a spherically symmetric state. This is clearly\nseen by the following expression\n|Ψ∝an}bracketri}ht=P(C†\nα)n|vac∝an}bracketri}ht, (8)\nwith\n∝an}bracketle{t1234|C†\nα|vac∝an}bracketri}ht= Φ(P)δP,p1+p2+p3+p4φα(1234)a†\n1a†\n2a†\n3a†\n4,\n(9)\nΦ(P) describing the c.o.m. motion of the αcluster, and\nφthe internal wave function of the four-nucleon cluster.\nThe operator Pis projecting out the total c.o.m. motion\nof the 4nnucleus. In the limit of infinite nuclear matter,\nthe Φ orbitals are plane waves, and the projection op-\neratorPcan be neglected. In the case considered here,\nthe use of Gaussians allows the explicit separation of the\nc.o.m. motion of the four-nucleon cluster as well as of\nthe whole 4 nnucleus. It should also be noted that Eq.\n(8) contains two limits exactly: the one of a pure Slater\ndeterminant relevant at higher densities and the one of\na 100 percent ideal α-particle condensate in the dilute\nlimit [10]. All intermediate scenarios are also correctly\ncovered.\nThe present authors extended the wave function so\nthat it can describe the α-cluster condensate with spatial\ndeformation [25]. They applied this new wave function\nto8Be and succeeded to reproduce not only the bind-\ning energy of the ground state but also the energy of\nthe excited 2+state. In addition, they found that al-\nthough the effect of the spatial deformation is not large,5\nthe introduction of the spatial deformation brought forth\na 100 % overlap of the THSR wave function with the\n“exact” wave function given by the microscopic 2 αclus-\nter model which solves the 2 α-cluster equation of mo-\ntion,∝an}bracketle{tφ2\nα|(H−E)|A{χ(r)φ2\nα}∝an}bracketri}ht= 0. This fact forces us\nto modify our understanding of the8Be structure from\nthe 2α“dumb-bell” structure to the 2 αdilute (gas-like)\nstructure. It was shown that the 0+\n2wave function of\n12C which was obtained long time ago by solving the full\nthree-body problem of the microscopic 3 αcluster model\nis almost completely equivalent to the wave function of\nthe3αTHSRstate. Thisresultgivesusstrongsupportto\nour opinion that the 0+\n2state of12C has a gas-like struc-\nture of 3αclusters with “Bose-condensation”. The rms\nradius for this THSR state was calculated as R(0+\n2)THSR\n= 4.3 fm which fits well with experimental data for the\nform factor of the Hoyle state, see Ref. [26]. It confirms\ntheassumptionoflowdensityasaprerequisiteforthefor-\nmation of an α-cluster structure for which the Bose-like\nenhancement of the occupation of the Sorbit is possible.\nRecently, a fermionic AMD calculationbased on nucle-\nons with effective interactions has been performed [26]\nwhich supports the applicability of the THSR state to\ndescribe the Hoyle state. It is found hat the form fac-\ntor calculated for the 0+\n2state of12C coincides with the\nform factor obtained from the THSR wave function. In\nparticular, the low density of nucleons, the formation of\nfour-nucleon clusters and the dominant contribution of\nthe gas-like distribution has been confirmed.\nA very interesting analysis of the applicability of the\nTHSR wave function can be performed by comparing\nwith stochastic variational calculations [27] and OCM\ncalculations [28]. The αdensity matrix ρ(r,r′) defined\nby integrating out of the total density matrix all intrin-\nsicα-particle coordinates, is diagonalized to study the\nsingle-αorbitsand occupationprobabilitiesin12Cstates.\nFig. 2 shows the occupation probabilities of the L-orbits\nwithS,DandGwaves belonging to the k-th largest oc-\ncupation number (denoted by Lk), for the ground and\nHoyle state of12C obtained by diagonalizing the density\nmatrixρ(r,r′). We found that in the Hoyle state the\nα-particleSorbit with zero node ( S1 in Fig. 2) is occu-\npied to more than 70 % by the three α-particles (see also\nRef.[27]andFig.1). Takingintoaccountthefinitesizeof\nthe nucleus, a reduction of the condensate fraction from\n100 % to about 70 % is not surprising, and the remaining\nfraction (about 30 %) is due to higher orbits originat-\ning from antisymmetrization among nucleons. This huge\npercentagemeansthatanalmostideal α-particleconden-\nsate is realized in the Hoyle state. One should remember\nthat superfluid4He has only 8 percent of the particles in\nthe condensate, what represents a macroscopic amount\nof particlesnonetheless. Pleasealsonote that the S-wave\noccupancy of the Hoyle state is at least by a factor ten\nlargerthan the occupancyofany otherstate (Fig. 2). In-\ndependent ofthe absoluteoccupancy ofthe S-wavestate,\nthisis aclearsignatureofquantumcoherence,i.e. ofcon-\ndensation.On the other hand, in the ground state of12C, the\nα-particle occupations are equally shared among S1,D1\nandG1 orbits, where they have two, one, and zero nodes,\nrespectively, reflecting the SU(3)( λµ) = (04) character\nof the ground state [28]. This fact thus invalidates a\ncondensate picture for the ground state.\nTo get a more extended analysis, OCM calculations\nhave been performed [28] for studying the density de-\npendence of the S-orbit occupancy in the 0+state of\n12C on the different densities ρ/ρ0∼(R(0+\n1)exp/R)3, in\nwhich the rms radius ( R) of12C is taken as a parame-\nter andR(0+\n1)exp=2.56 fm. A Pauli-principle respected\nOCM basis ΨOCM\n0+(ν) with a size parameter νis used,\nin which the value of νis chosen to reproduce a given\nrms radius Rof12C, and the αdensity matrix ρ(r,r′)\nwith respect to ΨOCM\n0+(ν) is diagonalized to obtain the S-\norbit occupancy in the 0+wave function. The results are\nshown in Fig. 3. The S-orbit occupancy is 70 ∼80 %\naroundρ/ρ0∼(R(0+\n1)exp/R(0+\n2)THSR)3= 0.21, while\nit decreases with increasing ρ/ρ0and amounts to about\n30∼40 % in the saturation density region. A smooth\ntransition of the S-orbit is observed from the zero-node\nS-wave nature ( ρ/ρ0≃0.2) to a two-node S-wave one\n(ρ/ρ0∼1) with increasing ρ/ρ0[28]. The feature of the\ndecrease of the enhanced occupation of the Sorbit is in\nstriking correspondence with the density dependence of\nthecondensatefractioncalculatedfornuclearmatter(see\nFig. 1).\nAn interesting item is whether there exist other nuclei\nshowing the Bose condensate-like enhancement of the S-\norbit occupation number. Then, the suppression of the\ncondensate with increasingdensity is alsoof relevance for\nthose nuclei. After we discussed the case of12C corre-\nsponding to n= 3 we will now shortly discuss the sit-\nuation in the next nucleus16O corresponding to n= 4,\nwhere great efforts are performed recently to investigate\nlow-density excitations in the 0+spectrum in theory as\nwell as in experiments.\nIn analogy to the aforementioned OCM calculation for\n12C [28], we recently performed a quite complete OCM\ncalculation also for16O, including many of the cluster\nconfigurations, just mentioned (a full account will be\ngiven in a separate publication [12]). We were able to\nreproduce the full spectrum of 0+states with 0+\n2at 6.4\nMeV, 0+\n3at 9.4 MeV, 0+\n4at 12.6 MeV, 0+\n5at 14.1 MeV,\nand 0+\n6at 16.5 MeV. Also the rms radii are obtained.\nThe largest values are found as R(0+\n6)OCM= 5.6 fm, fol-\nlowed byR(0+\n4)OCM= 4.0 fm. We tentatively make a\none to one correspondence of those states with the six\nlowest 0+states of the experimental spectrum. In view\nof the complexity of the situation, the agreement can be\nconsidered as very satisfactory. The analysis of the diag-\nonalization of the α-particle density matrix ρ(r,r′) (as\nwas done in Ref. [28]) showed that the newly discovered\n0+state at 13.6 MeV [31], as well as the well known 0+\nstate at 14.01 MeV, corresponding to our states at 12 .6\nMeV and 14 .1 MeV, respectively, have, contrary to what6\nwe assumed previously [32], very little condensate occu-\npancy of the zero-node S-orbit (about 20 percent). On\nthe other hand, the sixth 0+state at 16.5 MeV calcu-\nlated energy, to be identified with the experimental state\nat 15.1 MeV, has 61 percent of the αparticles being in\nthe zero-node S-orbit.\nTheseresultsconfirmourstatementthatthe α-particle\ncondensate in nuclear matter is suppressed with increas-\ning density and, consequently, a well developed con-\ndensate state in nuclei can be expected only at very\nlow densities. For16O, the relative densities ρ/ρ0\nare estimated as ( R(0+\n1)exp/R(0+\n4)OCM)3= 0.32 and\n(R(0+\n1)exp/R(0+\n6)OCM)3= 0.12. Therefore we expect a\nsignificant enhancement of the Sorbit occupation num-\nber only for the 0+\n6state, in full agreement with the\nOCM calculation cited above. The very large radius of\nthat state is again a clear indication of an α-particle gas\n(Hoyle)-like state, and the THSR wave function is ex-\npected to describe this state in a sufficient approxima-\ntion. Work in determining the complete spectrum of\nTHSR states in16O showing the relevance of a Bose-\ncondensate like state is in progress [11].\nIV. CONCLUSIONS\nMultiple successful theoretical investigations, concern-\ning the Hoyle state in12C, have established, beyond any\ndoubt, thatitisadilutegas-likestateofthree α-particles,\nheld together only by the Coulomb barrier, and describ-\nabletofirstapproximationbyawavefunctionoftheform\n(C†\nα)3|vac∝an}bracketri}htwhere the three bosons ( C†\nα) are condensed\ninto theS-orbital. There is no objective reason, why in\n16O,20Ne,···there should not exist similar ’Hoyle’-like\nstates. At least the calculations with THSR and OCM\napproaches show this to be the case, systematically. In\nthis work,wegivepreliminaryresultsofacomplete OCM\ncalculation which reproduces the six first 0+states of\n16O to rather good accuracy. In that calculation the 0+\n6\nstate at 16 .5 MeV, which might be identified with the\nexperimental 0+state at 15.1 MeV, shows the character-\nistics typical for a Hoyle-like state, that is high α-particle\nS-wave occupancy combined with an unusually large ra-\ndius.\nTherefore, the main quantity for the formation of an\nαcluster state is the density which should be low. Then,\nthe occurrence of a THSR state where all αparticles oc-\ncupy the same orbit with respect to the c.o.m. motion\nis an interesting effect which corresponds to the forma-\ntion of an α-particle condensate in symmetric nuclear\nmatter. The condensate fraction is decreasing with in-\ncreasing density because of correlations, as known from\ninteracting Bose systems. In addition, the internal struc-\nture of the four nucleon cluster is changed due to Pauli\nblocking if density is increasing.\nOnly in the very low density limit, the αparticles maybe considered as independent bosons moving relatively\nfree like quasi particles. A mean field approach of the\ninteraction which is assumed to be week would give a\nGross-Pitaevskii equation [33]. Then we can apply the\napproach of a non-interacting Bose gas where the αpar-\nticles may occupy the samec.o.m. orbital. The enhanced\noccupation of the ground state (plane wave) in infinite\nmatter is the standard description of Bose-Einstein con-\ndensation. This corresponds, in finite nuclei, to the en-\nhanced occupation of the same orbital for the c.o.m. mo-\ntionsothattheTHSRstatewillbeagoodapproximation\nfor the many-nucleon wave function. We stress the simi-\nlarity to two-particle pairing where the concept of a BCS\nstate was successfully applied to finite nuclei. The ques-\ntion of finite number of Cooper pairs in the nuclear BCS\nstate is also to be considered in analogy with the finite\nnumber ofα-particles in the THSR state.\nWith increasing contribution of the interaction, e.g.\nwith increasing density, the condensate state becomes\nmore complex. Calculations in infinite matter ( T= 0)\nshowthatthecondensatestatebecomesincreasinglynon-\nideal (the condensate fraction is smaller than one). The\nsame is also observed in OCM calculations for finite nu-\nclei where with increasing density the condensate state\nbecomes gradually depleted. We conclude that there are\nsimilarities between the structure of the ground state\nwave function of αmatter and the αgas like states in\nfinite nuclei.\nIn addition to the effect of interaction, mixing higher\nstates of c.o.m. orbits to the ground state wave function,\nthere is also the dissolution of the internal wave func-\ntion of the αparticle due to medium effects. The tran-\nsition from the cluster picture with well defined αstates\nto a shell model where nucleons move independently in\na mean field is also reproduced in harmonic oscillator\napproximation, but needs a first principle approach to\ncalculate the many-nucleon wave function.\nThese results are also of relevance for other phenom-\nena which ariseif the local density approachis used. Low\ndensity matter arises in the halo of heavy nuclei so that\npreformation of α-clusters is an interesting issue there,\nbut also in heavy ion reactions or during supernova ex-\nplosions. Cluster condensation very likely will soon also\nbecome an important subject in cold atom physics. The-\noretical investigations already have appeared [7]. So far\nnuclear physics is at the forefront of this subject.\nAcknowledgements\nThepresentauthors(G.R. andT.Y.) expressthanksto\nM.L.RistigandJ.W.Clark,andalsototheorganizersof\nthe International Workshop on Condensed Matter The-\nories (CMT31) at Bangkok, Thailand, (V. Sa-yakanit:\nchairperson), where this paper was completed.7\n[1] D.M. Brink, J.J. Castro, Nucl. Phys. A 216, 109 (1973).\n[2] S. Ali, A.R. Bodmer, Nucl. Phys. A 80, 99 (1966).\n[3] Y.C. Tang, K. Wildermut, A Unified Theory of the Nu-\ncleus(Academic, New York, 1977).\n[4] G. R¨ opke, L. M¨ unchow, H. Schulz, Nucl. Phys. A 379,\n536 (1982); A 399, 587 (1983).\n[5] M. Beyer, S. A. Sofianos, C. Kuhrts, G. R¨ opke, and P.\nSchuck, Phys. Lett. B 488, 247 (2000).\n[6] G. R¨ opke, A. Schnell, P. Schuck, and P. Nozieres, Phys.\nRev. Lett. 80, 3177 (1998).\n[7] S. Capponi, G. Roux, P. Lecheminant, P. Azaria, E.\nBoulat, S.R. White, arXiv: 0706.0609.; B. Doucot, J.\nVidal, Phys. Rev. Lett. 88, 227005 (2002).\n[8] A.Sedrakian, H.Muether, P. Schuck, Nucl.Phys. A 766,\n97 (2006).\n[9] O. Penrose, L. Onsager, Phys. Rev. 104, 576 (1956).\n[10] A. Tohsaki, H. Horiuchi, P. Schuck, and G. R¨ opke, Phys.\nRev. Lett. 87, 192501 (2001).\n[11] Y. Funaki, H. Horiuchi, G. R¨ opke, P. Schuck, A. Tohsaki ,\nT. Yamada, in preparation.\n[12] Y. Funaki, T. Yamada, P. Schuck, H. Horiuchi,\nA. Tohsaki, and G. R¨ opke, in preparation.\n[13] A. Tohsaki, H. Horuichi, P. Schuck and G. R¨ opke, Nucl.\nPhys.A 738, 259 (2004).\n[14] J. M. Lattimer, F. D. Swesty, Nucl. Phys. A 535, 331\n(1991).\n[15] M. T. Johnson and J. W. Clark, Kinam2, 3 (1980)\n(PDF available at Faculty web page of J. W. Clark\nat http://wuphys.wustl.edu); see also J. W. Clark and\nT. P. Wang, Ann. Phys. (N.Y.) 40, 127 (1966) and\nG. P. Mueller and J. W. Clark, Nucl. Phys. A 155, 561\n(1970).\n[16] K.A. Gernoth, M.L. Ristig, and T. Lindenau, Int. J.\nMod. Phys. B 21, 2157 (2007); G. Senger, M.L. Ristig,\nC.E.Campbell, and J.W. Clark, Ann. Phys. (N.Y.) 218,\n160 (1992); R. Pantfoerder, T. Lindenau, and M.L.\nRistig, J. Low Temp. Phys. 108, 245 (1997).[17] G. R¨ opke and P. Schuck, Mod. Phys. Lett. A 21, 2513\n(2006).\n[18] Z.F. Shehadeh et al., Int. J. Mod. Phys. B 21, 2429\n(2007); M.N.A. Abdullah et al., Nucl. Phys. A 775, 1\n(2006).\n[19] J. W. Clark, private communication.\n[20] For example, Y. Fujiwara, H. Horiuchi, K. Ikeda, M.\nKamimura, K. Kato, Y. Suzuki, and E. Uegaki, Prog.\nTheor. Phys. Suppl. No.68, 29 (1980).\n[21] E. Uegaki, S. Okabe, Y. Abe, and H. Tanaka, Prog.\nTheor. Phys. 57, 1262 (1977); E. Uegaki, Y. Abe, S. Ok-\nabe, and H. Tanaka, Prog. Theor. Phys. 59, 1031 (1978);\n62, 1621 (1979).\n[22] Y. Fukushima and M. Kamimura, Proc. Int. Conf.\non Nuclear Structure , Tokyo, 1977, ed. T. Marumori\n(Suppl. of J. Phys. Soc. Japan, Vol.44, 1978), p.225; M.\nKamimura, Nucl. Phys. A 351, 456 (1981).\n[23] B. R. Barrett, B. Mihaila, S. C. Pieper, and R. B.\nWiringa, Nucl. Phys. News, 13, 17 (2003).\n[24] H. Horiuchi, Prog. Theor. Phys. 51, 1266 (1974); 53, 447\n(1975).\n[25] Y. Funaki, H. Horiuchi, A. Tohsaki, P. Schuck, and G.\nR¨ opke, Prog. Theor. Phys. 108, 297 (2002).\n[26] M. Chernykh, H. Feldmeier, T. Neff, P. von Neumann-\nCosel, A. Richter, Phys. Rev. Lett. 98, 032501 (2007).\n[27] H. Matsumura, Y. Suzuki, Nucl. Phys. A 739, 238\n(2004).\n[28] T. Yamada, P. Schuck, Eur. Phys. J. A 26, 185 (2005).\n[29] H. Horiuchi and K. Ikeda, Prog. Theor. Phys. 40, 277\n(1968); Y. Suzuki, Prog. Theor. Phys. 55, 1751 (1976);\n56, 111 (1976).\n[30] F. Ajzenberg-Selove, Nucl. Phys. A 460, 1 (1986).\n[31] T. Wakasa et al., Phys Lett B 653, 173 (2007).\n[32] Y. Funaki, H. Horiuchi, G. R¨ opke, P. Schuck, A. Tohsaki ,\nT. Yamada, Nucl. Phys. News, 17(04), 11 (2007).\n[33] T. Yamada, P. Schuck, Phys. Rev. C 69, 024309 (2004)." }, { "title": "0803.0779v1.Ultimate_Energy_Densities_for_Electromagnetic_Pulses.pdf", "content": "arXiv:0803.0779v1 [quant-ph] 6 Mar 2008Ultimate Energy Densities for Electromagnetic Pulses\nMankei Tsang∗\nResearch Laboratory of Electronics, Massachusetts Institu te of Technology,\nCambridge, Massachusetts 02139, USA\n(Dated: November 14, 2018)\nThe ultimate electric and magnetic energy densities that ca n be attained by ban-\ndlimited electromagnetic pulses in free space are calculat ed using an ab initio quan-\ntized treatment, and the quantum states of electromagnetic fields that achieve the\nultimate energy densities are derived. The ultimate energy densities also provide\nan experimentally accessible metric for the degree of local ization of polychromatic\nphotons.\nPACS numbers:\nUltrafastlasershavebecome anindispensible toolinawide spectrum ofscience, including\nnonlinear optics[1, 2], metrology[3], laserfusion[4], biologicalimaging[5], biologicalsurgery\n[6], and chemistry [7]. A key to the success of such lasers is the extre mely high peak energy\ndensity that they can achieve, as the moderate energy of each las er pulse can be confined\nwithin femtoseconds or even attoseconds and focused to a micron -sized area. The high\nenergy density strongly enhances light-matter interactions, and is especially crucial to the\nstudy ofrelativistic nonlinear optics [2]. Given theimportance of anult rahigh optical energy\ndensity inabroadrangeofapplications, thelimit towhichonecanfocu sabroadbandoptical\npulse in three spatiotemporal dimensions and maximize the energy de nsity is a fundamental\nproblem.\nLocalization of electromagnetic pulses has been investigated both in the classical regime\n[8] and the quantum single-photon regime [9, 10]. While these seminal studies have con-\ntributed to our fundamental understanding of electromagnetic e nergy localization, all of\ntheir predictions have not yet been experimentally verified because of the difficulty in im-\nplementing their proposed electromagnetic pulse solutions. Most re quire a spectrum that\ncontains arbitrarily high frequency components [8, 9] and can be ex ceedingly difficult to\nrealize due to the finite laser gain bandwidth or finite transparency r ange of optical compo-2\nnents. The use of a spontaneously emitting atomproposedin Ref. [1 0] couples theproperties\nof the emitted photon to those of the atom and does not seem to be generalizable to other\nsituations. Moreover, the quantum studies [9, 10] are mainly conce rned with the decay of\nthe energy density far away from the center of localization at an ins tant of time for one\nphoton, and thus are not immediately relevant to the more practica l problem of maximizing\nthe energy density at the center of localization for a large number o f bandlimited photons.\nIn this Letter, the ultimate electric and magnetic energy densities t hat can be attained by\nbandlimited electromagnetic pulses in free space are calculated using anab initio quantized\ntreatment, and the quantum states that achieve the ultimate den sities are derived. By\ntaking into account all degrees of freedom of electromagnetic field s and explicitly limiting\nthebandwidth of thepulses, our result overcomes all theshortco mings of theaforementioned\nstudies and is more applicable to experimental situations. Measuring the energy densities\nat a fixed point is also considerably easier experimentally than measur ing the decay of the\nenergydensity ataninstant oftime, sothemaximum achievableener gydensities canbeused\nas an alternative and more accessible metric for the degree of localiz ation of polychromatic\nphotons. Most importantly, the ultimate energy densities impose a f undamental limit to\nwhich a bandlimited optical pulse can be focused spatially and tempora lly, so the presented\nresult should prove useful for designing ultrafast optics experime nts.\nThe procedure of calculating the maximum energy densities is similar to the one used to\ncalculate the multiphoton absorption rate limit for monochromatic ligh t in Ref. [11], except\nthatherewegeneralizetheproceduretopolychromaticlight, such thatalldegreesoffreedom\nare taken into account and the treatment can be considered ab initio. We also calculate the\nmaximum magnetic energy density and the corresponding quantum s tate, as the magnetic\nfield can also play a significant role in relativistic nonlinear optics [2].\nWe first derive the ultimate electric energy density, since it is more imp ortant for most\napplications. Consider the quantized electric field operator in free s pace [12]:\nˆE(r,t) =i\n(2π)3/2/summationdisplay\nσ/integraldisplay\nd3k/parenleftbigg¯hω\n2ǫ0/parenrightbigg1/2\nε(k,σ)ˆa(k,σ)eik·r−iωt+H.c., (1)\nwhereσdenotesthetwotransversepolarizations, ε(k,σ)istheunitelectric-fieldpolarization\nvector,ω=ck=c(k2\nx+k2\ny+k2\nz)1/2is the frequency, ˆ a(k,σ) is the annihilation operator\nsatisfying the commutation relation [ˆ a(k,σ),ˆa†(k′,σ′)] =δ3(k−k′)δσσ′, and H.c. is the\nHermitian conjugate. To impose a limit on the bandwidth, it is necessar y to describe3\nthe optical modes in terms of the frequency variable. This can be do ne by changing the\nmomentum-space coordinates ( kx,ky,kz) to normalized spherical coordinates (Ω ,θ,φ):\nω=ω0Ω, k x=k0Ωsinθcosφ,\nky=k0Ωsinθsinφ, k z=k0Ωcosθ,\ndkxdkydkz=dΩdθdφ/parenleftbig\nk3\n0Ω2sinθ/parenrightbig\n,ˆa(k,σ) = ˆa(Ω,θ,φ,σ)/parenleftbig\nk3\n0Ω2sinθ/parenrightbig−1/2.(2)\nwhereω0is a normalizationfrequency, k0≡ω0/c≡2π/λ0, andtheannihilation operatorhas\nbeen renormalized so that its commutator is [ˆ a(Ω,θ,φ,σ),ˆa†(Ω′,θ′,φ′,σ′)] =δ(Ω−Ω′)δ(θ−\nθ′)δ(φ−φ′)δσσ′. The positive-frequency electric field becomes\nˆE(+)(r,t) =i/parenleftbigg¯hω0\n2ǫ0λ3\n0/parenrightbigg1/2/integraldisplay∞\n0dΩ/integraldisplayπ\n0dθ/integraldisplay2π\n0dφ/parenleftbig\nΩ3sinθ/parenrightbig1/2ε(θ,φ,σ)ˆa(Ω,θ,φ,σ)eik·r−iωt.\n(3)\nIn the continuous Fock space representation [12], the N-photon momentum eigenstate is\ngiven by\n|Ω1,θ1,φ1,σ1,...,ΩN,θN,φN,σN/angbracketright ≡1√\nN!N/productdisplay\nn=1ˆa†(Ωn,θn,φn,σn)|0/angbracketright, (4)\nand the identity operator is\nˆ1 =∞/summationdisplay\nN=0|N/angbracketright/angbracketleftN|, (5)\n|N/angbracketright/angbracketleftN|=/summationdisplay\nσ1,...,σN/integraldisplay\ndΩ1dθ1dφ1...dΩNdθNdφN\n×|Ω1,θ1,φ1,σ1,...,ΩN,θN,φN,σN/angbracketright/angbracketleftΩ1,θ1,φ1,σ1,...,ΩN,θN,φN,σN|.(6)\nAn arbitrary quantum state of electromagnetic fields can thus be e xpressed as\n|Ψ/angbracketright=∞/summationdisplay\nN=0CN|N/angbracketright, CN≡ /angbracketleftN|Ψ/angbracketright, (7)\nand a Fock state as\n|N/angbracketright=/summationdisplay\nσ1,...,σN/integraldisplay\ndΩ1dθ1dφ1...dΩNdθNdφNΦN(Ω1,θ1,φ1,σ1,...,ΩN,θN,φN,σN)\n×|Ω1,θ1,φ1,σ1,...,ΩN,θN,φN,σN/angbracketright, (8)4\nwhere\nΦN(Ω1,θ1,φ1,σ1,...,ΩN,θN,φN,σN)≡ /angbracketleftΩ1,θ1,φ1,σ1,...,ΩN,θN,φN,σN|N/angbracketright(9)\nis theN-photon momentum-space probability amplitude, which must satisfy the normaliza-\ntion condition:\n/summationdisplay\nσ1,...,σN/integraldisplay\ndΩ1dθ1dφ1...dΩNdθNdφN|ΦN(Ω1,θ1,φ1,σ1,...,ΩN,θN,φN,σN)|2= 1,(10)\nand the symmetrization condition:\nΦN(...,Ωn,θn,φn,σn,...,Ωm,θm,φm,σm,...)\n= ΦN(...,Ωm,θm,φm,σm,...,Ωn,θn,φn,σn,...) for any nandm. (11)\nTo impose a limited bandwidth ( α≤Ω≤β) on the electromagnetic fields, we require the\nprobability of photons existing outside the bandwidth to vanish:\nΦN(Ω1,θ1,φ1,σ1,...,ΩN,θN,φN,σN) = 0 for any Ω n< αor Ωn> β. (12)\nWith the theoretical framework put forth, we now proceed to calc ulate a bound on the\nelectric energy density (minus the zero-point energy density) give n by\nUe≡/angbracketleftBig\n:ǫ0\n2ˆE·ˆE:/angbracketrightBig\n=/angbracketleftBig\nǫ0ˆE(−)·ˆE(+)/angbracketrightBig\n. (13)\nA bound on the electric energy density is equivalent to a bound on the energy density for\none component of the electric field:\nU′\ne≡/angbracketleftbigg\n:ǫ0\n2/parenleftBig\np·ˆE/parenrightBig2\n:/angbracketrightbigg\n=ǫ0/angbracketleftBig/parenleftBig\np·ˆE(−)/parenrightBig/parenleftBig\np·ˆE(+)/parenrightBig/angbracketrightBig\n, (14)\nwherepis an arbitrary real unit vector, because UeandU′\neare equivalent if we choose pto\nbe parallel to the electric field. In terms of the momentum-space re presentation,\nU′\ne=¯hω0\n2λ3\n0∞/summationdisplay\nN=0|CN|2N/summationdisplay\nσ2,...,σN/integraldisplay\ndΩ2dθ2dφ2...dΩNdθNdφN\n×/vextendsingle/vextendsingle/vextendsingle/vextendsingle/vextendsingle/summationdisplay\nσ/integraldisplayβ\nαdΩ/integraldisplayπ\n0dθ/integraldisplay2π\n0dφ/bracketleftBig\ni/parenleftbig\nΩ3sinθ/parenrightbig1/2p·ε(θ,φ,σ)eik·r−iωt/bracketrightBig\n×ΦN(Ω,θ,φ,σ,Ω2,θ2,φ2,σ2,...,ΩN,θN,φN,σN)/vextendsingle/vextendsingle/vextendsingle/vextendsingle/vextendsingle2\n, (15)5\nwhere the symmetric property of Φ Ngiven by Eq. (11) is used. By virtue of the Schwarz’s\ninequality and the normalization condition given by Eq. (10),\nU′\ne≤¯hω0\n2λ3\n0∞/summationdisplay\nN=0|CN|2N/summationdisplay\nσ1,...,σN/integraldisplay\ndΩ1dθ1dφ1...dΩNdθNdφN\n×|ΦN(Ω1,θ1,φ1,σ1,...,ΩN,θN,φN,σN)|2\n×/summationdisplay\nσ/integraldisplayβ\nαdΩ/integraldisplayπ\n0dθ/integraldisplay2π\n0dφ/parenleftbig\nΩ3sinθ/parenrightbig\n|p·ε(θ,φ,σ)|2\n=π\n3/angbracketleftN/angbracketright¯hω0\nλ3\n0/parenleftbig\nβ4−α4/parenrightbig\n, (16)\nwhere/angbracketleftN/angbracketright ≡/summationtext\nN|CN|2Nis the average photon number. As expected, the bound on U′\ne\ndoes not depend on p, and is therefore also applicable to the total electric energy densit y.\nDefining the actual lower and upper frequencies as ω1=αω0andω2=βω0, respectively,\nand the corresponding wavelengths as λ1,2= 2πc/ω1,2, we obtain the central result of this\nLetter:\n/angbracketleftBig\n:ǫ0\n2ˆE·ˆE:/angbracketrightBig\n≤π\n3/angbracketleftN/angbracketright/parenleftbigg¯hω2\nλ3\n2−¯hω1\nλ3\n1/parenrightbigg\n. (17)\nThis simple expression agrees with the intuition that the ultimate ener gy density is limited\nby the maximum energy of photons ( /angbracketleftN/angbracketright¯hω2) divided by the smallest volume that the\nphotons can occupy ( λ3\n2).\nIn the limit of a small bandwidth compared to the center frequency, we can let ∆ ω≡\nω2−ω1,ω0= (ω1+ω2)/2, ∆ω≪ω0, and obtain\n/angbracketleftBig\n:ǫ0c\n2ˆE·ˆE:/angbracketrightBig\n<∼2\n3/angbracketleftN/angbracketright¯hω0∆ω\nλ2\n0, (18)\nwhich is a bound on the peak intensity in the slowly-varying envelope re gime, and again\nagrees with the intuition that the highest intensity is achieved when t he mean energy of\nthe photons is focused to their minimum pulse width (2 π/∆ω) and beam size ( λ2\n0). This\napproximate bound also agrees with that derived in Ref. [11], where t he monochromatic\napproximation is made at the beginning. Beyond the slowly-varying en velope regime, the\nexact bound given by Eq. (17) depends on the upper frequency to the fourth power, un-\nderlying the importance of high-frequency components in maximizing the energy density, as\nthey have a higher energy as well as a smaller localization volume.\nThe use of the Schwarz’s inequality is reminiscent of the matched filte r concept in com-\nmunication theory [13]. In Eq. (15), the N-photon amplitude can be regarded as the input6\nsignal, and the expression in square brackets can be regarded as a filter transfer function in\nthe measurement of the electric field. An N-photon amplitude that achieves the Schwarz\nupper bound is one that is linearly dependent on the square-bracke ted expression, or in\nother words, when the input signal matches the filter. Assuming a f actorizable Φ N, the\nfollowing N-photon amplitude that achieves the ultimate electric energy densit y can then\nbe obtained:\nΦN=N/productdisplay\nn=1fe(Ωn,θn,φn,σn),\nfe=/braceleftbigg−iC−1/2(Ω3sinθ)1/2p·ε∗(θ,φ,σ)e−ik·r0+iωt0forα≤Ωn≤β,\n0 otherwise,\nC=2π\n3/parenleftbig\nβ4−α4/parenrightbig\n. (19)\nThis state produces the maximum electric energy density at ( r0,t0), with the electric field\nat (r0,t0) polarized along p.\nTo apply the above result to the classical regime, let\nCN=e−/angbracketleftN/angbracketright/2/angbracketleftN/angbracketrightN/2\n√\nN!. (20)\nTogether with a factorizable Φ Nin Eq. (19), the quantum state becomes a coherent state in\nthe continuous mode representation [12], and the mean electric fie ld is then given by\nE(r,t) =i/parenleftbigg/angbracketleftN/angbracketright¯hω0\n2ǫ0λ3\n0/parenrightbigg1/2/integraldisplayβ\nαdΩ/integraldisplayπ\n0dθ/integraldisplay2π\n0dφ\n×/parenleftbig\nΩ3sinθ/parenrightbig1/2ε(θ,φ,σ)fe(Ω,θ,φ,σ)eik·r−iωt+H.c., (21)\nwhich, incidentally, must be an exact solution of the Maxwell equation s. The Fourier trans-\nform ofE(r,t) is proportional to ω3, and the classical power spectrum is then proportional\ntoω6within the allowed frequency band.\nConsider now the magnetic field operator:\nˆB(r,t) =i\n(2π)3/2/summationdisplay\nσ/integraldisplay\nd3k/parenleftbiggµ0¯hω\n2/parenrightbigg1/2\nκ×ε(k,σ)ˆa(k,σ)eik·r−iωt+H.c.,(22)\nwhereκ≡k/k. While the total magnetic energy must be the same as the total elec tric\nenergyforphotonsinfreespace, itisnotdifficulttoshowthatthem agneticenergydensity at\n(r0,t0) is zero where the electric energy density is maximum for the state g iven by Eqs. (19).7\nTo maximize the magnetic energy density instead, we can simply apply t he same procedure\nas above to the magnetic energy density, which turns out to obey t he same bound as the\nelectric one:\n/angbracketleftbigg\n:1\n2µ0ˆB·ˆB:/angbracketrightbigg\n≤π\n3/angbracketleftN/angbracketright/parenleftbigg¯hω2\nλ3\n2−¯hω1\nλ3\n1/parenrightbigg\n. (23)\nThe quantum state with the ultimate magnetic energy density is also s imilar to the electric\ncase,\nΦN=N/productdisplay\nn=1fb(Ωn,θn,φn,σn),\nfb=/braceleftbigg−iC−1/2(Ω3sinθ)1/2m·[κ×ε∗(θ,φ,σ)]e−ik·r0+iωt0forα≤Ωn≤β,\n0 otherwise,(24)\nwheremis the unit vector of the magnetic field at ( r0,t0).\nThe ultimate electric and magnetic energy densities may be challenging to achieve ex-\nperimentally, as they require a power spectrum proportional to ω6within the allowed band,\nspatial focusing in all directions with a specific angular spectrum, an d polarization control.\nThat said, the results set forth impose fundamental limits to which t he energy densities\ncan reach regardless of technological advances in the control of electromagnetic fields, and\ntherefore should prove useful for designing ultrafast optics exp eriments.\nThe author would like to acknowledge financial support from the William M. Keck Foun-\ndation Center for Extreme Quantum Information Theory.\n∗Electronic address: mankei@mit.edu\n[1] T. Brabec and F. Krausz, Rev. Mod. Phys. 72, 545 (2000).\n[2] G. A. Mourou, T. Tajima, and S. V. Bulanov, Rev. Mod. Phys. 78, 309 (2006).\n[3] S. T. Cundiff and J. Ye, Rev. Mod. Phys. 75, 325 (2003).\n[4] M. Tabak et al., Phys. Plasmas 1, 1626 (1994).\n[5] W. Denk, J. H. Strickler, and W. W. Webb, Science 248, 73 (1990); W. R. Zipfel, R. M.\nWilliams, W. W. Webb, Nat. Biotechnol. 11, 1369 (2003).\n[6] T. Juhasz, F. H. Loesel, R. M. Kurtz, C. Horvath, J. F. Bill e, and G. A. Mourou, IEEE J.\nSel. Topics Quantum Electron. 5, 902 (1999).8\n[7] A. H. Zewail, J. Phys. Chem. A 104, 5660 (2000).\n[8] R. W. Ziolkowski, Phys. Rev. A 39, 2005 (1989); R. W. Hellwarth and P. Nouchi, Phys. Rev.\nE54, (1996).\n[9] W. O. Amrein, Helv. Phys. Acta 8, 2684 (1969); C. Adlard, E. R. Pike, and S. Sarkar, Phys.\nRev. Lett. 79, 1585 (1997); I. Bialynicki-Birula, Phys. Rev. Lett. 80, 5247 (1998).\n[10] K. W. Chan, C. K. Law, and J. H. Eberly, Phys. Rev. Lett. 88, 100402 (2002).\n[11] M. Tsang, e-print arXiv:0802.0516v1.\n[12] L. Mandel and E. Wolf, Optical Coherence and Quantum Optics (Cambridge University Press,\nCambridge, UK, 1995).\n[13] L. W. Couch, Digital and Analog Communication Systems (Prentice Hall, New Jersey, 2001)." }, { "title": "0803.0821v1.Finite_doping_of_a_one_dimensional_charge_density_wave__solitons_vs__Luttinger_liquid_charge_density.pdf", "content": "arXiv:0803.0821v1 [cond-mat.str-el] 6 Mar 2008Finite doping of a one-dimensional charge density wave: sol itons vs. Luttinger liquid\ncharge density\nYuval Weiss, Moshe Goldstein and Richard Berkovits\nThe Minerva Center, Department of Physics, Bar-Ilan Univer sity, Ramat-Gan 52900, Israel\nThe effects of doping on a one-dimensional wire in a charge den sity wave state are studied using\nthe density-matrix renormalization group method. We show t hat for a finite number of extra\nelectrons the ground state becomes conducting but the parti cle density along the wire corresponds\nto a charge density wave with an incommensurate wave number d etermined by the filling. We find\nthat the absence of the translational invariance can be disc erned even in the thermodynamic limit,\nas long as the number of doping electrons is finite. Luttinger liquid behavior is reached only for a\nfinite change in the electron filling factor, which for an infin ite wire corresponds to the addition of an\ninfinite number of electrons. In addition to the half filled in sulating Mott state and the conducting\nstates, we find evidence for subgap states at fillings differen t from half filling by a single electron or\nhole. Finally, we show that by coupling our system to a quantu m dot, one can have a discontinuous\ndependence of its population on the applied gate voltage in t he thermodynamic limit, similarly to\nthe one predicted for a Luttinger liquid without umklapp pro cesses.\nPACS numbers: 73.21.Hb,71.55.-i,71.45.Lr\nI. INTRODUCTION\nThe transition between a Mott insulator and a metal-\nlic phase in one-dimensional (1D) wires has attracted no-\ntable interest in recent years1. Different types of transi-\ntions are possible, controlled by various physical param-\neters, such as the electron-electron interaction range and\nitsstrength. SincetheMottstateexistsonlyforcommen-\nsurate fillings, the electron filling factor inside the wire is\nan important parameter, and it can be varied either by\ncontrolling the chemical potential, or by doping with a\nfinite number of electrons. However, these two methods\nare different: doping with an infinitesimally small num-\nber of electrons breaks immediately the Mott insulator\nstate, while due to the Mott gap, a finite change in the\nchemical potential is required in order to insert the first\nelectron and cause the transition2.\nThe metallic phase resulting from doping is usu-\nally described by the Tomonaga-Luttinger liquid (TLL)\ntheory2,3, although for a small finite doping, a large cur-\nvatureofthe elementaryexcitationspectrumis expected,\nand deviations from the TLL theory are possible. In the\nTLL, manyphysicalpropertiesaredetermined by the pa-\nrameterKwhich describes the interactions. Recently, a\ndiscontinuity in the occupation of a resonant level which\nis coupled to a TLL with K <1/2 was predicted by\nFurusaki and Matveev4. However, since the K <1/2\nregime corresponds to very strong repulsive interactions,\nneither experimental nor numerical evidence for such a\njump has been obtained so far.\nAs a generic model for the wire it is convenient to use\na tight-binding description of a 1D lattice with spinless\nelectrons. When repulsive interactions between nearest-\nneighbor electrons are considered, the Mott state, taking\nthe form of a charge density wave (CDW), occurs for\nstrong enough interactions. The Hamiltonian of such asystem can be written as\nˆHwire=−tL−1/summationdisplay\nj=1(ˆc†\njˆcj+1+H.c.) (1)\n+IL−1/summationdisplay\nj=1(ˆc†\njˆcj−1\n2)(ˆc†\nj+1ˆcj+1−1\n2),\nwhereIdenotes the nearest-neighbor interaction\nstrength, and the hopping matrix element between near-\nest neighbors, t, sets the energy scale. ˆ c†\nj(ˆcj) is the\ncreation (annihilation) operator of a spinless electron at\nsitejin theL-site wire, and a positive background is\nincluded in the interaction term.\nThe model of Eq. (1) is equivalent to that of the XXZ\nspin 1/2 chain by the Jordan-Wigner transformation5.\nThe XXZ model is exactly solvable using the Bethe\nansatz6,7. From this equivalence it is known that for\na half filled system with periodic boundary conditions, a\nphase transition between a TLL phase and a CDW one\noccurs at I= 2t. Of course, for sufficiently long wires\nthe type of the boundary conditions does not change this\nresult8,9. The interactionparameter Kis then givenby10\nK=π\n2cos−1(−I/2t), (2)\nso that in the TLL phase, I <2tandK >1/2, while for\nI >2t, the half filled wire is no longer in the TLL phase,\nbut is rather in a CDW state.\nThe competition between the TLL and the CDW\nphases is attributed to the presence ofumklapp processes\nin this model. For half filling the CDW phase wins over\nthe TLL phase once the interactions are strong enough.\nHowever, when the wire is not exactly half filled, a CDW\nstate cannot emerge since it demands a commensurate\nfilling, and thus the TLL description is valid even for in-\nteraction values which are greater than 2 t. As a result,2\nfor strong enough interactions, and sufficiently close to\nhalf filling, one can then get values of Kwhich are less\nthan 1/2.11\nAnother approach to treat the Hamiltonian of Eq. (1)\nis to map it, using bosonization2, into that of the sine-\nGordon model12. In this model, the elementary excita-\ntions are solitons, which carry half an electron charge2.\nIn the vicinity of the Mott state the value of the TLL\ninteraction parameter is K→1\n4+. The value K= 1/4,\nwhich cannot be obtained using this model, corresponds\nto the Luther-Emeryline, for which the solitonsare effec-\ntively non interacting13. For small deviations from half\nfilling the solitons are only weakly interacting.\nIn this paper we investigate the effects of a finite dop-\ning on a 1D Mott state. We show that although a small\nnumber of electrons results in an insulator-metal transi-\ntion, thechargedistributioninthe metallicsystemwith a\nfinite number of electrons is not uniform, but rather cor-\nresponds to an incommensurate CDW. The added charge\ncan also be considered as delocalized solitons in the orig-\ninal commensurate CDW, whose spatial distribution is\ndetermined by boundary conditions and symmetry con-\nsiderations. Nevertheless, the system is conducting, and\nonceabiasvoltageis appliedthe solitonsarefreeto move\nand transport charge across the wire.\nAn exceptional behavior is exhibited when a lattice\nwith an even number of sites L= 2pis occupied by p±1\nelectrons, i.e., there is a single additional electron (or\nhole) relative to half filling. As will be shown, in this\ncase the insulating behavior of the original Mott state is\nretained even though the filling is incommensurate. This\nunusualcasewill be explained asaresultofsubgapstates\nin the soliton energy spectrum.\nIn addition, we show that since the Mott state with a\nsmall doping allows for the interaction parameter in the\nrange 1/4< K < 1/2, it might exhibit the Furusaki-\nMatveev discontinuity. The appearance of that jump is\nnot a foreclosed conclusion, since the model used by Fu-\nrusaki and Matveev does not contain umklapp processes.\nNevertheless, we find that the jump indeed occurs, and\nshow its implications on the solitons present in the wire\nnear half filling.\nSince the CDW state is obtained for exactly half fill-\ning, it is convenient to denote the extra charge by Q=\nne−L/2, where neis the number of electrons in the sys-\ntem. As will be shown, the parity of Lplays a significant\nrole in the behavior when adding the first few electrons.\nUsing the finite-size version of the density-matrix renor-\nmalization group (DMRG) method14,15the Hamiltonian\nˆHwirewas diagonalized and the ground state was calcu-\nlated for different values of the charge Qand the lattice\nsizeL. In the current paper we present results obtained\nforI= 3t, which is deep inside the CDW regime (for\nhalf filling). Nevertheless, other interaction strengths in\nthe regime of I >2thave been checked as well, and were\nfound to lead to qualitatively similar results.\nThe outline of the paper is as follows: in the next sec-\ntion, Sec. II, we present the charge occupation alongnearly half filled wires, which demonstrates the presence\nof incommensurate CDW or solitons. In Sec. III we dis-\ncuss the excitation spectrum of the solitons, and the ad-\ndition spectrum of the system. For systems which are ex-\nactly half filled, or doped with only one electron or hole,\nwe show that the ground state is insulating; otherwise, it\nis conducting. In Sec. IV we discuss the existence of the\nFurusaki-Matveev jump when the doped wire is coupled\nto a resonant level. Finally we conclude in Sec. V.\nII. CHARGE DISTRIBUTION\nWe begin by presenting the occupation of electrons\nalong the wire for different numbers of extra electrons Q\nfor wires with an odd or an even number of sites (Fig. 1).\nTheQ= 0 case (for an even L) results in a flat distri-\nbution of nj= 1/2 for each lattice site j(dashed line).\nThisstateisalinearcombinationoftwodegenerateCDW\nstates, which are eigenstates of the Hamiltonian in the\nthermodynamic limit. These CDW states, which are also\nshown in Fig. 1, are numerically obtained by applying\nan infinitesimally small potential on the first site, which\nbreaks the degeneracy between them8. The population\nof such a CDW state along the wire (near the center) can\nbe written as\nnj=n+Acos(2πnj+φ), (3)\nwhere the amplitude Adepends on the interaction\nstrength, n= 1/2 is the filling, and φis a phase.\nWhenQ/negationslash= 0, the ground state is no longer degenerate\nand the electron occupation throughout the lead is not\nuniform. Let us start with the Q= 1/2 case (when Lis\nodd), which results in a true CDW state, i.e., for each\neven (odd) site the occupation is low (high). Unlike the\ncase of the degenerate ground states for Q= 0, the Q=\n1/2 ground state is not coupled by the Hamiltonian to\nthe similar CDW ground state of Q=−1/2 due to the\ndifference in their total population.\nIn order to explore the addition of electrons onto the\nCDW states, one can take the uniform distribution of the\nQ= 0 state as a reference, and investigate the difference\nfromitbycalculatingtheaccumulatedpopulation∆ nj=/summationtext\ni≤jni−j/2. As can be seen in Fig. 2, ∆ njfor each\nof the two clean CDW states of Q= 0 (presented in\nsolid lines), has an extra charge of e/4 localized at one of\nthe wire edges, and a compensating charge (i.e., −e/4)\nlocalized at the other. The difference between these two\nstates is thus a soliton (of charge e/2) located at one\nedge of the wire and an antisoliton (of charge −e/2) at\nthe other edge.\nBy taking one of the Q= 0 CDW states, and locating\na soliton at its negatively charged edge, while leaving\nthe other edge as it is, one gets a new CDW state in\nwhich both edges have a positive charge, i.e., a CDW\nstate having Q= 1/2. Similarly, placing an antisoliton\nat the edge with the positive charge of the Q= 0 state3\n0 100 200 300 400 500j0.20.30.40.50.60.70.8nj0\n1\n2\n3\n0 100 200 300 400 500j0.20.30.40.50.60.70.8nj1/2\n3/2\n5/2\nFIG. 1: (Color online) The electrons occupation in a lattice of\n500sites (upperpanel)and499sites (lower panel)for differ ent\nnumber of electrons. The values near the curves denote the\nextra charge Qwhich varies between Q= 0 for half filling and\nQ= 3.\nresults in the Q=−1/2 CDW state. These states are\nalso shown in Fig. 2 (dashed lines).\nFor theQ= 1 case (when Lis even), adding a local-\nized soliton near one of the edges is not sufficient, and it\nrequiresthe addition ofanotherchargeof e/2. This extra\ncharge is obtained by the formation of a delocalized soli-\nton, centered at the middle of the wire. Further increase\nof the filling above Q= 1/2 (Q= 1) for odd (even) wire\nlengths, results in the addition of two delocalized solitons\nfor everyelectron, sothat the total number ofdelocalized\nsolitons is 2 Q−1.\nThe delocalized solitons, which carry a charge of e/2\neach, are free to move along the wire. However, when\nno bias voltage is applied, the charge distribution across\nthe wire is fixed by the boundaries and by symmetry\nconsiderations. Practically these constraints lead to a\ndensitywavewith acosineterm similarto that ofEq.(3),\nin which the filling factor nis modified. This statement\nis true for much larger values of Qas well. For instance,\nFig. 3 shows the electron distribution for a 300-site wire\nwithQ= 14, i.e., with 27 delocalized solitons. It is easy\nto see that although the amplitude of the oscillations is50 100 150 200 250 300\nj-0.5 -0.5-0.25 -0.250 00.25 0.250.5 0.5∆nj\nFIG. 2: (Color online) The extra charge distribution along a\nwire of length L= 300 sites with Q= 0 (solid lines) and of\nL= 299 sites with Q=±1/2 (dashed lines).\n0 50 100 150 200 250 300\nj00.20.40.60.8nj\n100 120 140 160 180 2000.450.50.550.60.65\nFIG. 3: (Color online) The electrons occupation in a lattice of\n300 sites with additional 14 electrons. A fit to Eq. (3) (done\nover the middle region) is shown in the inset (symbols).\nreduced, it doesn’t vanish and it is rather constant near\nthe middle of the wire. Moreover, the oscillations can\nstill be fitted to Eq. (3) using nas a fitting parameter.\nAs can be expected, this results in a value of nwhich\nmatches the filling of the wire.\nThe non-uniform charge distribution for a fixed num-\nber of electrons is not a finite-size effect, and survives in\nthe thermodynamic limit as well. In Fig. 4 one can see\na comparison between wires of 300 and 600 sites, with\nQ= 14 and Q= 28. One can see weaker oscillations\nin theQ= 28 case, attributed to a smaller wave length.\nOn the other hand, increasing the size of the lead from\nL= 300 to L= 600 while maintaining the number of\nextra electrons Qconstant, results in an increase of the\noscillations.\nOne can thus conclude that on the one hand the oscil-4\n240 260 280 300 320 340 360\nj0.50.550.6njL=600 Q=28\nL=600 Q=14\nL=300 Q=14\nFIG. 4: (Color online) The electrons occupation near the mid -\ndle of the system in lattices of 300 or 600 sites with addition al\n14 or 28 electrons. Results for the 300-site wire are shown as\na function of 2 j.\nlations are expected to vanish if a finite filling fraction is\nconsidered, i.e., in the limit of Q→ ∞,L→ ∞, keeping\nQ/Lconstant. On the other hand, when a finite number\nof electrons is added, the oscillations are expected to be\nnoticed even in the limit of L→ ∞.\nIII. EXCITATION SPECTRUM\nUp to now we have shown that the density distribution\nof the electrons along the wire, when it is doped by a\nfinite number of electrons, is not flat, and preserves some\nfeatures of the Mott state. One can still wonder whether\nthese doped states retain the insulating behavior of the\nMott state or not. To clarify that point we study the size\ndependence of the addition spectrum, defined through\n∆2(Q) =E0(Q+1)−2E0(Q)+E0(Q−1),(4)\nwhereE0(Q) is the ground-state energy of the wire with\nQelectrons above half filling. For a Mott state the limit\nof ∆2, asL→ ∞, should be a finite value, corresponding\nto the gap size, while for a conducting state one expects\n∆2→0.\nThe dependence of ∆ 2(Q) on the wire length Lis pre-\nsented in Fig. 5 for even wire sizes (lines) and for odd\nsizes (filled symbols), for different number of electrons\nQ. As can be seen, for Q= 0 and 1 /2, the values of\n∆2converge to the value ∆ 2(∞) (presented by the as-\nterisk symbol at the left margin of the figure), given by\nthe Bethe ansatz6,7. On the other hand, for Q≥3/2\none gets ∆ 2→0, which indicates that these states are\nconducting.\nA deviation from this intuitive picture appears for\nQ= 1, in which an unexpected gap of ∆ 2(∞)/2 occurs.\nNevertheless, this result can be explained by the spec-\ntrum of soliton states near half filling. It is known that\nfor solitons in the sine-Gordon model with open bound-\nary conditions there are two degenerate subgap states at0.001 0.01\n1/L00.10.20.3∆2Q=0\nQ=1\nQ=2\nQ=3\n0.001 0.01\n1/L00.10.20.3∆2Q=1/2\nQ=3/2\n0.001 0.01\n1/L00.10.20.3∆2∆2(∞)\n∆2(∞)/2Even Odd Bethe ansatz\nFIG. 5: (Color online) The dependence of the addition spec-\ntrum on the lattice size L, for different fillings, comparing the\ncases of evenand oddsizes and theexact Bethe ansatz results .\nNote the semi-logarithmic scale.\nzero energy, positioned exactly between the conduction\nand the valence bands, each of them localized near one\nof the edges of the system16. In our case these subgap\nstates are thus the localized solitons, which we identified\nas the difference between the two CDW states of Q= 0\npresented in Fig. 1.\nThusthe energyspectrumofthe solitonsisrepresented\nby valence and conduction bands separated by a gap ∆,\nwhile the subgap states exist at ∆ /2 above the valence\nband (see schematic picture in Fig. 6). The filling of the\nsubgap states with the localized solitons does not change\nthe total energy. However, every additional soliton in-\ncreases the energy in ∆ /2 (in the limit L→ ∞).\n-∆/2∆/2\n0E\n∆\nFIG. 6: (Color online) A schematic picture of the energy\nbands of the solitons according to Ref. 16. The bands are\nseparated by energy ∆, and two discrete states exist at zero\nenergy.\nFor an odd L, the states with Q=±1/2, which have\nthe same energy due to the particle-hole symmetry of\nthe Hamiltonian, differ only in the occupation of the two\nlocalized solitons. The addition of any other electron is5\nequivalent to the addition of two delocalized solitons, so\nthat the energy difference between successive fillings is\n∆. Therefore, E0(1/2)−E0(−1/2) = 0 and E0(3/2)−\nE0(1/2) = ∆, resulting in ∆ 2(1/2) = ∆. For higher\nvalues of Q, one gets E0(Q+ 1)−E0(Q) = ∆, so that\n∆2(Q≥3/2) = 0. These results are summarized in\nTable I.\nQ−1\n21\n23\n25\n2Q′>5\n2\nE000∆2∆(Q′−1\n2)∆\n∆2∆∆000\nTABLE I: Addition spectrum for wires with an odd number\nof sites in the limit L→ ∞.\nOn the other hand, when Lis even, the Q= 0 state\ncorresponds to the occupation of only one localized soli-\nton (or, more precisely, to a linear combination of two\nstates, in each of them one edge soliton state is filled and\nthe other is empty). The addition of the first extra elec-\ntronfillstheadditionallocalizedsolitonstateandasingle\ndelocalized soliton state, so that E0(1)−E0(0) = ∆/2.\nEvery additional electron adds two delocalized solitons,\nthus forQ >1 one gets E0(Q+1)−E0(Q) = ∆. There-\nfore, asonecanseein TableII, ∆ 2(0) = ∆, ∆ 2(1) = ∆/2,\nand ∆ 2(Q≥2) = 0.\nQ−10123Q′>3\nE0∆\n20∆\n23∆\n25∆\n2(Q′−1\n2)∆\n∆2∆\n2∆∆\n2000\nTABLE II: Addition spectrum for wires with an even number\nof sites in the limit L→ ∞.\nBefore commencing with the observation of Furusaki-\nMatveev jump we briefly summarize the results so far.\nWe have demonstrated the existence of CDW states for\nQ= 0 and Q=±1/2, and shown that these states are\ninsulating. The ground states with Q=±1 were found\nto be insulating as well, with an excitation gap which is\nhalf of the Mott gap. States with Q >1 are conduct-\ning and thus should be generally described by the TLL\ntheory. Nevertheless, we have found that the spatial dis-\ntribution of the electron density is not uniform as in the\nregular TLL picture. Furthermore, this distribution can\nbe fitted to an incommensurate CDW form. It is there-\nfore interesting to explore some other predictions of the\nTLL theory for such states.\nIV. FURUSAKI-MATVEEV JUMP\nThe unconventionalbehavior of the gaplessstates with\nthe non-uniform density can be illustrated by studying\nthe discontinuity in the occupation of a resonant level\ncoupledtothewire4. Asmentionedabove,once K <1/2,the level is expected to show a jump in its occupation as\na function of its energy. Nevertheless, one should note\nthat umklapp processes, which are an essential ingredi-\nent in explaining the behavior of our system, were not\nconsidered in the theoretical framework of Ref. 4.\nFurthermore, since here the particle distribution in the\nuncoupledwireisnotflat, theoccurrenceoftheFurusaki-\nMatveev discontinuity raises an additional question. If\nthe electrons occupation profile along the uncoupled wire\nwas uniform, one should have observed Friedel oscilla-\ntions in the wire as soon as it is coupled to the resonant\nlevel. Thepotentialofthe resonantlevelcanthen be con-\ntinuously changed until the predicted jump in the level\noccupation occurs. At that point, an inversion of the\nFriedel oscillations along the wire is expected. On the\nother hand, once the wire is a Mott state doped by a\nfinite number of electrons and the electron density is not\nuniform, the situation is less clear.\nWe first calculate the value of Kfor the doped Mott\nstate, and show that it is below 1 /2. In order to have\nthe same value of Kfor different lengths of wires, one\nmust consider an equal density of additional electrons,\nand we choose to work with doping of Q=L/50, with\nwires of sizes 100sites and above. Kmay be obtained2,17\nby the ratio between ∆ E, the energy difference between\nthe groundstate and the first excited state, and the addi-\ntion spectrum ∆ 2, since for spinless electrons with open\nboundary conditions ∆ E=πvc\nLand ∆ 2=πvc\nKL. More ac-\ncurate result may be obtained by fitting both ∆ 2(L) and\n∆E(L) to a polynom in 1 /L, and then obtaining Kfrom\nthe ratio of the linear coefficients18. Using both methods\nwe find that the value of Kfor our wire is 0 .42.\nIn order to represent the coupling of a resonant level\nor an impurity of energy ǫ0to the left edge of the wire\nthe following term is added to the Hamiltonian:\nˆHimp=ǫ0ˆa†ˆa−V0(ˆa†ˆc1+H.c.), (5)\nwhere ˆa†(ˆa) is a creation (annihilation) operator of an\nelectron in the resonant level, and V0is the hopping ma-\ntrix element between the level and the first site of the\nwire. Interaction between the impurity and the wire is\nnot considered here since it does not change the result\nqualitatively, but only causes a renormalization of the\nhopping amplitude V0towards larger values4,8.\nThe level occupation n0as a function of the level en-\nergyǫ0is presented in Fig. 7 for different wire lengths\nusingV0= 0.2t. As mentioned above, wires in different\nlengths contain different number of additional electrons\nusingQ=L/50, andK= 0.42. When ǫ0is much larger\nthan the chemical potential in the wire µ, the impurity is\nalmost empty, and the wire contains most of the Qextra\nelectrons. On the other hand, for ǫ0≪µthe impurity\nis almost entirely occupied, and then only Q−1 extra\nelectrons are in the wire.\nAlthough the population of the impurity is found to\nbe continuous for all finite wire lengths studied, it is ex-\npected to display an abrupt jump in the thermodynamic\nlimit,L→ ∞. In order to demonstrate this we study the6\n0.35 0.4 0.45 0.5 0.55ε000.20.40.60.81n0\n100150200250300\nL-150-100-50slope\nFIG. 7: (Color online) The occupation of an impurity which\nis coupled to one end of 1D lattices of different sizes. The\nresults shown are for lattice sizes between 100 and 300 (from\nleft to right) in steps of 50. In order to compare cases with\nidentical values of K, the number of additional electrons is\ntaken as Q=L/50, so that K= 0.42. Inset: the slope near\nthe point where n0= 1/2, which shows a linear dependence\non the system size.\ndependence of the occupation slope near n0= 1/2 on the\nwire length L. As is clearly seen in the inset of Fig. 7,\nthe slope scales linearly with L, which is a clear sign of\na first order transition in the thermodynamic limit19.\n50 100 150jnj\nFIG. 8: (Color online) The transition of an electron from the\nresonant level into the wire, which leads to the formation of 2\nadditional solitons. The results shown are for a level coupl ed\nto a 150-site wire, with Q= 3. Different curves correspond\nto values of ǫ0between 0 .498 (top) to 0 .510 (bottom) with\noffsets in the vertical axis for clarity.\nThe finite size transition region, in which the electrontransfersfromthe resonantlevel into the wire, is interest-\ning by itself, since the addition of a single electron to the\nwire is related to the appearance of two additional delo-\ncalized solitons in the electron population inside it. As\ncan be seen in Fig. 8, as ǫ0increases the electron bound\nto the impurity tunnels into the wire and additional two\nsolitons are created in the wire. As mentioned above,\nthis is in sharp contrast to the case of a resonant level\ncoupled to a TLL having a uniform charge distribution,\nin which only a local change of the Friedel oscillations in\nthe charge distribution of the wire is expected.\nV. CONCLUSIONS\nIn this paper we have shown that doping a Mott state\nwith a finite number of electrons yields states which pre-\nserve the charge modulations of the density wave even\nfor an infinite system size. The charge carriers are soli-\ntons, whose spectrum fits the known predictions of the\nsine-Gordon model.\nWe have demonstrated how the solitons are added to\nthe wire. The first charge of e/2 above half filling results\nin the appearance of a localized soliton at one of the\nwire edges. Each additional charge of e/2 results in an\nadditional delocalized soliton, so that for charge Qone\ngets a wire with 2 Q−1 delocalized solitons and a single\nlocalized one.\nThe ground state of the pure Mott state, which is ex-\nactly half filled, is obviously insulating. For wires of odd\nsizes, the ground states for Q=±1/2 are insulating as\nwell, and the excitation spectrum in these three cases\nhas the regular Mott gap. Additional insulating states\nwith a non-trivial half gap were found for Q=±1. This\nunusual gap was explained according to the spectrum of\nthe soliton states in a 1D wire with open boundaries. For\nQ >1, the excitation spectrum was found to be gapless.\nThe unconventional behavior of wires with Q >1,\nwhich are gapless but do not preserve translational in-\nvariance of charge, is checked against a prediction for a\nTLL wire with an interaction parameter K <1/2 cou-\npled to a resonant level. We find that the TLL prediction\n(i.e., the Furusaki-Matveevjump) is indeed obtained also\nfor our system, although it involves the creation of two\nadditional delocalized solitons in the charge distribution\nof the wire, as opposed to the inversion of Friedel oscil-\nlations in the wire characterizing the conventional TLL\nscenario. As a final remark we note that it may be inter-\nesting to investigate, for this regime of parameters, some\nother TLL predictions as well.\nAcknowledgments\nWe thank M. Pepper and M. Kaveh for useful discus-\nsions, and the Israel Science Foundation (Grant 569/07)\nfor financial support.7\n1N. F. Mott, Metal Insulator Transitions (Taylor and Fran-\ncis, London, 1990).\n2T. Giamarchi, Quantum Physics in One Dimension (Ox-\nford University Press, New York, 2003).\n3J. Voit, Rep. Prog. Phys. 57, 977 (1994).\n4A. Furusaki and K. A. Matveev, Phys. Rev. Lett. 88,\n226404 (2002).\n5P. Jordan and E. Wigner, Z. Phys. 47, 631 (1928).\n6H. A. Bethe, Z. Phys. 71, 205 (1931).\n7R. J. Baxter, Exactly Solved Models In Statistical Mechan-\nics, (Academic Press, London, 1989).\n8Y. Weiss, M. Goldstein and R. Berkovits, J. Phys.: Con-\ndens. Matter 19, 086215 (2007).\n9Y. Weiss, M. Goldstein and R. Berkovits, Phys. Rev. B 76,\n024204 (2007).\n10F.WoynarovichandH.P.Eckle, J.Phys.A 20, L97(1987);\nC. J. Hamer, G. R. W. Quispel, and M. T. Batchelor, ibid.20, 5677 (1987).\n11F. D. M. Haldane, Phys. Rev. Lett. 45, 1358 (1980).\n12R. Rajaraman, Solitons and Instantons: An Introduc-\ntion to Solitons and Instantons in Quantum Field Theory\n(North Holland, Amsterdam, 1982).\n13A. Luther and V. J. Emery, Phys. Rev. Lett. 33, 589\n(1974).\n14S. R. White, Phys. Rev. B 48, 10345 (1993).\n15U. Schollw¨ ock, Rev. Mod. Phys. 77, 259 (2005).\n16M. Fabrizio and A. O. Gogolin, Phys. Rev. B 51, 17827\n(1995).\n17H. Schulz, Phys. Rev. Lett. 64, 2831 (1990).\n18S. Ejima, F. Gebhard, S. Nishimoto and Y. Ohta, Phys.\nRev. B72, 033101 (2005).\n19K.Binder andD.P.Landau, Phys.Rev.B 30, 1477(1984)." }, { "title": "0805.0542v1.Crossover_in_the_local_density_of_states_of_mesoscopic_SNS_junctions.pdf", "content": "arXiv:0805.0542v1 [cond-mat.supr-con] 5 May 2008Crossoverinthelocaldensity ofstatesofmesoscopic\nsuperconductor/normal-metal/superconductor junctions\nAlex Levchenko\nSchool of Physics and Astronomy, University of Minnesota, M inneapolis, MN, 55455, USA\n(Dated: May 5, 2008)\nAndreev levels deplete energy states above the superconduc tive gap, which leads to the peculiar non-\nmonotonouscrossoverinthelocaldensityofstatesofmesos copicsuperconductor /normal-metal/superconductor\njunctions. Thiseffect isespeciallypronounced inthe case whenthenormal meta l bridge length Lissmallcom-\npared to the superconductive coherence length ξ. Remarkable property of the crossover function is that it\nvanishes not onlyatthe proximityinduced gap ǫgbutalsoatthe superconductive gap ∆. Analytical expressions\nforthe density of states atthe both gapedges, as well as gene ral structure of the crossover are discussed.\nPACS numbers: 74.45. +c\nExperimental advances in probing systems at the meso-\nscopic scale1,2,3,4,5revived interest to the proximity related\nproblemsinsuperconductor–normalmetal(SN)heterostruc -\ntures.6The most simple physical quantity reflecting proxim-\nity effect is the local density of states (LDOS) ρ(ǫ,r), which\ncan be measured in any spatial point rat given energyǫus-\ning scanning tunneling microscopy. The e ffects of supercon-\nductive correlations on the spectrum of a normal metal are\nespecially dramatic in restricted geometries. For example ,\nin the case of superconductor–normal metal–superconducto r\n(SNS) junction, proximity e ffect induces an energy gap ǫg\nin excitation spectrum of a normal metal with the square\nroot singularityρ(ǫ,r)∝/radicalbigǫ/ǫg−1 in the density of states\njust above the threshold ǫ−ǫg≪ǫgRef. 7,8,9,10,11,12\n(here and in what follows, ρwill be measured in units of the\nbare normal metal density of states νat Fermi energy). The\nmost recent theoretical interest was devoted either to meso-\nscopic13,14,16,17orquantum18,19,20,21fluctuation effects on top\nof mean–field results7,8,9,10,11,12that smear hard gap below ǫg\nand lead to the so called subgaptail states with nonvanishing\nρ∝exp/bracketleftbig−g(1−ǫ/ǫg)(6−d)/4/bracketrightbigatǫg−ǫ/lessorsimilarǫg, where g is the\ndimensionlessnormalwire conductanceand dis the effective\nsystem dimensionality. The latter is essentially a nonpert ur-\nbative result that requires instantonlike approach within σ–\nmodel19,20or relies on methodsof random matrix theory.16,18\nSurprisingly, after all of these advances, there is somethi ng\ninteresting to discuss about proximity induced properties of\nthe SNS junctionseven at the level of quasiclassical approx i-\nmationbyemployingUsadelequations.22Thepurposeofthis\nworkistopointoutasubtlefeatureofthecrossoverinthelo -\ncal density of states of mesoscopic SNS junctions. The latte r\nwasseeninsomeearlyandrecentstudies,8,11,12,14,15however,\nneitheremphasizednortheoreticallyaddressed.\nTo this end, consider normal wire (N) of length Land\nwidthWlocated between two superconductive electrodes\n(S). In what follows, we concentrate on di ffusive quasi–one–\ndimensional geometry and the short wire limit L≪ξ, where\nξ=√DS/∆is superconductivecoherencelength, with DSas\nthediffusioncoefficientinthesuperconductorand ∆astheen-\nergygap(hereafter, /planckover2pi1=1). Thecenterofthewireischosento\nbeatx=0 andboundarieswith superconductorsat x=±L/2\ncorrespondingly,where xisthe coordinatealongthewire.\nUnder the condition L≪ξ, the proximity effect is es-/s40 /s44/s32 /s120 /s41\n/s84/s104/s47 \n/s49\n/s103 /s52 \n/s61\n/s103/s61/s66 /s67/s83 /s40 /s41\nFIG. 1: (Color online) Schematic plot for the local density o f states\ncrossover inthe short L≪ξdiffusive SNSjunction.\npecially strong — superconductive correlations penetrate the\nentire volume of the normal region. As the result, the in-\nduced energy gap in normal wire ǫgis large and turns out to\nbe of the same order as the gap in the superconductor itself\nǫg=∆−δ∆,withthefinitesize correction δ∆∼∆3/ǫ2\nTh≪∆,\nwhereǫTh=DN/L2is the Thoulessenergyand DNis the dif-\nfusion coefficient in the normal bridge. The latter should be\ncontrasted to the long wire limit L≫ξ, where the proxim-\nity effect is weak and the induced gap is ǫg∼ǫTh≪∆. Just\nabove the gapǫ−ǫg≪ǫg, density of states has square root\nsingularity, which is similar to the L≫ξcase,7,8,9,10,11,12see\nFig. 1, which is a robust property of quasiclassical approx-\nimation. However, the prefactor for L≪ξis significantly\nenhancedρ(ǫ,x)∝(ǫTh/∆)2/radicalbigǫ/ǫg−1(noteherethat propor-\ntionalitysignimpliescharacteristicenergydependence, while\nthe exact numerical coe fficientis differentfor a givencoordi-\nnatexalong the wire). The value of ǫgis a property of the\nspectrum,thusit is xindependent.\nOne would naturally expect that above the proximity in-\nduced gapǫg, local density of states goes through the max-\nimumρmax=ρ(ǫ=∆,x) and then crosses over to the BCS2\nlike DOSρBCS=ǫ/√\nǫ2−∆2atǫ >∆, which finally satu-\nrates to unityρ→1 atǫ→∞. Surprisingly, however, the\ncrossover scenario is di fferent. The density of states indeed\nreachesthemaximum,whichoccursat ǫ∼∆−δ∆/2,butthen\ndecreasesand vanishesto zeroat superconductivegap ∆with\nquarctic power law behavior ρ(ǫ,x)∝(ǫTh/∆)3/24√|ǫ/∆−1|\nfor|ǫ−∆|/lessorsimilarδ∆. Finally,ρ(ǫ,x) grows back forǫ>∆and at\nthe energyscaleǫ∗∼∆+δ∆crosses overto the ρBCS. Indeed,\nobserve thatρBCS(ǫ∗)∼ǫTh/∆∼ρ(ǫ∗,x), see Fig. 1. It is\nimportant to mention here that the discussed feature appear s\nonlyatthelevelof localdensityofstates. Energydependence\nof theglobaldensity of states, which is integrated over the\nentire volume, does not show dip at ǫ=∆(this point will be\ndiscussedlaterin thetext).\nIt seems that crossover picture presented in the Fig. 1 was\nalready numerically seen in previous studies,11,12,14,15how-\never the origin of the soft gap at ǫ=∆was never addressed.\nOneshouldalsonotethatnumericswasperformedalwaysfor\nthe not too short junctions L/greaterorsimilarξ. It is certainly unusual to\nfindzerointhedensityofstatesatthegapedge ∆. Inorderto\nget someinsightintothisbizarrefeature,let usconsidera toy\nmodel, which crudelymimicsthe system underconsideration\nand gives some hints for qualitative understanding. Quanti ta-\ntivetheoryoftheoutlinedcrossoverwillbedevelopedfoll ow-\ningthisdiscussion.\nImagine chaotic quantum dot (QD) sandwiched between\ntwo superconductors. A quantum dot is really a zero dimen-\nsional system, in the sense that L/ξ→0 (or equivalently,\n∆/ǫTh→0). However, one will momentarily see that finite\ntunnelingrateintothedot ΓplaysaneffectiveroleofThouless\nenergyǫTh. By following Ref. 23 one may construct scatter-\ning states and determine quasiparticles excitation spectr um.\nIf in addition we assume that QD supports only one trans-\nverse propagating mode, then the discrete spectrum obtaine d\nfromthepolesofthescatteringmatrix S(ǫ)consistsofasingle\nnondegenerateAndreev state at energy ǫo∈[0,∆], satisfying\nΩ(ǫ)+Γǫ2√\n∆2−ǫ2=0,where the function Ω(ǫ) is defined\nbyΩ(ǫ)=(∆2−ǫ2)(ǫ2−Γ2/4). The density of states in the\nsuperconductor–quantumdot–superconductorsystemisgiv en\nbyρ(ǫ)=ρBCS(ǫ)+δρ(ǫ),where\nδρ(ǫ)=1\n2πid\ndǫlnΩ(ǫ)+iΓǫ2√\nǫ2−∆2\nΩ(ǫ)−iΓǫ2√\nǫ2−∆2,(1)\nwhich follows from the relation ρ(ǫ)=ρBCS+\n(1/2πi)(d/dǫ)lnDetS(ǫ)betweenDOSandthescatteringma-\ntrix.24It is easy to check that in the limit Γ≫∆Andreev\nlevel is positioned at ǫo≈∆−8∆3/Γ2, which resembles the\nexpression forǫgin the case of the di ffusive wire discussed\nabove,whereΓindeedplaystheroleofThoulessenergy. The\npresence of an Andreev level changes density of states ρ(ǫ)\nin two ways. The first contribution δρ1(ǫ) originating from\ndlnDetS/dǫtermofEq.(1)belongstothesubgappartofthe\nspectrumǫ∈[0,∆]andhasstructureoftheform\nδρ1(ǫ)=1\n2πid\ndǫln/parenleftiggǫ−ǫo−i0\nǫ−ǫo+i0/parenrightigg\n=δ(ǫ−ǫo),(2)\nwhichis nothingelse but DOS associated with the singleAn-\ndreev state. Most interestingly, Andreev level changes ρ(ǫ)above the superconducting gap as well. Indeed, it follows\nfrom the Eq. (1) that at energies ǫ>∆the density of states\nρ(ǫ)getsthecorrection\nδρ2(ǫ)=−ρBCS(ǫ)Γ\nπ/parenleftbig2∆2−ǫ2/parenrightbigΓ2/4−ǫ2/parenleftbigǫ2+∆2/parenrightbig\n/parenleftbigǫ2−∆2/parenrightbig/parenleftbigǫ2−Γ2/4/parenrightbig2+Γ2ǫ4.(3)\nObservethatintheimmediatevicinityofthesuperconducti ve\ngapǫ−∆/lessorsimilar∆3/Γ2thecorrectionδρ2(ǫ)≈−(Γ/4π∆2)ρBCS(ǫ)\nisnegative, implying that Andreev level suppresses bulk su-\nperconductivedensity of states ρBCS(ǫ) abovethe gap. At en-\nergyǫ∗∼∆+∆3/Γ2the correction given by Eq. (3) reaches\nitsmaximum,while ρBCSisrecoveredatlargeenergies ǫ/greaterorsimilarΓ,\nwhereδρ2decays asδρ2(ǫ)≈Γ/πǫ2. Based on this example,\nit appears that Andreev level tends to depletebulk BCS den-\nsity of states at energies ǫ−∆/lessorsimilar∆3/Γ2. One shouldnote that\nthe Andreev level leads to pure redistribution of the energy\nstates — there are no additional states abovethe gap; indeed , /integraltext∞\n∆δρ2(ǫ)dǫ≡0. If QD supports not only one but a large\nnumberof the transversepropagatingchannels, then cummu-\nlativenegativeδρ2(ǫ) may compensate ρBCS(ǫ), which leads\nto the vanishing density of states ρ(ǫ) at the gap edgeǫ=∆,\nsee Fig.1.\nHaving discussed the qualitative picture of the crossover,\nlet us now turn to the quantitative description. The quasi-\nclassical approachto di ffusive SNS structuresis based on the\nUsadel equation22for the retarded Green’s function GR(r,ǫ).\nForthelatter,wewill employangularparametrization25GR=\nτzcosθ+τxsinθcosφ+τysinθsinφ, whereτiis the set of\nPaulimatrices. Intheabsenceofthephasedi fferencebetween\nthe S terminals, one can set φ≡0 and Usadel equations for\nquasi–one–dimensionSNS geometryacquiresthe form\n/braceleftigg\nDN/parenleftbigd2θN/dx2/parenrightbig+2iǫsinθN=0|x|/lessorequalslantL/2\nDS/parenleftbigd2θS/dx2/parenrightbig+2iǫsinθS+2∆cosθS=0|x|>L/2(4)\nwhereθN(S)(ǫ,x)aretheGreen’sfunctionanglesinN(S)parts\nof the junction, correspondingly. In writing Eq. (4), we as-\nsumedstepfunctionpair–potential ∆(x)=∆η(|x|−L/2),with\nη(x)=1 ifx>0andη(x)=0 otherwise. Theapplicabilityof\nthisapproximationreliesontheconditionthatthewidth Wof\nthejunctionissmallcomparedtothecoherencelength. Inth is\ncase,nonuniformitiesin ∆(r)extendonlyoverthedistanceof\norderWfrom the junction, which is due to the geometrically\nconstrained influence of the narrow junction on the bulk su-\nperconductor.26\nIn the absence of additional tunnel barriers at SN–\ninterfaces, Eqs. (4a) and (4b) are supplemented by the fol-\nlowingboundaryconditions:27\nθN(ǫ,x)|x=±L/2=θS(ǫ,x)|x=±L/2, (5a)\nσNdθN(ǫ,x)\ndx/vextendsingle/vextendsingle/vextendsingle/vextendsingle/vextendsinglex=±L/2=σSdθS(ǫ,x)\ndx/vextendsingle/vextendsingle/vextendsingle/vextendsingle/vextendsinglex=±L/2,(5b)\nwhereσN(S)aretheconductivitiesofnormalmetalandsuper-\nconductor. Knowing solutionsof Usadel equations, one finds\nlocaldensityofstates fromthegeneralexpression\nρ(ǫ,x)=Re[cosθ(ǫ,x)]. (6)3\nFor future convenience,we rotate the Green’s function an-\ngles asθN=π/2+iϑNandθS=θBCS+iϑS, whereθBCS=\nπ/2+iarsinhγwithγ=ε/√\n1−ε2,andintroducedimension-\nless variablesε=ǫ/∆,λ=x/L. After the rotation, Usadel\nEq. (4b) for superconducting sides of the junction becomes\nrealandcanbeeasilyintegrated,providing\nϑS(ε,λ)=4artanh/braceleftbigexp[−(|λ|−1/2)/ξS]/bracerightbig,(7)\nwhere we have introduced the energy depending supercon-\nductive coherence length ξ−1\nS=/radicalig\n2∆DN\nǫThDS4√\n1−ε2. Equation\n(4a) mayalso be exactly solvedin terms of elliptic function s;\nhowever, this exact solution is not needed for our purposes.\nIndeed, observe that it follows from the boundary condition\n[Eq. (5a)] that in the energy range 1 −ε≪1, the normal\nmetal phase Re/bracketleftbigϑN(ε,λ)/bracketrightbig∝arcsinhγ≈ln/radicalig\n2\n1−ε≫1 is large\neverywhere in the N part of the junction. Thus, one may ap-\nproximate sinθN≈exp(ϑN)/2 and solve Eq. (4a) in closed\nform\nϑN(ε,λ)=ϑ0(ε)−ln/bracketleftbigcosh2(λ/ξN)/bracketrightbig, (8)\nwith normal side coherence length being defined as ξ−1\nN=/radicalig\n2ε∆\nǫThandϑ0=ϑN(ε,0). We now introduce u0=expϑ0and\nuS=exp(arcsinhγ+ϑB\nS), whereϑB\nS=ϑS(ε,±1/2),to rewrite\ntheboundarycondition[Eq.(5b)]as\nuS/γ−2=κγ(u0−uS), (9)\nwhere the interface parameter κ=σ2\nNDS/σ2\nSDNmeasures\nthe mismatch of conductivities and di ffusion coefficients at\nthe SN boundaries. By using solutions (7) and (8), together\nwithEq.(9),oneeliminatestheunknown u0andarrivestothe\nalgebraicequationfor z=uS/γ−2,\nF(z,κ)=/radicaligg\nγε∆\n8ǫTh, (10)\nwherethesingleparameter κ=κγ2scalingfunction F(z,κ)is\ngivenby\nF(z,κ)=/radicalbiggκ\nz+(z+2)κarctanh/radicalbiggz\nz+(z+2)κ.(11)\nKnowing the solution of Eq. (10), one finds the density of\nstatesat theSN interfacesas\nρ(ǫ,±1/2)=γ\n2Im[z(ǫ)]. (12)\nAt the same time, uS(ǫ) togetherwith Eqs.(6)–(8)providean\nexplicitinformationaboutthe localdensityof states ρ(ǫ,x) at\nanyposition xalongthewire.\nBy looking at Eq. (10), one sees that its right hand side\ngrows to infinity when ε→1, while its left hand side has\nan absolute maximum for certain value of z. This implies\nthatforallenergiesbelowsomethreshold εg,Eq.(10)hasthe\nonly real solution for zprovidingρ(ǫ)≡0 as it follows fromEq. (12). The condition ε=εgwhen Eq. (10) has a complex\nsolution for zfor the first time defines the proximity induced\nenergygapεg. Forenergiesabovethe gap ε>εg, the density\nof states is nonzerosince Im[ z]/nequal0. It turnsout that Eq. (10)\npossesses two qualitatively di fferent solutions depending on\nthevalueofthe interfaceparameter κ.\nThe limit of strong superconductor κ≪∆2/ǫ2\nTh. In this\ncase,κ≪1 andFfunction determined by Eq. (11) has\nthe following asymptotes: F(z,k)≈√z/4κforz≪1 and\nF(z,k)≈√κ/4zln(2/κ) forz≫1. It means that the only\nrelevantz, which determine the maximum of Fare those\nz∼κ. In this region, F(z,κ) may be approximated as\nF(z,κ)≈/radicalig\nκ\nz+2κarctanh/radicalig\nz\nz+2κ.As a result, the absolute max-\nimumFm=F(z=zm) occurs at point zm≈4.5κ, which cor-\nresponds to Fm≈0.5. Then, the gap determining condition\ngives\nFm=/radicaligg\nγgεg∆\n8ǫTh⇒εg=1−1\n8/parenleftigg∆\nǫTh/parenrightigg2\n,(13)\nwhere we have used the notation γg=γ(εg). Just above the\ngapε−εg≪εg,onecanexpand FinTaylorseriesaroundthe\nmaximum F≈Fm+b(z−zm)2, withb=(1/2)(d2F/dz2)z=zm\nto findz≈zm+i/radicalig\n1\nb/radicalbigg/radicalig\nγε∆\n8ǫTh−/radicalig\nγgεg∆\n8ǫTh.By using now def-\ninition (12), one finds for the density of states just above th e\nproximityinducedgapat SN–interface,\nρ(ǫ,±1/2)∝/parenleftbiggǫTh\n∆/parenrightbigg2/radicaligg\nǫ−ǫg\nǫg, ǫ−ǫg≪ǫg,(14)\nwherethenumericalcoe fficientoftheorderofunitywasomit-\nted. For the other limiting case, in the vicinity of the super -\nconductivegapε∼1,Eq.(10)issolvedby z≈−π2p2(1−4ip)\nwithp=/radicalig\nǫTh\n∆4√\n1−ε,whichgivesforthe densityofstates,\nρ(ǫ,±1/2)∝/parenleftbiggǫTh\n∆/parenrightbigg3/24/radicalbigg\n|ǫ−∆|\n∆,|ǫ−∆|/lessorsimilarδ∆.(15)\nThisasymptoticresult holdsabove ∆aswell. Observethat at\nǫ∼ǫg+δ∆/2,Eqs.(14)and(15)crossovertoeachother,while\natǫ∼∆+δ∆, Eq. (15) crossoversto the BCS like density of\nstates.\nThe limit of weak superconductor κ≫∆2/ǫ2\nTh. This\nlimiting case corresponds to the situation when κ≫1 and\nthe expression for Fgreatly simplifies F(z,κ)≈/radicalig\n1\nκ√z\nz+2.At\nzm=2, the function Fhasthe maximum Fm=/radicalbig1/8κg, were\nκg=κγ2\ng, so that the gap determining condition is di fferent\nfromEq.(13) andreads\n/radicaligg\n1\n8κg=/radicaligg\nγgεg∆\n8ǫTh⇒εg=1−1\n2/parenleftiggκ∆\nǫTh/parenrightigg2/3\n.(16)\nTo calculate the asymptotes for density of states at both gap\nedges, one follows the same steps as in the previouscase and4\nfinds\nρ(ǫ,±1/2)∝/parenleftbiggǫTh\nκ∆/parenrightbigg2/3/radicaligg\nǫ−ǫg\nǫg, ǫ−ǫg≪ǫg,(17a)\nρ(ǫ,±1/2)∝/radicalbigg\nǫTh\nκ∆4/radicalbigg\n|ǫ−∆|\n∆,|ǫ−∆|/lessorsimilarδ∆∗,(17b)\nwhereδ∆∗∼∆(κ∆/ǫTh)2/3. Equations (14), (15) and (17)\ncomplement our qualitative considerations presented at th e\nbeginningofthispaper.\nAt this point, let us discuss the obtained results and limits\nof their applicability. (i) We have studied the energy depen -\ndence of the local density of states for short ( L≪ξ) dif-\nfusive SNS junctions. Although ρ(ǫ,x) was analytically cal-\nculated at SN interfaces only it turns out that its energy de-\npendence is generic for any x∈[−L/2,L/2] and given by\nEqs. (14), (15) and(17). The exactnumericalprefactor,how -\never, isxdependent and should be determined numerically.\n(ii) Let us stress that the discussed feature in the ρ(ǫ,x) at\nǫ∼∆disappears at the level of global ∝angbracketleftρ(ǫ)∝angbracketright, which is inte-\ngrated over the volume, density of states. Indeed, the spati alintegration brings an additional factor of ξd\nS, wheredis an\neffective dimensionality of superconductor, which is due to\nthelongspatialtailsofAndreevstatespenetratingdeepin side\nthe superconductor [note that because of these tails, it is n ot\ncorrect to integrate ρ(ǫ,x) overx∈[−L/2,L/2] only]. For\nquasi–one–dimensional geometry discussed here, a factor o f\nξS∝/parenleftbig1\n1−ε/parenrightbig1/4gained after xintegration exactly compensates\nthe dip inρ(ǫ,x) atǫ∼∆[see Eqs. (15) and (17b)], leading\nto the finite value∝angbracketleftρ(ǫ=∆)∝angbracketright∼ǫTh/∆. Ford>1, the density\nofstatesat superconductivegapedge ǫ∼∆shoulddivergeas\na certain power–law ∝angbracketleftρ(ǫ)∝angbracketright∼(ǫ−∆)−pfor p>0. (iii) It fol-\nlowsfromthenumericalanalysis12thatpresenceofadditional\ntunnelbarriersatSNinterfacesalterssoftgapandleadsto the\nnonzerodensity of states at the gap edge ∆. (iv) Finite super-\nconductivephaseφimposedacrossthejunctionshiftsposition\nof the proximityinducedgap9ǫg(φ), such thatǫg(φ=π)=0,\nand also changes shape of the crossover function. However,\nzero in the LDOS at ǫ=∆and asymptotes at both gap edges\npersist.\nNumerous useful discussions with A. Kamenev and\nL. Glazman, which initiated and stimulated this work are\nkindly acknowledged. This research is supported by DOE\nGrantNo. 08ER46482.\n1S.Gu´ eron,H.Pothier,NormanO.Birge,D.EsteveandM.H.De -\nvoret, Phys.Rev. Lett. 77, 3025 (1996).\n2M.F.Goffman,R.Cron,A.LevyYeyati,P.Joyez, M.H.Devoret,\nD.Esteve, and C.Urbina, Phys.Rev. Lett. 85, 170 (2000).\n3E. Scheer, W. Belzig, Y. Naveh, M. H. Devoret, D. Esteve, and\nC.Urbina, Phys. Rev. Lett. 86, 284 (2001).\n4A.Anthore,H.Pothier,andD.Esteve,Phys.Rev.Lett. 90,127001\n(2003).\n5Yong-Joo Doh, Jorden A. van Dam, Aarnoud L. Roest, Erik\nP. A. M. Bakkers, Leo P. Kouwenhoven, Silvano De Franceschi,\nScience309, 272(2005).\n6See, e.g., Mesoscopic Superconductivity , edited by P. F. Bagwell\nspecial issue of Superlattices Microstruct. 25, 5–6 (1999).\n7A.A.Golubov, E.P.Houwman, J.G.Gijsbertsen,V.M.Krasnov ,\nJ. Flokstra, H. Rogalla, and M. Yu. Kupriyanov, Phys. Rev. B 51,\n1073 (1995).\n8W.Belzig,C.Bruder,andG.Sch¨ on,Phys.Rev.B 54,9443(1996).\n9F. Zhou, P. Charlat, B. Spivak, and B. Pannetier, J. Low Temp.\nPhys.110, 841 (1998).\n10Ya. V. Fominov and M. V. Feigel’man, Phys. Rev. B 63, 094518\n(2001).\n11T. T. Heikkil¨ a, J. S¨ arkk¨ a, and F. K. Wilhelm, Phys. Rev. B 66,\n184513 (2002).\n12J. C. Hammer, J. C. Cuevas, F. S. Bergeret, and W. Belzig, Phys .\nRev. B76, 064514 (2007).\n13K. M. Frahm, P. W. Brouwer, J. A. Melsen, andC.W.J. Beenakker, Phys.Rev. Lett. 76, 2981 (1996).\n14A. Altland, B. D. Simons and D. Taras–Semchuk, Adv. Phys. 49,\n321 (2000).\n15F.K.Wilhelmand A.A. Golubov, Phys. Rev. B 62, 5353 (2000).\n16M. G. Vavilov, P. W. Brouwer, V. Ambegaokar, and\nC.W.J. Beenakker, Phys.Rev. Lett. 86, 874 (2001).\n17J.S.Meyer andB. D.Simons, Phys.Rev. B 64, 134516 (2001).\n18M. Titov, N. A. Mortensen, H. Schomerus, and\nC.W.J. Beenakker, Phys.Rev. B 64, 134206 (2001).\n19A.LamacraftandB.D.Simons,Phys.Rev.Lett. 85,4783(2000).\n20P. M. Ostrovsky, M. A. Skvortsov, and M. V. Feigel’man, Phys.\nRev. Lett. 87, 027002 (2001).\n21A.Silva,Phys. Rev. B 72, 224505 (2005).\n22K.D. Usadel,Phys. Rev. Lett. 25, 507(1970).\n23C. W. J. Beenakker and H. van Houten, Single-Electron Tunnel-\ning and Mesoscopic Devices , edited by H. Koch and H. L¨ ubbig,\nSpringer, Berlin,(1992), pp. 175-179.\n24E. Akkermans, A. Auerbach, J. E. Avron, and B. Shapiro, Phys.\nRev. Lett. 66, 76(1991).\n25W.Belzig,F.K.Wilhelm,C.Bruder,G.Sch¨ on, andA.D.Zaiki n,\nSuperlattices Microstruct. 25, 1251 (1999).\n26K.K. Likharev, Rev. Mod. Phys. 51, 101 (1979).\n27M. Yu. Kuprianov and V. F. Lukichev, Sov. Phys. JETP 67, 1163\n(1988)." }, { "title": "0805.2148v2.Density_of_states_of_disordered_graphene.pdf", "content": "arXiv:0805.2148v2 [cond-mat.mes-hall] 18 Oct 2008Density of states of disordered graphene\nBen Yu-Kuang Hu1,2, E. H. Hwang1, and S. Das Sarma1\n1Condensed Matter Theory Center, Department of Physics,\nUniversity of Maryland, College Park, MD 20742-4111 and\n2Department of Physics, The University of Akron, Akron, OH 44 325-4001\n(Dated: October 18, 2008)\nWe calculate the average single particle density of states i n graphene with disorder due to impurity\npotentials. For unscreened short-ranged impurities, we us e the non-self-consistent and self-consistent\nBorn and T-matrix approximations to obtain the self-energy. Among th ese, only the self-consistent\nT-matrix approximation gives a non-zero density of states at the Dirac point. The density of states\nat the Dirac point is non-analytic in the impurity potential . For screened short-ranged and charged\nlong-range impurity potentials, the density of states near the Dirac point typically increases in the\npresence of impurities, compared to that of the pure system.\nPACS numbers: 81.05.Uw; 72.10.-d, 72.15.Lh, 72.20.Dp\nI. INTRODUCTION\nThe recent experimental realization of a single layer\nof carbon atoms arranged in a honey-comb lattice has\nprompted much excitement and activity in both the ex-\nperimental and theoretical physics communities1,2. Car-\nriers in graphene (both electrons and holes) have a lin-\near bare kinetic energy dispersion spectra around the\nKandK′points (the “Dirac points”) of the Brillouin\nzone. The ability of experimentalists to tune the chemi-\ncal potential to lie above or below the Dirac point energy\n(by application of voltages to gates in close proximity to\nthe graphene sheets) allows the carriers to be changed\nfrom electrons to holes in the same sample. This sets\ngraphene apart from other two dimensional (2D) car-\nrier systems that have a parabolic dispersion relation,\nand typically have only one set of carriers, i.e. either\nelectrons or holes. Another unique electronic property,\nthe absence of back-scattering, has led to the specula-\ntion that carrier mobilities of 2D graphene monolayers\n(certainly at room temperature, but also at low temper-\nature) could be made to be much higher than any other\nfield-effect type device, suggesting great potential both\nfor graphene to be the successor to Si-MOSFET (metal-\noxide-semicondcutor field effect transistor) devices and\nfor the discovery of new phenomena that normally ac-\ncompanies any significant increase in carrier mobility3–6.\nIt is therefore of considerable fundamental and techno-\nlogical interest to understand the electronic properties o f\ngraphene2.\nGraphene samples that are currently being fabricated\nare far from pure, based on the relatively low electronic\nmobilities compared to epitaxially grown modulation-\ndoped two-dimensional electron gases (2DEGs) such as\nGaAs-AlGaAs quantum wells. It is therefore important\nto understand the effects of disorder on the properties of\ngraphene. Disorder manifests itself in the finite lifetimes\nof electronic eigenstates of the pure system. In the pres-\nence of scattering from an impurity potential that is not\ndiagonal in the these eigenstates, the lifetime scattering\nrate of the eigenstate, γ, is non-zero and can be measuredexperimentally by fitting the line-shape of the low-field\nShubnikov-de Haas (SdH) oscillations.7,8The effect of\ndisorder scattering on the SdH line shape is equivalent\nto increasing the sample temperature and one can there-\nfore measure this Dingle temperature ( TD) and relate it\nto the single-particle lifetime through /planckover2pi1γ= 2πkBTD. To\navoid potential confusion, we mention that the lifetime\ndamping rate γdiscussed in this paper is notequal to\nthetransport scattering rate which governs the electrical\nconductivity. The lifetime damping rate is the measure\nof the rate at which particles scatters out of an eigen-\nstate, whereas the transport scattering rate is a mea-\nsure of the rate of current decay due to scattering out\nof an eigenstate. In normal 2DEGs, the transport scat-\ntering time can be much larger than the impurity in-\nduced lifetime /planckover2pi1/2γ, particularly in high mobility mod-\nulation doped 2D systems where the charged impurities\nare placed very far from the 2DEG8,9. Recently, the issue\nof transport scattering time versus impurity scattering\nlifetime in graphene has been discussed.10\nThe single-particle level broadening due to the impu-\nrity potential changes many of the physical properties of\nthe system including the electronic density of states.11,12\nThe electronic density of states is an important property\nwhich directly affects many experimentally measurable\nquantities such as the electrical conductivity, thermoele c-\ntric effects, and differential conductivity in tunneling ex-\nperiments between graphene and scanning-tunneling mi-\ncroscope tips or other electron gases. Changes in the\ndensity of states also modify the electron screening,13\nwhich is an important factor in the determination of var-\nious properties of graphene. It is therefore imperative\nto take into account the effects of disorder on the den-\nsity of states, particularly since disorder is quite strong\nin currently available graphene samples.\nIn the present work, we present calculations of the av-\nerage density of states of disordered graphene. This prob-\nlem has been investigated using various models and tech-\nniques, both analytical and numerical.14?–20We take\ninto account scattering effects from long-range and short\nrange impurity potentials. We consider both unscreened2\nand screened short-ranged and screened charged impuri-\nties, using the Born approximation. In addition, for un-\nscreened short-ranged (USR) impurities, we go beyond\nthe Born approximation and include self-consistent ef-\nfects.\nThere is another class of disorder in graphene called\n“off-diagonal” or “random gauge potential” disorder, in\nwhich the hopping matrix elements of the electrons in the\nunderlying honeycomb lattice are random. In this paper\nwe do not consider in this type of disorder, which can\nresult from height fluctuations (ripples) in the graphene\nsheet and lead to qualitatively different results from the\nones presented in this paper.14,19,22,23\nThe rest of the paper is organized as follows: In Sec. II,\nwe describe the approximation schemes that we use.\nSecs. III and IV deal with unscreened short range impuri-\nties and screened short-range/charged impurities, respec -\ntively. In Sec. V, we compare our results to those from\nother workers, and we conclude in Sec. VI.\nII. APPROXIMATIONS FOR THE\nSELF-ENERGY\nThe single-particle density of states for a translation-\nally invariant 2DEG is given by24\nD(E) =−g\nπ/summationdisplay\nλ/integraldisplaydk\n(2π)2Im[Gλ(k, E)] (1)where gis the degeneracy factor (for graphene g= 4 due\nto valley and spin degeneracies) λis the band index, G\nis the retarded Green’s function, and the k-integration\nis over a single valley, which we assume to be a circle of\nradius |k|=kc.Gexpressed in terms of the retarded\nself-energy Σ λ(k, ω) is\nGλ(k, E) = [E−Ek,λ−Σλ(k, E) +iη]−1.(2)\nwhere Ek,λis the bare band energy of the state |kλ∝angbracketright\nandηis an infinitesimally small positive number. (In\nthis paper, the Green’s functions and self-energies are all\nassumed to be retarded.) Eqs. (1) and (2) show that if\nEk,λand Σ λ(k, E) are known, the density of states can\nbe obtained in principle from\nD(E) =g\nπ/summationdisplay\nλ/integraldisplaydk\n(2π)2−Im[Σ λ(k, E)] +η\n(E−Ek,λ−Re[Σ λ(k, E)])2+ (Im[Σ λ(k, E)]−η)2. (3)\nFor pure graphene systems, Σ( k, E) = 0 (ex-\ncluding electron–electron and electron–phonon inter-\nactions, which are not considered here), and hence\nIm[Gλ(k, E)] =−πδ(E−Ekλ). Close to the Dirac points\n(which we choose to the the zero of energy), the disper-\nsion for graphene is (we use /planckover2pi1= 1 throughout this paper)\nEk,λ=λvFk, (4)\nwhere λ= +1 and −1 for the conduction and valence\nbands, respectively, k=|k|is the wavevector with re-\nspect to the Dirac point, and vFis the Fermi velocity\nof graphene. Performing the k-integration in Eq. (1) for\nthe pure graphene case gives\nD0(E) =g\n2π|E|\nv2\nFθ(|E| −Ec), (5)\nwhere Ec=vFkcis the band energy cut-off.\nThe average density of states for a disordered 2DEG\ncan be obtained by averaging the Green’s function overimpurity configurations. The averaging procedure gives\na non-zero Σ which, in general, cannot be evaluated ex-\nactly. Various approximation schemes for Σ have there-\nfore been developed, four of which are described below.\nBorn Approximation — In the Born approximation,\nthe self-energy is given by the Feynman diagram shown\nin Fig. 1(a)24, and the expression for the self-energy is\nΣB,λ(k, E) =ni/integraldisplaydk′\n(2π)2|U(k−k′)|2\n×/summationdisplay\nλ′G0,λ′(k′, E)Fλλ′(k,k′) (6)\nwhere niis the impurity density, U(q) is the Fourier\ntransform of the impurity potential, G0is the bare\nGreen’s function and Fλλ′(k,k′) is square of the the over-\nlap function between the part of the wavefunctions of\n|kλ∝angbracketrightand|k′λ′∝angbracketrightthat are periodic with the lattice (here\nλ, λ′are band indices). For graphene states near the3\n(a)\n(b)\n+ + +...X\nX X X\nFIG. 1: Feynman diagrams for the (a) Born and (b) T-matrix\napproximations for the self-energy. The “x”, dotted line an d\nline with arrow signify the impurity, impurity potential, a nd\nGreen’s function, respectively. The Green’s functions are ei-\nther bare or self-consistent.\nDirac point,\nFλλ′(k,k′) =1\n2(1 +λλ′cosθkk′), (7)\nwhere θkk′is the angle between kandk′.\nSelf-Consistent Born Approximation — The Feynman\ndiagram for this self-energy is the same as Fig. 1(a),\nexcept that the bare Green’s function is replaced by the\nfull one. Consequently, the expression for Σ SBis the same\nas in Eq. (6), except with G0replaced by G.\nT-matrix Approximation — TheT-matrix approxima-\ntion is equivalent to the summation of Feynman diagrams\nshown Fig. 1(b). The expression for Σ Tis the same as\nin Eq. (6), except with Ureplaced by T, theT-matrix\nfor an individual impurity.\nSelf-consistent T-matrix Approximation — The Feyn-\nman diagram for this approximation is the same as in the\nT-matrix approximation, except that the bare Green’s\nfunctions are replaced by full ones. The expression for\nΣSTis the same as in Eq. (6), except with Ureplaced by\nT, and G0replaced by G.\nWhen the potentials for the impurities are not all iden-\ntical (for example, in the case where there is a distribu-\ntion of distances of charged impurities from the graphene\nsheet), one averages |U(k−k′)|2or|T(k−k′)|2over the\nimpurities.\nIII. UNSCREENED SHORT-RANGED (USR)\nDISORDER\nIn the present context, short-ranged impurities are im-\npurities which result from localized structural defects in\nthe honeycomb lattice, which are roughly on the length\nscale of the lattice constant. In this case, it is accept-\nable to approximate U(q) =U0, a real constant, for in-\ntravalley scattering processes. In this paper, we ignore\nthe intervalley processes. (However, we note that if thematrix element joining intervalley states is constant, in-\nclusion of intervalley scattering in our calculations is no t\ndifficult.) This simplification allows us to obtain some\nanalytic expressions for the self-energies in the approxi-\nmation schemes mentioned above.\nA. Self-energy for USR disorder\nFor USR disorder, the four approximation schemes we\nuse give self-energies that are independent of λandk.\n1. Born approximation\nThe self-energy for graphene with USR scatterers in\nthe Born approximation, using U(q) =U0, Eqs. (4) and\n(7) in Eq. (6), is\nΣ(usr)\nB(E) = ˜γBH0(E+iη); (8a)\n˜γB=niU2\n0\n2v2\nF; (8b)\nH0(ζ) = 2v2\nF/summationdisplay\nλ′/integraldisplaydk′\n(2π)2G0,λ′(k′, ζ)Fλλ′(k,k′)\n=/integraldisplayEc\n0dE′\n2π/parenleftbiggE′\nζ−E′+E′\n��+E′/parenrightbigg\n=−ζ\n2πln/parenleftbigg\n1−E2\nc\nζ2/parenrightbigg\n. (8c)\nAs a function of complex ζ,H0(ζ) is real and positive\n(negative) along the real axis from Ecto∞(−Ecto−∞).\nFurthermore, it has a branch cut in on the real axis of ζin\nbetween −EcandEc, so that for −Ec< E(real) < E c,\nH0(E±iη) =−(2π)−1/parenleftbigg\nEln/vextendsingle/vextendsingle/vextendsingle/vextendsingleE2\nc\nE2−1/vextendsingle/vextendsingle/vextendsingle/vextendsingle±iπ|E|/parenrightbigg\n.(9)\nFig. 2 shows H0(E+iη).\nFor real Eand|E| ≪Ec,\nΣ(usr)\nB(E)≈ −˜γB\n2/bracketleftbigg2E\nπln/vextendsingle/vextendsingle/vextendsingle/vextendsingleE\nEc/vextendsingle/vextendsingle/vextendsingle/vextendsingle+i|E|/bracketrightbigg\n.(10)\nThe Born approximation damping rate for state |kλ∝angbracketrightis\nγB(k) =−2Im[Σ(usr)\nB(Ek,λ)] = ˜γBvFk.\n2. Self-consistent Born approximation\nThe self-energy Σ(usr)\nSBfor unscreened short-ranged\nscatterers in the self-consistent Born approximation is\ngiven by the self-consistent equation\nΣ(usr)\nSB(E) = ˜γBHSB(E+iη); (11a)\nHSB(ζ) =−1\n2π/bracketleftBig\nζ−Σ(usr)\nSB(ζ)/bracketrightBig\nln/bracketleftBigg\n1−E2\nc\n[ζ−Σ(usr)\nSB(ζ)]2/bracketrightBigg\n=H0/bracketleftBig\nζ−Σ(usr)\nSB(ζ)/bracketrightBig\n, (11b)4\n-2-1012\nE/E-1.0-0.50.00.51.0H /E\ncc0\nIm[H ]0Re[H ]0\nFIG. 2: Real and imaginary parts of H0(E+iη) =\nΣ(usr)\nB(E)/˜γB[see Eq. (9)], where Σ(usr)\nB(E) is the self-energy\nfor USR impurities in the Born approximation.This shows that if |Σ(usr)\nSB(E)| ≪ |E|, then Σ(usr)\nSB(E)≈\nΣ(usr)\nB(E) (except possibly around E=±Ec, which is\nusually not experimentally relevant).\n3.T-matrix approximation\nIn general the impurity averaged T-matrix for a poten-\ntialU(q) is\nTk0λ0,kλ(E) =ni∞/summationdisplay\nn=1/bracketleftBigg/parenleftBiggn/productdisplay\ni=1/summationdisplay\nλi=±1/integraldisplaydki\n(2π)2U(ki−1−ki)G0,λi(ki, E)Fλi−1,λi(ki−1,ki)/parenrightBigg\nU(ki−k)Fλn,λ(kn,k)/bracketrightBigg\n.\n(12)\nIn this approximation, the self energy is\nΣT,λ(k, E) =Tkλ,kλ(E). (13)\nIfU(q) =U0, a constant, the term in the square paren-\ntheses in Eq. (12) is Un+1\n0Hn\n0(E+iη)/(2v2\nF)n. The sum\nthen gives\nΣ(usr)\nT(E) =˜γBH0(E+iη)\n1−U0\n2v2\nFH0(E+iη). (14)\n4. Self-consistent T-matrix approximation\nAs in self-consistent Born approximation, the H0(E) in\nEq. (14) is replaced by the self-consistent HST(E), giving\nΣ(usr)\nST(E) =˜γBHST(E+iη)\n1−U0\n2v2\nFHST(E+iη); (15a)\nHST(ζ) =H0/parenleftBig\nζ−Σ(usr)\nST(ζ)/parenrightBig\n. (15b)As in the self-consistent Born approximation, this shows\nthat if |Σ(usr)\nST(E)| ≪ | E|, then Σ(usr)\nST(E)≈Σ(usr)\nT(E)\n(except possibly around E=±Ec).\nB. Density of states for k-independent Σ\nWhen the self-energy is k- andλ-independent, the den-\nsity of states for graphene can be calculated analytically\nfrom Eq. (3). The result is\nD(E) =gsgv\n2π2v2\nF/braceleftBigg\nΓ\n2ln/parenleftbigg(E2\nc+ Ω2+ Γ2)2−4E2\ncΩ2\n(Ω2+ Γ2)2/parenrightbigg\n+ Ω/bracketleftbigg\ntan−1/parenleftbiggEc−Ω\nΓ/parenrightbigg\n−tan−1/parenleftbiggEc+ Ω\nΓ/parenrightbigg\n+ 2 tan−1/parenleftbiggΩ\nΓ/parenrightbigg/bracketrightbigg/bracerightBigg\n,\n(16)5\nwhere Γ( E) =−Im[Σ( E)] and Ω( E) =E−Re [Σ(E)].\nC. Density of states at the Dirac point\nThere has been considerable interest in the minimum\nDC electrical conductivity of disordered graphene as the\nFermi energy moves through the Dirac point.25,26There\nis still no consensus on whether the minimum conductiv-\nity is a universal value or not. Since the electrical con-\nductivity is directly proportional to the density of states\nat the Fermi energy, it is important to be able to deter-\nmine the density of states at the Dirac point of disordered\ngraphene. Because the minimum conductivity is non-zero\nas the Fermi energy passes through the Dirac point, the\ndensity of states should be nonzero.\nEq. (3) shows that if Σ( E)→0 when E→0, then\nthe density of states at the Dirac point D(0) = 0. (Note\nthat we have not included the term that is first order in\nthe impurity potential in our self-energy. Since this first-\norder term merely rigidly shifts the band by an amount\nniU0, ignoring this term is equivalent to shifting the zero\nof the energy by −niU0, and hence the Dirac point is still\natE= 0.) A non-zero density of states at the Dirac point\ndepends on a non-zero Im[Σ( E= 0)]. Since H0(E→\n0) = 0, it is clear from Eqs. (8a) and (14) that the Born\nandT-matrix approximations give zero density of states\nat the Dirac point.\nFor the self-consistent Born approximation, Eq. (11b)\ncan be rewritten as\nΣ(usr)\nSB,λ(E) =E\n1 +2π\n˜γBln/parenleftbigg\n1−E2c\n(E−Σ(usr)\nSB,λ(E))2/parenrightbigg\n−1\n,\n(17)\nwhich shows that Σ(usr)\nSB,λ(E→0) = 0, and therefore the\nself-consistent Born approximation also gives D(0) = 0.\nIn the case of the self-consistent T-matrix approxima-\ntion, re-writing Eq (15a) as\nΣ(usr)\nST(E) +niU0=niU0\n1−H0(E−Σ(usr)\nST(E))U0/(2v2\nF),\n(18)\nsetting E= 0, using Eq. (8c) and taking the imaginary\nparts of both sides of this equation gives\nIm[Σ(0)] ≡Γ(0) = Im\nniU0\n1−iU0Γ(0)\n2v2\nFln/parenleftbigg\n1 +E2\nc\nΓ(0)2/parenrightbigg\n.\n(19)\nIn the weak scattering limit, when ˜ γB≪1, the imag-\ninary term in the denominator of Eq. (19) is much less\nthan one (we check for self-consistency later), and thisgives\nΓw(0)≈Im/bracketleftbigg\niniΓw(0)U2\n0\nv2\nFln/parenleftbiggEc\nΓw(0)/parenrightbigg/bracketrightbigg\n= 2˜γBΓw(0)ln/parenleftbiggEc\nΓw(0)/parenrightbigg\n,\n(20)\nwhich implies that\nΓw(0) = Ecexp/parenleftbigg\n−1\n2˜γB/parenrightbigg\n. (21)\nNote that the result is non-analytic in U0. Inserting\nEq. (21) into the imaginary part of the denominator of\nEq. (19) (which we had assumed to be much smaller\nthan 1 in magnitude) gives the self-consistent criterion\nexp(−1/2˜γB)≪niU0/Ecfor the validity of Eq. (21).\nSubstituting this into Eq. (16) gives an average density\nof states at the Dirac point for weak scattering of approx-\nimately\nρw(0) =gsgvEc\nπ2niU2\n0exp/parenleftbigg\n−1\n2˜γB/parenrightbigg\n. (22)\nSimilar results to Eq. (22) have been reported15,19,27in\nstudies of disordered systems of fermions with linear dis-\npersions using other methods.\nWe mention that our calculation of the graphene den-\nsity of states at the Dirac point should only be consid-\nered as demonstrative since electron-electron interactio n\neffects are crucial28at the Dirac point, and the undoped\ngraphene system is not a simple Fermi liquid at the Dirac\npoint.\nIV. SCREENED SHORT-RANGED AND\nCHARGED IMPURITIES\nFree carriers will move to screen a bare impurity po-\ntential Vei(q), resulting in a screened interaction U(q) =\nVei(q)/ε(q), where ε(q) is the static dielectric function.\nTheε(q) results in an q-dependent effective electron-\nimpurity potential U, even in the case of short-ranged\n(q-independent) bare impurity potentials. This makes\nthe calculations much more involved than in the USR\ncase. Therefore, in this paper, we limit our investiga-\ntion of q-dependent screened potentials to the level of\nthe Born approximation.\nFor the dielectric function ε(q), we use the random\nphase approximation (RPA) for dielectric function appro-\npriate for graphene, given by ε(q) = 1−Vc(q)Π0(q) where\nVc(q) = 2πe2/(κq) is the two-dimensional Fourier trans-\nform of the Coulomb potential ( κis the dielectric con-\nstant of the surrounding material), and Π 0(q) is the static\nirreducible RPA polarizability for graphene.29We use\nVei(q) =Vc(q) for charged impurities, and Vei(q) =U0, a\nconstant, for short-range point defect scatterers.\nWe first look at the density of states at the Fermi sur-\nface;i.e., at energy E=λkFvF. To obtain this, we6\ncalculate the single-particle lifetime damping rate γin\nthe Born approximation, which is given by\nγλ(k) =−2Im[Σ λ(k, Ekλ)]\n=ni\n2πk\nvF/integraldisplayπ\n0dθ∝angbracketleft|Vei(q)|2∝angbracketright\nε(q)2(1 + cos θ),(23)\nwhere q= 2ksin(θ/2). Then, assuming that\nIm[Σ λ(k, E)] is relatively constant for kclose to kF,\nwe substitute1\n2γλ(kF) for−Im[Σ λ(k, EF)] into Eq. (3),\nwhich gives Eq. (16) with Γ =1\n2γλ(kF).\nWe assume that the charged or neutral impurities are\ndistributed completely at random on the surface of the\ninsulating substrate on which the graphene layer lies, the\nareal density for the charged and neutral impurities is\nnicandniδ, respectively, and the density of carriers in\nthe graphene layer is n=k2\nF/π, where kFis the Fermi\nwavevector relative to the Dirac point. (This relationship\nbetween nandkFtakes into account the spin and valley\ndegeneracy gs= 2 and gv= 2.) We use the RPA screen-\ning function at T= 0,29to obtain the effective impurity\npotential. The key dimensionless parameter that quan-\ntifies the screening strength is rs=e2/(κvF), which is\ncorresponding to the interaction strength parameter of a\nnormal 2D system ( i.e.the ratio of potential energy to ki-\nnetic energy). The Born approximation lifetime damping\nrates for screened charged-impurities γcandδ-correlated\nneutral impurities γδatkFare\nγc(kF) =nicEF\n4nIc(2rs); (24a)\nγδ(kF) =2EF˜γB\nπIδ(2rs). (24b)\nIn these equations,\nIc(x) =x−πx2\n2+x3f(x); (25a)\nIδ(x) =π\n4+ 3x/parenleftBig\n1−πx\n2/parenrightBig\n+x(3x2−2)f(x),(25b)\nwhere\nf(x) =\n\n1√\n1−x2ln/bracketleftBig\n1+√\n1−x2\nx/bracketrightBig\nforx <1;\n1 for x= 1;\n1√\nx2−1cos−11\nxforx >1.(26)\nIn Fig. 3 we show the calculated damping rates scaled\nby|EF|/=kFvFas a function the interaction param-\neterrs. For rs≪1 we have γc/EF≈nicrs/2nand\nγδ/EF≈˜γB/2. For rs≫1 we have γc/EF≈πnic/16n\nandγδ/EF≈˜γB/(2πrs). Thus, for small (large) rsthe\ndamping rate due to the short-ranged impurity domi-\nnates over that due to the long-ranged charged impu-\nrity. On the other hand, since γc(kF)∝k−1\nF∝n−1\n2and\nγδ(kF)∝kF∝n1\n2[Eq. (24)], in the low (high) carrier\ndensity limit the lifetime damping of single particle state s\nat the Fermi surface is dominated by charged impurity0.01 0.1 1r10-410-310-210-1100γ /E\nsF\nFIG. 3: Calculated damping rates scaled by Fermi energy\nγ/E Fas a function of rs. Graphene on a SiO 2(air) sub-\nstrate has an rs≈0.7 (2). Solid lines indicate damping\nrates ( γc) due to charged impurities with an impurity den-\nsitynic= 1011cm−2for different electron densities n= 1, 10,\n50×1011cm−2(from top to bottom), respectively. Dashed\nline indicate the damping rate ( γδ) due to short-ranged im-\npurity with impurity density niδ= 1011cm−2and potential\nstrength U0=1 KeV ˚A2, which correspond to ˜ γB= 0.11.\nNoteγδ/EFis independent on the electron density.\n(short-ranged impurity) scattering. The crossover takes\nplace around a density\nncross=nic\nniδπv2\nF\n4U2\n0Ic(2rs)\nIδ(2rs). (27)\nUsing Γ = γλ(kF)/2 in Eq. (16) gives (assuming EF,\nΓ≪Ec)\nD(EF)≈D0(EF)/bracketleftbigg1\n2+1\nπtan−1/parenleftbigg|EF|\nΓ/parenrightbigg\n+Γ\n2π|EF|ln/parenleftbiggE2\nc\nE2\nF+ Γ2/parenrightbigg/bracketrightbigg\n.(28)\nFor Γ/|EF| ≪1, this gives\nD(EF)≈D0(EF)/braceleftbigg\n1 +1\nπΓ\n|EF|/bracketleftbigg\nln/parenleftbiggEc\n|EF|/parenrightbigg\n−1/bracketrightbigg/bracerightbigg\n,\n(29)\nand for |EF|/Γ≪1\nD(EF)≈D0(EF)/bracketleftbigg1\n2+|EF|\nπΓ/bracketrightbigg\n+gsgv\n2π2Γ\nv2\nFln/parenleftbiggEc\nΓ/parenrightbigg\n.\n(30)\nWe can apply Eq. (29) for short-ranged impurity scat-\ntering and for charged impurity scattering in high carrier\ndensity limits, and Eq. (30) for charged impurity scat-\ntering in low density limits. Taking the limit |EF| →0\nin Eq. (30), it appears that for the case of screened\nCharged impurities, we obtain a finite density of states7\n-6 -4 -2 0 2 4 6\nω/Ε00.10.2-ImΣ (k ,ω)/E(a)\nFFF+\n-6 -4 -2 0 2 4 6\nω/E-0.100.1Re Σ (k ,ω)/E(b)\nFFF+\nFIG. 4: (a) Imaginary and (b) real parts of self energy of a dis -\nordered graphene at k=kFfor screened Coulomb scattering\npotential (solid lines) and for screened neutral short-ran ged\nscattering potential (dashed lines).\nat the Dirac point with the Born approximation (since\nΓ∝γc(kF)∝k−1\nF). However, recall that in deriving\nEqs. (29) and (30), we have assumed that Σ( k, EF) is\nconstant with respect to k, which is not necessarily the\ncase at the Dirac point.\nIn general, the damping rate (or the imaginary part of\nthe self energy) is a function of energy and wave vector\nrather than a constant. From Eq. (6), we calculated the\nself-energy of disordered graphene. Fig. 4 we show the\nself-energy of a conduction band electron ( λ= +1) for\nboth screened Coulomb scattering potential and screened\nneutral short-ranged potential. For Coulomb scatter-\ners we use the impurity density nic= 1012cm−2, and\nfor neutral short-ranged scatterers the impurity density\nniδ= 1011cm−2and potential strength U0=1 KeV\n˚A2. The self-energies in the valence band ( λ=−1)\nare related to the self-energy in the conduction band\nby ReΣ +(k, ω) =−ReΣ−(k,−ω) and ImΣ +(k, ω) =\nImΣ−(k,−ω). As ω→0,−ImΣ λ(kF, ω)→ |ω|, and\nReΣ λ(kF, ω)→ωln|ω|for both scattering potentials.\nHowever, for large value of |ω|the asymptotic behaviors\nare different, that is, as |ω| → ∞ − ImΣ(kF, ω)∝ |ω|−1\nfor Coulomb scattering potential and −ImΣ(kF, ω)∝ |ω|\nfor short-ranged potential. Note that by using only\nthe (non-self-consistent) Born approximation in this sec-\ntion, we assume weak scattering and ignoring multiple-\nscattering events in calculated Σ( k, ω). Therefore, the-3 -2 -1 0 1 2 3\nω/E0123D(ω)/D (E )\nFF(a)0\n-3 -2 -1 0 1 2 3\nω/E01234D(ω)/D (E )\nFF0(b)\nFIG. 5: The density of states in the presence of impurity (a)\nfor screened Coulomb potential and (b) for screened short-\nranged potential. In (a) we use the charged impurity densiti es\nnic= 0, 1, 5 ×1012cm−2(from bottom to top), and in (b)\nthe short-ranged impurity density nid= 0, 0.5, 1 ×1012cm−2\n(from bottom to top) and potential strength U0=1 KeV ˚A2.\nresults are unreliable in the strong disorder limit ( i.e.,\nwhen Σ( k, ω) is modified significantly from its lowest-\norder form).\nIn Fig. 5 the density of states in the presence of im-\npurity is shown for different impurity densities. In Fig.\n5(a) we show the density of states for Coulomb impurity\npotential with impurity densities, nic= 0, 1012cm−2,\n5×1012cm−2(from bottom to top), and in Fig. 5(b) we\nshow the density of states for neutral short-ranged impu-\nrity potential with densities, nid= 0, 5×1011cm−2, 1012\ncm−2(from bottom to top) and potential strength U0=1\nKeV˚A2. The calculated density of states is normalized\nbyD0(EF) = (gsgv/2π)EF/γ2. The density of states is\nenhanced near Dirac point ( E= 0), but as |E| →0 it\ngoes zero as D(E)→ |E|ln|E|for both types of screened\nimpurity scattering. Based on the results of Section III,\nwe expect that this result is an artifact of the Born ap-\nproximation, and that the density of states should in fact\nbe non-zero. The enhancement of the density of states\ncan be explained as follows. In normal 2D system with\nfinite disorder the band edge Eedge,0of the pure conduc-\ntion (valence) band is shifted to Eedge,imp<0 (>0) and8\na band tail forms below (above) the band edge of a pure\nsystem. Thus, the density of states in the presence of\nimpurities is reduced for E > 0 (E < 0) because the\nstates have been shifted by the impurity potential into\nthe band tail. However, for graphene since the conduc-\ntion band and the valence band meet at the Dirac point\nthe band tail (or shift of band edge) cannot be formed,\nwhich gives rise to enhancement of density of states near\nDirac point.\nBefore concluding we point out that our perturbative\ncalculation of the graphene density of states assumes that\nthe system remains homogeneous in the presence of impu-\nrities. It is, however, believed3,30that graphene carriers\ndevelop strong density inhomogeneous (i.e. electron-hole\npuddle) at low enough carrier densities in the presence of\ncharged impurities due to the breakdown of linear scat-\ntering. In such an inhomogeneous low-density regime\nclose to the Dirac point, our homogeneous perturbative\ncalculation does not apply.\nV. COMPARISONS TO OTHER WORKS\nIn this section, we compare and contrast our model of\ndisorder and results to other works in the field.\nPereset al.16studied the effect of disorder in graphene\nby considering the effect of vacancies on the honeycomb\nlattice. For a finite density of vacancies, they found that\nthe density of states at the Dirac point is zero for the\n“full Born approximation” (equivalent to our T-matrix\napproximation) and non-zero for the “full self-consistent\nBorn approximation” (equivalent to our self-consistent T-\nmatrix approximation). Our results are consistent with\ntheirs, even though the regimes that are studied are dif-\nferent. Vacancies correspond to the limit where the im-\npurity potential U0→ ∞ , whereas this work is more\nconcerned with the weak impurity-scattering limit.\nPereira at al.17considered, among several different\nmodels of disorder, both vacancies and randomness in the\non-site energy of the honeycomb lattice. They numeri-\ncally calculated the density of states for these models of\ndisorder. For compensated vacancies (same density of va-\ncancies in both sub-lattices of the honeycomb structure)\nthey found that the density of states increased around\nthe Dirac point. (Ref. 17 also studied the case of un-\ncompensated vacancies, but that has no analogue in our\nmodel of disorder.) For the case of random on-site impu-\nrity potential, they find that “there is a marked increase\nin the DOS (density of states) at ED(the Dirac point)”\nand “the DOS becomes finite at EDwith increasing con-\ncentration” of impurities. Our self-consistent T-matrix\napproximation result is consistent with their numerical\nresults, although it should be mentioned again that it\nstrictly does not apply to the case of vacancies.Wuet al.20numerically investigated the average den-\nsity of states of graphene for the case of on-site disorder.\nThey found that for weak disorder, the density of states\nat the Dirac point increased with both increasing den-\nsity of impurities and strength of the disorder. (In their\nwork, they did not absorb the shift in the band due to\nthe impurity potential in their definition of the energy,\nso the Dirac point had a shift ED=xvwhere xis the\nconcentration of impurities and vis the on-site impu-\nrity energy.) Their numerical results for the minimum in\nthe average density of states (Fig. 3(b) in Ref. 20) seem\nto indicate a non-linear dependence of the value of the\nminimum as a function of the strength of the disorder\npotential, and is at least not inconsistent with Eq. (22).\nThe issue of the effect of screening of the impurity in-\nteractions on the density of states discussed in Section IV,\nto the best of our knowledge has not yet been treated in\nthe literature. Qualitatively, the effect of screened im-\npurities away from the Dirac point is the increase the\ndensity of states, and is consistent with numerical results\nfor random on-site disorder.17,20(Since our treatment of\nscreened impurities is at the level of the Born approx-\nimation, we do not obtain either a non-zero density of\nstates at the Dirac point nor resonances in the density of\nstates.17,18,20)\nVI. CONCLUSION\nWe have calculated the density of states for disordered\ngraphene. In the case of unscreened short-ranged impuri-\nties, we utilized the non-self-consistent and self-consis tent\nBorn and T-matrix approximations to calculate the self-\nenergy. Among these, only the self-consistent T-matrix\napproximation gave a non-zero density of states at the\nDirac point, and the density of states is a non-analytic\nfunction of the impurity potential. We investigated the\ndensity of states in the case of screened short-ranged and\ncharged impurity potentials at the level of the Born ap-\nproximation. We find that, unlike the case of parabolic\nband 2DEGs, in graphene near the band-edge ( i.e., the\nDirac point) the density of states is enhanced by impu-\nrities instead of being suppressed. At very low carrier\ndensities, however, graphene develops strong carrier den-\nsity inhomogeneity in the presence of charged impurities,\nan effect not captured by the homogeneous many-body\ntheory in our description.\nAcknowledgments\nThis work is supported by U.S. ONR, NSF-NRI, and\nSWAN.\n1See, for example, the special issues of Solid State Com-\nmunications 143, 1-125 (2007) and Eur. Phys. J. SpecialTopics 148, 1-181 (2007).9\n2A. H. Castro Neto, F. Guinea, N. M. R. Peres, K. S.\nNovoselov, A. K. Geim, arXiv:0709.1163.\n3E. H. Hwang, S. Adam, and S. Das Sarma, Phys. Rev.\nLett.98, 186806 (2007); S. Adam, E. H. Hwang, V. M.\nGalitski, and S. Das Sarma, Proc. Natl. Acad. Sci. U.S.A.\n104, 18392 (2007).\n4T. Ando, J. Phys. Soc. Jpn. 75, 074716 (2006); K. No-\nmura and A. H. MacDonald, Phys. Rev. Lett. 98, 076602\n(2007); V. V. Cheianov and V. I. Fal’ko, Phys. Rev. Lett.\n97, 226801 (2006). I. Aleiner and K. Efetov, Phys. Rev.\nLett.97, 236801 (2006).\n5K. I. Bolotin, K. J. Sikes, Z. Jiang, G. Fudenberg, J. Hone,\nP. Kim, and H. L. Stormer, Solid State Commun. 146, 351\n(2008); S. Adam and S. Das Sarma, Solid State Commun.\n146, 356 (2008).\n6S.V. Morozov, K.S. Novoselov, M.I. Katsnelson, F.\nSchedin, D.C. Elias, J.A. Jaszczak, A.K. Geim, Phys.\nRev. Lett. 100, 016602 (2008); J. H. Chen, C. Jang, S.\nXiao, M. Ishigami, M. S. Fuhrer, Nature Nanotechnology\n3, 206(2008); E. H. Hwang and S. Das Sarma, Phys. Rev.\nB77, 115449 (2008).\n7I. A. Luk’yanchuk and Y. Kopelevich, Phys. Rev. Lett. 97,\n256801 (2006).\n8J. P. Harrang, R. J. Higgins, R. K. Goodall, P. R. Jay,\nM. Laviron, and P. Delescluse, Phys. Rev. B 32, 8126\n(1985); M. Sakowicz, J. Lusakowski, K. Karpierz, M. Gryn-\nberg, and B. Majkusiak, Appl. Phys. Lett. 90, 172104\n(2007).\n9S. Das Sarma and F. Stern, Phys. Rev. B 32, 8442 (1985).\n10E. H. Hwang and S. Das Sarma, Phys. Rev. B 77, 195412\n(2008).\n11S. Das Sarma and B. Vinter, Phys. Rev. B 24, 549 (1981).\n12T. Ando, A. B. Fowler, and F. Stern, Rev. Mod. Phys. 54,\n437 (1982).\n13P. G. de Gennes, J. Phys. Radium 23, 630 (1962); T. Ando,\nJ. Phys. Soc. Jpn 51, 3215 (1982); S. Das Sarma, Phys.\nRev. Lett. 50, 211 (1983); S. Das Sarma and X. C. Xie,\nPhys. Rev. Lett. 61, 738 (1988); Ben Yu-Kuang Hu and S.\nDas Sarma, Phys. Rev. B 48, 14388 (1993).\n14Wei Min Hu, John D. Dow, and Charles W. Myles, Phys.\nRev. B 30, 1720 (1984).\n15Matthew P. A. Fisher and Eduardo Fradkin, Nucl. Phys.\nB251 [FS13], 457 (1985); Eduardo Fradkin, Phys. Rev. B33, 3257 (1986); ibid.33, 3263 (1986).\n16N. M. R. Peres, F. Guinea, and A. H. Castro Neto, Phys.\nRev. B 73, 125411 (2006).\n17Victor M. Periera, F. Guinea, J. M. B. Lopes dos Santos,\nN. M. R. Peres, and A. H. Castro Neto, Phys. Rev. Lett.\n96, 036801 (2006); Victor M. Periera, J. M. B. Lopes dos\nSantos, and A. H. Castro Neto, Phys. Rev. B 77, 115109\n(2008).\n18Y. V. Skrypnyk and V. M. Loktev, Phys. Rev. B 73,\n241402(R) (2006); Low Temp. Phys. 33, 762 (2007); Phys.\nRev. B 75, 245401 (2007).\n19Bal´ aza D´ ora, Klauss Ziegler, and Peter Thalmeier, Phys.\nRev. B 77, 115422 (2008).\n20Shangduan Wu, Lei Jing, Quanxiang Li, Q. W. Shi, Jie\nChen, Haibin Su, Xiaoping Wang, and Jinlong Yang, Phys.\nRev. B 77, 195411 (2008).\n21Aur´ elein Lherbier, X. Blase, Yann-Michel Niquet, Fran coi s\nTriozon, and Stephan Roche, Phys. Rev. Lett. 101, 036808\n(2008).\n22Y. Morita and Y. Hatsugai, Phys. Rev. Lett. 79, 3728\n(1997); Shinsei Ryu and Yasuhiro Hatsugai, Phys. Rev. B\n65, 033301 (2001).\n23F. Guinea, Baruch Horovitz, and P. Le Doussal, Phys. Rev.\nB77, 205421 (2008).\n24Seee.g., G. Rickaysen, Green’s Functions and Condensed\nMatter (Academic Press, New York, 1980); S. Doniach\nand E. H. Sondheimer, Green’s Functions for Solid State\nPhysicists (Imperial College, London, 1998); G. D. Mahan,\nMany Particle Physics , 3rd ed. (Plenum, New York, 2000).\n25Y.-W. Tan, Y. Zhang, K. Bolotin, Y. Zhao, S. Adam, E. H.\nHwang, S. Das Sarma, H. L. Stormer, and P. Kim, Phys.\nRev. Lett. 99, 246803 (2007).\n26K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang,\nY. Zhang, M. I. Katsnelson, I. V. Grigorieva, S. V.\nDubonos, and A. A. Firsov, Nature 438, 197 (2005).\n27P. M. Ostrovsky, I. V. Gornyi, and A. D. Mirlin, Phys.\nRev. B 74, 235443 (2006).\n28S. Das Sarma, E. H. Hwang, and W. K. Tse, Phys. Rev. B\n75, 121406(R) (2007).\n29E. H. Hwang and S. Das Sarma, Phys. Rev. B 75, 205418\n(2007).\n30E. Rossi and S. Das Sarma, arXiv:0803.0963." }, { "title": "0805.3870v1.Stability_of_the_density_wave_state_of_a_dipolar_condensate_in_a_pancake_trap.pdf", "content": "arXiv:0805.3870v1 [cond-mat.other] 26 May 2008Stability of the density-wave state of a dipolar condensate in a pancake trap\nO. Dutta∗, R. Kanamoto†, and P. Meystre\nB2 Institute, Department of Physics and College of Optical S ciences,\nThe University of Arizona, Tucson, AZ 85721, USA\n(Dated: October 31, 2018)\nWe study a dipolar boson-fermion mixture in a pancake geomet ry at absolute zero temperature,\ngeneralizing our previous work on the stability of polar con densates and the formation of a density-\nwave state in cylindrical traps. After examining the depend ence of the polar condensate stability on\nthe strength of the fermion-induced interaction, we determ ine the transition point from a ground-\nstate Gaussian to a hexagonal density-wave state. We use a va riational principle to analyze the\nstability properties of those density-wave state.\nPACS numbers: 03.75.-b, 42.50.-p, 33.80.-b, 03.65.-w\nI. INTRODUCTION\nThe recent realization of a Bose-Einstein conden-\nsate of chromium atoms [1, 2] opens up the study of\nquantum-degenerate gases that interact via the long\nrange, anisotropic magnetic dipole interaction. This is\nan anisotropic and long-range interaction that leads to\nthe appearance of a wealth of new properties past those\ncharacteristic of systems with isotropic interactions [3].\nInthe caseof52Crs-waveinteractionsarenormallymuch\nstronger than the dipolar part, but it has been shown ex-\nperimentallythatbyapplyingamagneticfieldthe s-wave\nscattering length can be lowered to values comparable to\nthe Bohr radius [4], so that the dipole interaction domi-\nnates the system. In the regime where the dipole-dipole\ninteraction is dominant the condensate is characterized\nby the existence of a metastable state whose properties\ndepends on the trapping geometry [5, 6, 7, 9, 10, 11],\nas was experimentally demonstrated in Ref. [8]. A roton\nfeature has also been predicted to exist in these systems\nfor appropriate parameters. There would considerable\ninterest indeed in accessing the associated roton instabil-\nity [12, 13, 14], as it is characterized by the spontaneous\ngeneration of a periodic density modulation in the con-\ndensate [15]. Unfortunately that state is unstable against\ncollapse, but we found recently that this difficulty can\nbe circumvented by the addition of a small fraction of\nnon-interacting fermions, resulting in a significant stabi-\nlization of the bosonic system in cylindrical traps [16].\nThe simultaneous trapping of bosonic and fermionic iso-\ntopes of chromium [17] also points to an interesting di-\nrections, with the possibility to observenovel ground and\nmetastable phases in quantum degenerate polar boson-\nfermion mixtures.\nThis paper extends our previous study of dipolar\nboson-fermionmixturesfromcylindricaltrapstopancake\ngeometries,showingthat dipolarbosonscanbe stabilized\n∗Corresponding author. Electronic address:\ndutta@physics.arizona.edu\n†Present Address: Ochanomizu University, Tokyo 112-8610, J apanconsiderably by increasing the boson-fermion s-wave\nscattering length, and more importantly the density-\nwave states of dipolar bosons can be stable for typical,\nzero-field boson-fermion interaction strengths. Section II\ndescribes the induced potential created by fermions on\nbosons, and Sec. III discusses the general properties of\nthe energy functional as a function of boson-fermion in-\nteractionusing a variationalGaussianansatz. Section IV\npresentsananalysisofthetransitionpointtothedensity-\nwave state for various combinations of trap aspect ratios\nand boson-boson contact interaction strengths. Section\nV introducesa density-waveansatzin the form ofa series\nof shifted Gaussians. The stability of these density-wave\nstates is determined by varying the width of each Gaus-\nsian to seek minima in the interaction energy for various\nboson-boson s-wave interaction. We find a condition for\nthe stability of these types of density modulated states.\nFinally, Section VI is a summary and outlook.\nII. FERMION-INDUCED INTERACTION IN\nPOLAR CONDENSATES\nWe consider a mixture of Nbdipolar bosons of mass\nmbandNfsingle-component fermions of mass mfcon-\nfined in a pancake-shaped trap characterized by a tight\nharmonic potential of frequency ωzalong the z-axis and\na softer harmonic potential of frequency ω⊥in the trans-\nversedirection. Apolarizingexternalelectricormagnetic\nfield, is taken to be along the z-axis . The dipole-dipole\ninteraction between two bosonic particles separated by a\ndistance ris then\nVdd(r) =gdd/parenleftbigg\n1−3z2\nr2/parenrightbigg1\nr3, (1)\nwheregddis the dipole-dipole interaction strength. In\nthe mean-field approximation, the energy functional for\nthe order parameter φ(r) of the dipolar condensate can\nbe expressed as\nE=/integraldisplay\nφ∗(r)H0φ(r)d3r+gNb\n2/integraldisplay\n|φ(r)|4d3r(2)\n+Nb\n2/integraldisplay/integraldisplay\n|φ(r)|2Vdd(r−r′)|φ(r′)|2d3rd3r′+Eind,2\nwhere\nH0=−/planckover2pi12\n2mb∇2+mbω2\nz\n2/bracketleftbig\nλ2/parenleftbig\nx2+y2/parenrightbig\n+z2/bracketrightbig\n(3)\nis the sum of the kinetic energy and the trapping po-\ntential and λ=ω⊥/ωz. The second term in the energy\nfunctional (2) denotes the contact interaction between\nbosons, characterized by the strength g= 4π/planckover2pi12abb/mb\nwithabbbeing the s-wavescatteringlength, andthe third\nterm describes the nonlocal dipole-dipole interaction be-\ntween bosons. Finally, the last term Eindaccounts for\nthe fermion-induced interaction Vind(k) between bosons,\ngiven in linear response theory [18] as follows.\nWe assume a contact boson-fermion interaction of\nstrength gbf= 2π/planckover2pi12abf/mr, where abfis the boson-\nfermions-wavescattering length and mr=mbmf/(mb+\nmf) is the reduced mass. The boson-fermion inter-\naction energy has the form gbf/integraltext\nnf(k)n(−k)d3k, with\nn(k) andnf(k) being the bosonic and fermion densi-\nties in the momentum space. The linear response of the\nfermions to a bosonic density n(k) can be expressed as\nnf(k) =Vind(k)n(k),sothatfortheeffectofthefermions\non the bosonic energy functional is\nEind=1\n2gbfNb\n(2π)3/integraldisplay\nVind(k)n(k)n(−k)d3k.(4)\nThe explicit form of the induced potential is\nVind(k) =gbfχf(k), (5)\nwhereχfis the density response function [18]. It\nis related to the dynamical structure factor S(k,ω),\nwhich is the probability of exciting particle-hole pairs\nwith momentum kout of the Fermi sea, by χf(k) =\n−2/integraltext∞\n0dω′[S(k,ω′)/ω′]. For a non-interacting single-\ncomponent Fermi system we have\nS(k,ω) =∞/summationdisplay\npkfδ(ω−ω0\npk), (6)\nwherep=|p|,kfis the Fermi momentum, the excitation\nenergy is ω0\npk=pkcosθ/mf+k2/(2mf), andθis the\nrelative angle between pandk. This expression assumes\nthat the fermions are locally free, so that there is a localFermi sphere in momentum space. In this local density\napproximation, kfis given by [19]\nkfℓz= 1.9N1/6\nfλ1/3/radicalbiggmf\nmb, (7)\nwhereℓz=/radicalbig\n/planckover2pi1/(mbωz) is the oscillator length in the z\ndirection.\nFollowing Ref. [16] the induced potential Vind(k) is\ngiven by\nVind(k)=\n\ngbfν/bracketleftBigg\n−1+∞/summationdisplay\nn=1/parenleftbiggk\n2kf/parenrightbigg2n1\n4n2−1/bracketrightBigg\n, k <2kf,\n−gbfν∞/summationdisplay\nn=1/parenleftbigg2kf\nk/parenrightbigg2n1\n4n2−1, k > 2kf,(8)\nwhereν=kfmf/(π/planckover2pi1)2is the three-dimensional\nfermionic density of states, and k2=k2\nx+k2\ny+k2\nzis\nthe square of the fermionic momentum. The induced po-\ntential is attractive for very low momenta and goes to\nzero with increasing momenta.\nThroughout this paper, we consider the parameters\ncorresponding to a52Cr-53Cr mixture, and fix the num-\nber of fermions to be Nf= 103unless otherwise stated.\nThezero-fieldbosonic s-waveinteractionfor52Cristaken\nto beabb= 103a0, andabf= 70a0for the boson-fermion\nscattering length [20].\nIII. STABILITY OF THE GAUSSIAN DIPOLAR\nCONDENSATE\nWe now proceed to determine the stability of the dipo-\nlar condensate, using a variational ansatz in the parame-\nter space of the fermion-induced interaction and dipolar\nstrength. The variational wave function is taken as the\nGaussian\nφ(r) =1/radicalbig\nπ3/2d2dzexp/bracketleftbigg\n−x2+y2\n2d2−z2\n2d2z/bracketrightbigg\n,(9)\nwithd,dzbeing the variational parameters and/integraltext\nd3r|φ(r)|2= 1. Substituting this Gaussian ansatz and\nits Fourier transform into Eqs. (2) and (4) yields the en-\nergyEgof the condensate as\nmbℓ2\nz\n/planckover2pi12Eg=1\n2/parenleftbigg1\n2+η2/parenrightbigg/parenleftbiggℓz\ndz/parenrightbigg2\n+1\n2/parenleftbigg1\n2+λ2\nη2/parenrightbigg/parenleftbiggdz\nℓz/parenrightbigg2\n+g3d/parenleftbiggℓz\ndz/parenrightbigg3/bracketleftBigg\n2\n3η2−F(η−1)+/parenleftbiggdz\nℓz/parenrightbigg3/braceleftbig\ngE>\n1+gind(E<\n2−E>\n2+E<\n3)/bracerightbig/bracketrightBigg\n, (10)\nwhereη=dz/d, and˜kf=kfℓz. The deriva-\ntion of algebraic forms of the interaction-energy termsE<\n1,E>\n1,E<\n2,E>\n2,E<\n3and of the function Fare straight-3\nforward but lengthy, and their derivations are relegated\nto an Appendix. In Eq. (10) we have also introduced the\neffective three-dimensional dipole-dipole interaction\ng3d=mbNbgdd√\n2π/planckover2pi12ℓz.\nThe coefficients gfind their origin in the\nmomentum-independent contact interaction, given by\nthe summation of the s-wave boson-boson scattering and\nthe constant terms in the induced interaction (8). They\nare given explicitly by\ng<=g−g2\nbfν\n4πgdd,\ng>=g\n4πgdd. (11)\nFinally\ngind=g2\nbfν\n48πgdd˜k2\nf\nis due to the nonlocal induced interaction. We re-\nmark that the effective contact interaction consists of\nthe boson-boson contact interaction as well as the\nmomentum-independent part of the dipole-dipole inter-\naction and of the induced interaction. From Eqs. (10)\nand (11) we have that for low momenta\ngs=2\n3+g<. (12)\nIt is known that in boson-fermion mixtures without\ndipolar interaction, phase separation occurs when gsbe-\ncomes negative [22, 23, 24]. In this paper, in contrast, we\nrestrict our considerations to the case gs>0 by chang-\ning the boson-boson contact interaction g, so that phase\nseparation does not take place.\nThe energyfunctional (10) wasminimized with respect\ntoηanddz/ℓzfor various parameter values, with our re-\nsults summarized in Fig. 1. Without the fermion induced\ninteraction, gind= 0, the ground-state energy of the sys-\ntem is not bounded from below, and the strong trap-\nping inzdirection creates a local minimum in the energy\nlandscape as a function of dzandη[Fig.1 (a)]. For fi-\nnite induced interactions gind>0, we find in contrast\nthat the energy landscape is characterized by two min-\nima, as shown in Fig. 1 (b), (c), and (d). In Fig. 1(b) the\nglobal minimum, which occurs for ( η,dz/ℓz)≃(3,2.3),\nis a Gaussian state with narrow width in the trans-\nversex-yplane. The additional local minimum close to\n(η,dz/ℓz)≃(0.2,1.4) is a metastable state.\nWith increasing boson-fermion interaction gind,\nthough, the ground-state energy corresponding to the\nglobal minimum at η >1 approaches that of the local\nminimum [Fig. 1 (c)], and these minima eventually reach\nequal energies at a critical value of gind. The new ground\nstate past that point is characterized by the parameters\nFIG. 1: Energy landscape Egobtained by the variational\nGaussian ansatz Eq. (9) as a function of η=dz/danddz/ℓz\nat a constant dipolar interaction strength g3d= 30. In these\nfigures the energy is cut off at Eg= 1.5 from above and\nEg=−0.5 from below for viewability. (a) Behavior of Egin\nthe absence of the fermion-induced interaction gind= 0 (b)\nEnergyEgforgind= 0.02. The plot is characterized by pres-\nence of two minima: (i) the true, narrow Gaussian ground\nstate located at ( η,dz/ℓz)≃(3,2.3); and (ii) a pancake-\nshaped metastable state located at ( η,dz/ℓz)≃(0.2,1.4). (c)\nEnergy landscape for gind= 0.06. The energies of the ground-\nand metastable states approach each other. (d) The broad\nGaussian metastable state at ( η,dz/ℓz)≃(0.2,1.4) eventually\nbecomes the true ground state for gind= 0.07. In the figures\nthe global (ground state) and local minima (metastable stat e)\nare indicated by a circle and cross, respectively.\nη <1 anddz∼ℓz[Fig. 1 (d)], that is, it is a wide\nGaussian in x-yplane.\nExperimentally one may either vary the boson-boson\ns-wavescattering length while keeping the boson-fermion\nscattering length constant, or vary gbfwith constant\ns-wave interaction g. In a boson-fermion mixture of\nchromium isotopes, typically gind∼10−3is small, which\ncorresponds to an energy landscape that resembles that\nof Fig. 1(b) and the broad Gaussian corresponds to a\nmetastable state. Such a metastable state has been\nachieved experimentally in experiments by the Stuttgart\ngroup [8]. A system that offers the potential to reach the\nregime of Fig. 1(d) is provided by a mixture of bosonic\n87Rb and fermionic40K with the scattering length of\nrubidium atoms tuned close to zero. In that mixture,\na zero-field scattering length abf≈250a0[21] gives\ngind∼0.2.4\nIV. TRANSITION TO A DENSITY WAVE\nThe excitation spectrum of condensates dominated by\na dipolar interaction is predicted to exhibit a roton min-\nimum [12, 13]. As a consequence, a bosonic density-\nmodulatedstateinthe x-yplanemayariseasalocalmin-\nimum, and it may actually have a lower energy than themetastable Gaussian state and for high enough bosonic\ndensities [15]. This section discusses the transition to the\nappearance of such a state as a function of number of\nbosonic particles for various combinations of trap ratio.\nWe proceed by introducing the new variational wave\nfunction that describes a density-wave structure with tri-\nangular symmetry,\nφt\ndw(x,y,z) =φ(x,y,z)/bracketleftBigg\na0+∞/summationdisplay\nn=1an/braceleftBigg\ncos/parenleftBigg\nn˜k0x\ndz/parenrightBigg\n+2cos/parenleftBigg\nn˜k0x\n2dz/parenrightBigg\ncos/parenleftBigg\nn√\n3˜k0y\n2dz/parenrightBigg/bracerightBigg/bracketrightBigg\n, (13)\nwhereφ(x,y,z) is defined in Eq. (9), nis an integer, and ˜k0=k0dz≪η=dz/d. Substituting this trial wave\nfunction (13) into Eq. (1), we find that the scaled excess energy o f the density-modulated state relative to the\nGaussian state\nǫ(˜k0,a1,a2,a3) =2md2\nz\n/planckover2pi12(Edw−Eg)\nis approximately given by\nǫ(˜k0,a1,a2,a3)\n3≈˜k2\n0\n4(a2\n1+4a2\n2+9a2\n3)+g3dnd\n4/bracketleftBig\n(a2\n1+2a0a1+a1a2+a2a3)2Veff(˜k0)\n+ (a2\n1+2a1a2)2Veff(√\n3˜k0)+1\n4(a2\n1+2a2\n2+4a0a2+2a1a3)2Veff(2˜k0)\n+ 2(a2\n1a2\n2+a2\n1a2\n3+a2\n2a2\n3)Veff(√\n7˜k0)+(2a0a3+a2\n3)2Veff(3˜k0)/bracketrightBig\n, (14)\nwherend≡η2ℓz/dz[25]. The normalization condition reads a2\n0+3(a2\n1+a2\n2+a2\n3)/2 = 1, and the effective transverse\npotential is\nVeff(˜k0)≈\n\n2\n3+g<−/radicalbiggπ\n2˜k0erfcx/parenleftBigg˜k0√\n2/parenrightBigg\n+gind/parenleftbiggℓz\ndz/parenrightbigg2\n˜k2\n0, ˜k0<2kfdz,\n2\n3+g\n4πgdd−/radicalbiggπ\n2˜k0erfcx/parenleftBigg˜k0√\n2/parenrightBigg\n−gind/parenleftbiggℓz\ndz/parenrightbigg2(2kfdz)4\n˜k2\n0,˜k0>2kfdz,(15)\nIn evaluating Eq. (14) we kept only the first four terms\nn= 0,...,3 of Eq. (13) as the energy converges at the\ntransition point. The magnitude of the error in that ap-\nproximate expression is estimated to be of the order of\ne= exp(−˜k2\n0η2/4). In the subsequent calculations we\nconsider values of ˜k0such that this error is less than or\non the order of 10−6.\nWe numerically determine the variational parameters\nthatminimizethe excessenergy ǫ(˜k0,a1,a2,a3)asafunc-\ntion ofgE>\n1+gind(E<\n2−E>\n2+E<\n3)/bracerightbig\n+f(l,ξ,dz)/bracketrightBigg\n,\n(17)\nwhere we have assumed that ξ≪l, i.e., that neigh-\nboring Gaussian peaks in Eq. (16) have little overlap.\nHereE<\n1,E>\n1,E<\n2,E>\n2,E<\n3andFhave the same form as\nin the Appendix, but σis now a function of ( ξ,dz,θ),\nσ=/radicalBig\nd2zcos2θ+ξ2sin2θ. The function f(l,ξ,dz) is\nresponsible for the particular geometry of the density-\nmodulated state. Other terms, on the other hand, just\narise from the energy of each individual Gaussian and\nwe can thus apply the results of Sec. III to the present\nanalysis.\nAssuming a triangular crystal and including only the\ninteraction between nearest neighbors, we have\nf(l,ξ,dz) = 6/integraldisplay∞\n0Veff(˜k)exp/parenleftBigg\n−˜k2ξ2\n2d2z/parenrightBigg\nJ0/parenleftBigg˜kl\ndz/parenrightBigg\n˜kd˜k,\nwith˜k=kdz, the effective two-dimensional potential\nVeff(˜k) is defined in Eq. (15), and J0is the zeroth-order\nBessel function of the first kind.\nWe now discuss the minimum of the energy functionalEint(ξ,dz) as a function of the effective strength of the\ncontact interaction gs, which is varied by changing g<,\nsee Eq. (12). Without boson-fermion interaction, the in-\nteractionenergyis amonotonically increasingfunction of\nξwithEint(ξ→0,dz)→ −∞. As a result each gaussian\nof the density wave Eq. (16) collapses. A typical exam-\nple of the dependence of the interaction energy on ξis\nshown as the dotted curve in Fig. 3(a). With non-zero\nboson-fermion interaction energy, in contrast, the inter-\naction energy (17) exhibits a minimum at a finite ξas\nillustrated by the solid curve in Fig. 3(a) f or dz= 1.8ℓz\nandg<= 0.17. The existence of that minimum implies\nthe stability of each Gaussian, that is, the stability of the\nwave function (16).\nWe can determine the dependence of the gaussian\nwidthsξon thes-wave boson-boson scattering length by\nchanging g<. To do this we first minimize Eq. (17) as a\nfunction of g<, treating dzandξas variational parame-\nters. Figure3(b)showstheresultingvalue ξ/ℓzasafunc-\ntion ofg0.14 the effective po-\ntential is repulsive both at low and high momenta, and\nattractive in-between as illustrated by the solid curve in\nFig. 4. Subsequently the energy is minimized for a finite\nvalue ofξ.\nTo find the range of parameter space characterized by\nthe appearance of density-wave states, we need to con-\nsider, in addition to stability arguments, the transition\npoint discussed in Section III. Table 1 summarizes the\nrange of s-wave scattering lengths and critical numbers\nof52Cr atoms necessary to be inside stable density-wave\nregime.\nTABLE 1\nλℓz(µ)abb(a0)Nb(×104)\n0.15 16-21 ≥0.65\n0.25 0.2 16-19 ≥2.3\n0.25 15-18 ≥5.0\nTABLE I: Tabulation of the scattering length of boson-boson\ncontact interaction abband critical number of bosons Nbfor\ndifferentvalues of aspect ratio λand oscillator length ℓzinside\nthe density-wave regime with Nf= 103andgbf= 70a0for a\nmixture of chromium isotopes.\nVI. CONCLUSION\nInsummary, wehaveanalyzedthe stabilityofadipolar\nbosonic condensate mixed with non-interacting fermions\nin a pancake trap at T= 0. We found that the fermions\nhelp stabilize the condensate for a significant range of\nboson-boson and boson-fermion interaction strengths.\nWe then investigated the transition of the system from\na Gaussian-like to the density-wave ground state as a\nfunction of number of bosons, strength of the contact in-\nteraction, and trap aspect ratio. Our central result is the\nuse of a variational ansatz to show that while in a purely\nbosonic system the density-wave state is always unsta-\nble it can be stabilized by the admixture of even a small\nboson-fermion interaction.\nIn particular, this study leads us to the conclusion that\na pancake-shaped87Rb-40K mixture, which has a large\nboson-fermion s-wave scattering length, should be abso-\nlutely stable in the dipole dominated regime. By tuning\nthes-wave scattering length it is possible to reach a sit-\nuation characterized by the appearance of a roton insta-\nbility in the excitation spectrum, leading to the existence\nof a stable density-wave state. However, due to the small7\ndipole moment of rubidium atoms the transition to this\nstable density-wave regime needs a substantial number\nof atoms, of the order of 106.\nFuture work will discuss the effect of the dipolar na-\nture of the fermionic isotopes on the condensate – with\nchromium atoms in mind –, the existence and stability of\ndensity waves, as well as possible extensions to rotating\nsystems.\nAcknowledgments\nWe thank Prof. Tilman Pfau for several interesting\ndiscussions and his deep insight on dipolar condensates.This work is supported in part by the US Office of Naval\nResearch, by the National Science Foundation, and by\nthe US Army Research Office.\nAPPENDIX A: DERIVATION OF THE ENERGY\nFUNCTIONAL OF EQ. (10)\nIn this appendix we derive the energy functional\nEq. (10) by using the Gaussian ansatz of Eq. (9). First\nwe consider the dipolar interaction energy,\nEdd=gddNb\n3(2π)2/integraldisplay\ndk/parenleftbigg\n3k2\nz\nk2−1/parenrightbigg\nexp/bracketleftbigg\n−1\n2{k2\nzd2\nz+(k2\nx+k2\ny)d2}/bracketrightbigg\n, (A1)\nwherek2=k2\nx+k2\ny+k2\nz. Evaluating this integral gives\nthe form of dipolar energy\nEdd=gddNb\n3(2π)2d3z/bracketleftbigg2\n3η2−F(η−1)/bracketrightbigg\n,(A2)\nwithη=dz/dand\nF(y) =tan−1(/radicalbig\ny2−1)\n(y2−1)3/2−1\ny2(y2−1).\nNext we calculate the energy due to s-wave boson-boson\ninteraction and fermion-induced interaction. To achievethis goal we break the integral of the interaction energies\nexpressed in spherical coordinates into two parts, Es≡\nE<\ns+E>\nsandEdd≡E<\ndd+E>\ndd, according to\n/integraldisplay∞\n0dk=/integraldisplay2kf\n0dk+/integraldisplay∞\n2kfdk. (A3)\nFork <2kfand in spherical coordinate, the interac-\ntion energy including the s-wave and induced interaction\nis given by\nE<\ns+E<\nind=gddNb\n2π/integraldisplay2kf\n0/integraldisplayπ\n0/bracketleftBigg\ng<+g2\nbfν\n4πgdd∞/summationdisplay\nn=1/parenleftbiggk\n2kf/parenrightbigg2n1\n4n2−1/bracketrightBigg\nexp/bracketleftbigg\n−k2\n2(d2\nzcos2θ+d2sin2θ)/bracketrightbigg\nk2dksinθdθ\nwhereg2kfto the interac-\ntion energy that contains both the s-wave boson-bosoninteraction and the nonlocal induced interaction is\nE>\ns+E>\nind=gddNn\n2π/integraldisplay∞\n2kf/integraldisplayπ\n0/bracketleftBigg\ng>−g2\nbfν\n4πgdd∞/summationdisplay\nn=1/parenleftbigg2kf\nk/parenrightbigg2n1\n4n2−1/bracketrightBigg\nexp/bracketleftbigg\n−k2\n2(d2\nzcos2θ+d2sin2θ)/bracketrightbigg\nk2dksinθdθ,\nwhereg>defined in Eq. (11). In this equation the first\nterm in the bracket stems from the s-wave boson-boson\ninteraction and the last term from attractive nonlocal\npart of the induced interaction. Again after evaluating\nthe integral over kwe get\nE>\ns+E>\ndd=g>E>\n1−gindE>\n2 (A6)where\nE>\n1(σ) =/integraldisplayπ/2\n0erfc(√\n2kfσ)\nσ3sinθdθ,\nand\nE>\n2(σ) =3√\n2π∞/summationdisplay\nn=12kf2(n+1)\n4n2−1/integraldisplayπ/2\n0/bracketleftBigg\n−(−1)n/radicalbig\nπ/2\n(2n−1)!!σ2n−3erfc(√\n2kfσ)\n+n−2/summationdisplay\nm=0(−1)m2m+1(2kf)2m\n2kf2n−2m−3(2n−3)(2n−5)....(2n−2m−3)exp(−2k2\nfσ2)σ2m/bracketrightBigg\nsinθdθ, (A7)\nThe total interaction energy is the sum of the contri-\nbutions in Eq. (A2), (A4), (A6) and is given in Eq. (10).\nIn this present paper we took n= 1,...,3 in the series inEqs. (A5), (A7) as the energy Egin Eq. (10) converges\nwith this inclusion.\n[1] A. Griesmaier, J. Werner, S. Hensler, J. Stuhler, and\nT. Pfau, Phys. Rev. Lett. 94, 160401 (2005).\n[2] Q. Beaufils, R. Chicireanu, T. Zanon, B. Laburthe Tolra,\nE. Marechal, L. Vemac, J.C. Keller, and O. Gorceix,\narXiv/cond-mat: 0712.3521.\n[3] K. G´ oral, L. Santos and M. Lewenstein, Phys. Rev. Lett.\n88, 170406 (2002) .\n[4] T. Lahaye, T. Koch, B. Fr¨ ohlich, M. Fattori, J. Metz,\nA. Griesmaier, S. Giovanazzi, and T. Pfau, Nature 448,\n672 (2007).\n[5] L. Santos, G. V. Shlyapnikov, P. Zoller, and M. Lewen-\nstein, Phys. Rev. Lett. 85, 1791 (2000).\n[6] S. Yi and L. You, Phys. Rev. A 63, 053607 (2001).\n[7] C. Eberlein, S. Giovanazzi, D. H. J. O’Dell, Phys. Rev.\nA71, 033618 (2005).\n[8] T. Koch, T. Lahaye, J. Metz, B. Fr¨ ohlich, A. Griesmaier,\nT. Pfau, arXiv/cond-mat:0710.3643.\n[9] K. G´ oral, K. Rza¸˙ zewski, and T. Pfau, Phys. Rev. A 61,\n051601 (2000).\n[10] S. Ronen, D. C. E. Bortolotti, and J. L. Bohn, Phys. Rev.\nLett.98, 030406 (2007); S. Ronen, and J. L. Bohn, Phys.\nRev. A 76, 043607 (2007).[11] O. Dutta, and P. Meystre, Phys. Rev. A 75, 053604\n(2007).\n[12] L. Santos, G. V. Shlyapnikov and M. Lewenstein, Phys.\nRev. Lett. 90, 250403 (2003).\n[13] Uwe R. Fischer, Phys. Rev. A. 73, 031602 (2006).\n[14] S. Komineas, N.R. Cooper, Phys. Rev. A. 75, 023623\n(2007).\n[15] Y. PomeauandS. Rica, Phys.Rev.Lett. 72, 2426(1994);\nC. Josserand, Y. Pomeau, and S. Rica, Phys. Rev. Lett.\n98, 195301 (2007) and reference therein.\n[16] O. Dutta, R. Kanamoto, and P. Meystre, Phys. Rev.\nLett.99, 110404 (2007).\n[17] R. Chicireanu, A. Pouderous, R. Barb´ e, B. Laburthe-\nTolra, E. Mar´ echal, L. Vernac, J.-C. Keller, and O. Gor-\nceix, Phys. Rev. A 73, 053406 (2006).\n[18] P. Nozi` eres and D. Pines, The Theory of Quantum Liq-\nuids(Perseus, Cambridge, 1999).\n[19] C. J. Pethick, andH.Smith, Bose-Einstein Condensation\nin Dilute Gases , Cambridge University Press, (2002).\n[20] Private communication with Prof. Tilman Pfau.\n[21] J. Goldwin, S. Inouye, M.L. Olsen, B. Newman, B.D. De-\nPaola, and D.S. Jin, Phys. Rev. A 70, 021601 (2004) and9\nreference therein.\n[22] J. M. J. van Leeuwen and E. G. D. Cohen, Phys. Rev.\n176, 385 (1968).\n[23] H.P.B¨ uchler andG.Blatter, Phys.Rev.Lett. 91, 130404\n(2003).\n[24] Z. Akdeniz, P. Vignolo, A. Minguzzi, M.P. Tosi, J. Phys.\nB35, L105 (2002); P. Capuzzi, A. Minguzzi, M.P. Tosi,\nPhys. Rev. A 68, 033605 (2003).[25] For a fixed number of dipolar bosons, the quantity nd\ncan be changed by tuning the aspect ratio λ.\n[26] The qualitative behavior of the total-energy landscap e,\nhence the stability of the density-wave state, remains un-\nchanged by the inclusion of the kinetic energy and trap-\nping potential." }, { "title": "0806.1988v1.Statistical_Characterization_of_a_1D_Random_Potential_Problem___with_applications_in_score_statistics_of_MS_based_peptide_sequencing.pdf", "content": "arXiv:0806.1988v1 [q-bio.QM] 12 Jun 2008Statistical Characterization of a 1D Random\nPotential Problem – with applications in score\nstatistics of MS-based peptide sequencing\nGelio Alves and Yi-Kuo Yu1\nNational Center for Biotechnology Information, National Libr ary of Medicine,\nNational Institutes of Health, Bethesda, MD 20894\nAbstract\nWe provide a complete thermodynamic solution of a 1D hopping model in the pres-\nence of a random potential by obtaining the density of states . Since the partition\nfunction is related to the density of states by a Laplace tran sform, the density\nof states determines completely the thermodynamic behavio r of the system. We\nhave also shown that the transfer matrix technique, or the so -called dynamic pro-\ngramming, used to obtain the density of states in the 1D hoppi ng model may be\ngeneralized to tackle a long-standing problem in statistic al significance assessment\nfor one of the most important proteomic tasks – peptide sequencing using tandem\nmass spectrometry data.\nKey words: Statistical Significance, Dynamic Programming, Mass Spect rometry,\nDirected Paths in Random Media, Peptide Identification\n1 Introduction\nImportant in both fundamental science and numerous applications , optimiza-\ntion problems of various degrees of complexity are challenging (see [1 ] for\nan excellent introduction). Optimization conditioned by constraints that may\nvary fromevent to event is ofespecial theoretical andpractical importance. As\na first example, when dealing with a system under a random potential, each\nrealization of the random potential demands a separate optimizatio n result-\ning in a different ground state. The thermodynamic behavior of such a system\nin a quenched random potential crucially depends on the random pot ential\nrealized. A similar but practical problem may arise in routing passenge rs at\n1To whom correspondence should be addressed. E-mail address : yyu@ncbi.nlm.nih.gov\nPreprint submitted to Elsevier Science June 6th 2008various cities to reach their destinations. In the latter case, the o ptimal rout-\ning depends on the number of passengers at various locations, the costs from\none location to the others, which likely to vary from time to time. This t ype of\nconditional optimization also occurs in modern proteomics problem, t hat is,\nin the mass spectrometry (MS) based peptide sequencing. In this c ase, each\ntandem MS (MS2) spectrum constitute a different condition for optimization\nwhich aims to find a database peptide or a de novopeptide to best explain\nthe given MS2spectrum.\nWhen the cost function of an optimization problem can be expressed as a sum\nof independent local contributions, the problem usually can be solve d using\nthe transfer matrix method that is commonly employed in statistical physics.\nA well-studied example of this sort in statistical physics is the directe d poly-\nmer/path in a random medium (DPRM) [2,3,4]. Even when a small non-loca l\nenergetics is involved, the transfer matrix approach still proves u seful [5]. As\nan example, the close relationship between the DPRM problem and MS- based\npeptide sequencing, where a small nonlocal energetics is necessar y to enhance\nthe peptide identifications, was sketched in an earlier publication [5] a nd the\ncost value distribution from many possible solutions other than the o ptimal\none is explored. Indeed, obtaining the cost value distribution from allpossible\nsolutions in many cases is harder than finding the optimal solution alon e. In\nthis paper, we will provide the solution to a generic problem that enab les a full\ncharacterization of the peptide sequencing score statistics, inst ead of just the\noptimal peptide. The 1D problem considered is essentially a hopping mo del in\nthe presence of a random potential. The solution to this problem may also be\nuseful in other applications such as in routing of passengers and ev en internet\ntraffic.\nIn what follows, we will first introduce the generic 1D hopping model in a\nrandom potential, followed by its transfer matrix (or dynamic progr amming)\nsolution. Wethendiscuss theutilityofthissolutioninthecontext ofM S-based\npeptide sequencing, and demonstrate with real example from mass spectrum\nin real MS-based proteomics experiments. In the discussion sectio n, we will\nsketch the utility of the transfer matrix solution in other context a nd then\nconclude with a few relevant remarks.\n2 1D hopping in random potential\nAlongthe x-axis,letusconsideraparticlethatcanhopwithasetofprescribed\ndistances {mi}K\ni=1towards the positive ˆ xdirection. That is, if the particle is\ncurrently at location x0, it can move to location x0+m1,x0+m2, ...x0+mK\nin the next time step. At each hopping step, the particle will accumula te an\nenergy−s(x) from location xthat it just visited. The score s(x) (negative\n2of the on-site potential energy) is assumed positive and may only ex ist at a\nlimited number of locations. For locations that s(x) do not exist, we simply\nsets(x) = 0 there. The energy of a path starting from the origin specified\nby the sequential hopping events p≡ {mh1,mh2,...,m hL}would have visited\nlocations {x1,x2,...,x L}withxi≡/summationtexti\nj=1mhjand has energy\nEp(x=xL)≡ −L−1/summationdisplay\ni=1s(xi)≡ −Sp(x).\nIn general, there can be more than one path terminated at the sam e point.\nTreating each path as a state with energy given by Ep, one ends up having\nthe following recursion relation for the partition function Z(x)≡/summationtext\npe−βEp(x)\nZ(x) =K/summationdisplay\ni=1eβs(x−mi)Z(x−mi), (1)\nwhereβ= 1/Tplays the role of inverse temperature (with kB= 1 chosen).\nIf one were only interested in the best score terminated at point x, it will be\ngiven by the zero temperature limit β→ ∞and the recursion relation may be\nobtained by taking the logarithm on both sides of (1) and divided by βthen\ntakingβ→ ∞limit to reach\nSbest(x) = max\n1≤i≤K{s(x−mi)+Sbest(x−mi)}, (2)\nwhereSbest(x) records the best path score among all paths reaching position\nx. This update method, also termed dynamic programming, records t he lowest\nenergy and lowest energy path reaching a given point x. The lowest energy\namong all possible at position xis simply −Sbest(x) and the associated path\ncan be obtained by tracing backwards the incoming steps. It is inter esting to\nobserve thatonecanalsoobtaintheworst scoreateachpositionv iadynamical\nprogramming\nSworst(x) = min\n1≤i≤K{s(x−mi)+Sworst(x−mi)}. (3)\nThe full thermodynamic characterization demands more informatio n than the\nground state energy. In principle, one may obtain the full partition function\nusing eq. (1) evaluated at various temperatures. This procedure , however,\nhinders analytical property such as determination of the average energy\n/an}bracketle{tE/an}bracketri}ht ≡ −∂lnZ\n∂β.\n3A better starting point may be achieved if one can obtain the density of states\nD(E). In this case, we have\nZ≡/integraldisplay\ndEe−βED(E)\n/an}bracketle{tE/an}bracketri}ht=/integraltextdEe−βEED(E)/integraltextdEe−βED(E).\nNote that if the ground energy Egrdof the system is bounded from below,\nthe partition function is simply a Laplace transform of a modified dens ity of\nstates given by\nZ=e−βEgrd∞/integraldisplay\n0dEe−βE˜D(E)\nwhere˜D(E)≡D(E−Egrd) and\n/an}bracketle{tE/an}bracketri}ht=Egrd+/integraltext∞\n0dEe−βEE˜D(E)\n/integraltext∞\n0dEe−βE˜D(E)\nThis implies that the density of states D(E) together with the ground state\nenergyEgrddetermine all the thermodynamic behavior of the system. In the\nnext section, we will explain how to obtain the density of states using the\ndynamical programming technique as well as how to extend this appr oach\nto more complicated situations that will be useful in characterizing t he score\nstatistics in MS-based peptide sequencing.\n3 Obtaining the Density of States\nThe density of states is related to the energy histogram in a simple wa y. The\nnumber of states between energies EandE+η(withη≪1) is given by\nD(E)η. If we happen to use ηas the energy bin size for energy histogram,\nthe count C(E) in the bin with energy Eis simply D(E)ηand the density\nof states D(E) =C(E)/η. For simplicity, we will assume that the all the\non-site energies −s(x) are integral multiple of η. This implies that each path\nenergy/score is also an integral multiple of η. In the following subsections, we\nwill use score density of states instead of energy density of state s.\n43.1 The Simplest Case and its Application\nWe denote by C(x,N) the number of paths reaching position xwith score\nNη. With this notation, we can easily write down the recursion relation fo r\nC(x,N) as follows\nC(x,N) =K/summationdisplay\ni=1C(x−mi,N−s(x−mi)\nη). (4)\nThis recursion relation allows us to compute the density of states in t he same\nmanner as computing the partition function (1) except that we nee d to have\nan additional dimension for score at each position x. As an even simpler ap-\nplication of this recursion relation, suppose that one is only interest ed in the\nnumber of paths reaching position x, one may sum over the energy part on\nboth side of (4) and arrives at\nC(x) =/summationdisplay\niC(x−mi), (5)\nwhich enables a very speedy way to compute the total number of pa ths reach-\ning position x. In the context of de novopeptide sequencing [6], this number\ncorresponds to the total number of ⁀all possible de novopeptides within a given\nsmall mass range. Although simply obtained, this number may be usef ul for\nproviding rough statistical assessment in de novopeptide sequencing.\n3.2 The More Realistic Case\nIn general, one may wish to associate with each hop an energy hor one may\nwish to introduce some kind of score normalization based on the numb er of\nhopping steps. This is indeed the case when applying this framework t o MS-\nbased peptide sequencing where a peptide length factor adding or m ultiplying\nto the overall raw score is a common practice. In this case, it becom es impor-\ntant to keep track the number of hops made in each path. We may fu rther\ncategorize the counter C(x,N) into/summationtext\nLC(x,N,L). That is, we may separate\nthe paths with different number of steps from one another and arr ive at a\nfiner counter C(x,N,L) which records the number of paths reaching position\nxwith score Nηand with Lhopping steps.\nIt is rather easy to write down the recursion relation obeyed by this fine\n5counter\nC(x,N,L) =K/summationdisplay\ni=1C(x−mi,N−s(x−mi)\nη,L−1). (6)\nThis recursion relation allows us to renormalize the raw score based o n the\nnumber of steps taken. For example, for RAId DbS [7], a database search\nmethod we developed, we divide the raw score obtained by 2( L−1) for any\npeptide (path) of Lamino acids (hopping steps) to get better sensitivity in\npeptide identification.\nIn principle, the recursion relations given by (4-6) are all one-dimen sional up-\ndates. The only difference is the internal structure of counters a t each position\nx. For (5), the counter is just an integer and has no further struc ture. For (4),\nthe counter at each position has a 1D structure indexed by the sco re. For\n(6), the counter at each position xhas a 2D structure indexed by both the\nscore and the number of hopping steps. This means that in terms of solving\nthe problem using dynamical programming, it is always a 1D dynamical p ro-\ngramming with different degrees of internal structure that may len gthen the\nexecution time when shifting from the simplest case (5) to the more c ompli-\ncated case (6). Obviously at each position x, there is an upper bound and\na lower bound for score and also for the number of hopping steps ac cumu-\nlated. We shall call them Sbest(x),Sworst(x),Lmax(x) andLmin(x) respectively.\nThe first two quantities may be obtained by eqs. (2) and (3) respec tively. We\nprovide the recursions for the two latter quantities below\nLmax(x)= max\n1≤i≤K{Lmax(x−mi)}+1, (7)\nLmin(x)= min\n1≤i≤K{Lmin(x−mi)}+1. (8)\nEqs. (2-3) and (7-8) provides the ranges for both the scores an d the number\nof cumulative hopping steps at each position xvia simple dynamic program-\nming. As we will discuss later, this information enables a memory-efficie nt\ncomputations of score histograms.\n4 Application in MS-based Peptide Sequencing\nIn this section, we focus on an important subject in modern biology – using\nMS data to identify the numerous peptides/proteins involved in any g iven\nbiological process. Because of the peptide mass degeneracies and the limited\nmeasurement accuracyforthepeptidemass-to-chargeratio,u singMS2spectra\n6is more effective in peptide identifications. In a MS2setup, a selected peptide\nwith its mass identified by the first spectrometer is fragmented by n oble gas,\nand the resulting fragments are analyzed by a second mass spectr ometer.\nAlthough such MS2-based proteomics approaches promise high throughput\nanalysis, the confidence level assignment for any peptide/protein identified is\nchallenging.\nThe majority of peptide identification methods are so-called databa se search\napproaches. The main idea is to theoretically fragment each peptide in a\ndatabase to obtain the corresponding theoretical spectra. One then decides the\ndegree of similarity between each theoretical spectrum and the inp ut query\nspectrum using a scoring function. The candidate peptides from th e database\nare ranked/chosen according to their similarity scores to the quer y spectrum.\nAlthough one may assign relative confidence levels among the candida te pep-\ntides via various (empirical) means, an objective, standardized calib ration\nexists only recently [8]. In our earlier publications [5,9], we proposed to tackle\nthis difficulty by using a de novosequencing method to provide an objective\nconfidence measure that is both database-independent and take s into account\nspectrum-specific noise. In this paper, we will provide concrete alg orithms for\nsuch purpose.\nTo begin, consider a spectrum σwith parent ion mass range [ w−δ,w+δ],\nwe denote by Π( w,δ) the set of all “possible” peptides with masses in this\nrange. Given a peptide πfrom Π(w,δ), the associated quality score S(π,σ)\nis defined by a prescribed scoring system. The score distribution of S(π,σ)\nwithin Π( w,δ) provides naturally a likelihood measure for any given peptide\nπto the the correct one.\nHowever, as described earlier [5], this seemingly straightforward ide a faces two\ndifficulties in terms of implementations. First, unlike the DPRM problem f or\nwhich the function to be optimized is defined without ambiguity, the ch oice\nof the scoring function is somewhat empirical because the paramet ers used\nin the scoring must be trained using a training data set. Further, be cause of\ndifferent instruments and experimental setups, it seems impossible to design a\nscoring system such that the correct peptide for each spectrum has the highest\nscore among all possible peptides; the application of a given scoring function\nto general cases may require a leap of faith. Second, even after t he scoring\nfunction is chosen, it is not known how to find the peptide πothat maximizes\nS(π,σ) as well as the score distribution pdf( S) within Π( w,δ) other than by\nthe generally impractical procedure of examining all members of Π( w,δ).\nThe first difficulty can be alleviated by validating high scoring de novopep-\ntides via database searches [9] and is not the main focus of the curr ent paper.\nNote that a partial solution to the second problem via iterative mapp ing when\nnonlocal score contributions exist is provided earlier [5]. Here we tac kle the\n7second problem head on when the scoring function used does not co ntain\nnonlocal contribution other than a final renormalization with respe ct to the\npeptide length. Our algorithms contains two parts: computer memo ry alloca-\ntions anddynamical programming update. Prior to discussing these two parts,\nhowever, we first address the important issue of choosing a good m ass unit.\n4.1 Choosing a Good Mass Unit\nThe goal here is to choose a mass unit ∆ and expresses the molecular mass\nof each amino acid as an integral multiple of this unit. For example, one may\nchoose ∆ to be 0 .1 Dalton (Da), and round the molecular mass of each amino\nacid to be an integral multiple of 0 .1 Da. Once a mass unit is chosen, all the\nmasses under consideration are integral multiples of this unit. It tu rns out\nthat different choices of the mass unit leads to different maximum cum ula-\ntive mass error. As a specific example, consider using ∆ = 0 .1 Da as the mass\nunit. The mass of Alanine, with true mass 71 .03711538 Da, is now represented\nas 710∆. This molecular mass expression is 0 .03711538 Da smaller than the\ntrue molecular mass of Alanine. When this happens, the integral mas s rep-\nresentation has a mass smaller than the true mass, and we call such type\nof mass error a down-error. Now the amino acid Tryptophan with molecular\nmass 186 .07931613 Da will be assigned an integral mass of 1861∆, which has\nan extra of (0 .1−0.07931613) Da compared to the true mass. We call this\ntype of mass error the up-error.\nThe ratio of the mass error to the real molecular mass when multiplied by\n3000 Da provides the cumulative maximum error that can be induced b y a\nsingle amino acid at 3 ,000 Da mass. For a fixed mass unit, we went over this\nmass error analysis for each of the twenty amino acids and documen ted the\nlargest up-error and down-error. The larger one between the ma ximum up-\nerror and the maximum down-error is called the max-error. To sear ch for best\nmass units that minimize the max-error at 3 ,000 Da, we went over all possible\nmass unit ranging from 0 .005 Da to 1 .005 Da in step of 10−6Da. Interestingly\nenough, we found a discrete list of mass units that have smaller max- error\ncompared to their nearby mass units. These numerically found magicmass\nunits are summarized in table 1.\nOnce a mass unit is chosen, all the amino acid masses are effectively int egers.\nTo obtain the score histogram of all de novo peptides when queried by a\nspectrum σwith parent molecular mass w(with N- and C- terminal groups\nof the peptide stripped away), we first construct a mass array wh ere index\nkcorresponds a molecular mass k∆. To encode all possible peptides with\nmolecular mass up to w, we need to have an array of size w/∆+1.Apparently,\nwhen a larger mass unit is used, the size of the mass array is smaller an d thus\n8Table 1\nA list of best mass units in Da. The abbreviation “m.u.e.” sta nds for “maximum\nup-error,” while “m.d.e.” stands for “maximum down-error. ” The maximum up-\nerror, maximum down-error, and max-error are evaluated in e xtrapolation to 3 ,000\nDa as described in the text. The abbreviation “a.a.w.m.u.” s tands for “amino acid\nwith maximum up-error,”while “a.a.w.m.d.” stands for “ami no acid with maximum\ndown-error.”\nmass unit m.u.e. a.a.w.m.u. m.d.e. a.a.w.m.d. max-error\n0.006070 0.041980 Tryptophan 0.037455 Cysteine 0.041980\n0.007300 0.041495 Methionine 0.061276 Asparagine 0.06127 6\n0.017540 0.094183 Cysteine 0.121977 Proline 0.121977\n0.021500 0.199585 Arginine 0.182283 Asparagine 0.199585\n0.054470 0.453793 Asparagine 0.347792 Alanine 0.453793\n0.065400 0.553492 Lysine 0.536989 Alanine 0.553492\n0.109450 0.908287 Proline 0.900898 Lysine 0.908287\n0.110300 0.962781 Histidine 0.858742 Lysine 0.962781\n0.110320 0.960176 Aspartate 0.907801 Histidine 0.960176\n0.500208 0.980357 Cysteine 0.983149 (Iso)Leucine 0.98314 9\n1.000416 0.980357 Cysteine 0.983149 (Iso)Leucine 0.98314 9\nreduces computation time. However, as one may see from table 1, t he larger\nmassunitisalsoaccompaniedbyalargermax-errorandmightnotbep referred\nwhen high mass accuracy is the first priority.\n4.2 Efficient Memory Allocation\nThe basic idea of our algorithm is to encode all possible peptides in the m ass\narray by linking pointers, analogous to the consecutive hopping ste ps in the\n1D hopping model. For an amino acid a, letn(a) represents its corresponding\ninteger mass in unit of ∆. For a peptide made of [ a1,a2,...,a M], it will have\na hopping trajectory in the molecular array given by [0 ,x1,x2,...,x M] with\nxi≥1≡/summationtexti\nj=1n(aj). Let us also denote xMbyxFto indicate that it is the\nterminatingpointofthepath.Apparently, allpossiblepeptideswith molecular\nmasses equal to xF∆ will all have corresponding hopping paths starting at the\norigin and terminating at xF. Through appropriate pointer linking, one may\ntherefore encode allpossible peptides with molecular mass xF∆ in a one-\ndimensional mass array.\n9For a given spectrum σ, depending on the score function used, one may cal-\nculate local score contributions at each mass index. This step is don e once\nonly for the whole mass array, and need not be repeated for each c andidate\npeptide. In a typical MS2experimental spectrum, there always exists some\nlevel of parent ion mass uncertainty. Once the size of the mass unc ertainty\nis specified, we only need to examine de novopeptides whose corresponding\nhopping paths terminating at a few consecutive mass indices. This ind icates\nthat some of the mass indices of the aforementioned mass array ma y not even\nbe used in this context. Below we describe how to efficiently obtain relevant\nmass indices and only allocate computer memories for those masses.\nAssume that the possible terminating points are F1,F2,...,F kwithFj+1=\nFj+ 1. The update rules described in Eqs. (2-3), (5), and (7-8) will als o be\nused at this stage. The following pseudocode describes our algorith m.\nInitialize the mass index = 0 entry\nSbest=Sworst=Lmax=Lmin= 0;C=1;\nREMARK: Max aa is the maximum number of amino acids considered\nfor (aa index = 0; aa index=F1; massindex --) {\nbacktrack all possible paths →final occupied entries;\n}\nThelaststepinthealgorithmaboveidentifies relevantmassindices ,massindices\nthat will be traversed by the hopping paths of all peptides with molec ular\nmasses in the range [ F1∆,Fk∆]. We only need to allocate computer memory\nassociated with those sites. For each of these relevant sites, we a lso know the\nvalues of Sbest,Sworst,Lmax,Lmin, and the total number of peptides reaching\nthat site through the algorithm above. One may therefore allocate a 2D array\nof size (Sbest(i)−Sworst(i))/η×(Lmax(i)−Lmin(i)) for each relevant mass index\nifor later use.\n104.3 Main Algorithm and some Results\nOnce memory allocation for relevant mass indices is done, we can efficiently\ngo through those relevant sites to obtain the 2D score histogram t hat we\nmentioned. In the pseudocode below, update is performed using eq . (6). We\nnow demonstrate the very simple main algorithm\nInitialize all the fine counters C(x,N,L) = 0\nexceptC(x= 0,N= 0,L= 0)=1;\nfor (aa index = 0; aa indexSi·Sj(1)\nwhich has been used to describe many physical systems\nin the high-temperature superconductors32. The hop-\nping amplitude tij=t,t′, andt′′for sitesiandjbeing\nthe nearest-, the second-nearest, and the third-nearest-\nneighbors, respectively. We restrict the electron creation\noperators ˜c†\niσto the subspace with no-doubly-occupied\nsites.Siis the spin operator at site iand< i,j >\nmeans that the interaction occurs only for the nearest-\nneighboring sites. In the following, we mainly focus on\nthe caset′′=−t′/2 andJ/t= 0.3 at hole doping 1 /8.\nWe shall follow the work by Himeda et al.25to con-\nstruct the variational wave functions. In the mean-field\ntheory we assume a local AF order parameter, the stag-\ngeredmagnetization mi, and nearestneighborpairingor-\nder parameter ∆ ij. Thus the effective mean-field Hamil-\ntonian is reduced to\nHMF=/summationdisplay\ni,j/parenleftBig\nc†\ni↑ci↓/parenrightBig/parenleftbiggHij↑Dij\nD∗\nji−Hji↓/parenrightbigg/parenleftbiggcj↑\nc†\nj↓/parenrightbigg\n,(2)\nwhere the matrix elements\nHijσ=−\ntv/summationdisplay\nβ=N+t′\nv/summationdisplay\nβ=NN+t′′\nv/summationdisplay\nβ=NNN\nδj,i+β\n+/parenleftbig\nρi−µv+σ(−1)xi+yimi/parenrightbig\nδj,i, (3)\nDij=/summationdisplay\nβ=N∆ijδj,i+β. (4)\nHereβ=N,NN, andNNNcorrespond to the nearest-\n, the next-nearest, and the third-nearest-neighbors, re-\nspectively, and σ=↑(1) or↓(-1). The local charge den-\nsity is controlled by ρiandµvis the variational param-\neter for the chemical potential. For periodic stripes we3\nassume charge density ρiand staggered magnetization\nmiare anti-correlated, i.e.there are more holes at sites\nwith minimum staggered magnetization. For simplicity,\nwe assume these spatially varying functions with simple\nforms:\nρi=ρvcos[4πδ·(yi−y0)], (5)\nmi=mvsin[2πδ·(yi−y0)], (6)\nwhereδis the doping density and is 1 /8 in this paper.\nWe have also taken the lattice constant to be our unit.\nHere we assume the stripe extends uniformly along the x\ndirection.y0= 0 (1/2) corresponds to the site- (bond-)\ncentered stripe. In this paper, we will only focus on the\nbond-centered stripe since the variational energy differ-\nence between the site- and bond-centered stripes is very\nsmall (not shown) within the finite cluster33,34.\nBesides staggered magnetization and charge density,\nthe nearest neighbor pairing ∆ ijcan also have a spa-\ntial modulation. There are several different stripes we\ncan choose. If ∆ ijhas the same period 1 /δas the stag-\ngered magnetization but it is πphase shifted, this is\nthe so called ”antiphase” stripe studied by a number of\ngroups25,33,34,35. In this stripe state, the bond-average\n∆ijis zero and there is no net pairing. Hole density is\nmaximum at the sites with maximum pairing amplitude\n|∆ij|and minimum staggered magnetization |mi|.\nAnother moregeneralchoiceis to havethe spatialvari-\nation of ∆ ijof the form,\n∆i,i+ˆx= ∆M\nvcos[4πδ·(yi−y0)]−∆C\nv,\n∆i,i+ˆy=−∆M\nvcos[4πδ·(yi−y0)+2πδ]+∆C\nv.(7)\nThe bond-average∆ ijis determined by the constant∆C\nv.\nIf both ∆M\nvand ∆C\nvare positive, then the hole density is\nmaximum at sites with smallest pairing amplitude |∆ij|\nand smallest magnetization |mi|. This is similar to the\nphase diagram36,37predicted by the uniform RVB and\nAF states, when hole density is small both staggered\nmagnetizationandpairingamplitude arelarger. Thus we\nwilldenotethisstateastheAF-RVBstripe. ForAF-RVB\nstripe the period of ∆ ijis the same as the charge-density\nmodulation ρi. This period, 1 /2δ, is only half of the\nperiod for the antiphase or πphase stripe. Besides the\nantiphase stripe and AF-RVB stripes, we could also have\nthe AF stripe without both ∆M\nvand ∆C\nvor the charge-\ndensity stripe without any staggered magnetization but\nwith pairing amplitude modulation.\nIn general we have total seven variational parameters\nµv,t′\nv,t′′\nv,ρv,mv, ∆M\nv, and ∆C\nvwithtvset to be 1. Once\nthese parametersaregiven, wediagonalizethe mean-field\nHamiltonian in equation (2). By solving the Bogoliubov\nde Gennes (BdG) equations\n/summationdisplay\nj/parenleftbiggHij↑Dij\nD∗\nji−Hji↓/parenrightbigg/parenleftbiggun\nj\nvn\nj/parenrightbigg\n=En/parenleftbigg\nun\ni\nvn\ni/parenrightbigg\n,(8)and then obtain Npositive eigenvalues En(n= 1−\nN) andNnegative eigenvalues ¯Enwith corresponding\neigenvectors ( un\ni,vn\ni) and (¯un\ni,¯vn\ni). The eigenvectors are\nused to construct the mean-field wave function25|ψ∝angbracketrightby\nusing Bogoliubov transformation\n/parenleftbigg\nγn\n¯γn/parenrightbigg\n=/parenleftbigg\nun\nivn\ni\n¯un\ni¯vn\ni/parenrightbigg/parenleftbiggci↑\nc†\ni↓/parenrightbigg\n. (9)\nThe trial wave function P|ψ∝angbracketrightwith a Gutzwiller projec-\ntorPcan be constructed by creating all negative energy\nstates and annihilating all positive energy states on a\nvacuum of electrons |0∝angbracketright. Then we formulate the wave\nfunction in the Hilbert space with the fixed number of\nelectronsNe,\n|Φ∝angbracketright=PPNe|ψ∝angbracketright=PPNe/productdisplay\nnγn¯γ†\nn|0∝angbracketright (10)\n∝P\n/summationdisplay\ni,j(ˆU−1ˆV)ijc†\ni↑c†\nj↓\nNe/2\n|0∝angbracketright,\nwhereˆUij=ui\njandˆVij=vi\nj. We optimize the varia-\ntional energy E=∝angbracketleftΦ|H|Φ∝angbracketright/∝angbracketleftΦ|Φ∝angbracketrightby using the stochas-\ntic reconfiguration algorithm38. Additionally, to reduce\nthe boundary-condition effect in numerical studies45, we\naveragetheenergiesoverthe fourdifferent boundarycon-\nditions which is periodic or antiperiodic in either xory\ndirection.\nFig.1 shows the t′/tdependence of the variational en-\nergies for the hole density δ= 1/8. Firstly, we have\nshown that the AF-RVB stripe state (the empty trian-\ngles) becomes more stable than uniform d-wave RVB\nstate (the filled triangles) as decreasing t′/tfurther from\n−0.05. It is worth noting that the AF-RVB stripe state\nfort′/t <−0.05 has the vanishing pairing parameters\n∆C\nvand ∆M\nvbut largemvand finiteρv. Therefore, we\ncan consider this stripe state without SC order as the an-\ntiphase AF stripe state1. The hole density and the stag-\ngeredmagnetizationareplotted asafunction ofpositions\nfor a typical AF stripe (filled circles) in Fig.2(a) and (b),\nrespectively. The AF stripe pattern with a very large\nhole-density variation is essentially a nano-scale phase\nseparation with hole-rich and hole-poor regions alternat-\ning. This is consistent with what has been obtained by\nHimedaet al.25although they did not include t′′which\nis set to be −t′/2 here.\nIn our previous calculations to study the possibility\nof phase separation in the t−Jmodel46, we found the\ntendency to overestimate the strength of the pairing of\nholes. If we reduce this strength by going beyond the\nsimple trial wave functions used in our discussion above\nwe could push the phase separation boundary to a much\nhigher value of J/t. A simpler way to make this ad-\njustment is to introduce the hole-hole repulsion Jastrow\nfactor39,40,41:\nPJ=/productdisplay\ni\n-0.50-0.250.000.250.50\n2 4 6 8 10 12 14 160.0000.0010.0020.0030.0040.005Pxx(y,Ry)\ny(a)\n(b)\n(c)\nFIG. 2: The profiles of (a) hole density, (b) staggered mag-\nnetization and (c) pair-pair correlation function with Ry= 8\nfor the optimized states at 1 /8 doping in the extended t−J−\nmodel with t′/t=−0.3 andJ/t= 0.3. The filled (empty)\ncircles correspond to the AF-RVB stripe state without (with )\nhole-hole repulsion. In (c), the dashed (dashed-dotted) li ne\nindicates the uniform d-wave RVB state without (with) hole-\nhole repulsion. The red solid line corresponds to the random\nstripe state with hole-hole repulsive Jastrow correlation . All\nquantities are calculated on a 16 ×16 lattice system with the\nperiodic and antiperiodic boundary conditions along the x\nand y directions, respectively.\nthe AF-RVB stripe with a much more reduced staggered\nmagnetization as shown in Fig.2(b) and a larger pairing\nparameter ∆C\nv. Now the hole density has a much smaller\nvariation as shown by the empty circles in Fig.2(a).\nIt is surprising to find out that for all t′/t, the AF-\nRVB stripe state is almost degenerate in variational en-\nergy with the uniform d-wave RVB state. This is quite\nremarkable as the two trial wave functions are very dif-\nferent. As an illustration we show in Fig.2(a) and (b)\nthe variation of the hole density and the staggered mag-5\nnetization along the modulation direction, respectively,\nfor an AF-RVB stripe state (empty circles). It should\nbe noticed that not only the hole density is completely\nuniform for d-RVB state but there is also no staggered\nmagnetization at all. Instead of periodic stripes, we have\nalso examined the stripe state with 4 ×4 patches in the\n16×16 lattice system. For each patch, we choose a di-\nrection of the stripe, xory, randomly. We consider this\nstate as random stripe state. For simplicity, we still use\nthe same equations (5-7,11). The hole-density modula-\ntion of a random stripe state has been shown in the inset\nofFig.1. AsshowninFig.1,therandomstripestateswith\nthe Jastrow factors of equation (11) also have the opti-\nmized energies almost identical to uniform d-wave RVB\nand AF-RVB stripe states even though we have not op-\ntimized parameters on very bond or site independently.\nThese optimized random stripe states have finite mvand\n∆C\nvbut smaller ∆M\nv(less than one third of ∆C\nv).\nWe believe this energy degeneracy is caused mainly by\ntwo reasons. The first reason is that the terms in the\nt−JHamiltonian are all local within nearest neighbors\nor next nearest neighbors. The second reason is that the\nhopping terms and spin interaction not only are of the\nsame order of magnitude but they are competing against\neach other. We have found the energy competition be-\ntween the kinetic energy and spin interaction is very ro-\nbust among different states (not shown). Their competi-\ntion is enhanced by the no-doubly occupancy constraint\nas the presence of holes will suppress the spin interaction\nto zero. Thus it is possible to have locally different spin-\nhole configurationswith different emphasis on the kinetic\nenergy or the spin energy. Some of these patterns have\nlowerkinetic energy but higher magnetic energy than the\nuniformd-RVB state and some with opposite energetics.\nThere are at least two important implications of this\nenergy degeneracy. The first one is that the inhomoge-\nneousstatesarequite robustin the extended t−Jmodels\nwithout the need for introducing any other large inter-\nactions. States with different local arrangement of spin\nand holes may have very similar energies. The second\nimplication is that any small additional interaction could\nbreak the degeneracy. If we consider the ”realistic” situ-\nation of cuprates with large numbers of impurities, disor-\nders and electron-latticeinteractions, the inhomogeneous\nstates could be much more numerous and complex than\nwe have expected. Materials made with different pro-\ncessing conditions could also produce different inhomo-\ngeneous states. Since all these are presumably secondary\ninteractionssmallerthan the dominant tandJ, the mod-\nulations of charge density, staggered magnetization and\npairing are expected to be small. If one of the inter-\nactions, like electron-lattice interaction, becomes quite\nstrong as observed in LBCO−1/816, then the modula-\ntion will also become larger and longer ranged.\nWe have also investigated the pair-pair correlation\nfunction for the optimized states with/without hole-hole\nrepulsive Jastrow factors in the case of ( t′,t′′,J)/t=\n(−0.3,0.15,0.3). The singlet pair-pair correlation func-tion along the modulated direction (y direction) is de-\nfined as\nPxx(y,Ry) =1\nLx/summationdisplay\nx|∝angbracketleftΦ|∆†\nx(r)∆x(r+Ry)|Φ∝angbracketright|\n∝angbracketleftΦ|Φ∝angbracketright,(12)\nwherer= (x,y) and ∆†\nx(r) =c†\nr↑c†\nr+ˆx↓−c†\nr↓c†\nr+ˆx↑creates\na singlet pair of electrons among the nearest neighbors\nalongxdirection for each site r. Here, we focus on the\nlong range correlation, and thus set Ry= 8 to be the\nlargest distance on 16 ×16 lattice system. In Fig.2(c), it\nis shown that the hole-holerepulsion suppressesthe long-\nrange pair-pair correlation about three times in the uni-\nformd-waveRVBstate. Wehavealsofound that without\nhole-hole repulsive correlation, the long-range pair-pair\ncorrelation is much more reduced in the AF-RVB stripe\nstate than uniform d-wave RVB state, because the AF-\nRVB stripe state has almost vanishing ∆C\nv. Due to large\nmv, this AF-RVB stripe state shows much larger ampli-\ntude of the Pxxmodulation. However, after considering\nthe hole-hole repulsion, the AF-RVB stripe and uniform\nd-wave RVB states have almost the similar magnitude\nof pair-pair correlation as shown by the empty circles\nand dashed-dotted line in Fig.2(c), respectively. For the\nrandom stripe state shown in the inset of Fig.1 the aver-\nage valuePxxfor the whole system is shown as the red\nsolid line in Fig.2(c). It is essentially the same as the\nvalue of uniform d-wave RVB state and the periodic AF-\nRVB stripe state. Although the staggered magnetization\nfor the random stripe state has larger variation than the\nperiodic stripe state shown in Fig.2(b), the pairing cor-\nrelation is unchanged. These results indicate that the\nlong-range pair-pair correlation is mostly determined by\nthe value of ∆C\nvand is rather insensitive to the modu-\nlation of the hole density and staggered magnetization.\nThus we also expect to have a robust d-wave node ob-\nserved by recent experiments10,13. This will be shown\nbelow.\nIII. DENSITY OF STATES BY THE\nGUTZWILLER APPROXIMATION\nAccording to the VMC calculation for the extended\nt−Jmodel, it is likely that there are a number of inho-\nmogeneous states close in energy to the uniform ground\nstate. Then, some sort of small perturbation may choose\na particular stripe state as the ground state. Assum-\ning such a situation, here we regard a stripe state as the\nground state, and consider the projected quasi-particle\nexcitation spectra. However, calculation of the excited\nstates by the VMC method is computationally very ex-\npensive, and it takes too much time to investigate wide\nparameterrangetoobtaingeneralpropertiesofthestripe\nstates. Furthermore, one can take only a limited system\nsize and it is difficult to obtain dense spectra.\nTherefore, as a first step, we use a Gutzwiller mean-\nfield approximation for this purpose. The minimization6\n/Minus1012\nE/Slash1tv0.050.10.15LDOS/LParen1a/RParen1\n/Minus1012\nE/Slash1tv0.050.10.15LDOS/LParen1b/RParen1\nFIG. 3: Spatially averaged local DOS for the random stripe\nstates calculated bythe non-self-consistent BdG equation . (a)\nFor ∆C\nv/negationslash= 0, we use parameters optimized by the VMC, t′\nv=\n−0.35,t′′\nv= 0.16,m= 0.15,ρ= 0.03, ∆C\nv= 0.28 , ∆M\nv=\n0.02, butµ=−0.875tvis adjusted to realize 1 /8 filling, in\nunits oftv. (b) The same parameters except for ∆C\nv= 0.\nof the total energy yields a BdG equation49. Usually\nthe parameters in the BdG equations are solved self-\nconsistently to find an optimal solution. However, since\nwe already have the assumed inhomogeneous ground\nstate here, we shall use parameter sets obtained from\nVMC results and diagonalize the BdG Hamiltonian only\nonce, instead of solving self-consistently. Furthermore,\nfor convenience, we slightly modify the Gutzwiller pro-\njection by attaching fugacity factors. Namely, we assume\nthatPλ|ψ∝angbracketrightis the ground state and that Pλγ†\nn|ψ∝angbracketrightare the\nexcited states, where Pλ≡P/producttext\niσλniσ\niσ, andλiσis a fu-\ngacity factor to impose the local electron density conser-\nvationforeachspin, namely,/angbracketleftψ|PλˆniσPλ|ψ/angbracketright\n/angbracketleftψ|P2\nλ|ψ/angbracketright=/angbracketleftψ|ˆniσ|ψ/angbracketright\n/angbracketleftψ|ψ/angbracketrightfor\nanyiandσ. Thequasi-particleoperators γ†\nnareobtained\nby solvingthe BdG Hamiltonian. In addition, here we do\nnot take into account the Jastrow factor as it only affects\nthe hole-hole correlation slightly but not the local DOS\nstudied below. Then, under the assumption of the non-\nself-consistency, the BdG Hamiltonian is represented by\nequation (2). With this formulation, Pλγ†\nn|ψ∝angbracketrightof differ-\nentnare approximately orthogonal to each other49, and\nthus we expect that it is suitable to use Pλinstead ofP\nfor our purpose here. Since the result presented below\nis qualitatively not very sensitive to small change of the\nparameters, we expect that such a modification of the\nprojection should not affect the results qualitatively.\nThen, by taking the most dominant terms, the local\nDOS is represented by\nN↑(R,ω) =gt\nR↑/summationdisplay\nn|un\nR|2δ(ω−En),(13)\nN↓(R,ω) =gt\nR↓/summationdisplay\nn|vn\nR|2δ(ω+En),(14)/Minus1.5 /Minus1 /Minus0.500.5 11.5 2\nE/Slash1tv0.050.10.150.2LDOS\nFIG. 4: Local DOS at eight different positions for the ran-\ndom stripe states with ∆C\nv/negationslash= 0 calculated by the non-self-\nconsistent BdG equation.\nwhere index nruns for both positive and negative eigen-\nvalues. Note that only the position dependent constant\ngt\nRσ≡(1−nR↑−nR↓)/(1−nRσ) is multiplied in front\nof the local DOS by the standard BdG formalism. Since\nthe result of the site-centered stripe is very similar to\nthat of the bond-centered stripe, we show only the lat-\nter. Here we shall only discuss our results for the ran-\ndom stripe state. For the random stripe state consid-\nered above, each randomly oriented domain is assumed\nto have the same parameters so that the VMC calcula-\ntion is possible. Ideally we should have optimized these\nvariationalparameters on everybond or every site. Then\nwe expect to have a much broader distribution of these\nparameters. To simulate this effect, we simply replace\neach ∆ijby (1 +ξij)∆ij, whereξijis a random vari-\nable which has the Gaussian distribution around 0 with\nthe standard deviation of 1. We use a supercell of size\n32×32 sites, and the same configuration is repeated as\n20×20 supercells to obtain the local DOS. The Fourier\ntransformwith respect to the supercell index is similarto\na system of small clusters with many twisted boundary\nconditions.\nAs shown in Fig.3(a), the spatially averaged/summationtext\nσNσ(R,ω) for the AF-RVB stripe (∆C\nv∝negationslash= 0) has a V-\nshape at low energy. On the other hand for an antiphase\nstripe which has ∆C\nv= 0 there is no V-shape as shown in\nFig.3(b). The energy level is broadened with a width of\n0.05tv. In Fig.3(b), there is a very small dip at E= 0, it\nwould be bigger if ∆M\nvincreases. Our results are consis-\ntent with the very recent report by Baruch and Orgad31.\nIn general, there is no V-shape DOS for the antiphase\nstripe. For the same system as Fig.3(a), in Fig.4 we plot\nthe position dependence of local DOS at randomly cho-\nsen 8 sites. The low energy spectra seem less influenced\nby the disorder than high energy. This result shows that\nthe node and the low energy V-shape DOS are robust\nagainst this kind of inhomogeneity. This is possible be-\ncausenodal k-pointsdonothavemanystatestomix with\nand also the suppression of impurity scattering50. The\nsub-gap structures12,13also seem to be quite apparent.7\nIf we switch off the random variables ξij, the gap vari-\nations from site to site are small. These gap variations\ngrow larger as distributions of ξijbecome wider.\nThe key to understand the absence of V-shape in the\nLocal DOS for the antiphase stripe is the Fourier trans-\nform of the modulated ∆ ijterm written as,\n∆M\nv/summationdisplay\nkcoskx(e−iθc†\nk+q,↑c†\n−k,↓+eiθc†\nk−q,↑c†\n−k,↓+h.c.)\n−∆M\nv\n2/summationdisplay\nk/bracketleftbigg/parenleftBig\neiky+e−i(ky+qy)/parenrightBig\nei(qy\n2−θ)c†\nk+q,↑c†\n−k,↓\n+/parenleftBig\neiky+e−i(ky−qy)/parenrightBig\ne−i(qy\n2−θ)c†\nk−q,↑c†\n−k,↓+h.c./bracketrightbigg\n,\n(15)\nwhereθ= 0 for site-centered stripes and θ=qy/2 for\nbond-centered stripes; q= (0,π/4) for the antiphase\nstripe, and q= (0,π/2) for the AF-RVB stripe. What\nis important here is that it contains onlypairing with\nnonzerocenter-of-mass momentum as the FFLO state51.\nIn the case of zero-momentum pairing as the conven-\ntional BCS theory, the spin-up electron band couples\nwith the spin-down hole band (the upside-down down\nelectron band). These bands intersect at the Fermi level,\nand a gap opens if ∆ k∝negationslash= 0. In the case of finite- qpair-\ning, however, the spin-up electron band couples with ±q\nshifted spin-down hole bands, and thus the band inter-\nsections occur not at the Fermi level. Therefore, a gap\ndoes not open at the Fermi level. The constant ∆C\nvterm\nwhichformsthe usualCooperpairisnecessaryforhaving\nthe node and the V-shape DOS.\nTo compare with ARPES experiments, A(k,ω) is also\ncalculated. Since A(k,ω) is regarded as the local DOS in\nthek-space, Let us take the Fourier transform of renor-\nmalizedun\nR,vn\nR, namely,\n(˜un\nk,˜vn\nk)≡1√Nsite/summationdisplay\nRe−ikR/parenleftbig\ngt\nR↑un\nR,gt\nR↓vn\nR/parenrightbig\n.(16)\nThen,Aσ(k,ω) is written as\nA↑(k,ω) =/summationdisplay\nn|˜un\nk|2δ(ω−En), (17)\nA↓(k,ω) =/summationdisplay\nn|˜vn\nk|2δ(ω+En). (18)\nFig.5 shows/summationtext\nσAσ(k,ω) of the random stripe state\nwith ∆C\nv∝negationslash= 0, where eachof the δ-function spectraarere-\nplaced with a Lorentzian distribution with δE= 0.05tv.\nIn comparison with the uniform d-wave RVB state, the\nspectral weight around the antinodal region is much\nstrongly reduced than near the node.\nIV. CONCLUSIONS\nIn summary, we have used a variational approach to\nexamine the possibility of having inhomogeneous ground/Minus1.5 /Minus1/Minus0.50\nE/Slash1tv24681012A/LParen1k,E/RParen1/LParen1a/RParen1\nk/Equal/LParen13/SΠaceΠ\n/FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt\n16,3/SΠaceΠ\n/FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt\n16/RParen1k/Equal/LParen14/SΠaceΠ\n/FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt\n16,4/SΠaceΠ\n/FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt\n16/RParen1k/Equal/LParen15/SΠaceΠ\n/FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt\n16,5/SΠaceΠ\n/FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt\n16/RParen1k/Equal/LParen16/SΠaceΠ\n/FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt\n16,6/SΠaceΠ\n/FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt\n16/RParen1\n/Minus1.5 /Minus1/Minus0.50\nE/Slash1tv24681012A/LParen1k,E/RParen1/LParen1b/RParen1\nk/Equal/LParen1Π,2/SΠaceΠ\n/FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt\n16/RParen1k/Equal/LParen1Π,3/SΠaceΠ\n/FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt\n16/RParen1k/Equal/LParen1Π,4/SΠaceΠ\n/FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt\n16/RParen1k/Equal/LParen1Π,5/SΠaceΠ\n/FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt/FractionBarExt /FractionBarExt/FractionBarExt/FractionBarExt\n16/RParen1\nFIG. 5:P\nσAσ(k,E) for the random stripe state with\n∆C/negationslash= 0. Dotted lines are for uniform d-wave RVB state\n(ρv= 0,mv= 0,∆M\nv= 0). (a) Nodal and (b) near antinodal\nregion.\nstates within the extended t−Jmodel with 1 /8 doping.\nWe considered states with spatial modulation of charge\ndensity, staggered magnetization and pairing amplitude.\nBesides the antiphase or inphase stripes considered by\nmany groups we have proposed a new AF-RVB stripe. In\nthis stripe state, we assume there is a constant pairing\namplitude besides the various modulation. In addition\nto considering states with periodic stripes we also con-\nsider random stripe states to simulate the cluster glass\nstate observed by experiments10. By improving the trial\nwave functions with the introduction of hole-hole repul-\nsive correlation,we have greatlyimprovedthe variational\nenergiesbyseveralpercentsforbothuniformRVB d-wave\nSC state and states with periodic AF-RVB stripes for re-\nalistic values of t′/t. Most surprisinglythe random stripe\nstate essentially also has the same energy as the uniform\nstate in spite of our oversimplified assumption that all\nthe stripe domain has the same patterns of modulation\ninstead of each site or bond with different values. This\nrandom stripe state also has about the same long-range\npair-pair correlation as the uniform or periodic stripe\nstate even though there are significant staggered mag-\nnetization and charge variation from site to site. Then\nwe also examined the local DOS and the spectral weight\nof the random stripe state by using Gutzwiller approx-\nimation. We found the V-shape DOS and the node are\nstillpresentateverysite. ThelocalDOSmeasuredatdif-\nferent positions shows a broad variation of the gaps and\nalso it has sub-gap structures seen in experiments12,13.\nThe spectral weight at the antinodal direction is negli-\ngibly small but finite around the node. All these results\narequite consistentwith experimentsreportedfor cluster\nglass state in BSCCO10.\nOur result also resolves an inconsistency with ex-8\nperiments derived from previous theoretical calculations\nwithoutincludingthehole-holerepulsioninthetrialwave\nfunctions. Stripe is neither stabilized nor destabilized by\nthe long range hopping. In fact, due to the competition\nbetween the kinetic energy gain and magnetic interac-\ntion, it is very natural to have the spatial modulation, in\nperiodicorrandomconfiguration,ofchargedensity, mag-\nnetization and even pairing amplitude. The constraint of\ndisallowing doubly occupation of electrons at each lattice\nsite has significantly enhanced the competition. Many\nlocal arrangements of spin and hole configurations could\ngivealmostidenticaltotalenergyasthe uniformsolution.\nRecently, Capello et al.34have also found that the en-\nergy of the periodic RVB stripe state is very close to\nthat of the uniform RVB state by using a variational cal-\nculation. They have considered several possibilities for\nthe stability of the RVB stripe state, such as lattice dis-\ntortion, t’-effect, and long-range Coulomb repulsion with\nthe conclusion that uniform RVB state is still the lowest\nenergy state. This is very consistent with our conclusion\nalthough we haveincluded the antiferromagneticorderin\nthe stripe state. The issue about whether AF is present\nin the stripe will be discussed in the future. They have\nnot considered the cluster glass state with random AF-\nRVB stripe domains which is also a good candidate for\nthe ground state.\nThe presence of inhomogeneous or cluster glass states\nis apparently a very natural consequence of the t−Jmodel. There is no need for introducing additional in-\nteractions to generate such states. In fact we showed\nthat the RVB state with a finite constant pairing is quite\ncompatible with the local variations of charge density,\nmagnetization and even pairing amplitude. As long as\nthis modulation is not overlystrong, the superconductiv-\nity still survives as the node and V-shape DOS are still\npresent. In a realistic material, other interactionssuch as\nimpurity, disorder, and electron-lattice interactions, etc.,\nno doubt will help to determine the most suitable local\nconfigurationofspinsandholesbut they willnot produce\na globally ordered state unless there is a very strong and\ndominant interaction like the electron-lattice interaction\nseen inLa2−xBaxCuO4at 1/8 doping. The verification\nof this is left for future work.\nV. ACKNOWLEDGMENTS\nThisworkissupportedbytheNationalScienceCouncil\nin Taiwan with Grant no.95-2112-M-001-061-MY3. The\ncalculations are performed in the IBM Cluster 1350, For-\nmosaII Cluster and IBM P595 in the National Center\nfor High-performance Computing in Taiwan, and the PC\nFarm and Euler system of Academia Sinica Computing\nCenter in Taiwan.\n1J. Zaanen and O. Gunnarsson, Phys. Rev. B 40, 7391\n(1989).\n2D. Poilblanc andT. M. Rice, Phys.Rev.B 39, 9749 (1989).\n3S. A. Kivelson et al., Rev. Mod. Phys. 75, 1201 (2003).\n4J. M. Tranquada et al., Nature 375, 561 (1995).\n5H. A. Mook, P. Dai, and F. Dogan, Phys. Rev. Lett. 88,\n097004 (2002).\n6M. Fujita, H. Goka, K. Yamada, and M. Matsuda, Phys.\nRev. Lett. 88, 167008 (2002).\n7Y. S. Lee et al., Phys. Rev. B 60, 3643 (1999).\n8B. Lake et al., Nature 415, 299 (2002).\n9A. Lucarelli et al., Phys. Rev. Lett. 90, 037002 (2003).\n10Y. Kohsaka et al., Science 315, 1380 (2007).\n11C. Howald, H. Eisaki, N. Kaneko, M. Greven, and A. Ka-\npitulnik, Phys. Rev. B 67, 014533(2003).\n12M. C. Boyer et al., Nature Phys. 3, 802 (2007).\n13J. W. Alldredge et al., cond-mat/0801.0087.\n14W. S. Lee et al., Nature, 450, 81 (2007).\n15A. Kanigel et al., Phys. Rev. Lett. 99, 157001 (2007).\n16T. Valla et al., Science 314, 1914 (2006).\n17P. W. Anderson, Science 235, 1196 (1987).\n18P. W. Anderson et al., J. Phys. Condens. Matter 16, R755\n(2004).\n19P. A. Lee et al., Rev. Mod. Phys. 78, 17(2006).\n20K. Tanaka et al., Science 314, 1910 (2006).\n21C. S. Hellberg and E. Manousakis, Phys. Rev. Lett. 83,\n132 (1999).\n22S. R. White and D. J. Scalapino, Phys. Rev. B 57, 3031\n(1998); S. R. White and D. J. Scalapino, Phys. Rev. B 60,R753 (1999); S. R. White and D. J. Scalapino, Phys. Rev.\nLett.81, 3227 (1998).\n23T. Tohyama et al., Phys. Rev. B 59, R11649 (1999).\n24K. Kobayashi and H. Yokoyama, J. Low Temp. Phys. 117,\n199 (1999).\n25A. Himeda, T. Kato, and M. Ogata, Phys. Rev. Lett. 88,\n117001 (2002).\n26M. Miyazaki et al., J. Phys. Chem. Solids 63, 1403 (2002).\n27E. Pavarini, I. Dasgupta, T. Saha-Dasgupta, O. Jepsen,\nand O. K. Andersen, Phys. Rev. Lett. 87, 047003 (2001).\n28C. T. Shih, T. K. Lee, R. Eder, C. Y. Mou, and Y. C.\nChen, Phys. Rev. Lett. 92, 227002 (2004).\n29M. Hashimoto et al., cond-mat/0801.0782.\n30A. J. Millis and M. R. Norman, Phys. Rev. B 76,\n220503(R) (2007).\n31S. Baruch and D. Orgad, cond-mat/0801.2436.\n32K. Y. Yang et al., Phys. Rev. B 73, 224513 (2006).\n33M. Raczkowski, D. Poilblanc, R. Fresard, and A. M. Oles,\nPhys. Rev. B 75, 094505 (2007); M. Raczkowski, M.\nCapello, D. Poilblanc, R. Fresard, and A. M. Oles, Phys.\nRev. B76, 140505(R) (2007).\n34M. Capello, M. Raczkowski, and D. Poilblanc, Phys. Rev.\nB77, 224502 (2008).\n35A. Himeda and M. Ogata, J. Phys. Chem. Solids 631423\n(2002).\n36C. T. Shih et al., Low Temp. Phys. 31, 757 (2005).\n37C.P. Chou et al., J. Mag. Mag. Mat. 310, 474 (2007).\n38S. Sorella, Phys. Rev. B 64, 024512 (2001).\n39C. S. Hellberg and E. J. Mele, Phys. Rev. Lett. 67, 20809\n(1991).\n40R. Valenti and C. Gros, Phys. Rev. Lett. 68, 2402 (1992).\n41S. Sorella et al., Phys. Rev. Lett. 88, 117002 (2002).\n42C. P. Chou, T. K. Lee, and C. M. Ho, Phys. Rev. B 74,\n092503 (2006).\n43T. K. Lee, C. M. Ho, and N. Nagaosa, Phys. Rev. Lett.\n90, 67001 (2003).\n44M. Fujita, H. Goka, K. Yamada, J. M. Tranquada, and L.\nP. Regnault, Phys. Rev. B 70, 104517 (2004).\n45D. Poilblanc, Phys. Rev. B 44, 9562 (1991).46C. T. Shih et al., J. Phys. Chem. Solids 621797 (2001).\n47N. Doiron-Leyraud et al., Nature 447, 565 (2007).\n48M. Sera et al., Solid State Commun. 69, 851 (1989)\n49N. Fukushima, cond-mat/0801.2280.\n50Due to the projection operator, the impurity scattering\nmatrix element is strongly renormalized with a factor pro-\nportional to the hole denstiy.\n51Qian Wang, H. Y. Chen, C. R. Hu, and C. S. Ting, Phys.\nRev. Lett. 96, 117006 (2006)." }, { "title": "0807.4962v2.Negative_density_of_states__screening__Einstein_relation__and_negative_diffusion.pdf", "content": "arXiv:0807.4962v2 [cond-mat.str-el] 27 Aug 2008Negative density of states: screening, Einstein relation, and\nnegative diffusion.\nA. L. Efros∗\nDepartment of Physics, University of Utah, Salt Lake City UT , 84112 USA\nAbstract\nIn strongly interacting electron systems with low density a nd at low temperature the thermo-\ndynamic density of states is negative. It creates difficultie s with understanding of the Einstein\nrelation between conductivity and diffusion coefficient. Usin g the expression for electrochemical\npotential that takes into account the long range part of the C oulomb interaction it is shown that at\nnegative density of states Einstein relation gives a negati ve sign of the diffusion coefficient D, but\nunderthis condition thereis nothermodynamiclimitation o n the sign of D. It happensbecausethe\nunipolar relaxation of inhomogeneous electron density is n ot described by the diffusion equation.\nThe relaxation goes much faster due to electric forces cause d by electron density and by neutral-\nizing background. Diffusion coefficient is irrelevant in this c ase and it is not necessarily positive\nbecause process of diffusion does not contribute to the positi ve production of entropy. In the case\nof bipolar diffusion negative Dresults in a global absolute instability that leads to forma tion of\nneutral excitons. Graphene is considered as an example of a s ystem, where the density relaxation\nis expected to be due to electric force rather than diffusion. I t may also have a negative density of\nstates.\nPACS numbers: 71.27. +a,73.50.-h\n1I. INTRODUCTION\nThe idea of the Einstein relation was put forward by Einstein1and Smoluchowski2in\n1905-1906. Both scientists considered the Brownian motion in the p resence of gravitational\nforce. The result is the relation between mobility uin the field and diffusion coefficient D.\nIn case of electric field and particles with the charge eit has a form\neD=Tu, (1)\nwhereTis the temperature in energy units. The main idea was equivalence of a n exter-\nnal force and the density gradient. Of course, both Einstein and S moluchowski did not\ncare about negligible mutual gravitational or any other small intera ctions of the Brownian\nparticles.\nThe formulation of the Einstein relation for electrons is based upon e lectrochemical po-\ntential, thethermodynamic functionthat, like temperature andpr essure, shouldbethe same\nat all points of the system in the equilibrium state. The usual argume nts are asfollows. If an\nexternal potential ψis applied to the system, the condition of thermodynamic equilibrium\nreads\nEec=µ(n)+eψ=Const, (2)\nwhereµ(n) is the chemical potential as a function of inhomogeneous electron densityn.\nIn the equilibrium both nandψare function of coordinates while Eecis constant. The\ntemperature Tshouldalsobeconstant. Therefore, theelectrical currentdens ityjatconstant\nTcan be written in a form3\nj=−σ\ne∇Eec=σE−D∇en, (3)\nwhereσis conductivity and E=−∇ψ. Then one gets relation connecting σandD\nσ\ne2dµ\ndn=D, (4)\nwhich is also called Einstein relation. For the Boltzman gas dµ/dn=T/nand one gets\nEq.(1) ifσ=enu. It looks like derivation of Eq. (4) is independent of the properties o f the\nsystem and this equation can be consider as general thermodynam ic law.\nAsimple observationshows however thatinthe caseof non-ideal ele ctron gasthe Einstein\nrelation needs some comments. We discuss an electron gas on the po sitive background at\n2low temperatures and low densities when dimensionless parameter rsis not very small. Here\nr3\ns= 3/(4πna3\nB) for 3-d case and r2\ns= 1/πn2a2\nB, wherenandn2are 3- and 2-dimensional\nelectron densities respectively and aB=/planckover2pi12κ/me2is the Bohr radius, mis an effective\nelectronic mass, κis an effective permittivity.\nThe problems of dynamic screening and diffusion in slightly non-ideal ele ctron gas(rs<<\n1) with electron-electron interaction were considered in details abo ut 20 years ago (See\nRef.[4,5,6])Inthiscasethethermodynamicdensityofstatesislargeandposit ive. Iconcentrate\nhere on the strongly non-ideal case rs≥1.\nAn electron gas on the positive background at low temperatures an d low densities has\nenergyEof the order of −e2n1/dN/κ, whered= 2,3 is the space dimensionality and nis the\ndensity per area or volume respectively, Nis total number of electrons. Then µ∼ −e2n1/d/κ\nandE,µ, anddµ/dnare negative7,8. The first experimental confirmation of this idea was\ndone by Kravchenko et al9,10, but direct quantitative study of this effect was performed by\nEisenstein et al11,12.\nThe derivative dµ/dnis proportional to the reciprocal compressibility of the electron\ngas. Note that compressibility has to be positive due to the thermod ynamical condition of\nstability. However, this principle cannot be applied to the charged sy stems, like electron gas,\nbecause part of their energy is outside the system in a form of the e nergy of electric field.\nOn the other hand, in the case of a neutral electron-hole plasma, t he situation of negative\ncompressibility can arise leading to collapse of the system. Such a situ ation is considered at\nthe end of Sec. III.\nIt follows from Eq. 4 that if dµ/dnis negative, diffusion coefficient Dand conductivity σ\nhaveoppositesigns. Thisobservationneeds anexplanationbecaus enearthethermodynamic\nequilibrium both of them have to be positive to provide positive entrop y production due to\nthe Joule heat and due to the relaxation of inhomogeneous density.\nII. ELECTROCHEMICAL POTENTIAL AND STATIC SCREENING\nTo resolve this contradiction one should include the long-range part of the Coulomb\npotential created by inhomogeneous electron gas into the functio nEecin Eq. (2). This\ncontribution is a functional of n(r).\nTo findEectaking into account electron-electron interaction one should minimiz e the\n3Helmholtz energy Fwith respect to electron density n(r) at a given value of TandN. For\nlowTone gets\nF=e2\n2κ/integraldisplay /integraldisplayn′(r)n′(r′)d3rd3r′\n|r−r′|+/integraldisplay\nf(n+n′)d3r+\n/integraldisplay\nen′(r)ψd3r−Eec/integraldisplay\nn′(r)d3r, (5)\nwherefis the Helmholtz energy density of a homogeneous electron system t hat results from\nthe interaction in a neutral system, like the Wigner crystal or ”Wign er liquid”. Since this\ninteraction comes mainly from the nearest neighbors and n(r) is a smooth function, one\nmay assume that both fand chemical potential µ=df/dnare local functions of n(r) We\nassume also that n(r) =n+n′(r), wherenis average density and n′≪n.\nMinimization of this expression with respect to n′gives the equation\nEec=µ(n)+eψ+dµ\ndnn′+e2\nκ/integraldisplayn′(r′)d3r′\n|r−r′|. (6)\nIt differs from Eq. (2) by the potential of electrons in the right han d side. Note that\nthis potential is due to the violation of neutrality in a scale much larger than the average\ndistance between electrons. To check this equation we consider th ermodynamic equilibrium\nand find equations for the Thomas-Fermi static screening in 3- and 2-dimensional cases.\nSinceEecis independent of rin thermodynamic equilibrium one may take Eec−µ(n) as a\nreference point for the total potential ϕdefined as\nϕ=ψ+e\nκ/integraldisplayn′(r′)d3r′\n|r−r′|. (7)\nIt follows from Eq. (6) that\neϕ=−dµ\ndnn′. (8)\nThe Poisson equation has a form\n∇2ϕ=−4π(en′−ρext)\nκ, (9)\nwhereρextis density of external charge. Using Eq. (8) one gets final equatio n for the 3-d\nlinear screening\n∇2ϕ=−q2\n3ϕ−4πρext\nκ. (10)\nHere\nq2\n3=4πe2\nκdn\ndµ(11)\n4is the reciprocal 3-dimensional screening radius.\nConsider now a thin layer (x-y plane) with 2d electron gas separating two media with\ndielectric constants κ1andκ2. In this case one should substitute n⇒n2δ(z) andκ⇒¯κ=\n(κ1+κ2)/2. The results is13\n∇2ϕ=−q2ϕδ(z)−4πρext\n¯κ, (12)\nwhere\nq2=2πe2\n¯κdn2\ndµ. (13)\nIt is important that Eqs. (10), (12) are applicable only if the screen ing is linear ( n′≪\nn)14. There is another serious problem of applicability the Thomas-Fermi approximation\nin the case of the negative density of states. Indeed, the dielectr ic permittivity in this\napproximation has a form\nǫ(q) =κ(1−|q2\n3|\nq2) (14)\nin 3-d case and\nǫ(q) =κ(1−|q2|\nq) (15)\nin 2-d case. In both cases it has roots at q=|q3|,|q2|. The expression for the screened\npotentialϕhas a form\nϕ(r) =/integraldisplayϕ0(q)exp(iq·r)dq\nǫ(q), (16)\nwhereϕ0is a bare potential. Thus, the roots of ǫtransform into the first order poles without\nany reasonable way of the detour. Such a detour follows from the c asuality for the ω-plane\nbut not for the q-plane. Moreover, the electrostatic potential s hould be real and one cannot\nadd a small imaginary part in the denominator. Therefore I think tha t the poles do not\nhave any physical sense.\nThereasonisthatnegativesignofthedensityofstatesappearsw henq3,q2areoftheorder\nof average distance between electrons ¯ r. At such distances the very concept of macroscopic\nfield does not have sense. However, if the bare potential has only h armonics with q≪\n|q3|,|q2|, the Eqs.(10,12) have a sense. Consider, for example, the screen ing of the positive\nchargeZat a distance z0from the plane with 2-d gas(plane z= 0. The solution of Eq.(12)\nhas a form13\nϕ(ρ) =/integraldisplay∞\n0Zexp(−qz0)\nκ(q+q2)qJ0(qρ)dq, (17)\n5whereρis a polar radius in the plane z= 0. Suppose that |q2|z0≫1. Now the contri-\nbution to integral Eq.(17) from q≃ |q2|is exponentially small and one can ignore qin the\ndenominator. Then\nϕ(ρ) =Zz0\nκq2(z2\n0+ρ2)3/2. (18)\nNote that at q2<0 a positive charge creates a small negative potential in the plane with\nelectrons. That is what I call ”overscreening”.\nExtra electron density, as calculated from Eq. (8) is\nen′=−Zz0\n2π(z2\n0+ρ2)3/2(19)\nIt is negative and independent of the sign of q2. One can see that the total charge\n/integraldisplay∞\n0en′2πρdρ=−Z (20)\nDue to geometry of the problem electric field is zero below the plane wit h electrons. As\nfollows from Eq. (8), the signs of charge density and potential are opposite if the density of\nstates is negative.\nFor the case of two such planes (double quantum well structure) L uryi15has predicted\na small penetration of electric field through the first plane. He has c onsidered the case of\npositive density of states. Then the small penetrating field betwee n two planes has the same\ndirection as the incident field.\nEisenstein atal.11studiedthiseffectexperimentallyandfoundoutthatatnegativede nsity\nof states the propagating field is opposite to the incident field and th is is also a result of the\noverscreening (see the quantitative theory in Ref.12,16,17).\nNegative density of states was also used18for theexplanation ofmagnetocapacitance data\nby Smith at al.19.\nIII. CONDUCTIVITY VERSUS DIFFUSION\nNow I come back to the problem of the negative diffusion. If the syst em is not in\nequilibrium the electric current can be written in the same form as Eq. (3)\nj=−σ\ne∇Eec. (21)\n6Using Eq. (6) one gets\nj=σE−D∇en′−σe\nκ∇/integraldisplayn′(r′)d3r′\n|r−r′|. (22)\nHere D is connected to σby the Einstein relation Eq. (4). Considering relaxation of the\ncharge density one can ignore external field E. The relaxation is described by the continuity\nequation\n∂(en)\n∂t=−∇·j (23)\nor\n∂(en)\n∂t=σ/parenleftbigg1\ne2dµ\ndn∇2(en′)−4πen′\nκ/parenrightbigg\n. (24)\nThe ratioRof the first (diffusion) term in the right hand side to the second (field ) term is\nR= (q2\n3L2)−1, whereL−2=∇2n′/n′is the characteristic size of the extra charge and q2\n3is\ngiven by Eq. (11). If electron gas is non-ideal, q3∼1/¯r, where ¯ris the average distance\nbetween electrons. However, the very concept of diffusion equat ion is valid at L≫¯r. This\nmeans that for the non-ideal gas |R| ≪1 and the diffusion term in Eq. (24) should be\nignored. Then the equation has a simple solution\nn′(r,t) =n′(r,0)exp−(t/τM), (25)\nwhereτM=κ/(4πσ) is well-known Maxwell’s time. Coefficient Ddoes not enter in this\ncase in the entropy production and it does not have a physical sens e. Thus in 3-dimensional\nnon-ideal electron gas negative dµ/dndoes not create any contradiction with the Einstein\nrelation.\nIn the 3d gas of high density µ∼n2/3andR∼(¯r/L)2/rswithrs<1. In this case R\nmight be large and diffusion is possible. However dµ/dn> 0, andD>0.\nNow we consider the relaxation of the charge density in 2-dimensiona l case. Instead of\nEq.(24) one gets\n∂(en2)\n∂t=σ2(1\ne2dµ\ndn2∇2(en′\n2)\n−e\n¯κ∇2/integraldisplayn′\n2(r′)d2r′\n|r−r′|). (26)\nHeren2,σ2and∇are 2-dimensional density, conductivity, and 2-dimensional gradie nt re-\nspectively. To consider the ratio R2of the first (diffusion) term to the second (field) term it\n7is convenient to make the Fourier transformation. Then one gets\n∂(nq)\n∂t=−σ2(1\ne2dµ\ndn2q2nq+2πq\n¯κnq), (27)\nwherenqis the Fourier transformation of n′\n2.\nNow we find that the ratio of the first ( diffusion) term in the right han d side of Eq.\n(27) to the second (field) term R2=q/q2, whereq2is given by Eq. (13). Similar to the\n3d case in the non-ideal gas |q2| ∼1/¯rand diffusion should be ignored. Then we get the\nDyakonov-Furman equation20\n∂(nq)\n∂t=−vqnq, (28)\nwhere velocity v= 2πσ2/¯κ. The physical meaning of this equation is that extra density of\nelectrons localized initially at some spot propagates in all directions wit h velocityvconserv-\ning the total amount of extra electrons. Of course, this way of re laxation is more efficient\nthan diffusion (random walk), because r∼vtwhiler∼√\nDtin the case of diffusion. Thus,\ndiffusion coefficient Dis irrelevant and negative dµ/dndoes not create any contradiction\nwith the Einstein relationIn a high density electron gas R2=q¯r/rsand diffusion mechanism\nis possible. In this case dµ/dn> 0 andD>0.\nOne can consider this problem from a different point of view. In both 3 d and 2d cases the\nnegative diffusion coefficient Dappears in the term with the highest derivative that leads\nto the absolute instability even if Dis small21. Consider, for example Eq. (24) for 3d case.\nAfter the Fourier transformation the solution for the charge den sityρ=en′can be written\nin a form\nρq=ρ0\nqexp/parenleftbigg\n−4πσt\nκ−Dq2t/parenrightbigg\n, (29)\nwhereDis given by the Einstein relation Eq. (4). One can see that at D <0 solution\nincreases with time exponentially for harmonics with q¯r≥1.\nThe physical explanation is as follows. The Eqs.(24,26) contain avera ge distance between\nelectrons ¯r. So they contain information that the charged liquid has a discreet e lectronic\nstructure. This information comes from the negative density of st ates which originates\nfrom the interaction of the separate electrons. That is why macro scopic equations become\nunstable at small spacial harmonics. The message is that n(r) is rather a set of δ-functions\nthan a continuous function. The instability is absent if Dis positive.\nThe instability of small spatial harmonics at small negative Ddoes not affect larger\nharmonics because Eqs.(24,26) are linear. Due to the linearity differe nt harmonics are in-\n8dependent and transformation of energy from small spacial harm onics to large harmonics is\nforbidden (cp. phenomenon of turbulence in non-linear hydrodyna mics where the transfor-\nmation of energy is not forbidden, but the instability is initiated by larg e harmonics).\nTherefore, I think that at small Dapproximation D= 0 that gives Eqs.(25,28) is correct.\nOne should note that the problem of the non-physical roots of elec tric permittivity dis-\ncussed in the previous section is of the same nature.\nBefore we discussed the unipolar diffusion. Consider the simplest cas e of the ambipolar\ndiffusion assuming that at t= 0 the densities of electrons and holes are equal in some finite\nregion of space and are zero otherwise. Moreover we assume that the local macroscopic\ncharge density ρ(r,t) = 0 and a recombination of carriers is very slow. In this case Eq. (6)\ndescribes the electron-hole system in quasi-equilibrium. At large rsone getsE,µ,dµ/dn< 0\nbut the last term in Eq. (6) is absent. So the smearing of the density of particles is described\nby the equation of diffusion at all rs, but at small density ( rs≥1) coefficient D<0. Then\nthe absolute instability takes place for all harmonics that means a co llapse of the system.\nThus the electron-hole ”Wigner liquid” and crystal are unstable.\nThis result is very transparent. It happens because negative µjust means that the energy\nof the system decreases with increasing density. In bipolar case ne utrality is provided by\nthe particles and we do not consider any background. Thus the inst ability is a result of the\nnegative compressibility in a neutral system. At large enough rsthese particles are classical,\nand the absence of the mechanical equilibrium follows also from the Ea rnshaw theorem. In\nreality quantum mechanics becomes more important with increasing d ensity. As a result\nthe excitons are formed. These neutral particles have a positive d iffusion coefficient Daand\ntheir density smears with time through all available space. This proce ss is described by a\nregular diffusion equation. In the case of optical excitation the car riers may appear in the\nform of the excitons from the very beginning\nFor the coefficient of the ambipolar diffusion Daa textbook equation22\nDa=2DeDh\nDe+Dh(30)\nis often used, where De,hare diffusion coefficients of electron and holes in unipolar case.\nAs follows from the previous discussion, one should be careful with t his equation because\nfor the non-ideal electron (or hole) gas these unipolar coefficients might be negative and\nmeaningless. It happens because in unipolar case there is a deviation from neutrality that\n9creates electric field, while in bipolar case the system is neutral. In th is case Eq. (30) does\nnot work and one should calculate Dain a different way as a diffusion of the exciton.\nIn the recent paper by Zhao23the experimental results for the ambipolar diffusion in\nsilicon-on-insulator system are compared with Eq. (30). At high tem peratures a good\nagreement is found while at low temperatures the observed values o fDaare 6-7 times less.\nThe previously reported values24show similar temperature dependence.\nThe author’s explanation is that coefficients De,hare taken for the bulk silicon using\nEinstein relation and they might be larger than in the film at low tempera tures. However,\nthe reason discussed above cannot be excluded.\nIV. GRAPHENE AS A POSSIBLE EXAMPLE OF A NON-IDEAL ELECTRON\nSYSTEM\nIt is interesting to discuss the single layer graphene as an example of the system with\nnon-ideal electron gas. Graphene is a gapless material with the linea r spectrum of electrons\nand holes near the Dirac point. Due to some reasons, that are not q uite clear now, the\nvelocityvof electrons and holes in equation ǫ=±pvis of the order of e2//planckover2pi1. It follows\nthat at any Fermi energy inside this linear spectrum electron gas in g raphene is non-ideal\nin a sense mentioned above: the absolute value of the chemical pote ntial is of the order of\ninteraction energy e2n1/2. It means that unipolar density relaxation in this system should\nbe described by the Dyakonov-Furman equation rather than by diff usion equation.\nHowever,without magnetic field the electron gas in graphene is margin ally non-ideal. It\ncannot be classical, like an electron gas of a low density with quadratic spectrum. The\nmarginal situation makes theoretical calculations very difficult. Nev ertheless, it is accepted\nthat the Wigner crystal in single layer graphene is absent without ma gnetic field25,26. The\nsign ofdµ/dnis also an interesting question but very difficult for theoretical stud y. Recently\ntunneling microscopy experiment has been done by Martin et al.27. They claim that their\nmeasurement give the thermodynamic density of states and that it is positive. The last\nstatement might be a result of disorder.\n10V. CONCLUSION\nFinallyIarguethatthe negative signofdiffusion coefficient thatfollow s fromthe Einstein\nrelation at negative density of states does not lead to any contrad iction because diffusion\ncoefficient is irrelevant for the unipolar transport under this condit ion. The sign of the\ndiffusion coefficient in this case should not be definitely positive becaus e the diffusion is not\nthe main source of the entropy production. In bipolar situation neg ative diffusion means\nthe collapse of the system and formation of neutral excitons.\nI am grateful to Boris Shklovskii and Yoseph Imry for important dis cussion. I am espe-\ncially indebted to David Khmelnitskii and Emmanuel Rashba for multiple d iscussions and\ncriticism.\n∗Electronic address: efros@physics.utah.edu\n1A. Einstein, Annalen der Physik 17, 549 (1905).\n2M. von Smoluchowsky, Annalen der Physik 21, 756 (1906).\n3L. D. Landau and E. Lifshitz, Electrodynamics of Continuous Media (Butterworth-Heinenann,\n1984), chapter III.\n4B. L. Altshuler, A. G. Aronov, and P. A. Lee, Phys. Rev. Lett 44, 1288 (1980).\n5A. Y. Zyuzin, JETP Lett. 33, 360 (1981).\n6B. L. Altshuler and A. G. Aronov, Electron-Electron Interaction in Disordered Systems (ed. b y\nA. L. Efros and M. Pollak) (North-Holland, Amsterdam, 1985), p. 37.\n7B. I. Shklovskii and A. L. Efros, JETP Lett. 44, 669 (1987).\n8M. S. Bello, E. I. Levin, B. I. Shklovskii, and A. L. Efros, Sov . Phys JETP 53, 822 (1981).\n9S. V. Kravchenko, V. M. Pudalov, and S. G. Semenchinsky, Phys . Lett. A 141, 71 (1989).\n10S. V. Kravchenko, D. A. Rinberg, S. G. Semenchinsky, and V. M. Pudalov, Phys. Rev. B 42,\n3741 (1990).\n11J. P. Eisenstein, L. N. Pfeiffer, and K. West, Phys. Rev. Lett 68, 674 (1992).\n12J. P. Eisenstein, L. N. Pfeiffer, and K. West, Phys. Rev. B 50, 1760 (1994).\n13T. Ando, A. B.Fowler, and F. Stern, Rev. Mod. Phys. 54, 437 (1982).\n14A. L. Efros, Phys. Rev B 45, 11354 (1992).\n1115S. Luryi, Appl. Phys. Lett. 52, 501 (1988).\n16A. L. Efros, F. G. Pikus, and V. G. Burnett, Solid State Comm. 84, 91 (1992).\n17F. G. Pikus and A. L. Efros, Phys. Rev. B 47, 16395 (1993).\n18A. L. Efros, Phys. Rev. B 45, 11354 (1992).\n19T. P. Smith, W. I. Wang, and P. J. Stiles, Phys. Rev. B 34, 2995 (1986).\n20M. I. Dyakonov and A. S. Furman, Sov. Phys. JETP 65, 574 (1987).\n21I am grateful to E. I. Rashba for this comment.\n22K. Seeger, Semiconductor Physics. An Introduction (Springer, 1999), p. 124.\n23H. Zhao, Appl. Phys. Lett. 92, 112104 (2008).\n24M. Rosling, H. Bleichner, P. Jonsson, and E. Nordlander, App l. Phys. Lett. 76, 2855 (1994).\n25H. P. Dahal, T. O. Wehling, K. S. Bedell, J.-X. Zhu, and A. V.Ba latsky, arXiv:cond-\nmat/0706.1689.\n26R. Cote, J.-F. Jobidon, and H. A. Fertig, arXiv:cond-mat/08 06.0573.\n27J. Martin, N. Akrman, G. Ulbricht, T. Lohmann, J. H. Smet, K. v on Klitzing, and A. Yacoby,\nNature Physics 4, 144 (2008).\n12" }, { "title": "0809.0795v1.Impurity_states_in_antiferromagnetic_Iron_Arsenides.pdf", "content": "arXiv:0809.0795v1 [cond-mat.supr-con] 4 Sep 2008Impurity states in antiferromagnetic Iron Arsenides\nQiang Han\nDepartment of Physics, Renmin University, Beijing, China\nZ. D. Wang\nDepartment of Physics and Center of Theoretical and Computa tional Physics,\nThe University of Hong Kong, Pokfulam Road, Hong Kong, China\n(Dated: September 4, 2008)\nWe explore theoretically impurity states in the antiferrom agnetic spin-density wave state of the\niron arsenide. Two types of impurity models are employed: on e has only the intraband scattering\nwhile the other has both the intraband and interband scatter ing with the equal strength. Inter-\nestingly, the impurity bound state is revealed around the im purity site in the energy gap for both\nmodels. However, the impurity state is doubly degenerate wi th respect to spin for the first case;\nwhile the single impurity state is observed in either the spi n-up or spin-down channel for the second\none. The impurity-induced variations of the local density o f states are also examined.\nPACS numbers: 71.55.-i,75.30.Fv,75.10.Lp\nThe recent discovery of iron-based superconductors [1]\nhas triggered intensive efforts to unveil the nature of and\ninterplay between magnetism and superconductivity in\nthis family of materials. Series of iron arsenide have\nbeen synthesized, which possess many similar features\nof the normal and superconducting states. Experimental\nmeasurements have reported that the undoped ReFeAsO\n(where Re= rare-earth metals) and AFe 2As2(where\nA=divalent metals such as Ba, Ca, Sr) compounds ex-\nhibit a long-range antiferromagnetic spin-density-wave\n(SDW) order [2, 3, 4, 5, 6, 7]. Upon electron/hole dop-\ning the SDW phase is suppressed and superconductivity\nemerges with Tcup to above 50 K [8, 9, 10, 11, 12].\nAt present, there is likely certain controversy on\nthe understanding of the SDW state of the undoped\nFeAs-based parent compounds. Two kinds of theories\nhave been put forward: 1) the itinerant antiferromag-\nnetism, which takes advantage of proper Fermi surface\n(FS) nesting (or strong scattering) between different FS\nsheets [13, 14, 15, 16]; and 2) the frustrated Heisen-\nberg exchange model of coupled magnetic moments of\nthe localized d-orbital electrons around the Fe atoms\n[17, 18, 19, 20]. As for the itinerant electronic behav-\nior, first principle band structure calculations [21] based\non the density functional theory (DFT) indicate up to\nfive small Fermi pockets with three hole-like pockets cen-\ntered around the Γ point and two electron-like ones cen-\ntered around the Mpoint of the folded Brillouin zone\nof the FeAs layers, which have partially supported by\nthe angle-resolved photoemission spectroscopy (ARPES)\nfrom different groups [22, 23, 24, 25, 26]. Motivated by\nthe DFT calculation and experimental measurements, in\nRefs.[15, 16], the excitonic mechanism [27] of itinerant\ncarriers are employed taking account of the FS nesting\nbetween electron and hole pockets and the SDW phase\nare associated with triplet excitonic state, which can be\nunderstood as condensate of triplet electron-hole pairs[27].\nIn this paper, we explore theoretically the effect of a\nsingle impurity on the local electronic structure of an Fe-\nbased antiferromagnet in the triplet excitonic phase. It\nis shown that impurity bound states are formed inside\nthe SDW gap, which may be observed experimentally by\nlocal probes. Before introducing the impurity, we first\npropose an effective model Hamiltonian to address the\ntriplet excitonic state,\nˆHMF=/summationdisplay\ni,k,σεΓi(k)d†\nikσdikσ\n+/summationdisplay\nk,σεX(k+X)c†\nk+Xσck+Xσ\n+/summationdisplay\nik,σ,σ′/bracketleftBig\n∆∗\niσσ′(k)d†\nikσck+Xσ′+H.c./bracketrightBig\n,(1)\nwhereΓ= (0,0),X= (π,0). We use the index ito\nlabel different valence bands around Γ point. Around\nXandYpoints, there are two conduction bands. dikσ\nandck+Xσare the annihilation operators of electrons in\nthe ΓiandXbands. Theoretically XandYare two\nequivalent nesting directions. Note that, the structural\nphase transition occurred just above/onthe SDW transi-\ntion breaks this equivalency. Without loss of generality,\nit is assumed that only conduction band around the X\npoint couples with the valence bands around the Γ point,\nwhich is characterized by the mean-field order parame-\nters ∆ iσσ′. For the triplet excitonic phase (SDW), we\nhave real order parameters satisfying ∆ i↑↑=−∆i↓↓and\n∆i↑↓= ∆i↓↑= 0 [27].\nεΓi(k) andεX(k) are used to denote the band disper-\nsionsofthenonmagneticnormalstate. For kinthevicin-\nity of the Γ point (therefore, k+Xin the vicinity of the\nXpoint), the normal-state energy dispersions have ap-2\n-0.03 -0.02 -0.01 0.00 0.01 0.02 \n(a) \nε(k) (eV) \nX\nεX\n0εΓ2 \n0\nεΓ1 \n0\nΓY\nX Γ12\n(b) \nFIG. 1: Schematic plot of (a) the Fermi surfaces; and (b)\nthe band dispersions of the valence (hole) band and conduc-\ntance (electron) bands in the unfolded Brillouin Zone for th e\nundoped parent compound. See text for detail.\nproximately the 2D parabolic forms\nεΓi(k) =−¯h2(k2\nx+k2\ny)\n2mΓi+ǫΓi\n0, (2)\nεX(k+X) =¯h2(k2\nx+k2\ny)\n2mX−ǫX\n0,(3)\nas schematically shown in Fig. 1. Here mΓiandmX\nare the corresponding effective masses. In describing the\nXband, the elliptic FS is approximated by the circu-\nlar one for simplicity. ǫΓi\n0(ǫX\n0) denotes the top (bot-\ntom) of the hole (electron) bands. According to the\nARPES measurement[22], two hole-like Fermi pockets\nare revealed around the Γ point for undoped BaFe 2As2.\nThe band parameters extracted from the experimental\ndata are as follows. mΓ1≈2.8me,mΓ2≈7.4me,\nandmX≈6.5me, wheremeis the mass of bare elec-\ntron.εΓ1\n0≈4 meV,εΓ2\n0≈16 meV, and εX\n0≈24\nmeV. These parameters indicate that the nesting be-\ntween the Γ2 band and Xband is much better than\nthat of the Γ1 band. Therefore it is natural to assume a\nlarger order parameter ∆ 2and a vanishingly small ∆ 1.\nLetEg= (ǫΓ2\n0+ǫX\n0)/2 andµ0= (ǫΓ2\n0−ǫX\n0)/2. Here\nEG=−2Egdenotes the indirect gap between the top of\nthe Γ band and the bottom of the Xbands. Therefore,\nEg>0 describes a semimetal and Eg<0 a semiconduc-\ntor. Withthehelpof Egandµ0andafurtherassumption\nofmΓ2=mX=m≈7me. we can re-express the energy\ndispersions as\nεΓ2(k) =−ε(k)−µ0, (4)\nεX(k+X) =ε(k)−µ0, (5)\nε(k) =¯h2\n2mk2−Eg, (6)\nNotethat for µ0= 0, the holeandelectronbandsareper-\nfectly nested since εΓ2(k) =−εX(k+X) and the system\nis unstable with respect to infinitesimal Coulomb inter-\naction while for nonzero µ0finite strength of Coulomb\nrepulsion is needed.For the reason that the order parameter ∆ 1is set to\nzero, there is no coupling between the Γ1 band and the\nXband. The Hamiltonian of Eq. (1) is reduced to a\nmodel of two bands with one valence band (Γ2 band)\nand one conductionband (Xband). Introducing the two-\ncomponent Nambu operator, ˆψ†\nkσ= (d†\n2kσ,c†\nk+Xσ), the\nmodel Hamiltonian can be simplified as\nˆHMF=/summationdisplay\nkσˆψ†\nkσ/parenleftbigg\nεΓ(k) ∆σ\n∆σεX(k)/parenrightbigg\nˆψkσ+Himp,(7)\nwhere an impurity term has been added with the form,\nˆHimp=/summationdisplay\nk,k′,σˆψ†\nkσˆUk,k′ˆψk′σ, (8)\nwhereˆUk,k′represents a 2 ×2 matrix of the scattering\npotential associated with non-magnetic impurities. Here\nwe use ∆ σto denote ∆ σσfor short. The Green’s function\nmethodisappliedtostudythesingleimpurityeffect. The\nmatrix Greens functions are defined as\nˆGσσ(k,τ;k′,τ′) =−∝an}bracketle{tTτ[ψkσ(τ)ψ†\nk′σ(τ′])∝an}bracketri}ht,(9)\nˆGσσ(k,k′,iωn) =/integraldisplayβ\n0dτˆGσσ(k,τ;k′,0)eiωnτ(10)\nˆGσσ(k,k′,ω) =ˆGσσ(k,k′,iωn→ω+i0+).(11)\nFrom the Hamiltonian defined in Eq. (7) we can derive\nthe bare Green’s function\nˆG0\nσσ(k,ω) =/parenleftbigg\nω−εΓ(k)−∆σ\n−∆σω−εX(k)/parenrightbigg−1\n,(12)\n=˜ωˆτ0+∆σˆτ1−ε(k)ˆτ3\n˜ω2−ε(k)2−∆2σ, (13)\nwhere ˜ω=ω+µ0. ˆτ0is the 2×2unit matrix, and ˆ τ1,3are\nthe pauli matrices. The T-matrix approximation is em-\nployed to compute the Green’s function in the presence\nofimpurities. For a single impurity, the T-matrix exactly\naccountsfor the multiple scatteringoff the impurity. The\nsingle-particle Green’s function ˆGcan be obtained from\nthe following Dyson’s equation,\nˆGσσ(k,k′,ω) =ˆG0\nσσ(k,ω)δk,k′+ˆG0\nσσ(k,ω)\nˆTσσ(k,k′,ω)ˆG0\nσσ(k′,ω),(14)\nwhere the T matrix is given by\nˆTσσ(k,k′,ω) =ˆUk,k′+/summationdisplay\nk′′ˆUk,k′′ˆG0\nσσ(k′′,ω)ˆTσσ(k′′,k′,ω).\n(15)\nFor a point-like scattering potential interacting with itin-\nerant carriers just on the impurity site, the scattering\nmatrix is isotropic, ˆUk,k′′=ˆU. The above equation is\ngreatly simplified\nˆTσσ(ω) =ˆU+ˆUˆG0\nσσ(ω)ˆTσσ(ω), (16)3\nwhereˆG0\nσσ(ω) =/summationtext\nkˆG0\nσσ(k,ω). After some derivation\nwe obtain\nˆG0\nσσ(ω) =−πN0/bracketleftBigg\nα(˜ω)/radicalbig\n∆2σ−˜ω2(˜ωˆτ0+∆σˆτ1)+γ(˜ω)ˆτ3/bracketrightBigg\n,\n(17)\nwhere\nα(˜ω) =π−1/bracketleftBigg\narctan/parenleftBigg\nEc/radicalbig\n∆2σ−˜ω2/parenrightBigg\n+\narctan/parenleftBigg\nEg/radicalbig\n∆2σ−˜ω2/parenrightBigg/bracketrightBigg\n(18)\nγ(˜ω) = (2π)−1ln/parenleftbiggE2\nc+∆2\nσ−˜ω2\nE2g+∆2σ−˜ω2/parenrightbigg\n,(19)\nwithEcdenoting the high-energy cutoff and N0=\nma2/(2π¯h2)the densityofstatesperbandperspin. Note\nthatα(˜ω) andγ(˜ω) are independent of the spin index σ.\nThe first impurity model we study is the scattering-\npotential matrix with only intraband scattering terms,\ni.e.ˆU=Vimpˆτ0, which was adopted in Ref. [28] to study\neffect of many impurities. From Eq. (16), we obtain\nˆTσσ(ω) = [ˆτ0−ˆUˆG0\nσσ(ω)]−1ˆU= [V−1\nimp−ˆG0\nσσ(ω)]−1.(20)\nThe energy of the impurity bound state is determined by\nthe pole of ˆTσσ(ω), determined by det[ V−1\nimp−ˆG0\nσσ(Ω)] =\n0. Setting c≡(πN0Vimp)−1, we have the equation for\nthe energy of impurity bound state\nc2+2cα(˜Ω)˜Ω/radicalBig\n∆2σ−˜Ω2−α(˜Ω)2−γ(˜Ω)2= 0.(21)\nFor the spin triplet excitonic phase ∆ ↑=−∆↓, the\naboveequationgivesrisetoimpuritystateswiththesame\nbound energy, i.e. the impurity states are doubly degen-\nerate. Generally, the above equation has to be solved nu-\nmericallytoobtaintheboundenergy ˜Ω. However,wecan\nget some analytic results under certain approximations.\nUnder the wide-band approximation Ec,Eg≫ |∆σ|,\nα(˜Ω)≈1 andγ(˜Ω)≈γ0=π−1ln(Ec/Eg), so we have\n˜Ω\n|∆σ|= sgn(c)1−c2+γ2\n0/radicalbig\n(1−c2+γ2\n0)2+4c2,(22)\nand furthermore if the system has approximately the\nparticle-hole symmetry Ec≈Eg, thenγ0≈0 and\n˜Ω/|∆σ|= sgn(c)(1−c2)/(1+c2)fromtheaboveequation.\nFor the second impurity model, the four matrix ele-\nments of ˆUis assumed to be the same, i.e. the intra-\nand inter-band scattering terms are the same [29] with\nˆU=Vimp(ˆτ0+ ˆτ1)/2. Then the T-matrix according to\nEq. (16) is\nˆTσσ(ω) = [2V−1\nimp−ˆτ′ˆG0\nσσ(ω)]−1ˆτ′,(23)with ˆτ′= ˆτ0+ ˆτ1. The energy of the impurity bound\nstate is again determined by the pole of ˆTσσ(ω). From\ndet[2V−1\nimp−ˆτ′ˆG0\nσσ(Ω)] = 0 we have the equation for ˜Ω,\nc+˜Ω+∆ σ/radicalBig\n∆2σ−˜Ω2α(˜Ω) = 0. (24)\nFrom the above equation we find that ˜Ω is independent\nof the function of γ(˜ω) for this case, which reflects the\nparticle-hole asymmetry. Before solving the above equa-\ntion for the bound energy, we study the existence of the\nimpurity state. Because ˜Ω2<∆2\nσ,˜Ω+∆ σhas the same\nsign as that of ∆ σ. Therefore, the solution of Eq. (24)\nexists only if the sign of cis opposite to that of ∆ σ. For\nthe SDW state, i.e. the triplet excitonic phase, we have\n∆↑=−∆↓and so there is exactly one impurity bound\nstate in either the spin-up or spin-down channel. We\nmay assume ∆ ↑=−∆↓= ∆>0 as well, then for attrac-\ntive scattering Vimp<0, the impurity bound state only\nexists in the spin-up channel and its energy is given by\n˜Ω/∆ =−(1−c2)/(1+c2) under the wide-band approxi-\nmation. IfVimp>0, however, the impurity state will be\nin the spin-down channel, and ˜Ω/∆ = (1−c2)/(1+c2).\nIn general, the impurity bound-state energy is given by\n˜Ω\n∆σ= sgn(c)1−c2\n1+c2(25)\nin the valid regime of the wide-band approximation.\nTo apply the theoretical results to the iron arsenide,\nwe try to pin down the parameters of our model by ex-\ntracting them from the available experimental data for\nBaFe2As2[22].ǫΓ\n0≈16 meV and ǫX\n0≈24 meV so\nthatEg≈20 meV.mΓ≈mX≈7.0meand there-\nforeN0≈1.2 eV−1. ∆↑=−∆↓= ∆≈20 meV. The\nhigh-energycutoff is set as Ec= 500meV, which is ofthe\nsame order of magnitude as the band width. Note that\nEgextracted fromexperimental data is verysmall, which\nis in the same order of magnitude of the order parameter\n∆. Therefore, neither the wide-band approximation nor\nthe particle-hole symmetry can be applied to the present\ncase. Eqs. (21) and (24) have to be numerically solved.\nNow we examine the local characteristics induced by\nthe impurity by looking into the variation of the local\ndensity of states (LDOS), which can be probed by the\nscanning tunneling microscopy (STM). The LDOS is de-\nfined as\nN(r,ω) =−1\nπ/summationdisplay\nσIm{Tr[ˆGσσ(r,r′,ω)]},(26)\nwhereˆGσσ(r,r′,ω) the Green’s function in real space.\nApplying the T-matrix approximation we have,\nˆGσσ(r,r′,ω) =ˆG0\nσσ(r,r′,ω)+ˆG0\nσσ(r,0,ω)\nˆT(ω)ˆG0\nσσ(0,r′,ω), (27)4\nSubstituting Eq. (27) into Eq. (26) we may single out the\nvariation of LDOS due to the presence of the impurity\npotential,\nNimp(r,ω) =−1\nπ/summationdisplay\nσIm{Tr[ˆG0\nσσ(r,0,ω)ˆTσσ(ω)\nˆG0\nσσ(0,r,ω)]}. (28)\nFor the second impurity model[30], Fig. 2(a) shows the\nLDOS as a function of energy ˜ ωon the impurity site,\nnamelyN(0,˜ω), while Fig. 2(b) the impurity-induced\nLDOS at the bound energy as a function of radial dis-\ntanceroff the impurity site, i.e. Nimp(r,˜Ω).Vimphas\nbeen set as -0.36, -0.6, and -4.0 eV, giving rise to the\nimpurity bound states seen as the sharp peaks located\nrespectively at the energies ˜Ω/∆ = 0,−0.5, and−0.99 in\nFig. 2(a). The probability densities of these bound states\nexhibit a kind of exponential decay with the Friedel os-\ncillation, as seen in Fig. 2(b). Introducing two length\nscales,ξ1andξ2to characterize the oscillation and de-\ncay, we obtain the asymptotic behavior of Nimp(r,˜Ω) for\nlarger,\nNimp(r,˜Ω)∝r−1cos2(πr/ξ1)exp(−r/ξ2).(29)\nξ1/a=π//radicalbig\n4πN0Egwhich is approximately 5 .7 in con-\nsistence with the numerical results shown in Fig.2(b).\nξ2/ξ1=Eg/(π/radicalbig\n∆2−˜Ω2) = 0.32, 0.37, and 2.3 for the\nthree cases of impurity states. This explains why we see\nclear Friedel oscillation for impurity state with bound\nenergy near the gap edge.\n-15 -10 -5 0 5 10 15 20 010 20 30 40 50 60 \n-3 -2 -1 0 1 2 3010 20 30 40 50 60 \nNimp (r, Ω) \nr/a (b) N(0, ϖ)\nϖ/Δ(a) \nFIG. 2: LDOS on the impurity site as a function of energy ˜ ω,\ni.e.N(0,˜ω) (a), and impurity-induced LDOS at the bound\nenergy˜Ω as a function of radial distance roff the impurity\nsite, i.e.N(r,˜Ω) (b).ris in unit of awithathe lattice\nconstant of Fe-Fe plane. Black, red, and blue lines correspo nd\ntoVimp=−0.36,−0.6,−4 eV, respectively.\nThis work was supported by the NSFC grand under\nGrants Nos. 10674179 and 10429401, the GRF grant of\nHong Kong.\n[1] Y. Kamihara et al., J. Am. Chem. Sco. 130, 3296 (2008).[2] C. de la Cruz et al., Nature 453, 899 (2008); J. Zhao et\nal., arXiv:0806.2528; J. Zhao et al., arXiv:0807.1077.\n[3] Y. Chen et al., Phys. Rev. B 78, 064515 (2008), see also\narXiv.org:0806.0662.\n[4] Q. Huang et al., arXiv:0806.2776.\n[5] M. A. McGuire et al., arXiv.org:0806.3878.\n[6] A. A. Aczel et al., arXiv:0807.1044.\n[7] A. I. Goldman et al., arXiv.org:0807.1525.\n[8] H. Takahashi et al., Nature 453, 376 (2008).\n[9] Z. -A. Ren et al., Europhys. Lett. 82, 57002 (2008).\n[10] H. .H. Wen et al., Europhys. Lett. 82, 17009 (2008), see\nalso arXiv:0803.3021.\n[11] X. H. Chen et al., Nature 453, 761 (2008), see also\narXiv:0803.3603.\n[12] G. F. Chen et al., Phys. Rev. Lett. 100, 247002 (2008),\nsee also arXiv:0803.3790.\n[13] J. Dong et al., Europhys. Lett. , 83, 27006 (2008); see\nalso arXiv:0803.3426.\n[14] I. I. Mazin et al., Phys. Rev. Lett. 101, 057003 (2008),\nsee also arXiv:0803.2740; arXiv.org:0806.1869.\n[15] Q. Han, Y. Chen, and Z. .D .Wang, Europhys. Lett. 82\n37007 (2008), see also arXiv:0803.4346.\n[16] V. Barzykin and L. P. Gorkov, arXiv:0806.1933.\n[17] T. Yildirim, Phys. Rev. Lett. 101, 057010 (2008), see\nalso arXiv:0804.2252.\n[18] Q. Si and E. Abrahams, arXiv:0804.2480.\n[19] F. J. Ma, Z. Y. Li, and T. Xiang, arXiv.org:0804.3370.\n[20] C. Fang et al., Phys. Rev. B 77, 224509 (2008), see also\narXiv:0804.3843.\n[21] D. J. Singh and M. H. Du, Phys. Rev. Lett. 100\n237003 (2008), see also arXiv:0803.0429; G. Xu et\nal., Europhys. Lett., 82, 67002 (2008), see also\narXiv:0803.1282; K. Haule, J. H. Shim and G. Kotliar,\nPhys. Rev. Lett. 100226402 (2008), see also\narXiv:0803.1279; F. Ma, and Z. -Y. Lu, Phys. Rev. B 78\n033111 (2008), see also arXiv:0803.3286; I. A. Nekrasov,\nZ. V. Pchelkina, M. V. Sadovskii, JETP Lett., 88, 144\n(2008), see also arXiv:0806.2630; F. Ma, Z.-Y. Lu, and\nT. Xiang, arXiv:0806.3526; D. J. Singh, arXiv:0807.2643.\n[22] L. X. Yang et al., arXiv:0806.2627.\n[23] C. Liu et al., arXiv:0806.3453.\n[24] L. Zhao et al., arXiv:0807.0398.\n[25] H. Ding et al., Europhys. Lett. 83, 47001 (2008), see also\narXiv:0807.0419; P. Richard et al., arXiv:0808.1809.\n[26] D. H. Lu et al., arXiv:0807.2009.\n[27] B. I. Halperin and T. M. Rice, Solid State Phys. 21, 125\n(1968).\n[28] J. Zittartz, Phys. Rev. 164, 575 (1967).\n[29] The impurity Hamiltonian is expressed as Himp=R\nUimp(r)φ†(r)φ(r)dr=P\nk,k′ˆψ†(k)ˆUk,k′ˆψ(k) where\nˆψ†(k) = (d†\nk,c†\nk+X)φ(r) = 2−1/2P\nkϕΓk(r)dk+\nϕXk(r)ck+X, withϕΓk(r) andϕXk(r) Bloch functions\nof quasimomentum k. In this paper the scattering po-\ntential is assumed to be pointlike, Uimp=Vimpδ(r).\nWith further simplification of the Bloch functions of the\nform exp(ik·r),ˆUk,k′iskindependent and equal to\nVimp(ˆτ0+ ˆτ1)/2.\n[30] The following conclusions are also qualitatively corr ect\nfor the first impurity model." }, { "title": "0809.0976v1.Comparison_of_Raman_spectra_and_vibrational_density_of_states_between_graphene_nanoribbons_with_different_edges.pdf", "content": "arXiv:0809.0976v1 [cond-mat.mtrl-sci] 5 Sep 2008EPJ manuscript No.\n(will be inserted by the editor)\nComparison of Raman spectra and vibrational density of stat es\nbetween graphene nanoribbons with different edges\nSami Malola1,a, Hannu H¨ akkinen1,2, and Pekka Koskinen1\n1NanoScience Center, Department of Physics, FIN-40014 Univ ersity of Jyv¨ askyl¨ a, Finland\n2NanoScience Center, Department of Chemistry FIN-40014 Uni versity of Jyv¨ askyl¨ a, Finland\nReceived: date / Revised version: date\nAbstract. Vibrational properties of graphene nanoribbons are examin ed with density functional based\ntight-binding method and non-resonant bond polarization t heory. We show that the recently discovered\nreconstructed zigzag edge can be identified from the emergen ce of high-energy vibrational mode due to\nstrong triple bonds at the edges. This mode is visible also in the Raman spectrum. Total vibrational\ndensity of states of the reconstructed zigzag edge is observ ed to resemble the vibrational density of states\nof armchair, rather than zigzag, graphene nanoribbon. Edge -related vibrational states increase in energy\nwhich corroborates increased ridigity of the reconstructe d zigzag edge.\nPACS.61.46.-w Structure of nanoscale materials – 64.70.Nd Struc tural transitions in nanoscale materials\n– 63.22-m Phonons or vibrational states in low-dimensional structures and nanoscale materials – 78.30Na\nInfrared and Raman spectra, fullerenes and related materia ls\nInpastdecadescarbonnanomaterialshaveshowntheir\nrich properties in several applications[1,2]. Here graphene\nnanoribbons are not an exception, and their use in many\napplications, among which transistors in nanoelectronics,\nhavebeeninvestigatedconsideringdifferentedgestructures[3,\n4]. Low dimensionality and edges make ribbons particu-\nlarly fascinating both for theory and applications.\nThepreciseedgestructureingraphenenanoribbonsaf-\nfectsmanypropertieslikechemicalreactivity[5],electronic\nstructure[6] and vibrations[7]. Vibrational properties play\na role in structural stability [8], structure identification[9]\nandballistictransportthroughelectron-phononcoupling[9].\nInstructureidentificationscanningtunnelingmicroscopy[10,\n11] can reach near atom resolution but analysis of the full\nedge structure and properties is often ambiguous. In this\ncase Raman spectroscopy[12,13,9] is a valuable tool.\nVibrational properties of graphene nanoribbons have\nbeen recently studied for acoustic and optical phonons,\nsymmetries, and Raman activity as a function of ribbon\nwidth[12,8]. In this paper we consider recently reported\nself-passivating edge reconstruction of zigzag ribbon [14],\nandinvestigatehowthereconstructionaffectsRamanspec-\ntra and vibrational density of states (VDOS). Edge-loca-\nlized contributions to Raman spectra and vibrationalden-\nsity of states change. It turns out that high-energy modes\ndue to triple-bond vibrations, as well as overall changes in\nthe vibrational density of states and changes in rigidity of\nthe edges, make the reconstruction identifiable and visible\nin Raman spectra.\nasami.malola@phys.jyu.fi\nzigzag57\nW=20.12Å\nL=24.60Åzigzag\nW=19.88Å\nL=24.60Å\narmchair\nW=23.37Å\nL=21.30ÅL\nW\ntotal 200 atoms\nedge 80 atomstotal 200 atoms\nedge 80 atomstotal 200 atoms\nedge 80 atoms\nzx\ny\nFig. 1. Examined structures of zigzag, reconstructed zigzag\n(zigzag57) and armchair graphene nanoribbons. Selected ed ge\natoms are drawn in gray color. Coordinate axis is representa ted\nat lower right corner, where the ribbon direction (=periodi c\ndirection) is z-direction. W=width in non-periodic and L=\nlength in periodic direction.\nWe use density functional based tight-binding method\nto calculate forces[15], to optimize structure[16] and to2 S. Malola et al.: Comparison of Raman spectra and VDOS of gra phene nanoribbons with different edges\n500100015002000 500100015002000 500100015002000Intensity [arb. units]\nWavenumber [cm¯¹]armchair zigzag57 zigzagzz xz xx\n(1)\n(2)\n(L)\n(L)(L)\nFig. 2.Raman spectra of zigzag, reconstructed zigzag (zigzag57) a nd armchair graphene nanoribbons (different rows). Columns\ndenote different incident and scattered light polarization spectra (uppermost symbols xx, xz and zz). Edge-localized m odes are\nassigned with symbol (L). Symbols (1) and (2) are refered to i n the text. Raman spectra are drawn using Lorenz distributio n\nwith full width at half maximum of 5 cm−1.\ncalculate the vibrational eigenmodes for the systems. For\nRaman spectra calculationswe use non-resonantbond po-\nlarizationtheory[17,18], that has been used succesfully for\nribbons[12] and for non-identical carbon nanotubes[19,20,\n21]. The slight consistent energy shift in the high-energy\nmodes[19] does not affect the qualitative changes in Ra-\nman spectra or VDOS we want point out in this paper.\nFigure 1 shows the atom structures of zigzag, recon-\nstructedzigzag(zigzag57)andarmchairgraphenenanorib-\nbons. Widths and numbers of atoms of the ribbons are\nchosen to be easily comparable; also other sizes were sys-\ntematically studied and all qualitative results given below\nwere supported by these systematics. Edge atoms, as we\ndefine them, are drawn in gray and are used to deter-\nmine edge-weighted vibrational density of states, where\nthe edge-weight is ratio\nwµ\nedge=/summationtext\ni∈edgevµ\ni·vµ\ni/summationtext\nj∈allvµ\nj·vµ\nj(1)\nwhere(3-dimensionalvector) vµ\niisthecomponentonatom\nifor eigenmode µ. Hencewµ\nedge= 1 means complete lo-\ncalization to edge, wµ\nedge=80\n200means ”uniformly dis-\ntributed” mode, and wµ\nedge= 0 would mean a mode lo-calized in the central region. Contributions of different\ndirections are also analysed, analogously including weight\nof the components of the vibrational eigenvectors vµ\niin\neitherx-,y- orz-direction,\nwµ\nx=/summationtext\nivµ\ni,x·vµ\ni,x/summationtext\nj∈allvµ\nj·vµ\nj(2)\nand similarly for y and z. Hence wµ\nx+wµ\ny+wµ\nz= 1 and\nwµ\nx,edge+wµ\ny,edge+wµ\nz,edge=wµ\nedge;wµ\nxmeasureshowmuch\nofthe modeis in the x-directionand wµ\nx,edgemeasureshow\nmuch of the mode at the edge is in the x-direction.\nFigure2showsthexx,xzandzzpolarizedRamanspec-\ntra of the ribbons, where e.g. xz stands for incident light\npolarization in x- and scattered light polarization in z-\ndirection. The polarization pictures xx and zz show the\nsame modes, but with different intensities. Polarization\npictures that contain y-direction and out-of-plane modes\nyield tiny intensities and have been left out from figure 2.\nOrientation of the ribbons with respect to coordinate axis\nis shown in figure 1. The most pronounced modes in figure\n2 include breathing modes (intensive low-energy peaks\nin xx/zz spectra), G-band-related high-energy modes (in-\ntensive high-energy peaks in the xz and xx/zz spectra)S. Malola et al.: Comparison of Raman spectra and VDOS of grap hene nanoribbons with different edges 3\n00.500.500.5\n 0 500 1000 1500 2000 2500VDOS [1/cm-1]\nWavenumber [cm-1]zigzag\nzigzag57\narmchairy\nx\nz\ntotal\nFig. 3. Total vibrational density of states of zigzag, recon-\nstructed zigzag (zigzag57) and armchair graphene nanorib-\nbons. Contribution of different directions in vibrational s tates\nare shown in different patterned areas. Vibrational density\nof states is broadened with normal distribution with σ=\n45 cm−1, chosen to illuminate general changes in VDOS under\nreconstruction.\n00.500.500.5\n 0 500 1000 1500 2000 2500VDOS [1/cm-1]\nWavenumber [cm-1]zigzag\nzigzag57\narmchairy\nx\nz\ntotal\nFig. 4.As figure 3, but showing the edge weighted vibrational\ndensity of states of zigzag, reconstructed zigzag (zigzag5 7) and\narmchair graphene nanoribbons.and edge-localized modes (symbols (L) in xz and zz spec-\ntra). One of the main points is that only edge-localized\nmodes undergo observable changes under reconstruction,\nothersbeingwithin 2cm−1fromcorrespondingzigzagrib-\nbon modes. These changes in edge-localized Raman active\nmodes make the reconstruction visible. For zigzag57 the\nedge-localized mode of zigzag edge at 1550 cm−1disap-\npears in xz polarization spectrum in figure 2 and edge-\nlocalized triple-bond vibration becomes visible at energies\naround 2250 cm−1in zz polarization spectrum of zigzag57\nribbon, like in armchair ribbon. Intensity of the triple-\nbond vibration in reconstructed zigzag ribbon is smaller\nin our bond polarization theory than in armchair ribbon,\nmainly because of the differences in bond angles on the\nedge (the edge profile of zigzag57 is more linear).\nCouple of new Raman active modes can be noticed to\nappear for reconstructed zigzag edge, assinged with sym-\nbols (1) in xx and (2) in xz polarization spectrum in fi-\ngure 2. Mode (1) at 1500 cm−1includes stretching of edge\nbonds of pentagons. Mode (2) at 1600 cm−1is mainly a\nlongitudinal mode where first normal zigzag row of atoms\nin the zigzag57ribbon vibrates as the edge-localized mode\nof zigzag ribbon. Neither of the modes (1) nor (2) is fully\nlocalized on the edge.\nBreathing modes remain unchanged, because they are\ncollective vibrations which hold the edges internally un-\nchanged during the vibration. Similarly, because G-band-\nrelated high-energy modes vanish towards the edge, re-\nconstruction has only minor effect on them. Raman active\nout-of-plane vibrations have much lower intensity than in-\nplane vibrations and are not therefore considered here.\nNext we compare more generally the vibrational pro-\nperties of the different ribbon edges. Figure 3 shows the\ntotal vibrationaldensity of states, analysedwith contribu-\ntions from different directions. The surprising and signi-\nficant result is that zigzag57 ribbon’s vibrational density\nof states resembles armchair ribbon’s rather than zigzag\nribbon’s vibrational density of states. This means that for\nvibrational modes it is not the orientation of the under-\nlying honeycomb lattice that is important, but only the\nlocal structures of the edges . Comparison of changes bet-\nween total vibrational density of states in figure 3 and\nedge-weighted vibrational density of states in figure 4 re-\nveals that the most significant changes are seen in edge-\nlocalized modes. Triple-bond vibrations in z-direction on\nthe high energies appear as a separate peak around 2250\ncm−1in figures 3 and 4 for both reconstructed zigzag rib-\nbon and armchair ribbon. The triple-bond edge modes\nin z-direction in the zigzag57 ribbon, somewhat deplete\nmodes from the energy range 1700 −2000 cm−1of the\nG-band-related modes in z-direction. At the same time\nin the same energy range the number of edge-related in-\nplane vibrations in x-direction are increased due to recon-\nstruction - induced edge stiffening. Similar effects are seen\neven in lower energies. Some edge-localized out-of-plane\nmodes around 500 −650 cm−1are upshifted in energy\nfor zigzag57 ribbon for the same reason. Even the energy\nof the lowest energy out-of-plane modes increase, which4 S. Malola et al.: Comparison of Raman spectra and VDOS of gra phene nanoribbons with different edges\ncan dramatically alter low-temperature properties (such\nas heat capacity) of the ribbons.\nHence, due to edge stiffening, edge vibrations gene-\nrally up-shift in energy under reconstruction. This type\nof changes in mechanical properties can have an effect on\napplications such as sensors. The similarity of armchair\nand zigzag57 ribbon’s VDOS is general observation, as\nthe resemblance and also other qualitative features of the\nvibrational properties remained also for wider ribbons (up\ntoW= 40˚A).\nTo conclude, we have analysed the Raman spectra and\nvibrational density of states of the recently reported self-\npassivating edge reconstruction of zigzag ribbons. In the\nRaman spectra edge-localized triple-bond vibrations be-\ncome visible like the corresponding mode in armchair rib-\nbon’s spectra. These high-energy vibrations can be used\ntoidentify the reconstructionbyconcentratingthe Raman\nmeasurement on the edge.\nThepredictedintensitydifferencesbetweentriple-bond\nvibrations of the zigzag57ribbon and the armchair ribbon\ncould be used to separate the reconstructed edge struc-\nture from the armchair edge structure with help of scan-\nningtunneling microscopyimages.Scanningtunnelingmi-\ncroscopycanalsohelpinavoidingidentificationdifficulties\nwith ribbons that have combination of zigzag and arm-\nchair edges. For those ribbons the density of triple-bonds\nat the edges is lower and further intensity of the triple-\nbond vibrations is lower like in reconstructed zigzag rib-\nbon, which can lead to mix up between these structures.\nThe rest of the visible Raman active modes of the rib-\nbons do not change for zigzag57 and the breathing mode\nhas the same ribbon width dependence. Of course, in very\nnarrow ribbons with majority of atoms on the edge, ener-\ngies or spatial nature of the G-band-related high-energy\nmodes may still change.\nThis research is supported by the Academy of Finland through\nthe FINNANO consortium MEP (molecular electronics and\nnanoscale photonics) and project 121701. S. Malola acknow-\nledges a grant from the Finnish Cultural Foundation.\nReferences\n1. R.H. Baughman, A.A. Zakhidov, W.A. de Heer, Science\n297, 787 (2002)\n2. P. Avouris, Z. Chen, V. Perebeinos, Nature Nanotechnol-\nogy2, 605 (2007)\n3. Y. Ouyang, Y. Yoon, J. Fodor, J. Guo, Appl. Phys. Lett.\n89, 203107 (2006)\n4. Y. Yoon, J. Guo, Appl. Phys. Lett. 91, 073103 (2007)\n5. D. Jiang, B. Sumpter, S. Dai, The Journal of Chem. Phys.\n126, 134701 (2007)\n6. K. Nakada, M. Fujita, G. Dresselhaus, M. Dresselhaus,\nPhys. Rev. B 54, 17954 (1996)\n7. M. Yamada, Y. Yamakita, K. Ohno, Phys. Rev. B 77,\n054302 (2008)\n8. N. Dugan, S. Erkoc, Phys. Stat. Sol. b 245, 695 (2008)\n9. A. Ferrari, Solid State Comm. 143, 47 (2007)\n10. L. Tapaszto, G. Dobrik, P. Lambin, L. Biro, Nature Nan-\notechnology 149, 1 (2008)11. Y. Kobayashi, K. Fukui, T. Enoki, K. Kusakabe, Phys.\nRev. B73, 125415 (2006)\n12. J. Zhou, J. Dong, Appl. Phys. Lett. 91, 173108 (2007)\n13. L. Cancado, M. Pimenta, B. Neves, G. Medeiros-Ribeiro,\nT. Enoki, Y. Kobayashi, K. Takai, K. Fukui, M. Dressel-\nhaus, R. Saito et al., Phys. Rev. Lett. 93, 047403 (2004)\n14. P.Koskinen, S.Malola, H.H¨ akkinen, Phys. Rev. Lett.\n(in print), Preprint: arXiv:0802.2623 [cond-mat.mtrl-sc i]\n(2008)\n15. T. Frauenheim, G. Seifert, M. Elstner, T. Niehaus,\nC. K¨ ohler, M. Amkteutz, M. Sternbergand, Z. Hajnal,\nA.D. Carlo, S. Suhai, J. Phys.: Condens. Matter 14, 3015\n(2002)\n16. E. Bitzek, P. Koskinen, F. G¨ ahler, M. Moseler, P. Gumbsh ,\nPhys. Rev. Lett. 97, 170201 (2006)\n17. S. Guha, J. Menendez, J. Page, G. Adams, Phys. Rev. B\n53, 13106 (1996)\n18. R. Saito, T. Takeya, T. Kimura, G. Dresselhaus, M. Dres-\nselhaus, Phys. Rev. B 57, 4145 (1998)\n19. S. Malola, H. H¨ akkinen, P. Koskinen, Phys. Rev. B 77,\n155412 (2008)\n20. G. Wu, J. Zhou, J. Dong, Phys. Rev. B 72, 115411 (2005)\n21. S.Malola, H.H¨ akkinen, P.Koskinen, (submitted), Prep rint:\narXiv:0809.0769 [cond-mat.mtrl-sci] (2008)" }, { "title": "0811.3623v3.Superfluid_density_of_the_ultra_cold_Fermi_gas_in_optical_lattices.pdf", "content": "arXiv:0811.3623v3 [cond-mat.other] 16 Jul 2009Superfluid-density of the ultra-cold Fermi gas in\noptical lattices\nT. Paananen1,2\n1Department of Physics, P.O. Box 64, FI-00014 University of Helsink i, Finland\n2Laboratory of Physics, Helsinki University of Technology, P.O. Box 1100, 02015\nHUT, Finland\nE-mail:tomi.paananen@helsinki.fi\nAbstract. In this paper we study the superfluid density of the two component Fermi\ngas in optical lattices with population imbalance. Three different type of phases, the\nBCS-state (Bardeen, Cooper, and Schrieffer), the FFLO-state (Fulde, Ferrel, Larkin,\nand Ovchinnikov), and the Sarma state, are considered. We show t hat the FFLO\nsuperfluid density differs from the BCS/Sarma superfluid density in a n important way.\nAlthough there are dynamical instabilities in the FFLO phase, when th e interaction is\nstrong or densities are high, on the weak coupling limit the FFLO phase is found to\nbe stable.\nSubmitted to: J. Phys. B: At. Mol. Phys.Superfluid-density of the ultra-cold Fermi gas in optical la ttices 2\n1. Introduction\nThe field of ultra-cold Fermi gases offers a great tool to study man y different problems\nof correlated quantum systems. For example, in recent experimen ts [1, 2, 3, 4, 5, 6]\npolarized Fermi gases were considered. These systems make it pos sible to study physics\nin the presence of mismatched Fermi surfaces, and non-BCS type pairing such as that\nappearing in FFLO-states [7, 8] or Sarma-states [9, 10]. These pos sibilities have been\nconsidered extensively in condensed-matter, nuclear, and high-e nergy physics [11].\nIt is also possible to study many different physical problems with close analogs in\nthe field of solid state physics using optical lattices. However, unlike solid state systems,\nultra-coldgases inoptical lattices provide a very clean environment . These systems have\nvery few imperfections and if one is interested in imperfections, the y can be imposed\neasily on the system. Optical lattices are made with lasers, thus the lattice geometry is\neasy to modify [12, 13, 14, 15] by changing theproperties ofthe int ersecting laser beams.\nFor these reasons, in optical lattices one can study various quant um many-body physics\nproblems, such as Mott insulators, phase coherence, and superfl uidity. The possibility\nof a superfluid alkali atom Fermi gas in an optical lattice has been rec ently studied both\ntheoretically [16, 17, 18, 19, 20, 21, 22], as well as experimentally [23 ].\nThe pairing predicted by the BCS theory does not always mean that t he gas is\na superfluid [24]. When a linear phase is imposed on the order paramete r, the phase\ngradient corresponds to superfluid velocity, and a part of the gas , which has superfluid\nvelocity is called superfluid, and the coefficient of the inertia of this mo ving part is\ncalled superfluid density [25, 26, 27]. However, if the gas is normal th e order parameter\nvanishes and thus the phase shift does not affect to the normal ga s. If the superfluid\ndensity is positive in all directions the gas is a superfluid. Negative sup erfluid density\nimplies dynamical instability of the gas [28, 29].\nThere are a fair number of studies about superfluid density of Ferm i gas in the free\nspace as well as in a trap [30, 31, 32, 33], and few studies about supe rfluid density in\noptical lattices [34, 29]. However, these papers have focused on p opulation balanced\ncases i.e. the densities of the components are same. In this paper w e study the\nsuperfluid density of an ultra-cold two component Fermi gas in optic al lattices at finite\ntemperatures and with finite polarization. We are motivated by the f act by calculating\nthe superfluid density we can draw more conclusions on whether the gas is actually\nsuperfluid or not. Furthermore, we investigate if the superfluid de nsity can be used\nas an indicator of different superfluid phases. The phases we consid er are the BCS\nphase,i.e. the densities of the component are equal, the one mode FF LO phase, i.e.\nthe densities are not equal and the phase of the pairing gap modulat es as a function\nof position, and Sarma phase i.e. the densities are not equal and the phase of the\npairing gap is constant. We show that there are qualitatively differen ce between the\nBCS/Sarma superfluid density and the one mode FFLO superfluid den sity. We also\nstudy the stability of different phases. We also demonstrate that t here can be dynamical\ninstabilities in the one mode FFLO phase.Superfluid-density of the ultra-cold Fermi gas in optical la ttices 3\nThis paper is organized as follows. In Sec. 2 we discuss the physical s ystem and\npresent the Hamiltonian of the system. In Subsec. 2.1 the mean-fie ld approximation and\nthe ansatzes we consider are presented. In Sec. 3 we determine t he superfluid density\ntensor. In Sec. 4 we present the numerical results and we end with some concluding\nremarks in Sec. 5.\n2. Lowest band fermions in an optical lattice\nWe consider a two component Fermi gas, whose components are tw o different hyperfine\nstates of the same isotope, and we call them ↑-state and ↓-state. The Hamiltonian of\nthe system is given by\nˆH=/summationdisplay\nσ=↑,↓/integraldisplay\ndrˆΨ†\nσ(r)/parenleftbigg\n−/planckover2pi12∇2\n2m+Vσ(r)−µσ/parenrightbigg\nˆΨσ(r) (1)\n+g/integraldisplay /integraldisplay\ndrdr′ˆΨ†\n1(r)ˆΨ†\n2(r′)δ(r−r′)ˆΨ2(r′)ˆΨ1(r)\nwhere/planckover2pi1=h/2π,his Planck’s constant, Ψ†\nσ(r) and Ψ σ(r) are the fermionic creation-\nand annihilation field operators of the component σ,µσis the chemical potential of the\ncomponent σ, andδ(r−r′) is Dirac’s δ-function. The interaction strength is related to\nthe s-wave scattering length asthrough\ng=4π/planckover2pi12as\nm.\nThe lattice potential has a cubic structure and is given by\nVσ(r) =Er/summationdisplay\nαsσ,αsin2(kxα),\nwhereEr=/planckover2pi12k2/2mis the recoil energy ( k= (π/d) wheredis a lattice constant), and\nsσ,αis the lattice depth in the α-direction for the component σ.\nWhen we assume that only the lowest band is occupied the field operat ors can be\nexpanded by using the localized Wannier functions in the following way\nˆΨσ(r) =/summationdisplay\niwσ,i(r)ˆcσ,i (2)\nˆΨ†\nσ(r) =/summationdisplay\niw∗\nσ,i(r)ˆc†\nσ,i, (3)\nwherewσ,i(r) is a Wannier function, which is localized at a lattice point i, and ˆcσ,iis an\nannihilation operator.\nThe lowest band Hubbard model is valid, when the lattice is deep enoug h. In other\nwords, it is valid, when the Wannier functions decay within a single lattic e constant,\nand when the effective interaction between atoms is much smaller tha n the bandgap\nbetween bands. These two conditions imply that one has to take into account only the\non-site interactions. In this case overlap integrals of the kinetic en ergy operator between\nthe next nearest neighbor are small compared to overlap integrals of the kinetic energy\noperator between the nearest neighbor [35], and consequently, one needs to take intoSuperfluid-density of the ultra-cold Fermi gas in optical la ttices 4\naccount only hopping between the nearest neighbours. Then the lo west band Hubbard\nHamiltonian is given by\nˆH=−/summationdisplay\nσ=↑,↓\nJσ,x/summationdisplay\n/angbracketlefti,j/angbracketrightx+Jσ,y/summationdisplay\n/angbracketlefti,j/angbracketrighty+Jσ,z/summationdisplay\n/angbracketlefti,j/angbracketrightz\nˆc†\nσ,iˆcσ,j (4)\n−/summationdisplay\ni(µ↑c†\n↑,iˆc↑,i+µ↓c†\n↓,iˆc↓,i)+U/summationdisplay\niˆc†\n↑,iˆc†\n↓,iˆc↓,iˆc↑,i,\nwhere/angb∇acketlefti,j/angb∇acket∇ightαmeans sum over the nearest neighbours in the α-direction, and hopping\nstrength is defined by\nJσ,α=−/integraldisplay\ndrw∗\nσ,i(r)/parenleftbigg\n−/planckover2pi12∇2\n2m+Vσ(r)/parenrightbigg\nwσ,i±dˆxα(r).\nThe coupling strength of the above Hubbard model is given by\nU=g/integraldisplay\ndr|w↑,i(r)|2|w↓,i(r)|2.\n2.1. Mean-field approximation\nBecause the interaction term U/summationtext\niˆc†\n↑,iˆc†\n↓,iˆc↓,iˆc↑,iis hard to handle, we can approximate\nit by using mean-field approximation. Under this approximation the int eraction part\nbecomes\n/summationdisplay\ni∆(i)ˆc†\n↑,iˆc†\n↓,i+∆∗(i)ˆc↓,iˆc↑,i−|∆(i)|2\nU,\nwhere the pairing gap ∆( i) =U/angb∇acketleftˆc↓,iˆc↑,i/angb∇acket∇ight. With respect to position dependence, we only\nconsider the case, in which only the phase of the gap ∆( i) =|∆|eiq·Ri(Riis a lattice\nvector) candependonthelatticesite i. ThisansatziscalledtheonemodeFFLO(Fulde,\nFerrell, Larkin, and Ovchinnikov) [7, 8]. This ansatz includes also the B CS-ansatz and\nthe Sarma-state ansatz as special cases. In these phases the m omentum qis simply\nzero. Under this mean.field approximation the Hamiltonian is given by\nˆH0=−/summationdisplay\nσ=↑,↓\nJσ,x/summationdisplay\n/angbracketlefti,j/angbracketrightx+Jσ,y/summationdisplay\n/angbracketlefti,j/angbracketrighty+Jσ,z/summationdisplay\n/angbracketlefti,j/angbracketrightz\nˆc†\nσ,iˆcσ,j (5)\n−/summationdisplay\ni(µ↑c†\n↑,iˆc↑,i+µ↓c†\n↓,iˆc↓,i)+/summationdisplay\ni|∆|eiq·Riˆc†\n↑,iˆc†\n↓,i+|∆|e−iq·Riˆc↓,iˆc↑,i.\nBy minimizing the free energy F0= Ω0+µ↑N↑+µ↓N↓(Ω0is the grand canonical\npotential of the mean-field Hamiltonian) with respect |∆|andq, and solving the\nnumber equations ∂F0/∂µσ= 0 simultaneously, one finds the pairing gap, the chemical\npotentials, and the momentum qas functions of the temperature and the particle\nnumbers Nσ.Superfluid-density of the ultra-cold Fermi gas in optical la ttices 5\n3. Superfluid density\nLandau’s two component model for a superfluid gas says that the s uperfluid gas consists\ntwo components the normal component and the superfluid compon ent. In the free space\nthe superfluid density is density of the superfluid component. The c ase is a little bit\ndifferent inlatticesthustheenergydifferencebetween thetwisted systemandthesystem\nwithout thetwisting phase candepend onthedirectionofthephase shift. This isrelated\nthe fact that the effective masses can depend on the direction (th e effective masses are\nrelated to the hopping strengths). Thus in the lattice superfluid de nsity is more like a\ntensor than a scalar.\nIn the free space when the superfluid gas flows the kinetic energy o f the gas is given\nby\nEk=1\n2/integraldisplay\ndr˜ρs(r)vs(r)2,\nwhere ˜ρs(r) is the superfluid density and vs(r) is the superfluid velocity. The superfluid\nvelocity is defined by\nvs(r) =/planckover2pi1\n2m∇φ(r),\nwhereφ(r) is the phase of the order parameter.\nTo impose the linear phase variation to the order parameter, one ca n impose a\nlinear phase variation Θ·Ri= (Θx/(Mxd),Θy/(Myd),Θz/(Mzd))·Ri[26, 27, 36] to the\nHamiltonian. Here Mαindicates the number of lattice sites in direction alpha, i.e., the\ntotal volume the lattice V=MxMyMzd3This variation corresponds small (Θ αis small)\nsuperfluid velocity, which is given by\nvs=/planckover2pi1\n2m(Θx/(Mxd),Θy/(Myd),Θz/(Mzd)).\nThus this imposed phase gradient gives the system a kinetic energy, which corresponds\nto the free energy difference FΘ−F0, whereFΘis the free energy within the the phase\nvariation and F0is the free energy without the phase variation. The superfluid frac tion\nin this case can be determined as [26, 27]\nραα′= lim\nΘ→01\nNFΘ−F0\n¯JxΘαΘα′=1\nN¯Jx∂2FΘ\n∂Θα∂Θα′/vextendsingle/vextendsingle/vextendsingle\nΘ=0, (6)\nwhereNis the total number of particles, and ¯Jx= (J↑,x+J↓,x)/2. This¯Jxcorresponds\nthe effective mass\nmeff=/planckover2pi12\n2¯Jxd2.\nWhen one takes the limit Θ→0 the temperature, and the number of particles (and of\ncourse the interaction strength) should be kept as constants.\nOf course, one could define the superfluid fraction as\nραα= lim\nΘ→01\nNFΘ−F0\n¯Jα,αΘαΘα′(7)Superfluid-density of the ultra-cold Fermi gas in optical la ttices 6\nwhere¯Jα,α′= (J↑,α+J↓,α)/2, but in this case it is a little bit unclear how to define the\noff-diagonal elements. This definition is not directly connected to th e energy difference\nbetween the twisted system and the untwisted system.\nBy imposing the phase variation to the order parameter the one find s\nˆHΘ=−/summationdisplay\nσ=↑,↓\nJσ,x/summationdisplay\n/angbracketlefti,j/angbracketrightx+Jσ,y/summationdisplay\n/angbracketlefti,j/angbracketrighty+Jσ,z/summationdisplay\n/angbracketlefti,j/angbracketrightz\nˆc†\nσ,iˆcσ,j (8)\n−/summationdisplay\ni(µ↑c†\n↑,iˆc↑,i+µ↓c†\n↓,iˆc↓,i)+/summationdisplay\ni|∆|ei(q+2Θ)·Riˆc†\n↑,iˆc†\n↓,i+|∆|e−i(q+2Θ)·Riˆc↓,iˆc↑,i.\nOne can make the unitary transformation [26, 27, 36], which corres ponds the local\ntransformation ˆ cσ,i→ˆcσ,ieiΘ·Ri, and the twisted Hamiltonian becomes\nˆHΘ=−/summationdisplay\ni(µ↑c†\n↑,iˆc↑,i+µ↓c†\n↓,iˆc↓,i) (9)\n+/summationdisplay\ni|∆|eiq·Riˆc†\n↑,iˆc†\n↓,i+|∆|e−iq·Riˆc↓,iˆc↑,i\n−/summationdisplay\nσ/bracketleftBig/summationdisplay\nn/parenleftBig\nJσ,x(eiΘx/Mxˆc†\nσ,nˆcσ,n+dˆx+e−iΘx/Mxˆc†\nσ,nˆcσ,n−dˆx)\n+Jσ,y(eiΘy/Myˆc†\nσ,nˆcσ,n+dˆy+e−iΘy/Myˆc†\nσ,nˆcσ,n−dˆy)\n+Jσ,z(eiΘz/Mzˆc†\nσ,nˆcσ,n+dˆz+e−iΘz/Mzˆc†\nσ,nˆcσ,n−dˆz)/parenrightBig/bracketrightBig\n.\nThetwist anglesΘ αhave tobesufficiently small toavoid effects otherthanthecollective\nflow of the superfluid component. Since Θ α/Mαis small, we can expand up to second\norder\ne±iΘα/Mα≈1±iΘα\nMα−Θ2\nα\n2M2\nα.\nIn this way we can write the twisted Hamiltonian as\nˆHΘ=ˆH0+H′≈ˆH0+/summationdisplay\nσ/bracketleftBigΘ2\nx\n2M2\nx/summationdisplay\nnJσ,x(ˆc†\nσ,nˆcσ,n+dˆx+ˆc†\nσ,nˆcσ,n−dˆx)+ (10)\nΘ2\ny\n2M2\ny/summationdisplay\nnJσ,y(ˆc†\nσ,nˆcσ,n+dˆy+ˆc†\nσ,nˆcσ,n−dˆy)+Θ2\nz\n2M2\nz/summationdisplay\nnJσ,z(ˆc†\nσ,nˆcσ,n+dˆz+ˆc†\nσ,nˆcσ,n−dˆz)\n−iΘx\nMx/summationdisplay\nnJσ,x(ˆc†\nσ,nˆcσ,i+dˆx−ˆc†\nσ,nˆcσ,n−dˆx)−iΘy\nMy/summationdisplay\nnJσ,y(ˆc†\nσ,nˆcσ,i+dˆy−ˆc†\nσ,nˆcσ,n−dˆy)\n−iΘz\nMz/summationdisplay\nnJσ,z(ˆc†\nσ,nˆcσ,i+dˆz−ˆc†\nσ,nˆcσ,n−dˆz)/bracketrightBig\n.\nIn the momentum space this formula becomes\nˆHΘ≈ˆH0+/summationdisplay\nσ,α/bracketleftBigΘ2\nα\n2M2\nα/summationdisplay\nk2Jσ,αcos(kαd)ˆc†\nσ,kˆcσ,k (11)\n+Θα\nMα/summationdisplay\nk2Jσ,αsin(kαd)ˆc†\nσ,kˆcσ,k/bracketrightBig\n=ˆH0+ˆT+ˆJ,Superfluid-density of the ultra-cold Fermi gas in optical la ttices 7\nwhere\nˆT=/summationdisplay\nσ,α/bracketleftBigΘ2\nα\n2M2\nα/summationdisplay\nk2Jσ,αcos(kαd)ˆc†\nσ,kˆcσ,k/bracketrightBig\n(12)\nand\nˆJ=/summationdisplay\nσ,α/bracketleftBigΘα\nMα/summationdisplay\nk2Jσ,αsin(kαd)ˆc†\nσ,kˆcσ,k/bracketrightBig\n. (13)\nThese two terms are proportional to the current, and number op erator of the system,\nrespectively. These two terms commute with the number operator s, thus when one\nimposes the perturbation in the system the number of particles is co nserved. Using the\nsame method, what is used in reference [30], can be shown that in th e mean-field level\nlim\n|Θ|→0∂2F(Θ)\n∂Θα∂Θα′= lim\n|Θ|→0/parenleftbigg∂2Ω(Θ)\n∂Θα∂Θα′/parenrightbigg\n∆,µ↑,µ↓. (14)\nTherefore we can use the grand canonical potential.\nThe grand potential of the system can be written with a perturbing phase gradient\nas a series [37]\nΩΘ= Ω0−∞/summationdisplay\nn=1(−1)n\nβn/planckover2pi1n/integraldisplayβ/planckover2pi1\n0dτ1···/integraldisplayβ/planckover2pi1\n0dτn/angb∇acketleftTτˆH′(τ1)···ˆH′(τn)/angb∇acket∇ight0,(15)\nwhereβ= 1/kBT,kBis Boltzmann’s constant, Ω 0is the grand canonical potential in\nthe absence of the perturbation. The symbol Tτorders the operators such a way that τ\ndecreases from left to right. The brackets /angb∇acketleft.../angb∇acket∇ight0= Tr[exp( −β(ˆH0−µ↑ˆN↑−µ↓ˆN↓))...]\nmean the thermodynamic average evaluated in the equilibrium state o f the unperturbed\nsystem at temperature T. Because all the twisted angles Θ αare small, we can safely\nignore terms of higher order than Θ2\nα. With this approximation the grand potential\nbecomes\nΩΘ≈Ω0+1\nβ/planckover2pi1/integraldisplayβ/planckover2pi1\n0dτ/angb∇acketleftTτˆT(τ)/angb∇acket∇ight0−1\n2β/planckover2pi12/integraldisplayβ/planckover2pi1\n0/integraldisplayβ/planckover2pi1\n0dτ dτ′/angb∇acketleftTτˆJ(τ)J(τ′)/angb∇acket∇ight0,(16)\nThe first perturbation term is easy to calculate and it is given by\n1\nβ/planckover2pi1/integraldisplayβ/planckover2pi1\n0dτ/angb∇acketleftTτˆT(τ)/angb∇acket∇ight0=/summationdisplay\nσ,α/bracketleftBigΘ2\nα\n2M2α/summationdisplay\nk2Jσ,αcos(kαd)Nσ,k/bracketrightBig\n, (17)\nwhereNσ,kis number of particles of the σ-component in the momentum state k. Using\nWick’s theorem the second perturbation term is given by\n−1\n2β/planckover2pi12/integraldisplayβ/planckover2pi1\n0/integraldisplayβ/planckover2pi1\n0dτ dτ′/angb∇acketleftTτˆJ(τ)ˆJ(τ′)/angb∇acket∇ight0 (18)\n=−1\n2β/planckover2pi12/integraldisplayβ/planckover2pi1\n0/integraldisplayβ/planckover2pi1\n0dτ dτ′/bracketleftBig/summationdisplay\nσ,σ′,α,α′ΘαΘα′\nMαMα′/summationdisplay\nk,k′4Jσ,αJσ′,α′sin(kαd)sin(k′\nα′d)\n×/parenleftBig\n(1−δσσ′)δk,−k′+qF(k,τ,τ′)F†(k′,τ,τ′)−δσσ′δkk′Gσ(k,τ,τ′)Gσ′(k′,τ,τ′)/parenrightBig/bracketrightBig\n,Superfluid-density of the ultra-cold Fermi gas in optical la ttices 8\nwhere\nGσ(k,τ,τ′) =−/angb∇acketleftTτˆcσ,k(τ)ˆc†\nσ,k(τ′)/angb∇acket∇ight0=1√β/planckover2pi1∞/summationdisplay\nn=−∞e−iωn(τ−τ′)Gσ(k,ωn)(19)\nF(k,τ,τ′) =−/angb∇acketleftTτˆc↑,k+q(τ)ˆc↓,−k+q(τ′)/angb∇acket∇ight0=1√β/planckover2pi1∞/summationdisplay\nn=−∞e−iωn(τ−τ′)F(k,ωn).\nThe fermionic Matsubara frequencies are given by\nωn=π(2n+1)\n/planckover2pi1β,\nwherenis an integer. The individual Green’s functions of the mean-field theo ry can be\nwritten as\nG↑(k,ωn) =|uk,q|2\ni/planckover2pi1ωn−E+,k,q+|vk,q|2\ni/planckover2pi1ωn+E−,k,q(20)\nG↓(k,ωn) =|uk,q|2\ni/planckover2pi1ωn−E−,k,q+|vk,q|2\ni/planckover2pi1ωn+E+,k,q(21)\nF(k,ωn) =uk,qv∗\nk,q\ni/planckover2pi1ωn+E−,k,q−uk,qv∗\nk,q\ni/planckover2pi1ωn−E+,k,q, (22)\nwhere the quasiparticle dispersions and amplitudes are given by\nE±,k,q=/radicalBigg/parenleftbiggǫ↑,k+q/2+ǫ↓,−k+q/2\n2/parenrightbigg2\n+|∆|2±ǫ↑,k+q/2−ǫ↓,k+q/2\n2(23)\n|uk,q|2=1\n2/parenleftbigg\n1+ǫ↑,k+q/2+ǫ↓,−k+q/2\nE+,k,q+E−,k,q/parenrightbigg\n(24)\n|vk,q|2= 1−|uk,q|2, (25)\nand furthermore ǫσ,k=/summationtext\nα2Jσ,α(1−cos(kαd))−µσ. By calculating the integrals and\nthe Matsubara summations one finds a lengthy formula\n−1\n2β/planckover2pi12/integraldisplayβ/planckover2pi1\n0/integraldisplayβ/planckover2pi1\n0dτ dτ′/angb∇acketleftTτˆJ(τ)ˆJ(τ′)/angb∇acket∇ight0 (26)\n=/summationdisplay\nα,α′ΘαΘα′\nMαMα′/summationdisplay\nk/bracketleftBig\nJ↑,αJ↑,α′sin((kα+qα/2)d)sin((kα′+qα′/2)d)\n×/parenleftBig\n2|uk,q|2|vk,q|2f(E+,k,q)+f(E−,k,q)−1\nE+,k,q+E−,k,q−\nβ|uk,q|4f(E+,k,q)(1−f(E+,k,q))−β|vk,q|4f(E−,k,q)(1−f(E−,k,q))/parenrightBig\n+2J↓,αJ↓,α′sin((kα−qα/2)d)sin((kα′−qα′/2)d)\n×/parenleftBig\n2|uk,q|2|vk,q|2f(E+,k,q)+f(E−,k,q)−1\nE+,k,q+E−,k,q−\nβ|uk,q|4f(E−,k,q)(1−f(E−,k,q))−β|vk,q|4f(E+,k,q)(1−f(E+,k,q))/parenrightBig\n+4J↑,αJ↓,α′sin((kα+qα/2)d)sin((kα′−qα′/2)d)\n×|uk,q|2|vk,q|2/parenleftBig2(f(E+,k,q)−f(E−,k,q))\nE+,k,q−E−,k,qSuperfluid-density of the ultra-cold Fermi gas in optical la ttices 9\n−2f(E+,k,q)−1\n2E+,k,q−2f(E−,k,q)−1\n2E−,k,q/parenrightBig/bracketrightBig\n,\nwheref(E) is the Fermi-Dirac distribution. In order to handle the limit lim k′→k, we\nhave implicitly assumed in equation (26) that the lattice is large, i.e, Mα≫1.\nNow the twisted grand canonical potential can be written as\nΩΘ≈Ω0+/summationdisplay\nα,α′δΩαα′ΘαΘα′\nMαMα′. (27)\nIn all the cases we consider δΩαα′= 0 when α/negationslash=α′. The off-diagonal terms can be non-\nzero only when the single particle dispersion couples the different dire ctions together.\nIn our case where the directions are independent, the ksums in different directions\nare independent, and because sine is an antisymmetric function the se sums vanishes.\nTherefore the grand potential becomes\nΩΘ≈Ω0+/summationdisplay\nαδΩααΘ2\nα\nM2\nα. (28)\nWe can now determine the components of the dimensionless superflu id fraction tensor\nas\nραα′=δΩαα′\n¯Jx(N↑+N↓).\nAs a formula these components are given by\nραα=1\nN/summationdisplay\nk,σ˜Jσ,αcos(kαd)Nσ,k (29)\n+1\nN/summationdisplay\nk/bracketleftBig\n˜J2\n↑,αsin2((kα+qα/2)d)\n×/parenleftBig\n2|uk,q|2|vk,q|2f(E+,k,q)+f(E−,k,q)−1\nE+,k,q+E−,k,q−\nβ|uk,q|4f(E+,k,q)(1−f(E+,k,q))−β|vk,q|4f(E−,k,q)(1−f(E−,k,q))/parenrightBig\n+2˜J2\n↓,αsin2((kα−qα/2)d)/parenleftBig\n2|uk,q|2|vk,q|2f(E+,k,q)+f(E−,k,q)−1\nE+,k,q+E−,k,q−\nβ|uk,q|4f(E−,k,q)(1−f(E−,k,q))−β|vk,q|4f(E+,k,q)(1−f(E+,k,q))/parenrightBig\n+4˜J↑,α˜J↓,α′sin((kα+qα/2)d)sin((kα−qα/2)d)\n×|uk,q|2|vk,q|2/parenleftBig2(f(E+,k,q)−f(E−,k,q))\nE+,k,q−E−,k,q\n−2f(E+,k,q)−1\n2E+,k,q−2f(E−,k,q)−1\n2E−,k,q/parenrightBig/bracketrightBig\nραα′= 0,\nwhereN=N↑+N↓and˜Jσ,α=Jσ,α/¯Jx. When the momentum qis non-zero the\nsuperfluid density describes the one mode FFLO phase, and when th e momentum q= 0\nthesuperfluiddensitydescribestheBCSortheSarmaphase. ρααisasuperfluidfraction,Superfluid-density of the ultra-cold Fermi gas in optical la ttices 10\nit describes what fraction of the atoms is in the superfluid state. In the limit where\nJ↑,α=J↓,α′=Jfor allα,α′andN↑=N↓, the chemical potentials are same and q= 0.\nIn this limit superfluid density tensor can be simplified as\nραα=1\nN/summationdisplay\nk,σcos(kαd)Nσ,k−4βJ\nN/summationdisplay\nksin2(kαd)f(Ek)(1−f(Ek)) (30)\nραα′= 0 (31)\nwhere\nEk=/radicalBigg/parenleftbiggǫ↑,k+ǫ↓,−k\n2/parenrightbigg2\n+|∆|2.\nIn the continuum limit i.e. d→0, where Jd2remains constant, the superfluid\nfraction becomes\nραα=ρ= 1+2/planckover2pi12\nmanV/summationdisplay\nkk2\nz∂f(Ek)\n∂Ek,\nwhere the effective mass ma=/planckover2pi12/(2Jd2),n=n↑+n↓is the total number density and\nVis the size of the system. This continuum result is precisely the Landa u’s formula for\nthe BCS superfluid fraction [38].\n4. Results\n4.1. BCS/Sarma-phase results\nIn the continuum the formula of the BCS superfluid fraction is the we ll known Landau’s\nformula. Figure1(a)showstheBCSsuperfluidfractionasafunctio nofthetemperature,\nin the uniform case (the external potential is zero), without the la ttice. In figure 1 (b)\nwe show the BCS superfluid fraction divided by |∆|2as a a function of the temperature.\nOne sees from figure 1 (a) that the BCS superfluid fraction is one at zero temperature\nand from figure 1 (b) that the superfluid fraction is almost proport ional to |∆|2,\nbut the temperature dependence of the superfluid fraction differ s somewhat from the\ntemperaturedependence of |∆|2. ThestandardBCSresult isthatthesuperfluid fraction\nis proportional to |∆|2in limitT→Tc[39]. When we calculated the gap, we used the\nrenormalization, i.e., we have cancelled out the divergent part of the gap equation, in\nreference [39] is used the cutoff energy. This is the cause of little diff erence between the\nresults.\nWhen all the hopping strengths are the same i.e. J=J↑,α=J↓,α′, and filling\nfractions nσ=Nσ/MxMyMzare same and the average filling fraction nav= (n↑+n↓)/2\ngoes to zero, the superfluid fraction should go to one at zero temp erature. For low filling\nfractions the the atoms occupy only the lowest momentum states a nd the single particle\ndispersions ǫσ,k=/summationtext\nα2Jσ,α(1−cos(kαd))−µσapproaches to form J(kd)2−µσ, which\nis the free space dispersion. Therefore the results on the limit of low filling fractions\nbecome the free space result and we know from the Landau’s formu la that the BCS\nsuperfluid fraction in the free space is one at zero temperature (s ee figure 1 (a)) .Superfluid-density of the ultra-cold Fermi gas in optical la ttices 11\nIn figure 2 (a) we show the BCS superfluid fraction as a function of t he total filling\nfraction in the lattice. As we can see, the BCS superfluid fraction ap proaches one, when\nthe filling fractions become small. In the free space the superfluid fr action is one at zero\ntemperature, but in the lattice this no longer holds. In figure 2 (b) w e showρ(n↑+n↓)\nas a function of the total filling fraction and as one can see from it, t he BCS superfluid\ndensity, which is not divided by the total filling fraction is symmetric wit h respect to\nhalf filling. The cause of this is the particle-hole symmetry of the lattic e model. The\nparticle-hole symmetry of the lattice model can be seen also from fig ures 2 (c) and 2\n(d). In figures 2 (c) and 2 (d) we show ∆ /Jand|∆|2/J2, respectively, as functions of of\nthe total filling fraction. By comparing figures 2 (b)-(d) one can no tice that ρ(n↑+n↓)\nis not directly proportional to ∆ or |∆|2.\nFigure 3 (a) shows the BCS superfluid fraction as a function of the t emperature,\nwith three different average filling fractions. When one compares th is figure to figure 1\n(a), one notices that the lattice result is very similar compared to th e free space result.\nThe superfluid fraction goes to zero, when the paring gap goes to z ero as expected. The\nfigure is consistent with figure 2 (a), in that the total filling fraction increases when the\nsuperfluid fraction decreases. In figure 3 (b) we show the BCS sup erfluid fraction as a\nfunction of the dimensionless coupling strength −U/Jat zero temperature. As we can\nsee, when −U/Jincreases the superfluid fraction decreases contrary to the fre e space\nresult where superfluid fraction is a constant as a function of the c oupling strength at\nzero temperature. Of course at finite temperatures the free sp ace superfluid fraction\ndecreases with increasing coupling strength. This is due to the fact as−U/Jincreases\nthemovement oftheatomsdecreases inthelattice. Inotherword swhen−U/Jincreases\nthen Cooper pairs become smaller and the atoms are better localized . Of course in the\nlimit−U/J→ ∞the mean-field theory fails and in large values of −U/Jthe results\nare not reliable. Figure 3 (c) shows the BCS superfluid fraction divide d by|∆|2as a\nfunction of the temperature, with three different average filling fr actions and we see that\nthe BCS superfluid fraction is almost proportional to |∆|2(compare to figure 1 (b)) i.e\nρ=c(T,n↑=n↓)|∆|2, wherec(T,n↑=n↓) depends weakly on the temperature. We\nshow the pairing gap as a function of −U/Jat zero temperature in figure 3 (d). As one\ncan notice, the BCS superfluid fraction does not behave anything lik e the pairing gap\nas a function of −U/Jat zero temperature.\nSince it is experimentally easy to study anisotropic lattices and use th ese to explore\ndimensional crossovers, let us explore superfluid fractions in aniso tropic lattices. We\nshow in figure 4 (a) the BCS superfluid fraction as a function of the t emperature in\nthe case, where the hopping strengths are different in the differen t directions. As one\ncan see from the figure the components of the superfluid fraction tensor are different\nin different directions. Furthermore, we see that the superfluid fr action is larger in the\ndirection of large hopping strength. This is easy to understand, an d is due to the free\nenergy difference in α-direction being proportional to Jαas we see from equations (28)\nand (29). If the superfluid fraction was defined as in equation (7), the components of\nthe superfluid fraction would be equal in every direction. In other w ords the effectiveSuperfluid-density of the ultra-cold Fermi gas in optical la ttices 12\nmasses are different in different directions. figure 4 (c) shows that in the in the case,\nwhere the hopping strengths are different in the different direction s,ραα/|∆|2is almost\na constant as a function of the temperature.\nFigure 4 (b) shows the BCS superfluid fraction as a function of the t emperature in\nthecase, wherethehoppingstrengthsaredifferentforthecomp onents(but J↑+J↓iskept\nconstant). As it is clear from figure 4 (b), when the ratio J↑/J↓increases the superfluid\nfraction decreases, but the pairing gap, however, does not decr ease as a function of\nJ↑/J↓at zero temperature (it does as a function of the temperature), but it remains\nalmost a constant (at zero temperature it is a constant). This can be seen from figure 4\n(d), in which is shown the pairing gaps as as functions of the tempera ture in the cases,\nwhere the hopping strengths are different for the components. O ne can also notice that\nwhenJ↑/J↓→ ∞thenρ→0 ( in this case the critical temperature also goes to zero).\nWhenJ↑/J↓increases but 2 U/(J↑+J↓) remains constant, the lattice becomes deeper\nand deeper for ↓-component. Thus the atoms of the component ↓are more localized\nand the atoms do not move easily.\nFigure 5 (a) shows the superfluid fraction as a function of polarizat ionP=\n(n↑−n↓)/(n↑+n↓), ataconstant temperature( kBT/J= 0.75). Whenthepolarizationis\nlarger thanzero the stateis called the Sarma state. One sees, tha t when thepolarization\nincreases the superfluid fraction decreases. Figure 5 (b) shows t he superfluid fraction\ndivided by |∆|2as a function of the polarization at a constant temperature. When\nPis about 0 .35 the superfluid fraction divided by |∆|2drops to zero suddenly. This\nhappens because the pairing gap vanishes at this polarization. From figure 5 (b) is seen\nthat the superfluid fraction at a constant temperature is almost p roportional to |∆|2.\nFigure 5 (c) shows the Sarma state superfluid fraction as a functio n of the temperature,\nat a constant polarization ( P= 0.10). Figure 5 (d) shows the Sarma state superfluid\nfraction divided by |∆|2(ρ/|∆|2). We notice from figures 5 (b) and (d) that ρ/|∆|2is\nalmost a constant asa functionof thetemperature andpolarizatio ni.e.ρ=c(T,P)|∆|2,\nwherec(T,P) depends weakly on the temperature and the polarization.\n4.2. Results for FFLO-phase\nWhen the lattice is cubic and all the hopping strengths are equal the one mode FFLO\nsuperfluid fraction is symmetric like the BCS/Sarma superfluid fract ion. This due the\nfollowing fact, when one varies the q, it turns out that the free-energy is minimized,\nwhen the qlies along side the axis i.e. q=qαˆxα, but the system does not favour any of\nthese axis.\nThe BCS-andthe Sarmastatesuperfluid fractionsarealmost direc tly proportional\nto|∆2|i.eρ∼ |∆2|. This changes for the one mode FFLO superfluid fraction as we will\nnow demonstrate. We show in figure 6 (a) ρxx-component of the FFLO superfluid\nfraction as a function of the total filling fraction, with three differe nt interactions\nU/J=−4.0,−5.0,−6.0, at zero temperature. figure 6 (b) shows ρxx(n↑+n↓) as a\nfunction of the total filling fraction, when U/J=−6.0, at zero temperature. The dataSuperfluid-density of the ultra-cold Fermi gas in optical la ttices 13\nwhich is used in figure 6 (a) and (b) have been calculated just above t he Clogston\nlimit [40]. The Clogston limit is the limit for the chemical potential differenc e below\nwhich one can find only the BCS type solution when one minimizes the gra nd potential.\nThe numerical value of this limit is roughly δµ≈√\n2∆0, where ∆ 0is the pairing gap at\nzero temperature when δµ= 0. As one can see from figure 6 (a) when |U/J|increases\nthe superfluid fraction decreases and eventually becomes negativ e. One can also see,\nthat as the total filling fraction increases the superfluid fraction d ecreases but does\nnot necessarily become negative. Negative superfluid fraction implie s that the gas is\nunstable. If superfluid fraction is negative the state around which we expanded is not\nthe ground state of the system, and thus the state cannot be st able. In the case where\nsuperfluid fractionis negative, the real minimum of the free energy is beyond our ansatz.\nThe real minimum of the system may be a some kind of phase separatio n or modulating\ngap state (like the two-mode FFLO state) [41, 42, 43].\nThe superfluid density i.e. ρxx(n↑+n↓) is symmetric as a function of the total filling\nfraction, with the value n↑+n↓= 1.0, this is shown in figure 6 (b). This is again due\nto the particle-hole symmetry of the lattice model.\nFigure 7 (a) shows the superfluid fraction as a function of polarizat ion at a constant\ntemperature ( kBT/J≈0.23), and figure 7 (b) shows the superfluid fraction divided by\n|∆|2as a function of polarization. There is a second order phase transit ion between the\nSarma/BCS phase and the FFLO phase, when polarization P≈0.15, this can be seen\nfrom the bend in the superfluid fraction, clearly. We see also from fig ure 7 (b) that the\nFFLO superfluid fraction is not proportional to |∆|2. When one writes ρ=c(T,P)|∆|2,\nthe polarization dependence of c(T,P) is very different in the FFLO phase compared to\nthat in the Sarma phase.\n0.010.020.030.040.050.0600.10.20.30.40.50.60.70.80.91ρ\nT/TF(a)\n0.010.020.030.040.050.0601020304050607080ρ/|∆|2\nT/TF(b)\nFigure 1. Figure (a) we show the BCS superfluid fraction as a function of the\ntemperature in a free space (absence of the lattice). Figure (b) s hows the BCS\nsuperfluid fraction divided by |∆|2as a function of temperature. The interaction\nstrength kFas=−0.30, where kFis the Fermi wave vector.\nIn figures 8 and 9 we present phase diagrams with four different inte raction\nstrengths, when the average filling fractions are nav= 0.2 andnav= 0.5, respectively.\nThe figures show that there can be unstable regions in the FFLO pha se. In theSuperfluid-density of the ultra-cold Fermi gas in optical la ttices 14\n0.5 1 1.500.10.20.30.40.50.60.70.80.91ρ\nn↑+n↓(a)\n0.5 1 1.500.050.10.150.20.25ρ(n↑+n↓)\nn↑+n↓(b)\n0.5 1 1.500.511.52∆/J\nn↑+n↓(c)\n0.5 1 1.5012345|∆|2/J2\nn↑+n↓(d)\nFigure 2. In figure (a) we show the BCS superfluid fraction ρ=ρxx=ρyy=ρzzas\na function of the total filling fraction n↑+n↓= 2n↑= 2n↓at zero temperature. In\nfigure (b) is shown ρ(n↑+n↓) as a function of the total filling fraction. Figures (c) and\n(d) show ∆ /Jand|∆|2/J2, respectively as functions of the total filling fraction. All\nthe hopping strengths are same i.e. J=J↑,α=J↓,α′, and−U/J= 6.0.\nunstable FFLO region the superfluid density is negative, which implies t hat the phase\nis dynamically unstable. However, these instabilities occur only when t he interaction\nis relative high. By comparing the two figures one notices that the un stable regions\nincrease when the density increases. Thus when the density decre ases the unstable\nregions eventually disappear, and this implies that the FFLO phase re gion is stable\nwithout the lattice. Due to the particle-hole symmetry average filling fraction 0 .5 is\nthe worst case for the stability, but even when the average densit y is 0.5, the one\nmode FFLO is stable when |U/J|<4.4 for any polarization. The induced interactions\nmight make the FFLO phase more stable, since these induced interac tions make the\neffective interaction weaker [44]. References [22, 45] considered also the possibility of\nphase separation. However, this work we do not consider phase se paration, because\nwhen there is phase separation between the BCS phase and the nor mal gas, there is a\nsuperfluid gas in a part of the lattice and normal gas in another part of the lattice. In\nthis case the superfluid density is the standard BCS superfluid dens ity and the gas is\nstable. It should be notice that the region of the phase separation does not include the\nunstable regions of the FFLO phase [22].Superfluid-density of the ultra-cold Fermi gas in optical la ttices 15\n00.2 0.4 0.6 0.8 11.200.050.10.150.20.250.30.35\nkBT/Jρ(a)\n \nnav=0.3\nnav=0.5\nnav=0.7\n20 40 60 80 10000.050.10.150.20.250.30.350.4ρ\n−U/J(b)\n00.2 0.4 0.6 0.8 11.200.010.020.030.040.050.060.070.08\nkBT/Jρ/|∆|2(c)\n \nnav=0.3\nnav=0.5\nnav=0.7\n20 40 60 80 10001020304050∆/J\n−U/J(d)\nFigure 3. Figure (a) shows the BCS superfluid fraction ρ=ρxx=ρyy=ρzz\nas a function of the temperature,with three different average filling fractions nav=\n(n↑+n↓)/2. Figure (b) shows the BCS superfluid fraction as a function of a\ndimensionless coupling strength −U/Jat zero temperature. In figure (c) we show\nthe BCS superfluid fraction divided by |∆|2as a function of the temperature,with\nthree different average filling fractions. Figure (d) shows the pairin g gap as a function\nof−U/Jat zero temperature. In the all figures all the hopping strengths are same\ni.e.J=J↑,α=J↓,α′. In figures (a) and (c) −U/J= 6.0, and in figures (b) and (d)\nn↑=n↓= 0.5.\n5. Conclusions\nIn this paper we have presented, at the mean-field level, the super fluid density of the\ntwo component Fermi gas in an optical lattice. We have shown that t he BCS superfluid\ndensity in optical lattices differs from the free space results. We ha ve also shown that\nthe one-mode FFLO superfluid density differs crucially from the BCS/ Sarma superfluid\ndensity. In the BCS/Sarma phase the gas is always a stable superflu id, but in the\nFFLO phase dynamical instabilities can appear. However, the FFLO p hase is stable,\nwhen|U/J|<4.0 i.e, on the BCS limit. Even in the BCS phase the superfluid density\ncan be different in different directions depending on the lattice struc ture.\nAlthough the one mode FFLO phase is stable when |U/J|<4.0, it is unlikely that\nthe one mode FFLO is the free energy minimum the system; calculation s made by using\nthe Bogoliubov-de Gennes equations show that the spatially modulat ing gap might be\nmore favorable [41, 42, 43]. Such states are also, quite likely, more s table than the oneSuperfluid-density of the ultra-cold Fermi gas in optical la ttices 16\n0.2 0.4 0.6 0.8 11.200.050.10.150.20.250.30.350.40.45\nkBT/Jxρ(a)\n \nρxx\nρyy\nρzz\n0.20.40.60.8 11.200.050.10.150.20.25\n2kBT/(J↑+J↓)ρ(b)\n \nJ↑/J↓=1.2\nJ↑/J↓=5.7\nJ↑/J↓=20\nJ↑/J↓=200\n0.2 0.4 0.6 0.8 11.200.020.040.060.080.10.12\nkBT/Jxρ/|∆|2(c)\n \nρxx/|∆|2\nρyy/|∆|2\nρzz/|∆|2\n0.20.40.60.8 11.200.511.522.5\n2kBT/(J↑+J↓)2∆/(J↑+J↓)(d)\n \nJ↑/J↓=1.2\nJ↑/J↓=5.7\nJ↑/J↓=20\nJ↑/J↓=200\nFigure 4. [Colour online] In figure (a) we show the BCS superfluid fraction as a\nfunction of the temperature in the case, where the hopping stren gths are different\nin the different directions. In figure (b) we show the BCS superfluid f raction as a\nfunction of the temperature. Figure (c) shows the BCS superfluid fraction divided\nby|∆|2as a function of the temperature. In figure (d) we show the pairing gaps\nas functions of the temperature. In figures (a) and (c) Jx= (J↑,x+J↓,x)/2,\nJ↑,y=J↓,y= 0.8J↑,x= 0.8J↓,x,J↑,z=J↓,z= 0.5J↑,x= 0.5J↓,x,n↑=n↓= 0.2,\nand−U/J↑,x= 6.0. In figures (b) and (d) Jσ,α=Jσ′,α′=Jσ,−2U/(J↑+J↓) = 6.0,\nandn↑=n↓= 0.5\nmode FFLO phase.\nThe methods which have used in this paper to calculate thesuperfluid density, work\nin principle for many different systems. For example, one could use th e methods used\nhere to calculate the superfluid density in the system where anothe r component of the\ntwo component Fermi gas occupies the first excited band [46], or in the case where an\noptical lattice is in a trap. These methods can also been used calculat ing the superfluid\ndensity of modulating gap states.\nAcknowledgments I would like to thank J.-P. Martikainen and T. K. Koponen\nfor many illuminating discussions. This work was supported by Academ y of Finland\n(Project No. 1110191).Superfluid-density of the ultra-cold Fermi gas in optical la ttices 17\n00.050.10.150.20.250.300.020.040.060.080.10.120.140.160.180.2ρ\nP(a)\n00.050.10.150.20.250.300.0050.010.0150.020.0250.030.0350.040.045ρ/|∆|2\nP(b)\n0.7 0.8 0.9 1 1.1 1.200.020.040.060.080.10.120.140.16ρ\nkBT/J(c)\n0.7 0.8 0.9 1 1.1 1.200.0050.010.0150.020.0250.030.0350.040.045ρ/|∆|2\nkBT/J(d)\nFigure 5. [Colour online] The BP/Sarma phase superfluid fraction. In figure (a ) we\nshow the superfluid fraction as a function of the polarization P= (n↑−n↓)/(n↑−n↓).\nFigure (b) shows the superfluid fraction divided by |∆|2. Figure (c) shows the\nsuperfluid fraction as a function of the temperature. In figure (d ) we show the\nsuperfluid fraction divided by |∆|2. In figures (a) and (b) kBT/J= 0.75 and in\nfigures (c) and (d) P= 0.10. In all the figures all the hopping strengths are same i.e.\nJσ,α=Jσ′,α′=J, the average filling fraction nav= 0.50,U/J=−6.0, andq= 0.\n6. References\n[1] M. W. Zwierlein, A. Schirotzek, C. H. Schunck, and W. Ketterle. F ermionic superfluidity with\nimbalanced spin populations. Science, 311:492, 2006.\n[2] G. B. Partridge, W. Li, R. I. Kamar, Y. Liao, and R. G. Hulet. Pairin g and phase separation in\na polarized Fermi gas. Science, 311:506, 2006.\n[3] M. W. Zwierlein, C. H. Schunck, A. Schirotzek, and W. Ketterle. D irect observation of the\nsuperfluid phase transition in ultracold Fermi gases. Nature, 442:54–58, 2006.\n[4] Y. Shin, M. W. Zwierlein, C. H. Schunck, A. Schirotzek, and W. Ket terle. Observation of phase\nseparation in a strongly-interacting imbalanced Fermi gas. Phys. Rev. Lett. , 97:030401, 2006.\n[5] G. B. Partridge, W. Li, Y. Liao, R. G. Hulet, M. Haque, and H. T. C. Stoof. Deformation of a\ntrapped Fermi gas with unequal spin populations. Phys. Rev. Lett. , 97:190407, 2006.\n[6] Y. Shin, C. H. Schunck, A. Schirotzek, and W. Ketterle. Phase d iagramofa two-componentFermi\ngas with resonant interactions. Nature, 451:689 – 693, 2008.\n[7] P. Fulde and R. A. Ferrell. Phys. Rev. , 135:A550, 1964.\n[8] A. I. Larkin and Y. N. Ovchinnikov. Zh. Eksp. Teor. Fiz. , 47:1136, 1964.\n[9] G. Sarma. Sarma phase. J. Phys. Chem. Solids , 24:1029, 1963.\n[10] W. V. Liu and F. Wilczek. Interior gap superfluidity. Phys. Rev. Lett. , 90:047002, 2003.\n[11] R. Casalbuoni and G. Nardulli. Inhomogeneous superconductiv ity in condensed matter and qcd.Superfluid-density of the ultra-cold Fermi gas in optical la ttices 18\n0.2 0.4 0.6 0.8 1−0.2−0.100.10.20.30.40.5\nn↑+n↓ρxx(a)\n \nU/J=−6.0\nU/J=−5.0\nU/J=−4.0\n0 0.5 1 1.5 2−0.2−0.15−0.1−0.050ρ(n↑+n↓)\nn↑+n↓(b)\nFigure 6. The FFLO superfluid fraction at zero temperature. In figure (a) w e show\nthe FFLO superfluid fraction as a function of the total filling fractio n, with three\ndifferent interactions at zero temperature. In figure (b) we show ρxx(n↑+n↓) as a\nfunction of the total filling fraction at zero temperature, when U/J=−6.0. In the\nboth figures all the hopping strengths are same i.e. Jσ,α=Jσ′,α′=J,q=qxˆx, and\nP≈0.30. The horizontal lines in the figures show the value 0.\n0 0.1 0.2 0.3 0.4 0.500.10.20.30.40.5\nPρ(a)\n0 0.1 0.2 0.3 0.4 0.500.10.20.30.40.50.60.7\nPρ/|∆|2(b)\nFigure 7. Figure (a) shows the Sarma- and FFLO phase superfluid fraction as\na function of the polarization at constant temperature. In figure (b) we show the\nSarma- and FFLO phase superfluid fraction divided by |∆|2. The parameters, which\nwere used, are kBT/J≈0.23,U/J≈ −5.1, and the average filling fraction is\nnav= (n↑+n↓)/2 = 0.2, All the hopping strengths were equal. Below P≈0.15\nq= 0 and the phase is the Sarma phase and above P≈0.15q/negationslash= 0 and the phase is\nthe one mode FFLO phase. The sudden drop in figure (b) at P≈0.5 in indicates that\nsystem becomes normal for higher polarizations.\nRev. Mod. Phys. , 76:263, 2004.\n[12] C. Orzel, A. K. Tuchman, M. L. Fenselau, M. Yasuda, and M. A. K asevich. Squeezed states in a\nBose-Einstein condensate. Science, 291:2386, 2001.\n[13] M. Greiner, I. Bloch, O. Mandel, T. W. H¨ ansch, and T. Esslinger . Exploring phase coherence in\na 2d lattice of Bose-Einstein condensates. Phys. Rev. Lett. , 87:160405, 2001.\n[14] S. Burger, F. S. Cataliotti, C. Fort, F. Minardi, M. Inguscio, M. L. Chiofalo, and M. P. Tosi.\nSuperfluidanddissipativedynamicsofaBose-Einsteincondensatein aperiodicopticalpotential.\nPhys. Rev. Lett. , 86:4447, August 2001.\n[15] Z. Hadzibabic, S. Stock, B. Battelier, V. Bretin, and J. Dalibard . Interference of an array of\nindependent Bose-Einstein condensates. Phys. Rev. Lett. , 93:180403, 2004.Superfluid-density of the ultra-cold Fermi gas in optical la ttices 19\n00.050.10.150.20.250.30.3500.050.10.150.20.250.30.35\nkBT/JP\n00.1 0.2 0.3 0.4 0.500.10.20.30.40.5\nkBT/JP\n00.10.20.30.40.50.60.700.10.20.30.40.50.6\nkBT/JP\n00.2 0.4 0.6 0.8 100.20.40.60.8\nkBT/JP\nFigure 8. [Color online] Phase diagrams with four different interaction strengt hs,\nwhen the average filling fraction nav= 0.2. The interaction strengths are from up left\nto bottom right: U/J≈ −3.7,U/J≈ −4.4,U/J≈ −5.1, andU/J≈ −6.3. All the\nhopping strengths are equal. The color are as follows:BCS/Sarma=b lue(black), stable\nFFLO=yellow(light grey), unstable FFLO=red(dark grey), normal gas=white.\n[16] W. Hofstetter, J. I. Cirac, P. Zoller, E. Demler, and M. D. Lukin . High-temperature superfluidity\nof Fermionic atoms in optical lattices. Phys. Rev. Lett. , 89:220407, 2002.\n[17] G. Orso and G.V. Shlyapnikov. Superfluid Fermi gas in a 1d optical lattice. Phys. Rev. Lett. ,\n95:260402, 2005.\n[18] L. P. Pitaevskii, S. Stringari, and G. Orso. Sound propagation a nd oscillations of a superfluid\nFermi gas in the presence of a one-dimensional optical lattice. Phys. Rev. A , 71:053602, 2005.\n[19] M. Iskin and C. A. R. Sa de Melo. Imbalanced Fermi-Fermi mixture s in optical lattices. Phys.\nRev. Lett. , 99:080403, 2007.\n[20] T. Koponen, J.-P. Martikainen, J. Kinnunen, and P. T¨ orm¨ a. S ound velocity and dimensional\ncrossover in a superfluid Fermi gas in an optical lattice. Phys. Rev. A , 73:033620, 2006.\n[21] T. Koponen, J. Kinnunen, J.-P. Martikainen, L. M. Jensen, and P. T¨ orm¨ a. Fermion pairing with\nspin-density imbalance in an optical lattice. New J. Phys. , 8:179, 2006.\n[22] T.K. Koponen, T. Paananen,J.-P.Martikainen, andP. T¨ orm¨ a. Finite temperature phasediagram\nof a polarized Fermi gas in an optical lattice. Phys. Rev. Lett. , 99:120403, 2007.\n[23] J. K. Chin, D. E. Miller, Y. Liu, C. Stan, W. Setiawan, C. Sanner, K . Xu, and W. Ketterle.\nEvidence for superfluidity of ultracold Fermions in an optical lattice. Nature, 443:961–964,\n2006.\n[24] C. H. Schunck, Y. Shin, A. Schirotzek, M. W. Zwierlein, and W. Ke tterle. Pairing without\nsuperfluidity: The ground state of an imbalanced Fermi mixture. Science, 316:867–870, 2007.\n[25] M.E.Fisher, M.N. Barber,andD. Jasnow. Helicitymodulus, sup erfluidity, andscalinginisotropy\nsystems. Phys. Rev. A , 8:1111, 1973.\n[26] A. M. Rey, K. Burnett, R. Roth, M. Edwards, C. J. Williams, and C . W. Clark. BogoliubovSuperfluid-density of the ultra-cold Fermi gas in optical la ttices 20\n00.1 0.2 0.3 0.4 0.500.050.10.150.20.250.3\nkBT/JP\n00.10.20.30.40.50.60.700.10.20.30.4\nkBT/JP\n0 0.2 0.4 0.6 0.8 100.10.20.30.40.50.6\nkBT/JP\n00.20.40.60.811.200.10.20.30.40.50.6\nkBT/JP\nFigure 9. [Color online] Phase diagrams with four different interaction strengt hs,\nwhen the average filling fraction nav= 0.5. The interaction strengths are from up left\nto bottom right: U/J≈ −3.7,U/J≈ −4.4,U/J≈ −5.1, andU/J≈ −6.3. All the\nhopping strengths are equal. The color are as follows:BCS/Sarma=b lue(black), stable\nFFLO=yellow(light grey), unstable FFLO=red(dark grey), normal gas=white.\napproach to superfluidity of atoms in an optical lattice. J. Phys. B , 36:825, 2003.\n[27] R. Roth and K. Burnett. Superfluidity and interference patte rn of ultracold bosons in optical\nlattices. Phys. Rev. A , 67:031602, 2003.\n[28] B. Wu and Q. Niu. Landau and dynamical instabilities of the superfl ow of Bose-Einstein\ncondensates in optical lattices. Phys. Rev. A , 64:061603, 2001.\n[29] A.A. Burkov and A. Paramekanti. Stability of superflow for ultra cold Fermions in optical lattices.\nPhys. Rev. Lett. , 100:255301, 2008.\n[30] E. Taylor, A. Griffin, N. Fukushima, and Y. Ohashi. Pairing fluctua tions and the superfluid\ndensity through the BCS-BEC crossover. Phys. Rev. A , 74:063626, 2006.\n[31] N. Fukushima, Y. Ohashi, E. Taylor, and A. Griffin. Superfluid den sity and condensate fraction\nin the BCS-BEC crossover regime at finite temperatures. Phys. Rev. A , 75:033609, 2007.\n[32] E. Taylor. The josephson relation for the superfluid density in t he BCS-BEC crossover. Phys.\nRev. B, 77:144521, 2008.\n[33] V. M. Stojanovic, W. V. Liu, and Y. B. Kim. Unconventional inter action between vortices in a\npolarized Fermi gas. Annals of Physics , 323:989, 2008.\n[34] M. Iskin and C. A. R. S de Melo. Superfluidity of p-wave and s-wav eatomic Fermi gases in optical\nlattices. Phys. Rev. B , 72:224513, 2005.\n[35] I. Bloch, J. Dalibard, and W. Zwerger. Many-body physics with u ltracold gases. Rev. Mod. Phys ,\n80:885, 2008.\n[36] J.-P. Martikainen and H. T. C. Stoof. Spontaneous squeezing o f a vortex in an optical lattice.\nPhys. Rev. A , 70:013604, 2004.\n[37] G. D. Mahan. Many-Particle Physics, 3rd ed. Kluwer Academic/Plenum Publishers, New York,\n2000.Superfluid-density of the ultra-cold Fermi gas in optical la ttices 21\n[38] E. M. Lifshitz and L. P. Pitaevskii. Statistical Physics, Part 2 . Butterworth-Heinemann, Oxford,\n2002.\n[39] A. L. Fetter and J. D. Walecka. Quantum theory of many-particle systems . Dover, New York,\n2003.\n[40] A. M. Clogston. Upper limit for the critical field in hard supercond uctors.Phys. Rev. Lett. , 9:266\n– 267, 1962.\n[41] Meera M. Parish, Stefan K. Baur, Erich J. Mueller, and David A. H use. Quasi-one-dimensional\npolarized Fermi superfluids. Phys. Rev. Lett. , 99:250403, 2007.\n[42] Y. Chen, Z. D. Wang, F. C. Zhang, and C. S. Ting. Exploring exot ic superfluidity of polarized\nultracold Fermions in optical lattices. 2007.\n[43] M. Iskin and C. J. Williams. Population imbalanced Fermions in harmon ically trapped optical\nlattices. Phys. Rev. A , 78:011603(R), 2008.\n[44] D.-H. Kim, P. T¨ orm¨ a, and J.-P. Martikainen. Induced interact ions for ultra-cold Fermi gases in\noptical lattices. Phys. Rev. Lett. , 102:245301, 2009.\n[45] T. K. Koponen, T. Paananen, J.-P. Martikainen, M. R. Bakhtiar i, and P. T¨ orm¨ a. FFLO state\nin 1, 2, and 3 dimensional optical lattices combined with a non-uniform background potential.\nNew Journal of Physics , 10:045104, 2008.\n[46] J.-P. Martikainen, E. Lundh, and T. Paananen. Interband phy sics in an ultra-cold Fermi gas in\nan optical lattice. Phys. Rev. A , 78:023607, 2008." }, { "title": "0811.4513v2.Continuity_of_the_integrated_density_of_states_on_random_length_metric_graphs.pdf", "content": "arXiv:0811.4513v2 [math.SP] 25 Apr 2009CONTINUITY OF THE INTEGRATED DENSITY OF STATES ON\nRANDOM LENGTH METRIC GRAPHS\nDANIEL LENZ, NORBERT PEYERIMHOFF, OLAF POST, AND IVAN VESEL I´C\nAbstract. We establish several properties of the integrated density of stat es for\nrandom quantum graphs: Under appropriate ergodicity and amena bility assump-\ntions, the integrated density of states can be defined using an exh austion procedure\nby compact subgraphs. A trace per unit volume formula holds, similar ly as in the\nEuclidean case. Our setting includes periodic graphs. For a model wh ere the edge\nlengths are random and vary independently in a smooth way we prove a Wegner\nestimate and related regularity results for the integrated density of states.\nThese results are illustrated for an example based on the Kagome lat tice. In the\nperiodic case we characterise all compactly supported eigenfunct ions and calculate\nthe position and size of discontinuities of the integrated density of s tates.\n1.Introduction\nQuantum graphs are Laplace or Schr¨ odinger operators on metric graphs. As struc-\ntures intermediate between discrete and continuum objects they have received quite\nsome attention in recent years in mathematics, physics and materia l sciences, see e.g.\nthe recent proceeding volume [EKK+08] for an overview.\nHere, westudyperiodicandrandomquantumgraphs. Ourresultsc oncernspectral\nproperties which are related to the integrated density of states ( IDS), sometimes\ncalled spectral distribution function. As in the case of random Schr ¨ odinger operators\nin Euclidean space, disorder may enter the operator via the potent ial. Moreover, and\nthis is specific to quantum graphs, randomness may also influence th e characteristic\ngeometric ingredients determining the operator, viz.\n•the lengths of the edges of the metric graph and\n•the vertex conditions at each junction between the edges.\nIn the present paper we pay special attention to randomness in th ese geometric\ndata. Our results may be summarised as follows. For quite wide classe s of quantum\ngraphs we establish\n•the existence, respectively the convergence in the macroscopic lim it, of the\nintegrated density of states under suitable ergodicity and amenab ility con-\nditions (see Theorem 2.6),\n•a trace per unit volume formula for the IDS (see equation (2.9)),\n•a Wegner estimate for random edge length models (assuming indepen dence\nand smoothness for the disorder) (Theorem 2.9). This implies quant itative\ncontinuity estimates for the IDS (Corollary 2.10).\nDate: October 24, 2018, File:contids-qg20.tex .\n2000Mathematics Subject Classification. 35J10; 82B44.\nKey words and phrases. integrated density of states, periodic and random operators, me tric\ngraphs, quantum graphs, continuity properties.\n12 D. LENZ, N. PEYERIMHOFF, O. POST, AND I. VESELI ´C\nThese abstract results are illustrated by the thorough discussion of an example\nconcerning a combinatorial and a metric graph based on the Kagome lattice. In this\ncase we calculate positions and sizes of all jumps of the IDS. Our res ults show the\neffect of smoothing of the IDS via randomness.\nThe article is organised as follows: In the remainder of this section we summarise\nthe origin of results about the construction of the IDS and of Wegn er estimates and\npoint out aspects of the proofs which are different in the case of qu antum graphs\nin comparison to random Schr¨ odinger operators on L2(Rd) orℓ2(Zd). We mention\nbriefly recent results about spectral properties of random quan tum graphs which are\nin some sense complementary to ours. Finally, we point out some open problems in\nthis field of research. In the next section, we introduce the rando m length model\nand state the main results. In Section 3 we present the Kagome latt ice example. In\nSection 4 we prove Theorem 2.6 concerning the approximability of the IDS. Finally,\nin Section 5 we prove the Wegner estimate Theorem 2.9.\nIntuitively, the IDS concerns the number of quantum states per u nit volume below\na prescribed energy. From the physics point of view the natural de finition of this\nquantity is via a macroscopic limit. This amounts to approximating the ( ensemble-\naveraged) spectral distribution function of an operator on the w hole space by nor-\nmalised eigenvalue counting functions associated to finite-volume re strictions of the\noperator. For ergodic random and almost-periodic operators in Eu clidean space this\napproach has been implemented rigorously in [Pas71, Shu79], and dev eloped further\nin a number of papers, among them [KM82], [Mat93] and [HLMW01]. All t hese\noperators were stationary and ergodic with respect to a commuta tive group of trans-\nlations. For graphs and manifolds beyond Euclidean space the releva nt group is in\ngeneral no longer abelian. The first result establishing the approxim ability of the\nIDS of a periodic Schr¨ odinger operator on a manifold was [AS93]. An important\nassumption on the underlying geometry is amenability. Analogous res ults on tran-\nsitive graphs have been established e.g. in [MY02] and [MSY03]. For Sc hr¨ odinger\noperators with a random potential on a manifold with an amenable cov ering group\nthe existence of the IDS was established in [PV02], and for Laplace-B eltrami oper-\nators with random metrics in [LPV04]. For analogous results for discr ete operators\non amenable graphs see e.g. [Ves05] and [LV08]. A key ingredient of t he proofs of the\nabove results is the amenable ergodic theorem of [Lin01]. More recen ty, the question\nof approximation of the IDS uniformly with respect to the energy va riable has been\npursued, see for instance [LMV08] and the references therein.\nIndependently of the approximability by finite volume eigenvalue coun ting func-\ntions it is possible to give an abstract definition of the IDS by an avera ged trace per\nunit volume formula, see [Shu79, BLT85, Len99, LPV07]. In the ame nable setting,\nboth definitions of the IDS coincide.\nFor a certain class of metric graphs the approximability of the IDS ha s been\nestablished before. In [HV07, GLV07, GLV08] this has been carried out for random\nmetricgraphswithan Zdstructure. Astepoftheproofwhichisspecifictothesetting\nof quantum graphs concerns the influence of finite rank perturba tions on eigenvalue\ncounting functions. When one considers Laplacians on manifolds, on e would rather\nuse the principle of not feeling the boundary of heat kernels, cf. e.g . [AS93, PV02,\nLPV04], to derive the analogous step of the proof.CONTINUITY OF THE IDS ON RANDOM LENGTH METRIC GRAPHS 3\nNext we discuss the literature on Wegner estimates and on the regu larity of the\nIDS. Wegner gave in [Weg81] convincing arguments for the Lipschitz continuity of\nthe IDS of the discrete Anderson model on ℓ2(Zd). The proof is based on an esti-\nmate for the expected number of eigenvalues in a finite energy inter val of a restricted\nbox Hamiltonian. A rigorous proof of the latter estimate was given in [ Kir96] (for\nthe analogous alloy-type model on L2(Rd)). However, the bound of [Kir96] was not\nsufficient to establish the Lipschitz continuity of the IDS. In [CHN01 ] tools to prove\nH¨ older continuity were supplied, see also [HKN+06]. They concern bounds on the\nspectral shift function. Up to now the most widely applicable result c oncerning\nthe Lipschitz-continuity of the IDS is given in [CHK07]. An alternative a pproach\nto derive Lipschitz continuity of the IDS goes via spectral averagin g of resolvents,\nsee [KS87, CH94]. However, this method requires more assumptions on the under-\nlying model.\nWegner’s estimate and all references mentioned so far concern th e case where\nthe random variables couple to a perturbation which is a non-negativ e operator. If\nthis is not the case, additional ideas are necessary to obtain the de sired bounds,\nsee [Klo95, Ves02, HK02, KV06, Ves08]. In our situation, where the perturbation\nconcerns the metric of the underlying space, the dependence on t he random variables\nis not monotone. This is also the case for random metrics on manifolds studied\nin [LPPV08]. To deal with non-monotonicity, the proof of the Wegner estimate\n(Theorem 2.9) takes up an idea developed in [LPPV08], which is not unre lated\nto [Klo95]. The relevant formula used in the proof is (5.2). We need also a partial\nintegration formula whose usefulness was first seen in [HK02].\nIn the context of quantum graphs it is not necessary to rely on sop histicated\nestimates on the spectral shift function. It is sufficient to adapt a finite rank pertur-\nbation bound, which was used in [KV02] for the analysis of one-dimens ional random\nSchr¨ odingeroperators. Theseestimatesarecloselyrelatedtot hefiniterankestimates\nmentioned earlier in the context of the approximability of the IDS. Fo r Schr¨ odinger\noperators on metric graphs where the randomness enters via the potential, Wegner\nestimates have been proved in [HV07, GV08, GHV08]. In the recent p reprint [KP09]\na Wegner estimate for a model with Zd-structure and random edge lengths has been\nestablished. The proof is based on different methods than we use in t he present\npaper.\nNext we want to explain an application of Wegner estimates apart fro m the con-\ntinuity of the IDS. It concerns the phenomenon of localisation of wa ves in random\nmedia. More precisely, for certain types of random Schr¨ odinger o perators on ℓ2(Zd)\nand onL2(Rd) it is well known that in certain energy intervals near spectral boun d-\naries the spectrum is pure point. There are two basic methods to es tablish this fact\n(apart form the one-dimensional situation where specific methods apply). The first\none is called multiscale analysis and was invented in [FS83]. The second a pproach\nfrom [AM93] is called fractional moment method or Aizenman-Molchan ov method.\nA certain step of the localisation proof via multiscale analysis concern s the control of\nspectral resonances of finite box Hamiltonians. A possibility to achie ve this control is\nthe use of a Wegner estimate. In fact, the Wegner estimates need ed for this purpose\nare much weaker than those necessary to establish regularity of t he IDS. This has\nbeen discussed in the context of random quantum graphs in Section 3.2 of [GHV08].4 D. LENZ, N. PEYERIMHOFF, O. POST, AND I. VESELI ´C\nRecently localisation hasbeen proven for several types ofrandom quantum graphs.\nIn[EHS07,KP08,KP09]thishasbeendoneformodelswith Zd-structure, while[HP06]\nconsiders operators on tree-graphs. On the other hand, deloca lisation, i.e. existence\nof absolutely continuous spectrum, for quantum graph models on t rees has been\nshown in [ASW06a]. This result should be seen in the context of earlier , similar\nresults for combinatorial tree graphs [Kle96, Kle98, ASW06a, ASW0 6b, FHS06].\nNowletusdiscusssomeopenquestionsconcerningrandomquantum graphmodels.\nAs for models on ℓ2(Zd), proofs of localisation require that the random variables\nentering the operators should have a regular distribution. In part icular, if the law of\nthe variables is a Bernoulli measure, no known proof of localisation ap plies. This is\ndifferent for random Schr¨ odinger operators L2(Rd). Using a quantitative version of\nthe unique continuation principle for solutions of Schr¨ odinger equa tions, localisation\nwas established in [BK05] for certain models with Bernoulli disorder. T he proof\ndoes not carry over to the analogous model on ℓ2(Zd), since there is no appropriate\nversion of the unique continuation principle available. For random qua ntum graphs\nthesituationisevenworse, sincetheyexhibitingreatgeneralitycom pactlysupported\neigenfunctions, even if the underlying graph is Zd.\nLikeforrandom, ergodicSchr¨ odinger operatorson ℓ2(Zd)andonL2(Rd)thereisno\nproof of delocalisation for random quantum graphs with Zdstructure. In the above\nmentioned papers on delocalisation it was essential that the underly ing graph is a\ntree. An even harder question concerns the mobility edge. Based o n physical reason-\ning one expects that localised point spectrum and delocalised absolut ely continuous\nspectrumshouldbeseparatedindisjointintervalsbymobilityedges. Inthecontextof\nrandom operators where the disorder enters via the geometry th is leads to an intrigu-\ning question pointed out already in [CCF+86]. If one considers a graphover Zdwhich\nis diluted by a percolation process, the Laplacian on the resulting com binatorial or\nmetric graph has a discontinuous IDS. In fact, the set of jumps ca n be characterised\nrather explicitely and is dense in the spectrum [CCF+86, Ves05, GLV08]. Now the\nquestion is, where the eigenvalues of these strongly localised state s repell in some\nmanner absolutely continuous spectrum (if it exists at all).\nFor the interested reader we provide here references to textbo ok accounts of the\nissuesdiscussed above. Theyconcernthemoreclassical modelson L2(Rd)andℓ2(Zd),\nrather than quantum graphs. In [Ves07] one can find a detailed disc ussion and proofs\nof the approximability of the IDS by its finite volume analogues and of W egner\nestimates. The survey article [KM07] is devoted to the IDS in gener al, while the\nmultiscale proof of localisation is exposed in the monograph [Sto01]. T he theory of\nrandom Schr¨ odinger operators is presented from a broader per spective in the books\n[CL90, PF92] and in the summer school notes [Kir89, Kir07].\nAcknowledgements. The second author is grateful for the kind invitation to the\nHumboldt University of Berlin which was supported by the SFB 647. NP and OP\nalso acknowledge the financial support of the Technical University Chemnitz.\n2.Basic notions, model and results\nIn the following subsections, we fix basic notions (metric graphs, La placians and\nSchr¨ odinger operators with vertex conditions), introduce the r andom length modelCONTINUITY OF THE IDS ON RANDOM LENGTH METRIC GRAPHS 5\nand state our main results. For general treatments and further references on metric\ngraphs, we refer to [EKK+08].\n2.1.Metric graphs. Since our randommodel concerns aperturbationof the metric\nstructure of a graph, we carefully distinguish between combinatorial ,topological and\nmetric graphs . Acombinatorial graph G= (V,E,∂) is given by a countable vertex\nsetV, a countable set Eof edge labels and a map ∂(e) ={v1,v2}from the edge\nlabels to (unordered) pairs of vertices. If v1=v2, we callealoop. Note that this\ndefinition allows multiple edges, but we only consider locally finite combina torial\ngraphs, i.e., every vertex has only finitely many adjacent edges. A t opological graph\nXis a topologicalmodel of a combinatorial graphtogether with a choic e of directions\non the edges:\nDefinition 2.1. A(directed) topological graph is a CW-complex Xcontaining only\n(countably many) 0- and 1-cells. The set V=V(X)⊂Xof 0-cells is called the\nset of vertices . The 1-cells of Xare called (topological) edges and are labeled by the\nelements of E=E(X) (the(combinatorial) edges ), i.e., for every edge e∈E, there is\na continuous map Φ e: [0,1]−→Xwhose image is the corresponding (closed) 1-cell,\nand Φ e: (0,1)−→Φe((0,1))⊂Xis a homeomorphism. A 1-cell is called a loopif\nΦe(0) = Φ e(1). The map ∂= (∂−,∂+):E−→V×Vdescribes the direction of the\nedges and is defined by\n∂−e:= Φe(0)∈V, ∂ +e:= Φe(1)∈V.\nForv∈Vwe define\nE±\nv=E±\nv(X) :={e∈E|∂±e=v}.\nThe set of all adjacent edges is defined as the disjointunion1\nEv=Ev(X) :=E+\nv(X)·∪E−\nv(X).\nThedegreeof a vertex v∈VinXis defined as\ndegv= degX(v) :=|Ev|=|E+\nv|+|E−\nv|.\nAtopological subgraph Λ is a CW-subcomplex of X, and therefore Λ is itself a topo-\nlogical graph with (possible empty) boundary ∂Λ := Λ∩Λc⊂V(X).\nSinceatopologicalgraphisatopologicalspace, wecanintroduceth espaceC(X)of\nC-valued continuous functions and the associated notion of measur ability. A metric\ngraph is a topological graph where we assign a length to every edge.\nDefinition 2.2. A(directed) metric graph (X,ℓ) is a topological graph Xtogether\nwith alength function ℓ:E(X)−→(0,∞). The length function induces an identifi-\ncation of the interval Ie:= [0,ℓ(e)] with the edge Φ e([0,1]) (up to the end-points of\nthe corresponding 1-cell, which may be identified in Xifeis a loop) via the map\nΨe:Ie−→X,Ψe(x) = Φe/parenleftBigx\nℓ(e)/parenrightBig\n.\n1The disjoint union is necessary in order to obtain twodifferent labels in Ev(X) for a loop.6 D. LENZ, N. PEYERIMHOFF, O. POST, AND I. VESELI ´C\nNotethat every topological graph Xcanbe canonically regarded asa metric graph\nwhereall edges have length one . The corresponding length function /BDE(X)is denoted\nbyℓ0. In our random model, we will consider a fixed topological graph Xwith a\nrandom perturbation ℓωof this length function ℓ0.\nTo simplify matters, we canonically identify a metric graph ( X,ℓ) with the disjoint\nunionXℓof the intervals Iefor alle∈Esubject to appropriate identifications of the\nend-points of these intervals (according to the combinatorial str ucture of the graph),\nnamely\nXℓ:=·/uniondisplay\ne∈EIe/∼.\nThe coordinate maps {Ψe}ecan be glued together to a map\nΨℓ:Xℓ−→X. (2.1)\nRemark2.3.A metric graph is canonnically equpped with a metric and a measure.\nGiven the information about the lenght of edges, each path in Xℓhas a well defined\nlenght. Thedistancebetweentwo arbitrarypoints x,y∈Xℓisdefinedastheinfimum\nof the lenghts of paths joining the two points. The measure on Xℓis defined in the\nfollowing way. For each measurable Λ ⊂Xthe sets Λ ∩ψe(Ie) are measurable\nas well, and are assigned the Lebesgue measure of the preimage ψ−1\ne(Λ∩ψe(Ie)).\nConsequently, we define the volume of Λ by\nvol(Λ,ℓ) :=/summationdisplay\ne∈Eλ/parenleftbig\nψ−1\ne(Λ∩ψe(Ie))/parenrightbig\n(2.2)\nUsing the identification (2.1), we define the function space L2(X,ℓ) as\nL2(X,ℓ) :=/circleplusdisplay\ne∈EL2(Ie), f={fe}ewithfe∈L2(Ie) and\n/bardblf/bardbl2\nL2(X,ℓ)=/summationdisplay\ne∈E/integraldisplay\nIe|fe(x)|2dx.\n2.2.Operators and vertex conditions. For a given metric graph ( X,ℓ), we in-\ntroduce the operator\n(Df)e(x) = (Dℓf)e(x) =dfe\ndx(x),\nwhere the derivative is taken in the interval Ie= [0,ℓ(e)]. Note that both the\nnorm in L2(X,ℓ) andD=Dℓdepend on the length function. This observation is\nparticularly important in our random length model below, where we pe rturb the\ncanonical length function ℓ0= /BDE(X)and therefore have (a priori) different spaces\non which a function flives. Our point of view is that fis a function on the fixed\nunderlying topological graph X, and that the metric spaces are canonically identified\nvia the maps Ψ−1\nℓ0◦Ψℓ: (X,ℓ)−→(X,ℓ0). One easily checks that\n/bardblf/bardbl2\nL2(X,ℓ)=/summationdisplay\ne∈Eℓ(e)/integraldisplay\n(0,1)|fe(x)|2dx, (2.3a)\n(Dℓf)e(x) =1\nℓ(e)(Dℓ0f)e/parenleftBig1\nℓ(e)x/parenrightBig\n, (2.3b)CONTINUITY OF THE IDS ON RANDOM LENGTH METRIC GRAPHS 7\nwherefeandDℓ0fon the right side are considered as functions on [0 ,1] via the\nidentification Ψ−1\nℓ0◦Ψℓ.\nNext we introduce general vertex conditions for Laplacians ∆(X,ℓ)=−Dℓ2and\nSchr¨ odinger operators H(X,ℓ)= ∆(X,ℓ)+qwith real-valued potentials q∈L∞(X).\nThemaximal ordecoupled Sobolev space of order kon (X,ℓ) is defined by\nHk\nmax(X,ℓ) :=/circleplusdisplay\ne∈EHk(Ie)\n/bardblf/bardbl2\nHkmax(X,ℓ):=/summationdisplay\ne∈E/bardblfe/bardbl2\nHk(Ie).\nNote thatDℓ:Hk+1\nmax(X,ℓ)−→Hk\nmax(X,ℓ) is a bounded operator. We introduce the\nfollowing two different evaluation maps H1\nmax(X,ℓ)−→/circleplustext\nv∈VCEv:\nfe(v) :=/braceleftBigg\nfe(0),ifv=∂−e,\nfe(ℓ(e)),ifv=∂+e,andf− →e(v) :=/braceleftBigg\n−fe(0),ifv=∂−e,\nfe(ℓ(e)),ifv=∂+e,\nandf(v) ={fe(v)}e∈Ev∈CEv,f− →(v) ={f− →e(v)}e∈Ev∈CEv. It follows from\nstandard Sobolev estimates (see e.g. [Kuc04, Lem. 8]) that these e valuation maps\nare bounded by max {(2/ℓmin)1/2,1}, provided the minimal edge length\n0<ℓmin:= inf\ne∈Eℓ(e) (2.4)\nis strictly positive. The second evaluation map is used in connection wit h the deriva-\ntiveDfof a function f∈H2\nmax(X,ℓ). Note that Df− →is independent of the orientation\nof the edge.\nAsingle-vertex condition atv∈Vis given by a Lagrangian subspace L(v) of\nthe Hermitian symplectic vector space ( CEv⊕CEv,ηv) with canonical two-form ηv\ndefined by\nηv((x,x′),(y,y′)) :=/a\\}bracketle{tx′,y/a\\}bracketri}ht −/a\\}bracketle{tx,y′/a\\}bracketri}ht,\nwhere/a\\}bracketle{t·,·/a\\}bracketri}htdenotes the standard unitary inner product in CEv. The set of all La-\ngrangian subspaces of ( CEv⊕CEv,ηv) is denoted by Lvand has a natural manifold\nstructure (see, e.g., [Har00, KS99] for more details on these notio ns). A Lagrangian\nsubspaceL(v) can uniquely be described by the pair ( Q(v),R(v)) whereQ(v) is an\northogonal projection in CEvwith range G(v) := ranQ(v) andR(v) is a symmetric\noperator on G(v) such that\nL(v) :=/braceleftbig\n(x,x′)/vextendsingle/vextendsingle(1−Q(v))x= 0, Q(v)x′=R(v)x/bracerightbig\n(2.5)\n(see e.g. [Kuc04]).\nA field of single-vertex conditions L:={L(v)}v∈Vis called a vertex condition . We\nsay thatLisbounded, if\nCR:= sup\nv∈V/bardblR(v)/bardbl<∞, (2.6)\nwherethenormistheoperatornormon G(v). Foranysuchboundedvertexcondition\nL, a bounded potential qand a metric graph ( X,ℓ) withℓmin>0, we obtain a self-\nadjoint Schr¨ odinger operator H(X,ℓ),L= ∆(X,ℓ),L+q, by choosing the domain\ndomH(X,ℓ),L:={f∈H2\nmax(X,ℓ)|(f(v),Df− →(v))∈L(v) for allv∈V}.8 D. LENZ, N. PEYERIMHOFF, O. POST, AND I. VESELI ´C\nOf particular interest are the following vertex conditions with vanish ing vertex oper-\natorR(v) = 0 for all v∈V:Dirichlet vertex conditions (where L(v) ={0}⊕CEvor\nG(v) ={0}),Kirchhoff (also known as free) vertex conditions (where ( x,x′)∈L(v)\nif all components of xare equal and the sum of all components of x′add up to\nzero, or equivalently G(v) =C(1,...,1)) andNeumann vertex conditions (where\nL(v) =CEv⊕{0}or equivalently G(v) =CEv).\n2.3.Random length model. The underlying geometric structure of a random\nlength model is a random length metric graph. A random length metric graph is\nbased on a fixed topological graph XwithVandEthe sets of vertices and edges\nofX, a probability space (Ω ,P), and ameasurable mapℓ: Ω×E−→(0,∞), which\ndescribes the random dependence of the edge lengths. We also ass ume that there are\nω-independent constants ℓmin,ℓmax>0 such that ℓmin≤ℓω(e)≤ℓmaxfor allω∈Ω\nande∈E. We will use the notation ℓω(e) :=ℓ(ω,e).\nArandom length model associates to such a geometric structure ( X,Ω,P,ℓ) a\nrandom family of Schr¨ odinger operators Hω, by additionally introducing measurable\nmapsL(v): Ω−→Lvfor allv∈V, andq: Ω×X−→R, describing the random\ndependence of the vertex conditions and the potentials of these o perators. We will\nuse the notation Lω:={Lω(v)}v∈Vandqω(x) =q(ω,x). We assume that we have\nconstantsCR,Cpot>0 such that\n/bardblqω/bardbl∞≤Cpotand/bardblRω(v)/bardbl ≤CR (2.7)\nfor almost all ω∈Ω and allv∈V, whereRω(v) is the vertex operator associated to\nLω(v). From (2.7) and the lower length bound (2.4) it follows that the Schr ¨ odinger\noperatorsHω:= ∆ω+qωare self-adjoint and bounded from below by some constant\nλ0∈Runiformly in ω∈Ω (see Lemma 4.1). We call the tuple ( X,Ω,P,ℓ,L,q) a\nrandom length model with associated Laplacians and Schr¨ odinger operators ∆ωand\nHωand underlying random metric graphs ( X,ℓω).\n2.4.Approximation of the IDS via exhaustions. Let us describe the setting,\nfor which our first main result holds.\nAssumption 2.4. Let (X,Ω,P,ℓ,L,q) be a random length model with the following\nproperties:\n(i) The topological graph Xis non-compact and connected with underlying\n(undirected) combinatorial graph G= (V,E,∂). There is a subgroup Γ ⊂\nAut(G), acting freely on Vwith only finitely many orbits. Then Γ acts also\ncanonically on X(but does not necessarily respect the directions) by\nγΦe(x) =/braceleftBigg\nΦγe(x) if∂±(γe) =γ(∂±e),\nΦγe(1−x) if∂±(γe) =γ(∂∓e).\nThis action carries over to Γ-actions on the metric graphs ( X,ℓ0) and (X,ℓω)\nvia the identification (2.1). Note that Γ acts even isometrically on the equi-\nlateral graph ( X,ℓ0) withℓ0= /BDE. We can think of ( X,ℓ0) as acovering of\nthe compact topological graph ( X/Γ,ℓ0).\n(ii) We also assume that Γ acts ergodically on (Ω,P) by measure preserving\ntransformations with the following consistencies between the two Γ -actions\nonXand Ω:CONTINUITY OF THE IDS ON RANDOM LENGTH METRIC GRAPHS 9\nMetric consistency: We assume that\nℓγω(e) =ℓω(γe) (2.8a)\nfor allγ∈Γ,ω∈Ω ande∈E. This implies that for every γ∈Γ, the\nmap\nγ: (X,ℓω)−→(X,ℓγω)\nis an isometry between two (different) metric graphs. Moreover, t he\ninduced operators\nU(ω,γ):L2(X,ℓγ−1ω))−→L2(X,ℓω)\nare unitary.\nOperator consistency: The transformation behaviour of qωandLωis\nsuch that we have for all ω∈Ω,γ∈Γ,\nHω=U(ω.γ)Hγ−1ωU∗\n(ω,γ). (2.8b)\nSuch a random length model ( X,Ω,P,ℓ,L,q) is called a random length cov-\nering model with associated operators Hωand covering group Γ.\nRemark2.5.The simplest random length covering model is given when the probabil-\nity space Ω consists of only one element with probability 1. In this case , we have only\none length function ℓ=ℓω, one vertex condition L=Lω, and one potential q=qω.\nThe corresponding family of operators consists then of a single ope ratorH=Hω.\nMoreover, the metric consistency means that Γ acts isometrically o n (X,ℓ), and the\noperator consistency is nothing but the periodicity of H, i.e., the property that H\ncommutes with the induced unitary Γ-action on L2(X,ℓ).\nNext, weintroducesomemorenotation. Let F0bearelativelycompacttopological\nfundamental domain of the Γ-action on ( X,ℓ0) such that its closure F=F0is a\ntopological subgraph. (An example of such a topological fundamen tal domain is\ngiven in Figure 2 (a) below.) There is a canonical spectral distribution function\nN(λ), associated to the family Hω, given by the trace formula\nN(λ) :=1\nE(vol(F,ℓ•))E(tr•[ /BDFP•((−∞,λ])]), (2.9)\nwhereE(·) denotes the expectation in (Ω ,P), trωis the trace on the Hilbert space\nL2(X,ℓω), andPω(I)denotesthespectralprojectionassociatedto Hωandtheinterval\nI⊂R. Moreover, the volume vol( F,ℓ•) is defined in (2.2). The function Nis called\nthe(abstract) integrated density of states with abbreviation IDS.\nIn the case of an amenable group Γ the abstract IDS can also be obt ained via\nappropriate exhaustions. This is the statement of Theorem 2.6 belo w. A discrete\ngroup Γ is called amenable , if there exist a sequence In⊂Γ of finite, non-empty\nsubsets with\nlim\nn→∞|In∆Inγ|\n|In|= 0,for allγ∈Γ. (2.10)\nA sequence Insatisfying (2.10) is called a Følner sequence .10 D. LENZ, N. PEYERIMHOFF, O. POST, AND I. VESELI ´C\nFor every non-empty finite subset I⊂Γ, we define Λ( I) :=/uniontext\nγ∈IγF. A sequence\nIn⊂Γ of finite subsets is Følner if and only if the associated sequence Λ n= Λ(In)\nof topological subgraphs satisfies the van Hove condition\nlim\nn→∞|∂Λ(In)|\nvol(Λ(In),ℓ0)= 0. (2.11)\nTheproofofthisfactisanalogoustotheproofof[PV02, Lemma 2.4] intheRiemann-\nian manifold case. Note that (2.11) still holds if we replace ∂Λ(In) by∂rΛ(In) for any\nr≥1, where∂rΛ denotes the thickened combinatorial boundary {v∈V|d(v,∂Λ)≤\nr}andddenotes the combinatorial distance which agrees (on the set of vertices)\nwith the distance function of the unilateral metric graph ( X,ℓ0).\nA Følner sequence Inis called tempered, if we additionally have\nsup\nn∈N|/uniontext\nk≤nIn+1I−1\nk|\n|In+1|<∞. (2.12)\nTempered Følner sequences are needed for an ergodic theorem of Lindenstrauss\n[Lin01]. This ergodic theorem plays a crucial role in the proof of Theor em 2.6 pre-\nsented below. However, the additional property (2.12) is not very restrictive since it\nwas also shown in [Lin01] that every Følner sequence Inhas a tempered subsequence\nInj.\nForany compact topological subgraph Λ of X, we denote theoperator with Dirich-\nlet vertex conditions on the boundary vertices ∂Λ and with the original vertex con-\nditionsLω(v) on all inner vertices v∈V(Λ)\\∂Λ byHΛ,D\nω. The label D refers to\nthe Dirichlet conditions on ∂Λ. For a precise definition of the Dirichlet operator\nvia quadratic forms, we refer to Section 4. The spectral project ion corresponding to\nHΛ,D\nωis denoted by PΛ,D\nω. It is well-known that compactness of Λ implies that the\noperatorHΛ,D\nωhas purely discrete spectrum. The normalised eigenvalue counting\nfunction associated to the operator HΛ,D\nωis defined as\nNΛ\nω(λ) =1\nvol(Λ,ℓω)trω[PΛ,D\nω((−∞,λ])].\nThe function NΛ\nωis the distribution function of a (unique) pure point measure which\nwe denote by µΛ\nω.\nIf Λ = Λ(In) is associated to a Følner sequence In⊂Γ, we use the abbreviations\nHn,D\nω:=HΛ(In),D\nωfor the Schr¨ odinger operator with Dirichlet conditions on ∂Λ(In),\nNn\nω:=NΛ(In)\nωfor the normalised eigenvalue counting function and µn\nω:=µΛ(In)\nωfor\nthe corresponding pure point measure on Λ( In). We can now state our first main\nresult:\nTheorem 2.6. Let(X,Ω,P,ℓ,L,q)be a random length covering model as described\nin Assumption 2.4 with amenable covering group Γ. LetNbe the IDS of the operator\nfamilyHω. Then there exist a subset Ω0⊂Ωof fullP-measure such that we have,\nfor every tempered Følner sequence In⊂Γ,\nlim\nn→∞Nn\nω(λ) =N(λ)\nfor allω∈Ω0and all points λ∈Rat whichNis continuous.\nThe proof is given in Section 4.CONTINUITY OF THE IDS ON RANDOM LENGTH METRIC GRAPHS 11\nRemark2.7.The proof of Theorem 2.6 yields even more. Let µdenote the measure\nassociated to the distribution function N. Then we have\nlim\nj→∞µn\nω(f) =µ(f) (2.13)\nfor allω∈Ω0and all functions fof the form f(x) =g(x)(x+1)−1with a function\ngcontinuous on [0 ,∞) and with limit at infinity. (The behaviour of g(x) forx<0\nis of no importance since the spectral measures of all operators u nder consideration\nare supported on R+= [0,∞).)\n2.5.Wegner estimate. In this subsection, we state a linear Wegner estimate for\nLaplace operators of a random length model with independently dist ributed edge\nlengths and fixed Kirchhoff vertex conditions. This Wegner estimate is linear both\nin the number of edges and in the length of the considered energy int erval. As\nmentioned in the introduction, a similar result for the case Zdwas proved recently\nby different methods in [KP09]. In contrast to the previous subsect ion, we do not\nrequire periodicity of the graph Xassociated to a group action. More precisely, we\nassume the following:\nAssumption 2.8. Let (X,Ω,P,ℓ,L,q) be a random length model with the following\nproperties:\n(i) We have q≡0, i.e., the random family of operators are just the Laplacians\n(Hω= ∆ω)andwe have norandomness inthevertex conditionby fixing Lto\nbe Kirchhoff in all vertices. Thus it suffices to look at the tuple ( X,Ω,P,ℓ).\n(ii) We have a uniform upper bound dmax<∞on the vertex degrees deg v,\nv∈V(X).\n(iii) Since the only randomness occurs in the edge lengths satisfying\n0<ℓmin≤ℓω(e)≤ℓmaxfor allω∈Ω ande∈E(X),\nwe think of the probability space Ω as a Cartesian product/producttext\ne∈E[ℓmin,ℓmax]\nwith projections Ω ∋ω/ma√sto→ωe=ℓω(e)∈[ℓmin,ℓmax]. The measure Pis\nassumed to be a product/circlemultiplytext\ne∈EPeof probability measures Pe. Moreover,\nfor everye∈E, we assume that Peis absolutely continuous with respect\nto the Lebesgue measure on [ ℓmin,ℓmax] with density functions he∈C1(R)\nsatisfying\n/bardblhe/bardbl∞,/bardblh′\ne/bardbl∞≤Ch, (2.14)\nfor a constant Ch>0 independent of e∈E.\nRecall that tr ωis the trace in the Hilbert space L2(Λ,ℓω). In the next theorem\nPΛ,D\nωdenotes the spectral projection of the Laplacian ∆Λ,D\nωon (Λ,ℓω) with Kirchhoff\nvertex conditions on all interior vertices and Dirichlet boundary con ditions on∂Λ.\nUnder these assumptions we have:\nTheorem 2.9. Let(X,Ω,P,ℓ)be a random length model satisfying Assumption 2.8.\nLetu>1andJu= [1/u,u]. Then there exists a constant C >0such that\nE(trPΛ,D\n•(I))≤C·λ(I)·|E(Λ)|\nfor all compact subgraphs Λ⊂Xand all compact intervals I⊂Ju, whereλ(I)\ndenotes the Lebesgue-measure of I, and where |E(Λ)|denotes the number of edges in12 D. LENZ, N. PEYERIMHOFF, O. POST, AND I. VESELI ´C\nΛ. The constant C >0depends only the constants u,dmax,ℓmin,ℓmaxand the bound\nCh>0associated to the densities he(see(2.14)).\nThe proof will be given in Section 5. We finish this section with the followin g\ncorollary. Recall that the periodic situation is a special case of a ran dom length\ncovering model (see Remark 2.5):\nCorollary 2.10. Let(X,Ω,P,ℓ)be a random length covering model, satisfying both\nAssumptions 2.4 and 2.8, with amenable covering group Γ. Then the IDS Nof\nthe Laplacians ∆ωis a continuous function on Rand even Lipschitz continuous on\n(0,∞).\nProof.The Lipschitz continuity of Non (0,∞) follows immediately from Theo-\nrems 2.6 and 2.9. It remains to prove continuity of Non (−∞,0]. Note that our\nmodel is a special situation of the general ergodic groupoid setting given in [LPV07].\nThus,Nis the distribution function of a spectral measure of the direct inte gral op-\nerator/integraltext⊕\nΩ∆ωdP(ω). Since ∆ω≥0 for allω,N(λ) vanishes for all λ<0. Moreover,\nifNwould have a jump at λ= 0, then ker∆ωwould be non-trivial for almost all\nω∈Ω. But ∆ωf= 0 implies\n0 =/a\\}bracketle{tf,∆ωf/a\\}bracketri}ht=/integraldisplay\nX/vextendsingle/vextendsingle/vextendsingledf\ndx(x)/vextendsingle/vextendsingle/vextendsingle2\ndx\nsince ∆ωhas Kirchhoff vertex conditions. Thus fis a constant function. Now Xis\nconnected as well as non-compact, which implies that vol( X,ℓω) =∞by the lower\nboundℓminon the lengths of the edges. Hence constant functions are not in L2. This\ngives a contradiction. /square\nOur result on Lipschitz continuity of Non (0,∞) is optimal in the following sense:\nRemark 2.11.It is well-known that the IDS of the free Laplacian ∆RonRis pro-\nportional to the square root of the energy. Note that this does n ot change when\nadding Kirchhoff boundary conditions at arbitrary points. Therefo re, every model\nsatisfying Assumptions 2.4 and 2.8 for a metric graph isometric to Rhas in fact the\nabove IDS. Therefore, we cannot expect Lipschitz continuity of t he IDS at zero for\nrandom length models without further assumptions.\n3.Kagome lattice as an example of a planar graph\nIn this section, we illustrate the concepts of the previous section f or an explicit\nexample. We introduce a particular regular tessellation of the Euclide an plane ad-\nmitting finitely supported eigenfunctions of the combinatorial Lapla cian. We discuss\nin detail the discontinuities of the IDS of the combinatorial Laplacian and of the\nKirchhoff Laplacian of the induced equilateral metric graph. On the o ther hand, ap-\nplying Corollary2.10, we see that theIDSofa randomfamilyof Kirchho ff Laplacians\nfor independent distributed edge lengths is continuous. Thus, ran domness leads to\nan improvement of the regularity of the IDS in this example.\nWe consider the infinite planar topological graph X⊂Cas illustrated in Figure 1.\nThis graph is sometimes called Kagome lattice . Every vertex of Xhas degree four\nand belongs to a uniquely determined upside triangle. Introducing w1= 1 and\nw2= eπi/3, we can identify the lower left vertex of a particular upside triangle w ithCONTINUITY OF THE IDS ON RANDOM LENGTH METRIC GRAPHS 13\nthe origin in Cand its other two vertices with w1,w2∈C. Consequently, the vertex\nset ofXis given explicitly as the disjoint union of the following three sets:\nV(X) = (2Zw1+2Zw2)·∪(w1+2Zw1+2Zw2)·∪(w2+2Zw1+2Zw2).\nA pairv1,v2∈V=V(X) of vertices is connected by a straight edge if and only if\n|v2−v1|= 1. We write v1∼v2for adjacent vertices. The above realisation of the\nplanar graph X⊂Cis an isometric embedding of the metric graph ( X,ℓ0).\nPSfrag replacements w1w2\nFigure 1. Illustration of the planar graph X(Kagome lattice).\nThe group Z2acts onXvia the maps Tγ(x) := 2γ1w1+2γ2w2+x. A topological\nfundamental domain F0ofXis thickened in Figure 2 (a). The set of vertices of the\ntopological subgraph F=F0(obtained by taking the closure of F0considered as a\nsubset of the metric space ( X,ℓ)) is given by {a,b,c,a′,b′,b′′,c′′}.\nNote that we have to distinguish carefully between a topological and a combina-\ntorial fundamental domain. Let Gdenote the underlying combinatorial graph with\nsetVof vertices and Eof combinatorial edges. The maps Tγact also on the set of\nverticesVand a combinatorial fundamental domain is given by Q={a,b,c}. We\ndenote the translates Tγ(Q) ofQbyQγ.PSfrag replacements\n2w12w2\nab c\na′b′\nb′′c′′b′′′\n(a) (b)v1v2v3\nv4v5\nw1\nw2Hγ0\nFigure 2. (a) The periodic graph with thickened topological fun-\ndamental domain F0and combinatorial fundamental domain Q=\n{a,b,c}(b)Ifγ0isvertically extremal for F, allwhiteencircled vertices\nare zeroes of F.14 D. LENZ, N. PEYERIMHOFF, O. POST, AND I. VESELI ´C\n3.1.Spectrum and IDS of the combinatorial Laplacian. We first observe that\nGadmits finitely supported eigenfunctions of the combinatorial Laplacian ∆comb:\nChoose an arbitrary hexagon H⊂Xwith vertices {u0,u1,...,u 5}. Then there\nexists a centre w0∈CofHsuch that we have\n{u0,u1,...,u 5}={w0+ekπi/3|k= 0,1,...,5}.\nThe following function FH:V−→ {0,±1}on the vertices\nFH(v) :=/braceleftBigg\n0,ifv∈V\\{u0,...,u 5},\n(−1)k,ifv=w0+ekπi/3,(3.1)\nsatisfies\n∆combFH(v) =1\ndeg(v)/summationdisplay\nw∼v(FH(v)−FH(w)) =3\n2FH(v).\nThus, the vertices of every hexagon H⊂Xare the support of a combinatorial\neigenfunction FH:V−→R. The functions FHare the only finitely supported eigen-\nfunctions up to linear combinations:\nProposition 3.1. (a) LetF:V−→Rbe a combinatorial eigenfunction on Xwith\nfinite support suppF⊂V. Then\n∆combF=3\n2F\nandFis a linear combination of finitely many eigenfunctions FHof the above\ntype(3.1).\n(b) LetHi(i= 1,...,k)be a collection of distinct, albeit not necessarily disjoin t,\nhexagons, and Fi:=FHithe associatedcompactlysupported eigenfunctions. Then\nthe setF1,...,F kis linearly independent.\n(c) Ifg∈ℓ2(V)satisfies∆combg=µg, thenµ= 3/2.\n(d) The space of ℓ2(V)-eigenfunctions to the eigenvalue 3/2is spanned by compactly\nsupported eigenfunctions.\nProof.To prove (a), assume that F:V−→Ris a finitely supported eigenfunction.\nLetQ={a,b,c}be a combinatorial fundamental domain of Z2, as illustrated in\nFigure 2 (a) and Qγ:=Tγ(Q). LetHγbe the uniquely defined hexagon containing\nthe three vertices Qγ. Moreover, we define\nA0:={γ∈Z2|suppF∩Qγ/\\e}atio\\slash=∅}.\nLetε1= (1,0) andε2= (0,1). We say that γ0= (γ01,γ02)∈A0isvertically extremal\nforF, if the second coordinate γ02is maximal amongst all γ∈A0and ifγ0−ε1/∈A0.\nThis means that Fvanishes in the left neighbour of Qγ0and in all vertices vertically\naboveQγ0. Hence,γ0in Figure 2 (b) is vertically extremal if Fvanishes in all white\nencircled vertices and does not vanish in at least one of the black ver tices. Obviously,\nA0has always vertically extremal elements. Choosing such a γ0∈A0, we will show\nbelow that Fis an eigenfunction with eigenvalue 3 /2 and that the following facts\nhold:\n(i)γ0+ε1belongs toA0,\n(ii)γ0−ε2orγ0−ε2−ε1belong toA0,CONTINUITY OF THE IDS ON RANDOM LENGTH METRIC GRAPHS 15\n(iii) adding a suitable multiple of FHγ0toF, we obtain a new eigenfunction F1\nand a setA1:={γ∈Z2|suppF1∩Qγ/\\e}atio\\slash=∅}satisfying\nγ0/∈A1, A1\\A0⊂ {γ0−ε2,γ0+ε1−ε2}.\nTo see this, let γ0∈A0be vertically extremal and v1,...,v 5,w1,w2be chosen as in\nFigure 2 (b). The eigenvalue equation at the vertices v4andv5, in whichFvanishes,\nimply that we have F(v1) =−F(v2) =F(v3)/\\e}atio\\slash= 0. Applying the eigenvalue equation\nagain, now at v2, yields that the eigenvalue of Fmust be 3/2.\nIfγ0+ε1/∈A0,Fwould vanish in w1and all its neighbours, except for v3. This\nwould contradict to the eigenvalue equation at w1and (i) is proven. Similarly, if\nγ0−ε2,γ0−ε2−ε1/∈A0, we would obtain a contradiction to the eigenvalue equation\nat the vertex w2. This proves (ii).\nBy addingF(v1)FHγ0toF, we obtain a new eigenfunction F1(again to the eigen-\nvalue 3/2) which vanishes at all vertices of Qγ0={v1,v2,v3}. Thus we have γ0/∈A1.\nButFandF1differ only in the vertices Qγ0,Qγ0+ε1,Qγ0−ε2andQγ0+ε1−ε2, estab-\nlishing property (iii).\nThe above procedure can be iteratively (from left to right) applied t o the hexagons\nin the top row of A0: Step (iii) can be applied to the function F1and a vertically\nextremal element of A1. After a finite number nof steps the top row of hexagons in\nA0is no longer in the support of the function Fn. (Note that property (i) implies\nthat when removing the penultimate hexagon form the right, one ha s simultaneously\nremoved the rightermost one, too.) Again, this procedure can be it erated removing\nsuccessively rows of hexagons. This time property (ii) guarantees that the procedure\nstops after a finite number Nof steps with FN≡0. We have proven statement (a).\nNow we turn to the proof of (b). Since the graph is connected ther e exists a vertex\nvinA:=∪k\ni=1Hiwhich is adjacent to some vertex outside A. Thenvis contained in\nprecisely one hexagon Hi0. (In the full graph each vertex is in two hexagons.) Thus\nthe condition\nk/summationdisplay\ni=1αiFi= 0αi∈C (3.2)\nevaluated at the vertex vimpliesαi0= 0. This shows that all coefficients αiin (3.2)\ncorresponding to hexagons Hilying at the boundary of Avanish. This leads to an\nequation analogous to (3.2) where the indices in the sum run over a st rict subset of\n{1,...,k}. Now one iterates the pocedure and shows that actually all coefficie nts\nα1,...,α kin (3.2) are zero. We have shown linear independence of F1,...,F k.\nTo prove (c) we recall that the IDS ∆ combis a spectral measure (see e.g. [LPV07,\nProp. 5.2]). Thus the IDS jumps at the value µ. This in turn implies by [Ves05,\nProp. 5.2] that there is a compactly supported ˜ gsatisfying the eigenvalue equation.\nNow (a) implies µ= 3/2.\nStatement (d) follows from [LV08, Thm. 2.2], cf. also the proof of Pr oposition 3.3.\n/square\nWe are primarily interested in ℓ2-eigenfunctions of ∆comb, since their eigenvalues\ncoincide with the discontinuities of the corresponding IDS. For combinatorial cov-\nering graphs with amenable covering group Γ, every ℓ2-eigenfunction Fimplies the16 D. LENZ, N. PEYERIMHOFF, O. POST, AND I. VESELI ´C\nexistence of a finitely supported eigenfunction to the same eigenvalue which is im-\nplied, e.g., by [Ves05, Prop. 5.2] or [LV08, Thm. 2.2]. (Related, but diffe rent results\nhave been obtained before in [MY02]. If the group is even abelian, as is the case\nfor the Kagome lattice, the analogous result was proven even earlie r in [Kuc91].) It\nshould be mentioned here that the situation is very different in the sm ooth cate-\ngory of Riemannian manifolds. There, compactly supported eigenfu nctions cannot\noccur due to the unique continuation principle . In the discrete setting of graphs,\nnon-existence of finitely supported combinatorial eigenfunctions is — at present —\nonly be proved for particular examples or in the case of planar graph s of non-positive\ncombinatorial curvature; see [KLPS06] for more details. Hence, P roposition 3.1 tells\nus thatXdoes not admit combinatorial ℓ2-eigenfunctions associated to eigenvalues\nµ/\\e}atio\\slash= 3/2.\nNext, let us discuss spectral informations which can be obtained wit h the help of\nFloquet theory . Using a general result of Kuchment (see [Kuc91] or [Kuc05, Thm. 8 ])\nfor periodic finite difference operators (applying Floquet theory to such operators)\nwe conclude that the compactly supported eigenfunctions of ∆combassociated to the\neigenvalue 3 /2 are already dense in the whole eigenspace ker(∆comb−3/2). As for\nthe whole spectrum, we derive the following result:\nProposition 3.2. Denote by σac(∆comb)andσp(∆comp)the absolutely continuous\nand point spectrum of ∆combon ourZ2-periodic graph X. Then we have\nσac(∆comb) =/bracketleftBig\n0,3\n2/bracketrightBig\nandσp(∆comb) =/braceleftBig3\n2/bracerightBig\n.\nThe proof follows from standard Floquet theory (for a similar hexag onal graph\nmodel see [KP07]):\nProof.Note that we have the unitary equivalence\n∆comb∼=/integraldisplay⊕\nT2∆θ\ncombdθ,\nwhere ∆θ\ncombis theθ-equivariant Laplacian on Q,θ∈T2:=R2/(2πZ)2. This\noperator is equivalent to the matrix\n∆θ\ncomb∼=1\n4\n4−1−e−iθ2−e−iθ1−e−iθ2\n−1−eiθ24 −1−e−iθ1\n−eiθ1−eiθ2−1−eiθ1 4\n\nusingthebasis F∼=(F(a),F(b),F(c))forafunctionon Qandthefactthat F(Tγv) =\nei/a\\gbracketleftθ,γ/a\\gbracketrightF(v) (equivariance). The characteristic polynomial is\np(µ) =/parenleftBig\nµ−3\n2/parenrightBig/parenleftbigg/parenleftBig\nµ−3\n4/parenrightBig2\n−3+2κ\n16/parenrightbigg\n,\nwhereκ= cosθ1+cosθ2+cos(θ1−θ2), and the eigenvalues of ∆θ\ncombare\nµ1=3\n2andµ±=3\n4±1\n4√\n3+2κ.CONTINUITY OF THE IDS ON RANDOM LENGTH METRIC GRAPHS 17\nIn particular, we recover the fact that ∆combhas an eigenfunction, since µ1is inde-\npendent of θ, onlyµ±depend onθviaκ=κ(θ). Note that we have\n−3\n2=κ/parenleftBig2π\n3,4π\n3/parenrightBig\n≤κ(θ)≤κ(0,0) = 3,\ngiving the spectral bands B−= [0,3/4] andB+= [3/4,3/2]. /square\nThe next result discusses (dis)continuity properties of the IDS as sociated to the\ncombinatorial Laplacian on X:\nProposition 3.3. LetNcombbe the (abstract) IDS of the Z2-periodic operator ∆comb,\ngiven by\nNcomb(µ) =1\n|Q|tr[ /BDQPcomb((−∞,µ])],\nwheretris the trace on the Hilbert space ℓ2(V)andPcombdenotes the spectral pro-\njection of ∆comb. ThenNcombvanishes on (−∞,0], is continuous on R\\{3/2}and\nhas a jump of size 1/3atµ= 3/2. Moreover, Ncombis strictly monotone increasing\non[0,3/2]andNcomb(µ) = 1forµ≥3/2.\nProof.The following facts are given, e.g., in [MY02, p. 119]:\n(i) the points of increase of Ncombcoincide with the spectrum σ(∆comb) and\n(ii)Ncombcan only have discontinuities at σp(∆comb).\nTogether with Proposition 3.2, all statements of the proposition fo llow, except for\nthe size of the jump at µ= 3/2.\nLet us choose a Følner sequence In⊂Z2and define Λ n=/uniontext\nγ∈InQγ. Let∂Λn\ndenote the set of boundary vertices of the combinatorial graph in duced by the vertex\nset Λn, and\n∂rΛn:={v∈V(X)|d(v,∂Λn)≤r} (3.3)\nbe the thickened (combinatorial) boundary. Let\nD(µ) :=Ncomb(µ)−lim\nε→0Ncomb(µ−ε) =1\n|Λn|tr/bracketleftbig/BDΛnPcomb({µ})/bracketrightbig\n.(3.4)\nThe last equality in (3.4) holds for all nand follows easily from the Z2-invariance of\nthe operator ∆comb. It remains to prove that D(3/2) = 1/3. Let Λ′\nn= Λn\\∂1Λnand\nDn(µ) :=1\n|Λn|dimEn(µ),\nwhereEn(µ) :={F∈ker(∆comb−µ)|suppF⊂Λ′\nn}. Arguments as in [MSY03] or\nin [LV08] show that\nD(µ) = lim\nn→∞Dn(µ). (3.5)\nFor the convenience of the reader, we outline the proof of (3.5) be low. Using part\n(b) of Proposition 3.1 one can show that dim En(µ) equals up to a boundary term\nthe number of hexagons contained in Λ′\nn. Since every translated combinatorial fun-\ndamental domain Qγuniquely determines a hexagon Hγand|Q|= 3, we conclude\nthat dimEn(µ)≈1\n3|Λn|, up to an error proportional to |∂1Λn|. The van Hove prop-\nerty (2.11) (which holds also in the combinatorial setting) then implies the desired\nresultD(3/2) = lim n→∞Dn(3/2) = 1/3.18 D. LENZ, N. PEYERIMHOFF, O. POST, AND I. VESELI ´C\nFinally, we outline the proof of (3.5): Let E(µ) = ker(∆comb−µ) andSn(µ) =/BDΛnE(µ). Letbn:Sn(µ)−→R|∂1Λn|betheboundary map, i.e., bn(F) isthecollection\nof all values of Fassumed at the (thickened) boundary vertices ∂1Λn. Then kerbn=\nEn(µ)⊂Sn(µ), and we have\nDn(µ)≤D(µ)≤dimSn(µ)\n|Λn|=dimkerbn\n|Λn|+dimranbn\n|Λn|≤Dn(µ)+|∂1Λn|\n|Λn|,\nwhich yields (3.5), by taking the limit, as n→ ∞. /square\n3.2.Spectrum and IDS of the periodic Kirchhoff Laplacian. There is a well\nknown correspondence between the spectrum σ(∆comb) on a graph Gand the spec-\ntrumofthe(Kirchhoff)Laplacian∆0onthecorresponding (equilateral)metricgraph\n(X,ℓ0) withℓ0= /BDE(see e.g. [vB85, Nic85, Cat97, BGP08, Pos08] and the references\ntherein). Namely, any λ/\\e}atio\\slash=k2π2lies inσp(∆0) resp.σac(∆0) iffµ(λ) = 1−cos√\nλ\nlies inσp(∆comb) resp.σac(∆comb). Moreover, the eigenspace of the metric Laplacian\nis isomorphic to the corresponding eigenspace of the combinatorial Laplacian.\nLetF:V−→Cbe a finitely supported eigenfunction of ∆combas in the previous\nsection. In particular, the eigenvalue must be µ= 3/2. The above mentioned\ncorrespondence shows that, for every λ= (2k+2/3)2π2,k∈Z, (i.e.µ(λ) = 3/2)),\nthere is a Kirchhoff eigenfunction f:X−→Rof compact support associated to the\neigenvalueλ, satisfying f(v) =F(v) at all vertices v∈V. In addition, if λ=k2π2,\nthere are so-called Dirichlet eigenfunctions of ∆0, determined by the topology of the\ngraph (see e.g. [vB85, Nic85, Kuc05, LP08]), which are also generat ed by compactly\nsupported eigenfunctions.\nUsing the results [Cat97, BGP08], we conclude from Proposition 3.2:\nCorollary 3.4. Let∆0denote the Kirchhoff Laplacian of the equilateral metric gra ph\n(X,ℓ0). Letσpandσacdenote the point spectrum and absolutely continuous spectr um\nandσcompdenote the spectrum given by the compactly supported eigenf unctions. Then\nwe have\nσcomp(∆0) =σp(∆0) =/braceleftBig/parenleftBig\n2k+2\n3/parenrightBig2\nπ2/vextendsingle/vextendsingle/vextendsinglek∈Z/bracerightBig\n∪/braceleftbig\nk2π2/vextendsingle/vextendsinglek∈N/bracerightbig\nand\nσac(∆0) =/bracketleftBig\n0,/parenleftBig2\n3/parenrightBig2\nπ2/bracketrightBig\n∪/uniondisplay\nk∈N/bracketleftBig/parenleftBig\n2k−2\n3/parenrightBig2\nπ2,/parenleftBig\n2k+2\n3/parenrightBig2\nπ2/bracketrightBig\n. (3.6)\nSimilarly, as inthe discrete setting, we conclude thefollowing (dis)con tinuity prop-\nerties of the IDS:\nProposition 3.5. LetN0be the (abstract) IDS of the Z2-periodic Kirchhoff Lapla-\ncian∆0on the metric graph (X,ℓ0), given by\nN0(λ) =1\nvol(F,ℓ0)tr[ /BDFP0((−∞,λ])],\nwheretris the trace on the Hilbert space L2(X,ℓ0)andP0denotes the spectral pro-\njection of ∆0. Then all the discontinuities of N0:R−→[0,∞)are\n(i) atλ= (2k+2\n3)2π2,k∈Z, with jumps of size1\n6,\n(ii) atλ=k2π2,k∈N, with jumps of size1\n2.CONTINUITY OF THE IDS ON RANDOM LENGTH METRIC GRAPHS 19\nMoreover,N0is strictly monotone increasing on the absolutely continuo us spectrum\nσac(∆0)given in (3.6)andN0is constant on the complement of σ(∆0).\nProof.Our periodic situation fits into the general setting given in [LPV07], by choos-\ning the trivial probability space Ω = {ω}with only one element. Proposition 5.2\nin [LPV07] states that N0is the distribution function of a spectral measure for the\noperator∆0. Consequently, discontinuities of N0canonly occur atthe L2-eigenvalues\nof ∆0, and the points of increase of N0coincide with the spectrum σ(∆0), which is\ngiven in Corollary 3.4. Hence, it only remains to prove the statements about the dis-\ncontinuitiesof N0. Weknowfrom[Kuc05, Theorem11]thatthecompactlysupported\neigenfunctions densely exhaust every L2-eigenspace of ∆0.\nLetIn⊂Z2be a Følner sequence. This time, we look at the corresponding\ntopological graphs Λ( In) and their thickened topological boundaries ∂rΛ(In) ={x∈\nX|d(x,∂Λ(In))≤r}, and denote them by Λ nand∂rΛn, respectively. We are\ninterested in the jumps\nD(λ) :=N0(λ)−lim\nε→0N0(λ−ε) =1\nvol(Λn,ℓ0)tr/bracketleftbig/BDΛnP0({λ})/bracketrightbig\n,\nwhere the right hand side is, again, independent of the choice of n. Let Λ′\nnbe the\nclosure of Λ n\\∂1Λnand\nDn(λ) :=1\nvol(Λn,ℓ0)dimEn(λ),\nwithEn(λ) ={f∈ker(∆0−λ)|suppf⊂Λ′\nn}. Arguments analogously to the proof\nof (3.5) yield\nD(λ) = lim\nn→∞Dn(λ). (3.7)\nFor the proof of (3.7), however, we have to define the boundary m ap\nbn:Sn(λ)−→/circleplusdisplay\nv∈∂Λn(C⊕CEv) by (bnf)v:= (f(v),Df− →(v)).\nLetλ= (2k+2/3)2π2,k∈Z. We follow the same arguments as in the proof of\nProposition 3.3. Again, dim En(λ) is equal to the number of hexagons contained in\nΛnup to a boundary term and we have vol( F,ℓ0) = 6 (see Figure 2 (a)). Therefore,\nwe derive that the corresponding jump is of size 1 /6.\nLetλ=k2π2,k∈N. We know from [vB85, Nic85] or from [LP08, Lem. 5.1\nand Prop. 5.2] that the dimension of En(λ) is (up to an error proportional to |∂Λn|)\napproximately equal to\n|E(Λn)|−|V(Λn)| ≈1\n2vol(Λn,ℓ0).\nThis implies that N0has a discontinuity at λ=k2π2of size 1/2. /square\nRemark3.6.Note that Propositions 3.3 and 3.5 hold also for general covering gra phs\nX→X0with amenable covering group Γ and compact quotient X0∼=X/Γ, once we\nhave information about the shape of the support of elementary eig enfunctions (i.e.,\neigenfunctions, which generate the eigenspace by linear combinatio ns and transla-\ntions). In our Kagome lattice example the elementary eigenfunction is supported on20 D. LENZ, N. PEYERIMHOFF, O. POST, AND I. VESELI ´C\na hexagon. For example, the jump of size 1 /3 at the eigenvalue µ= 3/2 in the dis-\ncrete case is the number νof hexagons determined by a combinatorial fundamental\ndomain (ν= 1) divided by the number of vertices in a combinatorial fundamenta l\ndomain ( |Q|= 3).\nIn the metric graph setting, the jump at λ= (2k+ 2/3)2π2is of size 1/6 due to\nthe fact that we have six edges in one topological fundamental dom ain.\nFor the eigenvalues at λ=k2π2(also called topological , see [LP08]) we even have\na precise information for any r-regular amenable covering graph, namely\ndimEn(λ)≈ |E(Λn)|−|V(Λn)| ≈/parenleftBig\n1−2\nr/parenrightBig\n|E(Λn)|=/parenleftBig\n1−2\nr/parenrightBig\nvol(Λn,ℓ0),\nup to an error proportional to |∂Λn|, so that the jump of N0atλis (1−2/r).\n3.3.IDS of associated random length models. Finally, we impose a random\nlength structure ℓ: Ω×E−→[ℓmin,ℓmax] on the edges of ( X,ℓ0) with independently\ndistributed edge lengths, as described in Assumption 2.8. Then Coro llary 2.10 tells\nus that the associated integrated density of states N:R−→[0,∞) is continuous and\neven Lipschitz continuous on (0 ,∞). Hence, all discontinuities occurring for the IDS\nof the Kirchhoff Laplacian on the Z2-periodic graph ( X,ℓ0) disappear by introducing\nthis type of randomness.\n4.Proof of the approximation of the IDS via exhaustions\nIn this section, we prove Theorem 2.6, namely, that the non-rando m integrated\ndensity of states (2.9) can be approximated by suitably chosen nor malised eigenvalue\ncounting functions, for P-almost all random parameters ω∈Ω.\nFor the following considerations, we need the quadratic forms asso ciated to the\nSchr¨ odinger operators. Recall that for each Lagrangian subsp aceLv⊂CEv⊕CEv\ndescribing the vertex condition at v∈Vthere exists a unique orthogonal projection\nQvonCEvwith range Gv:= ranQvand a symmetric operator on Gvsuch that (2.5)\nholds.\nLet Λ⊂Xbe a topological subgraph. The quadratic form associated to the\noperator with vertex conditions given by ( Gv,Rv) at inner vertices V(Λ)\\∂Λ and\nDirichlet conditions at ∂Λ is defined as\ndomhΛ,D=/braceleftbig\nf∈H1\nmax(X,ℓ)/vextendsingle/vextendsinglef(v)∈Gv∀v∈V(Λ)\\∂Λ, f(v) = 0∀v∈∂Λ/bracerightbig\n,\nhΛ,D(f) =/bardblDf/bardbl2\nL2(Λ,ℓ)+/a\\}bracketle{tqf,f/a\\}bracketri}htL2(Λ,ℓ)+/summationdisplay\nv∈V(Λ)/a\\}bracketle{tRvf(v),f(v)/a\\}bracketri}htGv.\nIn particular, if Λ = Xis the full graph, then there is no boundary and h=hXis\nthe quadratic form associated to the operator H=H(X,ℓ),L.\nIfℓmin:= inf eℓ(e)>0,Cpot:=/bardblq/bardbl∞<∞and supv/bardblRv/bardbl=:CR<∞, thenhΛ,D\nis a closed quadratic form with corresponding self-adjoint operato rHΛ,D.CONTINUITY OF THE IDS ON RANDOM LENGTH METRIC GRAPHS 21\nLemma 4.1. For any subgraph ΛofX, the quadratic form hΛ,Dis closed. Moreover,\nthe associated self-adjoint operator HΛ,Dhas domain given by\ndomHΛ,D=/braceleftBig\nf∈H2\nmax(X,ℓ)/vextendsingle/vextendsingle/vextendsinglef(v) = 0∀v∈∂V,\nf(v)∈Gv, QvDf− →(v) =Rvf(v)∀v∈V(Λ)\\∂Λ/bracerightBig\n.\nMoreover,HΛ,Disuniformly bounded from below by −C0whereC0≥0depends only\nonℓ−,CRandCpot, but not on Λ.\nProof.Thefirstassertionfollowsfrom[Kuc04, Thm.17]. The uniformlowerboundis\naconsequence of[Kuc04, Cor.10]wherethelowerboundisgivene xplicitly. Basically,\nthe statements follow from a standard Sobolev estimate of the typ e/vextendsingle/vextendsingle/vextendsingle/summationdisplay\nv/a\\}bracketle{tRvf(v),f(v)/a\\}bracketri}ht/vextendsingle/vextendsingle/vextendsingle≤CR/summationdisplay\nv∈V(Λ)|f(v)|2≤η/bardblDf/bardbl2+Cη/bardblf/bardbl2\nforη>0, whereCηdepends only on η,CRandℓmin. /square\nThe Dirichlet operator will serve as upper bound in the bracketing ine quality (4.1)\nlater on. In order to have a lower bound we introduce a Neumann-ty pe operator\nHΛvia its quadratic form hΛ. Since the vertex conditions can be negative, we have\nto use the boundary condition ( CEv,−CR) instead of a simple Neumann boundary\ncondition ( CEv,0). The quadratic form hΛis defined by\ndomhΛ=/braceleftbig\nf∈H1\nmax(X,ℓ)/vextendsingle/vextendsinglef(v)∈Gv∀v∈V(Λ)\\∂Λ/bracerightbig\n,\nhΛ(f) =/bardblDf/bardbl2\nL2(Λ,ℓ)+/a\\}bracketle{tqf,f/a\\}bracketri}htL2(Λ,ℓ)+/summationdisplay\nv∈V(Λ)\\∂Λ/a\\}bracketle{tRvf(v),f(v)/a\\}bracketri}htGv−CR/summationdisplay\nv∈∂Λ|f(v)|2\nGv.\nNotethattheboundarycondition /tildewideRv=−CRtriviallyfulfillsthenormbound /bardbl/tildewideRv/bardbl ≤\nCR, and therefore by Lemma 4.1, the form hΛis uniformly bounded from below by\nthe same constant −C0ashΛ,D. By adding C0to the (edge) potential qwe may\nassume that w.l.o.g. HX,HΛ,DandHΛare all non-negative for all subgraphs Λ.\nWe can now show the following bracketing result:\nLemma 4.2. LetΛbe a topological subgraph of XandΛ′be the closure of the\ncomplement Λc. Then\nHΛ,D⊕HΛ′,D≥H≥HΛ⊕HΛ′≥0 (4.1)\nin the sense of quadratic forms.\nProof.Itisclearfromtheinclusions {0} ⊂Gv⊂CEvforallboundaryvertices v∈∂Λ\nthat the quadratic form domains fulfil\ndomhΛ,D⊕domhΛ′,D⊂domh⊂domhΛ⊕domhΛ′.\nMoreover, if f=fΛ⊕fΛ′is in the decoupled Dirichlet domain, then\nhΛ,D(fΛ)+hΛ′,D(fΛ′) =h(f)\nsincef(v) = 0 on boundary vertices, if f∈domh, then\nh(f)≥hΛ(fΛ)+hΛ′(fΛ′)22 D. LENZ, N. PEYERIMHOFF, O. POST, AND I. VESELI ´C\nsinceRv≥ −CR. In particular, we have shown the inequality for the quadratic\nforms. /square\nNext, we provide a useful lemma about the spectral shift function of two opera-\ntors. For a non-negative operator Hwith purely discrete spectrum {λk(H)|k≥0}\n(repeated according to multiplicity), the eigenvalue counting funct ion is given by\nn(H,λ) := tr /BD[0,λ)(H) =/vextendsingle/vextendsingle{k≥0|λk(H)≤λ}/vextendsingle/vextendsingle.\nThespectral shift function (SSF) of two non-negative operators H1,H2with purely\ndiscrete spectrum is then defined as\nξ(H1,H2,λ) :=n(H2,λ)−n(H1,λ).\nWe have the following estimate:\nLemma 4.3. Let(X,Ω,P,ℓ)be a random length metric graph (as described in Sub-\nsection 2.3) and Λ⊂Xbe a compact topological subgraph. Let L1,L2be two vertex\nconditions differing in the vertex set Vdiff⊂V(Λ)only, and such that the opera-\ntors∆(Λ,ℓω),Liare non-negative. Let 0≤qbe a bounded measurable potential and\nHi= ∆(Λ,ℓω),Li+q. Then we have\n|ξ(H1,H2,λ)| ≤2/summationdisplay\nv∈Vdiffdegv. (4.2)\nMoreover, if ρ:R+−→Ris a monotone function with ρ′∈L1(R+), then\n/vextendsingle/vextendsingletr[ρ(H1)−ρ(H2)]/vextendsingle/vextendsingle≤2|ρ(∞)−ρ(0)|/summationdisplay\nv∈Vdiffdegv, (4.3)\nwhere the trace is taken in the Hilbert space L2(Λ,ℓω).\nProof.LetD0= domH1∩domH2. Then D0has finite index in dom Hi, bounded\nabove by twice the number of all edges adjacent to vertices v∈Vdiff. This im-\nplies dim(dom Hi/D0)≤2/summationtext\nv∈Vdiffdegv. Inequality (4.2) follows now from [GLV07,\nLemma 9]. The second inequality (4.3) follows readily from Krein’s trace identity\n|trρ(H1)−ρ(H2)| ≤/integraldisplay∞\n0|ρ′(λ)|·|ξ(H1,H2,λ)|dλ. (4.4)\n/square\nThe following uniform resolvent boundedness holds in every random le ngth cover-\ning model:\nLemma 4.4. Let(X,Ω,P,ℓ,L,q)be a random length covering model with covering\ngroupΓ, as described in Assumption 2.4, and λ>0. Then there is a constant Cλ>0\nsuch that we have\ntr(HΛ\nω+λ)−1≤Cλvol(Λ,ℓ0)\nfor all compact subgraphs Λ⊂(X,ℓ)and allω∈Ω.CONTINUITY OF THE IDS ON RANDOM LENGTH METRIC GRAPHS 23\nProof.LetHΛ,0\nωdenote the restriction on Λ with Dirichlet vertex conditions at all\nvertices. ThenHΛ,0\nω=/circleplustext\ne∈E(Λ)He,D\nω, where we identify the edge ewith the topo-\nlogical subgraph consisting of this edge and its end vertices in X. From (4.2) of\nLemma 4.3 we conclude that\n|tr(HΛ,0\nω+λ)−1−(HΛ\nω+λ)−1| ≤4\nλ|E(Λ)|=4\nλvol(Λ,ℓ0).\nSince (He,D\nω+λ)−1is bounded from above by (∆e,D\nω+λ)−1, and since the edges\nare uniformly bounded from above by ℓmax, there is a constant cλ>0 such that\ntr(He,D\nω+λ)−1≤cλfor alle∈E(Λ) andω∈Ω. This implies the desired estimate\nwith constant Cλ= 4λ−1+cλ. /square\nThe proof of Theorem 2.6 will now be given in four lemmata. All of these lemmata\narebasedonagivenrandomlengthcovering model( X,Ω,P,ℓ,L,q)withan amenable\ncovering group Γ and a fixed tempered Følner sequence Inwith associated compact\ntopological graphs Λ n:= Λ(In).\nInthe first lemma, we prove the convergence (2.13) for a special f amilyof functions\nfλassociated to resolvents of the operators. Here, we need to app ly an ergodic\ntheorem of Lindenstrauss [Lin01].\nIn later lemmata we show that the convergence (2.13) carries over to the uniform\nclosure of finite linear combinations of the functions fλ, identify this closure with the\nhelp of the Stone-Weierstrass Theorem, and finally conclude the de sired convergence\nfor characteristic functions /BD[0,λ]at continuity points λ>0 of the IDS.\nLemma 4.5. Letλ >0andfλ: [0,∞)−→R,fλ(x) =1\nx+λ. Then there exists a\nsubsetΩ0⊂Ωof fullP-measure such that\nlim\nn→∞1\nvol(Λn,ℓω)tr[fλ(Hn,D\nω)] =1\nE(vol(F,ℓ•))E(tr[ /BDFfλ(H•)])\nfor allω∈Ω0.\nProof.We first consider a fixed ω∈Ω and a fixed Λ = Λ( In) and suppress the\nparameters ωandnin the notation. Recall the definitions of HΛ,DandHΛwith\nquadratic form domains given below. Let Λ′denote the closure of the complement\nΛcin the metric graph ( X,ℓ). By Lemma 4.2 we have (4.1) in the sense of quadratic\nforms. Since taking inverses is operator monotone, this implies\n(HΛ,D⊕HΛ′,D+λ)−1≤(H+λ)−1≤(HΛ⊕HΛ′+λ)−1\nfor allλ >0. In particular, we obtain inequalities for the following restricted qu a-\ndratic forms: Set ( H+λ)−1\nΛ=pΛ(H+λ)−1iΛ, whereiΛandpΛdenote the canonical\ninclusions and projections between L2(Λ,ℓ) andL2(X,ℓ). Then\n(HΛ,D+λ)−1≤(H+λ)−1\nΛ≤(HΛ+λ)−1. (4.5)\nConsequently, ( H+λ)−1\nΛ−(HΛ,D+λ)−1is non-negative and we have\n0≤trL2(Λ,ℓ)/bracketleftbig\n(H+λ)−1\nΛ−(HΛ,D+λ)−1/bracketrightbig\n≤trL2(Λ,ℓ)/bracketleftbig\nfλ(HΛ)−fλ(HΛ,D)/bracketrightbig\n≤2\nλdmax|∂Λ|,24 D. LENZ, N. PEYERIMHOFF, O. POST, AND I. VESELI ´C\nusing Lemma 4.3, where dmaxis a finite upper bound on the vertex degree of X,\nwhich exists due to the Γ-periodicity of X. Using the van Hove property (2.11) and\nthe estimate\nℓminvol(Λ,ℓ0)≤vol(Λ,ℓω)≤ℓmaxvol(Λ,ℓ0),\nwe conclude that\nlim\nn→∞1\nvol(Λn,ℓω)/parenleftbig\ntr[(Hω+λ)−1\nΛn]−tr[fλ(Hn,D\nω)]/parenrightbig\n= 0. (4.6)\nUsing additivity of the trace and the operator consistency (2.8b), we obtain\ntrL2(Λn,ℓω)(Hω+λ)−1\nΛn=/summationdisplay\nγ∈IntrL2(γF,ℓω)(Hω+λ)−1\nγF=/summationdisplay\nγ∈I−1\nngλ(γω),\nwhere\ngλ(ω) = trL2(F,ℓω)[(Hω+λ)−1\nF] = tr[ /BDFfλ(Hω)]. (4.7)\nSince, by monotonicity (4.5) and Lemma 4.4,\n0≤gλ(ω)≤trL2(F,ℓω)[(HF\nω+λ)−1]≤Cλvol(F,ℓ0),\nweconclude that gλ∈L1(Ω). Now, weargueasintheproofofTheorem 7in[LPV04]:\nApplying Lindenstrauss’ ergodic theorem separately to both expr essions\n1\n|In|/summationdisplay\nγ∈I−1\nngλ(γω) and1\n|In|/summationdisplay\nγ∈I−1\nnvol(F,ℓγω),\nwe conclude that\nlim\nn→∞1\nvol(Λn,ℓω)tr[(Hω+λ)−1\nΛn] =1\nE(vol(F,ℓ•))E(tr[ /BDFfλ(H•)]) (4.8)\nfor almost all ω∈Ω. The lemma follows now immediately from (4.6) and (4.8). /square\nLetusdenote by Ltheset offunctions {x/ma√sto→fλ(x) = (x+λ)−1|λ>0}andby A\nthe/bardbl·/bardbl∞-closure of the linear span of Land the constant function /BD: [0,∞)−→R,/BD(x) = 1. Note that, by monotonicity (4.5) and Lemma 4.4, both express ionsµn\nω(f1)\nandµ(f1) = (E(vol(F,ℓ•)))−1E(g1) (withg1defined in (4.7)) are bounded by a\nconstantK >0, independent of ωandn. Let Ω 0⊂Ω be the set of full P-measure\nfrom Lemma 4.5.\nLemma 4.6. Letω∈Ω0. Setνn=f1·µn\nω(forn∈N) andν=f1·µ. Then we\nhave, for all g∈A,\nlim\nn→∞νn(g) =ν(g).\nProof.By Lemma 4.5 we know that the statement holds for the function g= /BD. We\nnote thatfλ·f1=1\nλ−1(fλ−f1) forλ/\\e}atio\\slash= 1. Thus, by linearity and Lemma 4.5, the\nconvergence holds also for all functions g=fλwithλ>0,λ/\\e}atio\\slash= 1. To deal with the\ncaseλ= 1 note that f1+εconverges to f1uniformly, as ε→0. Thus\n|νn(f1)−νn(f1+ε)| ≤ /bardblf1−f1+ε/bardbl∞νn( /BD)≤Kε.\nAn analogous statement holds for νnreplaced by ν. Thus\n|ν(f1)−νn(f1)| ≤2Kε+/vextendsingle/vextendsingleν(f1+ε)−νn(f1+ε)/vextendsingle/vextendsingle→2Kε, (4.9)CONTINUITY OF THE IDS ON RANDOM LENGTH METRIC GRAPHS 25\nasn→ ∞. Sinceε >0 was arbitrary, we conclude that lim n→∞νn(f1) =ν(f1).\nBy linearity, the convergence statement of the Lemma holds for all functionsgin\nthe linear span of L∪ { /BD}. To show that is holds for all functions in the closure\nA, as well, one uses uniform approximation and an estimate of the same type as\nin (4.9). /square\nThe next lemma identifies the space Aexplicitly:\nLemma 4.7. The function space Acoincides with the set of continuous functions\non[0,∞)which converge at infinity.\nProof.The statement of the lemma is equivalent to A=C([0,∞]), where [0 ,∞] is\nthe one-point-compactification of [0 ,∞). We want to apply the Stone-Weierstrass\nTheorem. Any fλwithλ>0 separates points and /BDis nowhere vanishing in [0 ,∞].\nBy definition Ais a linear space. To show that it is an algebra we use again the\nformulafλ1·fλ2=1\nλ2−λ1(fλ1−fλ2), which shows that fλ1·fλ2∈Aforλ1/\\e}atio\\slash=λ2.\nSince Ais closed in the sup-norm, we can use an approximation as in the proof of\nthe Lemma 4.6 to show f2\nλ∈A. A similar argument shows that the product of two\nlimit points f,gof the linear span of L∪{ /BD}is inA. /square\nWe have established the convergence µn\nω(g)→µ(g) for all functions of the form\ng·f1withg∈A. The following lemma shows that this is sufficient to conclude the\nalmost sure convergence Nn\nω(λ)→N(λ) at continuity points λ, finishing the proof\nof Theorem 2.6. One has only to observe that every continuous fun ction of compact\nsupport on R+= [0,∞) can be written as g·f1, with an element g∈A.\nLemma 4.8. Forn∈N, letρn,ρbe locally finite measures on R+. Then\nlim\nn→∞ρn(g) =ρ(g)\nfor all continuous functions gof compact support implies that\nlim\nn→∞ρn([0,λ]) =ρ([0,λ])\nfor allλ>0which are not atoms of ρ.\nProof.The proof is standard. First note that locally finiteness of ρimplies\nlim\nε→0ρ([λ−ε,λ+ε]) =ρ({λ}) = 0.\nNow choose monotone functions g��\nε,g+\nε∈Cc(R+) satisfying/BD[0,λ−ε]≤g−\nε≤ /BD[0,λ]≤g+\nε≤ /BD[0,λ+ε].\nThen\nρ([0,λ])−ρn([0,λ])≤ρ(g+\nε)−ρ(g−\nε)+ρ(g−\nε)−ρn(g−\nε)\n≤ρ([λ−ε,λ+ε])+ρ(g−\nε)−ρn(g−\nε).\nFor anyδ >0 one can choose ε>0 such that ρ([λ−ε,λ+ε])<δ. Sinceδ >0 was\narbitrary, we have shown ρ([0,λ])≤liminf n→∞ρn([0,λ]). The opposite inequality is\nshown similarly. /square26 D. LENZ, N. PEYERIMHOFF, O. POST, AND I. VESELI ´C\n5.Proof of the Wegner estimate\nThis section is devoted to the proof of Theorem 2.9. Let ( X,Ω,P,ℓ) be a random\nlength model satisfying Assumption 2.8. We first introduce a new mea surable map\nα: Ω×E−→[ω−,ω+] withω−= lnℓmin,ω+= lnℓmax, defined by αω(e) :=α(ω,e) =\nlnℓω(e). The random variables α(·,e),e∈E, are independently distributed with\ndensity functions ge(x) = exhe(ex), and we have\n/bardblg′\ne/bardbl∞≤ℓmax/bardblhe/bardbl∞+ℓ2\nmax/bardblh′\ne/bardbl∞≤(ℓmax+ℓ2\nmax)Ch=:Dh<∞.(5.1)\nThus, we can re-identify Ω with the Cartesian product/producttext\ne∈E[ω−,ω+], and the maps\nα(·,e) are simply projections to the component with index e. The measure Pis\nnow given as the product/circlemultiplytext\ne∈E/tildewidePeof marginal measures /tildewidePewith density functions\nge∈C1(R) satisfying the above estimate (5.1). The advantage of the new “r escaled”\nidentification Ω =/producttext\ne∈E[ω−,ω+] is the following property of the eigenvalues of the\nLaplacian on any compact subgraph (Λ ,ℓω):\nλi(∆Λ,D\nω+s /BD) = e−2sλi(∆Λ,D\nω). (5.2)\nHere, the eigenvalues λiare counted with multiplicity and ω+s /BDdenotes the ele-\nment{ωe+s}e∈E(X)∈Ω. Property (5.2) is an immediate consequence of (2.3b),\nℓω+s /BD(e) = eαω(e)+s= esℓω(e), and the fact that a rescaling of all lengths by a fixed\nmultiplicative constant does not change the domain of the Kirchhoff L aplacian with\nDirichlet boundary conditions on ∂Λ. Property (5.2) is of crucial importance for the\nproof of the Wegner estimate.\nHenceforth, weusethisnewinterpretationofΩandrename /tildewidePebyPe, forsimplicity.\nLet Λ⊂Xbe a compact topological subgraph, λ∈Randε >0. We write\nthe interval Ias [λ−ǫ,λ+ǫ] and start with a smooth function ρ:R−→[−1,0]\nsatisfyingρ≡ −1 on (−∞,−ε], 0≤ρ′≤1/ε,ρ≡0 on [ε,∞). Moreover, we set\nρλ(x) =ρ(x−λ). Then we have/BD[λ−ε,λ+ε](x)≤ρλ(x+2ε)−ρλ(x−2ε) =/integraldisplay2ε\n−2ερ′\nλ(x+t)dt.\nUsing the spectral theorem, we obtain\nPΛ,D\nω([λ−ε,λ+ε]) = /BD[λ−ε,λ+ε](∆Λ,D\nω)≤/integraldisplay2ε\n−2ερ′\nλ(∆Λ,D\nω+t)dt,\nand, consequently,\ntrPΛ,D\nω([λ−ε,λ+ε])≤/integraldisplay2ε\n−2εtrρ′\nλ(∆Λ,D\nω+t)dt.\nDenote by (Ω(Λ) ,PΛ) the space Ω(Λ) =/producttext\ne∈E(Λ)[ω−,ω+] with probability measure\nPΛ=/circlemultiplytext\ne∈E(Λ)Pe, andEΛ(·) denote the associated expectation. E(·) means expecta-\ntion with respect to the full space (Ω ,P). Applying expectation yields\nE(trPΛ,D\n•([λ−ε,λ+ε])) =EΛ(trPΛ,D\n•([λ−ε,λ+ε]))\n≤/integraldisplay\nΩ(Λ)/integraldisplay2ε\n−2εtrρ′\nλ(∆Λ,D\nω+t)dtdPΛ(ω).(5.3)CONTINUITY OF THE IDS ON RANDOM LENGTH METRIC GRAPHS 27\nUsing the chain rule and scaling property (5.2), we obtain\n/summationdisplay\ne∈E(Λ)∂\n∂ωeρλ(λi(∆Λ,D\nω)+t) =ρ′\nλ(λi(∆λ,D\nω)+t)d\nds/vextendsingle/vextendsingle/vextendsingle\ns=0/parenleftbig\ns/ma√sto→λi(∆Λ,D\nω+s /BD)/parenrightbig\n=−2ρ′\nλ(λi(∆λ,D\nω)+t)λi(∆Λ,D\nω)≤0.\nNow, we use that [ λ−ε,λ+ε]⊂Ju= [1/u,u]. Since supp ρ′\nλ⊂[λ−ε,λ+ε], we\nderive\n0≤trρ′\nλ(∆Λ,D\nω+t)≤ −u\n2/parenleftBig/summationdisplay\ne∈E(Λ)∂\n∂ωetrρλ(∆Λ,D\nω+t)/parenrightBig\n. (5.4)\nFore∈E(Λ), denote by Λ ethe topological subgraph with vertex set Ve:=V(Λ)\nand edge set Ee:=E(Λ)\\{e}. Using the estimate (5.4), we obtain from (5.3)\nE(trPΛ,D\n•([λ−ε,λ+ε]))\n≤ −u\n2/summationdisplay\ne∈E(Λ)/integraldisplay\nΩ(Λe)/integraldisplay2ε\n−2ε/integraldisplayω+\nω−/parenleftBig∂\n∂ωetrρλ(∆Λ,D\n(ω′,x)+t)/parenrightBig\nge(x)dxdtdPΛe(ω′) (5.5)\nwith(ω′,x)∈Ω(Λe)×[ω−,ω+] = Ω(Λ). Next, wewanttocarryoutpartialintegration\nwith respect to xin (5.5). Before doing so, it is useful to observe, for fixed c∈\n[ω−,ω+],\n∂\n∂ωetrρλ(∆Λ,D\n(ω′,x)+t) =∂\n∂ωe/parenleftBig\ntrρλ(∆Λ,D\n(ω′,x)+t)−trρλ(∆Λ,D\n(ω′,c)+t)/parenrightBig\n.(5.6)\nUsing (5.6) and applying partial integration, we obtain\n/vextendsingle/vextendsingle/vextendsingle/integraldisplayω+\nω−/parenleftBig∂\n∂ωetrρλ(∆Λ,D\n(ω′,x)+t)/parenrightBig\nge(x)dx/vextendsingle/vextendsingle/vextendsingle\n≤ /bardblg′\ne/bardblL1sup\nc′∈[ω−,ω+]/vextendsingle/vextendsingletrρλ−t(∆Λ,D\n(ω′,c′))−trρλ−t(∆Λ,D\n(ω′,c))/vextendsingle/vextendsingle.(5.7)\nFor notational convenience, we identify the compact topological g raph consisting\nonly of the edge eand its end-points with e, and we denote by ∆e,D\ncbe the Dirichlet-\nLaplacian on the metric graph ( e,ℓc) defined by ℓc(e) = exp(c). Using (4.3) in\nLemma 4.3, we conclude that\n/vextendsingle/vextendsingletrρλ−t(∆Λ,D\n(ω′,c))−trρλ−t(∆Λe,D\nω′⊕∆e,D\nc)/vextendsingle/vextendsingle≤2|ρ(∞)−ρ(t−λ)|2dmax≤4dmax,\nforallvalues c∈[ω−,ω+]. Consequently, sup/vextendsingle/vextendsingletrρλ−t(∆Λ,D\n(ω′,c′))−trρλ−t(∆Λ,D\n(ω′,c))/vextendsingle/vextendsinglein(5.7)\ncan be estimated from above by\n8dmax+/vextendsingle/vextendsingletrρ(∆e,D\nc′+t−λ)−trρ(∆e,D\nc+t−λ)/vextendsingle/vextendsingle.\nNote that all eigenfunctions of the Dirichlet operator ∆e,D\ncare explicitly given sine\nfunctions. Therefore, since λ∈[1/u+ε,u−ε] andt∈[−2ε,2ε], there is a constant\nCu,ℓmax>0, depending only on u,ℓmax, such that\n/vextendsingle/vextendsingletrρ(∆e,D\nc+t−λ)/vextendsingle/vextendsingle≤Cu,ℓmax,28 D. LENZ, N. PEYERIMHOFF, O. POST, AND I. VESELI ´C\nfor all exp(c)∈[ℓmin,ℓmax]. This implies\n/vextendsingle/vextendsingle/vextendsingle/integraldisplayω+\nω−/parenleftbigg∂\n∂ωetrρλ(∆Λ,D\n(ω′,x)+t)/parenrightbigg\nge(x)dx/vextendsingle/vextendsingle/vextendsingle≤(8dmax+2Cu,ℓmax)/bardblg′\ne/bardblL1([ω−,ω+]).\nPlugging this into inequality (5.5), we finally obtain\nE(trPΛ,D\n•([λ−ε,λ+ε]))≤u(4dmax+Cu,ℓmax)Dhlnℓmax\nℓmin4ε|E(Λ)|,\nfinishing the proof of Theorem 2.9.\nReferences\n[AS93] T. Adachi and T. Sunada, Density of states in spectral geometry , Comment. Math. Helv.\n68(1993), no. 3, 480–493.\n[AM93] Michael Aizenman and Stanislav Molchanov, Localization at large disorder and at ex-\ntreme energies: an elementary derivation , Comm. Math. Phys. 157(1993), no. 2, 245–\n278.\n[ASW06a] M. Aizenman, R. Sims, and S. Warzel, Absolutely continuous spectra of quantum tree\ngraphs with weak disorder , Comm. Math. Phys. 264(2006), no. 2, 371–389.\n[ASW06b] ,Stability of the absolutely continuous spectrum of random S chr¨ odinger operators\non tree graphs , Probab. Theory Related Fields 136(2006), no. 3, 363–394.\n[BLT85] J. Bellissard, R. Lima, and D. Testard, Almost periodic Schr¨ odinger operators , Mathe-\nmatics + physics. Vol. 1, pages 1–64, World Sci. Publishing, Singapore , 1985.\n[BK05] J. Bourgain and C. E. Kenig, On localization in the continuous Anderson-Bernoulli\nmodel in higher dimension , Invent. Math. 161(2005), 389–426.\n[BGP08] J. Br¨ uning, V. Geyler, and K. Pankrashkin, Spectra of self-adjoint extensions and ap-\nplications to solvable Schr¨ odinger operators , Rev. Math. Phys. 20(2008), 1–70.\n[CL90] R. Carmona and J. Lacroix, Spectral theory of random Schr¨ odinger operators , Probabil-\nity and its Applications, Birkh¨ auser Boston Inc., Boston, MA, 1990 .\n[Cat97] Carla Cattaneo, The spectrum of the continuous Laplacian on a graph , Monatsh. Math.\n124(1997), no. 3, 215–235.\n[CCF+86] J. T. Chayes, L. Chayes, J. R. Franz, J. P. Sethna, and S. A. Trugman, On the density\nof states for the quantum percolation problem , J. Phys. A 19(1986), L1173–L1177.\n[CH94] J.-M. Combesand P. D. Hislop, Localization for some continuous, random Hamiltonians\nind-dimensions , J. Funct. Anal. 124(1994), 149–180.\n[CHK07] J. M. Combes, P. D. Hislop, and F. Klopp, An optimal Wegner estimate and its appli-\ncation to the global continuity of the integrated density of states for random Schr¨ odinger\noperators , Duke Math. J. 140(2007), no. 3, 469–498.\n[CHN01] J. M. Combes, P. D. Hislop, and Shu Nakamura, TheLp-theory of the spectral shift\nfunction, the Wegner estimate, and the integrated density o f states for some random\noperators , Comm. Math. Phys. 218(2001), 113–130.\n[EHS07] P. Exner, M. Helm, and P. Stollmann, Localization on a quantum graph with a random\npotential on the edges , Rev. Math. Phys. 19(2007), no. 9, 923–939.\n[EKK+08] P. Exner, J. P. Keating, P. Kuchment, T. Sunada, and A. Tepla yaev (eds.), Analysis on\ngraphs and its applications , Proc. Symp. Pure Math., vol. 77, Providence, R.I., Amer.\nMath. Soc., 2008.\n[FHS06] R. Froese, D. Hasler, and W. Spitzer, Absolutely continuous spectrum for the Anderson\nmodel on a tree: a geometric proof of Klein’s theorem , J. Funct. Anal. 230(2006), no. 1,\n184–221.\n[FS83] J. Fr¨ ohlich and T. Spencer, Absence of diffusion in the Anderson tight binding model\nfor large disorder or low energy , Comm. Math. Phys. 88(1983), 151–184.\n[GHV08] M. J. Gruber, M. Helm, and I. Veseli´ c, Optimal Wegner estimates for random\nSchr¨ odinger operators on metric graphs , in [EKK+08] (2008), 409–422.CONTINUITY OF THE IDS ON RANDOM LENGTH METRIC GRAPHS 29\n[GLV07] M. J. Gruber, D. Lenz, and I. Veseli´ c, Uniform existence of the integrated density of\nstates for random Schr¨ odinger operators on metric graphs o verZd, J. Funct. Anal. 253\n(2007), no. 2, 515–533.\n[GLV08] M. J. Gruber, D. Lenz, and I. Veseli´ c, Uniform existence of the integrated density of\nstates for combinatorial and metric graphs over Zd, in [EKK+08] (2008), 87–108.\n[GV08] M. Gruber and I. Veseli´ c, The modulus of continuity of the ids for random Schr¨ odinger\noperators on metric graphs , Random Oper. Stochastic Equations 16(2008), 1–10.\n[Har00] M.Harmer, Hermitian symplectic geometry and the factorization of the scattering matrix\non graphs , J. Phys. A 33(2000), no. 49, 9015–9032.\n[HV07] M. Helm and I. Veseli´ c, Linear Wegner estimate for alloy-type Schr¨ odinger operat ors on\nmetric graphs , J. Math. Phys. 48(2007), no. 9, 092107, 7.\n[HK02] P. D. Hislop and F. Klopp, The integrated density of states for some random operators\nwith nonsign definite potentials , J. Funct. Anal. 195(2002), no. 1, 12–47.\n[HP06] P.HislopandO.Post, Exponential localization for radial random quantumtrees ,Preprint\nmath-ph/0611022 , to appear in Waves in Random media (2006).\n[HKN+06] D. Hundertmark, R. Killip, S. Nakamura, P. Stollmann, and I. Ves eli´ c,Bounds on the\nspectral shift function and the density of states , Comm. Math. Phys. 262(2006), no. 2-3,\n489–503.\n[HLMW01] T. Hupfer, H. Leschke, P. M¨ uller, and S. Warzel, Existence and uniqueness of the\nintegrated density of states for Schr¨ odinger operators wi th magnetic fields and unbounded\nrandom potentials , Rev. Math. Phys. 13(2001), 1547–1581.\n[Kir89] W. Kirsch, Random Schr¨ odinger operators , H. Holden and A. Jensen (eds.), Schr¨ odinger\nOperators, Lecture Notes in Physics, 345, Springer, Berlin, 1989.\n[Kir96] W. Kirsch, Wegner estimates and Anderson localization for alloy-type potentials , Math.\nZ.221(1996), 507–512.\n[Kir07] ,An invitation to Random Sch¨ odinger operators , Preprint (2007).\n[KM82] W. Kirsch and F. Martinelli, On the density of states of Schr¨ odinger operators with a\nrandom potential , J. Phys. A: Math. Gen. 15(1982), 2139–2156.\n[KM07] W. Kirsch and B. Metzger, The integrated density of states for random Schr¨ odinger\noperators , Spectral Theory and Mathematical Physics, Proceedings of Sym posia in Pure\nMathematics, vol. 76, AMS, 2007, pp. 649–698.\n[KV02] W. Kirsch and I. Veseli´ c, Existence of the density of states for one-dimensional allo y-\ntype potentials with small support , MathematicalResultsin QuantumMechanics(Taxco,\nMexico, 2001), Contemp. Math., vol. 307, Amer. Math. Soc., Provid ence, RI, 2002,\npp. 171–176.\n[KLPS06] S. Klassert, D. Lenz, N. Peyerimhoff, and S. Stollmann, Elliptic operators on planar\ngraphs: unique continuation for eigenfunctions and nonpos itive curvature , Proc. Amer.\nMath. Soc. 134(2006), no. 5, 1549–1559 (electronic).\n[Kle96] Abel Klein, Spreading of wave packets in the Anderson model on the Bethe l attice,\nComm. Math. Phys. 177(1996), no. 3, 755–773.\n[Kle98] ,Extended states in the Anderson model on the Bethe lattice , Adv. Math. 133\n(1998), no. 1, 163–184.\n[Klo95] F. Klopp, Localization for some continuous random Schr¨ odinger oper ators, Comm.\nMath. Phys. 167(1995), no. 3, 553–569.\n[KP08] F. Klopp and K. Pankrashkin, Localization on quantum graphs with random vertex cou-\nplings, J. Statist. Phys. 131(2008), 561–673.\n[KP09] F. Klopp and K. Pankrashkin, Localization on quantum graphs with random edge length ,\nLett. Math. Phys., 87(2009), 99–114.\n[KS99] V. Kostrykinand R. Schrader, Kirchhoff’s rule for quantum wires , J. Phys. A 32(1999),\nno. 4, 595–630.\n[KV06] V. Kostrykin and I. Veseli´ c, On the Lipschitz continuity of the integrated density of\nstates for sign-indefinite potentials , Math. Z. 252(2006), no. 2, 367–392.\n[KS87] S. Kotani and B. Simon, Localization in general one-dimensional random systems. I I.\nContinuum Schr¨ odinger operators , Comm. Math. Phys. 112(1987), no. 1, 103–119.30 D. LENZ, N. PEYERIMHOFF, O. POST, AND I. VESELI ´C\n[Kuc91] P. Kuchment, On the Floquet theory of periodic difference equations , Geometrical and\nalgebraical aspects in several complex variables (Cetraro, 1989) , Sem. Conf., vol. 8,\nEditEl, Rende, 1991, pp. 201–209.\n[Kuc04] ,Quantum graphs: I. Some basic structures , Waves Random Media 14(2004),\nS107–S128.\n[Kuc05] ,Quantum graphs: II. Some spectral properties of quantum and combinatorial\ngraphs, J. Phys. A 38(2005), no. 22, 4887–4900.\n[KP07] P. Kuchment and O. Post, On the spectra of carbon nano-structures , Comm. Math.\nPhys.275(2007), no. 3, 805–826.\n[Len99] D. Lenz, Random operators and crossed products , Math. Phys. Anal. Geom. 2(1999),\nno. 2, 197–220.\n[LMV08] D. Lenz, P. M¨ uller, and I. Veseli´ c. Uniform existence of t he integrated density of states\nfor models on Zd.Positivity ,12(4):571–589, 2008.\n[LP08] F. Lled´ o and O. Post, Eigenvalue bracketing for discrete and metric graphs , J. Math.\nAnal. Appl. 348(2008), no. 2, 806–833.\n[LPPV08] D. Lenz, N. Peyerimhoff, O. Post, and I. Veseli´ c, Continuity properties of the integrated\ndensity of states on manifolds , Jpn. J. Math. 3(2008), no. 1, 121–161.\n[LPV04] D. Lenz, N. Peyerimhoff, and I. Veseli´ c, Integrated density of states for random metrics\non manifolds , Proc. London Math. Soc. (3) 88(2004), no. 3, 733–752.\n[LPV07] ,Groupoids, von Neumann algebras and the integrated density of states, Math.\nPhys. Anal. Geom. 10(2007), no. 1, 1–41.\n[LV08] D. Lenz and I Veseli´ c, Hamiltonians on discrete structures: jumps of the integrat ed\ndensity of states and uniform convergence. , to appear in Math. Z., arXiv:0709.2836\n(2008).\n[Lin01] E. Lindenstrauss, Pointwise theorems for amenable groups , Invent. Math. 146(2001),\nno. 2, 259–295.\n[MSY03] V. Mathai, Th. Schick, and S. Yates, Approximating spectral invariants of Harper oper-\nators on graphs. II , Proc. Amer. Math. Soc. 131(2003), no. 6, 1917–1923 (electronic).\n[MY02] V. Mathai and S. Yates, Approximating spectral invariants of Harper operators on\ngraphs, J. Funct. Anal. 188(2002), no. 1, 111–136.\n[Mat93] H. Matsumoto, On the integrated density of states for the Schr¨ odinger ope rators with\ncertain random electromagnetic potentials , J. Math. Soc. Japan 45(1993), 197–214.\n[Nic85] S. Nicaise, Some results on spectral theory over networks, applied to ne rve impulse trans-\nmission, Lecture Notes in Math., vol. 1171, Springer, Berlin, 1985, pp. 532 –541.\n[Pas71] L. A. Pastur, Selfaverageability of the number of states of the Schr¨ odin ger equation with\na random potential , Mat. Fiz. i Funkcional. Anal., 238(1971), 111–116.\n[PF92] L. Pasturand A. Figotin, Spectra of random and almost-periodic operators , Grundlehren\nder Mathematischen Wissenschaften, vol. 297, Springer-Verlag, Berlin, 1992.\n[Pos08] O. Post, Equilateral quantum graphs and boundary triples , in [EKK+08] (2008),469–490.\n[PV02] N.PeyerimhoffandI.Veseli´ c, Integrateddensity of states for ergodic random Schr¨ oding er\noperators on manifolds , Geom. Dedicata, 91(2002), 117–135\n[Shu79] M. A. Shubin, Spectral theory and the index of elliptic operators with alm ost-periodic\ncoefficients , Russ. Math. Surveys, 34(1979) 109–157.\n[Sto01] P. Stollmann, Caught by disorder: Bound states in random media , Progress in Mathe-\nmatical Physics, vol. 20, Birkh¨ auser Verlag, Basel, 2001.\n[Ves02] I. Veseli´ c, Wegner estimate and the density of states of some indefinite a lloy-type\nSchr¨ odinger operators , Lett. Math. Phys. 59(2002), no. 3, 199–214.\n[Ves05] ,Spectral analysis of percolation Hamiltonians , Math. Ann. 331(2005), no. 4,\n841–865.\n[Ves07] I. Veseli´ c, Lifshitz asymptotics for Hamiltonians monotone in the rand omness, Oberwol-\nfach Rep. 4(2007), no. 1, 380–382.\n[Ves08] I. Veseli´ c, Wegner estimates for sign-changing single site potentials , arXiv:0806.0482\n(2008).CONTINUITY OF THE IDS ON RANDOM LENGTH METRIC GRAPHS 31\n[vB85] J. von Below, A characteristic equation associated to an eigenvalue prob lem onC2-\nnetworks , Linear Algebra Appl. 71(1985), 309–325.\n[Weg81] F. Wegner, Bounds on the DOS in disordered systems , Z. Phys. B 44(1981), 9–15.\n(D. Lenz) Friedrich-Schiller-Universit ¨at Jena, Fakult ¨at f¨ur Mathematik & Infor-\nmatik, Mathematisches Institut, 07737 Jena, Germany\nE-mail address :daniel.lenz@uni-jena.de\nURL:www.tu-chemnitz.de/mathematik/mathematische physik/\n(N.Peyerimhoff) Department of Mathematical Sciences, Durham University, S cience\nLaboratories South Road, Durham, DH1 3LE, Great Britain\nE-mail address :norbert.peyerimhoff@durham.ac.uk\nURL:www.maths.dur.ac.uk/~dma0np/\n(O. Post) Institut f ¨ur Mathematik, SFB 647 “Space – Time – Matter”, Humboldt-\nUniversit ¨at zu Berlin, Rudower Chaussee 25, 12489 Berlin, Germany\nE-mail address :post@math.hu-berlin.de\nURL:www.math.hu-berlin.de/~post/\n(I. Veseli´ c) Fakult¨at f¨ur Mathematik,TU Chemnitz, D-09107 Chemnitz, Germany,\n& Emmy-Noether Programme of the DFG\nURL:www.tu-chemnitz.de/mathematik/schroedinger/members. php" }, { "title": "0812.2231v1.Supersolid_state_in_fermionic_optical_lattice_systems.pdf", "content": "arXiv:0812.2231v1 [cond-mat.supr-con] 11 Dec 2008Supersolid state in fermionic optical lattice systems\nAkihisa Koga,1Takuji Higashiyama,2Kensuke Inaba,2Seiichiro Suga,2and Norio Kawakami1\n1Department of Physics, Kyoto University, Kyoto 606-8502, J apan\n2Department of Applied Physics, Osaka University, Suita, Os aka 565-0871, Japan\n(Dated: November 8, 2018)\nWe study ultracold fermionic atoms trapped in an optical lat tice with harmonic confinement by\ncombiningthereal-space dynamicalmean-fieldtheorywitha two-site impuritysolver. Bycalculating\nthe local particle density and the pair potential in the syst ems with different clusters, we discuss\nthe stability of a supersolid state, where an s-wave superfluid coexists with a density-wave state of\ncheckerboard pattern. It is clarified that a confining potent ial plays an essential role in stabilizing\nthe supersolid state. The phase diagrams are obtained for se veral effective particle densities.\nI. INTRODUCTION\nSince the successful realization of Bose-Einstein con-\ndensation in a bosonic87Rb system1, ultracold atomic\nsystems have attracted considerable interest.2,3,4One of\nthe most active topics in this field is an optical lattice\nsystem,5,6,7,8which is formed by loading the ultracold\natoms in a periodic potential. This provides a clean sys-\ntem with quantum parameters which can be tuned in a\ncontrolledfashionfromweaktostrongcouplinglimits. In\nfact, remarkable phenomena have been observed such as\nthe phase transition between a Mott insulator and a su-\nperfluid in bosonic systems9. In addition, the superfluid\nstate10and the Mott insulating state11,12have been ob-\nserved in the fermionic optical lattices, which stimulates\ntheoretical investigations on the quantum states in the\noptical lattice systems. Among them, the possibilty of\nthesupersolidstatehasbeendiscussedasoneoftheinter-\nesting problems in optical lattice systems. The existence\nof the supersolid state was experimentally suggested in a\nbosonic4He system,13and was theoretically discussed in\nthestronglycorrelatedsystemssuchasbosonicsystems14\nand Bose-Fermi mixtures15. As for fermionic systems, it\nis known that a density wave (DW) state and an s-wave\nsuperfluid (SSF) state are degeneratein the half-filled at-\ntractiveHubbardmodelonthebipartitelatticeexceptfor\nonedimension,16,17whichmeansthatthesupersolidstate\nmight be realizable in principle. However, the degener-\nate ground states are unstable against perturbations. In\nfact, the hole doping immediately drives the system to a\ngenuine SSF state. Therefore, it is difficult to realize the\nsupersolid state in the homogeneous bulk system. By\ncontrast, in the optical lattice, an additional confining\npotential makes the situation different.18In our previ-\nous paper,19we studied the attractive Hubbard model\non square lattice with harmonic potential to clarify that\nthe supersolid state is indeed realized at low tempera-\ntures. However, we were not able to systematically deal\nwith large clusters to discuss how the supersolid state de-\npends on the particle density, the system size, etc. This\nmight be important for experimental observations of the\nsupersolid state in the optical lattice.\nIn this paper, we address this problem by combining\nthe real-space dynamical mean-field theory (R-DMFT)with atwo-site impurity solver. We then discuss how sta-\nble the DW, SSF and supersolid states are in the optical\nlattice system. We also clarify the role of the confining\npotential in stabilizing the supersolid state.\nThe paper is organized as follows. In Sec. II, we in-\ntroduce the model Hamiltonian and explain the detail of\nR-DMFT and its impurity solver. We demonstrate that\nthe supersolid state is indeed realized in a fermionic op-\ntical lattice with attractive interactions in Sec. III. In\nSec. IV, we discuss the stability of the supersolid state\nin large clusters. We also examine how the phase dia-\ngram depends on the particle number. A brief summary\nis given in Sec. V.\nII. MODEL HAMILTONIAN AND METHOD\nLetusconsiderultracoldfermionicatomsinthe optical\nlattice with confinement, which may be described by the\nfollowing attractive Hubbard model,16,17,20,21,22,23,24\nH=−t/summationdisplay\n/angbracketleftij/angbracketrightσc†\niσcjσ−U/summationdisplay\nini↑ni↓+/summationdisplay\niσv(ri)niσ,(1)\nwhereciσ(c†\niσ) annihilates (creates) a fermion at the ith\nsite with spin σandniσ=c†\niσciσ.t(>0) is a near-\nest neighbor hopping, U(>0) an attractive interaction,\nv(r) [=V(r/a)2] a harmonic potential and the term /an}bracketle{tij/an}bracketri}ht\nindicates that the sum is restricted to nearest neighbors.\nriis a distance measured from the center of the system\nandais lattice spacing. Here, we define the characteris-\ntic length of the harmonic potential as d= (V/t)−1/2a,\nwhich satisfies the condition v(d) =t.\nThe ground-state properties of the Hubbard model on\ninhomogeneous lattices have theoretically been studied\nby various methods such as the Bogoljubov-de Gennes\nequations25, the Gutzwiller approximation26, the slave-\nboson mean-field approach27, variational Monte Carlo\nsimulations28, local density approximation.29Although\nmagnetically ordered and superfluid states are described\nproperly in these approaches, it may be difficult to de-\nscribe the coexisting phase like a supersolid state in the\ninhomogeneous system. The density matrix renormal-\nization group method30and the quantum Monte Carlo\nmethod31are efficient for one-dimensional systems, but2\nit may be difficult to apply them to higher dimensional\nsystems with large clusters. We here use R-DMFT32,\nwhere local particle correlations are taken into account\nprecisely. This treatment is formally exact for the ho-\nmogeneous lattice model in infinite dimensions32and the\nmethod has successfully been applied to some inhomo-\ngeneous correlated systems such as the surface33or the\ninterface of the Mott insulators34, the repulsive fermionic\natoms35,36. Furthermore, it has an advantage in treating\nthe SSF state and the DW state on an equal footing in\nthe strong coupling regime, which allows us to discussthe supersolid state in the optical lattice.\nIn R-DMFT, the lattice model is mapped to an effec-\ntive impurity model, where local electron correlationsare\ntaken into account precisely. The lattice Green function\nisthen obtainedviaself-consistentconditionsimposedon\ntheimpurityproblem. When onedescribesthe superfluid\nstate in the framework of R-DMFT,32the lattice Green’s\nfunction for the system size Lshould be represented in\nthe Nambu-Gor’kov formalism. It is explicitly given by\nthe (2L×2L) matrix,\n/bracketleftBig\nˆG−1\nlat(iωn)/bracketrightBig\nij=−tδ/angbracketleftij/angbracketrightˆσz+δij/bracketleftBig\niωnˆσ0+{µ−v(ri)}ˆσz−ˆΣi(iωn)/bracketrightBig\n, (2)\nwhere ˆσα(α=x,y,z) is theαth component of the (2 ×\n2) Pauli matrix, ˆ σ0the identity matrix, µthe chemical\npotential, ωn= (2n+1)πTtheMatsubarafrequency, and\nTthe temperature. The site-diagonal self-energy at ith\nsite is given by the following (2 ×2) matrix,\nˆΣi(iωn) =/parenleftbigg\nΣi(iωn)Si(iωn)\nSi(iωn)−Σ∗\ni(iωn)/parenrightbigg\n,(3)\nwhere Σ i(iωn) [Si(iωn)] is the normal (anomalous) part\nof the self-energy. In R-DMFT, the self-energy at the\nith site is obtained by solving the effective impurity\nmodel, which is explicitly given by the following Ander-\nson Hamiltonian,23,24\nHimp,i=/summationdisplay\nkσEika†\nikσaikσ+/summationdisplay\nk(Dikaik↑aik↓+h.c.)\n+/summationdisplay\nkσVik/parenleftBig\nc†\niσaikσ+a†\nikσciσ/parenrightBig\n+ǫi/summationdisplay\nσc†\niσciσ−Uc†\ni↑ci↑c†\ni↓ci↓, (4)\nwhereaikσ(a†\nikσ)annihilates(creates)afermionwithspin\nσin the effective bath and ǫiis the impurity level. We\nhave here introduced the effective parameters in the im-\npurity model such as the spectrum of host particles Eik,\nthe pair potential Dikand the hybridization Vik. By\nsolving the effective impurity model eq. (4) for each\nsite, we obtain the site-diagonal self-energy and the lo-\ncal Green’s function. The R-DMFT self-consistent loop\nof calculations is iterated under the condition that the\nsite-diagonal component of the lattice Green’s function\nis equal to the local Green’s function obtained from the\neffective impurity model as/bracketleftBig\nˆGlat(iωn)/bracketrightBig\nii=ˆGimp,i(iωn).\nWhen R-DMFT is applied to our inhomogeneous sys-\ntem, it is necessary to solve the effective impurity mod-\nelsLtimes by iteration. Therefore, numerically powerful\nmethods such as quantum Monte Carlo simulations, theexact diagonalization method, and the numerical renor-\nmalization group method may not be efficient since they\nrequire long time to perform R-DMFT calculations. In\nthis paper, we use a two-site approximation38,39, where\nthe effective bath is replaced by only one site. In spite\nof this simplicity, it has an advantage in taking into ac-\ncount both low- and high-energy properties reasonably\nwell within restricted numerical resources34,38.\nIn the two-site approximation, a non-interacting\nGreen’s function for the impurity model at the ith site is\nsimplified as,\n/bracketleftBig\nˆG0\nimp,i(iωn)/bracketrightBig−1\n=iωnˆσ0−ǫiˆσz\n−Viˆσz1\niωnˆσ0−Eiˆσz−DiˆσxViˆσz,(5)\nwhere the index kwas omitted. The effective parameters\n{Ei,Di,Vi,ǫi}should be determined self-consistently so\nthat the obtained results properly reproduce the original\nlattice problem. Here, we use the following equations,\nǫi=−Re/bracketleftBig\nˆG0\nimp,i(iωn)/bracketrightBig−1\n11/vextendsingle/vextendsingle/vextendsingle/vextendsingle\nn→∞(6)\nVi=/radicalBig\n(a−1)(π2T2+E2\ni+D2\ni) (7)\nDi=b\n1−a, (8)\nwherea= Im[ˆG0\ni(iω0)]−1\n11/πTandb= Re[ˆG0\ni(iω0)]−1\n12.\nFurthermore, the number of particles is fixed in the non-\ninteracting Green’s function, as\nn(i)\n0= 2T/summationdisplay\nn=0Re/bracketleftBig\nˆG0\nimp,i(iωn)/bracketrightBig\n11+1\n2.(9)\nWecandeterminetheeffectiveparameters {Ei,Di,Vi,ǫi}\nin terms of these equations.\nHere, the effective particle density is defined as ˜ ρ=\nN/πd2, whereNis the total number of particles. This3\ndensity relates the systems with different sites, num-\nber of particles, and curvatures of the confining po-\ntentials in the same way as the particle density does\nfor periodic systems with different sites. We set t\nas a unit of energy and calculate the density profile\n/an}bracketle{tniσ/an}bracketri}ht= 2T/summationtext\nn=0Re[Giσ(iωn)]+1\n2and the distribution\nof the pair potential ∆ i= 2T/summationtext\nn=0Re[Fi(iωn)], where\nGiσ(iωn)[Fi(iωn)] is the normal (anomalous) Green’s\nfunction for the ith site. Note that ∆ irepresents the\norder parameter for the SSF state for the ith site.\nIn the following, we consider the attractive Hubbard\nmodel on square lattice with harmonic confinement as a\nsimple model for the supersolid. In this case, it is known\nthat the symmetry of the square lattice is not broken in\nthe DW, SSF, and supersolid states19,25,30. Therefore,\nthe point group C4vis useful to deal with the system\non the inhomogeneous lattice. For example, when the\nsystem with 5513 sites ( r <42.0) is treated, one can\ndeal with only 725 inequivalent sites. This allows us to\ndiscuss the low temperature properties in larger clusters,\nin comparison with those with ( d/a= 6.5,N∼300)\ntreated in our previous paper.\nIII. LOW TEMPERATURE PROPERTIES\nBy means of R-DMFT with the two-site impurity\nsolver,we obtainthe resultsforthe system with d/a= 10\nandN∼720(˜ρ∼2.3). Figures 1 and 2 show the profiles\nof the local density and the pair potential at T/t= 0.05.\nIn the non-interacting case ( U/t= 0), fermionic atoms\naresmoothlydistributed upto r/a∼21, asshownin Fig.\n1 (a). Increasing the attractive interaction U, fermions\ntend to gather around the bottom of the harmonic po-\ntential, as seen in Fig. 1 (b). In these cases, the pair po-\ntential is not yet developed, as shown in Figs. 2 (a) and\n(b), and thereby the normal metallic state with short-\nrange pair correlations emerges in the region ( U/t<∼2).\nFurther increase in the interaction Uleads to different\nbehavior, where the pair potential ∆ iis induced in the\nregion with /an}bracketle{tniσ/an}bracketri}ht /ne}ationslash= 0. Thus, the SSF state is induced by\nthe attractive interaction, which is consistent with the\nresults obtained from the Bogoljubov-de Gennes equa-\ntion25. In the case with U/t= 3, another remarkable\nfeature is found around the center of the harmonic po-\ntential (r/a <7), where a checkerboard structure ap-\npears in the density profile /an}bracketle{tniσ/an}bracketri}ht, as shown in Fig. 1\n(c). This implies that the DW state is realized in the\nregion. On the other hand, the pair potential ∆ iis not\nsuppressed completely even in the DW region, as shown\nin Fig. 2 (c). This suggests that the DW state coex-\nists with the SSF state, i.e. a supersolid state appears\nin our optical lattice system. The profile characteris-\ntic of the supersolid state is clearly seen in the case of\nU/t= 5. Figures 1(d) and 2(d) show that the DW state\nof checkerboard structure coexists with the SSF state in\nthe doughnut-like region (5 < r/a < 15). By contrast,\nthe genuine SSF state appears inside and outside of the\nFIG. 1: (Color online) The density profile /angbracketleftniσ/angbracketrightin the optical\nlattice system with d/a= 10 at T/t= 0.05 when U/t=\n0.0,2.0,3.0,5.0,7.0 and 10 .0 (from the top to the bottom).\nregion (r/a <5,15< r/a < 17). Further increase in\nthe interaction excludes the DW state out of the cen-\nter since fermionic atoms are concentrated around the\nbottom of the potential for large U. In the region, two\nparticles with opposite spins are strongly coupled by the\nattractive interaction to form a hard-core boson, giving\nrise to the band insulator with /an}bracketle{tniσ/an}bracketri}ht ∼1, instead of the\nSSF state. Therefore, the SSF state survives only in the\nnarrow circular region surrounded among the empty and\nfully occupied states. We see such behavior more clearly\nin Figs. 1 (f) and 2 (f). Note that in the strong coupling\nlimitU/t→ ∞, all particles are condensed in the region\nr < rc=/radicalbig\n˜ρ/2d∼1.07d= 10.7a.\nIn this section, wehavestudied the attractiveHubbard4\nFIG. 2: (Color online) The pair potential ∆ iin the optical\nlattice system with d/a= 10 at T/t= 0.05 when U/t=\n0.0,2.0,3.0,5.0,7.0 and 10 .0 (from the top to the bottom).\nmodel with the harmonic potential to clarify that the\nsupersolid state is realized in a certain parameter region.\nHowever, it is not clear how the supersolid state depends\non the system size and the number of particles. To make\nthis point clear, we deal with largeclusters to clarify that\nthe supersolid state is indeed realized in the following.\nIV. STABILITY OF THE SUPERSOLID STATE\nInthissection,wediscussthestabilityofthesupersolid\nstate in fermionic optical lattice systems, which may be\nimportant for experimental observations. First, we clar-\nifyhowlow-temperaturepropertiesdependonthesystemsize, byperformingR-DMFTforseveralclusterswithdif-\nferentd. We here fix U/t= 5 and µ/t∼ −1.58 to obtain\nthe profiles of the local particle density and the pair po-\ntential, which are shown in Fig. 3. We note that the\nS\u0010E \nS\u0010E \nFIG. 3: Profiles of particle density /angbracketleftniσ/angbracketrightand pair potential\n∆ias a function of r/dwith fixed d= 12.5,17.5 and 22 .5,\nwhenU/t= 5.\ndistance ris normalized by din the figure. It is found\nthat/an}bracketle{tniσ/an}bracketri}htand ∆ idescribe smooth curves for r/d<∼0.6\nand 1.4<∼r/d<∼1.7, where the genuine SSF state is re-\nalized. On the other hand, for 0 .6<∼r/d<∼1.4, two dis-\ntinct magnitudes appear in /an}bracketle{tniσ/an}bracketri}ht, reflecting the fact that\nthe DW state with two sublattices is realized. Since the\npair potential is also finite in the region, the supersolid\nstate is realized. In this case, we deal with finite systems,\nand thereby all data are discrete in r. Nevertheless, it is\nfound that the obtained results are well scaled by dal-\nthough some fluctuations appear due to finite-size effects\nin the small dcase. The effective particle density ˜ ρis\nalmost constant in the above cases. Therefore, we con-\nclude that when ˜ ρ∼2.3, the supersolid state discussed\nhere is stable in the limit with N,d→ ∞. This result\ndoes not imply that the supersolid state is realized in\nthe homogeneous system with arbitrary fillings. In fact,\nthe supersolid state might be realizable only at half fill-\ning16,17,20,21,23. Therefore, we can say that a confining\npotential is essential to stabilize the supersolid state in\nthe optical lattice system.\nNext, we focus on the system with U/t= 5 and\nd/a= 10 to discuss in detail how the supersolid state\ndepends on the effective particle density ˜ ρ. The DW\nstate is characterized by the checkerboard structure in5\nthe density profile /an}bracketle{tniσ/an}bracketri}ht, so that the Fourier transform\nnqatq= (π,π) is appropriate to characterize the exis-\ntence of the DW state. On the other hand, the Fourier\ntransform ∆ qatq= (0,0) may represent the rigidity of\nthe SSF state in the system. In Fig. 4, we show the\nʪO ππ ʫʪO \u0011\u0011 ʫ\nʪ∆\u0011\u0011 ʫʪO \u0011\u0011 ʫ \u0016\nρ∼\nFIG. 4: (Color online) /angbracketleftnππ/angbracketright//angbracketleftn00/angbracketrightand ∆ 00/5/angbracketleftn00/angbracketrightas a func-\ntion of the effective particle density ˜ ρ(=N/πd2) whenU= 5t\nandT= 0.05t. A broken line represents the local particle\ndensity at the center of the lattice in the noninteracting ca se.\nsemilog plots of the parameters normalized by /an}bracketle{tn00/an}bracketri}ht. It\nis found that the normalized parameter ∆ 00is always fi-\nnite although the increase in the attractive interaction\nmonotonically decreases it. This implies that the SSF\nstate appears in the system with the arbitrary particle\nnumber. In contrast to this SSF state, the DW state\nis sensitive to the effective particle density as shown in\nFig. 4. These may be explained by the fact that in the\nsystem without a harmonic confinement ( V0= 0), the\nDW state is realized only at half filling ( n= 0.5), while\nthe SSF state is always realized. To clarify this, we also\nshow the local particle density in the noninteracting case\nat the center of the system as the broken line in Fig. 4.\nIt is found that when the quantity approaches half-filling\n(∼0.5),/an}bracketle{tnππ/an}bracketri}ht//an}bracketle{tn00/an}bracketri}httakes its maximum value, where the\nSSF state coexists with the DW state. Therefore, we can\nsay that the supersolid state is stable around this condi-\ntion. Increasing the effective particle density, the band\ninsulating states become spread around the center, while\nthe DW and SSF states should be realized in a certain\ncircular region surrounded among the empty and fully-\noccupied regions. Therefore, the normalized parameters\n/an}bracketle{tnππ/an}bracketri}htand ∆ 00are decreased with increase in ˜ ρ. On the\nother hand, in the case with low density ˜ ρ<∼0.3, the\nlocal particle density niat each site is far from half fill-\ning even when U/t= 5. Therefore, the DW state does\nnot appear in the system, but the genuine SSF state is\nrealized. These facts imply that the condition n∼0.5\nis still important to stabilize the supersolid state even in\nfermionic systems confined by a harmonic potential.18\nBy performing similar calculations for the systems\nwith low, intermediate, and high particle densities (˜ ρ∼\n0.63,2.3 and 8.9), we end up with the phase diagrams, asshown in Fig. 5. We find that increasing the attractive\nFIG. 5: (Color online) The phase diagram of the attractive\nHubbard model on the optical lattice with ˜ ρ∼0.63,2.3 and\n8.9. The density plot represents the profiles of the s-wave pair\npotential as a function of the attractive interaction U/t. The\nDW state is realized in the shaded area. The broken lines\ngive a guide to eyes which distinguishes the region with a\nfractional particle density from the empty and fully-occup ied\nregions.\ninteraction, fermionic particles gradually gather around\nthe center of the system, where the empty state is stabi-\nlized away from the center and the band insulating state\nwith fully occupied sites is stabilized. It is found that\nthe region surrounded among these states strongly de-\npend on the effective particle density ˜ ρ. The increase in\nthe effective particle density shrinks the region, which af-\nfects the stability of the SSF, DW and their coexisting\nstates. In particular, the DW region, which is shown as\nthe shaded area in Fig. 5, is sensitive to the effective par-\nticle density, as discussed above. Namely, the local pair6\npotential ∆ itakes its maximum value around U/t∼15,\nwhich may give a rough guide for the crossover region\nbetween the BCS-type and the BEC-type states. We\nnote that the DW state appears only in the BCS region\n(U/t∼5). This implies that the condition n∼0.5 is not\nsufficient, but necessary to stabilize the supersolid state\nin the attractive Hubbard model with an inhomogeneous\npotential.\nWe wish to comment on the conditions to observe the\nsupersolid state in the fermionic optical lattice system.\nNeedless to say, one of the most important conditions\nis the low temperature.19Second is the tuning of the\neffective particle density ˜ ρ(∼1), which depends on the\ncurvature of the harmonic potential as well as the total\nnumber of particles. This implies that a confined poten-\ntial play a crucial role in stabilizing the supersolid state\nin fermionic optical lattice systems. In addition to this,\nan appropriate attractive interaction is necessary to sta-\nbilize the DW state in the BCS-type SSF state. When\nthese conditions are satisfied, the supersolid state is ex-\npected to be realized at low temperatures.\nV. SUMMARY\nWe haveinvestigatedthe fermionic attractiveHubbard\nmodel in the optical lattice with harmonic confinement.By combining R-DMFT with a two-site impurity solver,\nwe have obtained the rich phase diagram on the square\nlattice, which has a remarkable domain structure includ-\ning the SSF state in the wide parameter region. By per-\nforming systematic calculations, we have then confirmed\nthat the supersolid state, where the SSF state coexists\nwith the DW state, is stabilized even in the limit with\nN→ ∞,V0→0 and ˜ρ∼const. We have also elucidated\nthat a confining potential plays a key role in stabilizing\nthe supersolid state.\nAcknowledgments\nThis work was supported by the Grant-in-Aid for the\nGlobal COE Program ”The Next Generation of Physics,\nSpun from Universality and Emergence” and Scientific\nResearch [20740194 (A.K.), 20540390 (S.S.), 19014013\nand 20029013 (N.K.)] from the Ministry of Education,\nCulture, Sports, Science and Technology (MEXT) of\nJapan. Some of the computations were performed at the\nSupercomputer Center at the Institute for Solid State\nPhysics, University of Tokyo.\n1M. H. Anderson, J. R. Ensher, M. R. Matthews, C. E.\nWieman, and E. A. Cornell: Science 269(1995) 198.\n2For a review, see Nature (London) 416(2002) 205-246.\n3C. J. Pethick and H. Smith, Bose-Einstein Condensation\nin Dilute Gases (Cambridge University Press, Cambridge,\n2002).\n4L. Pitaevskii and S. Stringari, Bose-Einstein Condensation\n(Clarendon Press, Oxford, 2003).\n5I. Bloch and M. Greiner: Advances in Atomic, Molecu-\nlar, and Optical Physics , edited by P. Berman and C. Lin\n(Academic Press, New York, 2005), Vol. 52 p. 1.\n6I. Bloch: Nature Physics 1(2005) 23.\n7D. Jaksch and P. Zoller: Ann. Phys. (NY) 315(2005) 52.\n8O. Morsch and M. Oberhaler: Rev. Mod. Phys. 78(2006)\n179.\n9M. Greiner, O. Mandel, T. Esslinger, T. W. H¨ ansch and I.\nBloch: Nature 415(2003) 39.\n10J. K. Chin, D. E. Miller, Y. Liu, C. Stan, W. Setiawan, C.\nSanner, K. Xu, and W. Ketterle: Nature 443(2006) 961.\n11R. J¨ ordens, N. Strohmaier, K. G¨ unter, H. Moritz, T.\nEsslinger, arXiv:0804.4009.\n12U. Schneider, L. Hackerm¨ uller, S. Will, Th. Best, I. Bloch,\nT. A. Costi, R. W. Helmes, D. Rasch and A. Rosch,\narXiv:0809.1464.\n13E. Kim and M. H. W. Chan: Nature 427(2004) 225.\n14P. Sengupta, L. P. Pryadko, F. Alet, M. Troyer, and\nG. Schmid: Phys. Rev. Lett. 94(2005) 207202; D. L.\nKovrizhin, G. V. Pai, and S. Sinha: Europhysics Letters\n72(2005) 162; V. W. Scarola and S. Das Sarma: Phys.\nRev. Lett. 95(2005) 033003; S. Wessel and M. Troyer:Phys. Rev. Lett. 95(2005) 127205; D. Heidarian and K.\nDamle: Phys.Rev.Lett. 95(2005) 127206; R.G.Melko, A.\nParamekanti, A. A. Burkov, A. Vishwanath, D. N. Sheng,\nand L. Balents: Phys. Rev. Lett. 95(2005) 127207; G. G.\nBatrouni, F. H´ ebert, and R. T. Scalettar: Phys. Rev. Lett.\n97(2006) 087209.\n15H. P. B¨ uchler and G. Blatter: Phys. Rev. Lett. 91(2003)\n130404; I. Titvinidze, M. Snoek and W. Hofstetter: Rhys.\nRev. Lett., 100, 100401 (2008).\n16R. T. Scalettar, E. Y. Loh, J. E. Gubernatis, A. Moreo, S.\nR. White, D. J. Scalapino, R. L. Sugar, and E. Dagotto:\nPhys. Rev. Lett. 62(1989) 1407.\n17J. K. Freericks, M. Jarrell, and D. J. Scalapino: Phys. Rev.\nB48(1993) 6302.\n18F. Karim Pour, M. Rigol, S. Wessel, and A. Muramatsu,\nPhys. Rev. B 75, 161104(R) (2007).\n19A. Koga, T. Higashiyama, K. Inaba, S. Suga, and N.\nKawakami, J. Phys. Soc. Jpn. 77, 073602 (2008).\n20M. Capone, C. Castellani, and M. Grilli: Phys. Rev. Lett.\n88(2002) 126403.\n21R.Micnas, J. Ranninger, andS.Robaszkiewicz: Rev.Mod.\nPhys.62(1990) 113.\n22M. Keller, W. Metzner, and U. Schollw¨ ock: Phys. Rev.\nLett.86(2001) 4612.\n23A. Garg, H. R. Krishnamurthy, and M. Randeria: Phys.\nRev. B72(2005) 024517.\n24A. Toschi, M. Capone, and C. Castellani: Phys. Rev. B 72\n(2005) 235118.\n25Y. Chen, Z. D. Wang, F. C. Zhang, and C. S. Ting: cond-\nmat/0710. 5484.7\n26M. Yamashita and M. W. Jack: Phys. Rev. A76(2007)\n023606.\n27A. R¨ uegg, S. Pilgram, and M. Sigrist: Phys. Rev. B 75\n(2007) 195117.\n28Y. Fujihara, A. Koga, and N. Kawakami: J. Phys. Soc.\nJpn.76(2007) 034716.\n29T.-L. Dao, A. Georges, and M. Capone: arXiv:0704.2660.\n30M. Machida, S. Yamada, Y. Ohashi, and H. Matsumoto:\nPhys. Rev. A 74(2006) 053621.\n31M. Rigol, A.Muramatsu, G.G. Batrouni, andR.T.Scalet-\ntar: Phys. Rev. Lett. 91(2003) 130403.\n32W. Metzner and D. Vollhardt: Phys. Rev. Lett. 62(1989)\n324; E. M¨ uller-Hartmann: Z. Phys. B 74(1989) 507; T.\nPruschke, M. Jarrell, and J. K. Freericks: Adv. Phys. 42\n(1995) 187; A. Georges, G. Kotliar, W. Krauth and M. J.Rozenberg: Rev. Mod. Phys. 68(1996) 13.\n33M. Potthoff and W. Nolting: Phys. Rev. B 59(1999) 2549.\n34S. Okamoto and A. J. Millis: Phys. Rev. B 70(2004)\n241104(R); Nature (London) 428(2004) 630.\n35R. W. Helmes, T. A. Costi and A. Rosch: Phys. Rev. Lett.\n100(2008) 056403.\n36M. Snoek, I. Titvinidze, C. T¨ oke, K. Byczuk, and W. Hof-\nstetter: arXiv:0802.3211.\n37G. Xianlong, M. Rizzi, M. Polini, R. Fazio, M. P. Tosi, V.\nL. Campo Jr. and K. Capelle, Phys. Rev. Lett. 98, 030404\n(2007).\n38T. Higashiyama, K. Inaba, and S. I. Suga, Phys. Rev. A\n77, 043624 (2008).\n39M. Potthoff: Phys. Rev. B 64(2001) 165114." }, { "title": "0901.1087v2.Combinatorial_nuclear_level_density_model.pdf", "content": "arXiv:0901.1087v2 [nucl-th] 26 Jul 2009Combinatorial nuclear level-density model\nH. Uhrenholt1, S.˚Aberg1, P. M¨ oller2, T. Ichikawa3\n1Division of Mathematical Physics, LTH, Lund University, P. O. Box 118, S-221 00 Lund, Sweden\n2Theoretical Division, Los Alamos National Laboratory, Los Alamos, NM 87545, USA\n3RIKEN Nishina Center, RIKEN, Wako, Saitama 351-0198, Japan\nA microscopic nuclear level-density model is presented. Th e model is a completely combinatorial\n(micro-canonical) model based on the folded-Yukawa single -particle potential and includes explicit\ntreatment of pairing, rotational and vibrational states. T he microscopic character of all states\nenables extraction of level distribution functions with re spect to pairing gaps, parity and angular\nmomentum. The results of the model are compared to available experimental data: neutron separa-\ntion energy level spacings, data on total level-density fun ctions from the Oslo method, cumulative\nlevel densities from low-lying discrete states, and data on parity ratios.\nI. INTRODUCTION\nNuclear many-body level-density models arekey ingre-\ndients in nuclear reaction theories, where they, for exam-\nple, govern the rates and decay patterns of astrophysi-\ncal processes and nuclear fission. In statistical methods,\nfor example the Hauser-Feshbach formalism [1] for de-\nscribing nuclear reactions, a knowledge of the level den-\nsity is crucial [2, 3, 4]. How to calculate the nuclear\nlevel density (NLD) has been a long-standing challenge\n[5, 6, 7, 8, 9, 10]. Recently it has been subject to renewed\ninterest, theoretically as well as experimentally.\nThe simplest type of model is the Fermi-gas model,\nwhich is based on the partition-function method. It pro-\nvidessimpleanalyticalformulasfortheNLD[11]. Several\nphenomenologicalextensionshavebeenproposedinorder\nto reproduce experimental data. By adjusting free model\nparameters to data these models give unprecedented ac-\ncuracy in the region of the parameter fit [4, 12]. Also\nsemi-classical methods have been used to obtain expres-\nsions for the level density [13]. However, the Fermi-gas\nmodels are unreliable outside these regions, eg. when\nthey are extrapolated to higher excitation energies or\nto nucleon numbers far from stability. Ideally nuclear\nstructure should be included in NLD models. Several\ncombinatorial models based on nuclear mean-field the-\nory have been proposed, see eg. Refs. [14, 15]. Beyond\nmean-field methods have also been used to model the\nNLD, eg. the Shell-Model Monte-Carlo method [16, 17]\nand the interacting shell model [18]. These latter models\ntake into account effective nucleon-nucleon interactions,\nbut at the same time suffer from limitations due to the\nlimited size of the Hilbert space, and hence are presently\nunable to provide global predictions for level spacings at\nthe neutron-separation energy.\nExperimentally the NLD has been subject to renewed\ninterest in the last decade partly due to the development\nof the Oslo method, which has provided new experimen-\ntal data [19]. The Oslo method provides the level density\nover extended regions of excitation energy as opposed\nto the neutron-separation level-spacings data which only\nprovide one data point at relatively high excitation en-\nergy. Also recent measurements of separate level densi-tiesof2+and2−states[20]challengetheorytoreproduce\nthese observed parity ratios.\nFew nuclear-structure models have been used to si-\nmultaneously globally describe nuclear masses, fission\nbarriers, ground-state spins and decay rates. One such\nmodel is the microscopic-macroscopic FRLDM model\nwhich has previously been used to model these observ-\nables [3, 21, 22, 23], and here serve as the starting point\nfor calculating the nuclear level density. In this paper\na combinatorial (micro-canonical) nuclear level-density\nmodel based on the folded-Yukawa single-particle model\nis presented. The model is fully microscopic with pairing\ncorrelations, vibrations, and rotational excitations cal-\nculated for each many-particle-many-hole excited state.\nPresently, the lowest ten million or so states can be ac-\ncounted for. This implies that the excitation energy re-\ngion from the ground state to well above the neutron res-\nonance region is included. No additional parameters are\nintroduced in the model, and no refitting of parameters\nof the FRLDM is performed.\nIn Sec. II the combinatorial folded-Yukawa (CFY)\nlevel-density model is described. The model allows ex-\nplicit tracking of quantum numbers, and distributions\nof pairing gaps, parity and angular momentum are dis-\ncussed in Sec. III. Results from the CFY model are com-\npared to experimental data in Sec. IV, and in Sec. V the\nCFY model is compared to other theoretical NLD mod-\nels. Finally, a short summary is given in Sec. VI.\nII. THE COMBINATORIAL MODEL OF NLD\n(CFY)\nThe NLD is calculated by means of a combinatorial\ncountingofexcitedmany-particle-many-holestatesasde-\nscribed in Sec. IIA. In Sec. IIB we present how pairing\nis taken into account for excited states by explicitly solv-\ning the BCS equations for all individual configurations.\nRotations are taken into account combinatorially with\na pairing-dependent moment of inertia (Sec. IIC). The\nvibrational contribution to the NLD is investigated by\nincluding microscopically described phonons using the\nQuasi-particle Tamm-Dancoff Approximation (QTDA),2\nseeSec.IID. InSec.IIEitisdescribedhowwebrieflyac-\ncount for a general residual interaction causing a smear-\ning of the level-density distribution at higher excitation\nenergies.\nAll produced nuclearlevels, calculatedasdescribedbe-\nlow, are sorted into a binned level density, where the typ-\nical bin size is in the range ∆ E= 30–50 keV. The level\ndensity is calculated by counting the number of levels in\nthe energy bin,\nρ(Eb,I,π) =1\n∆E/integraldisplayEb+∆E\n2\nEb−∆E\n2/summationdisplay\niδ(E−Ei(I,π))dE,(1)\nwhereEbis the energy center of bin bandEi(I,π)\ndenotes the calculated state with energy Ei(given by\nEq. (15)), angular momentum Iand parity π. The total\nlevel density at a given excitation energy Eis\nρtot(E) =/summationdisplay\nI,πρ(E,I,π). (2)\nA. Combinatorial intrinsic level density\nA high-quality combinatorial level-density model re-\nquires a realistic description of the single-particle ener-\ngies. We shall here utilize a well-tested mean-field model,\nnamely the microscopic-macroscopic finite range liquid\ndrop model (FRLDM) [21], that provides a good global\ndescriptionofseveralnuclear-structurepropertiessuchas\nmasses, fission barriers and beta-decay properties. The\ngood agreement between ground-state spins calculated\nin the FRLDM model and experiment implies that the\nsingle-particle spectrum close to the Fermi surface is well\ndescribed [22].\nThe single-particle energies are thus obtained by solv-\ning the one-body Schr¨ odinger equation,\n(T+VFY(¯ε))|ν/an}bracketri}ht=eν|ν/an}bracketri}ht, (3)\nfor protons as well as for neutrons, where VFYis the\nFolded-Yukawasingle-particle potential. The parameters\nof the potential, as well as the deformation parameters,\n¯ε= (ε2,ε3,ε4,...), are taken from an extensive calcula-\ntion of nuclear masses [21]. All parameters of the single-\nparticle model are thus fixed from other studies. One\naim of the present study is to see how well highly excited\nstates (in terms of level-densities) are described for all\nnuclei heavier than16O.\nThe many-body ground state (the many-body particle\nvacuum) |0/an}bracketri}ht, isobtainedbyfillingthe lowest N(Z)states\n(including time-reversed states),\n|0/an}bracketri}ht=N(Z)/summationdisplay\nν=1a+\nν|−/an}bracketri}ht, (4)\nwhere|−/an}bracketri}htis the one-body vacuum. Intrinsic excitations\nare obtained by many-particle-many-hole excitations onthe many-body particle vacuum,\n|i/an}bracketri}ht=n/productdisplay\nα=1a+\nναaν′α|0/an}bracketri}ht, (5)\nwhereνandν′span all single-particle states (as well as\ntime-reversed states) in the potential. We include all n-\nparticle-n-hole states with n <10 for protons as well as\nfor neutrons. Corresponding energies are given by\nEi=E0+n/summationdisplay\nα=1(eνα−eν′\nα), (6)\nwhereE0is the ground-state energy. (When pairing\nis included this expression is modified as described in\nSec. IIB.) Nuclei are assumed to have constant defor-\nmation for all excitation energies. For spherical and\nwell-deformed nuclei this assumption is expected to be\na good approximation for excitation energies here con-\nsidered (mainly below the neutron separation energy).\nFor transitional nuclei this approximation might, how-\never, introduce some inaccuracies in the calculations.\nIn addition to the energy and the number of unpaired\nnucleons (seniority), we also register the total parity and\nK-quantum number for each excited state, where Kas\nusualisthe angularmomentum projectiononan intrinsic\nsymmetry axis. The identification of a K-quantum num-\nber can be done for all nuclei by assuming an infinites-\nimal deformation for spherical nuclei. The K-quantum\nnumber of the individual unpaired nucleons are allowed\nto couple in all different ways to build up the total K-\nquantum number.\nB. Pairing\nThe many-body wave-function of the excited states is\napproximated by the BCS wave-function with excited\nquasi-particles\n|τ/an}bracketri}ht=/productdisplay\nν′′∈τ2(−Vν′′+Uν′′a†\nν′′a†\n¯ν′′)×\n×/productdisplay\nν′∈τ1a†\nν′/productdisplay\nν∈τ0(Uν+Vνa†\nνa†\n¯ν)|0/an}bracketri}ht, (7)\nwhereτ2,τ1andτ0denote the spaces of double, single\nand zero quasi-particle excitations, respectively, used to\nbuild the many-body state iof Eq. 5. UνandVνare the\nstandard BCS vacancy and occupation factors and |0/an}bracketri}htis\nthe particle vacuum, see e.g. Ref. [24]. For the excited\npairs in the group τ2the effect (compared to zero-quasi-\nparticle states) is simply\nUν→ −Vν, V ν→Uν. (8)\nThepairinggap∆andtheFermienergy λareobtained3\nby solving the BCS-equations\n∆ =G/bracketleftBigg/summationdisplay\nν∈τ0UνVν−/summationdisplay\nν′′∈τ2Uν′′Vν′′/bracketrightBigg\n,(9)\nN= 2/summationdisplay\nν∈τ0V2\nν+/summationdisplay\nν′∈τ11+2/summationdisplay\nν′′∈τ2U2\nν′′(10)\nfor each state.\nThe pairing strength in the BCS model is governed\nby a single, primary BCS model parameter, namely the\nparameter of the effective-interaction pairing gap ( rin\nEq.˜(47) in [25] or the equivalent cpin Eq. (49) and\ncnin Eq. (50) in [26]). In an adjustment of macro-\nscopic, pairing, and other model parameters to optimize\na nuclear mass model r= 4.8 MeV was obtained for\nthe standard BCS pairing model [21]. Because the BCS\nmodel we use here, in contrast to Ref. [21, 25], includes\nblocking and particle-number-projection it is necessary\nto use a pairing strength optimized for this method. In\nRef. [26] cp=cn= 4.95 MeV were obtained in an ad-\njustment of calculated pairing gaps to odd-even mass\ndifferences. However, it has been pointed out [21, 25]\nthat odd-even mass differences are subject to numerous\nnon-smooth contributions, for example from deformation\nchanges and from irregularities in the microscopic level\nstructure, not just odd-even staggering due to pairing. It\nis thereforebetter to determine the pairing strength from\na full nuclear mass calculation that includes the particu-\nlar pairing model under consideration and the associated\nadjustment of all model parameters [21]. In the folded-\nYukawa macroscopic-microscopic mass model this yields\nthe value 4.5 MeV for the optimized strength parameter\nof the particle-number-projected BCS model [26]. The\npairing strength G, which is used to calculate pairing\ngaps for all excited states in the level-density calculation,\nis determined from this average pairing gap for each nu-\nclear system; for details see Ref. [26].\nIt follows that the excitation energy of the intrinsic\nmany-body configuration for one nucleon type is\nEmb,t= 2/summationdisplay\nν∈τ0eνV2\nν+/summationdisplay\nν′∈τ1eν+2/summationdisplay\nν′′∈τ2eνU2\nν′′−\nG/summationdisplay\nν∈τ0V4\nν−G\n2/summationdisplay\nν′∈τ11−G/summationdisplay\nν′′∈τ2U4\nν′′−∆2\nG−E0\nt,(11)\nwheretdenotes protons or neutrons and E0\ntis the proton\nor neutron part of the ground-state energy. The pairing\ngap and Fermi level are calculated by solving the BCS\nequationsEqs.(9)and(10). Thetotalintrinsicexcitation\nenergy of a many-body state iis\nEi\nmb=Emb,p+Emb,n. (12)\nThe combinatorial approach to calculate the level den-\nsity involves calculating all possible many-body states,\nEi\nmb, which thus means that the BCS equations are\nsolved about 107times for each nucleus.C. Rotations\nA general feature of deformed nuclei is the existence of\nrotational bands built on the ground state as well as on\nexcited states. The question is how to incorporate these\nstates in the NLD. In principle, rotational states may be\nmicroscopically treated by solving the cranking Hamil-\ntonian,hω=h0−ωrotjx, whereωrotis the rotational\nfrequency and jxis the angular momentum operator for\nrotations around the x-axis. Matrix elements of the jx-\noperator correspond to energy excitations of the order\n∆E=ε2¯hω≈ε2·41A−1/3MeV for a nucleus with mass\nnumberAand quadrupole deformation ε2[27]. This en-\nergy can be compared to a typical energy of lowest ro-\ntational state, E2+=¯h2\n2J2(2 +1) ≈90A−5/3MeV, and\nwe see that ∆ E >> E 2+for all considered nuclei. Since\nthecollectivestateandthe correspondingmatrixelement\nare well separated in energy, the risk of double-counting\nstates by adding a rotational band on a microscopically\ncalculated band-head is quite small, cf Ref. [7]. This is\ntrue as long as the excitation energies are smaller than\nenergies corresponding to the temperature, T= ∆E, i.e.\nfor excitation energies smaller than about 10 A1/3MeV\n(forε2= 0.25), giving about 30 MeV for mass A=25\nand 55 MeV for A=160. For higher excitation energies a\nsaturation of the rotational enhancement should set in.\nIn this study we concentrate on lower energies, say up to\nabout10MeV. It is thereforeareasonableapproximation\nto account for rotational enhancement by simply adding\na rotational band on each band head; double-counting of\nstates would not occur.\nRotational states are consequently taken into account\nbyaddingarotationalbandontopofeachintrinsicband-\nhead for deformed nuclei (here defined as nuclei with cal-\nculated quadrupole deformation |ε2| ≥0.05). The rota-\ntional energy of the different angular momentum states,\nI, in the rotational band is given by\nEi\nrot(I,Ki) =I(I+1)−K2\ni\n2J⊥(ε2,∆ip,∆in), (13)\nwhereKiis the angular momentum projection on the\nsymmetry axis of the intrinsic state iupon which the\nrotational band is built. The quadrupole deformation is\ndenoted by ε2while ∆i\npand ∆i\nndenote the proton and\nneutron pairinggapsof the intrinsic state i. The moment\nof inertia around an axis perpendicular to the symmetry\naxisJ⊥(ε2,∆i\np,∆i\nn) is approximated by the rigid-body\nmoment of inertia with deformation ε2, modified by the\ncalculated pairing gaps for the considered state, as given\nin Ref. [28]. Given the angular momentum projection\nKand parity πof the band-head the rotational band\nincludes the following levels\nIπ=\n\nKπ,(K+1)π,(K+2)π,...ifK/ne}ationslash= 0,\n0+,2+,4+,... ifK= 0+,\n1−,3−,5−,... ifK= 0−.(14)\nThe Coriolis anti-pairing effect is neglected and no vir-\ntual crossings of rotational bands are taken into account.4\n0 1 2 3 4 5 6 7 8\nExcitation Energy Eexc [MeV]012345678910Rotational enhancement Krot\n0 1 2 3 4 5 6 78\nExcitation Energy Eexc [MeV]0204060Krot\n162Dy\nFIG. 1: (Color online) Rotational enhancement, Krot, for the\nnucleus162Dy as a function of excitation energy. The insert\nshows the enhancement compared to the simple enhancement\nmodel of Ref. [29] (red dashed line).\nThus, the pairinggap and moment ofinertia areassumed\nto be unchanged from the band-head pairing gap for all\nstates in the rotational band. This approximation is rea-\nsonable since mainly low-spin states play a role in the\npresent study.\nThe total energy of the calculated level is thus given\nby\nEi=Ei\nmb+Ei\nrot(I,Ki) (15)\nwhereEi\nmbis the energy, Eq. (12), of the intrinsic many-\nbody configuration i, andEi\nrot(I,Ki) is the rotation en-\nergy, Eq. (13), of the level with angular momentum I\nbuilt from the intrinsic many-body configuration with\nangular momentum projection Ki. These energies are\nused to calculate the level density according to Eq. (1).\nFig. 1 shows the rotational enhancement, Krot, for the\nwell-deformed nucleus ( ε2=0.26)162Dy (cf. Fig. 12) cal-\nculated as the ratio of the level density when rotations\nare included or excluded. For low excitation energies\nthere are large fluctuations which are artifacts of the low\nlevel density combined with the smoothing procedure of\nSec. IIE. For higher excitation energies ( >∼3 MeV) the\nrotational enhancement is a slowly increasing function,\nof the order of a factor 5 at the neutron separation en-\nergy. This prediction is compared to the SU(3) model\nof Ref. [29], which is shown in the insert of Fig. 1. The\nSU(3) model gives almost an order of magnitude larger\nenhancement for excitation energies in the region of the\nneutron separation energy. We note that a combinatorial\nlevel-density model based on the Nilsson potential gives\na rotational enhancement for162Dy similar to our results\n[30].D. Vibrations\nMany nuclei exhibit low-lying states of vibrational\ncharacter, which are usually of quadrupole or octupole\ntype. Such low-lying vibrational states appear at exci-\ntation energies not much lower than the 2qp excitations\nwhich in a coherent way build up the collective state. We\ntherefore believe that the vibrational enhancement of the\nlevel density, contraryto the rotationalenhancement dis-\ncussed above, must be described microscopically.\nIn order to describe vibrational states the Quasi-\nparticle-Tamm-Dancoff-Approximation (QTDA) is used.\nAccording to the Brink-Axel hypothesis [31, 32] phonons\nare built on every intrinsic many-body configuration\nEi\nmb. The QTDA equation is solved for each state iin\norder to get all possible phonon excitation energies and\nwave-functions. Thismeans solvingthe QTDAequations\nmillions of times for every single nucleus.\nThe residual interaction is approximated by the dou-\nble stretched (isoscalar) Quadrupole-Quadrupole inter-\naction. This interaction is well defined in the case of a\nharmonicoscillatorpotential. Inthecaseofafinite-depth\npotential as the folded-Yukawa potential the interaction\nshould take into account additional finite size effects, for\nexample as is done in Ref. [33]. In the present work the\nfinite-depth effects are ignored and the double stretched\napproach is used as defined in Refs. [34, 35].\nThe QTDA secular equation can be written [36]\n1\nχ2K=/summationdisplay\nµν/vextendsingle/vextendsingle/angbracketleftbig\nµ/vextendsingle/vextendsingle¯Q2K/vextendsingle/vextendsingleν/angbracketrightbig/vextendsingle/vextendsingle2(UµVν+VµUν)2\n(Eqp,i\nµ+Eqp,i\nν)−(¯hω)i\nj,(16)\nwhere the effect of Eq. (8) has not been explicitly written\nout.¯Q2Kis the double stretched quadrupole operator,\nwhere the components K= 0 and K= 2 are considered.\nThe set of roots {(¯hω)i\nj}of this equation is the excitation\nenergies of the vibrational phonons jon top of the intrin-\nsic state i, whereas the poles Eqp,i\nµ=/radicalbig\n(eµ−λ)2+∆2\ni\nare the unperturbed two-quasiparticle excitations on the\nmany-body configuration Ei\nmbwith pairing gap ∆ ias\ncalculated by Eqs. (9) and (10), and eµare the single-\nparticle energies.\nThe self-consistent coupling strength is given by [34]\nχ2K=8π\n5Mω2\n0\nA/an}bracketle{t¯r2/an}bracketri}ht+g2K/radicalBig\n4π\n5A/angbracketleftbig¯Q20/angbracketrightbig,(17)\nwhereg20= 1 and g22=−1. The expectation-values/angbracketleftbig\n¯r2/angbracketrightbig\nand/angbracketleftbig¯Q20/angbracketrightbig\nare calculated in double stretched co-\nordinates [34]. To test that this gives overall reason-\nable result, we have verified for a large number of nuclei\nthat for the K= 0 and the K= 2 components of the\niso-scalar Giant Quadrupole Resonances the calculations\nagree well with the systematics of the Giant Resonance\nenergies ¯ hωGQR= 58A−1/3MeV [34].\nThe phonons are never repeated and double counting\nis explicitly avoided by the following procedure. The5\nphonon wave functions are given by\nO†=/summationdisplay\nµ,νXµ,νa†\nµaν, (18)\nwhereXµ,νare the wave-function components of all ex-\ncited quasi-particle states a†\nµaνon top of the configura-\ntionEi\nmb. The level density is increased by one state at\nthe energy of the phonon (¯ hω)i\nj, and decreased by the\namount given by the wave-function component X2\nµ,νat\nthe energy of the corresponding pole ( Eqp,i\nµ+Eqp,i\nν). The\nchange in level density due to the phonon jon top of in-\ntrinsic state iis thus\nδρ(E) =δ(E−Ei\nmb−(¯hω)i\nj)−\n/summationdisplay\nµ,νX2\nµ,νδ/parenleftbig\nE−Ei\nmb−(Eqp,i\nµ+Eqp,i\nν)/parenrightbig\n,(19)\nwhereEis the excitation energy relative to the ground-\nstate. This procedure increases the level density by re-\nducing the strength of the pure quasi-particle excitations\nand adding strength by means of the phonons (which are\npushed down in energy due to the QQ-interaction).\nMost phonons have little collectivity in the QTDA ap-\nproximation, as most of the phonon excitations lie very\nclose to a pure quasi-particle excitation, and hence the\nwave-function is completely dominated by that single\ncomponent. Even if there are a few very collective low-\nlying phonon states the vibrational enhancement of the\nlevel density becomes small, since most excited states\nprovide quite non-collective phonon excitations. Due to\nthis non-collectivityofmostphonons thereis no waythat\nthe phonons, in general, can be repeated to form two- or\neven three-phonon excitations.\nThe vibrational enhancement factor in this method is\nin general quite small, of the order of a few percent. This\nissubstantiallylowerthanpredictionsfromothermodels.\nFor example, the attenuated phonon method [12, 14, 37]\nand the Boson partition function method [38] both give\nup to an order of magnitude enhancement at the neu-\ntron separation energy. Fig. 2 shows the vibrational en-\nhancement as a function of excitation energy for162Dy.\nThe effect is very small, close to 1 % at 7 MeV excita-\ntion energy. For the same nucleus the attenuated phonon\nmethod gives an enhancement factor of about 3 as shown\nin the insert of Fig. 2.\nThe level-density enhancement due to quadrupole vi-\nbrations is thus found to be very small. Higher multipole\nvibrations (as octupole vibrations) are expected to con-\ntribute with enhancements of the same order of magni-\ntude or less and are neglected in the CFY model.\nE. Many-body damping width\nIn the mean-field approach all excited many-body\nstates are treated as non-interacting. A residual two-\nbody interaction will mix the many-body states obtained0 1 2 3 4 5 6 7 8\nExcitation Energy Eexc [MeV]11.0051.011.0151.02Vibrational enhancement Kvib\n0 1 2 3 4 5 6 78\nExcitation Energy [MeV]11.522.533.5Kvib162Dy\nFIG. 2: (Color online) Vibrational enhancement, Kvib, us-\ning the QTDA method for the nucleus162Dy as a function\nof excitation energy. The insert shows the enhancement com-\npared to the enhancement of the attenuated phonon model of\nRef. [12] (red dashed line).\nfrom the combinatorics. Smearing effects from the resid-\nual interaction can approximately be taken into account\nby assuming a spreading width of all excited states. The\nspreading width is implemented in terms of a Gaussian\nenvelope with width σ, i.e. the delta functions in Eq. 1\nare replaced by Gaussians. Estimates of the spreading\nwidth FWHM gives [39]\nΓ = 0.039/parenleftbiggA\n160/parenrightbigg−1/2\nE3/2MeV,(20)\nwhere the FWHM is related to the Gaussian spreading\nwidthσ=Γ\n2√\n2ln2.\nAssuming that all excited many-body states in an en-\nergy bin ∆ Eare uniformly distributed, the level density\nbecomes\nρ(Eb) =/summationdisplay\naρ(Ea)1\n2/bracketleftbigg\nerf/parenleftbiggEa+∆E/2−Eb√\n2σ/parenrightbigg\n−\n−erf/parenleftbiggEa−∆E/2−Eb√\n2σ/parenrightbigg/bracketrightbigg\n,(21)\nfor bin-point b. The method implies a smearing out of\nlevel-densitypropertiesovera rangeΓ, which is smoothly\nincreasing with excitation energy. As a consequence fluc-\ntuations of energies and wave-functions in the range Γ\nfollow GOE statistics of Random Matrix theory, that is\noften denoted as quantum chaos in the nucleus, see e.g\nRefs. [40, 41].\nIII. LEVEL DISTRIBUTIONS\nIn the present combinatorial approach it is possible\nto extract different distributions exhibiting details of the\nlevel-density function. We present here microscopically6\ncalculated distributions for pairing gaps (Sec. IIIA), par-\nity (Sec. IIIB), and angular momentum (Sec. IIIC).\nA. Pairing-gap distribution\nThe BCS equations, Eqs. (9) and (10), are solved for\nall individual many-body configurations and hence pair-\ning gaps for all states are obtained. In Fig. 3 the dis-\ntribution of proton pairing gaps for162Dy is shown for\na number of excitation energies. For the lowest excita-\ntion energies only the ground-state and the states in the\nground-state rotational band exist. The pairing gap of\nthe levels in the rotational band is fixed to be the same\nas the pairing gap of the band-head, see Sec. IIC. As the\nexcitation energy increases levels with reduced pairing\ngaps appear. However, no transition to a completely un-\npaired system is observed, and levels with non-collapsed\ngaps (∆>0) surviveto the highest considered excitation\nenergies. At the highest considered excitation energy for\n162Dy,Eexc= 8.4MeV, 36%ofthe levelsstill havea non-\nzeroprotonpairinggap,seeFig.3. Themeanvalueofthe\nproton pairing gap /an}bracketle{t∆/an}bracketri}htat different excitation energies is\nshown in Fig. 4 for162Dy. Between 2 MeV and 3.5 MeV\nexcitation energy there is a rapid decrease in the mean\npairing gap. At higher excitation energies the decrease\nis slower. At 8.4 MeV excitation energy the mean value\nis/an}bracketle{t∆/an}bracketri}ht= 0.2 MeV. The reason why pairing may survive\nand be of substantial size in states at these high ener-\ngies can be explained as follows: Several highly excited\nstates are built by exciting particles and holes far from\nthe Fermi energy, where blocking in the BCS equations\nhas no effect on the pairing gap.\nThe non-collapsed pairing gaps influence the moment\nof inertia and keep it reduced as compared to the rigid-\nbody value. This implies that even at excitation energies\nin the region of the neutron separation energy the mo-\nment of inertia is on average smaller than the rigid-body\nvalue. A similar observation is made in Ref. [15].\nB. Parity distribution\nIn Fermi-gas level-density models there is an implicit\nassumption of equal number of states with different par-\nity at any given excitation energy. However, models that\ntake into account microscopic effects show clear struc-\nture in the parity ratio, see eg. Refs. [16, 42, 43]. In\nconnection with astrophysical reaction rates the parity\nratio may play an important role, and can be included in\nthe Hauser Feshbach formalism, see eg Refs. [44, 45].\nFrom the single-particle point of view it is clear that\nthe parityratioshould showstructure. For84\n38Sr46the sit-\nuationisilluminating. Thenucleusisclosetosphericalin\nits ground-state ( ε2= 0.05) [21]. It has 2 proton holes in\nthepfshell and 6 neutrons in the g9/2shell. Since there\nare large gaps in the single-particle spectrum the energy\nto excite nucleons across the gaps, which cause changes0246\n 67 % Eexc= 0.7 MeV\n024 71 % Eexc= 1.8 MeVDistribution162Dy\n024 59 % Eexc= 2.9 MeV\n0 0.25 0.5 0.75024 49 % Eexc= 4.0 MeV 47 % Eexc= 5.1 MeV\n 42 % Eexc= 6.2 MeV\n 39 % Eexc= 7.3 MeV\n0 0.25 0.5 0.75 1 36 % Eexc= 8.4 MeV\nPairing Gap ∆ [MeV]\nFIG. 3: (Color online) Proton pairing-gap distributions at\ndifferent excitation energies Eexcfor162Dy shown in 20 keV\npairing-gap bins. The proportion of paired states (∆ >0) are\nshown in percent in the boxes.\n0 1 2 3 4 5 6 7 8 9\nExcitation Energy Eexc [MeV]00.10.20.30.40.50.6Average Proton Pairing Gap < ∆> [MeV]162Dy\nFIG.4: Average protonpairinggap /angbracketleft∆/angbracketrightat differentexcitation\nenergies Eexcfor162Dy. The mean values are given by the\ndistributions in Fig. 3.\nin parity, is quite large. The parity ratio displays long-\nrangeoscillations,seeupperleft panelofFig.5, whichare\ndirectly connected to the large single-particle gaps. The\nsingle-particle gaps effectively decrease with increasing\ndeformation, and the oscillations are therefore expected\nto decrease with deformation. To test these ideas the\nparityratio in84Sr is calculated at different deformations\nand shown in Fig. 5. Indeed, for larger deformations the\noscillatory pattern vanishes and the parity ratio equili-\nbrates to one.\nIn Fig. 6 the parity ratios for56Fe,60Ni and68Zn are\nshownandcomparedtotwocalculationsofRef.[16]. The\nsolid blue lines show the CFY model, the red dashed\nlines show the statistical parity projection model and\nthe black dots show calculations using the Shell-Model\nMonte-Carlo method. The parity ratio in the Fe and7\n0123\nε2=0.0584Sr ε2=0.20\n012Parity Ratio ρ-/ρ+\nε2=0.10 ε2=0.25\n0 5 10012 ε2=0.15\n0 5 10 15ε2=0.30\nExcitation Energy Eexc [MeV]\nFIG. 5: (Color online) Parity ratio versus excitation energ y\nfor84Sr calculated with different deformations. Large defor-\nmations imply a parity distribution close to 1 even at low\nexcitation energies. For small deformations the parity non -\nequilibrium can survive to high excitation energies.\n00.511.52\n56Fe, ε2=0.00\n60Ni, ε2=0.025\n68Zn, ε2=-0.1500.511.5Parity Ratio ρ-/ρ+\n5 10 15 20\nExcitation Energy Eexc [MeV]00.511.5\nFIG. 6: (Color online) Parity ratio versus excitation energ y.\nThe solid blue lines are calculated with the CFY model. The\nred dashed line and black dots are given by a statistical mode l\nand a Monte-Carlo method, respectively [16].\nNi isotopes is not equilibrated below 15 MeV in any of\nthe calculations, while68Zn equilibrates at much lower\nenergies (below 10 MeV). For60Ni and68Zn there is\na clear oscillatory behavior prior to equilibration in the\nCFY model, and in the case of Ni there is a good agree-\nment between the Monte-Carlo method and the CFY\nmodel, especially in the region of 8–12 MeV excitation\nenergy. Note, however, that fluctuations seen in the\npresent micro-canonical approach may be smeared out\nby the grand-canonical approach as used in Ref. [16].\nThe parity ratio has been calculated for Sr-isotopes\nwithin a statistical method in Ref. [45] where the im-\npact on astrophysical reaction rates was investigated and\nfound to be small. The statistical method gives at most\none oscillatorymaximum before it equilibrates, asseen in0123\n76Sr, ε2=0.37582Sr, ε2=0.05\n012Parity Ratio ρ-/ρ+\n78Sr, ε2=0.37584Sr, ε2=0.05\n0 5 10 1501280Sr, ε2=0.05\n0 5 10 15 2086Sr, ε2=0.05\nExcitation Energy Eexc [MeV]\nFIG. 7: (Color online) Parity ratio versus excitation energ y\nfor Sr isotopes. The solid blue lines are obtained from the\nCFY model and the dashed red lines are obtained from the\nstatistical parity projection of Ref. [45].\nFig. 7 (dashed lines). In the CFY model the parity ratio\nhassubstantiallymorestructure(solidlines). Nucleiwith\nsmall ground-state deformations show long-range oscil-\nlations which survive to high excitation energies before\nequilibration. The overall results are quite different from\nthe model of Ref. [45].\nC. Angular momentum distribution\nIn Fermi-gas models the distribution of angular mo-\nmentum is given by the Gaussian envelope in the spin\ncutoff model, see eg. Ref. [4],\nF(U,I) =2I+1\n2σ2exp/parenleftbigg−I(I+1)\n2σ2/parenrightbigg\n,(22)\nwhich is obtained by random coupling of uncorrelated\nspins of the nucleons. In Eq.(22) the spin cutoff factor,\nσ, is defined by\nσ2=Jrigid\n¯h2/radicalbigg\nU\na, (23)\nwhereJrigidis the rigid-body moment of inertia, U=\nE−δis the effective excitation energy shifted by the\nback-shift δ, andais the level-density parameter.\nIn Fig. 8 the angular-momentumdistributions for68Zn\nand162Dy are shown for several excitation-energy re-\ngions. The black lines with dots show the CFY model\nresults while the red solid lines show the Gaussian dis-\ntribution of Eq. (22) with the spin cutoff factors given in\nRef. [4], and the blue dashed lines show Gaussian distri-\nbutions fitted to the combinatorial calculation. For low\nexcitation energies there are clear deviations from the\nGaussian profiles while for higher excitations the combi-\nnatorial distribution tends to the Gaussian profile. How-\never, for68Zn the deviations from a Gaussian angular8\n00.050.10.150.20.25\nEexc = 3.02 MeV Eexc = 4.86 MeV\n0 5 10 1500.050.10.150.2\nEexc = 6.76 MeVAngular-momentum distribution\n0 5 10 15 20Eexc = 8.67 MeV\nAngular Momentum I68Zn\n00.050.10.150.2\nEexc = 2.63 MeV Eexc = 4.28 MeV\n0 5 10 1500.050.10.15\nEexc = 5.94 MeV\n0 5 10 15 20Eexc = 7.60 MeVAngular-momentum distribution\nAngular Momentum I162Dy\nFIG. 8: (Color online) Angular-momentum distribution for\n68Zn and162Dy in 4 different excitation-energy regions. The\nblack lines with dots show results from the CFY model. The\nred solid lines and blue dashed lines show the Gaussians give n\nby Eq. (22) using the spin cutoff factors from Ref. [4] and a\ndirect fit to CFY, respectively.\nmomentum distribution remain up to about the neutron\nseparation energy ( Sn=6.48 MeV).\nForlowexcitationenergiesin68Znthereisanodd-even\nspin staggering which is not explained by the spin cutoff\nmodel. This effect has also been observed in Fe-isotopes\nin the Shell Model Monte-Carlo Method [17].\nFig. 9 shows the spin cutoff factor deduced from the\nCFY model (black solid line) and from the statistical\nmodel [4] (red dashed line), as a function of excitation\nenergy for162Dy. For this particular nucleus the spin\ncutoff factor for excitation energies >∼3 MeV is similar in\nshape but ∼10 % larger than in the statistical model [4].\nThe fact that the statistical model is smaller than the\nCFY model, despite the presence of pairing in the CFY\nmodel, is accidental for this nucleus. Since both the rigid\nbody moment of inertia and the level density parameter\nare fitted to experiments using only three parameter, the\ndetailed description for a particular nucleus might give\nthis result when compared to a very different model as01234567Spin cutoff factor σ\n0 1 2 3 4 5 6 7 8\nExcitation Energy Eexc [MeV] 0.50.7511.25σCFY/σStat162Dy\nFIG. 9: (Color online) The top panel shows the spin cutoff\nfactor as a function of excitation energy for162Dy in the CFY\nmodel (black solid line) and the statistical model of Ref. [4 ]\n(red dashed line). The bottom panel shows the ratio of the\ncombinatorial and statistical spin cutoff factors of the top\npanel.\nthe CFY model.\nIV. COMPARISON WITH EXPERIMENTAL\nDATA\nA. Neutron separation-energy level spacings\nThe s-wave neutron resonance spacings constitute the\nmost comprehensive experimental database for compar-\nison with NLD calculations [46]. This database serves\nas a benchmark for all large-scale level-density models\n[4, 14, 15, 47].\nThe s-wave neutron resonance spacing D0at the\nneutron separation energy Snof the compound nu-\ncleus (Z,N) is obtained from calculated level densities\nρ(E,I,π) as,\n1\nD0=\n/braceleftbigg\nρ(Sn,I0+1/2,π0)+ρ(Sn,I0−1/2,π0) forI0>0,\nρ(Sn,1/2,π0) for I0= 0,\n(24)\nwhereI0is the ground-state spin and π0is the ground-\nstate parity of the target nucleus ( Z,N−1). In Fig. 10\nwe study the ratio of calculated and experimental level\nspacings at the neutron separation energy, DTh/DExp,\nfor all nuclei in the database.\nThe quality of a level-density model can be estimated\nby the rms-factor [15]\nfrms= exp/bracketleftBigg\n1\nNeNe/summationdisplay\ni=1ln2Di\nTh\nDi\nExp/bracketrightBigg1/2\n,(25)9\n0 50 100 150 200 250 300\nMass Number A0.010.1110100DTh/DExp\nFIG. 10: Ratio between theoretical and experimental level\nspacings at neutron separation energy versus mass number\nA. The rms-factor is frms= 4.2. The experimental data are\ntaken from Ref. [46].\nand the mean factor\nm= exp/bracketleftBigg\n1\nNeNe/summationdisplay\ni=1lnDi\nTh\nDi\nExp/bracketrightBigg\n, (26)\nwhereDi\nThandDi\nExpare the theoretical and experimen-\ntal level spacings and Neis the number of nuclei in the\ndatabase. The CFY model gives frms= 4.2 andm= 1.1,\nsee Fig. 10 of the error ratio versus A. These results are\ncompared to other statistical and combinatorial models\nin Table II and discussed in Sec. V.\nThereseems to be a clearresidual shellstructure in the\nlevel spacings, as seen in Fig. 10, especially in the doubly\nmagic208Pb region. A similar effect can be observed\nfor the Gogny model in Ref. [15], and in results of the\nSkyrme model with the BSk9 interaction, as shown in\nRef. [14]. In addition, there seems to be a drift with\nmass number where especially nuclei in the rare-earth\nandactinideregionsover-estimatetheneutronseparation\nlevel spacings. A similar drift can be seen in the Gogny\nmodel of Ref. [15] while it seems not to be present in the\nSkyrme-HFB model in Ref. [14].\nThe rare-earth and actinide regions, where too large\nlevel spacings are calculated, mainly correspond to de-\nformed nuclei. On the other hand, nuclei corresponding\nto regions with particularly low ratio DTh/DExpseen in\nFig. 10(around A=32, 132and 208)correspondto spher-\nical nuclei. The overestimation of Dfor deformed nuclei\n(too low calculated level density), and underestimation\nofDfor spherical nuclei (too high calculated level den-\nsity) seems to be approximatively systematic, as seen in\nFig. 11, where the ratio between theoretical and experi-\nmental level spacings is shown versus the absolute value\nof the deformation.\nHowever, as seen in the lower panel of Fig. 11 there\nis no clear correlation with the microscopic energy Emic\nof Ref. [21] as might be expected from the residual shell\nstructure seen in Fig. 10.0 0.1 0.2 0.3 0.4\nDeformation | ε2|0.010.1110100DTh/DExp\n-15 -10 -5 0 5\nMicroscopic Energy Emic [MeV]0.010.1110100DTh/DExp\nFIG.11: (Color online)Thetoppanelshowstheratiobetween\ntheoretical and experimental level spacings at neutron sep a-\nration energy versus the absolute value |ε2|of the calculated\nquadrupole deformation. The bottom panel shows this ratio\nas a function of the microscopic energy. The solid red line\nshows an exponential fit to illustrate the correlation. The m i-\ncroscopic energies and the quadrupole deformations are tak en\nfrom Ref. [21]. The experimental data are from Ref. [46].\nB. Detailed level-density functions\nThe Oslo method is a commonly used experimental\nmethod for extracting detailed level-density functions for\nlarge ranges of excitation energy [19]. It provides a valu-\nable test for NLD models. However, the approach is\nmodel dependent since the level density is extracted by\nuse of a back-shifted Fermi Gas approximation and a\nspin-cutoff factor, see also discussion in Ref. [38].\nIn Figs. 12 and 13 level densities obtained from the\nCFY model are compared to data for a number of nuclei\nwhere experimental data are available [48, 49, 50, 51, 52,\n53].\nIn Fig. 12 we show level densities of the rare-earth nu-\nclei148,149Sm,161,162Dy,166,167Er, and170,171,172Yb as\nfunctions of excitation energy. The data are in general\nwell reproduced by the CFY model, with an error of less\nthan a factor of 2. The good agreement between the cal-\nculatedand experimentalslopesindicatesthat the single-\nparticle structure and the moments of inertia in the rota-\ntional bands are sound. However, an observable trend in\nthismassregionisthatthe leveldensityforeven-evennu-\nclei is slightly over-estimated and for odd nuclei the level\ndensity is slightly under-estimated. The effective back-\nshift is to a large extent controlled by the ground-state\npairing gap. By fine-tuning the pairing gaps it is possi-\nble to get an almost perfect agreement with experiment.\nHowever, in this paper no efforts are made to adjust pa-\nrameters of the model to fit measured level-density data.\nIndeed, the ground-state pairing gaps are given by the\nmass model, see Sec. IIB. Neither are any renormaliza-\ntions of calculated level densities to fit data performed,\nas is sometimes done in other models, see Refs. [14, 38].10\n100102104106108\n100102104106\n100102104106\n0 1 2 3 4 5 6 78\nExcitation Energy Eexc [MeV]100102104106149Sm\n148Sm\n161Dy\n162Dy\n167Er\n166Er\n171Yb\n170Yb, 172YbLevel density ρ [MeV-1]\nFIG. 12: (Color online) Level densities ρas functions of ex-\ncitation energy for Sm, Dy, Er and Yb isotopes. The black\nsolid lines show the CFY model and the red dots show the\nexperimental data [50, 51, 52, 53].\nExperimental level-density functions are available also\nfor lighter mass regions, and in Fig. 13 data for V and\nMo isotopes are shown. The overall agreement between\nthe model and these experimental data is somewhat in-\nferior to what was obtained in the rare-earth region. For\nthe Mo isotopes in Fig. 13 the CFY model is roughly a\nfactor of 3 too large for high excitation energies while\nfor the V isotopes the over-estimation is roughly a fac-\ntor of 4. However, these errors are consistent with the\noverall results from the neutron separation level spac-100102104106Level density ρ [MeV-1]\n0 1 2 3 4 5 6 78\nExcitation Energy exc [MeV]100102104Level density ρ [MeV-1]50V\n51V\n97Mo\n98Mo\nFIG. 13: (Color online) Level densities as functions of exci ta-\ntion energy for50,51V and97,98Mo. The black solid lines show\nthe CFY model and the red dots show the experimental data\n[48, 49].\ning, see Fig. 10, which have an overall frms= 4.2. Also,\nthe slopes and the detailed structures are not as well de-\nscribed in these nuclei as in the rare-earths. Especially\n50V exhibits an oscillatory pattern in the CFY model,\nwhich is an effect of the small ground-state deformation\nε2= 0.05 [21]. A corresponding oscillatory pattern is not\nas clearly present in the experimental data. In51V the\nexperimental data shows oscillatory behavior while the\nCFY model is more smooth. The ground-state deforma-\ntion is slightly larger with ε2= 0.083 [21], which in the\nCFY model means that since the moment of inertia is\nlarger the rotational states have a larger influence and\nsmooths the level density.\nC. Low-lying discrete levels\nThe amount of experimental information on low-lying\ndiscrete energy levels is indeed substantial. In addition\nto the energy parity and angular momentum are often\nknown. In Ref. [22] a global comparison is made between\nmeasuredand FRLDM calculatedground-statespinsand\nparities for odd-mass nuclei, and the agreement was\nfound to be very good. The collection of low-energy data11\nprovides important information for level-density models,\nsince the accuracy of the model can be studied at the\nvery lowest energies. Cumulative distributions of known\nlevels can thus be constructed, and compared to calcula-\ntions. In Ref. [14] a selection of 15 different nuclei was\nmade to represent all different kinds of nuclear aspects,\nsuch as light-heavy, spherical-transitional-deformed, as\nwellasodd-odd,even-evenandodd-evennuclei(alsoused\nin Ref. [38]). We therefore compare our model low-lying\nlevel densities to experimental data for the identical set\nof 15 nuclei.\nThe cumulative level density for the 15 nuclei are\nshown in Fig. 14. The CFY model (solid blue line) is\ncompared to experimental data (black dash-dotted line)\nand to the HFB model (red dashed line) of Ref. [38]. For\na good agreement with data, the calculated line should\nfollow as precise as possible the experimental curve at\nthe lowest energies, and should then smoothly deviate,\nalways larger than data.\nIn general, the theoretical models seem to give very\ncomparable results for deformed nuclei while differences\nare larger for transitional and spherical nuclei. For the\nlight spherical nuclei, like42K and56Fe there are pro-\nnounced single-particle gap effects which cause jumps in\nthe level density. These effects are smoothly disappear-\ning at higher energies. A quite drastic deviation between\nour calculation and data is seen for208Pb. The calcu-\nlated (cumulative) level density is systematically larger\nstarting already at about 3 MeV excitation energy (in\nthe figure seen at 0.5 MeV; the energy scale is shifted by\n2.5 MeV for this nucleus). This is also seen in the to-\ntal level density at the neutron separation energy, and in\nTable I the ratio between theoretical and measured level\nspacingsat the neutron separationenergyis listed for the\n9 nuclei. For208Pb this ratio is 0.02, i.e. the calculated\nlevel density is about 50 times larger than experiment!\nWhen comparing to the HFB model calculation of the\nlevel density [38] one should note that interpolations be-\ntween different deformations are included in this model,\nwhile in our model the ground-state deformation is kept\nfixed for all excited states. In particular, this somewhat\nphenomenological way to treat deformation changes has\nstrong effects on the results for127Te.\nNuclDTh/DExpNuclDTh/DExpNuclDTh/DExp\n42K 0.9456Fe 0.5560Co 0.62\n94Nb 2.50107Cd 0.77127Te 0.26\n148Pm 1.72155Eu 2.95161Dy 2.61\n162Dy 1.27172Yb 3.10194Ir 4.58\n208Pb 0.02237U 5.18239Pu 3.46\nTABLE I: Table of neutron separation energy spacings com-\npared to experimental data for the nuclei showed in Fig. 14.100101102103104\n42K56Fe60Co\n100101102103\n94Nb107Cd127Te\n100101102103Accumulated Level Density N148Pm155Eu161Dy\n100101102103\n162Dy172Yb194Ir\n0 1 2 3 4 5 6100101102103\n208Pb(-2.5 MeV)\n0 1 2 3 4 5 6\nExcitation Energy [MeV]237U\n0 1 2 3 4 5 67239Pu\nFIG. 14: (Color online) Accumulated level density as a func-\ntion of excitation energy. The blue solid and red dashed line s\nshow the present CFY model and the HFB model [38]. The\nblack dash-dotted line shows experimental data collected i n\nRef. [38].\nD. Parity ratio\nThe parity ratio has been measured experimentally by\nKalmykov et.al. [20] for the two spherical nuclei58Ni\nand90Zr. For these nuclei the I= 2 angular-momentum\ncomponentofthe leveldensityismeasuredand separated\ninto parity components. In the top panel of Fig. 15(a)\nthe level-density component ρ(E,I= 2,π=±1) for90Zr\nis compared to CFY model calculations. The level den-\nsities from the CFY model are in good agreement with\nexperimental data, and in the lower panel of Fig. 15(a)\nthe parity ratio is shown. Predictions for the total level\ndensityaswellasthe I= 2componentareshownasblack\nsolid and red dashed lines, respectively. The difference\nbetweentheparityratioforthetotalleveldensityandthe\nI= 2 component decreases with excitation energy, be-\ncause the spin cutoff model becomes more realistic when\nthe excitation energy increases, see Sec. IIIC. For the\nparity ratio the CFY model result is within the experi-\nmental error-barsfor all excitation energies and seems to\nshow a similar pattern as the experiments: a high par-\nity ratio (excess of negative parity states) at 8 MeV and\na low ratio (excess of positive parity states) at 11 MeV\nexcitation energy. The blue dot-dashed line shows the12\n100102104106 Level density ρ [MeV-1]\n2+\n2-\n5 7.5 10 12.5 15\nExcitation Energy Eexc [MeV]012Parity Ratio ρ-/ρ+ 90Zr\n90Zr I = 2\n100102104106 Level Density [MeV-1]\n2+\n2-\n5 7.5 10 12.5 15 17.5 20\nExcitation Energy Eexc [MeV]00.511.5Parity Ratio ρ-/ρ+58Ni\n58Ni I = 2\nFIG. 15: (Color online) The top panels show CFY calculated\ncomponents of the 2+(black solid) and 2−(red dashed) level\ndensities for90Zr (top part of figure) and58Ni (lower part\nof figure), as functions of excitation energy. Experimental\ndata from Ref. [20] are shown as black dots and red diamonds\nwith error-bars for 2+and 2−, respectively. The lower panels\nshowforeachnucleustheparityratioversusexcitationene rgy.\nThe solid black and dashed red lines show the CFY model\nresults for the total level density and the I=2 component,\nrespectively. The blue dot-dashed lines show the Skyrme-\nHFB model of Ref. [14]. Experimental data from Ref. [20] are\nshown with error-bars.\nSkyrme-HFB model of Ref. [14]. It is also within the\nexperimental error-bars for all excitation energies except\nat 9 MeV where the HFB model gives a large positive\nratio (∼3) while the experiments and the CFY model\nare close to unity.\nTheI= 2 component of the level density in58Ni is\nshown in the top panel of Fig. 15(b). The agreement\nbetween experimental data and the CFY model is some-\nwhat inferior to what we obtained for90Zr. For both\nparities the level density seems to increase faster in the\nCFY model than what is seen in the experimental data.\nAt 14 MeV excitation energy the model overestimates\nthe level density by roughly a factor of 6. The parity\nratio is shown in the lower panel of Fig. 15(b). The ex-5 6 7 8 9 10\nExcitation Energy Eexc [MeV]102103104105Level density ρ [MeV-1]90Nb Iπ = 1+\nFIG. 16: (Color online) Level density of the 1+component of\nthe level density as a function of excitation energy for90Nb.\nThe black solid and red dashed lines show the CFY model\nand Skyrme-HFB model of Ref. [14], respectively. Data from\nRef. [20] are shown as black dots with error-bars.\nperimental parity ratio is close to unity at 8 MeV and\nthen decreases with increasing excitation energy. The\nCFY model shows a different pattern. It gives a very\nlow parity ratio at low excitation energies with an al-\nmost monotonic increase with increasing excitation en-\nergy. The ratio only becomes close to unity at 20 MeV.\nThe Skyrme-HFB model of Ref. [14] shows a similar pat-\ntern as the CFY model but with an even lower ratio for\nexcitation energies below 20 MeV.\nThe 1+component of the level density of90Nb has\nalso been measured in Ref. [20]. Fig. 16 shows the ex-\nperimental data compared to the CFY and Skyrme-HFB\nmodels. In the experimental data there is a clear os-\ncillating structure. The CFY model (black solid line)\nover-estimates the level density and the oscillating struc-\nture is much less pronounced. In the Skyrme-HFB model\n(red dashed line) there are long-range oscillations simi-\nlar to what is seen in the experimental data, but the\nenergy separation between consecutive minima is larger.\nThe level density is under-estimated in the Skyrme-HFB\nmodel, whereas the CFY model over-estimates the level\ndensity by a similar factor.\nV. COMPARISON WITH OTHER MODELS\nThe CFY model is here compared to other statistical\nand combinatorialNLD models that provideneutron res-\nonance level spacings. In general, the statistical models\nlistedinTableIIhaveanrmsdeviationjustbelow2. This\nlow rms deviation is obtained because severalparameters\nin the level-density formulas are directly adjusted to the\nneutron separation-energy level spacings and low-lying\ndiscrete energy levels.\nThe back-shifted Fermi model of Ref. [47] is based on\na simple Fermi-gas formula whereas the Constant Tem-13\nStatistical Models frmsRef.\nBack-shifted Fermi Model 1.71 [47]\nConst. Temp. Model 1.77 [47]\nBack-shifted Fermi + Const. Temp. Model 1.7 [4]\nGeneralized Superfluid Model 1.94 [47]\nCombinatorial Models frmsRef.\nSkyrme-HFB 2.35 [47]\nGogny-HFB 4.55 [15]\nCFY 4.2 Present\nTABLE II: Table of rms-factors frmsfor statistical and com-\nbinatorial models for neutron separation energy level spac ings\n[4, 15, 47].\n12345678\nAttenuated Phonon Method\n0 50 100 150 200 250 300\nMass Number A11.11.2 QTDAVibrational enhancement Kvib\nFIG. 17: Vibrational enhancements at the neutron separatio n\nenergy in the attenuated phonon method (top panel) and the\nQTDA method (bottom panel) versus of mass number A.\nperature model is based on the approach of Gilbert and\nCameron [10]. Both models contain explicit enhance-\nment factors for rotations and vibrations. The model\nby Rauscher, Thielemann and Kratz [4] is also based on\nthe approach of Gilbert and Cameron. In contrast to\nthe first two models this model has no explicit enhance-\nment factors for rotations and vibrations. It accounts for\nshell effects and thermal damping in terms of an effective\nlevel-density parameter. The model uses the microscopic\nenergy corrections of Ref. [21] together with three free\nparametersin the fitting procedure. The GeneralizedSu-\nperfluid Model is similar to the models mentioned above.\nIn addition it takesinto account howpairingevolveswith\nincreasing excitation energy. It also incorporates explicit\nrotational and vibrational enhancements [47].\nPredictions by statistical models in regions outside\nthe fitting region probably give much less accurate re-\nsults than in the adjustment region. On the other hand,\nsince the combinatorial models are all based on calcu-\nlated single-particle spectra they could in principle have\nbetter predictive power, in particular in regions wherethe single-particle model is sound. In contrast, they are\nnot equally flexible in terms of parameter fits to the level\nspacings database. Therefore, the rms deviation factor\nof combinatorial models with respect to known data is\nlarger than for the statistical models. The frms= 4.2\nfor the CFY model is about a factor of 2 larger than in\nstatistical models. However, the latter result is obtained\nwithnoparameters specifically fitted to the level-density\ndata. The mean deviation factor m= 1.1 indicates that\ntheleveldensityisonaveragereasonable. However,there\nseems to be a drift with mass number, as seen in Fig. 10,\npresent in the CFY model which is not seen in the HFB\nmodels of Refs. [15, 47].\nIn the lower part of Table II the CFY model is com-\npared to two other large-scale combinatorial NLD mod-\nels based on the HFB method. The model of Ref. [15]\nis based on the Gogny D1S interaction for the mean-\nfield and incorporates combinatorial rotations (similar to\nSec. IIC but no pairing dependence of the moment of\ninertia), and the attenuated phonon method is applied\nto account for vibrational states. Pairing is included\nexplicitly for the ground-state, and excited states are\nback-shiftedbyan energy-dependentgapprocedure. The\nmodel gives frms= 4.55 for the subset of even-even axi-\nally deformed nuclei. The erroris slightly largerthan the\nCFY model.\nFor the Skyrme-HFB NLD model the deviation factor\nisfrms= 2.35 [47], which is ���35% larger than the sta-\ntistical models and 44% smaller that in the CFY model.\nThis model is based on a Skyrme-HFB mean-field to-\ngether with combinatorial rotations and the attenuated\nphonon method for modeling vibrational states. In ad-\ndition a phenomenological deformation change from de-\nformed to spherical shape at some specified excitation\nenergy is incorporated.\nThe attenuated phonon method is a phenomenological\nway to model nuclear vibrations, see eg. Refs. [12, 14]. It\nassumes that the quadrupole and octupole phonon states\nin nuclei can be modeled by a gas of non-interacting\nbosons. The vibrational excitation energies are taken\nfrom systematics of the lowest non-rotational 2+and 3−\nstates. Themodel is formulatedasa multiplicativefactor\nKvibr= exp[δS−δU/T] which enhances the level den-\nsity.Tis the nuclear temperature and the δSandδUare\nthe entropy and internal energy change induced by the\nbosons. The occupation probabilities of the bosons are\ndescribed by damped Bose statistics, where the damping\nsets in at considerably higher energies than the neutron\nseparation energy. At excitation energies close to or be-\nlow the neutron separation energy the damping has neg-\nligible effect which implies that the phonons are allowed\nto be repeated several times.\nThe vibrational enhancement at the neutron separa-\ntion energy is quite small in the QTDA method as com-\npared to the attenuated phonon method, see Fig. 17.\nThe QTDA gives an enhancement of the order of a few\npercent compared to up to a factor 7 in the attenuated\nphonon method. The largest vibrational enhancements14\nare obtained for the A=75 region and for Cd isotopes in\ntheA=115 region, which is consistent with the strong\n(quadrupole) vibrational character of nuclei in these re-\ngions.\nThe QTDA phonons are microscopically built from\nquasi-particle excitations Eqp,i\nµwith energies not much\ndifferent than the QTDA phonon energies (¯ hω)i\nj. The\nwayto fully accountfordouble-countingofphononstates\n(Eq. 19), thus implies a small vibrational enhancement\nsince the existence of a few low-lying collective phonons\nwill not impact the level density if all other phonons\nare very non-collective. The non-collective character of\nmost of the phonons also implies that they cannot be re-\npeated. Phononsin theattenuatedphononmodelarenot\ndescribed microscopically, and double-counting of states\nmay clearly appear. In addition, it assumes that each\nphonon can be repeated several times.\nVI. SUMMARY\nA combinatorial model for the nuclear level density\nis presented. The model is based on the folded-Yukawa\nsingle-particlemodelwithground-statedeformationsand\nparameters from Ref. [21]. The model is used to calcu-\nlate the neutron resonancelevelspacings, yieldingan rms\ndeviation of frms= 4.2. It also compares favorably with\nexperimental level density data versus excitation energy\nfor several nuclei in the rare-earth region as measured\nby the Oslo method, as well as with the cumulative level\ndensity extracted from low-energy spectra.\nThe role of collective enhancements has been investi-\ngated in detail. Pairing is incorporated for each individ-\nual many-body configuration, and the distribution of the\npairinggapsisinvestigated. Nosharppairingphasetran-\nsition is observed. Instead, even at the highest excitation\nenergy considered, a non-negligible fraction of the states\nhave a considerable pairing gap.Rotational states are included combinatorially by a\nsimple rotor model with a moment of inertia dependent\non the deformation and pairing gap. Vibrational states\nare included using a Quasi-particle-Tamm-Dancoff Ap-\nproximation. It is found that the vibrational enhance-\nment in the QTDA model is very small, on the order of\na few percent at the neutron separation energy.\nThe parity distribution in the CFY model shows large\noscillatory patterns for nuclei which have large gaps in\nthe single-particle spectrum, separating shells with dif-\nferent parities. This is often the case for nuclei with\nsmalldeformations. Thepatternsarequitedifferent from\nthe smooth pattern of the statistical parity distribution\nmodel of Ref. [16]. The CFY model is compared to other\nmodels and to experimental data when available.\nThe angular-momentum distribution in the CFY\nmodel is compared to the spin cutoff model in Sec. IIIC.\nIt is found that the Gaussian envelope of the spin cutoff\nmodel is in good agreement with the CFY model for high\nexcitation energies.\nAcknowledgments\nWe acknowledge discussions with R. Capote, M. Gut-\ntormsen, K.-L. Kratz, T. Rauscher, A. Richter and F.-K.\nThielemann. S. Hilaire and S. Goriely are acknowledged\nfor providing data underlying Fig.14.\nH. U. is grateful for the hospitality of the Los Alamos\nNational Laboratory during several visits. S. ˚A. and H.\nU. thank the Swedish national research council (VR) for\nsupport. This work was supported by travel grants for\nP. M. to JUSTIPEN (Japan-U. S. Theory Institute for\nPhysics with Exotic Nuclei) under grant number DE-\nFG02-06ER41407 (U. Tennessee). This work was par-\ntially carried out under the auspices of the National\nNuclear Security Administration of the U. S. Depart-\nment of Energy at Los Alamos National Laboratory un-\nder Contract No. DE-AC52-06NA25396 and DE-FC02-\n07ER41457.\n[1] W. Hauser and H. Feshbach, Phys. Rev. 87(1952) 366.\n[2] T. Rauscher and F.-K. Thielemann, Atomic Data and\nNuclear Data Tables, 75(2000) 1.\n[3] P. M¨ oller, et. al., Phys. Rev. C 79(2009) 064304\n[4] T. Rauscher, F.-K. Thielemann and K.-L. Kratz, Phys.\nRev.C 56(1997) 1613.\n[5] H.A. Bethe, Rev. Mod. Phys. 9(1937) 69.\n[6] T. Ericson, Adv. Phys. 9(1960) 425.\n[7] S. Bjørnholm, A. Bohr and B. Mottelson, Physics and\nChemistry of Fission , Proceedings of a Conference at\nRochester (IAEA, Vienna), (1974) Vol. 1, p. 367.\n[8] A. Bohr and B. R. Mottelson, Nuclear Structure, Vol-\nume II(Benjamin, Reading, MA, 1975).\n[9] V. G. Soloviev, Nucl. Phys. 69(1965) 1.\n[10] A. Gilbert and A.G.W. Cameron, Can. J. Phys. 43\n(1965) 1446.\n[11] A. Bohr and B. R. Mottelson, Nuclear Structure, Vol-ume I(Benjamin, Reading, MA, 1969).\n[12] A.V. Ignatyuk, J.L. Weil, S. Raman and S. Kahane,\nPhys. Rev. C 47(1993) 1504.\n[13] P. Leboeuf and J. Roccia, Phys. Rev. Lett. 97(2006)\n010401.\n[14] S. Hilaire and S. Goriely, Nucl. Phys. A 779(2006) 63.\n[15] S. Hilaire, J. P. Delaroche and M. Girod, Eur. Phys. J.\nA 12(2001) 169.\n[16] Y. Alhassid, G. F. Bertsch, S. Liu and H. Nakada, Phys.\nRev. Lett. 84(2000) 4313.\n[17] Y. Alhassid, S. Liu and H. Nakada, Phys. Rev. Lett. 99\n(2007) 162504.\n[18] M. Horoi, J. Kaiser and V. Zelevinsky, Phys. Rev C 67\n(2003) 054309.\n[19] A.Schiller et. al., Nucl. Instr. Meth. Phys. Res. A 447\n(2000) 498.\n[20] Y. Kalmykov et.al., Phys. Rev. Lett. 99(2007) 202502.15\n[21] P. M¨ oller, J. R. Nix, W. D. Myers and W. J. Swiatecki,\nAtomic Data and Nuclear Data Tables, 59(1995) 185.\n[22] L. Bonneau, P. Quentin and P. M¨ oller, Phys. Rev. C 76\n(2007) 024320.\n[23] P. M¨ oller, J. R. Nix and K.-L. Kratz, Atomic Data and\nNuclear Data Tables, 66(1997) 131.\n[24] T. Døssing et. al., Phys. Rev. Lett. 75(1995) 1276.\n[25] P. M¨ oller and J.R. Nix, Nucl. Phys. A 536(1992) 20.\n[26] H. Olofsson, R. Bengtsson and P. M¨ oller, Nucl. Phys. A\n784(2007) 104.\n[27] S.G. Nilsson, Kgl. Danske Videnskab. Selskab. Mat.-Fy s.\nMedd.29: No. 16 (1955).\n[28] R. Bengtsson and S. ˚Aberg, Phys. Lett. B 172(1986)\n277.\n[29] G. Hansen and A. S. Jensen, Nucl. Phys. A 406(1983)\n236.\n[30] M. Guttormsen, private comm. (2009).\n[31] D. M. Brink, PhD-thesis (Oxford University, 1955).\n[32] P. Axel, Phys. Rev. 126(1962) 671.\n[33] V. O. Nesterenko, W. Kleinig, V. V. Gudkov and J.\nKvasil, Phys. Rev. C 53(1996) 1632.\n[34] H. Sakamoto and T. Kishimoto, Nucl. Phys. A 501\n(1989) 205.\n[35] S.˚Aberg, Phys. Lett. B 157(1985) 9.\n[36] D. J. Rowe, Nuclear Collective Motion - Models and The-\nory(Butler & Tanner Ltd, Frome and London, 1970).[37] A. V. Ignatyuk, Statistical Properties of excited atomic\nnuclei(Transl. from Russian) IAEA, INDC(CCP)-\n233(L), Vienna, IAEA 1985.\n[38] S. Goriely, S. Hilaire and A. J. Koning, Phys. Rev. C 78\n(2008) 064307.\n[39] B. Lauritzen, T. Døssing and R. A. Broglia, Nucl. Phys.\nA 457(1986) 61.\n[40] S.˚Aberg, Phys. Rev. Lett. 64(1990) 3119.\n[41] V. Zelevinsky, Ann. Rev. Nucl. Part. Sc. 46(1996) 237.\n[42] S. M. Grimes, Phys. Rev. C 38(1988) 2362.\n[43] N. Cerf, Nucl. Phys. A 554(1993) 85.\n[44] H. P. Loens et. al., Phys. Lett. B 666(2008) 395.\n[45] D. Mocelj et. al., Phys. Rev. C 75(2007) 045805.\n[46] RIPL2, Reference Input Parameter Li-\nbrary, IAEA-Tecdoc, 2005, available at\nhttp://www-nds.iaea.or.at/ripl2..\n[47] A. J. Koning, S. Hilaire and S. Goriely, Nucl. Phys. A\n810(2008) 13.\n[48] A. C. Larsen et. al., Phys. Rev. C 73(2006) 064301.\n[49] R. Chankova et. al., Phys. Rev. C 73(2006) 034311.\n[50] S. Siem et. al., Phys. Rev. C 65(2002) 044318.\n[51] M. Guttormsen et. al., Phys. Rev. C 68(2003) 064306.\n[52] E. Melby et. al., Phys. Rev. C 63(2001) 044309.\n[53] A. Schiller et. al., Phys. Rev. C 63(2001) 021306(R)." }, { "title": "0902.1850v1.Accurate_calculation_of_the_local_density_of_optical_states_in_inverse_opal_photonic_crystals.pdf", "content": "arXiv:0902.1850v1 [physics.optics] 11 Feb 2009Accurate calculation of the local density of optical states in\ninverse-opal photonic crystals\nIvan S. Nikolaev1, Willem L. Vos1,2, and A. Femius Koenderink1,∗\n1Center for Nanophotonics, FOM Institute for Atomic and Mole cular Physics (AMOLF), Kruislaan 407, NL-1098SJ\nAmsterdam, The Netherlands\n2Complex Photonic Systems, MESA+Institute for Nanotechnology, University of Twente, The Ne therlands\nCorresponding author: f.koenderink@amolf.nl\nSubmitted to OSA Dec 23, 2008\nWe have investigated the local density of optical states (LD OS) in titania and silicon inverse opals –\nthree-dimensional photonic crystals that have been realiz ed experimentally. We used the H-field plane-wave\nexpansion method to calculate the density of states and the p rojected local optical density of states, which\nare directly relevant for spontaneous emission dynamics an d strong coupling. We present the first quantitative\nanalysis of the frequency resolution and of the accuracy of t he calculated local density of states. We have\ncalculated the projected LDOS for many different emitter pos itions in inverse opals in order to supply a\ntheoretical interpretation for recent emission experimen ts and as reference results for future experiments and\ntheory by other workers. The results show that the LDOS in inv erse opals strongly depends on the crystal\nlattice parameter as well as on the position and orientation of emitting dipoles.\nOCIS codes: 160.5298,260.2510,270.5580,290.4210\n1. Introduction\nPhotonic crystals are metamaterials with periodic varia-\ntions of the dielectric function on length scales compara-\nble to the wavelength of light. These dielectric compos-\nites are of a keen interest for scientists and engineers\nbecause they offer exciting ways to manipulate pho-\ntons [1,2]. Of particular interest are three-dimensional\n(3D) photonic crystals possessing a photonic bandgap,\ni.e.a frequency range where no photon modes exist at\nall.Photonicbandgapmaterialspossessagreatpotential\nfor drastically changing the rate of spontaneous emission\nand for achieving localization of light [3–6]. Control over\nspontaneousemissionis importantformanyapplications\nsuch as miniature lasers [7], light-emitting diodes [8] and\nsolar cells [9]. According to Fermi’s golden rule, the rate\nof spontaneous emission from quantum emitters such as\natoms, molecules or quantum dots is proportional to the\n‘local radiative density of states’ (LDOS) [10,11], which\ncounts the number of electromagneticstates at given fre-\nquency, location and orientation of the dipolar emitters.\nIn addition, nonclassical effects can occur that are be-\nyond Fermi’s Golden Rule [2,11]. These strong coupling\nphenomena, such as fractional decay, rely on coherent\ncoupling of the quantum emitter to a sharp feature in\na highly structured LDOS. Whether these effects are in\nfact observable in real photonic crystals depends on how\nrapid the LDOS changes within a small frequency win-\ndow [12]. Therefore local density of states calculations\nfor experimentally realized photonic crystals are essen-\ntial to assess current spontaneous experiments and fu-\nturestrongcouplingexperimentsalike,providedthatthe\ncalculationsareaccurateandhaveacontrolledfrequency\nresolution.LDOS effects on spontaneous emission in photonic\ncrystals have been experimentally demonstrated in a\nvariety of systems. Since only three-dimensional (3D)\ncrystals promise full control over all optical modes with\nwhich elementary emitters interact, many groups have\npursued 3D photonic crystals. Fabrication of such pe-\nriodic structures with high photonic strengths is, how-\never, a great challenge [2,13]. Inverse opals are among\nthe most stronglyphotonic 3Dcrystalsthat can be fabri-\ncated relatively easily using self-assembly methods. Such\ncrystals consist of fcclattices of close-packed air spheres\nin a backbone material with a high dielectric constant\n[14–18]. In these inverse opals, continuous-wave exper-\niments on light sources with a low quantum efficiency\nrevealed inhibited radiative emission rates [19,20]. En-\nhanced and inhibited time-resolved emission rates have\nbeen observed from highly-efficient emitters in 3D in-\nverse opals [21]. Simultaneously, several groups have re-\nalized that emission enhancement and partial inhibition\ncan also be obtained in 2D slab structures [22–26].\nFollowing the first calculations by Suzuki and Yu [27],\nseveral papers have reported on calculations of the lo-\ncal density of optical states in photonic crystals using\nboth time domain [28–30] and frequency domain meth-\nods[31–34].Unfortunately,analysisofexperimentaldata\nfor emitters in 3D crystalsin terms ofthese LDOS calcu-\nlations is compounded by several problems in literature:\n1. Most prior calculations were performed for model\nsystems that don’t correspond to structures used in\nexperiments, and for emitter positions that are not\nprobed in experiments.\n2. The accuracy of the reported LDOS has never been\n1discussed, hampering comparison to experiments.\n3. The frequency resolution of the reported LDOS has\nremained unspecified. Therefore sharp features of\nrelevance for non-classical emission can not be as-\nsessed [11,12].\n4. ManypreviouslyreportedLDOScalculationsareer-\nroneous for symmetry reasons, as pointed out by\nWanget al.[33].\nIn this paper we aim to overcome all these problems.\nWe benchmarkthe accuracyandfrequencyresolutionfor\nour LDOS calculations to allow quantitative comparison\nwith experiments and with quantum optics strong cou-\npling requirements. Since prior LDOS calculations are\nscarce (and partly erroneous, item 1 above) we present\nsets of LDOS’s for experimentally relevant structures,\nand for spatial positions where sources can be practi-\ncally placed. Specifically, we model the spatial distribu-\ntion of the dielectric function ǫ(r) in such a way that it\nclosely resembles ǫ(r) in titania (TiO 2) [14,15] and sili-\ncon (Si) inverse-opal photonic crystals [17,18]. For these\ntwo structures we calculated the LDOS at various posi-\ntions in the crystal unit cell and for specific orientations\nof the transition dipoles. The results on the TiO 2inverse\nopals are relevant for interpreting recent emission exper-\niments [21,35].Toaidotherworkerstointerprettheir ex-\nperiments and to benchmark their codes, we make data\nsets that we report in figuresthroughout this manuscript\navailable as online material.\nThe paper is arranged as follows: in Section 2, we\npresent a detailed description of the method by which\nwe have calculated the photonic band structures and the\nLDOS.We discussthe accuracyandfrequencyresolution\nof our calculations. In Section 3 we compare our compu-\ntations with the known DOS in vacuum and with earlier\nresults on the DOS and LDOS [33] in 3D periodic struc-\ntures.Section4describestheLDOSininverseopalsfrom\nTiO2, and in Section 5 we present results of the LDOS\nin Si inverse opal photonic band gap crystals.\n2. Calculation of local density of states\nA. Introduction\nThe local radiative density of optical states is defined as\nN(r,ω,ed) =1\n(2π)3/summationdisplay\nn/integraldisplay\nBZdkδ(ω−ωn,k)|ed·En,k(r)|2,\n(1)\nwhere integration over k-vector is performed over the\nfirst Brillouin zone, nis the band index and edis the\norientation of the emitting dipole. The total density of\nstates (DOS) is the unit-cell and dipole-orientation av-\nerage of the LDOS defined as N(ω) =/summationtext\nn/integraltext\nBZdkδ(ω−\nωn,k). The important quantities that determine the\nLDOS are the eigenfrequencies ωn,kand electric field\neigenmodes En,k(r) for each k-vector. Calculation of\nthese parameters will be discussed below in Section B.The expression for the LDOS contains the term |ed·\nEn,k(r)|that depends on the dipole orientation ed. It is\nimportant to realize that in photonic crystals the vec-\ntor fields En,k(r) are not invariant under the lattice\npoint-group operations α, as first reported in Ref. [33].\nExplicitly, this means that the projection of the field\nof a mode at wave vector kon the dipole orientation\ned(i.e.|ed·En,k(r)|)is not identical to the projection\n|ed·En,α[k](r)|ofthe symmetry related modes with wave\nvectorsα[k] on the same ed. As a consequence one can\nnot calculate the LDOS for a specific dipole orientation\nby restricting the integral in Eq. (1) over the irreducible\npart (1/48th) of the Brillouin zone, since symmetry re-\nlated wave vectors do not give identical contributions.\nUnfortunately, in many previous reports on the LDOS,\nthis reduced symmetry for vector modes as compared to\nscalar quantities was overlooked, resulting in erroneous\nresults [33]. In general, the only symmetry that can be\ninvoked to avoid using the full Brillouin zone for LDOS\ncalculations is inversion symmetry, which correspondsto\ntime reversalsymmetry.Consequently,correctresultsre-\nquire that exactly half of the Brillouin zone is considered\nfor LDOS calculations, rather than the irreducible part\nof the Brillouin zone that was used in most previous lit-\nerature. We have explicitly verified that our implemen-\ntation (using the kz≥0 half of the Brillouin zone) re-\nsults in the same LDOS on symmetry related positions,\nprovided that one also takes the concomitant symmetry\nrelateddipole orientationintoaccount.Furthermore,our\ncalculations confirm the claim by Wang et. al. that this\nrequired symmetry is only recovered upon integration\nover half the Brillouin zone, rather than over just the\nirreducible part as considered in, e.g., Ref. [31,32].\nB. Plane-wave expansion\nWe use the H-field inverted plane wave expansion\nmethod [31,36,37] to solve for the electromagnetic field\nmodes in photonic crystals. For nonmagnetic materials,\nit is most convenient to solve the wave equation for the\nH(r) field [38]\n∇×/bracketleftbig\nǫ(r)−1∇×H(r)/bracketrightbig\n=ω2\nc2H(r).(2)\nbecause the operator ∇ ×ǫ(r)−1∇×is Hermitian, and\nconsequentlyhas real eigenvalues ω2/c2[1,31,36,37].Be-\ncause of the periodicity of the dielectric function ǫ(r) in\nphotonic crystals, the field modes Hk(r) of the eigen-\nvalue problem Eq. (2) satisfy the Bloch theorem [39]:\nHk(r) =eik·ruk(r). (3)\nTheseBlochmodes arefully described bythe wavevector\nkand the periodic function uk(r), which has the period-\nicity of the crystal lattice so that uk(r) =uk(r+R). To\nsolve the wave equation, the inverse dielectric function\nand the Bloch modes are expanded in a Fourier series\nover the reciprocal-lattice vectors G:\nǫ(r)−1=η(r) =/summationdisplay\nGηGeiG·rand (4)\n2Hk(r) =/summationdisplay\nGuk\nGei(k+G)·r, (5)\nwhereηGanduk\nGare the 3D Fourier expansion coeffi-\ncients of respectively η(r) anduk(r). Substituting these\nexpressions into the H-field wave equation in Eq. (2), we\nobtain a linear set of eigenvalue equations:\n−/summationdisplay\nG′ηG−G′(k+G)×[(k+G′)×un,k\nG′] =ω2\nn(k)\nc2un,k\nG.\n(6)\nThis infinite equation set with the known parameters G\nandηG−G′determines all allowed frequencies ωn(k) for\neach value of the wave vector k, subject to the transver-\nsality requirement ∇·Hk(r) = 0. Due to the periodicity\nofuk(r), we can restrict kto the first Brillouin zone. For\neach wave vector k, there is a countably infinite num-\nber of modes with discretely spaced frequencies. All the\nmodes are labeled with the band number nin order of\nincreasing frequency and are described as a family of\ncontinuous functions ωn(k) ofk.\nTo compute the eigenfrequencies ωn(k) and the ex-\npansion coefficients of the eigenmodes un,k\nG, the infinite\nequation set is truncated. By restricting the number of\nreciprocal-latticevectors Gto a finite set GwithNGele-\nments, Eq. (6) is limited to a 3 NGdimensional equation\nset. In ourimplementation wechoosethe truncatedset G\nto correspond to the set of all reciprocal lattice vectors\nwithin a sphere centered around the origin of k-space.\nThe transversality of the H-field gives an additional con-\ndition on the eigenmodes: ( k+G)·un,k\nG= 0, which elim-\ninates one vector component of un,k\nG. Following Ref. [31],\nforeachk+Goneneedstofindtwoorthogonalunitvec-\ntorse1,2\nk+Gthat form an orthogonaltriad with k+G. By\nexpressing the eigenmode expansion coefficients in the\nplane normal to k+Gasun,k\nG=un,k\nG,1e1\nk+G+un,k\nG,2e2\nk+G,\nwe remove one third of the unknowns. Then, Eq. (6) be-\ncomes\n/summationdisplay\nG′∈GηG−G′|k+G||k+G′|·/bracketleftBigg\n/parenleftbigge2\nk+G·e2\nk+G′−e2\nk+G·e1\nk+G′\n−e1\nk+G·e2\nk+G′e1\nk+G·e1\nk+G′/parenrightbigg/parenleftBigg\nun,k\nG′,1\nun,k\nG′,2/parenrightBigg/bracketrightBigg\n=ω2\nn(k)\nc2/parenleftBigg\nun,k\nG,1\nun,k\nG,2/parenrightBigg\n,∀G∈ G.\n(7)\nTo find the matrix of Fourier coefficients ηG−G′, we used\nthe method of Refs. [31,36]. The coefficients are com-\nputed by first Fourier-transforming the dielectric func-\ntionǫ(r), andthen truncatingandinvertingthe resulting\nmatrix. As first noted by Ho, Chan and Soukoulis [36],\nusing the inverse of ǫG−G′rather than ηG−G′dramati-\ncally improves the (poor) convergence of the plane-wave\nmethod that is associated with the discontinuous nature\nof the dielectric function [37]. A rigorous explanation\nfor this improvement was put forward by Li [40], whostudied the presence of Gibbs oscillations in the trun-\ncated Fourier expansion of products of functions with\ncomplementary jump discontinuities. Using the H-field\ninverted matrix plane wave method, the frequencies ob-\ntained with N G= 725 (for fccstructures) deviate by\nless than 0.5 % from the converged band structures [31].\nSolving Eqs. (7) gives the frequencies ωn(k) and the\nFourier expansion coefficients for the H-field eigenmodes\nHn,k(r) needed to calculate the LDOS in the photonic\ncrystal. The required E-fields En,k(r) are obtained using\nthe Maxwell equation ∂D/∂t=∇×H:\nEn,k(r) =1\nωn(k)ǫ0/summationdisplay\nG,G′∈GηG′−G|k+G|·/bracketleftbigg\n/parenleftBig\nun,k\nG,1e2\nk+G−un,k\nG,2e1\nk+G/parenrightBig\nei(k+G′)·r/bracketrightbigg\n.\n(8)\nFrom the orthonormality of the eigenvectors of Eq. (7)\nit follows that the Bloch functions Hn,k(r) andEn,k(r)\ndefined above satisfy the orthonormality relations:\n/integraldisplay\nBZHn,k(r)·H∗\nn′,k′(r)dr=δ(k−k′)δn,n′,(9)\n/integraldisplay\nBZǫ(r)En,k(r)·E∗\nn′,k′(r)dr=δ(k−k′)δn,n′.(10)\nIt should be noted in the definition of En,kthat the mul-\ntiplication by 1 /ǫ(r) to calculate EfromDis not in the\ndenominator in front of the Fourier expansion. Rather it\nappears as a matrix multiplying the D-field in Fourier\nspace, i.e., within the sum over reciprocal lattice vec-\ntors. This ordering ensures that complementary jumps\ninDand 1/ǫ(r) cancel, even for the truncated Fourier\nseries, as can be easily checked by plotting calculated\nmode profiles for TM modes in 2D crystals.\nC. Frequency resolution and accuracy of LDOS\nWe are not aware of any previous report that bench-\nmarks the accuracy of the calculated local radiative den-\nsity of states or that specifies the frequency resolution.\nMotivated by the requirements for accurate results for\ncomparison with experiments and for judging the utility\nof crystals for non-classical emission dynamics, we con-\nsider the accuracy and resolution of our approximation\nto the LDOS integral in Eq. (1). The main approxima-\ntion is to replace the integration over wave vector kby\nan appropriately weighted summation over a discrete set\nof wave vectors on a discretization grid [41]. Either in-\nterpolationschemes [31] orsimple histogramming(‘root-\nsampling’) methods are used to compute the LDOS. The\nk-point grid density sets the number of k-points in the\nBrillouin zone, Nk. The accuracy of the resulting LDOS\napproximation is set by the the density of grid points\nthat is used to discretize the wave vector integral. Due\nto the transparent relation between the accuracy of the\nDOS,thefrequencyresolutionandthe k-vectorsampling\nresolution, we will focus on the simple histogramming\n3Fig. 1. (Color online) DOS per volume in units 4 /a2cfor\nvacuum modeled as an empty fcccrystal. The calculated\nDOS shown by red histogram bars is compared to the\nanalytically derived ω2behavior (curve). In vacuum the\nDOS per volume equals the dipole-averaged LDOS.\nFig. 2. Averageabsolute deviation ofthe calculated DOS\nfrom the exact total DOS N(ω) for an ‘empty crystal.’\nThe average runs over the frequency range 0 < ω <\n2πc/aand the deviation is in units of the DOS N(ω)\nper volume at ωa/2πc= 1, i.e., in units 4 /a2c. In accor-\ndance with Eq. (11), the error is inversely proportional\nto the ratio ∆ ω/∆kof the histogram bin width ∆ ωto\nthe integration grid spacing ∆ k. Symbols correspond to\nintegration using the number of k-points Nk= 280, 770,\n1300, 2480, 2992 and 3570 ( /square,/squaresolid,◦,•,♦,/diamondsolid) respectively,\nwith various ∆ ω.\napproach. For the LDOS computations one chooses a\ncertain frequency bin width ∆ ωto build an LDOS his-\ntogram. For a desired frequency resolution ∆ ωand a\ndesired accuracy for the LDOS content N(r,ω,ed)∆ωin\neach frequency bin, one needs to choose an appropriate\nk-vector spacing ∆ kthat depends on the steepness of\nthe sampled dispersion relation ω(k). Indeed, the useful\nfrequency resolution ∆ ωof a histogram of the DOS and\nLDOS is limited by the resolution ∆ kof the grid in kspace to be\n∆ω≈∆k|∇kω|, (11)\nas detailed in Ref. [42] for the electronic DOS. This cri-\nterion relates the separation between adjacent k-grid\npoints to their approximate frequency spacing via the\ngroup velocity. If histogram bins are chosen too narrow\ncompared to the expected frequency spacing between\ncontributions to the discretized LDOS integral, unphysi-\ncal spikes appear in the approximation, especially in the\nlimit of small ω, where the group velocity |∇kω|is usu-\nally largest. Apart from full gaps in the LDOS, photonic\ncrystals also promise sharp lines at which the LDOS is\nenhanced, which are important for non-classical emis-\nsion dynamics. Hence it is especially important to dis-\ntinguish sharp spikes that are due to histogram binning\nnoise from true features. Unfortunately, many reports in\nliterature feature sharp spikes that are evidently binning\nnoise (wave vector undersampling errors), as they occur\nin the long wavelength limit, below any stop gap.\nTo improve the resolution without adding time-\nconsuming diagonalizations, several interpolation\nschemes have been suggested [42]. Within the his-\ntogramming approach, an interpolation scheme essen-\ntially improves binning statistics by adding histogram\ncontributions from intermediate grid points on the\nassumption that quantities vary linearly between grid\npoints. While interpolation decouples the binning noise\nfromthe frequencybin width ∆ ω, resultingin arbitrarily\nsmooth LDOS approximations, it has the disadvantage\nof obscuring the intrinsic relation between frequency\nresolution and wave vector resolution in Eq. (11).\nAgoodbenchmarkforthebinning noiseofthe k-space\nintegration,independent ofthe convergenceofthe plane-\nwave method, is to calculate the LDOS or DOS of an\n‘empty’ crystal, with uniform dielectric constant equal\nto unity. Such an ‘empty’ crystal represents a limit of\nzero photonic strength and the maximum possible group\nvelocity|∇kω|.InFigure1weshowtheDOSinanempty\nfcccrystal. As expected for a crystal with zero dielectric\ncontrast,thecalculatedDOSisindependentofdipolepo-\nsition andorientation.In agreementwith the well-known\nDOS in vacuum the calculated DOS increases paraboli-\ncally with frequency. Fluctuations around the parabola\nare due to binning noise associated with the finite k-\nspace discretization.\nWe have calculated the relative root-mean-square er-\nror in the calculated density of states (DOS) for an\nfcc‘empty’ crystal averaged over all histogram bins\nin the frequency range 0 < ωa/2πc <1 for several\ncombinations of histogram binwidth and k-grid resolu-\ntion, as specified in the caption of Fig. 2. As predicted\nby Eq. (11), the deviation of the approximation from\nthe analytic result is inversely proportional to the ratio\n∆ω/∆k. In most cases, one wants to calculate the LDOS\nwithagivenfrequencyresolution∆ ω,i.e.N(r,ω,ed)∆ω,\nto within a predetermined accuracy. For instance, calcu-\nlating the vacuum DOS for frequencies 0 < ωa/2πc <1\n4with a desired absolute accuracy better than 0 .01·4/a2c\n(1% of the vacuum DOS at ωa/2πc= 1) and a desired\nfrequencyresolutionof∆ ω= 0.01(2πc/a),requiresusing\n∆k∼∆ω/0.3c. The number of k-points corresponding\nto this wave vector sampling equals 2480 k-points in the\nirreducible wedge of the fccBrillouin zone, 59520 in the\nrequiredhalfBrillouinzone,orequivalently Nk= 119040\nin the full Brillouin zone. Photonic crystals with nonzero\nindex contrast cause a pronounced frequency structure\nof the LDOS. Due to the flattening of bands compared\nto the dispersion bands of vacuum-only crystals, the k-\nspace integration itself is at least as accurate, assuming\nthattheeigenfrequenciesandthefield-modepatternsare\nknown with infinite accuracy. In practice, one needs to\nadjust the number of plane waves to obtain all eigen-\nfrequencies to within the desired frequency resolution\n∆ω. In our computations for fcccrystals, we represented\nthe k-space of a half of the first Brillouin zone by an\nequidistant grid consisting of Nk/2 = 145708 k-points.\nThe frequency resolution of our LDOS histograms is\n∆ω= 0.01·2πc/a, and the spatial resolution of the\nLDOSis approximately a/40judgingfromthe maximum\n|G−G′| ≈120/ainvolved in the expansion with NG=\n725 plane waves in Eq. (4).\nD. Computation time required for LDOS\nCalculating the LDOS in 3D periodic structures is a\ntime-consuming task. The chosen degree of k-space dis-\ncretization ( Nk/2 = 145708 k-points) and the number\nof plane-waves ( NG= 725) are the result of a trade-off\nbetween the desired accuracy and tolerated duration of\nthe calculations. Essentially the computation consists of\nsolving for the lowest neigenvalues and eigenvectors of\na real symmetric 2 NG×2NGmatrix for each of the Nk\nk-points independently. To accomplish this calculation,\nwe use the standard Matlab ”eigs” implementation [43]\nof an implicitly restarted Arnoldi method (ARPACK)\nthat takes approximately 2.2 seconds to find the low-\nest 20 eigenvalues and eigenvectors of a single Hermitian\n1500x1500 (i.e., NG≈750 plane waves) matrix on a\n3 GHz Intel Pentium 4 processor, or on a 2.4 GHz In-\ntel Core Duo processor. For 145708 k-points the result-\ning computation time is hence on the order of 90 hours\n(4 days). Since the Matlab ARPACK routine is already\nhighlyoptimized,wedonotexpect thatthe computation\ntime per k-point can be significantly reduced for LDOS\ncalculations based on the standard H-field inverted ma-\ntrix plane wave method. In terms of the number of plane\nwavesNG, the computation time scales as N3\nG. As in\nRef. [44], the algorithm can be accelerated by realizing\nthat the iterative eigensolver does not require the full\nmatrixHmultiplying the vector uin Eq. 7, but rather\na function that quickly computes the image Huof any\ntrial vector u. Since calculating Hurepeatedly involves\na slow matrix-vector multiplication with ηG−G′, the al-\ngorithm is only accelerated for NG>1000.\nFig. 3. (Color online)DOS per volume in an fcccrys-\ntal consisting of spheres with ǫ= 7.35 in a medium\nwithǫ= 1.77 with a filling fraction of the spheres of\n25 vol %. The solid dotted curve represents calculations\nfrom Ref. [31]. Our result is plotted as a histogram(red).\n3. Comparison with previous results\nTo test the computations, we compare our results with\nearlier reports. We have calculated the DOS and LDOS\nin anfcccrystal consisting of dielectric spheres with ǫ1\n= 7.35 (TiO 2) in a medium with ǫ2= 1.77 (water) –\nthe same structure as was analyzed in Refs [31,33]. The\nspheresoccupy25vol%ofthe crystal.Figure3showsthe\ntotal DOS in this photonic crystal calculated by us and\nby Busch and John [31] . We reproduce the earlier cal-\nculations of the total DOS: both results shown in Fig. 3\nare in excellent agreement, with deviations less than 2%\nthroughout the frequency range 0 < ωa/2πc <1. In Fig-\nure 4 (Media 1)we demonstrate the LDOS in the same\nphotonic crystal at a specific location: at a point equidis-\ntant from two nearest-neighbor spheres. In this calcu-\nlation, we used the same number of reciprocal-lattice\nvectors N G= 965 as in the only available benchmark\npaper by Wang et al. [33] that does not contain sym-\nmetry errors. We find that our calculations (empty cir-\ncles) are in good agreement up to a/λ= 0.85 with the\nLDOS reported previously (solid curve): the deviations\nare smaller than 1%. Deviations at higher frequencies\nare either due to a difference in accuracy of k-space\nintegration (frequency binning resolution and k-space\nsampling density), or to a difference in accuracy of the\nplane wave methods (eigenmodes and eigenfrequencies).\nUnfortunately, neither the accuracy nor the frequency\nresolution is specified in Ref. [33]. The k-space sampling\ndensity in our work is approximately twice the density\nspecified in Ref. [33]. Based on the excellent agreement\nof our total DOS calculations with those of Busch and\nJohn[31],andonthefactthatWangetal.usedak-space\ndensity and plane wave number comparable to that of\nBusch and John, it is unlikely that inaccuracy of the k-\nspaceintegrationineithercalculationisthesourceofdis-\ncrepancy. We therefore surmise that deviations are due\nto a difference in the method of evaluation of En,k(r).\nConversion of DtoEby multiplication with 1 /ǫ(r) in\n5Fig. 4.(Coloronline) Dipole-averagedLDOS in the same\nphotonic crystal as in Fig. 3 at a position (1\n4,1\n4,0). Red\nhistogram: our calculations (Media 1). Solid curve: re-\nsults from Ref. [33]. This relative LDOS is the ratio of\nthe LDOS to that in vacuum at a/λ= 0.495.\nreal space as proposed in [33] can cause incorrect mode\namplitudes due to Gibbs oscillations, as opposed to the\nFourier space conversion using Eq. (8). In general our\ncalculations confirm the result by Wang et. al. [33] that\ntheLDOSisonlycorrectlycalculatedbyintegrationover\nhalf the Brillouin zone, rather than over the irreducible\npart as was incorrectly used in earlier reports.\n4. LDOS in TiO 2inverse opals\nIn recent time-resolved experiments, enhanced and in-\nhibited emission rates were demonstrated for quantum\ndots embedded inside strongly photonic TiO 2inverse\nopals [19,21,35]. In the framework of these experiments,\nit is highly relevant to calculate the LDOS inside such\ninverse-opal photonic crystals, especially for the source\npositions occupied in experiments. We model the posi-\ntion dependence of the dielectric function ǫ(r) as shown\nin Figure 5. This model assumes an infinite fcclattice\nof air spheres with radius r= 0.25√\n2a(ais the cubic\nlattice parameter). The spheres are covered by overlap-\nping dielectric shells ( ǫ= 6.5) with outer radius 1.09 r.\nNeighboringairspheresareconnectedbycylindricalwin-\ndows of radius 0.4 r. The resulting volume fraction of\nTiO2is equal to about 10.7%. These structural param-\neters are inferred from detailed characterization of the\ninverse opals using electron microscopy and small an-\ngle X-ray scattering [14,15]. Moreover, the stop gaps in\nthe photonic band structure (Fig. 6 (Media 2 and Me-\ndia 3)) calculated using this model agree well with re-\nflectivity measurements in the ranges of both the first-\norder (a/λ= 0.7) and second-order Bragg diffraction\n(a/λ= 1.2) [45].\nHenceforth, we will consider the relative LDOS, which\nis the ratio of the LDOS in a photonic crystal to that in\na homogeneous medium with the same volume-averaged\ndielectric function. This scaling is motivated by exper-\nimental practice, in which emission rate modifications\nare judged by normalizing measured rates to the emis-\nsion rate of the same emitter in crystals with much\nsmaller lattice constant a[19,21,35]. In these reference\nFig.5.Renderingofthedielectricfunctioninone fccunit\ncell that models the TiO 2inverse-opal structure in sec-\ntion 4: an fcclattice of air spheres of radius r= 0.25√\n2a\nwithabeing the cubic lattice parameter. The spheres\nare covered by shells with ǫ= 6.5 and outer radius\n1.09r. Neighboring air spheres are connected by win-\ndows of radius 0.4 r. The letters ( a–d) indicate four dif-\nferent positions at the TiO 2-air:a= (1,0,0)/(2√\n2),b=\n(1,1,2)/(4√\n3),c= (1,1,1)/(2√\n6) andd= (0.33,0.13,0)\n(points shown are symmetry-equivalents). The dash-\ndotted line shows the main diagonal of the cubic unit\ncell.\nFig. 6.Left: Photonic band structure (Media 2)for the\nTiO2inverse opal shown in Fig. 5. The grey rectangles\nindicate stopgaps in the ΓL direction and one stopgap\nin the ΓX direction for the inverse opal. The stopgaps\nresult in the decreased DOS ( right: circles, Media 3) at\ncorrespondingfrequencies compared to the DOS in a ho-\nmogeneous medium with nav= 1.27 (right: solid line).\n6Fig. 7. Relative LDOS in the inverse opal shown in Fig 5\nat three different positions: (a, Media 4) r= (0,0,0) [the\ncenterofan airsphere, solidcurve], r=1\n4(1,1,1)[among\nthree air spheres, dash-dotted curve] and r= (1\n2,0,0)\n[midwaybetweentwospheresalong[1,0,0]direction,dot-\nted curve]; (b, Media 5) r=1\n4(1,1,0) [in the window\nbetween two spheres] projected on [1,1,0], [-1,1,0] and\n[0,0,1] directions shown by solid, dash-dotted and dot-\nted curves, respectively.\nsystems, emission frequencies correspond to the effective\nmedium limit a/λ≪0.5 quantified by an average index\nnav=√ǫav= 1.27 for the TiO 2crystals [19]. In units\nof 4/a2c, the LDOS in a homogeneous medium is equal\ntonav(a/λ)2/3. In Figure 7a (Media 4) we plot the re-\nsulting LDOS at three positions in the unit cell: at r=\n(0,0,0),1\n4(1,1,1) and (1\n2,0,0). Due to the high symmetry\nof these points, the LDOS does not depend on the dipole\norientation, as we verified explicitly. A first main obser-\nvation from Fig. 7a (Media 4) is that the LDOS differs\nconsiderablybetween these threepositions at all reduced\nfrequencies. This observation illustrates the well-known\nstrong dependence of the LDOS on position within the\nunit cell of photonic crystals [10,27,31].\nA secondmain observationin Fig. 7a (Media 4) is that\ntheLDOSstronglyvarieswithreducedfrequency,reveal-\ning troughs and peaks caused by the pseudogap near\na/λ= 0.7, which is related to 1st-order stop gaps such\nas the L-gap. The effects of 2nd-order stopgaps appear\nbeyonda/λ >1.15 [45]. In the middle of the air region\nat position r= (0,0,0) there is a sharp, factor-of-three\nenhancement at a/λ≈1.35 within a narrow frequency\nrange. This feature could be probed by resonant atoms\ninfiltrated in the crystals [46]. At position r= (1\n2,0,0) in\nan interstitial, the mode density has a broad trough near\na/λ= 1.25 that will lead to strongly inhibited emission.\nA third main observation is that at spatial positions\nwith low symmetry, the LDOS clearly depends on the\norientation of the transition dipole moment. Figure 7b\n(Media 5) shows the frequency dependent LDOS for\nthree perpendicular dipole orientations at a position in\nthe center of a window that connects two neighboring\nair spheres (see Figure 5). The LDOS differs for all three\norientations, and is thus anisotropic. At low frequency\n(a/λ= 0.3), the emission rate is highest for a dipole\npointing in the (1 ,1,0) direction, with increasing fre-\nquency the highest rate shifts to the (0 ,0,1) and then to\nthe (−1,1,0) orientation. We emphasize that for emit-\nterswith fixedorslowlyvaryingdipole orientations,such\nFig. 8. Relative LDOS at four key frequencies, a/λ=\n0.535, 0.725, 0.865 and 1.295, in the inverse opal as a\nfunctionofposition rfrom(0,0,0)inthe[1,1,1]direction.\nThe hatched boxes indicate the position of the dielectric\nTiO2shell. The LDOS is projected on two dipole ori-\nentations: (a) [1,1,1] perpendicular to the dielectric-air\ninterface, and (b) [-1,1,0] parallel to the interface. The\nLDOS projected on the [-1,-1,2] and [-1,1,0] directions is\nequal. For r/a∈[√\n3\n2,√\n3] the LDOS is mirror-symmetric\nto that in the region [0 ,√\n3\n2].\nas dye molecules or quantum dots on solid interfaces, the\nemissionrateisdeterminedbytheopticalmodesthatare\nprojected on the dipole orientation. Therefore, knowl-\nedge of the projected LDOS is important for controlling\nspontaneous-emissionratesaswellasforinterpretingthe\ndata from experiments on emitters in photonic metama-\nterials.\nA fourth main observation from Fig. 7a and b (Media\n4 and 5) is that at low frequencies a/λ <0.5, the rela-\ntive LDOS hardly varies with frequency, which means\nthat the mode density is proportional to ω2, as in homo-\ngeneous media. Interestingly, there is a clear dependence\nofthe LDOS onboth the position andthe dipole orienta-\ntion even at these low frequencies, i.e., long wavelengths\nrelative to the crystal periodicity. The reason for these\neffects is that the photonic Bloch modes exhibit local\nvariations of the electric field related to local variations\nof the dielectric function in order to satisfy the conti-\nnuity equations at dielectric boundaries for the parallel\nEor perpendicular Dfield, respectively [34,47]. Conse-\nquently, the LDOS strongly varies on length scales much\nless than the wavelength, i.e., even in electrostatic or ef-\nfective medium limit. While such behavior may appear\nsurprising, its origin in electrostatic depolarization ef-\nfects has been discussed before [47,48].\nTo gain more insight in the spatial dependence of the\nLDOS in the inverse opals, we have performed calcula-\ntions for dipoles positioned along a characteristic axis in\nthe unit cell. Figure 8 shows the LDOS at four key fre-\nquencies for dipoles placed on the body diagonal of the\ncubic unit cell, that is, from (0 ,0,0) in the [1 ,1,1] direc-\ntion. The LDOS has clear maxima and minima along the\ndiagonal, varying by more than 10 ×. Figure 8(a) shows\nthat for a dipole oriented in the [1 ,1,1] direction (per-\npendicular to the dielectric-air interface), the LDOS is\nstrongly ( >5x) suppressed near the dielectric TiO 2shell\n(r/a≈0.353) at all frequencies. In contrast, no strong\n7/s48/s46/s48/s48/s46/s53/s49/s46/s48/s49/s46/s53/s50/s46/s48/s50/s46/s53/s51/s46/s48\n/s101\n/s100/s32/s61/s32/s91/s49/s44/s49/s44/s50/s93/s101\n/s100/s32/s61/s32/s91/s45/s49/s44/s45/s49/s44/s49/s93\n/s101\n/s100/s32/s61/s32/s91/s45/s49/s44/s49/s44/s48/s93/s80/s111/s115/s105/s116/s105/s111/s110/s32 /s98 /s32/s82/s101/s108/s46/s32/s76/s68/s79/s83/s32\n/s32/s32\n/s48/s46/s52 /s48/s46/s54 /s48/s46/s56 /s49/s46/s48 /s49/s46/s50 /s49/s46/s52 /s49/s46/s54/s48/s46/s48/s48/s46/s53/s49/s46/s48/s49/s46/s53/s50/s46/s48/s50/s46/s53\n/s101\n/s100/s32/s61/s32/s91/s45/s49/s44/s49/s44/s48/s93\n/s101\n/s100/s32/s61/s32/s91/s49/s44/s49/s44/s49/s93/s82/s101/s108/s46/s32/s76/s68/s79/s83\n/s97/s47\n/s32/s32\n/s80/s111/s115/s105/s116/s105/s111/s110/s32 /s99/s48/s46/s48/s48/s46/s53/s49/s46/s48/s49/s46/s53/s50/s46/s48/s50/s46/s53/s51/s46/s48\n/s101\n/s100/s32/s61/s32/s91/s48/s44/s49/s44/s48/s93\n/s101\n/s100/s32/s61/s32/s91/s49/s44/s48/s44/s48/s93/s82/s101/s108/s46/s32/s76/s68/s79/s83\n/s32/s32\n/s80/s111/s115/s105/s116/s105/s111/s110/s32 /s97\n/s48/s46/s52 /s48/s46/s54 /s48/s46/s56 /s49/s46/s48 /s49/s46/s50 /s49/s46/s52 /s49/s46/s54/s48/s46/s48/s48/s46/s53/s49/s46/s48/s49/s46/s53/s50/s46/s48/s50/s46/s53/s51/s46/s48/s51/s46/s53\n/s101\n/s100/s32/s61/s32/s91/s48/s46/s51/s51/s44/s48/s46/s49/s51/s44/s48/s93/s101\n/s100/s32/s61/s32/s91/s45/s48/s46/s49/s51/s44/s48/s44/s48/s46/s51/s51/s93/s101\n/s100/s32/s61/s32/s91/s48/s44/s48/s44/s49/s93\n/s97/s47/s80/s111/s115/s105/s116/s105/s111/s110/s32 /s100/s82/s101/s108/s46/s32/s76/s68/s79/s83/s32\n/s32/s32\n/s40/s100/s41/s40/s99/s41/s40/s98/s41 /s40/s97/s41\nFig.9.RelativeLDOSintheinverseopalatfourdifferent\npositions on the TiO 2-air interface shown in Fig. 5. At\neach position the LDOS is projected on three mutually\northogonal dipole orientations ed. (a, Media 6) point a\nfored= [1,0,0] and [0,1,0]. LDOS at ed= [0,0,1] and\n[0,1,0] is the same. (b, Media 7) point bfored= [1,1,2],\n[-1,1,0] and [-1,-1,1]. (c, Media 8) point cfored= [1,1,1]\nand [-1,1,0]. LDOS at ed= [-1,-1,2] is equal to that at\ned= [-1,1,0]. (d, Media 9) point dfored= [0.33,0.13,0],\n[-0.13,0,0.33] and [0,0,1].\nsuppression or enhancement occurs for dipole orienta-\ntions parallel to the dielectric interface, see Fig. 8(b).\nThe strongly differing LDOS for dipoles near the dielec-\ntric rationalize the broad distributions of emission rates\nthat were recently observed for quantum dots in inverse\nopals [35], see below. Enhancements and inhibitions also\noccur in the air regions: the LDOS is enhanced at r/a≈\n0.28 and 0.57 by up to 2.5 to 3 times, respectively, see\nFigure 8(a). At point r/a= 1/√\n3≈0.57 that lies in the\n(111) lattice plane in the air region, the LDOS is inhib-\nited at all frequencies for the dipole orientations [-1,1,0]\nand [-1,-1,2] that are perpendicular to the (111) plane\n(see Fig. 8(b)). Finally, for dipoles parallel to the body\ndiagonal (Figure 8(a)) the mode density shows pseudo-\noscillatory behavior on length scales much less than the\nwavelength, e.g., a period of 0 .2aat frequency 0 .535a/λ\n(period corresponds to 1 /10thof a wavelength). This ob-\nservation confirms that photonic crystals are bona fide\nmetamaterials where optical properties strongly vary on\nlength scales much less than the wavelength.\nIn recent time-resolved experiments [21,35], emission\nfrom quantum dot light sources distributed at the in-\nternal TiO 2-air interfaces of the inverse opals was in-\nvestigated. To analyze the experimental data, we have\ncalculated the LDOS at several symmetry-inequivalent\npositions on the TiO 2shells: at points a,b,candd\n(see Fig. 5) for three mutually orthogonal orientations\nof the emitting dipole, where the first orientation is cho-\nsen along the vector pointing from (0,0,0) toward the\ncorresponding point. Figures 9(a) through 9(d) (Media\n6 through 9) show that at all these positions the LDOSstrongly varies with reduced frequency and position, as\nexpected, and also with orientation of the dipole. In\nbroad terms, for a dipole parallel to the interface the\nLDOS is near 1 at low frequency, increases to a peak\nenhancement of 2 .5×ata/λ= 1.22 before strongly\nvarying at high frequencies. For reference, the frequency\na/λ= 1.22 is in the range where a band gap is expected\nfor more strongly interacting crystals. The plots also re-\nveal that for dipoles perpendicular to the TiO 2-air in-\nterface, the LDOS is strongly inhibited (more than 10 ×)\nover broad frequency ranges. For instance, Figure 9(a)\nshows that the emission rate for a dipole perpendicular\ntothe interfaceis16-foldinhibited toalevelof0.06,with\na maximum of 0.22 (5-fold inhibition) at a/λ= 1.22.\nIt is striking that the strong inhibition occurs over a\nbroad frequency range from 0.3 to 1.6, i.e., more than\ntwo octaves in frequency. While the inhibition is not\ncomplete, the bandwidth is much larger than the maxi-\nmum bandwidth of 12% for a full 3D bandgap in inverse\nopals [49] and far exceeds the bandwidth of the 2D gap\nforTEmodesinmembranephotoniccrystals,thatallows\na seven-fold inhibition [26,30] over a 30% bandwidth.\nIn the frequency range up to a/λ= 1.1, the depen-\ndence of the LDOS on frequency is quite similar at all\nthe positions and dipole orientations. This remarkable\nresult agrees with the experimental observation from\nRef. [35]: the complex decay curves of light sources in\ntheinverseopalsaredescribedbyoneand thesamefunc-\ntional shape (log-normal) of the decay-rate distribution\nfor all reduced frequencies studied.\nFor the interpretation of the time-resolved experi-\nments on ensembles of emitters in the inverse-opal pho-\ntonic crystals, the results presented above mean that\n1. thedecayrateofanindividualemitterisdetermined\nbyits frequency,positionandalsobytheorientation\nof its transition dipole in the photonic crystal;\n2. the measured spontaneous-emission decay depends\non how the emitters are distributed in the crystal;\n3. even in the low-frequency regime, ensemble\nmeasurements will reveal non-exponentional decay\ncurves;\n4. the similar shape of the reduced-frequency de-\npendence of the LDOS allows modeling of the\nnon-exponentional decay curves with a single type\nof decay-rate distributions. In other words, one\ndistribution function can be successfully used to\nmodel the multi-exponentional decay curves meas-\nured from crystals with different lattice parameters.\n5. LDOS in silicon inverse opals\nA complete inhibtion of spontaneous emission may only\nbe achieved in photonic crystals with a 3D photonic\nband gap. Therefore,there has been much effort to fabri-\ncate inverse opals from silicon rather than titania, since\nthe higher index of silicon allows for a photonic band\n8Fig. 10. Left: Photonic band structure (Media 10) for\nan inverse opal from silicon ( ǫ= 11.9). Right: The to-\ntal DOS in the Si inverse opal (Media 11). The DOS is\nstrongly depleted for frequencies near ΓL and ΓX stop-\ngaps (grey rectangles). A photonic band gap (grey bar)\noccurs between bands 8 and 9, as also reflected in the\nvanishing DOS.\nFig. 11. Relative LDOS in a Si inverse opal at: (a, Me-\ndia 12)r= (0,0,0) [the center of an air sphere, solid\ncurve],r=1\n4(1,1,1) [among three air spheres, dash-\ndotted curve] and r= (1\n2,0,0) [midway between two\nspheres along [1,0,0] direction, dotted curve]; (b, Media\n13)r=1\n4(1,1,0)[inthe windowbetween twoairspheres]\nprojected on [1,1,0], [-1,1,0] and [0,0,1] directions shown\nby solid, dash-dotted and dotted curves, respectively.\ngap [31,37]. For the calculation of the LDOS in such Si\ninverse opals, the dielectric function ǫ(r) was modeled\nsimilarly to that in the inverse opals shown in Fig. 5.\nFrom SEM observations we inferred the following struc-\ntural parameters [17,18]: the outer radius of the over-\nlapping dielectric shells with ǫ= 11.9 is about 1.15 r\n(wherer= 0.25√\n2ais the air-sphere radius, ais the\nlattice parameter). The cylindrical windows connecting\nneighboring air spheres have a radius of 0.2 r. The larger\nouter radius and the smaller window size compared to\nthe TiO 2inverse opals are commensurate with a higher\nvolume fraction of about 23% Si [18].\nThe band structure and total DOS for this system are\nshown in Fig. 10 (Media 10 and Media 11). The DOS is\nstrongly depleted in a pseudo-gap between the 2ndand\n3rdbands [36], at frequencies near 0.55. Near frequency\n0.85, both the band structures and the DOS reveal a\n3D photonic band gap of relative width ∆ ω/ω≈3 %between the 8thand 9thbands [31]. Compared to the\nTiO2structure, both the lowest-order L-gap and the 8th\nand 9thbands are shifted to lower reduced frequencies.\nThisshift isdue to the highereffective refractiveindex of\nthe Si inverse opals ( nav= 1.88), on account of a higher\nindex of the backbone and a higher filling fraction.\nFigure 11(a) (Media 12) presents the relative LDOS\nat three high-symmetry positions in the unit cell r=\n(0,0,0),1\n4(1,1,1) and (1\n2,0,0). The LDOS varies much\nmorestronglywith frequencythan in TiO 2inverseopals,\nas a result of the larger dielectric contrast that leads to\nstrongly modified dispersion relations and Bloch mode\nprofiles.While themaximaupto3.2arenotmuchhigher\nthan in TiO 2inverse opals, the minima in the mode den-\nsity are reduced and the slopes are steeper. As expected,\nthe LDOS is completely inhibited at all positions in the\nfrequency range of the band gap. Previously, it has been\nsuggested that the LDOS could be inhibited at salient\npositions in the unit cell for frequencies outside a com-\nplete gap. While Figure 11(a) (Media 12) reveals that\nthe mode density is strongly reduced above the band\ngap, e.g., at r= (1\n2,0,0), it is not truly inhibited. In\nfact, in the course of our study, we have not encoun-\ntered any ”sweet spots” where the LDOS is completely\ninhibited at frequencies outside a 3D photonic band gap.\nFigure 11(b) (Media 13) shows the LDOS at a lower\nsymmetry position r=1\n4(1,1,0), in the window between\ntwo nearest-neighbor air spheres. The mode density has\nbeen calculated for three perpendicular dipole orienta-\ntions. Figure 11(b) (Media 13) shows that the LDOS is\nhighly anisotropic since it differs for all three orienta-\ntions: at low frequencies ( a/λ <0.5), the mode density\nis highest for the ( −1,1,0) orientation, intermediate for\nthe (0,0,1) orientation, and lowest for the (1 ,1,0) ori-\nentation. With increasing frequency, the highest LDOS\nalso changes to other orientations, with the (0 ,0,1) ori-\nentation having the highest LDOS above the pseudo-gap\nand even a narrow peak at frequency 0.95. Therefore, if\nit is possible to orient a quantum emitter with its dipole\nparallel to (0 ,0,1), this frequency range is conducive to\nstrong emission enhancement and perhaps even QED ef-\nfects beyond weak-coupling. Interestingly, these frequen-\ncies are slightly higher than the upper edge of the band\ngap where strong coupling effects have recently been dis-\ncussed [12].\n6. Conclusions\nWe have performed intensive calculations of the local\ndensity ofstates in TiO 2and Si inverseopalswith exper-\nimentally relevant structural parameters. Since conflict-\ning and incorrect reports have appeared on the LDOS in\nphotoniccrystals,wehavesetouttovalidateourmethod\nof choice, i.e., the H-field plane-wave expansion method.\nThisvalidationreliedoncomparisontoliteratureresults,\non the explicit verification of required symmetries that\nprevious reports failed to satisfy, and on quantitative\nconsiderations of resolution and accuracy. Results for\n9each structure are made available for other workers in\nthe field, both as benchmarks and for comparison with\nexperimental data. With the help of these computations\nwe have obtained quantitative insight in the LDOS rel-\nevant for time-resolved ensemble fluorescence measure-\nments on photonic crystals, such as obtained in recent\nexperimental work [35]. The results of our numerical cal-\nculations reveal a surprisingly strong dependence of the\nLDOS on the orientation of the emitting dipoles.\nAcknowledgments\nWe thank Dries van Oosten for careful reading of the\nmanuscript. This work is part of the research program\nof the “Stichting voor Fundamenteel Onderzoek der Ma-\nterie (FOM),” which is financially supported by the\n“NederlandseOrganisatievoorWetenschappelijkOnder-\nzoek (NWO)”. AFK and WLV were supported by VENI\nand VICI fellowships funded by NWO. WLV also ac-\nknowledges funding by STW/NanoNed.\nReferences\n1. J. D. Joannopoulos, S. G. Johnson, J.N. Winn and R. D.\nMeade, Photonic Crystals: Molding the Flow of Light,\n2nd edition , (Princeton University Press, 2008).\n2. C. M. Soukoulis, editor, Photonic Crystals and Light Lo-\ncalization in the 21stCentury, (Kluwer,Dordrecht2001).\n3. V. P. Bykov, “Spontaneous emission in a periodic struc-\nture,” Sov. Phys. JETP 35, 269-273 (1972).\n4. “Spontaneous emission from a medium with a band\nspectrum,” Sov. J. Quant. Electron. 4, 861-871 (1975).\n5. E. Yablonovitch, “Inhibited spontaneous emission in\nsolid-state physics and electronics,” Phys. Rev. Lett. 58,\n2059-2062 (1987).\n6. S. John, “Strong localization of photons in certain dis-\nordered dielectric superlattices,” Phys. Rev. Lett. 58,\n2486-2489 (1987).\n7. O. Painter, R. K. Lee, A. Scherer, A. Yariv, J. D.\nO’Brien, P. D. Dapkus and I. Kim,“Two-dimensional\nphotonic band-gap defect mode laser,” Science 284,\n18191821 (1999).\n8. H.-G. Park, S.-H. Kim, S.-H. Kwon, Y.-G. Ju, J.-K.\nYang, J.-H. Baek, S.-B. Kim, Y.-H. Lee, “Electrically\ndriven single-cell photonic crystal laser,” Science, 305,\n1444-1447 (2004).\n9. M. Gr¨ atzel, “Photoelectrochemical cells,” Nature 414,\n338-344 (2001).\n10. R. Sprik, B.A. van Tiggelen, and A. Lagendijk, “Opti-\ncal emission in periodic dielectrics,” Europhys. Lett. 35,\n265-270 (1996).\n11. N. Vats, S. John and K. Busch, “Theory of fluorescence\nin photonic crystals,” Phys. Rev. A 65, 043808 (2002).\n12. P. Kristensen, A. F. Koenderink, P. Lodahl, B. Trom-\nborg, and J. Mørk, “Fractional decay of quantum dots in\nreal photonic crystals,” Opt. Lett. 33, 1557-1559 (2008).\n13. K. Busch, S. L¨ olkes, R. B. Wehrspohn and H. F¨ oll, ed-\nitors,Photonic Crystals. Advances in Design, Fabrica-\ntion, and Characterization ,(Wiley-VCH Verlag GmbH\n& Co., Weinheim 2004).14. J.E.G.J. WijnhovenandW. L.Vos,“Preparation ofpho-\ntonic crystals made of air spheres in titania,” Science\n281, 802-804, (1998).\n15. J.E.G.J. Wijnhoven, L. Bechger and W. L. Vos, “Fab-\nrication and characterization of large macroporous pho-\ntonic crystals in titania,” Chem. Mater. 13, 4486-4499\n(2001).\n16. A. A. Zakhidov, R. H. Baughman, Z. Iqbal, C. Cui, I.\nKhayrullin,S.O.Dantas,J. Marti, andV.G.Ralchenko,\n“Carbon structures with three-dimensional periodicity\nat optical wavelengths,” Science 282, 897-901 (1998).\n17. A. Blanco, E. Chomski, S. Grabtchak, M. Ibisate, S.\nJohn, S. W. Leonard, C. L´ opez, F. Meseguer, H. M´ ıguez,\nJ. P. Mondia, G. A. Ozin, O. Toader, and H. M. van\nDriel, “Large-scale synthesis of a silicon photonic crys-\ntal with a complete three-dimensional bandgap near 1.5\nmicrometres,” Nature 405, 437-440 (2000).\n18. Y. A. Vlasov, X. Z. Bo, J. C. Sturm, and D. J. Norris,\n“On-chip natural assembly of silicon photonic bandgap\ncrystals,” Nature 414, 289-293 (2001).\n19. A.F. Koenderink, L. Bechger, H. P. Schriemer, A. La-\ngendijk and W. L. Vos, “Broadband fivefold reduction\nof vacuum fluctuations probed by dyes in photonic crys-\ntals,” Phys. Rev. Lett. 88, 143903 (2002).\n20. S. Ogawa, M. Imada, S. Yoshimoto, M. Okato and S.\nNoda, “Control of light emission by 3D photonic crys-\ntals,” Science 305, 227-229 (2004).\n21. P. Lodahl, A. F. van Driel, I. S. Nikolaev, A. Irman, K.\nOvergaag, D.Vanmaekelbergh,andW.L.Vos,“Control-\nling the dynamics of spontaneous emission from quan-\ntum dots by photonic crystals,” Nature 430, 654 (2004).\n22. A.Badolato, K.Hennessy,M. Atat¨ ure, J. Dreiser, E.Hu,\nP. M. Petroff, A. Imamo˘ glu, “Deterministic coupling of\nsingle quantum dots to single nanocavity modes,” Sci-\nence308, 1158-1161 (2005).\n23. A. Kress, F. Hofbauer, N. Reinelt, M. Kaniber, H. J.\nKrenner, R. Meyer, G. B¨ ohm, and J. J. Finley, “Manip-\nulation of the spontaneous emission dynamics of quan-\ntum dots in two-dimensional photonic crystals,” Phys.\nRev. B71, 241304 (2005).\n24. M. Fujita, S. Takahashi, Y. Tanaka, T. Asano, S. Noda,\n“Simultaneous inhibition and redistribution of sponta-\nneous light emission in photonic crystals,” Science 308,\n1296-1298 (2005).\n25. D. Englund, D. Fattal, E. Waks, G. Solomon, B. Zhang,\nT. Nakaoka, Y. Arakawa, Y. Yamamoto, J. Vu˘ ckovi´ c,\n“Controlling the spontaneous emission rate of single\nquantum dots in a two-Dimensional photonic crystal,”\nPhys. Rev. Lett. 95, 013904 (2005).\n26. B. Julsgaard, J. Johansen, S. Stobbe, T. Stolberg-Rohr,\nT. S¨ unner, M. Kamp, A. Forchel and P. Lodahl, “Decay\ndynamics of quantum dots influenced by the local den-\nsity of optical states in two-dimensional photonic crystal\nmembranes,” Appl. Phys. Lett. 93, 094102 (2008).\n27. T. Suzuki and P. K. L. Yu, “Emission power of an elec-\ntric dipole in the photonic band structure of the fcc lat-\ntice,” J. Opt. Soc. Am. B 12, 570-582 (1995).\n28. R. K.Lee, Y. Xu,A.Yariv,“Modified spontaneous emis-\nsion from a two-dimensional photonic bandgap crystal\nslab,” J. Opt. Soc. Am. B 17, 1438–1442 (2000).\n29. C. Hermann and O. Hess, “Modified spontaneous emis-\nsion rate in an inverted-opal structure with complete\n10photonic bandgap,” J. Opt. Soc. Am. B 19, 3013–3018\n(2002).\n30. A. F. Koenderink, M. Kafesaki, C.M. Soukoulis and\nV. Sandoghdar, “Spontaneous emission control in two-\ndimensional photonic crystal membranes,” J. Opt. Soc.\nAm. B23, 1196-1206 (2006).\n31. K. Busch and S. John, “Photonic band gap formation in\ncertain self-organizing systems”, Phys. Rev. E 58, 3896-\n3908 (1998).\n32. Z.-Y.Li,L.-L.Lin,andZ.-Q.Zhang,“Spontaneousemis-\nsion from photonic crystals: full vectorial calculations, ”\nPhys. Rev. Lett. 84, 4341 (2000).\n33. R. Wang, X.-H. Wang, B.-Y. Gu and G.-Z. Yang, “Local\ndensity of states in three-dimensional photonic crystals:\ncalculation and enhancement effects,” Phys. Rev. B 67,\n155114 (2003).\n34. D. P. Fussell, R. C. McPhedran, and C. M. de\nSterke, “Three-dimensional Green’s tensor, local den-\nsity of states, and spontaneous emission in finite two-\ndimensional photonic crystals composed of cylinders,”,\nPhys. Rev. E 70, 066608 (2004).\n35. I. S. Nikolaev, P. Lodahl, A. F. van Driel, A. F. Koen-\nderink, and W. L. Vos, “Strongly nonexponential time-\nresolvedfluorescenceofquantum-dotensembles inthree-\ndimensional photonic crystals,” Phys. Rev.B 75, 115302\n(2007).\n36. K.M.Ho,C.T.ChanandC.M.Soukoulis,“Existenceof\na photonic gap in periodic dielectric structures”, Phys.\nRev. Lett. 65, 3152-3155 (1990).\n37. H.S. S¨ oz¨ uer, J.W. Haus, and R. Inguva, “Photonic\nbands: convergence problems with the plane-wave\nmethod,” Phys. Rev. B 45, 13962 (1992).\n38. J.D. Jackson, Classical Electrodynamics , (John Wiley &\nSons, New York 1975).\n39. N. W. Ashcroft and N. D. Mermin, Solid State Physics ,\n(Holt, Rinehart and Winston, New York, 1976).\n40. L. Li, “Use of Fourier series in the analysis of discon-\ntinuous periodic structures,” J. Opt. Soc. Am. A 13,\n1870-1876 (1996).\n41. H. J. Monkhorst and J. D. Pack, “Special points for\nBrillouin-zone integration,” Phys. Rev. B 13, 5188-5192\n(1976).\n42. G. Gilat, “Analysis of methods for calculating spec-\ntral properties in solids,” J. Comput. Phys. 10, 432-465\n(1972).\n43. R. B. Lehoucq, D. C. Sorensen, and C. Yang, ARPACK\nUsers Guide: Solution of Large-Scale Eigenvalue Prob-\nlems with Implicitly Restarted Arnoldi Methods , (SIAM\nPublications, Philadelphia 1998).\n44. S. G. Johnson and J. D. Joannopoulos, “Block-iterative\nfrequency-domain methods for Maxwell’s equations in a\nplanewave basis,” Opt. Express 8, 173-190 (2001).\n45. W. L. Vos and H. M. van Driel, “Higher order Bragg\ndiffraction by strongly photonic fcc crystals: onset of a\nphotonic bandgap,” Phys. Lett. A 272, 101 (2000).\n46. P.J. Harding, Photonic crystals modified by opti-\ncally resonant systems , (Ph.D. thesis, University of\nTwente (2008). ISBN 978-90-365-2683-8, available from\nwww.photonicbandgaps.com\n47. L. Rogobete, H. Schniepp, V. Sandoghdar and C.\nHenkel, “Spontaneous emission in nanoscopic dielectric\nparticles,” Opt. Lett. 28, 1736-1738 (2003).48. H. Miyazaki and K. Ohtaka, Phys. Rev. B 58, 6920\n(1998).\n49. A.F. Koenderink, Emission and transport of light in\nphotonic crystals , (Ph.D. thesis, University of Am-\nsterdam (2003)). ISBN 90-9016903-2, available from\nwww.photonicbandgaps.com\n11" }, { "title": "0902.1902v2.Local_density_of_states_of_electron_crystal_phases_in_graphene_in_the_quantum_Hall_regime.pdf", "content": "arXiv:0902.1902v2 [cond-mat.mes-hall] 14 Feb 2009\n/C4/D3 \r/CP/D0 /CS/CT/D2/D7/CX/D8 /DD /D3/CU /D7/D8/CP/D8/CT/D7 /D3/CU /CT/D0/CT\r/D8/D6/D3/D2/B9\r/D6/DD/D7/D8/CP/D0 /D4/CW/CP/D7/CT/D7 /CX/D2 /CV/D6/CP/D4/CW/CT/D2/CT /CX/D2 /D8/CW/CT /D5/D9/CP/D2 /D8/D9/D1 /C0/CP/D0/D0/D6/CT/CV/CX/D1/CT/C7/BA /C8 /D3/D4/D0/CP /DA/D7/CZ/DD/DD/B8\n/BD/B8 /BE/B8 /BF /B8∗/C5/BA /C7/BA /BZ/D3 /CT/D6/CQ/CX/CV/B8\n/BE/CP/D2/CS /BV/BA /C5/D3/D6/CP/CX/D7 /CB/D1/CX/D8/CW\n/BF/BD/BY /CP\r/D9/D0/D8/DD /D3/CU /C5/CP/D8/CW/CT/D1/CP/D8/CX\r/D7/B8 /CF/CX/D0/CQ /CT/D6/CU/D3/D6 \r /CT /CA /D3 /CP/CS/B8 /CD/D2/CX/DA/CT/D6/D7/CX/D8/DD /D3/CU /BV/CP/D1/CQ/D6/CX/CS/CV/CT/B8 /BV/BU/BF /BC/CF /BT/B8 /CD/D2/CX/D8/CT /CS /C3/CX/D2/CV/CS/D3/D1/BE/C4 /CP/CQ /D3/D6 /CP/D8/D3/CX/D6 /CT /CS/CT /C8/CW/DD/D7/CX/D5/D9/CT /CS/CT/D7 /CB/D3/D0/CX/CS/CT/D7/B8 /BV/C6/CA/CB /CD/C5/CA /BK/BH/BC/BE/B8 /CD/D2/CX/DA/CT/D6/D7/CX/D8/GH /C8/CP/D6/CX/D7 /CB/D9/CS/B8 /BY/B9/BL/BD/BG/BC/BH /C7/D6/D7/CP/DD /BV/CT /CS/CT/DC/B8 /BY /D6 /CP/D2\r /CT/BF/C1/D2/D7/D8/CX/D8/D9/D8/CT /CU/D3/D6 /CC/CW/CT /D3/D6 /CT/D8/CX\r /CP/D0 /C8/CW/DD/D7/CX\r/D7/B8 /CD/D2/CX/DA/CT/D6/D7/CX/D8/DD /D3/CU /CD/D8/D6 /CT \r/CW/D8/B8 /C4 /CT/D9/DA/CT/D2/D0/CP/CP/D2 /BG/B8 /BF/BH/BK/BG /BV/BX /CD/D8/D6 /CT \r/CW/D8/B8 /CC/CW/CT /C6/CT/D8/CW/CT/D6/D0/CP/D2/CS/D7/B4/BW/CP/D8/CT/CS/BM /C7\r/D8/D3/CQ /CT/D6 /BE/BH/B8 /BE/BC/BD/BK/B5/CF /CT \r/CP/D0\r/D9/D0/CP/D8/CT/B8 /DB/CX/D8/CW/CX/D2 /CP /D7/CT/D0/CU/B9\r/D3/D2/D7/CX/D7/D8/CT/D2 /D8 /C0/CP/D6/D8/D6/CT/CT/B9/BY /D3 \r /CZ /CP/D4/D4/D6/D3 /DC/CX/D1/CP/D8/CX/D3/D2/B8 /D8/CW/CT /D0/D3 \r/CP/D0 /CS/CT/D2/D7/CX/D8 /DD /D3/CU /D7/D8/CP/D8/CT/D7/CU/D3/D6 /CS/CX/AR/CT/D6/CT/D2 /D8 /CT/D0/CT\r/D8/D6/D3/D2 \r/D6/DD/D7/D8/CP/D0/D7 /CX/D2 /CV/D6/CP/D4/CW/CT/D2/CT /D7/D9/CQ /CY/CT\r/D8 /D8/D3 /CP /D7/D8/D6/D3/D2/CV /D1/CP/CV/D2/CT/D8/CX\r /AS/CT/D0/CS/BA /CF /CT /CX/D2 /DA /CT/D7/D8/CX/CV/CP/D8/CT /CQ /D3/D8/CW/D8/CW/CT /CF/CX/CV/D2/CT/D6 \r/D6/DD/D7/D8/CP/D0 /CP/D2/CS /CQ/D9/CQ/CQ/D0/CT \r/D6/DD/D7/D8/CP/D0/D7 /DB/CX/D8/CWMe\n/CT/D0/CT\r/D8/D6/D3/D2/D7 /D4 /CT/D6 /D0/CP/D8/D8/CX\r/CT /D7/CX/D8/CT/BA /CC/CW/CT /D8/D3/D8/CP/D0 /CS/CT/D2/D7/CX/D8 /DD /D3/CU/D7/D8/CP/D8/CT/D7 \r/D3/D2/D7/CX/D7/D8/D7 /D3/CU /D7/CT/DA /CT/D6/CP/D0 /D4/D6/D3/D2/D3/D9/D2\r/CT/CS /D4 /CT/CP/CZ/D7/B8 /D8/CW/CT /D2 /D9/D1 /CQ /CT/D6 /D3/CU /DB/CW/CX\r /CW /CX/D2 /D8/CW/CT /D2/CT/CV/CP/D8/CX/DA /CT /CT/D2/CT/D6/CV/DD /D6/CP/D2/CV/CT\r/D3/CX/D2\r/CX/CS/CT/D7 /DB/CX/D8/CW /D8/CW/CT /D2 /D9/D1 /CQ /CT/D6 /D3/CU /CT/D0/CT\r/D8/D6/D3/D2/D7 Me\n/D4 /CT/D6 /D0/CP/D8/D8/CX\r/CT /D7/CX/D8/CT/B8 /CP/D7 /CU/D3/D6 /D8/CW/CT \r/CP/D7/CT /D3/CU /CT/D0/CT\r/D8/D6/D3/D2/B9/D7/D3/D0/CX/CS /D4/CW/CP/D7/CT/D7/CX/D2 /D8/CW/CT \r/D3/D2 /DA /CT/D2 /D8/CX/D3/D2/CP/D0 /D8 /DB /D3/B9/CS/CX/D1/CT/D2/D7/CX/D3/D2/CP/D0 /CT/D0/CT\r/D8/D6/D3/D2 /CV/CP/D7/BA /BT/D2/CP/D0/DD/DE/CX/D2/CV /D8/CW/CT /D0/D3 \r/CP/D0 /CS/CT/D2/D7/CX/D8 /DD /D3/CU /D7/D8/CP/D8/CT/D7 /CP/D8 /D8/CW/CT /D4 /CT/CP/CZ/CT/D2/CT/D6/CV/CX/CT/D7/B8 /DB /CT /AS/D2/CS /D4/CP/D6/D8/CX\r/D9/D0/CP/D6 /D7\r/CP/D0/CX/D2/CV /D4/D6/D3/D4 /CT/D6/D8/CX/CT/D7 /D3/CU /D8/CW/CT /CS/CT/D2/D7/CX/D8 /DD /D4/CP/D8/D8/CT/D6/D2/D7 /CX/CU /D3/D2/CT /AS/DC/CT/D7 /D8/CW/CT /D6/CP/D8/CX/D3νN/Me/CQ /CT/D8 /DB /CT/CT/D2 /D8/CW/CT /AS/D0/D0/CX/D2/CV /CU/CP\r/D8/D3/D6νN\n/D3/CU /D8/CW/CT /D0/CP/D7/D8 /D4/CP/D6/D8/CX/CP/D0/D0/DD /AS/D0/D0/CT/CS /C4/CP/D2/CS/CP/D9 /D0/CT/DA /CT/D0 /CP/D2/CS /D8/CW/CT /D2 /D9/D1 /CQ /CT/D6 /D3/CU /CT/D0/CT\r/D8/D6/D3/D2/D7/D4 /CT/D6 /CQ/D9/CQ/CQ/D0/CT/BA /BT/D0/D8/CW/D3/D9/CV/CW /D8/CW/CT /D8/D3/D8/CP/D0 /CS/CT/D2/D7/CX/D8 /DD /D4/D6/D3/AS/D0/CT /CS/CT/D4 /CT/D2/CS/D7 /CT/DC/D4/D0/CX\r/CX/D8/D0/DD /D3/D2Me\n/B8 /D8/CW/CT /D0/D3 \r/CP/D0 /CS/CT/D2/D7/CX/D8 /DD /D3/CU /D7/D8/CP/D8/CT/D7/D3/CU /D8/CW/CT /D0/D3 /DB /CT/D7/D8 /D4 /CT/CP/CZ/D7 /D8/D9/D6/D2/D7 /D3/D9/D8 /D8/D3 /CQ /CT /CX/CS/CT/D2 /D8/CX\r/CP/D0 /D6/CT/CV/CP/D6/CS/D0/CT/D7/D7 /D8/CW/CT /D2 /D9/D1 /CQ /CT/D6 /D3/CU /CT/D0/CT\r/D8/D6/D3/D2/D7 Me\n/BA /CF/CW/CT/D6/CT/CP/D7 /D8/CW/CT/D7/CT/CT/D0/CT\r/D8/D6/D3/D2/B9/D7/D3/D0/CX/CS /D4/CW/CP/D7/CT/D7 /CP/D6/CT /D6/CT/D1/CX/D2/CX/D7\r/CT/D2 /D8 /D3/CU /D8/CW/D3/D7/CT /CT/DC/D4 /CT\r/D8/CT/CS /CX/D2 /D8/CW/CT \r/D3/D2 /DA /CT/D2 /D8/CX/D3/D2/CP/D0 /D8 /DB /D3/B9/CS/CX/D1/CT/D2/D7/CX/D3/D2/CP/D0 /CT/D0/CT\r/D8/D6/D3/D2/CV/CP/D7 /CX/D2 /BZ/CP/BT/D7 /CW/CT/D8/CT/D6/D3/D7/D8/D6/D9\r/D8/D9/D6/CT/D7 /CX/D2 /D8/CW/CT /D5/D9/CP/D2 /D8/D9/D1 /C0/CP/D0/D0 /D6/CT/CV/CX/D1/CT/B8 /D8/CW/CT /D0/D3 \r/CP/D0 /CS/CT/D2/D7/CX/D8 /DD /D3/CU /D7/D8/CP/D8/CT/D7 /CP/D2/CS /D8/CW/CT /D7\r/CP/D0/CX/D2/CV/D6/CT/D0/CP/D8/CX/D3/D2/D7 /DB /CT /CW/CX/CV/CW/D0/CX/CV/CW /D8 /CX/D2 /D8/CW/CX/D7 /D4/CP/D4 /CT/D6 /D1/CP /DD /CQ /CT/B8 /CX/D2 /CV/D6/CP/D4/CW/CT/D2/CT/B8 /CS/CX/D6/CT\r/D8/D0/DD /D1/CT/CP/D7/D9/D6/CT/CS /CQ /DD /D7/D4 /CT\r/D8/D6/D3/D7\r/D3/D4/CX\r /D1/CT/CP/D2/D7/B8/D7/D9\r /CW /CP/D7 /CT/BA/CV/BA /D7\r/CP/D2/D2/CX/D2/CV /D8/D9/D2/D2/CT/D0/CX/D2/CV /D1/CX\r/D6/D3/D7\r/D3/D4 /DD /BA/C8 /BT /BV/CB /D2 /D9/D1 /CQ /CT/D6/D7/BM /BJ/BF/BA/BG/BF/BA/B9/CU/B8 /BJ/BF/BA/BE/BC/BA/C9/D8/B8 /BJ/BF/BA/BE/BD/BA/B9/CQ/B8 /BI/BK/BA/BF/BJ/BA/B9/CS/C3/CT/DD/DB /D3/D6/CS/D7/BM /CF/CX/CV/D2/CT/D6 \r/D6/DD/D7/D8/CP/D0/BN /CT/D0/CT\r/D8/D6/D3/D2/B9/CQ/D9/CQ/CQ/D0/CT \r/D6/DD/D7/D8/CP/D0/BN /CW/CX/CV/CW /D1/CP/CV/D2/CT/D8/CX\r /AS/CT/D0/CS/BN /D0/D3 \r/CP/D0 /CS/CT/D2/D7/CX/D8 /DD /D3/CU /D7/D8/CP/D8/CT/D7/BN /D7\r/CP/D2/D2/CX/D2/CV/D8/D9/D2/D2/CT/D0/CX/D2/CV /D7/D4 /CT\r/D8/D6/D3/D7\r/D3/D4 /DD/C1/BA /C1/C6/CC/CA /C7/BW/CD/BV/CC/C1/C7/C6/BT/D7 /DB /CP/D7 /D7/CW/D3 /DB/D2 /CQ /DD /CF/CX/CV/D2/CT/D6 /CX/D2 /BD/BL/BF/BG/B8\n/BD/D8/CW/CT /CS/CT/CV/CT/D2/CT/D6/CP/D8/CT/BY /CT/D6/D1/CX /CV/CP/D7 /CX/D7 /D9/D2/D7/D8/CP/CQ/D0/CT /D8/D3 /DB /CP/D6/CS/D7 /D8/CW/CT /CU/D3/D6/D1/CP/D8/CX/D3/D2 /D3/CU /CP /D4 /CT/B9/D6/CX/D3 /CS/CX\r /D8/D6/CX/CP/D2/CV/D9/D0/CP/D6 /D0/CP/D8/D8/CX\r/CT /D3/CU /D0/D3 \r/CP/D0/CX/DE/CT/CS /CT/D0/CT\r/D8/D6/D3/D2/D7 /B4/CT/D0/CT\r/D8/D6/D3/D2\r/D6/DD/D7/D8/CP/D0/B5/B8 /D3/D2\r/CT /D8/CW/CT /BV/D3/D9/D0/D3/D1 /CQ /CT/D2/CT/D6/CV/DD /D4/D6/CT/DA /CP/CX/D0/D7 /D3 /DA /CT/D6 /D8/CW/CT /CZ/CX/B9/D2/CT/D8/CX\r /D3/D2/CT/BA /C0/D3 /DB /CT/DA /CT/D6/B8 /D8/CW/CT \r/D6/CX/D8/CX\r/CP/D0 /CT/D0/CT\r/D8/D6/D3/D2 /CS/CT/D2/D7/CX/D8 /DD /CP/D8 /DB/CW/CX\r /CW/D8/CW/CT /D8/D6/CP/D2/D7/CX/D8/CX/D3/D2 /D8/D3 /D8/CW/CT /CF/CX/CV/D2/CT/D6 \r/D6/DD/D7/D8/CP/D0 /B4/CF /BV/B5 /D3 \r\r/D9/D6/D7 /CX/D7 /D8/D3 /D3/D0/D3 /DB /CU/D3/D6 /D9/D7/D9/CP/D0 /D1/CT/D8/CP/D0/D7/BA /C6/CT/DA /CT/D6/D8/CW/CT/D0/CT/D7/D7/B8 /D8/CW/CT /D7/CX/D8/D9/CP/D8/CX/D3/D2 /CX/D7 /D1 /D9\r /CW/CX/D1/D4/D6/D3 /DA /CT/CS /CX/CU /D3/D2/CT /CP/D4/D4/D0/CX/CT/D7 /CP /D7/D8/D6/D3/D2/CV /D4 /CT/D6/D4 /CT/D2/CS/CX\r/D9/D0/CP/D6 /D1/CP/CV/D2/CT/D8/CX\r/AS/CT/D0/CS /D8/D3 /CP /D8 /DB /D3/B9/CS/CX/D1/CT/D2/D7/CX/D3/D2/CP/D0 /B4/BE/BW/B5 /CT/D0/CT\r/D8/D6/D3/D2 /CV/CP/D7 /B4/BE/BW/BX/BZ/B5/BA /C1/D2/D8/CW/CX/D7 \r/CP/D7/CT/B8 /D8/CW/CT /D7/CX/D2/CV/D0/CT/B9/D4/CP/D6/D8/CX\r/D0/CT \r/D3/D2 /D8/CX/D2 /D9/D3/D9/D7 /CT/D2/CT/D6/CV/DD /D7/D4 /CT\r/D8/D6/D9/D1/CX/D7 /D5/D9/CP/D2 /D8/CX/DE/CT/CS /CX/D2 /D8/D3 /CP /D7/CT/D5/D9/CT/D2\r/CT /D3/CU /CW /D9/CV/CT/D0/DD /CS/CT/CV/CT/D2/CT/D6/CP/D8/CT /C4/CP/D2/B9/CS/CP/D9 /D0/CT/DA /CT/D0/D7 /B4/C4/C4/B3/D7/B5/BA /C1/CU /D3/D2/CT /D6/CT/D7/D8/D6/CX\r/D8/D7 /D3/D2/CT/D7/CT/D0/CU /D8/D3 /D8/CW/CT /CT/D0/CT\r/D8/D6/D3/D2/D7/DB/CX/D8/CW/CX/D2 /D8/CW/CT /D0/CP/D7/D8 /D4/CP/D6/D8/CX/CP/D0/D0/DD/B9/AS/D0/D0/CT/CS /C4/C4/B8 /D3/D2/CT /AS/D2/CS/D7 /D8/CW/CP/D8 /D8/CW/CT/CX/D6 /CZ/CX/B9/D2/CT/D8/CX\r /CT/D2/CT/D6/CV/DD /CX/D7 /D5/D9/CT/D2\r /CW/CT/CS/B8 /CP/D2/CS /D8/CW/CT /D3/D2/D0/DD /CT/D2/CT/D6/CV/DD /D7\r/CP/D0/CT /CX/D7 /D8/CW/CT/BV/D3/D9/D0/D3/D1 /CQ /CT/D2/CT/D6/CV/DD /B8 /DB/CW/CX\r /CW /CU/CP /DA /D3/D6/D7 /D8/CW/CT /CU/D3/D6/D1/CP/D8/CX/D3/D2 /D3/CU /CP/D2 /CT/D0/CT\r/B9/D8/D6/D3/D2 \r/D6/DD/D7/D8/CP/D0 /CP/D8 /D7/D1/CP/D0/D0 /AS/D0/D0/CX/D2/CV /CU/CP\r/D8/D3/D6/D7/BA\n/BE /B8/BF/B8/BG/BT /D5/D9/CP/D2 /D8/D9/D1 /CT/D0/CT\r/D8/D6/D3/D2 \r/D6/DD/D7/D8/CP/D0 /CX/D2 /D8/CW/CT /D4/D6/CT/D7/CT/D2\r/CT /D3/CU /CP/CS/CX/D7/D3/D6/CS/CT/D6 /D4 /D3/D8/CT/D2 /D8/CX/CP/D0 /CX/D7 /CT/DC/D4 /CT\r/D8/CT/CS /D8/D3 /CQ /CT\r/D3/D1/CT \r/D3/D0/D0/CT\r/D8/CX/DA /CT/D0/DD/D4/CX/D2/D2/CT/CS /CP/D2/CS /D8/D3 /D1/CP/D2/CX/CU/CT/D7/D8 /CX/D8/D7/CT/D0/CU /CP/D7 /CP/D2 /CX/D2/D7/D9/D0/CP/D8/D3/D6/BA\n/BH/CF/CW/CX/D0/CT/CP/D8 /D7/D1/CP/D0/D0 /AS/D0/D0/CX/D2/CV /CU/CP\r/D8/D3/D6/D7 /D8/CW/CT /BE/BW /CF /BV /DB/CX/D8/CW /D8/D6/CX/CP/D2/CV/D9/D0/CP/D6 /D0/CP/D8/B9/D8/CX\r/CT /D7/DD/D1/D1/CT/D8/D6/DD\n/BF/B8/BG/CX/D7 /CT/DC/D4 /CT\r/D8/CT/CS /D8/D3 /DD/CX/CT/D0/CS /D8/CW/CT /CV/D0/D3/CQ/CP/D0 /CT/D2/B9/CT/D6/CV/DD /D1/CX/D2/CX/D1 /D9/D1/B8 /CX/D8 /DB /CP/D7 /D4/D6/CT/CS/CX\r/D8/CT/CS /D8/CW/CP/D8 /D8/CW/CT /D4/CW/CP/D7/CT /CS/CX/CP/B9/CV/D6/CP/D1 /D3/CU /D8/CW/CT /BE/BW/BX/BZ /CX/D2\r/D0/D9/CS/CT/D7 /CP/D0/D7/D3 /CT/D0/CT\r/D8/D6/D3/D2/B9/CQ/D9/CQ/CQ/D0/CT \r/D6/DD/D7/B9/D8/CP/D0/D7 /B4/CP /D4 /CT/D6/CX/D3 /CS/CX\r /D0/CP/D8/D8/CX\r/CT /DB/CX/D8/CW /D1/D3/D6/CT /D8/CW/CP/D2 /D3/D2/CT /CT/D0/CT\r/D8/D6/D3/D2 /D4 /CT/D6/D7/CX/D8/CT/B5/B8 /D7/D8/D6/CX/D4 /CT/D7/B8\n/BI/B8/BJ /B8/BK/CP/D2/CS /CT/DA /CT/D2 /D1/D3/D6/CT /CT/DC/D3/D8/CX\r /D5/D9/CP/D2 /D8/D9/D1 /C0/CP/D0/D0/D0/CX/D5/D9/CX/CS/B9\r/D6/DD/D7/D8/CP/D0 /D4/CW/CP/D7/CT/D7/BA\n/BL/CD/D2/D0/CX/CZ /CT /CT/D0/CT\r/D8/D6/D3/D2/B9\r/D6/DD/D7/D8/CP/D0 /D4/CW/CP/D7/CT/D7/B8 /D8/CW/CT/D4/D6/D3/D1/CX/D2/CT/D2 /D8 /D5/D9/CP/D2 /D8/D9/D1 /D0/CX/D5/D9/CX/CS/D7/B8 /DB/CW/CX\r /CW /CS/CX/D7/D4/D0/CP /DD /D8/CW/CT /CU/D6/CP\r/D8/CX/D3/D2/CP/D0/D5/D9/CP/D2 /D8/D9/D1 /C0/CP/D0/D0 /CT/AR/CT\r/D8 /CX/D2 /D8/CW/CT /D8 /DB /D3 /D0/D3 /DB /CT/D7/D8 /C4/C4/D7/B8\n/BD/BC/CP/D6/CT /D8/D6/CP/D2/D7/D0/CP/B9/D8/CX/D3/D2/CP/D0/D0/DD /CP/D2/CS /D6/D3/D8/CP/D8/CX/D3/D2/CP/D0/D0/DD /CX/D2 /DA /CP/D6/CX/CP/D2 /D8 /CP/D2/CS /D6/CT/D1/CP/CX/D2 \r/D3/D2/CS/D9\r/D8/CX/D2/CV\n/CT/DA /CT/D2 /CX/D2 /D8/CW/CT /D4/D6/CT/D7/CT/D2\r/CT /D3/CU /CS/CX/D7/D3/D6/CS/CT/D6/BA /CC/CW/CT/D6/CT/CU/D3/D6/CT/B8 /D8/CW/CT/D7/CT /CS/CX/AR/CT/D6/B9/CT/D2 /D8 /D5/D9/CP/D2 /D8/D9/D1 /D4/CW/CP/D7/CT/D7 /D1/CP /DD /CQ /CT /CS/CX/D7/D8/CX/D2/CV/D9/CX/D7/CW/CT/CS /CT/DC/D4 /CT/D6/CX/D1/CT/D2 /D8/CP/D0/D0/DD/DB/CX/D8/CW /D6/CT/D7/D4 /CT\r/D8 /D8/D3 /D8/CW/CT /CQ /CT/CW/CP /DA/CX/D3/D6 /CX/D2 /D8/D6/CP/D2/D7/D4 /D3/D6/D8 /D1/CT/CP/D7/D9/D6/CT/D1/CT/D2 /D8/D7/BA/BY /D3/D6 /CX/D2/D7/D8/CP/D2\r/CT/B8 /CP /D7/D9\r\r/CT/D7/D7/CX/D3/D2 /D3/CU /CX/D2/D7/D9/D0/CP/D8/CX/D2/CV /CP/D2/CS \r/D3/D2/CS/D9\r/D8/CX/D2/CV/D4/CW/CP/D7/CT/D7 /DD/CX/CT/D0/CS/D7 /CP /D6/CT/B9/CT/D2 /D8/D6/CP/D2 /D8 /CX/D2 /D8/CT/CV/CT/D6 /D5/D9/CP/D2 /D8/D9/D1 /C0/CP/D0/D0 /CT/AR/CT\r/D8/B4/C1/C9/C0/BX/B5 /CX/D2 /D8/CW/CT /AS/D6/D7/D8\n/BD/BD/B4N /BP/BD/B5 /CP/D2/CS /D7/CT\r/D3/D2/CS\n/BD/BE/B4N= 2 /B5 /CT/DC/B9\r/CX/D8/CT/CS /C4/C4/D7 /CP/D2/CS /CW/CP/D7 /CQ /CT/CT/D2 /CX/D2 /D8/CT/D6/D4/D6/CT/D8/CT/CS /CX/D2 /D8/CT/D6/D1/D7 /D3/CU /CP \r/D3/D1/D4 /CT/D8/CX/B9/D8/CX/D3/D2 /D3/CU /D7/D9\r /CW /CT/D0/CT\r/D8/D6/D3/D2/B9/D7/D3/D0/CX/CS /CP/D2/CS /D5/D9/CP/D2 /D8/D9/D1/B9/D0/CX/D5/D9/CX/CS /D4/CW/CP/D7/CT/D7/BA\n/BD/BF/BY /D9/D6/D8/CW/CT/D6 /CT/DA/CX/CS/CT/D2\r/CT /CU/D3/D6 /CT/D0/CT\r/D8/D6/D3/D2 \r/D6/DD/D7/D8/CP/D0/D7 /CX/D2 /C4/C4/D7 /D7/D8/CT/D1/D7 /CU/D6/D3/D1/D6/CP/CS/CX/D3/B9/CU/D6/CT/D5/D9/CT/D2\r/DD /D7/D4 /CT\r/D8/D6/D3/D7\r/D3/D4 /DD\n/BD/BG/CP/D2/CS /D8/D6/CP/D2/D7/D4 /D3/D6/D8 /D1/CT/CP/D7/D9/D6/CT/B9/D1/CT/D2 /D8/D7 /D9/D2/CS/CT/D6 /D1/CX\r/D6/D3 /DB /CP /DA /CT /CX/D6/D6/CP/CS/CX/CP/D8/CX/D3/D2/B8\n/BD/BH /B8/BD/BI/B8/BD/BJ/B8/BD/BK/B8/BD/BL/DB/CW/CX\r /CW /CT/DC/B9\r/CX/D8/CT/D7 /D8/CW/CT \r/D3/D0/D0/CT\r/D8/CX/DA /CT /D4/CX/D2/D2/CX/D2/CV /D1/D3 /CS/CT /D3/CU /D8/CW/CT /CT/D0/CT\r/D8/D6/D3/D2 \r/D6/DD/D7/D8/CP/D0/D7/BA/CF/CW/CX/D0/CT /CP/D0/D0 /D8/CW/CT/D7/CT /CT/DC/D4 /CT/D6/CX/D1/CT/D2 /D8/CP/D0 /D8/CT\r /CW/D2/CX/D5/D9/CT/D7 /CW/CP /DA /CT /CQ /CT/CT/D2 /DA /CT/D6/DD/D7/D9\r\r/CT/D7/D7/CU/D9/D0 /CX/D2 /CS/CX/D7\r/D3 /DA /CT/D6/CX/D2/CV /D2/CT/DB /CX/D2/D7/D9/D0/CP/D8/CX/D2/CV /D4/CW/CP/D7/CT/D7 /CP/D2/CS /CW/CP /DA /CT\r/D3/D2/AS/D6/D1/CT/CS /CP /D2 /D9/D1 /CQ /CT/D6 /D3/CU /D8/CW/CT/D3/D6/CT/D8/CX\r/CP/D0 /D4/D6/CT/CS/CX\r/D8/CX/D3/D2/D7/B8 /D8/CW/CT/DD /CP/D6/CT/CX/D2/CS/CX/D6/CT\r/D8 /CT/DA/CX/CS/CT/D2\r/CT /CU/D3/D6 /CW/CX/CV/CW/B9/AS/CT/D0/CS /CT/D0/CT\r/D8/D6/D3/D2 \r/D6/DD/D7/D8/CP/D0/D7 /CQ/CP/D7/CT/CS/D3/D2 /D8/D6/CP/D2/D7/D4 /D3/D6/D8 /D1/CT/CP/D7/D9/D6/CT/D1/CT/D2 /D8/D7 /AL /D8/CW/CT /BE/BW/BX/BZ /CX/D2 /BZ/CP/BT/D7 /CW/CT/D8/B9/CT/D6/D3/D7/D8/D6/D9\r/D8/D9/D6/CT/D7 /CX/D7 /CQ/D9/D6/CX/CT/CS /CS/CT/CT/D4 /CX/D2/D7/CX/CS/CT /D8/CW/CT /D7/D9/CQ/D7/D8/D6/CP/D8/CT/B8 /DB/CW/CX\r /CW/D6/CT/D2/CS/CT/D6/D7 /CX/D1/D4 /D3/D7/D7/CX/CQ/D0/CT /CP /CS/CX/D6/CT\r/D8 /D3/D4/D8/CX\r/CP/D0 /D3/CQ/D7/CT/D6/DA /CP/D8/CX/D3/D2 /D3/CU /CP /D4 /CT/B9/D6/CX/D3 /CS/CX\r /CT/D0/CT\r/D8/D6/D3/D2/B9\r/D6/DD/D7/D8/CP/D0 /D0/CP/D8/D8/CX\r/CT/B8 /CT/BA/CV/BA/B8 /CQ /DD /D1/CT/CP/D2/D7 /D3/CU /D7\r/CP/D2/D2/CX/D2/CV/D8/D9/D2/D2/CT/D0/CX/D2/CV /D1/CX\r/D6/D3/D7\r/D3/D4 /DD /B4/CB/CC/C5/B5/BA\n/BE/BC/C1/D2 \r/D3/D2 /D8/D6/CP/D7/D8 /D8/D3 /D8/CW/CT \r/D3/D2/B9/DA /CT/D2 /D8/CX/D3/D2/CP/D0 /BE/BW/BX/BZ/B8 /D7/D9\r /CW /D3/D4/D8/CX\r/CP/D0 /D7/D8/D9/CS/CX/CT/D7 /D1/CX/CV/CW /D8 /CQ /CT\r/D3/D1/CT /D4 /D3/D7/B9/D7/CX/CQ/D0/CT /CX/D2 /CV/D6/CP/D4/CW/CT/D2/CT/B8 /CP /D3/D2/CT/B9/CP/D8/D3/D1 /D8/CW/CX\r /CZ /D7/CW/CT/CT/D8 /D3/CU /CV/D6/CP/D4/CW/CX/D8/CT/B8/DB/CX/D8/CW /D9/D2/CX/D5/D9/CT /CT/D0/CT\r/D8/D6/D3/D2/CX\r /CP/D2/CS /D1/CT\r /CW/CP/D2/CX\r/CP/D0 /D4/D6/D3/D4 /CT/D6/D8/CX/CT/D7/BA\n/BE/BD /B8/BE/BE/C1/D2/B9/CS/CT/CT/CS/B8 /CV/D6/CP/D4/CW/CT/D2/CT /D1/CP /DD /CQ /CT /DA/CX/CT/DB /CT/CS /CP/D7 /CP /D4/CP/D6/D8/CX\r/D9/D0/CP/D6 /BE/BW/BX/BZ/B8/DB/CW/CT/D6/CT /D8/CW/CT /CT/D0/CT\r/D8/D6/D3/D2/D7 /CQ /CT/CW/CP /DA /CT /CP/D7 /CX/CU /D8/CW/CT/DD /DB /CT/D6/CT /D1/CP/D7/D7/D0/CT/D7/D7 /D4/CP/D6/B9/D8/CX\r/D0/CT/D7 /CS/CT/D7\r/D6/CX/CQ /CT/CS /CQ /DD /D8/CW/CT /D6/CT/D0/CP/D8/CX/DA/CX/D7/D8/CX\r /BE/BW /BW/CX/D6/CP\r /CT/D5/D9/CP/D8/CX/D3/D2/BA/C1/D2 /D8/CW/CX/D7 /BW/CX/D6/CP\r /CT/D5/D9/CP/D8/CX/D3/D2 /D8/CW/CT /BY /CT/D6/D1/CX /DA /CT/D0/D3 \r/CX/D8 /DD vF\n/D4/D0/CP /DD/D7 /D8/CW/CT/D6/D3/D0/CT /D3/CU /D8/CW/CT /D7/D4 /CT/CT/CS /D3/CU /D0/CX/CV/CW /D8c /B8 /CP/D0/D8/CW/D3/D9/CV/CW /CX/D8 /CX/D7 /D6/D3/D9/CV/CW/D0/DD /BF/BC/BC/D8/CX/D1/CT/D7 /D7/D1/CP/D0/D0/CT/D6 /D8/CW/CP/D2 /D8/CW/CT /D0/CP/D8/D8/CT/D6 /CX/D2 /DA /CP\r/D9/D9/D1/BA /C1/D2 /CP/CS/CS/CX/D8/CX/D3/D2/B8 /D8/CW/CT/BU/D6/CX/D0/D0/D3/D9/CX/D2 /DE/D3/D2/CT /D3/CU /CV/D6/CP/D4/CW/CT/D2/CT /CW/CP/D7 /D8 /DB /D3 /D2/D3/D2/B9/CT/D5/D9/CX/DA /CP/D0/CT/D2 /D8 \r/D3/D6/D2/CT/D6/BE/D4 /D3/CX/D2 /D8/D7 /B4\r/CP/D0/D0/CT/CS /BW/CX/D6/CP\r /D4 /D3/CX/D2 /D8/D7/B5 /DB/CW/CX\r /CW /DD/CX/CT/D0/CS /CP /D8 /DB /D3/CU/D3/D0/CS /DA /CP/D0/B9/D0/CT/DD /CS/CT/CV/CT/D2/CT/D6/CP\r/DD /CP/D2/CS /DB/CW/CX\r /CW /D1/CP /DD /CU/D3/D6/D1/CP/D0/D0/DD /CQ /CT /CS/CT/D7\r/D6/CX/CQ /CT/CS /CX/D2/D8/CT/D6/D1/D7 /D3/CU /CP/D2 /CB/CD/B4/BE/B5 /D4/D7/CT/D9/CS/D3/D7/D4/CX/D2 /CS/CT/CV/D6/CT/CT /D3/CU /CU/D6/CT/CT/CS/D3/D1/BA/C1/D2 /D7/D8/D6/D3/D2/CV /D1/CP/CV/D2/CT/D8/CX\r /AS/CT/D0/CS/D7/B8 /D8/CW/CT /CT/D2/CT/D6/CV/DD /D3/CU /BW/CX/D6/CP\r /CU/CT/D6/D1/CX/D3/D2/D7/CX/D2 /CV/D6/CP/D4/CW/CT/D2/CT /CX/D7 /D5/D9/CP/D2 /D8/CX/DE/CT/CS /CX/D2 /D8/D3 /C4/C4/B3/D7/B8 /D8/CW/CT /D7/D8/D6/D9\r/D8/D9/D6/CT /D3/CU /DB/CW/CX\r /CW/CX/D7 /CS/CX/AR/CT/D6/CT/D2 /D8 /CU/D6/D3/D1 /D8/CW/CP/D8 /D3/CU /D2/D3/D2/B9/D6/CT/D0/CP/D8/CX/DA/CX/D7/D8/CX\r /CT/D0/CT\r/D8/D6/D3/D2/D7 /CX/D2 /CP\r/D3/D2 /DA /CT/D2 /D8/CX/D3/D2/CP/D0 /BE/BW/BX/BZ/BA /BT/D4/CP/D6/D8 /CU/D6/D3/D1 /D8/CW/CT/CX/D6 /D9/D2\r/D3/D2 /DA /CT/D2 /D8/CX/D3/D2/CP/D0/B4/D7/D5/D9/CP/D6/CT /D6/D3 /D3/D8/B5 /D1/CP/CV/D2/CT/D8/CX\r/B9/AS/CT/D0/CS /CS/CT/D4 /CT/D2/CS/CT/D2\r/CT/B8 /D8/CW/CT/D6/CT /CT/DC/CX/D7/D8/D7 /CP/C4/C4 /CP/D8 /CT/DC/CP\r/D8/D0/DD /DE/CT/D6/D3 /CT/D2/CT/D6/CV/DD /B8 /CP/D2/CS /CT/CP\r /CW /C4/C4 /CX/D2 /D8/CW/CT \r/D3/D2/CS/D9\r/D8/CX/D3/D2/CQ/CP/D2/CS /CW/CP/D7 /CP \r/D3/D9/D2 /D8/CT/D6/D4/CP/D6/D8 /CX/D2 /D8/CW/CT /DA /CP/D0/CT/D2\r/CT /CQ/CP/D2/CS/BA /CC/CW/CX/D7 /D4/CP/D6/D8/CX\r/B9/D9/D0/CP/D6 /C4/C4 /D7/D8/D6/D9\r/D8/D9/D6/CT /D0/CT/CP/CS/D7 /D8/D3 /D8/CW/CT /CP/D2/D3/D1/CP/D0/D3/D9/D7 /CX/D2 /D8/CT/CV/D6/CP/D0 /D5/D9/CP/D2/B9/D8/D9/D1 /C0/CP/D0/D0 /CT/AR/CT\r/D8 /D3/CQ/D7/CT/D6/DA /CT/CS /CX/D2 /CV/D6/CP/D4/CW/CT/D2/CT/BA\n/BE/BF /B8/BE/BG/C1/D2 /CP/CS/CS/CX/D8/CX/D3/D2/B8/D8/CW/CT /CT/D2/CT/D6/CV/DD /CV/CP/D4 /CQ /CT/D8 /DB /CT/CT/D2 /D8/CW/CT /D7/D9/CQ/D7/CT/D5/D9/CT/D2 /D8 /C4/C4/B3/D7 /CX/D2 /CV/D6/CP/D4/CW/CT/D2/CT/CX/D7 /D7/D3 /D0/CP/D6/CV/CT/B8 /D8/CW/CP/D8 /CX/D8 /CX/D7 /D4 /D3/D7/D7/CX/CQ/D0/CT /D8/D3 /D3/CQ/D7/CT/D6/DA /CT /D8/CW/CT /C1/C9/C0/BX /CT/DA /CT/D2/CP/D8 /D6/D3 /D3/D1 /D8/CT/D1/D4 /CT/D6/CP/D8/D9/D6/CT/D7/BA\n/BE/BH/BT/D7 /D8/CW/CT /D1/D3/CQ/CX/D0/CX/D8 /DD /D3/CU /CV/D6/CP/D4/CW/CT/D2/CT /D7/CP/D1/D4/D0/CT/D7 /CX/D7 /CU/D9/D6/D8/CW/CT/D6 /CX/D1/B9/D4/D6/D3 /DA /CT/CS/B8 /D3/D2/CT /D1/CP /DD /CT/DC/D4 /CT\r/D8 /D8/D3 /D3/CQ/D7/CT/D6/DA /CT /D8/CW/CT /CU/D6/CP\r/D8/CX/D3/D2/CP/D0 /D5/D9/CP/D2/B9/D8/D9/D1 /C0/CP/D0/D0 /CT/AR/CT\r/D8/B8 /DB/CW/CX\r /CW /CW/CP/D7 /CQ /CT/CT/D2 /D7/D8/D9/CS/CX/CT/CS /D8/CW/CT/D3/D6/CT/D8/CX\r/CP/D0/D0/DD /CQ /DD/D7/CT/DA /CT/D6/CP/D0 /CP/D9/D8/CW/D3/D6/D7/B8\n/BE/BI /B8/BE/BJ/B8/BE/BK/B8/BE/BL/B8/BF/BC/CP/D2/CS /CP/D0/D7/D3 \r/D3/D0/D0/CT\r/D8/CX/DA /CT/D0/DD/B9/D4/CX/D2/D2/CT/CS/CX/D2/D7/D9/D0/CP/D8/CX/D2/CV /D4/CW/CP/D7/CT/D7 /D7/D9\r /CW /CP/D7 /CF/CX/CV/D2/CT/D6/B9\r/D6/DD/D7/D8/CP/D0 /CP/D2/CS /CQ/D9/CQ/CQ/D0/CT/D4/CW/CP/D7/CT/D7/B8 /CP/D7 /D4/D6/CT/CS/CX\r/D8/CT/CS /CX/D2 /CA/CT/CU/D7/BA /BF/BD /B8/BF/BE /B8/BF/BF /B8/BF/BG /B8/BF/BH /BA /C1/D2 \r/D3/D2/B9/D8/D6/CP/D7/D8 /D8/D3 /BZ/CP/BT/D7 /CW/CT/D8/CT/D6/D3/D7/D8/D6/D9\r/D8/D9/D6/CT/D7/B8 /D8/CW/CT/D7/CT /CT/D0/CT\r/D8/D6/D3/D2/CX\r /D4/CW/CP/D7/CT/D7/D3 \r\r/D9/D6 /CP/D8 /D8/CW/CT /D7/D9/D6/CU/CP\r/CT /D3/CU /D8/CW/CT /CV/D6/CP/D4/CW/CT/D2/CT /D7/CW/CT/CT/D8 /CP/D2/CS /CP/D6/CT/B8 /D8/CW /D9/D7/B8/CS/CX/D6/CT\r/D8/D0/DD /CP\r\r/CT/D7/D7/CX/CQ/D0/CT /CQ /DD /D7/D4 /CT\r/D8/D6/D3/D7\r/D3/D4/CX\r /D1/CT/CP/D2/D7/BA /C1/D2/CS/CT/CT/CS /CB/CC/C5/CW/CP/D7 /CQ /CT/CT/D2 /CP/D4/D4/D0/CX/CT/CS /D7/D9\r\r/CT/D7/D7/CU/D9/D0/D0/DD /D8/D3 /D4/D6/D3/CQ /CT /D8/CW/CT /CS/CT/D2/D7/CX/D8 /DD /CS/CX/D7/D8/D6/CX/B9/CQ/D9/D8/CX/D3/D2 /CX/D2 /CT/DC/CU/D3/D0/CX/CP/D8/CT/CS\n/BF/BI/CP/D2/CS /CT/D4/CX/D8/CP/DC/CX/CP/D0\n/BF/BJ/CV/D6/CP/D4/CW/CT/D2/CT/B8 /CP/D7 /DB /CT/D0/D0/CP/D7 /CX/D2 /CV/D6/CP/D4/CW/CT/D2/CT /D3/D2 /CP /CV/D6/CP/D4/CW/CX/D8/CT /D7/D9/CQ/D7/D8/D6/CP/D8/CT /CX/D2 /CP /D7/D8/D6/D3/D2/CV /D1/CP/CV/B9/D2/CT/D8/CX\r /AS/CT/D0/CS/BA\n/BF/BK/CC/CW/CX/D7 /CT/DC\r/CX/D8/CX/D2/CV /D4/D6/D3/D7/D4 /CT\r/D8 /D1/D3/D8/CX/DA /CP/D8/CT/CS /D9/D7 /D8/D3 \r/CP/D0/B9\r/D9/D0/CP/D8/CT /D8/CW/CT/D3/D6/CT/D8/CX\r/CP/D0/D0/DD /D4/CW /DD/D7/CX\r/CP/D0 /D5/D9/CP/D2 /D8/CX/D8/CX/CT/D7 /D3/CU /CP /BE/BW /CT/D0/CT\r/D8/D6/D3/D2\r/D6/DD/D7/D8/CP/D0 /DB/CW/CX\r /CW /D1/CX/CV/CW /D8 /CQ /CT /D1/CT/CP/D7/D9/D6/CT/CS /CX/D2 /CP/D2 /CB/CC/C5 /CT/DC/D4 /CT/D6/CX/D1/CT/D2 /D8/BM/D8/CW/CT /B4/CX/D2 /D8/CT/CV/D6/CP/D8/CT/CS/B5 /CS/CT/D2/D7/CX/D8 /DD /D3/CU /D7/D8/CP/D8/CT/D7 /B4/BW/C7/CB/B5 /CP/D2/CS /D8/CW/CT /D0/D3 \r/CP/D0/CS/CT/D2/D7/CX/D8 /DD /D3/CU /D7/D8/CP/D8/CT/D7 /B4/C4/BW/C7/CB/B5/BA /CF /CT /D7/CW/D3/D9/D0/CS /D2/D3/D8/CT /D8/CW/CP/D8 /D8/CW/CT /D5/D9/CP/D2/B9/D8/D9/D1 /C0/CP/D0/D0 /D6/CT/CV/CX/D1/CT /CX/D7 /D8/CW/CT /D3/D2/D0/DD \r/CP/D7/CT /CU/D3/D6 /DB/CW/CX\r /CW /D3/D2/CT /D1/CP /DD /CT/DC/B9/D4 /CT\r/D8 /D8/CW/CT /CU/D3/D6/D1/CP/D8/CX/D3/D2 /D3/CU /CT/D0/CT\r/D8/D6/D3/D2/B9\r/D6/DD/D7/D8/CP/D0 /D4/CW/CP/D7/CT/D7 /CX/D2 /CV/D6/CP/D4/CW/CT/D2/CT/BA/C1/D2/CS/CT/CT/CS/B8 /CX/D8 /CX/D7 /D4/D6/CT/CS/CX\r/D8/CT/CS /D8/CW/CP/D8 /D8/CW/CT /BE/BW /CF/CX/CV/D2/CT/D6 \r/D6/DD/D7/D8/CP/D0/D0/CX/DE/CP/B9/D8/CX/D3/D2 /CX/D7 \r/D3/D1/D4/D0/CT/D8/CT/D0/DD /CP/CQ/D7/CT/D2 /D8 /CX/D2 /CV/D6/CP/D4/CW/CT/D2/CT /CU/D3/D6 /CP/D2 /DD /CT/D0/CT\r/D8/D6/D3/D2/CS/CT/D2/D7/CX/D8 /DD /CX/D2 /D8/CW/CT /CP/CQ/D7/CT/D2\r/CT /D3/CU /CP /D1/CP/CV/D2/CT/D8/CX\r /AS/CT/D0/CS/B8\n/BF/BL/CS/D9/CT /D8/D3 /D8/CW/CT/D7\r/CP/D0/CT /CX/D2 /DA /CP/D6/CX/CP/D2\r/CT /D3/CU /D8/CW/CT /CS/CX/D1/CT/D2/D7/CX/D3/D2/D0/CT/D7/D7 /CX/D2 /D8/CT/D6/CP\r/D8/CX/D3/D2 /D4/CP/D6/CP/D1/CT/B9/D8/CT/D6rs=e2//planckover2pi1ǫvF≃2/ǫ /CU/D3/D6 /CP /BE/BW /D7/DD/D7/D8/CT/D1 /DB/CX/D8/CW /CP /D0/CX/D2/CT/CP/D6 /CS/CX/D7/B9/D4 /CT/D6/D7/CX/D3/D2 /D6/CT/D0/CP/D8/CX/D3/D2/BA /C0/CT/D6/CT/B8ǫ /CX/D7 /D8/CW/CT /CS/CX/CT/D0/CT\r/D8/D6/CX\r \r/D3/D2/D7/D8/CP/D2 /D8 /DB/CW/CX\r /CW/CS/CT/D4 /CT/D2/CS/D7 /D3/D2 /D8/CW/CT /CT/D2 /DA/CX/D6/D3/D2/D1/CT/D2 /D8 /DB/CW/CT/D6/CT /D8/CW/CT /CV/D6/CP/D4/CW/CT/D2/CT /D7/CW/CT/CT/D8 /CX/D7/CT/D1 /CQ /CT/CS/CS/CT/CS/BA/C1/D2 /D8/CW/CX/D7 /D4/CP/D4 /CT/D6/B8 /DB /CT /CS/CX/D7\r/D9/D7/D7 /D8/CW/CT /BW/C7/CB /CP/D2/CS /D8/CW/CT /C4/BW/C7/CB/CU/D3/D6 /D7/CT/DA /CT/D6/CP/D0 /CT/D0/CT\r/D8/D6/D3/D2 \r/D6/DD/D7/D8/CP/D0/D7 /CX/D2 /D8/CW/CTN= 2 /C4/C4 /DB/CX/D8/CW/CX/D2 /CP/C0/CP/D6/D8/D6/CT/CT/B9/BY /D3 \r /CZ /CP/D4/D4/D6/D3 /DC/CX/D1/CP/D8/CX/D3/D2/BA /CF /CT /CW/CP /DA /CT /D4 /CT/D6/CU/D3/D6/D1/CT/CS /D7/CX/D1/CX/D0/CP/D6\r/CP/D0\r/D9/D0/CP/D8/CX/D3/D2/D7 /CU/D3/D6N= 1 /B8 /BF /CP/D2/CS /BG/B8 /CQ/D9/D8 /DB /CT \r/D3/D2\r/CT/D2 /D8/D6/CP/D8/CT /CX/D2/D8/CW/CT /D4/D6/CT/D7/CT/D2 /D8 /D4/CP/D4 /CT/D6 /D3/D2N= 2 /CU/D3/D6 /D8 /DB /D3 /D6/CT/CP/D7/D3/D2/D7/BA /BY/CX/D6/D7/D8/B8 /D8/CW/CT/BW/C7/CB /CP/D2/CS /C4/BW/C7/CB /D6/CT/D7/D9/D0/D8/D7 /CU/D3/D6N= 2 /CP/D6/CT /D6/CT/D4/D6/CT/D7/CT/D2 /D8/CP/D8/CX/DA /CT /D3/CU/CW/CX/CV/CW/B9/AS/CT/D0/CS /CT/D0/CT\r/D8/D6/D3/D2 /D7/D3/D0/CX/CS/D7 /AL /D3/D9/D6 \r/CP/D0\r/D9/D0/CP/D8/CX/D3/D2/D7 /DD/CX/CT/D0/CS /CX/D2/CS/CT/CT/CS/D7/CX/D1/CX/D0/CP/D6 /D6/CT/D7/D9/D0/D8/D7 /CU/D3/D6 /D8/CW/CT /D3/D8/CW/CT/D6 /C4/C4/B3/D7/BA /CB/CT\r/D3/D2/CS/B8 /CU/D3/D6 /CW/CX/CV/CW/CT/D6 /C4/C4/B3/D7/D8/CW/CT/D6/CT /CW/CP /DA /CT /CQ /CT/CT/D2 /D2/D3 \r/D0/CT/CP/D6 /CX/D2/CS/CX\r/CP/D8/CX/D3/D2/D7 /D7/D3 /CU/CP/D6 /CU/D3/D6 /CT/D0/CT\r/D8/D6/D3/D2/B9\r/D6/DD/D7/D8/CP/D0 /D4/CW/CP/D7/CT/D7 /CX/D2 /BZ/CP/BT/D7 /CX/D2 /D8/CW/CT /D5/D9/CP/D2 /D8/D9/D1 /C0/CP/D0/D0 /D6/CT/CV/CX/D1/CT/BA /BY /D3/D6/D3/D9/D6 /D2 /D9/D1/CT/D6/CX\r/CP/D0 \r/CP/D0\r/D9/D0/CP/D8/CX/D3/D2/D7/B8 /DB /CT /CW/CP /DA /CT /CP/CS/D3/D4/D8/CT/CS /D8/CW/CT /CX/D8/CT/D6/CP/B9/D8/CX/DA /CT /D7\r /CW/CT/D1/CT /D4/D6/D3/D4 /D3/D7/CT/CS /CQ /DD /BV\u001d/D8/GH /CP/D2/CS /C5/CP\r/BW/D3/D2/CP/D0/CS/B8\n/BG/BC/B8/BG/BD/DB/CW/CX\r /CW/CW/CP/D7 /CP/D0/D7/D3 /CQ /CT/CT/D2 /CP/D4/D4/D0/CX/CT/CS /D8/D3 \r/CP/D0\r/D9/D0/CP/D8/CT /D8/CW/CT /CT/D2/CT/D6/CV/CX/CT/D7 /CP/D2/CS /D8/CW/CT/D6/CT/CP/D0/B9/D7/D4/CP\r/CT /D4/D6/D3/AS/D0/CT/D7 /D3/CU /DA /CP/D6/CX/D3/D9/D7 /CT/D0/CT\r/D8/D6/D3/D2/B9\r/D6/DD/D7/D8/CP/D0 /D4/CW/CP/D7/CT/D7 /CX/D2/CV/D6/CP/D4/CW/CT/D2/CT/BA\n/BF/BD/BT/D7 /CP /D8/CT/D7/D8 /D3/CU /D8/CW/CT /DA /CP/D0/CX/CS/CX/D8 /DD /D3/CU /D3/D9/D6 \r/D3 /CS/CT/B8 /DB /CT /CW/CP /DA /CT\r/D3/D6/D6/D3/CQ /D3/D6/CP/D8/CT/CS /D8/CW/CT /D6/CT/D7/D9/D0/D8/D7 /D3/CQ/D8/CP/CX/D2/CT/CS /CX/D2 /CA/CT/CU/BA /BF/BD /CP/D2/CS /D8/CW/CT/D2\n/CP/D4/D4/D0/CX/CT/CS /CX/D8 /CU/D3/D6 /D8/CW/CT \r/CP/D0\r/D9/D0/CP/D8/CX/D3/D2 /D3/CU /D8/CW/CT /BW/C7/CB /CP/D2/CS /D8/CW/CT /C4/BW/C7/CB/BA/C6/D3/D8/CX\r/CT /D8/CW/CP/D8/B8 /CS/CT/D7/D4/CX/D8/CT /D8/CW/CT /CW /D9/CV/CT /CP/D1/D3/D9/D2 /D8 /D3/CU /C0/CP/D6/D8/D6/CT/CT/B9/BY /D3 \r /CZ/D7/D8/D9/CS/CX/CT/D7 /D3/CU /D5/D9/CP/D2 /D8/D9/D1 /C0/CP/D0/D0 /CT/D0/CT\r/D8/D6/D3/D2/B9\r/D6/DD/D7/D8/CP/D0 /D4/CW/CP/D7/CT/D7/B8 /D2/D3/D2/CT /CX/D7/CS/CT/DA /D3/D8/CT/CS /D8/D3 /D7/D8/D9/CS/DD /D8/CW/CT /C4/BW/C7/CB /D3/CU /D8/CW/CT/D7/CT /D4/CW/CP/D7/CT/D7/BA /CF /CT /CW/CP /DA /CT\r/CP/D0\r/D9/D0/CP/D8/CT/CS /D8/CW/CT /C4/BW/C7/CB /CP/D8 /CT/D2/CT/D6/CV/CX/CT/D7 /DB/CW/CT/D6/CT /D8/CW/CT /CX/D2 /D8/CT/CV/D6/CP/D8/CT/CS/BW/C7/CB /CW/CP/D7 /DB /CT/D0/D0/B9/D4/D6/D3/D2/D3/D9/D2\r/CT/CS /D4 /CT/CP/CZ/D7/B8 /DB/CW/CX\r /CW /CU/CP/D0/D0 /CX/D2 /D8/D3 /D8 /DB /D3 /CS/CX/D7/B9/D8/CX/D2\r/D8 \r/D0/CP/D7/D7/CT/D7/BM /CQ /D3/D9/D2/CS /D7/D8/CP/D8/CT/D7 /CP/D8 /D2/CT/CV/CP/D8/CX/DA /CT /CT/D2/CT/D6/CV/DD /DB/CX/D8/CW /D6/CT/B9/D7/D4 /CT\r/D8 /D8/D3 /D8/CW/CT \r /CW/CT/D1/CX\r/CP/D0 /D4 /D3/D8/CT/D2 /D8/CX/CP/D0 /CP/D2/CS /CW/CX/CV/CW/B9/CT/D2/CT/D6/CV/DD /D4 /CT/CP/CZ/D7/CP/CQ /D3 /DA /CT/BA /CC/CW/CT /D2 /D9/D1 /CQ /CT/D6 /D3/CU /D2/CT/CV/CP/D8/CX/DA /CT/B9/CT/D2/CT/D6/CV/DD /D4 /CT/CP/CZ/D7 /CX/D7 /CX/CS/CT/D2 /D8/CX\r/CP/D0/DB/CX/D8/CW /D8/CW/CT /D2 /D9/D1 /CQ /CT/D6 /D3/CU /CT/D0/CT\r/D8/D6/D3/D2/D7 Me\n/D4 /CT/D6 /CQ/D9/CQ/CQ/D0/CT/B8 /CX/D2 /CP/CV/D6/CT/CT/B9/D1/CT/D2 /D8 /DB/CX/D8/CW /CQ/D9/CQ/CQ/D0/CT \r/D6/DD/D7/D8/CP/D0/D7 /CX/D2 /D8/CW/CT \r/D3/D2 /DA /CT/D2 /D8/CX/D3/D2/CP/D0 /BE/BW/BX/BZ/BA\n/BK/BY /D9/D6/D8/CW/CT/D6/D1/D3/D6/CT/B8 /DB /CT /AS/D2/CS /D8/CW/CP/D8 /D8/CW/CT /D7/D9/D1 /D3/CU /D8/CW/CT /C4/BW/C7/CB /CP/D8 /D8/CW/CT/D7/CT\nMe\n/D2/CT/CV/CP/D8/CX/DA /CT/B9/CT/D2/CT/D6/CV/DD /D4 /CT/CP/CZ/D7 /D6/CT/D4/D6/D3 /CS/D9\r/CT/D7 /D8/CW/CT /D6/CT/CP/D0/B9/D7/D4/CP\r/CT /CS/CT/D2/B9/D7/CX/D8 /DD /D4/D6/D3/AS/D0/CT /D3/CU /D8/CW/CTMe\n/B9/CT/D0/CT\r/D8/D6/D3/D2 /CQ/D9/CQ/CQ/D0/CT \r/D6/DD/D7/D8/CP/D0/BA/CC/CW/CX/D7 /D4/CP/D4 /CT/D6 /CX/D7 /D3/D6/CV/CP/D2/CX/DE/CT/CS /CP/D7 /CU/D3/D0/D0/D3 /DB/D7/BA /C1/D2 /CB/CT\r/D8/CX/D3/D2 /C1 /C1/B8 /DB /CT/D3/D9/D8/D0/CX/D2/CT /D8/CW/CT /CQ/CP/D7/CX\r /D7/D8/CT/D4/D7 /D3/CU /D8/CW/CT /C0/CP/D6/D8/D6/CT/CT/B9/BY /D3 \r /CZ /CP/D4/D4/D6/D3 /DC/CX/D1/CP/D8/CX/D3/D2/D8/D3 /D8/CW/CT /BE/BW/BX/BZ /CX/D2 /CV/D6/CP/D4/CW/CT/D2/CT/BA /C1/D2 /CB/CT\r/D8/CX/D3/D2 /C1 /C1 /C1 /B8 /DB /CT /D4/D6/CT/D7/CT/D2 /D8 /D2 /D9/B9/D1/CT/D6/CX\r/CP/D0 /D6/CT/D7/D9/D0/D8/D7 /CU/D3/D6 /D8/CW/CT /BW/C7/CB /CP/D2/CS /D8/CW/CT /C4/BW/C7/CB /CX/D2 /CV/D6/CP/D4/CW/CT/D2/CT/BA/BY/CX/D2/CP/D0/D0/DD /B8 /DB /CT /CS/D6/CP /DB /D3/D9/D6 \r/D3/D2\r/D0/D9/D7/CX/D3/D2/D7 /CX/D2 /CB/CT\r/D8/CX/D3/D2 /C1/CE /BA/C1 /C1/BA /C0/BT/CA /CC/CA/BX/BX/B9/BY /C7/BV/C3 /C0/BT/C5/C1/C4 /CC/C7/C6/C1/BT/C6/BY /D3/D6 /CP /D4/CP/D6/D8/CX/CP/D0/D0/DD /AS/D0/D0/CT/CS /C4/C4N /B8 /D8/CW/CT /D0/D3 /DB/B9/CT/D2/CT/D6/CV/DD /CT/D0/CT\r/D8/D6/D3/D2/CX\r/D4/D6/D3/D4 /CT/D6/D8/CX/CT/D7 /CP/D6/CT \r/CP/D4/D8/D9/D6/CT/CS /DB/CX/D8/CW/CX/D2 /CP /D1/D3 /CS/CT/D0 /D8/CW/CP/D8 /D8/CP/CZ /CT/D7 /CX/D2 /D8/D3/CP\r\r/D3/D9/D2 /D8 /D7/D8/CP/D8/CT/D7 /D3/D2/D0/DD /DB/CX/D8/CW/CX/D2 /D8/CW/CX/D7 /D0/CT/DA /CT/D0/BA /C1/D2 /D8/CW/CX/D7 \r/CP/D7/CT/B8 /D8/CW/CT/D7/CX/D2/CV/D0/CT/B9/D4/CP/D6/D8/CX\r/D0/CT /CZ/CX/D2/CT/D8/CX\r /CT/D2/CT/D6/CV/DD /CX/D7 /D8/CW/CT /D7/CP/D1/CT /CU/D3/D6 /CP/D0/D0 /D3/CU /D7/D8/CP/D8/CT/D7/B8/CP/D2/CS /D8/CW /D9/D7 /D3/D2/D0/DD /D8/CW/CT /CX/D2 /D8/CT/D6/CP\r/D8/CX/D3/D2 /D8/CT/D6/D1 /CX/D7 /D6/CT/D0/CT/DA /CP/D2 /D8/BA /BY /D9/D6/D8/CW/CT/D6/B9/D1/D3/D6/CT/B8 /DB /CT /D3/D1/CX/D8 /D8/CW/CT /D4/CW /DD/D7/CX\r/CP/D0 /D7/D4/CX/D2 /DB/CW/CX\r /CW /DB /CT \r/D3/D2/D7/CX/CS/CT/D6 /D8/D3 /CQ /CT\r/D3/D1/D4/D0/CT/D8/CT/D0/DD /D4 /D3/D0/CP/D6/CX/DE/CT/CS/B8 /CT/BA/CV/BA /CS/D9/CT /D8/D3 /CP /D7/D9Ꜷ\r/CX/CT/D2 /D8/D0/DD /D0/CP/D6/CV/CT /CI/CT/CT/B9/D1/CP/D2 /CT/AR/CT\r/D8/BA /CC/CW/CT /CS/CT/D6/CX/DA /CP/D8/CX/D3/D2 /D3/CU /D8/CW/CT /C0/CP/D6/D8/D6/CT/CT/B9/BY /D3 \r /CZ /C0/CP/D1/CX/D0/B9/D8/D3/D2/CX/CP/D2 /CU/D3/D6 /D8/CW/CT /BE/BW/BX/BZ /CX/D2 /BZ/CP/BT/D7 /CW/CP/D7 /CQ /CT/CT/D2 /CT/DC/D8/CT/D2/D7/CX/DA /CT/D0/DD /CS/CX/D7/B9\r/D9/D7/D7/CT/CS /CX/D2 /D8/CW/CT /D0/CX/D8/CT/D6/CP/D8/D9/D6/CT/BA\n/BG/BC/B8/BG/BD/C1/D2 /CV/D6/CP/D4/CW/CT/D2/CT/B8 /D8/CW/CT /CX/D2 /D8/CT/D6/CP\r/D8/CX/D3/D2/C0/CP/D1/CX/D0/D8/D3/D2/CX/CP/D2 /CU/D3/D6 /D8/CW/CT /BE/BW/BX/BZ /CX/D7 /D7/CX/D1/CX/D0/CP/D6 /D8/D3 /D8/CW/CP/D8 /CX/D2 /BZ/CP/BT/D7/B8 /CP/D0/B9/CQ /CT/CX/D8 /DB/CX/D8/CW /CS/CX/AR/CT/D6/CT/D2 /D8 /CU/D3/D6/D1 /CU/CP\r/D8/D3/D6/D7 /CS/D9/CT /D8/D3 /D8/CW/CT /D7/D4/CX/D2/D3/D6/CX/CP/D0 /CU/D3/D6/D1/D3/CU /D8/CW/CT /DB /CP /DA /CT /CU/D9/D2\r/D8/CX/D3/D2/D7/BA\n/BE/BI/B8/BG/BE/CC/CW/CX/D7 /D7/CX/D1/CX/D0/CP/D6/CX/D8 /DD /CP/D0/D0/D3 /DB/D7 /D3/D2/CT /D8/D3/D9/D7/CT /D8/CW/CT /D7/CP/D1/CT /D8/CW/CT/D3/D6/CT/D8/CX\r/CP/D0 /D1/CT/D8/CW/D3 /CS/D7 /DB/CW/CX\r /CW /DB /CT/D6/CT /D9/D7/CT/CS /D4/D6/CT/B9/DA/CX/D3/D9/D7/D0/DD /D8/D3 /D7/D8/D9/CS/DD /D8/CW/CT /BE/BW/BX/BZ /CX/D2 /BZ/CP/BT/D7/B8 /DB/CX/D8/CW /D8/CW/CT /CX/D1/D4 /D3/D6/D8/CP/D2 /D8/CS/CX/AR/CT/D6/CT/D2\r/CT /D8/CW/CP/D8 /DB /CT /D2/CT/CT/CS /D8/D3 /D8/CP/CZ /CT /CX/D2 /D8/D3 /CP\r\r/D3/D9/D2 /D8 /D8/CW/CT /D8 /DB /D3/CU/D3/D0/CS/DA /CP/D0/D0/CT/DD /CS/CT/CV/CT/D2/CT/D6/CP\r/DD /CX/D2 /D8/CW/CT /CU/D3/D6/D1 /D3/CU /CP/D2 /CB/CD/B4/BE/B5 /D4/D7/CT/D9/CS/D3/D7/D4/CX/D2/CS/CT/CV/D6/CT/CT /D3/CU /CU/D6/CT/CT/CS/D3/D1/B8 β=±1 /BA /C8/D6/D3 /DA/CX/CS/CT/CS /D8/CW/CP/D8 /CX/D2 /D8/CT/D6/B9/C4/C4 /D8/D6/CP/D2/B9/D7/CX/D8/CX/D3/D2/D7 /CP/D6/CT /D2/CT/CV/D0/CT\r/D8/CT/CS/B8 /DB /CT /D1/CP /DD /DB/D6/CX/D8/CT /D8/CW/CT /CX/D2 /D8/CT/D6/CP\r/D8/CX/D3/D2 /D4/CP/D6/D8/D3/CU /D8/CW/CT /CU/D9/D0/D0 /C0/CP/D1/CX/D0/D8/D3/D2/CX/CP/D2 /CU/D3/D6 /D8/CW/CT /BE/BW/BX/BZ /D3/CU /D7/D4/CX/D2/D0/CT/D7/D7 /CT/D0/CT\r/D8/D6/D3/D2/D7/CX/D2 /CV/D6/CP/D4/CW/CT/D2/CT /CP/D7\n/BE/BI\nˆHint=VC\n2/summationdisplay\nβ,β′, /D51\n| /D5|[FN( /D5)]2ˆρβ,β(− /D5)ˆρβ′,β′( /D5), /B4/BD/B5/DB/CW/CT/D6/CTVC≡e2/lBǫ /CX/D7 /D8/CW/CT /BV/D3/D9/D0/D3/D1 /CQ /CT/D2/CT/D6/CV/DD /D7\r/CP/D0/CT/B8 /DB/CX/D8/CW\nlB=/radicalbig\n/planckover2pi1/eB /D8/CW/CT /D1/CP/CV/D2/CT/D8/CX\r /D0/CT/D2/CV/D8/CW/B8B /D8/CW/CT /D1/CP/CV/D2/CT/D8/CX\r /AS/CT/D0/CS/B8/CP/D2/CSǫ /D8/CW/CT /CS/CX/CT/D0/CT\r/D8/D6/CX\r /D7/D9/D7\r/CT/D4/D8/CX/CQ/CX/D0/CX/D8 /DD /D3/CU /D8/CW/CT /D1/CT/CS/CX/D9/D1/B8 /CP/D2/CS/D5≡(qx,qy) /CX/D7 /CP /BE/BW /DB /CP /DA /CT/DA /CT\r/D8/D3/D6/BA /CC/CW/CT /B4/CV/D9/CX/CS/CX/D2/CV \r/CT/D2 /D8/CT/D6/B5/CS/CT/D2/D7/CX/D8 /DD /D3/D4 /CT/D6/CP/D8/D3/D6 /CX/D2 /D8/CW/CT /C4/CP/D2/CS/CP/D9 /CV/CP/D9/CV/CT /D6/CT/CP/CS/D7\nˆρβ1,β2( /D5) =N−1\nφ/summationdisplay\nXexp/parenleftbigg\n−iqxX−il2\nBqxqy\n2/parenrightbigg\n׈c†\nX,β1ˆcX+l2\nBqy,β2. /B4/BE/B5/BF/C0/CT/D6/CT/B8Nφ=S/2πl2\nB\n/D1/CT/CP/D7/D9/D6/CT/D7 /D8/CW/CT /C4/C4 /CS/CT/CV/CT/D2/CT/D6/CP\r/DD /B8 /DB/CX/D8/CW/D8/CW/CT /D7/D5/D9/CP/D6/CT /CP/D6/CT/CPS /D3/CU /D8/CW/CT /BE/BW/BX/BZ /D7/CP/D1/D4/D0/CT/B8ˆcX,β\n/CP/D2/CSˆc†\nX,β\n/CP/D6/CT/D8/CW/CT /CT/D0/CT\r/D8/D6/D3/D2/B3/D7 /CS/CT/D7/D8/D6/D9\r/D8/CX/D3/D2 /CP/D2/CS \r/D6/CT/CP/D8/CX/D3/D2 /D3/D4 /CT/D6/CP/D8/D3/D6/D7/B8 /D6/CT/D7/D4 /CT\r/B9/D8/CX/DA /CT/D0/DD /B8 /DB/CW/CT/D6/CTX /CS/CT/D2/D3/D8/CT/D7 /D7/CX/D2/CV/D0/CT/B9/D4/CP/D6/D8/CX\r/D0/CT /D5/D9/CP/D2 /D8/D9/D1 /D7/D8/CP/D8/CT/D7/DB/CX/D8/CW/CX/D2 /D8/CW/CTN /B9/D8/CW /C4/C4/BA /BY/CX/D2/CP/D0/D0/DD /B8 /CX/D2 /BX/D5/BA /B4/BD /B5 /D8/CW/CT /CV/D6/CP/D4/CW/CT/D2/CT/CU/D3/D6/D1 /CU/CP\r/D8/D3/D6FN( /D5) /D6/CT/CP/CS/D7\n/BE/BI /B8/BG/BE\nFN( /D5) =\n\n1\n2/bracketleftBig\nL|N|/parenleftBig\nq2\n2/parenrightBig\n+L|N|−1/parenleftBig\nq2\n2/parenrightBig/bracketrightBig\ne−q2\n4, N∝negationslash= 0;\ne−q2\n4, N = 0,/B4/BF/B5/DB/CW/CT/D6/CTq≡ | /D5| /B8 /CP/D2/CSLn(x) /CX/D7 /D8/CW/CT /C4/CP/CV/D9/CT/D6/D6/CT /D4 /D3/D0/DD/D2/D3/D1/CX/CP/D0 /D3/CU/D3/D6/CS/CT/D6n /BA /CF /CT /D2/D3/D8/CT /D8/CW/CP/D8 /D8/CW/CT /BE/BW/BX/BZ /CU/D3/D6/D1 /CU/CP\r/D8/D3/D6 /CX/D2 /BZ/CP/BT/D7 /CX/D7/CV/CX/DA /CT/D2 /CQ /DD\n/BG/BC\nFN( /D5) =LN/parenleftbiggq2\n2/parenrightbigg\ne−q2/4. /B4/BG/B5/C1/D8 /CX/D7 /CP/D4/D4/CP/D6/CT/D2 /D8 /CU/D6/D3/D1 /BX/D5/D7/BA /B4/BF/B5 /CP/D2/CS /B4/BG/B5 /D8/CW/CP/D8 /D8/CW/CT /CV/D6/CP/D4/CW/CT/D2/CT/B4/D6/CT/D0/CP/D8/CX/DA/CX/D7/D8/CX\r /BE/BW/BX/BZ/B5 /CU/D3/D6/D1 /CU/CP\r/D8/D3/D6 /CX/D7 /D7/CX/D1/D4/D0/DD /CP /D0/CX/D2/CT/CP/D6 \r/D3/D1/B9/CQ/CX/D2/CP/D8/CX/D3/D2 /D3/CU /CU/D3/D6/D1 /CU/CP\r/D8/D3/D6/D7 /CU/D3/D6 /CP/CS/CY/CP\r/CT/D2 /D8 /C4/C4/B3/D7 /D3/CU /D8/CW/CT /D2/D3/D2/B9/D6/CT/D0/CP/D8/CX/DA/CX/D7/D8/CX\r /BE/BW/BX/BZ /CX/D2 /BZ/CP/BT/D7/BA /CC/CW/CX/D7 /D4 /CT\r/D9/D0/CX/CP/D6 /CU/CP\r/D8 /D6/CT/D7/D9/D0/D8/D7/CU/D6/D3/D1 /D1/CX/DC/CX/D2/CV /D8/CW/CT /BW/CX/D6/CP\r /D4/CP/D6/D8/CX\r/D0/CT /DB /CP /DA /CT/CU/D9/D2\r/D8/CX/D3/D2/D7 /CQ /CT/D8 /DB /CT/CT/D2/D8/CW/CT /D7/CX/D8/CT/D7 /D3/CU /D8 /DB /D3 /D7/D9/CQ/D0/CP/D8/D8/CX\r/CT/D7 /CX/D2 /CV/D6/CP/D4/CW/CT/D2/CT/B8 /CP/D2/CS /CX/D7 /CP/D0/D7/D3 /CP\r/D3/D2/D7/CT/D5/D9/CT/D2\r/CT /D3/CU /D8/CW/CT /D7/D4/CX/D2/D3/D6/CX/CP/D0 /D2/CP/D8/D9/D6/CT /D3/CU /D8/CW/CT/D7/CT /DB /CP /DA /CT/CU/D9/D2\r/B9/D8/CX/D3/D2/D7/BA /BT/D4/CP/D6/D8 /CU/D6/D3/D1 /D8/CW/CT /CS/CX/AR/CT/D6/CT/D2\r/CT /CX/D2 /CU/D3/D6/D1 /CU/CP\r/D8/D3/D6/D7 /CV/CX/DA /CT/D2 /CQ /DD/BX/D5/D7/BA /B4/BF/B5 /CP/D2/CS /B4/BG/B5/B8 /D8/CW/CT /BE/BW/BX/BZ /CX/D2 /BZ/CP/BT/D7 /CP/D2/CS /CV/D6/CP/D4/CW/CT/D2/CT /CX/D7 /CS/CT/B9/D7\r/D6/CX/CQ /CT/CS /CT/D5/D9/CX/DA /CP/D0/CT/D2 /D8/D0/DD /B8 /CP/D7 /CU/D3/D0/D0/D3 /DB/D7 /CU/D6/D3/D1 /D8/CW/CT /D7/CP/D1/CT /CP/D2/CP/D0/DD/D8/CX\r/CP/D0/D7/D8/D6/D9\r/D8/D9/D6/CT /D3/CU /D8/CW/CT /BV/D3/D9/D0/D3/D1 /CQ /CX/D2 /D8/CT/D6/CP\r/D8/CX/D3/D2 /D8/CT/D6/D1 /CV/CX/DA /CT/D2 /CQ /DD /BX/D5/BA/B4/BD/B5/BA/BY/CX/D2/CP/D0/D0/DD /B8 /DB /CT /D2/D3/D8/CT /D8/CW/CP/D8 /D8/CW/CT /C0/CP/D1/CX/D0/D8/D3/D2/CX/CP/D2 /CX/D2 /BX/D5/BA /B4/BD/B5 /CX/D7/CB/CD/B4/BE/B5/B9/CX/D2 /DA /CP/D6/CX/CP/D2 /D8 /DB/CX/D8/CW /D6/CT/D7/D4 /CT\r/D8 /D8/D3 /D8/CW/CT /DA /CP/D0/D0/CT/DD /D4/D7/CT/D9/CS/D3/D7/D4/CX/D2/BA/C1/D2 \r/D3/D2 /D8/D6/CP/D7/D8 /D8/D3 /D8/CW/CT /D4/CW /DD/D7/CX\r/CP/D0 /CT/D0/CT\r/D8/D6/D3/D2 /D7/D4/CX/D2/B8 /D8/CW/CX/D7 /CB/CD/B4/BE/B5 /D7/DD/D1/B9/D1/CT/D8/D6/DD /CX/D7 /CP/D4/D4/D6/D3 /DC/CX/D1/CP/D8/CT/BA /C0/D3 /DB /CT/DA /CT/D6/B8 /CB/CD/B4/BE/B5/B9/D7/DD/D1/D1/CT/D8/D6/DD /CQ/D6/CT/CP/CZ/B9/CX/D2/CV /D8/CT/D6/D1/D7 /CP/D6/CT /D7/D9/D4/D4/D6/CT/D7/D7/CT/CS /D0/CX/D2/CT/CP/D6/D0/DD /CX/D2a/lB≪1 /DB/CW/CT/D6/CT\na= 0.14 /D2/D1 /CX/D7 /D8/CW/CT \r/CP/D6/CQ /D3/D2/B9\r/CP/D6/CQ /D3/D2 /CS/CX/D7/D8/CP/D2\r/CT /CX/D2 /CV/D6/CP/D4/CW/CT/D2/CT/CP/D2/CSlB= 26//radicalbig\nB[T] /D2/D1/B8 /CX/BA/CT/BA /CP/D8 /CP/D2 /CT/D2/CT/D6/CV/DD /D7\r/CP/D0/CT /D8/CW/CP/D8/CX/D7 /DB /CT/D0/D0 /CQ /CT/D0/D3 /DB /D8/CW/CT /CS/CX/D7/D3/D6/CS/CT/D6 /CQ/D6/D3/CP/CS/CT/D2/CX/D2/CV /D3/CU /D8/CW/CT /C4/C4/B3/D7/BA\n/BE/BI/B8/BG/BF/CC/CW/CX/D7 /D4/CW /DD/D7/CX\r/CP/D0 /D1/D3 /CS/CT/D0 /CX/D7 /D7/CX/D1/CX/D0/CP/D6 /D8/D3 /CP/D2/D3/D8/CW/CT/D6 /D8 /DB /D3/B9\r/D3/D1/D4 /D3/D2/CT/D2 /D8/D5/D9/CP/D2 /D8/D9/D1 /C0/CP/D0/D0 /D7/DD/D7/D8/CT/D1 /AL /CX/CU /D3/D2/CT /D6/CT/D4/D0/CP\r/CT/D7 /CX/D2 /BX/D5/BA /B4/BD/B5FN( /D5)/CQ /DD /D8/CW/CT /D2/D3/D2/B9/D6/CT/D0/CP/D8/CX/DA/CX/D7/D8/CX\r /CU/D3/D6/D1/B9/CU/CP\r/D8/D3/D6 FN( /D5) /B8 /D3/D2/CT /D3/CQ/D8/CP/CX/D2/D7/D8/CW/CT /C0/CP/D1/CX/D0/D8/D3/D2/CX/CP/D2 /CU/D3/D6 /D8/CW/CT /D2/D3/D2/B9/D6/CT/D0/CP/D8/CX/DA/CX/D7/D8/CX\r /BE/BW/BX/BZ /CX/D2\r/D0/D9/CS/CX/D2/CV/D8/CW/CT /CT/D0/CT\r/D8/D6/D3/D2/D7/B3 /D7/D4/CX/D2 /CX/D2 /D8/CW/CT /CP/CQ/D7/CT/D2\r/CT /D3/CU /CP /D4 /D3/D0/CP/D6/CX/DE/CX/D2/CV /CI/CT/CT/D1/CP/D2/CT/AR/CT\r/D8/BA /BT/D0/D8/CT/D6/D2/CP/D8/CX/DA /CT/D0/DD /B8 /D8/CW/CX/D7 /D1/D3 /CS/CT/D0 /D1/CP /DD /CS/CT/D7\r/D6/CX/CQ /CT /CP /D5/D9/CP/D2 /D8/D9/D1/C0/CP/D0/D0 /CQ/CX/D0/CP /DD /CT/D6 /CX/D2 /D8/CW/CT /D8/CW/CT/D3/D6/CT/D8/CX\r/CP/D0 /D0/CX/D1/CX/D8 /D3/CU /DE/CT/D6/D3 /D0/CP /DD /CT/D6 /D7/CT/D4/CP/D6/CP/B9/D8/CX/D3/D2/B8 /DB/CW/CT/D6/CT /D8/CW/CT /D8 /DB /D3 /AG/D7/D4/CX/D2/AH /D3/D6/CX/CT/D2 /D8/CP/D8/CX/D3/D2/D7 /CS/CT/D2/D3/D8/CT /D8/CW/CT /D8 /DB /D3/CS/CX/AR/CT/D6/CT/D2 /D8 /D0/CP /DD /CT/D6/D7/BA\n/BG/BD/C7/D2/CT /D1/CP /DD /CU/D9/D6/D8/CW/CT/D6 /D7/CX/D1/D4/D0/CX/CU/DD /D8/CW/CT /D1/D3 /CS/CT/D0 /CX/D2/BX/D5/BA /B4/BD/B5 /CQ /DD /D3/D1/CX/D8/D8/CX/D2/CV /D8/CW/CT /DA /CP/D0/D0/CT/DD /D4/D7/CT/D9/CS/D3/D7/D4/CX/D2 /CS/CT/CV/D6/CT/CT /D3/CU /CU/D6/CT/CT/B9/CS/D3/D1/B8 /CX/D2 /DB/CW/CX\r /CW \r/CP/D7/CT /D3/D2/CT /D4/D6/CT/D7/D9/D4/D4 /D3/D7/CT/D7 /CP \r/D3/D1/D4/D0/CT/D8/CT /DA /CP/D0/D0/CT/DD/D4 /D3/D0/CP/D6/CX/DE/CP/D8/CX/D3/D2 /D3/CU /D8/CW/CT /CT/D0/CT\r/D8/D6/D3/D2/CX\r /D4/CW/CP/D7/CT/D7/B8 /DB/CW/CX\r /CW /DB /D3/D9/D0/CS /D1/CP/DC/CX/B9/D1/CP/D0/D0/DD /D4/D6/D3/AS/D8 /CU/D6/D3/D1 /D8/CW/CT /CT/DC\r /CW/CP/D2/CV/CT /CX/D2 /D8/CT/D6/CP\r/D8/CX/D3/D2/BA /CC/CW/CX/D7 /CT/AR/CT\r/D8/CX/DA /CT/CD/B4/BD/B5 /D1/D3 /CS/CT/D0 /CX/D7 /CS/CT/D7\r/D6/CX/CQ /CT/CS /CQ /DD /D8/CW/CT /CX/D2 /D8/CT/D6/CP\r/D8/CX/D3/D2 /D8/CT/D6/D1\nˆHint=VC\n2/summationdisplay/D51\n| /D5|[FN( /D5)]2ˆρ(− /D5)ˆρ( /D5), /B4/BH/B5/DB/CW/CT/D6/CT /D8/CW/CT /CS/CT/D2/D7/CX/D8 /DD /D3/D4 /CT/D6/CP/D8/D3/D6 /D3/CU /D7/D4/CX/D2/D0/CT/D7/D7 /CT/D0/CT\r/D8/D6/D3/D2/D7 ˆρ( /D5) /CX/D7/D3/CQ/D8/CP/CX/D2/CT/CS /CU/D6/D3/D1 /BX/D5/BA /B4/BE/B5 /CQ /DD /D2/CT/CV/D0/CT\r/D8/CX/D2/CV /D8/CW/CT /D4/D7/CT/D9/CS/D3/D7/D4/CX/D2 /CX/D2/B9/CS/CX\r/CT/D7/BA /CC/CW/CX/D7 /D7/CX/D1/D4/D0/CX/AS/CT/CS /CD/B4/BD/B5 /D1/D3 /CS/CT/D0 /D3/CU /CU/D9/D0/D0/DD /DA /CP/D0/D0/CT/DD/B9/D4 /D3/D0/CP/D6/CX/DE/CT/CS\n/CV/D6/CP/D4/CW/CT/D2/CT /DB/CW/CX\r /CW /CX/D7 /CS/CT/D7\r/D6/CX/CQ /CT/CS /CQ /DD /BX/D5/BA /B4/BH/B5 /CX/D7 \r/CP/D0/D0/CT/CS /CD/B4/BD/B5/B9/CV/D6 /CP/D4/CW/CT/D2/CT /CX/D2 /D8/CW/CT /D6/CT/D1/CP/CX/D2/CS/CT/D6 /D3/CU /D8/CW/CT /D4/CP/D4 /CT/D6/BA /C6/D3 /DB/B8 /CX/CU /D3/D2/CT/D7/D9/CQ/D7/D8/CX/D8/D9/D8/CT/D7 /CX/D2 /D8/D3 /BX/D5/BA /B4/BH /B5 /D8/CW/CT /D2/D3/D2/B9/D6/CT/D0/CP/D8/CX/DA/CX/D7/D8/CX\r /CU/D3/D6/D1/B9/CU/CP\r/D8/D3/D6\nFN( /D5) /B8 /D3/D2/CT /D3/CQ/D8/CP/CX/D2/D7 /D8/CW/CT /D9/D7/D9/CP/D0 /D7/CX/D2/CV/D0/CT/B9/D0/CP /DD /CT/D6 /D5/D9/CP/D2 /D8/D9/D1 /C0/CP/D0/D0/BE/BW/BX/BZ /CU/D3/D6 /D7/D4/CX/D2/B9/D4 /D3/D0/CP/D6/CX/DE/CT/CS /CT/D0/CT\r/D8/D6/D3/D2/D7 /CX/D2 /BZ/CP/BT/D7/BA/CC/CW/CT /C0/CP/D6/D8/D6/CT/CT/B9/BY /D3 \r /CZ /CP/D4/D4/D6/D3 /DC/CX/D1/CP/D8/CX/D3/D2 /CP/D4/D4/D0/CX/CT/CS /D8/D3 /D8/CW/CT/CV/D6/CP/D4/CW/CT/D2/CT /CX/D2 /D8/CT/D6/CP\r/D8/CX/D3/D2 /D8/CT/D6/D1 /CX/D2 /BX/D5/BA /B4/BD/B5 /DD/CX/CT/D0/CS/D7\n/BF/BD/B8/BG/BD\nˆH(HF)\nint=NφVC/summationdisplay\nβ, /C9/braceleftBig/bracketleftbig\nH( /C9)−Xββ( /C9)/bracketrightbig\nˆρβ,β( /C9)\n−Xβ¯β( /C9)ˆρ¯β,β( /C9)/bracerightBig\n, /B4/BI/B5/DB/CW/CT/D6/CT¯β=−β /B8 /CP/D2/CS /C9 /B3/D7 /CP/D6/CT /D8/CW/CT /D6/CT\r/CX/D4/D6/D3 \r/CP/D0 /DB /CP /DA /CT/DA /CT\r/D8/D3/D6/D7 /D3/CU/D8/CW/CT /CF /BV /D0/CP/D8/D8/CX\r/CT/BA /CC/CW/CT /C0/CP/D6/D8/D6/CT/CT /CP/D2/CS /BY /D3 \r /CZ /CT/AR/CT\r/D8/CX/DA /CT /CX/D2 /D8/CT/D6/CP\r/B9/D8/CX/D3/D2 /D4 /D3/D8/CT/D2 /D8/CX/CP/D0/D7 /D6/CT/CP/CS/B8 /D6/CT/D7/D4 /CT\r/D8/CX/DA /CT/D0/DD /B8\nH( /C9) =e−Q2/2\nQ|FN( /C9)|2ρ(− /C9)(1−δ/C9,0), /B4/BJ/B5\nXββ′( /C9) =/integraldisplay∞\n0dxe−x2\n2|FN( /C9)|2J0(xQ)ρβ,β′(− /C9),/B4/BK/B5/DB/CW/CT/D6/CTQ≡ | /C9| /B8J0\n/CX/D7 /CP /BU/CT/D7/D7/CT/D0 /CU/D9/D2\r/D8/CX/D3/D2/B8 /CP/D2/CS /D8/CW/CT /CS/CT/D2/D7/CX/D8 /DD/CP /DA /CT/D6/CP/CV/CT/D7 /CP/D6/CTρβ,β′( /C9) =∝angbracketleftˆρβ,β′( /C9)∝angbracketright /B8ρ( /C9) =/summationtext\nβρβ,β( /C9) /BA/CF /CT /CP/D7/D7/D9/D1/CT /CP /D8/D6/CX/CP/D2/CV/D9/D0/CP/D6 /CT/D0/CT\r/D8/D6/D3/D2 /D0/CP/D8/D8/CX\r/CT /CU/D3/D6 /D8/CW/CT /CQ/D6/D3/CZ /CT/D2/B9/D7/DD/D1/D1/CT/D8/D6/DD /D7/D8/CP/D8/CT/B8 /DB/CX/D8/CW /D6/CT\r/CX/D4/D6/D3 \r/CP/D0 /D0/CP/D8/D8/CX\r/CT /DA /CT\r/D8/D3/D6/D7 /CV/CX/DA /CT/D2 /CQ /DD/C9=Q0/parenleftBigg\nn\n2,n\n2+m√\n3\n2/parenrightBigg\n, n,m ∈Z. /B4/BL/B5/C0/CT/D6/CTQ0\n/CX/D7 /D8/CW/CT /D0/CT/D2/CV/D8/CW /D3/CU /D8/CW/CT /CQ/CP/D7/CX/D7 /DA /CT\r/D8/D3/D6 /D3/CU /D8/CW/CT /D6/CT\r/CX/D4/D6/D3 \r/CP/D0/D0/CP/D8/D8/CX\r/CT/B8\nQ0=l−1\nB/parenleftbigg4πνN√\n3Me/parenrightbigg1/2\n, /B4/BD/BC/B5\nνN\n/CX/D7 /D8/CW/CT /AS/D0/D0/CX/D2/CV /CU/CP\r/D8/D3/D6 /D3/CU /D8/CW/CT /D0/CP/D7/D8 /D4/CP/D6/D8/CX/CP/D0/D0/DD /AS/D0/D0/CT/CS /C4/C4/B8/CP/D2/CSMe\n/CX/D7 /D8/CW/CT /D2 /D9/D1 /CQ /CT/D6 /D3/CU /CT/D0/CT\r/D8/D6/D3/D2/D7 /D4 /CT/D6 /D7/CX/D8/CT /B4Me= 1\r/D3/D6/D6/CT/D7/D4 /D3/D2/CS/D7 /D8/D3 /D8/CW/CT /CF /BV/B8 /CP/D2/CSMe≥2 /D8/D3 /CP/D2 /CT/D0/CT\r/D8/D6/D3/D2/B9/CQ/D9/CQ/CQ/D0/CT \r/D6/DD/D7/D8/CP/D0 /DB/CX/D8/CWMe\n/CT/D0/CT\r/D8/D6/D3/D2/D7 /D4 /CT/D6 /CQ/D9/CQ/CQ/D0/CT/B5/BA /CC/CW/CT /D7/CX/D2/CV/D0/CT/B9/D4/CP/D6/D8/CX\r/D0/CT /BZ/D6/CT/CT/D2/B3/D7 /CU/D9/D2\r/D8/CX/D3/D2 /CX/D2 /D8/CW/CT /CX/D1/CP/CV/CX/D2/CP/D6/DD/B9/D8/CX/D1/CT /C5/CP/D8/D7/D9/CQ/B9/CP/D6/CP /CU/D3/D6/D1/CP/D0/CX/D7/D1\n/BG/BG/D6/CT/CP/CS/D7\nGβ1,β2( /C9,iωn) =−N−1\nφ/integraldisplay/planckover2pi1/kBT\n0dτexp(iωnτ)\n×/summationdisplay\nXexp/bracketleftbigg\n−iQxX+l2\nBQxQy\n2/bracketrightbigg\n×∝angbracketleftTτˆcX−l2\nBQy,β1(τ)ˆc†\nX,β2(0)∝angbracketright, /B4/BD/BD/B5/DB/CW/CT/D6/CTT /CX/D7 /D8/CW/CT /D8/CT/D1/D4 /CT/D6/CP/D8/D9/D6/CT/B8 kB\n/CX/D7 /D8/CW/CT /BU/D3/D0/D8/DE/D1/CP/D2/D2\r/D3/D2/D7/D8/CP/D2 /D8/B8 Tτ\n/CS/CT/D2/D3/D8/CT/D7 /CX/D1/CP/CV/CX/D2/CP/D6/DD/B9/D8/CX/D1/CT /D3/D6/CS/CT/D6/CX/D2/CV/B8 /CP/D2/CS\nωn=π(2n+ 1)kBT//planckover2pi1 /CP/D6/CT /D8/CW/CT /C5/CP/D8/D7/D9/CQ/CP/D6/CP /CU/D6/CT/D5/D9/CT/D2\r/CX/CT/D7/BA\nGβ1,β2( /C9,iωn) /D1/CP /DD /CQ /CT /CS/CT/D8/CT/D6/D1/CX/D2/CT/CS /D7/CT/D0/CU/B9\r/D3/D2/D7/CX/D7/D8/CT/D2 /D8/D0/DD /CU/D6/D3/D1/D8/CW/CT /D5/D9/CP/CS/D6/CP/D8/CX\r /C0/CP/D1/CX/D0/D8/D3/D2/CX/CP/D2 /B4/BI/B5 /CQ /DD /D9/D7/CX/D2/CV /D8/CW/CT /C0/CT/CX/D7/CT/D2 /CQ /CT/D6/CV/CT/D5/D9/CP/D8/CX/D3/D2/D7 /D3/CU /D1/D3/D8/CX/D3/D2 /DB/CX/D8/CW/CX/D2 /D8/CW/CT /CX/D8/CT/D6/CP/D8/CX/DA /CT/B9/D7/D3/D0/D9/D8/CX/D3/D2 /D1/CT/D8/CW/D3 /CS/D4/D6/D3/D4 /D3/D7/CT/CS /CX/D2 /CA/CT/CU/BA /BG/BH /DB/CW/CX\r /CW /DB /CT /CP/CS/D3/D4/D8 /CX/D2 /D8/CW/CT /D4/D6/CT/D7/CT/D2 /D8 /DB /D3/D6/CZ/BA/BG/BT/CU/D8/CT/D6 /CP/D2/CP/D0/DD/D8/CX\r \r/D3/D2 /D8/CX/D2 /D9/CP/D8/CX/D3/D2 /D8/D3 /D6/CT/CP/D0 /CU/D6/CT/D5/D9/CT/D2\r/CX/CT/D7 iωn→\nω+i0+/B8Gβ1,β2( /C9,iωn) /DD/CX/CT/D0/CS/D7 /D8/CW/CT /D6/CT/D8/CP/D6/CS/CT/CS /BZ/D6/CT/CT/D2/B3/D7 /CU/D9/D2\r/B9/D8/CX/D3/D2 /DB/CW/CX\r /CW /D1/CP /DD /CQ /CT /D9/D7/CT/CS /D8/D3 \r/CP/D0\r/D9/D0/CP/D8/CT /D8/CW/CT /BW/C7/CBg(ω) /B8\ng(ω) =−N−1\nφπ/summationdisplay\nβ\n/C1/D1Gβ,β( /C9= 0,iωn→ω+i0+),/B4/BD/BE/B5/CP/D2/CS /D8/CW/CT /C4/BW/C7/CBA( /D6,ω) /B8\nA( /D6,ω) =−N−1\nφπ/summationdisplay\nβ\n/C1/D1Gβ,β( /D6,iωn→ω+i0+), /B4/BD/BF/B5/DB/CW/CT/D6/CT /D8/CW/CT /BZ/D6/CT/CT/D2/B3/D7 /CU/D9/D2\r/D8/CX/D3/D2 /CX/D2 /D6/CT/CP/D0 /D7/D4/CP\r/CT /D6/CT/CP/CS/D7\nGβ1,β2( /D6,iωn) = (2πl2\nB)−1/summationdisplay/C9,βexp(−i /C9· /D6)FN(− /C9)\n×Gβ1,β2( /C9,iωn). /B4/BD/BG/B5/C1 /C1 /C1/BA /CA/BX/CB/CD/C4 /CC/CB /BT/C6/BW /BW/C1/CB/BV/CD/CB/CB/C1/C7/C6/CB/C1/D2 /D8/CW/CX/D7 /D7/CT\r/D8/CX/D3/D2/B8 /DB /CT /CS/CX/D7\r/D9/D7/D7 /D8/CW/CT /D7/D4 /CT\r/D8/D6/D3/D7\r/D3/D4/CX\r /D4/D6/D3/D4 /CT/D6/D8/CX/CT/D7/D3/CU /D8/CW/CT /CT/D0/CT\r/D8/D6/D3/D2/B9/D7/D3/D0/CX/CS /D4/CW/CP/D7/CT/D7 /CU/D3/D6 /D8/CW/CT /C4/C4N= 2 /BA /BT/D7 /CP/D0/B9/D6/CT/CP/CS/DD /D1/CT/D2 /D8/CX/D3/D2/CT/CS /CX/D2 /D8/CW/CT /CX/D2 /D8/D6/D3 /CS/D9\r/D8/CX/D3/D2/B8 /DB /CT \r/D3/D2\r/CT/D2 /D8/D6/CP/D8/CT /D3/D2/D8/CW/CX/D7 /C4/C4 /CU/D3/D6 /CX/D0/D0/D9/D7/D8/D6/CP/D8/CX/D3/D2 /D4/D9/D6/D4 /D3/D7/CT/D7 /CP/D2/CS /CQ /CT\r/CP/D9/D7/CT /D8/CW/CT/DD /CP/D6/CT/D1/D3/D7/D8/D0/DD /D7/CX/CV/D2/CX/AS\r/CP/D2 /D8 /CU/D6/D3/D1 /D8/CW/CT /D4/CW /DD/D7/CX\r/CP/D0 /D4 /D3/CX/D2 /D8 /D3/CU /DA/CX/CT/DB/BA /CF /CT/CW/CP /DA /CT /D3/CQ/D8/CP/CX/D2/CT/CS /D7/CX/D1/CX/D0/CP/D6 /D6/CT/D7/D9/D0/D8/D7 /CU/D3/D6N= 1 /B8 /BF /CP/D2/CS /BG /B4/D2/D3/D8/CS/CX/D7\r/D9/D7/D7/CT/CS /CW/CT/D6/CT/B5/BA/CF /CT /CW/CP /DA /CT \r /CW/D3/D7/CT/D2 /D8/CW/D6/CT/CT /CS/CX/AR/CT/D6/CT/D2 /D8 /CT/D0/CT\r/D8/D6/D3/D2/B9/D7/D3/D0/CX/CS /D0/CP/D8/D8/CX\r/CT/D7/DB/CW/CX\r /CW /CW/CP /DA /CT /D8/CW/CT /D7/CP/D1/CT /D6/CP/D8/CX/D3νN/Me= 0.14≈1/7 /B8 /CP/D2/CS/CW/CT/D2\r/CT /D8/CW/CT /D7/CP/D1/CT /D0/CP/D8/D8/CX\r/CT /D4 /CT/D6/CX/D3 /CS /CV/CX/DA /CT/D2 /CQ /DD /BX/D5/BA /B4/BD/BC /B5 /BA /C7/D9/D6\r /CW/D3/CX\r/CT /CU/D3/D6 /D8/CW/CTνN/Me\n/D6/CP/D8/CX/D3 /CX/D7 /D6/CP/D8/CW/CT/D6 /CP/D6/CQ/CX/D8/D6/CP/D6/DD /BA /CF /CT /D2/D3/D8/CT/D8/CW/CP/D8 /D8/CW/CTMe= 1 /CP/D2/CSMe= 2 /D7/D8/CP/D8/CT/D7 /DD/CX/CT/D0/CS /CX/D2 /CV/D6/CP/D4/CW/CT/D2/CT /D8/CW/CT/CV/D0/D3/CQ/CP/D0 /CT/D2/CT/D6/CV/DD /D1/CX/D2/CX/D1/CP/B8 /DB/CW/CX/D0/CT /D8/CW/CTMe= 3 /CS/D3 /CT/D7 /D2/D3/D8/BA /C1/D8 /CW/CP/D7/CQ /CT/CT/D2 /D7/CW/D3 /DB/D2 /CX/D2 /CA/CT/CU/BA /BF/BD /D8/CW/CP/D8 /D8/CW/CT /CV/D6/D3/D9/D2/CS /D7/D8/CP/D8/CT /D3/CU /CV/D6/CP/D4/CW/CT/D2/CT/CP/D8νN≤0.43 /CX/D7 /CP/D2 /CP/D2/CX/D7/D3/D8/D6/D3/D4/CX\r /CF/CX/CV/D2/CT/D6 \r/D6/DD/D7/D8/CP/D0 /DB/CW/CT/D6/CT/CP/D7 /CP/D8\n0.28≤νN≤0.43 /D8/CW/CT /CV/D6/D3/D9/D2/CS /D7/D8/CP/D8/CT /CX/D7 /D8/CW/CTMe= 2 /CQ/D9/CQ/CQ/D0/CT\r/D6/DD/D7/D8/CP/D0/B8 /CP/D2/CS /CP/D8νN≤0.28 /D8/CW/CT /CF/CX/CV/D2/CT/D6 \r/D6/DD/D7/D8/CP/D0 /DD/CX/CT/D0/CS/D7 /D8/CW/CT/D0/D3 /DB /CT/D7/D8 /CT/D2/CT/D6/CV/DD /B4Me= 1 /B5/BA /CC/CW/CT/D6/CT/CU/D3/D6/CT/B8 /D8/CW/CTMe= 3 /D4/CW/CP/D7/CT/CX/D7 /D2/D3/D8 /D8/CW/CT /D0/D3 /DB /CT/D7/D8/B9/CT/D2/CT/D6/CV/DD /D7/D8/CP/D8/CT /CX/D2 /CV/D6/CP/D4/CW/CT/D2/CT/BN /D2/CT/DA /CT/D6/D8/CW/CT/D0/CT/D7/D7/B8/CX/D8 /CX/D7 /D9/D7/CT/CU/D9/D0 /D8/D3 /CP/D2/CP/D0/DD/DE/CT /D3/D2 /D8/CW/CT /D7/CP/D1/CT /CU/D3 /D3/D8/CX/D2/CV /CP/D0/D0 /D8/CW/D6/CT/CT \r/CP/D7/CT/D7\nMe= 1,2,3 /BA/BY /D3/D6 \r/D3/D1/D4/D0/CT/D8/CT/D2/CT/D7/D7/B8 /DB /CT /D1/CT/D2 /D8/CX/D3/D2 /D8/CW/CP/D8 /DB /CT /CW/CP /DA /CT /CP/D0/D7/D3 \r/CP/D0/B9\r/D9/D0/CP/D8/CT/CS /D8/CW/CT \r/D3/CW/CT/D7/CX/DA /CT /CT/D2/CT/D6/CV/CX/CT/D7 /D3/CU /D3/D8/CW/CT/D6 /D8 /DD/D4 /CT/D7 /D3/CU /CT/D0/CT\r/D8/D6/D3/D2/B9\r/D6/DD/D7/D8/CP/D0 /D4/CW/CP/D7/CT/D7 /B4/D2/D3/D8 /D3/D2/D0/DD /D8/D6/CX/CP/D2/CV/D9/D0/CP/D6 /CQ/D9/CQ/CQ/D0/CT /D4/CW/CP/D7/CT/D7/B8 /CQ/D9/D8/CP/D0/D7/D3 /CP/D2/CX/D7/D3/D8/D6/D3/D4/CX\r /CF/CX/CV/D2/CT/D6 \r/D6/DD/D7/D8/CP/D0/D7/B5/BA /C7/D9/D6 /D6/CT/D7/D9/D0/D8/D7 /CU/D3/D6 /D8/CW/CT/CT/D2/CT/D6/CV/CX/CT/D7 \r/D3/CX/D2\r/CX/CS/CT /DB/CX/D8/CW /D8/CW/D3/D7/CT /D3/CU /CA/CT/CU/BA /BF/BD /DB/CX/D8/CW /CT/DC\r/CT/D0/D0/CT/D2 /D8 /CP\r/B9\r/D9/D6/CP\r/DD /CP/D2/CS/B8 /D8/CW/CT/D6/CT/CU/D3/D6/CT/B8 \r/D3/D6/D6/D3/CQ /D3/D6/CP/D8/CT /D8/CW/CT /BW/C7/CB /CP/D2/CS /C4/BW/C7/CB/D6/CT/D7/D9/D0/D8/D7 /CS/CX/D7\r/D9/D7/D7/CT/CS /CQ /CT/D0/D3 /DB/BA/BT/BA /B4/C1/D2 /D8/CT/CV/D6/CP/D8/CT/CS/B5 /BW/CT/D2/D7/CX/D8 /DD /D3/CU /D7/D8/CP/D8/CT/D7/C7/D9/D6 /D6/CT/D7/D9/D0/D8/D7 /CU/D3/D6 /D8/CW/CT /BW/C7/CB /CX/D2 /CV/D6/CP/D4/CW/CT/D2/CT /CP/D8N= 2 /CP/D6/CT/D4/D6/CT/D7/CT/D2 /D8/CT/CS /CX/D2 /BY/CX/CV/BA /BD/BA/CF /CT /AS/D2/CS /D8/CW/CP/D8 /D8/CW/CT /BW/C7/CB \r/D3/D2/D7/CX/D7/D8/D7 /D3/CU /D8 /DB /D3 /DB /CT/D0/D0/B9/D7/CT/D4/CP/D6/CP/D8/CT/CS\r/D0/CP/D7/D7/CT/D7 /D3/CU /D4 /CT/CP/CZ/D7/BM /DB /CT/D0/D0/B9/CS/CT/AS/D2/CT/CS /D0/D3 /DB/B9/CT/D2/CT/D6/CV/DD /D4 /CT/CP/CZ/D7 /CP/D6/CT /CU/D3/D9/D2/CS/CQ /CT/D0/D3 /DB /D8/CW/CT \r /CW/CT/D1/CX\r/CP/D0 /D4 /D3/D8/CT/D2 /D8/CX/CP/D0 µ /B8 /DB/CW/CX\r /CW /CX/D7 /D7/CW/CX/CU/D8/CT/CS /D8/D3 /DE/CT/D6/D3\n/CT/D2/CT/D6/CV/DD /B8 /DB/CW/CT/D6/CT/CP/D7 /D8/CW/CT /D0/CP/D6/CV/CT /D2 /D9/D1 /CQ /CT/D6 /D3/CU /D4 /CT/CP/CZ/D7 /CP/CQ /D3 /DA /CTµ /CP/D6/CT/D2/D3/D8 /D8/CW/CP/D8 /CT/CP/D7/CX/D0/DD /CS/CX/D7/D8/CX/D2/CV/D9/CX/D7/CW/CT/CS/BA /CF /CT /D2/D3/D8/CT /CW/CT/D6/CT /D8/CW/CP/D8 /CX/D2/CV/D6/CP/D4/CW/CT/D2/CT /D8/CW/CT /D2 /D9/D1 /CQ /CT/D6 /D3/CU /D0/D3 /DB/B9/CT/D2/CT/D6/CV/DD /D4 /CT/CP/CZ/D7 /CX/D2 /CP/D0/D0 \r/CP/D7/CT/D7/CX/D7 /CT/D5/D9/CP/D0 /D8/D3Me\n/B8 /D8/CW/CT /D2 /D9/D1 /CQ /CT/D6 /D3/CU /CT/D0/CT\r/D8/D6/D3/D2/D7 /CX/D2 /CP /CQ/D9/CQ/CQ/D0/CT/BA/CC/CW/CT /D7/CP/D1/CT /D6/CT/D7/D9/D0/D8 /CW/CP/D7 /CQ /CT/CT/D2 /D3/CQ/D8/CP/CX/D2/CT/CS /CQ /CT/CU/D3/D6/CT /CX/D2 /C0/CP/D6/D8/D6/CT/CT/B9/BY /D3 \r /CZ /D7/D8/D9/CS/CX/CT/D7 /D3/CU /D8/CW/CT /D7/CX/D1/D4/D0/CT/D6 /D7/CX/D2/CV/D0/CT/B9/D0/CP /DD /CT/D6 /BE/BW /D5/D9/CP/D2 /D8/D9/D1/C0/CP/D0/D0 /D7/DD/D7/D8/CT/D1/BA\n/BK/CF /CT \r /CW/CT\r /CZ /CT/CS /D8/CW/CP/D8 /D8/CW/CT /D7/CP/D1/CT /D4/D6/D3/D4 /CT/D6/D8 /DD /CW/D3/D0/CS/D7/D8/D6/D9/CT /CP/D0/D7/D3 /CU/D3/D6 /CD/B4/BD/B5/B9/CV/D6/CP/D4/CW/CT/D2/CT/B8 /CP/D2/CS /CX/D2 /D2/D3/D2/B9/D6/CT/D0/CP/D8/CX/DA/CX/D7/D8/CX\r /D8 /DB /D3/B9\r/D3/D1/D4 /D3/D2/CT/D2 /D8 /D5/D9/CP/D2 /D8/D9/D1 /C0/CP/D0/D0 /D7/DD/D7/D8/CT/D1/D7/B8 /D7/D9\r /CW /CP/D7 /CP /CQ/CX/D0/CP /DD /CT/D6 /DB/CX/D8/CW/DE/CT/D6/D3 /D0/CP /DD /CT/D6 /D7/CT/D4/CP/D6/CP/D8/CX/D3/D2/BA/C1/D2 /D8/CW/CT /D7/CX/D1/D4/D0/CT/D6 /D7/CX/D2/CV/D0/CT/B9/D0/CP /DD /CT/D6 /BE/BW /D5/D9/CP/D2 /D8/D9/D1 /C0/CP/D0/D0 /D7/DD/D7/D8/CT/D1/CX/D2 /BZ/CP/BT/D7/B8 /D8/CW/CT /BW/C7/CB /CP/D8Me= 1 /CT/DC/CW/CX/CQ/CX/D8/D7 /D8/CW/CT /CU/CT/CP/D8/D9/D6/CT/D7 /D3/CU/D8/CW/CT /C0/D3/CU/D7/D8/CP/CS/D8/CT/D6 /CQ/D9/D8/D8/CT/D6/AT/DD /D7/D8/D6/D9\r/D8/D9/D6/CT/BA\n/BK/C1/D8 /D1/CT/CP/D2/D7 /D8/CW/CT /CU/D3/D0/D0/D3 /DB/B9/CX/D2/CV/BM /CV/CX/DA /CT/D2 /D8/CW/CP/D8 /D8/CW/CT /AS/D0/D0/CX/D2/CV /CU/CP\r/D8/D3/D6 /D1/CP /DD /CQ /CT /D6/CT/D4/D6/CT/D7/CT/D2 /D8/CT/CS /CQ /DD/CP /D6/CP/D8/CX/D3 /D3/CU /D8 /DB /D3 /CX/D2 /D8/CT/CV/CT/D6/D7 p, q /DB/CX/D8/CW/D3/D9/D8 /CP \r/D3/D1/D1/D3/D2 /CS/CX/DA/CX/D7/D3/D6/B8\nνN=p/q /B8 /D8/CW/CT/D7/CT /CX/D2 /D8/CT/CV/CT/D6/D7p, q /CS/CT/D8/CT/D6/D1/CX/D2/CT /D8/CW/CT/D2 /D8/CW/CT /D7/D8/D6/D9\r/B9/D8/D9/D6/CT /D3/CU /D8/CW/CT /D7/CX/D2/CV/D0/CT/B9/D4/CP/D6/D8/CX\r/D0/CT /CT/D2/CT/D6/CV/DD /D7/D4 /CT\r/D8/D6/D9/D1 /D3/CU /D8/CW/CT /D7/DD/D7/D8/CT/D1/BN/D2/CP/D1/CT/D0/DD /B8 /D8/CW/CT/D6/CT /D7/CW/D3/D9/D0/CS /CT/DC/CX/D7/D8p /D0/D3 /DB/B9/CT/D2/CT/D6/CV/DD /D0/CT/DA /CT/D0/D7 /CP/D2/CSq−p/CW/CX/CV/CW/B9/CT/D2/CT/D6/CV/DD /D0/CT/DA /CT/D0/D7 /B4/C0/D3/CU/D7/D8/CP/CS/D8/CT/D6 /CQ/D9/D8/D8/CT/D6/AT/DD \r/D3/D9/D2 /D8/CX/D2/CV /D6/D9/D0/CT/B5/BA/C1/D2 /D8/CW/CT /BW/C7/CB/B8 /DB/CW/CX\r /CW /CX/D7 /CP /CU/D9/D2\r/D8/CX/D3/D2 /D3/CU /CU/D6/CT/D5/D9/CT/D2\r/DD /B8 /D8/CW/CT/D7/CT /CT/D2/B9/CT/D6/CV/DD /D0/CT/DA /CT/D0/D7 /CP/D6/CT /D6/CT\r/D3/CV/D2/CX/DE/CT/CS /CP/D7 /D7/D1/D3 /D3/D8/CW/CT/CS /D4 /CT/CP/CZ/D7/BA /CC/CW/CT /C0/D3/CU/D7/B9/D8/CP/CS/D8/CT/D6 /CQ/D9/D8/D8/CT/D6/AT/DD \r/D3/D9/D2 /D8/CX/D2/CV /D6/D9/D0/CT /DB /CP/D7 \r/D3/D2/AS/D6/D1/CT/CS /CU/D3/D6Me= 1/CX/D2 /D8/CW/CT /D7/CX/D2/CV/D0/CT/B9/D0/CP /DD /CT/D6 /BE/BW/BX/BZ /CX/D2 /BZ/CP/BT/D7/B8 /DB/CW/CX/D0/CT /CU/D3/D6Me≥2 /CX/D8 /CX/D7\r/D0/CP/CX/D1/CT/CS /D8/CW/CP/D8 \r/D3/D9/D2 /D8/CX/D2/CV /D3/CU /D8/CW/CT /D7/CX/D2/CV/D0/CT/B9/D4/CP/D6/D8/CX\r/D0/CT /D0/CT/DA /CT/D0/D7 /CX/D7 /CS/CX/CU/B9/CU/CT/D6/CT/D2 /D8/BM /D8/CW/CT /D2 /D9/D1 /CQ /CT/D6 /D3/CU /D0/D3 /DB/B9/CT/D2/CT/D6/CV/DD /D4 /CT/CP/CZ/D7 /CX/D7 /CT/D5/D9/CP/D0 /D8/D3Me\n/B8/DB/CW/CT/D6/CT/CP/D7 /D2/D3/D8/CW/CX/D2/CV /CX/D7 /CZ/D2/D3 /DB/D2 /CP/CQ /D3/D9/D8 /DB/CW/CP/D8 /CX/D7 /D8/CW/CT /D4/D6/CT\r/CX/D7/CT /D6/D9/D0/CT/CU/D3/D6 \r/D3/D9/D2 /D8/CX/D2/CV /D8/CW/CT /D2 /D9/D1 /CQ /CT/D6 /D3/CU /CW/CX/CV/CW/B9/CT/D2/CT/D6/CV/DD /D4 /CT/CP/CZ/D7/BA\n/BK/B4/BT/B5 /B4/BU/B5 /B4/BV/B5/BU/D3/D9/D2/CS /D7/D8/CP/D8/CT/D7\n−0.4 −0.2 0.0\nω /CB/CX/D2/CV/D0/CT/B9/D4 /CP/D6/D8/CX\r/D0/CT /CT/DC\r/CX/D8/CP/D8/CX/D3/D2/D7\n/BY/C1/BZ/BA /BD/BM /C4/D3/CV/CP/D6/CX/D8/CW/D1/CX\r /D4/D0/D3/D8/D7 /CU/D3/D6 /D8/CW/CT /CS/CT/D2/D7/CX/D8 /DD /D3/CU /D7/D8/CP/D8/CT/D7g(ω) /D3/CU/CV/D6/CP/D4/CW/CT/D2/CT /CX/D2 /D8/CW/CTN= 2 /C4/CP/D2/CS/CP/D9 /D0/CT/DA /CT/D0 /CP/D8 /B4/BT/B5Me= 1, νN=\n0.14 /BN /B4/BU/B5Me= 2, νN= 0.28 /BN /B4/BV/B5Me= 3, νN= 0.42 /BA/CC/CW/CT /CU/D6/CT/D5/D9/CT/D2\r/DD ω /CX/D7 /CV/CX/DA /CT/D2 /CX/D2 /D9/D2/CX/D8/D7 /D3/CU /D8/CW/CT /BV/D3/D9/D0/D3/D1 /CQ /D7\r/CP/D0/CTVC\n/BA/CC/CW/CTω= 0 /CU/D6/CT/D5/D9/CT/D2\r/DD /CX/D7 /D8/CW/CT /D4 /D3/D7/CX/D8/CX/D3/D2 /D3/CU /D8/CW/CT \r /CW/CT/D1/CX\r/CP/D0 /D4 /D3/D8/CT/D2 /D8/CX/CP/D0/B4/BY /CT/D6/D1/CX /CT/D2/CT/D6/CV/DD/B5µ /BA /CC/CW/CT /CX/D2/AS/D2/CX/D8/CT/D7/CX/D1/CP/D0 /CX/D1/CP/CV/CX/D2/CP/D6/DD /CU/D6/CT/D5/D9/CT/D2\r/DD /D7/CW/CX/CU/D8\nω→ω+i0+/CX/D7 /CP/D4/D4/D6/D3 /DC/CX/D1/CP/D8/CT/CS /CQ /DDω→ω+iδω\n/B8 /DB/CX/D8/CWδω= 10−4/BA/BY/CX/CV/D9/D6/CT/D7 /B4/BT/B5/B9/B4/BV/B5 /CX/D2 /D8/CW/CT /AS/D6/D7/D8 /D6/D3 /DB /DD/CX/CT/D0/CS /D8/CW/CT /BW/C7/CB /CX/D2 /D8/CW/CT /D0/D3 /DB/B9/CT/D2/CT/D6/CV/DD /CU/D6/CT/D5/D9/CT/D2\r/DD /D6/CP/D2/CV/CT /B4/CQ /D3/D9/D2/CS /D7/D8/CP/D8/CT/D7 /D3/CU /CT/D0/CT\r/D8/D6/D3/D2/D7/B5/B8 /DB/CW/CX/D0/CT /CX/D2/D8/CW/CT /D7/CT\r/D3/D2/CS /D6/D3 /DB /D8/CW/CT /CW/CX/CV/CW/B9/CT/D2/CT/D6/CV/DD /D4 /CT/CP/CZ/D7 /D3/CU /D8/CW/CT /BW/C7/CB \r/D3/D6/D6/CT/D7/D4 /D3/D2/CS/D8/D3 /D7/CX/D2/CV/D0/CT/B9/CT/D0/CT\r/D8/D6/D3/D2 /CT/DC\r/CX/D8/CP/D8/CX/D3/D2/D7 /CP/CQ /D3 /DA /CT /D8/CW/CT /CV/D6/D3/D9/D2/CS /D7/D8/CP/D8/CT /D3/CU /D8/CW/CT/D0/CP/D8/D8/CX\r/CT/BA/CF/CW/CP/D8 /DB/CX/D0/D0 /CQ /CT /CX/D1/D4 /D3/D6/D8/CP/D2 /D8 /CX/D2 /D8/CW/CT /CU/D3/D0/D0/D3 /DB/CX/D2/CV /CS/CX/D7\r/D9/D7/D7/CX/D3/D2 /CX/D7/BH/D8/CW/CT /D3/D6 /CS/CT/D6 /D3/CU /CX/D2/CS/CT/DC/CX/D2/CV /D3/CU /D8/CW/CT /BW/C7/CB /D4 /CT /CP/CZ/D7 /BA /CF /CT /DB/CX/D0/D0 \r/D3/D9/D2 /D8/D8/CW/CT /D4 /CT/CP/CZ/D7 /CX/D2 /D8/CW/CT /BW/C7/CB /DB/CX/D8/CW /D6/CT/D7/D4 /CT\r/D8 /D8/D3 /CX/D2\r/D6 /CT /CP/D7/CX/D2/CV /D8/CW/CT /CU/D6 /CT/B9/D5/D9/CT/D2\r/DDω /BA /C1/D2 /BY/CX/CV/BA /BD /B4/BT/B5/B8 /D8/CW/CT /AS/D6/D7/D8 /BW/C7/CB /D4 /CT/CP/CZ /CX/D7 /D3/CQ /DA/CX/B9/D3/D9/D7/D0/DD /D8/CW/CT /D0/D3 /DB /CT/D7/D8/B9/CT/D2/CT/D6/CV/DD /D3/D2/CT /DB/CX/D8/CW /CT/D2/CT/D6/CV/DD≈ −0.24VC\n/BA /CC/CW/CT/D7/CT \r /D3/D2/CS /D4 /CT /CP/CZ /CX/D2 /B4/BT/B5 /CX/D7 /CP /CW/CX/CV/CW/CT/D6/B9/CT/D2/CT/D6/CV/DD /D3/D2/CT /DB/CX/D8/CW /CT/D2/CT/D6/CV/DD\n≈0.17VC\n/BA /CC/CW/CT /D2 /D9/D1 /CQ /CT/D6/CX/D2/CV /D3/CU /D4 /CT/CP/CZ/D7 \r/D3/D2 /D8/CX/D2 /D9/CT/D7 /D9/D2 /D8/CX/D0 /DB /CT/D6/CT/CP\r /CW /D8/CW/CT /D9/D8/D1/D3/D7/D8/B9/D6/CX/CV/CW /D8 /D4 /CT/CP/CZ /DB/CX/D8/CW /CT/D2/CT/D6/CV/DD≈0.3VC\n/BA /CC/CW/CT/D7/CP/D1/CT /CX/D2/CS/CT/DC/CX/D2/CV /D6/D9/D0/CT /CX/D7 /CP/D4/D4/D0/CX/CT/CS /D8/D3 /D8/CW/CT \r/CP/D7/CT/D7 /B4/BU/B5 /CP/D2/CS /B4/BV/B5/BA/C7/D2/CT /D7/CW/D3/D9/D0/CS /D2/D3/D8/CT /D8/CW/CP/D8 /CX/D2 /CP/D0/D0 /D8/CW/D6/CT/CT \r/CP/D7/CT/D7 /D8/CW/CT /BW/C7/CB /D4 /CT/CP/CZ/D7/DB/CX/D8/CW /D8/CW/CT /D7/CP/D1/CT /CX/D2/CS/CT/DC /D1/CP /DD /CW/CP /DA /CT /D6/CP/D8/CW/CT/D6 /CS/CX/AR/CT/D6/CT/D2 /D8 /CT/D2/CT/D6/CV/CX/CT/D7/BM/DB/CW/CX/D0/CT /CX/D2 /B4/BT/B5 /D8/CW/CT /D7/CT\r/D3/D2/CS /BW/C7/CB /D4 /CT/CP/CZ /CQ /CT/D0/D3/D2/CV/D7 /CP/D0/D6/CT/CP/CS/DD /D8/D3 /D8/CW/CT/CW/CX/CV/CW/B9/CT/D2/CT/D6/CV/DD /D6/CT/CV/CX/D3/D2/B8 /CX/D2 /B4/BU/B5 /CX/D8 /D7/D8/CX/D0/D0 /D0/CX/CT/D7 /CQ /CT/D0/D3 /DB /D8/CW/CT /BY /CT/D6/D1/CX/D0/CT/DA /CT/D0/BA/C1/D2 /CP/CS/CS/CX/D8/CX/D3/D2/B8 /D3/D2/CT /D7/CW/D3/D9/D0/CS /CW/CP /DA /CT /CP /D4/D6/D3 \r/CT/CS/D9/D6/CT /D3/CU /CT/DC/D8/D6/CP\r/D8/B9/CX/D2/CV /D8/CW/CT /CT/D2/CT/D6/CV/CX/CT/D7 /D3/CU /D8/CW/CT /BW/C7/CB /D4 /CT/CP/CZ/D7 /CU/D6/D3/D1 /D8/CW/CT /D7/D1/D3 /D3/D8/CW/CT/CS/BW/C7/CB /DA/D7 /CU/D6/CT/D5/D9/CT/D2\r/DD /CS/CT/D4 /CT/D2/CS/CT/D2\r/CT /D7/CW/D3 /DB/D2 /CX/D2 /BY/CX/CV/BA /BD /BA /C1/D2 /D8/CW/CT/D0/D3 /DB/B9/CT/D2/CT/D6/CV/DD /D6/CT/CV/CX/D1/CT /CX/D8 /D1/CP /DD /CQ /CT /CP/D0/DB /CP /DD/D7 /CS/D3/D2/CT /D6/CT/D0/CX/CP/CQ/D0/DD /BA /C1/D2 /D8/CW/CT/CW/CX/CV/CW/B9/CT/D2/CT/D6/CV/DD /D6/CT/CV/CX/D1/CT/B8 /CW/D3 /DB /CT/DA /CT/D6/B8 /D8/CW/CT/D6/CT /CX/D7 /CP /D0/CP/D6/CV/CT/D6 /D2 /D9/D1 /CQ /CT/D6/D3/CU \r/D0/D3/D7/CT/D0/DD/B9/D0/D3 \r/CP/D8/CT/CS /BW/C7/CB /D4 /CT/CP/CZ/D7/B8 /D8/CW/CT /D7/CW/CP/D4 /CT/D7/B8 /DB/CX/CS/D8/CW/D7/B8 /CP/D2/CS/CP/D1/D4/D0/CX/D8/D9/CS/CT/D7 /D3/CU /DB/CW/CX\r /CW /CS/CT/D4 /CT/D2/CS /D7/CT/D2/D7/CX/D8/CX/DA /CT/D0/DD /D3/D2 /D8/CW/CT /CX/D1/CP/CV/CX/D2/CP/D6/DD/CU/D6/CT/D5/D9/CT/D2\r/DD /D7/CW/CX/CU/D8δω\n/BA /CC/CW/CT /D0/CP/D8/D8/CT/D6 /CX/D7 /D9/D7/CT/CS /CU/D3/D6 /D8/CW/CT /CP/D2/CP/D0/DD/D8/CX/B9\r/CP/D0 \r/D3/D2 /D8/CX/D2 /D9/CP/D8/CX/D3/D2 /CX/D2 /D8/D3 /D8/CW/CT /D9/D4/D4 /CT/D6 \r/D3/D1/D4/D0/CT/DC /CW/CP/D0/CU/B9/D4/D0/CP/D2/CT /D3/CU /D8/CW/CT/BZ/D6/CT/CT/D2/B3/D7 /CU/D9/D2\r/D8/CX/D3/D2/B8 iωn→ω+iδω\n/BA /C8/CW /DD/D7/CX\r/CP/D0/D0/DD /B8 /D8/CW/CX/D7 /CX/D1/CP/CV/B9/CX/D2/CP/D6/DD /CU/D6/CT/D5/D9/CT/D2\r/DD /D7/CW/CX/CU/D8 /D6/CT/D4/D6/CT/D7/CT/D2 /D8/D7 /CP /D0/CT/DA /CT/D0 /CQ/D6/D3/CP/CS/CT/D2/CX/D2/CV /CS/D9/CT/B8/CT/BA/CV/BA/B8 /D8/D3 /CS/CX/D7/D3/D6/CS/CT/D6/BA /CF /CT /CW/CP /DA /CT /CU/D3/D9/D2/CS /D8/CW/CP/D8 /D8/CW/CT /CQ /CT/D7/D8 /DB /CP /DD /D8/D3 /CT/DC/B9/D8/D6/CP\r/D8 /D3/D2/D0/DD /D8/CW/D3/D7/CT /D4 /CT/CP/CZ/D7 /DB/CW/CX\r /CW /CP/D6/CT /D4/CW /DD/D7/CX\r/CP/D0 /CX/D7 /D8/D3 /D4/D0/CP\r/CT /CP \r/D9/D8/B9/D3/AR∆ /D3/D2 /D8/CW/CT /BW/C7/CB /D4 /CT/CP/CZ /CP/D1/D4/D0/CX/D8/D9/CS/CT/B8 /D7/D3 /D8/CW/CP/D8 /D4 /CT/CP/CZ/D7 /DB/CX/D8/CW /CP/D1/B9/D4/D0/CX/D8/D9/CS/CT /D0/CT/D7/D7 /D8/CW/CP/D2gmax∆ /CP/D6/CT /D2/CT/CV/D0/CT\r/D8/CT/CS/B8 /DB/CX/D8/CWgmax∝δ−1\nω/D8/CW/CT /D1/CP/DC/CX/D1 /D9/D1 /D4 /CT/CP/CZ /CP/D1/D4/D0/CX/D8/D9/CS/CT/BA /C1/D2 /D3/D9/D6 /D7/D8/D9/CS/DD /B8 /DB /CT /CW/CP /DA /CT\r /CW/D3/D7/CT/D2δω= 10−4/B8 /CP/D2/CS∆ = 0.5 /BA /CC/CW/CT /D2 /D9/D1 /CQ /CT/D6 /D3/CU /D7/CW/CT/D0/D0/D7 /D3/CU/D6/CT\r/CX/D4/D6/D3 \r/CP/D0 /D0/CP/D8/D8/CX\r/CT /DA /CT\r/D8/D3/D6/D7 /C9 /B3/D7 /CX/D7Nsh= 8 /B8 /D7/D3 /D8/CW/CP/D8 /D8/CW/CT /CP\r/B9/D8/D9/CP/D0 /D2 /D9/D1 /CQ /CT/D6 /D3/CU /DA /CT\r/D8/D3/D6/D7 /CX/D7NQ= 241 /BA /CB/CX/D2/CV/D0/CT/B9/D4/CP/D6/D8/CX\r/D0/CT /CT/D2/CT/D6/B9/CV/CX/CT/D7 /DB/CW/CX\r /CW /CP/D6/CT /CT/DC/D8/D6/CP\r/D8/CT/CS /CU/D6/D3/D1 /D8/CW/CT /D7/D1/D3 /D3/D8/CW/CT/CS /BW/C7/CB/B8 /DB/CX/D0/D0 /CQ /CT/D9/D7/CT/CS /CQ /CT/D0/D3 /DB /CX/D2 /D8/CW/CT \r/CP/D0\r/D9/D0/CP/D8/CX/D3/D2 /D3/CU /D8/CW/CT /C4/BW/C7/CB/BA /C1/D2 /D8/CW/CT /CD/B4/BD/B5/B9/CV/D6/CP/D4/CW/CT/D2/CT /CP/D2/CS /D8/CW/CT /D7/CX/D2/CV/D0/CT/B9/D0/CP /DD /CT/D6 \r/CP/D7/CT/D7 /CP/D8 /D8/CW/CT /D7/CP/D1/CT /CS/CT/D2/D7/CX/D8/CX/CT/D7\r/D3/D2/D7/CX/CS/CT/D6/CT/CS /CW/CT/D6/CT/B8 /DB /CT /CP/D6/CT /CP/CQ/D0/CT /D8/D3 /CT/DC/D8/D6/CP\r/D87 /BW/C7/CB /D4 /CT/CP/CZ/D7/BN /CP/D8\nνN= 0.14, Me= 1 /B8 /D8/CW/CT/D6/CT /CX/D7 /D3/D2/CT /D0/D3 /DB /CT/D7/D8/B9/CT/D2/CT/D6/CV/DD /D4 /CT/CP/CZ/B8 /CP/D2/CS\n6 /CW/CX/CV/CW/B9/CT/D2/CT/D6/CV/DD /D3/D2/CT/D7/BA /CC/CW/CT/D7/CT /CP/D6/CT /CT/DC/CP\r/D8/D0/DD /D8/CW/CT /D2 /D9/D1 /CQ /CT/D6/D7 /D3/CU/D7/CX/D2/CV/D0/CT/B9/D4/CP/D6/D8/CX\r/D0/CT /D0/CT/DA /CT/D0/D7 /CS/CX\r/D8/CP/D8/CT/CS /CQ /DD /D8/CW/CT /C0/D3/CU/D7/D8/CP/CS/D8/CT/D6 /CQ/D9/D8/D8/CT/D6/AT/DD\r/D3/D9/D2 /D8/CX/D2/CV /D6/D9/D0/CT/BA\n/BK/C1/D2 /CV/D6/CP/D4/CW/CT/D2/CT /CP/D2/CS /D8/CW/CT /D5/D9/CP/D2 /D8/D9/D1 /C0/CP/D0/D0 /CQ/CX/D0/CP /DD /CT/D6/B8 /D7/D8/D9/CS/CX/CT/CS/DB/CX/D8/CW/CX/D2 /D8/CW/CT /D8 /DB /D3/B9\r/D3/D1/D4 /D3/D2/CT/D2 /D8 /D1/D3 /CS/CT/D0/B8 /DB /CT /D3/CQ/D8/CP/CX/D2 /D8/CW/CT /D2 /D9/D1/B9/CQ /CT/D6 /D3/CU /CX/CS/CT/D2 /D8/CX/AS/CT/CS /BW/C7/CB /D4 /CT/CP/CZ/D7 /CP/D6/D3/D9/D2/CS14 /CJ/DB/CX/D8/CW /CS/CT/DA/CX/CP/D8/CX/D3/D2/D3/CU /D2/D3/D8 /D1/D3/D6/CT /D8/CW/CP/D2 /D3/D2/CT /DB/D6/D3/D2/CV/D0/DD /CX/CS/CT/D2 /D8/CX/AS/CT/CS /D4 /CT/CP/CZ℄/BA /BW/D9/CT /D8/D3/D8/CW/CT /CP/CS/CS/CX/D8/CX/D3/D2/CP/D0 /CB/CD/B4/BE/B5/B9/D7/DD/D1/D1/CT/D8/D6/DD /CX/D2 /D8/CW/CT /D0/CP/D8/D8/CT/D6 /D8 /DB /D3 \r/CP/D7/CT/D7/B8/CX/D8 /CX/D7 /D2/CP/D8/D9/D6/CP/D0 /D8/D3 /CT/DC/D4 /CT\r/D8 /D8/CW/CP/D8 /D8/CW/CT /D2 /D9/D1 /CQ /CT/D6 /D3/CU /D7/CX/D2/CV/D0/CT/B9/D4/CP/D6/D8/CX\r/D0/CT/D0/CT/DA /CT/D0/D7 /CX/D7 /D8/CW /D9/D7 /CS/D3/D9/CQ/D0/CT /CS /BA/BU/BA /CA/CT/CP/D0/B9/D7/D4/CP\r/CT /CS/CT/D2/D7/CX/D8 /DD /D4/D6/D3/AS/D0/CT/BY /D3/D6 /D0/CP/D8/CT/D6 \r/D3/D1/D4/CP/D6/CX/D7/D3/D2 /DB/CX/D8/CW /D3/D9/D6 /D6/CT/D7/D9/D0/D8/D7 /CU/D3/D6 /D8/CW/CT /C4/BW/C7/CB/B8/DB /CT \r/CP/D0\r/D9/D0/CP/D8/CT/CS /D8/CW/CT /D6/CT/CP/D0/B9/D7/D4/CP\r/CT /CT/D0/CT\r/D8/D6/D3/D2 /CS/CT/D2/D7/CX/D8 /DD /D4/D6/D3/AS/D0/CT\nn( /D6) =1\n2πl2\nB/summationdisplay/C9exp(−i /C9· /D6)FN(− /C9)ρ( /C9) /B4/BD/BH/B5/CU/D3/D6 /D8/CW/CT /D7/CP/D1/CT \r /CW/D3/CX\r/CT/D7 /D3/CU /B4νN, Me\n/B5 /CP/D7 /CX/D2 /BY/CX/CV/BA /BD /BA /CC/CW/CT /D6/CT/D7/D9/D0/D8/D7/B8/DB/CW/CX\r /CW /CP/D6/CT /D7/CW/D3 /DB/D2 /CX/D2 /BY/CX/CV/BA /BE/B8 /CP/CV/D6/CT/CT /DB/CX/D8/CW /D4/D6/CT/DA/CX/D3/D9/D7 \r/CP/D0\r/D9/D0/CP/B9\n/D8/CX/D3/D2/D7 /CU/D3/D6 /CV/D6/CP/D4/CW/CT/D2/CT /D4 /CT/D6/CU/D3/D6/D1/CT/CS /CQ /DD /CI/CW/CP/D2/CV /CP/D2/CS /C2/D3/CV/D0/CT/CZ /CP/D6/BA\n/BF/BD/B4/BT/B5 /B4/BU/B5 /B4/BV/B5\n[0.0,2.1] [0 .3,1.7] [0 .6,2.4]\n/BY/C1/BZ/BA /BE/BM /B4/BV/D3/D0/D3/D6 /D3/D2/D0/CX/D2/CT/B5 /CA/CT/CP/D0/B9/D7/D4/CP\r/CT /CS/CT/D2/D7/CX/D8 /DD /D4/D6/D3/AS/D0/CTn( /D6) /CX/D2/CV/D6/CP/D4/CW/CT/D2/CT /CX/D2 /D8/CW/CTN= 2 /C4/C4/BA /BV/CW/D3/CX\r/CT/D7 /B4/BT/B5/B9/B4/BV/B5 /CP/D6/CT /D8/CW/CT /D7/CP/D1/CT/CP/D7 /CX/D2 /BY/CX/CV/BA /BD /BA /C5/CX/D2/CX/D1/CP /CP/D2/CS /D1/CP/DC/CX/D1/CP /D3/CUn( /D6) /DB/D6/CX/D8/D8/CT/D2 /CX/D2/D7/CX/CS/CT /D8/CW/CT/D7/D5/D9/CP/D6/CT /CQ/D6/CP\r /CZ /CT/D8/D7 /CP/D7 /CJ/D1/CX/D2/B8 /D1/CP/DC℄ \r/D3/D6/D6/CT/D7/D4 /D3/D2/CS /D8/D3 /D8/CW/CT /DA /CP/D0/D9/CT/D7 /CJ/BC/BA/BC/B8/BD/BA/BC℄ /CX/D2 /D8/CW/CT \r/D3/D0/D3/D9/D6 /D4/D0/D3/D8/D7 /B4/CQ/D0/D9/CT /CP/D2/CS /DB/CW/CX/D8/CT \r/D3/D0/D3/D9/D6/D7/B8 \r/D3/D6/D6/CT/D7/D4 /D3/D2/CS/B9/CX/D2/CV/D0/DD/B5/BA/BV/BA /C4/D3 \r/CP/D0 /CS/CT/D2/D7/CX/D8 /DD /D3/CU /D7/D8/CP/D8/CT/D7/C7/D9/D6 /D6/CT/D7/D9/D0/D8/D7 /CU/D3/D6 /D8/CW/CT /C4/BW/C7/CB /CX/D2 /CV/D6/CP/D4/CW/CT/D2/CT /CP/D6/CT /D4/D6/CT/D7/CT/D2 /D8/CT/CS/CX/D2 /BY/CX/CV/D7/BA /BF/B8 /BG/B8 /CP/D2/CS /BH /BA /CC/CW/CT /C4/BW/C7/CB /D4/CP/D8/D8/CT/D6/D2/D7 /CX/D7 /D4/D0/D3/D8/D8/CT/CS /CU/D3/D6/CP/D0/D0 /D8/CW/D6/CT/CT \r/CP/D7/CT/D7 /B4/BT/B5/B9/B4/BV/B5/B8 /CP/D7 /CX/D2 /BY/CX/CV/BA /BD /B8 /CP/D2/CS /CP/D8 /D8/CW/CT /CT/D2/CT/D6/B9/CV/CX/CT/D7 /D3/CU /CP/D0/D0 /CT/DC/D8/D6/CP\r/D8/CT/CS /D7/CX/D2/CV/D0/CT/B9/D4/CP/D6/D8/CX\r/D0/CT /BW/C7/CB /D4 /CT/CP/CZ/D7 /D7/CX/D8/D9/CP/D8/CT/CS/CX/D2 /CX/D2\r/D6/CT/CP/D7/CX/D2/CV /D3/D6/CS/CT/D6/BA /CF /CT /D3/CQ/D8/CP/CX/D2 /D8/CW/CP/D8 /D8/CW/CT /D6 /CT/D7\r /CP/D0/CT /CS /B4/D8/D3 /D8/CW/CT/D6/CP/D2/CV/CT /D3/CU[0.0,1.0] /B5 /D6/CT/CP/D0/B9/D7/D4/CP\r/CT /D4/CP/D8/D8/CT/D6/D2/D7 /D3/CUA( /D6,ω) /B8 \r/CP/D0\r/D9/B9/D0/CP/D8/CT/CS /CP/D8 /D8/CW/CT /AS/D6/D7/D8 /CU/D3/D9/D6 /BW/C7/CB /D4 /CT/CP/CZ/D7 /CU/D3/D6 /CP/D0/D0 /D8/CW/D6/CT/CT \r /CW/D3/CX\r/CT/D7/D3/CU(νN, Me) /B8 \r/D3/CX/D2\r/CX/CS/CT /CP/D1/D3/D2/CV /D8/CW/CT/D1/D7/CT/D0/DA /CT/D7/BA /CC/CW/CT/D6/CT /CX/D7 /CP/D0/D7/D3/CP/D2 /CP/D4/D4/D6/D3 /DC/CX/D1/CP/D8/CT /D1/CP/D4/D4/CX/D2/CV /CQ /CT/D8 /DB /CT/CT/D2 /D8/CW/CT /C4/BW/C7/CB /D4/CP/D8/D8/CT/D6/D2/D7 /CP/D8/D8/CW/CT /AS/CU/D8/CW /BW/C7/CB /D4 /CT/CP/CZ/B8 /CP/D0/D8/CW/D3/D9/CV/CW /D0/CT/D7/D7 /D4/D6/D3/D2/D3/D9/D2\r/CT/CS /D8/CW/CP/D2 /CU/D3/D6/D8/CW/CT /AS/D6/D7/D8 /CU/D3/D9/D6 /D3/D2/CT/D7 /B4/D3/D2/CT /D7/CT/CT/D7 \r/D3/D6/D6/CT/D7/D4 /D3/D2/CS/CT/D2\r/CT /CQ /CT/D8 /DB /CT/CT/D2 /D8/CW/CT/D4 /D3/D7/CX/D8/CX/D3/D2/D7 /D3/CU /D1/CP/DC/CX/D1/CP /CP/D2/CS /D1/CX/D2/CX/D1/CP/B8 /CQ/D9/D8 /D8/CW/CT \r/D3/D0/D3/D6/D7 /CS/CT/DA/CX/CP/D8/CT/D7/D0/CX/CV/CW /D8/D0/DD /CX/D2 /CT/CP\r /CW \r/CP/D7/CT/B5/BA /BY /D3/D6 /D0/CP/D6/CV/CT/D6 /DA /CP/D0/D9/CT/D7 /D3/CU /D8/CW/CT /D4 /CT/CP/CZ /CX/D2/B9/CS/CT/DC/B8 /DB /CT /D7/D8/CP/D6/D8 /D8/D3 /D7/CT/CT \r/D3/D2/D7/CX/CS/CT/D6/CP/CQ/D0/CT /CS/CX/D7\r/D6/CT/D4/CP/D2\r/CX/CT/D7 /CQ /CT/D8 /DB /CT/CT/D2/D8/CW/CT /C4/BW/C7/CB /D4/CP/D8/D8/CT/D6/D2/D7/BA /BT/D0/D7/D3 /D8/CW/CT /D2 /D9/D1 /CQ /CT/D6 /D3/CU /CT/DC/D8/D6/CP\r/D8/CT/CS /D4 /CT/CP/CZ/D7\nNp\n/CX/D7 /CS/CX/AR/CT/D6/CT/D2 /D8 /CU/D3/D6 /CT/CP\r /CW \r/CP/D7/CT/BA /CC/CW/CT /D0/CP/D8/D8/CT/D6 /D4/D6/D3/D4 /CT/D6/D8 /DD /CX/D7 /CS/D9/CT/D8/D3 /D8/CW/CT /DA /CT/D6/DD /CP/D4/D4/D6/D3 /DC/CX/D1/CP/D8/CT /D2/CP/D8/D9/D6/CT /D3/CU /D3/D9/D6 /CT/DC/D8/D6/CP\r/D8/CX/D3/D2 /D4/D6/D3 \r/CT/B9/CS/D9/D6/CT/BM /DB/CW/CX/D0/CT /D0/D3 /DB/B9/CT/D2/CT/D6/CV/DD /D4 /CT/CP/CZ/D7 /CP/D6/CT /CP/D0/DB /CP /DD/D7 /CX/CS/CT/D2 /D8/CX/AS/CT/CS /D6/CT/D0/CX/B9/CP/CQ/D0/DD /B8 /D8/CW/CT /CW/CX/CV/CW/B9/CT/D2/CT/D6/CV/DD /D4 /CT/CP/CZ/D7 /CP/D6/CT /CS/CT/D8/CT/D6/D1/CX/D2/CT/CS /D3/D2/D0/DD /CP/D4/D4/D6/D3 /DC/B9/CX/D1/CP/D8/CT/D0/DD /BA /C0/D3 /DB /CT/DA /CT/D6/B8 /D8/CW/CT /CP\r\r/D9/D6/CP\r/DD /CX/D7 /D5/D9/CX/D8/CT /CV/D3 /D3 /CS/BA /CF /CT /CP/D0/D7/D3/D2/D3/D8/CT /CP /DA /CT/D6/DD /CX/D2 /D8/CT/D6/CT/D7/D8/CX/D2/CV /D4/D6/D3/D4 /CT/D6/D8 /DD /D3/CU /D8/CW/CT /C4/BW/C7/CB /CP/D8 /D8/CW/CT /D0/CP/D7/D8/D8 /DB /D3 /D4 /CT/CP/CZ/D7 /CU/D3/D6 /B4/BU/B5/B9/B4/BV/B5/BM /D8/CW/CT /C4/BW/C7/CB /D4/CP/D8/D8/CT/D6/D2/D7 /CP/D6/CT /CX/CS/CT/D2 /D8/CX\r/CP/D0/B8/CQ/D9/D8 /D8/CW/CT/CX/D6 /D4 /D3/D7/CX/D8/CX/D3/D2/D7 /CP/D6/CT /D7/DB/CP/D4/D4 /CT /CS /BA /CF /CT /CS/D3 /D2/D3/D8 /CW/CP /DA /CT /CP/D2 /DD/D4/CW /DD/D7/CX\r/CP/D0 /CP/D6/CV/D9/D1/CT/D2 /D8 /DB/CW /DD /D8/CW/CX/D7 /D7/CW/D3/D9/D0/CS /CQ /CT /D8/CW/CT \r/CP/D7/CT/B8 /CQ/D9/D8 /CX/D8\r/D3/D9/D0/CS /CQ /CT /CP /CU/CT/CP/D8/D9/D6/CT /D8/CW/CP/D8 /CP/D4/D4 /CT/CP/D6/D7 /DB/CW/CT/D2M > N /BA /C0/D3 /DB /CT/DA /CT/D6/B8/D8/CW/CX/D7 /D7/D8/CP/D8/CT/D1/CT/D2 /D8 /CX/D7 /CP /D1/CT/D6/CT /D7/D4 /CT\r/D9/D0/CP/D8/CX/D3/D2 /CP/D2/CS /CP /D1/D3/D6/CT /CS/CT/D8/CP/CX/D0/CT/CS/CX/D2 /DA /CT/D7/D8/CX/CV/CP/D8/CX/D3/D2 /CX/D7 /D6/CT/D5/D9/CX/D6/CT/CS /D8/D3 \r/D0/CP/D6/CX/CU/DD /D8/CW/CX/D7 /CP/D7/D4 /CT\r/D8/BA/BW/BA /BV/D3/D1/D4/CP/D6/CX/D7/D3/D2 /DB/CX/D8/CW /D8/CW/CT /D6/CT/CP/D0/B9/D7/D4/CP\r/CT /CS/CT/D2/D7/CX/D8 /DD/C6/D3 /DB/B8 /D8/D3 \r/D3/D1/D4/CP/D6/CT /D8/CW/CT /C4/BW/C7/CB /D4/CP/D8/D8/CT/D6/D2/D7 /D7/CW/D3 /DB/D2 /CX/D2 /BY/CX/CV/D7/BA/BF /B8 /BG /B8 /BH /DB/CX/D8/CW /D8/CW/CT /D6/CT/CP/D0/B9/D7/D4/CP\r/CT /CS/CT/D2/D7/CX/D8 /DD /D4/D6/D3/AS/D0/CTn( /D6) /CS/CT/AS/D2/CT/CS /CX/D2/BX/D5/BA /B4/BD/BH/B5 /CP/D2/CS /D4/D0/D3/D8/D8/CT/CS /CX/D2 /BY/CX/CV/BA /BE/B8 /DB /CT /CX/D2 /D8/D6/D3 /CS/D9\r/CT /D8/CW/CT /D6 /CT/D7/D9/D1/D1/CT /CS/BI/B4/BT/B5/BD /BE /BF /BG /BH\n/BI /BJ /BK /BL /BD/BC\n/BD/BD /BD/BE /BD/BF\n/BY/C1/BZ/BA /BF/BM /B4/BV/D3/D0/D3/D6 /D3/D2/D0/CX/D2/CT/B5 /C4/BW/C7/CBA( /D6,ω) /CU/D3/D6 /CV/D6/CP/D4/CW/CT/D2/CT /CP/D8νN=\n0.14, Me= 1 /CJ\r/CP/D7/CT /B4/BT/B5℄/BA /BV/D3/D2 /D8/D3/D9/D6 \r/D3/D0/D3/D9/D6/D7 /CP/D6/CT /CV/D6/CP/CS/CT/CS /CX/D2 /D8/CW/CT/D7/CP/D1/CT /DB /CP /DD /CP/D7 /CS/CT/AS/D2/CT/CS /CX/D2 /BY/CX/CV/BA /BE /BA /CC/CW/CT \r/D3/D2 /D8/D3/D9/D6 /D4/D0/D3/D8/D7 /CP/D6/CT /D3/D6/CS/CT/D6/CT/CS/DB/CX/D8/CW /D6/CT/D7/D4 /CT\r/D8 /D8/D3 /D8/CW/CT /CX/D2/CS/CT/DC /D3/CU /BW/C7/CB /D4 /CT/CP/CZ/D7 /CJ/CX/D2/CS/CX\r/CP/D8/CT/CS /CP/CQ /D3 /DA /CT /D8/CW/CT/D4/D0/D3/D8/D7℄/BA /CC/CW/CT /D2 /D9/D1 /CQ /CT/D6 /D3/CU /CT/DC/D8/D6/CP\r/D8/CT/CS /BW/C7/CB /D4 /CT/CP/CZ/D7 /CX/D7Np\n/BP/BD/BF/BA/B4/BU/B5/BD /BE /BF /BG /BH\n/BI /BJ /BK /BL /BD/BC\n/BD/BD /BD/BE /BD/BF /BD/BG /BD/BH\n/BY/C1/BZ/BA /BG/BM /B4/BV/D3/D0/D3/D6 /D3/D2/D0/CX/D2/CT/B5 /C4/BW/C7/CBA( /D6,ω) /CU/D3/D6 /CV/D6/CP/D4/CW/CT/D2/CT /CP/D8νN=\n0.28, Me= 2 /CJ\r/CP/D7/CT /B4/BU/B5℄/BANp= 15 /BA/C4/BW/C7/CB˜A( /D6,ω) /B8 /CS/CT/AS/D2/CT/CS /CU/D3/D6 /CP /AS/DC/CT/CS /D7/CX/D2/CV/D0/CT/B9/D4/CP/D6/D8/CX\r/D0/CT /CT/D2/CT/D6/CV/DD\nωi\n/CP/D7 /CP /D7/D9/D1 /D3/CU /CP/D0/D0 /C4/BW/C7/CB /D4/CP/D8/D8/CT/D6/D2/D7 /CP/D8 /D7/D1/CP/D0/D0/CT/D6 /D4 /CT/CP/CZ /CT/D2/CT/D6/CV/CX/CT/D7/B8\n˜A( /D6,ωi) =i/summationdisplay\nj=1A( /D6,ωj). /B4/BD/BI/B5/BZ/CX/DA /CT/D2 /D8/CW/CT /CT/DC\r/CT/D0/D0/CT/D2 /D8 \r/D3/CX/D2\r/CX/CS/CT/D2\r/CT /D3/CU /D8/CW/CT /C4/BW/C7/CB /D4/CP/D8/D8/CT/D6/D2/D7/D7/CW/D3 /DB/D2 /CX/D2 /BY/CX/CV/BA /BI /DB/CX/D8/CW /D8/CW/CT /D6/CT/CP/D0/B9/D7/D4/CP\r/CT /CS/CT/D2/D7/CX/D8/CX/CT/D7 /CX/D2 /BY/CX/CV/BA /BE /B8/D3/D2/CT /D1/CP /DD /CT/D1/D4/CX/D6/CX\r/CP/D0/D0/DD /DB/D6/CX/D8/CT\nn( /D6,Me=i)↔˜A( /D6,ωi), i= 1,2,3, /B4/BD/BJ/B5\n/B4/BV/B5/BD /BE /BF /BG /BH\n/BI /BJ /BK /BL /BD/BC\n/BD/BD /BD/BE /BD/BF /BD/BG\n/BY/C1/BZ/BA /BH/BM /B4/BV/D3/D0/D3/D6 /D3/D2/D0/CX/D2/CT/B5 /C4/BW/C7/CBA( /D6,ω) /CU/D3/D6 /CV/D6/CP/D4/CW/CT/D2/CT /CP/D8νN=\n0.42, Me= 3 /CJ\r/CP/D7/CT /B4/BV/B5℄/BANp\n/BP/BD/BG/BA/B4/BT/B5/BD /BE /BF\n/BY/C1/BZ/BA /BI/BM /B4/BV/D3/D0/D3/D6 /D3/D2/D0/CX/D2/CT/B5 /CA/CT/D7/D9/D1/D1/CT/CS /C4/BW/C7/CB˜A( /D6,ω) /CU/D3/D6 /CV/D6/CP/D4/CW/CT/D2/CT/CP/D8 /D8/CW/CT /D8/CW/D6/CT/CT /AS/D6/D7/D8 /BW/C7/CB /D4 /CT/CP/CZ/D7 /CU/D3/D6 /B4/BT/B5 /CJνN= 0.14, Me= 1 ℄/BA/BV/D3/D2 /D8/D3/D9/D6 \r/D3/D0/D3/D9/D6/D7 /CP/D6/CT /CV/D6/CP/CS/CT/CS /CX/D2 /D8/CW/CT /D7/CP/D1/CT /DB /CP /DD /CP/D7 /CS/CT/AS/D2/CT/CS /CX/D2 /BY/CX/CV/BA/BE/BA/DB/CW/CT/D6/CT /D8/CW/CT /D7/CX/CV/D2↔ /D1/CT/CP/D2/D7 /D1/CP/D4/D4/CX/D2/CV /CQ /CT/D8 /DB /CT/CT/D2 /D8/CW/CT /D6 /CT/D7\r /CP/D0/CT /CS/D8/D3 /D8/CW/CT /CJ/BC/BA/BC/B8/BD/BA/BC℄ /CX/D2/D8/CT/D6/DA/CP/D0 /D5/D9/CP/D2 /D8/CX/D8/CX/CT/D7/BA /CC/CW/CX/D7 /D1/CT/CP/D2/D7 /D8/CW/CP/D8 /D8/CW/CT/D6/CT/CP/D0/B9/D7/D4/CP\r/CT /CS/CT/D2/D7/CX/D8 /DD /D3/CU /D8/CW/CTMe\n/B9/CT/D0/CT\r/D8/D6/D3/D2 /CQ/D9/CQ/CQ/D0/CT \r/D6/DD/D7/D8/CP/D0 /CX/D7/CS/CT/D8/CT/D6/D1/CX/D2/CT/CS /CQ /DD /D8/CW/CT /D7/D9/D1 /D3/CU /D8/CW/CT /C4/BW/C7/CB /CP/D8 /D8/CW/CTMe\n/D2/CT/CV/CP/D8/CX/DA /CT/B9/CT/D2/CT/D6/CV/DD /D4 /CT/CP/CZ/D7/BA /C5/D3/D6/CT /D7/D9/D6/D4/D6/CX/D7/CX/D2/CV/D0/DD /B8 /CQ /CT\r/CP/D9/D7/CT /D3/CU /D8/CW/CT \r/D3/D6/D6/CT/B9/D7/D4 /D3/D2/CS/CT/D2\r/CT /CQ /CT/D8 /DB /CT/CT/D2 /D8/CW/CT /C4/BW/C7/CB /D4/CP/D8/D8/CT/D6/D2/D7 /D3/CU /D8/CW/CT /D0/D3 /DB/B9/CT/D2/CT/D6/CV/DD/D4 /CT/CP/CZ/D7 /CU/D3/D6 /CP/D0/D0 /CS/CX/AR/CT/D6/CT/D2 /D8 Me\n/CQ/D9/CQ/CQ/D0/CT \r/D6/DD/D7/D8/CP/D0/D7/B8 /D3/D2/CT /D1/CP /DD /CS/CT/B9/D8/CT/D6/D1/CX/D2/CT /D8/CW/CT /D6/CT/CP/D0/B9/D7/D4/CP\r/CT /CS/CT/D2/D7/CX/D8 /DD /D4/CP/D8/D8/CT/D6/D2 /D3/CU /D8/CW/CTMe\n/CQ/D9/CQ/CQ/D0/CT\r/D6/DD/D7/D8/CP/D0 /CQ /DD /D7/D9/D1/D1/CX/D2/CV /D8/CW/CT /C4/BW/C7/CB /D3/CU /D8/CW/CTMe\n/D4 /CT/CP/CZ/D7 /D3/CU /D0/D3 /DB /CT/D7/D8/CT/D2/CT/D6/CV/DD /CU/D3/D6 /CP/D2 /DD /D3/CU /D8/CW/CT /CT/D0/CT\r/D8/D6/D3/D2/B9/D7/D3/D0/CX/CS /D4/CW/CP/D7/CT/D7 /AL /D8/CW/CT /C4/BW/C7/CB/D4/CP/D8/D8/CT/D6/D2/D7 /D3/CU /D8/CW/CTMe= 1 /CF/CX/CV/D2/CT/D6 \r/D6/DD/D7/D8/CP/D0/B8 /CT/BA/CV/BA/B8 \r/D3/D2 /D8/CP/CX/D2/D7/D8/CW /D9/D7 /D8/CW/CT /CX/D2/CU/D3/D6/D1/CP/D8/CX/D3/D2 /D3/CU /D8/CW/CT /CS/CT/D2/D7/CX/D8 /DD /D3/CU /CP/D0/D0 /D3/D8/CW/CT/D6Me\n/CQ/D9/CQ/B9/CQ/D0/CT \r/D6/DD/D7/D8/CP/D0/D7/BA/BX/BA /CD/B4/BD/B5/B9/CV/D6/CP/D4/CW/CT/D2/CT/BM /D0/D3 \r/CP/D0 /CS/CT/D2/D7/CX/D8 /DD /D3/CU /D7/D8/CP/D8/CT/D7/C1/D2 /BY/CX/CV/BA /BJ /DB /CT /D4/D6/CT/D7/CT/D2 /D8 /D8/CW/CT /C4/BW/C7/CB /CU/D3/D6 /CD/B4/BD/B5/B9/CV/D6/CP/D4/CW/CT/D2/CT/BA /CC/CW/CT/D2 /D9/D1 /CQ /CT/D6 /D3/CU /CT/DC/D8/D6/CP\r/D8/CT/CS /BW/C7/CB /D4 /CT/CP/CZ/D7 /CU/D3/D6 /CP/D0/D0 /D8/CW/D6/CT/CT /CS/CT/D2/D7/CX/D8 /DD\r /CW/D3/CX\r/CT/D7 /B4/BT/B5/B9/B4/BV/B5 /CX/D7Np= 7 /BA /CC/CW/CX/D7 /CX/D7 /CX/D2 /CP\r\r/D3/D6/CS/CP/D2\r/CT /DB/CX/D8/CW/D8/CW/CT /C0/D3/CU/D7/D8/CP/CS/D8/CT/D6 /CQ/D9/D8/D8/CT/D6/AT/DD \r/D3/D9/D2 /D8/CX/D2/CV /D6/D9/D0/CT/BA /CF /CT /CP/D0/D7/D3 /D7/CT/CT /CP/D2/CT/DC\r/CT/D0/D0/CT/D2 /D8 \r/D3/D6/D6/CT/D7/D4 /D3/D2/CS/CT/D2\r/CT /D3/CU /D8/CW/CT /C4/BW/C7/CB /CU/D3/D6 /D8/CW/CT /AS/D6/D7/D8 /CU/D3/D9/D6/BW/C7/CB /D4 /CT/CP/CZ/D7/B8 /D8/CW/CT/D2/B8 /CP/D0/D7/D3 /CP /CV/D3 /D3 /CS \r/D3/CX/D2\r/CX/CS/CT/D2\r/CT /D3/CU /D8/CW/CT /C4/BW/C7/CB /CP/D8/BJ/D8/CW/CT /D4 /CT/CP/CZ/D7 /BI/B9/BJ /CU/D3/D6 /B4/BT/B5/B9/B4/BU/B5/B8 /DB/CW/CT/D6/CT/CP/D7 /D8/CW/CT/D7/CT /D8 /DB /D3 /C4/BW/C7/CB /D4/CP/D8/B9/D8/CT/D6/D2/D7 /CU/D3/D6 /B4/BV/B5 /CX/D2 /D8/CT/D6\r /CW/CP/D2/CV/CT /D4/D0/CP\r/CT/D7/B8 /CP/D7 \r/D3/D1/D4/CP/D6/CT/CS /DB/CX/D8/CW /D8/CW/CT/D4/CP/D8/D8/CT/D6/D2/D7 /BI/B9/BJ /CU/D3/D6 /B4/BT/B5/B9/B4/BU/B5/BA /CC/CW/CX/D7 /CX/D2 /D8/CT/D6\r /CW/CP/D2/CV/CT /D4/CW/CT/D2/D3/D1/CT/D2/D3/D2/CX/D7 /D8/CW/CT /D7/CP/D1/CT /CP/D7 /D3/CQ/D7/CT/D6/DA /CT/CS /CU/D3/D6 /CV/D6/CP/D4/CW/CT/D2/CT /CP/D2/CS /CX/D7 /D2/D3/D8 /DD /CT/D8 /D9/D2/B9/CS/CT/D6/D7/D8/D3 /D3 /CS/BA/C1/D2 /CV/CT/D2/CT/D6/CP/D0/B8 /D8/CW/CT /CD/B4/BD/B5/B9/CV/D6/CP/D4/CW/CT/D2/CT /D6/CT/D7/D9/D0/D8/D7 \r/D3/CX/D2\r/CX/CS/CT /D2 /D9/D1/CT/D6/CX/B9\r/CP/D0/D0/DD /DB/CX/D8/CW /D8/CW/D3/D7/CT /CU/D3/D6 /CV/D6/CP/D4/CW/CT/D2/CT /DB/CW/CT/D2 /D3/D2/CT /D8/CP/CZ /CT/D7 /CX/D2 /D8/D3 /CP\r\r/D3/D9/D2 /D8/D8/CW/CT /CB/CD/B4/BE/B5 /D7/DD/D1/D1/CT/D8/D6/DD /CU/D3/D6 /D8/CW/CT /DA /CP/D0/D0/CT/DD /D4/D7/CT/D9/CS/D3/D7/D4/CX/D2/BA /CC/CW/CX/D7 /CX/D2/B9/CS/CX\r/CP/D8/CT/D7 /D8/CW/CP/D8 /D3/D2/CT /D1/CP /DD /D9/D7/CT /D8/CW/CT /CD/B4/BD/B5/B9/D1/D3 /CS/CT/D0 /CX/D2/D7/D8/CT/CP/CS /D3/CU /D8/CW/CT/D1/D3/D6/CT \r/D3/D1/D4/D0/CT/DC /CB/CD/B4/BE/B5/B9/D7/DD/D1/D1/CT/D8/D6/CX\r /D3/D2/CT /CU/D3/D6 /D8/CW/CT /CS/CX/D7\r/D9/D7/D7/CX/D3/D2 /D3/CU/D8/CW/CT /CS/CT/D2/D7/CX/D8 /DD /D4/CP/D8/D8/CT/D6/D2/D7/B8 /CW/CT/D2\r/CT /D7/CX/D1/D4/D0/CX/CU/DD/CX/D2/CV /CU/D9/D6/D8/CW/CT/D6 \r/CP/D0\r/D9/D0/CP/B9/D8/CX/D3/D2/D7 /D3/D2 /CV/D6/CP/D4/CW/CT/D2/CT/BA /C5/D3/D6/CT/D3 /DA /CT/D6/B8 /CX/D8 /CX/D2/CS/CX\r/CP/D8/CT/D7 /D8/CW/CP/D8 /CX/D2 /D8/CW/CT/CT/D0/CT\r/D8/D6/D3/D2/B9\r/D6/DD/D7/D8/CP/D0 /D4/CW/CP/D7/CT/D7 \r/D3/D2/D7/CX/CS/CT/D6/CT/CS /CP/CQ /D3 /DA /CT /D8/CW/CT /DA /CP/D0/D0/CT/DD /CS/CT/B9/CV/D6/CT/CT /D3/CU /CU/D6/CT/CT/CS/D3/D1 /CX/D7 /CU/D9/D0/D0/DD /D4 /D3/D0/CP/D6/CX/DE/CT/CS/BA/B4/BT/B5/BD /BE /BF /BG /BH /BI /BJ\n/B4/BU/B5/BD /BE /BF /BG /BH /BI /BJ\n/B4/BV/B5/BD /BE /BF /BG /BH /BI /BJ\n/BY/C1/BZ/BA /BJ/BM /B4/BV/D3/D0/D3/D6 /D3/D2/D0/CX/D2/CT/B5 /C4/BW/C7/CBA( /D6,ω) /CU/D3/D6 /CD/B4/BD/B5/B9/CV/D6/CP/D4/CW/CT/D2/CT/BA /CC/CW/CT/D2 /D9/D1 /CQ /CT/D6 /D3/CU /CT/DC/D8/D6/CP\r/D8/CT/CS /BW/C7/CB /D4 /CT/CP/CZ/D7Np= 7 /BA/C1/CE/BA /BV/C7/C6/BV/C4/CD/CB/C1/C7/C6/CB/CC/CW/CT /CP/CX/D1 /D3/CU /D8/CW/CX/D7 /DB /D3/D6/CZ /CX/D7 /D8/D3 /D7/CW/D3 /DB /CW/D3 /DB /CP /CW/CX/CV/CW/B9/AS/CT/D0/CS/CT/D0/CT\r/D8/D6/D3/D2/B9/D7/D3/D0/CX/CS /D4/CW/CP/D7/CT /CX/D2 /D8/CW/CT /BE/BW/BX/BZ /D1/CP /DD /CQ /CT /CS/CT/D8/CT\r/D8/CT/CS /CQ /DD/D3/D4/D8/CX\r/CP/D0 /D1/CT/CP/D2/D7 /CX/D2 /CV/D6/CP/D4/CW/CT/D2/CT/BA /CF /CT /CW/CP /DA /CT \r/CP/D0\r/D9/D0/CP/D8/CT/CS /D8/CW/CT /BW/C7/CB/CP/D2/CS /D8/CW/CT /C4/BW/C7/CB /D3/CU /CT/D0/CT\r/D8/D6/D3/D2/B9/D7/D3/D0/CX/CS /D4/CW/CP/D7/CT/D7 /CX/D2 /D8/CW/CT /C0/CP/D6/D8/D6/CT/CT/B9/BY /D3 \r /CZ /CP/D4/D4/D6/D3 /DC/CX/D1/CP/D8/CX/D3/D2 /CX/D2 /D8/CW/CTN= 2 /C4/C4/BA/CF /CT /D7/CW/D3 /DB /D8/CW/CP/D8 /D8/CW/CT /D2 /D9/D1 /CQ /CT/D6 /D3/CU /D0/D3 /DB/B9/CT/D2/CT/D6/CV/DD /BW/C7/CB /D4 /CT/CP/CZ/D7 /CX/D2\n/CV/D6/CP/D4/CW/CT/D2/CT /CX/D7 /CV/CX/DA /CT/D2 /CQ /DD /D8/CW/CT /D2 /D9/D1 /CQ /CT/D6 /D3/CU /CT/D0/CT\r/D8/D6/D3/D2/D7 /D4 /CT/D6 /D7/CX/D8/CTMe\n/BA/CC/CW/CX/D7 /D6/CT/D7/D9/D0/D8 /CX/D7 /D7/CX/D1/CX/D0/CP/D6 /D8/D3 /D8/CW/CT /D4/D6/CT/DA/CX/D3/D9/D7 /BW/C7/CB \r/CP/D0\r/D9/D0/CP/D8/CX/D3/D2 /CX/D2/D8/CW/CT /C0/CP/D6/D8/D6/CT/CT/B9/BY /D3 \r /CZ /CP/D4/D4/D6/D3 /DC/CX/D1/CP/D8/CX/D3/D2 /CU/D3/D6 /BZ/CP/BT/D7/BA\n/BK/CF /CT /CU/D3/D9/D2/CS /D8/CW/CP/D8 /D8/CW/CT /D6/CT/D7\r/CP/D0/CT/CS /C4/BW/C7/CB /CX/D7 /CX/CS/CT/D2 /D8/CX\r/CP/D0 /CU/D3/D6 /CS/CX/AR/CT/D6/B9/CT/D2 /D8 /AS/D0/D0/CX/D2/CV /CU/CP\r/D8/D3/D6/D7νN\n/B8 /CP/D7 /D0/D3/D2/CV /CP/D7 /D8/CW/CT /D6/CP/D8/CX/D3νN/Me\n/B8 /DB/CW/CX\r /CW/CS/CT/D8/CT/D6/D1/CX/D2/CT/D7 /D8/CW/CT /D0/CP/D8/D8/CX\r/CT /D7/D4/CP\r/CX/D2/CV /D3/CU /D8/CW/CTMe\n/B9/CT/D0/CT\r/D8/D6/D3/D2 /CQ/D9/CQ/CQ/D0/CT\r/D6/DD/D7/D8/CP/D0/B8 /CX/D7 /CZ /CT/D4/D8 /AS/DC/CT/CS/B8 /CP/D2/CS /D8/CW/CT /C4/BW/C7/CB /CU/D6/CT/D5/D9/CT/D2\r/DD /CX/D7 /D8/CP/CZ /CT/D2/CP/D8 /D8/CW/CT /BW/C7/CB /D4 /CT/CP/CZ /DB/CX/D8/CW /D8/CW/CT /D7/CP/D1/CT /CX/D2/CS/CT/DC /B4/CU/D3/D6 /D8/CW/CT /AS/D6/D7/D8 /CU/D3/D9/D6/CX/D2/CS/CX\r/CT/D7/B5/BA /C1/D2 /D4/CP/D6/D8/CX\r/D9/D0/CP/D6/B8 /D8/CW/CX/D7 /D6/CT/D7/D9/D0/D8 /DD/CX/CT/D0/CS/D7 /CP/D2 /D9/D2/CT/DC/D4 /CT\r/D8/CT/CS\r/D3/D2\r/D0/D9/D7/CX/D3/D2 /D8/CW/CP/D8/B8 /CT/BA/CV/BA /CQ /DD /AS/DC/CX/D2/CV /D8/CW/CT /AS/D0/D0/CX/D2/CV /CU/CP\r/D8/D3/D6νN= 0.14/CX/D2 /D8/CW/CTN= 2 /C4/C4/B8 /CP/D2/CS /D9/D7/CX/D2/CV /CB/CC/C5/B8 /D3/D2/CT \r/D3/D9/D0/CS /D3/CQ/D7/CT/D6/DA /CT /CX/D2/D8/CW/CT /C4/BW/C7/CB /D8/CW/CT /DB/CW/D3/D0/CT /D7/D9\r\r/CT/D7/D7/CX/D3/D2 /D3/CU /CT/D0/CT\r/D8/D6/D3/D2/B9\r/D6/DD/D7/D8/CP/D0 /CS/CT/D2/B9/D7/CX/D8 /DD /D4/CP/D8/D8/CT/D6/D2/D7 /DB/CX/D8/CWMe= 1,2,3 /CQ /DD /AS/DC/CX/D2/CV /D8/CW/CT /CP/D4/D4/D0/CX/CT/CS/CB/CC/C5 /DA /D3/D0/D8/CP/CV/CT /CP/D8 /D8/CW/CT \r/D3/D2/D7/CT\r/D9/D8/CX/DA /CT /AS/D6/D7/D8 /D8/CW/D6/CT/CT /D7/CX/D2/CV/D0/CT/B9/D4/CP/D6/D8/CX\r/D0/CT/CT/DC\r/CX/D8/CP/D8/CX/D3/D2 /CT/D2/CT/D6/CV/CX/CT/D7/B8 /CP/D2/CS /D7/D9/D1/D1/CX/D2/CV /D9/D4 /D8/CW/CT /C4/BW/C7/CB /D8/D3 /D3/CQ/D8/CP/CX/D2/D8/CW/CT /D6/CT/D7/D9/D1/D1/CT/CS /C4/BW/C7/CB˜A( /D6,ω) /BA/CF /CT /CQ /CT/D0/CX/CT/DA /CT /D8/CW/CP/D8 /D8/CW/CX/D7 /C4/BW/C7/CB \r/D3/D6/D6/CT/D7/D4 /D3/D2/CS/CT/D2\r/CT /CW/D3/D0/CS/D7 /D8/D6/D9/CT/CU/D3/D6 /CP/D0/D0 /D7/CX/D2/CV/D0/CT/B9/D4/CP/D6/D8/CX\r/D0/CT /CT/DC\r/CX/D8/CP/D8/CX/D3/D2/D7 /D6/CT/D7/D3/D0/DA /CT/CS /CP/D7 /CX/D2/CS/CX/DA/CX/CS/D9/CP/D0/BW/C7/CB /D4 /CT/CP/CZ/D7 /D7/D3 /CU/CP/D6 /B4/CP\r\r/D3/D9/D2 /D8/CX/D2/CV /CU/D3/D6 /CX/D2 /D8/CT/D6\r /CW/CP/D2/CV/CX/D2/CV /D3/CU /D8/CW/CT/D0/CP/D7/D8 /D8 /DB /D3 /D4 /CT/CP/CZ/D7 /CX/D2 /D8/CW/CTMe= 2,3 \r/CP/D7/CT/D7/B5 /CP/D2/CS /CU/D3/D6 /CP/D0/D0 /C4/C4/B3/D7/B4/D7/CX/D1/CX/D0/CP/D6 \r/D3/D2\r/D0/D9/D7/CX/D3/D2/D7 /CU/D3/D0/D0/D3 /DB /CU/D6/D3/D1 /D3/D9/D6 \r/CP/D0\r/D9/D0/CP/D8/CX/D3/D2/D7 /CX/D2 /D8/CW/CT/C4/C4/D7N= 1,3,4 /B5/BA/CF /CT /CP/D0/D7/D3 /D3/CQ/D8/CP/CX/D2/CT/CS /D8/CW/CT /D7/CP/D1/CT /C4/BW/C7/CB \r/D3/D6/D6/CT/D7/D4 /D3/D2/CS/CT/D2\r/CT /CU/D3/D6/D3/D8/CW/CT/D6 /D1/D3 /CS/CT/D0/D7 /D3/CU /D8/CW/CT /BE/BW/BX/BZ/BM /B4/CX/B5 /CX/D2 /CP /D7/CX/D2/CV/D0/CT/B9/D0/CP /DD /CT/D6 /BZ/CP/BT/D7/CW/CT/D8/CT/D6/D3/D7/D8/D6/D9\r/D8/D9/D6/CT/BN /B4/CX/CX/B5 /CD/B4/BD/B5/B9/CV/D6/CP/D4/CW/CT/D2/CT/BN /B4/CX/CX/CX/B5 /CQ/CX/D0/CP /DD /CT/D6/BA /CC/CW/CX/D7/CX/D1/D4/D0/CX/CT/D7 /D8/CW/CP/D8 /D8/CW/CT /D3/CQ/D7/CT/D6/DA /CT/CS /C4/BW/C7/CBνN/Me\n/D7\r/CP/D0/CX/D2/CV /CX/D7 /CX/D2/CS/CT/B9/D4 /CT/D2/CS/CT/D2 /D8 /D3/CU /D8/CW/CT /D9/D2/CS/CT/D6/D0/DD/CX/D2/CV /CX/D2 /D8/CT/D6/CP\r/D8/CX/D3/D2 /D4 /D3/D8/CT/D2 /D8/CX/CP/D0 /CP/D2/CS /D8/CW/CT/D2 /D9/D1 /CQ /CT/D6 /D3/CU /CX/D2/D2/CT/D6 /CS/CX/D7\r/D6/CT/D8/CT /CS/CT/CV/D6/CT/CT/D7 /D3/CU /CU/D6/CT/CT/CS/D3/D1/BA /CC/CW/CT /CU/CP\r/D8/D8/CW/CP/D8 /D8/CW/CT /CD/B4/BD/B5/B9/CV/D6/CP/D4/CW/CT/D2/CT /D6/CT/D7/D9/D0/D8/D7 \r/D3/CX/D2\r/CX/CS/CT /D2 /D9/D1/CT/D6/CX\r/CP/D0/D0/DD /DB/CX/D8/CW/D8/CW/D3/D7/CT /CU/D3/D6 /CV/D6/CP/D4/CW/CT/D2/CT /CX/D2/CS/CX\r/CP/D8/CT/D7 /D8/CW/CP/D8 /D8/CW/CT /CT/D0/CT\r/D8/D6/D3/D2 \r/D6/DD/D7/D8/CP/D0/D7\r/D3/D2/D7/CX/CS/CT/D6/CT/CS /CW/CT/D6/CT /CP/D6/CT \r/D3/D1/D4/D0/CT/D8/CT/D0/DD /DA /CP/D0/D0/CT/DD/B9/D4/D7/CT/D9/CS/D3/D7/D4/CX/D2 /D4 /D3/D0/CP/D6/B9/CX/DE/CT/CS/BA/BT \r /CZ/D2/D3 /DB/D0/CT/CS/CV/D1/CT/D2 /D8/D7/CF /CT /D8/CW/CP/D2/CZ /C8 /BA /C4/CT/CS/CT/D6/CT/D6 /CU/D3/D6 /CU/D6/D9/CX/D8/CU/D9/D0 /CS/CX/D7\r/D9/D7/D7/CX/D3/D2/D7/BA /C7/BA /C8 /BA /CP\r/B9/CZ/D2/D3 /DB/D0/CT/CS/CV/CT/D7 /AS/D2/CP/D2\r/CX/CP/D0 /D7/D9/D4/D4 /D3/D6/D8 /CU/D6/D3/D1 /D8/CW/CT /BX/D9/D6/D3/D4 /CT/CP/D2 /BV/D3/D1/B9/D1/CX/D7/D7/CX/D3/D2 /D8/CW/D6/D3/D9/CV/CW /D8/CW/CT /C5/CP/D6/CX/CT/B9/BV/D9/D6/CX/CT /BY /D3/D9/D2/CS/CP/D8/CX/D3/D2 \r/D3/D2 /D8/D6/CP\r/D8/C5/BX/CB/CC /BV/CC /BE/BC/BC/BG/B9/BH/BD/B9/BG/BF/BC/BJ/B8 /CP/D2/CS /CU/D6/D3/D1 /D8/CW/CT /BZ/CP/D8/CT/D7 /BV/CP/D1 /CQ/D6/CX/CS/CV/CT/CB\r /CW/D3/D0/CP/D6/D7/CW/CX/D4 /CC /D6/D9/D7/D8/BA /CC/CW/CT /DB /D3/D6/CZ /D3/CU /BV/BA/C5/BA/CB/BA /DB /CP/D7 /D4/CP/D6/D8/CX/CP/D0/D0/DD /D7/D9/D4/B9/D4 /D3/D6/D8/CT/CS /CQ /DD /D8/CW/CT /C6/CT/D8/CW/CT/D6/D0/CP/D2/CS/D7 /C7/D6/CV/CP/D2/CX/DE/CP/D8/CX/D3/D2 /CU/D3/D6 /CB\r/CX/CT/D2 /D8/CX/AS\r /CA/CT/B9/D7/CT/CP/D6\r /CW /B4/C6/CF /C7/B5/BA\n∗/BX/D1/CP/CX/D0/D7/BM /CP/D4 /D3/D4/D0\b/D9/CZ/D6/BA/D2/CT/D8/B8 /D3/D4/BE/BE/BI\b\r/CP/D1/BA/CP\r/BA/D9/CZ/BD/BX/BA /C8 /BA /CF/CX/CV/D2/CT/D6/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /BD/BC/BC/BE /B8 /BG/BI /B4/BD/BL/BF/BG/B5/BA/BE/C0/BA /BY /D9/CZ/D9/DD /CP/D1/CP/B8 /C8 /BA /C5/BA /C8/D0/CP/D8/DE/D1/CP/D2/B8 /CP/D2/CS /C8 /BA /CF/BA /BT/D2/CS/CT/D6/D7/D3/D2/B8 /C8/CW /DD/D7/BA/CA/CT/DA/BA /BU /BD/BL /B8 /BH/BE/BD/BD /B4/BD/BL/BJ/BL/B5/BA/BF/BW/BA /CH /D3/D7/CW/CX/D3/CZ /CP /CP/D2/CS /C0/BA /BY /D9/CZ/D9/DD /CP/D1/CP/B8 /C2/BA /C8/CW /DD/D7/BA /CB/D3 \r/BA /C2/CP/D4/CP/D2 /BG/BJ /B8/BF/BL/BG /B4/BD/BL/BJ/BL/B5/BA/BG/BW/BA /CH /D3/D7/CW/CX/D3/CZ /CP /CP/D2/CS /C8 /BA /BT/BA /C4/CT/CT/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /BU /BE/BJ /B8 /BG/BL/BK/BI /B4/BD/BL/BK/BF/B5/BA/BH/C0/BA /BY /D9/CZ/D9/DD /CP/D1/CP /CP/D2/CS /C8 /BA /BT/BA /C4/CT/CT/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /BU /BD/BJ /B8 /BH/BF/BH /B4/BD/BL/BJ/BJ/B5/BN/BD/BK /B8 /BI/BE/BG/BH /B4/BD/BL/BJ/BK/B5/BA/BI/BT/BA /BT/BA /C3 /D3/D9/D0/CP/CZ /D3 /DA/B8 /C5/BA /BY /D3/CV/D0/CT/D6/B8 /CP/D2/CS /BU/BA /C1/BA /CB/CW/CZ/D0/D3 /DA/D7/CZ/CX/CX/B8 /C8/CW /DD/D7/BA\n/CA/CT/DA/BA /C4/CT/D8/D8/BA /BJ/BI /B8 /BG/BL/BL /B4/BD/BL/BL/BI/B5/BN /C5/BA /C5/BA /BY /D3/CV/D0/CT/D6/B8 /BT/BA /BT/BA /C3 /D3/D9/D0/CP/CZ /D3 /DA/B8/CP/D2/CS /BU/BA /C1/BA /CB/CW/CZ/D0/D3 /DA/D7/CZ/CX/CX/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /BU /BH/BG /B8 /BD/BK/BH/BF /B4/BD/BL/BL/BI/B5/BN /CA/BA/C5/D3 /CT/D7/D7/D2/CT/D6 /CP/D2/CS /C2/BA /CC/BA /BV/CW/CP/D0/CZ /CT/D6/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /BU /BH/BG /B8 /BH/BC/BC/BI /B4/BD/BL/BL/BI/B5/BA/BJ/C5/BA /C7/BA /BZ/D3 /CT/D6/CQ/CX/CV/B8 /C8 /BA /C4/CT/CS/CT/D6/CT/D6/B8 /CP/D2/CS /BV/BA /C5/D3/D6/CP/CX/D7 /CB/D1/CX/D8/CW/B8 /C8/CW /DD/D7/BA/CA/CT/DA/BA /BU /BI/BL /B8 /BD/BD/BH/BF/BE/BJ /B4/BE/BC/BC/BG/B5/BA/BK/CA/BA /BV\u001d/D8/GH/B8 /BV/BA /BU/BA /BW/D3/CX/D6/D3/D2/B8 /C2/BA /BU/D3/D9/D6/CP/D7/D7/CP/B8 /CP/D2/CS /C0/BA /BT/BA /BY /CT/D6/D8/CX/CV/B8 /C8/CW /DD/D7/BA/CA/CT/DA/BA /BU /BI/BK /B8 /BD/BH/BH/BF/BE/BJ /B4/BE/BC/BC/BG/B5/BA/BL/BX/BA /BY /D6/CP/CS/CZ/CX/D2 /CP/D2/CS /CB/BA /BT/BA /C3/CX/DA /CT/D0/D7/D3/D2/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /BU /BH/BL /B8 /BK/BC/BI/BH/B4/BD/BL/BL/BL/B5/BN /C5/BA /C5/BA /BY /D3/CV/D0/CT/D6/B8 /CX/D2 /BV/BA /BU/CT/D6/D8/CW/CX/CT/D6/B8 /C4/BA /C8 /BA /C4/GH/DA/DD /B8 /CP/D2/CS/BZ/BA /C5/CP/D6/D8/CX/D2/CT/DE /B4/CT/CS/D7/BA/B5/B8 /C0/CX/CV/CW /C5/CP/CV/D2/CT/D8/CX\r /BY/CX/CT/D0/CS/D7 /B8 /CB/D4/D6/CX/D2/CV/CT/D6 /CE /CT/D6/B9/BK/D0/CP/CV /B4/BE/BC/BC/BD/B5/BA/BD/BC/C0/BA /C4/BA /CB/D8/D3/D6/D1/CT/D6/B8 /BW/BA /BV/BA /CC /D7/D9/CX/B8 /CP/D2/CS /BT/BA /BV/BA /BZ/D3/D7/D7/CP/D6/CS/B8 /CA/CT/DA/BA /C5/D3 /CS/BA/C8/CW /DD/D7/BA /BJ/BD /B8 /CB/BE/BL/BK /B4/BD/BL/BL/BL/B5/BA/BD/BD/C2/BA /C8 /BA /BX/CX/D7/CT/D2/D7/D8/CT/CX/D2/B8 /C3/BA /BU/BA /BV/D3 /D3/D4 /CT/D6/B8 /C4/BA /C6/BA /C8/CU/CT/CX/AR/CT/D6/B8 /CP/D2/CS /C3/BA /CF/BA/CF /CT/D7/D8/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /C4/CT/D8/D8/BA /BK/BK /B8 /BC/BJ/BI/BK/BC/BD /B4/BE/BC/BC/BE/B5/BA/BD/BE/C3/BA /BU/BA /BV/D3 /D3/D4 /CT/D6/B8 /C5/BA /C8 /BA /C4/CX/D0/D0/DD /B8 /CP/D2/CS /C2/BA /C8 /BA /BX/CX/D7/CT/D2/D7/D8/CT/CX/D2/B8 /C4/BA /C6/BA /C8/CU/CT/CX/CU/B9/CU/CT/D6 /CP/D2/CS /C3/BA /CF/BA /CF /CT/D7/D8/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /BU /BI/BC /B8 /CA/BD/BD/BE/BK/BH /B4/BD/BL/BL/BL/B5/BA/BD/BF/C5/BA /C7/BA /BZ/D3 /CT/D6/CQ/CX/CV/B8 /C8 /BA /C4/CT/CS/CT/D6/CT/D6/B8 /CP/D2/CS /BV/BA /C5/D3/D6/CP/CX/D7 /CB/D1/CX/D8/CW/B8 /C8/CW /DD/D7/BA/CA/CT/DA/BA /BU /BI/BK /B8 /BE/BG/BD/BF/BC/BE/B4/CA/B5 /B4/BE/BC/BC/BF/B5/BA/BD/BG/C0/BA /BT/BA /BY /CT/D6/D8/CX/CV/B8 Ꜽ/C8/D6/D3/D4 /CT/D6/D8/CX/CT/D7 /D3/CU /D8/CW/CT /BX/D0/CT\r/D8/D6/D3/D2 /CB/D3/D0/CX/CSꜼ/B8 /CX/D2 Ꜽ/C8 /CT/D6/B9/D7/D4 /CT\r/D8/CX/DA /CT/D7 /CX/D2 /C9/D9/CP/D2 /D8/D9/D1 /C0/CP/D0/D0 /BX/AR/CT\r/D8/D7Ꜽ/B8 /CT/CS/D7/BA /CB/BA /BW/CP/D7 /CB/CP/D6/D1/CP /CP/D2/CS/BT/BA /C8/CX/D2\r/DE/D9/CZ/B8 /B4/CF/CX/D0/CT/DD /B8 /C6/CT/DB /CH /D3/D6/CZ/B8 /BD/BL/BL/BJ/B5/BA/BD/BH/CA/BA /C5/BA /C4/CT/DB/CX/D7/B8 /C8 /BA /BW/BA /CH /CT/B8 /C4/BA /CF/BA /BX/D2/CV/CT/D0/B8 /BW/BA /BV/BA /CC /D7/D9/CX/B8 /C4/BA /C6/BA/C8/CU/CT/CX/AR/CT/D6/B8 /CP/D2/CS /C3/BA /CF/BA /CF /CT/D7/D8/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /C4/CT/D8/D8/BA /BK/BL /B8 /BD/BF/BI/BK/BC/BG/B4/BE/BC/BC/BE/B5/BA/BD/BI/CA/BA /C5/BA /C4/CT/DB/CX/D7/B8 /CH/BA /C8 /BA /BV/CW/CT/D2/B8 /C4/BA /CF/BA /BX/D2/CV/CT/D0/B8 /BW/BA /BV/BA /CC /D7/D9/CX/B8 /C8 /BA /BW/BA/CH /CT/B8 /CP/D2/CS /C4/BA /C6/BA /C8/CU/CT/CX/AR/CT/D6/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /C4/CT/D8/D8/BA /BL/BF /B8 /BD/BJ/BI/BK/BC/BK /B4/BE/BC/BC/BG/B5/BA/BD/BJ/CA/BA /C5/BA /C4/CT/DB/CX/D7/B8 /CH/BA /C8 /BA /BV/CW/CT/D2/B8 /C4/BA /CF/BA /BX/D2/CV/CT/D0/B8 /BW/BA /BV/BA /CC /D7/D9/CX/B8 /C4/BA /C6/BA/C8/CU/CT/CX/AR/CT/D6/B8 /CP/D2/CS /C3/BA /CF/BA /CF /CT/D7/D8/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /BU /BJ/BD /B8 /BC/BK/BD/BF/BC/BD /B4/BE/BC/BC/BH/B5/BA/BD/BK/CH/BA /C8 /BA /BV/CW/CT/D2/B8 /CA/BA /C5/BA /C4/CT/DB/CX/D7/B8 /C4/BA /CF/BA /BX/D2/CV/CT/D0/B8 /BW/BA /BV/BA /CC /D7/D9/CX/B8 /C8 /BA /BW/BA/CH /CT/B8 /CI/BA /C0/BA /CF /CP/D2/CV/B8 /C4/BA /C6/BA /C8/CU/CT/CX/AR/CT/D6/B8 /CP/D2/CS /C3/BA /CF/BA /CF /CT/D7/D8/B8 /C8/CW /DD/D7/BA/CA/CT/DA/BA /C4/CT/D8/D8/BA /BL/BF /B8 /BE/BC/BI/BK/BC/BH /B4/BE/BC/BC/BG/B5/BA/BD/BL/CH/BA /C8 /BA /BV/CW/CT/D2/B8 /BZ/BA /CB/CP/D1 /CQ/CP/D2/CS/CP/D1 /D9/D6/D8/CW /DD /B8 /CI/BA /C0/BA /CF /CP/D2/CV/B8 /CA/BA /C5/BA/C4/CT/DB/CX/D7/B8 /C4/BA /CF/BA /BX/D2/CV/CT/D0/B8 /BW/BA /BV/BA /CC /D7/D9/CX/B8 /C8 /BA /BW/BA /CH /CT/B8 /C4/BA /C6/BA /C8/CU/CT/CX/AR/CT/D6/CP/D2/CS /C3/BA /CF/BA /CF /CT/D7/D8/B8 /C6/CP/D8/D9/D6/CT /C8/CW /DD/D7/CX\r/D7 /BE /B8 /BG/BH/BE /B4/BE/BC/BC/BI/B5/BA/BE/BC/G0/BA /BY/CX/D7\r /CW/CT/D6/B8 /C5/BA /C3/D9/CV/D0/CT/D6/B8 /C1/BA /C5/CP/CV/CV/CX/D3/B9/BT/D4/D6/CX/D0/CT/B8 /CP/D2/CS /BV/BA /BU/CT/D6/D8/CW/D3 /CS/B8/CA/CT/DA/BA /C5/D3 /CS/BA /C8/CW /DD/D7/BA /BJ/BL /B8 /BF/BH/BF /B4/BE/BC/BC/BJ/B5/BA/BE/BD/BT/BA /C3/BA /BZ/CT/CX/D1 /CP/D2/CS /C3/BA /CB/BA /C6/D3 /DA /D3/D7/CT/D0/D3 /DA/B8 /C6/CP/D8/D9/D6/CT /C5/CP/D8/CT/D6/CX/CP/D0/D7 /BI /B8 /BD/BK/BF/B4/BE/BC/BC/BJ/B5/BA/BE/BE/BY /D3/D6 /CP /D6/CT/DA/CX/CT/DB/B8 /D7/CT/CT /BT/BA /C0/BA /BV/CP/D7/D8/D6/D3 /C6/CT/D8/D3/B8 /C6/BA /C5/BA /CA/BA /C8 /CT/D6/CT/D7/B8 /C3/BA/CB/BA /C6/D3 /DA /D3/D7/CT/D0/D3 /DA/B8 /CP/D2/CS /BT/BA /C3/BA /BZ/CT/CX/D1/B8 /CA/CT/DA/BA /C5/D3 /CS/BA /C8/CW /DD/D7/BA /BK/BD /B8 /BD/BC/BL/B4/BE/BC/BC/BL/B5/BA/BE/BF/C3/BA /CB/BA /C6/D3 /DA /D3/D7/CT/D0/D3 /DA/B8 /BT/BA /C3/BA /BZ/CT/CX/D1/B8 /CB/BA /CE/BA /C5/D3/D6/D3/DE/D3 /DA/B8 /BW/BA /C2/CX/CP/D2/CV/B8 /C5/BA/C1/BA /C3/CP/D8/D7/D2/CT/D0/D7/D3/D2/B8 /C1/BA /CE/BA /BZ/D6/CX/CV/D3/D6/CX/CT/DA /CP/B8 /CB/BA /CE/BA /BW/D9/CQ /D3/D2/D3/D7 /CP/D2/CS /BT/BA /BT/BA/BY/CX/D6/D7/D3 /DA/B8 /C6/CP/D8/D9/D6/CT /B4/C4/D3/D2/CS/D3/D2/B5 /BG/BF/BK /B8 /BD/BL/BJ /B4/BE/BC/BC/BH/B5/BA/BE/BG/CH/BA /CI/CW/CP/D2/CV/B8 /CH/BA/B9/CF/BA /CC /CP/D2/B8 /C0/BA /C4/BA /CB/D8/D3/D6/D1/CT/D6/B8 /CP/D2/CS /C8 /BA /C3/CX/D1/B8 /C6/CP/D8/D9/D6/CT/BG/BF/BK /BE/BC/BD/B8 /B4/BE/BC/BC/BH/B5/BA/BE/BH/C3/BA /CB/BA /C6/D3 /DA /D3/D7/CT/D0/D3 /DA/B8 /CI/BA /C2/CX/CP/D2/CV/B8 /CH/BA /CI/CW/CP/D2/CV/B8 /CB/BA /CE/BA /C5/D3/D6/D3/DE/D3 /DA/B8 /C0/BA /C4/BA/CB/D8/D3/D6/D1/CT/D6/B8 /CD/BA /CI/CT/CX/D8/D0/CT/D6/B8 /C2/BA /BV/BA /C5/CP/CP/D2/B8 /BZ/BA /CB/BA /BU/D3 /CT/CQ/CX/D2/CV/CT/D6/B8 /C8 /BA /C3/CX/D1/B8/BT/BA /C3/BA /BZ/CT/CX/D1/B8 /CB\r/CX/CT/D2\r/CT /BF/BD/BH /B8 /BD/BF/BJ/BL /B4/BE/BC/BC/BJ/B5/BA/BE/BI/C5/BA /C7/BA /BZ/D3 /CT/D6/CQ/CX/CV/B8 /CA/BA /C5/D3 /CT/D7/D7/D2/CT/D6/B8 /CP/D2/CS /BU/BA /BW/D3/D9/GF/D3/D8/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA\n/BU /BJ/BG /B8 /BD/BI/BD/BG/BC/BJ/B4/CA/B5 /B4/BE/BC/BC/BI/B5/BA/BE/BJ/CE/BA /C5/BA /BT/D4/CP/D0/CZ /D3 /DA /CP/D2/CS /CC/BA /BV/CW/CP/CZ/D6/CP/CQ /D3/D6/D8 /DD /B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /C4/CT/D8/D8/BA /BL/BJ /B8/BD/BE/BI/BK/BC/BD /B4/BE/BC/BC/BI/B5/BA/BE/BK/BV/BA /CC/GU/CZ /CT/B8 /C8 /BA /BX/BA /C4/CP/D1/D1/CT/D6/D8/B8 /CE/BA /C0/BA /BV/D6/CT/D7/D4/CX/B8 /CP/D2/CS /C2/BA /C3/BA /C2/CP/CX/D2/B8/C8/CW /DD/D7/BA /CA/CT/DA/BA /BU /BJ/BG /B8 /BE/BF/BH/BG/BD/BJ /B4/BE/BC/BC/BI/B5/BN /BV/BA /CC/GU/CZ /CT /CP/D2/CS /C2/BA /C3/BA /C2/CP/CX/D2/B8/C8/CW /DD/D7/BA /CA/CT/DA/BA /BU /BJ/BH /B8 /BE/BG/BH/BG/BG/BC /B4/BE/BC/BC/BJ/B5/BA/BE/BL/BW/BA /CE/BA /C3/CW /DA /CT/D7/CW\r /CW/CT/D2/CZ /D3/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /BU /BJ/BH /B8 /BD/BH/BF/BG/BC/BH /B4/BE/BC/BC/BJ/B5/BA/BF/BC/C5/BA /C7/BA /BZ/D3 /CT/D6/CQ/CX/CV /CP/D2/CS /C6/BA /CA/CT/CV/D2/CP/D9/D0/D8/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /BU /BJ/BH /B8 /BE/BG/BD/BG/BC/BH/B4/BE/BC/BC/BJ/B5/BN /CA/BA /CS/CT /BZ/CP/CX/D0/B8 /C6/BA /CA/CT/CV/D2/CP/D9/D0/D8/B8 /CP/D2/CS /C5/BA /C7/BA /BZ/D3 /CT/D6/CQ/CX/CV/B8 /C8/CW /DD/D7/BA/CA/CT/DA/BA /BU /BJ/BJ /B8 /BD/BI/BH/BF/BD/BC /B4/BE/BC/BC/BK/B5/BA/BF/BD/BV/BA/B9/C0/BA /CI/CW/CP/D2/CV /CP/D2/CS /CH /D3/CV/CT/D7/CW /C6/BA /C2/D3/CV/D0/CT/CZ /CP/D6/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /BU /BJ/BH /B8/BE/BG/BH/BG/BD/BG /B4/BE/BC/BC/BJ/B5/BA/BF/BE/BV/BA/B9/C0/BA /CI/CW/CP/D2/CV /CP/D2/CS /CH /D3/CV/CT/D7/CW /C6/BA /C2/D3/CV/D0/CT/CZ /CP/D6/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /BU /BJ/BJ /B8/BE/BC/BH/BG/BE/BI /B4/BE/BC/BC/BK/B5/BA/BF/BF/C2/BA /CF /CP/D2/CV/B8 /BT/BA /C1/DD /CT/D2/CV/CP/D6/B8 /C0/BA /BT/BA /BY /CT/D6/D8/CX/CV/B8 /CP/D2/CS /C4/BA /BU/D6/CT/DD /B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA/BU /BJ/BK /B8 /BD/BI/BH/BG/BD/BI /B4/BE/BC/BC/BK/B5/BA/BF/BG/C0/BA /CF /CP/D2/CV/B8 /BW/BA /C6/BA /CB/CW/CT/D2/CV/B8 /C4/BA /CB/CW/CT/D2/CV/B8 /CP/D2/CS /BY/BA /BW/BA /C5/BA /C0/CP/D0/CS/CP/D2/CT/B8/C8/CW /DD/D7/BA /CA/CT/DA/BA /C4/CT/D8/D8/BA /BD/BC/BC /B8 /BD/BD/BI/BK/BC/BE /B4/BE/BC/BC/BK/B5/BA/BF/BH/CA/BA /BV\u001d/D8/GH/B8 /C2/BA/B9/BY/BA /C2/D3/CQ/CX/CS/D3/D2/B8 /CP/D2/CS /C0/BA /BT/BA /BY /CT/D6/D8/CX/CV/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /BU/BJ/BK /B8 /BC/BK/BH/BF/BC/BL /B4/BE/BC/BC/BK/B5/BA/BF/BI/C2/BA /C5/CP/D6/D8/CX/D2/B8 /C6/BA /BT/CZ /CT/D6/D1/CP/D2/B8 /BZ/BA /CD/D0/CQ/D6/CX\r /CW /D8/B8 /CC/BA /C4/D3/CW/D1/CP/D2/D2/B8 /C2/BA /C0/BA/CB/D1/CT/D8/B8 /C3/BA /DA /D3/D2 /C3/D0/CX/D8/DE/CX/D2/CV/B8 /CP/D2/CS /BT/BA /CH /CP\r/D3/CQ /DD /B8 /C6/CP/D8/D9/D6/CT /C8/CW /DD/D7/CX\r/D7 /BG /B8/BD/BG/BG /B4/BE/BC/BC/BK/B5/BA/BF/BJ/C8 /BA /C5/CP/D0/D0/CT/D8/B8 /BY/BA /CE /CP/D6\r /CW/D3/D2/B8 /BV/BA /C6/CP/D9/CS/B8 /C4/BA /C5/CP/CV/CP/D9/CS/B8 /BV/BA /BU/CT/D6/CV/CT/D6/B8/CP/D2/CS /C2/BA/B9/CH/BA /CE /CT/D9/CX/D0/D0/CT/D2/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /BU /BJ/BI /B8 /BC/BG/BD/BG/BC/BF /B4/BE/BC/BC/BJ/B5/BA/BF/BK/BZ/BA /C4/CX/B8 /BT/BA /C4/D9/CX\r/CP/D2/B8 /BX/BA /CH/BA /BT/D2/CS/D6/CT/CX/B8 /CP/D6/CG/CX/DA/BM/BC/BK/BC/BF/BA/BG/BC/BD/BI /BA/BF/BL/C0/BA /C8 /BA /BW/CP/CW/CP/D0/B8 /CH/BA /C6/BA /C2/D3/CV/D0/CT/CZ /CP/D6/B8 /C3/BA /CB/BA /BU/CT/CS/CT/D0/D0/B8 /CP/D2/CS /BT/BA /CE/BA /BU/CP/D0/B9/CP/D8/D7/CZ/DD /B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /BU /BJ/BG /B8 /BE/BF/BF/BG/BC/BH /B4/BE/BC/BC/BI/B5/BA/BG/BC/CA/BA /BV\u001d/D8/GH /CP/D2/CS /BT/BA /C0/BA /C5/CP\r/BW/D3/D2/CP/D0/CS/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /C4/CT/D8/D8/BA /BI/BH /B8 /BE/BI/BI/BE/B4/BD/BL/BL/BC/B5/BN /C8/CW /DD/D7/BA /CA/CT/DA/BA /BU /BG/BG /B8 /BK/BJ/BH/BL /B4/BD/BL/BL/BD/B5/BA/BG/BD/CA/BA /BV\u001d/D8/GH/B8 /C4/BA /BU/D6/CT/DD /B8 /CP/D2/CS /BT/BA /C0/BA /C5/CP\r/BW/D3/D2/CP/D0/CS/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /C4/CT/D8/D8/BA/BG/BI /B8 /BD/BC/BE/BF/BL /B4/BD/BL/BL/BE/B5/BA/BG/BE/C3/BA /C6/D3/D1 /D9/D6/CP /CP/D2/CS /BT/BA /C0/BA /C5/CP\r/BW/D3/D2/CP/D0/CS/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /C4/CT/D8/D8/BA /BL/BI /B8/BE/BH/BI/BI/BC/BE /B4/BE/BC/BC/BI/B5/BA/BG/BF/C2/BA /BT/D0/CX\r/CT/CP /CP/D2/CS /C5/BA /C8 /BA /BT/BA /BY/CX/D7/CW/CT/D6/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /BU /BJ/BG /B8 /BC/BJ/BH/BG/BE/BE/B4/BE/BC/BC/BI/B5/BA/BG/BG/BZ/BA /BW/BA /C5/CP/CW/CP/D2/B8 /C5/CP/D2 /DD/B9/D4/CP/D6/D8/CX\r/D0/CT /D4/CW /DD/D7/CX\r/D7/B8 /C1/D2/CS/CX/CP/D2/CP /CD/D2/CX/DA/BA /B4/BD/BL/BK/BC/B5/BA/BG/BH/CA/BA /BV\u001d/D8/GH /CP/D2/CS /BT/BA /C0/BA /C5/CP\r/BW/D3/D2/CP/D0/CS/B8 /C8/CW /DD/D7/BA /CA/CT/DA/BA /BU /BG/BG /B8 /BK/BJ/BH/BL/B4/BD/BL/BL/BD/B5/BA" }, { "title": "0902.4351v1.Superposition_rule_and_entanglement_in_diagonal_and_probability_representations_of_density_states.pdf", "content": "Superposition rule and entanglement in diagonal\nand probability representations of density states z\nVladimir I. Man'ko,1Giuseppe Marmo2and E. C. George\nSudarshan3\n1P. N. Lebedev Physical Institute, Leninskii Prospect 53, Moscow 119991,\nRussia\n2Dipartimento di Scienze Fisiche, Universit\u0012 a \\Federico II\" di Napoli\nand Istituto Nazionale di Fisica Nucleare, Sezione di Napoli\nComplesso Universitario di M. S. Angelo, Via Cintia, I-80126 Napoli, Italy\n3Department of Physics, University of Texas, Austin, Texas 78712, USA\nE-mail: mmanko@sci.lebedev.ru\nAbstract. The quasidistributions corresponding to the diagonal representation\nof quantum states are discussed within the framework of operator-symbol\nconstruction. The tomographic-probability distribution describing the quantum\nstate in the probability representation of quantum mechanics is reviewed. The\nconnection of the diagonal and probability representations is discussed. The\nsuperposition rule is considered in terms of the density-operator symbols. The\nseparability and entanglement properties of multipartite quantum systems are\nformulated as the properties of the density-operator symbols of the system states.\nPACS numbers: 03.65.-w, 03.65.-Wj\n1. Introduction\nThe pure quantum states are traditionally associated with the wave function [1] or\na vector in the Hilbert space [2]. The mixed quantum states are described by the\ndensity matrix [3] or the density operator [4]. There exist several representations of\nquantum states in terms of the quasidistribution functions like the Wigner function [5]\nand the Husimi{Kano function [6, 7]. The diagonal representation of quantum states\nwas suggested in [8] (see also [9]). It was studied and applied in [10, 11]. In this\nrepresentation, a quantum state is represented in terms of weighted sum of coherent-\nstatejziprojectors. The properties of all the quantum-state representations considered\nare associated with the properties of the density operator which is Hermitian, trace-\nclass nonnegative operator. This means, in particular, that all the eigenvalues of the\ndensity operators must be nonnegative. In the quantum domain, the multipartite\nsystems have a speci\fc property connected with strong correlations of the quantum\nsubsystems. This property provides the entanglement phenomenon [12].\nIn the diagonal representation of the density states, the weight function \u001e(z)\nis an analog of the probability-distribution function in the phase space. For some\nclass of states, this function is identical to the probability-distribution function like in\nzbased on the invited talk presented by one of us (V.I.M.) at the XV Central European Workshop\non Quantum Optics (Belgrade, Serbia, 30 May { 3 June 2008).arXiv:0902.4351v1 [quant-ph] 25 Feb 2009Diagonal and probability representations 2\nclassical statistical mechanics. In [13], the tomographic-probability representation of\nquantum states, where the quantum state is associated with the so-called symplectic\ntomogram, was introduced. The tomogram is a fair probability distribution containing\nthe same information on quantum state that the density operator does (or such its\ncharacteristics as the Wigner or Husimi{Kano functions). The aim of this work is to\n\fnd the explicit formulae realizing the connection of the diagonal and tomographic\nprobability representations. In [14], a review of the star-product-quantization schemes\nwas given in a uni\fed form. According to this scheme, the functions like the Wigner\nfunction, Husimi{Kano function and tomographic-probability-distribution function\nare considered as symbols of the density operators of a corresponding star-product\nscheme. The other goal of our work is to discuss in detail the diagonal representation\nwithin the framework of the star-product scheme along the lines of construction\ngiven in [14] and to \fnd mutual relations between the tomographic-probability\nrepresentation and the diagonal representation in this context. Using formulation\nof the superposition rule in terms of the density operator [15, 16, 17], we consider it\nwithin the framework of the density-state symbols. We focus on the superposition\nrule given in terms of tomograms and in terms of weight functions of the diagonal\nrepresentation where explicit kernels of the corresponding star-products are employed\nto obtain the addition rules for the tomograms and weight functions. We discuss also\nthe formulation of the separability and entanglement properties of composed system\nin the tomographic probability and diagonal representations.\nThe paper is organized as follows.\nIn Section 2, symplectic tomograms and the diagonal representation of quantum\nsates are reviewed. In Section 3, the superposition rule is considered. In Section 4, the\ndiagonal representation and the star-product formalism are compared. In Section 5,\nthe superposition rule for tomograms is presented. In Section 6, the entanglement\nin the tomographic and diagonal representations is studied. Conclusions are given in\nSection 7.\n2. Symplectic tomogram and diagonal representation\nBelow we review the approach where the quantum state associated with tomographic\nsymbol (called symplectic tomogram) of the density operator (density state) ^ \u001areads\n(see, for example, [17])\nw(X;\u0016;\u0017 ) = Tr ^\u001a\u000e(X^1\u0000\u0016^q\u0000\u0017^p) (\u0016h= 1): (1)\nHereX; \u0016; \u00172R, ^qand ^pare the position and momentum operators, respectively.\nFor the pure state, tomogram is expressed in terms of the wave function [18]\nw(X;\u0016;\u0017 ) =1\n2\u0019j\u0017j\f\f\f\fZ\n (y) exp\u0012i\u0016\n2\u0017y2\u0000iXy\n\u0017\u0013\ndy\f\f\f\f2\n: (2)\nThe tomogram is nonnegative normalized probability distribution function of a random\nvariableX, i.e.,\nw(X;\u0016;\u0017 )\u00150;Z\nw(X;\u0016;\u0017 )dX= 1: (3)\nIn the diagonal representation, the density state ^ \u001areads [8]\n^\u001a=Z\n\u001e(z)jzihzjdRez dImz; (4)Diagonal and probability representations 3\nwherejzi=^D(z)j0iis the coherent state ^ ajzi=zjziand the displacement operator\n^D(z) = exp\u0000\nz^ay\u0000z\u0003^a\u0001\nis called Weyl system. Here ^ a= 2\u00001=2(^q+i^p) is the\nboson annihilation operator and zis a complex number. The probability distribution\nw(X;\u0016;\u0017 ) is expressed in terms of the weight function \u001e(z) as follows:\nw(X;\u0016;\u0017 ) =Z\n\u001e(z)hzj\u000e(X^1\u0000\u0016^q\u0000\u0017^p)jzidRez dImz: (5)\nUsing in (1) the Fourier decomposition of delta-function and taking the density state\n^\u001ain form (4), we obtain for tomogram\nw(X;\u0016;\u0017 ) =1\n2\u0019Z\n\u001e(z)hzjeik(X\u0000\u0016^q\u0000\u0017^p)jzidRez dImz; (6)\nwhere the diagonal matrix element of the operator in the integral can be considered\nas the Weyl-system matrix element, i.e.,\nhzje\u0000k(i\u0016^q+i\u0017^p)jzi= exp\u0014\nz\u0003\u000b\u0000\u000b\u0003z\u0000j\u000bj2\n2\u0015\n; (7)\nwith\n\u000b=kp\n2(\u0017\u0000i\u0016): (8)\nEvaluating Gaussian integral (6), we arrive at\nw(X;\u0016;\u0017 ) =1p\n\u0019(\u00162+\u00172)\n\u0002Z\n\u001e(z) exp8\n><\n>:\u0000h\nX\u00002\u00001=2\u0010\nz\u0003(\u0016+i\u0017) +z(\u0016\u0000i\u0017)\u0011i2\n\u00162+\u001729\n>=\n>;dRez dImz: (9)\nThe above formula provides the relation of the weight function of the diagonal\nrepresentation of the density state and symplectic tomogram of the quantum state.\nFor example, the vacuum state j0ih0jhas the weight function\n\u001e0(z) =\u000e(Rez)\u000e(Imz):\nFormula (9) provides tomogram of the ground state\nw0(X;\u0016;\u0017 ) =1p\n\u0019(\u00162+\u00172)exp\u0012\n\u0000X2\n\u00162+\u00172\u0013\n: (10)\nThis expression can be obtained also by means of formula (2) with\n o(y) =\u0019\u00001=4exp\u0000\n\u0000y2=2\u0001\n:\n3. Superposition rule for density operators\nFor two orthogonal pure states j 1iandj 2i, the superposition rule provides the state\nj i=pp1j 1i+ei\u001epp2j 2i; (11)\nwhich can be realized in the nature as a Schr odinger cat state. Here the positive\nnumbersp1andp2satisfy the equality p1+p2= 1 and the phase factor ei\u001edeterminesDiagonal and probability representations 4\nthe interference picture. The density states ^ \u001a1=j 1ih 1jand ^\u001a2=j 2ih 2jprovide\nthe state ^\u001a=j ih j, if one uses the nonlinear addition rule [15]\n^\u001a=p1^\u001a1+p2^\u001a2+pp1p2^\u001a1^P0^\u001a2+ ^\u001a2^P0^\u001a1r\nTr\u0010\n^\u001a1^P0^\u001a2^P0\u0011; (12)\nwhere the operator ^P0is a projector (Tr ^P0= 1) which corresponds to the phase term\nei\u001ein (11).\nThe superposition rule can be formulated for any symbol of pure density states\n^\u001a1, ^\u001a2and ^\u001a.\n4. Diagonal representation and star-product formalism\nThe diagonal representation of density operators can be considered within the\nframework of star-product scheme [14]. Let us construct two families of operators,\nwhich are called dequantizer\n^U(z) =1\n\u00192Z\nexp\u00121\n2juj2+z\u0003u\u0000zu\u0003\u0013\n^D(u)dReu dImu (13)\nand quantizer\n^D(z) =jzihzj; (14)\nwhere ^D(u) is the Weyl system and z=x+iyis a complex number. One can check\nthat\nTrbU(z)bD(z0) =\u000e(x\u0000x0)\u000e(y\u0000y0): (15)\nIn view of this, one can construct the symbol of a density operator ^ \u001ain the diagonal\nrepresentation\n\u001e(z) = TrbU(z)b\u001a=1\n\u00192Z\nexp\u00121\n2juj2+z\u0003u\u0000zu\u0003\u0013\nTrb\u001abD(u)dReu dImu (16)\nand the reconstruction formula for the density operator reads\nb\u001a=Z\n\u001e(z)jzihzjdRez dImz: (17)\nAccording to [19, 20], one can construct dual symbol of the operator b\u001a\n\u001e(d)(z) = Trb\u001ajzihzj=hzjb\u001ajzi (18)\nand dual reconstruction formula\nb\u001a=Z\n\u001e(d)(z)bU(z)dRez dImz\n=1\n\u00192Z\n\u001e(d)(z) exp\u00121\n2juj2+z\u0003u\u0000zu\u0003\u0013\nbD(u)dReu dImu dRez dImz: (19)\nIf in (16) the operator b\u001ais replaced by some operator bA, the corresponding symbol\n\u001eA(z) provides the diagonal representation of the operator. The dual symbol (18)\nprovides the Husimi{Kano function Q(z). The reconstruction formula for the density\nstate in terms of the Husimi{Kano function is just formula (19) with the replacement\n\u001e(d)(z)!Q(z). The duality relation of the diagonal representation of the density\nstateb\u001aand the Husimi{Kano function was discussed in [19, 21].Diagonal and probability representations 5\nUsing the connection of an operator symbol with its dual [14], one has the\nconnection formula\n\u001e(z) =1\n\u00193Z\nQ(z1) exp\u0002\njuj2+ (z\u0003\u0000z\u0003\n1)u\u0000(z\u0000z1)u\u0003\u0003\ndReu dImu: (20)\nThe inverse formula reads\nQ(z) =Z\n\u001e(z)e\u0000jz1\u0000zj2dRez1dImz1: (21)\nAccording to the general formalism [14], the star-product of symbols related to the\ndiagonal representation is determined by the kernel\nK(z1;z2;z) =1\n\u00192Tr\u0014Z\njz1ihz1jjz2ihz2jbD(u) exp\u00121\n2juj2+z\u0003u\u0000zu\u0003\u0013\ndReu dImu\u0015\n;\n(22)\nwhich is generalized function of the form\nK(z1;z2;z) =1\n\u00192Z\nexp\u0010\n\u0000(x2\u0000x1)2\u0000(y2\u0000y1)2+ (x2\u0000x1)a+ (y2\u0000y1)b\n\u0000i(2y+y1+y2)a+i(2x+x1+x2)b\u0011\nda db; (23)\nwhere\nz=x+iy; z 1=x1+iy1; z 2=x2+iy2:\nThe star-product of symbols of arbitrary operators bAandbBin the diagonal\nrepresentation reads\n(\u001eA\u0003\u001eB)(z) =Z\nK(z1;z2;z)\u001eA(z1)\u001eB(z2)dx1dy1dx2dy2: (24)\nFor example, for the vacuum-state projector b\u001a0=j0ih0jwith the weight function {\nsymbol\u001e0(z) =\u000e(z), one has\n(\u001e0\u0003\u001e0)(z) =Z\n\u000e(z1)\u000e(z2)K(z1;z2;z)dx1dy1dx2dy2=\u000e(z); (25)\nwhich is equal to \u001e0(z) and corresponds to the pure-vacuum-state property b\u001a2\n0=b\u001a0.\nTomogram w(X;\u0016;\u0017 ) of the density state b\u001aprovides the following formula for the\ndiagonal representation of the density operator\n\u001e(z) =1\n2\u00192Z\nw(X;\u0016;\u0017 ) exp\u0014\niX+\u00162+\u00172\n4+z(\u0017+i\u0016)p\n2\u0000z\u0003(\u0017\u0000i\u0016)p\n2\u0015\ndX d\u0016 d\u0017:\n(26)\nFor example, the vacuum-state tomogram\nw0(X;\u0016; 0) =1\n\u0019(\u00162+\u00172)exp\u0012\n\u0000x2\n\u00162+\u00172\u0013\nprovides, by means of the above formula, the symbol of the state in the diagonal\nrepresentation, i.e., \u000e(z).Diagonal and probability representations 6\n5. Superposition rule for tomograms\nThe superposition of two pure states with their symbols \u001e1(z) and\u001e2(z) is described\nby the function\n\u001e(z) =p1\u001e1(z) +p2\u001e2(z) +pp1p2(\u001e1\u0003\u001e0\u0003\u001e2)(z) + (\u001e2\u0003\u001e0\u0003\u001e1)(z)qR\n(\u001e1\u0003\u001e0\u0003\u001e2\u0003\u001e0)(z)dxdy: (27)\nThe star-product in (27) is determined by the kernel (23).\nThe result obtained can be repeated also for tomographic symbols of the density\nstates. Thus, the addition rule of two tomographic probabilities of two pure states\nj 1iandj 2ireads\nw(X;\u0016;\u0017 ) =p1w1(X;\u0016;\u0017 ) +p2w2(X;\u0016;\u0017 )\n+pp1p2(w1\u0003w0\u0003w2)(X;\u0016;\u0017 ) + (w2\u0003w0\u0003w1)(X;\u0016;\u0017 )qR\n\u000e(\u0016)\u000e(\u0017)d\u0016d\u0017d\u00160d\u00170R\neiX(w1\u0003w0\u0003w2\u0003w0)(X;\u0016;\u0017 )dX: (28)\nThe kernel of tomographic star-product is given in [19, 20]. The order of integration\nin the denominator term is essential to obtain the correct result. In (27) and (28), \u001e0\nandw0are the corresponding symbols of projector bP0which determines the relative\nphase of statesj 1iandj 2iin their superposition.\n6. Entanglement in the diagonal and tomographic-probability\nrepresentations\nGiven bipartite system of a two-mode \feld.\nThe tomographic probability distribution is determined as follows:\nw(X1;\u00161;\u00171;X2;\u00162;\u00172) = Trh\nb\u001a(1;2)\u000e(X1b1\u0000\u00161bq1\u0000\u00171bp1)\u000e(X2b1\u0000\u00162bq2\u0000\u00172bp2)i\n:\n(29)\nThe density matrix in the diagonal representation is determined by the symbol of the\ndensity state b\u001a(1;2)\n\u001e(z1;z2) =1\n\u00194Z\nexp\u00141\n2\u0000\nju1j2+ju2j2\u0001\n+z\u0003\n1u1\u0000z1u\u0003\n1+z\u0003\n2u2\u0000z2u\u0003\n2\u0015\n\u0002Trb\u001a(1;2)bD1(u1)bD2(u2)dReu1dImu1dReu2dImu2: (30)\nThe stateb\u001a(1;2) is separable, if the density state can be written as a convex sum\nb\u001a(1;2) =X\nkPkb\u001ak(1)\nb\u001ak(2); Pk\u00150;X\nkPk= 1: (31)\nIn view of linearity property of tomographic map, one has the de\fnition of separability\nin terms of the state tomogram, i.e., the state is separable, if\nw(X1;\u00161;\u00171;X2;\u00162;\u00172) =X\nkPkw(1)\nk(X1;\u00161;\u00171)w(2)\nk(X2;\u00162;\u00172): (32)\nTomogram is the joint probability-distribution function of two random variables\nX1;X22R. Thus, the condition of the state separability is formulated as the above\nproperty (32) of the joint probability distribution.\nIf tomogram cannot be written as convex sum (32), the state is entangled. The\nseparability condition can be reformulated, in view of the standard characteristicDiagonal and probability representations 7\nfunction for the tomographic probability distribution (32). In fact, if the characteristic\nfunction can be written as\n\u001f(k1;\u00161;\u00171;k2;\u00162;\u00172) =X\nkPk\u001f(1)\nk(k1;\u00161;\u00171)\u001f(2)\nk(k2;\u00162;\u00172) (33)\nthe state is separable. Here \u001f(1)\nk(k1;\u00161;\u00171) and\u001f(2)\nk(k2;\u00162;\u00172) are the characteristic\nfunctions for tomographic probabilities w(1)\nk(X1;\u00161;\u00171) andw(2)\nk(X2;\u00162;\u00172), respec-\ntively.\nAn analogous de\fnition of the separability and entanglement of the density state\nb\u001a(1;2) can be formulated in the diagonal representation.\nThus the state is separable, if the function which is symbol of the density state\nin the diagonal representation can be written as\n\u001e(z1;z2) =X\nkPk\u001e(1)\nk(z1)\u001e(2)\nk(z2): (34)\nThus we formulated the problem of separability and entanglement in the diagonal\nrepresentation of the density state b\u001a(1;2). One can easily extend the de\fnition of\nseparable and entangled states to multipartite systems in both the tomographic and\ndiagonal representations of density states.\n7. Conclusions\nTo conclude, we resume the main results of this work.\nWe reviewed the diagonal and probability representations of quantum states using\nthe standard star-product scheme. We found mutual relations of the weight function\nof the diagonal representation and the tomographic-probability distribution associated\nwith the quantum state. We obtained the kernel of star-product of operator symbols in\nthe diagonal representation. The duality relation between the diagonal representation\nof the weight function and the Husimi{Kano function was obtained in the explicit\nform. The superposition rule was formulated in both the diagonal representation\nand probability representation of the density states. The problem of separability and\nentanglement was formulated in both the diagonal and probability representations.\nAcknowledgments\nV.I.M. thanks the Russian Foundation for Basic Research for partial support under\nProjects Nos. 07-02-00598 and 08-02-90300 and the Organizers of the XV Central\nEuropean Workshop on Quantum Optics (Belgrade, Serbia, 30 May { 3 June 2008)\nfor kind hospitality.\nReferences\n[1] Schr odinger E 1926 Ann. Phys. (Liepzig) 79489\n[2] Dirac P A M 1058 Principles of Quantum Mechanics 4th ed (London: Oxford University Press)\n[3] Landau L D 1927 Z. Physik 45430\n[4] von Neumann J 1932 Matematische Grundlagen der Quantenmechanyk (Berlin: Springer)\n[5] Wigner E 1932 Phys. Rev. 40749\n[6] Husimi K 1940 Proc. Phys. Math. Soc. Jpn 23264\n[7] Kano Y 1956 J. Math. Phys. 61913\n[8] Sudarshan E C G 1963 Phys. Rev. Lett. 10277\n[9] Glauber R J 1963 Phys. Rev. Lett. 1084Diagonal and probability representations 8\n[10] Klauder J R and Sudarshan E C G 1968 Fundamentals of Quantum Optics (New York:\nBenjamin)\n[11] Mehta C L and Sudarshan E C G 1965 Phys. Rev. 138B274\n[12] Schr odinger E 1935 Naturwissenchaften 23823\n[13] Mancini S, Man'ko V I and Tombesi P 1996 Phys. Lett. A 2131\n[14] Man'ko O V, Man'ko V I and Marmo G 2002 J. Phys. A: Math. Gen. 35699\n[15] Man'ko V I, Marmo G, Sudarshan E C G and Zaccaria F 2002 J. Phys. A: Math. Gen. 357137\n[16] Man'ko V I, Marmo G, Sudarshan E C G and Zaccaria F 2003 J. Russ. Laser Res. 24507\n[17] Man'ko V I, Marmo G, Simoni A, Sudarshan E C G and Ventriglia F 2008 Rep. Math. Phys.\n61337\n[18] Man'ko V I and Mendes R V 1999 Phys. Lett. A 26353\n[19] Man'ko V I, Marmo G and Vitale P 2005 Phys. Lett. A 3341\n[20] Man'ko O V, Man'ko V I, Marmo G and Vitale P 2007 Phys. Lett. A 360522\n[21] Klauder J R and Scagerstam B-S K 2007 J Phys. A: Math. Theor. 402093" }, { "title": "0903.2903v2.Measuring_Qutrit_Qutrit_Entanglement_of_Orbital_Angular_Momentum_States_of_an_Atomic_Ensemble_and_a_Photon.pdf", "content": "arXiv:0903.2903v2 [quant-ph] 12 Sep 2009Measuring Qutrit-Qutrit Entanglement of Orbital Angular M omentum States\nof an Atomic Ensemble and a Photon\nR. Inoue1,∗T. Yonehara1, Y. Miyamoto2, M. Koashi3, and M. Kozuma1,4\n1Department of Physics, Tokyo Institute of Technology,\n2-12-1 O-okayama, Meguro-ku, Tokyo 152-8550, Japan\n2Department of Information and Communication Engineering, The University of Electro-Communications,\n1-5-1 Chofugaoka, Chofu, Tokyo 182-8585, Japan\n3Division of Materials Physics, Graduate School of Engineer ing Science, Osaka University,\n1-3 Machikaneyama, Toyonaka, Osaka 560-8531, Japan and\n4CREST, Japan Science and Technology Agency, 1-9-9 Yaesu, Ch uo-ku, Tokyo 103-0028, Japan\n(Dated: December 6, 2018)\nThree-dimensional entanglement of orbital angular moment um states of an atomic qutrit and a\nsingle photon qutrit has been observed. Their full state was reconstructed using quantum state\ntomography. The fidelity to the maximally entangled state of Schmidt rank 3 exceeds the threshold\n2/3. This result confirms that the density matrix cannot be deco mposed into ensemble of pure\nstates of Schmidt rank 1 or 2. That is, the Schmidt number of th e density matrix must be equal to\nor greater than 3.\nPACS numbers: 03.65.Wj, 03.67.Mn, 32.80.-t, 42.50.Dv\nAn essential requirement towardpractical quantum in-\nformation systems is the capability to generate entangled\nstates between distant sites over quantum networks [1].\nInspired bya schemeto createlong-livedentanglementin\nscalable quantum networks proposed by Duan et al.[2],\nvariousexperimental resultsusingatomicensembles have\nbeen reported (such as [3]). Recently, d-dimensional\nquantum systems, or qudits, have been studied, and it\nhas been pointed out that qudits are better adapted for\ncertain purposes. As an example, they enable more effi-\ncientuseofcommunicationchannelsinquantumcryptog-\nraphy [4]. Following the pioneering experiment on para-\nmetric down-conversion [5], various protocols have been\ndemonstrated using orbital angular momentum (OAM)\nstates of photons (such as [6]). Photons are a promis-\ning carrier of quantum information. However, they are\ndifficult to store for appreciable periods of time. The re-\nalization of massive qudits and the ability to characterize\ntheir entanglement are therefore critical for applications.\nAnother recent landmark is the demonstration of the\nuseofOAMtogeneratearbitrarysuperpositionofatomic\nrotational states with the coherent transfer the OAM of\nlight to atoms in Bose-Einstein condensate [7, 8]. The\nexperiment illustrates the potential of OAM as a tool to\ngenerate and to control the atomic qudits.\nPrevious work [9] demonstrated entanglement associ-\nated with OAM in an ensemble of atoms and a photon.\nThe atomic OAM state is linked to the spatial degree\nof freedom of collective atomic excitations, and, in the\ncase of photons, the OAM state correspondsto Laguerre-\nGaussian (LG) modes. This result suggests that atomic\nensembles can be used as nodes of qudit-based quan-\ntum networks. However, previous observations were lim-\nited to two-qubit entanglement. In this letter, we report\nhigher-dimensionalityoftheentanglementofOAMstatesof an atomic ensemble and a photon, as confirmed by es-\ntimating the Schmidt number [10] of the reconstructed\ntwo-qutrit (i.e., qudits with d= 3) density matrix. For\npure states, the dimension of the range of the marginal\nstate is called the Schmidt rank, which describes how\nmany local levels are involved in the entanglement. Ex-\ntending this notion, the Schmidt number of a bipartite\nmixed state is defined to be the minimum Schmidt rank\nthat must appear in any decomposition of the state as a\nmixture of pure states. For example, any decomposition\nof a mixed state with Schmidt number 3 must include a\npure state of Schmidt rank 3 or greater.\nThe LG modes constitute a complete basis for describ-\ning the paraxial propagation of light [11, 12]. The in-\ntensity and phase distributions of several LG modes and\nsuperpositions of them are shown in Fig. 1. An LG mode\nis characterized by its two indices pandm, and by the\nGaussian beam waist w0. The integers pandmare the\nradial and azimuthal mode index, respectively, and the\nphase variation for a closed path around the optical axis\nis 2mπ. A single photon in the LGp,mmode carriesa dis-\nFIG.1: The intensity(left panelofeach pair)andphase(rig ht\npanel) distribution of several LG modes and their superposi -\ntions.2\ncrete OAM of m/planckover2pi1along its propagation direction. Here\nwe define |L/an}bracketri}ht,|G/an}bracketri}ht, and|R/an}bracketri}htto be the single photon states\nwith OAM of −/planckover2pi1, 0, and + /planckover2pi1, respectively.\nConsider an ensemble of atoms having a three-level\nstructure |a/an}bracketri}ht,|b/an}bracketri}ht, and|c/an}bracketri}ht, as shown in Fig. 2(a). Initially\nallofthe atomsarepreparedinlevel |a/an}bracketri}ht. Aclassicalwrite\npulsetunedtothe |a/an}bracketri}ht → |c/an}bracketri}httransitionwithproperdetun-\ning ∆ is incident on the atomic ensemble. In this process,\nthe|c/an}bracketri}ht → |b/an}bracketri}httransitionis stimulatedand a Stokesphoton\nis generated. The write pulse is weak and its interaction\ntime is short. Therefore the excitation probability of a\nStokes photon into a specified mode is much less than\nunity per pulse. The collective atomic excitation retains\nthe spatial distribution of the relative phase between the\nFIG. 2: (Color online) (a) Energy levels of87Rb, and the\nassociated laser frequencies. (b) Schematic of experimen-\ntal setup: SLM, spatial light modulator; SMF, single-mode\nfiber; SPCM, single photon counting module. Circularly\npolarized write and read pulses illuminate the87Rb MOT\n(Magnet-Optical Trap), and circularly polarized Stokes an d\nanti-Stokes photons are selectively directed onto the SPCM s\npassing through quarter-wave plates and polarizers. The pa ir\nof SLM and SMF serves as a spatial mode filter. (c) Co-\nincidence rate and time-resolved coincidence (inset) betw een\nthe Stokes and anti-Stokes photons of the LG0,0mode as a\nfunction of the time delay.write pulse and the Stokes photon. According to angular\nmomentum conservation, the state of the Stokes photon\nand collective atomic excitation will be entangled. We\nuse|l/an}bracketri}ht,|g/an}bracketri}ht, and|r/an}bracketri}htto denote the states of the collective\natomic excitation with OAM of −/planckover2pi1, 0, and + /planckover2pi1, respec-\ntively. In the present work, our measurement is sensitive\nto only the three-dimensional photonic and atomic OAM\nstates.\nWhen the write pulse carries zero OAM, The resultant\natoms-photon state will be\n|φ/an}bracketri}ht=CL|L/an}bracketri}ht|r/an}bracketri}ht+CG|G/an}bracketri}ht|g/an}bracketri}ht+CR|R/an}bracketri}ht|l/an}bracketri}ht,\nwhereCL,CG, andCRare the relative complex ampli-\ntudes. Theamplitudesdepend onthespatialshapeofthe\nregion where the write pulse interacts with the atoms.\nTheatoms-photonentanglementcanbetestedbymap-\nping the state of the atoms to that of an anti-Stokes\nphoton by illuminating the atomic ensemble with a laser\npulse (read pulse) resonant with the |b/an}bracketri}ht → |c/an}bracketri}httransition.\nThe efficiency of this transfer can be nearly unity be-\ncauseit correspondsto the retrievalprocessofthe atomic\nquantum memory based on electromagnetically induced\ntransparency [13].\nA schematic of the experimental setup is shown in\nFig. 2(b). An optically thick (optical depth of about\n5) cold atomic cloud is created using a magneto-optical\ntrap (MOT) for87Rb. Levels |a/an}bracketri}htand|b/an}bracketri}htcorrespond to\n5S1/2F= 1 and 2, respectively, and level |c/an}bracketri}htcorresponds\nto 5P1/2F′= 2. One cycle in the experiment comprises\na 6-ms loading period and a 4-ms measurement period.\nDuring the loading period, a gas of cold87Rb atoms is\ncooledandtrapped, andtheyareopticallypumped tothe\n5S1/2F= 1 level using a 50 µs depumping pulse tuned to\nthe 5S1/2F= 2→5P3/2F′= 2 transition. After the\nloading period, the magnetic field and the radiation re-\nsponsible for the cooling, trapping, and depumping are\nshut off. The vacuum cell is magnetically shielded using\na single-layer permalloy. The coil jig is non-metallic, and\nthus eddy currents, which prolong the decay of the mag-\nneticfield[9], aresuppressed. Theresidualmagneticfield\nis about 1 mG during the measurement period. During\nthat period, read and write pulses illuminate the atomic\nensemble with a 400-nsrepetition cycle. The 15-nsGaus-\nsian write pulse is tuned to the |a/an}bracketri}ht → |c/an}bracketri}httransition with\n10-MHz detuning and comprises 4 ×104photons. After a\n40-nsdelay, a 200-nsrectangularread pulse of300 µW in-\ntensity illuminates the MOT. As shown in Fig. 2(b), the\nreadandwritepulses,whosespatialmodesarecleanedup\nby passage through single-mode fibers SMF1 and SMF2,\nare counter-propagated along the same axis. The out-\nput beam from the fiber is focused into the MOT with\na Gaussian beam waist of 400 µm. Similarly, the Stokes\nandanti-Stokesphotonsshareasinglespatialmodewhen\ntheir conversions are appropriately chosen using spatial\nlight modulators SLM1 and SLM2 (Hamamatsu model\nX8267). Their Gaussian beam waist at the center of the3\nMOT is adjusted to 275 µm. The angle between the axis\nof the Stokes or anti-Stokes photons and that of SMF1\nor SMF2 is ∼3◦in order to spatially separate the weak\nStokes(anti-Stokes)photonsfromthe strongwrite (read)\npulses. The incident write and read pulses are circularly\npolarized, as are the Stokes and anti-Stokes photons.\nThe Stokes and anti-Stokes photons coupled into\nSMF3 and SMF4 are directed onto the single-photon-\ncounting modules SPCM1 and SPCM2 (Perkin-Elmer\nmodel SPCM-AQR-14). From the Stokes photon count-\ning, theexcitationprobabilityisestimated tobe5 ×10−4.\nTheir outputs are then fed into the start and stop inputs\nof the time interval analyzer. The experimental results\nfor the coincidence rate between the Stokes and anti-\nStokes photons in an LG0,0mode are displayed versus\ntime delay in Fig. 2(c), from which the normalized cross-\nintensity correlation is estimated to be g(2)\ns,as= 74.6±7.4,\nconfirming that the excitation probability of a Stokes\nphoton into a specified mode is much less than unity for\neach pulse. The coincidence count rate was 5 .2 s−1.\nTo determine the full state of the atoms and Stokes\nphoton, two-qutrit state tomography [14, 15] was per-\nformed, wherethe density matrixwasreconstructedfrom\nthe set of 81 measurements represented by the operators\nˆµi⊗ˆµj(withi,j= 0,1,···,8) and where ˆ µk≡ |k/an}bracketri}ht/an}bracketle{tk|.\nThe ket|k/an}bracketri}htforthe Stokesphoton waschosenfromamong\n{|L/an}bracketri}ht,|G/an}bracketri}ht,|R/an}bracketri}ht,(|G/an}bracketri}ht+|L/an}bracketri}ht)/√\n2,(|G/an}bracketri}ht+|R/an}bracketri}ht)/√\n2,(|G/an}bracketri}ht+\ni|L/an}bracketri}ht)/√\n2,(|G/an}bracketri}ht −i|R/an}bracketri}ht)/√\n2,(|L/an}bracketri}ht+|R/an}bracketri}ht)/√\n2,(|L/an}bracketri}ht+\nFIG. 3: (Color online) Gaussian components of applied phase\nmodulations T(x,y) to an incoming Gaussian beam of waist\nw0= 2.2 mm. (a) T(x,y) =eiarg(x−x0+i(y−y0)). The\nleft panel plots the simulation while the right panel shows\nthe experimental result. (b) T(x,y) =eiarg(x−x0+iy). (c)\nT(x,y) =eiπ\n2sgn(x−x0). In panels (b) and (c), the dots are\nexperimental results and the solid curves are obtained from\nthe numerical simulation.i|R/an}bracketri}ht)/√\n2}, and for the collective atomic excitation from\namong{|l/an}bracketri}ht,|g/an}bracketri}ht,|r/an}bracketri}ht,(|g/an}bracketri}ht+|l/an}bracketri}ht)/√\n2,(|g/an}bracketri}ht+|r/an}bracketri}ht)/√\n2,(|g/an}bracketri}ht−\ni|l/an}bracketri}ht)/√\n2,(|g/an}bracketri}ht+i|r/an}bracketri}ht)/√\n2,(|l/an}bracketri}ht+|r/an}bracketri}ht)/√\n2,(|l/an}bracketri}ht−i|r/an}bracketri}ht)/√\n2}.\nThese measurements are implemented using SLMs and\nSMFs; the SLMs produce spatial phase modulation and\nthe SMFs filter the LG0,0mode [16]. As reported in [17]\nin detail, arbitrary superpositions of an LG0,0and an\nLG0,±1mode can be converted into an LG0,0mode by\napplication of the phase modulation Tcorresponding to\nthe relative phase difference between the superposition\nmode and the LG0,0mode. As is clear from Fig. 1, such\nconversions can be achieved by moving the singularity in\nthe phase modulation to a particular location. Our re-\nflective SLMs have an active region of 768px ×768px.\nEven if the phase modulation is discrete, the fractional\nintensitydiffractedintohigherorderscanbedecreasedby\nadding the blazed phase grating structure. The spatial\nperiod of the grating is 4 px ∼100µm with a diffrac-\ntion efficiency of 25 %. In order to check the SLMs, the\nGaussian components of a beam diffracted with a spatial\nphase modulation of T(x,y) =eiarg(x−x0+i(y−y0))were\nmeasured. Here arg( z) is the argument of the complex\nnumberz. The results are plotted in Fig. 3(a) and (b),\nand are in good agreementwith numerical calculationsof\nthe superposition modes. The location of the singularity\nthatconvertsthesuperpositionmodeintoaGaussiancan\ntherefore be determined. At position ( x0,y0) = (0,0),\nwhere an incoming Gaussian mode is converted into an\nLG0,±1mode, the normalized intensity was measured to\nbe 3×10−3, indicating a high extinction ratio. Simi-\nlarly, the superposition mode ( LG0,−1+eiθLG0,+1)/√\n2\ncan be converted into an LG0,0mode by applying a dis-\ncontinuousphasemodulation. TheGaussiancomponents\nof the beam diffracted by an SLM with a spatial phase\nmodulation of T(x,y) =eiπ\n2sgn(x−x0)was also measured,\nwhere sgn( x) is the sign of x. The experimental results\nare shown in Fig. 3(c), and are again in good agreement\nwith numerical calculations.\nFig. 4 shows the graphical representation of density\nmatrixρexpreconstructed from the 81 coincidences. The\ntypical coincidence rate was roughly 5 s−1, and the data\nacquisition time of each measurement was 100 s. From\nthe density matrix, the fidelity to a maximally entangled\nstateFexp≡ /an}bracketle{tMES|ˆρexp|MES/an}bracketri}ht= 0.74±0.02 was ob-\ntained. Here |MES/an}bracketri}htwaschosenfromthesetofmaximally\nentangled states ( eiαπ|L/an}bracketri}ht|r/an}bracketri}ht+|G/an}bracketri}ht|g/an}bracketri}ht+eiβπ|R/an}bracketri}ht|l/an}bracketri}ht)/√\n3\nso as to maximize the fidelity, where the values of αand\nβwere 0.019πand−0.058π, respectively. The error in\nFexpis calculated by using Monte-Carlo method from\nthe statistical uncertainties in the coincidences. An op-\ntimal witness operator of Schmidt number 3 in C3⊗C3\nis given by ˆW3= 1−3|MES/an}bracketri}ht/an}bracketle{tMES|/2 [18], resulting in\nTr(ˆW3ˆρ)<0⇔ /an}bracketle{tMES|ˆρ|MES/an}bracketri}ht>2/3. The experimen-\ntal result Fexp>2/3 therefore confirms that the Schmidt\nnumber of the mixed state of the atomic ensemble and\nthe photon is greater than or equal to 3.4\nFIG. 4: (Color online) Graphical representation of the den-\nsity matrix ρexpof a state as estimated by quantum state\ntomography from the experimentally obtained coincidences .\nThe upper plot is the real part, and the lower plot is the\nimaginary part.\nThe OAM measurements were achieved using mode\nconversion by SLMs and mode filtering by SMFs in this\nexperiment. However, as frequently occurs in experi-\nments utilizing a spatial phase modulation, the mea-\nsurement bases cannot be realized completely accurately,\nresulting in unwanted radial and azimuthal components\ncomprising up to 20%. While this systematic effect in-\ncreases the error in Fexp, the resultant fidelity is never-\ntheless 0 .74+0.06\n−0.07, which is larger than 2 /3 even at the\nlowest error bound.\nThe major diagonal elements of the recon-\nstructed density matrix are /an}bracketle{tL|/an}bracketle{tr|ρexp|L/an}bracketri}ht|r/an}bracketri}ht=\n0.25,/an}bracketle{tG|/an}bracketle{tg|ρexp|G/an}bracketri}ht|g/an}bracketri}ht= 0.37,and/an}bracketle{tR|/an}bracketle{tl|ρexp|R/an}bracketri}ht|l/an}bracketri}ht=\n0.26. The summation of remaining elements\n1−0.25−0.37−0.26 = 0.12 suggests that compo-\nnents of non-zero total OAM, which may originate\nmainly from stray light, are one of the dominant factors\nlimiting the fidelity. The normalized cross intensity\ncorrelation g(2)\ns,as= 74.6±7.4 is significantly smaller thanthe value 2000 expected from the measured excitation\nprobability of 5 ×10−4. Therefore, stray light appears\nto have strongly affected the photon statistics and\ndecreased the fidelity. The imbalance of the diagonal\nelements may also affect the fidelity. However, even\nsupposing that the Stokes photons are locally filtered\nso as to balance the relative amplitude, the fidelity is\nstill expected to be 0 .74. This result confirms that the\nimbalance is not the dominant factor decreasing the\nfidelity in our case. Note that the diagonal elements can\nbe balanced by changing parameters such as the beam\nwaist of the write pulse since the relative amplitude\nis dependent on the spatial shape of the effective\ninteraction volume [19]. The other factor limiting the\nfidelity is the decoherence caused by the environmental\nnoise, such as the Larmor precession of the ground-state\nZeeman sublevels and the ballistic expansion of the\natomic ensemble.\nIn conclusion, higher-dimensionality of the entangle-\nment of OAM states has been observed for an atomic\nensemble and a photon by estimating the Schmidt num-\nber of the reconstructed density matrix. The experiment\ndescribed here enables one to communicate quantum in-\nformation encoded in the spatial degrees of freedom of a\nphoton and an atomic ensemble [20].\nWe gratefully acknowledge M. Ueda, K. Usami, and\nN. Kanai for valuable comments and stimulating dis-\ncussions. R.I. was supported by a Grant-in-Aid from\nJSPS. This work was supported by the Global Center\nof Excellence Program by MEXT, Japan through the\nNanoscience and Quantum Physics Project of the Tokyo\nInstitute of Technology, and by MEXT through a Grant-\nin-Aid for Scientific Research (B).\n∗inoue.r.aa@m.titech.ac.jp\n[1] H. J. Kimble, Nature 453, 1023 (2008).\n[2] L. Duan, M. D. Lukin, J. I. Cirac, and P. Zoller, Nature\n414, 413 (2001).\n[3] K. S. Choi, H. Deng, J. Laurat, and H. J. Kimble, Nature\n452, 67 (2008).\n[4] S. P. Walborn, D. S. Lemelle, M. P. Almeida, and\nP. H. Souto Ribeiro, Phys. Rev. Lett. 96, 090501 (2006).\n[5] A. Mair, A. Vaziri, G. Weihs, and A. Zeilinger, Nature\n412, 313 (2001).\n[6] N. K.Langford et al., Phys.Rev.Lett. 93, 053601 (2004).\n[7] M. F. Andersen et al., Phys. Rev. Lett. 97, 170406\n(2006).\n[8] K. Helmerson et al., Nucl. Phys. A, 790, 705c-712c\n(2007).\n[9] R. Inoue et al., Phys. Rev. A 74, 053809 (2006).\n[10] B. M. Terhal and P. Horodecki, Phys. Rev. A 61, 040301\n(2000).\n[11] L. Allen, M. W. Beijersbergen, R. J. C. Spreeuw, and\nJ. P. Woerdman, Phys. Rev. A 45, 8185 (1992).\n[12] G. F. Calvo, A. Picon, and E. Bagan, Phys. Rev. A 73,\n013805 (2006).5\n[13] M. Fleischhauer and M. D. Lukin, Phys. Rev. A 65,\n022314 (2002).\n[14] D. F. V. James, P. G. Kwiat, W. J. Munro, and A. G.\nWhite, Phys. Rev. A 64, 052312 (2001).\n[15] R. T. Thew, K. Nemoto, A. G. White, and W. J. Munro,\nPhys. Rev. A 66, 012303 (2002).\n[16] E. Yao et al., Optics Express 14, 13089 (2006).\n[17] A. Vaziri, G. Weihs, and A. Zeilinger, Journal of OpticsB: Quantum and Semiclassical Optics 4, S47 (2002).\n[18] A. Sanpera, D. Bruß, and M. Lewenstein, Phys. Rev. A\n63, 050301(R) (2001).\n[19] C. I. Osorio, S. Barreiro, M. W. Mitchell, and J. P. Tor-\nres, Phys. Rev. A 78, 052301 (2008).\n[20] G. Molina-Terriza, J. P. Torres, and L. Torner, Nature\nPhysics3, 305 (2007)." }, { "title": "0904.2396v2.Carrier_Density_and_Magnetism_in_Graphene_Zigzag_Nanoribbons.pdf", "content": "arXiv:0904.2396v2 [cond-mat.mes-hall] 25 Jun 2009CarrierDensity and Magnetism inGraphene ZigzagNanoribbo ns\nJ. Jung1,∗and A. H. MacDonald1\n1Department of Physics, University of Texas at Austin, USA\n(Dated: November 6, 2018)\nThe influence of carrier density on magnetism in a zigzag grap hene nanoribbon is studied in a π-orbital\nHubbard-model mean-field approximation. Departures from h alf-filling alter the magnetism, leading to states\nwith charge density variation across the ribbon and paralle l spin-alignment on opposite edges. Finite carrier\ndensities cause the spin-density near the edges to decrease steadily, leading eventually to the absence of mag-\nnetism. At low doping densities the system shows a tendency t o multiferroic order in which edge charges and\nspins are simultaneously polarized.\nINTRODUCTION\nGraphene sheets and related carbon based nanomaterials\nhave attracted attention recently after seminal experimen ts\n[1,2]revealednovelphysicsrelatedtotheiruniqueelectr onic\nstructure[3]. In graphene nanoribbons[4, 5, 6, 7, 8, 9, 10,\n11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23, 24] lateral\nconfinementleadstosizequantizationandtoone-dimension al\nconductionchannelswhosepropertiesdependqualitativel yon\nedgeterminationcharacter. Neutralzigzagterminatedrib bons\nhave attracted particular attention because they have a flat\nband, perflectly flat in simple π-band models, pinned to the\nFermi level. In self-consistent field (SCF) theories, inclu ding\nab initio spin-density-functional theories, the flat band leads\ntorobustmagneticorder. Ferromagneticalignmentofspins at\nthe zigzagedgesis predictedalso in treatmentsgoingbeyon d\nmean field [7, 8]. Although the reliability of SCF theories is\nuncertainandnot yet tested experimentally,interest in zi gzag\nedge magnetismhasremainedstrong becauseof potential for\ninterestingapplicationsin nano-electronics[18].\nMoststudiesoftheelectronicstructureofzigzagterminat ed\ngraphene ribbons have focused on properties of the neutral\nsystem or systems with substitutional doping [25]. We study\ntheroleofgatevoltageinducedchangesincarrierdensity, i.e.\ngate doping. A related work in the low carrier dopingregime\nwith an additional neutralizing background charge explore d\nthepossibilityofstable non-collinearmagneticstates[2 6]. In\nneutral systems, SCF theories predict edge magnetization i n\ngraphenenanoribbonswithoppositespinpolarizationsono p-\nposite edges [4, 9, 15]. In theoretical studies of locally ga ted\nzigzagribbonjunctionsusuallythe non-interactingelect ronic\nstructureisassumed[27,28,29,30,31,32,33],neglecting the\npossibility of doping-dependentinteraction-drivenrear range-\nments. Inthisworkweshowthatgatedopingleadstochanges\nin charge distribution, spin configuration, and total net sp in\npolarization, which are accompanied by important modifica-\ntions in electronic structure. Our study is based on the π-\norbital Hubbard-model SCF theory for the magnetic proper-\nties of graphenenanostructures[17, 34, 35, 36, 37], in whic h\nan electron of spin σin siteiexperiences a repulsive inter-\naction proportional to the density of opposite-spin electr ons\nniσ. The Hubbard-model SCF theory is broadly consistent\nwith DFT calculations when the interaction parameter UisW (˚A)δn\n100 90 80 70 60 50 40 30 20 0.6 \n0.4 \n0.2 \n0AFFb P\nF\nAFb NC \nFIG. 1: (Color online) Hubbard model SCF-theory phase diagr am\nas a function of doping per length δn(defined in the text) and rib-\nbon width Wfor nearest-neighbor hopping γ0=2.6eV, The on-site\nrepulsion strength was chosen to have a value U=2eVwhich re-\nproducetheribbonbandgapsobtainedintheLDA-DFTcalcula tions\nin reference [17]. We used 1200 k-points for Brillouin-zone sam-\npling. The energetic preference for opposite spins (AF) on o pposite\nedges isreplacedbya preference forparallel( F)spinsatlargerdop-\ning. Above a critical doping δn∼0.7 the SCF calculation does not\nfind magnetic states. Solutions at finite doping sometimes (A Fband\nFb)breaktheinversionsymmetryoftheribbon. Whennon-col linear\nspin (NC) is allowed canted spin solutions midway between AF and\nFconfiguration become energeticallyfavored atlow doping.\nchosen appropriately. π-orbital Hartree-Fock theory reduces\ntotheHubbardmodelwhenonlytheon-siteCoulombinterac-\ntions are retained. We have chosen to use a Hubbard interac-\ntionparameter U=2eVwhichreproducesintheundopedcase\ntheband-gapsobtainedbymicroscopicdensityfunctionalt he-\nory in the local density approximation. This value is smalle r\nthanotherestimates[34], buthasbeenadoptedwith asimila r\nmotivationin someotherrecentwork[35,36].\nThe Hubbard model mean-field Hamiltonian for each spin\nσis\nHσ=−γ0∑\n/angbracketlefti,j/angbracketrightc†\niσcjσ+U∑\niniσniσc†\niσciσ+vext∑\nic†\niσciσ2\nEP−EAFEFb−EAFEF−ENCEF−EAFEF−EAFb\nδn∆E (meV)\n0.5 0.4 0.3 0.2 0.1 04\n2\n0\n-2\nFIG. 2: (Color online) Total energy differences per edge ato m be-\ntweentheAF(orAFb)andF,Fb,Pstatesasafunctionof doping δn\nforarelativelynarrowribbonwith N=8atompairsperunitcell. For\nlowdopingAFbtypesolutionswithbrokenchargesymmetryar een-\nergeticallyfavored over AF solutions although the energy d ifference\nis very small. The non-collinear (NC) spin solutions are low est in\nenergy inthe weakly doped regime.\nconsist of a nearest neighbor tight-binding term with hop-\npingγ0=2.6eVconnecting lattice sites iandj, the Hubbard\ntermrepresentingelectron-electroninteractions,andan exter-\nnalpotentialtermaccountingfortheinteractionwiththec on-\nstant positivebackgroundchargeproportionalto a coeffici ent\nwe choose to be vext=−U. Given the uncertainty of predic-\ntionsimpliedbyparticularversionsofSCF-theory,theadv an-\ntagesofthisrelativelysimplemodeloftenoutweighdisadv an-\ntages. Because themagnetisminzigzagribbonsisessential ly\none-dimensional,we measuredoping δnin unitsof the num-\nberofexcesselectronsperrepeatdistance a=2.46˚Aalongthe\nedge. The corresponding areal density δn2D=δn/Wwhere\nthe ribbon width W=√\n3Na/2 andNis the number of atom\npairsperribbonunitcell.\nSCFSOLUTIONSAT FINITEDOPING\nThe main players in zigzag edge magnetism are the flat\nband states which occupy one-third of the one-dimensional\nribbonBrillouin-zone(BZ)andarelocalized[11]nearther ib-\nbonedges, most stronglyso nearthe BZ boundary |k|∼π/a.\nIn the undoped SCF ground state, electrons of opposite-spin\nare localized near opposite edgesof the ribbon and a gap[19]\nΔ ∝W−1separates occupied valence and empty conduction\nband ribbon states. By appealing to particle-hole symme-\ntry we can limit our discussion of doping to the n-type case\nin which electrons start filling the conduction band. Dop-\ning causes charge-density variation across the ribbon and t o\na complicated competition between band and interaction en-\nergies manifested by the variety of SCF equation solutions\nclassified below. We label solutions as AF (opposite) or F\n(parallel) to indicate the relative alignment of spins on op po-\nsite edges. The label NC is used indicate non-collinear spin\nsolutions. Thelabelbisappliedforsolutionswhichbreaki n-Fb FAFb \nδnζ\n0.4 0.2 00.03 \n0.02 \n0.01 \n0\nFIG. 3: (Color online) Net spin polarization obtained from t he total\nelectron spin densities ζ=/parenleftbig\nn↑−n↓/parenrightbig\n//parenleftbig\nn↑+n↓/parenrightbig\nfor AFb, F and Fb\nsolutions as a function of doping. AFb solutions collapse in to AF\nsolutions with zero net spin polarization for high enough do ping. F\nand Fb configurations also progressively lose net spin polar ization\nas they approach the non-magnetic Plimit. The shaded region rep-\nresents the doping regime where non-collinear solutions ar e favored\nenergetically.\nversion symmetry across the ribbon in a way which will be\nexplained in more detail later. Finally we use the letter P to\ndesignateaparamagneticstatewithnolocalspin-polariza tion.\nThephasediagraminFig. (1)illustratesthesequenceoftra n-\nsitions AF →NC→F→Fb→P in narrower ribbons. In\nwider ribbons we find an additional AF state region between\ntheF andFbregimes.\nThe total energy per unit cell consist of a sum over all the\noccupied single-particle eigenvalues εkmσlabeled with kand\nmthe band index divided by NKthe total number of k-points\nminusa termto accountforthe doublecountingcorrectionin\ntheinteraction\nE=1\nNKocc\n∑\nkmσεkmσ−U\n2∑\niσnlocal\niσnlocal\niσ\nwheretheoccupations nlocal\niσareevaluatedinthelocalframeat\nlatticesite iwherespinisdiagonal. Theirdifferencesbetween\ndifferentself-consistent solutionsare shown in Fig. ( 2) f or a\nparticular ( N=8) ribbon width when only collinear spin so-\nlutionsare considered. In the collinearscheme the energya s-\nsociatedwithbreakinginversionsymmetryacrosstheribbo ns\nisalwayssmallandthemaintrendisacrossoverfromantifer -\nromagnetic solutions at small δnto ferromagnetic solutions\nforδn/greaterorsimilar0.04 to non-magnetic solutions for δn/greaterorsimilar0.4. For\nthe F type solutions and those with broken charge symmetry\nthesystemhasanonzeronetspinpolarizationasafunctiono f\ndoping density. The doping dependence of spin-polarizatio n\nis illustrated for the same N=8 ribbon width in Fig. ( 3).\nEach of the solution types identified in Fig. (1) is associate d\nwith particular electronic structure features which are il lus-\ntrated in Fig. (4). For the AF solution, finite doping require s\nthatstatesabovetheinteractioninducedgapbeoccupied. F or\nsmalldopingelectronsstartoccupyingstatesnearthecond uc-\ntion band minima. (See Fig. (4).) These additional electron s3\nE↓E↑\nAFbδn= 0.03\nkaE↑/↓(eV)\nπ/2 2 π/3 π0.2\n0\n-0.2\n-0.4AFbδn= 0.03\nkaE↑/↓(eV)\nπ/2 2 π/3 π0.2\n0\n-0.2\n-0.4E↓E↑\nAFδn= 0.03\nka2π/3 πAFδn= 0.03\nka2π/3 πE↓E↑\nFδn= 0.1\nka2π/3 πFδn= 0.1\nka2π/3 πE↓E↑\nFbδn= 0.4\nka2π/3 πFbδn= 0.4\nka2π/3 πE↓E↑\nPδn= 0.46\nka2π/3 πPδn= 0.46\nka2π/3 π\nn↓n↑\nAFbδn= 0.03\ny(˚A)n↑/↓(y)\n16128400.6\n0.5\n0.4n↓n↑\nAFδn= 0.03\ny(˚A)1612840n↓n↑\nFδn= 0.1\ny(˚A)1612840n↓n↑\nFbδn= 0.4\ny(˚A)1612840n↓n↑\nPδn= 0.46\ny(˚A)1612840\nFIG.4: (Color online) Upper row. Bandstructures corresponding toAFb, AF,F,Fband Pspincol linear solutions of the Hubbard-model SCF\nequations forazigzagnanoribbon with N=8atompairsintheunitcell. Atfinitedopingtheenergygaind ue tothegappresent inthetheAFb\nand AF solutions is reduced, favoringthe Fsolution which do es not have a gap. Lower row. Up anddown spinelectron occupation per lattice\nsite in the unit cell across the ribbon. The AFb configuration has broken charge distribution symmetry relative to the rib bon center due to an\nunequal occupation of up and down spin bands. All mean-field b ands are invariant under k→−k. We show only the portion of the 1D BZ\nwithstatesclose tothe Fermilevel.\nsuffera largeenergypenaltyduetothe neutralsolutionban d-\ngapandhavelowerenergywhenspin-polarized. Theresultin g\nhalf metallic solution in the spin-collinear scheme implie s a\nnon-zerooverallspinpolarizationinthesystemandisacco m-\npanied by a breakingof chargedistributionsymmetry around\nthe ribbon center. This asymmetric charge distribution is a\ncombinedeffect of the net spin polarizationand the charact er\nof the AF solution at the neutrality point, in which electron s\nwith opposite spin polarizationsare concentratedon oppos ite\nedges[19]. Ifbothupanddownspinbandswereequallyoccu-\npiedtherewouldbenochargedistributionasymmetryaround\ntheribboncenter.\nIn the low doping regime a non-collinear spin-order that\ncontinuously bridges the intermediate situation between t he\nneutralAFconfigurationandmostlyFconfigurationathigher\ndoping is [26] a possibility. In the version of the Hubbard\nmodel mean-field theory which allows for non-collinear spin\ndenisties we must allow for the possibility that the average\nspin polarization on different lattice sites points in diff erent\ndirections [38]. This allows a larger variational space wit hin\nasingleSlaterdeterminantapproximationandcanpotentia lly\nleadtolowerenergysolutions,butthespinlabelbecomesun -\ndefined for each single-particle wave function. We verifiedthatnon-collinearspinsolutionsarefavoredenergetical ly[26]\nin the Hubbardmodel calculationsfor low dopingregionand\nthatthetransitiontoFconfigurationhappensatdopingdens i-\nties typically about 20% higher than when only collinear so-\nlutions are considered. The angle between the spin densitie s\non opposite edges and the band structure of the non-collinea r\nstate arerepresentedin Fig. 5.\nIn the intermediate doping regime the total energy is min-\nimized by solutions which are more similar to the F neutral-\nribbon configuration[19] which do not have an energy gap,\nand are thereforefavored by doping. This transition to F-li ke\nsolutions occurs already at a relatively small value of dopi ng\nδn≃0.06. The states that are occupied first at finite dop-\ning are those near the valley points |k|=2π/3athat are[19]\nspread across the ribbon and therefore control the exchange\ncouplingbetween opposite edges. The W-scaling rulesof the\nenergy bands near the valley points [19] are consistent with\ntheW−1decaylaw of the thresholddopingat which the tran-\nsition to F-like transition occurs in our numerical phase di a-\ngram. InelectronicstructureswithdominantlyFcharacter the\nchargedistributionsymmetryaroundtheribboncenterispr e-\nserved. In this case every occupied states, up or down spin\nand valenceor conductionedge band,has a symmetric distri-4\nnlocal \n↓nlocal \n↑\nE= 0 δn= 0 .05 \ny(˚A)nlocal ↑/↓(y)\n16 12 8400.6 \n0.5 \n0.4 NC δn= 0 .05 \nka E(eV )\nπ 2π/3 π/20.2 \n0\n-0.2 \n-0.4 NC δn= 0 .05 \nka E(eV )\nπ 2π/3 π/20.2 \n0\n-0.2 \n-0.4 \ny(˚A)nlocal ↑(y)−nlocal ↓(y)\n16 12 8 4 00.2 \n0.1 \n0δn= 0 .05 \nFIG. 5: (Color online) In the weakly doped region canted spin ori-\nentations develop in order to minimize the total energy when non-\ncollinear solutions are allowed. Upper row. Band structure and\nspinresolvedelectronoccupationper latticeforazigzagr ibbonwith\nN=8 andδn=0.05. The occupation and spin polarization at each\nlattice site are represented in a local frame where the spin i s diago-\nnal.Lowerrow. Spinpolarizationandrelativeorientationofthe spin\ndirection between different lattice sites represented wit h the arrow\nheads.\nbutionofelectrondensityaroundtheribboncenter. Whenth e\ndopingissufficientlylarge,however,we finda brokencharge\nsymmetry solution that we label as Fb. In this state one of\ntheoccupiedconductionbandshasAF(unbalancedacrossthe\nribbon) rather than F (balanced across the ribbon) characte r.\nIn addition to these solutions, we find that for wide ribbons\nthere is an intermediate doping region in which AF solutions\nhavelower total energythanthe F solutionsbeforethe Fb so-\nlutionisstabilized. Thedifferenceinenergybetweendiff erent\nmagneticsolutionsissmall atintermediateandlargedopin g.\nInthehighdopingregimethe magneticfeaturesofthe sys-\ntem progressively disappear as the edge state bands become\nfilled. The ribbon is found to turn paramagneticabove a crit-\nical value that increases with the ribbon width and saturate s\naroundδnc∼0.7. Considering that edge localized states in\nthe conduction bands with k-points near 2 π/3a≤ |k| ≤π/a\nspan approximately 1 /3 of the whole Brillouin zone we find\nthat the total amountof dopingelectronsrequiredto fill com -\npletelytheedgeforbothupanddownspinsis2 /3,anamount\nthat can be surpassed near the mentioned doping saturation\nlimit.DISCUSSION\nThe AF state of zigzag nanoribbons has the unusual fea-\nture that inversion symmetry across the ribbon is broken in\noppositesensesinthetwospinsub-systems[17,19]. Ourcal -\nculationsuggestthatinlowdopingregimethesystemcaneas -\nily develop solutions with a charge density that is distribu ted\nasymmetrically across the ribbon, creating an interesting and\nunusually strong type of multiferroic behavior [17, 39] in\nwhich spin polarization and charge density are coupled. We\nexpect that transport properties can correspondingly be ma -\nnipulated in interesting interrelated ways by both externa l\nmagneticfieldsandexternalelectric fieldsdirectedacross the\nribbon. Edgetransportshouldbe stronglysuppressed,fore x-\nample, when a transverse electric field is applied which has\noppositeorientationsonoppositeendsofa ribbon.\nAbove a certain critical doping density, which is inversely\nproportional to the ribbon width W−1, we find that the sys-\ntem undergoes a transition to a F configuration in which op-\nposite edgeshave parallel spin polarizations. Whendoping is\nincreased further the spin-configuration is altered yet aga in,\nrestoring inversion symmetry breaking across the ribbon. I n\nthis high-doping regime the total magnetic condensation en -\nergy is small and the energy differences between different\nmagnetic configurations is small. Eventually at sufficientl y\nhigh doping the Hubbard-model SCF equations have only\nparamagneticsolutions.\nTheSiO2substrates on which exfoliatedgraphenesamples\nare usually prepared have electron density inhomogeneitie s\n[40] of the order of nfluc∼1011cm−2that can extend over\nlengths of the order of L≃1µm. In the limiting case of rib-\nbonswiththissamewidthasthepuddlesizesaroughestimate\nofdopingperunitlatticeconstant awithineachpuddlecanbe\nevaluatedwiththeproductofthese twoquantities\nδnfluc∼L·nfluc∼0.25/a\nThis amount of doping can influence the spin configurations\nin the system and the presence of these randomperturbations\nis therefore expected to appreciably weaken the tendency to -\nwards magnetic order, especially for wide ribbons. For this\nreason we should expect a better chances of detecting edge\nmagnetism in suspended ribbons which have much weaker\nelectrondensityfluctuations.\nThe long-rangedcharacter of the Coulomb interaction, ne-\nglectedinthepresentwork,isexpectedtointroduceimport ant\nchangesin the details of electronic structure especially i n the\nregions in which states with different charge and spin con-\nfigurations compete closely. The discrepancies can be more\nacutethanintheneutralcasebecausetheinadequacyofshor t\nranged screening can be more relevant when the occupation\nofeachlatticesiteintheunitcellbecomesinhomogeneousa s\nwe depart from half filling. Nevertheless, it is also likely t hat\nseveral qualitative features of the solutions are still cor rectly\ncapturedbytheHubbardmodelandthereforecanprovideuse-\nfulhintsontheactualbehaviorofthemagneticconfiguratio ns5\nin ribbons as a function of doping. Even though fluctuation\neffects we have neglected may work against the formation\nof long-range order in these 1D-magnets, the unusually stif f\nferromagneticalignment of the spins predicted by mean field\ntheories for zigzag ribbons [16] suggests that magnetic ord er\ncould be possible and should be manifested in some way in\nexperiments.\nAcknowledgments. We gratefullyacknowledgehelpfuldis-\ncussions with P. M. Haney, T. Pereg-Barnea, K. T. Delaney,\nand P. Rinke. Financial support was receivedfromthe Welch\nFoundation, NRI-SWAN, ARO, DOE and the Spanish Min-\nistryofEducationthroughtheMEC-Fulbrightprogram.\n∗Electronicaddress: jeil@physics.utexas.edu\n[1] K.S.Novoselov etal.,Nature 438, 197(2005).\n[2] Y.Zhang et al.,Nature 438, 201 (2005)\n[3] A. H. CastroNeto, F. Guinea, N. M. R. Peres, K. S.Novoselo v\nand A. K. Geim, Rev. Mod. Phys. 81, 109 (2009); A. K. Geim\nand K. S. Novoselov et al., Nature Materials 6, 183 - 191\n(2007); A. K. Geim and A. H. MacDonald, Physics Today 60,\n35(2007).\n[4] M. Fujita, K. Wakabayashi, K. Nakada, K. Kusakabe, J. Phy s.\nSoc.Jpn. 65, 1920 (1996).\n[5] K.Nakada,M.Fujita,G.Dresselhaus,M.S.Dresselhaus, Phys.\nRev. B54, 17954 (1996).\n[6] K. Wakabayashi, M. Fujita, H. Ajiki, M. Sigrist, Phys. Re v. B\n59, 8271 (1999).\n[7] T.Hikihara, X. Hu, H.-H.Lin,and C.-Y.Mou Phys.Rev. B 68,\n035432 (2003).\n[8] S. Dutta, S. Lakshmi, and S. K. Pati, Phys. Rev. B 77, 073412\n(2008).\n[9] H.Lee,Y.-W.Son,N.Park,S.Han,andJ.Yu,Phys.Rev.B 72,\n174431 (2005).\n[10] M.Ezawa, Phys.Rev. B 73, 045432 (2006).\n[11] L.BreyandH.A. Fertig,Phys.Rev. B 73, 235411 (2006).\n[12] K.-I. Sasaki, S. Murakami, R. Saito, J. Phys. Soc. Jpn. 75,\n074713 (2006).\n[13] Y.-W. Son, Marvin L. Cohen, and Steven G. Louie, Phys. Re v.\nLett.97, 216803 (2006).\n[14] Y.-W.Son,MarvinL.Cohen, andStevenG.Louie,Nature 444,\n347(2006).\n[15] L. Pisani, J. A. Chan, B. Montanari, and N. M. Harrison Ph ys.\nRev. B75, 064418 (2007).\n[16] O. V. Yazyev, M. I. Katsnelson Phys. Rev. Lett. 100, 047209\n(2008).\n[17] J.Fern´ andez-Rossier, Phys.Rev. B 77, 075430 (2008).\n[18] W. Y. Kim and K. S. Kim, Nature Nanotechnology 3, 408\n(2008)\n[19] J. Jung. T. Pereg-Barnea, A. H. MacDonald, Phys. Rev. Le tt.102, 227205 (2009).\n[20] M. Y. Han, Barbaros ¨Ozyilmaz, Y. Zhang, and P. Kim, Phys.\nRev. Lett.98, 206805 (2007).\n[21] X. Li, X. Wang, L. Zhang, S. Lee, H. Dai, Science 319, 1229\n(2008).\n[22] S.S.Datta,D.R.Strachan,S.M.KhamisandA.T.C.Johns on\nNano Lett. 81912 (2008).\n[23] M. Zarea and N.Sandler, Phys.Rev. Lett.99, 256804 (200 7).\n[24] M. Zarea, C. Busser, and N. Sandler, Phys. Rev. Lett. 101 ,\n19680 (2008).\n[25] There have been several studies of the influence of subst itu-\ntionaldopingwhichchanges thechargedensityandatthesam e\ntime introduces structural modifications in the system: Bin g\nHuang, Qimin Yan, Gang Zhou, Jian Wu, Feng Liu, Bing-Lin\nGu, and Wenhui Duan, Appl. Phys. Lett. 91, 253122 (2007);\nF. Cervantes-Sodi, G. Csanyi, S. Piscanec, and A. C. Ferrari ,\nPhys.Rev.B77,165427(2008);N.Gorjizadeh,A.A.Farajian ,\nK. Esfarjani, and Y. Kawazoe Phys. Rev. B 78, 155427 (2008);\nS.S.Yua,W.T.Zheng,Q.B.WenaandQ.Jiang,Carbon 46,537\n(2008);T.Wassmannetal.Phys.Rev.Lett. 101,096402(2008);\nS. Dutta and S. K. Pati, J. Phys. Chem B 1221333 (2008); S.\nDutta,A.K.Manna,andS.K.PatiPhys.Rev.Lett. 102,096601\n(2009).\n[26] K. Sawada, F. Ishii, M. Saito, S. Okada, and T. Kawai, Nan o\nLett.9, 269 (2009).\n[27] J. M. Kinder, J. J. Dorando, H. Wang, G. K.-L. Chan,\narXiv:0901.3407v1\n[28] A. Cresti, G. Grosso, and G. P. Parravicini, Phys. Rev. B 77,\n233402 (2008).\n[29] A. Cresti, G. Grosso, and G. P. Parravicini, Phys. Rev. B 78,\n115433 (2008).\n[30] J. Nakabayashi, D. Yamamoto, S. Kurihara, Phys. Rev. Le tt.\n102, 066803 (2009).\n[31] A. Rycerz, J. Tworzydlo, and C.W.J. Beenakker, Nat. Phy s.3,\n172 (2007).\n[32] A. R. Akhmerov, J. H. Bardarson, A. Rycerz, and C. W. J.\nBeenakker, Phys.Rev. B77, 205416 (2008).\n[33] K. Wakabayashi and T. Aoki, Int. J. Mod. Phys. B 16, 4897\n(2002).\n[34] O. V.Yazyev, Phys.Rev. Lett.101, 037203 (2008).\n[35] J. Fern´ andez-Rossier and J. J. Palacios, Phys. Rev. Le tt. 99,\n177204 (2007).\n[36] J. Fern´ andez-Rossier, J.J.PalaciosPhys.Rev. Lett. 99, 177204\n(2007).\n[37] J. J. Palacios, J. Fern´ andez-Rossier, and L. Brey Phys . Rev. B\n77, 195428 (2008)\n[38] P. Haney, doctoral thesis ‘Spintronics in Ferromagnet s\nand Antiferromagnets From First Principles’, pages 79-81,\nhttp://www.ph.utexas.edu/ ∼haney411/PaulHaneyThesis.pdf\n[39] J. van den Brink et al. J. Phys. Cond. Matt. 20434217 (2008);\nN. A.Spaldinand M.Fiebig, Science 309, 391 (2005).\n[40] J. Martin, N. Akerman, G. Ulbricht, T. Lohmann, J. H. Sme t,\nK. vonKlitzing,and A.Yacobi, Nature Phys.4, 144 (2008)." }, { "title": "0905.1442v1.On_the_density_matrix_for_the_kink_ground_state_of_higher_spin_XXZ_chain.pdf", "content": "arXiv:0905.1442v1 [cond-mat.stat-mech] 11 May 2009Typeset with jpsj2.cls Short Note\nOn the density matrix for the kink\nground state of higher spin XXZ chain\nKohei Motegi\nInstitute of Physics, University of Tokyo, Meguro-ku, Toky o\n153\nKEYWORDS: XXZ chain, kink ground state, correlation func-\ntion\nThe exact computation of the correlation functions\nof 1D quantum integrable models has been one of the\nchallenging problems. For the spin-1/2 XXZ chain, the\ncorrelation functions in the antiferromagnetic regime\nwere found to be expressed in the multiple integral\nform.1,2)However, the exact evaluation of them is still\na hard work, and has been only successful for some spe-\ncial anisotropy parameters.3)\nIn theferromagnetic regime, there is a class of non-\ntranslationally invariant ground state which should be\ncalled as kinkground state4,5). We have studied the cor-\nrelation function of the kinkground state, and exactly\ncalculated the density matrix6)(see also ref 7).\nThe kink ground state also exists for arbitrary spin\nand dimension. In this paper, we extend the analysis de-\nveloped in our previous paper to higher spin 1D XXZ\nchain, and calculate the density matrix (note that the\nmodel we consider in this paper is not the integrable\nhigher spin XXZ chain8).\nTheHamiltonianofthe spin S1Dinfinite XXZ chain\nis\nH=−/summationdisplay\nm∈Z(Sx\nmSx\nm+1+Sy\nmSy\nm+1+∆(Sz\nmSz\nm+1−S2)).\n(1)\nWe shall consider the ferromagnetic regime ∆ >1. For\nlater convenience, we parametrize ∆ as ∆ = ( q1\n2+\nq−1\n2)/2.Then ∆ >1 corresponds to 0 < q <1. A kink\nground state is the superposition of kinks which have\nthe same center. For a (normalized) kink state ⊗x∈Z|mx/an}bracketri}ht\n(mx∈ {−S,···,S}), the center j−1/2 (j∈Z) is the\nposition where\n/summationdisplay\nxj(mx+S),\nholds. Let us denote the kink ground state whose center\nis atj−1\n2by|Ψj/an}bracketri}ht, and introduce the generatingfunction\nof|Ψj/an}bracketri}ht9)\n|Ψ(z)/an}bracketri}ht=/circlemultiplydisplay\nx∈Z<0(S/summationdisplay\nmx=−S(zq1\n2(1\n2+x))mx−Sc(mx)1\n2|mx/an}bracketri}ht)⊗/circlemultiplydisplay\ny∈Z≥0(S/summationdisplay\nmy=−S(zq1\n2(1\n2+y))my+Sc(my)1\n2|my/an}bracketri}ht),\n(2)\nwherec(mx) = (2S)!/(S−mx)!(S+mx)!.|Ψj/an}bracketri}htis the\ncoefficient of zjof the expansion of |Ψ(z)/an}bracketri}ht, i.e,\n|Ψ(z)/an}bracketri}ht=/summationdisplay\nj∈Zzj|Ψj/an}bracketri}ht. (3)\nLet us focus on one of the kink ground states |Ψ0/an}bracketri}ht, and\ncalculate the density matrix\n/an}bracketle{tn/productdisplay\nj=1Eǫ′\njǫj\nxj/an}bracketri}ht:=/an}bracketle{tΨ0|/producttextn\nj=1Eǫ′\njǫj\nxj|Ψ0/an}bracketri}ht\n/an}bracketle{tΨ0|Ψ0/an}bracketri}ht,(4)\nwhereEǫ′\nxǫx\nx|mx/an}bracketri}ht=δmx,ǫx|ǫ′\nx/an}bracketri}ht.xjis the position of the\nsite where the operator Eǫ′\njǫj\nxjacts on, and is assumed\nto bexj/ne}ationslash=xkforj/ne}ationslash=k. We only consider the case\nwhen/summationtextn\nj=1(ǫj−ǫ′\nj) = 0 is satisfied: otherwise, the\ndensity matrix is zero. We calculate /an}bracketle{tΨ(z)|Ψ(z)/an}bracketri}htand\n/an}bracketle{tΨ(z)|/producttextn\nj=1Eǫ′\njǫj\nxj|Ψ(z)/an}bracketri}htto obtain the exact expression\nof the density matrix since\n/an}bracketle{tΨ(z)|Ψ(z)/an}bracketri}ht=∞/summationdisplay\nj=−∞z2j/an}bracketle{tΨj|Ψj/an}bracketri}ht, (5)\n/an}bracketle{tΨ(z)|n/productdisplay\nj=1Eǫ′\njǫj\nxj|Ψ(z)/an}bracketri}ht=∞/summationdisplay\nj=−∞z2j/an}bracketle{tΨj|n/productdisplay\nj=1Eǫ′\njǫj\nxj|Ψj/an}bracketri}ht,\n(6)\nholds, and /an}bracketle{tΨ0|Ψ0/an}bracketri}ht,/an}bracketle{tΨ0|/producttextn\nj=1Eǫ′\njǫj\nxj|Ψ0/an}bracketri}htcan be ex-\ntracted from /an}bracketle{tΨ(z)|Ψ(z)/an}bracketri}htand/an}bracketle{tΨ(z)|/producttextn\nj=1Eǫ′\njǫj\nxj|Ψ(z)/an}bracketri}ht,\nwhich are easier to calculate. Let us first calculate\n/an}bracketle{tΨ(z)|Ψ(z)/an}bracketri}ht. Using the Jacobi triplet product identity\n(q;q)∞(−xq1\n2;q)∞(−x−1q1\n2;q)∞=∞/summationdisplay\nj=−∞xjqj2\n2,(7)\nwe have\n/an}bracketle{tΨ(z)|Ψ(z)/an}bracketri}ht=(−wq1\n2;q)2S\n∞(−w−1q1\n2;q)2S\n∞\n=1\n(q;q)2S∞/parenleftBig∞/summationdisplay\nj=−∞wjqj2\n2/parenrightBig2S\n=1\n(q;q)2S∞∞/summationdisplay\nj=−∞Ajwj, (8)\nwherew=z2, (a;q)∞:=/producttext∞\nj=0(1−aqj) and\nAj=/summationdisplay\nPjk=j,jk∈Z2S/productdisplay\nk=1qj2\nk\n2.\n1J. Phys. Soc. Jpn. Short Note\nUsing eq. (7) and the following identity6)\nn/productdisplay\nj=11\n1+xuj=∞/summationdisplay\nj=0(−x)jn/summationdisplay\nl=1uj+n−1\nl/producttext\ni/negationslash=l(ul−ui),(9)\n/an}bracketle{tΨ(z)|/producttextn\nj=1Eǫ′\njǫj\nxj|Ψ(z)/an}bracketri}htcan be calculated as\n/an}bracketle{tΨ(z)|n/productdisplay\nj=1Eǫ′\njǫj\nxj|Ψ(z)/an}bracketri}ht\n=n/productdisplay\nj=1{(wζj)[ǫ′\njǫj]c(ǫj)c(ǫ′\nj)}n/productdisplay\nj=11\n(1+wζj)2S\n×(−wq1\n2;q)2S\n∞(−w−1q1\n2;q)2S\n∞\n=1\n(q;q)2S∞n/productdisplay\nj=1{(wζj)[ǫ′\njǫj]c(ǫj)c(ǫ′\nj)}\n×∞/summationdisplay\nj=−∞/parenleftBig/summationdisplay\nj1∈Z,j2∈Z≥0\nj1+j2=jAj1Bj2/parenrightBig\nwj, (10)\nwhereζj=q1\n2+xj,/bracketleftbig\nǫ′\njǫj/bracketrightbig\n= (ǫj+ǫ′\nj+2S)/2 and\nBj=/summationdisplay\nPjk=j,jk∈Z≥02S/productdisplay\nk=1/braceleftBigg\n(−1)jkn/summationdisplay\nl=1ζjk+n−1\nl/producttext\ni/negationslash=l(ζl−ζi)/bracerightBigg\n.\nFrom eqs. (8) and (10), one has\n/an}bracketle{tΨ0|Ψ0/an}bracketri}ht=A0\n(q;q)2S∞, (11)\n/an}bracketle{tΨ0|n/productdisplay\nj=1Eǫ′\njǫj\nxj|Ψ0/an}bracketri}ht=1\n(q;q)2S∞n/productdisplay\nj=1{ζ[ǫ′\njǫj]\njc(ǫj)c(ǫ′\nj)}\n×/summationdisplay\nj1∈Z,j2∈Z≥0\nj1+j2=−Pn\nj=1[ǫ′\njǫj]Aj1Bj2,(12)\nleading to the exact expression of the density matrix\n/an}bracketle{tn/productdisplay\nj=1Eǫ′\njǫj\nxj/an}bracketri}ht=n/productdisplay\nj=1{ζ[ǫ′\njǫj]\njc(ǫj)c(ǫ′\nj)}\n×/summationdisplay\nj1∈Z,j2∈Z≥0\nj1+j2=−Pn\nj=1[ǫ′\njǫj]Aj1Bj2\nA0.(13)\nOne can check that for S= 1/2, eq. (13) recovers the\nresult in ref. 6. However, the expression of eq. (13) gets\nmore complicated as the spin becomes higher.\nConcentratingon the S= 1 case,we can derive some\nslightly easier expression by using\nn/productdisplay\nj=11\n(1+xuj)2=∞/summationdisplay\nj=0(−x)jXn,j,Xn,j= lim\nui+n→ui,i=1,···n2n/summationdisplay\nl=1uj+2n−1\nl/producttext\ni/negationslash=l(ul−ui),\nwhich is a special case of eq. (9). The result is\n/an}bracketle{tn/productdisplay\nj=1Eǫ′\njǫj\nxj/an}bracketri}ht=n/productdisplay\nj=1{ζ[ǫ′\njǫj]\njc(ǫj)c(ǫ′\nj)}\n×∞/summationdisplay\nj=0(−1)jXn,jq1\n4(j+Pn\nk=1[ǫ′\nkǫk])2\n×(δeven\nj+Pn\nk=1[ǫ′\nkǫk]+Cδodd\nj+Pn\nk=1[ǫ′\nkǫk]),(14)\nwhereC= 2/summationtext∞\nk=1q(k−1\n2)2/(1 + 2/summationtext∞\nk=1qk2). From eq.\n(14), the magnetization and the spin-spin correlation\nfunctions can be easily calculated.\n/an}bracketle{tSz\nx/an}bracketri}ht=∞/summationdisplay\nj=0(−1)j(j+1)ζjqj2\n4\n×(ζ2qj+1−1)(δeven\nj+Cδodd\nj), (15)\n/an}bracketle{tSz\nx1Sz\nx2/an}bracketri}ht=∞/summationdisplay\nj=0(−1)jX2,j(δeven\nj+Cδodd\nj)\n×(ζ2\n1ζ2\n2q(j+4)2\n4+qj2\n4−(ζ2\n1+ζ2\n2)q(j+2)2\n4),(16)\n/an}bracketle{tS+\nx1S−\nx2/an}bracketri}ht= 4ζ1\n2\n1ζ1\n2\n2∞/summationdisplay\nj=0(−1)jX2,j\n×{(δeven\nj+Cδodd\nj)(ζ1+ζ2)q(j+2)2\n4\n+(δodd\nj+Cδeven\nj)(q(j+1)2\n4+ζ1ζ2q(j+3)2\n4)},\n(17)\nwhere\nX2,j=(j+1)(ζj+3\n1−ζj+3\n2)−(j+3)ζ1ζ2(ζj+1\n1−ζj+1\n2)\n(ζ1−ζ2)3.\nFrom these expressions, one can easily show that the\nspin-spin correlation functions decay exponentially for\nlarge distances.\n/an}bracketle{tSz\nx1Sz\nx2/an}bracketri}ht−/an}bracketle{tSz\nx1/an}bracketri}ht/an}bracketle{tSz\nx2/an}bracketri}ht ∼Azz(x1)qx2+1\n2forx2≫1,\n(18)\nAzz(x1) = 2∞/summationdisplay\nj=0(−1)jqj2\n4(1−ζ2\n1qj+1)\n×(jζj−1\n1+(j+1)q1\n4Cζj\n1)(δeven\nj+Cδodd\nj),\n/an}bracketle{tS+\nx1S−\nx2/an}bracketri}ht ∼A+−(x1)q1\n2(x2+1\n2)forx2≫1, (19)\nA+−(x1) = 4ζ1\n2\n1∞/summationdisplay\nj=0(−ζ1)j(j+1)\n2/3J. Phys. Soc. Jpn. Short Note\n×{(δeven\nj+Cδodd\nj)ζ1q(j+2)2\n4+(δodd\nj+Cδeven\nj)q(j+1)2\n4}.\nAcknowledgment\nThe author thanks K. Sakai for useful discussions\nand comments on this work. This work was partially\nsupported by Global COE Program (Global Center of\nExcellence for Physical Sciences Frontier) from MEXT,\nJapan.\n1) M. Jimbo, K. Miki, T. Miwa and A. Nakayashiki, Phys. Lett.A168(1992) 256.\n2) N. Kitanine, J.M. Maillet, N.A. Slavnov and V. Terras, J Ph ys.\nA35(2002) L385.\n3) J. Sato and M. Shiroishi, Nucl.Phys. B 729(2005) 441.\n4) C.-T. Gottstein and R.F. Werner, e-print cond-mat/95011 23.\n5) F.C. Alcaraz, S.R. Salinas and W.F. Wreszinski, Phys. Rev .\nLett.75(1995) 930.\n6) K. Motegi and K. Sakai, Phys. Rev. E 79, (2009) 031108.\n7) R. Dijkgraaf, D. Orlando and S. Reffert, Nucl. Phys. B 811\n(2009) 463.\n8) L. A. Takhtajan, Phys. Lett. A 87(1982) 479.\n9) T. Komaand B.Nachtergaele, Adv.Theor.Math.Phys.2(199 8)\n533.\n3/3" }, { "title": "0905.2312v1.Spatial_distribution_of_local_density_of_states_in_vicinity_of_impurity_on_semiconductor_surface.pdf", "content": "arXiv:0905.2312v1 [cond-mat.mes-hall] 14 May 20096 pages, 4 figures\nSpatial distribution of local density of states in vicinity of impurity on semiconductor\nsurface\nV.N.Mantsevich∗and N.S.Maslova†\nMoscow State University, Department of Physics, 119991 Mos cow, Russia\n(Dated: December 4, 2018)\nWe present the results of detailed theoretical investigati ons of changes in local density of total\nelectronic surface states in 2 Danisotropic atomic semiconductor lattice in vicinity of im purity atom\nfor a wide range of applied bias voltage. We have found that ta king into account changes in density\nof continuous spectrum states leads to the formation of a dow nfall at the particular value of applied\nvoltage when we are interested in the density of states above the impurity atom or even to a series\nof downfalls for the fixed value of the distance from the impur ity. The behaviour of local density of\nstates with increasing of the distance from impurity along t he chain differs from behaviour in the\ndirection perpendicular to the chain.\nPACS numbers: 71.55.-i\nKeywords: D. Non-equilibrium effects; D. Many-particle int eraction; D. Tunneling nanostructures\nI. INTRODUCTION\nInfluence of different impurities on the semiconduc-\ntor local density of surface states was widely studied\nexperimentally and theoretically. Most of the experi-\nments were carried out with the help of scanning tun-\nneling microscopy/spectroscopy technique [1], [2], [3].\nTheoretical investigations of single impurities and clus-\nters influence on density of surface states deals with\nGreen’s functions formalism [4], [5] or based on the\ntotal-energy density-functional calculations using first-\nprinciple pseudo-potential [6]. Numerical calculations\nbased on the tight-binding model are also carried out\n[7].\nMost of the theoretical calculations don’t take into ac-\ncount modification of local density of continuous spec-\ntrum states due to the influence of impurity atom on\nsemiconductor surface which we consider to play an im-\nportant role in the formation of peculiarities in the local\ndensity of total surface states . So in the present work\nwe suggest a simple model of anisotropic atomic lattice\nwith impurity atom and we pay special attention to the\nchanges of local density of continuous spectrum states.\nThis model suits well for theoretical investigation of π-\nbonded chains on the reconstructed Ge or Si surfaces\n[8]. It can be also used for investigation of the sublat-\ntices on the cleaved planes of AIIIBVsemiconductors.\nWe have found that the view of local density of surface\nstates (amount of downfalls and their shape) strongly\ndiffers depending on the value of the distance from the\nimpurity atom position and from the direction of obser-\nvation (along the atomic chain or perpendicular to the\natomic chain). It will be shown that downfall in the res-\n∗vmantsev@spmlab.phys.msu.ru\n†Electronic address: spm@spmlab.phys.msu.ruonance when we are interested in the density of states\nabove the impurity atom can transforms to a series of\ndownfalls or even to a peak for different values of the\ndistance from the impurity.\nII. THE SUGGESTED MODEL AND MAIN\nRESULTS\nWe shall analyze 2 Danisotropic atomic lattice formed\nby the similar atoms with energy levels ε1and similar\ntunneling transfer amplitudes between the atoms talong\nthe atomic chain. The interaction between atomic chains\nis described by tunneling amplitude T, which has the\nsame value for all the similar atoms in the chain (Fig. 1).\nDistance between the atoms in the atomic chain is equal\ntoa, distance between the atoms in the neighboring\nchains is equal to b. Atomic lattice includes impurity\natom with energy level εd, tunneling transfer amplitude\nfrom impurity atom to the nearest atoms in the atomic\nchainτand to the nearest atoms in the neighbor chains\nℑ.\nThe model system can be described by the Hamilto-\nnian:ˆH:\nˆH=ˆH0+ˆHimp+ˆHtun\nˆH0=/summationdisplay\niεic+\nici+/summationdisplay\ntc+\nicj+/summationdisplay\nTc+\nkcl+h.c.\nˆHtun=/summationdisplay\ni,dτc+\nicd+/summationdisplay\nk,dℑc+\nkcd+h.c.\nˆHimp=/summationdisplay\ndεdc+\ndcd\n(1)\nˆH0is a typical Hamiltonian for atomic lattice with hop-\npings without any impurities. ˆHtundescribes transitions2\nbetween impurity atom and neighboring atoms of the\natomic lattice. ˆHimpcorresponds to the electrons in the\nlocalizedstateformedbytheimpurityatomintheatomic\nchain.\nIndexes i,j correspond to the direction along the chain;\nindexes k,l correspond to the direction perpendicular to\nthe chain.\nFIG. 1: Schematic diagram of 2 Datomic lattice with impu-\nrity atom.\nWe shall use diagramm technique in our investigation\nof atomic chain local density of states.\nThe dependence of local density of states on the dis-\ntance alongthe atomic chainand in the directionperpen-\ndicular to the atomic chain in the presence of impurity\natom is described by the equation:\nρ(ω,/vector r) =−1\nπSp/parenleftBig\nIm/summationdisplay\n/vector κ,/vector κ1ˆGR(/vector κ, /vector κ1,ω)ei/vector κ/vector rei/vector κ1/vector r/parenrightBig\n(2)\nWhere/vector r= (x,y),/vector κ= (κx,κy) and/vector κ1= (κx1,κy1).\nGreen function ˆGR(/vector κ, /vector κ1,ω) corresponds to the electron\ntransition from the impurity to the semiconductor con-\ntinuum states and can be found from the system of equa-\ntions:\nGR\nκx0dd=G0R\nκx0κx0τGR\ndd+G0R\nκx0κx0T/summationdisplay\nkyGR\nκxκydd\nGR\n0κydd=G0R\n0κy0κyℑGR\ndd+G0R\n0κy0κyt/summationdisplay\nkxGR\nκxκydd\nGR\ndd=G0R\ndd+G0R\nddτ/summationdisplay\nkxGR\nκx0dd+G0R\nddℑ/summationdisplay\nkyGR\n0κydd\nGR\n/vector κdd=G0R\n/vector κ/vector κτGR\nκx0dd+G0R\n/vector κ/vector κℑGR\n0κydd\nGR\n/vector κ/vector κ1=G0R\n/vector κ/vector κ+G0R\n/vector κ/vector κτGR\ndd/vector κ1+G0R\n/vector κ/vector κℑGR\ndd/vector κ1\nWhere zero Green function is evaluated for the 2 D\natomic lattice without any impurities and has the form:\nG0R\nκxκy(ω) =1\nω−ε1−2t·cos(kxa)−2T·cos(kyb)\n(3)\nSubstituting the expression for Green function\nˆGR(/vector κ, /vector κ1,ω) obtained from system into equation ( 2) and\nperforming summarization over wave vectors kxandkx1(kyandky1) we get the final expression for the local den-\nsity of continuous spectrum states along (perpendicular)\nthe atomic chain ρvolume(x) (ρvolume(y)):\nρvolume(ω,x) =ρ0(ω)·(ω−εd)2+γ2·(1−f(2kx(ω)x)\n(ω−εd)2+γ2\nρvolume(ω,y) =ρ0(ω)·(ω−εd)2+γ2·(1−f(2ky(ω)y)\n(ω−εd)2+γ2(4)\nWhere parameter γ= (τ2+ℑ2)·ρ0(ω) corresponds\nto relaxation rate of electron distribution at the local-\nized state formed by impurity atom, ρ0(ω) is a local den-\nsity of states for the atomic chain without any impuri-\nties. Functions f(2kx(ω)x) andf(2ky(ω)y) are periodi-\ncal and have the property: f(2kx(ω)x) = 1 ifx= 0 and\nf(2ky(ω)y) = 1 if y= 0. If both directions are equiv-\nalentf(2kx(ω)) =f(2ky(ω)y) =J0(2kx(ω)).Expression\nforkx(ω) orky(ω) can be found from the dispersion law\nof the 2Datomic lattice which has the form.\nω(kx,ky) = 2t·cos(kxa)+2T·cos(kyb)\n(5)\nImpurity atom density of states has lorentzian form\nline shape and can be evaluated as:\nρimpurity(ω) =−1\nπ·Im/summationdisplay\ndGR\ndd(ω) =γ2\n(ω−εd)2+γ2\n(6)\nLocaldensityoftotalsurfacestatesistheresultofsum-\nmarization between local density of continuous spectrum\nstates and impurity atom density of states.\nρ(ω) =ρvolume(ω)+ρimpurity(ω)\n(7)\nLet’s start from the 1 Dcase of the atomic chain. In this\ncase it is necessary to put in the Hamiltonian: b= 0,\nT= 0 and ℑ= 0. Final expression for the local density\nof continuous spectrum states ρvolumewill have the form:\nρvolume(ω) =��0(ω)·(ω−εd)2+γ2·(1−cos(2k(ω)r)\n(ω−εd)2+γ2\n(8)\nExpressionfor k(ω) can be found from the dispersion law\nof the 1Datomic chain.\nTypical numerical results for local density of total\nsurface states and local density of continuous spectrum\nstates calculated above the impurity atom in 1 Dcase\n(distance value is equal to zero ( r= 0)) are shown on\n(Fig. 2a,b). For the local density of continuous spectrum\nstates (Fig. 2a) a downfall exist in the resonance when\nenergy is equal to the impurity atom energy level depo-\nsition (ω=εd). Width of the downfall depends on the3\nFIG. 2: a) Local density of continuous spectrum states in the case of the distance from the impurity atom along the atomic\nchain equal to zero. b) Local density of surface spectrum sta tes in the case of the distance from the impurity atom along th e\natomic chain equal to zero. c)-f) Local density of continuou s spectrum states (black line) and local density of surface s pectrum\nstates (grey line) for the different values of the distance fr om the impurity atom along the atomic chain. For all the figure s\nvalues of the parameters a= 1,t= 1,5,εd= 0,6 are the same.\nparameters of the atomic chain, such as relaxation rate\nor tunneling transfer amplitude, it rises with the increas-\ning of tunneling transfer amplitude from impurity atom\nto the neighbor atoms of the atomic chain. Local den-\nsity of surface states (Fig. 2b) has lorentzian form with\na downfall in the resonance. With the increasing of re-\nlaxation rate (increasing of τ) the downfall depth at the\ntop of the peak decreases, resonance peak shape spreads\nand it’s amplitude falls down.\nNow let’s start to analyze the dependence of local den-\nsity of continuous spectrum states and local density of\nsurfacestatesatthefixedvalueofthedistance ralongthe\natomicchain fromthe impurity atomposition (Fig. 2c-f).\nWe shall again start from the local density of continuous\nspectrum states (black lines on Fig. 2).\nWhen the value of a distance is not equal to zero a se-\nries of downfalls in the local density of continuous spec-\ntrum exists. Amount of downfalls increases with the in-\ncreasing of distance value and downfalls amplitude de-\ncreases when energy aspire to the edges of the band. The\nmost significant amplitude of the downfalls corresponds\nto the vicinity of the resonance region. It is clearly evi-\ndent that positions of the downfalls on the energy scale\ncan be found from the equation 2 πn= 2k(ω)rwhere\nnis an integer number. This means that numerator ofthe equation ( 8) is equal to zero. When the distance is\nnot equal to zero not only a downfall in the resonance\n(Fig. 2d,f) but also a peak (Fig. 2c,e) can exist in the\nlocal density of continuous spectrum states. At the fixed\nparameters of the atomic chain existance of a downfall\nor a peak in the resonance is determined by the value\nof the distance. Local density of surface states is shown\nby the grey line on Fig. 2c-f. Comparison between local\ndensity of surface states and local density of continuous\nspectrum states for the fixed value of rmake it clearly\nevident that impurity atom density of states can drasti-\ncally influence on the local density of continuous spec-\ntrum states. Result depends on the value of tunneling\ntransfer amplitude from impurity atom to the nearest\natoms of the chain. We have found that the most signifi-\ncant influence corresponds to the situation when tunnel-\ning transfer amplitude between the atoms in the chain\ntsignificantly exceeds tunneling transfer amplitude from\nimpurity atom to the atoms of the chain τ. In this case\ndownfall in the resonance becomes a peak with a small\ndownfallonthe top ofthe peak(Fig. 2e). So onlyonesig-\nnificant downfall exists in the resonance region and there\nare no downfalls at the energies different from the reso-\nnance value (Fig. 2c)in the local density of surface states,\nwhileinthecontinuousspectrumdensityofstatesaseries4\nFIG. 3: a) Local density of continuous spectrum states (blac k line) and local density of surface spectrum states (grey li ne) for\nthe different values of the distance from the impurity atom al ong the atomic chain. For all the figures values of the paramet ers\na= 1,b= 2,t= 1,5,T= 1,2,τ= 0,6,ℑ= 0,3,εd= 0,6 are the same.\nof downfalls exists. With increasing of τlocal density of\nsurface states differs from the local density of continuous\nspectrum states only by the amplitude of the downfalls\n(Fig. 2e,f). In this case number of downfalls is the same\nin comparison with the continuous spectrum density of\nstates, downfalls don’t change their shape or position on\nthe energy scale and contribution to the local density of\nsurface states from the impurity atom can be considered\nto be a background. This effect can be qualitatively un-\nderstoodbythefollowingway: withincreasingoftransfer\namplitude from impurity atom to the atom of the chain\nincreases relaxation rate and lorentzian form peak of im-\npurity atom density of states spreads and it’s amplitude\ndecreases.\nNow let’s start to analyze 2 Datomic lattice. Nu-\nmerical results for local density of total surface states\nand local density of continuous spectrum states calcu-\nlated above the impurity atom in perpendicular direc-\ntions (along the atomic chain and perpendicular to the\natomic chain) are shown on Fig. 3a, Fig. 4a. It’s clear\nthat in this case all the results are equal for both direc-\ntions and downfall in the continuous spectrum density\nof states or a peak in the local density of surface states\nposesjust thesamepropertiesasin thecaseof1 Datomic\nchain.\nLet’s analyze the situation when the value of the dis-tances from the impurity atom position along the atomic\nchain (Fig. 3b-f) and perpendicular to the atomic chain\n(Fig. 4b-f) are not equal to zero. We shall start from the\nlocal density of continuous spectrum states (black lines\non Fig. 3,4).\nIn this case number of downfalls,their shape and po-\nsition on the energy scale in each of the perpendicular\ndirections can be found from the dispersion law just in\nthe same way as for 1 Datomic chain. When the distance\nis not equal to zero not only a downfall in the resonance\n(Fig. 3d,e; Fig. 4e,f) but also a peak (Fig. 3c,f; Fig. 4c,d)\ncan exists in the local density of continuous spectrum\nstates.\nWe have found distance interval for both directions\n(along the atomic chain and perpendicular to the atomic\nchain) where exists replacement of a peak by a down-\nfall (Fig. 3c-f; Fig. 4c-f) in local density of continuous\nspectrum states (and also in the local density of sur-\nface states). Moreover peak in one direction can cor-\nresponds to a downfall in the another direction. Let’s\nanalyze this interval carefully. We shall start from the\ndistance value when peaks in the resonance in local den-\nsity ofcontinuousspectrum states forboth directions can\nbe seen ( x(y) = 15·a) (Fig. 3c; Fig. 4c). With increasing\nof the distance value ( x(y) = 20·a) a resonance peak in\nthe direction perpendicular to the atomic chain still ex-5\nFIG. 4: Local density of continuous spectrum states (black l ine) and local density of surface spectrum states (grey line ) for\nthe different values of the distance from the impurity atom pe rpendicular to the atomic chain. For all the figures values of the\nparameters a= 1,b= 2,t= 1,5,T= 1,2,τ= 0,6,ℑ= 0,3,εd= 0,6 are the same.\nists (Fig. 4d) and in the direction along the atomic chain\na downfall appears (Fig. 3d). Further increasing of the\ndistance value ( x(y) = 25·a) shows that in the direction\nalong the atomic chain a downfall still exists (Fig. 3e)\nand in the perpendicular direction a downfall substitutes\npeak (Fig. 4e). Finally when the value of distances in\nboth directions becomes equal to x(y) = 30·ain the per-\npendicular direction a downfall still exists (Fig. 4f) and\nin the direction along the atomic chain a resonance peak\ncan be seen (Fig. 3f).\nLocal density of total surface states is shown by the\ngrey lines on Fig. 3,4. It is clearly evident that for the\nstudied parametersof the system taking into account im-\npurity atom density of states slightly changes local den-\nsity of total surface states in comparison with local den-\nsity of continuous spectrum states for both directions. In\nthis case number of downfalls is the same in comparison\nwith the continuousspectrum densityofstates, downfalls\ndon’t change their shape or position on the energy scale.\nIII. CONCLUSION\nIn this work we have shown that taking into account\nchangesofthelocaldensityofcontinuousspectrumstatesformed by the presence of the impurity atom in the 2 D\nanisotropic atomic lattice or even in the 1 Datomic chain\nleads to significant modification of the total local density\nof surface states and consequently to the modification of\nSTS spectra. We have found that a downfall exists in\nthe STS spectra measured just above the impurity when\nimpurity atom energy level is equal to the applied bias\nvoltage. With changingofthe distancefromthe impurity\na series of downfalls is formed on the energy scale both in\nthelocaldensityofcontinuousspectrumstatesandin the\nlocal density of total surface states. Number of downfalls\nand their position are determined by the atomic lattice\ndispersion law. It was shown that at some values of the\ndistance from the impurity a peak can exist in the reso-\nnance region instead of a downfall. We have found that\nbehaivour of local density of surface states depends on\nthe direction of the observation. Switching on and off\nof impurity atom in both directions was found. This ef-\nfect can be well observed experimentally with the help of\nSTM/STS technique.\nThis work was supported by RFBR grants and by the\nNational Grants for technical regulation and metrology\n01.648.12.3017 and 154 −6/259/4−08.6\n[1] R. Dombrowski, C. Wittneven, M. Morgenstern et al.,\nAppl. Phys. A ,66, S203-S206, (1998)\n[2] J. Inglesfield, M. Boon, S. Crampin, Condens. Matter ,12,\nL489-L496, (2000)\n[3] K. Kanisawa, M. Butcher, Y. Tokura, Phys. Rev. Letters ,\n87, 196804, (2001)\n[4] N. Sivan, N. Wingreen, Phys. Rev. B ,54, 11622, (1996)\n[5] V. Madhavan, W. Chen, M. Crommie et al., Phys. Rev.B,64, 165412, (2001)\n[6] M. Qian, M. Gothelid, B. Johnsson, Phys. Rev. B ,66,\n155326, (2002)\n[7] F. Marczinowski ,J. Weibe, J. Tang et al., Phys. Rev. Let-\nters,99, 1572002, (2007)\n[8] K. Pandey, Phys. Rev. Letters ,47, 1913, (1981)" }, { "title": "0905.2766v1.Gaps_and_tails_in_graphene_and_graphane.pdf", "content": "arXiv:0905.2766v1 [cond-mat.dis-nn] 17 May 2009Gaps and tails in graphene and graphane\nB. D´ ora1and K. Ziegler2\n1Max-Planck-Institut f¨ ur Physik Komplexer Systeme,\nN¨ othnitzer Str. 38, 01187 Dresden, Germany\n2Institut f¨ ur Physik, Universit¨ at Augsburg\nE-mail:klaus.ziegler@physik.uni-augsburg.de\nAbstract. We study the density of states in monolayer and bilayer graphene in t he\npresence of a random potential that breaks sublattice symmetrie s. While a uniform\nsymmetry-breaking potential opens a uniform gap, a random symm etry-breaking\npotential also creates tails in the density of states. The latter can close the gap again,\npreventing the system to become an insulator. However, for a suffi ciently large gap\nthe tails contain localized states with nonzero density of states. Th ese localized states\nallow the system to conduct at nonzero temperature via variable-r ange hopping. This\nresult is in agreement with recent experimental observations in gra phane by Elias et\nal..2\n1. Introduction\nGraphene is a single sheet of carbon atoms that is forming a honeyco mb lattice. A\ngraphene monolayer as well as a stack of two graphene sheets (i.e. a graphene bilayer)\nare semimetals with remarkably good conducting properties [1, 2, 3]. These materials\nhave been experimentally realized with external gates, which allow fo r a continuous\nchange in the charge-carrier density. There exists a non-zero min imal conductivity at\nthe charge neutrality point. Its value is very robust and almost una ffected by disorder\nor thermal fluctuations [3, 4, 5, 6].\nManypotentialapplicationsofgraphenerequireanelectronic gapt oswitch between\nconductingandinsulatingstates. Asuccessful stepinthisdirectio nhasbeenachieved by\nrecent experiments with hydrogenated graphene (graphane) [7] and with gated bilayer\ngraphene [8, 9, 10]. These experiments take advantage of the fac t that the breaking of\na discrete symmetry of the lattice system opens a gap in the electro nic spectrum at the\nFermi energy. In the case of monolayer graphene (MLG), a stagg ered potential that\ndepends on the sublattice of the honeycomb lattice plays the role of such symmetry-\nbreakingpotential(SBP).Forbilayergraphene(BLG)agatepote ntialthatdistinguishes\nbetween the two graphene layers plays a similar role.\nWith these opportunities one enters a new field in graphene, where o ne can switch\nbetween conducting and insulating regimes of a two-dimensional mat erial, either by a\nchemical process (e.g. oxidation or hydrogenation) or by applying a n external electric\nfield [11].\nTheopeningofagapcanbeobservedexperimentallyeitherbyadirec tmeasurement\nof the density of states (e.g., by scanning tunneling microscopy [12]) or indirectly by\nmeasuring transport properties. In the gapless case we observe a metallic conductivity\nσ∝ρD, whereDis the diffusion coefficient (which is proportional to the scattering\ntime)and ρisthedensityofstates(DOS).Thisgivestypicallyaconductivityoft heorder\nofe2/h. The gapped case, on the other hand, has a strongly temperatur e-dependent\nconductivity due to thermal activation of charge carriers [13]\nσ(T) =σ0e−T0/T(1)\nwith some characteristic temperature scale T0which depends on the underlying model.\nA different behavior was found experimentally in the insulating phase o f graphane [7]:\nσ(T)≈σ0e−(T0/T)1/3, (2)\nwhich is known as 2D variable-range hopping [14]. This behavior indicat es the\nexistence of well-separated localized states, even at the charge- neutrality point, where\nthe parameter T0depends on the DOS at the Fermi energy EFasT0∝1/ρ(EF).\nThe experimental observation of a metal-insulator transition in gra phane raises two\nquestions: (i) what are the details that describe the opening of a ga p and (ii) what is\nthe DOS in the insulating phase? In this paper we will focus on the mech anism of the\ngap opening due to a SBP in MLG and BLG. It is crucial for our study th at the SBP\nis not uniform in the realistic two-dimensional material. One reason fo r the latter is3\nthe fact that graphene is not flat but forms ripples [15, 16, 17]. An other reason is the\nincomplete coverage of a graphene layer with hydrogen atoms in the case of graphane\n[7]. The spatially fluctuating SBP leads to interesting effects, including a second-order\nphase transition due to spontaneous breaking of a discrete symme try and the formation\nof Lifshitz tails.\n2. Model\nQuasiparticles in MLG or in BLG are described in tight-binding approxima tion by a\nnearest-neighbor hopping Hamiltonian\nH=−/summationdisplay\ntr,r′c†\nrcr′+/summationdisplay\nrVrc†\nrcr+h.c., (3)\nwherec†\nr(cr) are fermionic creation (annihilation) operators at lattice site r. The\nunderlying lattice structure is either a honeycomb lattice (MLG) or t wo honeycomb\nlattices with Bernal stacking (BLG) [11, 18]. We have an intralayer ho pping rate tand\nan interlayer hopping rate t⊥for BLG. There are different forms of the potential Vr,\ndepending on whether we consider MLG or BLG. Here we begin with pot entials that are\nuniform on each sublattice, whereas random fluctuations are cons idered in subsection\n2.4.\n2.1. MLG\nVris a staggered potential with Vr=mon sublattice A and Vr=−mon sublattice\nB. This potential obviously breaks the sublattice symmetry of MLG. Such a staggered\npotential can be the result of chemical absorption of non-carbon atoms in MLG (e.g.\noxygen or hydrogen [7]). A consequence of the symmetry breaking is the formation of\na gap ∆ g=m: The spectrum of MLG consists of two bands with dispersion\nEk=±/radicalBig\nm2+ǫ2\nk, (4)\nwhere\nǫ2\nk=t2[3+2cos k1+4cos(k1/2)cos(√\n3k2/2)] (5)\nfor lattice spacing a= 1.\n2.2. BLG\nVris a biased gate potential that is Vr=m(Vr=−m) on the upper (lower) graphene\nsheet. The potential in BLG has been realized as an external gate v oltage, applied to\nthe two layers of BLG [8]. The spectrum of BLG consists of four band s [11] with two\nlow-energy bands\nE−\nk(m) =±/radicalbigg\nǫ2\nk+t2\n⊥/2+m2−/radicalBig\nt4\n⊥/4+(t2\n⊥+4m2)ǫ2\nk, (6)4\nwhereǫkis the monolayer dispersion of Eq. (5), and two high-energy bands\nE+\nk(m) =±/radicalbigg\nǫ2\nk+t2\n⊥/2+m2+/radicalBig\nt4\n⊥/4+(t2\n⊥+4m2)ǫ2\nk. (7)\nThe spectrum of the low-energy bands has nodes for m= 0 where E−\nk(0) vanishes\nin a (k−K)2manner, where Kis the position of the nodes, which are the same as\nthose of a single layer. For small m≪t⊥, a mexican hat structure develops around\nk=K, with local extremum in the low-energy band at E−\nk(m) =±m, and a global\nminimum/maximum inthe upper/lower low energy bandat E−\nk(m) =mt⊥//radicalbig\nt2\n⊥+4m2.\nFor small gating potential Vr=±mwe can expand E−\nk(m) under the square root\nnear the nodes and get\nE−\nk(m)∼ ±/radicalBig\n[1−4ǫ2\nkt⊥(t2\n⊥+4ǫ2\nk)−1/2]m2+E−\nk(0)2. (8)\nt⊥apparently reduces the gap. Very close to the nodes we can appro ximate the\nfactor in front of m2by 1 and obtain an expression similar to the dispersion of MLG:\nE−\nk(m)∼ ±/radicalbig\nm2+E−\nk(0)2. Here we notice the absence of the mexican hat structure\nin this approximation. The resulting spectra for MLG and BLG are sho wn in Fig. 1.\nPSfrag replacements\n|k|−KEk\nbilayermonolayer\nFigure 1. The energy spectra of MLG (blue) and BLG (red) are shown, with an d\nwithout a gap (dashed and solid line, respectively) for positive energ ies. Note the\ncharacteristic mexican hat structure of gapped BLG.\n2.3. Low-energy approximation\nThe two bands in MLG and the two low-energy bands in BLG represent a spinor-1/2\nwave function. This allows us to expand the corresponding Hamiltonia n in terms of\nPauli matrices σjas\nH=h1σ1+h2σ2+mσ3. (9)5\nNear each node the coefficients hjread in low-energy approximation [19]\nhj=i∇j(MLG), h1=∇2\n1−∇2\n2, h2= 2∇1∇2(BLG), (10)\nwhere (∇1,∇2) is the 2D gradient.\n2.4. Random fluctuations\nIn a realistic situation the potential Vris not uniform, neither in MLG nor in BLG,\nas discussed in the Introduction. As a result, electrons experienc e a randomly varying\npotential Vralong each graphene sheet, and min the Hamiltonian of Eq. (9) becomes a\nrandomvariableinspaceaswell. ForBLGitisassumed thatthegatevo ltageisadjusted\nat the charge-neutrality point such that in average mris exactly antisymmetric with\nrespect to the two layers: ∝angb∇acketleftm1∝angb∇acket∇ightm=−∝angb∇acketleftm2∝angb∇acket∇ightm.\nAt first glance, the Hamiltonian in Eq. (3) is a standard hopping Hamilto nian\nwith random potential Vr. This is a model frequently used to study the generic case of\nAnderson localization [20]. The dispersion, however, is special in the c ase of graphene\ndue to the honeycomb lattice: at low energies it consists of two node s (or valleys) K\nandK′[17, 19]. It is assumed here that randomness scatters only at small momentum\nsuch that intervalley scattering, which requires large momentum at least near the nodal\npoints (NP), is not relevant and can be treated as a perturbation. Then each valley\ncontributes separately to the DOS, and the contribution of the tw o valleys to the DOS\nρis additive: ρ=ρK+ρK′. This allows us to consider the low-energy Hamiltonian in\nEqs. (9), (10), even in the presence of randomness for each valle y separately. Within\nthis approximation the term mris a random variable with mean value ∝angb∇acketleftmr∝angb∇acket∇ightm= ¯mand\nvariance ∝angb∇acketleft(mr−¯m)(mr′−¯m)∝angb∇acket∇ightm=gδr,r′. The following analytic calculations will be\nbased entirely on the Hamiltonian of Eqs. (9),(10) and the numerical calculations on\nthe lattice Hamiltonian of Eq. (3). In particular, the average Hamilto nian∝angb∇acketleftH∝angb∇acket∇ightmcan be\ndiagonalized by Fourier transformation and is\n∝angb∇acketleftH∝angb∇acket∇ightm=k1σ1+k2σ2+ ¯mσ3 (11)\nfor MLG with eigenvalues Ek=±√\n¯m2+k2. For BGL the average Hamiltonian is\n∝angb∇acketleftH∝angb∇acket∇ightm= (k2\n1−k2\n2)σ1+2k1k2σ2+ ¯mσ3 (12)\nwith eigenvalues Ek=±√\n¯m2+k4.\n2.5. Symmetries\nLow-energy properties are controlled by the symmetry of the Ham iltonian and of the\ncorresponding one-particle Green’s function G(iǫ) = (H+iǫ)−1. In the absence of\nsublattice-symmetry breaking (i.e. for m= 0), the Hamiltonian H=h1σ1+h2σ2has a\ncontinuous chiral symmetry\nH→eασ3Heασ3=H (13)6\nwith a continuous parameter α, sinceHanticommutes with σ3. The term mσ3breaks\nthe continuous chiral symmetry. However, the behavior under tr ansposition hT\nj=−hj\nfor MLG and hT\nj=hjfor BLG in Eq. (10) provides a discrete symmetry:\nH→ −σnHTσn=H , (14)\nwheren= 1 for MLG and n= 2 for BLG. This symmetry is broken for the one-\nparticle Green’s function G(iǫ) by the iǫterm. To see whether or not the symmetry is\nrestored in the limit ǫ→0, the difference of G(iǫ) and the transformed Green’s function\n−σnGT(iǫ)σnmust be evaluated:\nG(iǫ)+σnGT(iǫ)σn=G(iǫ)−G(−iǫ). (15)\nFor the diagonal elements this is the DOS at the NP ρ(E= 0)≡ρ0in the limit ǫ→0.\nThus the order parameter for spontaneous symmetry breaking is ρ0. According to the\ntheory of phase transitions, the transition from a nonzero ρ0(spontaneously broken\nsymmetry) to ρ0= 0 (symmetric phase) is a second-order phase transition, and sho uld\nbe accompanied by a divergent correlation length at the transition p oint. Since our\nsymmetry is discrete, such a phase transition can exists in d= 2 and should be of Ising\ntype. A calculation, using the SCBA of ρ0, gives indeed a second-order transition at the\npoint where ρ0vanishes with a divergent correlation length ξfor the DOS fluctuations\nξ∼ξ0(m2\nc−¯m2)−1\nfor ¯m2∼m2\ncwith a finite coefficient ξ0[21]. Whether or not this transition is an\nartefact of the SCBA or represents a physical effect due to the a ppearence of two types\nof spectra (localized for vanishing SCBA-DOS and delocalized for non zero SCBA-DOS)\nis not obvious here and requires further studies.\n2.6. Density of states\nOur focus in the subsequent calculation is on the DOSof MLG and BLG. In the absence\nof disorder, the DOS of 2D Dirac fermions opens a gap ∆ ∝¯mas soon as a nonzero\nterm ¯mappears in the Hamiltonian of Eq. (9), since the low-energy dispersio n is\nEk=±√\n¯m2+k2for MLG and Ek=±√\n¯m2+k4for BLG, respectively (cf Fig. 2).\nHere we evaluate the DOS of MLG and BLG in the presence of a uniform gap. Given\nthe energy spectrum, the DOS is defined as\nρ(E) =/summationdisplay\nkδ(E−Ek). (16)\nBy using the MLG dispersion, this reduces to\nρ(E) =|E|Θ(|E|−m), (17)\nwhere Θ( x) is the Heaviside function. For BLG, this gives\nρ(E) =|E|\n2√\nE2−m2Θ(|E|−m), (18)7\n00.511.522.533.5400.511.522.533.54\nPSfrag replacements\nE/mmρ(E)\nBLGMLG\n0 1 2 3 4 5012345678910\nPSfrag replacements\nE/m\nmρ(E)\nBLG\nMLG\nt⊥=∞t⊥= 2m2t⊥=m\nE/mt⊥ρ(E)\nFigure 2. Density of states for a uniform symmetry-breaking potential for monolayer\ngraphene and bilayer graphene is shown in the left panel. The density of states for a\nuniform symmetry-breaking potential for BLG is shown for severa l values of t⊥. For\nsmallt⊥, the mexican hat structure influences the DOS by shifting the gap t o lower\nvalues, and by developing a kink at E=m.\nwhich are shown in Fig. 2. By retaining the full low-energy spectrum f or BLG, E−\nk, the\nDOS can still be evaluated in closed form, with the result\nρ(E) =|E|×\n\n(t2\n⊥+4m2)√\n(t2\n⊥+4m2)E2−t2\n⊥m2form >|E|>mt⊥√\nt2\n⊥+4m2/parenleftbigg\n(t2\n⊥+4m2)\n2√\n(t2\n⊥+4m2)E2−t2\n⊥m2+1/parenrightbigg\nfor|E|> m.(19)\nIn the limit of t⊥≫(E,m), this reduces to Eq. (18) after dividing it by t⊥, which was\nset to 1 in the low-energy approximation, and the DOS saturates to a constant value\nafter the initial divergence. For finite t⊥, however, the Dirac nature of the spectrum\nappears again, and the high energy DOS increases linearly even for t he BLG, similarly\nto the MLG case. For m= 0, and E≪t⊥, this lengthy expression gives\nρ(E≪t⊥) =t⊥\n2. (20)\nAn interesting question, from the theoretical as well as from the e xperimental point\nof view, appears here: What is the effect of random fluctuations ar ound ¯m? Previous\ncalculations, based on the self-consistent Born approximation (SC BA), have revealed\nthat those fluctuations can close the gap again, even for an avera ge SBP term ¯ m∝negationslash= 0\n[22]. Only if ¯ mexceeds a critical value mc(which depends on the strength of the\nfluctuations), an open gap was found in these calculations. This des cribes a special\ntransition from metallic to insulating behavior. In particular, the DOS at the Dirac\npointρ0vanishes with ¯ mlike a power law\nρ0(¯m)∼/radicalbig\n��m−m2\nc. (21)8\nThe exponent 1/2 of the power law is probably an artefact of the SC BA, similar to the\ncritical exponent in mean-field approximations.\n3. Self-consistent Born approximation\nThe average one-particle Green’s function can be calculated from t he average\nHamiltonian ∝angb∇acketleftH∝angb∇acket∇ightmby employing the self-consistent Born approximation (SCBA)\n[23, 24, 25]\n∝angb∇acketleftG(iǫ)∝angb∇acket∇ightm≈(∝angb∇acketleftH∝angb∇acket∇ightm+iǫ−2Σ)−1≡G0(iη,ms). (22)\nThe SCBA is also known as the self-consistent non-crossing approx imation in the Kondo\nand superconducting community. The self-energy Σ is a 2 ×2 tensor due to the spinor\nstructure of the quasiparticles: Σ = −(iησ0+msσ3)/2. Scattering by the random SBP\nproduces an imaginary part of the self-energy η(i.e. a one-particle scattering rate) and\na shiftmsof the average SBP ¯ m(i.e., ¯m→m′≡¯m+ms). Σ is determined by the\nself-consistent equation\nΣ =−gσ3(∝angb∇acketleftH∝angb∇acket∇ightm+iǫ−2Σ)−1\nrrσ3. (23)\nThe symmetry in Eq. (14) implies that with Σ also\nσnΣσn=−(iησ0−msσ3)/2 (24)\nis a solution (i.e. ms→ −mscreates a second solution).\nThe average DOS at the NP is proportional to the scattering rate: ρ0=η/2gπ.\nThis reflects that scattering by the random SBP creates a nonzer o DOS at the NP if\nη >0.\nNow we assume that the parameters ηandmsare uniform in space. Then Eq. (23)\ncan be written in terms of two equations, one for the one-particle s cattering rate ηand\nanother for the shift of the SBP ms, as\nη=gIη, m s=−¯mgI/(1+gI). (25)\nIis a function of ¯ mandηand also depends on the Hamiltonian. For MLG it reads with\nmomentum cutoff λ\nIMLG=1\n2πln/bracketleftbigg\n1+λ2\nη2+(¯m+ms)2/bracketrightbigg\n(26)\nand for BLG\nIBLG∼1\n4/radicalbig\nη2+(¯m+ms)2(λ∼ ∞). (27)\nA nonzero solution ηrequiresgI= 1 in the first part of Eq. (25), such that ms=−¯m/2\nfrom the second part. Since the integrals Iare monotonically decreasing functions for\nlarge ¯m, a real solution with gI= 1 exists only for |¯m| ≤mc. For both, MLG and BLG,\nthe solutions read\nη2= (m2\nc−¯m2)Θ(m2\nc−¯m2)/4, (28)9\na)\nb)E\nEDOS\nDOS\nFigure 3. Schematic shape of the density of states: full curves are the bulk\ndensity of states for uniform symmetry-breaking potential, dott ed curves represent the\nbroadening by disorder. The broadened density of states can ove rlap inside the gap\nfor ¯m < m c(a) or not for ¯ m > m c(b), depending on the average symmetry-breaking\npotential ¯ m.mcis given in Eq. (29).\nwhere the model dependence enters only through the critical ave rage SBP mc:\nmc=/braceleftBigg\n(2λ/√\ne2π/g−1)∼2λe−π/gMLG\ng/2 BLG. (29)\nmcis much bigger for BGL, a result which indicates that the effect of diso rder is much\nstrongerinBLG.Thisisalsoreflectedbythescatteringrateat ¯ m= 0whichis η=mc/2.\nA central assumption of the SCBA is a uniform self-energy Σ. The ima ginary\npart of Σ is the scattering rate η, created by the random fluctuations. Therefore, a\nuniformηmeans that effectively random fluctuations are densely filling the latt ice. If\nthe distribution of the fluctuations is too dilute, however, there is n o uniform nonzero\nsolution of Eq. (23). Nevertheless, a dilute distribution can still cre ate a nonzero DOS,\naswewill discuss inthefollowing: westudy contributions totheDOSdu etorareevents,\nleading to Lifshitz tails.\n4. Lifshitz tails\nInthesystem withuniformSBPthegapcanbedestroyed locallybyalo calchangeofthe\nSBPm→m+δmrdue to the creation of a bound state. We start with a translational-\ninvariant system and add δmron siter. To evaluate the corresponding DOS from the\nGreen’s function G= (H+iǫ+δmσ3)−1, using the Green’s function G0= (H+iǫ)−1\nwith uniform m, we employ the lattice version of the Lippmann-Schwinger equation [2 6]\nG=G0−G0TSG0= (1−G0Tr)G0 (30)\nwith the 2 ×2 scattering matrix\nTr= (σ0+δmrσ3G0,rr)−1σ3δmr. (31)10\nIn the case of MLG we have\nG0= [(E+iǫ)σ0−mσ3]1\n2π/integraldisplayλ\n0k\n(ǫ−iE)2+m2+k2dk (32)\n∼(Eσ0−mσ3)1\n4πlog[1+λ2/(m2−E2)]+o(iǫ)≡(g0+iǫs)σ0+g3σ3.(33)\n(remark: the DOS of BLG has the same form.) Then the imaginary par t of the Green’s\nfunction reads\nIm[G(η)] =−/parenleftBigg\nδǫs(g0+g3+δmr) 0\n0 δǫs(g0−g3−δmr)/parenrightBigg\n(34)\nwith\nδǫs(x) =ǫs\nx2+ǫ2s2. (35)\nThus the DOS is the sum of two Dirac delta peaks\nρr∝δǫs(g0+g3+δmr)+δǫs(g0−g3−δmr). (36)\nThe Dirac delta peak appears with probability ∝exp(−(g0±g3)2/g) for a Gaussian\ndistribution. This calculation can easily be generalized to δmron a set of several\nsitesr[26]. Then the appearance of the several such Dirac delta peaks de creases\nexponentially. Moreover, these contributions are local and form lo calized states. For\nstronger fluctuations δmr(i.e., for increasing g) the localized states can start to overlap.\nThis is a quantum analogue of classical percolation.\nThe localized states in the Lifshitz tails can be taken into account by a\ngeneralization of the SCBA to non-uniform self-energies. The main id ea is to search\nfor space-dependent solutions Σ rof Eq. (23). In general, this is a diffult problem.\nHowever, we have found that this problem simplifies essentially when w e study it in\nterms of a 1 /¯mexpansion. Using a Gaussian distribution, this method gives Lifshitz\ntails of the form [27]:\nρ0(¯m)∼¯m4\n32√πg5/2e−¯m2/4g. (37)\n5. Numerical approach\nTo understand to behavior of random gap fluctuations in graphene , and also the\nlimitations of the SCBA, we carried out extensive numerical simulation s on the\nhoneycomb lattice, allowing for various random gap fluctuations on t op of a uniform\ngapm. These fluctuations are simulated by box and Gaussian distributions . From the\nSCBA, the emergence of a second-order phase transition at a crit ical mean mcis obvious\nfor a given variance. This is best manifested in the behavior of the DO S, which stays\nfinite for ∝angb∇acketleftm∝angb∇acket∇ight< mc, and vanishes afterwards, and serves as an order parameter. D oes\nthis picture indeed survive, when higher order corrections in the flu ctuations are taken\ninto account?11\nTo start with, we take a fix random mass configuration with a given va riance and\nthe honeycomb lattice (HCL) with the conventional hoppings ( t), represented by H0.\nThen, wetakeaseparateHamiltonian, responsiblefortheuniform, non-fluctuatinggaps,\ndenoted by Hgap, and study the evolution of the eigenvalues of H0+mHgapby varying\nm for a 600x600 lattice. By using Lanczos diagonalization, we focus o ur attention\nonly to the 200 eigenvalues closest to the NP. Their evolution is shown in Fig. 4.\nThis supports the existence of a finite mc, but since it originates from a single random\ndisorder configuration, rare events can alter the result. As a pos sible definition of the\nrigid gap, we also show the maximum of the energy level spacing for th ese eigenvalues\nas a function of m. As seen, it starts to increase abruptly at a given value of m, which\ncan define mc.\n00.511.52−101−0.500.5−0.500.5\nPSfrag replacements\nm/teigenvalues /t\ng= 0.62g= 0.82g= 1\nFigure 4. (Color online) The evolution of the 200 lowest eigenvalues is shown for a\ngiven random mass configuration with Gaussian distribution (with var iance g) on a\n600x600 HCL, by varying the uniform gap. The red line denotes the m aximum of the\nlevel spacing of these eigenvalues, a possible definition of the avera ge gap.\nTo investigate whether a finite critical mcsurvives, we take smaller systems and\nevaluate the averaged DOSdirectly frommany disorder realizations . To achieve this, we\ntake a 200x200 HCL, and evaluate the 200 closest eigenvalues to th e NP, and count their\nnumber in a given small interval, ∆ E(smaller than the maximal eigenvalues) around\nzero. This method was found to be efficient in studying other types o f randomness [28].\nWe mention that large values of ∆ Etake contribution from higher energy states into\naccount, while too small values are sensitive to the discrete lattice a nd consequently the12\n0 0.5 1 1.5 200.020.040.060.080.10.12\n0.511.511.522.5\nPSfrag replacements\nm/tρ(0)t\ngexponent ( c)\nFigure 5. (Color online) The density of states at the NP is plotted for Gaussian\nrandom mass for a 200x200 HCL for g= 0.92, 1, 1.12, 1.22and 1.32from bottom to\ntop after 400 averages. The symbols denote the numerical data, solid lines are fits\nusingaexp(−bmc). The inset shown the obtained exponents, c, as a function of g,\nwhich is close to 1.5.\ndiscrete eigenvalue structure of the Hamiltonian. For lattices cont aining a few 104−105\nsites, ∆E/t∼10−2−10−4are convenient.\nThe resulting DOS is plotted in Figs. 5 and 6 for Gaussian (with variance g) and\nbox distribution (within [ −W..W], variance g=W2/3). This does not indicate a sharp\nthreshold, but rather the development of long Lifshitz tails due to r andomness, as we\nalready predicted in the previous section. To analyze them, we fitte d the numerical data\nby assuming exponential tails of the form\nρ(0) =texp(−a−bmc) (38)\nfor a Gaussian and\nρ(0) =texp(−a−b/|m−W|c) (39)\nfor a box distribution, as suggested by Ref. [29]. The obtained cvalues are visualized in\nthe insets of Figs. 5 and 6. Given the good agreement, we believe tha t the DOS at the\nNP is made of states that are localized in a Lifshitz tail. We mention that these results\nare not sensitive to finite size scaling at these values of the disorder and uniform gap,\nonly smaller systems (like the 30x30 HCL) require more averages ( ∼104), whereas for\nlarger ones (such as the 200x200 with 400 averages) fewer avera ges are sufficient.\nIn Fig. 7, the energy dependent DOS is shown for Gaussian random m ass with\ng= 1 and for several uniform gap values. With increasing m, the DOS dimishes rapidly\nat low energies, and develops a pseudogap. The logarithmic singularit y atE=tis\nwashed out for g= 1. We also show the inverse of the DOS, proportional to T0, the13\ncharacteristic temperature scale of variable range hopping as a fu nction of the carrier\ndensity (which is proportional to E2).\n00.20.40.60.8 11.200.020.040.060.080.10.120.14\n0.511.51.41.61.82\nPSfrag replacements\nm/tρ(0)t\ngexponent ( c)\nFigure 6. (Color online) The density of states at the NP is plotted for box distr ibuted\n([−W..W]) randomness for a 200x200 HCL for W= 1.5, 1.7 and 2 ( g=W2/3) from\nbottom to top after 400 averages. The symbols denote the numer ical data, solid lines\nare fits using aexp(−b/|m−W|c). The inset shown the obtained exponents, c, as a\nfunction of g.\n6. Discussion\nMLG and BLG consist of two bands that touch each other at two nod al points (or\nvalleys). Near the nodes the spectrum of MLG is linear (Dirac-like) an d the spectrum\nof BLG is quadratic. The application of a uniform SBP opens a gap in the DOS for both\ncases. For a random SPB, however, the situation is less obvious. Fir st of all, it is clear\nthat randomness leads to a broadening of the bands. If we have tw o separate bands\ndue to a small uniform SPB, randomness can close the gap again due t o broadening (cf.\nFig. 2a). The broadening of the bands depends on the strength of the fluctuations of\nthe random SBP. In the case of a Gaussian distribution there are en ergy tails for all\nenergies.\nNow we focus on the NP, i.e. we consider E= 0 and ρ0. Then we have two\nparameters in order to change the gap structure: the average S BP∝angb∇acketleftm∝angb∇acket∇ight ≡¯mand the\nvariance g. ¯mallows us to broaden the gap and ghas the effect of closing it due to\nbroadening of the two subbands. Previous calculations have shown that at the critical\nvaluemc(g) of Eq. (29) the metallic behavior breaks down for ¯ m > m c(g) [22]. On\nthe other hand, Gaussian randomness creates tails at all energies . Consequently, there\nare localized states for |¯m| ≥mc(g) at the NP, and there are delocalized states for14\n00.5 11.5 22.5 300.050.10.150.20.250.30.350.4\nPSfrag replacements\nE/tρ(E)t\n(E/t)2∼n\n1/ρ(E)t∼T0−0.1 −0.05 0 0.05 0.105101520253035404550\nPSfrag replacements\nE/t\nρ(E)t\n(E/t)2∼n1/ρ(E)t∼T0\nFigure 7. (Color online) The energy dependent density of states is plotted fo r\nGaussiandistributed randommassfora30x30HCLafter 104averagesfor g= 1,m= 2\n(cyan), 1 (blue), 0.5 (red), 0.3 (black), 0.2 (magenta) and 0 (gree n) in the left panel.\nThe right panel visualizes the inverse of the density of states, bein g proportional to T0\nin the variable range hopping model as a function of the energy squa red (proportional\nto the carrier density).\n|¯m|< mc(g) at the NP. The localized states in the tails are described, for instan ce, by\nthe Lippmann-Schwinger equation (30) . The SCBA with uniform self- energy is not able\nto produce the localized tails. An extension of the SCBA with non-unif orm self-energies\nprovides localized tails though, as an approximation for large ¯ mhas shown [27]. This is\nalso in good agreement with our exact diagonalization of finite system s up to 200 ×200\nsize.\nA possible interpretation of these results is that there are two diffe rent types of\nspectra. In a special realization of mrthe tails of the DOS represent localized states.\nOn the other hand, the DOS at the NP E= 0, obtained from the SCBA with uniform\nself-energy, comes from extended states [22]. The localized andth e delocalized spectrum\nseparate at the critical value mc, undergoing an Anderson transition.\nConductivity : Transport, i.e. the metallic regime, is related to the DOS trough the\nEinstein relation σ∝ρD, whereDis the diffusion coefficient. The latter was found in\nRef. [22] for E∼0 as\nD=ag/radicalbig\nm2c−¯m2\n2πm2cΘ(m2\nc−¯m2), (40)\nwherea= 1 (a= 2) for MLG (BLG). Together with the DOS ρ0=η/2gπand the\nscattering rate ηin Eq. (28), the Einstein relation gives us at the NP\nσ(ω∼0)∝ρ0De2\nh≈a\n8π2/parenleftbigg\n1−¯m2\nm2c/parenrightbigg\nΘ(m2\nc−¯m2)e2\nh. (41)\nIn the localized regime (i.e. for |¯m| ≥mc) the conductivity is nonzero only for\npositive temperatures T >0. Then we can apply the formula for variable-range hopping15\ninEq. (2), whichfitswell theexperimental result ingraphaneofRef . [7]. Theparameter\nT0is related to the DOS at the Fermi level as [14]\nkBT0∝1\nξ2ρ(EF), (42)\nwhereξis the localization length. T0has its maximum at the NP EF= 0, as shown in\nFig. 7 and decreases monotonically with increasing carrier density, a s in the experiment\non graphane [7].\nInconclusion, wehavestudiedthedensityofstatesinMLGandBLGa tlowenergies\nin the presence of a random symmetry-breaking potential. While a un iform symmetry-\nbreaking potential opens a uniform gap, a random symmetry-brea king potential also\ncreates tails in the density of states. The latter can close the gap a gain, preventing the\nsystem to become an insulator at the nodes. However, for a sufficie ntly large gap the\ntails contain localized states with nonzero density of states. These localized states allow\nthe system to conduct at nonzero temperature via variable-rang e hopping. This result\nis in agreement with recent experimental observations [7].\nAcknowledgement: This work was supported by a grant from the Deutsche\nForschungsgemeinschaft and by the Hungarian Scientific Researc h Fund under grant\nnumber K72613.\n[1] K.S. Novoselov, A.K. Geim, S.V. Morozov, D. Jiang, M.I. Katsnelson , I.V. Grigorieva, S.V.\nDubonos, A.A. Firsov, Nature 438, 197 (2005)\n[2] Y. Zhang, Y.-W. Tan, H.L. Stormer, P. Kim, Nature 438, 201 (2005)\n[3] A.K. Geim and K.S. Novoselov, Nature Materials, 6, 183 (2007)\n[4] Y.-W. Tan, Y. Zhang, K. Bolotin, Y. Zhao, S. Adam, E.H. Hwang, S. Das Sarma, H.L. Stormer,\nP. Kim, Phys. Rev. Lett. 99, 246803 (2007)\n[5] J.H. Chen, C. Jang, M.S. Fuhrer, E.D. Williams, M. Ishigami, Nature P hysics4, 377 (2008)\n[6] S.V. Morozov, K.S. Novoselov, M.I. Katsnelson, F. Schedin, D.C. E lias, J.A. Jaszczak, A.K. Geim,\nPhys. Rev. Lett. 100, 016602 (2008)\n[7] D.C. Elias, R.R. Nair, T.M.G. Mohiuddin, S.V.Morozov, P. Blake, M.P.H alsa ll, A.C. Ferrari,\nD.W. Boukhvalov, M.I. Katsnelson, A.K. Geim, K.S. and Novoselov, Scie nce323, 610 (2009)\n[8] O. Taisuke, A. Bostwick, T. Seyller, K. Horn, E. Rotenberg, Scie nce18, Vol. 313, 951\n[9] J.B. Oostinga, H.B. Heersche, X. Liu, A.F. Morpurgo, L.M.K. Vande rsypen, Nature Materials 7,\n151 (2008)\n[10] R.V. Gorbachev, F.V. Tikhonenkoa, A.S. Mayorova, D.W. Horsella and A.K. Savchenkoa, Physica\nE40, 1360 (2008)\n[11] E.V. Castro, N.M.R. Peres, J.M.B. Lopes dos Santos, F. Guinea, a nd A.H. Castro Neto, J. Phys.:\nConf. Ser. 129 012002 (2008)\n[12] G. Li, A. Luican, E.Y. Andrei, Phys. Rev. Lett. 102, 176804 (2009)\n[13] N.F. Mott, Metal-Insulator Transitions , (Taylor & Francis, London, 1990)\n[14] N.F. Mott, Philos. Mag. 19, 835 (1969)\n[15] S.V. Morozov et al., Phys. Rev. Lett. 97, 016801 (2006)\n[16] J.C. Meyer et al., Nature 446, 60 (2007)\n[17] A.H. Castro Neto, F. Guinea, N.M.R. Peres, K.S. Novoselov, and A .K. Geim, Rev. Mod. Phys.\n81, 109 (2009)\n[18] E. McCann and V.I. Fal’ko, Phys. Rev. Lett. 96, 086805 (2006); E. McCann, Phys. Rev. B 74,\n161403(R) (2006)\n[19] E. McCann et al., Phys. Rev. Lett. 97, 146805 (2006)16\n[20] P.W. Anderson, Phys. Rev. 109, 1492 (1958)\n[21] K. Ziegler, Phys. Rev. B 55, 10661 (1997)\n[22] K. Ziegler, Phys. Rev. Lett. 102, 126802 (2009); arXiv:0903.0740\n[23] H. Suzuura and T. Ando, Phys. Rev. Lett. 89, 266603 (2002)\n[24] N.M.R. Peres, F. Guinea, and A.H. Castro Neto, Phys. Rev. B 73, 125411 (2006)\n[25] M. Koshino and T. Ando, Phys. Rev. B 73, 245403 (2006)\n[26] K. Ziegler, J. Phys. A 18, L801 (1985)\n[27] S. Villain-Guillot, G. Jug, K. Ziegler, Ann. Phys. 9, 27 (2000)\n[28] K. Ziegler, B. D´ ora, P. Thalmeier, arXiv:0812.2790\n[29] J. L. Cardy, J. Phys. C: Solid State Phys. 11, L321 (1978)." }, { "title": "0905.3207v1.Density_functional_theory_of_two_component_Bose_gases_in_one_dimensional_harmonic_traps.pdf", "content": "arXiv:0905.3207v1 [cond-mat.str-el] 20 May 2009Density-functional theory of two-component Bose gases in o ne-dimensional harmonic\ntraps\nYajiang Hao1and Shu Chen2,∗\n1Department of Physics, University of Science and Technolog y Beijing, Beijing 100083, China\n2Beijing National Laboratory for Condensed Matter Physics,\nInstitute of Physics, Chinese Academy of Sciences, Beijing 100190, China\n(Dated: November 13, 2018)\nWe investigate the ground-state properties of two-compone nt Bose gases confined in one-\ndimensional harmonic traps in the scheme of density-functi onal theory. The density-functional\ncalculations employ a Bethe-ansatz-based local-density a pproximation for the correlation energy,\nwhich accounts for the correlation effect properly in the ful l physical regime. For the binary Bose\nmixture with spin-independent interaction, the homogeneo us reference system is exactly solvable\nby the Bethe-ansatz method. Within the local-density appro ximation, we determine the density\ndistribution of each component and study its evolution from Bose distributions to Fermi-like dis-\ntribution with the increase in interaction. For the binary m ixture of Tonks-Girardeau gases with\na tunable inter-species repulsion, with a generalized Bose -Fermi transformation we show that the\nBose mixture can be mapped into a two-component Fermi gas, wh ich corresponds to exact soluble\nYang-Gaudin model for the homogeneous system. Based on the g round-state energy function of the\nYang-Gaudin model, the ground-state density distribution s are calculated for various inter-species\ninteractions. It is shown that with the increase in inter-sp ecies interaction, the system exhibits\ncomposite-fermionization crossover.\nPACS numbers: 67.85.-d, 67.60.Bc, 03.75.Mn\nI. INTRODUCTION\nExperimental realization of the atomic mixtures and\nprogress in manipulating cold atom systems have opened\nexciting new possibilities to study many-body physics of\nlow-dimensional quantum gases beyond the mean-field\ntheory. In comparison with the single-component sys-\ntems, mixtures of quantum degenerate atoms may form\nnovel quantum many-body systems with richer phase\nstructures. Particularly, two-component Bose gases have\nbeen under intensive studies both experimentally [1, 2,\n3, 4, 5] and theoretically [6, 7, 8, 9, 10]. Due to the com-\npetition among the inter-species and intra-species inter-\nactions, many interesting phenomena such as phase sep-\naration, new quantum states and phase transitions arise\nin two-component cold atomic gases [6, 7, 8, 9, 10]. With\nstrong anisotropic magnetic trap or two-dimensional op-\ntical lattice the radial degrees of freedom of cold atoms\nare frozen and the quantum gas is described by an ef-\nfective one-dimensional (1D) model [11, 12, 13]. As\na textbook example the 1D Tonks-Girardeau (TG) gas\n[14] was observed firstly [12, 13]. Since then, not only\nthe single-component bosonic gas but also 1D multi-\ncomponent atomic mixtures have become the current re-\nsearch focus. In principle, both the intra-component and\ninter-component interactions can be tuned via the mag-\nnetic Feshbach resonance, which allows us to study the\nbosonic mixture in the whole interaction regime. Very\nrecently, two-species Bose gases with tunable interaction\n∗Electronic address: schen@aphy.iphy.ac.cnhave been successfully produced [4, 5].\nIn general, 1D quantum systems exhibit fascinating\nphysics significantly different from its three-dimensional\ncounterpart because of the enhanced quantum fluctua-\ntions in 1D[15, 16, 17, 18, 19, 20]. The mean-field theory\ngenerallyworksnotwellforthe1Dsystemsexceptinvery\nweakly interacting regime. When interaction is strong\nenough, non-perturbation methods, for instance, Bose-\nFermi mapping (BFM), Bethe ansatz and Bosonization\nmethod, have to be exerted to properly characterize the\nfeaturesofthesystem[9,21, 22, 23, 24, 25, 26,27, 28, 29].\nFor a single-component Bose gas, it has been shown that\nthe density profiles continuously evolve from Gaussian-\nlike distribution of bosons to shell-structured distribu-\ntion of fermions [30, 31, 32, 33] with the increase in re-\npulsion strength. Similar behavior of composite fermion-\nization has been observed in the 1D Bose-Bose mixtures\n[27, 28, 29]. For two-component Bose mixture with spin-\nindependent repulsions, the normalized density distribu-\ntion of each component displays the same behavior up to\na normalized constant which is proportional to respec-\ntive atomic number of each component [27]. While the\nextended Bose-Fermi mappings [21, 22, 23] are restricted\nto the infinitely repulsive limit, some sophisticate meth-\nods, such as multi-configuration self-consistent Hartree\nmethod [28], numerical diagonalization method [29], and\nBethe ansatz method [24, 25, 26, 27], have been applied\nto study the two-component Bose gases in the whole re-\npulsive regime. Nevertheless, the above methods either\nsuffer the small-size restriction [28, 29] or only limit to\nthe integrable models [24, 25, 26, 27] which generally\ndo not cover the realistic system confined in an external\nharmonic trap.2\nIn this work, we use the basic idea of density func-\ntionaltheory(DFT)toinvestigatetheground-stateprop-\nerties of the two-component Bose gases trapped in har-\nmonic potential in the full interacting regime. It is well\nknown that the DFT can handle the system with large\nsizes and will not encounter the exponential wall that\nthe above traditional multiparticle wave-function meth-\nods faced. The effects ofthe externalpotential will be ac-\ncounted by using the local-density approximation (LDA)\n[19, 34, 35, 36, 37, 38, 39, 40, 41, 42]. The combinationof\nDFTandLDAhadbeenappliedtodealwiththeconfined\n1D single-component Bose gas [19, 34, 35, 36] and Fermi\ngas [37, 38, 39, 40, 41, 42] successfully. However, the\ntwo-component Bose gas has rarely been studied. The\npresentworkwillstudytheground-statepropertiesofthe\nconfined Bose mixture and focus on two cases where the\ngroundstateenergyfunctionforthehomogeneoussystem\ncan be obtained exactly. In the first case, we consider the\ntwo-component Bose gas with spin-independent interac-\ntion. In the absence of external potential, this model is\nexactlysolvablebytheBethe-ansatzmethodandthusthe\nground-state energy density function can be extracted\nfromtheBethe-ansatzsolution[24, 25,26, 27]. Inthesec-\nond case, we consider the two-component Bose gas with\ninfinitely repulsive intra-component interactions and a\ntunable inter-component interaction. We will show that\nsuchamodelcanbemappedintotheYang-Gaudinmodel\nwith generalizedBose-Fermi transformationand thus the\nground-state energy function can be obtained from the\nBethe-ansatz solution of the Yang-Gaudin model [43]. In\nthe scheme of DFT, the ground-state density profiles are\nobtained by numerically solving the coupled nonlinear\nSchr¨ odinger equations (NLEs) for both the equal-mixing\nBose mixture and polarized Bose mixture.\nThe present paper is organized as follows. Sec-\ntion II investigates two-component Bose gas with spin-\nindependent interaction and introduce the method. Sec-\ntion III is devoted to the mixture of two TG gases with\ntunable inter-species repulsion. A summary is given in\nthe last section.\nII. BINARY BOSE MIXTURE WITH\nSPIN-INDEPENDENT INTERACTION\nWe consider a two-component Bose gas confined in the\nexternal potential Vext(x) independent of the species of\natoms, where the atomic mass of the i-component is mi.\nTheatomnumbersineachcomponentare N1andN2and\nN=N1+N2is the total atom number. The many-body\nHamiltonian in second quantized form can be formulatedas\nH=/integraldisplay\ndx/summationdisplay\ni=1,2/braceleftbigg\nˆΨ†\ni(x)/bracketleftbigg\n−¯h2\n2mi∂2\n∂x2+Vext(x)/bracketrightbigg\nˆΨi(x)\n+gi\n2ˆΨ†\ni(x)ˆΨ†\ni(x)ˆΨi(x)ˆΨi(x)/bracerightBig\n+g12/integraldisplay\ndxˆΨ†\n1(x)ˆΨ†\n2(x)ˆΨ2(x)ˆΨ1(x), (1)\nin which gi(i= 1,2) andg12denote the effective\nintra- and inter-species interaction that can be controlled\nexperimentally by tuning the corresponding scattering\nlengthes a1,a2anda12, respectively [4, 5]. Ψ†\ni(x)\n(Ψi(x)) is the field operator creating (annihilating) an i-\ncomponent atom at position x. Here we will consider the\ntwo-component Bose gas composed of two internal states\nofsamespeciesofatomssuchthatwehavethesamemass\nfor all atoms m1=m2=m. When the intra- and inter-\ncomponent interactions are equal ( g1=g2=g12=g),\nthe system is governed by the spin-independent Hamil-\ntonian with the first quantized form:\nH=N/summationdisplay\nj=1/bracketleftBigg\n−∂2\n∂x2\nj+Vext(xj)/bracketrightBigg\n+2c/summationdisplay\nj∆xt\ni, a vehicle must stop at the cell\nxt\ni+1−1.\nIn what follows, we assume that the maximum allowed\nvelocity is equal to 1, i.e., if m/greaterorequalslant2,wt\ni(m) = 0. Then,2\n✇/archrightdownp\n✇/archrightdownp\n✇✇\n(a)ASEP✇/archrightdownp(3)\n✇/archrightdownp(1)\n✇✇\n(b)ZRP\nFIG. 1: Schematic view of the tagged-particle model for\nASEP(a) and ZRP(b). The hopping probability of a parti-\ncle depends on the gap size in front of it in ZRP, while it is\nalways constant in ASEP. In both cases, hopping to an occu-\npied cell is prohibited.\nputtingvt\ni≡wt\ni(1) (accordingly wt\ni(0) = 1 −vt\ni), we\npropose a special type of fiin (1) as\nvt+1\ni= (1−ai)vt\ni+aiVi(∆xt\ni) (∀t≥0,∀i),(3)\nwhereai(0≤ai≤1) is a parameter and the function\nViis restricted to values in the interval [0 ,1] so that vt\ni\nalso should be within [0 ,1]. The intrinsic parameter ai,\na weighting factor of the optimal velocity Vi(∆xt\ni) to the\nintention wt+1\ni,correspondstothedriver’ssensitivitytoa\ntraffic condition. As long as the vehiclesmove separately,\nwe can also rewrite (2) simply as xt+1\ni=xt\ni+ 1 with\nprobability vt+1\ni. (Note that the original OV model does\nnot support the hard-core exclusion [8, 9].) Therefore,\nvt\nican be regarded as the average velocity, i.e., /an}bracketle{txt+1\ni/an}bracketri}ht=\n/an}bracketle{txt\ni/an}bracketri}ht+vt+1\niin the sense of expectation values. We call\nthe model expressedby (3) the stochastic optimal velocity\n(SOV) model because ofits formal similarityto a discrete\nversion of the OV model (or a coupled map lattice[11])\nxt+1\ni=xt\ni+vt+1\ni∆t, (4)\nvt+1\ni=(1−a∆t)vt\ni+(a∆t)V(∆xt\ni),(5)\nwhere ∆ tis a time interval and the OV model is recov-\nered in the limit ∆ t→0. In this special case that the\nmaximum allowed velocity equals to 1, we thus have an\nobvious correspondence of our stochastic CA model to\nan existent traffic model. As a matter of convenience, we\nsetai=aandVi=V(∀i) hereafter.\nFrom the viewpoint of mathematical interest, the SOV\nmodel includes two significant stochastic models. When\na= 0, (3) becomes vt+1\ni=vt\ni, i.e., the model reduces to\nASEP [12, 13, 14, 15] with a constant hopping probabil-\nityp≡v0\ni. When a= 1, (3) becomes vt+1\ni=V(∆xt\ni),\ni.e., the model reduces to ZRP [16] considering the head-\nways{∆xt\ni}as the stochastic variables of ZRP. In ZRP,\nthe hopping probability of a vehicle is determined ex-\nclusively by its present headway. Fig.1 illustrates the\ntwo stochastic models schematically. ASEP and ZRP are\nboth known to be exactly solvable in the sense that the\nprobability distribution of the configuration of vehicles\nin the stationary state can be exactly calculated [17, 18],\nand thus our model admits an exact calculation of the\nfundamental diagram in the special cases.\nIn order to investigate a phenomenological feature, we\ntake a realistic form of the OV function as\nV(x) =tanh(x−c)+tanh c\n1+tanh c, (6)00.050.10.150.2\n00.20.40.60.8100.050.10.150.2\n00.20.40.60.81t=10 t=100\nt=1000 t=10000\nDensityFlux\nFIG. 2: The fundamental diagram of the SOV model with the\nOV function (6) (c=1 .5 anda= 0.01) plotted at each time\nstaget, starting from uniform/random states with p(≡v0\ni) =\n0.5, including the exact curve(gray) of ASEP for comparison\n[13]. The system size is L= 1000.\nwhich was investigated in [8]. We find that the funda-\nmental diagrams simulated with (6) have a quantitative\nagreement with the exact calculation of ZRP ( a= 1)\nup toa∼0.6. In contrast, the fundamental diagram of\nthe SOV model does not come closer to that of ASEP\nasaapproaches to 0 although the SOV model coincides\nwith ASEP at a= 0. Fig. 2 shows that a curve sim-\nilar to the diagram of ASEP appears only for the first\nfew steps ( t= 10) and then changes the shape rapidly\n(t= 100,1000). Surprisingly, when the diagram be-\ncomes stationary, it allows a discontinuous point and two\noverlapping stable states around the density ρ∼0.14\n(t= 10000).\nLet us study the discontinuity of the flux in detail.\nFig. 3 shows the fundamental diagram expanded around\nthe discontinuous point. We have three distinct branches\nand then call them as follows: free-flow phase (vehicles\nmove without interactions), congested phase (a mixture\nof small clusters and free vehicles), and jam phase (one\nbig stable jam transmitting backward). These branches\nappearinthefundamentaldiagram,respectivelyasaseg-\nment of line with slope 1(free-flow), as a thick curve with\naslightpositiveslope(congested), andasathickline with\na negative slope(jam). Note that the congested and jam\nlines shows some fluctuations due to a randomness of\nthe SOV model. Fig. 3(a) shows snapshots of the flux\natt= 1000 and 5000. There exists midstream flux be-\ntween free-flow and congested phases at t= 1000, and\nbetween congested and jam phases at t= 5000. This\nsuggests that the high-density free-flow states can hold\nuntilt= 1000 but not until t= 5000, and that the\nhigh-density congested states have already started to de-\ncay into jam states before t= 5000. We have thereby\nrevealed the existence of two metastable branches lead-\ning out of the free-flow or congested lines. Comparing\nFig.3(a) with Fig. 3(b), we have six qualitatively dis-\ntinct regions of density(Fig.3(b)); free region S1(includ-3\n00.050.10.150.20.250.3\nDensity0.020.040.060.080.10.120.140.16Flux(a)\nFreeflow\nCongested\nJamS1B1T1T2B20.020.040.060.080.10.120.140.16Flux\n(b)\nFIG. 3: (a)The expanded fundamental diagram of the SOV\nmodel with a= 0.01 att= 1000 (gray) and t= 5000 (black)\nstarting from two typical states; the uniform state with equ al\nspacing of vehicles and p(≡v0\ni) = 1, and the random state\nwith random spacing and p= 1. We observe three distinct\nbranches, which we call the free-flow, congested, and jam\nbranch. They survive even in the stationary state, which is\nplotted at t= 50000 in (b). (b)The stationary states (black)\nand the averaged three branches (gray lines) are plotted at\nt= 50000. The vertical dotted lines distinguish the regions o f\ndensity from S1toB2. The arrows in T2indicate the trace of\na metastable free-flow state decaying to the lower branches.\n(see also Fig. 4).\ning only free-flow phase), bistable region B1(including\nfree-flow and congested phases, which are both stable),\ntristable region T1(including all the three phases, which\nare all stable), tristable region T2(including free-flow,\ncongested and jam phases. The former two phases are\nstable and the last one is metastable), bistable region\nB2(including congested and jam phases. The former is\nmetastable, and the latter is stable), and jam region S2\n(including only stable jam phase, which is not displayed\nhere). We stress that the tristable region T1is a novel\nand remarkable characteristic of many-particle systems,\nand that the successive phase transitions from a free-flow\nstate to a jam state via a congested state occur respec-\ntively on a short time scale.\nLet us study the dynamical phase transition especially\nat the density ρ= 0.14, indicated by successive arrowsin\nFig. 3(b). Fig. 4(upper) shows the flux plotted against\ntime. It is striking that there appear three plateaus\nwhich respectively correspond to a free-flow state, a con-\ngested state, and a jam state, and that the flux changes\nsharply from one plateau to another. In other words, the\nmetastable states have a remarkably long lifetime before\nundergoing a sudden phase transition. Stochastic mod-\nFIG. 4: The time evolution of flux at the density ρ= 0.14\nstarting from the uniform state. We observe two plateaus at\nthefluxQ= 0.14 with alifetime T≃5000, and Q≃0.08 with\nT≃7000 before reaching the stationary jam state(upper).\nThe lower figure shows the corresponding spatio-temporal di -\nagram, where vehicles (black dots) move from bottom up.\n(Note that the periodic boundary condition is imposed.)\nels, in general, arenot anticipated to havesuch long-lived\nmetastable states because stochastic fluctuations break a\nstability of states very soon [7]. Fig. 4(lower) shows the\nspatio-temporal diagram corresponding to the dynami-\ncal phase transition. Starting from a free-flow state, the\nuniform configuration stochastically breaks down at time\nt∼5000, and then the free-flow state is rapidly replaced\nby a congested state where a lot of clusters are forming\nand dissolving, moving forward and backward. Fig. 5\nshowsthe distributionofheadwaysatseveraltimestages.\nWe find that the distribution of headways changes signif-\nicantly after each phase transition. In particular, the\nratio of vehicles with null headway increases. As for a\ncongested state, it is meaningful to evaluate the average\nsize of clusters from a distribution of headways. If the\nratio of the vehicles with null headway is b0, the average\nclustersize ℓcanbeevaluatedas1 /(1−b0). Inthepresent\ncase, the average size of clusters is about 1 .2∼1.4 dur-\ningt= 6000∼12000. We also have some remarks, from\na microscopic viewpoint, on a single cluster: Since the\nsensitivity parameter ais set to a small value, the inten-\ntionvt\nidoes not change a lot before the vehicle catches\nup with the tail of a cluster(i.e. vt\ni∼1). Accordingly,\nthe aggregation rate αis roughly estimated at the den-\nsity of free region behind the cluster; α∼(1−b0)ρ. In\nthe present case, the aggregation rate averaged over the\nwhole clustersofa congestedstate is 0 .10∼0.12. For the\nsame reason, we can estimate the average velocity of the\nfront vehicle of a cluster at (1 −a)ℓ/δ, whereδdenotes\nthedissolutionrateand ℓ/δindicatesthe durationofcap-4\nt=40000.6\n0.5\n0.4\n0.3\n0.2\n0.1\n0t=6000\n0510 15 202530\nHeadwayt=14000\n0510 15 2025300.6\n0.5\n0.4\n0.3\n0.2\n0.1\n0Ratiot=12000\nFIG. 5: The distribution of the headways with which the\nvehicles move at each time stage t. It changes a lot after\nphase transitions( t∼5000,12000). The traffic states, free-\nflow(t <5000), congested(5000 < t <12000), and jam( t >\n12000) respectively show their specific pictures.\nture. Since δis also equivalent to the average velocity of\nthe front vehicle, it amounts roughly to 1 −aℓafter all.\nIn the present case, the dissolution rateaveragedoverthe\nwhole clusters of a congested state is 0 .86∼0.88. Then,\nthe average lifetime of the clusters ℓ/(δ−α) is estimated\nat 1.54∼1.89. The above estimations are appropriate\nonly when clusters are small and spaced apart. As many\nclusters arise everywhere and gather, a vehicle out of a\ncluster tends to be caught in another cluster again be-\nfore it recovers its intention at full value. Consequently,the clusters arising nearby reduce their dissolution rates,\nand finally grow into a jam moving steadily backward.\nThe steady transmitting velocity (the aggregation rate)\nis estimated at −0.055, coinciding with Fig. 4.\nIn this paper, beginning with a general scheme, we\nhave proposed a stochastic CA model to which we in-\ntroduce a probability distribution function of the vehi-\ncle’s velocity. It includes two exactly solvable stochastic\nprocesses, and it is also regarded as a stochastic gen-\neralization of the OV model. Moreover, it exhibits the\nfollowing features: In spite of a stochastic model, the\nfundamental diagram shows that there coexist two or\nthree stable phases (free-flow, congested, and jam) in\na region of density. As the density increases, the free-\nflow and congested states lose stability and change into\nmetastable states which can be observed only for a tran-\nsitional period. Moreoverthe dynamical phase transition\nfrom a metastable state to another metastable/stable\nstate, which is triggered by stochastic perturbation, oc-\ncurs sharply and spontaneously. We consider that the\nmetastable state may be relevant to the transient con-\ngested state observed in the upstream of on-ramp [5, 19].\nSuch a dynamical phase transition has not been observed\nin previous works [8, 9, 10] or among existent many-\nparticle systems. Further studies, e.g., on another choice\noftheOVfunction, underopenboundaryconditions, and\non the general (i.e. multi-velocity) version will be given\nin subsequent publications [20].\n[1] D. Helbing, Rev. Mod. Phys. 73, 1067 (2001).\n[2] D. Chowdhury, L. Santen and A. Schadschneider, Phys.\nRep.329, 199 (2000).\n[3] T. Nagatani, Rep. Prog. Phys. 65, 1331 (2002).\n[4] K. Nishinari and M. Hayashi, Traffic statistics in Tomei\nexpress way , (The Mathematical Society of Traffic Flow,\n1999, Nagoya).\n[5] B. S. Kerner, The Physics of Traffic , Springer-Verlag,\nBerlin, New York 2004.\n[6] M. Treiber, A. Hennecke and D. Helbing, Phys. Rev. E,\n62, 1805 (2000).\n[7] K. Nishinari, M. Fukui and A. Schadschneider, J. Phys.\nA37, 3101 (2004).\n[8] M. Bando, K. Hasebe, A. Nakayama, A. Shibata and\nY. Sugiyama, Phys. Rev. E 51, 1035 (1995).\n[9] M. Bando, K. Hasebe, K. Nakanishi, A. Nakayama,\nA. Shibata and Y. Sugiyama, J. Phys. I France 5, 1389\n(1995).\n[10] D. Helbing and M. Schreckenberg, Phys. Rev. E 59,R2505 (1999).\n[11] S. Yukawa and M. Kikuchi, J. Phys. Soc. Jpn. 64, 35\n(1995).\n[12] B. Derrida, E. Domany, and D. Mukamel, J. Stat. Phys.\n69, 667 (1992).\n[13] B. Derrida, M. R. Evans, V.Hakim, and V.Pasquier, J.\nPhys. A 26, 1493 (1993).\n[14] N. Rajewsky, L. Santen, A. Schadschneider, and M.\nSchreckenberg, J. Stat. Phys. 92, 151 (1998).\n[15] G. M. Sh¨ utz, J. Phys. A 36, R339 (2003).\n[16] F. Spitzer, Adv. Math. 5, 246 (1970).\n[17] M. R. Evans, J. Phys. A 30, 5669 (1997).\n[18] O. J. O’Loan, M. R. Evans, and M. E. Cates, Phys. Rev.\nE58, 1404 (1998).\n[19] N. Mitarai and H. Nakanishi, Phys. Rev. Lett. 85, 1766\n(2000).\n[20] M. Kanai, K. Nishinari and T. Tokihiro, to be published." }, { "title": "0905.4229v1.High_Precision_Thermodynamics_and_Hagedorn_Density_of_States.pdf", "content": "arXiv:0905.4229v1 [hep-lat] 26 May 2009MIT-CTP 4041\nHigh-Precision Thermodynamics and Hagedorn Density of Sta tes\nHarvey B. Meyer\nCenter for Theoretical Physics\nMassachusetts Institute of Technology\nCambridge, MA 02139, U.S.A.\n(Dated: May 26, 2009)\nWe compute the entropy density of the confined phase of QCD wit hout quarks on the lattice to\nvery high accuracy. The results are compared to the entropy d ensity of free glueballs, where we\ninclude all the known glueball states below the two-particl e threshold. We find that an excellent,\nparameter-free description of the entropy density between 0.7TcandTcis obtained by extending\nthe spectrum with the exponential spectrum of the closed bos onic string.\nPACS numbers: 12.38.Gc, 12.38.Mh, 25.75.-q\nI. INTRODUCTION\nThe phase diagram of quantum chromodynamics\n(QCD) is being actively studied in heavy ion collision\nexperiments as well as theoretically. A form of mat-\nter with remarkable properties [1] has been observed\nin the Relativistic Heavy Ion Collider (RHIC) experi-\nments [2, 3, 4, 5]. It appears to be a strongly coupled\nplasma of quarks and gluons (QGP), but no consensus\non a physical picture that accounts for both equilibrium\nand non-equilibrium properties has been reached yet. On\nthe other hand, below the short interval of temperatures\nwhere the transition from the confined phase to the QGP\ntakes place [6, 7, 8, 9], it is widely believed that the most\nprominent degrees of freedom are the ordinary hadrons.\nFrom this point of view, the zeroth order approximation\nto the properties of the system is to treat the hadrons\nas infinitely narrow and non-interacting. We will refer to\nthis approximation as the hadron resonance gas model\n(HRG). The HRG predictions were compared with lat-\ntice QCD thermodynamics data in [6, 10], and lately they\nhave been used to extrapolate certain results to zero tem-\nperature [7]. The HRG is also the basis of the statistical\nmodel currently applied to the analysis of hadron yields\nin heavy ion collisions [11], and recently the transport\nproperties of a relativistic hadron gas have been studied\nin detail [12].\nSince any heavy ion reaction ends up in the low-\ntemperature phase of QCD, it is important to under-\nstand its properties in detail in order to extract those of\nthe high-temperature phase with minimal uncertainty. In\nthis Letter we study whether the HRG model works in the\nabsence of quarks, in other words in the pure SU( N= 3)\ngauge theory, where the low-lying states are glueballs.\nThere are reasons to believe that if the HRG model is\nto work at any quark content of QCD, it is in the zero-\nflavor case. Firstly, the mass gap in SU(3) gauge theory\nis very large, M0/Tc≃5.3. As we shall see, the thermo-\ndynamic properties up to quite close to Tcare dominated\nby the states below the two-particle threshold, which\nare exactly stable. Furthermore, because of their large\nmass, neglecting their thermal width should be a good\napproximation. Secondly, the scattering amplitudes be-tween glueballs are parametrically 1 /N2suppressed while\nthose between mesons are only 1 /Nsuppressed [13]. This\nmeans that the glueballs should be free to a better ap-\nproximation than the hadrons of realistic QCD.\nAn additional motivation to study the thermodyna-\nmics of the confined phase of SU(3) gauge theory is that it\nis a parameter-free theory, simplifying the interpretatio n\nof its properties. Its spectrum is known quite accurately\nup to the two-particle threshold [14, 15]. By contrast,\nin full QCD calculations, lattice data calculated at pion\nmasses larger than in Nature are often compared out of\nnecessity to the HRG model based on the experimental\nspectrum [6, 7]. Finally, calculations in the pure gauge\ntheory are at least two orders of magnitude faster, which\nallows us to reach a high level of control of statistical and\nsystematic errors; in particular, we are able to perform\ncalculations in very large volumes.\nII. LATTICE CALCULATION\nWe use Monte-Carlo simulations of the Wilson action\nSg=1\ng2\n0/summationtext\nx,µ,νTr{1−Pµν(x)}for SU(3) gauge the-\nory [16], where Pµνis the plaquette. The lattice spacing\nis related to the bare coupling through g2\n0∼1/log(1/aΛ).\nWe calculate the thermal expectation value of θ≡Tµµ,\nthe (anomalous) trace of the energy-momentum tensor\nTµν, and of θ00≡T00−1\n4θ. In the thermodynamic limit,\nTs=e+p=4\n3/an}bracketle{tθ00/an}bracketri}htT, e−3p=/an}bracketle{tθ/an}bracketri}htT− /an}bracketle{tθ/an}bracketri}ht0.(1)\nHeree, p, s are respectively the energy density, pressure\nand entropy density. The operator θ00=1\n2(−Ea·Ea+\nBa·Ba) requires no subtraction, because its vacuum\nexpectation value vanishes. The choice of of θ00and\nθas independent linear combinations is convenient be-\ncause they both renormalize multiplicatively. We use the\n‘HYP-clover’ discretization of the energy-momentum ten-\nsor introduced in [17, 18]. The normalization of the θ00\noperator differs from its naive value by a factor that we\nparametrize as Z(g0)χ(g0). The factor Z(g0) is taken\nfrom [19] and rests on the results of [20]; its accuracy\nis about one percent. The factor χ(g0) is obtained by2\n 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8\n 4 5 6 7 8 9 10s / T3\nL TFinite-volume effects at 0.985Tc and 0.929Tc\n8x643 6/go2 = 6.053\n12x803 6/go2 = 6.3238\n8x643 6/go2 = 6.018\n12x803 6/go2 = 6.2822\nAsymptot. vol. correction\n0.239 + 7.4 exp(-0.82 LT)\nFIG. 1: Finite volume effects on the entropy density close to\nthe deconfining temperature Tc.\ncalibrating our discretization to the ‘bare plaquette’ dis -\ncretization in the deconfined phase at Nt= 6 [17]. We\nfind, for 6 /g2\n0between 5.90 and 6.41, χ(g0) = 0 .1306·\n(6/g2\n0)−0.1865 with an accuracy of half a percent. For\nthe lattice beta-function that renormalizes θ, we use the\nparametrization [21] of the data in [22] and the same ca-\nlibration method.\nOur results for the entropy density from Nt= 8 and\nNt= 12 simulations are shown on Fig. (3). The displayed\nerror bars do not contain the uncertainty on the normal-\nization factor, which is much smaller and would introduce\ncorrelation between the points. This factor varies by only\n7% over the displayed interval and so to a first approxi-\nmation amounts to an overall normalization of the curve.\nOur data is about five times statistically more accurate\nthan that of previous thermodynamic studies [23, 24],\nwhich were primarily focused on the deconfined phase.\nJust as importantly, we kept the finite-spatial-volume ef-\nfects under good control, in particular very close to Tc.\nFigure (1) shows the size of finite-volume effects. For\ninstance, at 0 .985Tcthe conventional choice LT= 4 leads\nto an overestimate of the entropy density by a factor\nthree. The fact that the Nt= 12 data fall on the same\nsmooth curve as the Nt= 8 is strong evidence that dis-\ncretization errors are small. We parametrize the volume\ndependence empirically by a A+Be−cLTcurve, and use it\nto convert the Nt= 12 data to LT= 8. At 0 .929Tc, there\nis no statistically significant difference between LT= 6\nand 8 and we do not apply any correction. It is the cor-\nrected Nt= 12 data that is then displayed on Fig. (3).\nIn [25], formulas for the leading finite-volume effects\non the thermodynamic potentials were derived in terms\nof the energy gap of the theory defined on a (1 /T)×L×L\nspatial hypertorus. Close to Tc, this gap corresponds to\nthe mass of the ground state flux loop winding around\nthe cycle of length 1 /T. Ifδs(T, L)≡s(T,∞)−s(T, L), 0 1 2 3 4 5 6\n 0 0.2 0.4 0.6 0.8 1(m(T) T / Tc2)2 \n(T / Tc)2Flux-Loop Mass (Nc=3, Nf=0)\nNt > 11\nNt = 8\nNt = 6\nNt = 5\nFit\nFIG. 2: The mass of the temporal flux loop as calculated from\nPolyakov loop correlators, and the fit (5). The Nt>11 data\nare from [26], the Nt= 5 data from [27].\nthe formula then reads\nδs(T, L) =e−m(T)L\n2πL/bracketleftbig\nm2(T) +3\n2T∂Tm2(T)/bracketrightbig\n.(2)\nUsing the calculation of m(T) described in the next sec-\ntion, the predicted asymptotic approach to the infinite-\nvolume entropy density for 0 .985Tcis displayed on\nFig. (1). While the sign is correct, the magnitude of the\nfinite-volume effects is not reproduced for LT≤8. We\nconclude that the asymptotic approach to infinite volume\nsets in for very large values of LT. Since m(T)Lis only\nabout 4 when LT= 6, it is not implausible that flux-loop\nstates with high multiplicity dominate the finite-volume\neffects at that box size.\nNext we obtain the correlation length ξ(T) of the\norder parameter for the deconfining phase transition,\nthe Polyakov loop. The method consists in computing\nthe two-point function of zero-momentum operators, de-\nsigned to have large overlaps with the ground state flux\nloop, along a spatial direction. We fit the lattice data for\nm(T)≡1/ξ(T) displayed on Fig. (2) with the formula\n/parenleftBig\nm(T)T\nT2c/parenrightBig2\n=a0−a1/parenleftBig\nT\nTc/parenrightBig2\n−a2/parenleftBig\nT\nTc/parenrightBig4\n(3)\nand find, either fitting a2or setting it to zero,\na0= 5.76(15) , a1= 4.97(65) , a2= 0.55(54) (4)\na0= 5.90(9), a1= 5.62(10) , a2= 0 (5)\nwith in both cases a χ2/dof of about 0.3. We remark\nthat the aiare not far from the Nambu-Goto string [28]\nvalues a1=2π\n3σ\nT2\nc= 5.02(5) [27] and a2= 0 (σis the ten-\nsion of the confining string). We extract the ‘Hagedorn’\ntemperature, defined as in [29] by m(Th) = 0, from the\nsecond fit,\nTh/Tc= 1.024(3) . (6)3\nThis extraction amounts to assuming mean-field expo-\nnents near Th(it is not clear which universality class\nshould be used [30]). The result is stable if the fit in-\nterval is varied, and also if a2is fitted with a0anda1\nconstrained to the known values of ( σ/T2\nc)2and2π\n3σ\nT2c.\nAs a check on the normalization of the operators θ00\nandθ, we calculate the latent heat in two different ways.\nThe latent heat is the jump in energy density at Tc. Since\nthe pressure is continuous, we obtain it instead from the\ndiscontinuity in entropy density or the ‘conformality mea-\nsure’e−3p. We obtain sande−3pon either side of\nTcby extrapolating LT= 10 data from the confined (de-\nconfined) phase towards Tc. The result is\n∆s\nT3c= 1.45(5)(5) ,∆(e−3p)\nT4c= 1.39(4)(5) ,(7)\nwhere the first error is statistical and the second comes\nfrom the uncertainty in the extrapolation (taken to be\nthe difference between a linear and quadratic fit). The\ncompatibility between these two estimates of Lh/T4\ncis\nstrong evidence that we control the normalization of our\noperators. They are in good agreement with previous cal-\nculations of the latent heat performed on coarser lattices\n[27, 31]. We have also verified more generally that the\nthermodynamic identity T∂T(s/T3) = (1 /T3)∂T(e−3p)\nis satisfied within statistical errors.\nIII. INTERPRETATION\nIn infinite volume the pressure associated with a single\nnon-interacting, relativistic particle species of mass M\nwithnσpolarization states reads\np=nσ\n2π2M2T2∞/summationdisplay\nn=11\nn2K2(nM/T ) (8)\nwhere K2is a modified Bessel function. By linearity,\nthe knowledge of the glueball spectrum leads to a simple\nprediction for the pressure and entropy density s=∂p\n∂T,\nwhich is expected to become exact in the large- Nlimit.\nSince only the low-lying spectrum of glueballs is known,\nit is useful to consider how the density of states might be\nextended above the two-particle threshold 2 M0, where\nM0is the mass of the lightest (scalar) glueball. The\nasymptotic closed bosonic string density of states in four\ndimensions is given by [32]\nρ(M) =(2π)3\n27Th/parenleftbiggTh\nM/parenrightbigg4\neM/T h. (9)\nIn the string theory, the Hagedorn temperature This re-\nlated to the string tension, T2\nh=3σ\n2π, corresponding to\nTh/Tc= 1.069(5) [33]. Below we use this value as an\nalternative to the more direct determination (6).\nOn Fig. 3, we show the entropy contribution of the\nglueballs lying below the two-particle threshold 2 M0. 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4\n 0.7 0.75 0.8 0.85 0.9 0.95 1s / T3\nT / TcEntropy of the confined phase (Nc=3, Nf=0)\n8 x 643\n12 x 803\nGlueballs (<2M0) + Hagedorn (>2M0)\nidem, with Th2 = 3σ/2π\nGlueballs (<2M0)\n0++ & 2++ contribution\nFIG. 3: The entropy density in units of T3forLT= 8. We\napplied a (modest) volume-correction to the Nt= 12 data.\nThe curve is just about consistent with the smallest tem-\nperature lattice data point, but clearly fails to reproduce\nthe strong increase in entropy density as T→Tc. The\nfigure also illustrates that the two lowest-lying states, th e\nscalar and tensor glueballs, account for about three quar-\nters of the stable glueballs’ contribution. We have used\nthe continuum-extrapolated lattice spectrum [14, 34].\nAdding the Hagedorn spectrum contribution, Eq. (9)\nwithThgiven by Eq. (6), leads to the solid curve on\nFig. 3. It describes the direct calculation of the entropy\ndensity surprisingly well, particularly close to Tc. The\ncurve tends to underestimate somewhat the entropy den-\nsity at the lower temperatures. This is likely to be a cut-\noff effect. Indeed, at fixed Ntlower temperatures corre-\nspond to a coarser lattice spacing, and the scalar glueball\nmass in physical units is known to be smaller on coarse\nlattices with the Wilson action [35]. If we use the stable\nglueball spectrum calculated at g2\n0= 1 instead of the con-\ntinuum spectrum, the agreement of the non-interacting\nglueball + Hagedorn spectrum with the lattice data at\nthe lower four temperatures is again excellent. This dif-\nference provides an estimate for the size of lattice effects.\nTo summarize, we have computed to high accuracy the\nentropy of the confined phase of QCD without quarks.\nThe low-lying states of the theory are therefore bound\nstates called glueballs, and their spectrum is well deter-\nmined [14, 15]. If the size Nof the gauge group is in-\ncreased, the interactions of the glueballs are expected to\nbe suppressed [13]. To what extent the glueballs really\nare weakly interacting at N= 3 is not known precisely.\nSome evidence for the smallness of their low-energy in-\nteractions was found some time ago by looking at the\nfinite-volume effects on their masses [26]. But it seems\nunlikely that glueballs well above the two-particle thresh -\nold would have a small decay width. We have neverthe-\nless compared the entropy density data to the entropy\ndensity of a gas of non-interacting glueballs. While re-4\nstricting the spectral sum to the stable glueballs leads to\nan underestimate by at least a factor two of the entropy\ndensity near Tc, extending the spectral sum with an ex-\nponential spectrum ρ(M)∼exp(M/T h), suggested long\nago by Hagedorn [36], leads to a prediction in excellent\nagreement with the lattice data for the entropy density\n(Fig. 3). This is remarkable, since the analytic form of\nthe asymptotic spectrum is completely predicted by free\nbosonic string theory, including its overall normalizatio n\n(Eq. 9). Therefore, since we also separately computed\nthe temperature (identified with Th) where the flux loop\nmass vanishes, no parameter was fitted in the comparison\nwith the thermodynamic data. By contrast, the entropy\ndensity is not nearly as well described if the Nambu-Goto\nvalue of This used, see Fig. (3).\nThe success of the non-interacting string density of\nstates in reproducing the entropy density suggests that\nonce the Hagedorn temperature has been determined di-\nrectly from the divergence of the flux-loop correlation\nlength, the residual effects of interactions on the thermo-\ndynamic potentials are small. It may be that thermody-\nnamic properties in general are not strongly influenced by\ninteractions when a large number of states are contribut-\ning. A well-known example is provided by the N= 4\nsuper-Yang-Mills theory, whose entropy density at very\nstrong coupling is only reduced by a factor 3/4 with re-spect to the free theory [37]. In this interpretation, the\nmain effect of interactions among glueballs on thermo-\ndynamic properties is to slightly shift the value of the\nHagedorn temperature Thwith respect to its free-string\nvalue. A possible mechanism is that the string tension\nthat effectively determines This an in-medium string ten-\nsion that is ∼8% lower than at T= 0.\nReturning to full QCD, our results lend support to the\nidea that the hadron resonance gas model can largely\naccount for the thermodynamic properties of the low-\ntemperature phase. Whether the open string density of\nstates reproduces the entropy calculated on the lattice\ncan also be tested at quark masses not necessarily as\nlight as in Nature using a simple open string model [38].\nAcknowledgments\nI thank B. Zwiebach for a discussion on the bosonic\nstring density of states. The simulations were done on\nthe Blue Gene L rack and the desktop machines of the\nLaboratory for Nuclear Science at M.I.T. This work was\nsupported in part by funds provided by the U.S. Depart-\nment of Energy under cooperative research agreement\nDE-FG02-94ER40818.\n[1] B. Mueller, Prog. Theor. Phys. Suppl. 174, 103 (2008).\n[2] I. Arsene et al. (BRAHMS), Nucl. Phys. A757 , 1 (2005),\nnucl-ex/0410020.\n[3] B. B. Back et al., Nucl. Phys. A757 , 28 (2005), nucl-\nex/0410022.\n[4] K. Adcox et al. (PHENIX), Nucl. Phys. A757 , 184\n(2005), nucl-ex/0410003.\n[5] J. Adams et al. (STAR), Nucl. Phys. A757 , 102 (2005),\nnucl-ex/0501009.\n[6] M. Cheng et al., Phys. Rev. D77, 014511 (2008),\n0710.0354.\n[7] A. Bazavov et al. (2009), 0903.4379.\n[8] Y. Aoki, Z. Fodor, S. D. Katz, and K. K. Szabo, Phys.\nLett.B643 , 46 (2006), hep-lat/0609068.\n[9] Y. Aoki et al. (2009), 0903.4155.\n[10] F. Karsch, K. Redlich, and A. Tawfik, Eur. Phys. J. C29,\n549 (2003), hep-ph/0303108.\n[11] A. Andronic, P. Braun-Munzinger, and J. Stachel, Phys.\nLett.B673 , 142 (2009), 0812.1186.\n[12] N. Demir and S. A. Bass (2008), 0812.2422.\n[13] E. Witten, Nucl. Phys. B160 , 57 (1979).\n[14] H. B. Meyer (2004), hep-lat/0508002.\n[15] Y. Chen et al., Phys. Rev. D73, 014516 (2006), hep-\nlat/0510074.\n[16] K. G. Wilson, Phys. Rev. D10, 2445 (1974).\n[17] H. B. Meyer and J. W. Negele, Phys. Rev. D77, 037501\n(2008), 0707.3225.\n[18] H. B. Meyer and J. W. Negele, PoS LATTICE2007 ,\n154 (2007), 0710.0019.\n[19] H. B. Meyer, Phys. Rev. D76, 101701 (2007), 0704.1801.\n[20] J. Engels, F. Karsch, and T. Scheideler, Nucl. Phys.B564 , 303 (2000), hep-lat/9905002.\n[21] S. Duerr, Z. Fodor, C. Hoelbling, and T. Kurth, JHEP\n04, 055 (2007), hep-lat/0612021.\n[22] S. Necco and R. Sommer, Nucl. Phys. B622 , 328 (2002),\nhep-lat/0108008.\n[23] G. Boyd et al., Nucl. Phys. B469 , 419 (1996), hep-\nlat/9602007.\n[24] Y. Namekawa et al. (CP-PACS), Phys. Rev. D64, 074507\n(2001), hep-lat/0105012.\n[25] H. B. Meyer (2009), 0905.1663.\n[26] H. B. Meyer, JHEP 03, 064 (2005), hep-lat/0412021.\n[27] B. Lucini, M. Teper, and U. Wenger, JHEP 02, 033\n(2005), hep-lat/0502003.\n[28] J. F. Arvis, Phys. Lett. B127 , 106 (1983).\n[29] B. Bringoltz and M. Teper, Phys. Rev. D73, 014517\n(2006), hep-lat/0508021.\n[30] L. G. Yaffe and B. Svetitsky, Phys. Rev. D26, 963 (1982).\n[31] B. Beinlich, F. Karsch, and A. Peikert, Phys. Lett. B390 ,\n268 (1997), hep-lat/9608141.\n[32] B. Zwiebach (2004), A First Course in String Theory,\nCambridge, UK: Univ. Pr. 558 p.\n[33] B. Lucini, M. Teper, and U. Wenger, JHEP 01, 061\n(2004), hep-lat/0307017.\n[34] H. B. Meyer, JHEP 01, 071 (2009), 0808.3151.\n[35] B. Lucini and M. Teper, JHEP 06, 050 (2001), hep-\nlat/0103027.\n[36] R. Hagedorn, Nuovo Cim. Suppl. 3, 147 (1965).\n[37] S. S. Gubser, I. R. Klebanov, and A. W. Peet, Phys. Rev.\nD54, 3915 (1996), hep-th/9602135.\n[38] A. Selem and F. Wilczek (2006), hep-ph/0602128." }, { "title": "0906.1036v1.Quantum_Hall_Systems_Studied_by_the_Density_Matrix_Renormalization_Group_Method.pdf", "content": "arXiv:0906.1036v1 [cond-mat.str-el] 5 Jun 20091\nQuantum Hall Systems Studied by the Density Matrix\nRenormalization Group Method\nNaokazu Shibata\nDepertment of Physicas, Tohoku University, Aoba, Aoba-ku, Sendai, Miyagi\n980-8578\nThe ground-state and low-energy excitations of quantum Hal l systems are studied by the\ndensity matrix renormalization group (DMRG) method. From t he ground-state pair corre-\nlation functions and low-energy excitions, the ground-sta te phase diagram is determined,\nwhich consists of incompressible liquid states, Fermi liqu id type compressible liquid states,\nand many kinds of CDW states called stripe, bubble and Wigner crystal. The spin transition\nand the domain formation are studied at ν= 2/3. The evolution from composite fermion\nliquid state to an excitonic state in bilayer systems is inve stigated at total filling factor ν= 1.\n§1. Introduction\nIntwodimensionalsystems, applyingperpendicularmagnet icfieldstronglymod-\nifies the wave function of electrons leading to many interest ing phenomena at low\ntemperatures. The fractional quantum Hall effects1)are typical example, where in-\ncompressible ground states are realized only at some fracti onal fillings of Landau\nlevels.2),3)Since fractional quantum Hall effects are observed only in hig h quality\nsamples, the Coulomb interaction between the electrons is t hought to be essential\nrather than random potentials from impurities. This is cont rasted with the case of\ninteger quantum Hall effect where random potentials are essen tial. The importance\nof the Coulomb interaction in high magnetic field is followed from the increase in\nthe energy scale of the Coulomb interaction. The wave functi on is scaled by the\nmagnetic length ℓ=/radicalbig\n/planckover2pi1/eB, which is equivalent to the classical cyclotron radius rc\nin the lowest Landau level. The increase in the magnetic field decreases the mag-\nnetic length and enhances the energy scale of the Coulomb int eraction between the\nelectrons, e2/εℓ.\nAt typical magnetic field of 10T, ℓis about 8nm, which is still much larger\nthan the atomic length of 0.1nm. Since the conduction electr ons are on the positive\nbackground charge from ions over length scale of ℓ, the positive charge may be\nsimplified to be uniform. Then the system is equivalent to the electron gas in a\nmagnetic field, and ℓbecomes unique length scale of the system.\nIn the magnetic field, the kinetic energy is scaled by the cycl otron frequency\nωc=eB/m, which is determined by the magnetic field B. The quantization of the\nwave function discretizes the classical cyclotron radius rc, which also discretizes the\nkinetic energy and makes Landau levels E=/planckover2pi1ωc(n+ 1/2). This means that the\nmacroscopic number of electrons have the same energy in each Landau level, and\nlarge-scale degeneracy appears in the ground state. This ma croscopic degeneracy is\nlifted by the Coulomb interaction between the electrons and various types of liquid\ntypeset using PTPTEX.cls/angbracketleftVer.0.9/angbracketright2 Naokazu Shibata\nstates4),5),6),7),8)and CDW states9),10),11),12)are realized depending on the filling of\nthe Landau levels.\nSince the ground state has macroscopic degeneracy in the lim it of weak Coulomb\ninteraction, standard perturbation theories are not usefu l. Thus numerical diagonal-\nizations of the many body Hamiltonian have been used to study this system. Since\nnumerical representation of the Hamiltonian needs complet e set of many body basis\nstates, we divide the system into unit cells with finite numbe r of electrons in each\ncell. The properties of the infinite system are obtained by th e finite size scalings.\nHowever, the number of many body basis states increases expo nentially with the\nnumber of electrons. For example, when we study the ground st ate atν= 1/3\nwith 18 electrons, each unit cell has 54 degenerated orbital s. The number of many\nbody basis states is given by the combination of occupied and unoccupied orbitals,\n54C18∼1014, which is practically impossible to manage by using standar d numerical\nmethod such as exact diagonalization.\nTo study systems with typically more than 10 electrons, we ne ed to reduce the\nnumberofmanybodybasisstates. For thispurpose,weusethe density matrixrenor-\nmalization group (DMRG) method, which was originally devel oped by S. White in\n1992.13),14)This method is a kind of variational method combined with a re al space\nrenormalization group method, which enables us to obtain th e ground-state wave\nfunction of large-size systems with controlled high accura cy within a restricted num-\nber of many body basis states. The DMRG method has excellent f eatures compared\nwith other standard numerical methods. In contrast to the qu antum Monte Carlo\nmethod, theDMRG method is freefrom statistical errors and t henegative sign prob-\nlem, whichinhibitconvergence ofphysical quantities atlo w temperatures. Compared\nwith the exact diagonalization method, the DMRG method has n o limitation in the\nsize of system. The error in the DMRG calculation comes from r estrictions of the\nnumber of basis states, which is systematically controlled by the density matrix cal-\nculated from the ground-state wave function, and the obtain ed results are easily\nimproved by increasing the number of basis states retained i n the system.\nThe application of the DMRG method to two-dimensional quant um systems\nis a challenging subject and many algorithms have been propo sed. Most of them\nuse mappings on to effective one-dimensional models with long -range interactions.\nHowever, the mapping from two-dimensional systems to one-d imensional effective\nmodels is not unique and proper mappingis necessary to keep h igh accuracy. In two-\ndimensional systems under a perpendicular magnetic field, a ll the one-particle wave\nfunctions ΨNX(x,y) are identified by the Landau level index Nand the x-component\nof the guiding center, X, in Landau gage. The guiding center is essentially the cente r\ncoordinate of the cyclotron motion of the electron and it is n atural to use Xas a\none-dimensional index of the effective model. More important ly,Xis discretized in\nfinite unit cell of Lx×Lythrough the relation to y-momentum, X=kyℓ2, which\nis discretized under the periodic boundary condition, ky= 2πn/Lywithnbeing\nan integer. Therefore, the two-dimensional continuous sys tems in magnetic field are\nnaturally mapped on to effective one-dimensional lattice mod els, and we can apply\nthe standard DMRG method.15)\nThis method was first applied to interacting electron system s in a high LandauQuantum Hall Systems Studied by the DMRG Method 3\nlevel and the ground-state phase diagram, which consists of various CDW states\ncalled stripe, bubble and Wigner crystal, has been determin ed.16),17)The ground\nstate and low energy excitations in the lowest and the second lowest Landau levels\nhave also been studied by the DMRG and the existence of variou s quantum liquid\nstates such as Laughlin state and charge ordered states call ed Wigner crystal have\nbeen confirmed and new stripe state has been proposed.18),19)\nIn the following, we first explain the effective one-dimension al Hamiltonian used\nin the above studies and then show the results obtained for th e spin polarized single\nlayer system. We next review recent study on the spin transit ion and domain forma-\ntion atν= 2/3,20)and finally explain the results on bilayer quantum Hall syste ms\natν= 1,21)where crossover from a Fermi liquid state to an excitonic inc ompressible\nstate occurs.\n§2. DMRG method\nHere we briefly describe how the effective 1D Hamiltonian is obt ained from 2D\nquantum Hall systems.15)To describe the many body Hamiltonian for a interacting\nsystem, we first need to define one-particle basis states. Her e, we use the eigenstates\nof free electrons in a magnetic field as one-particle basis st ates and represent the\nwave function ΨNX(x,y) in Landau gauge:\nΨNX(x,y) =CNexp/bracketleftbigg\nikyy−(x−X)2\n2ℓ2/bracketrightbigg\nHN/bracketleftbiggx−X\nℓ/bracketrightbigg\n, (2.1)\nwhereHNare Hermite polynomials and CNis the normalization constant. Then\nall the eigenstates ΨNX(x,y) are specified using two independent parameters Nand\nX;Nis the Landau level index and Xis thex-component of the guiding center\ncoordinates of the electron. Since the guiding center Xis related to the momentum\nkyasX=kyℓ2, andkyis discretized under the periodic boundary conditions, the\nguiding center Xtakes only discrete values\nXn= 2πℓ2n/Ly, (2.2)\nwhereLyis the length of the unit cell in the y-direction.\nIfwefixtheLandaulevel index N, all theone-particle states arespecifiedbyone-\ndimensional discrete parameter Xn. Since many body basis states are product states\nof one-particle states, they are also described by the combi nations of Xnof electrons\nin the system. Thus the system can be mapped on to an effective on e-dimensional\nlattice model.\nThe macroscopic degeneracy in the ground state of free elect rons in partially\nfilled Landau level is lifted by the Coulomb interaction\nV(r) =e2\nǫr. (2.3)\nThe Coulomb interaction makes correlations between the ele ctrons and stabilizes\nvarious types of ground states depending on the filling νof Landau levels. When4 Naokazu Shibata\nM M−1 n=1 243 n=1 2 7 8 65\n3M−2\n4 5\nM=8\nBL BR(a) (b) M=8\nBL BR\nFig. 1. Schematic diagrams for (a) infinite system algorithm and (b) finite system algorithm of the\nDMRG method. •represents a one-particle orbital in a given Landau level. B Land B Rare left\nand right blocks, respectively.\nthe magnetic field is strong enough so that the Landau level sp litting is sufficiently\nlarge compared with the typical Coulomb interaction e2/(ǫℓ), the electrons in fully\noccupied Landau levels are inert and the ground state is dete rmined only by the\nelectrons in the top most partially filled Landau level.\nThe Hamiltonian is then written by\nH=S/summationdisplay\nnc†\nncn+1\n2/summationdisplay\nn1/summationdisplay\nn2/summationdisplay\nn3/summationdisplay\nn4An1n2n3n4c†\nn1c†\nn2cn3cn4, (2.4)\nwhere we have imposed periodic boundary conditions in both x- andy-directions,\nandSis the classical Coulomb energy of Wigner crystal with a rect angular unit cell\nofLx×Ly.22)c†\nnis the creation operator of the electron represented by the w ave\nfunction defined in equation (2 .1) withX=Xn.An1n2n3n4are the matrix elements\nof the Coulomb interaction defined by\nAn1n2n3n4=δ′\nn1+n2,n3+n41\nLxLy/summationdisplay\nqδ′\nn1−n4,qyLy/2π2πe2\nǫq\n×/bracketleftbig\nLN(q2ℓ2/2)/bracketrightbig2exp/bracketleftbigg\n−q2ℓ2\n2−i(n1−n3)qxLx\nM/bracketrightbigg\n,(2.5)\nwhereLN(x) are Laguerre polynomials with Nbeing the Landau level index.17)\nδ′\nn1,n2= 1 when n1=n2(modM) withMbeing the number of one-particle states\nin the unit cell, which is given by the area of the unit cell 2 πMℓ2=LxLy.\nInordertoobtaintheground-statewavefunctionweapplyth eDMRGmethod.15)\nAs shown in Fig. 1 (a), we start from a small-size system consi sting of only four\none-particle orbitals whose indices nare 1, 2, M−1, andM, and we calculate the\nground-state wave function. We then construct the left bloc k containing one-particle\norbitals of n= 1 and 2, and the right block containing n=M−1 andMby using\neigenvectors of the density matrices which are calculated f rom the ground-state wave\nfunction. We then add two one-particle orbitals n= 3 and M−2 between the two\nblocks and repeat the above procedure until Mone-particle orbitals are includedQuantum Hall Systems Studied by the DMRG Method 5\n-10-8-6-4-20\n0 10010\n1010101010\n200 300Ne=18 M=54\n2D system\nw\nFig. 2. Eigenvalues wαof the density matrix for two-dimensional system of 54 orbit als with 18\nelectrons. Sumof wαis equivalentto the norm of theground-state wave function a nd normalized\nto be unity.\nin the system. We then apply the finite system algorithm of the DMRG shown in\nFig. 1 (b) to refine the ground-state wave function. After we h ave obtained the\nconvergence, we calculate correlation functions to identi fy the ground state.\nThe ground-state pair correlation function g(r) in guiding center coordinates is\ndefined by\ng(r) =LxLy\nNe(Ne−1)/angbracketleftΨ|/summationdisplay\ni/negationslash=jδ(r−Ri+Rj)|Ψ/angbracketright, (2.6)\nwhereRiis the guiding center coordinate of the ith electron, and it is calculated\nfrom the following equation\ng(r) =1\nNe(Ne−1)/summationdisplay\nq/summationdisplay\nn1,n2,n3,n4exp/bracketleftbigg\niq·r−q2ℓ2\n2−i(n1−n3)qxLx\nM/bracketrightbigg\n×\nδ′\nn1−n4,qyLy/2π/angbracketleftΨ|c†\nn1c†\nn2cn3cn4|Ψ/angbracketright, (2.7)\nwhereΨis the ground state and Neis the total number of electrons.\nThe accuracy of the results depends on the distribution of ei genvalues of the\ndensity matrix. A typical example of the eigenvalues of thed ensity matrix for system\nofM= 54 with 18 electrons is shown in Fig. 2, which shows an expone ntial decrease\nof eigenvalues wα. In this case accuracy of 10−4is obtained by keeping more than\none hundred states in each block.\n§3. Single layer system\nHere we present diverse ground states obtained by the DMRG me thod applied\nto the single layer quantum Hall systems. In the limit of stro ng magnetic field,\nthe electrons occupy only the lowest Landau level N= 0. In this limit, fractional\nquantum Hall effect (FQHE) has been observed at various fracti onal fillings.3)The\nFQHE state is characterized by incompressible liquid with a finite excitation gap.1)6 Naokazu Shibata\n 0 0.02 0.04 0.06 0.08\n 0.3 0.4 0.52/53/7\n4/9\n5/11\nFig. 3. The lowest excitation gap at various νin the lowest Landau level. Relatively large excitation\ngap is obtained at fractional fillings ν=n/(2n+1). The excitation gap is in units of e2/(ǫℓ).\nThese FQHE states are confirmed by the DMRG calculations, whe re relatively\nlarge excitation gaps are obtained at various fillings betwe enν= 1/2 and 3/1018)\nas shown in Fig. 3. We clearly find large excitation gaps at fra ctional fillings ν=\n1/3,2/5,3/7,4/9 and 5/11, which correspond to primary series of the FQHE at\nν=n/(2n+ 1). The pair correlation function at ν= 1/3 is presented in Fig. 4,\nwhich shows a circularly symmetric correlation consistent with the Laughlin’s wave\nfunction.1)\nIn the limit of low filling ν→0, mean separation between the electrons becomes\nmuch longer than the typical length-scale of the one-partic le wave function. In this\nlimit the quantum fluctuations are not important and electro ns behave as classical\npoint charges. The ground state is then expected to be the Wig ner crystal. The\nformation of the Wigner crystal is also confirmed by the DMRG c alculations at low\nfillings as shown in Fig. 5 (a). The ν-dependence of the low energy spectrum shows\nthat the first-order transition to Wigner crystal occurs at ν∼1/7.18)\nWith decreasing magnetic field, electrons occupy higher Lan dau levels. In high\nLandau levels, the one-particle wave function extends over space leading to effective\nlong range exchange interactions between the electrons. Th e long range interaction\nstabilizes CDW ground states and various types of CDW states called stripe and\nbubble are predicted by Hartree-Fock theory.9)These CDW states are confirmed by\nthe DMRG calculations as shown in Figs. 5 (b) and (c), where tw o-electron bubble\nstate and stripe state are obtained at ν= 8/27 and 3 /7, respectively, in the N= 2\nLandau level. Although the CDW structures are similar to tho se obtained in the\nHartree-Fock calculations, the ground state energy and the phase diagram are sig-\nnificantly different.16)The DMRG results are consistent with recent experiments,10)\nandthediscrepancyis dueto thequantumfluctuations neglec ted intheHartree-Fock\ncalculations.\nThe ground-state phase diagram obtained by the DMRG is shown in Fig. 6. In\nthe lowest Landau level, we find many liquid states at fractio nal fillings and around\nν= 1/2. Nevertheless, CDW states dominate over the whole range of filling inQuantum Hall Systems Studied by the DMRG Method 7\nhigher Landau levels. This difference in the ground state phas e diagram comes from\ndifferent effective interactions between the electrons. In the lowest Landau level,\nthe one particle wave function is localized within the magne tic length ℓ, that yields\nstrong short-range repulsion between the electrons. Since quantum liquid states such\nas Laughlin state are stabilized by the strong short-range r epulsion, liquid states are\nrealized in the lowest Landau level. In higher Landau levels , however, the wave\nfunction extends over space with the increase in the classic al cyclotron radius rc.\nThus the short-range repulsion is reduced and liquid states become unstable. As\nshown in Fig. 7 (a), the real space effective interaction betwe en the electrons in\nhigher Landau levels has a shoulder structure around the dis tance twice the classi-\ncal cyclotron radius. This structure of effective interactio n makes minimum in the\nCoulomb potential near the guiding center of the electron as shown in Fig. 7 (b) and\nstabilizes the clustering of electrons. This is the reason w hy stripe and bubble states\nare realized in higher Landau levels.19)\n0\n3.76-3.76\n7.52-7.520\n-3.763.76\n-7.527.52\nx\nyLaughlin state\n01.0\nFig. 4. Pair correlation function g(r) atν= 1/3 in the lowest Landau level. The length is in units\nofℓ.\n9.264.630-4.63-9.26\n-7.12-3.56\n0\n3.567.12(c) stripe\nx\ny12.706.350-6.35-12.70\n-6.68-3.34\n0\n3.34\n6.68(b) two-electron \n bubble (a)\nx x\nyy-11.65\n-5.82\n0 \n5.85\n11.65\n-9.71-4.85\n0 \n4.859.71Wigner crystal\nFig. 5. Pair correlation functions g(r) in guiding center coordinates. (a) Wigner crystal realize d in\nan excited state at ν= 1/6 in the lowest Landau level. The number of electrons in the un it\ncellNeis 12. (b) Two-electron bubble state at ν= 8/27 inN= 2 Landau level. Ne= 16. (c)\nStripe state at ν= 3/7 inN= 2 Landau level. Ne= 18.8 Naokazu Shibata\nstripe I2-electron \nbubble N=2Wigner crystal\nWigner crystal\nWigner crystal\n0 0.5 0.4 0.3 0.2 0.1stripe I N=1 pairing stripe II\n0 0.5 0.4 0.3 0.2 0.1FQHE\n1/5FQHE\n1/3N=0 liquidFQHEFQHE\nNN1/32/7\n2/5FQHE\n3/74/95/11\n0 0.5 0.4 0.3 0.2 0.11/5 stripe II\nFig. 6. The ground state phase diagram obtained by the DMRG me thod.Nis the Landau level\nindex and νNin the filling factor of the Nth Landau level.\n 0 1\n 0 2 4 6 8 10\nrVeff (r)\nN=0 231N = 1N = 0\nN = 2\nN = 3pure Coulomb\n2Rc 0 1\n 0 2 4 6xVeff(x) (b) (a)\nN=2\nx\nx x xx x\nFig. 7. (a) Effective interaction between the electrons in th eNth Landau level. Rcis the classical\ncyclotron radius. (b) Coulomb potential made by two electro ns separated by ∆x.\n§4. Spin transitions\nIn two dimensional systems, strong perpendicular magnetic field completely\nquenches the kinetic energy of electrons. Since the kinetic energy is independent\nof the spin polarization, the exchange Coulomb interaction easily aligns the electron\nspin. The ferromagnetic ground state at ν= 1/q(qodd) is thus realized even in\nthe absence of the Zeeman splitting.23)At the filling ν= 2/3 and 2/5, however, the\nparamagnetic ground states compete with the ferromagnetic state, and the Zeeman\nsplitting ∆z=gµBBinduces a spin transition.24)Such aspin transition in fractional\nquantum Hall states has been naively explained by the compos ite fermion theory.25)\nThe composite fermions are electrons coupled with even numb er of fluxes. These\nfluxes effectively reduces external magnetic field and the ν=p/(2p±1) fractional\nquantum Hall effect (FQHE) state is mapped on to the ν′=pinteger QHE state\nof composite fermions. The spin transitions at ν= 2/3 and 2/526)correspond to\nthe spin transition at ν= 2, where the Zeeman splitting corresponds to the effectiveQuantum Hall Systems Studied by the DMRG Method 9\nFig. 8. Lowest energies for fixed polarization ratio Pas a function of magnetic field Bat filling\nfactorν= 2/3 in units of e2/(ǫℓ). The total number of electron is 20. The aspect ratio is fixed\nat 2.0. The g-factor is 0.44.\nFig. 9. Charge gap of ν= 2/3 spin polarized states ( /square), unpolarized states ( •), and partially\npolarized states ( ◦) for various Neand aspect ratios Lx/Ly.∆cis in units of e2/(ǫℓ).\nLandau level separation, and theenergy levels of the minori ty spinstate in thelowest\nLandau level and the majority spin state in the second lowest Landau level coincide.\nExtensive experimental27),28),29),30),31),32),33),34),35)and theoretical36),37),38),39)\nstudies have been made on this transition. Nevertheless, th ere is no clear theoretical\nconsensus on this issue. This is due to the difficulties of stud ies in this system. A\nnumber of states possibly compete in energy, and large enoug h systems are needed to\nsee non-uniform structures in the partially polarized stat es. Here we use the DMRG\nmethod,15)and study the spin transition and the spin structures in larg e system to\nclarify the nature of the spin transition at ν= 2/3.\nWe first calculate the energy at various polarization Pas a function of the\nZeeman splitting, ∆z=gµB. The obtained results are shown in Fig. 8. In the\nabsence of the Zeeman splitting, the unpolarized state ( P= 0) is the lowest. The\nenergy of the polarized state ( P >0) monotonically increases as Pincreases. With10 Naokazu Shibata\n(a)\n(b)(c)\n(d)\nFig. 10. Pair correlation functions for minority spins g↓↓atν= 2/3 for several polarization ratios\n(a)P= 0.8, (b)P= 0.6, (c)P= 0.5, and (d) P= 0.4.\nthe increase in Zeeman splitting ∆z, however, the energy of polarized state decreases\nand the fully polarized state ( P= 1) becomes the lowest. Figure 8 shows that the\ntransition from the unpolarized state to the fully polarize d state occurs at B≃6T\nwhich is roughly consistent to the earlier work done in a sphe rical geometry.24)In\nthe present calculation on a torus, all partially polarized states (0 < P <1) are\nhigher in energy than the ground states ( P= 0 or 1). This feature is independent of\nthe size of the system and the aspect ratio Lx/Ly, and indicates phase separations\nofP= 0 and P= 1 in partially polarized states.\nThe unpolarized state of P= 0 and the fully polarized state of P= 1 are both\nquantum Hall states with finite charge excitation gap, which is defined by\n∆c(P) =E(Nφ+1,P)+E(Nφ−1,P)−2E(Nφ,P), (4.1)\nwhereNφis the number of one-particle states in the lowest Landau lev el. The filling\nfactorνis then given by Ne/Nφ. The charge gap ∆cfor various Nφand aspect\nratios of the unit cell is presented in Fig. 9. In this figure, t he gap∆cseems to\nvanish for partially polarized state P∼1/2 in the limit of Ne→ ∞. This result\nclearly indicates that partially polarized state with P∼1/2 is a compressible state\nin contrast to the incompressible states at P= 0 and 1, where ∆cremains to be\nfinite in the limit of Ne→ ∞.\nTo study the spin structure in the partially polarized state s, we next calculate\nthe pair-correlation function defined by\ngσσ(r) =LxLy\nNσ(Nσ−1)/angbracketleftΨ|/summationdisplay\nnmδ(r+Rσ,n−Rσ,m)|Ψ/angbracketright, (4.2)\nwhereσ=±1/2 is the spin index and Nσis the number of electrons with spin\nσ. The spin structures in partially spin polarized states are clearly shown in the\npair correlation function between minority spins. Namely, if unpolarized regions\nare formed in the partially polarized states, then electron s with minority spins are\nconcentrated in the unpolarized regions. This concentrati on of the minority spins isQuantum Hall Systems Studied by the DMRG Method 11\nFig. 11. Local densities of up spin, and down spin electrons f or various polarization ratios Pat\nν= 2/3. The number of electrons is 20.\nshown in Fig. 10, which shows g↓↓(x,y) for partially polarized states at (a) P= 0.8,\n(b)P= 0.6, (c)P= 0.5, and (d) P= 0.4. When Pis close to 1, for example P= 0.8\nshown in Fig. 10(a), a pair of minority spins is found only nea r the origin. As the\npolarization ratio Pdecreases, minority spins make a domain around the origin, a nd\ntwo domain walls along the y-direction are formed. These domain walls move along\nx-direction and the domain of minority spin finally covers ent ire unit cell in the limit\nofP= 0. This change in the size of the domain is consistent with th e expectation\nthat the domain in Fig. 10 corresponds to the unpolarized spi n singlet region where\nthe density of up-spin electrons and the down-spin electron s are the same.\nTo confirm the separation of the unpolarized and polarized sp in regions, we\nnext consider the local electron density of up-spin electro nsν↑(x) and down-spin\nelectrons ν↓(x). Figure 11 shows ν↑(x) andν↓(x) for partially polarized states with\nP= 0.2,0.4,0.6 and 0.7. Here ν↑(x) andν↓(x) are scaled to be the local filling\nfactor of the lowest LL. Thus, the total local electron densi tyν↑(x)+ν↓(x) is almost\n2/3. In this figure the separation to two regions is clearly seen ; the unpolarized spin\nregion around Lx/2, where both ν↑andν↓are close to 1/3, and the fully polarized\nspin region around x∼0 or equivalently x∼Lx, whereν↑is almost 2/3 while ν↓\nis close to 0. These results confirm the separation of the unpo larized and polarized\nspin regions as expected from the pair correlation function s shown in Fig. 11.\nThe polarized and unpolarized spin regions are separated by the domain walls\nwhose width is about 4 ℓ. This means that the phase separation is realized only\nfor systems whose size of the unit cell Lx,(Ly) is larger than twice the width of\ndomain wall; Lx,(Ly)>8ℓ. Indeed, exact diagonalization studies up to Ne= 8\nelectrons have never found the phase separation at ν= 2/3.39)We have found the\nphase separation only for large systems with Ne>12. We note that above behavior\nis generic over the aspect ratio. In an ideal system, the two s tates separate into\ntwo regions even when the system size is infinitely large. In e xperimental situations,\nhowever, multi-domain structuresarerealized duetothein homogeneity andcoupling\nwith randomly distributed nuclear spins.12 Naokazu Shibata\nThe DMRG study on the ground state energy for various polariz ationPshows\nthat the ground state at ν= 2/3 evolves discontinuously from the unpolarized P= 0\nstate to the fullypolarized P= 1state as the Zeeman splittingincreases. In partially\npolarized states 0 < P <1, the electronic system separates spontaneously into two\nstates; the P= 0 and the P= 1 states. These two states are separated by the\ndomain wall of width4 ℓ. Sincetheenergy of thedomain wall is positive, thepartial ly\npolarized states always have higher energy than that of P= 1 orP= 0 states. We\nthink this is the reason of the direct first order transition f romP= 0 toP= 1 state\nin the ground state.\nIt is useful to compare our result with the spin transition at ν= 2 which occurs\nwhen minority spin states in the lowest LL and majority spin s tates in the second\nlowest LL cross by varying the ratio of the Zeeman and Coulomb energy. The\nground state at ν= 2 is thus a fully polarized state or a spin singlet state. In\nanalogous to the case of ν= 2/3 the transition between them is first order,40)\nand spin domain walls have been found in high energy states.41)This analogy can\nbe expected, because the ν= 2 states and the ν= 2/3 states are connected in the\ncompositefermiontheory,25),26)althoughtheeffective interaction betweencomposite\nfermions is different from that for electrons.\n§5. Bilayer system\nThepropertiesofquantumHall systemssensitively dependo nthemagneticfield,\nand various types of ground states including incompressibl e liquids,1)compressible\nliquids,42),43)spin singlet liquid, CDW states called stripes, bubbles, an d Wigner\ncrystal are realized depending on the filling νof Landau levels. In bilayer quantum\nHall systems, additional length scale of the layer distance d, and the degrees of\nfreedom of layers make the ground state much more diverse and interesting.44)\nExcitonic phase, namely Haplerin’s Ψ1,1,1state, is one of the ground states real-\nized in bilayer quantum Hall systems at total filling ν= 1 at small layer separation\nd, where electrons and holes in different layers are bound with e ach other due to\nstrong interlayer Coulomb interaction. This excitonic sta te has recently attracted\nmuch attention because a dramatic enhancement of zero bias t unneling conductance\nbetween the two layers,45)and the vanishing of the Hall counterflow resistance are\nobserved.46),47)As the layer separation dis increased, the excitonic phase vanishes,\nand at large enough separation, composite-fermion Fermi-l iquid state is realized in\neach layer.\nSeveral scenarios have been proposed for the transition of t he ground state as\nthelayer separation increases.48),49),50),51),52),53),54)However howtheexcitonic state\ndevelops into independent Fermi-liquid state has not been f ully understood. In this\nsection we investigate the ground state of ν= 1 bilayer quantum Hall systems\nby using the DMRG method.15)We calculate energy gap, two-particle correlation\nfunction g(r) and excitonic correlation function for various values of l ayer separation\nd, and show the evolution of the ground state with increasing d.Quantum Hall Systems Studied by the DMRG Method 13\nNe=24 Lx/Ly=1.6Ne=18 Lx/Ly=1.0Ne=12 Lx/Ly=1. 5\nd/lgex(n)n=M /2\nn=M /2−1\n 0 0.2 0.4 0.6 0.8 1\n 0 1 2\nFig. 12. The exciton correlation of bilayer quantum Hall sys tems atν= 1. The solid line represents\ngex(M/2). The dashed line represents gex(M/2−1).\nThe Hamiltonian of the bilayer quantum Hall systems is writt en by\nH=/summationdisplay\ni1.6. This behavior is consistent with experiments.55)Although\nthe excitation gap for d/ℓ >1.7 is not presented in the figure for Ne= 24 because\nof the difficulty of the calculation of excited states in large system, we do not find\nany sign of level crossing in the ground state up to d/ℓ∼4, where two layers are\nalmost independent. These results suggest that the exciton ic state at small d/ℓcon-\ntinuously crossovers to compressible state at large d/ℓ, that is consistent with the\nbehaviors of exciton correlations gex(M/2) in Fig. 12, which shows gex(M/2) contin-\nuously approaches zero around d/ℓ∼1.6. In the present calculation it is difficult to\nconclude whether the gap closes at finite d/ℓ∼1.6 in the thermodynamic limit or\nexcitonic state survives with exponentially small finite ga p even for large d/ℓ >1.6.\nWe have calculated the excitation gap in different size of syst ems and aspect ratios,\nand obtained similar results as shown in the inset of Fig. 15.16 Naokazu Shibata\n 0 0.02 0.04 0.06 0.08\n 0 0.4 0.8 1.2 1.6 2 2.2\nd/ld/lNe=24\nNe=18\n 0 0.1\n 0 1 2\nFig. 15. The lowest excitation gap ∆of bilayer quantum Hall systems at the total filling factor\nν= 1.Ne= 24 and Lx/Ly= 1.6. The inset shows the result for Ne= 18 and Lx/Ly= 1.0.\nConcerning the first excited state, however, Fig. 15. shows a level crossing at\nd/ℓ∼1.2, where we can see sudden decrease in the excitation gap. We e xpect that\nthe lowest excitation at d/ℓ <1.2 is the pseudo-spin excitation whose energy gap\ndecreases with the increase in the size of system and tends to zero in the limit of\nlarge system. On the other hand, the lowest excitation at d/ℓ >1.2 shown in Fig. 15\nis expected to be the excitation to the roton minimum which co rresponds to the\nbound state of quasiparticle and quasihole excitatins, who se energy increases with\nthe decrease in d/ℓ. This change in the character of the low energy excitations a t\nd/ℓ∼1.2 will be confirmed in a clear change in correlation functions in the excited\nstate as shown later. We note that the position of the level cr ossing in the first\nexcited state itself depends on the size of system because th e pseudo-spin excitation\ngap decreases with the increase in the system size. However, the change in the\ncharacter of the low energy excitations of finite systems is e xpected to remain even\nin the limit of large system, since the spectrum weight of pse udo-spin waves transfers\nto high energy with the increase in d.\nWe next calculate pair correlation functions of the electro ns to see detailed evo-\nlution of the ground-state wave function. The interlayer pa ir correlation functions\nare defined by\ng12(r)≡LxLy\nN1N2/angbracketleftΨ|/summationdisplay\nn mδ(r+R1,n−R2,m)|Ψ/angbracketright, (5.4)\nwhere|Ψ/angbracketrightis the ground state. We present ∆g12(r) in Fig. 16, which is defined by\n∆g12(r) =/integraldisplay\n(g12(r′)−1)δ(|r′|−r) dr′, (5.5)\nwherer′is the two-dimensional position vector in each layer. ∆g12(r) represents the\ndifference from the uniform correlation of independent elect rons.Quantum Hall Systems Studied by the DMRG Method 17\n-0.3-0.2-0.1 0 0.1 0.2\n 0 1 2 3 4d/l=0.1d/l=0.10.2 0.40.60.8\n0.61.01.2\n1.21.41.61.81.82.02.22.4\n2.42.63.0\nr/lr/lNe=24\nNe=24\n-1 0\n 0 1 2 3 4\nFig. 16. The inter-layer pair correlation function of elect rons in the ground state of bilayer quantum\nHall systems at ν= 1.Ne= 24 and Ly/Lx= 1.6.\nAtd/ℓ= 0 we find clear negative ∆g12(r) around r/ℓ= 1, which is a characteris-\ntic feature of the excitonic state made by the binding of elec trons and holes between\nthe two layers. The binding of one hole means the exclusion of one electron caused\nby the strong interlayer Coulomb repulsion. The increase in the layer separation\nweakens Coulomb repulsion between the two layers and reduce s|∆g12(r)|around\nr/ℓ= 1.\nThe decrease in the interlayer correlation |∆g12(r)|opens space to enlarge corre-\nlation hole in the same layer and reduce the Coulomb energy be tween the electrons\nwithin the layer. This is shown in Fig. 17, which shows the pai r correlation functions\nof the electrons in the same layer defined by\ng11(r)≡LxLy\nN1(N1−1)/angbracketleftΨ|/summationdisplay\nnmδ(r+R1,n−R1,m)|Ψ/angbracketright, (5.6)\n∆g11(r) =/integraldisplay\n(g11(r′)−1)δ(|r′|−r) dr′. (5.7)\nThe obtained results indeed show that the correlation hole i n the same layer around\nr/ℓ∼1 is enhanced with the increase in d/ℓcontrary to the decrease in size of\ninterlayer correlation hole in Fig. 16. The correlation hol e in the same layer mono-\ntonically increases in size up to d/ℓ= 1.8, and then it becomes almost constant.\nThe correlation function g11(r) ford/ℓ >1.8 is almost the same to that of ν= 1/2\nmonolayer quantum Hall systems realized in the limit of d/ℓ=∞. This is consistent\nwith the almost vanishing excitation gap and exciton correl ation atd/ℓ >1.8 shown\nin Figs. 15 and 12.\nFigure 17 also shows that the growing of the correlation hole aroundr/ℓ∼1.5\nis accompanied with the increase in ∆g11(r) around r/ℓ∼4. The distance 4 ℓis\ncomparable to the approximate mean distance between the ele ctrons 3.54ℓestimated\nfrom (LxLy/N1)1/2= (2πL/N1)1/2ℓ. This means the electrons in the same layer18 Naokazu Shibata\n-0.4-0.2 0 0.2 0.4\n 0 1 2 3 4d/l=0.10.2\n0.4\n0.6\n0.8\n1.0\n1.2\n1.4\n1.61.82.02.22.42.63.0\nr/lNe=24 0\n -1 0 1 2 3 4r/ld/l=0.1\n0.6\n1.2\n1.8\n2.4Ne=24\nFig. 17. The intra-layer pair correlation function of elect rons in the ground state of bilayer quantum\nHall systems at ν= 1.Ne= 24 and Ly/Lx= 1.6.\ntend to keep distance of about 4 ℓfrom other electrons with the large correlation hole\naroundr/ℓ∼1.5 ford/ℓ>∼1. This is consistent with the formation of composite\nfermions at d=∞, where two magnetic flux quanta are attached to each electron ,\nwhich is equivalent to enhance the correlation hole around e ach electron in the same\nlayer to keep distance from other electrons.\nThe large correlation hole in g11(r) attracts electrons in the other layer as shown\nin Fig. 16, where we find a clear peak in ∆g12(r) atr/ℓ∼3. This peak at r/ℓ∼3\nis comparable to the neighboring correlation hole at r/ℓ∼1, which suggests that\nthe electrons excluded from the origin by strong interlayer Coulomb repulsion are\ntrapped by the correlation hole in g11(r) within r/ℓ∼4. Since the intra-layer\ncorrelation g11ford/ℓ >1.6 is almost the same to that of composite-fermion liquid\nstate,∆g12(r) represents the correlation of composite fermions between the layers.\nThe almost same amplitude of ∆g12(r) atr/ℓ∼1 and 3 for d/ℓ >1.6 actually shows\nthat the electrons in the other layer bind holes to form compo site fermions.\nWith decreasing d/ℓfrom infinity, the correlations of composite fermions in\ndifferent layers monotonically increases down to d/ℓ∼1.2 as shown in the enhance\nof|∆g12(r)|atr/ℓ∼1 and 3. But further decrease in d/ℓbroadens the peak at\nr/ℓ∼3 in∆g12(r) with the decrease in the correlation hole in ∆g11(r), and the peak\natr/ℓ∼3 in∆g12(r) finally disappears. This change in the correlation functio n\nshows how the composite-fermion liquid state evolves into e xcitonic state: The large\ncorrelation hole in the same layer, which is a characteristi c feature of the composite\nfermions, is transfered into the other layer to form exciton ic state. The correlation\nfunctions in Figs. 16 and 17 are continuously modified with th e decrease in d/ℓfrom\n∞to 0, which supports continuous transition from the compres sible liquid state to\nthe excitonic state. Fig. 16 also shows that the peak in ∆g12(r) atr/ℓ∼3 made\nby the binding of an electron to the hole around the origin gra dually disappears\nwith decreasing d/ℓfrom 1.2. This means the gradual break down of the concept of\ncomposite fermions.\nThebreakdownof thecompositefermionsaround d/ℓ∼1.2 affects thecharacterQuantum Hall Systems Studied by the DMRG Method 19\n-0.01-0.005 0 0.005 0.01\n 0 2 4 6 8d/l=0.2\nd/l=0.20.4\n0.41.3\n1.31.4\n1.40.6\n0.60.8\n0.81.0\n1.01.2\n1.2ij\nr/lNe=18\nFig. 18. The change in correlation function through the exci tation from the ground state to the\nfirst excited state. ν= 1 and Ne= 18 with Ly/Lx= 1.0.\nof the lowest excitations, which is clearly shown in the leve l crossing of the excited\nstate at d/ℓ∼1.2. The change in the character of excitation is confirmed by th e\ncorrelation functions in the excited state. Figure 18 shows the difference in the pair\ncorrelation functions gij(r) between the ground state and first excited state defined\nby\nδgij(r) =/integraldisplay\n(gE\nij(r′)−gG\nij(r′))δ(|r′|−r) dr′, (5.8)\nwheregG\nij(r) andgE\nij(r) are the pair correlation functions in the ground state and t he\nfirst excited state, respectively. δgij(r) in Fig. 18 show that there is a discontinuous\ntransition between d/ℓ= 1.2 and 1.3, which supports the level crossing in the first\nexcited state.\nBelowd/ℓ∼1.2,δg(r) have large amplitude at r/ℓ∼2 and 6, which shows elec-\ntrons are transfered between the inside of r/ℓ∼4 and its outside. Small singularity\natr/ℓ∼5.5 is due to finite size effects of square unit cell. Above d/ℓ∼1.2, only\nδg12(r) have large amplitude at r/ℓ∼1 and 2, which shows the electrons within\nr/ℓ∼4 in different layers are responsible for the lowest excitatio n. This result sug-\ngests that the low energy excitations are made by composite f ermions in different\nlayers for d/ℓ >1.2.\n§6. Conclusions\nInthis paperwehave reviewedthegroundstateandlow energy excitations of the\nquantum Hall systems studied by the DMRG method. We have appl ied the DMRG\nmethodtotwodimensionalquantumsystemsinmagneticfieldb yusingamappingon\ntoaneffective one-dimensional lattice model. SincetheCoul ombinteraction between\nthe electrons is long-range, all the electrons in the system interact with each other.\nThis fact seems to severely reduce the accuracy of the DMRG ca lculations. However,\nin the magnetic field, one-particle wave functions are local ized within the magnetic20 Naokazu Shibata\nlengthℓ, and the overlap of the one-particle wave functions exponen tially decreases\nwith increasing the distance between the two electrons. Thi s means the quantum\nfluctuations are restricted to short-range and the effective H amiltonian is suited for\nthe DMRG scheme. This is the reason why relatively small numb er of keeping states\nis enough for quantum Hall systems compared with usual two di mensional systems.\nIn quantum Hall systems, filling νof Landau levels is determined by ν=Ne/Nφ,\nwhereNφis the number of flux quanta and related to the magnetic field as Nφ=\n(e/h)LxLyB. Thus so many types of the ground state are realized only by ch ang-\ning the uniform magnetic field B. Since the ground state of free electrons in par-\ntially filled Landau level has macroscopic degeneracy, Coul omb interaction drasti-\ncally changes the wave function. The character of the ground state is sensitive to\nthe Landau level index Nand the filling ν, which modify the effective interaction\nand the mean distance between the electrons. This is the sour ce of many interesting\nlow temperature properties of quantum Hall systems and thei r inherent difficulties.\nAcknowledgments\nThe author would like to thank Prof. Yoshioka Daijiro and Dr. Kentaro Nomura\nfor valuable discussions. This work is supported by Grant-i n-Aid No. 18684012 from\nMEXT, Japan.\nReferences\n1) R. B. Laughlin: Phys. Rev. Lett. 50(1983), 1395\n2) D. C. Tsui, H. L. St¨ ormer and A.C. Gossard: Phys. Rev. Lett .48(1982), 1599\n3) R. R. Du, H. L. Stormer, D.C. Tsui, L. N. Pfeiffer and K. W. Wes t: Phys. Rev. Lett. 70\n(1993), 2944\n4) F. D. M. Haldane and E. H. Rezayi: Phys. Rev. Lett. 60(1988), 956\n5) M. Greiter, X. G. Wen, and F. Wilczek: Phys. Rev. Lett. 66(1991), 3205\n6) G. Moore and N. Read: Nucl. Phys. B 360(1991), 362\n7) W. Pan, J.-S. Xia, V. Shvarts, D. E. Adams, H. L. Stormer, D. C. Tsui, L. N. Pfeiffer, K.\nW. Baldwin, and K. W. West: Phys. Rev. Lett. 83(1999), 3530\n8) J. P. Eisenstein, K. B. Cooper, L. N. Pfeiffer, and K. W. West : Phys. Rev. Lett. 88(2002),\n076801\n9) A. A. Koulakov, M. M. Fogler, and B. I. Shklovskii: Phys. Re v. Lett.76(1996), 499\n10) M. P. Lilly, K.B. Cooper, J. P. Eisenstein, L. N. Pfeiffer, and K. W. West: Phys. Rev.\nLett.82(1999), 394\n11) R. R. Du, D. C. Tsui, H. L. Stormer, L. N. Pfeiffer, K. W. Bald win, and K. W. West:\nSolid State Commun. 109(1999), 389\n12) K. B. Cooper, M. P. Lilly, J. P. Eisenstein, L. N. Pfeiffer, and K. W. West: Phys. Rev. B\n60(1999), R11285\n13) S. R. White: Phys. Rev. Lett. 69(1992), 2863\n14) S. R. White: Phys. Rev. B 48(1993), 10345\n15) N. Shibata: J. Phys. A 36(2003), R381\n16) N. Shibata and D. Yoshioka: Phys. Rev. Lett. 86(2001), 5755\n17) D. Yoshioka and N. Shibata: Physica E 12(2002), 43\n18) N. Shibata and D. Yoshioka: J. Phys. Soc. Jpn. 72(2003), 664\n19) N. Shibata and D. Yoshioka: J. Phys. Soc. Jpn. 73(2004), 2169\n20) N. Shibata and D. Yoshioka: J. Phys. Soc. Jpn. 75(2006), 043712\n21) N. Shibata and K. Nomura: J. Phys. Soc. Jpn. 76(2007), 103711\n22) L. Bonsall and A. Maradudin: Phys. Rev. B 15(1977), 1959\n23) J. P. Eisenstein: Perspectives in Quantum Hall Effects , ed. S. Das Sarma and A. Pinczuk\n(John Wiley and Sons, New York 1997); S. M. Girvin and A. H. Mac Donald: in the sameQuantum Hall Systems Studied by the DMRG Method 21\nvolume.\n24) T. Chacraborty and P. Pietiainen: The Quantum Hall Effect (Springer, New York 1995)\n2nd ed.\n25) J. K. Jain: Phys. Rev. Lett. 63(1989), 199\n26) X. G. Wu, G. Dev, and J. K. Jain: Phys. Rev. Lett. 71(1993), 153\n27) J. P. Eisenstein, H. L. Stormer, L. N. Pfeiffer, and K. W. We st: Phys. Rev. B 41(1990),\n7910\n28) S. Kronm¨ uller, W. Dietsche, J. Weis, K. von Klizing, W. W egscheider, and M. Bichler:\nPhys. Rev. Lett. 81(1998), 2526\n29) I. V. Kukushkin, K. v. Klitzing, and K. Eberl: Phys. Rev. B 55(1997), 10607\n30) I. V. Kukushkin, K. v. Klitzing, and K. Eberl: Phys. Rev. L ett.82(1999), 3665\n31) N. Freytag, Y. Tokunaga, M. Horvati, C. Berthier, M. Shay egan, and L. P. L´ evy: Phys.\nRev. Lett. 87(2001), 136801\n32) K. Hashimoto, K. Muraki, T. Saku, and Y. Hirayama: Phys. R ev. Lett. 88(2002), 176601\n33) S. Kraus, O. Stern, J. G. S. Lok, W. Dietsche, K. von Klitzi ng, M. Bichler, D. Schuh, and\nW. Wegscheider: Phys. Rev. Lett. 89(2002), 266801\n34) J. H. Smet, R. A. Deutschmann, F. Ertl, W. Wegscheider, G. Abstreiter, and K. von\nKlitzing: Nature (London) 415(2002), 281\n35) J. H. Smet, R. A. Deutschmann, W. Wegscheider, G. Abstrei ter, and K. von Klitzing:\nPhys. Rev. Lett. 86(2001), 2412\n36) V. M. Apalkov, T. Chakraborty, P. Pietil¨ ainen, and K. Ni emel¨ a : Phys. Rev. Lett. 86\n(2001), 1311\n37) G. Murthy: Phys. Rev. Lett. 84(2000), 350\n38) E. Mariani, N. Magnoli, F. Napoli, M. Sassetti, and B. Kra mer: Phys. Rev. B 66(2002),\n241303(R)\n39) Karel Vyborny, Ondrej Certik, Daniela Pfannkuche, Dani el Wodzinski, Arkadiusz Wojs,\nand John J. Quinn: Phys. Rev. B 75(2007), 045434\n40) T. Jungwirth and A. H. MacDonald: Phys. Rev. B 63(2000), 035305\n41) K. Nomura, D. Yoshioka, T. Jungwirth, and A. H. MacDonald : Physica E 22(2004), 60\n42) J. K. Jain: Phys. Rev. Lett. 63(1989), 199; Phys. Rev. B 40(1989), 8079\n43) B. I. Halperin, P. A. Lee, and N. Read, Phys. Rev. B 47(1993), 7312\n44) J. P. Eisenstein and A. H. MacDonald: Nature 432(2004), 691\n45) I. B. Spielman J. P. Eisenstein L. N. Pfeiffer and K.W. West : Phys. Rev. Lett. 84(2000),\n5808\n46) M. Kellogg, J. P. Eisenstein L. N. Pfeiffer and K.W. West: P hys. Rev. Lett. 93(2004),\n036801\n47) E. Tutuc, M. Shayegan and D.A. Huse: Phys. Rev. Lett. 84(2004), 036802\n48) R. Cˆ ot´ e, L. Brey, and A. H. MacDonald: Phys. Rev. B 46(1992), 10239\n49) N. E. Bonesteel, I. A. McDonald, and C. Nayak: Phys. Rev. L ett.77(1996), 3009\n50) J. Schliemann, S. M. Girvin and A. H. MacDonald: Phys. Rev . Lett.86(2001), 1849\n51) Y. B. Kim, C. Nayak, E. Demler, N. Read, and S. Das Sarma: Ph ys. Rev. B 63(2001),\n205315\n52) A. Stern and B.I. Halperin: Phys. Rev. Lett. 88(2002), 106801\n53) K. Nomura and D. Yoshioka: Phys. Rev. B 66(2002), 153310\n54) S. H. Simon, E. H. Rezayi and M. V. Milovanovic: Phys. Rev. Lett.91(2003), 046803\n55) R. D. Wiersma, J. G. S. Lok, S. Kraus, W. Dietsche, K. vonKl itzing, D. Schuh, M. Bichler,\nH.-P. Tranitz and W. Wegscheider: Phys. Rev. Lett. 93(2004), 266805" }, { "title": "0906.2668v1.Schrödinger_equations_for_the_square_root_density_of_an_eigenmixture_and___the_square_root_of_an_eigendensity_spin_matrix.pdf", "content": "arXiv:0906.2668v1 [nucl-th] 15 Jun 2009Schr¨ odinger equations for the square root density\nof an eigenmixture and the square root of an\neigendensity spin matrix\nB. G. Giraud\nbertrand.giraud@cea.fr, Institut de Physique Th´ eorique,\nand\nP. Moussa\npierre.moussa@cea.fr, Institut de Physique Th´ eorique,\nDSM, CE Saclay, F-91191 Gif/Yvette, France\nOctober 28, 2018\nAbstract\nWe generalize a “one eigenstate” theorem of Levy, Perdew and Sahni\n(LPS) [1] to the case of densities coming from eigenmixture d ensity op-\nerators. The generalization is of a special interest for the radial density\nfunctional theory (RDFT) for nuclei [2], a consequence of th e rotational\ninvariance of the nuclear Hamiltonian; when nuclear ground states (GSs)\nhave a finite spin, the RDFT uses eigenmixture density operat ors to sim-\nplify predictions of GS energies into one-dimensional, rad ial calculations.\nWealsostudySchr¨ odingerequationsgoverningspineigend ensitymatrices.\nThe theorem of Levy, Perdew and Sahni [1] may be described as follo ws:\ni) letHAbe a Hamiltonian for Aidentical particles, with individual mass m,\nHA=A/summationdisplay\ni=1[−¯h2∆/vector ri/(2m)+u(/vector ri)]+A/summationdisplay\ni>j=1v(/vector ri,/vector rj), (1)\nii) consider a GS eigenfunction of HA, ψ(/vector r1,σ1,/vector r2,σ2,...,/vector rA,σA),whereσide-\nnotes the spin state of the particle with space coordinates /vector ri,\niii) use a trace of |ψ∝an}b∇acket∇i}ht∝an}b∇acketle{tψ|upon all space coordinates but the last one, and upon\nall spins, to define the density,\nρ(/vector r) =A/summationdisplay\nσ1...σA/integraldisplay\nd/vector r1d/vector r2...d/vector rA−1|ψ(/vector r1,σ1,/vector r2,σ2,...,/vector rA−1,σA−1,/vector r,σA)|2,(2)\niv) then there exists a local potential veff(/vector r) so that,\n[−¯h2∆/vector r/(2m)+veff(/vector r)]/radicalbig\nρ(/vector r) = (EA−EA−1)/radicalbig\nρ(/vector r), (3)\n1where the eigenvalue is the difference of the GS energy EAof theA-particle\nsystem and that, EA−1,of the (A−1)-particle one.\nCan this theorem be generalized for densities derived from eigenope rators,\nD ∝/summationtextN\nn=1wn|ψn∝an}b∇acket∇i}ht∝an}b∇acketle{tψn|,corresponding to cases where Hhas several ( N>1)\ndegenerate GSs ψn? The degeneracy situation is of a wide interest in nuclear\nphysics for doubly odd nuclei, the GSs of which often have a finite spin , and,\nif only because of Kramer’s degeneracy, for odd nuclei. In particula r, because\nof the rotational invariance of the nuclear Hamiltonian, the density operator of\ninterestforthe RDFT [2]reads, D=/summationtext\nM|ψJM∝an}b∇acket∇i}ht∝an}b∇acketle{tψJM|/(2J+1),whereJandM\nare the usual angular momentum numbers of a degenerate magnet ic multiplet\nof GSsψJM.Actually, more generally, it will easily be seen that the argument\nwhich follows holds for a degenerate multiplet of excited states as we ll.\nThis paper proves the generalization, by closely following the argume nt used\nfor one eigenstate only [1]. Furthermore, there is no need to assum e identi-\ncal particles. No symmetry or antisymmetry assumption for eigenf unctions is\nneeded. Let /vector piand/vector σibe the momentum and spin operators for the ith particle,\nat position /vector ri.Single out the A-th particle, with its degrees of freedom labelled\n/vector rand/vector σrather than /vector rAand/vector σA.For a theorem of maximal generality, with dis-\ntinct masses, one-body and two-body potentials, our Hamiltonian m ay become,\nHA=HA−1+VA+hA,with\nHA−1=A−1/summationdisplay\ni=1[−¯h2∆/vector ri/(2mi)+ui(/vector ri,/vector pi,/vector σi)]+A−1/summationdisplay\ni>j=1vij(/vector ri,/vector pi,/vector σi,/vector rj,/vector pj,/vector σj),\nVA=A−1/summationdisplay\nj=1vAj(/vector r,/vector rj,/vector pj,/vector σj), hA=−¯h2∆/vector r/(2mA)+uA(/vector r). (4)\nThe potentials acting upon the first ( A−1) particles may be non local and spin\ndependent, but, for a technical reason which will soon become obv ious, those\npotentials acting upon the A-th particle in VAandhAmust be strictly local\nand independent of the A-th spin. For notational simplicity, we choose units so\nthat ¯h2/(2mA) = 1 from now on.\nAs in the one eigenstate case [1] we select situations where there ex ists a\nrepresentation in which, simultaneously, the Hermitian Hamiltonian HAand all\nthe eigenfunctions ψ(/vector r1,σ1,...,/vector rA,σA) under consideration are real. This reality\ncondition does not seem to be restrictive, in view of time reversal inv ariance.\nLetEAbe a degenerate eigenvalue of HA.The degeneracy multiplicity being\nlarger than 1 ,selectN ≥2 of the corresponding eigenfunctions ψn,orthonor-\nmalized. Their set may be either complete or incomplete in the eigensub space.\nThe density operators,\nD=N/summationdisplay\nn=1|ψn∝an}b∇acket∇i}htwn∝an}b∇acketle{tψn|,N/summationdisplay\nn=1wn= 1, (5)\nwith otherwise arbitrary,positive weights wn,arenormalizedto unity, Tr D= 1,\nin theA-body space. They are eigenoperators of HA,namelyHAD=EAD.\n2The partial trace of a Dupon the first ( A−1) coordinates and all Aspins,\nτ(/vector r) =N/summationdisplay\nn=1wn/summationdisplay\nσ1...σA−1σ/integraldisplay\nd/vector r1...d/vector rA−1[ψn(/vector r1,σ1,...,/vector rA−1,σA−1,/vector r,σ)]2,(6)\ndefines a “density” τ, normalized so that/integraltext\nd/vector r τ(/vector r) = 1.Let nowφn/vector rσbe, in\nthe space of the first ( A−1) particles, an auxiliary wave function defined by,\nφn/vector rσ(/vector r1,σ1,...,/vector rA−1,σA−1) =ψn(/vector r1,σ1,...,/vector rA−1,σA−1,/vector r,σ)//radicalbig\nτ(/vector r).(7)\nNote that this auxiliary wave function depends on the choice of the w eightswn.\nNow the density operator in the space of the first ( A−1) particles, D′\n/vector r=/summationtext\nnσ|φn/vector rσ∝an}b∇acket∇i}htwn∝an}b∇acketle{tφn/vector rσ|,is normalized, Tr′D′\n/vector r= 1,where the symbol Tr′means\nintegration upon the first ( A−1) coordinates and sum upon the first ( A−1)\nspins. Since this normalization of D′\n/vector rin the (A−1)-particle space does not\ndepend on/vector r,two trivial consequences read, ∇/vector rTr′D′\n/vector r= 0 and ∆ /vector rTr′D′\n/vector r= 0.\nMore explicitly, this gives,\n/summationtext\nnσσ1...σA−1/integraltext\nd/vector r1...d/vector rA−1wnφn/vector rσ(/vector r1,σ1,...,/vector rA−1,σA−1)×\n∇/vector rφn/vector rσ(/vector r1,σ1,...,/vector rA−1,σA−1) = 0,(8)\nand\n/summationdisplay\nnσσ1...σA−1/integraldisplay\nd/vector r1...d/vector rA−1wn{[∇/vector rφn/vector rσ(/vector r1,σ1,...,/vector rA−1,σA−1)]2+\nφn/vector rσ(/vector r1,σ1,...,/vector rA−1,σA−1)∆/vector rφn/vector rσ(/vector r1,σ1,...,/vector rA−1,σA−1)}= 0.(9)\nThen one can rewrite the eigenstate property, ( HA−EA)ψn= 0,into,\n(HA−1+VA+hA−EA)√τ φn/vector rσ= 0. (10)\nThis also reads,\n√τ(HA−1+VA+uA−EA)φn/vector rσ−(∆/vector r√τ)φn/vector rσ=\n2(∇/vector r√τ)·(∇/vector rφn/vector rσ)+√τ(∆/vector rφn/vector rσ). (11)\nThe right-hand side (rhs) of Eq. (11) occurs because the Laplacia n, ∆/vector r,present\ninhA,acts also upon the parameter, /vector r,ofφn/vector rσ.This is where the local, multi-\nplicative nature of uAandvAjin the last particle space is used and avoids the\noccurrence of further terms, that would induce a somewhat unwie ldy theory.\nDefine, for any integrand Ψ n/vector rσ,the following expectation value in the first\n(A−1)-particle space,\n∝an}b∇acketle{t∝an}b∇acketle{tΨn/vector rσ∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht=/summationdisplay\nσ1...σA−1/integraldisplay\nd/vector r1...d/vector rA−1Ψn/vector rσ(r1,σ1,...,rA−1,σA−1).(12)\n3Multiply Eq. (11) by φn/vector rσand integrate out the first ( A−1) coordinates and\nspins, to obtain,\n∝an}b∇acketle{t∝an}b∇acketle{tφ2\nn/vector rσ∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht[Eexc\nnσ(/vector r)+EA−1+Unσ(/vector r)+uA(/vector r)−EA−∆/vector r]/radicalbig\nτ(/vector r)\n= 2∝an}b∇acketle{t∝an}b∇acketle{tφn/vector rσ(∇/vector rφn/vector rσ)∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht·[∇/vector r/radicalbig\nτ(/vector r)]+∝an}b∇acketle{t∝an}b∇acketle{tφn/vector rσ(∆/vector rφn/vector rσ)∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht/radicalbig\nτ(/vector r),(13)\nwhereEexc\nnσ(/vector r) is defined from,\n∝an}b∇acketle{t∝an}b∇acketle{tφ2\nn/vector rσ∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht[Eexc\nnσ(/vector r)+EA−1] =/summationtext\nσ1...σA−1/integraltext\nd/vector r1...d/vector rA−1\nφn/vector rσ(r1,σ1,...,rA−1,σA−1) [HA−1φn/vector rσ](r1,σ1,...,rA−1,σA−1),(14)\nandUnσ(/vector r) results from,\n∝an}b∇acketle{t∝an}b∇acketle{tφ2\nn/vector rσ∝an}b∇acket∇i}ht∝an}b∇acket∇i}htUnσ(/vector r) =/summationtextA−1\nj=1/summationtext\nσ1...σA−1/integraltext\nd/vector r1...d/vector rA−1\nφn/vector rσ(r1,σ1,...,rA−1,σA−1)vAj(/vector r,/vector rj,/vector pj,/vector σj)φn/vector rσ(r1,σ1,...,rA−1,σA−1).(15)\nThe square norm of φn/vector rσin the (A−1)-particle space results from Eq. (12)\nwith Ψ =φ2.In Eq. (14) the expectation value of HA−1is explicited as the\nsum of the GS energy EA−1ofHA−1and a positive, excitation energy Eexc\nnσ(/vector r).\nFrom Eq. (15), the Hartree nature of the potential Unσ(/vector r) is transparent.\nKeeping in mind that, ∀/vector r,the density operator D′is normalized to unity,\nnamely, that/summationtext\nnσwn∝an}b∇acketle{t∝an}b∇acketle{tφ2\nn/vector rσ∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht= 1,multiply Eq. (13) by wnand perform the\nsum uponnandσ.This gives,\n[Uexc(/vector r)+EA−1+UHrt(/vector r)+uA(/vector r)−EA−∆/vector r]√τ=/summationtext\nnσwn[2∝an}b∇acketle{t∝an}b∇acketle{tφn/vector rσ(∇/vector rφn/vector rσ)∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht·(∇/vector r√τ)+∝an}b∇acketle{t∝an}b∇acketle{tφn/vector rσ(∆/vector rφn/vector rσ)∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht√τ],(16)\nwhere the “mixture excitation potential”,\nUexc(/vector r) =/summationdisplay\nnσwn∝an}b∇acketle{t∝an}b∇acketle{tφ2\nn/vector rσ∝an}b∇acket∇i}ht∝an}b∇acket∇i}htEexc\nnσ(/vector r), (17)\nis local and positive and the “mixture Hartree-like potential”,\nUHrt(/vector r) =/summationdisplay\nnσwn∝an}b∇acketle{t∝an}b∇acketle{tφ2\nn/vector rσ∝an}b∇acket∇i}ht∝an}b∇acket∇i}htUnσ(/vector r), (18)\nis also local. Because of the frequent dominance of attractive term s invAj,\nit may show more negative than positive signs. Then notice that, bec ause of\nEqs. (8), and Eq. (12) with Ψ = φ∇φ,the sum in the rhs of Eq. (16),/summationtext\nnσwn∝an}b∇acketle{t∝an}b∇acketle{tφn/vector rσ(∇/vector rφn/vector rσ)∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht,vanishes. Note also, from Eqs. (9), and Eq. (12)\nwith Ψ =φ∆φ,that, again for the rhs of Eq. (16), the following equality holds,\n−/summationdisplay\nnσwn∝an}b∇acketle{t∝an}b∇acketle{tφn/vector rσ(∆/vector rφn/vector rσ)∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht=/summationdisplay\nnσwn∝an}b∇acketle{t∝an}b∇acketle{t(∇/vector rφn/vector rσ)·(∇/vector rφn/vector rσ)∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht,(19)\nwhere, again, the symbol ∝an}b∇acketle{t∝an}b∇acketle{t ∝an}b∇acket∇i}ht∝an}b∇acket∇i}htdenotes the trace T r′,an integration upon the\nfirst (A−1) coordinates together with summation upon their spins. The rhs o f\nthis equation, Eq. (19), defines a positive, local potential,\nUkin(/vector r) =/summationdisplay\nnσwn∝an}b∇acketle{t∝an}b∇acketle{t(∇/vector rφn/vector rσ)·(∇/vector rφn/vector rσ)∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht. (20)\n4Finally, according to Eq. (16), the sum of local potentials, Ueff=Uexc+\nUHrt+Ukin+uA,drives a Schr¨ odinger equation for√τ,\n[−∆/vector r+Ueff(/vector r)]/radicalbig\nτ(/vector r) = (EA−EA−1)/radicalbig\nτ(/vector r). (21)\nThis is the expected generalizationofthe LPS theorem. Note that, if the (A−1)\nparticles are not identical, then EA−1,the GS energy of HA−1, means here\nthe mathematical, absolute lower bound of the operator in all subsp aces of\narbitrarysymmetry or lack of symmetry. In practice, however, m ost cases imply\nsymmetriesin the ( A−1)-space, and EA−1means the groundstate energyunder\nsuch symmetries.\nFor nuclear physics, this generalization can be used in two ways:\ni) The first one consists in considering hypernuclei or mesonic nuclei, where the\nA-th particle is indeed distinct. Theoretical calculations with local inte ractions\nforthedistinctparticlemaybeattemptedwhilenonlocaland/orspin dependent\ninteractionsforthe A−1 nucleons areuseful, ifnot mandatory. Then, obviously,\nthe density τrefers to the hyperon or the meson and, given the neutron and\nproton respective numbers NandZ,wave functions ψnandφn/vector r,σbelong to\nbothN- andZ-antisymmetric subspaces. The energy EA−1is the GS energy of\nnucleus{N,Z},a fermionic GS energy, rather than the absolute lower bound\nof the mathematical operator HA−1in all subspaces.\nii) The secondone consists in setting all Aparticlesto be nucleons, at the cost of\nrestricting theoretical models to local interactions. Such models a re not without\ninterest indeed, although interactions which are spin dependent ar e certainly\nmore realistic. The antisymmetric properties of the functions ψnare requested\nin both N- and Z-spaces. If the singled out, A-th coordinate is a neutron one,\nthe density τdefined by Eq. (6) is the usual neutron density, divided by N; the\nfunctionsφn/vector rσareantisymmetricinthe( N−1)-neutronspaceandthe Z-proton\nspace; the energy EA−1now means the fermionic GS of nucleus {N−1,Z},not\nthat absolute, mathematical lower bound of operator HN−1,Z.Conversely,if the\nA-th coordinate is a proton one, then, mutatis mutandis ,τis the usual proton\ndensity, divided by Z,andEA−1is the GS energy of nucleus {N,Z−1}.\nIn both cases, the Hamiltonians to be used are scalars under the ro tation\ngroup, and, therefore [2], the density operators Dconsidered by the RDFT are\nalso scalars. Hence, all calculations defining Ueffreduce to calculations with a\nradial variable ronly.\nWe shall now extend our previous results to the case where we allow s pin\ndependence for all interactions, a most useful feature if all Aparticles are nucle-\nons. Polarized eigenmixtures are also interesting and need also be co nsidered.\nHence, a generalisation of our approach, which uses the “spin-den sity matrix”\n(SDM) formalism [3] [4], is in order. The Hamiltonian may become,\nA/summationdisplay\ni=1[−¯h2∆/vector ri/(2mi)+ui(/vector ri,/vector σi)]+A/summationdisplay\ni>j=1vij(/vector ri,/vector σi,/vector rj,/vector σj).(22)\nIt allows subtle differences between neutrons and protons, beside s the Coulomb\ninteractions between protons. More explicitly, there can be two dis tinct one-\n5body potentials, un,up,namely one forneutrons and one forprotons, but within\ntheneutronspacethefunction un(/vector ri,σi)obviouslywillnotread uni(/vector ri,σi).Simi-\nlarly in the proton space, the Hamiltonian contains terms up(/vector ri,σi) rather than\nupi(/vector ri,σi.The same subtlety allows terms vpp(/vector ri,σi,/vector rj,σj), vpn(/vector ri,σi,/vector rj,σj),\nvnp(/vector ri,σi,/vector rj,σj) andvpp(/vector ri,σi,/vector rj,σj),rather than vppij(/vector ri,σi,/vector rj,σj),... etc.\n(Of course, vpn=vnp.) But non localities of potentials and interactions, in the\nsense of explicit dependences upon momenta pi,remain absent.\nThen theA-th particle is again singled out, with degrees of freedom again\nlabelled/vector rand/vector σ,and the Hamiltonian is split as a sum, KA−1+WA+kA,some-\nwhat similar to the split described by Eqs. (4). For simplicity, we shall u se short\nnotations, K,Wandk,rather than KA−1,WAandkA.With two spin states,\nσ=±1/2,for theA-th nucleon, we represent eigenstates ψnas column vec-\ntors,ψn=/bracketleftbigg\nψn+\nψn−/bracketrightbigg\n,and operators as matrices such as, W=/bracketleftbigg\nW++W+−\nW−+W−−/bracketrightbigg\n,\nk=/bracketleftbigg\nk++k+−\nk−+k−−/bracketrightbigg\nand ¯u=/bracketleftbigg\nuA++uA+−\nuA−+uA−−/bracketrightbigg\n.The matrix, K=/bracketleftbigg\nK0\n0K/bracketrightbigg\n,is a\nscalar in spin space, since Kdoes not act upon the A-th particle.\nThe spin density matrix, ρn,results from an integration and spin sum over\nthe (A−1)-space of the matrix, ψn×ψT\nn,where the superscriptTdenotes\ntransposition,\nρn(/vector r) =∝an}b∇acketle{t∝an}b∇acketle{t/bracketleftbigg\nψn+\nψn−/bracketrightbigg\n×[ψn+ψn−]∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht=∝an}b∇acketle{t∝an}b∇acketle{t/bracketleftbigg\n(ψn+)2ψn+ψn−\nψn−ψn+(ψn−)2/bracketrightbigg\n∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht.(23)\nIt depends on the last coordinate, /vector r,and its matrix elements are labelled by\ntwo values, {σ,σ′},of the last spin. For an eigenmixture one defines, obviously,\n¯θ(/vector r) =/summationtext\nnwnρn(/vector r),and the trace in the last spin space, [ θ++(/vector r)+θ−−(/vector r)],is\nthat density, τ(/vector r),defined by Eq. (6).\nThe SDM, ¯θ,is symmetric andpositive semidefinite, ∀/vector r.Exceptfor marginal\nsituations, it is also invertible, in which case there exists a unique inver se square\nroot, also symmetric and positive. Define, therefore, a column vec torφn/vector rof\nstates in the ( A−1)-space according to,\nφn/vector r=¯θ−1\n2(/vector r)ψn. (24)\nThen the following property,\n/summationdisplay\nnwn∝an}b∇acketle{t∝an}b∇acketle{tφn/vector r×φT\nn/vector r∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht=¯θ−1\n2(/vector r)/parenleftBigg/summationdisplay\nnwn∝an}b∇acketle{t∝an}b∇acketle{tψn×ψT\nn∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht/parenrightBigg\n¯θ−1\n2(/vector r) =¯1,(25)\nholds∀/vector r.Here¯1denotes the identity matrix. Hence, the following gradient and\nLaplacian properties also hold,\n/summationdisplay\nnwn∝an}b∇acketle{t∝an}b∇acketle{t/parenleftbigg\n∇/vector r/bracketleftbigg\nφn/vector r+\nφn/vector r−/bracketrightbigg/parenrightbigg\n×[φn/vector r+φn/vector r−]∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht+\n/summationdisplay\nnwn∝an}b∇acketle{t∝an}b∇acketle{t/bracketleftbigg\nφn/vector r+\nφn/vector r−/bracketrightbigg\n×(∇/vector r[φn/vector r+φn/vector r−] )∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht= 0,∀/vector r,(26)\n6and\n/summationtext\nnwn∝an}b∇acketle{t∝an}b∇acketle{t/parenleftbigg\n∆/vector r/bracketleftbigg\nφn/vector r+\nφn/vector r−/bracketrightbigg/parenrightbigg\n×[φn/vector r+φn/vector r−]∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht+\n2/summationtext\nnwn∝an}b∇acketle{t∝an}b∇acketle{t/parenleftbigg\n∇/vector r/bracketleftbigg\nφn/vector r+\nφn/vector r−/bracketrightbigg/parenrightbigg\n·(∇/vector r[φn/vector r+φn/vector r−])∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht+\n/summationtext\nnwn∝an}b∇acketle{t∝an}b∇acketle{t/bracketleftbigg\nφn/vector r+\nφn/vector r−/bracketrightbigg\n×(∆/vector r[φn/vector r+φn/vector r−] )∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht= 0,∀/vector r . (27)\nThe eigenvector property,/parenleftbig\nK+W+k−EA¯1/parenrightbig\nψn= 0,also reads,\n/parenleftbig\nK+W+ ¯u−EA¯1/parenrightbig¯θ1\n2(/vector r)φn/vector r−/bracketleftBig\n∆/vector r¯θ1\n2(/vector r)/bracketrightBig\nφn/vector r=\n2/bracketleftBig\n∇/vector r¯θ1\n2(/vector r)/bracketrightBig\n·/parenleftbig\n∇/vector rφn/vector r/parenrightbig\n+¯θ1\n2(/vector r)/parenleftbig\n∆/vector rφn/vector r/parenrightbig\n. (28)\nRight-multiply Eq. (28) by the row vector, φT\nn/vector r,integrate and sum over the\n(A−1)-space, weigh the result by wnand sum upon n.Because of Eq. (25),\nthe weighted sum of averages over the ( A−1)-space simplifies into,\n/bracketleftBig\nUexc(/vector r)+(EA−1−EA)¯1+UHrt(/vector r)+ ¯u(/vector r)−∆/vector r/bracketrightBig\n¯θ1\n2(/vector r) =/summationtext\nnwn×\n/braceleftBig\n2/bracketleftBig\n∇/vector r¯θ1\n2(/vector r)/bracketrightBig\n· ∝an}b∇acketle{t∝an}b∇acketle{t/parenleftbig\n∇/vector rφn/vector r/parenrightbig\n×φT\nn/vector r∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht+¯θ1\n2(/vector r)∝an}b∇acketle{t∝an}b∇acketle{t/parenleftbig\n∆/vector rφn/vector r/parenrightbig\n×φT\nn/vector r∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht/bracerightBig\n,(29)\nwith\nUexc(/vector r) =/summationdisplay\nnwn∝an}b∇acketle{t∝an}b∇acketle{t/parenleftbig\nKφn/vector r/parenrightbig\n×φT\nn/vector r∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht−EA−1¯1, (30)\nand\nUHrt(/vector r) =/summationdisplay\nnwn∝an}b∇acketle{t∝an}b∇acketle{t/parenleftbig\nWφn/vector r/parenrightbig\n×φT\nn/vector r∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht. (31)\nThe rhs of Eq. (29) can be simplified, but less than that of Eq. (16). Indeed\nEq. (26) does not imply that the coefficient of ∇/vector r¯θ1\n2(/vector r) vanishes. In fact this\ncoefficient is,\n/summationdisplay\nnwn∝an}b∇acketle{t∝an}b∇acketle{t/parenleftbig\n∇/vector rφn/vector r/parenrightbig\n×φT\nn/vector r∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht=/summationdisplay\nnwn/bracketleftbigg\n0 ∝an}b∇acketle{t∝an}b∇acketle{t(∇/vector rφn/vector r+)φn/vector r−∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht\n∝an}b∇acketle{t∝an}b∇acketle{t(∇/vector rφn/vector r−)φn/vector r+∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht 0/bracketrightbigg\n,\n(32)\nand Eq. (26) shows that the matrix on the rhs is antisymmetric. In t urn, from\nEq. (27), the “Laplacian induced coefficient” in the rhs of Eq. (29) m ay be\nlisted as,\n/summationtext\nnwn∝an}b∇acketle{t∝an}b∇acketle{t/parenleftbig\n∆/vector rφn/vector r/parenrightbig\n×φT\nn/vector r∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht=\n−/summationtext\nnwn∝an}b∇acketle{t∝an}b∇acketle{t/parenleftbigg\n∇/vector r/bracketleftbigg\nφn/vector r+\nφn/vector r−/bracketrightbigg/parenrightbigg\n·(∇/vector r[φn/vector r+φn/vector r−])∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht+1\n2/summationtext\nnwn×\n∝an}b∇acketle{t∝an}b∇acketle{t/bracketleftbigg/parenleftbigg\n∆/vector r/bracketleftbigg\nφn/vector r+\nφn/vector r−/bracketrightbigg/parenrightbigg\n×[φn/vector r+φn/vector r−]−/bracketleftbigg\nφn/vector r+\nφn/vector r−/bracketrightbigg\n×(∆/vector r[φn/vector r+φn/vector r−])/bracketrightbigg\n∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht.(33)\n7In the rhs of Eq. (33), the similarity of its first term with potential Ukin,Eq.\n(19), is transparent. Also transparent is the antisymmetry of th e second term.\nWith the definitions,\nUkin(/vector r) =/summationdisplay\nnwn∝an}b∇acketle{t∝an}b∇acketle{t/parenleftbig\n∇/vector rφn/vector r/parenrightbig\n·/parenleftBig\n∇/vector rφT\nn/vector r/parenrightBig\n∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht, (34)\nUant(/vector r) =1\n2/summationdisplay\nnwn∝an}b∇acketle{t∝an}b∇acketle{t/bracketleftBig/parenleftbig\n∆/vector rφn/vector r/parenrightbig\n×φT\nn/vector r−φn/vector r×/parenleftBig\n∆/vector rφT\nn/vector r/parenrightBig/bracketrightBig\n∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht,(35)\nand\nUgrd(/vector r) = 2/summationdisplay\nnwn/bracketleftbigg\n0 ∝an}b∇acketle{t∝an}b∇acketle{t(∇/vector rφn/vector r+)φn/vector r−∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht\n∝an}b∇acketle{t∝an}b∇acketle{t(∇/vector rφn/vector r−)φn/vector r+∝an}b∇acket∇i}ht∝an}b∇acket∇i}ht 0/bracketrightbigg\n,(36)\nthe Schr¨ odinger equation for the square root of the spin density matrix reads,\n/bracketleftBig\nUexc+UHrt+ ¯u+Ukin−Uant−∆/bracketrightBig\n¯θ1\n2−Ugrd·∇¯θ1\n2= (EA−EA−1)¯θ1\n2.(37)\nThis concludes our generalizations of the LPS theorem. On the one h and,\nsee Eq. (21), we obtained for the square root density of an eigenm ixture a\nSchr¨ odinger equation, most similar to the LPS equation. On the oth er hand,\nat the cost of a slightly less simple result, we also obtained, see Eq. (3 7), an\nLPS-like equation that drives the square root of the spin density ma trix.\nThis opens a completely new zoology of local, effective potentials, of w hich\nvery little is known, but the interest of which is obvious, since they dr ive a\nreasonably easily measurable observable, the square root of an eig enmixture\ndensity, which depends on one degree of freedom /vector ronly. A connection of such\npotentials with optical potentials, or rather their real parts, is like ly, but yet\nremains an open problem.\nReferences\n[1] Mel Levy, John P. Perdue and Viraht Sahni, Phys. Rev. A 302745 (1984)\n[2] B. G. Giraud, Phys. Rev. C 78014307 (2008)\n[3] O. Gunnarson and B. J. Lundqvist, Phys. Rev. B 13, 4274 (1976)\n[4] A. G¨ orling, Phys. Rev. A 47, 2783 (1993)\n8" }, { "title": "0906.3300v1.Optimality_of_log_Hölder_continuity_of_the_integrated_density_of_states.pdf", "content": "arXiv:0906.3300v1 [math.SP] 17 Jun 2009OPTIMALITY OF LOG H ¨OLDER CONTINUITY OF THE\nINTEGRATED DENSITY OF STATES\nZHENG GAN AND HELGE KR ¨UGER\nAbstract. We construct examples, that log H¨ older continuity of the in te-\ngrated density of states cannot be improved. Our examples ar e limit-periodic.\n1.Introduction\nWe investigate optimality of the log H¨ older continuity of the integrat ed density\nof states. Let (Ω ,µ) be a probability space, T: Ω→Ω an invertible ergodic\ntransformation, and f: Ω→Ra bounded measurable function. Define a potential\nVω(n) =f(Tnω). The Sch¨ odinger operator Hω:ℓ2(Z)→ℓ2(Z) is defined by\n(1.1) Hωu(n) =u(n+1)+u(n−1)+Vω(n)u(n),\nand the integrated density of states kby\n(1.2) k(E) = lim\nN→∞/integraldisplay\nΩ/parenleftbigg1\nNtr(P(−∞,E)(Hω,[0,N−1]))/parenrightbigg\ndµ(ω),\nwhereHω,[0,N−1]denotes the restriction of Hωtoℓ2([0,N−1]). Using the Thouless\nformula, Craig and Simon showed that\nTheorem 1.1 (Craig and Simon, [4]) .There exists a constant C=C(/bardblf/bardbl∞)such\nthat\n(1.3) |k(E)−k(˜E)| ≤C\nlog|E−˜E|−1\nfor|E−˜E| ≤1\n2.\nThis is what is well known as log H¨ older continuity . We will be interested in the\noptimality of this statement in the sense that ε/mapsto→1\nlog(ε−1)cannot be replaced by\nanother function, which goes to zero faster. It was shown by Cra ig in [3], that the\nregularity cannot be improved to\nε/mapsto→1\nlog(ε−1)log(log(ε−1))β,\nwhereβ >1. However, in the case of specific dynamical systems (Ω ,µ,T), there\nexist many results, which improve the Craig–Simon result. We just me ntion two.\nFor quasi-periodic Schr¨ odinger operators, Goldstein and Schlag h ave shown in [7]\nthat the integrated density of states is H¨ older continuous and co mputed the H¨ older\nexponent, and shown that the integrated density of states is almo st everywhere\nDate: October 29, 2018.\n2000Mathematics Subject Classification. Primary 47B36; Secondary 47B80, 81Q10.\nKey words and phrases. Integrated density of states, limit periodic potentials.\nH. K. was supported by NSF grant DMS–0800100.\n12 Z. GAN AND H. KR ¨UGER\nLipschitz. For random Schr¨ odinger operators, the integrated d ensity of states is\neven everywhere Lipschitz. This is known as the Wegner estimate wh ich can be\nfound for example in the exposition of Kirsch in [6].\nOur interest in the question of optimality of the Craig–Simon results c omes from\nthe importance of the Wegner estimate in multiscale analysis (see the exposition of\nKirsch). If one could improve the result to a continuity of the form\nε/mapsto→1\nlog(ε−1)β\nfor some large enough β >1, one would be able to use this for multiscale analysis\n(see for example Theorem 3.12 in [8]). Already Craig’s result shows tha t this is\nimpossible, however one could hope that a combination of an improved continuity\nresultandanimprovementofmultiscaleanalysismightremovetheWeg nerestimate\nassumption. However, we will show that the continuity of integrate d density of\nstates cannot be improved for all potentials beyond log H¨ older con tinuity.\nA potential V∈ℓ∞(Z) is called almost-periodic, if the closure Ω of its translates\nis compact in the ℓ∞norm. Furthermore, then Ω can be made into a compact\ngroup, with an unique invariant Haar measure. For these our previo us definition of\nthe integrated density of states (1.2) can be replaced by\n(1.4) kV(E) = lim\nN→∞1\nNtr(P(−∞,E)(H[0,N−1])),\nwhereH[0,N−1]denotes now the restriction of ∆+ Vtoℓ2([0,N−1]) with ∆ the\ndiscrete Laplacian. This can be found for example as Theorem 2.9 in Av ron–Simon\n[2].\nNext,Vis calledpperiodic, if its p-th translate is equal to V. Furthermore, V\nis limit-periodic if it is the limit in the ℓ∞norm of periodic potentials. We denote\nbyσ(∆+V) the spectrum of the operator ∆+ V.\nTheorem 1.2. Given any increasing continuous function ϕ:R+→R+with\n(1.5) lim\nx→0ϕ(x) = 0\nand a constant C0>0, there is a limit-periodic Vsatisfying /bardblV/bardbl∞≤C0such that\nits integrated density of states satisfies\n(1.6) limsup\nE→E0|kV(E)−kV(E0)|log(|E−E0|−1)\nϕ(|E−E0|)=∞,\nfor anyE0∈σ(∆+V).\nThis result tells us, that with ϕas in the previous theorem, we cannot have\n|kV(E)−kV(E0)| ≤C·ϕ(|E−E0|)\nlog(|E−E0|−1)\nfor anyC >0 and allV. The proof of this theorem essentially happens in two\nparts. Given a periodic V0andε >0 satisfying /bardblV0/bardbl ≤C0−ε, we construct a\nsequenceVjof periodic potentials, with the following properties\n(i)Vjispj-periodic.\n(ii) The Lebesgue measure of σ(∆+Vj)\n(1.7) εj=|σ(∆+Vj)|OPTIMALITY OF LOG H ¨OLDER CONTINUITY OF THE IDS 3\nsatisfies\n(1.8) log( ε−1\nj)≥pj−1·pj·ϕ(2εj).\n(iii) We have that\n(1.9) /bardblVj−Vj−1/bardbl ≤min(ε,ε1,...,ε j−1)\n2j.\nHere/bardbl./bardbldenotes the ℓ∞norm. The construction of these Vjwill be given in the\nnext section and uses the tools developed by Avila in [1]. Before proce eding with\nthe proof of Theorem 1.2, recall that positivity of the trace implies t hat\nkW(E−/bardblV−W/bardbl)≤kV(E)≤kW(E+/bardblV−W/bardbl)\nfor any potentials VandW.\nProof of Theorem 1.2. By (1.9), we see that there exists a limiting potential V,\nsuch that for each j, we have that\n/bardblVj−V/bardbl ≤εj=|σ(∆+Vj)|.\nFurthermore, since /bardblV0−V/bardbl ≤ε, we have that /bardblV/bardbl ≤C0.\nNext, fixE0∈σ(∆+V) and letj≥1. By the previous equation, we have that\nthere exists E1∈σ(∆+Vj) such that\n|E0−E1| ≤εj=|σ(∆+Vj)|.\nDenote the band of σ(∆+Vj) containing E1by [E−,E+]. By a general fact about\nperiodic Schr¨ odinger operators, we know that\nkVj(E+)−kVj(E−) =1\npj.\nWe thus get that\nkV(E++εj)−kV(E−−εj)≥1\npj\nFurthermore, the interval [ E−−εj,E++εj] containsE0and we can choose Ej∈\n{E−−εj,E++εj}such that\n|kV(E0)−kV(Ej)| ≥1\n2pj\nand\n|E0−Ej| ≤2εj.\nThis implies the claim by (1.8), since jwas arbitrary. /square\nOne can slightly improve the above theorem, by for example showing t hat there\nis not only one V, that satisfies the conclusion, but that in fact the set is dense in\nthe limit-periodic operators. However, we have not done so, to kee p the statement\nas simple as possible.4 Z. GAN AND H. KR ¨UGER\n2.Construction of the periodic potentials\nWe will need the machinery developed by Avila in [1], in order to prove our\nresults. In the following, we let Ω be a totally disconnected compact g roup, known\nasCantor group . We furthermore let T: Ω→Ω be a minimal translation on this\ngroup. There is a decreasing sequence of Cantor subgroups\nX1⊇X2⊇...\nsuch that the quotients\nΩ/Xk\ncontainpkelements. We let Pkbe the subset of the continuous functions C(Ω) on\nΩ, which only depend on Ω /Xk.fis calledn-periodic if f(Tnω) =f(ω) for every\nω∈Ω. The elements of Pkwill bepkperiodic.\nWe now fix ω∈Ω. We have that {f(Tnω)}n∈Z∈ℓ∞(Z) is limit-periodic, since\nthe periodic fare dense in C(Ω). For a finite subset Fof the periodic potentials\nP=/uniontext\nk≥1Pk, we introduce the averaged Lyapunov exponent L(E,F) as\n(2.1) L(E,F) =1\n#F/summationdisplay\nf∈FL(E,f),\nwhere #Fdenotes the number of elements of F(with multiplicities) and\nL(E,f) = lim\nN→∞1\nNlog/vextenddouble/vextenddouble/vextenddouble/vextenddouble/vextenddouble1/productdisplay\nn=N/parenleftbiggf(Tnω)−E−1\n1 0/parenrightbigg/vextenddouble/vextenddouble/vextenddouble/vextenddouble/vextenddouble\nis the Lyapunov exponent of the periodic potential. For f∈C(Ω), we denote by\nΣ(f) the spectrum of the operator ∆ + f(Tnω). We will use the following two\nlemmas of Avila [1], see also [5].\nLemma 2.1 (Lemma 3.1. in [1]) .LetBbe an open ball in C(Ω), letF⊂P∩B\nbe finite, and let 0<ε<1. Then there exists a sequence FK⊂P∩Bsuch that\n(i)L(E,FK)>0wheneverE∈R,\n(ii)L(E,FK)→L(E,F)uniformly on compacts.\nLemma 2.2 (Lemma3.2. in[1]) .LetBbe an open ball in C(Ω), and letF⊂Pk∩B\nbe a finite family of sampling functions. Then for every N≥2andKsufficiently\nlarge, there exists FK⊂PK∩Bsuch that\n(i)L(E,FK)→L(E,F)uniformly on compacts,\n(ii)The diameter of FKis at mostp−10\nK,\n(iii)For everyλ∈R, if\n(2.2) inf\nE∈RL(E,F)≥δ#Fpk,\nthen for every f∈FK, the spectrum Σ(f)has Lebesgue measure at most\ne−δpK/2.\nThe construction of the Vjwill be accomplished by\nProposition 2.3. Given a continuous function ψ:R+→R+satisfying\n(2.3) lim\nx→0ψ(x) = 0,\nandp-periodicfandε>0, then there exists a ˜p-periodic function ˜f, such that\n(2.4) /bardblf−˜f/bardbl∞≤ε,OPTIMALITY OF LOG H ¨OLDER CONTINUITY OF THE IDS 5\nand\n(2.5) log( |Σ(˜f)|−1)≥˜p·ψ(|Σ(˜f)|).\nProof.By Lemma 2.1, we can find a finite family of p1-periodic potentials F1within\nB1\n2ε(f) such that\nδ1=1\n#F1p1L(E,F1) =1\n#F1p1·1\n#F1/summationdisplay\nf∈F1L(E,f)>0.\nApplying Lemma 2.2 to F1, we can get a finite family F2⊂Bε(f) of ˜p-periodic\npotentials, where we might require ˜ pto be arbitrarily large. Let ˜fbe any element\nofF2. By Lemma 2.2 (iii), we have that\n|Σ(f2)|0,wefindfrom\nEq. (20) that −π/2< ϑκ≤0. Using a trigonometric\nidentity we can transform Eq. (20) to\ntanϑκ=1\n2α/parenleftBig\n1−/radicalbig\n1+4α2/parenrightBig\n(25)\nFrom Eq. (20) we see that for finite Bandκ= 0 the\nangleϑκ=0=−π\n4. In order for ϑκto be a continuous\nfunction of κwith values in the correct range we invert\nEq. (25) as\nϑκ=/braceleftbigg\narctan/parenleftbig1\n2α/parenleftbig\n1−√\n1+4α2/parenrightbig/parenrightbig\nforκ≥0\n−π\n2+arctan/parenleftbig1\n2α/parenleftbig\n1−√\n1+4α2/parenrightbig/parenrightbig\nforκ <0\n(26)\nWith the single-particle states fully defined we can write\ndown the uniform electronic (ground state) density as\nn=/summationdisplay\nbn(b)=/summationdisplay\nb/integraldisplayd3k\n(2π)3θ(εF−ε(b)\nk)\n=1\n4π2/summationdisplay\nb/integraldisplay\ndκθ(εF−ε(b)\nκ)(εF−ε(b)\nκ) (27)\nwheren(b)is the density contribution of band b,θ(x) is\nthe Heaviside step function and the trivial integrals have\nbeen carried out in the last step. Using Eqs. (18) and\n(23), the integration limits can be determined analyti-\ncally and the remaining integral can easily be solved in\nclosed form but we refrain from giving the explicit ex-\npression here.\nSimilarly, we can compute the magnetization density.\nWe obtain for the x−andy−components\nmx(r) =m0cos(qz) (28)\nand\nmy(r) =m0sin(qz) (29)\nwhere the amplitude of the spin-density wave is\nm0=−µB\n2π2/summationdisplay\nbsign(b)\n/integraldisplay\ndκ θ(εF−ε(b)\nκ) (εF−ε(b)\nκ) sinϑκcosϑκ(30)\nand we have defined\nsign(b) =/braceleftbigg\n+1 forb= 1\n−1 forb= 2. (31)\nUsing the symmetry relation θ−κ=−π\n2−θκ(see\nEq. (26)), the z−component of the magnetization den-\nsity can be shown to vanish identically,\nmz(r) = 0. (32)\nWe point out that the vector of the magnetization den-\nsity here is parallel to the Kohn-Sham magnetic field.\nThis certainly is a consequence of the simplicity of the4\nsystem under study here. For more complicated systems\nit was shown in Ref. 9 that these quantities need not be\nparallel in noncollinear SDFT in EXX. This is an impor-\ntant difference to the noncollinear LSDA formulation of\nRef.3wherethemagnetizationdensityandtheexchange-\ncorrelationmagneticfieldarelocallyparallelbyconstruc-\ntion.\nWe now turn to the evaluation of the energy of Eq. (3)\nand note that for an electrically neutral system with uni-\nformionicandelectronicdensitiesthesumoftheionicen-\nergyEion, the electronic interactionwith the ionicpoten-\ntial/integraltext\nd3r v0(r)n(r), and the Hartree energy U[n] exactly\ncancels out. We study the system at vanishing external\nmagnetic field, B0(r)≡0, and use the exact exchange\nenergy of Eq. (10) as an approximation to the exchange-\ncorrelation energy functional. It is expected10,17that\ninclusion of correlation leads to SDW states higher in\nenergy than the paramagnetic states.\nIn our case the total energy per unit volume only con-\nsists of a kinetic and an exchange contribution, i.e.,\n˜etot=˜ts+ ˜eEXX. (33)\nThe kinetic energy per unit volume has contributions\nfrom the two bands,\n˜ts=Ts\nV=/summationdisplay\nb˜t(b)\ns, (34)\nwhere the contribution of the first band is given by\n˜t(1)\ns=1\n8π2/integraldisplay\ndκθ(εF−ε(1)\nκ)/parenleftBig\nεF−ε(1)\nκ/parenrightBig\n/parenleftbigg\nεF−ε(1)\nκ+κ2+q2\n4+2qκsin2ϑκ/parenrightbigg\n(35)\nwhile the contribution of the second band is\n˜t(2)\ns=1\n8π2/integraldisplay\ndκθ(εF−ε(2)\nκ)/parenleftBig\nεF−ε(2)\nκ/parenrightBig\n/parenleftbigg\nεF−ε(2)\nκ+κ2+q2\n4+2qκcos2ϑκ/parenrightbigg\n.(36)\nInserting the orbitals (17) and (22) into Eq. (10), the\nexchange energy per unit volume, ˜ eEXX, can also be ex-\npressed as sum of two terms,\n˜eEXX= ˜e(1)\nEXX+ ˜e(2)\nEXX. (37)\nThe first term which describes intraband exchange is,\nafter carrying out the angular integrals, given by\n˜e(1)\nEXX=−1\n32π3\n/summationdisplay\nb/integraldisplay\ndκθ(εF−ε(b)\nκ)/integraldisplay\ndκ′θ(εF−ε(b)\nκ′)\ncos2(ϑκ−ϑκ′)I(y(b)(κ),y(b)(κ′),(κ−κ′)2) (38)while the second term, describing interband exchange,\nreads\n˜e(2)\nEXX=−1\n16π3/integraldisplay\ndκθ(εF−ε(1)\nκ)/integraldisplay\ndκ′θ(εF−ε(2)\nκ′)\nsin2(ϑκ−ϑκ′)I(y(1)(κ),y(2)(κ′),(κ−κ′)2).(39)\nwhere we have defined\ny(b)(κ) = 2(εF−ε(b)\nκ). (40)\nIn Eqs. (38) and (39) we also have used the integral\nI(y1,y2,a) =/integraldisplayy1\n0dy/integraldisplayy2\n0dy′ 1/radicalbig\n(y−y′)2+2(y+y′)a+a2(41)\nwhich can be solved in closed form by transforming to\nnew integration variables z=y−y′andz′= (y+y′)/2\nand changing the integration limits accordingly. There-\nfore the calculation of the total energy only requires the\nnumerical calculation of a two-dimensional integral.\nWe have calculated the total energy per particle\netot=˜etot\nn(42)\nin the following way: we start by numerically calculating\nthe Fermi energy for a given value nof the density or,\nequivalently, the Wigner-Seitz radius\nrs=/parenleftbigg3\n4πn/parenrightbigg1/3\n, (43)\nandgivenvaluesoftheparameters BandqfromEq.(27).\nThe Fermi energy thus becomes a function of these three\nparameters,\nεF=εF(rs,q,B), (44)\nwhich is then used to evaluate the total energy per par-\nticle for these parameter values. We then have, for fixed\nrs, minimized etotas a function of the parameters qand\nBnumerically.\nIn Fig. 1 we show the total energy per electron at rs=\n5.4 for a few values of Bas function of q/kFwhere\nkF=/parenleftbigg9π\n4/parenrightbigg1/31\nrs(45)\nis the Fermi wavenumber of the uniform electron gas in\nthe paramagnetic state. The value rs= 5.4 was chosen\nbecause then i) the SDW phase is lower in energy than\nboth the paramagnetic and ferromagnetic phases and ii)\nthe amplitude of the SDW (or the KS magnetic field) is\nrelatively high such that the resulting energy differences\ncaneasilyberesolvednumerically. Weclearlyseethatfor\nthe given values of Bfor wavenumbers between q/kF≈5\n0 0.5 1 1.5 2\nq / kF-0.047-0.0469-0.0468-0.0467-0.0466etot (a.u.)µB B = 0.010 a.u.\nµB B = 0.011 a.u.\nµB B = 0.012 a.u.\nparamagnetic\n1.6 1.65 1.7 1.75\nq / kF-0.046976-0.046972-0.046968etot (a.u.)\nFIG. 1: Total energy per particle in EXX for the electron\ngas atrs= 5.4 with spin density wave as function of q/kF\nand different values of the amplitude Bof the Kohn-Sham\nmagnetic field. The straight line corresponds to the total\nenergy per particle of the paramagnetic state at this densit y.\nThe inset shows a magnification close to the minimum.\n1.5 andq/kF≈1.75the energyofthe SDW state is lower\nthan the energy of the paramagnetic state. The lowest\nenergy for this value of rsis achieved for the parameters\nµBB= 0.011 a.u. and q/kF= 1.68.\nIn Fig. 2 we show the KS single-particle dispersions\nof Eqs. (18) and (23) as well as the HF single particle\ndispersions. To obtain the latter ones we first calculate\nthe HF self energy (which is a 2 ×2 matrix in spin space)\nas\nΣHF(r,r′) =−/summationdisplay\nb/summationdisplay\nkθ(εF−ε(b)\nk)Φ(b)\nk(r)⊗Φ(b)†\nk(r′)\n|r−r′|\n(46)\nand then diagonalize the resulting HF Hamiltonian\nˆhHF=−∇2\n2+/integraldisplay\nd3r′ΣHF(r,r′)... (47)\nwhere the second term is a to be read as an integral oper-\nator. We would like to emphasize that we use the KS or-\nbitals and orbital energies to evaluate the HF self energy,\ni.e., we do notperform a selfconsistent HF calculation\nhere.\nIn Fig. 2 we show the KS and HF dispersions only\nfor theκ-coordinate, i.e., we set k/bardbl= 0. As expected,\nclose to κ/kF= 0 a direct gap opens up in the KS\nsingle-particledispersionsduetothe presenceofthe spin-\ndensity wave. The position of the Fermi energy is such\nthat not only states of the lower ( b= 1) band but also\nstates of the second ( b= 2) KS band are occupied in the\nground state.\nThe HF bands in Fig. 2 have been rigidly shifted by\na constant such that the lower HF band ( b= 1) and\nthe lower KS band equal the Fermi energy for the same\nvalue of κ/kF. It is evident that, as expected, the HF\nsingleparticledirectbandgapat κ/kF= 0ismuchlarger-2 -1 0 1 2\nκ / kF-0.100.10.20.30.40.5ε(b)(κ) (a.u.)KS band b=1\nKS band b=2\nHF band b=1\nHF band b=2\nFermi energy\nFIG. 2: Single-particle KS and HF bands at k/bardbl= 0 for the\noptimized parameter values ( µBB= 0.011 a.u. and q/kF=\n1.68 atrs= 5.4) mimizing the EXX total energy per particle\nin Fig. 1. The straight line indicates the Fermi energy and\nshows that close to κ/kF= 0 states of the second KS band\n[dashed (green) line] are occupied in the ground state. The\nKS orbitals have been used to compute the HF Hamiltonian\nand the resulting HF bands have been shifted rigidly such\nthat the lower HF and KS bands equal εFat the same value\nof|κ/kF|. The relative position of the second HF band [dash-\ndash-dotted (purple) line] indicates that also in HF states in\nboth bands will be occupied.\nthan the corresponding KS gap. Moreover, the position\nof the second HF band indicates that also in the HF\ncase there will be occupied states in the second band.\nWhile here we have calculated the HF bands using the\nDFT orbitals and orbital energies, we have confirmed13\nthat the above statement is true also for a HF energy\nminimization and the resulting HF bands are very close\nto the ones presented here.\nThe occupation of states in both single-particle bands\nis sometimes excluded in works on the SDW in the\nHartree-Fock approximation10,12and also in a numerical\ninvestigation13we have found that for the global energy\nminimum in Hartree-Fock only the lowest single-particle\nband is occupied. This has motivated us to do the mini-\nmization of the total energy in EXX also under the addi-\ntional constraint that only states of the lowest subband\nare occupied.\nSimilar to Fig. 1, in Fig. 3 we show the total energy\nper electron at rs= 5.4 for a few values of Bas function\nofq/kF. Of course, the constrained minimization leads,\nfor a given value of rs, to different optimized parameter\nvalues. Surprisingly, however, we found that the mini-\nmization constraining the occupation to the lower sub-\nbandleads to lower total energies than the ones obtained\nwith a two-band minimization. Moreover, this lower to-\ntal energy is achieved with a Slater determinant which\nhas empty states below the Fermi level.\nThis can be seen in Fig. 4 where we show the KS and\nHF energy bands at rs= 5.4 for this “one-band” mini-6\n0 0.5 1 1.5 2\nq / kF-0.047-0.0465-0.046-0.0455etot (a.u.)\nµB B = 0.019 a.u.\nµB B = 0.020 a.u.\nµB B = 0.021 a.u.\nparamagnetic1.25 1.3 1.35 1.4\nq / kF-0.04714-0.047136-0.047132etot (a.u.)\nFIG. 3: Same as Fig. 1 except that now only states in the\nlower band are allowed to be occupied. The total energy per\nparticle at the minimum is lower than when states in both\nbands are allowed to be occupied.\nmization for the optimized parameter values of µBB=\n0.020 a.u. and q/kF= 1.33. We see that there are states\nin the second KS band below the Fermi energy which,\ndue to the constraint in the minimization, remain un-\noccupied. We also note that for the one-band case the\namplitude of the minimizing Kohn-Sham magnetic field,\nand therefore also the “gap” between the two KS bands\natκ/kF= 0, is almost twice as large as in the two-band\ncase. Compared to Fig. 2, the intersection of the Fermi\nenergy with the bands ε(b)(κ) is shifted to a lower value\nof|κ|.\nTheHFbandsagainhaverigidlybeenshiftedsuchthat\nthe lower HF and KS bands intersect the Fermi energy\nat the same κ. Again, the direct HF gap at κ/kF= 0 is\nsignificantly larger than the KS gap. In contrast to the\ntwo-band case, the second HF band now is energetically\nhigher than the Fermi energy and the corresponding HF\nstate would, unlike the EXX state, have no unoccupied\nsingle-particle states below εF. Again here we have done\nonly a post-hoc evaluation of the HF bands but we have\nchecked that the statement remains valid for a selfconsis-\ntent HF calculation as well13.\nWe have optimized the EXX total energy per particle\nfor a range of rs-values once for single-particle occupa-\ntions in both energy bands and once for occupations re-\nstrictedtothelowerband. InFig.5weshowtheresulting\nphase diagram in the relevant density range. When al-\nlowing occupations in both bands, the SDW state (which\nis then a ground state Slater determinant) is lower in en-\nergy than both the paramagnetic and the ferromagnetic\nphase for rsin the range 5 .0<∼rs<∼5.46. In this case\nthe energiesareveryclosetothe energiesofthe paramag-\nnetic phase (energy differences of less than 4 ×10−5a.u.,\nsee lower panel of Fig. 5) and therefore the transition\nto the ferromagnetic phase occurs at an value of rsonly\nslightly higherthan the rs-value whereparamagneticand\nferromagnetic phase are degenerate.-2 -1 0 1 2\nκ / kF-0.100.10.20.30.40.5ε(b)(κ) (a.u.)KS band b=1\nKS band b=2\nHF band b=1\nHF band b=2\nFermi energy\nFIG. 4: Same as Fig. 2 except that the optimized parame-\nters are used which result from a minimization with occupied\nstates in the lower KS band only. For rs= 5.4 these values\nareµBB= 0.020 a.u. and q/kF= 1.33. Again, the straight\nline indicates the Fermi energy. Note that the states of the\nsecond KS band [dashed (green) line] remain unoccupied in\nthis calculation, even if their single-particle energies a re be-\nlow the Fermi level, i.e., the resulting Slater determinant is\nnota ground state of the Kohn-Sham problem. On the other\nhand, the post-hoc evaluation of the HF bands (for details se e\ncaption of Fig. 2 and the main text) indicates that the second\nHF band [dash-dash-dotted (purple) line] will remain unoc-\ncupied and the resulting HF wavefunction will be a ground\nstate Slater determinant.\nOn the other hand, restricting the single-particle occu-\npation to the lowest band, the SDW state is more stable\nthan para- and ferromagnetic state for 4 .78<∼rs<∼5.54.\nIn this case the energy differences between the param-\nagnetic and the SDW phase range to almost 4 ×10−4\na.u. (lower panel of Fig. 5), almost an order of magni-\ntude larger than in the two-band case. However, for all\nrsvalues in the stability range of the SDW phase, the\nminimizing Slater determinant in the one-band case is\nnota ground state of the Kohn-Sham problem.\nBoththe one-andtwo-bandcasesin EXXhavein com-\nmon that they predict the SDW phase to be lower in en-\nergy than the paramagnetic phase only for a restricted\nrangeofrsvalues. ThisisdifferentfromtheHartree-Fock\ncase10,12where the SDW phase is more stable than the\nparamagnetic phase for allvalues of rs. This is not com-\npletely surprising since due to the additional constraint\nof local Kohn-Sham potentials vsandBsin the EXX\nminimization, the resulting energies have to be higher\nthan the Hartree-Fock total energies. Since for small val-\nues ofrsthe SDW total energies in HF are extremely\nclose to the total energies of the paramagnetic phase13,\nthe higher EXX total energies can easily lead to a more\nstable paramagnetic phase.\nIn Fig.6 weshowthe SDW parameters q(upper panel)\nandB(middle panel) forwhich the EXXtotal energyper\nparticle is minimized in the one- and two-band cases for\nthosers-values for which the SDW phase is more stable7\n-0.0475-0.04725-0.047-0.04675etot (a.u.)paramagnetic\nferromagnetic\nSDW: two bands occupied\nSDW: one band occupied\n4.8 5 5.2 5.4\nrs (a0)1e-071e-061e-050.0001∆etot (a.u.)ePM-eSDW (2 bands)\nePM-eSDW (1 band)\nFIG. 5: Upper panel: Total energy per particle in EXX for\ndifferent phases of the uniform electron gas as function of\nWigner-Seitz radius rs. In the SDW phase, in one case the\noccupation of single-particle states in both bands is allow ed\nwhile in the other case the occupied states are restricted to\nthe lower band. Lower panel: energy difference between the\ntotal energies per particle of the paramagnetic phase and th e\nSDW phase for SDWs with occupied states in one and two\nbands. For the two-band case, the SDW phase is lower in\nenergy than both the paramagnetic and the ferromagnetic\nphase for 5 .0<∼rs<∼5.46. For the one-band case the range of\nstability of the SDW phase is 4 .78<∼rs<∼5.54.\nthan both paramagnetic and ferromagnetic phases. For\nthe one-band case, the wavenumber qof the spin-density\nwave covers almost the whole range between kFand 2kF\nwhile for the two-band case this range is much narrower.\nThe amplitudes Bandm0of the Kohn-Sham magnetic\nfield (middle panel) andthe magnetizationdensity (lower\npanel) of the SDW are significantly smaller in the two-\nband case as for the case with occupied single-particle\nstates in one band only. It is sometimes assumed10that\nthe wavenumber of the SDW is close to 2 kF. Our results\nshow that this need not be the case, as in the one-band\ncaseqapproaches kFfor densities at the lower end of the\nstability range of the SDW phase. However, neither in\nEXX nor in HF13have we ever found a stable SDW state\nwith wavenumber lower than kF.\nIV. SELFCONSISTENCY\nIn the previous Section we have used an ansatz for\nthe Kohn-Sham orbitals in the SDW phase which de-\npends on two parameters and then minimized the EXX\ntotal energy per particle with respect to these param-\neters. We have done this minimization once allowing\nsingle-particle states in both bands to be occupied and\nonce for occupations only in the lower band. This is dif-\nferent from the usual way of applying DFT where one\ncalculates the exchange-correlation potentials and solves\nthe Kohn-Sham equation self-consistently. In our case,11.251.51.75q/kFone band occupied\ntwo bands occupied\n012µB B (10-2 a.u.)\n4.8 5 5.2 5.4\nrs (a0)024m0 (10-4 a.u.)\nFIG. 6: Optimized values for the parameters q(upper panel)\nandB(middle panel) for which the EXXtotal energy perpar-\nticle of the SDW phase is minimized. The results are shown\nover the range of rs-values for which the SDW phase is lower\nin energy than both the paramagnetic and the ferromagnetic\nphases for the cases when both bands or only one band are\noccupied. Lower panel: amplitude of the SDW (Eq. (30)) for\nthe one- and two-band case.\nthe calculation of the EXX potentials requires solution\nof the OEP equations (13) and (14). In the present Sec-\ntion we still use our ansatz for the Kohn-Sham orbitals\nandinvestigateifitis consistentwith the OEPequations.\nWe start by calculating the orbital shifts of Eq. (15) in\nthe EXX approximation. Inserting our ansatz after some\nstraightforward algebra we obtain for the orbital shift of\nthe first band\nΨ(1)\nk(r) =G(k)\nε(1)\nκ−ε(2)\nκΦ(2)\nk(r) (48)\nwhile the shift for the second band reads\nΨ(2)\nk(r) =−G(k)\nε(1)\nκ−ε(2)\nκΦ(1)\nk(r) (49)\nwhere we have defined\nG(k) =−qκsinϑκcosϑκ+F(k) (50)\nas well as\nF(k) =/summationdisplay\nbsign(b)/integraldisplayd3k1\n(2π)3θ(εF−ε(b)\nk1)\n4π\n|k−k1|2sin(ϑκ1−ϑκ)cos(ϑκ1−ϑκ).(51)\nInserting the orbital shifts (Eqs. (48) and (49)) as well as\nthe orbitals (Eqs. (17) and (22)) it is straigthforward to\nsee that the first OEP equation (13) is satisfied by our\nansatz, i.e.,\n/summationdisplay\nb/summationdisplay\nkθ(εF−ε(b)\nk)/parenleftBig\nΦ(b)†\nk(r)Ψ(b)\nk(r)+c.c./parenrightBig\n= 0.(52)8\nThis can easily be understood from the physical content\nof the OEP equations: if we start from the KS Hamilto-\nnianasnon-interactingreferenceandperformaperturba-\ntion expansion of the interacting density in the perturba-\ntionˆWClb−ˆVxc−µBσˆBxc, whereˆWClbis the operatorof\nthe electron-electroninteraction and ˆVxc+µBσˆBxcis the\noperator of the KS exchange-correlation potentials, then\nthe OEP equation in EXX simply says that the density\nremains unchangedto first order. In our case we keep the\ndensity fixed and therefore the OEP equation (13) holds.\nAsimilarargumentcanbe usedfor the z-componentof\nthe OEP equation (14) which says that the z-component\nm(r) of the magnetization density remains unchanged\nunderthe sameperturbationtofirstorder. Thisequation\nreads explicitly\n/summationdisplay\nb/summationdisplay\nkθ(εF−ε(b)\nk)/parenleftBig\nΦ(b)†\nk(r)σzΨ(b)\nk(r)+c.c./parenrightBig\n=−2/summationdisplay\nbsign(b)\n/integraldisplayd3k\n(2π)3θ(εF−ε(b)\nk)sinϑκcosϑκG(k)\nε(1)\nκ−ε(2)\nκ= 0 (53)\nwhere the last equality can most easily be seen by noting\nthat the integrand is an odd function under the transfor-\nmationκ→ −κandalltheintegralsareoverasymmetric\nrange around κ= 0.\nFor thex−andy−component of Eq. (14) we obtain\n/summationdisplay\nb/summationdisplay\nkθ(εF−ε(b)\nk)/parenleftBig\nΦ(b)†\nk(r)σxΨ(b)\nk(r)+c.c./parenrightBig\n=J(εF,q,B)cos(qz) = 0 (54)\nand\n/summationdisplay\nb/summationdisplay\nkθ(εF−ε(b)\nk)/parenleftBig\nΦ(b)†\nk(r)σxΨ(b)\nk(r)+c.c./parenrightBig\n=J(εF,q,B)sin(qz) = 0 (55)\nwhere\nJ(εF,q,B) = 2/summationdisplay\nbsign(b)\n/integraldisplayd3k\n(2π)3θ(εF−ε(b)\nk)/parenleftbig\ncos2ϑκ−sin2ϑκ/parenrightbig\nG(k)\nε(1)\nκ−ε(2)\nκ(56)\nSince Eqs. (54) and (55) have to be satisfied for all values\nofz, we obtain only the condition\nJ(εF,q,B) = 0, (57)\ni.e., the two OEP equations are not independent.\nIn Fig. 7 we show Jof Eq. (56) for rs= 5.4 as function\nofq/kFfor different values of Bboth for the case of\noccupations in both bands (upper panel) as well as for\noccupations restricted to the lower band (lower panel).\nNote that for the latter case the sum over bands bboth\nin Eq. (56) as well as in Eq. (51) only extends over the\nlower band, b= 1.-2.5-2-1.5-1-0.50J (10-5 a.u.)\nµB B = 0.010 a.u.\nµB B = 0.011 a.u.\nµB B = 0.012 a.u.\n1.25 1.3 1.35 1.4\nq / kF-0.500.5J (10-5 a.u.)\n0 0.5 1 1.5 2\nq / kF-2.502.557.5J (10-5 a.u.)\nµB B = 0.019 a.u.\nµB B = 0.020 a.u.\nµB B = 0.021 a.u.\nFIG. 7: Upper panel: Jof Eq. (56) for rs= 5.4 as function of\nq/kFanddifferentvalues of Bfor thecase with occupations in\ntwo bands. The parameter values are the same as in Fig. 1 for\nwhich a minimum in the total energy per particle was found.\nSinceJnever crosses zero, in this case the minimum of the\ntotal energy is not consistent with the solution of the OEP\nequation. Lower panel: same as above but now for occupied\nsingle particle states only in the lower band. The parameter\nvalues are the same as in Fig. 3. In contrast to the two-\nband case, now Jnot only crosses zero but also does so at\nthose values of q/kFfor which a local minimum was found\nin Fig. 3 (see inset for magnification around the intersectio ns\nwith the zero axiss). Therefore, the OEP equation in this\ncase is consistent with the minimization of the total energy\nper particle.\nIn the upper panel of Fig. 7 we choose the same values\nfor the parameter Bas used in Fig. 1 which all had local\nminima for some value of q <2kF. For these values of\nB, however, Eq. (57) is not satisfied for any value of q\nin that range. We therefore have to conclude that in the\ntwo-band case the energy minimization is not consistent\nwith the OEP equations .\nIn the lower panel of Fig. 7 where only single-particle\nstates of the lower band are occupied we choose the pa-\nrameters as in Fig. 3. In this case, Jnot only crosses\nzero but also does so exactly for those values of q/kF\nfor which we found local minima in the total energy per\nparticle in Fig. 7 (see inset for a magnification of the re-\ngion where Jcrosses zero). We therefore conclude that\nin the one-band case the minimization of the total energy\nis consistent with the OEP equation, i.e., our ansatz is\nselfconsistent in this case. Again we emphasize that the\nresulting Slater determinant is notthe ground state of\nthe KS system.\nIt has been shown18that in unrestricted HF theory\nall the single-particle levels are fully occupied up to the\nFermi energy. To the best of our knowledge, a similar\nstatement has not been proven for DFT (even in EXX\napproximation)and our results indicate that it might not\nbe true in EXX. On the other hand, the proof of Ref. 18\nholds for the true, unrestricted HF ground state while in\nour case we have restricted the symmetry of our problem9\nto the SDW symmetry. It is quite conceivable that the\nfact that we find an excited-state Slater determinant as\nenergy-minimizing wavefunction hints towards an insta-\nbility of the SDW phase against further reduction of the\nsymmetry.\nV. SUMMARY AND CONCLUSIONS\nWe have investigated the SDW state of the uniform\nelectron gas within the EXX approximation of non-\ncollinear SDFT. While in the Hartree-Fock approxima-\ntion the SDW state is energetically more stable than the\nparamagneticstateforallvaluesof rs, inEXXthisisonly\ntrue for values of rslarger than a critical value. Using an\nexplicit ansatz for the spinor orbitals in the SDW state,\nwe have performed the energy minimization of the EXX\ntotal energy in two ways: (i) in the first case we used\nas non-interacting reference wavefunction a ground state\nSlater determinant with occupied single-particle orbitals\nbelonging to both single-particleenergy bands, as long as\ntheir energy is below the Fermi energy. Then the SDW\nphase is more stable than both paramagnetic and ferro-\nmagnetic phases for 5 .0<∼rs<∼5.46. (ii) In the second\ncase we required all the occupied single-particle orbitals\nin the Slater determinant to belong to the lower band.The minimizing Slater determinant in this case turns out\nto be an excited state, since orbitals with orbital ener-\ngiesbelowtheFermienergybelongingtothe secondband\nremain unoccupied. Nevertheless, for a given rsthe to-\ntal energies of the minizing SDW states are significantly\nlower than in case (i). The range of stabilty of the SDW\nphase with respect to both paramagnetic and ferromag-\nnetic phases is extended to 4 .78<∼rs<∼5.54.\nWe then have investigated if the self-consistency con-\nditions provided by the OEP equations for non-collinear\nSDFT are satisfied with our ansatz for the single-particle\norbitals. We have found that for case (i) the parameter\nvalues minimizing the EXX total energy are notconsis-\ntentwithasolutionoftheOEPequations. Incase(ii), on\nthe other hand, for the same parameter values for which\nthe EXX total energy is minimized also the OEP equa-\ntions are satisfied. In this case, the solution we found is\ntherefore selfconsistent.\nAcknowledgments\nWe would like to acknowledge useful discussions with\nIlya Tokatly, Giovanni Vignale, and Nicole Helbig.\nWe acknowledge funding by the ”Grupos Consolidados\nUPV/EHU del Gobierno Vasco” (IT-319-07).\n1U. von Barth and L. Hedin, J. Phys. C 5, 1629 (1972).\n2W. Kohn and L.J. Sham, Phys. Rev. 140, A1133 (1965).\n3J. K¨ ubler, K.-H. H¨ ock, J. Sticht, and A.R. Williams,\nJ. Phys. F 18, 469 (1988).\n4K. Capelle and L.N. Oliveira, Europhys. Lett. 49, 376\n(2000).\n5K. Capelle and L.N. Oliveira, Phys. Rev. B 61, 15228\n(2000).\n6J.D. Talman and W.F. Shadwick, Phys. Rev. A 14, 36\n(1976).\n7T. Grabo, T. Kreibich, S. Kurth, and E.K.U. Gross, in\nStrong Coulomb Correlations in Electronic Structure Cal-\nculations: Beyond Local Density Approximations , edited\nby V. Anisimov (Gordon and Breach, Amsterdam, 2000),\np. 203.\n8S. K¨ ummel and L. Kronik, Rev. Mod. Phys. 80, 3 (2008).\n9S. Sharma, J.K. Dewhurst, C. Ambrosch-Draxl, S. Kurth,\nN. Helbig, S. Pittalis, S. Shallcross, L. Nordstr¨ om, and\nE.K.U. Gross, Phys. Rev. Lett. 98, 196405 (2007).\n10G. Giuliani and G. Vignale, Quantum Theory of theElectron Liquid (Cambridge University Press, Cambridge,\n2005).\n11A.W. Overhauser, Phys. Rev. Lett. 4, 462 (1960).\n12A.W. Overhauser, Phys. Rev. 128, 1437 (1962).\n13F.G. Eich, S. Kurth, C.R. Proetto, S. Sharma, and\nE.K.U. Gross, in preparation.\n14S. Kurth and S. Pittalis, in Computational Nanoscience:\nDo It Yourself! , Vol. 31 of NIC Series , edited by J. Gro-\ntendorst, S. Bl¨ ugel, and D. Marx (John von Neumann In-\nstitute for Computing, J¨ ulich, 2006), p. 299.\n15S. K¨ ummel and J.P. Perdew, Phys. Rev. Lett. 90, 043004\n(2003).\n16S. K¨ ummel and J.P. Perdew, Phys. Rev. B 68, 035103\n(2003).\n17G.F. Giuliani and G. Vignale, Phys. Rev. B 78, 075110\n(2008).\n18V. Bach, E.H. Lieb, M. Loss, and J.P. Solovej, Phys. Rev.\nLett.72, 2981 (1994)." }, { "title": "0907.5129v1.Quantum_measures_for_density_correlations_in_optical_lattices.pdf", "content": "arXiv:0907.5129v1 [quant-ph] 29 Jul 2009Quantum measures for density correlations\nin optical lattices\nF. Benattia,b, R. Floreaniniband G. G. Guerreschia\naDipartimento di Fisica Teorica, Universit` a di Trieste, 34 014 Trieste, Italy\nbIstituto Nazionale di Fisica Nucleare, Sezione di Trieste, 34014 Trieste, Italy\nAbstract\nThedensity-densitycorrelationprofilesobtainedsuperim posingabsorptionimagesfrom\natomic clouds freely expanding after the release of the confi ning optical lattice can be\ntheoretically described in terms of a generalized quantum m easure based on coherent-\nlike states. We show that the corresponding density pattern s differ in a testable way\nfromthosecomputedusingstandardmany-bodymeanvalues, u suallyadoptedinfitting\nexperimental data.\n1 Introduction\nA standard technique used in experiments for extracting informat ion on the behavior of\nultracold atomic gases trapped in optical lattices1is based on analysis of interference phe-\nnomena (see [3]-[14] and references therein). Since the measure o f relevant observables inside\nthe optical lattice is problematic, the usual adopted procedure co nsists in the release of the\nconfining optical potentials, followed by the (free) expansion of th e atomic cloud up to meso-\nscopic sizes. At this point, the cloud is illuminated by a laser beam and th e corresponding\nabsorption image collected. The absorption process is destructive ; nevertheless, many pic-\ntures can be obtained by starting each time with a new system, prep ared in the same initial\nstate. By superimposing the various obtained pictures, informatio n on the atom density at\nthe moment of trap release can be inferred.\nTheimagethatisobtainedbysuperimposingallthese“photographs ”isusuallyinterpreted\nas the average of the density operator over the state of the sam ple. An alternate theoretical\ndescription is however possible. Indeed, in quantum mechanics, any averaging procedure\nobtained through a measuring process corresponds to a generaliz ed quantum measure, i.e.a\nso-called Positive Operator Valued Measure (POVM) [15]-[17]. In general, the choice of the\n1For recent reviews on the study of quantum many-body effects in t hese systems, see [1, 2].\n1POVMtobeusedissuggestedbytheexperimental evidences. Inth epresent caseofultracold\natoms in optical lattices, this evidence comes from experiments invo lving a two-well trapping\npotential [18]. The actual data show that interference effects ap pear in a single absorption\nimage even when the system is prepared in a totally incoherent state : more precisely, the\ninterference pattern seen in single pictures seems to always confo rm to what is expected for\na condensed, fixed-phase state, a state in which all atoms share t he same single particle wave\nfunction. By constructing a POVM in terms of these coherent-like s tates, one finds that the\ncorresponding generalized quantum measurement process leads t o predictions that, at least\nin line of principle, differ from those obtained through the simple avera ge of the density\noperator [19].2\nThe aim of the present investigation is to analyze a possible physical s cenario in which\nthose differences may become visible and experimentally detectable. To this end, we shall\nstudy the behavior of a system of cold atoms in bichromatic optical la ttices, where a second,\nlow-intensity laser is superimposed to the one forming the periodic po tential with the aim\nof obtaining an unbalanced filling of the lattice sites [20]-[23]. By exploit ing the properties\nof the density-density correlation function [24], one can show that there are experimentally\nrelevant instances in which the differences in the predictions of the t wo above mentioned\ntheoretical interpretations can be revealed. This result has been supported by a numerical\nsimulation, reproducing the situation of an actual experimental se tup. We are confident that\nthese results will stimulate further direct analysis and tests.\n2 Cold bosonic gases in optical lattices\nWe shall study the behavior of Nbosons confined in one-dimensional lattice with Msites,\neach separated by a fixed distance d.3In a suitable approximation, i.e.for a large enough\ninter-site barriers, their dynamics can be described by a Bose-Hub bard Hamiltonian [1, 2],\nH=−J/summationdisplay\nˆb†\niˆbj+/summationdisplay\niǫiˆni+1\n2U/summationdisplay\niˆni(ˆni−1), i,j= 1,2,...,M , (1)\nwhere< i,j >means nearest neighbor, ˆb†\ni,ˆbiare creation and annihilation operator for an\natom in site i, ˆni=ˆb†\niˆbiis the number operator on site i, whileJ,Uandǫiare parameters\n2The standard theoretical intepretation of these experimental f acts explains the appearence of the density\ninterference fringes as an “emergent phenomena”[7]-[11]; althoug h phenomenological in character, also this\naveraging procedure can be interpreted in terms of a quantum gen eralized measure, albeit through a rather\ninvolved, unnatural POVM.\n3In the actual experimental setups, the bosons are really confine d in an three-dimensional harmonic trap\nover which a periodic potential along say the xdirection is superimposed. Since the confining potentials are\nseparable, the single particle wave function describing the state of the atoms factorizes in a part depending\nonly on the variable xtimes a piece depending on the couple ( y,z). The dependence on the transverse\ncoordinates is irrelevant for the dynamics in the lattice, that there fore can be effectively described by a\none-dimensional Hamiltonian.\n2thatquantifythehopping, repulsionandsingle-sitedepthenergy, respectively. Theoperators\nˆb†\ni,ˆbiobey the standard Bose commutation relations: [ ˆbi,ˆb†\nj] =δij.\nThe totalnumber ofparticles Nis conserved by thedynamics generated by (1). Therefore,\nthe Hilbert space of the system is/parenleftbigN+M−1\nN/parenrightbig\n-dimensional and can be be spanned by the set\nof Fock states, |/vectork;N/an}bracketri}ht, describing the situation in which the occupation number of each site\nis fixed; the M-dimensional vector /vectork= (k1,k2,...,k M), with/summationtextM\ni=1ki=N, represents a\npossible distributions of the Natoms in the Msites. These states are obtained by acting\nwith the creation operators on the vacuum/vextendsingle/vextendsingle0/angbracketrightbig\n; explicitly, one has\n/vextendsingle/vextendsingle/vectork;N/angbracketrightbig\n≡1/radicalbig\n/vectork!(ˆb†\n1)k1(ˆb†\n2)k2...(ˆb†\nM)kM/vextendsingle/vextendsingle0/angbracketrightbig\n, (2)\nwhere in short /vectork! =k1!k2!...kM!.\nA different basis in the system Hilbert space is given by the collection of fixed-phase,\ncoherent-like states,\n/vextendsingle/vextendsingle/vectorξ,/vector ϕ;N/angbracketrightbig\n≡1√\nN!/parenleftBiggM/summationdisplay\nj=1eiϕj/radicalbig\nξjˆb†\nj/parenrightBiggN/vextendsingle/vextendsingle0/angbracketrightbig\n, (3)\nwhere/vectorξ,/vector ϕareM-dimensional vectors whose components ξj,ϕj,j= 1,2,...,M, represent sets\nof real parameters such that ϕj∈[0,2π],ξj∈[0,1] with/summationtext\njξj= 1. These states describe a\nphysical situation in which all Natoms are in a coherent superposition, where ξimeasures\nthe probability of finding an atom in the i-th site, while ϕigives the corresponding phase.\nOnly relative phases are relevant, so that one can arbitrarily fix the value of one of the ϕi.\nThe Fock states form an orthonormal set, while the coherent one s become orthogonal only\nin the large Nlimit;4nevertheless, they form an overcomplete set of states [19]:/BD=(N+M−1)!\nN!/integraldisplay2π\n0dϕ1\n2π...dϕM−1\n2π/integraldisplay1\n0dξ1.../integraldisplay1−ξ1−ξ2...−ξM−2\n1dξM−1/vextendsingle/vextendsingle/vectorξ,/vector ϕ;N/angbracketrightbig/angbracketleftbig/vectorξ,/vector ϕ;N/vextendsingle/vextendsingle.\n(4)\nBy expanding a coherent state over the Fock basis, one finds:\n/vextendsingle/vextendsingle/vectorξ;/vector ϕ;N/angbracketrightbig\n=/summationdisplay\n/vectork/radicalBigg\nN!\n/vectork!ei/vectork·/vector ϕ/parenleftBiggM/productdisplay\nj=1(ξj)kj\n2/parenrightBigg\n/vextendsingle/vextendsingle/vectork;N/angbracketrightbig\n, (5)\nwhere the sum runs over all possible M-vectors/vectork, whose components kiobey the constraint/summationtext\niki=N. The overlap between a Fock and a coherent state is then given by:\n/angbracketleftbig/vectork;N/vextendsingle/vextendsingle/vectorξ,/vector ϕ;N/angbracketrightbig\n=/radicalBigg\nN!\n/vectork!ei/vectork·/vector ϕ/parenleftBiggM/productdisplay\nj=1(ξj)kj\n2/parenrightBigg\n. (6)\n4Indeed, for an M-site lattice, one finds that: /an}bracketle{t/vectorξ, /vector ϕ;N|/vectorξ′, /vector ϕ′;N/an}bracketri}ht=/parenleftBig/summationtextM\ni=1/radicalbig\nξiξ′\niei(ϕi−ϕ′\ni)/parenrightBigN\n; using the\nCauchy-Schwartz inequality, one easily sees that the modulus of su m in the bracket is always less than one,\nunless/vectorξ=/vectorξ′and/vector ϕ=/vector ϕ′; as a consequence, its N-th power become vanishingly small as Nbecome large.\n3As well known [1, 2], the Hamiltonian (1) describes a cross-over bet ween a superfluid and\ninsulator phases, which becomes a true quantum phase transition, with order parameter\ndepending on the ratio J/U, in the limit of an infinite number Mof wells. Neglecting the\nshiftsǫi, for small J/U, the ground state of the system is given by a Fock state (Mott\ninsulator phase). On the other hand, for large J/U, the system shows phase coherence; all\nNparticles are in the same superposition and the ground state of (1) can be approximated\nby/vextendsingle/vextendsingle/vectorξ;/vector ϕ;N/angbracketrightbig\n, with definite relative phases and occupation probabilities (superflu id phase).\n3 Many-body states\nIn a typical experimental setup, the system of Natoms is first cooled to very low tempera-\ntures, of the order of few tens of nanokelvin, and then trapped in the optical lattice. Since\nmeasures of relevant observables directly in the lattice are difficult, indirect information on\nthe dynamics of theatoms areusually obtained by switching off the pe riodicconfining poten-\ntial and letting the atom gas expand freely up to mesoscopic dimensio ns. Absorption images\nof the expanded sample are then collected by shining it with a probe las er; by superimposing\nthe various obtained pictures, information on the atom density at t he moment of trap release\ncan be inferred.\nIn order to theoretically describe this process of measure, it is con venient to use a second\nquantized many-body formalism. Let us first introduce the field ope ratorˆψ†(x), creating\nfrom the vacuum an atom at position x,ˆψ†(x)|0/an}bracketri}ht=|x/an}bracketri}ht; it can be decomposed as\nˆψ†(x) =∞/summationdisplay\ni=1¯wi(x)b†\ni, (7)\nin terms of a complete set of single-particle wave functions wi(x)≡ /an}bracketle{tx|wi/an}bracketri}ht=/an}bracketle{tx|ˆb†\ni|0/an}bracketri}ht.\nAlthough the states |wi/an}bracketri}ht ≡ˆb†\ni|0/an}bracketri}ht,i= 1,2...,M, obtained by the action of the creation\noperators ˆb†\nionthevacuumareenoughtodescribeanatomconfinedinthelattice , acomplete,\ninfinite set is needed to properly represent its state when it moves f reely in space. These\nstates are orthonormal and therefore one can invert (7) and wr ite\nˆb†\ni=/integraldisplay\ndx wi(x)ˆψ†(x), (8)\nso that from [ ˆbi,ˆb†\nj] =/an}bracketle{twi|wj/an}bracketri}ht=δijone recovers the standard bosonic (equal-time) commu-\ntation relation,/bracketleftBig\nˆψ(x),ˆψ†(x′)/bracketrightBig\n=δ(x−x′).\nAssume now that the confining potential is released at time t= 0 and denote by ˆUtthe\nunitary operator that evolves freely in time the initial one-particle s tates:\n/vextendsingle/vextendsinglewi,t/angbracketrightbig\n≡ˆUt/vextendsingle/vextendsinglewi/angbracketrightbig\n=ˆUtˆb†\ni/vextendsingle/vextendsingle0/angbracketrightbig\n=ˆb†\ni(t)/vextendsingle/vextendsingle0/angbracketrightbig\n,ˆb†\ni(t)≡ˆUtˆb†\niˆU†\nt. (9)\n4The corresponding wave function is given by wi(x;t)≡ /an}bracketle{tx|wi(t)/an}bracketri}ht=/an}bracketle{tx|ˆb†\ni(t)|0/an}bracketri}ht, and coincides\nwith a transformed wi(x) under a ballistic expansion. At the moment of the release of the\nlattice, the wi(x,0),i= 1,2,...,M, are wave functions localized at the lattice sites xi:\nthey can be identified with one-dimensional Wannier functions. Afte r a large enough free\nexpansion time t, sufficient for the clouds coming from the various sites to overlap, o ne finds\nthrough (9) that these functions have a common envelope, differin g only by a phase [3, 4, 12]:\nwi(x,t) =|w(x,t)|eim\n2/planckover2pi1t(x−xi)2; (10)\nindeed, for those times, the scale over which |wi(x,t)|varies is larger than the product Md,\ngiving the original dimension of the lattice.\nSince the dynamics is free, every particle in a many-body state will ev olve independently\nwithˆUt; therefore, the evolution up to time tof thet= 0 coherent state/vextendsingle/vextendsingle/vectorξ,/vector ϕ;N/angbracketrightbig\nwill simply\nbe given by\n/vextendsingle/vextendsingle/vectorξ,/vector ϕ;N,t/angbracketrightbig\n≡1√\nN!/parenleftBiggM/summationdisplay\nj=1eiϕj/radicalbig\nξjˆb†\nj(t)/parenrightBiggN/vextendsingle/vextendsingle0/angbracketrightbig\n, (11)\nand analogously for Fock states (2)\n/vextendsingle/vextendsingle/vectork;N,t/angbracketrightbig\n=1/radicalbig\n/vectork!(ˆb†\n1(t))k1(ˆb†\n2(t))k2...(ˆb†\nM(t))kM/vextendsingle/vextendsingle0/angbracketrightbig\n. (12)\nIn this picture, only states evolve in time while operators, like ˆψ(x), remain fixed. Since\nat each instant of time tthe collection {|wi(t)/an}bracketri}ht}∞\ni=1is a complete set of single particle states\nobtained from the vacuum by the action of the creation operators ˆb†\ni(t),ˆψ(x) can be equiv-\nalently decomposed as ˆψ(x) =/summationtext∞\ni=1wi(x;t)ˆbi(t) for all times. Using this, one can compute\nthe action of the field operator ˆψ(x) on the two class of states, obtaining:\nˆψ(x)/vextendsingle/vextendsingle/vectorξ,/vector ϕ;N,t/angbracketrightbig\n=√\nN/parenleftBigg/summationdisplay\nj/radicalbig\nξjeiϕjwj(x,t)/parenrightBigg\n/vextendsingle/vextendsingle/vectorξ,/vector ϕ;N−1,t/angbracketrightbig\n, (13)\nˆψ(x)/vextendsingle/vextendsingle/vectork;N,t/angbracketrightbig\n=/summationdisplay\nj/radicalbig\nkjwj(x,t)/vextendsingle/vextendsinglek1,...,kj−1,...,kM;N−1,t/angbracketrightbig\n. (14)\n4 Generalized quantum measures\nIn order to apply the previous formalism to the theoretical interpr etation of the above men-\ntioned procedure of measuring density profiles, it is useful to reca ll some results deduced\nfrom the experiment.\nWhenM= 2 and the atoms in the lattice just before the release of the confin ing potential\nare prepared in a superfluid state described by (3), the picture th at is obtained at time\ntafter a free expansion shows a high visibility interference pattern, with fringe spacing\n5mediated by the wave vector Q=md//planckover2pi1t, wheremis the atom mass [18]. This is to be\nexpected, since in the superfluid phase essentially all Nparticles occupy the same quantum\nstate. The roughness and imperfection of the interference figur e in a single image, beside to\nexperimental errors, has to be ascribed to the finiteness of the p article number N. Indeed,\nin many-body physics, one can assimilate ensemble averages with mea n values with respect\nto macroscopically occupied many-body states, provided the numb er of particles involved is\nlarge enough. Therefore, the larger the number Nof atoms the system contains, the better\nasingleabsorption image will model the average n/vectorξ,/vector ϕ(x,t) of the density operator at point\nx,\nˆn(x)≡ˆψ†(x)ˆψ(x), (15)\nin the state/vextendsingle/vextendsingle/vectorξ,/vector ϕ;N,t/angbracketrightbig\n. Using (13), one can easily compute the theoretically predicted\ndensity profile to obtain\nn/vectorξ,/vector ϕ(x,t)≡/angbracketleftbig/vectorξ,/vector ϕ;N,t/vextendsingle/vextendsingleˆn(x)/vextendsingle/vextendsingle/vectorξ,/vector ϕ;N,t/angbracketrightbig\n=N/vextendsingle/vextendsingle/summationdisplay\nj/radicalbig\nξjeiϕjwj(x,t)/vextendsingle/vextendsingle2\n=N|w(x,t)|2/braceleftBig\n1+2/summationdisplay\nj2.\nIn other terms, the procedure of “taking a photograph” of the e xpanded sample seems to\nselect a coherent-like state/vextendsingle/vextendsingle/vectorξ,/vector ϕ;N,t/angbracketrightbig\nfor the state of the system, with definite amplitudes\nξiand phases ϕi, and as a consequence produce the profile n/vectorξ,/vector ϕ(x,t) in (16) for the average\ndensity.5As seen in experiments, amplitudes and phases nevertheless vary f rom shot to\nshot.6The occurrence of given values /vectorξand/vector ϕfor such parameters in a single shot will be\ndetermined by the initial state ρ. More precisely, the distribution of ξiandϕiover many\nabsorption pictures will be determined by the probability/angbracketleftbig/vectorξ,/vector ϕ;N,t/vextendsingle/vextendsingleρ(t)/vextendsingle/vextendsingle/vectorξ,/vector ϕ;N,t/angbracketrightbig\n, which\ngives the weight of the configuration/vextendsingle/vextendsingle/vectorξ,/vector ϕ;N,t/angbracketrightbig/angbracketleftbig/vectorξ,/vector ϕ;N,t/vextendsingle/vextendsinglein the expansion of ρ(t) in the\nbasis of coherent states.\nAs a result, in this scheme, the mean value of the density ˆ n(x), as for any other observable,\nis given by the sum over all possible configurations {/vectorξ,/vector ϕ}of the average n/vectorξ,/vector ϕ(x,t) weighted\nwith the above mentioned probability; explicitly, one then should write :\n˜nρ(x,t) =/integraldisplay\ndµ(ϕ)/integraldisplay\ndµ(ξ)/angbracketleftbig/vectorξ,/vector ϕ;N,t/vextendsingle/vextendsingleρ(t)/vextendsingle/vextendsingle/vectorξ,/vector ϕ;N,t/angbracketrightbig\nn/vectorξ,/vector ϕ(x,t), (19)\nwhere for the the properly normalized volume and integration measu re, we have used the\nshorthand notation [ cf.(4)]:\n/integraldisplay\ndµ(ϕ)/integraldisplay\ndµ(ξ)≡(N+M−1)!\nN!/integraldisplay2π\n0dϕ1\n2π...dϕM−1\n2π/integraldisplay1\n0dξ1.../integraldisplay1−ξ1−ξ2...−ξM−2\n0dξM−1.\n(20)\nThe expression (19) for the averaged density corresponds to a q uantum mechanical gen-\neralized measure. It can be described by an operation of trace of ˆ n(x) over the density\n5From a different perspective, this empirical fact has recently been the object of various investigations\n[3]-[14].\n6Unless, as mentioned before, one starts at t= 0 already with a coherent state/vextendsingle/vextendsingle/vectorξ, /vector ϕ;N/angbracketrightbig\n,i.e.in a\nsuperfluid phase.\n7matrix ˜ρobtained from the starting state ρthrough the action of a (completely positive)\nmap; explicitly:\nρ(t)/ma√sto→˜ρ(t) =/integraldisplay\ndµ(ϕ)/integraldisplay\ndµ(ξ)V(/vectorξ,/vector ϕ;N,t)ρ(t)V(/vectorξ,/vector ϕ;N,t), (21)\nwith\nV(/vectorξ,/vector ϕ;N,t) =/vextendsingle/vextendsingle/vectorξ,/vector ϕ;N,t/angbracketrightbig/angbracketleftbig/vectorξ,/vector ϕ;N,t/vextendsingle/vextendsingle; (22)\nit is a realization of a Positive Operator Valued Measure (POVM) [15]-[17].\nNoticethatthemeanvaluefortheatomdensitygivenin(19)obtaine dthroughthePOVM,\ni.e.\n˜nρ(x,t) = Tr/bracketleftBig\nˆn(x) ˜ρ(t)/bracketrightBig\n, (23)\nis clearly distinct from that obtained using the definition (17); never theless, they both repro-\nduce the profile (16) in the case of a superfluid initial state, ρ=/vextendsingle/vextendsingle/vectorξ,/vector ϕ;N/angbracketrightbig/angbracketleftbig/vectorξ,/vector ϕ;N/vextendsingle/vextendsingle, thanks\nto the (large N) orthogonality of the coherent states. On the other hand, in the case of a\nFock state, ρ/vectork=/vextendsingle/vextendsingle/vectork;N/angbracketrightbig/angbracketleftbig/vectork;N/vextendsingle/vextendsingle, the explicit evaluation of (23) gives\n˜n/vectork(x) =N\nN+MM/summationdisplay\nj=1(kj+1)|wj(x,t)|2, (24)\nwhich differs from the result (18) obtained applying (17) by an overa ll normalization factor,\nand in the weight assigned to the contribution of every single lattice s ite. Although in line\nof principle these differences might have experimental relevance, t heir actual detection looks\ntechnically problematic, except perhaps in the case of a double-well potential (M= 2) [19].\nInstead, as we will see in the next Section, density correlations app ear more suitable for\nstudying the predictions of the POVM prescription.\n5 Correlation functions\nBesides for density estimations, absorption images can also be used to extract information\non density correlations: by analyzing the absorption figure in distinc t points one can study\ncorrelations in atom positions. Let us then introduce the two-point correlation function as\nthe average of the following two-point operator7\nˆn(x,x′)≡ˆψ†(x)ˆψ†(x′)ˆψ(x)ˆψ(x′), (25)\nthat, according to the two measuring interpretations discussed in the previous Section, can\nbe either computed using the analog of the standard trace formula (17),\nnρ(x,x′,t) = Tr/bracketleftbig\nˆn(x,x′)ρ(t)/bracketrightbig\n, (26)\n7It differs from the density-density correlation operator by a δ(x−x′) contribution; on average, this term\nis suppressed by a factor 1 /Nand therefore negligible in the large Nlimit [24].\n8or by assuming that a single shot picture effectively projects the sy stem state into a coherent\none, so that operator averages should be taken using the transf ormed density matrix ˜ ρ(t) of\n(21), thus giving\n˜nρ(x,x′,t) = Tr/bracketleftbig\nˆn(x,x′) ˜ρ(t)/bracketrightbig\n. (27)\nIn order to make an easier comparison with the actual experimenta l data, a further in-\ntegration with respect to the barycenter coordinate, R= (x+x′)/2, is usually performed,\nso that the resulting integrated correlation function depends only on the relative coordinate\nr=x′−x[24]. Then, with a suitable normalization, one is lead to study the behav iour of\nthe following two functions\nGρ(r,t)≡/integraltext\ndR nρ(R−r\n2,R+r\n2,t)/integraltext\ndR nρ(R−r\n2,t)nρ(R+r\n2,t), (28)\nand, by replacing the standard averages with the POVM ones,\n˜Gρ(r,t)≡/integraltext\ndR˜nρ(R−r\n2,R+r\n2,t)/integraltext\ndR˜nρ(R−r\n2,t) ˜nρ(R+r\n2,t). (29)\nThese functions measure the conditional probability of finding two a toms in points separated\nby a distance r, averaged over all positions. In absence of correlations, they ta ke a constant\nvalue equal to one, while values greater than one would signal the te ndency of the atoms to\naggregate, a typical behaviour for bosons. As in the case of the a verage density discussed in\nthe previous Section, the expressions (28) and (29) take a partic ularly simple and compact\nformwhenthesystemisinitiallypreparedinaFockstate, ρ=/vextendsingle/vextendsingle/vectork;N/angbracketrightbig/angbracketleftbig/vectork;N/vextendsingle/vextendsingle. Indeed, recalling\n(10), that gives the free evolution of the Wannier functions, one e xplicitly finds\nGρ(r,t) =N(N−1)\nN2/braceleftbigg\n1 +1\nN(N−1)M/summationdisplay\ni/negationslash=j=1kikjeiQ(i−j)r/bracerightbigg\n, (30)\nwhile\n˜Gρ(r,t) =N(N−1)\nN2/braceleftbigg\n1 +1\n(N+M)(N+M−1)M/summationdisplay\ni/negationslash=j=1(ki+1)(kj+1)eiQ(i−j)r/bracerightbigg\n,(31)\nwhere, as before, kirepresents the initial occupation number of the i-th lattice site and\nQ=md//planckover2pi1t. It turns out that for a given initial state, the overall profile of (3 0) and (31)\nis very similar: both functions are essentially constant and close to o ne almost everywhere,\nexcept at the origin and at positions multiple of 2 π/Qwhere sharp peaks occur.\nNevertheless, the two expressions do differ in the normalization of t he oscillator terms, as\nwell as in their dependence on the occupation numbers. One can che ck that the effects of\nthese differences become more and more visible as the initial configur ation differs from that\ndescribed by a balanced Fock state, the one with an equal number N/Mof atoms in each\n9lattice site. Indeed, for such a state, up to terms of order 1 /N, one finds that both functions\n(30) and (31) reduce to:\nG(r,t) = 1 +sin2(πNQr)\nN2sin2(πQr)(32)\n≃1 ++∞/summationdisplay\nj=−∞δ/parenleftbiggQr\n2π−j/parenrightbigg\n. (33)\nIn order to appreciate the differences between the two theoretic al predictions Gρ(r,t) and\n˜Gρ(r,t), one has therefore to prepare the system in an unbalanced Fock state. The way\nin which this can be experimentally realized is by superimposing a second weaker periodic\npotential to the original one; this is done through the introduction of a second laser directed\nalong the optical lattice, whose wavevector κ2is chosen to be incommensurate with respect\ntoκ1=π/d, the wave vector of the first, original one [20]-[23]. When the amplitu de of the\nsecond laser is weak, thus introducing only small perturbations to t he original potential, the\ndynamics of the atoms in the lattice can still be described by the Bose -Hubbard Hamiltonian\n(1), where however thesitedepthenergy ǫigetsafurthersitedependent contributionoforder\nV2sin2(iκ2π/κ1), proportional to the strength V2of the second laser.\nThen, in order to prepare the system in a Fock state, one gradually increases the intensity\nof the lasers, so that the hopping term in the Hamiltonian (1) become s negligible. In this\nway, one drives the system into a ground state of the form/vextendsingle/vextendsingle/vectork;N/angbracketrightbig\n, characterized by the\nfilling configuration /vectork, minimizing the total energy E=/summationtext\njǫjkj+U\n2/summationtext\njkj(kj−1) with the\nconstraint/summationtext\njkj=N, forkjinteger. The advantage of working with a bichromatic lattice is\nnow apparent: it allows to change the distribution of the Natoms in the Mwells by varying\nthe amplitude V2of the second potential.\nFor a given value of the characteristic, physical parameters ente ring the Hamiltonian (1),\nthe configuration /vectorkthat minimizes the energy can be efficiently obtained using a numerical\nsimulation. We have used a Monte Carlo method implementing a simulated annealing algo-\nrithm, that better conforms to the experimentally adopted proce dure, since it considers all\natomsin the latticeat once.8Indeed, ina typical experiment, theatoms arefirst cooled inan\nharmonic potential, at the center of which the optical lattice is then slowly raised. Initially,\nthe hopping dynamics allows a redistribution of the atoms in the variou s sites; however, this\nbecomes highly suppressed at regime, leaving the system in a Fock gr ound state.\nTo be as closer as possible to an actual experimental situation, we h ave chosen to work\nwith the physical parameters that define the apparatus describe d in Ref.[21]. Although the\nprincipal lattice is in that case three-dimensional, the second, weak er potential is switched\non only along one direction, making the whole system effectively behav ing as a collection of\nindependent, separate one-dimensional bichromatic lattices. The whole system is filled with\n8This simulation procedure should be contrasted with the one adopte d in [22], where the Natoms are\ninserted in the lattice one by one, while minimizing the energy at each st ep.\n10about 3×105particles, but any single one-dimensional lattice is formed by about M= 130\nsites, filled with roughly N= 170 ultracold atoms.\nThe result of the simulation is summarized in Figure 1: it shows the plots of the corre-\nsponding correlation functions Gρ(r,t) (green line) and ˜Gρ(r,t) (red line). As expected, these\nfunctions are almost everywhere equal to one, except for the pr esence of periodic bumps.\nThe difference between the two is particularly visible in the lower, seco ndary peaks, whose\nheight is highly suppressed when the correlations are computed usin g the POVM prescrip-\ntion. This difference become more and more evident as the strength of laser giving rise to\nthe second potential is increased, as clearly shown in Figure 2. As a r esult, the suppression\nof the secondary peaks can be made visible well beyond any statistic al error, thus becoming\nexperimentally relevant.\n 0.8 1 1.2 1.4 1.6 1.8 2\n-250 -200 -150 -100 -50 0 50 100 150 200 250G(r)\nrG(r) with and without POVM\nG_POVM(r)\nG_without(r)\nFigure 1: Behavior of G(r,t) (green line) and of ˜G(r,t) (red line) with N= 170,M= 130,V2=\nh×9.9 kHz.\n6 Outlook\nThe measuring procedure commonly used in experiments with ultraco ld gases, consisting\nin extracting density profiles from absorption images taken after t he release of the optical\nlattice, suggests a theoretical interpretation in terms of genera lized quantum measurement\nprocesses. The experimental evidence regarding the presence o f interference fringes in single\n11 0.95 1 1.05 1.1 1.15 1.2 1.25 1.3 1.35 1.4 1.45\n 0 1 2 3 4 5 6 7 8 9height of the secondary peak\nV2 in units of energy of (h * 1.98 kHz)height of the secondary peak\nh_peak_POVM(r)\nh_peak_without(r)\nFigure 2: Height of the secondary peaks as a function of the intensity o f the secondary laser with\n(red line) and without POVM (green line).\nshot absorption pictures irrespective from the initial state of the system suggests the use of\na POVM based on coherent-like, fixed phase states. Within this fram ework, the averages\nof system observables in general differ from those obtained throu gh mean values of the\ncorresponding operator in the system state, the usually adopted paradigm in interpreting\nexperimental data.\nAs discussed in Section 4, these differences are hardly visible in the ca se of density mea-\nsures, since they are of the order of the inverse total number of atoms in the sample. Instead,\nthe situation appears quite different for density-density correlat ions. As explicitly discussed\nin the previous Section, by preparing the system in a suitable Fock st ate, the profile of the\nintegrated and normalized correlation function is characterized by a double series of peaks,\nthe smaller of which appear to be much lowered when the correlations are computed using\nthe POVM prescription. This effect is very pronounced and surely we ll beyond statistical\nerrors for situations close to the actual experiment: this may ope n the way to a direct test\nof the POVM assumption.\nAcknowledgements\nThis work is supported by the MIUR project “Quantum Noise in Mesos copic Systems”.\n12References\n[1] M. Lewenstein, A. Sanpera, V. Ahufinger, B. Damski, A. Sen and U. Sen,Adv. in Phys.\n56(2007) 243\n[2] I. Bloch, J. Dalibard and W. Zwerger, Rev. Mod. Phys. 80(2008) 885\n[3] L. Pitaevskii and S. Stringari, Bose-Einstein Condensation (Oxford University Press,\nOxford, 2003)\n[4] C.J. Pethick and H. Smith, Bose-Einstein Condensation in Dilute Gases , (Cambridge\nUniversity Press, Cambridge, 2004)\n[5] S. Haroche and J.-M. Raimond, Exploring the Quantum: Atoms, Cavities and Photons ,\n(Oxford University Press, Oxford, 2006)\n[6] A.J. Leggett, Quantum Liquids , (Oxford University Press, Oxford, 2006)\n[7] J. Javanainen and Sun Mi Ho, Phys. Rev. Lett. 76(1996) 161\n[8] M. Naraschewski, H. Wallis, A. Schenzle, J.I. Cirac and P. Zoller, Phys. Rev. A 54\n(1996) 2185\n[9] J.I. Cirac, C.W. Gardiner, M. Naraschewski and P. Zoller, Phys. R ev. A54(1996)\nR3714\n[10] T. Wong, M.J. Collett and D.F. Walls, Phys. Rev. A 54(1996) R3718\n[11] Y. Castin and J. Dalibard, Phys. Rev. A 55(1997) 4330\n[12] A.J. Leggett, Rev. Mod. Phys. 73(2001) 307\n[13] W.J. Mullin, R. Krotkov and F. Lalo¨ e, Am. J. Phys. 74(2006) 880\n[14] A. Imambekov, V. Gritsev andE. Demler, Fundamental noise in m atter interferometers,\ncond-mat/0703766\n[15] M. Takesaki, Theory of operator algebras , Vol. 1, (Springer-Verlag, New York, 1979)\n[16] A. Peres, Quantum Theory: Concepts and methods , (Kluwer, Dordrecht, 1993)\n[17] R. Alicki and M. Fannes, Quantum Dynamical Systems , (Oxford University Press, Ox-\nford, 2001)\n[18] M.R.Andrews, C.G.Townsend, H.-J.Miesner, D.SDurfee, D.M. Ku rnandW. Ketterle,\nScience275, 637 (1997)\n13[19] S. Anderloni, F. Benatti, R. Floreanini and A. Trombettoni, J. Phys. A 42(2009)\n035306\n[20] L. Fallani, L.E. Lye, V. Guarrera, C. Fort and M. Inguscio, Phys. Rev. Lett. 98, 130404\n(2007)\n[21] V. Guarrera, L. Fallani, J.E. Lye, C. Fortand M. Inguscio, New J. Physics 9, 107(2007)\n[22] V. Guarrera, N. Fabbri, L. Fallani, C. Fort, K.M.R. van der Stam a nd M. Inguscio,\nPhys. Rev. Lett. 100, 250403 (2008)\n[23] L. Fallani, C. Fort and M. Inguscio, Bose-Einstein condensates in disordered potentials,\narXiv:0804.2888\n[24] S. F¨ olling, F. Gerbier, A. Widera, O. Mandel, T. Gericke and I. Blo ch,Nature434, 481\n(2005)\n14" }, { "title": "0908.2743v1.Conductance_characteristics_of_current_carrying_d_wave_weak_links.pdf", "content": "arXiv:0908.2743v1 [cond-mat.supr-con] 19 Aug 2009Conductance characteristics of current-carrying d-wave weak links\nS.N. Shevchenko1\n1B. Verkin Institute for Low Temperature Physics and Enginee ring, 47 Lenin Ave., 61103, Kharkov, Ukraine\n(Dated: December 4, 2018)\nThe local quasiparticle densityofstates in thecurrent-ca rryingd-wavesuperconductingstructures\nwas studied theoretically. The density of states can be acce ssed through the conductance of the\nscanning tunnelling microscope. Two particular situation s were considered: the current state of the\nhomogeneous film and the weak link between two current-carry ingd-wave superconductors.\nPACS numbers: 74.50.+r, 74.78.-w, 74.78.Bz, 85.25.Cp\nINTRODUCTION\nUnconventional superconductors exhibit different fea-\nturesinterestingbothfromthefundamentalpointofview\nand for possible applications [1]. In particular, double\ndegenerated state can be realized in d-wave Josephson\njunctions [2]. If the misorientation angle between the\nbanks of the junction χis takenπ/4, the energy minima\nof the system appear at the order parameterphase differ-\nenceφ=±π/2. These degenerate states correspond to\nthe counter flowing currents along the junction bound-\nary. Such characteristics make d-wave Josephson junc-\ntions interesting for applications, such as qubits [3]. Our\nproposition was to make these qubits controllable with\nthe externallyinjected alongthe boundarytransportcur-\nrent [4]. It was shown that the transport current and the\nspontaneous one do not add up – more complicated in-\nterference of the condensate wave functions takes place.\nThis is related to the phenomena, known as the param-\nagnetic Meissner effect [1].\nIt was demonstrated both experimentally [5] and the-\noretically [6, 7] that at the boundary of some high- Tc\nsuperconductors placed in external magnetic field the\ncurrent flows in the direction opposite to the diamag-\nnetic Meissner supercurrent which screens the external\nmagnetic field. This countercurrent is carried by the\nsurface-induced quasiparticle states. These nonthermal\nquasiparticles appear because of the sign change of the\norder parameter along the reflected quasiparticle trajec-\ntory. Such a depairing mechanism is absent in the homo-\ngeneous situation. Note that in a homogeneous conven-\ntional superconductor at zero temperature the quasipar-\nticles appear only when the Landau criterion is violated,\natvs>∆0/pF. Herevsis the superfluid velocity which\nparameterizes the current-carrying state, ∆ 0stands for\nthe bulk order parameter, and pFis the Fermi momen-\ntum. The appearance of the countercurrent can be un-\nderstood as the response of the weak link with negative\nself-inductance to the externally injected transport su-\npercurrent. The state of the junction in the absence of\nthe transport supercurrent at zero temperature is unsta-\nble atφ=πfrom the point of view that small deviations\nδφ=±0 changethe Josephsoncurrent from 0 to its max-imal value [8]. The response of the Josephson junction\nto small transport supercurrent at φ=πproduces the\ncountercurrent [9]. It is similar to the equilibrium state\nwith the persistent current in 1D normal metal ring with\nstrong spin-orbit interaction: there is degeneracy at zero\ntemperature and φ=π, and the response of the ring is\ndifferent at δφ/negationslash= 0 orB/negationslash= 0, where Bis the effective\nmagnetic field which enters in the Hamiltonian through\nthe Zeeman term (which breaks time-reversal symmetry)\n[10]. The degeneracy is lifted by small effective magnetic\nfield so that the persistent current rapidly changes from\n0 to its maximum value. In the case of the weak link\nbetween two superconductors in the absence ofthe trans-\nport supercurrent there is degeneracy between + pyand\n−pyzero-energy states; both the time-reversal symme-\ntry breaking by the surface (interface) order-parameter\nand the Doppler shift (due to the transport supercurrent\nor magnetic field) lift the degeneracy and result in the\nsurface (interface) current [5].\nIn recent years mesoscopic superconducting structures\ncontinue to attract attention because of the possible ap-\nplication as qubits, quantum detectors etc. (e.g. [3],\n[11]). In particular, such structures can be controlled\nby the transport supercurrent and the magnetic flux\n(through the phase difference on Josephson contact).\nThis was in the focus of many recent publications, e.g.\n[4, 12, 13, 14, 15, 16]. Here we continue to study the\nmesoscopic current-carrying d-wave structures. Particu-\nlarly, we study the impact of the transport supercurrent\non the density of states in both homogeneous film and in\nthe film which contains a weak link.\nMODEL AND BASIC EQUATIONS\nWeconsideraperfectcontactbetweentwocleansinglet\nsuperconductors. The external order parameter phase\ndifference φis assumed to drop at the contact plane at\nx= 0. The homogeneoussupercurrent flows in the banks\nof the contact along the y-axis, parallel to the boundary.\nThe sample is assumed to be smaller than the London\npenetration depth so that the externally injected trans-\nport supercurrent can indeed be treated as homogeneous2\nfar from the weak link. The size of the weak link is as-\nsumed to be smaller than the coherence length. Such\na system can be quantitatively described by the Eilen-\nberger equation [8]. Taking transport supercurrent into\naccount leads to the Doppler shift of the energy variable\nbypFvs. The standard procedure of matching the solu-\ntions of the bulk Eilenberger equations at the boundary\ngives the Matsubara Green’s function /hatwideGω(0) at the con-\ntact atx= 0 [4]. Then for the component G11\nω≡g(ω,r)\nof/hatwideGω, which defines both the current density and the\ndensity of states (see below), we obtain in the left ( L)\nand right ( R) banks of the junction:\ngL,R(r) =gL,R(∞)+[g(0)−gL,R(∞)]exp/parenleftbigg\n−2|r|ΩL,R\n|vx|/parenrightbigg\n,\n(1)\ngL,R(∞) =/tildewideω\nΩL,R, (2)\ng(0) =/tildewideω(ΩL+ΩR)−isgn(vx)∆L∆Rsinφ\nΩLΩR+/tildewideω2+∆L∆Rcosφ.(3)\nHereω=πT(2n+ 1) are Matsubara frequencies, ∆ L,R\nstands for the order parameter in the left (right) bank,\nand\n/tildewideω=ω+ipFvs,ΩL,R=/radicalBig\n/tildewideω2+∆2\nL,R.(4)\nThe direction-dependent Doppler shift pFvsresults in\nthe modification of current-phase dependencies and in\ntheappearanceofthecountercurrentalongtheboundary.\nThe function g(ω,r) defines the current density, as fol-\nlowing:\nj= 4πeN0vFT/summationdisplay\nωn>0/angbracketleft/hatwidevImg/angbracketrightbv. (5)\nHereN0is the density of states at the Fermi level, /angbracketleft.../angbracketrightbv\ndenotes averaging over the directions of Fermi velocity\nvF,/hatwidev=vF/vFis the unit vector in the direction of vF.\nAnalytic continuation of g(ω),i.e.\ng(ε) =g(ω→ −iε+γ), (6)\ngives the retarded Green’s function, which defines the\ndensity of states:\nN(ε,r) = Reg(ε,r). (7)\nHereγis the relaxation rate in the excitation spectrum\nof the superconductor.\nThe local density of states can be probed with the\nmethod of the tunnelling spectroscopy by measuring the\ntunnelling conductance G=dI/dVof the contact be-\ntween our superconducting structure and the normal\nmetal scanning tunnelling microscope’s (STM) tip. Atlow temperature the dependence of the conductance on\nthe bias voltage Vis given by the following relation [17]:\nG(eV) =GN/angbracketleftD(pF)N(eV,pF)/angbracketright, (8)\nwhereGNis the conductance in the normal state; D(pF)\nis the angle-dependent superconductor-insulator-normal\nmetal barrier transmission probability. The barrier can\nbe modelled e.g.as in Ref. [7] with the uniform proba-\nbility within the acceptance cone |ϑ|< ϑc, whereϑis the\npolar angle and the small value of ϑcdescribes the thick\ntunnelling barrier:\nD(ϑ) =1\n2ϑcθ(ϑ2\nc−ϑ2), (9)\nwhereθ(...) is the theta function.\nCONDUCTANCE CHARACTERISTICS OF THE\nHOMOGENEOUS CURRENT-CARRYING FILM\nBeforestudying the current-carryingweaklink we con-\nsider the homogeneous situation. We will consider the\nd-wave film as shown in the left inset in Fig. 1. The\nmotivation behind this study is twofold: first, to demon-\nstrate the application of the theory presented above, and\nsecond, to describe recent experimental results [14].\nFIG. 1: Normalized (divided by GN) conductance dI/dVfor\nthe homogeneous current-carrying state in the d-wave film\nfor different values of the transport current. The curves are\nplottedwith γ/∆00= 0.15andϑc= 0.1π[∆00= ∆0(vs= 0)].\nLeft and right insets show the schemes for probing the densit y\nof states in the current-carrying d-wave film and in the weak\nlink (see text for details).\nThe system considered consists of the d-wave film,\nin which the current is injected along the y-axis, and\nthe STM normal metal tip (another STM contact is not\nshown in the scheme for simplicity; for details see [14]).\nFollowing the experimental work [14], we consider the\nc-axis along the x-axis and the misorientation angle be-\ntweena-axis and the direction of current ( y-axis) to be3\nπ/4. Such problem can be described with the equations\npresented in the previous section as following [7, 15].\nConsider the specular reflection at the border, when\nthe boundary between the current-carrying d-wave su-\nperconductor and the insulator can be modelled as the\ncontact between two superconductors with the order pa-\nrameters given by ∆ L= ∆(ϑ) = ∆ 0cos2(ϑ−χ) and\n∆R= ∆(−ϑ)≡∆ and with φ= 0. Then from Eq. (3)\nwe have the following:\ng(ω) =/tildewideω/parenleftbig\nΩ+Ω/parenrightbig\nΩΩ+/tildewideω2+∆∆, (10)\nwhere Ω =√\n/tildewideω2+∆2andΩ =/radicalBig\n/tildewideω2+∆2. This expres-\nsion is valid for any relative angle χbetween the a-axis\nand the normal to the boundary; in particular,\ng(ω) =/tildewideω\nΩ, χ= 0 (∆( ϑ) = ∆0cos2ϑ),(11)\ng(ω) =Ω\n/tildewideω, χ=π\n4(∆(ϑ) = ∆0sin2ϑ).(12)\nThe accurate dependence of the gap function ∆ 0=\n∆0(vs,γ) can be obtained from Ref. [4] with introducing\nγas following: pFvs−→pFvs−iγ(which is analogous\nto Eq. (6)). The energy values in this paper are made\ndimensionless with the zero-temperature gap at zero cur-\nrent: ∆ 00= ∆0(vs= 0).\nAnd now with Eqs. (12) and (6-8) we plot the STM\nconductance for the current-carrying d-wave film in Fig.\n(1). We obtain the suppression of the zero-bias conduc-\ntance peak by the transport supercurrent, as was studied\nin much detail in Ref. [14]. Our results are in agreement\nwith their Fig. 1. Also the authors of Ref. [14] devel-\noped the model based on phase fluctuations in the BTK\nformalismtoexplainthe suppressionofthe zero-biascon-\nductance peak. However, their theoretical result, Fig. 2,\ndescribes the experimental one only qualitatively, leaving\nseveral distinctions. They are the following: (i) position\nof the minima ( eV/∆00∼0.5 and 1 for the experiment\nand the theory respectively); (ii) height of the zero-bias\npeak atzerotransportcurrent( ∼2.5and 4respectively);\n(iii) height of the peak at maximal transport current\n(∼1.3 and 2.5 respectively); (iv) presence/absence of\nthe minima for all curves. Our calculations, Fig. (1),\ndemonstrate agreement with the experiment in all these\nfeatures. The agreement we obtained with two fitting\nparameters, γandϑc.\nTo further demonstrate the impact of the two fitting\nparametersofourmodel, γandϑc, inFig. (2)weplotthe\nnormalized conductance fixing one of them and chang-\ning another. The figure clearly demonstrates how they\nchange the shape of the curves: the position of the min-\nima, splitting of the zero-bias peak etc. Note that the\nsplitting is suppressed at small ϑcand high γ. This ab-\nsence of the splitting was observed in the experiment [14]\nand studied in several articles, e.g. [18].\nFIG. 2: Normalized conductance dI/dVfor the homogeneous\ncurrent-carrying state in the d-wave film for different values\nofγandϑcatpFvs/∆00= 0.5.\nCONDUCTANCE CHARACTERISTICS OF THE\nCURRENT-CARRYING WEAK LINK\nConsider now the weak link between two d-wave\ncurrent-carrying banks. For studying the effect of both\nthe transport current and the phase difference on the\ndensity of states in the contact, we propose the scheme,\npresented in the right inset in Fig. 1. The supercurrent\nis injected along y-axis in the superconducting film, as it\nwas discussed in the previous section. Besides, the weak\nlink is created by the impenetrable for electrons parti-\ntion atx= 0. The small break in this partition ( a < ξ0)\nplays the role of the weak link in the form of the pinhole\nmodel [8, 15, 19]. The STM tip in the scheme is posi-\ntioned above the weak link to probe the density of states\nin it. Two more contacts along the x-axis provide the\norder parameter phase difference φalong the weak link.\nThis can be done, for example, by connecting the con-\ntacts with the inductance, as shown in the scheme, and\napplying magnetic flux Φ eto this inductance. Then one\nobtainsthephasecontrolofthe contactwith the relation:\nφ= Φe/Φ0.\nThe two half-plains (for x <0 andx >0) play the\nrole of the two banks of the contact, which we also call\nleft and right superconductors. In our scheme the banks\ncarry the transport current along the boundary, and the\nJosephson current along the contact is created due to\nthe phase difference. The banks we consider to be d-\nwave superconductors with c-axis along the z-axis and\nwith the misorientation angles χL= 0 and χR=π/4.\nNow we can apply the equations presented in Sec. II to\ndescribe the conductance characteristics of the contact\nbetween current-carrying d-wave superconductors. This\nis done in Fig. 3, where the normalized conductance is\nplotted for two values of the phase difference, for φ=\n±π/2 and with γ/∆00= 0.1. The two values of the\nphase difference, φ=±π/2, are particularly interesting\nfor the application since they correspond to the double-4\ndegenerate states [3, 4]. So, the density of states is the\nsameinthe absenceofthetransportsupercurrentin both\npanelsin Fig. 3with mid-gapstates(at eV <∆00)which\ncreate the spontaneous current along the boundary. The\ntransport supercurrent ( vs/negationslash= 0) removes the degeneracy\nby significantly changing the mid-gap states (Fig. 3),\nwhich explains different dependencies of the current in\nthe contact on the applied transport current (i.e. on vs),\nstudied in [4, 9].\nFIG. 3: Normalized conductance dI/dVat the contact be-\ntween two current-carrying d-wave superconductors for dif-\nferent values of the transport supercurrent ( vs) and for two\nvalues of the phase difference φ=±π/2.\nCONCLUSION\nWe have studied the density of states in the current-\ncarrying d-wave structures. Namely, we have considered,\nfirst, the homogeneous situation and, second, the super-\nconducting film with the weak link. The former case was\nrelated to recent experimental work, while the latter is\nthe proposition for the new one. The local density of\nstates was assumed to be probed with the scanning tun-\nnellingmicroscope. The densityofstatesatthe weaklink\nand the current ( i.e.its components through the contact\nand along the contact plane) are controlled by the values\nofφandvs. Thesystemisinterestingbecauseofpossible\napplications: in the Josephsontransistorwith controlling\nparameters φandvsgoverned by external magnetic flux\nand the transport supercurrent [11], and in solid-state\nqubits, based on a contact of d-wave superconductors [3].\nThe author is grateful to Yu.A. Kolesnichenko and\nA.N. Omelyanchouk for helpful discussions. This work\nwas supported in part by the Fundamental Researches\nState Fund (grant number F28.2/019).\n[1] C.C. Tsuei and J.R. Kirtley, Rev. Mod. Phys. 72, 969\n(2000).[2] Yu.A. Kolesnichenko, A.N. Omelyanchouk, and A.M.\nZagoskin, Fiz. Nizk. Temp. 30, 714 (2004) (Low Temp.\nPhys.30, 535 (2004)).\n[3] A.M. Zagoskin, J. Phys.: Condens. Matter 9, L419\n(1997); L.B. Ioffe et al., Nature 398, 679 (1999); A.M.\nZagoskin, Turk. J. Phys. 27, 491 (2003).\n[4] Yu.A. Kolesnichenko, A.N. Omelyanchouk, and S.N.\nShevchenko, Fiz. Nizk.Temp. 30, 288(2004) (LowTemp.\nPhys.30, 213 (2004)).\n[5] W. Braunisch, N. Knauf, V. Kateev, S. Neuhausen,\nA. Grutz, A. Koch, B. Roden, D. Khomskii, and D.\nWohlleben, Phys. Rev. Lett. 68, 1908 (1992); H. Wal-\nter, W. Prusseit, R. Semerad, H. Kinder, W. Assmann,\nH. Huber, H. Burkhardt, D. Rainer, and J. A. Sauls,\nPhys. Rev. Lett. 80, 3598 (1998); E. Il’ichev, F. Tafuri,\nM. Grajcar, R.P.J. IJsselsteijn, J. Weber, F. Lombardi,\nand J.R. Kirtley, Phys. Rev. B 68, 014510 (2003).\n[6] S. Higashitani, J. Phys. Soc. Jpn. 66, 2556 (1997).\n[7] M. Fogelstr¨ om, D. Rainer, and J.A. Sauls, Phys. Rev.\nLett.79, 281 (1997).\n[8] I.O. Kulik and A.N. Omelyanchouk, Fiz. Nizk. Temp. 4,\n296 (1978) (Sov. J. Low Temp. Phys. 4, 142 (1978)).\n[9] Yu.A. Kolesnichenko, A.N. Omelyanchouk, and S.N.\nShevchenko, Phys. Rev. B 67, 172504 (2003); S.N.\nShevchenko, in Realizing Controllable Quantum States–\nIn the Light of Quantum Computation, Proceedings of\nthe International Symposium on Mesoscopic Supercon-\nductivity and Spintronics, Atsugi, Japan, edited by H.\nTakayanagi and J. Nitta, World Scientific, Singapore, p.\n105 (2005); arXiv:cond-mat/0404738.\n[10] T.-Z.Qian, Y.-S. Yi, and Z.-B. Su, Phys. Rev. B 55,\n4065 (1997); V.A. Cherkassky, S.N. Shevchenko, A.S.\nRozhavsky, I.D. Vagner, Fiz. Nizk. Temp. 25, 725 (1999)\n(Low Temp. Phys. 25, 541 (1999)).\n[11] F.K. Wilhelm, G. Sh¨ on, and A.D. Zaikin, Phys. Rev.\nLett.81, 1682 (1998); E.V. Bezuglyi, V.S. Shumeiko,\nand G. Wendin, Phys. Rev. B 68, 134606 (2003).\n[12] G. Rashedi and Yu.A. Kolesnichenko, Phys. Rev. B 69,\n024516 (2004).\n[13] D. Zhang, C.S. Ting, and C.-R. Hu, Phys. Rev. B 70,\n172508 (2004).\n[14] J. Ngai, P. Morales, and J.Y.T. Wei, Phys. Rev. B 72,\n054513 (2005).\n[15] S.N. Shevchenko, Phys. Rev. B 74, 172502 (2006).\n[16] V. Lukic and E.J. Nicol, Phys. Rev. B 76, 144508 (2007).\n[17] C. Duke, Tunneling in solids , Academic Press, NY\n(1969).\n[18] H. Aubin, L.H. Greene, S. Jian, and D.G. Hinks, Phys.\nRev. Lett. 89, 177001 (2002).\n[19] M. Fogelstr¨ om, S. Yip, and J. Kurkij¨ arvi, Czech. J. Ph ys.\n46, 1057 (1996).\n[20] G. Rashedi, J. Phys.: Condens. Matter 21, 075704\n(2009).\n[21] A.N. Omelyanchouk, S.N. Shevchenko, and Yu.A.\nKolesnichenko, J. Low Temp. Phys. 139, 247 (2005)." }, { "title": "0909.1362v1.Theory_of_the_Magnetic_Field_Induced_Insulator_in_Neutral_Graphene.pdf", "content": "arXiv:0909.1362v1 [cond-mat.mes-hall] 8 Sep 2009Theory oftheMagnetic-Field-Induced InsulatorinNeutral Graphene\nJ. Jung1and A. H. MacDonald1\n1Department of Physics, University of Texas at Austin, 78712 USA\n(Dated: November 14, 2018)\nRecent experiments have demonstrated that neutral graphen e sheets have an insulating ground state in the\npresence of an external magnetic field. We report on a π-band tight-binding-model Hartree-Fock calculation\nwhich examines the competition between distinct candidate insulating ground states. We conclude that for\ngraphenesheetsonsubstratesthegroundstateismostlikel yafield-inducedspin-density-wave,andthatacharge\ndensitywavestateispossible forsuspended samples. Neith erofthesedensity-wave statessupport gaplessedge\nexcitations.\nI. INTRODUCTION\nThe magnetic band energy quantization properties of\ngraphene sheets lead to quantum Hall effects1,2,3(QHEs)\nwithσxy=νe2/hat filling factors ν=4(n+1/2) =\n···,−6,−2,2,6,···for any integer value of n. The factor of\n4 in this expression accounts for a graphene sheet’s two-fol d\nvalley and spin degeneracies. When Zeeman spin-splittingo f\nLandau levels is included, quantum Hall effects are expecte d\nattheremainingevenintegervaluesof ν,includingtheneutral\ngraphene ν=0 case. The ν=0 quantum Hall effect of neu-\ntral graphenesystems is interesting from two-differentpo ints\nofview. Firstofall,thetransportphenomenologyofthequa n-\ntum Hall effect3is different at ν=0 because of the possi-\nble absence of edge states. Indeed the initial experimental\nindications3thataν=0quantumHalleffectoccursinneutral\ngraphenedidnotexhibiteithertheclearpleateauin ρxyorthe\ndeepminimumin ρxxwhicharenormallycharacteristicofthe\nQHE. Secondly, although a quantum Hall effect is expected\natν=0 even for non-interacting electrons, the large energy\ngaps identified experimentally suggest that interactions p lay\na substantial role in practice. Gaps due entirely to electro n-\nelectron interactions in ordered states are in fact common4,5\ninquantumHallsystemswhentwoormoreLandaulevelsare\ndegenerate. Partly for this reason, a number of different sc e-\nnarios have been proposed6,7,8,9,10,11,12,13,14,15,16in which the\ngapatν=0isassociatedwithdifferenttypesofbrokensym-\nmetrywithinthefourquasi-degenerateLandaulevelsneart he\nFermi level of a neutral graphene sheet. The prevailing view\nhas been that the ground state is spin-polarized, with parti al\nfilling factors νσequal to 1 and −1 for majority and minor-\nity spins respectively. This state has an interesting edge s tate\nstructureidenticaltothatofquantum-spin-Hallsystems,17and\ntransport propertiesin the quantum Hall regime that are con -\ntrolledbythepropertiesofcurrent-carryingspin-resolv edchi-\nraledge-states.8,18\nThe simplest picture of strong-field physics in nearly-\nneutral graphene sheets is obtained by using the Dirac-\nequationcontinuummodelandneglectinginteraction-indu ced\nmixing between Landau levels with different principal quan -\ntum number n. In this model, electron-electron interactions\nare valley and spin-dependent. When Zeeman interactions\nanddisorderareneglected,thebrokensymmetrygroundstat e\nconsists12of two-filled n=0 Landau levels with arbitrary\nspinors in the 4-dimensional spin-valley space. This famil yof states is favored by electron-electron interactions bec ause\nofFermistatisticswhichlowersCoulombinteractionenerg ies\nwhentheorbitalcontentofelectronsinthefermionsea ispo -\nlarized. When Zeeman coupling is included, it uniquely se-\nlects from this family the state in which both n=0 valley\norbitals are occupied for majority-spin states and empty fo r\nminority-spin states. The interacting system ground state is\nthen identical to the non-interacting system ground state, al-\nthoughinteractionsareexpected12todramaticallyincreasethe\nenergygapforchargedexcitations.\nThisargumentforthecharacterofthegroundstate appears\nto be compelling, but its conclusions are nevertheless unce r-\ntain. First of all, Landau-level mixing interactions are no r-\nmally stronger than Zeeman interactions, and could play a\nrole19. In addition, although corrections7,11,20to the contin-\nuum model for graphene are known to be small at experi-\nmentalfieldstrengths,theycouldstillbemoreimportantth an\nthe Zeeman interactions. Suspicions that the character of t he\ngroundstate couldbemisrepresentedbythe n=0continuum\nmodel theory have been heightened recently by the work of\nOngandcollaborators,whofoundasteepincreaseintheDira c\npointresistance21withmagneticfieldandevidenceforafield-\ninducedtransitiontoastronglyinsulatingstateatafinite mag-\nneticfieldstrength.22Somewhatlessdramaticincreasesinre-\nsistance at the Dirac point have also been reported by other\nresearchers.23,24,25\nIn this article we attempt to shed light on the ground\nstate of neutral graphene in a magnetic field by performing\nself-consistentHartree-Fockcalculationsfora π-orbitaltight-\nbinding model. In the continuum model, Hartree-Fock the-\nory is known12to yield the correct ground state. By us-\ning aπ-orbital tight-binding model we can at the same time\nconveniently account for Landau-level mixing effects and\nsystematically account for lattice corrections to the Dira c-\nequation continuum model. As we will discuss at length be-\nlow, it is essential to perform the mean-field-theory calcul a-\ntions with Coulombic electron-electron interactions, and not\ntheHubbard-likeinteractionscommonlyused10,26withlattice\nmodels. One disadvantage of our approach is that our cal-\nculations are feasible only at magnetic fields strengths whi ch\narestrongerthanthoseavailableexperimentally. Wethere fore\ncarefully examine the dependence on magnetic field strength\nandextrapolatetoweakerfields. Weconcludethatundertypi -\ncalexperimentalconditionsthemost-likelyfield-induced state\nof neutral graphene on a SiO 2substrate is an spin-density-2\nwave state, and that suspended samples might have a charge\ndensity wave state. Neither of these orderings support edge\nstates in the ν=0 gap. We also discuss the magnetic field\ndependence of different contributions to the total energy a nd\nestimate a critical value of perpendicular and tilted magne tic\nfield at which Zeemansplitting will bringabout a phase tran-\nsitiontoa solutionwith netspinpolarizationwhich doessup-\nportedgestates.\nAlthough our calculation captures some realistic features\nof graphene sheets that are neglected in continuum models,\nit is still not a complete all electron many-body theory. In\nparticularwe neglect the carbon σandσ∗orbitalswhose po-\nlarization is expected to screen the Coulomb interactions a t\nshort distances. Because of our uncertainty as to the streng th\nof this screening, our conclusions cannot be definitive. We\nnevertheless hope that our calculations, in combination wi th\nexperiment, will prove useful in identifying the character of\nthefield-inducedinsulatingstate in neutralgraphene.\nOurpaperisorganizedasfollows. InSection IIweexplain\nin detail the model which we study which has two parame-\nters, a relative dielectric constant εrwhich accounts for the\ndielectric environment of the graphene sheet, and on on-sit e\ninteraction Uwhich accountsfor short-distance screeningef-\nfects, for example by σ-band polarization. Our main results\non the competition between different ordered states are pre -\nsented in Section III. In Section IV we turn to a discussion\nof the electronic structure of neutral graphene ribbons, pa y-\ning particular attention to their edge states which play a ke y\nrole in most quantum Hall transport experiments. Finally in\nSection Vwe summarizeourmainconclusions.\nII. INTERACTING-ELECTRONLATTICEMODELFOR\nGRAPHENESHEETS\nA. Non-InteractingElectron π-bandmodel\nWefirstcommentbrieflyonthe π-bandtight-bindingmodel\nof graphene in the presence of a magnetic field.32,33,34,35,36\nEach carbonatom on graphene’shoneycomblattice hasthree\nnear-neighborswith π-orbitalhoppingparameter t=−2.6eV.\nMagnetic field effects are captured by a phase factor in the\nhopping amplitudes: t→t×e2πiφwhereφ= (e/ch)/integraltextAdl\ndependson line integral of the vector potential Aalong a tra-\njectory linking the two lattice sites. When the dimensionle ss\nmagneticfluxdensityis φ≡BShex/φ0=1/q,whereqisanin-\ntegerand φ0=ch/eis a magneticflux quantum,it ispossible\nto apply Bloch’s theorem in a unit cell which is enlarged by\na factor of qrelative to the honeycomblattice unit cell. (The\nhoneycomblatticeunitcellarea Shex=√\n3a2/2anda=2.46˚A\nfor a graphene sheet.) Lattice model Landau levels have a\nsmall width which increases with magnetic field strengthand\nreflects magnetic breakdown effects neglected in the contin -\nuummodel.\nThe ground state energy density differences discussed be-\nlow scale approximately as powers of the magnetic length\nℓB, defined by 2 πℓ2\nBB=φ0. (ℓBandqare related by ℓB=\n(Shexq/2π)1/2=0.371√qa=0.913√q˚A.) In a continuummodeldescriptionthedensitycontributedbyasinglefullL an-\ndau level is 1 /2πℓ2\nBand the energy of the nthLandau level is\ngiven by En=±2¯hvF/radicalbig\n|n|/ℓBwherevF=√\n3at/2¯his the\nFermi velocity of graphene. All energy levels evolve with\nmagneticfield except forthe n=0 level,E0=0.38When the\nnthLandaulevelisfullitcontributes En//parenleftbig\n2πℓ2\nB/parenrightbig\ntotheenergy\ndensity. From this we immediately see that in the weak-field\nlimitimportantenergiestendtoscaleas ℓ−3\nB∝B3/2. It iseasy\nto show, for example, that the magnetic-field dependence of\nthe total band energy of a neutral non-interacting graphene\nsheet is given by E(ℓB) =akin/ℓ3\nBwhereakin=2.65eV·˚A3.\nThis non-analytic field-dependence is responsible for the d i-\nvergent weak-field diamagnetic response ( (∂Etot/∂B)/B) of\ngraphene discussed some time ago by McClure39. We show\nbelow that when interactions are included, the energy diffe r-\nences between competing field-induced-insulator states al so\ntendto varyas ℓ−3\nB.\nB.π-bandModelEffective Interactions\nIt is clear from previous analysis of lattice-corrections t o\ncontinuum models7,11,13,20and from lattice-model calcula-\ntions based on extended Hubbard models10that conclusions\non the nature of the field-induced insulating ground state ar e\nvery dependenton the effective electron-electroninterac tions\nused in a π-band lattice model of graphene. In particular,\nit seems clear that the long-range 1 /rCoulomb interaction\ntail is essential. We approximate the interaction between π-\norbitals located at sites separated by a distance dbyV(d) =\n1/(εr/radicalbig\na2o+d2)whereao=a//parenleftbig\n2√\n3/parenrightbig\n, the bondingradiusof\nthe carbon atoms, accounts approximately for interaction r e-\nduction due to π-charge smearing on each lattice site,40and\nεraccounts for screening due to the dielectric environment\nof the graphene sheet. (Here energies are in Hartree ( e2/aB)\nunits and lengthsare in units of the Bohr radius aB.) The on-\nsite repulsiveinteractionparameter, U,isnotwell knownand\nwe take it to be a separate parameter. We motivate the range\nofvaluesconsideredforthisinteractionparameterbelow. The\nvalue chosen for εrcan also represent in part screening by σ\norbitals neglected in our approximation, or be understood a s\nanad-hoccorrection for overestimates of exchange interac-\ntions in Hartree-Fock theory. Although we study a range of\nvaluesforthisinteractionparametermodelinordertotest the\nrobustnessofourconclusions,webelievethatavalueof εr∼4\nisnormallyappropriateforgraphenesheetsplacedonadiel ec-\ntric substrate. For practical reasons we truncate the range of\nCoulomb interaction in real space at d=Lmax=6.5a. This\ntype of truncationis especially helpfulwhen treatingsyst ems\nwithout periodic boundary conditions and allows us to avoid\nproblemsduetoslowlyconvergingsumsinrealspacethatcan\notherwise be treated throughthe Ewald sum method.41Trun-\ncation of the Coulomb interaction at a reasonably large Lmax\nmust however be applied with utmost care in order to obtain\nsolutionsconsistent42,43withthelimit Lmax→∞.\nInconsideringappropriatevaluesfortheon-siteinteract ion\nUwe can start from the Coulomb interaction energy at the\ncarbon radius length scale which is ∼20eV, while this es-3\ntimate can be reduced if one considers a charge distribution\ncorrespondingtoa p-orbital. Infactanestimate fromthefirst\nionizationenergyandelectronaffinitygives U=9.6eV.27Itis\nknownthat the effectiveon-site interactionstrengthis gr eatly\nreducedfromthisbarevalueinthesolidstateenvironmentb e-\ncauseofscreeningbypolarizationofboundorbitalsonnear by\ncarbonatoms. Weconsidervaluesof Ubetween2eVand6eV,\nbracketingvaluesdeemedappropriatebyavarietyofdiffer ent\nresearchers7,28,29,30. A larger value of Uincreases the inter-\naction energy cost of any charge-density-wave (CDW) state\nwhichmightoccur. Thedirectinteractionenergyiszerowhe n\nall carbonsites stay neutral,butcan bepositiveornegativ ein\nCDW states. In the CDW states we discuss below electron\ndensityδnistransferredbetweenAandBhoneycombsublat-\ntices. Inthisstate thedirectinteractionenergypersite i s\nδEDI=(δn)2\n2/bracketleftBig\nU+∑\nj∈AV(dij)−∑\nj∈BV(dij)/bracketrightBig\n,(1)\nwheredijis the distance between lattice sites iandj,U=\nV(dii)andiis a fixed label belonging to sublattice A. The\nlargest terms in Eq.( 1) are the repulsive on-site interacti ons\nwhich are proportional in our model to Uand attractive ex-\ncitonicinteractionbetweenelectronsonneighboringoppo site\nsublattice sites which are inversely proportional to εr. Us-\ning an Ewald technique to sum over distant sites we find that\nδEDIispositivefor εr·U>13.05eV.(Thecorrespondingcri-\nterion for the truncated Coulomb interactions we use in our\nself-consitent-fieldcalculationsis εr·U>=12.23eV;thedif-\nferencebetween the right-hand-sideof these two equations is\none indicator of the inaccuracy introduced by truncating th e\nCoulombinteraction.) When εr·U<12.23eVtheCDWstate\nisstableunlessbandandexchangeenergiessupportaunifor m\ndensitystate43.\nGiven the band structure model and the interaction model,\nthe Hartree-Fock mean-field-theory calculations for bulk\ngraphene sheets with periodic boundary conditions and for\ngraphene ribbonsreported in the following sections are com -\npletelystandard.31Thebandquasiparticlesaredeterminedby\ndiagonalizing a single-particle Hamiltonian which includ es\ndirect and exchange interaction terms. The direct and ex-\nchange potentials are expressed in terms of the occupied\nquasiparticle states and must be determined self-consiste ntly.\n(We do not quote the detailed expressions for these terms\nhere.) Since the Hartree-Fock equations can be derived by\nminimizing the total energy for single Slater determinant\nwavefunctions, every solution we find corresponds to an ex-\ntremum of energy. The iteration procedure is stable only if\nthe extremum is a minimum so we can be certain that all the\nsolutions found below represent local energy minima among\nsingle Slater determinant wavefunctions with the same sym-\nmetryproperties.III. FIELD-INDUCEDINSULATINGGROUND STATES\nA. Identificationof CandidateStates\nAt zero-field band energy favors neutral graphene states\nwithout broken symmetries, and there is no compelling evi-\ndence from experiment that they are induced by interactions .\nIn a perpendicular magnetic field, however, the systems is\nparticularly susceptible to the formation of broken symme-\ntry ground states because of the presence of a half-filled set\nof four-fold spin (neglecting Zeeman) and valley degenerat e\nLandau levels with (essentially) perfectly quenched band e n-\nergy. Althoughthe final groundstate selection probablyres ts\nonconsiderationsthat it fails to capture,the n=0 continuum\nmodel captures the largest part of the interaction energy an d\nmostofthequalitativephysics. Thegroundstateisformedb y\noccupyingtwoofthefour n=0Landaulevels,selectedatran-\ndom from the four-dimensional orbital space, and producing\nagapforchargedexcitations.\nThreerepresentativebrokensymmetrystatesareillustrat ed\ninFig.1. Because n=0Landaulevelsorbitalsassociatedwith\ndifferentvalleys are completely localized on different ho ney-\ncomb sublattices, a charge density wave (CDW) solution re-\nsults when n=0 orbitals are occupied for both spins of one\nvalley. (WhenLandaulevelmixingisneglectedvalleyindic es\nand A or B sublattice indices are equivalent.) This state low -\ners the translational symmetry of the honeycomb lattice in a\nway which removes inversion symmetry. The other extreme\nis a spontaneouslyspin-polarizeduniformdensitystate (f erro\n- F) in which n=0 orbitals are occupied in both valleys but\nonlyforonespin-component. A thirdtypeof brokensymme-\ntry state, the spin-density-wave(SDW)state, hasbothbrok en\ninversionsymmetry and brokenspin-rotationalinvariance . In\nthe cartoon version of Fig. 1, n=0 electrons occupy states\nwith one spin-orientation on one sublattice and the opposit e\nspin-orientationontheothersublattice. Possible broken sym-\nmetrystates, some at other filling factors, had been discuss ed\npreviously by several authors.7,10These three states are all\ncontainedwithinthe n=0continuummodelfamilyofground\nstateswhosedegeneracyisliftedbybylatticendLandaulev el\nmixing effects. In the self-consistent mean-field-theory c al-\nculations described in detail below, the three states ident ified\naboveallappearasenergyextremainourcollinear-spinstu dy.\nB. EnergyComparisons\nInorderto examinethe physicsbehindthecompetitionbe-\ntweenthecandidategroundstateswedecomposethetotalen-\nergy for all three contributions into band, direct interact ion,\nand exchange interaction contributions. We have obtained\nself-consistentsolutionsforallthreestatesoverarange ofon-\nsite interaction Uand the dielectric screening εrvalues. Be-\ncause of kinetic-energy quenching in the (essentially dege n-\nerate)n=0 Landau level, the interaction strengths required\nto drive the system into an ordered state are essentially zer o.\nThe key question, then, is which state is favored. In Fig. ( 1)\nwe illustrate how the energy differences between the three4\nb. SDW \nc. F a. CDW Valley A Valley B a.b. c.\n0\n2 3 4 5 6 246810 \n-2E-3 -1E-3 -5E-4 03E-4 1E-3 2E-3 3E-3 \n0\n2 3 4 5 6 246810 \n-9E-4 -7E-4 -5E-4 -3E-4 -2E-4 02E-4 4E-4 6E-4 \n0\n2 3 4 5 6 246810 \n-2E-3 -2E-3 -1E-3 -7E-4 01E-4 6E-4 1E-3 \nCDW SDW F\nCDW SDW \nF/parenleftBig!!\n!(\"# ) !(\"# ) !(\"# )eV \nelec .eV \nelec .eV \nelec .\nFIG.1: (Coloronline) Upper panel: Schematic representationofvalleypolarizedchargedensi tywave (CDW),spindensitywave (SDW)and\nferromagnetic (F)spinpolarized broken symmetrysolution s thatcanbe obtained ina self-consistent meanfieldcalcula tionof graphene under\na perpendicular magnetic field at half filling. Each arrow rep resents the filling of one n=0 Landau level of a given spin and valley. Lower\npanel:From left to right ECDW−ESDW,ECDW−EFandESDW−EFtotal energy differences per electron in eVas a function of the onsite\nrepulsion Uandεrobtained from a data mesh of 9 ×10 points, calculated neglecting the Zeeman term and for a ma gnetic field of 792 Teslas\ncorresponding to 1 /100 of a flux quantum per honeycomb hexagon. For smaller value s ofUtheCDWsolutions are energetically favored\nwhereas for larger values of UtheSDWsolutions are favored in a wide range of εr. TheFsolutions are never the lowest in energy. The red\ncontour lines indicate degeneracy betweentwodifferent so lutions.\nF\nπ/L 0εr= 4U= 5eVSDW\nkπ/L 0CDWEband (eV)\nπ/L 01.2\n0.6\n0\n-0.6\n-1.2F\nπ/L 0εr= 2U= 5eVSDW\nkπ/L 0CDWEband (eV)\nπ/L 02\n1\n0\n-1\n-2\nFIG. 2: (Color online) One dimensional representation of th e dispersionless band structure of a graphene sheet under a s trong magnetic field\nB=440Trepresented inthe momentum coordinate kparallel tothe narrower directionof the unitcell. The band -gaps followthe B1/2scaling\nlawexpectedfromthecontinuummodel. Theredcolorisusedt orepresentupspinwhileblueisusedfordownspin. LeftPanel: Bandstructure\nfor CDW, SDW and F solutions obtained for U=5eVandεr=4. For these interaction parameters the SDW state has the low est energy and\nthe largest gap at the Fermi level. Right Panel: Band structures for U=5eVandεr=2. When the on site repulsion is sufficiently weak the\nenergeticallyfavoredsolution corresponds tothe CDWstat e andthis solution thenhas the largest gap.5\nstates depend on the model interaction parameters. The re-\nsultsinthisfigurewereobtainedfor q=100unitcellsperflux\nquantum, which corresponds to perpendicular field strength\nB=792 Tesla. The unit cells in which we can apply peri-\nodic boundary conditions in this case contain 100 ×2 lattice\nsites. The k-space integrations in the self-consistent Hartree-\nFockcalculationswereperformedusinga60 k-pointBrillouin\nzone sampling. The self-consistent field equations were ite r-\nateduntilthetotalenergieswereconvergedtoninesignific ant\nfigures. High accuracy is required because the three states\nare very similar in energy since the ordering occurs primar-\nily in the n=0 Landau level, involving only 1% or so of the\nelectrons for this value of q. This accuracy was sufficient to\nevaluate energy differences that typically have three sign ifi-\ncantfigures.\nThefirstpointtonoticein theseplotsofenergydifferences\nis that the two uniform charge density solutions, the F so-\nlution and the SDW solution, behave similarly. The largest\ncontrastthereforeisbetweentheCDWsolutionandtheSDW\nand F solutions. Focusing first on the CDW/SDW compari-\nson we notice that the SDW state is favored when Uis large\norεris large. The crossover occurs near εr·U∼12eV,\nvery close to the line along which δEDIchanges sign. The\nfact that the CDW/SDW phase boundaryoccursveryclose to\nthis line is expected because of kinetic energy quenching in\na magnetic field. When the non-uniform density CDW state\nis compared with the uniform-density spin-polarized F stat e\nthe phase boundarymovesvery close to a larger value of this\nproduct with εr·Uranging from ∼14 to∼18eV along the\nphase boundary. Evidently the competition between CDW\nandSDWstatesisbasedverycloselyonthedirectinteractio n\nenergy, with additional weaker elements of the competition\nenteringwhenthe F stateis considered.\nDirect comparison between the uniform density SDW and\nF solutions indicates that the latter is favored only at valu es\nofUandεrwhichareoutsidetherangeofmostlikelyvalues.\nAs discussed in more detail below, we find that the direct in-\nteraction energy in these two states is identical, and that t he\nmore negative exchange energy of the SDW state overcomes\nalargerbandenergy. Inthiscasethemaindifferencebetwee n\ntheenergiesofthetwostatesarisesfromLandaulevelmixin g\neffects. As we explain later, Landau level mixing leads to a\nlocal spin-polarization which is larger in the SDW state tha n\ninthe Fstate.\nInFig. (1)wehaveintroducedthemaintrendsin theener-\ngetic competition between CDW, SDW, and F states. How-\never, as we have explained, these calculations were under-\ntakenatfieldstrengthsthatexceedthoseavailableexperim en-\ntally. Inthefollowingsubsectionwedemonstratethatthefi eld\ndependence of the energy comparisons is extremely system-\naticsothatextrapolationsdowntophysicalfieldstrengths are\nreliable. So farwe havealso ignoredZeemancouplingwhich\nfavors F states. This coupling can be important and is also\naddressedinthe followingsubsection.−∆Ekin∆Etot∆EX\nU= 5eV, ε r= 4\nlB/aEF−EAF(meV)\n5 41\n0.4\n0.16−∆Ekin\nU= 5eV, ε r= 2\nlB/aEF−ECDW (meV)\n5 42.5\n1\n0.4∆Etot∆EX∆Ees\nU= 5eV, ε r= 2\nlB/aEF−ECDW (meV)\n5 42.5\n1\n0.4\nFIG. 3: (Color online) Total energy per site differences ΔEtot, sep-\narated into kinetic energy ΔEkin, electrostatic energy ΔEDIand ex-\nchange energy ΔEXIcontributions as a function of magnetic length\nℓBin lattice constant a=2.46˚Aunits. The total energy differences\nwere fitted to a C3/2B3/2+C2B2curve. The fitting parameters are\nlisted in Table I. Left panel: Energy differences between F and\nSDW solutions ΔEF/SDW=EF−ESDW. These results were ob-\ntained with interaction parameter values U=5eVandεr=4 for\nwhich SDW is the lowest energy configuration. The more negati ve\nvalues ofexchange energy intheSDWstatecompensates the ki netic\nenergy penalty related to the inhomogeneous accumulation o f the\nelectron wave functions at alternating lattice sites. The e lectrostatic\nenergy differences are zero thanks to the uniform electron d ensity\nfor both solutions. Right panel: Same as the previous figure but for\nΔEF/CDW=EF−ECDW. Theinteractionparameters inthiscaseare\nU=5eVandεr=2 for which CDW is the lowest energy configu-\nration. When the onsite repulsion Uis small enough that the elec-\ntrostatic energy penalty for the inhomogeneous charge dist ribution\nis small, exchange is the main contribution driving the CDW i nsta-\nbility. However, in the case illustrated here Uis so small that the\nelectrostatic part of the hamiltonian does play an importan t role in\nfavoringtheCDWstate. Theenergycontributionsfollowama gnetic\nfield decay law that deviates more from B3/2than in the previous\ncase because the on-site interaction Uplays anessential role.\nU= 5eV, ε r= 2U= 5eV, ε r= 4\nθBc(θ) ( Tesla )2000\n1000\n0\nπ/2 π/3 π/6 0200\n100\n0\nFIG. 4: Tilt angle θdependence of the critical magnetic field re-\nquiredtoinduce atransitiontotheFstate. Theblacksolidc urveand\ndashedblue curve represent the criticalmagnetic fields sta rtingfrom\nthe SDW and CDW states respectively. In the CDW curve we ob-\nserve a larger deviation from a simple cos (θ)law due to a stronger\ninfluence of lattice scale physics described by the C2coefficient in\nequation 3.6\nεr=2 εr=3 εr=4\nUCCDW\n3/2CCDW\n2CCDW\n3/2CCDW\n2CSDW\n3/2CSDW\n2CCDW\n3/2CCDW\n2CSDW\n3/2CSDW\n2\n2 NA NA 322 -1.89 — — 52.6 1.05 — —\n3 NA NA 130 0 — — 23.2 0.0314 — —\n4 1070 -13.6 54.2 -0.210 — — — — 51.6 0.0209\n5 417 -3.15 — — 111 -0.630 — — 92.8 0.839\n6 198 -1.67 — — 255 -2.10 — — 241 0\nU BCDW\nCBCDW\nCBSDW\nCBCDW\nCBSDW\nC\n2 NA 1200 — 70 —\n3 NA 300 — 10 —\n4 2600 52 — — 50\n5 1600 — 200 — 200\n6 490 — 740 — 1100\nTABLE I: In the upper table we list the values of C(SDW/CDW)\n3/2in units of 10−10eV/T3/2andC(SDW/CDW)\n2in units of 10−10eV/T2, obtained\nfrom fitstothe their energy difference withrespect toF stat esas given inequation (2) for twodifferent values of the int eractionparameters U\nandεr. Thecriticalfield Bcestimatesaredependent onthereliabilityofthesefits. Not ethatacrossoverbetweenSDWandCDWstatescanbe\ndriven by changes inthe dielectric screening environment c aptured by εr. The lower table lists Bcvalues at which the Fstates become critical\naccording toEq.(3).\nC. FieldStrengthandZeeman CouplingDependence\nWenowturnourattentiontothemagneticfielddependence\nof the solutions. For this purpose we found self-consistent\nsolutionsovera rangeofmagneticfieldsfortwo setsofinter -\naction parameters, U=5eVandεr=4 for which the SDW\nsolution has the lowest energy, and U=5eVandεr=2 for\nwhich the CDW solution has the lowest energy. The band\nstructures of the different possible solutions for these se t of\nparametersareshowninFig. (2)andthefielddependencesof\nthe three energy differences are plotted in Fig. (3). We see\nthat every contribution accurately follows a B3/2∝l−3\nBlaw\nwith small deviations that can be accounted for by allowing\na term proportional to B2. This is the same field-dependence\nlaw that we discussed earlier for the case ofa non-interacti ng\nelectron system. In the continuum model it is guaranteed in\nneutralgraphenewhenelectronsinteract via the Coulombin -\nteractions by the fact that both kinetic and interaction ene rgy\ndensities then scale as (length)−3; the magnetic field sim-\nply provides a scale for measuring density. The fact that we\nfindthisfielddependencesimplyshowsthatthecondensation\nenergies of all three ordered states are driven by continuum\nmodel physics. This is in agreement with the intuitive pic-\nture of the interactionenergyas the productof the numberof\nelectrons occupying a Landau level which is directly propor -\ntional to B, multiplied by the Coulomb interaction scale for\nelectrons in the n=0 Landau level which is proportional to\nB1/2. Thefactthatthedifferencesinenergybetweenthethree\nstatesfollowsthisrulesuggeststhatthemostimportantso urce\nof differences in energy between these states is Landau-lev el\nmixing, which should not violate the B3/2law. Small devia-\ntionsfromthislawareexpectedbecauseoflatticeeffects. The\ndeviations are stronger in CDW solutions than in the SDWsolutions because of the charge density inhomogeneityat th e\nlatticescale presentintheformer.\nWecandrawtwoimportantadditionalconclusionsfromthe\nB3/2behavior. First of all, lattice effects are not dominant\neffect at the field strengths for which we are able to perform\ncalculations,andshouldbelessimportantattheweakerfiel ds\nfor which experiments are performed because the magnetic\nlengthlBwillthenbeevenlongercomparedtothehoneycomb\nlattice constant. The difference in energy between the thre e\nstates should mainly vary as B3/2all the way down to zero\nfield, provided only that disorder is negligible. (We discus s\nthe role of disorder againin Sec. V.) Our calculationsshoul d\nthereforereliablypredicttheenergeticorderingofthest atesin\nthe experimental field range. The second conclusion we can\nmakeconcernsthe importanceofZeemancouplingwhichwe\nhave ignored to this point. First of all, Zeeman coupling wil l\nhaveanegligibleeffectontheenergiesoftheSDWandCDW\nstatessincetheyhaveavanishingspinmagneticsusceptibi lity.\nTheenergydifferencepersitebetweentheFstateandthetwo\ndensity-wavestates canbewrittenintheform\nΔE=E(SDW/CDW)−EF\n=B3/2C(SDW/CDW)\n3/2cos(θ)3/2\n+B2·/parenleftBig\nC(SDW/CDW)\n2cos(θ)2−CZcos(θ)/parenrightBig\n(2)\nwhereBis the total magnetic field strength and θis the field\ntilt angle relative to the graphene plane normal. Factors of\nBcos(θ)in this expression therefore account for the perpen-\ndicular field dependence. The second term in Eq. 2 contain\nthe contributions that scale with B2. The factor of Bcos(θ)\nwhichappearsintheZeemantermispresentbecausethespin-\npolarization of the F state is proportional to the Landau lev el\ndegeneracy. The coefficients C3/2andC2can be obtained7\nby fitting energy differences obtained from numerical solu-\ntions of the self-consistent field equations, like those plo t-\nted in Fig. 3, and depend on the interaction model parame-\nters as shown in Table I. The Zeeman coefficient in Eq. 2 is\nCZ=7.3·10−10eV/T2is independent of interaction param-\neters. From the above equation we find that the Fstate has\nlowerenergythanthespin-unpolarizedstatesfor\nB>Bc(θ)=C2\n3/2cos(θ)\n(Cz−C2cos(θ))2(3)\nThe fields required to achieve an energetic preference for th e\nspin-polarized state are smaller at larger tilt angles beca use\ntheorbitalenergyhasa stronger θdependence.\nIn table I we show the values of C3/2andC2for SDW and\nCDWconfigurationsfavoredwith respectto F fora set ofpa-\nrametersof Uandεr. We noticethatthecoefficientsdictating\nthe critical field transition to F solutions can be made rela-\ntivelysmall iftheparametersarenearthe crossoverbounda ry\nto F states. It is possible that a SDW or CDW to F transition\ncould be induced by varying magnetic field. If a transition\nwas observed, most likely by a change in transport proper-\nties asdiscussed in the nextsection,it couldprovidevalua ble\ninput on the effective interaction parameters of the π-orbital\ntight-bindingmodel.\nIV. QUANTUMHALL EDGESTATESIN GRAPHENE\nRIBBONS\nZigzag ribbon unit cell Armchair ribbon unit cell a=2.46 ˚A\nL=2a/√\n3Themagneticbandenergyquantizationprop-\nxy\nFIG. 5: Representation of the unit cell choices for armchair and\nzigzag edge terminated graphene nanoribbons.\nThe quantum Hall effect occurs when a two-dimensional\nelectron system has a chemical potential discontinuity (a g ap\nfor charged excitations) at a density which depends on mag-\nneticfield. Agapat afield-dependentdensitynecessarily50,51\nimpliesthepresenceofchiraledgestatesthatsupportaneq ui-\nlibrium circulating current. The current varies with chemi -\ncal potential at a rate defined by the field-dependence of thebulk gap density. Most quantum Hall measurements simply\nreflect the property51that separate local equilibria are estab-\nlished at opposite edges of a ribbon in systems with a bulk\nenergy gap or mobility gap. It is immediately clear therefor e\nthat theν=0 quantumHall effect is special since it is due to\nan energy gap at the neutrality point, i.e.at a density which\ndoes not depend on magnetic field. The issue of whether or\nnot theν=0 gap and associated phenomena should be re-\nferred to as an instance of the quantum Hall effect is perhaps\na delicate one. The ν=0 gap is intimately related to Landau\nquantization and in this sense is comfortably grouped with\nquantum Hall phenomena. This view supports the language\nwe use in referring to the ν=0 quantum Hall effect. On the\nother hand, since it occurs at a field-independent density, i ts\ntransportphenomenaaremorenaturallyviewedasthoseofan\nordinary insulator16which just happens to be induced by an\nexternalmagneticfield.\nAn exception occurs for the F state which does have edge\nstates8,18, and can be viewed as having ν=1 for majority\nspins and ν=−1 for minority spins. In the simplest case,\nit has two branches of edge state with opposite chirality for\nopposite spin, much like those of quantum-spin-Hall17sys-\ntems. In a Hall bar geometry most transport measurements\nare very stronglysensitive to the presence or absence of edg e\nstates. Inordertoaddressedgestate physicsdirectlyat ν=0\nwe have extended our study from the bulk graphene to the\ngraphenenanoribboncase. Tight-bindingmodelsolutionsf or\naribboninthepresenceofamagneticfieldcanbeobtainedin\nessentially the same way as for bulk graphene, with the sim-\nplification that any magnetic field strength preservesthe on e-\ndimensional ribbon wavevector kas a good quantum num-\nber when the gauge is chosen appropriately. This graphene\nribbon problem in a magnetic field was studied time ago by\nWakabayashi et. al.47andrevisitedrecentlywithinbothtight-\nbinding36,37and continuum8,48,49models in order to provide\na microscopic assessment of the relationship between Lan-\ndau levels and edge states. The general feature of the ribbon\nbandstructureinthepresenceofamagneticfieldisthatthos e\nstateslocalizedneartheedgeshavedispersivebands,wher eas\nthoseintheflat bandregionarelocatedmostlyinthe bulk. In\nthe case of zigzag edge termination, edge localized states a re\npresentevenintheabsenceofamagneticfield46. Inthequan-\ntum Hall regime these states are in the non-dispersive band\nregionlikethebulklocalizedstatesandtheydonotcontrib ute\nto edge currents, although they can interact with other edge\nlocalizedstates.\nFig. 6 explicitly illustrates how the character of the bulk\nbroken symmetry is manifested in ribbon edge state prop-\nerties. Because of practical limitations our calculations are\nrestricted to moderately narrow ribbons with widths of or-\nder 10nm. In order to properly reproduce bulk Landau level\nquantizationinthesenarrowsystemswehavetochoosemag-\nnetic field strengths strong enough to yield magnetic length s\nℓB∼25nm/(B[Tesla])1/2substantiallysmallerthantheribbon\nwidth,i.e.fieldsstrongerthantypicalexperimentalfields. On\ntheotherhandifthemagneticfieldsaretoostrong,say ℓB