[ { "title": "1910.14489v1.Gyrokinetic_investigation_of_the_damping_channels_of_Alfvén_modes_in_ASDEX_Upgrade.pdf", "content": "Gyrokinetic investigation of the damping channels of Alfv\u0013 en modes in\nASDEX Upgrade\nF. Vannini1, A. Biancalani1, A. Bottino1,\nT. Hayward-Schneider1, Ph. Lauber1, A. Mishchenko2, I. Novikau1, E. Poli1and the ASDEX\nUpgrade team1.\n1Max-Planck-Institut f ur Plasmaphysik, 85748 Garching, Germany\n2Max-Planck-Institut f ur Plasmaphysik, 17491 Greifswald, Germany\nfrancesco.vannini@ ipp.mpg.de\nAbstract\nThe linear destabilization and nonlinear saturation of energetic-particle driven Alfv\u0013 enic\ninstabilities in tokamaks strongly depend on the damping channels. In this work, the collision-\nless damping mechanisms of Alfv\u0013 enic modes are investigated within a gyrokinetic framework,\nby means of global simulations with the particle-in-cell code ORB5, and compared with the\neigenvalue code LIGKA and reduced models. In particular, the continuum damping and the\nLandau damping (of ions and electrons) are considered. The electron Landau damping is\nfound to be dominant on the ion Landau damping for experimentally relevant cases. As an\napplication, the linear and nonlinear dynamics of toroidicity induced Alfv\u0013 en eigenmodes and\nenergetic-particle driven modes in ASDEX Upgrade is investigated theoretically and compared\nwith experimental measurements.\n1arXiv:1910.14489v1 [physics.plasm-ph] 31 Oct 20191 Introduction\nIn burning plasmas relevant for magnetic fusion energy (MFE) research, an important role is\nplayed by energetic particles (EPs). With the term EPs, we refer to fusion reaction products\n(like alpha particles) or super-thermal ions or electrons resulting from plasma heating. Such\nparticles possess higher velocities compared to those typical of the background plasma. In typ-\nical tokamaks the frequency associated with the gyrocenter motion of the EPs falls inside the\nmagnetohydrodynamic (MHD) domain ( O(10\u00002\nci)), being \n cithe ion cyclotron frequency. Be-\ncause of that, through resonant wave-particle interactions, one of the three solution satisfying the\nMHD dispersion relation can be excited. Among them, the most detrimental are the shear Alfv\u0013 en\nwaves (SAWs), having a group velocity parallel to the equilibrium magnetic \feld and satisfying\nthe dispersion relation: !=kjjvA, where the Alfv\u0013 en speed has been de\fned vA=p\nB=(4\u0019\u001am0)\n(being\u001am;0the background plasma mass density and B the background magnetic \feld strength).\nThe excitation of these modes creates a transport channel for the EPs, which can lead to loss of\nEPs before their thermalization, causing a less e\u000bective heating and also possibly damaging the\nvessel of the machine. they are also believed to be responsible of large abrupt events (ALE) ob-\nserved in the Japanese tokamak (JT-60U) (see Ref.[1]). We can simply outline the whole zoology\nof SAWs basically in discrete Alfv\u0013 en eigenmodes (AEs) and energetic particle continuum modes\n(EPMs), (non-normal modes of the SAW continuum spectra, merging as discrete \ructuations at\nthe frequency that maximizes the wave-EP power exchange, above the threshold condition of the\ncontinuum damping [2]).\nIn this paper the attention will be mainly focused on toroidal Alfv\u0013 en eigenmodes (TAE, Alfv\u0013 en\neignemodes lying in the frequency gap caused by the tokamak toroidicity) and EPM. The main\ngoal of this paper will be to understand what are the damping mechanisms of the modes of\ninterest comparing (when possible) the simulations results with the prediction of MHD theory.\nThe proper domain to take into account all the nonlinear e\u000bects, like wave-wave and wave-\nparticle interaction, as well as \fnite-Larmor-radius and \fnite-orbit-width e\u000bects, is represented\nby the gyrokinetic theory. Because of that, the simulations have been principally carried with the\nglobal, nonlinear, electromagnetic, gyrokinetic, PIC code ORB5 [3, 4] whose model, if properly\nset, contains the MHD equations as a subset. Some comparisons with the linear gyrokinetic code\nLIGKA [5] have also been performed.\nThe paper is structured as follows. In Section 2 a description of the model implemented in the\ncode ORB5 is given. Section 3 and Section 4 will be dedicated to the description of two di\u000berent\ndamping mechanisms a\u000becting the SAWs: the continuum damping and the Landau damping.\nThey will be brie\ry explained analytically, starting from the MHD equations. The theory in\nuse will be compared with the results from the simulations carried with ORB5. In Section3\nthe numerical simulations will be performed in the cylinder limit using simpli\fed pro\fles. In\nthe simulations in Section 4, a small but \fnite inverse aspect ratio will be considered, using the\nequilibrium pro\fles of the International Tokamak Physics Activity (ITPA, see Ref.[6]). In Section\n5 the studies on the linear and nonlinear growth rate and frequency spectra conducted considering\nexperimental pro\fles from the NLED-AUG case [7] will be presented. Finally, summary and future\noutlook will be shown in Section 6.\n22 Model\nSince the Alfv\u0013 en waves have a frequency much smaller then the typical ion cyclotron frequency\n(\nci) and their amplitude in the core is small compared to the background quantities, a good\ndescription of their propagation and interaction with the bulk plasma can be given through the\ngyrokinetic theory. This allows to retain a kinetic description of the events under consideration,\nreducing the 6D problem to a 5D one, by averaging the fast gyromotion. In this way the numerical\ncosts are sensibly reduced.\nORB5 is a global, nonliner, gyrokinetic, electromagnetic, PIC code, which can take into\naccount collisions and sources [3, 8]. The gyrokinetic model of ORB5 [9] contains the reduced\nMHD equations as a subset [10]. In this section we give a brief description of the gyrokinetic model\nimplemented in ORB5 and brie\ry show how the implemented equations are solved. We refer for\nmore exhaustive explanations to Ref.[8], which also give a more complete description of the\nrecent updates in ORB5. Concerning the magnetic equilibrium in use, ORB5 can either upload\nan ideal-MHD equilibrium (solution of the Grad-Shafranov equation) from the CHEASE code\n[11], or consider ad-hoc equilibria constituted by circular, concentric magnetic surfaces. It deals\nwith a straight-\feld line set of coordinates. The magnetic surfaces are labeled by s=p\n = edge,\nwhich plays the role of radial coordinate. Here is the poloidal magnetic \rux function. The\nangular dependence is given by the toroidal coordinate 'and by the poloidal magnetic angle:\n\u001f=1\nq(s)Z\u0012\n0d\u00120B\u0001r'\nB\u0001r\u00120 (1)\nbeing\u00120the geometrical poloidal angle and q(s) the safety factor, de\fned as:\nq(s) =1\n2\u0019Z2\u0019\n0d\u00120B\u0001r'\nB\u0001r\u00120: (2)\nAll the quantities in the code are normalized through four reference parameters: the ion mass\n(mi), the ion charge ( qi=eZi, beingethe electric charge and Zithe atomic number), the\nvalue of the magnetic \feld strength on axis ( B0=jB(s= 0)j) and the value of the electron\ntemperature at a speci\fed reference position s0,Te(s0). All other normalized quantities are\nobtained through these: the time units are provided in the inverse of the ion-cyclotron fre-\nquency, \n ci=qiB0=(mic), the velocity units are normalized through the ions sound velocity\n(cs=p\nqsTe(s0)=mi, being the temperature measured in keV), the length units through the ion\nsound Larmor radius ( \u001as=cs=\nci) and the densities are normalized by means of their average\nin space. The Vlasov-Maxwell gyrokinetic equations are derived through variational principles\nfrom a discrete gyrokinetic Lagrangian. This choice allows to take into account all the simpli-\n\fcation needed in the model under consideration, directly in the Lagrangian and then derive\nthe gyrokinetic equation. An immediate consequence is that it is possible to consistently derive\nconserved quantities (like the energy, see Ref.[12]), that are also used in ORB5 to test the quality\nof the simulations performed. This choice also allows to derive naturally the weak gyrokinetic\nform of the \feld equations. In the Lagrangian an ordering is present, separating the e\u000bects given\nfrom the geometry of the non-uniform magnetic \feld, from those related to the \ructuations of\nthe electromagnetic perturbation. This means that (as can be derived, see Ref.[9, 13]) the small\nparameter related to the variation of the background magnetic \feld \u000fB=\u001athLB, (being\u001aththe\nthermal Larmor radius and LBthe typical variation length of the magnetic \feld) and the small\n3parameter related to the \ructuating electromagnetic \feld ( \u000f\u000e) are related through:\n\u000fB=O(\u000f2\n\u000e): (3)\nIn this way the action functional, written in \\ pz-formulation\" appears to be the following:\nA=Zt1\nt0Ldt=X\nsZ\ndtd\n\u0012qs\ncA\u0003\u0001_X+msc\nqs\u0016_\u0012\u0000H0\u0013\nfs+\n\u0000\u000f\u000eX\ns6=eZ\ndtd\nH1fs\u0000\u000f\u000eZ\ndtd\nHdk\n1fe+\n\u0000\u000f2\n\u000eX\ns6=eZ\ndtd\nH2feq;s\u0000\u000b\u000f2\n\u000eZ\ndtd\nHdk\n2feq;e\u0000\u000b\u000f2\n\u000eZ\ndtdV\f\fr?A1;jj\f\f2\n8\u0019(4)\nwhere\u000b= 0 gives the electrostatic model, while \u000b= 1 the electromagnetic one. In Eq.4 d\n =\ndV dW , beingdW =B\u0003\njjd\u0016dpz. A sum over the species \\ s\" also appears. The symplectic\nmagnetic \feld is de\fned through the symplectic magnetic potential A\u0003=A+ (c=qs)pz^bbeing ^b\nthe unitary vector parallel to the background magnetic \feld. The canonical gyrocenter momentum\nispz=msvk+\u000b\u000f\u000e(qs=c)A1k. In the action functional, some approximations have been done.\nThe quasi-neutrality allows to consider in Eq.4 only the contribution given from the magnetic\npotential, neglecting the one given from the perturbed electric \feld. Also the incompressibility\nof the parallel perturbed magnetic \feld is assumed B1;jj=o(B1;?) and only the perpendicular\ncomponent of the perturbed magnetic potential is retained: B1;jj=r\u0002(A1;jjb)\u0018rA1;jj\u0002b.\nIn Eq.4 it must be noted that while H0,H1multiply the total distribution functions fs;fe,H2\nis related only to the equilibrium distribution function. Thanks to this choice nonlinear second\norder terms do not appear in the gyrocenter dynamics and the \feld equations are linear. The\ngyrocenter hamiltonians appearing are:\nH0=p2\nz\n2ms+\u0016B H 1=qs\u001c\n\u001e1\u0000\u000bA 1;kpz\nmsc\u001d\n(5)\nH2=\u0000msc2\n2B2jr?\u001e1j2+\u000bq2\ns\n2msc2\nA1;k\u000b2\nwhere the gyroaveraging operator has been introduced hfi=1\n2\u0019R2\u0019\n0d\u0012f. The gyroaveraging is\nremoved for the electrons that are treated as drift-kinetic:\nHdk\n1=\u0000e\u0012\n\u001e1\u0000\u000bA 1;kpz\nmsc\u0013\nHdk\n2=\u000be2\n2mec2A2\n1;k (6)\nFor the distribution of the species sthe linear gyrokinetic Vlasov equation is:\ndfs\ndt=@fs\n@t+_X\u0001rfs+ _pz@fs\n@pz= 0 (7)\nwhere the gyrokinetic characteristics can be derived from Eq.4 and are:\n8\n>>>>><\n>>>>>:_X=c^b\nqsB?\nk\u0002rH+@H\n@pzB?\nB?\nk\n_pz=\u0000B?\nB?\nk\u0001rH(8)\n4The \feld equations, quasineutrality and Amp\u0012 ere, are both derived from Eq.4 via functional\nderivatives on the perturbed \feld. ORB5 splits the total distribution function in a background\ndistribution function f0and in a time dependent one \u000efand discretize this latter through numer-\nical particles (markers) used to sample the phase space. Through an operator splitting approach\nthe code solves \frst the conlisionless dynamics (using a 4th-order Runge-Kutta method) and\nthen treats the collisions with a Lanngevin approach. The quasineutrality and Amp\u0012 ere equations\nare solved using the Galerkin methods and the perturbed \felds are discretized through cubic\nB-splines \fnite elements de\fned on a grid ( Ns;N\u001f;N\u001e). Finally it is important to mention that\nfrom the numerical side, recently the mixed-representation (\\pullback\" scheme [14]) has solved\nthe so-called \\cancellation problem\" for electromagnetic simulations.\n53 Continuum damping\nIn the present section, the continuum damping will be studied. The tokamak con\fguration will be\nselected in order to have the continuum damping as main damping mechanism a\u000becting the Alfv\u0013 en\nwaves. In order to do so, it is \frst important to understand what are the equations governing\nthe Alfv\u0013 en waves. These will be obtained under the validity of the ideal magnetohydrodynamic\n(MHD) theory by treating MHD equations with a perturbative approach. The Alfv\u0013 en wave's\ndynamics can be expressed starting from the quasi-neutrality condition r\u0001\u000eJ= 0, (being\u000eJ\nthe perturbed current) that rewritten in terms of its components parallel and perpendicular to\nthe background magnetic \feld ( ^b=B=B) reads:\nr\u0001\u000eJ?+B\u0001r\u000eJk\nB= 0: (9)\nFollowing Ref.[15], in order to obtain a simpli\fed but relevant set of equations, modes with\nk?\u001dkkare considered, so that the time scale between incompressible shear Alfv\u0013 en waves and\ncompressional waves can be separated. To further simplify the problem, we consider a pressureless\nplasma (P= 0) obtaining the following vorticity equation:\nB\u0001r\u00141\nBr2\n?\u00121\nBB\u0001r\u000e\u001e\u0013\u0015\n\u0000r\u0001\u00121\nv2\nA@2\n@t2r?\u000e\u001e\u0013\n= 0: (10)\nA di\u000berential equation for the perturbed scalar potential \u000e\u001eis thus obtained. It is linked to the\nperturbed magnetic potential ( \u000eA\u0019\u000eA^b) through the condition \u000eEk= 0, derived from the\nideal Ohm's law. In this section a non-uniform plasma equilibrium with cylindrical limit, will be\nconsidered. awill denote the typical length scale perpendicular to the equilibrium magnetic \feld\nwhileR0will represent the typical length scale parallel to it. The equilibrium magnetic \feld,\nin a coordinate system ( r;\u0012;z ) will be assumed to be B= (0;B0;\u0012(r);B0;z(r)). The geometrical\nradiusrcan be used in the cylindrical limit instead of the \rux coordinate s. By assuming a shear\nAlfv\u0013 en oscillation of the scalar potential \u000e\u001e(r;\u0012;\u001e;t ) of the form:\n\u000e\u001e=X\nm;n\u000e\u001em;n(r)ei(m\u0012\u0000n z\nR0\u0000!t); (11)\nwheremis the poloidal mode number, we can now write Eq.10 in cylindrical coordinates:\n1\nr@\n@rr\"\u0012m\nq(r)\u0000n\u00132\n+R2\n0\nv2\nA@2\n@t2#\n@\n@r\u0012\u000e\u001e\nr\u0013\n=m2\nr2\"\u0012m\nq(r)\u0000n\u00132\n+R2\n0\nv2\nA@2\n@t2#\n\u000e\u001e ; (12)\nwhere the local safety-factor pro\fle has been de\fned:\nq(r) =rB 0;z\nR0B0;\u0012: (13)\nThe shear Alfv\u0013 en wave dispersion relation is then found to be:\n!2\nA=v2\nAk2\nm;n=v2\nA\nR2\n0\u0012m\nq(r)\u0000n\u00132\n(14)\nEquation 14 proves that the shear Alfv\u0013 en waves are local plasma oscillations, having a frequency\nspectrum that varies continuously throughout the plasma radial direction. Due to the hypothesis\n6k?\u001dkkthe local nature of the continuum plasma oscillation can be exploited by reducing Eq.12\nto:\n1\nr@\n@rr\"\u0012m\nq(r)\u0000n\u00132\n\u0000!2R2\n0\nv2\nA#\n@\u000e\u001e\n@r= 0; (15)\nwhich integrated in the radial domain, becomes a di\u000berential equation for the radial electric \feld\nEr:\u0012\n!2\nA+@2\n@t2\u0013\nEr= 0)Er=E0e\u0000i!A(r)t: (16)\nAssuming now a dispersion relation of the form !A(r) =!A0+!0\nA(r\u0000r0) and by Fourier\ntransforming the radial electric \feld in the radial coordinate the following relation is obtained:\n(FEr)(kr) =p\n2\u0019E 0e\u0000i(!A0\u0000!0r0)t\u000e(kr+!0\nAt)kr/\u0000!0\nAt (17)\nThe obtained linear dependence in time of the radial wave number is a proof of the phase mixing\ntogether with the fact that, being Er(r;t) =\u0000ikr(t)\u001e(r;t), the scalar potential exhibits the\ncharacteristic decay called continuum damping:\n\u000e\u001e/\f\f!0\nAt\f\f\u00001(18)\nas it was proved in Ref.[15] (see also Ref.[16, 17, 18] for the application to Geodesic Acoustic\nmodes, GAMs). To see evidence of this mechanism, the simulations presented in this section\nhave been run with simpli\fed geometries and pro\fles, without the presence of EPs. In order to\nbe in the cylinder limit the inverse aspect ratio has been chosen to be \u000f= 0:01 and \rat density\n(ne=ni= 2:22\u00011020m\u00003) and temperature ( Te=Ti= 0:01keV) pro\fles have been taken\ninto account. This leaves all the radial dependence of the dispersion relation in the safety factor\npro\fle. Moreover, in this temperature regime the Landau damping can be neglected [19]. In\nthe simulations under examination only one axysimmetric perturbation n= 0; m= 1 peaked\nat the radial position r= 0:6, has been considered, together with a linear safety-factor pro\fle\nq=q0+q1\u0001rso that:!A=vA\nR01\nq0+q1r. The other important parameters in the simulations (minor\nand major radius, value on axis of the equilibrium magnetic \feld, ion cyclotron frequency, Alfv\u0013 en\nfrequency on-axis and the ratio of the last two), are reported in Tab.1.\nTable 1: Main simulation's parameters of the reference case for the study of the continuum\ndamping.\na0[m]R0[m]B0[T]\nci[rad=s ]!A0[rad=s ]\nci=!A0\n0:1 10 3 2:87\u00011084:38\u0001105655\nIn Fig.1 on the left, the measured values for the wave numbers krare shown for a simulation\nhavingq0= 1:75 andq1= 0:5. They have been measured interpolating the mode structure with\na sinusoidal function at times where a maximum has been reached at the radial position r= 0:6.\nBy linearly inetrpolating the measured wave numbers it is possible to calculate the coe\u000ecient\nkr;1, which is found to be in reasonable agreement with the theoretical expectations Eq.17. In\nFig.1 the dynamic of the scalar potential at some radial positions is shown, together with the\npredicted decay, Eq.18.\n7Figure 1: Left: Radial wave number dependence on time. The results of ORB5 are given by dots.\nThe theoretical prediction for this simulation ( q0= 1:75 andq1= 0:5) is thatkr;1= 0:123!A0,\nwhile the measured value is kr;1= 0:107!A0. Right: Perturbation amplitude dependence on time.\nAnalytical estimation are given by the dashed lines (curves decaying in time as \u001e\u0018j!0\nAtj\u00001).\nThe scalar potential measured at di\u000berent radial positions is given by continuous lines. No EPs\nare present here.\nIn Fig.2 \fnally the obtained values of the coe\u000ecients kr;1have been plotted against di\u000berent\nvalues of the slope of the safety factor pro\fles ( q1) in use in the di\u000berent simulations and com-\npared with Eq.17. Given the reasonable agreement found between the results of the numerical\nsimulations and the theory, we can say to have veri\fed the relevance of the continuum damping\nas main damping mechanism for this speci\fc case and to have observed the presence of phase\nmixing.\nFigure 2: Dependence of kr;1on the slope of the safety factor pro\fle. Analytical estimation are\ngiven by the dashed lines and the results of ORB5 are given by dots. No EP are present here.\n84 Landau damping\nIn the present section bulk species temperatures high enough to make the continuum damping\nnegligible with respect to the Landau damping will be considered, so that the latter becomes the\nmain damping mechanisms.\nThe attention will be focused also on a particular Alfv\u0013 en eigenmode, the toroidal Alfv\u0013 en\neignemode (TAE). Its characteristic frequency lies in the gap created in the continuum spectra\nby two close poloidal modes ( m;m +1) which are coupled because of the \fnite tokamak toroidicity,\n[20]. A TAE is located at a radial position r0satisfying:q(r0) =2m+1\n2n. The theoretical derivation\nexposed in Ref.[20, 21] will be now followed. Here, a kinetic transverse part of the wave-induced\ncurrent\u000eJk\n?is added to the ideal MHD current, so that Eq.9 becomes:\nr\u0001(\u000eJMHD+\u000eJk\n?) = 0: (19)\nEquation 19 is then multiplied by \u000e\u001eand integrated in the overall plasma volume obtaining:\nZ\ndx\u000eJMHD\u0001r\u000e\u001e+Z\ndx\u000eJk\n?\u0001r?\u000e\u001e= 0 (20)\nwhere it was assumed as boundary conditionsR\ndx\u0001\u000eJ\u000e\u001e= 0. Calling !0the frequency of the\nwave solution of the ideal MHD vorticity equation, we can consider !=!0+\u000e!the solution\nof the new vorticity equation Eq.19, being \u000e!\u001c!0. Following a perturbative approach, an\nexpression for \r= Imf!gis obtained from Eq.20:\n\r=2\u0019\nc2P\nm;nR\nVd3x\u000eJk\n?m;n\u0001r\u000e\u001e\u0003\nm;n\nP\nm;nR\nVd3x1\nv2\nAh\f\f\u000e\u001e0m;n\f\f2+\u0000m\nr\u00012j\u000e\u001em;nj2i(21)\nwhere all the appearing perturbed quantities have been decomposed in Fourier components in\nthe poloidal plane. In order to obtain a simpli\fed equation for \r, some further calculations have\nbeen done and will be now described. Eq.21 is then written in cylindrical coordinates, after\nwriting the perturbed current in terms of the perturbed distribution function and this in terms of\nthe unperturbed distribution function F0. Assuming a Maxwellian distribution function F0and\nfocusing our attention on TAE (that is assuming to have a perturbation \u000e\u001estrongly peaked at\nthe radial position where we expect to have a TAE), we obtain:\n\r=X\nj\rj\rj=\u0000\fjq2\n0vA\n2q0R0\u0014\nGmj+nq0rL\u0012;j1\nn0;j@n0;j\n@r(Hmj+\u0011Jmj)\u0015\n: (22)\nBeing:\n\n\u0012;j=eBp\nmjcrL\u0012=vth;j\n\n\u0012;j\fj= 8\u0019n0;jTj\nB2\n0\u0011j=@log(Tj)\n@log(n0;j)\u0015j=vA=vth;j: (23)\n9And:\n8\n>>>>>>>>><\n>>>>>>>>>:gm;j(\u0015j) =\u0019\n2\u0015j(1 + 2\u00152\nj+ 2\u00154\nj)e\u0000\u00152\njGmj=gm;j(\u0015j) +gm;j(\u0015j=3)\nhm;j(\u0015j) =\u0019\n2(1 + 2\u00152\nj+ 2\u00154\nj)e\u0000\u00152\njHmj=hm;j(\u0015j) +1\n3hm;j(\u0015j=3)\njm;j(\u0015j) =\u0019\n2\u00123\n2+ 2\u00152\nj+\u00154\nj+ 2\u00156\nj\u0013\ne\u0000\u00152\njJmj=jm;j(\u0015j) +1\n3jm;j(\u0015j=3):(24)\nIn Eq.22,\rhas been decomposed in the species contributions (the sum over j). It is formally\nidentical to the one derived in Ref.[22]. The di\u000berence lies in the fact that in Ref.[22] the authors\nhave obtained the estimation for \rstarting from energy principles, while here everything has\nbeen done by adding a correction to the MHD quasi-neutrality equation and thus to the Alfv\u0013 en\ndynamics [23]. Since we are interested in the study of the Landau damping, we will not consider\nthe EPs contribution, which actually drives the mode unstable. Eq.22 depends on the ratio\nbetween the Alfv\u0013 en speed and the thermal velocity of the considered species. This means that\n\u0015j\u0018m1=2\njand, since the bulk ion mass is bigger than the electron mass (for Hydrogens mH\u0018\n2000me) one can understand that the ion contribution is negligible with respect to the electron\ncontributions (because of the presence of the exponential terms in the polinomia). Because of\nthat we will focus our attention on the electron Landau damping in this analytical derivation.\nIn this section an equilibrium with small, but \fnite value of inverse aspect ratio will be\nconsidered, \u000f= 0:1. The temperature pro\fles are \rat. When considered, the fast particles have a\ndensity pro\fle peaked on axis (see Fig.3 on the left). The magnetic equilibrium and pro\fles are\nthose of the ITPA-TAE international benchmark case [6] and the safety factor pro\fle is shown\nin Fig.3 on the right. The main results that will be displayed in this section, have been obtained\nconsidering heavier electrons: me=mH=200. This has been checked to be at convergence. In\nTab.2 other important details of the simulations are presented.\nTable 2: Main simulation's parameters of the ITPA-TAE case.\na0[m]R0[m]B0[T]\nci[rad=s ]!A0[rad=s ]\nci=!A0\n1 10 3 2:87\u00011081:46\u0001106196\nFigure 3: Density pro\fles and safety factor of the ITPA-TAE case ( q(r)'1:71 + 0:15r2).\n10Figure 4: Left:Initial mode structure.Right: Frequency spectra without energetic particles.\nThe chosen initial potential perturbation is peaked around r= 0:5 and is constituted by one\nsingle toroidal mode number n= 6 while the poloidal mode numbers 9 \u0014m\u001412 are considered\n(see Fig.4 on the left). A TAE is located at r= 0:5 in the gap of the continuum spectra created\nby the coupling of the poloidal modes m= 10 andm= 11, as it is shown in Fig.4 on the right.\nIn Fig.5 the dependence of the damping rate against the value of the (\rat) electron temperature\nis shown. The damping rate is found to increase with the increasing electron temperature. This\nis an evidence that the dominant damping is the electron Landau damping. The errorbar of\nthe measured points correspond to 20% of their value. This because, as it is shown in Fig.6\non the left, the damping rate value has a dependence on the chosen width of the perturbation.\nFor completeness in Fig.5 the approximated analytical electron Landau damping formula is also\nshown (dashed line). A reasonable qualitative agreement is found between the predicted decay and\nthe simulation results. Finally in Fig.6 on the right, the dependence of the measured damping\nrate of ORB5 simulations against the electron mass has been shown. For decreasing electron\nmasses, the absolute value of the damping rate is shown to decrease, consistently with theory of\nthe electron Landau damping. In summary, it has been proved that the bulk electrons provide\nthe main damping mechanism of the observed Alfv\u0013 en modes in this particular regime. Several\napproximations have been done in the analytical theory, inter alia only passing particles are\nconsidered thus neglecting the contributions of barely trapped electrons, which are thought to be\nimportant, and which are included in our numerical simulations.\nFigure 5: Landau damping dependence versus the electron temperature.\n11Figure 6: Left:Damping rate dependence on the width of the initial Gaussian beam. Right:\nDamping rate dependence on the electron mass for ORB5 simulations.\n125 NLED-AUG case\nIn the present section the results of numerical simulations involving a realistic scenario will be\npresented.\nThe shot number #31213 of ASDEX-Upgrade (AUG) has been selected within the Non-Linear\nEnergetic-particle Dynamics (NLED) Eurofusion enabling research project [7]. Here an early o\u000b-\naxis NBI (with TEP\u001893keV) occurs with an injection angle (angle between the horizontal\naxis and the beam-line) of 7 :13\u000e. The magnetic equilibrium measured at the time t= 0:84sis\nconsidered in the present simulations (see Fig.7 on the bottom left). This case is referred to as\n\\NLED-AUG case\". Further description of this case can be found in Ref.[7]. The NLED-AUG\ncase is found to be of great interest because of its rich linear and nonlinear dynamics arising from\nthe interaction of the modes with the EPs.\nTab.3 contains the details of the main parameters considered in the simulations. Tab.4 shows\nthe values of the bulk species pro\fles on axis in the absence of EPs. The bulk ions, as well as\nthe EPs (when considered), are constituted by deuterium. The EP temperature will be always\nconsidered to be radially \rat and equal to TEP= 113keV. For the EPs density pro\fles, an o\u000b-\naxis density pro\fle \ftting the experimental pro\fles is considered with Maxwellian distribution\nfunction. For comparison we also run simulations with an on-axis EPs density pro\fles. Note that\nwhen EPs are included the electron density pro\fles is changed in order to match quasi-neutrality\nne=Zi\u0001ni+ZEP\u0001nEP.\nIn Fig.7 the safety factor pro\fle is shown, together with the temperature pro\fles of the bulk\nspecies. The safety factor pro\fle has a reversed shear, with qmin(r= 0:5)'2:28. The density\npro\fles in use will be shown in the following subsection. For numerical reasons, the electron mass\nis chosen to be me=mD=500, beingmDthe deuterium mass.\nThis section is divided in two subsections. In the \frst, the results of numerical simulations\nonly involving the linear dynamics will be discussed. In particular, the dependence of \ragainst\nthe electron temperature will be shown, with and without EPs contribution. Also the results\nof a benchmark with LIGKA are presented. In the second subsection instead, the results of\nsimulations involving also the nonlinear dynamics will be presented. Finally, it is important to\nremind that, unless speci\fcally written, the bulk and energetic ions will be treated as drift-kinetic.\nThe initial perturbation considered will take into account just one toroidal mode number ( n= 1)\nand the poloidal mode number 0 \u0014m\u00147.\nTable 3: Main simulation's parameters of the NLED-AUG case.\na0[m]R0[m]B0[T]\nci[rad=s ]!A0[rad=s ]\nci=!A0\n0:482 1:666 2:202 1:0539\u00011084:98\u000110621:15\nTable 4: Pro\fle's parameters of the NLED-AUG case.\nTe(s= 0) [keV]Ti(s= 0) [keV]ne(s= 0) [m\u00003]ni(s= 0) [m\u00003]nf(s= 0) [m\u00003]\n0:709 2:48 1:672\u000110191:6018\u000110196:98\u00011017\n131.5 2.0\nR [m]0.8\n0.6\n0.4\n0.2\n0.00.20.40.60.8Z [m]\n1.82.02.22.42.62.83.03.2B [T]Figure 7: Top Left: Safety factor pro\fle. Top Right: Bulk species temperature pro\fles. Bottom\nleft: Poloidal view of the magnetic equilibrium in use.\n145.1 Linear simulations\nIn the present subsection the results of numerical simulations involving only the wave linear\ndynamics will be presented. Fig.8 and Fig.10 show the frequency spectra, mode structure and\npoloidal view of the scalar potential \u001e, obtained considering respectively o\u000b-axis and on-axis\ndensity pro\fle for the EPs and a concentration equal to 3%. The frequency spectra have been\nanalyzed in the same temporal domain, when a clearly growing mode is observed. In both Fig.8\nand Fig.10 the continuum spectra obtained with the linear gyrokinetic code LIGKA [5] is shown\n(red crosses), together with the analytical curve for the continuum spectra calculated in cylindrical\ncoordinates and including the toroidicity e\u000bects, [20] (green dotted line).\nWhen the EPs possess an o\u000b-axis pro\fle, a mode sitting at the radial position r'0:22 is\nobserved. The dominant poloidal component of the scalar potential appears to be that having\nm= 2. Due to the measured frequency lying on the branch of the continuum, this mode is\nidenti\fed as an EPM. The frequency is measured as:\nf= 129kHz (25)\nWhen \fnite Larmor radius are take into account, a slightly change in the frequency (from 129 to\n131kHz) is observed. The frequency measured in numerical simulations can be compared with\nthe experimental measurements. In Fig.9, the spectrogram obtained with Mirnov coils is shown.\nA big variety of EPs driven modes can be found. At t= 0:84s, the modes with frequencies around\n50kHz have been identi\fed as EGAMs (see Ref.[24, 25, 26]). We focus here on the Alfv\u0013 en modes\nwith frequency lying in the domain between 100 and 150 kHz. The numerical result shows that\nthese modes are indeed EPMs (see the white cross in Fig.9). Despite the approximation of the\nEPs distribution function we notice that the results of the numerical simulations appear to be in\ngood agreement with the modes observed in the spectrogram.\n15Figure 8: Numerical results for o\u000b-axis EPs pro\fle. EPs concentration of 3%, TEP= 113KeV .\nFigure 9: Experimental spectrogram obtained with Mirnov Coil compared with theoretical pre-\ndiction at one selected time. The theoretical prediction is obtained treating the fast ions as\ngyrokinetic.\n16Figure 10: Numerical results for on-axis EPs pro\fle with a concentration of 3%, TEP= 113KeV\nAn on-axis density pro\fle is also considered for the EPs. The radial dependence of the EPs\ndensity pro\fles is expressed by the formula: nEP'(1\u0000r\u000b)\f. The coe\u000ecients \u000b;\f have been\nchosen in order to have the second derivative of nEPequal to zero at the position where an Alfv\u0013 en\nmode is expected. The numerical analysis shows that a mode lying in the gap of the continuum\nspectra, created by the poloidal modes m= 2 andm= 3 is observed. It appears to be peaked at\nthe radial position r'0:738. Due to the radial localization and frequency this is identi\fed as a\nTAE.\nIn Fig.11 the dependence of \ragainst the value of the bulk species temperature is shown.\nIn the plot on the top left, the dependence of the growth rate against the electron temperature,\nkeeping the bulk ions temperature constant ( Ti(r= 0) = 3:5keV), is shown. In the plot on\nthe top right, the dependence of the growth rate against the bulk ion temperature, keeping the\nelectron temperature constant ( Te(r= 0) = 0:707keV), is shown. The EPs temperature is \rat\nTEP= 113KeV and the EPs have a concentration of 3%. In the plot on the bottom instead,\nthe dependence of the damping rate (simulations without fast particles) is shown, against the\nvalue of the electron temperature. This study of the dependence of the growth or damping rate\nagainst the electron temperature shows that the electrons are the main responsible of the damping\nof the Alfv\u0013 en modes even in this realistic scenario, which is identi\fed here as electron Landau\n17damping. Since a realistic scenario is considered here, the approximate theoretical predictions for\nthe Landau damping described in the previous sections is outside its validity regime and therefore\nis not shown.\nFigure 11: Top: respectively on the left and on the right, two scans in the electron and bulk ions\ntemperature for a growing mode are shown. EPs concentration 3%, TEP= 113KeV . Bottom:\nScan in the electron temperature for a damped mode.\nIn Fig.12 a theoretical estimation of the regions in the radial domain where the Landau damp-\ning is supposed to dominate over the continuum damping is presented for simulations without\nEPs. In order to perform this calculation, the characteristic radial structure has been measured\nin a simulation with an EPM, giving the value of kr;0. This allows us to calculate the analytical\nprediction for the half decay time of the continuum damping. The half decay time due to Landau\ndamping, on the other hands, can be measured in a simulation without EPs:\nt1=2;continuumdamping =j2\u0000kr;0j\f\f@!\n@s\f\ft1=2;Landaudamping = log 2=\r (26)\nThe radial regions where the Landau damping dominates over the continuum damping are those\nwhere the half decay time of the Landau damping is smaller than the half decay time of the\ncontinuum damping: t1=2;continuumdamping >t1=2;Landaudamping . Equivalently it can be said that\nthe Landau damping dominates over the continumm damping in those regions where:\n\r>\f\f\f\flog(2)\n2\u0000kr;0\f\f\f\f\u0001\f\f\f\f@!\n@s\f\f\f\f(27)\nIn Fig.12 on the left the continuum spectra calculated with the code LIGKA are shown. They\nare used to calculate the derivative of the frequency to obtain the estimation of the regions\n18where the Landau damping dominates over the continuum damping. To do so, the frequency has\nbeen divided in two branches, denoted as upper and lower branch. In Fig.12 on the right the\nregions where t1=2;continuumdamping >t1=2;Landaudamping are shown. Therefore the green regions\ncorrespond to the radial domain where the Landau damping is dominant over the continuum\ndamping if the frequency of the mode is sitting on the lower branch. The cyan regions instead\nrepresent the domains where the Landau damping is dominating on the continuum damping if\nthe frequency of the mode is sitting on the upper branch.\nFigure 12: Left: continuum spectra calculated with the code LIGKA. It has been\ndivided into two branches (denoted as upper and down). Right: regions where\nt1=2;continuumdamping >t1=2;Landaudamping .\nFinally in Fig.13, the results of a \frst benchmark between ORB5 and the code LIGKA are\nshown. Here a scan in the EPs temperature is depicted. The EPs have an on-axis pro\fles and\ntheir concentration is kept constant and equal to 3%. A reasonable agreement has been found\nbetween the two codes for the measured growth rate and frequency.\nFigure 13: Scan in TEP. TAE growth rate and frequency, calculated with LIGKA and ORB5 for\nan EPs concentration equal to 3 % (same density, temperature pro\fles in use).\n195.2 Nonlinear simulations\nIn this subsection results involving the nonlinear dynamics of the Alfv\u0013 en waves are presented,\nwhen both on-axis (Fig.14) and o\u000b-axis (Fig.15) density pro\fles for the EPs are considered. With\nan on-axis density pro\fles of the EPs, a mode sitting in the frequency gap is observed (TAE),\nFig.14. Its mode structure and frequency spectra are not observed to change passing from the\nlinear to the nonlinear phase, con\frming its nature of an eignemode of this system, which is only\nweakly perturbed by EPs.\nFigure 14: Mode structure and frequency spectra in the linear phase (left) and in the saturation\nphase (right). EPs have ah on-axis pro\fle.\nWhen an o\u000b-axis density pro\fle for the EPs is considered, a mode with dominant poloidal\nmode number m= 2 and peaked around r'0:22 is observed (see Fig.15). This is consistent with\nwhat was observed in the previous sections, when just the linear e\u000bects in the simulations were\ninvolved. Passing to the nonlinear phase a secondary mode with m= 2 andm= 3 is observed\nto grow around the radial position r'0:738. This second mode is identi\fed as the previously\ndescribed TAE. This happens, because in the \frst linear phase the EPs drive the EPM unstable,\nwhich appears in fact to be dominant. In the nonlinear phase, the coexistence of the EPM and\nTAE is observed, due to an earlier saturation of the EPM (see Fig.16) .\n20Figure 15: Mode structure and frequency spectra in the linear phase (left) and in the saturation\nphase (right). EPs have an o\u000b-axis pro\fles.\nFigure 16: Time evolution of the dominant poloidal modes of the scalar potential ( m= 2;3) at\nthe radial positions where the TAE and EPM are located. EPs have an o\u000b-axis pro\fles.\n216 Conclusion\nThe presence of Alfv\u0013 en modes in burning plasma can a\u000bect negatively the energy con\fnement\nand can also cause a damage in the con\fning machine. Because of their importance, the present\npaper has dealt with the main damping mechanisms a\u000becting the Alfv\u0013 en modes, trying to outline\nthem and to understand in which regime they are acting. Among the great zoology of existing\nAlfv\u0013 en modes, the attention has been focused on toroidal Alfv\u0013 en eigenmodes and energetic particle\nmodes. These studies have been mainly carried by means of numerical simulations conducted with\nthe code ORB5. The obtained results have been compared, when possible, with the presented\nanalytic theory developed in a simpli\fed geometry, and with the results of the linear code Ligka.\nIn Section 3 simulations with very small inverse aspect ratio ( \u000f= 0:01) have been considered,\nin order to lead the analysis in the cylinder limit. Simpli\fed pro\fles have been taken into account\nand very low electron temperature has been considered in order to have the continuum damping\ndominant over the Landau damping. The developed theory has been used to analyze the results\nof simulations without energetic particles. The dependence of the radial wave number of the\nmode against the time have been observed (phase mixing). Also the scalar potential has been\nfound to decay as \u000e\u001e/j!0\nAtj\u00001(continuum damping).\nIn Section 4 higher bulk ion and electron temperatures have been considered, in order to\nobserve the Landau damping to be dominant over the continuum damping. The numerical simu-\nlations have been conducted using plasma equilibrium and pro\fles from the ITPA-TAE interna-\ntional benchmark case. In order to separate the ions and electrons contributions to the damping,\na kinetic term to the ideal MHD vorticity equation has been added. Following a perturbative\napproach, a simpli\fed estimation for the Landau damping has been obtained using cylindrical\ncoordinates. The developed analytical theory has been compared with the dependence found in\nnumerical simulations of the damping rate against the bulk electron temperature. A reasonable\nagreement has been found and and this has also proved that the electron are the main responsible\nfor the damping.\nIn Section 5 a realistic plasma equilibrium taken from a shot in ASDEX Upgrade has been\nconsidered. The results of the linear numerical simulations have shown the dependence of the\ndamping rate against the bulk electron temperature describing, also in this case, the action of\nthe Landau damping. A theoretical estimation of the radial regions where the Landau damping\nis expected to be dominant over the continuum damping has been presented, for simulations\nnear marginal stability. A benchmark with the code LIGKA has shown good agreement for the\nfrequency and growth rate dependence on the EPs temperature. Finally, the nonlinear simulations\nhave shown the interaction of an EPM and a TAE in the scenario with o\u000b-axis EPs density pro\fle.\nThe future works will extend the developed theory in order to \fnd a better agreement be-\ntween the predicted estimation of the decaying mode and the numerical simulations. In Ref.[2]\nit was suggested that all the Alfv\u0013 en \ructuations can be explained within the framework of a\nsingle general \fshbone-like dispersion relation (GFLDR). This could represent the starting point\nto improve the anaytical prediction of the damping rate and it would be a very interesting ana-\nlytical and numerical task. Future and deeper benchmark with the code Ligka and the Hybrid\nMagnetohydrodynamics Gyrokinetic code HYMAGYC [27] would be of great interest in order to\nbetter understand the linear and nonlinear dynamics contained in the NLED-AUG case.\n227 Acknowledgments\nSimulations presented in this work were performed on the CINECA Marconi supercomputer\nwithin the ORBFAST and OrbZONE projects.\nOne of the authors, F. Vannini, would like to thank Xin Wang, Zhixin Lu and Gregorio Vlad\nfor useful, interesting discussions and for great help provided in understanding Alfv\u0013 en dynamics,\nGyrokinetic and MHD theory. The authors wish to acknowledge stimulating discussions with\nF. Zonca, G. Fogaccia, A. K onies, J. Gonzalez-Martin and A. Di Siena. This work was partly\nperformed in the frame of the \\Multi-scale Energetic particle Transport in fusion devices\" ER\nproject.\nThis work has been carried out within the framework of the EUROfusion Consortium and has\nreceived funding from the Euratom research and training program 2014-2018 and 2019-2020 under\ngrant agreement number 633053. The views and opinions expressed herein do not necessarily\nre\rect those of the European Commission.\n23References\n[1] A. Bierwage, Kouji Shinhoara, Yasushi Todo, Nobuyuki Aiba, Masao Ishikawa, Go Mat-\nsunaga, Manabu Takechi and Masatoshi Yagi, \\Simulation tackle abrupt massive migrations\nof energetic beam ions in tokamak plasma\", Nature Communications 9, (2018)\n[2] Liu Chen and Fulvio Zonca, \\Physics of Alfv\u0013 en waves and energetic particles in burning\nplasmas\", Rev. Mod. Phys. 88, (2016)\n[3] S. Jolliet, A. Bottino, P. Angelino, R. Hatzky, T.M. Tran, B. McMillan, O. Sauter, K. Appert,\nY. Idomura and L. Villard, \\A global collisionless PIC code in magnetic coordinates\", Comput.\nPhys. Comm. 177, (2007)\n[4] A. Bottino, T. Vernay, B. Scott, S. Brunner, R. Hatzky, S. Jolliet, B.F. McMillan, T.M.\nTran and L. Villard, \\Global simulations of tokamak microturbulence: \fnite- \fe\u000bects and\ncollisions\", Plasma Phys. Control. Fusion 53, (2011)\n[5] P. Lauber, S. G unter, A. K onies and S. D. Pinches, \\LIGKA: A linear gyrokinetic code for the\ndescription of background kinetic and fast particle e\u000bects on the MHD stability in tokamaks\"\nJournal of Computational Physics 226, (2007)\n[6] A. K onies, S. Briguglio, N. Gorelenkov, T. Feh\u0013 er, M. Isaev, Ph. Lauber, A. Mishchenko, D.A.\nSpong, Y. Todo, W. A. Cooper, R. Hatzky, R. Kleiber, M. Borchardt, G. Vlad, A. Biancalani,\nA. Bottino and ITPA EP TG, \\Benchmark of gyrokinetic, kinetic MHD and gyro\ruid codes\nfor the linear calculation of fast particle driven TAE dynamics\" Nucl. Fusion 58, (2018)\n[7] Ph. Lauber,http: //www2.ipp.mpg.de/ ~pwl/NLED_AUG/data.html\n[8] E. Lanti, N. Ohana, N. Tronko, T. Hayward-Schneider, A. Bottino, B.F. McMillan, A.\nMishchenko, A. Scheinberg, A. Biancalani, P. Angelino, S. Brunner, J. Dominski, P. Donnel,\nC. Gheller, R. Hatzky, A. Jocksch, S. Jolliet, Z. Lu, J Collar and L. Villard, \\ORB5: a global\nelectromagnetic gyrokinetic code using the PIC approach in toroidal geometry\", Computer\nPhysics Communications , (2019)\n[9] N. Tronko, A. Bottino and E. Sonnendr ucker, \\Second order gyrokinetic theory for particle-\nin-cell codes\", Phys. Plasmas 23, (2016)\n[10] N. Myato, B. Scott and Masatoshi Yagi, \\On the gyrokinetic model in long wavelength\nregime\", Plasma Physics and Controlled Fusion 55, (2013)\n[11] H. L utjens, A. Bondeson and O. Sauter, \\The CHEASE code for toroidal MHD equilib-\nria\",Comput. Phys. Commun. 97, (1996)\n[12] A. Bottino and E. Sonnendr ucker, \\Monte Carlo particle-in-cell methods for the simulation\nof Vlasov-Maxwell gyrokinetic equations\", J. Plasma Phys. 81, (2015)\n[13] N. Tronko, A. Bottino, C. Chandre and E. Sonnendr ucker, \\Hierarchy of second order gy-\nrokinetic Hamiltonian models for particle-in-cell codes\", Plasma Phys. Control. Fusion 59,\n(2017)\n24[14] A. Mishchenko, A. Bottino, A. Biancalani, R. Hatzky, T. Hayward-Schneider, N. Ohana,\nE. Lanti, S. Brunner, L. Villard, M. Borchardt et al., \\Pullback scheme implementation in\nORB5\", Comput. Phys. Commun. 238, (2019)\n[15] G. Vlad, F. Zonca and S. Briguglio, \\Dynamics of Alfv\u0013 en waves in tokamaks\", La Rivista\ndel Nuovo Cimento 22, (2008)\n[16] Fulvio Zonca and Liu Chen, \\Radial structures and nonlinear excitation of geodesic acoustic\nmodes\", Europhysics Letters 83, (2008)\n[17] F. Palermo, A. Biancalani, C. Angioni, F. Zonca and A. Bottino, \\Combined action of phase-\nmixing and Landau damping causing strong decay of geodesic acoustic modes\", Europhys.\nLett. 115, (2016)\n[18] A. Biancalani, F. Palermo, C. Angioni, A. Bottino and F. Zonca, \\Decay of geodesic acoustic\nmodes due to the combined action of phase mixing and Landau damping\", Phys. Plasmas 23,\n(2016)\n[19] Liu Chen and Akira Hasegawa, \\Plasma heating by spatial resonance of Alfv\u0013 en wave\",\nPhysics of Fluids 17, (1974)\n[20] G. Y. Fu and J. W. Van Dam, \\Excitation of the toroidicity-induced shear Alfv\u0013 en eigenmode\nby fusion alpha particles in an ignited tokamak\", Physics of Fluids B: Plasma Physics 10,\n(1989)\n[21] Ya. I. Kolesnichenko, V. V. Lutsenko, H. Wobig and V. Yakovenko, \\Alfv\u0013 en instabilities\ndriven by circulating ions in optimized stellarators and their possible consequences in a Helias\nreactor\" Physics of Plasmas 9, (2002)\n[22] R. Betti and J. P. Freidberg, \\Stability of Alfv\u0013 en gap modes in burinig plasmas\" Physics of\nFluids B: Plasma Physics 4, (1992)\n[23] Ya. I. Kolesnichenko, A. K onies, V. V. Lutsenko, M. Drevlak, Yu. Turkin and P. Helander,\n\\Isomon instabilities driven by energetic ions in Wendelstein 7-X\" Nucl. Fusion 56, (2016)\n[24] A. Di Siena, A. Biancalani, T. G orler, H. Doerk, I. Novikau, P. Lauber, A. Bottino and E.\nPoli, \\E\u000bect of elongation on energetic particle-induced geodesic acoustic mode\" Nucl. Fusion\n58, (2018)\n[25] I. Novikau, A. Biancalani, A. Bottino, A. Di Siena, P. Lauber, E. Poli, E. Lanti, L. Villard,\nN. Ohana and S. Briguglio, \\Implementation of energy transfer technique in ORB5 to study\ncollisionless wave-particle interactions in phase-space\" Comp. Phys. Comm. , (2019) submitted\n[26] P. Lauber, \\O\u000b-axis NBI heated discharges at ASDEX Upgrade: EGAMs, RSAEs, TAE\nbursts\" 13th Energetic Particle Physics TG Meeting, Padua, Italy, 21{23 October , (2014)\n[27] G. Fogaccia, G. Vlad and S. Briguglio, \\Linear benchmarks between the hybrid codes HY-\nMAGYC and HMGC to study energetic particle driven Alfv\u0013 enic modes\" Nucl. Fusion 56,\n(2016)\n25" }, { "title": "0710.4664v2.Confronting_the_trans_Planckian_question_of_inflationary_cosmology_with_dissipative_effects.pdf", "content": "arXiv:0710.4664v2 [hep-th] 20 Jun 2008Confronting the trans-Planckian question\nof inflationary cosmology\nwith dissipative effects\nRenaud Parentani\nLaboratoire de Physique Th´ eorique, UMR 8627 ,\nBˆ at. 210, Universit´ e Paris-Sud 11,\n91405 Orsay Cedex, France\nAbstract\nWe provideaclass ofQFTswhichexhibit dissipationaboveat hresholdenergy, thereby\nbreaking Lorentz invariance. Unitarity is preserved by cou pling the fields to additional\ndegrees of freedom (heavy fields) which introduce the rest fr ame. Using the Equivalence\nPrinciple, we define these theories in arbitrary curved spac etime. We then confront the\ntrans-Planckian question of inflationary cosmology. When d issipation increases with the\nenergy, the quantum field describing adiabatic perturbatio ns is completely damped at the\nonset of inflation. However it still exists as a composite ope rator made with the additional\nfields. And when these are in their ground state, the standard power spectrum obtains\nif the threshold energy is much larger that the Hubble parame ter. In fact, as the energy\nredshifts below the threshold, the composite operator beha ves as if it were a free field\nendowed with standard vacuum fluctuations. The relationshi p between our models and\nthe Brane World scenarios studied by Libanov and Rubakov dis playing similar effects is\ndiscussed. The signatures of dissipation will be studied in a forthcoming paper.\n11 Introduction and presentation of the settings\nEven though relativistic QFT provides an excellent description of par ticle physics, being non-\ncompact, Lorentz symmetry can only be tested up to a certain ene rgy scale. Thus one cannot\nexclude that some unknown high energy processes break the invar iance under boosts, thereby\nintroducing a threshold energy Λ LV, and a preferred frame. It is therefore of interest to deter-\nmine what would be the signatures when this possibility is realized [1].\nThis question is particularly relevant for inflationary cosmology beca use primordial density\nfluctuations arise from vacuum fluctuations which had very short w ave lengths (very large\nproper frequencies) at the onset of inflation[2]. Indeed, the initial frequencies of the modes we\nobserve today in the CMB anisotropies were all larger than\nΩin=HeNextra, (1)\nwhereHis the value of the Hubble parameter during inflation, and where Nextrais the number\nof e-folds from the onset of inflation till t0, the moment when the (comoving) scale of our\nvisible universe exited the Hubble radius. (The total number of e-fo lds is thus N0+Nextra\nwhereN0∼60 is the number of e-folds from t0till the end of inflation.) Irrespectively of the\ninflationary scenario, Ω inis thus larger than the Planck mass MPlwhenNextra>lnMPl/H.\nSinceHshould be of the order of 10−5MPl, the initial frequencies of observable modes were all\ntrans-Planckian when Nextra>5ln10∼12. Moreover, since in most scenarios Nextra≫12,\nΩinwas generically much larger than the Planck mass.\nIn the absence of Quantum Gravity, there is no understanding of t he nature of thedegrees of\nfreedom at these scales: we neither know what is their Hilbert space , nor how they propagate,\nand this, because the notions of differential geometry used when d escribing (quantum) fields\nin expanding universes might not apply at very short distances. In p articular eq. (1) which is\nbased on a relativistic mode equation in an accelerating universe might loose its validity above\na certain Ultra-Violet scale Λ LV. Since this scale might significantly differ from the Planck\nmass, we shall keep it distinct from the Planck mass. We shall also use the abbreviations LI\nandLVfor Lorentz Invariance and Lorentz Violation respectively.\nIn this context, it is instructive to phenomenologically parameterize the deviations from LI\nand determine their signatures on inflationary spectra. This line of t hought has been proposed\nin the context of Hawking radiation from black holes, wherein the asy mptotic quanta also arise\nfromvacuum fluctuations withexponentially growing frequencies [3 , 4]. The deviations from LI\nhave been characterized by non-linear dispersion relations imposed in a particular rest frame,\nΩ2=Fn(p2) =p2(1±p2n/Λ2n\nLV+O(p2n+2)), (2)\nwhere Ω and pare respectively the frequency and the norm of the spatial momen tum measured\nin that frame. The first deviation is characterized by a power of p/ΛLVand the sign defining\nsuper-luminous (+) and sub-luminous ( −) cases. When Λ LVis much larger than Hawking\ntemperature, and when the frame is freely falling, it was shown that the asymptotic properties\nof Hawking radiation are unmodified, even though the near horizon p ropagation is radically\nmodified when Ω is larger than Λ LV. The robustness of asymptotic properties relies on the\n2adiabatic character of the evolution of the vacuum (ground) stat e [5]. The same program was\nthen applied to the inflationary spectra [6, 7], and similar results were obtained because, in this\ncase as well, the ground state adiabatically evolves when σ=H/ΛLV≪1 [8]. However, so far,\nonly dispersive models have been systematically studied.\nThe aim of the present work is to extend this analysis to dissipative mo dels. To this end,\nwe shall first construct QFT displaying dissipative effects in the UV. I ndeed, unlike dispersion,\ndissipation requires enlarged dynamical settings because if one trie s to introduce dissipation\nfrom the outset in eq. (2), one looses both unitarity and predictab ility. To preserve them, we\nshall work with Hamiltoniantheories in which dissipative effects arecau sed by interactions with\nadditional fields. Doing so, we shall establish that dissipative effects aregeneric. That is, when\nstarting with a Lagrangian in which LIis broken in the UV, the effective theory unavoidably\ndevelops dissipation above a certain energy scale, simply because no thing can prevent this.\n(With relativistic QFT instead, LIdid prevent it).\nSince we want to construct generalized QFT, we should decide what t o keep. First, we want\nto preserve the unitarity of the description because the calculatio n of the power spectrum of\n(adiabatic) density fluctuations [2] rests on the identification of a s calar field, hereafter called\nφ, which obeys Equal Time Commutators (ETC). This identification is ne cessary to fix the rms\namplitude of density fluctuations in the vacuum. Explicitely the power spectrum is given\nPp(t)≡4πp3/integraldisplay/parenleftBigdx\n2π/parenrightBig3\neipxGa(t,x;t,0), (3)\nwhereGais the anti-commutator of φevaluated in the asymptotic (Bunch-Davies) vacuum\nmuch after horizon exit. In the absence of dissipation, Ppcan be deduced from the norm of\nφin\np(t), the Fourier mode of φwith asymptotic positive frequency. However in the presence\nof dissipation, this notion of (free) mode disappears. Hence the kn owledge of Gabecomes\nnecessary since eq. (3) gives the only way to obtain the power spec trum.\nTo get dissipation we shall thus introduce additional degrees of fre edom, hereafter called Ψ.\nThen the whole system ( φ+ Ψ) will evolve unitarily, by construction. This guarantees that the\nETC ofφis exactly preserved (in a non trivial way since φundergoes dissipation). Moreover,\nthe (dressed) 2pt function of φwill be given by the usual QM trace\nGW(x,y) = AR/bracketleftBig\nˆρTˆφ(x)ˆφ(y)/bracketrightbig\n, (4)\nwhere ˆρTis the initial matrix density of the total system, where ˆφ(x) is the Heisenberg field\noperator evolved with the time ordered exponential of the total H amiltonian, and where the\ntrace is taken over both Ψ and φ. The anti-commutator Gadetermining the power spectrum\nin eq. (3) is simply the symmetric part of GW.\nOur second requirement concerns the properties of dissipative eff ects. When considering\nthe theory in vacuum and in Minkowski space-time, we impose that th ese effects preserve\nthe stationary, the homogeneity, and isotropy of flat space-time . These requirements define\na preferred frame which is inertial and globally defined. Then irrespe ctively of the choice of\nΨ and Ψ- φinteractions, the Fourier transform of the retarded Green func tion,Gr(x,y) =\n3θ(tx−ty)2ImGW(x,y), is of the form\nGr(ω,/vector p) =1/parenleftBig\n−ω2+p2+Σr(ω,p)/parenrightBig. (5)\nIn the true vacuum, at the level of the 2 pt functions, the dissipat ive (dispersive) effects are\nthus completely characterized by the imaginary and odd part in ω(real and even part) of the\nself energy Σ r(ω,p).\nIn these expressions, the energy ωand the momentum square p2are defined in the preferred\nframe. The novelty is that Σ ris a function of ωandpseparately, and not only a function of\nthe relativistic invariant ω2−p2as it is the case in LIQFT. Therefore dissipation can become\nsignificant above a critical energy on the mass shell , i.e. along the minima of the denominator\nof eq. (5); a possibility forbidden in LItheories. For instance, one verifies that the following\nself-energies induce significant dissipation only above Λ LV\nImΣ(n)\nr(ω,p) =−ω\nΛLVp2/parenleftbiggp2\nΛ2\nLV/parenrightbiggn\n. (6)\nWiththisequation we have identified therelevant (lowest order) qua ntity governing dissipation,\nthe equivalent of the first order deviation in eq. (2). In the body of the paper, we shall provide\nLagrangians of Ψ and φwhich give this class of self-energies labeled by n. Rather than focusing\non a particular case, we shall describe the whole class of dissipative b ehaviors (at the level\nof 2-point functions). We shall thus follow a phenomenological appr oach based on minimal\nassumptions (i.e. unitarity), rather than some ”inspired” approac h (e.g. by string theory, or\nbranes scenarios [9, 10]) which would lead to a particular sub-class of models.\nOur thirdrequirement concerns theextension ofourmodels fromM inkowski space tocurved\nspace-times. TodefineourQFTincurvedspace, wesimplyimplementt heEquivalencePrinciple\n(EP). It fixes the action density our models to be the covariantized version of that we had in\nMinkowski space (up to some non-minimal coupling). To perform the covariantization, it is\nuseful to characterize the preferred frame in a coordinate invar iant way by lµ, a unit time-like\nvector field [11]. In terms of this vector, ωandp2are given by\nω≡lµpµ, p2≡⊥µνpµpν, (7)\nwhere⊥µν≡gµν+lµlνis the (positive definite) metric in the spatial sections orthogonal t olµ.\nThe covariant action will be a sum of scalar functions of the four loca l fieldsφ, Ψ,gµνandlν\nwhich reduces to the Minkowski model in the zero curvature limit. Ve ry importantly for us, we\nshall see that the EP guarantees the adiabaticity of the evolution o f the (interacting) ground\nstate as long the gradients of the metric are much smaller than the U V scale Λ LV.\nFor the interested reader, we mention that additional comments w hich place the present\nwork in broader contexts can be found in the introduction of [12]. T hese comments concern\nQuantum Gravity and the description of ”mode creation” in expandin g universes.\n42 Dissipation in Minkowski space from LVeffects\nIn this Section, we provide a class of models defined in Minkowski spac etime which exhibit\ndissipative effects above a certain energy scale Λ LV. Stationarity, homogeneity and isotropy\nwill be exactly preserved. Therefore, the only invariance of relativ ist QFT which is broken is\nthat under boosts. These theories define a preferred rest fram e which is globally defined, as it\nis the case in FLRW space-times. For the simplicity of the presentatio n, we shall first work in\nthe preferred frame. At the end of this Section we shall covariant ize our action and generalize\nit to curved spacetime.\n2.1 Free field settings\nWe start with a brief presentation of the free field quantization to in troduce notations and to\npoint out what are the properties which will be lost in the presence of dissipation.\nThe action of our free massless field φis the usual one:\nSφ=1\n2/integraldisplay\ndtd3x(∂tφ2−∂xφ·∂xφ), (8)\nwheretandxare Cartesian coordinates. Due to the homogeneity of space, the equation of\nmotion can be analyzed mode by mode:\nφ(t,x) =/integraldisplayd3p\n(2π)3/2eip·xφp(t). (9)\nThe Fourier mode φp(t) obeys\n(∂2\nt+ω2\np)φp= 0, (10)\nwhereω2\np=p2=p·pis the standard relativistic dispersion relation. Notice that eq. (10) is\nsecond order, homogeneous (no source term), and time reversib le (no odd power of ∂t), three\nproperties we shall loose when introducing interactions breaking LI.\nIn homogeneous space-times, the Equal Time Commutator betwee n the Heisenberg field\noperator and its momentum implies that the mode operator (q-numb er)φp(t) obeys\n[φp(t),∂tφ†\np′(t)] =iδ3(p−p′). (11)\nWhen decomposing this operator as\nφp(t) =apφp(t)+a†\n−pφ∗\np(t), (12)\nwhere the destruction and creation operators satisfy the usual commutators\n[ap,a†\np′] =δ3(p−p′),[ap,ap′] = 0, (13)\neq. (11) is verified because the Wronskian of the positive frequenc y (c-number) mode\nφp(t) =e−iωpt/(2ωp)1/2, (14)\n5is constant (and conventionally taken to be unity).\nHad an odd term like γ∂tbe present in eq. (10) the constancy of the Wronskian would have\nbeen lost. Hence the possibility of realizing the ETC (11) with the help o f eq. (13) would have\nbeen lost as well. This already indicates that, unlike dispersive (real) e ffects, dissipative effects\nrequire more general settings than the above.\n2.2 Interacting models breaking LI, general properties\nWe now introduce additional degrees of freedom, here after collec tively named Ψ, which induce\ndissipationabovetheenergyΛ LV. Weshallworkwithaparticularclassofmodelsinordertoget\nanexact (non-perturbative) expression for the two-point func tion of eq. (4). Before introducing\nthese models, we derive general results valid for all unitary QFT’s po ssessing dissipative effects\nabove Λ LVin the ground state (the interacting vacuum).\nWe assume that the total action decomposes as\nST=Sφ+SΨ+Sφ,Ψ, (15)\nwhere the first action is that of eq. (8), the second one governs t he evolution of the Ψfields,\nand the last one the coupling between φand these new fields. We also impose that the last two\nactions preserve the homogeneity and isotropy of Minkowski spac e but break the invariance\nunder boosts. From now on, the coordinates t,xare at rest with respect to the preferred frame\ndefined by SΨ+Sφ,Ψ, i.e.,∂t≡lµ∂µin the covariant notation of eq. (7).\nWhen the state of such system is homogeneous, the Fourier trans form of the 2pt function\nof eq. (4) is of the form\nGp,p′(t,t′) = Tr[ˆ ρTˆφp(t)ˆφ†\np′(t′)],\n=GW(t,t′;p)δ3(p−p′). (16)\nAt this point, an important remark should be made. In the presence of interactions, the notion\n(and the usefulness) of the time-dependent modes of eq. (14) ha s disappeared whereas the\nfunction GW(t,t′;p) of eq. (16) is always well-defined, for all choices of ΨandSΨ+Sφ,Ψ.\nWhen the situation is stationary, GW(p;t,t′) further simplifies in the frequency representa-\ntion:\nGW(t,t′;p) =/integraldisplaydω\n2πe−iω(t−t′)GW(ω,p). (17)\nWhen working in the ground state, one finds the simplest case, beca use whatever the Ψfields\nmay be, the Fourier transform of the time ordered (Feynman) pro pagator,\n2GF= ReGW+iImGWsign(t−t′), (18)\nis always of the same form as Grin eq. (5), and therefore characterized by a single (complex)\nfunction Σ F(ω,p). When restricting attentionto Gaussian models, Σ F(ω,p)is given by a 1-loop\ncalculation, whereas in general, it contains a series of 1PI graphs.\n6In non-vacuum states and in non-stationary situations, the 2pt f unctions and self-energies\nhave a more complicated structure [20]. To understand this struc ture it is useful to study\nseparately the commutator Gc=iImGWand the anti-commutator Ga= ReGW.\nLet us conclude with two remarks. First, the effective dispersion re lation of φisa posteriori\ndefined by the poles of eq. (5) which are governed by Σ( ω,p) [13, 14]. In this way, non-\ntrivial dispersion relations arise from dynamical processes rather than from being introduced\nfrom the outset. The present work therefore provides physical foundations (and restrictions,\nas later discussed) to the kinematical approach which is usually adop ted [1]. Second, from\nanalyzing dynamical models, we shall see that, even in the vacuum, on-shelldissipative effects\n(i.e. dissipation arising along the minima of the denominator of eq. (5)) are unavoidable when\nLIis broken in the UV by the action SΨ+Sφ,Ψ, in complete opposition with the fact that\non-shell dissipation is forbidden when working (in the vacuum) with LIactions.\n2.3 Gaussian models\nTo simplify the calculation of Σ( ω,p) and to get non perturbative expressions, we assume that\nthe action STis quadratic in all field variables. At first sight, this could be considere d as an\nartificial hypothesis. However, it should be recalled that we are not after computing Σ from\nfirst principles. Rather we aim to compute the signatures of the pow er spectrum (3) giventhe\nproperties of Σ, following the approach adopted in [3, 5, 6, 7].\nGiven that wearepreserving thehomogeneity of Minkowski space, theGaussianassumption\nimplies that the total action splits as\nST=/integraldisplay\nd3pST(p), (19)\nwhere each action ST(p) depends only on φpand the p-th Fourier component of Ψ. The\nstructure of these actions is\nST(p) =1\n2/integraldisplay\ndt φ∗\np(−∂2\nt−ω2\np)φp+1\n2Σi/integraldisplay\ndtΨ∗\ni(p)(−∂2\nt−Ω2\ni(p))Ψi(p)\n+Σi/integraldisplay\ndt gi(p)φpΨ∗\ni(p), (20)\nwhereiis a discrete (or continuous) index, where Ω i(p) is the energy of the quanta of the\noscillators Ψ i(p), and where gi(p) is the coupling constant at fixed p,i. Since φ†\np=φ−p,\nΨ†\ni(p) = Ψ i(−p),ST(p) +ST(−p) is real. Instead of choosing a priori one model, it is more\ninstructive to solve the equations of motion without specifying the s et of Ψ i(p), their energy\nΩi(p) and their coupling gi(p). We shall choose them in due course. To preserve stationarity\nin Minkowski space, the Ω’s and the g’s must be time independent. When Ω2\ni(p)∝ne}ationslash=M2\ni+p2,\nthe kinetic action of Ψ i(p) breaks LIand defines the preferred frame. On the contrary when\nΩ2\ni(p) =M2\ni+p2the preferred frame is only defined by SφΨthrough the p-dependence of the\ncoupling functions gi(p).\nModels of this type have been used for different purposes. They ha ve been introduced\n(in their continuous version) to study non-pertubatively atomic tr ansitions see [15] and refs.\n7therein, see also [16, 17] for an application to the Unruh effect. The y have been used in Quan-\ntum Optics [18], to model quantum Brownian motion, and to study dec oherence effects [19].\nDepending on the aims, they can be solved and analyzed by means of d ifferent methods. In\nwhat follows we shall use the simplest approach based on Heisenberg picture.1Since we shall\nwork at zero temperature, this approach offers a simple characte rization of the state of the\nsystem in terms of the ground states of the free modes before φ−Ψ interactions are turned on.\nTo prepare the application to inflation, we shall treat ω2\np, Ω2\niandgias arbitrary functions\nof time (in cosmology these quantities become time dependent throu gh their dependence in the\nscale factor a(t)). The equations of motion are\n(∂2\nt+ω2\np)φp= Σigi(p)Ψi(p), (21)\n(∂2\nt+Ω2\ni)Ψi(p) =gi(p)φp. (22)\nThe general solution of the second equation reads\nΨi(p,t) = Ψo\ni(p,t)+/integraldisplay\ndt′Ro\ni(t,t′;p)gi(t′;p)φp(t′), (23)\nwhere Ψo\ni(p,t) is a free solution which depends on initial conditions imposed on Ψ i(p). The\nsecond term contains Ro\ni(t,t′;p), the (free) retarded Green function of Ψ i(p). It obeys\n(∂2\nt+Ω2\ni(p))Ro\ni(t,t′;p) =δ(t−t′), (24)\nand vanishes for t < t′. Injecting eq. (23) in eq. (21) one gets\n(∂2\nt+ω2\np)φp= Σigi(t;p)Ψo\ni(p,t)+Σigi(t;p)/integraldisplay\ndt′Ro\ni(t,t′;p)gi(t′;p)φp(t′).(25)\nThe solution of this equation has always the following structure\nφp(t) =φd\np(t)+/integraldisplay\ndt′Gr(t,t′;p)[Σigi(t′;p)Ψo\ni(p,t′)]. (26)\nThe first term is the ”decaying” solution. It contains all the informa tion about the initial\nconditionof φp. Thesecondtermisthe”driven”solution. Itisgovernedbytheinitia lconditions\nof Ψi(p) and by the (dressed) retarded Green function, the solution of\n/integraldisplay\ndt′[δ(t−t′)(∂2\nt′+ω2\np)−Σigi(t;p)Ro\ni(t,t′;p)gi(t′;p)]Gr(t′,t1;p) =δ(t−t1).(27)\nNotice that φd\np(t′) is an homogeneous solution of this equation. Therefore the evolut ion of\nbothφdandGrfully takes into account, through the non-local term in the above b racket, the\n1Even though legitimate, we shall not use the general methods (Infl uence Functional, Master Equation,\nClosed Time Path Integral) [18] which have been developed to study ” open quantum systems” because they\nsomehow hide the simplicity of the present models. Moreover we are p lanning (in a subsequent work) to study\nthe correlations between φand Ψ i. Therefore we shall treat φand Ψ ion equal footing as in [15, 16, 17].\n8back-reaction due to the coupling to the Ψ i. In Gaussian models, it is quadratic in gi. Hence\nφdandGrare series containing all powers of gi. Moreover since gi(t;p) are arbitrary functions\nofpandt, at this point, there is no reason to consider non-Gaussian models.\nTo conclude this subsection, we notice that eq. (26) also furnishes the exact solution for the\n(Heisenberg) mode operator φp(t) because the equations we solved were all linear. Since the\npower spectrum in inflation is obtained from vacuum fluctuations, ins tead of further analyzing\nthe time dependence of the mode (as one would do in classical terms) , it is more relevant to\nstudy the correlation functions of φp(t).\n2.4 Structure of two-point correlation functions\nThis sub-section mainly contains well-known results which follow from t he linearity of eq. (26).\nThe key result we shall later use is given in eq. (31).\nSince our models are Gaussian, the 2pt function of eq. (16) govern sallobservables built\nwith the Heisenberg field operator φ. To analyse it, as already mentioned, it is appropriate to\nstudy separately the commutator and the anti-commutator.\nWe start with the simple part, the commutator\nGc(t,t′;p)δ3(p−p′)≡Tr[ρT[φp(t),φ†\np′(t′)]−]. (28)\nFrom eq. (26) one sees that it decomposes into two terms, one due to the non-commuting\ncharacter of φd, the other due to that of Ψ0\ni. In addition, since both commutators are c-\nnumbers, it is independent of ρT, the state of the system. Hence, for all Gaussian models, one\nhas\nGc(t,t′;p) = [φd(t),φd(t′)]−+/integraldisplay/integraldisplay\ndt1dt2Gr(t,t1)Gr(t′,t2)D(t1,t2), (29)\nwhere the ”dissipative” kernel D(t1,t2) is given by\nD(t1,t2) = ΣiΣjgi(t1)gj(t2)[Ψo\ni(t1),Ψo\nj(t2)]−= Σigi(t1)Go\nc,i(t1,t2)gi(t2).(30)\nNotice how this kernel combines the coupling giand the non-commuting properties of Ψ i.\nThe next property of Gcis more relevant. To all orders in giand for all sets of gi,Ωi(even\nwith arbitrary time dependence), one obtains\ni∂tGc(t,t′;p)|t=t′= 1. (31)\nThis identity corresponds to the ETC of eq. (11). The 1 on the rhsis guaranteed by the\nHamiltonian character of the evolution of the entire system φ+Ψ. It is therefore this equation\nwhich replaces the constancy of the Wronskian that was relevant in the case of free evolution.\nEq. (31) is crucial for us for two reasons. First, since the operat orφd(t) in eq. (26)\ndecays in t−tinwheretinis the moment when the interactions are turn on –because it is an\nhomogeneous solution of eq. (27)– the first term in eq. (29) decay s as exp−γ(t+t′−2tin).\nTherefore when γ(t+t′−2tin)≫1, the non-commuting properties φpareentirely due to those\nof the environment degrees of freedom, Ψo\ni. Secondly, these sum up exactly to 1, as if the driven\nterm of eq. (26) were a canonical degree of freedom.\n9We now analyze the anti-commutator,\nGa(t,t′;p)δ3(p−p′)≡Tr[ρT{φp(t),φ†\np′(t′)}+]. (32)\nWhen the (initial) density matrix factorizes, ρT=ρφρΨ, as it is the case in the ”free” vacuum\nbefore the interactions are turned on, Gaalso splits into two terms,\nGa(t,t′;p) = Tr[ρφ{φd(t),φd(t′)}+]+/integraldisplay/integraldisplay\ndt1dt2Gr(t,t1)Gr(t′,t2)N(t1,t2).(33)\nThe first term only depends on the initial state of φ. Similarly, the driven term only depends\non the state of the environment through the ”noise” kernel\nN(t1,t2) = ΣiΣjTr[ρΨ{gi(t1)Ψo\ni(t1),gj(t2)Ψo\nj(t2)}+]. (34)\nAs for the commutator, in the presence of dissipation, the first te rm exponentially decays,\nexpressing the progressive erasing on the information contained in the initial state of φ. At\nlate times therefore it is the state of Ψ which fixes the anti-commuta tor ofφ. This allows to\nremove the restriction that initially the density matrices factorizes . If one is interested by the\nlate time behavior, only Nmatters.\nIn brief, we have recalled two important results. First, at late time, the Heisenberg field\nφreduces to its driven term, the second term of eq. (26), since bot h its commutator and\nanti-commutator are determined by those of Ψo\ni. Second, only two (real) kernels determined\nby the environment govern the two-point functions of φ, namely DandNof eqs. (30, 34).\nTherefore the set of environments (Gaussian or notGaussian) possessing the same kernels will\ngive rise to the same 2pt functions. Hence they should be viewed as f orming an equivalent\nclass. The degeneracy can be lifted by considering correlations with observables containing the\noperators Ψ i, or higher order correlations functions of φ(for non-Gaussian environments), two\npossibilities we shall not discuss in this paper.\nTo compute GcandGa, two different routes can be adopted. When gi(p) and Ω( p) are\nconstant, one should work in Fourier transform because the equa tions can be algebraically\nsolved, in full generality. Instead when gi(p) and/or Ω( p) are time-dependent, as it will be the\ncase in expanding universes and in curved space-times, it is appropr iate (but not necessary) to\nexploit the above mentioned degeneracy by choosing the Ψ iand their frequencies Ω2\ni(p) so as\nto simplify the time dependence of the equations. In the text, we pr oceed with time dependent\napproach. In Appendix A, we present the Fourier analysis which is st raightforward. We invite\nthe reader unfamiliar with the Quantum Mechanical treatment of dis sipation to read it.\n2.5 Time dependent settings\nIn Appendix B, we provided a class of models characterized by the po wer ofp/ΛLVwhich\nspecifies how dissipative effects grow, see eq. (85). This class cove rs the general case and can\nbe used as a template to study the consequences of dissipative effe cts. In addition, as noticed\nafter eq. (76), the dissipative effects are governed by D= Σig2\niRo\ni. Therefore all environments\ndelivering the same kernel give rise to the same (stationary) pheno menology.\n10In this Section we exploit this freedom having in mind the transposition of our model from\nMinkowski space to cosmological, and thus time-dependent, metric s. Therefore the selected\nmodels should possess two-point functions with simple properties wh en expressed in the terms\nof time (as opposed to Fourier components). The core of the prob lem is that Gadepends, see\neq. (33), on the retarded Green function of φwhich is not known. Indeed Gris only implicitly\ndefined as a solution of eq. (27) which is, in general, a non-local differential equation. We\nare thus led to choose Ψ in order for this equation to be local. This implie s that the retarded\nGreen function of Ψ appearing in eq. (27) be proportional to δ(t−t′). (Since this requirement\nconcerns the Green function of the environment it does not restr ict the phenomenology of φ.)\nGiven this aim, we select the models defined by the action\nS(n)\nT(p) =1\n2/integraldisplay\ndt φ∗\np(−∂2\nt−ω2\np)φp+1\n2/integraldisplay\ndt/integraldisplay∞\n−∞dkΨ∗(p,k)(−∂2\nt−(πΛLVk)2)Ψ(p,k)\n+gΛLV/integraldisplay\ndt/integraldisplay∞\n−∞dk/parenleftbiggp\nΛLV/parenrightbiggn+1\nφp∂tΨ∗(p,k). (35)\nInSφΨhave factorized out a factor of Λ LVso that the coupling constant gis dimensionless.\nWhen compared with the action of eq. (20), we have replaced the dis crete index iby the\nintegral over the dimensionless variable k. As recalled in Appendix A, the spectrum of the\nenvironment must be continuous to have proper dissipation, see dis cussion after eq. (76). The\nvariablekcan be viewed as a momentum, in the units of Λ LV, in a flat extra spatial dimension.\nThe relationship with some Brane World Scenarios in then clear [9, 10]. I n the ’atomic’ version\nof this model [16, 17] which has inspired us, the radiation field Ψ is a ma ssless 2 dimensional\nfield propagating in the dimension associated with k.\nWe have also introduced an additional time derivative acting Ψ in Sφ,Ψ. This choice leads\nto the above mentioned δ(t−t′). Indeed, on the one hand, taking account this extra derivative,\nthe continuous character of k, and the fact that gis independent of k, eq. (25) becomes\n(∂2\nt+ω2\np)φp=gn∂t/integraldisplay\ndkΨo(p,k,t)−gn∂t/integraldisplay\ndt′/integraldisplay\ndkRo(t,t′;k,p)∂t′/parenleftbig\ngnφp(t′)/parenrightbig\n,(36)\nwheregn≡gΛLV(p/ΛLV)n+1. On the other hand, for each 3-momentum p,Ψ=/integraltext\ndkΨ(k) is a\nmassless 2-dimensional free field. In Fourier components, its reta rded Green function is given\nbyRo(ω,k) = 1/(−(ω+iǫ)2+(πΛLVk)2). Hence Ro(t,t′) obeys\n∂tRo(t,t′)≡∂t/integraldisplay∞\n−∞dω\n2π/integraldisplay∞\n−∞dkRo(ω,k)e−iω(t−t′)=δ(t−t′)\nΛLV, (37)\nwhich is the required property to simplify eq. (36).\nWhengis constant, the retarded Green function of φassociated to eq. (36) obeys the\nfollowing localequation\n[∂2\nt+g2\nn\nΛLV∂t+ω2\np]Gr(t,t′,p) =δ(t−t′), (38)\n11To make contact with Appendix A and B, let us rewrite this equation in F ourier transform:\n[−ω2−ig2ω\nΛLVp2/parenleftbigp\nΛLV/parenrightbig2n+ω2\np]Gr(ω,p) = 1. (39)\nWe thus see that ReΣ r= 0 and that ImΣ ris (exactly) given by g2times that of eq. (85).\nThus, even though we have chosen a simple form for Ro(t,t′), the above action delivers the n\ndissipative behaviors of Appendix B by choosing the appropriate pow er ofp/ΛLVinSφΨ.\nWhengandω2\nparearbitrarytime-dependent functions, theFourieranalysis loos esitspower.\nHowever, in time dependent settings, our Grstill obeys a local equation:\n/bracketleftbig\n∂2\nt+2˜γn∂t+[ω2\np+∂t˜γn]/bracketrightbig\nGr(t,t′) =δ(t−t′). (40)\nwhere the n-th decay rate ˜ γn(t) =g2(t)γnis now a definite time dependent function.\n2.6 Covariant description\nWe now provide the covariantized expression of the action ST=/integraltext\nd3pST(p), where ST(p) is\ngivenineq. (35). This expression will thenbeusedtodefineourtheo ryincurved backgrounds.2\nTwo steps should be done. We need to go from pconsiderations to a local description, and\nexpress the various actions in terms of the unit vector field lµand the spatial metric ⊥µν. Both\nare straightforward and, in arbitrary coordinates, the action re ads\nST=−1\n2/integraldisplay\nd4x√−ggµν∂µφ∂νφ\n+1\n2/integraldisplay\nd4x√−g/integraldisplay\ndk/bracketleftBig/parenleftbig\nlµlν−c2\nΨ⊥µν/parenrightbig\n∂µΨ(k)∂νΨ(k)−(πΛLVk)2Ψ2(k)/parenrightbig/bracketrightBig\n+gΛLV/integraldisplay\nd4x√−g/parenleftBig/parenleftbig∆\nΛ2\nLV/parenrightbig(n+1)/2φ/parenrightBig\nlµ∂µ/integraldisplay\ndkΨ(k), (41)\nwhere the symbol ∆ is the Laplacian on the three surfaces orthogo nal tolµ.\nWe have slightly generalized the action of eq. (35) by subtracting c2\nΨ⊥µνto the kinetic term\nof Ψ, where c2\nΨ≪1. With this new term, the Ψ( k) are now massive fields which propagate\nwith a velocity whose square is bounded by c2\nΨ. In addition they now possess a well defined\nenergy-momentum tensor which can be obtained by varying their ac tionwith respect to gµν. To\nobtain the simplified expressions we have used (and shall still use), t he (regular) limit c2\nΨ→0\nshould be taken.\n2This procedure perhaps requires further explanation since Ψ is not a fundamental field but ”nothing more\nthanaconvenientparameterizationofsomeenvironmentdegrees offreedom.”First,Ψhasnotbeenintroducedto\nparameterizethedissipativeeffectsarisingfromanytheory, buto nlythosefromtheoriesobeyingtheEquivalence\nPrinciple, see [9] for a prototype. Second, from the point of view of cond-mat physics, it could a priori seem\ninappropriate to proceed to a covariantization, since in most situat ions (heat bath) there is a preferred frame\nwhich is globally defined. However, when the system is non-homogene ous (e.g. a fluid characterized by a non-\nhomogeneous flow), low energy fluctuations effectively live in a curve d geometry[28]. Moreover, in this case,\nshort distance effects, i.e. dispersive (or dissipative) effects, are covariantly described when using this metric\nbecause they arise locally[3, 11]. Hence the covariantization is both a necessary step to implem ent the EP in\nour settings and a property that emerges in cond-mat physics.\n12In this limit, the (free) retarded Green function of the Ψfield obeys a particularly simple\nequation when expressed in space-time coordinates:\nlµ∂\n∂xµ/integraldisplay\ndkRo(x,y;k) =lµ∂\n∂xµRo(x,y) =1\nΛLVδ4(xµ−yµ)√−g. (42)\nOn the r.h.s, one finds the delta function with respect to the invarian t measure d4x√−g. This\nequation is nothing by the covariantized and ”localized” version of eq . (37). Its physical\nmeaning is clear. In our model, the back-reaction of φ(x) onto itself through Ψis local.\nIn spite of this, the equations of motions do not have a particularly s imple form. Using the\ncondensed notion Ψ=/integraltext\ndkΨ, one gets\n1√−g∂µ√−ggµν∂νφ(x) =gΛLV1√−g/parenleftbig∆\nΛ2\nLV/parenrightbign+1\n2√−glµ∂µΨ(x), (43)\nwhere the interacting Ψ(x) field is\nΨ(x) =Ψo(x)−gΛLV/integraldisplay\nd4y√−gRo(x,y)/bracketleftBig1√−g∂µ/parenleftBig\nlµ√−g/parenleftbig∆\nΛ2\nLV/parenrightbign+1\n2φ(y)/parenrightBig/bracketrightBig\n.(44)\nWhen inserting eq. (44) in eq. (43), using eq. (42), one verifies tha t the dissipative term is local\nand first order in lµ∂µ. Therefore, as expected, dissipation occurs along the preferre d direction\nspecified by the vector field lµ.\n2.7 Dissipative effects in curved background geometries\nTo define a dissipative QFT in an arbitrary curved geometry, one nee ds some principles. From\na physical point of view, we adopt the Equivalence Principle, or bette r what can be considered\nas its generalization in the presence of the unit time-like vector field lµ. We are in fact dealing\nwith two (set of) dynamical fields, the φfield we probe, and the Ψfield we do not; but also\nwith two background fields gµνandlµ. The Generalized Equivalence Principle means that the\naction densities of the dynamical fields be given by scalar functions ( under general coordinate\ntransformations) which coincide to those one had in Minkowski spac e time for a homogeneous\nand static lµfield, i.e. those of eq. (41).\nHowever, the densities are not completely fixed by the GEP. In this w e recover what was\nobtained withthe EP: Fora scalar field ina curved geometry, there w as always thepossibility of\nconsidering a non-minimal coupling to gravity by adding to the Lagran giana term proportional\ntoRφ2. In the present case, the ambiguity is larger because lµdefines new scalars, the first of\nwhich is the expansion Θ = ∇µlµ, where∇µis the covariant derivative with respect to gµν. The\nambiguity can only be resolved by adopting some additional principle, s uch as the principle of\nminimal couplings which forbids adding densities containing these scala rs.\nRather than adopting it, we shall choose the non-minimal coupling so as to keep eq. (42),\ni.e. so that the localityof the back-reaction effects of φthroughΨbe preserved. This choice\nmaintains the simplicity of the equations of motion in curved backgrou nds, but is by no means\nnecessary. Starting from eq. (41), the locality is preserved by re placing in SΨandSΨφ\nlµ∂µΨk→ DlΨk≡lµ∂µΨk+Ψk\n2Θ =1\n2(lµ∇µΨk+∇µ[lµΨk]). (45)\n13To simplify the forthcoming equations, we use the fact that one can always work in ”pre-\nferred” coordinate systems in which the shift livanishes and in which the preferred time is such\nthatl0= 1. (We assume that the set of orbits of lµis complete and without caustic. In this\ncase, every point of the manifold is reached by one orbit.) In these c oordinate systems, it is\nuseful to work with rescaled fields Ψr≡(−g)1/4Ψ because the above equation simplifies\nDlΨk= (−g)−1/4lµ∂µ((−g)1/4Ψk) = (−g)−1/4lµ∂µΨr\nk, (46)\nsince Θ = ( −g)−1/2∂µ[(−g)1/2lµ]. Notice also that there exists a subclass of background fields\n(g,l), for which one can find coordinate systems such that boththe shift liandgoivanish. In\nthesecomoving coordinate systems, the above equations further simplify since on ly the spatial\npart of the metric matters because −g=hcwhereh≡det(⊥ij).3\nHaving chosen this non-minimal coupling, one verifies that the kinetic term of the rescaled\nfields Ψ ris insensitive to the ”curvature” of both gµνandlµ(when the limit c2\nΨ→0 is taken).\nMoreover the differential operator which acts on the retarded Gr een function of Ψ rin the\nequation of motion of φ, see eqs. (43, 44), is also ”flat” thereby guaranteeing that the m odified\nversion eq. (42) still applies, that is\nDlRo(x,y) = (−g(x))−1/4lµ∂\n∂xµ/parenleftBig\nRo\nr(x,y)/parenrightBig\n(−g(y))−1/4\n=1\nΛLVδ4(xµ−yµ)√−g, (47)\nwhereRo\nr(x,y) is the retarded Green function of the rescaled field Ψ r. It obeys (in pre-\nferred coordinate systems) ∂tRo\nr=δ4/ΛLV, and ”defines” the retarded Green function Ro=\n(−g)1/4Ro\nr(−g)1/4which is a bi-scalar. Hence the equations of motion in an arbitrary bac k-\nground ”tensor-vector metric” specified by the couple ( gµν,lµ), are given by eqs. (43, 44) with\nthe substitution of eq. (45).\nSeveral remarks should be made. First, from the simplified equation ∂tRo\nr=δ4/ΛLVit\nmight seem that the background tensor metric gµνplays no role. This is not true, since it is\nused to normalize the field lµat every point.\nSecond, in the limit c2\nΨ→0, the (rescaled) Ψ kfields define a new kind of field. They\npropagate in an effective space-time given by the time development o f the 3-dimensional set\nof orbits of the lµfield. Indeed, at fixed k, Ψk(x) can be decomposed in non-interacting local\nfield-oscillators, each of them evolving separately along its orbit. Th is situation is similar to\nthe long wave length (gradient-free) expression of [32]. In the ab sence oflµ, the geometry must\nbe (nearly) homogenous for the action to posses this decompositio n. However, when lµis given,\n3It is clear that this is the case when lµcoincides with the cosmological frame and when one uses comoving\ncoordinates ds2=−dt2+a2dx2sincelµ= (1,0). However, in certaincases one should searchfor the ”comoving ”\ncoordinate system. To illustrate this point, consider the former sit uation in Lemaˆ ıtre coordinates X=ax. In\nthis case one has ds2=−dt2+ (dX−Vdt)2, where the velocity is V=HX. The spatial sections are now\nthe Euclidean space with h= 1, and the (contravariant) components of the unit vector field a relµ= 1,V.\nTo compute hcone should solve the equation of motion of comoving (free falling) obs erversdX−Vdt= 0,\nand use the initial position as new coordinates. This procedure is exp licitely done in [12] when starting with\nPainlev´ e-Gullstrand coordinates to describe the black hole metric a nd using a freely falling frame.\n14one can identify, even in non-homogeneous metrics, each space-t ime point in an invariant way\nby the spatial positionof the corresponding ”preferred” orbit at some time, andthe proper time\nalongtheorbit(aslongas lµhasnocaustic). Wecanthus buildcovariant actionsexploiting this\npossibility and consider fields composed of a dense set of local oscillat ors at rest with respect\ntolµ. The fields Ψ kwe use belong to this class of fields.\n3 Dissipative effects in cosmology\n3.1 The action\nOur aim is to describe dissipative effects in an expanding homogeneous universe when the\nvector field lµis aligned along the cosmological frame, when the dissipative effects a re known\nin Minkowski space, and when implementing the Equivalence Principle.\nIn this case, to get the action we simply consider eq. (41) (with the c urved metric modifi-\ncations discussed in the former subsection) in a FLRW metric. Using c omoving coordinates,\nds2=−dt2+a2(t)dx2, (48)\nthe components of lµare (1,0). To simplify the notations, we use the conformal time dη=dt/a\nand work with the rescaled fields φr=aφand Ψ r=h1/2\ncΨ =a3/2Ψ. Dropping these rindices,\nworking in Fourier transform with respect to x, the non-minimally coupled action associated\nwith the replacement of eq. (45) is\nS(n)\nT(p) =1\n2/integraldisplay\ndη φ∗\np(−∂2\nη−ω2\np(η))φp\n+1\n2/integraldisplay\ndt/integraldisplay\ndkΨ∗(p,k)(−∂2\nt−(πΛLVk)2)Ψ(p,k)\n+/integraldisplay\ndηgn(η)φp∂η/integraldisplay\ndkΨ∗(p,k). (49)\nThe conformal frequency ω2\np(η) =p2−∂2\nηa/ais that of a rescaled minimal coupled massless\nfield. In this expression, as everywhere in this Section, pis now the conformal (dimensionless\nand constant) wave vector. The time dependent coupling coefficien t is\ngn≡g a1/2ΛLV(p/aΛLV)n+1. (50)\nIts dependence in a(given that of p) follows from having implemented the Generalized Equiv-\nalence Principle (GEP) which determines the powers of afor each term in the action. As we\nshall see, this relative power guarantees that all φpwill be damped at a fixed and common\nproper scale (in the adiabatic approximation). Had we started with a n action of the type (49)\nwithout relying on the GEP, the dependence in awould have been arbitrary, and the proper\nscale at which modes would have been damped would have run as well.\nIn the same vein, one sees that the proper frequency of the Ψ field s stays constant. This\nfollowsfromtheGEPbut also fromourchoice ofnon-minimal couplings . Inthispaper, we want\n15indeed to analyse the phenomenology of dissipative effects when the new degrees of freedom\nare and stay in their ground state. One could have chosen more com plicated models in which\nΨ is parametrically excited. This would be the case when introducing a n on zero velocity cΨ,\nand with minimally coupling (by dropping Θ on the rhs of ≡in eq. (45)). However, when Ψ is\ntaken massive (i.e. by restricting the krange to |k|>1), or discretizing kas it would be the\ncase for Kaluza-Klein modes, and if Λ LV≫H, the phenomenology of all these models coincide\nsince the amplitude for parametric excitations will be exponentially da mped. In this regime\nthere is thus no gain in studying more complicated actions than that g iven in eq. (49).\n3.2 Equation of motion\nUsing the results of the subsection 2.5, the equation of motion of He isenberg operator φpis\n/parenleftbig\n∂2\nη+2γn∂η+(ω2\np(η)+∂ηγn)/parenrightbig\nφp=gn∂ηΨo(p), (51)\nwhere the decay rate in conformal time is,\nγn(η) =g2\nn\n2ΛLV=1\n2(aΛLV)/parenleftbigp\naΛLV/parenrightbig2n+2. (52)\nIt is dimensionless, as it should be. It should be compared with the com oving frequency pto\nget the relative strength of dissipation. One obtains\nγn(η)\np=1\n2/parenleftbigp\naΛLV/parenrightbig2n+1=1\n2/parenleftbigpphys\nΛLV/parenrightbig2n+1. (53)\nIn the last equality we have re-introduced the proper momentum pphys=p/a. With this\nequation we verify that, at any time in an expanding universe and for every mode φp, the\nrelativestrengthofdissipationissimplythatofMinkowski spaceeva luatedatthecorresponding\nenergy scale, see eq. (86). This directly results from having impleme nted the GEP.\nNotice however that the equation of motion in an expanding universe contains a frequency\nshift\ng2\nn\nΛLV∂η(lngn) =p2/parenleftbigaH\np/parenrightbig/parenleftbigp\naΛLV/parenrightbig2n+1. (54)\nWe have factorized out the unperturbed frequency square to ge t the relative value of the shift.\nIt vanishes both when dissipation is negligible and when aH/p≪1, i.e. when the expansion\nrateHis negligible with respect to the proper momentum. Fromthis express ion we can already\nconcludethatitcannotplayanyrolewhenthetworelevantscales HandΛLVarewell separated,\ni.e. when\nσ≡H\nΛLV≪1. (55)\nIndeed when the physical momentum is high and of the order of Λ LV, the relative frequency\nshift proportional to σ, and when the physical momentum is of the order of H(at horizon exit,\nsee Figure 1), it is proportional to σ2n+1.\nIn quantum settings however, it is not sufficient that the equation o f motion possesses its\nMinkowskian form because the state of the system, and therefor e the observables, might be\naffected by the combined effect of dissipative effects and the expan sion rate.\n163.3 Power Spectrum\nLet us consider the power spectrum of a scalar massless test field ( which is the relevant case\nfor gravitational waves and density perturbations) both in the st andard free field settings and\nin the presence of dissipative effects.\n3.3.1 Free settings\nThe equation of motion is the same as in eq. (10) with\nω2\np=p2−2\nη2. (56)\nFor simplicity we consider the inflationary background to be de Sitter wherea(η) =−1/Hη\nandη <0. At the onset of inflation, when pphys/H=p/ainH=p(−ηin)→ ∞, the positive\nfrequency modes are\nφin\np=1\n(2p)1/2/parenleftbigg\n1−i\npη/parenrightbigg\ne−ipη. (57)\nWhenp|η|= 1, the physical wave length λ=a/pbecomes larger than the Hubble radius. Near\nthat ”horizon-exit”, ω2\npflips sign, the mode stops oscillating and starts to grow like a. At late\ntime with respect to horizon exit, when |pη| ≪1, this implies that the power spectrum becomes\na constant, as we now recall.\nWhen inflation lasts long enough, i.e. when the number of extra e-fold s obeysNextra≫1,\nsee eq. (1), all observable modes φpare in their ground state at the onset of inflation (simply\nbecause this is the only state compatible with inflation [27]). In this cas e, the anti-commutator\nof the free field, see eq. (32), when evaluated at equal time ηis simply given by\nGfree\na(η,p) =|φin\np|2=1\n(2p)/parenleftbigg\n1+1\n(pη)2/parenrightbigg\n. (58)\nThen the power spectrum of the physical (un-rescaled) field given by\nPfree\np(η) =p3\n2π2Gfree\na(η,p)\na2(η)=/parenleftbigH2\np\n2π/parenrightbig2(1+(pη)2), (59)\nbecomes constant after horizon exit. We have added a psubscript to Hbecause in slow\nroll inflation, the relevant value of Hfor thep-mode is that evaluated at horizon exit, i.e.\nHp=H(tp), where tpis given by p/a=H. The above equation shows that Ppacquires\nsome scale dependence only through Hp. Similarly the deviations from this standard behavior\nstemming from some UV modification of the theory will also depend on pthroughHp(and its\nderivatives). For explicit examples, we refer to [30] where the mod ifications of the spectrum\nstem from the fact that pηinis taken large but finite, and also to [31] wherein the UV cutoff is\nendowed with a finite width. This last case bears many similarities with th e dissipative settings\nwe now study.\n173.3.2 Dissipative settings\nIn the presence of dissipation, the expression for Garadically differs from the above.\nAt the level of the Heisenberg operator, when dissipative effects g row with the energy (as\nwe suppose it is the case), the decaying solution of eq. (51) is comple tely erased (unless one\nfine-tunes Nextra, see eq. (1), so as to keep a residual amplitude). That is, in inflation the mode\noperator is entirely given by its driven term, the second term in eq. ( 26). Then the power\nspectrum is also purely driven and given by the second term of eq. (3 3):\nGdriven\na(η,p) =/integraldisplay\ndη1/integraldisplay\ndη2Gr(η,η1,p)Gr(η,η2,p)N(η1,η2,p), (60)\nwhereGris the retarded Green function, solution of eq. (51) with δ(η−η1) as a source, and\nwhere the kernel Nis the anti-commutator of gnlµ∂µΨp, the source of eq. (51).\nEq. (60) tells us that only the state of the environment matters. I n other words, because of\nthe strong dissipation at early times, the power spectrum is indepen dent of the initial state of\nφ, what ever it was. In spite of this, when eq. (55) is satisfied, and wh en the environment is in\nits ground state, the predictions are unchanged, i.e. the power sp ectrum obtained from Gdriven\na\ncoincides with that obtained with Gfree\na. To show this let us study GrandN.\n3.3.3 Dissipative Green functions and noise kernel\nThe retarded Green function is of the form\nGr(η,η0,p) =θ(η−η0)e−Rη\nη0dη′γ(η′)×2Im/parenleftBig\n˜φ∗\np(η)˜φp(η0)/parenrightBig\n, (61)\nwhere (in the under-damped regime) the modes ˜φpare unit Wronskian positive frequency\nsolutions of eq. (56) with a frequency square given by\n(ωeff\np)2=ω2\np−γ2\nn=p2(1−2\np2η2−γ2\nn\np2). (62)\nIn this, we obtain a time-dependent version of the stationary case , see eq. (87). With the\nsecond expression, we verify that when eq. (55) is satisfied, the t wo corrections terms are never\nsimultaneously relevant, which guarantees, as we shall discuss belo w, that dissipative effects\ncan be studied in the quasi-stationary approximation. Notice also th at the frequency shift ∂ηγ\npresent is eq. (54) drops out from ωeff.\nEq. (61) implies that\nGr(η,η0,p)→0 when η0→ηΛ\np, (63)\nwhereηΛ\npis the ”Λ LV-exit” time of the p-mode defined by p/a(ηΛ\np) = Λ LV, orγ/p= 1/2, see\neq. (53) and the Figure.\n18Ln a\nLn d\nFigure caption. We have represented in a log-log plot and by a dashed line the e volution of\ndH=RH/a, the comoving Hubble radius, as a function of a, both during inflation (dH∝1/a)\nand during the radiation era (dH∝a). We have represented by a thick line the trajectory\nof the cutoff length dΛ= 1/aΛLVof a fixed proper scale which obeys 1/ΛLV≪RHduring\ninflation. The dotted line corresponds to an intermediate fix ed proper length λwhich obeys\n1/ΛLV≪λ≪Rinfl.\nH. The vertical line represents a fixed comoving scale dp= 1/p. Below\nthe cutoff length, all modes are over-damped. When a mode exit s the cutoff length, it becomes\nunder-damped and starts propagating. When it crosses the in termediate length λ, it behaves\nas a free mode, and it gets amplified only when exiting the Hubb le radius. Adiabaticity, which\nis guaranteed by 1/ΛLV≪RH, guarantees in turn that, near λ, modes are all born in the\nBunch-Davies vacuum when the environment Ψis in its ground state.\nEq. (63) guarantees that no quantum coherence is left between la te timeη, e.g. horizon-\nexit, and what happened for times earlier than Λ LV-exit. The physical implication of this is\nthat, what ever transitions happened, what ever was the quantu m state, no record is kept at\nlate time.4Having understood the structure of the retarded Green functio n, let us now briefly\nconsider the commutator Gc(η,η0). It is given by\nGc(η,η0,p) =−i/parenleftBig\nGr(η,η0,p)−Gr(η0,η,p)/parenrightBig\n,\n=e−|Rη\nη0dη′γ(η′)|×2iIm/parenleftBig\n˜φp(η)˜φ∗\np(η0)/parenrightBig\n. (64)\nIt possesses two noteworthy properties. First, because of the absolute value in the damping\nexponential term, it obeys two different local differential equation s depending of the sign of\nη−η0. Whenη > η0,Gcis an homogeneous solution of eq. (51), whereas when η < η0, it obeys\nthe equation wherein the sign of the dissipative term has been chang ed. These two differential\n4This behaviorhasbeen described, drawn, andsometimes named anti-dissipation , see[5,11,7,33, 34, 35, 14].\nIn fact when tracing backwards in time vacuum configurations from the freely propagating regime down to the\ndissipative one, dissipation arises towards the past. When viewed as a function of η0,Gr(η,η0) indeed increases\nto the future, thereby showing anti-dissipation. From eq. (61) we understand the origin of this quantum\nproperty (without classical counterpart). As a function of η0, first,Grobeys eq. (51) with the ”wrong” sign for\nthe dissipative term, and second, the boundary condition is a final o ne and not an initial one, thereby explaining\nwhy when η0→η, the value of the current is ”adjusted” to become exactly one.\n19equations can be groupedinto a single non-local equation. Second, because of this ”quasi”-local\nproperties, Gcis independent of ηin, the moment when φwas put in contact with Ψ.\nThese properties of Gcfollow from the fact that its source term, the Dkernel of eq. (29),\nis ”ultra” local for the Ψfield of eq. (49):\nD(η1,η2)≡[gn∂η1Ψ(η1), gn∂η2Ψ(η2)] =gn(η1)i∂η2δ(η1−η2)\nΛLVgn(η2). (65)\nThis should be contrasted with the kernel N, the anti-commutator, which is non local. Indeed,\nin the vacuum, it is given\nN(η1,η2) =gn(η1)−a(η1)a(η2)\nπΛLV(t1−t2)2gn(η2), (66)\nwhere 1/(t1−t2)2should be understood as the derivative of the principal part of 1 /(t1−t2).\nOne gets a simple expression in terms of the proper time tbecause the proper frequency of Ψ k\nis constant in our model (49). We also remind the reader that it is only in the high temperature\nlimit that Nwould become proportional to δ(t1−t2).\nInthispaper weshallcomputethepower spectrumwhentheΨ kareallintheirgroundstate.\nThe reason for this choice is as follows: As pointed out in the footnot e 2, the Ψ kfields should\nbe conceived as parametrizing some degrees of freedom in a fundam ental theory obeying the\nEquivalence Principle. In these theories, when inflation lasts long eno ugh, all relevant degrees\nwill be in their ground state, as is the case when dealing with free mode s, see eq. (58). We see\nno reason why dissipation could possibly invalidate this conclusion.\nWhen the Ψ kare all in their ground state, the kernel Nis non-local. Therefore the anti-\ncommutator Gaobeys a non-local equation: eq. (51) with Nas the source. The physical\nmeaning is that the environment is driving φpin an non-instantaneous way (this is unavoidable\ninthevacuumsince onlyonesignofthefrequency ispresent). Thec onsequence isthatthevalue\nof power spectrum depends (to a certain extend) on the history o f the combined evolution of\nφ+Ψ. The mathematical implication is that anexact calculationof Gaiseffectively impossible.\nNevertheless when H/ΛLV≪1, these non local effects give only but sub-leading (non-\nadiabatic) corrections, because the evolution of φ+Ψconsists in a parametric (adiabatic)\nsuccession of stationary states ordered by the scale factor a.\n3.3.4 Scale separation and adiabatic evolution\nGiven the importance of this result, we shall spend some time to expla in its root and its\nvalidity. We proceed in two steps. We first show that adiabaticity is su fficient for obtaining\nthe standard power spectrum, and then show that scale separat ion, eq. (55), is sufficient to\nguarantee adiabaticity, thereby generalysing the results of [8].\nAdiabaticity means that the evolution proceeds slow enough for indu cing no (non-adiabatic)\ntransition (in molecular physics, they are called Landau-Zener tran sitions, in the present cos-\nmological context they correspond to pair creation of φquanta). When this is the case, the\nobservables take, at any time, the value they have in the correspo nding stationary situation.\n20As of retarded Green function and the commutator, this principle d oes not bring any new\ninformation because in the WKB approximation, they are local funct ions since these 2pt func-\ntions obey local equations. For the anti-commutator, adiabaticity guarantees that when ηand\nη′are close (in the sense that 1 −a(η)/a(η′)≪1), its value is well approximated by\nGdriven\na(η,η′,p)≃Gstatio\na(η−η′;ωp(a),gp(a)) =/integraldisplaydω\n2πeiω(η−η′)Ga(ω;ωp(a),gp(a)),(67)\nwhereGa(ω;ωp(a),gp(a)) is the Fourier component calculated with the values of the freque ncy\nωp(a) and the coupling gp(a) both evaluated with a=a(η). In Appendix A, these Fourier\ncomponents have been algebraically solved for all frequencies and a ll couplings, see eq. (81).\nMoreover, since by hypothesis, we are in the vacuum, eq. (79) also applies. In other words, the\nvalue ofGafollows from that of the commutator Gc. This is sufficient to guarantee that when\nthe mode φpbecomes free, i.e. much after Λ LV-exit but before horizon-exit ( H≪p/a≪ΛLV),\neq. (84) applies. Thus, irrespectively of what was the coupling with t he environment,\nGdriven\na(η,η,p)→Gfree\na(η,η,p) =1\n2ωp(a). (68)\nWith this equation we reach our first conclusion: in the adiabatic appr oximation and when\nthe environment is in its ground state, the modes φpare born in the usual in-vacuum (Bunch-\nDavies) as they become, one after the other, freely propagating after Λ LV-exit, see the Figure.5\nIt now behooves to usto show that scale separation guarantees t hat eq. (67) provides a valid\napproximationfor Gdriven\nabeforehorizon-exit, i.e. whenthesecondtermintheparenthesis in eq.\n(62)canstillbeneglected. Tothisend, weprovideanupperboundf ortheprobabilityamplitude\nof getting a non-adiabatic transition in the under-damped regime, i.e .H≪p/a≤ΛLV. This\namplitude is governed by the relative frequency change\n∂ηωeff\np\n(ωeff\np)2=−γ2\nn∂ηlnγn\n(p2−γ2n)3/2= (2n+1)Hγ2\nphys\n(p2\nphys−γ2\nphys)3/2, (69)\nwherepphys=p/aandγphys=γ/aare the physical (proper) momentum and decay rate.\nTherefore, going backwards in time from the free regime up to p2\nphys≥4γ2\nphys(i.e.pphys≤ΛLV),\nthe non-adiabatic parameter steadily grows but stays bounded by\n∂ηωeff\np\n(ωeff\np)2<3(2n+1)σ/parenleftBigpphys\nΛLV/parenrightBig4n+1\n<3(2n+1)σ≪1. (70)\nThis guarantees that the amplitude for the system to jump out of t he ground state is bounded\nbyσ(up to an overall factor which plays no role). There is no need to stu dy the stability\nof the ground state in the transitory regime from under-damped t o the over-damped modes\nbecause what ever transitions happened their residual impact at la te time would be suppressed\nby a factor e−Rγdt≃exp(−1/σ(2n+1))≪1. This completes the proof that scale separation\nguarantees adiabaticity, in agreement with the conclusion reached in [9, 10].\n5This behavior is in accord with the procedure of [29, 30, 31] wherein t he initial state is imposed on modes\nwhen their proper momentum reaches a given value. This radically diffe rs with the “initial slice” approach [36]\nwherein the state is imposed at the same time on all modes, irrespect ively of the value of their momentum.\nFrom what we see, the Equivalence Principle allows only the first to be d ynamically realized.\n213.3.5 Perspectives: Beyond the adiabatic approximation an d scale crossing\nIt is first interesting to point out that dissipative models are more ”r obust” than dispersive\nmodels in that the ”driven” power spectrum determined by eq. (60) is well defined even\nwhenH >ΛLV, contrary to the fact that most dispersive models make sense only when scale\nseparation ( H≪ΛLV) is realized. Therefore, with dissipative models, one can study scale\ncrossing, i.e. the crossing of Hpthrough Λ LVfrom above to below during slow roll inflation.\nThe power spectrum has been studied in this regime when using a part icular Brane World\nscenario in [10]. We agree with most of the conclusions but not with tha t reached after Eq.\n(35) according to which “towards the end of inflation, perturbatio ns on the brane are the sum\nof two independent fields”. We do not agree because in those settin gs as well the field operator\nis purely driven. Hence we do not see how to decompose it as a superp osition of two commuting\noperators. Moreover, with the approximation used in [10], it seems that the ETC, eq. (11), is\nviolated by a amount proportional to the reported change of the p ower spectrum.\nIn any case, the calculation of the modifications of the power spect rum introduced by\ndissipation is difficult because the non-local properties of eq. (60) c annot be neglected. We are\nnot aware of any analytical treatment, and we are currently stud ying the modifications using\nnumerical techniques [38].\nAdded note. Since the present paper was submitted, we have obtained several results [38].\nIn particular, the leading modification of the power spectrum with re spect to the standard\nresult in the regime σ=H/Λ≪1, i.e. the signature of UV dissipation, scales with a power of\nσwhich is equal to that of P/Λ inγ/p, see eq. (53). In this respect, dissipative models behave\nlike dispersive models where the leading modification scales with a power ofσwhich is that\nof the first non-linear term in the dispersion relation, i.e. 2 nusing the parameterization of eq.\n(2), see [39]. In the opposite regime, when H >Λ, no universal behavior is obtained.\n4 Conclusions\nIn this paper we have obtained the following results.\nFirst, we have provided a class of unitary models defined in Minkowski space which are\ncharacterized by the power of p/ΛLVwhich weighs the growth of dissipation in the preferred\nframe, see eq. (85). Unitarity is achieved by introducing a dense se t of additional fields\nΨwhich induce dissipation through interactions with φ. In Appendices A and B we have\ngiven a thorough analysis of the Green functions of these Gaussian models which cover the\nphenomenology of UV dissipative effects in Minkowski space at the lev el of 2pt functions.\nSecond, amongst the variousactions delivering the same stationar yphenomenology, we have\nselected one which gives rise to a local differential equation for the r etarded Green function\nofφ, see eqs. (36, 37). Using the Equivalence Principle, we have extend ed this class of\nmodels to arbitrary curved backgrounds in subsection 2.7, thereb y allowing to confront the\ntrans-Planckian question of inflationary cosmology and black hole ph ysics.\nThird, we have applied our dissipative models to inflationary cosmology . At early times,\nhence at very high energy, the dissipative effects are so strong th at all information about the\ninitial state of the φis erased, see eq. (63). Nevertheless, when the UV scale Λ LVis much\n22larger than the Hubble parameter, we have demonstrated that th e standard expression of the\npower spectrum is found when the environment is in its ground state . The reason for this is\nthe following: even though the field oscillator φpis purely driven by Ψp, i.e. it is given by the\nsecond term eq. (26), as its proper momentum p/aredshifts under Λ LV, the composite operator\nbehaves as if it were a free operator, see eqs. (31, 84), thereby guaranteeing eq. (68).\nLet us also note that our models can be used for phenomenological p urposes in the sense\nof [1], and that we are planning to apply them to study Hawking radiatio n in the presence of\ndissipation. We are presently completing the calculation of the power spectrum beyond the\nadiabatic approximation so as to determine the signatures of dissipa tion [38] and to compare\nthem to those of dispersion [39]. Finally it would be interesting to compu te the VEV of the\nstress-tensor of the Ψfields in non-trivial backgrounds. This would allow to take into accoun t\nthe back-reaction on the cosmological metric engendered by the fi eldsΨandlµ[40, 41, 42].\nAcknowledgments.\nI am grateful to Dani Arteaga and Enric Verdaguer for common wo rk allowing me to deepen\nmy understanding of dissipative effects. I am also grateful to Julian Adamek, David Campo,\nTed Jacobson, Jean Macher, Jens Niemeyer, and Valeri Rubakov f or interesting discussions. I\nwish to thank the organizers of the workshops on ”Micro and Macro structure of spacetime”\nheld inPeyresq in June 2005, 2006, and 2007where this work has bee n presented and discussed.\nThis work has been supported by the Agence Nationale de la Recherc he (projet 05-1-41810).\n5 Appendix A :\nStationary states and vacuum 2pt functions\nIn this Appendix, we recall the (well-known) relationships between Gc,Gawhich always hold\nin stationary states. In these states, Gc, Gaand the kernels D, Nare related to each other in\na universal way, generally referred as a Fluctuation-Dissipation re lation. We explain its origin\nand its physical implications in the present context. We start with th e most basic object: the\nretarded Green function Gr.\n5.1 The retarded Green function\nThe Fourier transform of eq. (25) gives\n(−ω2+ω2\np)φp(ω) = Σigi(p)Ψo\ni(p,ω)+Σig2\ni(p)Ro\ni(ω;p)φp(ω), (71)\nwhere\nRo\ni(ω;p) =/parenleftbig\n−(ω+iǫ)2+Ω2\ni(p))/parenrightbig−1, (72)\nis the Fourier transform (defined as in eq. (17)) of the retarded G reen function of Ψ i. As usual,\nits retarded character is enforced by the imaginary prescription o f the two poles to lay in the\nlower half plane ( ǫ >0). The solution of the above equation is\nφp(ω) =φd\np(ω)+Gr(ω,p)Σigi(p)Ψo\ni(p,ω), (73)\n23where the Fouriertransform ofthe retardedGreen functionof φ, thesolution ofeq. (27), always\ntakes the form\nGr(ω,p) =/parenleftbig\n−(ω+iǫ)2+p2+Σr(ω,p)/parenrightbig−1. (74)\nAll effects of the coupling to the Ψ i’s are thus encoded in the (retarded) self-energy Σ r(ω,p).\nFor Gaussian theories, it is algebraically given by\nΣr(ω,p) =−Σig2\ni(p)Ro\ni(ω,p). (75)\nThe dissipative effects are governed by the imaginary part of Σ r(ω,p). In the present case, one\nhas\n2ImΣr(ω,p) =−Σig2\ni(p)Go\nc,i(ω,p) =−D(ω,p). (76)\nTo get the first equality we have used the fact that in stationary st ates the retarded Green\nfunction and the commutator are related by 2Im Gr(ω) =Gc(ω) for all degrees of freedom, free\nor interacting. In the second equality, we have introduced D(ω), the Fourier transform of the\nkernel of eq. (30).\nSeveral observations should be made here. First, from eq. (72), we obtain that D(ω) is\nproportional to Σ ig2\niδ(ω−Ωi). Therefore there is no dissipation for lower frequencies than the\nlowest value of Ω i. This simply follows from energy conservation. Second, to obtain ”t rue”\ndissipation, D(ω,p) should be a continuous function and not a sum of delta. This can only\nhappen when the Ψ iform a dense ensemble. In Section 2.5, we shall thus replace the disc rete\nsum oniby an integral over a continuous variable, k. We shall not consider the discrete cases\neven though these could display interesting properties. Third, fro m a phenomenological point\nof view, only D(ω,p) matters. Hence we cannot disentangle the spectrum of the envir onment,\nwhich is given by Ro\ni(ω,p), from the coupling strength g2\ni(p). This is a good thing, because\nwhen working intime-dependent settings, we shall exploit this equiva lence to chose the simplest\nmodel of Ψ i’s which gives the kernel D(ω,p) we want.\nIt is also worth noticing that the dispersive (real) effects are not dir ectly related to D(or\nN). These are governed by the even part of Σ r(ω,p) which is given by\nReΣr(ω,p) =/integraldisplaydω′\n2πD(ω′,p)\nω−ω′, (77)\nwhere the integral should beunderstood asa principal value. This in tegral relationship explains\nwhy one often founds that dispersive effects appear before dissip ative effects (for increasing ω).\nWe also learn that the dispersive models studied in the literature violat e the above relation\nsince they assume ReΣ r∝ne}ationslash= 0 and ImΣ r≡0. Therefore these models are incoherent and cannot\nresult from dynamical processes.\n5.2 Fluctuation-Dissipation relations and vacuum self-en ergy\nIn this subsection, we derive the relationships between Gc,Gaand Σ Fwhich exist in the true\n(interacting) ground state.\n24In interacting theories, the only stationary states are thermal s tates, see e.g. [37]. In these\nstates, the Fourier transform of DandNare related by\nN(ω) =D(ω) coth(βω/2),\n=D(ω)sign(ω)[2n(|ω|)+1]. (78)\nIn the second line, n(ω) is the Planck distribution. It gives the mean occupation number of\nΨo\niquanta as a function of the frequency (measured in the rest fram e of the bath). The above\nrelation directly follows from the fact that the individual commutato rs and anti-commutators\nof the free fields Ψo\niobey this relation, as any free oscillator would do. It implies that the\nFourier transform of GcandGaare also related by\nGa(ω) =Gc(ω)sign(ω)[2n(ω)+1]. (79)\nIt should be stressed that this equation is exact, i.e. non-perturb ative, and valid for all theories,\nGaussian or not. (It indeed directly follows from the cyclic propertie s of the trace defining\nGβ(t,t′) =Tr[e−βHTφ(t)φ(t′)]).\nFor Gaussian models, there exists an alternative proof of eq. (79) . It suffices to note that\nin steady states the decaying terms of eqs. (29) and (33) play no r ole, and that the Fourier\ntransform of the driven terms are respectively given by\nGc(ω) =|Gr(ω)|2D(ω), (80)\nGa(ω) =|Gr(ω)|2N(ω), (81)\nsince the Fourier transform of the retarded Green function obey sGr(ω) =G∗\nr(−ω), see eq.\n(73). Irrespectively of the complexity of Gr, i.e. irrespectively of the functions gi(p), Ωi(p) and\nthe set of the Ψ ifields,GcandGaare thus related to each other by the FD relation (79).\nThese universal relations will be relevant for inflationary models whe rein only the ground\nstate contributes. In particular, they imply that in the true vacuum , i.e. when n(ω) = 0,Gc\nandGaare exactly related by Ga(ω) =Gc(ω)sign(ω). Hence the Wightman function\nGW=1\n2(Gc+Ga) =Gcθ(ω), (82)\nis determined by the commutator and contains only positive frequen cy, as in the free vacuum.\nEquations(80,81)alsoallowtocomputethevacuumself-energyof theFeynmanGreenfunction.\nFor Gaussian models it is given by\n2ImΣF(ω,p) =−D(ω,p)sign(ω). (83)\nWiththelastequalitywerecoverthefactthatinthevacuum, itissuffi cienttoconsiderFeynman\nGreen functions. In non-vacuum states, and in non-stationary s ituations, this is no longer true,\nthereby justifying the use of the in−inmachinery [20] (the Schwinger-Keldish formalism).\nBefore specializing to a specific class of models giving rise to dissipation at high frequency,\nwe make a pause by asking the following important question: What sho uld be known about\n25the Ψifields to get eqs. (80, 81, 82) ? We have proven that it is sufficient fo r the Ψ i’s to be\ncanonical fields, but is it necessary ?\nThe answer is two fold. On one hand, the Ψ icannot be stochastically fluctuating quantities\n(i.e. commuting variables) because this would lead to a violation of eq. ( 78) that would imply\nthe violation of eq. (79) and the ETC eq. (31).6They cannot be either a combination of\nquantum and stochastic quantities because this would still lead to a v iolation of the ETC.\nHence they must be built only from quantum (canonical) degrees of f reedom.\nOntheotherhand, theΨ i’scanbecompositeoperators, i.e. polynomialsofsome(unknown)\ncanonical fields. Indeed, their commutators would still be all relate d to their anti-commutators\nby the FD relation eq. (78), and this even though they depend non- linearly on n(ω) in non-\nvacuum states. The difference with Gaussian models is that these no n-linear operators posses\nnon-vanishing higher order correlation functions. Hence, the self -energies Σ r,ΣFwill be series\nin powers of g2\ni, and not just a single quadratic term as in eq. (75, 83). Neverthele ss these\nhigher loops corrections preserve the validity of eq. (82) in the gro und state, as well as that\nof eqs. (80, 81) in any thermal state, when properly understood , i.e. with Dnow defined by\n-2ImΣ r(as the effective dissipation kernel), and Nrelated to it by the FD relation.\nIn brief, we have reached/recalled the following results. Firstly, th e q-number combination\nΨ= ΣigiΨi, the fluctuating source term of φ, must obey the FD relation (78). This can\neither be postulated, or better, be viewed as resulting from the fa ct thatΨis entirely made\nout of quantum degrees of freedom. Secondly, to lowest order te rm ing, the self-energy can\nbe obtained by treating Ψas a Gaussian variable, what ever its composition may be. Thirdly,\nwhen dealing with non-Gaussian theories, once having computed Σ r(ω), the resulting equations\nfor the 2pt functions have the same structure and the same mean ing as in Gaussian theories,\nwithDreplaced by -2ImΣ r. Therefore, the phenomenology of two-point functions is entirely\ncovered by Gaussian settings .\n5.3 The double limit: g2T→ ∞followed by g2→0.\nTo perform a phenomenological analysis of dissipation, we need to un derstand how the theory\nbehaves intransitoryregime fromdissipative tofreepropagation. Similarly, to studyprimordial\nspectra in inflation or Hawking radiation, we also need to understand how free motion emerges\nas the proper frequency get red-shifted. It is therefore releva nt to study the behavior of the\ntwo-point function in the following double limit.\nOne first takes g2T→ ∞, whereT=t−tin,tinbeing the moment when the interactions\nare turn on, and tthe moment when one studies the field properties. This limit implies that\nthe decaying term in eq. (26) plays no role. Therefore, near time t, the Heisenberg field φ(t) is\na composite operator which only acts in the Hilbert space of Ψ.\nSecondly, one considers the ”free” limit g2→0 of that composite operator. One could\nnaively conclude that GcandGaof eqs. (80, 81) would vanish since both DandNare\n6This constitutes the simplest proof that it is inconsistent to couple q uantum variables to stochastic (or\nclassical) ones. If one does so, the ETC of the dressed quantum va riables will always be dissipated after a time\nof the order of γ−1. One can therefore view the experimental evidences for the ETC o f some degrees of freedom\nas a very strong indication that alldynamical variables in our world are quantum mechanical in nature.\n26proportional to g2. However, this is not the case, because the common prefactor, |Gr|2, is\nsingular in this second limit. In fact, one verifies that it scales in 1 /g2in such a way that, in\nthe (interacting) vacuum, one always recovers\nGW(ω)g2→0=1\n2ωp2πδ(ω−ωp). (84)\nThis is the standard vacuum fluctuations of a free massless mode of momentum p.\nTwo important lessons have been reached. First we learned is that e ven though φactsonly\non theΨ-Hilbert space, when g2→0, it behaves as if it were a free mode possessing its own\nHilbert space, with no reference to Ψ-dynamics. Secondly, the quantum state in this would be\nHilbert space is still exactly that of Ψ. Therefore, in stationary situations, the only ”souvenir”\nkept by the composite operator is the equilibrium distribution n(ω) inherited from its parents.\nLet us now emphasize that the above limit is relevant for non-Gaussia n models as well.\nIndeed, in the limit g2→0, there will always be a value of g2sufficiently small that the\nmodel can be well approximated by a Gaussian model. Therefore the behavior of the 2-point\nfunctions in the transitory regime from dissipation to free propaga tion can be analyzed by\nstudying Gaussian models (at least in the quasi-static limit).\n6 Appendix B:\nDissipative effects above ΛLV. The Phenomenology\nWe now have all the tools to construct models giving rise to dissipation in the vacuum above a\ncritical energy scale Λ LV. In this Appendix we work from a purely phenomenological point of\nview, and provide the class of dissipative models wherein the imaginary part of the self-energy\nis governed by a single term, in analogy with the dispersion relations of eq. (2).\nFromaphenomenologicalpointofview, ifoneconsidersonlystationa rysituations(i.e. static\nmetrics and stationary states), one can simply choose the functio nD(ω,p) entering eq. (76)\nand eq. (80) as one wishes . There is indeed no restriction on D(ω,p) besides its constitutive\nproperties, namely being odd in ωand giving rise to poles in Grall localized in the lower half ω\nplane. In this we have reached our first aim, namely identify how to ge neralize the free settings\nso as to incorporate some arbitrary dissipative effects.\nWe can thus consider the dispersive models which correspond to tho se defined by eq. (2).\nThey are characterized by a single term giving rise to dissipation abov e ΛLV. In the vacuum,\nthey are fully specified by the imaginary part of the (retarded) self -energy\n−ImΣ(n)\nr(ω,p) =ω\nΛLVp2/parenleftbiggp\nΛLV/parenrightbigg2n\n= 2ωγn. (85)\nIn these models, the decay rate (inverse life time) on the mass shell is\nγn=p\n2/parenleftbiggp\nΛLV/parenrightbigg2n+1\n. (86)\n27To verify it, assuming that ReΣ r= 0, the two poles of Gr(ω) in eq. (74) are located in\nω±(p) =±/radicalBig\nω2p−γ2−iγ. (87)\nFrom this, by inverse Fourier transform Gr(ω), one obtains that the decay rate is indeed γin\nthe under-damped regime, for ω2\np> γ2. In the overdamped regime, for γ2> ω2\np, the decay rates\nof the two independent solutions of G−1\nrφd= 0 are Γ ±=γ±/radicalbigγ2−ω2\np.\nOne thus have the following behavior as pgrows. For p≪ΛLV,ω±≃p, and one has\na free propagation which is slightly damped with a life time in the units of t he frequency\ngiven by (Λ LV/ω)n+1≫1. In the opposite regime of high momenta p≫ΛLV, deep in the\noverdamped regime, the two roots ω±are real and the notion of propagation (in space-time)\nis absent. In anticipation to what will occur in inflation (or in black hole p hysics), we invite\nthe reader to study the migration of the poles of Grwhen extrapolating backwards in time a\nmode, i.e. as pincreases. (Remember that the physical momentum of a mode in cos mology is\npphys(t)∝p/a(t) wherepis the norm of the conserved comoving wave vector (near a black ho le\none finds p(r)∝ω/xwherex=r−rSis the proper distance from the horizon measured in a\nfreely falling frame, and ωthe conserved Killing frequency.))\nOne could of course generalize the above class by considering in eq. ( 85) polynomials in\npdimensionalized by different UV scales. However, unless fine tuning, t he phenomenology of\nthe transition from the IR dissipation-free sector to the dissipativ e sector will be dominated a\nsingle term. One should also consider the possibility that ImΣ strictly v anishes below a certain\nfrequency Ω 1, as this would be the case when ever the spectrum of the Ψ fields pos sesses this\ngap, see the remarks after eq. (76).\nHaving the phenomenology of dissipative and unitary models under co ntrol (with dispersive\nand dissipative effects related by Kramers relation, eq. (77)) one c ould confront particle and\nastro-particle physics data and put lower bounds on Λ LVfor eachn, in analogy with what was\ndone for (pure) dispersion in [1].\nReferences\n[1] T. Jacobson, S. Liberati and D. Mattingly, Annals Phys. 321(2006) 150.\n[2] V. F. Mukhanov, H. A. Feldman and R. H. Brandenberger, Phys. Rept.215(1992) 203.\n[3] W. G. Unruh, Phys. Rev. D 51(1995) 2827.\n[4] T. Jacobson, Phys. Rev. D 48(1993) 728.\n[5] R. Brout, S. Massar, R. Parentani and P. Spindel, Phys. Rev. D 52(1995) 4559, and\nR. Balbinot, A. Fabbri, S. Fagnocchi and R. Parentani, Riv. Nuovo C im.28(2005) 1.\n[6] J. Martin and R. H. Brandenberger, Phys. Rev. D 63(2001) 123501.\n[7] J. C. Niemeyer, Phys. Rev. D 63(2001) 123502 .\n28[8] J. C. Niemeyer and R. Parentani, Phys. Rev. D 64(2001) 101301.\n[9] M. V. Libanov and V. A. Rubakov, JCAP 0509(2005) 005 .\n[10] M. V. Libanov and V. A. Rubakov, Phys. Rev. D 72(2005) 123503.\n[11] T. Jacobson, Phys. Rev. D 53(1996) 7082.\n[12] R. Parentani, “Constructing QFT’s wherein Lorentz Invarianc e is broken by dissipative\neffects in the UV,” arXiv:0709.3943 [hep-th].\n[13] D. Arteaga, R. Parentani and E. Verdaguer, Phys. Rev. D 70(2004) 044019, and Int. J.\nTheor. Phys. 46(2007) 2227.\n[14] R. Parentani, Int. J. Theor. Phys. 41(2002) 2175 [arXiv:0704.2563], and Phys. Rev. D 63\n(2001) 041503.\n[15] P. C. Aichelburg and R. Beig, Annals of Physics, 98(1976) 264.\n[16] D. J. Raine, D. W. Sciama and P. G. Grove, Proc. R. Soc. A 435(1991) 205.\n[17] S. Massar, R. Parentani and R. Brout, Class. Quant. Grav. 10(1993) 385.\n[18] C. Gardiner and P. Zoller, ”Quantum Noise,” Springer (2000).\n[19] W. G. Unruh and W. H. Zurek, Phys. Rev. D 40(1989) 1071.\n[20] D. Arteaga, arXiv:0711.0293 [quant-ph].\n[21] N.D.Birrell andP.C.W.Davies, “Quantumfieldsincurved space ,”CambridgeUniversity\nPress (1982).\n[22] R.M. Wald, “Quantumfield theoryincurved space-time andblack holethermodynamics,”\nChicago, USA: Univ. Pr. (1994).\n[23] B. Z. Foster and T. Jacobson, JHEP 0408(2004) 024.\n[24] A. Kempf, Phys. Rev. D 63(2001) 083514.\n[25] T. Jacobson, Phys. Rev. D 44(1991) 1731.\n[26] T. Jacobson and D. Mattingly, Phys. Rev. D 61(2000) 024017.\n[27] R. Parentani, C.R. Physique 4, 935 (2003), arXiv:astro-ph/04 04022.\n[28] W. G. Unruh, Phys. Rev. Lett. 46(1981) 1351.\n[29] U. H. Danielsson, Phys. Rev. D 66(2002) 023511.\n[30] J. C. Niemeyer, R. Parentani and D. Campo, Phys. Rev. D 66(2002) 083510.\n29[31] D. Campo, J. C. Niemeyer and R. Parentani, Phys. Rev. D 76(2007) 023513\n[32] D. S. Salopek and J. M. Stewart, Phys. Rev. D 47(1993) 3235.\n[33] R. Brout, S. Massar, R. Parentani, and Ph. Spindel, Phys. Rep .260(1995) 329.\n[34] T. Jacobson, Prog. Theor. Phys. Suppl. 136(1999) 1.\n[35] C. Barrab` es, V. Frolov, and R. Parentani, Phys. Rev. D 62(2000) 044020.\n[36] K. Schalm, G. Shiu and J. P. van der Schaar, JHEP 0404(2004) 076.\n[37] J. R. Anglin, “Thermal equilibrium from the Hu-Paz-Zhang maste r equation,”\narXiv:hep-th/9210034.\n[38] J. Adamek, D. Campo, J. Niemeyer, and R. Parentani, “Inflatio nary Spectra from Lorentz\nViolating Dissipative Models,” to appear .\n[39] J. Macher and R. Parentani, “Signatures of trans-Planckian d ispersion in inflationary\nspectra,” arXiv:0804.1920 [hep-th].\n[40] T. Jacobson and D. Mattingly, Phys. Rev. D 63(2001) 041502.\n[41] M. Lemoine, M. Lubo, J. Martin and J. P. Uzan, Phys. Rev. D 65(2002) 023510.\n[42] S. Shankaranarayanan and M. Lubo, Phys. Rev. D 72(2005) 123513.\n30" }, { "title": "2203.01632v1.Stability_results_of_locally_coupled_wave_equations_with_local_Kelvin_Voigt_damping__Cases_when_the_supports_of_damping_and_coupling_coefficients_are_disjoint.pdf", "content": "arXiv:2203.01632v1 [math.AP] 3 Mar 2022STABILITY RESULTS OF LOCALLY COUPLED WAVE EQUATIONS WITH LO CAL\nKELVIN-VOIGT DAMPING: CASES WHEN THE SUPPORTS OF DAMPING AN D\nCOUPLING COEFFICIENTS ARE DISJOINT\nMOHAMMAD AKIL1, HAIDAR BADAWI1, AND SERGE NICAISE1\nAbstract. In this paper, we study the direct/indirect stability of loc ally coupled wave equations with local\nKelvin-Voigt dampings/damping and by assuming that the sup ports of the dampings and the coupling coeffi-\ncients are disjoint. First, we prove the well-posedness, st rong stability, and polynomial stability for some one\ndimensional coupled systems. Moreover, under some geometr ic control condition, we prove the well-posedness\nand strong stability in the multi-dimensional case.\nContents\n1. Introduction 1\n2. Direct and Indirect Stability in the one dimensional case 4\n2.1. Well-Posedness 4\n2.2. Strong Stability 5\n2.3. Polynomial Stability 9\n2.3.1. Proof of Theorem 2.6 9\n2.3.2. Proof of Theorem 2.7 13\n3. Indirect Stability in the multi-dimensional case 16\n3.1. Well-posedness 16\n3.2. Strong Stability 17\nAppendix A. Some notions and stability theorems 20\nReferences 21\n1.Introduction\nThe direct and indirect stability of locally coupled wave equations with lo cal damping arouses many interests in\nrecent years. The study of coupled systems is also motivated by se veralphysicalconsiderationslike Timoshenko\nand Bresse systems (see for instance [ 10,6,3,2,1,15,14]). The exponential or polynomial stability of the wave\nequation with a local Kelvin-Voigt damping is considered in [ 20,23,13], for instance. On the other hand, the\ndirect and indirect stability of locally and coupled wave equations with lo cal viscous dampings are analyzed in\n[8,18,16]. In this paper, we are interested in locally coupled wave equations wit h local Kelvin-Voigt dampings.\nBefore stating our main contributions, let us mention similar results f or such systems. In 2019, Hayek et al.in\n[17], studied the stabilization of a multi-dimensional system of weakly cou pled wave equations with one or two\nlocally Kelvin-Voigt damping and non-smooth coefficient at the interfa ce. They established different stability\n1Universit ´e Polytechnique Hauts-de-France, CERAMATHS/DEMAV, Valencien nes, France\nE-mail address :Mohammad.Akil@uphf.fr, Haidar.Badawi@uphf.fr, Serge.N icaise@uphf.fr .\nKey words and phrases. Coupled wave equations, Kelvin-Voigt damping, strong stab ility, polynomial stability .\n1results. In 2021, Akil et al.in [24], studied the stability of an elastic/viscoelastic transmission problem of\nlocally coupled waves with non-smooth coefficients, by considering:\n\n\nutt−/parenleftbig\naux+b0χ(α1,α3)utx/parenrightbig\nx+c0χ(α2,α4)yt= 0,in (0,L)×(0,∞),\nytt−yxx−c0χ(α2,α4)ut= 0, in (0,L)×(0,∞),\nu(0,t) =u(L,t) =y(0,t) =y(L,t) = 0, in (0,∞),\nwherea,b0,L >0,c0/\\e}atio\\slash= 0, and 0 < α1< α2< α3< α4< L. They established a polynomial energy decay\nrate of type t−1. In the same year, Akil et al.in [5], studied the stability of a singular local interaction\nelastic/viscoelastic coupled wave equations with time delay, by consid ering:\n\n\nutt−/bracketleftbig\naux+χ(0,β)(κ1utx+κ2utx(t−τ))/bracketrightbig\nx+c0χ(α,γ)yt= 0,in (0,L)×(0,∞),\nytt−yxx−c0χ(α,γ)ut= 0, in (0,L)×(0,∞),\nu(0,t) =u(L,t) =y(0,t) =y(L,t) = 0, in (0,∞),\nwherea,κ1,L >0,κ2,c0/\\e}atio\\slash= 0, and 0 < α < β < γ < L . They proved that the energy of their system decays\npolynomially in t−1. In 2021, Akil et al.in [4], studied the stability of coupled wave models with locally\nmemory in a past history framework via non-smooth coefficients on t he interface, by considering:\n\n\nutt−/parenleftbigg\naux+b0χ(0,β)/integraldisplay∞\n0g(s)ux(t−s)ds/parenrightbigg\nx+c0χ(α,γ)yt= 0,in (0,L)×(0,∞),\nytt−yxx−c0χ(α,γ)ut= 0, in (0,L)×(0,∞),\nu(0,t) =u(L,t) =y(0,t) =y(L,t) = 0, in (0,∞),\nwherea,b0,L >0,c0/\\e}atio\\slash= 0, 0< α < β < γ < L , andg: [0,∞)/ma√sto−→(0,∞) is the convolution kernel function.\nThey established an exponential energy decay rate if the two wave s have the same speed of propagation. In\ncase of different speed of propagation, they proved that the ene rgy of their system decays polynomially with\nratet−1. In the same year, Akil et al.in [7], studied the stability of a multi-dimensional elastic/viscoelastic\ntransmission problem with Kelvin-Voigt damping and non-smooth coeffi cient at the interface, they established\nsome polynomial stability results under some geometric control con dition. In those previous literature, the\nauthors deal with the locally coupled wave equations with local dampin g and by assuming that there is an\nintersection between the damping and coupling regions. The aim of th is paper is to study the direct/indirect\nstability of locally coupled wave equations with Kelvin-Voigt dampings/d amping localized via non-smooth\ncoefficients/coefficient and by assuming that the supports of the d ampings and coupling coefficients aredisjoint.\nIn the first part of this paper, we consider the following one dimensio nal coupled system:\nutt−(aux+butx)x+cyt= 0,(x,t)∈(0,L)×(0,∞), (1.1)\nytt−(yx+dytx)x−cut= 0,(x,t)∈(0,L)×(0,∞), (1.2)\nwith fully Dirichlet boundary conditions,\n(1.3) u(0,t) =u(L,t) =y(0,t) =y(L,t) = 0, t∈(0,∞),\nand the following initial conditions\n(1.4) u(·,0) =u0(·), ut(·,0) =u1(·), y(·,0) =y0(·) and yt(·,0) =y1(·), x∈(0,L).\nIn this part, for all b0,d0>0 andc0/\\e}atio\\slash= 0, we treat the following three cases:\nCase 1 (See Figure 1):\n(C1)/braceleftiggb(x) =b0χ(b1,b2)(x), c(x) =c0χ(c1,c2)(x), d(x) =d0χ(d1,d2)(x),\nwhere 0< b1< b2< c1< c2< d1< d2< L.\nCase 2 (See Figure 2):\n(C2)/braceleftiggb(x) =b0χ(b1,b2)(x), c(x) =c0χ(c1,c2)(x), d(x) =d0χ(d1,d2)(x),\nwhere 0< b1< b2< d1< d2< c1< c2< L.\n2Case 3 (See Figure 3):\n(C3)/braceleftiggb(x) =b0χ(b1,b2)(x), c(x) =c0χ(c1,c2)(x), d(x) = 0,\nwhere 0< b1< b2< c1< c2< L.\nWhile in the second part, we consider the following multi-dimensional co upled system:\nb1b2c1c2d1d2Lb0c0\n0d0\nFigure 1. Geometric description of the functions b,canddin Case 1.\nb1b2 0 d1d2c1c2Lb0d0c0\nFigure 2. Geometric description of the functions b,canddin Case 2.\n0b1b2c1c2Lb0c0\nFigure 3. Geometric description of the functions bandcin Case 3.\nutt−div(∇u+but)+cyt= 0 in Ω ×(0,∞), (1.5)\nytt−∆y−cyt= 0 in Ω ×(0,∞), (1.6)\nwith full Dirichlet boundary condition\n(1.7) u=y= 0 on Γ ×(0,∞),\nand the following initial condition\n(1.8) u(·,0) =u0(·), ut(·,0) =u1(·), y(·,0) =y0(·) andyt(·,0) =y1(·) in Ω,\n3where Ω ⊂Rd,d≥2 is an open and bounded set with boundary Γ of class C2. Here,b,c∈L∞(Ω) are such\nthatb: Ω→R+is the viscoelastic damping coefficient, c: Ω→Ris the coupling function and\n(1.9) b(x)≥b0>0 inωb⊂Ω, c(x)≥c0/\\e}atio\\slash= 0 inωc⊂Ω and c(x) = 0 on Ω \\ωc\nand\n(1.10) meas( ωc∩Γ)>0 and ωb∩ωc=∅.\nIn the first part of this paper, we study the direct and indirect sta bility of system ( 1.1)-(1.4) by consider-\ning the three cases ( C1), (C2), and (C3). In Subsection 2.1, we prove the well-posedness of our system by using\na semigroup approach. In Subsection 2.2, by using a general criteria of Arendt-Batty, we prove the stron g\nstability of our system in the absence of the compactness of the re solvent. Finally, in Subsection 2.3, by using\na frequency domain approach combined with a specific multiplier metho d, we prove that our system decay\npolynomially in t−4or int−1.\nIn the second part of this paper, we study the indirect stability of s ystem (1.5)-(1.8). In Subsection 3.1,\nwe prove the well-posedness of our system by using a semigroup app roach. Finally, in Subsection 3.2, under\nsome geometric control condition, we prove the strong stability of this system.\n2.Direct and Indirect Stability in the one dimensional case\nIn this section, we study the well-posedness, strong stability, and polynomial stability of system ( 1.1)-(1.4).\nThe main result of this section are the following three subsections.\n2.1.Well-Posedness. In this subsection, we will establish the well-posedness of system ( 1.1)-(1.4) by using\nsemigroup approach. The energy of system ( 1.1)-(1.4) is given by\nE(t) =1\n2/integraldisplayL\n0/parenleftbig\n|ut|2+a|ux|2+|yt|2+|yx|2/parenrightbig\ndx.\nLet (u,ut,y,yt) be a regular solution of ( 1.1)-(1.4). Multiplying ( 1.1) and (1.2) byutandytrespectively, then\nusing the boundary conditions ( 1.3), we get\nE′(t) =−/integraldisplayL\n0/parenleftbig\nb|utx|2+d|ytx|2/parenrightbig\ndx.\nThus, if ( C1) or (C2) or (C3) holds, we get E′(t)≤0. Therefore, system ( 1.1)-(1.4) is dissipative in the sense\nthat its energy is non-increasing with respect to time t. Let us define the energy space Hby\nH= (H1\n0(0,L)×L2(0,L))2.\nThe energy space His equipped with the following inner product\n(U,U1)H=/integraldisplayL\n0vv1dx+a/integraldisplayL\n0ux(u1)xdx+/integraldisplayL\n0zz1dx+/integraldisplayL\n0yx(y1)xdx,\nfor allU= (u,v,y,z)⊤andU1= (u1,v1,y1,z1)⊤inH. We define the unbounded linear operator A:D(A)⊂\nH −→ H by\nD(A) =/braceleftbig\nU= (u,v,y,z)⊤∈ H;v,z∈H1\n0(0,L),(aux+bvx)x∈L2(0,L),(yx+dzx)x∈L2(0,L)/bracerightbig\nand\nA(u,v,y,z)⊤= (v,(aux+bvx)x−cz,z,(yx+dzx)x+cv)⊤,∀U= (u,v,y,z)⊤∈D(A).\nNow, ifU= (u,ut,y,yt)⊤is the state of system ( 1.1)-(1.4), then it is transformed into the following first order\nevolution equation\n(2.1) Ut=AU, U(0) =U0,\nwhereU0= (u0,u1,y0,y1)⊤∈ H.\n4Proposition 2.1. If (C1) or (C2) or (C3) holds. Then, the unbounded linear operator Ais m-dissipative in\nthe Hilbert space H.\nProof.For allU= (u,v,y,z)⊤∈D(A), we have\nℜ/a\\}b∇acketle{tAU,U/a\\}b∇acket∇i}htH=−/integraldisplayL\n0b|vx|2dx−/integraldisplayL\n0d|zx|2dx≤0,\nwhich implies that Ais dissipative. Now, similiar to Proposition 2.1 in [ 24] (see also [ 5] and [4]), we can prove\nthat there exists a unique solution U= (u,v,y,z)⊤∈D(A) of\n−AU=F,∀F= (f1,f2,f3,f4)⊤∈ H.\nThen 0∈ρ(A) andAis an isomorphism and since ρ(A) is open in C(see Theorem 6.7 (Chapter III) in [ 19]),\nwe easily get R(λI−A) =Hfor a sufficiently small λ >0. This, together with the dissipativeness of A, imply\nthatD(A) is dense in Hand that Ais m-dissipative in H(see Theorems 4.5, 4.6 in [ 22]). /square\nAccording to Lumer-Phillips theorem (see [ 22]), then the operator Agenerates a C0-semigroup of contrac-\ntionsetAinHwhich gives the well-posedness of ( 2.1). Then, we have the following result:\nTheorem 2.2. For allU0∈ H, system ( 2.1) admits a unique weak solution\nU(t) =etAU0∈C0(R+,H).\nMoreover, if U0∈D(A), then the system ( 2.1) admits a unique strong solution\nU(t) =etAU0∈C0(R+,D(A))∩C1(R+,H).\n2.2.Strong Stability. In this subsection, we will prove the strong stability of system ( 1.1)-(1.4). We define\nthe following conditions:\n(SSC1) ( C1) holds and |c0|0 independent\nofU0such that\n(2.44) E(t)≤C\nt4/ba∇dblU0/ba∇dbl2\nD(A), t >0.\nTheorem 2.7. Assume that ( SSC3) holds . Then, for all U0∈D(A) there exists a constant C >0 independent\nofU0such that\n(2.45) E(t)≤C\nt/ba∇dblU0/ba∇dbl2\nD(A), t >0.\nAccording to Theorem A.3, the polynomial energy decays ( 2.44) and (2.45) hold if the following conditions\n(H1) iR⊂ρ(A)\nand\n(H2) limsup\nλ∈R,|λ|→∞1\n|λ|ℓ/vextenddouble/vextenddouble(iλI−A)−1/vextenddouble/vextenddouble\nL(H)<∞withℓ=/braceleftigg1\n2for Theorem 2.6,\n2 for Theorem 2.7,\naresatisfied. Sincecondition( H1)isalreadyprovedin Subsection 2.2. We stillneedtoprove( H2), let usproveit\nbyacontradictionargument. Tothisaim,supposethat( H2)isfalse,thenthereexists/braceleftbig/parenleftbig\nλn,Un:= (un,vn,yn,zn)⊤/parenrightbig/bracerightbig\nn≥1⊂\nR∗\n+×D(A) with\n(2.46) λn→ ∞asn→ ∞and/ba∇dblUn/ba∇dblH= 1,∀n≥1,\nsuch that\n(2.47) ( λn)ℓ(iλnI−A)Un=Fn:= (f1,n,f2,n,f3,n,f4,n)⊤→0 inH,asn→ ∞.\nFor simplicity, we drop the index n. Equivalently, from ( 2.47), we have\niλu−v=f1\nλℓ, f1→0 inH1\n0(0,L), (2.48)\niλv−(aux+bvx)x+cz=f2\nλℓ, f2→0 inL2(0,L), (2.49)\niλy−z=f3\nλℓ, f3→0 inH1\n0(0,L), (2.50)\niλz−(yx+dzx)x−cv=f4\nλℓ, f4→0 inL2(0,L). (2.51)\n2.3.1.Proof of Theorem 2.6.In this subsection, we will prove Theorem 2.6by checking the condition ( H2),\nby finding a contradiction with ( 2.46) by showing /ba∇dblU/ba∇dblH=o(1). For clarity, we divide the proof into several\nLemmas. By taking the inner product of ( 2.47) withUinH, we remark that\n/integraldisplayL\n0b|vx|2dx+/integraldisplayL\n0d|zx|2dx=−ℜ(/a\\}b∇acketle{tAU,U/a\\}b∇acket∇i}htH) =λ−1\n2ℜ(/a\\}b∇acketle{tF,U/a\\}b∇acket∇i}htH) =o/parenleftig\nλ−1\n2/parenrightig\n.\nThus, from the definitions of bandd, we get\n(2.52)/integraldisplayb2\nb1|vx|2dx=o/parenleftig\nλ−1\n2/parenrightig\nand/integraldisplayd2\nd1|zx|2dx=o/parenleftig\nλ−1\n2/parenrightig\n.\nUsing (2.48), (2.50), (2.52), and the fact that f1,f3→0 inH1\n0(0,L), we get\n(2.53)/integraldisplayb2\nb1|ux|2dx=o(1)\nλ5\n2and/integraldisplayd2\nd1|yx|2dx=o(1)\nλ5\n2.\nLemma 2.8. The solution U∈D(A) of system ( 2.48)-(2.51) satisfies the following estimations\n(2.54)/integraldisplayb2\nb1|v|2dx=o(1)\nλ3\n2and/integraldisplayd2\nd1|z|2dx=o(1)\nλ3\n2.\n9Proof.We give the proof of the first estimation in ( 2.54), the second one can be done in a similar way. For\nthis aim, we fix g∈C1([b1,b2]) such that\ng(b2) =−g(b1) = 1,max\nx∈[b1,b2]|g(x)|=mgand max\nx∈[b1,b2]|g′(x)|=mg′.\nThe proof is divided into several steps:\nStep 1. The goal of this step is to prove that\n(2.55) |v(b1)|2+|v(b2)|2≤/parenleftigg\nλ1\n2\n2+2mg′/parenrightigg/integraldisplayb2\nb1|v|2dx+o(1)\nλ.\nFrom (2.48), we deduce that\n(2.56) vx=iλux−λ−1\n2(f1)x.\nMultiplying ( 2.56) by 2gvand integrating over ( b1,b2), then taking the real part, we get\n/integraldisplayb2\nb1g/parenleftbig\n|v|2/parenrightbig\nxdx=ℜ/parenleftigg\n2iλ/integraldisplayb2\nb1guxvdx/parenrightigg\n−ℜ/parenleftigg\n2λ−1\n2/integraldisplayb2\nb1g(f1)xvdx/parenrightigg\n.\nUsing integration by parts in the left hand side of the above equation , we get\n(2.57) |v(b1)|2+|v(b2)|2=/integraldisplayb2\nb1g′|v|2dx+ℜ/parenleftigg\n2iλ/integraldisplayb2\nb1guxvdx/parenrightigg\n−ℜ/parenleftigg\n2λ−1\n2/integraldisplayb2\nb1g(f1)xvdx/parenrightigg\n.\nUsing Young’s inequality, we obtain\n2λmg|ux||v| ≤λ1\n2|v|2\n2+2λ3\n2m2\ng|ux|2and 2λ−1\n2mg|(f1)x||v| ≤mg′|v|2+m2\ngm−1\ng′λ−1|(f1)x|2.\nFrom the above inequalities, ( 2.57) becomes\n(2.58) |v(b1)|2+|v(b2)|2≤/parenleftigg\nλ1\n2\n2+2mg′/parenrightigg/integraldisplayb2\nb1|v|2dx+2λ3\n2m2\ng/integraldisplayb2\nb1|ux|2dx+m2\ng\nmg′λ−1/integraldisplayb2\nb1|(f1)x|2dx.\nInserting ( 2.53) in (2.58) and the fact that f1→0 inH1\n0(0,L), we get ( 2.55).\nStep 2. The aim of this step is to prove that\n(2.59) |(aux+bvx)(b1)|2+|(aux+bvx)(b2)|2≤λ3\n2\n2/integraldisplayb2\nb1|v|2dx+o(1).\nMultiplying ( 2.49) by−2g/parenleftbig\naux+bvx/parenrightbig\n, using integration by parts over ( b1,b2) and taking the real part, we get\n|(aux+bvx)(b1)|2+|(aux+bvx)(b2)|2=/integraldisplayb2\nb1g′|aux+bvx|2dx+\nℜ/parenleftigg\n2iλ/integraldisplayb2\nb1gv(aux+bvx)dx/parenrightigg\n−ℜ/parenleftigg\n2λ−1\n2/integraldisplayb2\nb1gf2(aux+bvx)dx/parenrightigg\n,\nconsequently, we get\n(2.60)|(aux+bvx)(b1)|2+|(aux+bvx)(b2)|2≤mg′/integraldisplayb2\nb1|aux+bvx|2dx\n+2λmg/integraldisplayb2\nb1|v||aux+bvx|dx+2mgλ−1\n2/integraldisplayb2\nb1|f2||aux+bvx|dx.\nBy Young’s inequality, ( 2.52), and (2.53), we have\n(2.61) 2 λmg/integraldisplayb2\nb1|v||aux+bvx|dx≤λ3\n2\n2/integraldisplayb2\nb1|v|2dx+2m2\ngλ1\n2/integraldisplayb2\nb1|aux+bvx|2dx≤λ3\n2\n2/integraldisplayb2\nb1|v|2dx+o(1).\nInserting ( 2.61) in (2.60), then using ( 2.52), (2.53) and the fact that f2→0 inL2(0,L), we get ( 2.59).\n10Step 3. The aim of this step is to prove the first estimation in ( 2.54). For this aim, multiplying ( 2.49) by\n−iλ−1v, integrating over ( b1,b2) and taking the real part , we get\n(2.62)/integraldisplayb2\nb1|v|2dx=ℜ/parenleftigg\niλ−1/integraldisplayb2\nb1(aux+bvx)vxdx−/bracketleftbig\niλ−1(aux+bvx)v/bracketrightbigb2\nb1+iλ−3\n2/integraldisplayb2\nb1f2vdx/parenrightigg\n.\nUsing (2.52), (2.53), the fact that vis uniformly bounded in L2(0,L) andf2→0 inL2(0,1), and Young’s\ninequalities, we get\n(2.63)/integraldisplayb2\nb1|v|2dx≤λ−1\n2\n2[|v(b1)|2+|v(b2)|2]+λ−3\n2\n2[|(aux+bvx)(b1)|2+|(aux+bvx)(b2)|2]+o(1)\nλ3\n2.\nInserting ( 2.55) and (2.59) in (2.63), we get\n/integraldisplayb2\nb1|v|2dx≤/parenleftbigg1\n2+mg′λ−1\n2/parenrightbigg/integraldisplayb2\nb1|v|2dx+o(1)\nλ3\n2,\nwhich implies that\n(2.64)/parenleftbigg1\n2−mg′λ−1\n2/parenrightbigg/integraldisplayb2\nb1|v|2dx≤o(1)\nλ3\n2.\nUsing the fact that λ→ ∞, we can take λ >4m2\ng′. Then, we obtain the first estimation in ( 2.54). Similarly,\nwe can obtain the second estimation in ( 2.54). The proof has been completed. /square\nLemma 2.9. The solution U∈D(A) of system ( 2.48)-(2.51) satisfies the following estimations\n(2.65)/integraldisplayc1\n0/parenleftbig\n|v|2+a|ux|2/parenrightbig\ndx=o(1) and/integraldisplayL\nc2/parenleftbig\n|z|2+|yx|2/parenrightbig\ndx=o(1).\nProof. First, let h∈C1([0,c1]) such that h(0) =h(c1) = 0. Multiplying ( 2.49) by 2a−1h(aux+bvx),\nintegrating over (0 ,c1), using integration by parts and taking the real part, then using ( 2.52) and the fact that\nuxis uniformly bounded in L2(0,L) andf2→0 inL2(0,L), we get\n(2.66) ℜ/parenleftbigg\n2iλa−1/integraldisplayc1\n0vh(aux+bvx)dx/parenrightbigg\n+a−1/integraldisplayc1\n0h′|aux+bvx|2dx=o(1)\nλ1\n2.\nFrom (2.48), we have\n(2.67) iλux=−vx−λ−1\n2(f1)x.\nInserting ( 2.67) in (2.66), using integration by parts, then using ( 2.52), (2.54), and the fact that f1→0 in\nH1\n0(0,L) andvis uniformly bounded in L2(0,L), we get\n(2.68)/integraldisplayc1\n0h′|v|2dx+a−1/integraldisplayc1\n0h′|aux+bvx|2dx= 2ℜ/parenleftbigg\nλ−1\n2/integraldisplayc1\n0vh(f1)xdx/parenrightbigg\n/bracehtipupleft/bracehtipdownright/bracehtipdownleft/bracehtipupright\n=o(λ−1\n2)\n+ℜ/parenleftigg\n2iλa−1b0/integraldisplayb2\nb1hvvxdx/parenrightigg\n/bracehtipupleft/bracehtipdownright/bracehtipdownleft/bracehtipupright\n=o(1)+o(1)\nλ1\n2.\nNow, we fix the following cut-off functions\np1(x) :=\n\n1 in (0 ,b1),\n0 in ( b2,c1),\n0≤p1≤1 in (b1,b2),andp2(x) :=\n\n1 in ( b2,c1),\n0 in (0 ,b1),\n0≤p2≤1 in (b1,b2).\nFinally, take h(x) =xp1(x)+(x−c1)p2(x) in (2.68) and using ( 2.52), (2.53), (2.54), we get the first estimation\nin (2.65). By using the same argument, we can obtain the second estimation in (2.65). The proof is thus\ncompleted. /square\nLemma 2.10. The solution U∈D(A)of system (2.48)-(2.51)satisfies the following estimations\n(2.69) |λu(c1)|=o(1),|ux(c1)|=o(1),|λy(c2)|=o(1)and|yx(c2)|=o(1).\n11Proof.First, from ( 2.48) and (2.49), we deduce that\n(2.70) λ2u+auxx=−f2\nλ1\n2−iλ1\n2f1in (b2,c1).\nMultiplying ( 2.70) by 2(x−b2)¯ux, integrating over ( b2,c1) and taking the real part, then using the fact that\nuxis uniformly bounded in L2(0,L) andf2→0 inL2(0,L), we get\n(2.71)/integraldisplayc1\nb2λ2(x−b2)/parenleftbig\n|u|2/parenrightbig\nxdx+a/integraldisplayc1\nb2(x−b2)/parenleftbig\n|ux|2/parenrightbig\nxdx=−ℜ/parenleftbigg\n2iλ1\n2/integraldisplayc1\nb2(x−b2)f1uxdx/parenrightbigg\n+o(1)\nλ1\n2.\nUsing integration by parts in ( 2.71), then using ( 2.65), and the fact that f1→0 inH1\n0(0,L) andλuis uniformly\nbounded in L2(0,L), we get\n(2.72) 0 ≤(c1−b2)/parenleftbig\n|λu(c1)|2+a|ux(c1)|2/parenrightbig\n=ℜ/parenleftig\n2iλ1\n2(c1−b2)f1(c1)u(c1)/parenrightig\n+o(1),\nconsequently, by using Young’s inequality, we get\n|λu(c1)|2+|ux(c1)|2≤2λ1\n2|f1(c1)||u(c1)|+o(1)\n≤1\n2|λu(c1)|2+2\nλ|f1(c1)|2+o(1).\nThen, we get\n(2.73)1\n2|λu(c1)|2+|ux(c1)|2≤2\nλ|f1(c1)|2+o(1).\nFinally, from the above estimation and the fact that f1→0 inH1\n0(0,L), we get the first two estimations in\n(2.69). By using the same argument, we can obtain the last two estimation s in (2.69). The proof has been\ncompleted. /square\nLemma 2.11. The solution U∈D(A)of system (2.48)-(2.51)satisfies the following estimation\n(2.74)/integraldisplayc2\nc1|λu|2+a|ux|2+|λy|2+|yx|2dx=o(1).\nProof.Inserting ( 2.48) and (2.50) in (2.49) and (2.51), we get\n−λ2u−auxx+iλc0y=f2\nλ1\n2+iλ1\n2f1+c0f3\nλ1\n2in (c1,c2), (2.75)\n−λ2y−yxx−iλc0u=f4\nλ1\n2+iλ1\n2f3−c0f1\nλ1\n2in (c1,c2). (2.76)\nMultiplying ( 2.75) by 2(x−c2)uxand (2.76) by 2(x−c1)yx, integrating over ( c1,c2) and taking the real part,\nthen using the fact that /ba∇dblF/ba∇dblH=o(1) and/ba∇dblU/ba∇dblH= 1, we obtain\n(2.77)−λ2/integraldisplayc2\nc1(x−c2)/parenleftbig\n|u|2/parenrightbig\nxdx−a/integraldisplayc2\nc1(x−c2)/parenleftbig\n|ux|2/parenrightbig\nxdx+ℜ/parenleftbigg\n2iλc0/integraldisplayc2\nc1(x−c2)yuxdx/parenrightbigg\n=\nℜ/parenleftbigg\n2iλ1\n2/integraldisplayc2\nc1(x−c2)f1uxdx/parenrightbigg\n+o(1)\nλ1\n2\nand\n(2.78)−λ2/integraldisplayc2\nc1(x−c1)/parenleftbig\n|y|2/parenrightbig\nxdx−/integraldisplayc2\nc1(x−c1)/parenleftbig\n|yx|2/parenrightbig\nxdx−ℜ/parenleftbigg\n2iλc0/integraldisplayc2\nc1(x−c1)uyxdx/parenrightbigg\n=\nℜ/parenleftbigg\n2iλ1\n2/integraldisplayc2\nc1(x−c1)f3yxdx/parenrightbigg\n+o(1)\nλ1\n2.\nUsing integration by parts, ( 2.69), and the fact that f1,f3→0 inH1\n0(0,L),/ba∇dblu/ba∇dblL2(0,L)=O(λ−1),/ba∇dbly/ba∇dblL2(0,L)=\nO(λ−1), we deduce that\n(2.79) ℜ/parenleftbigg\niλ1\n2/integraldisplayc2\nc1(x−c2)f1uxdx/parenrightbigg\n=o(1)\nλ1\n2andℜ/parenleftbigg\niλ1\n2/integraldisplayc2\nc1(x−c1)f3yxdx/parenrightbigg\n=o(1)\nλ1\n2.\n12Inserting ( 2.79) in (2.77) and (2.78), then using integration by parts and ( 2.69), we get\n/integraldisplayc2\nc1/parenleftbig\n|λu|2+a|ux|2/parenrightbig\ndx+ℜ/parenleftbigg\niλc0/integraldisplayc2\nc1(x−c2)yuxdx/parenrightbigg\n=o(1), (2.80)\n/integraldisplayc2\nc1/parenleftbig\n|λy|2+|yx|2/parenrightbig\ndx−ℜ/parenleftbigg\niλc0/integraldisplayc2\nc1(x−c1)uyxdx/parenrightbigg\n=o(1). (2.81)\nAdding ( 2.80) and (2.81), we get\n/integraldisplayc2\nc1/parenleftbig\n|λu|2+a|ux|2+|λy|2+|yx|2/parenrightbig\ndx=ℜ/parenleftbigg\n2iλc0/integraldisplayc2\nc1(x−c1)uyxdx/parenrightbigg\n−ℜ/parenleftbigg\n2iλc0/integraldisplayc2\nc1(x−c2)yuxdx/parenrightbigg\n+o(1)\n≤2λ|c0|(c2−c1)/integraldisplayc2\nc1|u||yx|dx+2λ|c0|\na1\n4(c2−c1)a1\n4/integraldisplayc2\nc1|y||ux|dx+o(1).\nApplying Young’s inequalities, we get\n(2.82) (1 −|c0|(c2−c1))/integraldisplayc2\nc1(|λu|2+|yx|2)dx+/parenleftbigg\n1−1√a|c0|(c2−c1)/parenrightbigg/integraldisplayc2\nc1(a|ux|2+|λy|2)dx≤o(1).\nFinally, using ( SSC1), we get the desired result. The proof has been completed. /square\nLemma 2.12. The solution U∈D(A)of system (2.48)-(2.51)satisfies the following estimations\n(2.83)/integraldisplayc1\n0/parenleftbig\n|z|2+|yx|2/parenrightbig\ndx=o(1)and/integraldisplayL\nc2/parenleftbig\n|v|2+a|ux|2/parenrightbig\ndx=o(1).\nProof.Using the same argument of Lemma 2.9, we obtain ( 2.83). /square\nProof of Theorem 2.6.Using (2.53), Lemmas 2.8,2.9,2.11,2.12, we get /ba∇dblU/ba∇dblH=o(1), which contradicts\n(2.46). Consequently, condition (H2) holds. This implies the energy decay estimation ( 2.44).\n2.3.2.Proof of Theorem 2.7.In this subsection, we will prove Theorem 2.7by checking the condition ( H2),\nthat is by finding a contradiction with ( 2.46) by showing /ba∇dblU/ba∇dblH=o(1). For clarity, we divide the proof into\nseveral Lemmas. By taking the inner product of ( 2.47) withUinH, we remark that\n/integraldisplayL\n0b|vx|2dx=−ℜ(/a\\}b∇acketle{tAU,U/a\\}b∇acket∇i}htH) =λ−2ℜ(/a\\}b∇acketle{tF,U/a\\}b∇acket∇i}htH) =o(λ−2).\nThen,\n(2.84)/integraldisplayb2\nb1|vx|2dx=o(λ−2).\nUsing (2.48) and (2.84), and the fact that f1→0 inH1\n0(0,L), we get\n(2.85)/integraldisplayb2\nb1|ux|2dx=o(λ−4).\nLemma 2.13. Let0< ε 0. This, together with the dissipativeness of Ad,\nimply that D(Ad) is dense in Hand that Adis m-dissipative in H(see Theorems 4.5, 4.6 in [ 22]). According\nto Lumer-Phillips theorem (see [ 22]), then the operator Adgenerates a C0-semigroup of contractions etAdin\nHwhich gives the well-posedness of ( 3.3). Then, we have the following result:\nTheorem 3.1. For allU0∈ H, system ( 2.1) admits a unique weak solution\nU(t) =etAdU0∈C0(R+,H).\nMoreover, if U0∈D(A), then the system ( 2.1) admits a unique strong solution\nU(t) =etAdU0∈C0(R+,D(Ad))∩C1(R+,H).\n3.2.Strong Stability. In this subsection, we will prove the strong stability of system ( 1.5)-(1.8). First, we\nfix the following notations\n/tildewideΩ = Ω−ωc,Γ1=∂ωc−∂Ω and Γ 0=∂ωc−Γ1.\nLetx0∈Rdandm(x) =x−x0and suppose that (see Figure 4)\n(GC) m·ν≤0 on Γ 0= (∂ωc)−Γ1.\n17The main result of this section is the following theorem\nTheorem 3.2. Assume that (GC)holds and\n(SSC) /ba∇dblc/ba∇dbl∞≤min/braceleftigg\n1\n/ba∇dblm/ba∇dbl∞+d−1\n2,1\n/ba∇dblm/ba∇dbl∞+(d−1)Cp,ωc\n2/bracerightigg\n,\nwhereCp,ωcis the Poincarr´ e constant on ωc. Then, the C0−semigroup of contractions/parenleftbig\netAd/parenrightbig\nis strongly stable\ninH; i.e. for all U0∈ H, the solution of (3.3)satisfies\nlim\nt→+∞/ba∇dbletAdU0/ba∇dblH= 0.\nProof.First, let us prove that\n(3.4) ker( iλI−Ad) ={0},∀λ∈R.\nSince 0∈ρ(Ad), then we still need to show the result for λ∈R∗. Suppose that there exists a real number\nλ/\\e}atio\\slash= 0 and U= (u,v,y,z)⊤∈D(Ad), such that\nAdU=iλU.\nEquivalently, we have\nv=iλu, (3.5)\ndiv(∇u+b∇v)−cz=iλv, (3.6)\nz=iλy, (3.7)\n∆y+cv=iλz. (3.8)\nNext, a straightforward computation gives\n0 =ℜ/a\\}b∇acketle{tiλU,U/a\\}b∇acket∇i}htH=ℜ/a\\}b∇acketle{tAdU,U/a\\}b∇acket∇i}htH=−/integraldisplay\nΩb|∇v|2dx,\nconsequently, we deduce that\n(3.9) b∇v= 0 in Ω and ∇v=∇u= 0 in ωb.\nInserting ( 3.5) in (3.6), then using the definition of c, we get\n(3.10) ∆u=−λ2uinωb.\nFrom (3.9) we get ∆ u= 0 inωband from ( 3.10) and the fact that λ/\\e}atio\\slash= 0, we get\n(3.11) u= 0 in ωb.\nNow, inserting ( 3.5) in (3.6), then using ( 3.9), (3.11) and the definition of c, we get\n(3.12)λ2u+∆u= 0 in/tildewideΩ,\nu= 0 in ωb⊂/tildewideΩ.\nUsing Holmgren uniqueness theorem, we get\n(3.13) u= 0 in/tildewideΩ.\nIt follows that\n(3.14) u=∂u\n∂ν= 0 on Γ 1.\nNow, our aim is to show that u=y= 0 inωc. For this aim, inserting ( 3.5) and (3.7) in (3.6) and (3.8), then\nusing (3.9), we get the following system\nλ2u+∆u−iλcy= 0 in Ω , (3.15)\nλ2y+∆y+iλcu= 0 in Ω , (3.16)\nu= 0 on ∂ωc, (3.17)\ny= 0 on Γ 0, (3.18)\n∂u\n∂ν= 0 on Γ 1. (3.19)\n18Let us prove ( 3.4) by the following three steps:\nStep 1. The aim of this step is to show that\n(3.20)/integraldisplay\nΩc|u|2dx=/integraldisplay\nΩc|y|2dx.\nFor this aim, multiplying ( 3.15) and (3.16) by ¯yand ¯urespectively, integrating over Ω and using Green’s\nformula, we get\nλ2/integraldisplay\nΩu¯ydx−/integraldisplay\nΩ∇u·∇¯ydx−iλ/integraldisplay\nΩc|y|2dx= 0, (3.21)\nλ2/integraldisplay\nΩy¯udx−/integraldisplay\nΩ∇y·∇¯udx+iλ/integraldisplay\nΩc|u|2dx= 0. (3.22)\nAdding ( 3.21) and (3.22), then taking the imaginary part, we get ( 3.20).\nStep 2. The aim of this step is to prove the following identity\n(3.23) −d/integraldisplay\nωc|λu|2dx+(d−2)/integraldisplay\nωc|∇u|2dx+/integraldisplay\nΓ0(m·ν)/vextendsingle/vextendsingle/vextendsingle/vextendsingle∂u\n∂ν/vextendsingle/vextendsingle/vextendsingle/vextendsingle2\ndΓ−2ℜ/parenleftbigg\niλ/integraldisplay\nωccy(m·∇¯u)dx/parenrightbigg\n= 0.\nFor this aim, multiplying ( 3.15) by 2(m·∇¯u), integrating over ωcand taking the real part, we get\n(3.24) 2 ℜ/parenleftbigg\nλ2/integraldisplay\nωcu(m·∇¯u)dx/parenrightbigg\n+2ℜ/parenleftbigg/integraldisplay\nωc∆u(m·∇¯u)dx/parenrightbigg\n−2ℜ/parenleftbigg\niλ/integraldisplay\nωccy(m·∇¯u)dx/parenrightbigg\n= 0.\nNow, using the fact that u= 0 in∂ωc, we get\n(3.25) ℜ/parenleftbigg\n2λ2/integraldisplay\nωcu(m·∇¯u)dx/parenrightbigg\n=−d/integraldisplay\nωc|λu|2dx.\nUsing Green’s formula, we obtain\n(3.26)2ℜ/parenleftbigg/integraldisplay\nωc∆u(m·∇¯u)dx/parenrightbigg\n=−2ℜ/parenleftbigg/integraldisplay\nωc∇u·∇(m·∇¯u)dx/parenrightbigg\n+2ℜ/parenleftbigg/integraldisplay\nΓ0∂u\n∂ν(m·∇¯u)dΓ/parenrightbigg\n= (d−2)/integraldisplay\nωc|∇u|2dx−/integraldisplay\n∂ωc(m·ν)|∇u|2dx+2ℜ/parenleftbigg/integraldisplay\nΓ0∂u\n∂ν(m·∇¯u)dΓ/parenrightbigg\n.\nUsing (3.17) and (3.19), we get\n(3.27)/integraldisplay\n∂ωc(m·ν)|∇u|2dx=/integraldisplay\nΓ0(m·ν)/vextendsingle/vextendsingle/vextendsingle/vextendsingle∂u\n∂ν/vextendsingle/vextendsingle/vextendsingle/vextendsingle2\ndΓ andℜ/parenleftbigg/integraldisplay\nΓ0∂u\n∂ν(m·∇¯u)dΓ/parenrightbigg\n=/integraldisplay\nΓ0(m·ν)/vextendsingle/vextendsingle/vextendsingle/vextendsingle∂u\n∂ν/vextendsingle/vextendsingle/vextendsingle/vextendsingle2\ndΓ.\nInserting ( 3.27) in (3.26), we get\n(3.28) 2 ℜ/parenleftbigg/integraldisplay\nωc∆u(m·∇¯u)dx/parenrightbigg\n= (d−2)/integraldisplay\nωc|∇u|2dx+/integraldisplay\nΓ0(m·ν)/vextendsingle/vextendsingle/vextendsingle/vextendsingle∂u\n∂ν/vextendsingle/vextendsingle/vextendsingle/vextendsingle2\ndΓ.\nInserting ( 3.25) and (3.28) in (3.24), we get ( 3.23).\nStep 3. In this step, we prove ( 3.4). Multiplying ( 3.15) by (d−1)u, integrating over ωcand using ( 3.17), we\nget\n(3.29) ( d−1)/integraldisplay\nωc|λu|2dx+(1−d)/integraldisplay\nωc|∇u|2dx−ℜ/parenleftbigg\niλ(d−1)/integraldisplay\nωccy¯udx/parenrightbigg\n= 0.\nAdding ( 3.23) and (3.29), we get\n/integraldisplay\nωc|λu|2dx+/integraldisplay\nωc|∇u|2dx=/integraldisplay\nΓ0(m·ν)/vextendsingle/vextendsingle/vextendsingle/vextendsingle∂u\n∂ν/vextendsingle/vextendsingle/vextendsingle/vextendsingle2\ndΓ−2ℜ/parenleftbigg\niλ/integraldisplay\nωccy(m·∇¯u)dx/parenrightbigg\n−ℜ/parenleftbigg\niλ(d−1)/integraldisplay\nωccy¯udx/parenrightbigg\n= 0.\nUsing (GC), we get\n(3.30)/integraldisplay\nωc|λu|2dx+/integraldisplay\nωc|∇u|2dx≤2|λ|/integraldisplay\nωc|c||y||m·∇u|dx+|λ|(d−1)/integraldisplay\nωc|c||y||u|dx.\n19Using Young’s inequality and ( 3.20), we get\n(3.31) 2 |λ|/integraldisplay\nωc|c||y||m·∇u|dx≤ /ba∇dblm/ba∇dbl∞/ba∇dblc/ba∇dbl∞/integraldisplay\nωc/parenleftbig\n|λu|2+|∇u|2/parenrightbig\ndx\nand\n(3.32) |λ|(d−1)/integraldisplay\nωc|c(x)||y||u|dx≤(d−1)/ba∇dblc/ba∇dbl∞\n2/integraldisplay\nωc|λu|2dx+(d−1)/ba∇dblc/ba∇dbl∞Cp,ωc\n2/integraldisplay\nωc|∇u|2dx.\nInserting ( 3.32) in (3.30), we get\n/parenleftbigg\n1−/ba∇dblc/ba∇dbl∞/parenleftbigg\n/ba∇dblm/ba∇dbl∞+d−1\n2/parenrightbigg/parenrightbigg/integraldisplay\nωc|λu|2dx+/parenleftbigg\n1−/ba∇dblc/ba∇dbl∞/parenleftbigg\n/ba∇dblm/ba∇dbl∞+(d−1)Cp,ωc\n2/parenrightbigg/parenrightbigg/integraldisplay\nωc|∇u|2dx≤0.\nUsing (SSC) and (3.20) in the above estimation, we get\n(3.33) u= 0 and y= 0 in ωc.\nIn order to complete this proof, we need to show that y= 0 in/tildewideΩ. For this aim, using the definition of the\nfunction cin/tildewideΩ and using the fact that y= 0 inωc, we get\n(3.34)λ2y+∆y= 0 in/tildewideΩ,\ny= 0 on ∂/tildewideΩ,\n∂y\n∂ν= 0 on Γ 1.\nNow, using Holmgren uniqueness theorem, we obtain y= 0 in/tildewideΩ and consequently ( 3.4) holds true. Moreover,\nsimilar to Lemma 2.5 in [ 7], we can prove R(iλI− Ad) =H,∀λ∈R. Finally, by using the closed graph\ntheorem of Banach and Theorem A.2, we conclude the proof of this Theorem. /square\nLet us notice that, under the sole assumptions ( GC) and (SSC), the polynomial stability of system ( 1.5)-(1.8)\nis an open problem.\nAppendix A.Some notions and stability theorems\nIn order to make this paper more self-contained, we recall in this sh ort appendix some notions and stability\nresults used in this work.\nDefinition A.1. Assume that Ais the generator of C0−semigroup of contractions/parenleftbig\netA/parenrightbig\nt≥0on a Hilbert space\nH. TheC0−semigroup/parenleftbig\netA/parenrightbig\nt≥0is said to be\n(1) Strongly stable if\nlim\nt→+∞/ba∇dbletAx0/ba∇dblH= 0,∀x0∈H.\n(2) Exponentially (or uniformly) stable if there exists two positive co nstantsMandεsuch that\n/ba∇dbletAx0/ba∇dblH≤Me−εt/ba∇dblx0/ba∇dblH,∀t >0,∀x0∈H.\n(3) Polynomially stable if there exists two positive constants Candαsuch that\n/ba∇dbletAx0/ba∇dblH≤Ct−α/ba∇dblAx0/ba∇dblH,∀t >0,∀x0∈D(A).\n/square\nTo show the strong stability of the C0-semigroup/parenleftbig\netA/parenrightbig\nt≥0we rely on the following result due to Arendt-Batty\n[9].\nTheorem A.2. Assume that Ais the generatorof a C 0−semigroup of contractions/parenleftbig\netA/parenrightbig\nt≥0on a Hilbert space\nH. IfAhas no pure imaginary eigenvalues and σ(A)∩iRis countable, where σ(A) denotes the spectrum of\nA, then the C0-semigroup/parenleftbig\netA/parenrightbig\nt≥0is strongly stable. /square\nConcerning the characterization of polynomial stability stability of a C0−semigroup of contraction/parenleftbig\netA/parenrightbig\nt≥0we\nrely on the following result due to Borichev and Tomilov [ 12] (see also [ 11] and [21])\n20Theorem A.3. Assume that Ais the generator of a strongly continuous semigroup of contractio ns/parenleftbig\netA/parenrightbig\nt≥0\nonH. IfiR⊂ρ(A), then for a fixed ℓ >0 the following conditions are equivalent\n(A.1) limsup\nλ∈R,|λ|→∞1\n|λ|ℓ/vextenddouble/vextenddouble(iλI−A)−1/vextenddouble/vextenddouble\nL(H)<∞,\n(A.2) /ba∇dbletAU0/ba∇dbl2\nH≤C\nt2\nℓ/ba∇dblU0/ba∇dbl2\nD(A),∀t >0, U0∈D(A),for some C >0.\n/square\nReferences\n[1] F. Abdallah, M. Ghader, and A. Wehbe. Stability results o f a distributed problem involving Bresse system with histor y and/or\nCattaneo law under fully Dirichlet or mixed boundary condit ions.Math. Methods Appl. Sci. , 41(5):1876–1907, 2018.\n[2] M. Akil and H. Badawi. The influence of the physical coeffici ents of a Bresse system with one singular local viscous dampi ng\nin the longitudinal displacement on its stabilization. Evolution Equations & Control Theory , 2022.\n[3] M. Akil, H. Badawi, S. Nicaise, and A. Wehbe. On the stabil ity of Bresse system with one discontinuous local internal\nKelvin–Voigt damping on the axial force. Zeitschrift f¨ ur angewandte Mathematik und Physik , 72(3):126, May 2021.\n[4] M. Akil, H. Badawi, S. Nicaise, and A. Wehbe. Stability re sults of coupled wave models with locally memory in a past his tory\nframework via nonsmooth coefficients on the interface. Mathematical Methods in the Applied Sciences , 44(8):6950–6981, 2021.\n[5] M. Akil, H. Badawi, and A. Wehbe. Stability results of a si ngular local interaction elastic/viscoelastic coupled wa ve equations\nwith time delay. Communications on Pure & Applied Analysis , 20(9):2991–3028, 2021.\n[6] M. Akil, Y. Chitour, M. Ghader, and A. Wehbe. Stability an d exact controllability of a Timoshenko system with only one\nfractional damping on the boundary. Asymptotic Analysis , 119:221–280, 2020. 3-4.\n[7] M. Akil, I. Issa, and A. Wehbe. A N-dimensional elastic/v iscoelastic transmission problem with Kelvin-Voigt dampi ng and\nnon smooth coefficient at the interface, 2021.\n[8] F. Alabau-Boussouira and M. L´ eautaud. Indirect contro llability of locally coupled wave-type systems and applica tions.J.\nMath. Pures Appl. (9) , 99(5):544–576, 2013.\n[9] W. Arendt and C. J. K. Batty. Tauberian theorems and stabi lity of one-parameter semigroups. Trans. Amer. Math. Soc. ,\n306(2):837–852, 1988.\n[10] M. Bassam, D. Mercier, S. Nicaise, and A. Wehbe. Polynom ial stability of the Timoshenko system by one boundary dampi ng.\nJournal of Mathematical Analysis and Applications , 425(2):1177 – 1203, 2015.\n[11] C. J. K. Batty and T. Duyckaerts. Non-uniform stability for bounded semi-groups on Banach sp aces.J. Evol. Equ. , 8(4):765–\n780, 2008.\n[12] A. Borichev and Y. Tomilov. Optimal polynomial decay of functions and operator semigroups. Math. Ann. , 347(2):455–478,\n2010.\n[13] N. Burq and C. Sun. Decay rates for Kelvin-Voigt damped w ave equations II: The geometric control condition. Proc. Amer.\nMath. Soc. , 150(3):1021–1039, 2022.\n[14] L. H. Fatori and R. N. Monteiro. The optimal decay rate fo r a weak dissipative Bresse system. Applied Mathematics Letters ,\n25(3):600 – 604, 2012.\n[15] L. H. Fatori, R. N. Monteiro, and H. D. F. Sare. The Timoshenko system with history and Cattaneo law .Applied Mathematics\nand Computation , 228:128–140, Feb. 2014.\n[16] S. Gerbi, C. Kassem, A. Mortada, and A. Wehbe. Exact cont rollability and stabilization of locally coupled wave equa tions:\ntheoretical results. Z. Anal. Anwend. , 40(1):67–96, 2021.\n[17] A. Hayek, S. Nicaise, Z. Salloum, and A. Wehbe. A transmi ssion problem of a system of weakly coupled wave equations wi th\nKelvin-Voigt dampings and non-smooth coefficient at the inte rface.SeMA J. , 77(3):305–338, 2020.\n[18] C. Kassem, A. Mortada, L. Toufayli, and A. Wehbe. Local i ndirect stabilization of N-d system of two coupled wave equa tions\nunder geometric conditions. C. R. Math. Acad. Sci. Paris , 357(6):494–512, 2019.\n[19] T. Kato. Perturbation Theory for Linear Operators . Springer Berlin Heidelberg, 1995.\n[20] K. Liu and B. Rao. Exponential stability for the wave equ ations with local Kelvin-Voigt damping. Z. Angew. Math. Phys. ,\n57(3):419–432, 2006.\n[21] Z. Liu and B. Rao. Characterization of polynomial decay rate for the solution of linear evolution equation. Z. Angew. Math.\nPhys., 56(4):630–644, 2005.\n[22] A. Pazy. Semigroups of linear operators and applications to partial differential equations , volume 44 of Applied Mathematical\nSciences . Springer-Verlag, New York, 1983.\n[23] L. Tebou. Stabilization of some elastodynamic systems with localized Kelvin-Voigt damping. Discrete Contin. Dyn. Syst. ,\n36(12):7117–7136, 2016.\n[24] A. Wehbe, I. Issa, and M. Akil. Stability results of an el astic/viscoelastic transmission problem of locally coupl ed waves with\nnon smooth coefficients. Acta Applicandae Mathematicae , 171(1):23, Feb 2021.\n21" }, { "title": "1612.09264v1.Gravitational_Searches_for_Lorentz_Violation_with_Matter_and_Astrophysics.pdf", "content": "arXiv:1612.09264v1 [hep-ph] 29 Dec 2016Proceedings of the Seventh Meeting on CPT and Lorentz Symmet ry (CPT’16), Indiana University, Bloomington, June 20-24, 20 16\n1\nGravitational Searches for Lorentz Violation with\nMatter and Astrophysics\nJay D. Tasson\nPhysics Department, St. Olaf College, Northfield, MN 55057, USA\nThis contribution to the CPT’16 proceedings summarizes rec ent tests of\nLorentz violation in the pure-gravity sector with cosmic ra ys and reviews recent\nprogress in matter-gravity couplings.\n1. Introduction\nLorentz violation1as a signalof new physics at the Planckscale2is actively\nsought in a wide variety of tests3within the general test framework of the\ngravitational Standard-Model Extension (SME).4A subset of these tests\ninvolvegravitationalphysics,whichcanprobeLorentzviolationinth epure-\ngravitysectorassociatedwithbothminimal5andnonminimal6–8operators,\nas well as Lorentz violation in matter-gravity couplings.9,10While Lorentz\nviolation in gravity has been sought in a number of systems,3we focus\nhere on recent tests exploring gravitational ˇCerenkov radiation in Sec. 2,\nand review the status of searches for the α(aeff)µcoefficient in the matter\nsector in Sec. 3, including work with superconducting gravimeters.11\n2. Gravitational ˇCerenkov radiation\nTheˇCerenkov radiation of photons by charged particles moving faster than\nthe phase speed of light in ponderable media is a well-known phenomeno n\nin Nature. If the analogous situation of particles exceeding the pha se speed\nof gravity were to occur, as might be the case in General Relativity ( GR)\nin the presence of certain media,12the gravitational ˇCerenkov radiation of\ngravitons would be expected.13In the presence of suitable coefficients for\nLorentz violation in the SME, both electrodynamic14and gravitational6\nˇCerenkov radiation become possible in vacuum. In this section, we re view\nthe tight constraints that have been achieved by considering cosm ic rays\nand mention other possible implications of vacuum gravitational ˇCerenkov\nradiation.6Proceedings of the Seventh Meeting on CPT and Lorentz Symmet ry (CPT’16), Indiana University, Bloomington, June 20-24, 20 16\n2\nAs with all SME searches for Lorentz violation, our analysis begins wit h\nthe expansion about GR and the Standard Model provided by the SM E ac-\ntion. Here we consider the linearized pure-gravity sector, which ha s now\nbeen written explicitlyforoperatorsofarbitrarymassdimension dthat pre-\nserve the usual gauge invariance of GR.8Except where noted, we assume\nthat theothersectorsofthetheoryareconventional. Explorat ionofthe dis-\npersion relation generated by this action reveals that a class of coe fficients\nfor Lorentz violation manifest as a momentum-dependent metric pe rturba-\ntion that generates a momentum-dependent effective index of ref raction for\ngravity. With an appropriate sign for the coefficients for Lorentz v iolation,\nthis index is greater than 1 and particles may exceed the speed of gr avity\nand radiate gravitons. Hence each observation of a high-energy p article can\nplace a one-sided constraint on a combination of these coefficients.\nTo obtain constraints, we perform a calculation of the rate of grav iton\nemission that parallels standard methods assuming for simplicity and d ef-\niniteness that only coefficients at one arbitrary dimension dare nonzero.\nThe calculation is provided for photons, massive scalars, and fermio ns, the\nonly differences in the three cases being the details of the matrix elem ent\nfor the decay and the form of a dimensionless function of din the results.\nThe rate ofpowerloss canthen be integratedto generatearelatio nbetween\nthe time of flight tfor a candidate graviton-radiating particle, the energy\nof the particle at the beginning of its trip Ei, and the observed energy at\nthe end of its trip Ef. This relation takes the form\nt=Fw(d)\nGN(s(d))2/parenleftBigg\n1\nE2d−5\nf−1\nE2d−5\ni/parenrightBigg\n, (1)\nwhereGNis Newton’s constant, Fw(d) is a species wdependent function of\nd, ands(d)isacombinationofcoefficientsforLorentzviolationatdimension\ndthat depends in general on the direction of travel for the particle . This\nresult is distinguished from earlier work on the subject of gravitatio nal\nˇCerenkov radiation15by the connection to the field-theoretic framework\nof the SME, the consideration of anisotropic effects, the explorat ion of\narbitrary dimension d, and the treatment of photons and fermions.\nConservative constraints can be placed using Eq. (1) by setting\n1/E2d−5\ni= 0, solving for s(d),\ns(d)(ˆp)≡(s(d))µνα1...αd−4ˆpµˆpνˆpα1...ˆpαd−40. The potential V(BµBµ∓b2) is\nsuggested, by renormalizability reasons, to be either of the form V=f\n4!(BµBµ∓b2)2with\nfbeing a coupling constant, or of the form V=1\n2σ(BµBµ∓b2), withσbeing a Lagrange\nmultiplier field; for the (+ −−−) signature, the sign in the expression BµBµ∓b2is chosen\nto be (−) for the time-like Bµ, and (+) for the space-like one.\nIt is clear that for both these forms of the potential the minimum is a chieved for\nBµBµ=±b2. Therefore one concludes that the value of Bµequal to the constant vector\nbµ(satisfying the relation bµbµ=±b2), which can be chosen to be, e.g., parallel to one\nof coordinate axes, say x0orxi, introduces a privileged direction. Afterward, one can\nproceed in a manner similar to the Higgs mechanism, that is, we can exp and theBµnear\nthe minimum as Bµ→˜Bµ+bµ, and as a result, we get a mass term for the new field ˜Bµ\nlooking likef\n4!(4˜Bµ˜Bνbµbν+˜B2b2). Thus, we arrive at the essentially Lorentz-breaking\ncontribution to the quadratic action, looking like ˜Bµ˜Bνbµbν. Namely this mechanism is\nresponsible for arising of LV terms in a low-energy limit of string theor y [2], so that it\nis natural to treat LV theories as effective theories for a low-ener gy limit of some more\nfundamental theory. If the bumblebee action, including, what is es pecially important,\nits potential, is induced through radiative corrections, the Lorent z symmetry is said to\nbe broken dynamically. This is the simplest example of a situation where the Lorentz\nsymmetry breaking can be treated as an emergent phenomenon. A lso, in [63], it was\nshown that if the action (2.37) is extended to a curved space-time, one can introduce the\nbumblebee perturbation ˜Bµthrough the replacement Bµ→bµ+˜Bµ, so thatbµis the\nv.e.v. ofBµ, the induced metric is gµν=ηµν+βbµbν, withβis some constant, ˜Bµis\nsmall, and both the potential Vand its derivative vanish at Bµ=bµ(i.e. at˜Bµ= 0). As\na result, one gets the LV action for the ˜Bµ, composed by the sum of the usual Maxwell\nterm, the aether-like term (2.7), where κµνλρis constructed on the base of the vector bµ,\nand the corresponding potential. Therefore, this mechanism allows to generate a CPT-\neven LV terms for the vector field. In [64], this approach was gene ralized for the case of\nthe antisymmetric tensor field.\n21Let us consider now examples of dynamical Lorentz symmetry brea king, for theories\ndescribing vectors and antisymmetric tensors. Some first discuss ions of this approach can\nbe found in [65], and its further development, for the vector field, w as presented in [66],\nand for the antisymmetric tensor one, in [67]. Although these studie s involve calculations\nof quantum corrections, it is very natural to discuss their results namely here, within a\ngeneral context of spontaneous Lorentz symmetry breaking.\nWe start with the following four-fermion model, chosen to be massles s for the sake of\nthe simplicity [66]:\nL0=i¯ψ∂ /ψ−G\n2(¯ψγµγ5ψ)2. (2.38)\nAsusual, weintroducethevectorfield Bµinordertoavoidarisingofthenon-renormalizable\nfour-fermion vertex:\nL=g2\n2BµBµ+¯ψ(i∂ /−eB /γ5)ψ. (2.39)\nIt is immediate to see that, eliminating the Bµfrom this action through purely algebraic\nequations of motion, we recover (2.38) with G=e2\ng2. We can write down the one-loop\ncorrected effective potential for this theory as\nVeff=−g2\n2BµBµ+iTrln(p /−eB /γ5). (2.40)\nTo break theLorentz symmetry, we must first findthe minimum of th eVeff. It isobtained\nfrom the conditiondVeff\ndBµ|eBµ=bµ= 0. The resulting bµvector will introduce the privileged\nspace-time direction. As a result, we obtain\ndV\ndBµ|eB=b=−g2\nebµ−iΠµ= 0, (2.41)\nwhere Π µ= tr/integraltextd4k\n(2π)4i\nk /−b /γ5(−ieγµγ5). Calculating the trace in this expression with use of\nthe exact propagator of the fermionic field ( k /−b /γ5)−1instead of its expansion in power\nseries inbµ(we note that in the massless case this exact propagator has a ver y simple\nform; different expressions for the spinor propagator will be discu ssed in the next chapter,\ndetails of the calculation presented here can be found in [66]), we find that\nΠµ=ieb2\n3π2bµ. (2.42)\nWe note that the Πµis finite: during its calculation, instead of an expected UV divergence\none finds a removable singularity, poleterms cancel out, andno ren ormalization is needed.\nIn next chapters we will see that removable singularities sometimes a rise within one-loop\ncalculations in LV theories, providing finiteness of corresponding co ntributions.\n22So, the gap equation (2.41) yields\ndVeff\ndBµ/vextendsingle/vextendsingle/vextendsingle\neBµ=bµ=/parenleftBigg\n−1\nG+b2\n3π2/parenrightBigg\nebµ= 0, (2.43)\nand its integration gives the effective potential:\nVeff=e4\n12π2B4−e2b2\n6π2B2+α, (2.44)\nwhereαis an arbitrary constant, and for α=b4\n12π2, withb2=3π2\nG, one hasVeff=\n1\n12π2(e2B2−b2)2, that is, the positively defined potential perfectly reproducing th e bum-\nblebee form. If the finite temperature is introduced, one can find t hat this potential\ngenerates phase transitions [66].\nIt is interesting now to consider the dynamics of Bµexpanded about the non-zero\nvacuum< Bµ>=bµ\ne. In this case, we make a shift Bµ=bµ\ne+Aµ, withAµis the new\ndynamical field. In this case, the effective action of Aµis\nSeff=g2\n2(bµ\ne+Aµ)(bµ\ne+Aµ)−itr/integraldisplayd4k\n(2π)4ln(k /−b /γ5−eA /γ5). (2.45)\nAfter one-loop calculations (see [66] for details), we find\nL=−1\n4FµνFµν+e2\n24π2bµǫµνλρAνFλρ−\n−e2\n48π2(∂µAµ)2−e4\n12π2/parenleftbigg\nAµAµ+2\neA·b/parenrightbigg2\n. (2.46)\ni.e. we obtained an appropriate structure of the bumblebee Lagran gian, involving a\nMaxwell-like kinetic term (we note that such a form of the kinetic term is necessary\nto avoid ghosts, see [68]), a positively defined potential, and also, P roca-like and CFJ\nterms.\nNow, let us consider the antisymmetric tensor field case. We start w ith another four-\nfermion model:\nL=¯ψ(i∂ /−m)ψ−G\n2JµνJµν, (2.47)\nwhereJµν=i¯ψγ[µ∂ν]γq\n5ψ, withq= 1,2, is a (pseudo)tensor current which can be coupled\nto an antisymmetric tensor field in the minimal manner, i.e. through th e termBµνJµν.\nWe note that, unlike the above case of the vector bumblebee model, in this theory the\nchoice of zero mass will not simplify calculations essentially. Again, we in troduce the\nbumblebee field, in this case, tensorial one, Bµν, in order to eliminate the four-fermion\n23coupling:\nL=L0+g2\n2/parenleftBigg\nBµν−e\ng2Jµν/parenrightBigg2\n=g2\n2BµνBµν+¯ψ(i∂ /−ieBµνγ[µ∂ν]γq\n5−m)ψ. (2.48)\nSimilarly to (2.40), we can write the one-loop corrected effective pot ential of our theory\nas\nVeff=−g2\n2BµνBµν+itr/integraldisplayd4p\n(2π)4ln(p /−eBµνγ[µpν]γq\n5−m). (2.49)\nIn this case, we find that at q= 2, the one-loop effective potential can be obtained exactly\n[67]. However, for study of spontaneous Lorentz symmetry brea king, assuming that fields\nare small, it is sufficient to find only the terms up to the fourth order in dynamical fields.\nIn [67] our effective potential was shown to look like, after renorma lization:\nV(q=2)\neff=m4\n4(e2BµνBµν−bµνbµν)2+..., (2.50)\nwhere thebµνis the non-zero v.e.v. of Bµν.\nForq= 1, we can find the one-loop effective potential only order by order . The\nrelevant contribution looks like\nV(q=1)\neff=7m4\n12(e2BµνBµν−bµνbµν)2+7e4m4\n48(Bµν˜Bµν)2+..., (2.51)\nwhere˜Bµνis the dual of Bµν:˜Bµν=1\n2ǫµνρσBρσ. So, in both cases we have continuous\nsets of minima allowing for spontaneous symmetry breaking. The der ivative dependent\ncontributions to the effective action, both for q= 1 andq= 2, can be obtained as well.\nExplicitly, the low-energy effective Lagrangians for these theories , after wave function\nrenormalizations and adding some finite constants, look like\nLB,1=−1\n12HµνλHµνλ−7m4\n12(e2BµνBµν−bµνbµν)2−2\n5(∂αBµα)2−\n−7e4m4\n48(Bµν˜Bµν)2+O(B3), (2.52)\nand\nLB,2=−1\n12HµνλHµνλ−m4\n4(e2BµνBµν−bµνbµν)2−e4m4\n16(Bµν˜Bµν)2+\n+O(B3). (2.53)\nThus, we foundthateach of these Lagrangiansdisplays thesame s tructure asinthevector\nfield case (2.46), i.e. it includes a gauge invariant kinetic term, a poten tial allowing to\nbreak the Lorentz symmetry spontaneously, and some other ter ms.\n24To close the section, we note, first, that the Hamiltonian formulatio n of theories with\nspontaneous Lorentz symmetry breaking displays some peculiaritie s [69, 70] (for the dis-\ncussion of degrees of freedom in such theories, see also [71]), sec ond, that the spontaneous\nLorentz symmetry breaking plays a special role in a curved space-t ime where it appears\nto be the only consistent mechanism to break the Lorentz symmetr y, third, that in some\ncases, dynamical Lorentz symmetry breaking can generate a non trivial topology [72].\nThe detailed discussion of Lorentz symmetry breaking in gravity will b e presented in the\nChapter 6.\n2.5 Conclusions\nInthischapter, wediscussed themainclassical impactsoftheLore ntzsymmetry breaking,\nrelated with general structure of classical actions of various field theory models, corre-\nsponding propagators and plane wave solutions. It is clear that the list of results obtained\nat the classical level is much wider. For example, one can mention man y studies of exact\nsolutions and geometric phases. One of the first important results in this direction is the\nexplicit expression for static solutions of the modified Maxwell equat ions in the three-\ndimensional LV extended QED, whose action involves the Julia-Toulou se term (2.14). It\nwas shown in [73] that in this case the static equations are the fourt h order ones, and\nwhile at small distance the scalar potential displays the usual logarit hmic behavior, at\nlarge distance, instead of the usual exponential-like decay occurr ing in the usual Chern-\nSimons modified QED, the potential is dominated by the LV term and log arithmically\ngrows. Further, solutions of classical equations of motion in differe nt LV extensions of\nQED were studied also, e.g. in [74] (see also references therein). Th e paper [75], where a\npossibility of arising the Aharonov-Casher phase in the extended QE D, with the LV term\nwas introduced through the already mentioned magnetic coupling (2 .16), was studied,\nstarted a line of investigation of geometrical phases in LV theories. Further, the impact\nof Lorentz symmetry breaking in solutions of the Dirac equation has been also considered\nin [76, 77, 78] and many other papers; for the review of these stud ies, we recommend also\n[79].\nAll these results clearly show that the LV effects should have nontr ivial implications\nalready at the tree level. At the same time, one of the most importan t problems is –\nwhat possible mechanisms could be responsible for arising the LV modifi cations of known\nfield theory models? One of these mechanisms uses the spontaneou s Lorentz symmetry\nbreaking [2] discussed briefly in this chapter, with its key feature co nsists in arising of\nLV terms at the tree level as a consequence of choosing some minimu m of the potential\n25introducingtheprivilegedvectorortensor, sothattheLagrangia nofthetheoryconsidered\nnear this minimum displays dependence on this vector (tensor), and the another one,\nbased on the old idea of emergent dynamics [24], is a perturbative ge neration approach,\nintensively usedtojustifythepossibilityforarisingtheLVmodificatio nsofthefieldtheory\nmodels, and applied for the first time already in [25]. Following this idea, t here is some\nfundamental spinor matterfieldcoupledinaLVmanner withotherfie lds(scalar, vector or\ngravityones), sothatdifferent LVadditive termsforthesefieldsa riseasone-loopquantum\ncorrections. We note that, in a certain sense, dynamical symmetr y breaking, involving\nboth generating of a quantum correction and choosing of one minimu m, combines these\napproaches. At the same time we note that the coupling of a gauge fi eld to a scalar field\ninstead of a spinor one is less convenient for perturbative generat ion of LV terms since the\ncorresponding action will not include the Dirac matrices necessary t o generate the Levi-\nCivita symbol which is present in CPT-Lorentz breaking terms discus sed in this book.\nThe perturbative generation methodology allows to obtain many of t he terms proposed\nin [34], and we will discuss it in the next chapter.\n26Chapter 3\nPerturbative generation of LV terms\nIn this chapter, we demonstrate the methodology allowing for gene rating various LV\nterms, involving scalar and gauge fields. Conceptually, we follow the a lready mentioned\nidea of the emergent dynamics [24], suggesting that the new contrib utions to Lagrangians\nof different field models can arise as quantum corrections in some fun damental theories.\nNamely this method has been used in[25], where anappropriate LV cou pling of the spinor\nfield to a constant axial vector bµwithin a LV extension of QED was applied to generate\nthe CFJ term. Just this approach became further the main tool allo wing to obtain new\nLV terms. Besides of this, it is worth to mention also other methodolo gies allowing to do\nthis, such as: dimensional reduction [80], spontaneous Lorentz sy mmetry breaking, that\nwas briefly discussed in the previous chapter, and known, in some ca ses, like in [81], to\noccur in the one-loop corrected effective action of a correspondin g theory, so that one can\nspeak about the dynamical Lorentz symmetry breaking, and the n oncommutative fields\nmethod [82]. However, the perturbative generation seems to be th e most natural manner\nfor obtaining LV terms. We note nevertheless that this way is espec ially appropriate just\nin the case when the corresponding LV quantum corrections are fin ite, since otherwise one\nshould, inordertoensurethemultiplicativerenormalizability, introdu cethecorresponding\nLV term already at the tree level, which clearly contradicts to the int erpretation of the\ncorresponding new LV term as an emergent phenomenon.\n3.1 Problem of ambiguities\nFirst of all, we should discuss the reasons for the most controvers ial phenomenon charac-\nteristicforfour-dimensionalLVgaugecontributions, thatis, the arisingoftheambiguities.\nThe characteristic situation [26] is the following one: the one-loop int egral over the inter-\nnal momenta is an expression looking like a sum of two divergent terms , with the pole\n27parts of these terms mutually cancel. However, any divergent ter m represents itself as a\nsum of a pole part and a finite part. In principle, this finite part can be fixed with use of\nan appropriate normalization condition [83]. Nevertheless, in genera l, there is no reasons\nto prefer any of different normalization conditions in LV theories, es pecially taking into\naccount that the corresponding terms arising after the calculatio ns are actually finite and\nthus do not depend on any renormalization scale. Therefore, within different regulariza-\ntions, or, as is the same, different normalization conditions, the finit e parts turn out to\ndepend on the regularization scheme, and hence the result for the corresponding integral\nover internal momenta is ambiguous. From the formal viewpoint, th is ambiguity is a per-\nfect example of the indeterminate form ∞−∞, whose fixing requires special conditions.\nTypically, two most known ambiguous integrals are considered. The fi rst of them arises\nwithin study of the CFJ term, see e.g. [38], looking like\nJµν=/integraldisplayd4p\n(2π)4ηµνp2−4pµpν+3m2ηµν\n(p2−m2)3. (3.1)\nOne can easily verify that by carrying out the replacement pµpν→1\n4ηµνp2, with a sub-\nsequent integration in four dimensions, and the replacement pµpν→1\n4+ǫηµνp2, with a\nsubsequent integration in 4+ ǫdimensions, and, afterward, taking the limit ǫ→0, one\narrives at different results for Jµν[38]. This difference stays at the finite temperature\nas well, and, moreover, these two possibilities do not exhaust a list of other manners of\ncalculating this integral, with more results can be obtained.\nAnother ambiguous integral was found in [40]. It reads as\nIµν=/integraldisplayd4p\n(2π)4m2ηµν+2pµpν−ηµνp2\n(p2−m2)2, (3.2)\nand also yields different results under the same replacements as abo ve. Thus, we can\nconclude that in LV theories, in certain cases, formally divergent co ntributions can be\nactually finite, but display ambiguities as a consequence of arising the indeterminate form\n∞−∞.\nIn principle, more ambiguous integrals can exist. For example, in a man ner similar to\nthe above discussions, one can show that logarithmically divergent in tegrals of the form\nJ(n)\nµν=/integraldisplayd4p\n(2π)4p2n(ηµνp2−4pµpν+3m2ηµν)\n(p2−m2)n+3, (3.3)\nfor all non-negative n, will be ambiguous by the same reasons, and one of the possi-\nbilities for an analogous generalization for Iµνis described by the following superficially\nquadratically divergent integrals:\nI(n)\nµν=/integraldisplayd4p\n(2π)4p2n(m2ηµν+2pµpν−ηµνp2)\n(p2−m2)n+2. (3.4)\n28However, up to now, there is no known constructive examples of ph ysically motivated\nscenarios where both these ambiguous integrals really arise, excep t of the usual case n= 0\ncorresponding to eqs. (3.1,3.2) and occurring in the above-mention ed LV extensions of\nQED. In principle, other, more involved, ambiguous loop integrals can exist as well, see\nf.e. [84].\nAt the same time, it should be noted that the chiral QED, whose Lagr angian is given\nby [85]\nL=¯ψ(i∂ /+eA /(1−γ5)+b /(1+γ5)−m)ψ, (3.5)\ndisplays a highly nontrivial behavior. Indeed, since the spinor propa gator in this theory\nis the common one, that is,i\nk /−m, the CFJ-like contribution to the effective Lagrangian\nwill be given by the expression:\nS2(p) =−e2\n2tr/integraldisplayd4k\n(2π)4/bracketleftBig1\nk /+p /−mA /(−p)(1−γ5)1\nk /−mb /(1+γ5)×\n×1\nk /−mA /(p)(1−γ5)+ (3.6)\n+1\nk /+p /−mb /(1+γ5)1\nk /+p /−mA /(−p)(1−γ5)×\n×1\nk /−mA /(p)(1−γ5)/bracketrightBig\n.\nLet us now take into account only the first order in the external mo mentump, use the\nexpression1\nk /−m=k /+m\nk2−m2, and consider e.g. only the first contribution of S2. Then, we get\nS2,1(p) =e2\n2tr/integraldisplayd4k\n(2π)4k /+m\n(k2−m2)4p /(k /+m)A /(−p)×\n×(1−γ5)(k /+m)b /(1+γ5)(k /+m)A /(p)(1−γ5). (3.7)\nThis expression is superficially logarithmically divergent and involves th e factor (1 −\nγ5)(k /+m)b /(1 +γ5). However, γ5commutes with even degrees of γµand anticommutes\nwithoddones, hence, theUVleadingmomentumdependentfactor, whichonlycontributes\nto the divergence, turns out to be proportional to (1 −γ5)(1 +γ5) = 0. Thus, the\npotential divergence disappears. The similar situation occurs in ano ther contribution to\n(3.6), implying also vanishing of the leading (divergent) contribution. As a result, unlike\nthe usual calculation of the CFJ term, after the calculation of the t race, the resulting\nexpression is superficially finite and hence contains no ambiguity. The refore, in the chiral\nQED (3.5) the CFJ term is ambiguity-free. The explicit result for it is [85 ]:\nLCFJ=e2\n3π2ǫµνλρbµAν∂λAρ. (3.8)\n29However, one can show that in this case, the absence of an ambiguit y is paid by a price\nof existence of the gauge anomaly. Indeed, it is easy to see that in t his case the gauge-\nbreaking Proca-like divergent term is generated. Hence, the cons istent elimination of\nambiguities requires a strong extension of the model which has been performed in [86].\n3.2 LV contributions in the scalar sector\nIn this section, we discuss the simplest possible LV modifications for t he scalar field mod-\nels. There are two typical terms, namely, the aether-like term (2.1 0) and the essentially\ntwo-dimensional Chern-Simons-like term (2.15).\nThe simplest manner to generate the aether-like term is based on th e use of the LV\nYukawa-like action\nS=/integraldisplay\ndDx/bracketleftBig¯ψ(i∂ /−m)ψ+1\n2(∂µφ∂µφ+m2φ2)−g¯ψa /ψφ/bracketrightBig\n, (3.9)\nwhere the LV vector aµis dimensionless, and the new vertex, given by (2.18), is one of\nthe ingredients of the LV SME [6].\nThe lower contribution to the aether-like term is given by the Feynma n diagram de-\npicted by the Fig. 3.1 [40]:\n••\nFigure 3.1: Two-point function of the scalar field.\nAfter calculating thetraceanddisregarding thetermsproportion altoa2(denoted here\nby dots), which do not break the Lorentz symmetry, we arrive at t he following result:\nS2(p) =−d\n2g2φ(p)φ(−p)[ηµνηρσ+ηµσηρν]aµaρ×\n×/integraldisplaydDk\n(2π)Dkν(kσ+pσ)\n[k2−m2][(k+p)2−m2]+..., (3.10)\nwheredisadimensionofDiracmatricesinthecorrespondingspace-time, and Disaspace-\ntime dimension. It is clear that the aether-like contribution from this integral is finite\nin two- and three-dimensional space-times, and, within the framew ork of the dimensional\nregularization, also in all space-times with higher odd dimensions. Exp licitly, one gets,\nafter returning to the coordinate space:\nS2=−dg2Γ(2−D\n2)\n6(4π)D/2(m2)2−D/2φ(a·∂)2φ. (3.11)\n30This term evidently replays the structure (2.10). We see that at D= 4, the aether-\nlike term should be present already at the tree level, in order to intro duce the one-loop\ncounterterm of the form1\n2Zaetherφ(a·∂)2φ, withZaether= 1−g2\n12π2ǫ. We note that in this\ncase the renormalizability of the theory is not spoiled since no coupling s with a negative\nmass dimension are present. In principle, the aether term for the s calar field can be\ngenerated also through other couplings to a spinor matter, e.g. th e CPT-even one\nS=/integraldisplay\ndDx/bracketleftBig¯ψ(i∂ /+ikµνγµ∂ν−hφ−m)ψ+1\n2(∂µφ∂µφ+m2φ2)/bracketrightBig\n, (3.12)\nwithkµνis a constant tensor, and again, the aether-like term will be finite in t wo, three\nand all higher odd dimensions, and divergent in four dimensions.\nThereisonemorepossibleLVadditivetermforthescalartheory, th atis, theessentially\ntwo-dimensional term(2.15)(see[51]foritsapplicationsforstudie s oftopologicaldefects).\nTo generate it, we start with the following two-dimensional action, s ee [87]:\nS=/integraldisplay\nd2x¯ψ/bracketleftBig\ni∂ /−m−ga /φ−mgγ5χ/bracketrightBig\nψ, (3.13)\nwhereγ5=γ0γ1now is a two-dimensional analogue of the chirality matrix, φis a scalar\nfield (denoted by a solid line), and χis a pseudoscalar one (denoted by a dashed-and-\ndotted line).\n...◦•\nFigure 3.2: Two-point function of φandχfields.\nHere the dark circle is for a /insertion, and the light one is for γ5one. The contribution\nof the Fig. 3.2 looks like\nΓ =/integraldisplayd2p\n(2π)2φ(−p)Π(p)χ(p), (3.14)\nwith Π(p) being a self-energy tensor given by\nΠ(p) =−img2tr/integraldisplayd2k\n(2π)2/bracketleftBig\na /i\nk /−mγ5i\nk /+p /−m/bracketrightBig\n. (3.15)\nAfter straightforward calculations, we find that, in the first orde r in the external pµ,\nΠ(p) =−ig2\n2πǫµνaµpν. (3.16)\n31So, our final result for this contribution to the effective action is\nΓ =g2\n2π/integraldisplay\nd2xaµǫµνφ∂νχ. (3.17)\nThis is just the term (2.15). We note that it is ambiguity-free as is sho uld be, since the\nresult is superficially finite. In principle, such a term, for the special case of a space-like\naµ(namely, the case a0= 0), can be also generated through the noncommutative field\nmethod proposed in [82].\nIn principle, other LV additive terms, involving scalar fields only, can b e introduced,\ne.g., the Myers-Pospelov-like ones (cf. [42]):\nLMP,scal=1\nΛN¯φ(n·∂)Nφ, (3.18)\nwhere Λ Nis some scale factor of an appropriate mass dimension, nµis a dimensionless\nvector, and N≥3. The importance of these terms consists in the fact that for a sp ace-like\nnµ, there will be no higher time derivatives and hence no ghosts, so, on e has a consistent\nscheme to incorporate higher derivatives to a field theory model wit hout use of a Horava-\nLifshitz approach. These terms can be generated through schem es we presented above,\nat higher orders of a derivative expansion of the corresponding eff ective action.\n3.3 Mixed LV scalar-vector contribution\nIn three space-time dimensions, there is an unique possibility to cons truct the quadratic\nLVadditiveterm, involvingfieldsofdifferentnatures, scalaroneand gaugeone,thatis, the\nJulia-Toulouse term (2.14). In the section 2.3, we already noted tha t this term naturally\nemerges within the gauge embedding of the special Lorentz-CPT br eaking extension of\nthe self-dual model. However, this is not an unique way to generate the term (2.14).\nFollowing [88], we start with the following three-dimensional action:\nS=/integraldisplay\nd3x¯ψ/bracketleftBig\ni∂ /−m−eA /−ga /φ/bracketrightBig\nψ. (3.19)\nHereaµis a LV constant vector, and φis a scalar. The one-loop effective action is clearly\ngiven by the functional trace:\nΓ(1)=−iTrln/bracketleftBig\ni∂ /−m−eA /−ga /φ/bracketrightBig\n. (3.20)\nThe desired two-point “mixed” function is described by the Feynman diagram given by\nthe Fig. 3.3.\nHere the black dot is for the a /insertion, the dashed line is for a spinor propagator,\nand solid and wavy lines denote external scalar and gauge fields, res pectively.\n32•\nFigure 3.3: Mixed two-point vector-scalar function.\nThe contribution of this diagram yields\nI=−egtr/integraldisplayd3p\n(2π)3/integraldisplayd3k\n(2π)3A /(−p)(k /+m)φ(p)a /(k /+p /+m)×\n×1\n(k2−m2)[(k+p)2−m2]. (3.21)\nForthediag(+−−) signature, andDiracmatrices ( γ0)α\nβ=σ2,(γ1)α\nβ=iσ1,(γ0)α\nβ=iσ3,\nobeying relations: {γµ,γν}= 2ηµν, tr(γµγνγλ) = 2iǫµνλ, we arrive at\nI=−2iǫαµνegm/integraldisplayd3p\n(2π)3pαAµ(−p)φ(p)aν×\n×/integraldisplayd3k\n(2π)31\n(k2−m2)[(k+p)2−m2]. (3.22)\nThen, after integrating over internal momenta, in the first order inpµwe obtain\nI=ǫλµνegm\n4π|m|/integraldisplayd3p\n(2π)3pλAµ(−p)φ(p)aν. (3.23)\nThis expression can be represented as\nI=−eg\n8πsgn(m)/integraldisplay\nd3xǫλµνFλµaνφ. (3.24)\nThis is exactly the Julia-Toulouse term (2.14). We note that here the sign of the mass\narises. It is rather typical for the three-dimensional theories, s ee e.g. [89], where the\nsimilar situation was shown to take place for the Chern-Simons term.\nThis calculation can be performed through evaluating (3.20) not only with the Feyn-\nman diagram approach as we did now, but also through the proper-t ime method. To\nproceed with it, we add to the (3.20) the field independent term −iTrln(i∂ /+m), in order\nto have a trace of an operator of second order in derivatives, as t he proper-time method\nrequires. So, our one-loop effective action becomes:\nΓ(1)=−iTrln(−✷−m2−eA /(i∂ /+m)−gφa /(i∂ /+m)). (3.25)\nSince we need only the first order in aµ, we expand this expression up to the first order\ninaµand then employ the Schwinger proper-time representation A−1=1\ni/integraltext∞\n0dseisA. We\narrive at\nΓ(1)\n1=−gTr/bracketleftBig/integraldisplay∞\n0dseis(✷+m2+eA /(i∂ /+m))φa /(i∂ /+m)/bracketrightBig\n. (3.26)\n33Then we expand this expression up to first derivatives and to the fir st order in Aµ. We\nobtain\nΓ(1)\n1=−egTr/bracketleftBig/integraldisplay∞\n0dseism2/parenleftBig\nisA /(i∂ /+m)−s2(∂µA /)(i∂ /+m)∂µ/parenrightBig\n×\n×φa /(i∂ /+m)eis✷/bracketrightBig\n. (3.27)\nAfter keeping only the terms with three Dirac matrices, with the sub sequent Fourier\ntransform and integrating over momenta, we arrive at the result ( 3.24).\nWe note, that, in principle, our result for the Julia-Toulouse term co uld be obtained\nwithin usual calculations of the Chern-Simons term, where one could do the replacement\nAµ→Aµ+aµφ. Also, if one would like to consider the free scalar-vector theory wh ose\naction involves the usual kinetic term for the scalar field, the Maxwe ll term, and the\nJulia-Toulouse term, it is possible to show that the dispersion relation s are just those\nones given by (2.31).\nThere aresomeother mannerstogeneratethemixed term(2.14). Oneofthemconsists\nin a dual immersion of the LV extension of the self-dual theory, nam ely this way has been\nemployed in the section 2.3. Another one consists in a dimensional red uction of the\nelectrodynamics with the CFJ term carried out in [80] (see also [90]). W e start with its\nfour-dimensional Lagrangian looking like\nL=−1\n4FµνFµν+ǫµνλρkµAν∂λAρ, (3.28)\nand “freeze” the dynamics along the third spatial dimension (perha ps, this method can\nbe an useful tool to describe essentially two-dimensional quantum systems, for example,\ngraphene), so, one can split the indices like ν= (a,3), with the index atakes values 0 ,1,2,\nandA3=φ, and the dependence on x3will be suppressed (thus, ∂3φ=∂3Aa= 0), so,\none arrives at the following dimensionally reduced Lagrangian (with ǫabc≡ǫ3abc):\nLred=−1\n4FabFab−1\n2∂aφ∂aφ+ǫabc(µAa∂bAc−kaFbcφ). (3.29)\nWe see that within this reduction, the Chern-Simons term with a mass µ=k3and the\nJT term are generated.\nOne more way to generate this term is based on the dynamical Loren tz symmetry\nbreaking approach, presented in [81]. We proceed in a manner similar t o that one de-\nscribed in the section 2.4 (see also [65]). The key idea is the following one : in the\nthree-dimensional space-time, we consider several four-fermio n interactions which, as it\nis usually done within the Gross-Neveu approach, afterwards are p resented in terms of\nvertices involving corresponding different scalar or vector auxiliary fieldsAI, whereIis a\n34generalized summation index. As a result, we have the equivalent the ory represented by\nthe following Lagrangian:\nL=g2\nI\n2AIAI+¯ψ(i∂ /−m−eIAIΓI)ψ, (3.30)\nwhereAI={Vµ,Θ,Aµ,Φ,Tµ}is a set of the vector and scalar fields coupled to spinors\nwith use of matrices Γ I={γµ,γ3,γµγ5,γ3γ5,γµγ3γ5}playing here the role of generalized\nDirac matrices, and ψis the four-component spinor. Here we use 4 ×4 matrices, whose\nexplicitformisgivenin[81], so, ourspinorrepresentationoftheLore ntzgroupisreducible.\nThe one-loop corrected effective potential of the theory defined in a manner analogous\nto that one used in the section 2.4, i.e. as\nVeff=−g2\nI\n2AIAI+itr/integraldisplayd3k\n(2π)3ln(k /−m−eIAIΓI), (3.31)\nwhere only derivative independent terms are taken into account, e vidently possesses non-\ntrivial minima at some ∝an}b∇acketle{tAI∝an}b∇acket∇i}ht=bI/eI, withbI={0,0,bµ,φ0,0}. Now, let us make the\nshift of the vacuum by the rule AI→ AI+bI/eI, so that now ∝an}b∇acketle{tAI∝an}b∇acket∇i}ht= 0. As a result, the\none-loop contribution to the effective action (we note that here th e overall sign is changed\nas it must be!) takes the form\nSeff[A,b] =−iTrln(i∂ /−m−b·Γ−eIAIΓI). (3.32)\nIn this effective action we take into account only the lower, quadrat ic order in fields AI,\ngiven by\nS(2)\neff[A,b] =i\n2/integraldisplay\nd3xAIΠIJAJ, (3.33)\nwith the corresponding self-energy tensor ΠIJwritten within the derivative expansion\nmethodology[91](thedetaileddiscussionofthisapproachwillbegive ninthenextsection)\nlooks like\nΠIJ=−eIeJtr/integraldisplayd3p\n(2π)3i\ni∂ /−m−b·ΓΓIi\np /−i∂ /−m−b·ΓΓJ. (3.34)\nExpanding this expression up to the first order in derivatives and ca lculating correspond-\ning traces and integrals, we obtain the following mixed and convention al Chern-Simons\nterms:\nLCS=−eVeV3\n2πmΘǫµλνbµ∂λVν−eVeT\nπǫµλνVµ∂λTν+eV3eT\nπΘgµνbµTν\n+eAeA3\n6πmΦǫµλνbµ∂λAν+e2\nA\n6πmφ0ǫµλνAµ∂λAν. (3.35)\n35We note here also the presence of Julia-Toulouse terms, i.e., the firs t and fourth terms.\nByanalyzing theabove Chern-Simons-like Lagrangian, we candiscus s two models that\nexhibit charge fractionalization without the breaking of Tsymmetry. First, if we restrict\nourselves to the case Tµ=Aµ= Φ = 0, i.e. we have only the usual vector Vµand the\nscalar Θ, we arrive at\nLCS=eV3eV\n2πmϑǫµλν∂µΘ∂λVν, (3.36)\nwhere we have introduced ϑ(x) =xµbµ+C, withCbeing a constant. At the same way,\nif we choose Vµ= Θ =Tµ= 0, we have\nLCS=−eA3eA\n6πmϑǫµλν∂µΦ∂λAν, (3.37)\nwhich, upto anadjustment ofnumerical constants, isthe low-ene rgy mixed Chern-Simons\nterm forgraphene discussed in[92, 93]. Byconsidering the propert ies of the bilinear terms\nassociated to the fields AI, with respect to C,P, andTtransformations, we note that\nboth terms (3.36) and (3.37) violate the Csymmetry, whereas the PandTsymmetries\nare preserved. This, obviously, leads to a CPT symmetry violation, w ith theCsymme-\ntry violation is a direct consequence of the charge fractionalization [81], whereas the T\nsymmetry is preserved. This is another manner to generate the mix ed scalar-vector term.\nIn principle, it can be obtained also in an essentially non-perturbative way, that is, the\nJulia-Toulouse technique used for study of condensation of topolo gical defects [88]. To\nconclude this section, we note that this term, first, can arise by diff erent reasons, second,\napparently can be useful to study condensed matter systems.\n3.4 The Carroll-Field-Jackiw term\n3.4.1 General situation and zero temperature case\nThe CFJ term (1.1) is, without doubts, a paradigmatic example of the Lorentz-breaking\nterm. As we noted in the beginning of this chapter, it can be generat ed as a quantum cor-\nrection from the following in the appropriate LV extension of the spin or QED. Explicitly,\nthis theory is described by the Lagrangian:\nLQED=−1\n4FµνFµν+¯ψ(i∂ /−eA /−m+b /γ5)ψ. (3.38)\nThis possibility was for the first time described in [25] and, afterwar d, discussed in numer-\nous papers. The one-loop effective action is clearly given by the ferm ionic determinant:\nΓ(1)=−iTrln(i∂ /−eA /−m+b /γ5). (3.39)\n36The CFJ term is evidently generated by the expansion of this determ inant up to the\nsecond order in Aµ, and to the first order in derivatives. Its final form is\nLCFJ=Ce2ǫµνλρbµAν∂λAρ, (3.40)\nwithCbeing some numerical constant. To obtain the CFJ term, and conse quently the\nconstantC, one considers the contributions to the two-point function descr ibed by the\nfollowing Feynman diagrams:\n•\n•\nFigure 3.4: Contributions to the CFJ term.\nThe external lines are the gauge fields, and the •symbol here is for the b /γ5insertion.\nSo, let us describe some ways to calculate the contributions of thes e two graphs.\nWithin the first approach we use the usual spinor propagators. In this case, the sum\nof these graphs yields\nΓ2[A] =−e2\n2/integraldisplayd4p\n(2π)4Aµ(−p)Πµν(p)Aν(p), (3.41)\nwhere\nΠµν(p) = tr/integraldisplayd4k\n(2π)4/bracketleftBig\nγµ1\nk /−mb /γ51\nk /−mγν1\nk /+p /−m+\n+γµ1\nk /−mγν1\nk /+p /−mb /γ51\nk /+p /−m/bracketrightBig\n(3.42)\nis the self-energy tensor for this theory up to the first order in bµ. Following one manner,\nto obtain the CFJ term, one should expand the integrand up to the fi rst order in the\nexternalmomentum p. Namelythiswaywasemployedintheoriginalpaper[25]. However,\nalready inthat paper, it was arguedthat the CFJterm, being super ficially logarithmically\ndivergent, is actually finite but ambiguous, Indeed, this procedure yields the result\nΠµν=1\n2π2ǫµναβbαpβ/parenleftBiggθ\nsinθ−1\n4/parenrightBigg\n, (3.43)\nwhereθ= 2arcsin(√p2/2m). Therefore, in the low-energy limit p→0, one has Πµν=\n3\n8π2ǫµναβbαpβ, and, then, C=3\n16π2. One can develop as well another procedure, within\nwhich thebµbecomes an external non-constant field depending on one more ex ternal\n37momentum, so, the only gauge invariant result for the CFJ contribu tion to the effective\nLagrangian is zero. This suggestion was originally made in [94] and discu ssed in [25].\nFurther, in [95, 96] it has been argued that, in a certain sense, the zero result for the CFJ\nterm ispreferable since, if onewould treat bµasa dynamical field, even the contribution of\nthe CFJ term to the effective action will break the gauge invariance, so, the zero value for\nthe coefficient Cis necessary to maintain the gauge symmetry. The same argumenta tion\ncan be applied to the four-dimensional gravitational CS term as well [95].\nInstead of expanding the effective action in series in bµ, one can perform the sum over\nall possible b /γ5insertions and obtain the exact propagator of the spinor field [97]:\nS(p) =i(p /−m+b /γ5)−1=ip2+m2−b2−2(b·p+mb /)γ5\n(p2+b2−m2)2−4[(b·p)2−m2b2]×\n×(p /+m−b /γ5), (3.44)\ntogether with the dimensional regularization, one arrives at the re sult [97]:\nΠµν=−1\n2π2ǫµναβbαpβ\n1−θ(−b2−m2)/radicalBigg\n1+m2\nb2\n. (3.45)\nThus, forb2>0 (that is, time-like bµ), one has Πµν=−1\n2π2ǫµναβbαpβ, andC=1\n4π2,\nwhile for the zero mass and the space-like bµone finds a zero result, so, C= 0 (see\ndetails in [97]). We note that the exact propagator (3.44) is especially convenient for\nthe calculations in the massless fermions case since it displays no infra red singularities at\nm= 0. As we already mentioned in the section 3.1, the ambiguity is a price w e must pay\nfor a finiteness of the superficially (logarithmically) divergent result . It was argued in [26]\nthat this ambiguity is strongly related with the Adler-Bell-Jackiw anom aly.\nLet us also discuss some other ways to calculate the CFJ term. One o f the main\nmanners to do it is based on the derivative expansion method [91]. The essence of this\nmethod is as follows. We carry out a formal Fourier transformation of the determinant\n(3.39) which we then rewrite as\nΓ(1)=−iTrln(p /−eA /−m+b /γ5) =\n=−iTr(1−eA /1\np /−m+b /γ5)+···, (3.46)\nwhere dots are for the field-independent contribution. However, now, to extract possi-\nble contributions depending on derivatives of Aµ, we use the following identity which is\nstraightforward in the coordinate representation, it is the key ide ntity of this approach\nreflecting the fact that the fields and derivatives do not commute, withpµbeing an inte-\ngration momentum in the corresponding loop:\nAαpµ=pµAα−i(∂µAα). (3.47)\n38This identity can be used to put all internal momenta in positions allowin g to integrate\nover them (with the ”correct” position of internal momenta is at th e left-hand side from\nall fields). It can be straightforwardly generalized for higher orde rs in derivatives. Equiv-\nalently, we can start with the expression (3.46) with the subsequen t replacements: i∂µ→\npµ,Aµ(x)→Aµ(x−i∂\n∂p), andfurtherweexpand Aµ(x−i∂\n∂p) =Aµ(x)−i(∂νAµ)(x)∂\n∂pν+...,\nand the derivatives with respect to pµmust act on propagators1\np /−m+b /γ5. Keeping only\nthe terms of the second order in fields Aαand of the first order in their derivatives and\nin the constant bµvector, we arrive at the following expression for the one-loop effec tive\nLagrangian [98]:\nL(1)= 2ie2/integraldisplayd4p\n(2π)41\n(p2−m2)3[ǫνλµρ(3m2+p2)−4ǫνλµαpαpρ]bρAν∂λAµ\n=e2ǫνλµαkαAν∂λAµ, (3.48)\nwherekµ=Cbµ, andCis defined by the Eq. (3.40). The constant Cis easily shown to be\nfinite and ambiguous, just as we discussed in the Section 3.1: if we rep lacepµpν→p2\n4ηµν\nand integrate in 4 dimensions, we get C=3\n16π2, and if we replace pµpν→1\n4+ǫηµνp2and\nintegrate in 4+ ǫdimensions, we arrive at C=1\n4π2(the same values of Care presented\nin [38], arising not only within calculating the usual CFJ term but also with in computing\nits non-Abelian extension). Besides of these results, the coefficien tCwas calculated with\nuse of the exact propagator (3.44) in the massless case [99] and fo und to be equal to1\n8π2.\nWithin the Pauli-Villars regularization, the coefficient Cvanishes [6].\nOne more methodology to treat the CFJ term in the massless case wa s proposed in\n[100]. To do it, we express the exact propagator (3.44) for m= 0 as\nSexact(p) =i\nk /+b /γ5=i\np /+b /PL+i\np /−b /PR, (3.49)\nwherePR=1+γ5\n2andPL=1−γ5\n2are the left and right chiral projectors Actually, namely\nthese projectors were used in the section 3.1. The self-energy te nsor is expressed in terms\nof exact propagators in the simplest form,\nΠµν(p) =1\n2tr/integraldisplayd4q\n(2π)4γµSexact(q)γνSexact(q+p), (3.50)\nso, it remains to substitute here directly the exact massless propa gator (3.49) and extract\nonly the CFJ-like terms, that is, those ones involving only one γ5, which clearly can yield\nthe Levi-Civita symbol. The corresponding result is [100]:\nΠµν\nCFJ(p) = 4ibβǫµναβ/integraldisplayd4q\n(2π)4(pα+qα)\n(q+b)2(q+p+b)2, (3.51)\n39with the integral over kis supposed to be regularized in some way. Afterwards, we apply\nthe methodology of the implicit regularization (see e.g. [101]), which a llows us to find\nthat\nΠµν\nCFJ(p) =−4iα2bβpαǫµναβ, (3.52)\nwhereα2is some completely undetermined constant, finite but strongly depe nding on\nthe regularization prescription. It is clear that this is the only possib le one-derivative\ncontribution to the self-energy tensor proportional to ǫµναβ. In [100] the α2is treated as\na surface term arising within the integration, explicitly defined throu gh the relation\nα2gµν=/integraldisplayd4q\n(2π)4∂\n∂qµ/bracketleftBiggqν\n(q2−λ2)2/bracketrightBigg\n,\nwithλis a mass scale. From a formal viewpoint the arising of α2is related with the\nfact that integrals I0α=pα/integraltextd4q\n(2π)41\n(q+b)2(q+p+b)2andIα=/integraltextd4q\n(2π)4qα\n(q+b)2(q+p+b)2diverge, and\nthe Πµν\nCFJwill be proportional to their linear combination looking like a momentum pα\nmultiplied by some undetermined expression behaving as ∞−∞, withα2will be just the\nvalue of this expression, which is not fixed by any physical reasons.\nOnce more manner to argue that the CFJ term is strongly ambiguous being propor-\ntional to a completely undetermined constant isbased onthe funct ional integral argumen-\ntation [102]. To do it, we start with the functional integral corresp onding to the one-loop\neffective action (3.39):\nZ[A] =/integraldisplay\nDψD¯ψexp(iS[A]) =\n=/integraldisplay\nDψD¯ψexp(i/integraldisplay\nd4x¯ψ(iD /−m−b /γ5)ψ), (3.53)\nwhereDµ=∂µ−ieAµis a usual gauge covariant derivative. Then, after the changes of\nvariablesψ(x)→eiα(x)γ5ψ(x),¯ψ(x)→¯ψ(x)eiα(x)γ5, the integral measure is modified as\nDψD¯ψ→DψD¯ψexp/parenleftBig\n−i\n8π2/integraltextd4xα(x)˜FαβFαβ/parenrightBig\n, see details in [102, 103]. The action S[A]\nunder this changes of variables transforms to S′[A] =/integraltextd4x[¯ψ(iD /−me2iα(x)γ5−b /γ5)ψ−\n∂βα(x)Jβ\n5], withJβ\n5being a conserved axial current which, however, can be determine d\nin an ambiguous manner like Jβ\n5=¯ψγβγ5ψ+c˜FβγAγ, with an arbitrary constant c, and\nthe˜Fβγis a tensor dual to Fµν, since the only restriction for this axial current is that it\nshould satisfy the conservation law ∂βJβ\n5= 0. Choosing α(x) =−xαbα, one rests with\nZ[A] =ei(c−1\n4π2)/integraltext\nd4xbα˜FαβAβ×\n×/integraldisplay\nDψD¯ψexp(i/integraldisplay\nd4x¯ψ(iD /−me2iγ5xµbµ)ψ). (3.54)\nWe calculate the functional determinant of iD /−me2iγ5xµbµup to the first order in bµand\nthesecondorder inthefield Aµ. Inthemomentum space, multiplicationby thecoordinate\n40xµis equivalent to differentiation with respect to the corresponding mo menta, so we can\nperform the loop integration as above in this section, with the result for the exponential\nof the fermionic determinant will be [102]:\n/integraldisplay\nDψD¯ψexp(i/integraldisplay\nd4x¯ψ(iD /−me2iγ5xµbµ)ψ) =\n= exp(i\n4π2/integraldisplay\nd4xbα˜FαβAβ). (3.55)\nSubstituting this result to the generating functional Z[A] (3.54), we find that the only\nsurviving term is that one proportional to the constant c. Thus, the corresponding contri-\nbution to the effective action Γ[ A] =−ilnZ[A] is completely undetermined being equal\nto\nΓ[A] =c/integraldisplay\nd4xbα˜FαβAβ, (3.56)\nthat is, the CFJ coefficient Cfrom (3.40) within this prescription is equal to the factor c.\nWe note that this fact completely matches the conclusion achieved in [100] on the base\nof the implicit regularization. Regarding the argumentation present ed in [83], we note\nthat it is based on choosing appropriate normalization conditions, wh ereas there is no\nunique prescription to define these conditions once and forever, s o, the ambiguity can be\neliminated only within a separate calculation, but there is no profound reasons to say that\none value of the constant Cis correct but others are not. Actually, the papers [102, 100]\nargued that the ambiguity of the CFJ term is intrinsic. We note that t he resultC=1\n4π2\nhas been unambiguously obtained also within the Schwinger proper tim e approach [104].\nHowever, using of the proper time formulation itself involves some sp ecific regularization\nprescription.\n3.4.2 Finite temperature case\nThe ambiguity we found for the CFJ term can be naturally generalized for the finite\ntemperature case. To illustrate the situation, we study the contr ibution (3.48) at the\nfinite temperature within two different regularization schemes analo gous to those ones\nused in a previous subsection, that is, we suggest the purely spatia l dimension dto be\neither 3 or 3 + ǫwithǫ→0, so, actually, we again face the result which is finite but\nundetermined at ǫ= 0, i.e. a removable singularity.\nAs usual within the finite temperature approach, we follow the Mats ubara formalism,\nthat is, we assume that the system stays in a thermal equilibrium at t he temperature\nT=1\nβ. Then, we carry out the Wick rotation and suggest that the time-lik e component\nof the moment is discrete, p4= 2πT(n+1\n2), withnis integer. In this case we can replace\n41the integral over p0by the sum over n. It is instructive to consider (3.48) separately for\nthe time-like and space-like bµ.\nWithin one manner we start with the expression (3.48), and, after a pplying the Mat-\nsubara formalism, promote spatial integral to ddimensions, and choose bµto be time-like.\nWe arrive at\nk4=b4T∞/summationdisplay\nn=−∞/integraldisplaydd/vector p\n(2π)d3ω2\nn−/vector p2\n(/vector p2+ω2n)3=b4\n4π2, (3.57)\nthat is, in this case we have C=1\n4π2which matches one of zero-temperature results\nobtained in [97]. If within (3.48) we suggest the vector bµto be space-like, we replace\npipj→δij/vector p2\nd, withd= 3+ǫis a spatial dimension, and arrive at\nki=biT∞/summationdisplay\nn=−∞/integraldisplaydd/vector p\n(2π)d4m2−ω2\nn+(4\nd−1)/vector p2\n(/vector p2+ω2\nn)3, (3.58)\nwhich is not ambiguous within this calculation, and the final result is\nki=bi\n4π2(1+F(ξ)), (3.59)\nwhereξ=m\n2πT, and\nF(ξ) = 2π2/integraldisplay∞\n|ξ|dz(z2−ξ2)1/2tanh(πz)\ncosh2(πz)(3.60)\nis a temperature depending function vanishing at T→0, i.e.ξ→ ∞, and tending to 1\natT→ ∞.\nAlternatively, we could first replace pµpν→1\n4ηµνp2in (3.48), and only afterward apply\nthe Matsubara formalism. In this case we would arrive at\nkµ=3bµ\n16π2(1+F(ξ)), (3.61)\nwhereF(ξ) is given by (3.60). We conclude that the CFJ term continues to be am biguous\nat non-zero temperature as well, so, the finite temperature is not sufficient to eliminate\nthe ambiguity of the result.\nIt is clear that this list of possible finite-temperature values for kµwe gave here is\nnot complete since other regularizations exist as well. Moreover, it is interesting to note\nthat only within some specific schemes one finds that all components of thekµvector\ncharacterizing the CFJ term are proportional to the original LV ve ctorbµwith the same\ntemperature dependent coefficient α, that is,kµ=α(T)bµ, as occurs, for example, in\nthe second scheme we used here, whereas in general case it is not s o. The fact that\nin generalkµis not proportional to bµ, clearly reflects the intrinsic Lorentz symmetry\n42breaking occurring in a finite temperature case due to the existenc e of the privileged\nreference frame of the thermal bath, which could break the prop ortionality between kµ\nandbµ. In [105] the strong asymmetry between spatial and temporal co mponents of the\nkµstemming from this breaking of the proportionality has been obtaine d with use of\nnon-covariant splitting the integration momentum, applied by the ru lepµ→ˆpµ+δµ0p0\nso that ˆp0= 0, with subsequent arising the ”covariant” and ”non-covariant” CFJ-like\ncontributions (this asymmetry has been further confirmed in [106 ]). Actually, a lot of\nmanners to calculate the CFJ term, starting from the Lagrangian ( 3.38), on the base of\ndifferent regularization schemes, including the finite temperature c ase, is known (see e.g.\n[106] and references therein).\n3.4.3 Non-Abelian generalization\nThe natural generalization of the CFJ term consists in its non-Abelia n generalization\nwhichcanbeintroducedbyanalogywiththewell-knownnon-AbelianCS term. Originally,\nthe non-Abelian extension of the CFJ term (2.6) has been proposed in [37]. To show that\nit could emerge as a perturbative correction, we start with the exp ression for the one-loop\neffective action of the gauge field (3.39), where, however, we sugg est that the gauge field\nis Lie-algebra valued, Aµ=AµaTa, whereTaare generators of some Lie group satisfying\nthe relations [ Ta,Tb] =ifabcTcand tr(TaTb) =δab. To find the non-Abelian CFJ term we\nshould, besides of the contribution involving two external gauge fie lds and one derivative\nacting on one of these fields, as in the subsection 3.4.1, consider also the contribution with\nthree external fields and no derivatives.\nThis scheme of calculations has been pursued in [38]. In this case, the corresponding\ngeneralization of the fermionic determinant (3.39) is\nΓ(1)=−iTrln/bracketleftBig\n(i∂ /−m+b /γ5)δij−gγµAa\nµ(Ta)ij/bracketrightBig\n. (3.62)\nAfter expansion of this determinant up to relevant orders and calc ulating traces and\nintegrals, the non-Abelian CFJ term was found to be\nSCFJ,non −Abelian=/integraldisplay\nd4xǫνλρµkµ(Aa\nν∂λAa\nρ−2\n3igfabcAa\nνAb\nλAc\nρ), (3.63)\nwith the constant vector kµis just the same one that can be read off from the expression\n(3.48). Repeating the discussion of the previous subsections, we c onclude that this vector\nis finite and ambiguous. Moreover, in the finite temperature case it b ehaves just in the\nsame way as in the Abelian theory (see the subsection 3.4.2).\nAnother way to obtain this term is based on the Schwinger proper tim e approach\n[104]: we consider the same expression for the one-loop effective ac tion (3.39), but with\n43the Lie-algebra valued vector field Aµ=AµaTa. After adding the field-independent factor\n−iTrln(i∂ /−m−b /γ5) in order to deal with the second-order operator of the standar d\nform✷+...which is the most convenient one within the proper time method, and\nexpanding the effective action up to the first order in the LV vector bµ, we obtain the\ndesired contribution to the effective action in the form\nΓ =−iTr(✷+igA /∂ /+mgA /+m2)−1[(gA /+2m)b /γ5−2i(b·∂)γ5].(3.64)\nNow, we employ the Schwinger representation by the rule\n(✷+igA /∂ /+mgA /+m2)−1=/integraldisplay∞\n0dse−s(✷+igA /∂ /+mgA /+m2), (3.65)\nand, after manipulations described in [104], we see that the divergen t part of this effective\naction completely vanishes, and the result identically reproduces th e expression (3.63),\nwithkµ=g2\n4π2bµ, this is one of the results obtained in [38]. Formally this result is\nambiguity-free, however, we note that the Schwinger represent ation itself plays the role\nof the specific regularization eliminating the ambiguity. We note never theless that the\nnumber of methods to generate the CFJ term, both in Abelian and no n-Abelian versions,\nis not exhausted by the approaches described here, since it can be generated from much\nmore sophisticated theories, including non-minimal ones, see f.e. [84 ].\nAnother interesting question is the study of topological propertie s of the non-Abelian\nCFJ term. In principle, one could consider the simplest topological st ructure, that is,\nto suggest the vector kµto be directed along the zaxis, therefore, for any value of z\ncoordinate one has a topological structure of the three-dimensio nal non-Abelian Chern-\nSimons theory, and the topological structure of the whole four-d imensional theory is\nsome foliation obtained as a direct sum of the three-dimensional top ological structures.\nHowever, this issue certainly requires further studies.\n3.5 CPT-even term for the gauge field\nThe CPT-even extension for the free gauge field has been originally s uggested [6] to look\nlike\nLeven=−1\n4κµνρσFµνFρσ, (3.66)\nwhereκµνρσis a constant tensor possessing the symmetries of the Riemann cur vature\ntensor. One of the most interesting particular cases is that one wh en this tensor is\nexpressed intermsofonlyoneconstant vector(orpseudovecto r)uµthroughtheexpression\n(2.8). In this case the term (3.66) takes the form:\nLeven=uµuρFµνFρν. (3.67)\n44Namely this expression was discussed in [39] within the context of high er dimensions.\nThe natural question evidently consists in the possibility for its pert urbative generation.\nCertainly it could be done in different ways.\nThe most straightforward way to obtain this term is based on use of the magnetic\n(non-minimal) coupling, namely this way was followed in [40]. The magnetic coupling has\nbeen introduced in [75] in the four-dimensional space in the form (2.1 6). We have showed\ninSection2.4that just thiscoupling ariseswithin thedual embedding p rocedureoftheLV\nself-dual model (see (2.35)), where the current is the usual spin or one, that is, jµ=¯ψγµψ.\nHowever, it is clear that, first, it would be interesting to consider th e theory involving\nnot only the non-minimal coupling only but also the usual interaction v ertexe¯ψγµψAµ,\nsecond, it is possible to obtain not only a purely non-minimal contribut ion to the aether\nterm but other ones as well. Although the non-minimal coupling is non- renormalizable,\nwe note that if we restrict our study by the one-loop order, or, as is the same, by the\nfermionic determinant, the non-renormalizability does not give any p roblems.\nTo demonstrate how the aether term can arise, let us consider the extended QED with\nthe magnetic coupling whose action, in the spinor sector, is [107]:\nS=/integraldisplay\nd4x¯ψ(i∂ /−m−eA /−gǫµνρσγµbνFρσ−b /γ5)ψ, (3.68)\ni.e., the role of uµused in (3.67) is played by the axial vector bµ. The corresponding\nfermionic determinant would be\nΓ(1)=−iTrln(i∂ /−eA /−gǫµνρσγµbνFρσ−m−b /γ5). (3.69)\nThere will be three different one-loop aether-like contributions in th is theory: the first\nof them is a purely non-minimal one, the second one is mixed, involving b oth minimal\nand non-minimal vertices, and the third one involves two minimal vert ices. Let us study\nall these cases.\nFirst, the purely non-minimal contribution is given by the Feynman dia gram depicted\nat Fig. 3.5.\n••\nFigure 3.5: Contributions to the aether term for the electromagne tic field with non-\nminimal vertices.\nThe corresponding analytic expression for this diagram looks like [40 ]:\nS2(p) =−g2\n2ǫµνρσǫµ′ν′ρ′σ′bµFνρbµ′Fν′ρ′/integraldisplayd4k\n(2π)41\n[k2−m2]2×\n45×tr[m2γσγσ′+kαkβγαγσγβγσ′]. (3.70)\nThere are several manners to calculate this expression. Here we d iscuss two of them,\nthose ones developed in [107]. To start, we evaluate the trace in the four-dimensional\nspace-time and obtain\nS2(p) =−2g2ǫµνρσǫµ′ν′ρ′σ′bµFνρbµ′Fν′ρ′/integraldisplayd4k\n(2π)41\n[k2−m2]2[m2ησσ′+\n+kαkβ(ηασηβσ′−ηαβησσ′+ηασ′ηβσ)]. (3.71)\nActually, the integral in this expression has just the structure giv en by (3.2), up to the\noverall factor. Now, let us perform its calculation.\nWithin the first manner, we replace kαkβ→1\n4ηαβk2, as it should be done in the four-\ndimensional space-time, and only after that, we introduce the dime nsional regularization,\npromoting the integral to d= 4−ǫdimensions. Carrying out Wick rotation, we arrive at\nthe Euclidean result\nS2(p) =−ig2ǫµνρσǫµ′ν′ρ′σ′bµFνρbµ′Fν′ρ′ησσ′/integraldisplayd4−ǫkE\n(2π)4−ǫk2\nE+2m2\n[k2\nE+m2]2. (3.72)\nIn this case we have the mutual cancellation of divergences, repro ducing the scenario\ndiscussed in the Section 3.1, so, our integral turns out to be finite:\n/integraldisplayd4−ǫkE\n(2π)4−ǫk2\nE+2m2\n[k2\nE+m2]2=m2(1−ǫ/2)\n(4π)2−ǫ/2/bracketleftbigg\nΓ(ǫ\n2)+Γ(−1+ǫ\n2)/bracketrightbigg\n=\n=−m2\n16π2+O(ǫ). (3.73)\nCollecting all together, multiplying two Levi-Civita symbols, disregard ing the Lorentz-\ninvariant term proportional to b2, in order to keep track of the aether term only, and\nreturning to the Minkowski space, we arrive at\nS2(p) =g2m2\n4π2(bµFµν)2. (3.74)\nWithin another manner, we first promote our integral to d= 4−ǫdimensions, then we\nreplacekαkβ→1\ndηαβk2. In this case, in the Euclidean space we have\nS2(p) =−ig2ǫµνρσǫµ′ν′ρ′σ′bµFνρbµ′Fν′ρ′ησσ′/integraldisplayddkE\n(2π)d2(d−2\nd)k2\nE+2m2\n[k2\nE+m2]2.(3.75)\nIf one substitutes d= 4−ǫ, withǫ∝ne}ationslash= 0 in this expression, the integral identically vanishes.\nSo, the purely non-minimal contribution displays a removable singular ity atǫ= 0. In\nprinciple, otherresultscanbeobtainedfortheexpression(3.70), seethediscussionin[108],\n46where also the finite temperature behavior of the generated aeth er term is considered. So,\nfor these two examples one can write this result as\nSmixed=C0g2m2(bµFµν)2, (3.76)\nwithC0is a finite ambiguous constant.\nWenotethatinotherspace-timedimensions onecanintroducethen on-minimal vertex\nas well, it will look like Smagn=g¯ψǫνρσbνFρσψin 3DandSmagn=g¯ψǫλµνρσσλµbνFρσψin\n5D. In both cases, the aether term will be finite within the dimensional r egularization\nframework [40].\nSecond, we study the contributions involving onenon-minimal verte x and one minimal\none, and one insertion b /γ5in one of the propagators [107]. In this case one can consider\njust the same Feynman diagrams given by Fig. 3.4, used in the Section 3.4.1 for the\ncalculation of the CFJ term, with the only difference that one of the e xternalAµfields in\nthis case is replaced by ǫµνρσbνFρσ, and the coupling eaccompanying just this field – by\ng. The structure of the integral over the internal momenta contin ues to be the same as\nin the case of the CFJ term, that is, (3.48). As a result, we repeat a ll calculations from\n[98] and arrive at\nSmixed=−2Ceg(bµFµν)2, (3.77)\nwith the finite constant Cis just that one characterizing the value of the CFJ term, it is\ndefined in (3.40), and its value is discussed in details in the Sec. 3.4.1. We note that this\ncontribution arises only in the four-dimensional space-time, where the vectorǫµνρσbνFρσ\ncan be defined, and has no analogues in other space-time dimensions .\nThird, there is a purely minimal aether-like contribution. In this case we have two\nb /γ5insertions (some preliminary studies of this contribution were also ca rried out in [83]\nwithin the renormalization group context), and the explicit form of t his contribution is\nSAA(p) =ie2\n2/integraldisplayd4l\n(2π)4(γµ1\nl /−mγν1\nl /+p /−mγ5b /1\nl /+p /−mγ5b /1\nl /+p /−m+\n+γµ1\nl /−mγ5b /1\nl /−mγν1\nl /+p /−mγ5b /1\nl /+p /−m+\n+γµ1\nl /−mγ5b /1\nl /−mγ5b /1\nl /−mγν1\nl /+p /−m)Aµ(−p)Aν(p), (3.78)\nwherepis an external momentum.\nThis contribution is a sum of Feynman diagrams given by Fig. 3.6.\nTaking into account only the terms of the second order in p, we get for the sum of\nthese graphs:\nSAA=−e2\n6m2π2bµFµνbλFλν. (3.79)\n47••• •\n• •\nFigure 3.6: Contributions to the aether term for the electromagne tic field with minimal\nvertices.\nWe note that this expression is superficially finite and hence ambiguity -free.\nSo, the complete aether-like contribution is a sum of (3.76,3.77,3.79), and it is finite\nand involves two ambiguous constants C0andC. While the constant Cis related with\nthe CFJ anomaly, the question of relation of C0with any possible anomaly is still open.\nThis calculation can be straightforwardly generalized for a non-Abe lian case, see [109].\nFor the presence of a minimal coupling only, the result for the aethe r term is obtained\nthrough expansion of the functional determinant (3.62) up to rele vant orders, and the\nresult is a straightforward generalization of (3.79):\nSAA=−e2\n6m2π2bµFµνabλFa\nλν, (3.80)\nwhereFa\nλν=∂µAa\nν−∂νAa\nµ−igfabcAb\nµAc\nνis the non-Abelian field strength. It is clear\nthat including of the non-Abelian analogue of magnetic coupling will yield non-Abelian\nanalogues of (3.76) and (3.77).\nBesides of the study performed in [40, 107], a very interesting disc ussion of ambiguity\nof the aether term for the gauge field is presented in [110]. In this p aper, it was argued\nthat the same ambiguous integral over momenta as in (3.71), that is ,\nIσσ′=/integraldisplayd4k\n(2π)41\n[k2−m2]2×\n×[m2ησσ′+kαkβ(ηασηβσ′−ηαβησσ′+ηασ′ηβσ)]. (3.81)\nwill accompany not only the aether term, but also the CFJ term and t he Proca term, so\nthat the whole contribution will look like\nΓ2=−2(eAσ+gǫσλµνbλFµν)Iσσ′(eAσ′+gǫσ′λ′µ′ν′bλ′Fµ′ν′), (3.82)\nand it is natural to choose the value of the Iσσ′to be zero in order to rule out the AµAµ\nterm, thus avoiding the breaking of the gauge symmetry. Neverth eless, since this integral\nisambiguous, nobodyforbidstochooseitsvaluestobedifferent for allthreecontributions,\nthat is, to bezero when it accompanies theProca term, andnon-ze ro when it accompanies\nthe gauge invariant aether and CFJ terms.\n48To close the discussion of generating the aether term from the ext ended QED La-\ngrangian given by (3.68), we note its advantage consists in the fact that only in this\ntheory one can arrive at finite CPT-even contributions. All anothe r couplings will yield\ndivergent aether-like results in the four-dimensional space-time. There are two most im-\nportant examples of obtaining these results. In the paper [111], th e starting point is the\nfollowing extension for the spinor sector of QED:\nS=/integraldisplay\nd4x¯ψ(i∂ /−eA /+λ\n2κµνρσσµνFρσ−m)ψ. (3.83)\nWe see that the LV coupling is CPT-even. The corresponding one-loo p effective action of\nthe gauge field, that is, the fermionic determinant, is given by the ex pression:\nΓ(1)[A] =−iTrln(i∂ /−eA /+λ\n2κµνρσσµνFρσ−m). (3.84)\nIn this case, it is easy to calculate the contributions of first and sec ond orders in κµνρσto\nthe two-point function of the gauge field. They diverge, explicitly loo king like [111]:\nI1=meλ\n8π2ǫκµναβFµνFαβ;\nI2=−λ2\n16π2ǫκµνρσκρσ\nαβFµνFαβ. (3.85)\nIt is clear that if the constant LV tensor κµναβis characterized by only one vector, looking\nlike (2.8), both these results reproduce the aether term (2.9). We note that, first, to\nsubtract the arising divergence in a consistent manner, one should introduce the aether\nterm in the tree-level action from the very beginning, second, for the dimensionless κ, the\ncouplingλhas a negative mass dimension, so, this theory is actually non-renor malizable\nand can be treated only as a kind of some low-energy effective model.\nAnother interesting LV extension of the QED is characterized by th e following action\nof fermions:\nS=/integraldisplay\nd4x¯ψ[i(∂ /+ieA /)+icµνγµ(∂ν+ieAν)−m]ψ. (3.86)\nHere, thecµνis a constant tensor which is usually assumed to be symmetric (actua lly,\nifcµνis antisymmetric, it can be ruled out through an appropriate field red efinition\n[112]). Since it is dimensionless, the corresponding theory is all-loop re normalizable. The\ncorresponding fermionic determinant is\nΓ(1)[A] =−iTrln[i(∂ /+ieA /)+icµνγµ(∂ν+ieAν)−m]. (3.87)\nExpanding it up to the second order in Aµ, for the simplest choice cµν=uµuνwe arrive\nat the following result [113]:\nS(2)\neff=−e2\n6π2ǫ/integraldisplay\nd4x1\n4/parenleftBig\n1+u2/parenrightBig−1˜Fµν˜Fµν, (3.88)\n49where\n˜Fµν= (gµα+uµuα)(gνβ+uνuβ)Fαβ, (3.89)\nso, we arrive at the divergent aether-like contribution, and again t he aether term must be\npresent in the tree-level action from the very beginning for the sa ke of the multiplicative\nrenormalizability of the theory. The analogues of this result for oth er choices of cµνcan\nbe obtained as well [113]. In the five-dimensional case, the corresp onding result is finite\nwithin the dimensional regularization.\n3.6 Higher-derivative LV terms\nThe next step of our study consists in generating the higher-deriv ative terms, attracting\ngreat attention due to highly nontrivial dispersion relations allowing f or birefringence of\nelectromagnetic waves, rotation of polarization plane in a vacuum (s ee section 2.2) and\nmany other interesting effects, a general review on wave propaga tion in higher-derivative\nLV extensions of QED can be found in [114] for the gauge sector, an d in [115] for the\nspinor sector, more results on higher-derivative LV extensions of QED and their relation\nto massive photons are discussed in [116]. As we already mentioned, t he first known\nexample of such terms is the Myers-Pospelov (MP) term (2.13) which is very convenient\nsince it allows to conciliate higher derivatives with unitarity, for certa in choices of the LV\nvector. Another important example is the higher-derivative CFJ-lik e term\nLhd\nCFJ=α\nM2ǫµνρσbµAν✷∂νAρ. (3.90)\nHereαis some dimensionless constant, and Mis a mass scale. Both the MP term and the\nhigher-derivative CFJ-like term can be generated as perturbative corrections within the\nsameLVextended QEDwhose actionisgiven by(3.68). Westartagain withthefermionic\ndeterminant (3.69). Expanding it in power series in bµwe find that the three-derivative\ntermisdescribedbythreecontributions, thatis, thoseoneswitht wonon-minimalvertices,\nwith one minimal vertex and one non-minimal one, and, finally, with two minimal ones\n[117].\nFor the first contribution, we have only one insertion in one of propa gators and there-\nfore consider again the Feynman diagrams given by the Fig. 3.4, so, w e reproduce iden-\ntically the situation described in the Section 3.4.1, with the only differen ce is that in our\ncase we should replace both external eAµlegs by external gǫµνρσbνFρσlegs, while the\nintegral over momenta is given again by the (3.42). After extractin g only the first order\nin external momenta which is the only relevant one for this fragment , we arrive at the\nsame integral over momenta as in (3.48), whose result, in a pure ana logy with the Section\n503.4.1, is\nΓ(1)\n1= 2g2Cǫνλµαbα˜Aν∂λ˜Aµ, (3.91)\nwhere˜Aµ=ǫµνρσbνFρσ, andCis the same finite ambiguous constant arising within calcu-\nlations of the CFJ term (see Section 3.4) and defined by (3.40). Carr ying out contractions\nof the Levi-Civita symbols, we arrive at\nΓ(1)\n1= 2g2C/bracketleftBig\nbαFαµ(b·∂)bβǫβµνλFνλ+b2bβǫβµνλAµ✷Fνλ/bracketrightBig\n. (3.92)\nWe see that this contribution is, first, finite and ambiguous (which re flects the fact that\nthe Adler-Bell-Jackiw (ABJ) anomaly can be extended to involve highe r-derivative terms\n[118]), second, involves boththe MP term (the first one in this expre ssion) and the higher-\nderivative CFJ-like term (the second one in this expression). We not e that here the last\nterm is of third order in the vector bµwhile in principle one can obtain the analogous term\nwithout the constant b2multiplier, that is, the one given by (3.90), using the one-loop\neffective action of the minimal LV QED (3.39) as a starting point [119]. I n that case the\nhigher-derivative CFJ-like contribution is finite and ambiguity-free.\nFor the second contribution, we consider Feynman diagrams with tw o insertions in\npropagators. The corresponding contributions are again depicte d at Fig. 3.6, and in this\ncase we repeat the calculation from the previous section and replac e only oneeAµleg by\nthe external gǫµνρσbνFρσleg in the expressions (3.78) and (3.79). As a consequence, we\narrive at the expression analogous to (3.79), where this replaceme nt of one external leg is\nperformed. Hence, the corresponding result can be written as\nSb2=eg\n6π2m2/integraldisplay\nd4x/bracketleftBig\nbαFαµ(b·∂)bβǫβµνλFνλ+b2bβǫβµνλAµ✷Fνλ/bracketrightBig\n. (3.93)\nAgain, we obtain both the MP and the higher-derivative CFJ-like term s. We note that\nthis result is non-ambiguous.\nFor the third contribution, we have three insertions in the propaga tors. The corre-\nsponding Feynman diagrams are depicted at Fig. 3.7.\n• ••• •\n•• ••\n• ••\nFigure 3.7: Higher-derivative contributions with three insertions.\nThe corresponding result looks like\nSb3=4e2\n45π2m4/bracketleftbigg\nbαFαµ(b·∂)bβǫβµνλFνλ+5\n4b2bβǫβµνλAµ✷Fνλ/bracketrightbigg\n. (3.94)\n51The sum of these three contributions (3.92,3.93,3.94) gives our final result for the higher-\nderivative contribution to the effective Lagrangian. It reads as\nShd=/parenleftBigg\n2g2C+eg\n6π2m2+4e2\n45π2m4/parenrightBigg\nbαFαµ(b·∂)bβǫβµνλFνλ\n+/parenleftBigg\n2g2C+eg\n6π2m2+e2\n9π2m4/parenrightBigg\nb2bβǫβµνλAµ✷Fνλ. (3.95)\nWe see that this result is finite and gauge invariant. Also, one can obs erve that for the\nlight-likebµ, the CFJ-like term vanishes, and only the MP term survives. The finit e\ntemperature behavior of this term is studied in [120].\nThere are other possibilities to generate the higher-derivative LV t erms. First of\nall, it is worth to mention the study based on using the third-rank ten sorgµνρ(it is\nantisymmetric with respect to two first indices), so that the followin g extension of the\nQED is introduced [121]:\nS=/integraldisplay\nd4x¯ψ[i(γµ+1\n2gκλµσκλ)(∂µ+ieAµ)−m]ψ. (3.96)\nWe note that this theory is renormalizable since gκλµis dimensionless. In this case, the\nfirst-order LV contribution is given by the sum of Feynman diagrams given by Fig. 3.8,\nwhere the dark ball is for gµνλinsertion.\n•\n•• •\nFigure 3.8: Higher-derivative contributions with gµνλinsertion.\nThe propagator in this theory is usual, so, one can straightforwar dly sum the con-\ntribution of these diagrams. As a result, we arrive at the following hig her-derivative\ncontribution [121]:\nΓg=e2\n12mπ2Aµ(gµνα✷∂α−gµρα∂ρ∂α∂ν−gρνα∂ρ∂α∂µ)Aν. (3.97)\nWe note that a priorithe only restriction for the gµνλis its antisymmetry with respect\nto the first two indices. Therefore we can present the tensor gµναas a sum of irreducible\n(totally antisymmetric and partially symmetric) parts as gµνα=ǫµναβgβ+ ¯gµνα, where\ngγ=−1\n6gµναǫµναγ. In this case the first term in (3.97) will yield the higher-derivative\nCFJ-like result (3.90); evidently, if gµνλis totally antisymmetric, only the first term of\n52(3.97) is non-trivial. It was claimed in [121] that other components of (3.97) can describe\nsome MP-like corrections.\nAlso, we have noted above that the higher-derivative CFJ-like term (3.90) can arise\nalready in the first order in bµ. The corresponding prescription has been discussed in\n[119]. Our starting point is the usual fermionic determinant (3.39), w hich, as we already\nnoted, correspondstothesimplest LVextension oftheQED.Then weapplythederivative\nexpansion approach just as in the section 3.4, but in this case we exp and our fermionic\ndeterminant up to the third derivative of the background field Aµkeeping at the same\ntime only the first order in bµ. Effectively it means that we should expand the self-energy\ntensor given by(3.42) upto thethirdorder inthe external pµ. The result will befiniteand\nambiguity-free. Its explicit form is given by (3.90), withα\nM2=e2\n24m2π2. In principle, since\nthe constant vector bµis small, this term will dominate in comparison with the CFJ-like\npart of (3.95) which we calculated above. The finite temperature be havior of this term is\nalso described in [119].\n3.7 Divergent contributions in gauge and spinor sec-\ntors of LV QED\nUp to now, we, being motivated by the concept of emergent dynamic s, paid our main\nattention to finite contributions to the one-loop effective action. T he key idea of our\napproach consists in suggesting that various LV terms in the gauge sector are generated\nas a consequence of integrating over a spinor field in some fundamen tal LV extension of\nQED, which can be done consistently only if the results are finite, oth erwise one must\nintroduce these terms from the very beginning, jeopardizing the c oncept. At the same\ntime, possible divergent contributions to the effective action for diff erent LV extensions of\nQEDalso must beconsidered, inorder to obtaina morecomplete pert urbative description\nof a corresponding theory. We note that study of divergent corr ections is especially\nimportant for non-minimal LV theories because of their non-renor malizability.\nThe first step in study of divergent LV contributions has been perf ormed already in\n[35], where the one-loop renormalization of the minimal LV extension o f QED (2.1) in\nthe first order in LV parameters was carried out. Further discuss ions of these first-order\nresults, including the methodology allowing to adapt the known Lehma nn-Symanzik-\nZimmermann (LSZ) formalism to LV theories, are presented in [122]. C learly, the natural\ncontinuationsofthisstudyare, first, calculatingdivergent corre ctionsofsecondandhigher\norders in LV constant vectors (tensors), second, obtaining qua ntum corrections in the\nspinor sector of corresponding theories, third, investigation of h igher loop LV corrections.\n53Let us briefly review the main results achieved along these directions .\nThe divergent corrections in effective action, involving second and h igher orders in LV\nparameters, were studied in a number of papers. In [40], the aeth er-like contribution to\nthe effective action in the spinor sector of the non-minimal LV QED wit h the action\nS=/integraldisplay\nd4x¯ψ(i∂ /−m−eA /−gǫµνρσγµbνFρσ)ψ,\nsimilar to (3.68), but without ¯ψb /γ5ψterm, was found to look like\nΓ(1)\nsp∝ig2m2\nπ2ǫ¯ψb /(b·∂)ψ, (3.98)\nwhich matches the form of the CPT-even aether term for the spino r fieldiuµuν¯ψγµ∂νψ\nproposed in [39]. It worth mentioning that in three- and five-dimens ional space-times\nthe analogous aether-like corrections arise as well, being finite in the se cases within the\nframework of the dimensional regularization. It is clear that the ae ther term for the\nspinor field will arise as well in the case of presence of the ¯ψb /γ5ψadditive term. The\ndivergent aether-like results were also shown to arise in the spinor s ector in [123] where\nthe contributions of the second order in LV parameters of the minim al LV extension of\nQED[35], givenby(2.1),wereobtained. Also, divergentcontribution sinthespinorsector,\nup to the first order in derivatives, and proportional to the first o rder of non-minimal LV\nparameters listed in [53], were obtained in the paper [124] and in some o ther papers.\nIn the gauge sector, besides of divergent results for the CPT-ev en LV term discussed\nin the section 3.5, the most interesting divergent contributions of t he second order in LV\nparameters were obtained in the paper [125], where the minimal LV ex tension of QED\n[35], given by (2.1), was studied, and the one-loop divergent correc tions were shown to\nhavetheaether-likeform(3.66). Itisworthtomentionalsohigher- derivative divergent LV\ncontributions obtained in [111, 126] for the CPT-even case (of fou rth order in derivatives),\nand in [127], for the CPT-odd case (of third order in derivatives).\nAs for higher-loop calculations, it should be noted that, up to now, t here are a few\nhigher-loop results for LV theories. The examples are, first, two- loop renormalization in\nthe LV scalar field theory involving the aether-like LV term (2.10), pe rformed in [128],\nsecond, discussion ofhigher-loopcontributions inthesimplest LVex tension ofQED(3.38)\npresented in[129], whereit wasarguedthatnoCFJtermcanbegene ratedinhigherloops.\nThere are other interesting discussions of perturbative aspects of LV theories and\nrelated papers, we discussed only part of results obtained up to no w. Nevertheless, still\nthere are many perturbative calculations to be performed, espec ially, those ones treating\nimpacts of various non-minimal vertices listed in [53].\n543.8 Discussion and conclusions\nWe discussed the possibilities of perturbative generation of differen t LV terms. Explicitly,\nwe saw that these terms arise within the context of the emergent d ynamics in a theory\ninvolving coupling of scalar, gauge and, as it will be discussed in the Cha pter 6, gravi-\ntational fields to some primordial spinor field which afterwards is bein g integrated out.\nHere we were mostly interested in the calculation schemes allowing to o btain essentially\nfinite results, to avoid problems with the renormalization. Indeed, in an opposite case\nthe corresponding terms in gauge/scalar sector should be introdu ced already at the tree\nlevel, to get the multiplicatively renormalizable theory, hence these t erms cannot be more\nconsidered as an emergent phenomenon.\nWe showed that it is possible to obtain both the new LV terms defined in many space-\ntime dimensions, as e.g. the aether terms for gauge and scalar fields , and the new terms\nwell defined only in specific dimensions, such as the CFJ-like term for s calar fields defined\nonly in two dimensions, the mixed scalar-vector term defined only in th ree dimensions,\nand the CFJ term defined only in four dimensions. We demonstrated t hat there are\nmanners to generate various terms suggested in [53], which appar ently shows that the\nperturbative mechanism is a consistent way to get new LV terms. It is interesting to note\nthat in certain cases the LV couplings nevertheless can yield non-ze ro Lorentz-invariant\nquantum corrections. Besides of simple arising the terms proportio nal tobµbµ, as occurs\ne.g. within studies of the aether terms for scalar and gauge fields [40 ], a very interesting\nexample is presented in [52], where the Lorentz-invariant axion term has been generated\nin the LV Abelian gauge theory from the triangle Feynman diagram simila r to that one\nused in this chapter to obtain the CFJ term.\nAn interesting feature of some of our results is their ambiguity. Act ually, we showed\nthat at the one-loop order there are two typical ambiguities, the fi rst of them, being a\nparadigmatic example, arises within the calculation of the CFJ term, a nd of the terms\nwhichcanbeobtainedfromtheCFJtermthroughreplacingexterna llegs,whilethesecond\none arises in the purely non-minimal sector of the LV extended QED. We indicated that,\nin principle, other ambiguous integrals can exist, and the ambiguity pe rsists in the finite\ntemperature case as well. It is interesting to note that while the firs t ambiguity is closely\nrelated with the Adler-Bell-Jackiw (ABJ) anomaly (for details, see [26 ]), the possible\nanomaly related to the second ambiguity is not known yet, and, more over, the problem\nof its existence is still open, see also the discussion in [118].\nBesides of finite terms, which are of special interest, providing a po ssible mechanism\nfor generating LV terms in the gauge sector, there are many diver gent LV corrections\nwhich arise in various LV extensions of QED. We reviewed briefly some r esults related to\n55obtaining these terms presented in a number of papers. Still, many s tudies of divergent\nLV terms, as well as of finite ones, are to be done. It is natural to e xpect that one of\nmain lines in these studies will consist in explicit calculations of perturba tive impacts of\nvarious LV terms listed in the review [53]. In this context, we can ment ion the paper [84]\nwhere perturbative generation of the CFJ term in the extended sp inor QED involving\nall dimension-5 LV operators was studied, with the result was shown to be finite and\nambiguous.\nWe close the discussion claiming that the perturbative generation sh ows itself to be\nan appropriate mechanism for obtaining at least some of the LV exte nsions of known\ntheories, especially, of the spinor electrodynamics. In the Chapte r 6, we demonstrate that\nthis methodology can be applied in the case of gravity as well.\n56Chapter 4\nLorentz symmetry breaking and\nspace-time noncommutativity\nAs it was claimed in [28], one of important motivations for the Lorentz s ymmetry break-\ning stems from the space-time noncommutativity. Indeed, the initia l statement for the\nnoncommutative field theory is the suggestion that the commutato r of two space-time\ncoordinates ˆ xµ,ˆxνis the constant antisymmetric matrix Θµν(see [11]):\n[ˆxµ,ˆxν] =iΘµν. (4.1)\nThis matrix is clearly not Lorentz-invariant, so, this form of the spa ce-time noncommu-\ntativity breaks the Lorentz symmetry. At the same time, some of v arious additive LV\nterms proposed in [34, 53] are based on second-rank constant te nsors, which can be, in\nmany cases, antisymmetric as well. Hence, an action of a classical no ncommutative field\ntheory expanded up to a first order in Θµν, can be naturally treated as a particular case of\ncertain models discussed in [34]. Following [11], this relation emerges fro m the low-energy\nlimit of the superstring action in the presence of the constant antis ymmetric tensor field\nBµν, so that the Θµνis related with Bµνthrough the algebraic relation:\nΘµν= 2πα′/parenleftBigg1\ng+2πα′B/parenrightBiggµν\nA, (4.2)\nwhere the subscript Ais for the antisymmetric part, and α′is a constant describing a\ncoupling of the string to the Bµν. Here, the space-time metric gµνis assumed to be\nthe Minkowski one. We note that this manner of introducing the Lor entz symmetry\nbreaking within the string context essentially differs from that one u sed in [2], based\non the spontaneous Lorentz symmetry breaking. Therefore, th e problem of systematic\ndescription of noncommutative field theories is based on the method ology distinct from\n57that one described in previous chapters, but, nevertheless, the re are some strong analogies\nbetween noncommutative field theories and ”usual” LV theories, wh ich will be discussed\nin this chapter.\n4.1 Formulations for noncommutativity\nAs it was argued in [11], a systematic study of field theories defined in a noncommutative\nspace-time can be done through an appropriate mapping of these t heories to the usual\nspace-time. There are two known manners to map theories formula ted in a noncommuta-\ntive space-time to theories formulated in the usual space-time, th at is, the Seiberg-Witten\n(SW) map [11] and the Moyal product [11, 130].\nThe Moyal product is a universal formulation for noncommutative fi eld theories al-\nlowing to define noncommutative extensions for any field theory exc ept of gravity, since\nin that case one should generalize this product to the curved space -time as well, and the\nconsistent manner of such a generalization is still unknown. To intro duce the Moyal prod-\nuct, one assumes [11], that any field φ(ˆx) in a noncommutative D-dimensional space-time\ncan be expressed in the form of the Fourier integral:\nφ(ˆx) =/integraldisplaydDk\n(2π)D˜φ(k)eikˆx, (4.3)\nwithkare usual momenta, and ˜φ(k) is a Fourier transform of this field, so, using the\nHausdorff formula, for the product of two noncommutative fields o ne has\nφ1(ˆx)φ2(ˆx) =/integraldisplaydDk1dDk2\n(2π)2Dφ1(k1)φ2(k2)eik1ˆxeik2ˆx=\n=/integraldisplaydDk1dDk2\n(2π)2Dφ1(k1)φ2(k2)ei(k1+k2)ˆxe−i\n2Θµνk1µk2ν. (4.4)\nIt is clear that this product is associative and can be straightforwa rdly generalized to an\narbitrary number of fields, with the only difference with the usual co mmutative theories\nconsistsinthepresence ofanadditionalfactorlike e−iΘµνk1µk2ν. Thisallowsustointroduce\na following rule: the product of fields on the noncommutative space- time must be mapped\ninto their Moyal product on the usual space-time defined as\nφ1(x)∗φ2(x)∗...∗φn(x) =/integraldisplaydDk1dDk2...dDkn\n(2π)nDφ1(k1)×\n×φ2(k2)...φn(kn)ei(k1+k2+...+kn)xexp[−i\n2Θµν/summationdisplay\ni=−i\n✷D−m2δ8(z1−z2); (5.10)\n<Φ(z1)Φ(z2)>=<¯Φ(z1)¯Φ(z2)>∗=−im\n✷D−m2(D2\n4✷D)δ8(z1−z2),\nwithD2,¯D2factors are associated with the vertices by the same rules as in the usual\nWess-Zumino model [149], that is, a vertex with an external chiral fi eld carries the factor\n−¯D2\n4, and with an external antichiral one – the factor −D2\n4.\nHere, some words about dispersion relations are in order. Disregar ding the factor1\n✷D\nwhich is really needed only to rewrite integrals over a chiral (antichira l) subspace as those\nones over a whole superspace, and does not correspond to new de grees of freedom, one\nfinds the physical spectra are completely described by the denomin ator✷D−m2, or, after\nthe Fourier transform, ˜ p2+m2, where ˜p2= (pµ+αuµuαpα)(pµ+αuµuβpβ) is a twisted\nscalar square of the momentum. We find that for different choices o fuµ, we have different\ndispersion relations:\n(i) For the time-like uµ= (1,0,0,0), one has E2(1−α)2=/vector p2+m2.\n(ii) For the space-like uµ= (0,1,0,0), one has E2=/vector p2+m2+(2α+α2)(/vector u·/vector p)2.\n(iii)Forthelight-like uµ= (1,1,0,0),and/vector palongthexaxis, onehas E=1\n1−2α(−2αp±/radicalBig\np2(1+2α+4α2)+m2.\nIt is clear that the dynamics in all these cases is consistent if αis enough small.\nAs an example of quantum calculation in this theory we present here t he computation\nof the one-loop low-energy effective action which is completely descr ibed by the K¨ ahlerian\neffective potential, by the definition depending only on the superfield s themselves but not\non their derivatives. We employ the methodology of summation over a n infinite number\nof Feynman supergraphs discussed in details in [151].\nInsupergraphs given by theFig. 5.1, theFeynman rulesaremodified : weincorporated\nthe mass into background fields Ψ = m+λΦ,¯Ψ =m+λ¯Φ denoted here by double lines,\n74✧✦★✥\n✧✦★✥\n✧✦★✥\n\u0000\u0000\u0000\u0000 ❅❅❅❅❅❅❅❅ \u0000\u0000\u0000\u0000\n...\nFigure 5.1: Contributions to the K¨ ahlerian effective potential.\nso, we have\n<Φ(z1)¯Φ(z2)>=−i\n✷Dδ8(z1−z2);\n<Φ(z1)Φ(z2)>=<¯Φ(z1)¯Φ(z2)>∗= 0. (5.11)\nThese Feynman supergraphs can be summed, just as it is done for u sual superfield\ntheories, yielding the following expression for the one-loop K¨ ahleria n effective potential\nK(1)=−i\n2∞/summationdisplay\nn=11\nn/integraldisplay\nd4θ/bracketleftBigg\nΨ¯ΨD2¯D2\n16✷2\nD/bracketrightBiggn\nδ(z−z′)|z=z′, (5.12)\nWe simplify this expression taking into account that theD2¯D2\n16✷Dis a projecting operator\neven in our LV case, so, [D2¯D2\n16✷D]n=D2¯D2\n16✷D. Then, we shrink the loop into a point through\ntheknown identityD2¯D2\n16δ(θ−θ′)|θ=θ′= 1 [149], and, after Fourier transform, remembering\nthe structure of the deformed d’Alembertian operator, we arrive at\nK(1)=i\n2/integraldisplay\nd4θ/integraldisplayd4q\n(2π)41\n(qµ+kµνqν)2ln/parenleftBigg\n1−¯ΨΨ\n(qρ+kρσqσ)2/parenrightBigg\n. (5.13)\nTo integrate, we make a change of variables by the rule qµ+kµνqν→˜qµ, which implies\narising the Jacobian of this replacement through the rule d4q= Ξd4˜q, with Ξ = det∂qν\n∂˜qµ=\ndet−1(δµ\nν+kµ\nν). This Jacobian is a constant, it does not depend on momenta. We no te\nthat if one suggests kµνto beantisymmetric andsmall, which does not imply any inconsis-\ntencies, this Jacobian reduces to 1. In the case of the effective po tential, the only impact\nof the Lorentz symmetry breaking consists just in the presence o f this Ξ multiplier. After\nthis change of variables and the Wick rotation, we have\nK(1)=−1\n2Ξ/integraldisplay\nd4θ/integraldisplayd4˜q\n(2π)41\n˜q2ln(1+¯ΨΨ\n˜q2). (5.14)\nNamely this expression (of course, except of the Ξ factor) arises within the analogous\ncalculations in Lorentz-invariant theories (see e.g. [151]), so, we imm ediately can write\ndown the result:\nK(1)=−1\n32π2Ξ/integraldisplay\nd4θΨ¯ΨlnΨ¯Ψ\nµ2. (5.15)\n75We see that the result differs from that one in the Lorentz-invarian t theory only by the\nJacobian factor. This allows us to formulate the following rule: for an y supersymmetric\ntheory whose superfield action does not involve explicit space-time d erivatives, within\nthe Kostelecky-Berger deformation of the supersymmetry algeb ra, the results of quantum\ncalculations can be obtained from the usual ones through replacem ent of all derivatives\n(bothspinor supercovariant andusual ones) bytheir ”twisted” a nalogues, andmultiplying\nof theL-loop expression by ΞL, with Ξ is the above-mentioned Jacobian. This rule has\nbeen also explicitly verified for the coupling of the chiral matter with a gauge superfield\n[152]. Moreover, the same rule is valid in the three-dimensional case a s well, where there\nis only one type of Grassmannian variables θα, and the supercovariant derivatives are\ngiven by\nDα=∂α+iθβγµ\nβα∇µ, (5.16)\nand the Dirac matrices for the three-dimensional space are the fo llowing 2 ×2 matrices:\n(γ0)α\nβ=−iσ2,(γ1)α\nβ=σ1,(γ2)α\nβ=σ3. One can see that since in the three-dimensional\nspace the same deformed d’Alembertian ✷Darises, the dispersion relations in three- and\nfour-dimensional superfield theories within this methodology are th e same.\nIn this context, some interesting observations for supersymmet ric Lorentz-breaking\nfield theories within this approach can be done. First of all, this formu lation possesses a\nnontrivial geometrical interpretation. Indeed, if we consider the three-dimensional theory\nof the scalar superfield Φ, with the action\nS=/integraldisplay\nd5z/parenleftBigg1\n2Φ(D2+m)Φ+λ\n3!Φ3/parenrightBigg\n(5.17)\nand calculate the lower ”fish” contribution to the two-point functio n, in the low-energy\nlimit we arrive at [152]\nΓ2=λ2Ξ\n48π|m|/integraldisplay\nd5zΦ(D2−2m)Φ. (5.18)\nProjecting this expression to components, we have:\nΓ2=−λ2\n48π|m|/integraldisplay\nd3xΞ(ηµν∇µφ∇νφ+...), (5.19)\nwhereφis a lower component of the scalar superfield, and dots are for othe r terms.\nIf we remind that ∇µ=∂µ+kµν∂ν, we can introduce a new ”inverse” metric gρσ=\nηµν(δρ\nµ+kρ\nµ)(δσ\nν+kσ\nν), and it is clear that Ξ = |det(gρσ)|1/2≡/radicalBig\n|g|. So, this quantum\ncorrection can be represented as\nΓ2=−λ2\n48π|m|/integraldisplay\nd3x/radicalBig\n|g|gµν∂µφ∂νφ+..., (5.20)\n76so, this result formally replays the action of the scalar field in a curve d space-time, thus,\nactually the Lorentz symmetry breaking generates a new geometr y, which, however, also\ncorresponds to a flat space since the new metric gµνis composed by constant components,\nso, our new geometry is affine. The question about possibility of a fur ther generalization\nof this concept in order to get a space with a nontrivial curvature is still open since if\nthe tensorkµνwill not be constant, the possible extension of the supersymmetry algebra\nevidently would be much more involved. For a four-dimensional theor y of a chiral scalar\nsuperfield, the same situation occurs as well with the only difference that the lower scalar\ncomponent of this superfield is complex. However, a Lorentz-brea king gauge superfield\ntheory apparently does not admit a geometrical interpretation, s ince there is no way to\nobtain the modified metric gµνcontracted with vector fields.\nAnotherobservationisthatifweconsiderthegeneralizationofthe freesupersymmetric\nQED with an action of a gauge superfield V:\nS=−1\n16/integraldisplay\nd8zVDα¯D2DαV, (5.21)\nits component form will be\nS=−1\n4d4x¯Fµν¯Fµν, (5.22)\nwhere¯Fµν=∇µAν−∇νAµis a deformed field strength of the electromagnetic field. It\nis invariant under the Abelian transformations δAµ=∇µξ, so, the gauge transformation\nitself in this case depends on the LV parameter. We note that, unlike the cases of scalar\nand spinor fields, this action strongly differs from the QED with an add itive aether term\n(2.9) which is invariant under usual gauge transformations δAµ=∂µξ. The one-loop\nK¨ ahlerian effective potential for a super-QED deformed in this way was calculated also\nin [152]. The generalization of these results for a non-Abelian superg auge theory can be\neasily developed.\n5.2 Lorentz symmetry breaking through introducing\nextra superfields\nOne more approach to implement Lorentz symmetry breaking in supe rfield theories is\nbased on introducing Lorentz breaking parameters through some (extra) superfields. Ini-\ntially this idea was proposed in [147], and it got further development in [ 153, 154].\nThe key idea of this approach is the following one. Let us consider the gauge superfield\nwhose component form, in the gauge sector, is (see e.g. [149]):\nV(x,θ,¯θ) =¯θ˙ασµ\n˙ααθαAµ(x)+..., (5.23)\n77whereAµistheusualvectorgaugefield, θα,¯θ˙αaretheGrassmanniancoordinates, anddots\nare for terms irrelevant for our purposes. The corresponding Ab elian superfield strength\nWα=−1\n4¯D2DαV, in a purely vector sector, has the following component structure :\nWα(x,θ,¯θ) = (σµν)β\nαθβFµν−i\n2(σµν)β\nα(σλ)β˙αθ2¯θ˙α∂λFµν+.... (5.24)\nThen, let us consider the chiral scalar superfield S(i.e.¯D˙αS= 0):\nS(x,θ,¯θ) =s(x)+..., (5.25)\nwiths(x) is a scalar, and the lower component of the corresponding antichir al field is\ns∗(x). We suggest the superfield Sto be purely external, displaying no dynamics.\nLet us extend the super-QED described by the usual supersymme tric Maxwell action\n(cf. [149])\nSMaxw=1\n4/integraldisplay\nd6zWαWα, (5.26)\nthrough adding the following term:\nSodd=/integraldisplay\nd8z(SWαDαV+¯S¯W˙α¯DαV). (5.27)\nIn components, one has (cf. [147]):\nSodd=/integraldisplay\nd4x/parenleftBig\n−1\n2(s+s∗)FµνFµν+i\n2∂µ(s−s∗)ǫµνλρFνλAρ/parenrightBig\n+.... (5.28)\nHere the dots are for other terms which do not contribute to the p urely gauge sector of\nthis theory. We can choose s(x) =−ikµxµ, and, consequently, s∗(x) =ikµxµ, withkµis\na constant vector. Therefore, s+s∗= 0, and we rest with\nSodd=/integraldisplay\nd4xǫµνλρkµFνλAρ+.... (5.29)\nAs a result, we see that the CFJ term is the only purely gauge contrib ution to the Sodd\n(actually, it turns out to be that the gauge invariance of Soddin the component form is\nmuch more transparent than in the superfield form). Further, in [ 153], the dispersion\nrelations for the theory with such an additive term were studied in diff erent sectors. It\nis interesting to notice that, up to now, namely this approach is the o nly one allowing\nto generate the supersymmetric extension of the CFJ term, since the Kostelecky-Berger\napproachdiscussedintheprevioussectionisessentiallyCPTeven, a ndtheapproachbased\non applying extra derivatives to superfields, which we discuss in the n ext section, can\nonly generate the higher-derivative terms for the “main” compone nt of the corresponding\n78superfield, that is, for the scalar swithin the chiral superfield and the vector Aµwithin\nthe real one.\nFurther, the CPT-even term can also be introduced within this meth od. It was shown\nin [154] that, if we for the same superfields WαandSas above (and the conjugated ones)\nconsider the additive term\nSeven=/integraldisplay\nd8z(DαS)Wα(¯D˙α¯S)¯W˙α, (5.30)\nafter projecting this action to components we arrive at [154]\nSeven=/integraldisplay\nd4x/bracketleftBig\n−4FµνFµν∂λs∂λs∗−8FµλFλ\nν(∂µs∂νs∗+∂µs∗∂νs)/bracketrightBig\n.(5.31)\nFors(x) =ikµxµands∗(x) =−ikµxµ, we arrive at\nSeven=/integraldisplay\nd4x/bracketleftBig\n−4k2FµνFµν−16kµkνFµλFλ\nν/bracketrightBig\n, (5.32)\nwhere the first term replays the Maxwell term with the extra multiplie r depending on\nthe square of the LV vector, and the second one – the CPT-even c ontribution introduced\nin [40]. We see that the aether-like structure arises naturally within t his method. In\n[154], some applications of this term were studied. It is natural to ex pect that within\nthis approach some other LV terms can arise as well. Also, a natural problem consists in\nsearching for the possibility of emerging of such a term as a quantum correction from an\nappropriate coupling. Up to now this scenario never was demonstra ted.\n5.3 Straightforward Lorentz symmetry breaking in a\nsuperfield action\nThe third manner to break the Lorentz symmetry in superfield theo ries is based on\nstraightforward adding terms proportional to constant LV vect ors (tensors) to the su-\nperfield action, it was firstly proposed in [148]. The simplest way to do it consists in\nconsideration of additive terms like\nδS=/integraldisplay\nd8zkµν∂µΦ∂ν¯Φ, (5.33)\nwith the analogous extensions are possible as well for chiral Lagran gians (some examples\nwill be given further) and for terms depending on gauge superfields . Nevertheless, the\nexamples for LV extensions of superfield theories we consider here differ from those ones\nproposed in [148], moreover, unlike that paper, here we concentra te on calculating the\none-loop effective potential. It is clear that within this approach the supersymmetric LV\n79terms essentially involve higher derivatives, e.g. for the expression above, one gets terms\nlike/integraltextd4xkµν∂µφ✷∂νφ. However, for an appropriate choice of the LV parameters, in this\ncase it isk00=k0i=ki0= 0, one can avoid the presence of the higher time derivatives,\nand, hence, eliminate the ghosts, so, in a certain sense in this case w e can speak about an\nattempt of supersymmetric extension of Horava-Lifshitz-like the ories introduced in [4].\nOur first example will be the theory proposed in [155]:\nS=/integraldisplay\nd8zΦ(1+ρ∆z−1)¯Φ+(/integraldisplay\nd6zW(Φ)+h.c.), (5.34)\nwithρissome constant characterizing theenergy scale at which higher de rivatives become\nrelevant, andthenumber z≥2(typically, integer), inanalogywith[4], iscalledthecritical\nexponent. It is clear that for z= 2, the additive LVtermreproduces theexpression (5.33),\nwithk00=k0i=ki0= 0 andkij=−ρδij. We note that for any z, this LV term is CPT-\neven. One can consider W(Φ) =m\n2Φ2+λ\n3!Φ3as in the Wess-Zumino model, but it is not\nmandatory.\nOne can obtain the component structure of this action:\nS=/integraldisplay\nd4x/bracketleftBig¯φ✷(1+ρ∆z−1)φ−i¯ψ˙ασm\n˙αα∂m(1+ρ∆z−1)ψα+ (5.35)\n+¯F(1+ρ∆z−1)F−/parenleftBiggm\n2(φF+1\n2ψαψα)+λ\n2(φ2F+1\n2φψαψα+h.c./parenrightBigg/bracketrightBigg\n.\nWe see that in the scalar sector, one has two time derivatives and, m aximally, 2zspatial\nones, so, the critical exponent for the scalar component is equal toz. At the same time,\nthe critical exponent for the spinor field is 2 z−1, i.e. the critical exponents for different\ncomponents of the superfield do not coincide.\nThe propagators in this theory look like\n<Φ(z1)¯Φ(z2)>=i1+ρ∆z−1\n✷(1+ρ∆z−1)2−m2δ8(z1−z2); (5.36)\n<Φ(z1)Φ(z2)>=−im\n✷(1+ρ∆z−1)2−m2(−D2\n4✷)δ8(z1−z2).\nAs usual, Φ( z1)Φ(z2)>=<¯Φ(z1)¯Φ(z2)>∗. It is not difficult to show that for z≥2 the\ntheory is finite.\nTo find the K¨ ahlerian effective potential, we proceed just as in the S ec. 5.1: consider\nthe same sequence of the supergraphs given by Fig. 5.1, again we inc orporate the mass\ninto the background field Ψ = −W′′, and modify the propagators to be\n<Φ(z1)¯Φ(z2)>=i\n✷(1+ρ∆z−1)δ8(z1−z2);\n<Φ(z1)Φ(z2)>= 0. (5.37)\n80Doing the D-algebra transformations, Wick rotation and summation as in the Sec. 5.1\ntogether with the replacement ρ→ρ(−1)z−1, we arrive at\nK(1)=−1\n2/integraldisplay\nd4θ/integraldisplaydk0Ed3/vectork\n(2π)41\nk2ln/bracketleftBigg\n1+Ψ¯Ψ\nk2(1+ρ(/vectork2)z−1)2/bracketrightBigg\n. (5.38)\nUnfortunately, this integral cannot be evaluated exactly. Using t he scheme of approxi-\nmated computations based on disregarding of subleading orders in /vectork2, we find [155]:\nK(1)=1\n12πcsc(π\nz)(4ρ\n3)−1/z/integraldisplay\nd4θ(Ψ¯Ψ)1/z. (5.39)\nThis expression displays singularity at z→1 as it should be, since the Lorentz-invariant\ncase, where the theory displays a divergence, corresponds name ly to this value of the\ncritical exponent.\nIf we introduce the similar deformation into the chiral effective pote ntial, we have the\naction\nS=/integraldisplay\nd8zΦ¯Φ+/bracketleftBigg/integraldisplay\nd6z(1\n2Φ(m+a(−∆)z)Φ+λ\n3!Φ3)+h.c./bracketrightBigg\n. (5.40)\nProceeding in a similar way as above, after the Wick rotation we have t he one-loop\nK¨ ahlerian effective potential in the form\nK(1)=−1\n2/integraldisplay\nd4θ/integraldisplaydk0Ed3/vectork\n(2π)41\nk2ln\n1+(Ψ+a/vectork2z)(¯Ψ+a/vectork2z)\nk2\n0E+/vectork2\n, (5.41)\nand, under the same approximate manner, we get\nK(1)=−1\n8πa1/zcsc(π\n2z)/integraldisplay\nd4θ(Ψ¯Ψ)1/2z. (5.42)\nIn this case the singularity occurs at z= 1/2. However, this is not unusual since namely\nforz= 1/2 in this theory spatial and time derivatives in UV leading terms enter t he\ntheory in the same order making the behaviour of loop integrals to be similar to the\nLorentz-invariant case. Coupling of the chiral matter to gauge fie lds yields the analogous\nresults.\nA slightly different approach within this line is based on use of Myers-Po spelov-like\nterms (see Section 2.1) where some derivatives are contracted to constant LV vectors\n(tensors) whose specific choice allows to avoid arising the higher time derivatives. In this\ncase we start with the following action [156]:\nS=/integraldisplay\nd8z[Φ(1−1\nΛ2(n·∂)2)¯Φ+φ¯φ]+\n+/bracketleftbigg/integraldisplay\nd6z(M\n2Φ2+1\n2λΦφ2+1\n2fφΦ2)+h.c./bracketrightbigg\n. (5.43)\n81Here theφis a light (massless) chiral superfield which we treat as a purely backg round\none, and Φ is a heavy superfield which we assume to be a purely quantu m one. Here\nMis a large mass, and Λ is an energy scale at which the higher derivatives become\nimportant (from phenomenological estimations [156], it is reasonable to choose Λ to be\nthe Planck mass, and Mto be a characteristic string mass, so, M≃10−2Λ). Thenµis a\ndimensionless LV vector, in the Euclidean space nµnµ= 1. After summation of the same\ngraphs depicted at Fig. 5.1, we find the one-loop K¨ ahlerian effective potential to be\nK(1)=−1\n2/integraldisplay\nd4θ/integraldisplayd4k\n(2π)41\nk2ln\nk2/parenleftBigg\n1+(n·k)2\nΛ2/parenrightBigg2\n+|Ψ|2\n. (5.44)\nThis integral can be calculated approximately. Since, as we already s aid,M≃10−2Λ, we\ncan expand this expression in series in M/Λ and 1/M. As a result, we find\nK(1)=λ2\n32π2/integraldisplay\nd4θφ¯φ[3+lnM2\n4Λ2]+..., (5.45)\nwhere dots are for suppressed terms. We see that in this case the contributions arising\ndue to the Lorentz symmetry breaking are significant, so, we obse rve the effect of large\nquantum corrections. However, this is rather natural since the L V term here effectively\nplays the role of the higher-derivative regularization. The same situ ation occurs within\nthe first approach described in this section. Indeed, if we compare this result with (5.39)\nwith rescaling ρ=α\nΛ2z−2, in order to have a dimensionless αand a scale Λ, we see that\nthe result (5.39) is proportional to (M\nΛ)2\nz−2, thus, for z≥2 the quantum correction in\nthis case is also not suppressed, being large instead of this. To close this section, we\nconclude that within this “straightforward” approach, both the H orava-Lifshitz-like and,\nfor a space-like nµ, the Myers-Pospelov-like theories display no ghosts being therefo re\nperturbatively consistent.\n5.4 Conclusions\nNow, let us compare the results obtained within different approache s aimed to introduce\nLorentz symmetry breaking in superfield theories. First of all, it is int eresting to note\nthat all these ways allow for introducing CPT-even terms into classic al actions, and,\nfurther, into quantum corrections, whereas only the way based o n introducing of the\nextra superfield [147] can yield CPT-odd expressions, that is, first of all, the CFJ term.\nUp to now, no other manner to construct a supersymmetric exten sion of the CFJ term is\nknown.\nSecond, a nontrivial fact consists in a possibility to generate a nont rivial geometry\nfrom quantum corrections. A discussion on this fact is presented a lso in [157] where it\n82is noted that some LV modifications of Dirac matrices can be defined a s well. However,\nan open question is whether one can, starting from some LV superfi eld gauge theory,\nobtain the action of the gauge field coupled to the new geometry con sistently, so that not\nonly derivatives, but also vector component fields would be contrac ted to a new metric.\nAnother possible line of studies could consist in a generalization of this approach for the\ncase of a space-time with a non-zero curvature through suggest ing the tensor kµνto be\na fixed function of space-time coordinates rather than a constan t. However, in this case\nthe deformed supersymmetry algebra will clearly be much more comp licated, and other\ndifficulties, intrinsic for studies of Lorentz symmetry breaking in a cu rved space-time,\nwhich will be discussed in the next chapter of the book, will arise as we ll.\nThird, the problem of consistent supersymmetric extension of Hor ava-Lifshitz-like the-\nories continues to be open. The reason is that usual superfield mod els always involve\nlower (second) spatial derivatives in a kinetic term, therefore, on e cannot arrive at a the-\nory involving only terms with two time derivatives and just 2 zspatial derivatives, while\nimplementing of higher spatial derivatives into a spinor supercovaria nt derivative clearly\nbreaks the Leibnitz rule. Some attempts to proceed in this case are presented in [150],\nhowever, this methodology is still under development.\nIt must be mentioned that there is one more manner to implement Lor entz symmetry\nbreaking specific for supersymmetric field theories, that is, the fe rmionic noncommuta-\ntivity [158]. Following this approach, the anticommutation relations be tween fermionic\ncoordinates are deformed as {θα,θβ}= Σαβwhere Σαβis a constant symmetric matrix\nclearly breaking the Lorentz symmetry (e.g., in a three-dimensional space-time this ma-\ntrix is equivalent to the constant vector Σµ=1\n2(γµ)αβΣαβ). However, this methodology\nis known to meet its own difficulties. We close this section with a conclusio n that the\nproblem of construction of a consistent supersymmetric extensio n for LV theories is still\nopen.\n8384Chapter 6\nLorentz and CPT symmetry\nbreaking in gravity\nIn this chapter we describe some first steps in study of CPT and/or Lorentz breaking\nextensions of gravity. Implementation of the Lorentz symmetry b reaking within the grav-\nity context is certainly one of the most complicated issues within gene ral studies of LV\ntheories, facing many difficulties. The most important reason for th ese difficulties is the\nfollowing one. In the curved space-time, the symmetry group is tha t one of general co-\nordinate transformations like xµ=xµ(x′), which, at the same time, represents itself as\nthe extension both of the Lorentz group and of the gauge group. As we noted in previ-\nous chapters, in many cases the Lorentz symmetry breaking implies the CPT symmetry\nbreaking as well, whereas the gauge symmetry is not broken, the pa radigmatic example\nis the CFJ term. Therefore, it is natural to require for the most int eresting CPT, and\nin certain cases Lorentz breaking extensions of gravity to be cons istent with the general\ncoordinate (that is, gauge) invariance, and there is only a few know n examples where this\nconsistency is possible. The most important of such examples, is the four-dimensional\nChern-Simons modified gravity which we discuss in this chapter. Anot her possible ap-\nproach could consist in consideration of the weak (linearized) gravit y limit, where we\nconsider only the dynamics of the symmetric metric fluctuation tens orhµν, and we can\napply the methods similar to those ones used for studies of LV exten sions of QED. In this\nchapter we consider both these approaches within the Riemannian f ramework.\nBesides of this, we note that introducing Lorentz symmetry break ing in a curved\nspace-time faces one difficulty more. As we noted in previous chapte rs, in a flat space-\ntime the Lorentz symmetry can be violated explicitly through introdu cing new terms\nproportional to constant vectors (tensors) which cannot be int roduced consistently in\na non-zero curvature case. Indeed, let us consider for example a constant vector kµ.\n85In the flat space-time it satisfies the condition ∂νkµ= 0, but this condition evidently\nbreaks general covariance. A straightforward covariant gener alization of this condition,\nlooking like ∇νkµ= 0, imposes additional restrictions on space-time geometry (so-c alled\nno-go constraints), which are hard to satisfy in general (see the discussion in [159]). And\nif no such conditions are imposed, one faces an infinite tower of term s proportional to\n∇ν1...∇νnkµwhich makes all calculations to beextremely complicated (some lower t erms\nof this form are discussed in [160], where renormalization of LV QED in a curved space-\ntime is performed). Therefore, the most promising way of breaking Lorentz symmetry in\nthe presence of gravity is based on spontaneous Lorentz symmet ry breaking, instead of\nthe explicit one. The Einstein-aether and bumblebee models allowing to break Lorentz\nsymmetry spontaneously on a curved background, will also be discu ssed in this chapter.\n6.1 Motivations for 4Dgravitational Chern-Simons\nterm\nWe start our study of LV modifications in gravity with the paradigmat ic example of\nthe additive CPT-, and in certain cases Lorentz-breaking term, co nsistent at the same\ntime with the gauge symmetry, that is, the four-dimensional gravit ational CS term. It\nrepresents itself as a natural generalization of the three-dimens ional Lorentz-invariant\ngravitational CS term introduced in [36] and looking like\nSCS=1\n2µκ2/integraldisplay\nd3xǫµνα/bracketleftBig\n(∂µωνab)ωab\nα+2\n3ωc\nµaωνcbωab\nα/bracketrightBig\n, (6.1)\nwhereµis a mass,κ2is a gravitational constant (of mass dimension −1, in 3D), andωab\nµ\nis a connection. Here and further in this section, we follow notations adopted in [104], so,\nthe Greek indices are for the curved space-time, while the Latinindic es arefor the tangent\none. As our geometry is assumed to be Riemannian, the connection c an be expressed in\nterms of vielbeins or metrics through well-known expressions. One c an straightforwardly\nverify that the action (6.1) is gauge invariant. We note that there is no/radicalBig\n|g|factor in this\nintegral since the invariant Levi-Civita tensor in a curved space isǫµνα√\n|g|rather than ǫµνα.\nIntroducing a constant (pseudo)vector bµ, it is easy to generalize the CS action to\nthe four-dimensional case. From the formal viewpoint, this gener alization is similar to\npromoting the usual CS term to the CFJ term through the replacem entǫµνλ→bρǫρµνλ\ncorrespondingtoaddingofonemoredimension. Asaresult, treatin gωab\nµasaRiemannian\nconnection, wearriveatthe4 DCSaction(wenotethatinfourdimensions, no µmultiplier\n86is needed):\nSCS=/integraldisplay\nd4xǫρµναbρ/bracketleftBig\n(∂µωνab)ωab\nα+2\n3ωc\nµaωb\nνcωa\nαb/bracketrightBig\n. (6.2)\nOriginally, this term has been introduced in [32], and in this section we r eview some\nresults obtained in [32]. To prove gauge invariance of this term, that is, its invariance\nunder general coordinatetransformations, onecan writedown it s equivalent form: indeed,\nwe can define the vector\nKρ=ǫρµνα/bracketleftBig\n(∂µωνab)ωab\nα+2\n3ωc\nµaωb\nνcωa\nαb/bracketrightBig\n, (6.3)\nso that∂µKµ=1\n2∗RR, where\n∗RR=1\n2ǫαβγδRµν\nαβRγδµν (6.4)\nis a pseudoscalar Pontryagin density. So, suggesting that the vec torbµis expressed as\nbµ=∂µϑ, whereϑis a pseudoscalar called the CS coefficient, which, for the first step, is\ntreated as an external function, one can rewrite the 4 Dgravitational CS term as\nSCS=1\n4/integraldisplay\nd4xϑ∗RR. (6.5)\nWe see immediately that this term is both Lorentz and gauge invariant , but parity-\nbreaking. If we require the ϑto be linear in coordinates, just as we did in the section 3.3\nduring the study of three-dimensional CS terms, that is, ϑ=bµxµ, withbµbe constant,\nafter a simple integration by parts we arrive at the particular Loren tz-breaking CS term\n(6.2). We note that this is the lower gauge-invariant CPT-Lorentz b reaking term in\ngravity consistent withthegaugeinvariance requirement, andoth er possible CPT-Lorentz\nbreaking gauge-invariant terms for the gravity would necessarily in volve higher orders in\nderivatives.\nThe complete action of the CS modified gravity involving both the usua l Einstein-\nHilbert term and the CS term, looks like [32]:\nSCSMG=1\n16πG/integraldisplay\nd4x(/radicalBig\n|g|R−1\n2bµKµ). (6.6)\nComparing the theory (6.6) and the LV electrodynamics with the CFJ term (2.3), we can\nconclude that the 4 Dgravitational CS term is related with the gravitational anomalies\noriginallyintroduced in[27], without anyconcerning oftheLorentzsy mmetry breaking, in\nthe manner similar to that one relating the CFJ term with the Adler-Be ll-Jackiw anomaly\n(see the discussion of this correspondence in [26]).\n87The equations of motion for this theory also were obtained originally in [32], they have\nthe form\nGµν+Cµν=−8πGTµν, (6.7)\nwhereTµνis the energy-momentum tensor of a matter, Gµν=Rµν−1\n2Rgµνis the usual\nEinstein tensor, and Cµνis a Cotton tensor defined as\nCµν=−1\n2/radicalBig\n|g|ǫσµαβ/bracketleftBig\nbσ∇αRν\nβ+bστRτν\nαβ/bracketrightBig\n+(µ↔ν). (6.8)\nHere∇αis a covariant derivative, and bµ=∇µϑ,bµν=∇µ∇νϑare constructed on the\nbase of the CS coefficient.\nWe can find the divergence of the modified Einstein equations (6.7). I n the r.h.s. we\nobtain zero due to the conservation of the energy-momentum ten sor of the matter, then,\nthe identity ∇µGµν= 0 continues to be valid because of the Bianchi identities. Finally\nwe arrive at\n∇µCµν=1\n8/radicalBig\n|g|bν∗RR. (6.9)\nThis is a constraint for the possible solutions of Eq. (6.7). In many re levant cases, e.g.\nspherically symmetric metrics, one has∗RR= 0, so, the Cotton tensor will be conserved\n(see e.g. [161] for a general discussion).\nOne can suggest the CS coefficient to be not an external object bu t a dynamical field.\nIn this case one starts with the action of the dynamical CS modified g ravity [161]:\nSDCSMG=1\n16πG/integraldisplay\nd4x(/radicalBig\n|g|R+1\n4ϑ∗RR+1\n2/radicalBig\n|g|∇µϑ∇µϑ). (6.10)\nIn this case, one finds an additional contribution to the energy-mo mentum tensor of the\nmatter, that is, the energy-momentum tensor of the scalar field ϑ.\nIt remains to write down the linearized form of the gravitational CS t erm. Suggesting\nthat the metric looks like gµν=ηµν+hµν, wherehµνis a small metric fluctuation, and\nrelabelingbµbyvµ, we have [32, 162]:\nSCS=1\n4/integraldisplay\nd4xhµνvλǫαµλρ∂ρ(✷hα\nν−∂ν∂γhγα). (6.11)\nWe see that the gravitational CS term involves higher derivatives. H owever, it was proved\nin [32] that for physical degrees of freedom there is no higher time d erivatives in the\ncorresponding equation of motion, hence, neither unitarity nor ca usality are violated.\nSo,letusdiscusssomeaspectsrelatingtothisterm,namely, itsper turbativegeneration\nand the ambiguity of the result.\n886.2 Perturbative generation of the gravitational CS\nterm\nNow, letuspresentsomeschemestogeneratethegravitationalC Sterminfourdimensions.\nThefirst ofthese schemes hasbeenproposedin[162]. Thestarting pointofthecalculation\nis the following classical action of the spinor field ψon a curved background:\nS=/integraldisplay\nd4xeeµ\na¯ψ/parenleftBig1\n2iγa↔\n∂µ+1\n4iωµcdΓacd−bµγaγ5−m/parenrightBig\nψ. (6.12)\nHere Γacd=1\n6γaγcγd+...is the antisymmetrized product of three Dirac matrices. The\ncorresponding one-loop effective action of the gravitational field c an be cast as\nΓ(1)=−iTrln(i∂ /−m−b /γ5+ω /). (6.13)\nTo consider the weak field approximation, we write gµν=ηµν+hµν, so, the vielbein is\neµ\na=δµ\na+1\n2hµ\na, ande= 1 +1\n2ha\na(again, we note that Greek indices are for the curved\nspace-time, and Latin ones – for the tangent one). So, omitting th e irrelevant terms, in\nparticular, those ones proportional to hµ\nµ, we get the action of ψcoupled to hµν:\nS=/integraldisplay\nd4x¯ψ/parenleftBig1\n2iΓµ↔\n∂µ+hµνΓµν−bµγµγ5−m/parenrightBig\nψ, (6.14)\nwhere Γµ=γµ−1\n2hµνγν, and Γµν=1\n2bµγνγ5−i\n16(∂ρhαβ)ηβνΓρµα.\nThe nontrivial contributions to the two-point function of the metr ic fluctuation hµν\nare given by the Feynman diagrams depicted at Fig. 3.4, the only differ ence is that now\nthe external lines are for the metric fluctuation. It is clear that th e quartic vertex and\nthat one involving the bµhµνcontraction evidently will not yield CS-like results.\nThese Feynman diagrams are superficially divergent. However, aft er long and very\ninvolved calculations described in [162], adopting a special calculation s cheme based on\n’t Hooft-Veltman prescription [163], we find that all divergences can cel out, and, in the\nzero mass limit, where possible divergent contributions of the above mentioned Feynman\ndiagrams vanish, arrive at the finite result\nSCS=1\n192π2/integraldisplay\nd4xhµνbλǫαµλρ∂ρ(✷hα\nν−∂ν∂γhγα), (6.15)\nthat is, the vector vµ(6.11) is expressed as vµ=1\n48π2bµ. So, this result appears to be\nunique, at least within this approach. However, this is not sufficient t o conclude whether\nthe gravitational CS term we obtained is indeed unambiguous in gener al. First of all, we\nshould emphasize that the ’t Hooft-Veltman prescription adopted w ithin [162] is nothing\nmore that a fixing of the regularization scheme so that in this case on e obtains a definite\n89result. Besides, as we observed in the Chapter 3, the finiteness of a superficially divergent\nresult typically signalizes about its ambiguity.\nTo demonstrate that our result is indeed ambiguous, we use anothe r manner of calcu-\nlating the gravitational CS term based on the proper-time method w hich was for the first\ntime carried out in [104]. In this case, we can consider the complete ac tion of the spinor\non a curved background (6.12), thus avoiding use of the weak field a pproximation, so, we\ncan obtain the full-fledged CS term.\nTo compute the CS term, we can put e≃1. This approximation is sufficient for ob-\ntaining the gravitational CS term since all higher terms in expansion o fewill be irrelevant\nfor our calculations. Then, in order to complete the operator whos e trace we study, up\nto the quadratic one, as it is required by the proper-time technique (cf. Section 3.3), we\nadd to the one-loop effective action (6.13) the constant of the for m\nC0=−iTrln(i∂ /+m+b /γ5). (6.16)\nAs a result, after multiplication of determinants with use of the fact that detAB=\ndetAdetB, we arrive at the following expression for the one-loop effective act ion [104]:\nΓ(1)=−iTrln/bracketleftBig\n−✷+iω /∂ /+mω−m2+(ω /−2m)b /γ5+\n+ 2i(b·∂)γ5−b2/bracketrightBig\n. (6.17)\nNow, we expand this expression up to the first order in the LV vecto rbµand employ the\nSchwinger proper-time representation T−1=/integraltext∞\n0dse−sT.\nAfterward, the divergent parts easily can be found to cancel. The n, the finite contri-\nbution of the second order in connections ωlooks like\nS(2)\nfin= tr/integraldisplay\nd4x/integraldisplay∞\n0dse−sm2/bracketleftBig\ns2m2ω /(∂ /ω /)b /γ5+2m2s3ω /∂ /(∂αω /)∂αb /γ5\n+2m2s3ω /(∂αω /)∂α∂ /b /γ5/bracketrightBig\ne−s✷δ(x−x′)|x′=x. (6.18)\nTo perform calculations, we employ the following representation of t he geodesic bi-scalar\nσ(x,x′) [164]:\nδ(x−x′) =/integraldisplayd4k\n(2π)4eikα∇ασ(x,x′). (6.19)\nCalculating the trace (see details in [104]), we arrive at\nS(2)\nCS=i\n4/integraldisplay\nd4x/integraldisplay∞\n0dse−sm2bµωνab∂λωρcd/integraldisplayd4k\n(2π)4esk2s2m2ǫµνλρ(gacgbd−gadgbc)\n=1\n32π2/integraldisplay\nd4xǫµνλρbµ∂νωλabωρba, (6.20)\n90withǫµνλρ=eeµaeνbeλceρdǫabcd.\nThe relevant term of third order in connections ωis\nS(3)\nfin=itr/integraldisplay\nd4x/integraldisplay∞\n0dse−sm2/bracketleftBiggs2\n2m2ω /ω /ω /b /γ5+\n+s3\n3m2(ω /∇/ω /∇/ω /+ω /∇/ω /ω /∇/+ω /ω /∇/ω /∇/)b /γ5 (6.21)\n−s3\n3m2(ω /∇/ω /ω /+ω /ω /∇/ω /+ω /ω /ω /∇/)(b·∇)γ5−s3\n3m4ω /ω /ω /b /γ5/bracketrightBigg\n×\n×e−s✷δ(x−x′)|x′=x.\nAfter calculating the trace we arrive at\nS(3)\nCS=−i\n16/integraldisplay\nd4x/integraldisplay∞\n0dse−sm2bµωνabωλcdωρef/integraldisplayd4k\n(2π)4esk2/parenleftBiggs2\n2m2ǫµνλρ+\n+s3\n3m2ǫµνλρk2+s3\n3m2ǫανλρkαkµ−s3\n3m4ǫµνλρ/parenrightBigg\n×\n×/bracketleftBig\ngaf(gbcgde−gbdgce)+gae(gbdgcf−gbcgdf)+\n+gad(gbfgce−gbegcf)+gac(gbegdf−gbfgde)/bracketrightBig\n.\nBy integrating over the momenta kand the Schwinger parameter s, we find\nS(3)\nCS=−1\n48π2/integraldisplay\nd4xǫµνλρbµωνabωλbcωρca, (6.22)\nso that, summing this expression with (6.20), we find that our comple te result yields the\ngravitational CS term [32]:\nSCS=1\n32π2/integraldisplay\nd4xǫµνλρbµ/parenleftbigg\n∂νωλabωρba−2\n3ωνabωλbcωρca/parenrightbigg\n. (6.23)\nIt is easy to verify that in the weak field approximation this term does not reproduce\nthe value of the numerical coefficient presented in (6.15). We conclu de that the four-\ndimensional gravitational CS term is ambiguous. Unlike the approach developed in [162],\nin this case there is no need to use the zero mass limit to obtain our res ult.\nAt the same time, using the methodology based on transformations of the functional\nintegral, developed in [102], we can show that the gravitational CS te rm (6.23) is accom-\npanied by a completely undetermined constant C:\nSCS=C/integraldisplay\nd4xǫµνλρbµ/parenleftbigg\n∂νωλabωρba−2\n3ωνabωλbcωρca/parenrightbigg\n. (6.24)\nThis constant arises from the modification of a CS conserved curre nt which is totally\narbitrary [102]. Some more details of this calculation can be found in [1 65]. At the same\n91time, it has been argued in [95, 96], that the result C= 0, i.e. vanishing of the one-loop\ncontribution to the gravitational CS term, is preferable in a certain sense. Following\nthese arguments, if one assumes bµto be a dynamical field, instead of a given vector or a\ngradient of a given scalar, the gravitational CS term loses its gauge invariance, therefore,\nthe consistent manner of calculations should imply a zero result for it . As we noted in\nthe section 3.4.1, the CFJ term displays the analogous behaviour.\nAnother interesting discussion of ambiguity of the four-dimensiona l gravitational CS\nterm has been presented in [166]. In this paper, the methodology o f the implicit regu-\nlarization [100] is applied to generation of the gravitational CS term, where the starting\npoint is the action of the spinors on a weak gravity background (6.14 ) taken in the zero\nmass case, m= 0. Within this scheme of calculations, the self-energy tensor Πµναβis\nfound to be (see [166] for details):\nΠµναβ=−i\n8ǫλρβµbλpρ/bracketleftBig\n(i\n48π2−64σ0−4v0+4ξ0)pαpν− (6.25)\n−(i\n48π2+32σ0)ηανp2/bracketrightBig\n+(α↔β)+(µ↔ν)+(α↔β,µ↔ν).\nHereσ0,v0andξ0are finite but yet undetermined parameters of implicit regularization .\nThe next step consists in imposing the requirement of transversalit y for this self-energy\ntensor,pµΠµναβ= 0, which implies a requirement for these parameters to satisfy the\nrelationξ0−v0= 24σ0, so, one has\nΠµναβ=−i\n24ǫλρβµbλpρ(i\n16π2+96σ0)[pαpν−ηανp2]+ (6.26)\n+ (α↔β)+(µ↔ν)+(α↔β,µ↔ν).\nTherefore, we recover the explicit formof thegravitational CSte rm given by (6.11), where\nvµ= (1\n24π2−64iσ0)bµ. We see that the result unavoidably depends on an arbitrary regu-\nlarization parameter σ0, thus, this calculation is consistent with the above-mentioned fact\nthat the constant accompanying the gravitational CS term is comp letely undetermined,\ncf. [165].\nTo close the discussion of the four-dimensional gravitational CS te rm, we note again\nthat this term is related with the gravitational anomaly [27] just in t he same manner as\nthe CFJ term is related with the Adler-Bell-Jackiw anomaly (see [26] an d the Section 3.4\nof this book).\n926.3 ProblemofspontaneousLorentzsymmetry break-\ning in gravity\nAs we noted in the beginning of this chapter, constant vectors (te nsors) used to construct\nLV terms in a flat space-time, in general cannot be defined for a cur ved one – in [159],\npossible restrictions for geometry arising from requirement for su ch vectors and tensors\nto be constant, called no-go constraints, have been discussed in g reat details, and it was\nargued that in general, these restrictions cannot be satisfied, pe rhaps except of some fine-\ntuned situations. Moreover, these restrictions typically contrad ict to Bianchi identities\nand to conservation of the energy-momentum tensor. Hence, th e explicit Lorentz sym-\nmetry breaking, in general, does not appear to be an appropriate m echanism in gravity\n(nevertheless, some interesting results have been obtained within this approach as well,\nsee f.e. [167] where conservation laws for nondynamical backgrou nds were studied, and\n[168] where some schemes allowing to maintain Bianchi identities in the e xplicit Lorentz\nsymmetry breaking case were proposed), and one must employ the spontaneous Lorentz\nsymmetry breaking mechanism, within which, the vector (tensor) in troducing a privileged\nspace-time direction, is not required to be constant.\nSo, let us discuss models used to break the Lorentz symmetry spon taneously in a\ncurved space-time. There are two known theories considered in th is context, the Einstein-\naether gravity and the bumblebee gravity.\nLet us discuss first the Einstein-aether gravity. Originally, it was int roduced in [169].\nThe action of this theory is [170]:\nS=−1\n16πG/integraldisplay\nd4x/radicalBig\n|g|/bracketleftBig\nR+λ(uµuµ−1)+Kαβ\nµν∇αuµ∇βuν/bracketrightBig\n, (6.27)\nwhereuµis a dynamical vector satisfying the constraint uµuµ= 1, so, choosing of a\npossible value of this vector breaks the Lorentz symmetry sponta neously, the λis the\nLagrange multiplier field, the tensor Kαβ\nµνlooks like\nKαβ\nµν=c1gαβgµν+c2δα\nµδβ\nν+c3δα\nνδβ\nµ+c4uαuβgµν, (6.28)\nandc1,c2,c3,c4are constants.\nThe corresponding equations of motion are:\ngαβuαuβ= 1;∇αJα\nµ−c4˙uα∇µuα=λuµ;\nTαβ=−1\n2gαβLu+∇µ/parenleftBig\nJα\n(µuβ)−Jµ\n(αuβ)−J(αβ)uµ/parenrightBig\n+ (6.29)\n+c1[(∇µuα)(∇µuν)−(∇αumu)(∇βuµ)]+c4˙uα˙uβ+[uν∇µJµν−c4˙u2]uαuβ.\nHere ˙uµ=uα∇αuµ,Jα\nµ=Kαβ\nµν∇βuν, andLuisu-dependent part of the Lagrangian.\n93Consistency of some known metrics, in particular, cosmological and black hole ones, in\nthis theory has been verified in a number of papers (see e.g. [171] an d references therein).\nIt is clear that this theory can be extended through adding differen t terms where the\nvarious degrees of the uµvector are coupled to different fields, for example, scalar, spinor\nand gauge ones, for example, it is possible to implement aether terms discussed in the\nChapter 3, like that one for the gauge field, uαFαµuβFβµ. In the flat space limit, the uµ\nvectors can (but are not restricted to) be constant ones.\nIn the Einstein-aether theory, the spontaneous Lorentz symme try breaking is intro-\nducedwithuseoftheconstraint. Atthesametime, itisknownthatp resenceofconstraints\nmakes perturbative studies of such theories more complicated req uiring special method-\nologies like 1 /Nexpansion, which, moreover, cannot be applied in our case as the uµfield\npossesses only four components. The bumblebee gravity, where t he spontaneous Lorentz\nsymmetry breaking isimplemented in other manner, that is, through choosing a minimum\nfor some potential, proposed in [34], is free of this problem. The actio n of the bumblebee\ngravity looks like\nS=−/integraldisplay\nd4x/radicalBig\n|g|/parenleftbigg1\n16πG(R+ξBµBνRµν)−1\n4BµνBµν−V(BµBµ±b2)/parenrightbigg\n.(6.30)\nHereVisthepotential, typically onecanuse V=λ\n4(BµBµ±b2)2, andb2>0 isaconstant.\nTheBµν=∂µBν−∂νBµis the stress tensor, and ξis a coupling constant characterizing\nthe magnitude of the non-minimal coupling between the bumblebee fie ld and the Ricci\ntensor.\nLet usbriefly review themost importantresults obtainedwithinstud ies ofthebumble-\nbee gravity. As we have already noted, the main direction of resear ch in various modified\ngravity theories, including the bumblebee gravity, is study of consis tency of known results\nfound within general relativity, especially, cosmological solutions an d black holes, within\nthe new modified gravity. First we discuss the solutions originally trea ted within the\nbumblebee context in [172]. In this case, the action of the model is giv en by (6.30). In\nthis theory, the static spherically symmetric metric was considered and shown to imply\na modification of the Schwarzschild solution so that its 00 and 11 comp onents behave as\n−g00=g−1\n11= 1−2GLm\nr1−L, withGLis a modified gravitational constant, L≃b2\n0/2, and\nthe dimensionless b0characterizes the radial component of the vector implementing th e\nspontaneous Lorentz symmetry breaking. Another important so lution is the G¨ odel one,\nwhich, as it is known, closed timelike curves (CTCs) can arise. In the b umblebee gravity\nthis metric was shown to be consistent for certain vacua [173], so, t his Lorentz-breaking\nscenario does not exclude CTCs. The third solution usually tested fo r various modified\ngravity models is the Friedmann-Robertson-Walker cosmological me tric, and it has been\n94proved in[174] thatwithin thebumblebee gravity, cosmic accelerat ion, including late-time\nde Sitter expansion, can occur.\n6.4 Possible LV terms in gravity\nThe four-dimensional gravitational CS term we have considered ab ove is certainly a\nparadigmatic example of the CPT-Lorentz breaking term in gravity. Nevertheless, other\nLV additive terms in gravity, although perhaps less advantageous, can be considered as\nwell. The list of possible terms with dimensions up to 8 can be found in [15 9]. Let us\nbriefly characterize their general features.\nAs an example of a Lorentz-breaking extension for the Einstein gra vity, we can discuss\nthe aether-like term for the gravity [39] in a space-time of an arbit rary dimension D,\nlooking like\nSgrav\naether=α/integraldisplay\ndDx/radicalBig\n|g|uµuνRµν, (6.31)\nwithαis a small constant. This term is CPT-even. Actually, this is a particula r\nform of the CPT-even term Seven=/integraltextd4x/radicalBig\n|g|sµνRµν, whose different aspects, includ-\ning Hamiltonian formulation and cosmological issues, were studied in va rious papers, see\nf.e. [175, 176, 177]. As we noted in the previous section, to avoid the explicit breaking\nof the diffeomorphism invariance, one should suggest that the vect oruµis not a constant\nbut, instead of this, arises as a result of the spontaneous Lorent z symmetry breaking\noccurring for some potential. As an example, we can choose the bum blebee-like one, see\ne.g. [34]:\nV(uµ) =λn(uµuµ±v2)n(6.32)\nwherev2is some positive constant, so, we should add this potential to the ac tion (6.31),\ntherefore, the complete action would be\nS=MD−4\n∗/integraldisplay\ndDx/radicalBig\n|g|(1\nκR+αuµuνRµν+Lkin\nbumb[u]+λn(uµuµ±v2)n).(6.33)\nHere,Lkin\nbumb[u] =−1\n4Fµν[u]Fµν[u] is the Maxwell kinetic term for the bumblebee field,\ndenoted here as uµinstead ofBµused in the previous section, and the M∗is a parameter\nwith the mass dimension 1 introduced to ensure a correct dimension o f the action. We\nnote that this action essentially differs from the Enstein-aether th eory (6.27) discussed\nin the previous section. The term (6.31), proposed originally in [39] to couple the aether\nterm with gravity is, first, more convenient for studies within the we ak gravity framework\nsince it allows, in particular, to treat the vector uµasconstant one, second, is more similar\n95with other aether-like terms defined in [39, 40] for scalar, spinor an d gauge fields. One can\ncheck that the gauge symmetry in this theory can be maintained for special restrictions\nongµν(fot the weak gravity case, hµν) anduµ. For example, if we consider the five-\ndimensional space-time choosing a vector uµdirected along the extra (fifth) dimension,\nandremindthat,fortheweakgravity, thegaugetransformation sforthemetricfluctuation\nand the LV vector are δhµν=∂µξν+∂νξµandδuµ=∂5ξν, we can impose restrictions\nha5= 0 (witha= 0,1,2,3) and∂5ξa= 0 [39]. We note that for this term, as well as for\nmany other Lorentz and/or CPT-breaking terms in gravity propos ed in [34, 160], with\nthe only known exception in a full-fledged gravity is the gravitational CS term, the gauge\ninvariance requires imposing of special conditions (no-go constrain ts), e.g., the terms like\ntµνλρRµνλρare in general not gauge invariant by the same reasons as (6.31). M oreover,\nthe term of the first order in derivatives, being generated as a qua ntum correction in\nthe linearized theory described by the action (6.14), is essentially div ergent [145]. The\nsame is valid for terms linear in Riemann and Ricci tensors in a full-fledge d gravity with\nLV terms (i.e. the terms of second order in derivatives of hµνin the weak gravity case,\ncf. [160]. This emphasizes the importance of the gravitational CS te rm which, as we\nnoted above, is one-loop finite. In principle, one can abandon as well the restriction for\nthe geometry to be Riemannian, through introduction of the torsio n, but up to now,\nneither perturbative generation of any LV term involving torsion no r studies of impact of\nCPT-Lorentz breaking terms involving torsion were carried out.\nBesides of the aether-like term (6.31), other terms involving couplin gs of the LV vector\nuµwith various geometrical objects can be introduced, such as, e.g., sµνRµν,tµνλρCµνλρ,\nwithsµνandtµνλρare constructed on the base of the uµ, andCµνλρis the Weyl tensor,\nsee f.e. [178]. In that paper, the lower contributions from these te rms to the parametrized\npost-newtonian (PPN) expansion are found, and classical dynamic s of the vector field in\nthis theory is studied in [179].\nThese terms are the examples of dimension-4 LV terms in the gravita tional sector. A\nlist of possible LV terms in the gravitational sector with dimensions up to 8 is presented\nin [159]. Among most important terms, we can emphasize the following o nes: first, the\ngravitational Chern-Simons term, second, the terms proportion al to various orders of\ncovariant derivatives of the Riemann tensor, like e.g. (⌣\nk(n)\nD)αβγδµ 1...µnD(µ1...Dµn)Rαβγδ,\nthird, those ones proportional to direct products of various deg rees of Riemann tensor,\nlike e.g. (⌣\nk(6)\nR)α1β1γ1δ1α2β2γ2δ2Rα1β1γ1δ1Rα2β2γ2δ2, fourth, those ones involving various non-\nzero degrees both of the Riemann tensor and of its covariant deriv atives. Except of the\ngravitational Chern-Simons term, all these terms, being consider ed in the weak field limit,\nwith the LV vectores (tensors) assumed to be constant, break t he gauge symmetry (the\n96importance of gauge symmetry breaking within the gravity context has been discussed in\n[180]). The dispersion relations for linearized gravity models whose ac tion is given by the\nsum of the Einstein-Hilbert action and each of these terms, for a sim plified case when the\nconstant tensors, e.g. (⌣\nk(n)\nD)αβγδµ 1...µn, are completely characterized by only one constant\nvector, have been studied in [181], and were shown to display a usual formE2=/vector p2, being\nunaffected by the Lorentz-breaking vectors. Studies of dispers ion relations for a more\ngeneric form of Lorentz-breaking tensors are presented in [182].\nTo illustrate the problem with the gauge (non)invariance we turn aga in to the weak\ngravity limit, where the dynamical field is the metric fluctuation, and c onsider the term\ninitially introduced in [145]:\nLlingrav=−2ǫλµνρbρhνσ∂λhσ\nµ. (6.34)\nThe main motivation to define this term in [145] was the generalization o f the noncommu-\ntative field method discussed in the section 4.4 and based on deforma tion of the canonical\ncommutation algebra between fields and their canonically conjugate d momenta applied in\n[82] for the gauge fields, to the case of the (linearized) gravity, an d namely this structure\nrests when we impose a special gauge allowing to rule out the non-loca l contributions,\nwhose presence is a price we should pay for the gauge symmetry. If we consider the usual\ngauge transformations for this term, we will see that it is not invaria nt. Among other\nproperties of this term we should emphasize as well the birefringenc e of the gravitational\nwaves. Also, if one would try to generate this term from the usual t heory of spinors on\na curved background (6.14), the result will be divergent, which, ho wever, is natural since\nthe theory of spinors on a curved background (6.12) is non-renor malizable.\nNevertheless, many gauge invariant Lorentz-breaking terms in th e linearized gravity\ncase are possible as well. Besides of the gravitational Chern-Simons term discussed in the\nsection 6.2, there are other gauge invariant CPT-odd LV terms and CPT-even LV terms\npresented in [181]. However, all these terms are higher-derivative ones. Some examples\nare\nLeven=1\n2KαΠαβKβ, (6.35)\nwithKα=bµΠµνhνα, where Πµν=ηµν✷−∂µ∂νis the projection-like operator, and\nLodd=ǫαβγδbαKβ∂γKδ. (6.36)\nThe dispersion relations in both these cases are E2=/vector p2, being unaffected by the Lorentz-\nbreaking vectors. Other higher-derivative gauge invariant LV ter ms for the linearized\ngravity can be constructed as well, they involve more derivatives (e .g. one can consider\n97linear combinations of terms ǫαβγδbαKβ∂γ✷nKδ,1\n2KαΠαβ✷nKβfor various n≥1, being\nstraightforward generalizations of above terms). However, it is c lear that more studies of\nhigher-derivatives LV additive terms in gravity are still to be done.\n6.5 Conclusions\nIn this chapter, we described some first steps carried out to cons ider the CPT-Lorentz\nbreaking extensions in the gravity. We demonstrated that the fou r-dimensional gravita-\ntional CS term is apparently the best possible CPT-Lorentz breakin g extension of the\nEinstein gravity, since it, first, can be perturbatively generated in a consistent manner\nbeing finite, second, does not display any problems with unitarity and /or causality, third,\nis gauge invariant. Clearly, it calls the interest to studying of possible classical solutions\nof the CS modified gravity, that is, first of all, verification of consist ency of known general\nrelativity solutions within the CSMG, either in dynamical or non-dynam ical cases.\nForthefirst time, thisstudy was carriedout alreadyintheseminal w ork [32] where the\nconsistency of the Schwarzschild metric within the CSMG was proved . As a continuation,\nin [161] it was proved that the spherically symmetric solutions of gene ral relativity, whose\nmost important example are clearly the cosmological solutions, and t he static solutions\nwith axial symmetry, as well as their generalizations, so-called stat ionary solutions where\nthe metric depends on time only through the factor t−αφ, withαis some constant,\ncontinue to be consistent within the CSMG, where the Kerr metric is n o more consistent.\nFurther, it was shown in [183] that to achieve the consistency within the DCSMG, the\nKerr metric should be deformed by additive terms proportional to t he CS coefficient ϑ,\nwith the desired consistency can be shown order by order. A great review on different\nexact solutions in CSMG is presented in [33]. Other solutions, which do not match\nthese examples, are the G¨ odel-type solutions characterized by t he possibility of the closed\ntime-like curves [184, 185]. Their consistency within the CSMG was ch ecked both in\nnon-dynamical and dynamical cases in [186, 187], where it was shown that, in the CSMG,\nboth in non-dynamical and dynamical cases, besides of the known G ¨ odel-type solutions,\nthe new solutions, e.g., completely causal ones, can arise.\nHowever, it is clear that the problem of possible manners to implement the Lorentz\nsymmetry breaking within the gravity, and, further, to study the resulting theories at\nthe quantum level, is still open. As we already noted, one important is sue is the gauge\nsymmetry, that is, the general covariance, whose maintaining in an expected LV extension\nof a full-fledged gravity requires special efforts. For example, as w e already noted above,\none of the main difficulties is related to the development of an appropr iate manner for\n98introduction of privileged directions in a curved space-time without b reaking the general\ncovariance; it was claimed in [172] that a consistent manner to do it is b ased on a spon-\ntaneous Lorentz symmetry breaking introduced within the bumbleb ee model in a curved\nspace-time, however, up to now, only some first studies in this direc tion were carried\nout. Another problem is the search for the renormalizable model of the gravity itself.\nCertainly, constructing of Lorentz-breaking non-Riemannian gra vity models, which could\ninvolve torsion and nonmetricity, as well as development of LV massiv e gravity theories,\nis also an important direction of studies (some first steps along this w ay are discussed in\n[188, 189, 176]). Therefore we conclude that constructing of con sistent LV extension of\ngravity certainly requires more active studies.\n99100Chapter 7\nExperimental studies of the Lorentz\nsymmetry breaking\nIt isclear thatall new theories shouldbeverified throughsomeexpe riments. So, this book\nwould be clearly incomplete without a review of different experimental studies of possible\nLV phenomena and measurements of LV parameters characterizin g various LV extensions\nof known field theory models. Therefore, in this last chapter we pre sent this review, in\nwhich we discuss both macroscopic studies of LV effects, especially g ravitational ones,\nand the microscopic studies, especially quantum ones. Also, one of t he main reasons for\nexperimental studies of the Lorentz symmetry breaking is the det ermination of possible\nlimits of applicability for special relativity, which, as well as those ones for any physical\ntheory, require a careful study. Besides of this, there are stro ng cosmological motivations\nfor these studies, such as, e.g., the hypothetic possibility for cosm ic radiation anisotropy\ndiscussed in [190] where it was called the ”axis of evil”. While in [191] it was argued that\nit is more reasonable to attribute this anisotropy to inappropriate m ethods of statistical\nanalysis rather than to fundamental physical effects, the possib ility of this anisotropy\ncertainly requires further studies.\nOne more (and historically, the first) effect treated as a possible ma nifestation of the\nLorentz symmetry breaking was the Greisen-Zatsepin-Kuzmin (GZ K) effect [192, 193]\nwhich showed that the flux of the cosmic rays strongly decreases w ith the energy more\nthan about 1020eV, which naturally called the interest to the idea of deformed disper sion\nrelations which essentially involve a certain energy scale as it is sugges ted within the\ndouble special relativity [194] (see also the discussion in the section 2 .2 of this book\nand [56]). Other important effects within this context are the possib le birefringence of\na light in a vacuum and rotation of a plane of polarization of light in a vacu um which\ncan occur in certain LV extensions of electrodynamics (see discuss ions in Section 2.2)\n101and the Cherenkov radiation which in the LV case is possible even in a va cuum, see\nf.e. [195, 196] (various aspects of Cherenkov radiation in different LV extensions of QED\nare discussed also, e.g. in [41, 197, 198, 199], see also references t herein, in [200], the\nCherenkov radiation for massive photons is treated, and in [201], it is studied for LV\nfermions).\nThe most appropriate theory to treat all these issues related with Lorentz symmetry\nbreaking, is, clearly, the minimal LV Standard Model extension (LV S ME) [5, 6] which\nhas been discussed throughout this book and includes as ingredient s almost all terms\nconsidered here – the CFJ term (both in Abelian and non-Abelian form s), the aether\nterm for scalar, gauge and spinor fields, and different LV spinor-sc alar and spinor-vector\ncouplings. In [34], this model was generalized to include the gravity. T his model clearly\nallows various manners for experimental measurements of differen t LV effects. Treating\nnumerical estimations of the parameters of LV SME, we, first of all, should refer to the\nData Tables [1] which collect the results for estimated values of all LV parameters known\nup to now. Certainly, non-minimal extensions of this model discusse d in [53, 159] also\nmust be studied experimentally.\nWe note that, within verifying the Lorentz symmetry breaking and/ or measuring its\nparameters, two directions of studies can be emphasized: (i) non- gravitational studies,\nsuch as, first, the determinations of unusual behavior of electro magnetic waves which\ncan originate from the Lorentz symmetry breaking, such as birefr ingence and rotation\nof the plane of polarization, second, the possible modifications of dis persion relations,\nthird, measurements of the parameters of the Lorentz-violating extension of the Standard\nModel; (ii) gravitational and cosmological studies.\n7.1 Non-gravitational studies\nOne of the most known experiments aimed to study possible unusual behavior of elec-\ntromagnetic waves in a vacuum is the PVLAS experiment (see e.g. [202 ]) in which the\nrotation of plane of polarization of light in an external magnetic field w as measured. In\nthe paper [203] it was claimed that this rotation can be attributed to space-time noncom-\nmutativity, which, as we noted in the Chapter 4, represents itself a s one of the known\nforms of Lorentz symmetry breaking, with the noncommutativity p arameter is estimated\nas Θ≃(30GeV)−2. Nevertheless, further [204] it was argued that this rotation sho uld\nbe attributed to the axion-photon coupling, whereas the impact of the noncommutativity\nmust be unobservable.\nAnother important line of experimental studies for the possible Lor entz symmetry\n102breaking is based on studies of cosmic rays. Indeed, as we already n oted, one of inter-\npretations of the GZK effect was based on the idea that it can be exp lained by a strong\nLorentz symmetry breaking at some energy scale, as is suggested by the double special\nrelativity, therefore the dispersion relations should be modified by e xtra terms becoming\nimportant at very large energies. In the paper [205], one of the firs t reviews of studies\nof cosmic rays within this context was presented. Following [205], the ultra-high energy\ncosmic rays (UHECRs) with the energy E≥1018.5eV, mostly have the extragalactic\norigin. Actually they are frequently related with the γ-ray bursts, see e.g. [206]. The\nmain experimental studies of UHECRs presented in [205] are perfor med by the Pierre\nAuger Collaboration. While the experiments discussed in [205] showed that the Lorentz\ninvariance isvalid uptothe valueofthe usual Lorentz γfactorup to1011, we canconclude\nthat these experiments impose strong restrictions on the parame ters of Lorentz symmetry\nbreaking but do not rule it out, and, as it is claimed in [205], new studies in this direction\nclearly will establish new questions, and, among the suggestions give n in that paper, one\nshould emphasize the need to study, first, γ-ray bursts, second, high-energy neutrinos.\nWithin this context, the discussion of the γ-ray bursts, first time observed in 1967\nand representing themselves as processes characterized by ext remely high energy scale, is\nespecially important. In [207], it was suggested that namely by this r eason, it is natural\nto expect the LV effects within the γ-ray bursts observations, with the main line of study\ncould consist indetermination of polarizationof the electromagnetic waves emitted within\nthe bursts. Also, in [207], some spectroscopic experiments in order to determine the LV\nparameters kµandκµνλρaresuggested; theresultsofsometheseexperimentsareprese nted\nin [1] where the typical bound for kµobtained from different experiments is about 10−43\nGeV, and for the dimensionless κµνλρ– about 10−15÷10−17. More recent results for study\nof theγ-ray bursts in this context are presented in [208], where the resu lts are obtained\nwith use of the Fermisatellite, where it is shown on the base of these observations that\nthe characteristic energy of Lorentz symmetry breaking should b e no less than about\n7,6MPl, that is, larger than the Planck energy. Besides of this, one should mention as\nwell the paper [209], where the constraints for the LV coefficients, including the higher-\ndimensional coefficients, obtained on the base of the γ-ray bursts studies are given. Also,\nin [210], on the base of studies of γ-ray bursts some estimations for upper limits for\nthe vacuum birefringence effect were obtained, however, these lim its are too large, e.g.,\nforξcharacterizing the Myers-Pospelov term (2.13) they give ξ <2.6×108. Besides\nof this, in [206] estimations for modification of dispersion relations an d for rotation of\nthe polarization plane of the light in a vacuum based on study of the γ-ray burst GRB\n041219A were carried out, and the possibility of Horava-Lifshitz-lik e deformations of the\ndispersion relations, with additive αz/vectork2zterms, was estimated, and it was shown that\n103time delay expected from suggestions of existence of additive term s characterized by the\ncritical exponent z= 2, calculated on the base of results obtained for this γ-ray burst, is\nbounded by 10−21s.\nThe LV parameters can be estimated as well on the base of astroph ysical observations.\nIn this context, it is important to mention results obtained from obs ervations of astro-\nphysicalγ-ray sources which allowed to estimate the bounds for cµνcoefficients from the\nspinor sector of LV QED to be of the order of 10−16÷10−17[211].\nIt should benotedthat the LVmodifications of dispersion relations f or theelectromag-\nnetic field can imply as well in the Cherenkov radiation, see e.g. [41, 21 2]. For example,\nin [212] it was argued that the Cherenkov radiation can occur for th e Myers-Pospelov-like\ndispersion relationfor thephoton E2=p2±ξp3\nM, andfor theelectron E2=p2+m2+ηR,Lp3\nM,\nwithMis a Planck-order mass scale, ηR,Lcorrespond to different electron helicities, and\nthe sign±is for two photon helicities. It was argued in [212] that it follows from o bser-\nvations of the Crab nebula that |ξ| ≤10−3, and|ηL−ηR|<4.\nVarious experiments with elementary particles are used for tests a nd measurements of\nparameters of Lorentz symmetry breaking. In this context, the experiments with Penning\ntraps, allowing to detect a possible particle-antiparticle asymmetry , play the special role.\nIt was proved in these studies [213], that the bounds for componen ts of the LV axial\nvectorbµfrom the minimal LV extended QED (2.1) are about 10−24GeV. In the same\npaper, some estimations for non-minimal LV parameters are also pr esented. More results\nfor non-minimal LV parameters can be found in [214]. In [208], some o ther experiments\nwithin the particle physics allowing to measure the LV parameters wer e proposed.\nManyotherstudiesofelementaryparticlesaimedtodetecttheimpa ctsofLorentz-CPT\nsymmetry breaking were performed as well. Among them, experimen ts with neutrinos\nhave a special role. First of all, it should be noted that the neutrino o scillations also can\nbe explained with use of the Lorentz symmetry breaking, as it was ar gued in [215, 216].\nIt is interesting to notice that despite the famous “discovery of su perluminal neutrinos”\nwithin the OPERA experiment was proved to be an experimental erro r, it called much\nattention namely to studying of possible LV effects in neutrino physic s. Further, studies\nof neutrino oscillations were used to obtain estimations for LV param eters in [217], where\nit was shown that the bound for the parameter aµdefined in (2.1), estimated on the base\nof these oscillations, is about 10−23GeV. Various estimations for the Lorentz-breaking\nparameters were obtained also within the framework of the DUNE ex periment, see f.e.\n[218], and the IceCube experiments, see f.e. [219]. As another app lication of neutrinos, in\n[220] it was argued that the neutrinos are the very convenient obj ects for measurements\nallowing for imposing very strong bounds on LV parameters from gra vity experiments\nwhich we consider in the next section.\n1047.2 Gravitational effects\nPurely gravitational studies presented by an important line of expe riments have been\ndescribed in [221]. Within these studies, the LV extensions for theor y of fermions and\nbumblebee model coupled to the gravity were considered. Different tests, representing\nthemselves either as free-fall gravimeter tests, where falling cor ner cubes or matter in-\nterferometers were used as free-fall gravimeters, or as the te sts of the weak equivalence\nprinciple (WEP) were carried out. Other gravimeters used within the study performed\nin [221] were the force-comparison gravimeters based on comparin g of the gravitational\nforce with anappropriate electromagnetic force. As a result, e.g., for the force-comparison\ngravimeter tests, the experimental bounds were showed to be: a bout 10−8GeV for the\ncoefficientaµcharacterizing the term aµeeµ\na¯ψγaψ, and about 10−8for the coefficients cµν\nentering the term −1\n2icλνeeµ\naeνaeλ\nb¯ψ↔\nDµψ. In [221], also other tests for the WEP were\npresented, e.g., the satellite-based ones, the Solar system ones b ased on the precession\nof the perihelion, which yield the results: |aµ| ≤10−6GeV, and |cµν| ≤10−7, tests with\nantimatter, and a great number of photon tests. Some possible sa tellite tests have been\ndiscussed as well in [222].\nOne of the most important experimental discoveries of recent yea rs is, certainly, the\ndetection of gravitational waves [223]. Therefore, it is natural to suggest that the LV im-\npacts can be studied within the context of gravitational waves as w ell. In particular, as we\nalready mentioned in the Chapter 6, some LV extensions of gravity c an allow for birefrin-\ngence of gravitational waves (see e.g. [145]). The results of first s tudies in this direction\nbased on study of the gravitational-wave event GW150914 are pre sented in [224], for a\nmore detailed discussion see [182]. Later experimental studies of bir efringence of gravi-\ntational waves, for a certain higher-derivative extension of the E instein gravity, can be\nfound in [225]. It is interesting to note that some studies in that pape r suggest to consider\nhigher-derivative extensions of gravity, while up to now, except of the Chern-Simons mod-\nified gravity, no other examples of higher-derivative LV gravity mod els were sufficiently\nstudied. It worth mentioning that within this context, the gravitat ional Cherenkov ra-\ndiation can exist as well and will produce the energy losses, see the d iscussion in [226],\nwhere a typical extension of gravity implying in the modified Einstein eq uations\nGµν= 8πGTµν+ˆ¯sαβ˜Rαµβν,\nwith˜Rαµβνis a double dual of the Riemann tensor obtained through its contrac tion with\ntwo Levi-Civita symbols, and ˆ¯sαβis a differential operator involving LV parameters of\ndimensions d= 4,6,8,...likeˆ¯sµν=/summationtext\nd(¯s(d))α1...αd−4µν∂α1...∂αd−4, was considered, and\nobservations of cosmic rays were used to estimate the LV paramet ers (¯s(d))α1...αd−4µν , e.g.,\n105it was found that the bound for the dimensionless (¯ s(4))µνis of the order 10−14. In [227],\nit was estimated with use of the lunar laser range, to be about 10−9(the lunar laser\nrange was used to estimate LV parameters also in [228]), and in [229] this parameter was\ndetermined by gravity tests of the Solar system objects, where t he typical bound was\nfound to be about 10−9...10−11. Some aspects of gravitational Cherenkov radiation are\ndiscussed also in [230]. And in [231], estimations for the coefficients up t o (¯s(10)) are done\non the base of different cosmic ray observations, so that the typic al bound of (¯ s(10))ijturns\nout to be about 10−61GeV−6. As an aside observation, we once more see that the studies\nof cosmic rays and γ-ray bursts allow to obtain the most important information about th e\nscale of LV parameters. Further, in [232], a methodology allowing for estimation of LV\nparameters from observations of gravitational waves was prese nted, and it was claimed in\nthis paper that this analysis is currently on the way. Among other ex perimental studies\nof LV effects in gravity, the short-range gravity tests [233] are worth mentioning. The\nmost recent results of various experiments on Lorentz and CPT br eaking, both in flat and\ncurved spacetime, can be found in [234].\nTo close this book, we note that there are numerous experiments p erformed in order\nto study different LV extensions of various field theories, either of gravity, QED or other\nmodels, and apparently, more new experiments are to be performe d. And in principle, we\nnote that the complete picture of a possible LV extension of the Sta ndard Model is still\nvery far from its conclusion, and many problems in this context cont inue to be open and\nwill be open during many years, and the main attention in nearest yea rs certainly should\nbe paid to studying LV extensions of gravity.\n106Acknowledgements. Authors are grateful to J. F. Assun¸ c˜ ao, A. P. Baeta Scarpelli,\nH. Belich, L. H. C. Borges, F. A. Brito, L. C. T. Brito, M. Cvetic, A. G . Dias, J. C.\nC. Felipe, A. F. Ferrari, C. Furtado, M. S. Guimaraes, H. O. Girotti, M. Gomes, A. C.\nLehum, R. V. Maluf, C. Marat Reyes, E. Passos, P. Porfirio, A. A. R ibeiro, R. F. Ribeiro,\nV. O. Rivelles, A. F. Santos, M. Schreck, A. J. da Silva, A. R. Vieira, C . Wotzasek, C.\nA. D. Zarro, for fruitful collaboration and interesting discussions . The work has been\npartially supported by CNPq.\n107108Bibliography\n[1] V. A. Kostelecky, N. Russell, Rev. Mod. Phys. 83, 11 (2011), ar Xiv: 0801.0287.\n[2] V. A. Kostelecky, S. Samuel, Phys. Rev. D39, 683 (1989).\n[3] S. Carroll, G. B. Field, R. Jackiw, Phys. Rev. D41, 1231 (1990).\n[4] P. Horava, Phys. Rev. D 79, 084008 (2009), arXiv: 0901.3775.\n[5] D. Colladay, V. A. Kostelecky, Phys. Rev. D55, 6770 (1997), he p-ph/9703464.\n[6] D. Colladay, V. A. Kostelecky, Phys. Rev. D58, 116002 (1998), hep-ph/9809521.\n[7] M. Chaichian, K. Nishijima, A. Tureanu, Phys. Lett. B568, 146 (2 003),\nhep-th/0209008.\n[8] V. A. Kostelecky, “Topics in Lorentz and CPT violation”, hep-ph/ 0104227.\n[9] D. Mattingly, Living Rev. Rel. 8, 5 (2005), gr-qc/0502097.\n[10] A. Albrecht, J. Magueijo, Phys. Rev. D59, 043516 (1999), as tro-ph/9811018.\n[11] N. Seiberg, E. Witten, JHEP 9909, 032 (1999), hep-th/99081 42.\n[12] R. Gambini, J. Pullin, “A First Course in Loop Quantum Gravity”, Ox ford Univ.\nPress, 2001.\n[13] S. Hawking, Nucl. Phys. B144, 349 (1978).\n[14] S. Bernadotte and F. R. Klinkhamer, Phys. Rev. D75, 024028 ( 2007),\nhep-ph/0610216.\n[15] F. R. Klinkhamer and C. Rupp, Phys. Rev. D70, 045020 (2004), hep-th/0312032.\n[16] R. Jackiw, Nucl. Phys. Proc. Suppl. 108, 30 (2002), hep-th/0 110057.\n[17] M. Schreck, Phys. Rev. D89, 085013 (2014), arXiv:1311.0032 .\n109[18] A. G. Grushin, Phys. Rev. D86, 045001 (2012), arXiv: 1205.37 22.\n[19] P. D. S. Silva, M. M. Ferreira, M. Schreck and L. F. Urrutia, Phy s. Rev. D102,\n076001 (2020), arXiv:2006.02022.\n[20] V. A. Kosteleck´ y, R. Lehnert, N. McGinnis, M. Schreck and B. Seradjeh, Phys.\nRev. Res. 4, 023106 (2022), arXiv:2112.14293.\n[21] V. A. Kostelecky, R. Lehnert, Phys. Rev. D63, 065008 (2001 ), hep-th/0012060.\n[22] F. R. Klinkhamer and M. Schreck, Nucl. Phys. B848, 90 (2011), arXiv:1011.4258.\n[23] F. R. Klinkhamer and M. Schreck, Nucl. Phys. B856, 666 (2012) , arXiv:1110.4101.\n[24] J. D. Bjorken, Phys. Rev. 24, 174 (1963).\n[25] R. Jackiw, V. A. Kostelecky, Phys. Rev. Lett. 82, 357 (1999) , hep-ph/9901358.\n[26] R. Jackiw, Int. J. Mod. Phys. B14, 2011 (2000), hep-th/990 3044.\n[27] L. Alvarez-Gaume, E. Witten, Nucl. Phys. B234, 269 (1984).\n[28] S. Carroll, J. Harvey, V. A. Kostelecky, C. D. Lane, T. Okamot o, Phys. Rev. Lett.\n87, 141601 (2001), hep-th/0105082.\n[29] M. Schreck, J. Phys. Conf. Ser. 563, 012026 (2014).\n[30] Q. G. Bailey and C. D. Lane, Symmetry 10, 480 (2018), arXiv:181 0.05136.\n[31] M. S. Berger, V. A. Kostelecky, Phys. Rev. D65, 091701 (200 2), hep-th/0112243.\n[32] R. Jackiw, S.-Y. Pi, Phys. Rev. D68, 104012 (2003), gr-qc/03 08071.\n[33] S. Alexander, N. Yunes, Phys. Rept. 480, 1 (2009), arXiv: 09 07.2562.\n[34] V. A. Kostelecky, Phys. Rev. D69, 105009 (2004), hep-th/0 312310.\n[35] V. A. Kostelecky, C. D. Lane, A. G. M. Pickering, Phys. Rev. D6 5, 056006 (2002),\nhep-th/0111123.\n[36] S. Deser, R. Jackiw, S. Templeton, Ann. Phys. 140, 372 (1982 ).\n[37] D. Colladay, P. McDonald, Phys. Rev. D75, 105002 (2007), hep -ph/0609084.\n[38] M. Gomes, J. R. Nascimento, E. Passos, A. Yu. Petrov, A. J. d a Silva, Phys. Rev.\nD76, 047701 (2007), arXiv: 0704.1104.\n110[39] S. Carroll, H. Tam, Phys. Rev. D78, 044047 (2008), arXiv: 080 2.0521.\n[40] M. Gomes, J. R. Nascimento, A. Yu. Petrov, A. J. da Silva, Phys . Rev. D81, 045018\n(2010), arXiv: 0911.3548.\n[41] B. Altschul, Phys. Rev. Lett. 98, 041603 (2007), hep-th/06 09030.\n[42] R. Myers, M. Pospelov, Phys. Rev. Lett. 90, 211601 (2003), hep-ph/0301124.\n[43] C. M. Reyes, Phys. Rev. D82, 125036 (2010), arXiv:1011.2971 .\n[44] C. M. Reyes, Phys. Rev. D87, 125028 (2013), arXiv:1307.5340 .\n[45] M. Maniatis and C. M. Reyes, Phys. Rev. D89, 056009 (2014), a rXiv:1401.3752.\n[46] C. M. Reyes and L. F. Urrutia, Phys. Rev. D95, 015024 (2017) , arXiv:1610.06051.\n[47] J. Lopez-Sarrion, C. M. Reyes and C. Riquelme, “Tree-level un itarity, causality and\nhigher-order Lorentz and CPT violation,” arXiv:2206.05784.\n[48] M. M. Ferreira, L. Lisboa-Santos, R. V. Maluf and M. Schreck, Phys. Rev. D100,\n055036 (2019), arXiv:1903.12507.\n[49] H. Belich, M. M. Ferreira, J. A. Helayel-Neto, Eur. Phys. J. C38 , 511 (2005),\nhep-th/0401235.\n[50] P. Gaete and C. Wotzasek, Phys. Rev. D75, 057902 (2007), h ep-ph/0607321.\n[51] D. Bazeia, R. Menezes, Phys. Rev. D73, 065015 (2006), hep- th/0506262.\n[52] L. H. C. Borges, A. G. Dias, A. F. Ferrari, J. R. Nascimento, A. Yu. Petrov, Phys.\nRev. D89, 045005 (2014), arXiv: 1304.5484.\n[53] V. A. Kosteleck´ y and Z. Li, Phys. Rev. D99, 056016 (2019), a rXiv: 1812.11672\n[hep-ph].\n[54] G. Amelino-Camelia, Int. J. Mod. Phys. D11, 35 (2002), gr-qc/ 0012051.\n[55] S. Alexander, J. Magueijo, “Noncommutative geometry as a re alization of varying\nspeed of light cosmology”, hep-th/0104093.\n[56] J. Magueijo, L. Smolin, Phys. Rev. D67, 044017 (2003), gr-qc /0207085.\n[57] R. Casana, M. M. Ferreira, A. R. Gomes, F. E. P. dos Santos, P hys. Rev. D82,\n125006 (2010), arXiv: 1010.2776.\n111[58] S. Deser, R. Jackiw, Phys. Rev. B139, 371 (1984).\n[59] M. A. Anacleto, A. Ilha, J. R. S. Nascimento, R. F. Ribeiro, C. Wo tzasek, Phys.\nLett. B04, 268 (2001), hep-th/0104152.\n[60] M. A. Anacleto, C. Furtado, J. R. Nascimento, A. Yu. Petrov, Phys. Rev. D78,\n065014 (2008), arXiv: 0805.3517.\n[61] A. P. Baeta Scarpelli, J. R. Nascimento, A. Yu. Petrov, R. F. Rib eiro, Eur. Phys.\nJ. C75, 34 (2015), arXiv: 1412.6957.\n[62] M. S. Guimaraes, J. R. Nascimento, A. Yu. Petrov, C. Wotzase k, Europhys. Lett.\n95, 51002 (2011), arXiv: 1010.3666.\n[63] M. Seifert, Phys. Rev. D81, 065010 (2010), arXiv: 0909.3118 .\n[64] B. Altschul, Q. G. Bailey and V. A. Kostelecky, Phys. Rev. D81, 0 65028 (2010),\narXiv: 0912.4852.\n[65] M. Gomes, T. Mariz, J. R. Nascimento, A. J. da Silva, Phys. Rev. D77, 105002\n(2008), arXiv: 0709.2904.\n[66] J. F. Assuncao, T. Mariz, J. R. Nascimento and A. Y. Petrov, P hys. Rev. D96,\n065021 (2017), arXiv: 1707.07778.\n[67] J. F. Assun¸ c˜ ao, T. Mariz, J. R. Nascimento and A. Y. Petrov , Phys. Rev. D100,\n085009 (2019), arXiv: 1902.10592.\n[68] J. L. Chkareuli, J. Jejelava and Z. Kepuladze, Bled Workshops P hys. 19, 2, 74\n(2018), arXiv: 1811.09578.\n[69] M. D. Seifert, Phys. Rev. D100, 065017 (2019), arXiv: 1903.0 6140.\n[70] C. A. Escobar and R. Potting, Int. J. Mod. Phys. A35, 205017 4 (2020), arXiv:\n1810.01677.\n[71] M. D. Seifert, Phys. Rev. D99, 045003 (2019), arXiv: 1810.09 512.\n[72] M. D. Seifert, Phys. Rev. D82, 125015 (2010), arXiv: 1008.18 10.\n[73] H. Belich, M. M. Ferreira, J. A. Helayel-Neto, M. T. D. Orlando, P hys. Rev. D68,\n025005 (2003), hep-th/0301224.\n112[74] C. M. Cantanhede, R. Casana, M. M. Ferreira, E. da Hora, Phy s. Rev. D86, 065011\n(2012), arXiv: 1109.2931.\n[75] H. Belich, T. Costa-Soares, M. M. Ferreira, J. A. Helayel-Neto , Eur. Phys. J. C41,\n421 (2005), hep-th/0410104.\n[76] H. Belich, E. O. Silva, M. M. Ferreira, M. T. D. Orlando, Phys. Rev . D83, 125025\n(2011), arXiv: 1106.0789.\n[77] J. A. A. S. Reis and M. Schreck, Phys. Rev. D95, 075016 (2017 ), arXiv:1612.06221.\n[78] J. A. A. S. Reis and M. Schreck, Symmetry 11, 1197 (2019), ar Xiv:1909.11061.\n[79] K. Bakke, H. Belich, ”Spontaneous Lorentz Symmetry Violation and Low Energy\nScenarios”, Lambert Acad. Publ., 2015.\n[80] H. Belich, M. M. Ferreira, J. A. Helayel-Neto, M. T. D. Orlando, P hys. Rev. D67,\n125011 (2003), hep-th/0212330.\n[81] B. Charneski, M. Gomes, T. Mariz, J. R. Nascimento, A. J. da Silv a, Phys. Rev.\nD79, 065009 (2007), arXiv: 0809.1627.\n[82] J. Gamboa, J. Lopez-Sarrion, Phys. Rev. D71, 067702 (2005 ), hep-th/0501034.\n[83] G. Bonneau, Nucl. Phys. B593, 418 (2001), hep-th/0008210 .\n[84] T. Mariz, R. Martinez, J. R. Nascimento and A. Y. Petrov, Eur. Phys. J. C81, 974\n(2021), arXiv: 2104.05681.\n[85] F. A. Brito, J. R. Nascimento, E. Passos, A. Yu. Petrov, Phys . Lett. B664, 112\n(2008), arXiv: 0709.3090.\n[86] A. P. Baeta Scarpelli, M. Gomes, A. Yu. Petrov, A. J. da Silva, Ph ys. Rev. D93,\n025010 (2016), arXiv: 1509.05352.\n[87] E. Passos, A. Yu. Petrov, Phys. Lett. B662, 441 (2008), ar Xiv: 0801.4946.\n[88] J. R. Nascimento, A. Yu. Petrov, C. Wotzasek, C. A. D. Zarro , Phys. Rev. D89,\n065030 (2014), arXiv: 1403.0786.\n[89] A. N. Redlich, Phys. Rev. D29, 2366 (1984).\n[90] M. M. Ferreira, J. A. A. S. Reis and M. Schreck, Phys. Rev. D10 0, 095026 (2019),\narXiv:1905.04401.\n113[91] K. S. Babu, A. Das, P. Panigrahi, Phys. Rev. D36, 3725 (1987) .\n[92] B. Seradjeh, C. Weeks, M. Franz, Phys. Rev. B77, 033104 (2 008), arXiv: 0706.1559.\n[93] B. Seradjeh, M. Franz, Phys. Rev. Lett. 101, 146401 (2008 ), arXiv: 0709.4258.\n[94] S. Coleman, S. Glashow, Phys. Rev. D9, 116008 (1999), hep-p h/9812418.\n[95] B. Altschul, Phys. Rev. D99, 125009 (2019), arXiv: 1903.1010 0.\n[96] S. Karki and B. Altschul, Phys. Rev. D102, 035009 (2020), ar Xiv: 2007.00133.\n[97] A. A. Andrianov, P. Giacconi, R. Soldati, JHEP 0202, 030 (2002) , hep-th/0110279.\n[98] T. Mariz, J. R. Nascimento, E. Passos, R. F. Ribeiro, F. A. Brito , JHEP 0510, 019\n(2005), hep-th/0509008.\n[99] F. A. Brito, L. S. Grigorio, M. S. Guimaraes, E. Passos, C. Wotz asek, Phys. Lett.\nB681, 495 (2009), arXiv: 0906.2786.\n[100] A. P. Baeta Scarpelli, M. Sampaio, M. C. Nemes, B. Hiller, Eur. Ph ys. J. C56, 571\n(2008), arXiv: 0804.3537.\n[101] A. P. Baeta Scarpelli, M. Sampaio, B. Hiller, M. C. Nemes, Phys. R ev. D64, 046013\n(2001), hep-th/0102108.\n[102] J. M. Chung, Phys. Rev. D60, 127901 (1999), hep-th/9904 037.\n[103] K. Fujikawa, H. Suzuki, Path Integrals and Quantum Anomalies , Clarendon Press,\nOxford, 2004.\n[104] T. Mariz, J. R. Nascimento, A. Yu. Petrov, L. Y. Santos, A. J . da Silva, Phys. Lett.\nB661, 312 (2008), arXiv: 0708.3348.\n[105] L.Cervi, L.Griguolo,D.Seminara, Phys. Rev.D64,105003(2 001),hep-th/0104022.\n[106] F. Brito, J. R.Nascimento, E. Passos, A. Yu. Petrov, JHEP 0 706, 016(2007), arXiv:\n0705.1338.\n[107] A. P. Baeta Scarpelli, T. Mariz, J. R. Nascimento, A. Yu. Petro v, Eur. Phys. J.\nC73, 2526 (2013), arXiv: 1304.2256.\n[108] T. Mariz, J. R. Nascimento, A. Yu. Petrov, W. Serafim, Phys. Rev. D90, 045015\n(2014), arXiv: 1406.2873.\n114[109] A. J. G. Carvalho, D. R. Granado, J. R. Nascimento and A. Y. P etrov, Eur. Phys.\nJ. C79, 817 (2019), arXiv: 1908.04873.\n[110] G. Gazzola, H. G. Fargnoli, A. P. Baeta Scarpelli, M. Sampaio, M. C. Nemes, J.\nPhys. G39, 035002 (2012), arXiv: 1012.3291.\n[111] R. Casana, M. M. Ferreira, R. V. Maluf, F. E. P. dos Santos, P hys. Lett. B726, 815\n(2013), arXiv: 1302.2375.\n[112] D. Colladay and P. McDonald, J. Math. Phys. 43, 3554 (2002), hep-ph/0202066.\n[113] R. V. Maluf, T. Mariz, J. R. Nascimento, A. Yu. Petrov, Int. J . Mod. Phys. A33,\n1850018 (2018). arXiv: 1604.06647.\n[114] V. A. Kostelecky and M. Mewes, Phys. Rev. D80, 015020 (200 9), arXiv: 0905.0031.\n[115] A. Kosteleck´ y and M. Mewes, Phys. Rev. D88, 096006 (2013 ), arXiv: 1308.4973.\n[116] M. Cambiaso, R. Lehnert and R. Potting, Phys. Rev. D85, 085 023 (2012), arXiv:\n1201.3045.\n[117] T. Mariz, J. R. Nascimento, A. Yu. Petrov, Phys. Rev. D85, 1 25003 (2012), arXiv:\n1111.0198.\n[118] A. P. Baeta Scarpelli, T. Mariz, J. R. Nascimento, A. Yu. Petro v, Int. J. Mod. Phys.\nA31, 1650063 (2016), arXiv: 1505.04047.\n[119] T. Mariz, W. Serafim, J. Phys. G40, 075003 (2013), arXiv: 17 12.09675.\n[120] A. Celeste, T. Mariz, J. R. Nascimento, A. Yu. Petrov, Phys. Rev. D93, 065012\n(2016), arXiv: 1602.02570.\n[121] T. Mariz, Phys. Rev. D83, 045018 (2011), arXiv: 1010.5013.\n[122] M. Cambiaso, R. Lehnert and R. Potting, Phys. Rev. D90, 065 003 (2014), arXiv:\n1401.7317.\n[123] L.C. T.Brito, J.C. C.Felipe, J. R.Nascimento, A.Y.Petrov a ndA.P.B. Scarpelli,\nPhys. Rev. D102, 075017 (2020), arXiv: 2007.11538.\n[124] A. F. Ferrari, J. R. Nascimento and A. Y. Petrov, Eur. Phys. J. C80, 459 (2020),\n459, arXiv: 1812.01702.\n115[125] A. P. Baeta Scarpelli, L. C. T. Brito, J. C. C. Felipe, J. R. Nascim ento and\nA. Y. Petrov, EPL 123, 21001 (2018), arXiv: 1805.06256.\n[126] A. J.G.Carvalho, A.F.Ferrari, A.M. DeLima, J.R.Nasciment o andA.Y. Petrov,\nNucl. Phys. B942, 393 (2019), arXiv: 1803.04308.\n[127] T. Mariz, J. R. Nascimento, A. Y. Petrov and C. M. Reyes, Phy s. Rev. D99, 096012\n(2019), arXiv: 1804.08413.\n[128] A. Ferrero and B. Altschul, Phys. Rev. D84, 065030 (2011), arXiv: 1104.4778.\n[129] L. C. T. Brito, J. C. C. Felipe, A. Y. Petrov and A. P. B. Scarpe lli, Int. J. Mod.\nPhys. A36, 2150033 (2021), arXiv: 2005.04637.\n[130] T. Filk, Phys. Lett. B376, 53 (1996).\n[131] I. Ya. Aref’eva, D. M. Belov, A. S. Koshelev, Phys. Lett. B47 6, 431 (2000),\nhep-th/9912075.\n[132] A. F. Ferrari, A. C. Lehum, A. J. da Silva and F. Teixeira, J. Phy s. A40, 7803\n(2007), hep-th/0612223.\n[133] L. Alvarez-Gaume, J. L. F. Barbon, R. Zwicky, JHEP 0105, 05 7 (2001),\nhep-th/0103069.\n[134] S. Minwalla, M. van Raamsdonk, N. Seiberg, JHEP 0002, 020 (20 00),\nhep-th/9912072.\n[135] C.M.Reyes, S.Ossandon, C.Reyes, Phys.Lett.B746,190 (2015),arXiv: 1409.0508.\n[136] H. O. Girotti, M. Gomes, V. O. Rivelles, A. J. da Silva, Nucl. Phys. B587, 299\n(2000), hep-th/0005272.\n[137] A. A. Bichl, J. M. Grimstrup, L. Popp, M. Schweda, R. Wulkenha ar, Int. J. Mod.\nPhys. A17, 2212 (2002), hep-th/0102044.\n[138] Z. Guralnik, R. Jackiw, S.-Y. Pi, A. Polychronakos, Phys. Lett . B517, 450 (2001),\nhep-th/0106044.\n[139] T. Mariz, J. R. Nascimento, V. O. Rivelles, Phys. Rev. D75, 025 020 (2007),\nhep-th/0609132.\n[140] C. P. Martin, C. Tamarit, JHEP 0811, 087 (2008), arXiv: 0809 .2684.\n116[141] L. H. C. Borges, A. G. Dias, A. F. Ferrari, J. R. Nascimento, A . Yu. Petrov, Phys.\nLett. B756, 332 (2016), arXiv: 1601.03298.\n[142] J. M. Carmona, J.L. Cortez, J. Gamboa, F. Mendez, Phys. Le tt. B565, 222 (2003),\nhep-th/0207158.\n[143] J. R. Nascimento, A. Yu. Petrov, R. F. Ribeiro, Europhys. Le tt. 77, 51001 (2007),\nhep-th/0601077.\n[144] H. Falomir, J. Gamboa, J. Lopez-Sarrion, F. Mendez, A. J. da Silva, Phys. Lett.\nB632, 740 (2005), hep-th/0504032.\n[145] A. F. Ferrari, M. Gomes, J. R. Nascimento, E. Passos, A. Yu. Petrov, A. J. da Silva,\nPhys. Lett. B652, 174 (2007), hep-th/0609222.\n[146] C. Carlson, C. Carone, N. Zobin, Phys. Rev. D66, 075001 (20 02), hep-th/0206035.\n[147] H. Belich, J. L. Boldo, L. P. Colatto, J. A. Helayel-Neto, A. L. M . A. Nogueira,\nPhys. Rev. D68, 065030 (2003), hep-th/0304166.\n[148] P. Bolokhov, S. Nibbelink, M. Pospelov, Phys. Rev. D72, 01501 3 (2005),\nhep-ph/0505029.\n[149] I. L. Buchbinder, S. M. Kuzenko. Ideas and Methods of Supe rsymmetry and Super-\ngravity. IOP Publishing, Bristol and Philadelphia, 1998.\n[150] M. Gomes, J. Queiruga, A. J. da Silva, Phys. Rev. D92, 025050 (2015), arXiv:\n1506.01331.\n[151] A. Pickering, P. West, Phys. Lett. B383, 54 (1996), hep-th /9604147.\n[152] A. C. Lehum, J. R. Nascimento, A. Yu. Petrov, A. J. da Silva, P hys. Rev. D88,\n045022 (2013), arXiv: 1305.1812.\n[153] H. Belich, L. D. Bernard, P. Gaete, J. A. Helayel-Neto, Eur. P hys. J. C73, 2632\n(2013), arXiv: 1303.1108.\n[154] H. Belich, L. D. Bernard, P. Gaete, J. A. Helayel-Neto, F. J. L . Leal, Eur. Phys. J.\nC75, 291 (2015), arXiv: 1502.06126.\n[155] M. Gomes, J. R. Nascimento, A. Yu. Petrov, A. J. da Silva, Phy s. Rev. D90, 125022\n(2014), arXiv: 1408.6499.\n117[156] M. Cvetic, T. Mariz, A. Yu. Petrov, Phys. Rev. D92, 085041 ( 2015), arXiv:\n1509.00251.\n[157] D. Colladay, P. McDonald, Phys. Rev. D83, 025021 (2011), ar Xiv: 1010.1781.\n[158] N. Seiberg, JHEP 0306, 010 (2003), hep-th/0305248.\n[159] V. A. Kosteleck´ y, Z. Li, Phys. Rev. D103, 024059 (2021), a rXiv: 2008.12206.\n[160] G. de Berredo-Peixoto and I. L. Shapiro, Phys. Lett. B642 ( 2006), 153-159,\nhep-th/0607109.\n[161] D. Grumiller, N. Yunes, Phys. Rev. D77, 044015 (2008), arXiv : 0711.1868.\n[162] T. Mariz, J. R. Nascimento, E. Passos, R. F. Ribeiro, Phys. Re v. D70, 024014\n(2004), hep-th/0403205.\n[163] G. ’t Hooft and M. J. G. Veltman, Nucl. Phys. B44, 189 (1972).\n[164] S. Ojima, Progr. Theor. Phys. 81, 512 (1989).\n[165] M. Gomes, T. Mariz, J. R. Nascimento, E. Passos, A. Yu. Petr ov, Phys. Rev. D78,\n025029 (2008), arXiv: 0805.4409.\n[166] J. C. C. Felipe, R. R. Vieira, A. L. Cherchiglia, A.P. Baeta Scarpe lli, M. Sampaio,\nPhys. Rev. D89, 105034 (2014), arXiv: 1403.2690.\n[167] R. Bluhm and A. Sehic, Phys. Rev. D94, 104034 (2016), arXiv:1 610.02892.\n[168] R.Bluhm, H.Bossi andY.Wen, Phys. Rev.D100, 084022(201 9), arXiv:1907.13209.\n[169] T. Jacobson, D. Mattingly, Phys. Rev. D64, 024028 (2001), gr-qc/0007031.\n[170] T. Jacobson, D. Mattingly, Phys. Rev. D70, 024003 (2004), gr-qc/0402005.\n[171] T. Jacobson, PoS QG-PH (2007), 020, arXiv: 0801.1547.\n[172] O. Bertolami, J. Paramos, Phys. Rev. D72, 044001 (2005), h ep-th/0504215.\n[173] A. F. Santos, A. Y. Petrov, W. D. R. Jesus and J. R. Nascimen to, Mod. Phys. Lett.\nA30, 1550011 (2015), arXiv: 1407.5985.\n[174] D. Capelo and J. P´ aramos, Phys. Rev. D91, 104007 (2015), arXiv: 1501.07685.\n[175] C. M. Reyes and M. Schreck, Phys. Rev. D104, 124042 (2021 ), arXiv:2105.05954.\n118[176] C. M. Reyes and M. Schreck, “Modified-gravity theories with n ondynamical back-\nground fields,” arXiv:2202.11881.\n[177] C. M. Reyes, M. Schreck and A. Soto, “Cosmology in the prese nce of\ndiffeomorphism-violating, nondynamical background fields,” arXiv:22 05.06329.\n[178] Q. G. Bailey and V. A. Kostelecky, Phys. Rev. D74, 045001 (20 06), gr-qc/0603030.\n[179] M. D. Seifert, Phys. Rev. D79, 124012 (2009), arXiv: 0903.2 279.\n[180] V. A. Kosteleck´ y and M. Mewes, Phys. Lett. B779, 136 (201 8), arXiv:1712.10268.\n[181] J. R. Nascimento, A. Y. Petrov and A. R. Vieira, Galaxies, 9, 32 (2021), arXiv:\n2104.01651.\n[182] M. Mewes, Phys. Rev. D99, 104062 (2019), arXiv: 1905.0040 9.\n[183] K. Konno, T. Matsuyama, S. Tanda, Progr. Theor. Phys. 12 2, 561 (2009), arXiv:\n0902.4767.\n[184] K. Godel, Rev. Mod. Phys, 21, 747 (1949).\n[185] M. Rebou¸ cas, J. Tiomno, Phys. Rev. D28, 1251 (1983).\n[186] P. J. Porfirio, J.-B. Fonseca-Neto, J. R. Nascimento, A. Yu. Petrov, J. Ricardo, A.\nF. Santos, Phys. Rev. D94, 044044 (2016), arXiv: 1606.00743.\n[187] P. J. Porfirio, J.-B. Fonseca-Neto, J. R. Nascimento, A. Yu. Petrov, Phys. Rev. D94,\n104057 (2016), arXiv: 1610.01539.\n[188] V. A. Kosteleck´ y and Z. Li, Phys. Rev. D104, 044054 (2021) , arXiv:2106.11293.\n[189] V. A. Kosteleck´ y and R. Potting, Phys. Rev. D104, 104046 ( 2021),\narXiv:2108.04213.\n[190] K. Land, J. Magueijo, Phys. Rev. Lett. 95, 071301 (2005), astro-ph/0502237.\n[191] K. Land, J. Magueijo, Mon. Not. Roy. Astron. Soc. 378, 153 (2007),\nastro-ph/0611518.\n[192] K. Greisen, Phys. Rev. Lett. 16, 748 (1966).\n[193] G. T. Zatsepin, V. A. Kuzmin, JETP Lett. 4, 78 (1966).\n[194] J. Magueijo, L. Smolin, Class. Quant. Grav. 21, 1725 (2004), gr-qc/0305055.\n119[195] R. Lehnert and R. Potting, Phys. Rev. Lett. 93, 110402 (20 04), hep-ph/0406128.\n[196] R. Lehnert and R. Potting, Phys. Rev. D70, 125010 (2004) [e rratum: Phys. Rev.\nD70, 129906 (2004)], hep-ph/0408285.\n[197] B. Altschul, Phys. Rev. D75, 105003 (2007), hep-th/07012 70.\n[198] B. Altschul, Symmetry, 9, 11, 250 (2017).\n[199] F. R. Klinkhamer and M. Schreck, Phys. Rev. D78, 085026(20 08), arXiv:0809.3217.\n[200] D. Colladay, P. McDonald and R. Potting, Phys. Rev. D93, 1250 07 (2016), arXiv:\n1603.00308.\n[201] M. Schreck, Phys. Rev. D96, 095026 (2017), arXiv:1702.031 71.\n[202] E. Zavattini et al.[PVLAS Collaboration], Phys. Rev. D77, 032006 (2008), arXiv:\n0706.3419.\n[203] M. Chaichian, M. M. Sheikh-Jabbari, ”Noncommutativity of spa ce and rotation of\npolarization of light in a background magnetic field”, hep-ph/051132 3.\n[204] N. Chatillon, A. Pinzul, Nucl. Phys. B764, 95 (2007), hep-ph/0 607243.\n[205] W. Bietenholz, Phys. Rep. 505, 145 (2011), arXiv: 0806.3713 .\n[206] E. Passos, E. M. C. Abreu, M. A. Anacleto, F. A. Brito, C. Wot zasek, C. A. D.\nZarro, Phys. Rev. D93, 085022 (2016), arXiv: 1603.01558.\n[207] G. Shore, Nucl. Phys. B717, 86 (2004), hep-th/0409125.\n[208] V. Vasileiou, A. Jacholkowska, F. Piron et al, Phys. Rev. D87, 122001(2013), arXiv:\n1305.3463.\n[209] V. A. Kostelecky, M. Mewes, Phys. Rev. Lett. 110, 201601( 2013), arXiv: 1301.5367.\n[210] P. Laurent, D. Gotz, P. Binetruy, S. Covino, A. Fernandez- Soto, Phys. Rev. D83,\n121301 (2011), arXiv: 1106.1068.\n[211] B. Altschul, Symmetry 13, 688 (2021), arXiv: 2104.04587.\n[212] T. Jacobson, S. Liberati, D. Mattingly, F. W. Stecker, Phys. Rev. Lett. 93, 021101\n(2004), astro-ph/0309681.\n[213] Y. Ding and V. A. Kosteleck´ y, Phys. Rev. D94, 056008 (2016 ), arXiv:1608.07868.\n120[214] Y. Ding and M. F. Rawnak, Phys. Rev. D102, 056009 (2020), a rXiv:2008.08484.\n[215] V. A. Kostelecky, ”Neutrino Oscillations and Lorentz Violation” , astro-ph/0403088.\n[216] V. A. Kostelecky, M. Mewes, Phys. Rev. D85, 096005 (2012) , arXiv: 1112.6395.\n[217] U. Rahaman, Eur. Phys. J. C81, 792 (2021), arXiv:2103.0457 6.\n[218] G. Barenboim, M. Masud, C. A. Ternes and M. T´ ortola, Phys. Lett. B788, 308\n(2019), arXiv: 1805.11094.\n[219] A. Crivellin, F. Kirk and M. Schreck, JHEP 04, 082 (2021), arXiv :2009.01247.\n[220] J. S. Diaz, V. A. Kostelecky, M. Mewes, Phys. Rev. D89, 0430 05 (2014), arXiv:\n1308.6344.\n[221] V. A. Kostelecky, J. Tasson, Phys. Rev. 83, 016013 (2011) , arXiv: 1006.4106.\n[222] B. Altschul et al., Adv. Space Res. 55, 501 (2015), arXiv: 140 4.4307.\n[223] B.Abbott, R.Abbott, T.Abbott, M.Abernathy, F.Acerne se, K.Ackley, C.Adams,\nT. Adams, P. Addesso, R. Adhikari, et al. (LIGO Scientific Collaborat ion and Virgo\nCollaboration), Phys. Rev. Lett. 116, 061102 (2016), arXiv: 160 2.03837.\n[224] V. A. Kostelecky, M. Mewes, Phys. Lett. B757, 510 (2016), arXiv: 1602.04782.\n[225] Y. F. Wang, S. M. Brown, L. Shao and W. Zhao, arXiv: 2109.097 18.\n[226] V. A. Kostelecky, J. Tasson, Phys. Lett. B749, 551 (2015) , arXiv: 1508.07007.\n[227] A. Bourgoin, C. Le Poncin-Lafitte, A. Hees, S. Bouquillon, G. F rancou and\nM. C. Angonin, Phys. Rev. Lett. 119, 201102 (2017), arXiv: 1706 .06294.\n[228] A. Bourgoin, S. Bouquillon, A. Hees, C. Le Poncin-Lafitte, Q. G . Bailey,\nJ. J. Howard, M. C. Angonin, G. Francou, J. Chab´ e and C. Courde ,et al.Phys.\nRev. D103, 064055 (2021), arXiv:2011.06641.\n[229] Q. G. Bailey and D. Havert, Phys. Rev. D96, 064035 (2017), a rXiv: 1706.10157.\n[230] Y. Z. Chu and Y. W. Liu, Universe 7, 437 (2021), arXiv: 2108.13 463.\n[231] J. Tasson, Symmetry 8, 111 (2016), arXiv: 1610.05357.\n[232] K. O’Neal-Ault, Q. G. Bailey, T. Dumerchat, L. Haegel and J. Ta sson, Universe 7,\n10, 380 (2021), arXiv: 2108.06298.\n121[233] Q. G. Bailey, A. Kosteleck´ y and R. Xu, Phys. Rev. D91, 02200 6 (2015), arXiv:\n1410.6162.\n[234] E. Adelberger, D. Budker, R. Folman, A. A. Geraci, J. T. Hark e, D. M. Kaplan,\nD. F. J. Kimball, R. Lehnert, D. Moore and G. W. Morley, et al., “Snowmass White\nPaper: Precision Studies of Spacetime Symmetries and Gravitationa l Physics”,\narXiv: 2203.09691.\n122" }, { "title": "1206.3660v2.On_the_spectrum_of_the_pulsed_gamma_ray_emission_from_10MeV_to_400GeV_of_the_Crab_pulsar.pdf", "content": "arXiv:1206.3660v2 [astro-ph.HE] 9 Aug 2012Mon. Not. R. Astron. Soc. 000, 1–??(2011) Printed 7 November 2018 (MN L ATEX style file v2.2)\nOn the spectrum of the pulsed gamma-ray emission from\n10MeV to 400GeV of the Crab pulsar.\nN. Chkheidze1,⋆G. Machabeli1and Z. Osmanov2\n1Centre for Theoretical Astrophysics, ITP, Ilia State Unive rsity, Tbilisi, 0162, Georgia\n2Free University of Tbilisi, Tbilisi, 0183, Georgia\n7 November 2018\nABSTRACT\nIn the present paper a self-consistent theory, interpreting the VERITAS and the\nMAGIC observations of the very high energy pulsed emission from th e Crab pulsar is\nconsidered. The photon spectrum between 10MeV and 400GeV can be described by\ntwo power-law functions with the spectral indexes equal to 2 and 3 .8. The source of\nthe pulsed emission above 10MeV is assumed to be the synchrotronr adiation, which is\ngenerated near the light cylinder during the quasi-linear stage of th e cyclotron insta-\nbility. The emitting particles are the primary beam electrons with the L orentz factors\nup to 109. Such high energies by beam particles is supposed to be reached due to\nLandau damping of the Langmuir waves in the light cylinder region. This mechanism\nprovides simultaneous generation of low (radio) and high energy (10 MeV-400GeV)\nemission on the light cylinder scales, in one location of the pulsar magne tosphere.\nKey words: instabilities - plasmas - pulsars: individual (PSR B0531+21) - radiation\nmechanisms: non-thermal\n1 INTRODUCTION\nThe recent observations of the Crab pulsar in the very\nhigh energy (VHE) domain with the VERITAS array of\natmospheric Cherenkov telescopes revealed pulsed γ-rays\nabove 100GeV (Aliu et al. 2011), which was later con-\nfirmed by measurments of the MAGIC Cherenkov telescope\n(Aleksi´ c et al. 2011, 2012). Prior to the work Aliu et al.\n(2011) the highest energy of the pulsed emission from the\nCrab pulsar was 25GeV (Aliu et al. 2008). The detection\nof such a high energy pulsed γ-rays cannot be explained\non the basis of current pulsar emission models. It is gener-\nally assumed that the VHE emission is produced either by\nthe Inverse Compton scattering or by the curvature radi-\nation. By analyzing the aforementioned emission processes\n(Machabeli & Osmanov 2009, 2010), we have found that for\nCrab pulsar’s magnetospheric parameters even very ener-\ngetic electrons are unable to produce the photon energies\nof the order of 25GeV. Studying the curvature radiation, it\nwas shown that the curvature drift instability efficiently re c-\ntifies the magnetic field lines, leading to a negligible role o f\nthe curvature emission process in the observed VHE domain\n(Osmanov et al. 2009). In previous work Chkheidze et al.\n(2011) we have explained the origin and the measured spec-\ntrum of the Crab pulsar in the high energy (HE) domain\n(0.01−25)GeV, relying on the pulsar emission model first\n⋆E-mail: nino.chkheidze@iliauni.edu.gedeveloped by Machabeli & Usov (1979). According to these\nworks, in the electron-positron plasma of a pulsar magne-\ntosphere the low frequency cyclotron modes, on the quasi-\nlinear evolution stage create conditions for generation of the\nHE synchrotron radiation.\nIt is well known that the distribution function of rela-\ntivistic particles is one dimensional at the pulsar surface , be-\ncause any transverse momenta ( p⊥) of relativistic electrons\nare lost in a very short time ( /lessorequalslant10−20s) via synchrotron\nemission in very strong magnetic fields. But plasma with an\nanisotropic one-dimensional distribution function is uns ta-\nble which inevitably leads to the wave excitation process.\nThe main mechanism of the wave generation in plasmas of\nthe pulsar magnetosphere is the cyclotron instability, whi ch\ndevelops at the light cylinder length-scales (a hypotheti-\ncal zone, where the linear velocity of rigid rotation exactl y\nequals the speed of light). During the quasi-linear stage of\nthe instability a diffusion of particles arises as along, als o\nacross the magnetic field lines. Therefore, plasma particle s\nacquire transverse momenta and, as a result, start to radiat e\nin the synchrotron regime.\nIn Chkheidze et al. (2011), it was shown that near the\nlight cylinder the radio waves are generated, provoking the\nre-creation of the pitch angles and the subsequent syn-\nchrotron radiation in the HE domain. Thus, in the frame-\nwork of this model generation of low and high frequency\nwaves is a simultaneous process and it takes place in one\nlocation of the pulsar magnetosphere. This explains the ob-\nc/circlecopyrt2011 RAS2\nserved coincidence of the HE emission pulses with the ra-\ndio signals. The recent VERITAS and MAGIC observations\nhaveshownthattheaforementioned coincidencetakesplace s\nin the VHE domain ( >100GeV) as well (Aliu et al. 2008,\n2011; Aleksi´ c et al. 2011). According to Aliu et al. (2011)\ndetection of pulsed γ-ray emission of the order of 100GeV\nrequires that the emission should be produced far out in the\nmagnetosphere. Thus, we suppose that the pulsed high and\nthe very high energy radiation of the Crab pulsar is gener-\nated through the synchrotron mechanism at the light cylin-\nder length-scales, switched on due to the quasi-linear diffu -\nsion. The resonant particles are the primary beam electrons\nwith the Lorentz-factor γb∼108−9, giving the synchrotron\nemission in the (0 .01−400)GeV energy domain.\nAccordingtoAliu et al. (2008) ajoint fittothe EGRET\n(10MeV to 10GeV) and MAGIC ( >25GeV) data predicted\na power-law spectrum with a generalized exponential shape\nfor the cutoff, described as Fǫ∝ǫ−αexp(−(ǫ/ǫ0)β), where\nα= 2.022±0.014. We provided a theoretical confirmation\nof the measured spectrum, which yielded β= 1.6 and the\ncutoff energy ǫ0= 23GeV (Chkheidze et al. 2011). Recent\nVERITAS observations (100 −400)GeV combined with the\nFermi-LAT data (0 .1−10)GeV favor a broken power law as\na parametrization of the spectral shape. The good fit results\nare also obtained if one uses a log-parabola function, but it\nfails to describe the spectrum below 500MeV. Although, the\nFermi-LAT and Magic data below 60 GeV can be equally\nwell parameterized by broken power law and exponential\ncutoff. In the energy range between 100GeV and 400GeV\nmeasured by VERITAS and MAGIC, the spectrum is well\ndescribed by a simple power law with the spectral index\nequal to 3.8 (Aliu et al. 2011; Aleksi´ c et al. 2012).\nThe paper is organized as follows. In Sect. 2 we describe\nthe emission model, in Sect. 3 we derive the theoretical syn-\nchrotron spectrum for the high and the very high energy\nγ-ray emission of the Crab pulsar and in Sect. 4 we discuss\nour results.\n2 EMISSION MODEL\nAny well known theory of pulsar emission suggests that,\nthe observed radiation is generated due to processes tak-\ning place in the electron-positron plasma. It is generally a s-\nsumed that the pulsar magnetosphere is filled by dense rela-\ntivistic electron-positron plasma with an anisotropic one -\ndimensional distribution function (see Fig. 1 from Arons\n(1981)) and consists of the following components: a bulk of\nplasma with an average Lorentz-factor γ∼γp, a tail on the\ndistribution function with γ∼γt, and the primary beam\nwithγ∼γb. The distribution function is one-dimensional\nand anisotropic and plasma becomes unstable, which might\ncause a wave excitation in the pulsar magnetosphere. The\nmain mechanism of wave generation in plasmas of the pul-\nsar magnetosphere is thecyclotron instability. The cyclot ron\nresonance condition can be written as (Kazbegi et al. 1992):\nω−k/bardblV/bardbl−kxux+ωB\nγr= 0, (1)\nwhereux=cV/bardblγr/ρωBis the drift velocity of the particles\ndue to curvature of the field lines with the curvature radius\nρ,ωB≡eB/mcis the cyclotron frequency, eandmare theelectron’s charge and the rest mass, cis the speed of light,\nkxis the wave vector’s component along the drift and B\nis the magnetic field induction. During the wave generation\nprocess, one also has a simultaneous feedback of these waves\non the resonant electrons (Vedenov et al. 1961). This mech-\nanism is described by the quasi-linear diffusion (QLD), lead -\ning to a diffusion of particles as along as across the magnetic\nfield lines. Therefore, resonant particles acquire transve rse\nmomenta (pitch angles) and, as a result, start to radiate\nthrough the synchrotron mechanism.\nThe wave excitation leads to redistribution process of\nthe resonant particles via the QLD. The kinetic equation\nfor the distribution function of the resonant electrons can\nbe written as (Chkheidze et al. 2011):\n∂f0\n∂t+∂\n∂p/bardbl/braceleftbig\nF/bardblf0/bracerightbig\n+1\np⊥∂\n∂p⊥/braceleftbig\np⊥F⊥f0/bracerightbig\n=\n=1\np⊥∂\n∂p⊥/braceleftbigg\np⊥D⊥,⊥∂f0\n∂p⊥/bracerightbigg\n. (2)\nwhere\nF⊥=−αsp⊥\np/bardbl/parenleftbigg\n1+p2\n⊥\nm2c2/parenrightbigg\n, F /bardbl=−αs\nm2c2p2\n⊥,(3)\nare the transversal and longitudinal components of the syn-\nchrotron radiation reaction force, where αs= 2e2ω2\nB/3c2\nandD⊥,⊥is the transverse diffusion coefficient which is de-\nfined as follows (Chkheidze et al. 2011)\nD⊥,⊥=πe4np\n8mcω2\nBγ3p|Ek|2. (4)\nHere|Ek|2is the density of electric energy in the waves\nand its value can be estimated from the expression |Ek|2≈\nmc2nbγbc/2ωc, whereωcis the frequency of the cyclotron\nwaves. From Eq. (1) it follows that\nωc≈ωB\nδγr, (5)\nwhereδ=ω2\np/(4ω2\nBγ3\np),ωp≡/radicalbig\n4πnpe2/mis the plasma\nfrequency and npis the plasma density.\nThe transversal QLD increases the pitch-angle, whereas\nforceF⊥resists this process, leading to a stationary state\n(∂f/∂t= 0). The pitch-angles acquired by resonant electrons\nduring the process of the QLD satisfies ψ=p⊥/p/bardbl≪1.\nThus, one can assume that ∂/∂p⊥>> ∂/∂p /bardbl. In this\ncase the solution of Eq.(2) gives the distribution function\nof the resonant particles by their perpendicular momenta\n(Chkheidze et al. 2011)\nf(p⊥) =Cexp/parenleftbigg/integraldisplayF⊥\nD⊥,⊥dp⊥/parenrightbigg\n=Ce−/parenleftbigg\np⊥\np⊥0/parenrightbigg4\n, (6)\nwhere\np⊥0≈π1/2\nBγ2p/parenleftbigg3m9c11γ5\nb\n32e6P3/parenrightbigg1/4\n. (7)\nAnd for the mean value of the pitch angle we find ψ0≈\np⊥0/p/bardbl≃10−6. Synchrotron emission is generated as the\nresult of appearance of pitch angles.\nThe synchrotron emission flux of the set of electrons in\nthe framework of the present emission scenario is written as\n(see Chkheidze et al. (2011))\nc/circlecopyrt2011 RAS, MNRAS 000, 1–??3\nFǫ∝/integraldisplayp/bardblmax\np/bardblminf/bardbl(p/bardbl)Bψ0ǫ\nǫm/bracketleftBigg/integraldisplay∞\nǫ/ǫmK5/3(z)dz/bracketrightBigg\ndp/bardbl.(8)\nHeref/bardbl(p/bardbl) is the longitudinal distribution function of elec-\ntrons,ǫm≈5·10−18Bψ0γ2GeV is the photon energy of the\nmaximum of synchrotron spectrum of a single electron and\nK5/3(z) is the Macdonald function. After substituting the\nmean value of the pitch-angle in the above expression for\nǫm, we get\nǫm≃5·10−18π1/2\nγ2p/parenleftbigg3m5c7γ9\nb\n4e6P3/parenrightbigg1/4\n. (9)\nAccordingly, the beam electrons should have γb≃(6·\n108−109) to radiate the photons in the energy domain\n∼(10−100)GeV energy. The gap models provide the\nLorentz factors up to 107, which is not enough to explain\nthe detected pulsed emission. Consequently, the additiona l\nparticle acceleration mechanism should be invoked to accel -\nerate the fastest electrons to even higher energies.\nAccording to Aliu et al. (2011) the observations of the\nCrab pulsar indicates that the HE pulsed emission should\nbe produced far out in the magnetosphere. Therefore, we\nassume that the Langmuir waves (L) generated via the\ntwo-stream instability at the light cylinder length-scale s\nundergo Landau damping on the fastest beam electrons,\nwhich results in their effective acceleration. Excitation o f\nL waves through the two-stream instability in the rela-\ntivistic electron-positron plasma is considered in a serie s\nof works (see e.g., Lominadze & Mikhailovskii (1979); Usov\n(1987); Asseo & Melikidze (1998); Ursov & Usov (1988);\nGogoberidze et al. (2008)), where it is assumed that the in-\nstability develops due to overlapping of the fast and slow\npair plasma particles at distances r∼108cm from the star\nsurface. For typical parameters of pair plasma in the pulsar\nmagnetospehers the growth rate of the instability is quite\nsufficient (see Gogoberidze et al. (2008)), that inevitably\nprovides existence of L waves in the vicinity of the light\ncylinder zone. The phase velocity of the excited L waves is\nasymptotically close tothespeedoflight (Gogoberidze et a l.\n2008). Therefore, these waves can be only damped on the\nfastest electrons of the primary beam, which velocity equal s\nthe phase-velocity of the waves and the distribution functi on\nsatisfies∂fb0/∂p/bardbl<0. Taking into account the equipartition\nof energy among the plasma components npγp=nbγb/2,\none can estimate the energy density of L waves as the\nhalf of the energy density of the primary beam particles.\nThrough the Landau damping process the energy of the\nwaves is transferred to the small fraction of the beam elec-\ntrons with the highest Lorentz factors. If we assign the\ndensity of these electrons as n∗and the Lorentz factors as\nγ∗, and equate the energy densities of particles and the L\nwaves, we will get 2 n∗/nb≈γb/γ∗. The total density of\nthe beam electrons is equal to the Goldreich-Julian density\nnb=B/Pce≈2·107cm−3(Goldreich & Julian 1969) and\nalso if we take into account that γb∼107andγ∗∼109,\nwe findn∗∼105cm−3. As we see, if the wave energy is\ntransferred to the fastest beam electrons, which number is\ntwo orders of magnitude smaller than the total number of\nthe primary beam particles, they will gain the Lorentz fac-\ntors up to 109. The observational fact, that the emission fluxabove25GeV decreases shouldbe caused byreducednumber\nof emitting particles with the highest Lorentz factors.\nIt should be mentioned that during the Landau damp-\ning process the beam distribution function will be alongate d\nandwill form ahighenergy’tail’on thedistributionfuncti on.\nThe final shape of the distribution function after the quasi-\nlinear relaxation is the plateau, and the stationary state i s\nreached. But in our case this might not be achieved as in the\nsame region where the L waves are damped the cyclotron in-\nstabilityis developed,which involvesthebeam electrons i nto\nthe cyclotron resonance process. This complicated process\ncausing the redistribution of the resonant particles needs a\nmore detailed investigation, which we plan to perform in\nour future work. In the present paper, we estimate the final\nshape of the distribution function which provides explana-\ntion of the measured spectrum.\nThe beam particles lose their energy through the syn-\nchrotronradiation, whichsets theupperlimit ontheLorent z\nfactors that can be achieved during their acceleration pro-\ncess (de Jager et al. 1996). The maximum achievable value\nofγcan be estimated by equating the synchrotron radiative\nlooses to the power of the emitting particles gained through\nthe acceleration process, which in our case can be written\nas:\n2\n3e4B2ψ2\n0γ2\nm2c3=mc2γΓLD. (10)\nHere\nΓLD=nbγbωb\nnpγ5/2\np, (11)\nis the Landau damping rate, where ωb=/radicalbig\n4πe2nb/m\n(Volokitin et al. 1987). The estimations show, that the\nLorentz factor of the beam electrons that can be reached\nthrough the acceleration process γ/lessorsimilar1019. Consequently,\nthe limit on the energy of synchrotron photons in our case\nshould beǫmax∼1054eV (see Eq. (9)), which inevitably al-\nlows generation of the observed ∼100GeV photons from the\nCrab pulsar through the synchrotron emission mechanism.\n3 SPECTRUM OF THE SYNCHROTRON\nRADIATION\nTo obtain the synchrotron emission spectrum in our case,\nwe need to solve the integral (8). For this reason, first let us\nfind the parallel distribution function of the beam electron s\nf/bardblduring the QLD process of the cyclotron instability. By\nmultiplying both sides of Eq. (2) on p⊥, integrating it over\np⊥and taking into account that the distribution function\nvanishes at the boundaries of integration, Eq. (2) reduces t o\n∂f/bardbl\n∂t=∂\n∂p/bardbl/parenleftBigαs\nm2c2π1/2p2\n⊥0f/bardbl/parenrightBig\n. (12)\nForγψ≪1010, a magnetic field inhomogeneity does\nnot affect the process of wave excitation. The equation that\ndescribes the cyclotron noise level, in this case, has the fo rm\n(Lominadze et al. 1983)\n∂|Ek|2\n∂t= 2Γc|Ek|2, (13)\nwhere\nc/circlecopyrt2011 RAS, MNRAS 000, 1–??4\nΓc=π2e2\nk/bardblf/bardbl(pres), (14)\nis the growth rate of the instability. Here k/bardblcan be found\nfrom the resonance condition (1)\nk/bardblres≈ωB\ncδγres. (15)\nCombining Eqs. (10) and (11) one finds\n∂\n∂t/braceleftBigg\nf/bardbl−α∂\n∂p/bardbl/parenleftBigg\n|Ek|\np1/2\n/bardbl/parenrightBigg/bracerightBigg\n= 0, (16)\nα=/parenleftbigg4\n3e2\nπ5c5ω6\nBγ3\np\nω2p/parenrightbigg1/4\n. (17)\nConsequently, one can write\n/braceleftBigg\nf/bardbl−α∂\n∂p/bardbl/parenleftBigg\n|Ek|\np1/2\n/bardbl/parenrightBigg/bracerightBigg\n=const. (18)\nTakingintoaccountthatfor theinitial moment(themoment\nwhen the cyclotron instability arises) the energy density o f\ncyclotron waves equals zero, the corresponding expression\nwrites as\nf/bardbl−α∂\n∂p/bardbl/parenleftBigg\n|Ek|\np1/2\n/bardbl/parenrightBigg\n=f/bardbl0. (19)\nLet us assume that |Ek| ∝γ−m(as there is no direct way\nto calculate the dependence |Ek(γ)|, we can only make an\nassumption and check its plausibility by fitting the theoret -\nical emission spectrum with the observed one), in this case\nfor the parallel distribution function we will have\nf/bardbl∝f/bardbl0+γ−m−3\n2. (20)\nThe initial distribution f/bardbl0of the beam particles in this case\nis the redistributed one after the Landau damping process.\nThe final shape of the distribution function of resonant par-\nticles via the Landau damping when the stationary state is\nreached is the plateau. But in our case this might not be\nreached as in the same region develops the cyclotron insta-\nbility. Thus we consider f/bardbl0∝γ−n, wherenis not close\nto zero. Consequently, the parallel distribution of the bea m\nelectrons is proportional to two power-law function with th e\nindexesm+3/2 andn.\nThe effective value of the pitch angle depends on |Ek|2\nas follows (Chkheidze et al. 2011)\nψ0=1\n2ωB/parenleftBigg\n3m2c3\np3\n/bardblω2\np\nγ3p|Ek|2/parenrightBigg1/4\n. (21)\nUsing expressions (8), (18) and (19), and replacing the inte -\ngration variable p/bardblbyx=ǫ/ǫm, we will get the synchrotron\nemission spectrum\nFǫ∝ǫ−2m+4n−1\n5−2m/braceleftbigg\nG1/parenleftbiggǫ\nǫm/parenrightbigg\nmax−G1/parenleftbiggǫ\nǫm/parenrightbigg\nmin/bracerightbigg\n+\n+ǫ−6m+5\n5−2m/braceleftbigg\nG2/parenleftbiggǫ\nǫm/parenrightbigg\nmax−G2/parenleftbiggǫ\nǫm/parenrightbigg\nmin/bracerightbigg\n,(22)\nwhere\nG1(y) =/integraldisplay∞\nyx2m+4n−1\n5−2m/bracketleftbigg/integraldisplay∞\nxK5/3(z)dz/bracketrightbigg\ndxG2(y) =/integraldisplay∞\nyx6m+5\n5−2m/bracketleftbigg/integraldisplay∞\nxK5/3(z)dz/bracketrightbigg\ndx. (23)\nThe energy of the beam particles vary in a broad range\nγb∼106−109in which case, we have ( ǫ/ǫm)max≪1\nand (ǫ/ǫm)min≫1. Under such conditions the functions\nG1(y)≈G2(y)≈G(0). Consequently, we can assume that\nthe synchrotron emission spectrum (Eq. (20)) is propor-\ntional to two power-law functions ǫ−2m+4n−1\n5−2mandǫ−6m+5\n5−2m.\nAccording to VERITAS observations the spectrum mea-\nsured in the energy domain (100 −400)GeV is well described\nbypower-lawwiththespectral indexequalto3 .8(Aliu et al.\n2011). When m≈1, we have −(6m+5)/(5−2m) =−3.8.\nAt the same time, when m= 1 andn= 1.2, we find\n−(2m+4n−1)/(5−2m) =−2 that is in a good agreement\nwith the observations in the 10MeV - 25GeV energydomain,\nwhich shows the power-law spectrum F(ǫ)∝ǫ−2.022±0.014\n(Aliu et al. 2008).\n4 DISCUSSION\nThe interesting observational feature of the Crab pul-\nsar of the coincidence of pulse-phases from different fre-\nquency bands, ranging from radio to VHE gamma-rays\n(Manchester & Taylor 1980; Aliu et al. 2008, 2011) implies\nthat generation of these waves occur in one location of\nthe pulsar magnetosphere. This consideration automatical ly\nexcludes the generally accepted HE emission mechanisms,\nthe Inverse Compton (IC) scattering and the curvature\nradiation, which are not localized (Machabeli & Osmanov\n2009, 2010). These particular issues have been considered\nby Machabeli & Osmanov (2010). Studying the curvature\nradiation it was found that the curvature drift instability is\nefficient enough to rectify the magnetic field lines (curvatur e\ntends to zero) in the region of the generation of high and the\nVHE emission, making the curvature emission process neg-\nligible. At the same time by analyzing the IC scattering, it\nwas found that for Crab pulsar’s magnetospheric parame-\nters even very energetic electrons are unable to produce the\nobserved HE photons.\nIn Lyutikov et al. (2011) it is argued that the main gen-\nerationmechanism oftheVHEemission oftheCrabpulsar is\nthe IC scattering of the soft UV photons by the secondary\nplasma particles. Particularly, it is assumed that the sec-\nondary plasma, which is produced due to cascades in the\nouter gaps of the magnetosphere is responsible for the soft\nUV emission via the synchrotron mechanism. This radiation\nplays the target field role for the IC scattering process. As\na result, the VHE γ-ray emission is produced extending to\nhundreds of GeV. Let us consider the equation that gives\nthe frequency of the photon after the IC scattering (e.g.\nRybicki & Lightman (1979)):\nω′=ω(1−βcosθ)\n1−βcosθ′+(1−cosθ′′)/planckover2pi1ω/γmc2, (24)\nwhereωis the frequency before scattering, β≡υ/c,θ=\n(/hatwidestPK),θ′= (/hatwidestPK′) andθ′′= (/hatwidestKK′) (byPwe denoted the\nmomentum of relativistic electrons before scattering, Kand\nK′denotes the three momentum of photon before and af-\nter scattering, respectively). The emission maximum comes\nalong the magnetic field lines. The pitch angles of the rela-\nc/circlecopyrt2011 RAS, MNRAS 000, 1–??5\ntivistic electrons moving along the magnetic field lines are\nvery small ( ψ∼10−6see Eq.(7)), accordingly as the UV\nemission is generated through the synchrotron regime, the\nangleθshould also be very small. Since we observe the well-\nlocalized pulses of the VHE emission, the angle θ′≪1. Tak-\ning into account the observational fact of the phase coinci-\ndence of signals from different frequency bands, one should\nassume that θ≈θ′. At the same time the coincidence of\npulse signals of UV and the VHE emission gives θ′′= 0.\nConsequently, the IC scattering of the soft UV photons by\nthe secondary plasma electrons in case of the Crab pulsar\nshould only cause the redistribution of the electrons, butc an\nnot providethe increase in theemission frequency,especia lly\nup to the VHE band. For significantly increasing the photon\nenergy by the IC scattering processes without violating the\ncondition of the pulse-phase coincidence, the angle θmust\nbe large enough, which can not be provided by our model.\nApparently, the IC scattering should play the main role in\nthe generation of the high energy emission for pulsars with\nthe large pulse profiles in soft energy domains.\nThe emission model proposed in the present paper im-\nplicitly explains the observed pulse-phase coincidence of low\n(radio) and high frequency (10MeV-400GeV) waves, as their\ngeneration is a simultaneous process and it takes place in\nthe same place of the pulsar magnetosphere. In previous\nworks (Machabeli & Osmanov 2010; Chkheidze et al. 2011)\nwe applied this model to explain the HE (0.01-25GeV)\npulsed emission of the Crab pulsar observed by the MAGIC\nCherenkov Telescope (Aliu et al. 2008). It was found that\non the light cylinder length-scales the cyclotron instabil ity\nis arisen, which on the quasi-linear stage of the evolution\ncauses re-creation of the pitch angles and as the result the\nsynchrotron radiation mechanism is switched on. We assume\nthat the source of the high and the VHE pulsed emission of\nthe Crab pulsar is the synchrotron radiation of the ultrarel -\nativistic primary beam electrons. To explain the observed\nhigh frequency gamma-rays by synchrotron mechanism the\nLorentz factors of the emitting particles should be of the or -\nder of 108−109(see Eq. (9)). The highest Lorentz factor for\nthe typical pulsar is ∼107. Thus, we assume that such an\neffective particle acceleration is caused by existence of th e\nLangmuir waves with the phase velocities υϕ/lessorsimilarcclose to the\nvelocities of the fastest beam electrons in the region close to\nthe light cylinder. Consequently, the L (electrostatic) wa ves\nare efficiently damped on the most energetic primary beam\nelectrons. The Landau damping causes the growth of a HE\ntail on the distribution function of the resonant electrons\nand inevitably throws the most energetic particles to highe r\nLorentz factors (up to γ∼109). At the same time the beam\nelectrons acquire the pitch angles due to the cyclotron in-\nteraction with the transverse waves, which causes the syn-\nchrotron radiation processes giving the observed high and\nthe VHE emission up to 400GeV. The distribution func-\ntion tends to form the plateau (due to Landau damping),\nthough the number of processes impede this. The reaction\nforce of the synchrotron emission, scattering of L waves and\nthe Compton scattering of photons on the beam particles\nalso take place in process of formation of the distribution\nfunction of beam electrons. As the result, it is unlikely for\nthe distribution to reach the shape of plateau and thus, we\nrepresent it as f/bardbl0∝γ−n.\nThe calculation of the synchrotron emission spectrumby taking into account the processes described above (see\nEq. (20)) and matching it with the observations shows that\nthe emission spectrum in the 10MeV-25GeV energy domain\ndepends on the power-law index nof distribution function of\nthe beam particles. At the same time the emission spectrum\nin (100−400)GeV energy domain does not depend on nbut\nonly depends on m(see Eq. (18)). When m= 1 andn= 1.2\nthe emission spectrum well matches the measured one in\nboth high (0 .01−25GeV) and the very high (100 −400GeV)\nenergy domains. In previous paper (Chkheidze et al. 2011)\nexplaining the HE (10MeV-25GeV) spectrum we obtained\nthe power-law function with the exponential cutoff Fǫ∝\nǫ−2exp[−(ǫ/23)1.6], asγb∼108were the highest considered\nLorentz factors of the emitting electrons. We assume that\ndetection of the exponential cutoff at higher ( >400GeV)\nphoton energies can not be excluded, the exact location of\nthe cutoff energy is defined by the highest energy of the\nbeam electrons that can be reached through the Landau\ndamping before the particles reach the light cylinder. This\nparticular problem needs more detailed investigation, whi ch\nis the topic of our future work.\nREFERENCES\nAleksi´ c J. et al., 2011, ApJ, 744, 43\nAleksi´ c J. et al., 2012, A&A, 540, 69\nAliu E. et al., 2008, Sci, 322, 1221A\nAliu E. et al., 2011, Sci., 334, 69\nArons J., 1981, in Proc. Varenna Summer School and\nWorkhop on Plasma Astrophysics, ESA, 273\nAsseo E., Melikidze G. I., 1998, MNRAS, 301, 59\nChkheidze, N., Machabeli, G., Osmanov, Z., 2011, ApJ,\n730, 62\nde Jager, O. C., Harding, A. K., Michelson, P. F., Nel, H.\nI., Nolan, P. L., Sreekumar, P., & Thompson, D. J. 1996,\nApJ, 457, 253\nGogoberidze G., Machabeli G. Z., Usov V. V., 2008,\nPhRvE, 77, 7402\nGoldreich, P., & Julian, W. H. 1969, ApJ, 157, 869\nKazbegi A.Z., Machabeli G.Z & Melikidze G.I., 1992, in\nProc. IAUCollog. 128, TheMagnetospheric Structureand\nEmission Mechanisms of RadioPulsars, ed. T.H. Hankins,\nJ.M. Rankin & J.A. Gil (Zielona Gora: Pedagogical Univ.\nPress), 232\nLominadze D. G., Mikhailovskii A. B., 1979, ZhETF, 76,\n959\nLominadze J.G., Machabeli G. Z., & Usov V. V., 1983,\nAp&SS, 90, 19L\nLyutikov et al., 2011, arXiv:1108.3824\nMachabeli G.Z. & Usov V.V., 1979, AZhh Pis’ma, 5, 445\nMachabeli G. & Osmanov Z., 2009, ApJ, 700, 114\nMachabeli G. & Osmanov Z., 2010, ApJ, 709, 547\nManchester R.N. & Taylor J.H., 1980, Pulsars, F.H. Free-\nman and Company\nOsmanov, Z., Shapakidze, D., & Machbeli, G., 2009, A&A,\n509, 19\nRybicki G.B. & Lightman A. P., 1979, Radiative Processes\nin Astrophysics. Wiley, New York\nUrsov V. N., Usov V. V., 1988, Ap&SS, 140, 325\nUsov V. V., 1987, ApJ, 320, 333\nc/circlecopyrt2011 RAS, MNRAS 000, 1–??6\nVedenov A.A., Velikhov E.P. & Sagdeev R.Z., 1961, Soviet\nPhysics Uspekhi, Volume 4, Issue 2, 332\nVolokitin, A.S., Krasnoselskikh, V.V. & Machabeli, G.Z.,\n1987, Soviet Journal of Plasma Physics, 11, 310\nc/circlecopyrt2011 RAS, MNRAS 000, 1–??" }, { "title": "2203.06360v1.Asymptotic_expansion_of_solutions_to_the_wave_equation_with_space_dependent_damping.pdf", "content": "arXiv:2203.06360v1 [math.AP] 12 Mar 2022ASYMPTOTIC EXPANSION OF SOLUTIONS TO THE WAVE\nEQUATION WITH SPACE-DEPENDENT DAMPING\nMOTOHIRO SOBAJIMA AND YUTA WAKASUGI\nAbstract. We study the large time behavior of solutions to the wave equa tion\nwith space-dependent damping in an exterior domain. We show that if the\ndamping is effective, then the solution is asymptotically ex panded in terms\nof solutions of corresponding parabolic equations. The mai n idea to obtain\nthe asymptotic expansion is the decomposition of the soluti on of the damped\nwave equation into the solution of the corresponding parabo lic problem and\nthe time derivative of the solution of the damped wave equati on with certain\ninhomogeneous term and initial data. The estimate of the rem ainder term is\nan application of weighted energy methods with suitable sup ersolutions of the\ncorresponding parabolic problem.\nContents\n1. Introduction 2\n1.1. Problem and backgrounds 2\n1.2. Main result 5\n1.3. A rough descripsion of strategy 6\n1.4. Construction of the paper 6\n1.5. Notations 7\n2. Preliminaries 7\n2.1. Weight functions 7\n3. Justification of the decomposition 10\n3.1. Well-posedness and regularity of solutions for the damped wave\nequation 11\n3.2. Regularity of solutions for the corresponding heat equation 12\n3.3. A decomposition lemma 14\n3.4. Derivation of the asymptotic expansion 14\n4. Energy estimates for the heat equation 15\n5. Energy estimates for the damped wave equation 19\n5.1. First order energy estimates 19\n5.2. Higher order energy estimates 24\n6. Proof of the asymptotic expansion 28\nAcknowledgements 31\nReferences 31\n2020Mathematics Subject Classification. 35L20; 35C20; 35B40.\nKey words and phrases. wave equation, space-dependent damping, asymptotic expan sion.\n12 M. SOBAJIMA AND Y. WAKASUGI\n1.Introduction\n1.1.Problem and backgrounds. Let Ω be an exterior domain with a smooth\nboundary∂Ω inRNwithN≥2, or Ω = RNwithN≥1. We consider the initial-\nboundary value problem of the wave equation with space-dependen t damping\n\n\n∂2\ntu−∆u+a(x)∂tu= 0, x ∈Ω,t>0,\nu(x,t) = 0, x ∈∂Ω,t>0,\nu(x,0) =u0(x), ∂tu(x,0) =u1(x), x∈Ω.(1.1)\nHere,u=u(x,t) is a real-valuedunknownfunction, and a(x) denotes the coefficient\nof the damping term.\nWeassumethat a(x) isasmoothpositivefunctionon RNhavingboundedderiva-\ntives and satisfying\nlim\n|x|→∞|x|αa(x) =a0 (1.2)\nwith some constants α∈[0,1) anda0>0. Here, the precise meaning of (1.2)\nis limr→∞sup|x|>r||x|αa(x)−a0|= 0, that is, the convergence is uniform in the\ndirection. In this case, the damping is called effective, and, as we will s ee later,\nthe asymptotic behavior of the solution is closely related to a certain corresponding\nparabolic problem. Here, we remark that it is sufficient that a(x) is defined on ¯Ω,\nbecause we can extend it to RNso that it has the same property as above.\nThe initial data ( u0,u1) are assumed to belong to ( H2(Ω)∩H1\n0(Ω))×H1\n0(Ω).\nThen, it is known that (1.1) admits a unique solution\nu∈C([0,∞);H2(Ω))∩C1([0,∞);H1\n0(Ω))∩C2([0,∞);L2(Ω))\n(see [7, Theorem 2]). In our main result, we shall put stronger assu mptions on the\ndata.\nThe aim of this paper is to prove the asymptotic expansion of the solu tion as\ntime tends to infinity. In particular, we show that the solution uis asymptotically\nexpanded in terms of a sequence of solutions to corresponding par abolic equations\nwith certain inhomogeneous terms.\nThe asymptotic behavior of solutions to the damped wave equation h as long\nhistory after a pioneering work by Matsumura[15]. He studied the Ca uchy problem\nof the wave equation with constant damping\n∂2\ntu−∆u+∂tu= 0,(x,t)∈RN×(0,∞), (1.3)\nand applied the Fourier transform to obtain the L∞andL2estimates of solutions.\nIn particular, he showed that the decay rates are the same as tho se of the corre-\nsponding heat equation\n∂tv−∆v= 0,(x,t)∈RN×(0,∞). (1.4)\nAfter that, the precise asymptotic profile ofsolutions werestudie d by Hsiao and Liu\n[6] for the hyperbolic conservation laws with damping, and by Karch [1 3] and by\nYang and Milani [47] for (1.3), and the so-called diffusion phenomena was proved,\nthat is, the solution uof (1.3) is asymptotically approximated by a solution of the\nheatequation(1.4) astime tendstoinfinity. Moredetailedasymptot icbehaviorwas\nstudied by many mathematicians, and we refer the reader to [5, 14, 20, 21, 29] for\nthe asymptotic behavior involving the decomposition of solution into t he heat-part\nand wave-part.WAVE EQUATION WITH SPACE-DEPENDENT DAMPING 3\nFor the higher order asymptotic expansions of the Cauchy problem of (1.3),\nGallay and Raugel [4] determined the second order expansion when N= 1 by the\nmethod of scaling-variables. Moreover, by Fourier transform, Ta keda [37] studied\nthe caseN≤3 and obtained the expansion of any order in terms of the Gaussian.\nMichihisa [17] also gave another expression of expansion for any N≥1 by the\nFourier transform method.\nOn the other hand, the asymptotic behavior of solutions to the initia l-boundary\nvalue problem of the wave equation with constant damping in an exter ior domain\nis also well-studied. This problem firstly studied by Ikehata [8], and he p roved\nthe diffusion phenomena, that is, the asymptotic profile of solution t o the extorior\nproblem is given by the exterior heat semigroup with Dirichlet boundar y condition.\nAfterthat, thisresultwasextendedbyIkehataandNishihara[10], ChillandHaraux\n[3], and Radu, Todorova, and Yordanov [27] to the abstract proble m\nu′′(t)+Au(t)+u′(t) = 0, t>0, (1.5)\nwhereAis a nonnegative self-adjoint operator in a Hilbert space. Recently, the\nfirst author [31] proved the higher order asymptotic expansion of the solution to\n(1.5) in terms of the solutions of the corresponding first order equ ation. Radu,\nTodorova, and Yordanov [28] studied the diffusion phenomena for m ore general\nabstract equation\nCu′′(t)+Bu(t)+u′(t) = 0, t>0\nwith a nonnegative self-adjoint operator Band a positive bounded operator C\nby the method of diffusion approximation. Nishiyama [23] also studied a similar\nproblem by the method of resolvent estimates.\nWe also refer the reader to Wirth [41, 42, 43, 44, 45], Yamazaki [46], and [40],\nfor the diffusion phenomena of the wave equation with time dependen t damping.\nFortheinitial-boundaryvalueproblemofthewaveequationwithspac e-dependent\ndamping (1.1), under the assumption of (1.2), it is expected that th e damping is\nclassified in the following way:\n•(scattering)When α>1,thesolutionbehaveslikethatofthewaveequation\nwithout damping.\n•(effective) When α<1, the solution behaves like that of the corresponding\nheat equation.\n•(critical) When α= 1, the equation is formally invariant under hyperbolic\nscaling and the behavior of the solution may also depend on the const ant\na0.\nThe scattering case α >1 when Ω = RN(N/ne}ationslash= 2) or Ω ⊂RNis an exterior\ndomain (N≥3) was studied by Mochizuki [18], Mochizuki and Nakazawa [19], and\nMatsuyama [16], and they proved that there exist initial data such t hat the energy\nof the corresponding solution does not decay to zero, and it appro ached a solution\nof the wave equation without damping in the energy norm. The case N= 2 seems\nstill open.\nThe critical case α= 1 with the assumption on a(x) replaced by b0/an}b∇acketle{tx/an}b∇acket∇i}ht−1≤\na(x)≤b1/an}b∇acketle{tx/an}b∇acket∇i}ht−1(b0,b1>0) was studied by Ikehata, Todorova and Yordanov [11].\nThey proved that, when Ω = RNwithN≥3 and the initial data are in C∞\n0(RN),\nthe energy of the solution decays as O(t−b0) if 1< b0< N, andO(t−N+δ) with\narbitrarysmall δ>0 ifb0≥N. Moreover,the decayrate O(t−b0) when 10, and the initial data belong\ntoC∞\n0(RN), then we have m(a) =N−α\n2−α, and the following estimates hold:\n/ba∇dblu(t)/ba∇dblL2≤Cδ(1+t)−N−α\n2(2−α)+α\n2(2−α)+δ,\n/ba∇dbl(∂tu(t),∇u(t))/ba∇dblL2≤Cδ(1+t)−N−α\n2(2−α)−1\n2+δ,\nwhereδ >0 is an arbitrary small loss of decay. Radu, Todorova, and Yordano v\n[25, 26] studied the energy decay of higher order derivatives and e xtended the result\nto more general second-order hyperbolic equations. The assump tion of the radial\nsymmetry on a(x) was removed by the authors [33] by modifying the function A(x)\nabove. Moreover, the authors [39, 32, 34] proved the diffusion ph enomena in the\ncase ofα∈(−∞,1) and exterior domains. The asymptotic profile of solution uis\ngiven by the corresponding parabolic problem\n\na(x)∂tV0−∆V0= 0, x ∈Ω,t>0,\nV0(x,t) = 0, x ∈∂Ω,t>0,\nV0(x,0) =u0(x)+a(x)−1u1(x), x∈Ω.(1.6)\nRecently, the authors [35, 36] developed a different kind of weight ed energy\nmethod applicable to a wider class of initial data including polynomially dec aying\nfunctions. Roughly speaking, the suggested weight functions for m the inverse of\nthe self-similar solutions Φ βof the equation |x|−α∂tΦ−∆Φ = 0 given by\nΦβ(x,t) =t−βϕβ(ξ(x,t)),\nwhereβ∈[0,N−α\n2−α) is a parameter,\nϕβ(z) =e−zM/parenleftbiggN−α\n2−α−β,N−α\n2−α;z/parenrightbigg\n, ξ(x,t) =|x|2−α\n(2−α)2t\nwith the Kummer confluent hypergeometric function M(b,c;z) (see Section 2 for\nthe precise definition). Moreover, the relation between the order of the weight of\ninitial data and the decay rates of the solution was revealed. It is wo rthly noticing\nthat these weight functions have a polynomial growth which enables us to take\ninitial data having a polynomial decay, and the endpoint β=N−α\n2−αprovides the\nexponential type solution\nΦN−α\n2−α(x,t) =t−N−α\n2−αexp/parenleftbigg\n−|x|2−α\n(2−α)t/parenrightbigg\nwhich corresponds to the exponential type weight function introd uced in [38].\nAs mentioned above, the sharp decay estimates of solutions and th e diffusion\nphenomena for the effective case α <1 is now known very well. In contrast, the\nhigher order asymptotic expansion of the solution remains open.WAVE EQUATION WITH SPACE-DEPENDENT DAMPING 5\nHere, we mention a result by Orive, Zuazua, and Pazoto [24] and Joly and Royer\n[12] for periodic and asymptotically periodic coefficient cases from diff erent aspects.\nFor the exterior problem with decaying damping such as (1.1), it seem s difficult to\napply the Fourier analysis which is the strong tool for the whole spac e case, and\nto apply the spectral analysis because of the appearance of unbo unded diffusion\noperators and non-comutativity.\nIn the present paper we introduce a new method (inspired by [31]) t o reach the\nasymptotic expansion in terms of solutions of the corresponding pa rabolic equation\nwith certain inhomogeneous terms (see a description of the idea in Su bsection 1.3).\n1.2.Main result. The followingis our main result which desribes the higher order\nasymptotic expansion of the solution uof (1.1).\nTheorem 1.1. Letnbe a nonnegative integer. Assume (1.2)for someα∈[0,1)\nanda0>0. Ifn+10such that the following holds:\nSuppose the initial data u0andu1satisfy\nu0∈Hs+1,m(Ω)∩Hs,m\n0(Ω), u1∈Hs,m\n0(Ω). (1.7)\nThen there exist profiles /tildewideV1...,/tildewideVn∈C([0,∞);L2(Ω))and a positive constant C\nsuch that the solution uof(1.1)satisfies\n/vextenddouble/vextenddouble/vextenddouble/vextenddouble/vextenddouble/vextenddoubleu(t)−V0(t)−n/summationdisplay\nj=1/tildewideVj(t)/vextenddouble/vextenddouble/vextenddouble/vextenddouble/vextenddouble/vextenddouble\nL2(Ω)≤C(1+t)−λ\n2−(2n+1)(1−α)\n2−α+α\n2(2−α)(1.8)\nfort>0, whereV0is given by (1.6). Moreover, the profiles are successively deter-\nmined as /tildewideVj=∂j\ntVjwith the unique solutions Vjof\n\n\na(x)∂tVj−∆Vj=−∂tVj−1, x∈Ω,t>0,\nVj(x,t) = 0, x ∈∂Ω,t>0,\nVj(x,0) =−(−a(x))−j−1u1(x), x∈Ω.(1.9)\nforj= 1,...,n.\nRemark 1.2. For eachVj(j= 0,1,...,n), we also have\n/ba∇dbl∂j\ntVj(t)/ba∇dblL2(Ω)≤C(1+t)−λ\n2−2j(1−α)\n2−α+α\n2(2−α)\n(see Section 6). We remark that when a(x)≡1, the expansion in Theorem 1.1\ncoincides with the known result [31].\nRemark 1.3. One can also represent the profiles /tildewideVjforj= 1,...nin terms of the\nsemigroupetLgenerated by L=a−1∆. For instance, the second profile /tildewideV1can be\nwritten as\n/tildewideV1(t) =−LetL[a−2u1]−a−1LetL[u0+a−1u1]−/integraldisplayt\n0Le(t−s)L/bracketleftBig\na−1LesL[u0+a−1u1]/bracketrightBig\nds.\nIfa≡1, thenet∆anda−1commutes, and therefore, the above description can be\nsimplified to /tildewideV1(t) =−∆et∆u1−∆(1+t∆)et∆[u0+u1]as in[31], but the semigroup\netLanda−1do not commute in general. To avoid such a complicated situat ion, we\nhave chosen the parabolic equations (1.8)for the determinination of the profiles /tildewideVj.6 M. SOBAJIMA AND Y. WAKASUGI\nRemark 1.4. (i)About the explicit values of s=s(n)andm=m(n,α,λ)in\nTheorem 1.1, a rough computation shows that we can take s= 5(n+1)andm=\n(λ+2n+1)2−α\n2+(6n2+14n+8)α. However, we omit the detailed computation,\nand do not discuss the optimality of them here.\n(ii)Ifu0,u1∈C∞\n0(Ω), then the assumptions on the initial data of Theorem 1.1 are\nautomatically fulfilled.\n1.3.A rough descripsion of strategy. By the previous studies [39, 32, 36], the\nsolution of (1.6) is known to be the first asymptotic profile of the solu tion of (1.1).\nTo investigate the asymptotic behavior of solution to (1.1), we follow the idea of\n[31]. First, the fact that V0is the first asymptotic profile implies that u−V0is a\nremainderterm. In [31], it is found that the remainderterm u−V0can be expressed\nas the time derivative of the solution of the damped wave equation wit h a certain\ninhomogeneous term. More precisely, let U1be the solution of\n\n∂2\ntU1−∆U1+a(x)∂tU1=−∂tV0, x ∈Ω,t>0,\nU1(x,t) = 0, x ∈∂Ω,t>0,\nU1(x,0) = 0, ∂tU1(x,0) =−a(x)−1u1(x), x∈Ω.\nNext, we have the decomposition u=V0+∂tU1(see Lemma 3.9). Then we further\nconsider the asymptotic profile of U1. By experience, it is natural to choose V1\nvia (1.8) with n= 1 (V1andU1has the same inhomogeneous term in respective\nequations). Then, inasimilarway U1canalsobealsodecomposedas U1=V1+∂tU2\nwith the second auxiliary function U2via\n\n∂2\ntU2−∆U2+a(x)∂tU2=−∂tV1, x ∈Ω,t>0,\nU2(x,t) = 0, x ∈∂Ω,t>0,\nU2(x,0) = 0, ∂tU2(x,0) = (−a(x))−2u1(x), x∈Ω.\nThe relation\nu=V0+∂tU1=V0+∂tV1+∂2\ntU2\ncanbe expectedtodeterminethesecondexpansion. Continuously , usingthe ( n+1)-\nth auxiliary function Un+1given by\n\n∂2\ntUn+1−∆Un+1+a(x)∂tUn+1=−∂tVn, x ∈Ω,t>0,\nUn+1(x,t) = 0, x∈∂Ω,t>0,\nUn+1(x,0) = 0, ∂tUn+1(x,0) = (−a(x))−n−1u1(x), x∈Ω,(1.10)\none can obtain the relation\nu=V0+∂tV1+∂2\ntV2+···+∂n\ntVn+∂n+1\ntUn+1. (1.11)\nMore precise discussion will be given in Section 3. Note that even if the initial\ndata (u0,u1) are compactly supported, V0,...,V nandUn+1do not have compact\nsupports in general. Thereforethe finite propagatoinpropertyd oes not workin this\nsituation. Applying a weighted energy method developed by the auth ors’ previous\npapers [30, 36], we prove that ∂n+1\ntUn+1decays faster than the other terms in\n(1.11), and this implies that the solution uis asymptotically expanded by the sum\nofV0,∂tV1,...,∂n\ntVn.\n1.4.Construction of the paper. This paper is constructed as follows. In the\nnext section, we prepare the weight functions used in the energy m ethod in subse-\nquent sections. In section 3, we state the well-posedness and reg ularity of solutions\nof the problem (1.1) and discuss the validity of the decomposition (1.1 1) (formally\nexplained in Subsection 1.3) in a suitable weighted Sobolev space. In Se ction 4, weWAVE EQUATION WITH SPACE-DEPENDENT DAMPING 7\ndiscuss the weighted energy estimates for the corresponding par abolic equations.\nIn Section 5, we prove the weighted energy estimates for the damp ed wave equation\n(1.1) with an inhomogeneous term. Finally, in Section 6, we complete th e proof of\nTheorem 1.1 by adapting the energy estimates prepared in Sections 4 and 5 to the\noriginal problem (1.1).\n1.5.Notations. We finish this section with some notations used throughout this\npaper. The letter Cindicates a generic positive constant, which may change from\nline to line. We also express constants by C(∗,...,∗), which means this constant\ndepends on the parametersin the parenthesis. The symbol f/lessorsimilargstands forf≤Cg\nholds with some constant C >0, andf∼gmeans both f/lessorsimilargandg/lessorsimilarfhold.\nWe denote /an}b∇acketle{tx/an}b∇acket∇i}ht=/radicalbig\n1+|x|2forx∈Rn. LetL2(Ω) be the usual Lebesgue space\nwith the norm\n/ba∇dblf/ba∇dblL2(Ω)=/parenleftbigg/integraldisplay\nΩ|f(x)|2dx/parenrightbigg1/2\n,\nandC∞\n0(Ω) stands for the space of infinitely differentiable functions with co mpact\nsupport in Ω. For a nonnegative integer kandm∈R, we introduce the weighted\nSobolev spaces by\nHk,m(Ω) ={f: Ω→R;/an}b∇acketle{tx/an}b∇acket∇i}htm∂α\nxf∈L2(Ω)for any α∈ZN\n≥0with|α| ≤k},\n/ba∇dblf/ba∇dblHk,m(Ω)=/summationdisplay\n|α|≤k/ba∇dbl/an}b∇acketle{tx/an}b∇acket∇i}htm∂α\nxf/ba∇dblL2(Ω),\nwhere we used the notion of multi-index and the derivatives are in the sense of\ndistribution. When m= 0, we denote Hk(Ω) =Hk,0(Ω) for short. Also, Hk,m\n0(Ω)\nis the completion of C∞\n0(Ω) with respect to the norm /ba∇dblf/ba∇dblHk,m(Ω).\n2.Preliminaries\n2.1.Weight functions. Throughout this section, we slightly generalizethe condi-\ntions ona(x) and assume that a(x) is a smooth positive function on RNsatisfying\nlim\n|x|→∞|x|αa(x) =a0 (2.1)\nwith some constants α∈(−∞,min{2,N}) anda0>0.\nWe prepare weight functions constructed in [36]. First, we introduc e a suitable\napproximate solution of the Poisson equation ∆ A(x) =a(x).\nLemma2.1 ([33, Lemma2.1],[36, Lemma3.2]) .Leta(x)be asmooth positive func-\ntion onRNsatisfying the condition (2.1)with some constants α∈(−∞,min{2,N})\nanda0>0. Then for every ε∈(0,1), there exist a function Aε∈C2(RN)and\npositive constants cεandCεsuch that\n(1−ε)a(x)≤∆Aε(x)≤(1+ε)a(x), (2.2)\ncε/an}b∇acketle{tx/an}b∇acket∇i}ht2−α≤Aε(x)≤Cε/an}b∇acketle{tx/an}b∇acket∇i}ht2−α,\n|∇Aε(x)|2\na(x)Aε(x)≤2−α\nN−α+ε (2.3)\nhold forx∈RN.8 M. SOBAJIMA AND Y. WAKASUGI\nRemark 2.2. The above type function Aε(x)was firstly introduced by Ikehata [9],\nTodorova and Yordanov [38], and Nishihara [22]. In particular, in [38], a solution\nof∆A(x) =a(x), that is, the equation obtained by taking ε= 0in(2.2), was ap-\nplied for weighted energy estimates for the damped wave equa tion(1.1)with radially\nsymmetric a(x). Lemma 2.1 is a refinement of the method of [38]to remove the\nassumption of radial symmetry on a(x).\nThe following definitions are connected to the supersolution of a(x)vt−∆v= 0\nconstructed in [36], which plays a crucial role to obtain several estim ates verifying\nasymptotic expansion.\nDefinition 2.3 (Kummer’s confluent hypergeometric functions) .Forb,c∈Rwith\n−c /∈N∪{0}, Kummer’s confluent hypergeometric function of first kind is defined\nby\nM(b,c;s) =∞/summationdisplay\nn=0(b)n\n(c)nsn\nn!, s∈[0,∞),\nwhere(d)nis the Pochhammer symbol defined by (d)0= 1and(d)n=/producttextn\nk=1(d+\nk−1)forn∈N; note that when b=c,M(b,b;s)coincides with es.\nDefinition 2.4. (i)Forε∈(0,1/2), we define\n/tildewideγε=/parenleftbigg2−α\nN−α+2ε/parenrightbigg−1\n, γε= (1−2ε)/tildewideγε. (2.4)\n(ii)Forβ≥0andε∈(0,1/2), define\nϕβ,ε(s) =e−sM(γε−β,γε;s), s≥0.\nRemark 2.5. (i)We slightly modify the definition of ˜γεandγεfrom those of [36]\nin order to gain a positive term in the right-hand side of Prop osition 2.8 (iv). This\nmodification enables us to unify the proof of energy estimate s for the case N= 1\nandN≥2(see Sections 4 and 5).\n(ii)We note that ϕβ,ε(s)is a unique (modulo constant multiple) solution of\nsϕ′′(s)+(γε+s)ϕ′(s)+βϕ(s) = 0 (2.5)\nwith bounded derivative near s= 0.\nLemma 2.6. The function ϕβ,εdefined in Definition 2.4 satisfies the following\nproperties.\n(i)If0≤β <γε, thenϕβ,ε(s)satisfies the estimates\nkβ,ε(1+s)−β≤ϕβ,ε(s)≤Kβ,ε(1+s)−β\nwith some constants kβ,ε,Kβ,ε>0.\n(ii)For everyβ≥0, the estimate\n|ϕβ,ε(s)| ≤Kβ,ε(1+s)−β\nholds with some constant Kβ,ε>0.\n(iii)For everyβ≥0,ϕβ,ε(s)andϕβ+1,ε(s)satisfy the recurrence relation\nβϕβ,ε(s)+sϕ′\nβ,ε(s) =βϕβ+1,ε(s).WAVE EQUATION WITH SPACE-DEPENDENT DAMPING 9\n(iv)If0≤β <γε, then we have\nϕ′\nβ,ε(s) =−β\nγεe−sM(γε−β,γε+1;s)≤0,\nϕ′′\nβ,ε(s) =β(β+1)\nγε(γε+1)e−sM(γε−β,γε+2;s)≥0.\n(v)If0≤β <γε, thenϕ′\nβ,εsatisfies\n−ϕ′\nβ,ε(s)≥kβ,ε(1+s)−β−1\nholds with some constant kβ,ε>0.\nProof.The proof of the assertions (i)–(iv) are completely the same as tha t of [36,\nLemma 3.5], and we omit the detail. The property (v) follows from the e xpression\nin (vi) and the fact M(γε−β,γε+ 1;s)∼Γ(γε+1)\nΓ(γε−β)s−β−1esass→ ∞(see, for\nexample, [36, Lemma 2.2 (ii)] or [1, p.192, (6.1.8)]). /square\nHere, we give a family of supersolutions of a(x)vt−∆v= 0, which we use later.\nDefinition 2.7. Forβ≥0and(x,t)∈RN×[0,∞), we define\nΦβ,ε(x,t;t0) = (t0+t)−βϕβ,ε(z), z=/tildewideγεAε(x)\nt0+t,\nwhereε∈(0,1/2),/tildewideγεis the constant given in (2.4),t0≥1,ϕβ,εis the function\ndefined by Definition 2.4, and Aε(x)is the function constructed in Lemma 2.1.\nFort0≥1 and (x,t)∈RN×[0,∞), we also define\nΨ(x,t;t0) :=t0+t+Aε(x). (2.6)\nProposition 2.8. The function Φβ,ε(x,t;t0)defined in Definition 2.7 satisfies the\nfollowing properties:\n(i)For everyβ≥0, we have\n∂tΦβ,ε(x,t;t0) =−βΦβ+1,ε(x,t;t0)\nfor any(x,t)∈RN×[0,∞).\n(ii)Ifβ≥0, then there exists a constant Cα,β,ε>0such that\n|Φβ,ε(x,t;t0)| ≤Cα,β,εΨ(x,t;t0)−β\nfor any(x,t)∈RN×[0,∞).\n(iii)Ifβ∈[0,γε), then there exists a constant cα,β,ε>0such that\nΦβ,ε(x,t;t0)≥cα,β,εΨ(x,t;t0)−β\nfor any(x,t)∈RN×[0,∞).\n(iv)For everyβ≥0, there exists a constant cα,β,ε>0such that\na(x)∂tΦβ,ε(x,t;t0)−∆Φβ,ε(x,t;t0)≥cα,β,εa(x)Ψ(x,t;t0)−β−1.\nfor any(x,t)∈RN×[0,∞).10 M. SOBAJIMA AND Y. WAKASUGI\nProof.The properties (i)–(iii) are the same as [36, Lemma 3.8] and [36, Lemma\n5.1]. Thus, we omit the detail. For (iv), we put z= ˜γεAε(x)/(t0+t) and compute\n(t0+t)β+1(a(x)∂tΦβ,ε(x,t;t0)−∆Φβ,ε(x,t;t0))\n=−a(x)/parenleftbigg\nβϕβ,ε(z)+zϕ′\nβ,ε(z)+ ˜γε∆Aε(x)\na(x)ϕ′\nβ,ε(z)+ ˜γε|∇Aε(x)|2\na(x)Aε(x)zϕ′′\nβ,ε(z)/parenrightbigg\n.\nUsing the equation (2.5) with (2.4), we rewrite the right-hand side as\n˜γεa(x)/parenleftbigg\n1−2ε−∆Aε(x)\na(x)/parenrightbigg\nϕ′\nβ,ε(z)+a(x)/parenleftbigg\n1−˜γε|∇Aε(x)|2\na(x)Aε(x)/parenrightbigg\nzϕ′′\nβ,ε(z).\nBy (2.2) and (2.3) in Lemma 2.1, we have\n1−2ε−∆Aε(x)\na(x)≤ −ε,\n1−˜γε|∇Aε(x)|2\na(x)Aε(x)≥ε/parenleftbigg2−α\nN−α+2ε/parenrightbigg−1\n>0.\nCombining them to the properties (iv) and (v) in Lemma 2.6, we conclud e\na(x)∂tΦβ,ε(x,t;t0)−∆Φβ,ε(x,t;t0)≥ −ε˜γεa(x)(t0+t)−β−1ϕ′\nβ,ε/parenleftbigg˜γεAε(x)\nt0+t/parenrightbigg\n≥Ca(x)(t0+t)−β−1/parenleftbigg\n1+˜γεAε(x)\nt0+t/parenrightbigg−β−1\n≥Ca(x)(t0+t+Aε(x))−β−1\n=Ca(x)Ψ(x,t;t0)−β−1,\nwhich completes the proof. /square\nFinally, we prepare a useful lemma for our weighted energy method.\nLemma 2.9 ([30, Lemma 2.5]) .LetΦ∈C2(Ω)be a positive function and let\nδ∈(0,1/2). Then, for any u∈H2(Ω)∩H1\n0(Ω), we have\n/integraldisplay\nΩu∆uΦ−1+2δdx≤ −δ\n1−δ/integraldisplay\nΩ|∇u|2Φ−1+2δdx+1−2δ\n2/integraldisplay\nΩu2(∆Φ)Φ−2+2δdx,\nprovided that the right-hand side is finite.\n3.Justification of the decomposition\nIn this section, we justify the decomposition\nu=n/summationdisplay\nj=0∂j\ntVj+∂n+1\ntUn+1\nwhich is explained in Subsection 1.3. Here we need to clarify existence, uniqueness\nand also an expected regularity of respective components V0, ...VnandUn+1.\nTherefore we discuss it in the following way: we first prepare the well- posedness of\nthe initial-boundary value problem of the damped wave equation. Nex t, we show a\nkeydecompositionlemma which statesthat a solutionofthe damped w aveequation\ncan be decomposed into a solution of the corresponding parabolic eq uation and the\nderivative of a solution of the damped wave equation with another inh omogeneousWAVE EQUATION WITH SPACE-DEPENDENT DAMPING 11\nterm. Finally, using the decomposition lemma repeatedly, we explain ho w the\nhigher order asymptotic profiles are determined.\n3.1.Well-posedness and regularity of solutions for the damped w ave\nequation. We consider the initial-boundary value problem of the damped wave\nequation with a general inhomogeneous term\n\n\n∂2\ntw−∆w+a(x)∂tw=F, x ∈Ω,t>0,\nw(x,t) = 0, x ∈∂Ω,t>0,\nw(x,0) =w0(x), ∂tw(x,0) =w1(x), x∈Ω.(3.1)\nWe first prepare the well-posedness and the regularity of solutions for (3.1).\nWe recall the following well-posedness result by Ikawa [7].\nTheorem 3.1 ([7, Theorem 1]) .For any (w0,w1)∈(H2(Ω)∩H1\n0(Ω))×H1\n0(Ω)\nandF∈C1([0,∞);L2(Ω)), there exists a unique solution\nw∈C([0,∞);H2(Ω))∩C1([0,∞);H1\n0(Ω))∩C2([0,∞);L2(Ω))\nof(3.1).\nBy applying the above theorem to /an}b∇acketle{tx/an}b∇acket∇i}htmw, (/an}b∇acketle{tx/an}b∇acket∇i}htmw0,/an}b∇acketle{tx/an}b∇acket∇i}htmw1), and/an}b∇acketle{tx/an}b∇acket∇i}htmF, we\nhave the well-posedness of the problem (3.1) in weighted Sobolev spa ces.\nTheorem 3.2. For any (w0,w1)∈(H2,m(Ω)∩H1,m\n0(Ω))×H1,m\n0(Ω)andF∈\nC1([0,∞);H0,m(Ω)), there exists a unique solution\nw∈C([0,∞);H2,m(Ω))∩C1([0,∞);H1,m\n0(Ω))∩C2([0,∞);H0,m(Ω))\nof(3.1).\nNext, we discuss the regularity of the solution. We first recall the f ollowing\nregularity theorem by Ikawa [7]:\nTheorem 3.3 ([7, Theorem 2]) .Letk≥1be an integer and let m≥0,w0∈\nHk+2(Ω),w1∈Hk+1(Ω), andF∈/intersectiontextk\nj=0Cj+1([0,∞);Hk−j(Ω)). We successively\ndefine\nwp= ∆wp−2−a(x)wp−1+∂p−2\ntF(x,0)\nforp= 2,...,k+1, and assume the k-th order compatibility condition\n(wp,wp+1)∈(H2(Ω)∩H1\n0(Ω))×H1\n0(Ω)\nforp= 0,1,...,k. Then, the solution wto(3.1)obtained by Theorem 3.1 belongs\nto\nC([0,∞);Hk+2(Ω))∩\nk+1/intersectiondisplay\nj=1Ck+2−j([0,∞);Hj\n0(Ω))\n∩Ck+2([0,∞);L2(Ω)).\nFrom the above theorem and the same argument as Theorem 3.2, we have the\nfollowing regularity theorem in weighted Sobolev spaces.\nTheorem 3.4. Letk≥1be an integer and let m≥0,w0∈Hk+2,m(Ω),w1∈\nHk+1,m(Ω), andF∈/intersectiontextk\nj=0Cj+1([0,∞);Hk−j,m(Ω)). We successively define\nwp= ∆wp−2−a(x)wp−1+∂p−2\ntF(x,0)12 M. SOBAJIMA AND Y. WAKASUGI\nforp= 2,...,k+1, and assume the k-th order compatibility condition\n(wp,wp+1)∈(H2,m(Ω)∩H1,m\n0(Ω))×H1,m\n0(Ω)\nforp= 0,1,...,k. Then, the solution wto(3.1)obtained by Theorem 3.2 belongs\nto\nC([0,∞);Hk+2,m(Ω))∩\nk+1/intersectiondisplay\nj=1Ck+2−j([0,∞);Hj,m\n0(Ω))\n∩Ck+2([0,∞);H0,m(Ω)).\n3.2.Regularity of solutions for the corresponding heat equatio n.Follow-\ning our previous study [32, section 2], we prepare the well-posednes s and regularity\nof solutions for the initial-boundary problem of the corresponding h eat equation\nwith a general inhomogeneous term\n\n\na(x)∂tv−∆v=G, x∈Ω,t>0,\nv(x,t) = 0, x ∈∂Ω,t>0,\nv(x,0) =v0(x), x∈Ω.(3.2)\nLetdµ=a(x)dxand we define\nL2\ndµ(Ω) =/braceleftBigg\nf∈L2\nloc(Ω);/ba∇dblf/ba∇dblL2\ndµ=/parenleftbigg/integraldisplay\nΩ|f(x)|2dµ/parenrightbigg1/2\n<∞/bracerightBigg\n,\n(f,g)L2\ndµ:=/integraldisplay\nΩf(x)g(x)dµ.\nThe operator −a(x)−1∆ is formally symmetric in L2\ndµ(Ω), and its bilinear closed\nform is defined by\na(u,v) =/integraldisplay\nΩ∇u(x)·∇v(x)dx,\nD(a) =/braceleftbigg\nu∈L2\ndµ(Ω)∩˙H1(Ω);/integraldisplay\nΩ∂u\n∂xjϕdx=−/integraldisplay\nΩu∂ϕ\n∂xjdxfor allϕ∈C∞\n0(RN)/bracerightbigg\n.\nFrom [32], we have the Friedrichs extension −Lof the operator −a(x)−1∆ in\nL2\ndµ(Ω).\nLemma 3.5 ([32, Lemma 2.2]) .The operator −LinL2\ndµ(Ω)defined by\nD(L) =/braceleftBig\nu∈D(a);∃f∈L2\ndµ(Ω)s.t.a(u,v) = (f,v)L2\ndµfor anyv∈D(a)/bracerightBig\n,\n−Lu=f\nis nonnegative and selfadjoint in L2\ndµ(Ω). Therefore, Lgenerates an analytic semi-\ngroupT(t)onL2\ndµ(Ω)satisfying\n/ba∇dblT(t)f/ba∇dblL2\ndµ≤ /ba∇dblf/ba∇dblL2\ndµ,/ba∇dblLT(t)f/ba∇dblL2\ndµ≤1\nt/ba∇dblf/ba∇dblL2\ndµ\nfor anyf∈L2\ndµ(Ω).\nWe alsorecall the followingproperty proved in [32], which will be used in S ection\n6:WAVE EQUATION WITH SPACE-DEPENDENT DAMPING 13\nLemma 3.6 ([32, Lemma 2.3]) .We have\n/braceleftBig\nu∈H2(Ω)∩H1\n0(Ω);a(x)−1/2∆u∈L2(Ω)/bracerightBig\n⊂D(L).\nFor the inhomogeneous problem (3.2), applying [2, Lemma 4.1.1, Propo sition\n4.1.6], we have the following well-posedness result.\nTheorem 3.7. Assume that v0∈D(L)anda(x)−1G∈C1([0,∞);L2\ndµ(Ω)). Then,\nthe function vdefined by\nv(t) =T(t)v0+/integraldisplayt\n0T(t−s)(a(x)−1G(s))ds\nis the unique solution to the problem (3.2)satisfying\nv∈C([0,∞);D(L))∩C1([0,∞);L2\ndµ(Ω)).\nNext, we discuss the higher order regularity in time for the solution o f (3.2). We\nnote that, by a formal straightforward computation, the initial v alues of∂j\ntvfor\nj≥1 are given by\n∂j\ntv(x,0) = [a(x)−1∆]jv0(x)+j−1/summationdisplay\nl=0[a(x)−1∆]la(x)−1∂j−1−l\ntG(x,0).(3.3)\nTheorem3.8. Letk≥1be an integer, v0∈D(L), anda(x)−1G∈Ck+1([0,∞);L2\ndµ(Ω)).\nAssume that ∂j\ntv(x,0)defined by the right-hand side of (3.3)satisfies∂j\ntv(x,0)∈\nD(L)forj= 1,...,k. Then, the solution vto(3.2)obtained by Theorem 3.7\nbelongs to\nCk([0,∞);D(L))∩Ck+1([0,∞);L2\ndµ(Ω)). (3.4)\nProof.Whenk= 1, letψ=ψ(x,t) be the solution of (3.2) with the inhomogeneous\nterm∂tGand the initial data ψ(x,0) =a(x)−1∆v0(x) +a(x)−1G(x,0). Then, by\nTheorem 3.7, ψis given by\nψ(t) =T(t)[a(x)−1∆v0(x)+a(x)−1G(x,0)]+/integraldisplayt\n0T(t−s)(a(x)−1∂sG(s))ds\n=∂tT(t)v0+T(t)[a(x)−1G(x,0)]\n+/bracketleftbig\nT(t−s)(a(x)−1G(s))/bracketrightbigt\n0+/integraldisplayt\n0∂tT(t−s)(a(x)−1G(s))ds\n=∂tT(t)v0+a(x)−1G(t)\n+∂t/integraldisplayt\n0T(t−s)(a(x)−1G(s))ds−a(x)−1G(t)\n=∂tv(t).\nSinceψ∈C([0,∞);D(L))∩C1([0,∞);L2\ndµ(Ω)), we have\nv∈C1([0,∞);D(L))∩C2([0,∞);L2\ndµ(Ω)),\nthat is, the assertion when k= 1 is proved. The general case k≥1 can be proved\nin the same way with induction, and we omit the detail. /square14 M. SOBAJIMA AND Y. WAKASUGI\n3.3.A decomposition lemma. In the following two subsections, we give the idea\nof the asymptotic expansion of the solution of the damped wave equ ation (1.1). To\nsimplify the discussion, we only give formal computation here. The ju stification\nand the complete proof of the asymptotic expansion will be given in Se ction 6.\nRelated to the initial-boundary value problem of the damped wave equ ation\n(3.1), we consider the parabolic problem with the same inhomogeneou s termFand\nthe initial data w0(x)+a(x)−1w1(x):\n\n\na(x)∂tV−∆V=F, x ∈Ω,t>0,\nV(x,t) = 0, x ∈∂Ω,t>0,\nV(x,0) =w0(x)+a(x)−1w1(x), x∈Ω.(3.5)\nFor the solution Vof the above problem, we further consider the following initial-\nboundary value problem of the damped wave equation with the inhomo geneous\nterm−∂tVand the initial data ( U,∂tU)(x,0) = (0,−a(x)−1w1(x)):\n\n\n∂2\ntU−∆U+a(x)∂tU=−∂tV, x ∈Ω,t>0,\nU(x,t) = 0, x ∈∂Ω,t>0,\nU(x,0) = 0, ∂tU(x,0) =−a(x)−1w1(x), x∈Ω.(3.6)\nThen, we have the following decomposition of the solution wto (3.1).\nLemma 3.9. Letwbe the solution of the damped wave equation (3.1)with the\ninhomogeneous term Fand the initial data (w0,w1). LetVbe the solution of the\nparabolic problem (3.5), and letUbe the solution of the problem (3.6). Then, we\nhave\nw=V+∂tU.\nProof.Let ˜w=V+∂tU. Then, we have\n˜w(x,0) =V(x,0)+∂tU(x,0) =w0(x)+a(x)−1w1(x)−a(x)−1w1(x) =w0(x).\nAlso, by (3.6), we obtain\n∂t˜w=∂tV+∂2\ntU= ∆U−a(x)∂tU. (3.7)\nThis implies ∂t˜w(x,0) = ∆U(x,0)−a(x)∂tU(x,0) =w1(x). Finally, differentiating\n(3.7) again, and using the relation ∂2\ntU=∂t˜w−∂tV, we deduce\n∂2\nt˜w= ∆∂tU−a(x)∂2\ntU\n= ∆(˜w−V)−a(x)(∂t˜w−∂tV)\n= ∆˜w−a(x)∂t˜w+F.\nConsequently, ˜ wis the solution of (3.1), and hence, the uniqueness shows ˜ w=w.\nThis completes the proof. /square\n3.4.Derivation of the asymptotic expansion. Letube the solution of (1.1).\nTo expand uin terms of solutions of the corresponding parabolic problem, we con -\nsiderfunctions V0,V1,...,V nand the remainderterms U1,U2,...,U n+1successively\ndefined in the following way: first, we define V0by\n\n\na(x)∂tV0−∆V0= 0, x ∈Ω,t>0,\nV0(x,t) = 0, x ∈∂Ω,t>0,\nV0(x,0) =u0(x)+a(x)−1u1(x), x∈Ω.WAVE EQUATION WITH SPACE-DEPENDENT DAMPING 15\nandU1by\n\n\n∂2\ntU1−∆U1+a(x)∂tU1=−∂tV0, x ∈Ω,t>0,\nU1(x,t) = 0, x ∈∂Ω,t>0,\nU1(x,0) = 0, ∂tU1(x,0) =−a(x)−1u1(x), x∈Ω.\nThen, by Lemma 3.9 we have the first decomposition u=V0+∂tU1. According to\nthe experiences, we expect that this is the first-orderasymptot ic expansionof uand\ntherefore∂tU1can be regarded as a perturbation. Next, to obtain the second-o rder\nexpansion, we further consider the decomposition of U1in terms of corresponding\nparabolic problem. Namely, we define V1by\n\n\na(x)∂tV1−∆V1=−∂tV0, x∈Ω,t>0,\nV1(x,t) = 0, x ∈∂Ω,t>0,\nV1(x,0) =−a(x)−2u1(x), x∈Ω.\nandU2by\n\n\n∂2\ntU2−∆U2+a(x)∂tU2=−∂tV1, x ∈Ω,t>0,\nU2(x,t) = 0, x ∈∂Ω,t>0,\nU2(x,0) = 0, ∂tU2(x,0) =a(x)−2u1(x), x∈Ω.\nThen, by Lemma 3.9 again, we have the decomposition U1=V1+∂tU2, which\nimplies\nu=V0+∂tV1+∂2\ntU2.\nWe expect that this gives the second-order asymptotic expansion ofu. Repeating\nthis procedure for j= 2,3,...,nwe successively define Vjby\n\n\na(x)∂tVj−∆Vj=−∂tVj−1, x∈Ω,t>0,\nVj(x,t) = 0, x ∈∂Ω,t>0,\nVj(x,0) =−(−a(x))−j−1u1(x), x∈Ω\nandUn+1by\n\n\n∂2\ntUn+1−∆Un+1+a(x)∂tUn+1=−∂tVn, x ∈Ω,t>0,\nUn+1(x,t) = 0, x∈∂Ω,t>0,\nUn+1(x,0) = 0, ∂tUn+1(x,0) = (−a(x))−n−1u1(x), x∈Ω.(3.8)\nThen, we can have the expected higher order decomposition\nu=V0+∂tV1+∂2\ntV2+···+∂n\ntVn+∂n+1\ntUn+1. (3.9)\nBy Theorems 3.2 and 3.4, the existence, uniqueness, and regularity of the solution\nUn+1to (3.8) can be obtained from the assumptions on the initial data of T heorem\n1.1. The detail will be discussed in Section 6.\nIn the following sections, we give energy estimates for V0,V1,...,V nandUn+1\nto justify that (3.9) actually gives the n-th order asymptotic expansion of u.\n4.Energy estimates for the heat equation\nWe apply the weighted energy method to obtain the decay estimate o f the par-\nabolic problem\n\n\na(x)∂tv−∆v=G, x∈Ω,t>0,\nv(x,t) = 0, x ∈∂Ω,t>0,\nv(x,0) =v0(x), x∈Ω.(4.1)16 M. SOBAJIMA AND Y. WAKASUGI\nThe goal of this section is the following weighted energy estimates fo r higher\norder derivatives of solutions to (4.1).\nTheorem 4.1. Letk≥0be an integer, δ∈(0,1/2),ε∈(0,1/2),λ∈[0,(1−2δ)γε)\n(see(2.4)for the definition of γε),t0≥1andβ=λ/(1−2δ). Letv0∈D(L),\na(x)−1G∈Ck+1([0,∞);L2\ndµ(Ω)), and letvbe the corresponding solution of (4.1).\nMoreover, we assume that ∂j\ntv(x,0)given by the right-hand side of (3.3)satisfies\n∂j\ntv(x,0)∈D(L)∩H0,(λ+2j)2−α\n2−α\n2(Ω),\n∇∂j\ntv(x,0)∈H0,(λ+1+2j)2−α\n2(Ω),\nand/integraldisplay\nΩa(x)−1|∂j\ntG(x,t)|2Ψ(x,t;t0)λ+1+2jdx∈L1(0,∞)\nforj= 0,1,...,k. Then, we have/integraldisplay\nΩa(x)|∂j\ntv(x,t)|2Ψ(x,t;t0)λ+2jdx∈L∞(0,∞), (4.2)\n/integraldisplay\nΩ|∇∂j\ntv(x,t)|2Ψ(x,t;t0)λ+2jdx∈L1(0,∞), (4.3)\n/integraldisplay\nΩ|∇∂j\ntv(x,t)|2Ψ(x,t;t0)λ+1+2jdx∈L∞(0,∞), (4.4)\n/integraldisplay\nΩa(x)|∂j+1\ntv(x,t)|2Ψ(x,t;t0)λ+1+2jdx∈L1(0,∞) (4.5)\nforj= 0,1,...,k.\nWe note that Theorem 3.8 ensures the regularity property (3.4) fo r the solution\nv. Thus, it suffices to show the estimates (4.2)–(4.5).\nThe proof of Theorem 4.1 is based on an induction argument. The follo wing\nlemma is the first step.\nLemma 4.2. Under the assumptions on Theorem 4.1 with k= 0, we have (4.2)–\n(4.5)fork= 0.\nProof.ByLemma2.9, Proposition2.8(iii), andtheSchwarzinequality,wecalcu late\nd\ndt/integraldisplay\nΩa(x)v2\nΦ1−2δ\nβ,εdx= 2/integraldisplay\nΩv∆v\nΦ1−2δ\nβ,εdx−(1−2δ)/integraldisplay\nΩa(x)v2∂tΦβ,ε\nΦ2−2δ\nβ,εdx\n+2/integraldisplay\nΩvG\nΦ1−2δ\nβ,εdx\n≤ −2δ\n1−δ/integraldisplay\nΩ|∇v|2\nΦ1−2δ\nβ,εdx−(1−2δ)/integraldisplay\nΩ(a(x)∂tΦβ,ε−∆Φβ,ε)v2\nΦ2−2δ\nβ,εdx\n+C/parenleftbigg/integraldisplay\nΩa(x)v2Ψλ−1dx/parenrightbigg1/2/parenleftbigg/integraldisplay\nΩa(x)−1G2Ψλ+1dx/parenrightbigg1/2\n.\nFrom the Young inequality and Proposition 2.8 (iv), we obtain\nd\ndt/integraldisplay\nΩa(x)v2\nΦ1−2δ\nβ,εdx=−2δ\n1−δ/integraldisplay\nΩ|∇v|2\nΦ1−2δ\nβ,εdx+C/integraldisplay\nΩa(x)−1G2Ψλ+1dx.WAVE EQUATION WITH SPACE-DEPENDENT DAMPING 17\nIntegrating it over [0 ,t] and using Proposition 2.8 (iii), we deduce\n/integraldisplay\nΩa(x)v(x,t)2Ψ(x,t;t0)λdx∈L∞(0,∞),\n/integraldisplay\nΩ|∇v(x,t)|2Ψ(x,t;t0)λdx∈L1(0,∞), (4.6)\nThus, we have (4.2) and (4.3) in the case j= 0. We next compute\nd\ndt/integraldisplay\nΩ|∇v|2Ψλ+1dx= (λ+1)/integraldisplay\nΩ|∇v|2Ψλdx+2/integraldisplay\nΩ(∇∂tv·∇v)Ψλ+1dx.(4.7)\nBy (4.6), the first term of the right-hand side belongs to L1(0,∞), and by integra-\ntion by parts, the second term is estimated as\n2/integraldisplay\nΩ(∇∂tv·∇v)Ψλ+1dx\n=−2/integraldisplay\nΩ(a(x)∂tv−G)∂tvΨλ+1dx−(λ+1)/integraldisplay\nΩ∂tv(∇v·∇Aε(x))Ψλdx\n≤ −2/integraldisplay\nΩa(x)|∂tv|2Ψλ+1dx\n+1\n2/integraldisplay\nΩa(x)|∂tv|2Ψλ+1dx+2/integraldisplay\nΩa(x)−1G2Ψλ+1dx\n+1\n2/integraldisplay\nΩa(x)|∂tv|2Aε(x)Ψλdx+2(λ+1)2/integraldisplay\nΩ|∇Aε(x)|2\na(x)Aε(x)|∇v|2Ψλdx\n≤ −/integraldisplay\nΩa(x)|∂tv|2Ψλ+1dx\n+C/integraldisplay\nΩ|∇v|2Ψλdx+2/integraldisplay\nΩa(x)−1G2Ψλ+1dx. (4.8)\nHere, we have used the property (2.3) in Lemma 2.1 and the relation Aε(x)≤\nΨ(x,t;t0), which follows from the definition Ψ (see (2.6)). By (4.6) and the as-\nsumption on G, the last two terms of the right-hand side of above are in L1(0,∞).\nThus, integrating (4.7) over [0 ,t], we conclude\n/integraldisplay\nΩ|∇v(x,t)|2Ψ(x,t;t0)λ+1dx∈L∞(0,∞),\n/integraldisplay\nΩa(x)|∂tv(x,t)|2Ψ(x,t;t0)λ+1dx∈L1(0,∞),\nthat is, (4.4) and (4.5) in the case j= 0, and the proof is now complete. /square\nRemark 4.3. It should be noted that the integration by parts in the above p roof can\nbe justified completely by the approximation argument in the same as [36, Section\n4].\nNext, we prove the following lemma, which is the main part of the induct ion\nargument of the proof of Theorem 4.1.\nLemma 4.4. Assumea(x)satisfies (1.2). Letσ≥0,t0≥1,v0∈D(L)and\na(x)−1G∈C1([0,∞);L2\ndµ(Ω)). Letvbe the corresponding solution of (4.1). We18 M. SOBAJIMA AND Y. WAKASUGI\nfurther assume\nv0∈H0,2−α\n2σ−α\n2(Ω),∇v0∈H0,2−α\n2(σ+1)(Ω),\n/integraldisplay\nΩa(x)−1|G(x,t)|2Ψ(x,t;t0)σ+1dx∈L1(0,∞), (4.9)\nand also/integraldisplay\nΩa(x)|v(x,t)|2Ψ(x,t;t0)σ−1dx∈L1(0,∞). (4.10)\nThen, we have\n/integraldisplay\nΩa(x)|v(x,t)|2Ψ(x,t;t0)σdx∈L∞(0,∞), (4.11)\n/integraldisplay\nΩ|∇v(x,t)|2Ψ(x,t;t0)σdx∈L1(0,∞), (4.12)\n/integraldisplay\nΩ|∇v(x,t)|2Ψ(x,t;t0)σ+1dx∈L∞(0,∞), (4.13)\n/integraldisplay\nΩa(x)|∂tv(x,t)|2Ψ(x,t;t0)σ+1dx∈L1(0,∞). (4.14)\nRemark 4.5. In the proof of Theorem 4.1, we will choose σ=λ+2jforj∈N.\nProof of Lemma 4.4. Suppose (4.9) and (4.10). Similarly as the proof of Lemma\n4.2, we compute\nd\ndt/integraldisplay\nΩa(x)|v(x,t)|2Ψ(x,t;t0)σdx\n= 2/integraldisplay\nΩv(∆v)Ψσdx+σ/integraldisplay\nΩa(x)|v|2Ψσ−1dx\n+2/integraldisplay\nΩvGΨσdx (4.15)\nThe second term of the right-hand side is in L1(0,∞) due to the assumption (4.10).\nNoting the relation 2 v∆v=−2|∇v|2+∆(|v|2) and using the integration by parts,\nwe calculate the first term of the right-hand side as\n2/integraldisplay\nΩv(∆v)Ψσdx=−2/integraldisplay\nΩ|∇v|2Ψσdx+/integraldisplay\nΩ|v|2∆(Ψσ)dx.\nThe last term of the above is further estimated by\n/integraldisplay\nΩ|v|2|∆(Ψσ)|dx=σ/integraldisplay\nΩ|v|2/vextendsingle/vextendsingle∆AεΨ+(σ−1)|∇Aε|2/vextendsingle/vextendsingleΨσ−2dx\n≤C/integraldisplay\nΩa(x)|v|2Ψσ−1dx,\nwhere we have used (2.2), (2.3), and Aε≤Ψ. Therefore, this term also belongs to\nL1(0,∞) by the assumption (4.10). Finally, we apply the Schwarz inequality to the\nlast term of (4.15) and obtain\n2/integraldisplay\nΩvGΨσdx≤/integraldisplay\nΩa(x)|v|2Ψσ−1dx+/integraldisplay\nΩa(x)−1|G|2Ψσ+1dxWAVE EQUATION WITH SPACE-DEPENDENT DAMPING 19\nand these are in L1(0,∞) due to the assumptions (4.9) and (4.10). Consequently,\nintegrating (4.15) over [0 ,t], we have\n/integraldisplay\nΩa(x)|v(x,t)|2Ψ(x,t;t0)σdx∈L∞(0,∞),\n/integraldisplay\nΩ|∇v(x,t)|2Ψ(x,t;t0)σdx∈L1(0,∞).\nThus, we have (4.11) and (4.12).\nNext, we compute\nd\ndt/integraldisplay\nΩ|∇v(x,t)|2Ψ(x,t;t0)σ+1dx= (σ+1)/integraldisplay\nΩ|∇v|2Ψσdx\n+2/integraldisplay\nΩ(∇∂tv·∇v)Ψσ+1dx.(4.16)\nThe first term of the right-hand side belongs to L1(0,∞) by (4.12). The second\nterm can be estimated in completely the same way as (4.8), and we hav e\n2/integraldisplay\nΩ(∇∂tv·∇v)Ψσ+1dx≤ −/integraldisplay\nΩa(x)|∂tv|2Ψσ+1dx\n+C/integraldisplay\nΩ|∇v|2Ψσdx\n+C/integraldisplay\nΩa(x)−1|G|2Ψσ+1dx.\nBy using (4.9), and (4.12), the last two terms of above belong to L1(0,∞). Finally,\nintegrating (4.16) over [0 ,t], we conclude\n/integraldisplay\nΩ|∇v(x,t)|2Ψ(x,t;t0)σ+1dx∈L∞(0,∞),\n/integraldisplay\nΩa(x)|∂tv(x,t)|2Ψ(x,t;t0)σ+1dx∈L1(0,∞).\nThis completes the proof of (4.14) and (4.13). /square\nProof of Theorem 4.1. We note that the case k= 0 has been already proved by\nLemma 4.2. Let k≥1 be an integer. Then, by Lemma 4.2, we have (4.2)–(4.5) in\nthe casej= 0.\nNext, forj= 1, we apply Lemma 4.4 with σ=λ+2jand with the replacement\nofvandGby∂tvand∂tG, respectively. We remark that the condition (4.10) with\nσ=λ+2jis fulfilled by virtue of (4.5) with j= 0. Then, we obtain (4.11)–(4.14)\nforσ=λ+2jwith the replacement of vby∂tv, namely, we reach the conclusions\n(4.2)–(4.5) for j= 1.\nThe properties (4.2)–(4.5) for j= 1 allow us to apply again Lemma 4.4 with\nσ=λ+2j,j= 2andwith the replacementof vandGby∂2\ntvand∂2\ntG, respectively.\nThen, we can see that (4.2)–(4.5) for j= 2 hold. Repeating this argument until\nj=k, we complete the proof of Theorem 4.1. /square\n5.Energy estimates for the damped wave equation\n5.1.First order energy estimates. In this section, we discuss the energy esti-\nmate for the general damped wave equation (3.1).20 M. SOBAJIMA AND Y. WAKASUGI\nThe results of this section will be used in the next section by putting w=Un+1,\nF=−∂tVn,w0= 0, andw1= (−a(x))−n−1u1(x) (see (3.8)) to derive the energy\nestimate of ∂n+1\ntUn+1.\nWe start with the definition of the weighted energy of w.\nDefinition 5.1. Forδ∈(0,1/2),ε∈(0,1/2),λ∈[0,(1−2δ)γε)(see(2.4)for the\ndefinition of γε),β=λ/(1−2δ),t0≥1, andν >0, we define\nE1[w](t;t0,λ) :=/integraldisplay\nΩ/parenleftbig\n|∇w(x,t)|2+|∂tw(x,t)|2/parenrightbig\nΨ(x,t;t0)λ+1dx,\nE0[w](t;t0,λ) :=/integraldisplay\nΩ/parenleftbig\n2w(x,t)∂tw(x,t)+a(x)|w(x,t)|2/parenrightbig\nΦβ,ε(x,t;t0)−1+2δdx,\nE[w](t;t0,λ,ν) :=νE1[w](t;t0,λ)+E0[w](t;t0,λ)\nfort≥0.\nWe note that, for any ν >0, there exists t1>0 such that\nE[w](t;t0,λ,ν)∼E1[w](t;t0,λ)+/integraldisplay\nΩa(x)|w(x,t)|2Ψ(x,t;t0)λdx\nholds for any t0≥t1. Indeed, the Schwarz inequality implies\n|2w∂tw| ≤a(x)\n2|w|2+2a(x)−1|∂tw|2,\nand Proposition 2.8 (iii) and (iv) lead to\n2a(x)−1|∂tw|2Φ−1+2δ\nβ,ε≤CΨα\n2−α|∂tw|2Ψλ≤Ct−1+α\n2−α\n0|∂tw|2Ψλ+1≤ν\n2|∂tw|2Ψλ+1\nfor sufficiently large t0.\nThe main theorem of this subsection is the following:\nTheorem 5.2. Assume that a(x)satisfies(1.2). Letδ∈(0,1/2),ε∈(0,1/2),λ∈\n[0,(1−2δ)γε), andβ=λ/(1−2δ). Then, there exist constants ν=ν(N,α,δ,ε,λ )>\n0andt∗=t∗(N,α,δ,ε,λ,ν )≥1such that for any t0≥t∗, the following holds: Let\nm= (λ+1)2−α\n2and assume F∈C1([0,∞);H0,m(Ω))satisfies\n/integraldisplay\nΩa(x)−1|F(x,t)|2Ψ(x,t;t0)λ+1dx∈L1(0,∞).\nLetwbe the solution of (3.1)with the initial data (w0,w1)∈(H2,m(Ω)∩H1,m\n0(Ω))×\nH1,m\n0(Ω)given in Theorem 3.2. Then, we have\nE[w](t;t0,λ,ν)∈L∞(0,∞),\n/integraldisplay\nΩ|∇w(x,t)|2Ψ(x,t;t0)λdx∈L1(0,∞),\n/integraldisplay\nΩa(x)|∂tw(x,t)|2Ψ(x,t;t0)λ+1dx∈L1(0,∞).\nThe proof of Theorem 5.2 is a bit lengthy, but the outline is as follows. W e\nshall derive good terms −/integraldisplay\nΩ|∇w|2Ψλdxand−/integraldisplay\nΩa(x)|∂tw|2Ψλ+1dxfrom the\ncomputations ofd\ndtE0[w](t;t0,λ) andd\ndtE1[w](t;t0,λ), respectively (see the right-\nhand sides of Lemmas 5.3 and 5.4). Then, we sum up them with sufficient ly smallWAVE EQUATION WITH SPACE-DEPENDENT DAMPING 21\nνand sufficiently large t0so that the other bad terms are absorbed by these good\nterms.\nWe first give estimates of E0[w](t;t0,λ).\nLemma 5.3. Under the assumptions on Theorem 5.2, for any t0≥1andt >0,\nwe have\nd\ndtE0[w](t;t0,λ)≤ −η0/integraldisplay\nΩ|∇w(x,t)|2Ψ(x,t;t0)λdx\n−η0/integraldisplay\nΩa(x)|w(x,t)|2Ψ(x,t;t0)λ−1dx\n+C/integraldisplay\nΩ|∂tw(x,t)|2Ψ(x,t;t0)λdx\n+C/integraldisplay\nΩa(x)−1|F(x,t)|2Ψ(x,t;t0)λ+1dx\nwith some constants η0=η0(ε,δ)>0andC=C(N,α,δ,ε,λ )>0.\nProof.By the definition of E0[w](t;t0,λ) and using the equation (3.1), we calculate\nd\ndtE0[w](t;t0,λ) = 2/integraldisplay\nΩ|∂tw|2Φ−1+2δ\nβ,εdx+2/integraldisplay\nΩw/parenleftbig\n∂2\ntw+a(x)∂tw/parenrightbig\nΦ−1+2δ\nβ,εdx\n−(1−2δ)/integraldisplay\nΩ/parenleftbig\n2w∂tw+a(x)|w|2/parenrightbig\n(∂tΦβ,ε)Φ−2+2δ\nβ,εdx\n= 2/integraldisplay\nΩ|∂tw|2Φ−1+2δ\nβ,εdx+2/integraldisplay\nΩw(∆w+F)Φ−1+2δ\nβ,εdx\n−(1−2δ)/integraldisplay\nΩ/parenleftbig\n2w∂tw+a(x)|w|2/parenrightbig\n(∂tΦβ,ε)Φ−2+2δ\nβ,εdx.\nApplying Lemma 2.9, we have\nd\ndtE0[w](t;t0,λ)≤2/integraldisplay\nΩ|∂tw|2Φ−1+2δ\nβ,εdx−2δ\n1−δ/integraldisplay\nΩ|∇w|2Φ−1+2δ\nβ,εdx\n−(1−2δ)/integraldisplay\nΩ|w|2(a(x)∂tΦβ,ε−∆Φβ,ε)Φ−2+2δ\nβ,εdx\n−2(1−2δ)/integraldisplay\nΩw∂tw(∂tΦβ,ε)Φ−2+2δ\nβ,εdx\n+2/integraldisplay\nΩwFΦ−1+2δ\nβ,εdx. (5.1)\nBy Proposition 2.8 (iii) and (iv), the third term of the right-hand side o f (5.1) is\nestimated as\n−(1−2δ)/integraldisplay\nΩ|w|2(a(x)∂tΦβ,ε−∆Φβ,ε)Φ−2+2δ\nβ,εdx≤ −η/integraldisplay\nΩa(x)|w|2Ψλ−1dx\nwith someη >0. Moreover, by Proposition 2.8 (iii), the second term of the right-\nhand side of (5.1) is estimated as\n−2δ\n1−δ/integraldisplay\nΩ|∇w|2Φ−1+2δ\nβ,εdx≤ −η′/integraldisplay\nΩ|∇w|2Ψλdx (5.2)22 M. SOBAJIMA AND Y. WAKASUGI\nwith some η′>0. Moreover, we can drop the third term of the right-hand side,\nand we also have |∂tΦβ,ε|=|βΦβ+1,ε| ≤CΨ−β−1, which implies\n/vextendsingle/vextendsingle/vextendsinglew∂tw(∂tΦβ,ε)Φ−2+2δ\nβ,ε/vextendsingle/vextendsingle/vextendsingle≤C|w||∂tw|Ψλ−1.\nTherefore, from the above inequality with the Schwarz inequality, w e estimate the\nfourth term of the right-hand side of (5.1) as\n/vextendsingle/vextendsingle/vextendsingle/vextendsingle/integraldisplay\nΩw∂tw(∂tΦβ,ε)Φ−2+2δ\nβ,εdx/vextendsingle/vextendsingle/vextendsingle/vextendsingle\n≤C/integraldisplay\nΩ|w||∂tw|Ψλ−1dx\n≤C/parenleftbigg/integraldisplay\nΩa(x)|w|2Ψλ−1dx/parenrightbigg1/2/parenleftbigg/integraldisplay\nΩa(x)−1|∂tw|2Ψλ−1dx/parenrightbigg1/2\n≤η1/integraldisplay\nΩa(x)|w|2Ψλ−1dx+C(η1)/integraldisplay\nΩ|∂tw|2Ψλdx\nfor anyη1>0. Similarly, the last term of the right-hand side of (5.1) is estimated\nas\n/vextendsingle/vextendsingle/vextendsingle/vextendsingle2/integraldisplay\nΩwFΦ−1+2δ\nβ,εdx/vextendsingle/vextendsingle/vextendsingle/vextendsingle≤C/integraldisplay\nΩ|w||F|Ψλdx\n≤/parenleftbigg/integraldisplay\nΩa(x)|w|2Ψλ−1dx/parenrightbigg1/2/parenleftbigg/integraldisplay\nΩa(x)−1|F|2Ψλ+1dx/parenrightbigg1/2\n≤η2/integraldisplay\nΩa(x)|w|2Ψλ−1dx+C/integraldisplay\nΩa(x)−1|F|2Ψλ+1dx(5.3)\nfor anyη2>0. Therefore, by applying (5.2)–(5.3) to (5.1), we have the desired\nestimate. /square\nLemma 5.4. Under the assumptions on Theorem 5.2, there exists t2≥1such that\nfor anyt0≥t2andt>0, we have\nd\ndtE1[w](t;t0,λ)≤ −/integraldisplay\nΩa(x)|∂tw(x,t)|2Ψ(x,t;t0)λ+1dx+C/integraldisplay\nΩ|∇w(x,t)|2Ψ(x,t;t0)λdx\n+C/integraldisplay\nΩa(x)−1|F(x,t)|2Ψ(x,t;t0)λ+1dx\nwith some constant C=C(N,α,δ,ε,λ,t 2)>0.WAVE EQUATION WITH SPACE-DEPENDENT DAMPING 23\nProof.By the definition of E1[w](t;t0,λ) and the equation (3.1), we calculate\nd\ndtE1[w](t;t0,λ) = 2/integraldisplay\nΩ/parenleftbig\n∇∂tw·∇w+∂tw∂2\ntw/parenrightbig\nΨλ+1dx\n+(λ+1)/integraldisplay\nΩ/parenleftbig\n|∇w|2+|∂tw|2/parenrightbig\nΨλdx\n= 2/integraldisplay\nΩ∂tw/parenleftbig\n−∆w+∂2\ntw/parenrightbig\nΨλ+1dx\n−2(λ+1)/integraldisplay\nΩ∂tw(∇w·∇Ψ)Ψλdx\n+(λ+1)/integraldisplay\nΩ/parenleftbig\n|∇w|2+|∂tw|2/parenrightbig\nΨλdx\n=−2/integraldisplay\nΩa(x)|∂tw|2Ψλ+1dx+2/integraldisplay\nΩ∂twFΨλ+1dx\n−2(λ+1)/integraldisplay\nΩ∂tw(∇w·∇Ψ)Ψλdx\n+(λ+1)/integraldisplay\nΩ/parenleftbig\n|∇w|2+|∂tw|2/parenrightbig\nΨλdx. (5.4)\nHere, we note that the integration by parts in the second identity is justified, since\n∂tw∇wΨλ+1∈L1(Ω) for each t≥0. For the second term of the right-hand side of\n(5.4), we apply the Schwarz inequality to obtain\n|2∂twF| ≤a(x)\n4|∂tw|2+4a(x)−1|F|2. (5.5)\nNext, by the Schwarz inequality, the third term of the right-hand s ide of (5.4) is\nestimated as\n|−2(λ+1)∂tw(∇w·∇Ψ)| ≤a(x)\n4|∂tw|2Ψ+C|∇w|2|∇Ψ|2\na(x)Ψ\n≤a(x)\n4|∂tw|2Ψ+C|∇w|2,\nwhere we have also used\n|∇Ψ|2\na(x)Ψ≤|∇Aε(x)|2\na(x)Aε(x)≤2−α\nN−α+ε,\nwhich follows from (2.3). Moreover, for the last term of the right-h and side of\n(5.4), we note that Ψ−1≤t−1+α\n2−α\n0Aε(x)−α\n2−α≤Ct−1+α\n2−α\n0a(x)≤a(x)\n2(λ+1)holds for\nt0≥t2, provided that t2is sufficiently large. Thus, we have\n/vextendsingle/vextendsingle/vextendsingle/vextendsingle(λ+1)/integraldisplay\nΩ|∂tw|2Ψλdx/vextendsingle/vextendsingle/vextendsingle/vextendsingle≤1\n2/integraldisplay\nΩa(x)|∂tw|2Ψλ+1dx. (5.6)\nFinally, applying (5.5)–(5.6) to (5.4), we have the desired estimate. /square\nNow we are in the position to prove Theorem 5.2.24 M. SOBAJIMA AND Y. WAKASUGI\nProof of Theorem 5.2. Lett2be the constant given in Lemma 5.4. For t0≥t2, by\nLemmas 5.3 and 5.4, we calculate\nd\ndtE[w](t;t0,λ,ν)≤/integraldisplay\nΩ/parenleftbig\n−νa(x)+CΨ−1/parenrightbig\n|∂tw|2Ψλ+1dx\n+(νC−η0)/integraldisplay\nΩ|∇w|2Ψλdx\n+C/integraldisplay\nΩa(x)−1|F|2Ψλ+1dx. (5.7)\nBy takingν >0 sufficiently small so that νC−η0<0. After that, taking t∗≥t2\nsufficiently large so that\n−νa(x)+CΨ−1≤ −νa(x)+Ct−1+α\n2−α\n0a(x)≤ −ν\n2a(x)\nholds fort0≥t∗. Therefore, integrating (5.7) over [0 ,t], we have\nE[w](t;t0,λ,ν)+η∗/integraldisplayt\n0/integraldisplay\nΩa(x)|∂tw|2Ψλ+1dxdτ\n+η∗/integraldisplayt\n0/integraldisplay\nΩ|∇w|2Ψλdxdτ\n≤E[w](0;t0,λ,ν)+C/integraldisplayt\n0/integraldisplay\nΩa(x)−1|F|2Ψλ+1dxdτ\nwith some constant η∗>0. By the assumptions of the theorem, the right-hand\nside is bounded with respect to t>0. Thus, we complete the proof. /square\n5.2.Higher order energy estimates.\nDefinition 5.5. Letδ∈(0,1/2),ε∈(0,1),λ∈[0,(1−2δ)γε), whereγεis defined\nin(2.4), and letν >0. For an integer j≥1, we define\nE(j)\n1[w](t;t0,λ) :=/integraldisplay\nΩ/parenleftbig\n|∇w(x,t)|2+|∂tw(x,t)|2/parenrightbig\nΨ(x,t;t0)λ+1+2jdx,\nE(j)\n0[w](t;t0,λ) :=/integraldisplay\nΩ/parenleftbig\n2w(x,t)∂tw(x,t)+a(x)|w(x,t)|2/parenrightbig\nΨ(x,t;t0)λ+2jdx,\nE(j)[w](t;t0,λ,ν) :=νE(j)\n1[w](t;t0,λ)+E(j)\n0[w](t;t0,λ)\nfort≥0.\nWe remark that there exists t1>0 such that for any t0≥t1andt≥0, we have\nE(j)[w](t;t0,λ,ν)∼/integraldisplay\nΩ/bracketleftbig/parenleftbig\n|∇w|2+|∂tw|2/parenrightbig\nΨλ+1+2j+a(x)|w|2Ψλ+2j/bracketrightbig\ndx.\nThe main result of this subsection is the following energy estimates fo r higher\norder derivatives of the solution of the damped wave equation (3.1) .\nTheorem 5.6. Letk≥1be an integer. Assume that a(x)satisfies (1.2). Let\nδ∈(0,1/2),ε∈(0,1/2), andλ∈[0,(1−2δ)γε). Then, there exist constants\nν(j)=ν(j)(N,α,λ,j )>0andt(j)\n∗=t(j)\n∗(N,α,δ,ε,λ,ν(j),j)≥1forj= 1,...,kWAVE EQUATION WITH SPACE-DEPENDENT DAMPING 25\nsuch that for any t0≥max1≤j≤kt(j)\n∗, the following holds: let m= (λ+1+2k)2−α\n2\nand assume F∈/intersectiontextk\nj=0Cj+1([0,∞);Hk−j,m(Ω))satisfies\n/integraldisplay\nΩa(x)−1|∂j\ntF(x,t)|2Ψ(x,t;t0)λ+1+2jdx∈L1(0,∞)\nforj= 0,1,...,k. Letwbe the solution of (3.1)in Theorem 3.4 with the initial\ndata(w0,w1)∈Hk+2,m(Ω)×Hk+1,m(Ω)satisfying the k-th order compatibility\ncondition in the sense of Theorem 3.4. Then, we have\nE(j)[∂j\ntw](t;t0,λ,ν(j))∈L∞(0,∞),\n/integraldisplay\nΩa(x)|∂j+1\ntw(x,t)|2Ψ(x,t;t0)λ+1+2jdx∈L1(0,∞)\nforj= 1,...,k.\nThe proof of Theorem 5.6 is based on an induction argument, which is s imilar\nto that of Lemma 4.4. The main part of the induction argument is the f ollowing\nlemma.\nLemma 5.7. Letj∈N. Assumea(x)satisfies (1.2). Letλ≥0andm= (λ+\n1+2j)(2−α)/2. Then, there exist constants ν(j)=ν(j)(N,α,λ,j )>0andt(j)\n∗=\nt(j)\n∗(N,α,λ,ν(j),j)≥1such that for any t0≥t(j)\n∗, the following holds: Assume F∈\nC1([0,∞);H0,m(Ω))and letwbe the solution of (3.1)with initial data (w0,w1)∈\n(H2,m(Ω)∩H1,m\n0(Ω))×H1,m\n0(Ω). If\n/integraldisplay\nΩa(x)−1|F(x,t)|2Ψ(x,t;t0)λ+1+2jdx∈L1(0,∞),\n/integraldisplay\nΩa(x)|w(x,t)|2Ψ(x,t;t0)λ−1+2jdx∈L1(0,∞)\nare satisfied, then\nE(j)[w](t;t0,λ)∈L∞(0,∞),\n/integraldisplay\nΩa(x)|∂tw(x,t)|2Ψ(x,t;t0)λ+1+2jdx∈L1(0,∞)\nhold.\nFor the proof of Lemma 5.7, we further prepare the following two lem mas.\nLemma 5.8. Under the assumptions of Lemma 5.7, we have\nd\ndtE(j)\n0[w](t;t0,λ)≤C/integraldisplay\nΩ|∂tw|2Ψλ+2jdx−/integraldisplay\nΩ|∇w(x,t)|2Ψ(x,t;t0)λ+2jdx\n+C/integraldisplay\nΩa(x)|w(x,t)|2Ψ(x,t;t0)λ−1+2jdx\n+C/integraldisplay\nΩa(x)−1|F(x,t)|2Ψ(x,t;t0)λ+1+2jdx\nwith some constant C >0.26 M. SOBAJIMA AND Y. WAKASUGI\nProof.We compute\nd\ndtE(j)\n0[w](t;t0,λ) = 2/integraldisplay\nΩ|∂tw|2Ψλ+2jdx+2/integraldisplay\nΩw/parenleftbig\n∂2\ntw+a(x)∂tw/parenrightbig\nΨλ+2jdx\n+2(λ+2j)/integraldisplay\nΩ/parenleftbig\n2w∂tw+a(x)|w|2/parenrightbig\nΨλ−1+2jdx\n= 2/integraldisplay\nΩ|∂tw|2Ψλ+2jdx+2/integraldisplay\nΩw(∆w+F)Ψλ+2jdx\n+2(λ+2j)/integraldisplay\nΩ/parenleftbig\n2w∂tw+a(x)|w|2/parenrightbig\nΨλ−1+2jdx\n= 2/integraldisplay\nΩ|∂tw|2Ψλ+2jdx−2/integraldisplay\nΩ|∇w|2Ψλ+2jdx\n−2(λ+2j)/integraldisplay\nΩw(∇w·∇Ψ)Ψλ−1+2jdx\n+2(λ+2j)/integraldisplay\nΩ/parenleftbig\n2w∂tw+a(x)|w|2/parenrightbig\nΨλ−1+2jdx\n+2/integraldisplay\nΩwFΨλ+2jdx.\nThe Schwarz inequality implies\n/vextendsingle/vextendsingle/vextendsingle/vextendsingle2(λ+2j)/integraldisplay\nΩw(∇w·∇Ψ)Ψλ−1+2jdx/vextendsingle/vextendsingle/vextendsingle/vextendsingle\n≤/integraldisplay\nΩ|∇w|2Ψλ+2jdx+C/integraldisplay\nΩa(x)|w|2|∇Ψ|2\na(x)ΨΨλ−1+2jdx\n≤/integraldisplay\nΩ|∇w|2Ψλ+2jdx+C/integraldisplay\nΩa(x)|w|2Ψλ−1+2jdx,\nwhere we have used ∇Ψ =∇Aε(x), Ψ(x,t;t0)≥Aε(x), and (2.3) in Lemma 2.1.\nSimilarly, we have\n/vextendsingle/vextendsingle/vextendsingle/vextendsingle2(λ+2j)/integraldisplay\nΩ2w∂twΨλ−1+2jdx/vextendsingle/vextendsingle/vextendsingle/vextendsingle\n≤C/integraldisplay\nΩ|∂tw|2Ψλ+2jdx+C/integraldisplay\nΩa(x)|w|2a(x)−1Ψλ−2+2jdx\n≤C/integraldisplay\nΩ|∂tw|2Ψλ+2jdx+C/integraldisplay\nΩa(x)|w|2Ψλ−1+2jdx,\nand\n/vextendsingle/vextendsingle/vextendsingle/vextendsingle2/integraldisplay\nΩwFΨλ+2jdx/vextendsingle/vextendsingle/vextendsingle/vextendsingle\n≤C/integraldisplay\nΩa(x)|w|2a(x)λ−1+2jdx+C/integraldisplay\nΩa(x)−1|F|2Ψλ+1+2jdx.\nThis completes the proof. /squareWAVE EQUATION WITH SPACE-DEPENDENT DAMPING 27\nLemma 5.9. Under the assumptions of Lemma 5.7, there exists t2≥1such that\nfor anyt0≥t2andt>0, we have\nE(j)\n1[w](t;t0,λ)≤ −/integraldisplay\nΩa(x)|∂tw(x,t)|2Ψ(x,t;t0)λ+1+2jdx\n+C/integraldisplay\nΩ|∇w(x,t)|2Ψ(x,t;t0)λ+2jdx\n+C/integraldisplay\nΩa(x)−1|F(x,t)|2Ψ(x,t;t0)λ+1+2jdx\nwith some constant C=C(N,α,δ,ε,λ,j,t 2)>0.\nThe proof is completely the same as that of Lemma 5.4 by replacing λbyλ+2j.\nThus, we omit the detail.\nProof of Lemma 5.7. By Lemmas 5.8 and 5.9, taking ν(j)>0 sufficiently small,\nand then, taking t(j)\n∗≥t2sufficiently large depending on ν(j), we have\nd\ndtE(j)[w](t;t0,λ,ν(j)) =ν(j)d\ndtE(j)\n1[w](t;t0,λ)+d\ndtE(j)\n0[w](t;t0,λ)\n≤ −η1/integraldisplay\nΩa(x)|∂tw|2Ψλ+1+2jdx−η2/integraldisplay\nΩ|∇w|2Ψλ+2jdx\n+C/integraldisplay\nΩa(x)|w|2Ψλ−1+2jdx+C/integraldisplay\nΩa(x)−1|F|2Ψλ+1+2jdx\nfort0≥t(j)\n∗andt >0 with some constants η1,η2>0. By integrating the above\ninequality on [0 ,t] and using the assumptions, we conclude\nE(j)[w](t;t0,λ,ν(j))+/integraldisplayt\n0/integraldisplay\nΩa(x)|∂tw|2Ψλ+1+2jdxdτ+/integraldisplayt\n0/integraldisplay\nΩ|∇w|2Ψλ+2jdxdτ\n≤E(j)[w](0;t0,λ,ν(j))+C/integraldisplayt\n0/integraldisplay\nΩa(x)|w|2Ψλ−1+2jdxdτ\n+C/integraldisplayt\n0/integraldisplay\nΩa(x)−1|F|2Ψλ+1+2jdxdτ,\nand the proof is complete. /square\nProof of Theorem 5.6. By Theorem 5.2, there exist ν >0 andt∗≥1 such that\nE[w](t;t0,λ,ν)∈L∞(0,∞),\n/integraldisplay\nΩa(x)|∂tw(x,t)|2Ψ(x,t;t0)λ+1dx∈L1(0,∞) (5.8)\nhold fort0≥t∗. Now, thanks to the property (5.8), we apply Lemma 5.7 with\nj= 1 and the replacement of wandFby∂twand∂tF, respectively. Then, there\nexistν(1)>0 andt(1)\n∗≥1 such that\nE(1)[∂tw](t;t0,λ,ν(1))∈L∞(0,∞),\n/integraldisplay\nΩa(x)|∂2\ntw(x,t)|2Ψ(x,t;t0)λ+1+2dx∈L1(0,∞)28 M. SOBAJIMA AND Y. WAKASUGI\nhold fort0≥t(1)\n∗. The latter property allows us to apply again Lemma 5.7 with\nj= 2 and the replacement of wandFby∂2\ntwand∂2\ntF, respectively. Repeating\nthis argument until j=k, we reach the conclusion of Theorem 5.6. /square\n6.Proof of the asymptotic expansion\nIn this section, we give the estimates of the right-hand side of (3.9) and complete\nthe proof of Theorem 1.1.\nLetn∈Nbe fixed, and let δ∈(0,1/2),ε∈(0,1/2),λ∈[0,(1−2δ)γε), and\nt0≥1. Moreover, we define λj=λ−2α\n2−αjforj= 1,...,n+ 1, and we assume\nthatλn+1∈[0,(1−2δ)γε), that is,λ∈[2α\n2−α(n+1),(1−2δ)γε). Here, we note that\nthe assumption n+10 andt(n)\n∗≥1 such that for any t0≥t(n)\n∗we have\nE(n)[∂n\ntUn+1](t;t0,λn+1,ν(n))∈L∞(0,∞).\nIn particular, the above bound yields/integraldisplay\nΩ|∂n+1\ntUn+1(x,t)|2dx≤C(1+t)−λn+1−1−2n.\nNamely, we conclude/vextenddouble/vextenddouble/vextenddouble/vextenddouble/vextenddouble/vextenddoubleu(t)−n/summationdisplay\nj=0∂j\ntVj(t)/vextenddouble/vextenddouble/vextenddouble/vextenddouble/vextenddouble/vextenddouble\nL2(Ω)=/ba∇dbl∂n+1\ntUn+1(t)/ba∇dblL2(Ω)\n≤C(1+t)−λ\n2−(2n+1)(1−α)\n2−α+α\n2(2−α),\nwhich completes the proof of Theorem 1.1.\nAcknowledgements\nThis work was supported by JSPS KAKENHI Grant Numbers JP16K17 625,\nJP18H01132, 18K13445, JP20K14346.\nReferences\n[1]R. Beals, R. Wong , Special functions, A graduate text. Cambridge Studies in A dvanced\nMathematics 126, Cambridge University Press, Cambridge, 2 010.\n[2]T. Cazenave, A. Haraux , An Introduction to Semilinear Evolution Equation, Oxford Sci-\nence, 1998.\n[3]R. Chill, A. Haraux ,An optimal estimate for the difference of solutions of two abs tract\nevolution equations , J. Differential Equations 193(2003), 385–395.\n[4]Th. Gallay, G. Raugel ,Scaling variables and asymptotic expansions in damped wave equa-\ntions, J. Differential Equations 150(1998), pp. 42–97.\n[5]T. Hosono, T. Ogawa ,Large time behavior and Lp-Lqestimate of solutions of 2-\ndimensional nonlinear damped wave equations , J. Differential Equations 203(2004), 82–118.\n[6]L. Hsiao, T.-P. Liu ,Convergence to nonlinear diffusion waves for solutions of a s ystem of\nhyperbolic conservation laws with damping , Comm. Math. Phys. 43(1992), 599–605.\n[7]M. Ikawa ,Mixed problems for hyperbolic equations of second order , J. Math. Soc. Japan 20\n(1968), 580–608.\n[8]R. Ikehata ,Diffusion phenomenon for linear dissipative wave equations in an exterior do-\nmain, J. Differential Equations 186(2002), 633–651.\n[9]R. Ikehata ,Some remarks on the wave equation with potential type dampin g coefficients ,\nInt. J. Pure Appl. Math. 21(2005), 19–24.\n[10]R. Ikehata, K. Nishihara ,Diffusion phenomenon for second order linear evolution equa -\ntions, Studia Math. 158(2003), 153–161.\n[11]R. Ikehata, G. Todorova, B. Yordanov ,Optimal decay rate of the energy for wave equa-\ntions with critical potential , J. Math. Soc. Japan 65(2013), 183–236.\n[12]R. Joly, J. Royer ,Energy decay and diffusion phenomenon for the asymptoticall y periodic\ndamped wave equation J. Math. Soc. Japan 70(2018), 1375–1418.\n[13]G. Karch ,Selfsimilar profiles in large time asymptotics of solutions to damped wave equa-\ntions, Studia Math. 143(2000), 175–197.32 M. SOBAJIMA AND Y. WAKASUGI\n[14]P. Marcati, K. Nishihara ,TheLp-Lqestimates of solutions to one-dimensional damped\nwave equations and their application to the compressible flo w through porous media , J. Dif-\nferential Equations 191(2003), 445–469.\n[15]A. Matsumura ,On the asymptotic behavior of solutions of semi-linear wave equations , Publ.\nRes. Inst. Math. Sci. 12(1976), 169–189.\n[16]T. Matsuyama ,Asymptotic behavior of solutions for the wave equation with an effective\ndissipation around the boundary , J. Math. Anal. Appl. 271(2002), 467–492.\n[17]H. Michihisa ,L2-asymptotic profiles of solutions to linear damped wave equa tions, J. Dif-\nferential Equations 296(2021), 573–592.\n[18]K. Mochizuki ,Scattering theory for wave equations with dissipative term s, Publ. Res. Inst.\nMath. Sci. 12(1976), 383–390.\n[19]K. Mochizuki, H. Nakazawa ,Energy decay and asymptotic behavior of solutions to the\nwave equations with linear dissipation , Publ. RIMS, Kyoto Univ. 32(1996), 401–414.\n[20]T. Narazaki ,Lp-Lqestimates for damped wave equations and their applications to semi-\nlinear problem , J. Math. Soc. Japan 56(2004), 585–626.\n[21]K. Nishihara ,Lp-Lqestimates of solutions to the damped wave equation in 3-dime nsional\nspace and their application , Math. Z. 244(2003), 631–649.\n[22]K. Nishihara ,Decay properties for the damped wave equation with space dep endent potential\nand absorbed semilinear term , Comm. Partial Differential Equations 35(2010), 1402–1418.\n[23]H. Nishiyama ,Remarks on the asymptotic behavior of the solution to damped wave equations ,\nJ. Differential Equations 261(2016) 3893–3940.\n[24]R. Orive, E. Zuazua, A. F. Pazoto ,Asymptotic expansion for damped wave equations with\nperiodic coefficients , Math. Models Methods Appl. Sci. 11(2001), 1285–1310.\n[25]P. Radu, G. Todorova, B. Yordanov ,Higher order energy decay rates for damped wave\nequations with variable coefficients , Discrete Contin. Dyn. Syst. Ser. S. 2(2009), 609–629.\n[26]P. Radu, G. Todorova, B. Yordanov ,Decay estimates for wave equations with variable\ncoefficients , Trans. Amer. Math. Soc. 362(2010), 2279–2299.\n[27]P. Radu, G. Todorova, B. Yordanov ,Diffusion phenomenon in Hilbert spaces and appli-\ncations, J. Differential Equations 250(2011), 4200–4218.\n[28]P. Radu, G. Todorova, B. Yordanov ,The generalized diffusion phenomenon and appli-\ncations, SIAM J. Math. Anal. 48(2016), 174–203.\n[29]S. Sakata, Y. Wakasugi Movement of time-delayed hot spots in Euclidean space , Math. Z\n285(2017), 1007–1040.\n[30]M. Sobajima ,Global existence of solutions to semilinear damped wave equ ation with slowly\ndecaying initial data in exterior domain , Differential and integral equations 32(2019), 615–\n638.\n[31]M. Sobajima ,Higher order asymptotic expansion of solutions to abstract linear hyperbolic\nequations , Math. Ann. 380(2021), 1–19.\n[32]M. Sobajima, Y. Wakasugi ,Diffusion phenomena for the wave equation with space-\ndependent damping in an exterior domain , J. Differential Equations 261(2016), 5690–5718.\n[33]M. Sobajima, Y. Wakasugi ,Remarks on an elliptic problem arising in weighted energy\nestimates for wave equations with space-dependent damping term in an exterior domain ,\nAIMS Mathematics, 2 (2017), 1–15.\n[34]M. Sobajima, Y. Wakasugi ,Diffusion phenomena for the wave equation with space-\ndependent damping term growing at infinity , Adv. Differential Equations 23(2018), 581–614.\n[35]M. Sobajima, Y. Wakasugi ,Weighted energy estimates for wave equation with space-\ndependent damping term for slowly decaying initial data , Commun. Contemp. Math. 21\n(2019), no. 5, 1850035, 30 pp.\n[36]M. Sobajima, Y. Wakasugi ,Supersolutions for parabolic equations with unbounded or d e-\ngenerate diffusion and its applications to some classes of pa rabolic and hyperbolic equations ,\nJ. Math. Soc. Japan, 73(2021), 1091–1128.\n[37]H. Takeda ,Higher-order expansion of solutions for a damped wave equat ion, Asymptotic\nAnalysis 94(2015), pp. 1–31.\n[38]G. Todorova, B. Yordanov ,Weighted L2-estimates for dissipative wave equations with\nvariable coefficients , J. Differential Equations 246(2009), 4497–4518.\n[39]Y. Wakasugi ,On diffusion phenomena for the linear wave equation with spac e-dependent\ndamping , J. Hyp. Diff. Eq. 11(2014), 795–819.WAVE EQUATION WITH SPACE-DEPENDENT DAMPING 33\n[40]Y. Wakasugi ,Second order asymptotic expansion for wave equations with t ime- dependent\ndissipation in one-space dimension , in: Advanced Studies in Pure Mathematics 81 (2019),\nAsymptotic Analysis for Nonlinear Dispersive and Wave Equa tions, 401–419.\n[41]J. Wirth ,Solution representations for a wave equation with weak diss ipation, Math. Meth.\nAppl. Sci. 27(2004), 101–124.\n[42]J. Wirth ,Asymptotic properties of solutions to wave equations with t ime-dependent dissi-\npation, PhD thesis, TU Bergakademie Freiberg, 2005.\n[43]J. Wirth ,Wave equations with time-dependent dissipation I. Non-effe ctive dissipation , J.\nDifferential Equations 222(2006), 487–514.\n[44]J. Wirth ,Wave equations with time-dependent dissipation II. Effecti ve dissipation , J. Dif-\nferential Equations 232(2007), 74–103.\n[45]J. Wirth ,Scattering and modified scattering for abstract wave equati ons with time-dependent\ndissipation , Adv. Differential Equations 12(2007), 1115–1133.\n[46]T. Yamazaki ,Asymptotic behavior for abstract wave equations with decay ing dissipation ,\nAdv. Differential Equations 11(2006), 419–456.\n[47]H. Yang, A. Milani ,On the diffusion phenomenon of quasilinear hyperbolic waves , Bull.\nSci. Math. 124(2000), 415–433.\nDepartment of Mathematics, Faculty of Science and Technolo gy, Tokyo University\nof Science, 2641 Yamazaki, Noda-shi, Chiba, 278-8510, Japan\nEmail address :msobajima1984@gmail.com\nLaboratory of Mathematics, Graduate School of Engineering , Hiroshima University,\nHigashi-Hiroshima, 739-8527, Japan\nEmail address :wakasugi@hiroshima-u.ac.jp" }, { "title": "1207.6484v1.The_effect_of_non_uniform_damping_on_flutter_in_axial_flow_and_energy_harvesting_strategies.pdf", "content": "The e\u000bect of non-uniform damping on \rutter in axial \row and energy harvesting\nstrategies\nKiran Singh,1,\u0003S\u0013 ebastien Michelin,1,yand Emmanuel de Langre1,z\n1Department of Mechanics, LadHyX,\n\u0013Ecole Polytechnique, 91128, Palaiseau, France\nThe problem of energy harvesting from \rutter instabilities in \rexible slender structures in axial\n\rows is considered. In a recent study, we used a reduced order theoretical model of such a system to\ndemonstrate the feasibility for harvesting energy from these structures. Following this preliminary\nstudy, we now consider a continuous \ruid-structure system. Energy harvesting is modelled as\nstrain-based damping and the slender structure under investigation lies in a moderate \ruid loading\nrange, for which the \rexible structure may be destabilised by damping. The key goal of this work\nis to analyse the e\u000bect of damping distribution and intensity on the amount of energy harvested\nby the system. The numerical results indeed suggest that non-uniform damping distributions may\nsigni\fcantly improve the power harvesting capacity of the system. For low damping levels, clustered\ndampers at the position of peak curvature are shown to be optimal. Conversely for higher damping,\nharvesters distributed over the whole structure are more e\u000bective.\nI. INTRODUCTION\nIncreasing energy demands motivate the interest in energy harvesting concepts, where the idea is to harness the\nenergy of naturally occurring phenomena. At the scale of kilowatts, concepts include energy harvesting from tidal\ncurrents [1] and ocean waves [2]. At the lower end of the power spectrum, concepts based on photo/thermovoltaics\nand magneto/piezoelectrics show the scope for powering sensors and mobile electronic devices [3]; these include energy\nscavenging from ambient vibrations in structures such as buildings and bridges and oscillatory motion of wheels in\nautomobiles or turbines in engines [4, 5]. Energy harvesting from \ruid-structure interactions (FSI) includes concepts\nsuch as vortex induced vibrations (VIV) of blu\u000b bodies in cross-\row [6, 7], resonant vibrations induced in aerofoils\nmounted on elastic supports [8, 9], \rutter of \rexible plates [10{12], and combinations thereof, as for the coupled\nVIV-\rutter energy harvester examined by [13]. In this work we focus on harvesting energy from \rutter instabilities\nof slender structures in axial \row.\nThe classical description of \rutter instabilities is self-sustained oscillations that arise due to the unstable coupling of\n\ruid dynamic pressure and structural bending modes, where for undamped structures the critical speed at \rutter onset\ndepends on \ruid as well as structural properties [14]. Flutter instabilities are observed in a diversity of con\fgurations,\nwhich may be broadly classi\fed as internal or external \rows [14, 15]. Internal \row instabilities are observed in \rexible\npipes and channels and are invariably motivated by biological phenomena such as \row induced oscillations in airways\nand veins [16, 17]. The pipe-conveying \ruid is a canonical problem that yields deep insights into FSI; in particular\n[18] examined the relationship between local and global instabilities and showed that locally stable con\fgurations may\nbecome unstable due to wave re\rections at \fnite boundaries; this destabilising mechanism was recently predicted for\ncompliant channels as well [19]. External \row based instabilities include \rapping \rags [20{22] and panels [23] in a\nsteady \row. For all aspect ratios, the plate can become unstable due to \ruid loading, de\fned as the ratio of \ruid and\nstructure inertia, and may be represented as a nondimensional length [24] or mass ratio [25]. [26] and more recently\n[27] examined the occurrence of instabilities in slender structures. They used experimental and theoretical techniques\nto examine the role of inviscid and viscous drag on static and dynamic instabilities that arise in \rexible cylinders in\naxial \row. For a recent review on the \rutter dynamics of \rexible bodies in external \row, the reader is referred to [28]\nand references therein.\nIn this work, we analyse the scope for harvesting energy in slender elastic cantilevered structures that \rutter in\na steady \row. From the point of view of the \ruid-solid system, energy harvesters are essentially an energy sink,\ntherefore for the theoretical approach adopted here they are modelled as internal damping in the structure. As a \frst\nstep, we examined the feasibility of this concept using a nonlinear model of a reduced-order system, consisting of a\nslender cylinder pair connected by discrete springs and dampers [29]. It was shown that the optimal con\fguration for\n\u0003Electronic address: kiran.singh@cantab.net; Present address:OCCAM, The Mathematical Institute, 24-29 St Giles, OX13LB, Oxford\nyElectronic address: sebastien.michelin@ladhyx.polytechnique.fr\nzElectronic address: delangre@ladhyx.polytechnique.frarXiv:1207.6484v1 [physics.flu-dyn] 27 Jul 20122\nFIG. 1: Slender cantilevered structure placed in a steady axial \row of velocity U1exand \fxed in O. The instantaneous\ndeformation of the inextensible beam is measured by the tangent angle \u0012(s;t) with respect to exof the local tangent \u001c.\nthis two-degree-of-freedom system is one with energy harvesters positioned away from the instability source; such a\ncon\fguration maintains self-sustained \rapping in the presence of structural damping.\nIn this work we generalise this approach for a continuous system, and seek to maximise energy harvesting through\ncarefully-tailored distributions of structural damping. Conventionally, structural instabilities are stabilised by damping\n[16, 30] . However, damping may be destabilising under moderate to heavy \ruid loading conditions, as observed in\nin\fnite plates [31], \rags [11] and \ruid-conveying pipes [32]. [29] show that energy harvesting requires the presence\nof two traditionally competitive elements { \rutter oscillations and damping. A con\fguration for which \rutter is\ndestabilised by damping is especially interesting from an energy harvesting perspective. This idea has been exploited\nfor piezoelectric based energy harvesting from \rapping \rags: [11] used linear theory to show the gain in conversion\ne\u000eciency for \rags with high \ruid loading destabilised by piezoelectric based damping. In this work, we explore this\nidea using a general model for structural damping (equivalently energy harvesting) and a nonlinear model of the \ruid\nstructure interaction of a slender structure in a mean \row.\nIt is worth noting that existing insights on damping are generally based on the assumption of a constant distribution\nof damping in the structure [see for example 10], whilst little is known about how nonuniform damping distributions\na\u000bect the dynamics. In this work we seek clarity on the role of damping distribution on the \rutter response of slender\nstructures. The speci\fc motivation is to identify physical mechanisms that maximise this dissipated (i.e. harvested)\npower.\nThis paper is organised as follows: in Section II, the model used for the dynamics of slender structures with\nnon-uniform damping in an axial \row is presented. In Section III, the case of uniform damping is considered as a\nreference con\fguration and the role of \ruid loading on destabilisation by damping is discussed. Computations are\nthen performed for a neutrally buoyant slender cylinder with moderate \ruid loading, and the role of damping on the\nharvested power and \rutter response is examined. Section IV investigates non-uniform damping distributions and\nseeks optimals on two di\u000berent families of damping functions, either distributed over the whole structure or focused\non a particular region. In Section V, the impact of damping distribution on the \rutter dynamics is investigated further\nto understand the fundamental di\u000berence between optimal con\fgurations at low and intermediate damping.\nII. FLUID-SOLID MODEL\nWe consider a cantilevered (clamped-free, \fxed at O) slender structure of length Lwith crosswise dimension D,\ndensity\u001as, sti\u000bnessB, and nonuniform structural damping B\u0003(s). The slender solid is immersed in a stream of \ruid of\ndensity\u001amoving at mean speed U1, and the solid motion is con\fned to the ( ex;ey) plane (Figure 1). The equations\nof motion are nondimensionalised by the characteristic system scales: \u001a; L;U1.\nA. Nonlinear beam model\nThe \rexible structure is modelled as an inextensible Euler-Bernoulli beam, where r(s) is the position vector in the\n\fxed coordinate system ( ex;ey) andsis the curvilinear coordinate. At each point along the beam, the orientation\n\u0012(s;t) is de\fned as the angle of the tangent vector \u001c(s) with the horizontal; n(s) is the local normal. The nonlinear3\nequation of motion for the beam subjected to a \ruid force, f, is:\n1\nM\u0003@2r\n@t2=@\n@s\u001a\n\u0017\u001c\u00001\nM\u0003U\u00032\u0014@2\u0012\n@s2+@\n@s\u0012\n\u0018(s)@2\u0012\n@s@t\u0013\u0015\nn\u001b\n+f; (1)\nwhere the internal tension, \u0017(s;t), is essentially a Lagrange multiplier to satisfy the inextensibility condition @r=@s=\u001c\nand\u0018(s) =U1B\u0003(s)=(BL) is the non-dimensional damping distribution. A Kelvin-Voigt damping model is considered\nhere, generalising previous contributions [e.g. 10, 14, 32, 33] to nonuniform damping distributions.\nThe clamped-free boundary conditions must also be satis\fed, namely at the \fxed end ( s= 0):\n\u0012= 0; r= 0; (2)\nand at the free end ( s= 1):\n@\u0012\n@s+\u0018@2\u0012\n@s@t= 0;@2\u0012\n@s2+@\n@s\u0012\n\u0018@2\u0012\n@s@t\u0013\n= 0; \u0017 = 0: (3)\nConsistently with prior work [21, 25], the nondimensional mass ratio ( M\u0003) and \row speed ( U\u0003) are de\fned as:\nM\u0003=\u001aDL\n\u001asA; U\u0003=U1L\u0012\u001asA\nB\u00131=2\n; (4)\nwithAthe cross-sectional area of the structure.\nB. Fluid dynamic model\nIn the limit of slender structures ( D\u001cL), and for purely potential \row upstream of the structure's trailing edge,\nLighthill's large amplitude elongated-body theory [34] provides a leading order expression for the `reactive' force fi\napplied by the \row on the \rapping body, associated with the local transverse motion of each cross-section\nfi=\u0000ma\nM\u0003\u0012@(unn)\n@t\u0000@(unu\u001cn)\n@s+1\n2@(u2\nn\u001c)\n@s\u0013\n; (5)\nwhereu\u001c\u001c+unn=@r=@t\u0000exis the solid's local velocity relative to the incoming \row, and ma=Ma=\u001asA, where\nMais the dimensional added mass per unit length of the cross-section. [35] showed that Lighthill's theory compares\nwell with Reynolds-averaged Navier{Stokes (RANS) simulations to compute the forces on a swimming \fsh during\ntransient manoeuvres. This reactive force does not include any \row separation associated with the transverse motion\nof each cross-section, and as emphasised in [35], for freely \rapping bodies, an additional contribution for the \ruid\nforce must be included to account for such dissipative e\u000bects. Here, an empirical `resistive' force model is therefore\nadded following [36] and [33]:\nfv=\u00001=2Cdunjunjn; (6)\nwhereCDis the empirical drag coe\u000ecient, and CD= 1 is used in the following for circular cross-sections [see 29, for\na discussion of the impact of this coe\u000ecient on the \rapping dynamics]. [37] tuned the drag coe\u000ecients to show good\nagreement with direct numerical simulations.\nThus the \ruid force, f, in Eq. (1) is modelled as the sum of the reactive ( fi) and resistive ( fv) components. In the\nremainder of the paper we assume a neutrally buoyant circular cylinder, therefore \u001a=\u001asandA=\u0019D2=4. Unless\notherwise stated, we assume D=L = 0:1.\nNote that the \ruid force description is purely local here, and does not explicitly account for wake e\u000bects. When the\nslender body assumption is not veri\fed, and in particular, in the case of two-dimensional plates, an explicit description\nbecomes necessary [see for example 20, 21, 38].\nC. Energy harvester model\nAs noted earlier, energy harvesting is represented as a strain-based damping \u0018(s). We focus here on the mean\nnondimensional harvested power:\nP=P\n\u001aDLU31=1\nM\u0003U\u00032Z1\n0\u0018(s)\n_\u00142\u000b\nds; (7)4\n(a)\n (b)\nFIG. 2: (a) Maximum de\rection ymax(solid) and orientation \u0012max(dashed) at the free end as a function of the non-dimensional\n\row velocity U\u0003forM\u0003= 12:7. (b) Snapshots of the beam response for U\u0003= 10, 13 and 19 (from top to bottom).\nwherePis the mean dimensional harvested power, _ \u0014is the time derivative of the local curvature \u0014andh\u0001iis the time\naverage taken over a period Tof the limit-cycle oscillation. Pcan also be understood as the e\u000eciency of the system.\nAs discussed in Section I, the presence of structural damping can a\u000bect the \rutter response. The intensity and\ndistribution of damping are characterised by\n\u00180=Z1\n0\u0018(s)ds;and ~\u0018(s) =\u0018(s)=\u00180; (8)\nEquation (8) allows us to independently evaluate the impact on the system response of (i) the amount of damping \u00180\nand (ii) its spatial distribution.\nD. Numerical solution\nEquation (1) is solved numerically together with boundary conditions (2){(3) using an iterative second-order implicit\ntime-stepping scheme [39], and spatial derivatives are computed using Chebyshev collocation [40]. Conservation of\nenergy is ensured by verifying that _E=Wf\u0000Q;where\nE=1\n2Z1\n0 \nj_rj2\nM\u0003+\u00142\nM\u0003U\u00032!\nds; (9)\nQ=1\nM\u0003U\u00032Z1\n0\u0018(s) _\u00142ds; Wf=Z1\n0f\u0001_rds; (10)\nare respectively the mechanical energy of the system, the dissipated power and the rate of work of the \ruid forces,\nand is classically obtained by projecting the equation of motion (1) on the solid velocity _rand by integrating over\nthe entire beam. Also note from (7), P=hQi. In the numerical implementation the beam and the \row are initially\nat rest; the \row speed is ramped up to its steady state value and a small perturbation is applied to the vertical \row.\nE. Nonlinear response of an undamped beam\nPrior to analysing the energy harvesting properties of the system, we \frst examine the undamped \rutter response\nof the structure. In Figure 2, we plot the system response for increasing nondimensional \row speed, for a circular5\nFIG. 3: Variation of the critical \rutter speed, Ucfwith damping \u00180forM\u0003= 2:5 (dotted), 5 :1 (dash-dotted), 8 :5 (dashed), 12 :7\n(solid) and 20 (thick solid) obtained from linear stability analysis for uniform damping. Results of nonlinear computations for\nM\u0003= 12:7 are also presented (squares).\ncylinder (M\u0003\u001912:7). The critical \row speed at which \rutter ensues is veri\fed from linear stability analysis and\nis con\frmed with [26]. Consistent with \rutter in plates [33], we note that this instability is a supercritical Hopf\nbifurcation with \row speed. The jumps in the bifurcation curve correspond to the mode switching reported in [41],\nthis may also be discerned from the snapshots of the beam at three di\u000berent \row speeds. For the energy harvesting\ncomputations performed in the rest of the paper we set the \row speed at U\u0003= 13, corresponding to well-developed\noscillations and moderate de\rections for the undamped con\fguration.\nIII. UNIFORM DAMPING\nHere, a uniform damping distribution \u0018(s) =\u00180is considered. The dependence of critical \rutter speed on the\ndamping intensity \u00180is \frst analysed for varying values of \ruid loading M\u0003. Based on these results, a moderate \ruid\nloading is selected to study the power harvesting capacity of the con\fguration.\nA. Destabilisation by damping: impact on critical \rutter speed\nAs noted in Section I, damping can destabilise \rexible structures at su\u000eciently high \ruid loading. [32] shows that\ndestabilisation of long pipes with a high mass ratio is associated with a drop in the critical \rutter speed. Based on\nthis work, we analyse the linear stability of the system and in Figure 3 compare the variation of the critical \rutter\nspeed,Ucf, with damping, \u00180, at di\u000berent values of \ruid loading, M\u0003. For lightly loaded structures ( M\u0003<5:5)Ucf\nmonotonically increases with damping; at higher values ( M\u0003\u00158:5),Ucfdecreases with damping until a speci\fc value\n(\u0018m) above which Ucfincreases rapidly; note that \u0018mincreases with M\u0003.\nFor the remainder of the energy harvesting analysis, we settle on a value of M\u0003= 12:7 (corresponding to D=L =\n0:1): this choice allows us to analyse the scope for harvesting energy from a system at moderate \ruid loading with\ndestabilisation by damping.\nB. Harvesting power with constant damping\nFigure 4(a) presents the evolution of harvested power, P, with damping for 10\u00003<\u00180<10. Equation (7) becomes\nP=1\nM\u0003U\u00032\u00180kKk 1; (11)6\nFIG. 4: (a) Evolution of the harvested power P(solid) and curvature norm k\u0016Kk1(dashed) with the uniform damping intensity\n\u00180. (b) Distribution of curvature-change \u0016K(s) along the beam for \u00180= 0, 0:022, 0:22 and 0:47 (from left to right and top to\nbottom, respectively). ( M\u0003= 12:7; U\u0003= 13).\nwhereK(s) =h_\u00142iandkKk 1=R1\n0Kdsis theL1-norm ofK(s). The evolution of the \rutter response with damping\nis examined by plotting the rescaled curvature term, \u0016K=K=kK0k1(whereK0(s) is the value ofKfor\u00180= 0) for\nincreasing\u00180(Figure 4b). At small \u00180, the response is virtually no di\u000berent from the undamped case, but for \u00180>0:1\na perceptible change is observed. First, the curvature is redistributed along the entire beam, and as the damping is\nincreased further, the response of the beam is damped out globally. This can also be seen from the variation of kKk 1\nwith\u00180in Figure 4(a). Notably from (11) it is kKk 1that directly impacts the harvested power.\nAs a result, for small damping, the change in system response is virtually imperceptible from the undamped case,\nand power simply scales linearly with \u00180. However, damping has a strong e\u000bect on the \rutter response at larger \u00180: it\nreduces the amplitude of curvature-change signi\fcantly and causes a sharp reduction in P. The strategy to optimally\nharvest power is to \fnd the upper bound on \u00180below whichkKk 1can be maintained close to (or ideally enhanced\nabove) its undamped value.\nIV. NONUNIFORM DAMPING\nThe results for constant damping suggest that as long as kKk 1is maintained at undamped levels, the harvested\npower increases linearly with \u00180. Figure 4(b) clearly shows that K(s) varies signi\fcantly along the length of the beam.\nConcentrating harvesters around the zone of peak curvature could therefore enhance the harvested power. This idea\nis tested in this section on a reduced functional space for the damping distribution \u0018(s). Two families of damping\nfunctions are considered, namely (a) dispersed and (b) focused damping distributions. Optimisation of the harvested\npower is performed within each family, in anticipation of insights into the global optimal.\nA. Dispersed harvester distribution\nIn this section, we are interested in simple non-homogeneous distributions of damping of the form:\n\u0018(s) =\u00180\u0000\n1 +\u00181(s\u00001=2)\u0001\n(12)\ncharacterised by the total damping, 10\u00003< \u00180<10, and slope,\u00002\u0014\u00181\u00142. This function family corresponds to\na dispersed distribution, where the damping is signi\fcant over the entire length of the structure. Figure 5(a) shows\nthe variation of rescaled power, \u0016P=P=Pmax\nhwith (\u00180;\u00181), wherePmax\nhis the maximum harvested power for constant\ndamping (Section III). Figure 5 shows that for all \u00180, the optimal distribution corresponds to \u00181= 2, when damping\nis distributed increasingly from \fxed to free end, and that this linear distribution of damping leads to an increase7\nFIG. 5: Linear distribution (12): (a) rescaled power contours P=Pmax\nhfor varying \u00180and\u00181. (b) Rescaled harvested power\nP=Pmax\nhand (c)k\u0016Kk1as a function of the damping intensity \u00180for linearly decreasing ( \u00181=\u00002, dashed), constant ( \u00181= 0,\ndotted) and linearly increasing ( \u00181= 2, solid) damping distributions. ( M\u0003= 12:7; U\u0003= 13).\nof the maximum harvested power by 50% compared to the uniform distribution ( \u00181= 0). Through non-uniform but\nsimple distributions of damping on the structure, it is indeed possible to enhance the \rutter response above that of\nthe undamped con\fguration (Figure 5c).\nB. Focused harvester distribution\nThe results from the reduced order analysis [29] suggest that the optimal distribution ought to be localised at\nspeci\fc points on the beam. The chief drawback of the two parameter functions examined in the previous section is\nthe inability to consider localised distributions of damping in speci\fc regions. To consider such peaked distributions,\nwe now turn to the following three-parameter family of Gaussian damping distribution:\n\u0018(s) =\u00180\u0018g\nk\u0018gk1; \u0018g= e\u0000\u000b(s\u0000so)2; (13)\nwhere\u00180is the total damping in the system, sois the centre of the distribution and 1 =\u000bis a measure of its spread in\ns. For increasing \u000benergy harvesters are increasingly concentrated around so.\nFigure 6 presents the rescaled power P=Pmax\nhin the (\u000b;so)-space for small ( \u00180= 0:01) and moderate damping\n(\u00180= 0:47). For small damping, one sees that power is optimally harvested for dampers focused at so= 0:8, the\nposition of the maximum of K0along the beam (Figure 4b). This is quite distinct from the high damping optimal,\nwhich corresponds to a dispersed distribution ( \u000b= 2:7; so= 1). Of particular note is that for all \u00180, a uniform\ndistribution ( \u000b= 0) is preferred over a concentration of harvesters at so<0:5.\nWe next examine a wider range of damping, 10\u00003<\u00180<50, and plot the optimal values of ( \u000b;so), in Figure 7(a).\nFor moderate to large damping, \u00180>0:3, a dispersed distribution with peak at so= 1 is optimal. Conversely for\n\u00180<0:02, harvesters concentrated at so\u00190:8 (position of peak K) is the optimal con\fguration. Note that for these\ncomputations we set 0 <\u000b< 500, and at these values of \u00180the upper bound is reached; more detailed insights and\ncomputations on the small damping regime are presented in Section V. Figure 7(b) shows the evolution with \u00180of\nthe optimal normalised distribution and illustrates the transition from focused to dispersed pro\fles when the total\ndamping is increased.8\n(a)\n (b)\nFIG. 6: Gaussian distribution (13): Maps of rescaled power P=Pmax\nhfor varying \u000bandsoat (a)\u00180= 0:01 and (b) \u00180= 0:47.\n(M\u0003= 12:7; U\u0003= 13).\n(a)\n (b)\nFIG. 7: Gaussian function optimal con\fguration: (a) Spread parameter, \u000b, (open circles) and centre location, so, (\flled\nboxes) corresponding to the optimal Gaussian distribution for a given \u00180. (b) Rescaled damping distribution at discrete\n\u00180= 0:004;0:025;0:25;0:47 (solid, dashed, dashed-dot and dotted curves, respectively, indicated on (a) with vertical lines of\ncorresponding description); superimposed is the linear optimal distribution ( \u00181= 2; thick solid curve). ( M\u0003= 12:7; U\u0003= 13).\nIt is possible to determine the optimal harvested power for each value of \u00180; in Figure 8 these optimals are compared\nwith the results for uniform and linear distributions. Strikingly, the peak power values are coincident for linear\nand Gaussian distributions, which is consistent with the similarity in distributions (Figure 7(b)). Because of the\nfundamentally di\u000berent structures of the distribution used, this result suggests that the optimal obtained with such\nsimple functions represents a good approximation of the absolute optimal. Departing from this maximum of harvested\npower, in particular at small damping we see that concentrated harvesting is superior to the optimal linear distribution.\nThese results con\frm that nonuniform damping distributions can be advantageously employed to enhance the power\nharvesting capacity. Focused damping distributions are optimal for small damping, whereas dispersed distributions\nare preferential at higher damping. The following section investigates in more details this transition by considering\nthe impact of damping on the nonlinear dynamical response of the system.9\nFIG. 8: Optimal power values corresponding to Gaussian optimal con\fguration in Figure 7 (squares) compared to the optimal\npower obtained for linear distribution ( \u00181= 2; thick curve) and constant distribution ( \u00181= 0; dotted curve).\nV. DISCUSSION: IMPACT OF LOCALISED DAMPING\nThe above computations suggest a very di\u000berent impact of damping on the dynamics of the structure depending\non the magnitude of \u00180, thereby leading to quite di\u000berent optimal strategies to maximise the harvested power. In this\nsection, we \frst consider the limit of asymptotically small damping before studying the e\u000bect of focused dampers on\nthe body's dynamics.\nA. Optimal distribution for asymptotically small damping\nFor asymptotically small damping ( k\u0018k1\u001c1, withk\u0018k1the maximum value of \u0018(s) fors2[0 1]), the \rapping\ndynamics is not modi\fed at leading order so that K=K0+O(k\u0018k1). Thus\nP\u00181\nM\u0003U\u00032Z1\n0\u0018(s)K0ds\u0014\u00180\nM\u0003U\u00032kK0k1 (14)\nand this upper bound can be approached asymptotically using, for example, the Gaussian distribution (13) centred\non the maximum of K0and increasing \u000b. More precisely, any peaked distribution of damping of L1-norm\u00180centred\non this maximum and with a typical width (here \u000b\u00001=2) much smaller than the length scale \u0015associated with the K0\npeak width, should approach the theoretical upper bound Pth=\u00180kK0k1=(M\u0003U\u00032), provided that the maximum\ndamping remains small enough that the approximation for Pin Eq. (14) still holds (which in the present case is\nequivalent to keeping \u000b1=2\u00180bounded).\nThis conjecture is tested numerically using a Gaussian distribution, Eq. (13), with so= 0:8, and by computing the\nharvested power for small damping: 2 :2\u000210\u00005< \u00180<2:2\u000210\u00002. Figure 9 shows P=Pthand con\frms that this\nconjecture indeed holds since for increasing \u000b, the power asymptotically approaches its optimal for small enough \u00180.\nWhen\u00180\u00152:2\u000210\u00003, the asymptotic value Pthcan not be reached any more: as \u000bis increased, the e\u000bect of the\nmaximum damping \u000b\u00180on the \rutter dynamics is important even before the dampers are focused enough for Pto\napproachPth.\nTwo essential results are illustrated here. For su\u000eciently small damping, the nonlinear response of the structure\nremains unchanged so the best strategy is to focus all the harvesters in the region of maximum curvature-change.\nHowever \fnite levels of damping signi\fcantly modify the nonlinear dynamics of the beam, so at these levels a narrow\nfocusing of damping is sub-optimal; next we investigate further the behaviour at \fnite damping.10\nFIG. 9: Focused harvesters at so= 0:81 for small damping: Curves show the normalised power for increasingly focused\nharvesters for small damping values: \u00180= 2:2\u000210\u00005: 10\u00002(solid, dashed, dash-dot, dotted curves, respectively). ( M\u0003=\n12:7; U\u0003= 13).\nB. Impact of \fnite and localised damping\nThe computations from Section IV show that for Gaussian distributions and \u00180>0:02 a defocusing of harvesters is\npreferable, as the system response becomes increasingly in\ruenced by damping. In order to understand the impact on\nthe system's response, the Gaussian distribution (13) is used in the high damping range and \u000bis varied over a range\nthat transitions the distribution from dispersed to focused (0 <\u000b< 102) with\u00180= 0:47 andso= 0:45 (corresponding\nto the position of kKk1for\u000b= 0). Consistent with Figure 6, the harvested power in Figure 10(a) decreases with\n\u000b, and Figure 10(b) shows the evolution of K(s).\nFor dispersed distributions ( \u000b<10),K(so) decreases with \u000band the zone of maximum curvature-change shifts to\nregions of reduced damping. Nonetheless nonzero damping in this zone is adequate to retrieve some of the energy.\nFor\u000b>10 damping becomes increasingly focused and the \rexible body does not deform anymore near the damper\n(K(s0)\u00190): no energy is harvested anymore since no deformation occurs near the harvester's position, and this is\nre\rected in the sharp drop in power in Figure 10(b). This illustrates the main e\u000bect of a focused distribution for\n\fnite damping: we see a redistribution of the solid's deformation to regions with little or no damping. This is also\nthe reason why a focused damping distribution is not appropriate for optimal energy harvesting in the \fnite damping\nrange.\nVI. CONCLUSIONS\nIn this work, we considered the possibility of harvesting energy from a slender body \ruttering in an axial \row, and\nin particular the impact of harvester-distribution on the performance of the system as well as potential optimisation\nstrategies. To this end, a simpli\fed \ruid-solid model was proposed with energy harvesting represented as nonuniform\nstructural damping. Depending on the \ruid-to-solid inertia ratio, damping can actually enhance the \rutter response\nof the structure, as well as reduce the critical \row velocity above which the system can operate.\nFor uniform damping distributions, maximising the harvested power appears as a trade-o\u000b on the damping intensity:\nfor small damping, the \rutter response is only weakly modi\fed and the harvested power increases as more damping\nis added to the system. For higher damping however, the dynamical response can be strongly modi\fed and the self-\nsustained oscillations are eventually mitigated. The e\u000bect of a nonuniform distribution of harvesters along the structure\nwas considered next. We show that even simple nonuniform distributions such as linear and Gaussian functions, can\nlead to an increase in harvested peak power on the order of 50%. The similarity in the optimal distribution and\nperformance obtained through an optimisation on two fundamentally di\u000berent families of distributions suggests that\nsimple strategies can capture rather well the characteristics of the global optimal distribution.\nInvestigating further into the relationship between damping and \rutter response, we showed that for small damping,\nlocalised harvesting is optimal as it takes full advantage of the system's maximum curvature without impacting its11\n(a)\n (b)\nFIG. 10: Focused harvesters at so= 0:45 for large damping ( \u00180= 0:47): (a) power dependence on \u000band (b) \u0016K(s) for\n\u000b= 0;1;10;102(solid, dashed, dashed-dot, dotted curves, respectively). ( M\u0003= 12:7; U\u0003= 13).\ndynamics signi\fcantly. On the other hand, for \fnite damping, focused distributions perform rather poorly as the\nbeam response adapts to rigidify the damped region, leading to negligible harvested power. Instead, nonuniform\ndistributed damping over the entire length of the system becomes optimal.\nReturning to the result from the simple bi-articulated model [29], we note a clear di\u000berence in the optimal con-\n\fgurations. Whilst the reduced order model optimal has dampers focused at the \fxed end, the continuous optimal\nrequires a dispersed distribution with minimal damping at the \fxed end and increasing to a maximum at the free\nend. However, we \fnd a crucial di\u000berence between the two system con\fgurations: whilst the bi-articulated system\nhas a single moving region of curvature (the second articulation) that is responsible for driving the instability, defor-\nmations may occur all along the length of the beam in the continuous system. Therefore in the latter, curvature can\nbe displaced away from regions with focused damping while still maintaining the \rutter dynamics. In both cases, a\ncareful understanding of the dynamics of the system is necessary to determine the optimal nonuniform positioning of\nenergy harvesters.\nVII. ACKNOWLEDGEMENTS\nThe authors gratefully acknowledge the support of Electricit\u0013 e de France (EDF) for their support through the \\Chaire\nEnergies Durables\" at Ecole Polytechnique. S. M. was also supported by a Marie Curie International Reintegration\nGrant within the 7thEuropean Community Framework Program.\n[1] Adam Westwood. Ocean power: Wave and tidal energy review. Refocus , 5:50 { 55, 2004.\n[2] A.O.F Falcao. Wave energy utilization: A review of the technologies. Renew. Sust. Energ. Rev. , 10:899{918, 2010.\n[3] S.R. Anton and H.A. Sodano. A review of power harvesting using piezoelectric materials (2003-2006). Smart Mater.\nStruct. , 16:1{21, 2007.\n[4] J. A. Paradiso and T. Starner. Energy scavenging for mobile and wireless electronics. IEEE Pervas. Comput. , 4:18{27,\n2005.\n[5] A. Khaligh, P. Zeng, and C. Zheng. Kinetic energy harvesting using piezoelectric and electromagnetic technologies-state\nof the art. IEEE Trans. Ind. Electron , 57:850{860, 2010.\n[6] M. M. Bernitsas, K. Raghavan, Y. Ben-Simon, and E. M. H. Garcia. Vivace (Vortex Induced Vibration Aquatic Clean\nEnergy): A new concept in generation of clean and renewable energy from \ruid \row. J. O\u000bshore Mech. Arct. , 130:041101,\n2008.\n[7] C. Grouthier, S. Michelin, and E. de Langre. Energy harvesting using vortex-induced vibrations of tensioned cables.\npre-print, arXiv:1203.0236v1, 2012.\n[8] Z. Peng and Q. Zhu. Energy harvesting through \row-induced oscillations of a foil. Phys. Fluids , 21:123602, 2009.12\n[9] A Erturk, W G R Vieira, C. De Marqui, and D. J. Inman. On the energy harvesting potential of piezoaeroelastic systems.\nAppl. Phys. Letters , 96:184103, 2010.\n[10] L. Tang, M.P. Pa \u0010doussis, and J. Jiang. Cantilevered \rexible plates in axial \row: Energy transfer and the concept of\n\rutter-mill. Journal of Sound and Vibration , 326:263{276, 2009.\n[11] O Doar\u0013 e and S. Michelin. Piezoelectric enegy harvesting from \rutter instability: Local/global linear stability and e\u000eciency.\nJ. Fluids Struct. , 27:1357{1375, 2011.\n[12] M Bryant and E. Garcia. Modeling and testing of a novel aeroelastic \rutter energy harvester. J. Vib. Acoust. , 133:011012,\n2011.\n[13] S. Li, J. Yuan, and H. Lipson. Ambient wind energy harvesting using cross-\row \ruttering. J. Appl. Phys. , 109:026104,\n2011.\n[14] M. P. Pa \u0010doussis. Fluid-Structure Interactions: Slender structures and axial Flow, Volume 2 . London: Elsevier Academic\nPress., 2004.\n[15] M. P. Pa \u0010doussis. Fluid-Structure Interactions: Slender structures and axial \row, Volume 1 . London: Academic Press.,\n1998.\n[16] P.G. Larose and J.B. Grotberg. Flutter and long-wave instabilities in compliant channels conveying developing \rows. J.\nFluid. Mech. , 331:37{58, 1997.\n[17] P. W. Carpenter and T. J. Pedley, editors. Flow past highly compliant boundaries and in collapsible tubes . Kluwer, 2003.\n[18] O Doar\u0013 e and E. de Langre. The role of boundary conditions in the instability of one-dimensional systems. Eur. J. Mech.\nB/Fluids , 569:151{184, 2006.\n[19] P. S. Stewart, S. L. Waters, and O.E. Jensen. Local and global instabilities of \row in a \rexible-walled channel. Eur. J.\nMech. B/Fluids , 28:541{557, 2009.\n[20] S. Alben and M. Shelley. Flapping states of a \rag in an inviscid \ruid: bistability and the transition to chaos. Phys. Rev.\nLett., 100:074301, 2008.\n[21] S. Michelin, S. G Llewellyn Smith, and B. Glover. Vortex shedding model of a \rapping \rag. J. Fluid Mech. , 617:1{10,\n2008.\n[22] C. Eloy, R. Lagrange, C. Souilliez, and L. Schouveiler. Aeroelastic instability of a \rexible plate in a uniform \row. J. Fluid\nMech. , 611:97{106, 2008.\n[23] D. G. Crighton and J. E. Oswell. Fluid loading with mean \row. (I). Response of an elastic plate to localized excitation.\nPhilos. Trans. R. Soc. London, A , 335:557{592, 1991.\n[24] R. M. Howell, A. D. Lucey, P. W. Carpenter, and M. W. Pitman. Interaction between a cantilevered-free \rexible plate\nand ideal \row. J. Fluids Struct. , 25:544{566, 2009.\n[25] C. Eloy, C. Souilliez, and L. Schouveiler. Flutter of a rectangular plate. J. Fluids Struct. , 23:904{919, 2007.\n[26] M. P. Pa \u0010doussis, E Grinevich, D. Adamovic, and C. Semler. Linear and nonlinear dynamics of cantilevered cylinders in\naxial \row. Part 1. Physical dynamics. J. Fluids Struct. , 16:691{713, 2002.\n[27] E. de Langre, O. Doar\u0013 e, M. P Pa \u0010doussis, and Y. Modarres-Sadeghi. Flutter of long \rexible cylinders in axial \row. J.\nFluid Mech. , 571:371{389, 2007.\n[28] M.J. Shelley and J. Zhang. Flapping and bending bodies interacting with \ruid \rows. Ann. Rev. Fluid Mech. , 43:449{465,\n2011.\n[29] K Singh, S. Michelin, and E. de Langre. Energy harvesting from \ruid-elastic instabilities of a cylinder. J. Fluids Struct. ,\n30:159{172, 2012.\n[30] M. Gad-el Hak. Drag reduction using compliant walls. In P.W. Carpenter and T.J. Pedley, editors, Proceedings of the\nIUTAM Symposium, 2001: Flow past highly compliant boundaries and in collapsible tubes , 2003.\n[31] N. Peake. Nonlinear stability of a \ruid-loaded elastic plate with mean \row. J. Fluid Mech. , 434:101{118, 2001.\n[32] O Doar\u0013 e. Dissipation e\u000bect on local and global stability of \ruid-conveying pipes. J. Sound Vib. , 329:72{83, 2010.\n[33] C. Eloy, N. Kofman, and L. Schouveiler. The origin of hysteresis in the \rag instability. J. Fluid Mech. , 691:583{593, 2012.\n[34] M. J. Lighthill. Large-amplitude elongated-body theory of \fsh locomotion. Proc. R. Soc. B , 179:125{138, 1971.\n[35] F. Candelier, F. Boyer, and A Leroyer. Three-dimensional extension of Lighthill's large-amplitude elongated-body theory\nof \fsh locomotion. J. Fluid Mech. , 674:196{226, 2011.\n[36] G. I. Taylor. Analysis of the swimming of long and narrow animals. Proc. R. Soc. A , 214:158{183, 1952.\n[37] F Boyer, M Porez, A Leroyer, and M Visonneau. Fast dynamics of an eel-like robot- comparisons with Navier-Stokes\nsimulations. IEEE Transactions on Robotics , 24:1274{1288, 2008.\n[38] K. Singh and T.J. Pedley. Modelling lateral manoeuvres in \fsh. J. Fluid Mech. , 697:1{34, 2012.\n[39] S. Alben. Simulating the dynamics of \rexible bodies and vortex sheets. J. Comp. Phys. , 228:2587{2603, 2009.\n[40] J.P. Boyd. Chebyshev and Fourier Spectral Methods . Dover Publications New York, 2001.\n[41] C. Semler, J. L. Lopes, N. Augu, and M.P. Pa \u0010doussis. Linear and nonlinear dynamics of cantilevered cylinders in axial\n\row, Part 3: Nonlinear dynamics. J. Fluids Struct. , 16:739{759, 2002." }, { "title": "2312.09140v1.Nonlocal_damping_of_spin_waves_in_a_magnetic_insulator_induced_by_normal__heavy__or_altermagnetic_metallic_overlayer__a_Schwinger_Keldysh_field_theory_approach.pdf", "content": "Nonlocal damping of spin waves in a magnetic insulator induced by normal, heavy, or\naltermagnetic metallic overlayer: A Schwinger-Keldysh field theory approach\nFelipe Reyes-Osorio and Branislav K. Nikoli´ c∗\nDepartment of Physics and Astronomy, University of Delaware, Newark, DE 19716, USA\n(Dated: December 15, 2023)\nUnderstanding spin wave (SW) damping, and how to control it to the point of being able to\namplify SW-mediated signals, is one of the key requirements to bring the envisaged magnonic tech-\nnologies to fruition. Even widely used magnetic insulators with low magnetization damping in their\nbulk, such as yttrium iron garnet, exhibit 100-fold increase in SW damping due to inevitable con-\ntact with metallic layers in magnonic circuits, as observed in very recent experiments [I. Bertelli\net al. , Adv. Quantum Technol. 4, 2100094 (2021)] mapping SW damping in spatially-resolved\nfashion. Here, we provide microscopic and rigorous understanding of wavevector-dependent SW\ndamping using extended Landau-Lifshitz-Gilbert equation with nonlocal damping tensor , instead\nof conventional local scalar Gilbert damping, as derived from Schwinger-Keldysh nonequilibrium\nquantum field theory. In this picture, the origin of nonlocal magnetization damping and thereby in-\nduced wavevector-dependent SW damping is interaction of localized magnetic moments of magnetic\ninsulator with conduction electrons from the examined three different types of metallic overlayers—\nnormal, heavy, and altermagnetic. Due to spin-split energy-momentum dispersion of conduction\nelectrons in the latter two cases, the nonlocal damping is anisotropic in spin and space, and it can\nbe dramatically reduced by changing the relative orientation of the two layers when compared to\nthe usage of normal metal overlayer.\nIntroduction. —Spin wave (SW) or magnon damping\nis a problem of great interest to both basic and ap-\nplied research. For basic research, its measurements [1–4]\ncan reveal microscopic details of boson-boson or boson-\nfermion quasiparticle interactions in solids, such as:\nmagnon-magnon interactions (as described by second-\nquantized Hamiltonians containing products of three or\nmore bosonic operators [5, 6]), which are frequently en-\ncountered in antiferromagnets [4, 5] and quantum spin\nliquids [7], wherein they play a much more important\nrole [8] than boson-boson interactions in other condensed\nphases, like anharmonic crystalline lattices or superflu-\nids [5]; magnon-phonon interactions [3], especially rel-\nevant for recently discovered two-dimensional magnetic\nmaterials [2]; and magnon-electron interactions in mag-\nnetic metals [1, 9–12]. For the envisaged magnon-\nbased digital and analog computing technologies [13–\n17], understanding magnon damping makes it possible\nto develop schemes to suppress [18] it, and, further-\nmore, achieve amplification of nonequilibrium fluxes of\nmagnons [19–22]. In fact, overcoming damping and\nachieving amplification is the keyto enable complex\nmagnon circuits where, e.g., a logic gate output must\nbe able to drive the input of multiple follow-up gates.\nLet us recall that the concept of SW was introduced by\nBloch [23] as a wave-like disturbance in the local mag-\nnetic ordering of a magnetic material. The quanta [6] of\nenergy of SWs of frequency ωbehave as quasiparticles\ntermed magnons, each of which carries energy ℏωand\nspin ℏ. As regards terminology, we note that in magnon-\nics [13] SW is often used for excitations driven by an-\ntennas [24–27] and/or described by the classical Landau-\nLifshitz-Gilbert (LLG) equation [9, 10, 28, 29], whereas\nmagnon is used for the quantized version of the same ex-\ne\nee\nee\neFIG. 1. (a) Schematic view of bilayers where a metallic over-\nlayer covers the top surface of magnetic insulator, as often\nencountered in spintronics and magnonics [13, 30]. Three\ndifferent energy-momentum dispersion of conduction elec-\ntrons at the interface are considered, with their Fermi sur-\nfaces shown in panel (b)—normal metal (NM); heavy metal\n(HM) with the Rashba SOC [31, 32], and altermagnetic metal\n(AM) [33, 34]—with the latter two being spin-split. The rel-\native alignment of the layers is labeled by an angle θ[33, 34],\nmeaning that the wavevector qof SWs within FI is at an an-\ngleθaway from the kx-axis.\ncitation [5], or these two terms are used interchangingly.\nIn particular, experiments focused on SW damp-\ning in metallic ferromagnets have observed [1] its de-\npendence on the wavevector qwhich cannot be ex-\nplained by using the standard LLG equation [28,\n29],∂tMn=−Mn×Beff\nn+αGMn×∂tMn(where ∂t≡\n∂/∂t), describing dynamics of localized magnetic mo-\nments (LMMs) Mnat site nof crystalline lattice (also\nused in atomistic spin dynamics [28]) viewed as classi-\ncal vectors of unit length. This is because αG, as the\nGilbert damping parameter [35, 36], is a local scalar (i.e.,\nposition-independent constant). Instead, various forms\nof spatially nonuniform (i.e., coordinate-dependent) and\nnonlocal (i.e., magnetization-texture-dependent) damp-\ning due to conduction electrons have been proposed [9,arXiv:2312.09140v1 [cond-mat.mes-hall] 14 Dec 20232\n10, 37–39], or extracted from first-principles calcula-\ntions [40], to account for observed wavevector-dependent\ndamping of SWs, such as ∝q2(q=|q|) measured in\nRef. [1]. The nonlocal damping terms require neither\nspin-orbit coupling (SOC) nor magnetic disorder scatter-\ning, in contrast to αGwhich is considered to vanish [41]\nin their absence.\nThus, in magnonics, it has been considered [30] that\nusage of magnetic insulators, such as yttrium iron gar-\nnet (YIG) exhibiting ultralow αG≃10−4(achieved on\na proper substrate [42]), is critical to evade much larger\nand/or nonlocal damping of SWs found in ferromagnetic\nmetals. However, very recent experiments [24–27] have\nobserved 100-fold increase of SW damping in the segment\nof YIG thin film that was covered by a metallic overlayer.\nSuch spatially-resolved measurement [24] of SW damp-\ning was made possible by the advent of quantum sensing\nbased on nitrogen vacancy (NV) centers in diamond [43],\nand it was also subsequently confirmed by other meth-\nods [25–27]. Since excitation, control, and detection of\nSWs requires to couple YIG to metallic electrodes [13],\nunderstanding the origin and means to control/suppress\nlarge increase in SW damping underneath metallic over-\nlayer is crucial for realizing magnonic technologies. To\nexplain their experiments, Refs. [24–27] have employed\nthe LLG equation with ad hoc intuitively-justified terms\n(such as, effective magnetic field due to SW induced eddy\ncurrents within metallic overlayer [24]) that can fit the\nexperimental data, which is nonuniversal and unsatisfac-\ntory (many other examples of similar phenomenological\nstrategy exist [1, 44]).\nIn contrast, in this Letter we employ recently derived\n∂tMn=−Mn×Beff\nn+Mn×X\nn′(αGδnn′+λR)·∂tMn′,\n(1)\nextended LLG equation with all terms obtained [45]\nmicroscopically from Schwinger-Keldysh nonequilibrium\nquantum field theory [46] and confirmed [45] via exact\nquantum-classical numerics [47–50]. It includes nonlo-\ncal damping as the third term on the right-hand side\n(RHS), where its nonlocality is signified by dependence\nonR=rn−rn′, where rnis the position vector of lat-\ntice site n. Equation (1) is applied to a setup depicted in\nFig. 1 where conduction electron spins from three differ-\nent choices for metallic overlayer are assumed to interact\nwith LMMs of ferromagnetic insulator (FI) at the inter-\nface via sdexchange interaction of strength Jsd, as well\nas possibly underneath the top surface of FI because of\nelectronic evanescent wavefunction penetrating into it.\nNote that FI/normal metal (NM) bilayer directly mod-\nels recent experiments [24] where FI was a thin film of\nYIG and NM was Au, and SW damping within FI was\nquantified using quantum magnetometry via NV centers\nin diamond. Next, the FI/heavy metal (HM) bilayer,\nsuch as YIG/Pt [18, 27], is frequently encountered in\n0 2 4\nK/J0.51.0q (1/a)\n kF= 0\n.92kF= 0\n.99kF= 1\n.08kF= 1\n.15kF= 1\n.22kF= 1\n.30kF= 1\n.38(a)\n1.0 1.2 1.4\nkF0.81.01.21.4qmax(1/a)\n∝kF(b)FIG. 2. (a) Wavevector qof SW generated by injecting spin-\npolarized current in TDNEGF+LLG simulations of NM over-\nlayer on the top of 1D FI [Fig. 1(a)] as a function of anisotropy\nK[Eq. (3)] for different electronic Fermi wavevectors kF. (b)\nMaximum wavevector qmaxof SWs that can be generated by\ncurrent injection [21, 57] before wavevector-dependent SW\ndamping becomes operative, as signified by the drop around\nkFin curves plotted in panel (a).\nvarious spintronics and magnonics phenomena [13, 30].\nFinally, due to recent explosion of interest in altermag-\nnets [33, 34], the FI/altermagnetic metal (AM) bilay-\ners, such as YIG/RuO 2, have been explored experimen-\ntally to characterize RuO 2as a spin-to-charge conver-\nsion medium [51]. The Schwinger-Keldysh field theory\n(SKFT), commonly used in high energy physics and cos-\nmology [52–54], allows one to “integrate out” unobserved\ndegrees of freedom, such as the conduction electrons in\nthe setup of Fig. 1, leaving behind a time-retarded dis-\nsipation kernel [48, 55, 56] that encompasses electronic\neffects on the remaining degrees of freedom. This ap-\nproach then rigorously yields the effective equation for\nLMMs only , such as Eq. (1) [45, 56] which bypasses the\nneed for adding [1, 24, 44] phenomenological wavevector-\ndependent terms into the standard LLG equation. In\nour approach, the nonlocal damping is extracted from\nthe time-retarded dissipation kernel [45].\nSKFT-based theory of SW damping in FI/metal bilay-\ners.—The nonlocal damping [45] λRin the third term\non the RHS of extended LLG Eq. (1) stems from back-\naction of conduction electrons responding nonadiabati-\ncally [48, 59]—i.e., with electronic spin expectation value\n⟨ˆsn⟩being always somewhat behind LMM which gener-\nates spin torque [60] ∝ ⟨ˆsn⟩×Mn—to dynamics of LMMs.\nIt is, in general, a nearly symmetric 3 ×3 tensor whose\ncomponents are given by [45]\nλαβ\nR=−J2\nsd\n2πZ\ndε∂f\n∂εTr\u0002\nσαAnn′σβAn′n\u0003\n. (2)\nHere, f(ε) is the Fermi function; α, β =x, y, z ;σα\nis the Pauli matrix; and A(ε) = i\u0002\nGR(ε)−GA(ε)\u0003\nis\nthe spectral function in the position representation ob-\ntained from the retarded/advanced Green’s functions\n(GFs) GR/A(ε) =\u0000\nε−H±iη\u0001−1. Thus, the calcula-\ntion of λRrequires only an electronic Hamiltonian H\nas input, which makes theory fully microscopic (i.e.,3\n−5 0 5\nX−505Y\nλNM\nR(a)NM\n−1 0 1\nλα\nR\n−5 0 5\nX−505Y\nλxx\nR(b)HM t SOC= 0.3t0\n−5 0 5\nX−505Y\nλzz\nR(c)\n−5 0 5\nX−505Y\nλ⊥\nR(d)AM t AM= 0.5t0\n0.0 0.5 1.0 1.5\nq (1/a)0123Γq≡Im(ωq) (J/¯ h)×10−1\nEq.(6)(e)\nη= 0.1\nη= 0\n−5 0 5\nX−505Y\nλyy\nR(f)\n−5 0 5\nX−505Y\nλxy\nR(g)\n−5 0 5\nX−505Y\nλ/bardbl\nR(h)\nFIG. 3. (a)–(d) and (f)–(h) Elements of SKFT-derived nonlocal damping tensor in 2D FI, λRwhere R= (X, Y, Z ) is the\nrelative vector between two sites within FI, covered by NM [Eq. (5)], HM [Eqs. (8)] or AM [Eqs. (9)] metallic overlayer. (e)\nWavevector-dependent damping Γ qof SWs due to NM overlayer, where the gray line is based on Eq. (6) in the continuous\nlimit [58] and the other two lines are numerical solutions of extended LLG Eq. (1) for discrete lattices of LMMs within FI. The\ndotted line in (e) is obtained in the absence of nonlocal damping ( η= 0), which is flat at small q.\nHamiltonian-based). Although the SKFT-based deriva-\ntion [45] yields an additional antisymmetric term, not\ndisplayed in Eq. (2), such term vanishes if the system\nhas inversion symmetry. Even when this symmetry is\nbroken, like in the presence of SOC, the antisymmet-\nric component is often orders of magnitude smaller [56],\ntherefore, we neglect it. The first term on the RHS of ex-\ntended LLG Eq. (1) is the usual one [28, 29], describing\nprecession of LMMs in the effective magnetic field, Beff\nn,\nwhich is the sum of both internal and external ( Bextez)\nfields. It is obtained as Beff\nn=−∂H/∂Mnwhere His\nthe classical Hamiltonian of LMMs\nH=−JX\n⟨nn′⟩Mn·Mn′+K\n2X\nn(Mz\nn)2−BextX\nnMz\nn.(3)\nHere we use g= 1 for gyromagnetic ratio, which sim-\nplifies Eq. (1); Jis the Heisenberg exchange coupling\nbetween the nearest-neighbors (NN) sites; and Kis the\nmagnetic anisotropy.\nWhen nonlocal damping tensor, λRis proportional\nto 3×3 identity matrix, I3, a closed formula for the\nSW dispersion can be obtained via hydrodynamic the-\nory [58]. In this theory, the localized spins in Eq. (1),\nMn= (Re ϕn,Imϕn,1−m)T, are expressed using com-\nplex field ϕnand uniform spin density m≪1. Then,\nusing the SW ansatz ϕn(t) =P\nqUqei(q·rn−ωqt), we ob-\ntain the dispersion relation for the SWs\nωq= (Jq2+K−B)\u0002\n1 +i(αG+˜λq)\u0003\n, (4)\nwhere qis the wavevector and ωis their frequency. Thedamping of the SW is then given by the imaginary part\nof the dispersion in Eq. (4), Γ q≡Imωq. It is comprised\nby contributions from the local scalar Gilbert damping\nαGand the Fourier transform of the nonlocal damping\ntensor, ˜λq=R\ndrnλrneiq·rn.\nResults for FI/NM bilayer. —We warm up by extract-\ning Γ qfor the simplest of the three cases in Fig. 1, a\none-dimensional (1D) FI chain under a 1D NM over-\nlayer with spin-degenerate quadratic electronic energy-\nmomentum dispersion, ϵkσ=t0k2\nx, where t0=ℏ2/2m.\nThe GFs and spectral functions in Eq. (2), can be\ncalculated in the momentum representation, yielding\nλ1D\nR=2J2\nsd\nπv2\nFcos2(kFR)I3, where vFis the Fermi velocity,\nR≡ |R|, and kFis the Fermi wavevector. Moreover, its\nFourier transform, ˜λq=2J2\nsd\nv2\nF[δ(q) +δ(q−2kF)/2], dic-\ntates additional damping to SWs of wavevector q=\n0,±2kF. Although the Dirac delta function in this ex-\npression is unbounded, this unphysical feature is an ar-\ntifact of the small amplitude, m≪1, approximation\nwithin the hydrodynamic approach [58]. The features of\nsuch wavevector-dependent damping in 1D can be cor-\nroborated via TDNEGF+LLG numerically exact simu-\nlations [47–50] of a finite-size nanowire, similar to the\nsetup depicted in Fig. 1(a) but sandwiched between two\nNM semi-infinite leads. For example, by exciting SWs\nvia injection of spin-polarized current into the metallic\noverlayer of such a system, as pursued experimentally in\nspintronics and magnonics [21, 57], we find in Fig. 2(a)\nthat wavevector qof thereby excited coherent SW in-4\ncreases with increasing anisotropy K. However, the max-\nimum wavevector qmaxis limited by kF[Fig. 2(b)]. This\nmeans that SWs with q≳kFare subjected to additional\ndamping, inhibiting their generation. Although our an-\nalytical results predict extra damping at q= 2kF, finite\nsize effects and the inclusion of semi-infinite leads in TD-\nNEGF+LLG simulations lower this cutoff to kF.\nSince SW experiments are conducted on higher-\ndimensional systems, we also investigate damping\non SWs in a two-dimensional (2D) FI/NM bilayer.\nThe electronic energy-momentum dispersion is then\nϵkσ=t0(k2\nx+k2\ny), and the nonlocal damping and its\nFourier transform are given by\nλNM\nR=k2\nFJ2\nsd\n2πv2\nFJ2\n0(kFR)I3, (5)\n˜λNM\nq=kFJ2\nsdΘ(2kF−q)\n2πv2\nFqp\n1−(q/2kF)2, (6)\nwhere Jn(x) is the n-th Bessel function of the first kind,\nand Θ( x) is the Heaviside step function. The nonlo-\ncal damping in Eqs. (5) and (6) is plotted in Fig. 3(a),\nshowing realistic decay with increasing R, in contrast to\nunphysical infinite range found in 1D case. Addition-ally, SW damping in Eq. (6) is operative for wavectors\n0≤q≤2kF, again diverging for q= 0,2kFdue to arti-\nfacts of hydrodynamic theory [58]. Therefore, unphysical\ndivergence can be removed by going back to discrete lat-\ntice, such as solid curves in Fig. 3(e) obtained for n=1–\n100 LMMs by solving numerically a system of coupled\nLLG Eq. (1) where λRin 2D is used [45]. In this numer-\nical treatment, we use kF= 0.5a−1where ais the lattice\nspacing; k2\nFJ2\nsd/2πv2\nF=η= 0.1;K= 0; Bext= 0.1J;\nandαG= 0.1.\nResults for FI/HM bilayer. —Heavy metals (such as of-\nten employed Pt, W, Ta) exhibit strong SOC effects due\nto their large atomic number. We mimic their presence at\nthe FI/HM interface [31] by using 2D energy-momentum\ndispersion ϵk=t0(k2\nx+k2\ny) +tSOC(σxky−σykx), which\nincludes spin-splitting due to the Rashba SOC [31, 32].\nUsing this dispersion, Eq. (2) yields\nλHM\nR=\nλxx\nRλxy\nR0\nλxy\nRλyy\nR0\n0 0 λzz\nR\n, (7)\nfor the nonlocal damping tensor. Its components are, in\ngeneral, different from each other\nλxx\nR=J2\nsd\n4π\u0014\u0012kF↑\nvF↑J0(kF↑R) +kF↓\nvF↓J0(kF↓R)\u00132\n+ cos(2 θ)\u0012kF↑\nvF↑J1(kF↑R)−kF↓\nvF↓J1(kF↓R)\u00132\u0015\n, (8a)\nλyy\nR=J2\nsd\n4π\u0014\u0012kF↑\nvF↑J0(kF↑R) +kF↓\nvF↓J0(kF↓R)\u00132\n−cos(2 θ)\u0012kF↑\nvF↑J1(kF↑R)−kF↓\nvF↓J1(kF↓R)\u00132\u0015\n, (8b)\nλzz\nR=J2\nsd\n4π\u0014\u0012kF↑\nvF↑J0(kF↑R) +kF↓\nvF↓J0(kF↓R)\u00132\n−\u0012kF↑\nvF↑J1(kF↑R)−kF↓\nvF↓J1(kF↓R)\u00132\u0015\n, (8c)\nλxy\nR=−J2\nsdsin(2θ)\n4π\u0012kF↑\nvF↑J1(kF↑R)−kF↓\nvF↓J1(kF↓R)\u00132\n, (8d)\nwhere kF↑andkF↓are the spin-split Fermi wavevec-\ntors [Fig. 1(b)], and θis the relative orientation angle\n[Fig. 1(b)] between the SW wavevector qand the kxdi-\nrection. Thus, the nonlocal damping tensor in Eq. (7)\ngenerated by HM overlayer is anisotropic in spin due to\nits different diagonal elements, as well as nonzero off-\ndiagonal elements. It is also anisotropic in space due to\nits dependence on the angle θ. Its elements [Eqs. (8)]\nare plotted in Figs. 3(b), 3(c), 3(f), and 3(g) using\ntSOC= 0.3t0. They may become negative, signifying the\npossibility of antidamping torque [21] exerted by conduc-\ntion electrons. However, the dominant effect of nearby\nLMMs and the presence of local scalar αGensures that\nLMM dynamics is damped overall. Although there is no\nclosed expression for the SW dispersion in the presence of\nanisotropic λHM\nR, we can still extract SW damping Γ qin-duced by an HM overlayer from the exponential decay of\nthe SW amplitude in numerical integration of extended\nLLG Eq. (1) using SW initial conditions with varying q.\nFor an HM overlayer with realistic [31, 32] tSOC= 0.1t0\nthe results in Fig. (4)(a) are very similar to those ob-\ntained for NM overlayer with the same Fermi energy.\nAlso, the spatial anisotropy of λHM\nRdid not translate into\nθ-dependence of the SW damping.\nResults for FI/AM bilayer. —Altermagnets [33, 34] are\na novel class of antiferromagnets with spin-split elec-\ntronic energy-momentum dispersion despite zero net\nmagnetization or lack of SOC. They are currently in-\ntensely explored as a new resource for spintronics [51,\n61, 62] and magnonics [63, 64]. A simple model for\nan AM overlayer employs energy-momentum dispersion\nϵkσ=t0(k2\nx+k2\ny)−tAMσ(k2\nx−k2\ny) [33, 34], where tAMis5\n0.0 0.5 1.0 1.5\nq (1/a)1.01.52.02.5Γq≡Im(ωq) (J/¯ h)×10−1\n(a)NM\nHM\n0.0 0.5 1.0 1.5\nq (1/a)1234Γq≡Im(ωq) (J/¯ h)×10−1\n(b)NM\nAM :θ= 45◦\nAM :θ= 0◦\nFIG. 4. (a) Wavevector-dependent damping Γ qof SWs under\nNM or HM overlayer with the Rashba SOC of strength tSOC=\n0.1t0. (b) Γ qof SWs under AM overlayer with tAM= 0.8t0\nand for different relative orientations of FI and AM layers\nmeasured by angle θ[Fig. 1]. All calculations employ η= 0.1\nand Fermi energy εF= 0.25t0.\nthe parameter characterizing anisotropy in the AM. The\ncorresponding λAM\nR= diag( λ⊥\nR, λ⊥\nR, λ∥\nR) tensor has three\ncomponents, which we derive from Eq. (2) as\nλ⊥\nR=J2\nsd\n4πA+A−\u0014\nJ2\n0\u0012rϵF\nt0R+\u0013\n+J2\n0\u0012rϵF\nt0R−\u0013\u0015\n,\n(9a)\nλ∥\nR=J2\nsd\n2πA+A−J0\u0012rϵF\nt0R+\u0013\nJ0\u0012rϵF\nt0R−\u0013\n, (9b)\nwhere A±=t0±tAMandR2\n±=X2/A±+Y2/A∓is\nthe anisotropically rescaled norm of R. They are plot-\nted in Figs. 3(d) and 3(h), demonstrating that λAM\nRis\nhighly anisotropic in space and spin due to the impor-\ntance of angle θ[61, 65, 66]. Its components can also\ntake negative values, akin to the case of λHM\nR. It is inter-\nesting to note that along the direction of θ= 45◦[gray\ndashed line in Figs. 3(d) and 3(h)], λ⊥\nR=λ∥\nRso that\nnonlocal damping tensor is isotropic in spin. The SW\ndamping Γ qinduced by an AM overlayer is extracted\nfrom numerical integration of extended LLG Eq. (1) and\nplotted in Fig. (4)(b). Using a relatively large, but real-\nistic [33], AM parameter tAM= 0.8t0, the SW damping\nfor experimentally relevant small wavevectors is reduced\nwhen compared to the one due to NM overlayer by up to\n65% for θ= 0◦[Fig. 4(b)]. Additional nontrivial features\nare observed at higher |q|, such as being operative for a\ngreater range of wavevectors and with maxima around\n|q|= 2p\nϵF/t0and|q|= 3p\nϵF/t0. Remarkably, these\npeaks vanish for wavevectors along the isotropic direction\nθ= 45◦[Fig. 4(b)].\nConclusions. —In conclusion, using SKFT-derived non-\nlocal damping tensor [45], we demonstrated a rigorous\npath to obtain wavevector damping of SWs in magnetic\ninsulator due to interaction with conduction electrons\nof metallic overlayer, as a setup often encountered in\nmagnonics [13–17, 30] where such SW damping was di-\nrectly measured in very recent experiments [24–27]. Ouranalytical expressions [Eqs. (5), (7), and (9)] for nonlo-\ncal damping tensor—using simple models of NM, HM,\nand AM overlayers as an input—can be directly plugged\ninto atomistic spin dynamics simulations [28]. For more\ncomplicated band structures of metallic overlayers, one\ncan compute λRnumerically via Eq. (2), including com-\nbination with first-principles calculations [40]. Since we\nfind that λHM,AM\nR is highly anisotropic in spin and space,\nthe corresponding SW damping Γ qthus understood mi-\ncroscopically from SKFT allows us to propose how to\nmanipulate it [Fig. 4]. For example, by using HM or AM\noverlayer and by changing their relative orientation with\nrespect to FI layer, Γ qcan be reduced by up to 65%\nat small wavevectors q[Fig. 4], which can be of great\ninterest to magnonics experiments and applications.\n∗bnikolic@udel.edu\n[1] Y. Li and W. Bailey, Wave-number-dependent Gilbert\ndamping in metallic ferromagnets., Phys. Rev. Lett. 116,\n117602 (2016).\n[2] L. Chen, C. Mao, J.-H. Chung, M. B. Stone, A. I.\nKolesnikov, X. Wang, N. Murai, B. Gao, O. De-\nlaire, and P. Dai, Anisotropic magnon damping by\nzero-temperature quantum fluctuations in ferromagnetic\nCrGeTe 3, Nat. Commun. 13, 4037 (2022).\n[3] P. Dai, H. Y. Hwang, J. Zhang, J. A. Fernandez-Baca,\nS.-W. Cheong, C. Kloc, Y. Tomioka, and Y. Tokura,\nMagnon damping by magnon-phonon coupling in man-\nganese perovskites, Phys. Rev. B 61, 9553 (2000).\n[4] S. P. Bayrakci, D. A. Tennant, P. Leininger, T. Keller,\nM. C. R. Gibson, S. D. Wilson, R. J. Birgeneau, and\nB. Keimer, Lifetimes of antiferromagnetic magnons in\ntwo and three dimensions: Experiment, theory, and nu-\nmerics, Phys. Rev. Lett. 111, 017204 (2013).\n[5] M. E. Zhitomirsky and A. L. Chernyshev, Colloquium :\nSpontaneous magnon decays, Rev. Mod. Phys. 85, 219\n(2013).\n[6] U. Bajpai, A. Suresh, and B. K. Nikoli´ c, Quantum many-\nbody states and Green 's functions of nonequilibrium\nelectron-magnon systems: Localized spin operators ver-\nsus their mapping to Holstein-Primakoff bosons, Phys.\nRev. B 104, 184425 (2021).\n[7] R. L. Smit, S. Keupert, O. Tsyplyatyev, P. A. Maksimov,\nA. L. Chernyshev, and P. Kopietz, Magnon damping in\nthe zigzag phase of the Kitaev-Heisenberg-Γ model on a\nhoneycomb lattice, Phys. Rev. B 101, 054424 (2020).\n[8] M. Gohlke, A. Corticelli, R. Moessner, P. A. McClarty,\nand A. Mook, Spurious symmetry enhancement in lin-\near spin wave theory and interaction-induced topology\nin magnons, Phys. Rev. Lett. 131, 186702 (2023).\n[9] E. Hankiewicz, G. Vignale, and Y. Tserkovnyak, Inhomo-\ngeneous Gilbert damping from impurities and electron-\nelectron interactions, Phys. Rev. B 78, 020404(R) (2008).\n[10] Y. Tserkovnyak, E. M. Hankiewicz, and G. Vignale,\nTransverse spin diffusion in ferromagnets, Phys. Rev. B\n79, 094415 (2009).\n[11] P. Nikoli´ c, Universal spin wave damping in magnetic\nWeyl semimetals, Phys. Rev. B 104, 024414 (2021).6\n[12] M. Isoda, Spin-wave damping in itinerant electron ferro-\nmagnets, J. Phys.: Condens. Matter 2, 3579 (1990).\n[13] A. Chumak, V. Vasyuchka, A. Serga, and B. Hillebrands,\nMagnon spintronics, Nat. Phys. 11, 453 (2015).\n[14] G. Csaba, A. Papp, and W. Porod, Perspectives of using\nspin waves for computing and signal processing, Phys.\nLett. A 381, 1471 (2017).\n[15] A. V. Chumak, Fundamentals of magnon-based comput-\ning, arXiv:1901.08934 (2019).\n[16] A. V. Chumak, P. Kabos, M. Wu, C. Abert, C. Adel-\nmann, A. O. Adeyeye, J. Akerman, F. G. Aliev,\nA. Anane, A. Awad, et al. , Advances in magnetics\nroadmap on spin-wave computing, IEEE Trans. Magn.\n58, 1 (2022).\n[17] A. Mahmoud, F. Ciubotaru, F. Vanderveken, A. V. Chu-\nmak, S. Hamdioui, C. Adelmann, and S. Cotofana, In-\ntroduction to spin wave computing, J. Appl. Phys. 128,\n161101 (2020).\n[18] A. Hamadeh, O. d’Allivy Kelly, C. Hahn, H. Meley,\nR. Bernard, A. H. Molpeceres, V. V. Naletov, M. Viret,\nA. Anane, V. Cros, et al. , Full control of the spin-wave\ndamping in a magnetic insulator using spin-orbit torque,\nPhys. Rev. Lett. 113, 197203 (2014).\n[19] A. Akhiezer, V. Baryakhtar, and S. Peletminskii, Coher-\nent amplification of spin waves, Sov. Phys. JETP 18, 235\n(1964).\n[20] M. Evelt, V. E. Demidov, V. Bessonov, S. O. Demokritov,\nJ. L. Prieto, M. Mu˜ noz, J. B. Youssef, V. V. Naletov,\nG. de Loubens, O. Klein, et al. , High-efficiency control of\nspin-wave propagation in ultra-thin yttrium iron garnet\nby the spin-orbit torque, Appl. Phys. Lett. 108, 172406\n(2016).\n[21] V. E. Demidov, S. Urazhdin, A. Anane, V. Cros, and\nS. O. Demokritov, Spin–orbit-torque magnonics, J. Appl.\nPhys. 127, 170901 (2020).\n[22] D. Breitbach, M. Schneider, B. Heinz, F. Kohl,\nJ. Maskill, L. Scheuer, R. O. Serha, T. Br¨ acher, B. L¨ agel,\nC. Dubs, et al. , Stimulated amplification of propagating\nspin waves, Phys. Rev. Lett. 131, 156701 (2023).\n[23] F. Bloch, Zur theorie des ferromagnetismus, Z. Phys. 61,\n206 (1930).\n[24] I. Bertelli, B. G. Simon, T. Yu, J. Aarts, G. E. W. Bauer,\nY. M. Blanter, and T. van der Sar, Imaging spin-wave\ndamping underneath metals using electron spins in dia-\nmond, Adv. Quantum Technol. 4, 2100094 (2021).\n[25] S. Mae, R. Ohshima, E. Shigematsu, Y. Ando, T. Shinjo,\nand M. Shiraishi, Influence of adjacent metal films on\nmagnon propagation in Y 3Fe5O12, Phys. Rev. B 105,\n104415 (2022).\n[26] A. Krysztofik, N. Kuznetsov, H. Qin, L. Flajˇ sman,\nE. Coy, and S. van Dijken, Tuning of magnetic damping\nin Y 3Fe5O12/metal bilayers for spin-wave conduit termi-\nnation, Materials 15, 2814 (2022).\n[27] R. O. Serha, D. A. Bozhko, M. Agrawal, R. V. Verba,\nM. Kostylev, V. I. Vasyuchka, B. Hillebrands, and A. A.\nSerga, Low-damping spin-wave transmission in YIG/Pt-\ninterfaced structures, Adv. Mater. Interfaces 9, 2201323\n(2022).\n[28] R. Evans, W. Fan, P. Chureemart, T. Ostler, M. O. Ellis,\nand R. Chantrell, Atomistic spin model simulations of\nmagnetic nanomaterials, J. Phys.: Condens. Matter 26,\n103202 (2014).\n[29] S.-K. Kim, Micromagnetic computer simulations of spin\nwaves in nanometre-scale patterned magnetic elements,J. Phys. D: Appl. Phys. 43, 264004 (2010).\n[30] A. A. Serga, A. V. Chumak, and B. Hillebrands, YIG\nmagnonics, J. Phys. D: Appl. Phys. 43, 264002 (2010).\n[31] G. Bihlmayer, P. No¨ el, D. V. Vyalikh, E. V. Chulkov, and\nA. Manchon, Rashba-like physics in condensed matter,\nNat. Rev. Phys. 4, 642 (2022).\n[32] A. Manchon, H. C. Koo, J. Nitta, S. M. Frolov, and R. A.\nDuine, New perspectives for Rashba spin–orbit coupling,\nNat. Mater. 14, 871 (2015).\n[33] L. ˇSmejkal, J. Sinova, and T. Jungwirth, Emerging re-\nsearch landscape of altermagnetism, Phys. Rev. X 12,\n040501 (2022).\n[34] L. ˇSmejkal, A. B. Hellenes, R. Gonz´ alez-Hern´ andez,\nJ. Sinova, and T. Jungwirth, Giant and tunneling mag-\nnetoresistance in unconventional collinear antiferromag-\nnets with nonrelativistic spin-momentum coupling, Phys.\nRev. X 12, 011028 (2022).\n[35] T. Gilbert, A phenomenological theory of damping in\nferromagnetic materials, IEEE Trans. Magn. 40, 3443\n(2004).\n[36] W. M. Saslow, Landau-Lifshitz or Gilbert damping? that\nis the question, J. Appl. Phys. 105, 07D315 (2009).\n[37] S. Zhang and S. L. Zhang, Generalization of the Landau-\nLifshitz-Gilbert equation for conducting ferromagnets.,\nPhys. Rev. Lett. 102, 086601 (2009).\n[38] H. Yuan, Z. Yuan, K. Xia, and X. R. Wang, Influence\nof nonlocal damping on the field-driven domain wall mo-\ntion, Phys. Rev. B 94, 064415 (2016).\n[39] R. Verba, V. Tiberkevich, and A. Slavin, Damping of lin-\near spin-wave modes in magnetic nanostructures: Local,\nnonlocal, and coordinate-dependent damping, Phys. Rev.\nB98, 104408 (2018).\n[40] Z. Lu, I. P. Miranda, S. Streib, M. Pereiro, E. Sj¨ oqvist,\nO. Eriksson, A. Bergman, D. Thonig, and A. Delin, In-\nfluence of nonlocal damping on magnon properties of fer-\nromagnets, Phys. Rev. B 108, 014433 (2023).\n[41] K. Gilmore, Y. Idzerda, and M. Stiles, Identification of\nthe dominant precession-damping mechanism in Fe, Co,\nand Ni by first-principles calculations., Phys. Rev. Lett.\n99, 027204 (2007).\n[42] P. Trempler, R. Dreyer, P. Geyer, C. Hauser, G. Wolters-\ndorf, and G. Schmidt, Integration and characterization\nof micron-sized YIG structures with very low Gilbert\ndamping on arbitrary substrates, Appl. Phys. Lett. 117,\n232401 (2020).\n[43] F. Casola, T. van der Sar, and A. Yacoby, Probing\ncondensed matter physics with magnetometry based on\nnitrogen-vacancy centres in diamond, Nat. Rev. Mater.\n3, 17088 (2018).\n[44] I. N. Krivorotov, D. V. Berkov, N. L. Gorn, N. C. Emley,\nJ. C. Sankey, D. C. Ralph, and R. A. Buhrman, Large-\namplitude coherent spin waves excited by spin-polarized\ncurrent in nanoscale spin valves, Phys. Rev. B 76, 024418\n(2007).\n[45] F. Reyes-Osorio and B. K. Nikoli´ c, Gilbert damping in\nmetallic ferromagnets from Schwinger-Keldysh field the-\nory: Intrinsically nonlocal and nonuniform, and made\nanisotropic by spin-orbit coupling, arXiv:2306.13013\n(2023).\n[46] A. Kamenev, Field Theory of Non-Equilibrium Systems\n(Cambridge University Press, Cambridge, 2023).\n[47] M. Petrovi´ c, B. Popescu, P. Plech´ aˇ c, and B. Nikoli´ c,\nSpin and charge pumping by current-driven magnetic\ndomain wall motion: A self-consistent multiscale time-7\ndependent-quantum/time-dependent-classical approach,\nPhys. Rev. Appl. 10, 054038 (2018).\n[48] U. Bajpai and B. Nikoli´ c, Time-retarded damping and\nmagnetic inertia in the Landau-Lifshitz-Gilbert equation\nself-consistently coupled to electronic time-dependent\nnonequilibrium Green functions, Phys. Rev. B 99, 134409\n(2019).\n[49] M. D. Petrovi´ c, U. Bajpai, P. Plech´ aˇ c, and B. K. Nikoli´ c,\nAnnihilation of topological solitons in magnetism with\nspin-wave burst finale: Role of nonequilibrium electrons\ncausing nonlocal damping and spin pumping over ultra-\nbroadband frequency range, Phys. Rev. B 104, l020407\n(2021).\n[50] A. Suresh, U. Bajpai, and B. K. Nikoli´ c, Magnon-driven\nchiral charge and spin pumping and electron-magnon\nscattering from time-dependent quantum transport com-\nbined with classical atomistic spin dynamics, Phys. Rev.\nB101, 214412 (2020).\n[51] H. Bai, Y. C. Zhang, Y. J. Zhou, P. Chen, C. H. Wan,\nL. Han, W. X. Zhu, S. X. Liang, Y. C. Su, X. F. Han,\net al. , Efficient spin-to-charge conversion via altermag-\nnetic spin splitting effect in antiferromagnet RuO 2, Phys.\nRev. Lett. 130, 216701 (2023).\n[52] F. M. Haehl, R. Loganayagam, and M. Rangamani,\nSchwinger-Keldysh formalism. Part I: BRST symmetries\nand superspace, J. High Energy Phys. 2017 , 69 (2017).\n[53] F. M. Haehl, R. Loganayagam, and M. Rangamani,\nSchwinger-Keldysh formalism. Part II: thermal equivari-\nant cohomology, J. High Energy Phys. 2017 , 70 (2017).\n[54] J. Berges, Nonequilibrium quantum fields: From cold\natoms to cosmology, arXiv:1503.02907 (2015).\n[55] J. Anders, C. R. J. Sait, and S. A. R. Horsley, Quantum\nBrownian motion for magnets, New J. Phys. 24, 033020\n(2022).[56] S. Leiva M., S. A. D´ ıaz, and A. S. Nunez, Origin of the\nmagnetoelectric couplings in the spin dynamics of molec-\nular magnets, Phys. Rev. B 107, 094401 (2023).\n[57] M. Madami, S. Bonetti, G. Consolo, S. Tacchi, G. Car-\nlotti, G. Gubbiotti, F. B. Mancoff, M. A. Yar, and\nJ.˚Akerman, Direct observation of a propagating spin\nwave induced by spin-transfer torque, Nat. Nanotechnol.\n6, 635 (2011).\n[58] B. Halperin and P. Hohenberg, Hydrodynamic theory of\nspin waves, Phys. Rev. 188, 898 (1969).\n[59] M. Sayad and M. Potthoff, Spin dynamics and relax-\nation in the classical-spin Kondo-impurity model beyond\nthe Landau-Lifschitz-Gilbert equation, New J. Phys. 17,\n113058 (2015).\n[60] D. Ralph and M. Stiles, Spin transfer torques, J. Magn.\nMagn. Mater. 320, 1190 (2008).\n[61] C. Sun and J. Linder, Spin pumping from a ferromagnetic\ninsulator into an altermagnet, Phys. Rev. B 108, L140408\n(2023).\n[62] R.-W. Zhang, C. Cui, R. Li, J. Duan, L. Li, Z.-M. Yu, and\nY. Yao, Predictable gate-field control of spin in altermag-\nnets with spin-layer coupling, arXiv:2306.08902 (2023).\n[63] L. ˇSmejkal, A. Marmodoro, K.-H. Ahn, R. Gonzalez-\nHernandez, I. Turek, S. Mankovsky, H. Ebert, S. W.\nD’Souza, O. ˇSipr, J. Sinova, et al. , Chiral magnons in\naltermagnetic RuO 2, arXiv:2211.13806 (2022).\n[64] Z. Jin, H. Yang, Z. Zeng, Y. Cao, and P. Yan, Cavity-\ninduced strong magnon-magnon coupling in altermag-\nnets, arXiv:2307.00909 (2023).\n[65] M. Papaj, Andreev reflection at the altermagnet-\nsuperconductor interface, Phys. Rev. B 108, L060508\n(2023).\n[66] C. Sun, A. Brataas, and J. Linder, Andreev reflection in\naltermagnets, Phys. Rev. B 108, 054511 (2023)." }, { "title": "2211.00257v1.Modeling_of_three_dimensional_betatron_oscillation_and_radiation_reaction_in_plasma_accelerators.pdf", "content": "Modeling of three-dimensional betatron oscillation and radiation reaction in plasma\naccelerators\nYulong Liu and Ming Zeng\u0003\nInstitute of High Energy Physics, Chinese Academy of Sciences, Beijing 100049, China and\nUniversity of Chinese Academy of Sciences, Beijing 100049, China\n(Dated: November 2, 2022)\nBetatron oscillation is a commonly known phenomenon in laser or beam driven plasma wake\feld\naccelerators. In the conventional model, the plasma wake provides a linear focusing force to a rela-\ntivistic electron, and the electron oscillates in one transverse direction with the betatron frequency\nproportional to 1 =p\r, where\ris the Lorentz factor of the electron. In this work, we extend this\nmodel to three-dimensional by considering the oscillation in two transverse and one longitudinal\ndirections. The long-term equations, with motion in the betatron time scale averaged out, are\nobtained and compared with the original equations by numerical methods. In addition to the lon-\ngitudinal and transverse damping due to radiation reaction which has been found before, we show\nphenomena including the longitudinal phase drift, betatron phase shift and betatron polarization\nchange based on our long-term equations. This work can be highly valuable for future plasma based\nhigh-energy accelerators and colliders.\nI. INTRODUCTION\nThe new generation of accelerators, using plasma as\nthe acceleration media, o\u000ber high acceleration gradient\nin the order of 10{100 GV =m and strong transverse\nfocusing \feld [1{3]. Depends on the driver type, the\nplasma accelerators are named laser wake\feld acceler-\nators (LWFAs), which are driven by laser pulses, and\nplasma wake\feld accelerators (PWFAs), which are driven\nby charged particle beams. When a high intensity laser\npulse (>\u00181018W=cm2) or a high current particle beam\n(>\u00181 kA) propagates through an underdense plasma, the\nradiation pressure of the laser or the space charge of the\nbeam expels all plasma electrons away from axis radi-\nally, leaves behind a nearly uniform ion channel. This\nhigh-intensity three-dimensional (3D) regime has been\nreferred to as the blowout regime [4, 5]. In this regime,\nthe expelled electrons are pulled back by the ion chan-\nnel and thereby bubble-like plasma wake wave is created.\nThe wake consists of a longitudinal electric \feld that is\na function of distance behind the driver, and transversal\nelectromagnetic \felds that are proportional to the o\u000b-\naxis distance. Consequently, in addition to the longitu-\ndinal acceleration / deceleration, the electrons reside in\nthe wake also perform radial oscillation, called betatron\noscillation (BO), under the action of transverse focusing\n\feld, with the frequency !\f=!p\u0014=p\r, where!pis the\nplasma frequency, \ris the relativistic factor of the elec-\ntron, and\u0014is the focusing constant which takes 1 =p\n2\nfor the blow-out regime [6, 7].\nElectrons emit synchrotron radiation when performing\nBO [8{10], which a\u000bects the electrons in return. Such ef-\nfect is called the radiation reaction (RR) and its classical\nexpression is the Lorentz{Abraham{Dirac (LAD) equa-\ntion or the Landau-Lifshitz equation [11, 12]. Because\n\u0003Corresponding author: zengming@ihep.ac.cnthe RR force is proportional to the classical electron ra-\ndiusre\u00192:81\u000210\u000015m, it is generally negligible unless\nunder extreme conditions [13, 14] or for su\u000eciently long\ninteraction time [15]. The BO in a plasma accelerator is\nanother good case for such long interaction time. The ra-\ndiation leads to the energy loss of electrons and in return\na\u000bects the energy-dependent betatron frequency, as well\nas the other beam properties, such as the energy spread\nand emittance [16{23].\nAlthough there are many established theories on the\nlong-term RR damping e\u000bect of BO, their models assert\nthe electron only moves in one plane, spanned by the\nlongitudinal direction and one transverse direction, thus\nonly linear polarization is considered. Moreover, these\nmodels usually neglect the longitudinal and energy os-\ncillations during one betatron period. In this paper, we\nestablish a 3D BO model with RR e\u000bect, which gener-\nalizes the betatron polarization from linear to elliptical,\nand considers both the longitudinal and energy oscilla-\ntions. Long-term equations (LTEs), without resolving\nthe betatron period, are derived and veri\fed by numer-\nical methods. The LTEs reproduce the previous results,\nsuch as longitudinal and transverse damping due to RR,\nas the special cases, and meanwhile reveal new phenom-\nena such as betatron phase shift and polarization change.\nThe rest of this paper is organized as follows. Sec. II\ngives the original form of the force, and shows the equa-\ntions of motion expressed by the transverse motion only.\nSec. III derives the LTEs by averaging the equations of\nmotion through one betatron period. Sec. IV discusses\nthe di\u000berent phenomena in the RR dominant regime and\nthe betatron phase shift dominant regime. Sec. V nu-\nmerically veri\fes the LTEs by comparing with the code\nPTracker which solves the equations with the original\nform of force. Before start, it is worth noting that we use\nplasma normalization units described in Appx. A, and\nsome symbols and calculation rules used often during the\nderivation are described in Appx. B.arXiv:2211.00257v1 [physics.acc-ph] 1 Nov 20222\nII. THE ELECTROMAGNETIC FIELD AND\nTHE EQUATIONS OF MOTION\nConsider an electron with \r\u001d1 is trapped in a plasma\nwake\feld with the longitudinal co-moving coordinate \u0010=\nz\u0000\fwt, where the wake is moving in the z+ direction with\nthe phase velocity \fw. Neglect the interaction between\nthe beam particles, the electromagnetic \feld provided by\nthe wake can be modeled as [7]\nEz=Ez0+\u0015\u00101; (1)\n~E?=\u00142(1\u0000\u0015)~ r; (2)\nB\u0012=\u0000\u00142\u0015r; (3)whereEz0=Ezj\u0010=h\u0010i,\u0015=dEz=d\u0010j\u0010=h\u0010i, and~ r= (x;y)\nis the transverse o\u000bset. The force can be expressed as\nfz=\u0000Ez0\u0000\u0015\u00101+\u00142\u0015(x\fx+y\fy) +frad\nz;(4)\nfx=\u0000\u00142(1\u0000\u0015+\u0015\fz)x+frad\nx; (5)\nfy=\u0000\u00142(1\u0000\u0015+\u0015\fz)y+frad\ny; (6)\nwhere\fx= _x,\fy= _y,\fz= _z=\fz0+_\u00101,\fz0=\fw+\n_h\u0010i=h\fzi, and~fradis the RR force. The formulas of\n3D BO with RR can be written in the form of transverse\nterms only (see Appx. C)\n_\r=\u0000Ez0\fz0+\u0012\u0015\fz0\n4+\u00142\u0015\u0000\u00142\u0013\n(x\fx+y\fy)\u00002\n3re\r2\u00144\u0000\nx2+y2\u0001\n; (7)\n_pz=\u0000Ez0+\u0015\u00121\n4+\u00142\u0013\n(x\fx+y\fy)\u00002\n3re\r2\u00144\u0000\nx2+y2\u0001\n; (8)\n_px=\u0000\u00142x+\u00142\u0015\n2\u0010\nh\ri\u00002+\f2\nx+\f2\ny\u0011\nx\u00002\n3re\r2\u00144\u0000\nx2+y2\u0001\n\fx; (9)\n_py=\u0000\u00142y+\u00142\u0015\n2\u0010\nh\ri\u00002+\f2\nx+\f2\ny\u0011\ny\u00002\n3re\r2\u00144\u0000\nx2+y2\u0001\n\fy; (10)\nwhere\n~ p=\r~\f; (11)\n\fz0= 1\u00001\n2\u0010\nh\ri\u00002+\n\f2\nx\u000b\n+\n\f2\ny\u000b\u0011\n; (12)\nandreis also normalized to k\u00001\np. One may note the\nsecond terms in Eqs. (7) - (10), which come from the\noscillation of \fzand the modulation of \rdue to trans-\nverse oscillation, were neglected in previous works. In\nthe following sections we show these terms lead to new\nphenomena.\nIII. THE LONG-TERM EQUATIONS OF 3D\nBETATRON OSCILLATION\nIn this section we use the same averaging method as\nRef. [20]. We \frstly introduce two complex variables\nU=\u0010\nx\u0000i\u0014\u00001\r1\n2\fx\u0011\ne\u0000i'; (13)\nV=\u0010\ny\u0000i\u0014\u00001\r1\n2\fy\u0011\ne\u0000i'; (14)where\n'=Z\n!\fdt=\u0014Z\n\r\u00001\n2dt (15)\nis the betatron phase. Obviously jU1j\u001cjhUijandjV1j\u001c\njhVijare satis\fed, and we apply the rules in Appx. B\noften in the following. Because the equations for xandy\ndirections are symmetric, we may derive for xdirection\nonly, then exchange xandy,UandVfor theydirection.\nWith the help of Eqs. (7), (9) and (11), we may write the\ntime derivative of Eq. (13) as\n_U=\u0000i1\n2\u0014\u00001\r\u00001\n2Ez0\fz0\fxe\u0000i'+i1\n3re\u00143\r3\n2\u0000\nx2+y2\u0001\n\fxe\u0000i'\n+i1\n2\u0014\u00001\r\u00001\n2\u0014\u0015\fz0\n4+\u00142(\u0015\u00001)\u0015\n(x\fx+y\fy)\fxe\u0000i'\u0000i1\n2\u0014\u0015\r\u00001\n2\u0010\nh\ri\u00002+\f2\nx+\f2\ny\u0011\nxe\u0000i':(16)\nIn the following, we omit hionUandVfor conve- nience, so that all UandVactually meanhUiandhVi.3\nWe perform average on Eq. (16) to obtain (note only the terms with ei0'survive after averaging)\n_U=1\n4Ez0\fz0h\ri\u00001U\u00001\n24re\u00144h\ri\u0010\njUj2U+ 2jVj2U\u0000V2U\u0003\u0011\n+i1\n64\u0014\u0015\fz0h\ri\u00003\n2\u0010\njUj2U+V2U\u0003\u0011\n\u0000i1\n16\u00143h\ri\u00003\n2h\u0010\njUj2+ 2\u0015jVj2\u0011\nU\u0000(2\u0015\u00001)V2U\u0003i\n\u0000i1\n4\u0014\u0015h\ri\u00005\n2U:(17)\nBy asserting V= 0 and omitting the last three terms in\nEq. (17), which comes from the second terms in Eqs. (7)\n- (10), we can reproduce Eq. (19) in Ref. [20].\nThe average of Eq. (7) leads to\nh_\ri=\u0000Ez0\fz0\u00001\n3re\u00144h\ri2\u0010\njUj2+jVj2\u0011\n; (18)\nwith the second term reproduces Eq. (B2) in Ref. [18].\nEz0is a function ofh\u0010i, which obeys\n_h\u0010i=1\n2\r\u00002\nw\u00001\n2h\ri\u00002\u00001\n4\u00142h\ri\u00001\u0010\njUj2+jVj2\u0011\n;(19)\nwhere\rw=\u0000\n1\u0000\f2\nw\u0001\u00001=2and we have used Eq. (C2).The above averaged equations Eqs. (17), (18) and (19)\nare already enough to predict the long-term behavior of\nBO. However, the equations for the complex variables are\nnot explicit. To make them more physically meaningful,\nwe introduce\nU=jUjei\bx; (20)\nV=jVjei\by; (21)\n\u0001\b = \by\u0000\bx: (22)\njUjhas the meaning of the BO amplitude in the xdirec-\ntion, and \b xthe phase shift. For the ydirection they are\nsimilar. Thus \u0001\b is the phase di\u000berence of the two di-\nrections. By Applying djUj=dt=\u0010\n_UU\u0003+U_U\u0003\u0011\n=2jUj\nand _\bx=\u0010\n_UU\u0003\u0000_U\u0003U\u0011\n=2ijUj2we get\ndjUj\ndt=1\n4Ez0\fz0h\ri\u00001jUj\u00001\n24re\u00144h\rih\njUj3+jVj2jUj(2\u0000cos 2\u0001\b)i\n\u00001\n16\u0014\u00141\n4\u0015\fz0\u0000\u00142(1\u00002\u0015)\u0015\nh\ri\u00003\n2jVj2jUjsin 2\u0001\b;(23)\n_\bx=1\n24re\u00144h\rijVj2sin 2\u0001\b +1\n64\u0014\u0015\fz0h\ri\u00003\n2h\njUj2+jVj2cos 2\u0001\bi\n\u00001\n16\u00143h\ri\u00003\n2h\njUj2+ 2\u0015jVj2+ (1\u00002\u0015)jVj2cos 2\u0001\bi\n\u00001\n4\u0014\u0015h\ri\u00005\n2;(24)\nd\u0001\b\ndt=\u00001\n24re\u00144h\ri\u0010\njVj2+jUj2\u0011\nsin 2\u0001\b +1\n8\u0014\u00141\n4\u0015\fz0\u0000\u00142(1\u00002\u0015)\u0015\nh\ri\u00003\n2\u0010\njVj2\u0000jUj2\u0011\nsin2\u0001\b: (25)\nNote when doing exchange of UandVfor theydirection,\none also has to change the \u0006sign of \u0001\b.To further simplify we notice Eq. (23) can be rewritten\nwith the help of Eq. (18)\ndjUj\ndt=\u00001\n4h_\ri\nh\rijUj\u00001\n8re\u00144h\ri\u0014\njUj3+4\u0000cos 2\u0001\b\n3jVj2jUj\u0015\n\u00001\n16\u0014\u00141\n4\u0015\fz0\u0000\u00142(1\u00002\u0015)\u0015\nh\ri\u00003\n2jVj2jUjsin 2\u0001\b;\n(26)\nwhich reproduces Eq. (66) in Ref. [22] if V= 0. Introduce\nSx=\u0014h\ri1\n2jUj2; (27)\nSy=\u0014h\ri1\n2jVj2; (28)which have the physical meaning of the areas (divided by\n2\u0019) of the ellipses encircled by the particle trajectory in\nx-pxandy-pyphase spaces. Then Eqs. (18), (19), (23),4\n(24) and (25) can be rewritten as\nh_\ri=\u0000Ez0\fz0\u00001\n3re\u00143h\ri3\n2(Sx+Sy); (29)\n_h\u0010i=1\n2\r\u00002\nw\u00001\n2h\ri\u00002\u00001\n4\u0014h\ri\u00003\n2(Sx+Sy); (30)\n_Sx=\u00001\n4re\u00143h\ri1\n2\u0012\nS2\nx+4\u0000cos 2\u0001\b\n3SxSy\u0013\n\u00001\n8\u00141\n4\u0015\fz0\u0000\u00142(1\u00002\u0015)\u0015\nh\ri\u00002SxSysin 2\u0001\b; (31)\n_\bx=1\n24re\u00143h\ri1\n2Sysin 2\u0001\b +1\n64\u0015\fz0h\ri\u00002(Sx+Sycos 2\u0001\b)\n\u00001\n16\u00142h\ri\u00002[Sx+ 2\u0015Sy+ (1\u00002\u0015)Sycos 2\u0001\b]\u00001\n4\u0014\u0015h\ri\u00005\n2;(32)\nd\u0001\b\ndt=\u00001\n24re\u00143h\ri1\n2(Sy+Sx) sin 2\u0001\b +1\n8\u00141\n4\u0015\fz0\u0000\u00142(1\u00002\u0015)\u0015\nh\ri\u00002(Sy\u0000Sx) sin2\u0001\b: (33)\nIt is generally safe to take \fz0= 1 here. But Eq. (12), or\n\fz0= 1\u00001\n2h\nh\ri\u00002+1\n2\u0014h\ri\u00003=2(Sx+Sy)i\n, gives a bet-\nter accuracy. The above long-term equations, Eqs. (29)\n- (33), show that the BO experiences acceleration (for\nEz0<0) or deceleration (for Ez0>0), radiation damp-\ning, longitudinal phase drift, and betatron phase shift.\nThese equations may be used for the long-term behavior\nof BO without resolving the betatron period.\nIV. DISCUSSION ON TWO REGIMES\nFrom Eqs. (31) - (33) we note two regimes. One is the\nRR dominant regime, where reh\ri5=2\u001d1, so that the\n\frst terms in Eqs. (31) - (33) dominate. This regime has\nbeen discussed before [20], although only for the linearly\npolarized case \u0001\b = 0 (so that the ratio between Sx\nandSyis a constant). The other is the betatron phase\nshift dominant regime, where reh\ri5=2\u001c1, so that the\nremaining terms in Eqs. (31) - (33) dominate. These\nterms were previously proposed [22], but the betatron\nphase shift is found for the \frst time in the present work.\nIn the RR dominant regime, an interesting phe-\nnomenon is that an elliptical polarization (in the x-y\nplane) always approaches linear polarization, because\n\u0001\b always approaches the nearest integer multiple of\n\u0019according to Eq. (33). This phenomenon can also be\nviewed by rotating the xaxis to the major axis of the el-\nlipse, so that Sx>Syand \u0001\b =\u0019=2. De\fneR=Sy=Sx\nand perform time derivative with the help of Eq. (31)\n_R=\u00001\n6re\u00143h\ri1\n2R(Sx\u0000Sy)<0; (34)which suggests that the ellipse monotonically becomes\nthinner.\nThe betatron phase shift dominant regime requires a\nmoderate\r, or straightforwardly re!0, which corre-\nsponds to very dilute plasma case, leads to a constant\nS\u0011Sx+Sy. It can be proved that the time integral\nof Eq. (30) reproduces Eq. (6) in Ref. [21], which is the\nh\u0010idrift, in the case that h\rilinearly depends on t. In\nanother case that h\u0010idrifts around the zero point of Ez0,\nthe drift frequency can be obtained by using Eqs. (29)\nand (30), and asserting Ez0=\u0015h\u0010i\n!h\u0010i=s\n\u0015\fz0\u0012\n1 +3\n8\u0014h\ri1\n2S\u0013\nh\ri\u00003: (35)\nWe also note the angular momentum Lz=\rx\fy\u0000\ry\fx\nand its changing rate\nhLzi=\u0000S1\n2xS1\n2ysin \u0001\b; (36)\n_hLzi=\u00001\n3re\u00143h\ri1\n2(Sx+Sy)hLzi; (37)\nwhich obeys the law of conservation of angular momen-\ntum ifhLzi= 0 initially, or re!0. Especially for re!0,\nthe particle trajectory in the x-yplane generally encir-\ncles an ellipse with constant area and shape, which also\nhas precession leading to the rotation of the major and\nminor axes of the ellipse.5\nV. NUMERICAL COMPARISON OF\nLONG-TERM EQUATIONS AND THE\nORIGINAL ONES\nTo verify the LTEs, we solve them numerically us-\ning the backward-di\u000berentiation formulas (BDF) in the\nSciPy integration package [24]. Meanwhile, the original\nequations of motion with the force expressions Eqs. (4)\n- (6) are solved by Runge-Kutta 4th order method using\nthe code PTracker (PT) [25]. We choose four cases with\ntheir parameters and initial values listed in Tab. I, and\nthe comparison results are plotted from Fig. 1 to 4. Note\nthat \bxcannot be obtained directly from PT. Thus we\nperform the following treatment to the PT results\nx\u0001cos'=jUj\n2[cos \bx+ cos (2'+ \bx)]; (38)\nbecausex=jUjcos ('+ \bx), where'is obtained by\nnumerical integral based on Eq. (15). Then cos \b xcan\nbe obtained by a low-pass \flter. Similar treatment is\nperformed to obtain cos \u0001\b, according to\nx\u0001y=jUjjVj\n2[cos \u0001\b + cos (2 '+ \bx+ \by)]:(39)\nA case in the betatron phase shift dominant regime\nis shown in Fig. 1. We see Sx+Syis a constant, al-\nthoughSx,Syand \u0001\b change gradually. The approxi-\nmate \\phase-locking\" is chosen, i.e. \rw\u0019\rz0, thush\u0010i\noscillates near the zero point of Ezwith the drift fre-\nquency!h\u0010iaccording to Eq. (35). h\rioscillates with\nthe same drift frequency as shown in Fig. 1 (b).\nA second case during the transition of the two regimes\nis shown in Fig. 2. Sx+Syis approximately a constant\ninitially, and starts to decrease near the regime transition\n\r=r\u00002=5\ne.\nA third case in the RR dominant regime is shown in\nFig. 3. The initial values are chosen so that the particle\ntrajectory in the x-yplane is a ellipse with its major axis\nlaying on the xaxis. As shown in Fig. 3 (a), R=Sy=Sx\ndecreases monotonically, as predicted by Eq. (34).\nThe last case shown in Fig. 4 is also in the RR domi-\nnant regime, but the particle trajectory in the x-yplane\nis a oblique ellipse. As shown in Fig. 4 (c), \u0001\b gradually\napproaches \u0019, which is in accordance with the discussion\nin Sec. IV.\nIn all these plots, the results from PT and LTEs show\nagreement with high accuracy, demonstrating the cor-\nrectness of LTEs. Because the BO frequency is the high-\nest frequency in our physical process, the LTEs largely\nreduce the numerical complexity and meanwhile keep the\nlong-term accuracy.\nVI. CONCLUSIONS\nWe have established a three-dimensional betatron os-\ncillation model including radiation reaction to study the\n0 1 2 3 4\nt 1e52.55.07.510.012.5Sx, SySx, PT\nSx, LTE\nSy, PT\nSy, LTE\n0 1 2 3 4\nt 1e39698100102104\nPT\nLTE\n1 2 3 4\nt 1e51.0\n0.5\n0.00.51.0xy\nPT\nPT filtered\n|U||V|\n2cos, LTE\n1 2 3 4\nt 1e51.5\n1.0\n0.5\n0.00.51.0x cos\nPT\nPT filtered\n|U|\n2cosx, LTE\n(a) (b)\n(c)(d)\nFIG. 1. The numerical comparison of the LTEs and the orig-\ninal equations solved by PTracker in the betatron phase shift\ndominant regime. (a) SxandSychange with time, but Sx+Sy\nis a constant. (b) \rhas oscillation frequencies of 2 !\f\u00190:14\ndue to the BO and !h\u0010i\u00193:11\u000210\u00003due to the drift oscilla-\ntion ofh\u0010i. (c) The gray curve shows x\u0001yobtained from PT,\nand the black curve shows its low-pass \fltered result, which\nis compared with the LTE solution according to Eq. (39). (d)\nThe gray curve shows x\u0001cos'obtained from PT, and the\nblack curve shows its low-pass \fltered result, which is com-\npared with the LTE solution according to Eq. (38).\n0.0 0.5 1.0 1.5 2.0\nt 1e7010203040Sx, SySx, PT\nSx, LTE\nSy, PT\nSy, LTE\n0.0 0.5 1.0 1.5 2.0\nt 1e70.00.51.01.52.0\n1e4\nPT\nLTE\n0.5 1.0 1.5 2.0\nt 1e70.00.20.40.60.8xy\nPT\nPT filtered\n|U||V|\n2cos, LTE\n0.5 1.0 1.5 2.0\nt 1e70.25\n0.000.250.500.751.00xcos\nPT\nPT filtered\n|U|\n2cosx, LTE\n(a)(b)\n(c) (d)\nFIG. 2. The numerical comparison of the LTEs and the origi-\nnal equations solved by PTracker in the transition between the\nbetatron phase shift dominant and the RR dominant regimes.\n(a)Sx+Syis initially approximately a constant, but decreases\nlater. (b)\rincreases due to the acceleration \feld, and passes\nthe regime transition at \r=r\u00002=5\ne = 104. (c) and (d) show\nthe same treatments as in Fig. 1 (c) and (d).6\n0 2 4 6 8\nt 1e70.760.770.780.790.800.81RPT\nLTE\n0 1 2 3\nt 1e60.00.20.40.60.81.0\n1e6\nPT\nLTE\n1 2 3\nt 1e60.005\n0.0000.0050.0100.015xy\nPT\nPT filtered\n|U||V|\n2cos, LTE\n1 2 3\nt 1e60.000.050.100.150.20xcos\nPT\nPT filtered\n|U|\n2cosx, LTE\n(a)(b)\n(c) (d)\nFIG. 3. The numerical comparison of the LTEs and the origi-\nnal equations solved by PTracker in the RR dominant regime.\n\u0001\b =\u0019=2 andSx> Sy, thus the major axis of the particle\ntrajectory ellipse lays on the xaxis. (a)R=Sy=Sxdecreases\nwith time due to Eq. (34), thus the ellipse is getting thinner.\n(b)\rincreases due to the acceleration \feld. (c) and (d) show\nthe same treatments as in Fig. 1 (c) and (d).\n0 1 2 3 4\nt 1e70510152025Sx, SySx, PT\nSx, LTE\nSy, PT\nSy, LTE\n0 1 2 3 4\nt 1e70.500.751.001.251.501.75\n1e6\nPT\nLTE\n1 2 3 4\nt 1e72.0\n1.5\n1.0\n0.5\n0.02xy/|U||V|\nPT\nPT filtered\ncos, LTE\n1 2 3 4\nt 1e70.00.51.01.52xcos/|U|\nPT\nPT filtered\ncosx, LTE\n(a) (b)\n(c) (d)\nFIG. 4. The numerical comparison of the LTEs and the origi-\nnal equations solved by PTracker in the RR dominant regime.\nSx=Sy, thus the particle trajectory in the x-yplane is an\noblique ellipse. (a) SxandSydecrease with the same rate.\n(b)\rinitially decreases with time because the longitudinal\nRR damping is stronger than the acceleration. Later the RR\ndamping becomes weaker due to the decrease of SxandSy,\nand\rincreases. (c) and (d) show the same treatments as in\nFig. 1 (c) and (d), but the oscillation amplitudes are divided\nso that the changes of \u0001\b and \b xare clearer. \u0001\b is between\n\u0019=2 and\u0019, thus \u0001\b gradually approaches \u0019.TABLE I. The cases for comparing PT and LTEs\nCaseParameters Initial Values\nEz\u0015\u0014re\rwjUjjVj\bx\byh\rih\u0010i\nFig. 1\u0015\u00101\n41p\n20 141.12 0.87 0\u0019\n6102-0.05\nFig. 2 -0.001 01p\n210\u0000101041.12 0.87 0\u0019\n61030\nFig. 3\u0015\u00101\n21p\n210\u0000101040.20.18 0\u0019\n2103-0.1\nFig. 4 -0.1 01p\n210\u0000101040.2 0.2\u0019\n4\u00191060\nlong-term behavior of an electron in laser or beam driven\nplasma wake\feld. The original equations of motion have\nbeen expressed by the transverse oscillation terms as\nEqs. (7) - (10), and then averaged in one betatron pe-\nriod to obtain the long-term equations Eqs. (29) - (33).\nThe conditions of our model are r2\u001c\r,r2\r\u001d1 and\nr\rre=2\u000b\u001c1, as discussed in Appx. C. Our model, on\none hand, reproduces previous results such as longitu-\ndinal deceleration and transverse damping, and on the\nother hand reveals new phenomena such as longitudinal\nphase drift oscillation, betatron phase shift and betatron\npolarization change. Two regimes with distinct behav-\niors, determined by re\r5=2, are discussed in Sec. IV, and\nare demonstrated by numerical methods in Sec. V. The\nnumerical comparisons of the long-term equations and\nthe original equations of motion show the high accuracy\nof our model. This model can be fundamental for future\nplasma based high-energy accelerators and colliders [26].\nACKNOWLEDGMENTS\nMZ greatly appreciates the fruitful discussion on the\naveraging method with Igor Kostyukov and Anton Golo-\nvanov from Institute of Applied Physics RAS, Russia.\nThis work is supported by Research Foundation of Insti-\ntute of High Energy Physics, Chinese Academy of Sci-\nences (Grant Nos. E05153U1, E15453U2).\nAppendix A: Plasma normalization units\nThe plasma normalization units are used throughout\nthe paper, as listed in Tab. II, where cis the speed of\nlight in vacuum, !pis the plasma frequency, eis the\nelementary charge, and meis the electron mass. For\nexample, the time is normalized to !\u00001\np, means any time\nrelated quantity such as tin this paper actually means\n!ptin the unnormalized form.7\nTABLE II. The plasma normalization units\nPhysical quantities Variables Normalization units\ntime t !\u00001\np\nfrequency ! !p\nlength x;y;z;re c=!p\nvelocity v c\nmomentum p mec\nangular momentum L mec2=!p\nelectric \feld E mec!p=e\nmagnetic \feld Bme!p=e(in SI)\nforce f mec!p\nAppendix B: Symbols and rules\nIf any variable X, either real or complex, can be ex-\npressed asX=hXi+X1, wherehimeans taking average\nin the betatron period time scale, and X1is the BO term,\ntaking average and derivative can permute\n\u001cd\ndtX\u001d\n=d\ndthXi: (B1)\nWe use a dot on the top to express the time derivative if\nthere is no ambiguity. We have the order-of-magnitude\nestimation\n_X1\u0018!\fX1\u0018\r\u00001\n2X1: (B2)\nIfjX1j\u001cjhXij, for any power \u000bwe have\nhX\u000bi=hXi\u000b\"\n1 +O \nX2\n1\nhXi2!#\n: (B3)\nAnd if another variable Y=hYi+Y1also hasjY1j\u001c\njhYij,\nhXYi=hXihYi\u0014\n1 +O\u0012X1Y1\nhXihYi\u0013\u0015\n: (B4)\nIfXis a complex, taking average and modulus can per-\nmute\nhjXji=jhXij\"\n1 +O \njX1j2\njhXij2!#\n: (B5)\nHowever, taking modulus and derivative cannot permute.\nAppendix C: Equations of motion expressed by\ntransverse oscillations\nIn Eqs. (4) - (6), the longitudinal and transverse os-\ncillations are coupled. As shown in the following, thelongitudinal variables \u00101and\fzare dependent variables\nwhich can be expressed by the transverse ones.\nWe treat~fradas a perturbation and omit it \frst. On\none hand we have\n\r\u00002= 1\u0000\f2\nz\u0000\f2\nx\u0000\f2\ny\n=\r\u00002\nz0\u00002\fz0_\u00101\u0000\f2\nx\u0000\f2\ny+O\u0010\n_\u001012\u0011\n;(C1)\nwhere\rz0=\u0000\n1\u0000\f2\nz0\u0001\u00001=2. By taking average we get\n\r\u00002\nz0\u0019h\ri\u00002+\n\f2\nx\u000b\n+\n\f2\ny\u000b\n: (C2)\nWrite\r=h\ri+\r1in the form\n\r\u00002=h\ri\u00002\"\n1\u00002\r1\nh\ri+O \n\r2\n1\nh\ri2!#\n; (C3)\nwe have\n\r1\u0019\"\n\f2\nx\u0000\n\f2\nx\u000b\n2+\f2\ny\u0000\n\f2\ny\u000b\n2+\fz0_\u00101#\nh\ri3:(C4)\nOn the other hand,\n_\r=\u0000Ez0\fz0\u0000\u0015\fz0\u00101\u0000\u00142(1\u0000\u0015) (x\fx+y\fy) (C5)\nby applying _ \r=\u0000~\f\u0001~Eand Eqs. (1) and (2), or\n_\r1=\u0000\u0015\fz0\u00101\u0000\u00142(1\u0000\u0015) (x\fx+y\fy): (C6)\nNote Eq. (B2), Eq. (C4) seams incompatible with\nEq. (C6), unless\n_\u00101=\u0000\f2\nx\u0000\n\f2\nx\u000b\n2\u0000\f2\ny\u0000\n\f2\ny\u000b\n2; (C7)\nwhich leads to\n\u00101=\u0000x\fx+y\fy\n4; (C8)\nwhich is a general form of Eq. (18) in Ref. [22]. Then\n1\u0000\fz= 1\u0000\fz0\u0000_\u00101=1\n2\u0010\nh\ri\u00002+\f2\nx+\f2\ny\u0011\n;(C9)\nand the formulas of 3D BO with negligible RR are\n_\r=\u0000Ez0\fz0+\u0012\u0015\fz0\n4+\u00142\u0015\u0000\u00142\u0013\n(x\fx+y\fy);\n(C10)\n_pz=\u0000Ez0+\u0015\u00121\n4+\u00142\u0013\n(x\fx+y\fy); (C11)\n_px=\u0000\u00142x+\u00142\u0015\n2\u0010\nh\ri\u00002+\f2\nx+\f2\ny\u0011\nx; (C12)\n_py=\u0000\u00142y+\u00142\u0015\n2\u0010\nh\ri\u00002+\f2\nx+\f2\ny\u0011\ny: (C13)8\nFrom Eq. (C10) we may write\n\r=h\ri+\u0012\u0015\fz0\n4+\u00142\u0015\u0000\u00142\u0013x2\u0000\nx2\u000b\n+y2\u0000\ny2\u000b\n2;\n(C14)\nindicating the prerequisite of the above derivation, which\nhas used Eq. (B3), is r2\u001ch\ri.\nNow we consider RR as a perturbation. The LAD\nequation for the RR four-force is [11]\nFrad\n\u0016=2\n3re\u0014d2P\u0016\nd\u001c2+\u0012dP\u0017\nd\u001cdP\u0017\nd\u001c\u0013\nP\u0016\u0015\n: (C15)with the metric (1 ;\u00001;\u00001;\u00001), wherereis the classical\nelectron radius (also normalized to k\u00001\np),P\u0016is the four-\nmomentum, and \u001cis the proper time ( d\u001c=dt=\r). Use\nEqs. (C10) - (C13) we can verify\ndP\u0017\nd\u001cdP\u0017\nd\u001c=\r2\u0012\n_\r2\u0000\f\f\f_~ p\f\f\f2\u0013\n\u0019\u0000\r2\u0000\n_px2+ _py2\u0001\n(C16)\nas long as r2\r2\u001d1. We can also prove that the \frst\nterm in Eq. (C15) is negligible compared with the second\nterm as long as r2\r\u001d1. Finally the equations of motion\nexpressed by the transverse oscillations are obtained as\nEqs. (7) - (10). As it has been discussed in Ref. [22] and\n[23], this classical RR model is valid as long as r\rre=2\u000b\u001c\n1, where\u000bis the \fne structure constant.\n[1] T. Tajima and J. M. Dawson, Laser electron accelerator,\nPhys. Rev. Lett. 43, 267 (1979).\n[2] P. Chen, J. M. Dawson, R. W. Hu\u000b, and T. Katsouleas,\nAcceleration of electrons by the interaction of a bunched\nelectron beam with a plasma, Phys. Rev. Lett. 54, 693\n(1985).\n[3] I. Blumenfeld, C. E. Clayton, F.-J. Decker, M. J. Hogan,\nC. Huang, R. Ischebeck, R. Iverson, C. Joshi, T. Kat-\nsouleas, N. Kirby, W. Lu, K. A. Marsh, W. B. Mori,\nP. Muggli, E. Oz, R. H. Siemann, D. Walz, and M. Zhou,\nEnergy doubling of 42 gev electrons in a metre-scale\nplasma wake\feld accelerator, Nature 445, 10.1038/na-\nture05538 (2007).\n[4] J. B. Rosenzweig, B. Breizman, T. Katsouleas, and\nJ. J. Su, Acceleration and focusing of electrons in two-\ndimensional nonlinear plasma wake \felds, Phys. Rev. A\n44, R6189 (1991).\n[5] A. Pukhov and J. Meyer-ter Vehn, Laser wake \feld accel-\neration: the highly non-linear broken-wave regime, Ap-\nplied Physics B 74, 10.1007/s003400200795 (2002).\n[6] I. Y. Kostyukov, A. Pukhov, and S. Kiselev, Phenomeno-\nlogical theory of laser-plasma interaction in \\bubble\"\nregime, Physics of Plasmas 11, 5256 (2004).\n[7] W. Lu, C. Huang, M. Zhou, M. Tzoufras, F. S.\nTsung, W. B. Mori, and T. Katsouleas, A non-\nlinear theory for multidimensional relativistic plasma\nwave wake\felds, Physics of Plasmas 13, 056709 (2006),\nhttps://doi.org/10.1063/1.2203364.\n[8] E. Esarey, B. A. Shadwick, P. Catravas, and W. P.\nLeemans, Synchrotron radiation from electron beams\nin plasma-focusing channels, Phys. Rev. E 65, 056505\n(2002).\n[9] A. Rousse, K. T. Phuoc, R. Shah, A. Pukhov, E. Lefeb-\nvre, V. Malka, S. Kiselev, F. Burgy, J.-P. Rousseau,\nD. Umstadter, and D. Hulin, Production of a kev x-\nray beam from synchrotron radiation in relativistic laser-\nplasma interaction, Phys. Rev. Lett. 93, 135005 (2004).\n[10] S. Corde, K. Ta Phuoc, G. Lambert, R. Fitour, V. Malka,\nA. Rousse, A. Beck, and E. Lefebvre, Femtosecond x rays\nfrom laser-plasma accelerators, Rev. Mod. Phys. 85, 1\n(2013).\n[11] P. A. M. Dirac, Classical theory of radiating electrons,\nProc. R. Soc. Lond. A 167, 148 (1938).[12] L. D. Landau and E. M. Lifshitz, Chapter 9 - radiation of\nelectromagnetic waves, in The Classical Theory of Fields\n(Fourth Edition) , Course of Theoretical Physics, Vol. 2,\nedited by L. D. Landau and E. M. Lifshitz (Pergamon,\nAmsterdam, 1975) fourth edition ed., pp. 171{225.\n[13] J. M. Cole, K. T. Behm, E. Gerstmayr, T. G. Black-\nburn, J. C. Wood, C. D. Baird, M. J. Du\u000b, C. Harvey,\nA. Ilderton, A. S. Joglekar, K. Krushelnick, S. Kuschel,\nM. Marklund, P. McKenna, C. D. Murphy, K. Poder,\nC. P. Ridgers, G. M. Samarin, G. Sarri, D. R. Symes,\nA. G. R. Thomas, J. Warwick, M. Zepf, Z. Najmudin,\nand S. P. D. Mangles, Experimental evidence of radi-\nation reaction in the collision of a high-intensity laser\npulse with a laser-wake\feld accelerated electron beam,\nPhys. Rev. X 8, 011020 (2018).\n[14] K. Poder, M. Tamburini, G. Sarri, A. Di Piazza,\nS. Kuschel, C. D. Baird, K. Behm, S. Bohlen, J. M.\nCole, D. J. Corvan, M. Du\u000b, E. Gerstmayr, C. H. Keitel,\nK. Krushelnick, S. P. D. Mangles, P. McKenna, C. D.\nMurphy, Z. Najmudin, C. P. Ridgers, G. M. Samarin,\nD. R. Symes, A. G. R. Thomas, J. Warwick, and M. Zepf,\nExperimental signatures of the quantum nature of radi-\nation reaction in the \feld of an ultraintense laser, Phys.\nRev. X 8, 031004 (2018).\n[15] C. F. Nielsen, J. B. Justesen, A. H. S\u001crensen, U. I. Ug-\ngerh\u001cj, R. Holtzapple, and C. NA63, Experimental ver-\ni\fcation of the landau{lifshitz equation, New Journal of\nPhysics 23, 085001 (2021).\n[16] P. Michel, C. B. Schroeder, B. A. Shadwick, E. Esarey,\nand W. P. Leemans, Radiative damping and electron\nbeam dynamics in plasma-based accelerators, Phys. Rev.\nE74, 026501 (2006).\n[17] I. Y. Kostyukov, E. N. Nerush, and A. M. Pukhov,\nRadiative losses in plasma accelerators, Journal\nof Experimental and Theoretical Physics 103,\n10.1134/S1063776106110173 (2006).\n[18] C. B. Schroeder, E. Esarey, C. G. R. Geddes,\nC. Benedetti, and W. P. Leemans, Physics considerations\nfor laser-plasma linear colliders, Phys. Rev. ST Accel.\nBeams 13, 101301 (2010).\n[19] A. Deng, K. Nakajima, J. Liu, B. Shen, X. Zhang, Y. Yu,\nW. Li, R. Li, and Z. Xu, Electron beam dynamics and\nself-cooling up to pev level due to betatron radiation in9\nplasma-based accelerators, Phys. Rev. ST Accel. Beams\n15, 081303 (2012).\n[20] I. Y. Kostyukov, E. N. Nerush, and A. G. Litvak, Radia-\ntive damping in plasma-based accelerators, Phys. Rev.\nST Accel. Beams 15, 111001 (2012).\n[21] A. Ferran Pousa, A. Martinez de la Ossa, and R. W. Ass-\nmann, Intrinsic energy spread and bunch length growth\nin plasma-based accelerators due to betatron motion, Sci-\nenti\fc Reports 9, 10.1038/s41598-019-53887-8 (2019).\n[22] M. Zeng and K. Seto, Radiation reaction of betatron os-\ncillation in plasma wake\feld accelerators, New J. Phys.\n23, 075008 (2021).[23] A. A. Golovanov, E. N. Nerush, and I. Y. Kostyukov,\nRadiation reaction-dominated regime of wake\feld accel-\neration, New Journal of Physics 24, 033011 (2022).\n[24] https://docs.scipy.org/doc/scipy/reference/\ngenerated/scipy.integrate.BDF.html (2022), the\ndocumentation of BDF integration in SciPy.\n[25] M. Zeng, https://github.com/mingzeng7/PTracker\n(2022), the git repository for the code PTracker.\n[26] W. Leemans and E. Esarey, Laser-driven plasma-wave\nelectron accelerators, Physics Today 62, 44 (2009)." }, { "title": "1907.01045v1.Magnon_decay_theory_of_Gilbert_damping_in_metallic_antiferromagnets.pdf", "content": "Magnon decay theory of Gilbert damping in metallic antiferromagnets\nHaakon T. Simensen, Akashdeep Kamra, Roberto E. Troncoso, and Arne Brataas\nCenter for Quantum Spintronics, Department of Physics,\nNorwegian University of Science and Technology, NO-7491 Trondheim, Norway\n(Dated: July 3, 2019)\nGilbert damping is a key property governing magnetization dynamics in ordered magnets. We present a\ntheoretical study of intrinsic Gilbert damping induced by magnon decay in antiferromagnetic metals through\ns-dexchange interaction. Our theory delineates the qualitative features of damping in metallic antiferromagnets\nowing to their bipartite nature, in addition to providing analytic expressions for the damping parameters. Magnon-\ninduced intraband electron scattering is found to predominantly cause magnetization damping, whereas the Néel\nfield is found to be damped via disorder. Depending on the conduction electron band structure, we predict that\nmagnon-induced interband electron scattering around band crossings may be exploited to engineer a strong Néel\nfield damping.\nIntroduction.— The dynamical properties of a harmonic\nmode are captured by its frequency and lifetime [ 1,2]. While\nthe eigenfrequency is typically determined by the linearized\nequations of motion, or equivalently by a non-interacting de-\nscription of the corresponding quantum excitation, the lifetime\nembodies rich physics stemming from its interaction with one\nor more dissipative baths [ 1,3]. Dissipation plays a central\nrole in the system response time. In the context of magnetic\nsystems employed as memories, the switching times decrease\nwith increasing damping thereby requiring a stronger dissi-\npation for fast operation [ 4–6]. The dissipative properties of\nthe system also result in rich phenomena such as quantum\nphase transitions [ 7–10]. Furthermore, the formation of hybrid\nexcitations, such as magnon-polarons [ 11–18] and magnon-\npolaritons [ 19–24], requires the dissipation to be weak with\nrespect to the coupling strengths between the two participating\nexcitations [ 25]. Therefore, in several physical phenomena that\nhave emerged into focus in the recent years [ 12,16,26–30],\ndamping not only determines the system response but also the\nvery nature of the eigenmodes themselves. Understanding,\nexploiting and controlling the damping in magnets is thus a\nfoundational pillar of the field.\nThe success of Landau-Lifshitz-Gilbert (LLG) phenomenol-\nogy [ 31,32] in describing ferromagnetic dynamics has inspired\nvigorous e \u000borts towards obtaining the Gilbert damping param-\neter using a wide range of microscopic theories. The quantum\nparticles corresponding to magnetization dynamics - magnons\n- provide one such avenue for microscopic theories and form\nthe central theme in the field of magnonics [ 33,34]. While\na vast amount of fruitful research has provided a good under-\nstanding of ferromagnets (FMs) [ 35–54], analogous studies on\nantiferromagnets (AFMs) are relatively scarce and have just\nstarted appearing [ 55,56] due to the recently invigorated field\nof antiferromagnetic spintronics [ 57–62]. Among the ongoing\ndiscoveries of niches borne by AFMs, from electrically and\nrapidly switchable memories [ 63], topological spintronics [ 60],\nlong range magnonic transport [ 64] to quantum fluctuations\n[65], an unexpected surprise has been encountered in the first\nprinciples evaluation of damping in metallic AFMs. Liu and\ncoworkers [ 56] and another more recent first-principles study\n[66] both found the magnetization dissipation parameter to bemuch larger than the corresponding Néel damping constant,\nin stark contrast with previous assumptions, exhibiting richer\nfeatures than in FMs. An understanding of this qualitative\ndi\u000berence as well as the general AFM dissipation is crucial\nfor the rapidly growing applications and fundamental novel\nphenomena based on AFMs.\nHere, we accomplish an intuitive and general understanding\nof the Gilbert damping in metallic AFMs based on the magnon\npicture of AFM dynamics. Employing the s-d, two-sublattice\nmodel for a metallic AFM, in which the dandselectrons\nconstitute the magnetic and conduction subsystems, we derive\nanalytic expressions for the Gilbert damping parameters as\na function of the conduction electron density of states at the\nFermi energy and s-dexchange strength. The presence of spin-\ndegenerate conduction bands in AFMs is found to be the key\nin their qualitatively di \u000berent damping properties as compared\nto FMs. This allows for absorption of AFM magnons via s-\ndexchange-mediated intraband conduction electron spin-flip\nprocesses leading to strong damping of the magnetization as\ncompared to the Néel field [ 67]. We also show that interband\nspin-flip processes, which are forbidden in our simple AFM\nmodel but possible in AFMs with band crossings in the conduc-\ntion electron dispersion, result in a strong Néel field damping.\nThus, the general qualitative features of damping in metallic\nAFMs demonstrated herein allow us to understand the Gilbert\ndamping given the conduction electron band structure. These\ninsights provide guidance for engineering AFMs with desired\ndamping properties, which depend on the exact role of the\nAFM in a device.\nModel.— We consider two-sublattice metallic AFMs within\nthes-dmodel [ 35,36,44]. The delectrons localized at lat-\ntices sites constitute the magnetic subsystem responsible for\nantiferromagnetism, while the itinerant selectrons form the\nconduction subsystem that accounts for the metallic traits. The\ntwo subsystems interact via s-dexchange [Eq. (3)]. For ease of\ndepiction and enabling an understanding of qualitative trends,\nwe here consider a one-dimensional AFM (Fig. 1). The re-\nsults within this simple model are generalized to AFMs with\nany dimensionality in a straightforward manner. Furthermore,\nwe primarily focus on the uniform magnetization dynamics\nmodes.arXiv:1907.01045v1 [cond-mat.mes-hall] 1 Jul 20192\nFIG. 1: Schematic depiction of our model for a metallic AFM.\nThe red and blue arrows represent the localized delectrons\nwith spin up and down, respectively, thereby constituting the\nNéel ordered magnetic subsystem. The green cloud illustrates\nthe delocalized, itinerant selectrons that forms the conduction\nsubsystem.\nAt each lattice site i, there is a localized delectron with spin\nSi. The ensuing magnetic subsystem is antiferromagnetically\nordered (Fig. 1), and the quantized excitations are magnons\n[68,69]. Disregarding applied fields for simplicity and as-\nsuming an easy-axis anisotropy along the z-axis, the magnetic\nHamiltonian, Hm=˜JP\nhi;jiSi\u0001Sj\u0000KP\ni(Sz\ni)2, wherehi;ji\ndenotes summation over nearest neighbor lattice sites, is quan-\ntized and mapped to the sublattice-magnon basis [69]\nHm=X\nqh\nAq\u0010\nay\nqaq+by\nqbq\u0011\n+By\nqay\nqby\nq+Bqaqbqi\n; (1)\nwhere we substitute ~=1,Aq=(2˜J+2K)SandBq=\n˜JS e\u0000iq\u0001aP\nh\u000eieiq\u0001\u000e, where S=jSij,ais the displacement\nbetween the two atoms in the basis, and h\u000eidenotes sum-\nming over nearest neighbor displacement vectors. aqandbq\nare bosonic annihilation operators for plane wave magnons\non the A and B sublattices, respectively. We diagonalize\nthe Hamiltonian [Eq. 1] through a Bogoliubov transforma-\ntion [ 69] toHm=P\nq!q\u0010\n\u000by\nq\u000bq+\fy\nq\fq\u0011\n;with eigenenergies\n!q=q\nA2q\u0000jBqj2. In the absence of an applied field, the\nmagnon modes are degenerate.\nTheselectron conduction subsystem is described by a tight-\nbinding Hamiltonian that includes the “static” contribution\nfrom the s-dexchange interaction [Eq. (3)] discussed below:\nHe=\u0000tX\nhi;jiX\n\u001bcy\ni\u001bcj\u001b\u0000JX\ni(\u00001)i\u0010\ncy\ni\"ci\"\u0000cy\ni#ci#\u0011\n:(2)\nHere ci\u001bis the annihilation operator for an selectron at site\niwith spin\u001b.t(>0)is the hopping parameter, and J(>0)\naccounts for s-dexchange interaction [Eq. (3)]. The (\u00001)i\nfactor in the exchange term reflects the two-sublattice nature of\nthe AFM. The conduction subsystem unit cell consists of two\nbasis atoms, similar to the magnetic subsystem. As a result,\nthere are four distinct electron bands: two due to there being\ntwo basis atoms per unit cell, and twice this due to the two\npossible spin polarizations per electron. Disregarding applied\nfields, these bands constitute two spin-degenerate bands. We\nlabel these bands 1 and 2, where the latter is higher in energy.The itinerant electron Hamiltonian [Eq. (2)] is diagonalized\ninto an eigenbasis (c1k\u001b;c2k\u001b)with eigenenergies \u000f1k=\u0000\u000fk\nand\u000f2k= +\u000fk, where\u000fk=p\nJ2S2+t2j\rkj2, where\rk=P\nh\u000eie\u0000ik\u0001\u000e. The itinerant electron dispersion is depicted in Fig.\n2.\nThe magnetic and conduction subsystems interact through\ns-dexchange interaction, parametrized by J:\nHI=\u0000JX\niSi\u0001si; (3)\nwhere si=P\n\u001b\u001b0cy\ni\u001b\u001b\u001b\u001b0ci\u001b0is the spin of the itinerant elec-\ntrons at site i, where \u001bis the vector of Pauli matrices. The term\nwhich is zeroth order in the magnon operators, and thus ac-\ncounts for the static magnetic texture, is already included in He\n[Eq. (2)]. To first order in magnon operators, the interaction\nHamiltonian can be compactly written as\nHe\u0000m=X\n\u0015\u001aX\nkk0qcy\n\u0015k\"c\u001ak0#\u0010\nWA;\u0015\u001a\nkk0qay\n\u0000q+WB;\u0015\u001a\nkk0qbq\u0011\n+h.c.;(4)\nwhere\u0015and\u001aare summed over the electron band indices. As\ndetailed in the Supplemental material, WA;\u0015\u001a\nkk0qandWB;\u0015\u001a\nkk0q, both\nlinear in J, are coe \u000ecients determining the amplitudes for\nscattering between the itinerant electrons and the aqandbq\nmagnons, respectively. Specifically, when considering plane\nwave states, WA=B;\u0015\u001a\nkk0qbecomes a delta function, thereby enforc-\ning the conservation of crystal momentum in a translationally\ninvariant lattice. Inclusion of disorder or other many-body\ne\u000bects results in deviation of the eigenstates from ideal plane\nwaves causing a wave vector spread around its mean value [ 2].\nThe delta function, associated with an exact crystal momentum\nconservation, is thus transformed to a peaked function with\nfinite width (\u0001k). The\u0015\u001acombinations 11and22describe\nintraband electron scattering, while 12and21describe in-\nterband scattering. Intraband scattering is illustrated in Fig.\n2. Interband scattering is prohibited within our model due to\nenergy conservation, since the uniform q=0magnon energy\nis much smaller than the band gap.\nThe scattering described by He\u0000m[Eq. (4)] transfers spin\nangular momentum between the magnetic and conduction sub-\nsystems. The itinerant electrons are assumed to maintain a\nthermal distribution thereby acting as a perfect spin sink. This\nis consistent with a strong conduction electron spin relaxation\nobserved in metallic AFMs [ 70,71]. As a result, the magnetic\nsubsystem spin is e \u000bectively damped through the s-dexchange\ninteraction.\nGilbert damping.— In the Landau-Lifshitz-Gilbert (LLG)\nphenomenology for two-sublattice AFMs, dissipation is ac-\ncounted via a 2\u00022 Gilbert damping matrix [ 72]. Our goal here\nis to determine the elements of this matrix in terms of the\nparameters and physical observables within our microscopic\nmodel. To this end, we evaluate the spin current “pumped”\nby the magnetic subsystem into the sconduction electrons,\nwhich dissipate it immediately within our model. The angu-\nlar momentum thus lost by the magnetic subsystem appears\nas Gilbert damping in its dynamical equations [ 72,73]. The3\n-\n/2 -\n /4 0\n /2\n /4\ne\ne\nkF,1a/epsilon1=µ1\nm\ne\ne\nm\n/epsilon1=µ2\nkF,2a\nFIG. 2: The selectron dispersion in metallic AFM model,\nwhere the red and blue dispersions depict electron bands 1 and\n2, respectively. Illustrations of intraband electron-magnon\nscattering at two di \u000berent Fermi levels, \u00161and\u00162, are added.\nThe depicted momentum transfer is exaggerated for clarity.\nsecond essential ingredient in identifying the Gilbert damping\nmatrix from our microscopic theory is the idea of coherent\nstates [ 74,75]. The classical LLG description of the magne-\ntization is necessarily equivalent to our quantum formalism,\nwhen the magnetic eigenmode is in a coherent state [ 74–76].\nDriving the magnetization dynamics via a microwave field,\nsuch as in the case of ferromagnetic resonance experiments,\nachieves such a coherent magnetization dynamics [73, 77].\nThe spin current pumped by a two-sublattice magnetic sys-\ntem into an electronic bath may be expressed as [78]\nIz=Gmm(m\u0002˙m)z+Gnn(n\u0002˙n)z\n+Gmn\u0002(m\u0002˙n)z+(n\u0002˙m)z\u0003;(5)\nwhere mand nare the magnetization and Néel field nor-\nmalized by the sublattice magnetization, respectively. Here,\nGi j=\u000bi j\u0002(M=j\rj), where\u000bi jare the Gilbert damping co-\ne\u000ecients,\ris the gyromagnetic ratio of the delectrons\nandMis the sublattice magnetization. Considering the uni-\nform magnetization mode, Izis the spin current operator\nIz=i[He\u0000m;Sz][79], where Sz=P\niSz\ni. We get\nIz=iX\n\u0015\u001aX\nkk0qcy\n\u0015k\"c\u001ak0#\u0010\nWA;\u0015\u001a\nkk0qay\n\u0000q+WB;\u0015\u001a\nkk0qbq\u0011\n\u0000h.c.:(6)\nThe expectation value of this operator assuming the uniform\nmagnetization mode to be in a coherent state corresponds to\nthe spin pumping current [Eq. (5)].\nIn order to evaluate the spin pumping current from Eq. (6),\nwe follow the method employed to calculate interfacial spin\npumping current into normal metals in Refs. [ 73,77,78], and\nthe procedure is described in detail therein. Briefly, this method\nentails assuming the magnetic and conduction subsystems to\nbe independent and in equilibrium at t=\u00001, when the mu-\ntual interaction [Eq. (4)] is turned on. The subsequent timeevolution of the coupled system allows evaluating its physical\nobservables in steady state. The resulting coherent spin-current\ncorresponds to the classical spin current Izthat can be related\nto the motion of the magnetization and the Néel field [Eq. (5)].\nAs a last step, we identify expressions for (m\u0002˙m)z,(m\u0002˙n)z\nand(n\u0002˙n)zin terms of coherent magnon states, which enables\nus to identify the Gilbert damping coe \u000ecients\u000bmm,\u000bnnand\n\u000bmn.\nResults.— Relegating the detailed evaluation to Supplemen-\ntal Material, we now present the analytic expression obtained\nfor the various coe \u000ecients [Eq. (5)]. A key assumption that\nallows these simple expressions is that the electronic density of\nstates in the conduction subsystem does not vary significantly\nover the magnon energy scale. Furthermore, we account for a\nweak disorder phenomenologically via a finite scattering length\nlassociated with the conduction electrons. This results in an\ne\u000bective broadening of the electron wavevectors determined by\nthe inverse electron scattering length, (\u0001k)=2\u0019=l. As a result,\nthe crystal momentum conservation in the system is enforced\nonly within the wavevector broadening. By weak disorder we\nmean that the electron scattering length is much larger than\nthe lattice parameter a. Ifkandk0are the wave vectors of the\nincoming and outgoing electrons, respectively, we then have\n(k\u0000k0)a=(\u0001k)a\u001c1. This justifies an expansion in the wave\nvector broadening (\u0001k)a. The Gilbert damping coe \u000ecients\nstemming from intraband electron scattering are found to be\n\u000bmm=\u000b0(\u0018J)\u0000\u000b0(\u0018J)\n40BBBBBBBB@1+\u00182\nJ\u0010\n\u00182\nJ+8\u00004 cos2(kFa)\u0011\n\u0010\n\u00182\nJ+4 cos2(kFa)\u001121CCCCCCCCA[(\u0001k)a]2;\n\u000bnn=\u000b0(\u0018J)\n40BBBBBB@1+sin2(kFa)\ncos2(kFa)\u00182\nJ\u0010\n\u00182\nJ+4 cos2(kFa)\u00111CCCCCCA[(\u0001k)a]2:\n(7)\nwhere\u0018J=JS=t,kFis the Fermi momentum and ais the lattice\nparameter, and where\n\u000b0(\u0018J)=\u0019v2J2\n8g2(\u0016)j˜Vj24 cos2(kFa)\n\u00182\nJ+4 cos2(kFa): (8)\nHere, vis the unit cell volume, g(\u000f)is the density of states\nper unit volume, \u0016is the Fermi level, and !0is the energy of\ntheq=0magnon mode. ˜Vis a dimensionless and generally\ncomplex function introduced to account for the momentum\nbroadening dependency of the scattering amplitudes. It satisfies\n˜V(0)=1and0\u0014j˜V(\u0001k)j\u00141within our model. These analytic\nexpressions for the Gilbert damping parameters constitute one\nof the main results of this letter.\nDiscussion.– We straightaway note that \u000bnn=\u000bmm\u0018\n[(\u0001k)a]2\u001c1.\u000bnnis strictly dependent upon (\u0001k)a, and is non-\nzero only if there is disorder and a finite electron momentum\nbroadening. \u000bmmis large even when considering a perfectly\nordered crystal. This latter result is in good accordance with\nrecent first-principles calculations in metallic AFMs [ 56,66].\nWe moreover observe that both \u000bmmand\u000bnnare quadratic\ninJandg(\u0016). This result is shared by Gilbert damping ow-\ning to spin-pumping in insulating ferrimagnet |normal metal4\ne\n e\nm\nkFa/epsilon1=µ\nFIG. 3: A schematic depiction of magnon-induced interband\nscattering in a band crossing at the Fermi level.\n(NM) and AFM |NM bilayers with interfacial exchange cou-\npling [ 78]. Metallic AFMs bear a close resemblance to these\nbilayer structures. There are however two main di \u000berences:\nThes-dexchange coupling exists in the bulk of metallic AFMs,\nwhereas it is localized at the interface in the bilayer structures.\nAdditionally, the itinerant electron wave functions are qual-\nitatively di \u000berent in metallic AFMs and NMs, owing to the\nmagnetic unit cell of the AFM. Indeed, these di \u000berences turn\nout to leave prominent signatures in the Gilbert damping in\nmetallic AFMs.\nThe uniform mode magnon energy is much smaller than the\nelectron band gap within our simple model. Interband scat-\ntering is thus prohibited by energy conservation. However,\nin real AFM metals, the electron band structure is more com-\nplex. There may for instance exist band crossings [ 80–82].\nIn such materials, magnon-induced interband electron scatter-\ning should also contribute to Gilbert damping, depending on\nthe position of the Fermi surface. Motivated by this, we now\nconsider Gilbert damping stemming from interband scattering,\nwhile disregarding the energy conservation for the moment,\nlabeling the coe \u000ecients\u000bI\nmmand\u000bI\nnn. We then find the same\nexpressions as in Eq. (7) with the roles of \u000bI\nmm;nninterchanged\nwith respect to \u000bmm;nn. This implies that \u000bI\nnnis large and inde-\npendent of electron momentum broadening, whereas \u000bI\nmmis\nproportional to the electron momentum broadening squared.\nAlthough arriving at this result required disregarding the en-\nergy conservation constraint, the qualitative e \u000bect in itself is\nnot an artifact of this assumption. In materials with a band\ncrossing, as depicted in Fig. 3, \u000bI\nnn=\u000bI\nmm> \u000b nn=\u000bmmis a gen-\neral result. This generic principle derived within our simple\nmodel provides valuable guidance for designing materials with\nan engineered Gilbert damping matrix.\nWe now provide a rough intuitive picture for the damping\ndependencies obtained above followed by a more mathemati-\ncal discussion. Consider a conventional di \u000braction experiment\nwhere an incident probing wave is able to resolve the two\nslits only when the wavelength is comparable to the physical\nseparation between the two slits. In the case at hand, the wave-\nfunctions of electrons and magnon participating in a scatteringprocess combine in a way that the wavenumber by which the\nconservation of crystal momentum is violated becomes the\nprobing wavenumber within a di \u000braction picture. Therefore,\nthe processes conserving crystal momentum have vanishing\nprobing wavenumber and are not able to resolve the opposite\nspins localized at adjacent lattice sites. Therefore, only the aver-\nage magnetization is damped leaving the Néel field una \u000bected.\nWith disorder, the probing wavenumber becomes non-zero and\nthus also couples to the Néel field. The interband scattering,\non the other hand, is reminiscent of Umklapp scattering in a\nsingle-sublattice model and the probing wavenumber matches\nwith the inverse lattice spacing. Therefore, the coupling with\nthe Néel field is strong.\nThe Gilbert damping in metallic AFMs here considered is\ncaused by spin pumping from the magnetic subsystem into\nthesband. This spin pumping induces electron transitions\nbetween spin\"/#states among the selectrons. The Gilbert\ndamping coe \u000ecients depend thus on transition amplitudes pro-\nportional to products of itinerant electron wave functions such\nas y\n\u0015k\"(x) \u001ak0#(x). The damping e \u000bect on sublattice A depends\non this transition amplitude evaluated on the A sublattice, and\nequivalently for the B sublattice. Assuming without loss of gen-\nerality that site i=0belongs to sublattice A, we find in the one-\ndimensional model that the damping on sublattice A is a func-\ntion ofP\njcos2\u0010\u0019xj\n2a\u0011\n y\n\u0015k\"(xj) \u001ak0#(xj), whereas the damping\non sublattice B is a function ofP\njsin2\u0010\u0019xj\n2a\u0011\n y\n\u0015k\"(xj) \u001ak0#(xj).\nEquivalently, by straightforwardly using that m=(mA+mB)=2\nandn=(mA\u0000mB)=2, this analysis predicts that \u000bmmis a\nfunction ofP\nj y\n\u0015k\"(xj) \u001ak0#(x), whereas\u000bnnis a function of\nP\njcos\u0010\u0019xj\na\u0011\n y\n\u0015k\"(xj) \u001ak0#(x). Assuming plane wave solutions\nof the electron wave functions, and if we consider intraband\nscattering only, we more concretely find that \u000bmmis a function\nof(1\u0000i(\u0001k)a), where iis the imaginary unit, whereas \u000bnnis a\nfunction of ( \u0001k)a. This coincides well with Eq. (7).\nAbove, we presented a discussion of interband scattering in\nthe minimal model where the band gap artificially was set to\nzero. In this limit, the upper electron band is a continuation\nof the lower band with a \u0006\u0019=amomentum shift. We may then\nwrite 2k\u001b= 1;k+\u0019=a;\u001b. Under the assumption of a disappear-\ning band gap, momentum-conserving interband scattering at\nmomentum kis therefore equivalent to intraband scattering be-\ntween kandk\u0006\u0019=a. This is the exact phase shift which results\nin a small\u000bmmand a large \u000bnnconsistent with the discussion\nabove. In real metallic AFMs with complex band structures,\nthe exact wave function relations unveiled above do not apply.\nHowever, interband transition amplitudes will undoubtedly\ncarry a position dependent phase. This position dependence\nresults in a dephasing of transition amplitudes at neighboring\nlattice sites, which gives rise to a non-negligible \u000bnn. The pre-\ncise damping coe \u000ecients in real metallic AFMs depend on the\ndetailed electron wave functions. We may however generally\nconclude that \u000bI\nnn=\u000bI\nmm>\u000b nn=\u000bmm.\nConclusion.— We have provided a microscopic derivation\nof Gilbert damping resulting from magnon decay through s-d\nexchange interaction in metallic antiferromagnets. Analytic5\nexpressions for Gilbert damping coe \u000ecients resulting from in-\ntraband electron scattering are presented, while Gilbert damp-\ning resulting from interband electron scattering is discussed on\na conceptual level. We find that intraband electron scattering\ngives rise to a large magnetization damping and a negligible\nNéel field damping. The intraband Néel field damping is pro-\nportional to the inverse electron scattering length squared, and\ndisappears exactly if there is no crystal disorder. By relating\nGilbert damping to the degree to which transition amplitudes\nof the itinerant electron are in phase at neighboring lattice\nsites, we have argued for why interband electron scattering\nmay generate a large Néel field damping.\nAcknowledgments.— This work was supported by the Re-\nsearch Council of Norway through its Centres of Excellence\nfunding scheme, Project No. 262633 \"QuSpin\", the European\nUnion’s Horizon 2020 Research and Innovation Programme\nunder Grant DLV-737038 ”TRANSPIRE”, as well as by the\nEuropean Research Council via Advanced Grant No. 669442\n\"Insulatronics\".\n[1]R. P. Feynman and F. L. Vernon, “The theory of a general quan-\ntum system interacting with a linear dissipative system,” Annals\nof Physics 24, 118 (1963).\n[2]G. D. Mahan, Many-Particle Physics , 3rd ed. (Kluwer Aca-\ndemic /Plenum Publishers, New York, 2000).\n[3]A. O. Caldeira and A. J. Leggett, “Influence of Dissipation on\nQuantum Tunneling in Macroscopic Systems,” Phys. Rev. Lett.\n46, 211 (1981).\n[4]A. V . Kimel, A. Kirilyuk, A. Tsvetkov, R. V . Pisarev, and\nT. Rasing, “Laser-induced ultrafast spin reorientation in the anti-\nferromagnet TmFeO3,” Nature 429, 850 (2004).\n[5]A. Kirilyuk, A. V . Kimel, and T. Rasing, “Ultrafast optical\nmanipulation of magnetic order,” Rev. Mod. Phys. 82, 2731\n(2010).\n[6]A. Brataas, A. D. Kent, and H. Ohno, “Current-induced torques\nin magnetic materials,” Nature Materials 11, 372 (2012).\n[7]R. Fazio and H. van der Zant, “Quantum phase transitions and\nvortex dynamics in superconducting networks,” Physics Reports\n355, 235 (2001).\n[8]A. Biella, L. Mazza, I. Carusotto, D. Rossini, and R. Fazio,\n“Photon transport in a dissipative chain of nonlinear cavities,”\nPhysical Review A 91, 053815 (2015).\n[9]D. Maile, S. Andergassen, W. Belzig, and G. Rastelli, “Quantum\nphase transition with dissipative frustration,” Phys. Rev. B 97,\n155427 (2018).\n[10] G. Rastelli and I. M. Pop, “Tunable ohmic environment using\nJosephson junction chains,” Phys. Rev. B 97, 205429 (2018).\n[11] C. Kittel, “Interaction of Spin Waves and Ultrasonic Waves in\nFerromagnetic Crystals,” Physical Review 110, 836 (1958).\n[12] M. Weiler, H. Huebl, F. S. Goerg, F. D. Czeschka, R. Gross, and\nS. T. B. Goennenwein, “Spin Pumping with Coherent Elastic\nWaves,” Physical Review Letters 108, 176601 (2012).\n[13] A. Rückriegel, P. Kopietz, D. A. Bozhko, A. A. Serga, and\nB. Hillebrands, “Magnetoelastic modes and lifetime of magnons\nin thin yttrium iron garnet films,” Physical Review B 89, 184413\n(2014).\n[14] A. Kamra, H. Keshtgar, P. Yan, and G. E. W. Bauer, “Coherent\nelastic excitation of spin waves,” Physical Review B 91, 104409(2015).\n[15] B. Flebus, K. Shen, T. Kikkawa, K.-i. Uchida, Z. Qiu, E. Saitoh,\nR. A. Duine, and G. E. W. Bauer, “Magnon-polaron transport\nin magnetic insulators,” Physical Review B 95, 144420 (2017).\n[16] T. Kikkawa, K. Shen, B. Flebus, R. A. Duine, K.-i. Uchida,\nZ. Qiu, G. E. W. Bauer, and E. Saitoh, “Magnon Polarons in\nthe Spin Seebeck E \u000bect,” Physical Review Letters 117, 207203\n(2016).\n[17] L. Dreher, M. Weiler, M. Pernpeintner, H. Huebl, R. Gross, M. S.\nBrandt, and S. T. B. Goennenwein, “Surface acoustic wave\ndriven ferromagnetic resonance in nickel thin films: Theory and\nexperiment,” Phys. Rev. B 86, 134415 (2012).\n[18] H. T. Simensen, R. E. Troncoso, A. Kamra, and A. Brataas,\n“Magnon-polarons in cubic collinear antiferromagnets,” Physical\nReview B 99, 64421 (2019).\n[19] H. Huebl, C. W. Zollitsch, J. Lotze, F. Hocke, M. Greifenstein,\nA. Marx, R. Gross, and S. T. B. Goennenwein, “High Coopera-\ntivity in Coupled Microwave Resonator Ferrimagnetic Insulator\nHybrids,” Physical Review Letters 111, 127003 (2013).\n[20] Y . Tabuchi, S. Ishino, A. Noguchi, T. Ishikawa, R. Yamazaki,\nK. Usami, and Y . Nakamura, “Coherent coupling between a\nferromagnetic magnon and a superconducting qubit,” Science\n349, 405 (2015).\n[21] C. . Hu, B. M. Yao, S. Kaur, Y . S. Gui, and W. Lu, “Magnon\npolariton and pseudo-magnon-polariton,” in 2015 40th Interna-\ntional Conference on Infrared, Millimeter, and Terahertz waves\n(IRMMW-THz) (2015) pp. 1–3.\n[22] Ø. Johansen and A. Brataas, “Nonlocal Coupling between Anti-\nferromagnets and Ferromagnets in Cavities,” Physical Review\nLetters 121, 87204 (2018).\n[23] M. Harder and C.-M. Hu, “Cavity Spintronics: An Early Review\nof Recent Progress in the Study of Magnon–Photon Level Re-\npulsion,” in Solid State Physics 69 , edited by R. E. Camley and\nR. L. Stamps (Academic Press, Cambridge, 2018) pp. 47–121.\n[24] Y . Cao, P. Yan, H. Huebl, S. T. B. Goennenwein, and G. E. W.\nBauer, “Exchange magnon-polaritons in microwave cavities,”\nPhys. Rev. B 91, 94423 (2015).\n[25] A. Frisk Kockum, A. Miranowicz, S. De Liberato, S. Savasta,\nand F. Nori, “Ultrastrong coupling between light and matter,”\nNature Reviews Physics 1, 19 (2019).\n[26] J. Holanda, D. S. Maior, A. Azevedo, and S. M. Rezende,\n“Detecting the phonon spin in magnon–phonon conversion exper-\niments,” Nature Physics (2018), 10.1038 /s41567-018-0079-y.\n[27] S. Viola Kusminskiy, H. X. Tang, and F. Marquardt, “Coupled\nspin-light dynamics in cavity optomagnonics,” Physical Review\nA94, 33821 (2016).\n[28] N. Samkharadze, G. Zheng, N. Kalhor, D. Brousse, A. Sammak,\nU. C. Mendes, A. Blais, G. Scappucci, and L. M. K. Vander-\nsypen, “Strong spin-photon coupling in silicon,” Science 359,\n1123 (2018).\n[29] M. Harder, Y . Yang, B. M. Yao, C. H. Yu, J. W. Rao, Y . S.\nGui, R. L. Stamps, and C.-M. Hu, “Level Attraction Due to\nDissipative Magnon-Photon Coupling,” Phys. Rev. Lett. 121,\n137203 (2018).\n[30] P. F. Herskind, A. Dantan, J. P. Marler, M. Albert, and\nM. Drewsen, “Realization of collective strong coupling with\nion Coulomb crystals in an optical cavity,” Nature Physics 5,\n494 (2009).\n[31] L. Landau and E. Lifshitz, “On the theory of the dispersion of\nmagnetic permeability in ferromagnetic bodies,” Phys. Z. Sowjet.\n8(1935).\n[32] T. Gilbert, “Classics in Magnetics A Phenomenological Theory\nof Damping in Ferromagnetic Materials,” IEEE Transactions on\nMagnetics 40, 3443 (2004).6\n[33] V . V . Kruglyak, S. O. Demokritov, and D. Grundler, “Magnon-\nics,” Journal of Physics D: Applied Physics 43, 264001 (2010).\n[34] A. V . Chumak, V . I. Vasyuchka, A. A. Serga, and B. Hillebrands,\n“Magnon spintronics,” Nature Physics 11, 453 (2015).\n[35] A. H. Mitchell, “Ferromagnetic Relaxation by the Exchange\nInteraction between Ferromagnetic Electrons and Conduction\nElectrons,” Physical Review 105, 1439 (1957).\n[36] B. Heinrich, D. Fraitová, and V . Kamberský, “The Influence\nof s-d Exchange on Relaxation of Magnons in Metals,” physica\nstatus solidi (b) 23, 501 (1967).\n[37] V . Kamberský, “On the Landau–Lifshitz relaxation in ferromag-\nnetic metals,” Canadian Journal of Physics 48, 2906 (1970).\n[38] V . Kamberský, “On ferromagnetic resonance damping in metals,”\nCzechoslovak Journal of Physics 26, 1366 (1976).\n[39] J. Kuneš and V . Kamberský, “First-principles investigation of\nthe damping of fast magnetization precession in ferromagnetic\n3d metals,” Physical Review B 65, 212411 (2002).\n[40] A. Y . Dobin and R. H. Victora, “Intrinsic Nonlinear Ferromag-\nnetic Relaxation in Thin Metallic Films,” Physical Review Let-\nters90, 167203 (2003).\n[41] S. Zhang and Z. Li, “Roles of Nonequilibrium Conduction Elec-\ntrons on the Magnetization Dynamics of Ferromagnets,” Physi-\ncal Review Letters 93, 127204 (2004).\n[42] Y . Tserkovnyak, G. A. Fiete, and B. I. Halperin, “Mean-field\nmagnetization relaxation in conducting ferromagnets,” Applied\nPhysics Letters 84, 5234 (2004).\n[43] V . Kamberský, “Spin-orbital Gilbert damping in common mag-\nnetic metals,” Physical Review B 76, 134416 (2007).\n[44] H. J. Skadsem, Y . Tserkovnyak, A. Brataas, and G. E. W. Bauer,\n“Magnetization damping in a local-density approximation,” Phys-\nical Review B 75, 094416 (2007).\n[45] K. Gilmore, Y . U. Idzerda, and M. D. Stiles, “Identification of\nthe Dominant Precession-Damping Mechanism in Fe, Co, and\nNi by First-Principles Calculations,” Physical Review Letters 99,\n027204 (2007).\n[46] A. Brataas, Y . Tserkovnyak, and G. E. W. Bauer, “Scatter-\ning Theory of Gilbert Damping,” Physical Review Letters 101,\n37207 (2008).\n[47] M. C. Hickey and J. S. Moodera, “Origin of Intrinsic Gilbert\nDamping,” Physical Review Letters 102, 137601 (2009).\n[48] A. A. Starikov, P. J. Kelly, A. Brataas, Y . Tserkovnyak, and\nG. E. W. Bauer, “Unified First-Principles Study of Gilbert Damp-\ning, Spin-Flip Di \u000busion, and Resistivity in Transition Metal\nAlloys,” Physical Review Letters 105, 236601 (2010).\n[49] S. Mankovsky, D. Ködderitzsch, G. Woltersdorf, and H. Ebert,\n“First-principles calculation of the Gilbert damping parameter\nvia the linear response formalism with application to magnetic\ntransition metals and alloys,” Physical Review B 87, 014430\n(2013).\n[50] Z. Yuan, K. M. D. Hals, Y . Liu, A. A. Starikov, A. Brataas, and\nP. J. Kelly, “Gilbert Damping in Noncollinear Ferromagnets,”\nPhysical Review Letters 113, 266603 (2014).\n[51] E. Barati, M. Cinal, D. M. Edwards, and A. Umerski, “Gilbert\ndamping in magnetic layered systems,” Physical Review B 90,\n014420 (2014).\n[52] F. Mahfouzi, J. Kim, and N. Kioussis, “Intrinsic damping phe-\nnomena from quantum to classical magnets: An ab initio study\nof Gilbert damping in a Pt /Co bilayer,” Physical Review B 96,\n214421 (2017).\n[53] Y . Zhao, Y . Liu, H. Tang, H. Jiang, Z. Yuan, and K. Xia, “Gilbert\ndamping in FeCo alloy: From weak to strong spin disorder,”\nPhysical Review B 98, 174412 (2018).\n[54] Y . Li, F. Zeng, S. S.-L. Zhang, H. Shin, H. Saglam, V . Karakas,\nO. Ozatay, J. E. Pearson, O. G. Heinonen, Y . Wu, A. Ho \u000bmann,and W. Zhang, “Giant Anisotropy of Gilbert Damping in Epitax-\nial CoFe Films,” Physical Review Letters 122, 117203 (2019).\n[55] T. Moriyama, K. Hayashi, K. Yamada, M. Shima, Y . Ohya, and\nT. Ono, “Intrinsic and extrinsic antiferromagnetic damping in\nNiO,” Physical Review Materials 3, 51402 (2019).\n[56] Q. Liu, H. Y . Yuan, K. Xia, and Z. Yuan, “Mode-dependent\ndamping in metallic antiferromagnets due to intersublattice spin\npumping,” Physical Review Materials 1, 061401 (2017).\n[57] T. Jungwirth, X. Marti, P. Wadley, and J. Wunderlich, “An-\ntiferromagnetic spintronics,” Nature Nanotechnology 11, 231\n(2016).\n[58] V . Baltz, A. Manchon, M. Tsoi, T. Moriyama, T. Ono, and\nY . Tserkovnyak, “Antiferromagnetic spintronics,” Reviews of\nModern Physics 90, 015005 (2018).\n[59] A. H. MacDonald and M. Tsoi, “Antiferromagnetic metal spin-\ntronics,” Philosophical Transactions of the Royal Society A:\nMathematical, Physical and Engineering Sciences 369, 3098\n(2011).\n[60] L. Šmejkal, Y . Mokrousov, B. Yan, and A. H. MacDonald,\n“Topological antiferromagnetic spintronics,” Nature Physics 14,\n242 (2018).\n[61] E. V . Gomonay and V . M. Loktev, “Spintronics of antiferromag-\nnetic systems (Review Article),” Low Temperature Physics 40,\n17 (2014).\n[62] O. Gomonay, V . Baltz, A. Brataas, and Y . Tserkovnyak, “Anti-\nferromagnetic spin textures and dynamics,” Nature Physics 14,\n213 (2018).\n[63] P. Wadley, B. Howells, J. Elezny, C. Andrews, V . Hills,\nR. P. Campion, V . Novak, K. Olejnik, F. Maccherozzi, S. S.\nDhesi, S. Y . Martin, T. Wagner, J. Wunderlich, F. Freimuth,\nY . Mokrousov, J. Kune, J. S. Chauhan, M. J. Grzybowski, A. W.\nRushforth, K. W. Edmonds, B. L. Gallagher, and T. Jungwirth,\n“Electrical switching of an antiferromagnet,” Science 351, 587\n(2016).\n[64] R. Lebrun, A. Ross, S. A. Bender, A. Qaiumzadeh, L. Baldrati,\nJ. Cramer, A. Brataas, R. A. Duine, and M. Kläui, “Tunable\nlong-distance spin transport in a crystalline antiferromagnetic\niron oxide,” Nature 561, 222 (2018).\n[65] A. Kamra, E. Thingstad, G. Rastelli, R. A. Duine, A. Brataas,\nW. Belzig, and A. Sudbø, “Antiferromagnetic Magnons as\nHighly Squeezed Fock States underlying Quantum Correlations,”\n(2019), arXiv:1904.04553.\n[66] F. Mahfouzi and N. Kioussis, “Damping and antidamping phe-\nnomena in metallic antiferromagnets: An ab initio study,” Physi-\ncal Review B 98, 220410(R) (2018).\n[67] Such processes are forbidden in ferromagnets due to spin split-\nting between the opposite spin conduction electron bands.\n[68] P. W. Anderson, “An Approximate Quantum Theory of the Anti-\nferromagnetic Ground State,” Physical Review 86, 694 (1952).\n[69] R. Kubo, “The Spin-Wave Theory of Antiferromagnetics,” Phys-\nical Review 87, 568 (1952).\n[70] P. Merodio, A. Ghosh, C. Lemonias, E. Gautier, U. Ebels,\nM. Chshiev, H. Béa, V . Baltz, and W. E. Bailey, “Penetration\ndepth and absorption mechanisms of spin currents in Ir 20Mn80\nand Fe50Mn50 polycrystalline films by ferromagnetic reso-\nnance and spin pumping,” Applied Physics Letters 104(2014),\n10.1063 /1.4862971.\n[71] L. Frangou, S. Oyarzún, S. Au \u000bret, L. Vila, S. Gambarelli, and\nV . Baltz, “Enhanced Spin Pumping E \u000eciency in Antiferromag-\nnetic IrMn Thin Films around the Magnetic Phase Transition,”\nPhysical Review Letters 116, 077203 (2016).\n[72] A. Kamra, R. E. Troncoso, W. Belzig, and A. Brataas, “Gilbert\ndamping phenomenology for two-sublattice magnets,” Physical\nReview B 98, 184402 (2018).7\n[73] A. Kamra and W. Belzig, “Super-Poissonian Shot Noise of\nSqueezed-Magnon Mediated Spin Transport,” Physical Review\nLetters 116, 146601 (2016).\n[74] R. J. Glauber, “The Quantum Theory of Optical Coherence,”\nPhysical Review 130, 2529 (1963).\n[75] E. C. G. Sudarshan, “Equivalence of Semiclassical and Quantum\nMechanical Descriptions of Statistical Light Beams,” Physical\nReview Letters 10, 277 (1963).\n[76] N. Zagury and S. M. Rezende, “Theory of Macroscopic Excita-\ntions of Magnons,” Physical Review B 4, 201 (1971).\n[77] A. Kamra and W. Belzig, “Magnon-mediated spin current noise\nin ferromagnet |nonmagnetic conductor hybrids,” Physical Re-\nview B 94, 014419 (2016).\n[78] A. Kamra and W. Belzig, “Spin Pumping and Shot Noise in\nFerrimagnets: Bridging Ferro- and Antiferromagnets,” PhysicalReview Letters 119, 197201 (2017).\n[79] S. A. Bender and Y . Tserkovnyak, “Interfacial spin and heat\ntransfer between metals and magnetic insulators,” Physical Re-\nview B 91, 140402(R) (2015).\n[80] P. Tang, Q. Zhou, G. Xu, and S.-c. Zhang, “Dirac fermions in an\nantiferromagnetic semimetal,” Nature Physics 12, 1100 (2016).\n[81] S. Y . Bodnar, L. Šmejkal, I. Turek, T. Jungwirth, O. Gomonay,\nJ. Sinova, A. A. Sapozhnik, H.-J. Elmers, M. Kläui, and M. Jour-\ndan, “Writing and reading antiferromagnetic Mn2Au by Néel\nspin-orbit torques and large anisotropic magnetoresistance,” Na-\nture Communications 9, 348 (2018).\n[82] D. F. Shao, G. Gurung, S. H. Zhang, and E. Y . Tsymbal, “Dirac\nNodal Line Metal for Topological Antiferromagnetic Spintron-\nics,” Physical Review Letters 122, 77203 (2019)." }, { "title": "1803.09682v1.Radiation_reaction_of_charged_particles_orbiting_magnetized_Schwarzschild_black_hole.pdf", "content": "arXiv:1803.09682v1 [gr-qc] 26 Mar 2018Draft version March 28, 2018\nPreprint typeset using L ATEX style emulateapj v. 12/16/11\nRADIATION REACTION OF CHARGED PARTICLES ORBITING\nMAGNETIZED SCHWARZSCHILD BLACK HOLE\nArman Tursunov(1), Martin Kolo ˇs(1), Zdenˇek Stuchl ´ık(1), and Dmitri V. Gal’tsov(2,3)\n(1)Institute of Physics and Research Centre of Theoretical Phy sics and Astrophysics,\nSilesian University in Opava, Bezruˇ covo n´ am.13, CZ-7460 1 Opava, Czech Republic\n(2)Faculty of Physics, Moscow State University, 119899, Mosco w, Russia and\n(3)Kazan Federal University, 420008 Kazan, Russia\nDraft version March 28, 2018\nAbstract\nIn many astrophysically relevant situations radiation reaction forc e acting upon a charge can not\nbe neglected and the question arises about the location and stability of circular orbits in such regime.\nMotion of point charge with radiation reaction in flat spacetime is desc ribed by Lorenz-Dirac (LD)\nequation, while in curved spacetime – by DeWitt-Brehme (DWB) equat ion containing the Ricci term\nand the tail term. We show that for the motion of elementary partic les in vacuum metrics the DWB\nequation can be reduced to the covariant form of the LD equation w hich we use here. Generically,\nthe LD equation is plagued by runaway solutions, so we discuss compu tational ways to avoid this\nproblem in constructing numerical solutions. We also use the first ite ration of the covariant LD\nequation which is the covariant Landau-Lifshitz equation, comparin g results of these two approaches\nand showing smallness of the third-order Schott term in the ultrare lativistic case. We calculate the\ncorresponding energy and angular momentum loss of a particle and s tudy the damping of charged\nparticle oscillations around an equilibrium radius. We find that dependin g on the orientation of the\nLorentz force, the oscillating charged particle either spirals down t o the black hole, or stabilizes the\ncircular orbit by decaying its oscillations. The later case leads to an int eresting new result of shifting\nof the particle orbit outwards from the black hole. We also discuss th e astrophysical relevance of the\npresented approach and provide estimations of the main paramete rs of the model.\nSubject headings: black hole physics, magnetic fields, radiation reaction\n1.INTRODUCTION\nSynchrotron radiation emitted by a charged particle\nleads to appearance of the back-reaction force which can\nsignificantly affect its motion. The purpose of this pa-\nper is to study the motion of charged particles in the\ncombined magnetic and gravitational fields, taking into\naccount the radiation reaction force. There is convincing\nevidence that magnetic fields are indeed present in the\nvicinity of black holes. The observations of the Galac-\nticCentre(Eatough et al.(2013))havedemonstratedthe\nexistenceofstrongmagneticfieldofhundredGaussinthe\nvicinity of the supermassive black hole at the Galactic\ncentre. Recent studies of the quasi-periodic oscillations\n(QPOs) observed in the black hole microquasars have\nshown possible signatures of Galactic magnetic fields in\nthe vicinity of three microquasars (Koloˇ s et al. (2017)),\nif ”magnetic” generalizations of the geodesic models of\ntwin high frequency (HF) QPOs (Stuchl´ ık et al. (2013))\nare applied to data. Observed HF QPOs and relativistic\njets in microquasars indicate the presence of an external\nmagnetic field, influencing both oscillations in the accre-\ntion disk and creation of jets. The presence of radiation-\nreaction force and its relevance in the vicinity of black\nholes can make sufficient contribution for the shifts in\nthe QPO frequencies, which are usually observed. More-\nover, the radiation reaction can support the accretion of\ncharged particles from accretion disk towards the black\nhole. Depending on the magnitude of the external mag-\nnetic field, the radiation-reaction force can considerably\nshift the stable orbits of a particle which can sufficientlyinfluence the predictions of black hole parameters.\nThe weakness of magnetic field in our study is under-\nstood in the sense that it does not perturb the spacetime\nmetric; the corresponding condition on the field magni-\ntudeBin the vicinityofSchwarzschildblackholeofmass\nMreads (Gal’tsov & Petukhov (1978)):\nB << B G=c4\nG3/2M⊙/parenleftbiggM⊙\nM/parenrightbigg\n∼1019M⊙\nMGauss.(1)\nThis weakness is compensated by large ratio e/mfor\nelectrons and protons, whose motion will be essentially\naffected by magnetic fields already of the order of few\ngauss.\nStudy of the particle motion and electrodynam-\nics in magnetized black holes began long ago, see\ne.g. Gal’tsov & Petukhov (1978); Blandford & Znajek\n(1977), for a review of early results see Aliev & Gal’tsov\n(1989). Equatorial orbits are of primary interest for\ntheory of accretion disks. It was shown that the in-\nnermost stable circular orbits (ISCO) in the field of\nmagnetized black hole are shifted towards the hori-\nzon (Gal’tsov & Petukhov (1978)) for suitable direction\nof rotation. Recently, detailed studies of the equa-\ntorial motion of charged particles in weakly magne-\ntized Schwarzschild and Kerr black holes were per-\nformed in Frolov & Shoom (2010); Koloˇ s et al. (2015);\nTursunov et al. (2016), revealing a number of new types\nof motion.\nEstimates show that in many physically interesting\ncases the radiation reaction force is non-negligible, and2\nthis has prompted us to consider what happens once the\nreaction force is taken into account. The general prob-\nlem of the synchrotron radiation and treatment of the\nradiation-reaction force in curved spacetime was widely\nstudied in literature. Below, we remind only few of\nthem. The electromagnetic radiation emitted by rela-\ntivistic particles in the presence of strong gravity has\nbeen studied, e.g., in Poisson (2004); Johnston et al.\n(1973); Zerilli (2000); Price et al. (2013). The problem\nof synchrotron radiation in curved spacetime has been\nstudied, in Sokolov et al. (1983, 1978) using covarianti-\nzation of the flat space results. Actually, we will use here\nthe same approach to describe the reaction force. For\nthe recent treatment of radiation reaction in flat space\nsee Gal’tsov & Spirin (2006) where the derivation of the\nShott term was given explicitly along the lines of the\nTeitelboim idea to associate it with the Coulomb part of\nthe electromagnetic field of a point charge. It is worth\nnoting that the problem of the Shott contribution to re-\naction force (the third derivative term) still remains the\nsource of discussion (see e.g. Poisson (2004)), so we will\ntouch it in this paper too.\nGeneralizationofradiationreactionequationtocurved\nspace was given by DeWitt & Brehme (1960). Their re-\nsult (corrected by Hobbs (1968)) shows that radiation\nreaction in curved spacetime is essentially non-local and\nappeals to solve the integral equation. But contrary\nto gravitational radiation reaction in collisions of black\nholes, electromagnetic radiation from point particles can\nstill be simplified to a local theory.\nRecently, generalizationof the radiation reactionequa-\ntion to higher dimensions was discussed in Gal’tsov\n(2002); Gal’tsov & Spirin (2007) which may be rele-\nvant to extra-dimensional models. Radiation from hy-\npothetical massless charges was considered in Gal’tsov\n(2015) bringing new light on the relationship be-\ntween the ultrarelativistic and the massless limits in\nsynchrotron radiation and radiation reaction equa-\ntion. The motion of point particles with scalar,\nelectric and mass charges, has been reviewed in\nPoisson (2004). The problem of radiation-reaction\nfor extended charged particles has been studied in\nCremaschini & Tessarotto (2011), its quantum aspects\nare discussed in Cremaschini & Tessarotto (2015). Sta-\ntistical treatment of the radiation-reaction problem is\ngiven in Cremaschini & Tessarotto (2013). The problem\nof gravitational self-force of test particles is summarized\nin Barack (2014). Nevertheless, despite the active in-\nterest in the topic, and comprehensive literature review,\nthe successful attempts to integrate the equations of mo-\ntion in the curved background combined with external\nelectromagnetic fields is quite rare.\nIn spite of weakness of radiation reaction force in or-\ndinary stellar astrophysical situations, it was found re-\ncently that in high energy plasma processes in pulsar\nmagnetospheres, black-hole accretion disks, hot accre-\ntion flows in X-ray Binaries and Active Galactic Nu-\nclei, relativistic jets, and so on, which are dominated\nby magnetic reconnection in plasma, radiation reaction\nmayplayacrucialrole. The newtheory, named radiative\nmagnetic reconnection (for a recent review see Uzdensky\n(2016)), predicts notable radiative effects in astrophysi-\ncal reconnection such as radiation-reactionlimits on par-\nticle acceleration, radiative cooling, radiative resistivity,braking of reconnection outflows by radiation drag, radi-\nation pressure, viscosity, and even pair creation at high-\nest energy densities. The radiative reconnection theory\nis based on the flat-space description of radiation reac-\ntion adapted to phenomena in strong gravitational fields\nwhich is well justified in dense plasmas.\nHowever, in the case of diluted media, when individ-\nual particle motion is relevant, it seems more adequate\nto use description of radiation reaction in curved space-\ntime from the very beginning. This is the main goal of\nthe present paper. Here, we would like both to discuss\nconceptualproblemsofsuchadescription,andtoanalyse\nnumerically the modification of particle motion within a\nsimple model of weakly magnetized Schwarzschild black\nhole. Classifying the motion of charged particles by the\ngiven set of parameters, we find the explicit trajecto-\nries for each class of orbits and study the evolutions of\nthe particle energy, angular momentum, etc. Motion of\ncharged particles in the combined magnetic and gravi-\ntatational field without radiation reaction reveals inter-\nesting new types of trajectories. We intend to apply\nthe full machineryofcurvedspacetime radiationreaction\ntheory to these problems. We will also discuss possible\napplications of our results in some astrophysical scenar-\nios.\nThe paper is organized as follows. In Sec. 2, we test\nthe dynamical equations with radiation reaction force\nin flat spacetime using two main approaches and com-\npare the results. In Sec. 3, we present the properties of\nthe circular motion of non-radiating charged particles in\nweakly magnetized Schwarzschildblack hole, concentrat-\ning attention on the bounded orbits and charged parti-\ncle oscillations. In Sec. 4, we give the general relativistic\ntreatmentofthe radiation-reactionforcein curvedspace-\ntime, and under reasonable assumptions we give explicit\nform of the equations ofmotion ofchargedparticle in the\nvicinity of Schwarzschild black hole immersed in exter-\nnal asymptotically uniform magnetic field. In Sec. 5, we\nanalyse the trajectories of charged particles for different\nclasses of orbits and find the evolutions of the relevant\nparameters of the system during and after the decay of\nparticle oscillations. Switching to the Gaussian units in\nSec.6, we estimate the characteristic decay times of the\ncharged particle oscillations and discuss the relevance of\nthe model in astrophysicalsituations. We summarize our\nresults in Sec. 7.\nThroughout the paper, we use the spacetime signature\n(−,+,+,+), and the system of geometric units in which\nG= 1 =c. However, for expressions having an astro-\nphysicalrelevance, we use the constants explicitly. Greek\nindices are taken to run from 0 to 3.\n2.RADIATION REACTION IN FLAT SPACETIME\nFor completeness, we first summarize the results of the\nmotion of charged particle in the flat spacetime. The\nequation which describes the motion of charged particle\nin magnetic field in general contains two forces\nduµ\ndτ=fµ\nL+fµ\nR, (2)\nwherefµ\nL= (q/m)Fµνuνis the Lorentz force, Fµν=\n∂µAν−∂νAµis the tensor of electromagnetic field and\nuµ(τ) =dxµ/dτis a four-velocity of the particle. The3\nlast term fµ\nRis the radiation reaction force which in\nthe non-relativistic case has to lead to the expression\n3q2\n2md2uα\ndτ2. Moreover any force vector has to satisfy the\ncondition fµ\nRuµ= 0. This implies that the correct co-\nvariant form of the expression for the radiation reaction\nforce is\nfµ\nR=2q2\n3m/parenleftbiggd2uµ\ndτ2+uµuνd2uν\ndτ2/parenrightbigg\n. (3)\nThis expression was found by Dirac and motion equa-\ntion is sometimes called as Lorentz-Abraham-Dirac\nor Lorentz-Dirac (LD) equation. The first term\nin the parentheses, also known as the Schott term,\narises from the particle electromagnetic momentum\n(Gal’tsov & Spirin (2006)). The second term in paren-\ntheses is the radiation recoil term which corresponds to\nthe relativistic correction of the radiation reaction force.\nThe four-velocity of the charged particle satisfies the fol-\nlowing equations\nuαuα=−1, uα˙uα= 0, uα¨uα=−˙uα˙uα.(4)\nThus, the motion of radiating charge is described by a\nthird order differential equation in coordinates, rather\nthan habitual second order. This may lead to unphys-\nical solutions - the existence of pre-acceleration in the\nabsence of external forces. Even though the unphysi-\ncal solutions can be removed by properly chosen initial\nconditions, the integration of exact form of equations of\nmotion (2) and (3) is inconvenientdue to the exponential\nincrease of the computational error in practical calcula-\ntions.\nHowever, one can reduce the order of LD equation by\nthe method proposed in Landau & Lifshitz (1975), i.e.,\nby rewriting the self-force in terms of the external force\nand the four-velocity of a particle. Substituting higher\norder terms in ( 3) by the derivatives of the Lorentz force\nwe get the equation in the following form\nduµ\ndτ=fµ\nL+2q2\n3m(δµ\nα+uµuα)dfα\nL\ndτ. (5)\nThis equation, usually referred as Landau-Lifshitz (LL)\nequation, has important consequences: it is of the sec-\nond order, does not violate the principle of inertia, and\nthe self-force vanishes in the absence of the external\n(Lorentz) force, Rohrlich (2001); Poisson (1999). The\nself-contained derivation of the equation ( 5) in terms of\nretarded potentials is given in Poisson et al. (2011). The\nequation ( 5) can be applied for the cases with any ex-\nternal force acting on a charged particle instead of the\nLorentzforce. In case when fµ\nL=q\nmFµνuν, the radiation\nreaction force can be rewritten in the form\nfµ\nR=k˜q/bracketleftbig\nFµν\n,αuαuν+ ˜q(FµνFνρ−FβαFβρuαuµ)uρ/bracketrightbig\n,\n(6)\nwhere ˜q=q/mis the specific charge of the particle,\nk= (2/3)˜qq, and the comma in the first term denotes\nthe partial derivative with respect to the coordinate xα.\nItwasconcludedinSpohn(2000)thatusingtheLLequa-\ntionisidenticaltoimposingDirac’sasymptoticcondition\nlimτ→∞˙uµ= 0 to the LD equation. It was later con-\nfirmed by Rohrlich (2001) that the reduced form of the\nequation of motion is exact, rather than approximative,\nthough LL equation was proposed in Landau & Lifshitz(1975) as an approximative solution to the third order\nLD equation. More details on the treatment of radia-\ntion reaction of charged particles in flat spacetime can\nbe found in a book of Spohn (2004). In our numerical\nstudy we found that the LL approximation is perfectly\napplicable if the Schott term is small with respect to the\nradiation recoil term, which is the case we consider here.\nBelowweshowthe representativeexampleofthe charged\nparticle motion in external uniform magnetic field, inte-\ngrating both LD and LL equations. Results of numerical\nstudiesofLDand LLequationsforthe motion ofcharged\nparticle in uniform magnetic field in flat spacetime are\nin accord with the analytical treatment of the radiation\nreaction force performed in Spohn (2000).\n2.1.Charged particle in uniform magnetic field\nLetusconsiderthemotionofachargedparticleinaho-\nmogeneous magnetic field aligned along z-axis, such that\nthe independent nonvanishing component of the electro-\nmagnetic tensor is Fxy=B. We introduce the new pa-\nrameter in the form B=qB/(2m), where the factor 1 /2\nis added in order to correspond to the similar parameter\nintroduced in the curved spacetime case.\nBoth LD and LL equations, lead to equivalent results,\ndiffering in the number of initial conditions. In case of\nLD equation, one needs to set the values of 9 constants -\narbitrary independent components of initial position, ve-\nlocity and the acceleration of the charged particle. The\nthreeother constantsaregivenby the normalizationcon-\ndition (4). However, the direct integration of higher or-\nder equations leads to the exponential increase of the\ncomputational error in very short times. It is interest-\ning to note that the problem of the time dispersion er-\nror can be greatly reduced by integrating equations of\nmotion backwards in time. Similar method of solving\nLorentz-Dirac equations has been proposed in the past\nby Huschilt & Baylis (1976).\nOn the other hand, the reduced-orderequations of mo-\ntion (5) can be written explicitly in the form\ndux\ndτ=2Buy−4kB2/parenleftbig\n1+u2\n⊥/parenrightbig\nux, (7)\nduy\ndτ=−2Bux−4kB2/parenleftbig\n1+u2\n⊥/parenrightbig\nuy, (8)\nduz\ndτ=−4kB2u2\n⊥uz, (9)\ndut\ndτ=−4kB2u2\n⊥ut. (10)\nHereu2\n⊥= (ux)2+(uy)2is square of the particle veloc-\nity in the plane orthogonal to the magnetic field and z\naxis. The representative trajectory of radiating charged\nparticle corresponding to both LD and LL cases is shown\nin Fig.1. Initial conditions are chosen as follows: initial\nplane velocity is u⊥0= 0.8c, velocity in vertical direction\nisuz\n0= 0.5c, magnetic field is aligned along zaxis and\nmagnetic parameter is chosen to be B= 1, radiation pa-\nrameterk= 0.01. The estimations of the parameters in\nrealistic situations are given in Sec 6. Energy and angu-\nlar momentum loss leads to decay of the plane velocity of\nthe particle, while the vertical component remains con-\nstant. Apparent deceleration along zaxis shown in Fig. 1\n(solid thick curve in the middle plot), while no forces are4\nFig. 1.— Motion of radiating charged particle in flat spacetime. Left figure represents an example of 3D trajectory of charged part icle.\nMiddle plot represents the evolution of different component s of 3-velocity in proper time τ: the damping harmonic oscillations correspond\ntoux(solid thin) and uy(dashed) components of velocity, tangential to them (dotte d) is a plane velocity of a particle v⊥orthogonal to z\naxis, vertical uzcomponent of velocity is shown by solid thick line and dot-da shed curve of the middle plot shows the evolution of 3-veloci ty\n(/radicalBig\nu2x+u2y+u2z). Right figure shows the change of the energy, angular moment um and gyroradius of charged particle in time twith respect\nto the static observer. Energy decreases up to the certain va lue (12), while angular momentum and radius of gyration asym ptotically tend\nto zero.\nappliedinthe zdirection, appearsonlyin theframemov-\ning with the particle. The velocity with respect to the\nstatic observer, vz=dz/dt≡uz/ut=const, remains\nconstant.\n2.2.Energy and momentum loss\nThe rate of the energy loss of the particle can be eval-\nuated from Eq.( 10). Modifying it for a static observer,\nwe get\ndE\ndt=−kB2˜q2u⊥(t)2. (11)\nIntegrating this equation, one can obtain the energy of\nthe particle in a given moment of time. The evolution\nof the energy in time is shown in Fig. 1for the given tra-\njectory. Thus the energy loss will be given only by the\nchange of the plane velocity u⊥in time, while the kinetic\nenergy associated with the motion in the z-direction will\nbe conserved. The plane velocity u⊥of the particle de-\ncreases in time and asymptotically tends to zero as rep-\nresented by the dotted curve of the middle plot of Fig. 1.\nThis implies that there exists an irreducible (specific) en-\nergy of radiating charged particle which corresponds to\nthe final state of the particle, having the following simple\nform\nE0=/parenleftbig\n1−(vz\n0)2/parenrightbig−1\n2, (12)\nwherevz\n0is the vertical velocity of the particle along the\nz-axis, measured by the static observers, which is con-\nstant during the radiation process.\nIn orderto find the rate ofthe angularmomentum loss,\none can fix the motion in a plane by taking uz= 0. In\ngeneral, the specific angular momentum for the motion\nin a plane is defined by the formula L=ρ2dφ/dτ+˜qAφ,\nwhereρis gyroradius of the particle trajectory. Re-\nminding that dφ/dτ=u⊥γ/ρandρ=u⊥/ωL, where\nωL=qB/m≡2Bis the Larmorfrequency, one can write\nthe specific angular momentum of the radiating chargedparticle in given moment of time in the form\nL=u⊥(τ)2\n4B[2γ(τ)+1], γ= (1−u2\n⊥)−1\n2.(13)\nSolving first two equations of motion ( 7) and (8), and\nsubstituting into Eq.( 13), we get the evolution of the an-\ngular momentum in time, which is represented in Fig. 1.\nUnlike the energyof the particle, the angular momentum\nasymptotically tends to zero for large τ. This occurs due\nto the reason that the gyroradius ρof the charged parti-\ncle tends to zero as well, while Lis proportional to ρ.\nOne can find the ratio between the angular momentum\nloss˙L=dL/dtandthe energyloss ˙E=dE/dtin theform\n˙L\n˙E=r2Ωu2\n⊥+1\nu2\n⊥. (14)\nwhere Ω is the angular frequency of the charged particle\nmeasured by the observers at rest. In Cartesian coordi-\nnates Ω takes the form\nΩ =x˙y−y˙x\nx2+y2, (15)\nwhere dots denote the derivative with respect to the co-\nordinate time t.\n2.3.Decay time\nOne can find the characteristic time required to decay\nthe energyofa radiatingchargedparticle in the following\nway. Since the velocity in the magnetic field direction is\nconstant, one can consider only the planar motion of the\nparticle by taking uz= 0. This implies that according\nto condition uαuα=−1, we get u2\n⊥= (ut)2−1. Thus,\nequation ( 10) can be rewritten in the form\ndE\ndτ=−K/parenleftbig\nE3−E/parenrightbig\n,K= 4kB2.(16)5\nIntegrating this equation, we get the particle energy in a\ngiven moment of time\nE(τ) =EieKτ\n/radicalbig\n1+E2\ni(e2Kτ−1), (17)\nwhere the integration constant Eiis the initial specific\nenergy of the particle. Asymptotically in time the spe-\ncific energy tends to the particle rest energy, being equal\nto 1. Thus, the decay time during which the specific en-\nergy will be lowered from EitoEfdue to radiation takes\nthe following form\nT=1\n2KlnE2\nf(E2\ni−1)\nE2\ni(E2\nf−1). (18)\nwhereK= 4kB2. Since the energy in ( 17) is the expo-\nnentially decreasing function, the particle energy cannot\nbe reduced to 1 in practical calculations.\n3.CHARGED PARTICLE ORBITING BLACK HOLE\nWITHOUT RADIATION-REACTION\nIn this section we shortly summarize previous re-\nsults related to the particle motion in the field of mag-\nnetized black holes presented in Koloˇ s et al. (2015);\nStuchl´ ık & Koloˇ s(2016); Tursunov et al. (2016). The in-\nterval in the Schwarzschild black hole spacetime in sper-\nical coordinates ( t,r,θ,φ) reads\nds2=−f(r)dt2+f−1(r)dr2+r2(dθ2+sin2θdφ2),(19)\nwhereMis the black hole mass and the function f(r) is\nthe lapse function given by\nf(r) = 1−2M\nr. (20)\nHereafter, without loss of generality we put the mass of\nthe black hole to be equal to unity, M= 1. Let us\nconsider the black hole immersed into external asymp-\ntotically uniform magnetic field. In the case when this\nfield is weak implying that the metric of Schwarzschild\nblack hole is not violated (see Eq.( 1)), one can write the\nsolution of the Maxwell equation for the four-vector po-\ntential of electromagnetic field Aµin the following form\n(Wald (1984))\nAφ=B\n2gφφ=B\n2r2sin2θ, (21)\nwhich is the only nonzero component of four-vector po-\ntentialAµandBis the strength of the magnetic field\nat spatial infinity, which is taken to be constant. The\nantisymmetric tensor of the electromagnetic field Fµν=\nAν,µ−Aµ,νinthis casehasonlytwoindependent nonzero\ncomponents\nFrφ=Brsin2θ, Fθφ=Br2sinθcosθ.(22)\n3.1.Bounded motion around black hole\nThe motion of charged particles is described by the\nLorentz equation in curved spacetime. In case when the\nradiative processes can be neglected, one can write for\nthe particle of mass mand charge qthe Lorentz equation\nof motion\nDuµ\ndτ≡duµ\ndτ+Γµ\nαβuαuβ=q\nmFµ\nνuν,(23)whereuµis the four-velocity of the particle, normalized\nby the condition uµuµ=−1,τis the proper time of\nthe particle and components of Γµ\nαβare the Christoffel\nsymbols.\nSymmetry of the Schwarzschildblack hole ( 19) and ex-\nternal magnetic field give us a right to find the conserved\nquantities associated with the time and space compo-\nnents of the generalized four-momentum πα=pα+Aα.\nThus, the energy and angular momentum of the charged\nparticleinthepresenceofexternaluniformmagneticfield\ntake the following form\nE=−πt=mf(r)dt\ndτ, (24)\nL=πφ=mr2sin2θ/parenleftbiggdφ\ndτ+qB\n2m/parenrightbigg\n.(25)\nIn this case, the circular motion of charged particles is\nalways bounded in the plane orthogonal to the magnetic\nfield lines, which corresponds to the equatorial plane\nθ=π/2. The boundary of the motion is governed by\nthe shape of the effective potential, determined by the\nequation\nE2=Veff(r,θ;L,B). (26)\nUsing the notation\nE=E\nm,L=L\nm,B=qB\n2m. (27)\nwe can write the effective potential in the form\nVeff(r,θ)≡f(r)/bracketleftBigg\n1+/parenleftbiggL\nrsinθ−Brsinθ/parenrightbigg2/bracketrightBigg\n.(28)\nThe effective potential ( 28) shows clear symmetry\n(L,B)↔(−L,−B) that allows to distinguish the fol-\nlowing two situations\n-minus configuration , withL>0,B<0 (equiva-\nlent toL<0,B>0) - magnetic field and angular\nmomentum parametershaveoppositesigns and the\nLorentz force is attracting the charge towards the\nblack hole.\n+plus configuration , withL>0,B>0 (equivalent\ntoL<0,B<0) - magnetic field and angular mo-\nmentum parameters have the same signs and the\nLorentz force is repulsive, acting outward the black\nhole.\nThe last case appears only in the combination of gravity\nwith electromagnetic field and cannot exist in flat space-\ntime (for details, see Koloˇ s et al. (2015); Tursunov et al.\n(2016); Stuchl´ ık & Koloˇ s(2016)). Particular trajectories\nof charged particles in magnetized Schwarzschild black\nhole are compared with radiating particle motion in Sec-\ntion5.\n3.2.Charged particle oscillations\nCharged particles can undergo stable quasi-harmonic\noscillationsinradialandverticaldirectionsinthevicinity\nof magnetized black hole. In the case when the oscilla-\ntions are small enough in comparison with the radius of\nthe corresponding stable circular orbit, one can find the6\nFig. 2.— Dependence of angular momentum L, energy Eand azimuthal velocity uφof a charged particle on the location of the circular\norbitrwithout radiation for different values of magnetic paramete rB.\nlocally measured frequencies of vertical ωθand radial ωr\noscillations equal to\nω2\nθ=L2\nc\nr4−B2, (29)\nω2\nr=1\n(r−2)r5/bracketleftbig\n(r−2)2/parenleftbig\nB2r4+3L2\nc/parenrightbig\n−2r/parenleftbig\nBr2−Lc/parenrightbig2−2r3/bracketrightbig\n(30)\nwhereLcis the specific angular momentum at the circu-\nlar orbit, for details see Koloˇ s et al. (2015). In addition\nto the frequencies givenabove, one can find alsothe Kep-\nlerian axial frequency ωφand Larmor angular frequency\nωLwhich are given by\nωφ=dφ\ndτ=Uφ=Lc\ngφφ−B, ωL=qB\nm= 2B.(31)\nAs one can see, the frequency ωLdoes not depend on r\ncoordinate and plays crucial role in the regions detached\nfrom the black hole. The characteristic oscillations of\nchargedparticlesandrepresentativeplotsofchargedpar-\nticle trajectories around Schwarzschild black hole can be\nfound in Fig.7 of paper Koloˇ s et al. (2015). For the par-\nticle to oscillate in the vicinity of a black hole, its energy\nhas to be larger than the minimum of the effective po-\ntential, however, keeping the finite type of the motion.\nThus,theminimalenergyofachargedparticleintrapped\nstates corresponds to the stable circular orbits given by\nthe minimum of the effective potential ( 28). The maxi-\nmal energy at the trapped states is given by the unstable\ncircular orbit with the corresponding angular momen-\ntum. Energy of the charged particle at the circular orbit\nis given by\nE2\n±=rf2\n(r−3)2/parenleftBig\nr−3+2B2r3f±2Br/radicalbig\nr−3+B2r4f2/parenrightBig\n,\n(32)\nwhere the signs correspond to the maximal and minimal\nenergy of a particle at the circular orbit, and f= 1−2/r\nis the lapse function. The difference between E+and\nE−can be very large for large values of magnetic pa-\nrameter B, representing the charged particles acceler-atedup to ultrarelativisticvelocities(Koloˇ s et al.(2015);\nStuchl´ ık & Koloˇ s (2016)).\nSome types of the quasi-circular epicyclic motion are\npossible only in the presence of magnetic field. One of\nthe interesting and illustrative examples of the effect of\nmagnetic field on the charge particle motion in the black\nhole background is the appearance of curled trajectories,\nas demonstrated in Fig. 4which corresponds to the plus\nconfiguration with repulsive Lorentz force. The conser-\nvation of the angular momentum ( 25) in the absence of\nradiation leads to the equation\n˙φ=L\nr2−B. (33)\nInthe non-magneticcase, therighthandsideoftheequa-\ntion above is positive for positive L. The presence of\nmagnetic field with B>0 decreases the velocity in φ-\ndirection that can become negative, if\nL>L∗(r;B)≡ Br2. (34)\nFor the energy, this condition means\nE>E∗(L;B)≡/radicalBig\n1−2/radicalbig\nB/L. (35)\nThus, decreasingthe azimuthalvelocityofaparticlewith\nthe influence of magnetic field, and keeping the above\ngiven conditions, we get so called curled trajectories,\nwhich are not possible in the absence of magnetic field.\nThis type of orbits does not appear in the case of attrac-\ntive Lorentz force corresponding to the minus configura-\ntions.\nIn the next sections we show that the oscillations of\ncharged particles in magnetic field near a Schwarzschild\nblack hole will be damped due to the synchrotron radi-\nation and, in particular cases the charged particle orbits\nwill not be able to stay stable, bringing the particle to\nthe black hole due to the effect of the radiation reaction\nforce.\n4.RADIATION REACTION IN THE FIELD OF\nMAGNETIZED BLACK HOLES\n4.1.Radiation-reaction force7\nThe motion of a relativistic charged particle is gov-\nerned by the Lorentz-Dirac equation which includes the\ninfluence of the external electromagnetic fields and cor-\nresponding radiation-reaction force. The last force arises\nfrom the radiative field of the charged particle and the\nequations ofmotion in generalcan be written in the form\nDuµ\ndτ= ˜qFµ\nνuν+ ˜qFµ\nνuν, (36)\nwhere the first term on the right hand side of Eq.( 36)\ncorresponds to the Lorentz force with electromagnetic\ntensorFµνgiven by ( 22), while the second term is the\nself-force of charged particle with the radiative field\nFµν=Aν,µ− Aµ,ν. The vector potential of the self-\nelectromagnetic field of the charged particle satisfies the\nwave equation\n✷Aµ−Rµ\nνAν=−4πjµ, (37)\nwhere✷=gµνDµDν, andDµis the covariant differenti-\nation and Rµ\nνis the Ricci tensor. The retarded solution\nto Eq.(37) for the vector potential takes the form\nAµ(x) =q/integraldisplay\nGµ\n+λ(x,z(τ))uλdτ, (38)\nwhereGµ\n+λis the retarded Green function and the in-\ntegration is taken along the worldline of the particle z,\ni.e.,uµ(τ) =dzµ(τ)/dτ. For details, see, e.g. Poisson\n(2004). The covariant generalization of the dynamics of\nradiating charged particle in curved spacetime has been\nderived in DeWitt & Brehme (1960) and completed in\nHobbs (1968), using the tetrad formalism. The explicit\nform of Eq.( 36) for the motion of charged particle under-\ngoing radiation-reaction force in curved spacetime reads\nDuµ\ndτ= ˜qFµ\nνuν+2q2\n3m/parenleftbiggD2uµ\ndτ2+uµuνD2uν\ndτ2/parenrightbigg\n+q2\n3m/parenleftbig\nRµ\nλuλ+Rν\nλuνuλuµ/parenrightbig\n+2q2\nmfµν\ntailuν,(39)\nwhere the last term of Eq.( 39) is the tail integral\nfµν\ntail=/integraldisplayτ−0+\n−∞D[µGν]\n+λ′/parenleftbig\nz(τ),z(τ′)/parenrightbig\nuλ′dτ′.(40)\nDetailed derivation of the equations of motion of radi-\nating charged particles can be found in Hobbs (1968);\nPoisson (2004). The integral in the tail term is evalu-\nated over the past history of the charged particle with\nprimes indicating its prior positions. All other quanti-\nties are evaluated at the current position of the particle\nz(τ). The term containing the Ricci tensor vanishes in\nthe vacuum metrics, so this term is irrelevantin our case.\nThe existence of the ”tail” integral in ( 39) implies that\nthe radiation reaction in curved spacetime has non-local\nnature, because the motion of the charged particle de-\npends on its whole history and not only on its current\nstate. The radiation field Fµνin Eq.(36), emitted by the\ncharged particle, interacts with the curvature of back-\nground spacetime and comes back to the particle with\na delay corresponding to the tail integral in ( 39). In\nsuch sense, the radiated electromagnetic field of charged\nparticle carries the information about the history of the\nparticle. Even in the absence of external forces, such asthe Lorentz force, the free trajectory of the charged par-\nticle does not follow the geodesics, which is one of the\nmost important consequences of equation ( 39).\nHowever, for purposes of the present paper, the\ntail term can be neglected as we show below. The\ntail terms can be estimated based on the results by\nDewitt & Dewitt (1964), Smith & Will (1980), as well\nas multiple subsequent papers. For a particle with the\nchargeqand mass m, the ratio of the tail force Ftail∼\nGMq2/(r3c2) to the Newton-force FN∼GMm/r2in the\nvicinity of a black hole ( r∼rH= 2GM/c2) of the stellar\nmassM∼10M⊙is\nFtail\nFN∼q2\nmMG∼10−19/parenleftBigq\ne/parenrightBig2/parenleftBigme\nm/parenrightBig/parenleftbigg10M⊙\nM/parenrightbigg\n,(41)\nwhereeandmeare the charge and the mass of an elec-\ntron. For supermassive black holes (SMBH) with the\nmassM∼109M⊙this ratio is 8 orders lower. On the\nother hand, the radiation reaction force (second term\non the right hand side of ( 39)) depends on the presence\nof external force, arising in our case from the external\nmagnetic field. According to Piotrovich et al. (2011);\nBaczko et al.(2016), thecharacteristicvaluesofthemag-\nnetic fields near the stellar mass black holes and SMBH\nareB∼108G, forM= 10M⊙andB∼104G, for\nM= 109M⊙. Thus, for the particle with velocity com-\nparable to the speed of light, v∼c, the ratio of the radi-\nation reaction force FRR∼q4B2/(m2c4) to the Newton-\nforce gives an order\nFRR\nFN∼q4B2MG\nm3c8∼103/parenleftBigq\ne/parenrightBig4/parenleftBigme\nm/parenrightBig3/parenleftbiggB\n108G/parenrightbigg2/parenleftbiggM\n10M⊙/parenrightbigg\n.\n(42)\nThe ratio of FRRtoFNfor SMBH with M= 109M⊙\nand magnetic field B∼104G gives the same order of\nmagnitude.\nThe above estimations of the tail term apply in the\nnon-relativistic, or moderately relativistic case when the\nLorentzfactorisoftheorderofunity. Moregeneralargu-\nment is based on the treatment of the gravitational self-\nforce in the Schwarzschild and Kerr spacetimes. As was\nshown by one of the present authors in Gal’tsov (1982),\nthe radiative part of the self-force (based on the half-\ndifference of the retarded and advanced Green’s func-\ntion) in the Kerr metric satisfies the (averaged on time)\nbalance equations for the energy and angular momen-\ntum of radiation of all spins s= 0,1,2 in the sense that\nlocal work of the self-force is equal to radiated fluxes\nat infinity and the black hole horizon. This is valid\nindependently of the particle velocity and extends to\nthe ultrarelativistic case. Later on it was shown that\nsimilar balance is valid for the Carter’s constant (Mino\n(2003)), which completes the set of quantities determin-\ning geodesic motion in the Kerr field. The conservative\npart of the self-force (half-sum of the retarded and ad-\nvanced potentials) is more difficult to compute since this\ndemands proper elimination of divergences. This caused\na vivid discussion in the literature from the mid-90-ies\nto early 2000-ies, nicely reviewed in Tanaka (2006) (for\nmore recent work on the same subject containing fur-\nther references see Fujita et al. (2017)), resulting in a\nconsensus opinion that this contribution is small with\nrespect to to radiative self-force especially with growing8\nLorentz factor (Pound et al. (2005); Sago et al. (2005)).\nTherefore, if one is interested to estimate the gravita-\ntional and electromagnetic tail terms in the ultrarela-\ntivistic limit, using the above results one can revoke\nthe gravitational synchrotron radiation (GSR), which\nwas computed earlier for spins s= 0,1,2 most no-\ntably in Chrzanowski & Misner (1974) and in Breuer\n(1975) with comparison to flat-space synchrotron radi-\nation (SR). From these results one can see that GSR in\nthe ultrarelativistinc limit is suppressed by a square of\nthe Lorentz factor with respect to SR. These arguments\nlead us to conclude that for elementary particles mov-\ning with any velocity in both gravitational and electro-\nmagnetic field the purely gravitational tail term can be\nneglected in comparison with the electromagnetic (prop-\nerly covariantized) radiation reaction force. Thus, for\npurposes of the present paper, the equation of motion\n(39) can be simplified to the following covariant form of\nLD equation:\nDuµ\ndτ= ˜qFµ\nνuν+fµ\nR, (43)\nwith the radiation reaction force given by\nfµ\nR=2q2\n3m/parenleftbiggD2uµ\ndτ2+uµuνD2uν\ndτ2/parenrightbigg\n.(44)\nIntroducing the four-acceleration as a covariant deriva-\ntive of four-velocity, aµ=Duµ/dτ, one can rewrite the\ntermD2uµ/dτ2as follows\nD2uµ\ndτ2≡Daµ\ndτ=daµ\ndτ+Γµ\nαβuαaβ\n=d\ndτ/parenleftbiggduµ\ndτ+Γµ\nαβuαuβ/parenrightbigg\n+Γµ\nαβuα/parenleftbiggduβ\ndτ+Γβ\nρσuρuσ/parenrightbigg\n=d2uµ\ndτ2+/parenleftBigg\n∂Γµ\nαβ\n∂xγuγuβ+3Γµ\nαβduβ\ndτ+Γµ\nαβΓβ\nρσuρuσ/parenrightBigg\nuα.\n(45)\nThus, the general relativistic equations of radiating\ncharged particle motion are given by Eqs ( 44) and (45).\nHowever, the full form of the equations is plagued by\nrunaway solutions. One can avoid this problem in a sim-\nilar way, as in the flat spacetime case, namely reducingthe order of differential equations. In the absence of the\nradiation-reactionforce, the motion of the charged parti-\ncle in external electromagnetic field is governed by equa-\ntion (23). Taking the covariant derivative with respect\nto the proper time from both sides of Eq.( 23), we get\nD2uα\ndτ2= ˜qDFα\nβ\ndxµuβuµ+ ˜q2Fα\nβFβ\nµuµ,(46)\nSubstituting ( 46) into Eq.( 44), we get the radiation re-\naction force in the form\nfα\nR=k˜q/parenleftbiggDFα\nβ\ndxµuβuµ+ ˜q/parenleftbig\nFα\nβFβ\nµ+FµνFν\nσuσuα/parenrightbig\nuµ/parenrightbigg\n,\n(47)\nwhere the covariant derivative from the second rank ten-\nsor reads\nDFα\nβ\ndxµ=∂Fα\nβ\n∂xµ+Γα\nµνFν\nβ−Γν\nβµFα\nν.(48)\nEquations( 43) and (47) givea covariantform ofLandau-\nLifshitz equations. Below we test these equations in par-\nticular case of the motion of charged particles around\nSchwarzschild black hole immersed into external asymp-\ntotically uniform magnetic field.\nThe motion of charged particles in the vicinity of\nmagnetized Schwarzschild black hole is generally chaotic\nKop´ aˇ cek & Karas (2014); Kop´ aˇ cek et al. (2010). How-\never, close to the minimum of the effective potential,\nwhich corresponds to the stable circular orbit, the mo-\ntion of the charged particle is regular, being of harmonic\ncharacter. The motion is also regular, if the particle is\nmoving entirely in the equatorial plane and the chaotic\nbehaviour appears with increasing the inclination angle.\nHere, we focus our attention on the regular motion only.\nIn order to represent the equations of motion explicitly,\nwe fix the plane of the motion at the equatorial plane,\nθ=π/2, of a magnetized black hole. However, for ex-\nploring the trajectories of particles, we use the full set of\nequations of motion and solve them numerically. With-\nout loss of generality, one can again equalize the mass of\na black hole to unity, M= 1. The non-vanishing com-\nponents of equations of motion of radiating charged par-\nticles moving around Schwarzschild black hole immersed\ninto external asymptotically uniform magnetic field take\nthe following form\ndut\ndτ=2utur\nr(2−r)−2kBut\nr/braceleftbig\n2Brf/bracketleftbig\nf(ut)2−1/bracketrightbig\n−uφ/bracerightbig\n, (49)\ndur\ndτ=uφ(2Brf+r−1)−1\nr2−2kBur\nr/braceleftbig\n2Brf2(ut)2−uφ/bracerightbig\n, (50)\nduφ\ndτ=−2ur/parenleftbig\nuφ+B/parenrightbig\nr+2kB\nr3/bracketleftbig\nr2(uφ)2+2Br3f2(ut)2uφ+1/bracketrightbig\n, (51)\nwherefisthelapsefunctiongivenby( 20),B=qB/(2m)\nis the magnetic parameter and k= 2q2/(3m) is the ra-\ndiation parameter.\n4.2.Energy and angular momentum lossThe total energy-momentum radiated by the charged\nparticle is equal to the integral of the radiation reac-\ntion force taken along the worldline of the particle. In\nflat spacetime, the radiated four momentum of a particle\nwith charge qis given by dPµ/dτ=2\n3q2aαaαuµ. Syn-\nchrotron radiation in curved spacetime has been studied,9\nin Sokolov et al. (1983), Sokolov et al. (1978) using co-\nvariantization of the flat space results. The problem has\nbeen revisited more recently in Shoom (2015), where,\nhowever, the radiation reaction force is not taken into\naccount. The evolutions of the particular components of\nthe four-momentum of the particle with radiation reac-\ntion force can be found from equations ( 43) and (47).\nFor the motion at the equatorial plane, the energy loss\nis given by\ndE\ndτ=−2kB/bracketleftbigg\n2BE3−E/parenleftbigg\n2Bf+uφ\nr/parenrightbigg/bracketrightbigg\n.(52)\nFor ultrarelativisticparticle with E ≫1, the leading con-\ntribution to the energy loss is given by the first term in\nsquare brackets of ( 52). However, for small velocities\nclose to the stable circular orbits, the last two terms can\nplay significant role. Both situations will be studied in\ndetails in the following section. Similarly, one can find\nthe rate of angular momentum loss as\ndL\ndτ= 4B2kuφ/parenleftbig\nf2(ut)2−f/parenrightbig\n−2uruφ/parenleftbig\nr−4B2k2/parenrightbig\n+2rBur.\n(53)\nIn the next section we will test the equations of mo-\ntion numerically for general bounded motion of radiat-\ning charged particle without restriction to the equatorial\nplane.\n5.DAMPING OF OSCILLATIONS\nIn this section we study the damping of charged par-\nticle oscillations due to radiation reaction force for both\n”+” and ” −” configurations. In particular, we demon-\nstrate that the particles at ” −” configurations spiral\ndown to the black hole, while at ”+” configurations the\nmotion remains stable. In the subsection 5.3, we study\nthe evolution of circular orbits under the influence of self\nforce. We show that in such a case the circular orbits\ncan be shifted outwards from the black hole.\nAs pointed out in Section 3, depending on the direc-\ntion of the Lorentz force one can distinguish two qualita-\ntively different types of the motion. In a radiating case\none can see such differences as well. The motion along\na general worldline highly depends on the initial energy\nand the position of the particle. Radiation effect on the\ncharged particle motion also has different timescales de-\npending on whether the particle is initially oscillating or\nnot. The representative comparison of trajectories of os-\ncillating charged particle in the presence and absence of\nradiation reaction force is illustrated in Fig. 3for attrac-\ntive Lorentz force and in Fig. 4, for repulsive one. Note\nthat the trajectories represented in Figs. 3and4are\nplotted by integration of the full form of equations of\nmotion ( 43) and (47) without restriction to the equato-\nrial plane. In both cases, the motion of the charge is\ninitially bounded. The charged particle starts its motion\nslightly abovethe equatorialplane (keeping regularchar-\nacter of the motion) in the vicinity of weakly magnetized\nSchwarzschild black hole. The initial energy and angular\nmomentum of the particle correspond to the state close,\nbut above the minimum of the effective potential which\ngenerates barrier where the particle bounces. Due to the\nactionoftheradiationreactionforce,thechargedparticle\nchanges its oscillatory characterof the motion because ofthe loss of the energy and angular momentum. The final\nstate of the charged particle depends on the direction of\nthe Lorentz force. For the Lorentz force directed towards\nthe horizon of the black hole, the synchrotron radiation\nleads to the collapse of the initially stable particle into\nthe black hole (see Fig. 3). In the inverse case, when the\nLorentz force is directed outwards the horizon (compen-\nsating thus the ”gravitational attraction”), the particle\nremainsin the bounded regionin the vicinity ofthe black\nhole. On the other hand, loss of the energy of the par-\nticle leads to the disappearance of curled trajectories as\nwell as any oscillations, as shown in Fig. 4.\nAnother representative example of the dependence of\nthetrajectoriesonthealignmentoftheexternalmagnetic\nfield is shown in Figs. 5and6, where the initial condi-\ntions differ only in the sign of the magnetic parameter B,\nwhilethe otherparametersarechosentobe the same. As\nbefore, for the attractive Lorentz force, the charged par-\nticle spirals down to the black hole (Fig. 5), while in the\nrepulsive case, the motion is stable (Fig. 6). Stability of\ntheplus configurations against collapse can be explained\ndue to the reason that a radiating particle is assumed to\nbe accelerated by the external Lorentz force only. This\nimplies that the radiation reaction in fact decreases the\ngyroradiusof the radiatingparticle. In the casewhen the\nLorentz force is directed outwards the black hole, the gy-\nrocenter of oscillating charged particle is located near its\norbit (center of ”curls”), while in the case of minus con-\nfiguration , the gyrocenter of the orbit is located inside\nthe horizon (it coincides with a black hole singularity for\nthe circular motion). Indeed, from the analogy with the\nflat spacetime formula ( 5), the sign of the radiation reac-\ntion force depends on the alignment of the Lorentz force\ndue to the reduction of order procedure performed in the\nbeginning. Thus, the ”Lorentz-repulsive”case with radi-\nationreactionrepresentsdamping ofoscillations, turning\nthe oscillatory epicyclic type of motion to nearly circu-\nlar orbit. A strong non-linearity of equations of motion\nshifts the location of stable circular orbits as well, how-\never, as we will see below, this effect is relatively slow in\ncomparisontotheprocessofradiativedampingofoscilla-\ntions. One can conclude that for relatively short periods\nof time an oscillating charged particle, undergoing repul-\nsive Lorentz force and radiation reaction force will damp\nits oscillations settling down to the circular orbit. This\nresult is represented in Fig. 7, being in accord with the\nqualitative predictions given in Shoom (2015).\n5.1.Evolution of energy and angular momentum\nIn fact, the synchrotron radiation carries out the ki-\nnetic energy of the particle. One can see in E−τplots\nof Fig.5and Fig.6, that the energy slightly decreases in\nboth cases. However, the azimuthal angular momentum\nof the particle changes differently in dependence on the\ndirection of the Lorentz force. While in the minus con-\nfigurations the charged particle decreases its angular mo-\nmentum due to radiation, in the plus configurations the\nangular momentum quasi-periodically increases. Quasi-\nperiodicity is connected with the quasi-harmonic oscilla-\ntions of the particle in the motion with ”curls”. Increase\nof the angular momentum of the particles in the plus\nconfigurations becomes clear if one compares the angular\nmomentum of charged particles in the motion with curls,\nL∗given by ( 34), with those of the circular orbits given10\nFig. 3.— Representative comparison of trajectories of non-radiati ng (first line) and radiating (second line) charged particle s withminus\nconfiguration corresponding to the attractive Lorentz force from differen t view angles around black hole in magnetic field. Initial con ditions\nin both cases are chosen to be same and shown in the second row p lots. Starting point is indicated by black dot. A radiating c harged\nparticle escapes the initial boundary of the motion (dashed contours in the third row) governed by an effective potential (28) and collapses\nto the black hole due to the loss of angular-momentum. Magnet ic field is aligned with z- axis. Trajectory of radiating charged particle\ncorresponds to the integration of full set of equations of mo tion (43) and (47) without fixing the plane of the motion.\nFig. 4.— Similar to Fig.3 comparison of trajectories of non-radiati ng and radiating charged particles in case of repulsive Lore ntz force,\ni.e.plus configuration . A radiating particle stays in the region of initial boundar y of the motion and radiates its oscillatory part of\nenergy-momentum tending to the stable circular orbit.11\nFig. 5.— Spirallingdown to the black holeof charged particle due to r adiationreaction ofa minus configuration particle, and corresponding\nevolution in local time τof the radius of orbit r, angular momentum L, energy E, the radial velocity ur, the angular azimuthal velocity uφ\nand gamma factor ut≡dt/dτ≡γ. The angular momentum and energy are measured with respect t o observer at rest at infinity. Starting\npoint of the particle is indicated as black dot.\nFig. 6.— Decay of oscillations due to radiation reaction of a plus configuration particle, and corresponding evolution in time τof the radius\nof orbit r, angular momentum L, energy E, the radial velocity ur, the angular azimuthal velocity uφand gamma factor ut≡dt/dτ≡γ.\nThe angular momentum and energy are measured with respect to observer at rest at infinity. Starting point of the particle i s indicated as\nblack dot.\nFig. 7.— Damping of oscillations due to radiation reaction of a plus configuration particle, and transition from curled motion to the\nnearly circular orbit. Corresponding evolution of the radi us of orbit r, angular momentum L, energy E, the radial velocity ur, the angular\nazimuthal velocity uφand gamma factor ut≡dt/dτ≡γare represented as in previous plots. The angular momentum a nd energy are\nmeasured with respect to observer at rest at infinity. Starti ng point of the particle is indicated as black dot.12\nFig. 8.— Number of revolutions of the oscillatingcharged particle\naround black hole given as a ratio of the maximal decay time to\nthe orbital time in dependence on the magnetic parameter Band\ndifferent values of radiation parameter k.\nby Koloˇ s et al. (2015)\nLc=−Br2+r/radicalbig\nB2r2(r−2)2+r−3\nr−3.(54)\nLet us now assume that the initial angular momentum of\nthe particle is ≈ L∗, while the angular momentum at the\nfinal state is nearly equal to those of the stable circular\norbit,≈ Lc. From the fact that that Lc>L∗(see details\nin Koloˇ s et al. (2015)) one can conclude that the parti-\ncle is actually gaining an angular momentum by reduc-\ning radial oscillations due to radiation reaction force. In\ntheminus configurations , the trajectories with curls, i.e.\nthe regions with negative angular velocity cannot be ob-\nserved. Therefore, a charged particle initially located at\na bounded orbit around a black hole continuously loses\nits angular momentum due to radiation reaction force.\nWhen angular momentum of the particle becomes lower\nthan the one corresponding to the innermost stable cir-\ncular orbit (ISCO), L20 s\u00001), wave re\rection occurs mainly at\nthe entrance to the damping zone. Thus Hmean continually decreases\nwhen increasing f1, since less of the incoming wave can pass through\nthe damping zone, so the water surface will be virtually \rat near the\ndomain boundary. This is also visible later in Figs. 13 from Sect.\n13. The aforementioned increase in the amount of wave re\rections\ndetectable in the solution domain can be seen in the curve for CR. This\nshows that, for a given fade-in function and damping layer thickness,\n15there is an optimum for f1so that the wave re\rections propagating\nback into the solution domain are minimized. The best damping was\nachieved for f1= 10 s\u00001, in which case the e\u000bects of wave re\rections\non the wave height within the solution domain are less than 0 :7%.\nFigure 5: Mean wave height Hmean recorded at x= 5:75\u0015scaled by twice the height Hof\nthe undamped wave and re\rection coe\u000ecient CRover damping coe\u000ecient f1, whilef2= 0\nThis provides the following conclusions: Hmean is not a suitable indicator\nfor damping quality if the damping is stronger than optimal, however it is\nuseful to characterize where the re\rections originate from: if CRshows that\nnoticeable re\rections are present within the solution domain, while at the\nsame timeHmeanis negligibly small, then the re\rections cannot occur at the\ndomain boundary, but must occur closer to the entrance to the damping layer.\nSigni\fcant re\rections ( CR>2%) in form of partial standing waves appear for\nroughlyf1<5 s\u00001andf1>80 s\u00001. The height of the partial standing wave\nincreases the more f1deviates from the regime where satisfactory damping\nis observed; however, the increase is slower if the damping is stronger than\noptimal instead of weaker.\n9. Variation of Damping Coe\u000ecient for Quadratic Damping\nTo investigate which range for the quadratic damping coe\u000ecient accord-\ning to Eq. (5) produces satisfactory wave damping, waves with wavelength\n16\u0015= 4 m and height 0 :16 m are investigated. Only quadratic damping is con-\nsidered, so f1= 0. Simulations are performed for damping coe\u000ecient f2\nbetweenf22[0:625 m\u00001;10240 m\u00001]. The recorded surface elevations near\nthe end of the damping layer in Fig. 6 show that the damping is strongly\nin\ruenced by the choice of f2, while the wave phase and period remain nearly\nunchanged.\nFigure 6: Surface elevation scaled by height Hof the undamped wave over time scaled by\nthe wave period T; recorded at x= 5:75\u0015for simulations with f22[0:625 m\u00001;10240 m\u00001]\nandf1= 0\nFigure 7 shows how a variation of f2a\u000bects the re\rection coe\u000ecient CR\nand the mean wave height Hmeanrecorded in close vicinity to the boundary\nto which the damping layer is attached.\nThe results show the same trends as the ones in Sect. 8. For the given\nfade-in function and damping layer thickness, there is an optimum for f2\nso that the wave re\rections propagating back into the solution domain are\nminimized. The best damping was achieved for f2= 160 m\u00001, in which case\nthe e\u000bects of wave re\rections on the wave height within the solution domain\nare less than 0 :7%. The e\u000bects of wave re\rections increase the further f2\ndeviates from the optimum. For smaller f2values, the waves are re\rected\nmainly at the domain boundary, for larger f2values the re\rection occurs\nmainly at the entrance to the damping zone. Signi\fcant re\rections ( CR>\n2%) in form of partial standing waves appear for roughly f2<80 m\u00001and\nf2>640 m\u00001.\n17Figure 7: Mean wave height Hmean recorded at x= 5:75\u0015scaled by twice the height Hof\nthe undamped wave and re\rection coe\u000ecient CRover damping coe\u000ecient f2, whilef1= 0\nCompared to the linear damping functions from the previous section, the\nuse of quadratic damping functions does not o\u000ber a signi\fcant improvement\nin damping quality. With an optimal setup, both approaches provide roughly\nthe same damping quality if the setup is optimized. However, the range of\nwave frequencies that are damped satisfactorily is narrower for quadratic\ndamping.\n10. In\ruence of Computational Mesh on Achieved Damping\nThe simulations from Sect. 8 are rerun with same setup except for a grid\ncoarsened by factor 2. Thus whereas the \fne mesh simulations discretize the\nwave with 100 cells per wavelength and 16 cells per wave height, the coarse\nmesh simulations have 50 cells per wavelength and 8 cells per wave height.\nAll coarse grid re\rection coe\u000ecients di\u000ber from their corresponding \fne grid\nre\rection coe\u000ecient by CR;coarse =CR;\fne\u00060:8%. This is also visible in Fig.\n8. Thus for su\u000ecient resolution, the damping e\u000bectiveness can be considered\ngrid-independent. This is expected since the damping-related terms in Eqs.\n(2) to (4) do not depend on cell volume. The used grids in this study are\nthus adequate.\n18Figure 8: Re\rection coe\u000ecient CRover damping coe\u000ecient f1for coarse and \fne mesh\nsimulations, with f2= 0\n11. In\ruence of the Thickness of the Damping Layer\nThe simulation for \u0015= 4:0 m,H= 0:16 m andf1= 10 s\u00001from Sect.\n8 is repeated for xd= 0\u0015;0:25\u0015;0:5\u0015;0:75\u0015;1:0\u0015;1:25\u0015;1:5\u0015;2:0\u0015;2:5\u0015with\notherwise the same setup. The evolution of the free surface elevation in the\ntank over time in Fig. 9 shows that xdhas a strong in\ruence on the achieved\ndamping. Setting xd= 0\u0015deactivates the damping and produces at \frst\na nearly perfect standing wave, which then degenerates due to the in\ru-\nence of the pressure outlet boundary, since prescribing hydrostatic pressure\nestablishes an oscillatory in-/out\row of water through this boundary which\ndisturbs the standing wave. For xd= 0:5\u0015, a strong partial standing wave oc-\ncurs, and for xd= 1:0\u0015only slight re\rections are still observable CR\u00191:6%.\nFor largerxd, the in\ruence of wave re\rections continues to decrease. This is\nevident from the plot of CRfor the simulations shown in Fig. 10.\n19Figure 9: Free surface elevation over x-location in tank shown for 40 equally spaced\ntime instances over one period starting at t= 16 s; from top to bottom xd=\n0\u0015;0:5\u0015;1:0\u0015;1:5\u0015;2:0\u0015;2:5\u0015; the damping zone is depicted as shaded gray\n20Figure 10: Re\rection coe\u000ecient CRover damping zone thickness xd\nSubsequently, the simulations from Sect. 8 have been rerun with the\ndamping thickness set to xd= 1\u0015. Comparing the resulting curves for CR\noverf1shows that increasing xdnot only improves the damping quality; it\nalso widens the range of damping coe\u000ecients for which satisfactory damping\nis obtained; thus the wave damping then becomes less sensitive to !the more\nxdincreases. However, this also increases the computational e\u000bort.\nFigure 11: Re\rection coe\u000ecient CRover damping coe\u000ecient f1forxd= 1\u0015andxd= 2\u0015,\nwhilef2= 0\n21The results show that, if the damping coe\u000ecients are set up close to the\noptimum and it is desired that CR<2%, thenxd= 1\u0015su\u000eces. This knowl-\nedge is useful, since by reducing xdthe computational domain can be kept\nsmaller and thus the computational e\u000bort can be reduced. However, if better\ndamping is desired or when complex \row phenomena are considered, espe-\ncially when irregular waves or wave re\rections from bodies are present, then\nthe damping layer thickness should be increased to damp all wave compo-\nnents successfully. The present study suggests that a damping layer thickness\nof 1:5\u0015\u0014xd\u00142\u0015can be recommended.\n12. In\ruence of Wave Steepness on Achieved Damping\nThe simulations in this section are based on those from Sect. 8, i.e.\n\u0015= 4:0 m,H= 0:16 m and varying f1in the range [0 :625 s\u00001;1000 s\u00001]. The\nsimulations were rerun with the same setup except for two modi\fcations: The\nwave height was changed to H= 0:4 m, resulting in a steepness of H=\u0015 = 0:1\ninstead of the previous H=\u0015 = 0:04. Furthermore, the grid was adjusted to\nmaintain the same number of cells per wave height as well as per wavelength,\nso that both results are comparable.\nAs can be seen from Figure 12, the in\ruence of the increased wave steep-\nness is comparatively small, except for the cases with signi\fcantly smaller\nthan optimum damping ( f1\u00142:5 s\u00001). For the rest of the range, i.e. 5 \u0014f1\u0014\n1000 s\u00001, the di\u000berence in re\rection coe\u000ecients is only CR(H=\u0015 = 0:1) =\nCR(H=\u0015 = 0:04)\u00061:7%. Therefore, although the damping performs slightly\nbetter for waves of smaller steepness, the in\ruence of wave steepness can be\nassumed negligible for most practical cases. If stronger wave steepness is con-\nsidered and less uncertainty is required, then the thickness of the damping\nlayer can be further increased to decrease CR.\n22Figure 12: Re\rection coe\u000ecient CRover damping coe\u000ecient f1for waves of same period\nbut di\u000bering steepness H=\u0015 = 0:04 andH=\u0015 = 0:1, whilef2= 0\n13. The Scaling Law for Linear Damping\nIn order to verify the assumed scaling law for linear damping from Sect.\n5, the simulation setup with the best damping performance ( f1= 10:0 s\u00001)\nfrom Sect. 8 was scaled geometrically and kinematically so that the generated\nwaves are completely similar, for wavelengths \u0015= 0:04 m;4 m;400 m and thus\ncorresponding heights of H= 0:0016 m;0:16 m;16 m. This corresponds to a\nrealistic scaling, since geometrically scaling by 1 : 100 is common in both\nexperimental and computational model- and full scale investigations. As\nnecessary requirement to obtain similar damping (i.e. similarity of CRand\nsurface elevation), the damping length xdis scaled directly proportional to\nthe wavelength, according to Eq. (8).\nAt \frst, the simulations are run with no scaling of f1, so thatf1= 10 s\u00001in\nall cases. Figure 13 shows the free surface in the tank after the simulations. In\nthe vicinity of the inlet (0 ! p. The wave spectrum was discretized into 100 components.\nThe choice of xd= 2\u0015peak, with wavelength \u0015peakcorresponding to the\npeak wave frequency, seems adequate for such cases. No visible disturbance\ne\u000bects were noted in the simulation. The wave heights over time are shown\nrecorded close to the wave-maker in Fig. 17 and recorded close to the bound-\nary to which the damping zone is attached in Fig. 18. This shows that the\naverage wave height is reduced by roughly two orders of magnitude.\nFigure 17: Surface elevation over time recorded directly before the wave-maker\n31Figure 18: Surface elevation over time recorded in close vicinity to the boundary, to which\nthe damping layer is attached\nFrom the surface elevation in the tank in Fig. 19 no undesired wave\nre\rections can be noticed. Although a more detailed study of the error of\nthis approximation regarding di\u000berent parameters of the spectrum was not\npossible in the scope of this study, the proposed approach seems to work\nreasonably well for practical purposes. Further research in this respect is\nrecommended to reduce uncertainties regarding the re\rections.\nFigure 19: Surface elevation in the whole domain at t\u001915:6 s\n16. Application of Scaling Law to Ship Resistance Prediction\nThis section compares results from model and full scale resistance compu-\ntations in 3D of the Kriso Container Ship (KCS) at Froude number 0 :26. The\n32simulations in this section are based on the simulations reported in detail in\nEnger et al. (2010). For detailed information on the setup and discretiza-\ntion, the reader referred to Enger et al. (2010). In the following, only a brief\noverview of the setup and di\u000berences to the original simulations are given for\nthe sake of brevity. The present model scale simulation di\u000bers only in the\ndamping setup (and slight modi\fcations of the used grid) from the \fne grid\nsimulation in Enger et al. (2010). The KCS is \fxed in its \roating position\nat zero speed. To simulate the hull being towed at speed U, this velocity\nis applied at the inlet domain boundaries and the hydrostatic pressure of\nthe undisturbed water surface is applied at the outlet boundary behind the\nship. The domain is initialized with a \rat water surface and \row velocity U\nfor all cells. As time accuracy is not in the focus here, \frst-order implicit\nEuler scheme is used for time integration. Apart from the use of the k-\u000f\nturbulence model by Launder and Spalding (1974), the computational setup\nis similar to the one used in the rest of this work. The computational grid\nconsists of roughly 3 million cells. For comparability, we compare only the\npressure components of the drag and vertical forces, which are obtained by\nintegrating the x-component for the drag and z-component for the vertical\ncomponent of the pressure forces over the ship hull. The simulation starts\nat 0 s and is stopped at tmax= 90 s simulation time. The Kelvin wake of the\nship can be decomposed into a transversal and a divergent wave component.\nThe wave damping setup is based on the ship-evoked transversal wave (wave-\nlength\u0015t\u00193:1 m), the phase velocity of which equals the service speed U.\nWave damping according to Eqs. (5) and (6) has been applied to inlet, side\nand outlet boundaries with parameters xd= 2:3\u0015t,f1= 22:5 s\u00001,f2= 0,\nandn= 2. This setup provided satisfactory convergence of drag and vertical\nforces in model scale. The wake pattern for the \fnished simulation is shown\nin Fig. 20 and the results are in agreement with the \fndings from Enger et\nal. (2010). The obtained resistance coe\u000ecient CT;sim= 3:533\u000110\u00003compares\nwell with the experimental data ( CT;exp= 3:557\u000110\u00003, 0:68% di\u000berence to\nCT;sim) by Kim et al. (2001) and to the simulation results by Enger et al.\n(2001) (CT;Enger = 3:561\u000110\u00003, 0:11% di\u000berence to CT;exp).\n33Figure 20: Wave pro\fle for model scale ship with f1= 22:5 s\u00001att= 90 s\nAdditionally, full scale simulations are performed with Froude similarity.\nThe scaled velocity and ship dimensions are shown in Table 4. The grid is\nsimilar to the one for the model scale simulation except scaled with factor\n31:6. Assuming similar damping can be obtained with the presented scaling\nlaws,xdwas scaled according to Eq. 8 by factor 31 :6 as well. Otherwise the\nsetup corresponds to the one from the model scale simulation.\n34Table 4: KCS parameters\nscale waterline length L(m) service speed U(m=s)\nmodel 7:357 2 :196\nfull 232:5 12 :347\nFigure 21 shows drag and vertical forces over time when both model and\nfull scale simulations are run with the same value for damping coe\u000ecient\nf1. In contrast to the model scale forces, the full scale forces oscillate in a\ncomplicated fashion. Therefore without a proper scaling of f1, no converged\nsolution can be obtained for the full scale case.\n35Figure 21: Drag (top image) and vertical (bottom image) pressure forces on ship over time\nfor model (red) and full scale (grey); no scaling of f1, thusf1= 22:5 s\u00001is the same in\nboth simulations\nFinally, the full scale simulation is rerun with f1scaled according to Eq. 9\nto obtain similar damping as in the model scale simulation with f1= 22:5 s\u00001.\nThe correctly scaled value for the full scale simulation is thus f1;full=f1;model\u0001\n!full=!model = 22:5 s\u00001\u00011=p\n31:6\u00194 s\u00001. The resulting drag and vertical\nforces in Fig. 22 show that indeed similar damping is obtained, since in\nboth cases the forces converge in a qualitatively similar fashion. Note that a\nperfect match of the curves in Fig. 22 is not expected, since Froude-similarity\nis given, but not Reynolds-similarity. Thus although the pressure components\nof the forces on the ship will converge to the same values (if scaled by L3), the\nway they converge (amplitude and frequency of the oscillation) may not be\n36exactly similar, since this depends on the solution of all equations. However,\nthe tendency of the convergence will be qualitatively similar as shown in the\nplots.\nFigure 22: Drag (top image) and vertical (bottom image) pressure forces on ship over time\nfor model and full scale; the damping setup for the full scale simulation is obtained by\nscaling the model scale setup according to Eqs. 8 and 9\n17. Discussion and Conclusion\nIn order to obtain reliable wave damping with damping layer approaches,\nthe damping coe\u000ecients must be adjusted according to the wave parameters,\nas shown in Sects. 8, 9, 13 and 14. It is described in Sect. 7 that the\nprocedure to quantify the damping quality is quite e\u000bortful, and thus it\n37is seldom carried out in practice. This underlines the importance of the\npresent \fndings for practical applications, since unless the damping quality\ncan be reliably set to ensure that the in\ruence of undesired wave re\rections\nare small enough to be neglected, a large uncertainty will remain in the\nsimulation results. The optimum values for damping coe\u000ecients f1andf2\ncan be assumed not to depend on computational grid, wave steepness and\nthicknessxdof the damping layer as shown in Sects. 10, 11 and 12. As shown\nin Sect. 11, the damping layer thickness has the strongest in\ruence on the\ndamping quality. If it increases, the range of waves that will be damped\nsatisfactorily broadens and the re\rection coe\u000ecient for the optimum setup\nshrinks; unfortunately, the computational e\u000bort increases at the same time as\nwell, thus optimizing the damping setup is important. In contrast, Sect. 12\nshows that the wave steepness has a smaller e\u000bect on the damping, with the\ntendency towards better damping for smaller wave steepness. For su\u000eciently\n\fne discretizations, Sect. 10 shows that the damping can be considered not\na\u000bected by the grid. A practical approach for e\u000ecient damping of irregular\nwaves has been presented in Sect. 15. The scaling laws in Sect. 5 and\nrecommendations given in Sects. 13 and 14 provide a reliable way to set\nup optimum wave damping for any regular wave. Moreover, similarity of\nthe wave damping can be guaranteed in model- and full-scale simulations as\nshown in Sects. 13, 14 and 16. The \fndings can easily be applied to any\nimplementation of wave damping which accords to Sect. 3.\nReferences\n[1] Cao, Y., Beck, R. F. and Schultz, W. W. 1993. An absorbing beach for\nnumerical simulations of nonlinear waves in a wave tank, Proc. 8th Intl.\nWorkshop Water Waves and Floating Bodies , 17-20.\n[2] Choi, J. and Yoon, S. B. 2009. Numerical simulations using momentum\nsource wave-maker applied to RANS equation model, Coastal Engineer-\ning, 56, (10), 1043-1060.\n[3] Cruz, J. 2008. Ocean wave energy, Springer Series in Green Energy and\nTechnology , UK, 147-159.\n[4] Enger, S., Peri\u0013 c, M. and Peri\u0013 c, R. 2010. Simulation of Flow Around KCS-\nHull, Gothenburg 2010: A Workshop on CFD in Ship Hydrodynamics,\nGothenburg.\n38[5] Fenton, J. D. 1985. A \ffth-order Stokes theory for steady waves, J.\nWaterway, Port, Coastal and Ocean Eng. , 111, (2), 216-234.\n[6] Ferrant, P., Gentaz, L., Monroy, C., Luquet, R., Ducrozet, G., Alessan-\ndrini, B., Jacquin, E., Drouet, A. 2008. Recent advances towards the\nviscous \row simulation of ships manoeuvring in waves, Proc. 23rd Int.\nWorkshop on Water Waves and Floating Bodies , Jeju, Korea.\n[7] Ferziger, J. and Peri\u0013 c, M. 2002. Computational Methods for Fluid Dy-\nnamics, Springer.\n[8] Grilli, S. T., Horrillo, J. 1997. Numerical generation and absorption of\nfully nonlinear periodic waves. J. Eng. Mech. , 123, (10), 1060-1069.\n[9] Guignard, S., Grilli, S. T., Marcer, R., Rey, V. 1999. Computation of\nshoaling and breaking waves in nearshore areas by the coupling of BEM\nand VOF methods, Proc. ISOPE1999 , Brest, France.\n[10] Ha, T., Lee, J.W. and Cho, Y.S. 2011. Internal wave-maker for Navier-\nStokes equations in a three-dimensional numerical model, J. Coastal\nResearch , SI 64, 511-515.\n[11] Hasselmann, K., Barnett, T. P., Bouws, E., Carlson, H., Cartwright,\nD. E., Enke, K., Ewing, J. A., Gienapp, H., Hasselmann, D. E., Kruse-\nman, P., Meerburg, A., M uller, P., Olbers, D. J., Richter, K., Sell, W.\nand Walden, H. 1973. Measurements of wind-wave growth and swell de-\ncay during the Joint North Sea Wave Project (JONSWAP), Deutsche\nHydrographische Zeitschrift , (8), Reihe A.\n[12] Higuera, P., Lara, J. L., Losada, I. J. 2013. Realistic wave generation\nand active wave absorption for Navier-Stokes models Application to\nOpenFOAM r,Coastal Eng. , 71, 102-118.\n[13] Jha, D. K. 2005. Simple Harmonic Motion and Wave Theory, Discovery\nPublishing House.\n[14] Kim, J., O'Sullivan, J., Read, A. 2012. Ringing analysis of a vertical\ncylinder by Euler overlay method, Proc. OMAE2012 , Rio de Janeiro,\nBrazil.\n39[15] Kim, J., Tan, J. H. C., Magee, A., Wu, G., Paulson, S., Davies, B.\n2013. Analysis of ringing ringing response of a gravity based structure\nin extreme sea states, Proc. OMAE2013 , Nantes, France.\n[16] Kim, W. J., Van, S. H., Kim, D. H. 2001. Measurement of \rows around\nmodern commercial ship models, Experiments in Fluids , 31, Springer-\nVerlag, 567-578.\n[17] Kraskowski, M. 2010. Simulating hull dynamics in waves using a RANSE\ncode, Ship technology research , 57, 2, 120-127.\n[18] Lal, A., Elangovan, M. 2008. CFD simulation and validation of \rap type\nwave-maker, WASET , 46, 76-82.\n[19] Launder, B.E., and Spalding, D.B. 1974. The numerical computation of\nturbulent \rows, Comput. Meth. Appl. Mech. Eng. , 3, 269-289.\n[20] Lloyd, A. R. J. M. 1989. Seakeeping: Ship Behaviour in Rough Weather,\nEllis Horwood Limited.\n[21] Muzaferija, S. and Peri\u0013 c, M. 1999. Computation of free surface \rows us-\ning interface-tracking and interface-capturing methods, Nonlinear Water\nWave Interaction, Chap. 2, 59-100, WIT Press, Southampton.\n[22] Park, J. C., Kim, M. H. and Miyata, H. 1999. Fully non-linear free-\nsurface simulations by a 3D viscous numerical wave tank, International\nJournal for Numerical Methods in Fluids , 29, 685-703.\n[23] Peri\u0013 c, R. 2013. Internal generation of free surface waves and application\nto bodies in cross sea, MSc Thesis, Schriftenreihe Schi\u000bbau, Hamburg\nUniversity of Technology, Hamburg, Germany.\n[24] Peri\u0013 c, R., Abdel-Maksoud, M., 2015. Assessment of uncertainty due to\nwave re\rections in experiments via numerical \row simulations, Proc.\nISOPE2015 , Hawaii, USA.\n[25] Sch a\u000ber, H.A., Klopman, G. 2000. Review of multidirectional active\nwave absorption methods, Journal of Waterway, Port, Coastal, and\nOcean Engineering , 88-97.\n40[26] Ursell, F., Dean, R. G. and Yu, Y. S. 1960. Forced small-amplitude\nwater waves: a comparison of theory and experiment, Journal of Fluid\nMechanics , 7, (01), 33-52.\n[27] W ockner-Kluwe, K. 2013. Evaluation of the unsteady propeller perfor-\nmance behind ships in waves, PhD thesis at Hamburg University of\nTechnology, Schriftenreihe Schi\u000bbau , 667, Hamburg.\n41" }, { "title": "2010.14148v2.Legolas__a_modern_tool_for_magnetohydrodynamic_spectroscopy.pdf", "content": "Draft version June 2, 2021\nTypeset using L ATEX default style in AASTeX63\nLegolas: a modern tool for magnetohydrodynamic spectroscopy\nNiels Claes,1Jordi De Jonghe,1and Rony Keppens1\n1Centre for mathematical Plasma-Astrophysics, Celestijnenlaan 200B, 3001 Leuven, KU Leuven, Belgium\nABSTRACT\nMagnetohydrodynamic (MHD) spectroscopy is central to many astrophysical disciplines, ranging\nfrom helio- to asteroseismology, over solar coronal (loop) seismology, to the study of waves and insta-\nbilities in jets, accretion disks, or solar/stellar atmospheres. MHD spectroscopy quanti\fes all linear\n(standing or travelling) wave modes, including overstable (i.e. growing) or damped modes, for a\ngiven con\fguration that achieves force and thermodynamic balance. Here, we present Legolasa), a\nnovel, open-source numerical code to calculate the full MHD spectrum of one-dimensional equilibria\nwith \row, that balance pressure gradients, Lorentz forces, centrifugal e\u000bects and gravity, enriched\nwith non-adiabatic aspects like radiative losses, thermal conduction and resistivity. The governing\nequations use Fourier representations in the ignorable coordinates, and the set of linearised equations\nare discretised using Finite Elements in the important height or radial variation, handling Cartesian\nand cylindrical geometries using the same implementation. A weak Galerkin formulation results in a\ngeneralised (non-Hermitian) matrix eigenvalue problem, and linear algebraic algorithms calculate all\neigenvalues and corresponding eigenvectors. We showcase a plethora of well-established results, rang-\ning from p- and g-modes in magnetised, strati\fed atmospheres, over modes relevant for coronal loop\nseismology, thermal instabilities and discrete overstable Alfv\u0013 en modes related to solar prominences, to\nstability studies for astrophysical jet \rows. We encounter (quasi-)Parker, (quasi-)interchange, current-\ndriven and Kelvin-Helmholtz instabilities, as well as non-ideal quasi-modes, resistive tearing modes, up\nto magneto-thermal instabilities. The use of high resolution sheds new light on previously calculated\nspectra, revealing interesting spectral regions that have yet to be investigated.\n1.INTRODUCTION\nThe study of stability for plasmas and \ruids alike has been a major topic of research over the last century. Under-\nstanding how and why a given medium reacts to a linear perturbation is of central importance to many astrophysical\nphenomena. In incompressible or compressible \ruids, governed by hydrodynamic equations, notable instabilities are\nthe Kelvin-Helmholtz instability (KHI), which arises due to a velocity shear at the interface of two \ruids, and the\nRayleigh-Taylor instability (RTI) where gravitational strati\fcation can lead to an unstable con\fguration of layered\n\ruids of di\u000berent density (Chandrasekhar 2013; Choudhuri 1998). In plasmas, governed by the magnetohydrodynamic\n(MHD) equations, the study of waves and instabilities becomes much richer due to the inclusion of magnetic \felds,\nwith a modern overview provided in Goedbloed et al. (2019). Magnetic \felds modify the two aforementioned insta-\nbilities in various ways, and the combination of \row, magnetic \felds and pressure gradients introduces many new\nmodes, e.g. the magnetorotational instability (Balbus & Hawley 1991) relevant for (weakly magnetised) accretion\ndisks or the Trans-Slow-Alfv\u0013 en Continuum modes in disks of arbitrary magnetisation (Goedbloed et al. 2004). In the\nhighly magnetised solar corona, observed coronal loop oscillations (periods and damping times) are routinely used\nto infer loop parameters like their \feld strength (Nakariakov & Ofman 2001). Embedded in the hot solar corona,\nwe \fnd stable and long-lived quiescent prominences, with internal dynamics due to KHI (Hillier & Polito 2018) and\nRTI (Hillier 2018). The formation of prominences is due to the thermal instability (TI), as demonstrated in direct\nobservations by e.g. Berger et al. (2012) or in simulations by Xia & Keppens (2016); Claes et al. (2020). Together with\ncategorising all instabilities, knowing the stable eigenoscillations such as p-modes or g-modes in strati\fed atmospheres\nor stellar interiors, is of prime importance to link theoretical understanding with observed periodic phenomena. In\nall of these cases, we need to compute eigenoscillations and corresponding eigenfunctions from the linearised set of\ngoverning equations. Linear MHD spectroscopy, which encompasses the entirety of helio- and asteroseismology, but\na)The Legolas code is available on GitHub: https://github.com/n-claes/legolasarXiv:2010.14148v2 [astro-ph.SR] 1 Jun 20212\nincorporates laboratory fusion plasma MHD spectroscopy (Goedbloed et al. 1993), MHD spectroscopy of accretion\ndisks (Keppens et al. 2002) and jets, as well as solar coronal seismology (Roberts 2019), is thus a powerful tool for\nstudying many astrophysical processes.\nSince the advent of more powerful computational resources, the main focus of computational astrophysical research\nhas gradually shifted towards solving the fully non-linear MHD equations, where many non-adiabatic/non-ideal e\u000bects\nare incorporated, depending on the application at hand. While this approach successfully reproduced many physical\nphenomena, especially for realistic solar setups (e.g. sunspots (Rempel 2012), \rares (Ruan et al. 2019), or promi-\nnences (Xia & Keppens 2016)), it usually fails to answer which speci\fc perturbation produces the complex evolution\nas witnessed. At the same time, theoretical insight showed that MHD spectral theory actually governs the stabil-\nity of \rowing, (self-)gravitating single \ruid evolutions of nonlinear, time-dependent plasmas, and this at any time\nduring their nonlinear evolution (Demaerel & Keppens 2016). Hence, in order to predict the reaction of a certain\nphysical state to perturbations, we should really quantify all its waves and instabilities using linear theory. This has\nbeen recognised fully in laboratory fusion plasmas, where MHD spectroscopy is very successful for identifying waves\nand stability aspects of a given toroidal Grad-Shafranov equilibrium. That this can meaningfully be done for states\nthat include important non-adiabatic e\u000bects, like optically thin radiative losses, is important for investigations into\nprominences and their intriguing \fne structure, as revealed by means of direct observations (Engvold 1998; Mackay\net al. 2010; Ballester 2006) or through numerical simulations (Xia & Keppens 2016; Xia et al. 2017; Claes et al. 2020).\nIn that context, early analytical work by Van der Linden & Goossens (1991a) based on linear MHD suggests the\nhypothesis that \fnite perpendicular thermal conduction induces \fne structure in unstable linear eigenmodes. Since\nthis pioneering work of Van der Linden & Goossens (1991a), not much research has been done regarding the full MHD\nspectrum when non-adiabatic e\u000bects are at play, for the simple reason that to date, there existed no numerical tool to\nsolve the full system of linearised MHD equations with all the physical e\u000bects included. This is why we developed the\nnew and open-source Legolas solver.\nLegolas builds on the heritage of early numerical codes, most notably LEDA (Kerner et al. 1985), which allowed\nstudies of the ideal or resistive MHD spectrum for laboratory plasmas, approximated by a di\u000buse cylindrical plasma\ncolumn (or \rux tube) and CASTOR (Kerner et al. 1998), which applied to resistive spectra of general tokamak\ncon\fgurations. The latter has follow-up codes such as FINESSE (Beli en et al. 2002) and PHOENIX (Blokland\net al. 2007b), extending it to stationary and axisymmetric truly 2D con\fgurations. LEDA was later extended in\nVan der Linden et al. (1992), where non-adiabatic e\u000bects like anisotropic thermal conduction and optically thin\nradiative losses were added to the equations, using a simple analytic function to treat radiative cooling e\u000bects. A\ndi\u000berent branch of LEDA, called LEDAFLOW (Nijboer et al. 1997), was developed to investigate the resistive MHD\nspectrum, augmented with gravitational and \row e\u000bects, but omitting those non-adiabatic terms. Since these codes\nwere developed decades ago and focus shifted away from linear MHD, their further development was stalled, although\nin laboratory fusion context, tools to compute multidimensional equilibria and their linear modes are very important\nfor diagnosing experiments. The original codes, like LEDA(FLOW), were not \rexible, in the sense that adding di\u000berent\nequilibria or accounting for additional terms in the equations would be a major undertaking, as parts were hard-coded\nto (limited) computational resources of that time. Furthermore, programming languages and numerical tools like\nLAPACK (Anderson et al. 1999) to solve eigenvalue problems have come a long way. This prompted us to develop\na brand new, modern MHD spectral code which we named Legolas , short for \\Large Eigensystem Generator for\nOne-dimensional pLASmas\". The Legolas code is able to handle both Cartesian and cylindrical geometries, and\nintroduces many new features, e.g. selecting between modern cooling curves that treat optically thin radiative cooling\ne\u000bects. Furthermore, every aspect of the code is modularised, making it ready to be extended with additional physics\nor modern algorithmic requirements (such as mesh re\fnement). The main goal of this paper is to present the new\ncode in terms of its implementation details and to validate it against a plethora of test cases that ensure a correct\ntreatment of the governing equations.\nThese tests include eigenmode quanti\fcations of ideal, static MHD con\fgurations under adiabatic conditions, where\nthe static (that is, no equilibrium \row) and adiabatic linear MHD equations make the problem self-adjoint. When\nperforming a standard Fourier analysis in the ignorable directions, the resulting eigenvalue problem is then Hermitian,\nmeaning that all eigenfrequencies will be either fully real (stable waves) or fully complex (pure damped or unstable\nmodes), hence they are found on the real or imaginary axis of the complex eigenfrequency plane, and the full MHD\nspectrum will be both left-right and up-down symmetric. However, in nature, physical conditions may be far from\nideal. The inclusion of non-ideal e\u000bects like resistivity or thermal conduction lifts the self-adjointness of the eigenvalue3\nproblem, allowing the eigenmodes to move away from the axes into the complex plane, and the up-down symmetry\ngets broken. As long as the equilibrium con\fguration is static, all (adiabatic or non-adiabatic) modes will still\nhave a complementary mode that lies mirrored around the imaginary axis, making the entire spectrum left-right\nsymmetric. This is related to the forward and backward propagating mode symmetry, or the equivalent statement\non the parity-time (PT) symmetry. However, for typical astrophysical plasmas, the conditions are far from static:\ntokamak plasmas, astrophysical jets, solar coronal loops, accretion discs. . . all have equilibrium \rows. The inclusion\nof a background \row breaks the left-right symmetry of the MHD spectrum, resulting in an even more complicated\nstructure. However, the study of the ideal, linear MHD spectrum of \rowing plasmas is still governed by a pair of\nself-adjoint operators (Goedbloed 2011; Goedbloed et al. 2019), and it leaves the up-down symmetry of the spectrum\nintact (where every overstable mode has an equivalent damped counterpart at the same frequency). The combination\nof \row and non-adiabatic e\u000bects, where both left-right and up-down eigenfrequency symmetries are broken, has never\nbeen explored in earnest. All of the above makes it clear that a numerical approach becomes essential, especially\nwhen the equilibrium is no longer homogeneous. Since in reality, virtually no astrophysical con\fguration is spatially\nhomogeneous, we need a \rexible numerical tool to explore the spectrum systematically.\nLegolas solves the linearised MHD equations including various non-adiabatic e\u000bects, resistivity and gravity, assuming\na one-dimensional (1D) equilibrium pro\fle with the possibility of background \row. A standard Fourier analysis for\nthe perturbations is combined with a Finite Element representation of the eigenfunctions in the important coordinate.\nWe transform the original system into an eigensystem for the complex eigenfrequencies !. We use a general formalism\nto include two kinds of geometries, a plane Cartesian strati\fed slab or a (possibly also strati\fed) cylinder, through\nthe inclusion of a scale factor originating from the divergence, gradient and curl operators. Legolas can handle the\nhydrodynamic limit (where all equilibrium magnetic \feld components are set to zero), enabling us to investigate\nstability of hydrodynamic static and stationary equilibria. The resulting system of equations is solved in weak form,\ntransforming the original system into a non-Hermitian complex eigenvalue problem, which is solved using the QR\nalgorithm. This results in a calculation of all eigenfrequencies and corresponding eigenfunctions of the system, such\nthat a detailed analysis of mode stability, but also the entire overview on all supported linear wave modes, becomes\npossible.\nSection 2 introduces the system of equations, along with the linearisation procedure and Fourier mode representation.\nThe treatment of the \fnal system using the \fnite element method is given in Appendix A where we explain the basic\nmathematical formalism behind the FEM, with a complete treatment of the \fnite element matrix assembly process\nand the boundary conditions. Legolas is tested against a large amount of spectra found in the literature in Section\n3, which is subdivided into multiple categories. First we treat ideal MHD with only a gravitational term included,\nwhich has as main advantage that solutions can be obtained analytically. The \frst case handles gravito-MHD waves\nin a Cartesian slab, after which we quantify quasi-Parker instabilities in strati\fed atmospheres. Results obtained with\nLegolas are compared with results found in various modern textbooks, including results for cylindrical geometries,\ndiscussing modes for ideal \rux tubes and tokamak current equilibria where we show liability to interchanges. For\nthe inclusion of \row into the equations, we consider a non-trivial case related to astrophysical jet stability, looking\nat Kelvin-Helmholtz and current-driven instabilities. We demonstrate that we can compute Suydam cluster modes,\noriginating from a surface where the wave vector is locally perpendicular to the magnetic \feld. We then transit to the\ninclusion of non-ideal e\u000bects such as resistivity, looking \frst at the resistive spectrum for a homogeneous plasma. We\nthen extend to recover the resistive quasi-mode in a non-homogeneous case. This is further complemented by adding\nan inhomogeneous medium and background \row, in such a way as to give rise to resistive tearing modes, all of which\nin a Cartesian geometry. Additionally, by allowing for resistivity and current variation, we show that we can also\ncompute resistive rippling modes. Lastly, we treat non-adiabatic e\u000bects, including optically thin radiative losses and\nthermal conduction into the equations, revisiting some pioneering results of discrete Alfv\u0013 en waves and magnetothermal\ninstabilities. Along the way, we \fnd interesting extensions to the original published works, due to the much higher\nresolutions employed here. Since Legolas is the \frst, modern linear MHD code to investigate realistic astrophysical\nplasmas, this opens the door to further in-depth studies of non-ideal equilibrium con\fgurations at high resolutions,\nranging from loops, jets or accretion disks.4\n2.PROBLEM DESCRIPTION AND MODEL EQUATIONS\nThe MHD equations with the inclusion of non-adiabatic e\u000bects, resistivity and gravity can be written in a (nor-\nmalised) Eulerian representation as\n@\u001a\n@t=\u0000r\u0001 (\u001av); (1)\n\u001a@v\n@t=\u0000rp\u0000\u001av\u0001rv+ (r\u0002B)\u0002B+\u001ag; (2)\n\u001a@T\n@t=\u0000\u001av\u0001rT\u0000(\r\u00001)pr\u0001v\u0000(\r\u00001)\u001aL+ (\r\u00001)r\u0001(\u0014\u0001rT) + (\r\u00001)\u0011(r\u0002B)2; (3)\n@B\n@t=r\u0002(v\u0002B)\u0000r\u0002 (\u0011r\u0002B); (4)\nwhere\u001ais the plasma density, vis the velocity, Tis the temperature, pis the pressure, Bdenotes the magnetic \feld\n(satisfyingr\u0001B= 0),\u0011is the resistivity and gthe gravitational acceleration. To close the system the (normalised)\nideal gas law p=\u001aTis used, while \rdenotes the ratio of speci\fc heats, taken to be 5 =3. The symbol Lin Eq. (3)\nrepresents the heat-loss function, de\fned as energy losses minus energy gains due to optically thin radiative cooling\ne\u000bects (Parker 1953) and is given by\nL=\u001a\u0003(T)\u0000H; (5)\nwhereHrepresents the total energy gains. In general, Hcan be anything (for example heating through dissipative\nAlfv\u0013 en waves (van der Holst et al. 2014)), but since there is still no well-de\fned parametrisation for coronal heating\nto date this term is assumed to be as convenient as possible, that is, constant in time but possibly varying in space\nto ensure thermodynamic balance. Note that Hshould always be consistent with a given equilibrium pro\fle as to\nexactly balance out the radiative losses (and possibly the thermal conduction e\u000bects and/or ohmic heating e\u000bects),\nreaching a thermal equilibrium state. This indirectly implies that the heating term is not necessarily independent of\nlocation, but dependent on the connection between the radiative losses and equilibrium temperature pro\fle, which,\nin general, are both spatially dependent. The \frst term in (5) denotes the radiative losses, dependent on the cooling\ncurve \u0003(T). These curves are tabulated sets resulting from detailed calculations, and can hence be interpolated to high\ntemperature resolutions. Legolas has multiple cooling curves implemented, most notably those by Colgan et al. (2008)\nand Schure et al. (2009), where the latter one is extended to the low-temperature limit using Dalgarno & McCray\n(1972). In addition we also implemented a piecewise power law as described by Rosner et al. (1978), which is an explicit\n(piecewise) function over the entire temperature domain. In the solar and astrophysical literature, these cooling curves\ncollect detailed knowledge on radiative processes, which are all assumed to be in the optically thin regime. It is\nworth noting that the inclusion of time-dependent background heating or \rows can in\ruence the spectrum in itself,\nas shown in for example Barbulescu et al. (2019); Hillier et al. (2019), but due to the time-dependence of those e\u000bects\nthe assumption of a stationary background state is no longer applicable. A possibility however may be varying the\nbackground heating or \row in subsequent runs, provided the background state is known at every snapshot.\nThermal conduction in magnetised plasmas is highly anisotropic, as the e\u000bect is a few orders of magnitude stronger\nalong the \feld lines than across. We hence use a tensor representation to model this anisotropy, denoting the thermal\nconductivity tensor \u0014by\n\u0014=\u0014keBeB+\u0014?(I\u0000eBeB); (6)\nwhereIdenotes the unit tensor and eB=B=Bis a unit vector along the magnetic \feld. The coe\u000ecients \u0014kand\u0014?\ndenote the conductivity coe\u000ecients parallel and perpendicular to the local direction of the magnetic \feld. For typical\nastrophysical applications the Spitzer conductivity is used, given by\n\u0014k\u00198\u000210\u00007T5=2erg cm\u00001s\u00001K\u00001; \u0014?\u00194\u000210\u000010n2B\u00002T\u00003\u0014k;\n\u0014k\u00198\u000210\u000012T5=2W m\u00001K\u00001; \u0014 ?\u00194\u000210\u000030n2B\u00002T\u00003\u0014k;(7)\nwherendenotes the number density, given by \u001a=nmpwithmpthe proton mass. The \frst and second row in Eq.\n(7) give the thermal conduction coe\u000ecients in cgs and mks units, respectively. For the solar corona we typically \fnd\nthat\u0014kis about 12 orders of magnitude larger than \u0014?(Priest 2014), so perpendicular thermal conduction is usually\nignored. Nevertheless, in Legolas both parallel and perpendicular thermal conduction are implemented. For the5\nFigure 1. Unit vectors and examples of B0andv0for the Cartesian (left) and cylindrical case (right).\nresistivity\u0011one can in principle take any pro\fle. We implemented the Spitzer resistivity,\n\u0011=4p\n2\u0019\n3Zione2pmeln(\u0015)\n(4\u0019\u000f0)2(kBT)3=2; (8)\nwhereZiondenotes the ionisation taken to be unity, eandmedenote the electron charge and mass, respectively,\nand\u000f0andkBare the electrical permittivity and Boltzmann constant. The Coulomb logarithm is given by ln( \u0015)\nand is approximately equal to 22 for solar coronal conditions (Goedbloed et al. 2019). It is important to emphasise\nthat, since we will further linearise the governing non-linear equations, we can adopt fully realistic values for all the\nnon-ideal coe\u000ecients, such as the resistivity or thermal conduction coe\u000ecients. This is in contrast to fully nonlinear\ncomputations, which are severely restrained in reaching magnetic Reynolds numbers beyond 104\u0000105.\n2.1. Equilibrium conditions\nWe consider a general coordinate system denoted by ( u1;u2;u3), corresponding to three orthogonal basis vectors.\nThe main advantage of this approach is that it allows us to include two di\u000berent geometries with only one basic\nformalism (and implementation). First we consider a standard plane slab geometry in Cartesian coordinates, that is,\na plasma which is con\fned in height, and considered to be bounded by two horizontal, perfectly conducting walls at a\n\fxed distance apart, extending outwards to in\fnity in the other two ignorable coordinates. This case also approximates\nthe limit of a fully in\fnite free space when the walls are moved o\u000b to in\fnity. In Cartesian geometry the coordinate\nsystem can be written as ( x;y;z ) and the vectors fu1;u2;u3gare the standard Cartesian triad f^ex;^ey;^ezgalong the\naxes. This makes it quite convenient to include for example gravitational e\u000bects which will induce an equilibrium\nstrati\fcation in the u1coordinate. The second geometry is that of an in\fnitely long plasma cylinder encased by\na solid wall at a certain distance away from the cylinder axis, for which the coordinate system can be de\fned as\n(r;\u0012;z ). At each point the vectors fu1;u2;u3gare de\fned as the triad of tangent vectors, f^er;^e\u0012;^ezg, with ^eralong\nthe radial direction, ^ezin the direction of the cylinder axis and ^e\u0012=^ez\u0002^ertangent to the cylinder. A detailed view\nof both geometries and their corresponding coordinate systems is shown in Figure 1. The basic operators present in\nEquations (1)-(4), that is, the divergence, gradient and curl, introduce a scale factor \"=rfor cylindrical geometries,\nwhich is reduced to \"= 1 for a Cartesian coordinate system. Hence exploiting this scale factor in the mathematical\nformalism allows for one implementation, where one can conveniently switch between both cases. We note that the\ncylindrical setup is also applicable to the so-called cylindrical accretion disk limit, as for example exploited to study\nMHD instabilities in disks by Blokland et al. (2007a).\nLinearisation involves splitting variables into two parts: a time-independent part, usually denoted with subscript 0,\nand a perturbed part, denoted by a subscript 1. Legolas handles one-dimensional equilibria which depend only on\nu1, or, more speci\fcally, time-independent equilibria of the form\n\u001a0=\u001a0(u1);\np0=p0(u1);\nT0=T0(u1);v0=v02(u1)u2+v03(u1)u3;\nB0=B02(u1)u2+B03(u1)u3: (9)6\nIn general we have g=\u0000g(u1)u1, where the Cartesian case is for a strati\fed atmosphere or layer, and the cylindrical\ncase can also allow for gravitational strati\fcation of an accretion disk situated for u1=r2[1;R]. In the case of\na cylinder where u1=r2[0;R], this gravitational term is absent. Using these equations in combination with Eqs.\n(1)-(4) yields two conditions for the time-independent parts, given by\n\u0012\np0+1\n2B2\n0\u00130\n+\u001a0g\u0000\"0\n\"\u0000\n\u001a0v2\n02\u0000B2\n02\u0001\n= 0; (10)\n1\n\"(\"\u0014?T0\n0)0\u0000\u001a0L0= 0; (11)\nwhere the \frst condition originates from the momentum equation (2) and should always be satis\fed as it expresses a\nforce-balanced state. The second condition originates from the non-adiabatic terms in the energy equation and should\nbe accounted for if these terms are included, the prime denotes the derivative with respect to u1. It should be noted\nthat resistive terms are not considered here, which is justi\fed by considering that the time scales on which the magnetic\n\felds decay due to resistivity is much, much larger than the time scales of resistive modes. This is a consequence of\nlarge magnetic Reynolds numbers Rmin typical astrophysical cases, yielding magnetic decay time scales of \u001c\u0018Rm\u001cA\n(with\u001cAthe typical Alfv\u0013 en time in ideal MHD) compared to much faster resistive mode time scales of \u001c\u0018R\u000b\nm(where\ntypically 0<\u000b< 1). We can hence consider the equilibrium itself to be independent of resistivity, which removes some\nstringent extra conditions on the energy and induction equations. Also note that the third term in Eq. (10) is only\nincluded for a cylinder, since \"0= 0 in a Cartesian geometry. This translates to the well-known fact that the centrifugal\nand tensional part of the Lorentz force are absent for a Cartesian slab. Furthermore a cylindrical equilibrium pro\fle\nshould satisfy on-axis regularity conditions, meaning that v02;v0\n03;B02andB0\n03all have to be equal to zero at r= 0.\nWhen considering an accretion disk in the cylindrical limit, the inner edge of the disk is at r= 1, so lengths are then\nexpressed in this inner disk radius and no regularity conditions apply then.\n2.2. Linearised equations\nNow we linearise Eqs. (1)-(4) around the equilibrium speci\fed in (9), where the unperturbed time-independent parts\nare denoted with a subscript 0 and the perturbed time-dependent parts are denoted with a subscript 1. It follows from\nthe adopted equilibrium con\fguration that r\u0001v0= 0 andr\u0001B0= 0, such that the divergence-free condition on the\nmagnetic \feld is ful\flled and that the equilibrium \row \feld is incompressible. However, the perturbed quantities can\nrepresent both incompressible or compressible eigenoscillations. The no-monopole condition should also be taken into\naccount for the perturbed magnetic \feld. Therefore, we adopt a vector potential to write B1=r\u0002A1such that\nr\u0001B1= 0 is automatically satis\fed. The system of linearised equations is thus given by\n@\u001a1\n@t=\u0000\u001a0r\u0001v1\u0000v0\u0001r\u001a1; (12)\n\u001a0@v1\n@t=\u0000\u001a0v0\u0001rv1\u0000\u001a0v1\u0001rv0\u0000\u001a1v0\u0001rv0\u0000r(\u001a1T0+\u001a0T1)\n+ (r\u0002B0)\u0002(r\u0002A1) + [r\u0002(r\u0002A1)]\u0002B0+\u001a1g;(13)\n\u001a0@T1\n@t=\u0000\u001a0v1\u0001rT0\u0000\u001a0v0\u0001rT1\u0000(\r\u00001)\u001a0T0r\u0001v1\u0000(\r\u00001)\u001a1L0\u0000(\r\u00001)\u001a0(LTT1+L\u001a\u001a1)\n+ (\r\u00001)r\u0001(\u00140\u0001rT1) + (\r\u00001)r\u0001(\u00141\u0001rT0)\n+ 2(\r\u00001)\u0011(r\u0002B0)\u0001[r\u0002(r\u0002A1)] + (\r\u00001)d\u0011\ndTT1(r\u0002B0)2;(14)\n@A1\n@t=v1\u0002B0+v0\u0002(r\u0002A1)\u0000\u0011r\u0002(r\u0002A1)\u0000d\u0011\ndTT1(r\u0002B0); (15)\nwherep1is replaced by \u001a1T0+\u001a0T1resulting from the linearised ideal gas law. The perturbation \u00141of the thermal\nconduction tensor is obtained by linearising the expressions (6)-(7), while the derivatives of the heat-loss function with\nrespect to density and temperature are given by\nL\u001a=@L\n@\u001a\f\f\f\f\nT= \u0003(T0); LT=@L\n@T\f\f\f\f\n\u001a=\u001a0d\u0003(T0)\ndT; (16)\nwhich should be evaluated using the equilibrium quantities. The terms containing @\u0011=@T follow from a linearisation of\nthe resistivity parameter, which can be written in terms of the variable T1using the temperature dependence originating7\nfrom the assumed Spitzer resistivity in Eq. (8), that is, \u00111=T1@\u0011=@T . In addition, Legolas allows for an anomalous\nresistivity prescription in which we typically have \u0011(u1;j(u1;t)), that is, a resistivity pro\fle that is spatio-temporal\nin general, but for a \fxed time, depends on position and current pro\fle. This in turn implies that a total derivative\nshould be used for the resistivity, given by \u00110=@\u0011=@u 1+T0\n0@\u0011=@T . In most use cases a resistivity pro\fle \u0011(T(u1))\nis su\u000ecient, which is only temperature dependent (and hence indirectly spatially varying as well for inhomogeneous\ntemperature pro\fles). Also note that these linear equations (as also the nonlinear set above) assumed an external\ngravitational \feld, so we did not need to linearise the gravity term g(this is the so-called Cowling approximation). In\nthe future, we can extend the set of equations with the Poisson equation and also allow for self-gravity driven Jeans\ninstabilities.\nNext, we perform a Fourier analysis with an exponential time dependence, imposing standard Fourier modes on the\nu2andu3coordinates of a perturbed quantity f1, given by\nf1=bf1(u1) exp [i(k2u2+k3u3\u0000!t)]: (17)\nIn this form the wave numbers k2andk3correspond to kyandkzin Cartesian geometry, and to mandkin a cylindrical\ngeometry, respectively. Note that mis quanti\fed to integer values, since the \u0012-direction is periodic. Additionally, we\napply the following transformation to the perturbed quantities\n\"b\u001a1=e\u001a1; i\"bv1=ev1;bv2=ev2; \"bv3=ev3;\n\"bT1=eT1; ibA1=ea1; \"bA2=ea2;bA3=ea3;(18)\nThis particular transformation simpli\fes the resulting set of equations and has as additional e\u000bect that all terms are\nreal except for the non-adiabatic and resistive contributions, such that we are only dealing with imaginary terms\nwhen these physical e\u000bects are included. This is in analogy to the fact that the purely adiabatic case is governed by\nself-adjoint operators: one for the case without \row and two for the case with \row included (Goedbloed 2018a,b). The\n\fnal set of Fourier-analysed linearised equations is given below, where the tilde notation in Eqs. (18) is dropped for\nthe sake of simplicity. From now on, tildes will no longer be written explicitly since there is no confusion possible.\n!1\n\"\u001a1=\u00001\n\"\u001a0\n0v1\u00001\n\"\u001a0(v0\n1\u0000k2v2\u0000k3v3) +1\n\"\u00121\n\"k2v02+k3v03\u0013\n\u001a1; (19)\n!\u001a01\n\"v1=\u0012\u001a1T0+\u001a0T1\n\"\u00130\n+1\n\"B0\n03(a0\n2\u0000k2a1)\u0000(\"B02)0\n\"(a0\n3\u0000k3a1)\n\u0000B03(\nk2\n3\n\"a2\u0000k2k3\n\"a3\u0000\u00141\n\"(a0\n2\u0000k2a1)\u00150)\n+1\n\"\u0012k2v02\n\"+k3v03\u0013\n\u001a0v1\n+B02\u001ak2\n2\n\"2a3\u0000k2k3\n\"2a2\u00001\n\"h\n\"(a0\n3\u0000k3a1)i0\u001b\n\u00002\"0\n\"\u001a0v02v2\u0000\"0\n\"2v2\n02\u001a1+1\n\"g\u001a1;(20)\n!\u001a0v2=1\n\"2(\u001a1T0+\u001a0T1)k2+\u001a0\u0012k2v02\n\"+k3v03\u0013\nv2+1\n\"2(\"B02)0(k3a2\u0000k2a3)\n+B03\u0014\n\u0000\u0012\nk2\n3+k2\n2\n\"2\u0013\na1+k2\n\"2a0\n2+k3a0\n3\u0015\n\u00001\n\"2\u001a0(\"v02)0v1;(21)\n!\u001a01\n\"v3=1\n\"(\u001a1T0+\u001a0T1)k3+1\n\"\u001a0\u0012k2v02\n\"+k3v03\u0013\nv3+1\n\"B0\n03(k3a2\u0000k2a3)\n\u0000B02\u0014\n\u0000\u0012\nk2\n3+k2\n2\n\"2\u0013\na1+k2\n\"2a0\n2+k3a0\n3\u0015\n\u00001\n\"\u001a0v0\n03v1;(22)8\n!\u001a01\n\"T1=\u00001\n\"\u001a0T0\n0v1+1\n\"\u001a0\u0012k2v02\n\"+k3v03\u0013\nT1\u0000(\r\u00001)1\n\"\u001a0T0(v0\n1\u0000k2v2\u0000k3v3)\n\u0000i(\r\u00001)\u0000\n\u0014k;0\u0000\u0014?;0\u0001\n\"1\nB2\u0012k2B02\n\"+k3B03\u00132\nT1+i(\r\u00001)1\n\"\"\n\"\u0014?;0\u0012T1\n\"\u00130#0\n\u0000i(\r\u00001)\u0014?;0\n\"\u0012k2\n2\n\"2+k2\n3\u0013\nT1+i(\r\u00001)1\n\"(\"\u0014?;1T0\n0)0\n+i(\r\u00001)\u0000\n\u0014k;0\u0000\u0014?;0\u0001\n\"1\nB2T0\n0\u0014\u0012k2k3\n\"B02+k2\n3B03\u0013\na2\u0000\u0012k2\n2\n\"B02+k2k3B03\u0013\na3\u0015\n\u0000i(\r\u00001)L01\n\"\u001a1\u0000i(\r\u00001)1\n\"\u001a0(LTT1+L\u001a\u001a1)\n+ 2i(\r\u00001)\u0011(\nB0\n03\u0014\u00101\n\"(a0\n2\u0000k2a1)\u00110\n+k2k3\n\"a3\u0000k2\n3\n\"a2\u0015\n\u0000(\"B02)0\n\"2\u0014\u0010\n\"(a0\n3\u0000k3a1)\u00110\n+k2k3\n\"a2\u0000k2\n2\n\"a3\u0015)\n+i(\r\u00001)d\u0011\ndT1\n\"\"\n\u0000\nB0\n02\u00012+\u0000\nB0\n03\u00012+ 2\"0\n\"B02B0\n02+\u0012\"0\n\"B02\u00132#\nT1;(23)\n!a1=\u0000B03v2+1\n\"B02v3\u00001\n\"v02a0\n2\u0000v03a0\n3+\u0012k2v02\n\"+k3v03\u0013\na1+i\u0011\u0014\n\u0000\u0012k2\n2\n\"2+k2\n3\u0013\na1+k2\n\"2a0\n2+k3a0\n3\u0015\n; (24)\n!1\n\"a2=\u00001\n\"B03v1+k3v03\n\"a2\u0000k2v03\n\"a3+i\u0011\"\u00121\n\"(a0\n2\u0000k2a1)\u00130\n+k2k3\n\"a3\u0000k2\n3\n\"a2#\n+id\u0011\ndT1\n\"B0\n03T1; (25)\n!a3=1\n\"B02v1\u0000k3v02\n\"a2+k2v02\n\"a3+i1\n\"\u0011\u0014\u0010\n\"(a0\n3\u0000k3a1)\u00110\n+k2k3\n\"a2\u0000k2\n2\n\"a3\u0015\n\u0000id\u0011\ndT1\n\"2(\"B02)0T1: (26)\nThe perturbed thermal conductivity tensor \u0014?;1in Eq. (23) written in terms of the perturbed variables is given by\n\"\u0014?;1=@\u0014?;0\n@TT1+@\u0014?;0\n@\u001a\u001a1\u00002\"B02\u0000\na0\n3\u0000k3a1\u0001@\u0014?;0\n@(B2)+ 2B03\u0000\na0\n2\u0000k2a1\u0001@\u0014?;0\n@(B2): (27)\nAn interesting side note is that \u0014k;1does not appear in the equations, which is due to the fact that this term is\naccompanied by a B0\u0001rT0contribution, and this is zero due to the equilibrium pro\fle in Eq. (9) (Van der Linden &\nGoossens 1991a,b). We now have a system of eight ordinary di\u000berential equations in u1for the perturbed quantities\n\u001a1;v1;v2;v3;T1;a1;a2anda3.\n2.3. Boundary conditions\nThe above system of di\u000berential equations (19)-(26) has to be complemented by a set of boundary conditions on\nboth sides of the domain. For a Cartesian geometry we look at a domain enclosed by two conducting walls. Clearly,\nthe velocity component perpendicular to the walls has to be zero since there can not be any propagation into a solid\nboundary. Mathematically, this translates into v\u0001^en= 0, where ^enrepresents the normal vector to the wall. Following\nthe same reasoning we also require that B\u0001^en= 0, so in terms of a vector potential this implies ^en\u0002A= 0. Hence,\napplying this to the set of linearised equations this means that for the Cartesian case v1,a2anda3all have to be\nzero on the boundaries. Furthermore, since we are dealing with a perfectly conducting wall one has to take care when\nthermal conduction is included. In that case the rigid wall directly in\ruences the temperature, since it acts as an\nenergy reservoir essentially eliminating the temperature perturbation. Hence, if and only if perpendicular thermal\nconduction is taken into account we have to supplement the boundary conditions by the additional condition T1= 0\nat the boundary. In theory there is a second possibility, which is treating the wall as a perfect insulator instead of a\nperfect conductor. In that case there is no heat \rux, which translates to the boundary condition T0\n1= 0 instead of\nT1= 0. For now we only consider the latter condition, that is, the one corresponding to a perfectly conducting wall.9\nIn a cylindrical geometry we have the exact same boundary conditions as for the Cartesian case at the outer wall\nr=R, or at the outer edge of the accretion disk at R. The same is true at the inner disk edge, but for a \rux tube\nextending to r= 0 we have to take the regularity conditions into account when treating the cylinder axis r= 0, which\ncomes down to the fact that rvrshould go to zero when approaching r= 0. Looking back at the transformations (18)\nwe applied, it follows that this condition is equivalent to v1= 0. Analogously, the same holds true for a2anda3such\nthat these conditions are identical to the ones we applied for the Cartesian case, which is convenient implementation-\nwise. We again have to consider an additional condition if perpendicular thermal conduction is taken into account,\nsince thenrbT1= 0 should also hold on the cylinder axis, which, similarly as for v1, translates into T1= 0 atr= 0.\nIn the case of con\fnement by a perfectly conducting wall we thus have straightforward boundary conditions, that\nis,v1= 0;a2= 0;a3= 0 andT1= 0 for both the Cartesian and cylindrical geometries on both sides. This latter\nboundary condition should only be taken into account if and only if perpendicular thermal conduction is included.\n2.4. Solving the equations\nThe system of equations (19)-(26) is solved through usage of a Finite Element discretisation. Applying a weak\nGalerkin formalism turns this system of equations into a generalised matrix-eigenvalue problem. A detailed explanation\non how this is done can be found in Appendix A, where we describe the structure of the \fnite element approach and\nthe matrix assembly process, along with a detailed treatment of how the boundary conditions are handled.\n3.RESULTS\nAs is common practice when developing a new numerical code we tested Legolas against various results previously\nobtained in the literature. We divided this section into four subsections, each of which handles di\u000berent physical\ne\u000bects. To begin with, we discuss results for adiabatic equilibria where only gravity is included in Subsection 3.1. In\nthis case we can compare numerical spectra obtained through Legolas with analytical solutions acquired by solving\ndispersion relations, here we focus on strati\fed atmospheres containing p- andg-modes. We then move on to cylindrical\ngeometries in 3.2 where we \frst look at adiabatic \rux tubes, followed by the inclusion of \row e\u000bects by considering\nequilibria with Kelvin-Helmholtz instabilities and Suydam cluster modes. Next the focus shifts to non-adiabatic e\u000bects\nin 3.3 by looking at a resistive MHD computation for a case without gravity, where a quasi-mode is known analytically.\nResistive tearing modes are also discussed, combining the e\u000bects of \row and resistivity. The \fnal subsection 3.4 treats\nthe inclusion of thermal conduction and optically thin radiative cooling e\u000bects, where we look at non-adiabatic discrete\nAlfv\u0013 en waves and magnetothermal modes.\n3.1. Cartesian cases: waves in strati\fed atmospheres\nFirst of all we discuss multiple theoretical results for adiabatic equilibria in a Cartesian geometry, where only gravity\nis included. We consider p- andg-modes in strati\fed layers, and pay special attention to speci\fc unstable branches.\n3.1.1. Gravito-MHD waves\nThe \frst test case covers gravito-MHD waves as discussed in Goedbloed et al. (2019, \fg. 7.9), which handles an\nexponentially strati\fed atmosphere with constant sound and Alfv\u0013 en speeds. This magnetised atmosphere contains\nthe generalisation of the p- andg-modes of an unmagnetised layer, and the constancy of the sound and Alfv\u0013 en speed\nrenders it analytically tractable, since the slow and Alfv\u0013 en continua collapse to points. The geometry is Cartesian,\nwithx2[0;1] and an equilibrium con\fguration given by\n\u001a0=\u001ace\u0000\u000bx; p 0=pce\u0000\u000bx;B0=Bce\u00001\n2\u000bx^ez; \u000b =\u001acg\npc+1\n2Bc; (28)\nwherepcandBcare taken to be 0.5 and 1, respectively, as to yield a plasma beta equal to unity. The parameter\n\u000bis taken to be 20, which, together with g= 20, is used to constrain the value for the constant \u001ac. These four\nequations completely determine the equilibrium con\fguration, since the temperature is T0=p0=\u001a0, following the\nideal gas law. The spectrum discussed in Goedbloed et al. (2019) is actually the solution to the analytic dispersion\nrelation for gravito-MHD waves, which shows the squared eigenvalue as a function of wave number for a \fxed angle\n\u0012=\u0019=4 between the wave vector k0and the magnetic \feld B0. However, the spectrum as calculated by Legolas\ncorresponds to one single equilibrium con\fguration, meaning one value for kyandkz. In order to reproduce \fgure\n7.9 from Goedbloed et al. (2019) and compare the results, we performed 100 di\u000berent runs where the equilibrium10\nFigure 2. Spectrum of gravito-MHD modes, obtained through 100 Legolas runs of 351 gridpoints each. The fast (top), Alfv\u0013 en\n(middle) and slow (bottom) branches of the MHD spectrum are clearly visible. The inset shows unstable slow modes at low\nfrequencies.\nparameters in Eq. (28) remained unchanged, but kyandkztook on 100 di\u000berent values between 0 andp\n250 as to\nyield a wave number range for k2\n0between 0 and 500. Since the magnetic \feld is purely aligned with the z-axis we can\nwritekk=kz=k0cos(\u0012) andk?=ky=k0sin(\u0012). All runs were performed using 351 gridpoints, yielding a matrix\nsize of 5616\u00025616.\nOur results are shown in Figure 2, where every vertical collection of points at the same k2\n0value represents one single\nLegolas run. Since we are in an MHD regime with \f= 1, the three MHD subspectra can be clearly distinguished,\nshowing the fast p-modes (top-left branches), Alfv\u0013 en g-modes (middle branches) and slow g-modes (bottom branches).\nThe inset shows a zoom-in near the marginal frequency of the spectrum, showing unstable ( !2<0) slow MHD modes.\nThese long-wavelength unstable modes are related to the Parker instabilities, due to magnetic buoyancy, as we will\nshow in Section 3.1.2. Note that since this case is adiabatic and fully self-adjoint, every individual MHD spectrum\nis left-right and up-down symmetric in the complex eigenfrequency plane, but this aspect is hidden from the !2\u0000k2\n0\nview shown here.\n3.1.2. Quasi-Parker instabilities\nNext we discuss a modi\fed case of the gravito-MHD waves, namely a spectrum showing quasi-Parker instabilities as\ndone in Goedbloed et al. (2019, \fg. 12.2). The di\u000berence with the previous case is that a fully analytic description is\nno longer possible, since the introduction of magnetic shear leads to continuous ranges in the MHD spectrum. Instead\nof showing the spectrum for one single value for \u0012, we now vary the direction of the wave vector k0between 0 and \u0019.\nThe equilibrium con\fguration is similar to the one in Section 3.1.1, given in Cartesian geometry by\n\u001a0=\u001ace\u0000\u000bx; p 0=pce\u0000\u000bx; \u000b =\u001acg\npc+1\n2Bc;\nB02=Bce\u00001\n2\u000bxsin(\u0015x); B 03=Bce\u00001\n2\u000bxcos(\u0015x);(29)\nwhere magnetic shear was introduced through the parameter \u0015. The quantities \u000bandBcare assigned the same values\nas in Eq. (28), except that g= 0:5 andpc= 0:25 which yields a plasma beta \f= 0:5. The wave vectors are given by\nky=\u0019sin(\u0012) andkz=\u0019cos(\u0012), such that k2\n0\u001910. The angle \u0012was varied between 0 and \u0019for a total of 100 runs at\n351 gridpoints each, shown in Figure 3.11\nFigure 3. Spectrum showing Parker and quasi-Parker modes without (left) and with (right) magnetic shear. The slow and\nAlfv\u0013 en continua are shown in red and cyan, respectively, where the insets zoom into the region of quasi-interchange modes.\nThe bottom row of panels show the eigenfrequency view for the single case \u0012= 0:3\u0019. The continua are again annotated on the\n\fgures, visualising the collapsed single point values (left) as well as the genuine continuum ranges (right). The grey dashed line\nin the top two panels denotes != 0.\nThe left panels handle the case without magnetic shear, that is, \u0015= 0, which basically reduces to the one from\nthe previous subsection. In this case the slow and Alfv\u0013 en continua collapse into single point values, denoted in red\nand cyan, respectively. The right panels show the same con\fguration where \u0015= 0:3 was taken, introducing magnetic\nshear, which introduces genuine continua seen as bands. These continua a\u000bect the overall stability, and organise the\nentire MHD spectrum: all discrete modes are fully aware of the essential spectrum formed by these (slow and Alfv\u0013 en)\ncontinua and the (fast) accumulation points at in\fnite frequency. All features of the original \fgure in Goedbloed et al.\n(2019) are reproduced. The inset zooms into the region where both continua overlap, showing quasi-interchange and\ninterchange instabilities. Once more, each run shown here collectively in Figure 3 actually has a spectrum that is\nleft-right and up-down symmetric in the eigenfrequency plane. This is depicted on the bottom two panels, which show\nthe eigenfrequency view for one single case ( \u0012= 0:3\u0019). The continuum ranges separate nicely: the collapsed single\npoint values are denoted by cyan (Alfv\u0013 en) and red (slow) points on the left panel, the genuine continua are shown with\ncyan and red bands on the right panel. The instabilities themselves are situated on the (positive) imaginary axis, due\nto the self-adjointness of the eigenvalue problem mentioned earlier.\nAs explained in Goedbloed et al. (2019), we see from this eigenmode computation that the Parker instability, which is\nthere for k0parallel toB0, becomes a quasi-Parker instability away from perfect alignment, and connects smoothly to\nwell-known quasi-interchange instabilities that occur here (marginally) away from perpendicular orientation. Quantify-\ning how the equilibrium parameters in\ruence the growth rates of these unstable branches can only be done numerically,\ne.g. with Legolas .12\n3.2. Adiabatic, cylindrical cases\nNext we move on to cylindrical con\fgurations, which provide tests for the scale factor \"in the equations. Analytical\nresults from the literature are again well reproduced. Furthermore we look at di\u000berent spectra previously obtained by\nthe LEDA code, discussed in various papers, and compare those with the new spectra from Legolas .\n3.2.1. Magnetic \rux tubes\nThe \frst case that we describe in this subsection is a magnetic \rux tube embedded in a uniform magnetic envi-\nronment, discussed in Roberts (2019). The equilibrium con\fguration is simple, in the sense that we have a uniform\nmagnetic \feld aligned with the z-axis both inside and outside of the \rux tube, with a similar structure for the other\nequilibrium parameters:\nB0(r);\u001a0(r);p0(r);T0(r) =8\n<\n:B0;\u001a0;p0;T0; ra(30)\nwhere the subscripts 0 and erefer to values inside the tube and for the environment, respectively. The outer radius of\nthe tube is denoted by aand hence represents a discontinuous interface between the tube itself and the environment.\nSince total pressure balance should be preserved across the boundary, which is something that follows from Eq. (10),\nthis yields a relation between pressures and magnetic \feld components inside and outside of the tube, which in turn\nimplies a connection between the plasma densities, sound speeds and Alfv\u0013 en speeds across the boundary:\np0+1\n2B2\n0=pe+1\n2B2\ne;\u001ae\n\u001a0=c2\ns+1\n2\rc2\nA\nc2se+1\n2\rc2\nAe; (31)\nwherec2\ns=\rp0=\u001a0andc2\nA=B2\n0=\u001a0denote the sound speed and Alfv\u0013 en speed, respectively, in which the values outside\nof the \rux tube are used if there is a subscript epresent.\nIt should be noted that this extremely simple equilibrium con\fguration is the standard case used in many solar\ncoronal loop seismology e\u000borts. Since it simply has two uniform media (one inside the tube and one in its exterior),\nit has no continuous spectra (they reduce to point values), but the interface makes it possible to again have surface\nmodes that would be a\u000bected by true radial variation. Also note that these \rux tubes have only stable waves, but\nwe can distinguish between body and surface waves, depending on the variation of the eigenfunctions within the \rux\ntube. In the exterior of the \rux tube all eigenfunctions are exponentially varying.\nWe should also clarify here that the original dispersion relation as given in Roberts (2019) assumes a \rux tube\nembedded in an environment extending towards in\fnity, while Legolas on the other hand assumes a \fxed wall\nboundary at the outer edge of the domain. Hence, we assume here that the domain is situated in r2[0;10] with the\ninner \rux tube wall at r= 1, in order to minimise the outer wall in\ruence. However, this introduces an additional\ncomputational challenge, in the sense that we are (mainly) interested in the behaviour of the inner modes, since we\nknow that the outer modes all have exponentially varying eigenfunctions which decay to in\fnity (or towards our far-\naway outer wall). Hence, in order to resolve those inner waves huge resolutions are needed due to the 1 : 10 ratio. In\norder to circumvent this issue we used a simple prescription for mesh re\fnement, that is, a 60 \u000030\u000010 division of\nthe initial nodes. This means that 60% of the gridpoints are used for the inner tube region r2[0;0:95], 30% of the\ngridpoints are located near the transition region r2[0:95;1:05], and the remaining 10% are used for the environment\nr2[1:05;10].\na) Photospheric \rux tube. First we look at a \rux tube under photospheric conditions, that is, an equilibrium for\nwhichcAe0) denoted on the \fgure itself.\nThe real and imaginary parts of the \u001a,rvrandvzeigenfunctions for each of these modes are shown on the subsequent\npanels. Those for the KH mode (panels c\u0000d) are localised around the jet radius r= 1, which is also the point\nwhere the sonic and Alfv\u0013 enic Mach numbers drop to zero (panel b). Panelsethroughjdepict the eigenfunctions of\nthe \frst three CD modes, with the \frst, second and third shown on the \frst, second and third row of the bottom\npanels, respectively. The left column shows the real part, the right column the imaginary part of the eigenfunction.\nThe CD modes have an increasing number of nodes on r2[0;1], which is most clearly visible by looking at the rvr\neigenfunction (green): no nodes for the \frst CD mode (panels e\u0000f), one node for the second CD mode (panels g\u0000h)\nand two nodes for the third CD mode (panels i\u0000j).\nNote that here, we are still adiabatic such that the up-down symmetry (relating to time reversal) is still present\nin the eigenfrequency plane, but the introduction of equilibrium \row caused left-right symmetry breaking between\nforwards and backwards propagating modes. As pointed out in Goedbloed (2018a), the study of MHD spectra of\nstationary (with \row) equilibria is still governed by two self-adjoint operators, but as seen in Figure 7, modes can17\nFigure 7. Panela: MHD spectrum for the equilibrium con\fguration given in (36). The KH and CD instabilities are indicated\non the panel. Panel b: sonic and Alfv\u0013 enic Mach numbers. The other panels show the real and imaginary parts of the \u001a(blue),\nrvr(green) and vz(red) eigenfunctions, for the KH mode ( c\u0000d), \frst CD mode ( e\u0000f), second CD mode ( g\u0000h) and third\nCD mode ( i\u0000j), respectively.\nenter the complex plane at various locations (identi\fed by the spectral web (Goedbloed 2018b)). The correspondence\nwith the original \fgure in Baty & Keppens (2002) is one-to-one for the KH and CD modes, but here we have a lot\nmore detail near the axes due to the higher resolution. We will discuss this resolution aspect in Section 4.\n3.2.4. Suydam cluster modes\nNext we look at Suydam cluster modes in a cylindrical geometry, which arise from the presence of a Suydam surface in\nthe equilibrium con\fguration, that is, a location where k\u0001B0= 0. Shear \row e\u000bects are included, and the equilibrium\nis given by\n\u001a0= 1; v 02= 0; v 03=vz0\u0000\n1\u0000r2\u0001\n;\nB02=J1(\u000br); B 03=p\n1\u0000P1J0(\u000br); p 0=P0+1\n2P1J2\n0(\u000br);(38)\nwhere\u000b= 2,P0= 0:05,P1= 0:1 andvz0= 0:14, the functions J0andJ1denote the Bessel functions of the \frst kind.\nThe wave numbers were chosen to be k2=m= 1 andk3=k=\u00001:2, ensuring a Suydam surface at r\u00190:483.\nThe spectrum is calculated using 501 gridpoints for r2[0;1] and is shown in Figure 8. The resulting locations\nof the various o\u000b-axis outer modes are in agreement with results given in Nijboer et al. (1997). However, since this\nspectrum is calculated using a \fve times higher resolution, we have much more intricate detail near the Suydam surface,\nrevealing even more o\u000b-axis modes (inset on panel a). The top two panels on the right side of Figure 8 show the sonic\nand Alfv\u0013 enic Mach numbers (panel b), together with the k\u0001B0=B02k2=r+B03k3andk\u0001v0=B02v02=r+B03v03\npro\fles as a function of radius (panel c), respectively. The location of the Suydam surface at r\u00190:483 is denoted18\nFigure 8. Panela: Suydam cluster spectrum for the equilibrium given in (38). Inset: zoom-in near the Suydam surface,\nrevealing more detail. Panel bdepicts the sonic and Alfv\u0013 enic Mach numbers as a function of radius, panel cshows the k\u0001B0\nandk\u0001v0curves where the red cross denotes the Suydam surface. The bottom row of panels shows the real part of the \u001aand\nrvreigenfunctions, for the four modes in the Suydam sequence annotated on panel a. The dotted red line denotes the location\nof the Suydam surface.\nwith a red cross. The bottom row of panels ( d\u0000g) show the real part of the \u001aandrvreigenfunctions, for the four\nmodes annotated on panel awith the location of the Suydam surface annotated with a vertical red line. !1and!3\ncorrespond to modes on the left side of the Suydam surface, the other two correspond to modes on the right side. All\neigenfunctions shown here show the speci\fc variation associated with their location relative to the Suydam surface:\n!1and!3have their localised behaviour on the left side of the red dashed line, while it is vice-versa for the other two\nmodes. For visual purposes the horizontal axis in panels eandfis di\u000berent from the one in panels candd, since\nthe former correspond to the next modes in the Suydam sequence, showing stronger radial variation. Note that the\noriginal Suydam criterion (Goedbloed et al. 2019) is related to static equilibria, the generalisation of these Suydam\ncluster criteria is presented in Wang et al. (2004).\n3.3. Resistive, Cartesian cases\nAll cases discussed up to now handled an adiabatic equilibrium con\fguration with or without the inclusion of \row.\nThe up-down symmetry of all the MHD spectra shown so far is perfectly maintained, related to the fact that these\ncases are in essence time-reversible. Every instability (or overstability in the case with \row) has a damped counterpart.\nWe now move on to include additional e\u000bects. Hence, we now compute spectra for time-irreversible cases, where either\nresistivity or other non-adiabatic e\u000bects enter, which will break the up-down symmetry. Here we \frst focus on the\ninclusion of resistivity.\n3.3.1. Resistive homogeneous plasma19\nFigure 9. Panela: spectrum for a homogeneous medium with constant resistivity. Panel bzooms in near the origin, showing\nthe start of the fast mode sequence. Panel czooms in further, revealing the semi-circles traced out by the Alfv\u0013 en and slow\nmodes (outer and inner semi-circles, respectively).\nFirst we look at the most simple con\fguration, that is, a homogeneous plasma in a Cartesian geometry with resistivity\nincluded. The uniform equilibrium is given by\n\u001a0= 1; B 02= 0; B 03= 1; T 0=1\n2\fB2\n0; (39)\nwhere we take a plasma beta of \f= 0:25,k2=ky= 0,k3=kz= 1 andx2[0;1]. The value for the resistivity is\nassumed to be constant and given by \u0011= 10\u00003, as described in Goedbloed et al. (2019). This spectrum is calculated\nusing 1001 gridpoints, the result is shown in Figure 9. In ideal MHD the fast modes form a Sturmian sequence (that is,\nthe oscillation of the eigenfrequencies increases when the number of modes increases, which in turn implies a larger real\npart of!) of stable fast magneto-acoustic waves with frequencies accumulating to in\fnite frequency (related to the p-\nmodes in our strati\fed example). The slow modes have an anti-Sturmian sequence towards their accumulation point (in\nessence the collapsed slow continuum) and the Alfv\u0013 en modes are degenerate (Goedbloed et al. 2019). When resistivity\nis included the fast modes become damped and the Alfv\u0013 en and slow modes trace out semi-circles in the bottom-half\nof the complex plane. The semi-circles and initial fast mode sequence shown in Figure 9 are in perfect agreement with\nthe spectrum depicted in Goedbloed et al. (2019, \fg. 14.6). These semi-circles can be quanti\fed analytically, their\nradius does not depend on the resistivity. The magnetic Reynolds number, calculated as Rm= (x1\u0000x0)cA=\u0011, is equal\nto 1000.\nDue to the rather extreme resolution employed here we trace much further into the fast mode sequence, where we\nsee something interesting: the fast modes appear on curves that loop around, breaking the purely Sturmian behaviour\nfor a moment, after which they continue again towards in\fnity. This implies that initially the fast modes become more\ndamped at higher mode frequencies up to a certain turning point at which they achieve maximal damping (the bottom\nof the loop). After passing this turning point the oscillation frequency of the modes increases again and the damping\nfrequency seems to converge towards one single value.\nOf course, the strong damping for the fast modes as we go further into the fast mode sequence must have consequences\nfor the original uniform (ideal) equilibrium state. In what follows, we will adopt the common practice to compute\nresistive MHD spectra about an ideal MHD state, which itself will evolve when resistivity is acting.20\n3.3.2. Quasi-modes in resistive MHD\nWe now turn to a non-adiabatic case, where resistivity is important. We will compute the resistive MHD spectrum for\na case where the equilibrium varies across an interface which gives rise to so-called quasi-modes. These are essentially\nsurface waves undergoing damping, due to the fact that the global quasi-mode overlaps in frequency with the continuum\nrange, causing resonant absorption. Quasi-modes are quite important in solar physics, since they can be indirectly\nrelated to the coronal heating problem as discussed in for example Poedts et al. (1989); Poedts & Kerner (1991). A\ndetailed analytical treatment of quasi-modes including theoretical growth rates is given in Priest (2014), where they\nstart from an inhomogeneous layer of width lconnecting two regions of uniform plasma. This can be reproduced by\nintroducing a linear density pro\fle between two homogeneous regions in a Cartesian geometry, however, this would\nmean that the density derivative shows rather strong discontinuities near the edges of the transition layer. We therefore\nopt for a smooth pro\fle by introducing a sine dependence such that\n\u001a0(x) =8\n>>>><\n>>>>:\u001a1 ifx0\u0014x>>>>>><\n>>>>>>>:\u00110\n2+\u00110\n2 tanh\u0019tanh\u00122\u0019(x\u0000sL)\nl\u0013\nifsL\u0000l\n2\u0014x\u0014sL+l\n2;\n\u00110 ifsL+l\n2>>><\n>>>>:4 (x\u0000xj\u00001) (xj\u0000x)\n(xj\u0000xj\u00001)2;\n0;\n0;C1\nj(x) =8\n>>>>><\n>>>>>:\u0012x\u0000xj\u00001\nxj\u0000xj\u00001\u00132\u0012\n3\u00002x\u0000xj\u00001\nxj\u0000xj\u00001\u0013\nforxj\u00001\u0014x\u0014xj;\n\u0012xj+1\u0000x\nxj+1\u0000xj\u00132\u0012\n3\u00002xj+1\u0000x\nxj+1\u0000xj\u0013\nforxj\u0014x\u0014xj+1;\n0 elsewhere ;\nQ2\nj(x) =8\n>>>>><\n>>>>>:(2x\u0000xj\u0000xj\u00001) (x\u0000xj\u00001)\n(xj\u0000xj\u00001)2;\n(2x\u0000xj+1\u0000xj) (x\u0000xj+1)\n(xj+1\u0000xj)2;\n0;C2\nj(x) =8\n>>>>><\n>>>>>:(x\u0000xj)\u0012x\u0000xj\u00001\nxj\u0000xj\u00001\u00132\nforxj\u00001\u0014x\u0014xj;\n(x\u0000xj)\u0012xj+1\u0000x\nxj+1\u0000xj\u00132\nforxj\u0014x\u0014xj+1;\n0 elsewhere ;(A3)\nas in for example Kerner et al. (1998); Goedbloed et al. (2019), and are shown graphically along with their derivatives\nin Figure 16. Since we have two basis functions per subdomain, this allows for the approximation of both the original\nvariable and its derivative by di\u000berentiating Eq. (A2). To turn the problem algebraic in nature, we use the Galerkin\nmethod such that the eigenvalue problem is written as a set of integral equations by multiplying each of the eight\nequations by an appropriate element of the chosen basis, denoted by hj, and integrate over the relevant domain.\nMathematically, this can be written asZ\n\n(AX\u0000!BX)hjdu1= 0: (A4)31\nFigure 16. The quadratic (a) and cubic (b) basis functions for the interval [ xj\u00001;xj+1] along with their derivatives (c) and\n(d), respectively. The solid lines denote H1, the dashed lines H2.\nHowever, the matrix Acontains second-order derivatives with respect to u1. In order to reduce these to derivatives\nof \frst order, hence simplifying the integrals, we make use of the Galerkin weak formulation. This is achieved by\nperforming integration by parts, which introduces additional surface terms that have to be evaluated at the boundaries.\nThese surface terms can in turn be exploited to enforce boundary conditions, which will be discussed in Subsection\nA.3. Since there are eight unknowns in our eigenvalue system, together with two basis functions and Nsubdomains,\nthis implies that the \fnal matrix eigenvalue problem will have 16 Nequations, resulting in a 16 N\u000216Nsize matrix.\nBy extension, it becomes clear that using basis functions of even higher order will increase the size of this eigensystem\nconsiderably.\nA.2. Matrix assembly\nThe actual expression for the matrix elements can be obtained by applying Eq. (A4) to the system of di\u000berential\nequations (19)-(26). As an example we will look at the v1component of the linearised continuity equation (19),\ncorresponding to element (1 ;2) in the matrices. This \\number\" links to the indices of the state vector X, since the\ncontinuity equation is associated with \u001a1, or index 1 in the state vector (A1), and the v1component is associated with\nindex 2. How these indices translate to the actual position in each matrix will be discussed further in this section.\nIn the \fnite element representation adopted here, the ( \u001a1;v1) contribution can be expanded as\nNX\nj=0!Z1\n\"h1\njh1\nkdu1=\u0000NX\nj=0Z1\n\"\u001a0\n0h1\njh2\nkdu1\u0000NX\nj=0Z1\n\"\u001a0h1\njdh2\nk\ndu1du1; (A5)\nwhere theu1-dependence of the equilibrium density \u001a0and the basis functions h1andh2is implied. In this case h1\nis quadratic, since \u001a1is associated with a quadratic basis function; analogously h2is cubic. The matrix elements for\nthis particular contribution are hence given by\nBjk(1;1) =Z1\n\"h1\njh1\nkdu1;Ajk(1;2) =\u0000Z\u00121\n\"\u001a0\n0\u0013\nh1\njh2\nkdu1\u0000Z\u00121\n\"\u001a0\u0013\nh1\njdh2\nk\ndu1du1: (A6)\nIf this reasoning is applied to all equations in the linear system, it follows that Bwill only have equal-number elements,\nimplying thatBis fully symmetric and real. Aon the other hand will have cross-term elements such that it is, in\ngeneral, not symmetric. Furthermore, it might be complex, depending on the included physical e\u000bects. Terms that\ncontain derivatives of the state vector components, as for example ( v1;\u001a1) which corresponds to element (2, 1), will\nbe integrated by parts. It is exactly this integration that gives rise to the surface terms, that is, terms that do not\ncontain an integral which translate into the natural boundary conditions. These are discussed in the next subsection.\nThe actual assembly of both matrices AandBinLegolas is done by sequential iteration over the gridpoints. From\n(A3) we see that \fnite elements have a localised nature around every gridpoint j, meaning that only the elements\nassociated with a certain region (that is, the gridpoint jitself and its neighbours j\u00001 andj+ 1) yield a non-zero\ncontribution. However, it is actually easier implementation-wise to loop over the elements in the interval [ xj\u00001;xj]\ninstead of over those in [ xj\u00001;xj+1], since then the actual integration of the matrix elements can be done in the same\nway, independent of whether the basis functions are cubic or quadratic. If only the interval [ xj\u00001;xj] is considered we32\nend up with 16 possible combinations of the basis functions for every gridpoint. This translates into a 4 \u00024 sub-matrix\nfor every variable, where every one of the 16 elements corresponds to one speci\fc combination of the shape functions.\nAs there are eight variables in the state vector this implies a 32 \u000232 matrix block for every gridpoint, hereafter dubbed\na \\quadblock\" since it consists of four 16 \u000216 blocks (hereafter called \\subblocks\"). Every one of those subblocks\ninside a quadblock corresponds to a quarter section of the aforementioned sub-matrix, which is a 2 \u00022 block in every\nsubblock. Hence, to recap, a 2 \u00022 block times eight variables represents a 16 \u000216 subblock, of which four combined\nform a 32\u000232 quadblock for every gridpoint.\nOf course, every matrix element contains one or more integrals that still have to be calculated. Since the coe\u000ecient\nfunctions are in general complicated expressions depending on u1, this is done numerically using a 4-point Gaussian\nquadrature for which an integral in the interval [ xj\u00001;xj] can be expressed as\nZxj\nxj\u00001f(u1)du1\u00191\n2(xj\u0000xj\u00001)4X\ni=1wif\u00121\n2(xj\u0000xj\u00001)\u0018i+1\n2(xj\u00001+xj)\u0013\n; (A7)\nwhere\u0018iandwiare the evaluation points and weights of the Gaussian integration. These values can be found in various\ntextbooks, as for example given in Goedbloed et al. (2019). The function fdenotes the integral coe\u000ecients, which\nare essentially the equilibrium quantities and basis functions evaluated in the various evaluation points. This actually\nimplies that every grid interval [ xj\u00001;xj] is subdivided into four points, meaning that the equilibrium expressions are\nprobed using 4( N\u00001) points rather than Npoints. The way the matrices are then assembled is thus done on a\ndouble-loop basis, where the outer loop iterates over the intervals [ xj\u00001;xj] and the inner loop iterates over the four\nGaussian points. This inner loop will calculate the basis functions and matrix elements at every point, then multiply\nthe coe\u000ecients with the Gaussian weights and \fnally add them all together in a consistent manner.\nSince every quadblock corresponds to the interval [ xj\u00001;xj] we still have to account for the contribution of the xj+1\ngridpoint. This is done by partially overlapping the quadblock in the next gridpoint with the one from the previous\ngridpoint. Figure 17 shows a visual representation of the structure and assembly process for the Amatrix, using\nthe Kelvin-Helmholtz and current-driven equilibrium discussed in Section 3.2.3 with only six gridpoints here for the\npurpose of illustration. The left panel shows the general matrix, highlighting the block-tridiagonal structure. The\ndashed grey lines denote the 32 \u000232 quadblocks of the matrix, and every dot represents a non-zero value. In total\n\fve quadblocks can be distinguished for six gridpoints, one for every grid interval. The middle panel shows a zoom-in\nof the quadblock corresponding to the second grid interval as annotated on the left panel, where it can be seen that\nthe top-left corner of this quadblock overlaps with the bottom-right corner of the previous quadblock corresponding\nto the \frst grid interval.\nA single quadblock is further divided into four subblocks, with the location of the di\u000berent state variables annotated\non the middle panel of Figure 17. The matrix element A(2;5) is highlighted in every subblock which corresponds to\nthe (v1;T1) contribution, which are cubic ( v1) and quadratic ( T1) variables. The 2 \u00022 block in the top-left corner of\nthe quadblock corresponds to the top-left 2 \u00022 corner of the right panel, as indicated by the background colours. The\nright panel shows the various combinations of the regular basis functions for the A(2;5) contribution, where the blue\ncurve corresponds to the cubic basis functions and the orange curves to the quadratic ones. It should be noted that\nthe right panel shows one speci\fc case, that is, the regular h2\njh5\nkterms of the matrix element. If for example the h2\njh7\nk\nelement is calculated, corresponding to the a2term in (20), the quadratic basis functions have to be replaced by their\ncubic counterparts, since a2is also a cubic variable. A similar reasoning can be made for the other matrix elements.\nThe termpH\u000b\f\njH\u000b\f\njat the top of every sub-panel on the right of Figure 17 denotes which combination of the basis\nfunctions should be used, where pstands for the integral coe\u000ecients. The exponent \u000b\frefers to the expressions for\nthe basis functions in (A3), where \u000b= 1 stands for Q1\njorC1\njand\u000b= 2 stands for Q2\njorC2\nj, depending on the variable\nunder consideration. \fstands for which part of the basis function that should be taken, \f= 1 means the \frst equation\nin cases, while \f= 2 means the second one. As an example we can look at H12\njH21\nj: sinceA(2;5) represents a cubic\nand quadratic variable, this translates into C12\njQ21\nj. For the cubic part we therefore take the second equation of C1\nj,\ncorresponding to xj\u0014x\u0014xj+1. The quadratic part on the other hand is given by Q21\nj, which means that we take\nthe \frst equation of Q2\nj, corresponding to xj\u00001\u0014x\u0014xj. Boundary conditions are imposed after matrix assembly is\ncompleted.33\nFigure 17. General assembly and structure of the \fnite element (complex) Amatrix. Left: Example of a full matrix for six\ngridpoints, showing the block-tridiagonal structure where dots represent non-zero values. The middle panel zooms in on one\nquadblock, showing the dependence of di\u000berent subblock positions with respect to the variables. The 2 \u00022 blocks corresponding\ntoA(2;5) (cubic, quadratic) are highlighted. Right panel: 2 \u00022 block assembly for a general A(cubic, quadratic) matrix\nelement. Cubic elements are shown in blue, quadratic ones in orange. The dotted grey line denotes zero.\nA.3. Implementation of boundary conditions\nIntegration by parts on (A4) gives rise to additional surface terms, which should be evaluated at the boundaries.\nThese kind of conditions are called natural boundary conditions, since they emerge in a natural way by rewriting the\neigenvalue problem. The regularity and \fxed wall conditions considered earlier on the other hand are called essential\nboundary conditions, and have to be handled explicitly. Since the additional surface terms originate from reducing\nsecond-order derivatives to \frst order derivatives, we only have these terms for the variables v1;T1;a2anda3, as these\nare the only equations that contain derivatives of higher order. Hence, for the momentum equation, the additional\nsurface terms can be written as\nSv1=\"\nT0\n\"h2\njh1\nk+\u001a0\n\"h2\njh5\nk+\u0012\nB02k3\u00001\n\"B03k2\u0013\nh2\njh6\nk+B03\n\"h2\njdh7\nj\ndu1\u0000B02h2\njdh8\nk\ndu1#\n@\n=v1\u0014T0\n\"\u001a1+\u001a0\n\"T1+\u0012\nB02k3\u00001\n\"B03k2\u0013\na1+B03\n\"a0\n2\u0000B02a0\n3\u0015\n@\n;(A8)\nwhere the number in superscript on hj=kdenotes the index of the variable in the state vector X. The subscript @\nmeans that these terms should be evaluated at the left- or inner edge, as well as at the right- or outer edge. In a\nsimilar manner are the surface terms for the energy and induction equations given by\nST1=i(\r\u00001)\n\"T1\"\nT0\n0@\u0014?\n@\u001a\u001a1+\u0012\nT0\n0@\u0014?\n@T\u0000\"0\n\"\u0014?\u0013\nT1+\u0014?T0\n1\n+ 2\u0012\nT0\n0\u0000\n\"B02k3\u0000B03k2\u0001@\u0014?\n@(B2)\u0000\u0011B0\n03k2+\u0011k3(\"B02)0\u0013\na1\n+ 2\u0012\nT0\n0B03@\u0014?\n@(B2)+\u0011B0\n03\u0013\na0\n2\u00002\u0012\n\"T0\n0B02@\u0014?\n@(B2)+\u0011(\"B02)0\u0013\na0\n3#\n;(A9)\nSa2=i\u0011\n\"a2\u0010\n\u0000k2a1+a0\n2\u0011\n; (A10)\nSa3=i\u0011a3\u0010\n\u0000k3a1+a0\n3\u0011\n; (A11)\nwhich should all be evaluated at the boundaries. For the case of a solid wall we see that the natural boundaries\nsimplify considerably, since if v1;a2;a3are all zero at the wall, Sv1;Sa2andSa3are also zero and have to be omitted.\nThe natural boundary condition on T1is only relevant if resistivity or perpendicular thermal conduction is included.\nHowever, in the case of the latter, the additional essential boundary condition requires that T1= 0, in which case34\nST1drops out as well. The only combination in which the surface terms for the energy equation are non-zero is when\nresistivity is included, but perpendicular thermal conduction is omitted. In that case the resistive heating terms should\nbe included in the calculation, which is done by adding the appropriate terms to the matrix elements A(5;6);A(5;7)\nandA(5;8).\nCurrently only \fxed wall boundary conditions are implemented in Legolas . However, the surface terms described\nhere can be used to impose other types of boundary conditions as well. In the case of a plasma-vacuum-wall transition\nwe have for example Bessel functions at the outer boundary of a cylindrical geometry (Roberts 2019), which encode\nthe analytic vacuum solution for the electromagnetic \feld in the outer vacuum region. These expressions can then be\nused to rewrite the surface terms (A8)-(A11) in an appropriate way such that they can be added to their respective\nsubblock positions in the matrix. This is a planned extension to be included in future versions of Legolas . This\nfunctionality was previously available in some LEDA versions (Van der Linden et al. 1992).\nThe essential boundary conditions as described in Section 2.3 have to be implemented explicitly. This is done by\nomitting the relevant basis functions that do not satisfy the boundary conditions on the edges. Consider as an example\nthe variable v1, which is associated with a cubic basis function. From Figure 16 we see that the only cubic element\nwhich is non-zero at the left boundary is C12\nj, which implies that the matrix elements where it appears should be\nzeroed out. Looking back at how the quadblock is composed in Figure 17, the boundary condition v1= 0 corresponds\nto forcing the odd rows and columns of the v1contribution to zero, for subblocks 1, 2 and 3 on the left boundary (that\nis, the \frst node j) since these correspond to the 2 \u00022 blocks in blue, green and red on the right panel. Similarly, on\nthe right side (viz. the last node j) onlyC11\njis non-zero, which implies that the odd rows and columns of subblocks\n2, 3 and 4 have to be zeroed out (corresponding to green, red and yellow). Extending this reasoning to the essential\nboundary condition on T1, we see that in this case the even rows and columns should be handled since T1is associated\nwith a quadratic element. This is done for both matrices.\nOf course, \\just\" zeroing out rows and columns in a matrix has the unpleasant side-e\u000bect that the matrices become\nsingular. For the Amatrix this is not necessarily a problem, however, the Bmatrix can never be singular since it\nis inverted when solving the general eigenvalue problem using the QR algorithm. Therefore, we introduce a one on\nB's diagonal at the location that was zeroed out, and an element \u000eonA's diagonal. The e\u000bect of this is that one\nessentially \\forces\" the boundary condition, since this implies that \u000exi\nj=!xi\nj. By extension, if \u000eis taken to be a large\nnumber (we take \u000e= 1020), this means that xi\nj= 0, which corresponds to the essential boundary condition we wanted\nto impose. The only side-e\u000bect of this approach is that it introduces eigenvalues equal to \u000e. However, since \u000eis taken\nto be large, these will not in\ruence the spectrum in any way and they can be easily \fltered out during post-processing.\nThis method thus provides a relatively easy and straightforward way to impose Dirichlet boundary conditions at the\nedges. The imposed boundary conditions are noticeable on the left panel of Figure 17, especially for the \frst node.\nThe odd rows and columns that were zeroed out can clearly be seen, together with the large numbers introduced on\nthe main diagonal.\nB.ERRATUM: \\LEGOLAS: A MODERN TOOL FOR MAGNETOHYDRODYNAMIC SPECTROSCOPY\"\n(2020, APJS, 251, 25)\nIn this work, published in ApJS 251, 25 (2020), we reported on the Legolas code, where we gave a detailed overview\nof the code itself and discussed its application on various equilibria. The case in Section 3.3.4 treated so-called rippling\nmodes, which may arise whenever there is a spatially varying resistivity pro\fle present. Herein we imposed a hyperbolic\ntangent pro\fle for the resistivity \u0011(x), and showed a spectrum that has multiple unstable branches on the left and\nright side of the imaginary axis as depicted here in Fig. 18, middle left panel. Based on the very localized nature of\nthe eigenfunctions, we concluded that these were rippling modes alongside an already present tearing mode. However,\nduring an extension of the code we discovered a bug in one of the terms of the matrix elements A(7;5) andA(8;5),\nwhich both correspond to the T1resistive components in the linearized a2anda3equations. More speci\fcally, the\nresistivity derivative with respect to temperature d\u0011=dT was erroneously treated as a spatial derivative. This implies\nthat these two elements were nonzero, while they should vanish for the imposed temperature-independent resistivity\npro\fle. These nonzero values in turn led to a modi\fcation of the spectrum and gave rise to the two unstable branches.\nAfter these two terms were corrected, the rippling modes are no longer present as can be seen in the right-hand side\nof Figure 18, while the tearing mode and its associated eigenfunctions remain una\u000bected. The large-scale spectrum is\nbarely modi\fed, however on smaller scales it can be seen that there is a major e\u000bect on the damped slow and Alfv\u0013 en35\nFigure 18. Spectra before (left) and after (right) correction of the erroneous terms. The large-scale spectrum remains mostly\nuna\u000bected (top panels), as are the tearing mode and its eigenfunctions (bottom panels). The damped slow and Alfv\u0013 en sequences\n(middle panels) are in fact modi\fed, and there are no longer rippling modes present. The tearing mode is annotated with a\ncyan cross.\nsequences. After correction of the terms these sequences better trace out the semicircles in the complex plane, in much\ncloser resemblance to the original tearing mode spectra in Section 3.3.3 of the original work.\nIt should be noted that the general conclusions regarding rippling modes remain valid. For the spatially varying\nresistivity pro\fle that we imposed there are no rippling modes, but for a more general \u0011(x;T(x)) pro\fle the two matrix\nelements discussed here will indeed be nonzero. As such it is entirely possible that for some pro\fles rippling modes\nmay arise, and the question of rippling- versus tearing mode dominance in more realistic resistivity pro\fles remains\nrelevant.\nREFERENCES\nAnderson, E., Bai, Z., Bischof, C., et al. 1999, LAPACK\nUsers' Guide, 3rd edn. (Philadelphia, PA: Society for\nIndustrial and Applied Mathematics)\nBalbus, S. A., & Hawley, J. F. 1991, The Astrophysical\nJournal, 376, 214\nBallester, J. L. 2006, Space science reviews, 122, 129\nBarbulescu, M., Ruderman, M. S., Van Doorsselaere, T., &\nErd\u0013 elyi, R. 2019, The Astrophysical Journal, 870, 108\nBaty, H., & Keppens, R. 2002, The Astrophysical Journal,\n580, 800Beli en, A., Botchev, M., Goedbloed, J., van der Holst, B.,\n& Keppens, R. 2002, Journal of computational physics,\n182, 91\nBerger, T. E., Liu, W., & Low, B. 2012, The Astrophysical\nJournal Letters, 758, L37\nBlokland, J., Keppens, R., & Goedbloed, J. 2007a,\nAstronomy & Astrophysics, 467, 21\nBlokland, J., van der Holst, B., Keppens, R., & Goedbloed,\nJ. 2007b, Journal of computational physics, 226, 509\nChandrasekhar, S. 2013, Hydrodynamic and hydromagnetic\nstability (Courier Corporation)36\nChoudhuri, A. R. 1998, The Physics of Fluids and Plasmas:\nAn Introduction for Astrophysicists (Cambridge\nUniversity Press), doi: 10.1017/CBO9781139171069\nClaes, N., Keppens, R., & Xia, C. 2020, Astronomy &\nAstrophysics, 636, A112\nColgan, J., Abdallah Jr, J., Sherrill, M., et al. 2008, The\nAstrophysical Journal, 689, 585\nDalgarno, A., & McCray, R. 1972, Annual review of\nastronomy and astrophysics, 10, 375\nDemaerel, T., & Keppens, R. 2016, Physics of Plasmas, 23,\n122118\nEngvold, O. 1998, in Astronomical Society of the Paci\fc\nConference Series, Vol. 150, IAU Colloq. 167: New\nPerspectives on Solar Prominences, ed. D. F. Webb,\nB. Schmieder, & D. M. Rust, 23\nGoedbloed, H., Keppens, R., & Poedts, S. 2019,\nMagnetohydrodynamics of Laboratory and Astrophysical\nPlasmas (Cambridge University Press),\ndoi: 10.1017/9781316403679\nGoedbloed, J. 2018a, Physics of Plasmas, 25, 032109\n|. 2018b, Physics of Plasmas, 25, 032110\nGoedbloed, J., Beli en, A., van der Holst, B., & Keppens, R.\n2004, Physics of Plasmas, 11, 28\nGoedbloed, J., Holties, H., Poedts, S., Huysmans, G., &\nKerner, W. 1993, Plasma physics and controlled fusion,\n35, B277\nGoedbloed, J. H. 2011, Plasma Physics and Controlled\nFusion, 53, 074001\nHillier, A. 2018, Reviews of Modern Plasma Physics, 2, 1\nHillier, A., Barker, A., Arregui, I., & Latter, H. 2019,\nMonthly Notices of the Royal Astronomical Society, 482,\n1143\nHillier, A., & Polito, V. 2018, The Astrophysical Journal\nLetters, 864, L10\nKeppens, R., Casse, F., & Goedbloed, J. 2002, The\nAstrophysical Journal Letters, 569, L121\nKeppens, R., Van Der Linden, R. A., & Goossens, M. 1993,\nSolar physics, 144, 267\nKerner, W., Goedbloed, J., Huysmans, G., Poedts, S., &\nSchwarz, E. 1998, Journal of computational physics, 142,\n271\nKerner, W., Lerbinger, K., Gruber, R., & Tsunematsu, T.\n1985, Computer physics communications, 36, 225Mackay, D., Karpen, J., Ballester, J., Schmieder, B., &\nAulanier, G. 2010, Space Science Reviews, 151, 333\nMacTaggart, D. 2018, Journal of Plasma Physics, 84,\n905840501, doi: 10.1017/S0022377818001009\nNakariakov, V., & Ofman, L. 2001, Astronomy &\nAstrophysics, 372, L53\nNijboer, R., Holst, B., Poedts, S., & Goedbloed, J. 1997,\nComputer physics communications, 106, 39\nParker, E. N. 1953, The Astrophysical Journal, 117, 431\nPoedts, S., Goossens, M., & Kerner, W. 1989, Solar\nphysics, 123, 83\nPoedts, S., & Kerner, W. 1991, Physical review letters, 66,\n2871\nPriest, E. 2014, Magnetohydrodynamics of the Sun\n(Cambridge University Press),\ndoi: 10.1017/CBO9781139020732\nRappaz, J. 1977, Numerische Mathematik, 28, 15\nRempel, M. 2012, The Astrophysical Journal, 750, 62\nRoberts, B. 2019, MHD Waves in the Solar Atmosphere\n(Cambridge University Press),\ndoi: 10.1017/9781108613774\nRosner, R., Tucker, W. H., & Vaiana, G. 1978, The\nAstrophysical Journal, 220, 643\nRuan, W., Xia, C., & Keppens, R. 2019, The Astrophysical\nJournal Letters, 877, L11\nSchure, K., Kosenko, D., Kaastra, J., Keppens, R., & Vink,\nJ. 2009, Astronomy & Astrophysics, 508, 751\nvan der Holst, B., Sokolov, I. V., Meng, X., et al. 2014, The\nAstrophysical Journal, 782, 81\nVan der Linden, R., & Goossens, M. 1991a, Solar physics,\n134, 247\n|. 1991b, Solar physics, 131, 79\nVan der Linden, R., Goossens, M., & Hood, A. 1992, Solar\nphysics, 140, 317\nVan Doorsselaere, T., & Poedts, S. 2007, Plasma Physics\nand Controlled Fusion, 49, 261\nWang, C., Blokland, J., Keppens, R., & Goedbloed, J.\n2004, Journal of plasma physics, 70, 651\nXia, C., & Keppens, R. 2016, The Astrophysical Journal,\n823, 22\nXia, C., Keppens, R., & Fang, X. 2017, Astronomy &\nAstrophysics, 603, A42" }, { "title": "0707.0776v1.Damping_of_bulk_excitations_over_an_elongated_BEC___the_role_of_radial_modes.pdf", "content": "arXiv:0707.0776v1 [cond-mat.other] 5 Jul 2007Damping of bulk excitations over an elongated BEC - the role o f radial modes\nE. E. Rowen,∗N. Bar-Gill, R. Pugatch, and N. Davidson\nWeizmann Institute of Science, Rehovot, Israel, 76100\n(Dated: October 31, 2018)\nWe report the measurement of Beliaev damping of bulk excitat ions in cigar shaped Bose Einstein\ncondensates of atomic vapor. By using post selection, excit ation line shapes of the total population\nare compared with those of the undamped excitations. We find t hat the damping depends on the\ninitial excitation energy of the decaying quasi particle, a s well as on the excitation momentum. We\nmodel the condensate as an infinite cylinder and calculate th e damping rates of the different radial\nmodes. The derived damping rates are in good agreement with t he experimentally measured ones.\nThe damping rates strongly depend on the destructive interf erence between pathways for damping,\ndue to the quantum many-body nature of both excitation and da mping products.\nEver since Wigner and Weisskopf first calculated the\ndamping rate of an atom coupled to the vacuum [1], it\nis known that coupling to a bath affects the decay of\nan otherwise stable quantum system. The structure of\nthe bath plays a key role in determining the rate of the\ndamping, and may even cause the damping to be sub-\nexponential or super-exponential [2]. Gaseous Bose Ein-\nstein condensates (BEC) are usually well isolated from\ntheir surroundings, leading to coherent evolution. Three\nbody loss, which is the main cause for decoherence in the\ngroundstate,canbeontheorderofseconds[3]. However,\ndamping of the elementary excitations over the conden-\nsate is much more rapid, but still accessible experimen-\ntally. The bath, which is coupled to the excitation, is\nin this case a quasi-continuum of unoccupied excitations.\nThese excitations, known as Bogoliubov quasi particles,\nare approximate eigen states of the Bosonic many-body\nHamiltonian. Coupling of an excited quasi particle (QP)\nto initially unoccupied QP modes gives rise to decoher-\nence via the Beliaev damping mechanism, in which a\nQP decays into two new QPs while fulfilling energy and\nmomentum conservation [4]. Another damping process,\nthat takes place at a finite temperature, is called Lan-\ndau damping. This process involves the annihilation of\na thermally excited QP together with the damped one,\nand the creation of a more energetic QP, also conserving\nmomentum and energy [5].\nDiscrete Beliaev coupling of excitations was measured\nin [6, 7]. Since the spectrum is discrete, such Beliaev\ncouplingoflowenergyexcitationsis rare,and evenatlow\ntemperatures, damping is usually governed by Landau\nprocesses [5, 8, 9, 10, 11]. Damping of bulk excitations\nis due mainly to Beliaev coupling at low temperatures\n[11]. Such damping was measured [12], and found to\nhave strong momentum dependence, in agreement with\n[4]. However, the dependence of the Beliaev damping on\nthe excitation energy was not measured.\nIn this letter we study the effects of the damping on\nthe excitation line shape in an elongated condensate.\nBy comparing the overall response to Bragg excitations,\nwith the response of the undamped fraction, the damp-\ning rates are quantified as a function of both momentumand energy of the decaying QP. In addition to the mo-\nmentum dependence of the damping rate, we observe a\ndependence on the excitation energy. By modelling the\ncondensate as an infinite cylinder [13],[14], we calculate\nthe damping ratesofthe different elementarymodes, and\nfind good agreement with the measured line shapes. Ac-\ncording to the model, the spatial dependence of the QP\nwave functions plays a key role in the difference between\nthe damping rates. The different energy dependence at\ndifferent momentum excitations is a result of quantum\ninterference due to many-body effects.\nWe create a BEC of N= 4×105 87Rb atoms in\nthe internal state |F= 2,mF= 2/angb∇acket∇ightin a cylindrically sym-\nmetric Ioffe-Pritchard trap with trapping frequencies of\nωρ= 2π×350Hz and ωz= 2π×30Hz. The Thomas-\nFermi radii of the condensate are RTF≈3µm and\nZTF≈36µm, andthechemicalpotentialis µ/h≈4kHz.\nWe excite the condensate by shining it with two off res-\nonant laser beams detuned by 0 .2nm from the D2 tran-\nsition for a period of 2ms (Bragg excitation). Then we\nrapidly shut the magnetic trap off, and image the atomic\ncloud after 38ms time of flight. By varying both the\nfrequency detuning ωand the angle αbetween the two\nbeams, we can measure the response of the condensate\nto an excitation with momentum /planckover2pi1k= 2/planckover2pi1kLsin(α/2)\nin the axial direction ˆ z, and with energy /planckover2pi1ω. Here\nkL= 2π/780nm−1is the wave number of the lasers.\nSincekLZTF≫1, the wave vector of the excitation k\nis a good quantum number.\nA unique feature of experiments with BEC is the abil-\nity to observe the damping products in a single image.\nIn the insets of Fig. 1 we present time of flight images of\nan excited condensate. In both images roughly one third\nof the condensate atoms were excited to a QP mode with\nmomentum 2 /planckover2pi1kL, but in Fig. 1b, the damping is much\nlarger. This is evident in the plot of the normalized mo-\nmentum distribution in Fig. 1. For the lower energy ex-\ncitation (a), the peak in the momentum distribution at\nk= 2kLis more pronounced. For the larger energy ex-\ncitation (b), the damping products are dominant. The\nstrong reduction of the condensate fraction along with\nthe tendency of the damping products toward lower ax-2\n−1012300.020.040.060.080.1\n(a)\nmomentum [kL]momentum distribution [a.u.]\n−1012300.020.040.060.080.1\n(b)\nmomentum [kL]momentum distribution [a.u.]\nexcitation excitationBEC BEC\nFIG. 1: Beliaev damping of condensates excited with momen-\ntum 2/planckover2pi1kL, and different frequencies: 14 kHz (a) and 18 kHz\n(b). Insets - absorption images of excited condensates afte r\n38ms time of flight. The graphs show the normalized momen-\ntum distribution p(k) along the axial direction ˆ z, which is a\nvertical integral of the absorption images in the insets. Th e\nfree-particle resonance is at 15kHz and the local density re s-\nonance for our condensate is 17 .5kHz. In (a) the undamped\nexcitation is larger than that of (b), while the cloud of Beli aev\ndamping products is larger in (b). As the Beliaev damping\nextracts atoms from the condensate, the condensate peak is\ngreatly suppressed in (b).\nial momentum are evidence of multiple collisions.\nIn order to quantify the damping of the excitations\nat different frequencies we employ a post-selection tech-\nnique [15]. We measure the response in two ways. First\nwe measure the average momentum along the ˆ zaxis, of\nall atoms in the expanded atomic cloud: /angb∇acketleftk/angb∇acket∇ight/2kL=/integraltext\ndk k·p(k)/2kL, wherep(k) is the momentum distri-\nbution. Since momentum along ˆ zis conserved during\ndamping, this is a measure of the excitation fraction in-\ncluding damped excitations and will be referred to as\noverall response. The second way to measure response is\nto count the fraction of atoms that remain in the excita-\ntion mode alone: N2kL/Ntot, where the occupation N2kL\nand the total number of atoms Ntotare determined by\na fit to the momentum distribution. This is referred to\nas the undamped population. In the absence of collisions\nthe two measuring methods are equivalent. The differ-\nence between the two measurements is a measure of the\nnumber of excitations that were damped.\nWe repeat this measurement for different values of ω.\nIn Fig. 2a we compare the obtained line shape for ex-\ncitations with momentum 2 /planckover2pi1kL. Empty circles are the\nmeasurements of the overall response, while the filled cir-\ncles are the undamped results of the same images. The\noverall response displays a peak at 16 .6kHz, less than\npredicted by the local density approximation [16]. This\nis due to the relatively large excitation fraction, which\nshifts the resonance towards the free particle value [15].\nThe width of the resonance is due mainly to the inhomo-\ngeneous density of the condensate. The response of the/s49/s48 /s49/s50 /s49/s52 /s49/s54 /s49/s56 /s50/s48/s48/s46/s48/s48/s46/s49/s48/s46/s50/s48/s46/s51\n/s49/s48 /s49/s50 /s49/s52 /s49/s54 /s49/s56 /s50/s48/s48/s46/s48/s48/s46/s49/s48/s46/s50/s48/s46/s51/s114/s101/s115/s112/s111/s110/s115/s101\n/s32/s91/s107/s72/s122/s93/s40/s98/s41/s40/s97/s41\nFIG. 2: (a) Line shape at k= 2kLmeasured in two ways: ( ◦)\nThe overall response /angbracketleftk/angbracketright/2kL, and (•) the undamped pop-\nulationN2kL/Ntot. The dashed and solid lines are gaussian\nfits centered at 16 .6 kHz and 15 .3 kHz, with widths of 4 kHz\nand 2.6 kHz respectively. The resonance in the line shape of\nthe undamped population is shifted down in energy and is\nnarrower than that of the overall response. The dotted line\nmarks the center of the resonance of the undamped popula-\ntion. (b) The corresponding theoretical curves obtained fr om\nour model, for the line shape of the overall response (dashed\nline), and that of the undamped population (solid line). The\nradial modes leading to the line shape are marked by bars,\npositioned at their corresponding energy, and with a height\nproportional to their overlap with the condensate.\nundamped population (filled circles) peaks at 15 .3kHz,\nand is 5 times smaller than that of the total population,\ndue to Beliaev damping. There is a clear shift down\nin the resonance of the undamped population, implying\nthat the damping rate is larger for the more energetic\nexcitations. This can be understood intuitively in a local\ndensity approximation: more energetic excitations are\nin spatial regions of larger density, leading to a higher\ndamping rate.\nIn Fig. 3a we present the response of the condensate to\nexcitations with smaller momentum - 1 .1/planckover2pi1kL. A reduc-\ntion in the damping as compared to the case of 2 /planckover2pi1kLis\nevident. This is a result of the quantum interference due\nto the collective nature of the low momentum excitations\nand therefore is present in a homogeneous BEC as well\n[12]. A gaussian fit to the overall response is centered\nat 6.48±0.04 kHz. The fit to the undamped line shape\nhas a peak at 6 .63±0.07 kHz. Contrary to the 2 kLcase,\nthe line shape of the undamped population is not shifted\ndown in energy, and the naive linear dependence of the\ndamping rate on the local density fails.\nTo theoretically account for these effects we include3\n/s48 /s50 /s52 /s54 /s56 /s49/s48/s48/s46/s48/s48/s46/s49/s48/s46/s50/s48/s46/s51/s48/s46/s52\n/s48 /s50 /s52 /s54 /s56 /s49/s48/s48/s46/s48/s48/s46/s49/s48/s46/s50/s48/s46/s51\n/s40/s98/s41/s40/s97/s41/s114/s101/s115/s112/s111/s110/s115/s101\n/s91/s107/s72/s122/s93\nFIG. 3: (a) Line shape at k= 1.1kLmeasured both ways: ( ◦)\nThe overall response /angbracketleftk/angbracketright/1.1kL, and (•) the undamped popu-\nlationN1.1kL/Ntot. All circles are averages over 4 data points.\nThe dashed and solid lines are gaussian fits giving a central\nfrequency of 6 .5kHz for the overall response and 6 .6kHz for\nthe undamped population. The dotted line marks the cen-\nter of the line shape of the undamped population. There is\nno downward shift of the resonance, and no narrowing as are\npresent in Fig. 2. (b) The corresponding theoretical curves\nobtained from our model, for the line shape of the overall\nresponse (dashed line), and the undamped population (solid\nline). The bars mark the frequency and overlap with the con-\ndensate of the relevant radial modes.\nspatial dependence. We exploit the cigar shape of the\ncondensate to neglect the ˆ zdependence of the ground\nstate, thus reducing the problem to 1D in the radial di-\nrection [13]. A similar treatment is performed in [10]\nto describe Landau damping of quadrupole oscillations\nof an elongated BEC. The ground state ψ0(ρ) of the\nGross-PitaevskiiHamiltonian HGP=−∇2/2M+V(ρ)+\ngN/L|ψ0|2is calculated by imaginarytime evolution, for\na radially dependent potential V(ρ) =1\n2Mω2\nρρ2and our\nexperimental parameters. Excitations over the conden-\nsate are obtained by solving the Bogoliubov equations:\n/planckover2pi1ωνuν(r) =/bracketleftbigg\nHGP+gN\nL|ψ0|2/bracketrightbigg\nuν+gN\nLψ2\n0vν(1)\n−/planckover2pi1ωνvν(r) =/bracketleftbigg\nHGP+gN\nL|ψ0|2/bracketrightbigg\nvν+gN\nLψ∗\n02uν.(2)\nThe wave functions uν(r) of Eqs. 1,2 are decoupled to\nuν(r) =un,m,k(ρ)eimφeikz,and so arevν(r). The set of\nquantum numbers νare in this case: k- the momentum\nalong ˆz,m- the vorticity around ˆ z, andn- the number\nof radial nodes of the wave functions u(ρ) andv(ρ). The\nenergy of the radial modes /planckover2pi1ωk,m,nincreases with nas\nwell as|k|and|m|. There is a conservation law for k56700.511.522.53\nω/2π [kHz]Gamma [ms−1]\n14161800.511.522.53\nω/2π [kHz]Gamma [ms−1](a) (b)\nFIG.4: Dampingratesversustheenergyofthedifferentradia l\nmodes for the first few radial modes with k= 1.1kL(a), and\nk= 2kL(b). (•) - the calculated damping rate. ( ◦) - the\ndamping rate obtained by taking only the terms involving\nuνuν1uν2in Eq. 3.\nandm(kcan only decay into q,k−q, andmintom1,m2\nfulfillingm1+m2=m), butthequantumnumber nisnot\nconserved. We consider the damping of excitations with\nm= 0 as the Bragg pulse carries no angular momentum\nalong ˆz. Still, all mvalues need to be considered as\ndamping products.\nThe Beliaev coupling is part of the next order\nexpansion of the many-body Hamiltonian: HB=/summationtext\nν,ν1,ν2Aν\nν1;ν2ˆb†\nν1ˆb†\nν2ˆbν+H.C. The three QP overlap\nAν\nν1;ν2= 2g/radicalbig\nN0/integraldisplay\ndr(uνuν1uν2+uνuν1vν2\n+uνvν1uν2+vνuν1vν2+vνvν1uν2+vνvν1vν2) (3)\ninvolves interference of six different quantum pathways.\nUsing first order time dependent perturbation theory\nwe calculate the damping as a function oftime. Since the\ndamping is nearly linear, we can extract damping rates\nΓν=1\nt4\n/planckover2pi12/summationdisplay\nm,n1,n2/integraldisplay\ndq/vextendsingle/vextendsingleAν\nν1;ν2/vextendsingle/vextendsingle2sin2(∆ωt/2)\n∆ω2,(4)\nfrom the different initially excited radial modes ν=\n(n,0,k), to all possible pairs of modes ν1= (n1,m,q)\nandν2= (n2,−m,k−q), with an energy mismatch of\n∆ω=ων−ων1−ων2These rates are plotted in Fig. 4\nversus the corresponding eigen-frequencies ωn,0,k.\nThe radial mode dependence of the damping rate is\ndifferent for the different initial momenta. The damp-\ning rate from radial modes with momentum k= 2kL\nincreases with n. In this regime, /planckover2pi1ω≫µ, excitations\nare nearly single-particle in nature and v(r)→0, for\nboth the decaying excitation, and most of the damping\nproducts. Therefore the damping is mostly governed by4\nthe term involving only uνuν1uν2. This can be seen by\nthe fact that the damping rates in Fig. 4b (filled circles),\nfollow the term with only u’s (empty circles) [17]. The\nincrease of both rates with radial mode energy is a re-\nsult of the number of available energy-conservingpairs of\ndamping products. As the energy of the decaying QP is\nincreased, more pairs are available, and the damping in-\ncreases. Our calculations show that the n= 4 mode has\ntwice as many possible damping products as the mode\nn= 0, in agreement with the damping rate which is ap-\nproximately doubled.\nDamping of modes with momentum 1 .1/planckover2pi1kLis differ-\nent. First, the energy spacing between modes is com-\nparable to the 2 kLcase, but the energies are smaller.\nThis is manifested in a largerrelative increase of avail-\nable modes with radial mode. In fact, the n= 3 mode\nhas 5 times more possible damping target modes than\nthen= 0 mode. Alone, this would increase the damping\nfrom the higher radial modes as happens with the empty\ncircles in Fig. 4a. This effect is compensated by interfer-\nence between the different terms in Eq. 3. Nearly all the\ndamping products have /planckover2pi1ω/lessorsimilarµ, and therefore, as the ra-\ndial number increases, the wave functions overlap denser\nregions of the condensate and v(r) becomes more signif-\nicant. Together, the increase due to phase space, and\nthe reduction due to quantum interference between the\nterms in Eq. 3 cancel, and the damping is similar from\nall radial modes for k= 1.1kLas seen in filled circles of\nFig. 4.\nFor even smaller momentum excitations, our model\npredicts non-exponential damping. However, the damp-\ning is so slow, that the infinite cylindrical approximation\nis no longer valid for our trap parameters.\nThe line shapes of the overall response, and the un-\ndamped population obtained by the model are presented\nbelow the data in Figs. 2b, 3b. The bars are positioned\nat the eigen-frequencies of the radial modes with the cor-\nresponding k(form= 0). The height of the bars is pro-\nportional to the overlap with the condensate, that deter-\nmines the response to the Bragg pulse [13]. The dashed\nlinesaretheexpectedoverallresponsetotheBraggpulse,\nobtained by summation over the responses of the differ-\nent radial modes including Fourier broadening of each\nmode, as in [13]. The solid line is obtained in a similar\nmanner after multiplying each response function by the\ncorresponding damping e−Γn,0,kteff. We further convolve\nthe line shape with a gaussianof width 300Hz to account\nfor residual sloshing in the trap, that changes the effec-\ntive detuning [18]. The suitable effective damping time,\nteff= 5ms, is 3ms longer than the Bragg pulse. This\nmay be due to collisions during the expansion of the con-\ndensate and enhancement of the damping due to thermal\npopulation of the damping products.\nGaussian fits to the calculated line shapes give res-\nonances at 6 .3kHz for both overall response and un-\ndamped population for the 1 .1kLexcitations. For the2kLexcitations, the model gives a line shape centered at\n16.5kHz for the overall response and at 15 .6kHz for the\nundamped population in good agreement with the mea-\nsuredvalues. Themodeldisplaysthesamenarrowingand\ndownward shift in the line shape of the undamped popu-\nlation at high momentum, indicating the larger damping\nratefrommoreenergeticexcitations. Bothmodelandex-\nperiment also show a uniform, much smaller damping of\nthesmallmomentumexcitations. Thelowestradialmode\nis more separated from the others, and distinguishable\nfrom the rest both in the model and in the experimental\ndata.\nInconclusion,weuseapostselectiontechniquetomea-\nsure the Beliaev damping of QPs in an elongated conden-\nsate. We find a dependence of the damping rate on both\nthe excitation energy and the momentum of the decayed\nQP. We measure a shift downward in the excitation en-\nergy of the undamped excitation for high momentum,\nwhile for small momentum we measure no such shift.\nModelling our system as infinite along the axial direc-\ntion, we obtain quantitative agreement between theory\nand experiment. The overall damping is suppressed for\nlowmomentumexcitationsasaresultofinterferingquan-\ntum pathways within each damping channel.\nThe inhomogeneous treatment includes a quantization\nof the bath in the radial direction. Since our excitation\nenergiesare above /planckover2pi1ωρ, we are only indirectly sensitive to\nthis quantization. Beliaev damping becomes even more\nintriguingwhentheradialtrappingfrequencyisincreased\nand the system becomes one dimensional. In this regime\nit could be possible to observe effects such as quantum\nZeno[2], nonexponentialdampingandeventhecomplete\ninhibition of damping due to the convex excitation line\nshape.\nStudying damping in our system has the advantage\nof imaging which modes of the continuum are excited\nupon damping. This can serve as a spectroscopic tool\nto probe the QP spectrum of the damping products [19].\nOne can also differentiate between Beliaev damping and\nLandau damping according to the resulting momentum\ndistribution of damping products.\nThis work was supported by the DIP and Minerva\nfoundations.\n∗Electronic address: eitan.rowen@weizmann.ac.il\n[1] V. F. Weisskopf and E. Wigner, Z. Phys. 63, 54 (1930).\n[2] A. G. Kofman and G. Kurizki, Nature (London) 405,\n546 (2000).\n[3] E. A. Burt et al., Phys. Rev. Lett. 79, 337 (1997).\n[4] S. T. Beliaev, Sov. Phys. JETP 34, 299 (1958).\n[5] L. Pitaevskii and S. Stringari, Phys. Lett. A. 235, 398\n(1997).\n[6] E. Hodby et al., Phys. Rev. Lett. 86, 2196 (2001).\n[7] V. Bretin et al., Phys. Rev. Lett. 90, 100403 (2003).5\n[8] P. O. Fedichev et al., Phys. Rev. Lett. 80, 2269 (1998).\n[9] B. Jackson and E. Zaremba, New J. Phys. 5, 88 (2003).\n[10] M. Guilleumas and L. P. Pitaevskii, Phys. Rev. A 67,\n053607 (2003).\n[11] S. Giorgini, Phys. Rev. A 57, 2949 (1998).\n[12] N. Katz et al., Phys. Rev. Lett. 89, 220401 (2002).\n[13] C. Tozzo and F. Dalfovo, New J. Phys. 5, 54 (2003).\n[14] S. Stringari, Phys. Rev. A 58, 2385 (1998).[15] N. Katz et al., Phys. Rev. A 70, 033615 (2004).\n[16] F. Zambelli et al., Phys. Rev. A 61, 063608 (2000).\n[17] The discrepancy between the filled and empty circles is\ndue to destructive interference in damping events involv-\ning damping products with non-negligible v(r).\n[18] J. Steinhauer et al., Phys. Rev. Lett. 90, 060404 (2003).\n[19] N. Katz et al., Phys. Rev. Lett. 95, 220403 (2005)." }, { "title": "1004.4410v1.Asymptotic_behavior_of_Lorentz_violation_on_orbifolds.pdf", "content": "arXiv:1004.4410v1 [hep-ph] 26 Apr 2010OU-HET 665/2010\nAsymptotic behavior of Lorentz violation on orbifolds\nNobuhiro Uekusa\nDepartment of Physics, Osaka University\nToyonaka, Osaka 560-0043 Japan\nE-mail: uekusa@het.phys.sci.osaka-u.ac.jp\nAbstract\nMomentum dependenceof quantumcorrections with higher-di mensional Lorentz\nviolation is examined in electrodynamics on orbifolds. It i s shown that effects of the\nLorentz violation are not decoupled at high energy scales. D espite the loss of the\nhigher-dimensional Lorentz invariance, a higher-dimensi onal Ward identity is found\nto be fulfilled for one-loop vacuum polarization. This impli es that gauge invariance\nmay be prior to Lorentz invariance as a guiding principle in h igher-dimensional field\ntheory. As a universal application of electrodynamics, an e xtra-dimensional aspect\nfor Furry’s theorem is emphasized.1 Introduction\nField theory with extra dimensions provides an interesting framewor k for physics be-\nyond the standard model. As in the four-dimensional case, one of t he fundamental keys\nthat characterize theory is symmetry which is preserved or broke n. In models with ex-\ntra dimensions, a variety of symmetry breaking have been provided [1]-[11]. It has also\nbeen shown that combinations of sources for extra-dimensional s ymmetry breaking are\nrelatively accommodating and yield various possibilities [12]-[14]. Associa ted with non-\nrenormalizable properties, it is still controversial whether quantu m corrections are validly\nextracted in the field-theoretical context. Higher-dimensional fi eld theory can be regarded\nasahigh-energyeffectivetheorywithadistinctultravioletcompletio n. Whileattemptsfor\nrealistic models have been developed, most of models such as orbifold models with a min-\nimal setup require higher-dimensional Lorentz invariance as a basic symmetry. However,\nextra dimensions are clearly different than our four dimensions. It in cludes potentially an\nextra-dimensional Lorentz violation.\nIf the Lorentz invariance in extra dimensions is violated, it is importan t to be taken\ninto account whether the symmetry breaking is spontaneous or no t. When a symmetry\nbreaking is described to be spontaneous in a certain theory, the co rresponding symmetry\nis expected to be recovered at high energies in its framework. The H iggs mechanism is\nthis type of symmetry breaking. Symmetry breaking in orbifolding inv olves the extra-\ndimensional origin. It is nontrivial whether symmetry is recovered a t high energy scales.\nEven if the starting action is Lorentz invariant, loop effects can give rise to a Lorentz vio-\nlation. If a model in the standpoint of effective field theory beyond t he standard model al-\nlows that the Lorentz invariance is lost at high energy scales, the st arting action should be\ndescribed in a Lorentz-non-invariant manner or only approximately in a Lorentz-invariant\nmanner with respect to extra dimensions. The extra-dimensional L orentz violation has\nbeen found to affect spectra, Kaluza-Klein parity and parity violatio n [15]. Therefore in\nthe field-theoretical context it should be clarified if the extra-dime nsional Lorentz invari-\nance on orbifold models is asymptotically preserved.\nIn this paper, we study momentum dependence of Lorentz violating terms in elec-\ntrodynamics on an orbifold S1/Z2. With an explicit analysis for loop diagrams and\nrenormalization, it is shown that effects of the Lorentz violation are not decoupled at\nhigh energy scales. As another notable feature, despite the loss o f the higher-dimensional\nLorentz invariance, a higher-dimensional Ward identity is found to b e fulfilled for one-\nloop vacuum polarization. This implies that higher-dimensional gauge in variance may\nbe prior to higher-dimensional Lorentz invariance as a guiding princip le in a high-energy\nfield theory. We also discuss an extra-dimensional aspect for Furr y’s theorem.\nThe paper is organized as follows. In Sec. 2, our Lorentz violent act ion is given. In\nSec. 3, a formalism of a renormalization is shown in the orbifold model. I n Sec. 4, the\nasymptotic energy dependence of Lorentz violating terms is given. It is also shown that\nhigher-dimensional Ward identity is fulfilled for one-loop vacuum polar ization. In Sec. 5,\na discussion about Furry’s theorem is given. In Sec. 6, we conclude w ith some remark.\nThe detail of loop corrections is summarized in Appendix A.\n12 Five-dimensional electrodynamics and Lorentz vi-\nolation\nWe start with the action for five-dimensional quantum electrodyna mics,\nS=SLI+SLV+SGF, (2.1)\nwith the Lorentz invariant action,\nSLI=/integraldisplay\nd4x·1\n2/integraldisplayL\n−Ldy/parenleftbigg\n−1\n4FMNFMN+¯ψiγMDMψ/parenrightbigg\n, (2.2)\nand the Lorentz violating action\nSLV=/integraldisplay\nd4x·1\n2/integraldisplayL\n−Ldy/parenleftbigg\n−λ\n2FµyFµy+k¯ψiγ5Dyψ/parenrightbigg\n, (2.3)\nwhereλandkare dimensionless coupling constants and their nonzero values indica te\nthe violation of the five-dimensional Lorentz invariance. After a re normalization, both of\nλandkare momentum-dependent. The Lorentz violating terms such as ¯ψγ5ψcan be\nabsorbed by the terms in Eq. (2.3) via a field redefinition [15]. The actio ns (2.2) and (2.3)\nhave gauge invariance although its form is not in a Lorentz-invariant way. The gauge\nfixing action is denoted as SGF, whose explicit form will be given after a field redefinition\nwith respect to renormalization factors. The fifth-dimensional Lo rentz violation is only\ntaken into account while the four-dimensional Lorentz invariance is preserved. The five-\ndimensional indices are denoted as M. Greek indices µrun over 0,1,2,3 and the fifth index\nis denoted as y. The gamma matrices are given by\nγµ=/parenleftbigg\nσµ\n¯σµ/parenrightbigg\n, γ5=/parenleftbigg\n−i12\ni12/parenrightbigg\n, (2.4)\nwhere the Pauli sigma matrices are used as σµ= (12,σi) and ¯σµ= (−12,σi). The five-\ndimensional covariant derivative is defined as DM=∂M−igAM. The extra-dimensional\nspace is compactified on S1/Z2, where the fundamental region is 0 ≤y≤L. The five-\ndimensional spacetime is flat with the metric (1 ,−1,−1,−1,−1). The orbifold boundary\nconditions for gauge fields and fermions are\nAµ(x,−y) =Aµ(x,y), Aµ(x,L−y) =Aµ(x,L+y), (2.5)\nAy(x,−y) =−Ay(x,y), Ay(x,L−y) =−Ay(x,L+y), (2.6)\nψ(x,−y) =iγ5ψ(x,y), ψ(x,L−y) =iγ5ψ(x,L+y), (2.7)\nsuch that the photon and left-handed Weyl fermion have zero mod e.\nIn order to perform renormalized perturbation, we define renorm alized fields as\nAµ=Z1/2\nAAµr, Ay=Z1/2\n5Ayr, ψ=Z1/2\nψψr. (2.8)\nThe Lagrangian terms for the gauge field are rewritten as\n−1\n4FMNFMN−λ\n2FµyFµy\n=−1\n4FMNrFMN\nr−λr\n2FµyrFµy\nr\n−1\n4δ1FµνrFµν\nr+1\n2δ2∂µAyr∂µAyr−δ4∂µAyr∂yAµ\nr+1\n2δ3∂yAµr∂yAµ\nr,(2.9)\n2whereλris the renormalized coupling for λ. Among the counterterms in the equation\n(2.9), the cross term ∂µAyr∂yAµ\nralso appears. The renormalization factors are given by\nδ1=ZA−1, δ 2= (1+λ)Z5−(1+λr), (2.10)\nδ3= (1+λ)ZA−(1+λr), δ 4= (1+λ)Z1/2\n5Z1/2\nA−(1+λr).(2.11)\nThe part of the gauge field has the original three coefficients λ,ZAandZ5. One of the\nfour renormalization factors δ1,···,δ4can be written in terms of the other factors. For\nexample,δ4is\nδ4= (1+λ)(1+δ1)/braceleftBigg/bracketleftbiggδ2−δ3\n(1+λ)(1+δ1)+1/bracketrightbigg1/2\n−1/bracerightBigg\n+δ3. (2.12)\nThe equation (2.9) has gauge invariance although it is not the five-dim ensional Lorentz\ninvariant form. It is convenient to choose the gauge fixing action as\nSGF=/integraldisplay\nd4x·1\n2/integraldisplayL\n−Ldy/parenleftbigg\n−1\n2ξ(∂µAµ\nr−ξ(1+λr)∂yAyr)2/parenrightbigg\n. (2.13)\nFor the gauge ξ= 1, the kinetic term and λrterm in Eq. (2.9) and the gauge fixing yield\n−1\n2[∂µAνr∂µAν\nr−(1+λr)∂yAµr∂yAµ\nr]+1\n2/bracketleftBig\n∂µ˜Ayr∂µ˜Ayr−(1+λr)∂y˜Ayr∂y˜Ayr/bracketrightBig\n,(2.14)\nwhere the rescaling has been employed as ˜Ayr≡√1+λrAyrfor the canonical normaliza-\ntion. Unless 1+ λ>0, tachyonic degrees arise. At the moment its positivity is assumed.\nThe cross terms of AµandAyare gathered into a total derivative −(1+λr)∂y(Aµ\nr∂µAyr),\nwhich is vanishing due to periodicity. From Eqs. (2.1) and (2.8), the La grangian terms\nfor the fermion are rewritten as\n¯ψriγM∂Mψr+kr¯ψriγ5∂yψr+δ5¯ψriγµ∂µψr+δ6¯ψriγ5∂yψr, (2.15)\nwherekris the renormalized coupling for k. Correspondingly to the two coefficients kand\nZψ, the renormalization factors are given by δ5=Zψ−1 andδ6= (1+k)Zψ−(1+kr).\nThe Lagrangian terms of interactions are rewritten as\ngrAµr¯ψrγµψr+iNrgr˜Ayr¯ψrγ5ψr+δ7Aµr¯ψrγµψr+δ8˜Ayr¯ψrγ5ψr, (2.16)\nwith the rescaled field ˜AyrforAyr. HereNr≡ −i(1+kr)/√1+λr. The renormalization\nfactorsareδ7=gZ1/2\nAZψ−grandδ8=δ8(δ1,···,δ7,kr,λr,gr). Forcouplingsandfields, the\nsubscriptrand tilde to indicate renormalized and rescaled quantities will be suppr essed\nhereafter.\nIn order to calculate quantum loop corrections, we write down the f our-dimensional\nLagrangian based on a mode expansion. From the equations of motio n, the mode expan-\nsion of fields is given by\nAµ(x,y) =1√\nLAµ0(x)+∞/summationdisplay\nn=1/radicalbigg\n2\nLAµn(x)cos/parenleftBignπ\nLy/parenrightBig\n, (2.17)\n3Ay(x,y) =∞/summationdisplay\nn=1/radicalbigg\n2\nLAyn(x)sin/parenleftBignπ\nLy/parenrightBig\n, (2.18)\nψL(x,y) =1√\nLψL0(x)+∞/summationdisplay\nn=1/radicalbigg\n2\nLψLn(x)cos/parenleftBignπ\nLy/parenrightBig\n, (2.19)\nψR(x,y) =∞/summationdisplay\nn=1/radicalbigg\n2\nLψRn(x)sin/parenleftBignπ\nLy/parenrightBig\n. (2.20)\nAfter the integration of the fifth space, the four-dimensional La grangian is obtained as\nL4D=Lquad\nAµ+Lquad\nAy+Lquad\ncross+Lquad\nψ+Lint. (2.21)\nHere the quadratic Lagrangians are given by\nLquad\nAµ=−1\n2∂µAν0∂µAν\n0−1\n2∞/summationdisplay\nn=1/parenleftbig\n∂µAνn∂µAν\nn−m2\nAnAµnAµ\nn/parenrightbig\n−1\n4δ1Fµν0Fµν\n0−1\n2∞/summationdisplay\nn=1/parenleftbigg1\n2δ1FµνnFµν\nn−δ3\n(1+λ)m2\nAnAµnAµ\nn/parenrightbigg\n,(2.22)\nLquad\nAy=1\n2∞/summationdisplay\nn=1/parenleftbig\n∂µAyn∂µAyn−m2\nAnAynAyn/parenrightbig\n+δ2\n2(1+λ)∞/summationdisplay\nn=1∂µAyn∂µAyn,(2.23)\nLquad\ncross=δ4\n(1+λ)∞/summationdisplay\nn=1mAn(∂µAyn)Aµ\nn, (2.24)\nLquad\nψ=¯ψ0iγµPL∂µψ0+∞/summationdisplay\nn=1¯ψn(iγµ∂µ−mψn)ψn\n+δ5¯ψ0iγµPL∂µψ0+∞/summationdisplay\nn=1¯ψn/parenleftbigg\nδ5iγµ∂µ−δ6\n(1+k)mψn/parenrightbigg\nψn. (2.25)\nThe Lagrangian Lquad\nAµforAµhas counterterms for δ1andδ3. The Lagrangian Lquad\nAy\nforAyhas a counterterm for δ2. For the Lagrangian Lquad\ncross, there is a cross term only\nfor the counterterm. The renormalization factor δ4is not independent of δ1,δ2,δ3. The\nLagrangian Lquad\nψforψhas counterterms for δ5andδ6. Then-th masses of bosons and\nfermion are\nmAµn=√\n1+λnπ\nL=mAyn≡mAn, mψn= (1+k)nπ\nL. (2.26)\nWe have defined Dirac fermions as\nψ0≡/parenleftbiggψL0\n0/parenrightbigg\n, ψn≡/parenleftbiggψLn\nψRn/parenrightbigg\n, (2.27)\nand introduced the left-chiral projection operator PL≡(12+iγ5)/2. The interaction\nterms of the Lagrangian are\nLint=g√\nL¯ψ0γµPLAµ0ψ0+∞/summationdisplay\nn=1g√\nL¯ψnγµAµ0ψn\n4+∞/summationdisplay\nn=1g√\nL/parenleftbig¯ψnγµPLAµnψ0+¯ψ0γµPLAµnψn/parenrightbig\n+∞/summationdisplay\nn,m,ℓ=1g√\n2L/braceleftbig¯ψnγµAµmψℓ(δn+m,ℓ+δn,m+ℓ)+¯ψnγµiγ5Aµmψℓδn+ℓ,m/bracerightbig\n+∞/summationdisplay\nn=1Ng√\nL/parenleftbig¯ψnPLAynψ0−¯ψ0PRAynψn/parenrightbig\n+∞/summationdisplay\nn,m,ℓ=1Ng√\n2L/braceleftbig¯ψnAymψℓ(δn,m+ℓ−δn+m,ℓ)+¯ψniγ5Aymψℓδn+ℓ,m/bracerightbig\n,(2.28)\nwhere counterterms for interactions have been omitted. The sum of modes for three\nindices is denoted as/summationtext∞\nn,m,ℓ=1=/summationtext∞\nn=1/summationtext∞\nm=1/summationtext∞\nℓ=1. At tree level, λandkaffect the\nKaluza-Klein spectrum given in Eq. (2.26). The equation (2.23) means thatAynhas no\ncounterterm for the mass. As an explicit consistency check, it will b e shown that the one-\nloop two-point function for Aynhas the bulk divergence only for a four-momentum term.\nIn Eq. (2.28), the terms ¯ψnPLAynψ0and¯ψ0PRAynψnhave relative sign and ¯ψnAymψℓhas\nthe factor ( δn,m+ℓ−δn+m,ℓ). The importance of their signs will be emphasized in Sec. 5.\n3 Renormalization on orbifolds\nIn this section, we give a formalism of the renormalization for two-po int functions for Aµ\nandAy. The one-loop vacuum polarizations for AµandAyare diagonal with respect to\nKaluza-Klein modes. The detail of a calculation is summarized in Append ix A.\nThe tree level propagators for the s-th fieldsAµsandAysare\nDµν(p2) =−iηµν\np2−m2\nAs+iǫ, D 55(p2) =i\np2−m2\nAs+iǫ, (3.1)\nwherep2=pµpµ. For simplicity, iǫwill be omitted hereafter. Exact propagators can\nbe decomposed with one-particle irreducible amplitudes. At one-loop level, diagrams of\nthe decomposition are shown in Figure 1, where an unshaded circle de notes a one-loop\ndiagram. The corresponding equations are written as\nGµν=Dµν+DµρΠρσGσν+DµρΠρ5G5ν, (3.2)\nG55=D55+D55Π55G55+D55Π5σGσ5, (3.3)\nG5ν=D55Π5σGσν+D55Π55G5ν, (3.4)\nGµ5=DµρΠρσGσ5+DµρΠρ5G55. (3.5)\nThe one-loop vacuum polarizations have the tensor structure give n by\nΠµν= Π1ηµν+Π2pµpν,Πµ5=−Π5\nµ= Π3pµ= Π5µ, (3.6)\nwhere the explicit forms of Π 1, Π2and Π 3will be given later. With these quantities, the\none-loop exact propagators are solved as\nGµν=−D55\n1+D55Π1ηµν\n5Figure 1: One-loop decomposition of exact propagators.\n+D55[(D55Π2)(1−D55Π55)+(D55Π3)2]pµpν\n(1+D55Π1)[(1−D55Π55)(1+D55(Π1+Π2p2))+(D55Π3)2p2],(3.7)\nG55=D55(1+D55(Π1+Π2p2))\n(1−D55Π55)(1+D55(Π1+Π2p2))+(D55Π3)2p2, (3.8)\nG5ν=D55(D55Π3)\n(1−D55Π55)(1+D55(Π1+Π2p2))+(D55Π3)2p2pν, (3.9)\nwhereGν5=G5ν.\nNow we perform the renormalization. From the Lagrangians (2.22), (2.23) and (2.24),\nthe contributions of counterterms are led to\nΠct\nµν(p) =−i(p2ηµν−pµpν)δ1+iδ3\n1+λm2\nAsηµν, (3.10)\nΠct\nµ5(p) =δ4\n1+λmAspµ,Πct\n55(p) =iδ2\n1+λp2. (3.11)\nOnly three renormalization factors among δ1,···,δ4are independent. All the divergence\nassociated with Π 1,Π2,Π3,Π55must be removed with three renormalization factors. As\nthe first step, it is convenient to fix the renormalization condition fo r the off-diagonal\ncomponent, G5ν(m2\nAs) = 0. This condition yield\nΠ3(m2\nAs) = 0, (3.12)\nwhich corresponds to the fixing of δ4. For Π 3= 0, the other propagators are simplified as\nGµν=−D55\n1+D55(Π1+Π2p2)/bracketleftbigg\nηµν+D55Π2\n1+D55Π1(p2ηµν−pµpν)/bracketrightbigg\n,(3.13)\nG55=D55\n1−D55Π55. (3.14)\nFor Eq. (3.13), Gµν, the term of ( p2ηµν−pµpν) is renormalized with the counterterm for\nδ1. The corresponding renormalization condition can be imposed as\nΠ2(m2\nAs) = 0. (3.15)\n6Aswewillshowexplicitly, thedivergentpartforΠ 1andΠ55satisfy((Π 1+p2Π2)/m2\nAs)div=\n(Π55/p2)divat one-loop level. This reduces to δ2=δ3. Thus the renormalization can be\nchosen as\nΠ1(m2\nAs) = 0. (3.16)\nOn the other hand, the finite part is ((Π 1+p2Π2)/m2\nAs)/ne}ationslash= (Π55/p2). This means that the\npropagator for Ayreceives finite mass corrections with Π55(m2\nAs)/ne}ationslash= 0. For the divergent\npart, it will be found in the following sections that at one-loop level, δ2=δ3=δ4=\n(1+k)2δ1. Thus the momentum-dependent vacuum polarizations Π 1(p2), Π2(p2), Π3(p2)\nand Π55(p2) can be achieved after the divergent part is fixed with the renorma lization\nconditions (3.12), (3.15) and (3.16). Fromthese equations, we can identify the asymptotic\nbehavior of the Π 1(p2), Π2(p2), Π3(p2) and Π55(p2). It needs to be checked if Lorentz\ninvariance is preserved at high energy scales.\nRenormalization for fermion self-energies would be given in a similar pro cedure. It\nmay be technically complicated since one-loop self-energies are not d iagonal with respect\nto Kaluza-Klein modes. This can be found from explicit one-loop amplitu des summarized\nin Appendix A. A feasible way to treat off-diagonal components has b een developed in\nRef. [16]. At the first step to address asymptotic behavior of the L orentz violation, we are\ninterest innotonlyLorentzinvariancebutalsogaugeinvariance. Bo thoftheseinvariances\ncanbesimultaneously examined whenthevacuumpolarizationrather thantheself-energy\nis analyzed. Therefore we focus on the effects on the vacuum polar ization forAµandAy\nand the issue for determining momentum-dependent amplitudes with external fermions\nwill be left for future work.\n4 Energy dependence of Lorentz violating terms and\nhigher-dimensional Ward identity\nFollowing the formalism of the previous section, we analyze explicit one -loop results for\nthe Lorentz violation. The one-loop contributions for the vacuum p olarization, Π(1)\n1, Π(1)\n2,\nΠ(1)\n3and Π55(1)are given via the dimensional regularization by\nΠ(1)\n1(p2) =8ig2\n(4π)2(1+k)/integraldisplay1\n0dx\n\n\nz4−∞/summationdisplay\nnp=1z3e−2z4\nz3·cos(2πnpxs)\n\n×x(1−x)(p2−m2\nψs)\n−1\n4∞/summationdisplay\nnp=1z3(z3+2z4)e−2z4\nz3(1−2x)mψssin(2πnpxs)\n\n, (4.1)\nΠ(1)\n2(p2) =−8ig2\n(4π)2(1+k)/integraldisplay1\n0dx\n\n\nz4−∞/summationdisplay\nnp=1z3e−2z4\nz3·cos(2πnpxs)\n\n×x(1−x)}, (4.2)\nΠ(1)\n3(p2) =−8ig2N\n(4π)2(1+k)/integraldisplay1\n0dx\n\n\nz4−∞/summationdisplay\nnp=1z3e−2z4\nz3cos(2πnpxs)\n\n7×x(1−x)mψs\n+1\n4∞/summationdisplay\nnp=1z3(z3+2z4)e−2z4\nz3(1−2x)sin(2πnpxs)\n\n, (4.3)\nΠ55(1)(p2) =8ig2N2\n(4π)2(1+k)/integraldisplay1\n0dx/braceleftbig\nz4x(1−x)p2\n−1\n4∞/summationdisplay\nnp=1/bracketleftbig\n3z2\n3(z3+2z4)+2z3(2x(1−x)m2\nψs)/bracketrightbig\ne−2z4\nz3cos(2πnpxs)\n+1\n4∞/summationdisplay\nnp=1z3(z3+2z4)e−2z4\nz3(1−2x)mψssin(2πnpxs)\n\n, (4.4)\nwherez3≡(1 +k)/(npL) andz4≡/radicalBig\nx(1−x)(m2\nψs−p2). In the above equations, the\nnp-independent part is finite due to the dimensional regularization in sp acetime with\nodd dimensions but it is potentially divergent. For the np-independent part, (Π(1)\n1+\np2Π(1)\n2)/m2\nAs= Π55(1)/p2is satisfied.\nFrom the renormalization conditions (3.12), (3.15) and (3.16), the r enormalization\nfactorsδ1,···,δ4are fixed. Then the renormalized vacuum polarizations are given by\nΠj(p2) = Π(1)\nj(p2)−Π(1)\nj(m2\nAs), (4.5)\nΠ55(p2) = Π55(1)(p2)\n−ip2\n1−iΠ(1)\n1(m2\nAs)\nm2\nAs−1\n1+iΠ(1)\n2(m2\nAs)/parenleftBigg\n1−Π(1)\n3(m2\nAs)−iΠ(1)\n1(m2\nAs)\nm2\nAs/parenrightBigg2\n,(4.6)\nwherej= 1,2,3. At high energies, the vacuum polarizations behave as\nΠµν(p2)→Πµν\nas(p2) =/bracketleftbig\n−(p2ηµν−pµpν)−N2m2\nAsηµν/bracketrightbig\nΠas\n2(p2), (4.7)\nΠµ5(p2)→Πµ5\nas(p2) =−iN2(pµmAs)Πas\n2(p2), (4.8)\nΠ55(p2)→Π55\nas(p2) =−N2p2Πas\n2(p2), (4.9)\nwhere Πas\n2(p2)≡ −8ig2(4π)−2(1 +k)−1/integraltext1\n0dxz4x(1−x). In obtaining the asymptotic\nvalues (4.7), (4.8) and (4.9), we have employed the renormalization f actorsZAandZ5\nand the renormalized coupling constant λso as to satisfy the renormalization conditions.\nExplicitly these constants are given by\nZA= 1+iΠ(1)\n2(m2\nAs), (4.10)\nZ5=1\n1+iΠ(1)\n2(m2\nAs)/parenleftBigg\n1−Π(1)\n3(m2\nAs)−iΠ(1)\n1(m2\nAs)\nm2\nAs/parenrightBigg2\n, (4.11)\nλr=λ−(1+λ)iΠ(1)\n1(m2\nAs)\nm2\nAs. (4.12)\nwhere the subscript rindicates a renormalized quantity again to avoid confusion. The\nequations (4.8) and (4.9) include Nrin whichλrobeys Eq. (4.12) and is generally nonva-\nnishing. Thus the extra-dimensional Lorentz invariance is violated in a generic region in\nthe parameter space at high energy scales. Especially λ= 0 does not mean λr= 0.\n8To identify the effect of the violation of translation invariance due to the brane, we\nconsider the limit L→ ∞. For this limit, the factor z3approaches zero, z3→0 asL−1\nso that the vacuum polarization become Eqs. (4.7), (4.8) and (4.9). Thenλris given in\nEq. (4.12). In the representation (4.12), the limit yields Π(1)\n1(m2\nAs)→0 asL−3. Thus the\ncouping constant for L→ ∞isλr→λ. Therefore the infinite compactification radius\nand zero original λcan recover the higher-dimensional Lorentz invariance.\nNow we move on to the issue of Ward identity. We compare the Lorent z violating case\nwith a simple extension of the four-dimensional quantum electrodyn amics. In a simple\nextension, the vacuum polarization has the form ΠMN\n5D= (pMpN−pLpLηMN)Π. This is\ndecomposed as\nΠµν\n5D=/bracketleftbig\n−/parenleftbig\np2ηµν−pµpν/parenrightbig\n+p2\n5ηµν/bracketrightbig\nΠ,Πµ5\n5D= (pµp5)Π,Π55\n5D=p2Π.(4.13)\nwhich satisfy the identities,\npµΠµν\n5D+p5Π5ν\n5D= 0, pµΠµ5\n5D+p5Π55\n5D= 0. (4.14)\nOn the other hand, the asymptotic vacuum polarization given in Eq. ( 4.7), (4.8) and (4.9)\nhave the relation\npµΠµν\nas+imAsΠ5ν\nas= 0, pµΠµ5\nas−imAsΠ55\nas= 0. (4.15)\nFrom the correspondence Πµν\nas↔Πµν\n5D, Πµ5\nas↔iΠµ5\n5Dand Π55\nas↔Π55\n5D, we find that the\none-loop vacuum polarizations satisfy the five-dimensional Ward ide ntity even without\npreserving the five-dimensional Lorentz invariance.\n5 Furry’s theorem on orbifolds\nSo far we have examined the properties of the vacuum polarizations with an explicit\ndiagrammatic calculation. In this section, we give a formal aspect in h igher-dimensional\ngauge theory.\nIn the four-dimensional electrodynamics, the charge conjugatio n is a symmetry of the\ntheory,C|Ω/an}bracketri}ht=|Ω/an}bracketri}ht, whereCdenotes the charge conjugation and |Ω/an}bracketri}htis the vacuum state.\nThe electromagnetic current, jµ=¯ψγµψchanges sign under the charge conjugation,\nCjµ(x)C†=−jµ(x) so that its vacuum expectation value is vanishing, /an}bracketle{tΩ|Tjµ(x)|Ω/an}bracketri}ht= 0.\nFurry’s theorem states that any vacuum vacuum expectation valu e of an odd number of\nelectromagnetic currents is vanishing.\nNow we consider a two-current function Mµ\nY≡ /an}bracketle{tTjµ(x1)jY(x2)/an}bracketri}htby introducing an-\nother operator jY=¯ψψand by imaging gauge and Yukawa interactions for external lines.\nHere the ground state of the free theory with the symmetry of th e charge conjugation is\ndenoted as /an}bracketri}ht. Because of the charge conjugation CjY(x)C†= +jY(x), the two-current\nfunction Mµ\nYis vanishing. At the first sight, the function Mµ\nYwith gauge and Yukawa\ninteractions seems to look like the vacuum polarizations Πµ5and Πµ5\n5D. On the other hand,\nthe vacuum polarization Πµ5\n5Dis not vanishing for nonzero Πµν\n5Das seen from Eq. (4.13).\nWe have also explicitly derived a nonzero Πµ5. Thus the structure of Πµ5needs to be\nclarified from the viewpoint of Furry’s theorem.\n9The one-loop two-point function for Aµj(x) andAys(w) is given by\nNg2\n2Lδjs/integraldisplay\nd4x1d4x2Dj,µρ(x−x1)Ds(w−x2)/an}bracketle{tTOρ\nY(x1,x2)/an}bracketri}ht, (5.1)\nwith the two-current operator\nOρ\nY(x1,x2)≡jρ\ns,0(x1)j0,s(x2)−jρ\n0,s(x1)js,0(x2)\n−jρ\nn,ℓ(x1)jℓ,n(x2)(δn+s,ℓ−δn,s+ℓ)−jρ5\nn,ℓ(x1)j5\nℓ,n(x2)δn+ℓ,s.(5.2)\nHere the currents with zero mode are given by jρ\nI,J=¯ψIγρψJandjI,J=¯ψIψJ, where\nI,J= 0,1,···,∞and the currents with γ5are given by jρ5\nn,ℓ=¯ψnγρiγ5ψℓandj5\nℓ,n=\n¯ψℓiγ5ψn. The charge conjugation yields a change of the overall factor and the interchange\nof indices,\nCjρ\nI,J(x)C†=−jρ\nJ,I(x), Cj I,J(x)C†= +jJ,I(x), (5.3)\nCjρ5\nn,ℓ(x)C†= +jρ5\nℓ,n(x), Cj5\nn,ℓ(x)C†= +jρ5\nℓ,n(x). (5.4)\nFrom these equations, we obtain COρ\nY(x1,x2)C†= +Oρ\nY(x1,x2). Therefore that Πµ5is\nnot necessarily zero is consistent with Furry’s theorem. In Eq. (2.2 8),¯ψnPLAynψ0and\n¯ψ0PRAynψnhave relative sign. If they have the same sign, the contribution fro m the\nfirst line in Eq. (5.2) would vanish. The role of the relative sign in the ter m¯ψnAymψℓin\nEq. (2.28) is similar.\nApplication of Furry’s theorem in orbifold models may be given not only f or two-\npoint functions but also for other functions. For example, the vac uum expectation\nvalue of one current is vanishing, /an}bracketle{tTjµ\nn,n(x)/an}bracketri}ht= 0, where the indices are the identical\nn. Because a nonzero Πµ5is expected from a nonzero ΠMN\n5D, Furry’s theorem in effec-\ntive four-dimensional theory may be related to the discrete symme try in the original\nhigher-dimensional theory. Due to the dependence of Lorentz tr ansformation on the di-\nmensionality of spacetime, it is nontrivial to introduce discrete symm etry such as P,C,T\nin higher-dimensional theory [17, 18]. We leave further exploration o f this issue for future\nwork.\n6 Conclusion\nWe have studied the momentum dependence of Lorentz violating ter ms in the field-\ntheoreticalcontext inelectrodynamicsonorbifolds. Hereanexplic it analysishasbeenper-\nformed for loop diagrams and renormalization. We have found that t he extra-dimensional\nLorentz invariance is violated in a generic region in the parameter spa ce at high energy\nscales. In particular, even if the original action is higher-dimensiona l Lorentz invariant,\nit is violated by loop effects. While the higher-dimensional Lorentz inva riance is lost,\na higher-dimensional Ward identity has been found to be fulfilled for t he one-loop vac-\nuum polarization. Therefore higher-dimensional gauge invariance m ay be prior to higher-\ndimensional Lorentz invariance as a guiding principle in a high-energy fi eld theory. We\nhave also discussed Furry’s theorem in orbifold models to confirm the consistency about\nthe vacuum polarizations.\n10The four-dimensional Lorentz violation has also been studied as a dis tinct topic of\nLorentz violation. In the four-dimensional electrodynamics with Lo rentz violation, it\nhas been discussed that Pauli-Villars regularization is a useful choice associated with\ngauge invariance [19, 20, 21]. On the other hand, it has been shown t hat propagators\ncorresponding to Pauli-Villars are radiatively generated in an orbifold model [22]. In\nthis light, the Pauli-Villars regulator may be the necessity of an extra -dimensional model\nrather than a choice. These relations should be examined further.\nAcknowledgments\nThis work is supported by Scientific Grants from the Ministry of Educ ation and Science,\nGrant No. 20244028.\n11A Loop corrections\nIn this appendix, the details of loop corrections are given.\nA.1 Diagrams and four-momentum integrals\nWe evaluate loop corrections by calculating the sum of diagrams for e ach Kaluza-Klein\nmode. Propagators are defined for four-dimensional fields. The t ree-level propagators are\ndiagonal with respect to Kaluza-Klein modes and are given by\nDµν(x−w) =/an}bracketle{tTAµ\n0(x)Aν\n0(w)/an}bracketri}ht=/integraldisplayd4p\n(2π)4−iηµν\np2+iǫe−ip·(x−w)(A.1)\nDµν\nn(x−w) =/an}bracketle{tTAµ\nn(x)Aν\nn(w)/an}bracketri}ht=/integraldisplayd4p\n(2π)4−iηµν\np2−m2\nAn+iǫe−ip·(x−w)(A.2)\nDn(x−w) =/an}bracketle{tTAyn(x)Ayn(w)/an}bracketri}ht/integraldisplayd4p\n(2π)4i\np2−m2\nAn+iǫe−ip·(x−w),(A.3)\nfor bosons and\nS(x−w) =/an}bracketle{tTψ0(x)¯ψ0(w)/an}bracketri}ht=/integraldisplayd4p\n(2π)4PLip /\np2+iǫe−ip·(x−w)(A.4)\nSn(x−w) =/an}bracketle{tTψn(x)¯ψn(w)/an}bracketri}ht=/integraldisplayd4p\n(2π)4i(p /+mψn)\np2−m2\nψn+iǫe−ip·(x−w),(A.5)\nfor fermions.\nThe vacuum polarizations for AµandAyinvolve the following momentum integrals:\nI1(µ,ν;m1,m2)≡ −g2\n2L/integraldisplayd4p1\n(2π)4tr/parenleftbiggp /1\np2\n1−m2\n1γµp /1+p /2\n(p1+p2)2−m2\n2γν/parenrightbigg\n,(A.6)\nI2(µ,ν;m1,m2)≡ −g2\n2L/integraldisplayd4p1\n(2π)4tr/parenleftbiggm1\np2\n1−m2\n1γµm2\n(p1+p2)2−m2\n2γν/parenrightbigg\n,(A.7)\nfor two four-indices,\nI1(µ;m1,m2)≡ −g2\n2L/integraldisplayd4p1\n(2π)4tr/parenleftbiggm1\np2\n1−m2\n1γµp /1+p /2\n(p1+p2)2−m2\n2/parenrightbigg\n,(A.8)\nI2(µ;m1,m2)≡ −g2\n2L/integraldisplayd4p1\n(2π)4tr/parenleftbiggp /1\np2\n1−m2\n1γµm2\n(p1+p2)2−m2\n2/parenrightbigg\n,(A.9)\nfor one four-index and\nI1(m1,m2)≡g2\n2L/integraldisplayd4p1\n(2π)4tr/parenleftbiggp /1\np2\n1−m2\n1p /1+p /2\n(p1+p2)2−m2\n2/parenrightbigg\n, (A.10)\nI2(m1,m2)≡g2\n2L/integraldisplayd4p1\n(2π)4tr/parenleftbiggm1\np2\n1−m2\n1m2\n(p1+p2)2−m2\n2/parenrightbigg\n, (A.11)\nfor no four-indices. These satisfy a property I1(µ;m2,m1) =−I2(µ;m1,m2).\n12The self-energies for ψinvolve the following momentum integrals:\nB1(m1,m2)≡ −g2\n2L/integraldisplayd4p2\n(2π)4γµ1\np2\n2−m2\n1p /1−p /2\n(p1−p2)2−m2\n2γµ, (A.12)\nB2(m1,m2)≡ −g2\n2L/integraldisplayd4p2\n(2π)4γµ1\np2\n2−m2\n1m2\n(p1−p2)2−m2\n2γµ, (A.13)\nE1(m1,m2)≡ −g2\n2L/integraldisplayd4p2\n(2π)41\np2\n2−m2\n1p /1−p /2\n(p1−p2)2−m2\n2, (A.14)\nE2(m1,m2)≡ −g2\n2L/integraldisplayd4p2\n(2π)41\np2\n2−m2\n1m2\n(p1−p2)2−m2\n2. (A.15)\nWith these integral expressions, the vacuum polarizations are sum marized as follows:\n0 0\n=∞/summationdisplay\nn=−∞{I1(µ,ν;mψn,mψn)+I2(µ,ν;mψn,mψn)}, (A.16)\nj s\n=∞/summationdisplay\nn=−∞{I1(µ,ν;mψn,mψ,n+s)+I2(µ,ν;mψn,mψ,n+s)}δjs,(A.17)\nj s\n=N∞/summationdisplay\nn=−∞{I1(µ;mψn,mψ,n+j)+I2(µ;mψn,mψ,n+j)}δjs,(A.18)\nj s\n=N2∞/summationdisplay\nn=−∞{I1(mψn,mψ,n+j)+I2(mψn,mψ,n+j)}δjs.(A.19)\nThe vacuum polarizations for AµandAydo not give rise to one-loop corrections for brane\nterms. The Kaluza-Klein modes for external lines are diagonal.\nThe fermion self-energies are summarized as follows:\n0 0=∞/summationdisplay\nn=−∞B1(mAn,mψn)PL+B1(0,0)PL, (A.20)\n0s=√\n2∞/summationdisplay\nn=1{B1(mAn,mψn)PL+B2(mAn,mψn)PR}δ2n,s,(A.21)\ns 0=√\n2∞/summationdisplay\nn=1{B1(mAn,mψn)+B2(mAn,mψn)}PLδ2n,s, (A.22)\nj s={B1(0,mψs)+B2(0,mψs)}δjs\n13+B1(mAj,0)iγ5δjs\n+∞/summationdisplay\nn=1{B1(mAn,mψ,j+n)+B2(mAn,mψ,j+n)}δj+2n,s\n+∞/summationdisplay\nn=1{B1(mAn,mψ,n+s)+B2(mAn,mψ,n+s)}δj,s+2n\n+∞/summationdisplay\nℓ=1{B1(mA,ℓ+s,mψℓ)+B2(mA,ℓ+s,mψℓ)}iγ5δj,s+2ℓ\n+∞/summationdisplay\nℓ=1{B1(mA,j+ℓ,mψℓ)−B2(mA,j+ℓ,mψℓ)}iγ5δj+2ℓ,s\n+∞/summationdisplay\nn=−∞{B1(mAn,mψ,j+n)+B2(mAn,mψ,j+n)}δjs,(A.23)\n0 0=N2/bracketleftBigg∞/summationdisplay\nn=−∞E1(mAn,mψn)PL−E1(0,0)PL/bracketrightBigg\n, (A.24)\n0s=−√\n2N2∞/summationdisplay\nn=1{E1(mAn,mψn)PL+E2(mAn,mψn)PR}δ2n,s,(A.25)\ns 0=−√\n2N2∞/summationdisplay\nn=1{E1(mAn,mψn)+E2(mAn,mψn)}PLδ2n,s,(A.26)\nj s=N2/bracketleftBig\niγ5E1(mAj,0)δjs\n−∞/summationdisplay\nn=1{E1(mAn,mψ,n+s)+E2(mAn,mψ,n+s)}δj,s+2n\n−∞/summationdisplay\nn=1{E1(mAn,mψ,j+n)+E2(mAn,mψ,j+n)}δj+2n,s\n+∞/summationdisplay\nℓ=1iγ5{E1(mA,j+ℓ,mψℓ)+E2(mA,j+ℓ,mψℓ)}δj+2ℓ,s\n+∞/summationdisplay\nℓ=1iγ5{E1(mA,ℓ+s,mψℓ)−E2(mA,ℓ+s,mψℓ)}δj,s+2ℓ\n+∞/summationdisplay\nn=−∞{E1(mAn,mψ,j+n)+E2(mAn,mψ,j+n)}δjs\n−{E1(0,mψj)+E2(0,mψj)}δjs/bracketrightBig\n. (A.27)\nFor (E1+E2), the mode sum with −∞ ≤n≤ ∞is regarded as a formal equation because\n14Aynhas no zero mode.\nA.2 Evaluation of momentum integrals\nWe calculate themomentum integrals by introducing Feynman parame ters andemploying\nthe dimensional regularization and the Poisson resummation.\nThe momentum integrals for Π µνare given by\n∞/summationdisplay\nn=−∞(I1(µ,ν;mψn,mψ,n+s)+I2(µ,ν;mψn,mψ,n+s))\n=8ig2\n(4π)2(1+k)/integraldisplay1\n0dx\n\n\nz4−∞/summationdisplay\nnp=1z3e−2z4\nz3·cos(2πnpxs)\n\n×x(1−x)/parenleftbig\n(p2\n2−m2\nψs)ηµν−p2µp2ν/parenrightbig\n−1\n4∞/summationdisplay\nnp=1z3(z3+2z4)e−2z4\nz3(1−2x)mψssin(2πnpxs)ηµν\n\n.(A.28)\nFors= 0, thefour-dimensionalWardidentity issatisfied. Itisalsoseenfr omthefollowing\nequation,\nI1(µ,ν;m1,m2)+I2(µ,ν;m1,m2) =−8g2\nL/integraldisplay1\n0dx/integraldisplayddℓ\n(2π)d1\n[ℓ2−∆]2\n×/bracketleftbigg\nx(1−x)(p2\n2ηµν−p2µp2ν)+1\n2(m1−m2)(xm2−(1−x)m1)ηµν/bracketrightbigg\n,(A.29)\nwhereℓ=p1+xp2and ∆ =xm2\n2+(1−x)m2\n2−x(1−x)p2\n2. In the main text, the letter of\nthe external momentum is denoted as pinstead ofp2. The momentum integrals for Π µ5\nare given by\n∞/summationdisplay\nn=−∞(I1(µ;mψn,mψ,n+s)+I2(µ;mψn,mψ,n+s))\n=−8ig2\n(4π)2(1+k)p2µ/integraldisplay1\n0dx\n\n\nz4−∞/summationdisplay\nnp=1z3e−2z4\nz3cos(2πnpxs)\nx(1−x)mψs\n+1\n4∞/summationdisplay\nnp=1z3(z3+2z4)e−2z4\nz3(1−2x)sin(2πnpxs)\n\n. (A.30)\nThe momentum integrals for Π 55are given by\n∞/summationdisplay\nn=−∞(I1(mψn,mψ,n+s)+I2(mψn,mψ,n+s))\n=8ig2\n(4π)2(1+k)/integraldisplay1\n0dx/braceleftbig\nz4x(1−x)p2\n2\n15−1\n4∞/summationdisplay\nnp=1/bracketleftbig\n3z2\n3(z3+2z4)+2z3(2x(1−x)m2\nψs)/bracketrightbig\ne−2z4\nz3cos(2πnpxs)\n+1\n4∞/summationdisplay\nnp=1z3(z3+2z4)e−2z4\nz3(1−2x)mψssin(2πnpxs)\n\n. (A.31)\nFor fermion self-energies, the momentum integrals with −∞ ≤n≤ ∞are given by\n∞/summationdisplay\nn=−∞(B1(mAn,mψ,n+s)+B2(mAn,mψ,n+s)) =−4ig2\n(4π)2/integraldisplay1\n0dx1√w\n×\n\n\nw2−∞/summationdisplay\nnp=1w1e−2w2\nw1cos/parenleftbigg\n2πnp(1+k)2xs\nw/parenrightbigg\n(1−x)/parenleftbigg\np /1−2(1+λ)mψs\nw/parenrightbigg\n+∞/summationdisplay\nnp=1(1+k)√ww1(w1+2w2)e−2w2\nw1sin/parenleftbigg\n2πnp(1+k)2xs\nw/parenrightbigg\n\n. (A.32)\nHere\nw1≡√w\nnpL, w 2≡/radicalBigg\nx(1−x)/parenleftbigg(1+λ)\nwm2\nψs−p2\n1/parenrightbigg\n, (A.33)\nw≡x(1+k)2+(1−x)(1+λ). (A.34)\nThe momentum integrals for E1andE2are obtained as with the relations\nE1(m1,m2) =−1\n2B1(m1,m2), E 2(m1,m2) =1\n4B2(m1,m2). (A.35)\n16References\n[1] J. Scherk and J. H. Schwarz, Phys. Lett. B 82, 60 (1979).\n[2] J. Scherk and J. H. Schwarz, Nucl. Phys. B 153, 61 (1979).\n[3] Y. Hosotani, Phys. Lett. B 126, 309 (1983).\n[4] Y. Hosotani, Annals Phys. 190, 233 (1989).\n[5] E. A. Mirabelli and M. E. Peskin, Phys. Rev. D 58, 065002 (1998)\n[arXiv:hep-th/9712214].\n[6] D. E. Kaplan, G. D. Kribs and M. Schmaltz, Phys. Rev. D 62, 035010 (2000)\n[arXiv:hep-ph/9911293].\n[7] Z. Chacko, M. A. Luty, A. E. Nelson and E. Ponton, JHEP 0001, 003 (2000)\n[arXiv:hep-ph/9911323].\n[8] Y. Kawamura, Prog. Theor. Phys. 103, 613 (2000) [arXiv:hep-ph/9902423].\n[9] Y. Kawamura, Prog. Theor. Phys. 105, 999 (2001) [arXiv:hep-ph/0012125].\n[10] D. Diego, G. von Gersdorff and M. Quiros, JHEP 0511, 008 (2005)\n[arXiv:hep-ph/0505244].\n[11] N. Haba and N. Uekusa, arXiv:0911.2557 [hep-ph].\n[12] N. Uekusa, Mod. Phys. Lett. A 23, 603 (2008) [arXiv:0704.2490 [hep-th]].\n[13] N. Uekusa, Int. J. Mod. Phys. A 23, 3535 (2008) [arXiv:0803.1537 [hep-ph]].\n[14] N. Uekusa, Nucl. Phys. B 812, 81 (2009) [arXiv:0806.3229 [hep-ph]].\n[15] T. G. Rizzo, JHEP 0509, 036 (2005) [arXiv:hep-ph/0506056].\n[16] N. Uekusa, arXiv:1002.1904 [hep-ph].\n[17] Y. Adachi, C. S. Lim and N. Maru, Phys. Rev. D 80, 055025 (2009) [arXiv:0905.1022\n[hep-ph]].\n[18] C. S. Lim, N. Maru and K. Nishiwaki, arXiv:0910.2314 [hep-ph].\n[19] R. Jackiw and V. A. Kostelecky, Phys. Rev. Lett. 82, 3572 (1999)\n[arXiv:hep-ph/9901358].\n[20] M. Perez-Victoria, JHEP 0104, 032 (2001) [arXiv:hep-th/0102021].\n[21] B. Altschul, Phys. Rev. D 70, 101701 (2004) [arXiv:hep-th/0407172].\n[22] N. Uekusa, Nucl. Phys. B 827, 311 (2010) [arXiv:0909.0825 [hep-ph]].\n17" }, { "title": "1501.04551v2.Relativistic_Lagrangians_for_the_Lorentz_Dirac_equation.pdf", "content": "arXiv:1501.04551v2 [physics.class-ph] 13 Jun 2015Relativistic Lagrangians for the Lorentz-Dirac equation\nShinichi Deguchia,∗, Kunihiko Nakanoa, Takafumi Suzukib\naInstitute of Quantum Science, College of Science and Techno logy, Nihon University,\nChiyoda-ku, Tokyo 101-8308, Japan\nbJunior College Funabashi Campus, Nihon University, Narash inodai, Funabashi, Chiba\n274-8501, Japan\nAbstract\nWe present two types of relativistic Lagrangians for the Lorentz- Dirac equa-\ntion written in terms of an arbitrary world-line parameter. One of th e La-\ngrangians contains an exponential damping function of the proper time and\nexplicitly depends ontheworld-line parameter. Another Lagrangian includes\nadditional cross-terms consisting of auxiliary dynamical variables a nd does\nnotdependexplicitlyontheworld-lineparameter. Wedemonstratet hatboth\nthe Lagrangians actually yield the Lorentz-Dirac equation with a sou rce-like\nterm.\n1. Introduction\nA charged particle emitting electromagnetic radiation is subjected t o the\nreaction force caused by the particle’s own electromagnetic radiat ion. This\nphenomenon is well-known as the radiation reaction [1, 2, 3, 4, 5, 6]. It was\nfirst evaluated by Lorentz at the end of the 19th century [7] and s ubsequently\nargued by Abraham and Lorentz on the basis of the charged rigid sp here\nmodel of a charged particle [8, 9]. In the zero radius limit that this mod el\ntends to become a point charge, the classical non-relativistic equa tion of\nmotion for the charged particle located at the position x=x(t) is found to\nbe\nmd2x\ndt2=F+2\n3e2d3x\ndt3, (1.1)\n∗Corresponding author.\nEmail address: deguchi@phys.cst.nihon-u.ac.jp (Shinichi Deguchi)\nPreprint submitted to Elsevier April 4, 2018wheremis the physical mass of the particle, eits electric charge, and F\ndenotes the external Lorentz force. (In this paper, we employ u nits such\nthatc= 1.) Equation (1.1) is called the Lorentz-Abraham equation (or\nthe Abraham-Lorentz equation). A relativistic extension of the Lo rentz-\nAbraham equation was derived by Dirac in a manifestly covariant mann er\nby considering energy-momentum conservation [10], and is now oft en called\nthe Lorentz-Dirac equation [3, 6, 11, 12, 13]. With the spacetime co ordinates\nxµ=xµ(l) (µ= 0,1,2,3) of a charged particle propagating in 4-dimensional\nMinkowski space, the Lorentz-Dirac equation reads\nmduµ\ndl=eFµν(x)uν+2\n3e2/parenleftbig\nδµ\nν−uµuν/parenrightbigd2uν\ndl2. (1.2)\nHere,uµ:=dxµ/dl,Fµνisthefieldstrengthtensor ofanexternal electromag-\nnetic field, and ldenotes the proper time of the particle or, in other words,\nthe arc length of the world-line traced out by the particle. The metr ic ten-\nsor of Minkowski space is assumed to be ηµν= diag(1 ,−1,−1,−1), so that\nuµuµ=ηµνuµuν= 1 holds. Equations (1.1) and (1.2) are unusual ones in-\ncluding third-order time derivatives of the particle’s position coordin ates. In\nconnectionwiththisfact, theseequationsadmitphysicallyunaccep tablesolu-\ntions such as runaway and pre-acceleration solutions [1, 2, 4, 5, 6, 12, 13, 14].\nTo overcome this problem, various ideas have been proposed until r ecently\n[1, 6, 13, 14, 15, 16, 17, 18, 19, 20, 21]; however, it seems that an ultimate\nsolution to the problem has not been found yet.\nOnce the equations of motion (1.1) and (1.2) have been obtained, it is\nquite natural to seek Lagrangians corresponding to these equat ions in or-\nder to develop the Lagrangian and Hamiltonian formulations of a char ged\nparticle subjected to the radiation reaction force. If these form ulations are\nestablished, they might lead to a novel quantum-mechanical descr iption of a\ncharged particle undergoing radiation reaction and might give us new room\nto deal with the above-mentioned problem. As far as the present a uthors\nknow, there have been a few attempts to construct Lagrangians correspond-\ning to Eqs. (1.1) and (1.2) until now [22, 23, 24]. Carati constructe d an\nexplicitly time-dependent Lagrangian for Eq (1.1) with the use of aux iliary\ndynamical variables [22]. (Carati also considered a relativistic exten sion of\nthis Lagrangian in an extremely limited case.) Barone and Mendes deriv ed\nan explicitly time-independent Lagrangian for Eq. (1.1) by incorpora ting\nthe time-reversed copy of Eq. (1.1) into the original setting [23]. Ca rati’s\nand Barone-Mendes’s approaches are, respectively, based on lea rning from\n2the direct and indirect Lagrangian formulations of the damped harm onic os-\ncillator [25, 26, 27, 28, 29, 30].1It should be pointed out here that in these\napproaches, the external Lorentz force Fis assumed to be independent of\nthe velocity v:=dx/dt. Hence, it follows that in actuality, Carati’s and\nBarone-Mendes’s Lagrangians can describe only a charged particle being in\nthe particular situation in which the magnetic field vanishes or is paralle l to\nv.2\nIn this paper, we present two types of Lagrangians for the Loren tz-Dirac\nequation (1.2) that are constructed in such a fashion that the cor responding\nactions remain invariant under reparametrization of a world-line par ameter\nalong the particle’s world-line. These Lagrangians are completely rela tivis-\ntic and admit the general form of the external Lorentz force. Als o, the\nLagrangians are outside the scope of Kupriyanov’s proof [24], beca use they\ncontain auxiliary dynamical variables in addition to xµ. One of the La-\ngrangians contains an exponential damping function of the proper timel,\nwhile another Lagrangian includes additional cross-terms consistin g of two\nauxiliary dynamical variables. Both the Lagrangians include terms sim ilar\nto what can be seen in the Lagrangian that governs a certain model of a\nrelativistic point particle with rigidity [31, 32, 33]. We would like to em-\nphasize that our Lagrangians are not immediate extensions of Cara ti’s and\nBarone-Mendes’s Lagrangians.\nThis paper is organized as follows. In section 2, we introduce necess ary\ndynamicalvariablesanddefinetheirtransformationrulesunderre parametriza-\ntion of a world-line parameter. In section 3, we present a Lagrangia n that\ncontains an exponential damping function and show that the Lagra ngian ac-\ntually yields the Lorentz-Dirac equation with a source-like term. In s ection\n4, we consider a Lagrangian including additional cross-terms, inste ad of the\n1Thedirectformulationadoptsanexplicitlytime-dependent Lagrang ianofthedamped\nharmonic oscillator [25, 26, 27], while the indirect formulation adopts a n explicitly time-\nindependent Lagrangian for a system consisting of the damped har monic oscillator and its\ntime-reversed counterpart [25, 28, 29, 30].\n2In Ref. [24], Kupriyanov investigated the possibility of constructing Lagrangians\ncorresponding to Eqs. (1.1) and (1.2) and reached the conclusion t hat there exist no\ncorresponding Lagrangians. However, Kupriyanov’s proof of this conclusion considers\nLagrangians consisting only of the coordinate variables, such as xandxµ, and their\nfirst- and second-order time derivatives. Since Carati’s and Baron e-Mendes’s Lagrangians\ncontainextradynamicalvariables,theseLagrangiansareoutside thescopeofKupriyanov’s\nproof.\n3exponential damping function, and show that this Lagrangian also y ields the\nLorentz-Dirac equation with a source-like term. Section 5 is devote d to a\nsummary and discussion. Appendix A provides the Lorentz-Dirac eq uation\nwritten in terms of an arbitrary world-line parameter instead of the proper\ntimel.\n2. Preliminaries: dynamical variables andtheir transform ation rules\nLetτ(τ0≤τ≤τ1) be an arbitrary world-line parameter along the\nparticle’s world-line, being chosen in such a manner that dx0/dτ >0. The\nspacetime coordinates of a charged particle are now denoted as xµ=xµ(τ).\nUnder the reparametrization τ→τ′=τ′(τ) (dτ′/dτ >0), the coordinate\nvariables xµbehave as scalar fields on the 1-dimensional parameter space\nT:={τ|τ0≤τ≤τ1}:\nxµ(τ)→x′µ(τ′) =xµ(τ). (2.1)\nIn addition to xµ, we introduce auxiliary dynamical variables qµ\ni=qµ\ni(τ),\nλiµ=λiµ(τ) (i= 1,2), andξµ=ξµ(τ). They are assumed to transform\nunder the reparametrization as scalar-density fields of weight 1 on T:\nqµ\ni(τ)→q′µ\ni(τ′) =dτ\ndτ′qµ\ni(τ), (2.2)\nλiµ(τ)→λ′\niµ(τ′) =dτ\ndτ′λiµ(τ), (2.3)\nξµ(τ)→ξ′\nµ(τ′) =dτ\ndτ′ξµ(τ). (2.4)\nThe components of the vector resolute of the 4-vector (˙ qµ\ni) perpendicular to\n(qµ\ni) are given by\n˙qµ\ni⊥:= ˙qµ\ni−qi˙qi\nq2\niqµ\ni, (2.5)\nwhere ˙qµ\ni:=dqµ\ni/dτ,q2\ni:=qiµqµ\ni, andqi˙qi:=qiµ˙qµ\ni(no sum with respect to\ni). It can be shown by using Eq. (2.2) that unlike ˙ qµ\ni, the components ˙ qµ\ni⊥\ntransform homogeneously as\n˙qµ\ni⊥(τ)→˙q′µ\ni⊥(τ′) =/parenleftbiggdτ\ndτ′/parenrightbigg2\n˙qµ\ni⊥(τ). (2.6)\nWe thus see that under the reparametrization, ˙ qµ\ni⊥behave as scalar-density\nfields of weight 2 on T.\n43. A Lagrangian with an exponential damping function\nNow, from the dynamical variables xµ,qµ\ni,λiµ, andξµ, we construct the\nfollowing Lagrangian:\nLD=exp(−kl)\n(q2\n1q2\n2)1/4/bracketleftbigg1\n2/parenleftbigg˙q2\n1⊥\nq2\n1−˙q2\n2⊥\nq2\n2/parenrightbigg\n−λ1µ(qµ\n1−˙xµ)+λ2µ(qµ\n2−˙xµ)+ξµ(qµ\n1−qµ\n2)−3\n2eFµν(x)qµ\n1qν\n2/bracketrightbigg\n,\n(3.1)\nwherek:= 3m/2e2, ˙xµ:=dxµ/dτ, ˙q2\ni⊥:= ˙qi⊥µ˙qµ\ni⊥, andFµν(=−Fνµ) is again\nthe field strength tensor of an external electromagnetic field. In Eq. (3.1),\nthe proper time lis a function of τrepresented as\nl(τ) =/integraldisplayτ\nτ0d˜τ/radicalBig\n˙xµ(˜τ)˙xµ(˜τ). (3.2)\nHere,xµ(˜τ) is understood as a solution of the equation of motion for xµ\nobtained later, not as a dynamical variable whose variation is taken in to\naccount in varying the action\nSD=/integraldisplayτ1\nτ0dτLD. (3.3)\nThe Lagrangian LDexplicitly depends on τvia the exponential damping\nfunction exp( −kl). Sincel(τ) is geometrically the arc length of the particle’s\nworld-line, it is certainly reparametrization invariant.3Considering this fact\nand using the transformation rules in Eqs. (2.1), (2.2), (2.3), (2.4) , and (2.6),\nwe can show that the action SDremains invariant under the reparametriza-\ntionτ→τ′. We also see that LDremains invariant under the gauge trans-\nformation\nλ1µ→λ′\n1µ=λ1µ+θµ, λ2µ→λ′\n2µ=λ2µ+θµ, ξµ→ξ′\nµ=ξµ+θµ,\n(3.4)\n3Strictly speaking, l(τ) is a functional of xµas well as a function of τandτ0. In this\nsense,l(τ) should be read as l(τ,τ0;xµ). The reparametrization invariance of l(τ) can be\nexpressed as l(τ′,τ′\n0;x′µ) =l(τ,τ0;xµ).\n5with real gauge functions θµ=θµ(τ). The Lagrangian LDhas the antisym-\nmetric property\nLD(qµ\n1,˙qµ\n1,λ1µ;qµ\n2,˙qµ\n2,λ2µ) =−LD(qµ\n2,˙qµ\n2,λ2µ;qµ\n1,˙qµ\n1,λ1µ).(3.5)\nLet us derive the Euler-Lagrange equations for the dynamical var iables\nfromLD. Noting that xµ(˜τ) contained in l(τ) and hence l(τ) itself are not\nobjectsfortakingvariation,wecaneasilyobtaintheEuler-Lagran geequation\nforxµ:\nd\ndτ/bracketleftBigg\nexp(−kl)\n(q2\n1q2\n2)1/4(λ1µ−λ2µ)/bracketrightBigg\n+3exp(−kl)\n2e(q2\n1q2\n2)1/4∂µFνρ(x)qν\n1qρ\n2= 0.(3.6)\nThisequationincludesthegauge-invariantquantity λ1µ−λ2µasareflectionof\nthe gauge invariance of LD. Hence, λ1µandλ2µthemselves are not uniquely\ndetermined. The Euler-Lagrange equation for qµ\n1can be written as\nexp(−kl)\n(q2\n1q2\n2)1/4/bracketleftbigg1\n2/parenleftbiggd\ndτ∂K1\n∂˙qµ\n1−∂K1\n∂qµ\n1/parenrightbigg\n+λ1µ−ξµ+3\n2eFµν(x)qν\n2/bracketrightbigg\n+/parenleftBigg\nd\ndτexp(−kl)\n(q2\n1q2\n2)1/4/parenrightBigg\n1\n2∂K1\n∂˙qµ\n1+q1µ\n2q2\n1LD= 0, (3.7)\nwith\nK1:=˙q2\n1⊥\nq2\n1=q2\n1˙q2\n1−(q1˙q1)2\n(q2\n1)2. (3.8)\nApplying the formulas\n1\n2∂K1\n∂˙qµ\n1=˙q1⊥µ\nq2\n1=1/radicalbig\nq2\n1d\ndτq1µ/radicalbig\nq2\n1, (3.9)\nd\ndτ∂K1\n∂˙qµ\n1−∂K1\n∂qµ\n1=2\nq2\n1/parenleftbigg\n¨q1⊥µ−2q1˙q1\nq2\n1˙q1⊥µ/parenrightbigg\n, (3.10)\nd\ndτ1\n(q2\n1q2\n2)1/4=−1\n2(q2\n1q2\n2)1/4/parenleftbiggq1˙q1\nq2\n1+q2˙q2\nq2\n2/parenrightbigg\n(3.11)\nto Eq. (3.7) appropriately, we obtain\nkdl\ndτ1/radicalbig\nq2\n1d\ndτq1µ/radicalbig\nq2\n1=3\n2eFµν(x)qν\n2+(q2\n1q2\n2)1/4q1µ\n2q2\n1exp(−kl)LD\n+¨q1⊥µ\nq2\n1−/parenleftbigg5q1˙q1\nq2\n1+q2˙q2\nq2\n2/parenrightbigg˙q1⊥µ\n2q2\n1+λ1µ−ξµ.(3.12)\n6Here, ¨q1⊥µ, together with ¨ q2⊥µ, is defined by\n¨qi⊥µ:= ¨qiµ−qi¨qi\nq2\niqiµ, (3.13)\nwhere ¨qiµ:=d2qiµ/dτ2andqi¨qi:=qiµ¨qµ\ni(nosumwithrespectto i). Following\nthe same procedure as that used for deriving Eq. (3.12), we can de rive the\nEuler-Lagrange equation for qµ\n2as\nkdl\ndτ1/radicalbig\nq2\n2d\ndτq2µ/radicalbig\nq2\n2=3\n2eFµν(x)qν\n1−(q2\n1q2\n2)1/4q2µ\n2q2\n2exp(−kl)LD\n+¨q2⊥µ\nq2\n2−/parenleftbiggq1˙q1\nq2\n1+5q2˙q2\nq2\n2/parenrightbigg˙q2⊥µ\n2q2\n2+λ2µ−ξµ.(3.14)\nThe Euler-Lagrange equations for λ1µ,λ2µ, andξµare respectively found to\nbe\nqµ\n1= ˙xµ, (3.15)\nqµ\n2= ˙xµ, (3.16)\nqµ\n1=qµ\n2. (3.17)\nEquation (3.17) can also be found from Eqs. (3.15) and (3.16).\nSubstituting Eqs. (3.15) and (3.16) into Eq. (3.12) and noting\nLD(qµ\n1,˙qµ\n1,λ1µ;qµ\n2,˙qµ\n2,λ2µ) =LD(˙xµ,¨xµ,λ1µ; ˙xµ,¨xµ,λ2µ) = 0,(3.18)\nwe have\nkdl\ndτ1√\n˙x2d\ndτ˙xµ\n√\n˙x2=3\n2eFµν(x)˙xν+...xµ\n⊥\n˙x2−3(˙x¨x)¨xµ\n⊥\n(˙x2)2+λµ\n1−ξµ,(3.19)\nwhere ¨xµ:=d2xµ/dτ2and...xµ:=d3xµ/dτ3. Similarly, substituting Eqs.\n(3.15) and (3.16) into Eq. (3.14) and using (3.18), we have\nkdl\ndτ1√\n˙x2d\ndτ˙xµ\n√\n˙x2=3\n2eFµν(x)˙xν+...xµ\n⊥\n˙x2−3(˙x¨x)¨xµ\n⊥\n(˙x2)2+λµ\n2−ξµ.(3.20)\nComparing Eq. (3.19) with Eq. (3.20) leads to\nλµ\n1=λµ\n2. (3.21)\n7This equality is covariant under the gauge transformation (3.4). It follows\nfromEq. (3.21)thatEq. (3.6)isidenticallysatisfied, because ∂µFνρ(x)qν\n1qρ\n2=\n∂µFνρ(x)˙xν˙xρ= 0 holds by using Eqs. (3.15) and (3.16). Hereafter, taking\ninto account Eq. (3.21), we simply write λµ\n1andλµ\n2asλµ. Thereby, Eqs\n(3.19) and (3.20) can be written together as a single equation\nkdl\ndτ1√\n˙x2d\ndτ˙xµ\n√\n˙x2=3\n2eFµν(x)˙xν+...xµ\n⊥\n˙x2−3(˙x¨x)¨xµ\n⊥\n(˙x2)2+Λµ,(3.22)\nwhereΛµ:=λµ−ξµ. Obviously, Λµis invariant under the gauge trans-\nformation (3.4). Equation (3.22) is precisely the equation of motion f orxµ\nmentioned under Eq. (3.2). Since xµcontained in lhas been assumed to\nbe a solution of Eq. (3.22), it can be identified with xµin Eq. (3.22).\nUpon considering this fact, substituting the τ-derivative of Eq. (3.2), i.e.,\ndl(τ)/dτ=/radicalbig\n˙xµ(τ)˙xµ(τ), into Eq. (3.22) and recalling k:= 3m/2e2, we\nobtain\nmd\ndτ˙xµ\n√\n˙x2=eFµν(x)˙xν+2\n3e2/parenleftbigg...xµ\n⊥\n˙x2−3(˙x¨x)¨xµ\n⊥\n(˙x2)2/parenrightbigg\n+2\n3e2Λµ.(3.23)\nIfΛµ= 0, Eq. (3.23) is identical with the Lorentz-Dirac equation written\nin terms of the arbitrary world-line parameter τ; see Eq. (A.8) in Appendix\nA. For this reason, Eq. (3.23) can be said to be the Lorentz-Dirac e quation\nwith a source- liketerm 2e2Λµ/3. We thus see that the Lagrangian LDyields\nthe Lorentz-Dirac equation with a source-like term.\nNow we adopt the proper-time gauge τ=l, choosing τto be the proper\ntimel. Accordingly, ˙ xµ=uµ, ˙x2= 1, and ˙ x¨x= 0 are valid, so that Eq.\n(3.23) becomes\nmduµ\ndl=eFµν(x)uν+2\n3e2/parenleftbig\nδµ\nν−uµuν/parenrightbigd2uν\ndl2+2\n3e2Λµ. (3.24)\nThis is exactly what is defined by adding the source-like term 2 e2Λµ/3 to the\n(original) Lorentz-Dirac equation (1.2).\n84. A Lagrangian with additional cross-terms\nNext we consider an alternative Lagrangian defined by\nLA=1\n(q2\n1q2\n2)1/4/bracketleftBigg\n1\n2/parenleftbigg˙q2\n1⊥\nq2\n1−˙q2\n2⊥\nq2\n2/parenrightbigg\n−k\n2/parenleftBigg\n˙q1⊥µqµ\n2/radicalbig\nq2\n1−˙q2⊥µqµ\n1/radicalbig\nq2\n2/parenrightBigg\n−λ1µ(qµ\n1−˙xµ)+λ2µ(qµ\n2−˙xµ)+ξµ(qµ\n1−qµ\n2)−3\n2eFµν(x)qµ\n1qν\n2/bracketrightbigg\n.\n(4.1)\nHere it should be emphasized that LAincludes the additional cross-terms\nproportional to kinstead of the exponential damping function exp( −kl).\nAlso, it is worth noting that unlike LD, the Lagrangian LAdoes not depend\nexplicitly on τ. Using the transformation rules in Eqs. (2.1), (2.2), (2.3),\n(2.4), and (2.6), we can show that the action\nSA=/integraldisplayτ1\nτ0dτLA (4.2)\nremains invariant under the reparametrization τ→τ′. Just like LD, the\nLagrangian LAremains invariant under the gauge transformation (3.4) and\npossesses the antisymmetric property\nLA(qµ\n1,˙qµ\n1,λ1µ;qµ\n2,˙qµ\n2,λ2µ) =−LA(qµ\n2,˙qµ\n2,λ2µ;qµ\n1,˙qµ\n1,λ1µ).(4.3)\nWe now derive the Euler-Lagrange equations for the dynamical var iables\nfromLA. The Euler-Lagrange equation for xµis found to be\nd\ndτ/bracketleftBigg\n1\n(q2\n1q2\n2)1/4(λ1µ−λ2µ)/bracketrightBigg\n+3\n2e(q2\n1q2\n2)1/4∂µFνρ(x)qν\n1qρ\n2= 0.(4.4)\nThis equation includes the gauge-invariant combination λ1µ−λ2µowing to\nthegaugeinvarianceof LA, andhencecannotdetermine λ1µandλ2µuniquely.\nThe Euler-Lagrange equation for qµ\n1can be written as\n1\n(q2\n1q2\n2)1/4/bracketleftbigg1\n2/parenleftbiggd\ndτ∂K1\n∂˙qµ\n1−∂K1\n∂qµ\n1/parenrightbigg\n−k\n2/parenleftbiggd\ndτ∂J\n∂˙qµ\n1−∂J\n∂qµ\n1/parenrightbigg\n+λ1µ−ξµ+3\n2eFµν(x)qν\n2/bracketrightbigg\n+/parenleftBigg\nd\ndτ1\n(q2\n1q2\n2)1/4/parenrightBigg/parenleftbigg1\n2∂K1\n∂˙qµ\n1−k\n2∂J\n∂˙qµ\n1/parenrightbigg\n+q1µ\n2q2\n1LA= 0, (4.5)\n9whereK1andJare given by Eq. (3.8) and\nJ:=˙q1⊥µqµ\n2/radicalbig\nq2\n1−˙q2⊥µqµ\n1/radicalbig\nq2\n2. (4.6)\nApplying the formulas (3.9)–(3.11) and\n∂J\n∂˙qµ\n1=1/radicalbig\nq2\n1/parenleftbigg\nq2µ−q1q2\nq2\n1q1µ/parenrightbigg\n, (4.7)\nd\ndτ∂J\n∂˙qµ\n1−∂J\n∂qµ\n1=d\ndτq2µ/radicalbig\nq2\n2+1/radicalbig\nq2\n1/parenleftbigg\n˙q2µ−q1˙q2\nq2\n1q1µ/parenrightbigg\n(4.8)\nto Eq. (4.5) appropriately, we obtain\nk\n2/bracketleftBigg\nd\ndτq2µ/radicalbig\nq2\n2+1/radicalbig\nq2\n1/parenleftbigg\n˙q2µ−q1˙q2\nq2\n1q1µ/parenrightbigg/bracketrightBigg\n=3\n2eFµν(x)qν\n2+(q2\n1q2\n2)1/4q1µ\n2q2\n1LA+¨q1⊥µ\nq2\n1−/parenleftbigg5q1˙q1\nq2\n1+q2˙q2\nq2\n2/parenrightbigg˙q1⊥µ\n2q2\n1+λ1µ−ξµ\n+k\n4/radicalbig\nq2\n1/parenleftbiggq1˙q1\nq2\n1+q2˙q2\nq2\n2/parenrightbigg/parenleftbigg\nq2µ−q1q2\nq2\n1q1µ/parenrightbigg\n, (4.9)\nwhereq1q2:=q1µqµ\n2andq1˙q2:=q1µ˙qµ\n2. Similarly, the Euler-Lagrange equa-\ntion forqµ\n2is derived as\nk\n2/bracketleftBigg\nd\ndτq1µ/radicalbig\nq2\n1+1/radicalbig\nq2\n2/parenleftbigg\n˙q1µ−˙q1q2\nq2\n2q2µ/parenrightbigg/bracketrightBigg\n=3\n2eFµν(x)qν\n1−(q2\n1q2\n2)1/4q2µ\n2q2\n2LA+¨q2⊥µ\nq2\n2−/parenleftbiggq1˙q1\nq2\n1+5q2˙q2\nq2\n2/parenrightbigg˙q2⊥µ\n2q2\n2+λ2µ−ξµ\n+k\n4/radicalbig\nq2\n2/parenleftbiggq1˙q1\nq2\n1+q2˙q2\nq2\n2/parenrightbigg/parenleftbigg\nq1µ−q1q2\nq2\n2q2µ/parenrightbigg\n. (4.10)\nThe Euler-Lagrange equations for λ1µ,λ2µ, andξµare respectively found to\nbe\nqµ\n1= ˙xµ, (4.11)\nqµ\n2= ˙xµ, (4.12)\nqµ\n1=qµ\n2, (4.13)\n10which are compatible with one another.\nSubstituting Eqs. (4.11) and (4.12) into Eq. (4.9) and noting\nLA(qµ\n1,˙qµ\n1,λ1µ;qµ\n2,˙qµ\n2,λ2µ) =LA(˙xµ,¨xµ,λ1µ; ˙xµ,¨xµ,λ2µ) = 0,(4.14)\nwe have\nkd\ndτ˙xµ\n√\n˙x2=3\n2eFµν(x)˙xν+...xµ\n⊥\n˙x2−3(˙x¨x)¨xµ\n⊥\n(˙x2)2+λµ\n1−ξµ. (4.15)\nSimilarly, substituting Eqs. (4.11) and (4.12) into Eq. (4.10) and using\n(4.14), we have\nkd\ndτ˙xµ\n√\n˙x2=3\n2eFµν(x)˙xν+...xµ\n⊥\n˙x2−3(˙x¨x)¨xµ\n⊥\n(˙x2)2+λµ\n2−ξµ. (4.16)\nComparing Eq. (4.15) and Eq. (4.16) leads to\nλµ\n1=λµ\n2. (4.17)\nThen we see that Eq. (4.4) is identically satisfied owing to ∂µFνρ(x)qν\n1qρ\n2=\n∂µFνρ(x)˙xν˙xρ= 0. With Λµ:=λµ−ξµ(λµ:=λµ\n1=λµ\n2), Eqs. (4.15) and\n(4.16) can be written together as a single equation\nmd\ndτ˙xµ\n√\n˙x2=eFµν(x)˙xν+2\n3e2/parenleftbigg...xµ\n⊥\n˙x2−3(˙x¨x)¨xµ\n⊥\n(˙x2)2/parenrightbigg\n+2\n3e2Λµ,(4.18)\nafter the substitution of k= 3m/2e2. This equation is completely the same\nas Eq. (3.23). In this way, it is established that the Lagrangian LAalso\nyields the Lorentz-Dirac equation with a source-like term.\n5. Summary and discussion\nWe have presented two relativistic Lagrangians LDandLAand have\ndemonstrated that the Euler-Lagrange equations derived from LD, or those\nderived from LA, together lead to the Lorentz-Dirac equation with a source-\nlike term. This equation is a differential equation for xµhaving the inhomo-\ngeneous term 2 e2Λµ/3. Hence it follows that its solutions naturally depend\nonΛµ. The Lorentz-Dirac equation itself can be obtained in a particular\nsituation such that Λµ= 0. For this reason, LDandLAcan simply be said\nto be the Lagrangians for the Lorentz-Dirac equation.\n11Contracting both sides of Eq. (4.18) with ˙ xµ, we have the orthogonality\ncondition\n˙xµΛµ= 0. (5.1)\nThis condition can be written as Λ0=v·Λ, withΛ:= (Λr) (r= 1,2,3) and\nthe velocity vector v:= (dxr/dx0). Accordingly, the 4-vector ( Λµ) can be\nexpressed as ( v·Λ,Λ). As can be seen from Eq. (4.18), the source-like term\n2e2Λµ/3 is regarded as a component of the force 4-vector ( fµ) = (v·f,f),\nprovided that f:= (2e2/3)Λis identified with an external (non-Lorentzian)\nforce acting on the charged particle. In this way, the source-like t erm can be\ntreated as a component of the force 4-vector of an external fo rce.\nIn the indirect formulation of the damped harmonic oscillator [28], a pa ir\nof two coordinate variables is introduced to describe the motion for ward in\ntime and that backward in time. A pair of qµ\n1andqµ\n2does not correspond to\nsuchapairofcoordinatevariables. Infact, qµ\n1andqµ\n2areincludedeveninthe\nLagrangian with an exponential damping function LD. Also,qµ\n1=qµ\n2= ˙xµ\nis eventually found from the Lagrangians LDandLA. For these reasons, qµ\n1\nandqµ\n2should simply be regarded as auxiliary variables useful for deriving\nthe Lorentz-Dirac equation.\nAs has been emphasized above, LDexplicitly depends on the parameter\nτ, whereas LAdoes not depend explicitly on τ. In a consistent quantiza-\ntion of the damped harmonic oscillator [25, 29, 30], the indirect formu lation\nbased on an explicitly time-independent Lagrangian is adopted rathe r than\nthe direct formulation based on an explicitly time-dependent Lagran gian.\nReferring to this fact, we should choose LAas a desirable Lagrangian when\nwe consider quantum theory of a charged particle described by the Lorentz-\nDirac equation. The Lagrangian and Hamiltonian formulations based o nLA\nand the subsequent quantization procedure are interesting issue s that should\nbe addressed in the future.\nAcknowledgments\nWe are grateful to Shigefumi Naka, Takeshi Nihei and Akitsugu Miw a for\ntheir useful comments. We thank Keita Seto for valuable informatio n on the\nLorentz-Dirac equation. One of us (T.S.) thanks Kenji Yamada, Ka tsuhito\nYamaguchi and Haruki Toyoda for their encouragement. The wor k of S.D. is\nsupported in part by Grant-in-Aid for Fundamental Scientific Rese arch from\nCollege of Science and Technology, Nihon University.\n12Appendix A. Lorentz-Dirac equation written in terms of an ar bi-\ntrary world-line parameter\nLet us recall the Lorentz-Dirac equation Eq. (1.2), i.e.,\nmduµ\ndl=eFµν(x)uν+2\n3e2/parenleftbig\nδµ\nν−uµuν/parenrightbigd2uν\ndl2, (A.1)\nwhich is written in terms of the proper time l. Noting that the infinitesimal\nproper time can be expressed as dl=√\n˙x2dτ, we can show that\nuµ:=dxµ\ndl=˙xµ\n√\n˙x2, (A.2)\nduµ\ndl=¨xµ\n⊥\n˙x2, (A.3)\nd2uµ\ndl2=1√\n˙x2/bracketleftbigg...xµ\n⊥\n˙x2−3(˙x¨x)¨xµ\n⊥\n(˙x2)2−/parenleftbigg\n¨x2−(˙x¨x)2\n˙x2/parenrightbigg˙xµ\n(˙x2)2/bracketrightbigg\n,(A.4)\nwith\n¨xµ\n⊥:= ¨xµ−˙x¨x\n˙x2˙xµ,...xµ\n⊥:=...xµ−˙x...x\n˙x2˙xµ. (A.5)\nHere,\n˙xµ:=dxµ\ndτ,¨xµ:=d2xµ\ndτ2,...xµ:=d3xµ\ndτ3, (A.6)\n˙x2:= ˙xµ˙xµ,¨x2:= ¨xµ¨xµ,˙x¨x:= ˙xµ¨xµ,˙x...x:= ˙xµ...xµ.(A.7)\nSubstituting Eqs. (A.2) and (A.4) into Eq. (A.1), we obtain\nmd\ndτ˙xµ\n√\n˙x2=eFµν(x)˙xν+2\n3e2/parenleftbigg...xµ\n⊥\n˙x2−3(˙x¨x)¨xµ\n⊥\n(˙x2)2/parenrightbigg\n.(A.8)\nThis is the Lorentz-Dirac equation written in terms of anarbitrary w orld-line\nparameter τ. We can directly derive Eq. (A.8) by evaluating the reaction\nforce due to the particle’s own electromagnetic radiation without ad opting\nthe proper-time gauge τ=l.\n13References\n[1] L. D. Landau and E. M. Lifshitz, The Classical Theory of Fields , 4th\ned., Butterworth-Heinemann, Oxford, 1975.\n[2] J. D. Jackson, Classical Electrodynamics , 3rd ed., John Wiley & Sons,\nNew York, 1998.\n[3] A. O. Barut, Electrodynamics and Classical Theory of Fields and Par-\nticles, Dover, New York, 1980.\n[4] F. Rohrlich, Classical Charged Particles , 3rd ed., World Scientific, Sin-\ngapore, 2007.\n[5] F. Rohrlich, “The dynamics of a charged sphere and the electron ,” Am.\nJ. Phys. 65(1997) 1051-1056.\n[6] H. Spohn, Dynamics of Charged Particles and Their Radiation Field ,\nCambridge University Press, Cambridge, 2004.\n[7] H. A. Lorentz, “La th´ eorie ´ elecromagnetique de Maxwell et so n appli-\ncation aux corps mouvemants,” Arch. N´ eerland. Sci. Exactes Nat .25\n(1892) 363-552.\n[8] M. Abraham, Theorie der Elektrizit¨ at, Vol. II:Elektromagnetische The-\norie der Strahlung , Teubner, Leipzig, 1905.\n[9] H. A. Lorentz, The Theory of Electrons and Its Applications to the Phe-\nnomena of Light and Radiant Heat, 2nd ed., Dover, New York, 1952.\n[10] P. A. M. Dirac, “Classical theory of radiating electrons,” Proc. R. Soc.\nLondon Ser. A 167(1938) 148-169.\n[11] A. O. Barut, “Electrodynamics in terms of retarded fields,” Phy s. Rev.\nD10(1974) 3335-3336.\n[12] E. Poisson, “An introduction to the Lorentz-Dirac equation,” u npub-\nlished, arXiv:gr-qc/9912045.\n[13] C. W. Nakhleh, “TheLorentz-DiracandLandau-Lifshitzequat ions from\nthe perspective of modern renormalization theory,” Am. J. Phys. 81\n(2013) 180-185, arXiv:1207.1745 [physics.class-ph].\n14[14] A. D. Yaghjian, Relativistic Dynamics of a Charged Sphere , 2nd ed.,\nLect. Notes Phys. Vol. 686, Springer-Verlag, New York, 2006.\n[15] A. O. Barut, “Renormalization and elimination of preacceleration and\nrunaway solutions of the Lorentz-Dirac equation,” Phys. Lett. A 145\n(1990) 387-390.\n[16] G.W.FordandR.F.O’Connell, “Radiationreactioninelectrodyna mics\nand the elimination of runaway solutions,” Phys. Lett. A 157(1991)\n217-220.\n[17] G. W. Ford and R. F. O’Connell, “Relativistic form of radiation reac -\ntion,” Phys. Lett. A 174(1993) 182-184.\n[18] H. Spohn, “The critical manifold of the Lorentz-Dirac equation ,” Euro-\nphys. Lett. 50(2000) 287-292, arXiv:physics/9911027.\n[19] F.Rohrlich, “Thecorrect equationofmotionofaclassical point charge,”\nPhys. Lett. A 283(2001) 276-278.\n[20] F. Rohrlich, “Dynamics of a charged particle,” Phys. Rev. E 77(2008)\n046609 (1-4).\n[21] K.Seto, S.Zhang, J.Koga,H.Nagatomo,M.Nakai, andK.Mima , “Sta-\nbilization of radiation reaction with vacuum polarization,” Prog. Theor .\nExp. Phys. 2014(2014) 043A01 (1-10), arXiv:1310.6646 [physics.plasm-\nph].\n[22] A. Carati, “A Lagrangian formulation for the Abraham-Lorent z-Dirac\nequation,” in Symmetry and Perturbation Theory (Proceedings edited\nby D. Bambusi and G. Gaeta) Consiglio Nazionale delle Ricerche as\n“Quaderno GNFM-CNR, Vol. 54”, Roma, 1998.\n[23] P. M. V. B. Barone and A. C. R. Mendes, “Lagrangian descrip-\ntion of the radiation damping,” Phys. Lett. A 364(2007) 438-440,\narXiv:cond-mat/0607370.\n[24] V. G. Kupriyanov, “Hamiltonian formulation and action principle fo r\nthe Lorentz-Dirac system,” Int. J. Theor. Phys. 45(2006) 1129-1144.\n[25] H. Dekker, “Classical and quantum mechanics of the damped ha rmonic\noscillator,” Phys. Rep. 80(1981) 1-112.\n15[26] P. Caldirola, “Forze non conservative nella meccanica quantistic a,”\nNuovo Cimento 18(1941) 393-400.\n[27] E. Kanai, “Onthequantizationofthedissipative systems,” Prog .Theor.\nPhys.3(1948) 440-442.\n[28] H.Bateman, “Ondissipative systems andrelatedvariationalpr inciples,”\nPhys. Rev. 38(1931) 815-819.\n[29] H.FeshbachandY.Tikochinsky, “Quantizationofthedampedh armonic\noscillator,” Trans. N.Y. Acad. Sci., Ser. II 38(1977) 44-53.\n[30] R. Banerjee and P. Mukherjee, “A canonical approach to the quanti-\nzation of the damped harmonic oscillator,” J. Phys. A: Math. Gen. 35\n(2002) 5591-5598, arXiv:quant-ph/0108055.\n[31] M. Pavˇ siˇ c, “Classical motion of membranes, strings and poin t particles\nwith extrinsic curvature,” Phys. Lett. B 205(1988) 231-236.\n[32] M. Pavˇ siˇ c, “Thequantization ofa point particle with extrinsic curvature\nleads to the Dirac equation,” Phys. Lett. B 221(1989) 264-268.\n[33] M. S. Plyushchay, “Does the quantization of a particle with curv ature\nlead to Dirac equation?,” Phys. Lett. B 253(1991) 50-55.\n16" }, { "title": "2008.10381v3.Is_the_local_Lorentz_invariance_of_general_relativity_implemented_by_gauge_bosons_that_have_their_own_Yang_Mills_like_action_.pdf", "content": "Is the local Lorentz invariance of general relativity implemented by gauge bosons that\nhave their own Yang-Mills-like action?\nKevin Cahill\u0003\nDepartment of Physics and Astronomy\nUniversity of New Mexico\nAlbuquerque, New Mexico 87131\n(Dated: October 18, 2022)\nGeneral relativity with fermions has two independent symmetries: general coordinate invariance\nand local Lorentz invariance. General coordinate invariance is implemented by the Levi-Civita con-\nnection and by Cartan's tetrads both of which have as their action the Einstein-Hilbert action. It is\nsuggested here that local Lorentz invariance is implemented not by a combination of the Levi-Civita\nconnection and Cartan's tetrads known as the spin connection, but by independent Lorentz bosons\nLab\nithat gauge the Lorentz group, that couple to fermions like Yang-Mills \felds, and that have\ntheir own Yang-Mills-like action. A nonsingular 4 \u00024 hermitian scalar \feld his needed to make the\naction of the Lorentz bosons invariant under local Lorentz transformations. Lorentz bosons couple\nto fermion number and generate a spin-dependent static potential that violates the weak equiva-\nlence principle. If a Higgs mechanism makes them massive, then the static potential also violates the\ninverse-square law. Experiments put upper bounds on the strength of such a potential for masses\nmL<20 eV. These upper limits imply that Lorentz bosons, if they exist, are nearly stable and\ncontribute to dark matter.\nI. INTRODUCTION\nGeneral relativity with fermions has two indepen-\ndent symmetries: general coordinate invariance and local\nLorentz invariance. General coordinate invariance is the\nwell-known, de\fning symmetry of general relativity. It\nacts on coordinates and on the world indexes of tensors\nbut leaves Dirac and Lorentz indexes unchanged. It is im-\nplemented by the Levi-Civita connection and by Cartan's\ntetrads.\nLocal Lorentz invariance is a quite di\u000berent symmetry.\nIt acts on Dirac and Lorentz indexes but leaves coordi-\nnates and world indexes unchanged. In standard formu-\nlations [1{5], the derivative of a Dirac \feld is made co-\nvariant (@i+!i) by a combination of the Levi-Civita\nconnection \u0000j\nkiand Cartan's tetrads ca\njknown as the\nspin connection\n!i=1\n8!ab\ni[\ra;\rb] (1)\nin which\n!ab\ni=ca\njcbk\u0000j\nki+ca\nk@icbk; (2)\naandbare Lorentz indexes, and i;j;k are world indexes.\nBecause it acts on Lorentz and Dirac indexes but\nleaves world indexes and coordinates unchanged, local\nLorentz invariance is more like an internal symmetry than\nlike general coordinate invariance. In theories with local\nLorentz invariance and internal symmetry, the covariant\nderivativeDiof a vector of Dirac \felds has the spin\nconnection !iand a matrix Aiof Yang-Mills \felds side\nby side\nDi = (@i+!i+Ai) : (3)\n\u0003cahill@unm.eduJust as the Yang-Mills connection Aiis a linear combina-\ntionAi=\u0000it\u000bA\u000b\niof the matrices t\u000bthat generate the\ninternal symmetry group, so too the spin connection !i\nis a linear combination !i=1\n8!ab\ni[\ra;\rb]of the matrices\n\u0000i1\n4[\ra;\rb]that generate the Lorentz group.\nSo I ask: Does the independent symmetry of local\nLorentz invariance have its own, independent gauge \feld\nLi=1\n8Lab\ni[\ra;\rb] (4)\nwith its own \feld strength Fik=[@i+Li;@k+Lk]and\nYang-Mills-like action\nSL=\u00001\n4f2Z\nTr\u0010\nFy\nikhFikh\u00001\u0011pgd4x? (5)\nThis action is made suitably positive and invariant under\nthe local Lorentz transformations\nF0=D\u00001(\u0003)FD(\u0003) and h0=Dy(\u0003)hD(\u0003) (6)\nby a nonsingular 4 \u00024 hermitian scalar \feld h[6] whose\naction density is\nSh=\u0000M2Tr\u0002\n(Dih)h\u00001(Dih)h\u00001\u0003\n(7)\nin which its covariant derivative is [6]\nDih=@ih\u0000hLi\u0000Ly\nih: (8)\nThe covariant derivative of a Dirac \feld would then be\nDi =\u0000\n@i+1\n8Lab\ni[\ra;\rb]\u0001\n (9)\ninstead of the standard form (1{3)\nDi =\u0010\n@i+1\n8\u0000\nca\njcbk\u0000j\nki+ca\nk@icbk\u0001\n[\ra;\rb]\u0011\n :(10)arXiv:2008.10381v3 [gr-qc] 16 Oct 20222\nIf so, then the Lorentz bosons Lab\nicouple to fermion\nnumber and not to mass and lead to Coulomb and\nYukawa potentials that violate the inverse-square law\nand the weak equivalence principle. The hermitian scalar\n\feldhmust assume a nonzero value in the vacuum\nh0=h0jhj0ibecause it is a nonsingular matrix. It may\nplay the role of a new Higgs \feld. Its action (7) introduces\na new mass into the theory.\nExperiments [7{30] have set upper limits on the\nstrength of such Yukawa potentials for Lorentz bosons\nof mass less than 20 eV. These upper limits imply that\nLorentz bosons, if they exist, are nearly stable and con-\ntribute to dark matter. Whether fermions couple to\nLorentz bosons Liwith their own action SLor to the\nspin connection !iis an open experimental question.\nThis paper outlines a version of general relativity with\nfermions in which the six vector bosons of the spin con-\nnection!ab\niare replaced by six vector bosons Lab\nithat\ngauge the Lorentz group and have their own Yang-Mills-\nlike action. The theory is invariant under general co-\nordinate transformations and independently under local\nLorentz transformations.\nSection II sketches the traditional way of including\nfermions in a theory of general relativity. Section III\ndescribes the local Lorentz invariance of a theory with\nLorentz bosons. Section IV says why general-coordinate\ninvariance and local-Lorentz invariance are independent\nsymmetries. Section V describes the Yang-Mills-like ac-\ntion (5) of the gauge \felds Lab\niof the local Lorentz group.\nSections VI and VII suggest ways to make gauge the-\nory and general relativity more similar to each other.\nSection VIII discusses Higgs mechanisms that may give\nmasses to the gauge bosons Lab\niof the Lorentz group.\nSection IX describes some of the constraints that exper-\nimental tests [7{30] of the inverse-square law and of the\nweak equivalence principle place upon the proposed the-\nory. Section X discusses the stability and masses of L\nbosons and suggests that they may be part or all of dark\nmatter. Section XI summarizes the paper.\nII. GENERAL RELATIVITY WITH FERMIONS\nA century ago, Einstein described gravity by the action\nSE=1\n16\u0019GZ\nRpg d4x=1\n16\u0019GZ\ngikRikpg d4x\n(11)\nin whichG= 1=m2\nPis Newton's constant, the metric\nis (\u0000;+;+;+), letters from the middle of the alphabet\nare world indexes, g=jdetgikjis the absolute value of\nthe determinant of the space-time metric, and the Ricci\ntensorRik=R`\ni`kis the trace of the Riemann tensor\nRj\ni`k=@`\u0000j\nki\u0000@k\u0000j\n`i+ \u0000j\n`m\u0000m\nki\u0000\u0000j\nkm\u0000m\n`i(12)\nin which\n\u0000k\ni`=1\n2gkj(@`gji+@igj`\u0000@jgi`) = \u0000k\n`i (13)is the Levi-Civita connection which makes the covariant\nderivative of the metric vanish [31].\nThe standard action of general relativity with fermions\nis the sum of the Einstein-Hilbert action (11) and the\naction of matter \felds including the Dirac action\nZ\n\u0000\u0016 \raci\na(@i+!i+Ai) pgd4x: (14)\nIn what follows, it is proposed to replace the spin con-\nnection!iin the standard Dirac action (14) with an in-\ndependent gauge \feld Li=\u00001\n8Lab\ni[\ra;\rb]that has its\nown action (5) and to use\nSD=Z\n\u0000\u0016 \raci\na(@i+Li+Ai) pgd4x (15)\nas the action of a Dirac \feld. This change re\rects the\nindependence of general coordinate invariance and lo-\ncal Lorentz invariance and makes general relativity and\nquantum \feld theory somewhat more similar.\nIII. LOCAL LORENTZ INVARIANCE\nThe Einstein action (11) has a trivial symmetry under\nlocal Lorentz transformations that act on Lorentz indexes\nbut leave world indexes and coordinates unchanged. This\nsymmetry becomes apparent when Cartan's tetrads ca\ni\nandcb\nkare used to write the metric gikin a form\ngik(x) =ca\ni(x)\u0011abcb\nk(x) (16)\nthat is unchanged by local Lorentz transformations\nc0a\ni(x) = \u0003a\nb(x)cb\ni(x): (17)\nThe Levi-Civita connection (13) and the action (11) are\nde\fned in terms of the metric and so are also invariant\nunder local Lorentz transformations.\nMore importantly, the two Dirac actions (14) and (15)\nhave a nontrivial symmetry under local Lorentz trans-\nformations. Under such a local Lorentz transformation, a\nDirac \feld transforms under the (1\n2;0)\b(0;1\n2) represen-\ntationD(\u0003) of the Lorentz group with no change in its\ncoordinates x\n 0\n\u000b(x) =D\u00001\n\u000b\f(\u0003(x)) \f(x): (18)\nThe Lorentz-boson matrix Li=\u00001\n8Lab\ni[\ra;\rb]makes\n@i+Lia covariant derivative\n@i+L0\ni=D\u00001(\u0003)(@i+Li)D(\u0003): (19)\nIn more detail with \u0003 = \u0003( x), the matrix Litransforms\nas\nL0\ni=D\u00001(\u0003)@iD(\u0003) +D\u00001(\u0003)LiD(\u0003)\n=D\u00001(\u0003)@iD(\u0003)\u00001\n8D\u00001(\u0003)Lab\ni[\ra;\rb]D(\u0003)\n=D\u00001(\u0003)@iD(\u0003)\u00001\n8Lab\ni\u0003c\na\u0003d\nb[\rc;\rd] (20)3\nin which \u0003c\na= \u0003\u00001c\na. Since\nTr\u0000\n[\ra;\rb][\rc;\rd]\u0001\n= 16\u0000\n\u000ea\nd\u000eb\nc\u0000\u000ea\nc\u000eb\nd\u0001\n; (21)\nits components transform as\nL0ab\ni=\u00001\n2Tr(L0\ni[\ra;\rb]) (22)\n= \u0003a\nc\u0003b\ndLcd\ni\u00001\n2Tr\u0000\nD\u00001(\u0003)@iD(\u0003)[\ra;\rb]\u0001\n:\nUnder an in\fnitesimal transformation\n\u0003 =I+\u0015andD(\u0015) =I\u00001\n8\u0015ab[\ra;\rb]; (23)\nthe Lorentz bosons transform as\nL0ab\ni=Lab\ni+Lcb\ni\u0015a\nc+Lad\ni\u0015b\nd+@i\u0015ab: (24)\nThe components of the spin connection obey similar\nequations, and the conventional Dirac action (14) also\nis invariant under local Lorentz transformations.\nLocal Lorentz transformations operate on the Lorentz\nindexesa;b;c;::: of the tetrads, of the spin connection\n!ab\ni, and of the gamma matrices \ra[\rb;\rc], and also on\nthe Dirac indexes \u000b;\f;\r of the gamma matrices and of\nthe Dirac \felds \u0016 ; , but not upon the world index ior\nthe spacetime coordinates x. In this sense, the invariance\nof Dirac's action SDunder local Lorentz transformations\nis like an internal symmetry.\nIV. LOCAL LORENTZ INVARIANCE AND\nINVARIANCE UNDER GENERAL\nCOORDINATE TRANSFORMATIONS ARE\nINDEPENDENT SYMMETRIES\nThe Dirac action SDis invariant both under a local\nLorentz transformation \u0003( x) and under a general coordi-\nnate transformation x!x0. Under a local Lorentz trans-\nformation \u0003( x), the coordinates are unchanged, x0=x,\nand the \felds transform as\n 0\n\u000b(x) =D\u00001\n\u000b\f(\u0003(x)) \f(x)\nc0a\ni(x) = \u0003a\nb(x)cb\ni(x)\nL0ab\ni(x) = \u0003a\nc(x)\u0003b\nd(x)Lcd\ni(x)\n\u00001\n2Tr\u0000\nD\u00001(\u0003(x))@iD(\u0003(x))[\ra;\rb]\u0001\nh0(x) =Dy(\u0003(x))h(x)D(\u0003(x)) (25)\nin which \u0003a\nc(x) = \u0003\u00001a\nc. Under a general coordinate\ntransformation, the \felds transform as\n 0\n\u000b(x0) = \u000b(x)\nc0a\ni(x0) =@xk\n@x0ica\nk(x)\nL0ab\ni(x0) =@xk\n@x0iLab\nk(x)\nh(x0) =h(x): (26)The two transformations, \u000b(x)!D\u00001\n\u000b\f(\u0003(x)) \f(x) and\nx!x0, are di\u000berent and independent; the coordinates x0\nand \u0003(x)xare unrelated.\nEvery conventional, local Lorentz transformation is\na general coordinate transformation, so one might be\ntempted to imagine that every general coordinate trans-\nformation is a conventional, local Lorentz transformation.\nBut one can see that this is not the case by comparing the\nin\fnitesimal form of a general coordinate transformation\ndx0i=@x0i\n@xkdxk(27)\nwhich has 16 generators with that of a conventional, local\nLorentz transformation\ndx0a= \u0003a\nbdxb=dxa+ (\u000fr\u0001Ra\nb+\u000fb\u0001Ba\nb)dxb(28)\nwhich has only 6 [32].\nSpecial relativity o\u000bers another temptation. In special\nrelativity, global Lorentz transformations \u0003 act on the\nspacetime coordinates and on the indexes of a Dirac \feld\nx0a= \u0003a\nbxb\n 0\n\u000b(x0) =D\u00001\n\u000b\f(\u0003) \f(\u0003x):(29)\nThis global Lorentz transformation leaves the specially\nrelativistic Dirac action density unchanged\n\u0002\n\u0000i y\r0\ra@a \u00030=\u0000i yD\u00001y\r0\raD\u00001@0\na \n=\u0000i y\r0D\raD\u00001\u0003c\na@c \n=\u0000i y\r0\rb\u0003a\nb\u0003c\na@c (30)\n=\u0000i y\r0\rb\u000ec\nb@c =\u0000i y\r0\rb@b :\nBut in general relativity with fermions, Cartan's\ntetradsci\naallow the action to be invariant under a local\nLorentz transformation without a corresponding general\ncoordinate transformation. The matrix D\u00001\n\u000b\f(\u0003(x)) repre-\nsents a local Lorentz transformation and acts (18) on the\nspinor indexes of the Dirac \feld but not on its spacetime\ncoordinates. Since\nD\raD\u00001= \u0003a\na0\ra0andD\raD\u00001= \u0003a0\na\ra0;(31)\na local Lorentz transformation (20) does not change\nD\raD\u00001c0i\na=\rb\u0003a\nb\u0003c\naci\nc=\rb\u000ec\nbci\nc=\rbci\nb: (32)\nBut the e\u000bect of a local Lorentz transformation (20) on\nthe Lorentz matrix Liis\nD(@i+L0\ni)D\u00001\u0000@i=\u00001\n8L0ab\niD[\ra;\rb]D\u00001\n=\u00001\n8\u0003a\nc\u0003b\ndLcd\niD[\ra;\rb]D\u00001\n=\u00001\n8\u0003\u00001a\nc\u0003\u00001b\ndLcd\ni\u0003e\na\u0003f\nb[\re;\rf]\n=\u00001\n8\u000ee\nc\u000ef\ndLcd\ni[\re;\rf]\n=\u00001\n8Lcd\ni[\rc;\rd]=Li (33)4\nso that\nD(@i+L0\ni)D\u00001=@i+Li: (34)\nA local Lorentz transformation therefore leaves the Dirac\naction density invariant\n\u0002\n\u0000i y\r0\raci\na(@i+Li) \u00030\n=\u0000i yD\u00001y\r0\rac0i\na(@i+L0\ni)D\u00001 \n=\u0000i yD\u00001y\r0\rac0i\naD\u00001D(@i+L0\ni)D\u00001 \n=\u0000i y\r0D\rac0i\naD\u00001(@i+Li) (35)\n=\u0000i y\r0\raci\na(@i+Li) :\nThe symmetry under local Lorentz transformations is\nindependent of the symmetry under general coordinate\ntransformations. They are independent symmetries.\nV. ACTION OF THE GAUGE FIELDS Li\nSince local Lorentz symmetry is like an internal sym-\nmetry, its gauge \felds Li=1\n8Lab\ni[\ra;\rb]should have an\naction like that of a Yang-Mills \feld\nSL=\u00001\n4f2Z\nTr\u0010\nFy\nikhFikh\u00001\u0011pgd4x (36)\nin which\nFik=[@i+Li;@k+Lk]; (37)\nandhis a 4\u00024 nonsingular hermitian matrix [6] .\nUnder local Lorentz transformations D(\u0003(x)), the \feld\nstrengthsFikand the matrix htransform as\nF0ik(x) =D\u00001(\u0003(x))Fik(x)D(\u0003(x))\nh0(x) =Dy(\u0003(x))h(x)D(\u0003(x))(38)\nand so the action SLis invariant under local Lorentz\ntransformations as well as under general coordinate\ntransformations [6] .\nOne might think that it would be su\u000ecient to set h=\n\f=i\r0, sinceDy(\u0003)\fD(\u0003) =\fandD(\u0003)\fDy(\u0003) =\f,\nbut the resulting action SLis not bounded below.\nIn terms of the gamma matrices\n\r0=\u0000i\u0012\n0 1\n1 0\u0013\n; \ri=\u0000i\u0012\n0\u001bi\n\u0000\u001bi0\u0013\n;\nand\r5=\u0012\n1 0\n0\u00001\u0013\n;(39)\nthe commutators in Li=\u00001\n8Lab\ni[\ra;\rb]are for spatial\na;b;c = 1;2;3,\n[\ra;\rb]= 2i\u000fabc\u001bcIand [\r0;\ra]=\u00002\u001ba\r5:(40)\nSo setting\nra\ni=1\n2\u000fabcLbc\niandba\ni=La0\ni; (41)the matrix of gauge \felds Liis\nLi=\u00001\n8Lab\ni[\ra;\rb]=\u0000i1\n2ri\u0001\u001bI\u00001\n2bi\u0001\u001b\r5;(42)\nand its \feld strength (37) is\nFik=[@i+Li;@k+Lk] (43)\n=\u0000i1\n2(@irk\u0000@kri+ri\u0002rk\u0000bi\u0002bk)\u0001\u001bI\n\u00001\n2(@ibk\u0000@kbi+ri\u0002bk+bi\u0002rk)\u0001\u001b\r5:\nWith the abbreviations\nRik= (@irk\u0000@kri+ri\u0002rk\u0000bi\u0002bk)\nBik=(@ibk\u0000@kbi+ri\u0002bk+bi\u0002rk);(44)\nthe actionSLis the trace\nSL=\u00001\n16f2Z\nTrh\u0000\nRik\u0001\u001bI+iBik\u0001\u001b\r5\u0001\nh\n\u0000\nRik\u0001\u001bI\u0000iBik\u0001\u001b\r5\u0001\nh\u00001ipgd4x:(45)\nThe 4\u00024 nonsingular hermitian matrix hmay play a\nrole in the action (15) of a spin-one-half \feld because\nwe may take the quantity \u0016 either to be \u0016 = y\fas\nusual or to be \u0016 = yh. The covariant derivative (8) of\nhtransforms as\u0000\nDih\u00010=Dy(\u0003)DihD(\u0003) .\nVI. MAKING GENERAL RELATIVITY MORE\nSIMILAR TO GAUGE THEORY\nThere are three reasons to de\fne the covariant deriva-\ntive of a Dirac \feld in terms of Lorentz bosons\nLi=1\n8Lab\ni[\ra;\rb] (46)\nwith their own action (36) as\nDi = (@i+Li+Ai) (47)\nrather than in terms of the spin connection (1)\n!i=1\n8!ab\ni[\ra;\rb]\n=1\n8\u0010\nca\njcbk\u0000j\nki+ca\nk@icbk\u0011\n[\ra;\rb](48)\nas (3)\n(@i+!i+Ai) : (49)\nOne reason is that the symmetry of local Lorentz trans-\nformations is independent of the symmetry of general\ncoordinate transformations. So local Lorentz invariance\nshould have its own gauge \feld Liand action SLinde-\npendent of the tetrads and the Levi-Civita connection of\ngeneral coordinate transformations.\nA second reason to prefer the Lorentz connection Li\nto the spin connection !iis that the L-boson covariant\nderivative\n\u0000\n@i+1\n8Lab\ni[\ra;\rb]\u0001\n (50)5\nis simpler than the spin-connection covariant derivative\n\u0010\n@i+1\n8(ca\njcbk\u0000j\nki+ca\nk@icbk)[\ra;\rb]\u0011\n : (51)\nA third reason is that using the Lorentz connection\n(46), the Dirac covariant derivative (47), and the action\n(36) for the Lorentz connection, makes general relativity\nwith fermions more similar to the gauge theories of the\nstandard model.\nVII. MAKING GAUGE THEORY MORE\nSIMILAR TO GENERAL RELATIVITY\nUnder a local Lorentz transformation, the spin connec-\ntion!ichanges more naturally, more automatically than\ndoes the Lorentz connection Li. The automatic feature of\nthe spin connection is that its de\fnition (2) implies that\nunder in\fnitesimal (23) and \fnite local Lorentz transfor-\nmations it transforms as\n!0ab\ni=!ab\ni+!cb\ni\u0015a\nc+!ad\ni\u0015b\nd+@i\u0015ab(52)\nand as\n!0ab\ni= \u0003a\nc\u0003b\nd!cd\ni+ \u0003a\nc@i\u0003cb: (53)\nThe terms @i\u0015aband \u0003a\nc@i\u0003cboccur automatically\nwithout the need to put in by hand a term like\nD\u00001(\u0003)@iD(\u0003).\nTerms like D\u00001(\u0003)@iD(\u0003) are a common feature of\ngauge theories whether abelian or nonabelian. We can\nmake them occur automatically in local Lorentz trans-\nformations if we add to the Lorentz connection Lab\ni\nthe termua\n\u000b@iub\u000bin which the four Lorentz 4-vectors\nua\u000b(x) obey the condition\nua\u000b\u0011\u000b\fub\f=\u0011ab; (54)\nand\u000b= 0;1;2;3 is a label, not an index. It follows then\nfrom this condition (54) on the quartet of vectors ua\u000b\nthat the augmented Lorentz connection Lab\nnewi\nLab\nnewi=Lab\ni+ua\n\u000b@iub\u000b(55)\nautomatically changes under a local Lorentz transforma-\ntion \u0003a\nc= \u0003\u00001a\ncto\nL0ab\nnewi= \u0003a\nc\u0003b\ndLcd\ni+ \u0003a\ncuc\n\u000b@i(\u0003b\ndud\u000b)\n= \u0003a\nc\u0003b\nd(Lcd\ni+uc\n\u000b@iud\u000b) +uc\n\u000bud\u000b\u0003a\nc@i\u0003b\nd\n= \u0003a\nc\u0003b\ndLcd\nnewi+\u0011cd\u0003a\nc@i\u0003b\nd\n= \u0003a\nc\u0003b\ndLcd\nnewi+ \u0003a\nc@i\u0003cb\n= \u0003a\nc\u0003b\ndLcd\nnewi+ \u0003\u00001a\nc@i\u0003cb(56)\nwithout the need to explicitly add the last term\n\u0003\u00001a\nc@i\u0003cbby hand.\nIn matrix form, the condition (54) is the requirement\nua\u000b\u0011\u000b\fub\f=\u0011ab(57)that the matrix formed by the quartet of vectors ua\u000bbe\na Lorentz transformation\nxaua\u000b\u0011\u000b\fub\fyb=xa\u0011abyb: (58)\nThe augmentation of the Lorentz connection Lab\ni!\nLab\nnewiby the addition of the term ua\n\u000b@iub\u000b, which is\nsimilar to the tetrad term ca\nk@icbkof the spin connection\n(2), makes its change (56) under local Lorentz transfor-\nmations as automatic as that (53) of the spin connection.\nThe use of a more automatic connection makes gauge\ntheory more similar to general relativity with fermions.\nWe can extend the use of such terms to internal sym-\nmetries and so make the inhomogeneous terms appear\nautomatically rather than by hand or by \fat. For in-\nstance, we can augment the abelian connection Aito\nAnewi(x) =Ai(x) +e\u0000i\u001e(x)@iei\u001e(x)(59)\nin which\u001e(x) is an arbitrary phase. A U(1) transforma-\ntion\ne\u0000i\u001e(x)!e\u0000i(\u0012(x)+\u001e(x))and (x)!e\u0000i\u0012(x) (x) (60)\nwould then change the covariant derivative ( @i+Anewi) \nto\n[(@i+Anewi) ]0= (@i+Ai+e\u0000i(\u0012+\u001e)@iei(\u0012+\u001e))e\u0000i\u0012 \n=e\u0000i\u0012(@i\u0000i@i\u0012+Ai+i@i\u0012+e\u0000i\u001e@iei\u001e) (61)\n=e\u0000i\u0012(@i+Ai+e\u0000i\u001e@iei\u001e) =e\u0000i\u0012(@i+Anewi) :\nSimilarly, we can augment the nonabelian connection\nAi=\u0000it\u000bA\u000b\niforSU(n) to\nAnewi(x) =Ai(x) +u\u000b\f(x)@iu\u0003\n\u000b\r(x) (62)\nin which the n n-vectorsu\u000b\rare orthonormal\nu\f\u000bu\u0003\n\r\u000b=\u000e\f\r: (63)\nAnSU(n) transformation\nAi!gAigy; u!gu;and !g (64)\nwould then change the covariant derivative ( @i+Anewi) \nto\n[(@i+Anewi) ]0= (@i+gAigy+gu@i(uygy))g \n=g(@i+gy@ig+Ai+ (@igy)uuyg+u@iuy) \n=g(@i+gy@ig+Ai+ (@igy)g+u@iuy) \n=g(@i+Ai+u@iuy) =g(@i+Anewi) : (65)\nVIII. POSSIBLE HIGGS MECHANISMS\nThe 4\u00024 nonsingular hermitian matrix his needed to\nmake the action (36) gauge invariant. It also may replace6\n\fin the fermionic action (15). Since it is nonsingular, it\nmust assume a nonzero average value in the vacuum\nh0=h0jhj0i: (66)\nSo we have a new kind of Higgs mechanism that can give\nmasses to gauge \felds and fermions. This will be taken\nup in later papers.\nOther kinds of Higgs mechanisms are also possible with\ndi\u000berent Higgs \felds. The actions SLandSD(5 & 15)\nleave the gauge bosons Lmassless, but a Higgs mecha-\nnism is possible. An interaction with a \feld vathat is\na scalar under general coordinate transformations but a\nvector under local Lorentz transformations has as its co-\nvariant derivative\nDiva=@iva+La\nbivb: (67)\nIf the time component v0has a nonzero mean value in\nthe vacuumh0jv0j0i 6= 0, then the scalar \u0000DivaDiva\ncontains a mass term\n\u0000La\n0iv0L0i\nav0=\u0000La0\niv0La0iv0(68)\nthat makes the boost vector bosons bs\ni=Ls0\nimassive\nbut leaves the rotational vector bosons rs\ni=1\n2\u000fstuLtu\ni\nmassless. On the other hand, if the spatial components\nhave a nonzero mean value, h0jvj0i6= 0, then the mass\nterm is\n\u0000La\nsivsLsi\navs=\u0000Ls0s\nivsLs0sivs+L0s\nivsL0sivs:(69)\nAdding the two mass terms (68 and 69), we \fnd\n\u0000La\n0ivbL0i\navb=L0s\nivaL0siva\u0000Ls0s\nivsLs0sivs(70)\nwhich makes all six gauge bosons massive as long as the\nmean value is timelike\nh0jvavaj0i<0 (71)\nand at least two spatial components h0jvsj0i6= 0 have\nnonzero mean values in the vacuum. This condition holds\nin all Lorentz frames if three vectors va1;va2;andva3\nhave di\u000berent timelike mean values in the vacuum.\nIn the vacuum of \rat space, tetrads have mean values\nthat are Lorentz transforms of ca\ni=\u000ea\niand that produce\nthe Minkowski metric (16)\ngik=\u000ea\ni\u0011ab\u000eb\nk=\u0011ik: (72)\nSo it is tempting to look for a Higgs mechanism that\nuses the covariant derivatives D`ca\nkof the tetrads. For\n\u0000j\nk`= 0 andca\nk=\u000ea\nk, the term\n\u00001\n2m2\nL(Dick\na)Dica\nk=\u00001\n2m2\nLLbi\nack\nbLa\ncicc\nk\n=\u00001\n2m2\nLLi\nabLab\ni(73)\nmakes the rotational bosons rs\ni=1\n2\u000fstuLtu\nimassive but\nmakes the boost bosons bs\ni=Ls0\nitachyons. If weakly\ncoupled tachyons are unacceptable, then the Higgs mech-\nanism (67{70) that uses three world-scalar Lorentz vec-\ntors with di\u000berent time-like mean values in the vacuum\nva\n1,va\n2,va\n3is a more plausible way to make the gauge\nbosonsLab\nimassive.IX. TESTS OF THE INVERSE-SQUARE LAW\nIn the static limit, the exchange of six Lorentz bosons\nLab\niof massmLwould imply that two macroscopic bod-\nies ofFandF0fermions separated by a distance rwould\ncontribute to the energy a static Yukawa potential\nVL(r) =3FF0f2\n2\u0019re\u0000mLr: (74)\nThis potential is positive and repulsive (between fermions\nand between antifermions) because the L's are vector\nbosons. It violates the weak equivalence principle because\nit depends upon the number Fof fermions (minus the\nnumber of antifermions) as F= 3B+Land not upon\ntheir masses. The potential VL(r) changes Newton's po-\ntential to\nVNL(r) =\u0000Gmm0\nr\u0010\n1 +\u000be\u0000r=\u0015\u0011\n(75)\nin which the coupling strength \u000bis\n\u000b=\u00003FF0f2\n2\u0019Gmm0=\u00003FF0m2\nPf2\n2\u0019mm0; (76)\nand the length \u0015is\u0015=~=cmL. Couplings \u000b\u00181 are\nof gravitational strength. Experiments [7{30] that test\nthe inverse-square law and the weak equivalence principle\nhave put upper limits on the strength j\u000bjof the coupling\nfor a wide range of lengths 10\u00008< \u0015 < 1013m and\nmasses 2\u000210\u0000203:4\nbj\u000bj\u0002106t0 (93)10\nwhich forbj\u000bj<1:7\u0002106exceeds the age t0of the\nuniverse. The range \u0015L=h=mLcof the corresponding\nYukawa potential is\n\u0015L>4:5\u000210\u00007m: (94)\nPoints (\u0015;j\u000bj) below the dotted green line in Fig. 6 and\nbetween its vertical dashed blue-green lines denote L\nbosons constituting between 1 and 100% of the dark mat-\nter. The upper limit on the e\u000bective mass of the electron\nneutrino is m(e\u000b)\n\u0017e<1:1 eV [33]. The thin gray vertical\nline labelsLbosons of mass mL= 2me\u000b\n\u0017e.\nXI. CONCLUSIONS\nGeneral relativity with fermions has two indepen-\ndent symmetries: general coordinate invariance and local\nLorentz invariance. General general coordinate invariance\nacts on coordinates and on the world indexes of tensors\nbut leaves Dirac and Lorentz indexes unchanged. Local\nLorentz invariance acts on Dirac and Lorentz indexes but\nleaves world indexes and coordinates unchanged. It acts\nlike an internal symmetry.\nGeneral coordinate invariance is implemented by the\nLevi-Civita connection \u0000j\nkiand by Cartan's tetrads ca\ni.\nIn the standard formulation of general relativity with\nfermions, local Lorentz invariance is implemented by the\nsame \felds in a combination called the spin connection\n!ab\ni=ca\njcbk\u0000j\nki+ca\nk@icbk. These \felds all have the\nsame action, the Einstein-Hilbert action R.\nBecause local Lorentz invariance is di\u000berent from and\nindependent of general coordinate invariance, it is sug-\ngested in this paper that local Lorentz invariance is im-\nplemented by di\u000berent and independent \felds Lab\nithat\ngauge the Lorentz group and that have their own Yang-\nMills-like action.\nThe replacement of the spin connection with Lorentz\nbosons moves general relativity closer to gauge theory\nand simpli\fes the standard covariant derivative\n\u0010\n@i+1\n8\u0000\nca\njcbk\u0000j\nki+ca\nk@icbk\u0001\n[\ra;\rb]\u0011\n (95)\nto\n\u0000\n@i+1\n8Lab\ni[\ra;\rb]\u0001\n : (96)Whether the Dirac action has the spin-connection form\n(95) or the Lorentz-boson form (96) is an experimental\nquestion.\nBecause the proposed action (15) couples the gauge\n\feldsLab\nito fermion number and not to mass, it violates\nthe weak equivalence principle. It also leads to a Yukawa\npotential (74) that violates Newton's inverse-square law.\nExperiments [7{30] have put upper limits on the\nstrengthj\u000bjof the Yukawa potentials (75) that violate the\ninverse-square law and the weak equivalence principle for\ndistances 10\u00008<\u0015< 109m. The upper limit ranges from\nj\u000bj<1019at\u0015= 10\u00008m toj\u000bj<103at\u0015= 10\u00005m and\ntoj\u000bj<10\u000011at\u0015= 109m. There are no experimental\nlower limits on the coupling at any distance, so Lbosons\ncould have lifetimes that exceed the age of the universe.\nThere are no experimental upper limits on the masses of\nLbosons. Long lived, massive, weakly interacting, neu-\ntralLbosons would contribute to dark matter. From the\nobvious requirement that they could make up all of dark\nmatter but not more, we can infer a crude theoretical\nupper limit on their mass of mL.2:8 eV=c2if all 6 are\nstable and have the same mass.\nThe discovery of a violation of the inverse-square law\nby future experiments would not be enough to establish\nthe existence of Lbosons because the violation could be\ndue to the physics of a quite di\u000berent theory.\nIfLbosons are discovered, physicists will decide how\nto think about the force they mediate. The force might\nbe considered to be gravitational because it arises in a\ntheory that is a modest and natural extension of general\nrelativity. But the force is not carried by gravitons. It is\ncarried by Lbosons, and they implement a symmetry,\nlocal Lorentz invariance, that is independent of general\ncoordinate invariance. So the force is new and might be\ncalled a Lorentz force.\nACKNOWLEDGMENTS\nI am grateful to E. Adelberger, R. Allahverdi, D.\nKrause, E. Fischbach, and A. Zee for helpful email.\n[1] R. Utiyama, Invariant theoretical interpretation of inter-\naction, Phys. Rev. 101, 1597 (1956).\n[2] T. W. B. Kibble, Lorentz invariance and the gravitational\n\feld, J. Math. Phys. 2, 212 (1961).\n[3] S. Weinberg, Gravitation and Cosmology (John Wiley\nand Sons, New York, 1972) pp. 365{373.\n[4] S. Deser and C. J. Isham, Canonical Vierbein Form of\nGeneral Relativity, Phys. Rev. D14, 2505 (1976).\n[5] L. E. Parker and D. Toms, Quantum Field Theory inCurved Spacetime (Cambridge University Press, Cam-\nbridge, UK, 2009).\n[6] K. Cahill, General internal gauge symmetry, Phys. Rev.\nD18, 2930 (1978); Soluble gauge theory of a noncompact\ngroup, Phys. Rev. D 20, 2636 (1979); Nonlinear internal\nsymmetry, J. Math. Phys. 21, 2676 (1980); K. Cahill and\nS.Ozenli, Unitary gauge theories of noncompact groups,\nPhys. Rev. D 27, 1396 (1983).\n[7] B. W. Harris, F. Chen, and U. Mohideen, Precision11\nmeasurement of the Casimir force using gold surfaces,\nPhys. Rev. A62, 052109 (2000), arXiv:quant-ph/0005088\n[quant-ph].\n[8] Y.-J. Chen, W. Tham, D. Krause, D. Lopez, E. Fis-\nchbach, and R. Decca, Stronger Limits on Hypotheti-\ncal Yukawa Interactions in the 30{8000 nm Range, Phys.\nRev. Lett. 116, 221102 (2016), arXiv:1410.7267 [hep-ex].\n[9] J. Lee, E. Adelberger, T. Cook, S. Fleischer, and\nB. Heckel, New Test of the Gravitational 1 =r2Law at\nSeparations down to 52 \u0016m, Phys. Rev. Lett. 124, 101101\n(2020), arXiv:2002.11761 [hep-ex].\n[10] W.-H. Tan, A.-B. Du, W.-C. Dong, S.-Q. Yang, C.-G.\nShao, S.-G. Guan, Q.-L. Wang, B.-F. Zhan, P.-S. Luo,\nL.-C. Tu, and J. Luo, Improvement for Testing the Gravi-\ntational Inverse-Square Law at the Submillimeter Range,\nPhys. Rev. Lett. 124, 051301 (2020).\n[11] J. Berg\u0013 e, P. Brax, G. M\u0013 etris, M. Pernot-Borr\u0012 as,\nP. Touboul, and J.-P. Uzan, MICROSCOPE Mission:\nFirst Constraints on the Violation of the Weak Equiv-\nalence Principle by a Light Scalar Dilaton, Phys. Rev.\nLett.120, 141101 (2018), arXiv:1712.00483 [gr-qc].\n[12] W.-H. Tan, S.-Q. Yang, C.-G. Shao, J. Li, A.-B. Du, B.-\nF. Zhan, Q.-L. Wang, P.-S. Luo, L.-C. Tu, and J. Luo,\nNew Test of the Gravitational Inverse-Square Law at the\nSubmillimeter Range with Dual Modulation and Com-\npensation, Phys. Rev. Lett. 116, 131101 (2016).\n[13] S.-Q. Yang, B.-F. Zhan, Q.-L. Wang, C.-G. Shao, L.-C.\nTu, W.-H. Tan, and J. Luo, Test of the gravitational in-\nverse square law at millimeter ranges, Phys. Rev. lett.\n108, 081101 (2012).\n[14] E. Adelberger, J. Gundlach, B. Heckel, S. Hoedl, and\nS. Schlamminger, Torsion balance experiments: A low-\nenergy frontier of particle physics, Prog. Part. Nucl.\nPhys. 62, 102 (2009).\n[15] A. A. Geraci, S. J. Smullin, D. M. Weld, J. Chi-\naverini, and A. Kapitulnik, Improved constraints on non-\nNewtonian forces at 10 microns, Phys. Rev. D 78, 022002\n(2008), arXiv:0802.2350 [hep-ex].\n[16] D. Kapner, T. Cook, E. Adelberger, J. Gundlach, B. R.\nHeckel, C. Hoyle, and H. Swanson, Tests of the gravita-\ntional inverse-square law below the dark-energy length\nscale, Phys. Rev. Lett. 98, 021101 (2007), arXiv:hep-\nph/0611184.\n[17] S. Smullin, A. Geraci, D. Weld, J. Chiaverini, S. P.\nHolmes, and A. Kapitulnik, Constraints on Yukawa-type\ndeviations from Newtonian gravity at 20 microns, Phys.\nRev. D 72, 122001 (2005), [Erratum: Phys.Rev.D 72,\n129901 (2005)], arXiv:hep-ph/0508204.\n[18] C. Hoyle, D. Kapner, B. R. Heckel, E. Adelberger,\nJ. Gundlach, U. Schmidt, and H. Swanson, Sub-\nmillimeter tests of the gravitational inverse-square law,\nPhys. Rev. D 70, 042004 (2004), arXiv:hep-ph/0405262.\n[19] J. Long, H. Chan, A. Churnside, E. Gulbis, M. Var-\nney, and J. Price, Upper limits to submillimetre-range\nforces from extra space-time dimensions, Nature 421, 922\n(2003).\n[20] J. Chiaverini, S. Smullin, A. Geraci, D. Weld, and A. Ka-\npitulnik, New experimental constraints on nonNewtonian\nforces below 100 microns, Phys. Rev. Lett. 90, 151101\n(2003), arXiv:hep-ph/0209325.[21] J. Hoskins, R. Newman, R. Spero, and J. Schultz, Exper-\nimental tests of the gravitational inverse square law for\nmass separations from 2-cm to 105-cm, Phys. Rev. D 32,\n3084 (1985).\n[22] J. G. Williams, S. G. Turyshev, and D. H. Boggs,\nProgress in lunar laser ranging tests of relativistic gravity,\nPhys. Rev. Lett. 93, 261101 (2004), arXiv:gr-qc/0411113\n[gr-qc].\n[23] E. Adelberger, B. R. Heckel, and A. Nelson, Tests of the\ngravitational inverse square law, Ann. Rev. Nucl. Part.\nSci.53, 77 (2003), arXiv:hep-ph/0307284.\n[24] M. V. Moody and H. J. Paik, Gauss's law test of gravity\nat short range, Phys. Rev. Lett. 70, 1195 (1993).\n[25] R. Spero, J. K. Hoskins, R. Newman, J. Pellam, and\nJ. Schultz, Test of the Gravitational Inverse-Square Law\nat Laboratory Distances, Phys. Rev. Lett. 44, 1645\n(1980).\n[26] S. Schlamminger, K. Y. Choi, T. A. Wagner, J. H. Gund-\nlach, and E. G. Adelberger, Test of the equivalence prin-\nciple using a rotating torsion balance, Phys. Rev. Lett.\n100, 041101 (2008), arXiv:0712.0607 [gr-qc].\n[27] R. Decca, D. Lopez, H. Chan, E. Fischbach, D. Krause,\nand C. Jamell, Constraining new forces in the Casimir\nregime using the isoelectronic technique, Phys. Rev. Lett.\n94, 240401 (2005), arXiv:hep-ph/0502025.\n[28] L.-C. Tu, S.-G. Guan, J. Luo, C.-G. Shao, and L.-X. Liu,\nNull Test of Newtonian Inverse-Square Law at Submil-\nlimeter Range with a Dual-Modulation Torsion Pendu-\nlum, Phys. Rev. Lett. 98, 201101 (2007).\n[29] S.-Q. Yang, B.-F. Zhan, Q.-L. Wang, C.-G. Shao, L.-C.\nTu, W.-H. Tan, and J. Luo, Test of the Gravitational In-\nverse Square Law at Millimeter Ranges, Phys. Rev. Lett.\n108, 081101 (2012).\n[30] E. Fischbach and C. L. Talmadge, The Search for Non-\nNewtonian Gravity (New York, USA: Springer, 1999).\n[31] A. N. Bernal, B. Janssen, A. Jimenez-Cano, J. A. Ore-\njuela, M. Sanchez, and P. Sanchez-Moreno, On the\n(non-)uniqueness of the Levi-Civita solution in the Ein-\nstein{Hilbert{Palatini formalism, Phys. Lett. B768 , 280\n(2017), arXiv:1606.08756 [gr-qc].\n[32] K. Cahill, Physical Mathematics (Cambridge University\nPress, 2019) pp. 435{447, 2nd ed.\n[33] P. Zyla, The review of particle physics, Prog. Theor. Exp.\nPhys. 2020 , 083C01 (2020), neutrino Masses, Mixing,\nand Oscillations.\n[34] D. Clowe, M. Bradac, A. H. Gonzalez, M. Markevitch,\nS. W. Randall, C. Jones, and D. Zaritsky, A direct em-\npirical proof of the existence of dark matter, Astrophys.\nJ. Lett. 648, L109 (2006), arXiv:astro-ph/0608407 [astro-\nph].\n[35] S. Weinberg, Cosmology (Oxford University Press, 2010)\np. 186.\n[36] N. Aghanim et al. (Planck), Planck 2018 results. VI. Cos-\nmological parameters (2018), arXiv:1807.06209 [astro-\nph.CO].\n[37] K. Cahill, Physical Mathematics (Cambridge University\nPress, 2019) p. 518, 2nd ed.\n[38] S. Weinberg, Cosmology (Oxford University Press, 2010)\np. 152." }, { "title": "2005.05725v2.Effective_Viscous_Damping_Enables_Morphological_Computation_in_Legged_Locomotion.pdf", "content": "Effective Viscous Damping Enables\nMorphological Computation in Legged\nLocomotion\nAn Mo1, Fabio Izzi1;2Daniel F. B. Haeufle2and Alexander Badri-Spr ¨owitz1;\u0003\n1Dynamic Locomotion Group, Max Planck Institute for Intelligent Systems, Stuttgart,\nGermany\n2Hertie-Institute for Clinical Brain Research, University of T ¨ubingen, T ¨ubingen,\nGermany\nCorrespondence*:\nAlexander Badri-Spr ¨owitz\nsprowitz@is.mpg.de\nABSTRACT\nMuscle models and animal observations suggest that physical damping is beneficial for\nstabilization. Still, only a few implementations of physical damping exist in compliant robotic\nlegged locomotion. It remains unclear how physical damping can be exploited for locomotion\ntasks, while its advantages as sensor-free, adaptive force- and negative work-producing actuators\nare promising. In a simplified numerical leg model, we studied the energy dissipation from\nviscous and Coulomb damping during vertical drops with ground-level perturbations. A parallel\nspring- damper is engaged between touch-down and mid-stance, and its damper auto-decouples\nfrom mid-stance to takeoff. Our simulations indicate that an adjustable and viscous damper is\ndesired. In hardware we explored effective viscous damping and adjustability, and quantified the\ndissipated energy. We tested two mechanical, leg-mounted damping mechanisms: a commercial\nhydraulic damper, and a custom-made pneumatic damper. The pneumatic damper exploits a\nrolling diaphragm with an adjustable orifice, minimizing Coulomb damping effects while permitting\nadjustable resistance. Experimental results show that the leg-mounted, hydraulic damper exhibits\nthe most effective viscous damping. Adjusting the orifice setting did not result in substantial\nchanges of dissipated energy per drop, unlike adjusting the damping parameters in the numerical\nmodel. Consequently, we also emphasize the importance of characterizing physical dampers\nduring real legged impacts to evaluate their effectiveness for compliant legged locomotion.\nKeywords: damping, energy dissipation, legged locomotion, ground disturbance, drop test, rolling diaphragm\n1 INTRODUCTION\nWhile less understood, damping likely plays an essential role in animal legged locomotion. Intrinsic\ndamping forces can potentially increase the effective force output during unexpected impacts (M ¨uller et al.,\n2014), reduce control effort (Haeufle et al., 2014), stabilize movements (Shen and Seipel, 2012; Secer and\nSaranli, 2013; Abraham et al., 2015), and reject unexpected perturbations (Haeufle et al., 2010; Kalveram\net al., 2012), e.g., sudden variations in the ground level (Figure 1). Stiffness, in comparison, has been\nstudied extensively in legged locomotion. Its benefits have been shown both in numerical simulations, e.g.,\nthrough spring-loaded inverted pendulum (SLIP) models (Mochon and McMahon, 1980; Blickhan et al.,\n1arXiv:2005.05725v2 [cs.RO] 6 Jun 2020Mo et al. Effective damping in legged locomotion\nFigure 1. (A-C) Problem identification, and related research question. The limited nerve conduction\nvelocity in organic tissue (More et al., 2010) 2presents a significant hazard in legged locomotion.\nLocal neuromuscular strategies 6provide an alternative means of timely and tunable force and power\nproduction . Actuators like the indicated knee extensor muscle keep the leg extended during stance phase\n(muscle length L muscle ) by producing the appropriate amount of muscle force (F muscle ), correctly timed .\nNeuromuscular control 1plays a major role in initiating and producing these active muscle forces, but\nworks best only during unperturbed locomotion. Sensor information from foot contact travels via nerves\nbundles 2to the spinal cord, but with significant time delays in the range of t=40ms (More and Donelan,\n2018, for 1 mleg length) and more. Hence, the locomotion control system can become ‘sensor blind’ due\nto conduction delays, for half a stance phase , and can miss unexpected perturbations like the depicted\nstep-down. During step-down perturbations 3additional energy 4is inserted into the system. Viscous\ndamper-like mechanisms produce velocity dependent counter-forces, and can dissipate kinetic energy .\nLocal neuromuscular strategies 6producing tunable, viscous damping forces would act instantaneously\nand adaptively . Such strategies 6could also be robust to uncontrolled and harsh impacts of the foot after\nperturbations 5, better than sensor-based strategies. In this work (D), we are testing and characterizing\nspring-damper configurations mounted to a two-segment leg structure, during rapid- and slow-drop\nexperiments, for their feasibility to physically and instantaneously produce tunable, speed-dependent forces\nextending the leg. Work loops (E)will indicate how much effective negative work is dissipated, between\ntouch-down and mid-stance. Prior to impact 7and during the leg loading 8the spring-damper’s tendons\nact equally. Starting at mid-stance, the main spring extends the knee, leading to leg extension and leaving\nthe damper’s tendon slack 9.\n2007), and physical springy leg implementations (Hutter et al., 2016; Spr ¨owitz et al., 2013; Ruppert and\nBadri-Spr ¨owitz, 2019).\nWhat combines both mechanical stiffness and intrinsic, mechanical damping is their sensor- and\ncomputational-free action. A spring-loaded leg joint starts building up forces exactly at the moment\nof impact. Mechanical stiffness, or damping, acts instantaneously, and are not subject to delays from post-\nprocessing sensor data (Grimminger et al., 2020), delays from limited nerve conductive velocities (More\nand Donelan, 2018), or uncertainties in the estimation of the exact timing of swing-to-stance switching\n(Bledt et al., 2018).\nLegged robots commonly exploit virtual damping : actively produced and sensory-controlled negative\nwork in the actuators (Seok et al., 2015; Hutter et al., 2012; Havoutis et al., 2013; Kalouche, 2017;\nGrimminger et al., 2020). Virtual damping requires high-frequency force control, and actuators\nmechanically and electrically capable of absorbing peaks in negative work. In comparison, mechanical\ndamping based systems (Hu et al., 2019; Garcia et al., 2011) act instantaneously, share impact loads with\nthe actuator when in parallel configuration, and require no sensors or control feedback. The instantaneous\nmechanical response of a damper is especially relevant in biological systems, where the neuronal delay\nmay be as large as 5 %to40 % of the duration of a stance phase (More et al., 2010). In such a short\nThis is a provisional file, not the final typeset article 2Mo et al. Effective damping in legged locomotion\ntime-window, physical damping could help to reject the perturbation (Haeufle et al., 2010; Kalveram et al.,\n2012) by morphological computation, as it mechanically contributes to the rejection of the perturbation,\na contribution that otherwise would need to be achieved by a (fast) controller (Zahedi and Ay, 2013;\nGhazi-Zahedi et al., 2016). Hence, physical damping has the potential to contribute to the morphological\ncomputation (Zahedi and Ay, 2013; Ghazi-Zahedi et al., 2016) of a legged system.\nCompared to virtual damping with proprioceptive sensing strategies (Grimminger et al., 2020), a legged\nrobot with physical damping requires additional mechanical components, e.g., a fluidic cylinder, and the\nmechanics to convert linear motion to rotary output. In a cyclic locomotion task, the energy removed by\nany damper must also be replenished. Hence, from a naive energetic perspective, both virtual and physical\ndamping systems are costly.\nEnergy dissipation in the form of negative work has been quantified in running birds, and identified as a\npotential strategy to ‘... reduce the likelihood of a catastrophic fall.’ (Daley and Biewener, 2006, p. 185).\nIn virtual point-based control strategies for bipedal running, positive work is inserted into hip joints, and\nnegative work is then dissipated in equal amounts in the spring-damper leg (Drama and Spr ¨owitz, 2020).\nIn sum, either physical damping or virtual damping allows removing energy from a legged locomotion\nsystem. In this work, we focus on physical damping produced by a viscous damper. We aim towards an\nunderstanding of how physical damping can be exploited in legged locomotion and which requirements a\ndamper must fulfill.\nWe consider two damping principles: viscous damping and Coulomb damping. Viscous damping reacts\nto a system motion with a force that is linearly (or non-linearly) proportional to its relative acting speed.\nCoulomb damping generates a constant force, largely independent from its speed (Serafin, 2004). From a\ncontrol perspective, viscous damping can be beneficial for the negotiation of perturbations in locomotion\nas it approximates the characteristics of a differential, velocity-dependent term. Yet it is unknown how this\nintuition transfers into reality, where impact dynamics and non-linearities of the leg geometry alter the\nstance-phase dynamics of locomotion.\nDamping in legged locomotion can have other purposes, besides dissipating energy. The authors of\n(Werner et al., 2017, p. 7) introduced a damping matrix in the control scheme, which reduced unwanted\noscillations in the presence of modeling errors. Tsagarakis et al. (2013) mount compliant elements with\nsome damping characteristics, which also could reduce oscillations of the system’s springy components.\nIn this project, we focus our investigation on the effect of damping during the touch-down (impact)\nand mid-stance. We chose this simpler drop-down scenario as it captures the characteristics of roughly\nhalf a locomotion cycle. A complete cycle would require an active push off phase, and the leg’s swing\ndynamics. Hence, we study the effectiveness of physical damping on the leg’s energy dissipation within\none drop (touch-down to lift-off), by quantifying its effective dissipated energy Eeffective . We combine\ninsights from numerical simulations and hardware experiments (Figure 2). By studying the response\nof two damping strategies (viscous and Coulomb damping) in numerical drop-down simulations, we\ninvestigate how physical damping can influence the dynamics of the impact phase. We then examine\nhow these theoretical predictions relate to hardware experiments with two functionally different, physical\ndampers. Hence we explore and characterize the physical damper implementations in a robot leg for their\neffectiveness in drop-impacts.\nFrontiers 3Mo et al. Effective damping in legged locomotion\nHardware experiment: leg dropviscous vs. Coulomb friction\nvariation: ground level heightNumerical simulation: leg drop\nground level variation: valve weigh\nhydraulic diaphragm\n spring\nvariation:\nHardware experiment: hydraulic damper drop\nheight setting\ndamper only damper\nvalve\nsettingvalve\nsetting\n1\n 2\n3\n 4\n 5\n6\n 7\n 8\nFigure 2. Overview: We study the effective dissipated energy Eeffective in drop experiments, i.e., the energy\ndissipation within one drop cycle between touch-down and lift-off (Figure 6). We focus on a system\ndesign with a damper and a spring, both acting in parallel on the knee joint (Figure 1 Eand Figure 3).\nNo active motor is considered as it is not relevant for the drop scenario, but required for continuous\nhopping. In numerical simulations , we quantify the difference in energy dissipation between viscous 1\nand Coulomb damping for varying ground level heights 2(Section 2 and Figure 4). The first set of\nhardware experiments characterizes the industrial hydraulic damper. For this, we drop the isolated\ndamper (damper only, not mounted in the leg) on a force sensor and calculate the energy dissipation.\nWe vary the ground level height 3, the valve setting 4and the drop mass 5, to investigate its dynamic\ncharacteristics (Section 4.1 and Figure 7). For the second set of hardware experiments, we drop a\n2-segment leg with dampers mounted in parallel to knee springs . We investigate the energy dissipation\ndynamics of the hydraulic 6and diaphragm damper 7by comparing it to a spring-only condition 8, where\nthe damper cable is simply detached (Section 4.2 and Figure 8). We also vary the valve setting on the\ndampers to test the dynamic adjustability of damping (Section 4.3 and Figure 9).\n2 NUMERICAL SIMULATION\nWe use numerical simulations to theoretically investigate the energy dissipation in a leg drop scenario\n(Figure 2). In analogy to our hardware experiment (Section 3.3), a 2-segment leg with a damper and a\nspring in parallel on the knee joint is dropped vertically (Figure 3a). Once in contact with the ground, the\nknee flexes and energy is dissipated. We compare viscous vs. Coulomb damping to investigate which of\nthese two theoretical damping strategies may be more suited for the rejection of ground-level perturbations.\nAlso, we investigate how the adjustment of the damping characteristics influences the dissipated energy.\nIn all the damping scenarios investigated, the system is not energy conservative. As we investigate the\npotential benefit of damping in the initial phase of the ground contact, i.e., from touch-down to mid-stance,\nwe do not consider any actuation. Without actuation or control, the model’s dissipated energy is not refilled,\nunlike in, for example, periodic hopping (Kalveram et al., 2012).\nThis is a provisional file, not the final typeset article 4Mo et al. Effective damping in legged locomotion\nl0 = \nd\nk\nmβ\nλ1\nλ\n2\nFlegg\ny(t)α\nrkrd\n24.5cm\n(a)Leg model\n⑥\n⑦\n⑧\n⑨\n⑩\n⑪\n⑫⑬⑭⑮\n⑯ (b)Leg design\n (c)Drop test bench\nFigure 3. (a) 2-segment spring-damper-loaded leg model used for simulation. (b)Mechanical design of\nthe 2-segment leg. The knee pulley 11\ris fixed with the lower segment 12\r, coupled with the spring 8\rand the\ndiaphragm damper 15\ror hydraulic damper 16\rvia cables 9\r10\r.(c)Drop test bench with the 2-segment leg.\n2.1 Model\nThe numerical model is a modified version of the 2-segment leg proposed in Rummel and Seyfarth\n(2008) with an additional damper mounted in parallel to the knee-spring. The equation describing our leg\ndynamics is:\n¨y(t) =Fleg(t)\nm\u0000g (1)\nwhere gis the gravitational acceleration, mis the leg mass (lumped at the hip), and y(t)is the time-\ndependent vertical position from the ground. Fleg(t)is the force transmitted to the hip mass - and the\nground - through the leg structure. As such, the force depends on the current phase of the hopping cycle:\nFleg(t) =8\n<\n:0 , flight phase: y(t)>l0\ny(t)\nl1l2t(t)\nsin(b(t)), ground contact: y(t)\u0014l0(2)\nwith segment length liand knee angle b(t)(Figure 3a), l0is the leg length at impact. t(t)is the knee\ntorque which is produced by the parallel spring-damper element, as in\nt(t) =k r2\nk(b(t)\u0000b0)+td(t) (3)\nwith kandrkbeing the spring stiffness coefficient and lever arm, respectively. td(t)is the damping torque,\nwhich is set to zero during leg extension, i.e., the damper is only active from impact to mid-stance:\ntd(t) =0 if ˙b(t)>0 (4)\nThe modeled damper becomes inactive during leg extension, in accordance to our hardware: the tested\nphysical dampers apply forces to the knee’s cam via a tendon (Figure 1d, 9), and this tendon auto-decouples\nduring leg extension. By choosing different definitions of the damper torque td(t), we can analyse different\ndamper concepts. The model parameters are listed in Table 1.\nFrontiers 5Mo et al. Effective damping in legged locomotion\nSimulations were performed using MATLAB (the MathWorks, Natick, MA) with ODE45 solver (absolute\nand relative tolerance of 10−5, max step size of 10−5s). When searching for appropriate settings of the\nnumerical solver, we progressively reduced error tolerances and the maximum step size until convergence\nof the simulation results in Table 2 to the first non-significant digit.\nTable 1. Simulation and hardware parameters\nParameters Symbol Value Unit\nMass m 0.408 kg\nReference drop height h0 14 cm\nSpring stiffness k 5900 N/m\nLeg segment length l1;l2 15 cm\nLeg resting length l0 24.6 cm\nKnee resting angle b0 110 deg\nSpring lever arm rk 2.5 cm\nDamper lever arm rd 2 cm\n2.2 Damping characteristics\nWe compared two damping concepts in our numerical simulation: (1) pure Coulomb damping, i.e., a\nconstant resistance only dependent on motion direction, and pure viscous damping, i.e., a damper torque\nlinearly dependent on the knee angular velocity. Accordingly, we tested two different definitions of td:\ntd(t) =8\n<\n:\u0000dcrdsign(˙b(t)) , pure Coulomb damping\n\u0000dvr2\nd˙b(t) , pure viscous damping(5)\nwhere rdis the damper level arm, dc(inN) and dv(inNs/m ) the Coulomb damping and viscous damping\ncoefficients, respectively.\n2.3 Energy dissipation in numerical drop simulations\nWith this model, we investigate the difference in energy dissipation in response to step-up/down\nperturbations (cases 1and 2in Figure 2). For each drop test, the numerically modeled leg starts at\nrest ( ˙ y(t) =0) with a drop height\nh=y(t=0)\u0000l0 (6)\ncorresponding to the foot clearance at release. The total energy at release is ET(h) =mgh . Given that\nall model parameters in Table 1 are fixed, the energy dissipated in a drop becomes a function of the drop\nheight and the damping coefficients: ED=fED(h;dc;v).\nA simulated drop height hcan be seen as a variation Dhfrom a reference value h0:\nh=h0\u0006Dh (7)\nEqual to the hardware experiments, we use h0=14cm as reference drop height. In the reference drop\ncondition, i.e., h=h0, the energy dissipated by damping is ED0=ED(h0) = fED(h0;dc;v).ED0only\ndepends on the damping level, namely the chosen damping strategy (viscous or Coulomb damping) and\nassociated damping coefficient. We chose five different desired damping levels (set 1-5) as a means of\nThis is a provisional file, not the final typeset article 6Mo et al. Effective damping in legged locomotion\nscanning a range in which the damping could be adjusted: for each set, the amount of energy that is\ndissipated at the reference drop height ED0differs. The chosen ED0values (Table 2, column “Reference\nheight”) correspond to proportional levels ( [0:1;0:2;:::; 0:5]) of the systems potential energy in terms of\nthe leg resting length l0, as in\nED0=mg[0:1;0:2;:::; 0:5]l0 (8)\nThis corresponds to damping configurations that dissipate between \u001917% and\u001988% of the system’s\ninitial potential energy at the reference height ( ET0=ET(h0) =mgh 0=560mJ ), as shown in Table 2,\ncolumn “Reference height”. To achieve these desired damping levels, we adjusted the damper parameters\ndcanddvaccordingly (Table 2, column “Damping coeff.”). As an example: for set 3, both damping values\nwere adjusted such that at the reference height h0both dampers dissipate ED0=mg0:3l0=295mJ , which\ncorresponds to 53 % of the total energy ET0.\nIn the numerical simulations, we focus on the relation between a ground level perturbation Dhand the\nchange in energy dissipation – and their dependency on the damper characteristics. A drop from a height\nlarger than h0corresponds to a step-down ( Dh>0), and a drop from a height smaller than h0to a step-up\n(Dh<0). Each condition introduces a change of the total energy of DET=mgDh. The change in energy\ndissipation due to the perturbation is defined as\nDED(Dh) =ED(h0+Dh)\u0000ED0 (9)\nwhich is the difference between the dissipated energy when released from a perturbed height and the\ndissipated energy when released from the reference height. As a reference, we further define the full\nrejection case where\nDED(Dh) =DET=mgDh (10)\nIn human hopping a full recovery within a single hopping cycle is not seen during experimental drop down\nperturbations. Instead, a perturbation of Dh=0:1l0is rejected in two to three hopping cycles (Kalveram\net al., 2012, Figure 7a). In our results, this corresponds to the partial rejections observed with sets 2and3\nforDh=\u00062:5cm.\n2.4 Simulation results\nFigure 4a shows the relation between the change in drop height and the corresponding change in dissipated\nenergy by the simulated dampers for set 1,3and5(continuous line for pure viscous, dashed for pure\nCoulomb damping). For the range of simulated drop heights, pure viscous and Coulomb dampers change\nthe amount of dissipated energy with an almost linear dependence on the drop height. However, pure\nviscous damping has a slope closer to the full rejection scenario (blue line in Figure 4a), regardless of the\nset considered. In a step-down perturbation ( Dh>0in Figure 4a), pure viscous damping dissipates more\nof the additional energy DET, while in a step-up perturbation ( Dh<0) it dissipates less energy than pure\nCoulomb damping. As such, the results show that a viscous damper can reject a step-down perturbation\nfaster, e.g., within less hopping cycles, and it requires smaller correction by the active energy supply during\na step-up perturbation.\nAdjusting the damping parameters allows to change the reaction to a perturbation (Figure 4). Increasing\nthe damping intensity, i.e., dvanddcfrom set 1to5, allows to better match the full recovery behaviour (blue\nline in Figure 4a). However, this comes at the cost of a higher energy dissipation at the reference height,\nFrontiers 7Mo et al. Effective damping in legged locomotion\nTable 2. Numerical simulation Total dissipated energy ( ED) in one drop cycle for different drop heights\n(h).Reference height is the reference drop height h=h0=14cm . During step up(down) condition, the\ndrop height is reduced(increased) by Dh=2:5cm. Percentage values indicate the change in dissipated\nenergy ( DED) relative to the change in system total energy ( DET) due to the height perturbations. Each\nset simulates two separate mechanical dampers (pure viscous or pure Coulomb damping), with damping\ncoefficients chosen to dissipate the same energy at the reference condition, i.e., ED0. Results of set 1,3\nand5are further described in Figure 4. For all tested conditions, viscous damping outperforms Coulomb\ndamping, as indicated by the always higher percentage values.\nDamping coeff.Step up Reference height Step down\nh=h0\u0000Dh=11:5cm h=h0=14cm h=h0+Dh=16:5cm\ndv dc ED(DED=DET) ED0(ED0=ET0) ED(DED=DET)\nSet 1Viscous 29.5 Ns/m 0 N 82 mJ (15%) 97 mJ (17%) 112 mJ (15%)\nCoulomb 0 Ns/m 7.7 N 88 mJ (9%) 97 mJ (17%) 104 mJ (7%)\nSet 2Viscous 68 Ns/m 0 N 167 mJ (30%) 197 mJ (35%) 227 mJ (30%)\nCoulomb 0 Ns/m 17.3 N 178 mJ (19%) 197 mJ (35%) 214 mJ (17%)\nSet 3Viscous 119.4 Ns/m 0 N 249 mJ (46%) 295 mJ (53%) 341 mJ (46%)\nCoulomb 0 Ns/m 29.3 N 264 mJ (31%) 295 mJ (53%) 323 mJ (28%)\nSet 4Viscous 197.1 Ns/m 0 N 330 mJ (63%) 393 mJ (70%) 455 mJ (62%)\nCoulomb 0 Ns/m 46.1 N 346 mJ (47%) 393 mJ (70%) 436 mJ (43%)\nSet 5Viscous 349.4 Ns/m 0 N 411 mJ (81%) 492 mJ (88%) 572 mJ (80%)\nCoulomb 0 Ns/m 76.3N 423 mJ (69%) 492 mJ (88%) 556 mJ (64%)\ni.e., in absence of a ground perturbation (Table 2, column ‘reference height’). Increasing the damping rate\nalso affects the energetic advantage of viscous damping over Coulomb damping. Figure 4b shows this\nin detail for a specific step down perturbation (Dh=2:5cm): from set 1to set 3, the spread between the\nDEDvalues of the viscous damper and the Coulomb damper increases (from 8 mJ to18 mJ ). However, the\ndifference in dissipated energy DEDslightly reduces from set 3 to set 5 (from 18 mJ to 16 mJ).\nTable 2 quantifies the previous findings by indicating the percentage of energy perturbation DETthat\neach damping approach dissipates for Dh=\u00062:5cm and for all the tested sets of damping coefficients dv\nanddc. The data further confirms the observations from Figure 4, showing that:\n1.within each set, viscous damping outperforms Coulomb damping for all the simulated conditions - its\ndissipated energy is always the closest to 100 % of DET, which means the closest to full rejection ;\n2.the energetic benefit of viscous damping over Coulomb damping, i.e., the spread in percentage values\nofDED=DET,does not monotonically increase with higher damping rates, i.e., moving from set 1 to 5.\nFurthermore, Table 2 shows that for small damping rates, i.e., set 1, viscous damping introduces only\nmarginal benefits in energy management compared to Coulomb damping: <10% spread between the\ncorresponding DED=DETvalues.\n3 HARDWARE DESCRIPTION\nWith the previous results from our numerical simulation in mind, we tested two technical implementations\n(Figure 5) to produce adjustable and viscous physical damping. We implemented a 2-segment leg hardware\n(Figure 3b) and mounted it to a vertical drop test bench to investigate the role of physical damping. The\ndrop test bench produces velocity profiles during impact and stance phase similar to continuous hopping\nand allows us testing effective damping efficiently and repeatable.\nThis is a provisional file, not the final typeset article 8Mo et al. Effective damping in legged locomotion\nFigure 4. Numerical simulation Cases 1and 2from Figure 2, (a): Change of total energy vs. change\nof drop height for set 1,3and5, with damping coefficients as in Table 2. Continuous lines are viscous\ndamping results, dashed Coulomb damping. Positive perturbations, i.e., Dh>0, correspond to step-down\nperturbations; step-up perturbations, otherwise. The steepest line indicates the slope needed for a full\nrejection of aDhdeviation. For each set (1, 3, 5), the damping parameters are matched such that viscous\nand Coulomb damping dissipate the same energy at the reference height h0(see Table 2). Within each\nset, the viscous damping line is closer to the desired full rejection line than the corresponding Coulomb\ndamping line. This means that for the same cost (in the sense of dissipated energy at the reference height)\nviscous damping always rejects more of ground level perturbation than Coulomb damping. (b):DED\nforDh=2:5cm. The horizontal line indicates the amount of energy to dissipate for full rejection of\nDh. Energetic advantage of viscous damping over Coulomb damping, as indicated by the spread in the\ncorresponding DEDvalues, increases from set 1 to 3, and reduces from set 3 to 5.\n3.1 Rolling Diaphragm Damper\nThe most common designs of viscous dampers are based on hydraulic or pneumatic cylinders (viscous\ndamping) and can offer the possibility of regulating fluid flow by altering the orifice opening (adjustability).\nThese physical dampers can display high Coulomb friction, caused by the mechanical design of the sliding\nseal mechanisms. Typically, the higher the cylinder pressure is, the higher the Coulomb friction exists.\nIdeally, we wanted to test one physical damper concept with the least possible amount of Coulomb friction.\nInspired by the low-friction hydrostatic actuators (Whitney et al., 2014, 2016), we designed a low-Coulomb\ndamper based on a rolling diaphragm cylinder. Its cylinder is 3D printed from Onyx material. Figure 5a\nillustrates the folding movement of this rolling diaphragm mounted on a piston. The rolling diaphragm\nis made of an elastomer shaped like a top hat that can fold at its rim. When the piston moves out, the\ndiaphragm envelopes the piston. In the ideal implementation, only rolling contact between the diaphragm\nand the cylinder occurs, and no sliding contact. Hence, Coulomb friction between piston and cylinder is\nminimized. We measured FC\u00190:3Nof Coulomb friction for our rolling diaphragm cylinder, at low speed.\nFrontiers 9Mo et al. Effective damping in legged locomotion\n①\n②\n(a)Diaphragm damper\n (b)Hydraulic damper\nFigure 5. (a) -left-top: schematic of a diaphragm damper, illustrating the motion of rolling diaphragm,\nwhich includes an adjustable orifice 1\r, a cylinder 2\r, a piston 3\r, and a rolling diaphragm 4\r.(b)-left:\nschematic of a hydraulic damper, fluid is sealed inside the cylinder 2\rwith an recovery spring 5\rto reset\nthe piston 3\r.\nOur numerical simulation results promoted viscous and adjustable damping for use in vertical leg-drop.\nBy concept, both properties are satisfied by the diaphragm damper with an adjustable valve. When an\nexternal load Fextpulls the damper piston (Figure 5a), the fluid inside the cylinder chamber flows through\na small orifice, adjustable by diameter. This flow introduces a pressure drop DP(t), whose magnitude\ndepends on the orifice cross-section area Aoand piston speed v(t). As such, for a given cylinder cross\nsection area Ap, the diaphragm damper reacts to an external load Fextby a viscous force Fp(t)due to the\npressure drop DP(t):\nFp(t) =ApDP(t) =Apf(v(t);Ao) (11)\nWe mounted a manually adjustable valve (SPSNN4, MISUMI) to set the orifice size Ao. For practical\nreasons (weight, leakage, complexity of a closed circuit with two cylinders) we used air in the diaphragm\ncylinder as the operating fluid, instead of liquid (Whitney et al., 2014, 2016). Air is compressible, and\nwith a fully closed valve the diaphragm cylinder also acts as an air spring. This additional functionality\ncan potentially simplify the overall leg design. With the pneumatic, rolling diaphragm-based damper\nimplementation, we focused on creating a light-weight, adjustable damper with minimal Coulomb friction,\nand air as operating fluid.\n3.2 Hydraulic Damper\nIn the second technical implementation we applied an off-the-shelf hydraulic damper (1214H or 1210M,\nMISUMI, Figure 5b), i.e., a commercially available solution for adjustable and viscous damping. Tested\nagainst other hydraulic commercial dampers, we found these specific models to have the most extensive\nrange of adjustable viscous damping and the smallest Coulomb friction ( FC\u00190:7N). Similarly to the\ndiaphragm damper, these hydraulic dampers produce viscous damping by the pressure drop at the adjustable\norifice. The operating fluid is oil, which is in-compressible. Hence, the hydraulic damper should not exhibit\ncompliant behavior. Other than the diaphragm damper, the hydraulic damper produces damping force when\nThis is a provisional file, not the final typeset article 10Mo et al. Effective damping in legged locomotion\nits piston is pushed, not pulled. This design also includes an internal spring to recover the piston position\nwhen unloaded. In sum, the hydraulic damper features high viscous damping, no air-spring effect, and a\nhigher Coulomb friction compared the custom-designed pneumatic diaphragm damper.\n3.3 Articulated Leg Design\nThe characteristics of a viscous damper strongly depend on the speed- and force-loading profile imposed\nat its piston, because of the complex interaction of fluid pressure and compression, viscous friction, and\ncavitation (Dixon, 2008). We implemented a hardware leg to test our two physical dampers at loading\nprofiles (speed, force) similar to legged hopping and running.\nThe 2-segment hardware leg (Figure 3b) is designed with a constant spring and damper lever arm,\nparameters are provided in Table 3. In all experiments with the 2-segmented leg, the leg spring provides\nelastic joint reaction forces. Dampers are swapped in and out in a modular fashion, depending on the\nexperimental settings. The 2-segment leg design parameters are identical to those in our simulation model\n(Table 1). A compression spring 8\ris mounted on the upper leg segment 13\r. When the leg flexes, the spring\nis charged by a spring cap 7\rcoupled to a cable 10\rattached to the lower leg. Either damper 15\r16\ris fixed on\na support 6\ron the upper segment 13\r. The support 6\rcan be moved within the upper segment 13\r, to adjust\nthe cable 9\rpretension. Cables 9\r10\rlink the damper piston 3\rand the spring 8\rto the knee pulley 11\r, which\nis part of the lower segment 12\r.\nDuring the leg flexion, the cable under tension transmits forces instantly to the spring and damper. Spring\nand damper forces counteract the knee flexion. During leg extension, the spring releases energy, while\nthe damper is decoupled due to slackness of the cable. We included a hard stop into the knee joint to\nlimit the maximum leg extension, and achieve a fixed leg length at impact. At maximum leg flexion at\nhigh leg loading, segments can potentially collide. We ensured not to hit either hard stops during the\ndrop experiments. The hydraulic damper 16\rrequires a reverse mechanism 14\r, since its piston requires\ncompression to work. The piston of the diaphragm damper 15\rwas directly connected to the knee pulley.\nThe diaphragm damper 15\rincluded no recovery spring 5\r, hence we reset the piston position manually after\neach drop test. In sum, different spring-damper combinations can be tested with the 2-segment leg setup.\nNote that the here shown hardware leg has no actuation. If a motor would actuate the knee joint, in parallel\nmounted to the spring and the damper, the damper would share the external impact load, and consequently\nreduce an impact at the motor.\n3.4 Experimental set-up, data sampling and processing\nWe implemented an experimental setup for repetitive measurements (Figure 3c). A drop bench was\nused to constrain the leg motion to a single vertical degree of freedom, and linear motion. This allowed\nus to fully instrument the setup (slider position, and vertical ground reaction forces, GRF), and ensured\nrepeatable conditions over trials. Adjusting the drop height allowed us setting the touch-down speed. A\nlinear rail (SVR-28, MISUMI) was fixed vertically on a frame. The upper leg segment was hinged to a rail\nslider. The rail slider was loaded with additional, external weights, simulating different robot masses. We\nset the initial hip angle a0to align the hip and foot vertically. A hard stop ensured that the upper leg kept a\nminimum angle a>a0.\nFrontiers 11Mo et al. Effective damping in legged locomotion\nTwo sensors measured the leg dynamics: the body position yand the vertical ground reaction force are\nrecorded by a linear encoder (AS5311, AMS) and a force sensor (K3D60a, ME, amplified with 9326 ,\nBurster), respectively (Figure 3c). The duration from touch-down to mid-stance is very short, typically\nt\u0014100ms , and high-frequency data sampling was required. The encoder data was sampled by Raspberry\nPi 3B+ with f=8kHz sampling rate. Force data were recorded by an Arduino Uno, with a 10-bit internal\nADC at 1 kHz sampling rate. A high-speed camera (Miro Lab 110, Phantom) recorded the drop sequence\natf=1kHz sampling rate. We performed ten trials for each test condition. Sensor data was processed with\nMATLAB (the MathWorks, Natick, MA). Data was smoothed with a moving average filter, with a filter\nspan of 35samples for encoder data, and 200samples for force data. Repeated experiments of the same\ntest condition are summarized as an envelop defined by the average \u0006the standard deviation of the filtered\nsignals.\n4 HARDWARE EXPERIMENTS AND RESULTS\nIn the drop experiments, we characterize both the hydraulic and diaphragm dampers, and the 2-segment\nspringy leg (Figure 6). We chose three orifice settings (labeled as a, b, and c) for each damper, and focus on\nthe effects of viscous damping and adjustable dissipation of energy in the hardware setup. Table 3 lists an\noverview of the drop tests, and its settings (drop height, weight, orifice setting, damper type). To emphasize\nthe fundamental differences between the damper designs, we compare only one model of the hydraulic\ndamper (1214H) to the diaphragm damper (Section 4.1 - Section 4.3), and show the potential of the second\nhydraulic damper (1210M) in Section 4.4. Videos of the experiments can be found in the supplementary\nmaterial, and online1.\nTable 3. Drop test settings for experiments\nDrop test setup Drop height Drop weight Orifice\nFig. [cm] [g] [ \u0018]\nDamper (1214H)7a 3, 5, 7 280 b\n7b 5 280 a, b, c\n7c 3 280, 620 b\nDamper (1214H, diaphragm) & leg8a, 8b 14 408 c\n8c 14 408 damper detached\nDamper & leg (simulation)9a, 9b 14 408 a, c\n9c 14 408 viscous, Coulomb\nDamper (1210M) & leg 10 14 408 a, b\n4.1 Isolated damper drops, evaluation\nIn this experiment we characterized the hydraulic damper by dropping it under changing conditions of\nthe instrumented drop setup, without mounting it to the 2-segment leg. The experimental setup allows\ndifferentiating effects, compared to the 2-segment leg setup, and to emphasize the viscous damper behavior\nof the off-the-shelf component. We also applied the results to estimate the range of damping rates available\nwith changing orifice settings. The hydraulic damper was directly fixed to the rail slider into the drop bench\n(Section 3.4). The piston pointed downwards. We measure the vertical ground reaction force to determine\nthe piston force, and we recorded the vertical position of the slider over time, to estimate the piston speed\nafter it touches the force sensor.\n1https://youtu.be/F00Sma2BQ4c\nThis is a provisional file, not the final typeset article 12Mo et al. Effective damping in legged locomotion\n80 85 90 95 100\nLeg length [%]0123GRF vertical [BW]80 85 90 95 100\nLeg length [%]0123GRF vertical [BW]\nHydraulic damper Diaphragm damperTouch-down 215 msMid-stance 271 ms\nLift-off 364 ms\nTouch-down 216 msMid-stance 280 ms\nLift-off 371 ms0 ms 215 ms 271 ms 364 ms 400 ms 441 ms\n0 ms 216 ms 280 ms 371 ms 417 ms 468 ms\nFigure 6. High-speed snapshots of drop experiments starting from release to second touchdown. Leg with\nhydraulic damper is shown on the top row, leg with diaphragm damper the bottom row. Depicted are from\nleft to right: release, touchdown, mid-stance, lift-off, apex, second touchdown. The right plots illustrate the\ntiming of the events corresponding to the snapshots.\nFigure 7 shows the force-speed profiles for drop tests with different drop heights (Figure 7a), orifice\nsettings (Figure 7b), and drop loads (Figure 7c). Data lines in Figure 7 should be interpreted from high\nspeed (impact, right side of each plot) to low speed (end of settling phase, 0 m/s , left). The time from\nimpact to peak force (right slope of each plot) is ( \u001924 ms ), while the negative work (shown in legends)\nwas mainly dissipated along the falling slope in the much longer-lasting settling phase after the peak (left\nslope of each plot, \u0019200 ms).\n0 0.2 0.4 0.6 0.8 1\nPiston speed [m/s]0204060GRF vertical [N]3 cm: 56 mJ\n5 cm: 84 mJ\n7 cm: 116 mJ\n(a)Three drop heights\n0 0.2 0.4 0.6 0.8\nPiston speed [m/s]0204060GRF vertical [N]Orifice a: 89 mJ\nOrifice b: 82 mJ\nOrifice c: 81 mJ\n (b)Three valve settings\n0 0.2 0.4 0.6 0.8\nPiston speed [m/s]0204060GRF vertical [N] ~620 g: 241 mJ\n~280 g: 57 mJ\n (c)Two system weights\nFigure 7. Characterizing the hydraulic damper A single damper (not leg-mounted) drops onto the\nforce sensor. 10repeated experiments are plotted as an envelop, defined by the average \u000695 % of the\nstandard deviation data. The curves are read from right to left, i.e. from touch-down at maximum speed\nto zero speed at rest, also corresponding to the maximum damper compression. (a)280 g drop mass with\nmedium orifice in 3drop heights. (b)280 g drop mass with 5 cm drop height in 3orifice settings. (c)3 cm\ndrop height with medium orifice in 2 drop weights.\nThe results from tests with drop heights from 3 cm to7 cm show viscous damping behavior in the\nsettling phase after peak force (left slope), with higher reaction forces at higher piston speeds with higher\ndissipation, ranging from 45 N for maximum speeds of 0.6 m/s with 56 mJ to65 N at0.9 m/s with 116 mJ .\nThe piston force almost linearly depends on the piston speed (Figure 7a).\nFrontiers 13Mo et al. Effective damping in legged locomotion\nChanging the orifice setting at a constant drop height resulted in different settling slopes (Figure 7b).\nApplying a least-squares fit on the left-falling settling slope, we estimate an adjustable damping rate\nbetween 91 Ns/m and192 Ns/m . The dissipated energy changes from 89 mJ to81 mJ , respectively. Hence\nadjusting the orifice setting has an effect on the damping rate and the dissipated energy in the isolated\nhydraulic damper, but not as we intuitively expected.\nWe interpret the rising slope in the impact phase (right part of each curve, Figures 7a and 7b) as a build-up\nphase; the hydraulic damper takes time ( \u001924 ms ) to build up its internal viscous flow and the related piston\nmovement, after the piston impact. With heavier weights ( 620 g = heavy, 280 g = light, Figure 7c), the\nimpact phase equally lasts \u001924 ms . After the impact phase with heavy weight, the damper shows the same\ndamping rate in the settling phase, in form of an equal left slope.\nSimilar drop tests for the evaluation of the isolated diaphragm damper were not possible since the\norientation of the internal diaphragm only permits to pull the piston. In the following section, we test the\ndiaphragm (connected by a piston reverse mechanism) and the hydraulic damper directly on the 2-segment\nleg structure.\n4.2 Composition of dissipated energy\nWe performed drop tests of two damper configurations: one off-the-shelf hydraulic damper, and custom-\nmade pneumatic damper, each mounted in parallel to a spring at the 2-segment leg (Section 3.3, Figure 3b),\nto quantify the effect of viscous damping for drop dynamics similar to legged hopping.\nFor each drop, the effective dissipated energy Eeffective was computed by calculating the area enclosed\nby the vertical GRF-leg length curve from touch-down to lift-off (Josephson, 1985), i.e., the work-loop\narea. These work-loops are to be read counter-clockwise, with the rising part being the loading during leg\nflexion, and the falling part being the unloading, due to spring recoil. Eeffective does not only consist of the\nviscous loss Eviscous due to the damper, but also Coulomb friction loss in the leg ( Ecfriction ) and the impact\nlossEimpact due to unsprung masses:\nEeffective =Ecfriction +Eimpact +Eviscous : (12)\nWe propose a method to indirectly calculate the contribution of viscous damping, by measuring and\neliminating effects from Coulomb friction, and unsprung masses.\nTo quantify the Coulomb friction loss Ecfriction , we conducted ‘slow drop’ tests. The mechanical setup is\nidentical to ‘free drops’ test, where the leg is freely dropped from a fixed height. However, in the ‘slow\ndrop’ experiment the 2-segment leg is lowered manually onto the force plat, contacting and pressing the\nleg-damper-spring system onto the force plate. At slow speed only Coulomb friction in joints and damper\nact, but no viscous damping or impact losses occur. Consequently the dissipated energy calculated from the\nsize of the work loop is due to Coulomb friction losses Ecfriction .\nTo identify the impact loss Eimpact , we remove the viscous component first by detaching the damper\ncable on the setup. A ‘free drop’ test in this spring only condition measures the contribution of friction\nlossEcfriction and impact loss Eimpact combined. A ‘slow drop’ test of the same setup is able to quantify\nthe friction loss Ecfriction . The impact loss Eimpact is therefore estimated as the energy difference between\n‘free drop’ and ‘slow drop’ in the spring-only condition (Figure 8c). Since the effective dissipated energy\nThis is a provisional file, not the final typeset article 14Mo et al. Effective damping in legged locomotion\nEeffective is directly measured, and the friction loss Ecfriction and impact loss Eimpact are obtained separately,\nthe viscous loss Eviscous can be computed according to Equation (12).\nFigures 8a and 8b show the ‘free drop’ and ‘slow drop’ results of the hydraulic damper and diaphragm\ndamper, respectively. Both drop heights are 14 cm , at identical orifice setting. We calculated the negative\nwork of each work-loop (range indicated by the two vertical dash lines), as shown in Figure 8. To provide\nan objective analysis, the work-loop area of each ‘slow drop’ (manual movement) was cut to the maximum\nleg compression of the corresponding ‘free drop’ condition. The dissipated energy of the leg-mounted\nhydraulic damper is 150 mJ and60 mJ for ‘free drop’ and ‘slow drop’, respectively, and 100 mJ and67 mJ\nfor the diaphragm damper, respectively. According to Figure 8c, the impact loss Eimpact due to unsprung\nmasses play a large role, accounting for 31 mJ . The viscous loss Eviscous of the hydraulic and the diaphragm\ndamper are 59 mJ and 2 mJ, respectively.\n80 85 90 95 100\nLeg length [%]0123GRF vertical [BW]Free drop: 150 mJ\nSlow drop: 60 mJ\nleg flexion\n(a)Hydraulic damper and spring\n80 85 90 95 100\nLeg length [%]0123GRF vertical [BW]Free drop: 100 mJ\nSlow drop: 67 mJ\n (b)Diaphragm damper and spring\n80 85 90 95 100\nLeg length [%]0123GRF vertical [BW]Free drop: 91 mJ\nSlow drop: 60 mJ\n (c)Spring only\nFigure 8. Characterizing the contribution of velocity-dependent damping: Vertical GRF versus leg\nlength change, a 2-DOF leg with damper/spring drops onto the force sensor: Three different hardware\nconfigurations were tested, for slow and free drop speeds on the vertical slider. Yellow data lines indicate\nslow-motion experiments. Experiments ‘start’ bottom right, at normalized leg length 100 % . Reading goes\ncounter-clockwise, i.e. from touch-down to mid-stance is indicated by the upper part of the hysteresis curve,\nwhile the lower part indicates elastic spring-rebound, without damper contribution.\n4.3 Adjustability of dissipated energy\nWe tested the adjustability of energy dissipation during leg drops by the altering orifice setting for\neach leg-mounted damper, and quantified by calculating the size of the resulting work-loops. The drop\nheight was fixed to 14 cm and we used 2orifice settings. The identical same set-up but in spring-only\nconfiguration (damper cables detached) was tested for reference. Work-loop and corresponding effective\ndissipated energies are illustrated in Figures 9a and 9b. The hydraulic damper-mounted leg dissipated\n156 mJ and150 mJ energy on its two orifice settings, the pneumatic diaphragm damper dissipated 102 mJ\nand100 mJ . In Figure 9c, we display results from the numerical model introduced in Section 2 to estimate\nthe work-loop shape that either a pure viscous or pure Coulomb damper would produce, if dissipating the\nsame amount of energy as the hydraulic damper with orifice-a (Figure 9a). We set the damping coefficients\nof our numerical model to ED0\u0019156mJ , so that: (dv;dc) = ( 51Ns =m;0N)for pure viscous damping; and\n(dv;dc) = ( 0Ns=m;13:2N)for pure Coulomb damping. Work-loops from the numerical simulation differ\nnotably from the experimental data, suggesting that neither the hydraulic or diaphragm damper can easily\nbe approximated as pure viscous or pure Coulomb dampers. Both work loops in Figure 9c present about\nequal amount of dissipated energy. Yet, both differ greatly due to their underlying damping dynamics,\nvisible in their unique work-loop shapes. Their individual characteristics are different enough to uniquely\nidentify pure viscous or pure Coulomb dampers, from numerical simulation.\nFrontiers 15Mo et al. Effective damping in legged locomotion\n80 85 90 95 100\nLeg length [%]0123GRF vertical [BW]Spring only: 101 mJ\nOrifice a: 156 mJ\nOrifice c: 150 mJ\nleg flexion\n(a)Hydraulic damper and spring\n80 85 90 95 100\nLeg length [%]0123GRF vertical [BW]Spring only: 87 mJ\nOrifice a: 102 mJ\nOrifice c: 100 mJ\n (b)Diaphragm damper and spring\n75 80 85 90 95 100\nLeg length [%]0246GRF vertical [BW]Pure viscous\nPure Coulomb (c)Simulation\nFigure 9. Adjustability and tunability of damping: Vertical GRF vs leg length change, a 2-segment\nleg with damper and spring drops onto the force sensor. Two different hardware configurations were\ntested, for different orifice settings. (a) and (b) show the result from hydraulic damper and diaphragm\ndamper respectively, where the green data lines indicate the leg drop without damper for comparison. (c):\nSimulated approximation of hydraulic damper orifice a by a pure viscous and a Coulomb damper. Damping\ncoefficients are chosen to allows same dissipated energy, i.e., ED0=156mJ: respectively — pure viscous\ndamper: dc=0 Nanddv=51 Ns/m ; pure Coulomb damper: dc=13.2 N anddv=0 Ns/m . None of the two\ncurves can fully capture the work-loop of hydraulic damper.\n4.4 Damper selection choices\nIn accordance with the simulation results, we aim to use a viscous damper to dissipate energy introduced\nby a ground disturbance. How much energy could be dissipated by the damper, depended mainly on the\nselected viscous damper, and only to a limited degree on the orifice setting. Results from the hydraulic\ndamper 1214H showed significant energy dissipation capabilities: \u001911 % of the system’s total energy\n(59 mJ of560 mJ ) were dissipated (Figure 9a at orifice setting ’c’ and Table 4 ). At the drop, in sum\n150 mJ (27 % ) of the leg’s system energy were lost, due to Coulomb friction in the joints, impact dynamics,\nand viscous damping losses. Other dissipation dynamics are feasible, by selecting appropriate dampers.\nWe tested a second hydraulic damper (1210M, MISUMI) under equal conditions and compared it to\ndamper-1214H. The two applied orifice settings changed the observed work loop largely by shape, and\nlittle by area (Figure 10). The damper-1210M dissipated \u001960 % system energy, and the leg lost in sum\n(viscous+Coulomb+impact) 72 % of its system’s energy during that single drop. At other orifice settings,\nwe observed over-damping; the 1210M-spring leg came to an early and complete stop, and without rebound\n(data not shown here due to incomplete work loop).\nFor comparison, time plots of the vertical GRF and the impulse at stance phrase are shown in Figure 11.\nThe energy composition (Equation 12) is provided in Table 4. The ’spring only’ data correspond the\ncurves in Figure 8c. The diaphragm+spring data correspond to ‘orifice c’ in Figure 9b. The hydraulic\n(1214H)+spring data correspond to ‘orifice c’ in Figure 9a. The hydraulic (1210M)+spring data correspond\nto ‘orifice b’ in Figure 10. Among the tested dampers, the hydraulic 1210M damper showed the largest\nvertical GRF; peak vertical GRF of 6.3 BW are observed, almost twice as much as the ‘spring only’ case.\nThe viscous dampers 1214H and 1210M shifted the peak of their legs’ vertical GRF to an earlier point in\ntime, compared to the spring-leg and the spring+diaphragm-leg (Figure 11).\n5 DISCUSSION\nA primary objective of this study was to test how physical dampers could be exploited for locomotion\ntasks by characterizing multiple available technical solutions. Our numerical model predicted three crucial\naspects: (1) a pure viscous damper generally performs better than a pure Coulomb damper (Figure 4); (2)\nThis is a provisional file, not the final typeset article 16Mo et al. Effective damping in legged locomotion\n80 85 90 95 100\nLeg length [%]01234567GRF vertical [BW]Spring only: 101 mJ\nOrifice a: 395 mJ\nOrifice b: 401 mJ\nFigure 10. Higher energy dissipation with a different model of the hydraulic damper (1210M):\nVertical GRF vs. leg length change, a 2-DOF leg with a parallel damper and spring drops onto the\nforce sensor. Two damper orifice settings were tested (blue, red curves). The two resulting curves are\ncompared with the spring-only configuration, provided as reference.\n0 0.05 0.1 0.15 0.205 spring only\n0 0.05 0.1 0.15 0.205 diaphragm + spring\n0 0.05 0.1 0.15 0.205 Hydraulic (1214H) + spring\n0 0.05 0.1 0.15 0.2\nTime [s]05GRF vertical [BW]\nHydraulic (1210M) + spring\n(a)Vertical GRF over time\n0 0.05 0.1 0.15 0.201\nspring only\n0 0.05 0.1 0.15 0.201\ndiaphragm + spring\n0 0.05 0.1 0.15 0.201\nHydraulic (1214H) + spring\n0 0.05 0.1 0.15 0.2\nTime [s]01Impulse [Ns]\nHydraulic (1210M) + spring (b)Vertical Impulse over time\nFigure 11. Ground reaction forces and the corresponding, instantaneous impulse for leg drop\nexperiments . The corresponding work curves are provided in Figures 8 to 10.\nhigher damping rates result in better rejection of ground disturbances (Figure 4a), however at the cost\nof higher dissipation at reference height (Table 2); (3) characteristic work loop shapes for pure viscous\nand Coulomb damper during leg-drop (Figure 9c). Our hardware findings show that neither of the tested\nphysical dampers approximates as pure viscous or pure Coulomb dampers. The experiments also suggest\nthat the mapping between dissipated energy and damping rates is concealed by the dynamics of the impact\nand the non-linearity of the force-velocity characteristics of the leg in the stance phase. Therefore, it is vital\nto test damping in a real leg at impact because the behavior is not merely as expected from the data sheets\nand the simple model.\nFigure 7 characterizes how the hydraulic damper dissipates energy during a free drop. The experimental\nresults show that the dissipated energy of the hydraulic damper scales with drop height (Figure 7a) and\nweight (Figure 7c), but less intuitively, it reduces with increasing damping rates (Figure 7b). This can be\npartially interpreted in the context of an ideal viscous damper (as in Equation (5)), but linear) for which the\nFrontiers 17Mo et al. Effective damping in legged locomotion\nTable 4. Leg drop experiments and their individual energetic losses per drop. The system’s initial\npotential energy is 560 mJ .Eeffective : sum of all energetic losses visible as the area of the hysteresis curve,\ni.e. in Figure 8, Ecfriction : negative work dissipated by Coulomb friction, Eimpact : energetic losses from\nimpact (unsprung mass). The negative work dissipated by viscous damping in the physical damper is\nEviscous . The corresponding work curves are provided in Figures 8 to 10.\nDrop test setup Eeffective Ecfriction Eimpact Eviscous\n[mJ] [mJ] [mJ] [mJ]\nSpring only 91 60 31 0\nDiaphragm + spring 100 67 31 2\nHydraulic 1214H + spring 150 60 31 59\nHydraulic 1210M + spring 401 38 31 332\neffective dissipated energy Eeffective would be calculated as in,\nEeffective =Z\nFp(t)dyp=Z\n(dv\u0001vp(t))dyp (13)\nwhere Fp(t)is the damper piston force and ypis the piston displacement, vp(t)the corresponding velocity.\nWhen increasing the drop height, the velocity at impact is increased, so is vp(t). With the assumption of\nEquation (13), this results in higher damping forces Fp(t), and thus, dissipated energy Eeffective , as seen\nin Figure 7a. The heavier drop weight leads to slower deceleration. Therefore the velocity profile vp(t)is\nincreased, which also leads to higher dissipation Eeffective (Figure 7b). An orifice setting of high damping\nrate will increase the damping coefficient dv. However, the velocity profile vp(t)is expected to reduce due\nto higher resistance. This simple analogy shows that the coupling between damping coefficient dvand\nvelocity profile vp(t)makes it difficult to predict the energy dissipation by setting the orifice and serves as\nan interpretation of why adjusting the orifice generates a relatively small adjustment of 10 % (81 mJ -89 mJ )\nof the dissipated energy. Also, the impact phase (time for the damper to output its designed damping force\nunder sudden load) introduces additional non-linearity to the output force profile. Overall, the results in\nFigure 7 indicate that we can approximate the damping force produced by the hydraulic damper to be\nviscous and adjustable— as such dampers are typically designed (Dixon, 2008)—, but the mapping of\nenergy dissipation to orifice setting is difficult to predict in a dynamic scenario.\nThe approximation as a linear, velocity dependent damper allows us to rapidly estimate energy dissipation\nin simulation, over a range of parameters. However, the exact mapping of the hardware leg/spring/damper\nenergy dissipation to orifice setting is difficult to predict, when basing the estimation only on the isolated-\ndamper drop experiments from Figure 7. Instead, the leg/spring/damper experiments show that the energetic\nlosses from the impact remove 31 mJ energy, compared to 59 mJ damper losses. The high amount of force\noscillations at impact (up to 1 BW , Fig. 8a) during the first 3 %leg length change leads us to believe that\nthese impact oscillations move the damper’s dynamic working range, i.e., its resulting instantaneous force\nand velocity. The oscillations are likely caused by unsprung mass effects of the leg/spring/damper structure,\nand could not be captured in an isolated-damper setup, or—at least not easily—in a simulation.\nThe work loops of leg drop experiments (Figure 8) show the effects of our tested dampers on a legged\nsystem. From touch-down to mid-stance ( leg flexion ), the ‘free drop’ curves show a larger negative work\ncompared to the ‘slow drop’ curves, illustrating that the damper absorbs extra energy. The returning curves\n(mid-stance to lift-off) of the hydraulic damper aligns well with the ‘slow drop’ curve, indicating the\ndamper is successfully detached due to slackness cable while the spring recoil. Figure 8b shows that the\n‘free drop’ force of the diaphragm damper is slightly higher than ‘slow drop’ force in the first half of the leg\nThis is a provisional file, not the final typeset article 18Mo et al. Effective damping in legged locomotion\nextension phase. This discrepancy is likely caused by the elastic force component of the diaphragm damper\ndue to sudden expansion of the air chamber volume. The elastic component seems to dominate the damper\nbehavior, which thus acts mostly as an air spring. By separating its energetic components (Equation (12)),\nwe found that the hydraulic damper produces a viscous-like resistance higher than the diaphragm damper\n(59 mJ versus 2 mJ), indicating the hydraulic damper is more effective in dissipating energy under drop\nimpact. Hence, the hydraulic damper shows more viscous behavior, while the diaphragm damper is more\nelastic.\nPhysical damping in the system comes at the cost of energy loss, and to maintain periodic hopping, it\nbecomes necessary to replenish energy that is dissipated by damping ( ED0). Therefore, there is a trade-off\nto consider: simulation results show that higher damping results in faster rejection of ground perturbation\nat the price of more energy consumption at reference drop height (Table 2, Figure 4). An adjustable damper\nwould partly address this problem: on level ground, the damping rate could be minimal, and on rough\nterrain increased. The adjustability of the two dampers is illustrated in Figures 9a and 9b. We discuss the\nadjustability from both energy dissipation and dynamic behavior perspectives.\nCompared with the spring-only results, both the hydraulic and the diaphragm damper reduced the\nmaximum leg flexion and dissipated more energy. The orifice setting changes the shape of the work loop\ndifferently for the two setups. For the hydraulic damper (Figure 9a), orifice setting-c shrinks the work\nloop from left edge, indicating more resistance is introduced by the damper to reduce leg flexion. For\nthe diaphragm damper (Figure 9b), orifice setting-c not only shrinks the work loop, but also increases its\nslope. We interpret this as the elastic contribution of air compression: relatively fewer air leaves through\nthe smaller orifice, but instead acts as an in-parallel spring.\nConcerning energy dissipation, changes of orifice settings led to relatively small changes in effective\ndissipated energy Eeffective :150 mJ to156 mJ for the hydraulic damper, and 100 mJ to102 mJ for the\ndiaphragm damper. Even for the other damper model (1210M), which dissipates high amounts of energy,\nchanges in orifice setting change the work loop shape drastically, but not the dissipated energy ( 395 mJ\nversus 401 mJ).\nSimilar to the isolated damper drop, the data (Figures 9a and 9b) shows that specific orifice settings\nintroduce more resistance, but not necessarily lead to higher energy dissipation, for both hydraulic and\ndiaphragm damper. However, in our simplified numerical leg model, an increase in viscous damping\ncoefficients leads to a systematic increase of dissipated energy (Table 2), and a sharper tip at the left side of\nthe work loop (Figure 9c). The discrepancy is likely due to the non-linear coupling between the damper\nmechanics and the leg dynamics in the hardware setup: (1) The damping force generated by the fluid\ndynamics in the orifice only approximates a linear viscosity (Dixon, 2008). (2) The impact loading on\nboth the nonlinear leg structure and the damper. This makes the prediction of the energy dissipation not\nstraight-forward based on our simplified numerical leg model, and points towards the need of a combined\napproach between simulation and hardware testing to fully understand physical damping in a legged system.\nViscous, velocity dependent damping alters the leg’s loading characteristics, and leads to a peak force at\nthe instance of touch-down. As a result, the vertical GRF is increased in the early stance phase, shifting\nand increasing the peak vertical GRF before mid-stance (Figure 11a). When designing a legged system\nwith a viscous damper, its increasing load on the mechanical structure should be considered.\nThe selection of viscous dampers depends on the task. High damping can fully reject disturbances in a\nsingle cycle, but lower damping could have energetic benefits. Here we looked for a damper that would\ndissipate significant negative work (Eviscous\nET0\u001910%\u000015% ) in form of viscous damping. The air-filled\nFrontiers 19Mo et al. Effective damping in legged locomotion\ndiaphragm damper lead to insufficient energy losses ( 2 %), but the hydraulic dampers dissipated 10 % and\n60% of the system’s total energy (Table 4).\nDrawing conclusions about animal locomotion based on the here presented leg-drop experiments is\nsomewhat early. However, observations from (M ¨uller et al., 2014, Table 1, p. 2288) indicate that leg forces\ncan increase at unexpected step-downs during locomotion experiments. Further, Kalveram et al. (2012) Rev1,\nGC 1 suggests in a comparison of experimental human hopping and numerical simulations that damping may\nbethedriving ingredient in passive stabilization against ground-level perturbations. We are consequently\nexcited about the here presented results of viscous dampers mounted in parallel to a leg’s spring, producing\nadaptive forces without the need for sensing.\n6 CONCLUSION\nWe investigated the possibility to exploit physical damping in a simplified leg drop scenario as a template\nfor the early stance phase of legged locomotion. Our results from a) numerical simulation promote the\nuse of adjustable and viscous damping over Coulomb damping to deal with a ground perturbation by\nphysical damping. As such, we b) tested two technical solutions in hardware: a commercial, off-the-shelf\nhydraulic damper, and a custom-made, rolling diaphragm damper. We dissected the observed dissipated\nenergy from the hardware damper-spring leg drops, into its components, by experimental design. The\nresulting data allowed us to characterize dissipation from the early impact (unsprung-mass effects), viscous\ndamping, Coulomb damping, and orifice adjustments individually, and qualitatively . The rolling diaphragm\ndamper features low-Coulomb friction, but dissipates only low amounts of energy through viscous damping.\nThe off-the-shelf, leg-mounted hydraulic damper did exhibit high viscous damping, and qualitatively\nshowed the expected relationship between impact speed, output force and negative work. Changes in orifice\nsetting showed only minor changes in overall energy dissipation, but can lead to large changes in leg\nlength dynamics, depending on the chosen technical damper. Hence, switching between different viscous,\nhydraulic dampers is an interesting future option. Our results show how viscous, hydraulic dampers react\nvelocity-dependent, and create an instantaneous, physically adaptive response to ground-level perturbations\nwithout sensory-input.\nCONFLICT OF INTEREST STATEMENT\nThe authors declare that the research was conducted in the absence of any commercial or financial\nrelationships that could be construed as a potential conflict of interest.\nAUTHOR CONTRIBUTIONS\nAM contributed to concept, hardware design, experimental setup, experimentation, data discussion and\nwriting. FI contributed to concept, simulation framework, experimental setup, data discussion and writing.\nDH and ABS contributed to concept, data discussion and writing.\nACKNOWLEDGMENTS\nThe authors thank the International Max Planck Research School for Intelligent Systems (IMPRS-IS) for\nsupporting AM, FI, the China Scholarship Council (CSC) for supporting AM, and the Ministry of Science,\nResearch and the Arts Baden-W ¨urttemberg (Az: 33-7533.-30-20/7/2) for supporting DH, and the Max\nPlanck Society for supporting ABS.\nThis is a provisional file, not the final typeset article 20Mo et al. Effective damping in legged locomotion\nREFERENCES\nAbraham, I., Shen, Z., and Seipel, J. (2015). A Nonlinear Leg Damping Model for the Prediction of Running\nForces and Stability. Journal of Computational and Nonlinear Dynamics 10. doi:10.1115/1.4028751\nBledt, G., Wensing, P. M., Ingersoll, S., and Kim, S. (2018). Contact model fusion for event-based\nlocomotion in unstructured terrains. In 2018 IEEE International Conference on Robotics and Automation\n(ICRA), Brisbane, Australia, Submitted (IEEE), 1–8\nBlickhan, R., Seyfarth, A., Geyer, H., Grimmer, S., Wagner, H., and G ¨unther, M. (2007). Intelligence\nby mechanics. Philosophical Transactions of the Royal Society of London, Series A 365, 199–220.\ndoi:10.1098/rsta.2006.1911\nDaley, M. A. and Biewener, A. A. (2006). Running over rough terrain reveals limb control for intrinsic\nstability. Proceedings of the National Academy of Sciences 103, 15681–15686. doi:10.1073/pnas.\n0601473103\nDixon, J. C. (2008). The shock absorber handbook (John Wiley & Sons)\nDrama, O. and Spr ¨owitz, A. (2020). Trunk pitch oscillations for energy trade-offs in bipedal running birds\nand robots. Bioinspiration & Biomimetics doi:10.1088/1748-3190/ab7570\nGarcia, E., Arevalo, J. C., Munoz, G., and Gonzalez-de Santos, P. (2011). Combining series elastic actuation\nand magneto-rheological damping for the control of agile locomotion. Robotics and Autonomous Systems\n59, 827–839\nGhazi-Zahedi, K., Haeufle, D. F. B., Mont ´ufar, G., Schmitt, S., and Ay, N. (2016). Evaluating\nMorphological Computation in Muscle and DC-motor Driven Models of Human Hopping. Frontiers in\nRobotics and AI 3, 1–10. doi:10.3389/frobt.2016.00042\nGrimminger, F., Meduri, A., Khadiv, M., Viereck, J., W ¨uthrich, M., Naveau, M., et al. (2020). An Open\nTorque-Controlled Modular Robot Architecture for Legged Locomotion Research. IEEE Robotics and\nAutomation Letters 5, 3650–3657. doi:10.1109/LRA.2020.2976639\nHaeufle, D. F. B., Grimmer, S., and Seyfarth, A. (2010). The role of intrinsic muscle properties for stable\nhopping - stability is achieved by the force-velocity relation. Bioinspiration & Biomimetics 5, 016004.\ndoi:10.1088/1748-3182/5/1/016004\nHaeufle, D. F. B., G ¨unther, M., Wunner, G., and Schmitt, S. (2014). Quantifying control effort of biological\nand technical movements: An information-entropy-based approach. Physical Review E 89, 012716.\ndoi:10.1103/PhysRevE.89.012716\nHavoutis, I., Semini, C., Buchli, J., and Caldwell, D. G. (2013). Quadrupedal trotting with active\ncompliance. In Mechatronics (ICM), 2013 IEEE International Conference on (IEEE), 610–616\nHu, C.-J., Wang, T.-K., Huang, C.-K., and Lin, P.-C. (2019). A torque-actuated dissipative spring loaded\ninverted pendulum model with rolling contact and its application to hexapod running. Bioinspiration &\nbiomimetics 14. doi:10.1088/1748-3190/aafc4e\nHutter, M., Gehring, C., Bloesch, M., Hoepflinger, M. A., Remy, C. D., and Siegwart, R. (2012). Starleth:\nA compliant quadrupedal robot for fast, efficient, and versatile locomotion. In Adaptive Mobile Robotics\n(World Scientific), 483–490\nHutter, M., Gehring, C., Jud, D., Lauber, A., Bellicoso, C. D., Tsounis, V ., et al. (2016). Anymal - a highly\nmobile and dynamic quadrupedal robot. 2016 Ieee/Rsj International Conference on Intelligent Robots\nand Systems (Iros 2016) , 38–44\nJosephson, R. K. (1985). Mechanical power output from striated muscle during cyclic contraction. Journal\nof Experimental Biology 114, 493–512\nFrontiers 21Mo et al. Effective damping in legged locomotion\nKalouche, S. (2017). Goat: A legged robot with 3d agility and virtual compliance. In 2017 IEEE/RSJ\nInternational Conference on Intelligent Robots and Systems (IROS) . 4110–4117. doi:10.1109/IROS.\n2017.8206269\nKalveram, K. T., Haeufle, D. F. B., Seyfarth, A., and Grimmer, S. (2012). Energy management that\ngenerates terrain following versus apex-preserving hopping in man and machine. Biological Cybernetics\n106, 1–13. doi:10.1007/s00422-012-0476-8\nMochon, S. and McMahon, T. A. (1980). Ballistic walking: An improved model. Mathematical Biosciences\n52, 241–260\nMore, H. L. and Donelan, J. M. (2018). Scaling of sensorimotor delays in terrestrial mammals. Proc. R.\nSoc. B 285, 20180613. doi:10.1098/rspb.2018.0613\nMore, H. L., Hutchinson, J. R., Collins, D. F., Weber, D. J., Aung, S. K. H., and Donelan, J. M. (2010).\nScaling of Sensorimotor Control in Terrestrial Mammals. Proceedings of the Royal Society B: Biological\nSciences 277, 3563–3568. doi:10.1098/rspb.2010.0898\nM¨uller, R., Tschiesche, K., and Blickhan, R. (2014). Kinetic and kinematic adjustments during perturbed\nwalking across visible and camouflaged drops in ground level. Journal of Biomechanics 47, 2286–2291.\ndoi:10.1016/j.jbiomech.2014.04.041\nRummel, J. and Seyfarth, A. (2008). Stable Running with Segmented Legs. The International Journal of\nRobotics Research doi:10.1177/0278364908095136\nRuppert, F. and Badri-Spr ¨owitz, A. (2019). Series Elastic Behavior of Biarticular Muscle-Tendon Structure\nin a Robotic Leg. Frontiers in Neurorobotics 13, 64. doi:10.3389/fnbot.2019.00064\nSecer, G. and Saranli, U. (2013). Control of monopedal running through tunable damping. In 2013 21st\nSignal Processing and Communications Applications Conference (SIU) (IEEE), 1–4. doi:10.1109/SIU.\n2013.6531557\nSeok, S., Wang, A., Chuah, M. Y . M., Hyun, D. J., Lee, J., Otten, D. M., et al. (2015). Design principles\nfor energy-efficient legged locomotion and implementation on the mit cheetah robot. IEEE/ASME\nTransactions on Mechatronics 20, 1117–1129. doi:doi.org/10.1109/TMECH.2014.2339013\nSerafin, S. (2004). The sound of friction: real time models, playability and musical applications . Thesis,\nDepartment of Music\nShen, Z. and Seipel, J. (2012). A fundamental mechanism of legged locomotion with hip torque and leg\ndamping. Bioinspiration & biomimetics 7, 046010. doi:10.1088/1748-3182/7/4/046010\nSpr¨owitz, A., Tuleu, A., Vespignani, M., Ajallooeian, M., Badri, E., and Ijspeert, A. J. (2013). Towards\ndynamic trot gait locomotion: Design, control, and experiments with cheetah-cub, a compliant quadruped\nrobot. The International Journal of Robotics Research 32, 932–950\nTsagarakis, N. G., Morfey, S., Dallali, H., Medrano-Cerda, G. A., and Caldwell, D. G. (2013). An\nasymmetric compliant antagonistic joint design for high performance mobility. In 2013 IEEE/RSJ\nInternational Conference on Intelligent Robots and Systems (IEEE), 5512–5517\nWerner, A., Turlej, W., and Ott, C. (2017). Generation of Locomotion Trajectories for Series Elastic and\nViscoelastic Bipedal Robots. In 2017 IEEE/RSJ International Conference on Intelligent Robots and\nSystems (IROS) (IEEE), 5853–5860\nWhitney, J. P., Chen, T., Mars, J., and Hodgins, J. K. (2016). A hybrid hydrostatic transmission and\nhuman-safe haptic telepresence robot. In Proceedings of ICRA (IEEE), 690–695\nWhitney, J. P., Glisson, M. F., Brockmeyer, E. L., and Hodgins, J. K. (2014). A low-friction passive fluid\ntransmission and fluid-tendon soft actuator. In Proceedings of IROS (IEEE), 2801–2808\nZahedi, K. and Ay, N. (2013). Quantifying morphological computation. Entropy 15, 1887–1915. doi:10.\n3390/e15051887\nThis is a provisional file, not the final typeset article 22" }, { "title": "2210.09824v1.Concepts_in_Lorentz_and_CPT_Violation.pdf", "content": "arXiv:2210.09824v1 [hep-ph] 18 Oct 2022Proceedings of the Ninth Meeting on CPT and Lorentz Symmetry (CPT’22), Indiana University, Bloomington, May 17–26, 2022\n1\nConcepts in Lorentz and CPT Violation\nV. Alan Kosteleck´ y\nPhysics Department, Indiana University\nBloomington, IN 47405, USA\nThis contribution to the CPT’22 meeting provides a brief rev iew of some con-\ncepts in Lorentz and CPT violation.\n1. Introduction\nIn recent years, substantial advances have been made in the the ory and\nphenomenologyofLorentz and CPT violation, driven by the intriguing pos-\nsibility that experimental searches for the associated effects cou ld uncover\ntiny observable signals from an underlying unified theory such as str ings.1\nThis contribution to the CPT’22 proceedings summarizes a few conce pts\nin the subject, focusing on the approach that uses effective field t heory2to\nconstruct a general realistic framework describing physical effec ts.3–5\n2. Foundations\nA key property of Minkowski spacetime is its invariance under Loren tz\ntransformations. The Lorentztransformationsinclude spatialr otationsand\nvelocity boosts, which together can be viewed as generalized rotat ions in\nspacetime. Theories manifesting isotropy under these spacetime r otations\nhave Lorentz invariance (LI), so a theory with Lorentz violation (L V) incor-\nporates one or more spacetime anisotropies. Experiments testing LI seek\nto identify possiblespacetime anisotropiesby comparingphysicalqu antities\nat different spacetime orientations. For example, symmetry under spatial\nrotations can be explored by comparing at different spatial orienta tions the\nticking rates of two clocks or the lengths of two standard rulers.\nOur most successful fundamental theories describing nongravit ational\naspects of Nature are constructed on Minkowski spacetime. How ever, the\nUniverse contains gravitational interactions, which cannot be scr eened and\nhence are ubiquitous. Minkowski spacetime is therefore believed to be un-\nphysical in detail, with our Universe instead involving Riemann spacetim eProceedings of the Ninth Meeting on CPT and Lorentz Symmetry (CPT’22), Indiana University, Bloomington, May 17–26, 2022\n2\nor perhaps a generalization. A generic Riemann spacetime lacks globa l\nLI. Instead, the relevant spacetime symmetries are local Lorent z invariance\n(LLI) and diffeomorphism invariance (DI). A theory with LLI is isotro pic\nunder local spacetime rotations about every point, while one with DI is\nunchanged by local translations. Experiments can test these sym metries\nby comparing properties of objects at different orientations and lo cations\nin the neighborhood of a spacetime point.\nMost experiments and observations involve only weak gravitational\nfields and so take place in asymptotically flat spacetime, a limit of Rie-\nmann spacetime that reduces to Minkowski spacetime for zero gra vity. In\nasymptotically flat spacetime, the standard notion of LI turns out to be a\ncombination of LLI and DI.5Experiments searching for LV are therefore\nin reality sensitive to a combination of local Lorentz violation (LLV) an d\ndiffeomorphism violation (DV). The corresponding observables manif est a\nmixture of local spacetime anisotropy and spacetime-position depe ndence.\n3. Theory\nSince no compelling evidence for LV exists to date, a broad-based an d\nmodel-independent methodology is desirable in the search for possib le ef-\nfects. Any violations are expected to be small corrections to the k nown\nphysics of General Relativity (GR) and the Standard Model (SM), s o it\nis natural to study LV using the approach of effective field theory.2Typi-\ncally, integrating over high-energy degrees of freedom in a theory generates\na specific effective field theory applicable at low energies. In the cont ext\nof searches for LV, however, the approach can instead be used t o study\nsimultaneously a large class of underlying theories and determine pos sible\nobservable effects in a model-independent way.\nThe realistic coordinate-independent theory incorporating gener al LV\nthat is basedon GR coupled tothe SM is knownasthe Standard-Mode lEx-\ntension (SME).3,4In a realistic effective field theory, CPT violation comes\nwith LV both in Minkowski spacetime3,6and in asymptotically flat space-\ntime,4so the SME also describes CPT violation in a model-independent\nway. The Lagrange density Lof the theory includes LV operators of any\nmass dimension d, with the minimal theory defined as the subset of oper-\nators of renormalizable dimension d≤4. Each term in Lis constructed as\nan observer-scalarcontraction of a LV operator O(x) with a coupling coeffi-\ncientk(x) or its derivatives. The coefficients are expected to be suppresse d\neither by powers of a high-energy scale such as the Planck energy o r viaProceedings of the Ninth Meeting on CPT and Lorentz Symmetry (CPT’22), Indiana University, Bloomington, May 17–26, 2022\n3\na mechanism such as countershading.7The explicit forms of all minimal\nterms3,4and many nonminimal terms5,8–10are known.\n4. Backgrounds\nAny given coefficient k(x) inLcan be viewed as a prescribed background\nin spacetime,4which may arise as a vacuum expectation value of a field.\nAll indices carried by k(x) are contracted with those of the corresponding\noperator O(x), sok(x)iscovariantunderobserverlocalLorentztransforma-\ntions and general coordinate transformations. However, k(x) is unaffected\nby particle local Lorentz transformations and diffeomorphisms, wh ich act\nonly on dynamical fields. Covariant and contravariant local indices o nk(x)\nare physically equivalent because the local metric ηis Minkowski and non-\ndynamical. In contrast, covariant and contravariant spacetime in dices can\ngenerate physically distinct effects because the spacetime metric gis dy-\nnamical. Disregarding derivatives of k(x), a generic term in Lthus takes\nthe form L ⊃kµ...ν...a...(x)Oµ...ν...a...(x), wherespacetimeindicesareGreek\nand local indices areLatin. The term is LLVwhen k(x) carriesa localindex\nand is DV when k(x) carries a spacetime index or varies with x.\nIt is physically useful to distinguish two types of backgrounds k(x), de-\nnoted as /angbracketleftk/angbracketrightandk. Spontaneous backgrounds /angbracketleftk/angbracketrightarise dynamically from\nsolving equations of motion, so they satisfy the Euler-Lagrange eq uations\nand are thus on shell. Their dynamical origin means that small fluctua tions\naround/angbracketleftk/angbracketrightcan occur, so additional modes appear in the effective theory11\nincluding Nambu-Goldstone12and massive modes. Explicit backgrounds k\nare externallyprescribed and hence nondynamical. They areuncon strained\nby Euler-Lagrange equations and thus can be off shell, and no dynam ical\nfluctuations occur. The spontaneous or explicit nature of the LLV and DV\ncan be used to classify terms in Land determine their physical implica-\ntions.5The geometry of gravity is affected differently in the two scenarios.\nFor spontaneous violation, it can remain Riemann or Riemann-Cartan .4\nFor explicit violation, in contrast, the geometry typically cannot be R ie-\nmann and is conjectured instead to be Finsler,4,5,13,14leading to unique\ngravitational effects.15\n5. Applications\nBy virtue of its generality and model independence, the SME can be e x-\npected to contain the large-distance limit of any realistic theory with LV.\nThe background coefficients affect the behavior of experimentally k nownProceedings of the Ninth Meeting on CPT and Lorentz Symmetry (CPT’22), Indiana University, Bloomington, May 17–26, 2022\n4\nforce and matter fields and hence can be used to predict possible ob serv-\nable signals of Lorentz and CPT violation. The physical definition of a\ngiven particle species can be affected,16and both its free propagation and\ninteractions can be modified in a flavor-dependent way. The effects can\ndepend on the magnitude and orientation of momenta and spins and c an\ndiffer between particles and antiparticles. For spontaneous LV, th e dynam-\nical fluctuations around /angbracketleftk/angbracketrightcan also modify the physics and can even play\nthe role of the photon or graviton.11In any scenario, the coefficients are\nthe targets for experimental searches. Since a coefficient chang es under\nobserver frame transformations, the inertial frame used to pre sent results\nmust be specified. In practice no laboratory is inertial, and the Eart h’s\nrotation and revolution imply that LV measurements typically exhibit t ime\nvariations.17Instead, the canonical choice to report experimental results\nis the Sun-centered frame.18Many experiments have achieved impressive\nsensitivities to coefficients expressed in this frame.19\nThe SME framework also has applications in broader contexts. One\nconcernssearchesfornewLIphysicsbeyondGRandtheSM.Byvir tueofits\ninclusion of all effective background couplings, the framework can d escribe\nphysicaleffects fromanynewfieldproducingabackgroundoveras pacetime\nregionofexperimentalrelevance, andhenceitcanbeusedtodedu cebounds\non new LI physics from existing constraints on LV. This technique ha s\nled, for example, to tight constraints on torsion20and nonmetricity,21and\nsimilar ideas have been adopted for spacetime-varying couplings and ghost-\nfree massive gravity.22Another application in a different context involves\nthe description of emergent LI in certain condensed-matter syst ems. For\nexample, properties of Dirac and Weyl semimetals are calculable in the\nSME framework.23Future explorations in these and other contexts offer\nexcellent prospects for conceptual and practical advances.\nAcknowledgments\nThis work was supported in part by US DoE grant DE-SC0010120 and by\nthe Indiana University Center for Spacetime Symmetries.\nReferences\n1. V.A. Kosteleck´ y and S. Samuel, Phys. Rev. D 39, 683 (1989); V.A. Kost-\neleck´ y and R. Potting, Nucl. Phys. B 359, 545 (1991); Phys. Rev. D 51, 3923\n(1995).\n2. See, e.g., S. Weinberg, Proc. Sci. CD 09, 001 (2009).Proceedings of the Ninth Meeting on CPT and Lorentz Symmetry (CPT’22), Indiana University, Bloomington, May 17–26, 2022\n5\n3. D. Colladay and V.A. Kosteleck´ y, Phys. Rev. D 55, 6760 (1997); Phys. Rev.\nD58, 116002 (1998); V.A. Kosteleck´ y and R. Lehnert, Phys. Rev. D63,\n065008 (2001).\n4. V.A. Kosteleck´ y, Phys. Rev. D 69, 105009 (2004).\n5. V.A. Kosteleck´ y and Z. Li, Phys. Rev. D 103, 024059 (2021).\n6. O.W. Greenberg, Phys. Rev. Lett. 89, 231602 (2002).\n7. V.A.Kosteleck´ yand J.D. Tasson, Phys.Rev.Lett. 102, 010402 (2009); Phys.\nRev. D83, 016013 (2011).\n8. V.A. Kosteleck´ y and M. Mewes, Phys. Rev. D 80, 015020 (2009); Phys. Rev.\nD85, 096005 (2012); Phys. Rev. D 88, 096006 (2013); Y. Ding and V.A.\nKosteleck´ y, Phys. Rev. D 94, 056008 (2016).\n9. V.A. Kosteleck´ y and Z. Li, Phys. Rev. D 99, 056016 (2019).\n10. Q.G. Bailey et al., Phys. Rev. D 91, 022006 (2015); V.A. Kosteleck´ y and M.\nMewes, Phys. Lett. B 757, 510 (2016); Phys. Lett. B 766, 137 (2017); Phys.\nLett. B779, 136 (2018).\n11. R. Bluhm and V.A. Kosteleck´ y, Phys. Rev. D 71, 065008 (2005); R. Bluhm\net al., Phys. Rev. D 77, 065020 (2008); V.A. Kosteleck´ y and R. Potting, Gen.\nRel. Grav. 37, 1675 (2005); Phys. Rev. D 79, 065018 (2009); M.D. Seifert,\nPhys. Rev. D 81, 065010 (2010); B. Altschul et al., Phys. Rev. D 81, 065028\n(2010).\n12. Y. Nambu, Phys. Rev. Lett. 4, 380 (1960); J. Goldstone, Nuov. Cim. 19,\n154 (1961); J. Goldstone, A. Salam, and S. Weinberg, Phys. Re v.127, 965\n(1962).\n13. R. Bluhm, Phys. Rev. D 91, 065034 (2015); Phys. Rev. D 92, 085015 (2015);\nR. Bluhm and A. Sehic, Phys. Rev. D 94, 104034 (2016); R. Bluhm, H. Bossi,\nand Y. Wen, Phys. Rev. D 100, 084022 (2019).\n14. M. Schreck, Phys. Lett. B 793, 70 (2019); B.R. Edwards and V.A. Kost-\neleck´ y, Phys. Lett. B 786, 319 (2018); D. Colladay and P. McDonald, Phys.\nRev. D85, 044042 (2012); V.A. Kosteleck´ y, Phys. Lett. B 701, 137 (2011);\nV.A. Kosteleck´ y and N. Russell, Phys. Lett. B 693, 2010 (2010).\n15. V.A. Kosteleck´ y and Z. Li, Phys. Rev. D 104, 044054 (2021).\n16. V.A. Kosteleck´ y, E. Passemar, and N. Sherrill, arXiv:2 207.04545.\n17. V.A. Kosteleck´ y, Phys. Rev. Lett. 80, 1818 (1998).\n18. R.Bluhm et al., Phys.Rev.D 68, 125008 (2003); Phys.Rev.Lett. 88, 090801\n(2002); V.A. Kosteleck´ y and M. Mewes, Phys. Rev. D 66, 056005 (2002).\n19. V.A. Kosteleck´ y and N. Russell, arXiv:0801.0287v15.\n20. R. Lehnert, W.M. Snow, and H. Yan, Phys. Lett. B 730, 353 (2014); V.A.\nKosteleck´ y, N. Russell, and J.D. Tasson, Phys. Rev. Lett. 100, 111102\n(2008).\n21. R. Lehnert, W.M. Snow, Z. Xiao, and R. Xu, Phys. Lett. B 772, 865 (2017);\nJ. Foster et al., Phys. Rev. D 95, 084033 (2017).\n22. V.A. Kosteleck´ y, R. Lehnert, and M.J. Perry, Phys. Rev. D68, 123511\n(2003); V.A. Kosteleck´ y and R. Potting, Phys. Rev. D 104, 104046 (2021).\n23. V.A. Kosteleck´ y, R. Lehnert, N. McGinnis, M. Schreck, a nd B. Seradjeh,\nPhys.Rev.Res. 4, 023106 (2022); A.G´ omez, A. Mart´ ın-Ruiz, andL. Urrutia,\nPhys. Lett. B 829, 137043 (2022)." }, { "title": "2308.01575v1.Quasinormal_modes_of_the_spherical_bumblebee_black_holes_with_a_global_monopole.pdf", "content": "Quasinormal modes of the spherical bumblebee black holes with a\nglobal monopole\nRui-Hui Lin,1,∗Rui Jiang,1and Xiang-Hua Zhai1,†\n1Division of Mathematics and Theoretical Physics,\nShanghai Normal University, 100 Guilin Road, Shanghai 200234, China\nAbstract\nThe bumblebee model is an extension of the Einstein-Maxwell theory that allows for the sponta-\nneous breaking of the Lorentz symmetry of the spacetime. In this paper, we study the quasinormal\nmodes of the spherical black holes in this model that are characterized by a global monopole. We\nanalyze the two cases with a vanishing cosmological constant or a negative one (the anti-de Sitter\ncase). We find that the black holes are stable under the perturbation of a massless scalar field.\nHowever, both the Lorentz symmetry breaking and the global monopole have notable impacts on\nthe evolution of the perturbation. The Lorentz symmetry breaking may prolong or shorten the\ndecay of the perturbation according to the sign of the breaking parameter. The global monopole,\non the other hand, has different effects depending on whether a nonzero cosmological constant\npresences: it reduces the damping of the perturbations for the case with a vanishing cosmological\nconstant, but has little influence for the anti-de Sitter case.\n∗linrh@shnu.edu.cn\n†(corresponding author)zhaixh@shnu.edu.cn\n1arXiv:2308.01575v1 [gr-qc] 3 Aug 2023I. INTRODUCTION\nOne of the major challenges in theoretical physics is to reconcile Einstein’s general rel-\nativity (GR), the most successful theory of gravitation so far, with the standard model of\nparticle physics (SM), which unifies all other interactions. A possible clue to this problem is\nspontaneous symmetry breaking, which plays a key role in the elementary particle physics.\nIn the early universe, the temperature may have been high enough to trigger symmetry\nbreaking. Depending on the topology of the manifold Mof degenerate vacua, this process\nmay result in different types of topological defects[1, 2]. In particular, when second homo-\ntopy group π2MofMis nontrivial, point defects, i.e., monopoles are formed and can be the\nsource of inflation and seeds of the cosmic structures[3, 4]. A simplest model to describe a\nglobal monopole is to consider a triplet of scalar fields with a global O(3) symmetry sponta-\nneously broken to U(1). The gravitational field of a global monopole describes a spacetime\nwith a solid angle deficit[3, 5]. If such a global monopole is captured by an ordinary black\nhole, a new type of black hole with a global monopole charge will be formed. This charge\nmakes the black hole topologically different from the usual one, and may have interesting\nphysical implications (see, e.g., Refs. [6–14]).\nAnother possible symmetry breaking that may emerge in the quest of quantization of GR\nis the breaking of Lorentz symmetry. It is shown that this symmetry may be strongly vio-\nlated at the Planck scale ( ∼1019GeV) in some approaches to quantum gravity (QG)[15, 16],\nsuggesting that it may not be a fundamental symmetry of nature. Moreover, Lorentz symme-\ntry breaking (LSB) may also provide candidates of observable signals of the underlying QG\nframework at low energy scale [17–19]. Therefore, the occurrence of LSB has been explored\nin various scenarios, such as string theory (see, e.g., [17]), loop quantum gravity[20, 21],\nmassive gravity[22], Einstein-aether theory[23], and others[24–27]. The effective field theory\nto study violations of fundamental symmetries in SM and gravity is called the Standard-\nModel Extension (SME)[28, 29]. The gravitational couplings of SME is explored in a general\nRiemann-Cartan spacetime in Ref. [19], where it is shown that gravitational coupling spon-\ntaneous LSB can occur without geometrical incompatibilities.\nThe bumblebee model is a simple example to incorporate spontaneous LSB in the gravi-\ntational sector. It is a vector-tensor theory that generalizes the Einstein-Maxwell theory by\nintroducing a vector field Bµthat acquires a non-zero vacuum expectation value (VEV) [17–\n219, 30]. This vector field Bµcouples nonminimally to the spacetime and thus induces a spon-\ntaneous breaking of the local Lorentz symmetry by selecting a preferred spacetime direction\nin the local frames through its frozen VEV. When the bumblebee field is set to be a constant\nbackground, it can be viewed as the coefficients related to LSB in SME[31]. The bumblebee\nmodel has attracted recent interests since an exact Schwarzschild-like black hole solution\nwas found[32]. Other static[33–35] and rotating[36–38] solutions in the bumblebee model\nhave been obtained. The various physical aspects such as the gravitational lensing[39, 40],\nshadow[32, 36, 41], accretion[42], perturbations[43–49], and superradiance[50–52], have also\nbeen studied.\nIn particular, by considering the bumblebee model with a global monopole, G¨ ull¨ u and\n¨Ovg¨ un constructed a family of Schwarzschild-like black hole solutions and analyzed the in-\nfluence of the LSB parameter on the black hole shadow and the weak deflection angle[33].\nFurthermore, in the study of the generalized uncertainty principle correction of the bum-\nblebee black holes, Gogoi and Goswami proposed that a similar family of black hole so-\nlutions with both a global monopole and a non-vanishing cosmological constant can also\nbe constructed[49]. These two families of black holes, as strong field solutions, provide\noptimal environments for investigating spontaneous LSB and global monopoles within the\ngravitational context. When black holes are perturbed, the resulting oscillations and their\nevolution may reveal unique characteristics of the black hole, or equivalently, the spacetime\nand gravity. These oscillating modes are referred to as quasi-normal modes (QNMs) and\nhave become a widely utilized tool for analyzing the stability of black hole spacetime since\ntheir use in demonstrating the stability of the Schwarzschild black hole[53, 54]. In the cur-\nrent paper, we focus on the QNMs of aforementioned two families of black hole solutions,\nnamely the spherical bumblebee black holes with a global monopole in the presence or ab-\nsence of a cosmological constant. Our aim is to study how the spontaneous LSB and the\nglobal monopole affect the stability of the black hole spacetime.\nThis paper is structured as follows. In Section II, we provide a brief overview of spherical\nblack holes in the bumblebee model with a global monopole and establish the configuration\nof perturbation by considering an impinging massless scalar field. In Section III, we examine\nthe QNMs of these spherical bumblebee black holes with a global monopole for vanishing\nand non-vanishing cosmological constant. Our study is concluded in Section IV. Throughout\nthe paper, we follow the metric convention (+ ,−,−,−) and use the units G=ℏ=c= 1.\n3II. SPHERICAL BUMBLEBEE BLACK HOLES WITH A GLOBAL MONOPOLE\nAND THEIR SCALAR PERTURBATION\nA. Spherical bumblebee black holes with a global monopole\nThe Lagrangian of the bumblebee model is generally written as[32, 34]\nLB=√−g\u00141\n16π(R−2Λ + ξBµBνRµν)−1\n4BµνBµν−V\u0000\nBµBµ±b2\u0001\u0015\n+LM,(1)\nwhere Bµνdenotes the strength of the bumblebee field Bµand is defined as\nBµν=∂µBν−∂νBµ. (2)\nThe term with the parameter ξrepresents the quadratic coupling between the bumblebee\nfield and the Ricci tensor Rµν. Λ is the cosmological constant and LMis the Lagrangian of\nmatter. The term V(BµBµ±b2) refers to the potential of the bumblebee field that takes\na minimum at BµBµ±b2= 0 with b2>0. The ±signs here associate to a spacelike or\ntimelike bumblebee field, respectively. For a stable vacuum spacetime, one would expect\nthat the potential Vwill be minimized and the bumblebee field Bµwill be frozen at the\nVEV⟨Bµ⟩=bµwith bµbµ=∓b2. Due to the coupling featured by the parameter ξ, such a\nnonzero VEV of Bµthen will break Lorentz symmetry by selecting a preferred direction of\nspacetime. Variation of the Lagrangian (1) with respect to the metric gµνgives the equation\nof motion\n0 =Rµν+ Λgµν−κ\u0012\nT(M)\nµν−1\n2gµνT(M)\u0013\n+ξbµbαRαν\n+ξbνbαRαµ−ξ\n2gµνbαbβRαβ−ξ\n2∇α∇µ(bαbν)\n−ξ\n2∇α∇ν(bαbµ) +ξ\n2∇2(bµbν) +κ(gµνb2−2bµbν)V′,(3)\nwhere κ= 8πandV′denotes d V(x)/dx. For a global monopole field, the energy-momentum\ntensor T(M)\nµνhas the form[3]\nT(M)ν\nµ = diag\u0012η2\nr2,η2\nr2,0,0\u0013\n, (4)\nwhere ηis a constant related to the global monopole charge. For the usual spherical ansatz\nof metric\nds2=f(r)dt2−dr2\nh(r)−r2dθ2−r2sin2θdϕ2(5)\n4where f(r) and h(r) are the metric functions of the radial coordinate r, it is assumed that\nthe bumblebee field has a purely radial form bµ= (0, br(r),0,0), and hence, BµBµ=−b2\nr.\nIn the scenario with a vanishing Λ, a spherical solution can be constructed when the\npotential Vsatisfies V′= 0 at its minimum where BµBµ+b2= 0[32, 33]. This solution is\nshown to be Schwarzschild-like with a global monopole and can be given by[33]\nf(r) =1−κη2−2M\nr,\nh(r) =L+ 1\nf(r),\nbr(r) =|b|p\nh(r),(6)\nwhere L≡ξb2is the LSB parameter, and Mis an integration constant corresponding to\nthe Arnowitt-Deser-Misner mass. The event horizon can be easily found to be at rh=\n2M/(1−κη2), independent of the bumblebee field.\nFor the case where the cosmological constant Λ is nonzero, it is suggested that solutions\ncan be constructed by setting[34]\nV\u0000\nBµBµ+b2\u0001\n=λ\n2\u0000\nBµBµ+b2\u0001\n, (7)\nand viewing this term as a Lagrange multiplier term that ensures BµBµ+b2= 0 in the\nvariation of the Lagrangian (1). Following this scheme, we check that the Schwarzschild-\n(a)dS-like bumblebee black hole solution with a global monopole proposed in Ref. [49] can\nbe constructed, which is given by\nf(r) =1−κη2−2M\nr−Λ\n3r2,\nh(r) =L+ 1\nf(r),\nbr(r) =|b|p\nh(r),\nλ=ξΛ\nκ(L+ 1).(8)\nB. Scalar perturbation\nWe study the perturbation of the spherical bumblebee black holes (6) and (8) by con-\nsidering an impinging massless scalar field Φ. We restrict our attention to the cases with\n5Λ≤0, as the cases with Λ >0 have a cosmological horizon that limits the causal structure\nof the spacetime.\nWe assume that Φ is minimally coupled to gravity and satisfies the covariant Klein-Gordon\nequation,\n1√−g∂µ\u0000\ngµν√−g∂νΦ\u0001\n= 0. (9)\nUsing the separation of Φ in the spherical coordinates ( t, r, θ, ϕ ),\nΦ =ψl(r)\nrYlm(θ, ϕ)e−iωt, (10)\nwhere Ylm(θ, ϕ) is the spherical harmonic function, we can obtain the radial equation of Eq.\n(9) as\n1\n1 +Lfd\ndr\u0012\nfdψl\ndr\u0013\n+\u0002\nω2−U(r)\u0003\nψl= 0. (11)\nThe effective potential U(r) is defined as\nU(r) =f(r)\u0014l(l+ 1)\nr2+f′(r)\n(1 +L)r\u0015\n. (12)\nIntroducing the tortoise coordinate xdefined by\ndx\ndr=√\n1 +L\nf(r), (13)\none can rewrite the radial equation (11) as\nd2ψl\ndx2+\u0002\nω2−U(x)\u0003\nψl= 0. (14)\nTo keep track of the temporal evolution of Φ that is encoded in ωin the radial equation\n(11), one can, alternatively, adopt a different separation of Φ as\nΦ =Ψ(t, r)\nrYlm(θ, ϕ), (15)\nwhere its dependence on randtare kept in a single function Ψ( t, r). Then, the equation\nfor Ψ from Eq. (9) is\u0014∂2\n∂x2−∂2\n∂t2−U(x)\u0015\nΨ(x, t) = 0 . (16)\nWith the light cone coordinates\ndu=dt−dx, dv =dt+dx, (17)\n6L=-0.5L=0L=0.5L=1\n-20 0 20 40 60 800.000.020.040.060.080.100.120.14\nxUκη2=0.2\nκη2=0κη2=0.2κη2=0.4κη2=0.6\n-20 -10 0 10 20 30 40 500.000.050.100.150.200.25\nxUL=0.5FIG. 1. The effective potential for scalar perturbations with different Landκη2when l= 2, where\nwe have set M= 1 and Λ = 0.\nL=-0.5L=0L=0.5L=1\n0 2 4 6 8010203040\nxUκη2=0.2\nκη2=0κη2=0.2κη2=0.4κη2=0.6\n0 2 4 6 8010203040\nxUL=0.5\nFIG. 2. The effective potential for scalar perturbations with different Landκη2when l= 2, where\nwe have set M= 1 and Λ = −3.\none can rewrite Eq.(16) as\n\u0014\n4∂2\n∂u∂v+U(u, v)\u0015\nΨ(u, v) = 0 . (18)\nGenerally, based on different separations, either Eq. (14) or (18) can be solved numerically\nonce the boundary conditions are set, which we will do for the bumblebee black holes in\nthe next section. Here we concentrate on the effective potential Udefined in Eq. (12). The\nbehaviors of Ufor vanishing and negative Λ are shown in Figs. 1 and 2, respectively. The\ntortoise coordinate xranges from −∞ at the event horizon to + ∞at spatial infinity. For\nΛ = 0, the effective potential tends to zero both near the horizon and at large distances\nfrom the black hole. For Λ <0, however, the effective potential diverges at spatial infinity.\n7The effects of the LSB parameter Land the global monopole parameter κη2on the effective\npotential are evident.\nIII. QNMS OF THE BUMBLEBEE BLACK HOLES WITH A GLOBAL MONOPOLE\nA. The black hole with a vanishing cosmological constant\nAs seen in the previous section, the effective potential Ufor the scalar field approaches\nzero at x→ ±∞ in this case. Therefore, together with the physical consideration that the\nfield should be purely outgoing towards spatial infinity and purely ingoing at the horizon,\nthe asymptotic forms of the field can be written as\nψl∼\n\ne−iωx, x→ −∞ ,\neiωx, x→+∞.(19)\nThe radial equation (14) and the boundary condition (19) determine a discrete spectrum\nof frequencies {ωn}for a fixed pair of symmetry breaking parameters Landκη2. We use\nthe continued-fraction method[55] to solve this problem numerically. The frequency ωis\ngenerally complex, with a real part ωRcorresponding to the oscillation frequency of the field\nand an imaginary part ωIreflecting the evolution of the wave amplitude. A positive ωI\nimplies a growing scalar field with a growth rate of ωI. A negative ωIimplies a decaying\nscalar field with a damping rate of |ωI|.\nThe QNM frequencies as functions of κη2for different Lare shown in Fig. 3. The left\npanel shows that the real part of the frequencies, ωR, decreases with increasing Lorκη2. The\nright panel shows that the imaginary part of the frequencies, ωI, is negative for all values\nofLandκη2, indicating that the black hole is stable under the perturbation of a massless\nscalar field. However, we also observe that |ωI|becomes smaller as Lorκη2increases. This\nimplies that for the black hole with larger Lorκη2, the perturbation takes longer time to\ndecay away. The effect of Lon the stability in this case agrees with the results of other\nbumblebee black holes reported in Refs. [43, 44]. We note that κη2≥0, so the perturbation\nto the black hole without a global monopole ( κη2= 0) fades out more quickly than the case\nwith one ( κη2>0).\nWe also investigate the temporal evolution of the perturbation by solving Eq. (18). We\n8L=-0.5L=0L=0.5L=1\n0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.70.00.10.20.30.40.5\nκη2ωR\nL=-0.5L=0L=0.5L=1\n0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.70.000.020.040.060.080.100.120.14\nκη2-ωIFIG. 3. The real part (left) and imaginary part (right) of QNM frequencies of the spherical\nbumblebee black holes with a global monopole and a vanishing Λ, where we have set M= 1, l= 2\nandn= 0.\nset the initial conditions as\nΨ(u,0) = 0 , Ψ(0, v) = 1 . (20)\nThe numerical results are displayed in Fig. 4. They confirm that the black hole is stable\nunder the massless scalar field perturbation. The left panel shows the decay of the pertur-\nbation for different Lwith κη2= 0.2 and l= 2, in which it is obvious that the scalar field\ndecays slower for larger L. The right panel shows the decay of the perturbation for different\nκη2with L= 0.5 and l= 2, where one can see that the scalar field decays slower for larger\nκη2. This agrees with the results in the frequency domain. Another consistent observation\nfrom Figs. 3 and 4 is that the oscillation frequencies are more sensitive to the monopole\ncharges η2than to the LSB parameters L.\nWe use the Prony method[56] to extract the QNM frequencies from the time domain\nprofile and compare them with the results from the continued-fraction method. For instance,\nwe find the frequency ω= 0.342031 −0.0504228 iforκη2= 0.2, L= 0.5, l= 2 by using the\nProny method. The continued-fraction method gives ω= 0.342003 −0.0504359 ifor the\nsame case, which has an error less than one thousandth. Fig. 5 shows the comparison of the\nQNMs obtained by the two methods for different κη2, which demonstrates a good agreement\nbetween the two methods.\n9L=-0.5L=0L=0.5L=1150 200 250 300 35010-810-50.01\nt|Ψ|\nκη2=0κη2=0.2κη2=0.4κη2=0.6\n150 200 250 300 35010-810-50.01\nt|Ψ|FIG. 4. Temporal evolution of the scalar perturbations, with κη2= 0.2, l= 2 in the left panel and\nL= 0.5, l= 2 in the right panel.\n0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.70.00.10.20.30.40.5\nκη2ωR\n0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.70.000.020.040.060.08\nκη2-ωI\nFIG. 5. Comparison of the QNM frequencies obtained by the continued-fraction method (blue line)\nand those extracted from time domain analysis (red dots), with M= 1, L= 0.5, l= 2.\nB. The black hole with a negative cosmological constant\nFor Λ <0, the effective potential Uwill diverge at spatial infinity, as shown in Fig. 2.\nThis requires that Φ = 0 at spatial infinity. Therefore, the asymptotic forms of Ψ in this\ncase should be\nΨ∼\n\ne−iωx, x→ −∞ ,\n0 , x→+∞.(21)\nWe apply the Horowitz-Hubeny method[7] to solve Eq. (14) for the spectrum of frequencies\nwith these Dirichlet boundary conditions. The resulting QNM frequencies as functions of κη2\nfor various Lare depicted in Fig. 6. The left panel reveals that ωRdecreases with increasing\nLorκη2, implying that the scalar field oscillates at a lower frequency for larger Lorκη2.\n10L=-0.5\nL=0\nL=1\n0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.72.02.53.03.54.04.55.0\nκη2ωR\nL=-0.5\nL=0\nL=1\n0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.70.00.51.01.52.02.53.03.5\nκη2-ωIFIG. 6. The real part (left) and imaginary part (right) of QNM frequencies of the spherical\nbumblebee black holes with a global monopole and non-zero Λ, where we have set Λ = −3, l=\n2, n= 0 and a suitable Msuch that the horizon radius r+= 1.\nThe right panel demonstrates that ωIis negative for all values of Landκη2, indicating the\nstability of the bumblebee black hole under massless scalar field perturbations. Furthermore,\nwe observe that |ωI|decreases with increasing L, suggesting that the perturbation to the\nblack hole takes longer time to decay away for larger L. On the other hand, |ωI|is insensitive\ntoκη2, which differs significantly from the zero-Λ scenario.\nTo compute the temporal evolution of the QNMs in this setting, we need to limit the\ncalculation at a finite x=xmax, since the potential diverges at spatial infinity. This implies\nthat the calculation domain on the u-vplane is bounded by the line v−u= 2xmax[57], as\nillustrated in Fig. 7. Hence, we can impose the boundary condition\nΨ(v−u= 2xmax) = 0 . (22)\nFig. 8 illustrates the temporal evolution of scalar perturbations around the bumblebee\nblack hole with a global monopole and nonzero Λ for various values of Landκη2. The black\nhole is stable under massless scalar field perturbations, as evident from the decay of the\nscalar field. The left panel displays the evolution of the scalar perturbation for various L\nwith κη2= 0.2 and l= 2. The scalar field obviously decays slower for larger L, agreeing with\nthe results shown in Fig. 6. The right panel exhibits the evolution of scalar perturbation\nfor various κη2with L= 0.5 and l= 2. The decay speed of the scalar field is very similar\nfor different values of κη2.\n11FIG. 7. Schematic of the numerical grid and the calculation domain. The black dots denote the\ngrid points where the field values are given by the initial condition and the boundary condition.\nThe blue dots denote the grid points to be computed.[57]\nL=-0.5L=0L=1\n7.5 10.0 12.5 15.0 17.5 20.010-1010-810-610-4\nt|Ψ|\nκη2=0κη2=0.2κη2=0.4κη2=0.6\n6 7 8 9 10 11 1210-910-710-50.001\nt|Ψ|\nFIG. 8. Temporal evolution of the scalar perturbations in the scenario with nonzero Λ, where\nκη2= 0.2, l= 2 in the left panel and L= 0.5, l= 2 in the right panel.\nWe again employ the Prony method to extract the QNM frequencies from the time\ndomain profile. To verify the robustness of our calculations, in addition to the previously\nshown cases of l= 2, we use the cases of l= 0 as the examples to cross check the results\nfrom frequency and time domains, which are presented in Fig. 9. The lines correspond to\nthe Horowitz–Hubeny method, and the red dots represent the frequencies extracted by the\nProny method. The results exhibit good agreement between the two methods, confirming\nthe reliability of our results.\n12L=-0.5\nL=0\nL=1\n0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.701234\nκη2ωR\nL=-0.5\nL=0\nL=1\n0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.70123\nκη2-ωIFIG. 9. Comparison of the QNM frequencies obtained by the Horowitz–Hubeny method (solid\nlines) and those extracted from time domain analysis (red dots), with Λ = −3, l= 0 and Mchosen\nsuch that the horizon radius r+= 1.\nIV. CONCLUSION AND DISCUSSION\nWe have investigated the stability of the spherical bumblebee black holes with a global\nmonopole in the two cases with a zero or a negative cosmological constant by examining the\nQNMs of the black holes under the perturbations of a massless scalar field. The analyses are\nperformed in both frequency domain and time domain by using different numerical methods.\nWe find that both families of black holes are stable under the perturbations of a massless\nscalar field. In particular, the scalar field around the black holes decays slower for larger\nLSB parameter L. On the other hand, the influence of the global monopole on the stability\nof two types of black holes is different. For the black hole with zero Λ, the global monopole\nfield prolongs the decay of the perturbation. Whereas for the black hole with negative Λ,\nthe effect of the global monopole on the stability is negligible.\nSuch a significant difference may be understood as follows. QNM frequencies are eigen-\nvalues ωthat satisfy the boundary conditions corresponding to the outgoing wave at spatial\ninfinity and the ingoing wave at the horizon. Thus, computing QNM frequencies is an eigen-\nvalue problem that depends on the boundary condition at spatial infinity. When Λ = 0, the\nboundary condition (19) at spatial infinity depends on both Landκη2through the tortoise\ncoordinate (13). Hence, both Landκη2affect QNM frequencies in this case. However,\nwhen Λ <0, the boundary condition is a Dirichlet type that is independent of Landκη2\ndue to the divergent potential. Therefore, κη2has little effect on the QNM frequencies.\n13Nevertheless, Lstill affects the QNM frequencies because it breaks the symmetry between\nspace and time. Moreover, the QNMs are derived by analyzing the outgoing wave signals of\nthe perturbation, which propagates differently depending on Lbecause the temporal-radial\nsymmetry of the metric is broken. The global monopole, on the other hand, only manifests\nitself in a solid deficit angle in the spacetime, which does not contribute to the breaking of\nthe temporal-radial symmetry. This can also be seen by rewritting the effective potential\nU(r) in Eq. (12) for large ras\nU(r) = (1 −κη2−2M\nr+r2)\u0014l(l+ 1)\nr2+2M\n(1 +L)r3+2\n1 +L\u0015\n=2r2\n1 +L+\u00142(1−κη2)\n1 +L+l(l+ 1)\u0015\n−2M\n(1 +L)r+O(1\nr2).(23)\nOne can see that Lcontributes to the coefficient of the r2term, while κη2only affects the\nconstant term. Therefore, the global monopole does not alter the shape of U(r) significantly\nat large rand hence, the stability of the black hole is almost independent of the global\nmonopole.\nIn the high energy regime, the wavelength of the massless scalar field is negligible com-\npared to the horizon scale of the black hole. Thus, the field follows the null geodesics, which\nare influenced by the global monopole and the LSB parameter[33]. These factors may also\naffect the absorption and scattering cross sections of the bumblebee black holes. This is an\ninteresting topic for future research.\nACKNOWLEDGEMENT\nThis work is supported by the National Science Foundation of China under Grant No.\n12105179.\n[1] T. W. B. Kibble, “Topology of Cosmic Domains and Strings,” J. Phys. A 9, 1387–1398 (1976).\n[2] A. Vilenkin, “Cosmic Strings and Domain Walls,” Phys. Rept. 121, 263–315 (1985).\n[3] M. Barriola and A. Vilenkin, “Gravitational Field of a Global Monopole,” Phys. Rev. Lett.\n63, 341 (1989).\n14[4] D. P. Bennett and S. H. Rhie, “Cosmological evolution of global monopoles and the origin of\nlarge scale structure,” Phys. Rev. Lett. 65, 1709–1712 (1990).\n[5] X. Shi and X.-z. Li, “The Gravitational field of a global monopole,” Class. Quant. Grav. 8,\n761–767 (1991), arXiv:0903.3085 [gr-qc].\n[6] G. W. Gibbons, “Selfgravitating magnetic monopoles, global monopoles and black holes,”\nLect. Notes Phys. 383, 110–138 (1991), arXiv:1109.3538 [gr-qc].\n[7] G. T. Horowitz and V. E. Hubeny, “Quasinormal modes of AdS black holes and the approach\nto thermal equilibrium,” Phys. Rev. D 62, 024027 (2000), arXiv:hep-th/9909056.\n[8] X.-z. Li and J.-g. Hao, “Global monopole in asymptotically dS / AdS space-time,” Phys. Rev.\nD66, 107701 (2002), arXiv:hep-th/0210050.\n[9] H.-w. Yu, “Decay of massive scalar hair in the background of a black hole with a global\nmonopole,” Phys. Rev. D 65, 087502 (2002), arXiv:gr-qc/0201035.\n[10] H. Watabe and T. Torii, “Perturbations of global monopoles as a black hole’s hair,” JCAP\n02, 001 (2004), arXiv:gr-qc/0307074.\n[11] S. Chen, L. Wang, C. Ding, and J. Jing, “Holographic superconductors in the AdS black\nhole spacetime with a global monopole,” Nucl. Phys. B 836, 222–231 (2010), arXiv:0912.2397\n[gr-qc].\n[12] C. L. A. Rizwan, A. N. Kumar, and K. S. Ananthram, “Effect of Global Monopole on the\nMicroscopic Structure of RN-AdS Black Hole,” Springer Proc. Phys. 248, 81–85 (2020).\n[13] S. Soroushfar and S. Upadhyay, “Phase transition of a charged AdS black hole with a\nglobal monopole through geometrical thermodynamics,” Phys. Lett. B 804, 135360 (2020),\narXiv:2003.06714 [gr-qc].\n[14] Z. Luo, H. Yu, and J. Li, “Effects of a global monopole on the thermodynamic phase transition\nof a charged AdS black hole*,” Chin. Phys. C 46, 125101 (2022), arXiv:2206.09729 [gr-qc].\n[15] D. Mattingly, “Modern tests of Lorentz invariance,” Living Rev. Rel. 8, 5 (2005), arXiv:gr-\nqc/0502097.\n[16] G. Amelino-Camelia, “Quantum-Spacetime Phenomenology,” Living Rev. Rel. 16, 5 (2013),\narXiv:0806.0339 [gr-qc].\n[17] V. A. Kostelecky and S. Samuel, “Spontaneous Breaking of Lorentz Symmetry in String\nTheory,” Phys. Rev. D 39, 683 (1989).\n[18] V. A. Kostelecky and R. Potting, “CPT and strings,” Nucl. Phys. B 359, 545–570 (1991).\n15[19] V. A. Kostelecky, “Gravity, Lorentz violation, and the standard model,” Phys. Rev. D 69,\n105009 (2004), arXiv:hep-th/0312310.\n[20] J. R. Ellis, N. E. Mavromatos, and D. V. Nanopoulos, “Quantum gravitational diffusion and\nstochastic fluctuations in the velocity of light,” Gen. Rel. Grav. 32, 127–144 (2000), arXiv:gr-\nqc/9904068.\n[21] R. Gambini and J. Pullin, “Nonstandard optics from quantum space-time,” Phys. Rev. D 59,\n124021 (1999), arXiv:gr-qc/9809038.\n[22] S. Fernando and T. Clark, “Black holes in massive gravity: quasi-normal modes of scalar\nperturbations,” Gen. Rel. Grav. 46, 1834 (2014), arXiv:1411.6537 [gr-qc].\n[23] T. Jacobson, “Einstein-aether gravity: A Status report,” PoS QG-PH , 020 (2007),\narXiv:0801.1547 [gr-qc].\n[24] N. Arkani-Hamed, H.-C. Cheng, M. A. Luty, and S. Mukohyama, “Ghost condensation and\na consistent infrared modification of gravity,” JHEP 05, 074 (2004), arXiv:hep-th/0312099.\n[25] C. P. Burgess, J. M. Cline, E. Filotas, J. Matias, and G. D. Moore, “Loop generated bounds\non changes to the graviton dispersion relation,” JHEP 03, 043 (2002), arXiv:hep-ph/0201082.\n[26] J. M. Cline and L. Valcarcel, “Asymmetrically warped compactifications and gravitational\nLorentz violation,” JHEP 03, 032 (2004), arXiv:hep-ph/0312245.\n[27] A. R. Frey, “String theoretic bounds on Lorentz violating warped compactification,” JHEP\n04, 012 (2003), arXiv:hep-th/0301189.\n[28] S. M. Carroll, J. A. Harvey, V. A. Kostelecky, C. D. Lane, and T. Okamoto, “Noncommutative\nfield theory and Lorentz violation,” Phys. Rev. Lett. 87, 141601 (2001), arXiv:hep-th/0105082.\n[29] C. D. Lane, “Noncommutative Gravity and the Standard-Model Extension,” in 8th Meeting\non CPT and Lorentz Symmetry (2020) pp. 33–36, arXiv:1906.01008 [hep-ph].\n[30] R. Bluhm and V. A. Kostelecky, “Spontaneous Lorentz violation, Nambu-Goldstone modes,\nand gravity,” Phys. Rev. D 71, 065008 (2005), arXiv:hep-th/0412320.\n[31] V. A. Kostelecky and M. Mewes, “Electrodynamics with Lorentz-violating operators of arbi-\ntrary dimension,” Phys. Rev. D 80, 015020 (2009), arXiv:0905.0031 [hep-ph].\n[32] R. Casana, A. Cavalcante, F. P. Poulis, and E. B. Santos, “Exact Schwarzschild-like solution\nin a bumblebee gravity model,” Phys. Rev. D 97, 104001 (2018), arXiv:1711.02273 [gr-qc].\n[33] I. G¨ ull¨ u and A. ¨Ovg¨ un, “Schwarzschild-like black hole with a topological defect in bumblebee\ngravity,” Annals Phys. 436, 168721 (2022), arXiv:2012.02611 [gr-qc].\n16[34] R. V. Maluf and J. C. S. Neves, “Black holes with a cosmological constant in bumblebee\ngravity,” Phys. Rev. D 103, 044002 (2021), arXiv:2011.12841 [gr-qc].\n[35] C. Ding, X. Chen, and X. Fu, “Einstein-Gauss-Bonnet gravity coupled to bumblebee field in\nfour dimensional spacetime,” Nucl. Phys. B 975, 115688 (2022), arXiv:2102.13335 [gr-qc].\n[36] C. Ding, C. Liu, R. Casana, and A. Cavalcante, “Exact Kerr-like solution and its shadow\nin a gravity model with spontaneous Lorentz symmetry breaking,” Eur. Phys. J. C 80, 178\n(2020), arXiv:1910.02674 [gr-qc].\n[37] C. Ding and X. Chen, “Slowly rotating Einstein-bumblebee black hole solution and its grey-\nbody factor in a Lorentz violation model,” Chin. Phys. C 45, 025106 (2021), arXiv:2008.10474\n[gr-qc].\n[38] S. K. Jha and A. Rahaman, “Bumblebee gravity with a Kerr-Sen-like solution and its Shadow,”\nEur. Phys. J. C 81, 345 (2021), arXiv:2011.14916 [gr-qc].\n[39] A. Ovg¨ un, K. Jusufi, and I. Sakalli, “Gravitational lensing under the effect of Weyl and\nbumblebee gravities: Applications of Gauss–Bonnet theorem,” Annals Phys. 399, 193–203\n(2018), arXiv:1805.09431 [gr-qc].\n[40] I. D. D. Carvalho, G. Alencar, W. M. Mendes, and R. R. Landim, “The gravitational bending\nangle by static and spherically symmetric black holes in bumblebee gravity,” EPL 134, 51001\n(2021), arXiv:2103.03845 [gr-qc].\n[41] H.-M. Wang and S.-W. Wei, “Shadow cast by Kerr-like black hole in the presence of plasma\nin Einstein-bumblebee gravity,” Eur. Phys. J. Plus 137, 571 (2022), arXiv:2106.14602 [gr-qc].\n[42] C. Liu, C. Ding, and J. Jing, “Thin accretion disk around a rotating Kerr-like black hole in\nEinstein-bumblebee gravity model,” arXiv (2019), arXiv:1910.13259 [gr-qc].\n[43] S. Kanzi and I. Sakallı, “Greybody radiation and quasinormal modes of Kerr-like black hole\nin Bumblebee gravity model,” Eur. Phys. J. C 81, 501 (2021), arXiv:2102.06303 [hep-th].\n[44] R. Oliveira, D. M. Dantas, and C. A. S. Almeida, “Quasinormal frequencies for a black hole\nin a bumblebee gravity,” EPL 135, 10003 (2021), arXiv:2105.07956 [gr-qc].\n[45] Z. Wang, S. Chen, and J. Jing, “Constraint on parameters of a rotating black hole in\nEinstein-bumblebee theory by quasi-periodic oscillations,” Eur. Phys. J. C 82, 528 (2022),\narXiv:2112.02895 [gr-qc].\n[46] R. V. Maluf and C. R. Muniz, “Comment on “Greybody radiation and quasinormal modes\nof Kerr-like black hole in Bumblebee gravity model”,” Eur. Phys. J. C 82, 94 (2022),\n17arXiv:2202.01015 [gr-qc].\n[47] S. Kanzi and I. Sakallı, “Reply to “Comment on ‘Greybody radiation and quasinormal modes\nof Kerr-like black hole in Bumblebee gravity model”’,” Eur. Phys. J. C 82, 93 (2022),\narXiv:2202.11708 [gr-qc].\n[48] W. Liu, X. Fang, J. Jing, and J. Wang, “QNMs of slowly rotating Einstein–Bumblebee black\nhole,” Eur. Phys. J. C 83, 83 (2023), arXiv:2211.03156 [gr-qc].\n[49] D. J. Gogoi and U. D. Goswami, “Quasinormal modes and Hawking radiation sparsity of GUP\ncorrected black holes in bumblebee gravity with topological defects,” JCAP 06, 029 (2022),\narXiv:2203.07594 [gr-qc].\n[50] R. Jiang, R.-H. Lin, and X.-H. Zhai, “Superradiant instability of a Kerr-like black hole in\nEinstein-bumblebee gravity,” Phys. Rev. D 104, 124004 (2021), arXiv:2108.04702 [gr-qc].\n[51] M. Khodadi, “Black Hole Superradiance in the Presence of Lorentz Symmetry Violation,”\nPhys. Rev. D 103, 064051 (2021), arXiv:2103.03611 [gr-qc].\n[52] M. Khodadi, “Magnetic reconnection and energy extraction from a spinning black hole with\nbroken Lorentz symmetry,” Phys. Rev. D 105, 023025 (2022), arXiv:2201.02765 [gr-qc].\n[53] T. Regge and J. A. Wheeler, “Stability of a Schwarzschild singularity,” Phys. Rev. 108, 1063–\n1069 (1957).\n[54] C. V. Vishveshwara, “Stability of the schwarzschild metric,” Phys. Rev. D 1, 2870–2879 (1970).\n[55] E. W. Leaver, “An Analytic representation for the quasi normal modes of Kerr black holes,”\nProc. Roy. Soc. Lond. A 402, 285–298 (1985).\n[56] E. Berti, V. Cardoso, J. A. Gonzalez, and U. Sperhake, “Mining information from binary\nblack hole mergers: A Comparison of estimation methods for complex exponentials in noise,”\nPhys. Rev. D 75, 124017 (2007), arXiv:gr-qc/0701086.\n[57] B. Wang, C. Molina, and E. Abdalla, “Evolving of a massless scalar field in Reissner-\nNordstrom Anti-de Sitter space-times,” Phys. Rev. D 63, 084001 (2001), arXiv:hep-\nth/0005143.\n18" }, { "title": "1610.02799v1.A_Five_Freedom_Active_Damping_and_Alignment_Device_Used_in_the_Joule_Balance.pdf", "content": "IEEE TRANS. INSTRUM. MEAS., CPEM2016, JUNE 10, 2021 1\nA Five-Freedom Active Damping and Alignment\nDevice Used in the Joule Balance\nJinxin Xu, Qiang You, Zhonghua Zhang, Zhengkun Li and Shisong Li, Member, IEEE\nAbstract —Damping devices are necessary for sup-\npressing the undesired coil motions in the watt/joule\nbalance. In this paper, an active electromagnetic damp-\ning device, located outside the main magnet, is in-\ntroduced in the joule balance project. The presented\ndamping device can be used in both dynamic and static\nmeasurement modes. With the feedback from a detec-\ntion system, five degrees of freedom of the coil, i.e. the\nhorizontal displacement x,yand the rotation angles \u0012x,\n\u0012y,\u0012z, can be controlled by the active damping device.\nHence, two functions, i.e. suppressing the undesired coil\nmotions and reducing the misalignment error, can be\nrealized with this active damping device. The principle,\nconstruction and performance of the proposed active\ndamping device are presented.\nIndex Terms —watt balance, joule balance, the\nPlanck constant, electromagnetic damping, misalign-\nment error.\nI. Introduction\nSeveral national metrology institutes (NMIs) across the\nworld are working hard on redefining one of the seven\nSI base units, the kilogram (kg), in terms of the Planck\nconstanth, by means of either the watt/joule balance\n[1-9] or the X-ray crystal density (XRCD) method [10].\nThe watt balance, which was proposed by Kibble in 1975\n[11], has been adopted by the majority of NMIs. The\nNational Institute of Metrology (NIM, China) is focusing\non the joule balance, which can be seen as an alternative\nrealization of the watt balance [12].\nThe unwanted coil motions in the watt/joule balance,\ne.g., horizontal and rotational movements, will lower the\nsignal to noise ratio in the measurement. With these\nmotions, the measurement would take a much longer time\nin order to achieve the type A relative uncertainty at\nan order of 10\u00008. Moreover, the noticeable undesired coil\nmotions will introduce a systematic bias [13]. Therefore,\ndamping devices are necessary to be employed to suppress\nthese undesired coil motions in the watt/joule balance. In\nwatt balances, e.g., NIST-4 at the National Institute of\nStandards and Technology (NIST, USA) [14], an active\ndamping system has been used during the measurement.\nThis work is supported by the China National Natural Science\nFoundation (Grant Nos.91536224, 51507088) and the China National\nKey Research and Development Plan (Grant No. 2016YFF0200102).\nJ. Xu, Q. You are with Tsinghua University, Beijing 100084, China.\nZ. Zhang and Z. Li are with the National Institute of Metrology\n(NIM), Beijing 100029, China and the Key Laboratory for the\nElectrical Quantum Standard of AQSIQ, Beijing 100029, China.\nS. Li is currently with the International Bureau of Weights and\nMeasures (BIPM), Pavillon de Breteuil, F-92312 S` evres Cedex,\nFrance. E-mail:leeshisong@sina.com.The damping coils in such systems are conventionally\nlocated on the framework of the suspended coil. To avoid\nadditional force in the weighing mode and unwanted in-\nduced voltage in the velocity mode, the damping device\nis turned off during both measurement modes. During the\ndynamic measurement mode, the undesired coil motions,\nhowever, is not taken care of. In the joule balance at NIM\nto suppress unwanted motions during the measurement\nwithout introducing any flux into the main magnet, a novel\nactive electromagnetic damping device, located outside\nthe main magnet, is designed and used. The merit of\nthe system is that the field of the damping device has\nno effect on the main field, and hence can be used in\nboth the dynamic and static measurement modes. With\nthe feedback from a detection system, five degrees of\nfreedom of the coil, i.e. the horizontal displacement x,y\nand the rotation angles \u0012x,\u0012y,\u0012z, can be controlled by\nthis device. Two functions, i.e. suppressing the undesired\ncoil motions and reducing the misalignment error, then\ncan be realized in the measurements. The original idea was\npresented in [15]. In this paper, the principle, construction\nand performance of the active electromagnetic damping\ndevice are presented.\nThe rest of this article is organized as follows: the\nprinciple of the joule balance is introduced in section\n2; the principle and construction of the active damping\ndevice are presented in section 3; the performance of the\nactive damping device in the joule balance is presented\nin section 4; some potential systematic effects using the\nactive damping device, the vertical force and the flux\nleakage at the mass weighing position, are discussed in\nsection 5.\nII. Principles of Joule Balance\nThe joule balance is the integral of the watt balance. The\nmathematical details are presented in [12]. The equation\nis expressed as\nZ\nLF\u0001dl+Z\nL\u001c\u0001d\u0012=I[ (B)\u0000 (A)]; (1)\nwhereFdenotes the magnetic force, \u001cthe torque relative\nto the mass center of the coil, Lthe vector trajectory\nwhen the coil is moved from position A to B, Ithe current\nthrough the coil, (B)\u0000 (A)the flux linkage difference\nof the position B and A.\nOn the right side of equation (1), the product of the flux\nlinkage difference (B)\u0000 (A)and the current Iis the\nmagnetic energy change. On the left side of equation (1),arXiv:1610.02799v1 [physics.ins-det] 10 Oct 2016IEEE TRANS. INSTRUM. MEAS., CPEM2016, JUNE 10, 2021 2\nFig. 1. The schematic diagram of the misalignment errors.\nthe two integral terms are the work done by the magnetic\nforce and torque. They can be expressed as\nwAB=Z\nLFzdz+Z\nLFxdx+Z\nLFydy\n+Z\nL\u001czd\u0012z+Z\nL\u001cxd\u0012x+Z\nL\u001cyd\u0012y:(2)\nNote that in equation (2) only the vertical force, i.e.\nFz, can be measured precisely. The integral terms of the\nhorizontal forces Fx,Fyand torques \u001cx,\u001cy,\u001czare parts\nof the misalignment errors in the joule balance. Other\nparts of the misalignment errors come from the position\nchange of the coil between the two measurement modes\nin the joule balance. In equation (1), it is assumed that\nthe positions A and B in the measurement of the flux\nlinkage difference are the same as the starting point and\nending point of the integral trajectory L. However, it is\nnot the case in the actual measurement. As shown in Fig.\n1, when current pass through the coil, the positions A and\nB will actually shift to A’ and B’ . Then, equation (1) can\nbe rewritten as\nwAA0+wA0B0+wB0B=I[ (B)\u0000 (A)]; (3)\nwherewAA0,wA0B0,wB0Bare the work done by the mag-\nnetic force and torque when the coil is respectively moved\nfrom position A to A’, A’ to B’ and B’ to B. It can be seen\nfrom equation (3) that the misalignment error includes\nthree parts: wAA0,wB0Band the integral terms of the\nhorizontal forces and torques in wA0B0.\nThe active damping device used in the joule balance\nis able to produce auxiliary horizontal forces and torques\non the coil to change the position of the coil. Therefore,\nwith the feedback from the detection system of the five\ndirections of displacement, the position of the coil can\nbe easily controlled. As a result, the positions A and B\ncan be adjusted to be the same as A’ and B’ by the\nfeedback of the damping device. In the meanwhile, the\nhorizontal displacement x,yand the rotation angles \u0012x,\n\u0012y,\u0012z, can be kept unvaried when the coil is moved along\nthe vector trajectory Lin the weighting mode of the\nFig. 2. Construction of the damping device. (a) The overall structure\nof the damping device with the top cover 1 open. (b) The permanent\nmagnets (5, 6, 7, 11,12, 13) and the inner yokes (8, 9, 10). (c) Com-\nbination of auxiliary coils 14 and 15. (d) Combination of auxiliary\ncoils 14 and 16.\njoule balance. These two improvements will significantly\nreduce the misalignment error. For an ideal case, i.e. the\nalignment is adjusted to be good enough, only the work\ndone by the vertical force in wA0B0is left on the left side\nof equation (3) and other terms can be ignored. Then\nequation (3) can be rewritten as\nZB0\nA0Fzdz=I[ (B)\u0000 (A)]: (4)\nIII. Principle and Construction of the Damping\nDevice\nA. The practical structure\nFig. 2 shows the practical structure of the active damp-\ning device used in the joule balance.The outer yokes 1, 2,\n3 and the inner yokes 8, 9,10 are made of soft iron with\nhigh permeability. The component 4 shown in Fig. 2(a)\nis made of aluminum, for mechanically supporting three\ninner yokes. The component 17 in Fig. 2(c) and Fig. 2(d)\nis the framework of the auxiliary coils which is made up\nof polysulfone.\nComponents 5, 6, 7, 11, 12 and 13 are permanent\nmagnets magnetized along the zaxis. It should be noted\nthat the magnetization direction of the three permanent\nmagnets 5, 12, 13 is opposite to the magnetization direc-\ntion of the other three permanent magnets 6, 7, 11. As a\nresult, in this magnetic circuit, the magnetic flux in the\nair gap originates from the yoke 8 and enters the yokes 9,\n10.IEEE TRANS. INSTRUM. MEAS., CPEM2016, JUNE 10, 2021 3\nComponents 14, 15 and 16 are three orthogonal aux-\niliary coils. Fig. 2(c) and Fig. 2(d) show two different\ncombinations of the auxiliary coils. Coils 14 and 15 are\nglued together in Fig. 2(c) and coils 14 and 16 are glued\ntogether in Fig. 2(d). Since there are considerable mag-\nnetic gradients in x,yand the azimuth \u0012, torques and\nforces will be produced on the auxiliary coils when current\npasses through the auxiliary coils.\nThe cross sectional view of the inner yokes and the\nauxiliary coils in the active damping device is shown in\nFig. 3. The dotted lines denote the magnetic flux in the\nair gap. The coil 14 in the magnetic field shown in Fig.\n3(a) will produce a torque \u001cxaroundxaxis, relative to\nthe mass center. Coil 15 in Fig. 3(b) will produce a force\nFyalong theyaxis and coil 16 in Fig. 3(c) will produce\na forceFxalong thexaxis. The forces and torques of the\nthree coils along other directions are theoretically small.\nThe schematic diagram of the damping device is shown\nin Fig. 3(d), containing 3 damping segments. The coil is\nsuspended from the spider by three rods. Each damping\nsegment is fixed on a support with the auxiliary coils\nmechanically connected to the coil spider rod. In two\nof them, the combination of the auxiliary coils 14 and\n15 shown in Fig. 2(c) is used. In the other one, the\ncombination of the auxiliary coils 14 and 16 shown in Fig.\n2(d) is used. Hence, the torque produced by coil 14 in all\nthe three damping segments is used to adjust the rotation\nangle\u0012x,\u0012yof the suspended coil. The force of coil 15 in\ntwo segments points to the center of the suspended coil\nand can be applied to adjust the horizontal movement x,\ny. The force of coil 16 in the third segment is along the\ntangential of the suspended coil and can be employed to\nadjust the rotation along zaxis, i.e.\u0012z.\nThe auxiliary coils are connected to the current sources\nthrough the suspension system by fine wires called hair-\nsprings. To reduce the number of the hairsprings and keep\nthe symmetry of the suspended coil, six auxiliary coils are\ninstalled in the practical structure instead of nine in the\ninitial design [15]. Fig. 4(a) shows the six auxiliary coils\ninstalled on the suspended system. In theory, to suppress\nthe motions of the five degrees of freedom, the minimum\nnumber of auxiliary coils is five. Hence, only two of the\nthree coils 14 are connected to the current sources in\npractice. The three damping devices are placed at angles\nof 120 degrees around the magnet in the joule balance as\nshown in Fig. 4(b). The pallets of the damping devices are\nfixed with respect to the marble shelf in the joule balance\nwhich will be kept static.\nIn the joule balance, the suspended coil is kept static\nwhile the magnet is moved by a linear translation stage.\nFor the active damping device used in the joule balance,\nthe auxiliary coils are always kept static and do not require\ntoo much moving space. If such a damping device is used\nin the velocity mode of watt balances, the length of the\ninner yokes must be long enough for the movement of the\nauxiliary coils.\nFig. 3. The cross sectional view of the inner yokes and the auxiliary\ncoils. (a) The torque of coil 14. (b) The force of coil 15. (c) The force\nof coil 16. (d) The schematic diagram of three damping devices.\nFig. 4. (a) The auxiliary coils installed on the suspended system. (b)\nMechanical assembling of the damping device.\nB. The feedback control\nThe entire control circuit is shown in Fig. 5. The PID\ncontroller is realized with a LabVIEW program. Five\nchannels of a high-speed analog output (NI 6733) are used\nas the voltage control of the current sources. With an\nexternal voltage reference, the range of the analog output\nvoltage is \u00061 V. The circuit of a current source is shown in\nFig. 6, performing the U=Iconverter. A 10 \nfour-terminal\nresistor with a low temperature coefficient ( <1 ppm/\u000eC)\nis used as the sense resistor in the current source. Hence,\nthe range of the output current is \u0006100 mA.\nThe detection system of the relative position between\nthe coil and the main magnet is composed of a laser\ninterferometer and position sensitive devices (PSD). The\nrotation angles \u0012x,\u0012yare obtained with the laser interfer-\nometer. The resolution of \u0012xand\u0012yis less than 1 \u0016rad.\nThe horizontal displacements x,yand rotation angle \u0012zIEEE TRANS. INSTRUM. MEAS., CPEM2016, JUNE 10, 2021 4\nFig. 5. The block diagram of the active control circuit.\nFig. 6. The circuit of the current source.\nFig. 7. The setup of auxiliary coils and the coordinate axis of the\ndetection system.\nare obtained from four PSDs. The resolution in xandyis\nabout 2\u0016m and the resolution of \u0012zis about 5\u0016rad.\nThe five auxiliary coils and the coordinate axis of the\ndetection system are shown in Fig. 7. The large circle\ndenotes the suspended coil in joule balance. Five auxiliarycoils are set concentric to the suspended coil with every 120\ndegrees. Since the horizontal displacement and rotation\nangle of the suspended coil are in a small range, the\nrelationship between the detection and the auxiliary coil\ncurrent is close to be linear. In order to check the depen-\ndence between the coil motion and the current through\nthe auxiliary coils, the transfer matrix Mis measured by\napplying a 50 mA current individually to each of the five\nauxiliary coils. The measurement is written as\n2\n66664dx\ndy\nd\u0012x\nd\u0012y\nd\u0012z3\n77775=M2\n66664I16\nI15\u00002\nI15\u00001\nI14\u00002\nI14\u000013\n77775\n=2\n666640:12 0:32\u00001:04 \u00000:20 0:16\n0:16\u00001:44 0:88 \u00000:24\u00000:24\n0:10 0:12 0:16 1:78 1:30\n0:04 0:10\u00000:08\u00001:16 1:78\n6:30 0:20\u00000:30 0:00 0:003\n777752\n66664I16\nI15\u00002\nI15\u00001\nI14\u00002\nI14\u000013\n77775;\n(5)\nwhere dx,dy,d\u0012x,d\u0012yandd\u0012zare the variations of the\ndetection system with units of \u0016m,\u0016m,\u0016rad,\u0016rad,\u0016rad\nrespectively, I16,I15\u00002,I15\u00001,I14\u00002andI14\u00001the currents\npassing through the five auxiliary coils with unit of mA.\nThe determined matrix Min equation (5) matches the\nanalysis of the above section: The horizontal displacement\nx,yis mainly controlled by coils 15-1 and 15-2. The\nrotation angle \u0012x,\u0012yis mainly adjusted by coils 14-1 and\n14-2. The rotation angles \u0012zis mainly controlled by coil\n16. It can be seen that due to some imperfection of the\nmagnetic field and the misalignment of five auxiliary coils,\nthere are some weak couplings between different channels.\nIn the control, we should use the inverse of the transfer\nmatrix to decouple the correlation of different loops, i.e.\n2\n66664I16\nI15\u00002\nI15\u00001\nI14\u00002\nI14\u000013\n77775=M\u000012\n66664dx\ndy\nd\u0012x\nd\u0012y\nd\u0012z3\n77775\n=2\n66664\u00000:03 0:02 0:00 0:00 0:16\n\u00000:78\u00000:87\u00000:16 0:07 0:04\n\u00001:22\u00000:26\u00000:08 0:14 0:03\n0:12 0:04 0:39\u00000:29\u00000:01\n0:07 0:06 0:26 0:37\u00000:013\n777752\n66664dx\ndy\nd\u0012x\nd\u0012y\nd\u0012z3\n77775:\n(6)\nWith a known difference of the detection and the target,\nthe output voltage of auxiliary coils will be determined\nby both the PID controller and the inverse matrix M\u00001.\nNote that since the response of auxiliary coils can easily\nexcite the suspension mechanism, fast and large feedback\ncurrents should be avoided.\nIV. Experimental Results\nA. Damping performance\nWhen the suspended coil receives external shocks such\nas loading of the test mass, ground vibration, etc., theIEEE TRANS. INSTRUM. MEAS., CPEM2016, JUNE 10, 2021 5\nenergy blue transfer out through the suspension system is\nvery slow due to a high Qfactor. The suspended coil will\nswing for a long time without damping. Fig. 8 shows the\nmotions of the suspended coil when the mass is put on the\nmass pan with and without the proposed damper. Without\nany damping device, the maximal motion amplitudes of\nthe five degrees of freedom x,y,\u0012x,\u0012y,\u0012zare 250\u0016m,\n150\u0016m, 500\u0016rad, 550\u0016rad, 300\u0016rad respectively. After\nabout 5 minutes, the motion amplitudes will decrease\nto 10\u0016m, 10\u0016m, 15\u0016rad, 15\u0016rad, 50\u0016rad.The damping\nfactor is about 0.11 kg \u0001s\u00001. The steady state deviations\nbetween the mass on state and the mass off state of the\nfive degrees of freedom are 10 \u0016m, 35\u0016m, 50\u0016rad, 20\u0016rad,\n30\u0016rad respectively.\nWhen the active damping device is used, the motions of\nthe suspended coil will decay rapidly. As shown in Fig.\n8,x,y,\u0012x,\u0012y,\u0012zwill reduce to 8 \u0016m, 8\u0016m, 10\u0016rad,\n10\u0016rad, 20\u0016rad in 30 seconds. The damping factor is\nabout 1.1 kg \u0001s\u00001, which is about 10 times larger than that\nwithout the damper. The steady state deviations between\nthe mass on state and the mass off state of the five degrees\nof freedom are less than 2 \u0016m, 2\u0016m, 1\u0016rad, 1\u0016rad, 5\u0016rad.\nB. Reduction of the current dependence of the coil position\nWhen current passes through the suspended coil, the\npositions A and B in the flux linkage difference mea-\nsurement will change to A’ and B’, yielding the errors\nwAA0andwB0B. To generate a 5 N magnetic force, the\nsuspended coil current is about 7.5 mA. Fig. 9 shows the\nchanges of the five degrees of freedom x,y,\u0012x,\u0012y,\u0012z\nwhen the suspended coil current is reduced from 7.5 mA\nto 0 mA slowly. Without the active damping device, the\nmagnitude changes are 25 \u0016m, 5\u0016m, 15\u0016rad, 15\u0016rad,\n70\u0016rad respectively. Due to the restriction of the narrow\nair gap and the laser alignment in the present construction\nof the joule balance, it is very difficult to further adjust\nthe concentricity and horizontality of the suspended coil\nto reduce the magnitudes of the coil position change.\nTo reduce the errors wAA0andwB0B, the positions of A\nand B can be adjusted to be the same as A’ and B’ by the\nactive damping device. As shown in Fig. 9, when the active\ndamping device is added, the magnitude changes of x,y,\n\u0012x,\u0012y,\u0012zare less than 2 \u0016m, 2\u0016m, 1\u0016rad, 1\u0016rad, 5\u0016rad\nrespectively, which on average is one magnitude improved\nthan that without the electromagnetic damper.\nC. Improvement of the vertical movement\nIn both the weighing and dynamic modes of the joule\nbalance, the magnet is moved in the vertical direction zby\na translation stage [17]. However, due to the distortion of\nthe guide rail in the translation stage and other mechanical\nexcitations, the other five degrees of freedom x,y,\u0012x,\u0012y,\n\u0012zwill also change with the movement of the magnet.\nTherefore, the horizontal forces Fx,Fyand torques \u001cx,\n\u001cy,\u001czwill contribute to in the misalignment errors.\nFig. 10 shows the value of the five degrees of freedom\nx,y,\u0012x,\u0012y,\u0012zwith different zduring the movement of\nFig. 8. Coil motions when test mass is put on the mass pan with\nand without the proposed damper. The red lines (–) are without the\ndamper while the blue dash lines (- -) are with the damper.IEEE TRANS. INSTRUM. MEAS., CPEM2016, JUNE 10, 2021 6\nFig. 9. Coil motions when the coil current is reduced from 7.5 mA\nto 0 mA slowly. The red lines (–) are without the damper while the\nblue dash lines (- -) are with the damper.\nFig. 10. Coil motions with different zduring the movement of the\nmagnet. The red lines (–) are without the damper while the blue dash\nlines (- -) are with the damper.IEEE TRANS. INSTRUM. MEAS., CPEM2016, JUNE 10, 2021 7\nTABLE I\nThe unwanted vertical force of the five auxiliary coils\nwith 100 mA excitation.\nCoil No. Vertical force /mg uncertainty /mg\n16 0.08 0.01\n15-1 0.06 0.01\n15-2 0.06 0.01\n14-1 1.04 0.01\n14-2 0.85 0.01\nthe magnet. Without the active damping device, the peak-\nto-peak values of the five degrees of freedom are 15 \u0016m,\n15\u0016m, 10\u0016rad, 8\u0016rad, 40\u0016rad respectively. When the\nactive damping device is added during the movement, the\nunwanted motions of the magnet can be compensated by\nthe motions of the suspended coil. Hence, the changes of\nthe relative value of x,y,\u0012x,\u0012y,\u0012zcan be reduced. As\nshown in Fig. 10, the peak-to-peak values of the five de-\ngrees of freedom are reduced to 6 \u0016m, 6\u0016m, 4\u0016rad, 4\u0016rad,\n25\u0016rad with the active damping device. The measurement\nshows that all curves of the five degrees of freedom are\nnearly flat. Therefore, the misalignment errors, resulting\nfrom the work done by the horizontal forces and torques\ninwA0B0, are greatly reduced.\nV. Consideration of Potential Systematic\nEffects\nA. The vertical force\nWhen the active electromagnetic damping device is used\nin the weighing mode of the joule balance, any additional\nvertical force caused by the five auxiliary coils should be\nconsidered. Here the mass comparator has been used to\nmeasure the vertical force. With 100 mA current in the\nauxiliary coils, the measurement results are shown in Table\nI. As discussed in [15], the vertical forces of coil 16 and 15\nare very small when compared with that of coil 14. Since\nthe working current in the auxiliary coils is much less than\n100 mA in the weighing mode, the vertical forces will be\nmuch smaller than the results in Table I.\nWhen the force of the suspended coil is measured in the\nweighing mode, the suspended coil is static and the current\nin the auxiliary coils of the active damping device is nearly\nconstant. As the magnetic force is proportional to the cur-\nrent, the additional vertical force of the auxiliary coil can\nbe determined by measuring the current in the auxiliary\ncoil. Note that in this measurement, the uncertainty of the\nmass comparator is about 0.01 mg. A correction may then\nbe made and the effect of the unwanted vertical forces can\nbe reduced to several parts in 108.\nB. Magnetic flux leakage\nThe distance between the active damping devices and\nthe test mass is about 160 mm. Hence, the leakage mag-\nnetic field of the active damping devices should be consid-\nered. The interaction between the magnetic field and the\nmass is discussed in [16]. The equation of the vertical force\nFig. 11. The absolute magnitude of the magnetic field at the location\nof the mass. The measurement is fitted by a quadratic function.\non the mass with a volume susceptibility \u001fand permanent\nmagnetization Mis given by [18]\nFz=\u0000\u00160\n2@\n@zZ\n\u001fH\u0001HdV\u0000\u00160@\n@zZ\nM\u0001HdV: (7)\nThe magnetic field at the location of the mass is\nmeasured by a gauss meter. Fig. 11 shows the absolute\nmagnitude of the magnetic field as a function of the\ndistance from the top surface of the mass pan. Note that\nalthough the measured field is comparable to the earth’s\nmagnetic field, it produces a much larger field gradient,\nabout 1.3\u0016T/mm. The employed 500 g test mass is about\n50 mm in height and 40 mm in diameter. The magnetic\nsusceptibility \u001fof the mass material is less than 6\u000210\u00004\nand the permanent magnetization Mis about 0.1 A/m.\nWith the method used in [16], the calculated magnetic\nforce of the mass is less than 10 \u0016g. Hence, the relative\nsystematic error caused by the leakage magnetic field is\nless than 2\u000210\u00008.\nVI. Conclusions\nThe active electromagnetic damping device presented\nin this paper can be used in both the dynamic and\nstatic measurement modes of the joule balance. Five de-\ngrees of freedom of the suspended coil, i.e. the horizontal\ndisplacement x,yand the rotation angles \u0012x,\u0012y,\u0012z,\ncan be actively controlled by the damping device. The\nexperimental results show that the active damping device\nwell meets the design targets. The damping factor of the\nsystem is increased by a factor of 10 and the alignment\nis greatly improved by compensating for the non-vertical\nmovement of the coil. Two potential systematic effects, i.e.\nthe unwanted vertical force and the leakage magnetic field\nof the active damping device, are analyzed. It is shown that\nthese effects are small and compensable within few parts\nin108. At present, the misalignment error is about several\nparts in 107due to the restriction of the narrow air gap\nand the laser alignment in the joule balance project. The\npresented active damping device in this case is powerful\nto reduce misalignment uncertainties.IEEE TRANS. INSTRUM. MEAS., CPEM2016, JUNE 10, 2021 8\nReferences\n[1] Z. Zhang, Q. He and Z. Li, ”An approach for improving the\nwatt balance” in Digest of Conf. on Precision Electromagnetic\nMeasurement , Torino, Italy, Jul. 9-14, 2006, pp. 126-127.\n[2] A. Robinson, B. P. Kibble, ”An initial measurement of Planck’s\nconstant using the NPL Mark II watt balance. ” Metrologia ,\nvol.44, no. 6, pp.427-440, 2007.\n[3] A. G. Steele, J. Meija, C. A. Sanchez, et al, ”Reconciling Planck\nconstant determinations via watt balance and enriched-silicon\nmeasurements at NRC Canada.” , Metrologia , vol.48, pp.L8-L10,\n2012.\n[4] A. Picard, H. Fang, A. Kiss, et al, ”Progress on the BIPM watt\nbalance.” IEEE Transactions on Instrumentation and Measure-\nment , vol.58, pp.924-929, 2008.\n[5] S. Schlamminger et al, ”Determination of the Planck constant\nusing a watt balance with a superconducting magnet system at\nthe National Institute of Standards and Technology”, Metrologia ,\nvol.51, pp15-24, 2014.\n[6] A. Eichenberger, H. Baumann, B. Jeanneret, B. Jeckelmann, P.\nRichard,and W. Beer, ”Determination of the Planck constant\nwith the METAS wattbalance,” Metrologia , vol. 48, no. 3, pp.\n133ÂĺC141,2011.\n[7] M. Thomaset al, ”First determination of the Planck constantus-\ning the LNE watt balance”, Metrologia , vol.52, pp433-443, 2015.\n[8] M.C.Sutton, T.M.Clarkson. ”A magnet system for the MSL watt\nbalance.” Metrologia , vol.51, pp. 101-106,2014.\n[9] D.Kim, B.C.Woo, K.C.Lee, et al, ”Design of the KRISS watt\nbalance.” Metrologia , vol.51, pp. 96-100, 2014.\n[10] Y. Azumaet al, ”Improved measurement results forthe Avo-\ngadro constant using a28Si-enriched crystal”, Metrologia , vol.52,\npp360-375, 2015.\n[11] B. P. Kibble, ”A measurement of the gyromagnetic ratio of\nthe proton by the strong field method”, Atomic Masses and\nFundamental Constants5 , Springer US, pp545-551, 1976.\n[12] J. Xu et al, ”A determination of the Planck constant by the gen-\neralized joule balance method with a permanent-magnet system\nat NIM”, Metrologia , vol.53, pp86-97, 2016.\n[13] S. Li et al, ”Coil motion effects in watt balances: a theoretical\ncheck”, Metrologia , vol.53, pp817-828, 2016.\n[14] D. Haddadet al, ”A precise instrument to determine the Planck\nconstant, and the future kilogram”, Review of Scientific Instru-\nments ,vol.87, pp1-14, 2016.\n[15] J. Xu et al, ”A Magnetic Damping Device for Watt and Joule\nBalances”, To be published inProc. CPEM Dig., Jul. 2016 .\n[16] Seifert F, Panna A, Li S, et al,”Construction, Measurement,\nShimming, and Performance of the NIST-4 Magnet System”,\nIEEE Transactions on Instrumentation and Measurement ,vol.63,\npp.3027-3038, 2014.\n[17] Zhengkun Li, et al, ”The Improvement of Joule Balance NIM-\n1 and the Design of New Joule Balance NIM-2”, IEEE Trans-\nactions on Instrumentation and Measurement , vol.64, pp.1676-\n1684, 2015.\n[18] R. S. Davis, ”Determining the magnetic properties of 1 kg mass\nstandards”, J. Res. Nat. Inst. Standards Technol. , vol.100, no.3,\npp.209ÂĺC226, May. 1995.\nJinxin Xu was born in Yancheng, Jiangshu\nprovince, China, in 1990. He received the B.S.\ndegree from Harbin Institute of Technology,\nHarbin, China, in 2012. He is currently pursu-\ning the Ph.D. degree at Tsinghua University,\nBeijing, China. His dissertation will be a part\nof the joule Balance at the National Institute\nof Metrology, China.\nQiang You was born in Shandong province,\nChina. He is currently pursuing the Ph.D. de-\ngree with Tsinghua University, Beijing, China.\nHis dissertation will be a part of the joule bal-\nance with the National Institute of Metrology,\nBeijing, China.\nZhonghua Zhang was born in Suzhou,\nJiangshu, China in July 1940. He received his\nB.S. degree and M.S. degree in Electrical En-\ngineering from Tsinghua University, Beijing,\nChina separately in 1965 and 1967. In 1995, he\nbecame a member of the Chinese Academy of\nEngineering. In 1967, Zhonghua Zhang joined\nthe Electromagnetic Division of National In-\nstitute of Metrology (NIM), Beijing, China.\nSince then he has been involved in the re-\nsearch work on the Cross Capacitor standard,\nsuperconducting high magnetic field standard and Quantum Hall\nResistance standard. In 2006, he proposed the joule balance and led\nthe research work since then.\nZhengkun Li was born in Henan Province,\nChina in 1977. He received the B.S. degree in\ninstrument science and technology from China\nInstitute of Metrology, Hangzhou, China, in\n1999. He received the M.S. degree in quantum\ndivision of National Institute of Metrology\n(NIM), China in 2002. He received his Ph.D.\ndegree from Xi’an Jiaotong University, Xi’an,\nChina in 2006. In 2002, he became a perma-\nnent staff of NIM and worked on the establish-\nment of Quantum Hall Resistance standard.\nSince 2006, he has been involved in the research work on joule balance\nproject at NIM and focuses on the electromagnetic measurement\nincluding mutual inductance measurement for the joule balance.\nShisong Li (M’15) got the Ph.D. degree in\nTsinghua University, Beijing, China in July,\n2014. He then joined the Department of Elec-\ntrical engineering, Tsinghua University during\nJuly, 2014-September, 2016. He has been a\nguest researcher at the National Institute of\nMetrology in China since 2009, and he worked\nat the National Institute of Standards and\nTechnology, United States for 14 months dur-\ning 2013-2016. He is currently with the In-\nternational Bureau of Weights and Measures\n(BIPM), France. His research interests include modern precision\nelectromagnetic measurement and instrument technology." }, { "title": "2104.05789v1.Thermal_Radiation_Equilibrium___Nonrelativistic__Classical_Mechanics_versus__Relativistic__Classical_Electrodynamics.pdf", "content": "arXiv:2104.05789v1 [physics.class-ph] 12 Apr 2021Thermal Radiation Equilibrium: (Nonrelativistic) Classi cal\nMechanics versus (Relativistic) Classical Electrodynami cs\nTimothy H. Boyer\nDepartment of Physics, City College of the City\nUniversity of New York, New York, New York 10031\nAbstract\nPhysics students continue to be taught the erroneous idea th at classical physics leads inevitably\nto energy equipartition, and hence to the Rayleigh-Jeans la w for thermal radiation equilibrium.\nActually, energy equipartition is appropriate only for nonrelativistic classical mechanics, but has\nonly limited relevance for a relativistic theory such as classical electrodynamics. In this article,\nwe discuss harmonic-oscillator thermal equilibrium from t hree different perspectives. First, we\ncontrast the thermal equilibrium of nonrelativistic mecha nical oscillators (where point collisions\nare allowed and frequency is irrelevant) with the equilibri um of relativistic radiation modes (where\nfrequency is crucial). The Rayleigh-Jeans law appears from applying a dipole-radiation approx-\nimation to impose the nonrelativistic mechanical equilibr ium on the radiation spectrum. In this\ndiscussion, we note the possibility of zero-point energy fo r relativistic radiation, which possibility\ndoes not arise for nonrelativistic classical-mechanical s ystems. Second, we turn to a simple electro-\nmagneticmodelofaharmonicoscillator andshowthattheosc illator isfullyinradiationequilibrium\n(which involves allradiation multipoles, dipole, quadrupole, etc.) with clas sical electromagnetic\nzero-point radiation, but is notin equilibrium with the Rayleigh-Jeans spectrum. Finally, we\ndiscuss the contrast between the flexibility of nonrelativi stic mechanics with its arbitrary poten-\ntial functions allowing separate scalings for length, time , and energy, with the sharply-controlled\nbehavior of relativistic classical electrodynamics with i ts single scaling connecting together the\nscales for length, time, and energy. It is emphasized that wi thin classical physics, energy-sharing,\nvelocity-dependent damping is associated with the low-fre quency, nonrelativistic part of the Planck\nthermal radiation spectrum, whereas acceleration-depend ent radiation damping is associated with\nthe high-frequency adiabatically-invariant and Lorentz- invariant part of the spectrum correspond-\ning to zero-point radiation.\n1I. INTRODUCTION\nA. Erroneous Textbook Claim\nThe current textbooks of modern physics still make the erroneou s claim that classical\nphysics leads inevitably tothe Rayleigh-Jeansspectrum fortherma l radiationequilibrium.[1]\nIndeed the treatment of classical physics in these texts seems to have progressed only so far\nas James Jeans’ Report on Radiation and the Quantum-Theory in 1914, but no further.[2]\nIn this article, we make another[3] attempt to counteract the err oneous textbook claim.\nSpecifically, we point out that a simple electromagnetic-model harmo nic oscillator is notin\nequilibrium with the Rayleigh-Jeans spectrum. In the discussion, we e mphasize the con-\ntrasting points of view between nonrelativistic classical mechanics a nd relativistic classical\nelectrodynamics in connection with equilibrium for harmonic oscillators .\nOur earlier attempt[3] to correct the misinformation regarding the rmal radiation equilib-\nrium in classical physics involved an extended historical survey. It w as pointed out that\nthe introduction of classical electromagnetic zero-point radiation led to modifications of the\nhistorical arguments. The modified arguments provide natural cla ssical explanations for the\nPlanck spectrum within relativistic classical physics. The Planck spec trum with zero-point\nradiation corresponds to a radiation energy per normal mode\nUrad(ω,T) =1\n2/planckover2pi1ωcoth/parenleftbigg/planckover2pi1ω\n2kBT/parenrightbigg\n=/planckover2pi1ω\nexp[/planckover2pi1ω/(kBT)]−1+1\n2/planckover2pi1ω. (1)\nThe spectrum involves a transition from the relativistic high-freque ncy region,\n/planckover2pi1ω/(kBT)>>1, associated with adiabatically-invariant and Lorentz-invariant ze ro-point\nradiation where Urad(ω,T)→(1/2)/planckover2pi1ω, over to the low-frequency Rayleigh-Jeans re-\ngion,/planckover2pi1ω/(kBT)<<1, associated with nonrelativistic energy-sharing behavior where\nUrad(ω,T)→kBT.\nB. Three Analyses for Radiation Equilibrium\nThe present discussion is quite different from the previous historica l survey. Here we\npresent three different analyses of radiation equilibrium connected to harmonic oscillator\nsystems; these analyses include aspects of point-particle collisions , adiabatic invariance,\nscattering, and scaling which do not appear in earlier work.\n21. Point-Particle Collisions and Adiabatic Invariance\nAfter some basic preliminaries, we start our analysis by contrasting the equilibrium for\ntwo different situations; one is a box containing a finite number of mec hanical oscillators\nconnected through point collisions with a free particle, and the othe r involves a charged\nmechanical oscillator coupled to a divergent number of electromagn etic radiation modes.\nWhentheequipartitionideasofnonrelativisticmechanicsareimposed uponradiationtreated\nin the dipole approximation, the Rayleigh-Jeans spectrum appears. Our analysis also\nsuggests the possibility of classical zero-point energy for radiatio n associated with oscillator\nadiabatic invariance , something which is not possible for classical mec hanical systems which\nallow point-particle collisions.\n2. Equilibrium Under Scattering\nIn the second equilibrium analysis, we emphasize that a charged harm onic oscillator con-\nnected to radiation through the dipole approximation can come to eq uilibrium with any\nradiation spectrum. In order to obtain a specific radiation spectru m, some additional as-\nsumptionsareneeded. TheRayleigh-Jeansspectrumandzero-po intspectrumcorrespondto\nthe extremes allowed by Wien’s displacement theorem. If we turn to in variance under scat-\nteringasourcriterionforapreferredspectrum, thenagainther earetwo naturalpossibilities.\nIn one possibility, the mechanical behavior is modified to an anharmonic nonrelativistic os-\ncillator while maintaining the dipoleapproximation to radiation. The other possibility\ninvolves maintaining the harmonic mechanical behavior but going to electromagnetic in-\nteractions beyond the dipole approximation, giving an approximately relativistic oscillator\nsystem. We point out that the first nonrelativistic procedure leads to the Rayleigh-Jeans\nspectrum, while the second relativistic analysis leads to the classical zero-point radiation\nspectrum. For a harmonic oscillator, the Rayleigh-Jeans spectrum arising from the energy\nequipartition ideas of nonrelativistic classical mechanics is not tolera ted beyond the dipole\ncoupling to radiation.\n33. Scaling in Nonrelativistic Mechanics and in Relativisti c Electrodynamics\nFinally, in the third equilibrium analysis, we point out the sharp differenc es in scaling\nbehavior between nonrelativistic mechanics and relativistic electrod ynamics, and connect\nthese differences to the requirements for thermal equilibrium. The enormous flexibility of\nnonrelativistic classical mechanics contrasts with the strictly-con trolled radiation connec-\ntions of relativistic classical electrodynamics. And this contrast is c rucial in questions of\nthermal radiation equilibrium. We emphasize that, within classical phy sics, the full Planck\nspectrum requires both velocity-dependent damping and accelera tion-dependent radiation\ndamping.\nC. Missing Aspects in the Physics of 1900: Zero-Point Radiation and Relativity\nThe foundation for the erroneous claims regarding classical therm al radiation which per-\nvade the textbooks and the internet lies in the fact that the physic ists of the early 20th\ncenturywere(just as the physicists of today are) unaware of two crucial aspects of classical\nphysics existing within classical electrodynamics: 1)theexistence o fclassical electromagnetic\nzero point radiation, and 2)the importance of special relativity. In this article, we illustrate\nthe importance of these two aspects of classical physics by compa ring the ideas of thermal\nequilibrium which arise in classical mechanics and in classical electrodyn amics. The models\nfor our discussion are harmonic oscillators.\nII. BASIC PRELIMINARIES\nA. Some Contrasts Between (Nonrelativistic) Classical Mechanics and ( Relativ is-\ntic) Classical Electrodynamics.\nIn order to prepare the reader for our analysis, we first mention s ome of the important\ncontrasts between classical mechanics andclassical electrodyna mics. Classical mechanics al-\nlows point collisions between massive particles which provide velocity-d ependent damping,\nwhereas classical electrodynamics involves only long-range forces associated with electro-\nmagnetic fields. Classical mechanics allows particles to exist without a ny connection to\na spacetime-dependent field, whereas in classical electrodynamics , every charged particle is\n4connected to electromagnetic fields which introduce acceleration- dependent radiation forces.\nThereisnospecial roleforoscillation frequency when considering thermal equilibrium within\nnonrelativistic classical mechanics, whereas classical electrodyna mics determines thermal\nequilibrium through the frequency-dependent connection of char ges to radiation fields. Fi-\nnally, the relativistic theory of classical mechanical particles is an insignificant part of me -\nchanics; the relativistic theory is restricted to point collisions between particles and allows\nno potential functions.[4] Classical mechanics which allows the intera ction of particles be-\nyond point collisions satisfies Galilean symmetry, whereas classical ele ctrodynamics satisfies\nLorentz symmetry.\nB. Harmonic Oscillator in Action-Angle Variables\nA one-dimensional harmonic oscillator of natural (angular) freque ncyω0and mass mcan\nbe described by the Hamiltonian\nH(x,p) =p2/(2m)+(1/2)mω2\n0x2=Uosc. (2)\nWithin nonrelativistic mechanics, this Hamiltonian will give the equations of motion for the\nposition xand momentum pwhile the initial conditions will determine the actual motion,\nincluding theenergy Uoscandphase of the oscillation. When discussing adiabaticchanges in\nthe oscillator’s natural frequency ω0, it is convenient to reexpress the Hamiltonian in terms\nof action-angle variables wandJusing the canonical transformation with the connection\nx=/radicalbigg\n2J\nmω0sinwandp=/radicalbig\n2mω0Jcosw. (3)\nIn terms of the new variables wandJ, the Hamiltonian takes the form\nH=p2\n2m+1\n2mω2\n0x2=Jω0=Uosc. (4)\nIn terms of action-angle variables, the oscillator Hamiltonian no longe r refers to the particle\nmassmnor the amplitude of oscillation x. The initial conditions again give the oscillator\nenergyUosc=Jω0and the initial phase φinw=ω0t+φ.\n5C. Thermodynamics of the Harmonic Oscillator and Wien’s Theorem\nA harmonic oscillator (whether a mechanical oscillator or a radiation m ode) is such a\nsimple system that its thermodynamic behavior depends upononly tw o variables, its natural\nfrequency ω0and its temperature T. By considering the adiabatic change of the oscillator\nfrequency ω0, it is easy to derive Wien’s displacement theorem indicating that in ther mal\nequilibrium, the energy of the oscillator must take the form[5]\nUosc(ω0,T) =ω0f(ω0/T), (5)\nwheref(ω0/T)isanunknown functionoftheratio ω0/T. Theextremes ofenergy equiparti-\ntionandzero-point energy correspond to thelimiting situations whe re the energy Uosc(ω0,T)\nbecomes independent of one of the two variables. Thus if f(ω0/T)→const×(T/ω0), then\nwe have energy equipartition Uosc(ω0,T)→const×T; on the other hand, if f(ω0/T)→\nconst′, thenonefindstemperature-independent zero-pointenergy Uosci(ω0,T)→const′×ω0.\nIf we consider only the first two laws of thermodynamics, then the f unctionf(ω0/T) giving\nthe transition between these extremes is unknown. Only if we add fu rther criteria such\nas ”smoothness”[5] or inclusion of the third law,[6] do we obtain the P lanck spectrum with\nzero-point energy given inEq. (1). However, boththese additiona l criteria arenot currently\nin fashion in physics, and below we will consider other possibilities.\nD. Thermal Radiation Modes\nThe solutions of Maxwell’s equations involve electromagnetic fields aris ing both from\ncharge and current sources, and also from boundary conditions. The boundary-condition\nelectromagnetic fields satisfy the source-free Maxwell equations . Thermal radiation can be\ntreated as boundary-condition radiation, and can be expanded in t erms of plane waves with\nrandom phases satisfying periodic boundary conditions in a large cub ic box with sides of\nlengtha,\nE(r,t) =/summationdisplay\nk,λ/hatwideǫ(k,λ)/parenleftbigg8πUrad(ω)\na3/parenrightbigg1/21\n2{exp[ik·r−iωt+iθ(k,λ)]+cc},(6)\nB(r,t) =/summationdisplay\nk,λ/hatwidek×/hatwideǫ(k,λ)/parenleftbigg8πUrad(ω)\na3/parenrightbigg1/21\n2{exp[ik·r−iωt+iθ(k,λ)]+cc},(7)\n6where the sum over the wave vectors k=/hatwidex2πl/a+/hatwidey2πm/a+/hatwidez2πn/ainvolves integers\nl,m,n= 0,±1,±2,...running over all positive and negative values, there are two polar-\nizations/hatwideǫ(k,λ), λ= 1,2,and the random phases θ(k,λ) are distributed uniformly over the\ninterval (0 ,2π],independently for each wave vector kand polarization λ. The notation “ cc”\nrefers to the complex conjugate. We have assumed that the radia tion spectrum is isotropic,\nand that the energy per normal mode at radiation frequency ω=ckis given by Urad(ω).\nThere are a divergent number of radiation normal modes with no limit o n the frequency ω.\nE. Equilibrium Between a Charged Oscillator and Random Radiation in the\nDipole Approximation\nWhen treated in the dipole approximation, the interaction of a charg ed harmonic oscil-\nlator with the random radiation in Eqs. (6) and (7) can be written as N ewton’s second\nlaw\nm¨x=−mω2\n0x+mτ...x+eEx(0,t), (8)\ninvolving theharmonicrestoringforce −mω2\n0x, thedipoleradiationdampingforce mτ...xwith\nτ= 2e2/(3mc3), and the radiation driving force treated in dipole approximation eEx(0,t).\nThis equation of motion has been solved many times, beginning with Plan ck’s work at\nthe end of the 19th century.[7][8][9][10] It is found that in this dipole app roximation, the\noscillator acts essentially like a radiation mode at the oscillation freque ncyω0: the average\nenergy of the oscillator matches the average energy of the radiat ion modes at frequency ω0,\nUosc(ω0) =Urad(ω0), and the phase space distribution Posc(x,p) of the oscillator takes the\nform\nPosc(x,p) =const×exp[−H(x,p)/Urad(ω0)] (9)\nwhereH(x,p) is the Hamiltonian of the oscillator given in Eqs. (2) or (4).\nIn the dipole approximation, the harmonic oscillator is in equilibrium with anyspectrum\nUrad(ω) of isotropic random radiation. Thus, other than the requirement of isotropy in\ndirection, the spectrum Urad(ω) is arbitrary. The oscillator will enforce isotropic behav-\nior for the radiation spectrum, but the harmonic oscillator will not ch ange the frequency\ndistribution of radiation among the various frequencies.[9]\n7III. EXPLORING EQUILIBRIUM FOR OSCILLATORS IN MECHANICS AND\nIN ELECTROMAGNETIC RADIATION\nWe now turn from the preliminary aspects of our analysis over to spe cific treatments of\nradiation equilibrium for harmonic oscillators.\nA. Equilibrium for Systems of Oscillators\nSuppose that we have an uncharged free particle of mass Mand a special harmonic oscil-\nlator of (angular) frequency ω0at our disposal. We wish to contrast the equilibrium behav-\nior for two different systems. The first system involves a collection o fNnon-interacting,\nclassical-mechanical, one-dimensional oscillators of various freque nciesωin an elastic-walled\nbox; the second involves the electromagnetic radiation in a reflectin g-walled box. Both\nboxes can be described as containing harmonic oscillators inside, the one in terms of me-\nchanical oscillators and the other in terms of electromagnetic radia tion modes. We assume\nthat the oscillators in both boxes are in “thermal equilibrium,” and we w ish to explore this\nthermal equilibrium.\nB. Equilibrium for Mechanical Oscillators\n1. Equilibrium Through Point-Particle Collisions\nIfweintroduceourspecialoscillatorintotheboxof Nmechanicaloscillators, theoscillator\nwillnotcometoequilibriumintheboxunlessthereissomeinteractionbe tweenthisoscillator\nand the original oscillators in the box. However, if we now introduce o ur free particle M\ninto the box of mechanical oscillators, then the particle will collide with all the masses of the\nmechanical oscillators and will bring to equilibrium the entire collection ( including both our\nspecial oscillator, the Noscillators in the box, and also the mass M) by sharing the total\nsystem energy Utotalamong the finite number N+1 of mechanical oscillators and the mass\nM. Each oscillator of mechanical frequency ωwill come to equilibrium at the Boltzmann\ndistribution Posc(J,ω) =const×exp[−Jω/Uav] where here Jis the action variable of the\noscillator with Hamiltonian H=Jω, andUav=Utotal/(N+1+3/2) is the average energy,\n(from energy equipartition) with the average energy of each one- dimensional oscillator 2 /3\n8that of the mass Mwhich has three-dimensional motion. We notice that the natural\nfrequency ωof an oscillator is irrelevant for the average oscillator energy at equ ilibrium;\nour special oscillator at frequency ω0has the same average energy as any oscillator at any\nfrequency ω.However, we notice that the phase space distribution Posc(J,ω)for anoscillator\ndoesdependuponitsfrequency ωandalsoupontheaverageenergy Uavfortheentiresystem.\n2. Change of Equilibrium Due to Adiabatic Change of Oscillator Frequency\nIf we now isolate our special oscillator and carry out an adiabaticcha nge in the frequency,\nthe quantity of energy divided by frequency, Uav/ω0=J, is an adiabatic invariant and does\nnot change. When the frequency is increased from ω0toω′\n0, the energy of our special\noscillator is increased from Uavover toU′= (ω′\n0/ω0)Uav. This energy U′is now above the\naverage energy Uavof the other oscillators in the box. On reconnecting our special osc illator\nto the collection of the other mechanical oscillators and the mass M, the average energy of\neach oscillator will increase, and the phase space distribution of eac h oscillator will change\nto give a larger average value of its action variable J. Thus an adiabatic change in the\nfrequency of our special oscillator will indeed disturb the equilibrium o f the entire system.\nC. Equilibrium for Random Radiation\n1. Charged-Oscillator Equilibrium Depends on Oscillator Fr equency\nThisthermal-equilibrium situationfornonrelativistic mechanical oscilla tors(where point-\nparticlecollisionsareallowed, andthereisnoparticularroleregarding energyequilibrium for\ntheoscillatorfrequency) istobecontrastedwiththeelectromagn eticsituationwhichinvolves\na divergent number of massless time-harmonic radiation modes char acterized by frequency.\nIn analogy with the mechanical case, the introduction of a special h armonic oscillator of\nfrequency ω0intoaconducting-walled enclosure givesusnoinformationunless wes pecify the\ninteraction of the oscillator with the radiation modes of the enclosur e. A free but uncharged\nparticleMhas no interaction with the electromagnetic oscillators. However, w e canconnect\nour special mechanical oscillator to radiation by taking the oscillating particle as charged.\nAspointedoutbyPlanckattheendofthe19thcenturyandnoteda boveinourpreliminaries,\ntreated in the dipole approximation, this electromagnetic harmonic o scillator will interact\n9withtheradiationmodesatthesame frequency asthenaturalfre quency oftheoscillatorand\nwill come to equilibrium with these modes, but not with modes of differing frequency. Thus\nthe frequency-dependent connection between an oscillator and e lectromagnetic radiation is\na crucial aspect of electromagnetism.\n2. Possibility of Adiabatic Invariance in Zero-Point Radiat ion\nWe now consider an adiabatic change in the frequency ω0of our oscillator over to a new\nfrequency ω′\n0. During this adiabatic change, the ratio of energy to frequency, U0/ω0=J,\nof our special oscillator is unchanged. We can now have our oscillator (treated in the\ndipole approximation) interact with radiation at the new frequency ω′\n0. If the radiation\nmodes at ω′\n0have greater energy than our oscillator, then, on interaction, th ere will be\nenergy transferred from the radiation modes at ω′\n0over to our special oscillator. On the\nother hand, if the radiation modes at ω′\n0have less energy, then the transfer is reversed.\nHowever, we also have the possibility that our special oscillator has t he same average energy\nas the radiation modes at ω′\n0and so is already in equilibrium with these modes. This\nsituation of undisturbed equilibrium despite an adiabatic change of an oscillator’s frequency\nand energy is an entirely new possibility which does not arise for mecha nical oscillators\nwhere energy equipartition is involved. Here in the electromagnetic s ituation, we have\nthe possibility of making adiabatic changes in the frequency and ener gy of our oscillator\nand yet not transferring any average energy between the electr omagnetic radiation modes.\nIf this situation of thermal equilibrium holds for all radiation modes, t hen the classical\nelectromagnetic spectrum is Lorentz invariant and is termed “class ical electromagnetic zero-\npoint radiation.”[11]\n3. Introduction of Planck’s Constant /planckover2pi1\nThe scale of this zero-point radiation involves the ratio U/ωwhich is the same for all\nradiation frequencies. The scale which fits with nature correspond s to the introduction of\nPlanck’s constant /planckover2pi1as the scale invariant so that the average energy Urad(ω) of a radia-\ntion mode of frequency ωsatisfiesUrad(ω) = (1/2)/planckover2pi1ω. There is no question that, when\ninterpreted in terms of classical physics, nature contains classica l electromagnetic zero-point\n10radiation. Thus the introduction of Planck’s constant /planckover2pi1as the scale of classical zero-point\nradiation gives fully classical explanations for Casimir forces, van de r Waals forces, oscillator\nspecific heats, diamagnetism, and the Planck spectrum of thermal radiation.[12] Zero-point\nradiation involves the same phase space distribution for every radia tion mode no matter\nwhat its frequency,\nPradzp(J) =const×exp/bracketleftbigg\n−Jω\n(1/2)/planckover2pi1ω/bracketrightbigg\n=const×exp/bracketleftbigg\n−J\n/planckover2pi1/2/bracketrightbigg\n. (10)\n4. Planck’s Constant /planckover2pi1Can Appear in Classical or Quantum Theories\nThe appearance of Planck’s constant /planckover2pi1within classical physics provokes the indignation\nof some physicists. Some uninformed physicists continue to insist th at Planck’s constant /planckover2pi1\nis a “quantum constant,” and that any theory in which /planckover2pi1appears has a “quantum” element.\nThat this claim is untrue has been pointed out in work published in the Am erican Journal of\nPhysics.[13] Actually, Planck’s constant /planckover2pi1is a physical constant, like Cavendish’s constant\nG, and can appear in any theory which involves an appropriate scale, j ust as the constant G\ncan appear in both Newtonian gravity and in general relativity since Gsets the scale of the\ngravitational interaction. In modern theory, Planck’s constant a ppears both as the scale\nof action in quantum theory and as the scale of classical zero-point radiation in classical\nelectrodynamics.\nBecause Planck’s constant /planckover2pi1does not appear in Maxwell’s equations but only in the\nsource-free boundary condition[13] on the equations, there are two natural versions of clas-\nsical electrodynamics, one including /planckover2pi1and one not. The version which includes Planck’s\nconstant provides a natural classical explanation of the Planck sp ectrum[3] and of some\nother phenomena[12] with a scale set by /planckover2pi1.The version which appears in current textbooks\nof classical electromagnetism assumes that the source-free elec tromagnetic field vanishes and\nso offers no explanation of phenomena at the scale of /planckover2pi1.\nIt remains true that Planck’s constant /planckover2pi1has no place in classical mechanics because\nclassical mechanics (with point-particle collisions which share all ener gy) allows no zero-\npoint energy. Of course, classical mechanics also does not allow mas sless waves which carry\nenergy andmomentum. However, classical electromagnetism is quit e different fromclassical\n11mechanics. Classical electromagnetism includes massless waves, an d, indeed, has a natural\nplace for Planck’s constant as the scale of classical electromagnet ic zero-point radiation.\nIV. RADIATION EQUILIBRIUM FOR AN OSCILLATING CHARGED SYSTEM\nA. A Point Harmonic Oscillator Does Not Determine Radiation Equilibrium\nIn the discussion above, we have seen the reappearance of the sa me two oscillator ener-\ngies (equipartition and zero-point energy) which appeared from th e limits of Wien’s law (5),\narising from the thermodynamics of the harmonic oscillator. Again it m ust be emphasized\nthat since a charged harmonic oscillator treated in the dipole approx imation comes to equi-\nlibrium with anyradiation spectrum, the determination of the actual spectrum of thermal\nradiation equilibrium must be based upon other factors. Here we list s everal criteria.\nB. Specific Criteria for Radiation Equilibrium\n1. Adoption of the Nonrelativistic Oscillator Equilibrium\nIn the early years of the 20th century, just as in the modern phys ics texts of today, it\nwas assumed[1][2] that nonrelativistic statistical mechanics correc tly determines the phase\nspace of the harmonic oscillator according to the Boltzmann distribu tion\nPosc(x,p) =const×exp[−H(x,p)/(kBT)] (11)\nThen comparing equations (9) and (11), it was concluded that the s pectrum of random\nradiation in equilibrium with the oscillator must be Urad(ω) =kBT. In other words,\nnonrelativistic classical mechanics, which uses Boltzmann statistica l mechanics, must lead\nto the Rayleigh-Jeans spectrum for relativistic classical radiation.\nIn 1910, Einstein and Hopf[14][15] tried to avoid full Boltzmann statis tical mechanics\nand to use instead only the well-established average thermal kineticenergy of a massive\nnonrelativistic particle in one dimension as (1 /2)Mv2= (1/2)kBT. However, they again\narrived at the Rayleigh-Jeans spectrum for thermal radiation.\n122. Adoption of the Spectrum Invariant Under Adiabatic Change\nAn entirely different criterion involves adiabatic invariance. We saw ab ove that, under\nan adiabatic change of frequency, a charged harmonic oscillator of frequency ω0and energy\nU(ω0) (when treated in the dipole approximation) will not transfer energ y among radiation\nmodes if the radiation spectrum is Lorentz invariant. Thus the assu mption of adiabatic\ninvariance for the oscillator picks out a special radiation spectrum, that of classical electro-\nmagnetic zero-point radiation. This is also the spectrum allowed by th e limit of Wien’s law\n(5) which makes the equilibrium spectrum independent of the temper atureT.\n3. Adoption of the Spectrum Invariant Under Scattering\nA third criterion involves invariance under scattering. The charged harmonic oscillator\ntreated in the dipole approximation scatters radiation so as to make the radiation pattern\nisotropic; however, the scattering by the oscillator does not chan ge the frequency spectrum\nof the random radiation. In order to obtain a preferred spectrum based upon invariance\nunderscattering , we must go beyond use of the harmonic oscillator treated in the dipo le\napproximation.\na. Two Oscillator Extensions Leading to Equilibrium Radiat ion Spectra There are two\nnatural extensions of the harmonic oscillator model. The first poss ible extension retains the\nelectromagnetic aspect (continuing the use of the dipole approxima tion), but changes the\nmechanics (going beyond the harmonic oscillator over to an anharmo nic oscillator). The\nsecond possible extension retains the mechanics (continuing the us e of a purely harmonic\nmechanical oscillator), but changes the electromagnetic connect ion (going beyond the dipole\nradiation approximation over to full electromagnetic interaction at all harmonics). The\ntwo possible extensions give different preferred equilibrium spectra . The nonrelativistic\nmechanical extension while retaining the dipole radiation approximatio n gives the Rayleigh-\nJeans spectrum. The relativistic electromagnetic extension while re taining the harmonic\nmechanical behavior gives classical zero-point radiation. Both ext ensions will be discussed\nbelow.\n13C. Radiation Scattering by an Anharmonic Nonrelativistic Mechanical System in\nthe Dipole Approximation\nIn order to avoid the use of the ideas of classical statistical mecha nics, various physicists\nhave considered the transition beyond harmonic oscillators to nonr elativistic anharmonic\noscillators which will scatter radiation and change the spectrum of t he random radiation.\nIn 1924, Van Vleck[16] considered (and partially published) a genera l analysis of nonlin-\near nonrelativistic mechanical systems in scattering equilibrium wher e the radiation was\ntreated in the dipole approximation. Work by other authors has con sidered scattering by\na nonrelativistic nonlinear oscillator,[17] and scattering by nonrelativ istic potentials where\nthe charged particle momentum took relativistic form.[18] All of thes e classical scattering\ncalculations arrived at the Rayleigh-Jeans spectrum. None of thes e calculations treats a\nrelativistic scattering system.\nA nonrelativistic anharmonic oscillator with a nonlinear term in energy αx3and Hamil-\ntonian\nH(x,p) =p2/(2m)+(1/2)mω2\n0x2+αx3(12)\nprovides a simple example of a scattering system which will enforce th e Rayleigh-Jeans\nspectrum for radiation equilibrium. The equation of motion takes the form\nm¨x=−mω2\n0x+3αx2+mτ...x+eEx(0,t), (13)\nand continues the use of the dipole approximation in the radiation dam pingmτ...xand in\nthe electromagnetic force eEx(0,t). The mechanical motion of the oscillator now involves\na mechanical oscillation at frequency ω1which is different from ω0and which depends upon\nthe constant αgiving the scale of the nonlinear term.[19] The mechanical motion now\nincludes the harmonics[19] of the frequency ω1\nx(t) =D0+D1cos[ω1t+φ1]+D2cos[2ω1+φ2]+... (14)\nSince the mechanical motion has oscillations at all the harmonics nω1of the fundamental\nmechanical frequency ω1, the nonlinear oscillator will interact with radiation in the dipole\napproximation at all the harmonics of the oscillation frequency ω1. This interaction with\nradiation will transfer energy among the radiation modes, and so will lead to a unique\nequilibrium spectrum. We note that the radiation equilibrium is determin ed by the non-\nrelativistic mechanical system and not by any properties of the radiation with which the\n14system has minimal (dipole) interaction. In all treatments with nonr elativistic nonlinear\nmechanical scatterers connected to radiation through the dipole approximation, one finds\nthe Rayleigh-Jeans spectrum as the unique spectrum of radiation e quilibrium. When van\nVleck[16] partially published his analysis in 1924, the work was regarde d as confirming that\nthe Rayleigh-Jeans spectrum was the appropriate equilibrium radiat ion spectrum expected\nwithin classical physics.\nD. Radiation Scattering by a Charged Harmonic Oscillator with Full Electro-\nmagnetic Interactions\n1. One-Dimensional Electromagnetic Harmonic Oscillator\nIn order to emphasize electromagnetic interactions in contrast to arbitrary mechanical\npotentials, we mention here the possibility of a one-dimensional oscilla tor where the oscillat-\ning particle of mass mand charge emoves under purely electromagnetic forces. We imagine\na charge ewhich is constrained to move along the straight line between two char gesq(of\nthe same sign as e) separated by a distance l. The charge ewill then oscillate in one di-\nmension along the line between the charges q,but is in unstable equilibrium against motion\nto the side if the constraint were removed. For small oscillations, th e mechanical motion of\neis harmonic with (angular) frequency ω0= [32eq/(ml3)]1/2. For any finite amplitude of\noscillation, there will be non-harmonic contributions from the electr omagnetic forces on e\ndue to the charges q. However, as the amplitude of oscillation becomes ever smaller, the\noscillation becomes ever-more-nearly harmonic. Furthermore, on transformation to a new\ninertial frame S′, the electromagnetic system would have relativistic behavior in the lim it of\nzero-oscillation amplitude at a new Lorentz-transformed oscillation frequency ω′\n0.\n2. Relativistic Harmonic Oscillator\nIf we adopt this one-dimensional electromagnetic model for the os cillating charge ewhere\nthe restoring force arises from the electrostatic field in the rest f rame of the charges q, then\nthe relativistic equation for the constrained motion for the charge ein electromagnetic fields\nisdp/dt=eE(r,t).In the approximation that the velocity of the charge erelative to the\n15chargesqis small compared to c,one can approximate the relativistic change in momentum\nas mass times acceleration, accurate through first order in the ra tiov/c. Thus we have\ndp\ndt=d\ndt/bracketleftBigg\nmv\n(1−v2/c2)1/2/bracketrightBigg\n=d\ndt/bracketleftbigg\nmv/parenleftbigg\n1+v2\n2c2+.../parenrightbigg/bracketrightbigg\n≈d(mv)\ndt, (15)\nprovided that we can neglect terms of order ( v/c)2. Our electromagnetic oscillator can be\ntreated as a relativistic system provided that its oscillation velocity is small. The oscillator\nsystem is fully relativistic only in the limit v→0.\n3. Radiation Equilibrium Beyond the Dipole Approximation\nIn his textbook of classical electrodynamics, Jackson points out t hat[20] “Appreciable\nradiation in multiples of the fundamental [oscillatory frequency] can occur because of rela-\ntivistic effects ... even though thecomponents of velocity aretruly sinusoidal, or it can occur\nif the components of the velocity are not sinusoidal, even though pe riodic.” All of the previ-\nousscattering calculations have involved mechanical motionswhich are“not sinusoidal, even\nthough periodic,” as in our Eq. (14). Thus the harmonics appear in th e purely mechanical\nmotion of the charged particle. However, recent work by Huang an d Batelaan[21] regarding\nabsorption of radiation at harmonics due to relativistic effects sugg ested the possibility of\ndetermining radiation equilibrium not from nonrelativistic mechanical m otions but rather\nfrom purely relativistic effects in classical electromagnetism.\nInproblem14inChapter 14, Jackson[20]asks astudent tocalculat e theradiationemitted\nat the harmonics nω0due to a purely sinusoidal motion at ω0\nx=Dcos[ω0t+φ]. (16)\nThus the treatment of the multipole moments of the oscillating charg e can be extended\nbeyond the dipole term to include the quadrupole moment of the harm onic oscillator which\nwill emit radiation at frequency 2 ω0, and indeed can be extended to all higher multipoles.\nHowever, if the harmonic oscillator can emit quadrupole radiation at t he second harmonic\n2ω0of the oscillator’s natural frequency ω0, then it can also absorb energy at the sec-\nond harmonic. The force on the oscillator must be extended beyond the dipole inter-\nactioneEx(0,t) to include the next term in the expansion of the true force eEx(x,t) as\neEx(0,t) +ex/bracketleftBig\n∂Ex(/hatwideix′,t)/∂x′/bracketrightBig\nx′=0. Thus a purely sinusoidal mechanical motion of the\n16oscillator combined with the relativistic radiation analysis can lead to a specific equilibrium\nspectrum of random classical radiation. Here for the first time in th e classical physics\nscattering literature, radiation equilibrium is determined not by nonrelativistic mechanical\nconsiderations but by relativistic electromagnetic aspects.\n4. Zero-Point Radiation as the Oscillator-Scattering Equi librium Spectrum\nThe calculation for the equilibrium radiation spectrum of a electromag netic charged har-\nmonic oscillator of small amplitude has been carried out.[22] The equilibr ium spectrum is a\nLorentz-invariant radiation spectrum. The Rayleigh-Jeans spect rum is not Lorentz invari-\nant, and so is notthe spectrum of radiation equilibrium for a charged classical harmon ic\noscillator when treated beyond the dipole approximation for the rad iation interaction. Only\nLorentz-invariant zero-point radiation will serve as an equilibrium sp ectrum.\n5. Use of Nonrelativistic versus Relativistic Theory\nA harmonic oscillator is like a clock with frequency ω0. Thepointharmonic oscillator\nallows only a dipole connection to radiation, and is in equilibrium with anyspectrum of\nisotropic random radiation. In the dipole approximation, the radiatio n energy Urad(ω0)\nmerely determines the one number giving the oscillator energy Uosc(ω0). The point os-\ncillator clock itself can assume either Galilean or Lorentz transforma tion properties when\nviewed in a new inertial frame. However, if the oscillator motion has fin ite amplitude, then\nthe analysis of the oscillator must choose between the nonrelativist ic and the relativistic\ntransformations. If nonrelativistic mechanical motion beyond harmonic motion is involved,\nthen there is enormous flexibility in the choice of the mechanical inter action potential V(x);\nin equilibrium, the nonrelativistic nonlinear oscillator, with a dipole conne ction to radiation,\nenforces a frequency-independent constant energy on the rad iation modes, corresponding to\nthe frequency-independent energy kBTallowed by the Wien law (5). On the other hand, if\nrelativistic electromagnetic interactions are assigned to the harmonic oscillato r, then the ex-\ntension from the dipole moment to the quadrupole moment and beyon d allows no flexibility\nwhatsoever; the harmonic oscillator motion is completely determined , and the equilibrium\nradiationspectrumassumesaLorentz-invariantform, correspo ndingtothezero-pointenergy\n17limit (1/2)/planckover2pi1ω0allowed by the Wien law (5).\n6. The Relativistic Limit and Classical Zero-Point Radiation\nThe electromagnetic oscillator is fully relativistic only in the limit as its velo city goes to\nzero. However, it turns out that the radiation balance at each har monic does not depend\nupon the actual oscillation amplitude for small oscillations, provided t hat the amplitude is\nnon-vanishing.[22] Thus the radiation equilibrium continues to hold with out any change\nas the amplitude of oscillation goes toward zero, and the oscillator mo tion becomes ever\ncloser to the relativistic limit. We expect that the radiation equilibrium o btained from\napproximately relativistic oscillator motion will hold even in the relativist ic limit.\nIf one uses the action-angle variables wandJ, then the Hamiltonian in Eqs. (2) and (4)\nmakes no reference to the mass of the oscillator or to the amplitude of oscillation, and the\nphase space distribution for the oscillator given in Eq. (9) becomes\nPosc(w,J) =const×exp/bracketleftbigg\n−Jω0\nUrad(ω0)/bracketrightbigg\n. (17)\nThis phase space distribution gives equilibrium for the oscillator at all t he harmonics nω0\nin a Lorentz-invariant radiation spectrum Urad(ω). If we take the Lorentz-invariant radia-\ntion spectrum (required for full radiation equilibrium by the oscillator ) as that of classical\nelectromagnetic zero-point radiation Uzprzp(ω) = (1/2)/planckover2pi1ω, then the oscillator phase space\n(17) takes the form\nPosczp(w,J) =const×exp/bracketleftbigg\n−J\n/planckover2pi1/2/bracketrightbigg\n, (18)\nwhere the oscillator frequency ω0has cancelled out leaving a phase space which independent\nof frequency. This phase space distribution for the oscillator is the same as the phase space\nin Eq. (10) for each mode of electromagnetic radiationin the zero-p oint radiation field. The\nequilibrium phase space distribution in Eq. (18) is invariant under an ad iabatic change of\nfrequency ω0.\nV. SCALING IN NONRELATIVISTIC MECHANICS AND IN RELATIVISTIC\nELECTRODYNAMICS\nIn order to broaden our understanding of the contrasts arising in treatments of classical\nradiation equilibrium, we now turn to the matter of scaling. Scaling can be regarded as a\n18change in the fundamental unit used to evaluate some physical qua ntity, or as a multiplica-\ntive change in the physical quantity itself. The contrasting scaling a spects of nonrelativistic\nclassical mechanics and of relativistic classical electrodynamics are reflected in their deter-\nminations of thermal equilibrium.\nA. Scaling Aspects of Nonrelativistic Mechanics\n1. Separate Scalings in Length, Time, and Energy\nBecause nonrelativistic mechanics involves no fundamental consta nts, it allows separate\nscalings in length as σl, in time as σt, and in energy as σU. For example, any nonrelativistic\nclassical mechanical system can be reimagined as a system where all the lengths are twice\nas large ( σl= 2), the times are three times as long ( σt= 3), and the energies are four times\nas great ( σU= 4). This flexibility arises since the only forms of energy are kinetic en ergy\n(which can be scaled through the mass m), and potential energy V(x,y,z) (which can be\nrescaled through the constants connecting distance to energy) . For a nonlinear mechanical\noscillator of mass m, harmonic frequency ω0, and nonlinear parameter α, as in Eq. (12), the\nquantities m,ω0, andαare all freely adjustable, corresponding to allowing separate scalin gs\nσl,σt,σU.\n2. Thermal Equilibrium Reflects the Three Separate Scalings\nEquilibrium involving the interaction of classical mechanical systems m ust reflect the\nthree separate scalings allowed for nonrelativistic mechanical syst ems. In an equilibrium\nsituation for nonrelativistic physics, the energy kBTmust scale separately from the length\nand time parameters of the mechanical systems. If we rescale the energy (by a factor σU)of\na mechanical system which is in equilibrium, then the equili brium of the system should not\nbe disturbed by the energy rescaling. Furthermore, if we replace a member of a mechanical\nsystem by a new mechanical member involving new lengths and times bu t with the same\nenergy, then the mechanical equilibrium will be unchanged because t he total system energy\nremains unchanged. If these scaling ideas are applied to the Wien-law expression in Eq.\n(5), they pick out only the equipartition limit involving the energy-dep endent temperature\nTbut having no dependence upon the frequency-dependent ω. Indeed the Boltzmann\n19distribution reflects this idea of a separate energy scaling. The Bolt zmann probability on\nphase space involves the mechanical energies of the constituent s ystems with no other aspect\ninvolved. Interestingly enough, the Coulomb potential (which is the only potential which\ncan be extended to a fully relativistic classical electromagnetic theory) does not fit into the\nnonrelativistic Boltzmann analysis.\nTheRayleigh-Jeansspectrumreflectsthisnonrelativisticscaling pa tternwheretheenergy\nUrad(ω,T) =kBTof each radiation mode is the same, and is entirely independent of the\nfrequency ωof the radiation mode. It is interesting that in 1924, van Vleck[23] re marked\nwith surprise regarding his nonrelativistic calculations that “...in a field o f radiation whose\nspecific energy [ ρrad(ω,T) = [ω2/(π2c3)]Urad(ω,T)] does not vary with the frequency, we\nhave the rather surprising result that the mean absorption is indep endent of the form of\nthe force function V(x,y,z) which holds the electron in the atom.” Apparently, the idea\nofkineticenergy equipartition, completely independent of the force functio nV(x,y,z),\nwas a familiar and accepted idea of nonrelativistic mechanics, wherea s the directly-related\nresult associated with rates of energy absorption and loss from th e radiation field treated in\nthe dipole approximation was regarded as “surprising.” Van Vleck’s ex pressions for energy\nemission and absorption in the Rayleigh-Jeans spectrum can be show n to give kineticenergy\nequipartition.[24]\n3. No-Interaction Theorem in Relativistic Mechanics\nMany physicists are so accustomed to using nonrelativistic classical mechanics with its\nfreedom to choose interaction potentials V(x,y,z) at will, that they are surprised that rel-\nativity places very strong restrictions on systems. Indeed, most physicists have probably\nnever heard of the “no-interaction theorem” of Currie, Jordan, and Sudarshan[4] which\n“says that only in the absence of direct particle interaction can Lor entz invariant sys-\ntems be described in terms of the usual position coordinates and co rresponding canonical\nmomenta.”[26] Thestandardgraduate-level mechanics text then simply dismisses relativistic\nideas, continuing, “The scope of the relativistic Hamiltonian framewo rk is therefore quite\nlimited and so for the most part we shall confine ourselves to nonrela tivistic mechanics.”\n20B. Scaling Aspects of Relativistic Electrodynamics\n1. Single σltU−1-Scaling of Relativistic Electrodynamics\nIn contrast to the three separate scalings in length, time, and ene rgy appearing in non-\nrelativistic mechanics, electrodynamics allows only one single scaling co nnecting together\nlength, time, and energy. This situation arises because classical ele ctrodynamics involves\nseveral fundamental constants. Length and time are coupled th rough the speed of light c,\nwhile energy and length are coupled through the electronic charge e,or through Stefan’s\nconstant aS=U/(VT4) related to the total thermal-part Uof the radiation energy in a\nbox of volume V. Therefore (relativistic) classical electrodynamics allows only one s caling\nσltU−1which preserves the values of these fundamental constants.[25] We note that Planck’s\nconstant /planckover2pi1has the same dimensions as e2/cand so is also unchanged under σltU−1-scaling.\nSuchσltU−1-scaling should not disturb electromagnetic thermal equil ibrium.\nIn addition to preserving the fundamental constants c, e,andaS, theσltU−1-scaling also\npreserves the form of Maxwell’s equations. Thus invariance under σltU−1-scaling holds\nfor Gauss’s law ∇ ·E= 4πρprovided that the electric field Escales as charge divided by\nlength squared. Faraday’s law ∇×E=−(1/c)∂B/∂tsatisfies invariance under the scaling\nprovided that the magnetic field Balso scales as charge divided by length squared. Finally,\nthe scaling for the total energy Uin a volume V,U=uV, follows from the scaling for the\nenergy density u= [1/(8π)](E2+B2).\n2.σltU−1-Scaling Allows a Function of ω/Tfor Classical Thermal Radiation\nLorentz-invariant classical zero-point radiation is σltU−1-invariant, and has no preferred\nlength, time, or energy.[27] Thus under a σltU−1-scale transformation, zero-point radiation\nis mapped onto itself. For any radiation mode of frequency ω, and energy Uradzp(ω) =\n(1/2)/planckover2pi1ω, theσltU−1-scaling will carry the radiation mode into a new radiation mode where\nthe frequency is ω′=ω/σltU−1and the energy is U′=U′/σltU−1. But then we have\nU′\nradzp(ω′) =Uradzp(ω)/σltU−1= (1/2)/planckover2pi1ω/σltU−1= (1/2)/planckover2pi1ω′=Uradzp(ω′), so that the\nfunction connecting frequency and energy is completely unchange d. Note that under this\nσltU−1-scale transformation, the phase space distribution of each radia tion mode remains\nunchanged at Pradzp(Jrad) =const×exp[−Jrad/(/planckover2pi1/2)]. On the other hand, thermal\n21radiation at temperature T >0 is not invariant under σltU−1-scaling. Since temperature T\ntransforms as an energy, the information of Wien’s law in Eq. (5) und ergoes a σltU−1-scale\ntransformation from temperature Tto a new temperature T′=T/σltU−1, while making\nthe ratio ω/Tinvariant, ω/T=ω′/T′. Thus the σltU−1-scaling of relativistic classical\nelectrodynamics is consistent with Wien’s law.\n3.σltU−1-Scaling Allows Only Zero-Point Radiation for the Relativi stic Harmonic Oscillator\nBecause of the existence of the fundamental constants candewhich require the σltU−1-\nscaling of electromagnetism, a charged (relativistic) electrodynam ic system in thermal radi-\nation can have at most one freely-adjustable parameter for fixed temperature T. After the\none parameter of the charged system has been chosen, the cond itions of thermal equilibrium\nwill determine theotherphysical quantities. Forthe(relativistic) o ne-dimensional harmonic\noscillator with its one scaling parameter ω0, the oscillator energy Uosc(ω0) is determined by\nthe initial conditions in connection with the radiation spectrum at fre quencyω0.\nIn the calculation[22] for the full electromagnetic radiation equilibriu m of an oscillator\nof small amplitude, purely electromagnetic interactions are invoked to arrive at the same\nzero-point radiation spectrum which is associated with adiabatic tra nsformation of the os-\ncillator. However, the full Planck spectrum including zero-point rad iation does notappear.\nSome physicists have taken this limitation as a sign that there is somet hing wrong with the\nanalysis, and they feel justified in their contentment with the erro neous claim that classi-\ncal physics leads to the Rayleigh-Jeans spectrum. However, the r eason for the limitation\nin the calculation is that the small-amplitude oscillator becomes a relativ istic system only\nat zero oscillation velocity, and therefore does not show the variet y of behavior of a fully\nrelativistic electromagnetic system. Indeed, the restriction of th e (relativistic) oscillator to\nthe zero-oscillation-velocity limit necessarily excludes velocity-depe ndent damping which is\ncrucial for the low-frequency part of the Planck thermal spectr um.\n224.σltU−1-Scaling Allows a Function of mc2/(kBT)for the Relativistic Classical Hydrogen\nAtom\na. Scaling for the Classical Hydrogen Atom The limiting considerations for the (ap-\nproximately relativistic) harmonic oscillator do not appear for the re lativistic classical hy-\ndrogen atom. The classical hydrogen atom consists of a point char geeof massmin a\nCoulomb potential V(r) =−e2/rin the presence of random classical radiation. This\nsystem is fully relativistic when considered as part of classical electr odynamics. The one\nfreely-adjustable scaling parameter is the mass m.Using this mass and the fundamental\nconstants eandc, one can form the energy mc2, the length e2/(mc2), and time e2/(mc3).\nUnder a σltU−1-scale transformation, the mass mis carried into mass m′=m/σltU−1, but\nvelocities are unchanged. Since classical zero-point radiation is σltU−1-scale invariant, then\nthe classical hydrogen atom in zero-point radiation will have its equilib rium phase space\nunchanged by the transformation. Since velocities are unchanged under the transforma-\ntion, this fits with the absence of velocity-dependent damping in zer o-point radiation; only\nacceleration-dependent radiation damping is involved in equilibrium. It is familiar that for\ncircular orbits, the speed of a particle in a circular orbit is v=e2/JwhereJis the angular\nmomentum. Under a σltU−1-scale transformation, the quantities v, e,andJall remain\nunchanged.\nb. Scaling for Hydrogen in Zero-Point Radiation In zero-point radiation, the scale\nof the random radiation is given by /planckover2pi1which has the same dimensions as e2/c. Thus\nin the limit which removes all factors of c, we find that the classical hydrogen atom in\nzero-point radiation has a typical length [ e2/(mc2)][/planckover2pi1c/e2]2=/planckover2pi12/(me2), a typical time\n[e2/(mc3)][/planckover2pi1c/e2]3=/planckover2pi13/(me4), and a typical energy mc2[e2/(/planckover2pi1c)]2=me4//planckover2pi12. These quan-\ntities, which involve no factors of c, are familiar from the Bohr model of hydrogen. They\nalso appear in work involving both numerical simulations and the Fokke r-Planck equation\nfor the classical hydrogen atom in classical zero-point radiation.[28 ]\nc. Scaling for Hydrogen in Thermal Radiation with T >0 If the classical hydrogen\natom is in equilibrium with thermal radiation at temperature T >0, then there will be\nvelocity-dependent dampinganda σltU−1-scaletransformationwillindeedchangethesystem.\nThe relativistic particle in a Coulomb potential allows a phase space dist ribution dependent\nupon both the mass m, and also the fundamental constants eandc. Thus the phase space\n23for the Coulomb situation can involve the characteristic energy rat iomc2/(kBT) which is\nfreely adjustable in both mand inT.This freely-adjustable ratio allows the possibility of a\ntransition between a high-frequency region dominated by accelera tion-dependent radiation\ndamping and a low-frequency region dominated by velocity-depende nt damping.\nFor the Coulomb potential, when the orbital radius is small, the velocit y is high and\nthe frequency is high, corresponding to the zero-point radiation p art of the spectrum where\nacceleration-dependent radiation damping dominates and velocity- dependent damping is\nminor. On the other hand, when the orbital radius is large, the veloc ity and frequency\nare low, and acceleration-based radiation damping is also low. In this lo w-frequency re-\ngion, velocity-dependent damping due to thermal radiation for T >0 might well dom-\ninate. This situation can be illustrated[29] by considering relativistic circular orbits\ncharacterized by angular momentum Jwhere the velocity is v=e2/J,the energy is\nU(J) =mc2/radicalBig\n1−[e2/(Jc)]2, the radius is r= [e2/(mc2)](Jc/e2)2/radicalBig\n1−[e2/(Jc)]2, and\nthe frequency is ω=∂U/∂J= (mc3/e2)[e2/(Jc)]3//radicalBig\n1−[e2/(Jc)]2. The ratio of orbital\nfrequency ωto temperature Tcan be written as\n/planckover2pi1ω\nkBT=mc2\nkBT/parenleftbigg/planckover2pi1c\ne2/parenrightbigg/parenleftBigg\n1−/parenleftbigge2\nJc/parenrightbigg2/parenrightBigg−1/2/parenleftbigge2\nJc/parenrightbigg3\n. (19)\nWe see from Eq. (19), that for changing temperature Tand for fixed mass mand fixed\nangular momentum J(which corresponds to fixed frequency ω), the system changes from\nthe high-frequency region of the thermal spectrum /planckover2pi1ω/kBT >>1,to the low-frequency\nregion/planckover2pi1ω/kBT <<1 exactly in unison with the ratio mc2/(kBT).The transition between\nthe two regions of behavior is mediated by the ratio mc2/(kBT). An analogous ratio does\nnot exist for the (approximately relativistic) harmonic oscillator.\nVI. MISCELLANEOUS ASPECTS OF THERMAL EQUILIBRIUM\nA. Full Thermal Equilibrium Requires Both Velocity-Dependent Damping and\nAcceleration-Dependent Damping\nIndeed it is interesting to see that within classical physics, thermal radiation equilibrium\nfor non-zero temperature T >0 requires both acceleration-dependent radiation damping\nand velocity-dependent damping. The Einstein-Hopf analysis[14][15] of 1910 considered a\n24particle of large mass Mcontaining an electric dipole oscillator which experienced velocity-\ndependentdampingwhenmovingthroughrandomradiation,butthe rewerenonon-radiation\nforces and so the particle motion experienced no acceleration-dep endent damping. Assum-\ning that the average kinetic energy of the one-dimensional motion o f the large mass M\nwas (1/2)kBT, the Einstein-Hopf analysis led to the low-frequency, velocity-dep endent,\nRayleigh-Jeans part of the Planck law for thermal radiation equilibriu m. Indeed, the\nvelocity-dependent damping without acceleration-dependent rad iation damping is analo-\ngous to the situation which arises in nonrelativistic mechanics where thermal equilibrium\ndepends entirely upon velocity-dependent damping, and no acceler ation-dependent damp-\ning can appear. It was only later when the Einstein-Hopf analysis was extended[30] to\nalso include acceleration-dependent radiation damping of the partic le motion (arising from\nnon-radiation forces on the particle) that the classical Einstein-H opf analysis led to the full\nPlanckspectrumincluding zero-pointradiation. Furthermore, the inclusionofclassicalzero-\npoint radiation in the Planck spectrum suggests the possibility of sup erfluid-like behavior\nfor an Einstein-Hopf particle at low temperatures.[31]\nB. Information Contained within the Rayleigh-Jeans and Zero-Point Radiation\nSpectra\n1. Zero-Point Spectrum Has Least Information\nZero-point radiation involves the spectrum of random radiation with least possible infor-\nmation. The spectrum is Lorentz invariant and has no preferred ine rtial frame, and hence\nno velocity-dependent damping. The spectrum is also scale invariant , and has no preferred\nlength or time or energy. Even in curved spacetime, zero-point rad iation has correlation\nfunctions which involve only the geodesic separations between the s pacetime points where\nthe correlation is evaluated.\n2. Information in the Rayleigh-Jeans Spectrum\nThe Rayleigh-Jeans spectrum contains more information than the c lassical zero-point\nradiation spectrum. Both spectra are characterized by a single pa rameter: /planckover2pi1for zero-point\n25radiation and Tfor the Rayleigh-Jeans spectrum. For the Rayleigh-Jeans spectr um, the\ntemperature Tis freely-adjustable, and the energy per normal mode is UradRJ(ω,T) =kBT\nat anyfrequency ω.However, theRayleigh-Jeans spectrum determines exactly onepr eferred\ninertial frame in which it is isotropic. Furthermore, at fixed tempera tureT, one can\ndetermine whether one frequency is larger than another frequen cy by comparing their phase\nspace distributions. Thus the phase space distribution for electro magnetic radiation in the\nRayleigh-Jeans spectrum is PradRJ(J,ω,T) =const×exp[−Jω/(kBT)] so that a larger\nvalue of frequency ωcorresponds to a phase space distribution more concentrated ne ar\nJ= 0. Ontheother hand, theclassical zero-point radiationspectru mUradzp(ω) = (1/2)/planckover2pi1ω\nhas a phase space distribution which is the same for radiation modes o f any frequency ω,\nPradzp(J,ω) =const×exp[−J/(/planckover2pi1/2)].Zero-point radiation has no preferred inertial frame\nand is isotropic in every inertial frame. If a harmonic oscillator of sma ll amplitude were\nto scatter zero-point radiation so as to enforce any radiation spe ctrum other than the zero-\npoint radiation spectrum, the oscillator would impose a preferred ine rtial frame upon the\nscattered radiation.\n3. Thermal Radiation Spectrum as Giving Least Information in t he Presence of Zero-Point\nRadiation\nThermal radiation at T >0 involves a finite amount of energy spread over the modes\nof the zero-point radiation spectrum and so introduces a preferr ed inertial frame. The\nspectrum of zero-point radiation acts as a noise spectrum into whic h the finite amount of\nthermal energy is introduced, and the spectrum of thermal radia tion represents the mini-\nmal information consistent with the finite total available thermal en ergy and the divergent\nLorentz-invariant spectrum of zero-point radiation. The entrop y of thermal radiation at\nfrequency ωis determined by comparing the amount of thermal radiation to the a mount of\nzero-point radiation at that frequency.[3]\nC. Inconsistent Mixtures of Nonrelativistic and Relativistic Physics\nThe derivations of the Rayleigh-Jeans law for thermal radiation give n in the textbook\nliterature arise from attempts to combine nonrelativistic mechanics with relativistic elec-\n26trodynamics. Such inconsistent mixtures of nonrelativistic and rela tivistic theories satisfy\nneither Galilean invariance nor Lorentz invariance. Indeed, the res ults of the analysis de-\npend upon the inertial frame in which the analysis is done. For particle s undergoing point\ncollisions, this situation has already been exhibited explicitly in the litera ture.[32] Thus\nif one considers the point collision between two particles, using relativ istic physics for one\nand nonrelativistic physics for the other, one can solve for the mot ion of the particles after\nthe collision by using the energy and momentum conservation laws in an y inertial frame;\nhowever, the results satisfy neither the nonrelativistic conservation law for the uniform mo-\ntion of the center of mass nor the relativistic conservation law for the uniform motion of\nthe center of energy, and the results indeed depend upon the iner tial frame chosen for the\nanalysis.\nVII. CLOSING SUMMARY\nA. Two Natural Spectra Associated with Harmonic Oscillators\nAssociated with harmonic oscillators, there are two spectra of ran dom radiation which\narise in a natural manner. One of these is the Rayleigh-Jeans spect rum which is associated\nwith the equipartition ideas of nonrelativistic classical mechanics and is based uponvelocity-\ndependent damping. The other is the classical zero-point radiation spectrum which arises\nfrom an adiabatic transformation of oscillator frequency and is ass ociated with relativistic\nclassical electrodynamics which includes acceleration-dependent d amping.\nB. Oscillators in Translational Motion and Two Forms of Damping\nIf one considers not a stationary oscillator but rather an oscillator which is free to move in\nover-all translation, then one can obtain either the Rayleigh-Jean s spectrum (obtained for a\nsituation with velocity-dependent damping on the oscillator but no ac celeration-dependent\ndamping by Einstein and Hopf[14]) or the full Planck spectrum with ze ro-point radiation (if\none allows both types of damping on the oscillator[30]).\n27C. Thermal Radiation Equilibrium from Scattering\nA charged harmonic oscillator at rest with its electromagnetic intera ctions limited to the\ndipole approximation is in equilibrium with anyspectrum of radiation; the energy of the\noscillator merely matches the energy of the radiation modes at the n atural frequency of the\noscillator. If we wish to go beyond this ambiguous radiation situation a nd to determine\na preferred equilibrium radiation spectrum under scattering , then there are two natural\nscattering extensions. If we continue the limitation on the electromagnetic inte ractions\nto the dipole approximation but go to a more complex nonrelativistic mechanical motion,\nthen we arrive at the Rayleigh-Jeans spectrum. On the other hand , if we treat the charged\nharmonicoscillatorasafullyelectromagneticsystem but withonly approximately-relativistic\nharmonic oscillator mechanical behavior, then we arrive at classical zero-point radiation as\nthe preferred radiation spectrum. Scattering by a fully relativistic electromagnetic system,\nsuch as the classical hydrogen atom, has the possibility of giving the full Planck radiation\nspectrum with zero-point radiation as the equilibrium classical radiat ion spectrum.\nD. Relativity and Thermal Radiation\nIf one reads James Jeans’ Report on Radiation and the Quantum-Theory published in\n1914 or if one reads the modern physics textbooks published in the la st few years, one would\nfind no inkling that special relativity might have some relevance to the blackbody radia-\ntion problem. However, the equilibrium radiation spectrum is determin ed by assumptions\non both the mechanical motion (whether nonrelativistic or relativist ic) and the relativistic\nelectromagnetic interactions. In any case, the radiation equilibrium of the fully electromag-\nnetic charged harmonic oscillator of small amplitude does notlead to the Rayleigh-Jeans\nspectrum. It is an outright error to claim that classical physics lead s inevitably to the\nRayleigh-Jeans spectrum.\n[1] See for example, R. Eisberg and R. Resnick, Quantum Physics of Atoms, Molecules, Solids,\nNuclei, and Particles 2nd ed. (Wiley, New York 1985); K. S. Krane, Modern Physics , 2nd\ned. (Wiley, New York 1996); R. Taylor, C. D. Zafiratos, and M. A . Dubson, Modern Physics\n28for Scientists and Engineers , 2nd ed. (Pearson, New York, 2003); S. T. Thornton and A.\nRex,Modern Physics for Scientists and Engineers , 4th ed. (Brooks/Cole, Cengage Learning,\nBoston, MA, 2013).\n[2] J. H. Jeans, Report on Radiation and the Quantum Theory (www.ForgottenBooks.org, 2013).\nThis is a reproduction from the report originally published in 1914.\n[3] T. H. Boyer, “Blackbody radiation in classical physics: A historical perspective,” Am. J. Phys.\n86, 495-509 (2018).\n[4] D. G. Currie, T. F. Jordan, and E. C. G. Sudarshan, “Relati vistic Invariance and Hamiltonian\ntheories of interacting particles,” Rev. Mod. Phys. 35, 350-375 (1963).\n[5] T. H. Boyer, “Thermodynamics of the harmonic oscillator : Wien’s displacement law and the\nPlanck spectrum,” Am. J. Phys. 71, 866-870 (2003).\n[6] T. H. Boyer, “Thermodynamicsof the harmonic oscillator : derivation of the Planck blackbody\nspectrum from pure thermodynamics,” Eur. J. Phys. 40,025101(16pp) (2019).\n[7] See for example, M. Planck, The Theory of Heat Radiation (Dover, New York 1959).\n[8] T. W. Marshall, “Random electrodynamics,” Proc. R. Soc. A276, 475-491 (1963).\n[9] T. H. Boyer, “Random electrodynamics: Thetheory of clas sical electrodynamics with classical\nelectromagnetic zero-point radiation,” Phys. Rev. D 11, 790-808 (1975).\n[10] B. H. Lavenda, Statistical Physics: A Probabilistic Approach (Wiley, New York 1991), pp.\n73-74.\n[11] T. H. Boyer, “Understanding zero-point energy in the co ntext of classical electromagnetism,”\nEur. J. Phys. 37, 055206(14) (2016).\n[12] A review of the work on classical electromagnetic zero- point radiation up to 1996 is provided\nby L. de la Pena and A. M. Cetto, The Quantum Dice - An Introduction to Stochastic Elec-\ntrodynamics (Kluwer Academic, Dordrecht 1996). For a more recent short r eview, see T.\nH. Boyer, “Stochastic Electrodynamics: The Closest Classi cal Approximation to Quantum\nTheory,” Atoms 7(1), 29-39 (2019).\n[13] T. H. Boyer, “The contrasting roles of Planck’s constan t in classical and quantum theories,”\nAm. J. Phys. 86, 280-283 (2018).\n[14] A. Einstein and L. Hopf, “Statistische Untersuchung de r Bewegung eines Resonators in einem\nStrahlungsfeld,” Annalen der Physik (Leipzig) 33, 1105-1115 (1910).\n[15] For modernnotation, see P. W. Milonni, The Quantum Vacuum: An Introduction to Quantum\n29Electrodynamics (Academic Press, Boston1994), pp. 11-14.\n[16] J. H. van Vleck, “The absorption of radiation by multipl y periodic orbits, and its relation to\nthe correspondence principle and the Rayleigh-Jeans law: P art II. Calculation of absorption\nby multiply periodic orbits,” Phys. Rev. 24, 347-365 (1924); “A correspondence principle for\nabsorption,” Jour. Opt. Soc. Amer. 9, 27-30 (1924).\n[17] T. H. Boyer, “Equilibrium of random classical electrom agnetic radiation in the presence of a\nnonrelativistic nonlinear electric dipole oscillator,” P hys. Rev. D 13, 2832-2845 (1976).\n[18] R. Blanco, L. Pesquera, and E. Santos, “Equilibrium bet ween radiation and matter for clas-\nsical relativistic multiperiodic systems. Derivation of M axwell-Boltzmann distribution from\nRayleigh-Jeans spectrum,”Phys.Rev.D 27, 1254-1287 (1983); “Equilibriumbetweenradiation\nand matter for classical relativistic multiperiodic syste ms. II. Study of radiative equilibrium\nwith Rayleigh-Jeans radiation,” Phys. Rev. D 29, 2240-2254 (1984).\n[19] M. Born, The Mechanics of the Atom (Ungar, New York 1970), pp. 66-71.\n[20] J. D. Jackson, Classical Electrodynamics 3rd ed. (John Wiley & Sons, New York, 1999), p.\n704, problem 14.22.\n[21] W. C-W. Huang and H. Batelaan, “Discrete Excitation Spe ctrum of a Classical Harmonic\nOscillator in Zero-Point Radiation,” Found. Phys. 45, 333-353 (2015).\n[22] T. H. Boyer, “Equilibrium for classical zero-point rad iation: detailed balance under scattering\nby a classical charged harmonic oscillator,” J. Phys. Commu n.2, 105014(17) (2018).\n[23] See ref. 16, p. 359. Van Vleck’s comment involves radiat ion absorption but is related to\nnonrelativistic kinetic energy and to the nonrelativistic Larmor formula for radiation emission.\nThus for a general oscillation such as suggested by our Eq. (1 4), the average kinetic energy\nin thenthharmonic involves ( nω1)2D2\nnwhereas the average radiation emission by the nth\nharmonic is ( nω1)2times as large, involving ( nω1)4D2\nn= (nω1)2/bracketleftBig\n(nω1)2D2\nn/bracketrightBig\n.\n[24] T. H. Boyer, “Statistical equilibrium of nonrelativis tic multiply periodic classical systems and\nrandom classical electromagnetic radiation,” Phys. Rev. A 18, 1228-1237 (1978).\n[25] T. H. Boyer, “Scaling symmetries of scatterers of class ical zero-point radiation,” J. Phys. A:\nMath. Theor. 40, 9635-9642 (2007); “Scaling symmetry and thermodynamic eq uilibrium for\nclassical electromagnetic radiation,” Found. Phys. 19, 1371-1383 (1989).\n[26] H. Goldstein, C. Poole, and J. Safko, Classical Mechanics 3rd ed. (Addison-Wesley, New York,\n2002), p. 353.\n30[27] T. H. Boyer, “Conformal Symmetry of Classical Electrom agnetic Zero-Point Radiation,”\nFound. Phys. 19, 349-365 (1989).\n[28] For work on the classical hydrogen atom in classical zer o-point radiation, see D. C. Cole and\nY. Zou, “Quantum Mechanical Ground State of Hydrogen Obtain ed from Classical Electrody-\nnamics,” Phys. Lett. A 317, 14-20 (2003), and T. H. Boyer, “Relativity andRadiation Ba lance\nfor the Classical Hydrogen Atom in Classical Electromagnet ic Zero-Point Radiation,” Eur. J.\nPhys.42, 025205 (24pp) (2021).\n[29] T. H. Boyer, “Unfamiliar trajectories for a relativist ic particle in a Kepler or Coulomb poten-\ntial,” Am. J. Phys. 75, 992-997 (2004).\n[30] T. H. Boyer, “Derivation of the Blackbody Radiation Spe ctrum without Quantum Assump-\ntions,” Phys. Rev. 182, 1374-1383 (1969).\n[31] T. H. Boyer, “Particle Brownian motion due to random cla ssical radiation: Superfluid-like\nbehavior in classical zero-point radiation,” Eur. J. Phys. 41, 055103 (17pp) (2020).\n[32] T.H.Boyer, “Illustratingsomeimplications ofthecon servation lawsinrelativistic mechanics,”\nAm. J. Phys. 77, 562-569 (2009).\n31" }, { "title": "1709.03347v1.Comparison_of_damping_mechanisms_for_transverse_waves_in_solar_coronal_loops.pdf", "content": "Draft version November 8, 2018\nTypeset using L ATEX default style in AASTeX61\nCOMPARISON OF DAMPING MECHANISMS FOR TRANSVERSE WAVES IN SOLAR CORONAL LOOPS\nMar\u0013\u0010a Montes-Sol \u0013\u0010s1, 2and I ~nigo Arregui1, 2\n1Instituto de Astrof\u0013 \u0010sica de Canarias, E-38205 La Laguna, Tenerife, Spain\n2Departamento de Astrof\u0013 \u0010sica, Universidad de La Laguna, E-38206 La Laguna, Tenerife, Spain\n(Received; Revised; Accepted)\nSubmitted to ApJ\nABSTRACT\nWe present a method to assess the plausibility of alternative mechanisms to explain the damping of magnetohydro-\ndynamic (MHD) transverse waves in solar coronal loops. The considered mechanisms are resonant absorption of kink\nwaves in the Alfv\u0013 en continuum, phase-mixing of Alfv\u0013 en waves, and wave leakage. Our methods make use of Bayesian\ninference and model comparison techniques. We \frst infer the values for the physical parameters that control the wave\ndamping, under the assumption of a particular mechanism, for typically observed damping time-scales. Then, the\ncomputation of marginal likelihoods and Bayes factors enable us to quantify the relative plausibility between the alter-\nnative mechanisms. We \fnd that, in general, the evidence is not large enough to support a single particular damping\nmechanism as the most plausible one. Resonant absorption and wave leakage o\u000ber the most probable explanations in\nstrong damping regimes, while phase mixing is the best candidate for weak/moderate damping. When applied to a\nselection of 89 observed transverse loop oscillations, with their corresponding measurements of damping times scales\nand taking into account data uncertainties, we \fnd that only in a few cases positive evidence for a given damping\nmechanism is available.\nKeywords: magnetohydrodynamics (MHD) | methods: statistical | Sun: corona | Sun: oscillations\n| waves\nCorresponding author: Mar\u0013 \u0010a Montes-Sol\u0013 \u0010s\nmmsolis@iac.esarXiv:1709.03347v1 [astro-ph.SR] 11 Sep 20172 Montes-Sol \u0013\u0010s & Arregui\n1.INTRODUCTION\nThe damping of magnetohydrodynamic (MHD) waves in solar coronal structures is a commonly observed phe-\nnomenon and a source of information about their physical conditions, dynamics, and energetics. The study of the\ndamping of transverse waves has attracted particular attention, since the \frst imaging observations of decaying trans-\nverse coronal loop oscillations by Aschwanden et al. (1999) and Nakariakov et al. (1999). High resolution imaging and\nspectroscopic observations with ground- and space-based observatories such as Hinode, CoMP, SDO/AIA, STEREO,\nHi-C, and IRIS have enabled us to measure transverse wave dynamics with increasing precision in almost all layers of\nthe solar atmosphere (see e.g., Okamoto & De Pontieu 2011; White & Verwichte 2012; Verwichte et al. 2013; Morton\n& McLaughlin 2013; Thurgood et al. 2014; Morton et al. 2015; Okamoto et al. 2015). Recent observational analyses\nhave permitted the creation of databases containing the oscillation properties for a large number events and with the\ninclusion of measurement errors (Verwichte et al. 2013; Goddard et al. 2016). The increase in the number of measured\nevents and their properties including their uncertainty has led to advances in statistical seismology of coronal loops.\nVerwichte et al. (2013) considered a statistical approach to obtain information on the loop cross-sectional structuring\nparameters by forward modelling of scaling laws between periods and damping times, showing that restrictions can\nbe found to the loop's density contrast and inhomogeneity layer. Following Bayesian methods, a number of studies by\nArregui & Asensio Ramos (2011); Arregui et al. (2013a,b); Asensio Ramos & Arregui (2013); Arregui & Asensio Ramos\n(2014); Arregui et al. (2015) have shown how information on the uncertainty of the inferred coronal loop parameters\nand the plausibility between alternative models can be assessed.\nAlthough damping is not an omnipresent phenomenon, see e.g., An\fnogentov et al. (2013, 2015) for examples of\ndecay-less oscillations or Wang et al. (2012) for an example of growing oscillations, theoretical explanations for the\nphysical origin of the damping of transverse waves are abundant. Because direct viscous and resistive difussion time\nscales are too long in uniform plasmas, mechanisms based on the cross-\feld or \feld-aligned inhomogeneity of the wave\nguides become relevant. In the context of coronal loop oscillations, the discussion rapidly focused on mechanisms\nsuch as resonant absorption (Goossens et al. 2002; Ruderman & Roberts 2002; Goossens et al. 2006), phase mixing\nof Alfv\u0013 en waves (Heyvaerts & Priest 1983), lateral wave leakage (Spruit 1982; Cally 1986; Roberts 2000; Cally 2003)\nor foot-point leakage at the chromospheric density gradient (Berghmans & de Bruyne 1995; De Pontieu et al. 2001;\nOfman 2002). Methods to assess the plausibility between alternative damping mechanisms are discussed in Roberts\n(2000); Ruderman (2005); Nakariakov & Verwichte (2005); Aschwanden (2005).\nRough estimates of damping time scales predicted by the mechanisms under consideration, for typical coronal loop\nphysical properties, and their comparison to observed damping time scales point to resonant damping as the most\nplausible mechanism, with phase mixing and wave leakage producing damping time scales that seem to be too long in\ncomparison to those observed. The drawback to this method is the di\u000eculty in de\fning what a typical coronal loop\nis, since the physical parameters of observed loops cannot be directly measured.\nThe modeling and analysis of wave properties for equilibrium states that enable the simultaneous occurrence of more\nthan one mechanism is another alternative. The studies by e.g., Terradas et al. (2006), Rial et al. (2013) and Soler\net al. (2009) indicate that resonant absorption in the Alfv\u0013 en continuum is a far more e\u000ecient mechanism than lateral\nwave leakage in curved loops or than damping in the slow continuum in prominence threads. Resonant damping is\nfrequency selective, with low-frequency waves being favored in front of high frequency waves (Terradas et al. 2010).\nA comparison between the outward to inward power ratio measurements in CoMP observations of coronal waves by\nTomczyk et al. (2007) and the theoretical modeling by Verth et al. (2010) gives additional support to the resonant\ndamping model.\nOfman & Aschwanden (2002) proposed a method based on the use of scaling laws between the observed periods\nand damping times. Their suggestion is based on the assumption that each damping mechanism is characterized by\na particular power law, with a distinct power index between the damping time and the oscillation period. By \ftting\nthe observed time-scales to those predicted by each damping mechanism, one can compare the theoretically predicted\nand \ftted power indexes to discriminate between damping mechanisms. The application of this method has given\nresults that support both phase mixing (Ofman & Aschwanden 2002) and resonant absorption (White & Verwichte\n2012; Verwichte et al. 2013). It was pointed out by Arregui et al. (2008) that the use of scaling laws to discriminate\nbetween damping mechanisms is questionable. For example, the resonant damping model is able to produce data\nrealizations for which di\u000berent scaling laws with di\u000berent indexes may be obtained. The major drawback to this\nmethod is the faultiness of the assignment of a single power index to a particular damping mechanism, even more soDamping mechanisms for transverse coronal waves 3\nwhen the damping time depends on a number of loop parameters that might have an intrinsic variability and values\nthat are highly uncertain.\nWe adopt a di\u000berent approach and consider Bayesian methods to compute the level of plausibility among alternative\ndamping models, conditional on observed data and considering their uncertainty. Inference and model comparison\nmethods are applied to three particular damping mechanisms: resonant damping in the Alfv\u0013 en continuum; phase\nmixing of Alfv\u0013 en waves; and wave leakage of the principal kink mode. These damping models are selected so as to\ngive continuity to the discussion initiated by Ofman & Aschwanden (2002), but the methods here presented can be\nextended in the future to include additional damping models in a straightforward manner. The methods are \frst\napplied to hypothetical data and then to a sample of real observations from the databases compiled by Verwichte et al.\n(2013) and Goddard et al. (2016).\nThe layout of the paper is as follows. Section 2 gives a description of the damping models being compared. In\nSection 3, we present our methodology, based on the use of Bayes' rule for parameter inference and model comparison.\nOur results are presented in Section 4. We \frst apply Bayes' rule to the problem of inferring the relevant physical\nparameters, under the assumption that each of the considered damping models is true. Then, we assess the plausibility\nof each damping model conditional on both hypothetical and real observations of transverse loop oscillations. Our\nsummary and conclusions are presented in Section 5.\n2.DAMPING MODELS\nIn this work, we have studied the degree of plausibility of three damping mechanisms in explaining the observations of\ntransverse loop oscillations. The \frst considered damping model is resonant absorption in the Alfv\u0013 en continuum. This\nmechanism consists of an energy transfer between the global kink mode of a magnetic \rux tube to Alfv\u0013 en oscillations\nat the boundary of the tube, due to the transverse variation of Alfv\u0013 en velocity within a layer that separates the tube\nand the background corona. Resonant absorption has been studied extensively and in great detail and is known to be\nable to produce time and spatial damping scales comparable to those observed.\nFollowing Goossens et al. (2002) and Ruderman & Roberts (2002), we focus on the fundamental kink mode of a\ncylindrical density tube of length L and radius R with a uniform magnetic \feld along the axis of the tube and a\nnon-uniform variation of the cross-\feld density on a length-scale l. Under the thin tube ( L>>R ) and thin boundary\n(l=R<< 1) assumptions the expression for the damping ratio is\n\u001cd\nP=2\n\u0019R\nl\u0010+ 1\n\u0010\u00001; (1)\nwith\u001cdthe damping time, P the oscillation period, lthe thickness of the non-uniform layer, and \u0010=\u001ai=\u001aethe density\ncontrast between the internal ( \u001ai) and external ( \u001ae) densities. The factor 2 =\u0019is due to the assumption of a sinusoidal\npro\fle for the density at the non-uniform boundary layer.\nAccording to Equation (1), the damping ratio due to resonant absorption is a function of two parameters, \u0010andl=R.\nAs shown by Goossens et al. (2002), considering a typical contrast of \u0010= 10, the mechanism is able to explain observed\ndamping time scales for values of l=Rin between 0.1 and 0.5. Considering values of \u0010andl=Rwithin plausible ranges,\n\u00102(1;10] andl=R2(0;2], the values for the damping ratio predicted by theory are in a wide range \u001cd=Ps(0:5\u0000104).\nThe mechanism is therefore able to predict damping properties compatible with observations.\nThe next considered mechanism is associated with the phase mixing of Alfv\u0013 en waves, \frst discussed by Heyvaerts &\nPriest (1983) in the context of coronal heating. The basic idea is that Alfv\u0013 en waves propagating along the magnetic\n\feld in a medium with a transverse gradient of Alfv\u0013 en velocity become rapidly out of phase. As time progresses,\nincreasingly shorter spatial scales are created. The mechanism has been discussed as a possible reason to explain\nthe observed damping in Roberts (2000); Nakariakov & Verwichte (2005) and Ofman & Aschwanden (2002). In his\nanalysis, Roberts (2000) obtained an analytical expression for the damping ratio of the form\n\u001cd\nP=\u00123\n\u00192\u0017\u00131=3\nw2=3P\u00001=3; (2)\nwhere\u0017= 4\u0002103km2s\u00001is the coronal kinematic shear viscosity coe\u000ecient and wthe transverse inhomogeneity\nlength scale. Note that in Equation (2), the damping ratio is a function of the period. Considering a period in the range\nP2[150;1250]s and w2[0:5;20]Mm, damping ratios in the range \u001cd=Ps[0:3\u00006] are obtained, so this mechanism\ncould also explain the observed damping time scales, although it has serious limitations of applicability to transverse4 Montes-Sol \u0013\u0010s & Arregui\nloop oscillations. Ofman & Aschwanden (2002) considered phase mixing as a mechanism involving the friction between\nadjacent unresolved thin loop strands, a hypothesis that remains to be demonstrated. The mechanism is here included\nin the interest of continuity in the discussion by considering the same damping mechanisms discussed by Ofman &\nAschwanden (2002), but presenting an alternative method to the scaling law approach for their comparison.\nThe third considered mechanism is wave leakage of the principal leaky mode, as discussed by Cally (2003) and\nTerradas et al. (2007). This theoretical solution consists of a radiating wave which oscillates with the kink mode\nfrequency and looses part of its energy to the background corona. Following Cally (2003), we consider the analytic\napproximate solution for the damping ratio in the thin tube limit and valid for \u001ai>>\u001a eof the form\n\u001cd\nP=4\n\u00194\u0012R\nL\u0013\u00002\n: (3)\nConsidering values of R/L in the plausible range R/L 2[10\u00004;0:3], theory predicts damping ratios in the range\n\u001cd=Ps(0:5\u0000105).\nFor simplicity, strong assumptions are made in the adopted damping mechanisms and formulas. The resonant\ndamping formula assumes the thin tube and thin boundary approximation and the adoption of a particular radial\ndensity pro\fle. The in\ruence of considering di\u000berent density pro\fles in the inferences using resonant absorption was\nanalyzed by Arregui et al. (2015). This in\ruence was found to be important in strong damping regimes. The phase-\nmixing damping formula is valid in the strongly developed regime and the value of kinematic viscosity is kept constant.\nFor wave-leakage, the possible in\ruence of the density contrast was ignored. These assumptions lead to the existence\nof hidden variables that may vary from loop to loop. All loops do not necessarily have the same radial density pro\fle,\ntemperature, viscosity and density contrast.\nOnce the three damping mechanisms to be considered in this study are presented, it is worth clarifying that they\nare going to be compared on the basis of their ability to reproduce (in a statistical way) the observed periods and\ndamping time scales, taking into account those observed time scales and their associated uncertainties.\n3.BAYESIAN METHODOLOGY\nIn this paper, we adopt the methods of Bayesian analysis to perform parameter inference and model comparison.\nBayesian reasoning goes back to the essay by Bayes & Price (1763) and was given its current mathematical basis\nby Laplace (1774). It o\u000bers a principled way to assess the plausibility of statements conditional on the available\ninformation. In recent years, it has become widely used to perform scienti\fc inference in areas such as the physical\nsciences (von Toussaint 2011), cosmology (Trotta 2008) or exoplanet research (Gregory 2005).\nWe have used Bayesian analysis tools to perform parameter inference, assuming a particular damping model is true,\nand to compare the relative plausibilities of the three damping models described above.\n3.1. Parameter Inference\nIn the Bayesian framework the inference of a parameter set \u0012that characterizes a model M conditional on observed\ndatadis performed by making use of the Bayes' theorem\np(\u0012jM;d) =p(\u0012jM)p(djM;\u0012)\np(djM): (4)\nIn this expression, p(\u0012jM;d) is the posterior probability of the parameters conditional on the assumed model and the\nobserved data; p(\u0012jM) is their prior probability, before considering the data; and p(djM;\u0012) is the likelihood function.\nThe denominator in Equation (4) is a normalization factor called marginal likelihood given the model M, integrated\nlikelihood or model evidence. It represents the probability of the data given the model, and is computed by integrating\nthe likelihood times the prior over the parameter space\np(djM) =Z\n\u0012p(\u0012jM)p(djM;\u0012)d\u0012: (5)\nWhen interested in obtaining information on one particular parameter \u0012iof model M, we compute its marginal\nposterior as\np(\u0012ijM;d) =Z\np(\u0012jM;d)d\u00121:::d\u0012i\u00001d\u0012i+1:::d\u0012n; (6)\nwhich is an integral of the full posterior over the rest of the n model parameters.Damping mechanisms for transverse coronal waves 5\nTable 1. Kass & Raftery (1995) evidence classi\fcation.\n2lnBF kj Evidence\n0-2 Not Worth more than a bare Mention (NWM)\n2-6 Positive Evidence (PE)\n6-10 Strong Evidence (SE)\n>10 Very Strong Evidence (VSE)\n3.2. Model Comparison\nThe Bayesian framework also o\u000bers a straightforward method to compare the relative goodness of alternative models\nin explaining the observed data. Application of Bayes' theorem to N alternative models leads to\np(Mkjd) =p(Mk)p(djMk)\np(d)fork= 1;2;:::;N: (7)\nThe posteriors ratio gives us a measure of the relative plausibility between two models MkandMjas\np(Mkjd)\np(Mjjd)=p(djMk)\np(djMj)p(Mk)\np(Mj);k6=j: (8)\nIf we do not have any a priori reason to prefer one model over another before considering the data, p(Mk) =p(Mj),\nso that we can de\fne\nBFkj=p(djMk)\np(djMj);k6=j: (9)\nThis expression is called the Bayes factor (see Kass & Raftery 1995) which equals the ratio of marginal likelihoods in\nEquation (6) in the case of equal priors.\nOnce the Bayes factors are computed, the level of evidence for one model against the alternative is assessed by\nfollowing the criteria for evidence classi\fcation developed by Kass & Raftery (1995), see Table 1.\nThese general concepts and methods are applicable as long as we \frst specify priors and likelihood functions. In\nthis work, we have adopted independent priors for model parameters, so that the global prior is given by the product\nof individual priors\np(\u0012jM) =nY\ni=1p(\u0012ijM): (10)\nFurther, we consider that each parameter lies on a given plausible range, with all values being equally probable a\npriori. This de\fnes uniform priors of the form\np(\u0012ijM) =1\n\u0012max\ni\u0000\u0012min\ni;\u0012i2(\u0012min\ni;\u0012max\ni) (11)\nand zero otherwise.\nAs for the likelihood functions, we assume Gaussian pro\fles, so that the data and the theoretically predicted values\nare related as\np(djM;\u0012) =1p\n2\u0019\u001b\u0001e\u0000(\u0001obs\u0000\u0001theor)2\n2\u001b2\n\u0001; (12)\nwith \u0001 representing the observable and \u001b\u0001its associated uncertainty.\n3.3. Numerical Integration\nThe computation of marginal likelihoods and marginal posteriors using Equations (5) and (6) requires the compu-\ntation of integrals in the parameter space. In low-dimensional parameter spaces, such as the ones considered in this\nwork, the use of direct numerical integration techniques is still feasible from the point of view of computational cost.\nNevertheless, we have additionally used Markov Chain Monte Carlo sampling and Monte Carlo integration (Robert &6 Montes-Sol \u0013\u0010s & Arregui\nTable 2. Inferred parameters from Figure 2.\nMethod \u0010 l=R w R=L\n(Mm)\nNumeric 4 :1+3:9\n\u00002:40:4+0:5\n\u00000:27:5+1:1\n\u00001:10:12+0:01\n\u00000:01\nMCMC 4 :1+3:8\n\u00002:50:4+0:5\n\u00000:17:4+1:1\n\u00001:00:12+0:01\n\u00000:01\nCasella 2004) to compute marginal posteriors and marginal likelihoods, respectively. The comparison between both\ntypes of approaches gives robustness to the obtained results.\nThe marginal posteriors in Equation (6) are computed from the Markov Chain Monte Carlo (MCMC) sampling of\nthe global posterior over the space of parameters for each considered model. The marginal likelihoods in Equation\n(5) result from Monte Carlo (MC) integration. In this case, MC integration is carried out by the average of all\nvalues resulting from the evaluation of the likelihood function, at the points of the parameter space from the MCMC\nsampling of the normalized prior. These two integration methods are applied employing the emcee package of Python\n(Foreman-Mackey et al. 2013).\n4.RESULTS\nBefore considering the problem of assessing the relative plausibility of the three considered damping mechanisms in\nexplaining the observed time scales, we perform the inference of physical parameters under the assumption that each\nmechanism is true. Then, our model comparison tools are applied for the computation of the relative plausibility of\nthe three damping models, using possible values for the damping rate and its uncertainty. Finally, these techniques\nare applied to a large sample of transverse loop oscillations events for which the Bayes factors are computed.\n4.1. Parameter Inference\nFor each mechanism, the damping is determined by di\u000berent coronal loop physical parameters that cannot be\ndirectly measured. The theoretical predictions for the observable damping ratio are given by Equations (1), (2), and\n(3). The inference using observed damping ratios can in principle provide information about the cross-\feld density\ninhomogeneity, in the case of resonant absorption and phase mixing, and the coronal loop radius to length ratio in\nthe case of wave leakage. We consider di\u000berent possible values for the damping ratio and its uncertainty and perform\nBayesian inference using Equation (4), de\fning for each model Gaussian likelihoods according to Equation (12) and\nuniform priors, de\fned in Equation (11). Once the full posteriors are known, the marginalization in the corresponding\nparameter space (Equation 6) leads to the probability density function for each unknown parameter. The resulting\nmarginal posteriors are shown in Figure 1. The left hand side panels show results for di\u000berent values of the observable\ndamping rate with a \fxed uncertainty of 10%. The right hand side panels show results for a \fxed observable damping\nrate and three di\u000berent values for the uncertainty on the observable damping ratio.\nConsider \frst resonant absorption. For this mechanism, such inversion problem was \frst solved by Arregui &\nAsensio Ramos (2011) using a MCMC sampling of the posterior by using observed values for periods and damping\ntimes. Later on, Arregui & Asensio Ramos (2014) presented solutions akin to those presented here, by making use\nof the properties of the joint probability of data and parameters. Figures 1a-d show the results for this mechanism.\nFor this inference problem, \u0012=f\u0010;l=Rgandd=\u001cd=P. In principle, both the density contrast, \u0010, and the transverse\ninhomogeneity length scale, l=R, can be properly inferred, although as already noted by Arregui & Asensio Ramos\n(2014) the posteriors for the density contrast show long tails. The posterior densities peak at lower values for both\nunknowns the larger the damping ratio is (see Figures 1a and 1c) and their distributions have broader shapes the\nlarger the uncertainty in the observed damping ratio (see Figures 1b and 1d). These results indicate that, in principle,\nthe assumption that resonant absorption is the true mechanism that produces the damping of coronal loop oscillations\ntogether with measurements of the damping ratio enables us to infer the two parameters that control the cross-\feld\ndensity inhomogeneity.\nFor phase mixing, the theoretical prediction in Equation (2) does not enable to factorize the damping rate as a sole\nfunction of model parameters since the oscillation period is present in the right-hand side of this equation. For this\nreason, we perform the inference by assuming particular values for the period in the range P2[150;1250]s. This\nleaves us with an inversion problem with one observable, d=\u001cd=P, and one unknown, \u0012=fwg. Figures 1e and 1fDamping mechanisms for transverse coronal waves 7\n2 4 6 8 10\n0.00.20.40.60.81.01.2p(|r)\n(a) r=1.5\nr=3.0\nr=10.0\n2 4 6 8 10\n0.00.10.20.30.4p(|r=3)\n(b)=0.30\n=0.75\n=1.50\n0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00\nl/R02468101214p(l/R|r)(c) r=1.5\nr=3.0\nr=10.0\n0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00\nl/R012345p(l/R|r=3)(d) =0.30\n=0.75\n=1.50\n0.0 2.5 5.0 7.5 10.0 12.5 15.0 17.5 20.0\nw (Mm)0.00.20.40.60.81.0p(w|r)(e) r=1.5\nr=3.0\nr=10.0\n0.0 2.5 5.0 7.5 10.0 12.5 15.0 17.5 20.0\nw (Mm)0.000.050.100.150.200.250.300.35p(w|r=3)(f) =0.30\n=0.75\n=1.50\n0.00 0.05 0.10 0.15 0.20 0.25 0.30\nR/L020406080100120p(R/L|r)(g) r=1.5\nr=3.0\nr=10.0\n0.00 0.05 0.10 0.15 0.20 0.25 0.30\nR/L010203040506070p(R/L|r=3)(h)=0.30\n=0.75\n=1.50\nFigure 1. Marginal posterior distributions for coronal loop physical parameters under the assumption of resonant damping,\n(a)-(d); phase-mixing, (e) and (f); and wave leakage, (g) and (h). The left-hand side panels show the results for di\u000berent values\nof the damping ratio, r, with an uncertainty of 10%. The right-hand side panels show the results for a \fxed damping ratio,\nr=3, and three values for its uncertainty.8 Montes-Sol \u0013\u0010s & Arregui\n0.00 0.25 0.50 0.75 1.00 1.25 1.50 1.75 2.00\nl/R012345p(l/R|r=3.0)(a) numeric\nMCMC\n2 4 6 8 10\n0.00.10.20.30.4p(|r=3.0)\n(b) numeric\nMCMC\n0.0 2.5 5.0 7.5 10.0 12.5 15.0 17.5 20.0\nw0.000.050.100.150.200.250.300.35p(w|r=3.0)(c) numeric\nMCMC\n0.00 0.05 0.10 0.15 0.20 0.25 0.30\nR/L010203040506070p(R/L|r=3.0)(d)numeric\nMCMC\nFigure 2. A comparison between the marginal posterior distributions obtained by direct numerical integration (in red) and\nMCMC sampling (blue histograms) for a case with r=3, and an uncertainty of 10%.\nshow the results for this mechanism. The posteriors computed for the case P=150 s are only shown to not deface the\nplot. The posteriors shift toward larger values of wfor longer periods. The inference results for this mechanism show\nthat the length scale wcan be properly inferred although a larger upper limit in the parameter range is needed for\nthe case r=10. The posteriors shift towards larger values of wthe longer the damping ratio is (see Figure 1e) and\nshow broader distributions the larger the uncertainty in the observed damping ratio (see Figure 1f). These results\nindicate that the assumption that phase mixing is the true damping mechanism of coronal loop oscillations together\nwith measurements of the damping ratio enables us to infer w.\nRegarding wave leakage, the theoretical damping rate in Equation (3) is only a function of the coronal loop radius\nto length ratio, R=L. This enables us to solve the inference with one observable, d=\u001cd=P, and one parameter,\n\u0012=fR=Lg. Figures 1g and 1h show the results for this mechanism. They show that the ratio between the radius and\nthe length of coronal loops can be properly inferred. The posterior distributions shift toward smaller values of R=L\nthe larger the damping ratio is (see Figure 1g) and have broader shapes the larger the uncertainty in the observed\ndamping ratio (see Figure 1h). These results indicate that assuming that wave leakage is the true mechanism that\ncauses the damping of coronal loop oscillations together with measurements of the damping ratio, a relation among\ntwo structural features of coronal loops can be inferred.\nComparing these results obtained from direct numeric integration with the results from MCMC sampling, we can see\nthe degree of applicability of each method to the inference problem. An example of this comparison is show in Figure\n2, and the corresponding median values of damping model parameters within a credible interval of 68% are presented\nin Table 2. The agreement between the posterior distributions and the minimal di\u000berences between the mean values,Damping mechanisms for transverse coronal waves 9\nindicate the robustness of results and the applicability of both direct numerical integration and MCMC to solve this\ntype of inference problems.\n4.2. Model Comparison\nIn this section, we show the results from the application of Bayesian model comparison to the three considered\ndamping mechanisms. The comparison is performed in two principal ways. First, we consider plausible values for the\nobserved damping rate and compute the marginal likelihoods according to Equation (5), and the Bayes factors given\nby Equation (9) for each model as a function of this observable. Then, the Bayes factors between the damping models\nare computed as a function of periods and damping times on a wide range of values and for a given observational\nuncertainty. Finally, the method is applied to a large sample of transverse loop oscillation data to determine in how\nmany of the real events evidence supporting either of the considered mechanisms is found.\nFigure 3a shows the results from the computation of marginal likelihoods for the three considered damping mecha-\nnisms, as a function of the damping ratio. Resonant absorption and wave leakage o\u000ber more plausible explanations for\nlow damping rate values, while the evidence for phase mixing attains larger values from low to intermediate values of\nthe damping rate. We note that for phase mixing, as the analytic expression for the damping ratio is a function of the\nwave period, its marginal likelihood are computed for particular values of the period. These results are determined for\na \fxed value of the uncertainty on the observed damping ratio. Figures 3b-d show the marginal likelihoods, separately\nfor each mechanism, for three di\u000berent values for the error on the measured damping ratio. In all three cases, increasing\nthe uncertainty produces a decrease on the maximum value of the marginal likelihood, as well as some spread out of\nthe distributions. The discussed marginal likelihoods are computed by direct numerical integration, using Equation\n(5). Just as with the inference in the previous section, we check that they can equally well be computed by means\nof Monte Carlo integration. For each damping ratio, we carry out the MCMC sampling of the global prior given by\nEquation (10), just as we sample the posterior for the inference. Then, we evaluate the likelihood function in Equation\n(12) over the sampled parameters, and \fnally, we average the results to obtain the marginal likelihood. This is shown\nin Figure 4, where a comparison between both methods to evaluate marginal likelihoods is presented. On the one hand,\nthis result con\frms the goodness of direct numerical integration. On the other, it demonstrates that the developed\nMonte Carlo integration method can be con\fdently used when the \frst approach will not be feasible, for example, in\ncases with a larger dimension of the parameter space.\nThe quantitative assessment of the relative performance between damping mechanisms is given by the computation\nof the Bayes factors in the one-to-one comparison between the three damping mechanisms. In the following, we use\nthe subscripts 0, 1, and 2, to identify resonant absorption, phase mixing, and wave leakage, respectively. Figure 5\nshows the distribution of Bayes factors as a function of the damping ratio corresponding to one-to-one comparisons\nbetween the three mechanisms. An error measurement on the damping ratio of 10% is assumed. In Figure 5a, we see\nthat very strong evidence for resonant absorption in front of phase mixing is obtained for the lowest and the largest\nconsidered values of the damping ratio, in between 0 and 0.5 and for r >9. The \frst interval corresponds to extremely\nstrong damping regimes in which the damping time is shorter than the oscillation period. The second to rather weak\ndamping regimes. For intermediate damping ratio values, the evidence favors phase mixing, with values for 2 lnBF 10\nin between 2 and 5. These Bayes factor values correspond to positive evidence for phase mixing in front of resonant\nabsorption, according to the levels of evidence in Table 1. For the comparison between resonant absorption and wave\nleakage, Figure 5b shows very strong evidence in favor of the resonant absorption again for the lowest values of the\ndamping ratio, followed by a decrease of BF02until both Bayes factors intersect. For intermediate and large values of\nthe damping ratio, the evidence is inconclusive since the magnitude of both Bayes factors is below 2. Finally, Figure 5c\nshows the confrontation between phase mixing and wave leakage, with results that are akin to those between resonant\nabsorption and phase mixing in Figure 5a. Wave leakage is the favored model for strong and weak damping regimes,\nwhile positive evidence for phase mixing is obtained in the intermediate damping regime, with r2(2\u00007).\nIt is customary to address the comparison between damping mechanism using data on periods and damping times\nin a scatter plot where these data are then \ftted to a straight line, according to the scaling law hypothesis (Ofman\n& Aschwanden 2002). We proceed di\u000berently and next evaluate the plausibility between damping mechanisms by\ncalculating the Bayes factor distributions in such a two-dimensional plane with the two observables of interest, period\nand damping time. Hence d=fP;\u001cdg. To do so we consider the damping ratios resulting for di\u000berent combinations\nof\u001cdandPon a wide range of values and a \fxed uncertainty. Figure 6 shows the obtained Bayes factor distributions.10 Montes-Sol \u0013\u0010s & Arregui\n0 2 4 6 8 10\nr0.00.20.40.60.81.0p(r|M)(a)resonant\nwave leakage\nphase mixing\n0 2 4 6 8 10\nr0.00.20.40.60.81.0p(r|M)(b) =10%\n=25%\n=50%\n0 2 4 6 8 10\nr0.000.050.100.150.20p(r|M)(c)=10%\n=25%\n=50%\n0 2 4 6 8 10\nr0.00.10.20.30.40.50.60.70.8p(r|M)(d) =10%\n=25%\n=50%\nFigure 3. (a) Marginal likelihoods computed using Equation (5) for the three considered models as a function of the observable\ndamping ratio for a \fxed value of \u001b= 10%. For phase mixing, three \fxed values of period (150 ;500;and 1250sfrom right to left)\nhave been taken. (b)-(d): Marginal likelihoods for each mechanism separately for three di\u000berent values of \u001b= 10%;25%;50%.\nFor each mechanism the same color as in panel (a) is applied.\n0 2 4 6 8 10\nr0.00.20.40.60.81.0p(r|M)\nresonant\nphase mixing\nleakage\nFigure 4. A comparison between the marginal likelihoods obtained by direct numerical integration (continuous lines) and the\nMonte Carlo integration (\flled circles) for the three damping mechanisms as a function of the observable damping ratio with\nuncertainty of 10%. For phase mixing, P= 150 s is \fxed.\nThe general conclusion is that the evidence for any of the considered models against an alternative depends on the\nparticular combination of observed periods and damping times.\nIn particular, for resonant absorption against phase mixing, Figure 6a shows that at the upper-left corner of this\npanel, corresponding to large values of the damping ratio, we obtain strong (purple) and very strong (gray) evidence inDamping mechanisms for transverse coronal waves 11\n0 2 4 6 8 10\nr20\n10\n010202ln(BFkj)\nBF01BF10(a)\n0 2 4 6 8 10\nr15\n10\n5\n0510152ln(BFkj)BF20\nBF02(b)\n0 2 4 6 8 10\nr20\n10\n010202ln(BFkj) BF12\nBF21(c)\nFigure 5. Bayes factors for the one-to-one comparison between resonant absorption, phase mixing, and wave leakage mech-\nanisms, here represented with the subscripts 0, 1, and 2 respectively, in the range r2[0:01;10] with uncertainty of 10%. For\nphase mixing, P=150 s is \fxed.\n200 400 600 800 1000 1200\nP (s)5001000150020002500300035004000d(s)\n(a) 2lnB01\n10\n6\n2\n2610\n200 400 600 800 1000 1200\nP (s)5001000150020002500300035004000d(s)\n(b) 2lnB02\n10\n6\n2\n2610\n200 400 600 800 1000 1200\nP (s)5001000150020002500300035004000d(s)\n(c) 2lnB12\n10\n6\n2\n2610\nFigure 6. Bayes factors in the one-to-one comparison between resonant absorption, phase mixing, and wave leakage mechanisms\nas a function of the observables period and damping time with uncertainties of 10% for each. The dashed lines indicate \u001cd=P.\nThe di\u000berent levels of evidence of Table 1 are indicated in the color bars. NW (yellow), PE (green/red), SE (blue/purple), VSE\n(white/gray).\nfavor of resonant absorption. The lower-right corner corresponding to low damping ratios shows very strong evidence\n(white) in favor of the phase mixing model. In the remaining regions di\u000berent colored bands denote positive (pink for\nresonant and green for phase mixing) or insigni\fcant evidence (yellow), depending on the particular combination of\nthe two observables.\nIn the comparison between resonant absorption and wave leakage, Figure 6b, most of the ( \u001cd;P)- plane is colored in\nyellow resulting in a lack of evidence for a particular model. Grey and purple regions (very strong and strong evidence)\nindicate the dominance of the resonant mechanism for the lowest values of the damping ratio and positive evidence\n(green for wave leakage and pink for resonant) is found in very narrow regions of the plane of observables.\nFinally, Figure 6c shows the results from the comparison between phase mixing and wave leakage. The white\n(very strong evidence) and the blue (strong evidence) regions are located at period and damping time combinations\ncorresponding to large damping ratios and point out to the dominance of the wave leakage model over phase mixing.\nFor low damping ratios, we \fnd very strong evidence (gray) in favor of phase mixing. The remaining combinations\nlead to positive (pink for phase mixing and green for wave leakage) or inconclusive evidence (yellow) depending on the\nobservables.\nIn the three panels of Figure 6, focusing in observations with values of the damping ratio larger than 1, which are\nlocated in the upper-left side of the dashed line, resonant absorption and wave leakage are more plausible than phase\nmixing but no distinction can be made to favor any of them. Phase mixing seems to be the most plausible mechanism\nin the intermediate region of the observable ( \u001cd;P)-plane (green color in Figure 6a and pink color in Figure 6c).\nUntil now, model comparison is pursued by considering hypothetical values for observed periods, damping times,\nand their measurement errors. The full potential of the method here presented is discernible when applied to real\nevents of damped transverse loop oscillations. For this reason, we consider a selection of 89 loop oscillation events for12 Montes-Sol \u0013\u0010s & Arregui\n1 2 3 4 5\nr4\n3\n2\n1\n01232lnB01\n(a)NWM PE>0 PE<0\n1 2 3 4 5\nr0.6\n0.4\n0.2\n0.00.20.40.62lnB02\n(b)NWM\n1 2 3 4 5\nr2\n1\n01232 lnB12\n(c)NWM PE>0 PE<0\nFigure 7. Representation of the Bayes factors computed for the 89 events selected from Verwichte et al. (2013) and Goddard\net al. (2016). The di\u000berent panels correspond to the three one-to-one comparisons between resonant absorption, phase mixing,\nand wave leakage, here represented with the subscripts 0, 1, and 2 respectively.\nwhich periods and damping times are listed in the databases presented by Verwichte et al. (2013) and Goddard et al.\n(2016), discarding events for which errors were not reported. The considered events and their oscillation properties\nare presented in Table 3 of Appendix A. They are sorted in increasing value for their corresponding damping ratio\nin order to locate the events easily according to results. We further include for each event the inferred parameters,\ncomputed following the methods described in Section 4.1.\nFigure 7 shows scatter plots that display Bayes factor and damping ratio values for the selected 89 events for each\none-to-one model comparison between our three damping mechanisms. The colors indicate the level of evidence, based\non the magnitude of the corresponding Bayes factor. The Bayes factor values are also listed in Table 3. The conspicuous\nresult is that the blue color, which corresponds to absence of evidence for any damping mechanism, dominates in all\nthree panels.\nFor the comparison between resonant absorption and phase mixing (left panel), in approximately 78% of the events\nwe have not su\u000ecient evidence to favor one model or the other (NWM). Some events (colored in red and yellow) show\npositive evidence for either resonant absorption or for phase mixing. The evidence is positive for resonant absorption\nin 8% of the events, with strong damping (low values of r) and positive evidence for phase mixing in about 14% of\nthe events, with damping ratio values in between 3 and 6. In the middle panel the analysis for resonant absorption\nvs. wave leakage is shown. In all events the Bayes factor and therefore the evidence is not strong enough to support\nany of the two mechanisms. Finally, the right panel shows the comparison between phase mixing and wave leakage.\nFor 79% of the events the evidence is inconclusive, with the events distributed among the full range of damping ratios.\nOnly for 3% of the events we obtain positive evidence in favor of wave leakage, for loop oscillations with very strong\ndamping, and for another 18% positive evidence in favor of phase mixing.\n5.SUMMARY AND CONCLUSIONS\nIn the last years, high resolution observations have enabled us to better characterize the damping of transverse\noscillations in coronal loops. More accurate estimates of periods and damping times are now obtained and the size\nof the databases has increased. However, the discussion about the mechanisms involved in the quick damping of the\nobserved oscillations remains latent. The possible mechanisms were proposed a few decades ago but no self-consistent\nmethod exists yet to assess which one better accounts for the observed damping phenomenon. The widespread approach\nhas been to resort to the use of scaling laws that would apparently be an intrinsic property of each considered theoretical\nmechanism. This has proven to be of little use since, physical properties of coronal loops being unknown and probably\ndi\u000berent, mechanisms such as resonant absorption are known to predict not only one but many scalings with di\u000berent\npower indexes.\nIn this paper, we have presented a method to compute relative probabilities between alternative damping mechanisms\nfor transverse coronal loop oscillations. We considered three mechanisms among those that have been proposed:\nresonant absorption, phase mixing of Alfv\u0013 en waves, and wave leakage. They all are in principle likely to occur because\nof the highly inhomogeneous nature of the corona across the oscillating waveguides. Their direct applicability to\nthe damping of transverse loop oscillations is di\u000berent, with phase-mixing presenting serious limitations. Together\nwith model comparison, the coronal loop physical parameters that characterize each mechanism have been inferred.Damping mechanisms for transverse coronal waves 13\nThe analysis was carried out using Bayesian analysis tools. Bayesian inference enabled us to obtain probability\ndensity functions for the parameters of interest, with correct propagation of uncertainty from observables to physical\nparameters. The computation of marginal likelihoods informs us on the likelihood of each mechanism to generate the\nobserved data. Finally, Bayes factors quantify the level of evidence for one mechanism against another alternative.\nOur inference results indicate that physical parameters such as the density contrast, the transverse density inhomo-\ngeneity length-scales, and the aspect ratio of coronal loops can be properly inferred. Considering typically observed\ndamping ratio values, the obtained distributions peak at reasonable values of the unknown parameters. Our model\ncomparison results indicate that, as a general rule, a single damping mechanism cannot explain the observed damping\nof coronal loop oscillations. However, the method enables us to assign a level of evidence to each considered damping\nmodel. Considering hypothetical observed damping ratios over a range of plausible values, we found that resonant\nabsorption and wave leakage o\u000ber the most probable explanation in strong damping regimes, while phase mixing is\nthe best candidate for moderate/weak damping. The method was then applied to a large selection of loop oscillation\nevents compiled in databases that provide accurate measurements of period and damping time uncertainties. A fre-\nquentist analysis of the obtained Bayes factors indicates that only in very few cases the evidence is large enough to\nsupport a particular damping mechanism. For the rest of the cases, nothing can be stated. Since the uncertainty on\nthe measured times-scales is essential when translated into levels of evidence, the future increase in the precision of\ndata becomes relevant in order to determine the associated damping processes.\nThe results here presented have not given a de\fnitive answer to the question of what mechanism is responsible for\nthe quick damping of coronal loop oscillations, but the method makes use of all the available information - models,\nobserved data with their uncertainty, and prior information - in a consistent manner. The method is applicable to\nadditional damping models and formulas. Approximate answers have been obtained by considering the pertinent\nquestion of quantifying the evidence for each model in view of data, rather than obtaining exact answers to the more\nmisleading question of how well the data \ft to questionable theoretical power laws.\nWe acknowledge \fnancial support from the Spanish Ministry of Economy and Competitiveness (MINECO) through\nprojects AYA2014-55456-P (Bayesian Analysis of the Solar Corona), AYA2014-60476-P (Solar Magnetometry in the\nEra of Large Telescopes), and from FEDER funds. M.M-S. acknowledges \fnancial support through a Severo Ochoa\nFPI Fellowship under the project SEV-2011-0187-03. I.A. acknowledges \fnancial support through a Ram\u0013 on y Cajal\nFellowship.\nSoftware: emcee (Foreman-Mackey et al. 2013)\nAPPENDIX\nA.ANALYSED CORONAL LOOP DATA14 Montes-Sol \u0013\u0010s & ArreguiTable 3 . Transverse loop oscillations data, inference results and Bayes factors.\nRef Ev.ID Lp.ID P \u001cd r l/R \u0010 wR=L 2lnBF 01 2lnBF 02 2lnBF 12\n(s) (s) (Mm)\n1 24 \u0001 \u0001 \u0001 895\u00062 521 \u00068 0:6\u00060:0 1:5+0:3\n\u00000:16:5+2:4\n\u00002:21:6+0:1\n\u00000:00:27+0:00\n\u00000:003:45\u00030:48 \u00002:97\u0003\n2 10 1 687 :6\u000610:2 481:2\u000665:4 0:7\u00060:1 1:4+0:3\n\u00000:26:1+2:6\n\u00002:31:9+0:4\n\u00000:30:25+0:02\n\u00000:022:80\u00030:56 \u00002:24\u0003\n2 56 7 849 :6\u000633:0 818:4\u0006235:8 1:0\u00060:3 1:2+0:5\n\u00000:35:8+2:8\n\u00002:63:4+1:5\n\u00001:30:22+0:04\n\u00000:031:43 0:56 \u00000:87\n2 48 3 964 :8\u000612:6 945:6\u0006185:4 1:0\u00060:2 1:1+0:4\n\u00000:35:5+3:0\n\u00002:53:6+1:1\n\u00001:00:21+0:03\n\u00000:021:21 0:56 \u00000:65\n2 9 1 308 :4\u000610:2 305:4\u000658:8 1:0\u00060:2 1:1+0:4\n\u00000:35:5+3:0\n\u00002:52:2+0:6\n\u00000:60:21+0:03\n\u00000:022:31\u00030:56 \u00001:75\n1 40 \u0001 \u0001 \u0001 302\u000614 306 \u000643 1:0\u00060:1 1:0+0:4\n\u00000:25:3+3:1\n\u00002:52:2+0:5\n\u00000:40:21+0:02\n\u00000:022:18\u00030:56 \u00001:62\n1 44 \u0001 \u0001 \u0001 1170\u00066 1218 \u000648 1:0\u00060:0 0:9+0:5\n\u00000:15:2+3:2\n\u00002:54:2+0:3\n\u00000:20:20+0:00\n\u00000:000:62 0:55 \u00000:07\n1 25 \u0001 \u0001 \u0001 452\u00061 473 \u00066 1:0\u00060:0 0:9+0:5\n\u00000:15:2+3:2\n\u00002:52:8+0:0\n\u00000:10:20+0:00\n\u00000:001:57 0:55 \u00001:02\n2 43 1 428 :4\u00064:2 451:8\u000687:0 1:1\u00060:2 1:0+0:5\n\u00000:25:4+3:1\n\u00002:52:8+0:8\n\u00000:70:21+0:03\n\u00000:021:72 0:55 \u00001:18\n1 32 \u0001 \u0001 \u0001 216\u000627 230 \u000623 1:1\u00060:2 1:0+0:4\n\u00000:25:3+3:2\n\u00002:52:0+0:5\n\u00000:40:20+0:02\n\u00000:022:32\u00030:55 \u00001:78\n1 29 \u0001 \u0001 \u0001 225\u000640 240 \u000645 1:1\u00060:3 1:1+0:5\n\u00000:35:6+3:0\n\u00002:62:1+0:8\n\u00000:70:21+0:04\n\u00000:022:40\u00030:54 \u00001:85\n2 3 2 217 :2\u00064:8 247:2\u000628:2 1:1\u00060:1 0:9+0:4\n\u00000:25:1+3:3\n\u00002:52:2+0:4\n\u00000:30:19+0:01\n\u00000:011:97 0:52 \u00001:45\n2 16 2 141 :0\u00064:2 161:4\u000638:4 1:1\u00060:3 1:0+0:5\n\u00000:35:4+3:1\n\u00002:61:9+0:6\n\u00000:60:20+0:03\n\u00000:022:57\u00030:53 \u00002:04\u0003\n2 49 5 481 :8\u000610:8 562:2\u000673:2 1:2\u00060:2 0:9+0:5\n\u00000:25:1+3:3\n\u00002:53:3+0:6\n\u00000:60:19+0:02\n\u00000:011:09 0:51 \u00000:57\n1 31 \u0001 \u0001 \u0001 213\u00069 251 \u000636 1:2\u00060:2 0:9+0:5\n\u00000:25:1+3:3\n\u00002:52:3+0:5\n\u00000:50:19+0:02\n\u00000:011:88 0:51 \u00001:37\n1 41 \u0001 \u0001 \u0001 565\u00064 666 \u000642 1:2\u00060:1 0:8+0:5\n\u00000:15:1+3:3\n\u00002:53:6+0:3\n\u00000:40:19+0:01\n\u00000:010:82 0:50 \u00000:31\n2 23 1 921 :6\u000624:0 1151:4\u000693:0 1:2\u00060:1 0:8+0:5\n\u00000:15:0+3:3\n\u00002:55:0+0:6\n\u00000:60:18+0:01\n\u00000:010:09 0:48 0:39\n2 18 2 571 :2\u00066:6 732:0\u0006208:2 1:3\u00060:4 1:0+0:5\n\u00000:35:4+3:1\n\u00002:64:3+1:9\n\u00001:60:19+0:04\n\u00000:030:76 0:50 \u00000:26\n1 34 \u0001 \u0001 \u0001 596\u000650 771 \u0006336 1:3\u00060:6 1:1+0:6\n\u00000:45:7+2:9\n\u00002:84:9+2:9\n\u00002:40:20+0:06\n\u00000:040:71 0:51 \u00000:20\n2 44 3 417 :0\u00068:4 540:0\u0006180:0 1:3\u00060:4 1:0+0:6\n\u00000:45:5+3:1\n\u00002:73:8+1:9\n\u00001:60:20+0:05\n\u00000:031:08 0:50 \u00000:57\n2 9 2 537 :0\u00068:4 709:8\u0006285:6 1:3\u00060:5 1:1+0:6\n\u00000:45:6+3:0\n\u00002:84:7+2:6\n\u00002:20:20+0:05\n\u00000:040:76 0:51 \u00000:25\n2 40 3 331 :8\u00062:4 439:2\u000664:8 1:3\u00060:2 0:8+0:5\n\u00000:25:0+3:3\n\u00002:53:3+0:7\n\u00000:70:18+0:02\n\u00000:010:92 0:46 \u00000:46\n1 30 \u0001 \u0001 \u0001 215\u00065 293 \u000618 1:4\u00060:1 0:7+0:5\n\u00000:14:9+3:4\n\u00002:52:8+0:3\n\u00000:20:18+0:01\n\u00000:001:14 0:44 \u00000:70\n1 35 \u0001 \u0001 \u0001 212\u000620 298 \u000630 1:4\u00060:2 0:8+0:5\n\u00000:24:9+3:4\n\u00002:52:9+0:6\n\u00000:60:18+0:01\n\u00000:011:09 0:43 \u00000:66\n1 33 \u0001 \u0001 \u0001 520\u00065 735 \u000653 1:4\u00060:1 0:7+0:5\n\u00000:14:8+3:4\n\u00002:54:5+0:5\n\u00000:50:17+0:01\n\u00000:010:09 0:42 0:32\n2 48 1 916 :8\u00069:6 1318:8\u0006936:0 1:4\u00061:0 1:1+0:6\n\u00000:55:8+2:9\n\u00002:88:2+6:0\n\u00004:80:20+0:06\n\u00000:05\u00000:17 0:50 0:67\n1 46 \u0001 \u0001 \u0001 150\u00065 216 \u000660 1:4\u00060:4 0:9+0:6\n\u00000:35:2+3:2\n\u00002:62:7+1:1\n\u00000:90:18+0:04\n\u00000:021:62 0:46 \u00001:16\n2 19 2 676 :2\u00067:2 993:0\u000686:4 1:5\u00060:1 0:7+0:5\n\u00000:14:8+3:5\n\u00002:55:4+0:8\n\u00000:70:17+0:01\n\u00000:01\u00000:33 0:40 0:73\n2 49 4 627 :0\u000610:2 922:8\u0006154:8 1:5\u00060:2 0:8+0:5\n\u00000:24:9+3:4\n\u00002:55:3+1:4\n\u00001:20:17+0:02\n\u00000:01\u00000:15 0:42 0:57\nTable 3 continued on next pageDamping mechanisms for transverse coronal waves 15Table 3 (continued)\nRef Ev.ID Lp.ID P \u001cd r l/R \u0010 wR=L 2lnBF 01 2lnBF 02 2lnBF 12\n(s) (s) (Mm)\n2 44 2 586 :8\u000611:4 877:2\u0006297:6 1:5\u00060:5 0:9+0:6\n\u00000:45:3+3:2\n\u00002:75:6+2:8\n\u00002:30:19+0:05\n\u00000:030:19 0:46 0:28\n1 27 \u0001 \u0001 \u0001 2418\u00065 3660 \u000680 1:5\u00060:0 0:7+0:5\n\u00000:14:8+3:5\n\u00002:510:6+0:4\n\u00000:30:17+0:00\n\u00000:00\u00001:78 0:38 2:15\u0003\n1 38 \u0001 \u0001 \u0001 115\u00062 1750 \u000630 1:5\u00060:3 0:7+0:5\n\u00000:24:8+3:4\n\u00002:52:5+0:6\n\u00000:60:17+0:02\n\u00000:021:40 0:40 \u00001:00\n2 24 1 1071 :6\u000618:0 1645:8\u0006255:6 1:5\u00060:2 0:7+0:5\n\u00000:24:8+3:5\n\u00002:57:4+1:8\n\u00001:60:17+0:02\n\u00000:01\u00000:90 0:39 1:29\n1 45 \u0001 \u0001 \u0001 623\u00064 960 \u000660 1:5\u00060:1 0:7+0:5\n\u00000:14:8+3:5\n\u00002:55:6+0:5\n\u00000:50:16+0:01\n\u00000:00\u00000:48 0:37 0:86\n2 25 1 307 :8\u00066:6 480:0\u0006300:0 1:6\u00061:0 1:1+0:6\n\u00000:55:7+2:9\n\u00002:85:4+4:1\n\u00002:90:20+0:06\n\u00000:050:64 0:48 \u00000:16\n2 1 1 205 :2\u00063:6 320:4\u000667:2 1:6\u00060:3 0:8+0:5\n\u00000:24:9+3:4\n\u00002:53:4+1:1\n\u00001:00:17+0:02\n\u00000:020:79 0:40 \u00000:39\n2 26 1 717 :0\u00067:8 1122:6\u0006270:0 1:6\u00060:4 0:8+0:5\n\u00000:25:0+3:3\n\u00002:66:4+2:3\n\u00002:10:17+0:03\n\u00000:02\u00000:39 0:41 0:81\n1 26 \u0001 \u0001 \u0001 630\u000630 1000 \u0006300 1:6\u00060:5 0:9+0:6\n\u00000:35:2+3:3\n\u00002:66:3+2:8\n\u00002:40:18+0:05\n\u00000:03\u00000:18 0:43 0:61\n2 56 2 712 :8\u00067:8 1177:2\u0006177:6 1:7\u00060:2 0:7+0:5\n\u00000:24:8+3:5\n\u00002:56:7+1:6\n\u00001:40:16+0:02\n\u00000:01\u00000:83 0:35 1:18\n2 34 1 597 :0\u000616:2 1002:0\u000661:8 1:7\u00060:1 0:6+0:5\n\u00000:14:7+3:5\n\u00002:56:2+0:7\n\u00000:60:16+0:01\n\u00000:00\u00000:83 0:32 1:15\n2 48 2 945 :6\u00067:2 1598:4\u0006130:2 1:7\u00060:1 0:6+0:5\n\u00000:14:7+3:5\n\u00002:57:8+1:0\n\u00000:90:16+0:01\n\u00000:01\u00001:31 0:32 1:63\n1 50 \u0001 \u0001 \u0001 491\u000618 834 \u00066 1:7\u00060:1 0:6+0:5\n\u00000:14:6+3:6\n\u00002:55:7+0:4\n\u00000:30:16+0:00\n\u00000:00\u00000:70 0:31 1:02\n2 24 3 1227 :6\u000634:8 2100:6\u0006386:4 1:7\u00060:3 0:7+0:5\n\u00000:24:8+3:5\n\u00002:59:3+2:6\n\u00002:40:16+0:02\n\u00000:02\u00001:46 0:35 1:81\n1 48 \u0001 \u0001 \u0001 273\u000654 468 \u000636 1:7\u00060:4 0:7+0:5\n\u00000:24:8+3:5\n\u00002:54:5+1:4\n\u00001:30:16+0:02\n\u00000:020:09 0:36 0:26\n1 36 \u0001 \u0001 \u0001 256\u000622 444 \u0006105 1:7\u00060:4 0:7+0:6\n\u00000:24:9+3:4\n\u00002:64:5+1:7\n\u00001:50:16+0:03\n\u00000:020:22 0:37 0:16\n2 56 5 810 :0\u00069:6 1450:2\u0006307:8 1:8\u00060:4 0:7+0:5\n\u00000:24:8+3:5\n\u00002:58:1+2:6\n\u00002:30:16+0:02\n\u00000:02\u00001:19 0:34 1:52\n2 43 3 501 :0\u00064:8 902:4\u0006108:6 1:8\u00060:2 0:6+0:5\n\u00000:14:6+3:6\n\u00002:56:4+1:2\n\u00001:10:15+0:01\n\u00000:01\u00000:92 0:29 1:22\n2 31 2 575 :4\u00065:4 1054:2\u0006141:0 1:8\u00060:2 0:6+0:5\n\u00000:14:6+3:6\n\u00002:57:0+1:5\n\u00001:30:15+0:01\n\u00000:01\u00001:12 0:29 1:41\n1 42 \u0001 \u0001 \u0001 222\u000618 420 \u0006360 1:9\u00061:6 1:1+0:6\n\u00000:65:7+2:9\n\u00002:97:2+5:9\n\u00004:30:19+0:07\n\u00000:060:21 0:44 0:23\n1 43 \u0001 \u0001 \u0001 474\u000612 900 \u0006120 1:9\u00060:3 0:6+0:5\n\u00000:14:6+3:6\n\u00002:56:8+1:4\n\u00001:40:15+0:01\n\u00000:01\u00001:09 0:27 1:36\n2 40 8 259 :8\u00064:8 540:6\u0006129:6 2:1\u00060:5 0:6+0:5\n\u00000:24:7+3:6\n\u00002:55:9+2:2\n\u00001:90:15+0:03\n\u00000:02\u00000:65 0:27 0:92\n2 29 1 222 :6\u00063:0 469:8\u000637:2 2:1\u00060:2 0:5+0:5\n\u00000:14:4+3:7\n\u00002:55:4+0:6\n\u00000:70:14+0:01\n\u00000:01\u00000:89 0:19 1:07\n2 40 9 370 :8\u00063:0 789:0\u0006159:6 2:1\u00060:4 0:6+0:5\n\u00000:24:6+3:6\n\u00002:57:2+2:2\n\u00002:00:14+0:02\n\u00000:01\u00001:22 0:23 1:45\n2 4 2 208 :2\u00061:8 446:4\u000660:0 2:1\u00060:3 0:5+0:5\n\u00000:14:5+3:7\n\u00002:55:4+1:1\n\u00001:10:14+0:01\n\u00000:01\u00000:83 0:19 1:02\n1 49 \u0001 \u0001 \u0001 282\u00066 606 \u0006186 2:1\u00060:7 0:7+0:6\n\u00000:34:9+3:4\n\u00002:76:6+3:0\n\u00002:50:15+0:04\n\u00000:02\u00000:65 0:30 0:95\n2 44 1 433 :8\u00063:6 945:0\u0006185:4 2:2\u00060:4 0:6+0:5\n\u00000:24:5+3:6\n\u00002:58:0+2:4\n\u00002:20:14+0:02\n\u00000:01\u00001:50 0:22 1:72\n2 56 1 544 :2\u00068:4 1242:6\u0006282:6 2:3\u00060:5 0:6+0:5\n\u00000:24:6+3:6\n\u00002:59:7+3:3\n\u00002:90:14+0:02\n\u00000:02\u00001:85 0:21 2:06\u0003\n1 37 \u0001 \u0001 \u0001 135\u00069 311 \u000685 2:3\u00060:6 0:6+0:6\n\u00000:24:7+3:5\n\u00002:65:0+2:2\n\u00001:80:15+0:04\n\u00000:02\u00000:30 0:25 0:55\n1 39 \u0001 \u0001 \u0001 103\u00068 242 \u0006114 2:3\u00061:1 0:9+0:7\n\u00000:45:3+3:2\n\u00002:85:1+3:3\n\u00002:50:17+0:07\n\u00000:040:37 0:35 \u00000:02\n2 40 7 343 :2\u00063:6 850:2\u0006163:8 2:5\u00060:5 0:5+0:5\n\u00000:14:4+3:7\n\u00002:58:6+2:4\n\u00002:30:13+0:02\n\u00000:01\u00001:87 0:13 2:00\u0003\nTable 3 continued on next page16 Montes-Sol \u0013\u0010s & ArreguiTable 3 (continued)\nRef Ev.ID Lp.ID P \u001cd r l/R \u0010 wR=L 2lnBF 01 2lnBF 02 2lnBF 12\n(s) (s) (Mm)\n1 51 \u0001 \u0001 \u0001 348\u00067 906 \u0006288 2:6\u00060:8 0:6+0:7\n\u00000:34:8+3:5\n\u00002:79:7+4:4\n\u00003:80:14+0:05\n\u00000:02\u00001:61 0:22 1:83\n2 1 2 246 :6\u00063:0 645:6\u0006167:4 2:6\u00060:7 0:5+0:6\n\u00000:24:5+3:6\n\u00002:58:1+3:1\n\u00002:70:13+0:03\n\u00000:02\u00001:57 0:16 1:72\n2 43 2 216 :0\u00061:8 566:4\u000655:2 2:6\u00060:3 0:4+0:5\n\u00000:14:3+3:8\n\u00002:47:3+1:1\n\u00001:00:13+0:01\n\u00000:00\u00001:85 0:05 1:90\n2 8 1 224 :4\u00064:2 600:0\u000660:0 2:7\u00060:3 0:4+0:5\n\u00000:14:3+3:8\n\u00002:47:7+1:2\n\u00001:10:13+0:01\n\u00000:01\u00001:97 0:04 2:01\u0003\n1 52 \u0001 \u0001 \u0001 340\u00063 930 \u0006144 2:7\u00060:4 0:4+0:5\n\u00000:14:3+3:8\n\u00002:49:8+2:4\n\u00002:10:13+0:01\n\u00000:01\u00002:41\u00030:05 2:46\u0003\n1 13 \u0001 \u0001 \u0001 448\u000618 1260 \u0006500 2:8\u00061:1 0:7+0:7\n\u00000:45:0+3:3\n\u00002:811:6+5:2\n\u00005:20:15+0:07\n\u00000:03\u00001:58 0:25 1:83\n1 47 \u0001 \u0001 \u0001 122\u00066 348 \u0006360 2:9\u00063:0 1:0+0:7\n\u00000:65:6+3:0\n\u00002:99:1+6:7\n\u00005:60:18+0:08\n\u00000:07\u00000:21 0:36 0:57\n1 16 \u0001 \u0001 \u0001 358\u000630 1030 \u0006570 2:9\u00061:6 0:9+0:7\n\u00000:55:3+3:2\n\u00002:910:9+5:8\n\u00005:80:16+0:08\n\u00000:05\u00000:99 0:32 1:31\n2 38 2 312 :0\u00064:8 913:8\u0006330:0 2:9\u00061:1 0:7+0:7\n\u00000:34:8+3:5\n\u00002:710:8+5:0\n\u00004:60:14+0:06\n\u00000:03\u00001:72 0:20 1:92\n1 17 \u0001 \u0001 \u0001 326\u000645 980 \u0006400 3:0\u00061:3 0:7+0:8\n\u00000:45:1+3:3\n\u00002:811:2+5:3\n\u00005:30:15+0:08\n\u00000:03\u00001:46 0:25 1:71\n2 20 1 321 :6\u000613:8 971:4\u0006460:2 3:0\u00061:4 0:8+0:7\n\u00000:45:2+3:3\n\u00002:811:1+5:5\n\u00005:50:15+0:08\n\u00000:04\u00001:28 0:27 1:55\n2 43 5 270 :0\u00061:2 840:0\u0006120:0 3:1\u00060:4 0:4+0:5\n\u00000:14:2+3:8\n\u00002:410:6+2:3\n\u00002:20:12+0:01\n\u00000:01\u00002:80\u0003\u00000:05 2:76\u0003\n2 4 1 137 :4\u00061:8 430:8\u000690:0 3:1\u00060:7 0:4+0:5\n\u00000:14:3+3:8\n\u00002:57:8+2:4\n\u00002:30:12+0:02\n\u00000:01\u00002:00 \u00000:01 2:00\n2 45 1 148 :8\u00062:4 469:2\u000699:6 3:2\u00060:7 0:4+0:5\n\u00000:14:3+3:8\n\u00002:58:2+2:5\n\u00002:40:12+0:02\n\u00000:01\u00002:10\u0003\u00000:01 2:09\u0003\n2 31 1 460 :2\u00062:4 1453:2\u0006121:2 3:2\u00060:3 0:4+0:5\n\u00000:14:1+3:9\n\u00002:414:0+1:8\n\u00001:80:12+0:00\n\u00000:00\u00003:49\u0003\u00000:08 3:41\u0003\n2 11 3 156 :0\u00063:0 530:4\u000690:0 3:4\u00060:6 0:4+0:5\n\u00000:14:1+3:9\n\u00002:49:3+2:4\n\u00002:30:11+0:01\n\u00000:01\u00002:61\u0003\u00000:09 2:52\u0003\n1 15 \u0001 \u0001 \u0001 382\u000612 1330 \u0006528 3:5\u00061:4 0:7+0:8\n\u00000:34:9+3:4\n\u00002:812:8+4:8\n\u00005:70:13+0:07\n\u00000:03\u00001:81 0:16 1:97\n2 3 1 147 :6\u00061:8 528:0\u0006108:0 3:6\u00060:7 0:4+0:5\n\u00000:14:1+3:9\n\u00002:49:8+3:0\n\u00002:70:11+0:02\n\u00000:01\u00002:70\u0003\u00000:10 2:60\u0003\n2 32 1 256 :8\u00061:2 933:0\u000673:2 3:6\u00060:3 0:3+0:5\n\u00000:14:0+3:9\n\u00002:412:8+1:6\n\u00001:50:11+0:00\n\u00000:00\u00003:57\u0003\u00000:18 3:39\u0003\n1 12 \u0001 \u0001 \u0001 249\u000633 920 \u0006360 3:7\u00061:5 0:7+0:8\n\u00000:44:9+3:5\n\u00002:812:2+5:1\n\u00005:60:13+0:08\n\u00000:03\u00001:81 0:15 1:96\n1 18 \u0001 \u0001 \u0001 357\u000689 1320 \u0006720 3:7\u00062:2 0:9+0:7\n\u00000:55:3+3:2\n\u00002:911:9+5:6\n\u00006:40:16+0:09\n\u00000:05\u00001:01 0:26 1:27\n2 54 5 288 :0\u00066:0 1183:2\u0006193:8 4:1\u00060:7 0:3+0:5\n\u00000:14:0+3:9\n\u00002:415:6+2:8\n\u00003:30:10+0:01\n\u00000:01\u00003:68\u0003\u00000:23 3:45\u0003\n2 7 1 101 :4\u00061:2 433:8\u000678:0 4:3\u00060:8 0:3+0:5\n\u00000:14:0+3:9\n\u00002:410:6+2:8\n\u00002:60:10+0:01\n\u00000:01\u00003:24\u0003\u00000:25 2:99\u0003\n2 40 2 336 :6\u00061:8 1489:8\u0006204:6 4:4\u00060:6 0:3+0:5\n\u00000:13:9+4:0\n\u00002:317:4+2:0\n\u00002:70:10+0:01\n\u00000:01\u00003:35\u0003\u00000:30 3:06\u0003\n1 14 \u0001 \u0001 \u0001 392\u000631 1830 \u0006790 4:7\u00062:0 0:6+0:8\n\u00000:44:9+3:5\n\u00002:813:5+4:6\n\u00006:30:13+0:09\n\u00000:03\u00001:45 0:09 1:53\n2 17 1 124 :2\u00062:4 599:4\u0006275:4 4:8\u00062:2 0:7+0:8\n\u00000:45:0+3:4\n\u00002:912:2+5:2\n\u00005:90:13+0:09\n\u00000:04\u00001:78 0:10 1:88\n1 22 \u0001 \u0001 \u0001 436\u00064:5 2129 \u0006280 4:9\u00060:6 0:2+0:5\n\u00000:13:9+4:0\n\u00002:318:4+1:2\n\u00002:30:09+0:01\n\u00000:00\u00000:88 \u00000:37 0:51\n1 23 \u0001 \u0001 \u0001 243\u00066:4 1200 4 :9\u00060:1 0:2+0:5\n\u00000:03:8+4:1\n\u00002:319:4+0:5\n\u00000:60:09+0:00\n\u00000:00\u00004:20\u0003\u00000:38 3:82\u0003\n2 32 2 202 :8\u00061:2 1146:6\u0006291:0 5:7\u00061:4 0:3+0:5\n\u00000:13:9+4:0\n\u00002:415:9+2:9\n\u00004:50:09+0:02\n\u00000:01\u00002:89\u0003\u00000:35 2:53\u0003\nTable 3 continued on next pageDamping mechanisms for transverse coronal waves 17Table 3 (continued)\nRef Ev.ID Lp.ID P \u001cd r l/R \u0010 wR=L 2lnBF 01 2lnBF 02 2lnBF 12\n(s) (s) (Mm)\nNote |Columns present the reference (Ref), the event ID (Ev.ID) and loop ID (Lp.ID) from references, the observed period (P), the\ndamping time ( \u001cd), damping rate (r), the parameter inferred median values of the three selected theoretical models ( \u0010, l/R, w, R/L) with\ntheir uncertainties, and Bayes factor. The symbol (\u0003) indicate Positive Evidence according to the corresponding Bayes factor.\nReferences |(1) Verwichte et al. (2013); (2) Goddard et al. (2016)18 Montes-Sol \u0013\u0010s & Arregui\nREFERENCES\nAn\fnogentov, S., Nistic\u0012 o, G., & Nakariakov, V. M. 2013,\nA&A, 560, A107\nAn\fnogentov, S. A., Nakariakov, V. M., & Nistic\u0012 o, G. 2015,\nA&A, 583, A136\nArregui, I., & Asensio Ramos, A. 2011, ApJ, 740, 44\n|. 2014, A&A, 565, A78\nArregui, I., Asensio Ramos, A., & D\u0013 \u0010az, A. J. 2013a, ApJL,\n765, L23\nArregui, I., Asensio Ramos, A., & Pascoe, D. J. 2013b,\nApJL, 769, L34\nArregui, I., Ballester, J. L., & Goossens, M. 2008, ApJL,\n676, L77\nArregui, I., Soler, R., & Asensio Ramos, A. 2015, ApJ, 811,\n104\nAschwanden, M. J. 2005, Physics of the Solar Corona. An\nIntroduction with Problems and Solutions (2nd edition)\n(Springer-Praxis)\nAschwanden, M. J., Fletcher, L., Schrijver, C. J., &\nAlexander, D. 1999, ApJ, 520, 880\nAsensio Ramos, A., & Arregui, I. 2013, A&A, 554, A7\nBayes, M., & Price, M. 1763, Royal Society of London\nPhilosophical Transactions Series I, 53, 370\nBerghmans, D., & de Bruyne, P. 1995, ApJ, 453, 495\nCally, P. S. 1986, SoPh, 103, 277\n|. 2003, SoPh, 217, 95\nDe Pontieu, B., Martens, P. C. H., & Hudson, H. S. 2001,\nApJ, 558, 859\nForeman-Mackey, D., Hogg, D. W., Lang, D., & Goodman,\nJ. 2013, PASP, 125, 306\nGoddard, C. R., Nistic\u0012 o, G., Nakariakov, V. M., &\nZimovets, I. V. 2016, A&A, 585, A137\nGoossens, M., Andries, J., & Arregui, I. 2006, Philosophical\nTransactions of the Royal Society of London Series A,\n364, 433\nGoossens, M., Andries, J., & Aschwanden, M. J. 2002,\nA&A, 394, L39\nGregory, P. C. 2005, Bayesian Logical Data Analysis for\nthe Physical Sciences: A Comparative Approach with\n`Mathematica' Support (Cambridge University Press)\nHeyvaerts, J., & Priest, E. R. 1983, A&A, 117, 220\nKass, R. E., & Raftery, A. E. 1995, JASA, 90, 773\nLaplace, P. 1774, l'Academie Royale des Sciences, 6, 621\nMorton, R. J., & McLaughlin, J. A. 2013, A&A, 553, L10Morton, R. J., Tomczyk, S., & Pinto, R. 2015, Nature\nCommunications, 6, 7813\nNakariakov, V. M., Ofman, L., DeLuca, E. E., Roberts, B.,\n& Davila, J. M. 1999, Science, 285, 862\nNakariakov, V. M., & Verwichte, E. 2005, Living Reviews\nin Solar Physics, 2, 3\nOfman, L. 2002, ApJL, 568, L135\nOfman, L., & Aschwanden, M. J. 2002, ApJL, 576, L153\nOkamoto, T. J., Antolin, P., De Pontieu, B., et al. 2015,\nApJ, 809, 71\nOkamoto, T. J., & De Pontieu, B. 2011, ApJL, 736, L24\nRial, S., Arregui, I., Terradas, J., Oliver, R., & Ballester,\nJ. L. 2013, ApJ, 763, 16\nRobert, C. P., & Casella, G. 2004, Monte Carlo Integration\n(New York, NY: Springer New York), 79{122\nRoberts, B. 2000, SoPh, 193, 139\nRuderman, M. S. 2005, in ESA Special Publication, Vol.\n600, The Dynamic Sun: Challenges for Theory and\nObservations, 96.1\nRuderman, M. S., & Roberts, B. 2002, ApJ, 577, 475\nSoler, R., Oliver, R., Ballester, J. L., & Goossens, M. 2009,\nApJL, 695, L166\nSpruit, H. 1982, Solar Phys., 75, 3\nTerradas, J., Andries, J., & Goossens, M. 2007, SoPh, 246,\n231\nTerradas, J., Goossens, M., & Verth, G. 2010, A&A, 524,\nA23\nTerradas, J., Oliver, R., & Ballester, J. L. 2006, ApJL, 650,\nL91\nThurgood, J. O., Morton, R. J., & McLaughlin, J. A. 2014,\nApJL, 790, L2\nTomczyk, S., McIntosh, S. W., Keil, S. L., et al. 2007,\nScience, 317, 1192\nTrotta, R. 2008, Contemporary Physics, 49, 71\nVerth, G., Terradas, J., & Goossens, M. 2010, ApJL, 718,\nL102\nVerwichte, E., Van Doorsselaere, T., White, R. S., &\nAntolin, P. 2013, A&A, 552, A138\nvon Toussaint, U. 2011, Reviews of Modern Physics, 83, 943\nWang, T., Ofman, L., Davila, J. M., & Su, Y. 2012, ApJL,\n751, L27\nWhite, R. S., & Verwichte, E. 2012, A&A, 537, A49" }, { "title": "0805.1224v2.The_Lorentz_group_and_its_finite_field_analogues__local_isomorphism_and_approximation.pdf", "content": "arXiv:0805.1224v2 [math-ph] 4 Aug 2008The Lorentz group and its finite field analogues:\nlocal isomorphism and approximation\nStephan Foldes\nTampere University of Technology\nPL 553, 33101 Tampere, Finland\nsf@tut.fi\n4 August 2008\nAbstract\nFinite Lorentz groups acting on 4-dimensional vector space s coordinatized\nby finite fields with a prime number of elements are represente d as homomor-\nphic images of countable, rational subgroups of the Lorentz group acting on\nreal 4-dimensional space-time. Bounded subsets of the real Lorentz group are\nretractable with arbitraryprecisionto finitesubsets ofsu ch rational subgroups.\nThese finite retracts correspond, via local isomorphisms, t o well-behaved sub-\nsets of Lorentz groups over finite fields. This establishes a r elationship of\napproximation between the real Lorentz group and Lorentz gr oups over very\nlarge finite fields.\n1 Finite Minkowski spaces and statement of the theorem\nThe purpose of this paper is to establish a relationship of approximat ion\nbetween Lorentz transformations of Minkowski space R4and Lorentz trans-\nformations of 4-dimensional spaces over very large finite fields, in t he precise\nsense of Theorem 1 below. In relation to theoretical alternatives in particle\nphysics, Lorentztransformationsover finite fieldsseem to haveb een first con-\nsidered by Coish in [C], then by Shapiro [S], Ahmavaara [A1, A2], Yahia [Y],\nJoos [Jo], Beltrametti and Blasi [BB1], and Nambu [N]. Much of the algeb ra\nofLorentzgroupsoverfinitefieldshadalreadybeendevelopedbyD ickson[D],\nandinthe context ofphysical applications thistheory wasbrought uptodate\nby Beltrametti and Blasi [BB2]. A finite analogue of the Alexandrov-Z eeman\ncharacterization of the Lorentz group via optical causality was es tablished\n1by Blasi, Gallone, Zecca and Gorini [BGZG]. The approximation describe d\nin Theorem 1 relies on the one hand on modifying a relationship of appro xi-\nmation between positive real numbers and quadratic residues modu lo a large\nprime established by Kustaanheimo [JK, K1, K2], and modifying J¨ arne felt’s\napproximation of collinear ranges of points in Euclidean space by colline ar\nranges in finite geometries [J¨ a]. The proof of Theorem 1 also makes e ssen-\ntial use of rotation-boost decomposition, and of representing Lo rentz groups\nover finite fields as quotients of appropriate subgroups of the rea l Lorentz\ngroup. The algebraic background and terminology used in this theor em are\nas follows.\nWe consider fields Fin which −1 is not a square, and we shall assume\nthroughout that Fhas this property. (The main cases of interest are the\nfieldRof real numbers and the p-element finite fields Fpfor prime numbers\np≡7mod8, although many of the basic facts are also true for p-element\nfields with primes p≡3mod8.) For any such field Fthe 4-dimensional\nvector space F4is endowed with the ( F-valued)Minkowski norm\nµ(t,x,y,z) =t2−x2−y2−z2\nAproper Lorentz transformation over Fisalineartransformation Tfromthe\nvectorspace F4toitselfwhichpreserves theMinkowski norm( i.e.,µ(T(v)) =\nµ(v) for all v∈F4) and which has determinant 1. For any given field F,\nthese transformations form a group that we call the proper Lorentz group\noverF, denoted L+F. If the field Fis not specified, it is understood to be\nR.A Lorentz transformation over a field Fis said to be orthochronous if\nit maps (1 ,0,0,0) to a vector ( t,x,y,z) wheretis a non-zero square in F.\nForF=Rorthochronous Lorentz transformations form a subgroup of L+R\ncalled the orthochronous proper Lorentz group, denoted L↑\n+R.Over finite\nfields the composition of orthochronous transformations is not ne cessarily\northochronous.\nIfGandHare two (abstract) groups and A⊆G, Y⊆Hare arbitrary\nsets of group elements, a local isomorphism between AandYis a bijective\nmapσfromA∪AAtoY∪YY(whereAA={xy:x,y∈A}) such that\nσ[A] =Yand for all x,y∈A\nσ(xy) =σ(x)σ(y)\n2It follows that if 1 G∈Athenσ(1G) = 1Hand ifx,x−1∈Athenσ(x−1) =\nσ(x)−1.If such a local isomorphism exists, the sets AandYare said to be\nlocally isomorphic. (IfAandYare subgroups they are locally isomorphic\nif and only if they are isomorphic, but otherwise the sets AandYmay be\nlocally isomorphic even if the subgroups they generate are not isomo rphic.)\nBy thenorm/bardblT/bardblof any linear transformation T:R4−→R4with stan-\ndard matrix representation ( aij)1≤i,j≤4we mean (Σ a2\ni,j)1/2, i.e. the Euclidean\n(l2) norm of the matrix. By a retraction of L↑\n+Rto a subset A⊆ L↑\n+Rwe\nmean a map f:L↑\n+R−→Asuch that f(T) =Tfor allT∈A(i.e.,such\nthatf2=f).\nTheorem 1 For every ǫ >0andM >0the orthochronous proper\nLorentz group L↑\n+Rhas a retraction f:L↑\n+R−→Ato a finite subset\nA⊆ L↑\n+Rsatisfying\n/bardblT−f(T)/bardbl< ǫ\nfor all Lorentz transformations T∈ L↑\n+Rof norm not exceeding M,and such\nthat for some prime number p≡7mod8the subset Ais locally isomorphic to\na setYof orthochronous transformations in the proper Lorentz gro upL+Fp\nover the finite field Fp.\nThe proof is provided in Section 2, together with some elaborations o f the\nmathematical aspects that are needed for the sake of precision a nd clarity.\nThese are presented as a succession of intermediate propositions . A general-\nization to the extended Lorentz group is presented in Theorem 2 of Section 3,\nwith some additional properties of the local isomorphism linking the re spec-\ntive Lorentz groups over the real numbers and over finite fields. I n Section 4\nwe briefly comment on some of the particularities of finite analogues o f the\nLorentz group.\n2 Background developments and proof of Theorem 1\n2.1 Proper Lorentz groups over fields where -1 is not a square\nMuchofthebasictheoryoffiniteLorentzgroupsthatweneedisco ntained\nin or follows from the seminal work of Dickson [D] and the more recent study\nof Beltrametti and Blasi [BB2].\n3Lorentz groups are defined according to the active perspective a s groups\nof bijective transformations, and composition of transformation s is denoted\nsimply by juxtaposition T1T2, whereT2is the transformation first applied. If\nthe standard matrix representations of these transformations areT1andT2\nthen the matrix product T1T2represents the transformation T1T2.\nAproperLorentztransformation Tover afield Fwhere−1isnotasquare\nis called a space rotation (overF) if it fixes the first standard basis vector\n(1,0,0,0)∈F4. Equivalently, space rotations over Fare the linear trans-\nformations of F4with standard matrix representation of the block diagonal\nform /parenleftbigg1\nR3/parenrightbigg\nwhereR3is a 3-by-3 orthogonal matrix (i.e. R⊤\n3=R−1\n3) with determinant\n1. There are three space rotations which permute the standard b asis vectors,\nthey are given by\nR3=\n1\n1\n1\n,R3=\n1\n1\n1\n,R3=\n1\n1\n1\n\nThese particular space rotations are called rotations of standard space axes\nand they form a 3-element subgroup of L+F. All space rotations constitute\na subgroup of L+Fover any field F.If a space rotation fixes one of the\nthreestandard space unit vectors (0,1,0,0),(0,0,1,0) or (0,0,0,1) then it is\nsaid to be an elementary space rotation (aroundthat unit vector). For each\nof these three unit vectors, elementary space rotations around that vector\nconstitute a subgroup of the group of all space rotations. Aroun d (0,1,0,0)\nthis subgroup consists of the identity, the half-turn around (0,1,0,0) given\nby\n(t,x,y,z)/mapsto→(t,x,−y−z)\nand for each α/\\e}atio\\slash= 0 inFthe transformation Rαcalled abasic rotation whose\nstandard matrix representation is\n\n1\n1\nα−α−1\nα+α−12\nα+α−1\n−2\nα+α−1α−α−1\nα+α−1\n\n4Each of the two groups of elementary space rotations around (0 ,0,1,0) and\n(0,0,0,1) is obtained from the group of elementary space rotations aroun d\n(0,1,0,0) by conjugation with space axis rotations represented by appro pri-\nate permutation matrices. All space rotations are orthochronou s, over any\nfieldF.\nFor every non-zero square αinF(which means positive αinR, quadratic\nresidue class in Z/pZ∼=Fp) thebasic boost Bαis defined as the Lorentz\ntransformation over Fwith standard matrix representation\n\nα+α−1\n2α−α−1\n2\nα−α−1\n2α+α−1\n2\n1\n1\n\nFor any space rotation Randαany non-zero square in F, the conjugate\nRBαR−1is called a boost,and for any given space rotation Rthe set of\nboosts\n{RBαR−1:α/\\e}atio\\slash= 0, αsquare in F}\nis a subgroup of L+F. If the space rotation Ris a rotation of standard space\naxes, then RBαR−1issaidtobean elementary boost . Aboostiselementary if\nandonlyifitfixestwoofthethreebasisvectors(0 ,1,0,0),(0,0,1,0),(0,0,0,1).\nAll boostsareorthochronousover R,but boostsover a finitefield donot need\nto be orthochronous: consider over F7any of the two non-trivial boosts B2\nandB4.\nOver any field F, note that composing space-time reversal\n(t,x,y,z)/mapsto→(−t,−x,−y,−z)\nwith the half-turn around (0 ,1,0,0) yields reflection in the yz plane given by\n(t,x,y,z)/mapsto→(−t,−x,y,z).\nBasicbooststogetherwiththisreflectiongeneratethesubgroup ofL+Ffixing\n(0,0,1,0) and (0 ,0,0,1). (This subgroup is also referred to as the group of\nhyperbolic isometries of the txplane.)\nSince every space rotation in L+Ris the product of 3 elementary space\nrotations, elementary space rotations and basic boosts generat eL↑\n+R, and\n5together with reflection in the yzplane they generate L+R.On the other\nhand, a classical result on linear groups over finite fields due to Dicks on\n[D] states, when particularized to the case of 4-dimensional space s overFp\nfor primes p≡7mod8, that L+Fpis generated by elementary rotations,\nelementary boosts and reflections in the yzplane. Re-stated in a slightly\nstrengthened form, based on the observation that all elementar y boosts are\nconjugates of basic boosts by axis rotations which are in turn prod ucts of\nelementary space rotations, we have the following:\nRotation-Boost Lemma over Finite Fields (from Dickson [D]) For\nany prime number p≡7mod8,the proper Lorentz group L+Fpis generated\nby basic boosts, elementary space rotations and space-time reversal. /square\nIn fact the above lemma also holds for primes p≡3mod8 with the\nexception of p= 3 (see [D]) but this fact is not needed in what follows.\nUnlike over the real number field R, over finite fields not every proper\nLorentz transformation can be represented in the form RB(orBR) with\na rotation Rand a boost B.(Any such product of a rotation and a boost\ntransforms (1 ,0,0,0) to a vector ( t,x,y,z) wherex2+y2+z2is a square\nin the field of scalars. However, over finite fields not all proper Lore ntz\ntransformations, not even the orthochronous ones, have this p roperty.) For\nthe real case see Moretti’s study [M] for a comprehensive perspec tive on\nrotation-boost decomposition.\n2.2 Lorentz transformations over local rings\nWe represent the p-element field Fpas the quotient of the localization\nring\nZ(p)=Z·(Z/downslopepZ) ={st−1:s,t∈Z, t/\\e}atio\\slash= 0,(p,t) = 1}\nby its unique maximal ideal pZ(p), an approach also taken by Ahmavaara\n[A1]. There is a unique surjective ring homomorphism from the localizat ion\nringZ(p)onto the p-element field Fp, called the canonical map (orcanonical\nsurjection ) fromZ(p)toFp. (The restriction of the canonical map to Z⊆\nZ(p)is the unique surjective ring homomorphism Z−→Fpcorresponding to\nthe usual representation of Fpas a quotient of Z.) The canonical map from\nZ(p)toFpinducesa surjective ring homomorphism from the ring of 4-by-4\nmatrices with entries in Z(p)onto the ring of 4-by-4 matrices with entries in\n6Fpand thus inducesa map from any set of linear transformations of R4with\ncoefficients in Z(p)to the set of linear transformations of F4\np\nFor any integral domain Din which −1 is not a square, those D-module\nautomorphisms LofD4whose standard 4 ×4 matrix representation has\ndeterminant 1 and which leave invariant the ( D-valued) Minkowski norm\nµ(t,x,y,z) =t2−x2−y2−z2\narecalled Lorentz transformations over Dandtheyconstituteagroup, called\ntheproper Lorentz group over D, denoted L+D.A transformation LinL+D\nis said to be orthochronous if it maps (1 ,0,0,0) to a vector ( t,x,y,z) such\nthattis a non-zero square in D. The set L↑\n+Dof such orthochronous trans-\nformations may or may not be a subgroup of L+D. AD-module automor-\nphismLofD4having determinant 1 belongs to L+Dif and only if its stan-\ndard matrix representation Land the diagonal matrix Jwith main diagonal\n(1,−1,−1,−1) satisfy the equation LJL⊤=J.For any set Cof elements\nof the integral domain D,we denote by L+DC, respectively by L↑\n+DC,the\nset of those L∈ L+D,respectively L∈ L↑\n+D,for which all entries of the\nstandard matrix representation of Lare inC. IfCis a subring of D(con-\ntaining the unit 1 of D) thenL+DCis a subgroup of L+D. and restricting\ntransformations L∈ L+DCtoC4⊆D4yields a group isomorphism, called\ncanonical isomorphism,\nρ:L+DC−→ L +C\n(Theclosure of L+DCunder inversion isdue tothefact thatforall L∈ L+D\nwith matrix Lwe haveL−1=JL⊤J.) The isomorphism ρ:L+DC−→ L +C\nobviously maps orthochronoustransformationstoorthochrono ustransforma-\ntions,ρ[L↑\n+DC] =L↑\n+C,andL↑\n+DCandL↑\n+Care locally isomorphic subsets\neven if they are not subgroups of L+DCandL+C.They are isomorphic\nsubgroups, however, in the case of particular interest, which is D=Rand\nC=Z(p)(for any p≡3mod4).The following lemma provides a represen-\ntation of the proper Lorentz group over a finite field Fpas a quotient of the\nproper Lorentzgroupover the localizationring Z(p),or equivalently, asa quo-\ntient of the group of (real) proper Lorentz transformations with coefficients\ninZ(p).\nHomomorphism Lemma Letpbe any prime number congruent to 7\nmodulo8.The canonical map Z(p)−→Fpinduces a surjective group homo-\n7morphism L+Z(p)−→ L +Fp. Pre-composing it with the canonical isomor-\nphismρyields a surjective group homomorphism L+RZ(p)−→ L +Fp.\nProofIts is obvious that the map L+Z(p)−→ L +Fpinduced by the\ncanonical map Z(p)−→Fpis a group homomorphism. In L+Fpevery basic\nboost, every elementary rotation and also the space-time revers al transfor-\nmation is in the range of this homomorphism. Its surjectivity is then a con-\nsequence of the Rotation-Boost Lemma over Finite Fields. /square\nThe two homomorphisms onto L+Fpdefined by the above lemma will be\nalso referred to as canonical .\n2.3 Local isomorphisms\nThe next few lemmas are easily verified.\nInjection Lemma 1 Ifh:G1−→G2is a group homomorphism, A⊆\nG1and the restriction hAofhtoA∪AAis injective, then hAis a local iso-\nmorphismbetween Aandh[A]. /square\nAhmavaara [A1] seems to have been aware of the fact stated in Inj ection\nLemma 1, at least in the particular case where his the canonical surjection\nZ(p)−→Fpviewed as an additive group homomorphism, while apparently\ndisregarding (or treating differently) the case where his the restriction of\nthe same canonical map to Z∗\n(p)=Z(p)\\pZ(p)and viewed as a multiplicative\ngroup homomorphism onto F∗\np=Fp\\{0}.\nForeach positive integer kconsider theset Ckofrational numbers defined\nby\nCk={st−1:s,t∈Z, t/\\e}atio\\slash= 0,|s| ≤k,|t| ≤k}\nIfpis a prime number larger than k,thenCk⊆Z(p).\nInjection Lemma 2 For any prime number p≡3mod4letkbe a\npositive integer such that 2k2< p.Then the canonical map Z(p)−→Fpis\ninjective on Ckand the canonical map L+RZ(p)−→ L +Fpis injective on\nL+RCk./square\nInjection Lemma 3 For any prime number p≡3mod4letkbe a\npositive integer such that 32k16< p.LetA=L+RCk.Then the canonical\nhomomorphism L+RZ(p)−→ L +Fpis injective on A∪AA.\n8ProofA∪AA⊆ L+RC4k8and Injection Lemma 2. /square\nThese yield the following\nLocal Isomorphism Lemma For every positive integer k and prime\nnumber p ≡3mod4 such that p >32k16,the setL+RCkof proper Lorentz\ntransformations with coefficients in Ckis locally isomorphic to a subset of\nthe finite proper Lorentz group L+Fp.A local isomorphism is provided by\nrestricting the canonical homomorphism L+RZ(p)−→ L +FptoL+RCk./square\nUsing Dirichlet’s theorem on primes in arithmetic progressions, in [K1]\nKustaanheimo proved the following:\nKustaanheimo’s Chain Theorem [K1]For every positive integer k\nthere is a prime number p > k, p ≡7mod8,such that all the non-negative\nintegers up to kare quadratic residues modulo p. There are infinitely many\nsuch primes pfor any given k, they are necessarily larger than 2k./square\nFrom this we derive a key fact about the sets Ck, noting that these are\nobtained from {1,2,...,k}by symmetrization via the adjunction of negatives\nand reciprocals:\nSymmetrized Chain Lemma For every positive integer kthere is a\nprime number p > k, p ≡7mod8,such that all the positive numbers in\nCk⊆Z(p)are quotients a/bof quadratic residues a,bmodulop,i.e. all\npositive membersof Ckare mappedto non-zerosquares in Fpby the canonical\nsurjection Z(p)−→Fp.\nProofBy the Kustaanheimo Chain Theorem there is a prime number\np > k2, p≡7mod8,such that all the non-negative integers up to k2are\nquadratic residues modulo p. /square\n2.4 Density and conclusion of proof of the theorem\nBy thecoefficients of a linear transformation of Rnwe mean the entries\nof its standard matrix representation.\nIn this Section we deal mainly with the metric space structure on the\nset of all linear transformations of R4induced by the Euclidean norm of\nstandard matrix representations as briefly introduced in Section 1 . In this\n9metric space, the distance between two transformations T,Qis the norm of\ntheir difference, /bardblT−Q/bardbl.\nInL↑\n+=L↑\n+Rthe norm of any space rotation is 2 and the norm of the\nbasic boost Bαisα+α−1. IfTis any linear transformation of R4andR\nis any space rotation, then /bardblT/bardbl=/bardblRT/bardbl=/bardblTR/bardbl.The norm of a proper\northochronous Lorentz transformation T=RSBαSinL↑\n+, whereR,Sare\nspace rotations, is α+α−1≥2,and this is also the norm of the inverse\ntransformation T−1.\nIn any metric space, a set Aof elements is said to be ǫ-dense in a set\nof elements B(whereǫis any positive real) if for every b∈Bthere is an\na∈A∩Bat distance less than ǫfromb. Note that this does not require A\nto be a subset of B, butAisǫ-dense in Bif and only if A∩Bisǫ-dense in\nB.Note that if Aisǫ-dense in a set B, and also in a set C, thenAisǫ-dense\ninB∪C, but the converse implication generally fails. Density in the usual\nsense means ǫ-density for all ǫ >0.\nApproximation Lemma 1 For every positive integer kthe setCk2is\n(1/k)-dense in the real interval [−k,k].\nProofThis follows from the inclusion\n/braceleftbig\ni/k:i∈Z−k2≤i≤k2/bracerightbig\n⊆Ck2∩[−k,k]\n/square\nApproximation Lemma 2 Letǫ >0and let K be the union of a finite\nnumber of non-trivial compact real intervals. Then there is a positive integer\nl such that for all integers k≥lthe setCkisǫ-dense in K.\nProofApproximation Lemma 1. /square\nApproximation Lemma 3 Letǫ >0and letube any of the three\nstandard space unit vectors. Then there is a positive intege rlsuch that for\nall integers k≥lthe setL+RCkisǫ-dense within the set of space rotations\naroundu.\nProofSuppose u= (0,1,0,0). In the other cases the proof follows by\nconjugation.\nThe identity transformation and the half-turn around uhave their coef-\nficients in Ck.\nThere is some M >0 such that for all basic rotations RαwithM < α\nthe distance between Rαand the identity is less than ǫ.\n10There is some 0 < µ < M such that for all basic rotations Rαwith\n0< α < µ the distance between Rαthe and the half-turn is less than ǫ.\nLetK= [−M,−µ]∪[µ,M].By uniform continuity of the map α/mapsto→Rα\non the compact set K, there is δ >0 such that for all α,γ∈K,|α−γ|< δ\nimplies/bardblRα−Rγ/bardbl< ǫ.\nBy Approximation Lemma 2, there is a positive integer lsuch that for all\nintegersk≥lthe setCkisδ-dense in K.\nLetk≥land letRαbe a basic space rotation with α∈K.Byδ-density\nthere is a γ∈Ck∩Ksuch that |α−γ|< δ. We have /bardblRα−Rγ/bardbl< ǫ.\n/square\nApproximation Lemma 4 Letǫ >0andM >0.There is a positive\nintegerlsuch that for all integers k≥lthe setL+RCkisǫ-dense within\nthe set of basic boosts of norm not exceeding M.\nProofThere is some 0 < µ < M such that all basic boosts Bαwith\nα < µhave norm greater than M.\nLetK= [µ∪M].By uniform continuity of the map α/mapsto→Bαon the\ncompact interval K, there is δ >0 such that for all α,γ∈K,|α−γ|< δ\nimplies/bardblBα−Bγ/bardbl< ǫ.\nBy Approximation Lemma 2, there is a positive integer lsuch that for all\nintegersk≥lthe setCkisδ-dense in K.\nLetk≥land letBαbe a basic boost with α∈K.Byδ-density\nthere is a γ∈Ck∩Ksuch that |α−γ|< δ. We have /bardblBα−Bγ/bardbl< ǫ.\n/square\nThe above lemmas combine to yield the following\nApproximation Lemma 5 Letǫ >0andM >0.There is a positive\nintegerlsuch that for all integers k≥lthe setL+RCkisǫ-dense within\neach of the three elementary space rotation groups around th e standard space\nunit vectors, and also within the set of basic boosts of norm n ot exceeding M.\n/square\nApproximation Lemma 6 Letǫ >0andM >0.There is a positive\nintegerlsuch that for all integers k≥lthe setL+RCkisǫ-dense within the\nset of orthochronous proper Lorentz transformations of nor m not exceeding\nM.\n11ProofBy a compactness and uniform continuity argument, there is a\nδ >0 such that for all linear transformations\nT1,...,T10,Q1,...,Q10\nofR4of norm not exceeding M,if/bardblTi−Qi/bardbl< δfor all 1≤i≤10, then we\nhave for the distance of products\n/bardblT1...T10−Q1...Q10/bardbl< ǫ\nBy Approximation Lemma 5, there is a positive integer lsuch that for all\nintegersk≥lthe setLRCkisδ-dense within each of the three elementary\nspacerotationgroupsaroundthestandardspaceunitvectors, andalsowithin\nthe group of basic boosts of norm not exceeding M.\nLetTbeanyorthochronousproperLorentztransformation, T=RSBαS−1\nwhereR,Sare space rotations, and suppose that the norm α+α−1ofTis\nat mostM.We can factorize Ras a composition R1R2R3of three elemen-\ntary space rotations, and factorize also as Sas the composition R4R5R6of\nthree elementary space rotations. For convenience write R7=R−1\n6,R8=\nR−1\n5,R9=R−1\n4, so that\nT=R1...R6BαR7R8R9\nByδ-density, there are elementary space rotations R′\n1,...,R′\n9with coefficients\ninCkand a basic boost B′also inL+RCkand of norm at most M, such that\nfor each 0 ≤i≤9 the elementary rotations RiandR′\nihave a common fixed\nstandard space unit vector, and the distances\n/bardblR′\n1−R1/bardbl,...,/bardblR′\n9−R9/bardbl,/bardblB′−Bα/bardbl\narealllessthan δ.Thenthedistanceof Tfromthecomposition R′\n1...R′\n6BαR′\n7R′\n8R′\n9\nis at most ǫ./square\nConclusion of proof of Theorem 1 Givenǫ >0 andM >0,by\nApproximation Lemma 6 we can take a positive integer ksuch that the set\nL+RCkisǫ-dense within the set of orthochronous proper Lorentz transfo r-\nmations of norm at most M.\nBy the Symmetrized Chain Lemma, there is a prime p,\np >32k16> k, p≡7mod8\n12such that the canonical surjection Z(p)−→Fpmaps all of Ckinside the group\nof non-zero squares of Fp.Then the canonical surjection L+RZ(p)−→ L +Fp\nmaps all of L↑\n+RCkinsideL↑\n+Fp.LetA=L↑\n+RCk.\nDefinetheretraction f:L↑\n+R−→Abyassociating toeachorthochronous\nproper Lorentz transformation Ta member of Aat minimum distance from\nT.Then apply the Local Isomorphism Lemma of 2.3. /square\n3 Extended Lorentz groups and line reflections\nGivenǫ >0 andM >0 the retraction of the proper orthochronous\nLorentz group L↑\n+Rprovided by Theorem 1 can in fact be defined also on the\nother three connected components of the extended Lorentz gr oup, with the\nretractAbeinglocallyisomorphictoasubset Yoftheextended Lorentz group\nLFpof all bijective linear transformations of F4\nppreserving the Minkowski\nnorm. The local isomorphism between AandYcan also be seen to preserve\nseveral other properties of transformations besides the prope rty of being or-\nthochronous.\nAs overR, over any field F(where−1 is not a square) the proper Lorentz\ngroup is an index 2 subgroup of the extended group LF, and we denote by\nL−Fthe coset consisting of the transformations of determinant −1, called\nimproper transformations. Among these we have in particular the time re-\nversaltransformation τgiven by\nτ(t,x,y,z) = (−t,x,y,z)\nSimilarly toorthochronous transformationsin L+F,atransformationin L−F\nis also called o rthochronous if it maps (1 ,0,0,0) to a vector ( t,x,y,z) where\ntis a non-zero square in F. The set of orthochronous improper Lorentz\ntransformations is denoted by L↑\n−F.\nAmongallimproperLorentztransformations, afundamentalrole isplayed\nby linereflections intime-like orspace-like lines, i.e. 1-dimensional subs paces\nofF4consisting of the scalar multiples of a vector a∈F4with Minkowski\nnormµ(a) equal to 1 or −1. In the classical case F=Rthis role was re-\ncently emphasized by Urbantke in [U]. Over any field F,reflection in the\n13line ofa= (a0,a1,a2,a3)∈F4(whereµ(a) =±1) is the improper Lorentz\ntransformation with matrix\n2µ(a)[a⊤τ(a)−I]\nwhere in the square bracket the matrices a⊤,τ(a) andIare 4×1,1×4 and\n4×4,respectively, and τis time reversal.\nTheorem 2 For every ǫ >0andM >0the extended Lorentz group\nLRhas a retraction f:LR−→A, f2=f,onto a finite subset A⊆ LR\nsatisfying\n/bardblT−f(T)/bardbl< ǫ\nfor all Lorentz transformations T∈ LRof norm not exceeding M,and such\nthat for some prime number p≡7mod8the retract Ais locally isomorphic\nto a subset of LFpvia a local isomorphism λ,where the retraction fand the\nlocal isomorphism λsatisfy the following conditions:\n(i)fmaps each of the 4connected components of LRto its intersection\nwithA,\n(ii)λmaps each of the intersection sets\nA∩L↑\n+R, A∩L↓\n+R, A∩L↑\n−R, A∩L↓\n−R\nrespectively into the corresponding cosets\nL↑\n+Fp,L↓\n+Fp,L↑\n−Fp,L↓\n−Fp\n(iii)λmaps space rotations to space rotations, boosts to boosts, l ine re-\nflections to line reflections. /square\nThe extended retraction and local isomorphism of Theorem 2 are co n-\nstructed similarly to those of Theorem 1. Generating the extended finite\nLorentz group LFpby reflections now leads to a representation of the ex-\ntended finite Lorentz group LFpas a quotient of the group of all Minkowski\nnorm preserving automorphisms of the module Z4\n(p).\n4 Remarks on finiteness and approximation\n14It has been noted (see Beltrametti and Blasi [BB2]) that, as oppos ed to\nL↑\n+RandL↑R=L↑\n+R∪ L↓\n−R,the set of orthochronous Lorentz transforma-\ntions over a finite field does not constitute a group. For example, ta ke any\nprime number p≡7mod8 and such that 3 is a quadratic residue modulo p.\nThe first positive integer qsuch that q+1 is not a quadratic residue modulo\npis even, and q/2 is then a quadratic residue. Let αandγbe non-zero\nsquares in Fpsuch that the elements of Fpcorresponding to the integers\n2 andq/2 areα2andγ2respectively. Then the basic boosts BαandBγ\nare orthochronous but BαBγ=Bαγis not. However, for all orthochronous\ntransformations T1,T2in the set Y⊆ L+Fpof Theorem 1, the construction\nof the local isomorphism between A⊆ L↑\n+RandYimplies that T1T2is also\northochronous. Thus the phenomenon of an antichronous produ ct of two\northochronous parallel boosts arises only when the factors are n ot confined\nto the set Ywhich in the Lorentz group over Fprepresents, in the sense of\nlocal isomorphism, the retract A⊆ L↑\n+R.\nIn contrast to parallel boost groups in LR,inLFpthe boost group\n{RBαR−1:α/\\e}atio\\slash= 0, αsquare in Fp}\nis cyclic for any space rotation R, and it is isomorphic to the multiplica-\ntive subgroup of non-zero squares in Fpvia the map α/mapsto→RBαR−1. The\nfinite model then involves in each space direction the existence of a b oost\nfrom which all parallel boosts in that direction are obtainable by repe ated\napplication.\nAs in the real number based model, over a finite field Fpas well the\nvelocityvαassociated with the basic boost Bαis (α−α−1)/(α+α−1) and\nfor the velocities of boosts BαandBγand of their composition Bαγwe have\nvαγ=vα+vγ\n1+vαvγ\nIn finite field models as well as over R,velocity never equals 1 .Forp≡\n7mod8,ifα2+ 1 is a square in Fp(and this is always the case if Bαis in\nthe setY⊆ L↑\n+Fplocally isomorphic to the retract A⊆ L↑\n+Ras in Theorem\n1), then 1 −vαis a square in Fpi.e. in the non-transitive order 0 (permeability \u00161=\u00162=\u00160), and source/observation point above the interface\n(Fig. 1(a)).\nFigure 1: (a) Geometry under consideration: a plasmonic material is interfaced with a trans-\nparent dielectric. A vertically polarized point dipole source radiates close to the interface in\nthe dielectric region. The supported SPP dispersion surface is shown in (b) for a reciprocal\nisotropic plasmonic material, with no current bias, interfaced with free space, and (c) for\nthe same con\fguration but with the plasmonic material biased by a direct current with drift\nvelocity vd=\u0000c=10^x. (d,e) Source-excited in-plane distribution of the SPP electric \feld\n(time snapshot of the zcomponent) in the reciprocal and nonreciprocal cases: (d) without\nDC current at frequency !=!p= 0:6, and (e) with DC current at !=!p= 0:8. For panels\n(d) and (e) the source is located at z0=\u0015=10 above the interface, where \u0015is the free-\nspace wavelength at the radiation frequency. The black square at the center denotes the\nsource position. Animations of the calculated time-harmonic electric \feld distribution, for\nthe reciprocal and nonreciprocal cases, are included as Supplemental Material.\nFor the transparent dielectric region we consider vacuum and for the metallic/plasmonic\nregion, in the local non-biased case, we employ a standard Drude model for a free-electron\ngas with permittivity \u000f(!) = 1\u0000(!p=!)2=(1+i\u0000=!), where!pis the plasma frequency and \u0000\nis the damping rate. In the following, we \frst study the problem with \u0000 = 0 (lossless case),\nwhereas the impact of dissipation is discussed in details in Section 2.3. At a local, isotropic,\nmetal-vacuum interface, SPPs are known to exist for frequencies !\u0014!SPP=!p=p\n2. The\n63D dispersion surface of these SPP eigenmodes, calculated by tracking the poles of the Green\nfunction integrand in Eq. (33) of the Supplemental Material, is shown in Fig. 1(b) for the\nlossless case. The dispersion surface clearly exhibits a \rat dispersion asymptote at the surface\nplasmon resonance !=!SPP, as expected for a local/lossless Drude model56(see, e.g.,1,8and\nreferences therein, for a discussion of how losses and nonlocal e\u000bects modify this dispersion\ndiagram). The rotationally symmetric shape of the dispersion surface with respect to in-plane\nwavevectors indicates that the surface mode propagates reciprocally and isotropically on the\ninterface. The spatial \feld distribution of the SPPs excited by a vertically polarized source\nas in Fig. 1(a) was calculated using our 3D Green function formalism at != 0:6!p(see Fig.\n1(d)), con\frming that surface waves propagate symmetrically and omnidirectionally along\nthe interface.\nAs discussed in Ref.,1strong nonreciprocity and unidirectionality in a continuous plas-\nmonic platform of this type can be achieved if the mechanism that breaks reciprocity intro-\nduces an asymmetry in the dispersion asymptote for large wavevectors. While this is typically\nachieved with a static magnetic bias, a direct current bias, which can be directly carried by\nplasmonic materials due to their conductive nature, can also introduce a strong asymmetry,\nbut with qualitatively di\u000berent characteristics, as originally discussed in.47The presence of\na DC current means that conduction electrons move with average drift velocity vd=J0=ne\nwhereJ0is the current density, n=m!2\np=4\u0019e2is the electron density, and eis the negative\nelectron charge. Ref.47argues that the movement of electrons produces a Doppler frequency\nshift in the material permittivity, i.e., \u000f(!\u0000k\u0001vd) = 1\u0000!2\np=(!\u0000k\u0001vd)2, such that the\nmaterial response becomes spatially dispersive, with a wavevector-dependent isotropic per-\nmittivity. This point, and especially the isotropy assumption, is worth some clari\fcation. As\ndiscussed in,55,60in the presence of any non-trivial electron velocity distribution (due to a\ncurrent bias or just the thermal motion of free electrons), the permittivity of a free-electron\ngas becomes spatially dispersive (in other words, the material response is nonlocal due to the\nelectrons moving a certain distance during one period of the applied electromagnetic \feld).\n7Importantly, in the presence of spatial dispersion, the material permittivity is a tensor even\nin an isotropic medium. The general form of this wavevector-dependent permittivity tensor\ncan be written (in tensor notation in an arbitrary coordinate system) as,55\n\u000fij(!;k) =\u000fT(!;k)\u000eij+ (\u000fL(!;k)\u0000\u000fT(!;k))kikj=k2(1)\nwhere the scalar functions \u000fTand\u000fLare the transverse and longitudinal permittivities (for\nelectric \felds orthogonal or parallel to k, respectively), kis the wavevector magnitude, kikj\ndenotes a tensor product between the wavevector and itself, and \u000eijindicates the identity\nmatrix. In Supplemental Material (Notes I and II), we derive \u000fTand\u000fLstarting from the\nVlasov equation for plasmas, and we show that only the longitudinal permittivity is perfectly\nDoppler-shifted in the presence of a drift current, whereas the transverse permittivity exhibits\na di\u000berent, weaker form of spatial dispersion. Since the electric \feld of a SPP mode on a\n3D plasmonic material has both transverse and longitudinal components inside the material,\nits dispersion and propagation properties are expected to depend on both \u000fTand\u000fL. (in\na 2D electron gas, instead, it is clear that only the longitudinal conductivity is relevant\neven if the \feld has a transverse out-of-plane component). This implies that the modeling\napproach proposed in,47which uses the same equation for SPP dispersion as in the local\nunbiased case, but substituting the longitudinal Doppler-shifted permittivity, can only be\napproximately valid. To clarify this point, in the Supplemental Material (Note I), we derive\nthe exact modal solution for SPPs propagating along the electron current and compare it\nagainst the SPP dispersion obtained when only the longitudinal or transverse permittivity\nis considered. Our \fndings show that, while one has to be aware of this issue for more\naccurate calculations, the model proposed in47using only the longitudinal permittivity is\napproximately correct and captures the relevant physics especially in the large-wavevector\nregime. Based on these considerations, we can therefore safely use this approximate model to\ndrastically simplify the analysis and numerical calculations, and provide additional physical\ninsight. Furthermore, it should be mentioned that, while the nonlocal model based on\n8the Doppler e\u000bect is qualitatively valid for the 3D plasmonic materials considered here, its\nvalidity for 2D systems, such as graphene, has been the subject of debate in the literature,57\nand alternative models have been proposed in.43{45,49,58However, it has been argued that,\neven for 2D systems, this model is accurate when the electron-electron collisions predominate\nand force the electron gas to move with approximately constant velocity, which makes the\ndrift-biased medium act similarly to a moving medium.52,59\nWhile Ref.47focused on 1D propagation in the direction of the current, here we use\nour 3D Green function formalism to investigate drift-induced nonreciprocal e\u000bects over the\nentire two-dimensional interface of the three-dimensional geometry in Fig. 1(a), revealing\nanomalous and extreme forms of surface-wave propagation on this surface. The modi\fed\ndispersion and propagation properties of the SPPs supported by this nonreciprocal con\fg-\nuration can be readily analyzed using the same 3D Green function as for the reciprocal\ncase, but, as explained above, with the nonlocal, Doppler shifted, longitudinal permittivity,\n\u000fL(!;k) = 1\u0000!2\np=(!\u0000k\u0001vd)2. As a \frst example, we suppose the electron \row is along\nthe +x-axis, with drift velocity vd=\u0000c0=10 (wherec0is the speed of light in vacuum). We\nmust emphasize that this value of the drift velocity is unrealistically high and is considered\nhere just for illustrative purposes to describe the qualitative behavior of surface waves in this\nnonreciprocal platform. Later in the article, and in the Supplemental Material, we provide\nresults for much smaller values of drift velocity. The modi\fed SPP dispersion surface is\nshown in Fig. 1(c). The dispersion asymptote is now tilted as a result of the Doppler shift,\nas discussed in,47making the dispersion surface strongly asymmetric and nonreciprocal in\nthex-direction, with a frequency range where SPPs can only propagate toward the nega-\ntivex-axis. In addition, unidirectional SPPs are now supported at frequencies where SPPs\ncould not propagate at all in the reciprocal case (frequencies higher than the \rat asymptote\nin Fig. 1(b)). The \feld distribution of surface waves launched by a vertical dipole source\nis shown in Fig. 1(e) for a frequency higher than the original surface-plasmon resonance,\n!= 0:8!p>!SPP. We clearly see that, in this case, SPPs propagate preferentially toward\n9left, parallel to the electrons drift velocity, despite the homogeneity and isotropy of the sur-\nface. While this behavior is somewhat expected as an extension of the 1D results of Ref.47\nto a 2D surface, even more extreme wave-propagation e\u000bects emerge at di\u000berent frequencies\nand angles, as discussed in the next sections.\nIn\rexion Points and Slow-Light Beaming\nThe results in Fig. 1 clearly show that the dispersion properties of drift-biased SPPs strongly\ndepend on the angle \t between in-plane wavevector and current \row, as expected from the\nspatially dispersive form of the permittivity. To clarify this point further, Figure 2(a) shows\n1D dispersion curves for SPPs propagating along di\u000berent angles. SPPs propagating normal\nto the current \row (\t = 90 deg.) do not interact with the drifting electrons and the\ndispersion curve is the same as in the reciprocal case with \rat asymptotes at !=!SPP.\nConversely, SPPs propagating along the current \row (\t = 0) are maximally a\u000bected by the\npresence of the current, particularly in the large-wavevector asymptotic region where spatial\ndispersion becomes particularly strong. The tilted dispersion asymptote is associated with\nhighly localized SPPs dragged by the \row of electrons,47and its slope (group velocity) is\nindeed exactly equal to the electron drift velocity seen by the SPP mode, i.e., @!=@k =vd\u0001^uk,\nwhere ^ukis the in-plane unit wavevector. Such a tilted asymptote makes the dispersion curve\nnon-monotonic for wavevectors opposite to vd, with an in\rexion point (red dot in Figure\n2(a)) where the SPP group velocity vanishes and changes sign. The position of the in\rexion\npoint along the frequency axis depends on the direction of propagation, such that, for each\nangle \t, an in\rexion point exists at a di\u000berent frequency,\n10!inf=!pp\n2\"\n1\u00003\n2\u0012vdcos(\t)\n2c\u00132=3#\n)8\n>><\n>>:!min\ninf=!pp\n2h\n1\u00003\n2\u0000vd\n2c\u00012=3i\n;\t = 0\n!max\ninf!!pp\n2; \t!90 deg:(2)\nThus, while an in\rexion point exists for any angle \t 2[0;90) deg., its frequency is limited\nwithin the range !min\ninf\u0014!inf\u0014!max\ninf, as shown in Fig. 2(b). Minimum in\rexion-point\nfrequency occurs for SPP propagation along the current \row and maximum frequency is\nasymptotically reached for propagation normal to the current, in which case the in\rexion\npoint rapidly migrates to large values of wavevector as seen in Fig. 2(a). For any frequency\noutside this window, no in\rexion point appears, i.e., the SPP group velocity does not vanish\nalong any direction.\nThe existence of an angle-dependent in\rexion point in this frequency window leads to\nnontrivial propagation e\u000bects on the 2D surface of the considered nonreciprocal platform.\nThese e\u000bects, which have not been studied previously, can be elucidated by considering\nthe system's Green function under quasi-static approximation. Indeed, while the 3D Green\nfunction described above provides a complete formulation to study SPP propagation, it is\nusually evaluated numerically and tends to hide physical insight. A simpler, more reveal-\ning, approximate formulation can be obtained by neglecting retardation e\u000bects (quasi-static\napproximation) and assuming that the main radiation channel of the dipolar source is rep-\nresented by the excitation of a single, guided, surface mode. As was shown in Ref.,62under\nthese approximations, the source radiation intensity in a certain in-plane direction (radiated\npower per unit angle) is given by\nU(\t)\u0019!2\n16\u00191\njrkt!(kt)j1\nC(kt)j\r\u0003\u0001Ek(z0)j2; (3)\n11Figure 2: (a) Dispersion curves for SPPs propagating along di\u000berent angles with respect\nto the current \row. The angle \t is measured from the positive x-axis (see Fig. 1(a)).\nDi\u000berent curves correspond to the intersection of the dispersion surface in Fig. 1(c) with\nplanes at angle \t with respect to the kx-axis. The red dots indicate in\rexion points for\ndi\u000berent directions of propagation. (b) In\rexion-point frequencies for SPPs propagating\nalong di\u000berent angles. (c,d,e) Source-excited in-plane distribution of the SPP electric \feld\n(time snapshot of the zcomponent) for (c) !=!min\ninf, (d)!= 0:6!pand (e)!\u0019!SPP=\n!max\ninf. Thexandyaxes are normalized with respect to the free-space wavelength at the\nradiation frequency. The black squares indicate the in-plane position of the source, which\nis located in the vacuum region at z0=\u0015=10 above the interface for panels (c,d) and at\nz0=\u0015=30 for panel (e) (the source is closer to the interface in this panel to e\u000eciently excite\nthe high-wavevector components of the SPPs). The inset in panel (c) shows the amplitude\ndistribution of the z-component of the electric \feld, and the inset in panel (d) shows the real\nand imaginary parts of Ezalong the red dashed line crossing the narrow beam in the main\npanel.\nwhere ktis the in-plane wavevector, tan(\t) = k y=kx,Ek(z0) is the modal electric \feld at\nthe location z0of a source with polarization state \r,!(kt) is the dispersion relation of the\nrelevant mode (hence, rkt!(kt) is the group velocity), and C(kt) is the local curvature of the\nequifrequency contour (EFC) of the dispersion relation. Equation (3) gives the approximate\nin-plane radiation pattern of the dipole source, corresponding to the in-plane pattern of SPP\npropagation. This equation reveals that the SPP pattern can be controlled in two ways: (i)\n12by suitably selecting the polarization of the dipole source, which controls the coupling factor\nj\r\u0003\u0001Ek(z0)j2between the source and the surface mode of interest; or (ii) by engineering the\ndispersion relation of the relevant surface mode, namely, by controlling the local curvature\nC(kt) of the equifrequency contour and/or the wavevector dependence (angular dependence)\nof the group velocity, jrkt!(kt)j. In the considered current-biased plasmonic platform, the\npresence of in\rexion points where the group velocity vanishes is therefore a crucial factor\na\u000becting the in-plane propagation pattern. Indeed, at any frequency where an in\rexion point\nappears, the inverse of the group velocity in Eq. (3) diverges, 1 =jrkt!(kt)j!1 (in practice\nit becomes large but \fnite) along a speci\fc angle with respect to the current \row. Thus,\nthe in-plane propagation pattern is maximized and localized around this angle, resulting in\na peculiar \\beaming e\u000bect\" as further discussed below.\nFigure 2(c,d,e) show the in-plane \feld distributions, calculated using our Green func-\ntion formulation, for SPP propagation at three frequencies: (c) minimum and (e) maximum\nin\rexion-point frequencies and (d) another frequency in between. As shown in Fig. 2(a,b),\nminimum in\rexion-point frequency corresponds to vanishing SPP group velocity in the di-\nrection of the current \row. This leads to a diverging behavior in Eq. (3) at \t = 0, which\nimplies that the SPP \felds form a unidirectional narrow beam along the current \row, as con-\n\frmed by Fig. 2(c). Note that no energy actually propagates toward the + x-axis since the\ngroup velocity is zero (group velocity and energy velocity are virtually identical in a low-loss\nmedium even if material dispersion is high), and the narrow beam resembles a standing wave\nwith localized and enhanced \felds, as seen in the \feld amplitude distribution plotted in the\ninset of Fig. 2(c) (an animation of the calculated time-harmonic electric \feld, which further\nclari\fes this peculiar behavior, is included as Supplemental Material). Choosing a di\u000berent\nfrequency, for example != 0:6!p, an in\rexion point appears for SPPs propagating at an\nangle \t = 60 deg. with respect to the current \row (see Fig. 2(a), dashed blue line). This\nleads to narrow SPP beams along \t = \u000660 deg., as shown in Fig. 2(d). In this scenario, two\nbeams emerge because the system's response is reciprocal normal to the current, therefore\n13SPPs propagate symmetrically along the y-axis. As in the previous case, these beams are\nnarrow standing waves with enhanced and localized \felds. Interestingly, in this case the\n\feld distribution is even sharper around \t = \u000660 deg., especially toward smaller angles,\nwhere no surface-wave propagating solution exists at this frequency (see Fig. 1(c) and the\nequifrequency-contour animation in Supplemental Material). We note that, although the\n\feld pattern looks discontinuous around this angle, a closer inspection reveals that, rather\nthan an actual \feld discontinuity (which would be nonphysical on a homogeneous surface),\nthe \felds decay very sharply, within a small fraction of a wavelength, as demonstrated in\nthe inset of Fig. 2(d) which shows the real and imaginary parts of the electric \feld along the\nred dashed line. We speculate that such ultra-sharp \felds on a homogeneous surface may\nbe useful for surface-sensing applications or optical trapping near the surface. Finally, Fig.\n2(e) shows the \felds at !\u0019!SPP=!max\ninf, at which an in\rexion point appears for SPPs\npropagating close to \t = \u000690 deg. Indeed, in this case, the \feld distribution is dominated by\nthe presence of narrow beams almost normal to the current \row and with large wavenumber,\nconsistent with Fig. 2(a).\nThe emergence of these narrow, current/frequency-steerable, frozen-light beams with\nenhanced \felds is highly nontrivial, especially given the fact that the plasmonic platform\nthat supports them is homogeneous, and they may o\u000ber new opportunities for tunable,\ndirectional, enhanced, light-matter interactions on plasmonic surfaces.\nUnidirectional Propagation and Negative Group Velocity\nAs discussed above, for frequencies higher than !max\ninf=!SPPor lower than !min\ninf, no in\rexion\npoint appears, and hence the SPP group velocity never vanishes along any in-plane direc-\ntion. In these frequency regions, it is mostly the curvature of the equifrequency contour C(kt)\nthat determines, according to Eq. (3), how SPPs propagate on the two-dimensional interface.\nOther anomalous wave-propagation e\u000bects occur in this regime, qualitatively di\u000berent from\nthe behavior of SPPs in the !min\ninf!SPPand!= 0:52!p !inf, the poles rapidly migrate away from the real axis (Fig.\n4(d)), leading to an interesting modal transition further discussed below. This behavior\nindeed suggests that the in\rexion point could also be interpreted, in this 1D propagation\nscenario, as an exceptional point of degeneracy since it marks the coalescence and branching\nof modes.66,67\nTo further con\frm this interpretation, we considered the general conditions that a solu-\ntion of the dispersion equation, in a 1D wave-guiding system, must satisfy to be identi\fed as\nan EP, as originally derived in.63{65The \frst necessary condition to obtain a \frst-order EP\nis that two \frst-order roots of the dispersion equation, D(kEP;!EP) = 0, coalesce to form a\nsecond-order root, which implies that, at the EP,\nD(kEP;!EP) =@D(k;!)\n@k\f\f\f\fk=kEP!=!EP= 0 (4)\nwherekEPand!EPare the wavenumber and angular frequency at which the degeneracy\nemerges. However, an EP should also satisfy an additional condition,\n@D(k;!)\n@!\u0001@2D(k;!)\n@k2\f\f\f\fk=kEP!=!EP6= 0: (5)\nIn fact, only if Eq. (5) is satis\fed, then the second-order root becomes a branch point (and\nnot a saddle point) where various branches of the dispersion function k(!) merge.63{65Our\n18Figure 4: (a) Dispersion diagram for SPPs propagating in the direction of the current \row\n(\t = 0), similar to Fig. 2(a), but tracking each individual solution of the dispersion equation,\nand plotting its complex in-plane modal wavenumber, as frequency is varied. Absorption\nlosses are assumed to be zero, i.e., \u0000 = 0. See Fig. 5 for the dissipative case. Solid purple\nlines indicate the light cone. (b,c,d) SPP pole constellation, on the complex kx-plane (for\n\fxedky= 0), at di\u000berent frequencies: (a) !=!p= 0:54< !min\ninf, (b)!=!p= 0:5457 =!min\ninf\nand (c)!=!p= 0:55> !min\ninf, for SPP modes propagating along the drifting electrons with\nvd=\u0000c=10. The peculiar character of the two complex modes in panel (d) is discussed in\nthe text.\nnumerical tests applied to the case of SPPs propagating along the electron current verify that\nthe local 1D dispersion function, kx(!) (forky= 0), satis\fes conditions (4) and (5) around\nthe in\rexion frequency (red point in Fig. 2(a)), hence con\frming that in this 1D scenario,\nthe in\rexion point can be interpreted as an EP (branch point) where forward-propagating\nand backward-dragged SPPs coalesce. We also note that, unlike typical EPs in open non-\nHermitian systems, the EP observed here exists in a system with no balanced distribution\n19of loss and gain. Indeed, this dispersion feature is more similar to the EPs associated with\nthe band edges of lossless periodic structures (see, for example, Refs.61,67,68); however, in the\nconsidered current-biased system, the EP exists at a sort of unidirectional band edge (the\nin\rexion point) created not by a periodic modulation, but by the Doppler shift resulting\nfrom the applied current.\nThe modal transition occurring through this EP is particularly interesting as it leads to\na rich and subtle behavior at higher frequencies, and some additional observations should\nbe made. As discussed in Ref.,1in a system with continuous translational symmetry as\nthe one considered here, a direct transition from a purely propagating solution to a purely\nevanescent one, or viceversa, can only occur at band-edges or cut-o\u000b frequencies at zero\nor in\fnite wavenumbers (where the modal wavenumber can go directly from purely real to\npurely imaginary). This is clearly not the case in Fig. 4(a), where the band-edge (exceptional\npoint) exists in the dispersion diagram at a \fnite non-zero value of wavenumber and the\nsolutions above the EP frequency are indeed complex, with the same real part but opposite\nimaginary parts. However, since we assumed that the considered structure is lossless, \u0000 = 0,\na complex wavenumber is puzzling because it implies some form of energy dissipation or\ngain as the wave propagates. One of the two complex solution branches above the EP in\nFig. 4(a) represents an exponentially growing propagating wave, which, we argue, is a non-\nphysical solution since the considered system does not provide optical gain. Yet, due to\nthe presence of the drift current, this point deserves further clari\fcation. Indeed, a direct\ncurrent bias may in principle lead to optical gain through negative Landau damping in an\notherwise passive system: if the drift velocity is comparable and greater than the wave\nphase velocity, vd> !=k , the electrons \\trapped\" inside the moving potential wells of the\nwave lose kinetic energy to the wave (the opposite process is the regular Landau damping\ndiscussed in the following).52,60However, this cannot happen in the present situation since\nthe SPP phase velocity in the region of interest of the dispersion diagram is much larger\nthanvd. In addition, it was convincingly demonstrated in Ref.46that SPP ampli\fcation and\n20instabilities are not possible in a con\fguration with a single drift-biased plasmonic medium.\nThis also makes sense physically by considering the analogy with moving media: a single\nbody moving at a constant velocity cannot lead to wave instabilities; instead, the presence\nof another body in close proximity and in relative motion is necessary, for example two\ngraphene sheets with di\u000berent bias velocities,46or a single drift-biased graphene sheet on a\nnon-biased SiC substrate.52Based on these considerations, we can safely say that the case\nconsidered in this section, with a single drift-biased material, cannot lead to ampli\fcation\nand instabilities, and any unstable pole can be considered unphysical. Indeed, the presence\nof complex poles with opposite imaginary parts is common in even simpler passive systems,\nfor example beyond a band-edge in the gapped band diagram of a periodic structure, or\nbelow the cut-o\u000b frequency of a waveguide. Mathematically, poles on both sides of the real\naxis do exist in these scenarios, but only the exponentially decaying solution is retained on\nphysical grounds (i.e., the proper branch of a multivalued dispersion relation is chosen based\non considerations of physical realizability).\nGoing back to our discussion of the dispersion diagram in Fig. 4(a), we note that the other\nsolution branch above the EP corresponds, instead, to a physical, decaying, propagating wave\nthat becomes overdamped as it crosses the light line at higher frequencies, which suggests\nthat radiation leakage plays an important role in this behavior. In fact, analogous modal\ntransitions (albeit for reciprocal systems) were observed and investigated in the context of\nleaky-wave antennas a few decades ago.70,71In the transition region between bound- and\nleaky-waves, it was similarly observed that certain propagating solutions are non-physical\n(growing in the longitudinal direction), and the unexpectedly lossy solutions below the light\nline, in an otherwise lossless structure, can be explained as the \\winding down\" of the leaky-\nwave solution right under the light cone (with a virtually negligible contribution to the total\n\felds excited by an actual source71).\nFinally, we investigated how the dispersion diagram changes when absorption losses are\nincluded in the current-biased plasmonic material. This is particularly important consid-\n21Figure 5: Impact of dissipation on nonreciprocal current-biased SPPs. (a,b) Dispersion\ncurves for SPPs propagating along the current \row, with vd=\u0000c=10 in the presence of\nintrinsic (bulk) scattering losses: (a) \u0000 = 0 :02!pand (b) \u0000 = 0 :1!p. (c,d) Dispersion curves\nfor SPPs propagating along the current \row, with vd=\u0000c=40 and (c) \u0000 = 0 :1!pand (d)\n\u0000 = 0:15!p. For the forward-propagating mode with largest wavevector, the imaginary part\nis plotted over a limited frequency range due to numerical di\u000eculties in tracking this solution\nbeyond this range. Solid black lines denote the lossless case (showing only the real branches\nfor simplicity), and dotted green and dashed red lines indicate the real and imaginary part\nof the in-plane modal wavenumber, Re( k) and Im(k), respectively. Arrows indicate which\nimaginary and real branches correspond to the same mode. Solid purple lines denote the\nlight cone. (d) Phase velocity of SPP modes propagating parallel to the drifting electrons\nwithvd=\u0000c=10, compared with a Maxwell-Boltzman velocity distribution for a typical\nplasmonic material at room temperature and with the same electron drift velocity. The\ninset shows a zoomed-in view of the distribution function.\nering the virtually unavoidable presence of dissipation mechanisms in plasmonic media. In\nparticular, here we \frst studied the impact of the intrinsic (bulk) scattering losses of the\nmaterial, modeled by an intrinsic damping rate \u0000 which is assumed to be unaltered by the\npresence of the current bias, as done for example in.52However, how this damping rate\na\u000bects the propagation and attenuation of SPPs strongly depends on the presence and ve-\nlocity of the drift current, the SPP wavevector, and the speci\fc structure. As seen in Fig.\n225(a,b), increasing \u0000 determines an increase in the imaginary part of the modal wavenumber,\nespecially in the large-wavevector regime, as expected. Non-zero dissipation also \\opens\" the\nband bifurcation at the EP, such that the dispersion curves no longer \ratten out completely.\nThis behavior is therefore qualitatively di\u000berent from the way losses a\u000bect nonreciprocal\nSPPs in magnetically biased plasmas,1,8where band edges are located at diverging values\nof wavenumber. As a result of this modi\fed dispersion diagram, the e\u000bects described in\nSection 2.1, which mostly depend on a slowing down of the group velocity at the in\rexion\npoint, are expected to be sensitive to the presence of dissipation (which may actually be\nuseful for the sensing of absorbing layers on the plasmonic platform). Figure 5(b,c) com-\npare two cases where \u0000 is kept \fxed but the drift velocity is decreased. As seen here, by\nreducing the drift velocity, the imaginary part of the modal wavenumber generally increases,\nindicating that, in this con\fguration, drift-induced nonlocality and nonreciprocity may in-\ndirectly mitigate the detrimental impact of scattering losses by modifying the dispersion\ndiagram and the group velocity of the modes. Finally, if losses are su\u000eciently large and\nthe drift velocity is low (Fig. 5(d)), dissipation starts to dominate over the drift-induced\nnonlocal and nonreciprocal e\u000bects. In particular, the backward-propagating mode now ex-\nhibits the typical \\back-bending\" of a lossy SPP (real dispersion curve bends in the opposite\ndirection) and becomes overdamped at higher frequencies (Re[ k] !=c . The heat \rux density from body 2 to body 1 can be ob-\n26tained from Eqs. (7) and (8) by exchanging the subscripts 1 and 2 and changing the tempera-\nture in Eq. (6) from T1toT2. In the above equations, \u000bis the decay rate of evanescent waves\nalong thez-axis,Ri(kt;!) is the re\rectivity matrix of body i, whose elements are the re\rec-\ntion coe\u000ecients for light incident from vacuum, for both transverse-electric and transverse-\nmagnetic polarizations, and the parameter ^D12=h\n^I\u0000^Ry\n2(kt;!)^R1(kt;!)e\u0000i2kzdgi\nrepresents\nmultiple re\rections between the two bodies (see Ref.38for more details). This formalism al-\nlows us to calculate the near-\feld 3D heat transfer between two planar slabs with and without\na current bias. As in our 3D Green function formalism in Section 2 and Supplementary Note\nIII, the presence of a non-zero drift current modi\fes the permittivity, making it wavevector-\ndependent, and therefore changes the re\rection coe\u000ecients for plane-wave incidence.\nA nonzero net heat \row in a two-body geometry as in Fig. 6(a) requires a temper-\nature gradient, as dictated by the Second Law of Thermodynamics. In other words, if\nthe two bodies are at the same temperature, the total heat \rux densities between them\nare identical, S12=S21(integrated over all frequencies and wavevectors), independent of\nwhether the system is reciprocal or not. It is then possible to show that thermodynamics im-\nposes the same symmetry also for the individual contributions, i.e., S12(kt;!) =S21(kt;!),\nagain independent of reciprocity or the lack thereof. Interestingly, Ref.38demonstrated\nthat reciprocity imposes another symmetry constraint for opposite in-plane wavevectors:\nS12(kt;!) =S12(\u0000kt;!) =S21(\u0000kt;!). In a nonreciprocal system, this symmetry con-\nstraint can be violated since the dispersion diagram is no longer necessarily symmetric,\n!(kt)6=!(\u0000kt). As a result, S12(kt;!)6=S12(\u0000kt;!) =S21(\u0000kt;!), which implies that if\nheat transfer from body 1 to body 2 occurs through a certain wavevector-channel, k=kz+kt,\nthe transfer from body 2 to 1 does not have to occur symmetrically through the time-reversed\nchannel,kTR=\u0000kz\u0000kt=\u0000k, whose contribution can even be zero; the reverse transfer\ncan now occur entirely through a di\u000berent channel with the same kt, i.e., the specular one,\nkS=\u0000kz+kt. This results in a persistent heat current between the two bodies even when\nthey are kept at the same temperature.37,38In other words, while a nonreciprocal body may\n27emit strongly in a certain direction de\fned by k(at a certain frequency), it could be designed\nto absorb very weakly in that same direction (i.e., for \u0000k) and frequency, corresponding to\na breaking of Kirchho\u000b's law of thermal radiation (equal absorptivity and emissivity at each\ndirection and frequency).30,76{78We stress again that the existence of a persistent heat cur-\nrent for bodies at the same temperature or a violation of Kirchho\u000b's law in these systems\ndoes not break thermodynamic laws, since the total net heat \row into each body is still zero\nwhen the bodies have the same temperature. In the following, we provide the \frst theoret-\nical demonstration of these e\u000bects in a platform that does not require magneto-optical or\nspace-time-modulated media.\nFirst we consider the reciprocal scenario with no current bias in either the lower or upper\nbody. In this situation the material permittivity takes a local and isotropic form, where we\nassume the same plasma frequency, !p, and scattering rate, \u0000 = 0 :1!p, for both materials.\nFigure 6(b) shows the heat \rux density at each frequency and in-plane wavenumber. As\nexpected for reciprocal materials, the heat \rux density is symmetric with respect to in-\nplane wavenumber, S12(kt;!) =S12(\u0000kt;!), and the same plot is obtained for any in-plane\ndirection \t since the considered material is homogeneous and isotropic. We also veri\fed\nthat this wavevector- and frequency-resolved heat transfer map from body 1 to 2, S12(kt;!),\nis identical to the corresponding one from body 2 to 1, S21(kt;!), therefore respecting the\nSecond Law. We also note that the speci\fc shape of the heat \rux density map in Fig. 6(b)\nis a result of coupling between the surface waves supported by the two interfaces, which\nstrongly modi\fes their individual dispersion diagram. For subwavelength gaps between the\ntwo slabs, as in Fig. 6, the structure supports strongly con\fned SPPs, for which even\nrelatively small values of drift velocity determine large changes in their response, as seen in\nour results discussed in the following.\nWe make the system nonreciprocal by biasing the top conductor with a direct current\nwithvd=c=300 (much lower than in previous sections) along the + x-axis. The heat trans-\nfer spectra for this nonreciprocal system are shown in Figs. 6(c,d,e) for di\u000berent in-plane\n28Figure 6: (a) Illustration of the geometry considered for investigating the near-\feld radiative\nheat transfer between two planar bodies in the reciprocal and nonreciprocal (current-biased)\ncases. Two planar conducting slabs are placed at a distance dg=\u0015p=100 from each other\nand one of them is biased by a DC drift current. The thickness of the slabs is d=\u0015p, where\n\u0015pis the free-space wavelength at the plasma frequency. The two slabs are supposed to have\nidentical plasma frequency and damping rate, \u0000 = 0 :1!p. (b) Heat \rux density between the\ntwo bodies as a function of frequency and in-plane wavenumber when the current bias is set\nto zero. Identical heat \rux density maps are obtained in this case for any in-plane direction\nde\fned by the angle \t with respect to the + x-axis. (c,d,e) Heat \rux density for the same\ncon\fguration but with the top conducting body biased by a DC current with vd=\u0000c=300\nalong the + x-axis. The three panels show the heat \rux density map for di\u000berent in-plane\ndirections. While the system being studied is, strictly speaking, an active non-equilibrium\nsystem, we operate in a regime with relatively large separations and low velocities, where\nnon-equilibrium phenomena are still small, but nonreciprocal e\u000bects are clearly visible, as\nfurther discussed in the text.\nwavevector directions, de\fned by the angle \t with respect to the current \row. As seen\nfrom these results, S12(kt;!)6=S12(\u0000kt;!) for any in-plane direction except at \t = 90 deg.,\nwhich is consistent with our previous observation that there is no interaction between surface\nwaves and drifting electrons for wavevectors normal to the current \row. In this latter situa-\ntion, the heat \rux density map is exactly the same as in the reciprocal scenario in Fig. 6(b).\n29These results clearly demonstrate that such a magnet-free current-biased system exhibits a\nsignature of nonreciprocity for near-\feld radiative heat transfer, which implies, as mentioned\nabove, the presence of a persistent heat current between the two bodies even when they are\nkept at the same temperature. The asymmetric heat transfer spectra in Figs. 6(c,d) are\nsimilar, but not exactly identical, to the ones obtained with magneto-optical materials in.38\nIn particular, the heat transfer spectra in38follow the characteristic magnetic-\feld-induced\nasymmetry in the \rat dispersion asymptotes of surface magneto-plasmons, whereas our re-\nsults in Figs. 6(c,d) show an asymmetry that arises from a Doppler-shift-induced tilt of the\ndispersion asymptotes. Furthermore, we have again veri\fed that the current-biased system\nstill respects the constraint S12(kt;!) =S21(kt;!), which needs to be satis\fed regardless of\nthe local or nonlocal and reciprocal or nonreciprocal nature of the system.\nWe reiterate that the results and considerations above are valid as long as we operate\nin a regime where non-equilibrium e\u000bects due to the drifting electrons are negligible (the\nequilibrium \ructuation-dissipation theorem applies if !\u001dv=dg75). Indeed, the distance\nbetween the two slabs in Fig. 6 is much larger than the atomic distances considered in\nthe study of quantum friction in Ref.72(if graphene was used as the plasmonic material,\noperating at mid-infrared wavelengths, the distance in Fig. 6 would be on the order of\n100 nm, three orders of magnitude larger than in Ref.72). Moreover, while the two-body\nsystem being studied is active and could, in principle, exhibit optical gain, negative Landau\ndamping, and instabilities, as demonstrated for similar geometries in Refs.46,52, we have\nveri\fed that, in our case, these e\u000bects are negligible as the distance between layers is two\norders of magnitude larger ( dg\u001910\u00002\u0015for the case studied here, whereas dg\u001910\u00004\u0015in\nRef.46), for the same relative drift velocity. The analysis of small nonequilibrium corrections\nto our results and, more broadly, the full study of the nonequilibrium thermodynamics of\nsuch a nonreciprocal active system for smaller distances and higher velocities is potentially\na very interesting area of research and will be the subject of future investigations.\n30Conclusion\nThe possibility of breaking reciprocity using a direct drift-current bias in conducting materi-\nals has been known for some time, but is starting to receive more attention only recently. The\nexisting literature has been mostly focused on unidirectional propagation e\u000bects along the\ndirection of the current and/or on speci\fc platforms such as graphene. Here, instead, using a\n3D Green function formalism and microscopic considerations (starting from the Vlasov equa-\ntion, as further discussed in the Supplemental Material), we have studied current-induced\nnonreciprocal e\u000bects on the entire two-dimensional surface of a generic three-dimensional\nplasmonic platform, elucidating the dispersion and propagation properties of nonrecipro-\ncal SPPs and revealing a number of anomalous and extreme wave-propagation e\u000bects be-\nyond unidirectional propagation, including the possible excitation of frozen-light, steerable,\nsurface-wave beams with enhanced and localized \felds, and the presence of exceptional-point-\nlike modal transitions at in\rexion points. Since plasmonic media are inherently associated\nwith absorption losses, we have also clari\fed the impact of dissipation (due to collisions and\nLandau damping) on nonreciprocal current-biased SPPs and the associated tradeo\u000bs.\nWe have then focused on a particularly relevant example of application of these con-\ncepts, namely, the problem of breaking reciprocity for radiative heat transfer, which is usu-\nally achieved using magneto-optical materials. We have theoretically demonstrated a clear\nsignature of nonreciprocity in this context, namely, an asymmetry in the radiative heat \rux\ndensity between two planar bodies for opposite in-plane wavevectors, by relying, for the\n\frst time, on a direct drift-current bias instead of a magnetic bias. Our \fndings may open\nnew opportunities and directions toward the long-sought goal of controlling radiative heat\ntransfer without the constraints of time-reversal symmetry and reciprocity. Importantly,\nnonreciprocity in this context can be obtained with relatively low values of drift velocity,\nwhich could be practically realizable in certain high-mobility plasmonic media, such as high-\nmobility semiconductors and graphene in the THz, far-IR, and mid-IR spectral range. To\ndemonstrate the nonreciprocal e\u000bects discussed in this paper, these are much more promis-\n31ing material platforms than noble metals, such as gold and silver, which have relatively low\nelectron mobility. As an example, while in Ref.47a gold nanowire carrying a realistic elec-\ntric current on the order of tens of mA was predicted to support a modest drift velocity\nvd=c\u001910\u00006, producing very small nonreciprocal e\u000bects, if the same nanowire was made\nof high-mobility InSb, which exhibits plasmonic response in the low THz range,79drift ve-\nlocities up to vd=c\u001910\u00003would be achievable, comparable to the values considered in\nthis paper, for even smaller currents. Finally, thanks to its ultra-high electron mobility\nand mid-infrared plasmonic response, graphene is arguably the most interesting material\nfor drift-biased nonreciprocal plasmonics,46,50,54enabling drift velocities on the order of its\nFermi velocity, i.e., vd=c=300.\nIn summary, we believe that the rich physics of drift-biased plasmonic systems, which\nenable nonreciprocal, slow-light, active, tunable e\u000bects at the nanoscale, may open up novel\nopportunities, combining the advantages of plasmonics and nonreciprocal electrodynamics,\nfor both wave-guiding applications and thermal photonics.\nFunding\nThe authors acknowledge support from the Air Force O\u000ece of Scienti\fc Research with Grant\nNo. FA9550-19-1-0043 through Dr. Arje Nachman (whom we also thank for his suggestions\non this manuscript) and the National Science Foundation with Grant No. 1741694.\nReferences\n(1) F. Monticone, \\A Truly One-way Lane for Surface Plasmon Polaritons,\" Nature Pho-\ntonics 2020 , 14, 461-465.\n(2) A. K. Zvezdin and V. A. Kotov, \\Modern Magnetooptics and Magnetooptical Materi-\nals,\" (IOP publishing, Bristol 1997).\n32(3) Z. Wang, Y. D. Chong, J. D. Joannopoulos, M. Soljacic, \\Re\rectionfree One-way Edge\nModes in a Gyromagnetic Photonic Crystal,\" Phys. Rev. Lett. 2008 , 100, 013905.\n(4) Z. Wang, Y. D. Chong, J. D. Joannopoulos, M. Soljacic, \\Observation of Unidirec-\ntional Backscattering-immune Topological Electromagnetic States,\" Nature 2009 , 461,\n772{775.\n(5) Y. Hadad, B. Z. Steinberg, \\Magnetized Spiral Chains of Plasmonic Ellipsoids for One-\nway Optical Waveguides,\" Phys. Rev. Lett. 2010 , 105, 233904.\n(6) C. Sayrin, C. Junge, R. Mitsch, B. Albrecht, D. O'Shea, P. Schneeweiss, J. Volz, A.\nRauschenbeutel, \\Nanophotonic Optical Isolator Controlled by the Internal State of\nCold Atoms,\" Phys. Rev. X 2015 , 5, 041036.\n(7) S. Ali Hassani Gangaraj and F. Monticone, \\Topological Waveguiding near an Excep-\ntional Point: Defect-Immune, Slow-Light, and Loss-Immune Propagation,\" Phys. Rev.\nLett. 2018 , 121, 093901.\n(8) S. A. Hassani Gangaraj, F. Monticone, \\Do truly Unidirectional Surface Plasmon-\nPolaritons Exist?,\" Optica 2019 , 6, 9, 1158-1165.\n(9) M. I. Abdelrahman, and F. Monticone, \\Broadband and Giant Nonreciprocity at the\nSubwavelength Scale in Magneto-Plasmonic Materials,\" Physical Review B 2020 , 102,\n155420.\n(10) K. Shastri, and F. Monticone, \\Dissipation-Induced Topological Transitions in Weyl\nSemi-Metals, \" Physical Review Research 2020 , 2, 033065.\n(11) S. A. Hassani Gangaraj, F. Monticone, \\Physical Violations of the Bulk-Edge Corre-\nspondence in Topological Electromagnetics,\" Phys. Rev. Lett., 2020 , 124, 153901.\n(12) S. Pakniyat, A. M. Holmes, G. W. Hanson, S. A. Hassani Gangaraj, M. Antezza,\nM. G. Silveirinha, S. Jam, F. Monticone, \\Non-Reciprocal, Robust Surface Plasmon\n33Polaritons on Gyrotropic Interfaces,\" IEEE Trans. Antennas Propag., 2020 , 68, 5,\n3718-3729.\n(13) S. A. Hassani Gangaraj, B. Jin, C. Argyropoulos, F. Monticone, \\Broadband Field\nEnhancement and Giant Nonlinear E\u000bects in Terminated Unidirectional Plasmonic\nWaveguides,\" Phys. Rev. Applied, 2020 , 14, 054061.\n(14) S. A. Hassani Gangaraj, F. Monticone, \\Coupled Topological Surface Modes in Gy-\nrotropic Structures: Green's Function Analysis,\" IEEE Antennas and Wireless Prop-\nagation Letters 2018 , 17, 11, 1993-1997.\n(15) Z. Hayran, S. A. Hassani Gangaraj, F. Monticone, \\Topologically Protected Broadband\nRerouting of Propagating Waves around Complex Objects,\" Nanophotonics 2019 , 8,\n8, 1371-1378.\n(16) K. Gallo, G. Assanto, K. R. Parameswaran, M. M. Fejer, \\All-optical Diode in a Peri-\nodically Poled Lithium Niobate Waveguide,\" Appl. Phys. Lett. 2001 , 79, 314.\n(17) M. Soljacic, C. Luo, J. D. Joannopoulos, S. Fan, \\Nonlinear Photonic Crystal Microde-\nvices for Optical Integration,\" Opt. Lett. 2003 , 8, 28, 637-639.\n(18) I. V. Shadrivov, V. A. Fedotov, P. A. Powell, Y. S. Kivshar, N. I. Zheludev, \\Electro-\nmagnetic Wave Analogue of an Electronic Diode,\" New J. Phys. 2011 , 13, 033025.\n(19) L. Fan, J.Wang, L. T. Varhese, H. Shen, B. Niu, Y. Xuan, A. M. Weiner, M. Qi, \\An\nAll-Silicon Passive Optical Diode,\" Science 2012 , 335, 447, 447-450.\n(20) Y. Shi, Z. Yu, S. Fan, \\Limitations of Nonlinear Optical Isolators due to Dynamic\nReciprocity,\" Nat. Photonics 2015 , 9, 388, 388-392.\n(21) D. L. Sounas, J. Soric, A. Alu, \\Broadband Passive Isolators Based on Coupled Non-\nlinear Resonances,\" Nat. Electron. 2018 , 1, 113, 113-119.\n34(22) Y. Fan, J. Han, Z. Wei, C. Wu, Y. Cao, X. Yu, H. Li, \\Subwavelength Electromagnetic\nDiode: One-way Response of Cascading Nonlinear Meta-atoms,\" Appl. Phys. Lett.\n2011 , 98, 151903.\n(23) Z. Yu S. Fan, \\Complete Optical Isolation Created by Indirect Interband Photonic\nTransitions,\" Nat. Photonics 2009 , 3, 91-94.\n(24) H. Lira, Z. Yu, S. Fan, M. Lipson, \\Electrically Driven Nonreciprocity Induced by\nInterband Photonic Transition on a Silicon Chip,\" Phys. Rev. Lett. 2012 , 109, 033901.\n(25) R. Fleury, D. L. Sounas, C. F. Sieck, M. R. Haberman, A. Alu, \\Sound Isolation and\nGiant Linear Nonreciprocity in a Compact Acoustic Circulator,\" Science 2014 , 343,\n516-519.\n(26) M. Hafezi P. Rabl, \\Optomechanically Induced Non-reciprocity in Microring Res-\nonators,\" Opt. Express 2012 , 20, 7, 76-72.\n(27) M. A. Miri, F. Ruesink, E. Verhagen, A. Alu, \\Optical Nonreciprocity Based on Op-\ntomechanical Coupling,\" Phys. Rev. Appl. 2017 , 7, 064014.\n(28) N. R. Bernier, L. D. Toth, A. Koottandavida, M. A. Ioannou, D. Malz, A. Nunnenkamp,\nA. K. Feofanov, T. J. Kippenberg, \\Nonreciprocal Recon\fgurable Microwave Optome-\nchanical Circuit,\" Nat. Communication 2017 , 8, 604.\n(29) L. D. Landau, E. M. Lifshitz, L. P. Pitaevskij, \\Electrodynamics of Continuous Me-\ndia,\" 2nd ed. Elsevier: Amsterdam 2008 , 8.\n(30) S. Fan, \\Thermal Photonics and Energy Applications,\" Joule 2017 , 1, 264-273.\n(31) B. Zhao, G. Cheng, C. A. C. Garcia, P. Narang, S. Fan, \\Axion-Field-Enabled Nonre-\nciprocal Thermal Radiation in Weyl Semimetals,\" Nano Lett. 2020 , 20, 3, 1923-1927.\n(32) P. T. Landsberg G. Tonge, \\Thermodynamic Energy Conversion E\u000eciencies,\" J. Appl.\nPhys. 1980 , 51, R1.\n35(33) H. Ries, \\Complete and Reversible Absorption of Radiation,\" Appl. Phys. B 1983 , 32,\n153.\n(34) M. A. Green, \\Time-asymmetric Photovoltaics,\" Nano Lett. 2012 , 12, 11, 5985-5988.\n(35) W. Li, S. Buddhiraju, and S. Fan, \\Thermodynamic Limits for Simultaneous Energy\nHarvesting from the Hot Sun and Cold Outer Space, \" Light Sci. Appl. 2020 , 9.\n(36) P. Ben-Abdallah, \\Photon Thermal Hall E\u000bect,\" Phys. Rev. Lett. 2016 , 116, 084301.\n(37) L. Zhu, S. Fan, \\Persistent Directional Current at Equilibrium in Nonreciprocal Many-\nBody Near Field Electromagnetic Heat Transfer,\" Phys. Rev. Lett. 2016 , 117, 134303.\n(38) L. Fan, Y. Guo, G. T. Papadakis, B. Zhao, Z. Zhao, S. Buddhiraju, M. Orenstein,\nS. Fan, \\Nonreciprocal Radiative Heat Transfer Between Two Planar Bodies,\" Phys.\nRev. B 2020 , 101, 085407.\n(39) L. Zhu S. Fan, \\Near-complete Violation of Detailed Balance in thermal Radia-\ntion,\" Phys. Rev. B 2014 , 90, 220301(R).\n(40) B. Zhao, Y. Shi, J. Wang, Z. Zhao, N. Zhao, S. Fan, \\Nearcomplete Violation of\nKirchho\u000b's Law of Thermal Radiation with a 0.3 Tesla Magnetic Field,\" Opt. Lett.,\n2019 , 44, 17, 4203-4206.\n(41) D. A. B. Miller, L. Zhu, S. Fan, \\Universal Modal Radiation Laws for All Thermal\nEmitters,\" Proc. Natl. Acad. Sci. U. S. A. 2017 , 114, 4336-4341.\n(42) Y. Hadad, J. C. Soric, A. Alu, \\Breaking Temporal Symmetries for Emission and\nAbsorption,\" Proc. Natl. Acad. Sci. 2016 , 113, 3471.\n(43) M. Sabbaghi, H.-W. Lee, T. Stauber, K. S. Kim, \\Drift-induced Modi\fcations to the\nDynamical Polarization of Graphene,\" Phys. Rev. B 2015 , 92, 195429.\n36(44) D. S. Borgnia, T. V. Phan, L. S. Levitov, \\Quasi-Relativistic Doppler E\u000bect and Non-\nReciprocal Plasmons in Graphene,\" arXiv: 1512.09044, 2015 .\n(45) B. Van Duppen, A. Tomadin, A. N. Grigorenko, M. Polini, \\Current-induced Birefrigent\nAbsorption and Non-reciprocal Plasmons in Graphene,\" 2D Mater. 2016 , 3, 015011.\n(46) T. A. Morgado, M. G. Silveirinha, \\Negative Landau Damping in Bilayer\nGraphene,\" Phys. Rev. Lett. 2017 , 119, 133901.\n(47) K. Y. Bliokh, F. J. Fortuno, A. Y. Bekshaev, Y. S. Kivshar, F. Nori, \\Electric Current\nInduced Unidirectional Propagation of Surface Plasmon-Polaritons,\" Opt. Lett. 2018 ,\n43, 963.\n(48) T. A. Morgado, M. G. Silveirinha, \\Drift-induced Unidirectional Graphene Plas-\nmons,\" ACS Photonics 2018 , 5, 11, 4253-4258.\n(49) T. Wenger, G. Viola, J. Kinaret, M. Fogelstrom, P. Tassin, \\Current-controlled Light\nScattering and Asymmetric Plasmon Propagation in Graphene,\" Phys. Rev. B 2018 ,\n97, 085419.\n(50) D. Correas-Serrano and J. S. Gomez-Diaz, \\Nonreciprocal and Collimated Surface Plas-\nmons in Drift-biased Graphene Metasurfaces,\" Phys. Rev. B 2019 , 100, 081410(R).\n(51) T. A. Morgado, M. G. Silveirinha, \\Nonlocal E\u000bects and Enhanced Nonreciprocity in\nCurrent-driven Graphene Systems,\" Phys. Rev. B 2020 , 102, 075102.\n(52) T. A. Morgado, M. G. Silveirinha, \\Active Graphene Plasmonics with a Drift-current\nBias,\" ACS Photonics 2021 , 8, 4, 1129{1136.\n(53) F. R. Prudencio, M. G. Silveirinha, \\Asymmetric Electron Energy Loss in Drift-Current\nBiased Graphene,\" Plasmonics 2021 ,16, 19-26.\n(54) Y. Dong, L. Xiong, I. Y. Phinney, Z. Sun, R. Jing, A. S. McLeod, S. Zhang, S. Liu,\nF. L. Ruta, H. Gao, Z. Dong, R. Pan, J. H. Edgar, P. Jarillo-Herrero, L. S. Levitov,\n37A. J. Millis, M. M. Fogler, D. A. Bandurin, D. N. Basov, \\Fizeau Drag in Graphene\nplasmonics,\" Nature 2021 , 594, 513-516.\n(55) E. M. Lifshitz, L. P. Pitaevskii, \\Physical Kinetics,\" Butterworth-Heinemann, 1981 .\n(56) L. Novotny, B. Hecht, \\Principles of Nano-Optics,\" Cambridge: Cambridge University\nPress 2006 .\n(57) D. Svintsov and V. Ryzhii, \\Comment on Negative Landau Damping in Bilayer\nGraphene, \" Phys. Rev. Lett. 2019 , 123, 219401.\n(58) D. Svintsov, \\Emission of Plasmons by Drifting Dirac Electrons: A Hallmark of Hy-\ndrodynamic Transport,\" Phys. Rev. B 2019 , 100, 195428.\n(59) T. A. Morgado, M. G. Silveirinha, \\Morgado and Silveirinha Reply,\" Phys. Rev. Lett.\n2019 , 123, 219402.\n(60) J. A. Bittencourt, \\Fundamentals of plasma physics,\" 3rd ed., Springer-Verlag, 2010 .\n(61) M.A. Miri, A. Alu, \\Exceptional Points in Optics and Photonics,\" Science 2019 , 363,\n7709.\n(62) S. A. Hassani Gangaraj, G. W. Hanson, M. G. Silveirinha, K. Shastri, M. Antezza, F.\nMonticone, \\Unidirectional and Di\u000bractionless Surface Plasmon Polaritons on Three-\ndimensional Nonreciprocal Plasmonic Platforms,\" Phys. Rev. B 2019 , 99, 245414.\n(63) G. W. Hanson and A. B. Yakovlev, \\Investigation of Mode Interaction on Planar Di-\nelectric Waveguides with Loss and Gain,\" Radio Science, 1999 , 34, 1349-1359.\n(64) G. W. Hanson and A. B. Yakovlev, \\An analysis of Leaky-wave Dispersion Phenomena\nin the Vicinity of Cuto\u000b using Complex Frequency Plane Singularities,\" Radio Science,\n1998 , 33, 803-820.\n38(65) A. B. Yakovlev, G. W. Hanson, \\Fundamental Modal Phenomena on Isotropic and\nAnisotropic Planar Slab Dielectric Waveguides,\" IEEE Trans. Antennas Propag., 2003 ,\n51, 888-897.\n(66) M. Y. Nada, Mohamed A. K. Othman, F. Capolino, \\Theory of Coupled Resonator\nOptical Waveguides Exhibiting High-order Exceptional Points of Degeneracy,\" Phys.\nRev. B 2017 , 96, 184304.\n(67) G. W. Hanson, A. B. Yakovlev, M. A. K. Othman, F. Capolino, \\Exceptional Points of\nDegeneracy and Branch Points for Coupled Transmission Lines|Linear-Algebra and\nBifurcation Theory Perspectives,\" IEEE Trans. Antennas Propag., 2019 , 67, 2.\n(68) A. Figotin, I. Vitebskiy, \\Frozen Light in Photonic Crystals with Degenerate Band\nEdge,\" Phys. Rev. E, 2006 , 74, 066613.\n(69) S. Buddhiraju, Y. Shi, A. Song, C. Wojcik, M. Minkov, I. A.D. Williamson, A. Dutt,\nS. Fan, \\Absence of Unidirectionally Propagating Surface Plasmon-polaritons at Non-\nreciprocal Metal-dielectric Interfaces,\" Nat. Commun. 2020 , 11, 674.\n(70) P. Lampariello, F. Frezza, and A. A. Oliner, \\The Transition Region between Bound-\nwave and Leaky-wave Ranges for a Partially Dielectric-loaded Open Guiding Struc-\nture,\" IEEE Trans. Microwave Theory Tech. 1990 , 38, 1831.\n(71) H. Shigesawa and M. Tsuji, A. A. Oliner, \\The Nature of the Spectral Gap be-\ntween Bound and Leaky Solutions when Dielectric Loss is Present in Printed-circuit\nLines,\" Radio Science 1993 , 28, 6, 1235-1243.\n(72) J. B. Pendry, \\Shearing the vacuum { quantum friction,\" J. Phys.: Condens. Matter\n1997 , 9, 10301.\n(73) M. G. Silveirinha, \\Theory of Quantum Friction,\" New J. Phys., 2014 , 16, 063011.\n39(74) L. Landau, E. Lifshitz, and L. Pitaevskii, \\Statistical Physics,\" Pergamon: New York,\n1980 .\n(75) D. Reiche, F. Intravaia, J.-T. Hsiang, K. Busch, B. L. Hu, \\Nonequilibrium Thermo-\ndynamics of Quantum Friction,\" Phys. Rev. A, 2020 , 102, 050203(R).\n(76) G. Kirchho\u000b, \\Ueber das Verh altniss Zwischen dem Emissionsverm ogen und dem Ab-\nsorptionsverm ogen der K orper f ur W arme und Licht,\" Ann. Phys. 1860 , 185, 275.\n(77) M. Planck, \\The Theory of Heat Radiation,\" AIP-Press: New York, 1989 .\n(78) J. R. Howell, M. P. Menguuc, R. Siegel, \\Thermal Radiation Heat Transfer,\" 6th ed.;\nCRC Press: London, 2016 .\n(79) Hrostowski, H. J.; Morin, F. J.; Geballe, T. H.; Wheatley, G. H., \\Hall e\u000bect and\nconductivity of InSb, \" Phys. Rev. 1955 , 100, 1672-1676.\n40" }, { "title": "2112.00752v2.Numerical_Study_of_Cosmic_Ray_Confinement_through_Dust_Resonant_Drag_Instabilities.pdf", "content": "MNRAS 000, 000–000 (0000) Preprint 25 April 2022 Compiled using MNRAS L ATEX style file v3.0\nNumerical Study of Cosmic Ray Confinement through Dust\nResonant Drag Instabilities\nSuoqing Ji ( 季索清)12, Jonathan Squire3and Philip F. Hopkins2\n1Astrophysics Division & Key Laboratory for Research in Galaxies and Cosmology, Shanghai Astronomical Observatory, Chinese Academy of Sciences,\nShanghai 200030, China. E-mail:suoqing@shao.ac.cn\n2TAPIR & Walter Burke Institute for Theoretical Physics, Mailcode 350-17, California Institute of Technology, Pasadena, CA 91125, USA.\n3Physics Department, University of Otago, 730 Cumberland St., Dunedin 9016, New Zealand\nABSTRACT\nWe investigate the possibility of cosmic ray (CR) confinement by charged dust grains through resonant drag instabilities\n(RDIs). We perform magnetohydrodynamic particle-in-cell simulations of magnetized gas mixed with charged dust and\ncosmic rays, with the gyro-radii of dust and GeV CRs on \u0018AUscales fully resolved. As a first study, we focus on one\ntype of RDI wherein charged grains drift super-Alfvénically, with Lorentz forces strongly dominating over drag forces.\nDust grains are unstable to the RDIs and form concentrated columns and sheets, whose scale grows until saturating\nat the simulation box size. Initially perfectly-streaming CRs are strongly scattered by RDI-excited Alfvén waves, with\nthe growth rate of the CR perpendicular velocity components equaling the growth rate of magnetic field perturbations.\nThese rates are well-predicted by analytic linear theory. CRs finally become isotropized and drift at least at \u0018𝑣Aby\nunidirectionalAlfvénwavesexcitedbytheRDIs,withauniformdistributionofthepitchanglecosine 𝜇andaflatprofile\noftheCRpitchanglediffusioncoefficient 𝐷𝜇𝜇around𝜇=0,withoutthe“ 90\u000epitchangleproblem.”WithCRfeedback\non the gas included, 𝐷𝜇𝜇decreases by a factor of a few, indicating a lower CR scattering rate, because the backreaction\non the RDI from the CR pressure adds extra wave damping, leading to lower quasi-steady-state scattering rates. Our\nstudydemonstratesthatthedust-inducedCRconfinementcanbeveryimportantundercertainconditions,e.g.,thedusty\ncircumgalactic medium around quasars or superluminous galaxies.\nKey words: cosmic rays — plasmas — methods: numerical — MHD — galaxies: active — ISM: structure\n1 INTRODUCTION\nThe transport physics of cosmic rays (CRs) has been a subject of\nactive investigation since the 1960s (Kulsrud & Pearce 1969). As\ncharged ultra-relativistic particles coupled to magnetic fields via\nLorentz forces, CRs are fundamentally governed by particle-wave-\ninteractionswithmagneticfluctuations(e.g.,Alfvénwaves).Gener-\nally speaking, when Alfvén waves are excited with wavelengths\nbroadly similar to the CR gyro radii, CRs are scattered toward\nisotropy in the wave frame from small-scale irregularities of field\nlinesandthusbecome“confined”(thenetdrift/streamingspeedrel-\native to the plasma is suppressed). The relevant confining Alfvén\nwaves can be excited by the CR streaming instability when the CR\ndrift velocity 𝑣𝐷exceeds local Alfvén velocity 𝑣A\u0011j𝑩j√︁4𝜋𝜌gas\n(Kulsrud & Pearce 1969), and/or by extrinsic turbulence (Skilling\n1971;Jokipii1966).Ontheotherhand,CRsarelessconfinedwhen\nAlfvén waves are strongly damped, via ion-neutral damping, non-\nlinear Landau damping (Lee & Völk 1973) or through the magne-\ntohydrodynamic(MHD)turbulencecascade(Yan&Lazarian2002;\nFarmer & Goldreich 2004). Therefore, studying the excitation and\ndamping mechanisms of Alfvén waves around these gyro-resonant\nscales is crucial to understanding of CR transport.\nRecently, a number of numerical studies have attempted to\nmodel the effects of CRs around energies of \u0018GeV(which dom-\ninate the CR energy density) on ISM and galactic scales. These\nstudies suggest that CRs can have a significant influence on galac-\ntic “feedback processes” regulating star and galaxy formation (e.g.,\nPakmoretal.2016;Ruszkowskietal.2017;Farberetal.2018;Chan\net al. 2019; Hopkins et al. 2019; Buck et al. 2020; Su et al. 2020)\nand the phase structure and nature of the circumgalactic medium\n(CGM) (e.g., Salem et al. 2016; Butsky & Quinn 2018; Ji et al.\n2020, 2021). Meanwhile, classical models of Galactic CR trans-\nport which compare to Solar System CR experiments (Strong &\nMoskalenko 2001; Jóhannesson et al. 2016; Evoli et al. 2017) have\nsuggested the potential for new breakthroughs in particle physics\nfromdetailedmodelingofratiosofe.g.secondary-to-primaryratiosinCRpopulations.However,inallofthesestudiesafluidorFokker-\nPlanck description of CR transport is required, usually with some\nsimple assumption that CRs have a constant “diffusion coefficient”\nor“streamingspeed”oreffectivescatteringrate.Butthisintroduces\nsignificant uncertainties, as the micro-physical behavior and even\nthequalitativephysicaloriginofthesescatteringrates(andtherefore\ntheir dependence on plasma properties) remains deeply uncertain.\nThis uncertainty is vividly demonstrated in Hopkins et al. (2020a):\ncomparing different CR transport scalings based on local plasma\npropertiesaccordingtodifferentproposedscattering-ratemodelsin\nthe literature, Hopkins et al. (2020a) showed that existing models\ncould(1)differbyfactorslargerthan \u0018106intheirpredictedeffec-\ntive transport coefficients, and (2) even models with similar mean\ndiffusion coefficients could produce (as a consequence of different\ndetaileddependenceonplasmaproperties)qualitativelydifferentCR\ntransportoutcomesandeffectsonISMandCGMproperties.There-\nfore,itisparticularlyimportanttogobeyondsimplefluidtreatments\nofCRsbyinvestigatingexplicitCRscatteringphysicsanddynamics\non the scales of the CR gyro-radius 𝑟gyrocr(\u0018AUforGeVCRs in\ntypical Solar-neighborhood ISM conditions), where the interaction\nbetween CRs and magnetic fluctuations can be fully resolved.\nOne potentially important piece of physics on these scales,\nwhich has been almost entirely neglected in the historical CR lit-\nerature, is the role of dust grains. Recently, Squire et al. (2020)\nnoted a remarkable coincidence: under a broad range of ISM con-\nditions, the gyro-radii of charged dust grains in the ISM (with sizes\n\u001810\u00003\u00001𝜇𝑚) and\u001801\u000010GeV CRs overlap. While the grains\nhave much lower charge-to-mass ratios, they also have much lower\nvelocities, giving nearly coincident gyro radii. As a result, Squire\net al. (2020) proposed that dust grains can influence \u0018GeVCR\ntransport on gyro-resonant ( \u0018AU) scales in two ways: (1) as iner-\ntial particles, dust can damp parallel Alfvén waves excited by the\nCR streaming instability and thus reduce CR confinement, and (2)\ndust can be unstable to the so-called “resonant drag instabilities”\n(RDIs; Squire & Hopkins 2018), a recently-discovered, formally\n©0000 The AuthorsarXiv:2112.00752v2 [astro-ph.HE] 22 Apr 20222Ji et al.\ninfinitely-large family of instabilities that appear in different forms,\nwhenever dust grains move through a background fluid with some\nnon-vanishing difference in the force acting on dust vs. fluid. The\nRDIs can excite small-scale parallel Alfvén waves in magnetized\ngas (Hopkins & Squire 2018; Seligman et al. 2019), which can in\nturn scatter CRs and enhance their confinement. Which of these is\nthe dominant process, or whether still different forms of the RDIs\ncan be excited, depend on the local plasma conditions, with sce-\nnario(s) being more likely in regions where the gas is relatively\ndiffuse and external acceleration of dust grains is relatively strong\n(from e.g. absorbed photon momentum around a bright source). In\nanycase,theco-existenceofCRsandchargeddustgrainsintheISM\nis self-evident. Even in the CGM far from galaxies, where Squire\netal.(2020)suggestscenario(2)wouldbemorelikely,theexistence\nof significant dust populations is theoretically and observationally\nplausible: for instance, in the cool ( \u0018104\u0000105K) or warm-hot\n(105\u0000106K) phases of the CGM where the gas number density\nis low enough ( .10\u00003cm\u00002), the hadronic and Coulomb losses of\nCRsaresmall(Guo&Oh2008),andthetemperatureanddensityare\nlowenoughthatthedustsputteringtimeislong(Tielensetal.1994).\nInfact,asignificantamountofdusthasalreadybeenobservedinthe\nCGM (Ménard et al. 2010; Peek et al. 2015). And around quasars\nor superluminous galaxies, dust grains could in principle easily be\naccelerated by radiation pressure and become unstable to the spe-\ncificRDIswhichwouldsourcestrongCRscatteringon \u0018AUscales\naccording to Squire et al. (2020). Dust clumping can scatter and\nre-emitphotonsfromQSOs(e.g.,Hennawietal.2010;Martinetal.\n2010), and thus might produce observables in absorption lines of\nthe CGM. Nevertheless, this scenario remains largely unexplored;\neven for the pure dust RDIs without CRs, the only simulation study\nthus far, in Hopkins et al. (2020b), has simulated the CGM cases\nat a resolution of\u001810 pc, far above AUscales on which CR-dust\ncoupling occurs.\nMotivated by the above, in this paper we present the very first\nnumerical study of CR-dust-MHD coupling, with resolved CR and\ndust gyro-radii. In particular, here we focus on the proposed CR\nconfinement scenario in which CRs are scattered by magnetic field\nirregularities caused by the specific forms the RDI described in\nSquire et al. (2020). The paper is organized as follows. In §2, we\ndescribe the numerical methods and setup of our simulations. §3\npresents and analyzes the simulation results. We finally summarize\nour findings in §4.\n2 METHODS & SIMULATION SETUP\n2.1 Numerical Methods\nThe numerical methods adopted here consist of three main com-\nponents, each of which has been well-studied separately: the MHD\nsolverinthecode GIZMO(Hopkins2015),dustgrainsevolvedwith\na standard super-particle method (Hopkins & Squire 2018; Selig-\nman et al. 2019), and CR particles evolved with a hybrid MHD\nparticle-in-cell (MHD-PIC) method (Bai et al. 2015, 2019). De-\ntailed implementations are described in the above-cited references,\nand we only briefly review them here.\nThe background plasma/fluid/gas and magnetic fields follow\nthe equations of ideal MHD, solved in GIZMOwith the well-tested\nconstrained-gradient meshless finite-volume Lagrangian Godunov\nmethod (Hopkins & Raives 2016; Hopkins 2016). This has been\nshown to accurately reproduce a variety of detailed MHD phenom-\nena including amplification, shocks, detailed structure of the mag-\nnetorotationalandmagnetothermalinstabilities,andmore(Hopkins\n2017; Deng et al. 2019; Grudić et al. 2020). To this we add the\n“back-reaction” or feedback force of CRs and dust on gas, detailed\nbelow.IndividualgrainsandCRsareintegratedasPIC-like“super-particles” as is standard in the literature for both (Carballido et al.\n2008;Johansenetal.2009;Bai&Stone2010;Panetal.2011;McK-\ninnon et al. 2018; Bai et al. 2015, 2019; Holcomb & Spitkovsky\n2019; van Marle et al. 2019), sampling the distribution function\nstatistically by taking each super-particle to represent an ensemble\nof identical micro-particles (individual grains or CRs). Dust grains\nobey the equation of motion\n𝑑𝒗dust\n𝑑𝑡=𝒂extdust¸𝒂gasdust (1)\nwhere 𝒗dustis the grain velocity, 𝒂extdustand𝒂gasdustare grain\naccelerationsfromexternalforces(e.g.gravity,radiation)andback-\ngroundgasandmagneticfieldsrespectively.Thelatterincludesboth\ndrag and Lorentz forces:\n𝒂gasdust=\u0000𝒗dust\u0000𝒗gas\n𝑡𝑠dust\u0000\u0000𝒗dust\u0000𝒗gas\u0001\u0002ˆ𝑩\n𝑡𝐿dust(2)\nwhere ˆ𝑩=𝑩j𝑩jis the unit vector of the magnetic field 𝑩,𝑡𝑠dust\nthe dust drag drag or stopping time and 𝑡𝐿dustthe dust Larmor or\ngyro time. Since we are interested in regimes with super-sonic drift\nand microscopic dust grains (the Epstein drag limit), the drag and\ngyro timescales are given by:\n𝑡𝑠dust\u0011√︂𝜋𝛾\n8𝜌𝑖\ndust\n𝜌gas𝜖𝑑\n𝑐𝑠 \n1¸9𝜋𝛾\n128\f\f𝒗dust\u0000𝒗gas\f\f2\n𝑐2𝑠!\u000012\n(3)\n𝑡𝐿dust\u0011𝑚dust𝑐\n𝑞dust𝑩=4𝜋𝜌𝑖\ndust𝜖3\ndust𝑐\n3𝑒j𝑍dust𝑩j (4)\nwhere𝛾istheusualadiabaticindexofthegas,and 𝜌𝑖\ndust,𝜖dust,𝑚dust\nand𝑞dust\u0011𝑍dust𝑒are the internal density, radius, mass and charge\nof dust grains respectively.\nSimilarly, the equation of motion for CRs is:\n\u0010𝑐\n˜𝑐\u0011𝑑𝒖cr\n𝑑𝑡=𝒂extcr¸𝒂gascr=𝒂extcr¸\u0000𝒗cr\u0000𝒗gas\u0001\u0002ˆ𝑩\n𝑡𝐿0cr(5)\nwhere𝑐is the speed of light (and see below for ˜𝑐), and 𝒗cris the\nvelocityoftheCRparticles,whichisrelatedtotheCRfour-velocity\n𝒖cr\u0011𝛾𝐿𝒗crvia the usual Lorentz factor 𝛾𝐿\u0011¹1\u0000j𝒗crj2𝑐2º\u000012.\nThe𝑡𝐿0cr\u00111Ωcr0is the usual non-relativistic CR gyro time\n𝑡𝐿0cr\u0011¹𝑚cr𝑐º¹𝑞cr𝑩º, where𝑚crand𝑞crare the CR mass and\ncharge.\nIn the ideal MHD approximation, the “feedback” force from\ndust grains and CRs appears simply in the gas momentum equation\nas an equal-and-opposite force as from gas onto dust+CRs:\n𝜌gas\u0012𝜕\n𝜕𝑡¸𝒗gas\u0001r\u0013\n𝒗gas=\u0000r𝑃\u0000𝑩\u0002¹r\u0002 𝑩º\n4𝜋\n\u0000∫\n𝑑3𝒗dust𝑓dust¹𝒙𝒗dustº𝒂gasdust¹𝒗dustº\n\u0000∫\n𝑑3𝒗cr𝑓cr¹𝒙𝒗crº𝒂gascr¹𝒗crº\n(6)\nwhere𝑓dust¹𝒙𝒗dustºand𝑓cr¹𝒙𝒗crºare the phase-space density\ndistributions of dust and CRs respectively. Non-Lorentz forces are\nintegrated with a semi-implicit scheme and Lorentz forces using\na Boris integrator, with the back-reaction terms implemented in a\nmanner that ensures manifest machine-accurate total momentum\nconservation. These methods have been detailed and extensively\ntested in Hopkins & Lee (2016); Lee et al. (2017); Moseley et al.\n(2019); Seligman et al. (2019); Hopkins et al. (2020b).\nNotethatuptothedetailedformofthegyroaccelerationequa-\ntion, the expressions for dust and gas evolution are functionally\nidentical–indeedwecannumericallythinkofthedustasasecond,\n“heavy” (low charge-to-mass, non-relativistic) CR species which\nMNRAS 000, 000–000 (0000)CR Confinement by Dust 3\nalsoexperiencesdrag,orthinkoftheCRsas“relativistic,drag-free\ngrains.”\nFinally,toavoidtheCourantconditionfortheCRspeed 𝑐lead-\ning to computationally impractical timesteps in our simulation, we\nadopt the reduced speed-of-light (RSOL) approximation by defin-\ning the RSOL ˜𝑐in Eq. (5) ˜𝑐 𝑐(but still keeping ˜𝑐much larger\nthan other velocities in our simulation). As shown in Ji & Hop-\nkins (2021), the particularly form of the RSOL implemented here,\nwhich simply modifies the CR acceleration by the power of ˜𝑐𝑐\nand defines the CR advection speed over the (non-relativistic) grid\nto be 𝒗advectcr\u0011𝜕𝒙cr𝜕𝑡=¹˜𝑐𝑐º𝒗cr, ensures that all steady-state\npropertiesoftheCRdistributionfunction(number,momentum,en-\nergy density, pitch-angle distribution) are mathematically invariant\nto the choice of ˜𝑐. With this new implementation, we are able to\nsimultaneouslymatchtheCRenergydensity,massdensity, andmo-\nmentumdensityofanydesiredinitialconfiguration,andaccountfor\nthe“correct”CRgyro-radiusandback-reactionforceswhicharealso\nindependentofthechoiceof ˜𝑐.Thisalsomeansthatthroughoutthis\npaperwhenwerefertoe.g.CRgyro-frequenciesandotherstandard\nquantities,theyhavetheirusualmeaningandvalues(i.e.unlessoth-\nerwise specified, the RSOL ˜𝑐does not enter our expressions here).\n2.2 Simulation Parameters & Motivation\nIn what follows, we consider a system where dust grains are accel-\nerated by some large 𝒂extdust, for example from radiation pressure\nfrome.g.aquasarorluminousstarburstintheirhostgalaxy,inlow-\ndensity magnetized gas representative of e.g. the CGM or ionized\nISM bubbles. In these cases the fastest growing of the dust RDIs\nare generally the “aligned cosmic-ray like” (so named because the\neigenmodestructurebroadlyresemblestheBell2004instability)or\n“dust gyro-resonant” RDIs, which lead to rapid growth of parallel\nAlfvén waves which can scatter CRs. For the cases of interest, the\nexternal forces on the CRs 𝒂extcrare negligible.\nBecause this is a first study, to simplify the dynamics as much\nas possible and have a well-defined, resolved CR and dust gyro-\nradius, we consider just a single species of dust and single species\nofCRs.Inrealityofcoursetherewillexistabroadspectrumofdust\nsizes, with different charge and mass, e.g., and likewise spectrum\nof CR energies; but we will focus on parameters representative of\nthe grains that contain most of the dust mass and dominate the\n“feedback”force,aswellastheCRswhichdominatetheCRenergy\ndensity/pressure. We adopt a 3D box with a side-length of 𝐿box\nandperiodicboundaryconditions,filledwithinitiallyhomogeneous\ngas, dust, and CRs, with a uniform magnetic field whose initial\ndirectiondefinesthe 𝑧axis( 𝑩0=j𝑩0jˆ𝑧)anduniformvelocityfields\n(i.e. the initial CR and dust distribution functions are taken to be\n𝛿-functions, though we discuss relaxing this below), and impose an\nisothermal¹𝛾=1)equation-of-stateonthegas(motivatedbytypical\ncooling physics in the ISM/CGM). We adopt the same particle/cell\nnumbers for gas, dust and CRs, i.e., 𝑁gas=𝑁dust=𝑁cr=𝑁3\n1D,\nwhere𝑁1Dis the 1D resolution along each sides of the simulation\nbox,𝑁1D=64in our fiducial simulations (so the box contains\n3\u0002643elements),andwealsoperformedadditionalhigh-resolution\nsimulationswith 𝑁1D=128forconvergencetests.Weinitializethe\ndust drift velocity with its homogeneous equilibrium solution (see\nHopkinsetal.2020b),thegasvelocity 𝒗gas=0,andtheCRvelocity\n𝒗cr=𝑣crˆ𝑧, defined below.\nEvenwiththesesimplifications,writingthesimulationparam-\neters in units of the initial gas density 𝜌0g, sound speed 𝑐0𝑠and\nbox size𝐿box, our simulations require we specify ten dimension-\nless numbers: (1) The plasma beta 𝛽\u0011 ¹𝑐𝑠𝑣Aº2, where𝑣A\u0011\n𝑩√︁4𝜋𝜌gasis the Alfvén velocity; (2) The box-averaged dust-to-\ngasmassratio 𝜇dust\u0011𝜌dust𝜌gas;(3)Thedustexternalacceleration\n¯𝑎dust\u0011j𝒂extdustj𝐿box¹𝑐0𝑠º2; (4) The dust “size parameter” ¯𝜖dust\u0011𝜌𝑖\ndust𝜖dust𝜌0g𝐿box, which determines the grain drag forces; (5)\nThe dust “charge parameter” ¯𝜙dust\u00113𝑍dust𝑒¹4𝜋𝑐𝜖2\ndust¹𝜌0gasº12º,\nwhichdeterminesthegraincharge-to-massratioandLorentzforces;\n(6)Theangle cos¹𝜃dustº\u0011\f\fˆ𝑩0\u0001ˆ𝒂extdust\f\fbetweeninitial 𝐵-fielddi-\nrectionandexternaldustacceleration;(7)TheCR-to-gasmassratio\n𝜇cr\u0011𝜌cr𝜌gas(orequivalentlytheCRnumberratio 𝑛cr𝑛gas,since\nwe are interested in CR protons); (8) The CR “charge parameter”\n¯𝜙cr\u0011𝑞cr¹𝑐𝑚crº𝐿box¹𝜌0gasº12, which encodes the CR charge-to-\nmass ratio; (9) The angle cos¹𝜃crº\u0011\f\fˆ𝑩0\u0001ˆ𝒗cr\f\fbetween the initial\nmagnetic fields and CR velocity. (10) The initial CR Lorentz factor\n𝛾𝐿(or equivalently initial CR momentum 𝑝cr).\nThis forms an enormous parameter space, which is impos-\nsible to explore concisely. We therefore focus on one particular\nparameter set in this first study, motivated heuristically by the sce-\nnario proposed in Squire et al. (2020) and discussed above. Con-\nsider typical interstellar/intergalactic silicate or carbonaceous dust\n(𝜌𝑖\ndust\u00181 g cm\u00003, with𝜖dust\u001801𝜇mgrains containing most of\nthe mass, a standard ISM-like dust-to-gas ratio, and obeying the\nstandard collisional+photoelectric charge law from e.g. Draine &\nSutin 1987; Tielens 2005) and typical \u00181GeV kinetic energy CRs\nwhich dominate the CR energy density ( 𝛾𝐿\u00182protons). These\nproposesinamediumwithparameterstypicalofthewarm(volume-\nfilling) CGM outside of a galaxy (gas temperature 𝑇\u0018105K, den-\nsity𝑛\u001810\u00002cm\u00003,j𝑩j \u0018 01𝜇G, and CR-to-thermal pressure\n𝑃cr𝑃therm\u0018afew;Jietal.2020).Wewishtoresolvesomenumber\nof CR gyro radii in our box, so set 𝐿box\u001810𝑟𝐿cr.1This gives\nthe numerical parameters2𝛽\u00192\u0002103,𝜇dust\u0019001,¯𝜖dust\u0019105,\n¯𝜙dust\u001916\u0002106,𝜇cr\u001936\u000210\u00007,¯𝜙cr\u001913\u0002105,𝛾𝐿\u00192.3\nWhatremainsistheexternaldustacceleration ¯𝑎dust,whichistheul-\ntimatesourceofenergyfortheRDIs:alargervalueofthisparameter\ncorrespondstomorerapidRDIgrowthrates.Squireetal.(2020)con-\nsideredsuper-sonicdustdriftvelocities( 𝑣drift\u0011j𝒗dust\u0000𝒗gasj),such\nthatcertainRDIswereexcited,sowechoose ¯𝑎dust=12\u000210\u00003such\nthattheinitialequilibriumdustdriftvelocityis 𝑣drift\u001910𝑐0𝑠(safely\nsuper-sonic).4As discussed in Hopkins & Squire (2018); Hopkins\net al. (2020b) and Squire et al. (2020), while this is intentionally a\nsomewhat extreme choice, it is plausible given the observed radia-\ntive fluxes in CGM environments around super-luminous sources\nsuch as quasars and starburst galaxies, e.g., at \u001810\u0000100kpc from\na source with luminosity \u00181012\u00001013𝐿\f.\nIt is easy to verify that under these conditions, Lorentz forces\non dust grains strongly dominate over drag forces (by a factor of\n\u0018104), as desired for our problem of interest, and that both CRs\nanddusthaveverysimilargyro-radii, \u001806 AUinthephysicalunits\nabove.\n1More precisely, we set 𝐿box=p𝑁1D𝑟𝐿crsuch that each Larmor wave-\nlengthisresolvedwith \u0018p𝑁1Delementsandtheboxcontainsp𝑁1DLarmor\nwavelengths, which provides an optimal compromise for our purposes. For\nboxes with different resolution, we rescale the numerical parameters to cor-\nrespond to fixed physicalquantities (e.g. fixed gyro radii) while rescaling\nthe box size to follow this relation. Here and throughout, we define the\ngyro-radius specifically as the gyro radius for an equivalent circular orbit,\n𝑟𝐿cr=j𝒗crj𝑡𝐿cr.\n2For our low-resolution 𝑁1D=64boxes, the parameters which depend\nexplicitly on𝐿boxrescale to ¯𝜖dust\u001915\u0002105,¯𝜙cr\u00199\u0002104,¯𝑎dust\u0019\n85\u000210\u00004.\n3We also set reduced speed of light to ˜𝑐=005𝑐\u0019230𝑐0𝑠, ensuring it is\nlarger than other speeds in the problem, but explicitly examine convergence\nwith respect to this choice below.\n4As there is no particular reason to expect alignment between 𝑩and\n𝒂extdust, we set𝜃dust=45\u000e, but this is largely a nuisance parameter.\nMNRAS 000, 000–000 (0000)4Ji et al.\n2.3 On the Initial CR Pitch Angle Distribution\nNote that our initial condition for the CRs corresponds to all CRs\nhaving pitch angle =0(𝜇=1), i.e. free-streaming directly down\nmagnetic field lines at drift velocity 𝑣0\n𝐷\u0018𝑐with the maximum-\npossible anisotropy in the CR distribution function. This is almost\ncertainly unrealistic, but our purpose is to study how CRs would\nbe scattered away from this anisotropy by dust, hence the choice.5\nHowever one consequence of this choice is that for the parameters\nabove (with¹𝑒cr𝑣0\n𝐷º¹𝑒𝐵𝑐º¡1, where𝑒crand𝑒𝐵are the CR\nand magnetic energy densities), the non-resonant Bell instability\nwouldgrowmuchfasterthantheCRresonantinstability(Bell1978;\nHaggerty et al. 2019) and much faster than the RDIs (by a factor\n\u0018¹𝜇cr𝜇dustº12¹𝑣0\n𝐷𝑣0\ndustº¹𝑡0\n𝐿cr𝑡0\n𝐿dustº\u00001\u001d1). And indeed we\nhave verified this directly in test simulations, which also allow us\nto confirm that, in code, both the non-resonant and resonant CR\ninstabilities grow initially at their expected linear growth rates.\nA more realistic initial condition would feature a close-to-\nisotropic CR distribution function, which would strongly suppress\ntheseinstabilitiesrelativetotheRDIs,andwealsoconsiderthiscase\nbelow.Howeverforafirstsimulation,inordertoseehowCRswould\nbe scattered from an arbitrarily anisotropic distribution with 𝜇=1\n(whichisparticularlyusefulforunderstandingthescatteringphysics\nthemselves),weconsiderthe“ultra-lowCRdensity”or“test-particle\nCR” limit. This amounts to to ignoring the CR feedback on the gas\n(equivalently,takingtheCRnumberorenergydensitytobenegligi-\nbly small), i.e., dropping the last term on the right-hand side in Eq.\n(6). Thus, CR scattering via self-induced instabilities cannot occur\nand we can cleanly isolate the effects of the RDIs. We study this as\na test problem in §3.1 – §3.3. After the system reaches a saturation\nstate and CRs become more isotropized with 𝑣𝐷\u001c˜𝑐(a more re-\nalistic “initial condition” for the CRs), we turn on CR feedback and\ninvestigate its consequences in §3.4.\n3 RESULTS\n3.1 Time Evolution & Saturated States\nFig. 1 and 2 shows plots of the dust grain projections, gas density\nfluctuations, gas velocity perturbations superposed with velocity\nstreamlines,magneticfieldstrengthsuperposedwithfieldlines,and\ncosmic ray particle projections, viewed along the 𝑧-axis (the di-\nrection of the initial magnetic fields) and 𝑥-axis respectively. We\nchoose the snapshots at 𝑡=978𝑡𝐿dust,1022𝑡𝐿dust,1065𝑡𝐿dust\nand1304𝑡𝐿dust, which are representative samples spanning be-\ntween the end of the linear stage, the nonlinear stage and the satu-\nration stage. Dust grains are unstable to the RDI and non-linearly\nevolve into highly-concentrated columns along the 𝑧-axis, which\nmerge and form into fewer but thicker columns or sheets with time.\nAt the saturation stage, all dust grains form into one single col-\numn in our simulation box. This is expected since all wavelengths\nare unstable to RDIs, and the RDI growth rates decrease with in-\ncreasing wavelengths. Therefore, we see the merging of structures\nuntil the box-scale mode saturates. Compared to collisionless dust\ngrains which are strongly clumped, gas is only weakly compress-\nible, with density and velocity fluctuations at levels of \u00181%and\n10%respectively. As shown in Fig. 1 and 2, since field lines are\nstrongly stretched by dust grains, magnetic fields are amplified by\nuptooneorderofmagnitudefromtheirinitialvalues,andtheregions\nof field amplification tightly trace the location of clustered dust. At\nthe saturation stage, field lines become significantly distorted and\nwraparoundcolumnsofdustgrains,withthemagnetictensionforce\n5In contrast, because of drag, the only homogeneous stable equilibrium\nsolution for the dust grains is uniform streaming at the equilibrium drift\nvelocity, as we initialize.\u0018𝑩\u0002¹r\u0002𝛿𝑩º4𝜋balancingthedrivingforcefromthedustonthe\ngas\u0018𝜌dust𝒂extdust(Seligman et al. 2019). As illustrated in Fig. 3,\nthe morphology of the gas density, dust distribution and magnetic\nfield strength are highly correlated, and the system reaches approx-\nimately dynamical equilibrium between the gas pressure 𝑃gas, the\nmagnetic pressure\u0018j𝑩j28𝜋and the dust ram pressure 𝑃dust(es-\ntimated as the dust momentum flux across the surface of the dust\nfilament𝜌dust𝑣𝑧dust𝑣𝑥𝑦dust). Starting from a random distribution\ninitially, CRs strongly react to RDI-induced magnetic field pertur-\nbations, being scattered and transiently becoming highly clustered\ntogether with the dust before scattering leads to their returning to a\nnearly random spatial distribution again at the saturation stage (but\nnowwithanearly-isotropicpitchangledistribution,asweshowbe-\nlow). Initially the CRs react coherently small-scale modes (which\nhave𝜆\u001c𝑟𝐿CR) because CRs initially have a perfectly coherent\n(𝛿function) distribution function and magnetic fields are distorted\nby the RDIs; but as the CRs scatter, these local “concentrations”\ndisperse.\nFig. 4 shows the probability density functions (PDFs) of den-\nsities (top) and velocities / Lorentz factor (bottom) for gas (left),\ndust (middle) and CRs (right). Density PDFs are obtained by map-\nping particle mass onto grids, and velocity / Lorentz factor PDFs\nare weighted directly by particle masses. Compared with gas den-\nsities which only vary by \u001820%, the PDF of grain densities spans\nover two orders of magnitude and become highly non-Gaussian at\nthe non-linear and saturation stages, with a flat tail extending to\n𝜌dust𝜌0\ndust¡10(similar to other pure-RDI simulations in, e.g.,\nSeligmanetal.2019;Hopkinsetal.2020b).Therelativelyuniform\nPDFisqualitativelymaintainedinthehigh-resolutionrun,indicating\nthatifthereisacharacteristicclumpiness,it’snotyetbeingrecovered\nbythe simulation.TheCR densityPDFs arequalitativelysimilar to\nthedustdensityPDFsatthenonlinearstage,astheinitiallycoherent\nCRs are “dragged” by the RDIs, while at the saturation stage when\nCRsbecomefullyscattered,theCRdensityPDFrecovers itsinitial\nGaussianshape(forthereasonsabove).ThegasvelocityPDFssug-\ngest that gas is significantly accelerated by dust feedback up to rms\nvelocitiesh𝑣gas𝑐0𝑠i\u001810\u00002–10\u00001. Dust grains are gently deceler-\nated with time, and the velocity PDFs do not reach an equilibrium\nstatebytheendofthesimulation.TheCRLorentzfactorPDFsgrad-\nually broaden out with time, indicating CRs are mildly accelerated\nordeceleratedby\u00183%duringtheirinteractionwithlocalmagnetic\nfields (i.e., some “diffusive re-acceleration” effects with a non-zero\nCR momentum diffusion coefficient 𝐷𝑝𝑝, as will be discussed in\n§3.3).\n3.2 Growth Rates vs. Linear Theory\nWenextexaminetheamplificationofmagneticfields j𝑩𝑥j,j𝑩𝑦jand\nj𝑩𝑧j,andthegrowthoftheCRvelocitycomponentperpendicularto\nlocalmagneticfields 𝑣?cr(normalizedtothemagnitudeoftotalCR\nvelocityj𝒗crj).Bydefiningthepitchangle 𝜃astheanglebetweenCR\nvelocitiesandlocalmagneticfieldvectors,wesee 𝑣kcrj𝒗crj=𝜇and\n𝑣?crj𝒗crj=¹1\u0000𝜇2º12,where𝜇\u0011cos𝜃isthepitchanglecosine\nand𝑣kcrthe CR velocity component parallel with local magnetic\nfields. As shown in the top and middle panels of Fig. 5, j𝑩𝑥j,j𝑩𝑦j\nand𝑣?crgrow exponentially at almost the same growth rate until\nthey reach saturation at 𝑡\u001810𝑡𝐿dust. This indicates that, starting\nfrom highly anisotropic initial conditions with 𝑣?crj𝒗crj=0, CRs\nare strongly scattered by increasingly distorted magnetic fields due\nto the development of the dust RDI.\nTo estimate the magnitude of CR bulk drift velocity, we mea-\nsure the arithmetic mean values of the CR velocity components\n(the net drift velocity) over all CR particles at the saturation stage.\nWefindthat\n𝑣f𝑥𝑦𝑧gcr\u000b\n\u0018𝛼𝑐with𝛼fluctuatingbetween \u001810\u00004–\n10\u00002withtime,whichisthushighlyisotropiccomparedwiththeCR\nMNRAS 000, 000–000 (0000)CR Confinement by Dust 5\n0.50\n0.25\n0.000.250.50\nx0.4\n0.2\n0.00.20.4y9.78tL,dust\n0.50\n0.25\n0.000.250.50\nx10.22tL,dust\n0.50\n0.25\n0.000.250.50\nx10.65tL,dust\n0.50\n0.25\n0.000.250.50\nx13.04tL,dust\n123456\n100101102\n100101102\n100101\nDustParticleCount\n0.50\n0.25\n0.000.250.50\nx0.4\n0.2\n0.00.20.4y9.78tL,dust\n0.50\n0.25\n0.000.250.50\nx10.22tL,dust\n0.50\n0.25\n0.000.250.50\nx10.65tL,dust\n0.50\n0.25\n0.000.250.50\nx13.04tL,dust\n-10-3010-3\n-10-3010-3\n-10-2010-2\n-10-2010-2\n(ρgas−ρ0\ngas)/ρ0\ngas\n0.50\n0.25\n0.000.250.50\nx0.4\n0.2\n0.00.20.4y9.78tL,dust\n0.50\n0.25\n0.000.250.50\nx10.22tL,dust\n0.50\n0.25\n0.000.250.50\nx10.65tL,dust\n0.50\n0.25\n0.000.250.50\nx13.04tL,dust\n-10-3010-310-2\n-10-2010-2\n010-1\n010-1\n(vgas−¯vgas)/c0\ns\n0.50\n0.25\n0.000.250.50\nx0.4\n0.2\n0.00.20.4y9.78tL,dust\n0.50\n0.25\n0.000.250.50\nx10.22tL,dust\n0.50\n0.25\n0.000.250.50\nx10.65tL,dust\n0.50\n0.25\n0.000.250.50\nx13.04tL,dust\n5.96.06.16.26.36.4×103\n0.60.70.80.91.0×102\n1234×102\n0.020.040.060.080.100.12\n|B|//radicalbig\n4πP0\n0.50\n0.25\n0.000.250.50\nx0.4\n0.2\n0.00.20.4y9.78tL,dust\n0.50\n0.25\n0.000.250.50\nx10.22tL,dust\n0.50\n0.25\n0.000.250.50\nx10.65tL,dust\n0.50\n0.25\n0.000.250.50\nx13.04tL,dust\n100101\n100101\n100101\n100101\nCRParticleCount\nFigure1. Plotsof(fromtoptobottom)dustgrainprojections,gasdensityfluctuations,gasvelocityperturbationssuperposedwithvelocitystreamlines,magnetic\nfield strengths superposed with field lines, and CR particle projections, at (from left to right) 𝑡=978𝑡𝐿dust,1022𝑡𝐿dust,1065𝑡𝐿dustand1304𝑡𝐿dust,\nviewedalongthe 𝑧-axis(parallelwiththedirectionofinitialmagneticfields).DustgrainsareunstabletotheRDIandmodesgrowandmergeuntiltheysaturate\ninalargebox-scalesheetmode,whichsignificantlydistortsandamplifiesmagneticfields.CRsstronglyrespondtoandarescatteredbydust-inducedmagnetic\nfield irregularities.\nMNRAS 000, 000–000 (0000)6Ji et al.\n0.50\n0.25\n0.000.250.50\ny0.4\n0.2\n0.00.20.4z9.78tL,dust\n0.50\n0.25\n0.000.250.50\ny10.22tL,dust\n0.50\n0.25\n0.000.250.50\ny10.65tL,dust\n0.50\n0.25\n0.000.250.50\ny13.04tL,dust\n123456\n1234567\n100101\n100101\nDustParticleCount\n0.50\n0.25\n0.000.250.50\ny0.4\n0.2\n0.00.20.4z9.78tL,dust\n0.50\n0.25\n0.000.250.50\ny10.22tL,dust\n0.50\n0.25\n0.000.250.50\ny10.65tL,dust\n0.50\n0.25\n0.000.250.50\ny13.04tL,dust\n-10-3010-3\n-10-3010-3\n-10-20\n-10-2010-2\n(ρgas−ρ0\ngas)/ρ0\ngas\n0.50\n 0.25\n 0.00 0.25 0.500.4\n0.2\n0.00.20.49.78tL,dust\n0.50\n 0.25\n 0.00 0.25 0.5010.22tL,dust\n0.50\n 0.25\n 0.00 0.25 0.5010.65tL,dust\n0.50\n 0.25\n 0.00 0.25 0.5013.04tL,dust\n-10-3010-310-2\n-10-2010-2\n010-1\n010-1\n(vgas−¯vgas)/c0\ns\nyz\ny y y\n0.50\n0.25\n0.000.250.50\ny0.4\n0.2\n0.00.20.4z9.78tL,dust\n0.50\n0.25\n0.000.250.50\ny10.22tL,dust\n0.50\n0.25\n0.000.250.50\ny10.65tL,dust\n0.50\n0.25\n0.000.250.50\ny13.04tL,dust\n5.86.06.26.4×103\n6789×103\n1234×102\n0.020.040.060.080.100.120.14\n|B|//radicalbig\n4πP0\n0.50\n0.25\n0.000.250.50\ny0.4\n0.2\n0.00.20.4z9.78tL,dust\n0.50\n0.25\n0.000.250.50\ny10.22tL,dust\n0.50\n0.25\n0.000.250.50\ny10.65tL,dust\n0.50\n0.25\n0.000.250.50\ny13.04tL,dust\n123456\n12345678\n2468\n1234567\nCRParticleCount\nFigure 2. Plots as Fig. 1, but viewed along the 𝑥-axis (perpendicular to the direction of initial magnetic fields).\nMNRAS 000, 000–000 (0000)CR Confinement by Dust 7\n0.5\n0.00.5\ny0.4\n0.2\n0.00.20.4zPgas\n0.5\n0.00.5\nyPmag\n0.5\n0.00.5\nyPgas+Pmag\n0.5\n0.00.5\nyPdust\n0.5\n0.00.5\nyPgas+Pmag+Pdust\n0.9751.0001.0251.050\n10-310-210-1\n0.9751.0001.0251.050\n10-310-210-1\n0.9751.0001.0251.050\nFigure 3. Slice plots of (from left to right) gas thermal pressure 𝑃gas, magnetic pressure 𝑃mag, the sum of gas thermal pressure and magnetic pressure\n𝑃gas¸𝑃mag,dustrampressure 𝑃dustestimatedasdustmomentumfluxacrossthesurfaceofthedustfilament 𝜌dust𝑣𝑧dust𝑣𝑥𝑦dust,andthesumofallpressure\nterms𝑃gas¸𝑃mag¸𝑃dustin code units of 𝜌0gas¹𝑐0𝑠º2, at the saturation stage when 𝑡=1304𝑡𝐿dust. The sum of pressure terms is nearly a constant spatially,\nindicating the system is roughly in pressure balance.\n0.85 0.90 0.95 1.00 1.05\nρgas//angbracketleftbig\nρ0\ngas/angbracketrightbig10-410-310-210-1100PDF\n100101\nρdust//angbracketleftbig\nρ0\ndust/angbracketrightbig10-510-410-310-210-1PDF\n100101\nρcr//angbracketleftbig\nρ0\ncr/angbracketrightbig10-410-310-210-1PDF\n10-310-210-1100\nvgas/c0\ns10-310-210-1100101102PDF\n9.5 10.0 10.5 11.0\nvdust/c0\ns10-310-210-1100101102103PDF\n1.7 1.8 1.9 2.0 2.1 2.2 2.3\nγL10-310-210-1100101102103PDF\nt=0.0tL,dust t=9.8tL,dust t=10.2tL,dust t=10.9tL,dust t=12.0tL,dust t=13.0tL,dust\nFigure4. Thedensity(top)andvelocity/Lorentzfactor(bottom)PDFsofgas(left),dust(middle)andCRs(right).MostPDFsfeatureastrongasymmetry.The\ndensityPDFsofdustandCRsspanuptotwoordersofmagnitude,whilegasisonlyweaklycompressible.Gasanddustgrainsareacceleratedanddecelerated\nrespectively due to their momentum exchange, and the PDFs of CR Lorentz factor slightly broaden with time.\ninitial conditions of h𝑣𝑥cri0=h𝑣𝑦cri0=0andh𝑣𝑧cri0\u0018𝑐. Al-\nthoughthemagnitudeofthe“residual”driftvelocity√︃\nΣ𝑖𝑥𝑦𝑧h𝑣𝑖cri2\nis of the same order of magnitude with h𝑣Ai, given that we sample\ntheCRdistributionfunctionwith \u0018106particlesinourdefaultsim-\nulations,wecautionthatthisrangeof 𝛼ispreciselywhatwewould\nexpectfromMonteCarlosamplingnoiseforanintrinsicallyuniform\npitch-angle distribution. This sampling noise (which unfortunately\nconverges slowly, as 𝑁\u000012\ncr) dominates the “residual” drift velocity\nhere, thus we cannot draw a firm conclusion on the CR drift speed\nfromdirectmeasurement.However,thereisstillawaytoinvestigate\nthe CR drifting by Alfvén waves. As shown in the bottom panel of\nFig.5,the(negative)magnitudeofthegascrosshelicity \u0000𝛿𝒗gas\u0001𝛿𝑩,\nwhichisrelatedtotheasymmetryofAlfvénwaves,remainsasingle\nsignandgrowsexponentiallywithasquaredRDIgrowthratebefore\nsaturating. This indicates that the propagation of the Alfvén waves\nexcited by the RDIs is increasingly unidirectional (antiparallel to\nthe large-scale magnetic fields in this case when the cross helicity\nis negative); therefore, the CRs must at least drift at \u0018𝑣Aby the\nAlfvén waves all propagating in one direction.In Fig. 6, we plot the analytically predicted growth rates for\noursimulationparametersasafunctionofthewavenumber j𝒌j,for\na specific mode angle of ˆ𝒌\u0001ˆ𝒗dust=0.6We find that the growth\nratemeasuredfromoursimulationinFig.5replicatestheanalytical\nsolutionwell,withthegrowthratespeakingatwavenumbersaround\n𝑘𝐿box\u001810–40. This corresponds to a fastest growing wavelength\nof2𝜋𝑘\u0018016–06𝐿box, which is somewhat consistent with the\nstructures seen in Fig. 1 and 2.\nMNRAS 000, 000–000 (0000)8Ji et al.\n10-910-810-710-610-510-410-310-210-1100B\nexp(1.8t/tL,dust)\n|Bx|\n|By|\n|Bz|\n10-710-610-510-410-310-210-1100v,cr/|vcr|\nexp(1.8t/tL,dust)\n02468101214\nt/tL,dust10-2010-1810-1610-1410-1210-1010-810-610-410-2−δvgas·δB\nexp(3.6t/tL,dust)\nFigure 5. Time evolutions of averaged magnetic fields (top), the perpendic-\nularcomponentofCRvelocity(middle)andtheaveragednegativegascross\nhelicity (bottom). Both the magnetic fields and the CR perpendicular ve-\nlocity component follow almost the same analytically-predicted growth rate\n(and the squared growth rate for the cross helicity since it is the dot product\nof velocities and magnetic fields), indicating a strong correlation between\nmagnetic field fluctuations, Alfvén wave propagation and CR scattering.\n3.3 Pitch Angle Scattering & Transport/Scattering\nCoefficients\nFig.7showsthetimeevolutionofCRpitchanglePDFs(top)andthe\npitchanglediffusioncoefficients 𝐷𝜇𝜇(bottom),whichisdefinedas:\n𝐷𝜇𝜇¹𝜇0𝑡0º=\n¹𝜇\u0000𝜇0º2\u000b\n2¹˜𝑡\u0000˜𝑡0ºfor˜𝑡\u0000˜𝑡0=Δ˜𝑡not too large, (7)\nwhere𝜇0(˜𝑡0) and𝜇(˜𝑡) are initial and final pitch angle cosine (time)\nrespectively, Δ𝑡is the integration time, and we trace the change\nof pitch angles for all CR particles over a short time interval to\nkeep¹𝜇\u0000𝜇0ºsmall, following e.g., Beresnyak et al. (2011); Xu\n6We explicitly verified that the mode angle of ˆ𝒌\u0001ˆ𝒗dust=0, i.e., the wave\nvectoranddustvelocitiesarealignedoranti-aligned,givesthefastestgrowth\nrate,whichisconsistentwiththefindingsin(Seligmanetal.2019).Therefore,\nweonlyshowthegrowthratesforthisspecificmodeanglehere.Foradetailed\ndescription of calculating these growth rates, see Hopkins & Squire (2018).\n0.01 0.10 1 10 1000.0010.0100.100110\nk LboxIm(ωt L,dust )Figure 6. Analytical growth rates of RDI as a function of wave number\nj𝒌j, which perfectly predicts the growth rate measured directly from the\nsimulation as shown in Fig. 5.\n10-410-310-210-1100101PDF\nt=10.9tL,dust\nt=11.3tL,dustt=11.7tL,dust\nt=12.2tL,dustt=12.6tL,dust\nt=13.0tL,dust1.0\n 0.5\n 0.0 0.5 1.0\nµ10-410-310-210-1Dµµ(µ)/Ω0\ncr\nFigure7. TimeevolutionofPDFsoftheCRpitchanglecosine 𝜇(top),and\nthe CR pitch angle diffusion coefficient 𝐷𝜇𝜇(normalized by the initial CR\ngyro-frequency Ω0cr)asafunctionoftheCRpitchanglecosine(bottom).At\nthesaturationstage,CRsarefullyisotropizedwithauniformdistributionof\nthepitchanglecosine,andtheCRpitchanglediffusioncoefficientfeaturesa\nflat profile around 𝜇=0, without encountering the 90\u000epitch angle problem\npredicted by quasi-linear theories.\nMNRAS 000, 000–000 (0000)CR Confinement by Dust 9\n& Yan (2013). Here ˜𝑡\u0011¹˜𝑐𝑐º𝑡codedenotes the in-code time 𝑡code\nmultiplied by the factor ˜𝑐𝑐to correct for the reduced speed of\nlight (RSOL) approximation.7As shown in the top panel of Fig. 7,\nthe distribution of CR pitch angles becomes uniform at the satura-\ntion stage, indicating that CRs are nearly isotropized, with typical\n𝐷𝜇𝜇\u00180001\u0000001Ωcr(whereΩcr\u0011𝑞crj𝑩j¹𝛾𝐿𝑚cr𝑐ºis the\nusual relativistic CR gyro-frequency).\nForcomparison,ifweassumeisotropic/greyscatteringwiththe\nusual quasi-linear theory slab scattering expressions (Schlickeiser\n1989), then we would expect the average value of 𝐷𝜇𝜇to be given\nbyh𝐷𝜇𝜇i \u0018 ¹ 3𝜋16ºΩcrj𝛿𝑩j2j𝑩j2(Zweibel 2013), where 𝛿𝑩\nrepresentsthemagnitudeofmagneticfluctuationsongyro-resonant\nscales. We see in Fig. 7 that our typical 𝐷𝜇𝜇values correspond to\nj𝛿𝑩j\u001801j𝑩j, which is roughly what we see in Figs. 1-3 (in fact\nwe see slightly larger overall magnetic fluctuations, but what mat-\nters here is the gyro-resonant, parallel component, so the effective\nj𝛿𝑩j2entering the scattering-rate expressions we would expect to\nbe reduced by a factor of \u00182\u00003). Thus, at least to order of mag-\nnitude or better, the typical scattering rate we see is consistent with\nquasi-linear theory expectations.\nExamining𝐷𝜇𝜇¹𝜇º, we see that the CR pitch angle diffusion\ncoefficients saturates at a smooth distribution as a function of 𝜇\n(nearly𝜇-independent), which contradicts the usual prediction of\n𝐷𝜇𝜇¹𝜇=0º \u0018 0from quasi-linear theory (e.g., Skilling 1971;\nYan & Lazarian 2002). This prediction from quasi-linear theory is\nknownasthe 90\u000epitchangleproblem,since 𝜇=0correspondstoan\nextremely short resonance wave length which contains insufficient\nenergy to scatter CRs away from the 90\u000epitch angle, and thus CRs\nmight be naively expected to become “trapped” at 𝜃=90\u000ewithout\nbeing fully isotropized (e.g., Giacalone & Jokipii 1999; Felice &\nKulsrud 2001). However, in our simulation, since the RDI-induced\nmagnetic field perturbation 𝛿𝐵𝐵can grow up to a few 10\u00001, CRs\ncan be scattered across the 90\u000epitch angle and become isotropic\nwithoutanydifficulty,suggestingthedustRDIisefficientinexciting\nsmall-scaleparallelAlfvénwavesandconfineCRson \u0018AUscales.\nPrevious studies have suggested that resonance broadening due to\nnon-linear wave-particle interactions (e.g., Yan & Lazarian 2008;\nBai et al. 2019) or mirror scattering (e.g., Felice & Kulsrud 2001)\nmight help to avoid the 90\u000epitch angle problem, which may be\noccurring in our simulations as well.\nWith this, we can further estimate the CR parallel diffusion\ncoefficient𝜅kfrom our simulation in standard fashion (Earl 1974):\n𝜅k\u00191\n8∫1\n\u00001𝑑𝜇𝑣2cr¹1\u0000𝜇2º2\n𝐷𝜇𝜇¹𝜇º (8)\nCalculating this numerically in the saturation stage, we ob-\ntain𝜅k\u001817\u0002104𝑐0𝑠𝐿box\u0018700𝑐𝑟𝐿cr\u001807\u0002\n1027¹j𝑩j01𝜇Gº\u00001cm2s\u00001, i.e. a factor of\u00181000larger than\ntheBohmlimit.Thisvalueissignificantlylowerthantypicalvalues\nof𝜅\u00181029\u000030cm2s\u00001inferredfromSolarsystemmeasurementsor\n7Specifically, as shown in Ji & Hopkins (2021), the implementation of\nthe RSOL in our code is mathematically equivalent to taking the modified\nform of the general Vlasov equation for the CR distribution function to be:\n¹𝑐˜𝑐º𝜕𝑡𝑓cr¸𝒗cr\u0001r𝒙𝑓cr¸𝑭cr\u0001r𝒑𝑓cr=𝜕𝑡𝑓crjcoll, i.e. rescaling the time\nderivative of the distribution function in the simulation frame by ˜𝑐𝑐. This\nensuresthatoncetheCRdistributionfunctionreachessteady-state,alleffects\nof the choice of ˜𝑐 𝑐on its properties and on the plasma vanish, but also\nthat the CRs “respond” or evolve more slowly in time by a factor ˜𝑐𝑐to\nperturbations – effectively rescaling the units of time “as seen by” the CRs.\nThis is precisely what allows us to uniformly increase the CR timestep by\nthe factor𝑐˜𝑐, which is the purpose of the RSOL, but this means in Eq. 7,\nwe must rescale back to the “true” Δ˜𝑡=¹˜𝑐𝑐ºΔ𝑡codeto obtain the correct\n(˜𝑐-independent)valueof 𝐷𝜇𝜇.Weverifythis explicitlyinsimulationswith\nvaried ˜𝑐below.\n1.0\n 0.5\n 0.0 0.5 1.0\nµ10-410-310-210-1Dµµ(µ)/Ω0\ncr\nt=13.0tL,dust\nt=15.2tL,dustt=17.4tL,dust\nt=19.6tL,dustt=21.7tL,dustFigure 8. Time evolution of the CR pitch angle diffusion coefficient 𝐷𝜇𝜇\n(normalized by the initial CR gyro-frequency Ω0cr) as a function of the CR\npitch angle cosine 𝜇, as the bottom panel of Fig. 7 but with CR feedback\nturned on. CR feedback lowers the CR pitch angle diffusion coefficient by\nroughly one order of magnitude, and the 90\u000epitch angle problem does not\noccur either.\n𝛾-ray observations of Local Group galaxies (Blasi & Amato 2012;\nAmato & Blasi 2018; Chan et al. 2019; Hopkins et al. 2019), sug-\ngesting that dust-induced scattering near quasars or superluminous\ngalaxies can lead to strong CR confinement.\nIn Fig. 4, we clearly see that there is also some non-zero dif-\nfusion in CR momentum space. In quasi-linear theory again, the\neffectiveCRmomentum-spacediffusioncoefficient 𝐷𝑝𝑝istrivially\nrelated (to leading order in O¹𝑢𝑐º, where𝑢represents the back-\nground plasma velocities) to the pitch-angle-averaged h𝐷𝜇𝜇ias\nh𝐷𝑝𝑝i\u0018𝜒𝑝2cr𝑣2\nA\n𝑣2crh𝐷𝜇𝜇i (9)\nwhere the factor of 𝜒depends on the CR distribution function, and\nis\u001813fornearlyisotropicCRs(Hopkinsetal.2021b).Measuring\nh𝐷𝑝𝑝ieither “directly” (for an ensemble of CRs, as we estimated\n𝐷𝜇𝜇) or “indirectly” (by measuring the broadening of the PDF of\n𝑝cror𝛾𝐿inFig.4andcomparingtoananalyticdiffusionsolution),\nweconfirmthatthemomentum-spacediffusioncoefficientestimated\nnumerically is within tens of percent of the value one would infer\nfrom simply inserting our measured h𝐷𝜇𝜇iinto Eq. (9).\n3.4 Pitch Angle Scattering with CR Feedback\nWenowinvestigatehowCRfeedback(back-reactionfromCRsonto\ngas and magnetic fields) modifies the our previous findings. After\nthe CR population becomes nearly isotropized at 𝑡\u001813𝑡𝐿dust,\nwe continue evolving our simulation but now enable CR feedback.\nSinceatthisstagetheCRdriftvelocityismuchlessthanthespeedof\nlight(𝑣𝐷\u001c𝑐),andtheCR“initialconditions”aremorephysically\nrealistic, the CR non-resonant instability does not grow rapidly and\ndominatethesimulationbehavior(asitartificiallywouldifwebegan\nfrom𝑣𝐷\u0019𝑐with CR feedback included). Almost all of our quali-\ntative conclusions and the behaviors (in saturation) of gas, CR, and\ndust density fields remain qualitatively similar after turning on this\nback-reaction term, but quantitatively, there is some effect. Specif-\nically in Fig. 8, we plot the time evolution of the CR pitch angle\ndiffusioncoefficient 𝐷𝜇𝜇asafunctionoftheCRpitchanglecosine\n𝜇, after re-enabling the CR feedback. With CR feedback present,\n𝐷𝜇𝜇decreases somewhat (though it conserves the functional form\nof𝐷𝜇𝜇¹𝜇º)untilsaturatingatavaluesystematicallylowerbyafac-\ntor\u00184compared to that without CR feedback. This in turn implies\nMNRAS 000, 000–000 (0000)10Ji et al.\nfactor\u00184higher parallel diffusion coefficients. This is still more\nthan sufficient to keep the CRs isotropized and strongly-confined\n(andthe 90\u000epitch-angleproblemstilldoesnotappear),butitisnot\na negligible difference.\nPhysically, we showed in § 3.3 that the scattering rates fol-\nlowed approximately the quasi-linear theory expectation, 𝐷𝜇𝜇/\nj𝛿𝑩j2j𝑩j2. Thus a factor\u00184suppression of 𝐷𝜇𝜇by CR back-\nreaction corresponds to a factor of \u00182suppression of 𝛿𝑩on gyro-\nresonant scales. Indeed, we can directly verify that after we turn on\nthe CR feedback the fluctuations in the magnetic field are damped\nby roughly this factor. If the fluctuations are predominantly driven\nby the RDIs, then this change in their saturation amplitudes is not\nsurprising: recall from § 2.2, when we turn on back-reaction (i.e.\naccount for finite CR pressure effects on gas), we assume a fairly\nlargeCRpressurerelativetothermalpressure, 𝑃cr\u00185𝑃thermal(with\nboth𝑃crand𝑃thermalmuch larger than magnetic pressure). Thus if\nwe saturate as described in § 3.1 with the perturbations driven by\nthe RDIs (ultimately powered by the dust “ram pressure” or accel-\neration force per unit area 𝑃dust) compensated by gas thermal plus\nCR pressure, then we expect j𝛿𝑩j2to be a factor of a few lower\nin saturation (from the increase in the totalthermal+magnetic+CR\nbackgroundpressure).Effectively,thebackgroundmediumbecomes\n“stiffer” against perturbation by the dust.\nFromtheviewpointofthestandardquasi-lineartheorywherein\noneassumeslineargrowthofscatteringmodescompensatedbywave\ndamping settingthe quasi-steady-state scattering rates,with the CR\nfeedbackturnedon,thebackreactionontheRDIfromtheCRpres-\nsure plus the magnetic tension leads to stronger damping, and thus\nless CR confinement as found in the simulations. Quasi-linear the-\nories also predict that the CR feedback can generate small-scale\nAlfvén waves via the CR gyro-resonant and non-resonant instabili-\nties, i.e., the CR “self-confinement” modes which add linearly with\ntheRDIdrivenperturbationsandthus increasetheCRconfinement.\nHowever, this is in opposite to the decrease of the CR confinement\nin our simulations, because CRs are highly isotropic in the satu-\nration stage, thus the CR-excited Alfvén waves and their resulting\nCR confinement are negligible. But admittedly, it is plausible to\nimagine that if there is a continuous driving of the background CR\ngradients, the growth rates of Alfvén waves from dust and CRs do\nadd linearly when they both have drifts and small densities, and we\nmight see the enhancement of the CR confinement, as quasi-linear\nself-confinement theories predicted.\n3.5 Convergence Tests\nAlthoughpreviousstudieshaveinvestigatedhowsimulationsofjust\ntheRDIs(Moseleyetal.2019;Seligmanetal.2019;Hopkinsetal.\n2020b) or just CRs depend on resolution, those simulations did not\ncombine these physics nor investigate the same range of scales and\nparametersaswedohere.Therefore,weconsidera“high-resolution”\nsimulation with 𝑁gas=𝑁dust=𝑁cr=𝑁3\n1D=1283. Note that\nsinceweenforce 𝑟𝐿dust𝐿box=p𝑁1D,thehigh-resolutionrunwith\ndoubled 1D resolution fits a factor-of-p\n2larger number of Larmor\nwavelengthsalongeachsideofthesimulationdomain,andresolves\neach Larmor wavelength withp\n2times more elements.\nFig. 9 shows projections of dust grains, slices of magnetic\nfield strength superposed with field lines and projections of CR\nparticles along the 𝑧-axis for the high-resolution simulation. The\nhigh-resolutionrunisqualitativelysimilartothelow-resolutionrun\nshown in Fig. 1, but contains smaller-scale structures as expected.\nAt the saturation stage, the high-resolution run contains two large-\nscale coherent structures of dust columns and magnetic vortices,\nin contrast to the single sheet-like structure in the low-resolution\nrun. This is likely because the high-resolution run contains more\nLarmor wavelengths and thus it simulates a larger domain in realphysical units. The time evolution of the magnetic fields and the\nperpendicular CR velocity in the high-resolution run are shown\nin Fig. 10, where the growth rate is almost identical to the low-\nresolution run. Thus the linear growth rates of the relevant modes\nare not measurably dependent on resolution, and they agree well\nwith analytic theory; moreover the final properties (widths) of the\nPDFs are similar in both cases. Of greatest interest, the values of\n𝐷𝜇𝜇and(correspondingly) 𝜅kand𝐷𝑝𝑝differbylessthan 5%inthe\nhighresolutionrun.Thusitappearsthatthekeyresultsherearenot\nsensitivetoresolution.However,wecautionthatitisonlypossibleto\nnumerically resolve a tiny fraction of the interesting dynamic range\n(letaloneeffectsliketurbulentcascadesorfield-linewanderingfrom\nlarge-scale dynamics, which operate on orders-of-magnitude larger\nscales), so this should be taken with some caution.\nWe also examine the numerical convergence with respect\nto the RSOL by varying ˜𝑐over a factor of\u00185. For standard-\nresolution runs with ˜𝑐=¹00200501º𝑐, we find qualita-\ntively identical behavior in all properties studied here, with the\nnumerically-estimated in-saturation parallel diffusion coefficients\n𝜅k\u0018¹700700600º𝑐𝑟𝐿cr, respectively – i.e. nearly invariant to\n˜𝑐. Note there is a (very weak) hint that 𝜅kmay decrease with in-\ncreasing ˜𝑐here, perhaps owing to slightly stronger confinement at\nhigher ˜𝑐perhapsbecausewithlower ˜𝑐theCRsresponsemayslightly\nartificially lag the RDI growth rates; but if we extrapolate to ˜𝑐=𝑐\nby fitting our results, the resulting inferred 𝜅kis only decreased\nby\u001840%, a rather small correction compared to other theoretical\nuncertaintieshere(e.g.theeffectofback-reactiondiscussedbelow).\n3.6 Discussion: Damping & Saturation Scalings\nOwing to our limited numerical resolution and MHD-PIC assump-\ntions (where the plasma is treated as an MHD fluid), our sim-\nulations do not include certain plasma processes that can also\ndamp Alfvén waves, such as ion-neutral or Landau damping. Ion-\nneutral damping is not likely relevant under conditions of inter-\nest for our problem here (e.g. diffuse, warm, highly-illuminated\nCGM), as the expected neutral fractions are vanishingly small.\nBut Landau damping could be non-negligible, in principle. If\nwe consider e.g. the usual non-linear Landau damping rate with\nΓ\u0018¹p𝜋4º𝑐𝑠𝑘¹𝑘2\n?𝑘2\nkº\u0018¹p𝜋8º𝑐𝑠𝑘¹j𝛿𝑩j2j𝑩j2º, then at the\nscalewheretheRDIgrowthrateismaximized( \u0018𝑟𝐿cr)theimplied\nLandau damping rate (given the j𝛿𝑩j\u001801j𝑩jwe see) is generally\n\u001810%oftheRDIgrowthrateinFig.6.Thatsuggestsitmaynotbe\nnegligible, but it is also unlikely to qualitatively change the behav-\niors here, if included. However, the simulations do, given the finite\nresolution,havenon-zeronumericaldissipationwhichhappens(co-\nincidentally)tobesimilarinmagnitudetoLandaudamping:givena\nstandardnumericalMHDdissipationratein GIZMOwhichscalesas\n\u0018Δ𝑥𝑐𝑠𝑘2\u0018𝑐𝑠𝑘¹𝑘Δ𝑥º(whereΔ𝑥is the effective grid resolution\nfor the MHD, and the prefactor depends on the specific numerical\nproblem and details of the method, see Hopkins & Raives 2016),\nthen for𝑘\u00181𝑟𝐿cr\u00181¹10Δ𝑥ºthis is roughly similar in magni-\ntudetothephysicalLandaudamping.Ofcourse,othermechanisms\ncouldinprinciplecontributetodampingincludinginteractionswith\nextrinsic turbulence (Farmer & Goldreich 2004), which we cannot\ncaptureowingtoourlimitedrangeofscales.Dustitselfcouldactas\na damping mechanism in some circumstances (Squire et al. 2021)\nbuttheconditionswherethiswouldoccuraredramaticallydifferent\nfrom those here.\nThese limitations in physics, finite resolution, and simulation\nbox size preclude making detailed statements regarding the satura-\ntionmechanismsoftheRDIacrossabroaderparameterspace.How-\never, the fact that we see roughly isotropic 𝐷𝜇𝜇following approxi-\nmatelytheexpectedquasi-linearscalingwith j𝛿𝑩jj𝑩jgivesussome\nconfidence that the CR scattering rates induced by the RDI should\nMNRAS 000, 000–000 (0000)CR Confinement by Dust 11\n0.50\n0.25\n0.000.250.50\nx0.4\n0.2\n0.00.20.4y8.92tL,dust\n0.50\n0.25\n0.000.250.50\nx9.32tL,dust\n0.50\n0.25\n0.000.250.50\nx9.72tL,dust\n0.50\n0.25\n0.000.250.50\nx11.91tL,dust\n100101102\n100101102\n100101102\n100101102\nDustParticleCount\n0.50\n0.25\n0.000.250.50\nx0.4\n0.2\n0.00.20.4y8.92tL,dust\n0.50\n0.25\n0.000.250.50\nx9.32tL,dust\n0.50\n0.25\n0.000.250.50\nx9.72tL,dust\n0.50\n0.25\n0.000.250.50\nx11.91tL,dust\n5.56.06.57.0×103\n0.60.81.0×102\n12345×102\n0.020.040.060.080.100.120.14\n|B|//radicalbig\n4πP0\n0.50\n0.25\n0.000.250.50\nx0.4\n0.2\n0.00.20.4y8.92tL,dust\n0.50\n0.25\n0.000.250.50\nx9.32tL,dust\n0.50\n0.25\n0.000.250.50\nx9.72tL,dust\n0.50\n0.25\n0.000.250.50\nx11.91tL,dust\n100101\n100101\n100101102\n100101\nCRParticleCount\nFigure9. Plotsofdustgrainprojections(top),magneticfieldstrengthssuperposedwithfieldlines(middle)andCRparticleprojections(bottom),asFig.1,but\nfromthehigh-resolutionsimulation.Themorphologyofdustgrains,magneticfieldsandCRsinthehigh-resolutionrunisqualitativelysimilarwiththoseinthe\nfiducial run, but exhibits more detailed small-scale structures and contains more large-scale coherent structures (since the domain size of the high-resolution\nrun is effectively larger – see the text for a full explanation).\nindeedscalewiththesaturationamplitudeof 𝛿𝑩.Andevenifthesat-\nurationmechanismisuncertain,somebroadconclusionsarerobust.\nFor example, based on their idealized RDI-only simulations, Hop-\nkins et al. (2020b) discuss two possible saturation mechanisms, the\nfirstbeingbalancebetweenmagnetictensionanddustrampressure\nasdiscussedin§3.1,thesecondbeingascenariowherethecrossing\ntime of the RDI-generated modes ( \u0018𝛿𝒗gas𝜆) becomes faster than\nthe RDI growth time, producing non-linear dissipation. Non-linear\nLandaudampingisanotherpotentialsaturationmechanism,limiting\n𝛿𝑩whenthedampingbecomesfasterthanlinearRDIgrowthrates.\nSquireetal.(2021)consideryetanothersaturationscenario,assum-\ning a Kraichnan (1965)-like damping via self-interactions (which\ngivesadampingrateakintonon-linearLandaubutwith 𝑐𝑠replaced\nbytheAlfvénspeed).Crucially,the“driving”inallofthesescenarios\nscaleswiththedust-to-gasratio 𝑓dust\u0000gasandforce/acceleration/drift\nvelocityonthegrains(whichappearinboththedustrampressureand\nRDIgrowthrates).Thisisproportionaltotheincidentradiationflux\n𝐹rad.Combiningtheseestimatesforthesaturation 𝛿𝑩withtheusual\nscalings for gyro radii and scattering rates we obtain, for anyof the\nsaturationscenariosabove,ascatteringratewhichscalesdimension-\nallyas𝐷𝜇𝜇/𝑓dust\u0000gas𝐹07\u000015\nradj𝑩j\u0000¹08\u00003º(withaweakerresidual\ndependenceongasdensityand/ortemperature,andtheexactpower-\nlawscalingdependingonthesaturationmodel).Inotherwords,there\nisaqualitativelyrobustpredictionthatwithlower-dust-to-gas-ratiosand/or incident radiative fluxes and/or stronger magnetic fields, the\nconfinement of CRs by dust become weaker. All else equal, in or-\nder for the scattering rate from the dust RDIs to drop to below\nthat inferred for Milky Way ISM gas (and thus become relatively\nunimportant), the factor \u0018𝑓dust\u0000gas𝐹radwould need to be\u00181000\ntimessmallerthanthevalueweassumetomotivateourtests.Sofor\nthere to be “too little dust,” the metallicity or dust-to-metals ratio\nwould need to be 1000times lower than ISM values (which seems\nunlikely at least at low cosmological redshifts, even in the CGM,\ngiven the observations reviewed in § 1, which suggest this factor\nis perhaps something like \u001810times lower than in the ISM). But\nmoreplausibly,theincidentfluxcouldeasilybe 1000timessmaller,\nif for example the galaxy is a typical Milky Way-like or smaller\ndwarf galaxy with a star formation rate of \u001c10 M\fyr\u00001and has\nnegligible AGN luminosity (i.e. galaxy luminosity .1010L\f) or\nthe dust is at\u001d100kpc from the host. The magnetic field depen-\ndence is also interesting: it suggests these instabilities and ensuing\nconfinement would be easier to excite in the distant CGM or IGM\n(where nano-Gauss fields are expected), but may be suppressed in\ndenser, more highly-ionized super-bubbles near to galaxies.\n4 CONCLUSIONS\nIn this paper, we investigate the impact of the dust RDIs on cos-\nmic ray scattering, by performing the first numerical simulations of\nMNRAS 000, 000–000 (0000)12Ji et al.\n10-910-810-710-610-510-410-310-210-1100B\nexp(1.8t/tL,dust)\n|Bx|\n|By|\n|Bz|\n02468101214\nt/tL,dust10-810-710-610-510-410-310-210-1100v,cr/|vcr|\nexp(1.8t/tL,dust)\nFigure10. Timeevolutionofaveragedmagneticfields(top)andtheperpen-\ndicularcomponentofofCRvelocity,asFig.5,butfromthehigh-resolution\nsimulation.Theexponentialgrowthrateisalmostidenticaltothatinthefidu-\ncialrun(Fig.5),suggestingdifferentresolutionsdonotsignificantlyalterthe\nresults of our simulations.\nMHD-dust-CR interactions, where the charged dust and CR gyro-\nradii on\u0018AUscales are fully resolved. Since this is a first study,\nwe consider just one special case, where we might anticipate ef-\nficient dust-induced CR confinement, as compared to more typical\nSolar-neighborhood-likeISMconditions.Wefocuson oneregimeof\nthe RDIs, specifically conditions where the “cosmic ray-like” RDIs\nare rapidly growing and can produce “diffusive” dust behavior (see\nHopkins et al. 2020b), which as speculated in Squire et al. (2020)\ncould in turn lead to dust-induced CR confinement. This type of\nRDIrequiresparticularconditionstodominate: 𝑣dust\u001d𝑣A,plasma\n𝛽\u001d1,lowgasdensityandhighgraincharge,sothedustdragforce\nis substantially subdominant to the Lorentz force. These conditions\ncould arise in, e.g., the CGM around quasars or luminous galaxies.\nWe suspect that different RDIs might have different effects on CR\nscattering, which is a subject for future study. Under these condi-\ntions, we find that small-scale parallel Alfvén waves excited by the\nRDIs efficiently scatter CRs and significantly enhance CR confine-\nment.Therefore,dust-inducedCRscatteringcanpotentiallyprovide\na strong CR feedback mechanism on \u0018AUscales.\nWefirstexploretheultra-lowCRdensitylimitbyignoringCR\nfeedback to the gas. Dust quickly becomes unstable to the RDIs\nand forms high-density columns and sheets, until growth saturates\naround\u001810times the dust Larmor time 𝑡𝐿dust. The density and\nvelocity PDFs of both gas and dust grains show strong asymmetry,\nwhere the dust density spans over two orders of magnitude, while\nthe gas is only slightly compressible with \u001810%density fluctua-\ntions. Perpendicular magnetic field components are exponentially\namplified by the RDI, with the growth rates predicted by analyticalsolutions.InitiallyperfectlystreamingCRsarestronglyscatteredby\nmagnetic field fluctuations, with the growth rate of the perpendicu-\nlar CR velocity component (to local magnetic fields, 𝑣?crj𝒗crj, or√︁\n1\u0000𝜇2)equalingthegrowthrateofmagneticfields.Atthesatura-\ntionstage,theCRsareisotropizedwithanear-uniformdistributionof\nthepitchanglecosine 𝜇,andtheCRpitchanglediffusioncoefficient\n𝐷𝜇𝜇¹𝜇ºisnearlyindependentofpitch-angle 𝜇(inparticulararound\n𝜇=0).Thereisno 90\u000epitchangleprobleminoursimulations.The\nscattering rate is in order-of-magnitude agreement with the usual\nquasi-linear theory expectation 𝐷𝜇𝜇\u0018Ωcrj𝛿𝑩j2j𝑩j2, with the\nlargej𝛿𝑩j2j𝑩j2\u001810\u00003\u000010\u00002on gyro-resonant scales driven by\nthe dust RDIs (in part because the dust has broadly similar gyro-\nradiitotheCRs,underthesetypesofconditions).Thenumerically-\ncalculated CR parallel diffusion coefficient is \u0018500\u00001000times\nthe Bohm value: sufficient for strong confinement of the CRs.\nWhen the system reaches saturation and CRs become close to\nisotropic (with the CR drift velocity 𝑣𝐷\u001c˜𝑐), we turn on CR feed-\nback to the gas and study its consequences. We find that with CR\nfeedback, the CR pitch angle diffusion coefficient 𝐷𝜇𝜇decreases\nby a factor of\u00184(and thus this slightly reduces CR confinement).\nThis owes to the fact that the large assumed CR pressure (several\ntimes larger than thermal+magnetic) suppresses the saturation am-\nplitude of the magnetic field fluctuations 𝛿𝑩induced by the RDIs\nby a modest factor \u00182, and the scattering modes are thus damped\nmore by backreaction on the RDIs from the CR pressure, in ad-\ndition to the magnetic tension. Since CRs in the saturation stage\narehighlyisotropic,thequasi-linearself-confinementtheorywhich\npredictshigherscattering rates due to CR-excited Alfvén waves is\nnotapplicable here, unless there exists a continuous driving of CR\ngradients.\nWefinallystressthatseveralcaveatsapplytoourstudy.(1)Asa\nfirst experiment, we picked one particular initial condition to inves-\ntigate an interesting case, which is plausible for some conditions as\nnotedabovebutshouldnotbeconsideredtypicaleverywhere.There\nexist a variety of RDIs which can have totally different behaviors\nindifferentcircumstances(Hopkinsetal.2020b)andindeed,under\nsome conditions dust might even have the opposite effect, acting as\na wave-damping mechanism and reducing the confinement of CRs\n(Squire et al. 2020). (2) Although we perform a small resolution\nstudy, our simulations are still limited in resolution and dynamical\nrange. Even for a single grain size, the RDIs are unstable at all\nspatial wavelengths, so it is impossible to encompass their com-\npletedynamicrange(letaloneglobalscalesofstructureorextrinsic\nturbulence, which are vastly larger than our box sizes). (3) We do\nnot include any explicit wave-damping processes. While we do not\nexpect appreciable ion-neutral damping in environments of interest\n(as the neutral fractions are negligible), Landau damping could be\nimportant, as could damping from a turbulent cascade, and these\ncould reduce the efficacy of confinement. However at least for the\nextreme parameters considered here, it is unlikely these would sig-\nnificantlyreducethescatteringrates.(4)Ourperiodicboxesneglect\nlarge-scale (\u001d𝐿box) CR pressure gradients ( r𝑃cr) which act as a\nsource/driving term for super-Alfvénic CR drift. (5) For simplicity,\nweconsideronlyonegrainsize+chargeandoneCRenergy+species,\nratherthanafullspectrumofgrainsizesandCRenergies.Infuture\nwork it will be particularly interesting to see how a full spectrum\nof both modifies the dynamics here, as a broad range of gyro-radii\noverlapanddifferentgyro-resonantmodescaninteractnon-linearly\nand even linearly (when CRs+dust+gas are all combined), via their\nback-reaction on the gas.\nWith these limitations in mind, we consider this study to be a\nproofofconcept ,showingthatCRdust-gasinteractionsmightindeed\nbeveryimportantinsomeastrophysicalconditions.Forinstance,this\nmechanismmightbeabletoresolvesomeoftheincompatibilitybe-\nMNRAS 000, 000–000 (0000)CR Confinement by Dust 13\ntween CR confinement models and observations by preventing CR\n“runaway” (Hopkins et al. 2021a), at least under certain conditions\ninvestigated in this study. Considerable work remains to map out\nthe parameterspace, include additional physics,and understand the\nmacroscopic consequences of confining CR-dust interactions (for\neitherCRsthemselvesorforgalaxy/CGMevolution).Nevertheless,\nthesimulationshereclearlyarguethattheseeffectsareworthstudy-\ning in detail.\nDATA AVAILABILITY\nThe data supporting the plots within this article are available on\nreasonable request to the corresponding author. A public version\noftheGIZMOcodeisavailableat http://www.tapir.caltech.\nedu/~phopkins/Site/GIZMO.html .\nACKNOWLEDGMENTS\nSJ thanks E. Quataert for helpful discussions, and the referee\nfor constructive comments which improve this manuscript. SJ is\nsupported by a Sherman Fairchild Fellowship from Caltech, the\nNaturalScienceFoundationofChina(grants12133008,12192220,\nand 12192223) and the science research grants from the China\nMannedSpaceProject(No.CMS-CSST-2021-B02).SupportforJS\nwas provided by Rutherford Discovery Fellowship RDF-U001804\nand Marsden Fund grant UOO1727, which are managed through\nthe Royal Society Te Ap ¯arangi. Support for PFH was provided by\nNSF Research Grants 1911233 & 20009234, NSF CAREER grant\n1455342, NASA grants 80NSSC18K0562, HST-AR-15800.001-A,\nJPL 1589742. Numerical calculations were run on the Caltech\ncompute cluster “Wheeler,” allocations FTA-Hopkins/AST20016\nsupported by the NSF and TACC, and NASA HEC SMD-16-7592.\nWe have made use of NASA’s Astrophysics Data System. Data\nanalysis and visualization are made with Python 3, and its packages\nincluding NumPy(VanDerWaltetal.2011), SciPy(Oliphant2007),\nMatplotlib (Hunter 2007), and the ytastrophysics analysis software\nsuite (Turk et al. 2010).\nREFERENCES\nAmato E., Blasi P., 2018, Advances in Space Research, 62, 2731\nBai X.-N., Stone J. M., 2010, ApJ, 722, L220\nBai X.-N., Caprioli D., Sironi L., Spitkovsky A., 2015, The Astrophysical\nJournal, 809, 55\nBaiX.-N.,OstrikerE.C.,PlotnikovI.,StoneJ.M.,2019,TheAstrophysical\nJournal, 876, 60\nBellA.,1978,MonthlyNoticesoftheRoyalAstronomicalSociety,182,147\nBell A. R., 2004, MNRAS, 353, 550\nBeresnyak A., Yan H., Lazarian A., 2011, The Astrophysical Journal, 728,\n60\nBlasi P., Amato E., 2012, Journal of Cosmology and Astroparticle Physics,\n2012, 010\nBuck T., Pfrommer C., Pakmor R., Grand R. J., Springel V., 2020, Monthly\nNotices of the Royal Astronomical Society, 497, 1712\nButsky I. S., Quinn T. R., 2018, The Astrophysical Journal, 868, 108\nCarballido A., Stone J. M., Turner N. J., 2008, MNRAS, 386, 145\nChan T., Kereš D., Hopkins P., Quataert E., Su K., Hayward C., Faucher-\nGiguère C., 2019, Monthly Notices of the Royal Astronomical Society,\n488, 3716\nDengH.,MayerL.,LatterH.,HopkinsP.F.,BaiX.-N.,2019,ApJS,241,26\nDraine B. T., Sutin B., 1987, ApJ, 320, 803\nEarl J., 1974, The Astrophysical Journal, 193, 231\nEvoli C., Gaggero D., Vittino A., Di Bernardo G., Di Mauro M., Ligorini\nA., Ullio P., Grasso D., 2017, Journal of Cosmology and Astroparticle\nPhysics, 2, 015\nFarberR.,RuszkowskiM.,YangH.-Y.,ZweibelE.,2018,TheAstrophysical\nJournal, 856, 112\nFarmer A. J., Goldreich P., 2004, The Astrophysical Journal, 604, 671\nFelice G. M., Kulsrud R., 2001, The Astrophysical Journal, 553, 198Giacalone J., Jokipii J., 1999, The Astrophysical Journal, 520, 204\nGrudić M. Y., Guszejnov D., Hopkins P. F., Offner S. S. R., Faucher-\nGiguère C.-A., 2020, MNRAS, submitted, arXiv:2010.11254, p.\narXiv:2010.11254\nGuo F., Oh S. P., 2008, MNRAS, 384, 251\nHaggertyC.,CaprioliD.,ZweibelE.,2019,arXivpreprintarXiv:1909.06346\nHennawi J. F., et al., 2010, The Astrophysical Journal, 719, 1672\nHolcomb C., Spitkovsky A., 2019, ApJ, 882, 3\nHopkins P. F., 2015, MNRAS, 450, 53\nHopkins P. F., 2016, MNRAS, 462, 576\nHopkins P. F., 2017, MNRAS, 466, 3387\nHopkins P. F., Lee H., 2016, MNRAS, 456, 4174\nHopkins P. F., Raives M. J., 2016, MNRAS, 455, 51\nHopkins P. F., Squire J., 2018, Monthly Notices of the Royal Astronomical\nSociety, 479, 4681\nHopkins P. F., et al., 2019, arXiv preprint arXiv:1905.04321\nHopkinsP.F.,SquireJ.,ChanT.,QuataertE.,JiS.,KeresD.,Faucher-Giguère\nC.-A., 2020a, arXiv preprint arXiv:2002.06211\nHopkins P. F., Squire J., Seligman D., 2020b, Monthly Notices of the Royal\nAstronomical Society, 496, 2123\nHopkins P. F., Squire J., Butsky I. S., Ji S., 2021a, arXiv preprint\narXiv:2112.02153\nHopkins P. F., Squire J., Butsky I. S., 2021b, MNRAS, submitted,\narXiv:2103.10443, p. arXiv:2103.10443\nHunter J. D., 2007, Computing in science & engineering, 9, 90\nJi S., Hopkins P. F., 2021, Accurately Incorporating a Reduced-\nSpeed-of-Light in Magnetohydrodynamic-Particle-in-Cell Simulations\n(arXiv:2111.14704 )\nJi S., et al., 2020, Monthly Notices of the Royal Astronomical Society, 496,\n4221\nJi S., Kereš D., Chan T., Stern J., Hummels C. B., Hopkins P. F., Quataert\nE., Faucher-Giguère C.-A., 2021, Monthly Notices of the Royal Astro-\nnomical Society, 505, 259\nJóhannesson G., et al., 2016, ApJ, 824, 16\nJohansen A., Youdin A., Mac Low M.-M., 2009, ApJ, 704, L75\nJokipii J. R., 1966, The Astrophysical Journal, 146, 480\nKraichnan R. H., 1965, Physics of Fluids, 8, 1385\nKulsrud R., Pearce W. P., 1969, The Astrophysical Journal, 156, 445\nLee M. A., Völk H. J., 1973, Astrophysics and Space Science, 24, 31\nLee H., Hopkins P. F., Squire J., 2017, MNRAS, 469, 3532\nMartin C. L., Scannapieco E., Ellison S. L., Hennawi J. F., Djorgovski S.,\nFournier A. P., 2010, The Astrophysical Journal, 721, 174\nMcKinnon R., Vogelsberger M., Torrey P., Marinacci F., Kannan R., 2018,\nMNRAS, 478, 2851\nMénard B., Scranton R., Fukugita M., Richards G., 2010, Monthly Notices\nof the Royal Astronomical Society, 405, 1025\nMoseleyE.R.,SquireJ.,HopkinsP.F.,2019,MonthlyNoticesoftheRoyal\nAstronomical Society, 489, 325\nOliphant T. E., 2007, Computing in Science & Engineering, 9, 10\nPakmorR.,PfrommerC.,SimpsonC.M.,SpringelV.,2016,TheAstrophys-\nical Journal Letters, 824, L30\nPanL.,PadoanP.,ScaloJ.,KritsukA.G.,NormanM.L.,2011,ApJ,740,6\nPeek J., Ménard B., Corrales L., 2015, The Astrophysical Journal, 813, 7\nRuszkowskiM.,YangH.-Y.K.,ZweibelE.,2017,TheAstrophysicalJournal,\n834, 208\nSalem M., Bryan G. L., Corlies L., 2016, Monthly Notices of the Royal\nAstronomical Society, 456, 582\nSchlickeiser R., 1989, ApJ, 336, 243\nSeligman D., Hopkins P. F., Squire J., 2019, Monthly Notices of the Royal\nAstronomical Society, 485, 3991\nSkilling J., 1971, The Astrophysical Journal, 170, 265\nSquire J., Hopkins P. F., 2018, ApJ, 856, L15\nSquire J., Hopkins P. F., Quataert E., Kempski P., 2020, arXiv preprint\narXiv:2011.02497\nSquireJ.,HopkinsP.F.,QuataertE.,KempskiP.,2021,MNRAS,502,2630\nStrongA.W.,MoskalenkoI.V.,2001,AdvancesinSpaceResearch,27,717\nSu K.-Y., et al., 2020, Monthly Notices of the Royal Astronomical Society,\n491, 1190\nTielens A. G. G. M., 2005, The Physics and Chemistry of the Interstellar\nMedium. Cambridge, UK: Cambridge University Press\nTielens A., McKee C., Seab C., Hollenbach D., 1994, Astrophys J, p. 321\nMNRAS 000, 000–000 (0000)14Ji et al.\nTurk M. J., Smith B. D., Oishi J. S., Skory S., Skillman S. W., Abel T.,\nNorman M. L., 2010, The Astrophysical Journal Supplement Series,\n192, 9\nVan Der Walt S., Colbert S. C., Varoquaux G., 2011, Computing in Science\n& Engineering, 13, 22\nXu S., Yan H., 2013, The Astrophysical Journal, 779, 140\nYan H., Lazarian A., 2002, Physical review letters, 89, 281102\nYan H., Lazarian A., 2008, ApJ, 673, 942\nZweibel E. G., 2013, Physics of Plasmas, 20, 055501\nvan Marle A. J., Casse F., Marcowith A., 2019, MNRAS, p. 2249\nMNRAS 000, 000–000 (0000)" }, { "title": "2202.05295v1.Non_stationary_Anderson_acceleration_with_optimized_damping.pdf", "content": "arXiv:2202.05295v1 [math.NA] 10 Feb 2022Non-stationary Anderson acceleration with optimized\ndamping⋆\nKewang Chena,b,∗, Cornelis Vuikb\naCollege of Mathematics and Statistics, Nanjing University of Information Science and\nTechnology, Nanjing, 210044, China.\nbDelft Institute of Applied Mathematics, Delft University o f Technology, Delft, 2628XE, the\nNetherlands.\nAbstract\nAnderson acceleration (AA) has a long history of use and a strong r ecent inter-\nest due to its potential ability to dramatically improve the linear conve rgence\nof the fixed-point iteration. Most authors are simply using and analy zing the\nstationary version of Anderson acceleration (sAA) with a constan t damping\nfactor or without damping. Little attention has been paid to nonsta tionary\nalgorithms. However, damping can be useful and is sometimes crucia l for simu-\nlations in which the underlying fixed-point operator is not globally cont ractive.\nThe role of this damping factor has not been fully understood. In th e present\nwork, we consider the non-stationary Anderson acceleration algo rithm with op-\ntimized damping (AAoptD) in each iteration to further speed up linear and\nnonlinear iterations by applying one extra inexpensive optimization. W e an-\nalyze this procedure and develop an efficient and inexpensive implemen tation\nscheme. We also show that, compared with the stationary Anderso n accelera-\ntion with fixed window size sAA(m), optimizing the damping factors is related\nto dynamically packaging sAA(m) andsAA(1) in each iteration (alternating\nwindow size mis another direction of producing non-stationary AA). More-\n⋆Funding: This work was partially supported by the National Natural Sc ience Foundation\nof China [grant number 12001287]; the Startup Foundation fo r Introducing Talent of Nanjing\nUniversity of Information Science and Technology [grant nu mber 2019r106]\n∗Corresponding author\nEmail addresses: kwchen@nuist.edu.cn (Kewang Chen), c.vuik@tudelft.nl (Cornelis\nVuik)\nURL:https://homepage.tudelft.nl/d2b4e/ (Cornelis Vuik)\nPreprint submitted to Journal of Computational and Applied Mathematics.February 14, 2022over, we show by extensive numerical experiments that, in the cas e a larger\nwindow size is needed, the proposed non-stationary Anderson acc eleration with\noptimized damping procedure often converges much faster than s tationary AA\nwith constant damping or without damping. When the window size is ver y\nsmall (m≤3 was typically used, especially in the early days of application),\nAAoptD and AA are comparable. Lastly, we observed that when the system is\noverdamped (i.e. the damping factor is close to the lower bound zero ), incon-\nsistency may occur. So there is some trade-off between stability an d speed of\nconvergence. We successfully solve this problem by further restr icting damping\nfactors bound away from zero.\nKeywords: Anderson acceleration, fixed-point iteration, optimal damping.\n2010 MSC: 65H10, 65F10\n1. Introduction\nIn this part, we first give a literature review on Anderson Accelerat ion\nmethod. Thenwediscussourmainmotivationsandthe structurefo rthe present\npaper. To begin with, let us consider the nonlinearaccelerationfor t he following\ngeneral fixed-point problem\nx=g(x), g:Rn→Rn\nor its related nonlinear equations problem\nf(x) =x−g(x) = 0.\nThe associated basical fixed-point iteration is given in Algorithm 1.\nAlgorithm 1 Picard iteration\nGiven:x0.\nfork= 0,1,2,···do\nSetxk+1=g(xk).\nend for\n2The main concern related to this basic fixed-point iteration is that th e it-\nerates may not converge or may converge extremely slowly (only line ar conver-\ngent). Therefore, various acceleration methods are proposed t o alleviate this\nslow convergence problem. Among these algorithms, one popular ac celeration\nprocedure is called the Anderson acceleration method [1]. For the ab ove basic\nPicard iteration, the usual general form of Anderson acceleratio n with damping\nis given in Algorithm 2. In the above algorithm, fkis the residual for the kth\nAlgorithm 2 Anderson acceleration: AA(m)\nGiven:x0andm≥1.\nSet:x1=g(x0).\nfork= 0,1,2,···do\nSet:mk= min{m,k}.\nSet:Fk= (fk−mk,···,fk), where fi=g(xi)−xi.\nDetermine: α(k)=/parenleftig\nα(k)\n0,···,α(k)\nmk/parenrightigT\nthat solves\nmin\nα=(α0,···,αmk)T/bardblFkα/bardbl2s. t.mk/summationdisplay\ni=0αi= 1.\nSet:xk+1= (1−βk)mk/summationdisplay\ni=0α(k)\nixk−mk+i+βkmk/summationdisplay\ni=0α(k)\nig(xk−mk+i).\nend for\niteration; mis the window size which indicates how many history residuals will\nbe used in the algorithm. The value of mis typically no larger than 3 in the\nearly days of applications and now this value could be as large as up to 1 00,\nsee [2]. It is usually a fixed number during the procedure, varying mcan also\nmake the algorithm to be non-stationary. We will come back to this po int in\nsection Section 2; βk∈(0,1] is a damping factor (or a relaxation parameter) at\nkth iteration. We have, for a fixed window size m:\nβk=\n\n1, no damping,\nβ,(a constant independent of k) stationary AA,\nβk,(depending on k) non-stationary AA.\nThe constrained optimization problem can also be formulated as an eq uivalent\n3unconstrained least-squares problem [3, 4]:\nmin\n(ω1,···,ωmk)T/vextenddouble/vextenddouble/vextenddouble/vextenddouble/vextenddoublefk+mk/summationdisplay\ni=1ωi(fk−i−fk)/vextenddouble/vextenddouble/vextenddouble/vextenddouble/vextenddouble\n2(1)\nOne can easily recover the original problem by setting\nω0= 1−mk/summationdisplay\ni=1ωi.\nThis formulation of the linear least-squares problem is not optimal fo r imple-\nmentation, we will discuss this in more detail in Section 3.\nAnderson acceleration method dates back to the 1960s. In 1962, Anderson\n[1] developeda techniqueforacceleratingthe convergenceofthe Picarditeration\nassociated with a fixed-point problem which is called Extrapolation Algo rithm.\nThis technique is now called Anderson Acceleration (AA) in the applied m athe-\nmatics community and Anderson Mixing in the physics and chemistry co mmu-\nnities. This method is “essentially” (or nearly) similar to the nonlinear G MRES\nmethod or Krylov acceleration [5, 6, 7, 8] and the direct inversion on the itera-\ntive subspace method (DIIS) [9, 10, 11]. And it is also in a broad categ ory with\nmethods based on quasi-Newton updating [12, 13, 14, 15, 16]. Howe ver, unlike\nNewton-like methods, AA does not require the computation or appr oximation\nof Jacobians or Jacobian-vector products which could be an advan tage.\nAlthough the Anderson acceleration method has been around for d ecades,\nconvergence analysis has been reported in the literature only rece ntly. Fang\nand Saad [14] had clarified a remarkable relationship of AA to quasi-Ne wton\nmethods and extended it to define a broader Anderson family metho d. Later,\nWalker and Ni [17] showed that, on linear problems, AA without trunc ation\nis “essentially equivalent” in a certain sense to the GMRES method. Fo r the\nlinear case, Toth and Kelley [3] first proved the stationary version o f AA (sAA)\nwithout damping is locally r-linearly convergent if the fixed point map is a\ncontractionand the coefficients in the linear combination remain boun ded. This\nwork was later extended by Evens et al. [18] to AA with damping and t he\nauthors proved the new convergence rate is θk((1−βk−1)+βk−1κ), where κis\n4the Lipschitz constant for the function g(x) andθkis the ratio quantifying the\nconvergence gain provided by AA in step k. However, it is not clear how θk\nmay be evaluated or bounded in practice and how it may translate to im proved\nasymptotic convergence behavior in general. In 2019, Pollock et al. [19] applied\nsAA to the Picard iteration for solving steady incompressible Navier– Stokes\nequations(NSE) andprovedthat the accelerationimprovesthe co nvergencerate\nofthe Picarditeration. Then, De Sterck[20] extended the resultt o moregeneral\nfixed-point iteration x=g(x), given knowledge of the spectrum of g′(x) at\nfixed-point x∗and Wang et al. [21] extended the result to study the asymptotic\nlinear convergence speed of sAA applied to Alternating Direction Met hod of\nMultipliers (ADMM) method. Sharper local convergence results of A A remain\na hot research topic in this area. More recently, Zhang et al. [22] pr oved a\nglobal convergent result of type-I Anderson acceleration for no nsmooth fixed-\npointiterationswithoutresortingtolinesearchoranyfurtherass umptionsother\nthan nonexpansiveness. For more related results about Anderso n acceleration\nand its applications, we refer the interested readers to [2, 23, 24, 25, 26, 27, 28]\nand references therein.\nAs mentioned above, the local convergence rate θk((1−βk−1) +βk−1κ) at\nstagekis closely related to the damping factor βk−1. However, questions like\nhow to choose those damping values in each iteration [2] and how it will a ffect\nthe global convergence of the algorithm have not been deeply stud ied. Besides,\nAA is often combined with globalization methods to safeguard against erratic\nconvergence away from a fixed point by using damping. One similar idea in\nthe optimization context for nonlinear GMRES is to use line search str ategies\n[29]. This is an important strategy but not yet fully explored in the liter ature.\nMoreover, the early days of Anderson Mixing method (the 1980s, f or electronic\nstructure calculations) initially dictated the window size m≤3 due to the\nstorage limitations and costly gevaluations involving large N. However, in\nrecent years and a broad range of contexts, the window size mranging from\n20 to 100 has also been considered by many authors. For example, W alker\nand Ni [17] used m= 50 in solving the nonlinear Bratu problem. A natural\n5question will be should we try to further steep up Anderson acceler ationmethod\nor try to use a larger size of the window? No such comparison results have been\nreported. Motivated by the above works, in this paper, we propos e, analyze\nand numerically study non-stationary Anderson acceleration with o ptimized\ndamping to solve fixed-point problems. The goal of this paper is to ex plore the\nrole of damping factors in non-stationary Anderson acceleration.\nThe paper is organized as follows. Our new algorithms and analysis are in\nSection 2, the implementation ofthe new algorithmis in Section 3, expe rimental\nresults and discussion are in Section 4. Conclusions follow in Section 5.\n2. Anderson acceleration with optimized dampings\nIn this section, we focus on developing the algorithm for Anderson a cceler-\nation with optimized dampings at each iteration and studying its conve rgence\nrate explicitly.\nxk+1= (1−βk)mk/summationdisplay\ni=0α(k)\nixk−mk+i+βkmk/summationdisplay\ni=0α(k)\nig(xk−mk+i)\n=mk/summationdisplay\ni=0α(k)\nixk−mk+i+βk/parenleftiggmk/summationdisplay\ni=0α(k)\nig(xk−mk+i)−mk/summationdisplay\ni=0α(k)\nixk−mk+i/parenrightigg\n.(2)\nDefine the following averagesgiven by the solution αkto the optimization prob-\nlem by\nxα\nk=mk/summationdisplay\ni=0α(k)\nixk−mk+i,˜xα\nk=mk/summationdisplay\ni=0α(k)\nig(xk−mk+i). (3)\nThen (2) becomes\nxk+1=xα\nk+βk(˜xα\nk−xα\nk). (4)\nA natural way to choose “best” βkat this stage is that choosing βksuch that\nxk+1gives a minimal residual. This is similar to the original idea of Anderson\naccelerationwithwindowsizeequaltoone. Sowejustneedtosolvet hefollowing\nunconstrained optimization problem:\nmin\nβk/bardblxk+1−g(xk+1)/bardbl2= min\nβk/bardblxα\nk+βk(˜xα\nk−xα\nk)−g(xα\nk+βk(˜xα\nk−xα\nk))/bardbl2.(5)\n6Noting the fact that\ng(xα\nk+βk(˜xα\nk−xα\nk))≈g(xα\nk)+βk∂g\n∂x/vextendsingle/vextendsingle/vextendsingle\nxα\nk(˜xα\nk−xα\nk)\n≈g(xα\nk)+βk(g(˜xα\nk)−g(xα\nk)). (6)\nTherefore, (5) becomes\nmin\nβk/bardblxk+1−g(xk+1)/bardbl2\n= min\nβk/bardblxα\nk+βk(˜xα\nk−xα\nk)−g(xα\nk+βk(˜xα\nk−xα\nk))/bardbl2\n≈min\nβk/bardblxα\nk+βk(˜xα\nk−xα\nk)−[g(xα\nk)+βk(g(˜xα\nk)−g(xα\nk))]/bardbl2\n≈min\nβk/bardbl(xα\nk−g(xα\nk))−βk[(g(˜xα\nk)−g(xα\nk))−(˜xα\nk−xα\nk)]/bardbl2.(7)\nThus, we just need to calculate the projection\nβk=/vextendsingle/vextendsingle/vextendsingle(xα\nk−g(xα\nk))·[(xα\nk−g(xα\nk))−(˜xα\nk−g(˜xα\nk))]\n/bardbl[(xα\nk−g(xα\nk))−(˜xα\nk−g(˜xα\nk))]/bardbl2/vextendsingle/vextendsingle/vextendsingle. (8)\nSet\nrp= (xα\nk−g(xα\nk)), rq= (˜xα\nk−g(˜xα\nk)),\nwe have\nβk=/vextendsingle/vextendsingle/vextendsingle/vextendsingle(rp−rq)Trp\n/bardblrp−rq/bardbl2/vextendsingle/vextendsingle/vextendsingle/vextendsingle. (9)\nWe will discuss how much work is needed to calculate this βkin Section 3.\nFinally, our analysisleads to the followingnon-stationaryAnderson a cceleration\nalgorithm with optimized damping: AAoptD(m).\nRemark 2.1. As mentioned in Section 1, changing the window size mat each\niteration can also make a stationary Anderson acceleration to be non-stationary.\nComparing withthe stationary Anderson acceleration with fi xedwindow sAA(m),\nour proposed nonstationary procedure ( AAoptD(m)) of choosing optimal βkis\nsomewhat related to packaging sAA(m)andsAA(1)in each iteration in a cheap\nway. Combining sAA(m)withsAA(1)can provide really good outcomes, espe-\ncially in the case when larger mis needed. We will discuss this in detail for the\nnumerical results in Section 4.\n7Algorithm 3 Anderson acceleration with optimized dampings: AAoptD(m)\nGiven:x0andm≥1.\nSet:x1=g(x0).\nfork= 0,1,2,···do\nSet:mk= min{m,k}.\nSet:Fk= (fk−mk,···,fk), where fi=g(xi)−xi.\nDetermine: α(k)=/parenleftig\nα(k\n0,···,α(k)\nmk/parenrightigT\nthat solves\nmin\nα=(α0,···,αmk)T/bardblFkα/bardbl2s. t.mk/summationdisplay\ni=0αi= 1.\nSet:xα\nk=mk/summationdisplay\ni=0α(k)\nixk−mk+i,˜xα\nk=mk/summationdisplay\ni=0α(k)\nig(xk−mk+i).\nSet:rp= (xα\nk−g(xα\nk)), rq= (˜xα\nk−g(˜xα\nk)).\nSet: βk=(rp−rq)Trp\n/bardblrp−rq/bardbl2.\nSet:xk+1=xα\nk+βk(˜xα\nk−xα\nk).\nend for\nRemark 2.2. Here this optimized damping step is a “local optimal” strate gy at\nkth iteration. It usually will speed up the convergence rate c ompared with an\nundamped one, but not always. Because in (k+1)th iteration, it uses a combi-\nnation of all previous m history information. Moreover, whe nβkis very close\nto zero, the system is over-damped, which, sometimes, may al so slow down the\nconvergence speed. We may need to further modify our βk. See more discussion\nin our numerical results in Section 4.\nLastly, we summarize the convergence results with damping in Theor em 2.1.\nThe proof of this theorem can be found in [18].\nTheorem 2.1. [18] Assume that g:Rn→Rnis uniformly Lipschitz con-\ntinuously differentiable and there exists κ∈(0,1)such that/bardblg(y)−g(x)/bardbl2≤\nκ/bardbly−x/bardbl2for allx,y∈Rn. Suppose also that ∃Mandǫ >0such that for all\nk > m,/summationtextm−1\ni=0|αi|< Mand|αm|≥ǫ. Then\n/bardblf(xk+1)/bardbl2≤θk+1[(1−βk)+κβk]/bardblf(xk)/bardbl2+m/summationdisplay\ni=0O(/bardblf(xk−m+i)/bardbl2\n2),(10)\n8where\nθk+1=/bardbl/summationtextm\ni=0αif(xk−m+i)/bardbl2\n/bardblf(xk)/bardbl2.\n3. Implementation\nForimplementation, wemainlyfollowthepathin[4]andmodifyitasneed ed.\nWe first briefly review the implementation of AA without damping. Then we\nfocus on how to implement the optimized damping problem efficiently and ac-\ncurately.\nThe constrained linear least-squares problem in Algorithm 2 can be so lved\nin many ways. Here we rewrite it into an equivalent unconstrained for m which\ncan be solved efficiently by using QR factorizations. We define ∆ fi=fi+1−fi\nfor each iand setFk= (∆fk−mk,···,∆fk−1), then the least-squares problem\nis equivalent to\nmin\nγ=(γ0,···,γmk−1)T/bardblfk−Fkγ/bardbl2,\nwhereαandγare related by α0=γ0,αi=γi−γi−1for 1≤i≤mk−1, and\nαmk= 1−γmk−1.We assumeFhas a thin QRdecomposition i.e., Fk=QkRk\nwithQk∈Rn×mkandRk∈Rmk×mk, forwhichthesolutionoftheleast-squares\nproblem is obtained by solving the mk×mktriangular system Rkγ=QT\nkfk.\nAs the algorithm proceeds, the successive least-squares problem s can be solved\nefficiently by updating the factors in the decomposition.\nAssume that γk= (γk\n0,···,γk\nmk−1)Tis the solution to the above modified\nform of Anderson acceleration, we have\nxk+1=g(xk)−mk−1/summationdisplay\ni=0γk\ni[g(xk−mk+i+1)−g(xk−mk+i)] =g(xk)−Gkγk,\nwhereGk= (∆ggk−mk,···,∆gk−1) with ∆ gi=g(xi+1−g(xi)) for each i. For\nAnderson acceleration with damping\nxk+1= (1−βk)mk/summationdisplay\ni=0α(k)\nixk−mk+i+βkmk/summationdisplay\ni=0α(k)\nig(xk−mk+i)\n=mk/summationdisplay\ni=0α(k)\nixk−mk+i+βk/parenleftiggmk/summationdisplay\ni=0α(k)\nig(xk−mk+i)−mk/summationdisplay\ni=0α(k)\nixk−mk+i/parenrightigg\n.\n9Follow the idea in [4], we have\nmk/summationdisplay\ni=0α(k)\nig(xk−mk+i) =g(xk)−Gkγk, (11)\nmk/summationdisplay\ni=0α(k)\nixk−mk+i=/parenleftbig\ng(xk)−Gkγk/parenrightbig\n−/parenleftbig\nfk−Fkγk/parenrightbig\n. (12)\nThen this can be achieved equivalently using the following strategy:\nStep 1: Compute the undamped iterate xk+1=g(xk)−Gkγk.\nStep 2: Update xk+1again by\nxk+1←xk+1−(1−βk)/parenleftbig\nfk−QRγk/parenrightbig\n.\nNow we talk about how to efficiently calculate βkas described in Algorithm 3.\nTaking benefit of the QR decomposition in the first optimization proble m and\nnoting (11) and (12), we have\n˜xα\nk=mk/summationdisplay\ni=0α(k)\nig(xk−mk+i) =g(xk)−Gkγk,\nxα\nk=mk/summationdisplay\ni=0α(k)\nixk−mk+i= ˜xα\nk−/parenleftbig\nfk−Fkγk/parenrightbig\n.\nThen we could calculate optimized βkby doing two extra function evaluations\nand two dot products, which are not very expensive:\nrp= (xα\nk−g(xα\nk)), rq= (˜xα\nk−g(˜xα\nk)), βk=/vextendsingle/vextendsingle/vextendsingle/vextendsingle(rp−rq)Trp\n/bardblrp−rq/bardbl2/vextendsingle/vextendsingle/vextendsingle/vextendsingle.\nIn practice, when xkis very close to the fixed-point x∗, scientific computing\nerrors may arise in calculating these two high dimension vectors rpandrp−rq.\nThus we normalize these two vectors first, then calculate βkby simply doing a\ndot product.\n4. Experimental results and discussion\nInthissection,wenumericallycomparetheperformanceofthisnon -stationary\nAAoptD with sAA (with constant damping or without damping). The fir st part\n10containsexamples where largerwindow sizes mareneeded in orderto accelerate\nthe iteration. The second part consists of some examples where sm all window\nsizes are working very well. All these experiments are done in MATLAB 2021b\nenvironment. MATLAB codes are available upon request to the auth ors.\nThis first example is from Walker and Ni’s [17] paper, where a stationar y\nAnderson acceleration with window size m= 50 is used to solve the Bratu\nproblem. This problem has a long history, we refer the reader to Glow inski et\nal. [30] and Pernice and Walker [31], and the references in those pape rs. It is\nnot a difficult problem for Newton-like solvers.\nProblem 4.1. The Bratu problem. The Bratu problem is a nonlinear PDE\nboundary value problem as follows:\n∆u+λ eu= 0, in D = [0,1]×[0,1],\nu= 0, on ∂D.\nIn this experiment, we used a centered-difference discretization o n a 32×32,\n64×64and 128×128grid, respectively. We take λ= 6in the Bratuproblemand\nuse the zero initial approximate solution in all cases. We also applied pr econ-\nditioning such that the basic Picard iteration still works. The precon ditioning\nmatrix that we used here is the diagonal inverse of the matrix A, whereAis a\nmatrix for the discrete Laplace operator.\nThe resultsareshowninthe followingfigures. InFig.1, weplot there sultsof\napplying AA(m) andAAoptD(m) to acceleratePicarditeration with m= 5 and\nm= 10 on a grid of 32 ×32. As we see from the picture, AA(5) andAA(10) does\nnot accelerate the convergence speed very much. AAoptD(5) andAAoptD(10)\nperform much better than AA(5) andAA(10). However, we also notice that\nthere are some inconsistencies and stagnations in AAoptD(m). Thus we go\nfurther to plot the βkvalues that are used in each iteration, see Fig. 2. From\nFig. 2 we see that: for AAoptD(10), some optimized damping factors are below\n0.3(see the dashedline). As we know, the damping factor βk∈(0,1] andβk= 1\nmeans no damping. Thus small βkmay cause an over-damping phenomenon,\n110 20 40 60 80 100 120 140 160 180 200\niteration10-910-810-710-610-510-410-310-210-1residual\nAA(5)\nAA(10)\nAAoptD(5)\nAAoptD(10)\nFigure 1: Compare AA and AoptD for solving nonlinear Bratu pr oblems.\nwhich might be the reasonfor small inconsistencies observedin Fig. 1 ; Similarly,\nwe see that the residual of AAoptD(10) in Fig. 1 is not decreasing consistently\naround 10th iteration (see the read dashed square region in Fig. 1) , where the\ncorresponding βkvalues are super close to zero as shown in Fig. 2.\nTo balance the over-damping effect, we bound these βkaway from zero. The\nfirst strategy we propose is to use\nˆβk= max{βk,η}, (13)\nwhereηis a small positive number such that 0 < η <0.5. For example, to\nreduce the over-damping effect, we take η= 0.3 in (13) as a lower bound. We\nplot the new βkvalues at each iteration in Fig. 3. There are no βkvalues less\nthan 0.3 anymore. The corresponding results are in Fig. 4. Compared with t he\nresults in Fig. 1, we see that there is less stagnation (see the red da shed square\nregion in Fig. 4) and onvergence is also faster. We also note that the βkvalues\nin Fig. 3 differs a lot from the values of βkin Fig. 2. Because changing βkin\n120 10 20 30 40 50 60 70 80 90\niteration00.10.20.30.40.50.60.70.8k\nFigure 2: Optimal damping factors in each iteration for m= 10.\nprevious iterations will affect the later ones.\n0 10 20 30 40 50 60 70 80\niteration00.10.20.30.40.50.60.70.8k\nFigure 3: Modified optimal damping factors: ˆβk= max{βk,η}withη= 0.3\nAlthough the results in Fig. 4 are better than those in Fig. 1, we notic e that\nthere are still some inconsistencies in the red dashed square region . To further\nsmooth out these inconsistencies, we change these “bad” βkvalues further away\nfrom zero. Therefore, we propose our second strategy:\nˆβk=\n\nβkifβk≥η,\n1−βkifβk< η.(14)\nWe note here that there is some trade-off between stability and spe ed of conver-\ngence. This does not mean that larger βkwork better, since larger βkmay not\nspeed up the convergence if it is not appropriate. Therefore, dam ping is good,\n130 20 40 60 80 100 120 140 160 180 200\niteration10-910-810-710-610-510-410-310-210-1residual\nAA(5)\nAA(10)\nAAoptD(5)\nAAoptD(10)\nFigure 4: Solving nonlinear Bratu problems: ˆβk= max{βk,η}withη= 0.3\nbut over-damping may cause inconsistencies and stagnation. In ou r numerical\nexperiment, we take η= 0.3 in (14) as an example. The results are in Fig. 5.\nCompared with the results in Fig. 1 and Fig. 4, it becomes better. We s ee that\nthere are almost no inconsistencies and there is faster convergen ce. We also plot\nthe new βkin Fig. 5.\nTo compare with the results provided in [17], we go further to increas e the\nwindows until m= 50. Again, without bounding away from zero, there are\nsome stagnations and inconsistencies. To avoid strong over-damp ing, we apply\n(14) again with η= 0.3 and obtain our new results in Fig. 7. We easily see that\nAAoptD(20) works as well as AA(50). Moreover, to test its scaling properties,\nwe also solve the Bratu problem on larger grids. In Fig. 8, for a grid siz e 64×64,\nwe see that AAoptD(10) is already comparable with AA(60) and AAoptD(30)\nperforms better than AA(60). Similarly, for a grid size 128 ×128, Fig. 9 shows\nthatAAoptD(40) performs much better than AA(80).\n140 20 40 60 80 100 120 140 160 180 200\niteration10-910-810-710-610-510-410-310-210-1residual\nAA(5)\nAA(10)\nAAoptD(5)\nAAoptD(10)\nFigure 5: Solving nonlinear Bratu problems: ˆβk= 1−βkwhenβk<0.3.\n0 10 20 30 40 50 60 70 80 90 100\niteration0.10.20.30.40.50.60.70.8k\nFigure 6: Modified optimal damping factors: ˆβk= 1−βkwhenβk<0.3.\n150 20 40 60 80 100 120 140 160 180 200\niteration10-910-810-710-610-510-410-310-210-1residual\nPicard\nAA(10)\nAA(20)\nAA(30)\nAA(40)\nAA(50)\nAAoptD(10)\nAAoptD(20)\nFigure 7: Using larger size windows and bounding the damping factor away from zero.\n0 50 100 150 200 250 300\niteration10-910-810-710-610-510-410-310-210-1residual64x64\nPicard\nAA(10)\nAA(20)\nAA(30)\nAA(50)\nAA(60)\nAAoptD(10)\nAAoptD(20)\nAAoptD(30)\nFigure 8: Scaling: solve the Bratu problem on a 64 ×64 gird.\n160 50 100 150 200 250 300 350 400 450 500\niteration10-910-810-710-610-510-410-310-210-1residual128x128\nPicard\nAA(20)\nAA(40)\nAA(60)\nAA(80)\nAAoptD(20)\nAAoptD(40)\nFigure 9: Scaling: solve the Bratu problem on a 128 ×128 grid.\nProblem 4.2. The nonlinear convection-diffusion problem. Use AAand\nAAoptD to solve the following 2D nonlinear convection-diffu sion equation in a\nsquare region:\n(−uxx−uyy)+(ux+uy)+ku2=f(x,y),(x,y)∈D= [0,1]×[0,1]\nwith the source term\nf(x,y) = 2π2sin(πx)sin(πy)\nand zero boundary conditions: u(x,y) = 0on∂D.\nIn this numerical experiment, we use a centered-difference discre tization on\n32×32 and 64×64 grids, respectively. We take k= 3 in the above problem\nand use u0= (1,1,···,1)Tas an initial approximate solution in all cases. As\nin solving the Bratu problem, the same preconditioning strategy is us ed here so\nthat the basic Picard iteration still works. To bound βkaway from zero, we use\n(14) with η= 0.25. The results are shown in Fig. 10 and Fig. 11 for n= 32×32\n17andn= 64×64, respectively. From Fig. 10, we see that AAoptD(5) is already\nbetter than AA(15); From Fig. 11, we also observe that AAoptD(20) is better\nthanAA(50). In both cases, AAoptD(m) does a much better job than AA(m),\nwhich is consistent with our previous example.\n0 100 200 300 400 500\niteration10-1010-810-610-410-2100102residual32 32\nPicard\nAA(5)\nAA(10)\nAA(15)\nAAoptD(5)\nAAoptD(10)\nAAoptD(15)\nFigure 10: Solving the nonlinear convection-diffusion prob lem on a 32 ×32 gird.\nOur next example is about solving a linear system Ax=b. As proved by\nWalker and Ni in [17], AA without truncation is “essentially equivalent” in a\ncertain sense to the GMRES method for linear problems.\nProblem 4.3. The linear equations. Apply AA and AAoptD to solve the\nfollowing linear system Ax=b, whereAis\nA=\n2−1···0 0\n−1 2···0 0\n...............\n0 0···2−1\n0 0··· −1 2\n, A∈Rn×n\n180 100 200 300 400 500 600\niteration10-1010-810-610-410-2100102residual64 64\nPicard\nAA(10)\nAA(20)\nAA(30)\nAA(50)\nAAoptD(10)\nAAoptD(20)\nAAoptD(30)\nFigure 11: Solving the nonlinear convection-diffusion prob lem on a 64 ×64 gird.\nand\nb= (1,···,1)T.\nChoosen= 10andn= 100, respectively. Here, we choose a large nso that\na large window size mis needed in Anderson Acceleration. We also note that\nthe Picard iteration does not work for this problem.\nThe initial guess is x0= (0,···,0)T. Without bounding βkaway from zero,\nthe results are shown in Fig. 12 and Fig. 13. For small m,AA(1) does not work,\nbutAAoptD(1) works. Moreover, we obtain from Fig. 12 that AAoptD(m) still\ndoes better than AA(m). When n= 100, we need larger mvalues. In this case,\nas shown in Fig. 13, AAoptD(5) already performs much better than AA(25).\nThis example shows that AAoptD can also be used to solve linear problems.\nFinally, we consider cases where very small mworks. Our example is from\nToth and Kelley’s paper [3], where AA is applied to solve the Chandrase khar\nH-equation.\n190 50 100 150 200 250 300\niteration10-910-810-710-610-510-410-310-210-1100101residualn=10\nAA(2)\nAA(3)\nAA(4)\nAAoptD(1)\nAAoptD(2)\nAAoptD(3)\nAAoptD(4)\nFigure 12: Small m: solving a linear problem Ax=bwithn= 10.\n0 50 100 150 200 250 300 350 400 450 500\niteration10-810-710-610-510-410-310-210-1100101residualn=100\nAA(5)\nAA(10)\nAA(15)\nAA(20)\nAA(50)\nAAoptD(5)\nAAoptD(10)\nAAoptD(15)\nAAoptD(20)\nFigure 13: Large m: solving a linear problem Ax=bwithn= 100.\n20Problem 4.4. the Chandrasekhar H-equation, arising in Radiative Heat Tr ans-\nfer theory, is a nonlinear integral equation:\nH(µ) =G(H) =/parenleftbigg\n1−c\n2/integraldisplay1\n0µ\nµ+vH(v)dv/parenrightbigg−1\n,\nwherec∈[0,1)is a physical parameter.\nWe will discretize the equation with the composite midpoint rule. Here we\napproximate integrals on [0,1]by\n/integraldisplay1\n0f(µ)dµ≈1\nNN/summationdisplay\nj=1f(µj)\nwhereµj= (i−1/2)/Nfor1≤i≤N. The resulting discrete problem is\nF(x)i=xi−\n1−c\n2NN/summationdisplay\nj=1µixj\nµi+µj\n−1\n,\nwhich is a fully nonlinear system.\nIt is known [32] both for the continuous problem and its follo wing midpoint\nrule discretization, that if c <1\nρ(G′(H∗))≤1−√\n1−c <1,\nwhereρdenotes spectral radius. Hence the local convergence theor y and Picard\niteration works.\nIn our numerical experiment, we choose N= 500,c= 0.5,c= 0.99 and\nc= 1. The case c= 1 is a critical value (Picard does not work in this case, but\nAA does). The numerical results are in Fig. 14 to Fig. 16. Firstly, AA(m) and\nAAoptD(m), with very small m( ≤3) values, work for all cases including the\ncritical case c= 1 and their performances are comparable. Secondly, increasing\nmdoes not always increase the performance. Thirdly, AAoptD may no t always\nhave advantages over AA for small window size m. This result is reasonable\nsince AAoptD(m) is kind of like packaging AA(m) andAA(1). Ifmis small,\nthere is almost no difference between AA(m) andAA(1), thus packaging them\n(varying window sizes) may not give better results.\n210 1 2 3 4 5 6 7 8 9 10\niteration10-810-710-610-510-410-310-210-1100101102residualPicard\nAA(1)\nAA(2)\nAA(3)\nAAoptD(1)\nAAoptD(2)\nAAoptD(3)\nFigure 14: Solving Chandrasekhar H-equation with AA and AAo ptD:c= 0.5\n0 5 10 15 20 25 30 35 40 45 50\niteration10-1010-810-610-410-2100102residualPicard\nAA(1)\nAA(2)\nAA(3)\nAAopt(1)\nAAoptD(2)\nAAoptD(3)\nFigure 15: Solving Chandrasekhar H-equation with AA and AAo ptD:c= 0.99\n220 5 10 15 20 25 30 35 40 45 50\niteration10-810-710-610-510-410-310-210-1100101102residualPicard\nAA(1)\nAA(2)\nAAoptD(1)\nAAoptD(2)\nFigure 16: Solving Chandrasekhar H-equation with AA and AAo ptD:c= 1\n5. Conclusions\nWe proposedanon-stationaryAndersonaccelerationalgorithmwit h anopti-\nmized damping factor in each iteration to further speed up linear and nonlinear\niterations by applying one extra optimization. This procedure has a s trong con-\nnection toanotherperspective ofgeneratingnon-stationaryAA (i.e. varyingthe\nwindow size mat different iterations). It turns out that choosing optimal βk\nis somewhat similar to packaging sAA(m) and sAA(1) within a single itera tion\nin a cheap way. Moreover, by taking benefit of the QR decomposition in the\nfirst optimization problem, the calculation of optimized βkat each iteration is\ncheap if two extra function evaluations are relatively inexpensive. O ur numer-\nical results show that the gain of doing this extra optimized step on βkcould\nbe large. Moreover, damping is good but over damping is not good bec ause it\nmay slow down the convergence rate. Therefore, when the statio nary AA is not\nworking well or a larger size of the window is needed in AA, we recommen d to\n23use AAoptD proposed in the present work.\nAcknowledgments\nThis work was partially supported by the National Natural Science F ounda-\ntion of China [grant number 12001287]; the Startup Foundation for Introduc-\ning Talent of Nanjing University of Information Science and Technolo gy [grant\nnumber 2019r106]; The first author Kewang Chen also gratefully ac knowledge\nthe financial support for his doctoral study provided by the China Scholarship\nCouncil (No. 202008320191).\nReferences\n[1] D. G. Anderson, Iterative procedures for nonlinear integral e quations, J.\nAssoc. Comput. Mach. 12 (1965) 547–560. doi:10.1145/321296.321305 .\n[2] D. G. M. Anderson, Comments on “Anderson acceleration, mix-\ning and extrapolation”, Numer. Algorithms 80 (1) (2019) 135–234.\ndoi:10.1007/s11075-018-0549-4 .\n[3] A. Toth, C. T. Kelley, Convergence analysis for Anderson accele ration,\nSIAM J. Numer. Anal. 53 (2) (2015) 805–819. doi:10.1137/130919398 .\n[4] H. F. Walker, Anderson acceleration: Algorithms and implementat ions,\nWPI Math. Sciences Dept. Report MS-6-15-50.\nURLhttps://users.wpi.edu/ ~walker/Papers/anderson_accn_algs_imps.pdf\n[5] N. N. Carlson, K. Miller, Design and application of a gradient-weight ed\nmoving finite element code. I. In one dimension, SIAM J. Sci. Comput.\n19 (3) (1998) 728–765. doi:10.1137/S106482759426955X .\n[6] K. Miller, Nonlinear Krylov and moving nodes in the method\nof lines, J. Comput. Appl. Math. 183 (2) (2005) 275–287.\ndoi:10.1016/j.cam.2004.12.032 .\n24[7] C. W. Oosterlee, T. Washio, Krylov subspace acceleration of non linear\nmultigrid with application to recirculating flows, SIAM J. Sci. Comput.\n21 (5) (2000) 1670–1690. doi:10.1137/S1064827598338093 .\n[8] T.Washio,C.W.Oosterlee,Krylov subspace acceleration for no nlinear multigrid schemes,\nElectron. Trans. Numer. Anal. 6 (Dec.) (1997) 271–290.\nURLhttp://citeseerx.ist.psu.edu/viewdoc/summary?doi=10 .1.1.147.3799\n[9] L. Lin, C. Yang, Elliptic preconditioner for accelerating the self-c onsistent\nfield iteration in Kohn-Sham density functional theory, SIAM J. Sci. Com-\nput. 35 (5) (2013) S277–S298. doi:10.1137/120880604 .\n[10] P. Pulay, Convergence acceleration of iterative sequences. t he case\nof SCF iteration, Chemical Physics Letters 73 (2) (1980) 393–398 .\ndoi:https://doi.org/10.1016/0009-2614(80)80396-4 .\n[11] P. Pulay, Improved SCF convergence acceleration, Journal o f Computa-\ntional Chemistry 3 (4) (1982) 556–560. doi:10.1002/jcc.540030413 .\n[12] T. Eirola, O. Nevanlinna, Accelerating with rank-one updates, L inear Al-\ngebra Appl. 121 (1989) 511–520. doi:10.1016/0024-3795(89)90719-2 .\n[13] V. Eyert, A comparative study on methods for convergence a cceleration\nof iterative vector sequences, J. Comput. Phys. 124 (2) (1996) 271–285.\ndoi:10.1006/jcph.1996.0059 .\n[14] H.-r. Fang, Y. Saad, Two classes of multisecant methods for no nlin-\near acceleration, Numer. Linear Algebra Appl. 16 (3) (2009) 197–2 21.\ndoi:10.1002/nla.617 .\n[15] R. Haelterman, J. Degroote, D. Van Heule, J. Vierendeels, On t he similar-\nities between the quasi-Newton inverse least squares method and G MRES,\nSIAM J. Numer. Anal. 47 (6) (2010) 4660–4679. doi:10.1137/090750354 .\n[16] C. Yang, J. C. Meza, B. Lee, L.-W. Wang, KSSOLV—a MATLAB too lbox\nfor solving the Kohn-Sham equations, ACM Trans. Math. Software 36 (2)\n(2009) Art. 10, 35. doi:10.1145/1499096.1499099 .\n25[17] H. F. Walker, P. Ni, Anderson acceleration for fixed-point itera tions, SIAM\nJ. Numer. Anal. 49 (4) (2011) 1715–1735. doi:10.1137/10078356X .\n[18] C. Evans, S. Pollock, L. G. Rebholz, M. Xiao, A proof that Ander son\nacceleration improves the convergence rate in linearly converging fi xed-\npointmethods(but notinthoseconvergingquadratically),SIAMJ. Numer.\nAnal. 58 (1) (2020) 788–810. doi:10.1137/19M1245384 .\n[19] S. Pollock, L. G. Rebholz, M. Xiao, Anderson-accelerated conv ergence of\nPicard iterations for incompressible Navier-Stokes equations, SIA M J. Nu-\nmer. Anal. 57 (2) (2019) 615–637. doi:10.1137/18M1206151 .\n[20] H. De Sterck, Y. He, On the asymptotic linear convergence spe ed of An-\nderson acceleration, Nesterov acceleration, and nonlinear GMRES , SIAM\nJ. Sci. Comput. 43 (5) (2021) S21–S46. doi:10.1137/20M1347139 .\n[21] D. Wang, Y. He, H. De Sterck, On the asymptotic linear converg encespeed\nof Anderson acceleration applied to ADMM, J. Sci. Comput. 88 (2) (2 021)\nPaper No. 38, 35. doi:10.1007/s10915-021-01548-2 .\n[22] J. Zhang, B. O’Donoghue, S. Boyd, Globally convergent type-I Anderson\nacceleration for nonsmooth fixed-point iterations, SIAM J. Optim. 30 (4)\n(2020) 3170–3197. doi:10.1137/18M1232772 .\n[23] W. Bian, X. Chen, C. T. Kelley, Anderson acceleration for a class of nons-\nmooth fixed-point problems, SIAM J. Sci. Comput. 43 (5) (2021) S1 –S20.\ndoi:10.1137/20M132938X .\n[24] P. R. Brune, M. G. Knepley, B. F. Smith, X. Tu, Composing scal-\nable nonlinear algebraic solvers, SIAM Rev. 57 (4) (2015) 535–565.\ndoi:10.1137/130936725 .\n[25] Y.Peng,B.Deng,J.Zhang,F.Geng,W.Qin,L.Liu,Andersona cceleration\nfor geometry optimization and physics simulation, ACM Transactions on\nGraphics (TOG) 37 (4) (2018) 1–14. doi:10.1145/3197517.3201290 .\n26[26] A. Toth, J. A. Ellis, T. Evans, S. Hamilton, C. T. Kelley, R. Pawlows ki,\nS. Slattery, Local improvement results for Anderson acceleratio n with inac-\ncurate function evaluations, SIAM J. Sci. Comput. 39 (5) (2017) S 47–S65.\ndoi:10.1137/16M1080677 .\n[27] W. Shi, S. Song, H. Wu, Y.-C. Hsu, C. Wu, G. Huang,\nRegularized Anderson acceleration for off-policy deep reinforceme nt learning,\narXiv preprint arXiv:1909.03245.\nURLhttps://arxiv.org/abs/1909.03245\n[28] Y. Yang, Anderson acceleration for seismic inversion, Geophys ics 86 (1)\n(2021) R99–R108. doi:10.1190/geo2020-0462.1 .\n[29] H. De Sterck, A nonlinear GMRES optimization algorithm for canon ical\ntensor decomposition, SIAM J. Sci. Comput. 34 (3) (2012) A1351– A1379.\ndoi:10.1137/110835530 .\n[30] R. Glowinski, H. B. Keller, L. Reinhart, Continuation-conjugate gra-\ndient methods for the least squares solution of nonlinear boundary\nvalue problems, SIAM J. Sci. Statist. Comput. 6 (4) (1985) 793–83 2.\ndoi:10.1137/0906055 .\n[31] M. Pernice, H. F. Walker, NITSOL: a Newton iterative solver for\nnonlinear systems, SIAM J. Sci. Comput. 19 (1) (1998) 302–318.\ndoi:10.1137/S1064827596303843 .\n[32] C. T. Kelley, T. W. Mullikin, Solution by iteration of H-equations in multi-\ngroup neutron transport, J. Mathematical Phys. 19 (2) (1978) 500–501.\ndoi:10.1063/1.523673 .\n27" }, { "title": "2008.00551v1.Integration_of_Dirac_s_Efforts_to_construct_Lorentz_covariant_Quantum_Mechanics.pdf", "content": "arXiv:2008.00551v1 [quant-ph] 2 Aug 2020Integration of Dirac’s Efforts to construct\nLorentz-covariant Quantum Mechanics\nYoung S. Kim\nCenter for Fundamental Physics, University of Maryland, College P ark, MD 20742, USA\nMarilyn E. Noz\nDepartment of Radiology, New York University, New York, NY 10016 , USA\nAbstract\nThe lifelong efforts of Paul A. M. Dirac were to construct local ized quantum systems in\nthe Lorentz covariant world. In 1927, he noted that the time- energy uncertainty should be\nincludedintheLorentz-covariant picture. In1945, heatte mptedtoconstructarepresentation\nof the Lorentz group using a normalizable Gaussian function localized both in the space and\ntime variables. In 1949, he introduced his instant form to ex clude time-like oscillations. He\nalso introduced the light-cone coordinate system for Loren tz boosts. Also in 1949, he stated\nthe Lie algebra of the inhomogeneous Lorentz group can serve as the uncertainty relations\nin the Lorentz-covariant world. It is possible to integrate these three papers to produce the\nharmonic oscillator wave function which can be Lorentz-tra nsformed. In addition, Dirac, in\n1963, considered two coupled oscillators to derive the Lie a lgebra for the generators of the\nO(3,2) de Sitter group, which has ten generators. It is proven pos sible to contract this group\nto the inhomogeneous Lorentz group with ten generators, whi ch constitute the fundamental\nsymmetry of quantum mechanics in Einstein’s Lorentz-covar iant world.\nPublished in Symmetry, Vol. 12(8), 1720 (2020).1 Introduction\nSince 1973 [1], the present authors have been publishing pap ers on the harmonic oscillator\nwave functions which can be Lorentz-boosted, leading to a nu mber of books [2, 3, 4, 5].\nWe noticed that the Gaussian form of the wave function underg oes an elliptic space-time\ndeformation when Lorentz-boosted, leading to the correct p roton form factor behavior for\nlarge momentum transfers.\nIt was then noted that the harmonic oscillator functions exh ibit the contraction and\northogonality properties quite consistent with known rule s of special relativity and quantum\nmechanics [6, 7].\nIn1977, usingtheLorentz-covariant wave functions, thepr esentauthorsshowedthat Gell-\nMann’s quark model [8] for the hadron at rest and Feynman’s pa rton model for fast-moving\nhadrons [9] are two different manifestations of one Lorentz-c ovariant entity [10, 11].\nIn 1979 [12], it was shown that the oscillator system we const ructed can be used for a\nrepresentation of the Lorentz group, particularly Wigner’ sO(3)-like little group for massive\nparticles [13]. More recently, it was shown that these oscil lator wave functions can serve as\nsqueezed states of light and Gaussian entanglement [14, 15, 16].\nIn 1983, it became known that, if the speed of a spin-1 particl e reaches that of light, the\ncomponent of spin in the direction of its momentum remains in variant as its helicity, but the\nspin components perpendicular to the momentum become one ga uge degree of freedom [17,\n18]. Indeed, this result allows us to include our oscillator -based quark-parton picture as\nfurther content of Einstein’s E=mc2[11], as shown in Table 1.\nThepurposeofthepresentpaperistoshowthattheharmonico scillator wavefunctionswe\nhave studied since 1973 can serve another purpose. It is poss ibleto obtain this covariant form\nof relativistic extended particles by integrating the thre e papers Dirac wrote, as indicated in\nTable 1.\n1. In 1927, Dirac pointed out that the time-energy uncertain ty should be considered if the\nsystem is to be Lorentz-covariant [19].\n2. In 1945, Dirac said the Gaussian form could serve as a repre sentation of the Lorentz\ngroup [20].\n3. In 1949, when Dirac introduced both his instant form of qua ntum mechanics and his\nlight-cone coordinate system [21], he clearly stated that fi nding a representation of the\ninhomogeneous Lorentz group was the task of Lorentz-covari ant quantum mechanics.\n4. In 1963, Dirac used the symmetry of two coupled oscillator s to construct the O(3,2)\ngroup [22].\nIn thefourth paperpublishedin 1963, Dirac considered two c oupled oscillators usingstep-\nup and step-down operators. He then constructed the Lie alge bra (closed set of commutation\nrelations for the generators) of thede Sitter group, also kn own asO(3,2), usingten quadratic\n2Table 1: Lorentz covariance of particles both massive and massless . The little group of Wigner\nunifies, for massive and massless particles, the internal space-tim e symmetries. The challenge for\nus is to find another unification: that which unifies, in the physics of t he high-energy realm, both\nthe quark and parton pictures. In this paper, we achieve this purp ose by integrating Dirac’s three\npapers. A similar table was published in Ref. [11].\nMassive, Slow COVARIANCE Massless, Fast\nEnergy- Einstein’s\nMomentum E=p2/2m E =/radicalig\n(cp)2+(mc2)2E=cp\nInternal S3 S3\nSpace-time Wigner’s Gauge\nSymmetry S1, S2 Little Groups Transformation\nRelativistic Integration\nExtended Quark Model of Dirac’s papers Parton Model\nParticles 1927, 1945, 1949\n3forms of those step-up and step-down operators. The harmoni c oscillator is a language of\nquantum mechanics while the de Sitter group is a language of t he Lorentz group or Einstein’s\nspecial relativity. Thus, his 1963 paper [22] provides the fi rst step toward a unified view of\nquantum mechanics and special relativity.\nIn spite of all those impressive aspects, the above-listed p apers are largely unknown in the\nphysics world. The reason is that there are soft spots in thos e papers. Dirac firmly believed\nthat one can construct physical theories only by constructi ng beautiful mathematics [23].\nHis Dirac equation is a case in point. Indeed, all of his paper s are like poems [24]. They are\nmathematical poems. However, there are things he did not do.\nFirst, his papers do not contain any graphical illustration s. For instance, his light-cone\ncoordinate system could be illustrated as a squeeze transfo rmation, but he did not draw a\npicture [21]. When he talked about the Gaussian function [20 ] in the space-time variables,\nhe could have used a circle in the two-dimensional space of th e longitudinal and time-like\nvariables.\nSecond, Dirac seldom made reference to his own earlier paper s. In his 1945 paper [20],\nthere is a distribution along the time-like variable, but he did not mention his earlier paper\nof 1927 [19] where the time-energy uncertainty was discusse d. In his 1949 paper, when he\nproposed his instant form , he eliminated all time-like excitations, but he forgot to m ention\nhis c-number time-energy uncertainty relation he formulat ed in his earlier paper of 1927 [19].\nDirac’s wife was Eugene Wigner’s younger sister. Dirac thus had many occasions to meet\nhis brother-in-law. Dirac sometimes quoted Wigner’s paper on the inhomogeneous Lorentz\ngroup [13], but without making any serious references, in sp ite of the fact that Wigner’s\nO(3)-like little group is the same as his own instant form ment ioned in his 1949 paper [21].\nPaul A. M. Dirac is an important person in the history of physi cs. It is thus important\nto examine what conclusions we can draw if we integrate all of those papers by closing up\ntheir soft spots.\nIn Section 2, we list four important papers Dirac published f rom 1927 to 1963 [19, 20, 21,\n22]. We then point out his original ideas contained therein.\nIn 1971 [25], Feynman, Kislinger, and Ravndal published a pa per saying that although\nquantum field theory works for scattering problems with runn ing waves, harmonic oscillators\nmay be useful for studying bound states in the relativistic w orld. They then formulated\na Lorentz-invariant differential equation separable into a K lein–Gordon equation for a free\nhadron, and a Lorentz-invariant oscillator equation for th e bound state of the quarks. How-\never, theirsolutionoftheoscillator equationisnotnorma lizableandisphysicallymeaningless.\nIn Section 3 we discuss their paper.\nIn Section 4, we construct normalizable harmonic oscillato r wave functions. These wave\nfunctions are not Lorentz-invariant, because the shape of t he wave function changes as it\nis Lorentz-boosted. However, the wave function is Lorentz- covariant under Lorentz trans-\nformations. It is shown further that these Lorentz-covaria nt wave functions constitute a\nrepresentation of Wigner’s O(3)-like little group for massive particles [13].\n4In Section 5, the covariant oscillator wave functions are ap plied to hadrons moving with\nrelativistic speed. It is noted that the wave function becom es squeezed along Dirac’s light-\ncone system [21]. It is shown that this squeeze property is re sponsible for the dipole cut-off\nbehavior of the proton form factor [26]. It is shown further t hat Gell-Mann’s quark model [8]\nand Feynman’s parton model [9, 27] are two limiting cases of o ne Lorentz-covariant entity,\nas in the case of Einstein’s E=mc2[10, 11].\nIn Section 6, it is shown that the two-variable covariant har monic oscillator wave function\ncan serve as the formula for two-photon coherent states comm only called squeezed states of\nlight [14]. It is then noted that Dirac, in his 1963 paper [22] , constructed the Lie algebra of\ntwo-photon states with ten generators. Dirac noted further that these generators satisfy the\nLie algebra of the O(3,2) de Sitter group.\nTheO(3,2) de Sitter group is a Lorentz group applicable to three spac e-like dimensions\nand two time-like dimensions. Thereare ten generators for t his group. If we restrict ourselves\nto one of the time-variables, it becomes the familiar Lorent z group with three rotation and\nthree boost generators. These six generators lead to the Lor entz group familiar to us. The\nremainingfourgeneratorsareforthreeLorentzboostswith respecttothesecondtimevariable\nand one rotation generator between the two time variables.\nIn Section 7, we contract these last four generators into fou r space-time translation gener-\nators. Thus, it is possible to transform Dirac’s O(3,2) group to the inhomogeneous Lorentz\ngroup [28, 29]. In this way, we show that quantum mechanics an d special relativity share the\nsame symmetry ground. Based on Dirac’s four papers listed in this section, we venture to\nsay that this was Dirac’s ultimate purpose.\n2 Dirac’sEffortstoMake QuantumMechanicsLorentz-\nCovariant\nPaul A. M. Dirac made it his lifelong effort to formulate quantu m mechanics so that it would\nbe consistent with special relativity. In this section, we r eview four of his major papers\non this subject. In each of these papers, Dirac points out fun damental difficulties in this\nproblem.\nDirac noted, in 1927 [19], that the emission of photons from a toms is a manifestation of\nthe uncertainty relation between the time and energy variab les. He also noted that, unlike\nthe uncertainty relation of Heisenberg which allows quantu m excitations, the time or energy\naxis has no excitations along it. Hence, when attempting to c ombine these two uncertainty\nrelations in the Lorentz-covariant world, there remains th e serious difficulty that the space\nand time variables are linearly mixed.\nSubsequently in 1945 [20], Dirac, using the four-dimension al harmonic oscillator at-\ntempted to construct, using the oscillator wave functions, a representation of the Lorentz\ngroup. When he did this, however, the wave functions which re sulted did not appear to be\n5Lorentz-covariant.\nUsing the ten generators of the inhomogeneous Lorentz group Dirac in 1949 [21], con-\nstructed from them three forms of relativistic dynamics. Ho wever, after imposing subsidiary\nconditions necessitated by the existing form of quantum mec hanics, he found inconsistencies\nin all three of the forms he considered.\nIn 1963 [22], Dirac constructed a representation of the O(3,2) de Sitter group. To accom-\nplish this, he used two coupled harmonic oscillators in the f orm of step-up and step-down\noperators. Thus Dirac constructed a beautiful algebra. He d id not, however, attempt to\nexploit the physical contents of his algebra.\nInspiteoftheshortcomingsmentionedabove, itisindeedre markablethatDiracworkedso\ntirelessly on this important subject. We are interested in c ombiningall of his works to achieve\nhis goal of making quantum mechanics consistent with specia l relativity. Let us review the\ncontents of these papers in detail, by transforming Dirac’s formulae into geometrical figures.\n2.1 Dirac’s C-Number Time-Energy Uncertainty Relation\nIt was Wigner who in 1972 [30] drew attention to the fact that t he time-energy uncertainty\nrelation, known from the transition time and line broadenin g in atomic spectroscopy, existed\nbefore 1927. This occurred even before the uncertainty prin ciple that Heisenberg formulated\nin 1927. Also in 1927 [19], Dirac studied the uncertainty rel ation which was applicable to\nthe time and energy variables. When the uncertainty relatio n was formulated by Heisenberg,\nDirac considered the possibility of whether a Lorentz-cova riant uncertainty relation could be\nformed out of the two uncertainty relations [19].\nDirac then noted that the time variable is a c-number and thus there are no excitations\nalong the time-like direction. However, there are excitati ons along the space-like longitudinal\ndirection starting from the position-momentum uncertaint y. Since the space and time coor-\ndinates are mixed up for moving observers, Dirac wondered ho w this space-time asymmetry\ncould be made consistent with Lorentz covariance. This was i ndeed a major difficulty.\nDirac, however, neveraddressed,eveninhislaterpapers,t heseparationintimevariableor\nthe time interval. On the other hand, the Bohr radius, which m easures the distance between\nthe proton and electron is an example of Heisenberg’s uncert ainty relation, applicable to\nspace separation variables.\nIn his 1949 paper [21] Dirac discusses his instant form of relativistic dynamics. Thus\nDirac came back to this question of the space-time asymmetry , in his 1949 paper. There\nhe addresses indirectly the possibility of freezing three o f the six parameters of the Lorentz\ngroup, and hence only working with the remaining three param eters. Wigner, in this 1939\npaper [13, 2] already presented this idea. He had observed in that paper that his little groups\nwith three independent parameters dictated the internal sp ace-time symmetries of particles.\n62.2 Dirac’s Four-Dimensional Oscillators\nSince the language of special relativity is the Lorentz grou p, and harmonic oscillators pro-\nvide a start for the present form of quantum mechanics, Dirac , duringthe second World War,\nconsidered the possibility of using harmonic oscillator wa ve functions to construct represen-\ntations of the Lorentz group [20]. He considered that, by con structing representations of\nthe Lorentz group using harmonic oscillators, he might be ab le to make quantum mechanics\nLorentz-covariant.\nThus in his 1945 paper [20], Dirac considers the Gaussian for m\nexp/parenleftbigg\n−1\n2/bracketleftig\nx2+y2+z2+t2/bracketrightig/parenrightbigg\n. (1)\nThexandyvariables can be dropped from this expression, as we are cons idering a Lorentz\nboost only along the zdirection. Therefore we can write the above equation as:\nexp/parenleftbigg\n−1\n2/bracketleftig\nz2+t2/bracketrightig/parenrightbigg\n. (2)\nSince/parenleftbigz2−t2/parenrightbigis an invariant quantity, the above expression may seem stra nge for those who\nbelieve in Lorentz invariance.\nIn his 1927 paper [19] Dirac proposed the time-energy uncert ainty relation, but observed\nthat. because time is a c-number, there are no excitations al ong the time axis. Hence the\nabove expression is consistent with this earlier paper.\nIf we look carefully at Figure 1, we see that this figure is a pic torial illustration of Dirac’s\nEquation (2). There is localization in both space and time co ordinates. Dirac’s fundamental\nquestion, illustrated in Figure 2, would then be how to make t his figure covariant. Dirac\nstops there. However, this is not the end of his story.\n2.3 Dirac’s Light-Cone Coordinate System\nThe Reviews of Modern Physics, in 1949, celebrated Einstein ’s 70th birthday by publish-\ning a special issue. In this issue was included Dirac’s paper entitledForms of Relativistic\nDynamics [21]. Here Dirac introduced his light-cone coordinate syst em. In this system a\nLorentz boost is seen to be a squeeze transformation, where o ne axis expands while the other\ncontracts in such a way that their product remains invariant as shown in Figure 3.\nWhen boosted along the zdirection, the system is transformed into the following for m:\n/parenleftbiggz′\nt′/parenrightbigg\n=/parenleftbiggcosh(η) sinh(η)\nsinh(η) cosh(η)/parenrightbigg/parenleftbiggz\nt/parenrightbigg\n. (3)\nDirac defined his light-cone variables as [21]\nz+=z+t√\n2, z −=z−t√\n2. (4)\n7Figure 1: Quantum mechanics represented in terms of space-time. As can be seen, there are no\nexcitations along the time-like direction, but quantum excitations alo ng the space-like longitudinal\ndirection are allowed.\nFigure 2: Dirac’s four-dimensional oscillators localized in a closed spac e-time region. This is not\na Lorentz-invariant concept. How about Lorentz covariance?\n8A=4u'v'u\nv\nA=4uv\n=2(t2–z2)zt Dirac 1949\nFigure 3: The light-cone coordinate system pictured with a Lorentz boost. Not only does the\nboost squeeze the square into a rectangle, but it traces a point alo ng the hyperbola.\nThen the form of the boost transformation of Equation (3) bec omes\n/parenleftbiggz′\n+\nz′\n−/parenrightbigg\n=/parenleftbiggeη0\n0e−η/parenrightbigg/parenleftbiggz+\nz−/parenrightbigg\n. (5)\nIt is then apparent that uvariable becomes expanded, but the vvariable becomes contracted.\nWe illustrate this in Figure 3. The product then becomes:\nz+z−=1\n2(z+t)(z−t) =1\n2/parenleftig\nz2−t2/parenrightig\n(6)\nwhich remains invariant. The Lorentz boost is therefore, in Dirac’s picture, a squeeze trans-\nformation.\nDirac also introduced his 1949 paper his instant form of relativistic quantum mechanics.\nThis has the condition\nx0≈0. (7)\nWhat did his approximate equality mean? In this paper, we int erpret the nature of the\ntime-energy uncertainty relation in terms of his c-number. Furthermore, it could mean that\nit is, for the massive particle is the three-dimensional rot ation group, Wigner’s little group,\nwithout the time-like direction.\nAdditionally, Diracstatedthatconstructingarepresenta tionoftheinhomogeneousLorentz\ngroup, was necessary to construct a relativistic quantum me chanics. The inhomogeneous\n9Lorentz group has ten generators, four space-time translat ion generators, three rotation gen-\nerators, and three boost generators, which satisfy a closed set of commutation relations.\nIt is now clear that Dirac was interested in using harmonic os cillators to construct a\nrepresentation of the inhomogeneous Lorentz group. In 1979 , together with another author,\nthe present authors published a paper on this oscillator-ba sed representation [12]. We regret\nthat we did not mention there Dirac’s earlier efforts along thi s line.\n3 Scattering and Bound States\nFrom the three papers written by Dirac [19, 20, 21], let us find out what he really had in\nmind. In physical systems, there are scattering and bound st ates. Throughout his papers,\nDirac did not say he was mainly interested in localized bound systems. Let us clarify this\nissue using the formalism of Feynman, Kislinger, and Ravnda l [25].\nWe are quite familiar with the Klein–Gordon equation for a fr ee particle in the Lorentz-\ncovariant world. We shall use the four-vector notations\nxµ= (x,y,z,t),andx2\nµ=x2+y2+z2−t2. (8)\nThen the Klein–Gordon equation becomes\n\n−/bracketleftigg\n∂\n∂xµ/bracketrightigg2\n+m2\nφ(x) = 0. (9)\nThe solution of this equation takes the familiar form\nexp[±i(p1x+p2y+p3z±Et)]. (10)\nIn1971, Feynmanetal.[25]consideredtwoparticlesaandbb oundtogetherbyaharmonic\noscillator potential, and wrote down the equation\n\n\n−/bracketleftigg\n∂\n∂xaµ/bracketrightigg2\n−/bracketleftigg\n∂\n∂xbµ/bracketrightigg2\n+(xaµ−xbµ)2+m2\na+m2\nb\n\nφ(xaµ,xbµ) = 0.(11)\nThe bound state of these two particles is one hadron. The constituent particles are called\nquarks. We can then define the four-coordinate vector of the hadron a s\nX=1\n2(xa+xb), (12)\nand the space-time separation four-vector between the quar ks as\nx=1\n2√\n2(xa−xb). (13)\n10Then Equation (11) becomes\n\n\n−/bracketleftigg\n∂\n∂Xµ/bracketrightigg2\n+m2\n0+\n−/bracketleftigg\n∂\n∂xµ/bracketrightigg2\n+x2\nµ\n\n\nφ(X,x) = 0. (14)\nThis differential equation can then be separated into\n\n−/bracketleftigg\n∂\n∂Xµ/bracketrightigg2\n+m2\n0\nφ(X,x) =−\n−/bracketleftigg\n∂\n∂xµ/bracketrightigg2\n+x2\nµ\nφ(X,x), (15)\nwith\nφ(X,x) =f(X)ψ(x), (16)\nwheref(X) andψ(x) satisfy their own equations:\n\n−/bracketleftigg\n∂\n∂Xµ/bracketrightigg2\n+m2\na+m2\nb+λ\nf(X) = 0 (17)\nand\n1\n2\n−/bracketleftigg\n∂\n∂xµ/bracketrightigg2\n+x2\nµ\nψ(x) =λψ(x). (18)\nHere, the wave function then takes the form\nφ(X,x) =ψ(x)exp[±i(PxX+PyY+PzZ±ET)], (19)\nwherePx,Py,Pzare for the hadronic momentum, and\nE2=P2\nx+P2\ny+P2\nz+M2,withM2=m2\na+m2\nb+λ. (20)\nHere the hadronic mass Mis determined by the parameter λ, which is the eigenvalue of the\ndifferential equation for ψ(x) given in Equation (18).\nConsidering Feynman diagrams based on the S-matrix formali sm, quantum field theory\nhas been quite successful. It is, however, only useful for ph ysical processes where, after\ninteraction, one set of free particles becomes another set o f free particles. The questions\nof localized probability distributions and their covarian ce under Lorentz transformations is\nnot addressed by quantum field theory. In order to tackle this problem and address these\nquestions, Feynman et al. suggested harmonic oscillators [ 25]. In Figure 4, we illustrate this\nidea.\nHowever, for their wave function ψ(x), Feynman et al. uses a Lorentz-invariant exponen-\ntial form\nexp/parenleftbigg\n−1\n2/bracketleftig\nx2+y2+z2−t2/bracketrightig/parenrightbigg\n. (21)\n11Figure4: Feynman, in aneffort to combine quantum mechanics with sp ecial relativity, gave us this\nroadmap. Feynman’s diagrams provide, in Einstein’s world, a satisfac tory resolution for scattering\nstates. Thus they work for running waves. Feynman suggested t hat harmonic oscillators should\nbe used as a first step for representing standing waves trapped in side an extended hadron.\nThis wave function increases as tbecomes large. This is not an acceptable wave function.\nThey overlooked the normalizable exponential form given by Dirac in Equation (1). They\nalso overlooked the form in the paper of Fujimura et al. [31] w hich was quoted in their own\npaper.\nThus, we are interested in fixing this problem and thus constr ucting Lorentz-covariant\noscillator wave functions satisfying both the rules of quan tum mechanics and the rules of\nspecial relativity.\n4 Lorentz-CovariantPictureofQuantumBoundStates\nIn 1939, Wigner considered internal space-time symmetries of particles in the Lorentz-\ncovariant world [13]. For this purpose, he considered the su bgroups of the Lorentz group\nfor a given four-momentum of the particle. For the massive pa rticle, the internal space-time\nsymmetry is defined for when the particle is at rest, and its sy mmetry is dictated by the\nthree-dimensional rotation group, which allows us to define the particle spin as a dynamical\nvariable, as indicated in Table 1.\nLet us go to the wave function of Equation (19). This wave func tionψ(x) is for the\ninternal coordinates of the hadron, and the exponential for m defines the hadron momentum.\nThus, the symmetry of Wigner’s little group is applicable to the wave function ψ(x).\nThis situation is like the case of the Dirac equation for a fre e particle. Its solution is a\nplane wave times the four-component Dirac spinor describin g the spin orientation and its\nLorentz covariance. This separation of variables for the pr esent case of relativistic extended\nparticles is illustrated in Figure 4. With this understandi nglet us write the Lorentz-invariant\n12differential equation as\n1\n2/bracketleftigg\n−∂2\n∂x2−∂2\n∂y2−∂2\n∂z2+∂2\n∂t2+/parenleftig\nx2+y2+z2−t2/parenrightig/bracketrightigg\nψ(x,y,z,t) =λψ(x,y,z,t).(22)\nHere, the variables x, y, zare for the spatial separation between the quarks, like the B ohr\nradius in the hydrogen atom. The time variable tis thetime separation between the quarks.\nThis variable is very strange, because it does not exist in th e present forms of quantum\nmechanics and quantum field theory. Paul A. M. Dirac did not me ntion this time separation\nin any of his papers quoted here. Yet, it plays the major role i n the Lorentz-covariant world,\nbecause the spatial separation (like the Bohr radius) picks up a time-like component when\nthe system is Lorentz-boosted [34].\nIn his 1927 paper [19], Dirac mentioned the c-number time-en ergy uncertainty relation,\nand he used t≈0 in his 1949 paper for his instant form of relativistic dynam ics. When he\nwrote down a Gaussian function with/parenleftbigx2+y2+z2+t2/parenrightbigas the exponent, he should have\nmeantx, yandzare for the space separation and tfor the time separation, since otherwise\nthe system becomes zero in the remote future and remote past.\nWith this understanding, we are dealing here with the soluti on of the differential equation\nof the form\nψ(x) =f(x,y,z)exp/parenleftigg\n−t2\n2/parenrightigg\n, (23)\nwheref(x,y,z) satisfy the oscillator differential equation\n1\n2/parenleftigg\n−∂2\n∂x2−∂2\n∂y2−∂2\n∂z2+x2+y2+z2/parenrightigg\nf(x,y,z) =/parenleftbigg\nλ−1\n2/parenrightbigg\nf(x,y,z).(24)\nThe form of Equation (23) tells us there are no time-like exci tations. This equation is the\nSchr¨ odinger equation for the three-dimensional harmonic oscillator and its solutions are well\nknown.\nIf we use the three-dimensional spherical coordinate syste m, the solution will give the\nspin or internal angular momentum and the orientation of the bound-state hadron [12]. This\nspherical form is for the O(3) symmetry of Wigner’s little group for massive particles . The\nLorentz-invariant Casimir operators are given in Ref. [12] .\nIf we are interested in Lorentz-boosting the wave function, we note that the original wave\nequation of Equation (22) is separable in all four variables . If the Lorentz boost is made\nalong thezdirection, the wave functions along the xandydirections remain invariant, and\nthus can be separated. We can study only the longitudinal and time-like components. Thus\nthe differential equation of Equation (22) is reduced to\n1\n2/bracketleftigg\n−∂2\n∂z2+∂2\n∂t2+/parenleftig\nz2−t2/parenrightig/bracketrightigg\nψ(z,t) =λψ(z,t). (25)\n13The solution of this differential equation takes the form\nψ(z,t) =Hn(z)exp/parenleftigg\n−/bracketleftigg\nz2+t2\n2/bracketrightigg/parenrightigg\n, (26)\nwhereHn(z) istheHermitepolynomial. Therearenoexcitations in t, thereforeitisrestricted\nto the ground state. For simplicity, we ignore the normaliza tion constant.\nIf this wave function is Lorentz-boosted along the zdirection, the zandtvariables in\nthis expression should be replaced according to\nz→(coshη)z−(sinhη)t,andt→(coshη)t−(sinhη)z. (27)\nAccording to the light-cone coordinate system introduced b y Dirac in 1949 [20], this trans-\nformation can be written as\n(z+t)→e−η(z+t),and (z−t)→eη(z−t). (28)\nThus the Lorentz-boosted wave function becomes\nψη(z,t) =Hn/parenleftbigg1√\n2/bracketleftbige−η(z+t)+eη(z−t)/bracketrightbig/parenrightbigg\nexp/parenleftbigg\n−1\n4/bracketleftig\ne−2η(z+t)2+e2η(z−t)2/bracketrightig/parenrightbigg\n,(29)\nwithout the normalization constant.\nIt is possible to write this wave function using the one-dime nsional normalized oscillator\nfunctionsφn(z) andφk(t) as\nψη(z,t) =/summationdisplay\nnkAnkφn(z)φk(t). (30)\nThis problem has been extensively discussed in the literatu re [2, 6, 7, 35].\nThe most interesting case is the expansion of the ground stat e. The normalized wave\nfunction is then\nψη(z,t) =/parenleftbigg1\nπ/parenrightbigg1/2\nexp/parenleftbigg\n−1\n4/bracketleftig\ne−2η(z+t)2+e2η(z−t)2/bracketrightig/parenrightbigg\n. (31)\nThe Lorentz-boost property of this form is illustrated in Fi gure 5. The expansion in the\nharmonic oscillator wave functions takes the form\nψη(z,t) =1\ncoshη/summationdisplay\nn(tanhη)nφn(z)φn(t). (32)\nThis expression is the key formula for two-photon coherent s tates or squeezed states of\nlight [14]. We shall return to this two-photon problem in Sec tion 6.\n14As for the wave function ψ(x), this is the localized wave function Dirac was considering\nin his papers of 1927 and 1945. If Figures 2 and 3 are combined, we end up with an ellipse\nas a squeezed circle as shown in Figure 5.\nIndeed, one of the most controversial issues in high-energy physics is explained by this\nsqueezed circle. The bound-state quantum mechanics of prot ons, which are known to be\nbound states of quarks, are often assumed to be the same as tha t of the hydrogen atom.\nThen how would the proton look to an observer on a train become s the question. According\nto Feynman [9, 27], when the speed of the train becomes close t o that of light, the proton\nappears like a collection of partons. However, the properti es of Feynman’s partons are quite\ndifferent from the properties of the quarks. This issue shall b e discussed in more detail in\nSection 5.\n5 Lorentz-Covariant Quark Model\nEarly successes in the quark model include the calculation o f the ratio of the neutron and\nproton magnetic moments [32], and the hadronic mass spectra [25, 33]. These are based on\nhadrons at rest. We are interested in this paper how the hadro ns in the quark model appear\nto observers in different Lorentz frames.\nThese days, modern particle accelerators routinely produc e protons moving with speeds\nvery closetothat of light. Thus, thequestionis thereforew hetherthecovariant wave function\ndeveloped in Section 4 can explain the observed phenomena as sociated with those protons\nmoving with relativistic speed.\nThe idea that the proton or neutron has a space-time extensio n had been developed long\nbeforeGell-Mann’s proposal for the quark model [8]. Yukawa [36] developed this idea as early\nas 1953, and his idea was followed up by Markov [37], and by Gin zburg and Man’ko [38].\nSince Einstein formulated his special relativity for point particles, it has been and still\nis a challenge to formulate a theory for particles with space -time extensions. The most\nnaive idea would be to study rigid spherical objects, and the re were many papers on this\nsubjects. But we do not know where that story stands these day s. We can however replace\nthese extended rigid bodies by extended wave packets or stan ding waves, thus by localized\nprobability entities. Then what are the constituents withi n those localized waves? The quark\nmodel gives the natural answer to this question.\nHofstadter and McAllister [39], by using electron-proton s cattering to measure the charge\ndistribution inside the proton, made the first experimental discovery of the non-zero size of\nthe proton. If the proton were a point particle, the scatteri ng amplitude would just be a\nRutherford formula. However, Hofstadter and MacAllister f ound a tangible departure from\nthis formula which can only be explained by a spread-out char ge distribution inside the\nproton.\nInthissection, weareinterestedinhowwellthebound-stat epicturedevelopedinSection4\n15Figure 5: Lorentz covariance of the internal the wave function ψη(z,t) of Equation (31). Syn-\nthesizing Figures 2 and 3, we obtain a squeezed circle as shown in this fi gure. This figure thus\nintegrates Dirac’s three papers [19, 20, 21].\n16works in explaining relativistic phenomena of those proton s. In Section 5.1 we study in detail\nhow the Lorentz squeeze discussed in Section 4 can explain th e behavior of electron-proton\nscattering as the momentum transfer becomes relativistic.\nSecond, we note that the proton is regarded as a bound state of the quarks sharing the\nsame bound-state quantum mechanics with the hydrogen atom. However, it appears as a\ncollection of Feynman’s partons. Thus, it is a great challen ge to see whether one Lorentz-\ncovariant formula can explain both the static and light-lik e protons. We shall discuss this\nissue in Section 5.2\nOne hundred years ago, Einstein and Bohr met occasionally to discuss physics. Bohr was\ninterested in how the election orbit looks and Einstein was w orrying about how things look\nto moving observers. Did they ever talk about how the hydroge n atom appears to moving\nobservers? In Section 5.3, we briefly discuss this issue.\nIt is possible to conclude from the previous discussion that the Lorentz boost might\nincrease the uncertainty. To deal with this, in Section 5.4, we address the issue of the\nuncertainty relation when the oscillator wave functions ar e Lorentz-boosted.\n5.1 Proton Form Factor\nUsing the Born approximation for non-relativistic scatter ing, we see what effect the charge\ndistribution has on the scattering amplitude. When electro ns are scattered from a fixed\ncharge distribution with a density of eρ(r), the scattering amplitude becomes:\nf(θ) =−/parenleftigg\ne2m\n2π/parenrightigg/integraldisplay\nd3xd3x′ρ(r′)\nRexp(−iQ·x). (33)\nHere we use r=|x|,R=|r−r′|,andQ=Kf−Ki,which is the momentum transfer. We\ncan reduce this amplitude to:\nf(θ) =2me2\nQ2F(Q2). (34)\nThe density function’s Fourier transform is given by F(Q2) which can be written as:\nF/parenleftig\nQ2/parenrightig\n=/integraldisplay\nd3xρ(r)exp(−iQ·x). (35)\nThis above quantity is called the form factor and it describe s the charge distribution in terms\nof the momentum transfer which can be normalized by:\n/integraldisplay\nρ(r)d3x= 1. (36)\nTherefore, from Equation (35), F(0) = 1. the scattering ampl itude of Equation (33) be-\ncomes the Rutherford formula for Coulomb scattering if the d ensity function, corresponding\n17Proton\nPElectron\nExchange PhotonK\nKif\nPf\ni\nFigure 6: Electron-proton scattering in the Breit frame. The outg oing momentum of the proton\nis opposite in sign but equal in magnitude to that of the incoming proto n.\nto a point charge, is a delta function, F/parenleftbigQ2/parenrightbig= 1, for all values of Q2. Increasing values\nofQ2, which are deviations from Rutherford scattering, give a me asure of the charge distri-\nbution. Hofstadter’s experiment, which scattered electro ns from a proton target, found this\nprecisely [39].\nWhen the energy of the incoming electron becomes higher, it i s necessary to take into\naccount the recoil effect of target proton. It then requires th at the problem be formulated\nin the Lorentz-covariant framework. It is generally agreed that quantum electrodynamics\ncan describe electrons and their electromagnetic interact ion by using Feynman diagrams for\npractical calculations. When using perturbation, a power s eries of the fine structure constant\nα.is used to expand the scattering amplitude. Therefore, the l owest order in α,can, using\nthe diagram given in Figure 6, describe the scattering of an e lectron by a proton.\nMany textbooks on elementary particle physics [40, 41] give the corresponding matrix\nelement as the form:\n¯U(Pf)Γµ(Pf,Pi)/parenleftbigg1\nQ2/parenrightbigg\n¯U(Kf)γµU(Ki), (37)\nwhere the initial and final four-momenta of the proton and ele ctron, respectively, are given\nbyPi,Pf,KiandKf. The Dirac spinor for the initial proton is U(Pi), while the (four-\nmomentum transfer)2,Q2, is\nQ2= (Pf−Pi)2= (Kf−Ki)2. (38)\nIf the proton were a point particle like the electron, the Γ µwould beγµ, but for a\nparticle, like the proton, with space-time extension, it is γµF/parenleftbigQ2/parenrightbig. The virtual photon being\n18exchanged between the electron and the proton produces the/parenleftbig1/Q2/parenrightbigfactor in Equation (37).\nFor the particles involved in the scattering process this qu antity is positive for physical values\nof the four-momenta in the metric we use.\nFrom the definition of the form factor given in Equation (35), we can make a relativistic\ncalculation of the form factor. The density function, which depends only on the target\nparticle, is proportional to ψ(x)†ψ(x). The wave function for quarks inside the proton is\nψ(x). This expression is a special case of the more general form\nρ(x) =ψ†\nf(x)ψi(x), (39)\nwhere the initial and final wave function of the target atom is given byψiandψf. The form\nfactor of Equation (35) can then be written as\nF/parenleftig\nQ2/parenrightig\n=/parenleftig\nψf(x),e−iQ·rψi(x)/parenrightig\n. (40)\nThe required Lorentz generalization, starting from this ex pression, can be made using the\nrelativistic wave functions for hadrons.\nBy replacing each quantity in the expression of Equation (35 ) by its relativistic counter-\npart we should be able to see the details of the transition to r elativistic physics. If we go\nback to the Lorentz frame in which the momenta of the incoming and outgoing nucleons have\nequal magnitude but opposite signs, we obtain\npi+pf= 0. (41)\nThis kinematical condition is illustrated in Figure 6.\nWe call the Lorentz frame in which the above condition holds t he Breit frame. As illus-\ntrated in Figure 6, there is no loss of generality if the proto n comes in along the zdirection\nbefore the collision and goes out along the negative zdirection after the scattering process.\nThe four vector, Q= (Kf−Ki) = (Pi−Pf), n this frame, has no time-like component. The\nLorentz-invariant form, Q·x, can thus replace the exponential factor Q·r. The covariant\nharmonic oscillator wave functions discussed in this paper can be used for the wave func-\ntions for the protons, assuming that the nucleons are in the g round state. Then the integral\nin the evaluation of Equation (35), which includes the time- like direction and is thus four-\ndimensional, is the only difference between the non-relativi stic and relativistic cases. The\nexponential factor, which does not depend on the time-separ ation variable, therefore, is not\naffected by the integral in the time-separation variable.\nWe can now consider the integral:\ng/parenleftig\nQ2/parenrightig\n=/integraldisplay\nd4xψ†\n−η(x)ψη(x)exp(−iQ·x), (42)\nwhere tanh ηis the velocity parameter (tanh η=v/c) for the incoming proton, and the wave\nfunctionψη, from Equation (31), takes the form:\nψη(x) =1√πexp/parenleftbigg\n−1\n4/bracketleftig\ne−2η(z+t)2+e2η(z−t)2/bracketrightig/parenrightbigg\n. (43)\n19We can, now, after the above decomposition of the wave functi ons, perform the integra-\ntions in the xandyvariables trivially. We can, after dropping these trivial f actors, write the\nproduct of the two wave functions as\nψ†\n−η(x)ψη(x) =1\nπexp/parenleftig\n−[cosh(2η)]/bracketleftig\nt2+z2/bracketrightig/parenrightig\n. (44)\nNow thezandtvariables are separated. Since the tintegral in Equation (42), as the\nexponential factor in Equation (35), does not depend on t, it can also be trivially performed.\nThen integral of Equation (42) can be written as\ng/parenleftig\nQ2/parenrightig\n=1/radicalbig\nπcosh(2η)/integraldisplay\ne−2iPzexp/parenleftig\n−cosh(2η)z2/parenrightig\ndz. (45)\nHere thezcomponent of the momentum of the incoming proton is P. Thevariable Q2, which\nis the (momentum transfer)2, now become 4 P2. The hadronic material, which is distributed\nalong the longitudinal direction, has indeed became contra cted [42].\nWe note that tanh ηcan be written as\n(tanhη)2=Q2\nQ2+4M2, (46)\nwhereMis the proton mass. This equation tells us that β= 0 whenQ2= 0, whileit becomes\none asQ2becomes infinity.\nThe evaluation of the above integral for g/parenleftbigQ2/parenrightbigin Equation (45) leads to\ng/parenleftig\nQ2/parenrightig\n=/parenleftigg\n2M2\nQ2+2M2/parenrightigg\nexp/parenleftigg\n−Q2\n2(Q2+2M2)/parenrightigg\n. (47)\nForQ2= 0,the above expression becomes 1. It decreases as\ng/parenleftig\nQ2/parenrightig\n∼1\nQ2(48)\nasQ2assumes large values.\nSo far the calculation has been performed for an oscillator b ound state of two quarks.\nThe proton, however, consists of three quarks. As shown in th e paper of Feynman et al. [25],\nthe problem becomes a product of two oscillator modes. Thus, the three-quark system is a\nstraightforward generalization of the above calculation. As a result, the form factor F/parenleftbigQ2/parenrightbig\nbecomes;\nF/parenleftig\nQ2/parenrightig\n=/parenleftigg\n2M2\nQ2+2M2/parenrightigg2\nexp/parenleftigg\n−Q2\nQ2+2M2/parenrightigg\n, (49)\n20which is 1 at Q2= 0, and decreases as\nF/parenleftig\nQ2/parenrightig\n∼/bracketleftbigg1\nQ2/bracketrightbigg2\n(50)\nasQ2assumeslargevalues. Thisformfactorfunctionhastherequ ireddipole-cut-off behavior,\nwhich has been observed in high-energy laboratories. This c alculation was carried first by\nFujimura et al. in 1970 [31].\nLet us re-examine the above calculation. If we replace βby zero in Equation (45) and\nignore the elliptic deformation of the wave functions, g/parenleftbigQ2/parenrightbigwill become\ng/parenleftig\nQ2/parenrightig\n= exp/parenleftig\n−Q2/4/parenrightig\n, (51)\nwhich will lead to an exponential cut-off of the form factor. T his is not what we observe in\nlaboratories.\nIn order to gain a deeper understanding of the above-mention ed correlation, let us study\nthe case using the momentum-energy wave functions:\nφη(q) =/parenleftbigg1\n2π/parenrightbigg2/integraldisplay\nd4xe−iq·xψη(x). (52)\nIf we ignore, as before, the transverse components, we can wr iteg/parenleftbigQ2/parenrightbigas [2]\n/integraldisplay\ndq0dqzφ∗\nη(q0,qz−P)φη(q0,qz+P). (53)\nThe above overlap integral has been sketched in Figure 7. The two wave functions overlap\ncompletely in the qzq0plane ifQ2= 0 orP= 0. The wave functions become separated when\nP increases. Because of the elliptic or squeeze deformation , seen in Figure 7, they maintain a\nsmall overlapping region. In the non-relativistic case, th ere is no overlapping region, because\nthe deformation is not taken into account, as seen in Figure 7 . The slower decrease in Q2\nis, therefore, more precisely given in the relativistic cal culation than in the non-relativistic\ncalculation.\nAlthoughourinteresthasbeeninthespace-timebehaviorof thehadronicwavefunction, it\nmust be noted that quarks are spin-1/2 particles. This fact m ust be taken into consideration.\nThis spin effect manifests itself prominently in the baryonic mass spectra. Here we are\nconcerned with the relativistic effects, therefore, it is nec essary to construct a relativistic spin\nwave function for the quarks. The result of this relativisti c spin wave function construction\nfor the quark wave function should be a hadronic spin wave fun ction. In the case of nucleons,\nthe quark spinsshould becombined in a manner to generate the form factor of Equation (44).\nNaively wecouldusefreeDiracspinorsforthequarks. Howev er, itwasshownbyLipes[43]\nthat using free-particle Dirac spinors leads to a wrong beha vior of the form factor. Lipes’\n21P\nq\nq0\nz Wave Functions\n Squeezed \nNot S quue zed Not Squeezed Overlap\nFigure 7: The momentum-energy wave functions are Lorentz sque ezed in the form factor calcula-\ntion. The two wave functions become separated as the momentum t ransfer increases. However,\nin the relativistic case, the wave functions maintain an overlapping re gion. In the non-relativistic\ncalculation, the wave functions become completely separated. The unacceptable behavior of the\nform factor is caused by this lack of overlapping region.\n22result, however, does not cause us any worry, as quarks in a ha dron are not free particles.\nThus we have to find suitable mechanism in which quark spins, c oupled to orbital motion,\nare taken into account. This is a difficult problem and is a nont rivial research problem, and\nfurther study is needed along this direction [44].\nIn 1960, Frazer and Fulco calculated the form factor using th e technique of dispersion\nrelations[26]. Insodoingtheyhadtoassumetheexistenceo ftheso-called ρmeson, whichwas\nlater found experimentally, and which subsequently played a pivotal role in the development\nof the quark model.\nEven these days, the form factor calculation occupies a very important place in recent\ntheoretical models, such as quantum chromodynamics (QCD) l attice theory [45] and the\nFaddeev equation [46]. However, it is still noteworthy that Dirac’s form of Lorentz-covariant\nbound states leads to the essential dipole cut-off behavior o f the proton form factor.\n5.2 Feynman’S Parton Picture\nAs we did in Section 5.1, we continue using the Gaussian form f or the wave function of the\nproton. If the proton is at rest, the zandtvariables are separable, and the time-separation\ncan be ignored, as we do in non-relativistic quantum mechani cs. If the proton moves with a\nrelativistic speed, the wave function is squeezed as descri bed in Figure 5. If the speed reaches\nthat of light, the wave function becomes concentrated along positive light cone with t=z.\nThe question then is whether this property can explain the pa rton picture of Feynman when\na proton moves with a speed close to that of light.\nIt was Feynman who, in 1969, observed that a fast-moving prot on can be regarded as\na collection of many partons. The properties of these partons appear to be quite different\nfrom those of the quarks [9, 27, 2]. For example, while the num ber of quarks inside a static\nproton is three, the number of partons appears to be infinite i n a rapidly moving proton.\nThe following systematic observations were made by Feynman :\na. When protons move with velocity close to that of light, the parton picture is valid.\nb. Partons behave as free independent particles when the int eraction time between the\nquarks becomes dilated.\nc. Partons have a widespread distribution of momentum as the proton moves quickly.\nd. There seems to be an infinite number of partons or a number mu ch larger than that of\nquarks.\nThequestion is whethertheLorentz-squeezed wave function producedin Figure5 can explain\nall of these peculiarities.\nEach of the above phenomena appears as a paradox, when the pro ton is believed to be a\nbound state of the quarks. This is especially true of (b) and ( c) together. We can ask how a\nfree particle can have a wide-spread momentum distribution .\n23To resolve this paradox, we construct the momentum-energy w ave function corresponding\nto Equation (31). We can construct two independent four-mom entum variables [25] if the\nquarks have the four-momenta paandpb.\nP=pa+pb, q=√\n2(pa−pb). (54)\nSincePis the total four-momentum, it is the four-momentum of the pr oton. The four-\nmomentumseparation between thequarksis measuredby q. We can thenwritethelight-cone\nvariables as\nq+= (q0+qz)/√\n2, q−= (q0−qz)/√\n2. (55)\nThis results in the momentum-energy wave function\nφη(qz,q0) =/parenleftbigg1\nπ/parenrightbigg1/2\nexp/braceleftbigg\n−1\n2/bracketleftig\ne−2ηq2\n++e2ηq2\n−/bracketrightig/bracerightbigg\n. (56)\nSincetheharmonicoscillator isbeingusedhere, itiseasil yseenthattheabovemomentum-\nenergy wave function has the identical mathematical form to that of the space-time wave\nfunction of Equation (31) and that these wave functions also have the same Lorentz squeeze\nproperties. Though discussed extensively in the literatur e [2, 10, 11], these mathematical\nforms and Lorentz squeeze properties are illustrated again in Figure 8 of the present paper.\nWe can see from the figure, that both wave functions behave lik e those for the static\nbound state of quarks when the proton is at rest with η= 0. However, it can also be seen\nthat asηincreases, the wave functions become concentrated along th eir respective positive\nlight-cone axes as they become continuously squeezed. If we look at the z-axis projection\nof the space-time wave function, we see that, as the proton sp eed approaches that of the\nspeed of light, the width of the quark distribution increase s. Thus, to the observer in the\nlaboratory, the position of each quark appears widespread. Thus the quarks appear like free\nparticles.\nIf we look at the momentum-energy wave function we see that it is just like the space-time\nwave function. As the proton speed approaches that of light, the longitudinal momentum\ndistribution becomes wide-spread. In non-relativistic qu antum mechanics we expect that\nthe width of the momentum distribution is inversely proport ional to that of the position\nwave function. This wide-spread longitudinal momentum dis tribution thus contradicts our\nexpectation from non-relativistic quantum mechanics. Our expectation is that free quarks\nmust have a sharply defined momenta, not a wide-spread moment um distribution.\nHowever, as the proton is boosted, the space-time width and t he momentum-energy\nwidth increase in the same direction. This is because of our L orentz-squeezed space-time and\nmomentum-energy wave functions. If we look at Figures 5 and 8 we see that is the effect\nof Lorentz covariance described in these figures. One of the q uark-parton puzzles is thus\nresolved [2, 10, 11].\n24Figure 8: Lorentz-squeezed wave functions in space-time and in mo mentum-energy variables.\nBoth wave functions become concentrated along their respective positive light-cone axes as the\nspeed of the proton approaches that of light. All the peculiarities o f Feynman’s parton picture are\npresented in these light-cone concentrations.\n25Another puzzling problem is that quarks are coherent when th e proton is at rest but the\npartons appear as incoherent particles. We could ask whethe r this means that Lorentz boost\ncoherence is destroyed. Obviously, the answer to this quest ion is NO. The resolution to this\npuzzle is given below.\nWhen the proton is boosted, its matter becomes squeezed. The result is that the wave\nfunctionfortheprotonbecomes, alongthepositivelight-c one axis, concentrated intheelliptic\nregion. The major axis is expanded in length by exp( η), and, as a consequence, the minor\naxis is contracted by exp( −η).\nTherefore we see that, among themselves, the interaction ti me of the quarks becomes\ndilated. As the wave function becomes wide-spread, the ends of the harmonic oscillator well\nincrease in distance from each other. Universally observed in high-energy experiments, it\nwas Feynman [9] who first observed this effect. The oscillation period thus increase like\neη[27, 47].\nSince the external signal, on the other hand, is moving in the direction opposite to the\ndirection of the proton, it travels along the negative light -cone axis with t=−z. As the\nprotoncontracts alongthenegativelight-coneaxis, thein teraction timedecreasesbyexp( −η).\nThen the ratio of the interaction time to the oscillator peri od becomes exp( −2η). Each\nproton, produced by the Fermilab accelerator, used to have a n energy of 900 GeV. This then\nmeans the ratio is 10−6. Because this is such small number, the external signal cann ot sense,\ninside the proton, the interaction of the quarks among thems elves.\n5.3 Historical Note\nThe hydrogen atom played a pivotal role in the development of quantum mechanics. Niels\nBohr devoted much of his life to understanding the electron o rbit of the hydrogen atom.\nBohr met Einstein occasionally to talk about physics. Einst ein’s main interest was how\nthings look to moving observers. Then, did they discuss how t he hydrogen atom looks to a\nmoving observer?\nIf they discussed this problem, there are no records. If they did not, they are excused. At\ntheir time, the hydrogen atom moving with a relativistic spe ed was not observable, and thus\nbeyond the limit of their scope. Even these days, this atom wi th total charge zero cannot be\naccelerated.\nAfter 1950, high-energy accelerators started producing pr otons moving with relativistic\nspeeds. However, the proton is not a hydrogen atom. On the oth er hand, Hofstadter’s\nexperiment [39] showed that the proton is not a point particl e. In 1964, Gell-Mann produced\nthe idea that the proton is a bound-state of more fundamental particles called quarks. It is\nassumed that the proton and the hydrogen share the same quant um mechanics in binding\ntheir constituent particles. Thus, we can study the moving h ydrogen atom by studying the\nmoving proton. This historical approach is illustrated in F igure 9.\nPaul A. M. Dirac was interested in constructing localized wa ve functions that can be\n26Figure 9: Bohr and Einstein, and then Gell-Mann and Feynman. There are no records indicat-\ning that Bohr and Einstein discussed how the hydrogen looks to movin g observers. After 1950,\nwith particle accelerators, the physics world started producing pr otons with relativistic speeds.\nFurthermore, the proton became a bound state sharing the same quantum mechanics with the\nhydrogen atom. The problem of fast-moving hydrogen became tha t of the proton. How would the\nproton appear when it moves with a speed close to that of light? This is the quark-parton puzzle.\n27Figure 10: Scattering and bound states. These days, Feynman dia grams are used for scattering\nproblems. For bound-state problems, it is possible to construct Lo rentz-covariant harmonic oscil-\nlators by integrating the papers written by Dirac. Feynman diagram s and the covariant oscillators\nare both two different representations of the inhomogeneous Lor entz group. Then is it possible\nto derive Einstein’s special relativity from the Heisenberg brackets ? This problem is addressed in\nSections 6 and 7.\nLorentz boosted. He wrote three papers [19, 20, 21]. If we int egrate his ideas, it is possible\nto construct the covariant oscillator wave function discus sed in this paper. Figure 10 tells\nwhere this integration stands in the history of physics.\nIn constructing quantum mechanics in the Lorentz-covarian t world, the present form of\nquantum field theory is quite successful in scattering probl ems where all participating parti-\ncles are free in the remote past and the remote future. The bou nd states are different, and\nFeynman suggested Lorentz-covariant harmonic oscillator s for studying this problem [25].\nYes, quantum field theory and the covariant harmonic oscilla tors use quite different math-\nematical forms. Yet, they share the same set of physical prin ciples [56]. They are both\nthe representations of the inhomogeneous Lorentz group. We note that Dirac in his 1949\npaper [21] said that we can build relativistic dynamics by co nstructing representations of\nthe inhomogeneous Lorentz group. In Figure 10, both Feynman diagrams and the covariant\noscillator (integration of Dirac’s papers) share the Lie al gebra of the inhomogeneous Lorentz\ngroup. This figure leads to the idea of whether the Lie algebra of quantum mechanics can\nlead to that of the inhomogeneous Lorentz group. We shall dis cuss this question in Section 7.\nSince the time of Bohr and Einstein, attempts have been made t o construct Lorentz-\ncovariant bound states within the framework of quantum mech anics and special relativity.\nIn recent years, there have been laudable efforts to construct a non-perturbative approach\n28to quantum field theory, where particles are in a bound state a nd thus are not free in the\nremote past and remote future [48]. If and when this approach produces localized probability\ndistributions which can Lorentz-boosted, it should explai n both the quark model (at rest)\nand its parton picture in the limit of large speed.\nIn recent years, there have been efforts to represent the obser vation of movements inside\nmaterials using Dirac electric states [49] as well as to use r elativistic methods to under-\nstand atomic and molecular structure [50]. There have also b een efforts to provide covariant\nformulation of the electrodynamics of nonlinear media [51] .\n5.4 Lorentz-Invariant Uncertainty Products\nIn theharmonicoscillator regime, theenergy-momentum wav e functionstake the same math-\nematical form, and the uncertainty relation in terms of the u ncertainty products is well un-\nderstood. However, in the present case, the oscillator wave functions are deformed when\nLorentz-boosted, as shown in Figure 8. According to this figu re, both the space-time and\nmomentum-energy wave functions become spread along their l ongitudinal directions. Does\nthis mean that the Lorentz boost increases the uncertainty?\nIn order to address this question, let us write the momentum- energy wave function as a\nFourier transformation of the space-time wave function:\nφ(qz,q0) =1\n2π/integraldisplay\nψ(z,t)exp(i[qzz−q0t])dt dz. (57)\nThe transverse xandycomponents are not included in this expression. The exponen t of this\nexpression can be written as\nqzz−q0t=q+z−+q−z+, (58)\nwith\nq±=1√\n2(qz±q0), z ±=1√\n2(z±t), (59)\nas given earlier in Equations (4) and (55).\nIn terms of these variables, the Fourier integral takes the f orm\n1\n2π/integraldisplay\nψ(z,t)exp(i[q+z−+q−z+])dt dz. (60)\nIn this case, the variable q+is conjugate to z−, andq−toz+. Let us go back to Figure 8.\nThe major (minor) axis of the space-time ellipse is conjugat e to the minor (major) axis of\nthe momentum-energy ellipse. Thus the uncertainty product s\n/angbracketleftig\nz2\n+/angbracketrightig/angbracketleftig\nq2\n−/angbracketrightig\nand/angbracketleftig\nz2\n−/angbracketrightig/angbracketleftig\nq2\n+/angbracketrightig\n(61)\nremain invariant under the Lorentz boost.\n296 O(3,2)Symmetry DerivablefromTwo-PhotonStates\nIn this section we start with the paper Dirac published in 196 3 on the symmetries from two\nharmonic oscillators [22]. Since the step-up and step-down operators in the oscillator system\nare equivalent to the creation and annihilation operators i n the system of photons, Dirac was\nworking with the system of two photons which is of current int erest [5, 52, 53, 54].\nIn the oscillator system the step-up and step-down operator s are:\na1=1√\n2(x1+iP1), a†\n1=1√\n2(x1−iP1),\na2=1√\n2(x2+iP2), a†\n2=1√\n2(x2−iP2), (62)\nwith\niPi=∂\n∂xi. (63)\nIn terms of these operators, Heisenberg’s uncertainty rela tions can be written as\n/bracketleftig\nai,a†\nj/bracketrightig\n=δij, (64)\nwith\nxi=1√\n2/parenleftig\nai+a†\ni/parenrightig\n, P i=i√\n2/parenleftig\na†\ni−ai/parenrightig\n. (65)\nWith these sets of operators, Dirac constructed three gener ators of the form\nJ1=1\n2/parenleftig\na†\n1a2+a†\n2a1/parenrightig\n, J2=1\n2i/parenleftig\na†\n1a2−a†\n2a1/parenrightig\n, J3=1\n2/parenleftig\na†\n1a1−a†\n2a2/parenrightig\n,(66)\nand three more of the form\nK1=−1\n4/parenleftig\na†\n1a†\n1+a1a1−a†\n2a†\n2−a2a2/parenrightig\n,\nK2= +i\n4/parenleftig\na†\n1a†\n1−a1a1+a†\n2a†\n2−a2a2/parenrightig\n, (67)\nK3=1\n2/parenleftig\na†\n1a†\n2+a1a2/parenrightig\n.\nTheseJiandKioperators satisfy the commutation relations\n[Ji,Jj] =iǫijkJk,[Ji,Kj] =iǫijkKk,[Ki,Kj] =−iǫijkJk. (68)\nThis set of commutation relations is identical to the Lie alg ebra of the Lorentz group\nwhereJiandKiare three rotation and three boost generators respectively . This set of\n30commutators is the Lie algebra of the Lorentz group with thre e rotation and three boost\ngenerators.\nIn addition, with the harmonic oscillators, Dirac construc ted another set consisting of\nQ1=−i\n4/parenleftig\na†\n1a†\n1−a1a1−a†\n2a†\n2+a2a2/parenrightig\n,\nQ2=−1\n4/parenleftig\na†\n1a†\n1+a1a1+a†\n2a†\n2+a2a2/parenrightig\n, (69)\nQ3=i\n2/parenleftig\na†\n1a†\n2−a1a2/parenrightig\n.\nThey then satisfy the commutation relations\n[Ji,Qj] =iǫijkQk,[Qi,Qj] =−iǫijkJk. (70)\nTogether with the relation [ Ji,Jj] =iǫijkJkgiven in Equation (68), JiandQiproduce\nanother Lie algebra of the Lorentz group. Like Ki, theQioperators act as boost generators.\nIn order to construct a closed set of commutation relations f or all the generators, Dirac\nintroduced an additional operator\nS0=1\n2/parenleftig\na†\n1a1+a2a†\n2/parenrightig\n. (71)\nThen the commutation relations are\n[Ki,Qj] =−iδijS0,[Ji,S0] = 0,[Ki,S0] =−iQi,[Qi,S0] =iKi.(72)\nIt was then noted by Dirac that the three sets of commutation r elations given in Equa-\ntions (68), (70) and (72) form the Lie algebra for the O(3,2) de Sitter group. This group\napplies to space of ( x, y, z, t, s ), which is five-dimensional. In this space, the three space- like\ncoordinates are x, y, z, and the time-like variables are given by tands. Therefore, these\ngenerators are five-by-five matrices. The three rotation gen erators for the ( x, y, z) space-like\ncoordinates are given in Table 2. Thethree boost generators with respect to the time variable\ntare given in Table 3. Table 4 contains three boost operators w ith respect to the second\ntime variable sand the rotation generator between the two time variables tands.\nIt is indeed remarkable, as Dirac stated in his paper [22], th at the space-time symmetry\nof the (3 + 2) de Sitter group is the result of this two-oscilla tor system. What is even more\nremarkable is that we can derive, from quantum optics, this t wo-oscillator system. In the\ntwo-photon system in optics, where ican be 1 or 2, aianda†\niact as the annihilation and\ncreation operators.\n31Table 2: Three generators of the rotations in the five-dimensional space of (x, y, z, t, s ). The\ntime-likesandtcoordinates are not affected by the rotations in the three-dimens ional space of\n(x, y, z).\nGenerators Differential Matrix\nJ1 −i/parenleftig\ny∂\n∂z−z∂\n∂y/parenrightig\n0 0 0 0 0\n0 0−i0 0\n0i0 0 0\n0 0 0 0 0\n0 0 0 0 0\n\nJ2 −i/parenleftig\nz∂\n∂x−x∂\n∂z/parenrightig\n0 0i0 0\n0 0 0 0 0\n−i0 0 0 0\n0 0 0 0 0\n0 0 0 0 0\n\nJ3 −i/parenleftig\nx∂\n∂y−y∂\n∂x/parenrightig\n0−i0 0 0\ni0 0 0 0\n0 0 0 0 0\n0 0 0 0 0\n0 0 0 0 0\n\n32Table 3: Three generators of Lorentz boosts with respect the tim e variablet. Thescoordinate is\nnot affected by these boosts.\nGenerators Differential Matrix\nK1 −i/parenleftig\nx∂\n∂t+t∂\n∂x/parenrightig\n0 0 0i0\n0 0 0 0 0\n0 0 0 0 0\ni0 0 0 0\n0 0 0 0 0\n\nK2 −i/parenleftig\ny∂\n∂t+t∂\n∂y/parenrightig\n0 0 0 0 0\n0 0 0i0\n0 0 0 0 0\n0i0 0 0\n0 0 0 0 0\n\nK3 −i/parenleftig\nz∂\n∂t+t∂\n∂z/parenrightig\n0 0 0 0 0\n0 0 0 0 0\n0 0 0i0\n0 0i0 0\n0 0 0 0 0\n\n33Table 4: The O(3,2) group has four additional generators. Note that the generat ors in this table\nhave non-zero elements only in the fifth row and the fifth column. Th is is unlike those given in\nTables 2 and 3. Here the svariable is contained in every differential operator.\nGenerators Differential Matrix\nQ1 −i/parenleftig\nx∂\n∂s+s∂\n∂x/parenrightig\n0 0 0 0 i\n0 0 0 0 0\n0 0 0 0 0\n0 0 0 0 0\ni0 0 0 0\n\nQ2 −i/parenleftig\ny∂\n∂s+s∂\n∂y/parenrightig\n0 0 0 0 0\n0 0 0 0 i\n0 0 0 0 0\n0 0 0 0 0\n0i0 0 0\n\nQ3 −i/parenleftig\nz∂\n∂s+s∂\n∂z/parenrightig\n0 0 0 0 0\n0 0 0 0 0\n0 0 0 0 i\n0 0 0 0 0\n0 0i0 0\n\nS0 −i/parenleftig\nt∂\n∂s−s∂\n∂t/parenrightig\n0 0 0 0 0\n0 0 0 0 0\n0 0 0 0 0\n0 0 0 0 −i\n0 0 0i0\n\n34It is possible to construct, with these two sets of operators , two-photon states [5]. Yuen,\nas early as 1976 [14], used the two-photon state generated by\nQ3=i\n2/parenleftig\na†\n1a†\n2−a1a2/parenrightig\n. (73)\nThis leads to the two-mode coherent state. This is also known as thesqueezed state .\nIt was later that Yurke, McCall, and Klauder, in 1986 [55], in vestigated two-mode inter-\nferometers. In their study of two-mode states, they started withQ3given in Equation (73).\nThen they considered that the following two additional oper ators,\nK3=1\n2/parenleftig\na†\n1a†\n2+a1a2/parenrightig\n, S 0=1\n2/parenleftig\na†\n1a1+a2a†\n2/parenrightig\n, (74)\nwere needed in one of their interferometers. These three Her mitian operators from Equa-\ntions (73) and (74) have the following commutation relation s\n[K3,Q3] =−iS0,[Q3,S0] =iK3,[S0,K3] =iQ3. (75)\nYurke et al. called this device the SU(1,1) interferometer. Thegroup SU(1,1) is isomorphic\nto theO(2,1) group or the Lorentz group applicable to two space-like an d one time-like\ndimensions.\nIn addition, in the same paper [55], Yurke et al. discussed th e possibility of constructing\nanother interferometer exhibiting the symmetry generated by\nJ1=1\n2/parenleftig\na†\n1a2+a†\n2a1/parenrightig\n, J2=1\n2i/parenleftig\na†\n1a2−a†\n2a1/parenrightig\n, J3=1\n2/parenleftig\na†\n1a1−a†\n2a2/parenrightig\n.(76)\nThese generators satisfy the closed set of commutation rela tions\n[Ji,Jj] =iǫijkJk, (77)\ngiven in Equation (66). This is the Lie algebra for the three- dimensional rotation group.\nYurke et al. called this optical device the SU(2) interferometer.\nWe are then led to ask whether it is possible to construct a clo sed set of commutation\nrelations with the six Hermitian operators from Equations ( 75) and (76). It is not possible.\nWe have to add four additional operators, namely\nK1=−1\n4/parenleftig\na†\n1a†\n1+a1a1−a†\n2a†\n2−a2a2/parenrightig\n,\nK2= +i\n4/parenleftig\na†\n1a†\n1−a1a1+a†\n2a†\n2−a2a2/parenrightig\n, (78)\nQ1=−i\n4/parenleftig\na†\n1a†\n1−a1a1−a†\n2a†\n2+a2a2/parenrightig\n,\nQ2=−1\n4/parenleftig\na†\n1a†\n1+a1a1+a†\n2a†\n2+a2a2/parenrightig\n.\n35There are now ten operators. They are precisely those ten Dir ac constructed in his paper\nof 1963 [22].\nIt is indeed remarkable that Dirac’s O(3,2) algebra is produced by modern optics. This\nalgebra produces the Lorentz group applicable to three spac e-like and two time-like dimen-\nsions.\nThe algebra of harmonic oscillators given in Equation (64) i s Heisenberg’s uncertainty\nrelations in a two-dimensional space. The de Sitter group O(3,2) is basically a language of\nspecial relativity. Does this mean that Einstein’s special relativity can be derived from the\nHeisenberg brackets? We shall examine this problem in Secti on 7.\n7 Contraction of O(3, 2) to the Inhomogeneous\nLorentz Group\nAccording to Section 6, the group O(3,2) has anO(3,1) subgroup with six generators plus\nfour additional generators. The inhomogeneous Lorentz gro up also contains one Lorentz\ngroupO(3,1) as its subgroup plus four space-time translation generat ors. The question\narises whether the four generators in Table 4 can be converte d into the four translation\ngenerators. The purpose of this section is to prove this is po ssible according to the group\ncontraction procedure introduced first by In¨ on¨ u and Wigne r [57].\nIn their paper In¨ on¨ u and Wigner introduced the procedure f or transforming the Lorentz\ngroup into the Galilei group. This procedure is known as group contraction [57]. In this\nsection, we use the same procedure to contract the O(3,2) group into the inhomogeneous\nLorentz group which is the group O(3,1) plus four translations.\nLet us introduce the contraction matrix\nC=\n1/ǫ0 0 0 0\n0 1/ǫ0 0 0\n0 0 1/ǫ0 0\n0 0 0 1 /ǫ0\n0 0 0 0 ǫ\n. (79)\nThis matrix expands the z, y, z, t axes, and contracts saxis asǫbecomes small. Yet, the\nJiandKimatrices given in Tables 2 and 3 remain invariant:\nCJiC−1=Ji,andCKiC−1=Ki. (80)\nThese matrices have zero elements in the fifth row and fifth col umn.\nOn the other hand, the matrices in Table 4 are different. Let us c hooseQ3. The same\n36Table 5: Here the generators of translations are given in the four- dimensional Minkowski space.\nIt is of interest to convert the four generators in the O(3,2) group in Table 4 into the four\ntranslation generators.\nGenerators Differential Matrix\nQ1→P1 −i∂\n∂x\n0 0 0 0 i\n0 0 0 0 0\n0 0 0 0 0\n0 0 0 0 0\n0 0 0 0 0\n\nQ2→P2 −i∂\n∂y\n0 0 0 0 0\n0 0 0 0 i\n0 0 0 0 0\n0 0 0 0 0\n0 0 0 0 0\n\nQ3→P3 −i∂\n∂z\n0 0 0 0 0\n0 0 0 0 0\n0 0 0 0 i\n0 0 0 0 0\n0 0 0 0 0\n\nS0→P0 i∂\n∂t\n0 0 0 0 0\n0 0 0 0 0\n0 0 0 0 0\n0 0 0 0 −i\n0 0 0 0 0\n\n37algebra leads to the elements ǫ2and its inverse, as shown in this equation:\nCQ3C−1=\n1/ǫ0 0 0 0\n0 1/ǫ0 0 0\n0 0 1/ǫ0 0\n0 0 0 1 /ǫ0\n0 0 0 0 ǫ\n\n0 0 0 0 0\n0 0 0 0 0\n0 0 0 0 i\n0 0 0 0 0\n0 0i0 0\n\nǫ0 0 0 0\n0ǫ0 0 0\n0 0ǫ0 0\n0 0 0ǫ0\n0 0 0 0 1 /ǫ\n(81)\n=\n0 0 0 0 0\n0 0 0 0 0\n0 0 0 0 i/ǫ2\n0 0 0 0 0\n0 0 0iǫ20\n→P3=\n0 0 0 0 0\n0 0 0 0 0\n0 0 0 0 i\n0 0 0 0 0\n0 0 0 0 0\n. (82)\nThe second line of this equation tells us that ǫ2becomes zero in the limit of small ǫ. If\nwe make the inverse transformation, result is P3. We can perform the same algebra to arrive\nat the the results given in Table 5. According to this table, QiandS0become contracted\nto the generators of the space-time translations. In other w ords, the de Sitter group O(3,2)\ncan be contracted to the inhomogeneous Lorentz group.\nIfthismatrixisappliedtothefive-vectorof( x, y, z, t, s ), itbecomes( x/ǫ, y/ǫ, z/ǫ, t/ǫ, sǫ ).\nIfǫbecomes very small, the saxis contracts while the four others expand. As ǫbecomes\nvery small, ǫsapproaches zero and can be replaced by ǫ, since both of them are zero. We are\nreplacing zero by another zero, as shown here:\n\nx/ǫ\ny/ǫ\nz/ǫ\nt/ǫ\nǫs\n→\nx/ǫ\ny/ǫ\nz/ǫ\nt/ǫ\nǫ\n,and\nǫ0 0 0 0\n0ǫ0 0 0\n0 0ǫ0 0\n0 0 0ǫ0\n0 0 0 0 1 /ǫ\n\nx/ǫ\ny/ǫ\nz/ǫ\nt/ǫ\nǫ\n=\nx\ny\nz\nt\n1\n.(83)\nThis contraction procedure is indicated in Figure 11.\nIndeed, the five-vector ( x, y, z, t, 1) serves as the space-time five-vector of the inhomo-\ngeneous Lorentz group. The transformation matrix applicab le to this five-vector consists of\nthat of the Lorentz group in its first four row and columns. Let us see how the generators Pi\nandS0generate translations. First of all, in terms of these gener ators, the transformation\nmatrix takes the form\nexp(−i[aPx+bPy+zPz+dPt]) =\n1 0 0 0 a\n0 1 0 0 b\n0 0 1 0 c\n0 0 0 1 d\n0 0 0 0 1\n. (84)\n38Figure 11: Contraction of O(3,2) to the inhomogeneous Lorentz group. The extra time variable\nsbecomes a constant, as shown by a flat line in this figure.\n39If this matrix is applied to the five-vector ( x, y, z, t, 1), the result is the translation:\n\n1 0 0 0 a\n0 1 0 0 b\n0 0 1 0 c\n0 0 0 1 d\n0 0 0 0 1\n\nx\ny\nz\nt\n1\n=\nx+a\ny+b\nz+c\nt+d\n1\n. (85)\nThe five-by-five matrix of Equation (84) indeed performs the t ranslations.\nLet us go to Table 5 again. The four differential forms of the tra nslation generators\ncorrespond to the four-momentum satisfying the equation E2=p2\nx+p2\ny+p2\nz+m2. This is\nof course Einstein’s E=mc2.\n8 Concluding Remarks\nSince 1973, the present authors have been publishing papers on the harmonic oscillator wave\nfunctions which can be Lorentz-boosted. The covariant harm onic oscillator plays roles in\nunderstandingsomeofthepropertiesinhigh-energypartic lephysics. Thiscovariant oscillator\ncan also be used as a representation of Wigner’s little group for massive particles.\nMore recently, the oscillator wave function was shown to pro vide basic mathematical tools\nfor two-photon coherent states known as the squeezed state o f light [5].\nIn this paper, we have provided a review of our past efforts, wit h the purpose of integrat-\ning the papers Paul A. M. Dirac wrote in his lifelong efforts to m ake quantum mechanics\ncompatible with Einstein’s special relativity which produ ced the Lorentz-covariant energy-\nmomentum relation. It is interesting to note that Einstein’ s special relativity is derivable\nfrom Heisenberg’s expression of the uncertainty relation.\nReferences\n[1] Kim, Y.S.; Noz, M.E. Covariant harmonic oscillators and the quark model. Phys. Rev.\nD1973,8, 3521–3627.\n[2] Kim, Y.S.; Noz, M.E. Theory and Applications of the Poincar´ e Group ; Reidel: Dor-\ndrecht, The Netherlands, 1986.\n[3] Ba¸ skal, S.; Kim. Y.S.; Noz, M.E. Physics of the Lorentz Group, IOP Concise Physics ;\nMorgan & Claypool Publisher: San Rafael, CA, USA; IOP Publis hing: Bristol, UK,\n2015.\n[4] Kim, Y.S.; Noz, M.E. New Perspectives on Einstein’s E=mc2; World Scientific:\nSingapore, 2018.\n40[5] Ba¸ skal, S.; Kim, Y.S.; Noz, M.E. Mathematical Devices for Optical Sciences ; IOP\nPublishing: Bristol, UK, 2019.\n[6] Ruiz, M.J. Orthogonality relations for covariant harmo nic oscillator wave functions.\nPhys. Rev. D 1974,10, 4306–4307.\n[7] Kim, Y.S.; Noz, M.E.; Oh, S.H. A simple method for illustr ating the difference between\nthe homogeneous and inhomogeneous Lorentz groups. Am. J. Phys. 1979,47, 892–897.\n[8] Gell-Mann, M. A Schematic Model of Baryons and Mesons. Phys. Lett. 19648, 214–\n215.\n[9] Feynman, R. P. Very High-Energy Collisions of Hadrons. Phys. Rev. Lett. 1969,23,\n1415–1417.\n[10] Kim, Y.S.; Noz, M.E. Covariant harmonic oscillators an d the parton picture. Phys.\nRev. D1977,15, 335–338.\n[11] Kim, Y.S. Observable gauge transformations in the part on picture. Phys. Rev. Lett.\n1989,63, 348–351.\n[12] Kim, Y.S.; Noz, M.E.; Oh, S.H. Representations of the Po incar´ e group for relativistic\nextended hadrons. J. Math. Phys. 1979,20, 1341–1344.\n[13] Wigner, E. On unitary representations of the inhomogen eous Lorentz group. Ann.\nMath.1939,40, 149–204.\n[14] Yuen, H.P. Two-photon coherent states of the radiation field.Phys. Rev. A 1976,13,\n2226–2243.\n[15] Kim, Y.S.; Noz, M.E. Coupled oscillators, entangled os cillators, and Lorentz-covariant\nharmonic oscillators. J. Opt. B Quantum Semiclassical Opt. 2005,7, S458–S467.\n[16] Ba¸ skal, S.; Kim, Y.S.; Noz, M.E. Harmonic Oscillators and Space Time Entanglement.\nSymmetry 2016,8, 55–80.\n[17] Han, D.; Kim, Y.S.; Son, D. GaugetransformationsasLor entz-boosted rotations. Phys.\nLett. B1983,131, 327–329.\n[18] Kim, Y.S.; Wigner, E.P. Space-time geometry of relativ istic-particles. J. Math. Phys.\n1990,31, 55–60.\n[19] Dirac, P.A.M. The Quantum Theory of the Emission and Abs orption of Radiation.\nProc. R. Soc. (Lond.) 1927,A114, 243–265.\n[20] Dirac, P.A.M. Unitary Representations of the Lorentz G roup.Proc. R. Soc. (Lond.)\n1945,A183, 284–295.\n[21] Dirac, P.A.M. Forms of Relativistic Dynamics. Rev. Mod. Phys. 1949,21, 392–399.\n[22] Dirac, P.A.M. A Remarkable Representation of the 3 + 2 de Sitter Group. J. Math.\nPhys.1963,4, 901–909.\n41[23] Dirac, P.A.M. Can equations of motion be used in high ene rgy physics? Phys. Today\n1970,23, 29–31.\n[24] Farmelo, G. The Strangest Man, the Hidden Life of Paul Dirac, Mystic of Atom ; Basic\nBooks: New York, NY, USA, 2009.\n[25] Feynman, R.P.; Kislinger, M.; Ravndal, F. CurrentMatr ix Elements fromaRelativistic\nQuark Model. Phys. Rev. D 1971,3, 2706–2732.\n[26] Frazer, W.; Fulco, J. Effect of a Pion-Pion Scattering Res onance on Nucleon Structure.\nPhys. Rev. Lett. 1960,2, 365–368.\n[27] Bjorken, J.D.; Paschos, E.A. Electron-Proton and γ-Proton Scattering and the Struc-\nture of the Nucleon. Phys. Rev. 1969,185, 1975–1982.\n[28] Ba¸ skal, S.; Kim, Y.S.; Noz, M.E. Poincar´ e Symmetry fr om Heisenberg’s Uncertainty\nRelations. Symmetry 2019,11, 409.\n[29] Ba¸ skal, S.; Kim, Y.S.; Noz, M.E. Einstein’s E=mc2derivable from Heisenberg’s\nUncertainty Relations. Quantum Rep. 2019,1, 236–251.\n[30] Wigner, E.P. On the Time-Energy Uncertainty Relation. InAspects of Quantum The-\nory; Salam, A., Wigner, E.P., Eds.; Cambridge University Press : London, UK, 1972;\npp. 237–247.\n[31] Fujimura, K.; Kobayashi, T.; Namiki, M. Nucleon Electr omagnetic Form Factors at\nHigh Momentum Transfers in an Extended Particle Model Based on the Quark Model.\nProg. Theor. Phys. 1970,43, 73–79.\n[32] Beg, M.A.B.; Lee, B.W.; Pais, A. SU(6) and Electromagne tic Interactions. Phys. Rev.\nLett.1964,13, 514–517.\n[33] Greenberg, O.W.; Resnikoff, M. Symmetric Quark Model of B aryon Resonances. Phys.\nRev.1967,163, 1844–1851\n[34] Kim, Y.S.; Noz, M.E. The Question of Simultaneity in Rel ativity and Quantum Me-\nchanics. In Quantum Theory: Reconsideration of Foundations–3 ; Adenier, G., Khren-\nnikov, A., Nieuwenhuizen, T.M., Eds.; American Institute o f Physics: College Park,\nMD, USA, 2006; Volume 810, pp. 168–178.\n[35] Rotbart, F.C. Complete orthogonality relations for th e covariant harmonic oscillator.\nPhys. Rev. D 1981,23, 3078–3080.\n[36] Yukawa, H. Structure and Mass Spectrum of Elementary Pa rticles. I. General Consid-\nerations. Phys. Rev. 1953,91, 415–416.\n[37] Markov, M. On Dynamically Deformable Form Factors in th e Theory of Particles.\nSuppl. Nuovo Cim. 1956,3, 760–772.\n[38] Ginzburg, V.L.; Man’ko, V.I. Relativistic oscillator models of elementary particles.\nNucl. Phys. 1965,74, 577–588.\n42[39] Hofstadter, R.; McAllister, R.W. Electron Scattering from the Proton. Phys. Rev.\n1955,98, 217–218.\n[40] Frazer, W. Elementary Particle Physics ; Prentice Hall: Englewood Cliffs, NJ, USA,\n1966.\n[41] Itzykson, C.; Zuber, J.B. Quantum Field Theory ; McGraw-Hill: New York, NY, USA,\n1980.\n[42] Licht, A.L.; Pagnamenta, A. Wave Functions and Form Fac tors for Relativistic Com-\nposite Particles I. Phys. Rev. D 1970,2, 1150–1156.\n[43] Lipes, R. Electromagnetic Excitations of the Nucleon i n a Relativistic Quark Model.\nPhys. Rev. D 1972,5, 2849–2863.\n[44] Henriques, A.B.; Keller, B.H.; Moorhouse, R.G. Genera l three-spinor wave functions\nand the relativistic quark model. Ann. Phys. (NY) 1975,93, 125–151.\n[45] Matevosyan, H.H.; Thomas, A.W.; Miller, G.A. Study of l attice QCD form factors\nusing the extended Gari-Krumpelmann model. Phys. Rev. C 2005,72, 065204.\n[46] Alkofer, R.; Holl, A.; Kloker, M.; Karssnigg, A.; Rober ts, C.D. On Nucleon Electro-\nmagnetic Form Factors. Few-Body Syst. 2005,37, 1–31.\n[47] Hussar, P.E. Valons and harmonic oscillators. Phys. Rev. D 1981,23, 2781–2783.\n[48] Strocchi, F. An Introduction to Non-Perturbative. In Foundations of Quantum Field\nTheory Oxford Science Publications ; Oxford Science Publications: Oxford, UK, 2013.\n[49] Park, J.W.; Kim, H.S.; Brumme, T.; Heine T.; Yeom H.W. Ar tificial relativistic\nmolecules. Nat. Commun. 2020,11, 815.\n[50] Grant, I. Relativistic Atomic Structure. In Springer Handbook of Atomic, Molecular,\nand Optical Physics ; Drake, G., Ed.; Springer: New York, NY, USA, 2006.\n[51] Hartemann, F.; Toffano, Z. Relativistic electrodynamic s of continuous media. Phys.\nRev. A1990,41, 5066–5073.\n[52] Dodonov, V.V.; Man’ko, V.I. Theory of Nonclassical States of Light ; Taylor: London,\nUK; Francis: New York, NY, USA, 2003.\n[53] Saleh, B.E.A.; Teich, M.C. Fundamentals of Photonics , 2nd ed.; Wiley series in pure\nand applied optics; Wiley Interscience: Hoboken, NJ, USA, 2 007.\n[54] Walls, D.F.; Milburn, G.J. Quantum Optics , 2nd ed.; Springer: Berlin, Germany, 2008.\n[55] Yurke, B.S.; McCall, S.L.; Klauder, J.R. SU(2) andSU(1 ,1) interferometers. Phys. Rev.\nA1986,33, 4033–4054.\n[56] Han, D.; Kim, Y.S.; Noz, M.E. Physical principles in qua ntum field theory and in\ncovariant harmonic oscillator formalism. Found. Phys. 1981,11, 895–905.\n43[57] In¨ on¨ u, E.; Wigner, E.P. Onthe Contraction of Groups a nd their Representations. Proc.\nNatl. Acad. Sci. USA 1953,39, 510–524.\n44" }, { "title": "1605.08358v1.On__M_____terms_approximations_Besov_classes_in_Lorentz_spaces.pdf", "content": "arXiv:1605.08358v1 [math.CA] 26 May 2016ONM– TERMS APPROXIMATIONS BESOV CLASSES IN LORENTZ\nSPACES\nG. AKISHEV\nAbstract. In this paper we consider Lorentz space with a mixed norm\nof periodic functions of many variables. We obtain the exact estimat ion\nof the best M-term approximations of Nikol’ski’s, Besov’s classes in th e\nLorentz space with the mixed norm.\nKeywords: Lorentz space, and Besov’s class, and approximation\nMSC:41A10 and 41A25\n1.Introduction\nLetx= (x1,...,xm)∈Tm= [0,2π)mandθj,pj∈[1,+∞),j= 1,...,m. LetL¯p,¯θ(Tm)\ndenotes the space of Lebesgue measureable functions f(¯x) defined on Rm, which have 2 π\n– period with respect to each variable such that\n/bardblf/bardblp,θ=/bardbl.../bardblf/bardblp1,θ1.../bardblpm,θm<+∞,\nwhere\n/bardblg/bardblp,θ=\n\n2π/integraldisplay\n0(g∗(t))θtθ\np−1dt\n\n1\nθ\n,\nwhereg∗a non-increasing rearrangement of the function |g|(see. [1]).\nIt is known that if θj=pj,j= 1,...,m, thenL¯p,¯θ(Tm) =L¯p(Tm) the Lebesgue measur-\nable space of functions f(¯x) defined on Rm, which have 2 π– period with respect to each\nvariable with the norm\n/bardblf/bardbl¯p=/bracketleftBigg/integraldisplay2π\n0/bracketleftbigg\n···/bracketleftbigg/integraldisplay2π\n0|f(¯x)|p1dx1/bracketrightbiggp2\np1···/bracketrightbiggpm\npm−1dxm/bracketrightBigg1\npm\n<+∞,\nwherep= (p1,...,pm),1/lessorequalslantpj<+∞, j= 1,...,m(see [2],p. 128).\nAny function f∈L1(Tm) =L(Tm) can be expanded to the Fourier series\n/summationdisplay\nn∈Zman(f)ei/angbracketleftn,x/angbracketright,\nwherean(f) Fourier coefficients of f∈L1(Tm) with respect to multiple trigonometric\nsystem{ei/angbracketleftn,x/angbracketright}¯n∈Zm,andZmis the space of points in Rmwith integer coordinates.\nFor a function f∈L(Tm) and a number s∈Z+=N∪{0}let us introduce the notation\nδ0(f,¯x) =a0(f), δs(f,x) =/summationdisplay\nn∈ρ(s)an(f)ei/angbracketleftn,x/angbracketright,\nwhere/an}bracketle{t¯y,¯x/an}bracketri}ht=m/summationtext\nj=1yjxj,\nρ(s) =/braceleftbigg\nk= (k1,...,km)∈Zm: [2s−1]/lessorequalslantmax\nj=1,...,m|kj|<2s/bracerightbigg\n,\n12 G. AKISHEV\nwhere [a] is the integer part of the number a.\nLet us consider Nikol’skii, Besov classes( see [2], [3]). Let 1 < pj<+∞,1< θj<+∞,\nj= 1,...,m, 1/lessorequalslantτ/lessorequalslant∞, andr >0\nHr\n¯p,¯θ=/braceleftbigg\nf∈L¯p,¯θ(Tm) : sup\ns∈Z+2sr/bardblδs(f)/bardbl¯p,¯θ/lessorequalslant1/bracerightbigg\n,\nBr\n¯p,¯θ,τ=\n\nf∈L¯p,¯θ(Tm) :\n/summationdisplay\ns∈Z+2srτ/bardblδs(f)/bardblτ\n¯p,¯θ\n1\nτ\n/lessorequalslant1\n\n.\nIt is known that for 1 /lessorequalslantτ/lessorequalslant∞the following holds\nBr\n¯p,¯θ,1⊂Br\n¯p,¯θ,τ⊂Br\n¯p,¯θ,∞=Hr\n¯p,¯θ.\nLetf∈L¯p,¯θ(Tm) and/braceleftbig¯k(j)/bracerightbigM\nj=1be a system of vectors ¯k(j)= (k(j)\n1,...,k(j)\nm) with integer\ncoordinates. Consider the quantity\neM(f)¯p,¯θ= inf\n¯k(j),bj/vextenddouble/vextenddouble/vextenddouble/vextenddouble/vextenddoublef−M/summationdisplay\nj=1bje/angbracketlefti¯k(j),¯x/angbracketright/vextenddouble/vextenddouble/vextenddouble/vextenddouble/vextenddouble\n¯p,¯θ,\nwherebjare arbitrary numbers. The quantity eM(f)¯p,¯θis called the best M– term\napproximation of a function f∈L¯p,¯θ(Tm). For a given class F⊂L¯p,¯θ(Tm) let\neM(F)¯p,¯θ= sup\nf∈FeM(f)¯p,¯θ.\nThe best M– term approximation was defined by S.B.Stechkin [4]. Estimations of\nM– term approximations of different classes were provided by R.S. Ism agilov [5], E.S.\nBelinsky [6], V.E. Maiorov [7], B.S. Kashin [8], R. DeVore [9], V.N. Temlyakov [1 0], A.S.\nRomanyuk [11], Dinh Dung [12], Wang Heping and Sun Yongsheng [13], L. Q .Duan and\nG.S. Fang [14], W.Sickel and M. Hansen [15], S.A. Stasyuk [16], [17] and ot hers (see\nbibliography in [18], [19], [20]).\nFor the case p1=...=pm=pandq1=...=qm=θ1=...=θ1=qR.A. De Vore and\nV.N. Temlyakov [20] proved the following theorem.\nTheorem A. Let1/lessorequalslantp,q,τ/lessorequalslant∞andr(p,q) =m/parenleftBig\n1\np−1\nq/parenrightBig\n+if1/lessorequalslantp/lessorequalslantq/lessorequalslant2, or\n1/lessorequalslantq/lessorequalslantp <∞andr(p,q) = max/braceleftBig\nm\np,m\n2/bracerightBig\nin other cases. Then for r > r(p,q)the\nfollowing holds\neM(Br\np,τ)q≍M−r\nm+(1\np−max{1\nq,1\n2})+,\nwherea+= max{a;0}.\nMoreover, inthecaseof m(1\np−1\nq)< r > B.ONM– TERMS APPROXIMATIONS BESOV CLASSES IN LORENTZ SPACES 3\n2.Auxiliary results\nTo prove the main results the following auxiliary propositions are used .\nTheorem B ([21] ). Letp∈(1,∞).Then there exist positive numbers C1(p),C2(p)\nsuch that for any function f∈Lp(Tm)the following inequality holds:\n/bardblf/bardblp<1.\n3. Ifr >m/summationtext\nj=11\npj, then\neM/parenleftBig\nBr\n¯p,¯θ(1),τ/parenrightBig\n¯q,¯θ(2)≍M−1\nm(r+m/summationtext\nj=1(1\n2−1\npj))\n.4 G. AKISHEV\nProof.Firstly, we are going to consider the upper bound in the first item. Ta king into\naccount the inclusion Br\n¯p,¯θ(1),τ⊂Hr\n¯p,¯θ(1),1/lessorequalslantτ <+∞, it suffices to prove it for the class\nHr\n¯p,¯θ(1).\nLet 1< pj/lessorequalslant2< qj<∞, j= 1,...,m,andNbe the set of natural numbers. For a\nnumberM∈Nchoose a natural number nsuch that 2nm< M/lessorequalslant2(n+1)m.For a function\nf∈Hr\n¯p,¯θ(1), it is known that\n/bardblδs(f)/bardbl¯p,¯θ(1)/lessorequalslant2−sr,1< pj<∞, j= 1,...,m.\nWe will seek an approximation polynomial P(ΩM,¯x) in the form\nP(ΩM,¯x) =n−1/summationdisplay\ns=0δs(f,¯x)+/summationdisplay\nn/lessorequalslants<αnP(ΩNs,¯x), (1)\nwhere the polynomials P(ΩNs,¯x) will be constructed for each δs(f,¯x) in accordance with\nLemma 1, and the number α >1 will be chosen during the construction.\nLetm/summationtext\nj=1(1\npj−1\nqj)< r 0.\nConsider the function\nfn(¯x) =C32−nm(2m/summationtext\nj=11\nqj)−1(r−m/summationtext\nj=1(1\npj−1))\nF¯q,n(¯x).\nBy the inequality (8), we get\n∞/summationdisplay\ns=02sr/bardblδs(fn)/bardbl¯p,¯θ(1)<<\n<<2−nm(2m/summationtext\nj=11\nqj)−1(r−m/summationtext\nj=1(1\npj−1))[m(2m/summationtext\nj=11\nqj)−1]\n/summationdisplay\ns=02sr/bardblδs(F¯q,n)/bardbl¯p,¯θ(1)<<\n<<2−nm(2m/summationtext\nj=11\nqj)−1(r−m/summationtext\nj=1(1\npj−1))[m(2m/summationtext\nj=11\nqj)−1]\n/summationdisplay\ns=02sr2m/summationtext\nj=1(1−1\npj)\n/lessorequalslantC3.\nHence, the function C−1\n3fn∈Br\n¯p,¯θ(1),1.\nFor the functions (10) and (11), we have by the formula (7), the f ollowing\neM(fn)¯q,¯θ(2)>>inf\nΩM/vextendsingle/vextendsingle/vextendsingle/vextendsingle/integraldisplay\nTmfn(¯x)P1(¯x)d¯x/vextendsingle/vextendsingle/vextendsingle/vextendsingle>>\n>>2−nm(2m/summationtext\nj=11\nqj)−1(r−m/summationtext\nj=1(1\npj−1))\n2−nm\n2(/bardblF¯q,n/bardbl2\n2−M)>>ONM– TERMS APPROXIMATIONS BESOV CLASSES IN LORENTZ SPACES 7\n>>2−nm(2m/summationtext\nj=11\nqj)−1(r−m/summationtext\nj=1(1\npj−1))\n2−nm\n22nm(2m/summationtext\nj=11\nqj)−1\n=\n=C2−nm(2m/summationtext\nj=11\nqj)−1(r−m/summationtext\nj=1(1\npj−1\nqj))\n. (12)\nHence, it follows from (12) by the inclusion Br\n¯p,¯θ(1),1⊂Br\n¯p,¯θ(1),τthat\neM(fn)¯q,¯θ(2)>>2−nm(2m/summationtext\nj=11\nqj)−1(r−m/summationtext\nj=1(1\npj−1\nqj))\nin the case ofm/summationtext\nj=1(1\npj−1\nqj)< r 2 andLβ(Tm)⊂L¯q,¯θ(2)(Tm).\nTherefore by applying Theorem B and by the norm property we obta in\nJ1(n) =/vextenddouble/vextenddouble/vextenddouble/vextenddouble/vextenddouble/summationdisplay\nn/lessorequalslants<αn(δs(f)−P(ΩNs))/vextenddouble/vextenddouble/vextenddouble/vextenddouble/vextenddouble\n¯q,¯θ(2)/lessorequalslantC/vextenddouble/vextenddouble/vextenddouble/vextenddouble/vextenddouble/summationdisplay\nn/lessorequalslants<αn(δs(f)−P(ΩNs))/vextenddouble/vextenddouble/vextenddouble/vextenddouble/vextenddouble\nβ<<\n<m/summationtext\nj=11\npj. Suppose\nNs=/bracketleftBig\n2n(r−m/summationtext\nj=1(1\npj−1))\n2−s(r−m/summationtext\nj=11\npj)/bracketrightBig\n+1.\nThen\nn−1/summationdisplay\ns=0♯ρ(s)+/summationdisplay\nn/lessorequalslants<αnNs<<\n<<2nm+(α−1)n+2n(r−m/summationtext\nj=1(1\npj−1))/summationdisplay\nn/lessorequalslants<αn2−s(r−m/summationtext\nj=11\npj)\n<m/summationtext\nj=11\npjwe obtain\nJ1(n)/lessorequalslant/parenleftBigg/summationdisplay\nn/lessorequalslants<αnN−1\ns2sm22sm/summationtext\nj=1(1\npj−1\n2)\n/bardblδs(f)/bardbl2\n¯p,¯θ(1)/parenrightBigg1\n2\n<<\n<<2−n\n2(r−m/summationtext\nj=1(1\npj−1))/parenleftBigg/summationdisplay\nn/lessorequalslants<αn2s(r+m/summationtext\nj=11\npj)\n/bardblδs(f)/bardbl2\n¯p,¯θ(1)/parenrightBigg1\n2\n<<\n<<2−n\n2(r−m/summationtext\nj=1(1\npj−1))/parenleftBigg/summationdisplay\nn/lessorequalslants<αn2−s(r−m/summationtext\nj=11\npj)/parenrightBigg1\n2\n<<2−n(r+m/summationtext\nj=1(1\n2−1\npj))\n.\nThus,\nJ1(n)<< M−1\nm(r+m/summationtext\nj=1(1\n2−1\npj))\n(15)\nin the case of r >m/summationtext\nj=11\npj.\nTo estimate J2(n), we suppose α= (r+m/summationtext\nj=1(1\n2−1\npj))(r+m/summationtext\nj=1(1\nqj−1\npj))−1and get\nJ2(n)<m/summationtext\nj=11\npj.\nFromBr\n¯p,¯θ(1),τ⊂Hr\n¯p,¯θ(1)it follows that\neM/parenleftBig\nBr\n¯p,¯θ(1),τ/parenrightBig\n¯q,¯θ(2)<< M−1\nm(r+m/summationtext\nj=1(1\n2−1\npj))\nin the case of r >m/summationtext\nj=11\npj. It proves the upper bound estimation in the item 3.\nLet us consider the lower bound estimation inthe case r=m/summationtext\nj=11\npj. Consider the function\ng1(¯x) =n/summationdisplay\ns=1/summationdisplay\n¯k∈ρ(s)m/productdisplay\nj=1k−1\njcoskjxj. (17)\nThen\nδs(g1,¯x) =/summationdisplay\n¯k∈ρ(s)m/productdisplay\nj=1k−1\njcoskjxj.10 G. AKISHEV\nIt is known that for a function ds(¯x) =/summationtext\n¯k∈ρ(s)m/producttext\nj=1coskjxjthe following relation holds\n/bardblds/bardbl¯p,¯θ(1)≍2sm/summationtext\nj=1(1−1\npj)\n,1< pj,θ(1)\nj<+∞, j= 1,...,m.\nThereforebytheinequality ofdistinct metrics (seeTheoremC) and bytheMarcinkiewiczs\ntheorem on multipliers, we have\n/bardblδs(g1)/bardbl¯p,¯θ(1)<<2−sm/bardblds/bardbl¯p,¯θ(1)/lessorequalslantC2−sm/summationtext\nj=11\npj.\nHence, since r=m/summationtext\nj=11\npjwe obtain\n/parenleftBigg∞/summationdisplay\ns=02srτ/bardblδs(g1)/bardblτ\n¯p,¯θ(1)/parenrightBigg1\nτ\n/lessorequalslantC1n1\nτ.\nTherefore the function f1(¯x) =C−1\n1n−1\nτg1(¯x) belongs to the class Br\n¯p,¯θ(1),τ,1< pj<\n+∞,j= 1,...,m.\nNow, we are going to construct a function P1, which satisfies the conditions of the\nformula (7). Let\nv1(¯x) =n/summationdisplay\ns=1/summationdisplay\n¯k∈ρ(s)m/productdisplay\nj=1coskjxj\nand Ω Mbe an arbitrary set of vectors ¯k= (k1,...,km) inMwith integer coordinates.\nConsider the function\nu1(¯x) =n/summationdisplay\ns=1/summationdisplay\n¯k∈ρ(s)∩ΩMm/productdisplay\nj=1coskjxj.\nSuppose w1(¯x) =v1(¯x)−u1(¯x). Then, since 1 < q′\nj=qj\nqj−1<2, j= 1,...,m, we obtain,\nby the Perseval‘s equality, the following\n/bardblw1/bardbl¯q′,¯θ(2)′/lessorequalslant/bardblv1/bardbl¯q′,¯θ(2)′+/bardblu1/bardbl2/lessorequalslant/bardblv1/bardbl¯q′,¯θ(2)′+CM1\n2.\nBy the property of quasi-norm and the estimation of the norm of th e Dirichlet kernel in\nthe Lorentz space, we have\n/bardblv1/bardbl¯q′,¯θ(2)′<>/summationdisplay\nn1/lessorequalslants>\n/greaterorequalslantC(ln2)m/summationdisplay\nn1/lessorequalslants>(ln2)m2−nm\n2n1−1\nτ≍M−1\n2(log2M)1−1\nτ,\nwheren1is a natural number such that n1> M−1\n2(log2M)1−1\nτ\nin the case of r=m/summationtext\nj=11\npj. Hence\neM/parenleftBig\nBr\n¯p,¯θ(1),τ/parenrightBig\n¯q,¯θ(2)>> M−1\n2(log2M)1−1\nτ\nin the case of r=m/summationtext\nj=11\npj. It proves the lower bound estimation in the second item.\nLet us prove the lower bound estimation for the case r >m/summationtext\nj=11\npj. Since in this case\nan upper bound estimation of the quantity eM/parenleftBig\nBr\n¯p,¯θ(1),τ/parenrightBig\n¯q,¯θ(2)does not depend on τand\nBr\n¯p,¯θ(1),1⊂Br\n¯p,¯θ(1),τ, 1< τ <+∞, it suffices to prove the lower bound estimation for\nBr\n¯p,¯θ(1),1.\nFor a number M∈N, we choose a natural number nsuch that 2nm< M/lessorequalslant2(n+1)m\nand 2M/lessorequalslant♯ρ(n), where ♯ρ(n) denotes the number of elements in the set ρ(n).\nConsider the following function\nf3(¯x) =n−1n/summationdisplay\ns=12−sm/summationtext\nj=1(1−1\npj)/summationdisplay\n¯k∈ρ(s)m/productdisplay\nj=1k−r\nm\njcoskjxj.\nThen\n/bardblδs(f3)/bardbl¯p,¯θ(1)<<2−srn−1.\nHence\n∞/summationdisplay\ns=02sr/bardblδs(f3)/bardbl¯p,¯θ(1)/lessorequalslantC3\ni.e. the function C−1\n3f3∈Br\n¯p,¯θ(1),1.\nNext, consider the functions\nv3(¯x) =n/summationdisplay\ns=1/summationdisplay\n¯k∈ρ(s)m/productdisplay\nj=1coskjxj,\nu3(¯x) =n/summationdisplay\ns=1/summationdisplay\n¯k∈ρ(s)∩ΩMm/productdisplay\nj=1coskjxj.12 G. AKISHEV\nSuppose w3(¯x) =v3(¯x)−u3(¯x).By the Perseval‘s equality,\n/bardblu3/bardbl2/lessorequalslantM1\n2,\n/bardblv3/bardbl2= 2(n−1)m\n2.\nFrom these relations, we obtain, by the properties of the norm, th e following\n/bardblw3/bardbl2/lessorequalslant/bardblv3/bardbl2+/bardblu3/bardbl2/lessorequalslantC42nm\n2.\nTherefore the function P3(¯x) =C−1\n42−nm\n2w3(¯x) satisfies the conditions of the formula (7).\nSince 2< qjj= 1,...,m, we have eM(f3)2/lessorequalslantCeM(f3)¯q,¯θ(2). Now, by the formula (7), we\nget\neM(f3)¯q,¯θ(2)>> eM(f3)2>> n−12−nm\n2n/summationdisplay\ns=12−sm/summationtext\nj=1(1−1\npj)/summationdisplay\n¯k∈ρ(s)m/productdisplay\nj=1k−r\nm\nj>>\n>> n−12−nm\n2n/summationdisplay\ns=12−sm/summationtext\nj=1(1−1\npj)\n2s(m−r)=\n=Cn−12−nm\n2n/summationdisplay\ns=12−s(r−m/summationtext\nj=11\npj)\n>>2−n(r+m/summationtext\nj=1(1\n2−1\npj))\n.\nIt follows from the relation 2nm≍Mthat\neM(f3)¯q,¯θ(2)>> M−1\nm(r+m/summationtext\nj=1(1\n2−1\npj))\nin the case r >m/summationtext\nj=11\npjfor the function C−1\n3f3∈Br\n¯p,¯θ(1),1. Hence\neM/parenleftBig\nBr\n¯p,¯θ(1),1/parenrightBig\n¯q,¯θ(2)>> M−1\nm(r+m/summationtext\nj=1(1\n2−1\npj))\n.\nTherefore\neM/parenleftBig\nBr\n¯p,¯θ(1),τ/parenrightBig\n¯q>> M−1\nm(r+m/summationtext\nj=1(1\n2−1\npj))\nin the case r >m/summationtext\nj=11\npj. So Theorem 1 has been proved.\nTheorem 2. . Let¯p= (p1,...,pm),¯q= (q1,...,qm),1< pj< qj/lessorequalslant2,1< θ(1)\nj,θ(2)\nj<∞,\nj= 1,...,m,1/lessorequalslantτ/lessorequalslant+∞.\nIfr >m/summationtext\nj=1(1\npj−1\nqj),then\neM/parenleftBig\nBr\n¯p,¯θ(1),τ/parenrightBig\n¯q,θ(2)≍M−1\nm(r−m/summationtext\nj=1(1\npj−1\nqj))\n.\nProof.For a number M∈Nchoose a natural number nsuch that M≍2nm. By the\ninequality of distinct metrics and by Holder‘s inequality, we have\n/bardblf−n/summationdisplay\ns=0δs(f)/bardbl¯q,¯θ(2)/lessorequalslant∞/summationdisplay\ns=n/bardblδs(f)/bardbl¯q,¯θ(2)/lessorequalslant\n/lessorequalslant/bracketleftBig∞/summationdisplay\ns=02sτr/bardblδs(f)/bardblτ\n¯q,¯θ(2)/bracketrightBig1\nτ/lessorequalslant/bracketleftBig∞/summationdisplay\ns=n2sτ′(r−m/summationtext\nj=1(1\npj−1\nqj))/bracketrightBig1\nτ′<m\n2,then\neM/parenleftBig\nBr\n¯p,¯θ(1),τ/parenrightBig\n¯q,¯θ(2)≍M−r\nm.\nProof.By the inclusion Br\n¯p,¯θ(1),τ⊂Br\n¯2,¯θ(1),τ⊂Hr\n2,¯θ(1), we have\neM/parenleftBig\nBr\n¯p,¯θ(1),τ/parenrightBig\n¯q,¯θ(2)/lessorequalslanteM/parenleftBig\nBr\n2,¯θ(1),τ/parenrightBig\n¯q,¯θ(2)/lessorequalslanteM/parenleftBig\nHr\n2,¯θ(1)/parenrightBig\n¯q,¯θ(2).\nBy Theorem 1,\neM/parenleftBig\nHr\n2,¯θ(1)/parenrightBig\n¯q,¯θ(2)<< M−r\nm.\nforpj= 2,j= 1,...,m.Hence\neM/parenleftBig\nBr\n¯p,¯θ(1),τ/parenrightBig\n¯q,¯θ(2)<< M−r\nm.\nit proves the upper bound estimation.\nLet us consider the lower bound estimation. Consider Rudin-Shapiro s polynomial (see\n[24], p.155) of the type\nRs(x) =2s/summationdisplay\ns=2s−1εkeikx, x∈[0,2π], εk=±1.\nit is known that /bardblRs/bardbl∞= max\nx∈[0,2π]|Rs(x)|<<2s\n2(see [24], p. 155). For a given number M\nchoose a number nsuch that M≍2nm.Now consider the function\nf0(¯x) = 2−n(m\n2+r)n/summationdisplay\ns=1m/productdisplay\nj=1Rs(xj)\nThen,by the continuity, f0∈L¯p,¯θ(1)(Tm) and\n∞/summationdisplay\ns=02sτr/bardblδs(f0)/bardblτ\n¯p,¯θ(1)= 2−n(m\n2+r)n/summationdisplay\ns=12sτr/bardblm/productdisplay\nj=1Rs(xj)/bardblτ\n¯p,¯θ(1)/lessorequalslant14 G. AKISHEV\n/lessorequalslant2−n(m\n2+r)n/summationdisplay\ns=12s(m\n2+r)τ/lessorequalslantC0.\nHence, the function C−1\n0f0∈Br\n¯p,¯θ(1),τ. Now construct a function P(¯x), which would satisfy\nthe conditions in the formula (7). Suppose\nv0(¯x) =n/summationdisplay\ns=1m/productdisplay\nj=1Rs(xj), u0(¯x) =∗/summationdisplay\nsm/productdisplay\nj=1Rs(xj),\nwhere the sign ∗means that the polynomial u0(¯x) contains only those harmonics of v0\nwhich have indices in Ω M. Suppose w0(¯x) =v0(¯x)−u0(¯x).Then, since 1 < q′\nj=qj\nqj−1<\n2, j= 1,...,m, and by the Percevals equality, we have\n/bardblw0/bardbl¯q′,¯θ(2)′/lessorequalslant/bardblw0/bardbl2/lessorequalslantC12nm\n2.\nTherefore, for the function P0(¯x) =C−1\n12−nm\n2w0(¯x), the inequality holds /bardblP0/bardbl¯q′,¯θ(2)′/lessorequalslant1.\nNow using the formula (7), we obtain\neM/parenleftBig\nBr\n¯p,¯θ(1)τ/parenrightBig\n¯q,¯θ(2)>> eM(f0)¯q,¯θ(2)>>2−n(m\n2+r)2−nm\n2(2nm−M)>>\n>>2−n(m+r)2nm>> M−r\nm.\nSo\neM/parenleftBig\nBr\n¯p,¯θ(1),τ/parenrightBig\n¯q,¯θ(2)>> M−r\nm.\nIt proves Theorem 3.\nCorollary. Let 1< p/lessorequalslant2< q <∞, 1/lessorequalslantτ/lessorequalslant∞andr=m\np. Then\neM/parenleftbig\nBr\np,τ/parenrightbig\nq≍M−1\n2(logM)1−1\nθ.\nThe proof follows from the second item of Theorem 2.1 if pj=θ(1)\nj=p, qj=θ(2)\nj=\nq,j= 1,...,m.\nRemark. In the case pj=θ(1)\nj=p, qj=θ(2)\nj=q,j= 1,...,mandr > m(1\np−1\nq),,\nthe results of R.A. DeVore and V.N. Temlyakov [20] follow from Theore m 1 - 3. If\n1< p/lessorequalslant2< q <∞andm(1\np−1\nq)< r/lessorequalslantm\np,the results of S.A. Stasyuk [16], [17] follow\nfrom the first and second items of Theorem 1.\nThe cases pj=θ(1)\nj, qj=θ(2)\nj,j= 1,...,m.of Theorem 1 - 3 were announced in [25] and\nin of Theorem 1 the first item proved [26].\nReferences\n[1] Blozinski A.P. , Multivariate rearrangements and Banach functio n spaces with mixed norms, Trans.\nAmer. Math. Soc., 263(1) (1981), 146-167 .\n[2] Nikol’ski S. M., Approximation of classes of functions of several v ariables and embeding theorems,\nMoscow, 1977 (English transl. of lst ed., 1975, Springer - Verlag, Ne w York )\n[3] Besov O.V., Investigation of on family of functional spaces in conn ection with the theorems of\nimbedding and extension, Trudy Mat. Inst. Akad. Nauk SSSR, 60 (1 961), 42 – 61 (in russain).\n[4] Stechkin S.B., On the absolute convergence of orthogonal serie s. Doklad. Akadem. Nauk SSSR ,\n102(2) (1955), 37 – 40.\n[5] IsmagilovR.S., Widths ofsetsinlinearnormedspacesandtheappro ximationoffunctions bytrigono-\nmetric polynomials, Uspehi mathem. nauk, 29(3) (1974), 161 – 178 .\n[6] Belinsky E.S., Approximation by a ’floating’ system of exponents on the classes of smooth periodic\nfunctions. Mat. sb., 132(1) (1987), 20 – 27.\n[7] MaiorovV.E., On linear widths of Sobolev classes and chains of extre mal subspaces, Mat. sb., 113(3)\n(1980), 437 – 463.ONM– TERMS APPROXIMATIONS BESOV CLASSES IN LORENTZ SPACES 15\n[8] Kashin B.S., Approximation properties of complete orthnormal sy stems , Trudy Mat. Inst. Steklov,\n172 (1985), 187 – 201.\n[9] DeVore R.A., Nonlinear approximation, Acta Numerica, 7 (1998), 51 – 150.\n[10] Temlyakov V. N., Nonlinear methods approximation, Foundations Computational Mathematics, 3\n(2003), 33–107.\n[11] Romanyuk A.S., On the best M− −term trigonometric approximations for the Besov classes of\nperiodic functions of many variables, Izv. Ros. Akad. Nauk, Ser. M at., 67(2) (2003), 61 – 100.\n[12] Dinh Dung, On asymptotic order of n - term approximations and n on-linear - n widths, Vietnam\nJournal Math., 27(4) (1999), 363 - 367.\n[13] Wang Heping, Sun Yongsheng Representation and m- term appr oximation for anisotropic classes,\nin S.M. Nikol’skii, et al. (Eds), Theory of approximation of function and applications , Institute of\nRussian Academy, (2003), P. 250 - 268.\n[14] Duan L. Q., Fang G.S. Trigonometric widths and best N– term approximations of the generalized\nperiodic Besov classes BΩ\np,θ, Journal Math. Resear. and Expos., 31(1) (2011), 129 – 141.\n[15] Hansen M., Sickel W., Best m– term approximation and Lizorkin – Triebel spaces, Journal Appro x-\nimation Theory, 163 (2011), 923 – 954.\n[16] Stasyuk S.A., Best m– term trigonometric approximation for the classes Br\np,θof functions of low\nsmoothness, Ukrain. Mathem. Journal, 62(1) (2010), 114 –122.\n[17] Stasyuk S.A., Best m– term trigonometric approximation 0f periodic function of several variables\nfrom Nikol’skii - Besov classes for small smoothness, Journal of Ap proximation Theory, 177 (2014),\n1 - 16.\n[18] Akishev G., On the exact estimations of the best M− −term approximation of the Besov class,\nSiberian Electronic Mathematical Reports , 7 (2010), 255 – 274.\n[19] Akishev G., On the order of the M−−term approximation classes in Lorentz spaces, Matematical\nJournal. Almaty , 11(1) (2011), 5 - 29.\n[20] DeVore R.A., Temlyakov V.N., Nonlinear approximation by trigonome tric sums, Jour. Fourier Anal-\nysis and applications , 2(1) (1995), 29–48\n[21] Lizorkin P.I., Generalized Holder spaces B(r)\np,θand their relations with the Sobolev spaces L(r)\np, Sib.\nMat. Zh., 9(5) (1968) , 1127 - 1152.\n[22] Akishev G., Inequalities of distinct metric of polynomials in Lorentz spaces with mixed norm, First\nErjanov reading, Pavlodar state universuty, (2004), 211-215.\n[23] Korneichuk N.P., Extreme problem in the theory of approximation , Nauka, Moskow , 1976.\n[24] Kashin B.S., Sahakyan A.A., Orthogonal series, Nauka, Moscow, 1984\n[25] Akishev G., On the M−−term approximation Besov’s classes, International Conference “ Theory of\napproximation of functions and its applications” dedicated to the 7- th anniversary of corresponding\nmember of National Academy of Ukraine, professor A.I. Stepanet s (1942 – 2007) May 28 - June 3,\n(2012), Ukraine, Kamianets-Podiisky, 12.\n[26] Akishev G. Trigonometric widths of the Nikol‘skii - Besov classes in the Lebesgue space with mixed\nnorm. Ukr. Math. Zh. , 66(6) (2014), 723-732.\nDepartment of Mathematics and Information Technology, Buk etov Karaganda State\nUniversity, Universytetskaya 28 , 100028, Karaganda , Repu blic Kazakhstan" }, { "title": "1009.4871v1.Spatial_Damping_of_Propagating_Kink_Waves_in_Prominence_Threads.pdf", "content": "arXiv:1009.4871v1 [astro-ph.SR] 24 Sep 2010SPATIAL DAMPING OF PROPAGATING KINK WAVESIN\nPROMINENCE THREADS\nR. Soler, R. Oliver,and J. L.Ballester\nDepartamentdeF´ ısica,Universitatdeles IllesBalears,E -07122,PalmadeMallorca,Spain\nroberto.soler@uib.es\nABSTRACT\nTransverse oscillations and propagating waves are frequen tly observed in threads\nof solar prominences /filaments and have been interpreted as kink magnetohydrody-\nnamic (MHD) modes. We investigate the spatial damping of pro pagating kink MHD\nwavesintransverselynonuniformandpartiallyionizedpro minencethreads. Resonant\nabsorption and ion-neutral collisions (Cowling’s di ffusion) are the damping mecha-\nnismstakenintoaccount. Thedispersionrelationofresona ntkinkwavesinapartially\nionized magnetic flux tube is numerically solved by consider ing prominence condi-\ntions. Analytical expressions of the wavelength and dampin g length as functions of\nthe kink mode frequency are obtained in the Thin Tube and Thin Boundary approxi-\nmations. For typically reported periods of thread oscillat ions, resonant absorption is\nanefficientmechanismforthekinkmodespatialdamping,whileion -neutralcollisions\nhave a minor role. Cowling’s di ffusion dominates both the propagation and damping\nfor periods much shorter than those observed. Resonant abso rption may explain the\nobserved spatial damping of kink waves in prominence thread s. The transverse in-\nhomogeneity length scale of the threads can be estimated by c omparing the observed\nwavelengths and damping lengths with the theoretically pre dicted values. However,\nthe ignorance of the form of the density profile in the transve rsely nonuniform layer\nintroducesinaccuracies in thedeterminationoftheinhomo geneitylengthscale.\nSubject headings: Sun: oscillations – Sun: filaments, prominences – Sun: coron a –\nMagnetohydrodynamics(MHD)– Waves\n1. INTRODUCTION\nWaves and oscillatory motions are frequently reported in th e observations of solar promi-\nnences and filaments (see reviews by Oliver& Ballester 2002; Ballester 2006; Engvold 2008;– 2 –\nMackay et al. 2010). In high-resolution observations, the p rominence fine structures (threads) of-\ntendisplaytransverseoscillationsofsmallamplitude(e. g.,Linet al.2005,2007,2009;Okamotoet al.\n2007;Ninget al.2009),whichhavebeeninterpretedaskinkm agnetohydrodynamic(MHD)waves\n(e.g., D´ ıazet al. 2002; Dymova& Ruderman 2005; Terradas et al. 2008; Lin etal. 2009). The ob-\nserved threads in H αimages are between 3,000 km and 28,000 km long, and between 10 0 km\nand 600 km wide (Lin 2004; Linet al. 2008). The threads outlin e part of much larger magnetic\nflux tubes which are probably rooted in the solar photosphere . The majority of observed periods\nof transverse thread oscillations roughly range between 1 m in and 10 min, but a few detections\nof longer periods of about 20 min have been also informed (e.g ., Yiet al. 1991; Lin 2004). The\nwavelengths are usually between 700 km and 8,000 km, althoug h values up to 250,000 km have\nbeen reported (Okamotoet al. 2007). Recently, Soleret al. ( 2010a) pointed out that the short pe-\nriods and wavelengths are consistent with an interpretatio n in terms of propagating waves, while\nperiods larger than 10 min and wavelengths longer than 100,0 00 km could correspond to stand-\ning oscillations of the whole magnetic tube. In the case of st anding oscillations, the value of the\nwavelengthisnotstronglyinfluencedbythethreadproperti esbutismainlydeterminedbythetotal\nlength of the magnetic tube, since the fundamental kink mode wavelength is twice the length of\nthe tube, approximately (see details in Soler et al. 2010a). Although there are no direct measure-\nments of the length of prominence magnetic tubes, this param eter is estimated around 105km.\nThisroughestimationis inagreement withthewavelengthsr eported byOkamotoet al. (2007). In\naddition, a common feature of the observations is that the os cillations are strongly damped (e.g.,\nTerradas et al. 2002; Ninget al. 2009;Lin et al.2009).\nMotivatedbytheobservationalevidence,greate fforthasbeenrecentlydevotedtothetheoret-\nicalstudyofbothtemporalandspatialdampingofMHDwavesi nprominenceplasmas. Temporal\ndamping is investigated for waves with fixed wavelength, whi le spatial damping is studied for\npropagating waves with fixed frequency. Both phenomena have been extensively investigated in\nunbounded and homogeneous prominence plasmas by assuming d ifferent damping mechanisms\n(e.g.,Carbonell et al.2004,2006,2010;Forteza etal.2007 ,2008). ThereaderisreferredtoOliver\n(2009),Arregui &Ballester(2010),andreferencestherein foracompleteaccountofthetheoretical\nworks.\nInthecaseofprominencethreadoscillations,workssofarh avefocusedontemporaldamping\nbymechanismsas,e.g.,non-adiabatice ffects(Soleret al.2008),ion-neutralcollisions(Soleret a l.\n2009b), and resonantabsorption(Arreguiet al. 2008, 2010; Soleret al.2009a,c, 2010a). Thecon-\nclusions of these works indicate that resonant absorption i s efficient enough to provide realistic\nkink mode damping times consistent with the reported strong damping, whereas non-adiabatic\neffects are negligible and ion-neutral collisions are only imp ortant for shorter wavelengths than\nthoseobserved. Inthecaseofspatialdamping,P´ ecseli & En gvold(2000)studiedthee ffectofion-\nneutralcollisionsbutrestrictedthemselvestoAlfv´ enwa vesandkinkmodeswerenotinvestigated.– 3 –\nAlthough spatial damping of kink waves has been studied in th e context of coronal loops (e.g.,\nPascoeet al. 2010; Terradas etal. 2010a), to our knowledge n o detailed investigation taking into\naccount the peculiar properties of prominence threads can b e found in the existing literature. The\npurposeofthispaperistofillthisgapintheliterature,ast herecentobservationsofwavedamping\ninsolarprominencesneed tobeunderstood.\nHere, we study the spatial damping of propagating kink waves in prominence threads. Our\nmodel is composed of a cylindrical magnetic flux tube with par tially ionized prominence plasma,\nrepresenting a thread, surrounded by a fully ionized corona l environment. The thread is non-\nuniforminthetransversedirection. Resonantabsorptiona ndion-neutralcollisionsareassumedas\nthe damping mechanisms. We use the β=0 approximation, with βthe ratio of the gas pressure\nto the magnetic pressure, and the Thin Boundary approach to d escribe the effect of resonant ab-\nsorption in the Alfv´ en continuum using the connection form ulas for the perturbations across the\nresonantlayer(e.g.,Sakurai et al.1991;Goossenset al.19 92). Wedeterminethedominantdamp-\ningmechanismandobtainanalyticalexpressionsforthewav elength,thedampinglength,andtheir\nratio asfunctionsofthekinkmodefrequency.\nThis paper is organized as follows. Section 2 contains the de scription of the model configu-\nrationand thebasicEquations. First, theproblemisattack edanalyticallyinSection 3byadopting\nthethintubeapproximation. Lateron,thefulldispersionr elationisnumericallysolvedandapara-\nmetric study of the wavelength and damping length of thekink modeas functions of the period is\nperformed inSection 4. Finally,theconclusionsofthiswor k aregiveninSection 5.\n2. MODEL ANDDISPERSION RELATION\nFig. 1.—Sketch oftheprominencethread modeladoptedinthi swork.– 4 –\nThe equilibrium configuration is composed of a straight magn etic cylinder of radius aem-\nbedded in a homogeneous environment representing the coron al medium (see Figure 1). We use\ncylindrical coordinates, namely r,ϕ, andzfor the radial, azimuthal, and longitudinal coordinates,\nrespectively. The magnetic field is uniform and along the axi s of the cylinder, B=Bˆez, withB\nconstant everywhere. Hereafter, subscripts p and c denote p rominence and coronal quantities, re-\nspectively. The density within the prominence thread is den oted byρp, while the coronal density\nisρc. Bothρpandρcare homogeneous. A transverse transitional layer is includ ed in the radial\ndirection, where the density varies continuously between t he internal and external densities. We\ndo not specify the form of the density profile at this stage. Th e transverse inhomogeneous length\nscale in the transitional layer is given by the radio l/a, withlthe thickness of the layer. This ratio\nranges from l/a=0 if no transitional layer is present, to l/a=2 if the whole tube is radially\ninhomogeneous. Due to the presence of the transverse transi tional layer, the kink mode is reso-\nnantlycoupledtoAlfv´ encontinuummodes. Theresonancele adstothekinkmodedampingasthe\nenergyistransferredtoAlfv´ enmodesattheAlfv´ enresona nceposition. Thismechanismisknown\nas resonantabsorption.\nThe prominence plasma is partially ionized and we adopt the s ingle-fluid formalism (e.g.,\nBraginskii1965). Theionizationdegreeisarbitraryandis denotedherebythemeanatomicweight\nof the prominence material, ˜ µp. This parameter takes values in the range 0 .5≤˜µp≤1, where\n˜µp=0.5 corresponds to a fully ionized plasma and ˜ µp=1 to a fully neutral gas (see details\nin, e.g., Forteza etal. 2007; Soleret al. 2009b). The extern al coronal medium is assumed fully\nionized. The basic MHD Equations governing a partially ioni zed plasma can be found in, e.g.,\nFortezaet al. (2007); Pinto et al. (2008); Soler (2010). By a ssuming theβ=0 approximation and\nlinear perturbations from the equilibrium state, the basic MHD Equations discussed in this work\nare\nρ∂v\n∂t=1\nµ(∇×b)×B, (1)\n∂b\n∂t=∇×(v×B)+∇×/braceleftbiggηC\nB2[(∇×b)×B]×B/bracerightbigg\n, (2)\nwhereρisthelocaldensity,and v=/parenleftig\nvr,vϕ,vz/parenrightig\nandb=/parenleftig\nbr,bϕ,bz/parenrightig\narethevelocityandthemagnetic\nfieldperturbations,respectively. Notethat vz=0intheβ=0approximation. Inapartiallyionized\nplasma,theinductionequationcontainsatermaccountingf orCowling’sdiffusion,i.e.,thesecond\nterm on the right-hand side of Equation (2). Cowling’s di ffusion represents enhanced magnetic\ndiffusioncausedbyion-neutralcollisions,whichisseveralor dersofmagnitudemoree fficientthan\nclassical Ohm’s diffusionand is thedominante ffect in partially ionized plasmas (Cowling 1956).\nForthisreason,hereweneglectOhm’sdi ffusionandothertermsofminorimportancepresentinthe\ngeneralizedinductionequation(see,e.g.,Equation(14)o fFortezaet al.2007). Cowling’sdi ffusion\ncoefficient,ηC, depends on the ionization degree through ˜ µpas well as on the plasma physical\nconditions. The expression of ηCfor a hydrogen plasma can be found in, e.g., Pinto et al. (2008 )– 5 –\nand Soleret al. (2009b), whereas for a plasma composed of hyd rogen and helium see Soleret al.\n(2010b). Astheeffectofheliumisnegligibleforrealisticheliumabundances inprominences,here\nweconsiderapurehydrogenplasma. Thee ffectofCowling’sdi ffusionisneglectedintheexternal\nmediumbecause thecoronais assumedfullyionized.\nWe follow an approach based on normal modes. Since ϕandzare an ignorable coordinates,\nthe perturbations are expressed proportional to exp (imϕ+ikzz−iωt), whereωis the oscillatory\nfrequency, kzis the longitudinal wavenumber, and mis the azimuthal wavenumber ( m=1 for the\nkinkmode). Alternatively,theproblemcouldbeinvestigat edby meansoftime-dependentsimula-\ntionsof drivenwaves as in Pascoeet al. (2010). However, in t helinear regimethe di fferent values\nofmandkzare decoupled from each other, and a normalmodeanalysisis a simplerprocedure for\nlinear waves. If Cowling’s di ffusionis neglected, our configuration corresponds to that st udiedby\nTerradas et al. (2010a) for propagatingkinkwaves in corona l loops. Weextend theirinvestigation\nby incorporatingthee ffect ofCowling’sdi ffusionduetoion-neutralcollisions\nByusingtheThinBoundary(TB)approach(seedetailsin,e.g .,Goossenset al.2006;Goossens\n2008), theanalyticaldispersionrelationfor resonantlyd ampedkink wavespropagatingin atrans-\nversely nonuniform and partially ionized prominence threa d was obtained by Soleret al. (2009c,\nEquation (25)). If partial ionization is not considered and the effect of Cowling’s di ffusion is ab-\nsent, the dispersion relation of Soleret al. (2009c) reduce s to that investigated by Terradas et al.\n(2010a, Equation (28)) for kink waves in coronal loops. Sole ret al. (2009c) checked that the so-\nlutionsof theirdispersionrelation are in excellent agree ment with the solutionsobtained from the\nfull numerical integration of the MHD equations beyond the T B approximation. Therefore, the\ndispersion relation derived by Soleret al. (2009c) correct ly describes the kink mode behavior in\nourmodelandcomplicatednumericalintegrationsarenotne eded. Thedispersionrelationobtained\nby Soleret al.(2009c)inthecase ofa straightand homogeneo usmagneticfield is\nnc\nρc/parenleftig\nω2−k2zv2\nAc/parenrightigK′\nm(nca)\nKm(nca)−mp\nρp/parenleftig\nω2−k2zΓ2\nAp/parenrightigJ′\nm/parenleftig\nmpa/parenrightig\nJm/parenleftig\nmpa/parenrightig=−iπm2/r2\nA\nω2|∂rρ|rA, (3)\nwhereJmandKmaretheBesselfunctionandthemodifiedBesselfunctionofth efirstkindoforder\nm(Abramowitz& Stegun1972), respectively,and thequantiti esmpandncare defined as\nm2\np=/parenleftig\nω2−k2\nzΓ2\nAp/parenrightig\nΓ2\nAp,n2\nc=/parenleftig\nk2\nzv2\nAc−ω2/parenrightig\nv2\nAc, (4)\nwhereΓ2\nAp=v2\nAp−iωηCis the modified prominence Alfv´ en speed squared (Forteza et al. 2008),\nwithvAp=B√µρpandvAc=B√µρcthe prominence and coronal Alfv´ en speeds, respectively, a nd\nµ=4π×10−7N A−2the magnetic permeability. In addition, rAis the Alfv´ en resonance position,\nand|∂rρ|rAistheradialderivativeofthetransversedensityprofileat theAlfv´ enresonanceposition.– 6 –\nSoleret al.(2009c)studiedthetemporaldampingofkinkwav es,hencetheyassumedafixed,\nrealkzand solved Equation (3) to obtain the complex frequency. Her e, we investigate the spatial\ndampingandproceedtheotherwayround,i.e.,wefixareal ωandsolveEquation(3)toobtainthe\ncomplexwavenumber. Then, theperiod, P, wavelength,λ, and dampinglength, LD, are computed\nas follows,\nP=2π\nω, λ=2π\nkzR,LD=1\nkzI, (5)\nwithkzRandkzIthereal and imaginaryparts of kz, respectively.\n3. ANALYTICALAPPROXIMATIONS\nSomeanalyticalprogresscanbeperformedbeforesolvingEq uation(3)bymeansofnumerical\nmethods. To do so, we adopt the Thin Tube (TT) limit, i.e., λ/a≫1. A fist-order expansion of\nEquation(3)givesthedispersionrelationin boththeTT and TBapproximations,namely\nρp/parenleftig\nω2−k2\nzΓ2\nAp/parenrightig\n+ρc/parenleftig\nω2−k2\nzv2\nAc/parenrightig\n−iπm\nrAρpρc\n|∂rρ|rA/parenleftig\nω2−k2\nzΓ2\nAp/parenrightig/parenleftig\nω2−k2\nzv2\nAc/parenrightig\nω2=0.(6)\nIfbothCowling’sdi ffusionand resonant absorptionare omitted,thesolutiontoE quation(6)is\nk2\nz=ω2\nc2\nk, (7)\nwithc2\nk=2B2\nµ(ρp+ρc)the kink speed squared. Equation (7) corresponds to the idea l, undamped kink\nmode. The solutionsto Equation(6) considering thedi fferent dampingmechanisms are discussed\nnext.\n3.1. Damping by Cowling'sdi ffusion\nIn theabsence of transversetransitionallayer, i.e., l/a=0, resonant absorptiondoes not take\nplace and the damping is due to Cowling’sdi ffusion exclusively. In such a case, the third term on\nthe left-hand side of Equation (6) is absent. We write the wav enumber as kz=kzR+ikzIand use\nEquation(6)toobtaintheexactexpressionsfor k2\nzRandk2\nzI, namely\nk2\nzR=1\n2ω2c2\nk\nc4\nk+ω2¯η2\nC/radicaligg\n1+ω2¯η2\nC\nc4\nk+1, (8)\nk2\nzI=1\n2ω2c2\nk\nc4\nk+ω2¯η2\nC/radicaligg\n1+ω2¯η2\nC\nc4\nk−1, (9)– 7 –\nwith ¯ηC=ρp\nρp+ρcηC. By combining Equations (8) and (9), we compute the ratio of t he damping\nlengthtothewavelengthas\nLD\nλ=kzR\n2πkzI=1\n2πc2\nk+/radicalig\nc4\nk+ω2¯η2\nC\nω¯ηC. (10)\nEquations(8)–(10)areexactexpressionsthatcanbefurthe rsimplifieddependingonthevalue\noftheratioω2¯η2\nC\nc4\nk. Forω2¯η2\nC\nc4\nk≪1, i.e., in thelimitoflowfrequency ( ωsmall)and/orlargeionization\ndegree(¯ηCsmall),Equations(8)–(10)simplifyto\nk2\nzR≈ω2\nc2\nk/parenleftbigg\n1+ω2¯η2\nC\nc4\nk/parenrightbigg≈ω2\nc2\nk, (11)\nk2\nzI≈1\n4ω4¯η2\nC\nc6\nk/parenleftbigg\n1+ω2¯η2\nC\nc4\nk/parenrightbigg≈1\n4ω4¯η2\nC\nc6\nk, (12)\nLD\nλ≈1\nπc2\nk\nω¯ηC. (13)\nOn the contrary, ifω2¯η2\nC\nc4\nk≫1, i.e., high frequency and /or small ionization degree, the equivalent\nexpressionsare\nk2\nzR≈k2\nzI≈1\n2ω\nc2\nk¯ηC, (14)\nLD\nλ≈1\n2π. (15)\nThus, forω2¯η2\nC\nc4\nk≪1 the ratio of the damping length to the wavelength is inverse ly proportional to\nbothωand ¯ηC, andk2\nzRcoincides with the ideal value (Equation (7)). This case cor responds to a\nweaklydampedkinkmode. Ontheotherhand,forω2¯η2\nC\nc4\nk≫1,LD/λisindependentofωand ¯ηCand\nthe wave behavior is governed by di ffusion. By assuming typical values for the parameters in the\ncontext of oscillating prominence threads, e.g., P=3 min,B=5 G, andρp/ρc=200, we obtain\nω2¯η2\nC\nc4\nk≈6×10−17for ˜µp=0.5andω2¯η2\nC\nc4\nk≈1.6×10−4for ˜µp=0.99,meaning thatthecaseω2¯η2\nC\nc4\nk≪1\nismorerealisticin thecontextofoscillatingthreads even foran almostneutral plasma.\n3.2. Damping by resonant absorption andCowling's di ffusion\nNext,wetakethecase l/a/nequal0intoaccountandstudythecombinede ffectofresonantabsorp-\ntion and Cowling’s di ffusion. The third term on the left-hand side of Equation (6) is now present.– 8 –\nAsbefore, wewrite kz=kzR+ikzIand putthisexpressioninEquation(6). Sinceitisvery di fficult\ntogiveexactexpressionsfor kzRandkzIinthegeneralcase,wefocuson LD/λandrestrictourselves\ntoω2¯η2\nC\nc4\nk≪1. Following the procedure of Terradas et al. (2010a), we ass ume weak damping, i.e.,\nkzI≪kzR, and neglect terms with k2\nzI. The following process is long but straightforward, and we\nrefer the reader to Terradas et al. (2010a) for details. Fina lly, we arrive at the expression for the\nratio ofthedampinglengthtothewavelengthas\nLD\nλ≈/parenleftigg\nπω¯ηC\nc2\nk+m\nFl\naρp−ρc\nρp+ρc/parenrightigg−1\n, (16)\nwhere the first term within the parentheses accounts for Cowl ing’s diffusion and the second term\nfor resonant absorption. The factor Fin the second term takes di fferent values depending on\nthe density profile within the inhomogeneous layer. For exam ple,F=4/π2for a linear profile\n(Goossenset al. 2002), while F=2/πfor a sinusoidal profile with rA≈a(Ruderman &Roberts\n2002). If the term related to resonant absorption is absent, Equation (16) reverts to Equation (13).\nOn the other hand, if the term related to Cowling’s di ffusion is dropped, Equation (16) coincides\nwithEquation(13)ofTerradas et al.(2010a).\nTherelativeimportanceofthetwotermsinEquation(16)can beassessedbyperformingtheir\nratio as\nǫ≡(LD/λ)RA\n(LD/λ)C≈πFa\nlω¯ηC\nc2\nkmρp+ρc\nρp−ρc=πFa\nlωηC\nc2\nkmρp\nρp−ρc, (17)\nwhere(LD/λ)RAand(LD/λ)Cstand for the damping ratio by resonant absorption and Cowli ng’s\ndiffusion, respectively. By considering as before P=3 min,B=5 G, andρp/ρc=200, and\nadopting a linear profile with l/a=0.2, we obtainǫ≈8×10−8for ˜µp=0.5 andǫ≈0.12\nfor ˜µp=0.99, meaning that in the TT limit resonant absorption dominat es the kink mode spatial\ndamping for typical parameters of thread oscillations. Thi s result is equivalentto that obtained by\nSoleret al. (2009c) inthecaseoftemporaldamping.\n4. NUMERICALRESULTS\nNow, we solve the dispersion relation (Equation (3)) by mean s of standard numerical proce-\ndures. In the following figures, both the wavelength, λ, and the damping length, LD, are plotted\nin dimensionless form with respect to the thread mean radius ,a. The period, P, is computed in\nunits of the internal Alfv´ en travel time, τAp=a/vAp. Unless otherwise stated, the results have\nbeen computed with ρp=5×10−11kg m−3,ρp/ρc=200, and B=5 G. With these parameters,\nvAp≈63 kms−1and, fora=100km,τAp≈1.59 s.\nFigure 2 displaysλ/a,LD/a, andLD/λversusP/τApfor the case l/a=0, i.e., the damping– 9 –\nis due to Cowling’s di ffusion exclusively. We compute the results for di fferent values of ˜µp. The\nshaded areas in Figure 2 and in the other figures represent the range of observed periods of trans-\nversethread oscillations,i.e., 1min–10min,correspondi ngto40/lessorsimilarP/τAp/lessorsimilar400,approximately.\nRegardingthewavelength,weseethatthee ffectofCowling’sdi ffusionisonlyrelevantforperiods\nmuchshortedthanthoseobserved. ThisisinagreementwithE quations(11)and(14). Ontheother\nhand, an almost neutral plasma, i.e., ˜ µp→1, has to be considered to obtain an e fficient damping\nand to achieve small values of LD/λwithin the relevant range of periods. Although we do not\nknowtheexactionizationdegreein prominencethreads,suc hvery largevaluesof ˜ µpare probably\nunrealistic (see, e.g., Gouttebroze& Labrosse 2009). The a nalytical expressions for λ,LD, and\nLD/λintheTTcasegivenbyEquations(8),(9),and(10),respecti vely,areingoodagreementwith\nthefullresultsinthewholerange ofperiods(seesymbolsin Figure 2).\nNext, we study the general case l/a/nequal0. We adopt a sinusoidal density profile within the\ninhomogeneoustransitionallayer(Ruderman &Roberts2002 ). AstheAlfv´ enresonanceposition,\nrA,isneededforthecomputationsoftheresonantdamping,wef ollowatwo-stepprocedure. First,\nwe solve the dispersion relation for a fixed ωin the case l/a=0 and determine kzR. Then, we\nassume that the value of kzRis approximately the same in the case l/a/nequal0, meaning that the\nresonant condition is ω=kzRvA(rA). In the case of a sinusoidal profile, the expression of the\nresonantpositioncan beanalyticallyobtainedfrom theres onantconditionas\nrA=a+l\nπarcsinρp+ρc\nρp−ρc−2v2\nApk2\nzR\nω2ρp\nρp−ρc. (18)\nFinally, we compute |∂rρ|rAusing the previously determined rAby means of Equation (18) and\nsolvethedispersionrelationwith theseparameters to obta intheactual kzRandkzI. Figure3 shows\nthe results of these computations for di fferent values of l/awhen the ionization degree has been\nfixed to ˜µp=0.8, whereas Figure 4 displaystheequivalentcomputationsfo r different valuesof ˜µp\nwhen thetransverseinhomogeneitylengthscale hasbeen fixe d tol/a=0.2. Since thewavelength\nis not affected by the value of l/aand has the same behavior as in Figure 2(a), both Figures 3 and\n4 focus on LD/aandLD/λ. We obtain two di fferent behaviors of the solutions depending on the\nperiod. For small P/τAp, the damping length is independent of l/aand is governed by the value\nof ˜µp. On the contrary, for large P/τApthe damping length depends on l/abut is independent of\n˜µp. This result indicates that resonant absorptiondominates thedamping for large P/τAp, whereas\nCowling’sdiffusionismorerelevantforsmall P/τAp. Theapproximatetransitionalperiod,namely\nPtr,inwhichthedampinglengthbyCowling’sdi ffusionbecomessmallerthanthatduetoresonant\nabsorption can be estimated by setting ǫ≈1 in Equation (17) and writing Ptr=2π/ω. Then, one\nobtains\nPtr≈2π2Fa\nl¯ηC\nc2\nkmρp+ρc\nρp−ρc=2π2Fa\nlηC\nc2\nkmρp\nρp−ρc. (19)\nThistransitionalperiodisingoodagreementwiththenumer icalresults(seetheverticaldottedline– 10 –\nFig. 2.—Resultsforthekinkmodespatialdampinginthecase l/a=0: (a)λ/a,(b)LD/a, and(c)\nLD/λversusP/τApfor ˜µp=0.5,0.6,0.8,and0.95. Symbolsinpanels(a),(b),and(c)co rrespondto\ntheanalyticalsolutionintheTTapproximationgivenbyEqu ations(8),(9), and(10),respectively,\nwhile the horizontal dotted line in panel (c) corresponds to the limit of LD/λfor high frequencies\n(Equation(15)). Theshadedarea denotestherangeofobserv edperiodsofthread oscillations.– 11 –\nFig. 3.— Results for the kink mode spatial damping in the case l/a/nequal0: (a)LD/aand (b)LD/λ\nversusP/τApforl/a=0.05, 0.1, 0.2, and 0.4, with ˜ µp=0.8. Symbols in panel (b) correspond to\nthe analytical solution in the TT approximation given by Equ ation (16), while the vertical dotted\nline is the approximate transitional period given by Equati on (19) for l/a=0.1. The shaded area\ndenotestherangeofobservedperiodsofthread oscillation s.\nFig. 4.— Results for the kink mode spatial damping in the case l/a/nequal0: (a)LD/aand (b)LD/λ\nversusP/τApfor ˜µp=0.5, 0.6, 0.8, and 0.95, with l/a=0.2. Symbols in panel (b) correspond to\nthe analytical solution in the TT approximation given by Equ ation (16). The shaded area denotes\ntherangeofobservedperiodsofthread oscillations.– 12 –\nin Figure 3(b)). In addition, we see that Ptris much smaller than the typically observed periods,\nindicatingthat resonantabsorptionisthedominantdampin gmechanismin therelevantrange.\nFinally, we check that the analytical approximationof LD/λgivenby Equation (16) provides\nanaccuratedescriptionofthekinkmodespatialdampingint herelevantrangeofperiods(compare\nthesymbolsandthesolidlinesinFigures 3(b)and 4(b)).\n5. DISCUSSION AND CONCLUSION\nIn this paper, we have studied the spatial damping of kink wav es in prominence threads.\nResonant absorption and Cowling’s di ffusion are the damping mechanisms taken into account.\nBoth analytical expressions and numerical results indicat e that, in the range of typically observed\nperiods of prominence thread oscillations, the e ffect of Cowling’s di ffusion (and so the ionization\ndegree) is negligible. On the other hand, resonant absorpti on provides an efficient damping in\nagreement with the study of Terradas et al. (2010a) in the con text of coronal loop oscillations.\nThese conclusions are equivalent to those obtained by Soler etal. (2009c) in the case of temporal\ndamping.\nWe point out that small values of LD/λare obtained by resonant absorption in the observa-\ntionally relevant range of periods, which is consistent wit h the reported strong damping of the\noscillations. The analytical estimation of LD/λgiven by Equation (16) is very accurate in the ob-\nservationallyrelevantrangeofperiods,andthecontribut ionofCowling’sdi ffusioncanbedropped\nfromEquation(16)becausetheplasmaionizationdegreetur nsouttobeirrelevantforthedamping.\nTherefore, forkinkmodes( m=1)theradio LD/λsimplifiesto\nLD\nλ≈Fa\nlρp+ρc\nρp−ρc, (20)\nwhich coincides with the expression providedby Terradas et al. (2010a). Asρp+ρc\nρp−ρc→1 for typical\nprominenceandcoronaldensities,thisfactorcanbedroppe dfromEquation(20),meaningthatthe\nratioLD/λdepends almost exclusively on the transverse inhomogeneit y length scale, l/a, and the\nform ofthedensityprofilethrough Fas\nLD\nλ≈Fa\nl. (21)\nInthecaseofcoronallooposcillationsstudiedbyTerradas et al.(2010a),thefactorρp+ρc\nρp−ρccannotbe\ndroppedfromtheirexpressions,meaningthatincoronalloo pstheratio LD/λsignificantlydepends\non the density contrast. Therefore, information about the p arameters l/aandFin prominence\nthreads could be determined by using Equation (21) along wit h accurate measurements of the– 13 –\ndamping length and the wavelength provided from the observa tions. However, since the precise\nform of the transverse density profile in prominence threads is unknown, we have to assume an\nad hoc profile, i.e., a value of F, to infer the transverse inhomogeneity length scale from th e\nobservations,whichcan introducesomeuncertaintiesinth eestimationof l/a.\nFor example, let us assume that the ratio LD/λhas been determined from an observation of\ndampedkinkwavesinaprominencethreadandwewanttocomput ethetransverseinhomogeneity\nlength scale of the thread. For simplicity, we consider that the transverse density profile in the\ninhomogeneous layer is either linear or sinusoidal. Denoti ng as(l/a)linthe value of l/acomputed\nassuming a linear profile, and (l/a)sinthe corresponding value for a sinusoidalprofile, the relati on\nbetween bothofthemis(l/a)lin\n(l/a)sin=π\n2≈1.57, (22)\npointingoutthattherelativeuncertaintyof l/aislargerthan50%,andtheinaccuracycouldbeeven\nlargerifotherprofilesareconsidered. Thisfactshouldbet akenintoaccountinfutureseismological\ndeterminationsofthisparameter.\nThe present investigation is a first step for the study of the s patial damping of kink waves\nin prominence fine structures. Here, we have adopted a simple model of a prominence thread.\nSome effects that might influence the kink mode propagation and dampi ng are not included in\nthe present paper. Among them, plasma inhomogeneity along t he thread may affect somehow\nthe amplitude of a propagating kink mode, whereas the presen ce of flows affects the damping by\nresonantabsorption(seeTerradas et al.2010b). Theinfluen ceoftheseandothere ffectswillbethe\nsubjectofforthcomingworks.\nTheauthorsacknowledgethefinancialsupportreceivedfrom theSpanishMICINNandFEDER\nfunds (AYA2006-07637). The authors also acknowledge discu ssion within ISSI Team on Solar\nProminence Formation and Equilibrium: New data, new models . RS thanks the CAIB for a fel-\nlowship.\nREFERENCES\nAbramowitz,M.,&Stegun,I.A.1972,HandbookofMathematic alFunctions,DoverPublications\nArregui,I., Terradas, J., Oliver,R., & Ballester, J.L. 200 8,ApJ,682, L141\nArregui,I. et al. 2010,inpreparation\nArregui,I., &Ballester, J. L. 2010,Space Sci. Rev.,inpres s– 14 –\nBallester, J. L.2006,Phil. Trans.R. Soc. A, 364,405\nBraginskii,S. I. 1965,Rev. PlasmaPhys.,1, 205\nCarbonell, M.,Oliver,R., &Ballester, J. L.2004,A&A, 415, 739\nCarbonell, M.,Terradas, J., Oliver,R., & Ballester, J.L. 2 006,A&A,460, 573\nCarbonell, M.,Forteza, P., Oliver,R., &Ballester, J. L.20 10,A&A, 515,A80\nCowling,T.G. 1956,MNRAS, 116,114\nEngvold,O.2008,inIAUSymp.247,Waves&Oscillationsinth eSolarAtmosphere: Heatingand\nMagneto-Seismology, ed. R. Erd´ elyi & C. A. Mendoza-Brice˜ no (Cambridge: Cambridge\nUniv.Press), 152\nD´ ıaz, A J.,Oliver,R., &Ballester, J. L.2002, ApJ,580,550\nDymova,M.V., & Ruderman,M. S. 2005,Sol. Phys.,229,79\nForteza, P., Oliver,R., Ballester, J.L. &Khodachenko, M.L . 2007,A&A,461,731\nForteza, P., Oliver,R., &Ballester, J. L.2008,A&A, 492,22 3\nGoossens,M.,Hollweg,J. V., &Sakurai, T. 1992,Sol.Phys., 138, 233\nGoossens,M.,Andries,J., &Aschwanden,M. J.2002,A&A, 394 ,L39\nGoossens,M.,Andries,J., &Arregui,I. 2006,Phil. Trans.R . Soc. A, 364,433\nGoossens, M. 2008, in IAU Symp. 247, Waves & Oscillations in t he Solar Atmosphere: Heating\nand Magneto-Seismology, ed. R. Erd´ elyi & C. A. Mendoza-Bri ce˜ no (Cambridge: Cam-\nbridgeUniv.Press), 228\nGouttebroze,P., & Labrosse,N. 2009,A&A,503,663\nLin,Y. 2004,PhD Thesis,UniversityofOslo,Norway\nLin,Y.,Engvold,O.,RouppevanderVoort,L.H.M.,Wiik,J.E .,&Berger,T.E.2005,Sol.Phys.,\n226,239\nLin,Y., Engvold,O., RouppevanderVoort, L.H. M.,& vanNoor t,M. 2007,Sol. Phys.,246,65\nLin, Y., Martin, S. F., & Engvold, O. 2008, in ASP Conf. Ser. 38 3, Subsurface and Atmospheric\nInfluences on Solar Activity, ed. R. Howe, R. W. Komm, K. S. Bal asubramaniam, & G. J.\nD. Petrie(San Francisco: ASP), 235– 15 –\nLin,Y.,Soler,R.,Engvold,O.,Ballester,J.L.,Langangen ,Ø.,Oliver,R.,&RouppevanderVoort,\nL. H.M. 2009,ApJ,704,870\nMackay, D. H., Karpen, J. T.,Ballester, J. L., Schmieder, B. , & Aulanier, G. 2009,Space Sci Rev,\n151,333\nNing,Z., Cao, W., Okamoto,T. J.,Ichimoto,K.,& Qu,Z. Q. 200 9,A&A,499,595\nOliver,R. &Ballester, J. L.2002,Sol. Phys.,206,45\nOliver,R. 2009,Space Sci Rev,149,175\nOkamoto,T.J, etal. 2007,Science, 318,1557\nPascoe, D. J.,Wright,A. N., &DeMoortel,I. 2010,ApJ,711, 9 90\nP´ ecseli, H., &Engvold,O.2000,Sol. Phys.,194,73\nPinto,C., Galli,D., & Bacciotti,F. 2008,A&A, 484,1\nRuderman, M.,& Roberts,B. 2002,ApJ,577,475\nSakurai, T.,Goossens,M,& Hollweg,J. V. 1991,Sol. Phys.,1 33,227\nSoler, R. 2010,PhD Thesis,Universitatdeles IllesBalears , Spain\nSoler, R., Arregui,I., Oliver,R., &Ballester, J. L. 2010a, ApJ,in press\nSoler, R., Oliver,R., &Ballester, J. L. 2008,ApJ,684,725\nSoler, R., Oliver,R., &Ballester, J. L. 2009b,ApJ, 699,155 3\nSoler, R., Oliver,R., &Ballester, J. L. 2009c,ApJ,707,662\nSoler, R., Oliver,R., &Ballester, J. L. 2010b,A&A, 512,A28\nSoler, R., Oliver,R., Ballester, J.L., &Goossens,M. 2009a ,ApJ,695,L166\nTerradas, J., Molowny-Horas, R., Wiehr, E., Balthasar, H., Oliver, R., & Ballester, J. L. 2002,\nA&A,393,637\nTerradas, J., Arregui,I., Oliver,R., & Ballester, J.L. 200 8,ApJ,678, L153\nTerradas, J., Goossens,M., &Verth, G. 2010a,A&A,inpress\nTerradas, J., Goossens,M., &Ballai, I. 2010b,A&A,515,A46– 16 –\nYi,Z., Engvold,O., &Kiel,S. L.1991,Sol. Phys.,132,63\nThispreprintwaspreparedwiththeAAS L ATEXmacrosv5.2." }, { "title": "1708.08225v2.Magnetic_reheating.pdf", "content": "arXiv:1708.08225v2 [astro-ph.CO] 9 Dec 2017MNRAS 000, 000–000 (0000) Preprint 4 June 2022 Compiled using MNRAS L ATEX style file v3.0\nMagnetic reheating\nShohei Saga,1⋆Hiroyuki Tashiro,2and Shuichiro Yokoyama3,4\n1Yukawa Institute for Theoretical Physics, Kyoto University , Kyoto 606-8502, Japan\n2Department of Physics and Astrophysics, Nagoya University , Nagoya, 464-8602, Japan\n3Department of Physics, Rikkyo University, Tokyo 171-8501, Japan\n4Kavli IPMU (WPI), UTIAS, The University of Tokyo, Kashiwa, Chib a 277-8583, Japan\n4 June 2022\nABSTRACT\nWe provide a new bound on the amplitude of primordial magneti c fields (PMFs) by\nusing a novel mechanism, magnetic reheating . The damping of the MHD fluid motions\nin a primordial plasma brings the dissipation of the PMFs. In the early Universe with\nz/greaterorsimilar2×106, cosmic microwave background (CMB) photons are quickly the rmalized\nwith the dissipated energy and shift to a different Planck dis tribution with a new\ntemperature. In other words, the PMF dissipation changes th e baryon-photon number\nratio and we name such a process magnetic reheating . From the current baryon-photon\nnumber ratio obtained from the BBN and CMB observations, we p ut a strongest\nconstraint on the PMFs on small scales which CMB observation s can not access,\nB0/lessorsimilar1.0µGat the scales 104hMpc−1< k <108hMpc−1. Moreover, when the PMF\nspectrum is given in a blue power-law type, the magnetic rehe ating puts a quite strong\nconstraint, for example, B0/lessorsimilar10−17nG,10−23nG, and10−29nGat 1 comoving Mpc\nfornB= 1.0,2.0, and3.0, respectively. This constraint would give an impact on\ngeneration mechanisms of PMFs in the early Universe.\nKey words: cosmology: theory — cosmology: cosmic background radiatio n — cos-\nmology: early Universe\n1 INTRODUCTION\nMagnetic fields are ubiquitous in the universe and they have\nbeen observed in various astrophysical objects from planet s\nand stars to galaxies and galaxy clusters ( Ryu et al. 2012 ;\nWidrow et al. 2012 ). Recently there are some reports to\nsuggest the existence of magnetic fields in the intergalac-\ntic medium ( Neronov & Vovk 2010 ;Tavecchio et al. 2010 ;\nEssey et al. 2011 ;Tashiro et al. 2014 ;Chen et al. 2015 ).\nPrimordial magnetic fields (PMFs) are very attractive as\nthe origin of observed magnetic fields on cosmological\nscales. There are many proposals of the PMF generation\nmechanism in the early universe ( Durrer & Neronov 2013 ;\nSubramanian 2016 ). Therefore, observational constraints on\nPMFs can provide insights into the physics in the early uni-\nverse.\nPMFs can leave their signatures on many cosmologi-\ncal phenomena. As examples, the requirement of success-\nful Big Bang nucleosynthesis (BBN) gives a constraint on\nthe existence of PMFs at that time. The recent detailed\nstudy about the BBN with PMFs provides the upper limit\nB0<1.5µG where B0represents the present strength\nof PMFs ( Kawasaki & Kusakabe 2012 ). The measurements\n⋆shohei.saga@yukawa.kyoto-u.ac.jpof cosmic microwave background (CMB) anisotropies also\nyield constraints on PMFs, because PMFs can create\nthe additional features in the temperature and polar-\nization anisotropies of the CMB ( Subramanian & Barrow\n1998a ;Shaw & Lewis 2010 ). Recent Planck data gave\na limit on PMFs as B0/lessorsimilarO(1)nG around co-\nmoving 1Mpc, depending on the scale dependence of\nPMFs ( Planck Collaboration et al. 2016 ).\nMoreover, dissipation of PMFs in the early universe\nhas also drawn attention to obtain a constraint on PMFs\nat smaller scales, compared with the scales which could be\nconstrained from the CMB temperature and polarization\nanisotropies. Before the epoch of recombination, PMFs in-\nduce the fluid motions in a photon-baryon plasma through\nthe Lorentz force. From the viewpoint of the energy trans-\nfer, this process can be considered as a conversion of the\nmagnetic fields’ energy to the fluid kinetic energy. Since\nthere arises a viscosity in the plasma due to the finite mean\nfree path of photons, the induced motions on small scales\nare damped ( Jedamzik et al. 1998 ;Subramanian & Barrow\n1998b ). As a result, the magnetic fields’ energy on small\nscales would be dissipated into the photon-baryon plasma.\nIf the dissipation happens after the redshift z∼2×106(for\nrecent reviews see Chluba & Sunyaev 2012 ;Tashiro 2014 ),\nthese dissipated energy can be observed as CMB distortions,\nc/circlecopyrt0000 The AuthorsL2S. Saga et al.\nthat is, the spectral deviation of the CMB from the Planck\ndistribution ( Jedamzik et al. 2000 ;Miyamoto et al. 2014 ;\nKunze & Komatsu 2014 ;Ganc & Sloth 2014 ). Depending on\nthe redshift of the energy injections, the generated CMB\ndistortions are typically characterized by two parameters ,\nthe chemical potential µand the yCompton parameter.\nTherefore, the measurement of CMB distortions allows us\nto understand the thermal history of the universe from the\nredshift z∼2×106to the epoch of recombination. The\ncurrent constraint on CMB distortions has been obtained\nby COBE/FIRAS ( Fixsen et al. 1996 ). According to the\nCOBE/FIRAS constraint, the upper limit on the PMFs is\nset toB0<30 nG between comoving 400pc and0.5Mpc\n(Jedamzik et al. 2000 ).\nIn this Letter, we focus on the dissipation of the mag-\nnetic fields’ energy before the CMB distortion era ( z/greaterorsimilarzµ=\n2.0×106). Before the CMB distortion era, any kind of the\nenergy injections does not distort the energy spectrum of\nCMB photons. This is because the interaction processes for\nthe CMB thermalization, i.e.,the Compton scattering, dou-\nble Compton scattering, and bremsstrahlung have smaller\ntime scales than the cosmological time scale. In particular ,\nthe double Compton process keeps the distribution of CMB\nphotons as the Planck distribution by adjusting the number\nof CMB photons, that is, the CMB temperature. As a result,\nareheating of the CMB photons may occur.\nIf we measure the overall history of the absolute CMB\ntemperature or the number density of photons before the\nCMB distortion era, we can directly put a strong con-\nstraint on the amount of the injected energy. However, the\ncosmological observations, e.g., the measurement of CMB\nanisotropies, tell us the information only around the recom -\nbination epoch1. In this Letter, we therefore use the alter-\nnative observable, that is, a baryon-photon ratio η=nb/nγ.\nBefore the CMB distortion era, an energy injection into\nCMB photons can increase the CMB photon number due\nto the double Compton process and bremsstrahlung, while\nthe number of baryons does not change. As a result, the\nbaryon-photon number ratio decreases after the energy in-\njection occurs. Since the light element production in the\nBBN is sensitive to ηat the BBN epoch, the measurement\nof the light element amount can observationally determine\nthe value of η,ηBBN, at the BBN epoch with some errors.\nIndependently, CMB observations can provide the value of\nη,ηCMB, around the epoch of recombination with some er-\nrors. If the energy injection to the CMB photons occurred\nafter the BBN era and before the CMB distortion era, the\ndifference of the baryon-photon number ratios at between\nthe BBN era and CMB era can be reread in terms of the\ninjected energy density of photons ∆ργbetween these eras\nas\nηCMB\nηBBN= 1−3\n4∆ργ\nργ, (1)\nwhere we use the relations nγ∝T3\nγandργ∝T4\nγ.\nCombining these relations, Nakama et al. (2014) pro-\nvides a constraint on the density fraction of the injected\n1As we have mentioned, CMB distortions may tell us the infor-\nmation deeply before the recombination epoch.energy between these epochs as\n∆ργ\nργ<7.71×10−2, (2)\nwhere they adopted ηCMB,obs= (6.11−0.08)×10−10\n(Ade et al. 2014 ) andηBBN,obs= (6.19+0.21)×10−10\n(Nollett & Steigman 2014 ) as observation values for CMB\nand BBN, respectively. Here, we stress that the bound\nEq. (2) is intrinsically coming from the difference between\nthe baryon-photon number ratio at the BBN and CMB era,\nnot from the direct measurement of the CMB tempera-\nture or the CMB photon number. In order to obtain the\nabove constraint, we take into account for the negative and\npositive errors in the CMB and BBN bound, respectively,\nin order to obtain conservative limit. Note that, Eq. ( 2)\ncan be reread in terms of the CMB temperature by us-\ning the relation: ∆ργ/ργ= 4∆Tγ/Tγ+O((∆Tγ/Tγ)2), as\n∆Tγ/Tγ<1.93×10−2.\nThe dissipation of the PMFs can be considered to be\none of interesting heating sources, and we name the heat-\ning mechanism by the dissipation of PMFs magnetic re-\nheating . Before the epoch of recombination, PMFs induce\nthe magnetohydrodynamics (MHD) modes in the photon-\nbaryon plasma and these MHD modes are damped by the\nviscosity due to the diffusion process of CMB photons. In\nthis process, the energy of PMFs dissipates into the CMB\nphotons and subsequently, the baryon-photon number ratio\nchanges. Therefore, we can put a constraint on PMFs from\nEq. (2). In next section, we give the formulation for the\ninjected energy from the dissipation of PMFs and show a\nnew constraint on PMFs obtained from magnetic reheating\nprocess. The final section is devoted to the conclusion and\nsummary of our results.\n2 REHEATING BY DECAYING MAGNETIC\nFIELDS\nLet us consider the spatially-averaged injected energy int o\nthe CMB photons due to the decaying magnetic fields from\nthe BBN era to the CMB distortion era. The total injected\nenergy is represented as\n∆ργ\nργ=/integraldisplayzµ\nzidz/bracketleftbigg\n−1\nργ(z)(1+z)4\n8πd\ndz/angbracketleftbig\n|b(z,x)|2/angbracketrightbig/bracketrightbigg\n, (3)\nwhereργ(z)is the background energy density of photons\nwhich scales as ∝(1+z)4and the brackets denote ensem-\nble average. We also define b(z,x) = (1 + z)−2B(z,x)as\nthe magnetic fields without the adiabatic decay due to the\nexpansion of the universe. Here, we set zµ= 2.0×106and\nzi= 1.0×109, which eras correspond to the inefficient of\nthe double Compton scattering and the neutrino decoupling\nera, respectively ( Hu & Silk 1993 ).\nThe evolution of magnetic fields due to the MHD damp-\ning can be simply expressed in the Fourier space as\nbi(z,k) =˜bi(k)e−(k/kD(z))2, (4)\nwhere˜bi(k)denotes the initial PMFs and kD(z)represents a\nwave number of the magnetic fields which is damped at the\nredshiftz, caused by the photon viscosity, similar to the Silk\ndamping of the standard primordial plasma. The damping\nwave number can be evaluated by the mode analysis based\nMNRAS 000, 000–000 (0000)Magnetic reheating L3\non the magneto-hydrodynamics (MHD). Depending on the\nMHD modes and the scales, the damping wave number\nis different ( Jedamzik et al. 1998 ;Subramanian & Barrow\n1998b ). Here we consider the case for the fast-magnetosonic\nmode, where Alfvén and slow-magnetosonic modes are\ndamped in the photon diffusion limit in which the mode\nwavelength k−1is much larger than the photon mean free\npath. In such a case, the damping scale is similar to the Silk\ndamping scales and we adopt ( Jedamzik et al. 2000 )\nkD(z) = 7.44×10−6(1+z)3/2Mpc−1, (5)\nin the radiation dominated epoch. We will discuss the other\ncases later.\nWe assume that the initial PMFs are statistically ho-\nmogeneous and isotropic Gaussian random fields. The power\nspectrum of such fields can be written as\n∝angbracketleft˜bi(k)˜b∗\nj(k′)∝angbracketright= (2π)3δ3\nD(k−k′)δij−ˆkiˆkj\n22π2\nk3PB(k),(6)\nand then the bracket of the right-hand side in Eq. ( 3) can\nbe expressed as\n/angbracketleftbig\n|b(z,x)|2/angbracketrightbig\n=/integraldisplay\ndlnkPB(k)e−2/parenleftBig\nk\nkD(z)/parenrightBig2\n. (7)\nTherefore, Eq. ( 2) allows us to obtain the constraint on the\npower spectrum of magnetic fields. In our analysis, we em-\nploy two types of the power spectrum, one is a delta-function\ntype in logarithmic scale given by\nPB(lnk) =B2\ndeltaδD(ln(k/kp)), (8)\nwhereBdeltacorresponds to the amplitude at the peak wave\nnumberkp, and another is a power-law type given by\nPB(k) =B2/parenleftbiggk\nkn/parenrightbiggnB+3\n, (9)\nwhereBis the amplitude at the normalization wave number\nkn. In this Letter, we set the normalization scale as kn=\n1 Mpc−1.\nFirst we focus on the delta function type for the power\nspectrum. By substituting Eq. ( 8) into Eq. ( 3) with Eq. ( 7),\nwe obtain the injected energy fraction to the CMB energy\nby the decaying PMFs as\n∆ργ\nργ=B2\ndelta\n8πργ,0˜C(kp), (10)\nwhere˜C(kp)is the function of kpand which the explicit\nform is given as\n˜C(k)≡exp/bracketleftbigg\n−2k2\nk2\nD(zi)/bracketrightbigg\n−exp/bracketleftbigg\n−2k2\nk2\nD(zµ)/bracketrightbigg\n. (11)\nWhenkpis much larger than kD(zi)(> kD(zµ)), the energy\ninjection due to the decay of the PMFs becomes efficient\nbefore the BBN era. On the other hand, when kpis much\nsmaller than kD(zµ), the energy injection from the PMFs\noccurs during the CMB distortion era. We focus on the case\nwhere the energy injection from the decaying of the PMFs is\nefficient after the BBN era and before the CMB distortion\nera, that is, the case where kpis set between the two decay-\ning scales kD(zi)andkD(zµ). For such a case, we can approx-\nimately take ˜C(kp)→1(seeMiyamoto et al. 2014 ). Com-\nbining Eq. ( 2) and Eq. ( 10) yields a constraint on PMFs in\nthekp-Bdeltaplane. We present our new result obtained from101102103104\n102103104105106107108109Bdelta [nG]\nkp [Mpc-1]Magnetic reheating\nBBN\nCMB distortion\nFigure 1. The upper bound on the amplitude of the delta-\nfunction type for the power spectrum Bdelta as a func-\ntion of kpfrom the magnetic reheating. The limits from\nthe BBN ( Kawasaki & Kusakabe 2012 ) and CMB distortions\n(Jedamzik et al. 2000 ) in blue and magenta lines, respectively,\nare shown in the same plot.\nmagnetic reheating as a red line in Fig. 1. For comparison, we\nplot limits from the BBN and CMB distortions in magenta\nand blue lines, respectively. We can see that our magnetic\nreheating constraint gives a tight limit on the PMFs on small\nsales from kp= 104Mpc−1tokp= 108Mpc−1.\nNext we consider the power-law type of the power spec-\ntrum defined in Eq. ( 9). By using Eq. ( 3) with Eq. ( 7), we\nobtain\n∆ργ\nργ=B2\n8πργ,0Γ/parenleftbignB+3\n2/parenrightbig\n2(nB+5)/2/bracketleftBigg/parenleftbiggkD(zi)\nkn/parenrightbiggnB+3\n−/parenleftbiggkD(zµ)\nkn/parenrightbiggnB+3/bracketrightBigg\n.\n(12)\nHere, we focus on the magnetic reheating due to the fast-\nmagnetosonic mode given in Eq. ( 5). In Fig. 2, we plot a\nlimit for the amplitude of the power-law type as a function\nof the spectral tilt nBin a red line. We find that blue-tilted\nspectrum is strongly constrained compared with the scale-\ninvariant spectrum ( nB=−3.0). It is shown that the causal\nmechanism of PMFs generation predicts only a blue spec-\ntrum with nB/greaterorequalslant2(Durrer & Caprini 2003 ). Therefore, the\nconstraint on the blue-tilted spectrum is suggestive for su ch\na causal mechanism.\nSo far we have considered the case that the Alfvén\nand slow-magnetosonic modes are damped in the photon\ndiffusion limit. However, depending on the magnetic field\nstrength, the energy damping of these modes are insuffi-\ncient in the photon diffusion limit because the magnetic\nfield cannot accelerate the fluid efficiently. When it hap-\npens, the magnetic field can survive even below smaller scale\nthank−1\nD(z)given in Eq. ( 5). These modes are damped\nlater when the scale of the modes are smaller than the\nphoton free-streaming scale. This damping scale is roughly\ngiven by kA\nD(z)∼kD(z)/VA(z)in the radiation domi-\nnated epoch. Here VA(z)is the Alfvén velocity, VA(z) =\nBλ(z)//radicalbig\n16πργ,0/3, where we obtain the magnetic field\nBλ(z)by integrating the power spectrum over kwith a\nGaussian window function with a scale λ(z)corresponding\nMNRAS 000, 000–000 (0000)L4S. Saga et al.\n10-3010-2510-2010-1510-1010-5100105\n-3.0-2.0-1.00.01.02.03.0B [nG]\nnBFast mode\nAlfven mode\nPlanck constraint\nFigure 2. The upper bound on the amplitude of the power-\nlaw type for the power spectrum Bas the function of nBfrom\nthe magnetic reheating of the fast-magnetosonic mode (red)\nand Alfvén-magnetosonic mode (blue). The constraint from t he\nPlanck ( Planck Collaboration et al. 2016 ) is also shown. Note\nthatnB=−3.0corresponds to the scale-invariant power spec-\ntrum.\ntokA\nD(z)(Mack et al. 2002 ). Therefore, kA\nD(z)can be repre-\nsented as\nkA\nD(z)\nkn∼2π/bracketleftBigg\n3.5×105\nΓ/parenleftbignB+5\n2/parenrightbig/parenleftbiggB\n1 nG/parenrightbigg−2/parenleftbiggkD(z)\nkn/parenrightbigg2/bracketrightBigg1/(nB+5)\n.\n(13)\nReplacing kD(z)tokA\nD(z)in Eq. ( 12), we can evaluate the in-\njected energy fraction for the Alfvén and slow-magnetosoni c\nmodes in the photon free-streaming limit and plot the con-\nstraint on primordial magnetic fields as a blue line in Fig. 2.\nWe find that the magnetic reheating due to the Alfvén and\nslow-magnetosonic modes put a stronger constraint on the\namplitude of PMFs than the fast-magnetosonic mode. The\nobtained constraint can be roughly fitted as log(B/nG)/lessorsimilar\n−11−6nB(fornB/greaterorsimilar0). This constraint is tighter than\nother CMB constraints.\n3 CONCLUSION\nBefore the CMB distortion era, the additional energy injec-\ntion heats the CMB photons and due to the double Comp-\nton scattering, the number of the CMB photons increases.\nTherefore, the baryon-photon number ratio decreases by the\nreheating. By comparing the values of the baryon-photon\nnumber ratio between two epochs, we can determine the\nallowed energy injection or allowed changes of the CMB\ntemperature between them. In this Letter, we assume the\nreheating source as the decaying PMFs due to the photon\nviscosity, named the magnetic reheating. Since the baryon-\nphoton number ratio is well constrained by the current cos-\nmological observations such as the BBN and CMB, the mag-\nnetic reheating provides us a significant constraint on smal l-\nscale PMFs which we cannot access through the observa-\ntions of CMB anisotropies and distortions. From the state-\nof-the-art observations, we put the new constraint on theamplitude of small-scale PMFs as B0/lessorsimilar1µGin the range\n104hMpc−1< k <108hMpc−1.\nWe also apply the magnetic reheating constraint to\nPMFs with a power-law spectrum. Since the dissipation\nscale depends on the MHD modes, we consider two damping\nscales of not only the fast-magnetosonic mode but also the\nslow-magnetosonic and Alfvén modes. We find that for blue-\ntiled spectrum the slow-magnetosonic and Alfvén modes\ncan set stronger constraint, compared with the constraint\nfrom the fast-magnetosonic mode. In general, because the\ndamping scale of the slow-magnetosonic mode and Alfvén\nmode is smaller than that of the fast-magnetosonic mode\nthe slow-magnetosonic and Alfvén modes are sensitive to\nthe PMFs on small scales and can set a stronger constraint\nfor blue-tiled spectrum rather than the fast-magnetosonic\nmode. The obtained constraint can be roughly fitted as\nlog(B/nG)/lessorsimilar−11−6nB(fornB/greaterorsimilar0) where Bis the PMF\namplitude normalized at 1 comoving scale. This constraint\nis tighter than other CMB constraints in the case of the\nblue-tilted spectrum.\nAlthough we consider two damping scales separately in\nthis Letter, in reality, the energy dissipation of magnetic\nfields happens through the combination of these two damp-\ning scale channels. However it is not so trivial how much\npercentage of the magnetic field energy is dissipated throug h\nthe fast-, slow-magnetosonic or Alfvén modes. When the dis-\nsipated energy is in equipartition between these modes, our\nconstraint might be relaxed by a few factor. To investigate\nthe details, we will address this issues through the analysi s\nin higher order cosmological perturbations.\nAmong many mechanisms to generate PMFs in the\nearly universe, the causal mechanism can generate only blue\npower spectrum ( Durrer & Caprini 2003 ). Therefore, strong\nconstraints for the blue-tiled PMFs obtained in this Letter\nshould be a powerful tool to investigate a successful causal\nmechanism to generate PMFs in the early universe.\nACKNOWLEDGEMENTS\nThis work is supported in part by a Grant-in-Aid for JSPS\nResearch Fellow Number 17J10553 (S.S.), JSPS KAKENHI\nGrant Number 15K17646 (H.T.), 17H01110 (H.T.) and\n15K17659 (S.Y.), and MEXT KAKENHI Grant Number\n16H01103 (S.Y.).\nREFERENCES\nAde P. A. R., et al., 2014, Astron. Astrophys. , 571, A16\nChen W., Buckley J. H., Ferrer F., 2015, Phys. Rev. Lett. , 115,\n211103\nChluba J., Sunyaev R. A., 2012, MNRAS ,419, 1294\nDurrer R., Caprini C., 2003, J. Cosmology Astropart. Phys. ,\n11, 010\nDurrer R., Neronov A., 2013, Astron. Astrophys. Rev. , 21, 62\nEssey W., Ando S., Kusenko A., 2011, Astropart. Phys. , 35, 135\nFixsen D. J., Cheng E. S., Gales J. M., Mather J. C., Shafer R. A .,\nWright E. L., 1996, Astrophys. J. , 473, 576\nGanc J., Sloth M. S., 2014, JCAP , 1408, 018\nHu W., Silk J., 1993, Phys. Rev. , D48, 485\nJedamzik K., Katalinic V., Olinto A. V., 1998, Phys. Rev. , D57,\n3264\nMNRAS 000, 000–000 (0000)Magnetic reheating L5\nJedamzik K., Katalinic V., Olinto A. V., 2000, Phys. Rev. Lett. ,\n85, 700\nKawasaki M., Kusakabe M., 2012, Phys. Rev. D ,86, 063003\nKunze K. E., Komatsu E., 2014, J. Cosmology Astropart. Phys. ,\n1, 009\nMack A., Kahniashvili T., Kosowsky A., 2002, Phys. Rev. D ,\n65, 123004\nMiyamoto K., Sekiguchi T., Tashiro H., Yokoyama S., 2014,\nPhys. Rev. , D89, 063508\nNakama T., Suyama T., Yokoyama J., 2014, Phys. Rev. Lett. ,\n113, 061302\nNeronov A., Vovk I., 2010, Science , 328, 73\nNollett K. M., Steigman G., 2014, Phys. Rev. , D89, 083508\nPlanck Collaboration et al., 2016, A&A,594, A19\nRyu D., Schleicher D. R. G., Treumann R. A., Tsagas C. G.,\nWidrow L. M., 2012, Space Sci. Rev. ,166, 1\nShaw J. R., Lewis A., 2010, Phys. Rev. D ,81, 043517\nSubramanian K., 2016, Rept. Prog. Phys. , 79, 076901\nSubramanian K., Barrow J. D., 1998a, Physical Review Letters ,\n81, 3575\nSubramanian K., Barrow J. D., 1998b, Phys. Rev. , D58, 083502\nTashiro H., 2014, PTEP , 2014, 06B107\nTashiro H., Chen W., Ferrer F., Vachaspati T., 2014,\nMon. Not. Roy. Astron. Soc. , 445, L41\nTavecchio F., Ghisellini G., Foschini L., Bonnoli G., Ghirl anda\nG., Coppi P., 2010, Mon. Not. Roy. Astron. Soc. , 406, L70\nWidrow L. M., Ryu D., Schleicher D. R. G., Subramanian K.,\nTsagas C. G., Treumann R. A., 2012, Space Sci. Rev. ,166, 37\nMNRAS 000, 000–000 (0000)" }, { "title": "1505.07704v1.Damping_factors_for_head_tail_modes_at_strong_space_charge.pdf", "content": "DAMPING FACTORS FOR HEAD-TAIL MODES AT STRONG SPACE CHARGE Alexey Burov, Fermilab, Batavia, IL 60510 Abstract This paper suggests how feedback and Landau damping can be taken into account for transverse oscillations of bunched beam at strong space charge. MAIN EQUATION Space charge is known to be able to change dramatically collective modes [1-3]. For transverse oscillations of bunched beams, a parameter of the space charge strength is a ratio of the maximal space charge tune shift to the synchrotron tune. When this parameter is large, the transverse oscillations are described by a one-dimensional integro-differential equation derived in Ref. [2] and reproduced here for the sake of convenience: νy+1Qeffddτu2dydτ⎛⎝⎜⎜⎜⎞⎠⎟⎟⎟⎟=κNˆWy+Dy()ˆWy=W(τ−s)expiζτ−s()⎡⎣⎢⎤⎦⎥−∞∞∫ρ(s)y(s)dsDy=y(τ)D(τ−s)−∞∞∫ρ(s)dsdydττ→±∞=0 (1) Here y=y(τ) and ν are the eigenfunction and the eigenvalue to be found for the bunch transverse oscillations, τ and s are longitudinal positions within the bunch, W and D are the dipole and quadrupole (or the driving and detuning) wakes, ρ is the normalized line density ρds=1∫ , (2) N is the number of particles per bunch, κ=r0R4πβ2γQb , (3) with r0 as the classical radius, R as the average machine radius, β and γ as the relativistic factors and Qb as the bare betatron tune. The parameter ζ staying in the exponents of the wake integral is a negated ratio of the chromaticity ξ=pdQb/dp and the slippage factor η=γt−2−γ−2, i.e. ζ=−ξ/η. The symbol Qeff=Qeff(τ) stays for the space charge tune shift at the given position along the bunch τ, averaged over the both transverse action, see Ref. [2]. Thus, the effective space charge tune shift is proportional to the local line density: Qeff(τ)=Qeff(0)ρ(τ)/ρ(0). (4) For the transversely Gaussian bunch, Qeff(τ)=0.52Qmax(τ) (5) where Qmax(τ) is the space charge tune shift at the bunch axis. The symbol u2=u2(τ) stays for the average square of the particle longitudinal velocity υi=dτi/dθ, with time measured as the angle θ along the machine circumference, at the given position τ : u2=υ2=f(υ,τ)υ2dυ∫f(υ,τ)dυ∫ , (6) where f(υ,τ) is the longitudinal distribution function. For the longitudinally Gaussian distribution with the rms bunch length στ, the temperature function u2 is constant along the bunch: u2=Qs2στ2 , (7) where Qs is the synchrotron tune. In general, the wake term ˆWy is a sum of single-bunch (SB), coupled-bunch (CB) wakes and the damper (G) terms: ˆWy=ˆWSBy+ˆWCBy+ˆGy (8) The single-bunch term ˆWSBy is described exactly as in Eq. (1), where only s>τcontributes due to the causality, and the integral is taken along the single-bunch interval only: ˆWSBy=W(τ−s)expiζτ−s()⎡⎣⎢⎤⎦⎥SB∫ρ(s)y(s)ds (9) The coupled-bunch term results from summations of the fields left by preceding passages of the bunches through the given position of the ring. This summation is especially simple when the bunches are equidistant. In this case, due to the symmetry, the offsets of the neighbor bunches, being taken at the same time, differ only by the phase factors exp(iψµ): y(s+s0)=exp(iψµ)y(s). (10) For M bunches in the ring, ψµ=2πµM;µ=0,1,...,M−1 , (11) where integer µ is a counter of the coupled-bunch modes. After that, we have to take into account that the given reference bunch sees the fields left behind by other bunches not at the same time, but certain time ago, proportional to the distance between the bunches. This leads to an additional time-related factor to be taken into account together with the space-related phase factor: y(s+s0,θ−s0)=expiφµ()y(s,θ);φµ=ψµ+2πQbM. (12) Remember that both time and space are measured as the angles of revolution, and the leading particles have higher coordinate than the following ones. From here, the coupled-bunch contribution in Eq. (8) follows as a single-bunch integral: ˆWCBy=!Wµ(τ,s)expiζτ−s()⎡⎣⎢⎤⎦⎥SB∫ρ(s)y(s)ds;!Wµ(τ,s)=W(τ−s−ks0)exp(ikφµ)k=1∞∑. (13) For those cases when the wake function of the neighbour bunch does not change much along the reference bunch, i.e. the coupled-bunch wake is flat [4], W(τ−s−ks0)≈W(−ks0) (14) the result of summation in the right hand side of Eq. (13) does not depend on the specific positions s, τ within the bunches; thus, the effective coupled bunch wake !Wµ is a constant which can be taken out of the integral: ˆWCBy=!Wµeiζτe−iζsρ(s)y(s)dsSB∫. (15) In principle, the damper term ˆGy in Eq. (8) is similar to the coupled-bunch one. If the feedback bandwidth is much smaller than the inverse bunch length, the damper takes just one parameter per bunch. This parameter can be chosen as an offset of the centre of mass, and the kick can be designed to be flat along the bunch. Then the damper term is represented similar to Eq. (15): ˆGy=!Gµeiζτe−iζsρ(s)y(s)dsSB∫ , (16) If the damper is bunch-by-bunch, there is no coupled-bunch mode dependence in the feedback factor !Gµ, i.e. !Gµ=!G. Similarly to the driving wake factor, Eq. (8), there is a certain coupled-bunch contribution in the detuning wake as well: Dy=DSBy+DCBy;DSB=ρ(s)D(τ−s)dsSB∫;DCB=ρ(s)!D(τ,s)dsSB∫;!D(τ,s)=D(τ−s−ks0)k=1∞∑. (17) If the detuning wake is flat in the same sense as in Eq. (14), its coupled-bunch contribution is identical for all the particles, so it works as a constant quadrupole without any influence to the beam dynamics unless it leads to a dangerous resonance crossing. SOLUTION To find the spectrum of Eq. (1), its eigenfunctions can be expanded over its zero-wake solutions yk0(τ) satisfying the following equation: ν0y0+1Qeffddτu2dy0dτ⎛⎝⎜⎜⎜⎜⎞⎠⎟⎟⎟⎟=0;dy0dττ→±∞=0 (18) All the eigenfunctions are orthogonal and can be normalized, so that: dsρ(s)yk0(s)SB∫ym0(s)=δkm. (19) For the Gaussian distribution, the spectrum of this equation has been described in Ref.[2,3]; similarly, it can be found for any potential well and distribution function. Expansion of the eigenfunction y(τ) over the no-wake set y0(τ), y(τ)=Bkyk0(τ)k=0∞∑ (20) with the amplitudes B to be found, with the following multiplication of Eq.(1) on yl0(τ)ρ(τ) and its integration over the bunch length, leads to: κNˆW+κNˆD+Diag(ν0)⎡⎣⎢⎤⎦⎥B=νB . (21) Here ˆW and ˆD are the matrices of the driving and detuning wakes in the basis of the no-wake modes of Eq. (18), (19): yk0(τ) ˆWSB()lk=W(τ−s)ρ(τ)ρ(s)eiζ(τ−s)yl0(τ)SB∫∫yk0(s)dτds;ˆWCB()lk=!Wµ(τ,s)ρ(τ)ρ(s)eiζ(τ−s)yl0(τ)SB∫∫yk0(s)dτds; (22) ˆGlk=!GµIl(ζ)Ik*(ζ);Il(ζ)=ρ(τ)eiζτyl0(τ)SB∫dτ; (23) ˆDSB()lk=FSB(τ)SB∫ρ(τ)yl0(τ)yk0(τ)dτ;FSB(τ)=D(τ−s)SB∫ρ(s)ds; (24) ˆDCB()lk=FCB(τ)SB∫ρ(τ)yl0(τ)yk0(τ)dτ;FCB(τ)=!D(τ,s)SB∫ρ(s)ds. (25) DAMPER DETAILS In case when the feedback takes something different from the centre of mass and its kick is not flat over the bunch, the damper matrix has to be modified with provided pickup and kicker functions P(s),K(τ) : ˆGlk=!GµKl(ζ)Pk*(ζ);Kl(ζ)=K(τ)ρ(τ)eiζτyl0(τ)SB∫dτ;Pk*(ζ)=P(s)ρ(s)e−iζsyl0(s)SB∫ds. (26) Equation (21) is a standard linear algebra eigensystem problem which solution is straightforward as soon as the wake functions, the feedback properties, the potential well, and the beam distribution functions, longitudinal and transverse, are given. This equation allows computing the instability growth rates for fairly general situations when the Landau damping can be neglected. However, without Landau damping nothing can be said about the instability threshold, so the theory is significantly incomplete. INSTABILITY THRESHOLDS For the strong space charge, Landau damping rates were roughly estimated in Ref. [2]. Numerical simulations give a possibility for more accurate knowledge of the damping rates, with the numerical factors to be found with a good precision. This work has been started by V. Kornilov and O. Boine-Frankenheim several years ago with their PATRIC code [5.6], has been joined recently by A. Macridin et al. with the Synergia program [7]. As soon as the Landau damping rates are reliably established, they can be introduced in Eq. (21) as externally given imaginary parts of the vector ν0 in Eq. (21) with the following substitution: νk0→νk0−iλk , (27) where λk is Landau damping rate for the no-wake eigenfunction yk0(τ). When the collective tune shifts imposed by the wakes are small compared to the synchrotron tune, their influence on the Landau damping can be neglected. Potential importance of the image charges and currents for Landau damping was shown in Ref. [5,6]. As soon as Landau damping rates are included in Eq. (21), theory of transverse stability of bunched beams with strong space charge would be complete. Hopefully, this work will be done in a reasonable future. FNAL is operated by Fermi Research Alliance, LLC under Contract No. De–AC02–07CH11359 with the United States Department of Energy. REFERENCES [1] M. Blaskiewicz, Phys. Rev. ST Accel. Beams 1, 044201 (1998); Phys. Rev. ST Accel. Beams 6, 014203 (2003). [2] A. Burov, Phys. Rev. ST Accel. Beams 12, 044202 (2009). [3] V. Balbekov, Phys. Rev. ST Accel. Beams 12, 124402 (2009) [4] A. Burov, Phys. Rev. ST Accel. Beams 17, 021007 (2014) [5] V. Kornilov and O. Boine-Frankenheim, Phys. Rev. ST Accel. Beams 13, 114201 (2010) [6] V. Kornilov, O. Boine-Frankenheim, D.J. Adams, B. Jones, B.G. Pine, C.M. Warsop, R.E. Williamson, “Thresholds of the Head-Tail Instability in Bunches with Space Charge”, HB’2014 workshop, East Lansing, MI, USA. [7] A. Macridin, A. Burov, E. Stern, J. Amundson, and P. Spentzouris, “Simulation of transverse modes with intrinsic Landau damping for bunched beams with space charge”, to be published. " }, { "title": "1709.01425v1.Enhancement_of_space_charge_induced_damping_due_to_reactive_impedances_for_head_tail_modes.pdf", "content": "arXiv:1709.01425v1 [physics.acc-ph] 5 Sep 2017Enhancement of space-charge induced damping due to reactiv e\nimpedances for head-tail modes\nV. Kornilov,\nGSI Helmholtzzentrum, Planckstr.1, Darmstadt, Germany,\nO. Boine-Frankenheim,\nGSI Helmholtzzentrum, Planckstr.1, Darmstadt, Germany,\nTU Darmstadt, Schlossgartenstr.8, Darmstadt, Germany\nOctober 9, 2018\nAbstract\nLandau damping of head-tail modes in bunches due to spreads in the tune shift can\nbe a deciding factor for beam stability. We demonstrate that the co herent tune shifts\ndue to reactive impedances can enhance the space-charge induce d damping and change\nthe stability thresholds (here, a reactive impedance implies the imagin ary part of the\nimpedance of both signs). For example, high damping rates at stron g space-charge, or\ndamping of the k= 0 mode, can be possible. It is shown and explained, how the negative\nreactive impedances (causing negative coherent tune shifts similar ly to the effect of space-\ncharge) can enhance the Landau damping, while the positive cohere nt tune shifts have\nan opposite effect. It is shown that the damping rate is a function of the coherent mode\nposition in the incoherent spectrum, in accordancewith the concep t of the interaction of a\ncollective mode with resonant particles. We present an analytical mo del, which allows for\nquantitative predictions of damping thresholds for different head- tail modes, for arbitrary\nspace-charge and coherent tune-shift conditions, as it is verified using particle tracking\nsimulations.\n1 INTRODUCTION\nThe performance of many present and future hadron ring accel erators is limited by the effects\ndue to the self-field space-charge. The transverse head-tai l modes in bunches have been\ninitially described by a theory [1, 2] without space-charge . A significant progress have been\nrecently achieved in understanding of the effects of the betat ron tune shifts due to space-\ncharge for the collective eigenmodes in bunches. Similarly to other beam dynamics issues,\nthe spread in the space-charge tune shift appeared to be very important.\nAn ”airbag” bunch model, suggested in [3], allowed for an exa ct description of the head-\ntail modes for arbitrary space-charge. This model does not t ake into account a spread in the\nspace-charge tune shift which is present in realistic bunch distributions. Nevertheless, the\npredictions of the airbag model appeared to be rather accura te for the eigenfrequencies and\n1for the eigenmodes in the realistic Gaussian bunches [4, 5]. Experimental confirmations have\nbeen demonstrated in measurements on bunches with space-ch arge in [6, 7]. Theories for\nrealistic bunch distributions have been presented in [8, 9] , where an important implication of\nthe space-charge tune spread has been identified. This is a La ndau damping of the coherent\nmodes due to interactions with the incoherent particle spec trum. In contrast to the case\nof a coasting beam, the space-charge induced tune spread pro vides damping for the bunch\neigenmodes. Damping rates due to the tune spread produced by the longitudinal bunch\ndensity variations have been demonstrated and quantified in [5].\nAnimaginaryimpedanceoftheaccelerator facility represe ntsareactiveforceandshiftsthe\ncoherent frequencies [10, 11]. In real accelerators there a re many sources of the coherent and\nincoherent tune shifts. The detuning wakes can shift the inc oherent betatron frequencies [12].\nIn this work we focus on the driving dipole wakes and on their c oherent betatron frequency\nshifts. We thus specify ∆ Qcohas the real tune shift of a coherent eigenmode due to a reactiv e\nfacility impedance. Typically, the real coherent tune shif t is negative, which is also the case\nfor the self-field space-charge tune shift ∆ Qsc, so for the both notations we use the modulus\nof the negative value. As the both shifts depress the tunes, i t makes the intersections of the\ncoherent frequencies with the incoherent spectrum possibl e. This is the main reason for the\neffect of the reactive impedances on the space-charge induced damping in bunches which we\naddress in the present paper. The combination of space-char ge with reactive impedances has\nbeen discussed in [13, 7], and the enhancement effect of the rea ctive impedances on damping\nhas been suggested in [14, 15]. In this work we present an anal ytical model and simulations\nfor a systematic description of the space-charge induced da mping with reactive impedances.\nIn Sec.2 we discuss the nature of the space-charge induced da mping and how it can\ndepend on the beam distributions, especially regarding the distribution tails. The effect of\nreactive impedancesisexplainedinSec.3, withtheanalyti cal modelforarbitraryspace-charge\nconditions and mode index number. The predictions of the ana lytical model are analysed for\ndifferent cases and verified in Sec.4. The work is concluded in S ec.5.\n2 SPACE-CHARGE INDUCED DAMPING\nTherearetwobasicsourcesfortheself-fieldspace-charge t unespreadinabunch. Thefirstone\nis due to different synchrotron amplitudes of the individual p articles, i.e. in the longitudinal\nplane. The second one is due to different betatron amplitudes, i.e. in the transverse plane.\nIn the both cases, a non-uniform density distribution cause s a tune spread.\nThe line density λ(z) variation along the bunch gives the maximal space-charge t une shift\nfor the given longitudinal position,\n∆Qsc(z) =g⊥λ(z)rpR\n4γ3β2εx, (1)\nwhere(2πR)istheringcircumference, βandγaretherelativisticparameters, rp=q2\nion/(4πǫ0mc2)\nis the classical particle radius, εxis the transverse rms emittance. This tune shift is the mod-\nulus of the negative shift. The geometry factor g⊥depends on the transverse distribution, for\n2the Gaussian profile it is g⊥= 2, for the K-V beam it is g⊥= 1. In order to characterize the\nspace-charge force in a bunch, the space-charge parameter i s used,\nq=∆Qsc(0)\nQs(2)\nwhich is the tune shift for the rms-eqivalent K-V beam ( g⊥= 1) in the peak of the line\ndensity ( λ0), normalised by the synchrotron tune Qs. It was shown in [3, 8, 9, 5, 6] that\nthis parameter is the key value to describe the space-charge conditions in a bunch. This is\nespecially true for the analytical models, as it will be also shown below. The usage of the\nrms-eqivalent K-V beam for the comparisons between different beam distributions has been\nproven to be adequate, the same is true for comparisons with t he experiments. During the\nfurther discussion we use the notation ∆ Qscfor ∆Qsc=qQsfrom Eq.(2).\nAnother reason for the space-charge tune spread is the differe nt amplitudes of the trans-\nverse betatron oscillation of theindividual particles. Th eparticles with small amplitudes have\nthe tune shift close to the local maximal space-charge tune s hift from Eq.(1). The particles\nwith large amplitudes have smaller tune shifts, technicall y down to zero, depending on the\ntransverse distribution and on its truncation. A beam with t he transverse K-V distribution\nhas no space-charge tune spread of this kind.\nDamping of the head-tail modes exclusively due to the space- charge tune spread from\nthe longitudinal plane has been studied in [5]. There, a K-V d istribution has been assumed\nfor the transverse plane. A good agreement with the results f rom [5] has been recently\nreported in [16], in simulations with a different code and differ ent methods. One might\nexpect that an additional tune spread due to the transverse p rofile should enhance damping.\nThis is confirmed in our particle tracking simulations prese nted in Fig.1, where the space-\ncharge induced damping for the Gaussian bunch in the longitu dinal and transverse planes\nis compared with the results from [5]. Additional space-cha rge tune spread provides higher\ndamping rates and larger q−regions for a chosen damping level.\n 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45\n 0 1 2 3 4 5 6 7103 (−Im(∆Q))\nq = ∆Qsc / QsK-V Gaussiank=1\n 0 0.2 0.4 0.6 0.8 1 1.2 1.4\n 0 2 4 6 8 10 12 14103 (−Im(∆Q))\nq = ∆Qsc / QsK-V Gaussiank=2\nFigure1: Space-chargeinduceddampingratesfromparticle trackingsimulationsfor k= 1and\nk= 2 head-tail modes. Solid lines correspond to the longitudi nal Gaussian and transverse K-\nV bunches, as presentedin [5]. Dashedlines arefortheGauss ian (longitudinal andtransverse)\nbunches.\n3For the particle tracking simulations we use the code PATRIC [17], similarly to the simu-\nlations presented in [5]. The code has been validated using e xact analytical prediction [4], for\nthe space-charge effects and for the coherent effects. For the tr ansverse space-charge force,\na frozen slice-to-slice electric field model is used, i.e. a fi xed potential configuration which\nfollows the center of mass for each bunch slice. This approac h is justified in the rigid-slice\nregime [18, 8]. Thus, it is physically reasonable for modera te and strong space charge. The\nrecent study [15] suggests to confirm the applicability of th e frozen space-charge model. The\nmethod to excite the head-tail eigenmodes and to determine t he damping decrements is de-\nscribed in [5]. The simulations for Fig.1 have been performe d for the chromaticity equal zero.\nThe chromaticity does not affect damping, which is also clear i n the formalism of [8], and has\nbeen verified in our simulations.\n3 ANALYTICAL MODEL\nThe basic mechanism of Landau damping is the interaction bet ween a collective wave and the\nresonant particles. In the case of the head-tail modes in bun ches these roles are played by a\ncoherent buncheigenmodeandtheindividualparticlebetat ronoscillations. Thedampingrate\nis then proportional to the number of the resonant particles . It can be especially clear in the\nsimulations for different longitudinal and transverse distr ibutions in a bunch. Allowing more\nparticlesinthedistributiontails, bothlongitudinaland transverse, providesstrongerdamping.\nIn the experimental conditions, the bunch distributions ca n be very complicated, the tail\ntruncations depend on the transverse acceptance and on the m omentum acceptance of the\nmachine. It should be however possible to find general finding s for different distributions and\ntail truncations. In our simulations we consider the exampl e with the Gaussian distribution,\nand with the truncations at 2.5 of the standard deviation.\nA certain damping rate can be reached if the coherent oscilla tion interacts with a large\nenough number of resonant particles. The particle incohere nt spectrum is provided by tune\nshifts of different nature. In the case of space-charge in a Gau ssian bunch, this distribution\nis non-zero between ( −2qQs)<∆Q <0. The particles with tune shifts just below zero\n(∆Q <0), are in the distribution tails, which correspond to parti cles with large amplitudes.\nA growing beam intensity with a fixed beam distribution means a growing space-charge tune\nshift. If we consider resonant particles with the fixed oscil lation amplitudes, longitudinal\nand transverse, we see that the incoherent shift of these par ticles increases linearly with the\nintensity parameter,\n∆Qη=−η∆Qsc. (3)\nHere, the parameter ηis introduced in order to relate the coherent lines to the inc oherent\nspectrum in the upcoming discussion, and to indicate the fac t that space-charge tune shifts\ngrow linearly with the beam intensity (for a fixed emittance) . The value of ηin a specific\nbunch depends on the chosen damping rate and on the particle d istribution in the bunch.\nThe coherent tune shift dueto space-charge can be analytica lly estimated using the airbag\n4model [3],\n∆Qk\nQs=−q\n2±/radicalbigg\nq2\n4+k2, (4)\nwhere ”+” is for modes k≥0. As we discuss in the Introduction, this expression provid es a\nrather accurate prediction for the eigenfrequencies even i n the Gaussian bunches, especially\nfor strong space-charge. This has been confirmed in simulati ons, experiments, and theoretical\nstudies [8, 9, 5, 6].\nThe head-tail modes drop in frequency for stronger space-ch arge. Thus the coherent mode\ncan cross the border of the damping with a chosen damping rate . Since the head-tail modes\nwithk >0 are close to the upper tail of the incoherent spectrum (clos e to the bare tune\nQ0), there is a top border for the chosen damping. If the coheren t mode has a frequency\nbelow this border the damping rate will exceed the border dec rement. The conditions for a\nchosen damping rate should be estimated by the relation of th e border Eq.(3) to the coherent\nk−mode frequency which is modulated by the harmonics of the syn chrotron frequency [9, 5],\n∆Qk−kQs, (5)\nwhere ∆Qk=Qk−Q0is the coherent frequency of the mode k. The coherent line shifts are\npositive (∆ Qk>0) fork >0, and the incoherent tune shifts due to space-charge are neg ative.\nThis damping model thus also explains why damping is still po ssible.\nFigure2 illustrates the above discussion and shows the cohe rent head-tail frequencies\nmodulated by the harmonics of the synchrotron frequency for the modes k= 1 and k= 2.\nFrom the simulations with a Gaussian bunch, see Fig.1, we hav e determined η= 0.24, which\nis shown by the dashed line in Fig.2.\n-2-1.5-1-0.5 0\n 0 2 4 6 8 10Re(∆Qk)/Qs − k\nq = ∆Qsc / Qsk=0\nk=1\nk=2\nFigure 2: Positioning of the head-tail modes from Eqs.(4, 5) with respect to the incoherent\nspectrum border for the damping analysis. The dashed line in dicates the border Eq.(3) of\nthe chosen damping η= 0.24.\nThereis avalueofthe q−parameter wherethemodecrosses thedampingborder. Itcan b e\npredicted using the expression for the coherent head-tail e igenfrequency Eq.(4). Comparing\n5(∆Qk−kQs) with the damping border Eq.(3), we come to the conclusion th at a mode crosses\nthe border at q= 0, and at\nq=k1−2η\nη(1−η). (6)\nThis gives a prediction for the range of the space-charge par ameterqfor a chosen damping\nrate for arbitrary mode index. Figure2 illustrates this: fo r very small qeach head-tail mode\nenters the incoherent particle spectrum and thus experienc es a strong damping (see Fig.1).\nAs the coherent line crosses the damping border again, the da mping becomes weaker due to\na smaller number of the resonant particles. There is a smooth transition between damping\nregimes as the number of resonant particles changes. Estima tions for qwhere the modes cross\nthe damping borders Eq.(6) provide q≈2.8 fork= 1 and q≈5.7 fork= 2 (here η= 0.24),\nas we can also observe in Fig.2. All these observations are su pported by particle tracking\nresults for Gaussian bunches in Fig.1.\nThe damping rates decrease at strong space-charge ( q > k). For Gaussian bunches, the\nscaling Im(∆ Q)∼ −1/q3has been predicted in [12]. The drop of the damping rate for\nhigherqis in agreement with the concept of the coherent frequency sh ift Eq.(5) towards\nlower density in the incoherent spectrum. The fast decrease of the damping rate reflects the\nfact that the coherent lines are shifted further into the tai ls of the incoherent distribution\n. For realistic bunches, the specific scaling depends on the b unch distributions and the tail\ntruncations. Below we should demonstrate that the effect of re active impedances can lead to\nhigh damping rates also for strong space-charge.\n4 DAMPINGWITH REACTIVEIMPEDANCES: ANALYT-\nICAL MODEL\nThe additional effect of a coherent tune shift has been include d into the airbag theory in [13],\n∆Qk=−∆Qsc+∆Qcoh\n2±/radicalbigg\n(∆Qsc−∆Qcoh)2\n4+k2Q2s, (7)\nwhere ”+” is for modes k≥0. The coherent tune shift ∆ Qcohof the head-tail mode k\nis the result of the interaction with the imaginary part of th e facility impedance. It can\nbe calculated as an integral over the bunch spectrum with the frequency-dependent reactive\nimpedance[1,10,11]. For abroadband-typeimpedance(with respecttothebunchspectrum),\nthe neighboring modes have very close ∆ Qcoh, for example for the image charges. Figure3\nshows the combined effect of space-charge with the coherent tu ne shift in this case. For a\nnarrowband impedance (with respect to the bunch spectrum), the coherent tune shifts can be\nvery different and can contribute to the excitation of the tran sverse mode coupling instability\n[3, 8, 11] which is not in the scope of this paper.\nThe difference between the effect of an external impedance and sp ace-charge on head-\ntail modes is especially obvious for the k= 0 mode. Space-charge alone does not affect the\nk= 0 mode, ∆ Qk=0= 0. In contrast, a facility impedance alone shifts the k= 0 mode,\n∆Qk=0=−∆Qcoh.\n6-5-4.5-4-3.5-3-2.5-2-1.5-1-0.5 0\n 0 0.5 1 1.5 2 2.5 3Re(∆Qk)/Qs − k\n∆Qcoh / Qsk=0\nk=1\nk=2\nk=3\nFigure 3: Combined effect of a reactive impedances and space-c harge on the head-tail mode\npositioning according to Eq.(7) for q= 10. The dashed line indicates the damping border\n∆Qη.\nThere is an important implication from Eq.(7). For increasi ng ∆Qcoh(and a fixed ∆ Qsc),\neach head-tail mode crosses the damping border and enters th e regime with a stronger damp-\ning, see Fig.3. Some of the modes are damped shortly above ∆ Qcoh= 0, for example k= 3\nin Fig.3. Even the mode k= 0 is affected by the space-charge induced damping after certa in\n∆Qcoh. This implies that the coherent effects can change strongly th e damping conditions.\nHigh damping rates at strong space-charge, damping of the k= 0 mode, etc., are possible\nwith reactive impedances, in contrast to the previously dis cussed [8, 5] case of the self-field\nspace-charge only.\nThe coherent tune shift at which a head-tail mode Eq.(7) cros ses the damping border can\nbe determined using Eq.(3). The corresponding solution is\n∆Qcoh= ∆Qscη−kQs1−η\nk/q+1−η. (8)\nFor thek= 0 mode, this expression is\n∆Qcoh= ∆Qscη . (9)\nFor strong space-charge q≫kEq.(8) takes a simpler form,\n∆Qcoh= ∆Qscη−kQs. (10)\nThis means that the damping border grows linearly with ∆ Qsc(orq). For the higher-order\nhead-tail modes, the necessary ∆ Qcohis equidistantly shifted by kQs. Figure4 demonstrates\nthat the coherent tune shift needed for the chosen damping ra te can be easily predicted for\narbitrary space-charge conditions.\n7 0 2 4 6 8 10 12 14 16 18 20\n 0 10 20 30 40 50 60 70 80∆Qcoh / Qs\nq = ∆Qsc / Qsk=0\nk=1\nk=2\nFigure 4: Coherent tune shift ∆ Qcohneeded for a chosen damping rate for different head-tail\nmodes and for different strength of space-charge in bunches as predicted by Eq.(8).\n5 COMPARISONS BETWEEN THEORY AND SIMULA-\nTIONS\nWe verify the basic predictions of the analytical model Eq.( 8) using particle tracking simu-\nlations. Figure5 shows the positioning of the head-tail mod es with respect to the damping\nborder for η= 0.24. The damping conditions are compared for the moderate spa ce-charge\nq= 2 (left plot) and for the stronger space-charge q= 6 (right plot). This discussion\ncorresponds to a beam with a fixed intensity, but with an incre asing transverse imaginary\nimpedance. The mode k= 2 is always damped in the both plots of Fig.5. The mode k= 1 is\nalways damped for q= 2, but has a region outside the strong damping for q= 6. The mode\nk= 0 experiences no damping without coherent tune shifts, but it is affected by damping\nfor increasing ∆ Qcoh. Forq= 6, the border of the chosen damping is higher in ∆ Qcohthan\nforq= 2. The mode k= 1 forq= 6 has only a weak damping without coherent effect, and\naround ∆ Qcoh= 0.5Qsthe damping becomes stronger. Results of the simulations in Fig.6\nconfirm these observations.\nIn the next case we discuss an increasing ∆ Qscwith a fixed ∆ Qcoh/∆Qsc−ratio. This\ncorresponds to beams of an increasing intensity and with a co nstant transverse emittance.\nThe imaginary impedance is also fixed, for example, the vacuu m pipe. Figure7 shows the\npredictionoftheanalytical modelforthemode k= 1withthedifferentvaluesoftheparameter\nα= ∆Qcoh/∆Qsc. The mode crosses the damping boundary at higher q−values for larger\nα−ratios. The condition for this crossing for arbitrary kcan also be obtained from Eqs.(3,7),\nq=kα+1−2η\n(η−α)(1−η). (11)\nThis applies for α < ηand is plotted in Fig.8 for η= 0.24. Theq−value of the transfer across\nthe damping border goes to infinity as αapproaches η.\nPredictions of the analytical model in Figs.7,8 are compare d with the simulation results\ninFig.9. For thetwo cases ∆ Qcoh= 0and∆ Qcoh= 0.1∆Qsc, thereisaregion withthestrong\n8-2.5-2-1.5-1-0.5 0\n 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8Re(∆Qk)/Qs − k\n∆Qcoh / Qsk=0\nk=1\nk=2\n-2.5-2-1.5-1-0.5 0\n 0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8Re(∆Qk)/Qs − k\n∆Qcoh / Qsk=0\nk=1\nk=2\nFigure 5: Effect of the reactive impedances on damping for differ ent space-charge tune shifts\naccording to Eqs.(3, 7). Left plot: q= ∆Qsc/Qs= 2, right plot: q= 6. The dashed lines\nindicate the damping border.\n 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7\n 0 0.2 0.4 0.6 0.8 1103 (−Im(∆Q))\n∆Qcoh / Qsk = 0\n 0 0.2 0.4 0.6 0.8 1 1.2 1.4\n 0 0.5 1 1.5103 (−Im(∆Q))\n∆Qcoh / Qsk = 1\nk = 0\nFigure 6: Effect of the reactive impedances on damping obtaine d from the particle tracking\nsimulations. Left plot: q= ∆Qsc/Qs= 2, right plot: q= 6. Note that without reactive\nimpedances the k= 0 mode experiences no effect of space charge and no damping.\ndamping which transfers to a weaker damping by higher q. In the case of ∆ Qcoh= 0.1∆Qsc\nthe damping is stronger and has a larger q−region. Within the considered parameters, the\nmode stays in the strong damping for ∆ Qcoh= 0.2∆Qsc. Correspondingly, the coherent\nline remains below the damping border in Fig.7. Finally, a ne gative coherent shift ∆ Qcoh\npushes the head-tail mode out of the damping region and makes the damping weaker, and\nthe damping region smaller. This can be seen for the case with ∆Qcoh=−0.1∆Qscin\nthe analytical model Figs.7,8, and it is in a good agreement w ith the simulations in Fig.9.\nFigure9 also illustrates the enhancement of Landau damping due to an external impedance,\nespecially for stronger space-charge.\nUsing Fig.9 we can test the suggestion that the space-charge induced damping is directly\nrelated to the position of the coherent line with respect to t he damping border in the incoher-\nent spectrum. For every q−value and ∆ Qcohthe distance between the modulated theoretical\n9-4-3-2-1 0 1\n 0 2 4 6 8 10 12 14Re(∆Q/Qs − k\nq = ∆Qsc / Qs∆Qcoh = 0∆Qcoh = −0.1 ∆Qsc\n∆Qcoh = 0.1 ∆Qsc\n∆Qcoh = 0.2 ∆Qsc\nFigure 7: Effect of the reactive impedances on the k= 1 head-tail mode according to Eq.(7)\nfor different coherent tune shift in the case of a fixed α= ∆Qcoh/∆Qsc. The dashed line\nindicates the damping border Eq.(3) for η= 0.24.\n 0 5 10 15 20\n-0.1-0.05 0 0.05 0.1 0.15 0.2q\nα = ∆Qcoh / ∆Qsc\nFigure 8: The space-charge parameter at which the head-tail modek= 1 crosses the damping\nborder as given by Eq.(11) with η= 0.24.\ncoherent frequency Eq.(7) and the damping border Eq.(3),\n∆Qdamp= ∆Qk−kQs−∆Qη, (12)\ncanbecalculated andplottedasthehorizontal axis, seeFig .10. Thevalues∆ Qdamp<0imply\nthe position of the mode beyond the spectrum part ηsuch that the resulting damping rate\nshould behigher than that at the dampingborder. The points a tq= 0 in Fig.9 correspondto\n∆Qdamp= 0 in Fig.10. As we observe from Fig.7, the coherent lines div e below the damping\nborderfor small q, which correspondsto ∆ Qdamp<0 in Fig.10. With increasing q, the modes\ncome back to the damping border, and, in fact, the decrements return on the close path back\nto ∆Qdamp= 0 (see Fig.10). The exception is ∆ Qcoh= 0.2∆Qsc(blue), which stays well\n10 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7\n 0 1 2 3 4 5 6 7 8103 (−Im(∆Q))\nq = ∆Qsc / Qs∆Qcoh = 0∆Qcoh = 0.1 ∆Qsc∆Qcoh = 0.2 ∆Qsc\n∆Qcoh = −0.1 ∆Qsc\nFigure 9: Damping decrement of the k= 1 head-tail mode obtained from the particle tracking\nsimulations for a Gaussian bunch, Qs= 0.01, for different coherent tune shift due to reactive\nimpedances. The chosen parameters correspond to the analyt ical model predictions in Fig.7.\nbelow the damping border, its ∆ Qdampremain strongly negative and the decrements stay\nhigh. For the cases ∆ Qcoh= 0 (black) and ∆ Qcoh=−0.1∆Qsc(green), the modes gain\nlarge ∆Qdamp>0 and the damping becomes weak. All these damping rates are lo cated\nfairly close to a universal line of the space-charge induced damping. This demonstrates that\nthe damping rate is a monotonic function of the coherent mode position in the incoherent\nspectrum, irrespective the space-charge and ∆ Qcohconditions.\n 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7\n-0.6-0.4-0.2 0 0.2 0.4 0.6103 (−Im(∆Q))\n∆Qdamp∆Qcoh = 0∆Qcoh = 0.1 ∆Qsc∆Qcoh = 0.2 ∆Qsc\n∆Qcoh = −0.1 ∆Qsc\nFigure 10: Damping rates from Fig.9 vs. the mode-spectrum di stance Eq.(12). The plot\ndemonstrates that the space-charge induced damping depend s directly on the coherent mode\nposition with respect to the incoherent spectrum, irrespec tive the space-charge and ∆ Qcoh\nconditions.\n116 CONCLUSIONS\nSpace-charge inducedLandau damping in bunches has been stu died usingan analytical model\nand particle tracking simulations. It has been demonstrate d that two sources of the space-\ncharge tune spread, longitudinal and transverse, contribu te to the resulting damping. A\nspecific damping rate (for example, strong enough damping to stabilise an instability) can be\nconsidered as the damping border. The parameter ηfor the shift of the coherent mode into\nthe incoherent spectrum Eq.(3) can then be defined, dependin g on the distribution in bunch\ntails and on the chosen damping border.\nA model for the space-charge induced damping in bunches is pr esented. Using the airbag-\nbased analytical model [3] for the head-tail model with arbi trary space-charge Eq.(4), and the\nborder of the incoherent spectrum for a chosen damping Eq.(3 ), theq−region of the space-\ncharge induced damping can be predicted by Eq.(6). This resu lt is confirmed by the particle\ntracking simulations. Enhancement of Landau damping is dem onstrated for all considered\nhead-tail modes.\nWith the help of the analytical expression [Eq.(7)] from [13 ], a model for the damping\nborder has been suggested, Eq.(8). For the mode k= 0 and for strong space-charge, there\nare simple relations Eqs.(9,10). This model gives a univers al prediction for the coherent tune\nshift needed for the damping, Fig.4, for arbitrary space-ch arge conditions and for different\nhead-tail modes k.\nIt has been discussed [8, 9, 5, 6] that in the case of self-field space-charge only the mode\nk= 0cannotbeaffectedbyspace-chargeandbyrelateddamping. H erewehavedemonstrated\nthat theanalytical model(Fig.5)andthesimulations (Fig. 6) consistently show theconditions\nwhere the mode k= 0 can be damped.\nEffects of reactive impedances on different modes kare discussed for a fixed q= ∆Qsc/Qs\n(constant beam intensity and growing reactive impedance) a nd for a fixed α= ∆Qcoh/∆Qsc\n(constant reactive impedance and growing beam intensity). The both cases reveal different\ninteresting properties in the space-charge induced dampin g with reactive impedances. The\nanalytical model Eqs.(6,8,11) provide specific prediction s for the transition across the damp-\ning border for different modes k, which are confirmed by the particle tracking simulations.\nThe analysis of the simulation results has also demonstrate d (see Fig.10) that the space-\ncharge induced damping depends directly on the coherent mod e position with respect to the\nincoherent spectrum, irrespective the space-charge and th e reactive impedance conditions.\nIt has been shown that the coherent tune shifts can have a stro ng effect on the damping\ndue to space-charge in bunches. Landau damping can be dramat ically enhanced due to the\neffect of reactive impedances. Thus the reactive impedances a nd other ∆ Qcoh−sources should\nbetaken intoaccount inaninstability thresholdanalysis. Ontheother hand,thespace-charge\ninduceddampinginbunchescanbeenhancedbyemployingaddi tional reactive impedances, in\na passive (dedicated hardware elements) or in an active (a re active feedback [10, 19]) manner.\n12References\n[1] F.Sacherer, Proc. First Int. School of Particle Acceler ators, Erice, p.198 (1976)\n[2] F.Sacherer, CERN Report CERN/SI-BR/72-5 (1972)\n[3] M.Blaskiewicz, Phys. Rev. ST Accel. Beams 1, 044201 (1998)\n[4] V. Kornilov and O. Boine-Frankenheim, Proc. of ICAP2009 , San Francisco, USA, Aug\n31 - Sep 4 (2009)\n[5] V.Kornilov and O.Boine-Frankenheim, Phys. Rev. ST Acce l. Beams 13, 114201 (2010)\n[6] V.Kornilov and O.Boine-Frankenheim, Phys. Rev. ST Acce l. Beams 15, 114201 (2012)\n[7] R.Singh, et.al., Phys. Rev. ST Accel. Beams 16, 034201 (2013)\n[8] A.Burov, Phys. Rev. ST Accel. Beams 12, 044202 (2009); A.Burov, Phys. Rev. ST\nAccel. Beams 12, 109901(E) (2009)\n[9] V.Balbekov, Phys. Rev. ST Accel. Beams 12, 124402 (2009)\n[10] K.Y.Ng, Physics of Intensity Dependent Beam Instabili ties (World Scientific 2006)\n[11] A.Chao, Physics of Collective Beam Instabilities in Hi gh Energy Accelerators (John\nWiley Sons, Inc., New York, 1993).\n[12] A.Burov, V.Danilov, Phys. Rev. Letters 82, 2286 (1999)\n[13] O. Boine-Frankenheim, V. Kornilov, Phys. Rev. ST Accel . Beams, 12, 114201 (2009)\n[14] V.Kornilov, et.al., Proc. of HB2014, East Lansing, USA , November 10-14 (2014)\n[15] V.Kornilov, O.Boine-Frankenheim, Space-Charge 2015 Workshop, Oxford, March 23-27,\nhttps://www.cockcroft.ac.uk/events/SpaceCharge15/in dex.html (2015)\n[16] A.Macridin, A.Burov, E.Stern, J.Amundson, P.Spentzo uris, Phys. Rev. ST Accel.\nBeams18, 074401 (2015)\n[17] O.Boine-Frankenheim, V.Kornilov, Proc. of ICAP2006, 2-6 Oct., Chamonix Mont-\nBlanc, (2006)\n[18] A.Burov and V.Lebedev, Phys. Rev. ST Accel. Beams 12, 034201 (2009)\n[19] R.Ruth, CERN Report LEP-TH/83-22 (1983)\n13" }, { "title": "2209.14120v1.Tunable_nonlinear_damping_in_parametric_regime.pdf", "content": "Tunable nonlinear damping in parametric regime \nParmeshwar Prasad*, Nishta Arora, and A. K. Naik* \nCentre for Nano Science and Engineering, \nIndian Institute of Science, Bangalore, India, 560012 \n \nNonlinear damping plays a significant role in several area of physics and it is becoming \nincreasingly important to understand its underlying mechanism . However, microscopic origin \nof nonlinear damping is still a debatable topic. Here , we probe and report nonlinear damping \nin a highly tunable MoS 2 nano mechanical drum resonator using electrical homodyne \nactuation and detection technique . In our experiment, we achieve 2:1 internal resonance by \ntuning resonance frequency and observe enhanced non -linear damping . We probe the effect of \nnon-linear damping by characterizing parametric gain. Geometry and tunability of the device \nallow us to reduce the effect of other prominent Duffing non -linearity to probe the non -linear \ndamping effectively. The enhanced non -linear damping in the vicinity of internal resonance is \nalso observed in direct drive , supporting possible origin of non -linear damping . Our \nexperiment demonstrate s, a highly tunable 2D material based nano resonator offer s an \nexcellent platform to study the nonlinear physics and exploit nonlinear damping in parametric \nregime. \n \nDynamical systems in nature show ri ch physics demonstrating energy exchange. The systems \nfind its equilibrium m ediated by energy exchange. There are several ways through which a \nsystem accomplishes energy exchange with its surroundings. The energy exchange could be \ndissipative in nature and can be modelled as linear damping in the simplest case. However, \nthere are system which shows diverse nonlinear dissipative phenomena1–3. For example , nano \nmechanical systems (NEMS) offer such a system where vibration amplitude can be tuned to an \nextent to witness nonlinear dissipation2–5. In general, vibration amplitude comparable to the \nlength scale of its thickness drives the system in to nonlinear regime. In this regime, the \ndynamical systems offer rich physics dealing with nonlinear dynamics. As the nonlinear ity in \nthe system increases the presence of dissipative mechanics play an important role in dynamics \nof the nano -mechanical system s. \nFigure (1) Characterisation of the device: (a) Inset: SEM image of the device, the two thick \nyellow lines are source and dr ain electrodes, the suspended region is 3 𝜇𝑚 in diameter. The \ngate is 300 𝑛𝑚 below the suspended membrane . Schematic of the measurement setup : \nHomodyne capacitive actuation and detection technique. Membrane frequency is tuned using \nDC gate voltage 𝑉𝑔𝑑𝑐, actuated directly using 𝑉𝑔𝑎𝑐at frequency 𝜔 and parametrically at 2𝜔 \nusing 𝑉𝑝𝑎𝑐. AC voltages are combined using RF power combine r, further DC is combined with \nthe help of a bias-tee. Readout is amplified using low noise amplifier and measured at lock-in \namplifier locked at 𝜔. Frequency dispersion with 𝑉𝑔𝑑𝑐, two modes 𝜔1 (b) and 𝜔2 (c) with \nfrequency tuning > 1 MHz/V and 𝜔2~2 𝜔1. \nNonlinear dissipation has been of interest due to its presence in the applicat ion in biology6, \nmagnetization7, quantum optics8 and quantum information technologies9. However, source of \ntheir origin remains unclear and remains an important area to explore especially in low \ndimensional nanoresonators . Advancement in the nano -electromechanical system (NEMS) in \nthe last decade has led to explor ation of nonlinear dissipation mechanisms 2–4,10–12. It has been \nshown that the effect of nonlinear damping increases with vibration amplitude13. The nonlinear \nterm can be incorporated in the force equation as ∝𝜂𝑥2𝑧̇ where 𝜂 is the nonlinear damping \ncoefficient , 𝑧 is the displacement and 𝑧̇ is the velocity13. Though there are several models \nexplaining the microscopic origin of the nonlinear damping, the precise mechanism is not \nunderstood 10,14,15. This is partly due to the challeng es involved in isolat ing and decoupl ing the \nmechanism responsible for the nonli near damping from other nonlinear effects . Nonlinear \nphenomena are easily achievable in low -dimensional materials due to their extreme sensitiv ity \nto the external perturbation. In this context, 2D materials such graphene and transition metal \ndi-chalcogenid e (TMDC) provide an excellent platform to probe nonlinear damping. Recent ly, \nwork by Keskekler et. al4. shed light on the possible mechanism of the nonlinear damping \nthrough internal resonance (IR) in graphene using the opt ical actuation and detection method. \nThough optical method provides a sensitive detection techni que, it lacks the tunabi lity of the \nresonance frequency without heating the device . Heating is one of most common way to \ndissipate energy. \n \nFigure (2) (a) Amplitude response curve of mode1 (𝜔1) at multiple gate voltages. 𝑉𝑔𝑑𝑐×𝑉𝑔𝑎𝑐=\n0.2 𝑉2 is constant for the experiments. Shaded region shows the variation of amplitude. \nAmplitude is minimum at 𝑉𝑔𝑑𝑐=22 𝑉 (b) Inverse of quality factor at different 𝑉𝑔𝑑𝑐, it quantifies \nthe damping in the system. Damping is maximum at 𝑉𝑔𝑑𝑐=22 𝑉. \nIn this work, we demonstrat e the nonlinear damping mechanism in a highly tunable \nmolybdenum disulphide ( MoS2) device using the electrostatic homodyne capacitive actuation \nand detection method. The method provides a clean , fast and simple platform to study the \nnonlinear system16,17. Our device shows two prominent modes 𝜔1 and 𝜔2 the two modes are \n𝜔2∼2𝜔1. The modes provide favourable scenario to study the nonlinear damping in the \nparametric regime and probe the dissipation mechanism . In our experiment , we are able to tune \nthe nonlinear damping by changing t he applied gate voltage. We find the effect of nonlinear \ndamping considerably increases when the two modes are commensurate with each other in 2:1 \nratio. To probe the nonlinear damping , we use parametric pumping technique. Examining the \nparametric gain, we find that the effect of nonlinear damping reaches maximum in the vicinity \nof IR. Parametric gain decreases by as low as 8 3% when the nonlinear damping is maximum. \nWe observe , the enhancement in nonlinear damping is due to strong coupling between the two \nmodes near internal resonance . The experiment shed light into the mechanism of the nonlinear \ndamping in to nano mechanical system using a highly tunable NEMS . \nOur device is ~ 10 layer thick MoS2 drum resonator suspended over 3 μm diameter trench. The \ndevice is strained and actuated by applying voltage s (𝑉𝑔) at gate electro de. The gate electrode \nis 300 nm below the suspended membrane. Figure 1 a shows the schematic of the measurement \nused to measure the vibration of the membrane and for applying param etric pumping . The \nmembrane is strained by applying a DC gate voltage ( 𝑉𝑔𝑑𝑐) and actuated with an AC voltage \n(𝑉𝑔𝑎𝑐). The DC and the AC is combined using a bias -tee before applying to the gate electrode . \nApplied potential difference exerts f orce on the suspended membrane there by displacing it \nfrom the equilibrium position. The displacement of the membrane modulates the geometric \ncapacitance between gate and the membrane, leading to the change in voltage at the drain. The \nchange in drain volt age is amplifie d using a low -noise amplifier before measuring at the lock -\nin amplifier. Figure 1 b and c show the dispersion of the frequency with applied DC gate \nvoltage . We observe two distinct modes having similar tuning with each other. The two \nfrequenc ies are 𝜔1~ 56 MHz and 𝜔2 ~111 MHz at around 𝑉𝑔𝑑𝑐=23 V . The modes are \ninteresting due to 2:1 (i.e. 𝜔2~2𝜔1) ratio of the frequenc ies, which enables us to achieve \ninternal resonance by driving the fundamental mode in parametric regi me. We can achieve this \nIR ratio by tuning the gate voltage. \nTo obtain the elementary characteristics of the device, we drive it with fixed force (𝐹∝\n𝑉𝑔𝑑𝑐 𝑉𝑔𝑎𝑐) over a range of gate voltages ( 𝑉𝑔𝑑𝑐). A constant force allows us to observe and \ncompare the variation of amplitude response curve at different resonant frequency . We drive \nthe resonator with small force such that the response does not show nonlinear behaviour . The linear response can be quantified by the symmetry of the pe ak about the resonance frequency . \nIn a nonlinear regime a hysteresis in the frequency response curve is observed . Figure 2a shows \namplitude response with frequency at different gate voltages. We observe that amplitude \ndecreases with increase in gate voltage a nd reaches minimum value around 𝑉𝑔𝑑𝑐∼22 V. In a \nlinear system, a constant force results in a constant amplitude and linewidth due to linear \ndissipation. Reduction in amplitude for a constant applied force indicates enhanced dissipation. \nThe line wi dth of the response curve is shown in the figure 2b , it shows damping is maximum \nat 22 V. To probe the enhancement in damping , we perform parametric pumping by tuning \nstiffness of the membrane . Parametric pumping offers advantage in probing a resonance mod e \nby applying a force away from the resonant frequency and study the coupling of modes. \nParametric pump can be used alone or in a combination with direct drive . Parametric pump \nworks in a principle that it provides energy to overcome the linear damping of the system by \nmodulating the stiffness. \nWe use the following equation of motion to describe the parametric amplification in our \nsystem13: \n𝑚 𝑧̈+𝛾 𝑧̇+𝑚 𝜔02 𝑧+𝜂 𝑧̇ 𝑧2+𝛼𝑒𝑓𝑓 𝑧3=𝐹(𝜔 𝑡)+𝐹𝑝(2 𝜔 𝑡) 𝑧 (1) \nWhere m is mass of the system 𝛾 is the linear damping, 𝜔0 is the resonance frequency, 𝜂 is the \nnonlinear damping coefficient, 𝛼𝑒𝑓𝑓 is the effe ctive Duffing nonlinear ity coefficient associated \nwith geometric nonlinear ity. 𝐹 is the direct force applied at the frequency 𝜔, 𝐹𝑝 is the \nparametric force acting at the system at the frequency 2𝜔 and 𝑧 is the vibration amplitude. \nTo study parame tric amplification, we apply an alternating voltage (𝑉𝑝𝑎𝑐) at twice the resonance \nfrequency of the fundamental mode at the gate in addition to 𝑉𝑔𝑎𝑐 (see Figure 1 a). In a \nparametric system, there exists a critical pumping force beyond which the system overcomes \nlinear damping and reaches instability region also known as Mathew’s tongue13,18. The critical \npumping force can be measured by sweeping 𝑉𝑝𝑎𝑐 at 2𝜔. Figure 3 shows amplitude response \nusing parametric drive beyond critical pump voltage ( 𝑉𝑝𝑎𝑐0) for different 𝑉𝑔𝑑𝑐. Beyond the \ncritical pumping , the amplitude response is governed by the nonlinear parameters and offers \nan excellent regime to probe the nonlinear coefficients. There are two dominant nonlinear ities \npresent in our device namely Duffing nonline arity ( 𝛼𝑒𝑓𝑓) and nonlinear damping ( 𝜂). We \nobserve that 𝛼𝑒𝑓𝑓 changes from negative to positive , it changes sig n around 𝑉𝑔𝑑𝑐~24 V. It is \ninteresting for the fact that this allows us to probe the effect of nonlinear damping by minimi zing the effect of the prominent geometric nonlinear ity. Figure 3d shows the computed \nvalue of 𝛼𝑒𝑓𝑓 using continuum model (see supplementary for derivation and computation ). The \nvariation of computed 𝛼𝑒𝑓𝑓 matches well with our experimental fi ndings. Figure 3 a also shows \ndecreases in amplitude response at around 𝑉𝑔𝑑𝑐~24 V. The variation in amplitude response \nsimilar to the experimental findings shown in figure 2a. In the past, i t has been shown that the \namplitudes in parametric regime are saturated by nonlinear ities present in the system19. The \ncurrent device provides favourable platform to investigate the eff ect of the nonlinear damping \nwhile suppressing the effect of 𝛼𝑒𝑓𝑓. To probe the nonlinear damping, we perform parametric \namplification and observe the amplitude gain in the system. \n \nFigure 3 Tuning of geometric nonlinear ity (a) Amplitude response with upward frequency \nsweep for different 𝑉𝑔𝑑𝑐 with 𝑉𝑝𝑎𝑐=400 𝑚𝑉 and 𝑉𝑔𝑎𝑐=0. Change in the s hape of amplitude \nresponse curve shows variation in geometric nonlinear ity. Amplitude response at (b) 𝑉𝑔𝑑𝑐=\n16 V and (c) 𝑉𝑔𝑑𝑐=28 V. (d) Blue line is effective Duffing nonlinearity (𝛼𝑒𝑓𝑓) calculated \nusing continuum model, it crosses the 𝛼𝑒𝑓𝑓=0 (red line) around 𝑉𝑔𝑑𝑐=24 V, which matches \nthe experimental observation of in (a) . \nIn parametric amplification , the gain ( 𝐺=𝑧𝑜𝑛\n𝑧𝑜𝑓𝑓) is defined as the ratio of amplitude when 𝑉𝑝𝑎𝑐 \nis ON to OFF. Figure 4 a shows parametric gain with 𝑉𝑝𝑎𝑐 for different 𝑉𝑔𝑑𝑐. It shows that gain \nincreases steadily and get saturated after the critical pumping voltage. The maximum gain 𝐺𝑚𝑎𝑥 \nobtained before the critical pumping is shown in the figure 4b. It shows that the 𝐺𝑚𝑎𝑥 decreases \nsignificantly in the region 𝑉𝑔𝑑𝑐= 22 V to 𝑉𝑔𝑑𝑐= 26 V. The decrease in the 𝐺𝑚𝑎𝑥 is as low as \n83% as compared the 𝐺𝑚𝑎𝑥 at 𝑉𝑔𝑑𝑐= 20 V. To find the nonlinear damping coefficient we use \nthe relation 𝜂=2(𝐹𝑝𝑄−2)\n𝑧𝑚𝑎𝑥2 , where 𝐹𝑝 is parametric force, 𝑄 quality factor and 𝑧 is the maximum \namplitude13. The relation allows us to estimate the coefficient of nonlinear damping \nindependent of any nonlinear parameter in the lowest order . Figure 4e shows the coefficient of \nthe nonlinear damping extra cted for different gate voltages. \n \nFigure (4) Parametric amplification: (a) Amplitude gain (𝐺) vs applied pump strength 𝑉𝑝𝑎𝑐 for \nmultiple 𝑉𝑔𝑑𝑐. Gain increases as pump strength increases, eventually saturating due to \nnonlinear effects. (b) Maximum gain (𝐺𝑚𝑎𝑥) at the start of instability tongue vs 𝑉𝑔𝑑𝑐. 𝐺𝑚𝑎𝑥 \ndecreases in the region 𝑉𝑔𝑑𝑐=22 𝑉 to 𝑉𝑔𝑑𝑐=26 V. (c)(d) Self oscillation of mode1 , response \nmeasured at 𝜔 is plotted against pump 𝑉𝑝𝑎𝑐 applied at frequency 2𝜔. (c) and (d) are upward \nand downward frequency sweep respectively. Dashed rectangle shows the region of frequency \nlocking, resulting from 1:2 internal resonance . (e) Coefficient of nonlinear damping 𝜂 \ncalculated using the parametric model using the equation 1 . Higher value of 𝜂 is observed in \nthe region 𝑉𝑔𝑑𝑐=22 𝑉 to 𝑉𝑔𝑑𝑐=26 V. It explains the low gain observed in (b) and matches \nthe experimental findings. \n \nWe observe that the nonlinear damping increases significantly around 𝑉𝑔𝑑𝑐=24 𝑉. Around the \ngate voltage , we observe the minima of the amplitude and increased damping in case of direct \ndriving (See figure 2a, b). Enhancement of nonlinear damping is evident , as suggested by \ndecrease in amplitude during direct drive and the gain during parametric amplification . Figure \n4c,d shows instability region, 𝑉𝑝𝑎𝑐 is swept at frequency 2𝜔 and amplitude response is \nmeasured at 𝜔. As we increase the parametric pump strength, the direction of peak shifts from \nlower frequency and to higher frequency, indicating change in the value of 𝛼𝑒𝑓𝑓, which \nmatches our theoretical prediction using continuum model of the drum resonator. Upon further \nincreasing the pump, we observe the frequency locking , we note that this is in the vicinity of \nfavourable 2:1 internal resonance. In the internal resonance, the two m odes are coupled to each \nother, it makes pathway to exchange the energy. In the vicinity of IR, the mode 1 dissipates \nenergy to mode2 enhancing the nonlinear damping. \nIn summary, we study rich nonlinear physics in MoS 2 drum resonator in parametric regime . \nWe observe two prominent tunable modes in 50 MHz to 200 MHz range and they are coupled \nto each other through tension in the membrane . Dominant nonlinear coefficients 𝛼𝑒𝑓𝑓 and 𝜂 \nare tunable in our device using DC gate voltage . We tune the modes such that it facilitates 2:1 \ninternal resonance and exchange energy with each other. We observe that the nonlinear \ndamping increases significantly in the vicinity of internal resonance , 𝜂 reaching as high as \n60×1015 kg m−2s−1 in MoS 2 drum resonator . In our experiment, minimizing the effect of \n𝛼𝑒𝑓𝑓 helps us to probe nonlinear damping efficiently. A highly tunable NEMS device offers \nversatile control on the nonlinear coefficient s and in probing them. Our experiment show s \nevidence of the microscopi c origin of nonlinear damping and shed light in understan ding the \nunderlying physics . Parametric regime provides excellent approach to probe the nonlinear \nphysics by tuning the fundamental parameters such as stiffness. \n Acknowledgements: \nWe acknowledge fu nding support from Nano Mission, Department of Science and Technology \n(DST), India through Grants SR/NM/NS -1157/2015(G) and SR/NMITP -62/2016(G) and from \nBoard of Research in Nuclear Sciences (BRNS), India through Grant 37(3)/14/25/2016 -BRNS. \nP.P. acknowled ges scholarship support from CSIR, India. N.A. acknowledges fellowship \nsupport under Visvesvaraya Ph.D.Scheme, Ministry of Electronics and Information \nTechnology(MeitY), India. We also acknowledge funding from MHRD, MeitY, and DST \nNano Mission for supporti ng the facilities at CeNSE. We gratefully acknowledge the usage of \nthe National Nanofabrication Facility (NNfC) and the Micro and Nano Characterization \nFacility (MNCF) at CeNSE, IISc, Bengaluru . \n \nReferences : \n1 A. Eichle r, J. Moser, J. Chaste, M. Zdrojek, I. Wilson -Rae, and A. Bachtold, Nat. \nNanotechnol. 6, 339 (2011). \n2 J. Güttinger, A. Noury, P. Weber, A.M. Eriksson, C. Lagoin, J. Moser, C. Eichler, A. \nWallraff, A. Isacsson, and A. Bachtold, Nat. Nanotechnol. 12, 631 (2 017). \n3 A. Eichler, J. Chaste, J. Moser, and A. Bachtold, Nano Lett. 11, 2699 (2011). \n4 A. Keşkekler, O. Shoshani, M. Lee, H.S.J. van der Zant, P.G. Steeneken, and F. Alijani, \nNat. Commun. 12, 1099 (2021). \n5 J. Atalaya, T.W. Kenny, M.L. Roukes, and M.I. Dykman, Phys. Rev. B 94, 195440 (2016). \n6 M. Amabili, P. Balasubramanian, I. Bozzo, I.D. Breslavsky, G. Ferrari, G. Franchini, F. \nGiovanniello, and C. Pogue, Phys. Rev. X 10, 011015 (2020). \n7 B. Divinskiy, S. Urazhdin, S.O. Demokritov, and V.E. Demidov, Nat. Commun. 10, 5211 \n(2019). \n8 L. Mandel and E. Wolf, Optical Coherence and Quantum Optics (Cambridge University \nPress, 1995). \n9 Z. Leghtas, S. Touzard, I.M. Pop, A. Kou, B. Vlastakis, A. Petrenko, K.M. Sliwa, A. Narla, \nS. Shankar, M.J. Hatridge, M. Reagor, L. Frunzio, R.J. Schoelkopf, M. Mirrahimi, and M.H. \nDevoret, Science 347, 853 ( 2015). 10 S. Zaitsev, O. Shtempluck, E. Buks, and O. Gottlieb, Nonlinear Dyn. 67, 859 (2012). \n11 V. Singh, O. Shevchuk, Y.M. Blanter, and G.A. Steele, Phys. Rev. B 93, 245407 (2016). \n12 M. Imboden, O. Williams, and P. Mohanty, Appl. Phys. Lett. 102, 103502 (2013). \n13 R. Lifshitz and M.C. Cross, Reviews of Nonlinear Dynamics and Complexity (Wiley -VCH \nVerlag GmbH & Co. KGaA, Weinheim, Germany, 2008). \n14 X. Dong, M.I. Dykman, and H.B. Chan, Nat. Commun. 9, 3241 (2018). \n15 A. Croy, D. Midtvedt, A. Isacsson, and J.M. Kinaret, Phys. Rev. B 86, 235435 (2012). \n16 P. Prasad, N. Arora, and A.K. Naik, Nano Lett. 19, 5862 (2019). \n17 J.P. Mathew, R.N. Patel, A. Borah, R. Vijay, and M.M. Deshmukh, Nat. Nanotechnol. 11, \n747 (2016). \n18 A.H. Nayfeh and D.T. Mook, Nonlinear O scillations (Wiley -VCH Verlag GmbH, \nWeinheim, Germany, 1995). \n19 R. Lifshitz and M.C. Cross, Phys. Rev. B 67, 134302 (2003). \n \n \n \n \n \n \n \n \n \n \n \n Supplementary Information \nTunable nonlinear damping in parametric regime \n \nCalculation of Effective Duffing Non -linearit y (𝜶𝒆𝒇𝒇): \nFigure S1 describes the MoS 2 resonator clamed at the edge of a circular trench. An electrostatic \npotential ( 𝑉𝑔𝑑𝑐) applied at gate bends the membrane and tune the frequency. The applied strain \nusing the electrostatic force also modif ies the non -linear coefficients. The gate voltage can be \nused as a knob to control the non -linear coefficient which is discussed the following section. \n \n. \nFigure S 1: MoS 2 NEMS: A DC voltage 𝑉𝑔𝑑𝑐 is applied at the gate to de form the membrane. \nDistance between the trench and the gate is 𝑧0. \n \nWe consider continuum model of the drum resonator1: \n𝜕𝜏2𝜁𝛼+Λ𝛼𝜁𝛼+∑∑𝑄𝛽𝛾𝛼𝜁𝛽𝜁𝛾∞\n𝛾≥𝛽∞\n𝛽=1+∑∑∑𝐶𝛽𝛾𝜂𝛼𝜁𝛽𝜁𝛾𝜁𝜂∞\n𝜂≥𝛾∞\n𝛾≥𝛽∞\n𝛽=1=𝑓𝛼(𝜏) (𝑆1.1) \nWhere, 𝜁 are the coordinates, 𝑄𝛽𝛾𝛼 and 𝐶𝛽𝛾𝜂𝛼 are the coupling constants, 𝑓𝛼(𝜏) is the applied \nforce and Λ𝛼 are the eigen frequencies squared. \nThe maximum scaled deflection ( 𝜁0) at the center of the resonator is given by: \n𝜁0=𝑏 (2\n3𝜃)1\n3\n−(𝜃\n18𝑐3)1\n3\n(𝑆1.2) \n \n𝜃=√3(4(𝑏 𝑐)3+27𝑎2𝑐4)12⁄−9𝑎 𝑐2(𝑆1.3) \nWhere 𝑎= −𝑅 𝜖0 𝑉𝑔𝑑𝑐2\n4𝑌ℎ 𝑑 𝑧02, 𝑏=2𝑇0\n𝑌ℎ−𝑅2𝜖0 𝑉𝑔𝑑𝑐2\n3𝑌 ℎ 𝑧03 and 𝑐=7−𝜈\n6(1−𝜈) \nWhich translates the Stress ( 𝑇) in the membrane as: \n𝑇=𝑇0(1+𝑧2\n4) (𝑆1.4) \nWhere 𝑧 is scaled static center deflection and given by: \n𝑧=𝜁0(𝑌 ℎ\n𝑇0)1\n2√(3−𝜈\n1−𝜈) (𝑆1.5) \nThe above information is used to find the non -linear coefficients. \n𝛼3=𝐶1𝑇\n𝑅4𝜌0 (𝑆1.6) \n \n𝛼2=𝑄1𝑇\n𝑅3𝜌0 (𝑆1.7) \n \nWhere 𝐶1 and 𝑄1 are the coupling constant determined as follows: \n𝐶1=𝑌 ℎ\n𝑇(3.92+3.68 𝛿𝜈) (𝑆1.8) And \n𝑄1=𝜁0𝑌 ℎ\n𝑇(11.7+11.3 𝛿𝜈) (𝑆1.9) \nAfter the determination of the 𝛼2 and 𝛼3, 𝛼𝑒𝑓𝑓 can be defined as:2 \n𝛼𝑒𝑓𝑓=𝛼3−10\n9(𝛼2\n𝜔0)2\n(𝑆1.10) \nWe used the following parameters for the calculation of 𝛼𝑒𝑓𝑓 \nParameters Value \nYoung’s modulus 0.33×1012 Pa \nMass density of MoS2 (ρ) 5.06×103 kg.m−3 \nPoisson ratio ( ν) 0.27 \n \n \n(a) (b) (c) \n \n(d) \nFigure S3: Variation (a) Static deflection (b) DC force (c) AC force (d) 𝛼𝑒𝑓𝑓 with applied \nDC gate voltage 𝑉𝑔𝑑𝑐. \n \nCalculat ion of non -linear damping coefficient: \nWe consider the following equation to describe the parametric drive3: \n𝑚 𝑧̈+𝛾 𝑧̇+𝑚 𝜔02 𝑧+𝜂 𝑧̇ 𝑧2+𝛼𝑒𝑓𝑓 𝑧3=𝐹(𝜔 𝑡)+𝐹𝑝(2 𝜔 𝑡) 𝑧 (𝑆2.1) \nEquation S2.1 describes nano mechanical Duffing resonator with linear and non -linear \ndamping driven by a direct force 𝐹 and a parametric force 𝐹𝑝 acting at 𝜔 and 2𝜔 respectively. \nThe resonator has mass 𝑚 and resonance frequency 𝜔0. 𝛼𝑒𝑓𝑓, 𝛾, and 𝜂 are Duffing, linear and \nnon-linear damping coefficients respectively. Displacement amplitude ( 𝑧) is a time ( 𝑡) varying \nfunction and dot represents derivative with respect to time. \nCoefficients of non -linear damping ( 𝜂) can be derived from equation S1 and written as follows: \n𝜂=2 𝐹𝑝 𝑄−4\n𝑧2(𝑆2.2) \nWhere the vibration amplitude ( 𝑧) is scaled as follows: \n(𝑚 𝜔02\n𝛼̃)1\n2\n𝑧=𝑧̃ (𝑆2.3) \n𝜂̃𝜔0\n𝛼= 2 𝐹𝑝 𝑄−4\n𝑧̃2𝛼\n𝑚 𝜔02(𝑆2.4) \n𝜂̃=(2 𝐹𝑝 𝑄−4) 𝑚 𝜔0\n𝑧̃2(𝑆2.5) \nWhere: \n𝐹𝑝=2 𝐴 𝜖0 𝑉𝑔𝑑𝑐 𝑉𝑝𝑎𝑐\n𝑧̃03(𝑆2.6) \n \nNon-linear damping constant can be estimated using equation S2.5. \n \nReferences: \n1 A.M. Eriksson, D. Midtvedt, A. Croy, and A. Isacsson, Nanotechnology 24, 395702 (2013). 2 A.H. Nayfeh and D.T. Mook, Nonlinear Oscillations (Wiley -VCH Verlag GmbH, Weinheim, \nGermany, 1995 ). \n3 R. Lifshitz and M.C. Cross, Reviews of Nonlinear Dynamics and Complexity (Wiley -VCH \nVerlag GmbH & Co. KGaA, Weinheim, Germany, 2008). \n \n " }, { "title": "0710.4610v1.Damping_of_Condensate_Oscillation_of_a_Trapped_Bose_Gas_in_a_One_Dimensional_Optical_Lattice_at_Finite_Temperatures.pdf", "content": "arXiv:0710.4610v1 [cond-mat.other] 25 Oct 2007Damping of Condensate Oscillation of a Trapped Bose Gas in a\nOne-Dimensional Optical Lattice at Finite Temperatures\nEmiko Arahata and Tetsuro Nikuni\nDepartment physics, Faculty of science, Tokyo University o f Science,\n1-3 Kagurazaka, Shinjuku-ku, Tokyo 162-8601, Japan\n(Dated: November 5, 2018)\nAbstract\nWe study damping of a dipole oscillation in a Bose-Condensed gas in a combined cigar-shaped\nharmonic trap and one-dimensional (1D) optical lattice pot ential at finite temperatures. In order\nto include the effect of thermal excitations in the radial dire ction, we derive a quasi-1D model of\nthe Gross-Pitaeavskii equation and the Bogoliubov equatio ns. We use the Popov approximation to\ncalculate the temperature dependence of the condensate fra ction with varying lattice depth. We\nthen calculate the Landau damping rate of a dipole oscillati on as a function of the lattice depth\nand temperature. The damping rate increases with increasin g lattice depth, which is consistent\nwith experimental observations. The magnitude of the dampi ng rate is in reasonable agreement\nwith experimental data. We also find that the damping rate has a strong temperature dependence,\nshowing a sharp increase with increasing temperature. Fina lly, we emphasize the importance of\nthe radial thermal excitations in both equilibrium propert ies and the Landau damping.\n1I. INTRODUCTION\nRecently the dynamics of ultracold atomic gases in optical lattices ha ve attracted atten-\ntion both theoretically and experimentally [1]. In particular, center- of-mass dipole oscilla-\ntions of Bose-Einstein condensates in a combined cigar shaped trap and one dimensional\n(1D) optical lattice potential have been experimentally studied in de tail [2, 3, 4]. In the\npresence of the periodic lattice potential, a decrease of the dipole m ode frequency was ob-\nserved [3]. This decrease can be understood in terms of the increas e of the effective mass\ndue to the lattice potential [5]. On the other hand, at finite tempera tures where an ap-\npreciable number of atoms are thermally excited out of the condens ate, strong damping of\nthe dipole oscillation of the condensate was observed in the presenc e of the lattice potential\n[4]. In a pure harmonic potential, a Bose gas exhibits undamped dipole o scillations even at\nfinite temperatures, since the condensate and noncondensate a toms oscillate with the same\nfrequency without changing their density profiles [6]. In contrast, in the presence of the\nperiodic lattice potential, only the condensate component can cohe rently tunnel through\nthe potential barriers, while the thermal component is locked by th e lattice potential [4]. It\nis clear that this incoherent thermal conponent gives rise to the da mping of the condensate\noscillations. In Bose-condensed gases trapped in harmonic potent ials, Landau damping is\nknown to be the dominant contribution to damping of the condensat e oscillations in the\ncollisionless regime [7, 8, 9, 10]. In the analysis of damping of dipole oscilla tions in opti-\ncal lattice, the authors of Ref. [4] gave a rough estimate of the La ndau damping rate and\ncompared it with their experimental data. However, quantitative c alculations of the Landau\ndamping rate of dipole condensate oscillations in optical lattices have not given in any detail\nso far. In fact, even equilibrium properties of a Bose gas in an optica l lattice at finite tem-\nperatures in connection with the experimentas of Ref. [4] have not been studied in detail.\nAlthough several papers have discussed finite-temperature pro perties of ultracold atoms in a\n1D optical lattice, most theoretical studies have concentrated o n the first Bloch band using\nthe Bose-Hubbard model, and have ignored the effect of radial exc itations [11, 12, 13]. This\napproximation is only valid when kBT≪ERandkBT≪¯hω⊥, whereERis the recoil energy\nthat is roughly the highest energy in the first Bloch band, and ¯ hω⊥is the first excitation\nenergy in the radial direction. In order to obtain more quantitative results that is applicable\nto the experiments of Refs. [3, 4], however, it is important to consid er thermal excitations in\n2the higher bloch bands as well asa the radial direction, since the exp eriments do not always\nsatisfykBT≪ERandkBT≪¯hω⊥.\nIn this paper, we study the Landau damping of condensate oscillatio ns of a trapped Bose\ngas in a 1D optical lattice, with explicitly including the effect of the radia l excitations. In\nSec. II, we derive a quasi-1D model of the Gross-Pitaeavskii equa tion and the Bogoliubov\nequations that include the effect of the excitations in the radial dire ction. Using the Hatree-\nFock-Bogoliubov-Popov (HFB-Popov) approximation, we solve the se equations to calculate\nthe temperature dependence of the condensate fraction.\nIn Sec. III, we calculate the Landau damping of the dipole condensa te oscillation in an\noptical lattice. We calculate the damping rate as a function of the lat tice depth with a fixed\ntemperature, and compare it with the experimental data [4]. The ma gnitude of the damping\nrate is found to be in reasonable agreement of the experimental da ta [4]. The damping rate\nincreases with increasing lattice depth, which is consistent with the e xperimental result [4].\nWe also calculate the temperature dependence of damping rate with a fixed lattice depth.\nWefindthatthedampinghasastrongtemperature, showing ashar pincreasewithincreasing\ntemperature.\nII. QUASI 1D MODELING OF A TRAPPED BOSE GAS\nWe consider a Bose condensed gas in a combined potential of highly-e longated harmonic\ntrap and 1D optical lattice. Our system is described by the following H amiltonian:\nˆH=/integraldisplay\ndr/braceleftBigg\nˆψ†(r)/bracketleftBigg\n−¯h2\n2m∇2+Vext(r)/bracketrightBigg\nˆψ(r)\n+g\n2ˆψ†(r)ˆψ†(r)ˆψ(r)ˆψ(r)/bracerightBigg\n, (1)\nwhereg=4π¯h2a\nmis the coupling constant determined by the s-wave scattering length\na. The external potential Vextis given by Vext(r) =Vtrap(r) +Vop(z), whereVtrap(r) =\nm\n2[ω2\n⊥(x2+y2)+ω2\nzz2] is an anisotropic harmonic potential and Vop(z) =sERcos2(kz) is an\noptical lattice potential. Here sis a dimensionless parameter describing the intensity of the\nlaser beam creating the 1D lattice in units of the recoil energy ER≡¯h2k2\n2m, wherek=2π\nλis\nfixed by the wavelength λof the laser beam.\nInthis paper we consider ahighly-anisotropic cigar shaped harmonic trappotential ω⊥≫\nωz. In order to take into account this quasi-1D situation, we expand t he field operator in\n3terms of the radial wave function [14]:\nˆψ(r) =/summationdisplay\nαˆψα(z)φα(x,y), (2)\nwhereφα(x,y) is the eigenfunction of the radial part of the single-particle Hamilto nian,\n/bracketleftBigg\n−¯h2\n2m∇2\n⊥+m\n2ω2\n⊥(x2+y2)/bracketrightBigg\nφα(x,y)\n=ǫαφα(x,y), (3)\nwhich satisfy the orthonormality condition/integraltextdxdy φ∗\nα(x,y)φβ(x,y) =δαβ. Hereα= (nx,ny)\nis the index of the single-particle state with the eigenvalue ǫ(nx,ny)= ¯hω⊥(nx+ny+ 1).\nInserting Eq. (2) into Eq. (1) and using Eq. (3), we obtain\nˆH=/summationdisplay\nα/integraldisplay\ndzˆψ†\nα(z)/bracketleftBigg\n−¯h2\n2m∂2\n∂z2+m\n2ω2\nzz2+Vop(z)+ǫα/bracketrightBigg\nˆψα(z)\n+/summationdisplay\nαα′ββ′gαα′ββ′\n2/integraldisplay\ndzˆψ†\nαˆψ†\nβˆψβ′ˆψα′, (4)\nwhere the renormalized coupling constant is defined by\ngαα′ββ′≡g/integraldisplay\ndxdyφ∗\nαφ∗\nβφβ′φα′. (5)\nFollowing the procedure described in Ref. [15] , we separate out the condensate wavefunc-\ntionfromthe field operatoras ˆψα=/an}b∇acketle{tˆψα/an}b∇acket∇i}ht+˜ψα≡Φα+˜ψα, where Φ α=/an}b∇acketle{tˆψα/an}b∇acket∇i}htis the condensate\nwavefunction and ˜ψαis the noncondensate field operator. Within the HFB-Popov approx i-\nmation, we obtain the generalized Gross-Pitaevskii (GP) equation,\nµΦα=/bracketleftBigg\n−¯h2\n2m∂2\n∂z2+m\n2ω2\nzz2+Vop+ǫα/bracketrightBigg\nΦα\n+/summationdisplay\nα′ββ′gαα′ββ′/parenleftBig\nΦ∗\nβΦβ′Φα′+2Φβ′/angbracketleftBig˜ψ†\nβ˜ψα′/angbracketrightBig/parenrightBig\n. (6)\nThePopovapproximationneglectstheanomalouscorrelation/angbracketleftBig˜ψ˜ψ/angbracketrightBig\n[15]. Fromthenumerical\nsolutions of the GP equation (6) using the trap frequencies relevan t to the experiment [4],\nwe find that |Φα|2/|Φ0|2≪10−6(α/ne}ationslash= 0), where we have denoted the lowest radial state\nasα= 0 = (0,0), and thus the contribution from higher radial modes to the cond ensate\nwavefunction isnegligiblesmall. Forthisreason, wewill henceforthap proximateΦ α≈Φδα,0.\nTaking the usual Bogoliubov transformations for the noncondens ate,\n˜ψα(z) =/summationdisplay\nj/parenleftBig\nujαˆαj−v∗\njαˆα†\nj/parenrightBig\n, (7)\n4we obtain the coupled Bogoliubov equations,\nˆLαujα+/summationtext\nα′/bracketleftBigg\n(2gα\nα′n0+gαα′ββ′˜nββ′)ujα′−gα\nα′n0vjα′/bracketrightBigg\n=Ejujα, (8)\nˆLαvjα+/summationtext\nα′/bracketleftBigg\n(2gα\nα′n0+gαα′ββ′˜nββ′)vjα′−gα\nα′n0ujα′/bracketrightBigg\n=−Ejujα, (9)\nwhere we have introduce the operator\nˆLα≡ −¯h2\n2m∂2\n∂z2+m\n2ω2\nzz2+Vop(z)+ǫα−µ. (10)\nWehavealsointroducethesimplified notations n0(z) =|Φ(z)|2,gα\nα′=gαα′00,˜nββ′=/angbracketleftBig˜ψ†\nβ˜ψβ′/angbracketrightBig\n.\nAs noted above, we only include the lowest mode ( α= 0) in the condensation wavefunction\nΦ. Sums over the repeated indices β,β′are implied in Eqs. (8), and (9). These equations\ndefine the quasi-particle excitation energies Ejand the quasi-particle amplitudes ujαand\nujα. The orthonormality of the quasi-particle amplitudes is specified by t he relation\n/integraldisplay\ndz/parenleftBig\nu∗\njαuiα−uiαu∗\njα/parenrightBig\n=δij (11)\nUsing the solutions of Eqs. (8) and (9), one can obtain the noncond ensate density from\n˜n=/summationtext\nα˜nαα, where\n˜nαβ=/summationdisplay\nj/bracketleftbigg\n(ujαujβ+vjαvjβ)N(Ej)+vjαvjβ/bracketrightbigg\n(12)\nwithN(Ej) = 1/[exp(βEj)−1].\nSolving the coupled equations (6), (8) and (9), we self-consistent ly determine the exci-\ntations spectrum Ejand the condensate fraction at finite temperatures. Our calculat ion\nprocedure is summarized as follows. Eq. (6) is first solved self-cons istently for µand Φ ne-\nglecting the interaction terms. Once Φ is known, Ej,ujαandujβare obtained from Eqs. (8)\nand (9) with ˜ nββ′set to zero. This is inserted into Eq. (6) and the process is repeate d until\nconvergence is reached. At each step, we define the normalization of the condensate wave-\nfunctionsby/integraltextdz|Φ(z)|2=N−˜N, where˜N=/integraltextdz˜n(z)isthetotalnumber ofnoncondensate\natoms.\nThroughout this paper we use the following parameters of the expe riment of Ref. [4]:\nm(87Rb)=1.44 ×10−25kg,ωz/2π= 9.0 Hz,ω⊥/2π= 92 Hz, scattering length a= 5.82 nm\n5and the wavelength of the optical lattice λ=795 nm. We fix the total number of atoms as\nN= 4×105. InFig.1, weplotthecondensatefraction Nc/Nasafunctionofthetemperature\nfor various values of the lattice depth sby solving GP equation Eq. (6) and Bogolibov\nequations Eqs. (8) and (9). It can be seen that each line falls to zer o at approximately\nT=140 nK, which is close to the semiclassical prediction of the BEC tran sition temperature\nof an ideal Bose gas in a 3D harmonic trap T0\nc= 0.94¯h(ω2\n⊥ωz)1/3N1/3/kB= 141 nK [17]. In\ncontrast,Tcof an ideal Bose gas in a 1D harmonic trap is Tc∼¯hωzN\nkBlnN=13.4µK [17].\nThis crearly shows that one must explicitly take into account the exc itations in the radial\ndirection in order to obtain the correct thermodynamic behavior at finite temperatures\n[14]. The transition temperature deceases with increasing lattice de pth, but there is no\nsignificant change in the temperature dependence of the condens ate fraction, as long as the\nlattice potential is not so deep, i. e., s≤2,\nFIG. 1: The condensate fraction Nc/Nas a function of the temperature. “ s=0” represents the\nideal Bose gas result without a lattice potential Nc/N= 1−(T/T0\nc)3, whereT0\nc=141 nK.\nWe turn to the detailed structure of the excitation spectrum Ej. From Eqs. (8) and\n(9), one sees that the different radial modes are coupled due to int eractions. However the\ncoupling is so small that there are still distinct branches correspon ding the radial modes.\nWe thus label each branch with the index α.For example, the lowest branch ( α= 0) can be\nidentified with the branch corresponding to the lowest radial mode. In Fig 2, we plot the\nexcitation spectrum. Frequency of the condensate collective mod e is related the excitation\nenergy through Ej= ¯hωj. In particular, the dipole mode frequency can be identified with\nthe lowest frequency. In Fig. 3, we plot the dipole-mode frequency as a function of the\nlattice depth with a fixed temperature T= 40nK. One can see the decrease of the dipole-\n6FIG. 2: The Bogoiubov excitation energy spectrum for s=1.2, where iis the energy label for each\nbranch\nmode frequency with increasing lattice depth. The same behavior wa s also observed in\nthe experiment [3]. The negative energy shift also found for other excitation modes. The\ndecrease of Tccan be attributed to these negative shifts.\nFIG. 3: The dipole mode frequency as a function of the lattice depthswith a fixed temperature\nT= 40nK.\nIII. LANDAU DAMPING RATE OF DIPOLE MODE\nIn this Section, we calculate Landau damping of the dipole mode. Land au damping is\nthe dominant damping mechanism for low-lying collective modes in trapp ed Bose-condensed\ngases in the collisionless regime at finite temperatures [8]. Landau da mping originates from\nthe coupling between single-particle excitations and the collective os cillations. Damping\noccurs because the thermal bath of the elementary excitations c an absorb quanta of the\n7collective oscillations. A general expression for the Landau damping rate is given in Refs. [7,\n8, 10, 17]. In our quasi-1D model, Landau damping rate can be expre ssed in terms of the\nquasi-particle excitation energies Ejand quasi-particle amplitudes ujαandvjαcalculated in\nthe previous Section:\nγL= 4π/summationdisplay\nα,α′gαα′00/summationdisplay\ni/negationslash=j|Aαα′\nij|2/bracketleftBig\nN(Ei)−N(Ej)/bracketrightBig\nδ(¯hω+Ei−Ej), (13)\nwhereωis the eigenfrequency of a condensate dipole oscillation. The matrix e lementAαα′\nij\nis defined by,\nAαα′\nij≡/integraldisplay\ndzΦ{u10[u∗\niαujα′+v∗\niαvjα′−v∗\niαujα′]\n−v10[u∗\niαujα′+v∗\niαvjα′−u∗\niαvjα′]} (14)\nwhereu10andv10are the quasi-particle amplitudes corresponding the dipole mode. On e\ndifficulty in calculating the damping rate in Eq. (13) using the discrete e xcitation spectrum\nEjis that it involves the energy-conserving delta functions. This difficu lty can be overcome\nby replacing each delta function by a function with a finite width. In th is paper, we use the\nfollowing replacement:\nδ(¯hω+Ei−Ej)→1\n2∆Θ/parenleftBig/vextendsingle/vextendsingle/vextendsingle∆−¯hω+Ei−Ej/vextendsingle/vextendsingle/vextendsingle/parenrightBig\n. (15)\nThe width factor ∆ is somewhat arbitrary, and the result for γLwill vary with ∆. In Fig. 4,\nFIG. 4: The variation of the Landau damping rate γL(arbitrary unit) with the width factor ∆\nfor lattice depth s= 1.0 and temperature T= 120 nK. To capture the ∆ dependence of γLwhile\nsaving computational time, we take the matrix element Aαα′\nijas a constant in this figure.\nwe plot a typical behavior of the ∆ dependence of γL. One can see that the variation of\n8γLis very weak when ∆ /¯hωzlies between 1 ×10−4and 3×10−4. The same behaviors are\nalso found for other temperatures and lattice heights. We thus ta ke ∆/¯hωz= 2.5×10−4in\ncalculating the damping rate γL.\nFIG. 5: The damping rate as a function of lattice depth swith a fixed temperature T= 120nK.\nFIG. 6: The damping rate as a function of temperature with a fix ed lattice depth s= 1.0\nInFig.5,weplotthedampingrateasafunctionoflatticedepth swithafixedtemperature\nT= 120nK calculated from Eq. (13). We see that the magnitude of the damping rate is\n∼1s−1, which is in reasonable agreement with the experimental data [4]. We also see\nthat the damping rate increases with increasing lattice depth, which is consistent with the\nexperimental result [4]. This increase of the damping rate can be att ributed to the increase\nof the number of the elementary excitations because of a reductio n of the excitation energy\nwith increasing lattice depth.\nWe next investigate the temperature dependence of the damping r ate. In Fig. 6, we plot\nthedampingrateasafunctionoftemperaturewithafixedlatticede pths= 1.0. Wefindthat\n9damping rate decreases rapidly with decreasing temperature. This is due to the reduction\nof thermal excitations with decreasing temperature. Because of this strong temperature\ndependence of the damping rate, it is very difficult to make a quantita tive comparison with\nthe experimental data without precise knowledge of the experimen tal temperature.\nHerewecommentontheeffectofadiabaticloadingofanopticallattic eonthetemperature\nof a gas. The usual path for preparing for condensed Bose gases in an optical lattice consists\nof first forming a ultracold bosons in a weak magnetic trap, to which a 1D lattice potential\nis adiabatically applied by slowly rating up the light field intensity. In this c ase, the initial\nand final temperature are not usually equal since the energy spec trum changes during lattice\nloading [16]. We used the entropy-temperature curves to consider the effect of the adiabatic\nloading into a lattice. For the initial ( s=0) temperature T=120nK, the temperature shift in\nthe final state s= 1.6 is only about 2nK. Thus, one can ignore the effect of the adiabatic\nloading in shallow lattice regime.\nWe note that the radial excitations are important in Landau damping . If we calculated\nthe the damping rate ignoring the radial thermal excitations, the d amping rate would be\norder of magnitude smaller. This means that the radial thermal exc itations make significant\ncontributions to the Landau damping.\nIV. CONCLUSION\nIn this paper, we studied the Landau damping of dipole oscillations of B ose-condensed\ngases in a combined potential of highly-elongated harmonic trap and 1D optical lattice, with\nexplicitly including the effect of the radial excitations. While we treate d the condensate\nwavefunction only with the lowest radial mode, we took into account the radial excitations\nfor thermal cloud.\nFirst, we studied equilibrium properties of Bose-condensed gases in a combined harmonic\ntrap and 1D optical lattice potentials. We have presented a detailed calculation of the\ncondensate fraction in a 1D optical lattice at finite temperatures. We find the negative shift\noftheexcitationenergieswithincreasinglatticedepth. Weobtainth edipole-modefrequency\nas a function of the lattice depth. The negative shift of the dipole-m ode frequency was also\nobserved in the experiment.\nSecond, we calculated Landau damping rate of dipole-modes with var ying lattice depth\n10and temperature. Result for the damping rate is consistent with ex perimented data. There-\nfore, the experimentally observed damping can be understood as L andau damping. We also\nshowed that the radial thermal excitations are important in both e quilibrium condensate\nfractions and Landau damping rate.\nACKNOWLEDGMENTS\nThis research was support by Academic Frontier Project (2005) o f MEXT.\n[1] For a recent review, see for example O. Morsch and M. Obert haler Rev. Mod. Phys. 78, 179\n(2006)\n[2]S. Burger et al., Phys. Rev. Lett. 86, 4447(2001)\n[3]F. S. Cataliotti et al., Science 293, 843(2001)\n[4]F. Ferlaino et al., Phys. Rev. A 66, 011604(2002)\n[5]M. Kr¨ amer et al., Eur. Phys. J. D 27, 247 (2003)\n[6]E. Zaremba, T. Nikuni and A. Griffin, J. Low. Temp. Phys. 116277 (1999)\n[7]S.Giorgini, Phys. Rev. A 57, 2949(1998)\n[8]Pitaevskii,L.P.and Stringari,S. Phys. Rev. Lett. 235, 398 (1997)\n[9]M. Guilleumas and L. P. Pitaevskii, Phys. Rev. A 61013602 (2000)\n[10]B. Jackson and E. Zaremba New J. Phys. 588(2003)\n[11]S. Tsuchiya and A. Griffin, Phys. Rev. A 70, 023611(2004)\n[12]A. M. Rey, et al., Phys. Rev. A 72, 033616(2005)\n[13]B. G. Wild, et al., Phys. Rev. A 73, 023604(2006)\n[14]E. Arahata and T. Nikuni J. Low Temp. Phys. 148, 345 (2007)\n[15]A. Griffin, Phys. Rev. B 53, 9341(1995)\n[16]P.B.Blakie and J.V.Porto, Phys. Rev. A 69, 013603(2004)\n[17]C. J. Pethich and H. Smith, Bose-Einstein Condensation in Dilu te Bose Gass (UNIVERSITY\nPRESS CAMBRIDGE)\n11" }, { "title": "0709.0539v2.Causal_sets_and_conservation_laws_in_tests_of_Lorentz_symmetry.pdf", "content": "arXiv:0709.0539v2 [gr-qc] 24 Jun 2008Causal sets and conservation laws in tests of Lorentz symmet ry\nDavid Mattingly\nUniversity of New Hampshire∗\nMany of the most important astrophysical tests of Lorentz sy mmetry also assume that energy-\nmomentum of the observed particles is exactly conserved. In the causal set approach to quantum\ngravity a particular kind of Lorentz symmetry holds but ener gy-momentum conservation may be\nviolated. We show that incorrectly assuming exact conserva tion can give rise to a spurious signal\nof Lorentz symmetry violation for a causal set. However, the size of this spurious signal is much\nsmaller than can be currently detected and hence astrophysi cal Lorentz symmetry tests as currently\nperformed are safe from causal set induced violations of ene rgy-momentum conservation.\nI. INTRODUCTION\nThe search for a complete and experimentally veri-\nfied theory of quantum quantum gravity is one of the\nmost important open questions in physics today. Unfor-\ntunately, despite the efforts of numerous eminent physi-\ncists, we do not yet have a theoretically complete model\nfor quantum gravity. Since the natural energy scale of\nquantum gravity is the Planck scale ( MP= 1019GeV)\nit is also extremely difficult to perform direct experi-\nments that support one candidate model for quantum\ngravity over another. Fortunately, various ideas about\nquantum gravity have suggested that the defining sym-\nmetry of special relativity, Lorentz symmetry, is not an\nexact symmetry even at low energies. In string theory\nthis can occur perhaps by tensor VEV’s [1] or noncritical\nstrings [2, 3] while in loop quantum gravity the ultimate\nfate of Lorentz symmetry and how it’s implemented is an\nopenquestion(c.f. [4, 5, 6, 7]). Emergentspacetimemod-\nels, for example analog spacetime [8], contain Lorentz\nsymmetry violation (LV) at high energies. Canonical\nnoncommutative field theories also contain Lorentz vi-\nolation [9, 10, 11] although they are invariant under the\ntwisted Poincare group [12]. Note that there are also\nnoncommutative models where Lorentz invariance is de-\nformed rather than explicitly violated (see [13, 14] and\nreferences therein).\nOf course any violation of Lorentz invariance must be\nvery, very small and therefore for any model with LV\nthere is a severe hierarchy problem. For example, in an\neffective field theory context, mass dimension three op-\nerators that generate Lorentz symmetry violation must\nbe less than 10−32GeV while the dimension four oper-\nators are constrained at the level of 10−28[15]. These\nextreme bounds mean that any Lorentz violating theory\nmust answer the question of why we have such good ap-\nproximate low energy Lorentz invariance. One could fine\ntune operators, but this is unnatural. In an attempt to\navoidthis, manyauthorshavelookedathigherdimension\noperators where the magnitude of the operator is sup-\npressed by MPand so is “hidden”. However, standard\n∗Electronic address: davidmmattingly@comcast.neteffective field theory arguments require that the higher\ndimension operators dimensionally transmute into lower\ndimension ones [16]. This means that dimension five and\nsix operators are constrained at the same order as the\ndimension three and four operators and so LV alone is\nextremely unlikely. It is possible to naturally suppress\nthe renormalizable operators by introducing new physics\nat scales just above that currently accessible by particle\ndetectors. For example, imposing supersymmetry [17]\nas well as Lorentz violation stabilizes the generation of\nlower dimension operators at the SUSY breaking scale.\nIf the SUSY breaking scale is given by mSUSY, and the\noriginal higher dimension operators are suppressed by\nMP, then SUSY will naturally suppress the dimension\nthree and four operators to the level of m2\nSUSY/MPand\nm2\nSUSY/M2\nPrespectively. For mSUSY<100 TeV, this\ngives a suppression to a level of approximately 10−9GeV\nand 10−28. The dimension three operators are still in\nconflictwithobservationalbounds, butthesecanbeelim-\ninated by further imposing CPT invariance. This would\nmake a model with low energy SUSY and CPT invari-\nance and LV sourced by Planck scale quantum gravity\nviable, but requires significant new low energy physics as\nwell (in this case SUSY and the assumption of CPT).\nThere is, of course, nothing a priori wrong with intro-\nducing new low energy physics and SUSY certainly has\na number of other compelling features unrelated to LV.\nIf we restrict ourselves just to the question of LV, how-\never, then Occam’s razor suggests that adding twonew\nphysical ideas is disfavored. Hence, instead of either fine\ntuning or imposing new low energy physics in addition\nto LV, it would be better perhaps to have a quantum\ngravity theory that respects Lorentz invariance exactly.\nThe discrete model we will discuss for the rest of this\npaper, causal sets, is singled out among discrete models\nas it is constructed to be Lorentz invariant by definition.\nRecently, Sorkin, Bombelli, and Henson [18] proved that\na causal set is Lorentz invariant for an abstract operator\nthat represents a measurement of a preferred frame for a\nsection of the causal set (the proof is that such operators\ncannot exist). In this workwe arguethat this operator D\ndoes not quite reflect the way that we currently analyze\nmany real Lorentz violating experiments and that such\nexperiments (in particular astrophysical tests) may the-\noretically show “spurious” Lorentz violating effects if the2\nunderlying spacetime is a causal set. However, we also\nshow that any effect is much smaller than our current\nexperimental sensitivities and can be safely ignored.\nThe essence of our argument is that swerves, a hypo-\nthetical effect on particle propagation in causal set the-\nory [19], can mimic a signal of LV in time of flight or\nthreshold astrophysics experiments where the expected\nflux of incoming high energy particles is very low. This\nmay seem strange, as causal sets are supposed to not\ngive any LV signal. However, swerves violate energy-\nmomentum conservation, which we usually assume holds\nexactly in the vacuum when we analyze LV experiments.\nThis mismatch between theory and assumption leads to\na spurious LV signal. Violations of energy-momentum\nconservation are tightly constrained from cosmology and\nthe constraints are strong enough that this effect is irrel-\nevant at current sensitivities in LV experiments but may\nnot be in the future. Furthermore, other ideas about\nquantum gravity have raised the spectre of a small vi-\nolation of energy-momentum conservation for particles\ntraveling in the vacuum(c.f [20, 21, 22]), so exploring\nhow this changes LV experiments is required if we are to\nanalyze the experiments properly.\nAn added benefit to this study is that, for a population\nof particles, the violation of energy-momentum conserva-\ntion predicted from causal sets satisfies a low energy dif-\nfusion equation in energy-momentum space. This equa-\ntion is unique, and therefore our result implies that we\ncan rather generically neglect the effects of any stochas-\ntic, Lorentzinvariantviolationofenergy-momentumcon-\nservation in Lorentz symmetry tests, whether or not we\nbelievein causalsets. Hencewhile causalsetsarethe mo-\ntivation, the results apply more broadly (although causal\nsets are the only model at this point that hypothesizes\nsuch an effect). To put it differently, assuming energy-\nmomentum conservation is not a priori warranted when\nsearching for quantum gravity induced LV. However, the\nconstraints on Lorentz invariant conservation violation\nare so tight that we can safely ignore it at our current\nlevel of sensitivity in LV tests, independent of whatever\nthe fundamental theory turns out to be.\nThroughout the following discussion we choose metric\nsignature −2 and units such that /planckover2pi1=c= 1.\nII. CAUSAL SETS AND THE POINCARE\nGROUP\nBefore we can discuss the fate of the Poincare group\nand conservation laws in the context of causal sets, we\nneed to know what exactly a causal set is. At its basic\nlevel, a causal set is a partially ordered set (a poset),\nconsisting of ‘points’ x,y,...and relations x≺ywhich\nencode causal ordering, i.e. x≺yimpliesxis to the\npast ofy. The ordering obeys two other rules, x≺y,\ny≺z⇒x≺zandx/ne}ationslash≺x. The first is simple transitivity\nsuch that if xis in the past of yandyis in the past\nofz,xhad better also be in the past of z. The othercondition forbids closed causal curves. Causal sets are\nusually considered to be finite, so for any ordered pair\n{x,y}withx≺y, there are a finite number of points z\nthat satisfy x≺zandz≺y. While it is certainly not\nnecessaryfor a spacetime to be present for a causal set to\nbe defined, a useful picture is to think of a causal set as\na random lattice “sprinkled” in a Lorentzian spacetime\nwhere there is an average of one point per every volume\nV0(which will be assumed here to be a Planck volume\nL4\nPl). The causal set sprinkling therefore approximates\nM. For more information on causal sets, see [23] and\nreferences therein.\nIt has been argued that an individual causal set is\nLorentz invariant [18], even locally, in a particular sense:\nthere is no experiment that can assign an intrinsic vio-\nlation of Lorentz invariance to a causal set at any point.\nHere “intrinsic violation” means a frame that depends\nonly on the sprinkling. We now briefly outline the proof\nin [18], restricting ourselves to the simplest type of LV -\na preferred frame. The question is then, can a causal set\ndefine such a frame? Certainly a regular discrete struc-\nture for spacetime does, but a causal set is a random\nstructure where the points are distributed in spacetime\nfrom a Lorentz invariant probability distribution. A pre-\nferred frame is specified by the existence, in vacuum, of a\nunit future pointing timelike vector field uaeverywhere\non the spacetime manifold M. Consider now a subset\nΩ of a causal set sprinkling that approximates M. An\nexperiment that assigns a LV preferred frame can be rep-\nresented as an operator Dthat maps Ω to the unit hy-\nperboloid of future timelike vectors H, i.e.ua=D◦Ω.\nuatransforms as an ordinary four vector under a Lorentz\ntransform Λ. Dmust not posses any intrinsic preferred\ndirection(to ensurethat anyLVcomesfrom Ωitself) and\nthereforeDmust commute with Λ so that DΛΩ = Λua.\nThe proof is then that no such operator Dcan exist and\nthere is no experiment that can assign a preferred frame\nto a local patch Ω of a causal set. The implication is that\nno LV experiment can ever assign a preferred frame to a\ncausal set intrinsically.\nThis result, on its face, is incompatible with another\nhypothesized effect of causal sets, a violation of transla-\ntion invariance due to the so-called “swerve” effect [19].\nThe swerve effect, which we discuss in more detail in\nthe next section, manifests itself at low energies via a\nLorentz invariant diffusion equation in momentum space.\nAn initial collection of particles with some momentum\ndistribution ρ(p) therefore evolves into a different state\nover (proper) time, which obviously violates translation\ninvariance. The immediate question is then, how does\nthis not directly yield LV? After all, translation violation\nallows us to immediately define a frame by taking, for\nexample, the gradient of the evolving quantity as the op-\neratorD. The answer is that the frame so defined is not\nintrinsic tothecausalset, butalsoinvolvestheinitialdis-\ntributionρ(p). In particular, if ρ(p) is a Lorentz invari-\nant function (which is unrealistic physically but useful\nfor this discussion) it is preserved by the diffusion equa-3\ntion. Hence there is neither a violation of translation or\nLorentz invariance in this case. If, instead, ρ(p) is not\nLorentz invariant then there can be translation invari-\nance violation, but both it and the corresponding LV are\nfunctions ofthe causalset andthe initial LVdistribution.\nThere is a “signal” of LV, but it is not the intrinsic LV\nwhich is forbidden by the proof outlined above because\nour operator Dcontains a preferred direction, which vi-\nolates one of the assumptions in the proof.\nThe difference between these two types of LV signal -\na real signal intrinsic to the underlying spacetime vs. a\nspurious signal due to the combination of the preferred\nframe of an experiment with another property of space-\ntime (in this case swerves) is a possibility which has not\narisen before in analysis of LV experiments. Since all of\nour observations and experiments that search for LV are\nnot LI in and of themselves (as they are local and hap-\npen in a particular frame) we must consider it. Hence\nthe question of LV in causal sets for a practical experi-\nment is not quite as straightforward as presented in [18].\nThe particular example we examine in this paper is how\ntheassumption of translation invariance, which is cus-\ntomary in LV experiments, can generate a spurious LV\nsignal from a causal set.\nTranslation invariance is a bad assumption due to the\nswerve effect, as we a) single out a preferred Killing vec-\ntor which is not an intrinsic property of the causal set\nitself and b) neglect momentum space diffusion due to\nswerves. The most common astrophysical tests of LV\nare time of flight tests, where we compare the arrival\ntime of two high energy particles, and anomalous scat-\ntering/decay phenomena, where we look for a modifica-\ntion to the scattering amplitude caused by LV dispersion\nrelations for the in/out free particle states. As will be-\ncome clear below, when analyzing these phenomena we\nmustassume something about translationinvariance and\nconservation of energy and momentum in order to put\nconstraints on any possible LV. Usually, translation in-\nvariance has been assumed to hold exactly in flat space,\nasthis is most consistentwith astraightforwardfield the-\noretical approach if the background LV tensor fields are\nconstant.1With causal sets, this assumption isn’t right\nand we must verify that a LV signal is really due to LV\nand not the swerve effect.\n1There are two major exceptions to this, the doubly special re la-\ntivity program (see [13] for an introduction) which deforms both\nthe translational and Lorentz subgroups of the Poincare gro up,\nand spacetime foam ideas which couple LV to a fluctuating dis-\npersion term [24, 25]. These approaches, however, have a sig nif-\nicant a priori modification of bothLorentz invariance and trans-\nlation invariance which makes them distinct from the causal set\napproach.A. Swerves\nConsider the intuitive picture for swerves in [19], that\nof a classical particle propagating on a random space-\ntime lattice with mass mand velocity v. The particle\nis constrained to move from point to point, which poses\na problem as generically there is no probability of a lat-\ntice point lying on the future worldline of the particle.\nThe particle must therefore “swerve” slightly so that it\nremains on the lattice and slightly change its velocity. A\nchange in velocity is equivalent to the particle moving\nto a different point on its mass shell, which is assumed\nto be unchanged since causal sets are Lorentz invariant.\nThe net result of swerving is that a collection of parti-\ncles initially with an energy-momentum distribution ρ(p)\nwill diffuse in momentum space along their mass shell\naccording to the unique Lorentz invariant diffusion equa-\ntion [19, 26],\n∂ρ\n∂τ=k∇2\nPρ−m−1pµ∂µρ. (1)\nHerekis the diffusion constant, ∇2\nPis the Laplacian\nin momentum space on the mass shell of the particle, τ\nis the proper time, ∂µis an ordinary spacetime deriva-\ntive, andmis the mass. While the underlying classi-\ncal picture is almost certainly incorrect,(1) is the unique\nLorentz invariant diffusion equation. Therefore if there\nis any type of stochastic violation of Poincare invariance\ndue to a random discrete structure underlying spacetime\na la causal sets, at lowest order it should be described by\n(1). There is a very tight limit on kfor neutrinos from\ncosmology[27], k<10−61GeV3, which comes from limits\non the amount relic neutrinos can contribute to hot dark\nmatter (and hence how much energy they can gain from\nswerves). While theoretically each particle species could\nhave a different k, it would be rather unnatural if they\nwere too far apart, especially as the source of the diffu-\nsion is supposed the same discrete spacetime structure.\nTherefore we shall take k<10−61GeV3to be our rough\nconstraint for all particles. As it turns out, any value\nofkeven close to 10−61GeV3makes energy-momentum\nconservation violation irrelevant for LV searches, so this\nis a perfectly safe assumption.\nEnergy is bounded below and so a particle initially at\nrest will increase its kinetic energy due to swerves. Since\nkis solow, we canmake asimplifying assumptionfor any\ncollection of particles that are not of cosmological age.\nIn the initial rest frame of the particle the swerves can\nfirst be treated in the non-relativistic limit for a certain\nperiod of time that depends on k. This is obvious as over\ntime energy is being added to the particle via swerves,\nbut as long as the total energy is less than the rest mass\nthe appropriate limit is the non-relativistic one. In the\nnon-relativistic limit, (1) simplifies to\n∂ρ\n∂t=k∇2\nPρ (2)\nwheretis now the coordinate time and ∇2\nPis the stan-4\ndard Laplacian operator on momentum space R3. The\nsolution for a collection of particles all initially at rest\nhas been derived in [27] and is given by\nρ(p) = (4πkt)−3/2e−p2\n4kt (3)\nwhich is the thermal distribution for a non-relativistic\ngas at temperature T= 2kt/m.\nWe can now ask how long a collection of initially cold\nparticles must exist for the non-relativistic approxima-\ntion to break down. This happens when the temperature\nis roughly equal to the mass, which gives t≈m2/(2k).\nFor a neutrino of mass 10−1eV, the approximation\nbreaks down after 1017seconds, while for electrons and\nprotons the approximation is always good as the break-\ndown time is far longer than the age of the universe.\nAfter a population becomes relativistic it will still gain\nenergybut notasquicklysincethepropertime isshorter.\nTherefore we know that the energy gained (per neutrino)\nby a population of initially cold neutrinos over the life-\ntime of the universe (also approximately 1017seconds) is\nno more than about the rest massof the neutrino and the\nenergy gain for other species is far less. This neglects the\neffects of cooling due to expansion, etc. which are dealt\nwith in [27], however we can ignore these secondary ef-\nfects for our purposes. Now consider a population of\nparticles with a high gamma factor γthat have traveled\nto earth from a source at one Gpc. The lifetime of the\nparticle in our frame Ois 109years = 1016seconds. How-\never, the time in the (initial) rest frame O′of the particle\nis much shorter, 1016/γseconds. Since any particle can\ngain no more than mνin energy over 1017seconds and\nthe propagation time in O′is much shorter, very little\nenergy is gained in O′due to swerves. Hence in O′the\nparticles are all still very non-relativistic once we include\nthe swerveeffect aslongas γ≫1and the non-relativistic\napproximation is valid for their entire lifetime.\nOf particular interest is the averagedeviation from the\ninitial energy Eiof a particle. If we define ∆ E/Ei=\n|(Ef−Ei)|/Ei, whereEfis the energy at time of mea-\nsurement, then from the discussion above we know that\n∆E/Eiis at least less than γ−1(on average) for any\nspecies. We can see this more explicitly by noting from\nabove that for a neutrino with a lifetime of the age of\nthe universe at rest in our frame, the total energy gain is\nat best the mass of the neutrino. If instead the neutrino\nis boosted with respect to our frame, the proper time is\nreduced by a factor of γ−1and so the total energy gain\nis reduced also by γ−1from its initial energy Ei=γm.\nThereforeforneutrinos the ratio∆ E/Ei≤γ−1. ForTeV\nand above astrophysical neutrinos, which we are primar-\nily interested in, ∆ E/Ei≤γ−1= 10−13at worst. The\nswerverateisslowerforotherspecies, sothisboundholds\nforthemaswell. Thisgainisverysmall,howeverLVtests\ncan be sensitive to fractional changes in energy of order\n10−28so it is not a priori obvious that swerves are irrele-\nvant. We now turn to how this diffusion affects these LV\ntests.III. SWERVES AND ASTROPHYSICAL TESTS\nOF LV\nA. Time of flight\nTime offlight tests areperhaps the simplest type ofLV\nexperiment. In these experiments one looks for delays in\nthe arrival time of high energy particles from distant as-\ntrophysical events. A time of flight experiment compares\nthe arrival time of at least two high energy particles and\nthe time delay between arrivals can be caused by three\ndistinct effects. The first is source effects - the particles\nare not produced at the same time or location in the as-\ntrophysical event. The source delay can be quite long, in\nthe case of neutrinos from gamma ray bursts the delay\ntime of a neutrino associated with the burst can be days.\nThe second type of delay is detector response. These de-\nlaysareusuallysmallandknownandwewill notconsider\nthem further. Finally, an unexplained time delay is usu-\nally considered evidence that the speed of propagation\nof the high energy particles is different than the speed\nof light. This is the “interesting” signal of a violation of\nLorentz invariance.\n1. Time of flight in field theory\nIt will be useful to discuss in detail how these experi-\nments work in a veryconcrete and established framework\nfirst, before we consider the causal set scenario, so let us\nanalyze time of flight in a field theory context first. LV\noccurs when fields are coupled in vacuum to a non-zero\ntensor field other than the metric. If there exists a pre-\nferred frame in nature, the fields couple to a unit time-\nlike vectoruawhich describes the preferred frame. There\nare many ways a field could couple in both the matter\nand gravitational sectors [28, 29, 30, 31], for a review of\nboth the renormalizable and non-renormalizable opera-\ntors see [32]. The free field mass dimension five couplings\nand below are already tightly constrained, so we consider\nhere an unconstrained operator - that of a dimension six\nCPT even operator. While terms like this may be able\nto be tested in ultrahigh energy neutrino observatories in\nthe future and hence are intrinsically relevant, we choose\nthis term for another reason: if the swerve effect is ir-\nrelevant for tests of this term it is certainly irrelevant\nfor any LV time of flight test we could conceivably per-\nform in the near future. The dimension six CPT even LV\nmodification to the kinetic term for a fermion is\nLf=ψ(i/∂−m)ψ−i\nE2\nPlψ(u·∂)3(u·γ)(αLPL+αRPR)ψ\n(4)\nwherePR,Lare the usual right and left chiral projec-\ntion operators and αR,Lare coefficients. Usually αR,L\nare assumed to be O(1). We choose the operator to be\nsuppressed by the Planck scale EPlas this would be the\nnatural scale if the term was generated by some theory5\nof quantum gravity at EPl.\nThe Hamiltonian corresponding to (4) commutes with\nthe helicity operator, hence the eigenspinors of the modi-\nfied Dirac equation will also be helicity eigenspinors. We\nnow solve the free field equations for the positive fre-\nquency eigenspinor ψ. Assume the eigenspinor is of the\nformψse−ip·xwhereψsis a constant four spinor and\ns=±1 denotes positive and negative helicity. Then the\nDirac equation becomes the matrix equation\n/parenleftBigg\n−m E −sp−α(6)\nRE3\nE2\nPl\nE+sp−α(6)\nLE3\nE2\nPl−m/parenrightBigg\nψs= 0 (5)\nThe dispersion relation, given by the determinant of (5),\nis\nE2−(α(6)\nRE3)(E+sp)\n−(α(6)\nLE3)(E−sp) =p2+m2(6)\nwhere we have dropped terms quadratic in α(6)\nR,Las they\nare small relative to the first order corrections for those\nterms.\nAtE≫m, as appropriate for high energy astrophys-\nical particles, the helicity states are almost chiral, with\nmixing due solely to the particle mass. Note also that at\nenergiesE >>m , we can replace Ebypat lowest order,\nwhich yields the approximate dispersion relation\nE2=p2+m2+2αR,Lp4\nE2\nPl. (7)\nPositive coefficients correspond to superluminal propa-\ngation, i.e. v=∂E/∂p > 1, while negative coefficients\ngive subluminal propagation. In either case, the speed\nof astrophysical particles does not asymptote to cas the\nenergy increases.\nThe group velocity ∂E/∂pis\nv= 1−m2\n2p2+3αR,Lp2\nE2\nPl= 1+∆v. (8)\nFor a source at distance dfrom earth the arrival time\ntaof a high energy particle is ta=d/v. The difference\n∆tLVbetween a light pulse emitted from the source at\nthe same time as the particle is\n∆tLV=tlight−ta=d−d\nv=d(v−1\nv)≈d∆v(9)\n=d/parenleftbig\n−m2\n2E2+3αR,LE2\nE2\nPl/parenrightbig\nwhere we have replaced pbyEin the high energy limit.\nTheαR,Lterm in (9) grows with energy. In order to\nestablish the best possible constraints we therefore need\nto look at the highest energy particles. The best chance\nwehaveinatimeofflightexperimenttoseeLVatthis or-\nder is in the comparisonofthe arrivaltime ofhigh energy\nneutrinos produced by GRB’s with the prompt emissionarrival. The flux at these energies is very low which has\nboth positive and negative ramifications. On the neg-\native side we require large detectors like ICECUBE to\nsee any appreciable flux of ultra high energy GRB neu-\ntrinos. However the background flux is also very low.\nRecently, it has been proposed by Jacob and Piran [33]\nthat since the background is so low even the detection of\na single neutrino event days or weeks after a GRB can be\nassociated with the GRB and used to bound αR,Lin a\ntime of flight experiment. While this approach has other\nproblems, primarily long source delays [34] due to the\nGRB fireball mechanics, it raises an interesting question\nwith regards to causal sets - what happens in the swerve\npicture when there are very, very low statistics?\n2. Time of flight with swerves\nIn time of flight experiments with large fluxes, swerves\naren’t a problem. For a strong multiparticle signal the\naveragearrival time is still that as predicted by special\nrelativity. However, for a single particle signal with mea-\nsured energy Ethere is no concept of averaging and the\narrival time will not be that predicted by special rela-\ntivity. The reason is simple - during propagation the\nparticle’s energy is not Eand the particle’s velocity is\nnotv= 1−m2/(2E2). Therefore to conclusively ascribe\na time delay to a LV dispersion as in (9) we need to make\nsure the same delay cannot be due to swerves.\nWe can overestimate the time of flight delay for a typi-\ncal particle compared to special relativity by considering\na particle propagating with energy Ef+∆E, whereEf\nis the measured (final) energy and ∆ Eis the average\ndeviation for a single particle introduced previously. It\nis an overestimate since we apply the energy difference\nover the entire propagation of the particle. The disper-\nsion relation is simply the relativistic dispersion relation,\nso ∆tSis\n∆tS=d/parenleftbigg\n−m2\n2(Ef+∆E)2/parenrightbigg\n=d/parenleftBigg\n−m2\n2E2\nf(1+∆E/Ef)2/parenrightBigg\n(10)\n∆E/Ef<γ−1≪1 and we therefore have\n∆tS=d/parenleftBigg\n−m2\n2E2\nf(1−2∆E\nEf)/parenrightBigg\n. (11)\nComparing(11)with (9)weseethatanexperimentsensi-\ntive to LV at our chosen order is also sensitive to swerves\nif\n3αR,LE2\nf\nE2\nPl≈m2\nE2\nf∆E\nEf. (12)\nRewriting this equation for ∆ E/Efwe have\n∆E\nEf= 3αR,LE4\nf\nm2E2\nPl. (13)6\nThere is no intrinsic size to the LV term in (13) and\ndepending on how accurate a LV time of flight measure-\nment is, it could theoretically also probe swerves. How-\never, the duration of a long GRB can be of order 1000\nseconds and it is unclear when in the burst emission the\nneutrinos will occur. Therefore there is an intrinsic un-\nknown source delay of at least 1000 seconds2. Any sig-\nnificant time delay must be greater than this value and\nhence there is a lower limit on a meaningful ∆ t(swerve\nor LV induced). For a GRB with this value of ∆ tand a\ndistance of 1 Gpc, ∆ t/d≈10−14seconds. With (9) this\nestablishes a lower limit that the LV dispersion term of\n|αR,LE2\nf\nE2\nPl| ≥10−14(14)\nwhich must be satisfied if we are to see anything mean-\ningful in a time of flight GRB experiment. If we are con-\nservative and take our high energy neutrinos to be above\n1 TeV (actual proposed energies are much higher) and a\nneutrino mass of approximately 0.1 eV, the gamma fac-\ntor is 1013. From (13), ∆ E/Effrom swerves must then\nbe greater than 1012, which is 25 orders of magnitude\nlargerthanthe upper limit forswervingneutrinos. Hence\nswervesare completely and totally irrelevant. This result\ncan easily be generalized to other forms of LV dispersion\nand in none are astrophysical time of flight experiments\nsensitive to swerves by many orders of magnitude.\nB. Anomalous particle interactions\nAnomalous particle interactions are much more sensi-\ntive tests ofLV than time offlight tests and alsoof course\nrequire assumptions about energy-momentum conserva-\ntion. Therefore they will also be sensitive to the swerve\neffect at some level. There are two types of anomalous\ninteractions. The first type is when the interaction oc-\ncurs only in the LV model. The second type is when the\ninteraction begins to occur at a certain energy and this\nenergy is different for Lorentz invariant versus LV mod-\nels. We deal with each type of interaction separately as\nthe effect of swerves is different.3\n2The actual value is of course dependent on the exact mechanic s\nof the GRB and can be shorter. We use this value for illustrati ve\npurposes as an order of magnitude estimate.\n3We are not considering particle creation due to the time depe n-\ndence of the spectrum of an initial flux of particles due to swe rves\nor how to correctly define initial/final states in a model with out\nasymptotic translation invariance. These questions must a lso be\nanswered in regards to the swerve effect but are outside the sc ope\nof this paper.1. New particle interactions\nConsider a proton with dispersion relation like that\ngiven in (7). If αR,L>0 then at high energies protons\nbecome unstable and emit photons in what is known as\nthe “vacuum Cerenkov effect”. They rapidly lose energy\nvia this process and hence the existence of ultra high en-\nergy cosmic ray protons (for which this process cannot\nbe occurring) limits how positive αR,Lcan be. This is\nan example of a test that uses a particle interaction com-\npletely absent in usual Lorentz invariant physics to limit\npossible LV dispersion relations. The key to these tests\nis the energy-momentum conservation equations, which\ntell us how much parameter space is available for the\nreaction - zero in the Lorentz invariant case and non-\nzero with LV. Naively then, it seems reasonable that a\nfluctuating energy-momentum of the initial and/or final\nstates may allow for new reactions to also occur. We now\nshow that in the case of causal sets, this is not true. If\nwe consider just a swerve induced change in the energy-\nmomentum of the initial and final states then if a reac-\ntion does not occurin straightforwardspecial relativityit\ndoes not occur in causal sets. Note that in the following\ndiscussion we have implicitly assumed that the swerve\neffect for massless particles yields diffusion in a null cone\nin energy-momentum space (the m→0 limit of a mass\nshell).\nThe idea is very simple. Let us consider a generic\nmulti-particle reaction i1+i2+...→f1+f2+...whereia\nare the incoming particles and fbare the outgoing parti-\ncles. If the reaction does not occur in Lorentz invariant\nphysics it means that there is no solution to the conser-\nvation equation\npµ\n1+pµ\n2+...=qµ\n1+qµ\n2+... (15)\nwherepµ\na(theincoming4-momenta)and qµ\nb(theoutgoing\n4-momenta) are subject to the on-shell constraints p2\na=\nm2\na,q2\nb=m2\nb. In a LV theory the on-shell constraints\nchange, in causal sets they do not. In a causal set, the\nincoming and outgoing momenta are, however, modified\nby a fluctuation term ∆ pa,∆qbas the particle will swerve\nduringthe courseofthe interaction. We thereforerewrite\nthe conservation equation in the causal set approach as\npµ\n1+∆pµ\n1+pµ\n2+∆pµ\n2+...=qµ\n1+∆qµ\n1+qµ\n2+∆qµ\n2...(16)\nsubject to the on-shell constraints ( pa+ ∆pa)2=\nm2\na,(qb+ ∆qb)2=m2\nbsince the fluctuations must keep\nall particles on-shell. This though is just a relabeling of\nmomenta and doesn’t change any physics - we can define\nnew momenta ¯ pµ\na=pµ\na+∆pµ\naetc. and the conservation\nand constraint equations take the exact same form as be-\nfore. Therefore there is still no solution and the reaction\ndoesn’t happen with swerves either.\nThere is a caveat to the above argument. In keeping\nwith other LV threshold tests, where one looks at only\ninitial and final states, we have not considered the ef-\nfect of swerves in the interaction region. This is more7\ndangerous for causal sets, as field theory on causal sets\nis not well developed enough to know what the swerve\neffect might do to virtual states that are not necessarily\non-shell. Hence this conclusion can not be considered ab-\nsolutely concrete until we understand more of quantum\nfield theory on a causal set. However, since no LV reac-\ntions have been seen to date it is likely that causal set\nQFT does respect LI to a very good approximation even\nwhen quantum effects are taken into account.\n2. Modification of existing energy thresholds\nThe situation is different for reactions where there is\na Lorentz invariant solution. Here, the swerve effect can\nchange the energy the reaction occurs at, although the\nshift is tiny. Let us take a specific reaction, pion pro-\nduction by proton-photon scattering, p+γ→p+π0.\nThis reaction is important in LV tests as the high en-\nergy cosmic ray spectrum should exhibit a cutoff (the\nGreisen-Zapsetin-Kuzmin cutoff) around 5 ×1019eV due\nto the scattering of cosmic ray protons off the cosmic mi-\ncrowave background. LV tests can shift this cutoff, and\nthe recent confirmation of the GZK cutoff by HiRes [35]\nand the Pierre Auger observatory [36] constrains some\nLV models. Again, we have a limited number of events\n(although with Auger the statistics are getting rapidly\nbetter) and so one might wonder if the random nature of\nswerves can cause a spurious signal. Theoretically this is\ntrue, but the effect of swerves on GZK protons is negli-\ngible, even if we wildly overestimate the length of time\nswerves have to act. A GZK proton has a gamma fac-\ntor of near 1010, which means that its maximum proper\nlifetime is 1017/γ= 107seconds if it was generated very\nearly in the universe. The maximum amount of energy\ntheprotoncangaininitsinitialrestframe,usingourcon-\nstraintfork, is 10−30GeV, whichin turn implies that the\nshift in energy possible for a GZK proton in our frame is\n10−20GeV. The GZK cutoff will hence be broadened byat best the insignificant amount of 10−20GeV and there-\nfore swerves are irrelevant in the GZK reaction. This\ntype of analysis can be done for many different thresh-\nold reactions, for example the scattering of high energy\nTeV photons off the infrared background. In all cases\nthe limits on kare strong enough that swerves have no\nappreciable effect.\nIV. CONCLUSION\nLV searches in astrophysics rely not only on assump-\ntions about the nature of the LV model being explored,\nbut also on energy-momentum conservation of the parti-\ncles involved. In this paper we have argued that causal\nsets, while intrinsically Lorentz invariant, can technically\nintroducespuriousLVsignalsifwedealwith practicalex-\nperiments and make the usual assumption that energy-\nmomentum is conserved due to the swerve effect. How-\never, we have also shown that existing limits on swerves\nimply that any errors introduced are negligible at cur-\nrent experimental sensitivity. A bonus is that the dif-\nfusion equation (1) that describes the swerve effect is\nthe unique low energy Lorentz invariant diffusion equa-\ntion. Any statistical process that respects Lorentzinvari-\nancebut causesfluctuations in energy-momentumshould\nthereforebe described bythis equation andwe canrather\ngenerically conclude that any signal of LV in a time of\nflight or anomalous particle interaction experiment can-\nnot be due instead to a Lorentz invariant modification of\ntranslation invariance.\nV. ACKNOWLEDGEMENTS\nWe thank Stefano Liberati for useful comments on a\ndraft of this paper.\n[1] V. A. Kostelecky and S. Samuel, Phys. Rev. D 39, 683\n(1989).\n[2] J. R. Ellis, N. E. Mavromatos and D. V. Nanopoulos,\narXiv:gr-qc/9909085.\n[3] N. E. Mavromatos, arXiv:0708.2250 [hep-th].\n[4] R. Gambini and J. Pullin, Phys. Rev. D 59, 124021\n(1999) [arXiv:gr-qc/9809038].\n[5] L. Smolin, Nucl. Phys. B 742, 142 (2006)\n[arXiv:hep-th/0501091].\n[6] C. Rovelli and S. Speziale, Phys. Rev. D 67, 064019\n(2003) [arXiv:gr-qc/0205108].\n[7] M. Bojowald, H. A. Morales-Tecotl and H. Sahlmann,\nPhys. Rev. D 71, 084012 (2005) [arXiv:gr-qc/0411101].\n[8] C. Barcelo, S. Liberati and M. Visser, Living Rev. Rel.\n8, 12 (2005) [arXiv:gr-qc/0505065].\n[9] S. M. Carroll, J. A. Harvey, V. A. Kostelecky, C. D. Lane\nand T. Okamoto, Phys. Rev. Lett. 87, 141601 (2001)[arXiv:hep-th/0105082].\n[10] A. Anisimov, T. Banks, M. Dine and M. Graesser, Phys.\nRev. D65, 085032 (2002) [arXiv:hep-ph/0106356].\n[11] I. Hinchliffe, N. Kersting and Y. L. Ma, Int. J. Mod.\nPhys. A 19, 179 (2004) [arXiv:hep-ph/0205040].\n[12] M. Chaichian, P. Presnajder and A. Tureanu, Phys. Rev.\nLett.94, 151602 (2005) [arXiv:hep-th/0409096].\n[13] J. Kowalski-Glikman, Lect. Notes Phys. 669, 131 (2005)\n[arXiv:hep-th/0405273].\n[14] L. Freidel and J. Kowalski-Glikman, arXiv:0710.2886\n[hep-th].\n[15] V. A. Kostelecky and C. D. Lane, Phys. Rev. D 60,\n116010 (1999) [arXiv:hep-ph/9908504].\n[16] J. Collins, A. Perez, D. Sudarsky, L. Urrutia and\nH. Vucetich, Phys. Rev. Lett. 93, 191301 (2004)\n[arXiv:gr-qc/0403053].\n[17] P. A. Bolokhov, S. G. Nibbelink and M. Pospelov, Phys.8\nRev. D72, 015013 (2005) [arXiv:hep-ph/0505029].\n[18] L. Bombelli, J. Henson and R. D. Sorkin,\narXiv:gr-qc/0605006.\n[19] F. Dowker, J. Henson and R. D. Sorkin, Mod. Phys. Lett.\nA19, 1829 (2004) [arXiv:gr-qc/0311055].\n[20] J. R. Ellis, N. E. Mavromatos and D. V. Nanopoulos,\nPhys. Rev. D 63, 024024 (2001) [arXiv:gr-qc/0007044].\n[21] R. Parentani, arXiv:0709.3943 [hep-th].\n[22] G. Amelino-Camelia, Y. J. Ng and H. Vanm Dam, As-\ntropart. Phys. 19, 729 (2003) [arXiv:gr-qc/0204077].\n[23] J. Henson, arXiv:gr-qc/0601121.\n[24] R. Aloisio, A. Galante, A. F. Grillo, S. Liberati, E. Luz io\nand F. Mendez, Phys. Rev. D 74, 085017 (2006)\n[arXiv:gr-qc/0607024].\n[25] R. Aloisio, P. Blasi, A. Galante, P. L. Ghia and\nA. F. Grillo, Astropart. Phys. 19, 127 (2003)\n[arXiv:astro-ph/0205271].\n[26] R. D. Sorkin, Annals Phys. 168, 119 (1986).\n[27] N. Kaloper and D. Mattingly, Phys. Rev. D 74, 106001(2006) [arXiv:astro-ph/0607485].\n[28] D. Colladay and V. A. Kostelecky, Phys. Rev. D 58,\n116002 (1998) [arXiv:hep-ph/9809521].\n[29] R. C. Myers and M. Pospelov, Phys. Rev. Lett. 90,\n211601 (2003) [arXiv:hep-ph/0301124].\n[30] T. Jacobson and D. Mattingly, Phys. Rev. D 64, 024028\n(2001) [arXiv:gr-qc/0007031].\n[31] P. A. Bolokhov and M. Pospelov, arXiv:hep-ph/0703291.\n[32] D. Mattingly, Living Rev. Rel. 8, 5 (2005)\n[arXiv:gr-qc/0502097].\n[33] U. Jacob and T. Piran, Nature Phys. 3, 87 (2007)\n[arXiv:hep-ph/0607145].\n[34] M. C. Gonzalez-Garcia and F. Halzen, JCAP 0702, 008\n(2007) [arXiv:hep-ph/0611359].\n[35] R. Abbasi et al. [HiRes Collaboration],\narXiv:astro-ph/0703099.\n[36] T. Yamamoto [Pierre Auger Collaboration],\narXiv:0707.2638 [astro-ph]." }, { "title": "1010.3990v1.Lorentz_Symmetry_and_Matter_Gravity_Couplings.pdf", "content": "arXiv:1010.3990v1 [hep-ph] 19 Oct 20101\nLORENTZ SYMMETRY AND MATTER-GRAVITY\nCOUPLINGS\nJAY D. TASSON\nDepartment of Physics, Whitman College\nWalla Walla, WA 99362, USA\nE-mail: tassonjd@whitman.edu\nThis proceedings contribution summarizes recent investig ations of Lorentz vi-\nolation in matter-gravity couplings.\n1. Introduction\nIn spite of the many high-sensitivity investigations of Lorentz violat ion1\nperformed in the context of the fermion sector of the minimal Stan dard-\nModel Extension (SME) in Minkowski spacetime,2only about half of the\ncoefficientsforLorentzviolationinthatsectorhavebeeninvestiga tedexper-\nimentally. Reference 3 establishes a methodology for obtaining sens itivities\nto some of these open parameters by considering gravitational co uplings in\nthe fermions sector of the SME,4extending pure gravity work.5Of partic-\nular interest are the aµcoefficients for baryons and charged leptons, which\nare unobservable in principle in Minkowski spacetime, but could be rela -\ntively large due to gravitational countershading.6\nThe first half of Ref. 3 develops the necessary theoretical result s for\nthe analysis of Lorentz violation in matter-gravity couplings. Those results\nare summarized in Sec. 2 below, while Sec. 3 summarizes the experimen tal\npredictions provided in the second half of that work.\n2. Theory\nThe theoretical portion of Ref. 3 addresses a number of useful c onceptual\npoints prior to developing the necessary results for experimental analysis.\nThis includes a discussion of the circumstances under which relevant types\nof Lorentz violation are observable in principle. It turns out that th eaµco-\nefficient,which canbe removedfromthe singlefermiontheoryinMinko wski2\nspacetime via a spinor redefinition cannot typically be removed in the p res-\nenceofgravity.4Thismakesit aninterestingcaseforstudyin theremainder\nof Ref. 3. A coordinate choice that can be used to fix the sector of the the-\nory that defines isotropy is also discussed and ultimately used to tak e the\nphoton sector to have ηµνas the background metric.\nAnother issue is the development of general perturbative techniq ues\nto treat the fluctuations in the coefficient fields in the context of ma tter-\ngravity couplings. Two notions of perturbative order are introduc ed. One,\ndenoted O( m,n), tracks the orders in Lorentz violation and in gravity,\nwhere the first entry represents the order in the coefficients for Lorentz\nviolation and the second represents the order in the metric fluctua tionhµν.\nThe secondary notion of perturbative order, denoted PNO( p), tracks the\npost-newtonianorder.ThegoalofRef. 3istoinvestigatedominan tLorentz-\nviolating implications in matter-gravity couplings, which are at O(1,1).\nReference 3 provides the necessary results to analyze experimen ts at a\nvariety of levels while workingtoward the classical nonrelativistic equ ations\nof motion, which are most relevant for many of the experiments to b e con-\nsidered. Development of the quantum theory of the gravity-matt er system\nprovides the first step. Starting from the field-theoretic action, the rela-\ntivistic quantum mechanicsinthe presenceofgravitationalfluctua tionsand\nLorentzviolationisestablishedafterinvestigatingmethods ofident ifying an\nappropriatehamiltonianinthe presenceofaneffectiveinversevierb einE0\nµ.\nThe explicit form of the relativistic hamiltonian involving all coefficients f or\nLorentz violation in the minimal QED extension is provided. Attention is\nsubsequently specialized to the study of spin-independent Lorent z-violating\neffects, which are governed by the coefficient fields ( aeff)µ,cµνand the met-\nric fluctuation hµν. Analysis then proceeds to the nonrelativistic quantum\nhamiltonian via the standard Foldy-Wouthuysen procedure.\nWhile the quantum mechanics above is useful for analysis of quan-\ntum experiments, most measurements of gravity-matter coupling s are per-\nformed at the classical level. Thus the classical theory7associated with\nthe quantum-mechanical dynamics involving nonzero ( aeff)µ,cµν, andhµν\nis provided at leading order in Lorentz violation both for the case of t he\nfundamental particles appearing in QED and for bodies involving many\nsuch particles. These results enable the derivation of the modified E instein\nequation and the equation for the trajectory of a classical test p article.\nSolving for the trajectory requires knowledge of the coefficient an d metric\nfluctuations. A systematic methodology for calculating this informa tion is\nprovided, and general expressionsfor the coefficient and metric fl uctuations3\nto O(1,1) in terms of various gravitational potentials and the backg round\ncoefficient values ( aeff)µandcµνare obtained. Bumblebee models are con-\nsidered as an illustration of the general results.\n3. Experiments\nA major class of experiments that can achieve sensitivity to coefficie nts\n(aeff)µandcµνinvolve laboratory tests with ordinary neutral matter. Tests\nof this type are analyzed via the PNO(3) lagrangian describing the dy nam-\nics of a test body moving near the surface of the Earth in the prese nce of\nLorentz violation. The analysis reveals that the gravitational forc e acquires\ntiny corrections both along and perpendicular to the usual free-f all trajec-\ntory near the surface of the Earth, and the effective inertial mas s of a test\nbody becomes a direction-dependent quantity. Numerous laborat ory exper-\nimentssensitiveto theseeffects areconsidered.The tests canbe classifiedas\neither gravimeter or Weak Equivalence Principle (WEP) experiments a nd\nas either force-comparison or free-fall experiments for a total of 4 classes.\nFree-fall gravimeter tests monitor the acceleration of freely fallin g ob-\njects and search for the characteristic time dependence associa ted with\nLorentz violation. Falling corner cubes8and matter interferometry9,10pro-\nvide examples of such experiments and are discussed in Ref. 3. Forc e-\ncomparison gravimeter tests using equipment such as supercondu cting\ngravimeters are also studied.11Note that the distinction, force comparison\nversesfree fall, is importantdue to the potentialLorentz-violatin gmisalign-\nment of force and acceleration. Making direct use of the flavor dep endence\nassociated with Lorentz-violating effects implies signals in WEP tests. A\nvariety of free-fall WEP tests are considered including those using falling\ncornercubes,12atom interferometers,10,13tossed masses,14balloondrops,15\ndrop towers,16and sounding rockets,17along with force-comparison WEP\ntests with a torsion pendulum.18For all of the tests considered, the pos-\nsible signals for Lorentz violation are decomposed according to their time\ndependence, and estimates of the attainable sensitivities are obta ined.\nSatellite-based WEP tests,19which offer interesting prospects for im-\nproved sensitivities to Lorentz violation, are also discussed in detail. The\nsignal is decomposed by frequency and estimated sensitivities are o btained.\nThe experimental implications of Lorentz violation in the gravitationa l\ncouplingsofchargedparticles,antimatter,andsecond-andthird -generation\nparticles are also studied. These tests are experimentally challengin g, but\ncan yield sensitivities to Lorentz and CPT violation that are otherwise dif-\nficult or impossible to achieve. Possibilities including charged-particle in -4\nterferometry,20ballistic tests with charged particles,21gravitational experi-\nmentswithantihydrogen,22andsignalsinmuoniumfreefall23arediscussed.\nSimple toymodelsareusedtoillustratesomefeaturesofantihydrog entests.\nSolar-system tests of gravity including lunar and satellite laser rang ing\ntests24and measurements of the precession of the perihelion of orbiting\nbodies25are also considered. The established advance of the perihelion for\nMercury and for the Earth is used to obtain constraints on combina tions of\n(aeff)µ,cµν, andsµν, which provides the best current sensitivity to ( aeff)J.\nA final class of tests involves the interaction of photons with gravit y.\nSignals arising in measurements of the time delay, gravitational Dopp ler\nshift, and gravitational redshift, are considered along with compa risons of\nthe behaviors of photons and massive bodies. Implications for a var iety of\nexisting and proposed experiments and space missions are consider ed.26\nExisting and expected sensitivities from the experiments and obser va-\ntionssummarizedabovearecollectedinTablesXIVandXVofRef.3.T hese\nsensitivities reveal excellent prospects for using matter-gravity couplings to\nseek Lorentz violation. The opportunities for measuring the count ershaded\ncoefficients ( aeff)µare particularly interesting in light of the fact that these\ncoefficients typically cannot be detected in nongravitational searc hes. Thus\nthe tests proposed in Ref. 3 offer promising new opportunities to se arch for\nsignals of new physics, potentially of Planck-scale origin.\nReferences\n1.Data Tables for Lorentz and CPT Violation, 2010 edition, V.A. Kosteleck´ y\nand N. Russell, arXiv:0801.0287v3.\n2. D. Colladay and V.A. Kosteleck´ y, Phys. Rev. D 55, 6760 (1997); Phys. Rev.\nD58, 116002 (1998).\n3. V.A. Kosteleck´ y and J.D. Tasson, arXiv:1006.4106.\n4. V.A. Kosteleck´ y, Phys. Rev. D 69, 105009 (2004).\n5. Q.G. Bailey and V.A. Kosteleck´ y, Phys. Rev. D 74, 045001 (2006); Q.G.\nBailey, Phys. Rev. D 80, 044004 (2009); arXiv:1005.1435.\n6. V.A. Kosteleck´ y and J.D. Tasson, Phys. Rev. Lett. 102, 010402 (2009).\n7. V.A. Kosteleck´ y and N. Russell, Phys. Lett. B 693, 443 (2010); N. Russell,\nthese proceedings, arXiv:1009.1383.\n8. I. Marson and J.E. Faller, J. Phys. E 19, 22 (1986).\n9. A. Peters, K.Y. Chung, and S. Chu, Nature 400, 849 (1999); J.M. McGuirk\net al., Phys. Rev. A 65, 033608 (2002); N. Yu et al., Appl. Phys. B 84, 647\n(2006); B. Canuel et al., Phys. Rev. Lett. 97, 010402 (2006); H. Kaiser et al.,\nPhysica B 385-386 , 1384 (2006).\n10. S. Dimopoulos et al., Phys. Rev. D 78, 042003 (2008).\n11. R.J. Warburton and J.M. Goodkind, Astrophys. J. 208, 881 (1976); S. Sh-\niomi, arXiv:0902.4081; L. Carbone et al., in T. Damour, R.T. Jantzen, and5\nR. Ruffini, eds., Proceedings of the Twelfth Marcel Grossmann Meeting on\nGeneral Relativity , World Scientific, Singapore, 2010.\n12. K. Kuroda and N. Mio, Phys. Rev. D 42, 3903 (1990); T.M. Niebauer, M.P.\nMcHugh, and J.E. Faller, Phys. Rev. Lett. 59, 609 (1987).\n13. S. Fray et al., Phys. Rev. Lett. 93, 240404 (2004).\n14. R.D. Reasenberg, in V.A. Kosteleck´ y, ed., CPT and Lorentz Symmetry II ,\nWorld Scientific, Singapore, 2005.\n15. V. Iafolla et al., Class. Quantum Grav. 17, 2327 (2000).\n16. H. Dittus and C. Mehls, Class. Quantum Grav. 18, 2417 (2001).\n17. R.D. Reasenberg and J.D. Phillips, Class. Q. Grav. 27, 095005 (2010).\n18. Y. Su et al., Phys. Rev. D 50, 3614 (1994); S. Schlamminger et al., Phys.\nRev. Lett. 100, 041101 (2008).\n19. P. Touboul et al., Comptes Rendus de l’Acad´ emie des Sciences, Series IV,\n2, 1271 (2001); T.J. Sumner et al., Adv. Space Res. 39, 254 (2007); A.M.\nNobiliet al., Exp. Astron. 23, 689 (2009); G. Amelino-Camelia et al., Exp.\nAstron.23, 549 (2009); P. Worden, these proceedings.\n20. F. Hasselbach and M. Nicklaus, Phys. Rev. A 48, 143 (1993); B. Neyenhuis,\nD. Christensen, and D.S. Durfee, Phys. Rev. Lett. 99, 200401 (2007).\n21. F.S. Witteborn and W.M. Fairbank, Phys. Rev. Lett. 19, 1049 (1967).\n22. G. Gabrielse, Hyperfine Int. 44, 349 (1988); N. Beverini et al., Hyperfine\nInt.44, 357 (1988); R. Poggiani, Hyperfine Int. 76, 371 (1993); T.J. Phillips,\nHyperfineInt. 109,357(1997); AGECollaboration, A.D.Cronin et al.,Letter\nof Intent: Antimatter Gravity Experiment (AGE) at Fermilab ,February2009;\nD. Kaplan, these proceedings, arXiv:1007.4956; J. Walz and T.W. H¨ ansch,\nGen. Rel. Grav. 36, 561 (2004); P. P´ erez et al.,Letter of Intent to the CERN-\nSPSC,November 2007; F.M. Huber, E.W. Messerschmid, and G.A. Smit h,\nClass. QuantumGrav. 18,2457 (2001); AEGISCollaboration, A.Kellerbauer\net al., Nucl. Instr. Meth. B 266, 351 (2008).\n23. K. Kirch, arXiv:physics/0702143; B. Lesche, Gen. Rel. G rav.21, 623 (1989).\n24. J.G. Williams, S.G. Turyshev, and H.D. Boggs, Phys. Rev. Lett.93, 261101\n(2004); T.W. Murphy et al., Pub. Astron. Soc. Pac. 120, 20 (2008).\n25. C.M. Will, Living Rev. Relativity 4, 4 (2001).\n26. B. Bertotti, L. Iess, and P. Tortora, Nature 425, 374 (2003); T. Appour-\nchauxet al., Exp. Astron. 23, 491 (2009); L. Iess and S. Asmar, Int. J. Mod.\nPhys. D 16, 2117 (2007); P. Wolf et al., Exp. Astron. 23, 651 (2009); B.\nChristophe et al., Exper. Astron. 23, 529 (2009); S.G. Turyshev et al., Int.\nJ. Mod. Phys. D 18, 1025 (2009); S.B. Lambert and C. Le Poncin-Lafitte,\nAstron. Astrophys. 499, 331 (2009); R. Byer, Space-Time Asymmetry Re-\nsearch, Stanford University proposal, January 2008; L. Cacciapuo ti and C.\nSalomon, Eur. Phys. J. Spec. Top. 172, 57 (2009); S.G. Turyshev and M.\nShao, Int. J. Mod. Phys. D 16, 2191 (2007); S.C. Unwin et al., Pub. Astron.\nSoc. Pacific 120, 38 (2008)." }, { "title": "1212.4448v4.Coupled_scalar_fields_Oscillons_and_Breathers_in_some_Lorentz_Violating_Scenarios.pdf", "content": "arXiv:1212.4448v4 [hep-th] 27 Feb 2015Coupled scalar fields Oscillons and Breathers in some Lorent z Violating Scenarios\nR. A. C. Correa1,∗and A. de Souza Dutra2,†\n1CCNH, Universidade Federal do ABC, 09210-580, Santo Andr´ e , SP, Brazil\n2UNESP-Campus de Guaratinguet´ a, 12516-410, Guaratinguet ´ a, SP, Brazil\n(Dated: February 23, 2015)\nIn this work we discuss the impact of the breaking of the Loren tz symmetry on the usual oscillons,\nthe so-called flat-top oscillons, and the breathers. Our ana lysis is performed by using a Lorentz vio-\nlation scenario rigorously derived in the literature. We sh ow that the Lorentz violation is responsible\nfor the origin of a kind of deformation of the configuration, w here the field configuration becomes\noscillatory in a localized region near its maximum value. Fu rthermore, we show that the Lorentz\nbreaking symmetry produces a displacement of the oscillon a long the spatial direction, the same\nfeature is present in the case of breathers. We also show that the effect of a Lorentz violation in\nthe flat-top oscillon solution is responsible by the shrinki ng of the flat-top. Furthermore, we find\nanalytically the outgoing radiation, this result indicate s that the amplitude of the outgoing radia-\ntion is controlled by the Lorentz breaking parameter, in suc h away that this oscillon becomes more\nunstable than its symmetric counterpart, however, it still has a long living nature.\nKeywords: oscillons, breathers, Lorentz, symmetry, break ing, violation\n1. INTRODUCTION\nThe study of nonlinear systems is becoming an area of\nincreasing interest along the last few decades [ 1,2]. In\nfact, such nonlinear behavior of physical systems is found\nin a broad part of physical systems nowadays. This in-\ncludescondensedmatter systems, field theoreticalmodels,\nmodern cosmology and a large number of other domains\nof the physical science [ 3]-[28]. One of the reasons of this\nincreasing interest is the fact that many of those systems\npresent a countable number of distinct degenerate mini-\nmal energy configurations. In many cases that degenerate\nstructure can be studied through simple models of scalar\nfields possessing a potential with two or more degenerate\nminima. For instance, in two or more spatial dimensions,\none can describe the so called domain walls [ 4] connecting\ndifferent portions of the space were the field is at differ-\nent values of the degenerate minima of the field poten-\ntial. In other words, the field configuration interpolates\nbetween two of those potential minima. At this point, it\nis important to remark that a powerful insight to solve\nnonlinear problems analytically was introduced by Bogo-\nmolnyi, Prasad and Sommerfield [ 5,6]. In this case, the\nmethod shown by Bogomolnyi, Prasad and Sommerfield\nis now called of BPS approach, and it is based in ob-\ntaining a first-order differential equation from the energy\nfunctional. By using this method, it is possible to find so-\nlutions that minimize the energy of the configuration and\nthat ensure their stability.\nIn the context of the field theory it is quite common the\nappearance of solitons [ 7], which are field configurations\npresenting a localized and shape-invariant aspect, having\na finite energy density as well as being capable of keep-\n∗rafael.couceiro@ufabc.edu.br\n†dutra@feg.unesp.bring their shape unaltered after a collision with another\nsolitons. The presence of those configurations is nowa-\ndays well understood in a wide class of models, presenting\nor not topological nature. As examples one can cite the\nmonopoles, textures, strings and kinks [ 8].\nAn important feature of a large number of interesting\nnonlinear models is the presence of topologically stable\nconfigurations, which prevents them from decaying due\nto small perturbations. Among other types of nonlinear\nfield configurations, there is a specially important class of\ntime-dependent stable solutions, the breathers appearing\nin the Sine-Gordon like models. Another time-dependent\nfield configuration whose stability is granted for by charge\nconservation are the Q-balls as baptized by Coleman [ 9]\nor nontopological solitons [ 10]. However, considering the\nfact that many physical systems interestingly may present\na metastable behavior, a further class of nonlinear sys-\ntems may present a very long-living configuration, usually\nknown as oscillons. This class of solutions was discovered\nin the seventies of the last century by Bogolyubsky and\nMakhankov [ 29], and rediscovered posteriorly by Gleiser\n[30]. Thosesolutions,appearedinthestudyofthedynam-\nics of first-order phase transitions and bubble nucleation.\nSince then, more and more works were dedicated to the\nstudy of these objects [ 30]-[48].\nOscillons are quite general configurations and are found\nin inflationary cosmological models [ 30], in the Abelian-\nHiggsU(1) models [ 31], in the standard model SU(2)×\nU(1) [32], in axion models [ 33], in expanding universe sce-\nnarios [34] and in systems involving phase transitions [ 35].\nTheusualoscillonaspectistypicallythatofabellshape\nwhich oscillates sinusoidally in time. Recently, Amin and\nShirokoff [ 36] have shown that depending on the intensity\nof the coupling constant of the self-interacting scalar field,\nit is possible to observe oscillons with a kind of plateau\nat its top. In fact, they have shown that these new oscil-\nlons are more robust against collapse instabilities in three\nspatial dimensions.2\nAt this point it is interesting to remark that Segur and\nKruskal[ 37] haveshownthat the asymptoticexpansiondo\nnot represent in general an exact solution for the scalar\nfield, in other words, it simply represents an asymptotic\nexpansion of first order in ǫ, and it is not valid at all\norders of the expansion. They have also shown that in\none spatial dimension they radiate [ 37]. In a recent work,\nthe computation of the emitted radiation of the oscillons\nwas extended to the case of two and three spatial dimen-\nsions [38]. Another important result was put forward by\nHertzberg [ 39]. In that work he was able to compute the\ndecayingrateofquantizedoscillons,anditwasshownthat\nits quantum rate decayis verydistinct of the classicalone.\nOn the other hand, some years ago, Kostelecky and\nSamuel [ 49] started to study the problem of the Lorentz\nandCPT(charge conjugation-parity-time reversal) sym-\nmetry breaking. This was motivated by the fact that\nthe superstring theories suggest that Lorentz symmetry\nshould be violated at higher energies. After that seminal\nwork, a theoretical framework about Lorentz and CPT\nsymmetry breaking has been rigorously developed. As an\nexample, the effects on the standard model due to the\nCPTviolation and Lorentz breaking were presented by\nColladay and Kostelecky [ 50]. Recently, a large amount\nof works considering the impact of some kind of Lorentz\nsymmetry breaking have appeared in the literature [ 51]-\n[66]. As one another example, recently Belich et. al.[52]\nstudied the Aharonov-Bohm-Casher problem with a non-\nminimal Lorentz-violating coupling. In that reference the\nauthorshaveshown that the Lorentz-violationis responsi-\nblebytheliftingoftheoriginaldegeneraciesintheabsence\nof magnetic field, even for a neutral particle.\nBy introducing a dimensional reduction procedure to\n(1+2) dimensions presented in Ref. [ 53], Casana, Car-\nvalho and Ferreira applied the approach to investigate the\ndimensional reduction of the CPT-even electromagnetic\nsector of the standard model extension. Another impor-\ntant work was presented by Boldo et al.[54], where the\nproblem of Lorentz symmetry violation gauge theories in\nconnection with gravity models was analyzed. In a very\nrecentwork, Kosteleckyand Mewes[ 55], alsoanalyzedthe\neffects of Lorentz violation in neutrinos.\nIn recent years, investigations about topological defects\nin the presence of Lorentz symmetry violation have been\naddressed in the literature [ 56]-[58]. Works have also been\ndone on monopole and vortices in Lorentz violation sce-\nnarios [59]. For instance, in Ref. [ 59], a question about\nthe Lorentz symmetry violation on BPS vortices was in-\nvestigated. In that paper, the Lorentz violation allows a\ncontrol of the radial extension and of the magnetic field\namplitude of the Abrikosov-Nielsen-Olesen vortices.\nIn fact, Lorentz invariance is the most fundamental\nsymmetry of the standard model of particle physics and\nthey have been very well verified in several experiments.\nBut, it is important to remark that we can not be sure\nthat this, or any other, symmetry is exact apart from\nan experimental accuracy. This affirmation is encouraged\ndue to the fact that there exists some experimental testsof the Lorentz invariance being carried in low energies, in\nother words, energies smaller than 14 Tev. Thus, from\nthis fact, we can suspect that at high energies the Lorentz\ninvariance could not be preserved. As an example, in the\nstring theory there is a possibility that we could be living\nin an Universe which is governed by noncommutative co-\nordinates [ 67]. In this scenario it was shown in Ref. [ 68]\nthat the Lorentz invariance is broken.\nFurthermore, in a cosmologicalscenario, the occurrence\nof high energy cosmic rays above the Greisen-Zatsepin-\nKuzmin (GZK) cutoff [ 69] or super GZK events, has been\nfound in astrophysical data [ 70]. This event indicate the\npossibility of a Lorentz violation [ 71].\nThe impact of Lorentz violation on the cosmological\nscenario is very important, because several of its weak-\nnesses could be easily explained by the Lorentz violation.\nFor instance, it was shown by Bekenstein [ 72] that the\nproblem of the dark matter is associated with the Lorentz\nviolating gravity and in Ref.[ 73] Lorentz violation also is\nused to clarify the dark energy problem. Nowadays, the\nbreaking of the Lorentz symmetry is a fabulous mecha-\nnism for description of several problems and conflicts in\ncosmology, such as the baryogenesis, primordial magnetic\nfield, nucleosynthesis and cosmic rays [ 74].\nIn the inflationary scenario with Lorentz violation,\nKannoandSoda[ 75]haveshownthatLorentzviolationaf-\nfects the dynamics of the inflationary model. In this case,\nthatauthorsshowedthat, usingascalar-vector-tensorthe-\nory with Lorentz violation, the exact Lorentz violation\ninflationary solutions are found in the absence of the in-\nflaton potential. Therefore, the inflation can be connected\nwith the Lorentz violation.\nHere, it is convenient to us to emphasize that the in-\nflation is the fundamental ingredient to solve both the\nhorizon as the flatness problems of the standard model\nof the very early universe. Approximately 10−33seconds\nafter the inflation, the inflaton decays to radiation, where\nquarks, leptons and photons were coupled to each other.\nInthiscase,thebaryonicmatterwaspreventedfromform-\ning. Therefore, approximately 1 .388×1012seconds after\nthe Big Bang, the universe has cooled enough to allow\nphotons to freely travel through the universe. After that,\nmatter has became dominant in the universe.\nAt this point, it is important to remark that the post-\ninflationaryuniverseisgovernedbyrealscalarfieldswhere\nnonlinear interactions are present. Thus, it was shown\nin Ref. [ 76] that oscillons can easily dominate the post-\ninflationary universe. In that work, it was demonstrated\nthatthe post-inflationaryuniversecancontainaneffective\nmatter-dominated phase, during which it is dominated by\nlocalized concentrations of scalar field matter. Further-\nmore, in a very recent work [ 77], a class of inflationary\nmodels was introduced, giving rise to oscillons configu-\nrations. In this case, it was argued that these oscillons,\ncould dominate the matter density of the universe for a\ngiven time. Thus, one could naturally wonder about the\neffect of Lorentz violation over this scenario.\nThus, in this work we are interested in answer the fol-3\nlowing issues: Can oscillons and breathers exist in sce-\nnarios with Lorentz violation symmetry? If oscillons and\nbreathers exists in these scenarios how their profile is\nchanged? Furthermore, what happens with the lifetime\nof the oscillons?\nTherefore, in this paper, we will show that oscillons\nand breathers can be found in Lorentz violation scenarios,\nour study is performed by using Lorentz violation theories\nrigorously derived in the literature [ 50,78]. As a conse-\nquence, the principal goal here is to analyze the case of\ntwo nonlinearly coupled scalar fields case. However, we\nuse a constructive approach, so that we start by studying\nthe cases of one scalar field models and, then, use those\nresults in the study we are primarily interested in.\nThis paper is organized as follows. In section 2 we\npresent the description of the Lagrangian density for a\nrealscalarfield in presenceofa Lorentzviolationscenario.\nIn section 3 we calculate the respective commutation rela-\ntions of the Poincar´ egroup in the Standard-Model Exten-\nsion (SME) in a 1+1 dimensional flat Minkowski space-\ntime. The approach of the equation of motion is given in\nsection4. Usualoscillonsinthe backgroundoftheLorentz\nviolation is analyzed in section 5. In section 6 we will find\nthe flat-top oscillons which violates the Lorentz symme-\ntry. The breathers solutions are presented in section 7.\nWe discuss the outgoing radiation by oscillons in section\n8. In the section 9 we will present the oscillons in a two\nscalar field theory. Finally, we summarize our conclusions\nin section 10.\n2. STANDARD-MODEL EXTENSION\nLAGRANGIAN\nIn this section, we present a scalar field theory in a\n3+1-dimensional flat Minkowski space-time, but here we\nconsider a break of the Lorentz symmetry. In low energy,\nLorentz and CPTsymmetries the standard model (SM)\nofparticlephysicsisexperimentallywellsupported, but in\nhighenergiesthesuperstringtheoriessuggestthatLorentz\nsymmetry should be violated, in this context, the frame-\nwork to study Lorentz and CPTviolation is the so-called\nstandard-model extension. In the description of the SME,\nthe Lagrangian density for a real scalar field containing\nLorentz violation (LV), which can be read as a simplified\nversion of the Higgs model, is given by [ 50,78]\nL=1\n2∂µϕ∂µϕ+1\n2kµν∂µϕ∂νϕ−V(ϕ),(1)\nwhereϕis a real scalar field, kµνis a dimensionless tensor\nwhich controls the degree of Lorentz violation and V(ϕ)\nis the self-interaction potential. It is important to remark\nthat, some years ago [ 56], this Lagrangian density was\nused to study defect structures in Lorentz and CPTvio-\nlating scenarios. In that case the authors showed that the\nviolation of Lorentz and CPTsymmetries is responsible\nby the appearance of an asymmetry between defects andantidefects. This was generalized in [ 56]. Furthermore,\none similar Lagrangian density have been applied in the\nstudy on the renormalization of the scalar and Yukawa\nfield theories with Lorentz violation. In that case, it was\nshown that a LV theory with Nscalar fields, interacting\nthrough a φ4interaction, can be written as\nLK=1\n2(∂µϕi)(∂µϕi)+1\n2N/summationdisplay\ni=1Ki\nµν∂µϕi∂µϕi−1\n2λ2ϕ2\ni\n+N/summationdisplay\ni=1uβ\niϕi∂βϕi+N/summationdisplay\nj=1ϕ2\nivβ\nj∂βϕj−g\n4!(ϕ2\ni)2.(2)\nAs a simple example, that authors showed for Ki\nµν=\nKi\n00δ0\nµδ0\nνthat the dispersion relation is given by E=/radicalbig\np2−Ki\n00(p0)2+λ2, which implies in a LV. Therefore,\nusing explicit calculations, the quantum correctionsin the\nabove LV theory was studied, and these results show that\nthe theory is renormalizable.\nNow, returning to the equation ( 1), we can write the\nLagrangian density in the form\nL=1\n2(ηµν+kµν)∂µϕ∂νϕ−V(ϕ). (3)\nIn this case, the Minkowsky metric is modified from\ngµνtoηµν+kµν, which is responsible for the breaking\nof the Lorentz symmetry [ 50,78,79]. At this point it\nis possible to apply an appropriate linear transformation\nof the space-time variable xµ, in order to map the above\nLagrangiandensity into a Lorentz-likecovariantform, but\nthisleadstochangesinthefieldsandcouplingconstantsof\nthe potential. Thus, the coupling constants and the fields\nare rescaled in function of the kµνparameters.\nClearly, asafinalproduct, theLVandLorentzinvariant\nLagrangians have the same equation of motion. The fun-\ndamental difference between these two equations comes\nfrom the fact that the new variables xµcarry informa-\ntion of the Lorentz violations through of the kµνparam-\neters. In other words, in the transformed variables, the\nsystem looks to be covariant (under boosts of the trans-\nformed space-time variables). However, as a consequence\nof the fact that the resulting couplings become not invari-\nant when one changes from a reference frame to another,\nthere is no real Lorentz invariance. For instance, such be-\nhavior would be analogous to a change of the value of the\nelectrical charge when one moves from an inertial refer-\nence frame to another one, which is forbidden.\nIn the Lagrangian density ( 1),kµνis a constant tensor\nrepresented by a 4 ×4 matrix. It is the term which can\nbe responsible for the breaking of the Lorentz symmetry.\nThus, we write the tensor kµνin the form\nkµν=\nk00k01k02k03\nk10k11k12k13\nk20k21k22k23\nk30k31k32k33\n, (4)4\nIngeneral kµνhasarbitraryparameters,butit isimpor-\ntant to remark that if this matrix is real, symmetric, and\ntraceless, the CPTsymmetry is kept [ 50,78]. Here, we\ncomment that under CPToperation, ∂µ→ −∂µ, the term\nkµν∂µϕ∂νϕgoes askµν∂µϕ∂νϕ→+kµν∂µϕ∂νϕ. Thus,\none notices that kµνis always CPT-even, regardless its\nproperties. Furthermore, the tensor kµνshould be sym-\nmetric in order to avoid a vanishing contribution.\nIn a recent work, Anacleto et al.[80] also analyzed\na similar process to break the Lorentz symmetry, where\nthe tensor kµνwas used to study the problem of acous-\ntic black holes in the Abelian Higgs model with Lorentz\nsymmetry breaking. In another work by Anacleto et al.\n[80] the tensor kµνwas used to study the superresonance\neffect from a rotating acoustic black hole with Lorentz\nsymmetry breaking. Finally, in a very recent work [ 57],\nit was introduced a generalized two-fields model in 1+1\ndimensions which presents a constant tensor and vector\nfunctions. In that case, it was found a class of traveling\nsolitons in Lorentz and CPTbreaking systems.\nHowever,wecantofindsystemswithLorentzsymmetry\nbreak which has an additional scalar field [ 79]\nL=1\n2∂µϕ1∂µϕ1+1\n2∂µϕ2∂µϕ2+1\n2kµν∂µϕ1∂νϕ1(5)\n−m1ϕ2\n1\n2−m2ϕ2\n2\n2−V(ϕ1,ϕ2).\nIn the above Lagrangian density, we have a different\ncoefficient correcting the metric, but the coefficients for\nLorentz violation cannot be removed from the Lagrangian\ndensity using variables or fields redefinitions and observ-\nable effects of the Lorentz symmetry break can be de-\ntected in the above theory. Therefore, theories with fewer\nfields and fewer interactions allow more redefinitions and\nobservable effects.\n3. SME LAGRANGIAN: ONE FIELD THEORY\n(OFT)\nIn this section, we will work in a 1 + 1-dimensional\nMinkowski space-time. Here, we study a scalar field the-\nory in the presence of a Lorentz violating scenario. The\ntheory that we will study is given by the Lagrangian den-\nsity (1). Thus, in this case, the corresponding Lagrangian\ndensity must\nL1+1=1\n2α1(∂tϕ)2−1\n2α2(∂xϕ)2+1\n2α3∂tϕ∂xϕ−V(ϕ),\n(6)\nwhere\nα1≡(1+k00), α2≡(1−k11), α3≡(k01+k10),\n(7)\n∂t≡∂/∂t, ∂ x≡∂/∂x.\nAt this point, it is important to remark that the La-\ngrangiandensity clearly has not manifest covariance. Fur-\nthermore, it is possible to observe that the covariance isrecovered by choosing k00=k11= 0 and k01=−k10\n(ork01=k10= 0). Another possibilities that does\nnot represent a LV are k00=−k11andk01=−k10(or\nk01=k10= 0).\nNow, from the above, we can easily construct the cor-\nresponding Hamiltonian density\nH=β1Π2+β2(∂xϕ)2+β3Π(∂xϕ)+V(ϕ),(8)\nwhereβ1= 1/(2α1),β2= [2α1α2+α3(α3−1)]/(4α1),\nβ3=−α3/(2α1) and Π is the conjugate momentum,\nwhich is given by\nΠ =α1∂tϕ+(α3/2)∂xϕ. (9)\nLet us now see how the Poincar` e algebra is modified in\nthisscenario. Theideaofthepresentanalysisistoseehow\nthe Poincar´ e invariance is broken. In other words, ver-\nify how this scenario has the Lorentz symmetry violated.\nTherefore, for this we write down the three Poincar` e gen-\nerators, the Hamiltonian H, the total momentum Pand\nthe Lorentz boost M\nH=/integraldisplay\ndxH, (10)\nP=/integraldisplay\ndx/bracketleftbiggΠ(∂xϕ)\nα1−α3(∂xϕ)2\n2α1/bracketrightbigg\n, (11)\nM=/integraldisplay\ndx/braceleftbigg\nt/bracketleftbiggΠ(∂xϕ)\nα1−α3(∂xϕ)2\n2α1/bracketrightbigg\n−xH/bracerightbigg\n.(12)\nWith this, we can calculate the commutation relations\nof the Poincar` e group. Thus, after straightforward calcu-\nlations of the usual commutation relations, it is not diffi-\ncult to conclude that\n[H,P] =−i/parenleftbiggα3\nα2\n1/parenrightbigg/integraldisplay\ndx(∂xϕ)(∂xΠ), (13)\n[M,H] =−i/integraldisplay\ndx/parenleftbig\n(4β1β2+β2\n3)Π(∂xϕ) (14)\n−α3\nα2\n1(∂xϕ)(∂xΠ)+2β2β3(∂xϕ)2+2β1β3(Π)2/parenrightbigg\n,\n[M,P] =−iH\nα1+iα3\n2α2\n1/integraldisplay\ndx(Π(∂xϕ) (15)\n+x(∂xϕ)(∂xΠ)+β3\n4β1(∂xϕ)2/parenrightbigg\n.\nFrom the above relations, we can see that the Poincar` e\nalgebra is not closed, since that the usual commutations\nare not recovered. As a consequence, in this scenario we\nhaveoneviolationoftheLorentzsymmetry. However,itis\npossible torecoverthe complete commutation relationsby\ntakingk00=k11= 0 and k01=−k10(ork01=k10= 0).\nFor instance, making k00=k11= 0 and k01=−k10we\nhave\n[H,P] = 0,[M,H] =−iP,[M,P] =−iH. (16)5\nAt this pointwe canverifythat, forthe case k00=−k11\nandk01=−k10(ork01=k10= 0), the commutation\nrelations ( 13)-(15) lead to\n[H,P] = 0,[M,H] =−iα1P,[M,P] =−iH/α1.(17)\nHowever, in the above case, we can recover the usual\nPoincar` e algebra using the re-scale P=˜P/α1. Thus,\nsuch commutation relations indicates that there is no LV\nin this tensor configuration.\nIn summary, the Lagrangian density ( 6) has explicit\ndependenceontheparameters k00,k11,k01andk10, which\nis responsible for the violation of the Lorentz symmetry.\nThis happens due to the fact that the Poincar´ einvariance\nis not preserved, as one can see from ( 13)-(15).\n4. EQUATION OF MOTION IN LORENTZ\nVIOLATION SCENARIOS: OFT\nIn this section, we will study the equation of motion in\nthe presence of the scenario with Lorentz violation of the\nprevious section. Here, our aim is to study the case in\nthe 1+1 -dimensional Minkowski space-time. As a conse-\nquence, we will study the theory that is governed by the\nLagrangian density ( 6). Consequently, the corresponding\nclassical equation of motion can be written as\nα1∂2ϕ(x,t)\n∂t2−α2∂2ϕ(x,t)\n∂x2+α3∂2ϕ(x,t)\n∂x∂t+Vϕ= 0,(18)\nwhereVϕ≡∂V/∂ϕ. Note that the above equation is\ncarrying information about the symmetry breaking of the\ntheory.\nHere, if one applies the transformation involving the\nLorentz boost in the above equation of motion, one gets\nq1∂ϕ2(x,,t,)\n∂t,2−q2∂ϕ2(x,,t,)\n∂x,2+q3∂ϕ2(x,,t,)\n∂x,∂t,+Vϕ= 0,\n(19)\nwhere\nx,=γ(x−vt),t,=γ(t−vx/c2),γ= 1//radicalbig\n1−(v/c)2,(20)\nand\nq1=γ2/parenleftbiggα1c2−α2v2−α3cv\nc4/parenrightbigg\n,\nq2=γ2/parenleftbigg−α1v2+α2c2+α3cv\nc2/parenrightbigg\n, (21)\nq3=γ2/parenleftbigg−2vα1c+2cα2v−α3(c2+v2)\nc3/parenrightbigg\n.\nFollowing the above demonstration, we can see clearly\nthatthisequationisnotinvariantunderboosttransforma-\ntions. For instance, we can conclude that the possibilities\n[k00=−k11,k01=−k10] or [k00=−k11,k01=k10= 0]leads to the equations\nα1\nc2∂ϕ2(x,t)\n∂t2−α1∂ϕ2(x,t)\n∂x2+Vϕ= 0,(22)\nα1\nc2∂ϕ2(x,,t,)\n∂t,2−α1∂ϕ2(x,,t,)\n∂x,2+Vϕ= 0.(23)\nNote that there is no modification of the equations, in\nother words, the possibilities [ k00=−k11,k01=−k10] or\n[k00=−k11,k01=k10= 0] does not represent a genuine\nfactor for LV.\nIn order to solve analytically the differential equation\n(18) and simultaneously keep the breaking of the Lorentz\nsymmetry ,we must decouple the equation. For this, we\napply the rotation\n/parenleftbiggx\nt/parenrightbigg\n=/parenleftbiggcos(θ)−sin(θ)\nsin(θ) cos(θ)/parenrightbigg/parenleftbiggX\nT/parenrightbigg\n,(24)\nwhereθis an arbitrary rotation angle. Thus, the equation\n(18) in the new variables is rewritten as\nh1∂2ϕ(X,T)\n∂T2−h2∂2ϕ(X,T)\n∂X2+Vϕ= 0,(25)\nwith the definitions\nθ≡ −1\n2arctan/parenleftbiggα3\nα1+α2/parenrightbigg\n,\nh1≡α2\n1−α2\n2+[α2\n3+(α1+α2)2]cos(2θ)\n2(α1+α2),(26)\nh2≡α2\n2−α2\n1+[α2\n3+(α1+α2)2]cos(2θ)\n2(α1+α2).\nNote that the rotation angle θhas been chosen in or-\nder to eliminate the dependence in the term ∂2ϕ/∂X∂T .\nNow, performingthedilations T=√h1ΥandX=√h2Z,\none gets\n∂2ϕ(Z,Υ)\n∂Υ2−∂2ϕ(Z,Υ)\n∂Z2+Vϕ= 0.(27)\nFrom now on we will use the above equation to de-\nscribe the profile of oscillons and breathers. It is of great\nimportance to remark that the above equation has all the\ninformation about the violation of the Lorentz symmetry.\nIn fact, the field ϕ(Z,Υ) carries on the dependence of the\nparameters that break the Lorentz symmetry, this infor-\nmation arises from the fact that new variables Zand Υ\nhave explicit dependence on the kµνelements.\n5. USUAL OSCILLONS WITH LORENTZ\nVIOLATION: OFT\nNow, we study the case of a scalar field theory which\nsupports usual oscillons in the presence of Lorentz violat-\ning scenarios. The profile of the usual oscillons is one in\nwhich the spatial structure is localized in the space and,6\nin the most cases, is governed by a function of the type\nsech(x). On the other hand, the temporal structure is like\ncos(t), which is periodic. The theory that we will study\nis given by the Lagrangian density ( 6). In this case, we\nshowed in the last section that the corresponding classical\nequation of motion, after some manipulations, can be rep-\nresented by the equation ( 27). Thus, in order to analyze\nusual oscillons in this situation, we choose the potential\nthat was used in [ 36], which is written as\nV(ϕ) =1\n2ϕ2−1\n4ϕ4+g\n6ϕ6, (28)\nwheregrepresents a free coupling constant and we will\nconsider a regime where g >>1.\nSince our primordial interest is to find periodic and lo-\ncalized solutions, it is useful, as usual in the study of the\noscillons, to introduce the following scale transformations\nintandx\nτ=ωΥ, y=ǫZ, (29)\nwithω=√\n1−ǫ2. Thus, the equation of the motion ( 27)\nbecomes\nω2∂2ϕ(y,τ)\n∂τ2−ǫ2∂2ϕ(y,τ)\n∂y2+ϕ−ϕ3+gϕ5= 0.(30)\nNow we are in a position to investigate the usual oscil-\nlons. But it is important to remark that the fundamental\npoint is that here we have the effects of the Lorentz sym-\nmetry breaking. We can see this by inspecting the above\nequation of motion, which is carrying information about\nthe terms of the Lorentz breaking through the variables\nyandτ. We observe that it is possible to recover the\noriginal equation of motion for usual oscillons choosing\nk00=k11= 0 and k01=−k10(ork01=k10= 0). In this\ncase the Lorentz symmetry is recovered.\nNext we expand ϕas\nϕ(y,τ) =ǫϕ1(y,τ)+ǫ3ϕ3(y,τ)+ǫ5ϕ5(y,τ)+....(31)\nNote that the above expansion has only odd powers of\nǫ, this occurs because the equation is odd in ϕ. Let us\nnow substitute this expansion of the scalar field into the\nequation of motion ( 30). This leads to\n∂2ϕ1\n∂τ2+ϕ1= 0, (32)\n∂2ϕ3\n∂τ2+ϕ3−∂2ϕ1\n∂τ2−∂2ϕ1\n∂y2−ϕ3\n1= 0.(33)\nTherefore, the solution of equation ( 32) is of the form\nϕ1(y,τ) = Φ(y)cos(τ), (34)\nHere we call attention to the fact that the solution must\nbe smooth at the origin and vanishing when ybecomes\ninfinitely large.In order to find the solution of Φ( y), let us substitute\nthe solution obtained for ϕ1(y,τ) into the equation ( 33).\nThus, it is not hard to conclude that\n∂2ϕ3\n∂τ2+ϕ3=/parenleftbiggd2Φ\ndy2−Φ+3\n4Φ3/parenrightbigg\ncos(τ)+1\n4Φ3cos(3τ).\n(35)\nSolving the above equation we find a term which is\nlinear in the time-like variable τ, resulting into a non-\nperiodical solution, and we are interested in solutions\nwhich are periodical in time. Then to avoid this we shall\nimpose that\nd2Φ\ndy2−Φ+3\n4Φ3= 0. (36)\nAt this point, one can verify that the above equation\ncan be integrated to give\n/parenleftbiggdΦ\ndy/parenrightbigg2\n+U(Φ) =E, (37)\nwhereU(Φ) =−Φ2+ (3/8)Φ4. Note that in the above\nequation, the arbitrary constant Eshould be set to zero\nin order to get solitonic solution. This condition allows\nthe fieldconfigurationtogoasymptoticallytothevacuaof\nthe field potential U(Φ). Now, we must solvethe equation\n(37) withE= 0. In this case one gets\nΦ(y) =4/radicalbigg\n8\n3[sech(y)]1/2. (38)\nAs one can see, up to the order O(ǫ), the corresponding\nsolution for the field in the original variables is given by\nϕosc(x,t) =ǫ4/radicalbigg\n8\n3/parenleftigg/radicaligg\nsech/bracketleftbiggǫ[xcos(θ)+tsin(θ)√h2/bracketrightbigg/parenrightigg\n(39)\n×cos/bracketleftbiggω[−xsin(θ)+tcos(θ)]√h1/bracketrightbigg\n+O(ǫ3).\nThe profile of the above solution is plotted in Fig. 1\nfor some values of the kµνparameters. In the Figure 1\nwe see the profile of the usual oscillon in the presence of\nthe backgroundofthe Lorentzbreakingsymmetry. In this\ncase, one can check that the dependence ofthe solution on\nthe Lorentz breaking parameters is responsible for a kind\nofdeformation ofthe configuration, where the field config-\nuration becomes oscillatory in a localized region near its\nmaximumvalue. Furthermore,inthecourseofthetime, it\nis possible to observethat the Lorentz breakingsymmetry\nproduces a displacement of the oscillon along the spatial\ndirection. In this case we will call these configurations as\n”envelopedoscillons”, since in t= 0 the new configuration\nis enveloped by the oscillon with Lorentz symmetry.\nMoreover, one can note that if one wants to recover the\nLorentz symmetry, it is necessary to impose that k00=\nk11= 0 and k01=−k10(ork01=k10= 0).7\n6. FLAT-TOP OSCILLONS WITH LORENTZ\nVIOLATION: OFT\nSome years ago, a new class of oscillons, which is char-\nacterized by a kind of plateau at its top, was presented by\nAmin and Shirokoff [ 36]. In that work, the authors have\nshown that this configuration has an important impact\non an expanding universe. Thus, in this section, we will\ndescribe the impacts of the Lorentz violation overthe flat-\ntop oscillons. We will study the case in 1+1-dimensional\nMinkowski space-time where the classical equation of mo-\ntion is given by ( 27). Also, in order to analyze the flat-top\noscillons in this scenario, we choose the potential that was\nused in [36], which is represented in ( 28).\nNow, we begin a direct attack to the problem of find-\ning the flat-top oscillons. Likewise to the procedure pre-\nsented in [ 36], we introduce a re-scaled scalar field by\nϕ(Z,Υ) =φ(y,τ)/√g, where Z=√gy,τ=̟Υ and\n̟=/radicalbig\n1−α2/g. It is important to remark that the con-\nstantα2is responsible by the change in the frequency,\nits presence comes from the nonlinear potential. Thus, it\nis not difficult to conclude that the classical equation of\nmotion can be rewritten as\n(∂2\nτφ+φ)+g−1[−α2∂2\nτφ−∂2\nyφ−φ3+φ5] = 0.(40)\nSo, we are in a position to investigate the so-called flat-\ntop oscillons. But it is important to remark that the fun-\ndamental point is that all the effects of the Lorentz sym-\nmetry breaking are present implicitly in the classical field.\nOf course, it is possible to recover the original equation\nof motion presented by Mustafa [ 36] through a suitable\nchoice of kµν.\nLet us go further on our search for flat-top oscillons.\nFor this, we expand φas\nφ(y,τ) =φ1(y,τ)+g−1φ3(y,τ)+.... (41)\nIf we substitute the above expansion of the scalar field\ninto the equation of motion ( 40), and collect the terms in\norderO(1) andO(g−1), we find\n∂2φ1\n∂τ2+φ1= 0, (42)\n∂2φ3\n∂τ2+φ3−α2∂2φ1\n∂τ2−∂2φ1\n∂y2−φ3\n1+φ5\n1= 0.(43)\nTherefore, the solution of equation ( 42) is of the form\nφ1(y,τ) = Ψ(y)cos(τ), (44)\nIn order to find the solution of Ψ( y) let us substitute\nthe solution obtained for φ1(y,τ) into the equation ( 43).\nThus, it is not hard to conclude that\n∂2φ3\n∂τ2+φ3=/parenleftbiggd2Ψ\ndy2−α2Ψ+3\n4Ψ3−5\n8Ψ5/parenrightbigg\ncos(τ)\n(45)\n+/parenleftbigg3\n4Ψ3−5\n16Ψ5/parenrightbigg\ncos(3τ)−Ψ5\n16cos(5τ).whose solution can be written as\nφ3(y,τ) =1\n8[4G(y)−2H(y)+8c1]cos(τ)\n−H(y) cos(3τ)+4[G(y)τ+2c2]sin(τ),(46)\nwhere we defined that G(y) ≡/parenleftig\nd2Ψ\ndy2−α2Ψ+3\n4Ψ3−5\n8Ψ5/parenrightig\nandH(y)≡/parenleftbig3\n4Ψ3−5\n16Ψ5/parenrightbig\n.\nFurthermore, c1andc2are arbitrary integration con-\nstants.\nSince that the solution of the function φ3has a term\nwhich is linear in the variable τ, resulting into a non-\nperiodical solution, and we are interested in solutions\nwhich are periodical in time, we shall impose that G(y)\nvanishes. As a consequence we get\nd2Ψ\ndy2=/parenleftbigg\nα2Ψ−3\n4Ψ3+5\n8Ψ5/parenrightbigg\n, (47)\nAt this point, one can verify that the above equation\nhas the same profile ofthe equationpresented in Ref. [ 36].\nTherefore, this equation can be integrated to give\n1\n2/parenleftbiggdΨ\ndy/parenrightbigg2\n+U(Ψ) =E, (48)\nwhereU(Ψ) =−(1/2)α2Ψ2+(3/16)Ψ4−(5/48)Ψ6. Note\nthat in the above equation, the arbitrary constant E\nshould be set to zero in order to get solitonic solution.\nThis condition allows the field configuration to go asymp-\ntotically to the vacua of the field potential U(Ψ). On\nthe other hand, it is usual to impose that the profile of\nΨ(y) be smooth at y= 0, then it is necessary to make\ndΨ(0)/dy= 0. As a consequence E=U(Ψ0) = 0, which\nimplies\nα2=3\n8Φ2\n0−5\n24Φ4\n0, (49)\nwith Ψ 0≡Ψ(0). Thus, solving the above equation in Ψ 0,\nwe have a critical value a≤αc=/radicalbig\n27/160. Above this\ncritical value, Ψ 0becomes imaginary.\nNow, we must solve the equation ( 48) withE= 0. In\nthis case, we have\ndΨ/radicalig\nα2Ψ2−3\n8Ψ4+5\n24Ψ6=dy. (50)\nFrom this it follows that\nΨ(y) =(u4√\n4vu)/radicalig\n2√v+cosh[2y/radicalbig\nuv(α2c−α2)],(51)\nwherev= 27/[160(α2\nc−α2)] andu= (v−1)/v.8\nAs one can see, up to the order O(1), the corresponding\nsolution for the field in the original variables is given by\nϕFT(x,t) = (52)\nu4√\n4vu/radicaligg\n2g√v+gcosh/braceleftbigg\n2[xcos(θ)+tsin(θ)]√\nuv(α2c−α2)√gh2/bracerightbigg\n×cos/braceleftbigg̟[−xsin(θ)+tcos(θ)]√h1/bracerightbigg\n+O(g−3/2).\nThe profile of the above solution is plotted in Fig. 2.\nIn the Figure 2 we see the profile of the flat-top oscillon\nin the presence of the background of the Lorentz breaking\nsymmetry. Inthiscase,onecancheckthatthedependence\nof the solution on the Lorentz breaking parameters is re-\nsponsible for a control of the size of the oscillon plateau.\nThus, by measuring the width of the oscillon one could be\nable to verify the existence and the degree of the breaking\nof the symmetry. In Fig. 3 we see the typical profile of\nthe flat-top oscillon.\nThere one can note that the effect of the Lorentz break-\ning over the energy density, it is to becoming it more and\nmore localized around the origin.\n7. BREATHERS WITH LORENTZ VIOLATION:\nOFT\nWe will now construct the profile ofa breather in a 1+1\ndimensionalMinkowskispace-time. Again, we will usethe\nclassical equation of motion ( 27). The breather solutions\narise from the sine-Gordon model\nV(ϕ) =γ\nβ[1−cos(βϕ]. (53)\nThe sine-Gordonmodel is invariantunder ϕ→ϕ+2nπ,\nwherenis an integer number. In this case, the classical\nequation of motion is\n∂2ϕ(Z,Υ)\n∂Υ2−∂2ϕ(Z,Υ)\n∂Z2+γsin(βϕ) = 0.(54)\nThe above equation can be solved by the inverse-\nscattering method [ 81]. Thus, after straightforward cal-\nculations we conclude that the breather solution is given\nby\nϕB(Z,Υ) =4\nβarctan/bracketleftigg/radicalbig\nγ−w2sin(wΥ)\nwcosh(Z/radicalbig\nγ−w2)/bracketrightigg\n,(55)\nwherewis the frequency of oscillation and describe dif-\nferent breathers. In Figures 4 and 5 we show the behavior\nof the above solution.8. RADIATION OF OSCILLONS WITH LORENTZ\nVIOLATION SYMMETRY: OFT\nAn important characteristic of the oscillons is its radi-\nation emission. In a seminal work by Segur and Kruskal\n[37] it was shown that oscillons in one spatial dimension\ndecay emitting radiation. Recently, the computation of\nthe emitted radiation in two and three spatial dimensions\nwas did in [ 38]. On the other hand, in a recent paper by\nHertzberg [ 39], it was found that the quantum radiation\nis very distinct of the classic one. It is important to re-\nmark that the author has shown that the amplitude of the\nclassical radiation emitted can be found using the ampli-\ntude of the Fourier transform of the spatial structure of\nthe oscillon.\nThus, inthis section, wedescribetheoutgoingradiation\nin scenarios with Lorentz violation symmetry. Here, we\nwill establish a method in 1 + 1 dimensional Minkowski\nspace-time that allows to compute the classical radiation\nof oscillons in scenarios with Lorentz symmetry breaking.\nThis is done by following the method presented in [ 39].\nThis method suggests that we can write the solution of\nthe classical equation of motion in the following form\nϕsol(x,t) =ϕosc(x,t)+η(x,t), (56)\nwhereϕosc(x,t) is the oscillon solution and η(x,t) repre-\nsents a small correction. Let us substitute this decompo-\nsition of the scalar field into the equation of motion ( 18).\nThis leads to\nα1∂2ϕosc\n∂t2−α2∂2ϕosc\n∂x2+α3∂2ϕosc\n∂x∂t+α1∂2η\n∂t2\n(57)\n−α2∂2η\n∂x2+α3∂2η\n∂x∂t+U(ϕosc,η) = 0,\nwhereU(ϕosc,η) is a function which depends on the form\nofVϕsol(ϕsol). In order to decouple the above equation we\napply the rotation ( 24) and the dilations T=√h1Υ and\nX=√h2Z. Thus, we find\n∂2ϕosc(Z,Υ)\n∂Υ2−∂2ϕosc(Z,Υ)\n∂Z2+∂2η(Z,Υ)\n∂Υ2(58)\n−∂2η(Z,Υ)\n∂Z2+U(ϕosc,η) = 0.\nFrom the above equation it is possible to find the so-\nlution for η(Z,Υ) which carries the dependence on the\nparameters that break the Lorentz symmetry. We want\nto investigate the model given by ( 28), then we have\nU(ϕosc,η) =ϕosc+η−ϕ3\nosc−η3+3ϕ2\noscη+3ϕoscη2\n(59)\n+g(ϕ5\nosc+η5+10ϕ2\noscη3+10ϕ3\noscη2\n+5ϕoscη4+5ϕ4\noscη).9\nAsηrepresents a small correction, we assume that the\nnonlinear terms η2,η3,η4,η5and the parametric driving\nterms 3ηϕ2\nosc, 5gηϕ4\nosccan be neglected. At this point, it\nis important to remark that the parametric driven terms\nwere not considered because we are working in an asymp-\ntotic regime where ϕoscis also small. In this case, the\nequation ( 58) takes the form\n∂2η(Z,Υ)\n∂Υ2−∂2η(Z,Υ)\n∂Z2+η(Z,Υ) =−J(Z,Υ),(60)\nwhere\nJ(Z,Υ) =∂2ϕosc(Z,Υ)\n∂Υ2−∂2ϕosc(Z,Υ)\n∂Z2(61)\n+ϕosc(Z,Υ)−ϕ3\nosc(Z,Υ)+gϕ5\nosc(Z,Υ).\nWe can use the Fourier transform for solving the differ-\nential equation ( 60) whereJ(Z,Υ) acts as a source. With\nthis in mind, we write down the Fourier integral trans-\nforms\nη(R,w) =1√\n2π/integraldisplay\ndZdΥη(Z,Υ) (62)\n×exp[−i(RZ−wΥ)],\nJ(R,w) =1√\n2π/integraldisplay\ndZdΥJ(Z,Υ) (63)\n×exp[−i(RZ−wΥ)].\nThen, we have the corresponding solution\nη(Z,Υ) =1√\n2π/integraldisplay\ndRdwη(R,w)(64)\n×exp[i(RZ−wΥ)],\nwhere\nη(R,w) =−J(R,w)\nR2−(w2+1). (65)\nFrom the above approach it is possible to find the ra-\ndiation field for the oscillons. As a consequence of the\nmethod, the oscillons expansion must be truncated.\n8.1. SME Usual Oscillons Radiation: OFT\nIn this subsection we will study the outgoing radiation\nof the usual oscillons in a Lorentz violation scenario. In\nthis case, the oscillon expansion truncated in order Nis\ngiven by\nϕ(y,τ) =ǫϕ1(y,τ)+ǫ3ϕ3(y,τ)+ǫ5ϕ5(y,τ)(66)\n+...+ǫNϕN(y,τ).As an example, we will consider N= 1. This is the case\nwhere the field configuration corresponds to the oscillon\nϕosc(y,τ) =ǫϕ1(y,τ). (67)\nSubstituting ( 67) in (61), we obtain\nJ(Z,Υ) =/parenleftigg\n4/radicalbigg\n8\n3/parenrightigg\nǫ3[sech(ǫZ)]3/2cos(3ωΥ).(68)\nThus, for N= 1 we can solve easily the integral ( 64)\nwhich allows to find η(Z,Υ). Therefore, we can gener-\nalize the result to Nsubstituting the expansion ( 66) in\n(61), and using the differential equation ( 30). After the\ncalculations, the result is\nJ(Z,Υ) =CNǫN+2[sech(ǫZ)]N+1/2cos(¯nωΥ)+...,(69)\nwhereCNare constant coefficients. For instance, for N=\n1 we have C1=4/radicalbig\n8/3. Next we calculate η(Z,Υ) as\ngiven by ( 64). After straightforward computations, one\ncan conclude that\nη(Z,Υ) =π√πCNǫN+2\nkradcos(ωradΥ) (70)\n×sin(kradZ)/integraldisplay\ndZsech(ǫZ)]N+1/2cos(kradZ).\nwhere\nωrad= ¯nω, k rad=/radicalig\nω2\nrad−1. (71)\nOn the expression ( 70), we note that there is an out-\ngoing radiation which has an amplitude described by the\nintegral\nA(krad) =π√πCNǫN+2\nkrad(72)\n×/integraldisplay\ndZsech(ǫZ)]N+1/2cos(kradZ),\nwe also note that the radiation has frequency ωradand\nwave number krad. We can make use of the above gen-\neralization to calculate the amplitude of radiation of the\nusual oscillons in Lorentz violation scenario. For instance,\nforN= 1, we have\nA(krad) =4π√\n2πC1ǫ3\nkrad(73)\n×[b1F(a1,a2,a3,−1)+b∗\n1F(a1,a∗\n2,a∗\n3,−1)],\nwhereF(a1,a2,a3,−1) andF(a1,a∗\n2,a∗\n3,−1) are hyperge-\nometric functions with\nb1=1\n3ǫ−2ikrad, a1=3\n2, (74)\na2=3\n4−ikrad\n2ǫ, a3=7\n4−ikrad\n2ǫ.10\nIn Fig. 6 we see how the amplitude of the outgoing\nradiation changes with the parameters of kµν. From that\nFigure one can see that the amplitude of the outgoing ra-\ndiation of the oscillons is controlled by the terms of the\nLorentz breaking of the model, in such way that the radi-\nation amplitude will decay faster when the Lorentz break-\ning increases.\n8.2. SME Flat-top oscillons radiation: OFT\nWe will now present the outgoing radiation by the Flat-\ntop oscillons in Lorentz violation scenario. Here, the as-\nsociated oscillon expansion truncated in Nis defined as\nϕ(y,τ) =ϕ1(y,τ)+1\ngϕ3(y,τ) (75)\n+1\ng2ϕ5(y,τ)+...+1\ngN−1ϕ2N−1(y,τ).\nSubstituting the above expansion in ( 61), we have that\nJ(Z,Υ) =¯CN (76)\nׯCN\n\n(u4√\n4vu)/radicalig\n2g√v+gcosh[2Z/radicalbig\nuv(α2c−α2)/√g]\n\nN+2\n×cos(¯n¯ωΥ)+...,\nwhere¯CNare constant coefficients. Now we calculate\nη(Z,Υ) as given by ( 64). After straightforward compu-\ntations, one can conclude that\nη(Z,Υ) =π√π¯CN\n¯kradcos(¯ωradΥ)sin(¯kradZ) (77)\n×/integraldisplay\nd˜Z\n\n(u4√\n4vu)/radicalig\n2g√v+gcosh[2˜Z/radicalbig\nuvg(α2c−α2)]\n\nN+2\n×cos(¯krad˜Z).\nwhere\n¯ωrad= ¯n¯ω,¯krad=/radicalig\n¯ω2\nrad−1. (78)\nFrom the above expression, we see that there is an out-\ngoing radiation which has its amplitude described by the\nintegral\nA(krad) =π√π¯CN\n¯krad/integraldisplay\ndZcos(¯kradZ) (79)\n×\n\n(u4√\n4vu)/radicalig\n2√v+cosh[2Z/radicalbig\nuv(α2c−α2)/√g]\n\nN+2\n.We can make use the above generalization to calcu-\nlate the amplitude of radiation of the Flat-top oscillons\nin Lorentz violation scenario. For instance, for N= 1, we\nhave\nA(krad) =4π¯CN\nA0¯krad/parenleftigg\nu4√\n4vu√g/parenrightigg3\n(ξ1Fa+ξ∗\n1Fb),(80)\nwhereFa=F(Ω1;Ω2;Ω2;Ω3,Ω4,Ω5) andFb=\nF(Ω∗\n1;Ω2;Ω2;Ω∗\n3,Ω4,Ω5) are the Appell hypergeometric\nfunctions of two variables, and\nA0= 2/radicaligg\nuv(α2c−α2)√g, ξ1= 3+2ikrad\nA0,\nΩ1=3\n2−ikrad\nA0,Ω2=3\n2,Ω3=5\n2−ikrad\nA0,(81)\nΩ4=/radicalig\nA2\n0−1−A0,Ω5=1/radicalbig\nA2\n0−1−A0.\nIn this case we see that the amplitude of the outgoing\nradiation changes with the parameters kµν. We can see\nthat the amplitude of the outgoing radiation of the oscil-\nlons is controlled by the terms of the Lorentz breaking of\nthe model, in such way that the radiation amplitude will\ndecay faster when the Lorentz breaking increases.\n9. OSCILLONS WITH LV: TWO FIELD THEORY\n(TFT)\nWe have seen in section 2 that the most important sce-\nnariowithLVisthatdescribedbyatheorywithtwoscalar\nfields, because it is possible to find observable effects of\nthe LV. Then, in this section, we study a two scalar field\ntheory in the presence of a LV scenario. The theory that\nwewillstudyissimilartothatgivenbyPotting[ 79]. Here,\nwe will work with the corresponding Lagrangian density\nL=1\n2∂µϕ1∂µϕ1+1\n2∂µϕ2∂µϕ2 (82)\n+1\n2kµν∂µϕ1∂νϕ2−V(ϕ1,ϕ2).\nwhereV(ϕ1,ϕ2)istheinteractionpotential. Forexample,\nin order to find oscillons solutions, we can to choose the\npotential in the form\nV(ϕ1,ϕ2) =g\n3/parenleftbig\nϕ6\n1+ϕ6\n2/parenrightbig\n−1\n2/parenleftbig\nϕ4\n1+ϕ4\n2/parenrightbig\n(83)\n+ϕ2\n1+ϕ2\n2+5g/parenleftbig\nϕ4\n1ϕ2\n2+ϕ2\n1ϕ4\n2/parenrightbig\n−3ϕ2\n1ϕ2\n2.\nIn order to decouple the Lagrangian density ( 82), we\napply the rotation\n/parenleftbigg\nϕ1\nϕ2/parenrightbigg\n=1\n2/parenleftbigg\n1 1\n1−1/parenrightbigg/parenleftbigg\nσ1\nσ2/parenrightbigg\n. (84)11\nAfter straightforward computations, one can conclude\nthat\nL=1\n2∂µσ1∂µσ1+1\n2kµν\n1∂µσ1∂νσ1 (85)\n+1\n2∂µσ2∂µσ2+1\n2kµν\n2∂µσ2∂νσ2−V(σ1,σ2),\nwhere\nkµν\n1=1\n4kµν,kµν\n2=−1\n4kµν, (86)\nand the potential is\nV(σ1,σ2) =V(σ1)+V(σ2), (87)\nwith\nV(σi) =g\n6σ6\ni−1\n4σ4\ni+1\n2σ2\ni, i= 1,2.(88)\nIt is important to note that applying the rotations in\nthe fields, the Lagrangian density was decoupled into two\nindependent Lagrangians L=2/summationtext\ni=1Li, where\nLi=1\n2∂µσi∂µσi+1\n2kµν\ni∂µσi∂νσi−V(σi),(89)\nWe can see that all the preceding approaches and re-\nsults can be used here to find the fields σ1andσ2. An-\nother important point that it is convenient to remark at\nthis point, comes from the fact that any variable xµredef-\ninition will carry information of the parameter kµνwhich\nit is responsible by LV.\nAs we are working in 1+1-dimensions, the Lagrangians\n(89) become\nL(1+1)\ni=1\n2ai(∂tσi)2−1\n2bi(∂xσi)2(90)\n+1\n2di∂tσi∂xσi−V(σi),i= 1,2.\nIn this case, we have\nai≡(1+k00\ni), bi≡(1−k11\ni), di≡(k01\ni+k10\ni).(91)\nNow it is quite clear why the Lagrangian density ( 82)\nis more important and general than the one described\nby (1). First, because the commutation relations of the\nPoincar` e group is not closed, indicating a Lorentz Viola-\ntion. Second, because it is impossible to perform coor-\ndinate changes to eliminate the LV parameters in ( 85),\nbecause if we apply a coordinate change in order to write\nthe Lagrangian in an covariant form, only one of the sec-\ntors will stay invariant.Now, by using the approaches described in section 4,\nwe find the equations\n∂2σi(Zi,Υi)\n∂Υ2\ni−∂2σi(Zi,Υi)\n∂Z2\ni+Vσi= 0,(92)\nwhere\nZi=xcos(θi)+tsin(θi)√Li, (93)\nΥi=−xsin(θi)+tcos(θi)√Hi. (94)\nwith the set\nθi=−1\n2arctan/parenleftbiggdi\nai+bi/parenrightbigg\n, (95)\nLi=b2\ni−a2\ni+[d2\ni+(ai+bi)2]cos(2θi)\n2(ai+bi),(96)\nHi=a2\ni−b2\ni+[d2\ni+(ai+bi)2]cos(2θi)\n2(ai+bi).(97)\nFortunately, we can find periodical solutions for the\nfieldsσ1andσ2from the equation ( 92). In this case,\nwe are looking oscillons-like solutions. These solutions\nwere presented in the sections 5 and 6. Thus, from those\nsections we can show that\nσ(USUAL )\ni(x,t) = (98)\nǫi4/radicalbigg\n8\n3/parenleftigg/radicaligg\nsech/braceleftbiggǫi[xcos(θi)+tsin(θi)]√Li/bracerightbigg/parenrightigg\n×cos/braceleftbiggωi[−xsin(θi)+tcos(θi)]√Hi/bracerightbigg\n+O(ǫ3\ni),\nand\nσ(FLAT−TOP)\ni (x,t) = (99)\nui4√4viui/radicaligg\n2g√vi+gcosh/braceleftbigg\n2[xcos(θi)+tsin(θi)]√\nuivi(α2c−α2\ni)√gLi/bracerightbigg\n×cos/braceleftbigg̟i[−xsin(θi)+tcos(θi)]√Hi/bracerightbigg\n+O(g−3/2).\nIn the above solutions σ(USUAL )\ni represents the usual\noscillons and σ(FLAT−TOP)\ni are the Flat-Top ones. Fur-\nthermore, we have\nωi=/radicalig\n1−ǫ2\ni,̟i=/radicalig\n1−α2\ni/g,\n(100)\nvi= 27/[160(α2\nc−α2\ni)],ui= (vi−1)/vi.12\nAs above asserted, the original scalar fields ϕ1andϕ2\nare obtained from the fields σ1andσ2in the following\nform\nϕ1=σ1+σ2\n2,ϕ2=σ1−σ2\n2. (101)\nIt is important to remarkthat the resulting solutionsdo\nnot present merely algebraic relation between σiand the\noriginal parameters of the theory, but essentially lead to\nphysical consequences. As one can see, there are two kind\nof frequencies which can be combined for each scalar field\nϕi. This means that their solutions can be considered\nas a superposition of two independent fields and, as a\nconsequence, we can have an interference phenomena in\nthe structure of the oscillon.\nAn important question concerns the stability of the so-\nlutions, given that each field ϕiis a combination of the\nfieldsσi, the stability and longevity of the oscillons are\nguaranteed. From a mathematical point of view, one can\nthink that the originalfieldsconsistoflinearcombinations\nofσi. The same occurs when we calculated the outgoing\nradiation, in that case we have two radiation fields η1and\nη2, which are independent solutions with small resulting\namplitudes. As a consequence, their linear combinations,\n¯η1=η1+η2and ¯η2=η1−η2, will give the radiation field\nof solutions ϕi. Therefore, as ηiare very small solutions,\nwe still have the stability and longevity of the solutions\nguaranteed.\n10. CONCLUSIONS\nIn this work we have investigated the so-called flat-top\noscillons in the case of Lorentz breaking scenarios. We\nhave shown that the Lorentz violation symmetry is re-\nsponsible for the appearance of a kind of deformation of\nthe configuration. On the order hand, from inspection\nof the results coming from the flat-top oscillons in 1+1-\ndimensions with Lorentz breaking in comparison with the\nflat-topgivenin[ 36], onecanseethatthe oscillonsarecar-\nryinginformationabout the termsofthe Lorentzbreaking\nof the model, in this case by taking k00=k11= 0 and\nk01=−k10(ork01=k10= 0) one recovers the solution\npresented in Ref. [ 36]. Furthermore, this can lead one\nto obtain the degree of symmetry breaking by measuring\nthe width of the oscillon in 1 + 1 dimensions. One im-\nportant question about the non-linear solution is related\nto its stability. Thus, we studied the solutions found here\nby using the procedure introduced by Hertzberg [ 36,39].\nWe concluded that the radiation emitted by these oscil-\nlons is controlled by the terms of the Lorentz breaking\nof the model, in such way that the radiation will decay\nmore quickly as the terms become larger. Finally, all the\nresults obtained for the case of one scalar field models are\npromptlyextendedforthe caseofdoubletsofnounlinearly\ncoupled scalar fields.\nMoreover, it is important to highlight that the bounds\nin Lorentz violation theories in the Standard Model arevery small, and are compatible with the stability observed\nfor the oscillons here introduced. On the other hand, ob-\nservableeffects ofthese oscillonsinthe realworld, arepos-\nsible, for instance, in a cosmologicalcontext. In that case,\nthe life time of these oscillons can be decisive in the gener-\nation of coherent structures after cosmic inflation [ 82,83],\nwhere it was shown that oscillons can contribute up to\n20% of the energy density of the Universe. Thus, in this\nscenario, one should find bounds on the Lorentz violation\nwhich will open a new window to detect observable effects\nof breaking Lorentz symmetry. This possibility is encour-\naged by the fact that the break of the Lorentz symmetry\ninduces a kind of beat phenomenon in the structure of\nthe outgoing radiation, in contrast with the Lorentz in-\nvariant case (see Fig. 6). In this way, in a real world, one\ncan detect the difference in the frequency of the outgoing\nradiation, effect that would indicate the presence of a vi-\nolation of the Lorentz symmetry. Therefore, in order to\ndealing with these questions, we are presently working in\na future workwhere oscillonsin cosmologicalbackgrounds\nwith Lorentz symmetry breaking are presented.\nAcknowledgments\nThe authors thank Professor D. Dalmazi for useful dis-\ncussions. R.A.C.CorreathankP.H.R.S.Moraesfordis-\ncussions regarding cosmology. R. A. C. Correa also thank\nAlan Kostelecky for helpful discussions about Lorentz\nbreaking symmetry. The authors also thank CNPq and\nCAPES for partial financial support. Furthermore, the\nauthors also are grateful to the anonymous referees for\nthe comments that lead to improving the results and con-\nclusions of this work.\n[1]G. B. Whitham, Linear and Non-Linear Waves , John Wi-\nley and Sons, New York, (1974).\n[2]A.C. Scott, F. Y. F. Chiu, and D. W. Mclaughlin, Proc.\nI.E.E.E. 61, 1443 (1973).\n[3]R. Rajaraman and E. J. Weinberg, Phys. Rev. D 11, 2950\n(1975).\n[4]T. Vachaspati, Kinks and Domain Walls: An Introduction\nto Classical and Quantum Solitons (Cambridge University\nPress, Cambrifge, England, 2006).\n[5]M. K. Prasad and C. M. Sommerfield, Phys. Rev. Lett.\n35, 760 (1975).\n[6]E. B. Bolgomol ’nyi, Sov. J. Nucl. Phys. 24, 449 (1976).\n[7]R. Rajaraman, Solitons and Instantons (North-Holand,\nAmsterdam, 1982).\n[8]A. Vilenkin and E. P. S. Shellard, Cosmic Strings and\nOther Topological Defects (Cambridge University, Cam-\nbridge, England, 1994).\n[9]S. Coleman, Nucl. Phys. B262, 263 (1985).\n[10]T. D. Lee and Y. Pang, Phys. Rep. 221251 (1992).\n[11]H. Arodz, Nucl. Phys. B450, 174 (1995).\n[12]A. Strumia and N. Tetradis, Nucl. Phys. B542, 719\n(1999).\n[13]C. Csaki, J. Erlich, C. Grojean, and T. J. Hollowood,\nNucl. Phys. B584, 359 (2000).\n[14]M. Gremm, Phys. Lett. B 478, 434 (2000).13\n[15]A. de Souza Dutra and A. C. Amaro de Faria, Jr., Phys.\nLett. B642, 274 (2006).\n[16]D. Bazeia, W. Freire, L. Losano, and R. F. Ribeiro, Mod.\nPhys. Lett. A 17, 1945 (2002).\n[17]A. Campos, Phys. Rev. Lett. 88, 141602 (2002).\n[18]A. de Souza Dutra, Phys. Lett. B 626, 249 (2005).\n[19]V. I. Afonso, D. Bazeia, and L. Losano, Phys. Lett. B 634,\n526 (2006).\n[20]M. Cvetic and H. H. Soleng, Phys. Rep. 282, 159 (1997).\n[21]L. J. Boya and J. Casahorran, Phys. Rev. A 39, 4298\n(1989).\n[22]D. Bazeia, M. J. dos Santos, and R. F. Ribeiro, Phys.\nLett. A208, 84 (1995).\n[23]A. de Souza Dutra and R. A. C. Correa, Phys. Lett. B\n679, 138 (2009).\n[24]A. de Souza Dutra and R. A. C. Correa, Phys. Lett. B\n693, 188 (2010).\n[25]C. Adam, L. A. Ferreira, E. da Hora, A. Wereszczynski,\nand W.J. Zakrzewski, J. High Energy Phys. 1308(2013)\n062.\n[26]L. A.Ferreira and W. J. Zakrzewski, J. High Energy Phys.\n1309(2013) 097.\n[27]G. P. de Brito, R. A. C. Correa, and A. de Souza Dutra,\nPhys. Rev. D 89(2014) 6, 065039.\n[28]A. A. Izquierdo, J. M. Guilarte, N. G. de Almeida, J.\nPhys. B 48(2015) 015501.\n[29]I.L.BogolyubskyandV.G.Makhankov, Pis’maZh.Eksp.\nTeor. Fiz. 24, 15 (1976).\n[30]M. Gleiser, Phys. Rev. D 49, 2978 (1994).\n[31]M. Gleiser and J. Thorarinson, Phys. Rev. D 79, 025016\n(2009).\n[32]N. Graham, Phys. Rev. Lett. 98, 101801 (2007).\n[33]E. W. Kolb and I. I. Tkachev, Phys. Rev. D 49, 5040\n(1994).\n[34]N. Graham and N. Stamatopoulos, Phys. Lett. B 639, 541\n(2006).\n[35]E. J. Copeland, M. Gleiser, and H. -R. M¨ uller, Phys. Rev.\nD52, 1920 (1995).\n[36]M. A. Amin and D. Shirokoff, Phys. Rev. D 81, 085045\n(2010).\n[37]H. Segur and M. D. Kruskal, Phys. Rev. Lett. 58, 747\n(1987).\n[38]G. Fodor, P. Forg´ acs, Z. Horv´ ath, and M. Mezei, Phys.\nLett. B674, 319 (2009).\n[39]M. P. Hertzberg, Phys. Rev. D 82, 045022 (2010).\n[40]M. Gleiser and R. C. Howell, Phys. Rev. Lett. 94, 151601\n(2005).\n[41]P. M. Saffin and A. Tranberg, J. High Energy Phys. 01,\n30 (2007).\n[42]T. Romanczukiewicz and Ya. Shnir, Phys. Rev. Lett. 105,\n081601 (2010).\n[43]R. A. C. Correa and A. de Souza Dutra, arXiv:1411.1795.\n[44]K. McQuighan and B. Sandstede, Nonlinearity 27(2014)\n3073.\n[45]S.-Y. Zhou, E. J. Copeland, R. Easther, H. Finke, Z.-G.\nMou, and P. M. Saffin, J. High Energy Phys. 10(2013)\n026.\n[46]V. Achilleos, F. K. Diakonos, D. J. Frantzeskakis, G.\nC. Katsimiga, X. N. Mainta, E. Manousakis, C.E.\nTsagkarakis, and A. Tsapalis, Phys. Rev. D 88(2013)\n045015.\n[47]P. M. Saffin, P. Tognarell, and Anders Tranberg, J. High\nEnergy Phys. 08(2014) 125.\n[48]F. K. Diakonos, G. C. Katsimiga, X. N. Maintas, and C.\nE. Tsagkarakis, Phys. Rev. E 91023202 (2015).\n[49]A.KosteleckyandS.Samuel, Phys.Rev.D 39, 683(1989).[50]D. Colladay and V. A. Kostelecky, Phys. Rev. D 55, 6760\n(1997).\n[51]L. Maccione, S. Liberati, and G. Sigl, Phys. Rev. Lett.\n105, 021101 (2010).\n[52]H. Belich, E. O. Silva, M. M. Ferreira, Jr., and M. T. D.\nOrlando, Phys. Rev. D 83, 125025 (2011).\n[53]H. Belich, M. M. Ferreira, Jr., and J. A. Helayel-Neto,\nEur. Phys. J. C 38, 511 (2005).\n[54]J. L. Boldo, J. A. Helay¨ el-Neto, L. M. de Moraes, C. A.\nG. Sasaki, and V. J. V´ asquez Otoya, Phys. Lett B 689,\n112 (2010).\n[55]V. A. Kostelecky and Matthew Mewes, Phys. Rev. D 85,\n096005 (2012).\n[56]A. de Souza Dutra, M. Hott, and F. A. Barone, Phys.\nRev. D74, 085030 (2006).\n[57]A. de Souza Dutra and R. A. C. Correa, Phys. Rev. D 83,\n105007 (2011).\n[58]R.A.C.Correa, Rold˜ aodaRocha, andA.deSouzaDutra,\narXiv:1501.02000.\n[59]R. Casana, M. M. Ferreira Jr., E. da Hora, and C. Miller,\nPhys. Lett. B 718, 620 (2012).\n[60]M. V. Romalis, D. Sheng, B. Saam, and T. G. Walker,\nPhys. Rev. Lett. 113(2014) 18, 188902.\n[61]L. Zhou and B.-Q. Ma, Adv. in High Energy Phys. 2014\n(2014) 374572.\n[62]R.Casana, M. M.Ferreira, Jr., E.daHora, A.B.F.Neves,\nEur. Phys. J. C 74(2014) 9, 3064.\n[63]J. I. Aranda, F. Ram´ ırez-Zavaleta, D. A. Rosete, F. J.\nTlachino, J. J. Toscano and E. S. Tututi, J. Phys. G:\nNucl. Part. Phys. 41(2014) 055003.\n[64]F. Tamburini and M. Laveder, Phys. Scr. 85(2012)\n035101.\n[65]J. Polchinski, Class. Quantum Grav. 29(2012) 088001.\n[66]K. Bakke and H. Belich, Annals Phys. 354, 1 (2015).\n[67]A.Connes, M.R. Douglas, andASchwarz, J. HighEnergy\nPhys.02, 003 (1998).\n[68]S. M. Carrol, J. A. Hervey, V. A. Kostelecky, C. D. Lane,\nand T. Okamoto, Phys. Rev. Lett. 87, 141601 (2001).\n[69]G. T. Zatsepin and V. A. Kuzmin, Zh. Eksp. Teor. Fiz. 4,\n114 (1966).\n[70]M. Takeda et. al., Phys. Rev. Lett. 81, 1163 (1998).\n[71]S. R. Coleman and S. L. Glashow, Phys. Rev. D 59,\n116008 (1999).\n[72]J. D. Bekenstein, Phys. Rev. D 70, 083509 (2004).\n[73]O. Bertolani, R. Lehnert, R. Potting, and A. Ribeiro,\nPhys. Rev. D 69, 083513 (2004).\n[74]O. Bertolani, Gen. Relativ. Gravit. 34, 707 (2002).\n[75]S. Kanno and J. Soda, Phys. Rev. D 74, 063505 (2006).\n[76]M. A. Amin, R. Easther, and H. Finkel, J. Cosmol. As-\ntropart. Phys. 12, 001 (2010).\n[77]M. A. Amin et. al., Phys. Rev. Lett. 108, 241302 (2012).\n[78]V.A.KosteleckyandR.Lehnert, Phys.Rev.D 63, 065008\n(2001).\n[79]R. Potting, Phys. Rev. D 85, 045033 (2012).\n[80]M. A. Anacleto, F. A. Brito, and E. Passos, Phys. Lett B\n694, 149 (2010).\n[81]C.S.Gardner, J. Greene, M.Kruskal, andR.Miura, Phys.\nRev. Lett. 19, 1095 (1967).\n[82]M. Gleiser, Phys. Lett. B 600, 126 (2004).\n[83]M.Gleiser, N.Graham, andN.Stamatopoulos, Phys.Rev.\nD83, 096010 (2011).\n[84]M. A. Amin, R. Easther, H. Finkel, R. Flauger, and M.\nP. Hertzberg, Phys. Rev. Lett. 108, 241302 (2012).\n[85]M. A. Amin, M. P. Hertzberg, D. I. Kaiser, and J.\nKarouby, Int. J. Mod. Phys. D 241530003 (2015).14\n/Minus1000/Minus500 500 1000x\n/Minus0.02/Minus0.010.010.02/CurlyPΗi/LParen1x,0/RParen1\n/Minus1000/Minus500 500 1000x\n/Minus0.02/Minus0.010.010.02/CurlyPΗi/LParen1x,1250/RParen1\nFIG. 1: Profile of the usual oscillons in 1+ 1-dimensions with Lorentz and CPTbreaking for t= 0 (left) and t= 1250 (right)\nwithǫ= 0.01. The thin line corresponds to the case with k00= 0.12,k11= 0.30,k01= 0.27 andk10= 0.21 and the thick line to\nthe case with kµν= 0.\n/Minus60/Minus40/Minus20 204060x\n/Minus0.50.5/CurlyPΗi/LParen1x,0/RParen1\n/Minus100/Minus50 50 100x\n/Minus1.0/Minus0.50.51.0/CurlyPΗi/LParen1x,200/RParen1\nFIG. 2: Profile of the Flat-Top oscillons in 1+1-dimensions w ith Lorentz symmetry breaking for t= 0 (left) and t= 200 (right)\nwithg= 5. The thin line corresponds to the case with k00= 0.12,k11= 0.30,k01= 0.27 andk10= 0.21 and the thick line to\nthe case with kµν= 0.\nxt/CurlyPhi/LParen1x,t/RParen1\nx xt/CurlyPhi/LParen1x,t/RParen1\nx\nFIG. 3: Typical profile of the Flat-Top oscillon. The left-ha nd figure corresponds to the case with Lorentz breaking symme try\nand the right-hand figure to the one with Lorentz symmetry.15\n/Minus6/Minus4/Minus2 246x\n/Minus0.50.51.01.52.02.53.0/CurlyPΗi/LParen1x,0/RParen1\n/Minus6/Minus4/Minus2 2 4 6x\n/Minus3/Minus2/Minus112/CurlyPΗi/LParen1x,10/RParen1\nFIG. 4: Profile of the Breathers 1+1-dimensions with Lorentz symmetry breaking for t= 0 (left) and t= 10 (right) with v= 2,\nw= 1,β= 1. The thin line corresponds to the case with k00= 0.28,k11= 0.30,k01= 0.27 andk10= 0.37 and the thick line to\nthe case with kµν= 0.\n-2024\nx05101520t\n-4-2024\nx05101520t\nFIG. 5: Density plot of a Breather. Solution with Lorentz sym metry breaking (left) and to the one Lorentz symmetry (right ).\n50100150200250300350x\n/Minus4./Multiply10/Minus9/Minus2./Multiply10/Minus92./Multiply10/Minus94./Multiply10/Minus9Η/LParen1x,t/RParen1\n1020304050x\n/Minus6./Multiply10/Minus13/Minus4./Multiply10/Minus13/Minus2./Multiply10/Minus132./Multiply10/Minus134./Multiply10/Minus136./Multiply10/Minus13Η/LParen1x,t/RParen1\nFIG. 6: Amplitude of the outgoing radiation determined by th e Fourier transform. The left-hand figure corresponds to the case\nwith Lorentz breaking symmetry and the right-hand to the cas e with Lorentz symmetry." }, { "title": "1508.05265v1.Radiative_damping_in_wave_guide_based_FMR_measured_via_analysis_of_perpendicular_standing_spin_waves_in_sputtered_Permalloy_films.pdf", "content": "Radiative damping in wave guide based FMR measured via\nanalysis of perpendicular standing spin waves in sputtered\nPermalloy \flms\nMartin A. W. Schoen,1Justin M. Shaw,2Hans T.\nNembach,2Mathias Weiler,3and Thomas J. Silva2\n1Institute of Experimental and Applied Physics,\nUniversity of Regensburg, 93053 Regensburg, Germany\n2Electromagnetics Division, National Institute of\nStandards and Technology, Boulder, CO, 80305\n3Walther-Meiner-Institut, Bayerische Akademie\nder Wissenschaften, D-85748 Garching, Germany\n(Dated: August 20, 2021)\nAbstract\nThe damping \u000bof the spinwave resonances in 75 nm, 120 nm, and 200 nm -thick Permalloy\n\flms is measured via vector-network-analyzer ferromagnetic-resonance (VNA-FMR) in the out-of-\nplane geometry. Inductive coupling between the sample and the waveguide leads to an additional\nradiative damping term. The radiative contribution to the over-all damping is determined by\nmeasuring perpendicular standing spin waves (PSSWs) in the Permalloy \flms, and the results are\ncompared to a simple analytical model. The damping of the PSSWs can be fully explained by\nthree contributions to the damping: The intrinsic damping, the eddy-current damping, and the\nradiative damping. No other contributions were observed. Furthermore, a method to determine\nthe radiative damping in FMR measurements with a single resonance is suggested.\nContribution of NIST, not subject to copyright\n1arXiv:1508.05265v1 [cond-mat.mtrl-sci] 21 Aug 2015I. INTRODUCTION\nExcited magnetic moments relax towards their equilibrium orientation due to damp-\ning. Several physical mechanism can cause damping. Many mechanisms, like eddy current\ndamping1in conducting ferromagnets, were already identi\fed in the 1950s. More recently,\nenhanced damping due to spin pumping2from a ferromagnet into an adjacent metallic layer\nwas identi\fed, and remains a topic of ongoing investigation3{6. Furthermore, wavenumber-\ndependent contributions to the damping caused by intralayer spin pumping have been theo-\nretically predicted7,8and currently are the subject of experimental investigation9,10. Another\ndamping process, referred to as radiative damping11,12, has been known to exist since the\n1970s and is purely due to inductive coupling between the sample and the waveguide in ferro-\nmagnetic resonance (FMR) experiments. More recently, this phenomenon has been further\ninvestigated in the context of strong magnon-photon coupling experiments, with possible\napplications in quantum information processing13. In these quantum-coherent experiments,\nradiative damping was identi\fed as a manifestation of non-resonant magnon-photon cou-\npling, and it was determined that such coupling14is indeed a source of extrinsic line width\nin cavity-based FMR studies. In a radiative damping process, the time-varying magnetic\n\rux associated with the dynamic magnetization generates microwave-frequency currents in\nthe proximate conductor of a microwave waveguide that carries the resultant power away\nfrom the sample. This process is similar to that exploited in eddy-current brakes and can\nbe seen as a non-local counterpart to the eddy-current damping in conductive ferromagnets.\nTo determine the magnitude of radiative damping in magnetic thin \flms, we used broad-\nband vector-network-analyzer ferromagnetic resonance (VNA-FMR) to measure damping in\nNi0:8Fe0:2Permalloy (Py) \flms with thicknesses \u000evarying between 70 nm and 200 nm. By\nuse of the geometry sketched in Fig. 1(a), we determine the total damping for each mode\n\u000bnas a sum of intrinsic damping \u000bint, eddy-current damping \u000beddyand radiative damping\n\u000brad\nn. We then perform a quantitative analysis of the PSSW resonance \felds, amplitudes and\ndamping to extract the di\u000berent contributions to \u000bn. We \fnd that eddy current damping is\nonly signi\fcant for the lowest order mode, the radiative damping strongly a\u000bects the \frst\n\fve modes, and no additional contributions to the damping are detectable for spin waves up\ntok= 1:75\u0002106cm\u00001. This last \fnding is in contrast with reports of exchange mediated\ndamping in both nanostructures10and thin \flms9.\n2II. DAMPING MODELS\nAccording to Faraday's Law, the time-varying \rux of a precessing magnetic moment\ngenerates an ac voltage in any conducting material that passes through the \rux. As shown\nin Fig. 1, spin wave precession in a conducting ferromagnet on top of a coplanar waveguide\n(CPW) induces ac currents both in the ferromagnet and the CPW. The dissipation of these\neddy currents in the sample, and the \row of energy away in the CPW give rise to two\ncontributions to magnetic damping. Historically, the damping caused by eddy currents in\nthe ferromagnet \u000beddyis called eddy current damping, while the damping caused by the eddy\ncurrents in the waveguide is called radiative damping \u000brad\nn.\nEddy current damping has been recognized since the 1950s1,15. For the lowest order mode\nin FMR,\n\u000beddy=C\n16\r\u00162\n0Ms\u000e2\n\u001a; (1)\nwith the resistivity \u001a, saturation magnetization Ms, the vacuum permeability \u00160, the gyro-\nmagnetic ratio \r, and the sample thickness \u000e(see derivation in Appendix Sec. B). We\nintroduce a correction factor Cto account for details of the eddy current spatial pro\fle. As\nshown in a later section, \u000beddyfor all higher order PSSW modes investigated in this study\nis much smaller than that of the lowest order mode.\nWe now turn to the radiative damping \u000brad\nn. We consider the experimental geometry\nsketched in Fig. 1 (a). A ferromagnetic sample with thickness \u000eand length lis placed on\ntop of the center conductor of a coplanar waveguide with width W. The sample dimension\nalongxis much larger than W. The sample and CPW are separated by a gap of height d.\nAn external dc magnetic \feld H0is applied perpendicular to the sample plane, and the spin\nwave resonances (SWR) are driven by microwaves in the CPW at resonance frequency f. A\nfraction of the ac magnetic induction Bdue to the dynamic component of the magnetization\nmx\nn(H0;I;z) wraps around the center conductor.\nTo derive a quantitative expression for \u000brad\nn, we start by calculating hx(I;x;z), the x\ncomponent of the driving \feld hmwthat is generated by an excitation current Iin the\ncenter conductor. We assume hx(I;x;z) is uniform along y, but we allow for variation along\nxandz. To estimate hx(I;x;z) we use the Karlqvist equation16\nhx(I;x;z) =I\n2\u0019W\u0014\narctan\u0012x+W=2\nz\u0013\n\u0000arctan\u0012x\u0000W=2\nz\u0013\u0015\n: (2)\n3n=0 123\nxz\nwdδ\nq31 -1z\nH\n0Mm\nx\nBdl\nδxyz\nCPW\nw(a)\n(b)h\nmwFigure 1: Schematic of the radiative damping process. (a) Mis the dynamic magnetization, H0\nthe applied external \feld and Bis the magnetic inductance due to the x-component mxof the\ndynamic magnetization. Wis the width of the center conductor, lthe length of the sample on the\nwaveguide, \u000ethe thickness of the sample and dthe spacing between sample and wave guide. (b)\nSimpli\fed depiction of the PSSW eigenfunctions qnfor mode numbers n= 0, 1, 2, 3. We exemplarily\nused boundary conditions that are completely pinned on one side and completely un-pinned on the\nother side. The origin of the coordinate system is indicated.\n4This microwave \feld can excite PSSWs in the sample. Schematic mode pro\fles for the\nfundamental mode ( n= 0) and the \frst three PSSW modes are shown in Fig. 1(b), where\nwe use unpinned boundary conditions at the top surface and pinned boundary conditions at\nthe bottom surface. As shown in Fig. 1(b), the mode pro\fles describe a z-dependence of the\ndynamic magnetization components mxandmy. In the perpendicular geometry used here,\njmxj=jmyjeverywhere, i.e. the precession is circular. In what follows, we will only discuss\nmx, the dynamics of which are inductively detected in the measurement. For a PSSW with\nmode number n,emx\nn(x;z) =qn(z)\u001fnhqn(z)hx(I;x;z)iwherehidenotes spatial averaging\ninxandzdirections, as de\fned in the Appendix, \u001fn=\u001fxx\nnis the diagonal component of\nthe magnetic susceptibility of the n-th order mode, and \u00001\u0014qn(z)\u00141 is the normalized\nmode pro\fle (eigenmode), an example of which is sketched in Fig. 1 (b) for n= 3. The\nmode inductance Lnis given by Ln=\u001fneLn, where, as detailed in the Appendix, we de\fne\na normalized mode inductance eLnfor thenth PSSW mode,\neLn=\u00160l\nI2hqn(z)hx(I;x;z)i2W\u000e: (3)\neLn, as explained in the appendix, no longer has any dependence on magnetic \feld or\nexcitation frequency. In the simplest case of a uniform magnetization pro\fle q0(z) = 1\n(FMR-mode) and uniform excitation \feld hx(I;x;z) =hx(I; 0;0) =I=(2W), the normalized\ninductance is eLn=\u00160\u000el=(4W).\nThex-component of the dynamic magnetization mx\nn(x;z) produces a net \rux \b n=\u001fnIeLn\nthat threads around CPW center conductor, leading to a power dissipation\nPn=!2\n2Z0\u0010\n\u001fnIeLn\u00112\n; (4)\nwhereZ0is the waveguide impedance (in our case Z0= 50 \n) and !is the angular frequency\nof the magnetization precession. With Eq. (4), the power dissipation rate\u0010\n1\nT1\u0011\nn=Pn=En\ncan be calculated, where Enis the energy of the dynamic component of the magnetization\nderived in the Appendix. This power \row from the sample to the waveguide leads to the\nradiative damping contribution\n\u000brad\nn=1\n2!\u00121\nT1\u0013\nn=\u0011\r\u0016 0MseLn\nZ0;\n(5)\n5where\u0011=\u000e=\u0010\n4R\u000e\n0dzjqn(z)j2\u0011\nis a dimensionless parameter that accounts for the actual\nmode pro\fle in the sample, see Appendix Sec. A. In the case of sinusoidal PSSWs, \u0011= 1=2,\nand for a completely uniform mode pro\fle, i.e. qn(z) = 1,\u0011= 1=4. From Eq. (5), it is\nevident that \u000brad\nnis proportional to eLnforn > 0. In the simplest case of uniform driving\n\feldhx=I=(2W), the radiative contribution is given by\n\u000brad\n0=\u0011\r\u0016 0Ms\nZ0eLn\u0018=\u0011\rM s\u00162\n0\u000el\n2Z0W: (6)\nNote that the radiative damping thus depends on the sample and waveguide dimensions, in\nparticular linearly on the sample thickness. Unlike eddy-current damping, \u000brad\nnis indepen-\ndent of the conductivity of the ferromagnet, hence this damping mechanism is also operative\nin ferromagnetic insulators.\nIII. SAMPLES AND METHOD\nWe deposit Ta(3)/Py( \u000e)/Si 3N4(3), Ta(3)/Py( \u000e)/Ta(5), Ta(3)/Py( \u000e) and Py(\u000e) layers on\n100\u0016m thick glass substrates by DC magnetron sputtering at a Ar pressure of 0 :7 Pa (\u0019\n5\u000110\u00003Torr) in a chamber with a base-pressure of less than 5 \u000110\u00006Pa (\u00194\u000110\u00008Torr);\nwhere\u000e= 75 nm, 120 nm and, 200 nm is the Permalloy thickness. The Py thickness was\ncalibrated by x-ray re\rectivity. We estimate that the damping enhancement due to spin\npumping into the Ta layer is two orders of magnitude smaller than the intrinsic damping of\nthe Permalloy layer for Permalloy samples of these thicknesses. The various combinations of\ncapping and seed layers are chosen to determine the sensitivity of our results on the spinwave\nboundary conditions and the resultant mode pro\fles. Prior to deposition, the substrates are\ncleaned by Ar plasma sputtering. The samples are coated with approximately 150 nm of\nPMMA in order to avoid electrical shorting when samples are placed directly on the CPW.\nThe CPW has a center conductor width of W= 100 \u0016m. The SWR are characterized using\n\feld-swept VNA-FMR17{19in the out-of-plane geometry (see Fig. 1) with an external static\nmagnetic \feld H0applied perpendicular to the sample plane. The excitation microwave \feld\nhx(x;y) is applied over a frequency range of 10 GHz to 30 GHz. A VNA is used to measure\nthe complex S21transmission parameter (ratio of voltage applied at one end of the CPW\nto voltage measured at the other end) for the waveguide/sample combination. The change\ninS21due to the FMR of the sample is then \ftted with a linear superposition of complex\n6susceptibility tensor components \u001fn,\n\u0001S21\nn(H0) =NX\nn=0An\u001fn(H0)ei\u001en+ linear background (7)\nwith the mode number n, phase\u001en, and dimensionless mode amplitude An, as de\fned in the\nAppendix. A complex linear background and o\u000bset is included in the \ft. The susceptibility\ncomponents are derived from the Landau-Lifshitz equation for the perpendicular geometry;\nin the \fxed-frequency, swept-\feld con\fguration, we obtain20\n\u001f(H0)n=Ms(H0\u0000Me\u000b\nn\u0000Hex\nn)\n(H0\u0000Me\u000b\nn\u0000Hex\nn)2\u0000(He\u000b)2\u0000i\u0001Hn(H0\u0000Me\u000b\nn\u0000Hex\nn)(8)\nwithHe\u000b=!=(\r\u00160) andMe\u000b\nn=Ms\u0000Hk, whereHkis the perpendicular anisotropy, Hex\nnis\nthe exchange \feld (de\fned below), and \u0001 Hnis the linewidth. An example of the resulting\n\fts for the complex S21data is shown in Fig. 2 (a) and (b).\nIV. EXPERIMENT\nWe detect both even and odd PSSW modes. If we assume a uniform excitation \feld\nand Dirichlet boundary conditions (completely pinned), only odd modes would be detected.\nAlternatively if we assume Neumann boundary conditions (completely unpinned), only the\nfundamental mode would be detected.\nTwo e\u000bects can contribute to our ability to detect all the PSSW modes. First, the\nexcitation \feld pro\fle might not be uniform due to eddy current shielding21,22. Second, the\ninterfacial boundary conditions might be asymmetrical, as alluded to above. According to\nthe criterion in Ref. [ 22], the threshold sheet resistance for the onset of eddy current shielding\nat 20 GHz is 0 :065 \n=\u0003. We estimate that the sheet resistance for our 200 nm is in excess of\n0:345 \n=\u0003, so we conclude that the eddy current shielding is relatively weak for our samples.\nOn the other hand, all modes are in principle detectable if we assume asymmetric in-\nterfacial anisotropy. For the sake of simplicity of the analysis, we will assume interfa-\ncial anisotropy for a single interface, then use an optimization approach to determine the\nwavenumber of the modes that is consistent with such a hypothesis. However we must\nemphasize that this approach does not provide a unique \ft for the measured distribution\nof resonance \felds for the PSSW spectrum, but simply allows us to accommodate for the\nwavenumber values required to be consistent with the measured spectrum. As such, the\n71.51 .61 .7-0.20-0.15-0.10-0.05-0.45-0.40-0.35-0.300\n1x1062x1060.00.51.0(\nb)4\n235 10 \nIm(S21)µ\n0H0 [T] 53412 \nRe(S21)0\n(a) \n(c)µ0H exn\n[T] \nk\nn [cm-1]Figure 2: Measured S21transmission parameter (black circles) at 20 GHz and the multi-peak\n-susceptibility \ft (red line) for the (a) real part and (b) imaginary part obtained with the Ta(3)-\nPy(200)-Si 3N4(3) sample. The \frst 6 modes are shown. (c) The exchange \feld Hex\nn(black squares)\nand exchange \feld \ft, from Eqs. (9) and (10) (red crosses) for all 13 detected modes plotted as a\nfunction of the \ftted wave numbers kn.\n\ftted value for Ksis to be interpreted as no more than a self-consistent value associated\nwith only one of many possible scenarios.\nIf we assume negligible magnetocrystalline perpendicular anisotropy Hk,Hres\nnis related\nto the exchange \feld via\nHres\nn=Hex\nn+Ms;\nwithHex\nn=2Aex\n\u00160Msk2\nn:(9)\nHere,knis the spinwave wavevector, and Aexis the exchange energy that is related to the\nspinwave sti\u000bness DviaD=2Aexg\u0016B\nMs. On the other hand, if we want to include interfacial\nanisotropy for a single interface in our analysis, we can numerically solve the transcendental\nequation23\u0012\n\u00001\n2kna+Ks\n2Aexkn+ 1\u0013\ntan(kn\u000e) =Ks\n2Aexkn; (10)\n8whereKsis the interfacial anisotropy, and a= 0:3547 nm is the lattice constant24. We\nminimize the residue of the \ft of Eq. (9) to Hres\nnwith the \ftting parameters Ms,Aex, and\nKsfrom Eq. (10) by use of a Levenberg-Marquardt optimization algorithm. This yields the\npairs (kn,Hex\nn) shown in Fig.2 (c) for all modes.\nFrom the \ft, we obtain a saturation magnetization of \u00160Ms= 1:02\u00060:01 T, in agree-\nment with that determined by magnetometry. The exchange sti\u000bness constant of D=\n3:22\u00060:04 meVnm2is close to a value of D\u00193:1 meVnm2reported by Maeda et al.25.\nThe exchange \ft also yields a single surface anisotropy Ksthat depends on the\ncap and seed layer con\fgurations. For the Ta(3)-Py( \u000e)-Si 3N4(3) sample series,\nKs= (5:1\u00060:8)\u000210\u00004J=m2, while all the other samples have a higher Ksof\n(7\u00061)\u000210\u00004J=m2. All values for Ksare in the range of other reported interface\nanisotropies for Permalloy layers of these thicknesses26.\nWe now turn to the linewidth \u0001 Hnand the amplitude Anfor the individual modes. The\nGilbert damping parameter \u000bnis extracted from the slope of the linewidth vs. frequency f\nplot10shown in Fig. 3(a) via\n\u0001Hn=4\u0019\u000bnf\nj\rj\u00160+ \u0001H0\nn; (11)\nwhere \u0001H0\nnis the inhomogeneous broadening that gives rise to a nonzero linewidth in the\nlimit of zero frequency excitation. The normalized inductance of the modes eLnis extracted\nin a similar fashion from the dependence of the mode amplitude Anon the frequency f, see\nFig. 3(b) and Eq. (38) in the Appendix\nAn= 2\u0019feLn\nZ0+A0\nn; (12)\nwhereA0\nnis an o\u000bset for each mode. A0\nnis a phenomenological \ftting parameter, which is\nnot yet fully understood.\nWe plot\u000bnandeLnas a function of mode number nin Fig. 3 (c). The damping and\nthe normalized mode inductance are found to be proportional. In order to explore this\ncorrelation, we plot \u000bnvs.eLnin Fig. 4(a). Here, the data for \u000bnvs.eLnare linearly\ncorrelated for all modes except for n= 0, as seen by the linear \ft (line) to the data for\nn\u00151. This is as expected for the radiative damping model, as summarized in Eq. (5). The\nadditional damping of the fundamental mode is interpreted as the result of eddy current\ndamping, as quanti\fed in Eq. (1). In Fig. 4(b), we plot the residual \u0001 \u000bnof the linear \ft\n901 2 3 4 0.0070.0080.0090.0100.011~ α( c)Ln [pH]αnm\node number n0n0\n123 \nLn \n01 02 03 0010203040n4213\nµ0ΔH [mT]f\n [GHz](a)0\n4\n213\n0\n1 02 03 0-0.0020.0000.0020.0040.0060.0080.010 (b) \n f\n [GHz]AnFigure 3: Parameter extraction for the \frst 5 PSSWs of the Ta(3)-Py(200)-Si 3N4(3) sample. (a)\nExtraction of \u000bfrom the linewidth \u00160\u0001H(data points) via linear \fts (lines); staggered for display.\n(b) Extraction of the normalized mode inductance eLn(data points) from the resonance amplitude\nAn(f) via linear \fts (lines). (c) \u000bn(black squares) and eLn(red crosses) for each PSSW.\nshown in Fig. 4(a) for all modes. \u0001 \u000bnis negligible for all modes except for n= 0. We\nextract \u0001\u000bn=0for all the samples and plot \u0001 \u000bn=0vs.\u000e2in Fig. 4(c).\nIt appears that \u0001 \u000bn=0for all the samples scales linearly with \u000e2, as expected from Eq. (1)\nfor eddy current damping. Simultaneous weighted \fts of all the data to Eq. (1) yields\nC= 0:4\u00060:1. This value suggests a localization of eddy currents, since Ccorrects for the\n1001 2 3 4 0.00000.00030.00060.00090.00.51.01.52.02.53.00.0070.0080.0090.0100.0110\n1 x1042x1043x1044x1040.00000.00030.00060.0009 \n Δαnn\n(b)~4321n =0 \nαnL\nn[pH](a) \n Δαn=0δ\n2[nm2](c)Figure 4: Damping \u000bnand inductance eLnfor the Ta(3)-Py(200)-Si 3N4(3) sample. (a) Linear \ft\nof\u000bto Eq. (5) where the \ft is constrained to the n= 1, 2, 3, 4 modes. (b) The residual of the\nlinear \ft, showing enhanced damping for the 0-th order mode (black). We attribute the enhanced\ndamping to an eddy current contribution. (c) Enhanced 0-th order mode damping for all samples.\nThe red line is a \ft of the data points to the eddy current damping model from Eq. (1).\neddy current distribution in the sample.\nFor then\u00151 modes, it can be shown27that\u000beddy\nn/1=k2\nn. The calculated wavevectors\nfrom Eq. (10) for the n= 1 mode of all the samples is at least a factor three larger than that\nof then= 0 mode and, therefore, the eddy current damping of the n= 1 mode is predicted\nto be approximately one order of magnitude smaller than the eddy current damping of the\nn= 0 mode. Thus, the eddy current damping of the n\u00151 modes is negligible to within\n11the error bars, i.e., \u000beddy\nn\u00190 forn\u00151. This supports the analysis of the data in Ref. [ 9],\nwhich also neglects the eddy current damping in higher order modes.\nV. EXTRACTION OF THE RADIATIVE CONTRIBUTION TO THE DAMPING\nBy use of Eq. (1) and our \ftted value of C= 0:4, we subtract the eddy current contri-\nbution to the damping of all the n= 0 modes to obtain a corrected damping value \u000b0\nn=0,\nwhere\u000b0\nn=0=\u000bn=0\u0000\u000beddy(C= 0:4). The corrected data for all the modes are plotted in\nFig. 5.\nFigures 5 (a) to (c) group all data obtained for a set of samples with identical Py thickness\n\u000e. The lines are linear \fts to Eq. (5). For each thickness \u000e, we observe a signi\fcant correlation\nof\u000bnandeLnfor all seed and cap layer con\fgurations, as expected for a radiative damping\nmechanism.\nFurthermore, by use of Eq. (6) for the n= 0 mode of the 75 nm thick sample, using a\nvalue of\u0011\u00190:46 as determined in the Appendix, we estimate \u000brad\n0\u00190:00023\nThe experimentally determined value is \u000brad\n0\u00190:00035\u00060:0001. The deviance from\nthe calculated value is possibly due to non-uniformities of both the excitation \feld and\nmagnetization pro\fle in Eq. (6), that requires the solution of the integral in Eq. (25).\nNevertheless the estimated value for \u000brad\n0is of the correct order of magnitude.\nWe determine the intrinsic damping \u000bintfrom theeLn= 0 intercept of the linear \fts in\nFig. 5. We plot \u000bintfor the three values of \u000ein Fig. 6 (right scale). We \fnd that \u000bintis\napproximately constant to within \u00065% for all samples. In addition the average value over\nall the \flm thicknesses is in reasonable agreement to the previously reported value of \u000bint=\n0.006 (dotted red line)28.\nThe other \ftting parameter \u0011, extracted from the slope of \u000bnvs.eLn, is also plotted in\nFig. 6 (left scale). For anti-symmetric boundary conditions, \u0011= 1=2 is expected, whereas\nfor the uniform mode, \u0011= 1=4.\nWe see that the \ftted values lie exclusively within these extremes, within error bars.\nThere have been recent reports of a non-zero, wavenumber-dependent component for\ndamping for both localized eigenmodes in magnetic nanostructures10and PSSWs in thick\n120.00.51.01.52.02.53.00.0060.0070.0080.0090.010η\n=0.24~\n(c)δ\n = 75 nmαnαn \n L\nn[pH]0.0060.0070.0080.0090.010η\n=0.36(b)δ\n = 120 nm \n 0.0060.0080.0100.012(\na)δ\n = 200 nmαn \nη=0.45Figure 5: Dependence of damping \u000bon the normalized mode inductance eLnafter correction for\nthe eddy current damping \u0001 \u000b0of the fundamental mode, plotted for all sample con\fgurations\nand di\u000berent thicknesses: (a) \u000e= 200 nm, (b) \u000e= 120 nm and (c) \u000e= 75 nm. The red lines are\nweighted linear \fts to the data by use of Eq. (5), that describes the radiative component of the\ndamping.\n13501 001 502 000.00.10.20.30.40.50.6u\nniform ηδ\n [nm]αintsinusoidal0\n.0000.0020.0040.0060.0080.010 \nFigure 6: Mode pro\fle parameter \u0011(black squares, left axis) and intrinsic damping \u000bint(red circles,\nright axis) as a function of Py thickness \u000e. The mode pro\fle parameter \u0011lies between the value\nof 1=2 for sinusoidal PSSWs with anti-symmetric boundary conditions (dashed black line) and the\nvalue of 1=4 for the uniform mode (both dashed black lines). The intrinsic damping is close to\n\u000bint= 0:006 (dotted red line).\nPermalloy \flms9. Such exchange-mediated damping of the form \u000bex:=Aexk2was originally\npredicted by Baryakhtar based on symmetry alone8. Nembach, et al.10, obtained a value of\nAex= 1:4 nm\u00002, whereas Li, et al.8, found a much smaller value of 0 :09 nm\u00002. To determine\nwhether wavenumber-dependent damping is apparent in our data, we examined the residual\ndamping after subtraction of both the intrinsic damping \u000bintand the radiative damping \u000brad\nn\nfrom all the modes, as well as subtraction of the eddy current damping from the n= 0 mode.\nThe residual damping \u000bresis plotted in Fig. 7 (b). Within the scatter of \u0018=\u00060:001,\u000bresdoes\nnot have any clear dependence on k. Thus, we obtain an upper bound of Aex\u00140:045 nm\u00002\nfor this particular system, given the sensitivity of our measurements. For comparison and to\nensure that the subtraction of \u000bint,\u000brad\nn, and\u000beddydid not hide a potential k2contribution\nthe measured damping of the Ta(3)-Py(200)-Si 3N4(3) sample up to the n= 10 mode is\nshown in Fig. 7 (a). For n\u00155 the measured damping scatters around the for the 200 nm\nsamples determined intrinsic damping \u000bintand no trend for higher mode numbers (larger k\nvalues) is discernible.\n14Tserkovnyak, et al., calculated the damping coe\u000ecient Aexin terms of a microscopic\nmodel for the di\u000busive transport of dissipative transverse spin current within a ferromagnetic\nmetal2. The theory in Ref. [ 2] framed the exchange-mediated damping in terms of a so-called\ntransverse spin conductivity \u001b?,\nAex=\u0012\r\nMs\u0013\u0012~\n2e\u00132\n\u001b?; (13)\nwhere\n\u001b?:=\u0010\u001b\n\u001c\u0011 \n\u001c?\u0000\n1 + (!ex\u001c?)2\u0001!\n; (14)\nwith the exchange splitting ~!ex, the conductivity \u001b, the spin scattering time \u001c, and\ntransverse spin scattering time \u001c?. Given that ~!ex\u00191 eV for Permalloy, the maximum\nvalue forAexpredicted by the transverse spin current theory is 0 :001 nm\u00002. Insofar as\nwe are not able to observe any such wavenumber-dependent damping down to the level of\n0:045 nm\u00002, our results are consistent with the predictions of the microscopic theory.\nWhile the theory in Ref. [ 9] is speci\fc to the microscopic mechanism of transverse spin\naccumulation in a metallic ferromagnet, the phenomenology of exchange mediated damping,\nas described in Ref. [ 8], is not limited to such a microscopic mechanism. As such, it remains\nplausible that extrinsic material-speci\fc parameters that have not yet been identi\fed could\nbe responsible for the previously reported values for k2damping. For example the presence\nof anti-symmetric exchange at interfaces, i.e. the Dzyaloshinskii-Moriya interaction (DMI)\ncould enhance the coupling between magnons and Stoner-excitations insofar as the DMI\ngives rise to exotic spin textures29with nanometer length scales, that are comparable to\nthe wavelength of low energy Stoner-excitations30. Thus, the results of Ref. [ 10] could be\na manifestation of interfacial enhancement for Aex\nn, insofar as the magnetic \flms used in\nRef. [ 10] are only 10 nm thick.\nIn another experiment, we further validate the presence of radiative damping and demon-\nstrate an alternative method to determine \u000brad\nnby varying the distance din Eq. (2) between\nthe sample and waveguide. To this end, we insert a d= 200 \u0016m glass spacer between the\nsample and waveguide. By comparing h(0;0) toh(0;200\u0016m) via Eq. (2), we estimate that\nthe insertion of the spacer decreases the microwave magnetic \feld by about a factor of 6.25.\nReferring to Eq. (3), the normalized mode inductance eLndecreases by a factor of \u0018=40. To\n1502 4 6 8 1 00.0050.0060.0070.0080.0090.0100.0110\n1 x1062x106-0.0010.0000.0010.002( b) \n αresn\nk\n [cm-1] \nαnn\nαint(a)Figure 7: (a) The measured damping for the \frst 11 PSSW modes of the Ta(3)-Py(200)-Si 3N4(3)\nsample. The enhanced damping due to inductive coupling to the waveguide and eddy currents\nin the sample only a\u000bects the \frst \fve modes at wavevectors \u00147\u0002105cm\u00001. (b) The residual\ndamping for all detected modes for all samples is plotted against their respective wave vector k.\nWithin the scatter, no dependence of the residual damping on kis observed.\ndetermine the e\u000bect of the reduced inductive coupling on the radiative damping, we used\nVNA-FMR to measure the \frst 4 modes for the Ta(3)-Py(120) sample with and without\nthe spacer. The e\u000bect of the spacer can be seen in the raw data, reducing the linewidth\nof the \frst two modes measured at 10 GHz in the 120 nm samples by approximately 6 Oe,\nwell outside error bars. The \ftted values of eLnare shown in Fig. 8 (a). Indeed, eLnde-\ncreases on average for all modes by a factor of \u0018=50 after inserting the spacer, in good\n16agreement with the predictions of Eq. (2) and (3). Thus, we will assume that \u000brad\nnis neg-\nligible when the spacer is used. The data for the damping \u000bnof the \frst four modes, both\nwith and without the spacer, are plotted in Fig. 8 (b). Indeed, the damping determined\nfrom the measurement with the spacer layer (circles) is consistently lower than that found\nwithout the spacer layer (squares). The line in Fig. 8 (b) is the previously determined in-\ntrinsic damping. Under the assumption that the radiative damping contribution is given by\n\u000brad\nn=\u000bn(d= 0)\u0000\u000bn(d= 200 \u0016m), we plot \u000brad\nnvseLn(d= 0) in Fig. 8 (c). The line is\nthe calculated \u000brad\nn, where we used Eq. (5) with \u0011= 0:35 and\u000e= 120 nm, as determined\nfrom the \fts in Fig. 5. Good agreement between the calculated and measured values for \u000brad\nn\nare obtained, which demonstrates the self-consistency of our analysis. Of great importance\nis that the spacer-layer approach can also be used to determine the radiative contribution\nto the damping in the absence of PSSWs (single resonance). By measuring \u000bfor varying\ndistancedbetween sample and waveguide and extrapolating \u000btod!1 , both the intrinsic\nvalue for the damping and the radiative contribution can be determined, under conditions\nwhere eddy current damping is negligible.\nVI. SUMMARY\nIn summary, we identi\fed three contributions to the damping in PSSWs: Intrinsic damp-\ning\u000bint, eddy current damping \u000beddy, and radiative damping \u000brad\nn. The latter exhibits a\nlinear dependence on the normalized sample inductance eLnin a waveguide based FMR\nmeasurement. We attribute this linear dependence to radiative losses that stem from the\ninductive coupling between the sample and the waveguide. The radiative damping term is\ninherent to the measurement process and is thus present in all FMR measurements. The\nradiative damping constitutes up to 40 % of the total damping of the spin wave modes\nin our 200 nm thick Permalloy \flms. Furthermore, the radiative damping can be already\nimportant for much lower \flm thicknesses, in materials with small intrinsic damping.\nAs an example, the radiative damping calculated from Eq. (6) for a 20 nm thick and\n1 cm long sample of Yttrium-Iron-Garnet (YIG), measured on a 100 \u0016m wide wave guide is\n\u000brad\n0\u00191:26\u000110\u00004. When compared to the reported value for the damping of \u000b= 2:3\u000110\u00004,31\nwe see, that the radiative part of the damping, among others32, can substantially in\ruence\nthe determination of \u000bint. As such, careful analysis of \u000bvs. inductance is required to isolate\n1701 2 3 1E-30.010.110\n.00 .51 .01 .52 .00.0000.0010.0020.00301 2 3 0.0050.0060.0070.0080.009~d=200µm \nLn[pH]n\n(a)d=0αradn\n~\n \n L\nn(d=0)[pH](c)αint(b) αn n\nFigure 8: Measurement of the \frst four PSSWs of the Ta(3)-Py(120) sample with and without a\nspacer inserted between sample and CPW. (a) Inductance eLndetermined for the sample directly\non the CPW (black squares) and for a 200 \u0016m spacer between sample and CPW (red circles). (b)\nThe resulting damping constants for both measurements (same symbols and colors). The red line is\nthe previously extracted intrinsic damping \u000bint. (c) The di\u000berence between the damping with and\nwithout the spacer (black squares) is in good agreement with the radiative damping from Fig. 5(b)\n(gray line).\nthe radiative damping contribution.\nVII. AKNOWLEDGEMENT\nThe authors are grateful for the assistance of Mikhail Kostylev in the development of our\ntheoretical analysis.\n18VIII. APPENDIX\nA. Derivation of \u000brad\nn\nIn this section, we derive model equations for the normalized mode inductance eLn, mode\namplitudeAn, and the radiative damping \u000brad\nn. We are restricting our analysis to the case\nof ideal perpendicular standing spin wave modes that only vary through the \flm thickness\nwithout any lateral variation. It is assumed that the excitations are in the perpendicular\ngeometry with the magnetization saturated out of the \flm plane. As such, the response to\nthe z-coordinate component of the microwave excitation \feld above the waveguide can be\nneglected. In addition, the magnetization precession is always circular. As such, the magne-\ntization dynamics in the x- and y-coordinates in response to the microwave \feld generated\nby the waveguide are degenerate, outside of a phase factor of \u0019=2. This eliminates the need\nto explicitly consider the full Polder susceptibility tensor in the calculation of the sample\nresponse to the excitation \feld. The sample dimensions are lalong the waveguide direction,\n\u000ein thickness, but in\fnite in the lateral direction.\nWe begin by introducing the concepts of a spin wave mode susceptibility \u001fn, and the di-\nmensionless, normalized spin wave amplitude qn(z) for the nth spin wave mode, such that\nthe magnetic excitation of amplitude in x-direction ~mx\nn(H0;I;z) that results from the ap-\nplication of a microwave magnetic \feld of amplitude hx(I;x;z), given in Eq. (2), driven by\nan ac current I=Vin=Z0in an applied \feld H0is given by\n~mx\nn(H0;I;z) := ~mx\nn(z) =qn(z)\u001fn(H0)hqn(z)hx(I;x;z)i (15)\nwhere the quantity in brackets is simply the overlap integral of the excitation \feld and\nthe spatial pro\fle of the nth spin wave mode. The magnetic excitation of amplitude in\ny-direction ~my\nn(z) can be written in a similar way. In the trivial case of a uniform excitation\n\feld and uniform spin wave mode, we recover the usual relation between the excitation\n\feld and the magnetization dynamics via the Polder susceptibility tensor component, \u001fxx.\nHowever, if the product of the mode pro\fle and excitation \feld has odd spatial symmetry,\ndynamics are not excited, as we expect. The overlap integral is nothing more than the\nspatial average of the mode/excitation product:\nhqn(z)hx(I;x;z)i=1\nW\u000eZ1\n\u00001dxZ\u000e+d\nddzqn(z)hx(I;x;z): (16)\n19First, the power transferred to the waveguide via inductive coupling with the spin wave\ndynamics is given by\nPn=j@t\bn(H0;I)j2\n2Z0; (17)\nwhere\n@t\bn(H0;I) =\u00160`1Z\n\u00001dx\u000e+dZ\nddz(@tmx\nn(z))~hx(x;z); (18)\nwith ~hx(x;z) =hx(I;x;z)=I.\nIt is important to recognize at this point that the power dissipation is not constant with\ntime, given that Pnis proportional only to @tmx\nn. As such, the damping associated with the\nre-radiation of the microwave energy back into the waveguide is best characterized with an\nanisotropic damping tensor, to be elaborated upon more fully later in this Appendix. To\ncalculate the energy of the spin wave mode, we start by de\fning a spatially averaged spin\nwave excitation density33,\n\n\u00162\nn(H0;I)\u000b\n=R1\n\u00001dxR\u000e+d\nddz[(@tmx\nn(z)) (my\nn(z))\u0003\u0000(@tmy\nn(z)) (mx\nn(z))\u0003]\n4!\u000eW: (19)\nWe can then calculate the magnon density Nnassociated with the nth spin wave excitation\nas\nNn=h\u00162\nn(H0;I)i\n2g\u0016BMs(20)\nThe total energy associated with the spin wave mode is given by\nEn=!h\u00162\nn(H0;I)i\n\rMs\u000e`W (21)\nThe energy dissipation rate (1 =T1)nfor the nth mode is therefore\n\u00121\nT1\u0013\nn=Pn\nEn=2\u00160`!M\nZ0\f\f\f\f1R\n\u00001dx\u000e+dR\nddz(@tmx\nn(z))~hx(x;z)\f\f\f\f2\n1R\n\u00001dx\u000e+dR\nddz\u0002\n(@tmx\nn(z)) (my\nn(z))\u0003\u0000(@tmy\nn(z)) (mx\nn(z))\u0003\u0003;(22)\n20where!M=\r\u00160Ms. We then apply the Fourier transform to move into the frequency do-\nmain, where @tmx\nn(H0;I;z)$i!~mx\nn(H0;I;z), such that the energy relaxation rate (1 =T1)x\nn\nfor magnetization oscillations along the x-axis is\n\u00121\nT1\u0013x\nn=!\u00160`!M\nZ0Kn; (23)\nwhere\nKn:=\f\f\f\f1R\n\u00001dx\u000e+dR\nddz( ~mx\nn(z))~hx(x;z)\f\f\f\f2\n1R\n\u00001dx\u000e+dR\nddzIm [ ~mx\nn(z) ( ~mx\nn(z))\u0003](24)\nis a dimensionless inductive coupling parameter. In the limiting case of the n= 0 (i.e.,\nuniform) mode with a uniform excitation \feld due to current \rowing only through the\nwaveguide center conductor, and an in\fnitesimal spacing between the waveguide and the\nsample, we have K0=\u000e=4w. Substituting Eq. (15) into Eq. (24), we obtain the general\nresult\nKn=\f\f\f\f1R\n\u00001dx\u000e+dR\nddzqn(z)~hx(x;z)\f\f\f\f2\n\u000f\u000e+dR\nddzjqn(z)j2; (25)\nwith\u000f=j~mz\nnj=j~mx\nnj.\nSince the energy dissipation rate for the case of radiative damping is anisotropic, it must be\ngenerally treated in the damping tensor formalism, where the Gilbert damping torque ~Tis\ngiven by\nTk=\"ijk\u000bij^mi(@t^m)j: (26)\nThe equation of motion is\n@t^m=\u0000\r\u00160^m\u0002~H+~T (27)\nand ^m=~M=Msis the normalized magnetization. For the coordinates in Fig 1, the\nonly nonzero radiative damping tensor components are \u000bzxand\u000byx. For the perpendicular\n21FMR geometry, the relationship between the energy relaxation rate and the Gilbert damping\ncomponents is\n\u00121\nT1\u0013x\n=\u000bzx!x; (28)\nand\n\u00121\nT1\u0013y\n=\u000bzy!y; (29)\nwhere!xand!yare the respective sti\u000bness frequencies, de\fned as\n!i:=\r\nMs@2Um\n@^mi(30)\nandUmis the magnetic free energy function. The frequency-swept linewidth \u0001 !=\n\r\u00160\u0001H, where \u0001His the \feld-swept linewidth in Eq. (11), is given by\n\u0001!=\u0010\n1\nT1\u0011x\n+\u0010\n1\nT1\u0011y\n2(31)\n=\u000bzx!x+\u000bzy!y (32)\nFor perpendicular FMR, !x=!y=!, and the speci\fc case of anisotropic radiative\ndamping,\u000bzx=\u000brad\nn,\u000bzy= 0, and we obtain\n\u000brad\nn=1\n2!\u00121\nT1\u0013x\nn=\u00160l!M\n2Z0Kn (33)\nand \u0001!rad\nn=\u000brad\nn!. This is in contrast to the case of isotropic damping processes, such\nas eddy currents and intrinsic damping, where we obtain \u0001 !iso\nn= 2\u000biso\nn!instead. Thus,\nthe net damping due to the sum of anisotropic radiative damping, and any other isotropic\nprocesses, is given by\n\u000bn=\u000bint+\u000beddy\nn+\u000brad\nn\n2(34)\nwhere\u000bnis the damping parameter in Eq. (11) for the \feld-swept linewidth.\nWe use a vector network analyzer (VNA) to measure the two-port S-parameter matrix\nelement for the nth spin wave mode, \u0001 S21\nn. The matrix element is de\fned as the ratio of the\nvoltage induced in the waveguide by the nth spin wave mode Vn(H0) in an applied magnetic\n\feldH0, and the excitation voltage Vin,\n22\u0001S21\nn:=Vn(H0)\nVin: (35)\nIf we model the reactance of the nth spin wave mode as nothing more than a purely induc-\ntive element of inductance Lnin series with an impedance matched transmission line, and\nif we assume the sample inductance is much smaller than the transmission line impedance,\nwe can approximate \u0001 S21\nnas\n\u0001S21\nn(H0)\u0018=\u0000i!Ln(H0)\nZ0; (36)\nwhereLn(H0) = \bn(H0;I)=I.\nWe de\fne a normalized, \feld-independent mode-inductance eLnas\neLn:=Ln(H0)\n\u001fn(H0); (37)\nand a dimensionless, \feld-independent mode-amplitude An,\nAn:=i!eLn\nZ0: (38)\nsuch that\n\u0001S21\nn(H0) =\u0000An\u001fn(H0): (39)\nThus,Anis the dimensionless amplitude parameter that we obtain when \ftting data for\n\u0001S21\nn(H0). By use of Eqs. 18, 38, and 37, we can rewrite the mode-amplitude as\nAn:=i!\u00160l\nW\u000eZ 0\u0012Z1\n\u00001dxZ\u000e+d\nddzqn(z)~hx(x;z)\u00132\n: (40)\nRemembering that the normalized mode inductance has a factor identical to the numer-\nator of Eq. (25), we can rewrite the radiative damping in terms of the normalized mode\ninductance,\n\u000brad\nn\neLn=!M\u0011n\nZ0; (41)\nwhere\n\u0011n:=\u000e\n4R\u000e+d\nddzjqn(z)j2(42)\n23We emphasize that Eq. (41) is a very general result, regardless of the details of the\nexcitation \feld pro\fle . Thus, even if the \feld pro\fle is highly non-uniform due to the\ncombination of eddy current and capacitive coupling e\u000bects22,34, there should still be a \fxed\nscaling between the radiative damping and the normalized inductance.\nIn the case of the uniform mode, \u0011= 1=4 and Eq. (41) reduces to\n\u000brad\nn\neLn=!M\n4Z0: (43)\nHowever, for a sinusoidal mode of the form\nqn(z) = cos\u0012(2n+ 1)\u0019z\n2\u000e\u0013\n(44)\nthat is expected in the case of a pinned boundary condition at one interface and an open\nboundary condition at the other interface, as shown in Fig. 1 (b), we obtain \u0011= 1=2 and\n\u000brad\nn\neLn=!M\n2Z0; (45)\nFor the case of the wavenumber values extracted from the data shown in Fig. 2 (c) for a\n200 nm Py \flm, we can determine the value for \u0011nand the degree to which it can vary with\nmode number. We use the following form for the spin wave pro\fle:\nqn(z) = cos (knz) (46)\nconsistent with our assumption, when extracting knfrom our PSSW data, that an un-\npinned boundary condition applies to only one of the interfaces, i.e. at z= 0. Using these\nextracted values for the wavenumber, we obtain values for \u0011nshown in Fig. 9.\nWe see in Fig. 9 that the variation in \u0011nwith varying mode number is less than 10%.\nThus, to within \frst order, we can treat \u0011nas a constant for the purposes of \ftting our data,\ni.e.\u0011n\u0018=\u0011.\nB. Derivation of \u000beddy\nFor the derivation of the eddy current damping \u000beddyuniform magnetization dynamics\nare assumed. The notation stays the same as for the radiative damping.\nThen the total \rux passing through the magnetic \flm is\n2402 4 6 8 10120.00.20.40.60.81.0ηnn\nFigure 9: Dependence of calculated values for \u0011non mode number for the case of the spectral data\npresented in Fig. 2 (c) and 5 (a). Wavenumbers are extracted from those data via the procedure\noutlined in the main text of the paper, based upon a model with a single surface with interfacial\nanisotropy, and an unpinned boundary condition at the other interface. Within the context of that\nparticular model, and the expected quadratic dependence of spin wave resonance frequency with\nwavenumber that it produces, we see that \u0011nhas a weak dependence on mode number, justifying\nour presumption that \u0011ncan be treated as a constant for the purposes of \ftting the damping data.\n@t\b =\u00160`\u000e(@t~ m)x; (47)\nwhere (@t~ m)x= ^x\u0001@t~ m. The electrical power dissipated by the eddy-currents is\nPind=1\n2j@t\bj2\n\u0010\n2\u001a`\n\u000ee\u000bW\u0011 (48)\n=C\n8\u00162\n0\u000e3`W\n\u001aj(@t~ m)xj2; (49)\nwith\u000ee\u000b:=C\u000e=2, where 0\u0014C\u00141 is a phenomenological parameter that accounts for\ndetails of the non-uniform eddy-current distribution in the ferromagnet. Analogous to the\nderivation of the radiative damping, we now need the energy of the magnetic excitations.\nThe number of magnons in the system is given by\nNmag=Ms\ng\u0016B!2j@t~ mj2: (50)\n25Thus, the total magnon energy is\nEmag=~!NmagW`\u000e (51)\n=Ms\n\r!j@t~ mj2W`\u000e: (52)\nThe rate of energy dissipation is then given by\n1\nT1=Pind\nEmag=C\n8\r!\u00162\n0Ms\u000e2\n\u001aj(@t~ m)xj2\nj@t~ mj2: (53)\nThe maximum energy decay rate occurs when ( @t^m)x=j@t^mj, in which case\n\u00121\nT1\u0013x\n=C\n8\r!\u00162\n0Ms\u000e2\n\u001a; (54)\nwhere the superscript indicates that this is the maximum decay rate for magnetization\noscillations along the x-axis. For the case of a perpendicular applied \feld su\u000ecient to\nsaturate the static magnetization out of the \flm plane, the damping process is isotropic,\ni.e.,\n\u00121\nT1\u0013y\n=C\n8\r!\u00162\n0Ms\u000e2\n\u001a: (55)\nTherefore, analogous to Eq. (30), the frequency-swept linewidth \u0001 !is simply\n\u0001!=\u0010\n1\nT1\u0011x\n+\u0010\n1\nT1\u0011y\n2(56)\n= 2\u000beddy!; (57)\nthen\n\u000beddy=C\n16\r\u00162\n0Ms\u000e2\n\u001a: (58)\n1J. M. Lock, British Journal of Applied Physics 17, 1645 (1966).\n2Y. Tserkovnyak, A. Brataas, and G. E. W. Bauer, Phys. Rev. Lett. 88, 117601 (2002).\n263G. B. Y. Tserkovnyak, A. Brataas and B. Halperin, Rev. Mod. Phys. 77, 1375 (2005).\n4A. B. Y. Tserkovnyak and G. Bauer, Phys. Rev. Lett. 88, 117601 (2002).\n5C. T. Boone, H. T. Nembach, J. M. Shaw, and T. J. Silva, Journal of Applied Physics 113,\n153906 (2013).\n6J. M. Shaw, H. T. Nembach, and T. J. Silva, Phys. Rev. B 85, 054412 (2012).\n7Y. Tserkovnyak and C. H. Wong, Phys. Rev. B 79, 014402 (2009).\n8V. Baryakhtar, E. Krotenko, and D. Yablonski, Sov. Phys. JETP (1986).\n9Y. Li and W. Bailey, ArXiv : 1401.6467 (2014).\n10H. T. Nembach, J. M. Shaw, C. T. Boone, and T. J. Silva, Phys. Rev. Lett. 110, 117201 (2013).\n11R. W. Sanders, D. Paquette, V. Jaccarino, and S. M. Rezende, Phys. Rev. B 10, 132 (1974).\n12G. Wende, physica status solidi (a) 36, 557 (1976).\n13B. Bhoi, T. Cli\u000b, I. S. Maksymov, M. Kostylev, R. Aiyar, N. Venkataramani, S. Prasad, and\nR. L. Stamps, Journal of Applied Physics 116, (2014).\n14L. Bai, M. Harder, Y. P. Chen, X. Fan, J. Q. Xiao, and C.-M. Hu, Phys. Rev. Lett. 114, 227201\n(2015).\n15C. Scheck, L. Cheng, and W. E. Bailey, Applied Physics Letters 88, 252510 (2006).\n16J. Mallinson, The Foundations of Magnetic Recording (Academic Press, San Diego, 1993), 2nd\ned.\n17Y. Ding, T. J. Klemmer, and T. M. Crawford, Journal of Applied Physics 96, 2969 (2004).\n18S. S. Kalarickal, P. Krivosik, M. Wu, C. E. Patton, M. L. Schneider, P. Kabos, T. J. Silva, and\nJ. P. Nibarger, Journal of Applied Physics 99, 093909 (2006).\n19I. Neudecker, G. Woltersdorf, B. Heinrich, T. Okuno, G. Gubbiotti, and C. Back, Journal of\nMagnetism and Magnetic Materials 307, 148 (2006).\n20H. T. Nembach, T. J. Silva, J. M. Shaw, M. L. Schneider, M. J. Carey, S. Maat, and J. R.\nChildress, Physical Review B (2011).\n21M. Kostylev, Journal of Applied Physics 106, 043903 (2009).\n22M. Bailleul, Applied Physics Letters 103, 192405 (2013).\n23R. F. Soohoo, Phys. Rev. 131, 594 (1963).\n24S. Mathias, C. La-O-Vorakiat, P. Grychtol, P. Granitzka, E. Turgut, J. M. Shaw, R. Adam,\nH. T. Nembach, M. E. Siemens, S. Eich, C. M. Schneider, T. J. Silva, M. Aeschlimann, M. M.\nMurnane, and H. C. Kapteyn, Proceedings of the National Academy of Sciences 109, 4792\n27(2012).\n25T. Maeda, H. Yamauchi, and H. Watanabe, Journal of the Physical Society of Japan 35, 1635\n(1973).\n26G. C. Bailey and C. Vittoria, Phys. Rev. B 8, 3247 (1973).\n27P. Pincus, Phys. Rev. 118, 658 (1960).\n28J. M. Shaw, T. J. Silva, M. L. Schneider, and R. D. McMichael, Phys. Rev. B 79, 184404 (2009).\n29N. Romming, C. Hanneken, M. Menzel, J. E. Bickel, B. Wolter, K. von Bergmann, A. Kubetzka,\nand R. Wiesendanger, Science 341, 636 (2013).\n30H. Ibach, Physics of Surface and Interfaces (Springer, 2006).\n31O. Allivy Kelly, A. Anane, R. Bernard, J. Ben Youssef, C. Hahn, A. H. Molpeceres, C. Carretero,\nE. Jacquet, C. Deranlot, P. Bortolotti, R. Lebourgeois, J.-C. Mage, G. de Loubens, O. Klein,\nV. Cros, and A. Fert, Applied Physics Letters 103, 082408 (2013).\n32M. L. Schneider, T. Gerrits, A. B. Kos, and T. J. Silva, Applied Physics Letters 87, 072509\n(2005).\n33V. Lvov, Wave turbulence under parametric excitation (Springer-Verlag, 1994).\n34I. Maksymov and M. Kostylev, Journal of Applied Physics 113, 043927 (2013).\n28" }, { "title": "2206.06353v1.The_unambiguous_Lorentz_violating_induced_terms.pdf", "content": "arXiv:2206.06353v1 [hep-th] 13 Jun 2022Proceedings of the Ninth Meeting on CPT and Lorentz Symmetry (CPT’22), Indiana University, Bloomington, May 17–26, 2022\n1\nThe unambiguous Lorentz-violating induced terms\nA.R. Vieira,1\n1Universidade Federal do Triˆ angulo Mineiro - Campus Ituram a,\nIturama, MG 38280-000, Brasil\nIn this work, we show that surface terms, which map dependenc e on regular-\nization, can be fixed requiring momentum routing invariance of tadpoles or\ndiagrams with more external legs. This condition makes the L orentz-violating\nterms induced by quantum corrections determined and unique s.\nLorentz and CPT violating operators reappear in quantum correct ions\nand are called induced terms. If the coupling generates a loop with a fi nite\nintegral, the induced term is well defined1. On the other hand, if the loop\ncorrection is divergent, it requires a regularization and a renormaliz ation\nscheme. Thechoiceoftheformermightleadtospurioussymmetryb reaking\ntermsandtoambiguousinducedterms. Sincethefinitepiecesofamp litudes\nof the Standard Model Extension (SME) are in principle observable, they\nshould not depend on this choice. However, symmetries of the built m odel\nare used to decide what kind of induced term is allowed and we can also\nfind out if spurious terms appeared because of the choice of regula rization2.\nThe question if a symmetry of the classical theory is also a symmetry of\nthe quantum theory is non-trivial. The breaking of a classical symme try at\nthe quantum level is called an anomaly and the ideal regularization sho uld\npreserve all symmetries so that it can tell us if the anomaly is physica l\nor spurious. In the SME, this question is even more interesting beca use\nLorentz or CPT symmetries are broken at the classical theory by o ne of the\nSME coefficients but the same coefficient might not appear at the qua ntum\nlevel or it may be a correction to a tree level term from a different se ctor.\nFor instance, the bµcoefficient of the fermion sector appears as a correction\nto the CS-like term ( kAF)µfrom the photon sector.\nThere is no such ideal regularization but the spurious breaking of th e\nsymmetries is usual related to the assumption of a regulator. If it is not\nassumed, the so called implicit regularization3, the breaking of the symme-\ntries is mapped on surface terms. These surface terms manifest t hemselves\nas differences between two infinities. They can be any number and de pendProceedings of the Ninth Meeting on CPT and Lorentz Symmetry (CPT’22), Indiana University, Bloomington, May 17–26, 2022\n2\non the regularization scheme. They are born zero in dimensional reg ular-\nization and if explicit computed, with a hard cutoff for instance, furn ish a\nfinite ora divergentresult. There are severalexamples where the y show up,\nespecially when computing the breaking of a classical symmetry by qu an-\ntum corrections. Let is consider the QED vacuum polarization tenso r as\nan example:\nΠµν(p) =4\n3(p2ηµν−pµpν)Ilog(m2)−4υ2ηµν−\n+4\n3(p2ηµν−pµpν)υ0−4\n3(p2ηµν+2pµpν)(ξ0−2υ0)−\n−8i\n(4π)2(p2ηµν−pµpν)/integraltext1\n0x(1−x)logm2−p2x(1−x)\nm2, (1)\nwhereIlog(m2) =/integraltextd4k\n(2π)41\n(k2−m2)2is a logarithmic basic divergent integral,\nυ2is a quadratic surface term, υ0andξ0arelogarithmic surface terms. The\nbasic divergent integrals are defined as\nIµ1···µ2n\nlog(m2)≡/integraldisplayd4k\n(2π)4kµ1···kµ2n\n(k2−m2)2+n(2)\nand\nIµ1···µ2n\nquad(m2)≡/integraldisplayd4k\n(2π)4kµ1···kµ2n\n(k2−m2)1+n, (3)\nwhile the surface terms are defined according to the relations∗\nυ2wgµν=gµνI2w(m2)−2(2−w)Iµν\n2w(m2),\nξ2wg{µνgαβ}=g{µνgαβ}I2w(m2)−4(3−w)(2−w)Iµναβ\n2w(m2),\nσ2wg{µνgαβgγδ}=g{µνgαβgγδ}I2w(m2)−\n−8(4−w)(3−w)(2−w)Iµναβγδ\n2w(m2), (4)\nwhere 2wis the degree of divergence and we substitute the subscripts log\nandquadby 0 and 2, respectively.\nThe Ward-Takahashi identity ( pµΠµν= 0) we get with equation (1)\nreveals the quadratic surface term υ2is zero and the relation ξ0= 2υ0.\nThus, it is not always possible to fix all the surface terms requiring a s ym-\nmetry. Another example similar to this one is the induced CS-like term\nin the massless case. The result is Πµν\n5(p) = 4iυ0bαpβǫναµβ4, where we\ncan see the logarithmic surface term υ0makes the induced term undeter-\nmined. Also, the Ward-Takahashi identity is compatible with any value of\nthis surface term.\n∗The curly brackets stand forpermutation of indices, A{µνBαβ}=AµνBαβ+AµαBνβ+\nAµβBνα, for instance.Proceedings of the Ninth Meeting on CPT and Lorentz Symmetry (CPT’22), Indiana University, Bloomington, May 17–26, 2022\n3\nAt the same time, there is a freedom of choosing the momentum rout ing\nin the internal lines of loop diagrams. If we set an arbitraryrouting in these\nlines, constrained by momentum conservation at the vertices, this arbitrary\nrouting is always accompanied by a surface term. For example, in the\nequation Πµν(p,l) = Πµν(p,l′), where landl′are arbitrary routing. A\ntextbook example where we can apply that is the axial anomaly. The u sual\nprocedure in its diagrammatic computation is choosing the routing so as to\nfulfill the gauge Ward-Takahashi identities and break the axial cur rent by\nthe known finite amount. However, this does not necessarily mean t hat the\nanomaly depends on the routing. It is possible to show in a diagrammat ic\ncomputationthatthisresultisvalidforanarbitraryrouting5. Furthermore,\nthere is a one-to-one diagrammatic relation between gauge symmet ry and\nmomentum routing invariance in abelian and non-abelian theories. Figu re\n1 shows the pictorial representation of a gauge Ward-Takahashi identity\nwhere we can see an external photon leg being inserted wherever is possible\nina2−pointfunctiondiagram,generatingadifferencebetweentwo1 −point\nfunction diagramswith differentrouting. Asimilarrelationcanbe obta ined\nbeyond one-loop and for diagrams with more external legs.\nFig. 1. The gauge and momentum routing invariance relation f or a 2-point function\ndiagram. The dot refers to the QED extension matrix Γν=γν+cµνγµ+dµνγ5γµ+\neν+ifνγ5+1\n2gλµνσλµ.\nThe gauge and momentum routing invariance relation can be compute d\nwith implicit regularization and reveals a set of equations for the surf ace\nterms. In the QED extension this result yields for the cµνcoefficient:\npµΠµν(p) =τν(p)−τν(0) =−16(3pνcpp+p2cνp+p2cpν)(ξ0−υ0)+\n+4(p2meν+2mpνe·p)(2υ0−ξ0)+4σ0(2pνcpp+p2cνp+p2cpν)+\n+4(ξ2−2υ2)(cνp+cpν) = 0, (5)\nwhereτν(p) is the tadpole for the full fermion propagator and cpν≡pµcµν.Proceedings of the Ninth Meeting on CPT and Lorentz Symmetry (CPT’22), Indiana University, Bloomington, May 17–26, 2022\n4\nWe see in equation (5) that the only possible solution is to make all\nsurface terms zero. Thus, it is possible to fix the arbitrariness req uir-\ning momentum routing invariance and this automatically fulfill the gaug e\nWard-Takahashi identity. If the surface terms were computed w ith an ex-\nplicit regulator like a hard cutoff, we would find spurious breaking of ga uge\ninvariance at one-loop. However, QED extension is gauge invariant b eyond\ntree level6. In this sense, we have determined induced terms. Some ex-\namples of the one-loop vacuum polarization tensor of the QED exten sion\nare\nΠµν\nb(p) =/parenleftbigg\n4ie2υ0+e2m2\nπ2ι0/parenrightbigg\nbαpβǫαβνµ=e2m2\nπ2ι0bαpβǫαβνµ\nΠµν\nd(p) = 2ie2(dλα+dαλ)ǫλµνβpβpα(υ0−ξ0) = 0\nΠµν\ne,a(p) =−4e2(ηµν(me·p−a·p)+pµ(meν−aν)+\n+pν(meµ−aµ))(ξ0−2υ0) = 0, (6)\nwhereι0=/integraltext1\n0dx(1−x)\nm2−p2x(1−x).\nWe see in eq. (6) that we do not expect the induction of dµν,eµor\naµcoefficients and the mass term in the induced CS-like term avoids the\ninfrared divergence. This is in agreement with recent results7where only\ngµναandbµcoefficients appear in the low-energy limit of the one-loop\nvacuum polarization tensor.\nReferences\n1. F. A. Brito, J. R. Nascimento, E. Passos, and A. Yu. Petrov, Phys. Lett. B\n664, 112 (2008).\n2. R. Jackiw, Int. J. of Mod. Phys. B 14, 2011 (2000); B. Altschul, Phys. Rev.\nD.99, 125009 (2019).\n3. O.A. Battistel, A.L. Mota, M.C. Nemes, Mod. Phys. Lett. A 13, 1597 (1998).\n4. J.C.C. Felippe, A.R. Vieira, A.L. Cherchiglia, A.P.B. Sc arpelli and M. Sam-\npaio, Phys. Rev. D 89, 105034 (2014).\n5. A.C.D. Viglioni, A.L. Cherchiglia, A.R. Vieira, B. Hille r and M. Sampaio,\nPhys. Rev. D 94, 065023 (2016).\n6. T.R.S. Santos, R.F. Sobreiro, Phys. Rev. D 94, 125020 (2016); A.R. Vieira,\nA.L. Cherchiglia and M. Sampaio, Phys. Rev. D 93, 025029 (2016).\n7. V.A. Kostelecky, R. Lehnert, N. McGinnis, M. Schreck, B. S eradjeh, Phys.\nRev. Res. 4, 023106 (2022)." }, { "title": "2104.04596v2.New_binary_pulsar_constraints_on_Einstein_æther_theory_after_GW170817.pdf", "content": "New Binary Pulsar Constraints on Einstein-æther\nTheory after GW170817\nToral Gupta1, Mario Herrero-Valea2;3, Diego Blas4, Enrico\nBarausse2;3, Neil Cornish1, Kent Yagi5and Nicolás Yunes6\n1eXtreme Gravity Institute, Department of Physics, Montana State University,\nBozeman, Montana 59717, USA.\n2SISSA, Via Bonomea 265, 34136 Trieste, Italy & INFN, Sezione di Trieste.\n3IFPU - Institute for Fundamental Physics of the Universe,\nVia Beirut 2, 34014 Trieste, Italy.\n4Theoretical Particle Physics and Cosmology Group, Department of Physics, King’s\nCollege London, Strand, London WC2R 2LS, UK.\n5Department of Physics, University of Virginia, Charlottesville, Virginia 22904, USA.\n6Illinois Center for Advanced Studies of the Universe, Department of Physics,\nUniversity of Illinois at Urbana-Champaign, Urbana, Illinois, 61820 USA.\nE-mail: toralgupta@montana.edu\nAbstract. The timing of millisecond pulsars has long been used as an exquisitely\nprecise tool for testing the building blocks of general relativity, including the strong\nequivalence principle and Lorentz symmetry. Observations of binary systems involving\nat least one millisecond pulsar have been used to place bounds on the parameters\nof Einstein-æther theory, a gravitational theory that violates Lorentz symmetry\nat low energies via a preferred and dynamical time threading of the spacetime\nmanifold. However, these studies did not cover the region of parameter space that\nis still viable after the recent bounds on the speed of gravitational waves from\nGW170817/GRB170817A. The restricted coverage was due to limitations in the\nmethods used to compute the pulsar “sensitivities”, which parameterize violations of\nthe strong-equivalence principle in these systems. We extend here the calculation\nof pulsar sensitivities to the parameter space of Einstein-æther theory that remains\nviable after GW170817/GRB170817A. We show that observations of the damping of\nthe period of quasi-circular binary pulsars and of the triple system PSR J0337+1715\nfurther constrain the viable parameter space by about an order of magnitude over\nprevious constraints.\nPACS numbers: 04.30Db,04.50Kd,04.25Nx,97.60Jd\nSubmitted to: Class. Quantum Grav.arXiv:2104.04596v2 [gr-qc] 27 Aug 2021Updated Binary Pulsar Constraints on Einstein-æther theory 2\n1. Introduction\nLorentz symmetry has been the foundation of the magnificent edifice of theoretical\nphysics for more than a century, playing a central role in special and general relativity\n(GR), as well as in the quantum theory of fields. Because of its special status,\nLorentz invariance has been tested to exquisite precision in the matter sector via\nparticle physics experiments [49, 50, 54, 45]. More recently, this experimental program\nhas been extended to the matter-gravity [48], dark matter [16, 13], and pure-gravity\nsectors [42, 53], where bounds on Lorentz violations (LVs) have been historically looser\n(because of the intrinsic weakness of the gravitational interaction).\nCompelling theoretical reasons to seriously consider the possibility of LVs in the purely\ngravitational sector were provided by the realization that they could generate a better\nbehavior in the ultraviolet (UV) limit. In particular, P. Hořava [39] showed that\nby allowing for a non-isotropic scaling between space and time, one can construct\na theory that is power-counting renormalizable in the UV. Renormalizability beyond\npower counting (i.e. pertubative renormalizability) in special (“projectable”) versions of\nHořava gravity has also been proven [10].\nThe low-energy limit of Hořava gravity reduces to “khronometric theory” [15, 43], which\nconsists of GR plus an additional hypersurface-orthogonal and timelike vector field,\noften referred to as the “æther”. Because this vector field is hypersurface orthogonal,\nit selects a preferred spacetime foliation, which makes LVs manifest. A more general\nboost-violating low-energy gravitational theory, however, can be obtained by relaxing\nthe assumption that the æther be hypersurface-orthogonal, in which case it selects a\npreferredtime threading ofthespacetimeratherthanapreferredfoliation. Theresulting\ntheory is known as Einstein-æther theory [46].\nDespite allowing for an improved UV behavior, LVs in gravity face long-standing\nexperimental challenges, particularly when it comes to their percolation into the matter\nsector, where particle physics experiments are in excellent agreement with Lorentz\nsymmetry. While some degree of percolation is inevitable, because of the coupling\nbetweenmatterandgravity, mechanismssuppressingithavebeenputforward, including\nsuppression by a large energy scale [58], or the effective emergence of Lorentz symmetry\nat low energies as a result of renormalization group flows [23, 12, 11] or accidental\nsymmetries [38].\nAt the same time, purely gravitational bounds on LVs are becoming increasingly\ncompelling. The parameters (“coupling constants”) of both Einstein-æther and\nkhronometric theory have been historically constrained by theoretical considerations\n(absence of ghosts and gradient instabilities [18, 41, 37], well-posedness of the\nCauchy problem [60]), by the absence of vacuum Cherenkov cascades in cosmic-\nray experiments [30]), by solar-system tests [69, 34, 18, 19, 55], by observations ofUpdated Binary Pulsar Constraints on Einstein-æther theory 3\nthe primordial abundances of elements from Big-Bang nucleosynthesis [22], by other\ncosmological tests [8], and by precision timing of binary pulsars (where LVs generically\npredict violations of the strong equivalence principle) [33, 32, 73, 74, 9]. More recently,\nthe coincident detection [3, 2] of gravitational waves (GW170817) and gamma rays\n(GRB170817A) emitted by the coalescence of two neutron stars and the subsequent\nkilonova explosion has allowed extremely strong constraints on the propagation speed\nof gravitational waves, which must equal that of light to within z10\u000015[1], which in turn\nplaces even more stringent bounds on the couplings of both theories [31, 59, 60, 56].\nThe bounds from the coincident GW170817/GRB170817A observations force us to\nrethink the parameter spaces of both Einstein-æther and khronometric theory, as the\nonly currently allowed regions appear to be ones that were previously thought to be\nof little interest, and which were not explored extensively. In the case of khronometric\ntheory, Refs. [59, 9] found that the couplings that remain viable after GW170817 and\nGRB170817 produce exactly no deviations away from the predictions of GR, not only\nin the solar system, but also in binary systems of compact objects, be they black holes\n(BHs) or neutron stars (NSs), to leading post-Newtonian (PN) order. Reference [35]\nextended this result to the quasinormal modes of spherically symmetric black holes and\nto fully non-linear (spherical) gravitational collapse, where again no deviations from\nthe GR predictions are found. It would therefore seem that the most promising avenue\nto further test khronometric theory may be provided by cosmological observables (e.g.\nBig-Bang nucleosynthesis abundances or CMB physics), where the viable couplings do\nproduce non-vanishing deviations away from the GR phenomenology.\nLike for khronometric theory, the parameter space where detailed predictions for\nisolated/binary pulsars were obtained in Einstein-æther theory [73, 74] does notinclude\nthe region singled out by the combination of the GW170817/GRB170817A bound and\nexisting solar-system constraints (see Ref. [60] for a discussion). The goal of this paper\nis therefore to extend the previous analysis of binary/isolated-pulsar data by some of\nus [73, 74] to this region of parameter space. This will require a significant modification\nof the formalism that Refs. [73, 74] utilized to calculate pulsar “sensitivities”, i.e. the\nparameters that quantify violations of the strong-equivalence principle in these systems.\nMoreover, we will extend our analysis to include additional data over that considered\nin Refs. [73, 74], namely the triple system PSR J0337+1715 [7]. Overall, we find that\nobservations of the damping of the period of quasi-circular binary pulsars, and that of\nthe triple system PSR J0337+1715, reduce the viable parameter space of Einstein-æther\ntheory by about an order of magnitude over previous constraints.\nWe will also amend an error (originally pointed out in Ref. [70]) in the calculation of\nthe strong-field preferred-frame parameters ^\u000b1and^\u000b2for isolated pulsars, which were\npresented in Refs. [73, 74]. While we have checked that this error does not impact the\nbounds presented in Refs. [73, 74], we present in Appendix A a detailed derivation of\nzSee also [24] for looser bounds coming from mergers of black holes.Updated Binary Pulsar Constraints on Einstein-æther theory 4\n^\u000b1and ^\u000b2for possible future applications, also correcting a few typos present in the\noriginal calculation of Ref. [70].\nThis paper is organized as follows. In Sec. 2 we give a succinct introduction to Einstein–\næther theory, including the modified field equations and the current observational\nbounds on the coupling constants. In Sec. 3 we introduce the concept of stellar\nsensitivities as parameters regulating violations of the strong equivalence principle.\nSolutionsdescribingslowlymovingstarsarederivedinSec.4, andtheyareusedinSec.5\nto compute the sensitivities. Section 6 uses the sensitivities to obtain the constraints on\nEinstein–æther theory resulting from observations of binary and triple pulsar systems.\nWe summarize our conclusions in Sec. 7. Appendix A contains a calculation of the\nstrong-field preferred-frame parameters ^\u000b1and ^\u000b2in Einstein-æther theory, fixing an\noversight in [73], which was pointed out by [70], and correcting also a few typos present\nin [70] itself. We will adopt units where c= 1and a signature +\u0000\u0000\u0000, in accordance\nwith most of the literature on Einstein-æther theory.\n2. Einstein æther theory\nIn order to break boost (and thus Lorentz) symmetry, Einstein-æther theory introduces\na dynamical threading of the spacetime by a unit-norm, time-like vector field U. This\nvector field, often referred to as the æther, physically represents a preferred “time\ndirection” at each spacetime event. Requiring the action to also include the usual\nspin-2 graviton of GR, to be quadratic in the æther derivatives, and to feature no direct\ncoupling between the matter and the æther (so as to enforce the weak equivalence\nprinciple, i.e. the universality of free fall, and the absence of matter LVs at tree level),\none obtains the action [46, 44]\nS=\u00001\n16\u0019GZh\nR+1\n3c\u0012\u00122+c\u001b\u001b\u0016\u0017\u001b\u0016\u0017+c!!\u0016\u0017!\u0016\u0017+caA\u0016A\u0016\n+\u0015(U\u0016U\u0016\u00001)ip\u0000gd4x+Smat( ;g\u0016\u0017); (1)\nwhereRis the four-dimensional Ricci scalar, gthe determinant of the metric, G\nthe bare gravitational constant (related to the value GNmeasured locally by GN=\nG=(1\u0000ca=2)[22, 42]), collectively denotes the matter degrees of freedom, \u0015is a\nLagrange multiplier enforcing the æther’s unit norm, c\u0012,c\u001b,c!andcaare dimensionless\nconstantsx, and we have decomposed the æther congruence into the expansion \u0012, the\nxNote that much of the earlier literature on Einstein-æther theory uses a different set of coupling\nconstantsci(i= 1;:::; 4), which are related to our parameters by c1= (c!+c\u001b)=2,c2= (c\u0012\u0000c\u001b)=3,\nc3= (c\u001b\u0000c!)=2andc4=ca\u0000(c\u001b+c!)=2.Updated Binary Pulsar Constraints on Einstein-æther theory 5\nshear\u001b\u0016\u0017, the vorticity !\u0016\u0017and the acceleration A\u0016as follows:\nA\u0016=U\u0017r\u0017U\u0016; (2)\n\u0012=r\u0016U\u0016; (3)\n\u001b\u0016\u0017=r(\u0017U\u0016)+A(\u0016U\u0017)\u00001\n3\u0012h\u0016\u0017; (4)\n!\u0016\u0017=r[\u0017U\u0016]+A[\u0016U\u0017]; (5)\nwithh\u0016\u0017=g\u0016\u0017\u0000U\u0016U\u0017the projector onto the hyperspace orthogonal to U.\nBy varying the action with respect to the metric, the æther and the Lagrange multiplier,\nand by eliminating the latter from the equations, one obtains the generalized Einstein\nequations\nE\u000b\f\u0011G\u000b\f\u0000TÆ\n\u000b\f\u00008\u0019GTmat\n\u000b\f= 0 (6)\nand the æther equations\nÆ\u0016=\u0014\nr\u000bJ\u000b\u0017\u0000\u0012\nca\u0000c\u001b+c!\n2\u0013\nA\u000br\u0017U\u000b\u0015\nh\u0016\u0017= 0; (7)\nwhereG\u000b\fis the Einstein tensor, the æther stress-energy tensor is\nTÆ\n\u000b\f=r\u0016\u0000\nJ(\u000b\u0016U\f)\u0000J\u0016\n(\u000bU\f)\u0000J(\u000b\f)U\u0016\u0001\n+c!+c\u001b\n2[(r\u0016U\u000b)(r\u0016U\f)\u0000(r\u000bU\u0016)(r\fU\u0016)]\n+U\u0017(r\u0016J\u0016\u0017)U\u000bU\f\u0000\u0012\nca\u0000c\u001b+c!\n2\u0013\u0002\nA2U\u000bU\f\u0000A\u000bA\f\u0003\n+1\n2M\u001b\u001a\n\u0016\u0017r\u001bU\u0016r\u001aU\u0017g\u000b\f;\n(8)\nwith\nJ\u000b\n\u0016\u0011M\u000b\f\n\u0016\u0017r\fU\u0017;\nM\u000b\f\n\u0016\u0017=\u0012c\u001b+c!\n2\u0013\nh\u000b\fg\u0016\u0017+\u0012c\u0012\u0000c\u001b\n3\u0013\n\u000e\u000b\n\u0016\u000e\f\n\u0017+\u0012c\u001b\u0000c!\n2\u0013\n\u000e\u000b\n\u0017\u000e\f\n\u0016+caU\u000bU\fg\u0016\u0017;\nand the matter stress-energy tensor is defined as usual by\nT\u000b\f\nmat\u0011\u00002p\u0000g\u000eSmat\n\u000eg\u000b\f: (9)\nAs already mentioned, a number of experimental and theoretical results constrain\nEinstein-æthertheoryandthecouplings ci. Inmoredetail,perturbingthefieldequations\nabout Minkowski space yields propagation equations for spin-0 (i.e. scalar), spin-1\n(i.e. vector) and spin-2 (i.e. tensor gravitons) modes. Their propagation speeds are\nrespectively given by [41]\nc2\nT=1\n1\u0000c\u001b; (10)\nc2\nV=c\u001b+c!\u0000c\u001bc!\n2ca(1\u0000c\u001b); (11)\nc2\nS=(c\u0012+ 2c\u001b)(1\u0000ca=2)\n3ca(1\u0000c\u001b)(1 +c\u0012=2): (12)Updated Binary Pulsar Constraints on Einstein-æther theory 6\nIn order to ensure stability at the classical level (i.e. no gradient instabilities) and at\nthe quantum level (i.e. no ghosts) one needs to have c2\nT>0,c2\nV>0andc2\nS>0[41, 37].\nIf we also require the modes to carry positive energy, we get ca>0andc!>0[29].\nFurthermore, significantly subluminal graviton propagation would cause ultrarelativistic\nmatter to lose energy to gravitons via a Cherenkov-like process [30]. Since this effect\nis not observed e.g. in ultrahigh energy cosmic rays, one must have c2\nI&1\u0000O(10\u000015)\n(withI=T;V;S). More recently, the coincident detection of a neutron-star merger in\nGW170817 (gravitational waves) and GRB170817A (gamma rays) had led to the bound\n\u00003\u000210\u000015\u0000\u000b1\n2; (46)\nwithj\u000b1j.10\u00004andj\u000b2j.10\u00007, in the limit c\u001b!0. Upon simplification, we get the\nmodified Tolman-Oppenheimer-Volkoff (TOV) equations\ndM\ndr=1\n\u000b1(\u000b1+ 8)rn\n\u00004p\nr\u00002M(\u000b1+ 8)p\n(\u0000\u000b1+ 8)M\u00004P\u000b1\u0019r3+ 4r\n\u0000\u0000\n\u000b2\n1+ 24\u000b1+ 128\u0001\nM\u000016r\u0010\n\u000b1\u0019r2(\u000b1+ 2)P\u00002\u0019r2\u000b1\u001a\u0000\u000b1\n2\u00004\u0011o\n;(47)\nd\u0017\ndr=1\n\u000b1(r\u00002M)rh\n\u00008p\nr\u00002Mp\n(\u0000\u000b1\u00008)M\u00004P\u000b1\u0019r3+ 4r+ 16r\u000032Mi\n;\n(48)\ndp\ndr=1\n\u000b1(r\u00002M)r4(P+\u001a)hp\nr\u00002Mp\n(\u0000\u000b1\u00008)M\u00004P\u000b1\u0019r3+ 4r\u00002r+ 4Mi\n:\n(49)Updated Binary Pulsar Constraints on Einstein-æther theory 14\nModifications to the GR TOV equations can be singled out by expanding the above\nequations(41)–(43)inasmallcouplingapproximation,i.e., ca\u001c1or\u000b1\u001c1[66,57,73].\n4.3. First order in velocity\nWe derive field equations at first order in velocity from Eqs. (6) and (7), which include\nthe potentials as functions of rand\u0012, at first order in velocity. We can separate variables\ninrand\u0012using a Legendre decomposition [73] to obtain\nV(r;\u0012) =X\nnKn(r)Pn(cos\u0012); (50)\nS(r;\u0012) =X\nnSn(r)Pn(cos\u0012)\nd\u0012; (51)\nW(r;\u0012) =X\nnWn(r)Pn(cos\u0012); (52)\nwherePnis the Legendre polynomial of order n. More details on tensor harmonic\ndecomposition can be found in [65]. By separation of variables, we arrive at O(v)\nequations, where only the (t;r)and(t;\u0012)components of the modified Einstein equations\nand therand\u0012components of the æther field equations are non-trivial. We are\nonly interested in n= 1component of Legendre decomposition since these functions\ndetermine sensitivities and consequently the change in orbital period.\nSincec\u001b\u001c10\u000015(c.f. Sec. 2), we proceed with calculations in the limit c\u001b!0to\nobtain [73]\ndS1\ndr=1\n\u000b1r(r\u00002M)n\n\u00002S1p\nr\u00002M(\u000b1+ 4)p\n(\u0000\u000b1\u00008)M\u00004P\u000b1\u0019r3+ 4r\u0000\n(r\u00002M)\u0000\nJ1e\u0017=2c!\u000b1\u0000(3\u000b1+ 16)S1+K1\u000b1(c!\u00001)\u0001\t\n; (53)\ndK1\ndr=1\nc!(r\u00002M)(\u000b2\n1+ (2\u00002\u000b2)\u000b1\u000016\u000b2)\u000b1(\u000b1+ 8)r\b\n2\u0002\n((c!+ 1)\u000b1+ 2c!)J1e\u0017=2\n\u0000(6\u000b1+ 32)S1+c!K1\u000b1] (\u000b2\n1+ (2\u00002\u000b2)\u000b1\u000016\u000b2)(\u000b1+ 8)p\nr\u00002M\n\u0002p\n(\u0000\u000b1\u00008)M\u00004P\u000b1\u0019r3+ 4r+ [\u0000(\u000b1+ 8)(r\u00002M)\u000b1r\n((c!+ 1)\u000b2\n1\u00002c!(\u000b2\u00001)\u000b1\u000016\u000b2c!)\u0012@J1\n@r\u0013\n\u000016\u0012\n\u00001\n8\u0012\n3\u0012\u0012\nc!+5\n3\u0013\n\u000b3\n1\n+\u0012\u0012\n\u00002\u000b2+14\n3\u0013\nc!\u00008\u000b2\n3+8\n3\u0013\n\u000b2\n1+\u0012\u0012\n\u000064\u000b2\n3+16\n3\u0013\nc!\u000064\u000b2\n3\u0013\n\u000b1\n\u0000128\u000b2c!\n3\u0013\n(\u000b1+ 8)M\u0013\n+\u0000\n\u0019(\u000b1+ 2)\u000b1r2((c!+ 1)\u000b2\n1\u00002c!(\u000b2\u00001)\u000b1\n\u000016\u000b2c!)P\u00002\u0019\u000b1r2((c!+ 1)\u000b2\n1\u00002c!(\u000b2\u00001)\u000b1\u000016\u000b2c!)\u001a\n+1\n4\u0012\n(\u000b1+ 8)\u0012\u0012\nc!+3\n2\u0013\n\u000b3\n1+ ((\u00002\u000b2+ 4)c!\u00002\u000b2+ 2)\u000b2\n1Updated Binary Pulsar Constraints on Einstein-æther theory 15\n+((\u000020\u000b2+ 4)c!\u000016\u000b2)\u000b1\u000032\u000b2c!)))r)J1]e\u0017=2\n+ 6\u0012\n\u00002(\u000b1+ 8)2S1\n3+c!K1\u000b1\u0013\n(\u000b1+ 8)(\u000b2\n1+ (2\u00002\u000b2)\u000b1\u000016\u000b2)M\n\u000016\u0014\u0012\n\u0019(\u000b2\n1+ (2\u00002\u000b2)\u000b1\u000016\u000b2)(\u000b1+ 4)\u000b1r2P\u000011\u000b3\n2\n8\n+(3\u000b2\u000011)\u000b2\n1+ (40\u000b2\u000016)\u000b1+ 128\u000b2\u0001\n(\u000b1+ 8)S1+c!K1(\u000b2\n1\n+(2\u00002\u000b2)\u000b1\u000016\u000b2)\u000b1\u0000\n\u0019r2(\u000b1+ 2)P\u00002\u0019r2\u001a+\u000b1=4 + 2\u0001\u0003\nr\t\n; (54)\nd2J1\ndr2=1p\n(\u0000\u000b1\u00008)M\u00004P\u000b1\u0019r3+ 4r(r\u00002M)3=2\u000b3\n1r2(\u000b1+ 8)(4f[(\u000b1+ 8)\n\u0012\n(\u000b2\n1\u00004\u000b1\u000b2\u000032\u000b2)S1+\u0012\n\u0000\u000b2\n1\n2+c!(\u000b2\u00001)\u000b1+ 8\u000b2c!\u0013\nK1\u0013\n\u000b1re\u0000\u0017=2\n\u000012\u000b2\n1\u0012\u00121\n24\u000b2\n1+\u000b1+16\n3\u0013\nM+\u0012\n\u000b1\u0019r2(\u000b1+ 2)P\u00002\u0019r2\u000b1\u001a+\u000b2\n1\n24\u00008\n3\u0013\nr\u0013\n\u0002r\u0012@J1\n@r\u0013\n+\u0012\n8\u0012@\u001a\n@r\u0013\n\u000b3\n1\u0019r4+1\n2\u0000\n(\u000b1+ 8)\u0000\n\u000b3\n1+ (\u00008\u000b2+ 56)\u000b2\n1\n+ (\u0000192\u000b2+ 128)\u000b1\u00001024\u000b2)M) +\u0000\n8\u0019(\u000b3\n1+ (\u00005\u000b2+ 14)\u000b2\n1\n+ (\u000056\u000b2+ 24)\u000b1\u0000128\u000b2)\u000b1r2P+ 4\u0019\u000b2\n1r2\u0000\n\u000b2\n1+ (\u00002\u000b2+ 16)\u000b1\u000016\u000b2\n+ 16)\u001a+\u0012\u000b3\n1\n2+ (\u000b2c!\u0000c!\u000012)\u000b2\n1+ (\u000032 + (8c!+ 32)\u000b2)\u000b1+ 256\u000b2\u0013\n\u0002(\u000b1+ 8)r)J1]p\nr\u00002Mp\n(\u0000\u000b1\u00008)M\u00004P\u000b1\u0019r3+ 4r\n+ 64\u0010\u0010\u000b1\n4+ 2\u0011\nM+P\u000b1\u0019r3\u0000r\u0011\n\u0002\u00141\n8\u0012\n\u0000S1\u000b1\u000b2r(\u000b1+ 8)2e\u0000\u0017=2+\u000b2\n1r(\u000b1+ 8)(r\u00002M)@J1\n@r\u0013\n+J1\u00123(\u000b1+ 8)\n4\n\u0002\u0012\n\u000b2\n1+\u0012\u00008\u000b2\n3+8\n3\u0013\n\u000b1\u000064\u000b2\n3\u0013\nM+\u0000\nr2\u0019\u000b2\n1(\u000b1+ 2)P\n+\u000b2\n1\u0019r2(\u000b1+ 2)\u001a\u00003\u000b3\n1\n8+ (\u000b2\u00004)\u000b2\n1+ (16\u000b2\u00008)\u000b1+ 64\u000b2\u0013\nr\u0013\u0015\u001b\u0013\n;(55)\nwhere we have defined Jn=Wn+e\u0000\u0017=2Kn[73]. With the above set of equations at\nhand, the next section describes the methods of solving these equations at each order\nin velocity.\n5. The calculation of the sensitivities\nThe sensitivities are calculated by solving the coupled differential equations in Eqs. (47)-\n(49) and Eqs. (53)-(55), which are obtained from the modified Einstein and the æther\nfield equations in a v\u001c1expansion atO(v0)andO(v)respectively [73]. In Secs. 5.1\nand 5.2 we describe and apply two methods to solve these equations and find the NS\nsensitivities. The first method, outlined in Sec. 5.1, was used previously in Ref. [73],Updated Binary Pulsar Constraints on Einstein-æther theory 16\nbut we will explain how it leads to unstable solutions in particular regions of parameter\nspace. A second method outlined in Sec. 5.2 provides stable results in all regions of\nparameter space.\nTheO(v0)solutions are common to both methods, as they both involve solving O(v0)\ndifferential equations (47)–(49) numerically once in the interior and then in the exterior\nof the NS. The initial and boundary conditions to it are obtained by imposing regularity\nat the NS center, while imposing asymptotic flatness at spatial infinity respectively. The\ndifferential equations are solved from a core radius (i.e. some small initial radius) to the\nstellar surface radius, where the pressure goes to zero. These numerical solutions at the\nNS surface are now used as initial conditions to solve the exterior evolution equations\nfrom the stellar surface to an extraction radius rb. Using continuity and differentiability\nofthesolutions, theasymptoticsolutionsatspatialinfinityarematchedtothenumerical\nsolutions evaluated at rb. This gives the observed mass of the NS and the integration\nconstantcorrespondingto \u0017(0)(obtainedbysolvingthe O(v0)differentialequations[73])\nwhich will be used in solving the O(v)equations discussed further.\n5.1. Method 1: Direct Numerical Solutions\nIn this method, the aforementioned O(v)differential equations are solved in two regions,\nthe interior of the star, and the exterior. The initial conditions at O(v)are obtained\nby solving the corresponding differential equations asymptotically about a core radius,\nwhile imposing regularity at the core, and asymptotically about spatial infinity, while\nimposing asymptotic flatness [73]. In both cases, the solutions depend on integration\nconstants – ~Cand ~Din the interior asymptotic solution and ~Aand ~Bin the exterior\nasymptotic solution – that must be chosen so as to guarantee that the numerical interior\nand exterior solutions are continuous and differentiable at the stellar surface, where\npressure becomes significantly smaller than their core values.\nAs defined above, the global solution reduces to finding the right constants (~A;~B;~C;~D),\nwhich in turn is a shooting problem. In practice, Ref. [73] solved this shooting prob-\nlem by first picking two sets of values for interior constants, ~ c(1)= (~C(1);~D(1))and\n~ c(2)= (~C(2);~D(2)), andthensolvingtheinteriorequationstwicefromthecoreradius rcto\nthe NS surface R?to find the solutions ~fint\n(1)(r) = [S(1;int)\n1(r);K(1;int)\n1(r);J(1;int)\n1;J0(1;int)\n1 (r)]\nand~fint\n2(r) = [S(2;int)\n1(r);K(2;int)\n1(r);J(2;int)\n1(r);J0(2;int)\n1 ]. Then, each interior nu-\nmerical solution is evaluated at the stellar surface and used as initial con-\nditions for a numerical evolution in the exterior, leading to two exte-\nrior solutions ~fext\n1(r) = [S(1;ext)\n1(r);K(1;ext)\n1 (r);J(1;ext)\n1 (r);J0(1;ext)\n1 (r)]and~fext\n2(r) =\n[S(2;ext)\n1(r);K(2;ext)\n1 (r);J(2;ext)\n1 (r);J0(2;ext)\n1 (r)].\nThe global solutions ~fglo\n1;2(r) =~fint\n1;2(r)[~fext\n1;2(r)are then automatically continuous and\ndifferentiable at the surface, but in general they will not satisfy the boundary conditionsUpdated Binary Pulsar Constraints on Einstein-æther theory 17\nat spatial infinity. Because of the linear and homogeneous structure of the differential\nsystem, one can find the correct global solution through linear superposition\n~fglo(r;C0;D0) =C0~fglo\n1(r) +D0~fglo\n2(r); (56)\nwhereC0andD0are new constants, chosen to guarantee that ~fglosatisfies the correct\nasymptotic conditions near spatial infinity, which in turn depend on (~A;~B), i.e.\n~fglo(rb;C0;D0) =~fglo;1(rb;~A;~B); (57)\nwhererb\u001dR?is the matching radius, ~fglo(rb;C0;D0)is given by Eq. (56) evaluated at\nr=rb(which depends on (C0;D0)) and~fglo;1(rb;~A;~B)is the asymptotic solution to\nthe differential equations near spatial infinity evaluated at the matching radius (which\ndepends on (~A;~B)).\nWith this at hand, one can calculate the NS sensitivities via [73]\n\u001b= 2~A\u000b1\n\u000b1+ 8; (58)\nwhere ~Ais the coefficient of 1=rin the near-spatial infinity asymptotic solution of Wext\n1\nsuch that\nWext\n1(r) =~AM?\nr+O\u0012M2\n?\nr2\u0013\n; (59)\nwhile we recall that \u000b1.10\u00004(c.f. Sec. 2). Because of the latter constraint, it is\nobvious that the sensitivities are essentially controlled by \u001b\u0019~A\u000b1, so the numerical\nstability of its calculation relies entirely on the numerical stability of the calculation of\nthis coefficient. Unfortunately, as we show below, this calculation is not numerically\nstable in the region of parameter space we are interested in.\nFigure 1 shows S1as a function of radius, assuming (\u000b1;\u000b2;c!) = (10\u00004;4\u000210\u00007;\u00000:1),\nand setting (rc;rb) = (102;2\u0002107)cm. Observe that both S(1;glo)\n1andS(2;glo)\n1diverge at\nspatial infinity, so in order to find an Sglo\n1that is finite at spatial infinity, a very delicate\ncancellation of large numbers needs to take place. This cancellation needs to lead to\n~A\u000b1\u00190but in general ~A\u000b16= 0, since\u001b\u0019~A\u000b1=4\u001c16= 0, and precisely by how\nmuch ~A\u000b1deviates from 0is what determines the value of the sensitivity. We find in\npractice that \u001bis highly sensitive to the accuracy of the numerical algorithm used to\nsolve for~fglo\n(1;2), as well as the choice of rc,rband the value of p(R?)that defines the\nstellar surface. Figure 1 is in the parameter region that is outside of interest but it\nindicates how sensitive the calculations are to the aforementioned cancellation, making\nit difficult to find numerically stable solutions.\n5.2. Method 2: Post-Minkowskian Approach\nGiven that the first method does not allow us to robustly compute the sensitivities in\nthe regime ofinterest, we developed a newpost-Minkowskian method, which we describeUpdated Binary Pulsar Constraints on Einstein-æther theory 18\nFigure 1: The metric function | S1| is plotted against the radius in the entire numerical\ndomain, where the radius of the star is at 11.1 km (vertical dashed line). Observe that\nboth of the trial solutions S(1;glo)\n1andS(2;glo)\n1diverge at spatial infinity. Hence, the global\nlinearly combined solution S(glo)\n1shows a diverging behaviour representing numerical\ninstabilities in the calculation of sensitivities.\nhere. In this method, the background O(v0)equations are solved by direct integration,\nas done in method 1. The differential equations at O(v), however, are expanded in\ncompactnessCand solved order by order. This is a post-Minkowskian approximation\nbecause the compactness always appears multiplied by G=c2, so in this sense it is a\nweak-field expansion. NSs are not weak-field objects, but their compactness is always\nsmaller than\u00181=3(and usually between [0:1;0:3]), so provided enough terms are kept\nin the series, this approximation has the potential to be valid. Moreover, and perhaps\nmore importantly, we will show below that such a perturbative scheme stabilizes the\nnumerical solution for the NS sensitivities.\nThe procedure presented above is not technically a standard post-Minkowskian series\nsolution because the background equations (or their solutions) are not expanded in\npowers ofC. Had we expanded in powers of Ceverywhere, we would have encountered\nterms in the differential equations with derivatives of the equation of state (EoS). Such\nderivatives would introduce numerical noise because “realistic” (tabulated) EoSs are not\nusually smooth functions, potentially introducing steep jumps, see, e.g., [75]. By not\nexpanding theO(v0)differential equations, we are implicitly adding higher order terms\nin compactness, so this procedure could be seen as a resummation technique.Updated Binary Pulsar Constraints on Einstein-æther theory 19\nThrough this approach, the differential equations at O(v1)turn intonsets of differential\nequations for an expansion carried out to O(Cn), withntherefore labeling the\ncompactness order. In order to derive these equations, however, one must first establish\nthe order of the background solutions, which can be shown to satisfy\nM(r) =O(C); P (r) =O(C2); (60)\n\u001a(r) =O(C);and\u0017(r) =O(C); (61)\nby looking at the differential equations these functions obey at O(v0). The metric\nperturbation functions at O(v1)are then expanded in powers of compactness through\nYi(r) =nX\nj=1Yij\u000fj; (62)\nwhereYi\u0011(S1;K1;W1),\u000fis a bookkeeping parameter of O(C)andjindicates the order\nofCto be summed over. We work with W1(r)instead ofJ1(r)to avoid introducing\nnumerical error during the conversion between these two functions.\nUsing these expansions in the differential equations at O(v), and re-expanding them in\npowersofcompactness,onefinds nsetsofdifferentialequations. At O(C),thedifferential\nsystem becomes\ndS11\ndr=2rK11\u00002r(S11+\u000b1W11)\u0000(4 +\u000b1)M\n2r2; (63)\ndK11\ndr=\u00001\n2c!r2\u000b1\u0002\n8c!\u0019r3\u000b1\u001a+ 4c!r\u000b1(K11\u0000S11)\u00004r\u000b2\n1W11\n+c!(8\u000b1+\u000b2\n1\u000016\u000b2\u00002\u000b1\u000b2)M\u00004c!r2\u000b1W0\n11\n\u00002r2(\u000b2\n1+c!\u000b2\n1\u000016c!\u000b2\u00002c!\u000b1\u000b2)W0\n11\u0003\n; (64)\nd2W11\ndr2=2W11\nr2; (65)\nwhereS11(r),K11(r)andW11(r)are metric functions at O(C), and recall that the\ndensity\u001a(r)is related to pressure through the EoS as \u001a(r)=\u001a(P(r)). We note that\nW11(r)is decoupled from S11(r)andK11(r), and so we can solve Eq. (65) separately and\nanalytically in the regions r\u0014rbandr\u0015rb. The solutions to them are W11(r) =~D1r2\nforr\u0014rbandW1\n11(r) = ~A1=rforr\u0015rb, where ~A1and ~D1are integration constants.\nBy requiring continuity and differentiability of metric functions W11(r)andW0\n11(r), we\nmatch the solutions at the extraction radius rb. This fixes the values of two integration\nconstants ~A1= 0and ~D1= 0.\nThe remaining two equations, namely Eqs. (63) and (64), are solved numerically with\ninitial conditions obtained using regularity at the NS center and asymptotic flatness at\nspatial infinity. At the center, we have\nS11(r) =~C1\u00001\n120\u000b1\u0019\u0002\n(240\u000b1+ 40\u000b2\n1\u0000128\u000b2\u000016\u000b1\u000b2)\u001acUpdated Binary Pulsar Constraints on Einstein-æther theory 20\n+(48\u000b1\u000b2\u000072\u000b2\n1\u000015\u000b3\n1+ 6\u000b2\n1\u000b2)pc\u0003\nr2\n+1\n1260\u00192\u0002\n(\u0000160\u000b1\u000035\u000b2\n1+ 80\u000b2+ 10\u000b1\u000b2)\u001a2\nc\n+ (288\u000b2\u0000576\u000b1\u0000126\u000b2\n1+ 36\u000b1\u000b2)p2\nc\n+(336\u000b2\u0000672\u000b1\u0000147\u000b2\n1+ 42\u000b1\u000b2)\u001acpc\u0003\nr4+O(r6); (66)\nK11(r) =~C1\u00001\n120\u000b1\u0019\u0002\n(400\u000b1+ 40\u000b2\n1\u0000384\u000b2\u000048\u000b1\u000b2)\u001ac\n+(144\u000b1\u000b2\u000096\u000b2\n1\u000015\u000b3\n1+ 18\u000b2\n1\u000b2)pc\u0003\nr2\n+1\n1260\u00192\u0002\n(400\u000b2\u0000240\u000b1\u000035\u000b2\n1+ 50\u000b1\u000b2)\u001a2\nc\n+ (1440\u000b2\u0000864\u000b1\u0000126\u000b2\n1+ 180\u000b1\u000b2)p2\nc\n+(1680\u000b2\u00001088\u000b1\u0000147\u000b2\n1+ 210\u000b1\u000b2)\u001acpc\u0003\nr4+O(r6);(67)\nwhere ~C1is an integration constant P. At spatial infinity, we have\nS1\n11(r) =\u0000~B1\n2r3\u00001\n2c!r\u000b1h\n~A1(\u000b2\n1\u00002c!\u000b1\u00002\u000b2\n1c!+ 16c!\u000b2+ 2c!\u000b1\u000b2)\n+c!(8\u000b2\u00006\u000b1\u0000\u000b2\n1+\u000b1\u000b2)M?\u0003\n+ (16\u000b2\u00008\u000b1\u0000\u000b2\n1+ 2\u000b1\u000b2)M2\n?\n64r2\n+ (16\u000b2\u000016\u000b1\u00003\u000b2\n1+ 2\u000b1\u000b2)M3\n?\n192r3\n+ (8\u000b2\u00002\u000b1+\u000b1\u000b2) ln\u0012r\nM?\u0013M3\n?\n48r3+O\u0012M4\n?\nr4\u0013\n; (68)\nK1\n11(r) =~B1\nr3+4M?+ 2~A1\u000b1+M?\u000b1\n2r\u0000(\u000b2\n1+ 16\u000b2+ 2\u000b1\u000b2)M2\n?\n64r2\n+ (2\u000b1\u00008\u000b2\u0000\u000b1\u000b2) ln\u0012r\nM?\u0013M3\n?\n24r3+O\u0012M4\n?\nr4\u0013\n(69)\nwhere ~B1is an integration constant and M?is the mass of the star.\nWenextexplainhowtosolveEqs.(63)and(64)toconstructthesolutionfor S11andK11.\nFirst, homogeneous solutions are given by Shom\n11=Khom\n11=~C1. Next, one can construct\nparticular solutions Spart\n11andKpart\n11by setting ~C1= 0and numerically integrate the\nequations from rctoR?. We then use the numerically calculated interior solutions,\nevaluated at R?, as the initial conditions to solve the exterior evolution equations with\nzero pressure and density from R?torb. The true solutions are simply the sum of the\nhomogeneous and particular solutions, namely\nS11(r) =~C1+Spart\n11; (70)\nK11(r) =~C1+Kpart\n11: (71)\nPEquations (66) and (67) (and also Eqs. (68) and (69)) contain terms higher than O(C)because the\nbackground functions are not expanded in a series of Cand thus contain higher order contributions.Updated Binary Pulsar Constraints on Einstein-æther theory 21\nBy requiring continuity and differentiability of all metric functions, we match the true\nnumerical solution to the analytic asymptotic solution in Eqs. (68) and (69) at rb.\nApplying this matching condition gives the values of ~B1and ~C1.\nNow let us focus on the solution to O(C2)differential equations. The equation for the\nmetric function W12(r)is\nd2W12\ndr2=2W12\nr2+ 4\u0019rW 11\u001a0\u00002\u0019\u001a\u0012\n\u00002K11+ 2S11+ (4 +\u000b1+2\u000b1\nc!)W11\u00006rW0\n11\u0013\n\u00002\u0019\u001a(6 +\u000b1)M\nr\u0000W0\n11\u0012\n4 +\u000b2+8\u000b2\n\u000b1\u0013M\nr2\n\u0000\u0012\n3K11\u00003S11\u000010W11\u00002\u000b1W11\u00002\u000b1W11\nc!\u0013M\nr3\n+\u0012\n5 +\u000b1+\u000b2\n2+4\u000b2\n\u000b1\u0013M2\nr4; (72)\nwhere\u001a0(r)is the derivative of \u001a(P(r))obtained from the EoS. This equation is\ndecoupled from the remaining metric functions at O(C2),S12andK12, and can be\nsolved numerically on its own. The initial condition obtained at the center of the NS is\nW12(r) =~D2r2+1\n529200\u000b1\u00192r4\u0002\n560\u001a3\nc\u0019r2\u000b1(\u000b1+ 8)(5\u000b1+ 4\u000b2)\n\u000027p2\nc\u000b2\n1(70(8\u0019pcr2\u00007)\u000b2\n1\u00008\u0019pcr2\u000b1(\u000b2\u0000422) + 49\u000b1(\u000b2\u000062)\n+ 8(49\u00008\u0019pcr2)\u000b2)\u0000252\u001acpc\u000b1\b\n70pc\u0019r2\u000b3\n1+\u000b2\n1(70\u0000pc\u0019r2(\u000b2\u0000382))\n+64(7\u00004\u0019pcr2)\u000b2\u00008\u000b1(5\u0019pcr2(\u000b2+ 8)\u00007(\u000b2+ 10))\t\n\u000012\u001a2\nc\u0000\n350\u0019pcr2\u000b4\n1\u00005\u0019pcr2\u000b3\n1(\u0000226 +\u000b2)\u000031360\u000b2\n\u0000784\u000b1(\u000040 + (5 + 8\u0019pcr2)\u000b2)\u00008\u000b2\n1(\u0000490 +\u0019pcr2(980 + 103\u000b2))\u0001\u0003\n+O(r6);\n(73)\nwhere ~D2is an integration constant. The boundary condition to Eq. (72) at spatial\ninfinity is\nW1\n12(r) =~A2\nr+1\n320r3\u000b1c!h\n~A1M?(40r\u0000M?\u000b1)(\u000b2\n1+c!(4\u000b2\n1\n+\u000b1(34\u00004\u000b2)\u000032\u000b2)) +c!(80M2\n?r(\u000b1+ 8)(\u000b1\u0000\u000b2)\n+M3\n?\u000b1(\u00004\u000b2\n1+ 56\u000b2+\u000b1(\u000038 + 7\u000b2)))\u0003\n+O\u0012M4\n?\nr4\u0013\n; (74)\nwhere ~A2is an integration constant. To construct the solution, we first note that the\nhomogeneous solution is given by Whom\n12=~D2r2. Next, we set ~D2= 0and find the\nparticular solution Wpart\n12(r)numerically in the interior of the NS by solving Eq. (72).\nThis interior solution evaluated at the NS surface now serves as initial conditions to\nsolve the differential equations in the exterior up to the boundary radius rb. The correct\nsolution in the entire numerical domain is then\nW12(r) =~D2r2+Wpart\n12(r); (75)Updated Binary Pulsar Constraints on Einstein-æther theory 22\nwhere the values of ~A2and ~D2are obtained using the matching condition at rb. The\nequations for S12andK12are solved similar to the way Eqs. (63) and (64) are solved,\nso we omit a more detailed description here for brevity. We can use the above method\nto solve differential equations at higher order in C.\n5.3. Comparison between numerical and analytical approaches\n5.3.1. Tabulated APR4 EoS The sensitivity in the æther theory \u001bfor an isolated NS\ndepends on the EoS chosen, and here we perform the calculations of the previous section\nfor the APR4 tabulated EoS [4]. The results are representative of what one finds with\nother EoSs.\nEq. (58) gives the expression of sensitivity in terms of the integration constant ~A[73]\nwhere ~Acan be expressed as\n~A\u00111\nM?nX\nj=2~Aj\u000fj; (76)\nwherenis the order of the compactness expansion, with ~A1= 0. The coefficients ~Aj\ncan be calculated numerically as described in the previous subsection. Notice that the\nleading contribution to ~A(and hence to the sensitivities) is of O(\u000f), sinceM?=O(\u000f).\nThe calculation of the sensitivity as described above requires one to choose the\ntruncation order nof the post-Minkowskian expansion. We will choose nby the\nsensitivities computed from methods 1 and 2 in a regime of parameter space where\nmethod 1 yields stable results [73]. In particular, we will focus on the choice\n(\u000b1;\u000b2;c!;c\u001b) = (10\u00004;4\u000210\u00007;10\u00004;0). Figure 2 compares the sensitivities computed\nwith the two methods with this parameter choice. Observe that as the order of post-\nMinkowskian approximation increases (i.e. as nincreases), the curves approach the\nmethod 1 solution, but in an oscillatory manner. The bottom panel of Fig 2 shows the\nstability of post-Minkowskian method at order n= 3.\nIn the weak field limit, the sensitivity can be well-approximated as the ratio of the\nbinding energy to the NS mass (\n=M\u0003)through [33]\nswf=\u0012\n\u000b1\u00002\n3\u000b2\u0013\nM?; (77)\nwhere the stellar binding energy \nis [32]\n\n =\u00001\n2Z\nd3x\u001a(r)Z\nd3x0\u001a(r0)\njx\u0000x0j; (78)\nwithr=jxjandr0=jx0j. We can use a Legendre expansion of the Green’s function to\nevaluate this integral, and to leading order in C, we find a result that is identical to thatUpdated Binary Pulsar Constraints on Einstein-æther theory 23\nFigure 2: (Top) Sensitivity as a function of compactness using the APR4 EoS, for\nvarious post-Minkowskian truncation orders at (\u000b1;\u000b2;c!;c\u001b) = (10\u00004;4\u000210\u00007;10\u00004;0).\nAt leading order in C, the sensitivity curve overlaps with that computed analytically\nin the weak field limit in [32] (Eq. (77)). As the compactness order is increased, the\nsensitivity curve starts to converge toward the solution found with method 1. (Bottom)\nFractional difference between the sensitivities at different order of compactness and\nthose found from method 1. Observe that when n= 3, the truncated post-Minkowskian\nseries is already an excellent approximation. The vertical dashed line corresponds to\nthe compactness of a 1M\fNS.\ncomputed in the weak field limit by [32]. This can also be seen numerically in Fig. 2,\nwhere the weak field curve coincides with O(C1)post-Minkowskian approximation.\n5.3.2. Tolman VII EoS We now focus on the sensitivity of a NS using the Tolman VII\nEoS. The latter is an analytic model that accurately describes non-rotating NSs [66] by\nthe energy density profile\n\u001a(r) =\u001ac\u0012\n1\u0000r2\nR2\n?\u0013\n: (79)Updated Binary Pulsar Constraints on Einstein-æther theory 24\nThe advantage of using the Tolman VII EoS is that the background solution is known\nanalytically in GR [66, 47]. We expand analytically both the O(v0)andO(v)equations\norder by order in compactness. The sensitivity obtained is then\ns=5\n21C(\u00003\u000b1+ 2\u000b2)\n+ 5\u0012573\u000b3\n1+\u000b2\n1(67669\u0000764\u000b2) + 96416\u000b2\n2+ 68\u000b1\u000b2(\u00002632 + 9\u000b2)\n252252\u000b1\u0013\nC2\n+1\n1801079280 c!\u000b2\n1n\n(4\u000b1)2(8 +\u000b1)(36773030\u000b2\n1\u000039543679\u000b1\u000b2\n+ 11403314 \u000b2\n2) +c!\u0002\n\u00001970100\u000b5\n1+ 13995878400 \u000b3\n2\n+ 640\u000b1\u000b2\n2(\u000049528371 + 345040 \u000b2) + 5\u000b4\n1(\u000019596941 + 788040 \u000b2)\n+\u000b3\n1(\u00002699192440 + 440184934 \u000b2\u00005974000\u000b2\n2)\n16\u000b2\n1\u000b2(1294533212\u000029152855\u000b2+ 212350\u000b2\n2)\u0003o\nC3+O(C4): (80)\nNote that the above expression is not regular in the limit of \u000b1!0while keeping \u000b2\nfinite orc!!0while keeping \u000b1or\u000b2finite. This is a known feature of Einstein-æther\ntheory, which recovers GR only when a certain combination of coupling constants is\ntaken to zero at a specific rate.\nWith this EoS, the compactness can be expressed as a function on \n=M?as\nC=\u00007\n5M?+35819\u000b1\n3\n85800M3\n?+O\u0012\n4\nM4\n?\u0013\n: (81)\nHereCandM?are the observed values with æther corrections included. With this at\nhand, we can rewrite the sensitivity as a function of \nto find\ns=(3\u000b1+ 2\u000b2)\n3\nM?\n+\u0012573\u000b3\n1+\u000b2\n1(67669\u0000764\u000b2) + 96416\u000b2\n2+ 68\u000b1\u000b2(9\u000b2\u00002632)\n25740\u000b1\u0013\n2\nM2\n?\n+1\n656370000c!\u000b2\n1n\n\u00004\u000b2\n1(\u000b1+ 8)\u0002\n36773030\u000b2\n1\u000039543679\u000b1\u000b2\n+11403314\u000b2\n2\u0003\n+c!\u0002\n1970100\u000b5\n1\u000013995878400 \u000b3\n2\n\u0000640\u000b1\u000b2\n2(\u000049528371 + 345040 \u000b2)\u00005\u000b4\n1(19548109 + 788040 \u000b2)\n\u000016\u000b2\n1\u000b2(1294533212\u000029152855\u000b2+ 212350\u000b2\n2)\n+\u000b3\n1(2699192440\u0000309701434\u000b2+ 5974000\u000b2\n2)\u0003o\n3\nM3\n?+O\u0012\n4\nM4\n?\u0013\n:(82)\nNote that this expression matches identically to that of [33] when working to leading\norder in the binding energy.\nOne may wonder whether the above analytic expression is capable of approximating\nthe sensitivity when using other EoSs. Figure 3 shows the absolute magnitude ofUpdated Binary Pulsar Constraints on Einstein-æther theory 25\nFigure 3: Top panel shows the plot of sensitivity as a function of binding energy for\ndifferent EoS including Tolman VII (Eq. (82)) valid to O(C3). The bottom panel shows\nthe relative fractional difference between the EoS from data and the Tolman case, which\nrepresents the EoS variation in the relations. Observe that the universality holds to\nbetter than 3%.\nthe sensitivity as a function of \ncomputed analytically with Eq. (82), as well as\nnumerically with six other EoSs. Here we have chosen to work in a different region\nof parameter space, namely (\u000b1;\u000b2;c!) = (\u000010\u00004;\u00004\u000210\u00007;10\u00003), where we obtain a\nstable smoothly varying sensitivity curve. Observe that the sensitivities differ by less\nthan 3 %, exhibiting an approximate universality already discovered in [73] as a function\nof compactness. Given these results, in all future calculations we will use the analytic\nsensitivities computed with the Tolman VII EoS.\n6. Constraints from binary pulsars and triple systems\nThe majority of millisecond pulsars are found in binary and triple systems. The orbital\ndynamics of these systems modulate the time of arrival of radio waves and allow for\nprecise measurements of the orbital parameters [27, 40, 63, 64]. In this section, weUpdated Binary Pulsar Constraints on Einstein-æther theory 26\ndiscuss the use of precise orbital parameter data to place constraints on the Lorentz-\nviolating Einstein-æther theory. In GR energy is carried away at quadrupolar order\ndue to propagation of tensor modes whereas in this theory (and many of other modified\ntheories of gravity), one usually finds radiation from extra scalar and vector modes\nwhich are responsible for energy loss at dipole order, i.e., \u00001PN order (c.f. the term\nproportional toEin Eq. (30)) as compared to GR. Hence, energy is radiated faster than\nwhat is predicted in GR. This results in a decrease in the orbital separation and orbital\nperiod (Pb) of the binary. The modified orbital period decay rate ( _Pb) relates to the\ntotal energy of the binary, i.e., Eq. (30) via\n_Pb\nPb=\u00003\n2_Eb\nEb; (83)\nsuggesting a strong dependence of _Pbon the sensitivity of the NS [17]. Since the GR\npredictions agree with the observed value of _Pbwithin observational uncertainty, this\nallows for stringent constraints to be placed on æther theory.\n6.1. Observations\nFollowing from Eqs. (30) (with LV terms set to zero) and (83), one can relate the\npost-Keplerian parameter _Pbin GR to the Keplerian parameter Pbvia [73]\n _Pb\nPb!\nGR=\u0000384\u0019\n522=3\u0012\u0019(m1+m2)\nPb\u00135=3m1m2\n(m1+m2)21\nPb: (84)\nHerem1andm2are the masses in the binary system. In principle, if we can measure the\nmasses, orbitalperiodsandtheorbitalperioddecayrateswithsomeuncertaintyandfind\nthat they are consistent with GR predictions, then we can place constraints on Einstein-\næther theory. In this section we focus on data from the measurements of Keplerian\nand post Keplerian parameters of four different pulsar systems PSR J1738+0333, PSR\nJ0348+0432, PSR J1012+5307 and PSR J0737-3039 (Table 1) and a stellar triple\nsystem [67]. The first three are pulsar-white dwarf binaries in orbits with O(10\u00007)\neccentricity and 8.5-hour period, 0.17 eccentricity and 4.74-hour period and O(10\u00007)\neccentricity and 14.5-hour period respectively. The fourth is the double pulsar binary\nsystem with 0.088 eccentricity and 2.45-hour period. Because of the small eccentricity\nof these systems, we will ignore it in the following, i.e. we will consider quasi-circular\nbinaries.\n6.2. Parameter Estimation and Bayesian Analysis\nOur goal is to constrain the theory parameters using measurements of _Pb. We discuss\nbriefly the Bayesian formalism with Markov-Chain Monte-Carlo (MCMC) explorationUpdated Binary Pulsar Constraints on Einstein-æther theory 27\nTable 1: Orbital parameters as measured for the binary systems studied in this paper.\nTable shows the estimated values of the parameters and the 1- \u001buncertainty in the last\ndigits in parentheses. Here, _Pbobsis the observed value of _Pb.\nPulsar System m1(M\f)m2(M\f) Pb(days) _Pbobs\nPSR J1738+0333[36] 1:46+0:06\n\u00000:05 0:181+0:008\n\u00000:007 0:3547907398724(13) \u000025:9(3:2)\u000210\u000015\nPSR J0348+0432 [6] 2.01(4) 0.172(3) 0:102424062722(7) \u00000:273(45)\u000210\u000012\nPSR J1012+5307 [21] [52] 1:64(0:22) 0:16(0:02) 0:60467272355(3) \u00001:5(1:5)\u000210\u000014\nPSR J0737-3039 [51] 1.3381(7) 1.2489(7) 0.10225256248(5) \u00001:252(17)\u000210\u000012\nused to calculate the posteriors on the model parameters [c.f. Sec. 6.2.1] and derive\nrobust constraints.\nFor the parameter estimation, we need the expression for the orbital period decay, which\ndepends on both the relative velocity of the binary constituents v21and the center-of-\nmass velocity wof the binary’s center of mass with respect to the æther field. A natural\nchoice for the æther field direction is provided by the cosmic microwave background (i.e.\nthe æther is expected to be approximately aligned with the cosmological background\ntime direction). In this case, a typical value for the center-of-mass velocity is w\u0018\n10\u00003[33], which for binary pulsar observations is of the same order as v21. If so, the w-\ndependent corrections in the rate of change of the binding energy [Eq. (30)] and orbital\nperiod [Eq. (85)] enter at the same PN order (0PN) as the quadrupole emission terms\nof GR, but multiplied by either (s1\u0000s2)or powers of it. As such, these w-dependent\ncorrections are negligible for both white dwarf-pulsar systems (for which the dominant\nterm is the\u00001PN dipole emission) and also for the relativistic double pulsar system\n(for which s1\u0000s2\u00190as a result of the similar pulsar masses, which kills both the\ndipole emission and the w-dependent corrections to quadrupole emission). Therefore,\nour results are independent of the exact value of was long as that is of order w\u001810\u00003\nor smaller [73]\nFrom the above assumption and using Eqs. (83) and (30), the orbital period decay rate\nin Einstein-æther theory is a function of the individual masses (m1;m2)+, pulsar radii\n(R?;1;R?;2), orbital period ( Pb) and coupling constants (\u000b1;\u000b2;c!)as shown below\n_Pb\nPb=1\n5(m1+m2)4Pbq\n\u000b1\n(\u000b1\u00008\u000b2)q\n\u0000c!\n\u000b1\u000b3\n1c2\n!(\n3 28=3\u0019 \n\u0019(m1+m2)\nPb!5=3\nm1m2\u0014\n\u00001\n12 \n5\u000b1c!(m1+m2)2(s1\u0000s2)2\n+Thesearetheactivemasses, whosefractionaldifferencefromthe“real” masses ( ~m1;~m2)isoftheorder\nof the sensitivities and thus negligible. In the following we will therefore typically identify (m1;m2)and\n( ~m1;~m2). Note that this could however introduce correlations not captured by our sufficient statistics\napproach, but as we show, even large correlations would have little impact on the results.Updated Binary Pulsar Constraints on Einstein-æther theory 28\n \n25=6\u000b2\n1(\u000b1+ 8)r\u000b1\n(\u000b1\u00008\u000b2)\u0000213=3r\u0000c!\n\u000b1c!(\u000b1\u00008\u000b2)!\n \nPb\n\u0019m!2=3!\n+ \n(s1\u00001)(s2\u00001)!2=3 \n(\u000b1+ 8)\u000b3\n1\n \n\u00004c2\n!(m1+m2)2r\u0000c!\n\u000b1+p\n2\u000b1(m1s2+m2s1)2!r\u000b1\n(\u000b1\u00008\u000b2)+\n2q\n\u0000c!\n\u000b1(\u000b1\u00008\u000b2)2c2\n!((m1+m2)\u000b1+ 8m1s2+ 8m2s1)2\n3!#)\n; (85)\nwheres1ands2arefunctionsofCandcouplingconstants. Onemayworrythattheterms\ninside the square roots in the above expression may be negative, leading to a complex\norbital decay rate, but this is not the case because when \u000b1>0, thenc!\u0014\u0000\u000b1=2, while\nwhen\u000b1<0, thenc!\u0015\u0000\u000b1=2. As noted in Fig. 3, sensitivities are independent of the\nEoS. Here we choose to work with the Tolman VII EoS since it gives stable analytic\nsolutions for the sensitivities.\nThere are some phenomenological constraints on the æther coupling constants as\ndiscussed in Sec. 2, i.e., j\u000b1j.10\u00004,j\u000b2j.10\u00007(Solar system constraints), \u000b1<0,\n\u000b1<8\u000b2<0andc!>\u0000\u000b1=2(positive energy, absence of vacuum Cherenkov\nradiation and gradient instabilities) and c\u001b.10\u000015(GW constraint).\u0003Using these\npre-existing constraints and by determining if the estimated value of _Pblies within the\nrange _Pbobs\u0006\u000e_Pbobs(Table 1), we determine the consistency of points in the parameter\nspace with observations.\nOne important point is that we are not using the pulsar timing data directly [5], but\ninstead we are using existing constraints on _Pb; m 1andm2derived from the primary\npulsar timing data as a sufficient statistic . Unfortunately, the published results only\nquote values for the individual parameters and their uncertainties, so we do not have\naccess to the joint posterior distributions. For simplicity we assume that the parameter\ncorrelations are negligible. To check the impact of this assumption, we compared results\nwith zero correlations with a case with 90% correlation between the parameters, and\nfound that it only changed the results by a maximum of 17%.\n6.2.1. Bayesian Analysis We are interested in constructing a posterior distribution on\na set of model parameters ~\u0015= (m1,m2,Pb,R?;1,R?;2,\u000b1,\u000b2,c!) and using an MCMC\nalgorithm to explore the parameter space. According to Bayes’ theorem, the probability\n\u0003Note that we cannot use existing bounds on ^\u000b1[62, 69] and ^\u000b2[61, 62, 69] as priors. This is because\nthose quantities depend on the derivatives of the sensitivities (c.f. Appendix A), which are currently\nunknown.Updated Binary Pulsar Constraints on Einstein-æther theory 29\ndensity for parameters ~\u0015given data D and hypothesis H (the theory) is\nP(~\u0015jD;H) =P(Dj~\u0015;H)P(~\u0015jH)\nP(DjH); (86)\nwhereP(~\u0015jH)is called the prior which represents the state of knowledge about the\nparametersbeforeweanalyzethedata. P(Dj~\u0015;H)iscalledthelikelihoodwhichdescribes\nthe probability of measuring data D given the model H and a set of parameters ~\u0015.\nP(DjH)is called the model evidence which represents the overall normalization factor.\nIn practice it is better to work with log probability densities to better cover the dynamic\nrange of the densities.\nWe assumed uniform priors on \u000b2andc!such that\u00004\u000210\u00007\u0014\u000b2\u00144\u000210\u00007and\n\u0000105\u0014c!\u0014105and a Gaussian prior for \u000b1,m1,m2,_Pbwith mean and standard\ndeviation given by the existing bounds listed in Table 1. We use Gaussian priors on\nR?;1andR?;2with mean and standard deviation given 12:4\u00061:1km based on LIGO\nand NICER measurements [25]. While these bounds are derived assuming GR, the\ncorrections due to LV effects are sub-dominant compared to those impacting _Pb(c.f. e.g.\nfootnoteP). Using lunar laser ranging experiments, the bounds on \u000b1were obtained to\nbe\u000b1= (\u00000:7\u00060:9)\u000210\u00004[55] (c.f. also Sec. 2).\nThe log likelihood function is\nln(P(Dj~\u0015;H))/\u00001\n2\u0012\u0010\n_Pb=Pb\u0011obs\n\u0000\u0010\n_Pb=Pb\u0011th\u00132\n\u001b2\n(_Pb=Pb); (87)\nwhere (_Pb=Pb)this the theoretically predicted value of (_Pb=Pb)from the model. With the\nlikelihood and the priors in place, we can find the posterior using a MCMC algorithm.\nWestarttheMCMCsimulationnearthemeanvaluesforthemodelparameters,calculate\nthe posterior and iterate through these steps. Model parameters are allowed to explore\nthe entire range of parameter space and that gives the joint posterior distribution on all\nparameters ~\u0015. For the proposal distribution we use the prior distribution for a certain\nset of parameters, and a relative jump from the current position for the remaining.\nProposed jumps are accepted or rejected based on the Metropolis-Hastings acceptance\nprobability\nH=min \nP(~\u0015newjH)P(Dj~\u0015new;H)Q(~\u0015oldj~\u0015new)\nP(~\u0015oldjH)P(Dj~\u0015old;H)Q(~\u0015newj~\u0015old);1!\n: (88)\nA random number u\u0018U[0;1]is drawn, and if H > uthe proposed jump is accepted,\notherwise it is rejected. This process is repeated multiple times to ensure convergence.\nWe begin by considering a single observation from the pulsar-white dwarf system PSR\nJ1738+0333. Since the sensitivity of a white dwarf (WD) is negligible compared to theUpdated Binary Pulsar Constraints on Einstein-æther theory 30\nFigure4: Priorandposteriordistributiononthemodelparameters ~\u0015from _Pbconstraints\nfor PSR J1738+0333. The pre-existing constraints from solar system, Big-Bang\nnucleosynthesis and stability requirements are applied to uniform priors shown in blue\nwhere, for \u000b1negative it results in a negative \u000b2and positive c!. The posterior\ndistributions depend on the _Pbconstraints for PSR J1738+0333. The three shades\nof contours in the prior and posterior distribution in the off-diagonal cross-correlation\npanels represent 1- \u001b, 2-\u001band 3-\u001buncertainty on model parameters starting from the\ncenter(weonlyshow1- \u001bshadedregionsfortheone-dimensionalmarginaldistributions).\nThere is a small dip at very small magnitudes of \u000b1, as that is the region where the\npre-existing constraints come into play, while for larger magnitudes the constraints on\n\u000b1are automatically satisfied. Observe that the value of \u000b1is further constrained by a\nfactor of 2 compared to existing solar system constraints (prior).Updated Binary Pulsar Constraints on Einstein-æther theory 31\nFigure 5: Joint prior and posterior distribution on \u000b1,\u000b2andc!from _Pbconstraints for\nall four pulsars listed in (Table 1) The constraint on \u000b1is improved by a factor of 2.\nNS we can set s2=sWD= 0(thusR?;2is excluded from ~\u0015) but for a double pulsar binary\nwe should have s26= 0. Figure 4 shows the prior and posterior distribution on the model\nparameters. We are recovering our priors on the masses and radii, given that these are\nwell constrained as can be noted from Table 1. The distribution on \u000b1and\u000b2is such\nthat\u000b1<0and\u000b2<0from existing constraints. This pulsar system further constrains\nthe value of parameter \u000b1by approximately a factor of 2 while the coupling constants\n\u000b2andc!remain unconstrained. Notice that the posteriors on \u000b2andc!are flat and\nvery similar to the priors. Therefore, one cannot model them as Gaussian and construct\nconfidence region, as no information is gained for the values of these parameters.\nWe then consider constraints on the coupling parameters by stacking all four different\nbinary systems from Table 1 and computing the joint constraints (Fig. 5). These\njoint constraints also restrict the region of \u000b1by a factor of 2 better than the existing\nconstraints.\nIn the coming years we expect to have more observations, as the sensitivities of radio\ntelescopes will improve as a result of larger collecting areas (e.g. the Square Kilometre\nArray (SKA) project [68, 20]), which will allow for discovering more pulsars. Moreover\nthelongerobservationtime( T)willreducetheerrorinmeasurementsof _PbbyT\u00005=2[14],\nallowing for more precise measurements of the orbital parameters. One may wonder,Updated Binary Pulsar Constraints on Einstein-æther theory 32\nFigure 6: Prior and posterior distribution on the model parameters \u000b1,\u000b2andc!from\n_Pbconstraints for PSR J0348+0432, in a scenario where the observational uncertainties\ntighten by a factor of 10 and the value of _Pbmatches the GR prediction. It shows that\nthe GR values are favoured and the value of \u000b1is very closely centered around zero.\nwhether we can get tighter constraints from finding Nsimilar systems or a single system\nmeasuredwithhigherSNR(signal-to-noiseratio). TheSNR2growslinearlywithnumber\nof sourcesNand the observing time T, and quadratically with the effective collecting\narea of the radio telescope. Significant improvements in the measurement sensitivity are\nmore likely to come from some combination of larger telescopes and additional observing\ntime than from discovering large numbers of systems similar to those known. Figure 6\nillustrates the kind of bounds we will get for a PSR J0348+0432 system if _Pbobsmatches\nthe GR prediction _PbGRand the uncertainties are tightened by a factor of 10. The\nimproved constraints on _Pbtranslate directly into similarly improved bounds on \u000b1. We\nalso considered an alternative scenario, in which the uncertainties in _Pbimproved by a\nfactor of 10, but stayed centered on the current observed value. As shown in Figure 7,\nthis leads to a value for \u000b1bounded away from zero. In other words, in a scenario where\nthe observed period derivative stays at the current value while the uncertainty drops by\na factor of ten, we would find that Einstein-æther theory would be favored over GR!Updated Binary Pulsar Constraints on Einstein-æther theory 33\nFigure7: Priorandposteriordistributiononthemodelparameters \u000b1,\u000b2andc!from _Pb\nconstraints for PSR J0348+0432, in a scenario where the uncertainties in measurements\nare reduced by a factor of 10, and the value of _Pbstays at the currently observed value.\nThe 1-\u001buncertainty shows that \u000b1= 0, i.e., the GR value is disfavoured. It can also be\nnoted from Table 2 that posterior does not include \u000b1= 0:\n6.3. Constraints from the triple system\nNextwehaveconstraintscomingfromapulsarinastellartriplesystemPSRJ0337+1715\nconsisting of an inner millisecond pulsar-white dwarf binary and a second white dwarf\n(WD)inanouterorbit[7]. DuetothegravitationalpulloftheouterWD,thepulsarand\nthe inner WD experience accelerations that differ fractionally. If the strong equivalence\nprinciple is violated (as a result of the sensitivities), the triple system constrains the\nfractional acceleration difference parameter \u000eato(+0:5\u00061:8)\u000210\u00006[67]. The relation\nbetween\u000eaand the sensitivity parameter \u001bpulsar(before rescaling) in Einstein-æther\ntheory is [70, 9]\nj\u000eaj=\f\f\f\f\u001bpulsar\n1 +\u001bpulsar=2\f\f\f\f\u0019j\u001bpulsarj; (89)\nas can be obtained directly from Eq. (29) (in the Newtonian limit).\nWe use MCMC simulations in Bayesian analysis similar to that for the _PbconstraintUpdated Binary Pulsar Constraints on Einstein-æther theory 34\nFigure 8: Joint prior and posterior distribution on the model parameters ~\u0015from _Pb\nconstraints for all four pulsars listed in (Table 1) and stellar triple system. Observe\nthat inclusion of stellar triple system improves the constraints and parameter \u000b1is now\nconstrained by a factor of 10 better than lunar laser ranging experiments.\nand with the likelihood\nP(Dj~\u0015;H)/exp \n\u00001\n2\u0000\n\u001bobs\npulsar\u0000\u001bth\npulsar\u00012\n\u001b2\n(\u001bpulsar )!\n; (90)\nwhere\u001bobs\npulsar = (+0:5\u00061:8)\u000210\u00006from Eq. (89) and \u001bth\npulsaris given by Eq. (58),\nto constrain the model parameters. Figure 8 shows the joint pulsar and triple system\nconstraints on the model parameters ~\u0015assuming uniform distribution in \u000b2andc!.\nThe 95% upper limit on \u000b1, which was\u00002:4\u000210\u00004(from the prior constraints) has\nnow shifted to \u000b1=\u00002:4\u000210\u00005. It shows that the preferred frame parameter \u000b1is\nconstrained by a factor of 10 better than the lunar laser ranging experiments.\nTable 2 shows bounds on \u000b1from binary and triple systems mentioned in this paper.\nThe data from joint binary + triple system allows us to put a stringent constraint on\n\u000b1, which is an order of magnitude stronger than the bounds from lunar laser ranging\nexperiments [71, 55].Updated Binary Pulsar Constraints on Einstein-æther theory 35\nTable 2: Our bounds on \u000b1from different pulsar systems shown in Figs. 4–8 with 1- \u001b\nuncertainity. The first half shows the bounds from existing measurements, while the\nsecond half shows projected future bounds assuming that the measurement error on\n_Pobs\nbreduces by a factor of 10 with the central value of _Pobs\nbat the GR predicted value\n(\u000027:3\u000210\u000014) and at the current measured value ( \u000025:3\u000210\u000014).\nPulsar System \u000b1\nPSR J1738+0333 (\u00003:975\u00062:968)\u000210\u00005\nJoint binary system (\u00004:073\u00062.936)\u000210\u00005\nJoint binary + triple system (\u00001:111\u00060.674)\u000210\u00005\nPSR J0348+0432 ( _Pbobs=\u000027:3\u000210\u000014)(\u00008:119\u00064.622)\u000210\u00005\nPSR J0348+0432 ( _Pbobs=\u000025:3\u000210\u000014)(\u00001:729\u00061.805)\u000210\u00005\n7. Conclusions\nWe have investigated Einstein-æther theory in the context of binary pulsars and NSs.\nWe have recalculated the sensitivities in the regime of coupling parameter space that\nstill survives after the recent measurement of the speed of GWs. This required the\ndevelopment of a new post-Minkowskian approach that allows for stable numerical\nevaluation of the sensitivities, in addition to the derivation a closed form analytic\nsolution for the Tolman VII EoS. We used these results to place a constraint on certain\ncoupling constants of Einstein-æther theory using Bayesian analysis of binary pulsar\nobservations, including recent observations on the triple system. We find that these\ndata allows for constraints on a certain combination of the coupling constants, \u000b1, of\nO(10\u00005), improving current Solar System constraints by one order of magnitude.\nThe work carried out here opens the door to several avenues for future research. One\nsuch avenue is to use gravitational wave data directly to place constraints on Einstein-\næther theory, now that the sensitivites have been analytically calculated. This can be\ndone today to leading post-Newtonian order in the inspiral, and it remains to be seen\nwhether it is enough to lead to interesting constraints. To include the very late inspiral\nand merger phase, numerical simulations of coalescing NSs would have to be carried out\nin Einstein-æther theory. However, since the parameter space of the theory is already\nquite well constrained, it is not clear whether stronger bounds can be achieved with\ngravitational wave data.\nAnotheravenueforfutureresearchconcernscomputingsensitivitiesforblackholes. This\nhas been done in khronometric theory [59] but not yet in Einstein-æther theory. Once\nthe black hole sensitivities are in hand, and assuming they do not vanish, one could useUpdated Binary Pulsar Constraints on Einstein-æther theory 36\nthe existing GW data for binary black hole mergers to constrain Einstein-æther theory,\nincluding the dipole radiation effect in the gravitational waveform.\nOne more avenue for future work would be along the lines of improving the analysis\nin this paper by directly analyzing binary pulsar data and carrying out a parameter\nestimation and model selection study with a GR and a non-GR timing model. For this,\nit would be ideal to compute the derivative of the NS sensitivities that enter in the\nconservative post-Keplerian parameters, such as the periastron precession and Shapiro\ntime delay.\nAcknowledgements\nWe would like to thank Clifford Will and Ted Jacobson for many insightful discussions,\nwhich motivated part of this work. TG and NJC acknowlege the support of NASA\nEPSCoR grant MT-80NSSC17M0041 and NSF grant PHY-1912053. KY acknowledges\nsupport from NSF Grant PHY-1806776, NASA Grant No. 80NSSC20K0523, a Sloan\nFoundation Research Fellowship and the Owens Family Foundation. KY would\nlike to also thank the support by the COST Action GWverse CA16104 and JSPS\nKAKENHI Grants No. JP17H06358. NY acknowledges support from NASA grant\nNo. NNX16AB98G, No. 80NSSC17M0041, No. 80NSSC18K1352 and NSF grant PHY-\n1759615. EB and MHV acknowledge financial support provided under the European\nUnion’s H2020 ERC Consolidator Grant “GRavity from Astrophysical to Microscopic\nScales” grant agreement no. GRAMS-815673.\nAppendix A. Modified EIH technique\nIn this Appendix, we start from the 1PN acceleration (29) and analyze the effect of the\n1PN conservative dynamics on the orbital parameters of a binary of compact objects.\nWe will follow the osculating-orbits technique of [70], which will lead us to amend the\ncalculation of the preferred frame parameters ^\u000b1and^\u000b2presented in [73]. In doing so,\nwe will also correct a few typos that we found in the expressions of [70]. ]\nThe relative acceleration between the two gravitating bodies is obtained by simply\nletting\na=dv1\ndt\u0000dv2\ndt; (A.1)\nwhile the position of the center of mass is not accelerated. Thus, we can set _X=X= 0\nat Newtonian order without any loss of generality, getting\nx1=\u0010m2\nm+O(\u000f)\u0011\nx; (A.2)\n]We have double checked the correctness of our expressions with the authors of [70].Updated Binary Pulsar Constraints on Einstein-æther theory 37\nx2=\u0000\u0010m1\nm+O(\u000f)\u0011\nx; (A.3)\nwhere\u000f\u0018m=r\u0018v2\n21is a book-keeping parameter that counts PN order.\nHere the acceleration of every individual body is given by (29). Hereinafter we will\nborrow the notation of [70] and thus we define\nm=m1+m2; \u0011 =m1m2\nm2;\u0001 =m2\u0000m1\nm; (A.4)\nand the functions of the sensitivities\nG=G12;B+=B(12);B\u0000=B[12];D=m2\nmD122+m1\nmD211;\nE=E12;A(n)=\u0010m2\nm\u0011n\nA1\u0000\u0010\n\u0000m1\nm\u0011n\nA2: (A.5)\nUsing this, the relative acceleration can be written in a compact form\na=aL+aPF; (A.6)\nwhere we have separated the purely local contributions and those coming from preferred\nframe effects. The former reads\naL=m\nr2h\nn\u0010\n^A1v2\n21+^A2_r2+^A3m\nr\u0011\n+ _r^Bv21i\n; (A.7)\nwhere\nv21=_x2\u0000_x1; (A.8)\n^A1=1\n2\u0002\nG(1\u00006\u0011)\u00003B+\u00003\u0001B\u0000\u0000\u0011(C12+ 2E) +GA(3)\u0003\n; (A.9)\n^A2=3\u0011\n2(G+E); (A.10)\n^A3=D+G[2\u0011G+ 3B++\u0011(C12+E) + 3\u0001B\u0000]; (A.11)\n^B=G(1\u00002\u0011) + 3B++ 3\u0001B\u0000+\u0011G+GA(3): (A.12)\nThese expressions agree with those of [70]. However, we find a difference in the\nacceleration due to preferred frame effects\naPF=m\nr2\u001a\n\u0000n\u0014\u0012^\u000b1\n2+ 2GA(2)\u0013\n(!\u0001v21) +3\n2\u0000\n^\u000b2+GA(1)\u0001\n(!\u0001n)2\u0015\n(A.13)\n\u0000!\u0014^\u000b1\n2(n\u0001v21) + ^\u000b2(n\u0001!)\u0015\n+GA(2)v21(n\u0001!)\u001b\n\u0000m!2\n2r2\u0000\nC12+GA(1)\u0001\nn:\nwhere we have already specified the generic boost velocity win (29) to match the\nvelocity of the preferred frame !.Updated Binary Pulsar Constraints on Einstein-æther theory 38\nThis differs from the result of [70] in a sign multiplying the first whole line, as well as in\nthe last term, which is absent in [70]. Here we have defined the following compact-body\neffective PPN parameters\n^\u000b1= \u0001(C12+E)\u00006B\u0000\u00002GA(2); ^\u000b2=E\u0000GA(1): (A.14)\nThese are the strong field versions of the parameters \u000b1and\u000b2, which are contained\ninside the definition of the calligraphic objects in (A.14), so that ^\u000b1and^\u000b2are implicit\nfunctions of them. They are directly proportional to each other only in the case in which\nthe sensitivities vanish exactly.\nIn the absence of PN corrections, the motion of the two-body system describes a\nKeplerian orbit, parametrized by x=rnwith\nn= [\u0000cos \n cos(!+f)\u0000cos\u0013sin \n sin(!+f)]eX+ [sin \n cos( !+f)\n+ cos\u0013cos \n sin(!+f)]eY+ sin\u0013sin(!+f)eZ; (A.15)\nwhere the orbital elements are: inclination \u0013, longitude of the ascending node \nand\npericenter angle !. The element f=!\u0000\u001eis the true anomaly, with \u001ethe orbital phase\nmeasured from the ascending node. The reference vectors eiform an orthonormal basis.\nWhen the extra force (A.6) is included, Keplerian orbits are not solutions to the\nequations of motion anymore. However, provided that the force is small enough relative\nto the Newtonian force, we can use perturbation theory and translate the dependence\non time of the motion to the orbital parameters. This is the method of osculating orbits\ndescribed in [71], which leads to a secular variation of the orbital elements under the\neffect of a. In order to parametrize this change in terms of the velocity vector of the\npreferred frame, we decompose the latter by projecting it onto the orbital plane by\ndefining\n!P=!\u0001eP; !Q=!\u0001eQ; !Z=!\u0001z; (A.16)\nas well as onto the angular momentum vector\n!h=!\u0001h=!Zp\nGmp; (A.17)\nFollowing the computation in [71], we thus find that the local terms in aLinduce a\nchange only on the pericenter angle !, which in an orbit changes by\n\u0001L!=6\u0019m\nGp\u0014\nGB++1\n6\u0000\nG2\u0000D\u0001\n+1\n6G\u0000\n6\u0001B\u0000+\u0011(2C12+E) +GA(3)\u0001\u0015\n;(A.18)\nandisofcourseindependentof !. Again,thisagreeswith[70]uptoatypographicalerror\nin their result. The rest of secular changes vanish, either because they are identically\nzero or because they compensate along the orbit.Updated Binary Pulsar Constraints on Einstein-æther theory 39\nOn the other hand, the force induced by preferred frame effects produces a secular\nchange in all orbital parameters\n\u0001PFa=2\u0019e!P\n(1\u0000e2)2\u0012mp\nG\u00131\n2\u0000\n^\u000b1+ 4A(2)G\u0001\n; (A.19)\n\u0001PF\u0013=\u0019^\u000b1\u0012m\nGp\u00131\n2\n!hsin(!)eF(e)\u00002\u0019^\u000b2!h!RF(e)\nGp\n1\u0000e2; (A.20)\n\u0001PF\n =\u0000\u0019^\u000b1\u0012m\nGp\u00131\n2!h\nsin(\u0013)cos(!)eF(e)\u0000)2\u0019^\u000b2!h!SF(e)\nGsin(\u0013)p\n1\u0000e2; (A.21)\n\u0001PF$=\u0000\u0019^\u000b1\u0012m\nGp\u00131\n2\n!Qp\n1\u0000e2F(e)\ne\u0000\u0019^\u000b2(!2\nP!2\nQ)F(e)2\n+\u0019!Q\ne\u0012m\nGp\u00131\n2\n(^\u000b1+ 4A(2)G); (A.22)\n\u0001PFe=\u0000\u0019^\u000b1\u0012m\nGp\u00131\n2\n!P(1\u0000e2)F(e) + 2\u0019^\u000b2!P!Qep\n1\u0000e2F(e)2\n+\u0019!P\u0012m\nGp\u00131\n2\n(^\u000b1+ 4A(2)G); (A.23)\nwhere \u0001PF$= \u0001 PF!+ cos(\u0013)\u0001PF\nand\nF(e) =1\n1 +p\n1\u0000e2; (A.24)\n!R=!Pcos!\u0000!Qp\n1\u0000e2sin!; (A.25)\n!S=!Psin!+!Qp\n1\u0000e2cos!: (A.26)\nOut of these deviations, the most relevant one is the variation of the semimajor axis,\nwhich can be related to the change in the period of the orbit by using Kepler’s third law\n\u0001T\nT=3\n2\u0001a\na: (A.27)\nNote however that this change is sub-leading with respect to the change expected from\nemission of gravitational radiation in a binary system like the one considered throughout\nthis paper [c.f. Eq. (30)], which is actually the dominant factor.Updated Binary Pulsar Constraints on Einstein-æther theory 40\nReferences\n[1] B. P. Abbott et al. Gravitational Waves and Gamma-rays from a Binary Neutron Star Merger:\nGW170817 and GRB 170817A. Astrophys. J. , 848(2):L13, 2017.\n[2] B. P. Abbott et al. Gw170817: Observation of gravitational waves from a binary neutron star\ninspiral. Phys. Rev. Lett. , 119:161101, Oct 2017.\n[3] B. P. Abbott et al. Multi-messenger Observations of a Binary Neutron Star Merger. Astrophys.\nJ., 848(2):L12, 2017.\n[4] A. Akmal, V. R. Pandharipande, and D. G. Ravenhall. Equation of state of nucleon matter and\nneutron star structure. Phys. Rev. C , 58:1804–1828, Sep 1998.\n[5] David Anderson, Paulo Freire, and Nicolás Yunes. Binary pulsar constraints on massless\nscalar–tensor theories using bayesian statistics. Classical and Quantum Gravity , 36(22):225009,\nOct 2019.\n[6] J. Antoniadis, P. C. C. Freire, N. Wex, T. M. Tauris, R. S. Lynch, M. H. van Kerkwijk, M. Kramer,\nC. Bassa, V. S. Dhillon, T. Driebe, and et al. A massive pulsar in a compact relativistic binary.\nScience, 340(6131):1233232–1233232, Apr 2013.\n[7] Anne M. Archibald, Nina V. Gusinskaia, Jason W. T. Hessels, Adam T. Deller, David L. Kaplan,\nDuncan R. Lorimer, Ryan S. Lynch, Scott M. Ransom, and Ingrid H. Stairs. Universality of\nfree fall from the orbital motion of a pulsar in a stellar triple system. Nature, 559(7712):73–76,\n2018.\n[8] B. Audren, D. Blas, M.M. Ivanov, J. Lesgourgues, and S. Sibiryakov. Cosmological constraints on\ndeviations from Lorentz invariance in gravity and dark matter. JCAP, 03:016, 2015.\n[9] Enrico Barausse. Neutron star sensitivities in Horava gravity after GW170817. Phys. Rev. D ,\n100(8):084053, 2019.\n[10] Andrei O. Barvinsky, Diego Blas, Mario Herrero-Valea, Sergey M. Sibiryakov, and Christian F.\nSteinwachs. Renormalization of Horava gravity. Phys. Rev. , D93(6):064022, 2016.\n[11] Andrei O. Barvinsky, Diego Blas, Mario Herrero-Valea, Sergey M. Sibiryakov, and Christian F.\nSteinwachs. Horava Gravity is Asymptotically Free in 2 + 1 Dimensions. Phys. Rev. Lett. ,\n119(21):211301, 2017.\n[12] Grigory Bednik, Oriol Pujolas, and Sergey Sibiryakov. Emergent Lorentz invariance from Strong\nDynamics: Holographic examples. JHEP, 11:064, 2013.\n[13] Dario Bettoni, Adi Nusser, Diego Blas, and Sergey Sibiryakov. Testing Lorentz invariance of dark\nmatter with satellite galaxies. JCAP, 05:024, 2017.\n[14] R. Blandford and S. A. Teukolsky. Arrival-time analysis for a pulsar in a binary system. ApJ,\n205:580–591, April 1976.\n[15] D. Blas, O. Pujolas, and S. Sibiryakov. Consistent Extension of Horava Gravity. Phys. Rev. Lett. ,\n104:181302, 2010.\n[16] Diego Blas, Mikhail M. Ivanov, and Sergey Sibiryakov. Testing Lorentz invariance of dark matter.\nJCAP, 10:057, 2012.\n[17] Diego Blas, Oriol Pujolas, and Sergey Sibiryakov. Models of non-relativistic quantum gravity:\nThe Good, the bad and the healthy. JHEP, 04:018, 2011.\n[18] Diego Blas and Hillary Sanctuary. Gravitational Radiation in Horava Gravity. Phys. Rev. ,\nD84:064004, 2011.\n[19] Matteo Bonetti and Enrico Barausse. Post-Newtonian constraints on Lorentz-violating gravity\ntheories with a MOND phenomenology. Phys. Rev. , D91:084053, 2015. [Erratum: Phys.\nRev.D93,029901(2016)].\n[20] RobertBraun, AnnaBonaldi, TylerBourke, EvanKeane, andJeffWagg. AnticipatedPerformance\nof the Square Kilometre Array – Phase 1 (SKA1). 12 2019.\n[21] Paul J. Callanan, Peter M. Garnavich, and Detlev Koester. The mass of the neutron star in\nthe binary millisecond pulsar PSR J1012 + 5307. Monthly Notices of the Royal Astronomical\nSociety, 298(1):207–211, 07 1998.Updated Binary Pulsar Constraints on Einstein-æther theory 41\n[22] Sean M. Carroll and Eugene A. Lim. Lorentz-violating vector fields slow the universe down. Phys.\nRev., D70:123525, 2004.\n[23] S. Chadha and Holger Bech Nielsen. LORENTZ INVARIANCE AS A LOW-ENERGY\nPHENOMENON. Nucl. Phys. , B217:125–144, 1983.\n[24] Neil Cornish, Diego Blas, and Germano Nardini. Bounding the speed of gravity with gravitational\nwave observations. Phys. Rev. Lett. , 119(16):161102, 2017.\n[25] Michael W Coughlin, Tim Dietrich, Ben Margalit, and Brian D Metzger. Multimessenger bayesian\nparameter inference of a binary neutron star merger. Monthly Notices of the Royal Astronomical\nSociety: Letters , 489(1):L91–L96, Aug 2019.\n[26] T Damour and G Esposito-Farese. Tensor-multi-scalar theories of gravitation. Classical and\nQuantum Gravity , 9(9):2093, 1992.\n[27] ThibaultDamourandJosephH.Taylor. Strongfieldtestsofrelativisticgravityandbinarypulsars.\nPhys. Rev. , D45:1840–1868, 1992.\n[28] D. M. Eardley. Observable effects of a scalar gravitational field in a binary pulsar. Astrophys. J.\nLett., 196:L59–L62, March 1975.\n[29] ChristopherEling. EnergyintheEinstein-aethertheory. Phys. Rev. , D73:084026, 2006. [Erratum:\nPhys. Rev.D80,129905(2009)].\n[30] Joshua W. Elliott, Guy D. Moore, and Horace Stoica. Constraining the new Aether: Gravitational\nCerenkov radiation. JHEP, 08:066, 2005.\n[31] A. Emir Gumrukcuoglu, Mehdi Saravani, and Thomas P. Sotiriou. Hořava gravity after\nGW170817. Phys. Rev. , D97(2):024032, 2018.\n[32] Brendan Z. Foster. Radiation damping in Einstein-aether theory. Phys. Rev. , D73:104012, 2006.\n[Erratum: Phys. Rev.D75,129904(2007)].\n[33] Brendan Z. Foster. Strong field effects on binary systems in Einstein-aether theory. Phys. Rev. ,\nD76:084033, 2007.\n[34] Brendan Z. Foster and Ted Jacobson. Post-Newtonian parameters and constraints on Einstein-\naether theory. Phys. Rev. , D73:064015, 2006.\n[35] Nicola Franchini, Mario Herrero-Valea, and Enrico Barausse. On distinguishing between General\nRelativity and a class of khronometric theories. 3 2021.\n[36] Paulo C. C. Freire, Norbert Wex, Gilles Esposito-Farèse, Joris P. W. Verbiest, Matthew Bailes,\nBryan A. Jacoby, Michael Kramer, Ingrid H. Stairs, John Antoniadis, and Gemma H. Janssen.\nThe relativistic pulsar-white dwarf binary psr j1738+0333 - ii. the most stringent test of scalar-\ntensor gravity. Monthly Notices of the Royal Astronomical Society , 423(4):3328–3343, Jun 2012.\n[37] David Garfinkle and Ted Jacobson. A positive energy theorem for Einstein-aether and Hořava\ngravity. Phys. Rev. Lett. , 107:191102, 2011.\n[38] Stefan Groot Nibbelink and Maxim Pospelov. Lorentz violation in supersymmetric field theories.\nPhys. Rev. Lett. , 94:081601, 2005.\n[39] Petr Hořava. Quantum Gravity at a Lifshitz Point. Phys. Rev. , D79:084008, 2009.\n[40] R.A.HulseandJ.H.Taylor. Discoveryofapulsarinabinarysystem. Astrophys. J. , 195:L51–L53,\n1975.\n[41] T. Jacobson and D. Mattingly. Einstein-Aether waves. Phys. Rev. , D70:024003, 2004.\n[42] Ted Jacobson. Einstein-aether gravity: A Status report. PoS, QG-PH:020, 2007.\n[43] Ted Jacobson. Extended Horava gravity and Einstein-aether theory. Phys. Rev. , D81:101502,\n2010. [Erratum: Phys. Rev.D82,129901(2010)].\n[44] Ted Jacobson. Undoing the twist: The hořava limit of einstein-aether theory. Physical Review D ,\n89(8), Apr 2014.\n[45] Ted Jacobson, Stefano Liberati, and David Mattingly. Lorentz violation at high energy: Concepts,\nphenomena and astrophysical constraints. Annals Phys. , 321:150–196, 2006.\n[46] Ted Jacobson and David Mattingly. Gravity with a dynamical preferred frame. Phys. Rev. ,\nD64:024028, 2001.\n[47] Nan Jiang and Kent Yagi. Improved Analytic Modeling of Neutron Star Interiors. Phys. Rev. ,Updated Binary Pulsar Constraints on Einstein-æther theory 42\nD99(12):124029, 2019.\n[48] Alan V. Kostelecky and Jay D. Tasson. Matter-gravity couplings and Lorentz violation. Phys.\nRev., D83:016013, 2011.\n[49] V. Alan Kostelecky. Gravity, Lorentz violation, and the standard model. Phys. Rev. , D69:105009,\n2004.\n[50] V. Alan Kostelecky and Neil Russell. Data Tables for Lorentz and CPT Violation. Rev. Mod.\nPhys., 83:11–31, 2011.\n[51] M. Kramer et al. Tests of general relativity from timing the double pulsar. Science, 314:97–102,\n2006.\n[52] K. Lazaridis, N. Wex, A. Jessner, M. Kramer, B. W. Stappers, G. H. Janssen, G. Desvignes, M. B.\nPurver, I. Cognard, G. Theureau, A. G. Lyne, C. A. Jordan, and J. A. Zensus. Generic tests of\nthe existence of the gravitational dipole radiation and the variation of the gravitational constant.\nMonthly Notices of the Royal Astronomical Society , 400(2):805–814, 11 2009.\n[53] Stefano Liberati. Tests of Lorentz invariance: a 2013 update. Class.Quant.Grav. , 30:133001, 2013.\n[54] David Mattingly. Modern tests of Lorentz invariance. Living Rev. Rel. , 8:5, 2005.\n[55] Jurgen Muller, James G. Williams, and Slava G. Turyshev. Lunar laser ranging contributions to\nrelativity and geodesy. Astrophys. Space Sci. Libr. , 349:457–472, 2008.\n[56] Jacob Oost, Shinji Mukohyama, and Anzhong Wang. Constraints on Einstein-aether theory after\nGW170817. Phys. Rev. , D97(12):124023, 2018.\n[57] J. R. Oppenheimer and G. M. Volkoff. On massive neutron cores. Phys. Rev. , 55:374–381, Feb\n1939.\n[58] Maxim Pospelov and Yanwen Shang. On Lorentz violation in Horava-Lifshitz type theories. Phys.\nRev., D85:105001, 2012.\n[59] Oscar Ramos and Enrico Barausse. Constraints on Hořava gravity from binary black hole\nobservations. Phys. Rev. , D99(2):024034, 2019.\n[60] Olivier Sarbach, Enrico Barausse, and Jorge A. Preciado-López. Well-posed Cauchy formulation\nfor Einstein-æther theory. Class. Quant. Grav. , 36(16):165007, 2019.\n[61] Lijing Shao, R. Nicolas Caballero, Michael Kramer, Norbert Wex, David J. Champion, and Axel\nJessner. AnewlimitonlocalLorentzinvarianceviolationofgravityfromsolitarypulsars. Class.\nQuant. Grav. , 30:165019, 2013.\n[62] Lijing Shao and Norbert Wex. New tests of local Lorentz invariance of gravity with small-\neccentricity binary pulsars. Class. Quant. Grav. , 29:215018, 2012.\n[63] J. H. Taylor and J. M. Weisberg. A new test of general relativity: Gravitational radiation and\nthe binary pulsar PS R 1913+16. Astrophys. J. , 253:908–920, 1982.\n[64] Joseph H. Taylor and J. M. Weisberg. Further experimental tests of relativistic gravity using the\nbinary pulsar PSR 1913+16. Astrophys. J. , 345:434–450, 1989.\n[65] Kip S. Thorne. Multipole expansions of gravitational radiation. Rev. Mod. Phys. , 52:299–339,\nApr 1980.\n[66] Richard C. Tolman. Static solutions of Einstein’s field equations for spheres of fluid. Phys. Rev. ,\n55:364–373, 1939.\n[67] G. Voisin, I. Cognard, P. C. C. Freire, N. Wex, L. Guillemot, G. Desvignes, M. Kramer, and\nG. Theureau. An improved test of the strong equivalence principle with the pulsar in a triple\nstar system. Astronomy & Astrophysics , 638:A24, Jun 2020.\n[68] A. Weltman et al. Fundamental physics with the Square Kilometre Array. Publ. Astron. Soc.\nAustral., 37:e002, 2020.\n[69] Clifford M. Will. The Confrontation between General Relativity and Experiment. Living Rev.\nRel., 17:4, 2014.\n[70] Clifford M. Will. Testing general relativity with compact-body orbits: a modified\nEinstein–Infeld–Hoffmann framework. Class. Quant. Grav. , 35(8):085001, 2018.\n[71] Clifford M. Will. Theory and Experiment in Gravitational Physics . Cambridge University Press,\n2 edition, 2018.Updated Binary Pulsar Constraints on Einstein-æther theory 43\n[72] Clifford M. Will and Helmut W. Zaglauer. Gravitational Radiation, Close Binary Systems, and\nthe Brans-dicke Theory of Gravity. Astrophys. J. , 346:366, 1989.\n[73] Kent Yagi, Diego Blas, Enrico Barausse, and Nicolás Yunes. Constraints on Einstein-Æther\ntheory and Hořava gravity from binary pulsar observations. Phys. Rev. , D89(8):084067, 2014.\n[Erratum: Phys. Rev.D90,no.6,069902(2014); Erratum: Phys. Rev.D90,no.6,069901(2014)].\n[74] Kent Yagi, Diego Blas, Nicolas Yunes, and Enrico Barausse. Strong Binary Pulsar Constraints on\nLorentz Violation in Gravity. Phys. Rev. Lett. , 112(16):161101, 2014.\n[75] Kent Yagi, Leo C. Stein, Nicolas Yunes, and Takahiro Tanaka. Isolated and Binary Neutron Stars\nin Dynamical Chern-Simons Gravity. Phys. Rev. D , 87:084058, 2013. [Erratum: Phys.Rev.D\n93, 089909 (2016)].\n[76] Xiang Zhao, Chao Zhang, Kai Lin, Tan Liu, Rui Niu, Bin Wang, Shaojun Zhang, Xing Zhang,\nWen Zhao, Tao Zhu, and Anzhong Wang. Gravitational waveforms and radiation powers of the\ntriple system psr J0337 + 1715 in modified theories of gravity. Phys. Rev. D , 100:083012, Oct\n2019." }, { "title": "1909.08360v1.Global_smooth_solutions_of_the_damped_Boussinesq_equations_with_a_class_of_large_initial_data.pdf", "content": "arXiv:1909.08360v1 [math.AP] 18 Sep 2019Global smooth solutions of the damped Boussinesq\nequations with a class of large initial data\nJinlu Li1∗, Xing Wu2†, Weipeng Zhu3‡\n1School of Mathematics and Computer Sciences, Gannan Normal U niversity,\nGanzhou, Jiangxi, 341000, China\n2College of Information and Management Science, Henan Agricul tural University,\nZhengzhou, Henan, 450002, China\n3School of Mathematics and Information Science, Guangzhou Un iversity, Guangzhou 510006, China\nAbstract: The globalregularityproblemconcerning theinviscid Boussinesq equ ationsremains\nan open problem. In an attempt to understand this problem, we exa mine the damped Boussinesq\nequations and study how damping affects the regularity of solutions . In this paper, we consider the\nglobal existence tothedampedBoussinesq equations withaclass of largeinitial data, whose Bs\np,ror\n˙Bs\np,rnorms can be arbitrarily large. The idea is splitting the linear Boussines q equations from the\ndampedBoussinesq equations, theexponentially decaying solutiono ftheformerequationstogether\nwith the structure of the Boussinesq equations help us to obtain th e global smooth solutions.\nKeywords: Boussinesq equations; Global existence; Large initial data.\nMSC (2010): 35Q35; 76B03\n1 Introduction\nThis paper considers the global smooth solutions for the incompres sible Boussinesq equations with\ndamping\n\n\n∂tu+u·∇u+νu+∇p=θed, x∈Rd,t >0,\n∂tθ+u·∇θ+λθ= 0, x ∈Rd,t >0,\ndivu= 0, x ∈Rd,t≥0,\n(u,θ)|t=0= (u0,θ0), x ∈Rd,(1.1)\nwhereuis the velocity vector, pis the pressure, θdenotes the temperature or density which is a\nscalar function. ed= (0,0,...,0,1)Tandν,λare two positive parameters, standing for kinematic\nviscosity and thermal diffusivity, respectively.\nThe Boussinesq equations model large scale atmospheric and ocean ic flows that are responsible\nfor cold fronts and the jet stream [ 12,14] and mathematically has received significant attention,\n∗E-mail: lijinlu@gnnu.cn\n†E-mail: ny2008wx@163.com\n‡E-mail: mathzwp2010@163.com\n1since it has a vortex stretching effect similar to that in the 3D incompr essible flow. When νu\nis replaced by −ν∆u,λθby−λ∆θ, (1.1) becomes the standard viscous Boussinesq equations,\nthe global in time regularity in two dimension is well understood even in t he zero diffusivity\n(ν >0,λ= 0) or the zero viscosity case( ν= 0,λ >0) [1,3,4,5,8,9,10], however the global\nregularity in dimension three appears to be out of reach. while ν= 0 and λ= 0, (1.1) is reduced\nto the inviscid Boussinesq equations, due to the absence of dissipat ive terms, the global solution or\nfinite-time singularity evoluting from general initial data remains uns olved in spite of the progress\non the local well-posedness and regularity criteria [ 6,7,11,15,16,17]. Recently, following the\nconvex integration method, Tao and Zhang [ 16] obtained the H¨ older continuous solution with\ncompact support both in space and time for inviscid 2D Boussinesq eq uations.\nWhen adding velocity damping term νuand temperature damping term λθ, Adhikar et al. [ 2]\nproved (1.1) admits a unique global small solution with the initial data satisfying\n/ba∇dbl∇u0/ba∇dbl˙B0\n∞,11020eVwereobserved[1,2,3]. Theenergyoftheseprotonsarebeyon dtheGZKcutoff\n5×1019eV, which was derived independently by Greisen[4] and Zatsepin , Ku zmin[5]\nshortly after the discovery of the cosmic microwave background r adiation(CMBR).\nAccording to the particle theory, the cosmic-ray nucleons with suffi cient energy will\ninelastic collide with photons of CMBR to produce baryons or pions as f ollows\nP+γ(CMB)→∆, P+γ(CMB)→p+Nπ, (1.1)\nso the nucleons with energies beyond the GZK limit cannot reach us fr om a source\nfurther than a few dozen Mpc. It should be mentioned that, this GZ K limit and the\nphysical process are well understood and measured in the laborat ory[6].\nAnother similar paradox is the TeV γ-rays. Two photons with energy over 2 mec2\ncan produce an electron-positron pair γ+γ→e−+e+. Photons of 10 TeV are most\nsensitive to 30 µminfrared photons, so a photon with enough energy propagating in t he\nintergalacticspacewillinteractwithinfraredbackgroundphotons , andbeexponentially\nsuppressed. However the 10 TeV photons from Mkn501, a BL Lac o bjects at distance\nabout 150Mpc[7, 8], could reach the Earth.\nTo explain the above anomalies, numerous solutions have been propo sed. Most of\nthese solutions are related to slight violation of Lorentz invariance t o get a shift of the\nthresholds of the energy. Glashow believe that the limit speed of a pa rticle depends\non its species and should be the eigenvalue of the velocity eigenstate s[9]-[15]. Then for\nparticlePthe dispersion relation becomes\nE2=p2c2\nP+m2\nPc4\nP. (1.2)\nIn [16], the authors suggested a non-quadratic dispersion relation for photons\np2c2=E2(1+f(δ)), δ=E\nEQG→0 (1.3)\nwhereEQG∼1019GeV is an energy scale caused by quantum gravity effect. Then the\nlight speed is perturbed by E as\n/tildewidec=∂E\n∂p≈c(1+kδ). (1.4)\nThis assumption causes the Lorentz violation and deformed special relativity, which is\nrelatedto the κ-Poincar´ esuperalgebra[17]. The Casimir of the κ-Poincar´ e superalgebra\nhas a structure similar to (1.3). Nowadays, the deformed relativity or noncommutative\ngeometry is greatly developed[18]-[26].\nIn theoretical aspect[27, 28, 29], the Standard Model Extension and quantum grav-\nity suggest that Lorentz invariance may not be an exact symmetry . The possibility\n2Lorentz violation has been investigated in different quantum gravity models, including\nstring theory[30, 31], warped brane worlds[32], and loop quantum gr avity[33]. These\nmodels adopt the Lagrangian like the following[34]\nL=¯ψ/parenleftbigg1\n2eµ\naΓa↔\n∂µ−M/parenrightbigg\nψ, (1.5)\nwhereψis Dirac spinor, eµ\nais the vierbein, and ea\nµis the inverse,\nΓa=γa−cµνeνaeµ\nbγb−fµeµa−ikµeµaγ5+···, (1.6)\nM=m+iµγ5+aµeµ\naγa+bµeµ\naγaγ5+···. (1.7)\nThe first right terms of (1.6) and (1.7) correspond to the normal L orentz invariant\nkinetic term and mass for the Dirac spinor. But the other coefficient s are Lorentz\nviolating coefficients arising from nonzero vacuum expectation value s of the coupling\ntensor fields, which seem to be introduced quite arbitrarily.\nIn contrast with the above theories of Lorentz violation, the tors ion theory is the\nmost natural one in logic[35, 36], which is derived from the fact that t he connection\ncan compatibly introduce an antisymmetric part, namely, the torsio n.\n/tildewideΓµ\nαβ= Γµ\nαβ+Tµ\nαβ, (1.8)\nwhere Γµ\nαβis the Christoffel symbol, and Tµ\nαβ=−Tµ\nβαis the torsion of the spacetime.\nDifferent from all matter fields like electromagnetic field, torsion is a g eometrical inter-\naction similar to gravity, which uniformly interacts with all matter in an accumulating\nmanner. So to test torsion, it seems more effective to measure the movement of heaven\nbody rather than atoms.\nHowever, in contrast with the natural essence and deep philosoph ical meanings\nof the Lorentz invariance, the violation theories seem to be someho w artificial[37].\nSome experiments have been elaborated to test the Lorentz violat ion, but all results\ngave negative answers in high accuracy[38]-[44], so one should take a little conservative\nattitudetowards theLorentz violation. Then how toexplain thethr eshold anomalies of\nUHECR andTeV γ-rays? Herewepresent anotherscenario basedonthenonlinearfi eld\ntheory, which also provides non-quadratic mass-energy relations or dispersion relations\nsimilar to (1.3), but strictly keeps theLorentz invariance[45, 46, 47 ]. In what follows we\nexamine the local Lorentz transformation for some classical para meters defined from\nnonlinear fields and establish their relations.\n32 Local Lorentz transformation and non-quadratic\ndispersion relation\n2.1 Local Lorentz transformation for classical parameters\nTakingtheMinkowski metricas ηµν= diag[1,−1,−1,−1], weconsider thefieldsystems\nof nonlinear spinor ψand scalarφ. For the nonlinear spinor ψ, the dynamic equation\nis given by[48]-[52]\nαµ(¯hi∂µ−eAµ)ψ= (µ−F′)γψ, (2.1)\nwhere the 4 ×4 Hermitian matrices are defined by\nαµ=\n\n\nI0\n0I\n,\n0/vector σ\n/vector σ0\n\n\n, γ=\nI0\n0−I\n (2.2)\nwith Pauli matrices\n/vector σ= (σk) =/braceleftBigg/parenleftBigg0 1\n1 0/parenrightBigg\n,/parenleftBigg0−i\ni0/parenrightBigg\n,/parenleftBigg1 0\n0−1/parenrightBigg/bracerightBigg\n. (2.3)\nF=F(ˇγ) is a positive function of the quadratic scalar ˇ γ≡ψ+γψ.\nFor scalar field φ(xµ), the dynamic equation is given by\n∂µ∂µφ=Kφ, (2.4)\nwhereK=K(|φ|2,∂µφ+∂µφ) is a smooth real function.\nFor both spinor ψand scalarφ, the current conservation law holds due to the gauge\ninvariance of their dynamic equations,\n∂µρµ= 0, (2.5)\nwhere the current is defined respectively by\nρµ=\n\nψ+αµψ for spinor,\niκℑ(φ+∂µφ) for scalar .(2.6)\nκis a normalizing constant. By (2.5) we have the normalizing condition\n/integraldisplay\nR3ρ0d3x= 1. (2.7)\nFor the nonlinear equations (2.1) and (2.4), their solutions have par ticle-wave\nduality[48]-[52]. In [46, 47, 53, 54], the local Lorentz transformatio ns were widely\nused for the classical parameters without proof. Here we set the transformations on\na solid base at first, and then derive the non-quadratic dispersion r elations for some\ncases. The conditions for these results are helpful to understan d the relation between\n4classical mechanics and quantum theory. To clarify the status of t he field system, we\ndefined\nDefinition 1 For the field system ψorφ, we define the central coordinate /vectorX\nand drifting speed/vector vrespectively by\n/vectorX(t) =/integraldisplay\nR3/vector xρ0d3x, /vector v=d\ndt/vectorX, (2.8)\nwheret=x0. The coordinate system with central coordinate /vectorX= 0 is called the\ncentral coordinate system of the field.\nDefinition 2 If a field is a localized wave pack drifting smoothly without emitting\nand absorbing energy quantum, that is, it is at the energy eigensta te in its central\ncoordinate system, we call it at the particle state . Otherwise, the field is in the\nprocess ofexchanging energy quantum with itsenvironment, we ca ll it inthe quantum\nprocess.\nBy the current conservation law (2.5), we have\n/vector v=/integraldisplay\nR3/vector x∂0ρ0d3x=−/integraldisplay\nR3/vector x∇·/vector ρd3x=/integraldisplay\nR3/vector ρd3x. (2.9)\nFor the field at particle state with mean radius much less than the cha racteristic length\nof its environment, by (2.8) and (2.9) we have the classical approxim ation\nρµ→uµ√\n1−v2δ(/vector x−/vectorX), (2.10)\nwhere\nuµ≡(ξ,ξ/vector v), ξ=1√\n1−v2. (2.11)\n(2.10) is the precondition for validity of classical mechanics[46, 47, 5 3]. Forsuch system\nat particle state, we can clearly define the classical parameters su ch as “momentum”,\n“energy” and “mass”, and derive the classical mechanics. From [46 , 47], we learn that\na system at particle state can be described by classical mechanics in high accuracy,\nwhereas for the system in the quantum process, we must describe it by quantum theory\nor by the original equation (2.1) or (2.4). The quantum process is an unstable state,\nwhich is usually completed in a very short time.\nIn what follows, we examine the local Lorentz transformation for t he classical pa-\nrameters. Since the rotation transformation is trivial, we only cons ider the boost one.\nFor the case of flat spacetime, assume xµis the Cartesian coordinate. Consider the\ncentral coordinate system of the field with coordinate ¯Xµ, which moves along x1at\nspeedvwith¯Xk(k/negationslash= 0) parallel to xk, and¯Xk= 0 corresponds to the central coordi-\nnate of the field Xk(t). Then the Lorentz transformation between xµand¯Xµin the\nform of matrix is given by\nx=L(v)¯X,¯X=L(v)−1x=L(−v)x (2.12)\n5wherex= (t,x1,x2,x3)T,¯X= (¯X0,¯X1,¯X2,¯X3)Tand\nL(v) = diag/bracketleftBigg/parenleftBiggξ ξv\nξv ξ/parenrightBigg\n,1,1/bracketrightBigg\n= (Lµ\nν). (2.13)\nAssumeS,PµandTµνareanyscalar, vectorandtensordefinedbytherealfunctions\nofφorψand their derivatives, such as S=|φ|2,Tµν=ℜ(ψ+αµ∂νψ). For the field\nat particle state, all these functions are independent of proper t ime¯X0. Thus in the\ncentral coordinate system, the spatial integrals of these funct ions define the proper\nclassical parameters of the field, and these proper parameters a re all constants. Their\nLorentz transformation laws are given by\nTheorem 1 For a field system at particle state, the integrals of covaria nt functions\nS,PµandTµνsatisfy the following instantaneous Lorentz transformati on laws under\nthe boost transformation (2.12) between xµand¯Xµatdt= 0,\nI≡/integraldisplay\nR3S(x)d3x=√\n1−v2¯I. (2.14)\nIµ≡/integraldisplay\nR3Pµ(x)d3x=√\n1−v2Lµ\nν¯Iν. (2.15)\nIµν≡/integraldisplay\nR3Tµν(x)d3x=√\n1−v2Lµ\nαLν\nβ¯Iαβ. (2.16)\nWhere¯I,¯Iµ,¯Iµνare the proper parameters defined in the central system\n¯I=/integraldisplay\nR3S(¯X)d3¯X,¯Iµ=/integraldisplay\nR3¯Pµd3¯X,¯Iµν=/integraldisplay\nR3¯Tµνd3¯X. (2.17)\nProofWe take (2.15) as example to show therelations. Forthe field at the p article\nstate, by the transformation law of vector, we have\nPµ(x) =Lµ\nν¯Pν(¯X) =Lµ\nν¯Pν(¯X1,¯X2,¯X3) =Pµ(ξ(x1−vt),x2,x3).(2.18)\nSo the integral can be calculated as follows\nIµ=/integraldisplay\nR3Pµ(x)d3x|dt=0\n=/integraldisplay\nR3Pµ(ξ(x1−vt),x2,x3)√\n1−v2d(ξ(x1−vt))dx2dx3(2.19)\n=/integraldisplay\nR3Lµ\nν¯Pν(¯X)√\n1−v2d3¯X=√\n1−v2Lµ\nν¯Iν.\nThe proof is finished.\nRemarks 1 The Lorentz transformation laws (2.14), (2.15) and (2.16) a re valid\nfor the varying speed v(t), because the integrals are only related to the simultaneous\nconditiondx0=dt= 0, and the relations only related to algebraic calculations.\nRemarks 2 When the field is not at the particle state, the covariant inte grands\nwill depend on the proper time ¯X0, then the calculation (2.19) can not pass through,\n6and the relations (2.14)-(2.16) are usually invalid, excep t the integrand satisfies some\nconservation law similar to (2.5).\nIn some text books, the relations (2.14)-(2.16) are directly derive d via Lorentz\ntransformation of the integrands and volume element relation\nd3x=√\n1−v2d3¯X. (2.20)\nAs mentioned by remark 2, the calculation will provides wrong result if the field is not\nat the particle state, because (2.20) is just spatial volume, it suffe rs from the problem\nof simultaneity.\nUsually, the proper parameters have very simple form. For any tru e vector, it\nalways takes ¯Iµ= (¯I0,0,0,0), then\nIµ=√\n1−v2¯I0uµ. (2.21)\nFor some cases of tensor Iµν, it can be expressed as\nIµν=√\n1−v2/parenleftBig¯Kuµuν+¯Jηµν/parenrightBig\n, (2.22)\nwhere¯K,¯Jare constants.\nIn the curved spacetime with orthogonal time coordinate, if the ra dius of curvature\ninthe neighborhoodof the center of the field is much larger thanthe mean radius of the\nfield, the above calculations and relations can be parallel transform ed into the curved\nspacetime. In this case, let ¯Xµbe the central Cartesian coordinate of the tangent\nspacetime at the center of the field. Xµis an inertial coordinate system in the tangent\nspacetime fixed with the curved spacetime. ¯Xµinstantaneously moves along X1at\nspeedv. Then it is easy to check the local Lorentz transformation laws (2.1 4), (2.15)\nand (2.16) also hold, as long as the field is at the particle state.\n2.2 Non-quadratic dispersion relation for spinor\nFor the dark spinor, e= 0 in (2.1). From [46, 47], we get the momentum and mass-\nenergy relation of the spinor as follows\npµ=/parenleftBigg\nm0+Wln1√\n1−v2/parenrightBigg\nuµ, (2.23)\nE=m0√\n1−v2+W/parenleftBigg1√\n1−v2ln1√\n1−v2+√\n1−v2/parenrightBigg\n, (2.24)\nwhereW≪m0is the proper energy provided by the nonlinear potential. The mass-\nenergy relation or dispersion relation in the usual form is given by\n/parenleftBig\nE−W√\n1−v2/parenrightBig2=/vector p2+/parenleftBigg\nm0+Wln1√\n1−v2/parenrightBigg2\n. (2.25)\n7Referring to electron, by estimation we haveW\nm0∼10−6. For a spinor moving at speed\n1−10−n, forn>1 we have\nln1√\n1−v2≈1.15n−0.35. (2.26)\nThe Lagrangian of the particle becomes\nL=−\n(m0+W)+Wln1/radicalBig\ngµν˙xµ˙xν\n/radicalBig\ngµν˙xµ˙xν, (2.27)\nwhere ˙xµ=d\ndtxµ. So the nonlinear term leads to a tiny departure from the geodesic.\nFor a spinor with interaction like electromagnetic field, the mass-ene rgy relation of\nthe particle will be much complicated. More generally, we consider the system with\nfollowing Lagrangian[47]\nL=ψ+αµ(i∂µ−eAµ)ψ−µˇγ+F(ˇγ)−sˇγG\n−1\n2κ(∂µAν∂µAν−a2AµAµ)−1\n2λ(∂µG∂µG−b2G2),(2.28)\nwhereAµandGare potentials produced by ψitself,κ=±1 andλ=±1 stand for\nthe repulsive or attractive self interaction. Then the 4-dimensiona l momentum pµand\nenergyEof the system at particle state are respectively defined by\npµ≡/integraldisplay\nR3ψ+\nk(i∂µ−eAµ)ψd3x, (2.29)\nE≡/integraldisplay\nR3\n/summationdisplay\n∀f∂L\n∂(∂tf)∂tf−L\nd3x=p0+EF+EA+EG,(2.30)\nwhere the classical parameters are given by\npµ=/parenleftBigg\nm0+sWγG+WFln1√\n1−v2/parenrightBigg\nuµ, (2.31)\nEF=WF√\n1−v2, (2.32)\nEA=WA1√\n1−v2/parenleftBigg\n1+v2\n3/parenrightBigg\n+Wav2\n√\n1−v2, (2.33)\nEG=WG1√\n1−v2/parenleftBigg\n1−v2\n3/parenrightBigg\n−Wbv2\n√\n1−v2, (2.34)\nin which the proper parameters are calculated by\nWF=/integraldisplay\nR3(F′ˇγ−F)d3¯x>0, Wγ=/integraldisplay\nR3ˇγd3¯x, (2.35)\nWA=κe2\n8π/integraldisplay\nR6e−ar\nr/parenleftBig\n|ψ(¯x)|2|ψ(¯y)|2+/vector ρ(¯x)·/vector ρ(¯y)/parenrightBig\nd3¯xd3¯y, (2.36)\nWG=−λs2\n8π/integraldisplay\nR6e−br\nrˇγ(¯x)ˇγ(¯y)d3¯xd3¯y, (2.37)\nWa=κ\n3/parenleftbiggae\n4π/parenrightbigg2/integraldisplay\nR3/parenleftBigg/integraldisplay\nR3e−ar\nr|ψ(¯y)|2d3¯y/parenrightBigg2\nd3¯x, (2.38)\nWb=λ\n3/parenleftBiggbs\n4π/parenrightBigg2/integraldisplay\nR3/parenleftBigg/integraldisplay\nR3e−br\nrˇγ(¯y)d3¯y/parenrightBigg2\nd3¯x, (2.39)\n8andr=|¯x−¯y|. By (2.30)-(2.34) we get the dispersion relation as follows\n(E−EF−EA−EG)2=/vector p2+/parenleftBigg\nm0+sWγG+WFln1√\n1−v2/parenrightBigg2\n.(2.40)\nBy (2.25) and (2.40), we find that the interaction term can result in v ery complicated\ndispersion relation. How to use these relation to explain the thresho ld paradoxes of the\nhigh energy comic rays is involved in the interpretation of parameter s, which will be\ndiscussed elsewhere. By the way, the scalar interaction may be abs ent in the nature,\nbecause it manifestly appears in proper mass of feimions, see (2.31) and (2.40).\n3 Discussion and conclusion\nFrom the above analysis, we find that nonlinear field theories do includ e the non-\nquadraticdispersion relationsuch as(2.25)and(2.40). Sotheexpla nationforthreshold\nparadoxes of UHECR and TeV γ-rays by dispersion relation does not definitely require\nLorentz violation. It should be mentioned that, all interactions are actually related to\nnonlinearity, for instance, the charge density of the electromagn etic interaction ψ+αµψ\nisnonlinear, whichcontributesproperenergy WAfortheparticleasdescribedby(2.36).\nThis part of energy satisfy the energy-speed relation (2.33).\nBy (2.31)-(2.34), we learn different interaction term leads to differe nt energy-speed\nrelation. So the experiments towards such relations should bring us important in-\nformation from each interaction, and the high energy cosmic rays m ay be the useful\nmaterials. On the other hand, the disturbance of nonlinear effect m ay influence our\nastronomical observation. For the movement of a heaven body, b y (2.27), we learn\nthat the order of the relative deviation from geodesic is Err=Wv2\nmc2, whereWis the\nenergy contribution of all interactions, mis the usual mass. The typical values for a\nproton are W∼1MeV and mp∼103MeV, so we haveW\nm∼10−3. In a galaxy, the\ntypical speed of heaven bodies relative to the CMB is 300km/s[55], so we have the\ntypical nonlinear deviation for galactic system Err|galaxy∼10−9. In the solar system,\nthe typical speed of the planets is about v∼30km/s, so we have the typical nonlinear\ndeviation for solar system Err|planet∼10−11. Thus, before the nonlinear effects are\nclearly worked out, an astronomical measurement with relative err or less than Erris\ndifficult. The precession of the perihelion of Mercury is 43 seconds of arc per century\n43/(100×360×3600)≈3.3×10−7, so the nonlinear effects have not influence on this\nresult.\nLorentz invariance includes two fold meanings: One is the covariance of the uni-\nversal physical laws, which is a problem of philosophy. The other is pr operty of the\nspacetime, namely the measurement of line element. Whether the sp acetime manifold\n9is measurable is also a philosophical problem, but how to measure the d istances is a\nproblem of geometry. The philosophical problem involves the most fu ndamental and\nuniversal postulates which could only be acceptable as faith, becau se we can neither\ntest all particles whether they satisfy the covariant equation eve rywhere and every\ntime, nor can we check whether all parts of the spacetime are meas urable, including\nthe singularity inside a black hole and each points at a line of Planck lengt h. Even\nthough the Lorentz violation theories can not manifestly violate the covariance, see\nthe forms of (1.5)-(1.7) as example. One may argue the thermodyn amics is not covari-\nant, the answer is that it is just a conditional theory but not a unive rsal one, thus to\nabuse its concepts and laws, such as entropy and the second law, w ithout restriction is\ninadequate and leads to confusion.\nWe once measured the spacetime with rule dst=|dt|,ds2\ns=dx2+dy2+dz2\nand solved infinite practical problems. Today we measure the univer se withds2=\ngµνdxµdxνthen we achieved beauty and harmony. There are infinite rules to me asure\nlength consistently, but only the spacetime with rule of quadratic fo rm has wonderful\nproperties and potential. If accepting the quadratic rule, the loca l Lorentz invariance\ncertainly holds for the spacetime.\nThere are also fields satisfy the covariance but violate the Lorentz invariance. Their\nLagrangian always includes term as follows[27]\nLLV= (gµν+τµν)∂µ∂νφ+ψ+(αµ+aµ)∂µψ+···.\nFor such fields, the propagating speed is not pure geometrical, whic h depends on fields\nτµνandaµ. Similar to the case of Navier-Stockes equation in fluid mechanics, th e\nnonlinearity is much worse than that of (2.28). Although for adequa tely small value\nτµνandaµ, the solutions to the dynamical equation will also be finite, but the wo rld\nincluding such fields will become a mess, because each atom of the sam e element have\ndifferent spectrum depending on coordinates, crystal lattices ar e distorted, the solar\nsystem has turbulence and chaos, and the double-helix of DNA has d isordered knots.\nSo‘the coefficients of the highest order derivatives in the Lagr angian must\nbe constants’ should be a fundamental postulate. This postulate is related to the\nquaternion structure of the world, only of such excellent structu re the world becomes\nso luxuriant but so harmonious. In percipience of such opinion, the m odification of\ngeneral relativity with terms φR,f(R) is also doubtful.\nSomeconfusionsinphysicsarecausedbyambiguousconceptsorcir cularrelations[56,\n57]. For example, to the relation between classical mechanics(CM) a nd quantum me-\nchanics(QM), the common answer must be that: “transform the c lassical parameters,\nHamiltonian and the energy equation of CM into operators, we get QM . Contrarily,\nthe limit of QM as ¯ h→0 provides CM.”\n10At first, by the answer, it seems that both CM and QM are equivalent theories\nin logic. In fact they are different theory suitable for different stat us of a system.\nThe basic concepts such as ‘coordinate’, ‘momentum’ have different meanings in each\nmechanics, although they have close relations. The uncertainty re lation is the typical\nplausible laws making puzzles and paradoxes[56, 57]. Clearly they sho uld be unified\nin a higher level theory[45, 46, 50, 52]. secondly, ¯ his a universal constant acting as a\nunit to measure other physical parameters, we have not a concep t ¯h→0? Whether\na physical system should be described by CM or QM is obviously not det ermined by\nexternal conditions such as ¯ h→0 and CM or QM, but determined by the status of\nthe system itself. This is the meaning of the Definition 2 .\nSpacetimeandfieldsaredifferentcomponentsoftheworld. Theypa lydifferent roles\nwith different characteristics, and satisfy completely different pos tulates and measure-\nment rules. So it is hard to understand the motivation of the quantu m gravity. Why\nwe should modify a well defined and graceful theory, without any de finite violation of\nexperiments, by an ambiguous and incomplete theory? Why not modif y the quantum\nfield theory by general relativity? The mission of a physical theory is to find out the\nintrinsic truth and beauty and harmony of the nature. But at its be st, besides some\nill defined concepts as foam, wormhole, Lorentz violation, what can quantum gravity\nactually provide us?\nInthespinortheoryofgeneralrelativitycontext, thereexistst hepseudo-violation\nof Lorentz invariance[58, 59], which is caused by the derivatives of t he vierbein or local\nframe. The vierbein is defined in the tangent spacetime of a fixed poin t in spacetime\nmanifold, and the Lorentz transformation is just an algebraic oper ation in this fixed\ntangent spacetime. Whereas the derivatives of the vierbein must in volve the tangent\nspacetimes of different points in some sense, so it violates the local L orentz invariance.\nIn strict sense, the equivalence principle only holds for the linear ten sors, where the in-\nfluences of vierbein andnonlinear fields are absent. However, the f undamental dynamic\nequations of the field system should be intrinsically Lorentz invariant .\nAcknowledgments\nThe author is grateful to his supervisor Prof. Ta-Tsien Li for his e ncouragement and\nguidance.\n11References\n[1] M. Takeda, et al, Extension of the Cosmic-Ray Energy Spectrum Beyond the Pre-\ndicted Greisen-Zatsepin-Kuzmin Cutoff , Phys. Rev. Lett. 81 (1998) 1163.\nastro-ph/9807193\n[2] N. Hayashida, et al, Updated AGASA event list above 4×1019eV, Astron. J. 120\n(2000) 2190, astro-ph/0008102\n[3] A. A. Watson, Observations of Ultra-High Energy Cosmic Rays , Astron. Astro-\nphys. 441 (2005) 465-472, astro-ph/0507207\n[4] K. Greisen, Phys. Rev. Lett. 16 (1966) 748.\n[5] G. T. Zatsepin and V. A. Kuz’min, JETP Lett. 41 (1966) 78.\n[6] G. Amelino-Camelia, T. Piran, Planck-scale deformation of Lorentz symmetry as\na solution to the UHECR and the TeV- γparadoxes ,Phys. Rev. D64(2001)036005,\nastro-ph/0008107\n[7] F. A. Aharonian, et al, Reanalysis of the high energy cutoff of the 1997 Mkn 501\nTeV energy spectrum , astro-ph/0011483\n[8] M. Harwit, R. J. Protheroe, P. L. Biermann, TeV Cherenkov Events as Bose-\nEinstein Gamma Condensations , astro-ph/9908338\n[9] S. Coleman, S. L. Glashow, Cosmic Ray and Neutrino Tests of Special Relativity ,\nPhys. Lett. B405 (1997) 249-252, hep-ph/9703240\n[10] S. L. Glashow, et al, Comments on Neutrino Tests of Special Relativity , Phys.\nRev. D56 (1997) 2433-2434, hep-ph/9703454\n[11] S. Coleman, S. L. Glashow, Evading the GZK Cosmic-Ray Cutoff ,\nhep-ph/9808446\n[12] S. L. Glashow, Cosmological Searches for Photon Velocity Oscillations , Phys. Lett.\nB430 (1998) 54-56, astro-ph/9803202\n[13] S. Coleman, S. L. Glashow, High-Energy Tests of Lorentz Invariance , Phys. Rev.\nD59 (1999) 116008, hep-ph/9812418\n[14] F. W. Stecker, S. L. Glashow, New Tests of Lorentz Invariance Following from\nObservations of the Highest Energy Cosmic Gamma Rays , Astropart. Phys. 16\n(2001) 97-99, astro-ph/0102226\n12[15] A. G. Cohen. S. L. Glashow, A Lorentz-Violating Origin of Neutrino Mass? ,\nhep-ph/0605036\n[16] G. Amelino-Camelia, et al, Tests of quantum gravity from observations of γ-ray\nbursts, Nature 393 (1998) 763, astro-ph/9712103\n[17] P. Kosinski, J. Lukierski, P. Maslanka, J. Sobczyk, The classical basis for κ-\ndeformed Poincar´ e (super)algebra and the second κ-deformed supersymmetric\nCasimir, Mod. Phys. Lett. A10 (1995) 2599, hep-th/9412114\n[18] R. Aloisio1, PBlasi, P. L. Ghia, A. F. Grillo, Probing The Structure of Space-Time\nwith Cosmic Rays , Phys. Rev. D62 (2000) 053010, astro-ph/0001258\n[19] N. R. Bruno, G. Amelino-Camelia, J. Kowalski-Glikman, DEFORMED BOOST\nTRANSFORMATIONS THAT SATURATE AT THE PLANCK SCALE , Phys.\nLett. B522 (2001) 133-138, hep-th/0107039\n[20] Alex Granik, Magueijo-Smolin Transformation as a Consequence of a Speci fic\nDefinition of Mass,Velocity, and the Upper limit on Energy , hep-th/0207113\n[21] R. J. Szabo, Quantum Field Theory on Noncommutative Spaces , Phys. Rept. 378\n(2003) 207-299, hep-th/0109162\n[22] S. Mignemi, Transformations of coordinates and Hamiltonian formalism in de-\nformed Special Relativity , Phys. Rev. D68 (2003) 065029, gr-qc/0304029\n[23] S. Hossenfelder, Deformed Special Relativity in Position Space , Phys. Lett. B 649,\n310 (2007), gr-qc/0612167\n[24] A. Agostini, et al, Generalizing the Noether theorem for Hopf-algebra spaceti me\nsymmetries , hep-th/0607221\n[25] M. Daszkiewicz, J. Lukierski and M. Woronowicz, NONCOMMUTATIVE\nTRANSLATIONS AND *-PRODUCT FORMALISM , hep-th/0701152\n[26] S. Ghosh, P. Pal, Deformed Special Relativity and Deformed Symmetries in a\nCanonical Framework , hep-th/0702159\n[27] D. Mattingly, Modern Tests of Lorentz Invariance , Living Rev. Rel. 8 (2005) 5,\ngr-qc/0502097\n[28] R. Bluhm, Overview of the SME: Implications and Phenomenology of Lore ntz\nViolation , hep-ph/0506054\n13[29] D. Colladay, V. A. Kostelecky, Lorentz-violating extension of the standard model ,\nPhys. Rev. D58, 116002\n[30] V.A.Kostelecky, S.Samuel, Spontaneous Breaking Of Lorentz Symmetry In String\nTheory, Phys. Rev. D, 39, 683, (1989).\n[31] J. R. Ellis, N. E. Mavromatos, D. V. Nanopoulos, Probing models of quantum\nspacetime foam , ACT-10/99, CTP-TAMU-40/99, gr-qc/9909085\n[32] C. P. Burgess, et al, Loop-generated bounds on changes to the graviton dispersio n\nrelation, JHEP, 03, 043, (2002), hep-th/0201082\n[33] R. Gambini, J. Pullin, Nonstandard optics from quantum spacetime , Phys. Rev.\nD, 59, 124021, (1999), gr-qc/9809038\n[34] V. A. Kostelecky, TOPICS IN LORENTZ AND CPT VIOLATION ,\nhep-ph/0104227\n[35] I. L. Shapiro, Physical aspects of the spaceCtime torsion , Physics Reports 357\n(2002) 113C213\n[36] A. S. Belyaev, I. L. Shapiro, Torsion action and its possible observables , Nucl.\nPhys. B543 (1999) 20-46, hep-ph/9806313\n[37] A. Iorio, Three questions on Lorentz violation , physics/0612171\n[38] Paul L. Stanwix, M. E. Tobar, P. Wolf, C. R. Locke, E. N. Ivano v,Improved\ntest of Lorentz invariance in electrodynamics using rotati ng cryogenic sapphire\noscillators , Phys. Rev. D74, 081101(R) (2006)\n[39] P. Wolf, S. Bize, A. Clairon, G. Santarelli , Improved test of Lorentz invariance in\nelectrodynamics , Phys. Rev. D70, 051902(R)(2004)\n[40] P. Wolf, S. Bize, A. Clairon, A. N. Luiten, G. Santarelli, M. E. Toba r,Tests of\nLorentz Invariance using a Microwave Resonator , Phys. Rev. Lett. Vol.90(No.6),\n060402(2003)\n[41] H. Muler, S. Herrmann, C. Braxmaier, S. Schiller, A. Peters, Modern Michelson-\nMorley Experiment using Cryogenic Optical Resonators , Phys. Rev. Lett.\nVol.91(No.2), 020401(2003)\n[42] C. J. Berglund, et al, New Limits on Local Lorentz Invariance from Hg and Cs\nMagnetometer , Phys. Rev. Lett. Vol.75(No.10), 1879(1995)\n14[43] T. E. Chupp, et al, Results of a New Test of the Local Lorentz Invariance , Phys.\nRev. Lett. Vol.63(No.15), 1541(1989)\n[44] E. Fischbach, etal, Lorentz Noninvariance and the E¨ otv¨ os experiments , Phys. Rev.\nD32(No.1), 154(1985)\n[45] Y. Q. Gu, Canonical Form of Field Equations , Adv in Appl. Clif. Alg. V7(1), 13-\n24(1997), http://www.clifford-algebras.org/v7/v71/GU71.pdf, h ep-th/0610189\n[46] Y. Q. Gu, New Approach to N-body Relativistic Quantum Mechanics , IJMPA\nVol.22(No.11) 2007-2019(2007), hep-th/0610153\n[47] Y. Q. Gu, Mass-Energy Relation of the Nonlinear Spinor , hep-th/0701030\n[48] R. Finkelsten, et al, Phys. Rev. 83(2), 326-332(1951)\n[49] M. Soler, Phys. Rev. D1(10), 2766-2767(1970)\n[50] Y. Q. Gu, Some Properties of the Spinor Soliton , Adv in Appl. Clif. Alg. 8(1),\n17-29(1998), http://www.clifford-algebras.org/v8/81/gu81.pdf\n[51] Y. Q. Gu, Spinor Soliton with Electromagnetic Field , Adv in Appl. Clif. Alg.\nV8(2), 271-282(1998), http://www.clifford-algebras.org/v8/82/gu82 .htm\n[52] Y. Q. Gu, Characteristic Functions and Typical Values of the Nonline ar Dark\nSpinor, hep-th/0611210\n[53] Y. Q. Gu, A Cosmological Model with Dark Spinor Source , to appear in IJMPA,\ngr-qc/0610147\n[54] Y. Q. Gu, Accelerating Expansion of the Universe with Nonlinear Spin ors,\ngr-qc/0612176\n[55] Y. Sofue, Y. Tutui, M. Honma, et al, CENTRAL ROTATION CURVES OF SPI-\nRAL GALAXIES , astro-ph/9905056\n[56] M. Pavsi, The Landscape of Theoretical Physics , Published by Kluwer Academic\nPublishers, 2001, gr-qc/0610061\n[57] M. Tegmark, The Mathematical Universe , arXiv:0704.0646\n[58] Y. Q. Gu, Representation of the Vierbein Formalism , gr-qc/0612106\n[59] Y. Q. Gu, Simplification of the covariant derivatives of spinors , gr-qc/0610001\n15" }, { "title": "1707.05192v2.Damping_of_gravitational_waves_by_matter.pdf", "content": "arXiv:1707.05192v2 [gr-qc] 17 Oct 2017Damping of gravitational waves by matter\nGordon Baym,a,bSubodh P. Patil,band C. J. Pethickb,c\naDepartment of Physics, University of Illinois, 1110 W. Gree n Street, Urbana, IL 61801-3080\nbThe Niels Bohr International Academy, The Niels Bohr Instit ute,\nUniversity of Copenhagen, Blegdamsvej 17, DK-2100 Copenha gen Ø, Denmark\ncNORDITA, KTH Royal Institute of Technology and Stockholm Un iversity,\nRoslagstullsbacken 23, SE-10691 Stockholm, Sweden\n(Dated: October 18, 2017)\nWe develop a unified description, via the Boltzmann equation , of damping of gravitational waves\nby matter, incorporating collisions. We identify two physi cally distinct damping mechanisms –\ncollisional and Landau damping. We first consider damping in flat spacetime, and then generalize\nthe results to allow for cosmological expansion. In the first regime, maximal collisional damping of a\ngravitational wave, independentofthedetails ofthecolli sions inthematteris, as weshow, significant\nonly when its wavelength is comparable to the size of the hori zon. Thus damping by intergalactic or\ninterstellar matterfor all butprimordial gravitational r adiationcanbeneglected. Althoughcollisions\nin matter lead to a shear viscosity, they also act to erase ani sotropic stresses, thus suppressing\nthe damping of gravitational waves. Damping of primordial g ravitational waves remains possible.\nWe generalize Weinberg’s calculation of gravitational wav e damping, now including collisions and\nparticles of finite mass, and interpret the collisionless li mit in terms of Landau damping. While\nLandau damping of gravitational waves cannot occur in flat sp acetime, the expansion of the universe\nallows such damping by spreading the frequency of a gravitat ional wave of given wavevector.\nPACS numbers:\nI. INTRODUCTION\nThe opening of a new window on the universe through\nthe ongoing observations of gravitational waves [ 1] un-\nderlines the importance of reexamining how they prop-\nagate through the matter in the universe, and asking\nwhat gravitational wave measurements can teach one\nabout this matter. Half a century ago, Hawking showed\nthat if matter could be treated in the hydrodynamic\nlimit the damping rate of a gravitational wave would be\nγ= 16πGη, whereGis Newton’s gravitational constant\nandηthe viscosity of the matter [ 2,3]. Using this result,\nGoswami et al. [ 4] argued that gravitational wave obser-\nvations could be used to constrain the viscosity of dark\nmatter between the source and Earth. But, as Hawking\nfirst pointed out, there are in general too few collisions\nin matter for hydrodynamics to be valid, and the damp-\ning would be less than the hydrodynamic result. Ref-\nerence [5] estimated damping in the almost collisionless\nlimit by investigating the response of individual particles\nto a gravitational wave and found that the damping rate\nof the wave by nonrelativistic particles is\nγ∼Gnm\nω2/parenleftBig¯v\nc/parenrightBig21\nτ. (1)\nHereωisthefrequencyofthewave, ntheparticledensity,\nmthe particle mass, ¯ vthe typical particle velocity, and\nτthe particle-particle collision time; the damping is ∼\n1/(ωτ)2smaller than the viscous result.\nIn additiontodamping bycollisionsin matter, gravita-\ntional waves can also be attenuated by Landau damping,\nin which particles surf the gravitational wave and ex-\ntract its energy, first proposed for gravitational waves inRef.[6]. Thiseffectwasoriginallyinvestigatedinthecon-\ntext of plasma physics [ 7], then in galactic dynamics [ 8],\nand later in quantum chromodynamic plasmas [ 9]. In a\nstaticflat universemassiveparticlescannotproduceLan-\ndaudampingsincetheymovemoreslowlythan agravita-\ntional wave. In an expanding universe, however, Landau\ndamping becomes possible, as we show, since the expan-\nsion in the presence of matter effectively spreads the fre-\nquency of a gravitational wave. Indeed the damping of\ncosmological gravitational waves by non-interacting neu-\ntrinos, as first proposed by Weinberg [ 10] and expanded\nupon in Refs. [ 11,12], can in fact be understood in terms\nof Landau damping, as we indicate below.\nOur aim in this paper is to present a unified treatment\nof the damping of gravitational waves by matter, for ar-\nbitrary collision rates, thus encompassing the hydrody-\nnamic and nearly collisionless limits studied earlier, as\nwell as cosmological expansion. We begin, in Sec. II,\nby considering a weak gravitational wave propagating\nthrough a dilute gas of colliding particles of arbitrary\nmass in an otherwise flat spacetime, and calculate, in\nSec.III, the response of the matter to the wave using\nthe Boltzmann equation. For simplicity we work in the\nrelaxation or collision time approximation.\nAs we show, in Sec. IV, the maximum damping of\nthe amplitude of a gravitational wave with frequency\nωis or order 1 /(ωτU) where τUis the age of the uni-\nverse. Thus collisional damping by matter of gravita-\ntional waves generated by astrophysical sources cannot\nprovide useful information about the nature of matter in\nthe universe. Furthermore, damping by dense environ-\nments surrounded localized sources of gravitational ra-\ndiation is, as we estimate, insignificant. After a general2\ndiscussion of Landau damping in Sec. VI, we generalize\nthe Boltzmann equation results in Sec. VIIto describe\ncollisional damping by particles of arbitrary mass in the\npresence of an expanding cosmological background.\nII. STATIC SPACETIME\nInitially, we do not include the expansion of the uni-\nverse, and consider rather the Minkowski space metric\nwith a gravitational wave superimposed:\nds2=−dt2+gijdxidxj, (2)\nwhere\ngij=δij+hij(/vector r,t), (3)\nwithhijthe weak metric perturbation caused by a grav-\nitational wave. We work with hijin the transverse–\ntraceless gauge, and generally set c= 1.\nThe effects on a gravitational wave passing through\nmatter are given in terms of the gravitational wave equa-\ntion in the transverse traceless gauge\n∂µ∂µhij=/parenleftbigg\n−∂2\n∂t2+∇2/parenrightbigg\nhij=−16πGπij,(4)\nwhereπijis the transverse traceless part of the matter\nstress tensor, Tij,M, defined by\nπi\nj≡Ti\nj,M−δi\nj\n33/summationdisplay\nk=1Tk\nk,M. (5)\nIn equilibrium,1\n3/summationtext3\nk=1Tk\nk,Mis simply the pressure Pof\nthe matter.\nThe effect of a gravitational wave on a particle is given\nin terms of the dispersion relation,\npµpνgµν+m2= 0, (6)\nwhich in the present case implies that the particle energy\nǫis given by\nǫ2=gijpipj+m2. (7)\nThus a weak gravitational wave changes the particle dis-\npersion relation from ǫ0=/radicalbig\np2+m2to\nǫ=ǫ0+δǫ. (8)\nTo first order in hij\nδǫ=1\n2hijpipj\nǫ0=−1\n2hijpipj\nǫ0, (9)\nsince to this order, hij=−hij.III. BOLTZMANN EQUATION\nWe treat the matter as a dilute gas and calculate πij\nfrom the Boltzmann equation for the matter. We first\nwrite the non-linear Boltzmann equation for the particle\ndistributionfunction f(ri,pj)asafunctionoftheparticle\npositions and canonical momenta,\n/parenleftbigg∂\n∂t+/vector∇pǫ·/vector∇r−/vector∇rǫ·/vector∇p/parenrightbigg\nf=C,(10)\nwhereCis the collision term. Here position gradients\nare taken with respect to ri, and momentum gradients\nwith respect to pi. This form of the equation is valid for\nrelativistic as well as non-relativistic particles.\nThe conservation laws of energy and momentum are\nfound by taking the moments of Eq. ( 10) with respect\ntopiandǫ; assuming that collisions conserve the total\nenergy and momentum of the particles, we find (as in\nstandard Fermi liquid theory)\n∂\n∂t/integraldisplay\npǫf+∇i/integraldisplay\npǫvif=/integraldisplay\np∂ǫ\n∂tf, (11)\nand\n∂\n∂t/integraldisplay\nppif+∇jTj\ni,M=−/integraldisplay\np(∇riǫ)f, (12)\nwhere/integraltext\np≡g/integraltext\nd3p/(2π)3, withgthe number of internal\nstates, e.g., spin, and\nTj\ni,M=/integraldisplay\nppivjf=gjk/integraldisplay\nppipk\nǫf (13)\nis the matter stress tensor.\nWith Eq. ( 9) the right side of Eq. ( 11) becomes\n/integraldisplay\np∂ǫ\n∂tf=1\n2∂hij\n∂t/integraldisplay\nppipj\nǫf=1\n2∂hij\n∂tTij,M,(14)\nso that the change in energy of the matter is given by\n∂E\n∂t=1\n2/integraldisplay\nd3r∂hij\n∂tπij. (15)\nOnly the transverse-traceless part of the stress tensor,\nEq. (5), enters Eq. ( 14).\nNote that, for a wave of the form H(z−ct), say, the\nright side of Eq. ( 12) becomes\n−/integraldisplay\np∂ǫ\n∂zf=1\n2∂hij\n∂tπij, (16)\nindicating that as momentum qis transferred from the\ngravitational wave, energy qis also transferred. Since\nπijis itself at least of first order in hijthe energy and\nmomentum transfers are second order and higher in the\namplitude of the gravitational wave.3\nWe turn now to calculating the transverse-traceless\npart of the matter stress tensor, Eq. ( 13); to linear order\ninhij,\nδTij,M=/integraldisplay\nppipj\nǫ0/bracketleftbigg\nδf−δǫ\nǫ0f0/bracketrightbigg\n(17)\nwhereδf=f−f0,withf0thedistributionfunctioninthe\nabsence of hij. Theδǫterm arises from the dependence\nofǫin the denominator on hij, Eq. (9). Subtracting out\nthe trace, we find, after using the vanishing trace of hij\nand writing the equilibrium pressure of the matter as/integraltext\np(p2/3ǫ0)f0, that\nπij=/integraldisplay\nppipj\nǫ0/bracketleftbigg\nδf−δǫ\nǫ0f0/bracketrightbigg\n+hij/integraldisplay\npp2\n3ǫ0f0.(18)\nThesecondtermofthisexpressionismanifestlytraceless;\nthe trace of the first term vanishes since the integrations\nover both δfandδǫ, being symmetric in angles, vanish.\nThe latter two terms that contain f0can be sim-\nply combined, with an integration by parts using the\ntransverse-traceless structure of hij, into a term propor-\ntional to ∂f0/∂ǫso that Eq. ( 18) becomes,\nπij=/integraldisplay\nppipj\nǫ0/bracketleftbigg\nδf−δǫ∂f0\n∂ǫ/bracketrightbigg\n. (19)\nThis combination of terms falls out naturally, as we shall\nsee, from the Boltzmann equation.\nCollisions between the particles, prior to freeze-out,\ntend to bring the distribution function into a local equi-\nlibrium in the presence of hij:\nf→fh=1\neβ(ǫ−µ)∓1, (20)\nwhereǫ, givenby Eq.( 7), depends on hij;βis the inverse\ntemperature, and µthe particle chemical potential. Note\nthat to first order in hij,\nfh=f0+δǫ∂f\n∂ǫ. (21)\nFor simplicity we employ a collision time approximation.\nSince the only disturbances relevant here involve spher-\nical harmonics of degree greater than one, we can write\nthe collision term as\nC=−f−fh\nτ=−1\nτ/parenleftbigg\nδf−δǫ∂f\n∂ǫ/parenrightbigg\n,(22)\nwhereτis the collision time, and δf−δǫ∂f/∂ǫ is the\ndeviation of the distribution from local equilibrium; the\nadditional terms commonly introduced to ensure conser-\nvation of particle number and total momentum (which\ninvolve spherical harmonics of degree zero and one) are\nnot relevant [ 7]. The linearized Boltzmann equation then\nreduces to\n/parenleftbigg∂\n∂t+1\nτ+/vector v·/vector∇/vector r/parenrightbigg\nδf=/parenleftbigg\n/vector v·∇rδǫ+1\nτδǫ/parenrightbigg∂f0\n∂ǫ.\n(23)Withhij(/vector r,t) =ei(/vector q·/vector r−ωt)hij, and Fourier transforming\nin space and time we find the solution of Eq. ( 23),\nδf=∂f0\n∂ǫ/parenleftbigg−/vector q·/vector v+i/τ\nω−/vector q·/vector v+i/τ/parenrightbigg\nδǫ. (24)\nThe deviation from local equilibrium is thus given by\nδf−∂f0\n∂ǫδǫ=−/parenleftbiggω\nω−/vector q·/vector v+i/τ/parenrightbigg∂f0\n∂ǫδǫ.(25)\nWe then find the general result,\nπij(q,ω) =/integraldisplay\nppipj\nǫ0δǫ/parenleftbiggω\n/vector q·/vector v−ω−i/τ/parenrightbigg∂f0\n∂ǫ.\n(26)\nFourier transformed back to time, the stress tensor is\nπij(q,t) = (27)\n−/integraldisplay\nppipj\nǫ0/integraldisplayt\n−∞dt′e−(iq·v+1/τ)(t−t′)∂f0\n∂ǫ˙δǫ(t′).\nThe response can be written in the more general form\nπij(q,ω) =−ωhij/integraldisplaydω′\n2πA(q,ω′)\nω−ω′+iξ,(28)\nwhere the spectral function is\nA(q,ω) =−2/integraldisplay\np/parenleftbiggpxpy\nǫ0/parenrightbigg21/τ\n(ω−/vector q·/vector v)2+1/τ2∂f0\n∂ǫ,\n(29)\nandξis a positive infinitesimal. In the collisionless limit,\n1/τ→0, and for relativistic particles,\nA(q,ω) =πρ∝angb∇acketleft(1−ζ2)2δ(ω−qζ)∝angb∇acket∇ight, (30)\nwhereρis the energy density of the excitations, and the\nangularbracketsdenotetheaverageover ζ≡ˆq·ˆv. Fourier\ntransformed back to time,\nπij(q,t) =−/integraldisplayt\n−∞dt′/integraldisplaydω\n2πe−iω(t−t′)A(q,ω)˙hij(t′).\n(31)\nThe damping of gravitational waves is governed by the\nimaginary part of the response,\nℑ/parenleftbiggπij\nhij/parenrightbigg\n=−ω/integraldisplay\np/parenleftBigpipj\nǫ/parenrightBig2 1/τ\n(ω−/vector q·/vector v)2+1/τ2∂f0\n∂ǫ.(32)\nIn the collision-dominated regime, τ≪1/ω; doing the\nangular averages in the integrals we have\nπij=iτω/integraldisplay\nppipj\nǫ0δǫ∂f0\n∂ǫ=−iτω\n15/integraldisplay\npp4\nǫ2\n0∂f0\n∂ǫhij\n=−η˙hij. (33)4\nThe viscosity calculated in the relaxation time approxi-\nmation is\nη=−/integraldisplay\np/parenleftbiggpipj\nǫ0/parenrightbigg2∂f0\n∂ǫ0τ, (34)\nwithi∝negationslash=j. In this limit the damping rate of a gravita-\ntional wave, from Eq. ( 4), is 16πGη, in agreement with\nearlier hydrodynamic treatments [ 2,3].\nFor non-relativistic matter, T≪mc2, the/vector q·/vector vin the\ndenominator of Acan be neglected, and we have\nπij≃ω\nω+i/τP hij. (35)\nFrom the imaginary part of Eq. ( 4), the dispersion re-\nlation of gravitational waves is\nω≃q+8πG\nωπij\nhij, (36)\nso that the damping of a wave is given by\nℑω=8πG\nωℑ/parenleftbiggπij\nhij/parenrightbigg\n. (37)\nFor fully relativistic matter, in the nearly collisionless\nlimit, to first order in 1 /τω,\nπij=−P/parenleftbigg\n1−2i\nτω/parenrightbigg\nhij, (38)\nwhile in the collision-dominated regime, πijis given by\nEq. (33).\nIV. MAXIMUM COLLISIONAL DAMPING\nAs one can see from Eq. ( 32),ℑ(πxy/hxy) has its max-\nimum magnitude for ωτ∼1. The possible damping is\nthus limited by\nMax(|ℑω|)/lessorsimilar−8πG\nω/integraldisplay\npp4\n15ǫ2\n0∂f0\n∂ǫ≤8πGP\nω,(39)\nwherePis the total local pressure of the matter under\nconsideration, which gives rise to the damping. It is in-\nstructive to write this bound in terms of the age of the\nuniverse, defined by\n1\nτ2\nU=8πG\n3¯ρ=/parenleftbigg˙a\na/parenrightbigg2\n, (40)\nwhereais the cosmological scale parameter and ¯ ρthe\nmean mass density of the universe. Since the mean pres-\nsure obeys ¯P≤¯ρ/3, we find\nMax(|ℑω|)/lessorsimilarP\n¯P1\nωτ2\nU. (41)A wave traversing matter will average the local pressure,\nand thus we can conclude,\nMax(|ℑω|)/lessorsimilar1\nωτ2\nU, (42)\nindicating that damping of a gravitational wave by mat-\nter in the universe can only be significant for a wave of\nfrequency of order 1 /τU. This bound includes all contri-\nbutions from dark matter particles as well [ 4].\nTo express the result ( 42) in another way, the colli-\nsionaldamping ofagravitationalwavewithin the charac-\nteristic expansion time of the universe is of order 1 /ωτU.\nForω∼103s−1, as in the recent gravitational wave de-\ntections [ 1], andτU∼1018s, this ratio is ∼10−21. Col-\nlisional damping in intergalactic or interstellar matter of\ngravitational waves produced by astrophysical sources is\nnot useful to determine the nature of matter in the uni-\nverse. This result is valid for any particle-like form of\ndark matter, including that in a possible shadow uni-\nverse [13] or matter that only interacts with gravitation-\nallysuppressedinteractions[ 14]. Furthermore, collisional\ndamping in locally high dense environments, e.g., in the\nneighborhood of mergers of black holes or neutron stars,\nis also negligible for gravitational waves produced by as-\ntrophysical sources, as we argue in the next section. On\nthe other hand for primordial gravitational waves with\nω∼10−16−10−15, one has 1 /ωτU∼10−3−10−2, an\neffect that could play a role in interpretation of future\nprecision measurements of the spectrum of primordial\ngravitational radiation.1\nAs we discuss in the following sections, Landau damp-\ning cannot occur in a flat spacetime. Even in an expand-\ning universe, the Landau damping rate is ∼1/ω2t3\nU, so\nthat the total damping within the expansion time of the\nuniverse is ∼1/(ωtU)2, a factor 1 /ωtUsmaller than the\nmaximum collision damping.\nOne can write the contribution to the damping from a\nparticular component, s, of the matter, e.g., neutrinos or\ndark matter, in the form\n|ℑω|s≡nsσGW,s, (43)\nwhereσGW,sis the graviton scattering cross section on\nparticles of species s, andnsis the number density of\nspeciess. The ratio of the cross section to the Planck\nlength,ℓPl, squared is essentially bounded above by\nσGW,s\nℓ2\nPl/lessorsimilar∝angb∇acketleftpv∝angb∇acket∇ights\nω(44)\nwhere∝angb∇acketleftpv∝angb∇acket∇ightsis the mean product of the particle momen-\ntum and velocity of species s, which is of order the tem-\nperature for a species in thermal equilibrium, or the tem-\nperature at which the species froze out. Thus in general,\nσGW,i\nℓ2\nPl/lessorsimilarT\nω, (45)\n1We thank Vicky Kalogera and Chris Pankow for this observatio n.5\nwith the above understanding of T.\nV. MAXIMAL COLLISIONAL DAMPING IN\nDENSE ENVIRONMENTS\nWe looknow at the damping ofgravitationalwavepro-\nduced by binary astrophysical sources, as the waves pass\nthrough the dense medium surroundingthe sources. Col-\nlisional damping is limited by |ℑω|< γmax, where from\nEq. (39),\nγmax∼1\nM2\nplP\nω=w\nM2\nplρ\nω; (46)\nwe have introduced the Planck mass M−2\npl= 8πGand\nwritten the relation between the pressure and the energy\ndensity by the equation of state parameter w=P/ρ.\nAssuming the gravitational wave source to a binary\nsystem inside a regionsurroundedby matter with a given\ndensity profile with equation of state parameter w, we\nfind collisional damping along a line of sight to be signif-\nicant if\n/integraldisplayR\n0drγmax∼1 (47)\nwhereRis a physical radius enclosing the ambient mat-\nter. A reasonable first estimate is simply to associate\nthe integral with the characteristic size Rcof the dense\nregion:\n/integraldisplayR\n0drγmax∼w\nM2\nplρ\nωRc∼w\nM2\nplR2cM\nω(48)\nwhereMis the total mass in the region with character-\nistic size Rc. To go beyond this estimate, one could take\nthe density profile from detailed calculations, e.g. the\nprofile of a typical dark matter halo (determined via phe-\nnomenological models that fit N-body simulations such\nas the Navarro-Frenk-White or Einasto density profiles\n[15]), and find numerical factors that little affect the con-\nclusion. In “natural” units, one solar mass M⊙∼1066\neV, 1 Hz ∼10−15eV,M−1\npl=ℓPl∼10−35m, and 1 kpc\n∼1019m, one has\n/integraldisplayR\n0drγmax∼w10−27\n(Rc/kpc)2/parenleftbiggM\nM⊙/parenrightbigg1\n(ν/Hz),(49)\nwhereν=ω/2π. We first consider a typical galactic halo\nsurrounding a binary system source of the gravitational\nwave. Here, typically M∼1012M⊙,Rc∼100 kpc, so\nthat\n/integraldisplayR\n0drγmax∼w10−19\n(ν/Hz)(50)\nwhich is feeble for all astrophysical sources within the\nhalo.We next consider the ambient region surrounding a bi-\nnarysystem similar to that which gaveriseto GW150914\n–containingadensedistributionof(notnecessarilydark)\nmatter. If the ambient matter has a mass comparable\nto that of the binary system localized within some re-\ngionRc≫Rs, the Schwarzschild radius associated with\nthe total mass of the binary system, we find that at\nthe lowest frequencies of the binary system, the factor\nM/M⊙×(ν/Hz)−1isO(1); furthermore for the expected\nnon-relativisticlow pressure surroundingmatter, w∼c2\ns,\nthe square of the adiabatic sound velocity. Thus\n/integraldisplayR\n0drγmax∼c2\ns1011\n(Rc/m)2. (51)\nwhich can only be of order unity if the ambient matter\nis localized to within a radius Rc/lessorsimilarcs×300km around\nthe binary system, which even for mildly non-relativistic\nambientmatter(e.g. cs∼ O(10−1))wouldrequireamass\ncomparable to that of the binary system to be crammed\ninto a region comparable to the Schwarzschild radius of\nthe final merged black hole ( ∼70 km∼Rs); such a high\ndensity is contrary to our initial assumption that Rc≫\nRs. Requiring that this matter be distributed within\na region an order of magnitude larger than the binary\nsystem yields/integraltextR\n0drγmax∼c2\ns≪1. We conclude that\na distribution of non-relativistic matter of high density\nsurrounding the source of gravitational radiation is not\ncapable of significantly damping gravitational radiation.\nMore realistically, one is in general far from the con-\ndition of maximal collisional damping, that the collision\nrateτ−1, be comparable to the frequency of the gravita-\ntional wave. Maximal collisional damping is a highly un-\nlikely prospect even for relativistic matter jets and lobes\nclose to the mergerof neutron star/blackhole binary sys-\ntems [16,17]. To see this we write roughly, τ−1=nσv,\nwherenis the density of particles, σis a particle-particle\nscattering cross section, and va mean particle velocity.\nIn terms of the mass, M, and characteristic radius, Rc,\nof the dense environment,\n1\nτ∼M\nM⊙1035\n(Rc/m)3/parenleftbiggσ\nfm2/parenrightbiggv\ncs−1, (52)\nClearly, for a typical nuclear or particle physics cross sec-\ntion, the above is much larger than the typical frequency\nofanastrophysicalbinarysystembymanyordersofmag-\nnitude, so that ωτ≪1.\nVI. LANDAU DAMPING: GENERAL\nCONSIDERATIONS\nIn flat space in the collisionless limit ( τ→ ∞) Eq. (32)\nreduces to\nℑ/parenleftbiggπij\nhij/parenrightbigg\n=πω/integraldisplay\np/parenleftbiggpipj\nǫ0/parenrightbigg2\nδ(ω−/vector q·/vector v)∂f0\n∂ǫ,(53)6\na result describing Landau damping, the decay of the\nmode into a single particle–hole pair.2The particle-hole\nexcitationsarespacelike. Foragravitationalwave, ω=q,\nthe integral vanishes except possibly for massless parti-\ncles moving in the same direction, say ˆ z, as the grav-\nitational wave. However, for such particles, the factor\np2\nip2\nj→p2\nxp2\nyvanishes; Landau damping is forbidden in\nthe absence of cosmologicalexpansion. Following general\nremarks on Landau damping in this section we show in\nthe following section how the collisionless damping pro-\ncess described by Weinberg [ 10] can be understood as a\ngeneralization of Landau damping, driven by the expan-\nsion of the universe.\nIn an expanding universe, the gravitational wave en-\nergy changes during expansion, i.e., the frequency of the\ngravitational wave is not constant, since the expansion\nabsorbs energy from the wave. This energy loss is differ-\nent from Landau damping by the matter traversedby the\nwave. When the phase velocity of the wave is different\nfrom the group velocity of the excitations in the matter,\nenergy in flat spacetime is pumped to and fro between\nthe wave and the matter, but the net rate of transfer\nis zero because the energy transferred in one half-cycle\nof the wave is exactly cancelled by the loss in the other\nhalf-cycle. In an expanding universe, however, the can-\ncellation is incomplete.\nWe recall the energy loss caused by expansion. A weak\ngravitational wave of period small compared with the\nage of the universe behaves in the absence of matter as\nhij=χ(u)e−iqu, whereuis conformal time, related to\ncoordinate time by du=dt/a(t); as we see in the next\nsection, χ(u)∝1/a(u). This structure is expected on\nthe basis of simple arguments: the energy of a gravita-\ntional wave packet is proportional to the energy density\nin the wave packet times the volume of the packet. The\nenergy density of the wave varies as gii(∂χ/∂xi)2∼a−4\nand the volume of the packet varies as a3, so the total\nenergydecreasesas 1 /a. This result alsoagreeswith sim-\nple redshift arguments: The energy of a massless particle\nvaries as 1+ z∼1/aowing to the expansion of the uni-\nverse, and thus the energy density measured in locally\nMinkowskian spacetime ( ds2=−dt2+ (d/vector r)2) varies as\n1/(1+z)4. For example, the redshift of the first LIGO\nevent GW150914 was z= 0.09+0.03\n−0.04[1], leading to an en-\nergy density reduction by a factor ≃1−1/(1.09)4≈30%\nfrom cosmological expansion. By comparison, even were\nthe intervening matter collisionless, Landau damping of\nthe wave would be totally negligible, ∼1/(ωtU)2.\n2In the language of quantum mechanics, the damping may be re-\ngarded as the creation, with amplitude ∝1/(ω−/vector q·/vector vs), of a\nvirtual single particle–hole pair, which subsequently dec ays into\ntwo real particle–hole pairs. Equation ( 32) can be understood,\nwhen 1/τ→0, as this amplitude squared, summed over all mo-\nmenta.VII. GRAVITATIONAL WAVE DAMPING\nWITH COSMOLOGICAL EXPANSION\nWe turn now to relate the gravitational radiation\ndamping derived by Weinberg [ 10] to the calculations\nabove, and to Landau damping in collisionless plasmas,\ndriven by the expansion of the universe. We first gener-\nalize the treatment of [ 10] to allow for massive particles\nand collisions, working in conformal time, u, related to\ncoordinate time by dt=adu, whereais the scale pa-\nrameter of the expansion. The metric in the presence of\nexpansion and a gravity wave is given by\nds2=a(u)2[−du2+(δij+hij)dxidxj].(54)\nOwing to the explicit a2, upper and lower components of\nvectors are related by xµ=a2xµto zeroth order in h.\nIn addition, the energy of a particle in the metric ( 54) is\ngiven, in the absence of hij, by\nǫ2\n0=p2\ni+a2m2. (55)\nWe study, following [ 10], the evolution of the coupled\ngravitational wave – matter system, after an initial time\nu0at which the matter distribution function is given as\nf0, essentially the fhin Eq. (20). Since f0includes the\nmetric perturbations hij(u0) at that time, the additional\nperturbations of the energy that modify the distribution\nfunction by δf(u) at later time depend only on the devi-\nation from hij(u0), that is,\nδǫ=−pipj\n2ǫ0[hij(u)−hij(u0)]. (56)\nThe Boltzmann equation in conformal time, Fourier\ntransformed in space (cf. Eq. ( 23)) is\n/parenleftbigg∂\n∂u+1\nτc+i/vector q·/vector v/parenrightbigg\nδf=∂f0\n∂ǫ/parenleftbigg1\nτc+i/vector q·/vector v/parenrightbigg\nδǫ,(57)\nThe particle velocity, v, the distribution function f0, and\nthe conformalcollision time τc=τ/a, aredirectly depen-\ndent on the scale factor. (More generally τwill depend\non the cosmological epoch and thus contain further de-\npendence on the scalefactor, a question wedo not pursue\nhere.) In particular\nvi=∂ǫ0/∂pi=pi/radicalbig\npjpj+m2a(u)2. (58)\nSimilarly, an equilibrium distribution function,\nf0=1\neǫ0/T0(u)∓1(59)\n(with∓for bosons or fermions) depends on athrough\nthe term m2a2inǫ, andT0(u)/a(u) is the temperature\nofthe darkmatter. Formasslessparticles, T0is constant.\nIn addition τcis a function of the ambient density along\nthe trajectoryofthe gravitationalwave, andso in general7\ndepends on time through its dependence on the scale fac-\ntor, aswell as throughthe evolvingparticle distributions.\nEquation ( 57) has the general solution\nδf(u,p) =/integraldisplayu\nu0du′∂e−Φ(u,u′)\n∂u′∂f0\n∂ǫ(u′)δǫ(u′); (60)\nwhere we write\nΦ(u,u′)≡/integraldisplayu\nu′du′′/parenleftbigg1\nτc(u′′)+i/vector q·/vector v(u′′)/parenrightbigg\n=ℓ(u,u′)+iˆq·ˆps(u,u′), (61)\nin terms of\nℓ(u,u′) =/integraldisplayu\nu′du′′\nτc(u′′), (62)\nand\ns(u,u′) =q/integraldisplayu\nu′du′′v(u′′), (63)\nwhich is the displacement of the particle in the interval\nu′toutimes the wavevector. The deviation from local\nequilibrium in ( 65) is therefore\nδf−∂f\n∂ǫδǫ=−/integraldisplayu\nu0du′e−Φ(u,u′)∂\n∂u′/bracketleftbigg∂f0\n∂ǫδǫ(u′)/bracketrightbigg\n.\n(64)\nOn a cosmological background,\nπij=/integraldisplay\nppipj√−gǫ0/bracketleftbigg\nδf−∂f\n∂ǫδǫ/bracketrightbigg\n, (65)\nthe generalization of Eq. ( 19), where in the absence of a\ngravitational wave,√−g=a4. With Eq. ( 64), we then\nhave\nπij(q,u) =−/integraldisplay\nppipj\na(u)4ǫ0/integraldisplayu\nu0du′e−Φ(u,u′)\n×∂\n∂u′/bracketleftbigg∂f0\n∂ǫδǫ(u′)/bracketrightbigg\n. (66)\nThis equation is the direct generalization of Eq. ( 28) to\nan expanding spacetime.\nUsingǫdǫ=pdpwe have\nπij=1\na(u)4/integraldisplay\nppipjpkpl\n2pǫ0(u)/integraldisplayu\nu0du′e−Φ(u,u′)\n×∂\n∂u′/bracketleftbigg∂f0\n∂p(hkl(u′)−hkl(u0))/bracketrightbigg\n.(67)\nSincek∝negationslash=lthe angular average above has the form\n(δikδjl+δilδjk)K(s) where in terms of spherical Bessel\nfunctions K(s) =j2(s)/s2, and explicitly,\nK(s)=/integraldisplaydΩ\n4πe−iζ s(1−ζ2)2sin2ϕcos2ϕ\n=−sins\ns3−3coss\ns4+3sins\ns5, (68)withζ= cosθ; thus\nπij(q,u) =1\na(u)4/integraldisplay\npp3\nǫ0(u)/integraldisplayu\nu0du′e−ℓ(u,u′)K(s)\n×∂\n∂u′/bracketleftbigg∂f0\n∂p(hij(u′)−hij(u0))/bracketrightbigg\n.(69)\nIn the masslesslimit, we integratethe momentum deriva-\ntive by parts, using ǫdǫ=pdp, and noting that s→\nq(u−u′), to obtain\nπij=−4¯ρ/integraldisplayu\nu0du′e−ℓ(u,u′)K(q(u−u′))h′\nij(u′),(70)\nwhere the prime denotes d/du, and\n¯ρ=1\na4/integraldisplay\nppf0 (71)\nis the mass density of the matter. Away from the mass-\nless limit, generalizing Ref. [ 10], we find extra contribu-\ntions from the pdependence of sinK. We see from Eq.\n(69) or (70), that, as expected, the net effect ofcollisional\ninteractionsistoefficientlyeraseanisotropicstresses,and\nhence limit their ability to damp gravitational waves.\nSince astrophysical sources of gravitationalwaves have\ncharacteristic frequencies much greater than the inverse\nHubble scale at late times, we can expand the time de-\npendence of the mode functions in powers of a′/(aq). For\nsuch a wave, the spatial Fourier component /vector qobeys the\nequation of motion,\nh′′\nij+2a′\nah′\nij+q2hij= 16πGa2πij. (72)\nThe solution for hij(q,u) in the absence of matter is, to\nlowest order in a′/(aq),\nhij(q,u)∝e−iqu\na(u)(73)\n(during radiation domination, this result is exact). We\nassume that the gravitational wave is in the form of a\nwavepacket for which\nhij(/vector r,u) =/integraldisplayd3q\n(2π)3e−iqu\na(u)F(q), (74)\nwhereF(q) is localized about a wave vector /vector q0. We con-\nsider the absorption by a region of matter much smaller\nthan the horizon size.\nTo see the mechanism of Landau damping in the ab-\nsence of collisions, we calculate the damping of the wave\ndirectly in terms of the energy transfer to the matter,\nwriting, from Eq. ( 15), using conformal time\n∂E\n∂u=−1\n2/integraldisplay\nd3rh′\nij(/vector r,u)πij(/vector r,u)\n=−1\n2ℜ/integraldisplayd3q\n(2π)3h′\nij∗(q,u)πij(q,u). (75)8\nWe work in the massless limit, in order to illustrate the\nphysics with the fewest complications. Then f0does not\ndepend on a, and Φ(u,u′)→iqζ(u−u′), and one has,\n∂E\n∂u=−1\n4a(u)4ℜ/integraldisplayd3q\n(2π)3h′\nij∗(q,u)/integraldisplay\nppipjpkpl\np2\n×/integraldisplayu\nu0du′e−i/vector q·ˆp(u−u′)∂f0\n∂ph′\nkl(q,u′)\n=−1\n16a(u)4ℜ/integraldisplayd3q\n(2π)3h′\nij∗(q,u)/integraldisplay\npp2∂f0\n∂p\n×(1−ζ2)2/integraldisplayu\nu0du′e−iqζ(u−u′)h′\nij(q,u′).(76)\nFrom Eq. ( 73), we see that h′\nij(u) =−(iq+\nH(u))hij(u), where H ≡a′(u)/a(u). The explicit H(u),\nwhichissmallrelativeto the qtermandforanastrophys-\nical gravitational wave produces only a small correction\nto the Landau damping, can be neglected. Over the time\nspan of a gravitational wavepacket transversing a given\nregion of matter, the scale factor a(u′) inhij(u′) can be\nexpanded as a(u′) =a(u)+a′(u)(u′−u)≃a(u)eH(u′−u).\nThus the u′integral can be written as\n−iqe−iquF(q)\na(u)/integraldisplayu\nu0du′e(iq(1−ζ)+H(u))(u−u′)\n≃1\nH+iq(1−ζ)/parenleftBig\nh′\nij(q,u)−e−iqζ(u−u0)h′\nij(q,u0)/parenrightBig\n.\n(77)\nSince the characteristic frequencies are large compared\nwith 1/(u−u0) the phase factor in the final term will\naverage to zero inside the qandζintegrals in Eq. ( 76).\nWe find then\n∂E\n∂u=¯ρ\n4/integraldisplayd3q\n(2π)3|h′\nij(q,u)|2\n×/integraldisplay1\n−1dζ\n2H(1−ζ2)2\nH2+q2(1−ζ)2.(78)\nWe see here how expansion of the universe introduces\na spread in frequencies ∼ ±Haboutq, thus allowing\nLandau damping; in the absence of expansion, H= 0,\nand Landau damping vanishes.\nTo lowest order in H/qthe integral is simply 4/3, so\nthat\n∂E\n∂u=¯ρ\n3a′\na/integraldisplayd3q\n(2π)3|h′\nij(q,u)|2\nq2. (79)\nThe energy density of the gravitational wave is\nEgw=/integraldisplayd3q\n(2π)3|h′\nij(q,t)|2\n32πG, (80)\nso that for a wavepacket centered about a frequency ¯ q\n∂E\n∂u=32πG¯ρ\n3¯q2HEgw. (81)Finally we note that 8 πG¯ρ/3∼(a′/a2)2and thus\n∂E\n∂u∼4H3\n(a¯q)2Egw. (82)\nThe characteristic absorption time via Landau damping\nis thus∼ω2t3\nU(withω= ¯q), which is thoroughly negligi-\nble. The corresponding fractional change in energy over\nan expansion time of the universe is ∼1/(ωtU)2.\nVIII. CONCLUDING REMARKS\nIn this paper we have laid out a framework for eval-\nuating the damping of gravitational radiation by matter\nwith arbitrary mass particles and collision strengths. By\nconsidering the damping of gravitational waves in both\nflat spacetime and in an expanding universe, we identify\ntwo distinct mechanisms through with damping can oc-\ncur – the first in which collisions produce the damping,\nand the second, via Landau damping.\nWhen the expansion of spacetime can be neglected,\nthe damping of a wave of a given frequency, propor-\ntionaltothe relaxation rate 1/τinthecollisionlessregime\n(ωτ≫1) and to the collision time ,τ, in the hydro-\ndynamic regime ( ωτ≫1), is maximal when ωτ≃1.\nFor the frequencies to which LIGO is sensitive and for\nplausible models of dark matter, calculations of damping\nbased on hydrodynamical considerations are gross over-\nestimates, and we conclude that it is impossible from\ncurrent observations of gravitational waves to put use-\nful bounds on the properties of dark matter. Landau\ndamping in this case is not possible because particles\nhave velocities less than c. As we estimate in Sec. V,\ncollisional damping of gravitational waves of frequencies\nproduced by astrophysical binary systems, propagating\nthrough dense local environments, is also insignificant.\nCollisionless damping is possible in an expanding uni-\nverse since the frequency of the gravitational wave and\nthe energies of particles depend on time. Damping of\ngravitational waves by free-streaming relativistic parti-\ncles, proposed by Weinberg [ 10], may, as we have shown,\nbe regarded as a generalization of Landau damping; we\nhave also generalized Weinberg’s formalism to allow for\ncollisions in the matter. We note in passing that one can\nstraightforwardlyextend the present framework to incor-\nporate scenarios of non-thermal dark matter, since it was\nnot essential to assume a specific functional form for the\ndistribution, see, e.g., Eq. ( 20).\nIn the future we will apply the present framework to\nstudy the processing of stochastic gravitational waves of\nprimordial origin, e.g., from (first order) phase transi-\ntions in the early universe during matter domination, in\nscenarios of ultralight dark matter. Such scenarios are\nsimilar to damping by neutrinos during radiation dom-\nination [10,11], and might describe damping by axions\n[18–20].9\nAcknowledgments\nWe are grateful to Stu Shapiro and Subir Sarkar\nfor very helpful remarks. The research of author GB\nwas supported in part by NSF Grant PHY1305891 and\nPHY1714042. He is grateful to the Aspen Center forPhysics, supported in part by NSF Grants PHY1066292\nand PHY1607611, and the Niels Bohr International\nAcademy where parts of this research were carried out.\nAuthor SP is supported by funds from DanmarksGrund-\nforskningsfond under Grant No. 1041811001.\n[1] B. P. Abbott et al. (LIGO Scientific Collab. and Virgo\nCollab.) Phys. Rev. Lett. 116, 061102 (2016); ibid. 116,\n241103 (2016); arXiv:1706.01812.\n[2] S. W. Hawking, Astrophys. J. 145, 544 (1966).\n[3] S. Weinberg, Gravitation and Cosmology: Principles and\nApplications of the General Theory of Relativity , (Wiley,\nNY, 1972). Ch. X.\n[4] G. Goswami, G. K. Chakravarty, S. Mohanty, and\nA. R. Prasanna, Phys. Rev. D 95, 103509 (2017).\n[5] A. P. Lightman, W. H. Press, R. H. Price, and S. A.\nTeukolsky, Problem Book in Relativity and Gravitation\n(Princeton Univ. Press, 1975), Problem 18.15.\n[6] S.GayerandC. F.Kennel, Phys.Rev.D 19, 1070 (1979).\n[7] A. A. Abrikosov and I. M. Khalatnikov, Rep. Prog. Phys.\n22, 329 (1959), Eq. (10.1).\n[8] D. Lynden-Bell, MNRAS 124, 279 (1962).\n[9] G. Baym, H. Monien, C. J. Pethick and D. G. Ravenhall,\nPhys. Rev. Letters 64, 1867 (1990).\n[10] S. Weinberg, Phys. Rev D 69, 023503 (2004).\n[11] T. Watanabe and E. Komatsu, Phys. Rev. D 73, 123515(2006).\n[12] B. A. Stefanek and W. W. Repko, Phys. Rev. D 88,\n083536 (2013).\n[13] K. Nishijima and M. H. Saffouri, Phys. Rev. Lett. 14,\n205 (1965).\n[14] M. Garny, M. Sandora and M. S. Sloth, Phys. Rev. Lett.\n116, 101302 (2016)\n[15] A. W. Graham, D. Merritt, B. Moore, J. Diemand and\nB. Terzic, Astron. J. 132, 2685 (2006).\n[16] V. Paschalidis, M. Ruiz and S. L. Shapiro, Astrophys. J.\nLetters806, L14:1-5, (2015).\n[17] M. Ruiz, R. Lang, V. Paschalids and S. L. Shapiro, As-\ntrophys. J. Letters 824, L1:1-5 (2016).\n[18] L. Hui, J. P. Ostriker, S. Tremaine, and E. Witten, Phys.\nRev. D95, 043541 (2017).\n[19] N. Banik and P. Sikivie, in Universal Themes of Bose-\nEinstein Condensation, (eds. D. Snoke, N. Proukakis and\nP. Littlewood, Cambridge Univ. Press, 2017).\n[20] D. J. E. Marsh, Phys. Repts. 6431-79 (2016)." }, { "title": "1805.08022v1.Critical_damping_in_nonviscously_damped_linear_systems.pdf", "content": "arXiv:1805.08022v1 [math.DS] 21 May 2018Critical damping in nonviscously damped linear systems\nMario L´ azaro1\n1Dep. of Continuum Mechanics and Theory of Structures. Universit ` at Politecnica de\nVal` encia. 46022, Valencia, Spain\nAugust 25, 2021\nAbstract\nIn structural dynamics, energy dissipative mechanisms with non-v iscous damping are character-\nized by their dependence on the time-history of the response veloc ity, mathematically represented\nby convolution integrals involving hereditary functions. Combination of damping parameters in\nthe dissipative model can lead the system to be overdamped in some ( or all) modes. In the domain\nof the damping parameters, the thresholds between induced oscilla tory and non–oscillatory motion\nare called critical damping surfaces (or manifolds, since we can have a lot of parameters). In this\npaper a general method to obtain critical damping surfaces for no nviscously damped systems is\nproposed. The approach is based on transforming the algebraic eq uations which defined implicitly\nthe critical curves into a system of differential equations. The der ivations are validated with three\nnumerical methods covering single and multiple degree of freedom sy stems.\n1 Introduction\nIn this paper, nonviscously damped linear systems are under consideration. Nonviscous (also named\nby viscoelastic) materials have been widely used for vibrat ing control in mechanical, aerospace, au-\ntomotive and civil engineering applications. This paper de als precisely with those applications where\nvibrations are tried to be disappeared, that is, designing o f damping devices which are able to avoid\noscillatory motion at dynamical systems. In nonviscous mod els, damping forces are assumed to be\ndependent on the history of the response velocity via kernel time functions. As far as the motion\nequations concerned, this fact is represented by convoluti on integrals involving the velocities of the\ndegrees–of–freedom (dof) and affected by the hereditary kern els. Denoting by u(t)∈Rnto the array\nwith the degrees of freedom of the system, this vector verifie s the dynamicequilibrium equations which\nin turn has an integro-differential form\nM¨u+/integraldisplayt\n−∞G(t−τ)˙udτ+Ku=f(t) (1)\nwhereM,K∈Rn×nare the mass and stiffness matrices assumed to be positive defin ite and positive\nsemidefinite, respectively; G(t)∈Rn×nis thenonviscous dampingmatrix in the time domain, assumed\nsymmetric, which satisfies the necessary conditions of Goll a and Hughes [1] for a strictly dissipative\nbehavior. As known, the viscous damping is just a particular case of Eq. (1) with G(t)≡Cδ(t),\nwhereCis the viscous damping matrix and δ(t) the Dirac’s delta function. The time–domain system\nof motion equations are then reduced to the well known expres sions\nM¨u+C˙u+Ku=f(t) (2)\nConsidering now the free motion case f(t)≡0in Eq. (1), we test exponential solutions of the type\nu(t) =uest, withuandsto be found. Then, the classical nonlinear eigenvalue probl em associated\nto viscoelastic vibrating structures yields\n/bracketleftbig\ns2M+sG(s)+K/bracketrightbig\nu≡D(s)u=0 (3)\niwhereG(s) =L{G(t)} ∈CN×Nis the damping matrix in the Laplace domain and D(s) is the dy-\nnamical stiffness matrix or transcendental matrix.\nResponse of Eqs. (1) is closely related to the eigensolution s of the eigenvalue problem (3). Ad-\nhikari [2] derived modal relationships and closed form expr essions for the transfer function in the\nLaplace domain. Due to the non–linearity, induced by a frequ ency-dependent damping matrix, the\nsearch eigensolutions is in general much more computationa lly expensive than that of classical viscous\ndamping [3]. A survey of the different viscoelastic models can be found on the references [4, 5] al-\nthough as far as this paper concerned we work with hereditary damping models based on exponential\nkernels [6].\nSince the present paper is devoted on critical damping and th is field has been deeply studied in\nthe bibliography for viscously damped systems, we consider relevant to review the main works on\nit. Duffin [7] defined an overdamped system in terms of the quadr atic forms of the coefficient ma-\ntrices. Nicholson [8] obtained eigenvalue bounds for free v ibration of damped linear systems. Based\non these bounds, sufficient condition for subcritical dampin g were derived. M¨ uller [9] characterized\nan underdamped system in similar terms to Duffin’s work derivi ng a sufficient condition expressed\nas function of the definiteness of the system matrices. Inman and Andry [10] proposed sufficient\nconditions for underdamped, overdamped and critically dam ped motions in terms of the definiteness\nof the system matrices. These conditions are valid for class ically damped systems although Inman\nand Andry shown that they also could work for non–classical s ystems. Inman and Orabi [11] and\nGray and Andry [12] proposed more efficient method for computi ng the critical damping condition.\nHowever, Barkwell and Lancaster [13] pointed out some deffici encies in the Inman and Andry criterion\nof ref. [10] presenting a counterexample and they provided s ome reasons explaining why this crite-\nrion had been usually adopted to check criticality in damped systems. Additionally, Barkwell and\nLancaster [13] obtained necessary and sufficient conditions for overdamping in gyroscopic vibrating\nsytems. Bhaskar [14] presented a more complete overdamping condition which somehow corrected\nthat of Inman and Andry [10] giving a generalization to a clas s of non–conservative systems. Beskos\nand Boley [15] established conditions for finding critical d amping surfaces from the determinant of\nthe system and its derivative. They proved that a critically damped eigenvalue was simultaneously\nroot of the characteristic equation and its derivative, som ething that can be used to detect critical\ndamping surfaces. In the work [16] the same authors studied c onditions for critical damping in con-\ntinuous systems. Beskou and Beskos [17] presented an approx imate method computationally efficient\nto find critical damping surfaces separating overdamping or partially overdamping regions from those\nof underdamping for viscously damped systems.\nAs far as nonviscous systems concerned, research on critica l damping has not been as exhaustive as\nthat ofviscousdamping. Mainly, investigations havebeenf ocusedonsingledegree–of–freedom systems\non the discussion of the type of response attending at the dam ping parameters of a single–exponential\nhereditary kernel. Muravyov and Hutton [18] and Adhikari [1 9] analyzed the conditions under which\nsingle degree–of–freedom nonviscously damped systems by o ne exponential kernel becomes critically\ndamped. He carried out an exhaustive analysis of the roots na ture of the resulting third order char-\nacteristic polynomial. Adhikari [20] studied the dynamic r esponse of nonviscously damped oscillators\nand discussed the effect of the damping parameters on the frequ ency response function. M¨ uller [21]\nperformed a detailed analysis on the nature of the eigenmoti ons of a single degree of freedom Zener\n3–parameter viscoelastic model. Muravyov [18] obtained cl osed–form solutions for forced nonviscoulsy\ndamped beams studying the conditions for overdamping or und erdamping time response. As known,\nviscoelastic systems modeled by hereditary exponential fu nctions are characterized by having extra\nreal overdamped modes associated to those kernels. As far as these type of modes concerned the\nreferences [22, 23, 24] provide a mathematical characteriz ation and some numerical methods to their\nevaluation.L´ azaro [5] observed that certain recursive me thod to obtain eigenvalues in proportionally\ndamped viscoelastic systems always converges under a linea r rate except just in the critical surfaces\nwhere the scheme is underlineal.\niiIn this paper, critical damping surfaces of nonviscously da mped linear systems are presented.\nCritical damping is refereed to the set of damping parameter s within the threshold between induced\noscillatory and non–oscillatory motion (for all or for some modes). The general procedure to extract\nthese manifolds in the domain of the damping parameters is to eliminate a parameter of a system of\ntwo algebraical equations. Encouraged by the fact that this elimination is not possible for polynomials\nwith order greater than four, a new method to construct criti cal curves is developed. This method is\nbased on to transform the algebraical equations into two ord inary differential equations. The method\nis validated with three numerical examples. The two first are devoted on single degree–of–freedom\nsystems with one and two hereditary exponential kernesl, re spectively. The application of the current\napproach for multiple degree–of–freedom sytems is present ed in the third example.\n2 Conditions of criticality in terms of determinant of the sy stem\nIn this section we will extend the main results derived by Bes kos and Boley [15] on critical viscous\ndamping to nonviscously damped systems. In order to establi sh the basis of our work, we will describe\nthe type of damping model adopted in its most general form. We will consider a damping matrix\nbased on hereditary Biot’s exponential kernels. Mathemati cally, this model adopts the following form\nin time and in frequency domain\nG(t) =N/summationdisplay\nk=1Ckµke−µkt,G(s) =L{G(t)}=N/summationdisplay\nk=1µk\ns+µkCk (4)\nwhereµk>0, 1≤k≤Nrepresent the relaxation or also called nonviscous coefficie nts andCk∈Rn×n\nare the (symmetric) matrices of the limit viscous damping mo del, defined as the limit\nN/summationdisplay\nk=1Ck= lim\nµ1...µN→∞G(s) (5)\nCoefficients µkcontrol the time and frequency dependence of the damping mod el while the spatial\nlocation andthelevelofdampingarecontrolledbycoefficien tswithinmatrices Ck. Itisstraightforward\nthat the following relationships hold\nN/summationdisplay\nk=1Ck=/integraldisplay∞\n0G(t)dt=G(0) (6)\nHenceforth, we will consider the damping matrix, and for ext ension the transcendental matrix, as\ndepending not only on the frequency via s, but also on set of parameters controlling the dissipative\nbehavior. In the most general case, let us see that the symmet ric damping model presented in (4)\ndepends on pmax=N+Nn(n+1)/2 parameters. Indeed, Nnonviscous coefficients µ1,...,µ Nplus\nn(n+1)/2 possible independent entrees in every symmetric matrix Ck, with 1 ≤k≤N. Thus, the\ncomplete set of parameters can be listed as\nµ1,...,µ N,C111,...,C 1nn,...,C N11,...,C Nnn (7)\nwhereCkij=Ckjiis the entree ijof matrix Ck. Real applications depend in general on much less\nparameters, say p << p max. In the sake of clarity, we will denote by θ={θ1,...,θp}the set of\nindependent damping parameters and consequently we can wri te the damping matrix as G(s,θ).\nAccording to the said above, we can denote as D(s,θ) = det[D(s)] the determinant associated to\nthe nonlinear eigenvalue problem (3). Eigenvalues are then roots of the equation\nD(s,θ) = det/bracketleftBigg\ns2M+sN/summationdisplay\nk=1µk\ns+µkCk+K/bracketrightBigg\n= 0 (8)\niiiAttending to the values of θ∈Rp, the algebraic structure of the spectrum of this problem can vary.\nIf the level of damping induced by matrix G(s,θ) is light, we will have 2 ncomplex eigenvalues with\noscillatory nature and rreal eigenvalues with non–oscillatory nature and associat ed to the nonviscous\nhereditary kernels (hence they are also usually named as non viscous eigenvalues). Furthermore, the\ntotal number of these real eigenvalues is [3]\nr=r1+···+rN=N/summationdisplay\nj=1rank(Cj) (9)\nAs longas thereexist 2 ncomplex eigenvalues and rnonviscous eigenvalues, wewill say that thesystem\nis completely underdamped. As thedampinglevel increases, the real part of eigenvalues (not necessary\nall) becomes higher (in absolute value) and the imaginary pa rt decreases. For certain value of the\ndamping parameters a conjugate–complex pair could merge in to a double real negative root. The set\nof damping parameters is said then to be on a critical surface , which in turn represents the threshold\nbetween underdamping and overdamping. If oscillatory mode s coexist with those non–oscillatory,\nthen we say the system is partially overdamped (or mixed over damping). The system is said to be\ncompletely overdamped if all modes are so. For mixed or compl ete overdamping, some (or maybe all)\nof the roots of (8) are negative real numbers, say s=λ, withλ <0 so that\nD(λ,θ) = det/bracketleftbig\nλ2M+λG(λ,θ)+K/bracketrightbig\n= 0 (10)\nFor each value of λ, Eq. (10) defines a p–dimensional surface in the space where the parameters arra y\nθcan take values. Since for light damping we have initially npairs of conjugate–complex eigenvalues,\nwe will have as much as ncritical surfaces because, as Beskos and Boley [15] point ou t: “there are at\nmost as many partial critical damping possibilities as the n umber of the pairs in (8) of roots swith\nzero imaginary part”. The mathematical principle which cha racterizes a critical damping surfaces can\nbe extrapolated to non–vicous damping and therefore these o nes can be found imposing a minimum\namong all possible values of λin Eq. (10), that is\n∂\n∂λD(λ,θ) =∂\n∂λdet[D(λ,θ)] = 0 (11)\nThis condition is consistent with the fact that a critical ro ot is double just under critical condition.\nTherefore, Eqs. (8) and (11) define a set of critical surfaces resulting after eliminating parameter λ\nfrom both equations. This process, although well defined fro m a theoretical point of view, can only be\ncarried out if an analytical closed–form of D(λ,θ) is provided something that only occurs for small to\nmoderately sized systems. For nonvisocusly damped systems , this procedure has not been used yet,\nto the authors knowledge. Instead, they have been found for s ingle degree–of–freedom systems and\nforN= 1 kernels since this particular problem leads to a three ord er polynomial, which as known\nallows radicals–based analytical solution (Cardano’s for mulas). Additionally, in the present paper we\nalso attempt to improve the numerical evaluation of critica l damping surfaces proposing a numerical\nmethod which will be described in the following paragraphs.\nNumerical evaluation of critical surfaces consists in solv ing Eqs. (8) and (11) simultaneously for\na prefixed range of values of damping parameters. This proces s becomes in general computationally\ninefficient sincefor each value oftheprefixedparameters, as ystem of two non–linear equations must be\nsolved. Attempting to improve the numerical procedure for c onstructing critical surfaces we propose\na method valid to build critical curves formed by two paramet ers, assuming as fixed the rest of them.\nThe method is able to find critical overdamped regions in two d imensional cross sections of the p–\ndimensional real critical manifolds. Thus, from the comple te set of parameters θ={θ1,...,θp}, we\nchose two of them, which will be named as design parameters. W ithout loss of generality, we can take\nθ1andθ2while the rest of parameters remain fixed, say θ30,...,θp0. The challenge is to draw the\ncritical damping curves in the plane ( θ1,θ2). For a sake of clarity in the notation, we will denote by\np=θ1andq=θ2and will assume then that the critical curve(s) are function s of the form q=q(p).\nFor each value of p, both equations\nD(λ,p,q) = 0,∂\n∂λD(λ,p,q) = 0 (12)\nivallow to find a pair ( λ,q) (or several, since λis within a polynomial). Let us consider a point p0for\nwhichq0andλ0are solutions of Eqs. (12) and let us assume around p=p0the functions q(p) andλ(p)\nexist. The three numbers ( p0,q0,λ0) form a initial point of the proposed approach. The derivati ves\nλ′(p) = dλ/dpandq′(p) == dq/dpcan be evaluated just applying the chain rule in Eqs. (12). In deed,\nD,λλ′(p)+D,qq′(p)+D,p= 0 (13)\nD,λλλ′(p)+D,λqq′(p)+D,λp= 0 (14)\nwhere subscripts denoting partial derivatives. Since from Eq. (12), we have D,λ= 0, then we can solve\nforλ′(p) andq′(p)\nq′(p) =−D,p\nD,q\nλ′(p) =D,pD,λq\nD,qD,λλ−D,λp\nD,λλ(15)\nThese two equations form a system of two ordinary differential equations whose solutions are be well\ndefinedprovided that thederivatives D,λλandD,qdonot vanishat ( p0,q0,λ0). Existence of thecritical\ncurve beyond a close interval around the initial point will b e subordinate to the existence of those\nderivatives along the curve. Critical curves arise now as th e numerical solution of a system of ordinary\ndifferential equations, forwhichRunge–Kuttabasedmethods canbeused. Before, themethodrequires\nsolving the system of two algebraical equations (12) and two unkowns, say λ0, q0, which in general\nresults in several solutions because of the polynomial form . Pairs ( λ0,q0) both reals and verifying\nλ0<0 andq0≥0 (we will assume a positive range for parameters) will be app ropriate solutions lying\non a critical surface. Taking derivatives repeatedly respe ctpin (14) also lead us to obtain higher order\nderivatives, allowing to find the Taylor expansion of the cri tical curve around p=p0. This procedure\nis used to find an approximation of a critical curve in the one– kernel single–dof numerical example.\n3 Numerical examples\n3.1 Single degree of freedom systems, N= 1exponential kernel PSfrag replacements\nF(t) F(t)R(t)\nG(t)m mku(t)\nFigure 1: A single degree–of–freedom viscoelastic oscilla tor\nFirst, wewillconsiderthesingledofnonviscous systemwit honehereditaryexponential kernel. The\ndof represents the displacement of certain mas mattached to ground by the viscoelastic constraint.\nFig. 1 shows the schematic configuration mass–spring–viscoelastic damper and the corresponding free\nbody diagram. Hence, the internal force is related to the dis placement by\nR(t) =/integraldisplayt\n−∞G(t−τ)˙u(τ)dτ+ku(t) (16)\nkis the constant of the linear–elastic spring and G(t) is the dissipative kernel or damping function\nwith the general form, both in time and frequency domain\nG(t) =cµe−µt, G(s) =µc\ns+µ(17)\nvwhereµandcare respectively is the nonviscous and the viscous coefficien ts. The free motion equation\ncan be deduced from the dynamic equilibrium for F(t)≡0\nm¨u+/integraldisplayt\n−∞G(t−τ)˙u(τ)dτ+ku(t) = 0 (18)\nAnd the associated characteristic equation\nms2+sG(s)+k=ms2+sµc\ns+µ+k= 0 (19)\nIt is quite appropriate to board this problem using dimensio nless variables in order to compare with\nexisting results presented in the bibliography. Thus, we de fine the following non–dimensional variables\nx=s\nωn, ν=ωn\nµ, ζ=c\n2mωn(20)\nwhereωn=/radicalbig\nk/mis he natural frequency of the undamped system. After straig ht operations and\nmultiplying Eq. (19) bythe denominator of thedampingfunct ionwe obtain the characteristic equation\nin non–dimensional form as the third order polynomial\nD(x,ν,ζ) = (1+ νx)(x2+1)+2xζ=νx3+x2+(ν+2ζ)x+1 = 0 (21)\nAs known, the three roots are available as function of the coe fficients so that a detailed discussion of\nthe nature of the three roots can be addressed as function of t he values of ζ >0 andν >0. This work\nwas carried out by Adhikari in the references [19, 20] where c losed form expressions of the critical\ncurves enclosing the overdamped region were derived. For th e sake of our exposition we consider\ninteresting transcript here the Adhikari’s results of the c ritical curves, since later we will present also\napproximations of them. Thus, the overdamped region can be d efined as the set\n/braceleftbig\n(ζ,ν)∈R+:ζL(ν)≤ζ≤ζU(ν)/bracerightbig\n(22)\nwhere the critical damping curves ζL(ν), ζU(ν) are\nζL(ν) =1\n24ν/bracketleftbigg\n1−12ν2+2/radicalbig\n1+216ν2+cos/parenleftbigg4π+θ\n3/parenrightbigg/bracketrightbigg\nζU(ν) =1\n24ν/bracketleftbigg\n1−12ν2+2/radicalbig\n1+216ν2+cos/parenleftbiggθ\n3/parenrightbigg/bracketrightbigg\n(23)\nwith\nθ= arccos/bracketleftBigg\n−5832ν4+540ν2−1\n(1+216ν2)3/2/bracketrightBigg\n(24)\nLet us apply the proposed method to find critical damping curv es based on the solution of the\nsystem of differential equations (15). According to the theor etical derivations, the critical surfaces\narises from eliminating xfrom the two following equations\nD(x,ν,ζ) =νx3+x2+(ν+2ζ)x+1 = 0\n∂D\n∂x= 3νx2+2x+ν+2ζ= 0 (25)\nFrom the second equation we can obtain the two roots x1,2= (−1±/radicalbig\n−6ζν−3ν2+1)/3νand then\nplug them into the first one. After some simplifications, we ob tain the critical surfaces in implicit form\n8ζ3ν+12ζ2ν2−ζ2+6ζν3−10ζν+ν4+2ν2+1 = 0 (26)\nwhich coincides with the third order polynomial obtained by Adhikari [19].\nviWe will attempt now to obtain curves of the form ζ=ζ(ν), therefore our independent parameter\nisp=νand the dependent variable is q=ζ. We need to find the partial derivatives of D(x,ν,ζ) and\nD,x(x,ν,ζ) respect to x,ζandν, obtaining\nD,ν=x+x3,D,ζ= 2x\nD,xν= 1+3 x2,D,xζ= 2,D,xx= 2+6νx (27)\nAfter some math, the two differential equations are set as\nζ′(ν) =−2\n1+x2\nx′(ν) =2x2\n(1+x2)(1+3νx)(28)\nThese equations must becompleted with initial conditions. Takingζ0= 1 equations (25) can besolved\nobtaining four pairs of roots\n(x=−1, ν= 0) ; ( x=−3.38298, ν= 0.134884) ; ( x= 0.191±0.508i, ν=−3.067±2.327i)\n(29)\nOnly real solutions with x <0, ν≥0 are of interest as initial values. The first pair results\nin the initial values ν0= 0,ζ0= 1,x0=−1 of the critical curve ζL(ν) while the second pair\nν0= 0.134884,ζ0= 1,x0=−3.38298 gives as a result ζU(ν). Both curves have been plotted in\nFig. (2). Results fit perfectly with those of exact solutions of Eqs. (23). Although from an analytical\n0.000.050.100.150.200.25\n0.0 0 .5 1 .0 1 .5 2 .0 2 .5 3 .0 3 .5 4 .0\nViscous damping factor, ζNon–viscous damping factor, ν\nζL\n≈\n1−\nν−\nν2/2ζU≈1\n8ν−ν\n2xL≈ −1−ν−5ν2/2\nxU≈ −1\n2νOverdamped region\nExact critical curves\nProposed approximations\nFigure2: Example1: singledegree–of–freedom with N= 1exponential kerkel. Exactandapproximate\n(proposed) overdamped region. xLandxUare the overdamped critical eigenvalues along the critical\n(approximate) curves\npoint of view, the problem of determining critical curves in this case is solved, we will go further\nproposing two closed–form approximations of ζL(ν) andζU(ν). We consider these derivations of in-\nterest on one hand, by their simplicity respect to those of th e exact expressions. And, on the other\nhand, by the procedure to deduce them.\nEncouragedby thefact that theinitial point of thecritical curveζL(ν) isas simpleas ζL(0) = 1and\nalso by its regularity and low curvature (information alrea dy known since the exact result is available\nin Fig. 2, we think that the Taylor series expansion around ν0= 0 can provide accurate results and\nviiin turn simple in form. Indeed, the first derivative can be det ermined just from Eq. (28) for x0=−1\nandν0= 0, resulting\nζ′\nL(0) =−1, x′(0) =−1\nNow, taking again derivatives respect to νin Eqs. (27) and after some operations, second derivatives\nζ′′\nL(0) andx′′(0) can be found, so that\nζ′′\nL(0) =−1, x′′(0) =−5\nHence, Taylor series expansions up to the second order of the critical damping curve ζL(ν), and its\nassociated critical eigenvalue xL(ν) are then\nζL(ν)≈1−ν−ν2/2 (30)\nxL(ν)≈ −1−ν−5ν2/2 (31)\nSimilar procedure could be followed to find a Taylor based app roximation around upper critical\ncurveζU, however, this function presents higher changes of curvatu res and a wider domain of ζ(in fact\nan infinite range). In is expected that a polynomial based app roximation only will work around the\ninitial point and of course it will not be able to represent th e asymptotic behavior. To undertake this\napproach a recent result on asymptotic behavior of polynomi al roots proposed by L´ azaro et al. [25]\nwill be used. Given a polynomial\na0+a1X+···+an−2Xn−2+an−1Xn−1+Xn(32)\nthen the numbers (called polynomial pivots)\n−an−1\n2±/radicalbigg/parenleftBigan−1\n2/parenrightBig2\n−an−2 (33)\npresent the property of lying close to one (or two) roots prov ided that they are not relatively smaller\nthan the rest of the polynomial coefficients. The exact mathem atical conditions describing this state-\nment are given in form of several theorems in the reference [2 5]. We check if the so defined pivots\nof the third order polynomial D(x,ν,ζ) can give us valuable information respect to the nature of th e\nroots. Thus, the pivots are\n−a2\n2±/radicalbigg/parenleftBiga2\n2/parenrightBig2\n−a1=−1\n2ν±/radicalbigg\n1\n4ν2−2ζ\nν−1 (34)\nLooking for critical damping curves, we know that along them , the roots have double multiplicity\n(double roots). Therefore, if we admint that the pivots are c lose to two roots of the problem and we\nforce the discriminant of Eq. (34) to be zero, then the obtain ed root will be double and we will lie on\na critical damping curve. Hence, to vanish the discriminant results in the approximation of the upper\ncurveζU(ν). Indeed,\n1\n4ν2−2ζ\nν−1 = 0→ζU(ν)≈1\n8ν−ν\n2(35)\nThe associated damped eigenvalue can be approximated by\nxU(ν)≈ −1\n2ν(36)\nAccording to the results of [25] the bigger the pivots the clo ser to the roots of the polynomial. There-\nfore, we can predict that the lower the nonviscous parameter νthe more accurate the results, as can\nbe appreciated in the Fig. 2.\nviii3.2 Single degree of freedom systems, N= 2exponential kernels\nNow we will attempt to find the critical damping surfaces of a s ingle dof system with N= 2 hereditary\nkernels. Theapproachcaneasilybeextrapolatedtothegene ralcaseof Nkernels. AccordingtoEq.(4),\nthe damping function is\nG(t) =c1µ1e−µ1t+c2µ2e−µ2t, G(s) =L{G(t)}=µ1c1\ns+µ1+µ2c2\ns+µ2(37)\nAnd the characteristic equation yields\nms2+sG(s)+k=ms2+s/parenleftbiggµ1c1\ns+µ1+µ2c2\ns+µ2/parenrightbigg\n+k= 0 (38)\nAs noticed, the dissipative model has four parameters, c1,c2,µ1,µ2. Our procedure allows to draw\ncritical curves of two parameters, hence before finding the s olution of the proposed differential equa-\ntions two parameters must be fixed. For a sake of the solution r epresentation, we fusion the damping\ncoefficients into only one, considering the case c1=c2=c. Let us define the following dimensionless\nparameters\nx=s\nωn, ν1=ωn\nµ1, ν2=ωn\nµ2, ζ=c\nmωn(39)\nWhereωn=/radicalbig\nk/mis the natural frequency of the system. The Eq. (38) can be exp ressed now in\ndimensionless form\nx2+xζ/parenleftbigg1\n1+ν1x+1\n1+ν2x/parenrightbigg\n+1 = 0 (40)\nNote that ν1=ν2= 0 yields the particular case of viscous damping leading to ζcr= 1. Multiplying\nEq. (40) by (1+ ν1x)(1+ν2x) we transform the characteristic equation into a four order polynomial\nequation\nD(x,ν1,ν2,ζ) = (1+ ν1x)(1+ν2x)(x2+1)+xζ(2+ν1x+ν2x) = 0 (41)\nwhich together with\nD,x(x,ν1,ν2,ζ) = 2ζ+2x+2xν1ν2(1+2x3)+(1+2 xζ+3x3)(ν1+ν2) (42)\nallow to find the critical surfaces after eliminating x. Observe that the last equation is a third order\npolynomial, therefore the Cardano formulas can be used to ob tain the three roots. Plugging them into\nEq. (41) would lead the exact critical curves. These derivat ions will not be carried out here since the\nresulting expressions would be hardly handled and together with the difficulty to follow properly the\nexposition. On the other hand, they can easily be programmed in a symbolic software and results be\ncompared to those of the present method.\nAccording to the proposed approach, the critical surfaces a re defined in terms of two parameters,\nleaving fixed the rest. For the current example we will consid er critical curves in the plane ( ζ,ν2) (i.e.\nfunctions ν2=f(ζ) withν1fixed) and also in the plane ( ν1,ν2) (i.e. functions ν2=f(ν1) withζfixed).\nTwo systems of differential equations must be assembled, one i nvolving the functions x(ζ),ν2(ζ) and\nthe other one the functions x(ν1),ν2(ν1). From Eq. (14) the two problems can be written in matrix\nform as\n•Critical curves in plane ( ζ,ν2). Parameter ν1fixed.\n/bracketleftbigg0D,ν2\nD,xxD,xν2/bracketrightbigg/braceleftbiggx′(ζ)\nν′\n2(ζ)/bracerightbigg\n=−/braceleftbiggD,ζ\nD,xζ/bracerightbigg\n,/braceleftbiggx(ζ0)\nν2(ζ0)/bracerightbigg\n=/braceleftbiggx0\nν20/bracerightbigg\n(43)\n•Critical curves in plane ( ν1,ν2). Parameter ζfixed.\n/bracketleftbigg0D,ν2\nD,xxD,xν2/bracketrightbigg/braceleftbiggx′(ν1)\nν′\n2(ν1)/bracerightbigg\n=−/braceleftbiggD,ν1\nD,xν1/bracerightbigg\n,/braceleftbiggx(ν10)\nν2(ν10)/bracerightbigg\n=/braceleftbiggx0\nν20/bracerightbigg\n(44)\nixwhere\nD,ν1=x/bracketleftbig\n1+x(x+ζ+ν2+x2ν2)/bracketrightbig\nD,ν2=x/bracketleftbig\n1+x(x+ζ+ν1+x2ν1)/bracketrightbig\nD,ζ=x[2+x(ν1+ν2)]\nD,xν1= 1+x[2(ζ+ν2)+x(3+4xν2)]D,xν2= 1+x[2(ζ+ν1)+x(3+4xν1)]D,xζ= 2[1+x(ν1+ν2)]\nThe initial conditions come from solving Eqs. (41) and (42) f or prescribed values of two of the pa-\nrameters. Table 3.2 lists a complete set of initial values wh ich allow to address the solution of the\ndifferential equations. From the solution of the aforementio ned algebraic equations one can find simul-\ntaneously two different initial conditions. Thus, for instan ce, for the case ν10= 0.00 andζ0= 1.00,\namong other complex solutions we find ( ν20,x0) = (0,−1) and (ν20,x0) = (0.1916,−2.7693) as the\ninitial conditions of the two first curves shown in Table 3.2. Notice that this fact also takes place for\nother cases in the table, always if we somewhat are fortunate in our election of the prescribed pair of\nparameters. Otherwise, we could not obtain satisfactory so lutions, for example taking ν10= 0.25 and\nζ0= 5.00 we only find complex solutions to the system (41) and (42).\nINITIAL VALUES\nValue of the fixed parameter Curve ζ0ν20 x0\nν1: fixed ν1= 0.00 C1 1.00000 0.00000 -1.00000\nC2 1.00000 0.19160 -2.76929\nC3 4.00000 1.80565 -2.22076\nν1= 0.05 C1 0.95000 0.04736 -1.05573\nC2 0.95000 0.19197 -2.72450\nC3 5.20000 0.79903 -4.84245\nC4 5.20000 1.93559 -9.57166\nν1= 1.50 C1 4.00000 0.03207 -2.61688\nC2 4.00000 0.06738 -6.74685\nValue of the fixed parameter ν10 ν20 x0\nζ: fixed ζ= 0.90 C1 0.00000 0.17053 -1.14479\nC2 0.00000 0.22155 -2.26820\nζ= 5.00 C1 0.00000 0.03443 -16.97040\nC2 0.00000 1.32591 -2.83001\nC3 0.05200 2.06935 -9.20408\nC4 1.20000 0.01599 -3.10458\nC5 1.20000 0.05344 -8.61426\nζ= 8.00 C1 0.00000 0.02148 -27.25100\nC2 0.00000 0.76381 -4.62400\nC3 0.03400 0.56548 -13.0963\nC4 0.70000 0.00952 -5.03922\nC5 0.70000 0.03329 -13.89100\nTable 1: Initial conditions used to the computation of the cr itical damping curves shown in Fig. 3\nxIn Fig. 3, the different obtained critical curves have been plo tted. If we read overdamped regions\nas certain volumes enclosed by critical surfaces in the para metric space ( ζ,ν1,ν2), then the left curves\nare cross sections of these volumes for certain values of the damping ratios (in Fig. 3 left cases\nζ= 0.9,5.0,8.0 are shown). On the other hand, right plots have the same inte rpretation but as\ncross sections of the planes ν1= 0.0,0.054,1.50. The critical curves are in correspondence with\nthe notation used in Table 3.2. As expected overdamped regio n in the plane ( ν1,ν2) are symmetric\nrespect to line ν1=ν2since the physical model has inherently this symmetry. In th e plane ( ν2,ζ)\nforν1= 0 we observe an overdamped region for values ν2≪1 very similar to that obtained in the\nfirst example, see Fig. 2. Now, due to the commented symmetry o f the problem respect to ν1and\nν2, this long–triangle–like of the top–right plot is in fact a s ection of a conoid–like form in the space\n(ζ,ν1,ν2). Actually, it is a quarter of conoid because ν1,ν2≥0. Another section of this conoid is\nobtained just shifting the cross section to the value ν1= 0.05 (middle–right plot). It is also interesting\nthe new overdamped region arising in the corner of the right– top plot (section ν1= 0). Presumably\nhigh values of ζandν2(simultaneously) lead the system to non–oscillatory motio n. Let us see that\nthe damper model in this case is formed by two dampers in paral lel, one viscous and the other one\nnonviscous with parameter µ1. Indeed, for ν1= 0 (which is equivalent to µ1→ ∞) the damping\nfunction is transformed into\nlim\nµ1→∞G(t) =c1δ(t)+c2µ2e−µ2tlim\nµ1→∞G(s) =c1+µ2c2\ns+µ2(45)\nSomehow, we can interpret this overdamped region as the effect producedin the nature of the response\nof both the nonviscous parameter µ2and the viscous coefficcient c1. This is the reason because such\ncritical surface did not appear in the example 1. Again, from the symmetry respect to the nonviscous\nparameters, this overdamped region also appears in the the p lane (ζ,ν1) forν2= 0. Furthermore,\nthis effect is extended as an narrow volume in the approximate r ange 0≤ν1≤0.07 (respectively for\nsymmetry in 0 ≤ν2≤0.05), see middle–left and bottom–left plots. It seems clear t hat adding new\ndamping parameters will make more difficult how to read into th e form of the overdamped manifolds,\nspecially since they do not follow regular geometrical stru ctures, as seen in this example. However,\nthe proposed method could be applied sequentially to extrac t those more interesting curves for our\nanalysis, for instance in artificial dampers design problem s. Let us see now in a final example how to\nextract critical damping curves for a multiple dof system.\n3.3 Multiple degree of freedom systems\nm\nk\nmk\nmk\nmk\nG (t) G (t)u1u2u3u4\nFigure 4: Example 3: The four degrees–of–freedom discrete s ystem.G(t) represents the hereditary\nfunction of nonviscous dampers\nIn order to validate the proposed approach to find critical da mping curves for multiple dof systems a\ndiscrete lumped mass dynamical system with four dof is analy zed. The Fig. 4 represents the distri-\nbution of masses m, rigidities kand viscoelastic dampers with a hereditary function G(t). The mass\nmatrix of the system is M=mI4while, according to the rigidities and dampers distributio n, stiffness\nximatrix yields\nK=k\n2−1 0 0\n−1 2−1 0\n0−1 2−1\n0 0−1 2\n=kK (46)\nWe will assume a dampingfunction formed by one hereditary ex ponential kernel of dampingcoefficient\ncand nonviscous parameter µ. Hence, te damping matrix can be expressed as G(t) =µCe−µtwhere\nC=c\n0 0 0 0\n0 1−1 0\n0−1 2−1\n0 0 1 1\n≡cC (47)\nWith help of these dimensionless matrices, say KandC, we can express the non–linear eigenvalue\nproblem associated to this problem under dimensionless for m as\n/bracketleftbigg\nx2I4+2xζ\n1+νxC+K/bracketrightbigg\nu=0 (48)\nwhere\nx=s\nω0, ν=ω0\nµ, ζ=c\n2mω0, ω0=/radicalbig\nk/m (49)\nSincer= rank(C) = 2, the system has 2 n+r= 10 eigenvalues. Therefore, the determinant of the\ntranscendental matrix can be transformed into a 10th order p olynomial multiplying by the factor\n(1+νx)2. We define then our function D(x,ν,ζ) as\nD(x,ν,ζ) = (1+ νx)2det/bracketleftbigg\nx2I4+2xζ\n1+νxC+K/bracketrightbigg\n= 5+2(8 ζ+5ν)x+[(2ζ+ν)(6ζ+5ν)+20]x2+(54ζ+40ν)x3\n+/parenleftbig\n32ζ2+54ζν+20ν2+21/parenrightbig\nx4+(40ζ+42ν)x5+/parenleftbig\n12ζ2+40ζν+21ν2+8/parenrightbig\nx6\n+8(ζ+2ν)x7+/parenleftbig\n1+8νζ+8ν2/parenrightbig\nx8+2νx9+νx10(50)\nWe are looking for overdamped regions enclosed by critical c urves of type ν=ν(ζ), therefore we con-\nstruct our system of differential equations following the met hodology described in Section 2, resulting\n/bracketleftbigg0D,ν\nD,xxD,xν/bracketrightbigg/braceleftbiggx′(ζ)\nν′(ζ)/bracerightbigg\n=−/braceleftbiggD,ζ\nD,xζ/bracerightbigg\n,/braceleftbiggx(ζ0)\nν(ζ0)/bracerightbigg\n=/braceleftbiggx0\nν0/bracerightbigg\n(51)\nFor a sake of clarity in the exposition the expressions of the partial derivatives will not be written.\nInitial conditions can be found solving the system of algebr aical equations for a particular value of ν0\nandζ0\nD(x0,ν0,ζ0) = 0,D,x(x0,ν0,ζ0) = 0 (52)\nTesting for ν0= 0.06 we obtain four different pairs ( ζ0,x0), listed in Table 2. After solving Eqs. (51)\nINITIAL VALUES\nCurve ζ0 ν0 x0\nC1 0.53258 0.06000 -2.05512\nC2 0.72949 0.06000 -7.88861\nC3 1.25218 0.06000 -1.45645\nC4 2.14493 0.06000 -8.07994\nTable 2: Initial conditions used for the critical damping cu rves shown in Fig. 5\nwe plot the solutions in Fig. 5. The four found curves enclose two overdamped regions which in turn\nxiiintersect each other. The solid–filled region represents th e set of values ζ,νwhich lead the fourth\nmode to overdamping. On the other hand, lines–filled area cor responds to the overdamped region\nof the second mode. This can be checked following a root–locu s plot varying parameters ζ,νfrom\nundamping to overdamping. Moreover, 2nd and 4th mode are pre cisely those modes for which degrees\nof freedom linked to the viscoelastic dampers are most activ ated. Obviously, it follows then that the\noverlapping zone (with both types of shading in Fig. 5) corre sponds with the overdamping of both\nmodes, simultaneously.\n0.000.050.100.150.20\n0.0 0 .5 1 .0 1 .5 2 .0 2 .5 3 .0 3 .5 4 .0 4 .5 5 .0\nViscous damping factor, ζNon–viscous damping factor, νOverdamped Region (4th mode)\nOverdamped Region (2nd mode)\nCurve C1\nCurve C2\nCurve C3\nCurve C4\nInitial values\nFigure 5: Example 3: Critical damping curves and the corresp onding enclosing overdamped regions\nfor the four degrees–of–freedom system\nA deeper inspection of Fig. 5 leads us to ask ourselves about t he existence of singularities in the\ndomain ( ζ,ν) which can impede the application of our approach. Well, it i s known that a system\nof differential equations like that shown in (51) has solution (and it is unique) provided that the\ndeterminantofthematrixdoesnotvanishinaneighborhooda roundtheinitialvalue. Thisdeterminant\nis equal to −D,νD,xx, whence it follows that these two system of algebraic equati ons\nS1:\n\nD= 0\nD,x= 0\nD,ν= 0S2:\n\nD= 0\nD,x= 0\nD,xx= 0(53)\nallow us to find singularities. For every solution of the first system S1, some of the variables, x,ζor\nνis within the complex plain. On the other hand, we do find valid solutions of system S2, verifying\nx <0, ζ,ν≥0, say\nx=−3.1059 ζ= 0.46473 ν= 0.10658\nx=−2.3221 ζ= 1.06113 ν= 0.14088 (54)\nThese two points are precisely the vertexes of the two overda mped regions, namely, intersection points\nof curves C1–C2 and C3–C4, respectively. These points can no t be used as initial points since ac-\ncording to the implicit function theorem, equations D= 0 and D,x= 0 do not define x(ζ) andν(ζ)\nunequivocally. Furthermore, both points verifies D,xx= 0, hence they are triple roots.\nWe wonder now how does the solution behave in the intersectio n point between curves C2 and C3,\nlocated approximately at ζ= 1.30465, ν= 0.03239. This point satisfies −D,νD,xx/ne}ationslash= 0, therefore it\nshould be valid as initial value of our method. However, it be longs to both curves simultaneously, so\nxiiithat at a first sight the solution would not seem to be well defin ed. However, in this point we find\ntwo solutions for the variable x, sayx=−1.38042 and x=−15.2148, which leads to two different\ninitial values. Somehow this point does not correspond with a intersection point of the curves in the\n3D domain ( x,ν,ζ), while the tripleroots of (54) do.\nAs far as multiple degree-of-freedom systems concern, the s uccess of the method lies on the avail-\nability of the transcendental matrix determinant and their derivatives. Therefore, from a numerical\npoint of view, large systems will require high computationa l effort which limits the range of applica-\nbility for small or moderate order systems. Currently, our e fforts are addressed to find out numerical\nprocedures allowing to construct approximate critical cur ves but for larger systems, something that is\nunder research.\n4 Conclusions\nIn this paper critical damping of nonviscously damped linea r systems is studied. Nonviscous or vis-\ncoelastic vibrating structures are characterized by dissi pative mechanisms depending on the history\nof response through hereditary functions. For certain valu es of the damping parameters, the response\ncan become non–oscillatory. It is said then that some (or all ) modes are overdamped. Particular\nvalues of the damping parameters which establish the limit b etween oscillatory and non–oscillatory\nmotion are said to be within a critical surface (or critical m anifold). In the present paper a general\nprocedure to build critical damping surfaces is developed. In addition, a numerical method based on\nthe transformation of the algebraical equations into a syst em of two ordinary differential equations is\nproposed. This approach allow to find critical curves of two p arameters for certain fixed values of the\nrest of parameters.\nTo validate the theoretical results three numerical exampl es are analyzed. In the first example,\nthe well–known overdamped region of a single degree–of–fre edom system with one exponential kernel\nis resolved, shown perfect fitting between our curves, obtai ned from the differential equations, and\nthose of the analytical expressions. In order to give added v alue to this problem, we propose sim-\nplified approximate expressions for the critical curves. Th e second example is devoted to construct\noverdamped regions for a two exponential kernels based damp ing function. This problem involves\nthree parameters, so that the critical damping curves are pl otted along several cross sections defined\nby the free parameter. The third example shows how the method can be applied for multiple degrees\nof freedoms systems deriving overdamping regions for differe nt modes and interpreting the obtained\noverlapping regions. Since this method is based on the evalu ation and derivation of the determinant\nof the transcendental matrix, its range of validity is reduc ed to small or moderately sized systems.\nEncouraged by this limitation, the author is currently inve stigating how to extrapolate this method\nfor larger systems.\nReferences\n[1] D. Golla, P. Hughes, Dynamics of Viscoelastic Structure s - A Time-domain, Finite-element For-\nmulation, Journal of Applied Mechanics-Transactions of th e ASME 52 (4) (1985) 897–906.\n[2] S. Adhikari, Dynamics of Non-viscously Damped Linear Sy stems, Journal of Engineering Me-\nchanics 128 (3) (2002) 328–339.\n[3] N. Wagner, S. Adhikari, Symmetric state-space method fo r a class of nonviscously damped sys-\ntems, AIAA Journal 41 (5) (2003) 951–956.\n[4] S. Adhikari, A Reduced Second-Order Approach for Linear Viscoelastic Oscillators, Journal of\nApplied Mechanics-Transactions of the ASME 77 (4) (2010) 1– 8.\nxiv[5] M. L´ azaro, J. L. P´ erez-Aparicio, M. Epstein, Computat ion of eigenvalues in proportionally\ndampedviscoelastic structuresbased onthefixed-pointite ration, AppliedMathematics andCom-\nputation 219 (8) (2012) 3511–3529.\n[6] M. Biot, Variational Principles in Irreversible Thermo dynamics with Application to Viscoelastic-\nity, Physical Review 97 (6) (1955) 1463–1469.\n[7] R. Duffin, A MINIMAX THEORY FOR OVERDAMPED NETWORKS, JOURN AL OF RA-\nTIONAL MECHANICS AND ANALYSIS 4 (2) (1955) 221–233.\n[8] D. Nicholson, EIGENVALUE BOUNDS FOR DAMPED LINEAR-SYST EMS, MECHANICS\nRESEARCH COMMUNICATIONS 5 (3) (1978) 147–152.\n[9] P. Muller, OSCILLATORY DAMPED LINEAR-SYSTEMS, MECHANI CS RESEARCH COM-\nMUNICATIONS 6 (2) (1979) 81–85.\n[10] D. Inman, A. Andry, SOME RESULTS ON THE NATURE OF EIGENVA LUES OF DISCRETE\nDAMPEDLINEAR-SYSTEMS,JOURNALOFAPPLIEDMECHANICS-TRAN SACTIONSOF\nTHE ASME 47 (4) (1980) 927–930.\n[11] D. Inman, I. Orabi, AN EFFICIENT METHOD FOR COMPUTING TH E CRITICAL DAMP-\nING CONDITION,JOURNALOFAPPLIEDMECHANICS-TRANSACTIONS OFTHEASME\n50 (3) (1983) 679–682.\n[12] J. Gray, A. Andry, A SIMPLE CALCULATION FOR THE CRITICAL DAMPING MATRIX\nOF A LINEAR MULTIDEGREE OF FREEDOM SYSTEM, MECHANICS RESEAR CH COM-\nMUNICATIONS 9 (6) (1982) 379–380.\n[13] L. Barkwell, P. Lancaster, Overdamped and Gyroscopic V ibrating Systems, Journal of Applied\nMechanics-Transactions of The ASME 59 (1) (1992) 176–181.\n[14] A. Bhaskar, Criticality of damping in multi-degree-of -freedom systems, Journal of Applied Me-\nchanics, Transactions ASME 64 (2) (1997) 387–393, cited By 7 .\n[15] D. BESKOS, B. BOLEY, CRITICAL DAMPING IN LINEAR DISCRET E DYNAMIC-\nSYSTEMS, JOURNAL OF APPLIED MECHANICS-TRANSACTIONS OF THE ASME 47 (3)\n(1980) 627–630.\n[16] D. BESKOS, B. BOLEY, CRITICAL DAMPING IN CERTAIN LINEAR CONTINUOUS\nDYNAMIC-SYSTEMS, INTERNATIONAL JOURNAL OF SOLIDS AND STRU CTURES\n17 (6) (1981) 575–588.\n[17] S. Papargyri-Beskou, D. Beskos, On critical viscous da mping determination in linear discrete\ndynamic systems, ACTA MECHANICA 153 (1-2) (2002) 33–45.\n[18] A. Muravyov, S. Hutton, Free vibration response charac teristics of a simple elasto-hereditary\nsystem, Journal of Vibration and Acoustics-Transactions o f the ASME 120 (2) (1998) 628–632.\n[19] S.Adhikari, Qualitative dynamiccharacteristics ofa non-viscouslydampedoscillator, Proceedings\nof the Royal Society A-Mathematical Physical and Engineeri ng Sciences 461 (2059) (2005) 2269–\n2288.\n[20] S. Adhikari, Dynamic response characteristics of a non viscously damped oscillator, Journal of\nApplied Mechanics-Transactions of the ASME 75 (1) (2008) 01 1003.01–011003.12.\n[21] P. Muller, Are the eigensolutions of a l-d.o.f. system w ith viscoelastic dampingoscillatory or not?,\nJournal of Sound and Vibration 285 (1-2) (2005) 501–509.\nxv[22] M. L´ azaro, J. L. P´ erez-Aparicio, Characterization o f real eigenvalues in linear viscoelastic oscilla-\ntors and the non-viscous set, Journal of Applied Mechanics ( Transactions of ASME) 81 (2) (2014)\nArt. 021016–(14pp).\n[23] M. L´ azaro, Nonviscous modes of nonproportionally dam ped viscoelastic systems, Journal of Ap-\nplied Mechanics (Transactions of ASME) 82 (12) (2015) Art. 1 21011 (9 pp).\n[24] M. L´ azaro, C. F. Casanova, C. L´ azaro, Nonviscous Mode s of Viscoelastically Damped Vibrating\nSystems, InTech, 2016, Ch. Viscoelastic and Viscoplastic M aterials, pp. 165–187.\n[25] M. L´ azaro, P. Mart´ ın, A. Ag¨ uero, I. Ferrer, The polyn omial pivots as initial values for a new\nroot-finding iterative method, Journal of Applied Mathemat ics Vol. 2014 (Special Issue: Iterative\nMethods and Applications) (2015) Article ID 413816, 14 page s.\nxviOverdamped regions for ζ≡const Overdamped regions for ν1≡const\n0.000.250.500.751.001.251.500.00 0.25 0.50 0.75 1.00 1.25 1.50Non–viscous parameter, ν2Cross section\nζ= 0.90\nCurve C1\nCurve C2\n0.000.250.500.751.001.251.500.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0\nCross section\nν1= 0.00\nCurve C1\nCurve C2\nCurve C3\n0.000.250.500.751.001.251.50Non–viscous parameter, ν2Cross section\nζ= 5.00\nCurve C1\nCurve C2\nCurve C3\nCurve C4\nCurve C5\n0.000.250.500.751.001.251.50\nCross section\nν1= 0.05\nCurve C1\nCurve C2\nCurve C3\nCurve C4\n0.000.250.500.751.001.251.50\n0.00 0.25 0.50 0.75 1.00 1.25 1.50\nNon–viscous parameter, ν1Non–viscous parameter, ν2Cross section\nζ= 8.00\nCurve C1\nCurve C2\nCurve C3\nCurve C4\nCurve C5\n0.000.250.500.751.001.251.50\n0.0 1.0 2.0 3.0 4.0 5.0 6.0 7.0 8.0\nViscous damping factor, ζCross section\nν1= 1.50\nCurve C1\nCurve C2\nFigure 3: Example 2: Single degree–of freedom nonviscous sy stem with N= 2 kernels. Critical\ndamping curves and overdamped regions\nxvii" }, { "title": "1206.1695v2.A_realisation_of_Lorentz_algebra_in_Lorentz_violating_theory.pdf", "content": "arXiv:1206.1695v2 [hep-th] 14 Nov 2012A realisation of Lorentz algebra in Lorentz violating theor y\nOindrila Ganguly∗\nS. N. Bose National Centre for Basic Sciences, Kolkata 70009 8, India\nJune 16, 2018\nAbstract\nA Lorentz non-invariant higher derivative effective action in flat spacetime, characterised by a\nconstant vector, can be made invariant under infinitesimal L orentz transformations by restricting\nthe allowed field configurations. These restricted fields are defined as functions of the background\nvector in such a way that background dependance of the dynami cs of the physical system is no longer\nmanifest. We show here that they also provide a field basis for the realisation of Lorentz algebra and\nallow the construction of a Poincar´ e invariant symplectic two form on the covariant phase space of\nthe theory.\n1 Introduction\nAny departure from exact Lorentz symmetry is expected to be a t elltale footprint of quantum gravity,\nthe theory of physics beyond the Planck scale (denoted by Planck m assMPl). This possibility is tremen-\ndouslyexcitingowingtothefactthatquantumgravityeffectsfallla rgelyoutsidethedomainofourcurrent\nexperiments and observations. Moreover, it is always challenging to find the limits of validity of any sym-\nmetry. These factorshavestimulated a lot of workon the theoret ical, phenomenologicaland experimental\naspects of Lorentz violation since the last couple of decades [1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14].\nIn the present report, we shall focus on a low energy effective field theory of scalar, vector and spinor\nfields developed by Myers and Pospelov [10] that incorporates devia tion from Lorentz symmetry through\nthe inclusion of an extra term. This new term contains a Planck suppr essed dimension five operator with\na constant background vector nthat essentially assigns a preferred direction to spacetime. Howev er, in\n[15], the authors demonstrate that there exist quite general field configurations which can restore Lorentz\nsymmetryin thelimit ofinfinitesimal transformationsin modified action of[10]. TheseLorentzpreserving\nfields, as they have been named in [15], locally conserve the N¨ other current for Lorentz transformations.\nNotice though that the Lorentz violation that had been woven into t he Myers Pospelov theory is not\nlost. It manifests itself through the dependance of the Lorentz p reserving fields on the directionality of\nthe background i.e. on n. Given this scenario, a natural question to ask is whether Lorentz Lie algebra\ncan be realised on a basis of such special field configurations.\nWe shall consider the case of the Lorentz preserving scalar fields [1 5] for the modified action of [10] as\nan illustrative example. The extended action of a complex scalar field φproposed by Myers and Pospelov\nis,\nSMPφ=/integraldisplay\nd4xLMPφ,\n=/integraldisplay\nd4x/bracketleftbig\n|∂φ|2−m2|φ|2/bracketrightbig\n+/integraldisplay\nd4xiκ\nMPlφ∗∂3\nnφ . (1)\n∗oindrila@bose.res.in\n1ThefirstintegralcontainstheusualLagrangiandensityofacomp lexscalarfield withmass mwhilethe\nterm under the second integral is the Lorentz violating contributio n.κis a real, dimensionless parameter\nandn·∂≡∂n,nbeing a constant four vector. According to [15], this action shall be symmetric under\ninfinitesimal Lorentz boosts or rotations if the complex scalar field c an be split as,\nφ(x) =φ/bardbl(x/bardbl)+φ⊥(x⊥) (2)\nwhere,φ/bardblandφ⊥are arbitrary functions of their respective arguments x/bardblandx⊥defined by n.\nx/bardbl≡x·n\nn2nandn·x⊥= 0, so that x=x/bardbl+x⊥. So, the derivative operator can be written as\n∂=∂x/bardbl+∂x⊥=∂/bardbl+∂⊥where the notation ∂/bardbl≡∂x/bardbland∂⊥≡∂x⊥.\nA key aspect of the additional term in (1) is that it contains third ord er derivatives of the field unlike\nstandard first order Lagrangians. But higher derivative Lagrang ians are not new to physics [16]. Back\nin 1961, Ostrogradskii [17] had developed a canonical formalism for dealing with them. Reviews and\nmodifications of his technique [18, 19, 20, 21] study mostly systems having finite number of degrees of\nfreedom and higher time derivatives of the generalised coordinates . An extension to special relativistic\ncontinuous systems is presented in [22], though it relies on the exist ence of Lorentz symmetry. In the\nnext section, we shall briefly review Ostrogradskii’s original constr uction and its generalisation to field\nsystems, tailored to suit the Lagrangian (1) with fields that decoup le as (2)\nUsually, the canonical formalism is understood to be non-covariant because it involves the choice of a\nspacelikehypersurfaceand its orthogonaltime direction. In fact , all the peculiaritiesofahigherderivative\ntheory are due only to higher order time derivatives. Spatial deriva tives are quite benign, staying within\nthe scope of standard first order canonical approach. So, in sec tion 3, the study of the modified scalar\nfield theory will be split up into two cases distinguished by n2being timelike and spacelike. We will not\ncomment on n2= 0 in this paper. We shall also take the liberty of working in certain Lor entz frames\nthat simplify calculations.\nHowever, a covariant framework for the canonical formulation of a relativistic theory may also be\ndeveloped [23, 24, 25, 26, 27] through the construction of a cova riant phase space and a symplectic two-\nform on it. Reference [28] has presented a natural extension of t his technique to higher derivative field\ntheories. Section 4 will contain a summary of the main features of th e covariant phase space for first\norder as well as higher order derivative field systems. We shall conc lude by showing how the Lorentz\npreserving fields facilitate the construction of a covariant phase s pace structure for the Lorentz violating\neffective field theory under consideration.\n2 Canonical formalism in the presence of higher derivatives\n2.1 Non-relativistic systems with finite degrees of freedom\nLet us consider a system described by the Lagrangian L(qa,dtqa,...,dl\ntqa) which is a function not only of\nthe generalised coordinate qa(t) and the velocity dtqa(t) but of all derivatives dtqa(t),d2\ntqa(t),...,dl\ntqa(t)\nupto order l. Herealabels the different degrees of fredom and we adopt the notation qa(j)≡dj\ntqa, j=\n0,...,lso thatqa(j+1)= ˙qa(j). Each of these qa(j)uptoj=l−1 will have a conjugate momentum\npa(j). Remember though that the superscript jofpa(j)only denotes that it is conjugate to qa(j).\nThe system is now specified by a point in the phase space spanned by qa(j),pa(j);j= 0,...,l−1. We\nhave here assumed that the highest derivative qa(l)can be written as a function of the other variables\nqa(l)(qa(0),pa(0);...;qa(l−1),pa(l−1)). The condition for extremisation of the action S[q(t)] =/integraltext\ndt Lis,\nδqS= 0 =/integraldisplay\ndtl/summationdisplay\nj=0(−dt)j∂L\n∂qa(j)δqa+/integraldisplay\ndtdtl/summationdisplay\ni=j+1l−1/summationdisplay\nj=0(−dt)i−(j+1)∂L\n∂qa(i)δqa(j).\nIf the boundaries are such that the variations of qaand its derivatives upto order l−1 vanish, the\nsecond integral will go to zero and the equation of motion will be\n2(−dt)j∂L\n∂qa(j)= 0. (3)\nOn the other hand, if this system undergoes a symmetry transfor mation,δS= 0 and substitution of\nequation of motion (3) yields,\ndt\nl/summationdisplay\ni=j+1l−1/summationdisplay\nj=0(−dt)i−(j+1)∂L\n∂qa(i)δqa(j)\n= 0.\nThe quantity within the square brackets is the conserved N¨ other currentJ. From its structure, we\nmay read off the conjugate momenta\np(j)\na≡l/summationdisplay\ni=j+1(−dt)i−(j+1)∂L\n∂qa(i), j= 0,...,l−1. (4)\nThe N¨ other current may then be cast into the standard form J=l−1/summationtext\nj=0p(j)\naδqa(j).\n2.2 Relativistic continuous systems\nHere, in place of the generalised coordinates, we shall be working wit h special relativistic fields hav-\ning infinite number of degrees of freedom. Let us consider a system of scalar fields φa(x). If\nL(φa,φa,ρ1,φa,ρ1ρ2,...,φa,ρ1...ρl) be the Lagrangian density (where φa,ρ1...ρj≡∂ρ1...∂ρjφa), then by ex-\ntremising the action S[φ(x)] =/integraltext\nd4xLwith respect to the fields φa(x), we get the equation of motion\nl/summationdisplay\nj=0(−1)j∂ρ1...∂ρj∂L\n∂φa,ρ1...ρj= 0. (5)\nThe necessary boundary conditions are that the fields and their de rivatives upto φρ1...ρl−1fall off at\ninfinity. It would be convenient if we could write down a general form o f the N¨ other current for such\nhigher derivative systems. This is achieved by applying a symmetry tr ansformation to the fields and\nconsequently employing the equation of motion (5) to get the N¨ oth er current,\nJρ1=l/summationdisplay\ni=j+1l−1/summationdisplay\nj=0(−1)i−(j+1)∂ρj+2...∂ρi∂L\n∂φa,ρ1...ρiδφρ2...ρj+1, (6)\nwhich is locally conserved i.e. ∂ρ1Jρ1= 0. It may so happen that instead of δLbeing zero, the\nLagrangian density varies by a total derivative. This would then con tribute to the N¨ other current.\nHowever, as we have already mentioned, a canonical formalism of re lativistic field theories requires\nus to foliate spacetime into spacelike hypersurfaces. This in turn inv olves the separation of temporal and\nspatial derivatives of the fields1. It is most often possible to arrange terms in the Lagrangian such t hat\nmixed derivatives of the fields as in ∂2\nt∂kφdo not survive. This is true not only when the Lagrangian is\nLorentz invariant [22] but also when the Lorentz violating action (1) is written in terms of the Lorentz\n1At this stage, it is imperative that we sort the indices. Lati n letters from the middle of the alphabet set viz. i,j,k,l are\nbeing used as summation indices while those from the end like r,s,...,z will denote spatial components. Different fields w ill\nbe labelled by the alphabets a,b,c,d. Greek letters are rese rved for spacetime indices.\n3preserving fields, as will be illustrated in the next section. In such sit uations, the canonical momenta will\nbe given by\nπa(j)≡l/summationdisplay\ni=j+1(−dt)i−(j+1)∂L\n∂φa(i), j= 0,...,l−1. (7)\nThe canonical variables will satisfy the Poisson bracket\n{φa(i)(t,/vector x),πb(j)(t,/vector x′)}=δb\naδj\niδ(3)(/vector x−/vector x′). (8)\n3 Myers Pospelov theory\n3.1 Constant timelike background vector\nWithout loss of generality, we are going to work in a Lorentz frame de fined by n= (1,/vector0). Then the\nLorentz preserving fields φ/bardbl(t),φ⊥(/vector x) become spatially homogeneous and static, spatially inhomogeneous\nrespectively. This greatly simplifies the Lagrangian density:\nLMPφ=˙φ∗\n/bardbl˙φ/bardbl−/vector∇φ∗\n⊥·/vector∇φ⊥+iκ\nMPl(φ∗\n/bardbl+φ∗\n⊥)...φ/bardbl, (9)\n=LMPφ(φ/bardbl,˙φ/bardbl,¨φ/bardbl,...φ/bardbl,φ∗\n/bardbl,˙φ∗\n/bardbl,φ⊥,/vector∇φ⊥,φ∗\n⊥,/vector∇φ∗\n⊥). (10)\nHere, we have neglected the masses of the fields as we are interest ed in behaviour of the system at\nenergiesmuch higher than the field masses. It is now evident why our chosenLorentz frame is particularly\nuseful. Eq.(9) has only higher order time derivatives of φ/bardbl(t) while/vector∇φ/bardbl= 0 =∂tφ⊥. Thus all mixed\nderivatives in the sense described above will vanish. This permits us t o safely use eq.(7) to determine the\ncanonical momenta. We list them in the following table.\nGeneralised Generalised\ncoordinate momentum\nφ/bardbl(0)=φ/bardblπ/bardbl(0)=˙φ∗\n/bardbl+iκ\nMPl¨φ∗\n/bardbl\nφ/bardbl(1)=˙φ/bardblπ/bardbl(1)=−iκ\nMPl˙φ∗\n/bardbl\nφ/bardbl(2)=¨φ/bardblπ/bardbl(2)=iκ\nMPl(φ∗\n/bardbl+φ∗\n⊥)\nφ∗\n/bardbl(0)=φ∗\n/bardblπ/bardbl∗(0)=˙φ/bardbl\nφ⊥(0)=φ⊥π⊥(0)= 0\nφ∗\n⊥(0)=φ∗\n⊥π⊥∗(0)= 0\nTable 1: Canonically conjugate phase space variables\nNote that π⊥(0)= 0 and π⊥∗(0)= 0 are constraint relations that will be imposed weakly. The equal\ntime Poisson Bracket, eq. (8), will hold for the canonical variables.\nUnder an infinitesimal Lorentz transformation, δαβφ=x[α∂β]φand the Lagrangian density being a\nscalar function, also changes δαβL=x[α∂β]L=∂ρ1(x[αδρ1\nβ]L). The Lorentz preserving fields conserve the\nN¨ other current Jµ\nαβ[15]. The N¨ other charge is given by\nQαβ=/integraldisplay\nΣd3σρ1Jρ1\nαβ.\n4Here, Σ is a three dimensional hypersurface. If we orient it orthog onal to the time axis then,\nQαβ=/integraldisplay\nΣd3/vector x/parenleftBig\nπ(0)\n/bardblδαβφ/bardbl(0)+π(1)\n/bardblδαβφ/bardbl(1)+π(2)\n/bardblδαβφ/bardbl(2)+π/bardbl∗δαβφ∗\n/bardbl−x[αδ0\nβ]L/parenrightBig\n,\n=/integraldisplay\nΣd3/vector x/parenleftBig\nπ(j)\naδαβφa(j)−x[αδ0\nβ]L/parenrightBig\n. (11)\nHereφa=φ/bardbl,φ∗\n/bardbl. The other variables do not contribute. From the structure of Qαβwe can deduce\nthat\n{Qαβ(t),φb(k)(t,/vector x)}=−δαβφb(k)(t,/vector x), for φ b=φ/bardbl,φ∗\n/bardbl;\n= 0, for φ b=φ⊥,φ∗\n⊥. (12)\nThe final step is to evaluate the algebra of the charges. A simple calc ulation using eq.,(11), the basic\nPoisson bracket (8) and the relation δαβφ=x[α∂β]φgives,\n{Qαβ(t),Qρσ(t)}=ηασQβρ(t)−ηαρQβσ(t)+ηβρQασ(t)−ηβσQαρ(t). (13)\nEqs. (12), (13) confirm that N¨ other charges defined in terms of the Lorentz preserving fields are\ngenerators of Lorentz transformation and satisfy the standar d Lorentz algebra. Thus, the special field\nconfigurations provide a valid basis for the realisation of Lorentz Lie algebra which determines the local\nstructure of Lorentz group near the identity.\n3.2 Constant spacelike background vector\nNext, we go over to the Lorentz frame where n= (0,/vector1). Then the Lorentz preserving fields will be\nφ/bardbl(/vector x),φ⊥(t) and the Lagrangian will take the form\nLMPφ=˙φ∗\n⊥˙φ⊥−/vector∇φ∗\n/bardbl·/vector∇φ/bardbl+iκ\nMPl(φ∗\n/bardbl+φ∗\n⊥)|/vector∇|3φ/bardbl. (14)\nThis Lagrangian density contains only first order time derivatives of the fields. Thus, no extra phase\nspace variables will be required. The table below lists the pairs of cano nical coordinates and momenta.\nGeneralised Generalised\ncoordinate momentum\nφ/bardblπ/bardbl= 0\nφ∗\n/bardblπ/bardbl∗= 0\nφ⊥ π⊥=˙φ∗\n⊥\nφ∗\n⊥ π⊥∗=˙φ⊥\nTable 2: Canonically conjugate phase space variables for n= (0,/vector1)\nThe fundamental Poisson Bracket is,\n{φa(t,/vector x),πb(t,/vector x′)}=δabδ(3)(/vector x−/vector x′). (15)\n5Integration of the zeroth component of the N¨ other current ov er an appropriately chosen three dimen-\nsional spatial slice normal to the time axis gives the N¨ other charge for Lorentz transformation,\nQαβ=/integraldisplay\nd3/vector x/parenleftBig\nπ⊥δαβφ⊥+π⊥∗δαβφ∗\n⊥−x[αδ0\nβ]L/parenrightBig\n. (16)\nThis is of the same form as the standard N¨ other current obtained for a first order system of scalar\nfields. Hence, the N¨ other charges would generate infinitesimal Lo rentz transformation and satisfy the\nLorentz algebra.\n4 Overview of covariant phase space formulation\nThe covariant construction of phase space keeps intact Poincar´ e invariance of a physical system [23, 24,\n25, 26, 27]. As opposed to the standard decomposition of phase sp ace in the 3+1 framework, the classical\ncovariant phase space Zof a physical theory is defined as the space of classical solutions of the dynamical\nequations of the theory. Functions φ(x), tangent vectors δφ, one forms δφ(x) and exterior derivatives δ\ncan be defined on Zfollowing [25, 26].\nThe phase space Zis naturally endowed with a closed, non-degenerate two-form ˜ ω( ˜ωis non-\ndegenerate provided the one-form ˜ ω(V\n/tildewide) = 0 if and only if V\n/tildewide= 0) called a symplectic structure. It\ncan be written as an integral of some closed, conserved symplectic two-form current ˜ ωµ(note that ˜ ωµis\na two-form in phase space Zand a vector current in spacetime; ∂µ˜ωµ= 0,δ˜ωµ= 0) over a hypersurface\nΣ,\n˜ω=/integraldisplay\nΣdσµ˜ωµ(17)\nHence, the task now is to find a suitable symplectic current, given an y Lagrangian. Owing to our\npresent interest in the covariant description of canonical formalis m for higher derivative theories [28], we\nshall straight away take the general example of a Lagrangian dens ityL(φa,φa,ρ1,φa,ρ1ρ2,...,φa,ρ1...ρl).\nForφabelonging to Zand an arbitrary linear transformation δφa(x) that takes φa(x)→φa(x)+δφa(x)\non the phase space, we have\nδL=∂µl/summationdisplay\ni=j+1l−1/summationdisplay\nj=0(−1)i−(j+1)∂ρj+2...∂ρi∂L\n∂φa,µρ2...ρiδφρ2...ρj+1\n=∂µjµ, (18)\nafter substituting the Euler Lagrange equation of motion (5). jµis interpreted as a pre-symplectic\ncurrent because it is used to define the symplectic current ˜ ωµ=δjµ. From eq. (18) one can see that\nit is obviously closed in phase space and conserved through its depen dance on spacetime. Hence, the\nsymplectic structure\n˜ω=/integraldisplay\nΣdσµ˜ωµ=δ/integraldisplay\nΣdσµjµ(19)\nis not only closed but also exact. Moreover, the local conservation of the symplectic current in\nspacetime guarantees that ˜ ωwill not change with the choice of the surface of integration Σ in (19) and\nin particular, will be Poincar´ e invariant [26].\n65 Symplectic structure with Lorentz preserving fields\nThecovariantversionofthe canonicalformalismismeaningful onlyw hen thephysicaltheoryhasPoincar´ e\ninvariance. The original Myers Pospelov Lorentz violating model doe sn’t meet this requirement. But\nwhen the allowed field configurations are restricted to the Lorentz preserving fields, the dynamics of\nthe theory becomes invariant under infinitesimal Lorentz transfo rmations enabling us to construct its\ncovariant phase space. Only in this context, eq. (18) holds for LMPφ.\nOne must have observed that the presymplectic current (18) and the N¨ other current (6) have identical\nforms. Infact, with theinterpretationof δφa(x) asaone-formonphasespace,the N¨ othercurrentbecomes\nthepre-symplecticcurrentoneform. Itmustalsobestressedth attheentireconstructionofthe symplectic\nstructure (19) follows from (18), the validity of which is ensured by the Lorentz preserving fields. This is\nturn guarantees the existence of a Lorentz invariant (and Poinca r´ e invariant) symplectic stucture on the\ncovariant phase space of Lorentz preserving solutions of the equ ation of motion of Myers Pospelov action\n[15].\nAcknowledgements\nI am grateful to D. Gangopadhyay and P. Majumder for suggestin g the problem and having numerous\ninsightful discussions. I also thank A. V. Sleptsov and P. I. Dunin-B arkowski for pointing out reference\n[29] to us.\nReferences\n[1]S. M. Carroll ,G. B. Field , andR. Jackiw ,Phys.Rev. D41, 1231 (1990).\n[2]J. I. Latorre ,P. Pascual , andR. Tarrach ,Nuclear Physics B 437, 60 (1995).\n[3]D. Colladay andV. A. Kostelecky ,Phys.Rev. D55, 6760 (1997).\n[4]D. Colladay andV. A. Kostelecky ,Phys.Rev. D58, 116002 (1998).\n[5]G. Amelino-Camelia ,J. R. Ellis ,N. Mavromatos ,D. V. Nanopoulos , andS. Sarkar ,\nNature393, 763 (1998).\n[6]S. R. Coleman andS. L. Glashow ,Phys.Rev. D59, 116008 (1999).\n[7]T. Jacobson andD. Mattingly ,Phys.Rev. D64, 024028 (2001).\n[8]D. Colladay andV. A. Kostelecky ,Phys. Lett. B511, 209 (2001).\n[9]S. Sarkar ,Mod.Phys.Lett. A17, 1025 (2002).\n[10]R. C. Myers andM. Pospelov ,Phys.Rev.Lett. 90, 211601 (2003).\n[11]T. Jacobson ,S. Liberati , andD. Mattingly ,Springer Proc.Phys. 98, 83 (2005).\n[12]T. Jacobson ,S. Liberati , andD. Mattingly ,Annals Phys. 321, 150 (2006).\n[13]D. Mattingly ,Living Rev.Rel. 8, 5 (2005).\n[14]L. Maccione ,A. M. Taylor ,D. M. Mattingly , andS. Liberati ,JCAP0904, 022 (2009).\n[15]O. Ganguly ,D. Gangopadhyay , andP. Majumdar ,Europhys.Lett. 96, 61001 (2011).\n[16]B. Podolsky andP. Schwed ,Rev. Mod. Phys. 20, 40 (1948).\n7[17]M. V. Ostrogradskii ,Complete Collected Works , volume 2, Akad. Nauk Ukrain, SSR, Kiev,\n1961, in Russian.\n[18]D. Musicki ,J.Phys.A A11, 39 (1978).\n[19]J. Barcelos-Neto andN. R. Braga ,Acta Phys.Polon. B20, 205 (1989).\n[20]J. Z. Simon ,Phys.Rev. D41, 3720 (1990).\n[21]A. Morozov ,Theor.Math.Phys. 157, 1542 (2008).\n[22]F. de Urries andJ. Julve ,J.Phys.A A31, 6949 (1998).\n[23]G. J. Zuckerman ,Conf. Proc. C8607214 , 259 (1986).\n[24]E. Witten ,Nucl.Phys. B276, 291 (1986).\n[25]C. Crnkovic andE. Witten ,Covariant description of canonical formalism in geometric al theories ,\nchapter 16, pp. 676–684, Cambridge University Press, Cambridge , 1987.\n[26]C. Crnkovic ,Class.Quant.Grav. 5, 1557 (1988).\n[27]A. Ashtekar ,L. Bombelli , andO. Reula ,The covariant phase space of asymptotically flat\ngravitational fields , pp. 417–450, Elsevier Science Publishers B. V., North-Holland, Ams terdam\n(Netherlands), 1991.\n[28]V. Aldaya ,J. Navarro-Salas , andM. Navarro ,Phys.Lett. B287, 109 (1992).\n[29]P. Dunin-Barkowski andA. Sleptsov ,Theor.Math.Phys. 158, 61 (2009).\n8" }, { "title": "1507.04509v1.Lorentz_Dispersion_Law_from_classical_Hydrogen_electron_orbits_in_AC_electric_field_via_geometric_algebra.pdf", "content": "Lorentz Dispersion Law from classical Hydrogen electron orbits\nin AC electric \feld via geometric algebra\nUzziel Perez\nNational Institute of Physics, University of the Philippines,\nDiliman, Quezon City, Philippines\nAngeleene S. Ang\nAteneo de Manila University, Department of Physics,\nLoyola Heights,Quezon City, Philippines 1108\u0003\nQuirino M. Sugon, Jr. and Daniel J. McNamara\nManila Observatory, Upper Atmosphere Division,\nAteneo de Manila University Campusy\nAkimasa Yoshikawa\nDepartment of Earth and Planetary Sciences, Faculty of Sciences,\nKyushu University, Fukuoka, Japan\n(Dated: July 2, 2018)\nWe studied the orbit of an electron revolving around an in\fnitely massive nucleus of a large\nclassical Hydrogen atom subject to an AC electric \feld oscillating perpendicular to the electron's\ncircular orbit. Using perturbation theory in geometric algebra, we show that the equation of motion\nof the electron perpendicular to the unperturbed orbital plane satis\fes a forced simple harmonic\noscillator equation found in Lorentz dispersion law in Optics. We show that even though we did\nnot introduce a damping term, the initial orbital position and velocity of the electron results to a\nsolution whose absorbed energies are \fnite at the dominant resonant frequency !=!0; the electron\nslowly increases its amplitude of oscillation until it becomes ionized. We computed the average power\nabsorbed by the electron both at the perturbing frequency and at the electron's orbital frequency.\nWe graphed the trace of the angular momentum vector at di\u000berent frequencies. We showed that at\ndi\u000berent perturbing frequencies, the angular momentum vector traces epicyclical patterns.\nPACS numbers: 45.10.Hj, 45.10.Na, 37.10.Vz\nI. INTRODUCTION\nIn standard optics texts, the position xof an electron\nof chargeqand massmunder the time-varying electric\n\feldE=Eei!tof light is given by[1{3]\nx+ \u0000 _x+!2\n0x=q\n2mE; (1)\nwhere \u0000 is the damping coe\u000ecient, !0is the natural fre-\nquency of oscillation of the electron. The complex solu-\ntion ~xto this equation is shown by Akhmanov and Nikitin\n[4] to be\n~x=q\nm1\n!2\n0\u0000!2+i!\u0000E; (2)\nso that the power absorbed by the atom is\nhPi=hqE_xi=q2\n2m!2\u0000\n(!2\n0\u0000!2)2+!2\u00002jEj2;(3)\n\u0003Electronic address: angeleene.ang@gmail.com\nyAlso at Ateneo de Manila University, Department of Physics,\nLoyola Heights,Quezon City, Philippines 1108which in complex space becomes\nP=e\n4i!\u0000\nE\u0003~x+E~xe2i!t\u0001\n: (4)\nNotice that the damping term \u0000 makes the power ab-\nsorbed \fnite at the resonant frequency !=!0.\nAt present it is still not clear why an atom can be\ndescribed as a simple harmonic oscillator subject to si-\nnusoidal electric \feld of light, e.g. what is responsible\nfor the restoring force constant k=m!2\n0and what is the\ncause of damping force \u0000 _ x?\nIn this paper we wish to show that a forced harmonic\noscillator equation can arise for a large Hydrogen atom\nwith a circular orbital of frequency !0subject to a lin-\nearly polarized light of frequency !whose corresponding\nwavelength 2 \u0019c=! is much larger than the electron's or-\nbital radius r0, e.g. microwave frequencies, so that the\nphase of light is approximately the same at any point in\nthe orbital path within a certain time period. That is, if\nthe electron is initially in circular orbit in the xy-plane,\nthe position sof the electron along the z-axis is given by\ns+!2\nos=\u0000qE\nmcos (!t); (5)\nas similarly given in Born and Wolf [5]. Even though\nthis equation does not have a damping term, we shallarXiv:1507.04509v1 [physics.atom-ph] 16 Jul 20152\nshow that the energy absorbed at the resonant frequency\n!=!0remains \fnite, provided we take into account the\nposition rof the electron in 3D and use the vector form\nof the electrical energy dissipation expression [1]:\nP=qE\u0001_ r: (6)\nWe shall show that hPiis \fnite at the resonant frequency\n!=!0even though there is no damping.\nIn 1974, Bay\feld and Koch experimentally studied the\nionization of hydrogen atoms under microwave frequen-\ncies [6]. Since then, many tried to study the interaction of\nmicrowave radiation with classical hydrogen atom within\nthe context of Rydberg atoms.[7] Some authors stud-\nied the interaction with circularly polarized light [8{12],\nwhile others such as Leopold [13], Grosfeld and Friedland\n[14] and Neishtadt [15] focused on the linearly polarized\ncase.\nFor Leopold, his Hamiltonian is of the form:\nH(~ r;~ p) =1\n2p2\u0000r\u00001+zFmaxcos!t; (7)\nand the equations are solved using Monte-Carlo tech-\nniques. For Grosfeld and Friedland, their Hamiltonian\nis of the form:\nH=1\n2p2\u0000r\u00001+Z\u0016cos \t; (8)\nwhere the frequency !(t) =d\t=dtand the authors used\naction angle variables. This Hamiltonian is the same one\nused by Neishtadt and Vasiliev, except that the latter\nauthors used Delaunay elements.\nIn our work, we shall not use the Hamiltonian ap-\nproach. Instead, we shall use the force equation\nr=\u0000kq2\nmr\njrj3\u0000\u0015e3q\nmE0cos(!t+'); (9)\nand use linear perturbation theory to simplify the equa-\ntion to a simple harmonic oscillator equation in (5) for a\nmotion perpendicular to the initial circular orbital plane\nof the electron. This method is simpler than those of the\nprevious authors because the solution to the simple har-\nmonic oscillator equation is well-known. Just as in the\noptical dispersion theory, we computed for the average\npower absorption by the atom and showed that it only\ndepends on the z-coordinate as in the standard theory.\nhPi\u001c=1\n\u001cZ\u001c\n0Pdt: (10)\nWe shall show that the orbit of the electrons at integral\nfrequency ratios are similar to De Broglie waves, except\nthat the oscillation is perpendicular to the electron's or-\nbital plane.\nWe shall divide the paper into six sections. Section 1\nis Introduction. In Section 2, we shall discuss the Ge-\nometric Algebra formalism applied to planar rotations.In Section 3, we shall describe the unperturbed circu-\nlar orbit of the electron around the nucleus. After this,\nwe shall introduce an oscillating electric \feld perturba-\ntion and derive the equations of motion of the electron's\noscillation perpendicular to its orbital plane, using the\ngeometric algebra framework in our previous paper on\nCopernican epicyclical orbits[16]. In Section 4, we shall\ncompute the electron's orbital angular momentum and\ndetermine its limiting form at the resonant frequency. In\nSection 5, we shall compute the electrical power dissi-\npation of the electron and plot the results for di\u000berent\nvalues of the ratio between the orbital and light frequen-\ncies. We shall show that the average power, either over\nthe perturbing or orbital period, is approximately simi-\nlar to the standard absorption resonance curve with \fnite\npeak. Finally, we graph the angular momentum of the\nelectron at di\u000berent frequency ratios and show that the\nangular momentum vector traces epicyclical patterns.\nII. GEOMETRIC ALGEBRA\nA. Scalars, Vectors, Bivectors, and Trivectors\nIn Cli\u000bord (Geometric) Algebra Cl3;0, also known as\nthe Pauli Algebra, the product of the three unit vectors\ne1,e2, ande3satis\fes the orthonormality relation [17{19]\nejek+ekej= 2\u000ejk; (11)\nwhere\u000ejkis the Kronecker delta function. In other words,\nthe square of the length of the vectors is equal to one and\nthe product of two perpendicular vectors anticommute.\nLetaandbbe two vectors spanned by e1,e2, and\ne3. We can show that their product satis\fes the Pauli\nidentity[20, 21]\nab=a\u0001b+i(a\u0002b); (12)\nwherei=e1e2e3is the unit trivector which behaves like\nan imaginary scalar that transforms vectors to bivectors.\nThe Pauli identity states that the geometric product of\ntwo vectors is equal to the sum of their scalar dot product\nand their imaginary cross product.\nB. Exponential Function and Rotations\nLetie3\u0012be the product of a bivector ie3=e1e2with\nthe scalar\u0012. Since the square of ie3\u0012is negative, then\nthe exponential of ie3\u0012is given by Euler's theorem\neie3\u0012= cos\u0012+ie3sin\u0012: (13)\nFrom this we can see that\ncos\u0012=1\n2(eie3\u0012+e\u0000ie3\u0012); (14a)\nsin\u0012=1\n2ie3(eie3\u0012\u0000e\u0000ie3\u0012); (14b)3\ne1eie3\u0012e2eie3\u0012e2\ne1\u0012\u0012\nFIG. 1: Rotation of e1ande2about e3counterclockwise by\nan angle\u0012\nwhich are the known exponential de\fnitions of cosine and\nsine functions.\nMultiplying Eq. (13) by e1,e2, and e3, we obtain\ne1eie3\u0012=e1cos\u0012+e2sin\u0012=e\u0000ie3\u0012e1; (15a)\ne2eie3\u0012=e2cos\u0012\u0000e1sin\u0012=e\u0000ie3\u0012e2; (15b)\ne3eie3\u0012=e3cos\u0012+e3ie3sin\u0012=eie3\u0012e3: (15c)\nNotice that e1eie3\u0012is a rotation of e1counterclockwise\nabout e3by an angle \u0012, while e2eie3\u0012is a rotation of e2\ncounterclockwise about the same direction and the same\nangle. Notice, too, that the argument of the exponen-\ntial changes sign when e1ore2trades places with the\nexponential, while e3commutes with the exponential.\nA vector ain 2D can be expressed in both rectangular\nand polar forms:\na=axe1+aye2=ae1eie3\u0012: (16)\nExpanding the exponential using Eq. (15a) and separat-\ning the e1ande2components, we arrive at the standard\ntransformation equations for polar to rectangular coor-\ndinates:\nx=acos\u0012; (17a)\ny=asin\u0012: (17b)\nWe may also factor out e1in Eq. (16) either to the left\nor to the right to get\na=e1^a=e1(x+ie3y) =e1aeie3\u0012; (18a)\na= ^a\u0003e1= (x\u0000ie3y)e1=ae\u0000ie3\u0012e1: (18b)\nFactoring out e1yields the de\fnition of the complex num-\nber ^aand that of its complex conjugate ^ a\u0003:\n^a=ax+ie3ay=aeie3\u0012; (19a)\n^a\u0003=ax\u0000ie3ay=ae\u0000ie3\u0012: (19b)In general, we have the following relations:\ne1^a= ^a\u0003e1; (20a)\ne2^a= ^a\u0003e2; (20b)\ne3^a= ^ae3: (20c)\nThat is, e1ande2both changes the complex number ^ a\nto its conjugate ^ a\u0003after commutation, while e3simply\ncommutes with ^ a[17{19, 22].\nIII. LIGHT-ATOM INTERACTION\nA. Unperturbed Electron Orbit\nClassically, the position rof an electron of mass m\nand charge\u0000qas it revolves around a massive proton of\nchargeqis given by Coulomb's law:\nr=\u0000kq2\nmr\njrj3; (21)\nwherekis the electrostatic force constant. We claim that\na solution to Eq. (21) is given by\nr=r0=e1^r0^ 0; (22)\nwhere\n^r0=r0eie3'0(23a)\n^ 0=eie3!0t(23b)\nare the complex amplitude and the rotation operator,\nrespectively. Substituting these back to Eq. (22), we get\nr0=e1r0eie3(!0t+'0); (24)\nwhich yields\nr0=e1r0cos(!0t+'0) +e2r0sin(!0t+'0);(25)\nafter expanding the exponential and distributing e1.\nEquation (22) states that the electron moving around the\nproton in circular orbit of radius r0with angular velocity\n!0and rotational phase angle '0.\nTo verify that Eq. (22) is indeed a solution to the\nCoulomb's law in Eq. (21), we \frst compute the \frst\nand second time derivatives of Eq. (22):\n_r=e1^r0ie3!o^ 0; (26a)\nr=\u0000e1^r0!2\n0^ 0: (26b)\nNow, substituting Eqs. (22) and (26b) to the Coulomb's\nlaw in Eq. (21), we obtain\n!2\no=kq2\nmr3\n0; (27)\nafter cancelling out e1^r0^ 0. Equation (27) is the familiar\ncircular orbit condition.4\nr0e1eie3('+!t)\nr0e1eie3'\nr0e1'!t\nFIG. 2: Uniform circular motion of an electron with a distance\nr0from the nucleus. The orbital angular frequency is !and\nthe phase angle is '.\n+jqj\n\u0000jqjFcFp\nv0r0\nFIG. 3: The Coulomb force Fcand the perturbing force Fp\non an electron moving with velocity v0and radius r0\nB. Perturbation by an Oscillating Field\nSuppose that the electron is subject not only to the\nCoulomb force\u0000qEcdue to the proton, but also to the\nforce\u0000qEpdue to an oscillating perturbing \feld. The\nequation of motion of the electron then becomes\nmr=\u0000qEc\u0000\u0015qEp; (28)\nwhere\u0015is a perturbation parameter that shall later be\nset equal to unity. More speci\fcally, we write\nr=\u0000kq2\nmr\njrj3\u0000\u0015e3q\nmE0cos(!t+'); (29)\nwhereE0,!, and'are the amplitude, angular frequency,\nand phase of the perturbing electric \feld. Our aim is\nto determine the position rof the electron that satis\fes\nEq. (29).\nTo \fnd the solution to the perturbed orbit equation\nin Eq. (29), we assume that the solution is a sum of the\nelectron's unperturbed circular orbit in Eq. (22) and a\nslight perturbation ^ sperpendicular to this orbit. So we\nwrite\nr=r0+\u0015r1=e1^r0^ 0+\u0015e3^s: (30)The \frst and second time derivatives of rare\n_r=_r0+\u0015_r1=e1^r0ie3!0^ 0+\u0015e3_^s; (31a)\n r= r0+\u0015 r1=\u0000e1^r0!2\no^ 0+\u0015e3^s: (31b)\nEquation (31b) shall take care of the left side of Eq. (29).\nTo expand the right-hand side, we need \frst to take\nthe square of the position vector rin Eq. (30) and retain\nonly the terms up to \frst order in \u0015:\nr2=r2\n0+ 2\u0015(r0\u0001r1): (32)\nSince r0lies on the unperturbed orbital plane of the elec-\ntronxyplane and r1=e3is perpendicular to this plane,\nthenr0\u0001r1= 0, so that Eq. (32) reduces to\nr2=r2\n0=e1^r0^ 0e1^r0^ 0=e1e1^r\u0003\n0^ \u0003\n0^r0^ 0\n= ^r\u0003\n0^r0=r2\n0; (33)\nwhere we used the de\fnitions of ^ r0and ^ 0in Eqs. (23a)\nand (23b). Thus, jrj=r0, so that\nr\njrj3=1\nr3\n0e1^r0^ 0: (34)\nEquation (34) shall take care of the Coulomb term on the\nright side of Eq. (29).\nNow, substituting Eqs. (31b) and (34) back to equation\nof motion in Eq. (29), we obtain\n\u0000e1^r0!2\no^ 0+\u0015e3s=\u0000!2\n0(e1^r0^ 0+\u0015e3^s)\n\u0000\u0015e3q\nmE0cos(!t+');(35)\nwhere we used the circular orbit condition in Eq. (27).\nThe term zeroth order in \u0015cancels out, so we are left\nwith the term \frst order in \u0015. Hence,\n^s+!2\n0^s=\u0000q\nmE0cos(!t+'); (36)\nafter rearranging the terms. Notice that Eq. (36) is a\nsimple harmonic oscillator equation with sinusoidal forc-\ning, which is the standard model for classical light-atom\ninteraction.\nC. Solving the Forced Harmonic Oscillator\nEquation\nLet ^shand ^spbe the homogeneous and particular so-\nlutions of the Eq. (36). That is,\n^s= ^sh+ ^sp; (37)\nand\n^sh+!2\n0^sh= 0; (38)\n^sp+!2\n0^sp=\u0000q\nmE0cos(!t+'): (39)5\nThe solution to the homogenous equation in Eq. (38)\nis a sum of a sines and cosines:\n^sh=ch1cos(!0t) +ch2sin(!0t); (40)\nwherech1andch2are scalar constants that will be de-\ntermined from the boundary conditions. On the other\nhand, the solution to the particular equation in Eq. (39)\nis of the same form as the perturbing \feld:\n^sp=cpcos(!t+'); (41)\nwherecpis a scalar constant. Substituting Eq. (41) back\nto the particular equation in Eq. (39) and solving for cp,\nwe get\ncp=qE0\nm!2\n01\n(\u000b2\u00001); (42)\nwhere\n\u000b=!\n!0(43)\nis the ratio of the perturbing frequency !to the electron's\norbital frequency !0. Hence,\n^sp=q\nm!2\n01\n(\u000b2\u00001)E0cos(\u000b!0t+'): (44)\nAdding the homogenous solution ^ shin Eq. (40) to the\nparticular solution ^ spin Eq. (41) yields the total solution:\n^s=ch1cos(!0t) +ch2sin(!0t)\n+q\nm!2\n01\n(\u000b2\u00001)E0cos(\u000b!0t+'): (45)\nIts time derivative is\n_^s=\u0000ch1!0sin(!0t) +ch2!0cos(!0t)\n\u0000q\nm! 0\u000b\n(\u000b2\u00001)E0sin(\u000b!0t+'): (46)\nTo determine the unknown constants ch1andch2, we\n\frst substitute the expressions for sand _sin Eqs. (45)\nand (46) back to the expressions for the position rand\nvelocity _rin Eqs. (30) and (31a) to get\nr=e1^r0^ 0+e3(ch1cos(!0t) +ch2sin(!0t))\n+e3q\nm!2\n01\n(\u000b2\u00001)E0cos(\u000b!0t+'); (47a)\n_r=e1^r0ie3!0^ 0+e3(\u0000ch1!0sin(!0t) +ch2!0cos(!0t))\n\u0000e3q\nm! 0\u000b\n(\u000b2\u00001)E0sin(\u000b!0t+'); (47b)\nafter setting the perturbation parameter \u0015= 1. If we\nassume that at t= 0, the electron is in its unperturbed\ncircular orbit around the nucleus, then\nr(0) = e1^r0; (48a)\n_r(0) = e1^r0ie3!0: (48b)Substituting these to Eqs. (47a) and (47b), and setting\nt= 0, we arrive at the expressions for the parameters ch1\nandch2:\nch1=\u0000q\nm!2\n01\n(\u000b2\u00001)E0cos'; (49a)\nch2=q\nm!2\n01\n(\u000b2\u00001)E0sin': (49b)\nSubstituting Eqs. (49a) and (49b) back to the expres-\nsion for the position rin Eq. (47a), we get\nr=e1^r0^ 0+e3qE0\nm!2\n01\n(\u000b2\u00001)(\u0000cos'cos(!0t)\n+ sin'sin(!0t) + cos(\u000b!0t+')): (50)\nUsing the identity for the cosine of a sum of two angles,\nEq. (50) reduces to\nr=e1^r0^ 0+e3qE0\nm!2\n01\n(\u000b2\u00001)\u0002\n(cos(\u000b!0t+')\u0000cos(!0t+')): (51)\nIts time derivative is\n_r=e1^r0ie3!0^ 0+e3qE0\nm! 01\n(\u000b2\u00001)\u0002\n(\u0000\u000bsin(\u000b!0t+') + sin(!0t+'); (52)\nwhere we used the de\fnition of \u000b=!=! 0. Equations (51)\nand (52) are the position and velocity of the electron\ninitially orbiting at radius r0, angular frequency !0, and\nphase'0, and perturbed by an oscillating electric \feld\nwith amplitude E0, frequency !, and phase '.\nD. Orbit Equations and Limiting Conditions\nTo convert Eqs. (51) and (52) into rectangular coor-\ndinates, we use the expansions in Eq. (15a) and (15b),\ntogether with the identity e1ie3=e2to arrive at\nx=r0cos(!0t+'0); (53a)\ny=r0sin(!0t+'0); (53b)\nz=qE0\nm!2\n01\n(\u000b2\u00001)(cos(\u000b!0t+')\u0000cos(!0t+'));\n(53c)\nand\n_x=\u0000r0!0sin(!0t+'0); (54a)\n_y=r0!0cos(!0t+'0); (54b)\n_z=qE0\nm! 01\n(\u000b2\u00001)(\u0000\u000bsin(\u000b!0t+') + sin(!0t+')):\n(54c)\nEquations (53a) to (53c) are the equations for plotting\nthe orbit of the electron as a function of time. Equa-\ntions (54a) to (54c) are for plotting the corresponding\nvelocities.6\nFIG. 4: The unperturbed orbit lies \rat along the e1\u0000e2plane. When \u000b= 0, or when the electric \feld is constant in time, the\norbit slants. When \u000b!1 , more waves are observed.\nWhen the perturbing frequency != 0, corresponding\nto\u000b= 0, the expressions for zand _zin Eqs. (53c) and\n(54c) reduces to\nz=\u0000qE0\nm!2\n0(cos'\u0000cos(!0t+')); (55a)\n_z=\u0000qE0\nm! 0sin(!0t+'): (55b)\nIf'= 0, the perturbing \feld is E=E0e3, so that\nz=\u0000qE0\nm!2\n0(1\u0000cos(!0t)); (56a)\n_z=\u0000qE0\nm! 0sin(!0t): (56b)\nOn the other hand, if '=\u0019, the perturbing \feld is E=\nE0e3, so that\nz=qE0\nm!2\n0(1 + cos(!0t)); (57a)\n_z=qE0\nm! 0sin(!0t): (57b)\nThese are the behavior of the electron's orbit along the\nz\u0000direction when the perturbing electric \feld is constant,\nalso known as the DC electric \feld.\nNow, when the perturbing frequency !=!0, corre-\nsponding to \u000b= 1, the \feld resonates with the electron'sorbit. The only terms a\u000bected are zand _zin Eqs. (53c)\nand (54c). Since both their numerators and denomina-\ntors approach zero as \u000b!1, we apply L'hopital's rule by\ndi\u000berentiating the numerators and denominators prior to\nevaluation of the limits:\nlim\n\u000b!1z=qE0\nm!2\n0lim\n\u000b!1\u0012\n\u0000!0tsin(\u000b!0t)\n2\u000b\u0013\n; (58a)\nlim\n\u000b!1_z=\u0000qE0\nm! 0lim\n\u000b!1\u0012sin(\u000b!0t+')\n2\u000b\u0013\n\u0000qE0\nm! 0lim\n\u000b!1\u0012\u000b!0tcos(!t+')\n2\u000b\u0013\n: (58b)\nHence,\nlim\n\u000b!1z=\u0000qE0\n2m! 0tsin(\u000b!0t); (59a)\nlim\n\u000b!1_z=\u0000qE0\n2m! 0(sin(!0t+') +!0tcos(!0t+')):\n(59b)\nNotice that the amplitude of the oscillations along zand\nits corresponding velocity are linearly increasing in time.\nOnce the amplitudes of the oscillations becomes so large,\nour perturbation approximations breaks down. Thus, our\ntheory cannot really say what happens during ionization\nor whether ionization will really happen at all at the res-\nonant frequency. (See Figs. 4 and 5)7\n\u000b= 1\n\u000b= 0:5\n\u000b= 2\ntz\nFIG. 5: Height zof the electron from its unperturbed circular\norbit with respect to time.\nIV. ANGULAR MOMENTUM\nA. Product Form\nLet us compute the product of the position rin\nEq. (30) and its velocity _rin Eq. (31a):\nr_r= (e1^r0^ 0+\u0015e3^s)(e1^r0ie3!0^ 0+\u0015e3_^s): (60)\nDistributing the terms, we get\nr_r=e1e1^r\u0003\n0^ \u0003\n0^r0ie3!0^ 0\n+e3e1(^s^r0ie3!0^ 0\u0000_^s^r\u0003\n0^ \u0003\n0) +e3e3s_^s; (61)\nafter setting the perturbation parameter \u0015= 1. Since\ni=e1e2e3andie3=e1e2, then Eq. (61) reduces to\nr_r=ie3!0r2\n0\u0000ie1s^r0!0^ 0\u0000ie2s^r\u0003\n0^ \u0003\n0+ ^s_^s: (62)\nSeparating the scalar and bivector parts of Eq. (62), we\narrive at\nr\u0001_r= ^s_^s; (63a)\nr\u0002_r=e3!0r2\n0\u0000e1s^r0!0^ 0\u0000e2s^r\u0003\n0^ \u0003\n0; (63b)\nafter factoring out the trivector iin the second equation.\nMultiplying Eq. (63b) by the electron's mass myields\nthe the electron's angular momentum:\nL=mr\u0002_r=e3!0r2\n0\u0000e1s^r0!0^ 0\u0000e2s^r\u0003\n0^ \u0003\n0;(64)\nwheresand _sare the e3components of the electron's\nposition rand velocity _rin Eqs. (51) and (52):\ns=qE0\nm!2\n01\n(\u000b2\u00001)(cos(\u000b!0t+')\u0000cos(!0t+'));\n(65a)\n_s=qE0\nm! 01\n(\u000b2\u00001)(\u0000\u000bsin(\u000b!0t+') + sin(!0t+'):\n(65b)Using the de\fnitions ^ r0=r0eie3'and ^ 0=eie3!0tin\nEq. (64), and separating the e1,e2, and e3components,\nwe arrive at\nL1=\u0000m! 0sr0cos(!0t+'0)\u0000m_sr0sin(!0t+'0);\n(66a)\nL2=\u0000m! 0sr0sin(!0t+'0)\u0000m_sr0cos(!0t+'0);\n(66b)\nL3=m! 0r2\n0; (66c)\nwhich are the parametric expressions for the angular mo-\nmentum in rectangular coordinates.\nB. Harmonic Form\nThe vertical oscillation sand its derivative _ sin\nEqs. (65a) and (65b) may be expressed in exponential\nforms:\ns=qE0\n2m!2\n01\n(\u000b2\u00001)(eie3(\u000b!0t+')+e\u0000ie3(\u000b!0t+')\n\u0000eie3(!0t+')\u0000e\u0000ie3(!0t+'));(67a)\n_s=qE0\n2m! 0\u0000ie3\n(\u000b2\u00001)(\u0000\u000beie3(\u000b!0t+')+\u000be\u0000ie3(\u000b!0t+')\n+eie3(!0t+')\u0000e\u0000ie3(!0t+')):(67b)\nThese may be rewritten as\ns=qE0\n2m!2\n01\n(\u000b2\u00001)(^\u0011^ \u000b\n0+ ^\u0011\u0003^ \u0000\u000b\n0\n\u0000^\u0011^ 0\u0000^\u0011\u0003^ \u00001\n0); (68a)\n_s=qE0\n2m! 0\u0000ie3\n(\u000b2\u00001)(\u0000\u000b^\u0011^ \u000b\n0+\u000b^\u0011\u0003^ \u0000\u000b\n0\n+ ^\u0011^ 0\u0000^\u0011\u0003^ \u00001\n0); (68b)\nwhere\n^\u0011=eie3': (69)\nSubstituting Eqs. (68a) and (68b) back to Eq. (64) and\nnoting that e2ie3=\u0000e1, we obtain\nL=e3m! 0r2\n0\u0000e1qE0\n2!01\n(\u000b2\u00001)^ L; (70)\nwhere\n^ L= ^\u0011^r0^ \u000b+1\n0+ ^\u0011\u0003^r0^ \u0000\u000b+1\n0\u0000^\u0011^r0^ 2\n0\u0000^\u0011\u0003^r0\n\u0000\u000b^\u0011^r\u0003\n0^ \u000b\u00001\n0+\u000b^\u0011\u0003^r\u0003\n0^ \u0000\u000b\u00001\n0 + ^\u0011^r\u0003\n0\u0000^\u0011\u0003^r\u0003\n0^ \u00002\n0:\n(71)\nExpanding the terms of ^ Linto exponential form, we get\n^ L=r0(eie3((\u000b+1)!0t+'+'0)+eie3((\u0000\u000b+1)!0t\u0000'+'0)\n\u0000eie3(2!0t+'+'0)\u0000eie3(\u0000'+'0)\n\u0000\u000beie3((\u000b\u00001)!0t+'\u0000'0)+\u000beie3(\u0000(\u000b+1)!0t\u0000'\u0000'0)\n+eie3('\u0000'0)\u0000eie3(\u00002!0t\u0000'\u0000'0)); (72)8\nSubstituting the result back to Eq. (70) and separating\nthee1,e2, and e3components, we arrive at\nL1=\u0000qE0r0\n2!01\n(\u000b2\u00001)\r1; (73a)\nL2=\u0000qE0r0\n2!01\n(\u000b2\u00001)\r2; (73b)\nL3=m! 0r2\n0; (73c)\nwhere\n\r1= (1 +\u000b) cos((\u000b+ 1)!0t+'+'0)\n+ (1\u0000\u000b) cos((\u000b\u00001)!0t+'\u0000'0)\n\u00002 cos(2!0t+'+'0); (74a)\n\r2= (1\u0000\u000b) sin((\u000b+ 1)!0t+'+'0)\n\u0000(1 +\u000b) sin((\u000b\u00001)!0t+'\u0000'0)\n+ 2 sin('\u0000'0): (74b)\nThus, since \u000b=!=! 0, we see that the orbit of the tip of\nthe angular momentum vector Lis a linear combination\nof circular motions with the following orbital frequencies:\n!L=f\u0006(!+!0);\u0006(!\u0000!0);\u00062!0;0g: (75)\nC. Limiting Conditions\nIn the DC \feld limit, \u000b!0, so that\nlim\n\u000b!0L1=qE0r0\n2!0(cos(!0t+'+'0)\n+ cos(\u0000!0t+'\u0000'0)\n\u00002 cos(2!0t+'+'0)); (76a)\nlim\n\u000b!0L2=qE0r0\n2!0(sin(!0t+'+'0)\n\u0000sin(\u0000!0t+'\u0000'0)\n+ 2 sin('\u0000'0)): (76b)\nOn the other hand, in the resonance frequency limit, \u000b!\n1, both the numerators and denominators of L1andL2\napproach zero, so that we apply the L'hopital's rule:\nlim\n\u000b!0L1=\u0000qE0r0\n4!0(cos(2!0t+'+'0)\n\u00002!0tsin(2!0t+'+'0)\u0000cos('\u0000'0));\n(77a)\nlim\n\u000b!0L2=\u0000qE0r0\n4!0(\u0000sin(2!0t+'+'0)\n\u0000sin('\u0000'0)\u00002!0tcos('\u0000'0)):(77b)\nNotice that at the resonant frequency !=!0, theL1\nandL2components of the angular momentum increases\nin time; in our perturbative approximation, the atom will\nbe ionized.V. POWER AND ENERGY ABSORPTION\nA. Work-Energy Theorem\nThe work-energy theorem states that\nZr\nr0F\u0001dr=1\n2mv2\u00001\n2mv02: (78)\nThis may be rewritten as\nZt\n0Pdt =1\n2mv2\u00001\n2mv02; (79)\nwhere the power Pis de\fned as\nP=F\u0001_r: (80)\nThat is, the integral of the power Pexpended by a force\nFacting to move a mass mfrom timet= 0 totis equal\nto the change in the mass's kinetic energy between these\ntimes.\nIn our model, there are two forces acting on the elec-\ntron: the Coulomb force Fcand the perturbing force Fp.\nBut since the sum of these two forces is mr, as given in\nEq. (28), then the left side of Eq. (79) becomes\nP=mr\u0001_r: (81)\nWe shall use this equation to compute the power ab-\nsorbed by the atom.\nB. Power: Product Form\nTo evaluate the left side of Eq. (81), we \frst multiply\nthe expressions for rand_rin Eqs. (31b) and (31a):\nr_r= (\u0000e1^r0!2\n0^ 0+\u0015e3s)(e1^r0ie3!0^ 0+\u0015e3_s) (82)\nDistributing the terms, we get\nr_r=\u0000e1e1^r\u0003\n0!2\n0^ \u0003\n0^r0ie3!0^ 0+e3e1s^r0ie3!0^ 0\n\u0000e1e3^r0!2\n0^ 0_s+e3e3s_s; (83)\nafter setting \u0015= 1. Since i=e1e2e3andie3=e1e2,\nthen Eq. (83) reduces to\nr_r=\u0000ie3r3\n0!3\n0\u0000ie1_s^r0!0^ 0+ie2_s^r0!2\n0^ 0+ s_s:(84)\nSeparating the scalar and bivector parts, we get\nr\u0001_r= s_s; (85a)\nr\u0002_r=\u0000e3r3\n0!3\n0\u0000e1_s^r0!0^ 0+e2_s^r0!2\n0^ 0;(85b)\nafter factoring out iin the second equation.\nEquation (85a) leads to a very simple expression for\nthe power absorbed by the atom:\nP=mr\u0001_r=ms_s: (86)9\nFIG. 6: Projection of the tip of the angular momentum vector Lin thexyplane for di\u000berent values of \u000b.\nTaking the time derivative of _ sin Eq. (65b),\ns=qE0\nm1\n(\u000b2\u00001)(\u0000\u000b2cos(\u000b!0t+') + cos(!0t+'));\n(87)\nand substituting this and that of _ sto Eq. (86), we get\nP=q2E2\n0\n2m!2\n01\n(\u000b2\u00001)2\u0002\n(\u0000\u000b2cos(\u000b!0t+') + cos(!0t+'))\u0002\n(\u0000\u000bsin(\u000b!0t+') + sin(!0t+')): (88)\nWe can show that this is equivalent to\nP=q2E2\n0\n2m! 01\n(\u000b2\u00001)2\u0000\n\u000b3sin(2\u000b!0t+ 2')\n\u0000(\u000b2+\u000b) sin((\u000b+ 1)!0t+ 2')\n+ (\u000b2\u0000\u000b) sin((\u000b\u00001)!0t)\n+ + sin(2!0t+ 2')); (89)\nwhich is the desired harmonic form of the power Pab-\nsorbed by the atom. Notice that power is not constant\nbut \ructuating in time.\nhPi\u001c\n\u000bhPi\nFIG. 7: This is the graph of the average power over the per-\nturbing period10\nC. Average Power over Perturbing Period\nLet us de\fne the average power over the perturbing\nperiod as\nhPi\u001c=1\n\u001cZ\u001c\n0Pdt; (90)\nwhere\n\u001c=2\u0019\n!=2\u0019\n\u000b!0: (91)\nNow, let us take the time average of the power Pin\nEq. (98) over the perturbing period:\nhPi\u001c=q2E2\n0\n2m! 01\n(\u000b2\u00001)2\u0000\n\u000b3hsin(2\u000b!0t+ 2')i\u001c\n\u0000(\u000b2+\u000b)hsin((\u000b+ 1)!0t+ 2')i\u001c\n+ (\u000b2\u0000\u000b)hsin((\u000b\u00001)!0t)i\u001c\n+hsin(2!0t+ 2')i\u001c); (92)\nwhere\nhsin(2\u000b!0t+ 2')i\u001c\n= 0; (93a)\nhsin((\u000b+ 1)!0t+ 2')i\u001c\n=\u00001\n2\u0019\u000b\n(\u000b+ 1)(cos((1 + 1=\u000b)2\u0019+ 2')\u0000cos(2'))\n(93b)\nhsin((\u000b\u00001)!0t)i\u001c\n=\u00001\n2\u0019\u000b\n(\u000b\u00001)(cos((1\u00001=\u000b)2\u0019)\u00001); (93c)\nhsin(2!0t+ 2')i\u001c\n=1\n2\u0019\u000b(cos(4\u0019=\u000b+ 2')\u0000cos(2')): (93d)\nSubstituting these back to Eq. (98), we get\nhPi\u001c=q2E2\n0\n4\u0019m! 01\n(\u000b2\u00001)2\u0002\n(\u0000\u000b2(cos((1 + 1=\u000b)2\u0019+ 2')\u0000cos(2'))\n\u0000\u000b2(cos((1\u00001=\u000b)2\u0019)\u00001)\n+\u000b(cos(4\u0019=\u000b+ 2')\u0000cos(2')): (94)\nIf'= 0, we have\nhPi\u001c=q2E2\n0\n4\u0019m! 01\n(\u000b2\u00001)2\u0002\n(\u0000\u000b2(cos((1 + 1=\u000b)2\u0019)\u00001)\n\u0000\u000b2(cos((1\u00001=\u000b)2\u0019)\u00001)\n+\u000b(cos(4\u0019=\u000b)\u00001)): (95)\nEquation (95) is graphed in Fig. 7. Notice that despite\nthe small oscillations in the interval \u000b= 0 and\u000b= 1,\nthe power absorption curve is similar to that in Lorentz\ndispersion theory.\nhPi\u001c0\n\u000bhPi\nFIG. 8: This is the graph of the average power over the orbital\nperiod\nD. Average Power over Orbital Period\nLet us de\fne the average power over the perturbing\nperiod as\nhPi\u001c0=1\n\u001c0Z\u001c0\n0Pdt; (96)\nwhere\n\u001c0=2\u0019\n!0: (97)\nNow, let us take the time average of the power Pin\nEq. (98) over the orbital period:\nhPi\u001c0=q2E2\n0\n2m! 01\n(\u000b2\u00001)2(\u000b3hsin(2\u000b!0t+ 2')i\u001c0\n\u0000(\u000b2+\u000b)hsin((\u000b+ 1)!0t+ 2')i\u001c0\n+ (\u000b2\u0000\u000b)hsin((\u000b\u00001)!0t)i\u001c0\n+hsin(2!0t+ 2')i\u001c0); (98)\nwhere\nhsin(2\u000b!0t+ 2')i\u001c0\n=\u00001\n2\u0019cos(4\u0019\u000b+ 2')\u0000cos(2')\n2\u000b; (99a)\nhsin((\u000b+ 1)!0t+ 2')i\u001c0\n=\u00001\n2\u0019cos(2\u0019(\u000b+ 1))\u0000cos(2')\n\u000b+ 1; (99b)\nhsin((\u000b\u00001)!0t)i\u001c0\n=\u00001\n2\u0019cos(2\u0019\u000b)\u00001\n(\u000b\u00001); (99c)\nhsin(2!0t+ 2')i\u001c0\n= 0: (99d)11\nSubstituting these back to Eq. (98), we get\nhPi\u001c0=q2E2\n0\n8\u0019m! 01\n(\u000b2\u00001)2\u0002\n(\u0000\u000b2(cos(4\u0019\u000b+ 2')\u0000cos(2'))\n\u00002\u000b(cos(2\u0019(\u000b+ 1))\u0000cos(2'))\n\u00002\u000b(cos(2\u0019\u000b)\u00001)): (100)\nIf'= 0, we have\nhPi\u001c0=q2E2\n0\n8\u0019m! 0\u00001\n(\u000b2\u00001)2(\u000b2(cos(4\u0019\u000b)\u00001)\n\u00002\u000b(cos(2\u0019(\u000b+ 1))\u00001)\u00002\u000b(cos(2\u0019\u000b)\u00001)):\n(101)\nEquation (101) is graphed in Fig. 8. Notice that unlike\nin Fig. 7, there are periodic oscillations after \u000b= 2.\nVI. CONCLUSION AND RECOMMENDATION\nWe modelled the classical Hydrogen atom as an elec-\ntron revolving in circular orbit around an immovable pro-\nton subject to the Coulomb force. We subjected this\natom to a perturbing oscillating electric \feld perpendicu-\nlar to the electron's initial orbital plane. We showed that\nthe resulting equations of motion of the electron along\nthe axis of the perturbing electric \feld is similar to that\nof a simple harmonic oscillator with sinusoidal forcing.\nFurthermore, the absorbed energy averaged over the pe-\nriod of the perturbing \feld or over the orbital frequencyof the electron is approximately similar to a resonance\ncurve with one dominant frequency with \fnite peak at\n!=!o; other small resonance peaks occur to the left or\nto the right of the major resonant frequency.\nThe Lorentz dispersion model of the light-atom inter-\naction assumes that the electron is subject to Hooke's\nforce and the force due to the oscillating electric \feld\nof the light. Interestingly, even if our initial assumption\nis an electron in circular orbit around the nucleus, we\nstill obtained the same forced harmonic oscillator equa-\ntion as that of the standard model. We also obtained\nthe same resonant frequency, though the actual peak is\nat a frequency slightly smaller than !0. But what is new\nis that even though we did not put a damping term in\nour harmonic oscillator equation, we still obtained a \f-\nnite energy absorption at the resonant frequency !. We\nalso computed the electron's angular momentum vector\nand showed that its tip traces rosette patterns similar to\nepicycles[23].\nIn the future work, we shall extend our work to the in-\nteraction of the hydrogen atom with elliptically polarized\nradiation.\nAcknowledgements\nThis work was supported by the Loyola Schools Schol-\narly Work Faculty Grants of the Ateneo de Manila Uni-\nversity.\n[1] J. Jackson, Classical Electrodynamics (Wiley, 1975).\n[2] A. Zangwill, Modern Electrodynamics (Cambridge Uni-\nversity Press, 2013).\n[3] M. Klein and T. Furtak, Optics (John Wiley and Sons,\n1986).\n[4] S. Akhmanov and S. Nikitin, Physical Optics (Oxford\nUniversity Press, 1997).\n[5] M.Born and E. Wolf, Principles of Optics (Oxford Uni-\nversity Press, 1964).\n[6] J. Bay\feld and P. Koch, Physical Review Letters 33, 258\n(1974).\n[7] H. Haken and H. C. Wolf, The Physics of Atoms\nand Quanta: Introduction to Experiments and Theory\n(Springer-Verlag, 2000).\n[8] D. Cole and Y. Zou, Journal of Scienti\fc Computing 21,\n145 (2004).\n[9] A. Brunello, D. Farelly, and T. Uzer, Physical Review A\n55, 3730 (1996).\n[10] D. Farelly and T. Uzer, Physical Review 74, 1720 (1995).\n[11] D. Farelly and J. Gri\u000eths, Physical Review A 45, R2678\n(1992).\n[12] M. Gajda, B. Piraux, and K. Rzazweski, Physical Review\nA50, 2528 (1994).\n[13] J. Leopold and I. Percival, Physical Review Letters 41,\n944 (1978).[14] E. Grosfeld and L. Friedland, Physical Review Letters E\n65, 1 (2002).\n[15] A. Neishtadt and A. Vasiliev, Physical Review Letters E\n71, 1 (2005).\n[16] Q. M. Sugon Jr., S. Bragais, and D. J. McNamara, Coper-\nnicus's epicycles from newton's gravitational force law via\nlinear perturbation theory in geometric algebra (2008),\nmath/0807.2708v1.\n[17] T. Vold, Am. J. Phys. 61, 505 (1993).\n[18] D. Hestenes, Am. J. Phys. 71, 104 (2003).\n[19] Q. M. Sugon Jr. and D. J. McNamara, Am. J. Phys. 72,\n104 (2004).\n[20] W. E. Baylis, Cli\u000bord (Geometric) Algebras with Ap-\nplications in Physics, Mathematics, and Engineering\n(Birkh auser, 1996).\n[21] P. Lounesto, Cli\u000bord Algebra and Spinors (Cambridge\nUniversity Press, 2001).\n[22] B. Jancewicz, Multivectors and Cli\u000bord Algebras in Elec-\ntrodynamics (World Scienti\fc, 1989).\n[23] G. Gallavotti, ATTI-Accademia Nazionale Dei Lincei\nRendiconti Lincei Classe di Scienze Fisiche Matemaiche\ne Naturali Serie 9 Matematica e Applicazioni 12, 125\n(2001)." }, { "title": "1211.4781v1.Damping_rates_of_surface_plasmons_for_particles_of_size_from_nano__to_micrometers__reduction_of_the_nonradiative_decay.pdf", "content": "arXiv:1211.4781v1 [physics.optics] 20 Nov 2012Damping rates of surface plasmons for particles of size from nano- to\nmicrometers; reduction of the nonradiative decay\nK. Kolwas∗∗, A. Derkachova∗\nInstitute of Physics, Polish Academy of Sciences, Al. Lotni k´ ow 32/46, Warszawa, Poland\nAbstract\nDamping rates of multipolar, localized surface plasmons (SP) of gold a nd silver nanospheres\nof radii up to 1000 nmwere found with the tools of classical electrodynamics. The significa nt\nincrease in damping rates followed by noteworthy decrease for larg er particles takes place along\nwith substantial red-shift of plasmon resonance frequencies as a function of particle size. We\nalso introduced interface damping into our modeling, which substant ially modifies the plasmon\ndamping rates of smaller particles. We demonstrate unexpected re duction of the multipolar SP\ndamping rates in certain size ranges. This effect can be explained by t he suppression of the\nnonradiative decay channel as a result of the lost competition with t he radiative channel. We\nshow that experimental dipole damping rates [H. Baida, et al., Nano Le tt. 9(10) (2009) 3463,\nand C. S¨ onnichsen, et al., Phys. Rev. Lett. 88 (2002) 077402], an d the resulting resonance\nquality factors can be described in a consistent and straightforwa rd way within our modeling\nextended to particle sizes still unavailable experimentally.\nKeywords : plasmon damping, radiation damping, interface damping, surface p lasmon res-\nonance, multipolar plasmons, multipolar plasmon modes, Mie theory, g old nanoparticles, sil-\nver nanoparticles, nanosphere, nanoantenna, nanophotonics, plasmonics, size dependent optical\nproperties, SERS, SP.\n1. Introduction\nThe response of noble-metal nanoparticles to electromagnetic (E M) excitation is dominated\nbyresonantexcitationsoflocalizedsurfaceplasmons(SPs)(see: [1,2,3,4,5]forreviews). When\na metal particle is illuminated at corresponding resonance frequenc y, notably strong surface-\nconfined optical fields can be generated. This property is applied in s urface-enhanced Raman\n∗Corresponding author\n∗∗Principal corresponding author\nEmail addresses: Krystyna.Kolwas@ifpan.edu.pl (K. Kolwas), Anastasiya.Derkachova@ifpan.edu.pl\n(A. Derkachova)\nPreprint submitted to Elsevier November 21, 2012spectroscopy (SERS) [6, 7, 8], high-resolution microscopy [9] or imp rovement of plasmonic solar\ncells [10, 11, 12, 13, 14, 15]. Excitation of SPs at optical frequencie s, non-diffraction-limited\nguidingofthem(e.g. viaalinearchainofgoldnanospheres[8,16,17, 18,19,20])andtransferring\nthem back into freely propagating light are processes of great impo rtance in many applications.\nGold and silver nanoparticles are important components for subwav elength integrated optics\nand high-sensitivity biosensors [4, 21, 22] due to their chemical iner tness and unique optical\nproperties in the visible to near-infrared spectral range.\nMetal nanospheres are the simplest and the most fundamental st ructures for studying the\nbasis of plasmon phenomena. Understanding the resonant interac tion of light with plasmonic\nnanoparticles is also essential for applications, e.g. for designing us eful photonic devices. The\nmodeling of plasmon properties under the simplest electrostatic (qu asistatic) approximation is\nvalid only for particles much smaller than the wavelength of light; than size dependence of\nplasmon properties can be neglected. If a nanosphere is larger, th e properties of the surface\nplasmons supported by this structure become dependent on the s ize of the particle [1, 23, 24,\n25, 26, 27]. Size dependence of the retarded electronic response of a plasmonic particle is\ndetermined by the properties of metal at the SP resonance wavele ngth and the presence of the\nmetal-dielectric interface.\nNanoparticles much smaller than optical wavelength first of all exhib it dipolar surface plas-\nmon oscillations. Higher order multipolar resonances appear as new f eatures in the optical\nspectra, at frequencies higher than that of the dipolar plasmon fr equency. The spectral sig-\nnatures of these higher order plasmon resonances were observe d in elongated gold and silver\nnanoparticles [28, 29, 30] first. Fewer experimental investigation s directly demonstrated the\nmultipolar character of the observed resonances in spherical par ticles [26, 31, 32].\nControlling the spectral properties of a plasmonic nanosphere for technological or diagnostic\napplications is not possible without knowing the direct dependence of plasmon properties on\nparticlesize. Stillitisbelieved(e.g. [5,25])thatexistingtheoriesdono tallowtherigorousdirect\ncalculation of the frequencies of the multipolar SP modes. However, apart from the plasmon\nresonance frequencies, also the plasmon decay rates (damping tim es) are essential in controlling\nthespectralresponseoftheplasmonicparticle. Totaldampingra tesofmultipolarSPs[27]define\nabsorbing and emitting properties of a plasmonic nanoantennas whic h can be tuned by particle\nsize. Nanoantenna of the right size can serve as an effective trans itional structure which is able\nto absorb or to transmit electromagnetic radiation in plasmonic mech anism. The knowledge\nand modeling of plasmon damping effect as a function of particle size ar e thus of key interest for\nthe development of plasmonic nanosystems. The central feature here is the actual enhancement\nof the electromagnetic field. It served as a stimulus for us for deve loping a strict and direct\nsize characterization of multipolar plasmons resonance frequencie s and plasmon damping rates.\n2Our extended modeling up to the radius R= 1000nm, and up to SP multipolarity l= 10\nallows us to determine the intrinsic properties of plasmonic spheres in the size range which has\nnever been studied before. We do not apply any restrictions on the particle radius in relation\nto the wavelength of light; our study goes far beyond the particle s ize for which the quasistatic\napproximation is justified (e.g. [2, 5]).\nThe main reason of size sensitivity of SP which we discuss (usually refe rred to as the retar-\ndation effect ([33] and references therein)) is responsible for impo rtant red-shift of the plasmon\nresonance frequency for larger nanoparticles. It is accompanied by the significant increase in\ndamping rate followed by noteworthy decrease for larger particles . In addition, we discuss the\neffect called the surface electron scattering effect [1] or interfac e damping [34] which causes the\nsubstantial modificationoftheplasmondamping ratesinparticles of sizes comparableorsmaller\nthan the mean electron free path. Sometimes called the ”intrinsic siz e effect”, it is described by\nthe 1/Rdependence added to the electron relaxation rate γof the dielectric function [1, 5, 35].\nWe perform both types of modeling including or neglecting the effect o f surface scattering, in\norder to examine the causes of the size-dependent plasmon featu res.\nOur extended modeling for large particles reveals new features of t he total plasmon damping\nrate. We demonstrated, for the first time in spherical particles, t he effect of reducing of mul-\ntipolar SP damping rates below its low size limit which is equal to the nonra diative damping\nrate. Suppression of the total damping rates proven to exist in ce rtain size ranges allowed a new\ninsight into the role of radiation damping in the plasmon decay mechanis m. We also describe\nchanges of the quality Q-factor of SP multipolar resonances with pa rticle size. Q-factor is used\nfor determination of the local field enhancement [34] and effective s usceptibility in nonlinear\noptical processes [36].\nWe also confront the size characteristics resulting from our modelin g with the experimental\nresults obtained for the dipole plasmon for gold [32, 34] and silver [37 ] particles of different sizes.\nWe show that the measured dipole damping rates and quality factors [32, 34] can be described\nin a consistent and straightforward way within our modeling which deliv er also predictions for\nparticle of size range still unavailable experimentally.\n2. Size dependence of SP parameters derived from optical spe ctra\nIt is widely accepted, that excitation of surface plasmons (collectiv e free-electron oscilla-\ntions) is responsible for the commonly observed pronounced peaks in the optical absorption or\nextinction spectra. The spectra collected for nanoparticles of va rious sizes are used as a source\nof data allowing to reconstruct the change of the resonance (usu ally dipole only) position with\nsize (e.g. [1, 34, 38]. The width of the peak is related to the (dipole) pla smon damping rate\n[32, 34, 37, 39, 40, 41].\n3Mie theory delivers an indispensable formalism enabling the description of scattering of a\nplane monochromatic wave by a homogeneous sphere of known radiu s, surround by a homoge-\nneous medium [42, 43, 44, 45]. Mie formulas allow to predict the intensit ies of light scattered in\na given direction or the absorption and scattering cross-section s pectra for particles of a chosen\nsize. The Mie spectra can be compared with the spectra (intensities of a bsorbed or scattered\nlight) measured experimentally for particles of the same size .\n/s49/s44/s50 /s49/s44/s54 /s50/s44/s48 /s50/s44/s52 /s50/s44/s56/s48/s53/s48/s49/s48/s48/s49/s53/s48/s50/s48/s48\n/s49/s44/s50 /s49/s44/s54 /s50/s44/s48 /s50/s44/s52 /s50/s44/s56/s48/s53/s48/s49/s48/s48/s49/s53/s48/s32/s32\n/s115/s99/s97/s116\n/s32/s32\n/s97/s98/s115\n/s39\n/s108/s61 /s51/s39\n/s108/s61/s50/s39\n/s108/s61 /s49\n/s98/s41\n/s115/s99/s97/s116/s44/s32/s32\n/s97/s98/s115/s82/s32 /s61/s32/s56/s32/s110/s109/s97/s41\n/s32/s91/s101/s86/s93/s32\n/s115/s99/s97/s116/s44\n/s97/s98/s115\n/s39\n/s108/s61 /s52/s39\n/s108/s61 /s51/s39\n/s108/s61 /s49/s39\n/s108/s61 /s50/s82/s32 /s61/s32/s56/s48/s32/s110/s109\nFigure 1: Absorption Cabs(ω) and scattering Cscat(ω) cross-sectionsfor nanospheres ofradii a) 8nm and b) 80nm\n(Mie theory). Figures illustrate suppression of absorption in larger particles. Broadening, deformation and shift\nof the maxima in the scattering spectra of larger particles in compar ison with the SP resonance eigenfrequencies\nω′\nlis also shown.\nHowever, predicting the size dependence of multipolar plasmon resonance frequencies, Mie\nscattering theory is an inconvenient tool. SP size dependence can b e determined only indirectly,\nby laborious derivation of positions of maxima in the consecutive spec tra collected for various\nparticleradii. Thesameappliestoderivationofthesizedependence o fthe(dipole)SPfrequency\n4corresponding to the peak position from experimental data. In ad dition, the maximum in\nthe spectrum ascribed to SP resonance can be broadened, defor med and shifted in respect\nto the frequency of plasmon oscillation mode [26, 27], as illustrated in F igure 1. That can\naffect the plasmon resonance position obtained from the experimen tal or calculated scattering\nspectra. It is even more difficult to determine the SP damping times fr om the width of the SP\nresonances, manifested in the optical spectra of larger particles , and to understand the derived\nsize dependence then (e.g. [34, 37, 38]). The processes leading to plasmon damping are the\nsubject of extensive research and debate (e.g. [1, 34, 39, 40, 46 , 47, 48]. The homogeneous\nlinewidth of the SP resonance was connected to the SP damping time. The properties of the\nSP are strongly influenced by this parameter. Mie scattering theor y is not sufficient to define\na general rule describing the size dependence of SP parameters; t he damping rates and the\nresonance frequencies are not parameters of this theory.\n3. Direct size characterization of multipolar SP modes\nIn order to derive parameters of SP as a function of particle size, w e use an accurate elec-\ntrodynamic approach based on [49], and described in more details in e .g. [50]. The formalism\nof Mie theory is used. However, the problem is formulated in absence of the illuminating light\nfield; we are interested in intrinsic eigenproperties of the spherical particle (the analogy to the\ncavity eigenproblem) [26, 27]. In the present paper we extend the n umerical calculations up to\nthe radius of 1000 nmand plasmon polarity from l= 1 up tol= 10. That allowed us to explore\nnew features of localized SP excitations and reconsider the results of [27].\nWe consider continuity relations at the spherical metal/dielectric int erface for the electro-\nmagnetic (EM) fields, which arethe solutions of the Helmholtz equatio n in spherical coordinates\n(r,θ,φ) inside and outside the sphere. The transverse magnetic (TM) mod es of EM waves lo-\ncalized on the interface possess a nonzero component of the elect ric field normal to the surface\nEr, which can couple with free-electron charge oscillations at the boun daryr=R. The condi-\ntions for the nontrivial solutions of the continuity relations for TM m ode define the dispersion\nrelation:\n/radicalBig\nεin(ω)ξ′\nl(kout(ω)R)ψl(kin(ω)R)−/radicalBig\nεout(ω)ξl(kout(ω)R)ψ′\nl(kin(ω)R) = 0 (1)\nwhich is fulfilled for the complex eigenfrequencies of the field Ω lin successive multipolar\nmodesl= 1,2,3,... . ψl(z) andξl(z) are Riccati-Bessel spherical functions, the prime symbol\n(′) indicates differentiation with respect to the argument, kin=ω\nc/radicalBig\nεin(ω) andkout=ω\nc√εout,\nεin(ω) andεoutare the dielectric functions of a metal and of the particle dielectric s urrounding,\nrespectively. The dispersion relation (Equation(1)) is solved numer ically for complex values of\nΩl(R) =ω′\nl(R) +iω′′\nl(R) (ω′′\nl(R)<0) forlin the range of 1 ÷10. We solved the problem\n5for successive Rvalues from 1 to 1000 nmin 1nmsteps. The fsolvefunction of the MatLab\nprogram, utilizing the Trust-region dogleg algorithm was used.\n3.1. Material properties of plasmonic particles\nThe simplest analytic function often used to describe a wavelength d ependence of optical\nproperties of metals like gold or silver [51, 52, 53, 54, 55] results fro m the Drude-Lorentz-\nSommerfeld model:\nεD(ω) =ε0−ω2\np/(ω2+iγω) (2)\nwithε0= 1 (e.g. [1, 5]). The effective parameter ε0>1 describes contribution of bound\nelectrons,ωpis the effective bulk plasma frequency which is associated with effectiv e concentra-\ntion of free-electrons, γis the phenomenological damping constant of electron motion. For b ulk\nmetalsγ=γbulkis related to the electrical resistivity of the metal and is supposed t o include all\nmicroscopic damping processes due to photons, phonons, impuritie s and electron-electron inter-\nactions. In the present modeling we accept the following effective pa rameters of the dielectric\nfunction of the bulk gold: ε0= 9,84,ωp= 9,010eV,γbulk= 0,072eV.\nThe collision time 1 /γbulkdetermines the electron mean free path in bulk metals. At room\ntemperatures the electron mean free path in gold is 42 nm[1]. It can be comparable or larger\nthan a dimension of a particle. An additional relaxation term added to the relaxation rate γ\naccounts for the effect of scattering of free electrons by the su rface [1, 35, 41, 56, 57, 58]:\nγR(R) =γbulk+AvF\nR(3)\nwherevFis the Fermi velocity ( vF= 1.4·10−6m/sin gold), and Ais the theory dependent\nquantity of the order of 1 [1]. We accept the value A= 0.33 in our modeling, according to [57].\nThen the size-adopted bulk dielectric function is:\nεγ(R)(ω,R) =ε0−ω2\np/(ω2+iγR(R)ω) (4)\nThe real part of εγ(R)is not modified by surface scattering. Frequency (and radius) dep en-\ndence of the dielectric function (Equation 2 or 4) couples to the ove rall frequency (and radius)\ndependence of the plasmon dispersion relation (Equation(1)).\nTo underline the crucial role of interface damping in smaller particles, we will compare the\nresults obtained with the dielectric function εin=εD(ω) given by Equation (2) (with γ=γbulk,\nsurface scattering neglected) and εin=εγ(R)(ω,R) given by Equation (4) (with γ=γR(R)).\nThe dielectric function of the particle surrounding (vacuum/air) is a ssumed to be εout= 1, or\nis chosen to reflect the index of refraction of the particle environm ent, as described in Section\n7, where we compare the results of our modeling with some experimen tal results of [34, 37].\n63.2. Resonance frequencies and damping rates vs radius; the role of the interface damping\nThe solutions of the dispersion relation (Equation(1)) define the siz e dependence of multipo-\nlarplasmonoscillationfrequencies ω′\nl(R) =Re(Ωl(R))anddampingrates |ω′′\nl(R)|=|Im(Ωl(R))|\n(and SP damping times Tl(R) = ¯h/|ω′′\nl(R)|) of the EM surface modes oscillations. We express\nω′\nl,ω′′\nl,ωandγin electronvolts (eV) for convenience.\nThe size dependence of ω′\nl(R) and|ω′′\nl(R)|of SP modes for the first ten multipolar plasmons\nfor gold spheres in vacuum/air ( εout= 1,γ=γbulk= 0.072eV, the radius up to 1000 nm) is\npresented in Figure 2. Black line ( l= 1) represents the dipole resonance frequency. Resonant\nexcitation of SP oscillations takes place when the frequency of incom ing light field ωapproaches\nthe eigenfrequency of a plasmonic nanoantenna of a given radius R:ω=ω′\nl(R),l= 1,2,3,.... If\nexcited, plasmon oscillations are damped at corresponding rates |ω′′\nl(R)|(Figure 2b)). Damping\nof surface plasmon oscillations is the inherent property of SP modes that defines absorbing and\nscattering properties of the plasmonic particle as a function of size .\n/s48 /s50/s53/s48 /s53/s48/s48 /s55/s53/s48 /s49/s48/s48/s48/s48/s44/s48/s48/s44/s53/s49/s44/s48/s49/s44/s53/s50/s44/s48/s50/s44/s53/s51/s44/s48/s51/s44/s53\n/s48 /s50/s53/s48 /s53/s48/s48 /s55/s53/s48 /s49/s48/s48/s48/s48/s44/s48/s48/s44/s49/s48/s44/s50/s48/s44/s51/s48/s44/s52/s48/s44/s53/s48/s44/s54/s48/s44/s55/s48/s44/s56\n/s48 /s51/s48/s48 /s54/s48/s48/s48/s44/s48/s51/s50/s48/s44/s48/s51/s52/s48/s44/s48/s51/s54/s39\n/s108/s32/s32/s91/s101/s86/s93/s32\n/s82 /s32/s91/s110/s109/s93/s97/s41 /s98/s41\n/s47/s50/s39/s39\n/s108/s124/s32/s32/s91/s101/s86/s93/s32\n/s82 /s32/s91/s110/s109/s93/s32/s108/s32/s61/s32/s49\n/s32/s108/s32/s61/s32/s50\n/s32/s108/s32/s61/s32/s51\n/s32/s108/s32/s61/s32/s52\n/s32/s108/s32/s61/s32/s53\n/s32/s108/s32/s61/s32/s54\n/s32/s108/s32/s61/s32/s55\n/s32/s108/s32/s61/s32/s56\n/s32/s108/s32/s61/s32/s57\n/s32/s108/s32/s61/s32/s49/s48/s47/s50/s124\n/s108/s39/s39/s124/s32/s91/s101/s86/s93/s32\n/s32/s82/s32 /s91/s110/s109/s93\n/s32/s32\n/s32/s84\n/s108/s32/s32/s91/s102/s115/s93\n/s49/s56/s44/s51/s54/s44/s54/s51/s44/s51/s50/s44/s50/s49/s44/s54/s49/s44/s51/s49/s44/s49/s48/s44/s57/s48/s44/s56\nFigure 2: a) Multipolar plasmon resonance frequencies ω′\nl(R) and b) plasmon damping rates |ω′′\nl(R)|(left axis)\nand corresponding damping times Tl(right axis) vs particle radius R. Reduction of the plasmon damping rates\n|ω′′\nl(R)|below the nonradiative limit γ/2) is demonstrated in the inset.\nThe SP frequencies ω′\nl(R) decrease with size monotonically, as illustrated in Figure 2a). For\ngiven particle radius R,ω′\nl(R) increase with plasmon polarity l. The multipolar SP resonances\nfor larger particles are better spectrally resolved than those for the smaller ones.\nIn the limit of small size, the analytic expressions for ω′\nl(R) andω′′\nl(R) can be found from\nthe relation (1) after applying the power series expansion of the sp herical Bessel and Hankel\nfunctions. Keeping only the first terms of the power series one can get:\n7ω′\n0,l=/bracketleftBiggω2\np\nε0+l+1\nlεout−/parenleftbiggγ\n2/parenrightbigg2/bracketrightBigg1/2\n, (5)\nω′′\n0,l=−γ\n2. (6)\nFor a perfect free-electron metal ( ε0= 1,γ= 0) andεout= 1, Equation (5) leads to\nthe well-known plasmon frequencies within the ”quasistatic approxim ation” [1, 50, 59]: ω′\n0,l=\nωp//radicalBig\n1+εout(l+1)/l, and in particular to the giant Mie resonance frequency ω′\n0,l=1=ωp/√\n3\nforl= 1.\nIn the smallest particles, the plasmon damping rates |ω′′\nl(R)|fall to pure nonradiative damp-\ning rates |ω′′\n0,l|=γnrad=γ/2, as demonstrated in Figure 2b). The values of |ω′′\n0,l|are the same\nfor all multipolar plasmon modes l= 1,2,3...(Equation (6)). The nonradiative damping rates\nγnradare due to the ohmic losses if interface damping is neglected ( γ=γbulk), or are equal to\nγR/2 (Equation 3) if interface damping is included in the modeling.\n/s48 /s53/s48 /s49/s48/s48 /s49/s53/s48/s49/s44/s48/s49/s44/s53/s50/s44/s48/s50/s44/s53/s51/s44/s48\n/s48 /s53/s48 /s49/s48/s48 /s49/s53/s48/s48/s44/s48/s48/s44/s50/s48/s44/s52/s48/s44/s54/s48/s44/s56\n/s53/s48 /s49/s48/s48 /s49/s53/s48/s48/s44/s48/s51/s53/s48/s44/s48/s52/s48/s48/s44/s48/s52/s53/s48/s44/s48/s53/s48\n/s48 /s49/s48 /s50/s48 /s51/s48 /s52/s48/s50/s44/s53/s53/s50/s44/s54/s48/s50/s44/s54/s53/s50/s44/s55/s48\n/s32/s108/s32/s61/s32/s49\n/s32/s108/s32/s61/s32/s50\n/s32/s108/s32/s61/s32/s51\n/s32/s108/s32/s61/s32/s49\n/s32/s108/s32/s61/s32/s50\n/s32/s108/s32/s61/s32/s51\n/s40/s82 /s41 /s48/s32/s45/s32/s50\n/s112/s47/s40/s50\n/s32/s45/s32/s105\n/s82/s40/s82 /s41 /s41/s68 /s48/s32/s45/s32/s50\n/s112/s47/s40/s50\n/s32/s45/s32/s105 /s41/s108/s39/s32/s91/s101/s86/s93/s32\n/s82 /s32/s91/s110/s109/s93/s97/s41 /s98/s41\n/s47/s50/s108/s39/s39/s124/s32/s91/s101/s86/s93/s32\n/s82 /s32/s91/s110/s109/s93/s47/s50/s32/s124\n/s108/s39/s39/s124/s32/s91/s101/s86/s93\n/s32/s82 /s32/s91/s110/s109/s93/s82/s40/s82 /s41/s47/s50\n/s49/s56/s44/s51\n/s32/s84\n/s108/s32/s91/s102/s115/s93\n/s54/s44/s54/s51/s44/s51/s50/s44/s50/s49/s44/s54/s49/s44/s51/s49/s44/s49/s48/s44/s57/s48/s44/s56/s108/s39/s32/s91/s101/s86/s93 /s32\n/s82 /s32/s91/s110/s109/s93/s97/s41\nFigure 3: a) Dipole, quadrupole and hexapole plasmon resonance fre quencies ω′\nl(R) and b) corresponding plas-\nmon damping rates |ω′′\nl(R)|calculatedwithout (solid lines) and with (lines with hollow circles)interfa ce damping\ntaken into account. Insets are magnifications of figures a) and b) for particles of small radii.\nInterface damping practically does not influence the size dependen ce ofω′\nl(R) as illustrated\nin Figure 3a). The minute red shift for the smallest sizes due to interf ace damping is shown\nin inset of Figure 3a). A similar finding results from the experimental o bservations [32, 34].\nTherefore, the red shift of the (dipole) SP resonance frequency sometimes observed in small\n8particles versus decreasing size must result from some other phen omena due to the complex\nchemical effects and uncertainties in experimental samples [60].\nIn extended size range that we study, plasmon damping rates |ω′′\nl(R)|are not simple mono-\ntonicfunctionsoftheradius(Figure2b)). Theinitialfastincrease of|ω′′\nl(R)|withsizeisfollowed\nby gradual decrease for sufficiently large radii R. The increase in |ω′′\nl+1(R)|in the subsequent\nl+1 SP modes is followed by the decrease in the SP damping rates |ω′′\nl(R)|of lower polarity\nmodes.\nThe surface scattering effect affects the total damping rate |ω′′\nl(R)|, as illustrated in Figure\n3b) and in the magnification of the part of this graph presented in Fig ure 4. The surface\nscattering contribution to the total relaxation rate γR(R) looses its importance for particles of\nradius larger than ∼300nm(using the criterion ( γR−γ)/γR≈1%).\n/s48 /s53/s48 /s49/s48/s48 /s49/s53/s48 /s50/s48/s48 /s50/s53/s48 /s51/s48/s48/s48/s44/s48/s51/s53/s48/s44/s48/s52/s48/s48/s44/s48/s52/s53/s48/s44/s48/s53/s48\n/s32/s84\n/s108/s32/s91/s102/s115/s93\n/s82/s124\n/s108/s39/s39/s124/s32/s91/s101/s86/s93\n/s82 /s32/s91/s110/s109/s93/s68 /s48/s32/s45/s32/s50\n/s112/s47/s40/s50\n/s32/s45/s32/s105 /s41/s32/s32/s32/s32\n/s40/s82 /s41 /s48/s32/s45/s32/s50\n/s112/s47/s40/s50\n/s32/s45/s32/s105\n/s82/s40/s82 /s41 /s41/s32\n/s32/s32/s32 /s32/s108/s32/s61/s32/s49 /s32/s32/s32 /s32/s108/s32/s61/s32/s49\n/s32/s32/s32 /s32/s108/s32/s61/s32/s50 /s32/s32/s32 /s32/s108/s32/s61/s32/s50\n/s32/s32/s32 /s32/s108/s32/s61/s32/s51 /s32/s32/s32 /s32/s108/s32/s61/s32/s51\n/s32/s32/s32 /s32/s108/s32/s61/s32/s52 /s32/s32/s32 /s32/s108/s32/s61/s32/s52\n/s32/s32/s32 /s32/s108/s32/s61/s32/s53 /s32/s32/s32 /s32/s108/s32/s61/s32/s53\n/s49/s54/s44/s53/s49/s52/s44/s54/s49/s51/s44/s50\n/s49/s56/s44/s56\nFigure 4: Demonstration of the reduction effect of plasmon damping rates|ω′′\nl(R)|calculated without ( γ=γbulk,\nsolid lines) and with ( γ=γR(R), lines with hollow circles) interface damping; the magnification of a pa rt of\nFigure 3.\n4. Radiative and nonradiative contributions to the total SP damping rates\nThe processes leading to damping of SP have been the subject of wid e discussion and ex-\ntensive studies, e.g. [1, 32, 34, 61, 62, 63] which, however, were lim ited to the dipole case. It\nwas accepted that (dipole) SP in metal nanoparticles decays throu gh both inelastic processes\nand elastic dephasing process which were usually neglected. Inelast ic processes can be further\ndivided into radiative and non-radiative decay processes [33, 34, 61 ]. The SP damping time\nTwas determined from the homogeneous linewidth Γ = 2¯ h/Tof a maximum in the spectrum\n9due to the SP dipole resonance. The observed spectral broadenin g of the maxima with size due\nto radiation is sometimes expected to be proportional to the partic le volume [34, 57, 64]. Our\nresults show (Figure 2b)), that the multipolar plasmon damping rate s|ω′′\nl(R)|are not simple,\nmonotonic functions of particle radius. The initial fast increase of t he total SP damping rate\n|ω′′\nl(R)|with size is followed by gradual decrease for sufficiently large radii fo r the given mode\nl. Our modeling suggest that the proportionality of the multipolar rad iation damping rate to\nR3is not a general rule (see Figure 2b)).\nOur extended modeling allows us to consider the higher order multipola r damping rates\n|ω′′\nl(R)|l=1,2...10, as well. The total multipolar damping times ¯ h/Tlcan be related to the\ncorresponding damping rates (homogeneous linewidths of the abso rption [27] spectra Γ l(R))\nand their size dependencies in a natural way:\nΓl(R) = 2|ω′′\nl(R)|= 2¯h/Tl (7)\nand\n|ω′′\nl|=γrad\nl+γnrad(8)\nwhereγnrad=γ/2. The assumption, that the nonradiative and radiative processes (in\ndipoleplasmon mode) areadditive andareindependent, is commonly ac cepted (see for example:\n[27, 41, 57, 58, 61]). However, expectation that the size depende nce of the total damping rate\n|ω′′\nl(R)|is due to the size dependence of the radiative damping rate γrad\nl(R) only (|ω′′\nl(R)|=\nγrad\nl(R)+γnrad, [27]) must be reconsidered (see Section 5 bellow).\nIf the surface scattering effect is included, |ω′′\nl(R)|becomes a decreasing function of size,\nstarting fromsmaller particles (see Figure3b) and4, lines with symbo ls). The total SPdamping\nrates follow the size dependence of nonradiative damping: γR(R)/2:|ω′′\nl(R)| ≈γR(R)/2 =\nγnrad(R) in the range of radii which extends to larger Rfor growing SP polarity l. After\nreaching theminimum, |ω′′\nl(R)|tendstofollowtheradiusdependence unaffectedbytheinterface\ndamping.\nA contribution of the radiation damping γrad\nl(R) to the total damping rate |ω′′\nl(R)|can be\ntreated as a measure of the ability of the particle to couple to the inc oming field and to emit\nlight in plasmonic mechanism. As long as |ω′′\nl(R)| ≃γnrad, the SP mode lcan only weakly\ncouple with the incoming radiation and has weak radiative abilities. Cons equently, the weakly\nradiative plasmons appear in the absorption spectra with a smaller am plitude and in scattering\nspectra (see Figure 1) they hardly manifest. With increasing l, SP plasmons gain the radiative\ncharacter starting from larger particles due to the fact that γrad\nl>1(R)>γrad\nl−1(R). Therefore, the\nhigher order SP modes can be excited (and appear in the scattering spectra) for larger particles.\n10This fact is known from Mie scattering theory, but its physical sens e have not been explained\nby Mie solutions.\n5. Effect of reduction of multipolar plasmon damping rates\nIn the easier case (interface damping omitted in the modeling), redu ction of the total SP\ndumping rates γnradbellow the γ/2 value (see the inset in Figure 2b)) manifests for plasmon\nmodes with l>1. This reduction effect can be clearly distinguished from other size d ependent\nprocesses in some size ranges which grow with increasing l; for the quadrupole ( l= 2) plasmon\nmode, the reduction effect extends up to R≃40nm, for the hexapole ( l= 3) plasmon mode up\ntoR≃92nm, forl= 4 up toR≃155nmand so on.\nWe have checked, the reduction of the total damping rate is not pr esent ifγ= 0 (2). On\nthe contrary, it does not disappear if γ/negationslash= 0 andε0= 1. That suggests, that if the additive form\nof the Expression (8) holds, reduction of the total plasmon dampin g rate is due to the decrease\nof nonradiative decay γnradas compared with its low size limit γnrad=γ/2 resulting from\nthe energy dissipation of meatal (absorption). Consequently, th e radiative and nonradiative\ncontributions have to be coupled by their size dependence. The mod el with interface damping\nincluded, reproduces the SP rate reduction below γnrad=γR/2 (see Figure 4), as well.\nWe can conclude, that reduction of |ω′′\nl(R)|below theγ/2 =γbulk/2 (orγ/2 =γR/2) value\ntakes place in the regions of sizes where the nonradiative damping is s till not dominated by the\nfast radiative damping. Reduction in the total plasmon damping rate must be connected with\nsuppression of the nonradiative decay channel. It can be inferred then, that the radiative and\nnonradiativeprocessesarenotindependent. Theeffective rateo fthenonradiativedampingmust\nbe size dependent: γnrad=γnrad(R) with the value for the small particle limit: γnrad\n0=γ/2.\nSuppression of the nonradiative damping is not restricted to the siz e ranges for which the effect\nwas demonstrated, but influences the optical properties of plasm onic particles in a large range\nof sizes. Reduction of γnrad(R) manifests in the absorption spectrum of particles; absorptive\nabilities of large particles are poor as compared with small particles. T his fact is described by\nsolutions of Mie scattering theory, but its physical meaning have no t been explained.\nThe effect of the total damping rate reduction with particle dimensio n was observed for\nthe first time in the experiment with gold nanorods [34]. The size depe ndence of the dipole\nplasmon damping rate was deduced from the homogeneous linewidth Γ = 2¯h/Tof the maxima\nin the scattered intensities both for nanorods with various aspect ratios and spherical nanopar-\nticles with radii in the range from 10 to 75 nm. For nanorods, (dipole) plasmon damping rates\ndecreased significantly for lower dipole plasmon oscillation frequencie s. However, the decrease\nin the damping rate was not ascribed to the radiative processes. Th e authors of explain this\n11effect as reduction nonradiative plasmon decay by the fact that int erband excitations in gold\nrequire a threshold energy of about 1 .8eVand expect, that suppression of the damping rate\nin such mechanism would also be present in gold spheres, but for plasm on resonance energies\nbelow 1.8eV. Suppression of the plasmon damping rates in spheres had not been observed\nexperimentally, as far as we know; the conclusive experimental dat a are limited to the plas-\nmon dipole only, while according to our study the effect manifests for l >1. Importance of\nthe threshold energy at 1 ,8eVin gold and exclusion of the radiative damping as a reducing\nmechanism was not confirmed by our study. On the contrary, the r eduction of multipolar SP\nnonradiative damping rates results from competition between radia tive damping γrad\nl(R) and\nall other damping processes included in γnrad(R) and is present in plasmonic particles of any\nmaterial.\n6. Quality factor of multipolar plasmon resonances\nEnhancement of optical response in resonance is usually described by the quality factor\nQ, defined as the product of the resonance center frequency and the bandwidth. In case of\nplasmonic particles the quality factor is interpreted as a measure of the local field enhancement\n[34] and is expected to define the effective susceptibility in nonlinear o ptical processes [36].\n/s48/s44/s48 /s48/s44/s56 /s49/s44/s54 /s50/s44/s52/s48/s44/s48/s48/s44/s50/s48/s44/s52/s48/s44/s54\n/s48/s44/s48 /s48/s44/s56 /s49/s44/s54 /s50/s44/s52/s48/s49/s48/s50/s48/s51/s48/s52/s48/s50/s44/s53 /s50/s44/s54 /s50/s44/s55/s48/s44/s48/s51/s50/s48/s44/s48/s51/s52/s48/s44/s48/s51/s54/s48/s44/s48/s51/s56/s48/s44/s48/s52/s48\n/s68 /s48/s32/s45/s32/s50\n/s112/s47/s40/s50\n/s32/s45/s32/s105 /s41/s32/s32/s32/s32 /s32/s108/s32/s61/s32/s49\n/s32/s108/s32/s61/s32/s50\n/s32/s108/s32/s61/s32/s51\n/s32/s108/s32/s61/s32/s52\n/s32/s108/s32/s61/s32/s53\n/s32/s108/s32/s61/s32/s54\n/s32/s108/s32/s61/s32/s55\n/s32/s108/s32/s61/s32/s56\n/s32/s108/s32/s61/s32/s57\n/s32/s108/s32/s61/s32/s49/s48/s110\n/s111/s117/s116/s61 /s49/s124\n/s108/s39/s39/s124/s32/s91/s101/s86/s93\n/s32/s84\n/s108/s32/s91/s102/s115/s93\n/s49/s56/s44/s51\n/s49/s57/s44/s52\n/s50/s48/s44/s54\n/s108/s39/s32/s91/s101/s86/s93\n/s108/s39/s32/s91/s101/s86/s93/s81\n/s108/s61\n/s108/s39/s47/s40/s50/s124\n/s108/s39/s39/s124/s41/s32\n/s50/s44/s53 /s50/s44/s54 /s50/s44/s55/s51/s48/s51/s51/s51/s54/s51/s57/s52/s50/s32/s97/s41\n/s98/s41\nFigure 5: a) Damping rates |ω′′\nl|vsω′\nlfor successive radii Rand b) quality factors Ql=ω′\nl(R)/(2|ω′′\nl(R)|) for\nmultipolar plasmon modes of gold nanoparticles ( nout= 1, surface scattering neglected). Graphs on the right\nare magnifications of circled regions of the graphs on the left.\nThe dependence of the ω′′\nlvsω′\nlfor successive radii Rand the resulting quality factors\n12Ql(R) =ω′\nl(R)/(2|ω′′\nl(R)|) for multipolar plasmon modes of gold nanoparticles ( nout= 1) are\npresented in Figures 5 (surface scattering neglected) and 6 (sur face scattering included). The\ngraphs on the right are the magnifications of the graphes (circled r egions) on the left.\n/s48/s44/s48 /s48/s44/s56 /s49/s44/s54 /s50/s44/s52/s48/s44/s48/s48/s44/s50/s48/s44/s52/s48/s44/s54\n/s48/s44/s48 /s48/s44/s56 /s49/s44/s54 /s50/s44/s52/s48/s49/s48/s50/s48/s51/s48/s52/s48/s50/s44/s52 /s50/s44/s53 /s50/s44/s54 /s50/s44/s55/s48/s44/s48/s53/s48/s44/s49/s48/s48/s44/s49/s53/s48/s44/s50/s48/s124\n/s108/s39/s39/s124/s32/s91/s101/s86/s93\n/s32/s84\n/s108/s32/s91/s102/s115/s93\n/s52/s44/s52\n/s54/s44/s54\n/s49/s51/s44/s50/s110\n/s111/s117/s116/s61 /s49\n/s40/s82 /s41 /s48/s32/s45/s32/s50\n/s112/s47/s40/s50\n/s32/s45/s32/s105\n/s82/s40/s82 /s41 /s41/s32\n/s108/s39/s32/s91/s101/s86/s93\n/s108/s39/s32/s91/s101/s86/s93/s81\n/s108/s61\n/s108/s39/s47/s40/s50/s124\n/s108/s39/s39/s124/s41/s32\n/s32/s108/s32/s61/s32/s49\n/s32/s108/s32/s61/s32/s50\n/s32/s108/s32/s61/s32/s51\n/s32/s108/s32/s61/s32/s52\n/s32/s108/s32/s61/s32/s53\n/s50/s44/s52 /s50/s44/s53 /s50/s44/s54 /s50/s44/s55/s49/s48/s50/s48/s51/s48/s52/s48/s32/s98/s41/s97/s41\nFigure 6: a) Damping rates |ω′′\nl|vsω′\nlfor successive radii Rand b) quality factors Ql=ω′\nl(R)/(2|ω′′\nl(R)|) for\nmultipolar plasmon modes of gold nanoparticles ( nout= 1, surface scattering included). Graphs on the right are\nmagnifications of circled regions of the graphs on the left.\nFigure 5a) illustrates the effect of the total plasmon damping rate r eduction to the value\nbellowγnrad= 0,032eVfor modes with l>1. Figure 5b) illustrates the resulting quality factors\nQlfor successive SP modes which we use as a measure of energy store d in the SP oscillation of\nplasmon mode lat the resonance frequency ω′\nl(R). In some ranges of SP resonance frequencies,\nthe higher polarity plasmon modes ( l>1) are more efficient in storing the SP oscillation energy.\nThe similar conclusion comes from Figure 6b) (interface damping includ ed). Quality factor of\nthe dipole plasmon resonance reaches maximum value Ql=1≈29 forω′\nl(R)≈2,606eVfor gold\nnanoparticle of the radius R≈16nm(see Figure 7) and is the fast decreasing function of size\nfor both smaller and larger particles. If a gold particle is embedded in a medium of higher\noptical density (see Figure 9, nout= 1,5), the maximum quality factor of the dipole plasmon\nresonance is even smaller, Ql=1≈26. Such value corresponds to the resonance frequency of a\ngold particle of radius R≈11nm.\n13/s48 /s49/s48/s48 /s50/s48/s48 /s51/s48/s48 /s52/s48/s48 /s53/s48/s48/s48/s49/s48/s50/s48/s51/s48/s52/s48\n/s32/s108/s32/s61/s32/s49\n/s32/s108/s32/s61/s32/s50\n/s32/s108/s32/s61/s32/s51\n/s32/s108/s32/s61/s32/s52\n/s32/s108/s32/s61/s32/s53/s40/s82 /s41 /s48/s32/s45/s32/s50\n/s112/s47/s40/s50\n/s32/s45/s32/s105\n/s82/s40/s82 /s41 /s41/s32/s81\n/s108/s61\n/s108/s39/s47/s40/s50/s124\n/s108/s39/s39/s124/s41/s32\n/s82 /s32/s91/s110/s109/s93\nFigure 7: Quality factors Qlvs radius Rfor multipolar plasmon modes of gold nanoparticles ( nout= 1, surface\nscattering included).\n7. Dipole plasmon damping rates vs size: comparison with exp erimental results for\nsilver and gold nanoparticles [34, 37]\nExperimental investigations of spectral properties of plasmonic p articles are usually per-\nformed on large particle ensembles, where inevitable variations in size , shape and surface prop-\nerties tend to mask the spectral properties of the individual part icles. The effects of surface\nchemistry (especially in Ag particles) and uncertainties in sample para meters affect linewidths\nand maxima positions which define damping times and resonance frequ encies of SP. We have\nchosen two experiments [34, 37], which provide the spectroscopic d ata for individual spherical\nparticles of silver and gold. The experiments were performed in well-c ontrolled particle environ-\nments as a function of size for nanoparticles up to R= 25nmfor silica-coated silver (Ag@SiO 2)\nand for gold nanoparticles immersed in a index matching fluid ( nout= 1.5) for relatively large\nrange of radii (up to R= 75nm). However, well resolved data was reported only for the dipole\nplasmon resonance. In Figures8 and9 we compare theexperimenta l results presented in[34, 37]\nwith the results of our modeling.\nFigure 8 illustrates the spectral width Γ measured in different single A g@SiO 2nanoparticles\nvs the inverse of their equivalent diameter Deq, optically determined by fitting the extinction\nspectra (see Figure 4 in [37]). Dashed line represents γR((2R)−1) dependence (Equation (3) for\nA= 0.7 andγbulk= 0.125eV). These parameters [37] result from a linear fit to the experimenta l\ndata. Solid line results from our modeling in the extended range of size s forωp= 9.10eV,\n14/s65/s103 /s32/s32/s32 /s110\n/s111/s117/s116/s32/s61/s32/s49/s44/s52/s53/s32/s40/s83/s105/s79\n/s50/s41\n/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32 /s32/s32/s32/s32 /s115/s112/s101/s99/s116/s114/s97/s108/s32/s119/s105/s100/s116/s104/s32 /s32/s91/s51/s55/s93\n/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32 /s32/s32\n/s108/s32/s61/s49/s50 /s82 /s61/s32/s50/s124\n/s108/s32/s61/s49/s39/s39/s40 /s82 /s41/s124/s32/s40/s116/s104/s105/s115/s32/s119/s111/s114/s107/s41 /s32\n/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32/s32 /s32/s32/s32/s32\n/s82/s40/s82 /s41/s32/s61/s32\n/s98/s117/s108/s107/s43 /s50 /s65/s118\n/s70/s47/s50 /s82\n/s48/s44/s48/s48 /s48/s44/s48/s50 /s48/s44/s48/s52 /s48/s44/s48/s54 /s48/s44/s48/s56 /s48/s44/s49/s48 /s48/s44/s49/s50/s48/s44/s48/s48/s48/s44/s50/s53/s48/s44/s53/s48/s48/s44/s55/s53/s49/s44/s48/s48/s49/s44/s50/s53/s32/s91/s101/s86/s93\n/s40/s50 /s82 /s41/s45/s49\n/s32/s91/s110/s109/s45/s49\n/s93\nFigure 8: Spectral width Γ of plasmon resonances measured [37] in single spherical Ag@SiO 2particles vs the\ninverse of their equivalent diameter. Dashed line represents γR((2R)−1) dependence (equation (3)), with the\nparameters A= 0.7 andγbulk= 0.125eVresulting from linear fit the line to the experimental data [37]. Solid\nline results from our model with the parameters ωp= 9.10,ε0= 4eV, γbulk= 0.125eVandA= 0.7.\nε0= 4eV,γbulk= 0.125eVandA= 0.7. Our model with such input parameters reproduces the\nexperimental data for smallest particles perfectly and describes t he departure of ωl=1((2R)−1)\ndependence from linear. The ωl(R) model functions describe consistently the size dependence\nof the SP damping rates for experimentally available particles and pre dicts damping rates (and\nthe resulting damping times) for larger sizes (experimentally unavaila ble so for) and for higher\nplasmon multipolarities.\nFigure 9a) illustrates spectral widths Γ vs resonance energy ω′\nl=1(R) measured for different\nsingle Au nanoparticles (up to 2 R= 150nm) (Figure 4 of [34]). The experimental data seems\nto suggest that Γ decays linearly with the SP resonance frequency . However, it is not the case.\nThe dependence of Γ l=1(ωl=1(R)) = 2|ω′′\nl=1(R)|(line in figure 9a)) reproduces the experimental\ndata and predicts the decrease in Γ l=1for particles larger than those studied experimentally.\nSuch decrease in damping rate |ω′′\nl=1(R)|of the dipole SP is accompanied by the increase in\nthe damping rate of the quadrupole |ω′′\nl=2(R)|SP and of the following higher polarity SPs, as\ndiscussed in Section 5 (see Figure 2b)).\nFigure 9b) illustrates the quality factor of the dipole resonance Ql=1=ω′\nl=1(R)/2|ω′′\nl=1(R)|\nvs resonance energy ω′\nl=1(R) (solid lines). Our modeling shows that interface damping (dashed\nline) substantially reduces the quality factor for particles from the smallest size range. The\nmaximum factor Ql=1≈26 is found in nanoparticles of radii R= 12nm. The agreement with\n15the experimental data of [34] is very good.\n/s48/s44/s53 /s49/s44/s48 /s49/s44/s53 /s50/s44/s48 /s50/s44/s53/s48/s44/s48/s48/s44/s50/s48/s44/s52/s48/s44/s54/s48/s44/s56/s49/s44/s48\n/s48/s44/s53 /s49/s44/s48 /s49/s44/s53 /s50/s44/s48 /s50/s44/s53/s48/s49/s48/s50/s48/s51/s48/s52/s48/s32/s91/s101/s86/s93\n/s32\n/s108/s61/s49/s32/s61/s32/s50/s124 \n/s108/s61/s49/s34 /s124/s32/s40 /s116/s104/s105/s115/s32/s119/s111/s114/s107 /s44/s32\n/s98/s117/s108/s107/s41\n/s32\n/s108/s61/s49/s32/s61/s32/s50/s124 \n/s108/s61/s49/s34 /s124/s32/s40 /s116/s104/s105/s115/s32/s119/s111/s114/s107 /s44/s32\n/s98/s117/s108/s107/s82 /s41/s41\n/s32/s49/s48/s32/s110/s109\n/s32/s50/s48/s32/s110/s109\n/s32/s51/s48/s32/s110/s109\n/s32/s52/s48/s32/s110/s109\n/s32/s53/s48/s32/s110/s109\n/s32/s55/s53/s32/s110/s109/s65/s117\n/s110\n/s111/s117 /s116/s61/s49/s46/s53/s97/s41\n/s98/s41\n/s81/s117/s97/s108/s105/s116/s121/s32/s102/s97/s99/s116/s111/s114\n/s108/s61 /s49/s39/s40 /s82 /s41/s32/s91/s101/s86/s93\nFigure 9: a) Spectral widths Γ of plasmon resonances vs resonanc e energy measured [34] in single spheri-\ncal Au particles. b) Quality factor resulting from the experimental data. Lines represent the spectral width\nΓl=1(ωl=1(R)) = 2|ω′′\nl=1(R)|and the quality factor Ql=1=ω′\nl=1(R)/2|ω′′\nl=1(R)|vsω′\nl=1(R) for the dipole mode\nobtained from our modeling with surface damping neglected (dashed line) and included (solid line).\n8. Conclusions\nThedependence oftheSPresonancefrequencies ω′\nl(R)anddampingrates |ω′′\nl(R)|onparticle\nradius determine the intrinsic optical properties of plasmonic spher es (illuminated or not). Our\nstudy provides direct, accurate size characteristics for a broad range of particle radii (up to\nR= 1000nm) and plasmon polarities (up to l= 10). At present, this exceeds the range\nexperimentally explored.\nSize dependence of SP damping rates |ω′′\nl(R)|allows to distinguish the size ranges in which\nefficient transfer of radiation energy into heat takes place (large c ontribution of the nonradiative\ndecay) and those in which the particles are effective radiating anten nas (dominant contribution\nof the radiative damping). As long as the contribution of the radiativ e decay is negligible, the\nparticle is not able to couple to the incoming field effectively and has wea k radiative abilities.\nIf the contribution of the radiative damping prevails, the particle is a ble to emit light within\n16plasmonic mechanism efficiently. Effective radiating enables efficient int eraction of SP near-\nfield with other structures at a desired resonance frequency ω=ω′\nl(R). The knowledge of\nsize dependence of both: the multipolar SP resonance frequencies ω′\nl(R) and the corresponding\ndamping rates |ω′′\nl(R)|= ¯h/Tl(R) is indispensable to shape the particle plasmonic features\neffectively.\nOur study, extended toward large particle sizes and plasmon multipo larities, revealed new\nfeatures of the total plasmon damping rates. In certain ranges o f radii, reduction of the multi-\npolar SP damping rates, as compared with its small size limit, takes plac e. The small size limit\nis equal to the nonradiative damping rate resulting from absorption and heat dissipation. The\nsuppression of nonradiative damping is not limited to the size range, in which the effect was\ndirectly demonstrated. It is present when the radiative damping br ings the dominant contribu-\ntion to the total plasmon damping. The reduction of γnrad(R) with the growing contribution of\nradiation damping is revealed in the absorption spectra of particles; absorptive abilities of large\nparticles are poor. This fact is described by solutions of Mie scatter ing theory, but its physical\nbackground have not been explained, as far as we know. Our study led us to conclusion, that\nthe reduction of multipolar SP nonradiative damping rates results fr om competition between\nradiative damping and all other damping processes included in γnrad(R). As far as we know,\nsuch hypothesis has not been proposed before. Our study provid es a new starting point for\nbetter understanding of rules that define contributions of plasmo n radiative and nonradiative\ndecay ratesto thetotalSP decay rateasafunction ofparticle siz e. The independent modeling is\nnecessary for better understanding of the role of SP radiative an d nonradiative decay channels.\nThe proposed SP characteristics vs particle size provide a consiste nt, uniform description\nof the experimental SP damping rates measured for single spherica l particles in [34] and [37].\nNot only dipole damping rates (found experimentally) but also multipole damping rates (and\nresulting damping times) in a broad range of sizes can be predicted.\nThe derived size dependence of plasmon decay rates |ω′′\nl(R)|at resonance frequencies ω′\nl(R\n), and the quality factors Ql(R) of SP multipolar modes define not only the spectral scattering\nabilities of the plasmonic spheres, but also reflect changes in the str ength of coupling of SP\nmodeswiththe external field. Such characteristics canserve asa toolforcontrolling thespectral\nfeatures of plasmonic nanospheres in technological or diagnostic a pplications, by optimizing the\nsize at the desired resonance frequency ω=ω′\nl(R).\nAcknowledgement We would like to acknowledge the financial support of this work by the\nMinistry of Science and Higher Education (No N N202 126837).\n17References\n[1] U. Kreibig, M. Vollmer, Optical Properties of Metal Clusters (Spr inger Series in Material\nScience), Vol. 25, Berlin: Springer, 1995.\n[2] K.L.Kelly, E.Coronado,L.L.Zhao,G.C.Schatz, Theopticalp ropertiesofmetalnanopar-\nticles: the influence of size, shape, and dielectric environment, J. P hys. Chem. B 107 (2003)\n668 – 677.\n[3] J. Z. Zhang, C. Noguez, Plasmonic optical properties and applica tions of metal nanostruc-\ntures, Plasmonics 3 (2008) 127 – 150.\n[4] K. A.Willets, R. P. V. Duyne, Localized surface plasmon resonance spectroscopy and sens-\ning, Ann. Rev. of Phys. Chem. 58 (2007) 267 – 297.\n[5] C. Noguez, Surface plasmons on metal nanoparticles: The influe nce of shape and physical\nenvironment, J. Phys. Chem. C 111 (2007) 3806 – 3819.\n[6] H. G. A Otto, I Mrozek, W. Akemann, Surface-enhanced raman scattering, J. Phys.: Con-\ndens, Matter 4 (1992) 1143 – 1212.\n[7] J. J. Mock, M. Barbic, D. R. Smith, D. A. Schultz, S. Schultz, Sha pe effects in plasmon\nresonance of individual colloidal silver nanoparticles, J. Chem. Phys . 116 (2002)6755–6759.\n[8] S.A.Maier, H.A.Atwater, Plasmonics: Localizationandguidingof electromagneticenergy\nin metal/dielectric structures, J. Appl. Phys. 98 (2005) 011101.\n[9] N. Fang, H. Lee, C. Sun, X. Zhang, Subdiffraction-limited optical imaging with a silver\nsuperlens, Science 308 (2005) 534 – 537.\n[10] B. P. Rand, P. Peumans, S. R. Forrest, Long-range absorpt ion enhancement in organic\ntandem thin-film solar cells containing silver nanoclusters, J. App.Phy s. 96 (2004) 7519 –\n7526.\n[11] K. Catchpole, A. Polman, Plasmonic solar cells, Optics Express 16 (2008) 21793 – 21800.\n[12] S. Pillai, K. R. Catchpole, T. Trupke, M. A. Green, Surface plasm on enhanced silicon solar\ncells, J. of Appl. Phys. 101 (2007) 093105.\n[13] K.Nakayama, K.Tanabe, H.A.Atwater, Plasmonicnanopartic leenhancedlightabsorption\nin gaas solar cells, App. Phys. Lett. 93 (2008) 121904.\n18[14] A. J. Morfa, K. L. Rowlen, T. H. R. III, M. J. Romero, J. van de Lagemaat, Plasmon-\nenhanced solar energy conversion in organicbulk heterojunction p hotovoltaics, Appl. Phys.\nLett. 92 (2008) 013504.\n[15] S.-S. Kim, S.-I.Na, J. Jo, D.-Y.Kim, Y.-C.Nah, Plasmon enhanced performanceoforganic\nsolar cells using electrodeposited ag nanoparticles, Appl. Phys. Let t. 93 (2008) 073307.\n[16] M. Quinten, A. Leitner, J. R. Krenn, F. R. Aussenegg, Electro magnetic energy transport\nvia linear chains of silver nanoparticles, Opt. Lett. 23 (1998) 1331– 1333.\n[17] S.A.Maier, M.L.Brongersma, P.G.Kik, H.A.Atwater, Obser vationofnear-fieldcoupling\nin metal nanoparticle chains using far-field polarization spectrosco py, Phys. Rev. B 65\n(2002) 193408.\n[18] R. M. Dickson, L. A. Lyon, Unidirectional plasmon propagation in metallic nanowires, J.\nPhys. Chem. B 104 (2000) 6095 – 6098.\n[19] M. L. Brongersma, J. W. Hartman, H. A. Atwater, Electromag netic energy transfer and\nswitching in nanoparticle chain arrays below the diffraction limit, Phys. Rev. B 62 (2000)\nR16356 – R16359.\n[20] J. R. Krenn, A. Dereux, J. C. Weeber, E. Bourillot, Y. Lacrout e, J. P. Goudonnet,\nG. Schider, W. Gotschy, A. Leitner, F. R. Aussenegg, C. Girard, S queezing the optical\nnear-field zone by plasmon coupling of metallic nanoparticles, Phys. R ev. Lett. 82 (12)\n(1999) 2590 – 2593.\n[21] B. Liedberg, I. Lundstrom, E. Stenberg, Principles of biosens ing with an extended coupling\nmatrix and surface plasmon resonance, Sensors and Actuators B : Chemical 11 (1993) 63 –\n72.\n[22] K. Aslan, J. R. Lakowicz, C. D. Geddes, Plasmon light scattering in biology and medicine:\nnew sensing approaches, visions and perspectives, Current Opinio n in Chemical Biology 9\n(2005) 538 – 544.\n[23] A. Derkachova, K. Kolwas, Size dependence of multipolar plasmo n resonance frequencies\nand damping rates in simple metal spherical nanoparticles, Eur. J. P hys. ST 144 (2007) 93\n– 99.\n[24] F. Tam, A. L. Chen, J. Kundu, H. Wang, N. J. Halas, Mesoscopic nanoshells: Geometry-\ndependent plasmon resonances beyond the quasistatic limit, J. Che m. Phys. 127 (2007)\n204703.\n19[25] S. Link, M. El-Sayed, Shape and size dependence of radiative, n on-radiative and photother-\nmal properties of gold nanocrystals, International Reviews in Phy sical Chemistry 19 (3)\n(2000) 409–453.\n[26] K. Kolwas, A. Derkachova, M. Shopa, Size characteristics of s urface plasmons and their\nmanifestationinscattering propertiesofmetalparticles, J.Q.S.R.T. 110(2009)1490–1501.\n[27] K. Kolwas, A. Derkachova, Plasmonic abilities of gold and silver sph erical nanoantennas\nin terms of size dependent multipolar resonance frequencies and pla smon damping rates,\nOpto-Electr. Rev. 18 (2010) 429 – 437.\n[28] J.R.Krenn, G.Schider, W.Rechberger, B.Lamprecht, A.Le itner, F.R.Aussenegg, Design\nof multipolar plasmon excitations in silver nanoparticles, Appl. Phys. L ett. 77 (2000) 3379\n– 3381.\n[29] K. L. Shuford, M. A. Ratner, G. C. Schatz, Multipolar excitatio n in triangular nanoprisms,\nJ. Chem. Phys. 123 (2005) 114713.\n[30] E. K. Payne, K. L. Shuford, S. Park, G. C. Schatz, C. A. Mirkin , Multipole plasmon\nresonances in gold nanorods, J. Phys. Chem. B 110 (2006) 2150 – 2 154.\n[31] K. Kolwas, S. Demianiuk, M. Kolwas, Dipole and quadrupole plasmon resonances in large\nsodium clusters observed in scattered light, J. Chem. Phys. 106 (1 997) 8436 – 8441.\n[32] C. S¨ onnichsen, T. Franzl, T. Wilk, G. von Plessen, J. Feldmann, Plasmon resonances in\nlarge noble-metal clusters, New J. Phys. 4 (2002) 93.1–93.8.\n[33] R. Krahne, G. Morello, A. Figuerola, C. George, S. Deka, L. Man na, Physical properties\nof elongated inorganic nanoparticles, Phys. Rep. 501 (2011) 75 – 2 21.\n[34] C. S¨ onnichsen, T. Franzl, T. Wilk, G. von Plessen, J. Feldmann, O. Wilson, P. Mulvaney,\nDrasticreduction of plasmon damping ingoldnanorods, Phys. Rev. L ett. 88 (2002)077402.\n[35] S. Link, M. A. El-Sayed, Size and temperature dependence of t he plasmon absorption of\ncolloidal gold nanoparticles, J. Phys. Chem. B 103 (1999) 4212 – 421 7.\n[36] V. Shalaev, E. Poliakov, V. Markel, Small-particle composites. ii. n onlinear optical prop-\nerties, Physical Review B 53 (5) (1996) 2437.\n[37] H. Baida, P. Billaud, S. Marhaba, D. Christofilos, E. Cottancin, A . Crut, J. Lerm´ e,\nP. Maioli, M. Pellarin, M. Broyer, et al., Quantitative determination of t he size depen-\ndence of surface plasmon resonance damping in single ag@sio2 nanop articles, Nano Letters\n9 (10) (2009) 3463–3469.\n20[38] P. K. Jain, X. Huang, I. H. El-Sayed, M. A. El-Sayed, Review of s ome interesting surface\nplasmon resonance-enhanced properties of noble metal nanopar ticles and their applications\nto biosystems, Plasmonics 2 (2007) 107 – 118.\n[39] H. H¨ ovel, S. Fritz, A. Hilger, U. Kreibig, M. Vollmer, Width of clust er plasmon resonances:\nBulk dielectric functions and chemical interface damping, Physical R eview B 48 (24) (1993)\n18178.\n[40] S. Link, M. El-Sayed, Spectral properties and relaxation dyna mics of surface plasmon\nelectronic oscillations in gold and silver nanodots and nanorods, The J ournal of Physical\nChemistry B 103 (40) (1999) 8410–8426.\n[41] D. E. Charles, M. Gara, D. Aherne, D. M. Ledwith, J. M. Kelly, W. J. Blau, M. E.\nBrennan-Fournet, Scaling of surface plasmon resonances in trian gular silver nanoplate sols\nfor enhanced refractive index sensing, Plasmonics 6 (2011) 351 36 2.\n[42] M. Born, E. Wolf, Principles of Optics, Oxford: Pergamon, 1975 .\n[43] C. F. Bohren, D. R. Huffmann, Absorption and scattering of ligh t by small particles, New\nYork: Wiley-Interscience, 1983.\n[44] J. A. Stratton, Electromagnetic Theory, Intarnational Ser ias in Physics, McGRAW-HILL\nBook Company, Inc., 1941.\n[45] M. I. Mishchenko, L. D. Travis, A. A. Lacis, Scattering, Absor ption and Emission of Light\nby Small Particles, Cambridge University Press: Cambridge, 2002.\n[46] T. Klar, M. Perner, S. Grosse, G. Von Plessen, W. Spirkl, J. Feld mann, Surface-plasmon\nresonances in single metallic nanoparticles, Physical Review Letters 80 (19) (1998) 4249–\n4252.\n[47] E.Heilweil, R.Hochstrasser, Nonlinearspectroscopyandpicos econdtransientgratingstudy\nof colloidal gold, The Journal of chemical physics 82 (11) (1985) 47 62.\n[48] Y. Liau, A. Unterreiner, Q. Chang, N. Scherer, Ultrafast dep hasing of single nanoparticles\nstudied by two-pulse second-order interferometry, The Journa l of Physical Chemistry B\n105 (11) (2001) 2135–2142.\n[49] R. Rupin, Electromagnetic Surface Modes, Wiley: Chichester, 1 982.\n[50] R. Fuchs, P. Halevi, Basic Concepts and Formalism of Spatial Disp ertion, in Spatial Dis-\npertion in Solids and Plasmas, North-Holland, 1992.\n21[51] P. G. Etchegoin, E. C. L. Ru, M. Meyer, An analytic model for th e optical properties of\ngold, J. Chem. Phys. 125 (2006) 164705.\n[52] K.-S. Lee, M. A. El-Sayed, Goldandsilver nanoparticles insensing andimaging: sensitivity\nof plasmon response of size shape and metal composition, J. Phys. Chem. B 110 (2006)\n19220 – 19225.\n[53] C. Oubre, P. Nordlander, Optical properties of metallodielectr ic nanostructures calculated\nusing thefinitedifference timedomainmethod, J.Phys. Chem. B 108( 2004)17740–17747.\n[54] C. Noguez, C. E. Roman-Velazquez, Dispersive force between dissimilar materials: geomet-\nrical effects, Phys. Rev. B 70 (2004) 195412.\n[55] M. A. Ordal, R. J. Bell, R. W. Alexander, Jr, L. L. Long, M. R. Que rry, Optical properties\nof fourteen metals in the infrared and far infrared: Al, co, cu, au, fe, pb, mo, ni, pd, pt,\nag, ti, v, and w, Appl. Opt. 24 (1985) 4493 – 4499.\n[56] M. Quinten, Optical constants of gold and silver clusters in the s pectral range between 1.5\nev and 4.5 ev, Zeitschrift f¨ ur Physik B Condensed Matter 101 (2) ( 1996) 211–217.\n[57] C.Novo, D.Gomez, J.Perez-Juste, Z.Zhang, H.Petrova, M .Reismann, P.Mulvaney, G.V.\nHartland, Contributions from radiation damping and surface scatt ering to the linewidth\nof the longitudinal plasmon band of gold nanorods: a single particle st udy, Phys. Chem.\nChem. Phys. 8 (2006) 3540 3546.\n[58] M. Hu, C. Novo, A. Funston, H. Wang, H. Staleva, S. Zou, P. Mu lvaney, Y. Xiae, G. V.\nHartland, Dark-field microscopy studies of single metal nanopartic les: understanding the\nfactors that influence the linewidth of the localized surface plasmon resonance, J. Mater.\nChem. 18 (2008) 1949 1960.\n[59] A. D. Boardman, B. V. Paranjape, The optical surface modes of metal spheres, J. Phys.\nF: Metal Phys. 7.\n[60] S. Peng, J. M. McMahon, G. C. Schatz, S. K. Gray, Y. Sun, Rev ersing the size-dependence\nof surface plasmon resonances, PNAS 107 (2010) 14530 14534.\n[61] E. J.Heilweil, R. M.Hochstrasser, Nonlinear spectroscopy and picosecond transient grating\nstudy of colloidal gold, J. Chem. Phys. 82 (1985) 4762 – 4770.\n[62] B. Lamprecht, J. Krenn, A. Leitner, F. Aussenegg, Particle- plasmon decay-time determi-\nnation bymeasuring the optical near-fields autocorrelation: influe nce of inhomogeneous line\nbroadening, Appl. Phys. B 69 (1999) 223 227.\n22[63] F. Stietz, J. Bosbach, T. Wenzel, T. Vartanyan, A. Goldmann, F. Tr¨ ager, Decay times\nof surface plasmon excitation in metal nanoparticles by persistent spectral hole burning,\nPhys. Rev. Lett. 84 (2000) 5644 – 5647.\n[64] A. Melikyan, H. Minassian, Onsurfaceplasmon damping inmetallic na noparticles, Applied\nPhysics B: Lasers and Optics 78 (3) (2004) 453–455.\n23" }, { "title": "0910.4560v2.Free_energy_of_Lorentz_violating_QED_at_high_temperature.pdf", "content": "arXiv:0910.4560v2 [hep-th] 25 Feb 2010Free energy of Lorentz-violating QED at high temperature\nM. Gomes,1T. Mariz,2J. R. Nascimento,3A. Yu. Petrov,3A. F. Santos,3and A. J. da Silva1\n1Instituto de F´ ısica, Universidade de S˜ ao Paulo\nCaixa Postal 66318, 05315-970, S˜ ao Paulo, SP, Brazil∗\n2Instituto de F´ ısica, Universidade Federal de Alagoas, 5707 2-270, Macei´ o, Alagoas, Brazil†\n3Departamento de F´ ısica, Universidade Federal da Para´ ıba\nCaixa Postal 5008, 58051-970, Jo˜ ao Pessoa, Para´ ıba, Brazi l‡\nAbstract\nIn this paper we study the one- and two-loop contribution to the fr ee energy in QED with the Lorentz\nsymmetry breakingintroduced via constantCPT-evenLorentz-b reakingparametersat the high temperature\nlimit. We find the impact of the Lorentz-violating term for the free en ergy and carry out a numerical\nestimation for the Lorentz-breaking parameter.\n∗Electronic address: mgomes,ajsilva@fma.if.usp.br\n†Electronic address: tmariz@if.ufal.br\n‡Electronic address: alesandro,jroberto,petrov@fisica.u fpb.br\n1I. INTRODUCTION\nNowadays the Lorentz symmetry breaking is treated as an impo rtant ingredient of field theory\nmodels for probing quantum gravity phenomena. In fact, Lore ntz-violating theories have been\nstudied in various contexts, such as string theory [1], nonc ommutative field theory [2], and, more\nrecently, Horava-Lifshitz gravity [3].\nMany of the effects observed for the Lorentz-breaking field the ories at zero temperature are\nknown to persist also at finite temperatures. For example, th e dependence of the loop corrections\non the regularization scheme in the finite temperature case w as shown to arise in the one-loop order\nin the Lorentz-breaking quantum electrodynamics (QED) [4] . Analogous situation takes place in\nthe Lorentz-breaking Yang-Mills theory [5]. Different issue s related to the Lorentz-violating QED\nin the finite temperature case were considered in [6]. In [7] t he finite temperature properties in the\nCPT-odd Lorentz-breaking extension of QED for a purely spac elike background were studied, and\nin [8] these properties were analysed for the CPT-even Loren tz-breaking extension.\nMost of the previous studies of the Lorentz-breaking theori es, including those described in [4, 5],\nwere based on the use of couplings which violate not only Lore ntz symmetry but also the CPT\nsymmetry, although CPT-even Lorentz breaking interaction s are also possibleand certainly require\nmore detailed study. In this paper, we will see that, at the hi gh temperature regime, the linear\ncontribution in the Lorentz-breaking parameter may arise.\nThe aim of this paper is the study of one and two-loop correcti ons to the free energy (second\norder in the coupling constant e) of QED in the presence of Lorentz breaking terms, at finite\ntemperature. As it is known thefree energy provides an impor tant information to different physical\nissues such as plasma behavior, solar interior and Big Bang n ucleosynthesis (BBN) (see, [9–11]). In\nthis context, some aspects of thecorrections to the freeene rgy in theories without Lorentz-breaking\nare discussed in [12, 13].\nFor estimating bounds of the Lorentz-breaking parameter, w e use information of the BBN (for\na review, see [14]), which is one of the observational pillar s of the standard cosmology. Note that\nthe Lorentz-violating parameter can explain for the light e lements abundance. In particular, the\ndifference between the theoretical and observational result s can be understood as come from a\ncontribution of the Lorentz-breaking parameter to the prim ordial helium abundance.\nThe structure of the paper is as follows. In Sec. II we present the basic features of the Lorentz-\nviolating QED in the regime of high temperature. In Sec. III w e calculate the one and two-loop\ncontributions to the free energy for QED involving the Loren tz-breaking fermion coupling. In\n2Sec. IV, by using information about the primordial helium ab undance, we obtain a numerical\nestimation for the Lorentz violation parameter. A summary i s presented in Sec. V.\nII. LORENTZ-VIOLATING QED AT HIGH TEMPERATURE\nLet us start by considering the Lagrangian of the Lorentz- an d CPT-violating QED extension\n[15]\nL=−1\n4FµνFµν−1\n4(kF)µνλρFµνFλρ+1\n2(kAF)µǫµνλρAνFλρ+¯ψ(iΓµDµ−M)ψ+Lgf+Lgh,(1)\nwhere Γµ=γµ+Γµ\n1,M=m+M1, with\nΓµ\n1=cµνγν+dµνγ5γν+eµ+ifµγ5+1\n2gλνµσλν (2)\nM1=aµγµ+bµγ5γµ+1\n2Hµνσµν, (3)\nandDµ=∂µ+ieAµ.Lgfis the gauge fixing term and Lghis the ghost field term, which decouples\nfrom the rest of the Lagrangian. The coefficients carrying an o dd (even) number of Lorentz indices\nare CPT-odd (-even).\nThe two Lorentz-violating terms of the photon sector, the Ch ern-Simons-like (CPT-odd) and\n(kF)µνλρFµνFλρ(CPT-even) terms, can be induced by radiative corrections f rom the terms with\nthe coefficients bµandcµνof the fermion sector, respectively, so that ( kAF)µ∝bµ[16] and [17]\n(kF)µνλρ∝1\n2gµλ(cνρ+cρν)+1\n2gνρ(cµλ+cλµ)−1\n2gµρ(cνλ+cλν)−1\n2gνλ(cµρ+cρµ).(4)\nThe coefficients aµ,bµ,Hµν, and (kAF)µhave dimensions of mass, while cµν,dµν,eµ,fµ,\ngλνµ, and (kF)µνλρare dimensionless. In the high temperature regime ( T≫M) the dimensionful\ncoefficients may be neglected. For instance, the bµ-corrections to the free energy are observed to\nbe proportional to b2T2, similarly to the scenario occurring with the corrections s temming from\nthe Chern-Simions-like term [7]. Thus both bµand (kAF)µare negligible at high temperature, as\nwell asaµandHµν.\nAmong the dimensionless coefficients, eµ,fµ, andgλνµare expected to bemuch smaller than the\nother, because their terms cannot be obtained directly from the standard model extension [15] (for\nmore details, see also [18]). Moreover, if we require that th e theory at high temperature is invariant\nunder chiral transformations, these terms are ruled out, si nce{γ5,eµ+ifµγ5+1\n2gλνµσλν} ∝ne}ationslash= 0.\nTherefore, the remaining coefficients are cµν,dµν, and (kF)µνλρ. Now, in order to get a Clifford\nalgebra for Γµ=γµ+cµνγν+dµνγ5γν, we must choose dµν=Q(δµν+cµν), whereQis a constant\n3[19]. With this assumptions, the theory (1) becomes\nLhigh=−1\n4FµνFµν−1\n4(kF)µνλρFµνFλρ+¯ψ[i∂µ(gµν+cµν)˜γν−eAµ(gµν+cµν)˜γν]ψ\n+Lgf+Lgh, (5)\nwhere ˜γµ= (1+Qγ5)γµ.\nWe now assume rotational invariance, such that the coefficien ts in (5) may be reduced to\nproducts of a given unit timelike vector uµ, which describes the preferred frame (see [20], for more\ndetails). Proceeding in this way, we write cµν=κuµuν, and\n(kF)µνλρ= ˜κ(gµλuνuρ+gνρuµuλ−gµρuνuλ−gνλuµuρ), (6)\nsee Eq. (4), where uµ= (1,0,0,0) and now κand ˜κare the coefficients that determine the scale of\nLorentz violation. By choosing α= 1 (Feynman gauge) and ˜ κ= (1+κ\n2)κ, which is convenient to\nkeep the Lagrangian (5) formally covariant, we get\nL=−1\n4˜Fµν˜Fµν+i¯ψ˜∂µ˜γµψ−e¯ψ˜Aµ˜γµψ+1\n2(˜∂µ˜Aµ)2+(˜∂µ¯C)(˜∂µC), (7)\nwhere˜Fµν=˜∂µ˜Aν−˜∂ν˜Aµ, with˜∂µ= ((1+κ)∂0,∂i) and˜Aµ= ((1+κ)A0,Ai). The corresponding\nFeynman rules are shown in Fig. 1.\nFIG. 1: Feynman Rules. Continuous, wavy, and dashed lines repres ent the fermion, photon, and ghost\npropagators, respectively, with momenta ˜ pµ= ((1+κ)p0,pi). The fermion-photon vertex is the usual one,\n−ie˜γµ.\nIII. THE FREE ENERGY\nLet us now compute the free energy per unit of volume (pressur e) as a function of the temper-\natureTand of the coupling constant e, in the regime of high temperature. We shall calculate the\nexpression for the pressure to order e2, which has the form\nP=TlnZ\nV=P0+P2, (8)\nwhereP0is the zero order contribution in the coupling constant and P2is the second order one.\nThe free energy density is minus the expression (24).\n4A. One-loop contribution\nThe lowest-order contributions are given by the three one-l oop vacuum diagrams, displayed in\nFig. 2, and written in the imaginary time formalism as\nP0= tr/summationtext/integraldisplay\n{dp}ln∝ne}ationslash˜p+4(−1\n2)/summationtext/integraldisplay\ndpln˜p2+2(1\n2)/summationtext/integraldisplay\ndpln ˜p2, (9)\nrespectively, where we have introduced the shorthand notat ion\n/summationtext/integraldisplay\n{dp}=T/summationdisplay\np0=2π(n+1\n2)T/integraldisplayd3p\n(2π)3(10)\nfor fermionic loop momenta and\n/summationtext/integraldisplay\ndp=T/summationdisplay\np0=2πnT/integraldisplayd3p\n(2π)3(11)\nfor bosonic loop momenta, and ∝ne}ationslash˜p= ˜pµ˜γµ.\nAs usual, the bosonic contributions of Eq. (9) have four degr ees of freedom for the gauge field\n(Fig. 2(b)) and two degrees of freedom for the ghost field (Fig . 2(c)). The fermionic contribution,\n(a) ( b) ( c)\nFIG. 2: One loop vacuum diagrams.\nafter the calculation of the trace,\nPf\n0= 2/summationtext/integraldisplay\n{dp}ln[(1−Q2)(p2\n0(1+κ)2+p2)], (12)\nhas the same form as the bosonic contributions,\nPb\n0=−/summationtext/integraldisplay\ndpln(p2\n0(1+κ)2+p2). (13)\nIn order to evaluate these expressions we proceed similarly to [21], such that\nP0=/integraldisplayd3p\n(2π)3/bracketleftbigg|p|\n1+κ+4\nβln/parenleftbigg\n1+e−|p|β\n1+κ/parenrightbigg\n−2\nβln/parenleftbigg\n1−e−|p|β\n1+κ/parenrightbigg/bracketrightbigg\n, (14)\nwhich is independent of the parameter Q. The zero temperature (divergent) part is absorbed in a\nrenormalization of the vacuum energy [12]. Therefore, afte r integrating in the angular variables,\nwe get\nP0=2\nπ2β/integraldisplay∞\n0drr2ln/parenleftBig\n1+e−rβ\n1+κ/parenrightBig\n−1\nπ2β/integraldisplay∞\n0drr2ln/parenleftBig\n1−e−rβ\n1+κ/parenrightBig\n, (15)\n5and finally, we obtain\nP0=11π2\n180T4(1+κ)3. (16)\nTherefore we conclude that the only impact of the Lorentz sym metry breaking in the case of\nthe CPT-even Lorentz-breaking parameter consists in the mo dification of a constant factor.\nB. Two-loop contribution\nIn this subsection we study the two-loop contributions to th e free energy. The Lorentz violating\ncontribution to free energy up to the two-loop order (which c orresponds to the second order in e,\nis given by Fig. 3), which can be written as\nFIG. 3: Two-loop diagram.\nP2=1\n2e2/summationtext/integraldisplay\n{dp}/summationtext/integraldisplay\n{dq}tr/bracketleftbigg\n˜γµ1\n∝ne}ationslash˜p˜γµ1\n∝ne}ationslash˜q1\n(˜p+ ˜q)2/bracketrightbigg\n. (17)\nNote that, due to the fact that\n˜γµ1\n˜pα˜γα=γµ1\n˜pαγα, (18)\nthe above second order contribution is also independent of t he parameter Q, and so on for the\nother contributions. Thus, we can rewrite (17) as\nP2=1\n2e2/summationtext/integraldisplay\n{dp}/summationtext/integraldisplay\n{dq}tr/bracketleftbigg\nγµ˜pαγα\n˜p2γµ˜qβγβ\n˜q21\n(˜p+ ˜q)2/bracketrightbigg\n, (19)\nso that, after calculating the trace, we get\nP2= 2e2/summationtext/integraldisplay\n{dp}/summationtext/integraldisplay\n{dq}/bracketleftbigg\n−1\n˜p2˜q2+1\n˜q2(˜p+ ˜q)2+1\n˜p2(˜p+ ˜q)2/bracketrightbigg\n. (20)\nUsing the results for the sum-integrals,\n/summationtext/integraldisplay\n{dp}/summationtext/integraldisplay\n{dq}1\n˜p2˜q2=1\n576T4(1+κ)2, (21)\n/summationtext/integraldisplay\n{dp}/summationtext/integraldisplay\n{dq}1\n˜q2(˜p+ ˜q)2=−1\n288T4(1+κ)2, (22)\n6we arrive at the following Lorentz violating contribution t o the free energy,\nP2=−5\n288e2T4(1+κ)2. (23)\nThus, the modification in the free energy due to the Lorentz-b reaking parameter for the two-loop\norder consists only in multiplying by a constant, just as in t he one-loop order.\nTherefore, the expression for the pressure to order e2in the presence of Lorentz symmetry\nbreaking looks like\nP=π2T4\n9/bracketleftbigg11\n20(1+κ)3−5e2\n32π2(1+κ)2/bracketrightbigg\n. (24)\nIV. NUMERICAL ESTIMATIONS\nIn order to estimate a bound for the Lorentz violating parame terκ, we use the theoretical\npredictions of the primordial helium abundance Ydeveloped in the references [9–11]. A way to\ndetermineYconsists in the analysis of the change in thermodynamics qua ntities such as the energy\ndensityρ, the pressure Pand the neutrino temperature Tν. Using the result (24), let us now study\nthese changes in the presence of the Lorentz-breaking.\nThe contribution to the energy density can be found from the s tandard thermodynamic relation\nρ=−P+T(∂P/∂T), so that\nρ=π2T4\n15(N+δN), (25)\nwhereN=11\n4andδN≈ −0.007+8.236κ. Thus, we obtain\n∆ρ\nρ≈ −2.5×10−3+2.994κ. (26)\nThe fraction ∆ Yis affected by QED in several ways, as can be seen in [10]. The tot al effect is\napproximately\n∆Y≈2.9×10−4+0.15∆Tν\nTν+0.07∆ρ\nρ, (27)\nwhereTνdepends on energy density ρ(for more details, see [22]).\nThe usual theoretical result, without parameter κ, is ∆Y≈10−4, whereas the experimental one\nis ∆Y≈10−3[14]. Therefore, an upper boundfor κnecessary for the coincidence of the theoretical\nand experimental results must be κ∼10−2. This result agrees with the value found in [23, 24].\n7V. SUMMARY\nWe have calculated the contributions to the free energy in th e rotationally invariant Lorentz-\nviolating QED in one- and two-loop approximations at high te mperature. The corresponding\ncorrection to the pressure was then determined.\nWe also observe that Lorentz violation can be used to explain the difference between theoretical\nand experimental predictions of the primordial helium abun dance. By matching these predictions,\nwe have estimated the Lorentz-breaking parameter κ∼10−2, which agrees with that obtained in\n[23].\nAcknowledgements. This work was partially supported by Conselho Nacional de De sen-\nvolvimento Cient´ ıfico e Tecnol´ ogico (CNPq) and Funda¸ c˜ a o de Amparo ` a Pesquisa do Estado de\nS˜ ao Paulo (FAPESP), Coordena¸ c˜ ao de Aperfei¸ coamento do Pessoal do Nivel Superior (CAPES:\nAUX-PE-PROCAD 579/2008) and CNPq/PRONEX/FAPESQ.\n[1] V. A. Kostelecky and S. Samuel, Phys. Rev. D 39, 683 (1989).\n[2] S. M. Carroll, J. A. Harvey, V. A. Kostelecky, C. D. Lane and T. O kamoto, Phys. Rev. Lett. 87, 141601\n(2001) [arXiv:hep-th/0105082].\n[3] P. Horava, Phys. Rev. D 79, 084008 (2009) [arXiv:0901.3775 [hep-th]].\n[4] J. R. Nascimento, E. Passos, A. Yu. Petrov and F. A. Brito, JHE P0706, 016 (2007) [arXiv:0705.1338\n[hep-th]]\n[5] M. Gomes, J. R. Nascimento, E. Passos, A. Yu. Petrov and A. J. da Silva, Phys. Rev. D 76, 047701\n(2007), arXiv:0704.1104.\n[6] L. Cervi, L. Griguolo, D. Seminara, Phys. Rev. D 64, 105003 (2001), [arXiv:hep-th/0104022]; F. A.\nBrito, T.Mariz, J. R.Nascimento, E.PassosandR. F.Ribeiro, JHEP 10, 019(2005), [hep-th/0509008].\n[7] R. Casana, M. M. Ferreira and J. S. Rodrigues, Phys. Rev. D 78, 125013 (2008) [arXiv:0810.0306\n[hep-th]].\n[8] R. Casana, M. M. Ferreira and J. S. Rodrigues, M. R. O. Silva, “Fin ite Temperature CPT-even Elec-\ntrodynamics of the Standard Model Extension”, arXiv:0907.1924 [h ep-th].\n[9] D. A. Dicus, E. W. Kolb, A. M. Gleeson, E. C. G. Sudarshan, V. L. T eplitz, M. S. Turner, Phys. Rev.\nD26, 2694 (1982).\n[10] A. Heckler, Phys. Rev. D 49, 611 (1994).\n[11] R. E. Lopez and M. S. Turner, Phys. Rev. D 59, 103502 (1999).\n[12] J. Kapusta, ”Finite-temperature field theory”, Cambridge Un iv. Press, 1989.\n[13] J. O. Andersen, ”On effective field theories at finite temperatu re”, arXiv:hep-ph/9709331.\n8[14] F. Iocco, G. Mangano, G. Miele, O. Pisanti and P. D. Serpico, Ph ys. Rept. 472, 1 (2009)\n[arXiv:0809.0631 [astro-ph]].\n[15] D. Colladay and V. A. Kosteleck´ y, Phys. Rev. D 55, 6760 (1997) [arXiv:hep-ph/9703464]; Phys. Rev.\nD58, 116002 (1998) [arXiv:hep-ph/9809521].\n[16] R. Jackiw and V. A. Kosteleck´ y, Phys. Rev. Lett. 82, 3572 (1999) [arXiv:hep-ph/9901358].\n[17] V. A. Kosteleck´ y, C. Lane, A. Pickering, Phys. Rev. D 65, 056006 (2002) [arXiv:hep-th/0111123].\n[18] D. Mattingly, Living Rev. Rel. 8, 5 (2005) [arXiv:gr-qc/0502097].\n[19] P. Arias, H. Falomir, J. Gamboa, F. Mendez and F. A. Schaposnik , Phys. Rev. D 76, 025019 (2007)\n[arXiv:0705.3263 [hep-th]].\n[20] T. Jacobson, S. Liberati and D. Mattingly, Ann. Phys. 321, 150 (2006) [arXiv:astro-ph/0505267].\n[21] L. Dolan and R. Jackiw, Phys. Rev. D 9, 3320 (1974).\n[22] S. Weinberg, Gravitation and Cosmology , J. Wiley, New York, 1972.\n[23] C. D. Lane, Phys. Rev. D 72, 016005 (2005) [arXiv:hep-th/0505130].\n[24] V. A. Kostelecky, N. Russell, “ Data tables for Lorentz and CPT violation”, arXiv:0801.0287 [hep-ph].\n9" }, { "title": "1009.4717v1.Lorentz_violation_in_solar_neutrino_oscillations.pdf", "content": "arXiv:1009.4717v1 [hep-ph] 23 Sep 2010November 9, 2018 5:7 WSPC - Proceedings Trim Size: 9in x 6in pr oceedings\n1\nLORENTZ VIOLATION IN\nSOLAR-NEUTRINO OSCILLATIONS\nJONAH E. BERNHARD\nDepartment of Physics and Astronomy, Swarthmore College,\nSwarthmore, PA 19081, United States\nE-mail: jbernha2@swarthmore.edu\nSolar-neutrino oscillations are considered using a massiv e model with pertur-\nbative Lorentz violation. The adiabatic approximation is u sed to calculate the\neffects of Lorentz violation to leading order. The results ar e more compact\nthan previous work involving vacuum oscillations, and are a ccurate for small\nLorentz-violating coefficients.\n1. Introduction\nNeutrinos are convenient for studying Lorentz violation due to the ir small\nmass and high velocities. Previous research has considered neutrin o oscilla-\ntions with short1and long2baselines, but mostly zero or constant matter\neffects. Very little has been done with variable matter effects, such as those\nin the Sun. We attempt to develop a perturbative technique for eas ily cal-\nculating Lorentz-violating effects in solar-neutrino oscillations.\n2. Theory\nWe assume neutrinos have the standard mass differences, and tre at Lorentz\nviolation as a perturbative effect. The neutrino Hamiltonian thus has three\nterms\nH=1\n2EU†∆m2U+V+δH, (1)\nwhere the first term on the right-hand side is the standard vacuum Hamil-\ntonian,Vis the Sun’s matter potential, and δHis a small Lorentz-violating\nHamiltonian. The diagonal mass squared matrix ∆ m2and the unitary mix-November 9, 2018 5:7 WSPC - Proceedings Trim Size: 9in x 6in pr oceedings\n2\ningmatrix Utakethestandardform,usingtheapproximateobservedvalues\n∆m2\n⊙≃8.0×10−17MeV2,∆m2\natm≃2.5×10−15MeV2,\nθ12≃34◦, θ23≃45◦, θ13≃0◦, δ≃0.\nSinceθ23≃45◦andθ13≃0◦, muon and tau neutrinos mix equally and\nPeµ=Peτ. Given this similarity, it is convenient to group muon and tau\ntogether and consider only two flavors. In this case, there are on ly two\nparameters: a mixing angle θand a mass difference ∆ m2, equivalent to\nθ12and ∆m2\n⊙, respectively, in the three-generation case. This significantly\nsimplifies many calculations. The two-generationcase will be consider ed for\nthe remainder of this work.\nThe Sun produces very large quantities of electron neutrinos in its c ore.\nAs they propagate to the surface, they interact with matter. Th is effect can\nbe described by the Sun’s matter potential V, given approximately by\nV=/parenleftbigg√\n2GFne0\n0 0/parenrightbigg\n, GFne≃1.32×10−17e−10.54R/R⊙MeV,(2)\nwhereGFis the Fermi coupling constant, neis the number density of\nelectrons in the Sun, and R⊙is the solar radius.3\nThe variable matter potential precludes an exact solution; howeve r,pro-\nvided the Hamiltonian changes sufficiently slowly, approximate analytic ex-\npressions can be obtained using the adiabatic approximation. This ap proxi-\nmation assumes energy eigenstates change slowly enough that neu trinos do\nnot transition between them as they propagate. High-frequency oscillations\nbetween flavors can then be averaged away, resulting in simpler exp ressions\nthan in previous work, such as reference 2. The averages can be c alculated\nfrom\n/angbracketleftP(0)\neb/angbracketright=/angbracketleftbig\n|νb|2/angbracketrightbig\n=/summationdisplay\na/vextendsingle/vextendsingle(U∗)ab(U0)a1/vextendsingle/vextendsingle2(3)\nwhere the superscript (0) indicates that these are zeroth-orde r (Lorentz-\ninvariant) probabilities, Uis the vacuum mixing matrix, and U0is the\neffective mixing matrix at R= 0. This evaluates to\n/angbracketleftP(0)\nee/angbracketright=s2\nθs2\nθ0+c2\nθc2\nθ0,/angbracketleftP(0)\neµ/angbracketright=s2\nθc2\nθ0+c2\nθs2\nθ0, (4)\nwheresθ≡sinθ,cθ≡cosθ,θis the vacuum mixing angle, and θ0is the\neffective mixing angle at R= 0. Note that the matter potential only needs\nto be known at the center and surface of the Sun.\nFigure 1 shows the electron neutrino survival probability as a funct ion\nof energy. The adiabatic approximation agrees with the numerical s olution\nalmost exactly.November 9, 2018 5:7 WSPC - Proceedings Trim Size: 9in x 6in pr oceedings\n3\nFig. 1. Average electron (left) and muon (right) neutrino su rvival probability (solid)\nand envelope function (dashed, given by 2 sθsθ0cθcθ0) as a function of energy in the\nadiabatic limit. For comparison, the probability was calcu lated numerically at 10000\nlogarithmically spaced energies from 0.1 to 100 MeV (grey).\n3. Lorentz violation\nLorentz violation is a natural extension to the theory. Assuming Lo rentz\nviolation is sufficiently small, it can be treated as a first-order pertur bation\nto the zeroth-order probabilities from Eq. (4).\nThe Lorentz-violating term has the form4\nδHab=1\nE/bracketleftbig\n(aL)αpα−(cL)αβpαpβ/bracketrightbig\nab(5)\nwhere (aL)α\nab, (cL)αβ\nabare complex Lorentz-violating coefficients and pαis\nthe neutrino energy-momentum four-vector, pα= (E,−/vector p)≃E(1,−ˆp). The\ncoefficients ( aL)T\nab, (cL)TT\nabare isotropic and hence introduce changes to\nenergy dependence, while the remaining coefficients are anisotropic and\ncause annual variations. The CPT-odd ( aL)α\nabare constant with energy;\nthe CPT-even ( cL)αβ\nabare linear.\nThe probabilities from the adiabatic approximation become\n/angbracketleftPeb/angbracketright=/angbracketleftP(0)\neb/angbracketright+2E\n∆m2Mij\nebδHij, (6)\nwhereMij\nebis a unitless, order one effective scale factor for δH, used to\ncalculate perturbed probabilities via simple linear combinations of coeffi -\ncients. It is determined by expanding the Lorentz-violating sine and cosine\nto leading order, which gives\nMij\nee=∆m2\nE∆λ/parenleftbigg−1\n4/parenleftbig\ns2\n2θc2θ0+s2\n2θ0c2θ/parenrightbig\n−sθc3\nθc2θ0−sθ0c3\nθ0c2θ\ns3\nθcθc2θ0+s3\nθ0cθ0c2θ1\n4/parenleftbig\ns2\n2θc2θ0+s2\n2θ0c2θ/parenrightbig/parenrightbigg\n(7)\nandMij\neµ=−Mij\nee, where ∆ λis the difference between the two eigenvalues\natR= 0.November 9, 2018 5:7 WSPC - Proceedings Trim Size: 9in x 6in pr oceedings\n4\nFig. 2. Comparison of the energy dependence of electron neut rino survival probability\nwith no Lorentz violation (solid), first-order expansion (d ashed, calculated from Mij\nee),\nand exact (dotted, calculated numerically from the adiabat ic approximation). The values\nused are ( aL)T\n11= 2×10−19MeV in the left plot, and ( cL)TT\n11= 2×10−19on the right.\nFigure 2 shows the effects of the isotropic coefficients on the energ y de-\npendence of the electron neutrino survival probability. The first- order ex-\npansion is effective with nonzero ( aL)T\n11, especiallyat lowenergies. It begins\nto fail with nonzero ( cL)TT\n11at high energies, since these effects grow with\nenergy. However, the Sun does not produce neutrinos with energ ies higher\nthan approximately 10 MeV, so this shortcoming is of limited relevance .\n4. Discussion\nSolar neutrinos are convenient for the study of Lorentz violation d ue to\nthe large quantities and significantly lower energy and longer baseline than\naccelerator experiments. The adiabatic approximation allows simple a nd\naccurate calculations of Lorentz-violating effects in solar-neutrin o oscilla-\ntions.\nFurther work will generalize the analysis to the three-generation c ase\nand explore additional consequences of Lorentz violation, such as an-\nnual variations caused by nonzero anisotropic coefficients and neu trino-\nantineutrino mixing.\nReferences\n1. V. A. Kostelecky and M. Mewes, Phys. Rev. D 70, 076002 (2004).\n2. J. S. Diaz, V. A. Kostelecky, and M. Mewes, Phys. Rev. D 80, 076007 (2009).\n3. J. N. Bahcall, M. H. Pinsonneault and S. Basu, Astrophys. J .555, 990 (2001).\n4. V. A. Kostelecky and M. Mewes, Phys. Rev. D 69, 016005 (2004)." }, { "title": "1106.6317v2.An_Osserman_type_condition_on__g_f_f__manifolds_with_Lorentz_metric.pdf", "content": "An Osserman-type condition on g:f:f-manifolds\nwith Lorentz metric\u0003\nLetizia Brunetti\nAbstract\nA condition of Osserman type, called '-null Osserman condition, is introduced and stud-\nied in the context of Lorentz globally framed f-manifolds. An explicit example shows the\nnaturalness of this condition in the setting of Lorentz S-manifolds. We prove that a Lorentz\nS-manifold with constant '-sectional curvature is '-null Osserman, extending a result stated\nfor Lorentz Sasaki space forms. Then we state a characterization for a particular class of\n'-null Osserman S-manifolds. Finally, some examples are examined.\n2000 Mathematics Subject Classi\fcation. 53C25, 53C50, 53B30.\nKeywords and phrases. Lorentz metrics, Osserman condition, g:f:f -structure.\n1 Introduction\nThe study of the behaviour of the Jacobi operators is an important topic in Riemannian and, more\ngenerally, in semi-Riemannian geometry. More precisely, let ( M;g) be a Riemannian manifold with\ncurvature tensor Rand consider a point pinM. For any unit vector X2TpM, the symmetric\nendomorphism RX=Rp(\u0001;X)X:X?!X?is called the Jacobi operator with respect to X.\nIf the eigenvalues of RXare independent of the choices of Xandp, one says that ( M;g) is an\nOsserman manifold ([15]).\nSeveral results have been obtained looking for the solution of the Osserman Conjecture ([8, 22]),\nwhich states that an Osserman manifold is \rat or it is locally a rank-one symmetric space ([8, 9,\n10, 18, 19, 20]). Osserman manifolds have been studied in the Lorentzian context ([4, 13, 14]),\nwhere a complete solution for the Osserman conjecture has been found. Recently, in [1], Atindogbe\nand Duggal have introduced and studied suitable operators of Jacobi type associated with a semi-\nRiemannian degenerate metric.\nIn ([14]) the authors de\fned the Jacobi operator \u0016Ru,ubeing a null (or lightlike) vector tangent\nto a Lorentz manifold M. Given a unit timelike vector ztangent to M, they introduced and\ninvestigated the so-called null Osserman condition with respect to z(see also [15]).\nObviously, Lorentz almost contact manifolds are studied in this context. In particular, a Lorentz\nSasaki space form, whose characteristic vector \feld \u0018is timelike, is globally null Osserman with\nrespect to\u0018([15]). This result does not hold in the context of Lorentz globally framed f-manifolds\n(M2n+s;';\u0018\u000b;\u0011\u000b;g),s\u00152, as we will see with a counterexample.\nThis motivates the introduction of a more general condition of Osserman type, which we will\ncall'-null Osserman condition .\nThe main results of this paper state the links between the '-null Osserman condition and\nthe behaviour of the '-sectional curvature in Lorentz S-manifolds. After a preliminary section,\n\u0003The author wishes to express her thanks to professors A.M. Pastore and M. Falcitelli for helpful comments and\nfor many stimulating conversations. The work was supported by the Research Program n. 01.08 of University of\nBari\n1arXiv:1106.6317v2 [math.DG] 27 Sep 2012where we gather some facts about g:f:f -manifolds, needed in the rest of the paper, in Section 3\nwe discuss the relationship between the null Osserman condition and the Lorentz S-structures,\ngiving an example of Lorentz S-space form which does not satisfy the null Osserman conditions.\nWe endow the compact Lie group U(2) with a Lorentz S-structure of rank 2. This manifold is an\nS-space form with two characteristic vector \felds \u00181and\u00182,\u00181timelike, that does not satisfy the\nnull Osserman condition with respect to \u00181.\nIn Section 4 we introduce the notion of '-null Osserman manifold, and we state that a Lorentz\nS-manifold with constant '-sectional curvature is '-null Osserman with respect to the timelike\ncharacteristic vector \feld. We prove, in Section 5, an algebraic characterization for the Riemannian\ncurvature tensor \feld in a particular class of '-null Osserman Lorentz S-manifolds. Namely, we\ndivide this section in two parts. In the \frst subsection we deal with technical results which are\nvery useful in the second subsection where we state the main result. Moreover, we look at the\nbehaviour of the '-sectional curvature when the number of the eigenvalue of the Jacobi operator\nis one.\nIn particular, it is interesting to note that the existence of the only eigenvalue 1 of the Jacobi\noperator is related to the '-sectional \ratness of the manifold.\nFinally in the case of 4-dimensional '-null Osserman manifolds we \fnd a compact example,\nusing the Lie group U(2), and also a non compact example.\nAll manifolds, tensor \felds and maps are assumed to be smooth, moreover we suppose all\nmanifolds are connected. We will use the Einstein convention omitting the sum symbol for repeated\nindexes. Following the notations of S. Kobayashi and K. Nomizu ([12]), for the curvature tensor\nRwe haveR(X;Y )Z=rXrYZ\u0000rYrXZ\u0000r [X;Y]Z, andR(X;Y;Z;W ) =g(R(Z;W )Y;X),\nfor anyX;Y;Z;W2X(M). The sectional curvature Kp(\u0019) atpof a non-degenerate 2-plane\n\u0019=spanfX;Ygis given by\nKp(\u0019) =Kp(X;Y ) =Rp(X;Y;X;Y )\n\u0001(\u0019)=gp(Rp(X;Y )Y;X)\n\u0001(\u0019);\nwhere \u0001(\u0019) =g(X;X )g(Y;Y)\u0000g(X;Y )26= 0.\n2 Preliminaries\nFollowing [3, 6, 23], we recall some de\fnitions. An almost contact manifold is a (2 n+1)-dimensional\nmanifoldMendowed with an almost contact structure, i.e. M2n+1has a (1;1)-tensor \feld fsuch\nthatrank (f) = 2n, a 1-form\u0011and a vector \feld \u0018satisfyingf2(X) =\u0000X+\u0011(X)\u0018and\u0011(\u0018) = 1.\nMoreover, if gis a semi-Riemannian metric on M2n+1such that, for any X;Y2X(M2n+1),\ng(fX;fY ) =g(X;Y )\u0000\"\u0011(X)\u0011(Y);\nwhere\"=\u00061 according to the causal character of \u0018, thenM2n+1is called an inde\fnite almost\ncontact manifold. Such a manifold is said to be an inde\fnite contact manifold if d\u0011= \b, where\n\b is de\fned by \b( X;Y ) =g(X;fY ). Furthermore, if the structure ( f;\u0018;\u0011 ) is normal, that is\nN= [f;f] + 2d\u0011\n\u0018= 0, then the inde\fnite contact structure is called an inde\fnite Sasaki\nstructure and, in this case, the manifold ( M2n+1;f;\u0018;\u0011;g ) is called inde\fnite Sasaki.\nIn the Riemannian case a generalization of these structures have been studied by Blair in [2],\nby Goldberg and Yano in [17]. In [6] we studied such structures in semi-Riemannian context.\nA manifold Mis called a globally framed f-manifold (brie\ry g:f:f -manifold) if it is endowed\nwith a nowhere-vanishing (1 ;1)-tensor \feld 'of constant rank, such that ker 'is parallelizable i.e.\nthere exist global vector \felds \u0018\u000b,\u000b2f1;:::;sg, and 1-forms \u0011\u000b, satisfying\n'2=\u0000I+\u0011\u000b\n\u0018\u000band\u0011\u000b(\u0018\f) =\u000e\u000b\n\f:\n2Ag:f:f -manifold ( M2n+s;';\u0018\u000b;\u0011\u000b),\u000b2f1;:::;sg, is said to be an inde\fnite g:f:f -manifold\nifgis a semi-Riemannian metric satisfying the following compatibility condition\ng('X;'Y ) =g(X;Y )\u0000\"\u000b\u0011\u000b(X)\u0011\u000b(Y)\nfor any vector \felds X;Y , being\"\u000b=\u00061 according to whether \u0018\u000bis spacelike or timelike. Then,\nfor any\u000b2f1;:::;sgandX2X(M2n+s), one has\u0011\u000b(X) =\"\u000bg(X;\u0018\u000b).\nAn inde\fnite g:f:f -manifold is an inde\fnite S-manifold if it is normal and d\u0011\u000b= \b, for any\n\u000b2f1;:::;sg, where \b(X;Y ) =g(X;'Y ) for anyX;Y2X(M2n+s). The normality condition is\nexpressed by the vanishing of the tensor \feld N=N'+ 2d\u0011\u000b\n\u0018\u000b,N'being the Nijenhuis torsion\nof'.\nFurthermore, as proved in [6], the Levi-Civita connection of an inde\fnite S-manifold satis\fes:\n(rX')Y=g('X;'Y )e\u0018+e\u0011(Y)'2(X);\nwheree\u0018=Ps\n\u000b=1\u0018\u000bande\u0011=\"\u000b\u0011\u000b. Note that, for s= 1, we reobtain the notion of inde\fnite Sasaki\nmanifold.\nWe recall thatrX\u0018\u000b=\u0000\"\u000b'Xand ker'is an integrable \rat distribution since r\u0018\u000b\u0018\f= 0, for\nany\u000b;\f2f1;:::;sg. Anyway, an inde\fnite S-manifold is never \rat and it is never a real space\nform since, for example, K(X;\u0018\u000b) =\"\u000bfor any non lightlike X2Im'p.\nFor more details we refer to [6], where we describe three examples of non compact inde\fnite S-\nmanifolds. More precisely we construct two di\u000berent inde\fnite S-structures with metrics of index\n\u0017= 2 on R6and an inde\fnite S-structure with Lorentz metric on R4. Moreover, in [7] we give\nexplicit examples of compact inde\fnite g:f:f -manifolds and inde\fnite S-manifolds.\nWe also remark that every g:f:f -manifold is subject to the following topological condition: it\nhas to be either non compact or compact with vanishing Euler characteristic, since it admits never\nvanishing vector \felds. This implies that such a g:f:f -manifold always admits Lorentz metrics.\nLet us \fx few notation about curvature tensor \feld. As usual, a 2-plane \u0019=spanfX;'Xgin\nTpM, withp2MandX2Im'p, is said to be a '-plane and the sectional curvature at pof such\na plane, with Xa non lightlike vector, is called the '-sectional curvature atpand is denoted by\nHp(X).\nAn inde\fniteS-manifold ( M;';\u0018\u000b;\u0011\u000b;g) is said to be an inde\fnite S-space form if the '-\nsectional curvature Hp(X) is constant, for any point and any '-plane. In particular, in [6] it is\nproved that an inde\fnite S-manifold (M;';\u0018\u000b;\u0011\u000b;g) is an inde\fniteS-space form with Hp(X) =c\nif and only if the Riemannian (0 ;4)-type curvature tensor \feld Ris given by\nR(X;Y;Z;W ) =\u0000c+ 3\"\n4fg('Y;'Z )g('X;'W )\u0000g('X;'Z )g('Y;'W )g (1)\n\u0000c\u0000\"\n4f\b(W;X )\b(Z;Y)\u0000\b(Z;X)\b(W;Y )\n+ 2\b(X;Y )\b(W;Z )g\u0000fe\u0011(W)e\u0011(X)g('Z;'Y )\n\u0000e\u0011(W)e\u0011(Y)g('Z;'X ) +e\u0011(Y)e\u0011(Z)g('W;'X )\n\u0000e\u0011(Z)e\u0011(X)g('W;'Y )g;\nfor any vector \felds X,Y,ZandWonM, where\"=Ps\n\u000b=1\"\u000b.\nIn regard to the curvature tensor of an inde\fnite S-manifold, it is important to recall the\n3following formulas, for any X;Y;Z;W2Im'and any\u000b;\f;\r;\u000e2f1;:::;sg:\nR(X;\u0018\u000b;X;Y ) =\"\u000bg(X;X )g(e\u0018;Y) = 0;\nR(\u0018\u000b;X;\u0018\f;Y) =\"\u000b\"\fg(X;Y );\nR(\u0018\u000b;X;\u0018\f;\u0018\r) =\"\u000b\"\fg(X;\u0018\r) = 0; (2)\nR(\u0018\u000b;\u0018\u000e;\u0018\f;\u0018\r) = 0;\nR(X;Y;'Z;W ) +R(X;Y;Z;'W ) =\"P(X;Y ;Z;W );\nwhereP(X;Y ;Z;W ) = \b(X;Z)g(Y;W )\u0000\b(X;W )g(Y;Z)\u0000\b(Y;Z)g(X;W ) + \b(Y;W )g(X;Z).\nFinally, we recall some useful properties for a curvature-like algebraic tensor. Let ( V;g) be a\npseudo-Euclidean real vector space of index \u0017, 0<\u0017 < dimV. A multilinear map F:V4!Ris\ncalled a curvature-like map (or curvature-like algebraic tensor) if it satis\fes the following conditions\nF(y;x;z;w ) =\u0000F(x;y;z;w );\nF(z;w;x;y ) =F(x;y;z;w );\nF(x;y;z;w ) +F(x;z;w;y ) +F(x;w;y;z ) = 0:\nFor any non-degenerate 2-plane \u0019=spanfz;wginVit is possible to de\fne the number\nk(z;w) =F(z;w;z;w )\n\u0001(\u0019):\nIfk(z;w) is constant for any non-degenerate 2-plane and k(z;w) =kthen one gets F(x;y;z;w ) =\nk(g(x;z)g(y;w)\u0000g(y;z)g(x;w)). Now, arguments similar to those in Proposition 28 ([21, page\n229]), can be used to prove the following result.\nLemma 2.1. Let(V;g)be a Lorentz real vector space and F:V4!Ra curvature-like map. Then\nthe following conditions are equivalent.\na)F(x;y;z;w ) =k(g(x;z)g(y;w)\u0000g(y;z)g(x;w)),\nb)F(x;y;y;x ) = 0 for any degenerate plane \u0019=spanfx;yginV.\n3 Null Osserman condition and Lorentz S-manifolds\nIt is well-known that a Lorentz manifold has constant sectional curvature at a point pif and only\nif it satis\fes the Osserman condition at p.\nContrary to this, no Lorentz S-manifold can satisfy the Osserman condition since, as remarked\nin Section 2, a Lorentz S-manifold can not have constant sectional curvature.\nIn [14] the authors introduce another Osserman condition, named the null Osserman condition.\nNamely, let ( M;g) be a Lorentz manifold, p2Mandua null vector in TpM. Then the orthogonal\ncomplement u?ofuis a degenerate vector space since spanfug \u001au?. So one considers the\nquotient space \u0016 u?=u?=spanfugand the canonical projection \u0019:u?!\u0016u?. It is possible to\nde\fne a positive de\fnite inner product \u0016 gon \u0016u?putting\n\u0016g(\u0016x;\u0016y) =g(x;y);\nwhere, for any x;y2u?, \u0016x=\u0019(x) and \u0016y=\u0019(y).\nFrom now on, every bar-object will stand for geometrical objects related to \u0016 u?. So, \fxed a\nnull vector u2TpM, the Jacobi operator with respect to ucan be de\fned by the linear map\n\u0016Ru: \u0016u?!\u0016u?such that \u0016Ru\u0016x=\u0019(R(x;u)u) ([14] and De\fnition 3.2.1 in [15]).\n4Clearly, \u0016Ruis self-adjoint with respect to \u0016 g, hence \u0016Ruis diagonalizable.\nIn Lorentzian geometry it is well-known that a null vector uand a timelike vector zare never\northogonal. Hence, in a Lorentz manifold ( M;g), the null congruence set determined by a timelike\nvectorz2TpMatp, denoted by N(z), is de\fned by\nN(z) =fu2TpMjg(u;u) = 0; g(u;z) =\u00001g:\nA Lorentz manifold ( M;g) is called null Osserman with respect to a unit timelike vector z2TpM\nat a pointpif the characteristic polynomial of \u0016Ruis independent of u2N(z). LetLbe a timelike\nline subbundle of TM. If (M;g) is null Osserman with respect to each unit timelike vector z2L,\nthen (M;g) is called pointwise null Osserman with respect to L. Moreover, if ( M;g) is pointwise\nnull Osserman with respect to Land the characteristic polynomial of \u0016Ruis independent of the\nchoice of a unit z2L, then (M;g) is said to be globally null Osserman with respect to L.\nAnother set associated to a unit timelike vector zinTpMis the celestial sphere S(z) ofzgiven\nby\nS(z) =fx2z?jg(x;x) = 1g:\nAccording to a result in [15], using the celestial sphere of z, one can obtain all the elements of\nN(z). In fact one has\n8u2N(z)9jx2S(z) such that u=z+x:\nIt is very natural to use this de\fnition in the context of Lorentz contact manifolds. In particular, as\nstated in [15], Lorentz Sasaki space forms are globally null Osserman with respect to the timelike\ncharacteristic vector \feld. An easy example shows that in a Lorentz S-space form the null Osserman\ncondition with respect to a timelike characteristic vector does not hold.\nIndeed, considering the 4-dimensional manifold U(2) and the Lie algebra u(2), we denote by\n\u00181;\u00182;X;Y the left-invariant vector \felds on U(2), determined, in the same order, by the basis\nf{E11;\u0000{E22;E12\u0000E21;{(E12+E21)gofu(2), where ( Eij)i;j2f1;2gis the canonical basis of gl(2;C).\nThen, we get:\n[X;Y ] = 2\u00181+ 2\u00182;[X;\u0018\u000b] =\u0000Y;[Y;\u0018\u000b] =X; [\u0018\u000b;\u0018\f] = 0\nfor any\u000b;\f2f1;2g. Let us consider the left-invariant 1-forms \u00111and\u00112determined by the dual\n1-forms of{E11and\u0000{E22, respectively, and the left-invariant tensor \feld 'such that'(X) =Y,\n'(Y) =\u0000Xand'(\u00181) ='(\u00182) = 0. The manifold U(2) is compact, connected, with Euler number\n\u001f(U(2)) = 0, thus we can de\fne a left-invariant Lorentz metric gsuch that the vector \felds \u00181,\u00182,\nXandYform an orthonormal basis with g(\u00181;\u00181) =\u00001. Such a structure on U(2) is constructed\nin the Riemannian context ([11]) and then it is adapted to the Lorentzian case ([7]).\nThis structure is a normal inde\fnite g:f:f -structure and its associated Sasaki 2-form \b veri\fes\n\b =d\u0011\u000b, for any\u000b2f1;2g, so that it turns out to be a Lorentz S-structure on U(2). Moreover,\none sees at once that U(2) has constant '-sectional curvature 4. We see that U(2) does not verify\nthe null Osserman condition with respect to ( \u00181)p, for anyp2U(2). In fact, \fxing p2U(2) and\nputting\nu1=Xp+ (\u00181)p; u2=Yp+ (\u00181)p; u3= (\u00182)p+ (\u00181)p;\none hasu1;u2;u32N((\u00181)p). By (1), we have\nR(Yp;u1)u1=Yp+ 3g(Yp;'u 1)'u1+e\u0011(u1)e\u0011(u1)Yp= 5Yp;\nR((\u00182)p;u1)u1=2X\n\u000b=1(\u0018\u000b)p+Xp= (\u00182)p+u1:\n5Analogously, for u2, we obtain\nR(Xp;u2)u2=Xp+ 3Xp+Xp= 5Xp;\nR((\u00182)p;u2)u2=2X\n\u000b=1(\u0018\u000b)p+Yp= (\u00182)p+u2:\nFor anyz2u?\n3, we have\nR(z;u3)u3=\u0000e\u0011(u3)e\u0011(u3)'2z= 0;\nsincee\u0011(u3) = 0.\nThen it is evident that the eigenvalues of \u0016Ru1and \u0016Ru2are 5 and 1 whereas \u0016Ru3= 0.\n4 The '-Null Osserman Condition\nIn this section, inspired by the example of U(2), we introduce a new Osserman condition that will\nbe applied to Lorentz g:f:f -manifolds.\nLet (M;';\u0018\u000b;\u0011\u000b;g),\u000b2f1;:::;sg, be a Lorentz g:f:f -manifold, it is easy to check that the\ntimelike vector \feld must be a characteristic vector \feld. Without loss of generality we can assume\nthat\u00181is the timelike vector \feld.\nTaking in mind the example in Section 3, we claim that, if s\u00152, then the \ratness of ker '\nin\ruences the behaviour of the Jacobi operators \u0016Ru\u000bwithu\u000b= (\u00181)p+(\u0018\u000b)p, for any\u000b2f2;:::;sg\nandp2M. Since the matter is related to the null vector u\u000b, we give the following Osserman\ncondition.\nGiven a point pofM, the set\nS'((\u00181)p) =S((\u00181)p)\\Im'p;\nis called the '-celestial sphere of (\u00181)patp. We de\fne the analogous of the null congruence set,\ncalled the'-null congruence set, denoted by N'((\u00181)p), putting\nN'((\u00181)p) =fu2TpMju= (\u00181)p+x; x2S'((\u00181)p)g:\nNow, we are ready to state the de\fnition of '-null Osserman condition with respect to the\ntimelike vector ( \u00181)pat a pointp2M.\nDe\fnition 4.1. Let (M;';\u0018\u000b;\u0011\u000b;g) be a Lorentz g:f:f -manifold, dim M= 2n+s,n;s\u00151, with\ntimelike vector \feld \u00181and consider p2M.Mis called'-null Osserman with respect to ( \u00181)pat\na pointp2Mif the characteristic polynomial of \u0016Ruis independent of u2N'((\u00181)p), that is the\neigenvalues of \u0016Ruare independent of u2N'((\u00181)p).\nRemark 4.2. If (M;';\u0018;\u0011;g ) is a Lorentz almost contact manifold, then it can be considered as a\nLorentzg:f:f -manifold with s= 1. Obviously one has S((\u0018)p) =S'((\u0018)p) andN((\u0018)p) =N'((\u0018)p),\nfor anyp2M. It follows that the null Osserman condition with respect to \u0018pat a pointpcoincides\nwith the'-null Osserman condition at the same point.\nIt is clear that U(2) veri\fes the '-null Osserman condition with respect to ( \u00181)pat a point\np2U(2). In fact, we consider an arbitrary unit vector zof Im'pputtingz=aXp+bYp. Setting\nu4=z+ (\u00181)p, we haveu42N'((\u00181)p) and\nu?\n4=spanfXp+a(\u00181)p;Yp+b(\u00181)p;(\u00182)pg=spanf'u4;u4;(\u00182)pg:\n6Then, we get\nR('u4;u4)u4='u4+ 3'u4+'u4= 5'u4;\nR((\u00182)p;u4)u4=2X\n\u000b=1(\u0018\u000b)p\u0000'2u4= (\u00182)p+ (\u00181)p+z= (\u00182)p+u4:\nIt follows that, for any u=z+ (\u00181)pinN'((\u00181)p) withz2Im'pandg(z;z) = 1, the eigenvalues\nof\u0016Ruare 5 and 1, hence the eigenvalues of \u0016Ruare independent of the choice of u2N'((\u00181)p).\nTaking into account the classical de\fnitions for the Osserman manifolds, we introduce the\nglobally'-null Osserman condition.\nDe\fnition 4.3. Let (M;';\u0018\u000b;\u0011\u000b;g) be a Lorentz g:f:f -manifold, dim M= 2n+s,n;s\u00151, with\ntimelike vector \feld \u00181. IfMis'-null Osserman with respect to ( \u00181)p, for anyp2M, and the\ncharacteristic polynomial of \u0016Ruis independent of the choice of p2M, thenMis said to be globally\n'-null Osserman with respect to \u00181.\nLooking again at the example of U(2) one can see at once that it is a globally '-null Osserman\nmanifold with rispect to \u00181. In fact, it is clear that the eigenvalues of \u0016Ruare independent of the\npointp.\nIn the next theorem we prove, more generally, that each Lorentz S-space form satis\fes the\n'-null Osserman condition.\nTheorem 4.4. Let(M;';\u0018\u000b;\u0011\u000b;g),dimM= 2n+s, be a LorentzS-manifold with \u00181timelike\nand constant '-sectional curvature, Let p2M. ThenMveri\fes the '-null Osserman condition\nwith respect to the timelike characteristic vector at a point p.\nProof. Letp2M. Denoting by cthe'-sectional curvature, (1) holds with \"=s\u00002.\nLetube a vector in N'((\u00181)p), henceu= (\u00181)p+x1withx12S'((\u00181)p), and consider x2u?.\nWe have:\ng('u;'u ) =g(u;u)\u0000sX\n\u000b=1\"\u000b\u0011\u000b(u)\u0011\u000b(u) =\u00111(u)\u00111(u) = 1; (3)\ng('x;'u ) =g(x;u)\u0000sX\n\u000b=1\"\u000b\u0011\u000b(x)\u0011\u000b(u) =\u00111(x): (4)\nBy (1), (3) and (4) we compute R(x;u;u;w ) for anyw2TpM, obtaining\nRp(x;u;u;w ) =\u0000c+ 3(s\u00002)\n4fg('x;'w )\u0000\u00111(x)g('u;'w )g (5)\n\u00003\n4(c\u0000s+ 2)g(x;'u )g(w;'u )\n\u0000fe\u0011(w)e\u0011(x) +e\u0011(w)\u00111(x) +g('w;'x ) +e\u0011(x)g('w;'u )g:\nNow, being dim M= 2n+s, we considerfx1;'x 1;x3;:::;x 2ngas an orthonormal base of\nIm'p, which determines the bases B=fu;'x 1;(\u00182)p;:::; (\u0018s)p;x3;:::;x 2ngofu?andB=\nf'x1;(\u0016\u00182)p;:::; (\u0016\u0018s)p;\u0016x3;:::; \u0016x2ngof \u0016u?. For brevity, we also denote them by B=feig1\u0014i\u0014m,\n\u0016B=f\u0016eig1\u0014i\u0014m\u00001, beingm= 2n+s\u00001. In general, for any x2u?, one has\n\u0016Ru(\u0016x) =\u0000m\u00001X\ni=1Rp(x;u;u;ei)\u0016ei: (6)\n7By (5) and (6) we obtain\n\u0016Ru('x1) =fc+ 3(s\u00002)\n4+3\n4(c\u0000s+ 2)g'x1+'x1= (c+ 1)'x1;\n\u0016Ru(\u0016xj) =c+ 3(s\u00002)\n4\u0016xj+ \u0016xj=c+ 3s\u00002\n4\u0016xj;8j2f2;:::; 2ng\n\u0016Ru((\u0016\u0018\f)p) =sX\n\r=2e\u0011((\u0018\f)p)e\u0011((\u0018\r)p)(\u0016\u0018\r)p=sX\n\r=2(\u0016\u0018\r)p;8\f2f2;:::;sg:\nIt follows that the representation matrix of \u0016Ruwith respect to \u0016Bis independent of the choice of\nu2N'((\u00181)p). In particular, it is easy to compute that the other eigenvalues are 0 and s\u00001,\nhaving eigenvectors \u0016 x\u000b= (\u0016\u00182)p\u0000(\u0016\u0018\u000b)p,\u000b2f3;:::;sg, and \u0016x=Ps\n\f=2(\u0016\u0018\f)p, respectively. This\ncompletes the proof.\nBy the above proof we note that, as for U(2), each Lorentz S-manifold ( M;';\u0018\u000b;\u0011\u000b;g),\ndimM= 2n+s, with constant '-sectional curvature is globally '-null Osserman with respect\nto\u00181.\nFrom now on, since the Osserman conditions are formulated pointwise, to simplify the notation\nwe omit any reference to the point, there is no ambiguity.\n5 The '-null Osserman condition on Lorentz S-manifolds\nwith additional assumptions\nIn this section we proceed with the study of '-null Osserman manifolds and we will \fnd an expres-\nsion for the curvature tensor \feld of a '-null Osserman Lorentz S-manifold with two characteristic\nvector \felds, using a suitable expression for null vectors. An analogous statement can be found in\ndi\u000berent contexts ([15]). In the \frst part of this section we collect the technical issues needed for\nthe main result, which will be provided in the second subsection.\n5.1 Technical results\nIn [16] the authors have given the esplicit construction of a complex structure on a (4 m+ 2)-\ndimensional globally Osserman manifold with exactly two distinct eigenvalues of the Jacobi opera-\ntors with multiplicities 1 and 4 m(see also [15]). We will use such a construction, adapting it when\nthe manifold veri\fes the '-null Osserman condition at a point.\nFollowing [14, 15], we recall that if ( M;g) is a Lorentz manifold and uis a null vector of\nTpMthen a non-degenerate subspace W\u001au?such that dim W= dim \u0016u?is called a geometric\nrealization of \u0016u?. Let\u0019jW: (W;g)!(\u0016u?;g) be an isometry where, to simplify, we use the same\nlettergfor non-degenerate metrics on Wand \u0016u?. A vector x2Wis said to be a geometrically\nrealized eigenvector of\u0016RuinWcorresponding to an eigenvalue \u0015if\u0019jW(x) = \u0016xis an eigenvector\nof\u0016Ruwith eigenvalue \u0015([15]).\nRemark 5.1. Let (M;';\u0018\u000b;\u0011\u000b;g) be a (2n+s)-dimensional '-null Osserman Lorentz S-manifold\nat a pointp2Mandu2N'(\u00181). We suppose that the Jacobi operator \u0016Ru, restricted to u?\\Im',\nhas exactly two eigenvalues, c1andc2, with multiplicities 1 and 2 n\u00002.\nSinceu=\u00181+x,x2S'(\u00181), using (2), it is easy to see that the eigenvalues and the eigenvectors\nof the Jacobi operator \u0016Ruare connected with those of Rxjx?\\Im'. Namely, one can prove that\nv2x?\\Im'is an eigenvector of Rxrelated to the eigenvalue \u0015if and only if it is a geometrically\nrealized eigenvector of \u0016Rurelated to the eigenvalue \u0015+ 1 ([5]).\n8Now, let us \fx p2Mand, following [16], identify S'(\u00181)\u0018=S2n\u00001. For anyx2S2n\u00001consider\nthe operator Rx:x?\\Im'!x?\\Im'and the line bundle over the sphere S2n\u00001, de\fned by\nthe eigenspace corresponding to the eigenvalue c1\u00001 ofRx. Since any line bundle over a sphere\nis trivial, we have a map J:S'(\u00181)!S'(\u00181) such that Jx=vxfor anyx2S'(\u00181), wherevis a\nglobal unit section of the line bundle. To simplify the writing, we put \u0015=c1\u00001 and\u0016=c2\u00001.\nThen, with the following sequence of claims, we proceed along the same lines as the authors made\nin [16], which the reader is referred to for details.\nClaim (a). The mapJsatis\fesJ2(x) =\u0000xandJ(\u0000x) =\u0000J(x) for anyx2S'(\u00181).\nConsidered the 2-plane Vx=spanfx;Jxg, ifwis a unit vector in Vxthen there exists \u00122[0;2\u0019[\nsuch thatw= cos(\u0012)x+ sin(\u0012)Jx. De\fning z(w) =\u0000sin(\u0012)x+ cos(\u0012)Jx, one proves that z(w)\nis eigenvector of Rwcorresponding to \u0015, thenz(w) =\u0006Jw. Using this last formula, it follows\nJ2(x) =\u0000xandJ(\u0000x) =\u0000J(x) for anyx2S'(\u00181).\nClaim (b). J: Im'!Im'is linear.\nThe mapJis extended to Im 'puttingJ(ax) =aJ(x), wherea2R. Assuming that J(cos(\u0012)x+\nsin(\u0012)y) = cos(\u0012)Jx+sin(\u0012)Jyfor all angles \u0012and anyx,yunit vectors such that y?Vx, we obtain\nJ(x0+y0) =J(x0) +J(y0), for anyx0andy0such thaty0?V0\nx, which implies the claim.\nClaim (c). J(cos(\u0012)x+ sin(\u0012)y) = cos(\u0012)Jx+ sin(\u0012)Jyfor all angles \u0012and anyx,yunit vectors\nsuch thaty?Vx.\nLet us de\fne J0=\u0006Jand consider A\u0012= cos(\u0012)x+ sin(\u0012)y,B\u0012= cos(\u0012)Jx+ sin(\u0012)J0y. Assuming\nthatB\u0012is an eigenvector of RA\u0012, then one has B\u0012=\u0006JA\u0012, for any angle \u0012. For\u0012= 0 one has\nB\u0012=JA\u0012then the plus sign occurs. For \u0012=\u0019\n2it followsJ0y=B\u0012=JA\u0012=Jy, i.e.J0=J, that\nimplies the claim.\nClaim (d). RA\u0012(B\u0012) =\u0015B\u0012.\nNote that the claim is equivalent to proving that R(B\u0012;A\u0012;A\u0012;B\u0012) =\u0000\u0015. Expanding this last\nformula in term of x,Jx,yandJ0yone \fndsR(Jx;x;y;J0y) +R(J0y;x;y;Jx ) =\u0016\u0000\u0015. After the\nfollowing two technical lemmas one obtains the claim.\nLemma 5.2 ([16]).(1)R(z;v)w=\u0000R(z;w)vwhenv,wandzare unit vectors such that v?w\nandw;v?Vz.\n(2)R(z;v)w= 0whenv,wandzare unit vectors such that z?Vvandz;v?Vw.\n(3)2R(x;y;J0y;Jx ) =R(Jx;x;y;J0y).\n(4)2R(J0y;x;y;Jx ) =R(Jx;x;y;J0y).\nLemma 5.3 ([16]).The curvature tensor satis\fes R(Jx;x;y;J0y) =\u00062(\u0016\u0000\u0015)\n3.\nNow we give some remarks about a null vector of a Lorentz S-manifold with two characteristic\nvector \felds and then we prove a lemma.\nRemark 5.4. Let (M;';\u0018\u000b;\u0011\u000b;g),\u000b2f1;2g, be a LorentzS-manifold with timelike vector \feld\n\u00181andua null vector in TpM,p2M. SinceTM= Im'\bker', one can write uin the following\nway\nu=\u0015x+a\u00181+b\u00182;\nwherex2Im'withg(x;x) = 1. Being ua null vector, we have \u00152+b2=a2therefore there exists\n\u00122[0;2\u0019[ such that ucan be written as follows\nu=a(cos\u0012x+\u00181+ sin\u0012\u00182);\n9and it is not a restriction to use\nu= cos\u0012x+\u00181+ sin\u0012\u00182: (7)\nFor cos\u00126= 0 consider the vector w= tan\u0012\u00181+1\ncos\u0012\u00182. It is easy to check that wis a unit vector\northogonal to u, therefore\nu?=spanfu;'x;x 2;'x 2;:::xn;'xn;wg:\nAnyy2u?can be written as\ny=\u001au+\u0017y0+\u0014w; (8)\nwherey02spanf'x;x 2;'x 2;:::xn;'xng\u001aIm'p\\u?and\u001a;\u0014;\u00172R.\nWe need to de\fne two (1 ;3)-type tensors S\u0003andS\u0003putting\nS\u0003(x;y)v=e\u0011(y)e\u0011(v)x\u0000e\u0011(x)e\u0011(v)y+g(y;v)e\u0011(x)e\u0018\u0000g(x;v)e\u0011(y)e\u0018;\nS\u0003(x;y)v=\u0000g('y;'v )'2x+g('x;'v )'2y:\nRemark 5.5. Ifu2N'(\u00181) andy2Im'\\u?, then\ng(S\u0003(u;y)u;y)\u0000g(S\u0003(u;y)u;y) = 0:\nThe following lemma allows to state the expression of a curvature-like map FwhenFvanishes\non a particular type of degenerate 2-plane and it has a suitable behaviour with respect to the\ncharacteristic vector \felds.\nLemma 5.6. Let(M;';\u0018\u000b;\u0011\u000b;g),\u000b2f1;2g, be a Lorentz g:f:f -manifold with timelike vector\n\feld\u00181. Fixed a point p2M, letF: (TpM)4!Rbe a curvature-like map such that, for any\nx;y;v2Im'and any\u000b;\f;\r2f1;2g,\nF(x;\u0018\u000b;y;v) = 0; F(\u0018\u000b;x;\u0018\f;y) =\"\u000b\"\fg(x;y); F(\u0018\u000b;x;\u0018\f;\u0018\r) = 0; F(\u00181;\u00182;\u00181;\u00182) = 0:(9)\nThen the following statements are equivalent.\na)Fvanishes on any degenerate 2-plane\u0019=spanfu;yg, withu2N'(\u00181)andy2u?\\Im',\nb)F(x;y;v;z ) =g(S\u0003(x;y)v;z)\u0000g(S\u0003(x;y)v;z).\nProof. An easy computation, using Remark 5.5, shows that b))a).\nConversely, \fx p2Mand consider the curvature-like map Hsuch that, for any x;y;z;v2TpM,\nH(x;y;v;z ) =F(x;y;z;w )\u0000g(S\u0003(x;y)v;z) +g(S\u0003(x;y)v;z): (10)\nConditiona) and Remark 5.5 imply that Hvanishes on any degenerate 2-plane spanfu;yg, for\nanyu2N'(\u00181) andy2u?\\Im'. We start proving that Hvanishes on any degenerate 2-plane.\nTo see this, let ube a null vector of TpMas in (7) such that cos \u00126= 0. By the hypotheses and\nusing (8), for any y2u?we have\ng(S\u0003(u;y)u;y) = (\u001ag('u;'u ) +\u0017g('u;'y0))2\u0000g('u;'u )\u0000\n\u001a2g('u;'u ) +\u00172g('y0;'y0)\u0001\n=\u001a2g('u;'u )2\u0000\u001a2g('u;'u )2\u0000\u00172g('y0;'y0)g('u;'u ) =\u0000\u00172g(y0;y0)g('u;'u );\ng(S\u0003(u;y)u;y) =\u0000e\u0011(u)e\u0011(u)g(y;y);\nF(u;y;u;y ) =\u00172F(u;y0;u;y0) + 2\u0014\u0017F(u;y0;u;w ) +\u00142F(u;w;u;w ) =\u00172cos2\u0012F(x;y0;x;y0)\n+ (1\u0000sin\u0012)2(\u00172g(y0;y0) +\u00142) =\u00172g('u;'u )F(x;y0;x;y0) +e\u0011(u)e\u0011(u)g(y;y)\n=\u00172g('u;'u )F(u0;y0;u0;y0)\u0000\u00172g('u;'u )g(y0;y0) +e\u0011(u)e\u0011(u)g(y;y);\n10whereu0=x+\u00181which belongs to N'(\u00181). Hence one obtains\nH(u;y;u;y ) =\u0017g('u;'u )F(u0;y0;u0;y0); (11)\nwithu0=x+\u00181andy2u?\\Im'.\nIf cos\u0012= 0, thenu=\u00181\u0006\u00182andu?=spanfug\bIm'. By direct computation, it is easy to\nverify that\nH(u;y;u;y ) = 0; (12)\nfor anyy2u?.\nEquations (11) and (12) clearly imply that Hvanishes on any degenerate 2-plane. Applying\nLemma 2.1 to Hone has\nF(x;y;v;z ) =k(g(x;v)g(y;z)\u0000g(y;v)g(x;z)) +g(S\u0003(x;y)v;z)\u0000g(S\u0003(x;y)v;z): (13)\nBy de\fnition of k, using the hypotheses and (10), we deduce\nk=H(\u0018\u000b;x;\u0018\u000b;x)\n\"\u000bg(x;x)=F(\u0018\u000b;x;\u0018\u000b;x)\u0000g(x;x)\n\"\u000bg(x;x)= 0:\nThen, substituting in (13), we obtain our assertion.\n5.2 Main results\nNow, we consider the following two standard tensor \felds of type (1 ;3), evaluating them at the\npointp:\nR0(x;y)v=g(\u0019I(y);\u0019I(v))\u0019I(x)\u0000g(\u0019I(x);\u0019I(v))\u0019I(y);\nRJ(x;y)v=g\u0000\nJ(\u0019I(y));\u0019I(v)\u0001\nJ(\u0019I(x))\u0000g\u0000\nJ(\u0019I(x));\u0019I(v)\u0001\nJ(\u0019I(y))\n+ 2g\u0000\n\u0019I(x);J(\u0019I(y))\u0001\nJ(\u0019I(v));\nwhere\u0019I:TpM!Im'is the projection on Im 'andJis an almost Hermitian structure on Im '.\nIt is useful to note that RJandR0vanish on the triplets containing a characteristic vector and\nthat they are orthogonal to \u00181and\u00182, for anyx;y;v2TpM.\nNow we are ready to prove the following result.\nTheorem 5.7. Let(M;';\u0018\u000b;\u0011\u000b;g),\u000b2f1;2gandn > 1, be a (2n+ 2) -dimensional Lorentz\nS-manifold with timelike vector \feld \u00181. The following three statements are equivalent.\na)Mis'-null Osserman with respect to \u00181and for any u2N'(\u00181)the Jacobi operator\n\u0016RujIm'\\u?has exactly two distinct eigenvalues c1andc2with multiplicities 1and2(n\u00001),\nrespectively.\nb) There exist an almost complex structure JonIm'pandc1;c22Rsuch that, for any x;y;v2\nTpM,\nR(x;y)v=S\u0003(x;y)v\u0000S\u0003(x;y)v+c2R0(x;y)v+c1\u0000c2\n3RJ(x;y)v:\nc) 1. For any v2spanf\u00181g,x2\u0018?\n1we have\nR(x;v)v= (\u00111(v))2(x\u0000e\u0011(x)\u00182):\n112. There exist an almost complex structure JonIm'pandc1;c22Rsuch that, for any\nv;y;x2\u0018?\n1, we have\nR(x;y)v=\u00112(v)\u0000\n\u00112(y)x\u0000\u00112(x)y\u0001\n+\u0000\ng(y;v)\u00112(x)\u0000g(x;v)\u00112(y)\u0001e\u0018\n+g('y;'v )'2x\u0000g('x;'v )'2y+c2R0(x;y)v+c1\u0000c2\n3RJ(x;y)v:\nProof. We begin proving a))b). Under the assumption a)by Remark 5.1 we know that Im 'pis\nendowed with an almost complex structure Jsuch thatJxis an eigenvector of \u0016Rurelated to the\neigenvaluec1. To prove b), we consider the curvature-like map FonTpMgiven by\nF(x;y;v;z ) =R(x;y;v;z ) +\u0016g(R0(x;y)v;z) +\u001cg(RJ(x;y)v;z); (14)\nwhere\u0016;\u001c2R.\nWe want to apply Lemma 5.6 to F. About the hypotheses of Lemma 5.6, we see at once that F\nsatis\fes (9) since F=Rif one of its four arguments is a characteristic vector and (2) hold. Thus\nwe must only compute F(u;y;u;y ), for any degenerate vector u2N'(\u00181) andy2u?\\Im'.\nNamely, considering a null vector u2N'(\u00181) and a vector y2u?\\Im', we \fnd the suitable\nvalues of\u0016and\u001cinRfor whichFvanishes on degenerate 2-plane \u0019=spanfu;yg.\nPuttingy1=Jx12u?, one computes\nF(y1;u;u;y 1) =\u0000g(R(y1;u)u;y1) +\u0016g(R0(y1;u)u;y1) +\u001cg(RJ(y1;u)u;y1) (15)\n=\u0000c1+\u0016+ 3\u001c:\nAnalogously, if y2andy0\n2are orthonormal eigenvectors of \u0016Ruwith respect to the eigenvalue c2,\nthen we have\nF(y2;u;u;y 2) =\u0000g(R(y2;u)u;y2) +\u0016g(R0(y2;u)u;y2) +\u001cg(RJ(y2;u)u;y2)) (16)\n= (\u0000c2+\u0016);\nF(y2;u;u;y0\n2) =\u0000g(R(y2;u)u;y0\n2) +\u0016g(R0(y2;u)u;y0\n2) +\u001cg(RJ(y2;u)u;y0\n2)) = 0; (17)\nF(y2;u;u;y 1) =\u0000g(R(y2;u)u;y1) +\u0016g(R0(y2;u)u;y1) +\u001cg(RJ(y2;u)u;y1)) = 0: (18)\nNow, imposing F= 0, we get\n\u0016=c2and\u001c=c1\u0000c2\n3: (19)\nSo, since a vector yofu?\\Im'can be written as y=ay1+bjyj\n2, wherey1andyj\n2are eigenvectors\nof\u0016Ruinu?\\\u0018?\n1corresponding to c1andc2, respectively. By (15), (16), (17) and (18) we have\nF(y;u;u;y ) =a2F(y1;u;u;y 1) +abjF(y1;u;u;yj\n2) +abkF(yk\n2;u;u;y 1)\n+bkbjF(yk\n2;u;u;yj\n2) = 0:\nTherefore, applying Lemma 5.6, we obtain F(x;y;v;z ) =g(S\u0003(x;y)v;z)\u0000g(S\u0003(x;y)v;z), for any\nx;y;v;z2TpM. Then, by (14) and (19), we get\nR(x;y;v;z ) =g(S\u0003(x;y)v;z)\u0000g(S\u0003(x;y)v;z)\u0000c2g(R0(x;y)v;z)\u0000c1\u0000c2\n3g(RJ(x;y)v;z):\nThus, one obtains\nR(x;y)v=\u0000S\u0003(x;y)v+S\u0003(x;y)v+c2R0(x;y)v+c1\u0000c2\n3RJ(x;y)v:\n12The proofb))c) is straightforward. In fact, for any v2spanf\u00181g,x2\u0018?\n1, we have\nR(x;v)v=S\u0003(x;v)v= (\u00111(v))2(x+e\u0011(x)\u00181+\"1e\u0011(x)e\u0018) = (\u00111(v))2(x\u0000e\u0011(x)\u00182);\nso obtaining c)1.\nFor anyv;y;x2\u0018?\n1, by b)one gets\nR(x;y)v=\u00112(y)\u00112(v)x\u0000\u00112(x)\u00112(v)y+ (g(y;v)\u00112(x)\u0000g(x;v)\u00112(y))e\u0018\n+ (\u0000S\u0003+c2R0+c1\u0000c2\n3RJ)(x;y)v;\nthat is c)2.\nFinally, we prove c))a). Consider u2N(\u00181),u=\u00181+x1and puty1=Jx1. One has\nR(y1;u)u=R(y1;\u00181)\u00181+R(y1;x1)\u00181+R(y1;\u00181)x1+R(y1;x1)x1:\nSo, using c)1. and c)2., we have\nR(y1;\u00181)\u00181=y1andR(y1;x1)x1= (c1\u00001)y1:\nByc)2., for anyv2\u0018?\n1, it is clear that\ng(R(y1;x1)\u00181;v) =\u0000g(R(y1;x1)v;\u00181) = 0; g(R(y1;\u00181)x1;v) =g(R(x1;v)y1;\u00181) = 0:\nOn the other hand, if v=\u00181, then\ng(R(y1;x1)\u00181;\u00181) = 0; g(R(y1;\u00181)x1;\u00181) =\u0000g(y1;x1) = 0:\nHence, \u0016Ru(y1) =c1y1.\nAnalogously, considering y22(spanfx1;y1g)?\\Im', then\nR(y2;u)u=R(y2;\u00181)\u00181+R(y2;x1)\u00181+R(y2;\u00181)x1+R(y2;x1)x1:\nAs fory1, using c), it is easy to check that R(x1;v)y2= 0 andR(y2;x1)v= 0. Moreover, applying\nc)1., we get\nR(y2;\u00181)\u00181=y2:\nThe relation c)2. implies\nR(y2;x1)x1= (c2\u00001)y2:\nTherefore we have \u0016Ru(y2) =c2y2.\nFinally, to prove the '-null Osserman condition, we have to check that every eigenvalue does\nnot depend on u2N'(\u00181). In fact, by c)we \fnd\nR(\u00182;\u00181)\u00181= 0;\nR(\u00182;x1)x1=g(x1;x1)e\u0018=\u00181+\u00182:\nIt is easy to see that, for any v2\u0018?\n1\ng(R(\u00182;\u00181)x1;v) +g(R(\u00182;x1)\u00181;v) =\u00002g(R(\u00182;x1)v;\u00181) +g(R(\u00182;v)x1;\u00181)\n= 2g(x1;v)\u0000g(x1;v) =g(x1;v):\nMoreover, since\ng(R(\u00182;\u00181)x1;\u00181) +g(R(\u00182;x1)\u00181;\u00181) =\u0000g(R(\u00182;\u00181)\u00181;x1) = 0;\none obtains R(\u00182;\u00181)x1+R(\u00182;x1)\u00181=x1. Then one gets R(\u00182;u)u=\u00182+\u00181+x1=\u00182+u, so\n\u0016Ru(\u00182) =\u00182. This proves a).\nThis concludes the proof.\n13Remark 5.8. SinceRhas to satisfy the last formula in (2), for any x;y;v;z2Im'one gets\n(1\u0000c2)P(x;y;v;z) +c1\u0000c2\n3\u0000\ng(RJ(x;y)'v;z ) +g(RJ(x;y)v;'z )\u0001\n= 0: (20)\nIf'x1is an eigenvector of \u0016Ru, withu=\u00181+x12N'(\u00181), related to the eigenvalue c1, then\n'=\u0006Jand (20) yields c1\u00004c2+ 3 = 0.\nBy Theorem 4.4, it is a simple matter to prove the following result in the particular case of the\nJacobi operator with exactly one eigenvalue.\nProposition 5.9. Let(M;';\u0018\u000b;\u0011\u000b;g),\u000b2f1;2g,n > 1be a (2n+ 2) -dimensional Lorentz S-\nmanifold with timelike vector \feld \u00181. ThenMis'-null Osserman with respect to \u00181, and the\nJacobi operator \u0016Ruju?\\Im'has a single eigenvalue \u0015, if and only if it is a Lorentz S-space form\nwith'-sectional curvature c= 0. Moreover, \u0015= 1.\nNow we end dealing with the case n= 1, which is a special case because it is clear that any\n4-dimensional Lorentz g:f:f -manifold is '-null Osserman with respect to \u00181. More precisely, for\nanyu=\u00181+x12N'(\u00181) the only eigenvector of the Jacobi operator \u0016Ruju?\\Im'is realized\ngeometrically by 'x1inu?\\\u0018?\n1. Unlike before, the eigenvalue of the Jacobi operator does not\nnecessarily have to be one, as in the case of U(2), but, when it is one, the '-sectional curvature\nwill be zero. About this case we have a non compact example. It is carried out by R4endowed\nwith the Lorentz S-structure, constructed as follows ([6]). Denoting the standard coordinates with\nfx;y;z1;z2g, we de\fne on R4two vector \felds and two 1-forms putting\n\u0018\u000b=@\n@z\u000b; \u0011\u000b=dz\u000b+ydx;\nfor any\u000b2f1;2g. The tensor \felds 'andgare given in the standard basis by\nF:=0\nBB@0\u00001 0 0\n1 0 0 0\n0y0 0\n0y0 01\nCCAG:=0\nBB@1\n20\u0000y y\n01\n20 0\n\u0000y0\u00001 0\ny0 0 11\nCCA\nrespectively. It is easy to check that ( R4;';\u0018\u000b;\u0011\u000b;g),\u000b2f1;2g, is a LorentzS-manifold with\ndi\u000berent causal type of the characteristic vector \felds. Moreover it is a Lorentz space form with\n'-sectional curvature c= 0. Therefore, by (1), one obtains\nR(X;Y;V ) =e\u0011(X)g('V;'Y )2X\n\u000b=1\u0018\u000b\u0000e\u0011(Y)g('V;'X )2X\n\u000b=1\u0018\u000b\u0000e\u0011(Y)e\u0011(V)'2X\n+e\u0011(V)e\u0011(X)'2Y;\nfor anyX;Y;V2X(R4). Since Im '=hX;YiwhereX=p\n2(@\n@x\u0000y\u00181\u0000y\u00182) andY=p\n2@\n@y, one\nhas\n\u0016Ru'Z='Z; \u0016Ru\u00182=\u00182;\nfor anyZ=aX+bYandu=\u00181+Zwherea2+b2= 1. Then the only eigenvalue of \u0016Ru,\nu2N'(\u00181), is 1.\n14References\n[1] C. Atindogbe and K.L. Duggal, Pseudo-Jacobi operators and Osserman lightlike hypersur-\nfaces, K\u0016 odai Math. J. 32(2009), no. 1, 91-108.\n[2] D.E. Blair, Geometry of manifolds with structural group U(n)\u0002O(s), J. Di\u000berential Geom.\n4(1970), 155{167.\n[3] D.E. Blair, Riemannian Geometry of Contact and Symplectic Manifolds, 2ed. , Progress in\nMathematics 203, Birkh auser, Boston, 2010.\n[4] N. Bla\u0014 zi\u0013 c, N. Bokan and P. Gilkey, A note on Osserman Lorentzian manifolds, Bull. London\nMath. Soc. 29(1997), 227-230.\n[5] L. Brunetti and A.V. Caldarella, Principal torus bundles of Lorentzian S-manifolds and the\n'-null Osserman condition, accepted for publication in K\u0016 odai Math. J. Preprint available at\narXiv:1207.5786.\n[6] L. Brunetti and A.M. Pastore, Curvature of a class of inde\fnite globally framed f-manifolds,\nBull. Math. Soc. Sci. Math. Roumanie (N.S.) 51(99) (2008), no. 3, 183{204.\n[7] L. Brunetti and A. M. Pastore, Examples of inde\fnite globally framed f-structures on compact\nLie groups, Publ. Math. Debrecen 80/1-2 (2012), 215{234.\n[8] Q.S. Chi, A curvature characterization of certain locally rank-one symmetric space, J. Di\u000ber-\nential Geom. 28(1988), 187{202.\n[9] Q.S. Chi, Quaternionic K ahler manifolds and a curvature characterization of two-point homo-\ngeneous space, Illinois J. Math. 35(1991), 408{418.\n[10] Q.S. Chi, Curvature characterization and classi\fcation of rank-one symmetric space, Paci\fc\nJ. Math. 150(1991), 31{42.\n[11] L. Di Terlizzi and J.J. Konderak, Examples of a generalization of contact metric structures\non \fbre bundles, J. Geom. 87(2007), no. 1-2, 31{49.\n[12] S. Kobayashi and K. Nomizu Foundations of Di\u000berential Geometry , Vol. I, II Interscience\nPublish., New York, 1963,1969.\n[13] E. Garc\u0013 \u0010a-R\u0013 \u0010o and D. Kupeli, 4 -Dimensional Osserman Lorentzian manifolds , New Develop.\nin Di\u000b. Geom. (Debrecen, 1994), 201{211, Math. Appl., 350, Kluwer Acad. Publ., Dordrecht,\n1996.\n[14] E. Garc\u0013 \u0010a-R\u0013 \u0010o, D. Kupeli and M.E. V\u0013 azquez-Abal, One problem of Osserman in Lorentzian\ngeometry, Di\u000berential Geom. Appl. 7(1997), 85{100.\n[15] E. Garc\u0013 \u0010a-R\u0013 \u0010o, D. Kupeli and R. V\u0013 azquez-Lorenzo, Osserman manifolds in semi-Riemannian\ngeometry , Lecture Notes in Mathematics, Vol. 1777, Springer-Verlag, Berlin, 2002.\n[16] P. Gilkey, A. Swann and L. Vanhecke, Isoparametric geodesic spheres and a conjecture of\nOsserman concerning the Jacobi operator, Quart. J. Math. Oxford 46(1995), 299{320.\n[17] S.I. Goldberg and K. Yano, On normal globally framed f-manifolds, T^ ohoku Math. J. 22\n(1970), 362{370.\n[18] Y. Nikolayevsky, Osserman manifolds of dimension 8, Manuscripta Math. 115(2004), 31{53.\n15[19] Y. Nikolayevsky, Osserman Conjecture in dimension n6= 8;16, Math. Ann. 331(2005), 505{\n522.\n[20] Y. Nikolayevsky, On Osserman manifolds of dimension 16, Contemporary Geometry and Re-\nlated Topics, Proc. Conf. Belgrade, 2005 (2006), 379-398.\n[21] B. O'Neill, Semi-Riemannian geometry , Academic Press, New York, 1983.\n[22] R. Osserman, Curvature in the eighties, Amer. Math. Monthly 97(1990), 731-756\n[23] T. Takahashi, Sasakian manifold with pseudo-Riemannian metric, T^ ohoku Math. J. (2) 21\n(1969), 271{290.\n16" }, { "title": "1905.11562v1.Lorentz_Violation_and_Riemann_Finsler_Geometry.pdf", "content": "arXiv:1905.11562v1 [hep-ph] 26 May 2019Proceedings of the Eighth Meeting on CPT and Lorentz Symmetr y (CPT’19), Indiana University, Bloomington, May 12–16, 2019\n1\nLorentz Violation and Riemann-Finsler Geometry\nBenjamin R. Edwards\nPhysics Department, Indiana University,\nBloomington, Indiana 47405, USA\nThe general charge-conserving effective scalar field theory incorporating viola-\ntions of Lorentz symmetry is presented. The dispersion rela tion is used to infer\nthe effect of spin-independent Lorentz violation on point pa rticle motion. A\nlarge class of associated Finsler spaces is derived, and the properties of these\nspaces are explored.\n1. Introduction\nConnections between Riemann-Finsler spaces and theories with Lor entz\nviolation have recently been uncovered.1A lack of physical examples is\nan obstacle on the path toward developing a strong intuition about F insler\nspaces. Inthe firstsection, thegeneraleffectivequadraticsca larfieldtheory\nincorporating violation of Lorentz symmetry will be developed. In th e next\nsection, a method to generate the lagrangian describing the motion of an\nanalogue point particle experiencing spin-independent Lorentz viola tion is\nderived. The last section explores the properties of these Finsler s paces.\nThese proceedings are based on results in Ref. 2.\n2. Field theory\nFor a complex scalar field φpropagating in an n-dimensional Minkowski\nspacetime with metric ηµν, the quadratic Lagrange density incorporating\nLorentz violation is\nL(φ,φ†) =∂µφ†∂µφ−m2φ†φ+1\n2/bracketleftbig\n∂µφ†(/hatwidekc)µν∂νφ−iφ†(/hatwideka)µ∂µφ+ h.c./bracketrightbig\n.(1)\nThe Lorentz violation is realized by the CPT-odd operator ( /hatwideka)µ, and the\nCPT-evenoperator( /hatwidekc)µν, eachofwhichcanincludecoefficientsforLorentz\nviolation associated with operators of arbitrarily large mass dimensio nd.\nThe hermiticity of Limplies these operators are themselves hermitian. InProceedings of the Eighth Meeting on CPT and Lorentz Symmetr y (CPT’19), Indiana University, Bloomington, May 12–16, 2019\n2\nthe special case of hermitian scalar fields, the term involving ( /hatwideka)µis pro-\nportional to a total derivative. It follows that CPT symmetry is gua ranteed\nwhenφ=φ†.\nField redefinitions can eliminate any traces present in the coefficients\nfor Lorentz violation by absorbing them into the terms with lower mas s\ndimension. We can therefore take them to be traceless without loss of\ngenerality. The commutativity of derivatives implies that they are to tally\nsymmetric in all their indices. From these considerations, it is found t hat\nthe coefficients contain (2 d−n+2)(d−1)!/(d−n+2)(n−2)!independent\ncomponents.\nThe dispersion relation for this theory is found to be\np2−m2+(/hatwidekc)µνpµpν−(/hatwideka)µpµ= 0, (2)\nwhere the operators ( /hatwidekc)µνand (/hatwideka)µare expressed in momentum space as\n(/hatwidekc)µν=∞/summationdisplay\nd=n(k(d)\nc)µνα1α2···αd−npα1pα2···pαd−n,\n(/hatwideka)µ=/summationdisplay\nd=n−1(k(d)\na)µα1···αd−n+1pα1pα2···pαd−n+1, (3)\nwith the sums running over even powers of p. For brevity, both types of\ncoefficients will be expressed without the aorcsubscripts in what follows,\nand the appropriate sign difference will be absorbed into the k(d)coefficient\nwhere the CPT properties will be determined by the mass dimension d.\n3. Classical kinematics\nA method has been developed to extract point particle lagrangians f rom a\ngiven field theory.3Using the three equations\nR(p) = 0, (4)\n∂p0\n∂Pj=−uj\nu0, (5)\nL=−uµpµ, (6)\nthe idea is to identify the centroid of a localized wave packet with the\npoint particle. The method starts with the dispersion relation R(p). Next,\nthe components of the classical velocity uµof the particle are related to\nthe group velocity of the corresponding wave packet. The last equ ation is\nrequired by translation invariance of L, along with the requirement that L\nbe one-homogeneousin the velocity, L(λu) =λL(u). The first two relationsProceedings of the Eighth Meeting on CPT and Lorentz Symmetr y (CPT’19), Indiana University, Bloomington, May 12–16, 2019\n3\ncan then be used to eliminate the components of the n-momentum to write\nLonly as a function of the velocity uµ.\nThese equations can be combined to produce a quadratic polynomial in\nL. For the case d=n, solving this quadratic leads to the exact lagrangian.\nFor the nonminimal cases d≥5, corrections to the usual lagrangian can\nbe determined through an iterative procedure. The process begin s with\nan expansion in ( k(d)) of the roots of the quadratic. Call this root L=\nL(uµ,pµ,(k(d))). Define L0=L(uµ,pµ,0) =−√uµηµνuν, and then Lj=\nL(uµ,p(j−1)\nµ(uµ),(k(d))) where p(j)\nµ=−∂Lj/∂uµ. This leads to\nL(d)\n3=L(d)\n0/bracketleftbig\n1−1\n2/tildewidek(d)−1\n8(d−n+1)2(/tildewidek(d))2\n+1\n8(d−n+2)2/tildewidek(d)\nα/tildewidek(d)α−1\n16(d−n+1)4(/tildewidek(d))3\n+1\n16(d−n+1)(d−n+2)2(2d−2n+1)/tildewidek(d)/tildewidek(d)α/tildewidek(d)α\n−1\n16(d−n+1)(d−n+2)3/tildewidek(d)α/tildewidek(d)αβ/tildewidek(d)β/bracketrightbig\n, (7)\nwhere the\n/tildewidek(d)=mn−d(k(d))α1α2···αd−n+2ˆuα1ˆuα2···ˆuαd−n+2,\n/tildewidek(d)\nα1···αl=mn−d(k(d))α1···αlαl+1···αd−n+2ˆuαl+1···ˆuαd−n+2,(8)\ncontain the directional dependence ˆ uµ=uµ/u,u=√uµηµνuν. This La-\ngrangian matches the first order correction found by Reis and Sch reck for\nthe nonminimal fermion sector using an ansatz-based technique.4\n4. Finsler geometry\nThe Finsler structure (or Finsler norm) plays a central role in the st udy of\nFinsler spaces. Classical lagrangians satisfy many of the requireme nts in\nthe definition of Finsler structures. Though there are important d ifferences\npreventing the lagrangians derived above from fulfilling the definition of a\nFinsler structure, a method exists to generate associated Finsler structures\nfrom a given lagrangian.5\nFor the lagrangians developed from the scalar field theories discuss ed\nabove, the prescription to generate a Finsler structure in this cas e is given\nbypx(u)→(−i)npx(y),k(d)x→ink(d)x,L→ −F=−y·p,ux→(i)nyx.\nAs a demonstration of the kinds of geometrical quantities one can c alcu-\nlate in these spaces, we use the Finsler space associated with the fir st order\nlimit of the lagrangian given in Eq. (7). The Finsler structure associat ed\nwith this lagrangian is\nF(d)=y−1\n2y/tildewidek(d). (9)Proceedings of the Eighth Meeting on CPT and Lorentz Symmetr y (CPT’19), Indiana University, Bloomington, May 12–16, 2019\n4\nThe fundamental tensor of a Finsler space determines the metric a nd\ntherefore also the geodesics. The definition of this tensor is gjk=\n1\n2∂2F2/∂yj∂yk. For the limit under consideration, the fundamental ten-\nsor is given by\ng(d)\njk=rjk(1+1\n2(d−n)/tildewidek(d))−1\n2(d−n+2)(d−n+1)/tildewidek(d)\njk\n+1\n2(d−n)(d−n+2)[/tildewidek(d)jˆyk+/tildewidek(d)kˆyj−/tildewidek(d)ˆyjˆyk].(10)\nInspection shows gjkreduces to a purely riemannian one for the cases d=n\nandd=n−2. This is consistent with the fact that the d=ncoefficient\ncan be absorbed into the metric at the level of the field theory, while a\nd=n−2 coefficient would correspond to a rescaling of the mass term.\nThe situation is not as straightforward for other values of mass dim en-\nsion. It has been demonstrated that the nonvanishing of the Cart an torsion\nimplies non-riemmannian nature of the underlying space.6Calculation of\nthis tensor shows it vanishes for d=nandd=n−2, and also in the case\nofn= 1, which represents a Riemann curve, but is nonvanishing in other\ncases. Calculation of the Matsumoto torsion7shows only d=n−1 reduces\nto a Randers metric.\nThe geodesics are obtained from the geodesic equation Fd\ndλ(yj/F) +\nGj= 0. A calculation shows the geodesic spray coefficients Gjare\n1\ny2Gj=1\n2/tildewideDj/tildewidek(d)+1\n2(d−n)ˆyj/tildewideD•/tildewidek(d)\n−1\n2(d−n+2)rjl/tildewideD•/tildewidek(d)\nl+/tildewideγj\n••, (11)\nwhere a•subscript denotes contraction of ˆ yjwith a lower jindex, with all\ncontractions exterior to any derivatives.\nIt is clear from this expression that if the backgroundfield is covaria ntly\nconstant with respect to the riemannian metric, /tildewideDj/tildewidek(d)l= 0, then the\ngeodesics are unaffected. This situation was conjectured to hold in general\nand is confirmed here, but remains unproved.\nReferences\n1. V.A. Kosteleck´ y, Phys. Rev. D 69, 105009 (2004).\n2. B.R. Edwards and V.A. Kosteleck´ y, Phys. Lett. B 786, 319 (2018).\n3. V.A. Kosteleck´ y and N. Russell, Phys. Lett. B 693, 443 (2010).\n4. J.A.A.S. Reis and M. Schreck, Phys. Rev. D 97, 065019 (2018).\n5. V.A. Kosteleck´ y, Phys. Lett. B 701, 137 (2011).\n6. A. Deicke, Arch. Math. 4, 45 (1953).\n7. M. Matsumoto, Tensor 24, 29 (1972),\nM. Matsumoto and S. H¯ oj¯ o , Tensor 32, 225 (1978)." }, { "title": "1807.09730v1.Regularity_and_asymptotic_behaviour_for_a_damped_plate_membrane_transmission_problem.pdf", "content": "arXiv:1807.09730v1 [math.AP] 25 Jul 2018REGULARITY AND ASYMPTOTIC BEHAVIOUR FOR A\nDAMPED PLATE-MEMBRANE TRANSMISSION PROBLEM\nBIENVENIDO BARRAZA MART ´INEZ, ROBERT DENK,\nJAIRO HERN ´ANDEZ MONZ ´ON, FELIX KAMMERLANDER, AND MAX NENDEL\nAbstract. We consider a transmission problem where a structurally\ndamped plate equation is coupled with a damped or undamped wa ve equa-\ntion by transmission conditions. We show that exponential s tability holds\nin the damped-damped situation and polynomial stability (b ut no exponen-\ntial stability) holds in the damped-undamped case. Additio nally, we show\nthat the solutions first defined by the weak formulation, in fa ct have higher\nSobolev space regularity.\n1.Introduction\nIn this paper, we study a coupled plate-membrane system, whe re we assume\nstructural damping for the plate and damping / no damping for the wave equa-\ntion. More precisely, we consider the following geometric s ituation: Let Ω ⊂R2\nbe a bounded C4-domain with boundary Γ, and let Ω 2⊂Ω be a non-empty\nboundedC4-domain satisfying Ω2⊂Ω. We set Ω 1:= Ω\\Ω2andI:=∂Ω2. Then\nIis the interface between Ω 1and Ω2(see Figure 1for the geometric situation).\nByν, we denote the outer unit normal with respect to Ω 1both on Γ and on I.\nIn Ω1∪Ω2, we consider the plate-membrane (plate-wave) system\nutt+∆2u−ρ∆ut= 0 in(0,∞)×Ω1, (1.1)\nwtt−∆w+βwt= 0 in(0,∞)×Ω2, (1.2)\nwhereρ≥0 andβ≥0 are fixed constants. For ρ >0, we have structural\ndamping for the plate equation ( 1.1), whereas the coefficient β≥0 describes\nthe damping (or the absence of damping) for the wave equation (1.2). On the\nouter boundary Γ, we impose clamped (Dirichlet) boundary co nditions\n(1.3) u=∂νu= 0 on(0,∞)×Γ.\nOn the interface I, we have transmission conditions of the form\nu=won(0,∞)×I, (1.4)\nB1u= 0 on(0 ,∞)×I, (1.5)\nFinancial Support through DAAD, COLCIENCIAS via Project 12 1571250194 and the\nGermanResearchFoundationviaCRC1283“TamingUncertaint y”isgratefullyacknowledged.\nDate: July 26, 2018.\n2010Mathematics Subject Classification. 74K20; 74H40; 35B40; 35Q74.\nKey words and phrases. Plate-membrane equation, transmission problem, asymptot ic\nbehaviour.\n12 B. BARRAZA MART ´INEZ ET AL.\nνν\nΩ2Ω1\nΓ\nI\nFigure 1. The set Ω = Ω 1∪I∪Ω2.\nB2u−ρ∂νut=−∂νwon(0,∞)×I (1.6)\nwith\nB1u:= ∆u+(1−µ)B1uand B2u:=∂ν∆u+(1−µ)∂τB2u,\nwhere\nB1u:=−/a\\}b∇acketle{tτ,(∇2u)τ/a\\}b∇acket∇i}htandB2u:=/a\\}b∇acketle{tτ,(∇2u)ν/a\\}b∇acket∇i}ht.\nHere,µ∈/parenleftbig\n0,1\n2/parenrightbig\nis Poisson’s ratio and τ:= (−ν2,ν1)⊤. As we have a coupling of\na fourth-order equation with a second-order equation, we ha ve two transmission\nconditions (( 1.4) and (1.6)) and one boundary condition ( 1.5) on the interface\nI.\nFinally, the boundary-transmission problem ( 1.1)–(1.6) is endowed with ini-\ntial conditions of the form\nu|t=0=u0, ut|t=0=u1in Ω1, (1.7)\nw|t=0=w0, wt|t=0=w1in Ω2. (1.8)\nThe aim of the present paper is to investigate well-posednes s as well as reg-\nularity and stability of the solution of ( 1.1)–(1.8). Note that we omitted all\nphysical constants for simplicity. Concerning the modelli ng of plate-membrane\nsystems and more detailed models including physical consta nts, we refer to,\ne.g., [6], [15], and [19].\nIt is well known that the structurally damped plate equation itself has expo-\nnential stability and leads to the generation of an analytic C0-semigroup even in\ntheLp-setting, see [ 12] and the references therein. Due to the hyperbolic struc-\nture of the wave equation ( 1.2),Lp-theory is not feasible for the coupled system,\nand we will consider the plate-membrane system in an L2-framework. It is not\nhard to see the for all ρ≥0 andβ≥0 we have well-posedness, i.e. generation\nof aC0-semigroup in the corresponding L2-Sobolev spaces (see Theorem 2.2\nbelow). The main results of the present paper state that we ha ve exponential\nstability if both dampings are present ( ρ >0 andβ >0) but no exponential\nstability if the wave equation is undamped ( β= 0), see Theorems 3.1and3.2.\nIn the case of a structurally damped plate equation and an und amped wave\nequation (ρ >0 andβ= 0) we obtain polynomial stability (Theorem 5.2).\nMoreover, the “good” parabolic structure of the damped plat e equation implies\nhigh elliptic regularity for uandw(Theorem 4.5). In particular, the transmis-\nsion conditions ( 1.4)–(1.6) hold in the sense of boundary traces.A PLATE-MEMBRANE TRANSMISSION PROBLEM 3\nThere is a huge amount of literature on transmission problem s for elastic\nsystems, most of them dealing with wave-wave systems. For wa ve-plate trans-\nmission problems, we mention [ 17], where Kelvin-Voigt damping for the plate\nequation is considered (see also [ 18] for the one-dimensional case). In [ 5] expo-\nnential stability was obtained for a damped wave / damped pla te transmission\nproblem under some geometric condition which leads to a flat i nterface. This\nwas generalized in [ 26] to a model with curved middle surface by virtue of geo-\nmetric multiplier method. In [ 16], stabilization of a damped wave / damped\nplate system with variable coefficients is studied by means of a Riemannian\ngeometrical approach. For stability of coupled wave-plate systems within the\nsame domain, we mention, e.g., [ 21].\nWhereas the above mentioned results show exponential stabi lity for many\ncases of damped-damped systems, this cannot be expected in t he damped-\nundamped situation where we have, from a mathematical point of view, a\nparabolic-hyperbolic coupled system (see, e.g., [ 7], [8], [13] for heat-wave sys-\ntems).\nFor transmission problems in (thermo-)viscoelasticity, w e mention, e.g., [ 3],\n[4], [14], [23], and [24]. In particular, in [ 24] polynomial stability for a (thermo-)\nviscoelastic damped-undamped system with Kelvin-Voigt da mping has been\nshown. The proof is based on an extended version of a characte rization of\npolynomial stability due to Borichev and Tomilov [ 9]. It turns out that some\narguments in [ 24] can be adapted to the plate-wave situation considered in\nthe present paper to show that the system is not exponentiall y but polyno-\nmially stable (Section 5). We remark that our proof of polyno mial stability is\nbased on rather general methods which should be applicable f or other transmis-\nsion problems. However, by this method we do not obtain optim al polynomial\nrates. The proof of higher regularity (Section 4) uses argum ents similar to [ 11]\nwheredampedplate/undampedplatetransmissionproblemsw ereinvestigated.\nIn particular, we apply the classical theory of parameter-d ependent boundary\nvalue problems (see [ 2]) to obtain sufficiently good estimates in the damped\npart.\nThe structure of the paper is as follows: In Section 2, we defin e the basic\nspacesandoperatorsandshowthegenerationofa C0-semigroupofcontractions.\nExponential stability for ρ >0 andβ >0 and non-exponential stability for\nβ= 0 is shown in Section 3, whereas the proof of higher regulari ty based on\nparameter-elliptic theorycanbefoundinSection4. Finall y, polynomialstability\nforρ>0 andβ= 0 is proven in Section 5.\n2.Well-posedness\nWe denote by H2\nΓ(Ω1) the space of all u∈H2(Ω1) withu|Γ=∂νu|Γ= 0. On\nH2\nΓ(Ω1) we consider the inner product\n/a\\}b∇acketle{tu,v/a\\}b∇acket∇i}htH2\nΓ(Ω1):=/integraldisplay\nΩ1∇2u:∇2v+µ[u,v]dx,\nwhere\n∇2u:∇2v:=ux1x1vx1x1+ux2x2vx2x2+2ux1x2vx1x24 B. BARRAZA MART ´INEZ ET AL.\nand\n[u,v] :=ux1x1vx2x2+ux2x2vx1x1−2ux1x2vx1x2\nfor allu,v∈H2\nΓ(Ω1). We thus have that\n/a\\}b∇acketle{tu,v/a\\}b∇acket∇i}htH2\nΓ(Ω1)=µ/a\\}b∇acketle{t∆u,∆v/a\\}b∇acket∇i}htL2(Ω1)+(1−µ)/a\\}b∇acketle{t∇2u,∇2v/a\\}b∇acket∇i}htL2(Ω1)4\nfor allu,v∈H2\nΓ(Ω1). By Poincar´ e’s inequality, we have that\n/ba∇dblu/ba∇dbl2\nH2(Ω1)≤C/parenleftbig\n/ba∇dbl∇u/ba∇dbl2\nL2(Ω1)2+/ba∇dbl∇2u/ba∇dbl2\nL2(Ω1)4/parenrightbig\n=C/parenleftbig\n/ba∇dblux1/ba∇dbl2\nL2(Ω1)+/ba∇dblux2/ba∇dbl2\nL2(Ω1)+/ba∇dbl∇2u/ba∇dbl2\nL2(Ω1)4/parenrightbig\n≤C/parenleftbig\n/ba∇dbl∇ux1/ba∇dbl2\nL2(Ω1)2+/ba∇dbl∇ux2/ba∇dbl2\nL2(Ω1)2+/ba∇dbl∇2u/ba∇dbl2\nL2(Ω1)4/parenrightbig\n≤C/ba∇dbl∇2u/ba∇dbl2\nL2(Ω1)4\nforu∈H2\nΓ(Ω1). Here and in the following, Cdenotes a generic constant which\nmay change at each appearance. The above estimate shows that /ba∇dbl· /ba∇dblH2\nΓ(Ω1)is\nequivalent to the H2(Ω1)-norm onH2\nΓ(Ω1). In particular,/parenleftbig\nH2\nΓ(Ω1),/a\\}b∇acketle{t·,·/a\\}b∇acket∇i}htH2\nΓ(Ω1)/parenrightbig\nis a Hilbert space.\nWe will also use the following result on integration by parts .\nLemma 2.1. (See[10], p. 27.) For u∈H4(Ω1)∩H2\nΓ(Ω1)andv∈H2\nΓ(Ω1)it\nholds\n(2.1) /a\\}b∇acketle{t∆2u,v/a\\}b∇acket∇i}htL2(Ω)=/a\\}b∇acketle{tu,v/a\\}b∇acket∇i}htH2\nΓ(Ω1)−/a\\}b∇acketle{tB1u,∂νv/a\\}b∇acket∇i}htL2(I)+/a\\}b∇acketle{tB2u,v/a\\}b∇acket∇i}htL2(I).\nLet\nH:=/braceleftbig\nU= (u1,v1,u2,v2)⊤∈H2\nΓ(Ω1)×L2(Ω1)×H1(Ω2)×L2(Ω2):u1|I=u2|I/bracerightbig\nbe endowed with the inner product\n/a\\}b∇acketle{tU,/tildewideU/a\\}b∇acket∇i}htH:=/a\\}b∇acketle{tu1,/tildewideu1/a\\}b∇acket∇i}htH2\nΓ(Ω1)+/a\\}b∇acketle{tv1,/tildewidev1/a\\}b∇acket∇i}htL2(Ω1)\n+/a\\}b∇acketle{t∇u2,∇/tildewideu2/a\\}b∇acket∇i}htL2(Ω2)2+/a\\}b∇acketle{tv2,/tildewidev2/a\\}b∇acket∇i}htL2(Ω2)\nforU,/tildewideU∈H. Then ( H,/a\\}b∇acketle{t·,·/a\\}b∇acket∇i}htH) is a Hilbert space. Note that we can omit the\nterm/a\\}b∇acketle{tu2,/tildewideu2/a\\}b∇acket∇i}htL2(Ω2)by Poincar´ e’s inequality as u1χΩ1+u2χΩ2∈H1\n0(Ω). Here,\nχΩjstands for the characteristic function of Ω j.\nWe introduce the operator matrix Agiven by\nA:=\n0 1 0 0\n−∆2ρ∆ 0 0\n0 0 0 1\n0 0 ∆ −β\n.\nBy (2.1), we have that\n/a\\}b∇acketle{tAU,/tildewideU/a\\}b∇acket∇i}htH=/a\\}b∇acketle{tv1,/tildewideu1/a\\}b∇acket∇i}htH2\nΓ(Ω1)−/a\\}b∇acketle{t∆2u1−ρ∆v1,/tildewidev1/a\\}b∇acket∇i}htL2(Ω1)\n+/a\\}b∇acketle{t∇v2,∇/tildewideu2/a\\}b∇acket∇i}htL2(Ω2)2+/a\\}b∇acketle{t∆u2−βv2,/tildewidev2/a\\}b∇acket∇i}htL2(Ω2)\n=/a\\}b∇acketle{tv1,/tildewideu1/a\\}b∇acket∇i}htH2\nΓ(Ω1)+/a\\}b∇acketle{t∇v2,∇/tildewideu2/a\\}b∇acket∇i}htL2(Ω2)2−/a\\}b∇acketle{tu1,/tildewidev1/a\\}b∇acket∇i}htH2\nΓ(Ω1)\n−/a\\}b∇acketle{tB2u1,/tildewidev1/a\\}b∇acket∇i}htL2(∂Ω1)+/a\\}b∇acketle{tB1u1,∂ν/tildewidev1/a\\}b∇acket∇i}htL2(∂Ω1)\n−ρ/a\\}b∇acketle{t∇v1,∇/tildewidev1/a\\}b∇acket∇i}htL2(Ω1)2+ρ/a\\}b∇acketle{t∂νv1,/tildewidev1/a\\}b∇acket∇i}htL2(∂Ω1)A PLATE-MEMBRANE TRANSMISSION PROBLEM 5\n−/a\\}b∇acketle{t∇u2,∇/tildewidev2/a\\}b∇acket∇i}htL2(Ω2)2−β/a\\}b∇acketle{tv2,/tildewidev2/a\\}b∇acket∇i}htL2(Ω2)−/a\\}b∇acketle{t∂νu2,/tildewidev2/a\\}b∇acket∇i}htL2(I)\nfor all sufficiently smooth U,/tildewideU. This leads us to the following interpretation of\nthe transmssion conditions ( 1.5) and (1.6): we say that Usatisfies the trans-\nmission conditions (1.5)and(1.6)weaklyif the equality\n(2.2)/a\\}b∇acketle{tAU,Φ/a\\}b∇acket∇i}htH=/a\\}b∇acketle{tv1,ϕ1/a\\}b∇acket∇i}htH2\nΓ(Ω1)+/a\\}b∇acketle{t∇v2,∇ϕ2/a\\}b∇acket∇i}htL2(Ω2)2−/a\\}b∇acketle{tu1,ψ1/a\\}b∇acket∇i}htH2\nΓ(Ω1)\n−ρ/a\\}b∇acketle{t∇v1,∇ψ1/a\\}b∇acket∇i}htL2(Ω1)2−/a\\}b∇acketle{t∇u2,∇ψ2/a\\}b∇acket∇i}htL2(Ω2)2−β/a\\}b∇acketle{tv2,ψ2/a\\}b∇acket∇i}htL2(Ω2)\nholds true for all Φ = ( ϕ1,ψ1,ϕ2,ψ2)⊤∈H2\nΓ(Ω1)×H2\nΓ(Ω1)×H1(Ω2)×H1(Ω2)\nsatisfyingϕ1=ϕ2andψ1=ψ2onI.\nNow, we consider the linear operator A:D(A)⊂H→H, U/mapsto→AUwith\nD(A) :=/braceleftbig\nU∈H:v1∈H2\nΓ(Ω1),v2∈H1(Ω2),∆2u1∈L2(Ω1),∆u2∈L2(Ω2),\nv1=v2onIand (1.5),(1.6) are weakly satisfied/bracerightbig\n.\nAs\n(2.3) Re /a\\}b∇acketle{tAU,U/a\\}b∇acket∇i}htH=−ρ/ba∇dbl∇v1/ba∇dbl2\nL2(Ω1)2−β/ba∇dblv2/ba∇dbl2\nL2(Ω2)≤0\nfor allU∈D(A), the operator Ais dissipative. The same argument shows\nthat for any smooth solution ( u,w) of (1.1)-(1.6), the energy\nE(t) :=1\n2/integraldisplay\nΩ1µ|∆u(t)|2+(1−µ)|∇2u(t)|2+|ut(t)|2dx\n+1\n2/integraldisplay\nΩ2|∇w(t)|2+|wt(t)|2dx\nis decreasing and the dissipation is caused by the damping bo th in Ω 1and Ω2.\nMoreover, the system is still dissipative if only one of the d amping terms is\nactive (ρ+β >0) and the system is conservative if there is no damping at all\n(ρ=β= 0).\nIn what follows, we show that the system ( 1.1)-(1.6) is well-posed for any\nchoice ofρ≥0 andβ≥0.\nTheorem 2.2. The operator A:H⊃D(A)→Hgenerates a strongly con-\ntinuous semigroup (S(t))t≥0of contractions on H.\nProof.First, we show that 1 −Ais surjective. Let F= (f1,g1,f2,g2)⊤∈H.\nWe need to show that there exists a U= (u1,v1,u2,v2)⊤∈D(A) such that\n(1−A)U=F,i.e.\nu1−v1=f1,\nv1+∆2u1−ρ∆v1=g1,\nu2−v2=f2,\nv2−∆u2+βv2=g2.\nPlugging in vi=ui−fifori= 1,2, we have to solve\nu1+∆2u1−ρ∆u1=f1+g1−ρ∆f1, (2.4)\nu2−∆u2+βu2=f2+g2+βf2. (2.5)6 B. BARRAZA MART ´INEZ ET AL.\nMotivated by the notion of the weak transmission conditions , we introduce the\nspace\nV:={u= (u1,u2)⊤∈H2\nΓ(Ω1)×H1(Ω2) :u1=u2onI}.\nEndowed with the scalar product\n/a\\}b∇acketle{tu,/tildewideu/a\\}b∇acket∇i}htV=/a\\}b∇acketle{tu1,/tildewideu1/a\\}b∇acket∇i}htH2\nΓ(Ω1)+/a\\}b∇acketle{t∇u2,∇/tildewideu2/a\\}b∇acket∇i}htL2(Ω2)2(u,/tildewideu∈ V),\n(V,/a\\}b∇acketle{t·,·/a\\}b∇acket∇i}htV) becomes a Hilbert space.\nIn order to solve ( 2.4), (2.5), we will use the theorem of Lax-Milgram in the\nHilbert space V.Letb:V ×V → Rbe defined by\nb(u,ϕ) :=/a\\}b∇acketle{tu1,ϕ1/a\\}b∇acket∇i}htL2(Ω1)+/a\\}b∇acketle{tu1,ϕ1/a\\}b∇acket∇i}htH2\nΓ(Ω1)+ρ/a\\}b∇acketle{t∇u1,∇ϕ1/a\\}b∇acket∇i}htL2(Ω1)2\n+(1+β)/a\\}b∇acketle{tu2,ϕ2/a\\}b∇acket∇i}htL2(Ω2)+/a\\}b∇acketle{t∇u2,∇ϕ2/a\\}b∇acket∇i}htL2(Ω2)2.\nObviously, bis bilinear and continuous. Since\nb(u,u) =/ba∇dblu1/ba∇dbl2\nL2(Ω1)+/ba∇dblu1/ba∇dbl2\nH2\nΓ(Ω1)+ρ/ba∇dbl∇u1/ba∇dbl2\nL2(Ω1)2\n+(1+β)/ba∇dblu2/ba∇dbl2\nL2(Ω2)+/ba∇dbl∇u2/ba∇dbl2\nL2(Ω2)2\n≥ /ba∇dblu1/ba∇dbl2\nH2\nΓ(Ω1)+/ba∇dbl∇u2/ba∇dbl2\nL2(Ω2)2\nholds for all u∈ V,the bilinear form bis coercive on V.Hence, there exists a\nuniqueu∈ Vsatisfying\nb(u,ϕ) = Λ(ϕ) (2.6)\nfor allϕ∈ V, where the linear functional Λ: V →Ris given by\nΛ(ϕ) :=/a\\}b∇acketle{tf1+g1,ϕ1/a\\}b∇acket∇i}htL2(Ω1)+ρ/a\\}b∇acketle{t∇f1,∇ϕ1/a\\}b∇acket∇i}htL2(Ω1)2\n+/a\\}b∇acketle{tg2+(1+β)f2,ϕ2/a\\}b∇acket∇i}htL2(Ω2).\nNote that for ϕ1∈C∞\n0(Ω1) we have\n/a\\}b∇acketle{tu1,ϕ1/a\\}b∇acket∇i}htH2\nΓ(Ω1)=/a\\}b∇acketle{tu1,∆2ϕ1/a\\}b∇acket∇i}htL2(Ω1)−/a\\}b∇acketle{tu1,B2ϕ1/a\\}b∇acket∇i}htL2(∂Ω1)+/a\\}b∇acketle{t∂νu1,B1ϕ1/a\\}b∇acket∇i}htL2(∂Ω1)\n=/a\\}b∇acketle{t∆u1,∆ϕ1/a\\}b∇acket∇i}htL2(Ω1).\nIn particular, for any ( ϕ1,ϕ2)∈C∞\n0(Ω1)×C∞\n0(Ω2)⊂ V, we have that ( 2.4)\nand (2.5) are satisfied in L2(Ω1) andL2(Ω2), respectively. This implies that\n∆2u1∈L2(Ω1) and ∆u2∈L2(Ω2).We set\nU:=\nu1\nu1−f1\nu2\nu2−f2\n∈H.\nFinally, using ( 2.4), (2.5) and (2.6), we calculate\n/a\\}b∇acketle{tAU,Φ/a\\}b∇acket∇i}htH=/a\\}b∇acketle{tv1,ϕ1/a\\}b∇acket∇i}htH2\nΓ(Ω1)−/a\\}b∇acketle{t∆2u1−ρ∆v1,ψ1/a\\}b∇acket∇i}htL2(Ω1)\n+/a\\}b∇acketle{t∇v2,∇ϕ2/a\\}b∇acket∇i}htL2(Ω2)2+/a\\}b∇acketle{t∆u2−βv2,ψ2/a\\}b∇acket∇i}htL2(Ω2)\n=/a\\}b∇acketle{tv1,ϕ1/a\\}b∇acket∇i}htH2\nΓ(Ω1)−/a\\}b∇acketle{tg1+f1,ψ1/a\\}b∇acket∇i}htL2(Ω1)+/a\\}b∇acketle{tu1,ψ1/a\\}b∇acket∇i}htL2(Ω1)\n+/a\\}b∇acketle{t∇v2,∇ϕ2/a\\}b∇acket∇i}htL2(Ω2)2−/a\\}b∇acketle{tf2+g2,ψ2/a\\}b∇acket∇i}htL2(Ω2)\n=/a\\}b∇acketle{tv1,ϕ1/a\\}b∇acket∇i}htH2\nΓ(Ω1)+/a\\}b∇acketle{t∇v2,∇ϕ2/a\\}b∇acket∇i}htL2(Ω2)2A PLATE-MEMBRANE TRANSMISSION PROBLEM 7\n−ρ/a\\}b∇acketle{t∇v1,∇ψ1/a\\}b∇acket∇i}htL2(Ω1)2−β/a\\}b∇acketle{tv2,ψ2/a\\}b∇acket∇i}htL2(Ω2)\n−/a\\}b∇acketle{tu1,ψ1/a\\}b∇acket∇i}htH2\nΓ(Ω1)−/a\\}b∇acketle{t∇u2,∇ψ2/a\\}b∇acket∇i}htL2(Ω2)2\nfor any Φ = ( ϕ1,ψ1,ϕ2,ψ2)⊤∈H2\nΓ(Ω1)×H2\nΓ(Ω1)×H1(Ω2)×H1(Ω2) satisfying\nϕ1=ϕ2andψ1=ψ2onI.Therefore,Usatisfies the transmission conditions\nweakly. Hence, U∈D(A) and (1 −A)U=F.\nAsAis dissipative and 1 −Ais surjective, Agenerates a C0-semigroup of\ncontractions by the Lumer-Phillips Theorem. /square\nRemark 2.3. In the same way as in the previous proof, one can show that the\noperator Ais continuously invertible, i.e. 0 belongs to the resolvent setρ(A).\nTo show this, we now have to consider\n∆2u1=g1−ρ∆f1, (2.7)\n−∆u2=g2+βf2 (2.8)\ninstead of ( 2.4) and (2.5). The sesquilinear form Band the functional Λ are\nnow defined by B(u,ϕ) :=/a\\}b∇acketle{tu,ϕ/a\\}b∇acket∇i}htVand\nΛ(ϕ) :=/a\\}b∇acketle{tg1,ϕ1/a\\}b∇acket∇i}htL2(Ω1)+ρ/a\\}b∇acketle{t∇f1,∇ϕ1/a\\}b∇acket∇i}htL2(Ω1)2+/a\\}b∇acketle{tg2,ϕ2/a\\}b∇acket∇i}htL2(Ω2)+β/a\\}b∇acketle{tf2,ϕ2/a\\}b∇acket∇i}htL2(Ω2)\nforu= (u1,u2),ϕ= (ϕ1,ϕ2)∈ V. The Riesz Representation Theorem implies\nthat there exists a unique solution u= (u1,u2)∈ Vsatisfying\n(2.9) B(u,ϕ) = Λ(ϕ) for all ϕ∈ V.\nIn particular, choosing ( ϕ1,ϕ2)∈C∞\n0(Ω1)×C∞\n0(Ω2)⊂ Vwe see that ( 2.7)\nand (2.8) hold in the sense of distributions in Ω 1and Ω2,respectively. As the\nright-hand side of ( 2.7) belongs to L2(Ω1), the same holds for the left-hand\nside, i.e. ∆2u1∈L2(Ω1). In the same way, we see that ( 2.8) holds as equality\ninL2(Ω2) and therefore ∆ u2∈L2(Ω2). Now, set\n(2.10) vi:=−fifori= 1,2.\nThenU:= (u1,v1,u2,v2)⊤∈Hand\n(2.11) ∆2u1−ρ∆v1=g1,−∆u2+βv2=g2.\nIn the same way as in the proof of Theorem 2.2, one sees that Usatisfies the\ntransmission conditions weakly. Therefore Ubeolongs to D(A) and satisfies\n−AU=F.\nOn other hand, if /tildewideU∈D(A) solves −A/tildewideU=F, thenB(/tildewideu,ϕ) = Λ(ϕ)\nholds for all ϕ∈ Vdue to the definition of D(A) and the weak transmission\nconditions. Therefore U=/tildewideU, and Ais a bijection. Since Ais the generator of\naC0−semigroup by Theorem 2.2,Ais closed and hence 0 ∈ρ(A).\n3.Results on exponential stability\nIn this section, we study exponential stability of the semig roup (S(t))t≥0\ngenerated by A. First, we consider the case where we have damping in both\nsub-domains, i.e., ρ>0 andβ >0. Itis nosurprisethat in this case exponential\nstability holds.8 B. BARRAZA MART ´INEZ ET AL.\nTheorem 3.1. Letρ >0andβ >0. Then the semigroup (S(t))t≥0is expo-\nnentially stable, i.e., for any U0∈D(A)andU(t) :=S(t)U0(t≥0)we have\nE(t)≤Ce−κtE(0)with positive constants Candκ, whereE(t) :=1\n2/ba∇dblU(t)/ba∇dbl2\nH.\nProof.LetU(t) = (u1(t),v1(t),u2(t),v2(t))⊤=S(t)U0withU0∈D(A). For\nthe energy E(t) we obtain\n(3.1)E′(t) = Re/a\\}b∇acketle{tAU(t),U(t)/a\\}b∇acket∇i}htH=−ρ/ba∇dbl∇v1(t)/ba∇dbl2\nL2(Ω1)2−β/ba∇dblv2(t)/ba∇dbl2\nL2(Ω2).\nWe defineF(t) :=/a\\}b∇acketle{tu1(t),v1(t)/a\\}b∇acket∇i}htL2(Ω1)+/a\\}b∇acketle{tu2(t),v2(t)/a\\}b∇acket∇i}htL2(Ω2)fort≥0. Then\n|F(t)| ≤1\n2/parenleftig\n/ba∇dblu1(t)/ba∇dbl2\nL2(Ω1)+/ba∇dblv1(t)/ba∇dbl2\nL2(Ω1)+/ba∇dblu2(t)/ba∇dbl2\nL2(Ω2)+/ba∇dblv2(t)/ba∇dbl2\nL2(Ω2)/parenrightig\n.\nBy definition of H, we haveu1(t) =u2(t) on the interface I, and therefore\nthe function u(t) :=u1(t)χΩ1+u2(t)χΩ2belongs to H1\n0(Ω) for all t≥0. An\napplication of Poincar´ e’s inequality yields\n/ba∇dblu1(t)/ba∇dbl2\nL2(Ω1)+/ba∇dblu2(t)/ba∇dbl2\nL2(Ω2)=/ba∇dblu(t)/ba∇dbl2\nL2(Ω)≤C/ba∇dbl∇u(t)/ba∇dbl2\nL2(Ω)2\n=C/parenleftig\n/ba∇dbl∇u1(t)/ba∇dbl2\nL2(Ω1)2+/ba∇dbl∇u2(t)/ba∇dbl2\nL2(Ω2)2/parenrightig\n≤C/parenleftig\n/ba∇dblu1(t)/ba∇dbl2\nH2\nΓ(Ω1)+/ba∇dbl∇u2(t)/ba∇dbl2\nL2(Ω2)2/parenrightig\n.\nTherefore, for some constant c1>0 we get\n(3.2) |F(t)| ≤c1\n2/ba∇dblU(t)/ba∇dbl2\nH=c1E(t).\nUsingU′(t) =AU(t), we obtain\nF′(t) =/a\\}b∇acketle{tu′\n1(t),v1(t)/a\\}b∇acket∇i}htL2(Ω1)+/a\\}b∇acketle{tu1(t),v′\n1(t)/a\\}b∇acket∇i}htL2(Ω1)\n+/a\\}b∇acketle{tu′\n2(t),v2(t)/a\\}b∇acket∇i}htL2(Ω2)+/a\\}b∇acketle{tu2(t),v′\n2(t)/a\\}b∇acket∇i}htL2(Ω2)\n=/ba∇dblv1(t)/ba∇dbl2\nL2(Ω1)−/a\\}b∇acketle{tu1(t),∆2u1(t)−ρ∆v1(t)/a\\}b∇acket∇i}htL2(Ω1)\n+/ba∇dblv2(t)/ba∇dbl2\nL2(Ω2)+/a\\}b∇acketle{tu2(t),∆u2(t)−βv2(t)/a\\}b∇acket∇i}htL2(Ω2).\nNow we use the fact that U(t)∈D(A) and take Φ := (0 ,u1(t),0,u2(t))⊤in\nthe weak transmission conditions ( 2.2). We obtain\nF′(t) =/ba∇dblv1(t)/ba∇dbl2\nL2(Ω1)+/ba∇dblv2(t)/ba∇dbl2\nL2(Ω2)+/a\\}b∇acketle{tΦ,AU(t)/a\\}b∇acket∇i}htH\n=/ba∇dblv1(t)/ba∇dbl2\nL2(Ω1)+/ba∇dblv2(t)/ba∇dbl2\nL2(Ω2)−/ba∇dblu1(t)/ba∇dbl2\nH2\nΓ(Ω1)−/ba∇dbl∇u2(t)/ba∇dbl2\nL2(Ω2)2\n−ρ/a\\}b∇acketle{t∇u1(t),∇v1(t)/a\\}b∇acket∇i}htL2(Ω1)2−β/a\\}b∇acketle{tu2(t),v2(t)/a\\}b∇acket∇i}htL2(Ω2).\nBy Young’s inequality and Poincar´ e’s inequality in Ω 1, for every δ >0 there\nexists aCδ>0 such that\n−ρ/a\\}b∇acketle{t∇u1(t),∇v1(t)/a\\}b∇acket∇i}htL2(Ω1)2≤ρδ/ba∇dbl∇u1(t)/ba∇dbl2\nL2(Ω1)2+ρCδ/ba∇dbl∇v1(t)/ba∇dbl2\nL2(Ω1)2\n≤c2ρδ/ba∇dbl∇2u1(t)/ba∇dbl2\nL2(Ω1)2+ρCδ/ba∇dbl∇v1(t)/ba∇dbl2\nL2(Ω1)2\n≤c3ρδ/ba∇dblu1(t)/ba∇dbl2\nH2\nΓ(Ω1)+ρCδ/ba∇dbl∇v1(t)/ba∇dbl2\nL2(Ω1)2. (3.3)A PLATE-MEMBRANE TRANSMISSION PROBLEM 9\nIn the same way, using Poincar´ e’s inequality in Ω,\n−β/a\\}b∇acketle{tu2(t),v2(t)/a\\}b∇acket∇i}htL2(Ω2)≤βδ/ba∇dblu2(t)/ba∇dbl2\nL2(Ω2)+βCδ/ba∇dblv2(t)/ba∇dbl2\nL2(Ω2)\n≤/tildewidec3βδ/parenleftbig\n/ba∇dblu1(t)/ba∇dbl2\nH2\nΓ(Ω1)+/ba∇dbl∇u2(t)/ba∇dbl2\nL2(Ω2)2/parenrightbig\n+βCδ/ba∇dblv2(t)/ba∇dbl2\nL2(Ω2).(3.4)\nChoosingδsmall enough such that ( c3ρ+/tildewidec3β)δ≤1\n2, we get from ( 3.3) and\n(3.4) (again using Poincar´ e’s inequality for v1(t) in Ω1)\nF′(t)≤c4/parenleftbig\n/ba∇dbl∇v1(t)/ba∇dbl2\nL2(Ω1)2+/ba∇dblv2(t)/ba∇dbl2\nL2(Ω2)/parenrightbig\n−1\n2/parenleftbig\n/ba∇dblu1(t)/ba∇dbl2\nH2\nΓ(Ω1)+/ba∇dbl∇u2(t)/ba∇dbl2\nL2(Ω2)2/parenrightbig\n. (3.5)\nNow letL(t) :=c5E(t) +F(t), where the constant c5satisfiesc5≥2c1and\nmin{ρ,β}c5≥c4+1\n2. By (3.1) and (3.5) we see that\n(3.6)L′(t)≤ −1\n2/parenleftig\n/ba∇dblu1(t)/ba∇dbl2\nH2\nΓ(Ω1)+/ba∇dbl∇v1(t)/ba∇dbl2\nL2(Ω1)2\n+/ba∇dbl∇u2(t)/ba∇dbl2\nL2(Ω2)2+/ba∇dblv2(t)/ba∇dbl2\nL2(Ω2)/parenrightig\n≤ −CE(t).\nAs|F(t)| ≤c1E(t)≤c5\n2E(t), we obtain\nc5\n2E(t)≤L(t)≤3c5\n2E(t).\nTherefore, ( 3.6) yieldsL′(t)≤ −κL(t) with some positive constant κ. By Gron-\nwall’s lemma, L(t)≤e−κtL(0) which yields\nE(t)≤CL(t)≤Ce−κtL(0)≤Ce−κtE(0). /square\nNow let us consider the case where the membrane is not damped, i.e.,β= 0.\nIn this situation, we show that the system is not exponential ly stable, nomatter\nifρ >0 orρ= 0. The proof of the following theorem follows an idea of [ 24,\nTheorem 3.5].\nTheorem 3.2. Forβ= 0andρ≥0, the system is not exponentially stable.\nProof.We consider the closed subspace\nH0:={0}×{0}×H1\n0(Ω2)×L2(Ω2)\nofH. OnH0we consider the C0-semigroup ( /tildewideS(t))t≥0with the generator\n/tildewiderA:H0⊃D(/tildewiderA)→H0, U/mapsto→\n1 0 0 0\n0 1 0 0\n0 0 0 1\n0 0 ∆ 0\nU\nwhereD(/tildewiderA) :={0} × {0} ×(H2(Ω2)∩H1\n0(Ω2))×H1\n0(Ω2). In the sequel, we\nwill show that S(t)−/tildewideS(t):H0→His compact. For U0∈H0, we consider\nE(t) :=1\n2/vextenddouble/vextenddoubleS(t)U0−/tildewideS(t)U0/vextenddouble/vextenddouble2\nH10 B. BARRAZA MART ´INEZ ET AL.\nfort≥0. Then, we denote by ( u,ut,w,wt)⊤:=S(t)U0the solution of the\ntransmission problem ( 1.1)–(1.6) and (0,0,/tildewidew,/tildewidewt)⊤:=/tildewideS(t)U0the solution of\nthewave equationinΩ 2withhomogeneousDirichlet boundaryconditions.Then\nz:=w−/tildewidewsolves the wave equation ztt−∆z= 0 in Ω 2withz|I=w|I=u|I.\nTherefore, applying the weak transmission conditions to\n/a\\}b∇acketle{tAS(t)U0,S(t)U0−/tildewideS(t)U0/a\\}b∇acket∇i}htH\nand using integration by parts for /a\\}b∇acketle{t∆/tildewidew(t),zt(t)/a\\}b∇acket∇i}htL2(Ω2), we obtain\nE′(t) = Re/parenleftig\n/a\\}b∇acketle{tu(t),ut(t)/a\\}b∇acket∇i}htH2\nΓ(Ω1)+/a\\}b∇acketle{tut(t),utt(t)/a\\}b∇acket∇i}htL2(Ω1)\n+/a\\}b∇acketle{t∇z(t),∇zt(t)/a\\}b∇acket∇i}htL2(Ω2)2+/a\\}b∇acketle{tzt(t),ztt(t)/a\\}b∇acket∇i}htL2(Ω2)/parenrightig\n= Re/parenleftig\n/a\\}b∇acketle{tu(t),ut(t)/a\\}b∇acket∇i}htH2\nΓ(Ω1)+/a\\}b∇acketle{tut(t),−∆2u(t)+ρ∆ut(t)/a\\}b∇acket∇i}htL2(Ω1)\n−/a\\}b∇acketle{t∂νz(t),zt(t)/a\\}b∇acket∇i}htL2(I)+/a\\}b∇acketle{tB2u(t)+ρ∂νut(t),ut(t)/a\\}b∇acket∇i}htL2(I)/parenrightig\n=−ρ/ba∇dbl∇ut(t)/ba∇dbl2\nL2(Ω1)2+Re(/a\\}b∇acketle{t∂ν/tildewidew(t),ut(t)/a\\}b∇acket∇i}htL2(I)).\nThis implies that\n(3.7)E(t)+/integraldisplayt\n0ρ/ba∇dbl∇ut(s)/ba∇dbl2\nL2(Ω1)2ds=/integraldisplayt\n0Re(/a\\}b∇acketle{t∂ν/tildewidew(s),ut(s)/a\\}b∇acket∇i}htL2(I))ds.\nNow, let (Uk\n0)k∈N⊂H0beaboundedsequence. We define /tildewidewkandukas/tildewidewandu\nbutwithU0beingreplaced by Uk\n0fork∈N.Then,asthesequence/parenleftbig\n∂ν/tildewidewk/parenrightbig\nk∈N⊂\nL2/parenleftbig\n[0,t];L2(I)/parenrightbig\nis uniformly bounded, there exists a subsequence of ( /tildewidewk)k∈N\nwhich will again be denoted by ( /tildewidewk)k∈Nsuch that/parenleftbig\n∂ν/tildewidewk/parenrightbig\nk∈Nconverges weakly\ninL2/parenleftbig\n[0,t];L2(I)/parenrightbig\n. Moreover, the sequences ( uk\nt)k∈N⊂L2/parenleftbig\n[0,t];H2(Ω1)/parenrightbig\nand\n(uk\ntt)k∈N⊂L2/parenleftbig\n[0,t];L2(Ω1)/parenrightbig\nare both uniformly bounded. By the Aubin-Lions\nLemma, there exists a subsequence of ( uk)k∈N, which will again be denoted by\n(uk)k∈Nsuch that (uk\nt)k∈N⊂L2/parenleftbig\n[0,t];H1(Ω1)/parenrightbig\nconverges. As the trace\nH1(Ω1)→L2(I), v/mapsto→v|I\nis continuous, we obtain that ( uk\nt)k∈N⊂L2/parenleftbig\n[0,t];L2(I)/parenrightbig\nis convergent. For\nk,l∈Nwe now denote by\nEkl(t) :=1\n2/vextenddouble/vextenddoubleS(t)(Uk\n0−Ul\n0)−/tildewideS(t)(Uk\n0−Ul\n0)/vextenddouble/vextenddouble2\nH.\nThen, by ( 3.7) we get that\nEkl(t)≤/integraldisplayt\n0/angbracketleftbig\n∂ν/tildewidewkl(s),ukl\nt(s)/angbracketrightbig\nL2(I)ds=/angbracketleftbig\n∂ν/tildewidewkl,ukl\nt/angbracketrightbig\nL2/parenleftbig\n[0,t];L2(I)/parenrightbig→0\nask,l→ ∞,where/tildewidewklanduklaredefinedas /tildewidewandubutwithU0beingreplaced\nbyUk\n0−Ul\n0fork,l∈N. Therefore, (( S(t)−/tildewideS(t))Uk\n0)k∈Nis a Cauchy sequence in\nHand thus convergent. This shows the compactness of S(t)−/tildewideS(t):H0→H.\nAs/tildewideS(t) is the semigroup related to the wave equation, its essentia l spectral\nradius equals 1. An application of [ 24, Theorem 3.3] gives that the essentialA PLATE-MEMBRANE TRANSMISSION PROBLEM 11\nspectral radius of S(t) equals 1, too, and thus ( S(t))t≥0is not exponentially\nstable. /square\n4.Higher regularity\nIn this section, we show that the functions in the domain of Ahave higher\nregularity, which implies that the transmission condition s hold in the strong\nsense of traces. For this, we need some results from the theor y of parameter-\nelliptic boundary value problems developed in the 1960’s ([ 2], see also [ 1]).\nLet Ω⊂R2be a domain, and let A(D) =/summationtext\n|α|≤2maα∂αbe a linear differential\noperatorinΩoforder2 m.ThenA(D)iscalledparameter-ellipticiftheprincipal\nsymbolA(iξ) :=/summationtext\n|α|=2maα(iξ)αsatisfies\nλ−A(iξ)/\\e}atio\\slash= 0 (Reλ≥0, ξ∈R2,(λ,ξ)/\\e}atio\\slash= 0).\nLetB1(D),...,B m(D) be linear boundary operators on ∂Ω of the form\nBj(D) =/summationtext\n|β|≤mjbjβ∂βof ordermj<2mwith principal symbols Bj(iξ) :=/summationtext\n|β|=mjbjβ(iξ)β. Then we say that the boundary value problem is parameter-\nelliptic ifA(D) is parameter-elliptic and if the following Shapiro-Lopat inskii\ncondition holds:\n(SL)Letx0∈∂Ω, and rewrite the boundary value problem ( A(D),\nB1(D),...,B m(D)) in the coordinate system associated with x0, which is ob-\ntained from the original one by a rotation after which the pos itivex2-axis has\nthe direction of the interior normal to ∂Ω atx0. Then the trivial solution w= 0\nis the only stable solution of the ordinary differential equat ion on the half-line\n(λ−A(iξ1,∂2))w(x2) = 0 (x2∈(0,∞)),\nBj(iξ1,∂2)w(0) = 0 (j= 1,...,m)\nfor allξ1∈Rand Reλ≥0 with (ξ1,λ)/\\e}atio\\slash= 0.\nIt was shown in [ 2] that the operator corresponding to a parameter-elliptic\nboundary value problem generates an analytic C0-semigroup in L2(Ω). We will\napply these results to ∆2and ∆ in Ω 1and Ω 2, respectively, with different\nboundary operators.\nLemma 4.1. The operator −∆2inΩ1, supplemented with the boundary oper-\nators B1andB2on∂Ω1, is parameter-elliptic. The same holds for −∆2with\nclamped boundary conditions u=∂νu= 0on∂Ω1and for−∆2with boundary\nconditionsu=B1u= 0on∂Ω1.\nProof.Obviously, the operator −∆2with symbol −(ξ2\n1+ξ2\n2)2is parameter-\nelliptic. Let x0∈∂Ω1, and choose a coordinate system associated with x0.\nThen thex1-axis is in tangential direction, while the positive x2-axis coincides\nwith the inner normal direction. In these coordinates, we ha ve to solve the\nordinary differential equation\n(4.1)/parenleftbig\nλ+(∂2\n2−ξ2\n1)2/parenrightbig\nw(x2) = 0 (x2∈(0,∞)),\n(B1(iξ1,∂2)w)(0) = 0,\n(B2(iξ1,∂2)w)(0) = 0.12 B. BARRAZA MART ´INEZ ET AL.\nBy the definition of the boundary operators B1andB2, we obtain the local\nsymbols B1(iξ1,∂2)w= (∂2\n2−µξ2\n1)wandB2(iξ1,∂2)w=/parenleftbig\n−∂3\n2+(2−µ)ξ2\n1∂2/parenrightbig\nw.\nNow we use the following identity for w∈H2((0,∞)), which is obtained by\nintegration by parts in (0 ,∞):\n(4.2)/a\\}b∇acketle{t(∂2\n2−ξ2\n1)2w,w/a\\}b∇acket∇i}htL2((0,∞))=µ/ba∇dbl(∂2\n2−ξ2\n1)w/ba∇dbl2\nL2((0,∞))\n+(1−µ)/parenleftig\n/ba∇dblξ2\n1w/ba∇dbl2\nL2((0,∞))+/ba∇dbl∂2\n2w/ba∇dbl2\nL2((0,∞))+2/ba∇dblξ1∂2w/ba∇dbl2\nL2((0,∞))/parenrightig\n+/parenleftbig\nB1(iξ1,∂2)w/parenrightbig\n(0)∂2w(0)+/parenleftbig\nB2(iξ1,∂2)w/parenrightbig\n(0)w(0).\nNote that this can be seen as a localized version of ( 2.1).\nLetwbe a stable solution of ( 4.1). We multiply the firstline in ( 4.1) byw(x2)\nand integrate over x2∈(0,∞). Due to the boundary conditions, all boundary\nterms in ( 4.2) disappear, and we obtain\n0 =/a\\}b∇acketle{t(λ+(∂2\n2−ξ2\n1)2)w,w/a\\}b∇acket∇i}htL2((0,∞))\n=λ/ba∇dblw/ba∇dbl2\nL2((0,∞))+µ/ba∇dbl(∂2\n2−ξ2\n1)w/ba∇dbl2\nL2((0,∞))\n+(1−µ)/parenleftig\n/ba∇dblξ2\n1w/ba∇dbl2\nL2((0,∞))+/ba∇dbl∂2\n2w/ba∇dbl2\nL2((0,∞))+2/ba∇dblξ1∂2w/ba∇dbl2\nL2((0,∞))/parenrightig\n.\nAs Reλ≥0 andµ∈(0,1), we can take the real part and obtain\n/ba∇dblξ2\n1w/ba∇dblL2((0,∞))= 0 and therefore w= 0 in the case ξ1/\\e}atio\\slash= 0. Ifξ1= 0, then\nλ/\\e}atio\\slash= 0, and we obtain λ/ba∇dblw/ba∇dbl2L2((0,∞))= 0 which again implies w= 0. Therefore,\nthe Shapiro-Lopatinskii condition (SL) holds.\nThe statement for the other combinations of boundary condit ions follows\nexactly in the same way, as in all cases the boundary terms in ( 4.2) disappear.\n/square\nWe will apply parameter-elliptic theory to a boundary value problem in Ω 1\nwith clamped boundary conditions on Γ and free boundary cond itions onI. In\nthe next lemma, we show that the resolvent of such boundary va lue problems\nwith ‘mixed’ boundary conditions exists and satisfies a unif orm estimate.\nLemma 4.2. Consider the boundary value problem\n(4.3)(λ+∆2)u=finΩ1,\nu=∂νu= 0onΓ,\nB1u=B2u= 0onI.\nThen there exists a λ0>0such that for all λ≥λ0and for all f∈L2(Ω1)\nthere exists a unique solution u∈H4(Ω1)of(4.3). Moreover, for all λ≥λ0\nthe uniform a priori-estimate\n(4.4) /ba∇dblu/ba∇dblH4(Ω1)+λ/ba∇dblu/ba∇dblL2(Ω1)≤C1/ba∇dblf/ba∇dblL2(Ω1)\nholds with a constant C1depending on λ0but not on λorf.\nProof.(i)We first show the existence of a solution. Let f∈L2(Ω1). We choose\nϕ1∈C∞(Ω1) with 0 ≤ϕ1≤1,ϕ1= 1 in a neighbourhood of Γ, and supp ϕ1∩\nI=∅. We setϕ2:= 1−ϕ1onΩ1. Further, let ψj∈C∞(Ω1),j= 1,2, with\n0≤ψj≤1,ψj= 1 on supp ϕj, suppψ1∩I=∅, and suppψ2∩Γ =∅.A PLATE-MEMBRANE TRANSMISSION PROBLEM 13\nBy Lemma 4.1, the boundary value problem given by −∆2and clamped\nboundary conditions is parameter-elliptic. Therefore (se e [2, Theorem 5.1]) for\nλ≥λ0with sufficiently large λ0there exists a unique solution u(1)=R1(λ)ψ1f\nof\n(λ+∆2)u(1)=ψ1fin Ω1,\nu(1)=∂νu(1)= 0 on∂Ω1.\nIn the same way, using parameter-ellipticity of the boundar y value problem\n(−∆2,B1,B2), there exists a unique solution u(2)=R2(λ)ψ2fof\n(λ+∆2)u(2)=ψ2fin Ω1,\nB1u(2)=B2u(2)= 0 on∂Ω1.\nMoreover, the a priori-estimate\n(4.5) /ba∇dblu(j)/ba∇dblH4(Ω1)+λ/ba∇dblu(j)/ba∇dblL2(Ω1)≤c2/ba∇dblψjf/ba∇dblL2(Ω1)\nholds for all λ≥λ0with a constant c2independent of λandf(see [2, Theo-\nrem 4.1]).\nForλ≥λ0, we define\nR(λ)f:=ϕ1R1(λ)ψ1f+ϕ2R2(λ)ψ2f.\nBy the product rule,\n(λ+∆2)R(λ)f=ϕ1(λ+∆2)R1(λ)ψ1f+ϕ2(λ+∆2)R2(λ)ψ2f\n+S1(D)R1(λ)ψ1f+S2(D)R2(λ)ψ2f,\nwhereS1(D) andS2(D) are linear partial differential operators of order 3 de-\npending on the choice of ϕ1, but not on λorf. As (λ+ ∆2)Rj(λ)ψjf=ψjf\nandϕjψj=ϕj,j= 1,2, we obtain\n(4.6) ( λ+∆2)R(λ)f= (1+T(λ))f\nwithT(λ)f:=S1(D)R1(λ)ψ1f+S2(D)R2(λ)ψ2f. AsSj(D) are bounded linear\noperators from H3(Ω1) toL2(Ω1), we can estimate\n/ba∇dblSj(D)Rj(λ)ψjf/ba∇dblL2(Ω1)≤C/ba∇dblRj(λ)ψjf/ba∇dblH3(Ω1)\n≤δ/ba∇dblRj(λ)ψjf/ba∇dblH4(Ω1)+Cδ/ba∇dblRj(λ)ψjf/ba∇dblL2(Ω1)\n≤c2δ/ba∇dblf/ba∇dblL2(Ω1)+c2\nλCδ/ba∇dblf/ba∇dblL2(Ω1).\nHere we used the interpolation inequality and ( 4.5). Now we first choose δ>0\nsmall enough such that c2δ≤1\n4and thenλ≥λ0withλ0large enough such\nthatc2\nλCδ≤1\n4. Therefore, the norm of T(λ) as a bounded operator in L2(Ω1)\nis not larger than1\n2, and 1+T(λ) is invertible. So we can define\nu:=R(λ)(1+T(λ))−1f∈H4(Ω1).\nFrom (4.6) we see (λ+∆2)u=f, and by definition of R(λ) we have\nu|Γ= (R1(λ)ψ1f)|Γ= 0, ∂νu|Γ= 0\nas well as\nBju|I=Bj/parenleftbig\nϕ2R2(λ)ψ2f/parenrightbig\n|I=BjR2(λ)ψ2f|I= 0 (j= 1,2).14 B. BARRAZA MART ´INEZ ET AL.\nTherefore,uis a solution of the boundary value problem ( 4.3).\n(ii)Now we show that every solution of ( 4.3) satisfies the a priori-estimate\n(4.4). Letu∈H4(Ω1) be a solution of ( 4.3). Thenu(1):=uϕ1is a solution of\nthe boundary value problem\n(λ+∆2)u(1)=ϕ1f+/tildewideS1(D)uin Ω1,\nu(1)=∂νu(1)= 0 on∂Ω1,\nwhere/tildewideS1(D) is a linear partial differential operator of order 3. By param eter-\nelliptic theory [ 2, Theorem 4.1], u(1)satisfies\n/ba∇dblu(1)/ba∇dblH4(Ω1)+λ/ba∇dblu(1)/ba∇dblL2(Ω1)≤C/parenleftbig\n/ba∇dblf/ba∇dblL2(Ω1)+/ba∇dblu/ba∇dblH3(Ω1)/parenrightbig\n.\nThe same holds for u(2):=ϕ2uby parameter-ellipticity of ( −∆2,B1,B2). For\nthe sumu=u(1)+u(2), we get\n/ba∇dblu/ba∇dblH4(Ω1)+λ/ba∇dblu/ba∇dblL2(Ω1)≤C/parenleftbig\n/ba∇dblf/ba∇dblL2(Ω1)+/ba∇dblu/ba∇dblH3(Ω1)/parenrightbig\n.\nNow, by interpolation inequality again, we can estimate\n/ba∇dblu/ba∇dblH4(Ω1)+λ/ba∇dblu/ba∇dblL2(Ω1)≤C/ba∇dblf/ba∇dblL2(Ω1)+δ/ba∇dblu/ba∇dblH4(Ω1)+Cδ/ba∇dblu/ba∇dblL2(Ω1).\nChoosingδ≤1\n2and thenλ0>2Cδ, we can absorb the u-dependent terms on\nthe right-hand side and obtain\n/ba∇dblu/ba∇dblH4(Ω1)+λ/ba∇dblu/ba∇dblL2(Ω1)≤C/ba∇dblf/ba∇dblL2(Ω1)(λ≥λ0).\nThis also yields uniqueness of the solution. /square\nCorollary 4.3. Letf∈L2(Ω1),g1∈H7/2(Γ),g2∈H5/2(Γ),h1∈H3/2(I),\nandh2∈H1/2(I). Then for sufficiently large λ0>0, the boundary value prob-\nlem\n(4.7)(λ0+∆2)u=finΩ1,\nu=g1onΓ,\n∂νu=g2onΓ,\nB1u=h1onI,\nB2u=h2onI\nhas a unique solution u∈H4(Ω1). Moreover, the a priori-estimate\n(4.8)/ba∇dblu/ba∇dblH4(Ω1)≤C2/parenleftig\n/ba∇dblf/ba∇dblL2(Ω1)+/ba∇dblg1/ba∇dblH7/2(Γ)+/ba∇dblg2/ba∇dblH5/2(Γ)\n+/ba∇dblh1/ba∇dblH3/2(I)+/ba∇dblh2/ba∇dblH1/2(I)/parenrightig\nholds with a constant C2>0which depends on λ0but not onuor on the data.\nProof.We defineG:= (g1,g2,0,0)⊤on Γ andH:= (0,0,h1,h2−(divν)h1)⊤\nonI. By [25], Section 4.7.1, p. 330, the map\nR:u/mapsto→/parenleftbig\nu|∂Ω1,∂νu|∂Ω1,∂2\nνu|∂Ω1,∂3\nνu|∂Ω1,/parenrightbig⊤\nisaretractionfrom H4(Ω1)to/producttext3\nj=0H4−j−1/2(∂Ω1).LetEdenoteacoretraction\ntoR, and set\nu(1):=E/parenleftbig\nχΓG+χIH/parenrightbig\n∈H4(Ω1).A PLATE-MEMBRANE TRANSMISSION PROBLEM 15\nThe boundary operators B1andB2can be expressed in terms of normal and\ntangential derivatives as (see [ 20], Propositions 3C.7 and 3C.11)\nB1u(1)=∂2\nνu(1)+µ∂2\nτu(1)+µ(divν)∂νu(1),\nB2u(1)=∂3\nνu(1)+∂ν∂2\nτu(1)+(1−µ)∂τ∂ν∂τu(1)+∂ν[(divν)∂νu(1)].\nAsu(1)=∂νu(1)= 0 onIdue to the definition of u(1), we obtain ∂k\nτu(1)=\n∂k\nτ∂νu(1)= 0 onIfor allk∈N. Moreover, applying the identity\n∂ν∂τw=∂τ∂νw−(divν)∂τw\n(see [20, Corollary 3C.10]) to w:=u(1)and tow:=∂τu(1), respectively, we see\nthat\n∂ν∂2\nτu(1)=∂τ∂ν∂τu(1)= 0 onI.\nTherefore,\nB1u(1)=∂2\nνu(1)=h1,\nB2u(1)=∂3\nνu(1)+(divν)∂2\nνu(1)=h2\nonI. By continuity of E, we have\n(4.9)/ba∇dbl(λ0+∆2)u(1)/ba∇dblL2(Ω1)≤C/ba∇dblu(1)/ba∇dblH4(Ω1)\n≤C/parenleftig\n/ba∇dblg1/ba∇dblH7/2(Γ)+/ba∇dblg2/ba∇dblH5/2(Γ)+/ba∇dblh1/ba∇dblH3/2(I)+/ba∇dblh2/ba∇dblH1/2(I)/parenrightig\nwithCdepending only on λ0.\nConsidering u(2):=u−u(1), we see that usolves (4.7) if and only if u(2)\nsolves the boundary value problem\n(λ0+∆2)u(2)=/tildewidefin Ω1,\nu(2)=∂νu(2)= 0 on Γ ,\nB1u(2)=B2u(2)= 0 onI.\nHere,/tildewidef:=f−(λ0+ ∆2)u(1). By Lemma 4.2, this is uniquely solvable, and\nthe a priori estimate /ba∇dblu(2)/ba∇dblH4(Ω1)≤C/ba∇dbl/tildewidef/ba∇dblL2(Ω1)in connection with ( 4.9) yields\n(4.8). /square\nRemark 4.4. a) The statement and the proof of Lemma 4.2and Corollary 4.3\nare independent of the particular equation. We have shown un ique solvabil-\nity and uniform a priori-estimates for boundary value probl ems where we have\ndifferent boundary operators on disjoint and not connected pa rts of the bound-\nary, given that on each part of the boundary the Shapiro-Lopa tinskii condition\nholds.\nb) From elliptic theory, it is well known that the analog stat ement of Corol-\nlary4.3also holds (with λ0= 0) in the much easier situation of the Dirich-\nlet Laplacian in Ω 2: For every f∈L2(Ω2) andg∈H3/2(I) there exists\na uniqueu∈H2(Ω2) with ∆u=fin Ω2andu|I=g, and/ba∇dblu/ba∇dblH2(Ω2)≤\nC(/ba∇dblf/ba∇dblL2(Ω2)+/ba∇dblg/ba∇dblH3/2(I)).\nThe elliptic regularity results above are the key for the str ong solvability of\nthe transmission problem, i.e. for higher regularity of the weak solution.16 B. BARRAZA MART ´INEZ ET AL.\nTheorem 4.5. LetU= (u1,v1,u2,v2)⊤∈D(A). Thenu1∈H4(Ω1)and\nu2∈H2(Ω2). In particular, the transmission conditions hold in the stro ng sense\nof traces on the interface I.\nProof.LetU∈D(A) andF= (f1,g1,f2,g2)⊤:=AU. Thenv1=f1∈\nH2\nΓ(Ω1),v2=f2∈H1(Ω2), ∆u2=g2+βf2, and ∆2u1=ρ∆f1−g1.\nBy Remark 4.4b), there exists a unique /tildewideu2∈H2(Ω2) such that ∆ /tildewideu2=\ng2+βf2in Ω1and/tildewideu2|I=u1|I. Asu2−/tildewideu2belongs to H1\n0(Ω2) and is a weak\nsolution of ∆( u2−/tildewideu2) = 0, we immediately obtain /tildewideu2=u2which already yields\nu2∈H2(Ω2).\nSimilarly, by Corollary 4.3there exists a unique solution /tildewideu1∈H4(Ω1) of the\nboundary value problem\n(4.10)(λ0+∆2)/tildewideu1=λ0u1+ρ∆f1−g1in Ω1,\n/tildewideu1=∂ν/tildewideu1= 0 on Γ ,\nB1/tildewideu1= 0,B2/tildewideu1=ρ∂νf1−∂νu2onI.\nNote here that λ0u1+ρ∆f1−g1∈L2(Ω1) andρ∂νf1+∂νu2∈H1/2(I), and\nthat all boundary conditions hold in the trace sense.\nLetψ1∈H2\nΓ(Ω1). Then ( 2.1) in combination with the boundary conditions\nabove yields\n(4.11)/a\\}b∇acketle{t(λ0+∆2)/tildewideu1,ψ1/a\\}b∇acket∇i}htL2(Ω1)=λ0/a\\}b∇acketle{t/tildewideu1,ψ1/a\\}b∇acket∇i}htL2(Ω1)+/a\\}b∇acketle{t/tildewideu1,ψ1/a\\}b∇acket∇i}htH2\nΓ(Ω1)\n+/a\\}b∇acketle{tρ∂νf1−∂νu2,ψ1/a\\}b∇acket∇i}htL2(I).\nWe compare /tildewideu1with the weak solution u1. For this, we consider Φ :=\n(0,ψ1,0,ψ2)⊤withψ1∈H2\nΓ(Ω1),ψ2∈H1(Ω2), andψ1=ψ2onI. By defi-\nnition ofD(A), we obtain\n/a\\}b∇acketle{tAU,Φ/a\\}b∇acket∇i}htH=/a\\}b∇acketle{t−∆2u1+ρ∆v1,ψ1/a\\}b∇acket∇i}htL2(Ω1)+/a\\}b∇acketle{t∆u2−βv2,ψ2/a\\}b∇acket∇i}htL2(Ω2)\n=−/a\\}b∇acketle{tu1,ψ1/a\\}b∇acket∇i}htH2\nΓ(Ω1)−ρ/a\\}b∇acketle{t∇v1,∇ψ1/a\\}b∇acket∇i}htL2(Ω1)2\n−/a\\}b∇acketle{t∇u2,∇ψ2/a\\}b∇acket∇i}htL2(Ω2)2−β/a\\}b∇acketle{tv2,ψ2/a\\}b∇acket∇i}htL2(Ω2).\nFrom this,v1=f1and integration by parts (as we already know u2∈H2(Ω2)),\nwe see that\n/a\\}b∇acketle{t∆2u1,ψ1/a\\}b∇acket∇i}htL2(Ω1)=/a\\}b∇acketle{tρ∆v1,ψ1/a\\}b∇acket∇i}htL2(Ω1)+/a\\}b∇acketle{t∆u2,ψ2/a\\}b∇acket∇i}htL2(Ω2)\n+/a\\}b∇acketle{tu1,ψ1/a\\}b∇acket∇i}htH2\nΓ(Ω1)+/a\\}b∇acketle{tρ∇v1,∇ψ1/a\\}b∇acket∇i}htL2(Ω1)2+/a\\}b∇acketle{t∇u2,∇ψ2/a\\}b∇acket∇i}htL2(Ω2)2\n=/a\\}b∇acketle{tρ∂νf1,ψ1/a\\}b∇acket∇i}htL2(I)−/a\\}b∇acketle{t∂νu2,ψ2/a\\}b∇acket∇i}htL2(I)+/a\\}b∇acketle{tu1,ψ1/a\\}b∇acket∇i}htH2\nΓ(Ω1)\n=/a\\}b∇acketle{tρ∂νf1−∂νu2,ψ1/a\\}b∇acket∇i}htL2(I)+/a\\}b∇acketle{tu1,ψ1/a\\}b∇acket∇i}htH2\nΓ(Ω1).\nIn the last step we used ψ1=ψ2onI. Therefore, ( 4.11) also holds with /tildewideu1\nbeing replaced by u1.\nBy definition of /tildewideu1, we have (λ0+∆2)/tildewideu1= (λ0+∆2)u1=λ0u1+ρ∆f1−g1.\nTherefore, we can insert the difference w:=/tildewideu1−u1into (4.11) and obtain\n0 =λ0/a\\}b∇acketle{tw,ψ1/a\\}b∇acket∇i}htL2(Ω1)+/a\\}b∇acketle{tw,ψ1/a\\}b∇acket∇i}htH2\nΓ(Ω1)for allψ1∈H2\nΓ(Ω1). But by construction\nw∈H2\nΓ(Ω1), so we can set ψ1:=wand getw= 0, i.e.,u1=/tildewideu1∈H4(Ω1)./squareA PLATE-MEMBRANE TRANSMISSION PROBLEM 17\n5.Polynomial stability\nAs we saw in Section 3, the system is not exponentially stable when β= 0.\nWhenβ=ρ= 0, (2.3) shows that the system is conservative. In this section\nwe consider the case β= 0 andρ>0 and show that polynomial decay is still\nguaranteed under certain geometrical conditions. More pre cisely, we assume\nthat there exists some x0∈R2such that\nq·ν=q⊤ν≤0 (5.1)\nonI, whereq(x) :=x−x0. Note that νis the inner normal w.r.t. to Ω 2, which\nis why we require q·ν≤0 instead of q·ν≥0. In order to prove the polynomial\nstability, we use the following result by Borichov and Tomil ov (Theorem 2.4 in\n[9])\nTheorem 5.1. Let(T(t))t≥0be a bounded C0-semigroup on a Hilbert space\nHwith generator Asuch thatiR⊂ρ(A). Then, for fixed α >0the following\nconditions are equivalent:\n(i)There exist C >0andλ0>0such that for all λ∈Rwith|λ|> λ0and\nallF∈Hit holds\n/ba∇dbl(iλ−A)−1F/ba∇dbl ≤C|λ|α/ba∇dblF/ba∇dbl.\n(ii)There exists some C >0such that for all t >0and allU0∈D(A)it\nholds\n/ba∇dblT(t)U0/ba∇dbl ≤Ct−1\nα/ba∇dblAU0/ba∇dbl.\nWe now state the main result of this section: we show polynomi al stability for\nthe transmission problem in the case where only the plate equ ation is damped\nbut the wave equation is undamped. Using rather general meth ods, it is very\nlikely that the rate of decay is not optimal. On the other hand , the approach\nmight be versatile enough to be applicable to different transm ission problems\nof a similar form, i.e. transmission problems where the equa tion in the outer\ndomain is parameter-elliptic, whereas the equation in the i nner domain simply\nis of lower order.\nTheorem 5.2. Letβ= 0andρ>0and assume that the geometrical condtion\n(5.1)is satisfied. Then the semigroup (S(t))t≥0decays polynomially of order at\nleast1/30, i.e. there exists some constant C >0such that\n/ba∇dblS(t)/ba∇dblH≤Ct−1\n30/ba∇dblAU0/ba∇dblH\nfor allt>0andU0∈D(A).\nThroughout the remainder of this section, let λ0>0,λ∈Rwith|λ|>\nλ0,F= (f1,g1,f2,g2)⊤∈HandU= (u1,v1,u2,v2)⊤∈D(A) such that\n(iλ−A)U=F. We first observe that ( iλ−A)U=Fimplies\nv1=iλu1−f1, (5.2)\n−λ2u1+∆2u1−iλρ∆u1=g1+iλf1−ρ∆f1, (5.3)\nv2=iλu2−f2, (5.4)\n−λ2u2−∆u2=g2+iλf2. (5.5)18 B. BARRAZA MART ´INEZ ET AL.\nMultiplying ( 5.3) by−u1and (5.5) by−u2, integrating and adding yields\nλ2/parenleftbig\n/ba∇dblu1/ba∇dbl2\nL2(Ω1)+/ba∇dblu2/ba∇dbl2\nL2(Ω2)/parenrightbig\n−/a\\}b∇acketle{t∆2u1,u1/a\\}b∇acket∇i}htL2(Ω1)\n+iλρ/a\\}b∇acketle{t∆u1,u1/a\\}b∇acket∇i}htL2(Ω1)+/a\\}b∇acketle{t∆u2,u2/a\\}b∇acket∇i}htL2(Ω2)\n=−/a\\}b∇acketle{tg1+iλf1−ρ∆f1,u1/a\\}b∇acket∇i}htL2(Ω1)−/a\\}b∇acketle{tg2+iλf2,u2/a\\}b∇acket∇i}htL2(Ω2).\nUsing Lemma 2.1, integration by parts and plugging in theboundaryand trans -\nmission conditions we obtain\nλ2/parenleftbig\n/ba∇dblu1/ba∇dbl2\nL2(Ω1)+/ba∇dblu2/ba∇dbl2\nL2(Ω2)/parenrightbig\n−/ba∇dblu1/ba∇dbl2\nH2\nΓ(Ω1)−iλρ/ba∇dbl∇u1/ba∇dbl2\nL2(Ω1)2−/ba∇dbl∇u2/ba∇dbl2\nL2(Ω2)2\n=−ρ/a\\}b∇acketle{t∇f1,∇u1/a\\}b∇acket∇i}htL2(Ω1)2−/a\\}b∇acketle{tg1+iλf1,u1/a\\}b∇acket∇i}htL2(Ω1)−/a\\}b∇acketle{tg2+iλf2,u2/a\\}b∇acket∇i}htL2(Ω2).\nTaking the real part in the above equality we see that\n/ba∇dblu1/ba∇dbl2\nH2\nΓ(Ω1)+/ba∇dbl∇u2/ba∇dbl2\nL2(Ω2)2≤λ2/parenleftbig\n/ba∇dblu1/ba∇dbl2\nL2(Ω1)+/ba∇dblu2/ba∇dbl2\nL2(Ω2)/parenrightbig\n+/parenleftbig\n|λ|/ba∇dblf1/ba∇dblL2(Ω1)+/ba∇dblg1/ba∇dblL2(Ω1)/parenrightbig\n/ba∇dblu1/ba∇dblL2(Ω1)\n+/parenleftbig\n|λ|/ba∇dblf2/ba∇dblL2(Ω2)+/ba∇dblg2/ba∇dblL2(Ω1)/parenrightbig\n/ba∇dblu2/ba∇dblL2(Ω2)\n+ρ/ba∇dbl∇f1/ba∇dblL2(Ω1)2/ba∇dbl∇u1/ba∇dblL2(Ω1)2\n≤λ2/parenleftbig\n/ba∇dblu1/ba∇dbl2\nL2(Ω1)+/ba∇dblu2/ba∇dbl2\nL2(Ω2)/parenrightbig\n+C|λ|/ba∇dblU/ba∇dblH/ba∇dblF/ba∇dblH,\nwhere we used the fact that |λ| ≥λ0. Moreover, due to ( 5.2) and (5.4), we have\nthat\n/ba∇dblvj/ba∇dbl2\nL2(Ωj)=/ba∇dbliλuj−fj/ba∇dbl2\nL2(Ωj)≤2/parenleftbig\nλ2/ba∇dbluj/ba∇dbl2\nL2(Ωj)+/ba∇dblfj/ba∇dbl2\nL2(Ωj)/parenrightbig\nforj= 1,2. Hence,\n(5.6)/ba∇dblv1/ba∇dbl2\nL2(Ω1)+/ba∇dblv2/ba∇dbl2\nL2(Ω2)≤C/parenleftig\nλ2/parenleftbig\n/ba∇dblu1/ba∇dbl2\nL2(Ω1)+/ba∇dblu2/ba∇dbl2\nL2(Ω2)/parenrightbig\n+/ba∇dblF/ba∇dbl2\nH/parenrightig\nand therefore, combining ( 5.6) with the estimate for u1andu2, we get that\n/ba∇dblU/ba∇dbl2\nH≤C/parenleftig\nλ2/parenleftbig\n/ba∇dblu1/ba∇dbl2\nL2(Ω1)+/ba∇dblu2/ba∇dbl2\nL2(Ω2)/parenrightbig\n+/parenleftbig\n|λ|/ba∇dblU/ba∇dblH/ba∇dblF/ba∇dblH+/ba∇dblF/ba∇dbl2\nH/parenrightbig/parenrightig\n.\nIt remains to estimate /ba∇dblu1/ba∇dbl2\nL2(Ω1)and/ba∇dblu2/ba∇dbl2\nL2(Ω2). In order to estimate\n/ba∇dblu1/ba∇dbl2\nL2(Ω1), we observe that, due to ( 2.3), it holds\n/ba∇dbl∇v1/ba∇dbl2\nL2(Ω1)2≤1\nρ/ba∇dblU/ba∇dblH/ba∇dblF/ba∇dblH,\nand therefore, using Poincar´ e’s inequality, we obtain tha t\nλ2/ba∇dblu1/ba∇dbl2\nH1(Ω1)=/ba∇dblf1+v1/ba∇dbl2\nH1(Ω1)≤2/parenleftbig\n/ba∇dblf1/ba∇dbl2\nH1(Ω1)+/ba∇dblv1/ba∇dbl2\nH1(Ω1)/parenrightbig\n≤C/parenleftbig\n/ba∇dblF/ba∇dbl2\nH+/ba∇dbl∇v1/ba∇dbl2\nL2(Ω1)2/parenrightbig\n≤C/parenleftbig\n/ba∇dblU/ba∇dblH/ba∇dblF/ba∇dblH+/ba∇dblF/ba∇dbl2\nH/parenrightbig\n. (5.7)\nUsing the fact that |λ| ≥λ0and/ba∇dblu1/ba∇dblL2(Ω1)≤ /ba∇dblu1/ba∇dblH1(Ω1), we thus get that\n(5.8) /ba∇dblU/ba∇dbl2\nH≤C/parenleftig\nλ2/ba∇dblu2/ba∇dbl2\nL2(Ω2)+/parenleftbig\n|λ|/ba∇dblU/ba∇dblH/ba∇dblF/ba∇dblH+/ba∇dblF/ba∇dbl2\nH/parenrightbig/parenrightig\nand it remains to estimate /ba∇dblu2/ba∇dbl2\nL2(Ω2), which will be done in the following lem-\nmas.A PLATE-MEMBRANE TRANSMISSION PROBLEM 19\nLemma 5.3. It holds\nλ2/ba∇dblu2/ba∇dbl2\nL2(Ω2)≤C/parenleftbig\n|λ|/ba∇dblU/ba∇dblH/ba∇dblF/ba∇dblH+/ba∇dblF/ba∇dbl2\nH/parenrightbig\n+/integraldisplay\nI/vextendsingle/vextendsingle∂νu2(q∇u2)/vextendsingle/vextendsingledS.\nProof.Using Rellich’s identity (cf. [ 22], Eq. (2.5)), we have that\n(5.9) Re/integraldisplay\nΩ2∆u2(q∇u2)dx=−Re/integraldisplay\nI∂νu2(q∇u2)−1\n2(q·ν)|∇u2|2dS.\nWe multiply ( 5.5) byq∇u2, integrate over Ω 2, take the real part and use ( 5.9)\nin order to obtain that\nRe/parenleftbigg\n−λ2/integraldisplay\nΩ2u2(q∇u2)dx/parenrightbigg\n+Re/parenleftbigg/integraldisplay\nI∂νu2(q∇u2)−1\n2(q·ν)|∇u2|2dS/parenrightbigg\n= Re/integraldisplay\nΩ2(g2+iλf2)(q∇u2)dx.\nAsq∇u2= div(qu2)−2u2, integration by parts and taking the real part shows\nRe/integraldisplay\nΩ2u2(q∇u2)dx=−/ba∇dblu2/ba∇dbl2\nL2(Ω2)−1\n2/integraldisplay\nI(qν)|u2|2dS\nand we obtain\nλ2/ba∇dblu2/ba∇dbl2\nL2(Ω2)=−λ2\n2/integraldisplay\nI(q·ν)|u2|2dS+1\n2/integraldisplay\nI(q·ν)|∇u2|2dS\n−Re/integraldisplay\nI∂νu2(q∇u2)dS+Re/integraldisplay\nΩ2(g2+iλf2)(q∇u2)dx.\nSinceq·ν≤0 onIandu1=u2onI, we arrive at\nλ2/ba∇dblu2/ba∇dbl2\nL2(Ω2)≤C/parenleftbig\nλ2/ba∇dblu1/ba∇dbl2\nH1(Ω1)+|λ|/ba∇dblU/ba∇dblH/ba∇dblF/ba∇dblH/parenrightbig\n+/integraldisplay\nI/vextendsingle/vextendsingle∂νu2(q∇u2)/vextendsingle/vextendsingledS\n≤C/parenleftbig\n|λ|/ba∇dblU/ba∇dblH/ba∇dblF/ba∇dblH+/ba∇dblF/ba∇dbl2\nH/parenrightbig\n+/integraldisplay\nI/vextendsingle/vextendsingle∂νu2(q∇u2)/vextendsingle/vextendsingledS,\nwhere in the first step we used the trace theorem and in the last step we used\n(5.7) as well as |λ| ≥λ0. /square\nLemma 5.4. For anyε>0, there exists a constant C(ε)>0such that/integraldisplay\nI/vextendsingle/vextendsingle∂νu2(q∇u2)/vextendsingle/vextendsingledS≤ε/ba∇dblU/ba∇dbl2\nH+C(ε)|λ|60/ba∇dblF/ba∇dbl2\nH.\nProof.Using the transmission conditions, we can estimate/integraldisplay\nI/vextendsingle/vextendsingle∂νu2(q∇u2)/vextendsingle/vextendsingledS=/integraldisplay\nI/vextendsingle/vextendsingleB2u1−iλρ∂νu1+ρ∂νf1/vextendsingle/vextendsingle/vextendsingle/vextendsingleq∇u2/vextendsingle/vextendsingledS\n≤C/ba∇dblB2u1−iλρ∂νu1+ρ∂νf1/ba∇dblL2(I)/ba∇dbl∇u2/ba∇dblL2(I)2\n≤C/parenleftbig\n/ba∇dblu1/ba∇dblH7/2(Ω1)+|λ|/ba∇dblu1/ba∇dblH3/2(Ω1)+/ba∇dblF/ba∇dblH/parenrightbig\n/ba∇dblu2/ba∇dblH3/2(Ω2).\nIn order to estimate the terms on the right-hand side, we will use interpolation\ntheory for both the terms /ba∇dblu1/ba∇dblH7/2(Ω1)and/ba∇dblu2/ba∇dblH3/2(Ω2). Hence, we start with\nan estimate for /ba∇dblu2/ba∇dblH2(Ω2).\nBy (5.5),u2satisfies the equation\n∆u2=−(λ2u2+g2+iλf2).20 B. BARRAZA MART ´INEZ ET AL.\nTherefore, Remark 4.4b) andu1=u2onIyield the estimate\n/ba∇dblu2/ba∇dblH2(Ω2)≤C/parenleftbig\n/ba∇dblλ2u2+g2+iλf2/ba∇dblL2(Ω2)+/ba∇dblu1/ba∇dblH3/2(I)/parenrightbig\n≤C/parenleftbig\nλ2/ba∇dblu2/ba∇dblL2(Ω2)+|λ|/ba∇dblF/ba∇dblH+/ba∇dblu1/ba∇dblH2\nΓ(Ω1)/parenrightbig\n≤C|λ|/parenleftbig\n|λ|/ba∇dblU/ba∇dblH+/ba∇dblF/ba∇dblH/parenrightbig\n. (5.10)\nUsing interpolation inequality and the equivalence of the p-norms on R2, we get\nthat\n/ba∇dblu2/ba∇dblH3/2(Ω2)≤C/ba∇dblu2/ba∇dbl1/2\nH2(Ω2)/ba∇dblu2/ba∇dbl1/2\nH1(Ω2)≤C/ba∇dblu2/ba∇dbl1/2\nH2(Ω2)/ba∇dblU/ba∇dbl1/2\nH\n≤C/parenleftbig\n|λ|/ba∇dblU/ba∇dblH+|λ|1/2/ba∇dblU/ba∇dbl1/2\nH/ba∇dblF/ba∇dbl1/2\nH/parenrightbig\n. (5.11)\nIn the next step, we will estimate the term /ba∇dblu1/ba∇dblH7/2(Ω1). By (5.3),u1satisfies\nthe equation\n(λ+∆2)u1=λu1+λ2u1+iλρ∆u1+g1+iλf1−ρ∆f1.\nHence, Corollary 4.3states\n/ba∇dblu1/ba∇dblH4(Ω1)≤C/parenleftbig\n/ba∇dblλu1+λ2u1+iλρ∆u1+g1+iλf1−ρ∆f1/ba∇dblL2(Ω1)\n+/ba∇dblB1u1/ba∇dblH7/2(I)+/ba∇dblB2u1/ba∇dblH1/2(I)/parenrightbig\ndue to the homogeneous boundary conditions on Γ .Using the trace theorem,\nthe transmission conditions, ( 5.2) and (5.3) as well as ( 5.10), we obtain\n/ba∇dblu1/ba∇dblH4(Ω1)≤C/parenleftig\n|λ|/parenleftbig\n|λ|/ba∇dblU/ba∇dblH+/ba∇dblF/ba∇dblH/parenrightbig\n+/ba∇dbl∂νu2/ba∇dblH1/2(I)/parenrightig\n≤C|λ|/parenleftbig\n|λ|/ba∇dblU/ba∇dblH+/ba∇dblF/ba∇dblH/parenrightbig\n.\nMoreover, note that ( 5.7) reformulates to\n/ba∇dblu1/ba∇dblH1(Ω1)≤C\n|λ|/parenleftbig\n/ba∇dblU/ba∇dblH/ba∇dblF/ba∇dblH+/ba∇dblF/ba∇dbl2\nH/parenrightbig1/2.\nAgain, by interpolation inequality and the equivalence of t hep-norms on R2,\nwe thus get that\n/ba∇dblu1/ba∇dblH7/2(Ω1)≤C/ba∇dblu1/ba∇dbl5/6\nH4(Ω1)/ba∇dblu1/ba∇dbl1/6\nH1(Ω1)\n≤C|λ|5/6/parenleftbig\n|λ|/ba∇dblU/ba∇dblH+/ba∇dblF/ba∇dblH/parenrightbig5/6|λ|−1/6/parenleftbig\n/ba∇dblU/ba∇dblH/ba∇dblF/ba∇dblH+/ba∇dblF/ba∇dbl2\nH)1/12\n≤C|λ|2/3/parenleftbig\n|λ|5/6/ba∇dblU/ba∇dbl5/6\nH+/ba∇dblF/ba∇dbl5/6\nH/parenrightbig/parenleftbig\n/ba∇dblU/ba∇dbl1/12\nH/ba∇dblF/ba∇dbl1/12\nH+/ba∇dblF/ba∇dbl1/6\nH/parenrightbig\n≤C/parenleftbigg\n|λ|3/2/parenleftbig\n/ba∇dblU/ba∇dbl11/12\nH/ba∇dblF/ba∇dbl1/12\nH+/ba∇dblU/ba∇dbl5/6\nH/ba∇dblF/ba∇dbl1/6\nH/parenrightbig\n+|λ|2/3/parenleftbig\n/ba∇dblU/ba∇dbl1/12\nH/ba∇dblF/ba∇dbl11/12\nH+/ba∇dblF/ba∇dblH/parenrightbig/parenrightbigg\n. (5.12)\nYoung’s inequality\na2−αbα≤εa2+C(ε)b2\nfor fixedα∈(0,2) andε>0 arbitrary, yields\n|λ|5/2/ba∇dblU/ba∇dbl23/12\nH/ba∇dblF/ba∇dbl1/12\nH=/ba∇dblU/ba∇dbl23/12\nH/parenleftbig\n|λ|30/ba∇dblF/ba∇dblH/parenrightbig1/12≤ε/ba∇dblU/ba∇dbl2\nH+C(ε)|λ|60/ba∇dblF/ba∇dbl2\nH.A PLATE-MEMBRANE TRANSMISSION PROBLEM 21\nConsidering the powers of |λ|, this is the worst term appearing in the estimate\nof/parenleftbig\n/ba∇dblu1/ba∇dblH7/2(Ω1)+|λ|/ba∇dblu1/ba∇dblH3/2(Ω1)+/ba∇dblF/ba∇dblH/parenrightbig\n/ba∇dblu2/ba∇dblH3/2(Ω2). This is due to the fact\nthat in any other term appearing, the power of /ba∇dblU/ba∇dblHis less than23\n12which\nresults in lower powers of |λ|after applying Young’s inequality. Now, using\n|λ|>λ0, we can conclude\n/integraldisplay\nI/vextendsingle/vextendsingle∂νu2(q∇u2)/vextendsingle/vextendsingledS≤C/parenleftbig\n/ba∇dblu1/ba∇dblH7/2(Ω1)+|λ|/ba∇dblu1/ba∇dblH3/2(Ω1)+/ba∇dblF/ba∇dblH/parenrightbig\n/ba∇dblu2/ba∇dblH3/2(Ω2)\n≤ε/ba∇dblU/ba∇dbl2\nH+C(ε)|λ|60/ba∇dblF/ba∇dbl2\nH,\nwhereε>0 is arbitrary and C(ε)>0 is a constant only depending on ε.\n/square\nWe are now able to finish the proof of Theorem 5.2.\nProof of Theorem 5.2.By (5.8), Lemma 5.3and Lemma 5.4together with\nYoung’s inequality applied to the term |λ|/ba∇dblU/ba∇dblH/ba∇dblF/ba∇dblH,we get\n/ba∇dblU/ba∇dbl2\nH≤ε/ba∇dblU/ba∇dbl2\nH+C(ε)|λ|60/ba∇dblF/ba∇dbl2\nH\nfor anyε>0 and a constant C(ε)>0 only depending on ε.This shows\n/ba∇dblU/ba∇dblH≤C|λ|30/ba∇dblF/ba∇dblH.\nTakingF= 0,this estimate also shows that iR∩σp(A) =∅.Since A−1is\ncompact, the spectrum σ(A) ofAcoincides with the point spectrum σp(A)\nofAand we may conclude that iR⊂ρ(A).Now, the assertion follows from\nTheorem 5.1. /square\nReferences\n[1] M.Agranovich,R.Denk,andM. Faierman.Weaklysmoothno nselfadjoint spectral elliptic\nboundary problems. In Spectral theory, microlocal analysis, singular manifolds , volume 14\nofMath. Top. , pages 138–199. Akademie Verlag, Berlin, 1997.\n[2] M. S. Agranovich and M. I. Vishik. Elliptic problems with a parameter and parabolic\nproblems of general type. Russian Math. Surveys , 19(3):53–157, 1964.\n[3] M. Alves, J. Mu˜ noz Rivera, M. Sep´ ulveda, O. Vera Villag r´ an, and M. a. Zegarra Garay.\nThe asymptotic behavior of the linear transmission problem in viscoelasticity. Math.\nNachr., 287(5-6):483–497, 2014.\n[4] M. Alves, J. Mu˜ noz Rivera, M. Sep´ ulveda, and O. V. Villa gr´ an. The lack of exponential\nstability in certain transmission problems with localized Kelvin-Voigt dissipation. SIAM\nJ. Appl. Math. , 74(2):345–365, 2014.\n[5] K. s. Ammari and S. Nicaise. Stabilization of a transmiss ion wave/plate equation. J.\nDifferential Equations , 249(3):707–727, 2010.\n[6] J. A. Arango, L. P. Lebedev, and I. I. Vorovich. Some bound ary value problems and\nmodels for coupled elastic bodies. Quart. Appl. Math. , 56(1):157–172, 1998.\n[7] G. Avalos, I. Lasiecka, and R. Triggiani. Heat-wave inte raction in 2–3 dimensions: optimal\nrational decay rate. J. Math. Anal. Appl. , 437(2):782–815, 2016.\n[8] C. Batty,L.Paunonen,andD.Seifert. Optimal energydec ayin aone-dimensional coupled\nwave-heat system. J. Evol. Equ. , 16(3):649–664, 2016.\n[9] A. Borichev and Y. Tomilov. Optimal polynomial decay of f unctions and operator semi-\ngroups.Math. Ann. , 347(2):455–478, 2010.\n[10] I. Chueshov and I. Lasiecka. Von Karman evolution equations . Springer Monographs in\nMathematics. Springer, New York, 2010. Well-posedness and long-time dynamics.22 B. BARRAZA MART ´INEZ ET AL.\n[11] R. Denk and F. Kammerlander. Exponential stability for a coupled system of damped-\nundamped plate equations. IMA J. Appl. Math. , 83:302–322, 2018.\n[12] R. Denk and R. Schnaubelt. A structurally damped plate e quation with Dirichlet-\nNeumann boundary conditions. J. Differential Equations , 259(4):1323–1353, 2015.\n[13] T. Duyckaerts. Optimal decay rates of the energy of a hyp erbolic-parabolic system cou-\npled by an interface. Asymptot. Anal. , 51(1):17–45, 2007.\n[14] H. D. Fern´ andez Sare and J. E. Mu˜ noz Rivera. Analytici ty of transmission problem to\nthermoelastic plates. Quart. Appl. Math. , 69(1):1–13, 2011.\n[15] R. K. Gazizullin and V. N. Paimushin. The transmission o f an acoustic wave through a\nrectangular plate between barriers. J. Appl. Math. Mech. , 80(5):421–432, 2016.\n[16] B. Gong, F. Yang, and X. Zhao. Stabilization of the trans mission wave/plate equation\nwith variable coefficients. J. Math. Anal. Appl. , 455(2):947–962, 2017.\n[17] F. Hassine. Asymptotic behavior of the transmission Eu ler-Bernoulli plate and wave\nequation with a localized Kelvin-Voigt damping. Discrete Contin. Dyn. Syst. Ser. B ,\n21(6):1757–1774, 2016.\n[18] F. Hassine. Energy decay estimates of elastic transmis sion wave/beam systems with a\nlocal Kelvin-Voigt damping. Internat. J. Control , 89(10):1933–1950, 2016.\n[19] J. Hern´ andez Monz´ on. A system of semilinear evolutio n equations with homogeneous\nboundary conditions for thin plates coupled with membranes . InProceedings of the 2003\nColloquium on Differential Equations and Applications , volume 13 of Electron. J. Differ.\nEqu. Conf. , pages 35–47. Southwest Texas State Univ., San Marcos, TX, 2 005.\n[20] I. Lasiecka and R. Triggiani. Control theory for partial differential equations: continuo us\nand approximation theories. I, Abstract parabolic systems , volume 74 of Encyclopedia of\nMathematics and its Applications . Cambridge University Press, Cambridge, 2000.\n[21] H. Liu and N. Su. Existence and uniform decay of solution s for a class of generalized\nplate-membrane-like systems. Int. J. Math. Math. Sci. , pages Art. ID 83931, 24, 2006.\n[22] E. Mitidieri. A Rellich type identity and applications .Comm. Partial Differential Equa-\ntions, 18(1-2):125–151, 1993.\n[23] J. E. Mu˜ noz Rivera and H. Portillo Oquendo. A transmiss ion problem for thermoelastic\nplates.Quart. Appl. Math. , 62(2):273–293, 2004.\n[24] J. E. Mu˜ noz Rivera and R. Racke. Transmission problems in (thermo)viscoelasticity with\nKelvin-Voigt damping: nonexponential, strong, and polyno mial stability. SIAM J. Math.\nAnal., 49(5):3741–3765, 2017.\n[25] H. Triebel. Interpolation theory, function spaces, differential operat ors, volume 18 of\nNorth-Holland Mathematical Library . North-Holland Publishing Co., Amsterdam-New\nYork, 1978.\n[26] W. Zhang and Z. Zhang. Stabilization of transmission co upled wave and Euler-Bernoulli\nequations on Riemannian manifolds by nonlinear feedbacks. J. Math. Anal. Appl. ,\n422(2):1504–1526, 2015.A PLATE-MEMBRANE TRANSMISSION PROBLEM 23\nB. Barraza Mart ´ınez, Universidad del Norte, Departamento de Matem ´aticas y\nEstad´ıstica, Barranquilla, Colombia\nE-mail address :bbarraza@uninorte.edu.co\nR. Denk, Universit ¨at Konstanz, Fachbereich f ¨ur Mathematik und Statistik,\nKonstanz, Germany\nE-mail address :robert.denk@uni-konstanz.de\nJ. Hern ´andez Monz ´on, Universidad del Norte, Departamento de Matem ´aticas\ny Estad ´ıstica, Barranquilla, Colombia\nE-mail address :jahernan@uninorte.edu.co\nF. Kammerlander, Universit ¨at Konstanz, Fachbereich f ¨ur Mathematik und\nStatistik, Konstanz, Germany\nE-mail address :felix.kammerlander@uni-konstanz.de\nM. Nendel, Universit ¨at Bielefeld, Institut f ¨ur Mathematische Wirtschafts-\nforschung, Bielefeld, Germany\nE-mail address :max.nendel@uni-bielefeld.de" }, { "title": "2402.08519v2.Hyperballistic_transport_in_dense_ionized_matter_under_external_AC_electric_fields.pdf", "content": "Hyperballistic transport in dense ionized matter under external AC electric fields\nDaniele Gamba1, Bingyu Cui2,3, Alessio Zaccone1,4∗\n1Department of Physics “A. Pontremoli”, University of Milan, via Celoria 16, 20133 Milan, Italy\n2School of Science and Engineering, The Chinese University of Hong Kong, Shenzhen, Guangdong, 518172, P. R. China\n3Department of Chemistry, University of Pennsylvania, Philadelphia, Pennsylvania 19104, USA and\n4Institute for Theoretical Physics, University of G¨ ottingen,\nFriedrich-Hund-Platz 1, 37077 G¨ ottingen, Germany.\n(Dated: February 21, 2024)\nThe Langevin equation is ubiquitously employed to numerically simulate plasmas and dusty plas-\nmas. However, the usual assumption of white noise becomes untenable when the system is subject\nto an external AC electric field. This is because the charged particles in the plasma, which provide\nthe thermal bath for the particle transport, become themselves responsive to the AC field and the\nthermal noise is field-dependent and non-Markovian. We theoretically study the particle diffusiv-\nity in a Langevin transport model for a tagged charged particle immersed in a dense plasma of\ncharged particles that act as the thermal bath, under an external AC electric field, by properly\naccounting for the effects of the AC field on the thermal bath statistics. We analytically derive the\ntime-dependent generalized diffusivity D(t) for different initial conditions. The generalized diffusiv-\nity exhibits damped oscillatory-like behaviour with initial very large peaks, where the generalized\ndiffusion coefficient is enhanced by orders of magnitude with respect to the infinite-time steady-state\nvalue. The latter coincides with the Stokes-Einstein diffusivity in the absence of external field. For\ninitial conditions where the external field is already on at t= 0 and the system is thermalized under\nDC conditions for t≤0, the short-time behaviour is hyperballistic, MSD∼t4(where MSD is the\nmean-squared displacement), leading to giant enhancement of the particle transport. Finally, the\ntheory elucidates the role of medium polarization on the local Lorentz field, and allows for estimates\nof the effective electric charge due to polarization by the surrounding charges.\nI. INTRODUCTION\nThe Langevin equation has become one of the most\nwidely employed mathematical tools to numerically sim-\nulate plasmas and dusty plasmas, let alone colloidal sys-\ntems and the molecular dynamics of liquids and solids\n[1]. The Langevin equation in its standard form reads\nas:\nmdv\ndt=−γ0v− ∇V(x) +Fp(t) (1)\nwhere vandmare the particle’s velocity and mass. re-\nspectively, γ0is a viscous friction coefficient, Vis a con-\nservative field, and Fpis a stochastic force. The latter\nis always assumed to obey the statistics of a white noise,\nas given by the fluctuation-dissipation theorem (FDT):\n⟨Fi,p(t)Fj,p(t′)⟩= 2γ0kBTδi,jδ(t−t′) (2)\nwhere i, jdenote Cartesian components and δ(t−t′) is\na Dirac delta function.\nMorfill and co-workers [2] established a computational\nscheme to simulate plasmas and dusty plasmas via the\nLangevin equation, with a delta-correlated noise as in\nthe above Eq. (2). All subsequent studies in the area\nof plasmas and dusty plasmas have adopted the same\nscheme [3–5], including dusty plasmas in oscillating (AC)\nexternal fields [6–9], lane and band formation in dusty\n∗alessio.zaccone@unimi.it; daniele.gamba1@studenti.unimi.itplasmas under AC fields [10–12], and machine learning\nfor searching phase transitions in dusty plasmas [13].\nIn Ref. [14], it was mathematically proved that, for\nsystems of charged particles in an external AC electric\nfield, the (generalized) Langevin equation is intrinsically\nnon-Markovian. This is reflected in the fact that the\nfluctuation-dissipation theorem (i.e. the statistics of the\nthermal noise) is given (in 1D) by\n⟨Fp(t)Fp(t′)⟩=mkBTK(t−t′) +Q2E(t)E(t′) (3)\nwhere E(t) represents the external AC electric field, e\nis the electron charge and Qis a renormalized effective\ncharge related to FPvia⟨FP⟩=QE(t).\nThis modified FDT has an extra term ∼E(t)E(t′) in\naddition to the usual term K(t−t′). While the latter can,\nunder certain conditions, be reduced to a Dirac delta as\ndiscussed e.g. in [15], the term ∼E(t)E(t′) obviously\ncannot be reduced to a Dirac delta under any conditions,\nbecause E(t)∼sin(Ω t) is a sinusoidal function. This\neffect, therefore, arises only in presence of an external\ntime-dependent field, whereas, if the external field is DC,\nthe effect is not there and Eq. (2) remains valid.\nTherefore, the correct FDT for a plasma (and any\nother system of charged particles such as e.g. liquid elec-\ntrolytes [16, 17]) is always non-Markovian and is given\nby Eq. (3). This fact is new and has been systematically\noverlooked in the literature on Langevin simulations of\nplasmas, where the standard Markovian FDT Eq. (2) is\nused even when the plasma is under an AC electric or\nelectromagnetic field.\nWe present a mathematical model of particle transport\nin dense ionized matter under an external AC electricarXiv:2402.08519v2 [physics.plasm-ph] 20 Feb 20242\nfield that is relevant to dense plasmas and dusty plasmas,\nhigh energy density matter, solute colloidal systems, liq-\nuid metals and ferrofluids, and liquid electrolytes.\nIn this model, the generalized Langevin equation of\nmotion for the charged particle is derived from a mi-\ncroscopic Caldeira-Leggett particle-bath Hamiltonian,\nwhere, crucially, the bath is formed of charged harmonic\noscillators which also respond to the external AC field.\nThis is a trick to effectively decompose the thermal kicks\non the tagged particle due to innumerable collisions by\ndecomposing the corresponding thermal bath into the vi-\nbrational eigenmodes of the system.\nIn this way, upon setting suitable initial conditions, we\nanalytically derive the time-dependent particle diffusiv-\nityD(t), which we study as a function of the AC elec-\ntric field parameters, in particular the field amplitude E0\nand frequency Ω. The magnitude of the effective polar-\nization charge is estimated, and a giant enhancement of\nthe polarization field is found when the field frequency\napproaches the characteristic vibration frequency of the\ncharged particles forming the bath. Predictions concern-\ning the behaviour of the generalized diffusion coefficient\nD(t) are obtained including the prediction of superdiffu-\nsive transport when the external field frequency becomes\nvery small. In particular, for the case of initial conditions\nwhere the AC field is switched on at t= 0, this short-\ntime superdiffusive regime is simply ballistic. Instead,\nif the system has thermalized under DC conditions for\ntimes t≤0 and then the field becomes AC for t >0,\na hyperballistic superdiffusive regime is predicted. It is\nalso established that D(t→ ∞ ) coincides with the par-\nticle diffusivity in the absence of the electric field under\nall initial conditions, which is a non-trivial result.\nII. MODEL OF PARTICLE-BATH SYSTEM\nWe consider the same model as in Ref. [14], in which an\nexternal, time-dependent electric field, E(t), is applied to\na dense medium. The motion of the tagged particle of\nmass mis embedded in a dense medium (thermal bath)\nof other particles. In order to effectively decompose the\neffect of the thermal bath onto the normal modes of the\nsystem, we use the Caldeira-Leggett trick of representing\nthe thermal environment as fictitious harmonic oscilla-\ntors with natural frequencies corresponding to the eigen-\nmodes of the medium. In our theory, these fictitious har-\nmonic oscillators are electrically charged, and thus are\nsubject to the effect of the external electric field via the\nLorentz force. The latter acts on all the charged parti-\ncles which, in reality, form the medium and the thermal\nbath. Hence, the thermal bath responds to the external\nfield and its statistics is modified by the external field,\nas it should be, in the physical reality. We shall neglect\nmagnetic forces throughout.\nThe Hamiltonian H(in 1D) of the total system consistsof the Hamiltonian of the tagged particle\nHp=p2\n2m−q x E (t), (4)\nwhere pis the momentum, xthe position and qthe\nparticle’s charge, plus the Hamiltonian Hbof the bath.\nThe latter is effectively modeled by Nindependent\nelectrically-charged harmonic oscillators of coordinates\n{xi, pi}, with mass mi, electric charge qi, and oscillating\nat some eigenfrequency ωi[15, 18]. Since each oscillator\nicarries a charge qi, it thus feels the electric force due\nto the external AC field, and the bath as a whole is re-\nsponsive to the AC field. The Hamiltonian of the bath\noscillators can be written as [14]\nHb=1\n2NX\ni=1 \np2\ni\nmi+miω2\ni\u0012\nxi−ν2\ni\nω2\nix\u00132\n−2qixiE(t)!\n=H0\nb−1\n2E(t)NX\ni=1qixi, (5)\nwith\nH0\nb=1\n2NX\ni=1 \np2\ni\nmi+miω2\ni\u0012\nxi−ν2\ni\nω2\nix\u00132!\n. (6)\nNote that we have completed the square to put the ex-\npression in a more convenient form for future calcula-\ntions, and we have also supposed that displacements x\nandxiare small so that the i-th oscillator couples to\nthe Brownian particle linearly. The coupling strength is\ndenoted as ν2\ni.\nThe external electric field oscillates with frequency Ω\nand amplitude E0. We consider two different initial con-\nditions (ICs): field-off initial conditions , given by\nE(t) =(\n0, t < 0\nE0sin(Ω t), t≥0,, (7)\nandfield-on initial conditions , given by\nE(t) =(\nE0, t < 0\nE0cos(Ω t), t≥0, (8)\nThe field behavior in the two cases is shown in Fig. 1.\nFollowing the canonical formalism, the derivation of\nRef. [14] gives the equation of motion of the tagged par-\nticle as\ndp\ndt=−Zt\n0K(t′)p(t−t′)dt′−∇V(x)+qE(t)+Fp(t),(9)\nwhere\nK(t) =NX\ni=1miν4\ni\nmω2\nicos(ωit) (10)3\n0 1 2 3 4-1.5-1.0-0.50.00.51.01.5\ntE(t)/E0\n(a)\n0 1 2 3 4-1.5-1.0-0.50.00.51.01.5\ntE(t)/E0\n(b)\nFigure 1: Different initial conditions for the\nparticle-bath Hamiltonian (5) under the action of an\nexternal AC electric field E(t). (a): Field-off initial\nconditions, where the field is switched on at t= 0\n(Eq. (7)). (b): Field-on initial conditions, where the\nfield is already on at t= 0 (Eq. (8)).\nis the memory kernel describing the time accumulation\neffect on the tagged particle due to the coupling to the\nbath, and\nFp(t) =NX\ni=0\"\nmiν2\ni\u0012\nxi(0)−ν2\ni\nω2\nix(0)\u0013\ncos(ωit) +\n+ν2\ni\nωipi(0) sin( ωit) +qiE′\ni(ωi, t)#\n(11)\nwhere\nE′\ni(ωi, t) =ν2\ni\nωiZt\n0E(t′) sin(ωi(t−t′))dt′. (12)\nThe “noise” (11) depends on the distribution of the\nbath oscillators and on the electric field, and is a well-\ndefined function of time. It describes the effect of a huge\nnumber of irregular kicks on the tagged particle due to\nthe collisions with the particles of the bath. The forcePN\ni=1qiE′\ni(t) acting on the Brownian particle is due to\nthe internal polarization of the medium under the exter-\nnal AC field E(t). The function E′\n0(ω0, t) (Eq. (12)) isplotted in Fig. 2 as a function of time, and in Fig. 3 as\na funcion of frequency.\nIn most cases, the medium contains a large number\nof oscillating modes, so we can regard the collection of\noscillating eigenfrequencies as continuous. Therefore, the\nsum might be replaced with the integralR\ng(ω)dω, where\ng(ω) =P\niδ(ω−ωi) is the density of vibrational states.\nThus, in the continuum limit, the friction kernel (10) may\nbe written as\nK(t) =m0\nmZωD\n0ν(ω)4\nω2cos(ωt)g(ω)dω, (13)\nwhere we have set mi=m0for all bath oscillators i.\nFollowing Zwanzig [15], we assume a Debye form for the\ndensity of states of the bath:\ng(ω) =(\ng0ω2,0< ω < ω D,\n0, otherwise,(14)\nwhere g0is a constant. This choice is motivated by the\nfact that the main low-energy excitations in the plasma\nare sound waves with some dispersion relation ω(k) =ck,\nwhere cis the sound speed. The corresponding density\nof states of these low-energy eigenmodes ωis thus given\nby the quadratic Debye form [19].\nEq. (13) becomes\nK(t) =g0m0\nmZωD\n0ν(ω)4cos(ωt)dω (15)\nWe also introduce, for later convenience, the renormal-\nized effective charge\nQ=NX\ni=1qiν2\ni\nω2\ni=g0q0ZωD\n0ν(ω)2dω≡g0q0f(0),(16)\nwhere we have set qi=q0for all oscillators i, and\nf(t) =ZωD\n0ν(ω)2cos(ωt)dω. (17)\nThe average noise Fpand its autocorrelation function can\nbe computed explicitly in the two situations described\nbelow.\nA. Pseudo-resonance of the polarization function\nConsider the situation described by Eq. (7), where the\nexternal electric field Eis switched on at time t= 0.\nHowever, for very short times, the bath doesn’t have time\nto instantaneously respond to the change, and is still in\nthe same equilibrium distribution ∼e−H0\nb/kBTas at t= 0\nand earlier times. Therefore, the average noise becomes\n⟨Fp(t)⟩=NX\ni=0qiν2\niE0\nωiωisin(Ω t)−Ω sin( ωit)\nω2\ni−Ω2.(18)4\n0 5 10 15 20 25 30-2-1012\ntE'0(t)\n(a)\n0 5 10 15 20 25 30-2-1012\ntE'0(t)\n(b)\nFigure 2: Panel (a): field-off ICs. Panel (b): field-on\nICs. Plot of the effective (Lorentz) field as a function of\ntime. Blue, yellow and green lines correspond to values\nΩ/ω0= 0.8,0.99,1.13: the oscillations are giantly\nenhanced at the resonant frequency Ω ≃ω0, and their\namplitude increases with time (Eq. (12)).\nFor the large values of Ω, that are used for instance in\nTHz spectroscopy, it can become comparable with the\nsmallest of the ωivalues, say ω0, then\n⟨Fp(t)⟩ ∼q0E′\n0(ω0, t) =q0ν2\n0E0\n2ω0\u0014sin(ω0t)\nω0−tcos(ω0t)\u0015\n(19)\nand the tagged particle is subject to an oscillating force\nwhose amplitude grows linearly in time.\nThe trend of E′\n0(ω0, t) (Eq. (12)) is shown in Figs. 2, 3,\nfor both the field-off and the field-on ICs.\nNow we consider the situation described by Eq. (8) in\nwhich at t= 0, the electric field has already been on for a\ntime sufficiently long for the bath to equilibrate under the\naction of the external field. The usual relaxation time for\natomic/molecular systems is of order 10−13seconds, so\npossibly much shorter then the time of observation. The\nstate of the medium is therefore thermalized according\n0 2 4 6 8 10-2-1012\nω0E'0(ω0)(a)\n0 2 4 6 8 10-2-1012\nω0E'0(ω0)\n(b)\nFigure 3: Panel (a): field-off ICs. Panel (b): field-on\nICs. Plot of the effective (Lorentz) field as a function of\nthe resonant frequency ω0, fort= 7. Blue, yellow and\ngreen lines correspond to values Ω /ω0= 0.8,0.99,1.13.\nOne can see peaks where the value of Ω is very close to\nω0(Eq. (12)).\nto the distribution ∼e−Hb(t=0)/kBT, with which we find\n⟨Fp(t)⟩=NX\ni=1qiν2\niE0(\ncos(ωit)\nω2\ni+cos(Ω t)−cos(ωit)\nω2\ni−Ω2)\n.\n(20)\nAgain, if Ω is comparable with some resonant frequency\nω0, then\n⟨Fp(t)⟩ ∼q0E′\n0(ω, t) =q0ν2\n0E0\n2sin(ω0t)\nω0t, (21)\nand we have an oscillating force whose amplitude in-\ncreases linearly in time.\nAs we see, Eqs. (18, 20) exhibit a pole in coincidence\nof Ω≃ω0. As explained in [20], this pole represents a\nresonance peak, when the experimental frequency Ω is\nprobing exactly a characteristic frequency ω0of the ma-\nterial medium, and this gives rise to spectra of emission\nand absorption. One usually solves the problem of inte-\ngrating the singularity, by displacing the pole by a little\namount in the complex plane: this physically corresponds5\nto recognizing that each mode ωialso suffers some dis-\nsipation. In that case however, the resonance is made\npossible when the probing oscillating field has been on\nfor a long amount of time. In our case, the divergence\nis avoided because we have been probing the medium on\nits resonance frequency only for a time t.\nThese frequency sums can be evaluated analytically.\nOne should note that the sum contains diverging quan-\ntities, which, however, when summed together give rise\nto a finite quantity, due to a negative interference effect.\nThe result is a quantity that is oscillating in time, and\nthe giant amplitude enhancement that grows linear in t\nhas no effect on the quantity integrated in time, because\nall the giant oscillating terms average to quantities of or-\nder 1. Interesting effects might appear if we consider the\nprobing frequency Ω comparable with the highest eigen-\nfrequency ωDof the medium, but that goes beyond the\nscope of the present paper and may be the subject for\nfuture work.\nHaving made the above considerations, we can keep\nworking in the limit Ω ≪ωifor every mode i, and in this\nlimit the sum in Eq. (18) can be evaluated and gives\n⟨Fp(t)⟩=QE0sin(Ω t) (22)\nfor field-off ICs; similarly Eq. (20) gives\n⟨Fp(t)⟩=QE0cos(Ω t) (23)\nfor field-on ICs, where\nQ=g0q0ν2\n0ωD (24)\n(see App. A for the detailed derivation of\nEqs. (22, 23, 24)). So, as we can see, the quantity\nQhas the role of an effective or renormalized charge ac-\nquired by the tagged particle, due to the polarization of\nthe surrounding medium. The physical picture is that, in\ndense or supercooled liquid ionized matter, each particle\nis trapped in a cage by the surrounding particles. If the\noscillatory AC field drives the particle in one direction\nand opposite charges in the opposite direction, then we\nwill have a resonance effect with the normal mode of the\nparticle in the cage. The effect adds up to the electric\nfield on the particle, because in general the positive ions\ndraw the particle in the opposite direction of the field,\nand if q=−Qthey would completely screen the effect\nof the external AC field. See Fig. 4 for a schematic\nvisualization of this effect. It is also important to note\nthat results (22, 23) show insensitivity of the system\nto the choice of initial conditions, i.e. whether field-off\nor field-on, because memory of the initial conditions is\nlost in a time proportional to ω−1\nD, which is a very short\ntime, e.g. if ωDcoincides with the plasma frequency.\nIndeed, in (20), in the limit Ω ≪ωD, the first term,\nintegrated on the frequency, is proportional tosin(ωDt)\nt,\nbecomes negligible in comparison with the second term,\nproportional to ωD, because the time of observation is\nt≫ω−1\nD. Hence, the extra term due to field-on initial\nconditions becomes negligible after very short times, and\nthe dependence on the initial conditions is lost.\n+\n+++\n+\n------------------\n-+\n+\n+\n+\n+\n+\n+\n+\n+\n+V(x)\n-+Figure 4: Schematic depiction of the physical system\n(dense ionized matter) considered in the theoretical\nmodel. The tagged particle (green) moves in a thermal\nbath (dense medium) of other charges schematically\nrepresented as charged harmonic oscillators in the\ntheoretical model. The frequencies of these fictitious\noscillators schematically represent the eigenfrequencies\nof the sound modes of the system. These oscillators are\ncoupled to the tagged particle by different coupling\nstrengths, represented as red lines. The effective charge\nof the tagged particle, Q(represented as a shaded area),\naccounts for the local charge distribution and polarizing\nLorentz field around the tagged particle. The system is\nunder an AC electric field (depicted in figure at a\ncertain instant of time) exerted by the electrodes\n(represented by two opposite charged plates in the\ngraph) and might be subjected to an overall static field\nV(x), which is not explicitly taken into account in our\nderivations but the presence of which would not\nqualitatively change our results.\nB. Autocorrelation functions\nThe influence of the external field is also exhibited in\nthe autocorrelation function of the random force,\n⟨Fp(t)Fp(t′)⟩=mkBTK(t−t′) + (QE0)2sin(Ω t) sin(Ω t′),\n(25)\nas has been proposed in [14], for field-off initial condi-\ntions, and\n⟨Fp(t)Fp(t′)⟩=mkBTK(t−t′) + (QE0)2cos(Ω t) cos(Ω t′)\n(26)\nfor field-on initial conditions (here derived for the first\ntime; the details of the calculation are presented in Ap-\npendix A). In contrast to the previous case of the field-off\nICs, the effective Lorentz field arising inside the medium\nis phase-shifted according to the external electric field.6\n0 5 10 15 2002468\ntD˜(t)\n(a)\n0 5 10 15 2002468\ntD˜(t)\n(b)\nFigure 5: Time evolution of the generalized diffusivity\n˜D=D2mγ0\nkBTfor (a) field-off (32) and (b) field-on (34)\ninitial conditions, obtained with different field\namplitudes E0(Eq. (33)), for Ω = 1. Yellow, green and\norange lines correspond to E0= 1,2,3, whereas the blue\nline is the Stokes-Einstein result obtained for E0= 0.\nWe can also calculate the momentum autocorrelation\nfrom the force correlation in the overdamped Brownian\nlimit, i.e., when γ0≫m(see Eq. (1)). Setting ν(ω) =ν0,\nEq. (15) becomes K(t) =γ0δ(t) with\nγ0=πg0m0\nmν4\n0. (27)\nWe find\n⟨p(t1)p(t2)⟩=1\nγ2\n0⟨F(t1)F(t2)⟩, (28)\nwith\nF(t) =Fp(t) +qE(t). (29)\n0 100 200 300 400 500 600020406080\ntD˜(t)(a)\n0 100 200 300 400 500 600020406080\ntD˜(t)\n(b)\nFigure 6: Time evolution of diffusivity ˜D=D2mγ0\nkBTfor\n(a) field-off (32) and (b) field-on (34) initial conditions\nfor different Ω values, comparable to the vibrational\nfrequencies of the bath oscillators. Yellow, green and\norange lines correspond to Ω /ωD= 0.06,0.03,0.01, for\nE0= 1, whereas the blue line is the Einstein result\nobtained for E0= 0. For small Ω the diffusivity is\ngiantly augmented in comparison with the DC\ndiffusivity, which coincides with the Stokes-Einstein\nresult (blue line).\nIII. PARTICLE DIFFUSIVITY UNDER\nEXTERNAL AC FIELD\nUsing the time correlation of the velocity v=p/m, we\nare able to calculate the diffusivity\nD(t) =1\n2t⟨{x(t)−x0}2⟩, (30)\nwhere x0is the position of the Brownian particle at t= 0,\nsince\n⟨{x(t)−x0}2⟩=Zt\n0dt1Zt\n0dt2⟨v(t1)v(t2)⟩. (31)7\nFor the field-off ICs, the diffusivity becomes\nD(t) =kBT\n2mγ0\u0012\n1 +E2\n0(cos(Ω t)−1)2\nΩt\u0013\n(32)\nwith\nE0=(q+Q)E0√mkBTγ0Ω, (33)\nwhile for the field-on ICs, we find\nD(t) =kBT\n2mγ0\u0012\n1 +E2\n0sin2(Ωt)\nΩt\u0013\n(34)\n(see App. B for a detailed derivation).\nInterestingly, we shall note that both the above expres-\nsions for D(t) asymptotically recover the particle diffu-\nsivity in the absence of the external field in the limit\nt→ ∞ :\nlim\nt→∞D(t) =kBT\n2mγ0, (35)\nwhich is the standard Stokes-Einstein result, independent\nof the AC field parameters. Recalling the discussion fol-\nlowing Eq. (21), the generalized diffusivity D(t) doesn’t\nshow a divergent behavior for frequencies near the reso-\nnant frequencies of the system: in fact, in Eqs. (32, 34),\nthe resonant frequencies of the system affect only the pa-\nrameter E0, trough the renormalized charge Q.\nThe non-dimensional diffusivity ˜D(t) =2mγ0\nkBTD(t) as-\nsociated with the motion of the tagged particle under\nthe field-off ICs is shown in Figs. 5 and 6. In Fig. 5, we\ncompare D(t) for different values of the non-dimensional\nfield amplitude Eq. (33). In Fig. 6, we compare D(t) for\ndifferent values of the probing frequency Ω.\nOverall, the oscillatory diffusivity ultimately decays to-\nwards the t→ ∞ Stokes-Einsten value (35) as if there\nwas no field applied. The effect of the AC field under the\nfield-off ICs is more significant at short times, resulting\nin a first peak of D(t) which is larger for larger values of\nE0. Similar behaviour is also reported for the response\nof a colloidal particle to a time-dependent quadratic po-\ntential due to an optical trap being dragged through the\nfluid [21]. In our case, these oscillations are augmented\nby a factor Qdue to the polarization of the surrounding\nmedium (Eq. (33)). If the external frequency Ω is very\nsmall, in the case of field-on ICs we recover for times\nt <Ω−1the limit of ballistic motion under a constant\nfield, that is, the limit for small Ω of (34), which is\nD(t)≈kBT\n2mγ0\u0000\n1 +E2\n0Ωt\u0001\n(36)\nThis asymptotic behavior for small tis visible in the ini-\ntial slope in Figs. 5b and 6b.\nEquation (34) shows that the first peak in D(t) is at\nΩt= 1.16:\nDmax≃D(1.16/Ω) =kBT\n2mγ0\u0000\n1 + 0 .72E2\n0\u0001\n. (37)SinceE2\n0∝1/Ω, when Ω is decreased, the peak is shifted\nforward in time and amplified. In the DC limit of con-\nstant field, i.e. for small Ω, and for finite times, Eq. (36)\nreduces to the Stokes-Einstein value, plus a standard drift\ncontribution, linear in time, which contributes to the\ndressed or generalized diffusivity D(t), while the naked\ndiffusivity is given simply by the Stokes-Einstein value.\nIn the case of field-off ICs, instead, the short-time\nbehaviour is given by taking the limit of small Ω tin\nEq. (32), which gives:\nD(t)≈kBT\n2mγ0\u0000\n1 +E2\n0Ω3t3\u0001\n(38)\nand, in this case, there is an inflection point at t= 0.\nImportantly, in this case, a hyperballistic behaviour is\nobserved at short times, D(t)∼t3, giving a faster and\nlarger growth of the first peak in D(t) compared to the\nprevious case. This corresponds to a mean squared dis-\nplacement (MSD) that grows as ⟨x(t)−x02⟩ ∼t4, i.e.\nwith the fourth power of time.\nThis is a consequence of the AC field ∼sin(Ω t) being\nswitched on at t= 0 following thermalization under DC\nconditions. In the overdamped regime, this gives a par-\nticle displacement cos(Ω t)−1∼t2. Eq. (32) shows that\nthe first peak of D(t) is at Ω t= 2.78:\nDmax≃D(2.78/Ω) =kBT\n2mγ0\u0000\n1 + 1 .35E2\n0\u0001\n. (39)\nWhen Ω is decreased, the peak is shifted forward in time\nand amplified. In the Ω t≪1 limit, Eq. (39) shows that,\nfor finite times, the Stokes-Einstein value (blue line in\nFig. 6a) is recovered. This is the first prediction, to our\nknowledge, of hyperballistic superdiffusive transport in\ndense systems of charged particles under an AC electric\nfield.\nConsistent with results in Ref. [22], while the charge is\naccelerating under the external field, large velocities are\nsmeared out by the damping γ0. Charges moving faster\nwill be dragged by the friction to a larger extent. For\na lump of charge density moving in the field direction,\nthe charges in the front are faster and undergo major\ndamping, while those on the rear do not undergo much\nfrictional resistance and do not lose much momentum.\nAs a result, charges accumulate in the front, increasing\nthe local charge density. The overall effect is that the\npulse gets compressed in the direction of the field, which\nreduces diffusivity parallel to the field. On the other\nhand, the transverse diffusivity might increase. In the\nexample of diffusion-convection of a solute in a liquid,\ncfr. Ref. [23], there is complete uniformity in Taylor dis-\npersion transverse to the applied field, so there is no rise\nin gradients of velocity or concentration. However, in this\npaper, we are only considering the longitudinal case for\nthe 1D Langevin equation along the field direction. The\ntransverse equation might be coupled to what happens\nin the longitudinal one, but this would require a full 3D\ntreatment which may not be analytically tractable within\nthe Langevin equation framework.8\nAs we have seen from the above theory, the correction\nto the (naked) diffusivity coefficient due to the external\nAC electric field is proportional to ωD(this is because\nE0∝Q∝ωD, see Eqs. (33, 24)). Since this is the highest\nfrequency of the bath, it can be taken as the plasma\nfrequency, and hence,\nωD∝1√m(40)\nwhere mis the mass of the electron, or in general, of the\nlightest charged particle in the medium. This implies\nthat, in general, the diffusivity enhancement effect pre-\ndicted by the above theory will be larger the smaller the\nmass of the lightest constituent present in the medium.\nIV. CONCLUSIONS\nWe studied a Caldeira-Leggett particle-bath model\nwhere both the tagged particle and the oscillators form-\ning the bath are electrically charged and respond to an\nexternal AC electric field. Since the bath is responding to\nthe external AC field, the thermal noise is no longer white\nnoise, but is non-Markovian and strongly influenced by\nthe external time-dependent electric field. Based on this\nmodel, we analytically derived results for the autocorre-\nlation functions of force and momentum, and used the\nlatter to evaluate the time-dependent generalized parti-\ncle diffusivity D(t). Analytical close-formed expressions\nforD(t) as a function of the physical parameters of the\nsystem are obtained for two different initial conditions\n(AC field off and AC field on at t= 0, respectively),\nand are given by Eq. (32) and (34). We found that the\ngeneralized diffusivity in the asymptotic t→ ∞ steady\nstate is the same as in the absence of the external field\nand is given by the Stokes-Einstein formula, which is a\nrather remarkable result. Also, the generalized diffusiv-\nity exhibits time-dependent damped fluctuations in the\ntransient regime which, for low AC field frequency Ω,\nproduce a transient giant enhancement of the general-\nized diffusivity with respect to the steady-state value.\nFor low enough frequency of the external AC field, the\nenhancement can be up to few orders of magnitude with\nrespect to the steady-state value. In particular, our the-\nory predicts: (i) ballistic superdiffusion for the field-off\ninitial conditions (the AC field is switched on at t= 0\nfollowing equilibration at zero external field), and (ii) hy-\nperballistic superdiffusion with quartic power of the mean\nsquared displacement, MSD ∼t4, for the case of field-\non initial conditions (the AC field is switch on at t= 0\nfollowing thermalization under DC field).\nThis effect, predicted here for the first time, can be\ninterpreted as the result of a memory accumulation pro-\ncess, due to the non-Markovianity of the noise (Eqs. (25)\nand (26)). This is made evident if we take the DC limit of\nconstant field (Ω →0): in this case, both the naked diffu-\nsivities corresponding to (36) and (38) are reduced to the\nStokes-Einstein value. From a physical point of view, thismemory accumulation means that the stochastic force\nacting on the particle, representing collisions with other\nparticles, retains memory of previous collisions which in\nthe field-on ICs case are continuously enhanced by the\ndrift over past times, and hence grows as a function of\ntime. Also, the enhancement of the transient generalized\ndiffusivity is predicted to be inversely proportional to the\nsquare root of the lightest particle in the medium (e.g.\nthe electron mass in the case of plasmas).\nThe theory also allows one to evaluate the local\n(Lorentz) field acting on the tagged particle due to po-\nlarization in the surrounding medium. Interestingly, a\nresonance-type behaviour of the local effective field is un-\nveiled corresponding to a frequency value of the external\nAC field that matches the value of a pole in one of the\nterms constituting the average random force ⟨Fp(t)⟩act-\ning on the tagged particle. Also interestingly, however,\na destructive interference with other oscillating terms in\n⟨Fp(t)⟩conspires to remove this resonance in the overall\nbehaviour of ⟨Fp(t)⟩which turns out to be given by the\nsame sinusoidal function of the external AC field, with\na renormalized charge due to charge renormalization by\nthe surrounding charged particles.\nSince the diffusivity is a key parameter in plasma\nphysics, e.g. it is a control parameter for the formation of\nplasma blobs in tokamaks [24, 25] and plasma turbulence\n[26], the current predictions may be useful for optimiz-\ning transport phenomena and overall efficiency of fusion\nreactors.\nFrom the fundamental point of view of nonequilibrium\nstatistical mechanics, we have discovered and predicted\na new physical effect which represents a new example\nof anomalous diffusion [27, 28], namely of hyperballis-\ntic superdiffusion [29, 30]. In particular, we showed that\ndynamic accumulation of the stochastic force, which rep-\nresents random collisions enhanced by an external drift,\nleads to hyperballistic transport [31]. Our theoretical\nfindings may connect to previous experimental reports of\nhyperballistic transport in photons propagating through\ndynamic disorder [32, 33]. It may also connect to the\nphenomenon of secondary Fermi acceleration, proposed\nto explain the origin and acceleration of cosmic rays [34].\nThis is the mechanism by which a plasma is heated and\naccelerated by time-dependent stochastic driving fields.\nIn the original setup, this was a model for cosmic rays be-\ning scattered by magnetized interstellar clouds that move\nrandomly and act as magnetic mirrors.\nAll in all, the hyperballistic dynamics under an exter-\nnal AC field discovered here could potentially lead to a\nnew controlled setup for the acceleration of ionized mat-\nter [35]. Future extensions of the current theory will in-\nclude the effect of magnetic fields for magnetized plasmas\nand possibly magneto-hydrodynamic effects. Also, the\npresent theory may represent the starting point of linear\nresponse theories to predict material response properties\nunder external fields [36–38] or within molecular simula-\ntions [1].9\nAppendix A: Average of noise and its\nautocorrelations\nThe average force fluctuations on the particle is the\nsum of all the incoherent contributions on some finite\ntime span T,\n⟨Fp(t)⟩=1\nTZt+T\ntFp(t′)dt′. (A1)\nSimilarly, the autocorrelation of Fv, i.e. the second mo-\nment, is expressed as\n⟨Fp(t)Fp(t+τ)⟩=1\nTZt+T\ntFp(t′)Fp(t′+τ)dt′.(A2)\nIf the bath is in thermal equilibrium at t= 0, since\nthe equilibrium system is ergodic, the average, Eq. (A1),\nis equivalent to a Boltzmann average on every possible\nbath configuration {x1, .., x N, p1, .., p N}:\n⟨Fp(t)⟩H0\nb=NY\ni=1Z+∞\n−∞dpie−p2\ni\n2mikBT×\n×Z+∞\n−∞dxie−miω2\ni\n2kBT\u0012\nxi−ν2\ni\nω2\nix\u00132\nFv(t).(A3)\nCompleting the squares in the Gaussian integrals, we ob-tain\n⟨pi⟩= 0, (A4a)\n⟨p2\ni⟩=mikBT, (A4b)*\nxi−ν2\ni\nω2\nix+\n= 0, (A4c)\n*\u0012\nxi−ν2\ni\nω2\nix\u00132+\n=kBT\nmiω2\ni. (A4d)\nWhen the bath oscillators are subject to the action of\nthe external field, they respond and relax fast (on times\nγ−1\n0≃10−13s, much shorter than the time-scale Ω−1on\nwhich the transient is observed) relative to the dynamics\nof the tagged particle, so the oscillatory field can be ap-\nproximated with its value in a neighborhood of t= 0 [15]:\nHb(t= 0)≈\n1\n2NX\ni=1 \np2\ni\nmi+miω2\ni\u0012\nxi−ν2\ni\nω2\nix−qi\nmiω2\niE0\u00132!\n(A5)\nplus a constant term that depends on x. Since the latter\nremains inert in the computation of bath averages, it is\nirrelevant for our results. Then, the average force is\n⟨Fp(t)⟩=NY\ni=1Z+∞\n−∞dxie−miω2\ni\n2kBT\u0012\nxi−ν2\ni\nω2\nix−qi\nmiω2\niE0\u00132\nFv(t).\n(A6)\nFurther, we have\n*\nxi−ν2\ni\nω2\nix−qi\nmiω2\niE0+\n= 0 (A7)\nand\n*\u0012\nxi−ν2\ni\nω2\nix−qi\nmiω2\niE0\u00132+\n=kBT\nmiω2\ni. (A8)\nTherefore, referring to field-on (7) and field-off (8) ICs,\nwe find the effective field (12) is equal to\nE′\ni(ωi, t) = ν2\niE0(1\nωiωisin(Ω t)−Ω sin( ωit)\nω2\ni−Ω2 ,off,\ncos(Ω t)−cos(ωit)\nω2\ni−Ω2 , on.(A9)\nthen for the average force we have\n⟨Fp(t)⟩=NY\ni=1Z+∞\n−∞dxiFp(t)\n\nexp\"\n−miω2\ni\n2kBT \nxi−ν2\ni\nω2\nix!2#\n, off,\nexp\"\n−miω2\ni\n2kBT \nxi−ν2\ni\nω2\nix−qiE0\nmiω2\ni!2#\n,on.(A10)10\nUsing formula (11) for the noise,\n⟨Fp(t)⟩=NX\ni=1(\nqiE′\ni(t), off,\nmiν2\niqiE0\nmiω2\nicos(ωit) +qiE′\ni(t),on,\n=NX\ni=0qiν2\niE0(1\nωiωisin(Ω t)−Ω sin( ωit)\nω2\ni−Ω2 , off,\ncos(ωit)\nω2\ni+cos(Ω t)−cos(ωit)\nω2\ni−Ω2 ,on,\n=q0g0ν2\n0E0ZωD\n0dω(\nωωsin(Ω t)−Ω sin( ωt)\nω2−Ω2 , off,\ncos(ωt) +ω2cos(Ω t)−cos(ωt)\nω2−Ω2,on,(A11)\nand in the small Ω limit,\n⟨Fp(t)⟩=(\nQE0sin(Ω t),off,\nQE0cos(Ω t),on,(A12)\nwhich gives Eqs. (22, 23) of the main text. Moreover,\nsetting ν(ω) =ν0in Eq. (16), one finds\nQ=g0q0ν2\n0ωD (A13)\nwhich is Eq. (24) of the main text.\nThe autocorrelation function can be similarly evalu-\nated, leading to Eqs. (25, 26) of the main text, which for\nE0= 0 reduce to Kubo’s result [39]:\n⟨Fp(t)Fp(t′)⟩=mkBTK(t−t′). (A14)\nAppendix B: Diffusivity\nThe mean square displacement of the Brownian parti-\ncle is\n⟨{x(t)−x0}2⟩=Zt\n0dt1Zt\n0dt2⟨v(t1)v(t2)⟩. (B1)\nUnder field-off ICs, with a Markovian kernel, we get\n⟨{x(t)−x0}2⟩=mkBT\nγ0+\u0012(q+Q)E0\nΩγ0\u00132\n(cos(Ω t)−1)2.\n(B2)The second term is a new contribution due to the po-\nlarization of the bath, which may be observable even for\nlarge times tif the field’s intensity E2\n0is large enough.\nThe diffusivity, D(t) =⟨{x(t)−x0}2⟩/(2t), is\nD(t) =mkBT\nγ0\u0012\n1 +E2\n0(cos(Ω t)−1)2\nΩt\u0013\n(B3)\nwhere we have introduced the dimensionless quantity\nE0=(q+Q)E0√γ0mΩkBT. (B4)\nEqs. (B3, B4) are Eqs. (32, 33) of the main text.\nUnder field-on ICs, with the Markovian kernel, we get\n⟨{x(t)−x0}2⟩=mkBT\nγ0t+\u0012(q+Q)E0\nγ0\u00132\nsin2(Ωt).\n(B5)\nThe diffusivity becomes\nD(t) =mkBT\nγ0\u0012\n1 +E2\n0sin2(Ωt)\nΩt\u0013\n, (B6)\ncorresponding to Eq. (34) of the main text.\n[1] M. Ceriotti, G. Bussi, and M. Parrinello, Phys. Rev.\nLett. 102, 020601 (2009).\n[2] A. V. Ivlev, S. K. Zhdanov, B. A. Klumov,\nand G. E. Morfill, Physics of Plasmas 12(2005),\n10.1063/1.2041327.\n[3] P. Mabey, S. Richardson, T. G. White, L. B. Fletcher,\nS. H. Glenzer, N. J. Hartley, J. Vorberger, D. O. Gericke,\nand G. Gregori, Nature Communications 8, 14125 (2017).\n[4] F. R. Graziani, V. S. Batista, L. X. Benedict, J. I. Castor,\nH. Chen, S. N. Chen, C. A. Fichtl, J. N. Glosli, P. E.\nGrabowski, A. T. Graf, S. P. Hau-Riege, A. U. Hazi, S. A.Khairallah, L. Krauss, A. B. Langdon, R. A. London,\nA. Markmann, M. S. Murillo, D. F. Richards, H. A. Scott,\nR. Shepherd, L. G. Stanton, F. H. Streitz, M. P. Surh,\nJ. C. Weisheit, and H. D. Whitley, High Energy Density\nPhysics 8, 105 (2012).\n[5] A. Melzer, Lecture Notes in Physics (2019), 10.1007/978-\n3-030-20260-6.\n[6] I. G. Marchenko, V. Aksenova, I. I. Marchenko,\nJ. Luczka, and J. Spiechowicz, Physical Review E 107\n(2023), 10.1103/physreve.107.064116.\n[7] T. Colla, P. S. Mohanty, S. N¨ ojd, E. Bialik, A. Riede,11\nP. Schurtenberger, and C. N. Likos, ACS Nano 12,\n4321–4337 (2018).\n[8] F. Katzmeier, B. Altaner, J. List, U. Gerland, and\nF. C. Simmel, Physical Review Letters 128 (2022),\n10.1103/physrevlett.128.058002.\n[9] T. Hoang Ngoc Minh, G. Stoltz, and B. Roten-\nberg, The Journal of Chemical Physics 158 (2023),\n10.1063/5.0139258.\n[10] T. Vater, M. Isele, U. Siems, and P. Nielaba, Physical\nReview E 106(2022), 10.1103/physreve.106.024606.\n[11] T. Vissers, A. van Blaaderen, and A. Imhof,\nPhysical Review Letters 106 (2011), 10.1103/phys-\nrevlett.106.228303.\n[12] K. R. S¨ utterlin, A. Wysocki, A. V. Ivlev, C. R¨ ath, H. M.\nThomas, M. Rubin-Zuzic, W. J. Goedheer, V. E. Fortov,\nA. M. Lipaev, V. I. Molotkov, O. F. Petrov, G. E. Mor-\nfill, and H. L¨ owen, Physical Review Letters 102(2009),\n10.1103/physrevlett.102.085003.\n[13] H. Huang, V. Nosenko, H.-X. Huang-Fu, H. M.\nThomas, and C.-R. Du, Physics of Plasmas 29(2022),\n10.1063/5.0096938.\n[14] B. Cui and A. Zaccone, Phys. Rev. E 97, 060102 (2018).\n[15] R. W. Zwanzig, Nonequilibrium Statistical Mechanics\n(Oxford University Press, New York, NY, 2001).\n[16] T. Hoang Ngoc Minh, G. Stoltz, and B. Roten-\nberg, The Journal of Chemical Physics 158,\n104103 (2023), https://pubs.aip.org/aip/jcp/article-\npdf/doi/10.1063/5.0139258/16790539/104103 1online.pdf.\n[17] A. Barnaveli and R. van Roij, Soft Matter 20, 704 (2024).\n[18] A. Caldeira and A. Leggett, Annals of Physics 149, 374\n(1983).\n[19] S. Khrapak, Physics Reports 1050 , 1 (2024), elementary\nvibrational model for transport properties of dense fluids.\n[20] J. D. Jackson, Classical Electrodynamics , 1st ed. (John\nWiley & Sons, New York, NY, 1967).\n[21] I. Di Terlizzi, F. Ritort, and M. Baiesi, J. Stat. Phys.\n181, 1609 (2020).[22] R. Robson, R. White, and T. Makabe, Annals of Physics\n261, 74 (1997).\n[23] R. F. Probstein, Physicochemical Hydrodynamics (Else-\nvier, 1989).\n[24] N. Bisai and A. Sen, Reviews of Modern Plasma Physics\n7, 22 (2023).\n[25] I. Furno, B. Labit, M. Podest` a, A. Fasoli, S. H. M¨ uller,\nF. M. Poli, P. Ricci, C. Theiler, S. Brunner, A. Diallo,\nand J. Graves, Phys. Rev. Lett. 100, 055004 (2008).\n[26] P. Ricci, F. D. Halpern, S. Jolliet, J. Loizu, A. Mosetto,\nA. Fasoli, I. Furno, and C. Theiler, Plasma Physics and\nControlled Fusion 54, 124047 (2012).\n[27] R. Metzler, J.-H. Jeon, A. G. Cherstvy, and E. Barkai,\nPhys. Chem. Chem. Phys. 16, 24128 (2014).\n[28] J.-P. Bouchaud and A. Georges, Physics Reports 195,\n127–293 (1990).\n[29] E. Ilievski, J. De Nardis, S. Gopalakrishnan, R. Vasseur,\nand B. Ware, Phys. Rev. X 11, 031023 (2021).\n[30] P. Siegle, I. Goychuk, and P. H¨ anggi, Phys. Rev. Lett.\n105, 100602 (2010).\n[31] E. G. Flekkøy, A. Hansen, and B. Baldelli, Frontiers in\nPhysics 9(2021), 10.3389/fphy.2021.640560.\n[32] L. Levi, Y. Krivolapov, S. Fishman, and M. Segev, Na-\nture Physics 8, 912 (2012).\n[33] M. Peccianti and R. Morandotti, Nature Physics 8, 858\n(2012).\n[34] E. Fermi, Phys. Rev. 75, 1169 (1949).\n[35] T. Tajima and J. M. Dawson, Phys. Rev. Lett. 43, 267\n(1979).\n[36] B. Cui, R. Milkus, and A. Zaccone, Phys. Rev. E 95,\n022603 (2017).\n[37] K. J. Willis, S. C. Hagness, and I. Kneze-\nvic, Applied Physics Letters 102, 122113\n(2013), https://pubs.aip.org/aip/apl/article-\npdf/doi/10.1063/1.4798658/13220140/122113 1online.pdf.\n[38] A. Zaccone, Phys. Rev. E 108, 044101 (2023).\n[39] R. Kubo, Reports on Progress in Physics 29, 255 (1966)." }, { "title": "1908.06862v1.Spectral_determinant_for_the_damped_wave_equation_on_an_interval.pdf", "content": "arXiv:1908.06862v1 [math-ph] 19 Aug 2019SPECTRAL DETERMINANT FOR THE DAMPED WAVE\nEQUATION ON AN INTERVAL\nPEDRO FREITAS\nDepartamento de Matem´ atica, Instituto Superior T´ ecnico , Universidade de Lisboa,\nAv. Rovisco Pais 1, P-1049-001 Lisboa, Portugal andGrupo de F´ ısica Matem´ atica,\nFaculdade de Ciˆ encias, Universidade de Lisboa, Campo Gran de, Edif´ ıcio C6,\nP-1749-016 Lisboa, Portugal\nJIˇR´I LIPOVSK ´Y\nDepartment of Physics, Faculty of Science, University of Hr adec Kr´ alov´ e,\nRokitansk´ eho 62, 50003 Hradec Kr´ alov´ e, Czechia\nAbstract. We evaluate the spectral determinant for the damped wave equat ion on\nan interval of length Twith Dirichlet boundary conditions, proving that it does not\ndepend on the damping. This is achieved by analysing the square of th e damped\nwave operator using the general result by Burghelea, Friedlander , and Kappeler on\nthe determinant for a differential operator with matrix coefficients .\nPACS: 46.40.Ff, 03.65.Ge\n1.Introduction\nWe consider the simple mathematical model of wave propagationon a damped string\nfixed at both ends given by\n∂2v(t,x)\n∂t2+2a(x)∂v(t,x)\n∂t=∂2v(t,x)\n∂x2, (1.1)\nwith the space variable xon an interval [0 ,T],v(0) =v(T) = 0 and a(x)∈C([0,T]).\nDespite its apparent simplicity, the problem is nontrivial and interest ing and has re-\nceived much attention over the last two decades – see, for instanc e, [GH11, FL17,\nCFN+91, CZ94, BF09].\nThe operator associated with (1.1) is non-selfadjoint and the asym ptotical location\nof its eigenvalues was determined to first order in [CFN+91, CZ94], wh ere it was shown\nthat eigenvalues λconverge to the vertical line Re λ=−/a\\}bracketle{ta/a\\}bracketri}htas their imaginary part\ngoes to±∞, where/a\\}bracketle{ta/a\\}bracketri}htdenotes the average of the damping function. The general\nasymptotic behaviour was analysed in [BF09], where further spectr al invariants were\ndetermined.\nComingfromothersourcesintheliterature, thenotionofdetermin antofamatrixhas\nbeen generalised to operators. In this analogy, we would like to obta in a regularisation\ncorresponding to the product of eigenvalues of the given operato r. If the considered\nE-mail addresses :psfreitas@fc.ul.pt, jiri.lipovsky@uhk.cz .\n12 DETERMINANT FOR THE DAMPED WAVE EQUATION\noperator Shas eigenvalues {λj}∞\nj=1, in agreement with [RS71] (see also [GY60, MP49])\nwe define the generalised zeta function associated with the operat orSby\nζS(s) =∞/summationdisplay\nj=1λ−s\nj,\nfor complex sin a half-plane such that the above Dirichlet series converges. The\nspectral determinant may then be defined by the formula\nDetS= e−ζ′\nS(0), (1.2)\nwhere the prime denotes the derivative with respect to the variable s. Note that the\nseries defining the zeta function will not, in general, be convergent fors= 0. We use\nthe definition of ζSfor the real part of slarge enough and understand the formula in\nthe sense of the analytic continuation of the generalized zeta func tion to the complex\nplane.\nThe spectral determinant was computed for the Sturm-Liouville op erator in [LS77],\nwhere an elegant expression using the solution of a corresponding C auchy problem was\npresented. This was extended to the case of quantum graphs in [AC D+00, Fri06]. We\npoint out that spectral determinants have several applications e .g. in quantum field\ntheory [Dun08].\nTo the best of our knowledge the spectral determinant for the da mped wave equation\nhad not been studied previously, so in this note we bring together th ese two topics and\nevaluate this object. From a mathematical perspecive there is also what we believe to\nbe the interesting feature of applying the concept of the determin ant of an operator to\na non-selfadjoint operator. Furthermore, and as we will see, the determinant does not,\nin fact, depend on the damping. This may be expected form a formal analysis, and\nour purpose is to give a rigorous justification of this fact.\nThis note is structured as follows. In the next section, we ellaborat e on the math-\nematical description of the model and state the main result. In Sec tion 3 we then\naddress the problem for the case without damping, as this already d isplays some of the\nimportant features which we will need to consider later, namely, the fact that the asso-\nciated zeta function will depend on the branch cut which is chosen fo r the logarithm.\nIn Section 4 we recall the general result of Burghelea, Friedlander , and Kappeler. In\nSection 5 we apply this result to the square of our operator, since a direct application\nis not possible. Finally, we find the sought determinant for the dampe d wave equation\nin Section 6.\n2.Basic setting and formulation of the main result\nEquation (1.1) may be written in a different form, namely,\n∂\n∂t/parenleftbigg\nv0(t,x)\nv1(t,x)/parenrightbigg\n=/parenleftbigg0 1\n∂2\n∂x2−2a(x)/parenrightbigg/parenleftbigg\nv0(t,x)\nv1(t,x)/parenrightbigg\nwhich will prove to be more convenient for our purposes. Using the a nsatzv0(t,x) =\neλtu0(x),v1(t,x) = eλtu1(x), we can translate the initial value problem into the follow-\ning spectral problem\nH/parenleftbigg\nu0(x)\nu1(x)/parenrightbigg\n=λ/parenleftbigg\nu0(x)\nu1(x)/parenrightbiggDETERMINANT FOR THE DAMPED WAVE EQUATION 3\nwhereHdenotes the matrix operator\nH=/parenleftbigg0 1\n∂2\n∂x2−2a(x)/parenrightbigg\n.\nThe domain of this operator consists of functions u(x) =/parenleftbigg\nu0(x)\nu1(x)/parenrightbigg\nwith components\nin the Sobolev spaces uj(x)∈W2,2([0,T]),j= 0,1 satisfying the Dirichlet boundary\nconditions\nuj(0) =uj(T) = 0, j= 0,1.\nOur main result is the following.\nTheorem 2.1. Assumea(x)∈C([0,T]), and let εbe a positive number such that there\nare no eigenvalues with phase on the interval [π−ε,π), Then the spectral determinant\nof the operator Hdoes not depend on the damping and equals ±2T, where the plus and\nminus signs correspond to whether we define λ−s\nj= e−slogλjin such a way that the\nbranch cut of the logarithm is λ=tei(π−ε), orλ=tei(2π−ε),t∈[0,∞), respectively.\nRemark 2.2. Note that a value of εas above always exists, since on any compact set\nthere are only a finite number of eigenvalues.\nRemark 2.3. The case of the damped wave equation where a potential is adde d to\nthe right-hand side of (1.1)may be treated in a similar fashion and the corresponding\ndeterminant also turns out to be independent of the damping t erm. We discuss this\nsituation in Remark 6.2.\n3.The case of a(x) = 0\nWe begin by considering the case without damping, and denote the co rresponding\noperator by H0. It is a simple exercise that its eigenvalues are of the form λj=ijπ\nT,\nj∈Z\\{0}. To obtain the spectral determinant in this instance, we start fro m the zeta\nfunction resulting from the definition (1.2). However, one must pro ceed carefully here,\nas the result depends on the definition of λ−s\njand, in particular, on which branch of\nthe logarithm we use when defining i−sand (−i)−s.\nFirst, we consider that the logarithm has the cut in the negative rea l axis, i.e. the\neigenvalues of H0in the upper half-plane are λj=jπ\nTeiπ\n2,j∈Nand the eigenvalues in\nthe lower half-plane are λj=jπ\nTe−iπ\n2,j∈N.\nThe generalized zeta function for this operator is\nζH0(s) =∞/summationdisplay\nj=1/bracketleftBigg/parenleftbiggjπ\nTeiπ\n2/parenrightbigg−s\n+/parenleftbiggjπ\nTe−iπ\n2/parenrightbigg−s/bracketrightBigg\n=∞/summationdisplay\nj=1/parenleftBig\ne−iπ\n2s+eiπ\n2s/parenrightBig/parenleftbiggjπ\nT/parenrightbigg−s\n= 2eslogT\nπcos/parenleftBigπs\n2/parenrightBig\nζR(s),\nwhereζR(s) =∞/summationdisplay\nj=1j−sis the Riemann zeta function. We obtain\n−ζ′\nH0(0) =−2logT\nπζR(0)−2ζ′\nR(0) = logT\nπ+log(2π) = log(2 T),4 DETERMINANT FOR THE DAMPED WAVE EQUATION\nwhere we have used ζR(0) =−1\n2andζ′\nR(0) =−1\n2log(2π). Hence the spectral determi-\nnant for the operator H0is given by\nDetH0= e−ζ′\nH0(0)= 2T .\nNow we are going to compute the determinant in the case where we ch oose the cut\nto be the positive real axis. The eigenvalues are λj=jπ\nTeiπ\n2,j∈N, for the upper half-\nplane and λj=jπ\nTe3iπ\n2,j∈Nfor the lower half-plane. The generalized zeta function is\nnow\nζH0(s) =∞/summationdisplay\nj=1/bracketleftBigg/parenleftbiggjπ\nTeiπ\n2/parenrightbigg−s\n+/parenleftbiggjπ\nTe3iπ\n2/parenrightbigg−s/bracketrightBigg\n= e−iπs∞/summationdisplay\nj=1/parenleftBig\neiπ\n2s+e−iπ\n2s/parenrightBig/parenleftbiggjπ\nT/parenrightbigg−s\n= 2e−iπseslogT\nπcos/parenleftBigπs\n2/parenrightBig\nζR(s).\nHence we have\n−ζ′\nH0(0) =−2logT\nπζR(0)+2iπζR(0)−2ζ′\nR(0) = logT\nπ−iπ+log(2π) =−iπ+log(2T).\nThe spectral determinant for the operator H0is\nDetH0= e−ζ′\nH0(0)=−2T .\n4.A general result\nThe starting point for finding the determinant for the damped wave equation is a\ngeneral result by Burghelea, Friedlander and Kappeler [BFK95]. Th is result gives a\nformula for the determinant of a more general matrix-valued oper ator on an interval.\nFor convenience, we state this result here, together with the nec essary definitions.\nDefinition 4.1. Let us for n∈Ndefine the operator A=2n/summationdisplay\nk=0ak(x)(−i)kdk\ndxk, where\nakarer×rmatrices, in general smoothly dependent on x∈[0,T]. We assume that\nthe leading term a2nis nonsingular and that there exist an angle θso thatspeca2n∩\n{ρeiθ,0≤ρ <∞}=∅. We assume the following boundary conditions at the end poin ts\nof the interval.\nαj/summationdisplay\nk=0bjku(k)(T) = 0,βj/summationdisplay\nk=0cjku(k)(0) = 0,1≤j≤n.\nHere,bjkandcjkare for each j,kconstant r×rmatrices and bjαj=cjβj=I(Idenotes\nther×ridentity matrix). The integer numbers αjandβjsatistfy\n0≤α1< α2<···< αn<2n−1,\n0≤β1< β2<···< βn<2n−1.\nMoreover, we define |α|=/summationtextn\nj=1αj,|β|=/summationtextn\nj=1βj. We define the 2n×2nmatrices\nB= (Bjk)andC= (Cjk), whose entries are r×rmatrices. Here 1≤j≤2n,DETERMINANT FOR THE DAMPED WAVE EQUATION 5\n0≤k≤2n−1and\nBjk:=/braceleftbigg\nbjkfor 1≤j≤nand 0≤k≤αj\n0 otherwise,\nCjk:=/braceleftbigg\nbj−n,kforn+1≤j≤2nand 0≤k≤αj−n\n0 otherwise.\nWe define a 2n×2nmatrixY(x) = (ykℓ(x)),0≤k,ℓ≤2n−1whose entries are\nr×rmatrices ykℓ(x)defined by\nykℓ(x) :=dkyℓ(x)\ndxk,\nwhereyℓ(x)is the solution of the Cauchy problem Ayℓ(x) = 0with the initial conditions\nykℓ(0) =δkℓI. We are interested in the value of the matrix Yat the point T.\nFinally, we introduce\ngα:=1\n2/parenleftbigg|α|\nn−n+1\n2/parenrightbigg\n,\nhα:= det\nwα1\n1... wα1n.........\nwαn\n1... wαnn\n,\nwherewk= exp/parenleftbig2k−n−1\n2nπi/parenrightbig\n. Similarly, we define gβandhβ. We denote by γj,j=\n1,...,rthe eigenvalues of the matrix a2nand define\n(deta2n)gα\nθ:=r/productdisplay\nj=1|γj|gαexp(igαarg(γj))\nwithθ−2π 10G (the strong field limit of the Hanle effect). To quantify the magne tic field\nchanges we turn to inversions for the Si IIQUVdata.\n2Other photospheric lines, not shown, also show emission cores in the FIRS flare footpoint spectra:\n1081.83 nm (Fe I), 1083.91 nm (Ca I), 1084.40 nm (Si I?).– 8 –\nFirst we will use properties of the Si Iline and continuum to constrain the depth of\npenetration of flare energy that is sufficient to change the temper ature structure in the\ndeep chromosphere and photosphere. At a first glance the contin uum intensity appears\nto change very little, if at all. But both p modes and granulation modify the continuum\nintensity at a level of a few percent at angular resolutions similar to t hose of FIRS (e.g.\nS´ anchez Cuberes et al.2000), making relatively small changes difficult to see in slit rasters.\nCareful examination of the FIRS spectra obtained beginning at 17:4 6:16 UT, show a signif-\nicant brightening of 4 ±1% above levels in the neighboring spectra taken 13 seconds before\nand after, close to the flare kernel observed in RHESSI and line dat a. Evidence for this is\nshown in Figure 5. These data include variations in the transparency and seeing conditions\nof the atmosphere and hence vary significantly from row to row in th e figure. As is obvious\nin the figure, transparency variations were strongest in the first scan, getting progressively\nweaker in the second and third scans. However, relative intensities alongeach row in each\npanel can be fairly compared. The uncertainty quoted above is a 1 σstatistical variation of\nthe detrended intensity measured along the rows immediately adjac ent to the row containing\nthe flare.\nThe coherent streak of brightness in continuum data from 17:46:16 has all the charac-\nteristics of a genuine brightness increase associated with a white ligh t flare. This picture is\nsupported by HMI continuum data from the SDO spacecraft, which shows a≈5% increase\nin continuum intensity during the flare. Its appearance only in one FI RS scan indicates a\nvery rapid evolution, characteristic of an origin from dense photos pheric material which has\na radiative relaxation time of 1-2 seconds (Spiegel 1957).\n3. Analysis\nThe spatio-temporal behavior of the flare as obtained by FIRS is su mmarized in Fig-\nures 3 and 4. Remarkably, the slit happened to scan across the flar e footpoint ribbons at\nthe flare peak, 17:46 UT (Table 1). Also shown on the figure is the cor e of the flare-related\nacoustic source (Donea and others 2014). It is clear that FIRS ma naged to capture those\nlocations in solar- yheliographic coordinate that correspond to the time and place of th e\nacoustic source.– 9 –\n3.1. Helioseismic holography\nSeismic transients from solar flares can be detected by pre-proce ssing solar data and\napplying the analytical technique of helioseismic holography to Dopple r measurement of\nthe active region hosting the solar flare. Donea and others (2014) analyzed Doppler maps\nfrom the Helioseismic and Magnetic Imager instrument (HMI; Schou et al.2012) on board\nthe Solar Dynamics Observatory satellite (SDO). HMI measures pro perties of photospheric\ndynamics and magnetic fields every 45 seconds. Donea and others ( 2014) generated Postel\nprojection maps of the seismic emission of NOAA 12017. We refer disc ussion of the prin-\nciples of seismic holography to Section 4 of Lindsey and Braun (2000) , with application to\nflare observations to Donea et al.(1999). Briefly, the seismic responses to the flare pertur-\nbations are identified through an excess of the emission power, |H+(r,t)|2. Each pixel in\nan image of |H+(r,t)|2is a representation of the coherent acoustic power for waves tha t\nhave propagated downward from the focus, traveled thousands of kilometers beneath the\nsolar surface, and re-emerged into a pupil a significant distance fr om the focus. With this\ntechnique Donea and others (2014) uncovered a weak but significa nt seismic source at the\nfootpoint shown in Figure 1. Hard X-ray emission, magnetic transien ts and strong UV foot-\npoint emission were analyzed by Donea and others(2014), confirmin g thattheseismic source\nis indeed associated with the flare.\n3.2. The depth of penetration of flare energy\nBy comparing the brightness of models of continuum and Si I1082.7 nm line to ob-\nservations, we can in principle constrain the depth in the atmospher e to which significant\nheating from above can penetrate. Our approach is simple. We ask: what are the deepest\nand shallowest layers in the atmosphere heated by the flare that ar e compatible with the\ndata?\nTo preface the model calculations below, we note that the flare Si Iprofile (Figure 2)\nresembles classical Ca IIHandKself-reversed profiles (Linsky and Avrett 1970), but with\nfar weaker line absorption wings. The Ca IIline cores, much more opaque than the line\nof silicon, form in the chromosphere with a source function dominate d by scattering. The\nsimple observation of a self-reversed profile of Si Iimplies a significant column mass, much\nhigher than that for the calcium line. The breadth of a Doppler-broa dened, self-reversed\nline is larger than an optically thin line formed under the same conditions by the factor\n≈√lnτ0whereτ0is the line center optical depth. For τ0= 100 this factor is over 2.1. The\nself-reversal is also very narrow (FWHM ≈0.015 nm, see Figure 2), indicating turbulent\nspeeds of FWHM/1.66 ≈2.5 kms−1where the line core forms. A profile averaged along the– 10 –\nregion with obvious Si Iemission in Figure 2 is shown in Figure 6. The averaging washes\nout the self-reversal in the latter plot.\nWe model the Si Iline and the neighboring IR continuum both during and outside\nof the impulsive phase. We performed nLTE radiative transfer calcu lations, in several 1D\nmodels of the solar atmosphere, following the tradition of Vernazza et al.(1973, 1976, 1981,\nhenceforth VAL81). We solved nLTE statistical equilibrium equation s for atoms of H, C,\nSi and Fe using the program RH (Uitenbroek 2000). These atoms we re chosen because\nUV radiation controlling the Si Ispectral line at 1082.7 nm is dependent on the nLTE\nsolutions of these abundant elements. We considered using one of s everal flare models (e.g.\nMachado et al.1989). However, these models were constructed to try to identif y the origin\nof white light emission in flares. Our goal is different, to try to see if mo deling can provide a\ndepth of penetration of flare energy into the photosphere. Ther efore we adopted a different,\nmore straightforward strategy. We started with the model “C” o f VAL81 and explored the\neffects of introducing temperatures plateaus of the form\nTe=T0−T′\n1logm, m 2> m,\nwheremis the column mass of the atmosphere, T0,T′\n1are non-negative constants, and m2\nis a column mass above which temperatures are changed. We have th ree free parameters,\nand so our results will not be unique. But such plateaus, with small gr adientsT′\n1, have\njustification at least during some phases of flaring plasmas seen in ra diation hydrodynamic\ncalculations (see the 50s panel of Figure 3 of Allred et al.2005, for example). The main\nsensitivity of the emerging spectra is to the two parameters aandm2. Given an estimate of\nm2the height of the energy penetration follows from the m(z) relationship for the model.\nWe made calculations in two limits: in the calculations shown in the Figures below\nwe allowed the atmosphere to relax to a state of hydrostatic equilibr ium; in the other limit\nwe merely solved the statistical equilibrium equations with no such adj ustment. The sound\ncrossing time of the photosphere is on the order of a few scale heigh ts divided by 7 kms−1, a\nminute or so, comparable to the duration of the flare impulsive phase . These limits probably\nspan the behavior of intensities from an evolving atmosphere. The d ifferences between the\ncalculationsaresmallinphotosphericlayersbutaresignificantforr egionsandspectraformed\nabove 600 km above the photosphere. Such differences do not affe ct our conclusions which\ndepend only on the photospheric Si Iline.\nThese calculations are not state-of-the art in terms of dynamics, our focus is instead on\na careful treatment of the formation of the Si I1082.7 nm line and of the continua formed\nbetween 125 and 180 nm for later comparisons with SDO/AIA data. W e therefore took\ncare to use modern and complete atomic data for the Si and Fe neut rals. We used atomic\nenergy levels and transition probabilities from NIST up to and including the 4plevels in– 11 –\nSi, and we used photoionization cross sections from the OPACITY pr oject (Seaton 1987),\ntreated as outlined in Judge (2007). Collisions with electrons were tr eated using the impact\napproximation for permitted transitions (Seaton 1962), Seaton’s semi empirical formula for\ndirect ionization (Allen 1973), and a collision strength of 0.1 for forbid den transitions.\nFigure 6 shows, in the right panels, computed and observed profiles of SiI1082.7 nm,\nwith all intensities normalized to quiet Sun values. These calculations a re representative of\ntwo limits of the value of m2– and hence height of penetration – used in the models.\nThe first class (upper panel) allows penetration of energy and enha nced temperatures\ndown to photospheric layers - we allowed temperatures to rise down to 0 km height by adding\nvarious plateaus at such depths. Remarkably, the model shown pr oduces an acceptable\nmatch to the observed profiles and continuum (the He Iline is not modeled here). Exploring\ndifferent temperature plateaus we determined that a reasonable a greement with the line and\ncontinuum observations requires the flare energy to penetrate a nd heat down to a height of\n∼>100±100 km above the photosphere. The “error bar” comes from the n eed to produce\nthe 4% enhancement in continuum emission ( <200 km) with temperatures that can match\nthe SiIprofile, both features spanning the region between 0 and 700 km.\nThe second limiting case is one where flare energy penetrates only to the mid-upper\nchromosphere. Downward propagating radiation enhances the co res of lines, a typical calcu-\nlation is shown in the lower panels. The line width is very narrow even tho ugh we adopted\nnon-thermal speeds (microturbulence) of up to 8 km/s in the middle chromosphere (close\nto the sound speed). The continuum, formed predominantly in the p hotosphere with a tiny\ncontribution from optically thin emission in the plateau, is close to the p re-flare level. The\ncomputed continuum includes thermal photospheric emission as well as hydrogen recombi-\nnation from the plateaus. These two contributions have been discu ssed by Machado et al.\n(1989); Kerr and Fletcher (2014), among others. The contribut ion from the latter is small\nin our calculations, the Balmer continuum originating from an optically t hick layer near 350\nkm and the longer-wavelength ( >364 nm) H−and Paschen continua near 0 km.\nThe core of the Si Iline during the flare is broad compared with a thermal width near\n1.8 kms−1(Figure 6), and like the well-studied Ca IIHandKlines the origin of this width\nappears most naturally explained through scattering (see above) . Some decades ago there\nwas a discussion of the Wilson-Bappu effect, an empirical relationship between the width\nof the core of the Ca IIlines and stellar luminosity, in favor of line formation in terms of\nscattering (Ayres 1979) and not optically thin micro- or macro-tur bulence (Fosbury 1973).\nThe presence of the narrow self-reversed core seems irrefutab le evidence for the presence\nof scattering and argues strongly for a deep formation of the cor e. Only calculations of\npenetration of flare energy to the photosphere produce lines bro adened by scattering and– 12 –\nself-reversals, the latter happen to be weak in the case shown in Fig ure 6, but not in obvious\ndisagreement with the observed profile.\nWe stress that the detailed structure of our calculations is not uniq ue and should only\nbe viewed as an attempt to find the depth of penetration of significa nt heating during the\nimpulsive phase of the flare. Overall, our comparisons with observat ions of the Si I1082.7\nnm line, and taking into consideration the difficulties of tying down the c ontinuum intensity\nduring the flare, we conclude that heating sufficient to change detectably the photospheric\ntemperature occurs at least to about 100 ±100km above the visible photosphere . Based on an\nexploration of values of T0,m2in our model, we believe that this aspect of our calculations\nis robust.\n3.3. Inversions\nWe used the code MELANIE (Socas-Navarro 2003) to invert the Si IStokesIQUV\nprofiles to derive the vector magnetic field in the photosphere. This was done only for\nscans obtained before and after the impulsive phase. Codes exist f or inversion of the He I\nmultiplet (e.g. L´ opez Ariste and Casini 2002; Lagg et al.2004; Asensio Ramos et al.2008),\nbutwehavenotattemptedsuchinversions yetbecausewemustde alwithsignificant crosstalk\nin the He IQUprofiles during the flare, and because outside of the flare these pr ofiles are\nmostly of low signal-to-noise ratio.\nThe observed Si Iline – 3p4s3Po\n2−3p4p3Pe\n2(lower and upper levels respectively) –\nformsbetween ≈100km (wings) and600km (core) above thephotosphere inour 1Dm odels.\nMELANIEsolvesforasolutiontotheMilne-Eddingtonequations(sou rcefunctionlinearwith\noptical depth) for lines with Zeeman-induced polarization, minimizing d ifferences between\nobserved and computed profiles. The solution includes the vector m agnetic field (with its\n180◦ambiguity), opacity, Doppler width and shift, damping parameter, n on-magnetic filling\nfactor. The Milne-Eddington approximation is a simplification that sur ely is invalid during\nthe flare itself. But before and after the flare its use appears rea sonable, outside of bright\nflare ribbons and below say 600 km in the atmosphere. Our conclusion s will be based only\non the non-flaring atmosphere.\nWe inverted all five scans. We set the statistical uncertainty of ea ch data point to\n10−3ICto evaluate values of χ2, estimated using the measured fluctuations in QUVat\ntypical continuum wavelengths. For comparison, some of the best vector polarimetric data,\nthe “deep mode high S/N” observations from the SP instrument on t he Hinode spacecraft\nhave rms noise of 3 ×10−4ICin the 630 nm region, for integrations of 67 s (Lites et al.2008).– 13 –\nOutside of the flare scan, the distribution of χ2peaks near 40, showing that systematic errors\nare large and/or the model parameterization is poor. Given the res idual fringing and other\nartifacts evident in the data, this does not by necessity imply that t he model is poor. The\nreproducibility of the inversions was tested by initializing the same dat aset with two different\nrandom initial guesses. The resulting rms variations in the magnetic fi eld strength Bare\n140 G, inclination 18◦, azimuth 41◦, and the LOS B30 G.\nFigures 7 and 8 show results of inversions of the scans obtained bef ore and after the\nflare, begun at 16:29:26, 16:55:58, and 18:01:55 UT. No attempt at a r esolution of the 180◦\nambiguity in the field azimuth has been made, and the angles are define d relative to the local\nvertical3(inclination) and in the plane of sky (azimuth, zero and 180◦being along the E-W\ndirection). Circles show the location of the center of the acoustic s ource. Figure 7 shows\nmeasured changes in magnetic parameters in the two scans obtaine d before the flare. There\nare detectable differences across the bulk of the field of view in all ma gnetic parameters.\nFocusing on data in the circled region of the acoustic source, we see a significant increase\nin the field strength in this region, accompanied by becoming more inclin ed to the vertical\ndirection (data shown in the first two rows of the Figure). Note tha t the circled region is\nsome 5′′from the polarity inversion line. The maps of Bsuggest that a channel of weak\nfield is moved to the west by an arcsecond. Initially the field is inclined at some 130◦to\nthe vertical. But by 17:05 UT two bands of field connected in a “Y-sha pe” on its side in\nthe image appear more inclined to the vertical. The field azimuth in the “ Y” shape departs\nsignificantly from initially E-W to more N-S. The LOS field within at the circ le’s center\nshows an increase that results from increases in Bdespite the decrease in inclination. It\nis unclear from our data if these changing fields arise from motions of field vertically (flux\nemergence) or horizontally (flows). There is little evidence for vert ical motions from the LOS\nvelocity measurements shown in Figure 3, but the inversion data (no t shown) reveal a very\nsmall (-0.3 kms−1) blue-shift pattern in the Si Idata in the 10:55:58 UT scan that might\nconceivably be associated with the upper part only of the “Y” patte rn seen in the magnetic\ndata.\nThe scans upon which the inversions are based are26 minutes apart . The above changes\nare unremarkable when compared to the larger field of view, except that they are within the\ncircle encompassing the acoustic source and they are significant in a ll magnetic parameters.\nFigure8showsmeasuredchangesbeforeandaftertheflareitself , scansbegun66minutes\napart. The difference panels show again an increase of Band azimuth, and a weak reduction\nof inclination, in a band in the E-W direction cutting through the circled region, flanked\n3The vertical direction of center of the region is rotated 39.6◦E-W and 15.8◦S-N relative to the LOS.– 14 –\nby regions of increased inclination just to the S and N. This sheared r egion (differentially\nchanging field inclinations with time) appears aligned with the bright foo tpoint emission\nseen in the core intensities of the Si Iand HeIlines. The data are noisy, however.\nThus, our analysis hints that magnetic fields associated with the par ticular acoustic\nsource evolve to become more sheared (i.e. inclination angles divergin g in time), stronger\n(perhaps due to flux emergence) and rotated relative to the EW dir ection, during the flare.\nThese results appear to correspond to a mixture of earlier results . Wang et al.(2012a)\nfound penumbral fields which became more vertical after flaring. I n contrast, for some flares\nMart´ ınez-Oliveros et al.(2008); Wang et al.(2012b) reported field lines highly inclined to\nthe vertical after a flare-associated seismic transient. We note t hat the seismic source we\nhave analyzed is unusual. It is found near a magnetic pore, emerging from a magnetically\nquieter area somewhere between the main two sunspots of the AR1 2017 (Donea and others\n2014), instead of in a penumbra.\nLastly, if flux emergence were responsible for these measured cha nges in magnetic field,\nin 1 hour the plasma and magnetic field moving vertically through the co mpressible sub-\nphotosphere with a surface velocity∼<0.3 kms−1could have emerged from depths no deeper\nthan≈200 km. If advected by granules with 1 kms−1speeds, the flux could have emerged\nfrom no deeper than 600km. It is interesting to consider how such c hanges to the immediate\nsubsurface structure might or might not affect the generation of sunquakes.\n3.4. The mode of transfer of flare energy down through the atmo sphere\nArmed with a unique dataset, we have studied the depth of penetra tion of flare energy\ndown into the solar photosphere. We have shown that the detecte d changes in thermal\nstructure in the atmosphere reach the photospheric level, but ba rely. Here we examine\npossible modes by which the energy might be transported through t he photosphere into the\ndeeper solar layers, thereby exciting the sunquake.\nThe power in the main kernel of the acoustic source measured using seismology from\nHMI is (Donea and others 2014):\nP= 1.3±0.05×1026erg s−1.\nThis power is distributed over an area including the main kernel cente red at (X,Y) =\n(518,264) (see Figure 1), the source just to the SW requiring an addition al 1.0×1026erg s−1.\nThe mainsource’s spatialdistribution isnearly bi-Gaussianwith ageom etricmeanfull width\nat half-maximum (FWHM) of w= 4.2 HMI pixels, w≡1.5×108cm. The peak of the power– 15 –\nper unit area is F=P/(A=πa2) witha=w/2√\nln2, or\nF= 5×109erg cm−2s−1.\nThis should be regarded as a lower limit since both the holographic tech nique and HMI have\nnon-negligible angular resolutions. The area A= 2.6×1016cm2is strictly an upper limit\nfor the same reasons.\nLet us consider first “non-magnetic mechanisms” by which energy is transported to the\nacoustic source. In this picture the changing magnetic field genera tes thermal perturbations\nindirectly via the end product of large coronal magnetic restructu ring (conduction, particles,\nlocal downward radiative heating), channeling some flare energy int o the photosphere. We\ncan estimate energy fluxes into the acoustic source region that ar e compatible with our\nobservations in several ways. First, we note that the excess the rmal energy radiated from\nthe photosphere during the few minutes of the rise phase is roughly 4-5% (i.e. the measured\ncontinuum enhancement) of the unperturbed solar radiative flux d ensityF⊙= 6.33×1010\nergscm−2s−1:\nPRAD≈0.04F⊙A∼<7×1025erg s−1.\nThe radiative cooling time of photospheric plasma is 1-2 s (Spiegel 195 7). Curiously then,\nalthough PRAD∼P, this excess thermal energy is simply radiated into space on such\ntimescales, and is unavailable to contribute to P. We can look at the enthalpy fluxFenth\nassociated with bulk flows into the photosphere, for this we need a m easurement of plasma\nmotions and we turn to the Si Iline core emission which forms between 200 and 500 km in\nour models. We use pressures p= 2×104dyne cm−2and densities ρ= 4×10−8g cm−2.\ncorresponding to 300 km height. These are conservatively high valu es for average ther-\nmal properties of the plasma where this line is formed, favoring highe r estimates of energy\ntransport.\nA careful comparison of the flare emission core and the pre-flare a bsorption profile of\nthe SiIline reveals an upper limit to differential flows of roughly 0.5 wavelength pixels,\n0.5 kms−1. This is equivalent to 2.5 σwhereσis the sensitivity of the Doppler shifts\nfrom our FIRS spectra. A Doppler photospheric signature of the fl are is present in HMI\ndata at the location of the seismic source with a shift equivalent to u≈0.3−0.5 kms−1\n(Donea and others 2014). However, such filtergram data, scann ing wavelengths in time,\ncannot be trusted during flaring and so we adopt the upper limit abov e. We then find an\nupper limit to the enthalpy energy flux of\nFenth∼<5\n2puA≈6×1025erg s−1.\nThe close agreement of the upper limit for FenthwithFradmeans that the excess energy\nradiated by the photosphere during the flare can be supplied by a bu lk flow of energy– 16 –\nassociated with a subsonic downflow of 0.5 kms−1induced (somehow) by the flare. The\npower in the acoustic pulse is a factor of at least 2 larger than our op timistic estimate of\nFenth.\nIf however the pressure pulse involves high frequency phenomena (ν > cS/H≈60 mHz,\nwhereHis the pressure scale height and cSthe sound speed), the pulse would be invisible\nto observation except as a broadening of spectral lines to at most the sound speed (for linear\nwaves), the lines being formed over a length ≈Hin a stratified atmosphere. The WKB\nexpression for the energy flux density (propagating both upward s and downwards) at the\nsound speed is\nFwave=ρcS/angbracketleftξ2/angbracketrighterg cm−2s−1,\nwhereξis the velocity amplitude of the wave. We can set limits on ξthrough the measured\nline broadening and line profiles during the flare itself. Before and aft er the flare, the\ninversions yield ξ∼<2.8 kms−1. During the flare the measured line wings are similar in\nshape to the pre- and post- flare profiles. The emission core of the profile has a FWHM of\n0.05 nm (Figures 2 and 6). Treated as optically thin emission, this FWHM is equivalent\nto an e-folding Doppler broadening speed of 8 kms−1. In the presence of scattering this\nis a strict upper limit. To estimate the energy flux available in such mode s we again use\nρ= 4×10−8g cm−2, and the upper limit of 8 kms−1. Assuming that only half of the waves\nare emitted downwards, we find\nFwaveA∼<2.5×1026erg s−1.\nBut this, we believe, is a gross over-estimate. Firstly, the scatter ing leads to emission profiles\na factor of ≈√lnτ0broader than mere Doppler broadening where τ0is the line center optical\ndepth. We are able to reproduce the core Si Iemission using microturbulent speeds of 1-\n2 kms−1and the full Voigt profile, reducing the above estimate Fwaveby a factor of at\nleast 16! Secondly, the line profile shows, within a broad emission core , a very narrow self-\nreversal during the flare (lower left panel of Figure 2), indicating b oth scattering-induced\nline broadened profiles and values of ξin the line core far smaller than those adopted above.\nLastly, any high frequency waves with frequencies of a few Hz or les s are rapidly damped\nin the photosphere by the continuum radiative exchange processe s first modeled by Spiegel\n(1957). The thermal perturbations associated with high frequen cy wave energy rapidly\nradiatethis energy fromthephotosphere ona timescale of1-2s. O nlywaves withfrequencies\nin excess of several Hz could propagate down into the interior unmo dified by radiation\ndamping. All things considered, it seems unlikely that the power of th e sunquake can be\nprovided by such high frequency waves.\nWe conclude that non-magnetic modes of energy transport int o the interior are very\nunlikely to be sound waves . More likely is a coherent downward-moving plug of plasma– 17 –\ncarrying enthalpy of almost the right magnitude, but we have above already set an upper\nlimit to this process that is optimistically a factor of two smaller than ne eded.\nConsider in turn the energetics of the Lorentz force picture. The field strength from\ninversions from the Si Iline, before and after the flare, is of order 800 G from which the\nmagneticenergydensityis B2/8π= 2.5×104ergcm−3,about1/3ofthephotosphericthermal\nenergy density,3\n2pph(the latter is a lower limit since we neglect latent heat of ionization).\nThe Alfv´ en speed cAfor a photospheric density of 2 .6×10−7g cm−3is 4.3 kms−1, so the\nlocal magnetic energy flux is at most\nFMA∼1, so that the same conclusions are\ndrawn as for the non-magnetic case. The magnetic forces are either (a) incompatible with\nthe observations of line profiles revealing down-flowing mat erial with insufficient energy flux\nto account for the acoustic source, or (b) are in the form of mo dified high frequency sound\nwaves with the same problems that the sound waves have.– 18 –\n4. Discussion and Conclusions\nWehave reportedsome unusual spectral andpolarimetricprofiles oflines ofSi IandHeI\nobtained at a flare footpoint at infrared wavelengths. Several ph otospheric lines revealed line\ncore emission in excess of non-flaring conditions, and the helium 1083 nm multiplet almost\ndoubled in brightness. Using the enhanced brightness of the Si I1082.7 nm line and the\nneighboring continuum, we have demonstrated via 1D radiative tran sfer models that flare-\nrelated heating can be detected all the way to the photosphere, t o 100±100 km. This is\ndeeperbyseveralscaleheightsthantheexpecteddepthofpene trationofhardXrayemission.\nOur models merely show the depth to which heating is observed to occ ur through the flare\nmechanism(s), they shed no light on the nature of this transport f rom the corona into the\nSun’s deeper atmosphere. It could be a direct mechanical effect or a two-stage mechanism\nof mechanical transport followed by radiative back warming (Macha doet al.1989). Using\ndynamical models in a stratified atmosphere, Allred et al.(2005) found the flare energy\npenetrated only to the mid chromosphere (800 km). Mart´ ınez Olive roset al.(2012) found\nheights of 305 ±170 km and 195 ±70 km, respectively, for the centroids of the hard X ray\nand white light footpoint sources of a flare observed stereoscopic ally.\nHow robust is our new result? We can produce significant emission in th e core of\nthe SiIline using both deep and shallow penetration models (enhanced tempe ratures only\nabove 500 km). But the shallow models can be eliminated: the continuu m intensities are\nunchanged from pre-flare conditions; the Si Iline cores are far too narrow; also, only models\nwith energy penetrating down into the photosphere have sufficient opacity to produce the\nopacity broadening and subtle narrow self-reversal observed in t he very center of the line.\nThe combination of broad, self reversed line emission and a brighter c ontinuum appears to\nbe a clear signature of flare heating down to 100 ±100 km. This picture also eliminates\nthe possibility that hydrogen recombination radiation contributes t o the visible and infrared\ncontinuum emission for this flare (Kerr and Fletcher 2014). One sho uld expect that our\nconclusions depend strongly on the adopted model of the photosp here/chromosphere: surely\n1D models are inadequate for studies of the Sun’s atmosphere in gen eral? It is sometimes\nforgotten that the photosphere/low chromosphere are charac terized by subsonic motion, and\nthat therefore the atmosphere is strongly stratified4. It is this essential property of these\nplasmas, in which the Si Idiagnostic line we have used is formed, that is most critical for\ndetermining the depth of penetration of flare energy. There appe ars to be no credible way to\nreconcile the salient line profile and continuum observations with heat ing occurring purely\n4Very dynamic phenomena seen at UV wavelengths, for example with t he HRTS or IRIS instruments, are\nformed mostly in less dense structures above the stratified layer t hat is optically thick to most UV radiation.– 19 –\nabove 200 km. Whatever model we would choose to use, we would still require continuum\nand line formation within the photosphere, not the chromosphere.\nWe found significant pre- and post-flare changes in the magnetic co nditions within and\noutside of the acoustic source region. Within the source region, th ese include an increase in\nmagnetic field strength, a rotation of the magnetic azimuth and a re duction of inclination\nwith respect to the solar vertical. The further interpretation of t hese changes, part of the\nevolution of the entire active region, is beyond the scope of this pap er.\nWe detected no signature of downward energy transport capable of carrying 2 .3×1026\nerg s−1needed to account for the acoustic source, that is compatible with our ground-based\nobservations. Curiously, both line and continuum emission is present along an extended\nribbon, but the acoustic emission is confined to a smaller region only as sociated with the\nbrightest parts of the ribbon. At some future time, the flux of ene rgy insomething propa-\ngating downwards through the Sun’s atmosphere must be detecte d. We have shown using\ncurrent spectroscopic capabilities that macroscopic motions drive n magnetically or otherwise\ncarry insufficient energy by at least an order of magnitude, and tha t high frequency acoustic\nand magnetic waves are also are very likely to fall short. Spectropo larimetric analysis reveals\nmagnetic fields that are too small at the acoustic source for magne tic wave modes to carry\nthe energy ( cA∼0∀bfor\nboth massive and massless particles. We can rescale the supe rcharges defining Q′\nα=√aQαand\ndropping the primes we obtain\n/braceleftbig\nQα,¯Q˙α/bracerightbig\n= 2σˆµ\nα˙αPˆµ+2cσ¯µ\nα˙αP¯µ. (3)\nTo be consistent with the definition of weighted power counti ng (1) the constant chas to be\ndimensionless and to weight [c] = 1−1/n. The weight of supercharges is fixed by the commutation\nrelations and is equal to their dimension:\n[Q] =/bracketleftbig¯Q/bracketrightbig\n=1\n2.\nWe can reabsorb the modification of the SUSY algebra by redefin ing the Minkowski metric ηµν\nand theσmatrices [ 22]:\nσ′ˆµ=σˆµ, σ′¯µ=cσ′¯µ,\n/parenleftbig\nσ′µ¯σ′ν+σ′ν¯σ′µ/parenrightbigα\nβ=−2δα\nβη′µν=−2δα\nβ/parenleftbig\nηˆµˆν+c2η¯µ¯ν/parenrightbig\n,\n✷′=η′µν∂µ∂ν=∂ˆµ∂ˆµ+c2∂¯µ∂¯µ,\np′2=η′µνpµpν=pˆµpˆµ+c2p¯µp¯µ (4)\nand after these redefinitions the algebra ( 3) looks like the usual one. Therefore the structure of\nthe superspace is identical to the Lorentz invariant case as was observed in [ 29]. In particular we\ncan define, as usual, the covariant derivatives so that {D,Q}=/braceleftbig\nD,¯Q/bracerightbig\n= 0. In our construction\nthe crucial requirement is the linearity of the supercharge s in the space momenta P¯µ. This\nassumption makes possible the usual definitions of the N= 1superspace as a coset space defined\nby the set of variables zπ= (ˆx,¯x,θ,¯θ). The operator U(ˆy,¯y,η,¯η) =ei(ˆyˆP+¯y¯P+iηQ+i¯η¯Q)is a\nwell defined translation in superspace and we can define a scal ar superfield S(ˆx,¯x,θ,¯θ)so that\nδUS=−iˆyˆP−i¯y¯P+ηQS+¯η¯QS. From the Leibniz rule for the supercharges and the definitio n of\ncovariant derivatives it follows that every polynomial in t he superfields and its covariant derivatives\nis still a superfield.\n52.2 Higher-momenta superalgebras\nThe Lorentz-violating case admits a larger class of superal gebras compatible with the Coleman-\nMandula theorem and the weighted power counting. The antico mmutator {Qα,¯Q˙α}is in the\n(1/2,1/2)representation of the Poincaré algebra and has dimension 1. In the Lorentz-invariant\ncase this implies that it has to be proportional to σµPµ, which is the only operator of dimension\n1in the same representation. In the Lorentz-violating case w e can construct new operators ¯O¯µof\nweight[¯O] = 1and dimension 1in the vector representation of the so(¯d)algebra. These operators\nare simply weighted polynomials of odd degree in the momentu m¯P\n¯O¯µ=/summationdisplay\nkak¯P2k¯P¯µ\nΛ2k\nL,withk≤/bracketleftbiggn−1\n2/bracketrightbigg\n,\nwhere the constants akweight[ak] = 1−(2k+1)/nand[ak]≥0. The parameter nshould be\nunderstood as the highest power of ¯Pthat appears in the quadratic terms of the lagrangian, as\nexplained in [ 9]. Therefore we can construct new superalgebras allowed by t he Coleman-Mandula\ntheorem and by the weighted power counting:\n/braceleftbig\nQα,¯Q˙α/bracerightbig\n= 2σ′µ\nα˙αPµ+2σ¯µ\nα˙α¯O¯µ,/braceleftbig\nDα,¯D˙α/bracerightbig\n=−2σ′µ\nα˙αPµ−2σ¯µ\nα˙α¯O¯µ. (5)\nThe new operators ¯O¯µhave the same commutation rules as ¯Pµin the SUSY algebra because,\nremovingJˆµ¯νfrom the original super-algebra sp,¯P2becomes a Casimir operator for the new\nsuper-algebra sp′. If we take the low-energy limit ΛL→ ∞ in (5) we obtain again ( 3). In\nprinciple sp′has the same superspace as spbecause we have not introduced new supercharges,\nbut now our supercharges will be non-linear operators in ¯P:\nQα=∂\n∂θα−iσ′µ\nα˙α¯θ˙α∂µ−iσ¯µ\nα˙α¯θ˙α¯O¯µ, (6)\nso the operator U(ˆy,¯y,η,¯η) =ei(ˆyˆP+¯y¯P+iηQ+i¯η¯Q)is not a simple translation in superspace any-\nmore. We can still define a superfield Sso thatδUS=−iˆyˆP−i¯y¯P+ηQS+ ¯η¯QS, but now it\nis in general not true that every polynomial in Sis still a superfield because the supercharges\n(6) do not respect Leibniz rule: δU(S1S2)/ne}ationslash=δUS1S2+S1δUS2. For this reason the superfield\nformalism loses its usefulness and it is not possible to prom ote directly the usual supersymmetric\nlagrangians for matter and gauge fields to lagrangians which are symmetric under the modified\nSUSY agebra ( 5). The construction of a recent publication [ 28] is completely invalidated by this\nobservation. Actually, that construction works only for fr ee theories because the Leibniz rule\nfor the supercharges ( 6) is fulfilled up to a total derivative if we consider only bili nears in the\nsuperfields. As an example of our general argument we conside r the same theory as [ 28], namely\nfor (1,3) splitting we take the superalgebra ( 5) forn= 3. For the chiral supermultiplet we can\nwrite a kinetic lagrangian:\nLkin=ˆ∂φ†ˆ∂φ+(c1¯∂+¯∂2¯∂\nΛ2\nL)φ†(c1¯∂+¯∂2¯∂\nΛ2\nL)φ+χ†(iˆ/∂+c1i¯/∂+i¯/∂3\nΛ2\nL)χ+F†F . (7)\n6The lagrangian is invariant under the transformations\nδηφ=√\n2ηχ ,\nδηχ=i√\n2(σˆµ¯η∂ˆµ+σ¯µ¯η(c1∂¯µ+¯∂2∂¯µ\nΛ2\nL))φ+√\n2ηF ,\nδηF=i√\n2¯η(¯σˆµ∂ˆµ+c1¯σ¯µ(c1∂¯µ+∂2∂¯µ\nΛ2\nL))χ , (8)\nwhich are clearly generated by supercharges that satisfy th e superalgebra ( 5) forn= 3. The\nauthor [ 28] claims that if we want to add interactions invariant under ( 8) to this theory it is\nsufficient to rephrase the usual superfield formalism for the n ew supercharges. Following his\nderivation, from the definition of the chiral superfield ¯D˙αΦ = 0 we obtain:\nΦ(x,θ) =φ(x)+iθσˆµ¯θ∂ˆµφ(x)−iθσ¯µ¯θ(c1∂¯µ+¯∂2∂¯µ\nΛ2\nL)φ(x)+1\n4θ2¯θ2(ˆ∂2+2c1¯∂4\nΛ2\nL+¯∂6\nΛ4\nL)φ(x)+\n+√\n2θχ(x)+i√\n2θ2¯θ(¯σˆµ∂ˆµ+ ¯σ¯µ(c1∂¯µ+¯∂2∂¯µ\nΛ2\nL))χ(x)+θ2F(x).(9)\nIf the superfield formalism worked, then each holomorphic fu nction of the superfield ( 9) would be\ninvariant under ( 8). As an example of a possible interaction lagrangian let us c onsider the usual\ncubic interaction in the superfields:\nLint=g\n3/integraldisplay\nd2θΦ3+h.c=gFφ2−gφψψ+h.c. .\nNow varying Lintwith respect to ( 8) we obtain\nδLint=i√\n2¯η[∂ˆµ(¯σˆµχφ2)+∂¯µ(c1¯σ¯µχφ2)+∂¯µ\nΛ2\nL(¯∂2χφ2−2¯σ¯ν∂¯µχ∂¯νφφ+2¯σ¯ν(∂¯µ∂¯νφφ+∂¯µφ∂¯νφ))]\n+i2√\n2¯η¯σ¯µχ∂¯ν∂¯µφ∂¯νφ .\nBecause of the last term in the previous equation δLint/ne}ationslash=∂ˆµAˆµ+∂¯µB¯µand therefore the action\nis not invariant under the transformations ( 8). This was expected, since the supercharges are\nnon-linear in the spatial derivatives.\nClearly, the possibility of constructing interacting quan tum field theories which are invariant\nunder the super-algebra ( 5) is not ruled out by our observation and could be an interesti ng open\nfield of research which, however, needs more involved constr uctions.\nIn the rest of this paper we will restrict ourselves to the cas e in which the supercharges are\nassumed to be linear in the spatial momenta and we will work wi th the superymmetric algebra\n(3).\n3 Renormalizable Lorentz-Violating Theories for Chiral Su per-\nfields\n3.1 General Discussion\nThe integration measure in four dimensions for a general spl itting4 =ˆd+¯dweights−đ≡ −(ˆd+¯d\nn),\nso a renormalizable lagrangian has to be a weighted polynomi al in the momenta of weight [L] =đ\n7and dimension 4, as we want to keep the action dimensionless and weightless.\nConsidering the super-algebra ( 3) we take a chiral superfield Φ(ˆx,¯x,θ,¯θ)in theN= 1super-\nspace defined by the constraint ¯DΦ = 0 :\nΦ(ˆx,¯x,θ,¯θ) =φ+iθσ′µ¯θ∂µφ+θ2¯θ2\n4✷′φ+√\n2θχ+i√\n2θ2¯θ¯σ′µ∂µχ+θ2F .\nThe complex scalar field φ(x)and the Weyl spinor χ(x)are propagating quantum fields, and we\ncan derive their weights from their kinetic lagrangian [ 9]. The chiral superfield then weights:\n[Φ] = [φ] =đ−2\n2= [θ]+[χ] = [θ]+đ−1\n2. (10)\nThe equation ( 10) fixes the weight of the spinor coordinates to [θ] =−1/2∀n, which is equal\nto the difference of weight between scalars and fermions and i s constant with respect to n. The\nspinor coordinates in the superspace do not rescale with nand their weights are equal to their\ndimensions so that, as in the Lorentz invariant case,/bracketleftbig\nd4θ/bracketrightbig\n= 2and/bracketleftbig\nd2θ/bracketrightbig\n=/bracketleftbig\nd2¯θ/bracketrightbig\n= 1∀n.\nWe want to construct the most general Lorentz-violating lag rangian for Mdifferent chiral\nsuperfields Φiwithi= 1...M. The Kähler potential K[Φ,¯Φ]weights[K] =đ−2and the\nsuperpotential f[Φ]weights[f] =đ−1. If we demand polynomialilty of the lagrangian, [Φ]>0,\nwe obtain:\nL(ˆd,¯d)=/integraldisplay\nd4θ¯ΦiΦi+/summationdisplay\nα,N/integraldisplay\nd2θλN,α\nNΛN(d−2)/2+p1+p2−3\nL/bracketleftBig\nˆ∂p1¯∂p2ΦN/bracketrightBig\nα+h.c., (11)\nwhereαlabels possible different derivative structures and a combi natorical factor can appear in\nthe denominator if there are identical superfields. The most general Kähler potential which is\nrenormalizable by weighted power counting has the same form as the Lorentz invariant one. This\nobservation severely restricts the possibility of constru cting supersymmetric Lorentz-violating\nmodels and at the same time will have important implications in the study of the RG flow at low\nenergies of the cparameter. In the superpotential, the derivative structur e of the vertex defines\na monomial in the superfield momenta of weight δα\nN=p1+p2/n. If we want to preserve CPT\ninvariance and symmetry under rotation in the submanifold M¯dwe have to assume that p1and\np2are even numbers. The coupling constant λα,Nassociated to a vertex with Nsuperfields is a\nsymmetric tensor with Ninternal indices i1...iN, whereik= 1...M∀k, and by power counting\nit has to weight\n[λN,α] =đ−1−N[Φ]−δα\nN. (12)\nWe will call a vertex weighted marginal when its coupling con stant weights [λN,α] = 0, weighted\nrelevant when [λN,α]>0and weighted irrelevant when [λN,α]<0. As it has been shown in\n[9], the renormalization rules in the Lorentz-violating case work as in the Lorentz invariant one\nafter the substitution d→đ, so that we can express the renormalizabilty condition imp osing that\nthe weight of the coupling constant has to be greater or equal to zero,[λN,α]≥0. As we need\n[Φ]>0for polynomiality, taking N= 2in this inequality we can derive an upper bound on\nthe weight of the monomial in the momenta δα\nN≤1, that ensures perturbative unitarity of the\ntheory and forbids the presence of terms with time derivativ es in the superpotential. We write\n8L(ˆd,¯d)=Lkin+Lint, where the kinetic term of the lagrangian ( 11) is\nLkin=/integraldisplay\nd4θ¯ΦiΦi+/integraldisplay\nd2θ\n/summationdisplay\nl≤[n/2](al)ij\n2Λ2l−1\nL¯∂lΦi¯∂lΦj\n+h.c.. (13)\nOnly terms with an even number of space derivatives are allow ed and the index lin the sum is\nan integer that goes from zero to the integer part of the ratio [n/2]. The higher space derivatives\nterms generalize the mass term in the Wess-Zumino model [ 30] and are regulated by the coupling\nconstants (al)ij, which are M×Msymmetric matrix of weight [al] = 1−2l/n. The diagonal\nterms ofalbehave as Majorana mass terms, whereas the off-diagonal term s behave as Dirac mass\nterms. In order to simplify the notation we omit the internal indicex structure and take the free\npartition function of a theory with only one chiral superfiel d, in which the coupling constant al\nbecomes a coefficient and we can construct only Majorana mass t erms.\nZ0/bracketleftbig\nJ,¯J/bracketrightbig\n=/integraldisplay\nDΦD¯Φexp−/braceleftBigg/integraldisplay\nd8z/bracketleftBigg\n1\n2/parenleftbig\nΦ¯Φ/parenrightbig\nA/parenleftbiggΦ\n¯Φ/parenrightbigg\n−/parenleftbig\nΦ¯Φ/parenrightbig/parenleftBigg\nD2\n4✷′J\n¯D2\n4✷′¯J/parenrightBigg/bracketrightBigg/bracerightBigg\nandA=/parenleftBigg\nA11D2\n4✷′1\n1A22¯D2\n4✷′/parenrightBigg\nwhereA11=A22=\n/summationdisplay\nl≤[n/2](−)lal\nΛ2l−1\nL¯∂2l\n.\nFrom the partition function we can derive the propagators fo r the chiral superfield using the\nmethods of [ 31]:\n/angbracketleftbig\nΦ(1)¯Φ(2)/angbracketrightbig\nJ=0=δ12\np′2+/parenleftbigg/summationtext\nl≤[n/2]al\n2Λ2l−1\nL(¯p2)l/parenrightbigg2, (14)\n/an}bracketle{tΦ(1)Φ(2) /an}bracketri}htJ=0=D2\n4\n/summationdisplay\nl≤[n/2]al(¯p2)l\nΛ2l−1\nL\nδ12\np′2/parenleftBigg\np′2+/parenleftbigg/summationtext\nl≤[n/2]al(¯p2)l\nΛ2l−1\nL/parenrightbigg2/parenrightBigg, (15)\nwhere we have reabsorbed the −D2\n4factors in the Feynman rules for the vertices and p′2is defined\nin (4). If we differentiate the propagators ( 14) and ( 15) with respect to any coefficient alwith\nl<[n/2]the weight of the denominator increases by 1−2l/nand differentiating with respect to\ncincreases by 1−1/n. Hence we can make any Feynman graph convergent by differenti ating it\na suitable number of times with respect to the coefficients alorc, and the counterterms will be\npolynomials in alandc. We can consider the super-renormalizable operators assoc iated with the\ncoefficients alandcas vertices with two external lines and treat them perturbat ively. Doing that,\nwe can study the UV behavior of the Lorentz-violating theori es keeping in the propagators ( 14)\n9and (15) only the terms with the maximum number of spatial derivativ es:\nPΦ¯Φ=δ12\nˆp2+a2\n[n/2](¯p2)2[n/2]\nΛ4[n/2]−2\nL, (16)\nPΦΦ=D2\n4a[n/2](¯p2)[n/2]\nΛ2[n/2]−1\nLδ12\np′2/parenleftbigg\nˆp2+a2\n[n/2](¯p2)2[n/2]\nΛ4[n/2]−2\nL/parenrightbigg. (17)\nThe weight of the coefficient a[n/2]is zero for even nand strictly positive for odd n, and in fact\nthe operators/parenleftbig¯∂[n/2]Φ/parenrightbig2in the free lagrangian ( 13) are strictly renormalizable for even nand\nsuper-renormalizable for odd n. Therefore, for odd nwe cannot construct propagators which\nare the inverse of homogeneous polynomials of weight 2 and th is fact completely invalidate our\nconstruction. To understand what does not work in the odd ncase we compare the kinetic terms\nof the fermionic and the scalar lagranians in the non-supers ymmetric case for an arbitrary n[9]:\nLs=1\n2(ˆ∂φ)2−c2\n2(¯∂φ)2−n/summationdisplay\nl=2a2\nl\n2Λ2l−2\nL(∂lφ)2−m2\n2φ2, (18)\nLf=¯ψ(iˆ/∂+vi¯/∂+n/summationdisplay\nloddb′\nl\nΛl−1\nL(i¯/∂)l+n/summationdisplay\nlevenbl\nΛl−1\nL(i¯/∂)l−M)ψ . (19)\nFrom the equations of motion associated with the lagrangian s (18) and ( 19) we derive the corre-\nsponding dispersion relations1:\nE2\ns(¯p) =c2¯p2+n/summationdisplay\nl=2a2\nl¯p2l\nΛ2l−2\nL+m2, (20)\nE2\nf(¯p) = ¯p2(v+n/summationdisplay\nloddb′\nl\nΛl−1\nL¯pl−1)2+(M+n/summationdisplay\nlevenbl\nΛl−1\nL¯pl)2. (21)\nIn the Lorentz invariant case the dispersion relation among energy and spatial momentum is\nuniversal for all particles E2(¯p) =c2¯p2+m2. Conversely, we see from ( 21) that in the Lorentz-\nviolating case the dispersion relation for fermions contai ns two different kind of contributions\nthat are related respectively to terms with an even or an odd n umber of derivatives in the kinetic\nlagrangian ( 19). This happens because the terms with an even number of deriv atives behave like\nmass terms from the point of view of the spin 1/2representation of the Lorentz group, while\nthe terms with an odd number of derivatives behave like ¯/p. A necessary condition for the theory\nto be supersymmetric is that all the dispersion relations of particles in the same supermultiplet\nhave to be the same. We have shown that all the higher spatial d erivatives terms that behave\nlike masses can be supersymmetrized by adding appropriate F -terms to the superpotential. On\nthe contrary, all terms with an odd number of higher spatial d erivatives would correspond to\n1For simplicity we write the dispersion relations in the case of(1,3)splitting, in which there is a natural\nidentification of the energy with the only component of the fo ur momentum whose weighted dimension coincides\nwith the usual one. The extension to different splittings is s traightforward.\n10modifications of the Kälher potential. However, a Kähler pot ential which is renormalizable by\nweighted power counting should have the same form as the Lore ntz invariant one and therefore\nwe cannot construct a supersymmetric version of a theory wit h an odd number of higher spatial\nderivatives in the fermionic kinetic term. If nis odd this means that it is not possible to construct\na supersymmetric version of a free theory with scalars and fe rmions. For even nwe can construct\nsupersymmetric theories in which dispersion relations for fermions will be of the form ( 21) with\nb′\nl= 0∀l.\nThe Feynman rules for the vertices remain unchanged with res pect to the Lorentz invariant\ncase [31] because the Lorentz-violating terms do not modify the θ-structure of Feynman graphs.\nTherefore the divergent contributions to the effective ener gy are polynomials in the external\nmomenta of the form\nΓ∞=/integraldisplay\nd4xd4θF(Φ,¯Φ,DΦ...,¯∂Φ,...). (22)\nBy power counting we obtain [F] =đ−2, so that the non-renormalization theorem [ 32] for the\ndivergent contributions still works in the Lorentz-violat ing case and the divergent contributions\naffect only the Kähler potential. We can calculate the superfi cial divergence for a generic Feynman\ngraphGatLloops, with Ipropagators, Vvertices and Eexternal lines2\nω(G) = (đ−2)L−2I−E+/summationdisplay\nN,αvN(N−1+δα\nN)\n=d(E)−2−/summationdisplay\nN,αvα\nN[λα,N],whered(E) =đ−E[Φ](23)\nand the weights of the chiral superfield Φand of the coupling constant are defined in ( 10) and ( 12).\nFor renormalizable theories, taking E= 2yields an upper bound for the superficial divergence,\nω(G)≤0, that is in agreement with the result of the non-renormaliza tion theorem ( 22). Therefore\nin a supersymmetric Lorentz-violating theory the Kähler po tential can receive radiative corrections\nwith logarithmic divergences if and only if there are strict ly renormalizable interactions. If the\ntheory contains only super-renormalizable interactions, then the theory is finite. In the Lorentz\ninvariant Wess-Zumino model, the non-renormalization the orem ensures that the behavior of the\ntheory at different energies is regulated only by the wave fun ction renormalization constant. In our\nmodels this is not true anymore. However, we can derive relat ions among different renormalization\nconstants. The most general renormalizable interaction la grangian contains two different kinds of\ncomposite operators:\nLint⊃/integraldisplay\nd2θ/braceleftBigg\nλk\nΛk−3\nLΦk+λ′\nk,l,α\nΛk+2l−3\nL/bracketleftBig\n¯∂2lΦk/bracketrightBig\nα/bracerightBigg\n+h.c..\nDemanding the renormalizability of the theory it is easy to s ee thatkis an integer number\n20andc/ne}ationslash= 0the kinetic term in the Kähler potential always introduces a non-homogeneous\nterm in the propagator, which breaks the weighted scale inva riance.\nWe can define Lorentz-violating homogeneous theories if [c] = 0 andn= 1. In this case we\nobtain the model proposed in [ 22], in which the interaction sector is equal to the Wess-Zumin o\nmodel. However, we can obtain another class of homogeneous t heories by taking c= 0. In this\ncase the supercharges algebra ( 3) becomes:\n/braceleftbig\nQα,¯Q˙α/bracerightbig\n= 2σˆµ\nα˙αPˆµ. (27)\n12The resulting algebra ( 27) is the usual N= 1supersymmetry algebra in d= 4, but projected\non the submanifold Mˆd. It is clear that the Kähler potential for the chiral superfie lds associated\nwith this algebra does not introduce non-homogeneous terms in the propagators. Therefore we\ncan define a class of free homogeneous lagrangians for every e ven value of n, inserting in the\nsuperpotential ( 13) only the bilinear with the maximum number of spatial deriva tives, regulated\nby the weightless constant an/2. The propagators are the inverse of homogeneous polynomial s of\nweight2:\n/an}bracketle{tΦ(1)¯Φ(2)/an}bracketri}ht=δ12\nˆp2+a2\nn/2¯p2n\nΛ2n−2\nL,\n/an}bracketle{tΦ(1)Φ(2) /an}bracketri}ht=D2\n4an/2(¯p2)n/2\nΛn−1\nLδ12\nˆp2/parenleftBig\nˆp2+a2\nn/2(¯p2)n\nΛ2n−2\nL/parenrightBig. (28)\nFrom these free theories we can construct interacting lagra ngians by adding all the renormalizable\nterms in the superpotential. When c= 0, if we consider only the strictly renormalizable interac-\ntions in the superpotential, we then obtain homogenous inte racting theories. At the quantum level\nthe fixed points of the renormalization group for these theor ies are still invariant under weighted\nscale transformations, but far from the fixed points the symm etry is anomalous. In the low-energy\nlimit we cannot restore the usual N= 1supersymmetry algebra in d= 4and we cannot obtain a\nLorentz-invariant theory with usual propagators because o f the weighted scale invariance; in fact\nsuper-renormalizable terms cannot be produced by renormal ization because they would break the\nweighetd scale invariance. In the IR limit the propagators ( 28) become\nlim\nΛL→∞/an}bracketle{tΦ(1)¯Φ(2)/an}bracketri}ht=δ12\nˆp2,lim\nΛL→∞/an}bracketle{tΦ(1)Φ(2) /an}bracketri}ht= 0. (29)\nThe propagators ( 29) do not depend on ¯p, so that all the diagrams constructed with these propa-\ngators are not computable, because they contain divergence s that no counterterms can eliminate.\nHence the IR limit is singular. From homogenous theories we c an construct non-homogeneous the-\nories invariant under the algebra ( 27), by adding to the superpotential all the super-renormaliz able\nterms. Doing that, we are breaking the weighted scale invari ance, but terms fundamental for the\nLorentz symmetry recovery such as φ∗∂¯µ∂¯µφori∂¯µ¯ψ¯σ¯µψare not generated by renormalization\nbecause they break the symmetry of the lagrangian under supe rsymmetry transformations ( 27).\nThis means that for c= 0Lorentz symmetry cannot be restored and the IR limit of these theories\nis still singular.\n3.3 Classification of neutral chiral superfield’s models\nWe want to classify all the possible theories invariant unde r the superalgebra ( 3) for all possible\nsplittings of the four dimensional spacetime manifold. The basic ingredient of such classification\nwill be the maximum number of legs for a chiral vertex defined i n (24).\n1.For splitting (0,4) we have ¯N=/bracketleftBig\n24−n\n4−2n/bracketrightBig\n.The only solution is the n= 1and¯N= 3and\nwe thus recover the Wess-Zumino model.\n132.For (1,3) splitting we obtain ¯N=/bracketleftBig\n23\n3−n/bracketrightBig\n. Forn= 1we find again the Lorentz-violating\nWess-Zumino model proposed in [ 22,29]. Takingn= 2we find ¯N= 6, which is the only\nnon-trivial solution of the condition ( 24). Hence, for (1,3)splitting the only theory with\nstrictly renormalizable interactions has n= 2and đ= 5/2. The Lagrangian, requiring\nsymmetry under the transformation Φ→ −Φ, will be\nL=/integraldisplay\nd4θ¯ΦΦ+/integraldisplay\nd2θ/bracketleftbigg(¯∂Φ)2\n2ΛL+mΦ2\n2/bracketrightbigg\n+/integraldisplay\nd2θ/bracketleftbigg\nλ4Φ4\n4!ΛL+λ6Φ6\n6!Λ3\nL/bracketrightbigg\n+h.c..(30)\nWe can derive the propagators for the superfields and expand t hem for high momenta in\norder to study the UV behavior of the theory:\nPΦ¯Φ=δ12\nˆp2+(c2+2m\nΛL)¯p2+¯p4\nΛ2\nL+m2≃δ12\nˆp2+¯p4\nΛ2\nL, (31)\nPΦΦ=D2\n4¯p2\nΛLδ12\nˆp2+(c2+2m\nΛL)¯p2+¯p4\nΛ2\nL+m2≃D2\n4¯p2\nΛLδ12\np′2/parenleftBig\nˆp2+¯p4\nΛ2\nL/parenrightBig. (32)\nThe first divergent radiative correction to the Kähler poten tial is at 4loops:\nΦ ¯Φ=λ2\n6\n5!/integraldisplaydˆk\n2πd3¯k\n(2π)3/integraldisplay\nd4θ1d4θ2Φ(−k,θ1)Φ(k,θ2)BD.\nThe covariant derivatives algebra is easy to compute if we re call the usual identities\nD=δ12D2¯D2\n16δ12¯D2D2\n16δ12D2¯D2\n16δ12D2¯D2\n16δ12=δ12.\nTherefore, the superfields computation is reduced to the com putation of the bosonic integral\nB, which is not easily computable because of the modified form o f the propagators ( 31) and\n(32).\n3.For(2,2)splitting ¯N=n+ 2and we can construct theories with strictly renormalizable\ninteractions for any even n. For example we can choose n= 2and write the complete\ntheory:\nL=/integraldisplay\nd4θ¯ΦΦ+/integraldisplay\nd2θ/bracketleftbigg(¯∂Φ)2\n2ΛL+m\n2Φ2/bracketrightbigg\n+/integraldisplay\nd2θ/bracketleftbiggλ3\n3!Φ3+λ4\n4!ΛLΦ4/bracketrightbigg\n.\nThe kinetic term is the same of the theory n= 2for(1,3)splitting and the propagators are\n(31) and ( 32).\n14The first divergent radiative correction to the Kähler poten tial is at 2loops:\nΦ ¯Φ=λ2\n4\n3!/integraldisplayd2ˆk\n(2π)2d2¯k\n(2π)2/integraldisplay\nd4θ1d4θ2Φ(−k,θ1)Φ(k,θ2)BD\nAgain, for this graph D=δ12but the bosonic integral Bis again very hard to compute.\n4.In the(3,1)case we obtain ¯N=/bracketleftBig\n22n+1\nn+1/bracketrightBig\nthat has only one integer solution for n= 1that\nis the trivial one. Therefore we can construct only super-re normalizable theories.\n5.For splitting (4,0)we obtain the Lorentz invariant Wess-Zumino model.\n4 Gauge invariant theories\nWe want to study the problem of finding a gauge invariant versi on of the Lorentz-violating su-\npersymmetric theories that we have found in section 3. First of all we apply a result of [ 24] to\nshow that it is not possible to generalize the theories of sec tion 3 to the case of charged chiral\nsuperfields. The basic observation in our construction was t hat it is possible to insert higher space\nderivatives as mass terms in the superpotential in ( 13) because they preserve the chirality of Φ,\nwhich is clear recalling that [Dα,∂¯µ] = 0. Charged chiral superfields transform with a phase under\nthe action of a general SU(N)gauge group\nΦi→eΛΦi,\nwhereΛis a chiral superfield. We can define a gauge invariant version of the supersymmetric\ncovariant derivative:\nDαΦi=e−VDα(eVΦi),\nwhereV(ˆx,¯x,θ,¯θ)is the vector superfield with its usual gauge trasformation l aw:eV→e¯ΛeVe−Λ.\nIt is clear that explicit spatial derivatives in the action b reak gauge invariance. In principle, as was\nobserved in [ 24], we could still introduce higher spatial derivatives in th e superpotential promoting\n∂¯µto a covariant derivative:\nD¯µΦ =−i¯σ˙αα\n¯µ\n4¯D˙αDαΦ. (33)\nThe problem, however, is that D¯µdoes not preserve the chirality condition. In fact, as it was\nchecked in [ 24]:\n¯D˙α¯D˙β/parenleftbig\ne−VDαeVΦi/parenrightbig\n= 2ε˙β˙αWαΦ/ne}ationslash= 0. (34)\nThis argument shows that the theories that we have construct ed are not generalizable to the case\nof charged chiral superfields.\nNow we will see directly in the gauge sector that, requiring g auge invariance and supersym-\nmetry, the weighted power counting has to coincide with the u sual one. We briefly review the\n15derivation of the weights for the gauge fields referring to [ 7,8] for a complete study of Lorentz-\nviolating gauge theories. In the Lorentz-violating case th e gauge field Aµhas to be decomposed\nas all the other vectors into time and space components A= (ˆA,¯A). Therefore the covariant\nderivative is decomposed as\nD= (ˆD,¯D) = (ˆ∂+eˆA,¯∂+e¯A), (35)\nwhereeis the gauge coupling constant. To be consistent with the defi nition of covariant derivative\n(35) we have to weight [eˆA′] = [ˆ∂] = 1and[e¯A] = [¯∂] = 1/n. The field strength is split into three\nparts:\nˆFµν≡Fˆµˆν,˜Fµν≡Fˆµ¯ν,¯Fµν≡F¯µ¯ν.\nThe kinetic lagrangian has to contain (ˆ∂ˆA)2, so we can obtain the weight of ˆA, that is equal to the\nweight of the scalar field, and from the definition of covarian t derivatives we derive the weights\nfor¯Aand[e] = 2−đ/2. Summarizing we have:\n[ˆA] =đ−2\n2,[¯A] =đ\n2−2+1\nn,[ˆF] =đ\n2,[˜F] =đ\n2−1+1\nn,[¯F] =đ\n2−2+2\nn.(36)\nThe requirement of absence of spurious subdivergences [ 7] implies that\nˆd= 1,đ<2+2\nn, d=even, n=odd,\nand the only acceptable splitting is (1,3). In this case we have ˆF= 0so it is possible to rearrange\nthe weights of the gauge field and the gauge coupling so that th e producteAmaintains the same\nweight and at the same time [˜F] =đ/2:\n[ˆA] =đ\n2−1\nn,[¯A] =đ\n2−1,[˜F] =đ\n2,[¯F] =đ\n2−1+1\nn,[e] = 1+1\nn−đ\n2.(37)\nIn the supersymmetric case both weight assignments ( 36) and ( 37) have to be consistent with\nthe relations among the weights of the fields imposed by the su persymmetric transformations\ngenerated by the supercharges ( 3). For the vector multiplet the supersymmetric transformat ion\nare:\nδηAµ= ¯η¯σµ′λ+¯λ¯σµ′η ,\nδηλ=i\n2σµ′¯σν′ηFµν+ηD ,\nδηD= ¯η¯σµ′∂µλ−∂µ¯λ¯σµ′η , (38)\nwhere the gaugino λis a propagating Majorana fermion, Dis an auxiliary field and ηis the\nspinorial parameter of the supersymmetry transformation. Since we know the weights of λ,ηand\nof the weighted constant c, we can obtain the weights of the other fields of the supermult iplet,\napplying the weighted power counting to the relations ( 38) yelds\n[ˆA] =đ−2\n2,[¯A] =đ\n2−1\nn,[D] =đ\n2,\n[ˆF] =đ\n2,[˜F] =đ\n2−(1−1\nn),[¯F] =đ\n2−2(1−1\nn). (39)\n16Any gauge theory that has a supersymmetric extension invari ant under the supersymmetry algebra\n(3) has to satisfy the constraint on the weight of the fields ( 39). Therefore, we can take the two\npossible weight assignments for Lorentz-violating gauge fi eld theories ( 36) and ( 37) and check for\nwhich value of nthese theories can admit a supersymmetric extension. Doing that we found that\nthe only possible value is n= 1in both cases and hence, as long as we require supersymmetry a nd\ngauge invariance, we have to weight time and space in the same way, regardless of the condition\nof absence of spurious divergences. The only Lorentz-viola ting operators are introduced by the\nweighted constant c. These operators are renormalizable in the usual sense and c orrespond to the\nCPT preserving operator in the gauge sector of the SME [ 13]. In particular they can be expressed\nintroducing a twisted derivative ˜∂µ=∂µ+kν\nµ∂ν[38], where in our case kµν= (c−1)δ¯µ¯ν. We will\nshow in the next section that the constant cdoes not renormalize in the supersymmetric case. As\nwe need to fine tune cin order to recover Lorentz symmetry at low energies, the Lor entz-violating\nparameters will be extremely small also at high energies. Th is argument shows that demanding\nrenormalizability and gauge invariance for supersymmetri c theories, the Lorentz invariance follows\nas a consequence.\n5 Low-energy limit and Lorentz symmetry recovery\nIf we consider Lorentz-violating theories as candidates to describe high-energy physics, then\nLorentz invariant theories are effective field theories whic h describe physics at low energies with\nrespect to ΛL. In the framework of renormalizable Lorentz-violating the ories it is still true that\nthe renormalization procedure commutes with the low-energ y limit. Therefore it is a general re-\nsult that a Lorentz-violating high-energy theory renormal izable by weighted power counting tend\nto a low-energy theory which is renormalizable by usual powe r counting and contains Lorentz-\nviolating parameters. The low-energy theory is then obtain ed from the high-energy one simply\nby taking the limit ΛL→ ∞. The recovery of Lorentz invariance at low energies is regul ated by\nthe RG behavior of the Lorentz-violating parameters at low e nergies that correspond to operators\nof dimension less than or equal to four, which are not suppres sed by any power of ΛL.\nIn this section we want to study how the Lorentz recovery prob lem at low energies is modified\nin presence of supersymmetry. If supersymmetry is an exact s ymmetry of nature at high energies\nthen in the low-energy limit supersymmetry has to be broken f or several phenomenological reasons\n[33]. Ignoring the exact mechanism of SUSY breaking we can param etrize the supersymmetry\nbreaking at low energies by adding explicit breaking terms t o the supersymmetric lagrangian.\nWe require the breaking to be soft, in the sense that the super symmetry breaking terms should\nnot generate quadratic divergences. It has been pointed out in [34] that the natural setting for\nstudying the low-energy limit of supersymmetric theories i n presence of soft breaking terms is the\nsuperfield formalism. The soft breaking terms are parametri zed as couplings among dynamical\nsuperfields and external spurion superfields. The possible t ypes of spurion superfields which break\nsupersymmetry softly for Lorentz invariant theories are cl assified in [ 34].\nOn this basis we can compute the low-energy limit for the gene ral supersymmetric Lorentz-\nviolating lagrangian ( 11). Assuming that the supersymmetry breaking soft terms are g enerated\n17at energyµs<<ΛLwe obtain the standard Lorentz-invariant soft terms:\nL=/integraldisplay\nd4θ¯ΦiΦi+/integraldisplay\nd2θ(mij\n2ΦiΦj+λijk\n3ΦiΦjΦk)+h.c.\n+/integraldisplay\nd4θUij¯ΦiΦj+/integraldisplay\nd2θij(χΦiΦj+ηijkΦiΦjΦk)+h.c.,(40)\nwhereUij=µ2\nsuijθ2¯θ2,χij=µ2\nsxijθ2andηijk=µsnijkθ2are the soft breaking spurion superfields\nandu, x, n are dimensionless matrices in the generations indices. We c an write the resulting\nsoft-breaking terms in components:\nLbreak=µ2((u+x)ijAiAj+(u−x)ijBiBj)+µnijk(AiAjAk−3AiBjBk). (41)\nAll the terms in ( 41) are super-renormalizable and they introduce new divergen ces in addition to\nthe usual wave function renormalization of the Wess-Zumino model, which are studied in [ 34,35].\nSince they all are super-renormalizable, these terms do not affect the divergent part of the wave\nfunction renormalization. Therefore, even for softly brok en supersymmetry, it is easy to see that\nthe Lorentz-violating parameter cdoes not renormalize, because it does not appear explicitly in\nthe superfield lagrangian. In the IR limit we need to fine tune caccording to the experimental\nbounds on Lorentz violation, but the low-energy value of cwill be also the value of the constant c\nat high energy, because cdoes not renormalize. Hence the deviation from the speed of l ight of the\nlimiting speed of elementary particle is negligible for Lor entz-violating supersymmetric models.\nWe will consider an explicit model in components in order to u nderstand how the behavior\nof the renormalization group equations is modified by the fac t that the effective theory at low\nenergies is the low-energy limit of a supersymmetric Lorent z-violating theory with soft breaking.\nFor this specific model we will explicitly show at one loop how the renormalization properties\nof a softly broken supersymmetric theory imply the non renor malization of c. Let us consider\nthe most general low-energy limit of a renormalizable super symmetric Lorentz-violating theory\nfor an interacting chiral multiplet. Since supersymmetry i s softly broken at low-energy, from the\nprevious discussion it is clear that we can parametrize this breaking considering as independent the\ndimensionfull parameters of the lagrangian. Therefore the low-energy lagrangian will be described\nby six independent parameters:\nL=(ˆ∂A)2\n2+c2(¯∂A)2\n2+(ˆ∂B)2\n2+c2(¯∂B)2\n2+m2A2\n2+m′2B2\n2+1\n2¯ψ(ˆ/∂+v¯/∂+M)ψ\n+λ3\n3!A3+λ3\n2′\nAB2+g2\n4!(A4+B4+6A2B2)+gA¯ψψ+igB¯ψγ5ψ .\nIn the approximation of small deviations from the speed of li ght we can put c2≃1 +δc2and\nv≃1+δv. The bare quantities are defined as:\nAb=Z1\n2\nAA , Bb=Z1\n2\nBB , ψb=Z1\n2\nψψ , δc2b=δc2+∆δc2, δc2b=δv+∆δv,\nm2\nb=m2+∆m2, m′2\nb=m′2+∆m′2\nB, Mb=M+∆M ,\nλ3b=µǫ\n2(λ3+∆λ3), λ′\n3b=µǫ\n2(λ′\n3+∆λ′\n3), gb=µǫ(g+∆g).\n18Recalling the Feynman rules for Majorana fermions [ 36] we obtain at one loop:\nZA=ZB= 1−g2\n2π2ǫ(1−¯dδv), Zψ= 1−g2\n2π2ǫ(1−¯d\n3δc2−¯d\n3δv),\n∆δc2=g2\n2π2ǫ(δc2\n2−δv),∆δv=g2\n2π2ǫ(2δv−δc2).\nIf supersymmetry is softly broken the divergent part of the w ave function renormalization has to\nbe the same for all particles in the same supermultiplet. The refore, imposing Zψ=ZAwe obtain\n2δv=δc2, which implies ∆δc2= ∆δv= 0. We thus conclude that the Lorentz-violating parameter\ndoes not renormalize. The only renormalization constant le ft is a common renormalization for the\nwave functions ZA=ZB=Zψ= 1−g2\n2π2ǫ(1−¯dδv), that at the zeroth order in δvis in agreement\nwith the well known result for the Wess-Zumino model.\nAn interesting problem to address concerns the nature of the constant which parametrizes\nthe deviation from the speed of light in supersymmetric Lore ntz-violating theories. In particular\nwe want to understand whether or not the weighted constant cis physically observable. As was\nalready noted in [ 24], if we consider a supersymmetric Lorentz-violating theor y with one single\nsector the parameter cappears both in the kinetic lagrangian ( 13) and in the supersymmetry\ntransformations ( 3) and can therefore be reabsorbed by a rescaling of the spatia l coordinates\nx′\n¯µ=cx¯µ. Now, as far as supersymmetry is an exact symmetry of our theo ry, all interacting\nsupermultiplets have the same limiting speed c, because this parameter explicitly appears in the\nsupersymmetry transformations. In this type of theories th e parameter cis physically unobserv-\nable because we can always set it to one by suitably choosing t he length units. However, as was\nsuggested in [ 37], we can construct more complicated situations with two or m ore sectors which\nare separately invariant under supersymmetry transformat ions (3) with different limiting speeds\nci, where the lower index labels the number of sectors. For exam ple let us consider two different\nsectorsS1andS2, separately invariant under the supersymmetry transforma tions\n/braceleftbig\nQα,¯Q˙α/bracerightbig\n= 2σˆµ\nα˙αPˆµ+2c1σ¯µ\nα˙αP¯µ, (42)\n/braceleftbig\nQα,¯Q˙α/bracerightbig\n= 2σˆµ\nα˙αPˆµ+2c2σ¯µ\nα˙αP¯µ. (43)\nIfS1andS2are completely decoupled, the supersymmetry algebras ( 42,43) are exactly realized\nin their respective sectors.3We can still rescale the spatial coordinates in order to reab sorbc1or\nc2but, after the rescaling, we will have a Lorentz invariant se ctorS1withc1/ne}ationslash= 1and an Lorentz\nviolating hidden sector S2completely decoupled with c2/ne}ationslash= 0. Moreover, we can make S1andS2\ninteracting by adding super-renormalizable interactions which will softly break supersymmetry in\nboth sectors. The deviation from the speed of light in S2then becomes experimentally observable\nbecause any rescaling performed in order to remove the c2factor will produce Lorentz-violating\neffects inS1, as was already pointed out in [ 37].\nIn conclusion, there is always the possibility to set one lim iting speed to one by rescaling the\nspatial coordinates. This fact can make the Lorentz-violat ing supersymmetric models presented\nin [22,37] physically equivalent to the Lorentz invariant ones if we r estrict our attention to models\n3The Lorentz-violating theories that we are considering are rigid supersymmetric theories with very good ap-\nproximation. Indeed the scale ΛLhas to be around 1014GeV in order to explain the neutrino masses [ 14]. Therefore\nwe can neglect gravitational effects and consider the two sec torsS1andS2completely decoupled.\n19with a single sector. Besides that we have shown that the param etercdoes not renormalize in\nany softly broken supersymmetric theory, so that even if the deviation from the the speed of light\nis indeed observable it would be extremely small at any energ y scale.\n6 Conclusion\nIn this paper we have investigated the possibility of constr ucting supersymmetric Lorentz-violating\ntheories that can be renormalized by a weighted power counti ng. Our analysis starts from the\nobservation that supersymmetry and Lorentz violation are c ompatible at the level of the alge-\nbra [22].\nMoreover, we have shown that in the Lorentz-violating case i t is possible to construct new\nsuperalgebras with supercharges non-linear in the spatial momenta. However, the non-linearity\nof the supercharges makes the problem of finding interacting theories invariant under the new\nsuperalgebras very involved, because the superfield formal ism looses its usefulness. As an exam-\nple of this general difficulty we have shown that the interacti ng theory constructed in [ 28] is not\ninvariant under this new class of superalgebras.\nAssuming linearity of the supercharges in the spatial momen ta we find renormalizable super-\nsymmetric Lorentz-violating models for even nand classify them. It is straightforward to verify\nin the superspace formalism that the non-renormalization t heorem is still valid in the Lorentz-\nviolating case and as a consequence our models exhibit the im proved ultraviolet behavior typical\nof supersymmetric theories. Moreover, the weighted consta ntcappearing in the supercharges\nalgebra which parametrizes the limiting speed of the multip let does not renormalize at high ener-\ngies because the Kähler potential of a renormalizable Loren tz-violating theory has the same form\nas in the Lorentz invariant case. Furthermore, the low-ener gy recovery of Lorentz symmetry is\nparametrized only by this parameter c, which does not renormalize even at low energies if we\nassume supersymmetry to be broken softly.\nIn the case of gauge theories we show that demanding supersym metry implies that the weighted\npower counting has to coincide with the usual one. The only Lo rentz-violating operators are then\nintroduced by the weighted constant c, which does not renormalize and has to be very close to\nthe speed of light at low energies in order to satisfy the expe rimental bounds on Lorentz violation\n[1]. Therefore, if we demand renormalizabilty, gauge invaria nce and supersymmetry, the Lorentz\ninvariance follows as a consequence.\nOur analysis agrees with the conjecture that supersymmetry can solve the Lorentz fine tuning\nproblem for Lorentz-violating theories, but at the same tim e it reveals that the requirement of su-\npersymmetry restricts drastically the possibility of cons tructing renormalizable Lorentz-violating\ntheories at high energies. Indeed, the final picture which em erges from our investigation is that the\nonly possible models with non trivial Lorentz-violating op erators involve neutral chiral superfields\nand do not have a gauge invariant extension. Therefore, if we want to construct renormalizable\nLorentz-violating extensions of the Standard Model which h ave new interesting phenomenological\nconsequences, the Lorentz fine tuning problem [ 21,23] does not seem solvable by the requirement\nof supersymmetry.\n20Acknowledgements\nThe author is grateful to D. Anselmi for suggesting the probl em and for useful discussions. He also\nthanks R. Argurio for useful comments. The research of D.R. i s supported by the “Communauté\nFrançaise de Belgique” through the ARC program, and also in pa rt by IISN-Belgium (conventions\n4.4511.06, 4.4505.86 and 4.4514.08) and by the Belgian Feder al Science Policy Office through the\nInteruniversity Attraction Pole IAP VI/11.\nReferences\n[1]V. Kostelecky and N. Russell, “Data Tables for Lorentz and CP T Violation,”\nRev.Mod.Phys. 83(2011) 11, arXiv:0801.0287 [hep-ph] . —2,20\n[2]S. Chadha and H. B. Nielsen, “Lorentz Invariance As a Low-Ener gy Phenomenon,”\nNucl.Phys. B217 (1983) 125 . —2\n[3]V. Kostelecky and S. Samuel, “Spontaneous Breaking of Lorent z Symmetry in String\nTheory,” Phys.Rev. D39(1989) 683 . —2\n[4]N. E. Mavromatos, “Lorentz Invariance Violation from Strin g Theory,” PoSQG-PH\n(2007) 027, arXiv:0708.2250 [hep-th] . —2\n[5]R. Gambini and J. Pullin, “Nonstandard optics from quantum s pace-time,”\nPhys.Rev. D59(1999) 124021 ,arXiv:gr-qc/9809038 [gr-qc] . —2\n[6]D. Anselmi, “Weighted scale invariant quantum field theorie s,”JHEP 0802 (2008) 051 ,\narXiv:0801.1216 [hep-th] . —2\n[7]D. Anselmi, “Weighted power counting and Lorentz violating gauge theories. I. General\nproperties,” Annals Phys. 324(2009) 874–896 ,arXiv:0808.3470 [hep-th] . —2,16\n[8]D. Anselmi, “Weighted power counting and Lorentz violating gauge theories. II.\nClassification,” Annals Phys. 324(2009) 1058–1077 ,arXiv:0808.3474 [hep-th] . —2,16\n[9]D. Anselmi and M. Halat, “Renormalization of Lorentz violat ing theories,”\nPhys.Rev. D76(2007) 125011 ,arXiv:0707.2480 [hep-th] . —2,5,6,8,10,12\n[10]A. Dhar, G. Mandal, and S. R. Wadia, “Asymptotically free fou r-fermi theory in 4\ndimensions at the z=3 Lifshitz-like fixed point,” Phys.Rev. D80(2009) 105018 ,\narXiv:0905.2928 [hep-th] . —2\n[11]A. Dhar, G. Mandal, and P. Nag, “Renormalization group flows i n a Lifshitz-like four fermi\nmodel,” Phys.Rev. D81(2010) 085005 ,arXiv:0911.5316 [hep-th] . —2\n[12]R. Bluhm, “Overview of the SME: Implications and phenomenolo gy of Lorentz violation,”\nLect.Notes Phys. 702(2006) 191–226 ,arXiv:hep-ph/0506054 [hep-ph] . —2\n21[13]D. Colladay and V. Kostelecky, “Lorentz violating extensio n of the standard model,”\nPhys.Rev. D58(1998) 116002 ,arXiv:hep-ph/9809521 [hep-ph] . —2,17\n[14]D. Anselmi, “Weighted power counting, neutrino masses and L orentz violating extensions of\nthe Standard Model,” Phys.Rev. D79(2009) 025017 ,arXiv:0808.3475 [hep-ph] . —3,19\n[15]D. Anselmi, “Renormalization of Lorentz violating theorie s,”. — 3\n[16]D. Anselmi, “Standard Model Without Elementary Scalars And High Energy Lorentz\nViolation,” Eur.Phys.J. C65(2010) 523–536 ,arXiv:0904.1849 [hep-ph] . —3\n[17]D. Anselmi and E. Ciuffoli, “Low-energy Phenomenology Of Sca larless Standard-Model\nExtensions With High-Energy Lorentz Violation,” Phys.Rev. D83(2011) 056005 ,\narXiv:1101.2014 [hep-ph] . —3\n[18]D. Colladay and P. McDonald, “One-Loop Renormalization of P ure Yang-Mills with\nLorentz Violation,” Phys.Rev. D75(2007) 105002 ,arXiv:hep-ph/0609084 [hep-ph] . —3\n[19]S. R. Coleman and S. L. Glashow, “High-energy tests of Lorent z invariance,”\nPhys.Rev. D59(1999) 116008 ,arXiv:hep-ph/9812418 [hep-ph] . —3\n[20]R. Iengo, J. G. Russo, and M. Serone, “Renormalization group in Lifshitz-type theories,”\nJHEP 0911 (2009) 020 ,arXiv:0906.3477 [hep-th] . —3\n[21]J. Collins, A. Perez, D. Sudarsky, L. Urrutia, and H. Vucetic h, “Lorentz invariance and\nquantum gravity: an additional fine-tuning problem?,” Phys.Rev.Lett. 93(2004) 191301 ,\narXiv:gr-qc/0403053 [gr-qc] . —3,20\n[22]M. Berger and V. Kostelecky, “Supersymmetry and Lorentz viol ation,”\nPhys.Rev. D65(2002) 091701 ,arXiv:hep-th/0112243 [hep-th] . —3,5,12,14,19,20\n[23]P. Jain and J. P. Ralston, “Supersymmetry and the Lorentz fine tuning problem,”\nPhys.Lett. B621 (2005) 213–218 ,arXiv:hep-ph/0502106 [hep-ph] . —3,20\n[24]S. Groot Nibbelink and M. Pospelov, “Lorentz violation in su persymmetric field theories,”\nPhys.Rev.Lett. 94(2005) 081601 ,arXiv:hep-ph/0404271 [hep-ph] . —3,4,15,19\n[25]P. A. Bolokhov, S. G. Nibbelink, and M. Pospelov, “Lorentz vio lating supersymmetric\nquantum electrodynamics,” Phys.Rev. D72(2005) 015013 ,\narXiv:hep-ph/0505029 [hep-ph] . —3\n[26]S. R. Coleman and J. Mandula, “All Possible Symmetries Of The S Matrix,”\nPhys.Rev. 159(1967) 1251–1256 . —4\n[27]R. Haag, J. T. Lopuszanski, and M. Sohnius, “All Possible Gen erators of Supersymmetries\nof the s Matrix,” Nucl.Phys. B88(1975) 257 . —4\n[28]W. Xue, “Non-relativistic Supersymmetry,” arXiv:1008.5102 [hep-th] . —4,6,7,20\n22[29]M. Berger, “Superfield realizations of Lorentz and CPT violat ion,”\nPhys.Rev. D68(2003) 115005 ,arXiv:hep-th/0308036 [hep-th] . —5,14\n[30]J. Wess and B. Zumino, “A Lagrangian Model Invariant Under Sup ergauge\nTransformations,” Phys.Lett. B49(1974) 52 . —9\n[31]M. T. Grisaru, W. Siegel, and M. Rocek, “Improved Methods for Supergraphs,”\nNucl.Phys. B159 (1979) 429 . —9,11\n[32]S. Gates, M. T. Grisaru, M. Rocek, and W. Siegel, “Superspace Or One Thousand and One\nLessons in Supersymmetry,” Front.Phys. 58(1983) 1–548,\narXiv:hep-th/0108200 [hep-th] . —11\n[33]S. P. Martin, “A Supersymmetry primer,” arXiv:hep-ph/9709356 [hep-ph] . —17\n[34]L. Girardello and M. T. Grisaru, “Soft Breaking of Supersymme try,”\nNucl.Phys. B194 (1982) 65 . —17,18\n[35]J. Iliopoulos and B. Zumino, “Broken Supergauge Symmetry and R enormalization,”\nNucl.Phys. B76(1974) 310 . —18\n[36]A. Denner, H. Eck, O. Hahn, and J. Kublbeck, “Compact Feynman rules for Majorana\nfermions,” Phys.Lett. B291 (1992) 278–280 . —19\n[37]D. Colladay and P. McDonald, “Vector Superfields and Lorentz Violation,”\nPhys.Rev. D83(2011) 025021 ,arXiv:1010.1781 [hep-ph] . —19\n[38]D. Colladay and P. McDonald, “Lorentz Violation and Extende d Supersymmetry,”\narXiv:1008.1279 [hep-ph] . —17\n23" }, { "title": "1801.08838v1.Ads_spacetime_in_Lorentz_covariant_gauges.pdf", "content": "arXiv:1801.08838v1 [gr-qc] 26 Jan 2018AdS spacetime in Lorentz covariant gauges\nP. Valtancoli\nDipartimento di Fisica, Polo Scientifico Universit´ a di Fir enze\nand INFN, Sezione di Firenze (Italy)\nVia G. Sansone 1, 50019 Sesto Fiorentino, Italy\nAbstract\nWe show how to generate the AdS spacetime metric in general Lo rentz covariant\ngauges. In particular we propose an iterative method for sol ving the Lorentz gauge.1 Introduction\nRecently there has been a renewed interest in the cosmological con stant of general relativity,\nsince it could be useful as a cheap explanation for the antigravity fo rce which drives the\naccelerated expansion of the universe, better known as dark ene rgy [1]-[2]. It is now widely\naccepted that the cosmological constant is a very small paramete r, analogously to what\nhappens for the neutrino mass, and related to the vacuum energy of the quantum fields [3].\nTheAdSspace-time [4] is usually represented in a Lorentz non-covariant fo rm, but there\nare physical applications like for example the study of cosmological g ravitational waves [5],\nin which it is useful to reformulate it in a Lorentz covariant form. It is the purpose of this\npaper to generate all the Lorentz covariant solutions for AdSspace-time. We start from\nsolving perturbatively the Einstein equations up to the second orde r in Λ in the Lorentz\ngauge, which turns out to be a tedious calculation.\nTo simplify the discussion we are going to present a general 5-dimens ional representation\nof theAdSspace-time in terms of an arbitrary function f(Λx2), where x2=ηµνxµxν. We\nthen show that the Lorentz gauge fixes this arbitrary function wit h a recursive method and\nthe result agrees with the direct perturbative solution of the Einst ein equations.\n2 Perturbative calculation of the Einstein equations\nWe are going to discuss all the aspects of the AdS space-time in the L orentz gauge. We\nstart performing the perturbative calculation directly using the Ein stein equations, and only\nafterwards we will introduce a faster method. We recall the known perturbative solution of\ntheAdSspace-time at the first order in Λ\ngµν=ηµν+h(1)Λ\nµν+O(Λ2)\nh(1)Λ\nµν=−Λ\n9(xµxν+2ηµνx2) (2.1)\nIt is straightward to show that eq. (2.1) satisfies the Lorentz gau ge:\nηµµ′∂µ′h(1)Λ\nµν=1\n2∂ν(ηµµ′h(1)Λ\nµµ′) (2.2)\nThe associated connection is, at this order, given by:\nΓµ(1)Λ\nαβ=−Λ\n9( 2δµ\nαxβ+2δµ\nβxα−ηαβxµ) (2.3)\n1We can easily compute the curvature tensor, limited at the linear ter m in the connection\nRα(1)Λ\nµβν=Λ\n3(ηβνδα\nµ−ηβµδα\nν) (2.4)\nThe corresponding Ricci tensor and the curvature are then\nR(1)Λ\nβν= ΛηβνR(1)Λ=ηβνR(1)Λ\nβν= 4Λ (2.5)\nsolving perturbatively the Einstein equations.\nNow we are going to compute the perturbative metric at the second order in Λ\ngµν=ηµν+h(1)Λ\nµν+h(2)Λ\nµν+O(Λ3) (2.6)\nWe can again impose the Lorentz gauge\nηµµ′∂µ′h(2)Λ\nµν=1\n2∂ν(ηµµ′h(2)Λ\nµµ′) (2.7)\nWe must be careful that the connection is made by several pieces\nΓµ(2)Λ\nαβ= Γµ(2)ΛI\nαβ+ Γµ(2)ΛII\nαβ\nΓµ(2)ΛI\nαβ=1\n2ηµµ′(∂αh(2)Λ\nµ′β+∂βh(2)Λ\nµ′α−∂µ′h(2)Λ\nαβ)\nΓµ(2)ΛII\nαβ=−1\n2hµµ′\n(1)Λ(∂αh(1)Λ\nµ′β+∂βh(1)Λ\nµ′α−∂µ′h(1)Λ\nαβ) (2.8)\nwherehµν\n(1)Λ=ηµµ′ηνν′h(1)Λ\nµ′ν′.\nLet us work out the second term\nΓα(2)ΛII\nβν=−Λ2\n81( 4xαxβxν−3ηβνxαx2+4δα\nβxνx2+4δα\nνxβx2) (2.9)\nIts contribution to the curvature tensor is given by\nRα(2)ΛII\nβµν =∂µΓα(2)ΛII\nβν−(µ↔ν) =\n=−Λ2\n81[ 10ηβµxνxα−10ηβνxµxα+4δα\nνxβxµ\n−4δα\nµxβxν+7δα\nνηβµx2−7δα\nµηβνx2] (2.10)\n2The Ricci tensor\nR(2)ΛII\nβν=δµ\nαRα(2)ΛII\nβµν=Λ2\n81( 2xβxν+31ηβνx2) (2.11)\nand its contribution to the Einstein equation is\nR(2)ΛII\nβν−1\n2R(2)ΛIIηβν=2Λ2\n81(xβxν−16ηβνx2) (2.12)\nThe first term of the connection produces the following curvature tensor Γµ(2)ΛI\nαβ\nR(2)ΛI\nβν=−1\n2ηρρ′∂ρ∂ρ′h(2)Λ\nβν+ gauge terms (2.13)\ngiving rise to the another contribution to the Einstein equations\nR(2)ΛI\nβν−1\n2R(2)ΛIηβν=−1\n2ηρρ′∂ρ∂ρ′/parenleftbigg\nh(2)Λ\nβν−1\n2ηβνh(2)Λ/parenrightbigg\n(2.14)\nTill now we have worked out in the curvature tensor only linear terms in the connections.\nLet us add now the non linear ones:\nRα(2)ΛIII\nβµν = Γα(1)Λ\nµα′Γα′(1)Λ\nβν−(µ↔ν) =\n=Λ2\n81[4δα\nµxνxβ−2δα\nµηβνx2−4δα\nνxµxβ+\n+2δα\nνηβµx2−ηβµxαxν+ηβνxαxµ] (2.15)\nThe associated Ricci tensor is\nR(2)ΛIII\nβν=Λ2\n81( 11xνxβ−5ηβνx2) (2.16)\ngiving rise to the third contibution to the Einstein equations\nR(2)ΛIII\nβν−1\n2R(2)ΛIIIηβν=Λ2\n81( 11xνxβ−1\n2ηβνx2) (2.17)\nWe have not yet finished. We must add some residual extra contribu tions, given by\n1\n2R(2)ΛIV=h(1)ΛβνR(1)Λ\nβν→ −1\n2R(2)ΛIVηβν=−Λ2x2\n2ηβν(2.18)\n3and finally\n−1\n2R(1)Λh(1)Λ\nβν+Λh(1)Λ\nβν=−Λh(1)Λ\nβν (2.19)\nBy collecting all the various ( five ) terms\n−1\n2ηρρ′∂ρ∂ρ′/parenleftbigg\nh(2)Λ\nβν−1\n2ηβνh(2)Λ/parenrightbigg\n+2Λ2\n81(xβxν−16ηβνx2)\n+Λ2\n81( 11xνxβ−1\n2ηβνx2)−Λ2x2\n2ηβν+Λ2\n9(xβxν+2ηβνx2) = 0 (2.20)\nwe arrive at the final equation\nηρρ′∂ρ∂ρ′/parenleftbigg\nh(2)Λ\nβν−1\n2ηβνh(2)Λ/parenrightbigg\n=22\n81(2xβxν−5ηβνx2) (2.21)\nThe compatibility between this equation and the Lorentz gauge is str aightforward since\napplying the operator ∂β( or∂ν) to both sides of this equation we simply get zero.\nFinally we obtain the solution of this equation at the second order in Λ:\nh(2)Λ\nβν=11\n16·81Λ2( 4xβxνx2+5ηβνx4)\nh(2)Λ=ηβνh(2)Λ\nβν=11\n54Λ2x4x4≡(x2)2(2.22)\nWe have just learned that the Lorentz gauge imposes strong cons traints on the general\nsolution. Let us suppose to make the perturbative calculation at th e generic order n:\ngµν=ηµν+h(1)Λ\nµν+....+h(n)Λ\nµν+.... (2.23)\nthe final equation, similar to eq. ( 2.21 ), at the order nis of the type\nηρρ′∂ρ∂ρ′/parenleftbigg\nh(n)Λ\nβν−1\n2ηβνh(n)Λ/parenrightbigg\n= (Anxβxνx2(n−2)+Bnηβνx2(n−1)) (2.24)\nLet us apply the operator ∂β( or∂ν) to both sides of this equation\n∂β(...) = 0 →Bn=−2n+1\n2(n−1)An (2.25)\n4the Lorentz gauge implies the constraint\nηρρ′∂ρ∂ρ′/parenleftbigg\nh(n)Λ\nβν−1\n2ηβνh(n)Λ/parenrightbigg\n=cnΛn[ 2(n−1)xβxνx2(n−2)−(2n+1)ηβνx2(n−1)] (2.26)\nTherefore the solution at the order nis of the type\nh(n)Λ\nβν=cnΛn/bracketleftbiggxβxνx2(n−1)\n2(n+2)+(n+3)ηβνx2n\n4n(n+2)/bracketrightbigg\nh(n)Λ=ηβνh(n)Λ\nβν=3\n2ncnΛnx2nx2n≡(x2)n(2.27)\nUnfortunately the cncoefficients cannot be computed only from the Lorentz gauge and\nmust be determined by solving the Einstein equation.\n3 General solution of the Einstein equations\nTo solve the Einstein equation, we are going to embed the AdSspace-time in the following\n5dspace\nds2=dX2\n0+dX2\n1−dX2\n2−dX2\n3−dX2\n4 (3.1)\nsubject to the constraint\nX2\n0+ηijXiXj=3\nΛi= 1,...,4 (3.2)\nWecanautomaticallygeneratesolutionsoftheEinstein equationssim ply determining the\nmapping X0=X0(xµ),Xi=Xi(xµ). In this choice we must respect the Lorentz covariance\nof the 4dspace-time and we are forced to choose this mapping\nX0=/radicalbigg\n3\nΛ/radicalbigg\n1−Λ\n3x2f2(Λx2)\nXi=xif(Λx2) (3.3)\nwith the 4 dmetric defined as usual by the formula\n5gµν=ηabdXa\ndxµdXb\ndxν(3.4)\nAll these mappings produce solutions of the Einstein equation in a cov ariant gauge and\ndepend on an arbitrary function f(Λx2). First let us study the basic solution with f(Λx2) =\n1:\nX0=/radicalbigg\n3\nΛ/radicalbigg\n1−Λ\n3x2Xi=xi\ng(0)\nµν=ηµν+Λ\n3xµxν\n1−Λ\n3x2(3.5)\nWe can easily verify that it solves the Einstein equations exactly. The inverse metric is given\nby\ngµν(0)=ηµν−Λ\n3xµxν(3.6)\nWe can compute the associated connection\nΓµ(0)\nαβ=Λ\n3xµg(0)\nαβ (3.7)\nand the complete curvature tensor\nRα(0)\nβµν=∂µΓα(0)\nβν+ Γα(0)\nσµΓσ(0)\nβν−(µ↔ν) =\n=Λ\n3(δα\nµg(0)\nβν−δα\nνg(0)\nβµ) (3.8)\nAnalogously the Ricci tensor and the curvature have simple forms\nR(0)\nβν= Λg(0)\nβνR(0)= 4Λ (3.9)\nsolving exactly the Einstein equations. Let us note that this solution ( similarly to those\nwith an arbitrary function f(Λx2) ) is singular for large events of the order\nx2≃1\nΛ(3.10)\nThis singularity is present both for positive and negative cosmologica l constant.\n6It is possibile fixing the arbitrary function f(Λx2) adding into this scheme the Lorentz\ngauge condition\nηµµ′∂µ′hµν=1\n2∂ν(ηµµ′hµµ′) (3.11)\nBy substituting the definition of the metric in terms of the mapping Xa=Xa(xi) we get\nηabηµµ′∂µ∂µ′Xa·∂νXb= 0 (3.12)\nThe variables Xaare constrained leading to the property ηabXa·∂νXb= 0, therefore\nwe must impose that\nηµµ′∂µ∂µ′Xa=λ(Λx2)Xa (3.13)\nwhere we have introduced a second function λ(Λx2). We are going to show that this\nequation (3.13) determines univocally both functions f(Λx2) andλ(Λx2).\nLet us introduce the following power series\nλ(Λx2) =∞/summationdisplay\nn=1λnΛnx2(n−1)\nf(Λx2) = 1 +∞/summationdisplay\nn=1fnΛnx2n(3.14)\nThe coefficients λnare determined by the eq. for X0while the coefficients fnare deter-\nmined by the eq. for Xiin a recursive scheme. We start from the lowest order\nX0≃/radicalbigg\n3\nΛ/parenleftbigg\n1−Λ\n6x2/parenrightbigg\n→λ1=−4\n3(3.15)\nThe coefficient λ1allows to compute the coefficient f1by using the equation for Xi\nηρρ′∂ρ∂ρ′Xi≃ηρρ′∂ρ∂ρ′(xi·f1Λx2) =−4\n3Λxi (3.16)\nfrom which we obtain\nf1=−1\n9→f(Λx2) = 1−Λ\n9x2+O(Λ2x4) (3.17)\n7Going back to the eq. for X0at the next order in Λ, we can use f1to determine λ2as\nfollows\nX0≃/radicalbigg\n3\nΛ/parenleftbigg\n1−Λ\n6x2+5\n9·24Λ2x4+O(Λ3x6)/parenrightbigg\n(3.18)\nfrom which we compute λ2=1\n3and therefore\nλ(Λx2) =−4\n3Λ+1\n3Λ2x2+O(Λ3x4) (3.19)\nRepeating the iteration, the coefficient λ2determines f2at the next order in Λ by using\nthe equation for Xi\nf2=13\n27·32(3.20)\nfrom which we know Xiup to the second order in Λ\nXi≃xi/parenleftbigg\n1−Λ\n9x2+13\n27·32Λ2x4+O(Λ3x6)/parenrightbigg\n(3.21)\nBy substituting these findings in the equation (3.4), defining the met ric in terms of the\nmapping Xa(xi)\ngµν=ηabdXa\ndxµdXb\ndxν=ηµν+h(1)Λ\nµν+h(2)Λ\nµν+.... (3.22)\nwe recover the perturbative solutions (2.1) e (2.22)\nh(1)Λ\nµν=−Λ\n9(xµxν+2ηµνx2)\nh(2)Λ\nµν=11\n16·81Λ2( 4xµxνx2+5ηµνx4) (3.23)\nobtained with a very laborious calculation from the Einstein equations .\n4 Conclusion\nTheAdSspace-time is an example of curved space-time in the absence of mat ter. In this\npaper, we have searched how to describe it with a Lorentz covarian t coordinate system,\n8which is better suited for discussing the propagation of gravitation al waves in such a curved\nspace.\nWe first built the perturbative solution of the Einstein equations with a cosmological\nconstant Λ intheLorentz gauge, andthenintroduced amoresyst ematic methodtoconstruct\nthe solution by embedding the AdSspace-time into a 5 dspace with a quadratic constraint\nbetween the coordinates.\nWe have succeeded to identify the right mapping Xa=Xa(xi) which allows us to reach\nthe Lorentz gauge and constructed an iterative method to calcula te theAdSsolution to the\nvarious perturbative orders in Λ. This method is the easiest way to s olve the Lorentz gauge\ncondition if compared to the direct perturbative calculation of the E instein equations.\nOnce the structure of the AdSspace-time has been accurately identified in a Lorentz\ncovariant coordinate system, we believe that it will be easier to stud y the propagation of a\ngravitational wave in the presence of such a background.\nReferences\n[1] Planck collaboration, Astronomy and Astrophysics 594(2016) A13,\narXiv:1502.01589v3.\n[2] Frieman J. A., Turner M. S.and Huterer D., Annal Review of Astron omy and Astro-\nphyics46(2008) 385, arXiv:0803.0982.\n[3] Weinberg S., Phys. Rev. Lett. 59(1987) 2607.\n[4] Ellis G. F. R., Hawking S. W. (1973) ”The large scale structure of sp ace-time”, Cam-\nbridge University Press pp. 124-134.\n[5] Bernabeu J., Espriu D. and Puigdomenech D., Phys. Rev. D 84(2012) 063523,\narXiv:1106.4511 [hep-th].\n9" }, { "title": "1502.00431v1.Evolution_of_Higgs_mode_in_a_Fermion_Superfluid_with_Tunable_Interactions.pdf", "content": "arXiv:1502.00431v1 [cond-mat.quant-gas] 2 Feb 2015Evolution of Higgs mode in a Fermion Superfluid with Tunable I nteractions\nBoyang Liu,1Hui Zhai,1and Shizhong Zhang2\n1Institute for Advanced Study, Tsinghua University, Beijin g, 100084, China\n2Department of Physics and Center of Theoretical and Computa tional Physics,\nThe University of Hong Kong, Hong Kong, China\n(Dated: October 4, 2018)\nIn this letter we present a coherent picture for the evolutio n of Higgs mode in both neutral\nand charged s-wave fermion superfluids, as the strength of attractive int eraction between fermions\nincreases from the BCS to the BEC regime. In the case of neutra l fermionic superfluid, such as\nultracold fermions, the Higgs mode is pushed to higher energ y while at the same time, gradually\nloses its spectral weight as interaction strength increase s toward the BEC regime, because the\nsystem is further tuned away from Lorentz invariance. On the other hand, when damping is taken\ninto account, Higgs mode is significantly broadened due to co upling to phase mode in the whole\nBEC-BCS crossover. Inthechargedcase ofelectron supercon ductor, theAnderson-Higgsmechanism\ngaps out the phase mode and suppresses the coupling between t he Higgs and the phase modes, and\nconsequently, stabilizes the Higgs mode.\nThe experimental search for Higgs boson in particle\nphysics has made remarkable progresses [1, 2]. On the\nother hand, Higgs mode has also generated considerable\ninterest in condensed matter and cold atom systems.\nEarly in 1980s’, Raman scattering experiment has re-\nvealed an unexpected peak in a superconducting charge\ndensity wave compound NbSe 2[3], which was later at-\ntributed to the Higgs mode [4, 5]. Signal of Higgs mode\nhas also been observed in antiferromagnet TlCuCl 3by\nthe neutron scattering [6], and recently in superconduct-\ning NbN sample by terahertz pump probe spectroscopy\nin a nonadiabatic excitation regime [7, 8]. In cold atom\nsystem, Higgs mode has been observed near the super-\nfluid to Mott insulator phase transition of bosonic atoms\nin optical lattices at integer filling [9, 10].\nTheoretically, the simplest field theory where Higgs\nmode emerges is a relativistic U(1) field theory with\nLorentz invariance in the symmetry broken phase. This\noccurs, for example, in the weak coupling BCS super-\nconductor [11, 12] or in the Mott-superfluid transition of\nBose-Hubbard model at integer filling [13, 14]. However,\nin most condensed matter systems, Lorentz invariance\nonly emerges with fine tuning and the generic symmetry\nis usually Galilean [15]. Thus, it is an interesting ques-\ntion to investigate how the Higgs mode evolves as the\nsystem is tuned away from the Lorentz invariance point.\nMoreover, in condensed matter systems, further compli-\ncations often occur because the Higgs mode is usually\ncoupled to other elementary excitations which leads to\nits damping [16–19]. In this Letter, we investigate these\nissues in the context of the BEC-BCS crossover model.\nIn the BCS limit, the system obeys approximate Lorentz\nsymmetry due to particle-hole symmetry and is expected\nto host Higgs mode. In the BEC limit, it is a condensate\nof molecular bosons and obeys the Galilean invariance.\nIt thus provides a unique system to describe the fate of\nHiggs mode as the system is tuned away from Lorentz\ninvariant limit. In addition, due to tunable interactions/Minus4/Minus3/Minus2/Minus10120100200300400500600700\nΖ/Equal1/Slash1kFas/Minus4/Minus3/Minus2/Minus1000.050.1\nΖ/Equal1/Slash1kFasv''/CapDelta0/Slash1u''\nu''/Slash1u'v'/CapDelta0/Slash1u'\nFIG. 1: (Color online) v′∆0/u′andu′′/u′as functions of the\nscattering length ζ= 1/kFas. In the inset we show v′′∆0/u′′\nas a function of ζ.\nin the BEC-BCS crossover, it also provides a great plat-\nform to investigate the interaction effects on the Higgs\nmode due to coupling to collective and quasi-particle ex-\ncitations [20].\nWe investigate these questions based on the time-\ndependent Ginzburg-Landau formulation of the BEC-\nBCS crossover,\nS=/integraldisplay\ndtd3x/bracketleftbig\nφ∗(−iu∂t+v∂2\nt−∇2\n2m∗−r)φ+b\n2|φ|4/bracketrightbig\n,(1)\nwhereφis the Ginzburg-Landau order parameter. The\nvarious parameters u,v,r,b andm∗can be computed\nalongBEC-BCScrossoverintermsofthechemicalpoten-\ntialµ, temperature Tandζ= 1/(kFaS), whereaSis the\ns-wave scattering length. Within the Nozi` eres-Schmit-\nRink [21] framework, this can be calculated as detailed\nin the supplementary material [22]. The coefficients of\nthe time derivative terms u=u′+iu′′andv=v′+iv′′\nare complex in general. The real parts u′andv′describe2\nthe propagating behavior of the cooper pair field, while\nthe imaginary parts u′′andv′′describe its damping due\nto coupling to the fermionic quasi-particles. A plot of\nvarious parameters are given in Fig.1. We note the fol-\nlowing features.\n(i) Consider the real parts u′andv′in the BEC-BCS\ncrossover. In the BCS limit, u′/v′∆0→0 because of\nthe approximate particle-hole symmetry in the weak-\ncoupling BCS theory while ∆ 0=/radicalbig\nr/bis the mean field\nvalue of order parameter. As a result, the system ac-\nquires an emergent Lorentz invariance, and one expects\nthe emergenceofHiggs mode, togetherwith the standard\nAnderson-Bogoliubov mode for neutral fermion super-\nfluid. In the BEC limit, however, v′∆0/u′∼∆0/|µ| ≪1,\nand we can neglect the v′-term. This leads to a Galilean\ninvariantneutralbosontheory,forwhichonlyBogoliubov\nmode exists.\n(ii) The damping terms ( u′′) becomes important as\none moves to the BCS side, because of the decreasing\nfermionic excitation gap and as a result, a stronger cou-\npling of the pairing field to the quasi-particle excitations.\nThis corresponds to finite lifetime of Cooper pairs at fi-\nnite temperature. We will show that the damping u′′-\nterm itself will generate considerable effect for the ap-\npearance of the Higgs mode different from that in a pure\nLorentz invariance theory. In the BEC limit, the imagi-\nnary parts vanishes within NSR. On the other hand, we\nfind that whenever they are nonzero, v′′∆0/u′′≪1 for\nthe entire crossover regime and we shall thus neglect v′′-\nterm altogether in the following discussion.\nSpectral Weight Transfer without Damping. To inves-\ntigate the evolution of Higgs mode as the system is tuned\ngradually from its Lorentz-invariant BCS limit towards\nthe Galilean invariant BEC limit, we shall first neglect\nthedampingtermsinEq.4andstudythetransferofspec-\ntral weight between the Higgs and Goldstone modes. In\nthe symmetry broken state, we can write the order pa-\nrameterφ= ∆0+δa+iδp, whereδaandδpdescribe\namplitude and phase fluctuations, respectively. In terms\nofδaandδpand withu′′=v′′= 0, we can write the\naction Eq. 4 in the Fourier space as\nS=/integraldisplaydω\n2πd3k\n(2π)3¯Φ(−ω,−k)G−1Φ(ω,k),(2)\nwith¯Φ(ω,k) = (δa(ω,k),δp(ω,k)) and the kernel Gis\ngiven by\nG−1=/parenleftbigg−v′ω2+ξk+2r iu′ω\n−iu′ω−v′ω2+ξk/parenrightbigg\n,(3)\nwithk=|k|andξk=k2/2m∗. Two branches of spec-\ntrum can be identified, with mode frequencies given by\nω2\n±=ξk+r\nv′+u′2\n2v′2±/radicalbigg\nr2\nv′2+u′4\n4v′4+u′2\nv′3(ξk+r).(4)\nIn the BCS limit, v′∆0≫u′and solutions can be writ-\nten asω−(k) =k/√\n2m∗v′andω+(k) =/radicalbig\n(ξk+2r)/v′;\nFIG. 2: (Color online) Spectral function Aaa(k,ω) in the ab-\nsence of damping term. Aaa(k,ω) as a function of k(in unit\nof 1/ξ) andω(in unit of ∆ 0) for three different interaction\nstrength ζ=−1/(kFas),ζ=−7 for (a), ζ=−3 for (b)\nandζ=−1 for (c), corresponding to different gaps ∆ 0/EF=\n10−5, ∆0/EF= 4×10−3and ∆ 0/EF= 7×10−2, respectively.\n(a2-c2): Aaa(k,ω) as a function of ωfork= 0.1/ξ(purple\ndashed line) and k= 0.01/ξ(blue solid line). T/Tc= 0.9 and\nδis taken as 10−4∆0.\nthe first being the Goldstone mode with linear dis-\npersion and the second Higgs mode, with Higgs gap\nω+(0) =/radicalbig\n2r/v′=/radicalbig\n2b∆2\n0/v′. Using the facts that\nv′= 7β2ζ(3)ν0/16π2andb= 7β2ζ(3)ν0/8π2in the BCS\nlimit, one finds ω+(0) = 2∆ 0, as expected for a Lorentz-\ninvariant theory. Here β= 1/kBTis the inverse temper-\nature,ζ(n) is the Riemann-Zeta function and ν0is the\ndensity of state at the Fermi energy ǫF.\nIn the BEC limit, u′≫v′∆0, we find ω−(k) =/radicalbig\nξk(ξk+2r)/u′is the Bogoliubov mode while the other\nmodeω+(k) =/radicalbig\n2ξk/v′+2r/v′+(u′/v′)2has a gap\n∼ |µ|, of order of binding energy of the molecule in the\nBEC limit. The existence of the gapped mode is a reflec-\ntion of the fact that our bosonic field φis a composite of\ntwo fermions and disappears in the infinite binding limit\nwhere only Bogoliubov mode exists as it should.\nIn between these two limits, Lorentz invariance is bro-\nken and the coupling between the amplitude and phase\ndegrees of freedom becomes stronger, as characterized\nby the off-diagonal term iu′ω. We note that for low en-\nergy Bogoliubov excitations, such coupling is small, but\nfor gapped Higgs mode, it provides significant mixing of\nthe amplitude and phase. To characterize such mixing,\nwe calculate the spectral function for the amplitude δa,\ngiven byAaa(k,ω) =−1\nπImGaa(k,ω+iδ). Explicitly,\nthis can be written as\nAaa(k,ω) =A+(k)δ(ω−ω+(k))+A−(k)δ(ω−ω−(k))\n(5)\nwithA+(k) andA−(k) being the spectral weight den-3\nsities associated with two modes ω−andω+[22]. In\nFig.2 (a,b,c), we plot the spectral function Aaa(k,ω) for\nthree representative values of ζ(corresponding to differ-\nent ∆0/EF). Two features can be noticed immediately.\nFirst, the Higgs gap increases beyond 2∆ 0of the BCS\nlimit as interaction strength increases. Secondly, there is\nincreasingspectralweighttransferfromthegappedHiggs\nmode to the gapless mode. One can show explicitly that\nA+/A−= 4v′r2(u′2/radicalbig\nk2/2m∗(k2/2m∗+2r))−1, which\nindicates the gradual increasing of the mixing between\nphase and amplitude degrees of freedom.\nIncluding Damping Term. Due to the presence of\ndamping term, the time-dependent Ginzburg-Landau\ntheory is not a pure Lorentzinvariant U(1) theory. Thus,\natanyfinitetemperature, evenintheBCSlimit, thepeak\nof Higgs excitation ω+(k) will not as sharp as discussed\nabove. To calculate the equilibrium spectral weight in\nthe presence of damping, we need to introduce the so-\ncalled Langevin force η(t,x), which satisfies the follow-\ning conditions, /angb∇acketleftη(t′,x′)η(t,x)/angb∇acket∇ight=/angb∇acketleftη∗(t′,x′)η∗(t,x)/angb∇acket∇ight= 0,\nand/angb∇acketleftη∗(t′,x′)η(t,x)/angb∇acket∇ight= 2u′′kBTδ(t−t′)δ(x−x′). In-\ncluding the corresponding term in the action as SL=/integraltextdtd3x(φ∗η+φη∗), we obtain the equations of mo-\ntion forδaandδp, by setting ∂(S+SL)/∂δa= 0 and\n∂(S+SL)/∂δp= 0,\n/parenleftbig\n−vω2+ξk+2r/parenrightbig\nδa−iuωδp+η′= 0,(6)\n/parenleftbig\n−vω2+ξk/parenrightbig\nδp+iuωδa+η′′= 0,(7)\nwhereη′andη′′are the real and imaginary parts of the\nLangevin force η, respectively. The spectral functions for\nthe amplitude fluctuation is given by, using fluctuation\ndissipation theorem,\nAaa=u′′ω\n2|−vω2+ξk|2+|uω|2\n|−(uω)2+(−vω2+ξk)(−vω2+ξk+2r)|2.\n(8)\nBy comparing Fig. 3 with Fig. 2, one can see three\nimportantfeaturesbroughtaboutbyincludingthedamp-\ning term. First, the spectral weight transfer is enhanced.\nFor instance, for ζ=−7, there is almost no spectral\nweight transfer in the absence of damping (Fig. 2(a))\nwhile in the presence ofdamping, for verysmall k≪1/ξ,\nAaa(k,ω) exhibits a clear peak at the energy of Bogoli-\nubov mode, with a weight proportional to u′′[22]. Sim-\nilar enhancement of spectral weight transfer can also be\neasily seen in Fig. 3(b) for ζ=−3. Secondly, also for\nk≪1/ξ, in the BCS limit, the location of Higgs peak is\nsubstantially reduced from/radicalbig\n2r/v′to/radicalbig\n2r/v′−u′′2/v′2,\nas shown for ζ=−7 and−3 in Fig. 3(a) and (b), re-\nspectively [22]. Thirdly, as kstarts to derivate from zero,\nthe Higgs mode quickly loses its identity, due to strong\nhybridization with the Bogoliubov mode. For instance,\neven fork= 0.1/ξ, as displayedby the purple dashed line\nin Fig. 3(a2-c2), no feature of sharp peak is observed in\nFIG. 3: (Color online) Spectral function Aaa(k,ω)in presence\nof damping term. Aaa(k,ω) as a function of k(in unit of 1 /ξ)\nandω(in unit of ∆ 0) for three different interaction strength\nζ=−1/(kFas),ζ=−7for (a), ζ=−3for (b) and ζ=−1for\n(c), corresponding to different ∆ 0/EF= 10−5, ∆0/EF= 4×\n10−3and ∆ 0/EF= 7×10−2, respectively. (a2-c2): Aaa(k,ω)\nas a function of ωfork= 0.1/ξ(purple dashed line) and\nk= 0.01/ξ(blue solid line). T/Tc= 0.2 andδis taken as\n10−4∆0.\nAaa(k,ω). And for ζ=−1, no sharp peak exists even\nfork= 0.01/ξ.\nEffects of Coupling to External Gauge Fields . Now we\nunderstand that, in the weakly interacting BCS side of a\nneutral superfluid, the appearance of Higgs mode suffers\nsignificant broadeningdue to finite u′′-term at finite tem-\nperature, which couples the Higgs mode to the collective\nBogoliubov excitations. Therefore, if we further consider\nthe presence of coupling to external electromagnetic field\nfor the case of charged fermions, the Bogoliubov mode\nis gapped out by the Anderson-Higgs mechanism. Thus,\nwe expect that the Higgs mode is easier to observe in\nthe charged case. To incorporate this effect of external\nelectromagnetic field, we introduce the gauge potential\nϕ(t,x) and extend the action as\nSc=/integraldisplay\ndtd3x/braceleftbig\nφ∗[−iu(∂t−2eϕ)+v(∂t−2eϕ)2(9)\n−∇2\n2m∗−r]φ+b\n2|φ|4−1\n8πϕ∇2ϕ/bracerightbig\n,\nwhereeis the charge of the electron. Following the same\nprocedure as before, we find that the coupling between\nδaandδpis modified and is now proportional to k2\ni2uωk2\nk2+32πve2∆2\n0δa(ω,k)δp(−ω,−k).(10)\nAs a result, the original gapless phase mode is gapped to\na finite frequency, ω(k) =/radicalbig\nξk/v′+16πe2∆2\n0/m∗, which\nis known as the Anderson-Higgs mechanism. Thus, the4\nFIG. 4: (Color online) Spectral function Aaa(k,ω) for the\ncharged case. Aaa(k,ω) as a function of k(in unit of 1 /ξ)\nandω(in unit of ∆ 0) for three different interaction strength\nζ=−1/(kFas),ζ=−7 for (a), ζ=−3 for (b) and ζ=−1for\n(c), corresponding to different ∆ 0/EF= 10−5, ∆0/EF= 4×\n10−3and ∆ 0/EF= 7×10−2, respectively. (a2-c2): Aaa(k,ω)\nas a function of ωfork= 0.1/ξ(purple dashed line) and\nk= 0.01/ξ(blue solid line). T/Tc= 0.9 andδis taken as\n10−4∆0.\nlarge energy separation between this gapped phase mode\nand Higgs mode strongly suppresses their coupling. A\nfurther consequence of the modification is that at long\nwave length k→0, the coupling between phase and\namplitude mode becomes small. The spectral function\nAaa(k,ω) for charged case is plotted in Fig. 4 [22]. In\nsharp contrast to neutral case Fig. 3, the presence of\ndamping term has almost no effect on Higgs mode, and\nthere is always a peak located at ω= 2∆0. In this case,\nas attractive interaction increases and the system gradu-\nally loses its Lorentz invariance, the peak becomes more\nand more broad.\nConclusion. Insummary, wehaveinvestigatedtheevo-\nlution of Higgs mode in the BEC-BCS crossover for both\nneutral and charged Fermi superfluid. Our main con-\nclusions include: i) Towards the BEC side, as the sys-\ntem gradually loses the Lorentz invariance, the Higgs\nmode is pushed to very high energy and the spectral\nweight is transferred to Bogoliubov mode. ii) In the BCS\nside, damping terms arises in the Ginzburg-Landau the-\nory, due to coupling between Cooper pair field and the\nfermionic quasi-particles, and strongly couples the Higgs\nmodetothegaplessphasemodein theneutralsuperfluid,\nwhich enhances the spectral weight transfer and washes\nout features of Higgs mode at finite momentum. (iii)\nFor the charged case, the phase mode is gapped out by\ncoupling to external electromagnetic field, and the Higgs\nmode becomes much more stable.\nOur results also deepen our understandings of Higgs\nmode in superconductor. The physical picture behindthe observation of Higgs mode in a BCS superconduc-\ntor is much more subtle and its observability is not\nmerely guaranteed by Lorentz symmetry. While the\ndamping terms broadens the Higgs peak, the Anderson-\nHiggs mechanism alleviate the coupling between Higgs\nand phase mode and as a result, Higgs mode remains\nat energy 2∆ 0. As for cold atom system, because of the\ncooling limit, so far we can not reachFermi superfluid for\nζ <−1. However, our results show no Higgs feature in\nspectralfunction for ζ >−1. On the otherhand, with re-\ncentdevelopmentofsyntheticgaugefield, therearemany\nproposals to generate a synthetic dynamic gauge field in\ncold atom system [29]. If such a dynamic gauge field can\nbe experimentally realized and coupled to fermions, the\nAnderson-Higgsmechanism will be activated and a Higgs\nmode will be observed. This can be used as a way to test\nour theory.\nAcknowledgements. BY and HZ are supported by\nTsinghua University Initiative Scientific Research Pro-\ngram, NSFC Grant No. 11174176, No. 11325418 and\nNKBRSFC under Grant No. 2011CB921500. SZ is sup-\nportedbyastart-upgrantfromUniversityofHongKong,\nthe Collaborative Research Fund HKUST3/CRF/13G\nandrgc-grf 17306414. HZwouldliketothankHongKong\nUniversity for hospitality where this work was initiated.\n[1] CMS collaboration, Phys. Lett. B 716, 30 (2012).\n[2] ATLAS collaboration, Phys. Lett. B 716, 1 (2012).\n[3] R. Sooryakumar, and M. V. Klein, Phys. Rev. Lett. 45,\n660 (1980).\n[4] P.B. Littlewood and C.M. Varma, Phys. Rev. B 26, 4883\n(1982).\n[5] P.B. Littlewood and C.M. Varma, Phys. Rev. Lett. 47,\n811 (1981).\n[6] Ch. R¨ uegg, B. Normand, M. Matsumoto, A. Furrer, D.F.\nMcMorrow, K.W. Kramer, H.U. Gudel, S.N. Gvasaliya,\nH. Mutka, and M. Boehm, Phys. Rev. Lett. 100, 205701\n(2008).\n[7] R. Matsunaga, Y. I. Hamada, K. Makise, Y. Uzawa, H.\nTerai, Z. Wang, and R. Shimano, Phys. Rev. Lett. 111,\n057002 (2013).\n[8] R. Matsunaga, N. Tsuji, H. Fujita, A. Sugioka, K. Makise,\nY. Uzawa, H. Terai, Z. Wang, H. Aoki and R. Shimano,\nScience345, 1145 (2014).\n[9] U. Bissbort, S. G¨ otze, Y. Li, J. Heinze, J. S. Krauser, M.\nWeinberg, C. Becker, K. Sengstock, and W. Hofstetter,\nPhys. Rev. Lett. 106, 205303 (2011) .\n[10] M. Endres, T. Fukuhara, D. Pekker, M. Cheneau, P.\nSchaub, C.Gross, E. Demler, S.KuhrandI.Bloch, Nature\n487, 454-458 (2012).\n[11] C. M. Varma, J. Low Temp. Phys. 126, 901 (2002).\n[12] Y. Barlas, and C. M. Varma, Phys. Rev. B 87, 054503\n(2013).\n[13] S.D.Huber, B.Theiler, E.Altman, andG.Blatter, Phys.\nRev. Lett. 100, 050404 (2008).\n[14] L. Pollet, andN. Prokof’ev, Phys. Rev.Lett. 109, 010401\n(2012).1\n[15] D. Pekker, and C. M. Varma, arXiv:1406.2968.\n[16] D. Podolsky, A. Auerbach, and D. P. Arovas, Phys. Rev.\nB84, 174522 (2011).\n[17] D. Podolsky, and S. Sachdev, Phys. Rev. B 86, 054508\n(2012).\n[18] S. Gazit, D. Podolsky, and A. Auerbach, Phys. Rev. Lett.\n110, 140401 (2013).\n[19] A. Ran¸ con, and N. Dupuis, Phys. Rev. B 89, 180501(R)\n(2014).\n[20] G. M. Bruun, Phys. Rev. A 90, 023621 (2014).\n[21] P. Nozi` eres, and S. Schmitt-Rink, J. Low Temp. Phys.\n59, 195 (1985).\n[22] See supplementary materials for a brief derivation of t he\ntime-dependent Ginzburg-Landau equation and the ex-\npression for various parameters in it. We also discuss de-tails of spectral weight for various cases.\n[23] S. Sachdev, Quantum Phase Transition , Chapter 9, 2ed\nedition, Cambridge University Press, 2011.\n[24] A. Larkin, and A. Varlamov, Theory of Fluctuations in\nSuperconductors , Oxford University Press, 2005.\n[25] A. Altland, and B. Simons, Condensed Matter Field The-\nory, Chapter 10, 2ed edition, Cambridge University Press,\n2010.\n[26] P. W. Anderson, Phys. Rev. 130, 439 (1963).\n[27] P. W. Higgs, Phys. Lett. 12, 132 (1964).\n[28] R. G. Scott, F. Dalfovo, L. P. Pitaevskii, and S. Stringa ri,\nPhys. Rev. A 86, 053604 (2012).\n[29] N. Goldman, G. Juzeli¯ unas, P. ¨Ohberg and I B Spielman,\nReports on Progress in Physics, 77126401 (2014)\nSUPPLEMENTARY MATERIALS\nTime-dependent Ginzburg-Landau theory of BEC-BCS crossov er\nA time-dependent Ginzburg-Landau theory can be constructed f or the entire BEC-BCS crossover in the vicinity of\nTc[1]. The partition function takes the form Z=/integraltext\nD[¯ψσ,ψσ]e−S[¯ψσ,ψσ], with\nS[¯ψσ,ψσ] =/integraldisplay\ndτd3x/braceleftBig\n¯ψσ(∂τ−∇2\n2m−µ)ψσ−g¯ψ↑¯ψ↓ψ↓ψ↑/bracerightBig\n, (1)\nwhereψσare Grassman fields and gis the contact interaction between fermions of opposite spins. µis the chemical\npotential which is determined by requiring the number density to be e qual ton. To investigate the fluctuation effects\nin the Cooper channel, we use a Hubbard-Stratonovich transform ation to decouple the interaction term in the Cooper\nchannel and then integrating out the fermions. We obtain an effect ive theory for the bosonic field ∆( τ,x), which\nrepresents the cooper pair field. Straightforward calculations yie ld the partition function in terms of field ∆ as\nZ=/integraldisplay\nD(¯∆,∆)exp/bracketleftBig\n−1\ng/integraldisplay\ndτdx|∆|2+lndet ˆG−1/bracketrightBig\n, (2)\nwhere\nˆG−1=/parenleftBigg\n−∂τ+∇2\n2m+µ ∆\n¯∆ −∂τ−∇2\n2m−µ/parenrightBigg\n(3)\nis the Gor’kov Green function.\nIn the vicinity of the phase transition the gap parameter ∆ is small an d an expansion in terms of ∆ becomes\npossible. Including both the spatial and time derivatives (after Wick rotation) and retaining the parameter ∆ up to\nthe forth order we obtain an effective action as\nS[¯∆,∆] =/integraldisplay\ndtd3x/braceleftBig\n¯∆/bracketleftbig\n−iu∂t+v∂2\nt−∇2\n2m∗−r/bracketrightbig\n∆+b\n2¯∆¯∆∆∆/bracerightBig\n, (4)\nwhereu=u′+iu′′andv=v′+iv′′are complex in general and all the parameters can be expressed in t erms of2\nmicroscopic parameters as\nu′=(2m)3/2\n16π2/bracketleftBigg\n2√\n2β/radicalbig\n|µ|\nπ∞/summationdisplay\nn=0/radicalbigg/radicalBig\n1+(2n+1)2(π\nβµ)2−sgn(µ)\n(2n+1)2−πβ\n2/radicalbig\n|µ|θ(−µ)/bracketrightBigg\n, (5)\nu′′=m3/2\n8√\n2πβ/radicalbig\n|µ|Θ(µ), (6)\nv′=(2m)3/2\n32π2/bracketleftBigg\n2√\n2β2/radicalbig\n|µ|\nπ2∞/summationdisplay\nn=0/radicalbigg/radicalBig\n1+(2n+1)2(π\nβµ)2+sgn(µ)\n(2n+1)3−πβ\n4/radicalbig\n|µ|θ(−µ)/bracketrightBigg\n, (7)\nv′′=−m3/2\n32√\n2πβ/radicalbig\n|µ|Θ(µ), (8)\n1\n2m∗=1\n2m/integraldisplayd3k\n(2π)3/braceleftBigg\n1−2N(ξk)\n8ξ2\nk+∂N(ξk)\n∂ξk\n4ξk+∂2N(ξk)\n∂ξ2\nk·k2\n2m\n6ξk/bracerightBigg\n, (9)\nr=m\n4πa+/integraldisplayd3k\n(2π)3/braceleftBigg\n1−2N(ξk)\n2ξk−1\n2ǫk/bracerightBigg\n, (10)\nb=/integraldisplayd3k\n(2π)3/braceleftBigg\n1−2N(ξk)\n4ξ3\nk+βN(ξk)[N(ξk)−1]\n2ξ2\nk/bracerightBigg\n. (11)\nIn the above equations, N(ξk) = 1/(exp(βξk)+1) is the Fermi distribution function and ξk=ǫk−µwithǫk=k2/2m.\nFunction Θ(2 µ) is the heaviside step function. Explicitly, the parameter bis the result of one-loop calculation with\nfour fermion propagators\nb=−1\nβ2/summationdisplay\nωn/integraldisplayd3k\n(2π)31\n(−iωn+k2/2m−µ)21\n(iωn+k2/2m−µ)2. (12)\nThe other parameters u,v,1\n2m∗andrare all derived from the inverse vertex function Γ−1(ωn,k), which after the\nstandard renormalization by replacing gwith the two-body scattering length as, is given by\nΓ−1(ωn,k) =−m\n4πas−/integraldisplayd3k\n(2π)3/braceleftbigg1−N(ǫk−µ)−N(ǫk−q−µ)\n−iωn+ǫk+ǫk−q−2µ−1\n2ǫk/bracerightbigg\n. (13)\nTo derive the time-dependent Ginzburg-Landau equation, we first analytically continue vertex function to real fre-\nquencyiωn→ω+i0+. This procedure generates a time-dependent term with paramete ruandv. The detailed\nderivation is as following.\nThe frequency dependent part of Γ−1(ω,k) is\nΓ−1(ω,0)−Γ−1(0,0) =−m\n4πa−/integraldisplayd3k\n(2π)3/braceleftBigg\n1−2N(ǫk−µ)\n−ω−iη+2ǫk−2µ−1\n2ǫk/bracerightBigg\n−Γ−1(0,0). (14)\nThen we expand it in series of small ωas\nΓ−1(ω,0)−Γ−1(0,0)≃ −ω·/integraldisplayd3k\n(2π)31−2N(ǫk−µ)\n(2ǫk−2µ−iη)2−ω2·/integraldisplayd3k\n(2π)31−2N(ǫk−µ)\n(2ǫk−2µ−iη)3. (15)\nWe define the parameters as u≡/integraltextd3k\n(2π)31−2N(ǫk−µ)\n(2ǫk−2µ−iη)2andv≡/integraltextd3k\n(2π)31−2N(ǫk−µ)\n(2ǫk−2µ−iη)3. They both can be calculated by\ncontour integration.\nu≡/integraldisplayd3k\n(2π)31−2N(ǫk−µ)\n(2ǫk−2µ−iη)2\n=ν(ǫF)\n4√ǫF/integraldisplay∞\n0dǫ√ǫ1−2N(ǫ−µ)\n(ǫ−µ−iη)23\n=ν(ǫF)\n4√ǫF·1\n2/integraldisplay\ncdz√z1−2N(z−µ)\n(z−µ−iη)2, (16)\nwherecdenotes the contour in Fig. 1. There are infinite first-order poles zn=µ+(2n+1)πi\nβand one second order\npolezη=µ+ηi. The contour integration can be evaluated in the summation of the r esiduals as\nFIG. 1: The contour “c” in the calculation of Eq. (16). The dot s “·” denote the first-order poles zn=µ+(2n+1)πi\nβand the\ncross “×” denotes the second order pole zη=µ+ηi.\n/integraldisplay\ncdz√z1−2N(z−µ)\n(z−µ−iη)2\n= 2πi/bracketleftBigg\nlim\nz→zn2√z/β\n(z−µ−ηi)2+ lim\nz→zηd\ndz/bracketleftBig√z/parenleftbig\n1−2\nexp(β(z−µ))+1/parenrightbig/bracketrightBig/bracketrightBigg\n= 2πi/bracketleftBigg\n−2β√µ\nπ2+∞/summationdisplay\nn=−∞/radicalbig\n1+(2n+1)πi/βµ\n(2n+1)2+√µβ\n2/bracketrightBigg\n. (17)\nCalculation shows that/summationtext+∞\nn=−∞√\n1+(2n+1)πi/βµ\n(2n+1)2is pure imaginary due to the symmetry of the znpole\nlocations with respect to the horizontal axes. Then it can be writte n as/summationtext+∞\nn=−∞√\n1+(2n+1)πi/βµ\n(2n+1)2=\n√\n2i/summationtext+∞\nn=0/radicalBig√\n1+((2n+1)π/βµ)2−sgn(µ)\n(2n+1)2.Hence, the parameter uis calculated as\nu=(2m)3/2\n16π2/bracketleftBigg\n2√\n2β/radicalbig\n|µ|\nπ∞/summationdisplay\nn=0/radicalbigg/radicalBig\n1+(2n+1)2(π\nβµ)2−sgn(µ)\n(2n+1)2−πβ\n2/radicalbig\n|µ|θ(−µ)+iπβ\n2/radicalbig\n|µ|θ(µ)/bracketrightBigg\n.(18)\nIn the same manner, the parameter vcan also be calculated as shown in Eq. (6) and (7). We should note tha t while\nthe expressions for uand others look different from the standard expression, as given in ref. [1], they in fact reduce\nto the same expressions. We found that this form is more convenien t to use the above expression when dealing with\nhigher order time-derivative terms.\nIn the BCS and BEC limits all the parameters can be analytically derived as shown in Table I.\nSpectral weight function in the case without damping term\nIf we ignore the damping term by taking u′′= 0 the action can be written as\nS=/integraldisplaydω\n2πd3k\n(2π)3¯Φ(−ω,−k)G−1Φ(ω,k), (19)4\nParameters BCS limit BEC limit\nu′0πν(ǫF)\n8√\nǫF|µ|\nu′′ν(ǫF)·π\n8kBT0\nv′ 7ν(ǫF)\n16π2(kBT)2·ζ(3)πν(ǫF)\n64√ǫF|µ|3/2\nv′′−ν(ǫF)·π\n32kBTǫF0\n1\n2m∗1\n2m·7ν(ǫF)ǫF\n12π2(kBT)2ζ(3)1\n2m·πν(ǫF)\n16√\nǫF|µ|\nr ν(ǫF)lnTc\nTπν(ǫF)\n2√\n2√ǫF·(1√mas−/radicalbig\n2|µ|)\nb7ν(ǫF)\n8π2(kBT)2·ζ(3)πν(ǫF)\n32√ǫF|µ|3/2\nTABLE I: Asymptotic behaviors of the parameters in the time- dependent Ginzburg-Landau theory in the BCS and BEC limits.\nwith¯Φ(ω,k) = (δa(ω,k),δp(ω,k)) and the kernel Gis given by\nG−1=/parenleftbigg\n−v′ω2+ξk+2r iu′ω\n−iu′ω−v′ω2+ξk/parenrightbigg\n, (20)\nwithk=|k|andξk=k2/2m∗. Then the amplitude-amplitude correlation function can be easily calc ulated as\nGaa(ω,k) =−v′ω2+ξk\n−u′2ω2+(−v′ω2+ξk)(−v′ω2+ξk+2r). (21)\nStraight forward calculation yields the spectral function as\nAaa(ω,k) =−1\nπImGaa(ω+iδ,k)\n=A+(k)δ(ω−ω+(k))+A−(k)δ(ω−ω−(k)), (22)\nwhere the mode frequencies are given as\nω2\n±=ξk+r\nv′+u′2\n2v′2±/radicalbigg\nr2\nv′2+u′4\n4v′4+u′2\nv′3(ξk+r) (23)\nand the spectra weight density\nA+(k) =v′ω2\n+−ξk\n2v′2ω+(ω2\n+−ω2\n−),\nA−(k) =−v′ω2\n−+ξk\n2v′2ω−(ω2\n+−ω2\n−). (24)\nAt BCS limit the ratio of the two spectral weight densities can be appr oximately calculated as\nA−(k)\nA+(k)=u′2/radicalbig\nk2/2m∗(k2/2m∗+2r)\n4v′r2. (25)\nAt BCS limit we have u′/v′→0, this ratio vanishes. This spectral weight transfer is shown in Fig. 2 in the main\ntext.\nSpectral weight function in the case with damping term\nThe spectral weight function of the amplitude mode in the case with d amping term u′′is\nAaa=u′′ω\n2·|−vω2+k2\n2m∗|2+|uω|2\n|−(uω)2+(−vω2+k2\n2m∗)(−vω2+k2\n2m∗+2r)|2\n=u′′ω\n2·|−vω2+k2\n2m∗|2+|uω|2\n|v′2(ω2−˜ω2\n+)(ω2−˜ω2\n−)−2iu′u′′ω2|2, (26)5\nwhere the eigen mode frequencies are\n˜ω2\n±=ξk+r\nv′+u′2−u′′2\n2v′2±/radicalbigg\nr2\nv′2+(u′2−u′′2)2\n4v′4+u′2−u′′2\nv′3(ξk+r). (27)\nFor small momentum they can be approximated as\n˜ω−=/radicalbigg\n2rξk\n2v′r+u′2−u′′2,\n˜ω+=/radicalBigg\n2v′r+u′2−u′′2\nv′2+2v′r+2u′2−2u′′2\nv′(2v′r+u′2−u′′2)ξk. (28)\nCompared with the case without damping term we see that the gap of the Higgs mode is reduced from/radicalbig\n2r/v′to/radicalbig\n2r/v′−u′′2/v′2at BCS limit.\nFor smallξkthe spectral weight on the Goldstone mode can be calculated as\nAaa(˜ω−,k) =u′′\n8u′2˜ω−. (29)\nDifferent from the case without damping term, we see that in the cas e with damping term the spectral function has\na weight proportional to u′′on the Goldstone mode.\nThe spectral weight function in the case with Coulomb intera ction\nA time-dependent Ginzburg-Landau theory with Coulomb interactio n can be cast as [2]\nF=/integraldisplay\ndtd3x/braceleftBig\n−1\n8πφ∇2φ+¯∆/parenleftBig\n−iu(∂t−2eφ)+v(∂t−2eφ)2−∇2\n2m∗−r/parenrightBig\n∆+b\n2¯∆¯∆∆∆/bracerightBig\n, (30)\nwhereeis the electric charge and φ(t,x) is the electric field. By taking a symmetry breaking ∆ →∆0+δa+iδpwe\ncan have a free energy for the low energy excitations in the moment um space as\nF=/integraldisplaydω\n2πd3k\n(2π)3/braceleftBigg\n2uiωδa(ω,k)δp(−ω,−k)+δa(−ω,−k)(−vω2+k2\n2m∗+2r)δa(ω,k)+δp(−ω,−k)\n(−vω2+k2\n2m∗)δp(ω,k)+4iue∆0φ(−ω,−k)δa(ω,k)−4veω∆0φ(−ω,−k)δp(ω,k)+4ve2∆2\n0φ(−ω,−k)φ(ω,k)/bracerightBigg\n.\n(31)\nWe integrate out the electric field φand obtain\nF=/integraldisplaydω\n2πd3k\n(2π)3/braceleftBigg\n2uiωk2/8π\nk2/8π+4ve2∆2\n0δa(ω,k)δp(−ω,−k)+δp(−ω,−k)(−vω2k2/8π\nk2/8π+4ve2∆2\n0+k2\n2m∗)δp(ω,k)/bracerightBigg\n+δa(−ω,−k)(−vω2+k2\n2m∗+2r)δa(ω,k). (32)\nThen the spectral functions can be calculated as\nImχaa=u′′ω\n2·(−vω2k2/8π\nk2/8π+4ve2∆2\n0+k2\n2m∗)2+|uωk2/8π\nk2/8π+4ve2∆2\n0|2\n|−(uωk2/8π\nk2/8π+4ve2∆2\n0)2+(−vω2k2/8π\nk2/8π+4ve2∆2\n0+k2\n2m∗)(−vω2+k2\n2m∗+2r)|2,\nImχpp=u′′ω\n2·(−vω2+k2\n2m∗+2r)2+|uωk2/8π\nk2/8π+4ve2∆2\n0|2\n|−(uωk2/8π\nk2/8π+4ve2∆2\n0)2+(−vω2k2/8π\nk2/8π+4ve2∆2\n0+k2\n2m∗)(−vω2+k2\n2m∗+2r)|2. (33)\n[1] C. A. R. S´ a de Melo, M. Randeria, and J. R. Engelbrecht, Ph ys. Rev. Lett. 71, 3202 (1993).\n[2] Adriaan M. J. Schakel, Boulevard of Broken Symmetries: Effective Field Theories of C ondensed Matter , World Scientific,\nSingapore, 2008." }, { "title": "1309.2741v1.Initial_versus_tangent_stiffness_based_Rayleigh_damping_in_inelastic_time_history_seismic_analyses.pdf", "content": "arXiv:1309.2741v1 [physics.class-ph] 11 Sep 2013INITIAL VS. TANGENT STIFFNESS-BASED RAYLEIGH\nDAMPING IN INELASTIC TIME HISTORY SEISMIC\nANALYSES\nP. Jehel12, P. L´ eger3and A. Ibrahimbegovic4\npierre.jehel[at]ecp.fr\nPaper accepted for publication in Earthquake Engineering and Structural Dynamics\nPublished online (wileyonlinelibrary.com). DOI: 10.1002 /eqe.2357\nAbstract. In the inelastic time history analyses of structures in seismic mo-\ntion, part of the seismic energy that is imparted to the structure is absorbed\nby the inelastic structural model, and Rayleigh damping is commonly us ed in\npractice as an additional energy dissipation source. It has been ac knowledged\nthat Rayleigh damping models lack physical consistency and that, in t urn, it\nmust be carefully used to avoid encountering unintended conseque nces as the\nappearance of artificial damping. There are concerns raised by th e mass pro-\nportional part of Rayleigh damping, but they are not considered in t his paper.\nAs far as the stiffness proportional part of Rayleigh damping is conc erned, ei-\nther the initial structural stiffness or the updated tangent stiffn ess can be used.\nThe objective of this paper is to provide a comprehensive compariso n of these\ntwo types of Rayleigh damping models so that a practitioner i) can obj ectively\nchoose the type of Rayleigh damping model that best fits her/his ne eds and ii)\nis provided with useful analytical tools to design Rayleigh damping mod el with\ngood control on the damping ratios throughout inelastic analysis. T o that end,\na review of the literature dedicated to Rayleigh damping within these la st two\ndecades is first presented; then, practical tools to control the modal damping\nratios throughout the time history analysis are developed; a simple e xample is\nfinally used to illustrate the differences resulting from the use of eith er initial or\ntangent stiffness-based Rayleigh damping model.\nkeywords: Rayleigh damping, inelastic structure, modal analysis, damping\nratio time history, upper and lower bounds for damping ratios.\n1Laboratoire MSSMat /CNRS-UMR 8579, ´Ecole Centrale Paris, Grande voie des Vignes,\n92295 Chˆ atenay-Malabry Cedex, France\n2Department of Civil Engineering and Engineering Mechanics, Columbia University, 630 SW\nMudd, 500 West 120th Street, New York, NY, 10027, USA\n3Department of Civil Engineering, ´Ecole Polytechnique de Montr´ eal, P.O. Box 6079, Station\nCV, Montreal, QC, H3C 3A7, Canada\n4LMT-Cachan(ENS Cachan/CNRS/UPMC/PRESUniverSud Paris), 61 avenue du Pr´ esident\nWilson, 94235 Cachan Cedex, France\n12 RAYLEIGH DAMPING IN INELASTIC ANALYSES\n1.Introduction\nSeismic analyses were first developed for elastic structures. An ela stic struc-\nture does not absorb energy and, therefore, damping was added to represent all\nthe energy dissipation sources at the boundary conditions or at sm all scales not\nexplicitly considered for civil engineering applications (a list of such en ergy dissi-\npating phenomena can be found in [ 15]). For this purpose, Rayleigh damping is\nmathematicallyveryconvenient because, onceprojectedontoth eundampedmodal\nbasis, the set of equations of motion of the discrete structures a re uncoupled. This\ntype of damping is herein referred to as elastic damping . Then, to account for\nyielding mechanisms in the overall structural response, elastic visc ous-equivalent\nseismic analyses appeared [ 7, p.74]. In this case, the structural model still is elastic\nbut the damping properties are enhanced so as to account for bot h elastic damp-\ning and energy dissipation due to the actual, but not explicitly modeled , yielding\nresponse. This latter type of damping is herein referred to as hysteretic damping .\nHowever, inelastic timehistoryanalysis(ITHA), whicharebasedont hebuilding\nof an inelastic structural model, is the most appropriate way to pro perly account\nfor hysteretic mechanisms in seismic analyses. In this case, the str uctural modes\nare not uncoupled anymore in the modal basis. The modal basis could be updated\nafter each inelastic event to maintain uncoupled modes, but the com putational\nbenefit would be counterbalanced by the additional cost resulting f rom the succes-\nsive modal analyses required. Consequently, there is few mathema tical advantage\nin using Rayleigh damping for ITHA (the main advantage is that there is no need\nto explicitly build and store a damping matrix because mass and stiffnes s matrices\nalready are stored for other purposes). In spite of this, Rayleigh damping still is\ncommonly used in ITHAs. Ideally, Rayleigh damping should be added to m odel\nelastic damping only, and hysteretic damping should arise from the ex plicit mod-\neling of the energy dissipation mechanisms in the inelastic structural model. In\npractice, this can hardly be achieved because, on the one hand, co ntrolling the\namount of elastic damping in ITHA is challenging due to the intrinsic natu re of\nRayleigh damping models and, on the other hand, inelastic structura l models only\nprovide an approximation of the numerous inelastic phenomena that actually con-\ntribute to the seismic energy absorption in the structure. This pap er only focusses\non the issue of maintaining good control on elastic Rayleigh damping th roughout\ninelastic analysis.\nIn its most general form, Rayleigh damping consists in adding viscous forces of\nthe form fD(t) =C(t)˙u(t) in the discrete structural equations of motion, where\n˙uis the displacements rate and the damping matrix C(t) is built as a linear\ncombination of the structural mass and stiffness matrices MandK:\n(1) C(t) =α(t)M+β(t)K(t).\nThe Rayleigh coefficients αandβare computed so that the critical damping ratios\nξAandξBareobservedatthefrequencies ωAandωB.αandβcanbeeithersetonceRAYLEIGH DAMPING IN INELASTIC ANALYSES 3\nand then frozen or updated throughout ITHA. The stiffness matr ix is commonly\nbuilteitherfromtheinitialorfromtheupdatedtangentstiffnesses . Bothmass-and\nstiffness-proportional terms of Rayleigh damping models can gener ate difficulties\nin controlling elastic damping throughout an inelastic analysis. Contra ry to the\nstiffness-proportional term, which is difficult to control in inelastic a nalyses only,\nproblems can arise from the mass-proportional component no mat ter whether it\nis an elastic(-equivalent) or an inelastic time-history analysis. Indee d, assuming\nthat the mass matrix is diagonal, the physical counterpart of a mas s-proportional\nterm is a viscous damper connecting a structural degree of freed om to the frame\nwhich contains the reference point for measuring displacements in t he structure.\nThen, the mass-proportional term can lead to unrealistically high da mping forces\nwhenever the whole or parts of the structure behave like a rigid bod y. This issue\nencountered with the mass-proportional term has already been w ell explained and\nillustrated [ 10]. Consequently, in this paper, although both mass- and stiffness-\nproportional terms are used to construct the damping matrix, on ly the stiffness-\nproportional component is focussed on.\nHow to avoid spurious damping forces andimprove control on dampin g through-\nout inelastic analyses has been widely studied and led to practical rec ommenda-\ntions (e.g.[10,14,5]). However, when it comes to dealing with the stiffness-\nproportional term, one can find in the literature differing viewpoints on whether\nto use initial or tangent stiffness. The objective of this paper is to p rovide a\ncomprehensive comparison of Rayleigh damping based on either initial or tangent\nstiffness, and to answer the question: “which of initial and tangent stiffness-based\nRayleigh damping provides better control on damping ratio through out ITHA?”\n(the mass-proportional term being also present in the Rayleigh dam ping models).\nTo that aim, we first present in the next section a review of the litera ture where\nthe different strategies that have been proposed to build the Rayle igh damping\nstiffness-proportional term are exposed. Then, in section 3, mat hematical de-\nvelopments lead to formulas that allow quantifying damping ratios shif ts due to\nstiffness degradations. These relations are useful to design a Ray leigh damping\nmodel before running an analysis and, once the analysis has been ru n, to assess\nthe validity of the elastic damping modeling. To the best of our knowled ge, this is\nthe first time such formulas are provided for initial stiffness-based Rayleigh damp-\ning. For the sake of completeness, we already mention here that th e formula we\nderive in section 3 are only valid when the two following assumptions are made: i)\nthe damping ratios ξAandξBchosen to identify the Rayleigh coefficients αandβ\nare taken as equal ( ξA=ξB), and ii) the structural equations of motion are uncou-\npled when expressed in modal coordinates, which is in most of the cas es only an\napproximation for inelastic structures with initial stiffness-based R ayleigh damp-\ning. Those two assumptions are common for the type of problems we consider in\nthis work and they do not reduce the contributions of this paper co mparing to4 RAYLEIGH DAMPING IN INELASTIC ANALYSES\nprevious publications on the subject. Before closing the paper with some conclu-\nsions, we compare the performances of initial and tangent stiffnes s-based Rayleigh\ndamping models in the analysis of a simple structure with stiffness degr adations.\n2.In the literature\nAccording to Charney [ 5], one can go back to the eighties to find the first papers\ndedicated to problems pertaining to modeling damping in inelastic struc tures with\nthe work of Chrisp in 1980 [ 6] and Shing and Mahin in 1987 [ 16].\nIn the nineties, L´ eger and Dussault [ 14] investigate, in the case of inelastic frame\nstructures in seismic loading, the variation of nonlinear response ind icators (aver-\nage ductility, hysteretic-to-input energy ratio, average number of yield incursions)\naccording to the additional damping model used. A bilinear hysteres is model is\nassigned to the structural elements. Amongst other mass- or st iffness-proportional\ndamping models, Rayleigh damping models either with elastic or tangent stiffness\nand either with frozen or updated coefficients ( α,β) are considered. The authors\nhave developed a computer program to update αandβat each time step. This\nallows maintaining constant critical damping ratio throughout the se ismic analy-\nses for the two frequencies used to identify the Rayleigh coefficient s, which are, in\nthis work, the first natural frequency and the frequency for wh ich 95% of effective\nmodalmass isrepresented by thetruncated eigenbasis. Notetha t these frequencies\nare not the initial elastic ones but that they also are updated at eac h time step.\nL´ eger and Dussault show that, while having little effects on the amou nt of energy\nimparted to a structure by an earthquake, the choice of an additio nal damping\nmodel significantly influences the amount of hysteretic energy due to damage in\nthe structure. The term “additional damping” is used here to refe r to a source of\ndamping that comes in addition to the damping resulting from the abso rption of\nhysteretic energy in the structural model. L´ eger and Dussault u se for instance the\ndisplacement ductility averaged over all the stories as an indicator o f the inelastic\nstructural response and, for the El Centro ground motion cons idered, variations\nof this latter indicator of up to 40% from one damping model to the ot her are\nobserved. Considering the Rayleigh model with updated coefficients as a base-\nline, recommendations on which additional damping model to use given the elastic\nfundamental period of the structure close the paper.\nFrom the statement that the number of degrees of freedom (DOF s) needed to\nassemble the stiffness matrix is often much larger than that needed for building\nan adequate mass matrix, Bernal [ 3] shows that spurious damping forces are likely\nto arise from the presence of massless – or with relatively small inert ia – DOFs in\ninelastic structural systems. Indeed, massless DOFs have the te ndency to undergo\nabrupt changes in velocity when stiffness changes, leading to unrea listically large\nviscous damping forces. In this work, Bernal adds proportional d amping in the\nequilibrium equations using the Caughey series. It is stated that the most impor-\ntant effect of spurious damping forces is found in the distortion the y introduceRAYLEIGH DAMPING IN INELASTIC ANALYSES 5\nin the maximum internal forces rather than the displacements. As a solution to\nprevent spurious damping due to massless DOFs, it is suggested to a ssemble the\ndamping matrix using the stiffness matrix condensed to the size of th e DOFs with\nmass, followed by expansion to the full set of coordinates with colum ns and rows\nof zeros. This procedure is equivalent to what is proposed in other w orks where\nit is expressed in the other terms: assembling the damping matrix, ele ment by\nelement, with zero damping assigned to the DOFs where local abrupt changes of\nstiffness can occur (see e.g.the procedure advocated in [ 8] for structures with fiber\nelements, or the issues raised by local stiffness changes in [ 5]).\nIn [10], Hall focuses on practical situations where the use of Rayleigh dam ping\ncan lead to damping forces that are unrealistically large compared to the restoring\nforce, and then proposes a capped viscous damping formulation to overcome some\nof the problems pointed out. It is stated in this paper that the tang ent stiffness\nshould not beused to build theRayleigh damping matrix, especially beca use ofthe\nconvergence difficulties it can cause. To illustrate this assertion, th e author gives\nthe example of the local damping stress, resulting from the stiffnes s-proportional\npart ofRayleigh damping, that canjump fromzero toa possibly large value incase\nof crack closing as soonascontact ismade. Then, Hall distinguishes between prob-\nlems arising from the mass- and stiffness-proportional part of Ray leigh damping\nand quantifies the likely undesirable effects. The mass-proportiona l term can lead\nto unrealistically high damping forces whenever the whole or parts of the structure\nbehave like a rigid body: formulation of an earthquake analysis in term s of total\nmotion, superstructure on a relatively flexible base, portions of a s tructure that\nbreak loose like in a dam undergoing sliding at its base,... As far as the stiff ness-\nproportional term of Rayleigh damping is concerned, very large dam ping forces\ncan be generated when Rayleigh damping is based on the initial stiffnes s, while\nstructural elements yield, leading to an increase in the velocity grad ient: gravity\ndam undergoing cracking, presence of penalty elements,... Finally, as a remedy\nto the problems listed in the paper, the author proposes a capped v iscous damp-\ning formulation in which the mass-proportional contribution is eliminat ed and the\nstiffness-proportional contribution is limited by bounds defined in ac cordance with\nthe actual physical mechanism that limits the structural restorin g forces.\nIn [5], Charney first investigates the effects of global stiffness change s on the\nseismic response of a 5-story structure when Rayleigh damping is us ed. Two cases\nare considered: a reduction of the story stiffness by the same fac tor for each story,\nwhich doesnotchangethemodeshapes, andanonuniformstoryst iffness reduction\nalong the height, which leads to mode shapes that are different from those of the\ninitial structure. In both cases, it is shown that, when structure yields, i) artificial\ndamping is generated when the Rayleigh damping matrix is computed ac cording\nto the initial structural properties; ii) significant but reduced art ificial damping is\ngeneratedwhenRayleighdampingisbuiltwithtangentstiffnessandfix edRayleigh\ncoefficients computed fromtheinitial stiffness; iii) virtually no artificia l damping is6 RAYLEIGH DAMPING IN INELASTIC ANALYSES\ngenerated when Rayleigh damping is based on both tangent stiffness and updated\ncoefficients from the tangent stiffness.\nThen, Charney [ 5] also investigates the effects of local stiffness changes on\nthe seismic structural time-history response when initial stiffness -based Rayleigh\ndamping is used. The same 5-story structure as previously is consid ered, but this\ntime with inelastic rotational springs at beam-to-column connection s, which rep-\nresent local yielding mechanisms. There is no rotational mass and ze ro hardening\nafteryielding. Under these conditions, threedifferent pairsofbea ms/columns stiff-\nness and connections elastic stiffness are defined so that, in each c ase, the overall\nstiffness matrix of the structure be unchanged. The response his tories are identi-\ncal when a linear analysis is performed and when the damping matrix is b uilt so\nthat the rows and columns pertaining to the rotational DOFs are fille d with zeros.\nHowever, when the Rayleigh damping matrix is built with non-zero initial stiffness\nassociated to the massless DOFs, artificial viscous damping forces develop. These\nforces can be extremely high, especially as the initial spring stiffness is large. Yet\nthey are not easy to detect.\nTo Charney, the best strategy to model damping would be of cours e to eliminate\nthe use of viscous damping, because it is not physically sound, and re place it by\nfrictional or hysteretic devices. Nevertheless, amongst other r ecommendations,\nCharney advocates that Rayleigh damping based on tangent stiffne ss be used.\nIndeed, with this method, the problems associated with local stiffne ss changes are\nalways eliminated, and reduced artificial damping is generated becau se of global\nstiffness changes, which, moreover, can still be reduced by anticip ating the shift\nof the structural natural frequencies. If elastic stiffness-bas ed Rayleigh damping\nonly is implemented in the computer program used, the damping matrix should\nbe computed with zero stiffness associated to the elements that ha ve large initial\nstiffness and that are likely to yield.\nThe more recent papers which also focus on modeling structural da mping in in-\nelastic time history seismic analyses are those of Zareian and Medina [ 17] and Er-\nduran [9]. In these papers, practical strategies are presented to cope w ith the\nproblems encountered with Rayleigh damping in the context of perfo rmance-based\nseismic design. In the approach presented in [ 17], each structural element is mod-\neled with an equivalent combination of an elastic element with initial stiffn ess-\nproportional damping and yielding springs at the two ends without st iffness-\nproportional damping. With this strategy, i) numerical solution inst abilities when\nsignificant changes in stiffness values occur are avoided because init ial stiffness\nmatrix is used in the damping model, and, also, no local spurious dampin g forces\nare generated in the structural parts that yield because there is no stiffness-\nproportional part in the damping model pertaining to these parts.\nIn [9], Erduran recalls that it has been shown that designing Rayleigh damp ing\nmodels based on the initial stiffness matrix results in unreasonably hig h dampingRAYLEIGH DAMPING IN INELASTIC ANALYSES 7\nforces after yielding and consequently decides to exclusively use ta ngent stiffness-\nbased Rayleigh damping models in his study. This study consists in analy zing the\nstory drift ratios, floor accelerations and damping forces in two 3- and 9-story steel\nmoment-resisting frame buildings for three seismic hazard levels. It is concluded\nin [9] that, as long as they are designed according to reduced modal fr equencies\n(comparing to the elastic properties), Rayleigh damping models with b oth mass-\nand tangent stiffness-proportional components lead to reasona ble damping forces\nand floor acceleration demands without suppressing higher modes e ffects in the\n9-story building.\nEither following or initiating these research efforts, there are comp uter program\nuser manuals and technical reports who directly provide practition ers with ad-\nvanced tools and guidelines for proper implementation of Rayleigh dam ping in\ninelastic time history seismic analyses. For instance: eight types of d amping mod-\nels are implemented in the computer program RUAUMOKO, whose user manual\nalso comes with comprehensive discussion on damping [ 4, pp. 9-10]; visualization\ntools for damping effects in the simulations are available in the compute r program\nPERFORM-3D, and an entire chapter is dedicated to modeling damping in the\naccompanying user guide [ 8,§18]; a large share is dedicated to viscous damping for\ninelastic seismic analyses in the technical reports [ 2,§2.4.2] and, to a more limited\nextent, in [ 1,§6.4.4].\nFrom the review of the literature above, it is obvious that some auth ors recom-\nmend using initial stiffness-based Rayleigh damping models while others recom-\nmend using tangent stiffness. In the following section, we provide ma thematical\ndevelopment for a rational comparison of the two approaches, re lying on an anal-\nysis of the damping ratios time history throughout inelastic analyses .\n3.Mathematical developments\n3.1.Modal analysis for inelastic structures. The equilibrium equations of a\nstructure in seismic loading, with additional viscous damping, discret ized in space,\nand transformed to undamped modal coordinates at time t, read:\n(2)ΦT\ntMΦt¨U(t)+ΦT\ntC(t)Φt˙U(t)+ΦT\ntK(t)ΦtU(t) =ΦT\nt/parenleftbig\nFsta(t)+Fsei(t)/parenrightbig\nwhereU(t) ={Um(t)}m=1,..,Nmis the undamped modal coordinates vector, Fsta(t)\nandFsei(t) are the pseudo-static and seismic forces (the static force is app lied step\nby step prior to the application of the seismic load to grasp possible no nlinear\nmechanisms), Φt= (φ1(t)... φNm(t)) is the matrix composed by the undamped\nmodal shape vectors, and /squareTdenotes the transpose operator. Mode shapes can\nalso be computed from the damped system, but the undamped assu mption is\ncommonly retained for systems with low damping ratios, as it is the cas e in this\nwork. Although cumbersome, the explicit dependence of every qua ntity on time\ntis indicated ( /square(t) or/squaret) to emphasize that each of these quantities can change\nwithin the inelastic time history of the structure.8 RAYLEIGH DAMPING IN INELASTIC ANALYSES\nThere are two ways to compute the viscous damping matrix. If modal damping\nis used, the modal damping ratios ξmare chosen at time twhen the modal analysis\nis performed, and the damping matrix is built as diagonal:\n(3) ΦT\ntC(t)Φt= diag(Cm(t)) with Cm(t) = 2ξm(t)ωm(t)Mm(t).\nIn this relation, we have introduced the undamped modal circular fr equencies ωm\nasω2\nm(t) =Km(t)/Mm(t), along with the notation Am=φT\nmAφmwhereA=M,\nCorK. IfRayleigh damping is used, a set of coefficients ( α,β) is computed, and\nthe damping matrix is built as:\n(4)C(t) =α(t)M+β(t)K(t)⇒ C m(t) =α(t)Mm(t)+β(t)Km(t).\nFrom equations ( 3) and (4), we thus have the relation\n(5)α(t)φT\nm(t)Mφm(t)+β(t)φT\nm(t)K(t)φm(t) = 2ξm(t)ωm(t)φT\nm(t)Mφm(t)\nbetween – right-hand side – the modal quantities that would arise fr om a modal\nanalysis at any time tin thehistory of thestructure and– left-handside – what ac-\ntually results from the use of Rayleigh damping at this specific time t. Relation ( 5)\nprovides a definition of the modal damping ratio.\nIn equation ( 2), bothΦT\ntMΦtandΦT\ntK(t)Φtare diagonal matrices. ΦT\ntC(t)Φt\nhowever can be non-diagonal, which implies that ( 2) is a set of coupled equa-\ntions. A damping model that is not diagonalized in the modal basis is qua lified as\nnon-classical ornon-proportional . As shown in relation ( 3), modal damping inher-\nently is classical. Rayleigh damping is however not necessarily classical: if initial\nstiffness is used, off-diagonal terms appear in ΦT\ntCΦtin most of the cases when\nstructure yields. Relation ( 5) provides exact values of ξm(t) forclassical damping\nonly; otherwise, not all the damping energy is included in the damping r atios.\nConsequently, in the presence of non-classical damping, the defin ition ofξmin (5)\nis only valid if the off-diagonal terms in ΦT\ntCΦtcan be neglected. For systems\nwith low damping ratios (a few percents of critical), as it is the case in t his work,\nthe assumption that the equations in ( 2) can approximated as decoupled in the\npresence of non-classical damping generally is, explicitly or implicitly, r egarded as\nvalid in the literature ( e.g.in [10,5]). Accordingly, we will hereafter retain this\nassumption.\n3.2.Implementing Rayleigh damping. In practice, there are several methods\nto build a Rayleigh damping matrix, which all have different consequenc es on the\ntime history of the damping ratios throughout the inelastic analysis:\na. The Rayleigh coefficients, as well as the stiffness matrix, are set on ce for all\nbefore the beginning of the dynamic analysis. We refer to these qua ntities\nas (α0,β0) andK0. These quantities are not necessarily identified consid-\nering the state of the structure at the beginning of the dynamic an alysis:\nbased on an approximation of the reduced stiffness, the state of t he struc-\nture at any time within or at the end of the seismic loading can be used f orRAYLEIGH DAMPING IN INELASTIC ANALYSES 9\nthe identification of the Rayleigh coefficients before running the dyn amic\nanalysis. At a given time tin the structure time history, equation ( 5) thus\ntakes the following expression:\n(6) ξa\nm(t) =1\n2ωm(t)/parenleftbigg\nα0+β0φT\nm(t)K0φm(t)\nMm(t)/parenrightbigg\nwhich we rewrite as\n(7)ξa\nm(t) =1\n2/parenleftbiggα0\nωm(t)+β0ha\nm(t)ωm(t)/parenrightbigg\nwithha\nm(t) =φT\nm(t)K0φm(t)\nKm(t).\nNotethat ΦT\ntK0Φtgenerallyisnotdiagonalandthattheoff-diagonalterms\nare neglected in this definition of ξa\nm(t).\nb. TheRayleighcoefficientsaresetonceforallatthebeginningofthe dynamic\nanalysis, asincase a,andthetangentstiffnessmatrix K(t),updatedateach\ntime step, is used to build C(t). With this definition, Rayleigh damping is\nclassical throughout the analysis and we have:\n(8) ξb\nm(t) =1\n2/parenleftbiggα0\nωm(t)+β0ωm(t)/parenrightbigg\n.\nc. Both the Rayleigh coefficients and the stiffness matrix are updated at each\ntime step in the numerical analysis. In this case, the modal damping r atio\nξc\nm(t) can be controlled throughout the analysis, e.g.maintained constant\nby updating ( α,β) accordingly:\n(9) ξc\nm(t) =1\n2/parenleftbiggα(t)\nωm(t)+β(t)ωm(t)/parenrightbigg\n.\nd. There is no risk of generating spurious local damping forces when m eth-\nods based on the updated tangent stiffness are used to implement R ayleigh\ndamping [ 5]. However, this risk arises when, as in case a, the Rayleigh\ncoefficients along with the stiffness matrix are set once for all befor e the\nbeginning of the dynamic analysis, and there are some elements in the\nstructure with artificially large initial stiffness (plastic hinges, gap or con-\ntact elements). It is then advocated, e.g.in [5], to associate reduced β0\nfactors to this latter type of element ( βe\n0< β0). However, this method can\npotentially generate the same type of global effects as method a. Indeed,\ndefining a reduced initial stiffness matrix Kr\n0as:\n(10) C=Ae(α0Me+βeKe\n0) =α0M+β0Kr\n0withβ0Kr\n0=Aeβe\n0Ke\n0,\nwhereAedenotes the finite element assembly procedure [ 11], it comes from\nequation ( 5):\n(11)ξd\nm(t) =1\n2/parenleftbiggα0\nωm(t)+β0hd\nm(t)ωm(t)/parenrightbigg\nwithhd\nm(t) =φT\nm(t)Kr\n0φm(t)\nKm(t).10 RAYLEIGH DAMPING IN INELASTIC ANALYSES\nIf the reduced initial stiffness Kr\n0better approximates the tangent stiffness\nmatrix than the initial stiffness K0, the global effects with method dare re-\nduced comparing to method a. Also, using this method, off-diagonal terms\ncan appear in the generalized damping matrix; those terms are negle cted\nin this definition of ξd\nm(t).\nFor methods bandc, the updated stiffness matrix could loose its positiveness if\nsignificant softening or second-order effects would occur. Mathe matically, it would\nimply that the modal frequencies be null or imaginary ( ω2≤0), which, physically,\nwould not make any sense. However, this would only happen if the str ucture\nbecame unstable and, consequently, this is an ultimate potential pr oblem that we\nlet out of the scope of this work.\nMethodcis rarely used in practice because a modal analysis has to be carried\nout after each inelastic event. Therefore, hereafter, we do not consider case cany-\nmore and the Rayleigh coefficients ( α0,β0) are set once for all before the dynamic\nanalysis.\n3.3.Rayleigh coefficients computation. For method i=a,bord, we have,\naccording to what precedes:\n(12) ξi\nm(t) =1\n2/parenleftbiggα0\nωm(t)+β0hi\nm(t)ωm(t)/parenrightbigg\nwithhb\nm(t) = 1,∀m ,∀t .\nTwo independent equations can be written from equation ( 12) with two different\npairs of damping ratios and circular frequencies hereafter referr ed to as ( ξA,ωA)\nand (ξB,ωB). The resulting set of two equations can then be solved to obtain α0\nandβ0. For inelastic analyses, the practitioner can choose the pairs ( ξA,ωA) and\n(ξB,ωB) either for two different modes mAandmBat the same time t=tA=tB,\nor for two different modes mAandmBat two different times tAandtB, or for the\nsame mode m=mA=mBat two different times tAandtB.AandBthus refer to\na mode and a particular time in the history of the structure and, in th e following,\nAhas to be understood as the pair ( t=tA,m=mA), and so for B.\nFor the sake of conciseness, we abandon the superscript ithat explicitly refers\nto method a,bord. Yet, one should keep in mind that the quantities ξA/B,ωA/B,\nhA/B,α0andβ0all depend on the method used. Rewriting equation ( 12) forA\nandB, we then have:\n(13)/braceleftbigg\nξA\nξB/bracerightbigg\n=1\n2/parenleftbigg\n1/ωAhAωA\n1/ωBhBωB/parenrightbigg/braceleftbigg\nα0\nβ0/bracerightbigg\n.\nTheRayleighcoefficientscanthenbecomputedbyinvertingrelation( 13),which,\nas long as hAω2\nA/ne}ationslash=hBω2\nB, is always possible:\n(14)/braceleftbigg\nα0\nβ0/bracerightbigg\n=2ωAωB\nhBω2\nB−hAω2\nA/parenleftbigg\nhBωB−hAωA\n−1/ωB1/ωA/parenrightbigg/braceleftbigg\nξA\nξB/bracerightbigg\n.RAYLEIGH DAMPING IN INELASTIC ANALYSES 11\nWe assume that ξA=ξB=ξ0, which will simplify the following developments.\nBecause this assumption is commonly done in practice, it does not red uce the\ncontributions of the present work. We then have:\n(15)α0=2ωAωB(hBωB−hAωA)\nhBω2\nB−hAω2\nAξ0andβ0=2(ωB−ωA)\nhBω2\nB−hAω2\nAξ0.\nNote that the coefficients hgenerally are absent in the definition of the Rayleigh\ncoefficients. Common definitions for α0andβ0indeed are:\n(16) α0=2ωAωB\nωA+ωBξ0andβ0=2\nωA+ωBξ0,\nbut this actually is only exact for case bwherehb\nm(t) = 1 for all tandm. There\nis another important notion that is generally not explicitly mentioned in the def-\ninition of the Rayleigh coefficients in ITHA and that we recall here: the indexA\n(respectively B) does not only refer to a specific mode but to a mode at a given\ntime in the history of the structure.\n3.4.Upper and lower bounds for the damping ratios. We now seek to\npredict the shift of the damping ratiospertaining to modal frequen cies that remain\nin a pre-determined range, throughout the inelastic time history an alysis. In other\nwords, we seek an answer to the problem illustrated in figure 1: How to compute\n∆>0 so that, for ωA,ωB, and a targeted damping ratio ˆξgiven, for all tand for\nallm:\n(17) ωA≤ωm(t)≤ωB⇒ˆξ−∆< ξm(t)<ˆξ+∆ ?\nWe denote ξmax=ˆξ+ ∆ and ξmin=ˆξ−∆. We moreover set ξmax=ξ0(=\nξA=ξB) and introduce R >1 so that ωB=R×ωA. With these notations and\nconsidering the expression of the Rayleigh coefficients in equations ( 15), we rewrite\nrelation ( 12) as:\n(18)ξm(t)\nξmax=1\nR2hB−hA/parenleftBigg\nR(R hB−hA)1\nωm(t)\nωA+(R−1)hm(t)ωm(t)\nωA/parenrightBigg\n.\nTo guarantee ξm(t)>0, withhm(t)>0,ωm(t)/ωA>0, andR >1, we require\nthathA>0 andhB>0 be chosen so that they satisfy the condition:\n(19) R hB−hA≥0,\nwhich implies that R2hB−hA≥0 because R≥1.\nWe now study relation ( 18) for methods a,banddexplicitly. Because method\nbis the easiest to deal with, we start with it and methods aanddare considered\ntogether right after.12 RAYLEIGH DAMPING IN INELASTIC ANALYSES\nFigure 1. We seek to characterize the grey box within which the\ndampingratiosremainallalongtheinelasticanalysis( t∈[0,T]).ˆξis\nthe targeted damping ratio in the range [ ωA,ωB], ∆ is the deviation\nof the actual damping ratio from the targeted one. The black point s\nindicate two possible points to identify the Rayleigh coefficients\n(α0,β0).\nMethod b:hA=hB= 1 and hm(t) = 1 for all tandm, leading to the function\n(20)ξm(t)\nξmax=1\n1+R/parenleftBigg\nR1\nωm(t)\nωA+ωm(t)\nωA/parenrightBigg\n,\nwhich we write asξm(t)\nξmax=f/parenleftBig\nωm(t)\nωA/parenrightBig\n. To study the variation of ξm(t) with respect\ntoωm(t) in the range [ ωA,ωB], it is useful to compute the derivative f′of the\nfunction fand to analyze its sign:\n(21) f′/parenleftbiggωm(t)\nωA/parenrightbigg\n=d(ξm(t)/ξmax)\nd(ωm(t)/ωA)=1\n1+R/parenleftBigg\n1−R1/parenleftbigωm(t)\nωA/parenrightbig2/parenrightBigg\n,\nwhich implies that f′/parenleftBig√\nR/parenrightBig\n= 0 and also that f′/parenleftBig\nωm(t)\nωA/parenrightBig\n<0 forωA≤ωm(t)<\n√\nR ωA, andf′/parenleftBig\nωm(t)\nωA/parenrightBig\n>0 for√\nR ωA< ωm(t)≤ωB, because R >1. Conse-\nquently, ξm(t) decreases for ωm(t)∈[ωA,√\nR ωA] an then increases for ωm(t)∈\n[√\nR ωA,ωB] after reaching the minimum:\n(22)ξm(t)\nξmax/vextendsingle/vextendsingle/vextendsingle/vextendsingle\nmin=ξm(t)\nξmax/vextendsingle/vextendsingle/vextendsingle/vextendsingleωm(t)\nωA=√\nR=2√\nR\n1+R.\nBecause stiffness degradations lead to a decrease of the modal fr equencies, the\ndamping ratios in the frequency range [ ωA,√\nR ωA] (respectively [√\nR ωA,ωB])\nwill necessarily increase (respectively necessarily decrease) durin g the inelastic\nanalysis. As will be illustrated in the applications that follow this section , this\nresult is useful to design the damping model because it provides info rmation onRAYLEIGH DAMPING IN INELASTIC ANALYSES 13\nhow to choose the modes and times AandBfor which the maximum damping\nratioξmaxare observed.\nThen, one can compute ∆bas follows. By definition, 2∆b=ξmax−ξmin, with\n(see equation ( 22))ξmin=2√\nR\n1+Rξmaxandξmax=ˆξ+∆b, which gives:\n(23) ∆b=ˆξ1+R−2√\nR\n1+R+2√\nR.\nThis result is mode- and time-independent. It is a useful measure of the damping\nratios drifts in the range [ ωA,ωB], throughout ITHA. Being able to guarantee\nlow ∆ for large Rwould be an ideal situation to provide good control on the\ndamping ratios. It is however not possible for Rayleigh damping model because\n∆bmonotonically grows with R(d∆b\ndR=2(R−1)ˆξ\n(1+R+2√\nR)2≥0).\nAn analogous development is presented in [ 10] where the same result is given for\n∆ (equation (5) in [ 10]). But, it should be noted that it is only valid for method\nb, not for methods aord.\nMethods aandd:Finding a mode- and time-independent expression for ∆ is in\nthese cases not as straightforward as for method bbecause the value of hm(t) is not\nknowna priori. First, we define Has a mode- andtime-independent variable, such\nthathm(t)≥H≥1, which, assuming that the stiffness matrix remains positive,\nis true for all mandt(see equations ( 7) and (11)). Then, replacing hm(t) byH,\nwe can rewrite equation ( 18) as:\n(24)\nξm(t)\nξmax≥1\nR2hB−hA/parenleftBigg\nR(R hB−hA)1\nωm(t)\nωA+(R−1)Hωm(t)\nωA/parenrightBigg\n=g/parenleftbiggωm(t)\nωA/parenrightbigg\nandstudythevariationsofthefunction gwithrespectto ωm(t). Fromananalogous\nprocedure as for method b, we conclude that gdecreases for ωm(t)< G ω A, is\nminimum for ωm(t) =G ωA, and then increases for ωm(t)> G ω A, whereG=/radicalBig\nR(RhB−hA)\n(R−1)H. Consequently:\n(25)ξm(t)\nξmax≥g(G) =2/radicalbig\nR(R−1)(R hB−hA)H\nR2hB−hA.\nBecause H≥1, it follows that a mode- and time-independent lower bound for\nξminis:\n(26) ξmin≥ξmax2/radicalbig\nR(R−1)(R hB−hA)\nR2hB−hA.\nIn contrast to what was found for method b, we only have here an upper bound\nforξmin. Besides, for method b, it is certain that the minimum of ξm(t) is attained\nforωm(t)∈[ωA,ωB] whereas, for methods aandd, we cannot determine a priori\nwhere the minimum will be observed. Therefore, when designing the d amping14 RAYLEIGH DAMPING IN INELASTIC ANALYSES\nmodel, there is in these cases no indication on how to efficiently choose the modes\nand times AandBfor which ξmaxis reached.\nConsidering that AandBhave been chosen, and using the same procedure as\nfor method b, we obtain the following upper bound for ∆a,d:\n(27) ∆a,d≤∆a,d\nmax=ˆξR2hB−hA−2/radicalbig\nR(R−1)(RhB−hA)\nR2hB−hA+2/radicalbig\nR(R−1)(RhB−hA).\nNote that for hA=hB= 1, we recover ∆a,d\nmax= ∆b.\n3.5.Is∆a,d\nmax≤∆b?If this inequality were true, ∆a,dwould necessarily be lower\nthan ∆bforRandωAgiven and no matter what hAandhBwould be equal\nto. Unfortunately, we show in figure 2that, for ha,d\nA≥1 andha,d\nB≥1, we have\n∆a,d\nmax≥∆b, with ∆a,d\nmax= ∆bforha,d\nA=ha,d\nB= 1. It is thus impossible to con-\nclude anything about whether using initial stiffness-based Rayleigh d amping can\nprovide better control on the damping ratios than tangent stiffne ss-based Rayleigh\ndamping. Besides, one can see in figure 2that, in certain area of the ( ha,d\nA,ha,d\nB)\nplane, ∆a,d\nmaxcan be large comparing to ∆b.\nFigure 2. ∆a,d\nmaxwith respect to ( ha,d\nA,ha,d\nB). ∆a,d\nmax= ∆bfor\n(ha,d\nA,ha,d\nB) = (1,1). For the purpose of illustration, we use here\nR= 10 and ˆξ= 2%.\n3.6.Summary. In table1, we summarize the formula we derived in this section\nto calculate the damping ratios time histories and to evaluate their ma ximum\ndrift. Obviously, Rayleigh damping models designed according to meth odbhave\ntwo advantages over models designed with methods aord:RAYLEIGH DAMPING IN INELASTIC ANALYSES 15\n(i) The sole quantities that are a priori unknown with method bareωAand\nR, which is much easier to approximate with fair accuracy with some ex-\nperience or simplified analyses than the quantities hwith methods aor\nd;\n(ii) We have access to the exact value of ∆b, whereas we can only calculate\nan upper bound for ∆a,dand it is not straightforward to identify the pair\n(tB,mB) that defines ξB=ξmax.\nNevertheless, we recall that damping models designed from method bcan cause\nsolution convergence problems, which is reported in the review of th e literature\nabove. It is important to stress here that those models only have t he potential\nto provide practitioners with a better control on the damping ratio s time history\nthroughout inelastic seismic analysis. Indeed, as will be illustrated in t he next\nsection, it does not imply that a damping model designed from method aordis\nnecessarily poor.\nTable 1. Initial vs. tangent stiffness-based Rayleigh damping:\nwhat the calculus provides. ξmax=ξA=ξBwithξA/B=\nξm=mA/B(t=tA/B) andξmax=ˆξ+ ∆ with ˆξthe targeted damp-\ning ratio in [ ωA,R×ωA].\ni ξi\nm(t)/ξmax ∆i/ˆξ Comments\na1\n1+R/parenleftbigg\nR\nωm(t)\nωA+ha\nm(t)ωm(t)\nωA/parenrightbigg\n≤R2ha\nB−ha\nA−2√\nR(R−1)(Rha\nB−ha\nA)\nR2ha\nB−ha\nA+2√\nR(R−1)(Rha\nB−ha\nA)·ha\nm(t) =φT\nm(t)K0φm(t)\nφT\nm(t)K(t)φm(t)\n·AandBa prioriunknown\nb1\n1+R/parenleftbigg\nR\nωm(t)\nωA+ωm(t)\nωA/parenrightbigg\n=1+R−2√\nR\n1+R+2√\nR/\nd1\n1+R/parenleftbigg\nR\nωm(t)\nωA+hd\nm(t)ωm(t)\nωA/parenrightbigg\n≤R2hd\nB−hd\nA−2√\nR(R−1)(Rhd\nB−hd\nA)\nR2hd\nB−hd\nA+2√\nR(R−1)(Rhd\nB−hd\nA)·hd\nm(t) =φT\nm(t)Kr\n0φm(t)\nφT\nm(t)K(t)φm(t)\n·AandBa prioriunknown\n4.Illustrative applications\n4.1.Example 1: Rayleighdamping designed from elastic structur al prop-\nerties.The inelastic structural models considered in this example are the sa me\nas in Charney’s work [ 5], which we briefly present here. The structure is a five-\nstory building modeled as a system of five DOFs – the horizontal displa cements\n– connected by inelastic columns which all have the same elastic prope rties. At\neach DOF same mass is lumped. Concerning the inelastic response, tw o different\nyielding scenarios are imagined to occur during a hypothetical ITHA:\n(i) The entire stiffness matrix is assumed to uniformly reduce to 50% o f its\noriginal value;16 RAYLEIGH DAMPING IN INELASTIC ANALYSES\n(ii) Thestructuralelements nonuniformlyyieldalongthebuildingheigh t:Nth-\nstory stiffness is reduced to 10%+( N−1)×20% of its original value, that\nis 10% for the the 1ststory, 30% for the 2nd,..., 90% for the 5th.\nFor this first example, we also design the same Rayleigh damping models as in\nCharney’s work [ 5], including with method cfor a complete comparison. Accord-\ningly, Rayleigh damping models are designed so that a critical damping r atio of\n2% is observed for the modes 1 and 3 of the structure in its initial sta te (t= 0).\nAdopting theanalytical frameworkpresented intheprevious sect ion, itmeans that\nwe design damping models with the following parameters: ωA=ω1(0),ωB=ω3(0)\nandξ0=ξA=ξB=ˆξ= 2%. To clearly show the time history of the damping\nratios, we arbitrarily set the total duration of the inelastic analysis T= 1 s and\ndivide the analysis into 5 time steps (0 ,0.2 s,...,1 s). Each time step corresponds\nto immediate degradations in the structure: e.g., in the case of the uniform yield-\ning scenario described above (scenario 1), at t= 0 the structural stiffness matrix\nKis equal to the initial stiffness matrix K0, then at t= 0.2 s the stiffness matrix\nsuddenly changes to K= 90%×K0, att= 0.4 s there is another sudden degrada-\ntion toK= 80%×K0,..., finally, at t= 1 s,K= 50%×K0. The time histories\nof the damping ratios for methods a,bandcare shown in figure 3for uniform\nstiffness degradations (yielding scenario 1) and in figure 4for nonuniform stiffness\ndegradations (yielding scenario 2).\nFrom figures 3and4, the best control on the damping ratios time histories is\nobtainedformethod c,althoughthedampingratioofmode5significantlyincreases\nfor the structure with nonuniform stiffness degradations. In par ticular, and as\nexpected for method c, modes 1 and 3 are perfectly controlled: in the ( ξ,t)-plane\nin figures 3[bottom, right] and 4[bottom, right], ξ1(t) andξ3(t) both describe the\nsame straight line ξ1(t) =ξ3(t) =ξ0. We recall that method cis rarely used in\npractice. Itisalsoobviousthat, inthisparticularcasewheretheRa yleighdamping\nmodels are designed according to the initial structural modal prop erties, method\nbprovides better control on damping than method a. For both methods aand\nb, the damping models designed lead to overdamped first modes espec ially when\nnonuniform stiffness degradations are assumed. This is not accept able because\nfirst mode generally is important to reconstruct the overall struc tural behavior. In\nthe case of uniform stiffness degradations, one can also observe in figure3[top,\nright] that ξ1(t) andξ3(t) describe the same curve and consequently have the same\ntime history.\nIn table2, we gather the time history of the modal properties of the struct ure\nwith nonuniform stiffness degradations. In particular, this table sh ows that the\nfactorha\nmcan become significantly large comparing to their initial unit value.\n4.2.Example 2: Design of optimal Rayleigh damping models. In this ex-\nample, thesamestructureasforexample1isconsidered, butonlyw ithnonuniform\nstiffness degradations (yielding scenario 2), which is the more realist ic case.RAYLEIGH DAMPING IN INELASTIC ANALYSES 17\n051015202530354000.511.522.5\nωξa/ξ0\n0 0.2 0.4 0.6 0.8 100.511.522.5\ntξa/ξ0\n051015202530354000.511.522.5\nωξb/ξ0\n0 0.2 0.4 0.6 0.8 100.511.522.5\ntξb/ξ0\n051015202530354000.511.522.5\nωξc/ξ0\n0 0.2 0.4 0.6 0.8 100.511.522.5\ntξc/ξ0\nFigure 3. Time histories of the damping ratios associated to the\nfive modes of the structure with uniform stiffness degradations (s ce-\nnario 1). ×is used for mode 1, ◦for mode 2, /squarefor mode 3, ⋆for\nmode 4, and ⋄for mode 5. Time histories are illustrated in ( ξ,ω)- and\n(ξ,t)-planes. For method b[center], the curve ξb(ω) is the same for all\nmodes. Conversely, for methods a[top] and c[bottom], ξ(ω) depends\non the mode (the ×’s,◦’s,...all belong to a different trajectory in the\n(ξ,ω)-plane). The curves plotted in the ( ξ,ω)-plane are commonly used\nin the literature to illustrate Rayleigh damping. The dashe d lines are\nplotted at t= 0.\nWhen choosing ωA,ωB,ξAandξB, the practitioner has to seek for the best con-\ntrol on all the damping ratiospertaining to the most important mode s, throughout\nthe ITHA. When structure yields, the modal frequencies decreas e, which can lead\nto overdamping modes, especially for the fundamental one which ca n increase a\nlot from its initial value to its actual value after yielding (see figure 4where it\nincreases by a factor of up to 2 .5). In this second example, we present how to18 RAYLEIGH DAMPING IN INELASTIC ANALYSES\n051015202530354000.511.522.53\nωξa/ξ0\n0 0.2 0.4 0.6 0.8 100.511.522.53\ntξa/ξ0\n051015202530354000.511.522.53\nωξb/ξ0\n0 0.2 0.4 0.6 0.8 100.511.522.53\ntξb/ξ0\n051015202530354000.511.522.53\nωξc/ξ0\n0 0.2 0.4 0.6 0.8 100.511.522.53\ntξc/ξ0\nFigure 4. Time histories of the damping ratios associated to the\nfive modes of the structure with nonuniform stiffness degradation s\n(scenario 2). Further details provided in the caption of figure 3also\napply here.\ndesign optimal Rayleigh damping models with methods a(ord) andbto provide\nthe best control possible on the damping ratios time histories.\nTo that purpose, and according to the analytical developments pr esented in the\nprevious section, we now design Rayleigh damping models accounting f or stiffness\ndegradations:\n(i) As above, we consider that assigning a damping ratio to modes 1 an d 3\nis relevant to control the damping ratio of the most important mode s and\nthus setmA= 1andmB= 3. Weremarkthat ω1andω3necessarily remain\nin the range [ ω1(T),ω3(0)] throughout ITHA because ω1(T)< ω1(0) and\nω3(0)> ω3(T) due to stiffness degradations, where Tis the total duration\nof the simulation. Note that the quantities at time Tarea priori not\nknown and need some preliminary analysis results to be efficiently chos en;RAYLEIGH DAMPING IN INELASTIC ANALYSES 19\nTable 2. Time history of the eigenfrequencies [rad.s−1] andha\nm(t)\n[-] factors for the structure with nonuniform stiffness degradat ions\n(see also figure 4).\ntω1ha\n1ω2ha\n2ω3ha\n3ω4ha\n4ω5ha\n5\n0.05.56 1.00 16.23 1.00 25.58 1.00 32.87 1.00 37.49 1.00\n0.25.17 1.16 15.42 1.11 24.34 1.11 31.27 1.11 35.87 1.09\n0.44.72 1.41 14.49 1.28 22.90 1.27 29.45 1.26 34.42 1.16\n0.64.19 1.84 13.37 1.56 21.18 1.54 27.42 1.46 33.15 1.22\n0.83.51 2.85 11.94 2.13 19.05 2.00 25.29 1.68 32.02 1.27\n1.02.39 8.10 9.81 3.82 16.41 2.75 23.18 1.89 31.00 1.31\nhereafter, wewillusetheresultspreviously obtainedwithexample1 . Then,\nthe choice of tAandtBdepends on the method used:\n–Formethod a: accordingtofigure 4,ξa\n1(t)andξa\n3(t)constantlyincrease\nthroughout the analysis. Consequently, the maximum damping ratio s\nfor modes 1 and 3 will be observed at time t=Tand this is why we\nsettA=tB=T. Then, ωA=ω1(T) = 2.39 rad.s−1,ωB=ω3(T) =\n16.41 rad.s−1(R= 6.87),hA=h1(T) = 8.10, andhB=h3(T) = 2.75;\n–For method b: we set tA=TandtB= 0, because, according to\nfigure4,ξ1,max=ξ1(T) andξ3,max=ξ3(0). Then, ωA=ω1(T) =\n2.39 rad.s−1andωB=ω3(0) = 25.58 rad.s−1(R= 10.70).\n(ii) With a targeted damping ratio in the range [ ωA,R ωA] (R ωA=ωB) of\nˆξ= 2%, we compute:\n–For method a: ∆a\nmax= 1.06% from equation ( 27). ∆a\nmaxis an upper\nbound and a lower value is therefore likely to be more appropriate.\nNevertheless, we take ξmax=ξ0=ξA=ξB= 2% + 1 .06% = 3 .06%,\nwhich has to be validated once the analysis has been run;\n–For method b: ∆b= 0.57% from equations ( 23) and we set ξmax=\nξ0=ξA=ξB= 2%+0 .57% = 2.57%.\nThe histories of the damping ratios resulting from the Rayleigh dampin g models\ndesigned according to the procedure presented just above are s hown in figure 5\nwhen method ais used and in figure 6for method b. For the sake of comparison,\nthedampingratiostimehistoriesshowninfigure 4,viz.whenstiffnessdegradations\nare not accounted for in the damping models design, are reproduce d in the top\nhalf of figures 5and6.\nAccording to figure 5, one can make the following observations for method a:\n(i) The damping ratios pertaining to the first three modes remain in th e range\n[1.11%,2.98%] which is, as expected after the analytical developments in\nthe previous section, included in the range [0 .94%,3.06%] = [ ˆξ−∆a\nmax,ˆξ+\n∆a\nmax]. The choice of ξa\nmax=ξa\n1(T) =ξa\n3(T) thus is validated.20 RAYLEIGH DAMPING IN INELASTIC ANALYSES\n(ii)ˆξ= 2% is well centered in [1 .11%,2.98%], which means that the choice of\nξa\nmax=ξa\nA=ξa\nB=ˆξ+∆a\nmaxwas satisfying.\n(iii) Control on the damping ratios is significantly improved when stiffne ss\ndegradations are anticipated, especially for the first mode.\n(iv) Althoughmodes4and5areoutoftheselectedcontrolrange[ ω1(T),ω3(T)],\nthey are less overdamped than when the damping model does not ac count\nfor stiffness degradations.\n051015202530354000.511.522.5No anticipation of stiffness degradations\nωξa/ˆξ\n0 0.2 0.4 0.6 0.8 100.511.522.5\ntξa/ˆξ\n051015202530354000.511.522.5With anticipation of stiffness degradations\nωξa/ˆξ\n0 0.2 0.4 0.6 0.8 100.511.522.5\ntξa/ˆξ\nFigure 5. Structure with nonuniform stiffness degradations. [bot-\ntom] Rayleigh damping model of type awith anticipation of stiffness\ndegradations. [top]Forthesakeofcomparison, thefiguresrepr oduce\nthe results shown in figure 4where stiffness degradations are not ac-\ncounted for in the design of the damping model.\nLooking now at figure 6, we can remark the following for method b:\n(i) The damping ratios pertaining to the first three modes remain in th e range\n[1.47%,2.57%]whichis, asexpected, includedintherange[1 .43%,2.57%] =\n[ˆξ−∆b,ˆξ+∆b]. Wecoulddecrease ξb\nmaxof0.02%tohave ˆξperfectlycentered\nin [1.47%−0.02%,2.57%−0.02%] = [1 .45%,2.55%]: the correction isminorRAYLEIGH DAMPING IN INELASTIC ANALYSES 21\nin this example because the minimum of the ξb(ω) curve is almost reached\nbyξb\n3(T) (see figure 6[bottom])5.\n(ii) Control on the first damping ratio is significantly improved when st iffness\ndegradations are anticipated.\n(iii) Higher modes are overdamped. We could increase Rto have a better\ncontrol on these modes too, but this would increase ∆band consequently\nalter the control on the first three modes.\n051015202530354000.511.522.5No anticipation of stiffness degradations\nωξb/ˆξ\n0 0.2 0.4 0.6 0.8 100.511.522.5\ntξb/ˆξ\n051015202530354000.511.522.5With anticipation of stiffness degradations\nωξb/ˆξ\n0 0.2 0.4 0.6 0.8 100.511.522.5\ntξb/ˆξ\nFigure 6. Structure with nonuniform stiffness degradations. [bot-\ntom] Rayleigh damping model of type bwith anticipation of stiffness\ndegradations. [top]Forthesakeofcomparison, thefiguresrepr oduce\nthe results shown in figure 4where stiffness degradations are not ac-\ncounted for in the design of the damping model.\nFinally, comparing figure 5[bottom] to 6[bottom], we remark that:\n(i) A Rayleigh damping model designed with method awith anticipation of\nstiffness degradations can be more efficient than designed with meth odb\nwithout accounting for stiffness degradations. In particular, it av oids the\nstrong increase of the damping ratio pertaining to the first mode.\n5Such a simple a posteriori correction of ξmaxis possible here because hA=hB= 1 and\nξA=ξB.22 RAYLEIGH DAMPING IN INELASTIC ANALYSES\n(ii) As far as the first three modes are concerned, the control pr ovided by\nmethodbis better: [1 .43%,2.57%]⊂[1.11%,2.98%].\n(iii) Better control on the modes 4 and 5 is observed with method a. However,\nRayleigh damping models were not designed to control those two mod es.\nFormethod b, onecouldhavebettercontrolonmodes4and5, byincreasing\nωB, and soR, until ∆breaches the actual ∆a= 0.98%.\n4.3.Some additional guidelines for practical use. As illustrated above, it is\nnecessary to have some hint of the overall nonlinear structural b ehavior to design a\nRayleighdampingmodelthatefficientlycontrolsthedampingratiotime historyfor\nthe structural modes of interest. This issue is inherent to every R ayleigh damping\nmodel in ITHA. In practice, the time history of the modal propertie s are however\nnot known a priori and designing optimal Rayleigh damping model is thus not\nstraightforward, especially with method a, and might need some iterations. The\nfollowing practical guidelines can therefore be useful:\n(i) It isrecommended in[ 14]thatωAbethefirstmodeand ωBthelowest mode\nfor which the cumulative effective mass exceeds 90%-95%of the tot al mass.\n(ii) The following possible methods are recommended in the user manua l of\nthe computer program PERFORM-3D [ 8,§18.2] to define ωAandωB:\n(28) ( i)/braceleftbigg\nωA= 1.10ω1(0)\nωB= 4.00ω1(0)or (ii)/braceleftbigg\nωA= 1.10ω1(0)/√µ\nωB= 0.85ω1(0),\nwhereµis the ductility of the structure. Method (i) is the same as for\na linear analysis and method (ii) is particularly effective for structure s\ndominated by their first mode.\n5.Conclusions\nFrom the review of the literature proposed in the second section of this paper, it\nis obvious that some researchers or practitioners advocate using the initial stiffness\nmatrix in the design of Rayleigh damping models, whereas others advo cate using\nthe tangent stiffness matrix. Controlling the damping ratios throug hout inelastic\ntime history analyses is an important issue to avoid the appearance o f spurious\ndamping forces. To that purpose, useful analytical formulas are developed in\nthe third section of this paper for both Rayleigh models based on initia l and\ntangent stiffness. Whereas there exists a simple relation that allows controlling\nthe damping ratios time histories when tangent stiffness is used, the re is no such\nrelation when initial stiffness is used and controlling damping ratios is, a lthough\nachievable, not straightforward.\nFrom these latter analytical relations and from the examples shown in section\n4, we can conclude that it is easier to design a Rayleigh damping model w ith well-\ncontrolled damping ratiostime histories throughout the inelastic ana lysis when the\ntangent stiffness is used than with the initial stiffness. However, co ntrolling theRAYLEIGH DAMPING IN INELASTIC ANALYSES 23\ndamping ratios with initial stiffness-based damping models can be achie ved. That\nis why, when convergence difficulties are experienced with the solutio n algorithm,\ninitialstiffnesscanbeusedwithoutnecessarily leadingtoveryhighda mpingratios.\nWhether initial ortangent stiffness isused inthedesignoftheRayleig hdamping\nmodel used for inelastic time history seismic analysis, we advocate th at:\n(i) Figures like figures 5and6in this paper be plotted to keep control on\nthe damping ratios time histories. To that purpose, the required an alytical\nrelations are summarized in table 1of this paper. These later relations also\nhelp improving control on the damping ratios;\n(ii) Thedampingforcesbecomputedandcomparedtotheotherres istingforces\nso as to guarantee there are no spurious damping forces generat ed in the\nsystem. Indeed, even if the damping ratios are well-controlled thro ughout\nthe inelastic time history seismic analysis, Rayleigh damping still lacks\nphysical evidence and there is no guarantee that the actual damp ing forces\nare properly modeled.\nConsidering the difficulty to control Rayleigh damping in inelastic struc tures\nthat is illustrated in this paper, strategies that rely on using models w ith nonlin-\nearities expected to develop in clearly identified structural parts w ith the other\nparts remaining elastic appear as promising. Such strategies are me ntioned in the\nsecond section of this paper but are out of the scope of the develo pments that\nfollow in sections 3 and 4. However, there are some obvious argumen ts that would\nmotivate further investigations for this recent class of methods in future work: i)\nin elastic parts, there is no concerns about damping ratio shift; ii) a r educed – in\ncomparison to the number of structural DOFs – eigenbasis could be defined ( e.g.\nwith the procedures proposed in [ 13] and [12]), which would increase computa-\ntional efficiency; iii) this would simplify the damping model design becaus e the\nnumber of modes, which a proper damping ratio has to be associated to, would be\nreduced; iv) the inelastic parts could be separately treated with pr actical methods\nspecifically developed to avoid the problems likely to be encountered w hen using\nRayleigh damping.\nAcknowledgements This research was supported by a Marie Curie Interna-\ntional Outgoing Fellowship within the 7th European Community Framew ork Pro-\ngramme (proposal No. 275928). The financial support provided b y the Fond\nQu´ eb´ ecois de Recherche sur la Nature et les Technologies (FQRNT ), the Natural\nScience and Engineering Research Council of Canada (NSERC), and the ENS-\nCachan Invited Professor Program also are gratefully acknowledg ed.\nReferences\n[1] Quantification of building seismic performance factors. Technica l Report FEMA P695, Fed-\neral Emergency Management Agency, Washington, DC, June 2009 .24 RAYLEIGH DAMPING IN INELASTIC ANALYSES\n[2] Modeling and acceptance criteria for seismic design and analysis of tall buildings. Techni-\ncal Report PEER 2010/111 or PEER/ATC-72-1, Pacific Earthquak e Engineering Research\nCenter, Richmond (CA), October 2010.\n[3] Dionisio Bernal. Viscous damping in inelastic structural response. ASCE Journal of Struc-\ntural Engineering , 120(4):1240–1254, April 1994.\n[4] Athol J. Carr. Ruaumoko manual, Volume 2: User manual for the 2-dimensional version\nRuaumoko2D. Technical report, University of Canterburry, Chr istchurch, New Zealand,\nNovember 2008.\n[5] Finley A Charney. Unintended consequences of modeling damping in structures. Journal of\nStructural Engineering , 134(4):581–592, April 2008.\n[6] D.J. Chrisp. Damping models for inelastic structures. Master’s th esis, University of Canter-\nbury, Christchurch, New Zealand, 1980.\n[7] Ray W Clough and Joseph Penzien. Dynamics of structures . McGraw-Hill, Inc., 1975.\n[8] CSI. Perform3D User’s manual. Technical report, California, 20 07.\n[9] Emrah Erduran. Evaluation of Rayleigh damping and its influence on engineering demand\nparameterestimates. Earthquake Engineering and Structural Dynamics , 41:1405–1919,2012.\n[10] J F Hall. Problems encountered from the use (or misuse) of Rayle igh damping. Earthquake\nEngineering and Structural Dynamics , 35:525–545, 2006.\n[11] Adnan Ibrahimbegovic. Non-linear solid mechanics: theoretical formulations and finite ele-\nment solution methods . Springer, Berlin, 2009.\n[12] Adnan Ibrahimbegovicand EdwardL Wilson. Simple numerical algor ithms for the mode su-\nperposition analysis of linear structural systems with non-propor tionaldamping. Computers\nand Structures , 33(2):523–531, 1989.\n[13] Adnan Ibrahimbegovic and Edward L Wilson. Efficients solution pro cedures for systems\nwith local non-linearities. Engineering Computations , 9:385–398, 1992.\n[14] Pierre L´ eger and Serge Dussault. Seismic-Energy Dissipation in MDOF Structures. ASCE\nJournal of Structural Engineering , 118(6):1251–1267, May 1992.\n[15] A. D. Nashif, D. I. G. Jones, and J. P. Henderson. Vibration damping . John Wiley Inter-\nscience Publications, New York, 1985.\n[16] P. B. Shing and S. A. Mahin. Elimination of spurious higher-mode re sponse in pseudody-\nnamic tests. Earthquake Engineering and Structural Dynamics , 15:425–445, 1987.\n[17] Farzin Zareian and Ricardo A Medina. A practical method for pro per modeling of structural\ndamping in inelastic plane structural systems. Computers and Structures , 88:45–53, 2010." }, { "title": "2311.12357v2.Electrodynamics_with_violations_of_Lorentz_and_U_1__gauge_symmetries_and_their_Hamiltonian_structure.pdf", "content": "arXiv:2311.12357v2 [hep-th] 10 Apr 2024Electrodynamics with violations of Lorentz and U(1) gauge s ymmetries and their\nHamiltonian structure\nXiu-Peng Yanga,b,1,∗Bao-Fei Lia,b,1,†and Tao Zhua,b1,‡\n1aInstitute for Theoretical Physics and Cosmology,\nZhejiang University of Technology, Hangzhou, 310032, China\nbUnited Center for Gravitational Wave Physics (UCGWP),\nZhejiang University of Technology, Hangzhou, 310032, China\nThis article aims to study the Lorentz/U(1) gauge symmetry- breaking electrodynamics in the\nframework of the Standard-Model Extension and analyze the H amiltonian structure for the the-\nory with a specific dimension d≤4 of Lorentz breaking operators. For this purpose, we con-\nsider a general quadratic action of the modified electrodyna mics with Lorentz/gauge-breaking\noperators and calculate the number of independent componen ts of the operators at different di-\nmensions in gauge invariance and breaking. With this genera l action, we then analyze how the\nLorentz/gauge symmetry-breaking can change the Hamiltoni an structure of the theories by con-\nsidering Lorentz/gauge-breaking operators with dimensio nd≤4 as examples. We show that the\nLorentz-breaking operators with gauge invariance do not ch ange the classes of the constraints of the\ntheory and the number of the physical degrees of freedom of th e standard Maxwell’s electrodynam-\nics. When the U(1) gauge symmetry-breaking operators are pr esent, the theories in general lack\nfirst-class constraint and have one additional physical deg ree of freedom, compared to the standard\nMaxwell’s electrodynamics.\nI.Introduction\nThe Standard Model (SM) successfully describes the\nfundamental constituents of matter using quarks, lep-\ntons, gauge bosons, and Higgs bosons. It effectively\nexplains phenomena involving elementary particles and\ntheir interactions. The last predicted elementary parti-\ncle of the SM, the Higgs boson, was experimentally con-\nfirmed in 2012 [ 1,2], giving a happy ending to the de-\nvelopment of this great theory. General Relativity (GR)\nlinks gravity with the curvature of spacetime and pro-\nvides a highly successful description of gravity-related\nphenomena. Since itsinception, it hasbeen recognizedas\na great theory and continues to be corroborated through\nvarious experiments. Its basic principles and predictions\nhave been confirmed through multiple observations and\nmeasurements, establishing its foundational position in\nmodern physics. In 2016, gravitational waves were ob-\nserved in experiments [ 3], which established the great\nposition of general relativity. These two theories are ex-\npected to be unified at the Planck scale and potentially\nexhibit observable quantum gravity effects at accessible\nlow-energy scales. This signal may be related to Lorentz\nsymmetry breaking and can be described by an effective\nfield theory [ 4].\nTo construct a consistent effective field theory that in-\ncorporates both GR and SM, Kosteleck´ y and Colladay\nproposed an effective field theory called the Standard\nModel Extension (SME) [ 5,6], which features general\nLorentz and CPT violations. The measured and derived\n∗youngxp@zjut.edu.cn\n†libaofei@zjut.edu.cn\n‡corresponding author: zhut05@zjut.edu.cnvalues of coefficients for Lorentz and CPT violations in\nthe SME can be found in the data table organized by\nKosteleck´ y and Russell[7]. In recent years, research\non the Cosmic Microwave Background (CMB) radiation\nand the ultra high energy cosmic rays (UHECR) have\nprovided new opportunities for studying the pure pho-\nton sector of the SME. One reason is that any predic-\ntion involving the pure photon part that deviates from\nSM could potentially indicate Lorentz violation originat-\ning from the pure photon sector of the SME. A large\namount of research has been conducted in this area [ 8–\n23]. Compared to conventional Maxwell electrodynam-\nics, the pure-photon sector of the SME includes addi-\ntional Lorentz-breaking terms, which can be classified\nas CPT-even and CPT-odd. The inclusion of these\nterms leads to the emergence of new effects, which has\nspurred extensive research in this field. Studying the\ngeneral aspects of the pure photon sector is a challeng-\ning task, Ref. [ 24] proposed a general electrodynamics\nextension theory with a quadratic action, which can be\nused to describe many related studies , including pho-\nton interactions [ 25], optical activity of media [ 26], the\nLorentz-invariance-Violating term [ 27,28], the Chern-\nSimons term [ 29], nonminimal SME [ 30], and other re-\nlated phenomena involving photons [ 31]. Theories with\nLorentz-breaking operators of dimension 4 have received\nconsiderable attention, including the Lorentz-invariance-\nviolating(LIV)[ 32], Carroll-Field-Jackiw(CFJ)[ 33], and\nProca electrodynamics [ 34], with the first two being U(1)\ngauge invariant theories and the last one involving U(1)\ngauge symmetry breaking. In Refs. [ 27,28,35,36], they\nconducted Hamiltonian analyses of the aforementioned\nthree theories and identified the constraint structures of\nthe theories, which motivates us to study the constraint\nstructure of more general theories with a specific dimen-\nsiond≤4.2\nThe Lagrangian density of quadratic electrodynamics,\ngiven in [ 24], is a quadratic polynomial in the photon\nfieldAα1and its higher-order derivatives ∂α3...∂αdAα2\nwithd≥2.Since the constant coefficients Kα1α2α3...αd\n(d),\nwhich are contracted with Aα1∂α3...∂αdAα2, remain in-\nvariant under coordinate transformations, it leads to a\nviolation of the Lorentz symmetry of the theory . These\nconstant coefficients can be regarded as originating from\nthe vacuum expectation value of an operator in the un-\nderlying theory, or the dominant component of dynamic\nbackground fields, or an averaged effect. Through di-\nmensional analysis, it can be found that the constants\nKα1α2α3...αd\n(d)associated with Aα1∂α3...∂αdAα2which is\nof dimension dmust have dimension 4 −d. Some re-\nsearchers believe that the theories with power series that\nare renormalizable have mass dimension d≤4 [24,37],\nand it is these theories with mass dimension d≤4 that\nare mainly studied in this work. Usually they are also\nthose theories that contain only the first time deriva-\ntive of the field Aµ, which makes the theories healthy\nas this avoids the potential Ostrogradsky instability that\nthe theory could possess [38,39].\nIn this work, we will extend the U(1) gauge-invariant\nLagrangian density of quadratic electrodynamics de-\nscribed in [ 24] to the one that includes U(1) gauge sym-\nmetry breaking terms. We also perform Hamiltonian\nanalysis for renormalizable specific dimension cases of\nd≤4. Our purpose is to clarify how the U(1) gauge\nbreaking terms affect the constraint structure and phys-\nical degrees of freedom of the theory.\nOur results indicate that the Lagrangian density of\ngeneral quadratic electrodynamics is the combination of\nfive terms. One corresponds to the Lagrangian density\nofstandard electrodynamics, one is U(1) gaugeinvariant,\nonecontainsbothU(1)gaugeinvariantandbreaking,and\nthe remaining two are U(1) gauge breaking. The Hamil-\ntoniananalysisofthegeneralgauge-breakingsystemwith\na specific dimension d≤4 reveals that the system with\nd= 4 requires additional conditions to become a con-\nstrained system, while systems with d= 2 and d= 3 do\nnot. It also shows that the Lorentz-breaking operators\nwith gauge invariance do not change the classes of the\nconstraints of the theory and the number of the phys-\nical degrees of freedom of the standard Maxwell’s elec-\ntrodynamics. When the U(1) gauge symmetry-breaking\noperators are presented, the theories in general lack first-\nclass constraint and have one additional physical degree\nof freedom, compared to the standard Maxwell’s electro-\ndynamics.\nThestructureofthis paperis asfollows. The basicthe-\nory is discussed in Sec. II, where we give the number of\nindependent components of the Lorentz-breaking opera-\ntors with U(1) gauge violation and extend it to the gen-\neral Lagrangian density containing U(1) gauge-breaking\nterms. Sec. IIIis about the Hamiltonian structure and\ndegrees of freedom of theories with a specific dimension\nd≤4. In Sec. IV, We apply the obtained results to some\nspecific models and derive the results for LIV, CFJ, andProca electrodynamics. Our summary is in Sec. V.\nFor the sake of clarity and conciseness, we will use two\nconventions:\n1.The Greek indices range from 0 to 3 and the Latin\nindices run from 1 to 3. The metric of the back-\nground spacetime ηµν≡(1,−1,−1,−1);\n2.The time argument of the vector field Aµis\nsuppressed throughout the manuscript, namely\nAµ(x)≡Aµ(t,x).\nII.Electrodynamics with violations of Lorentz and\nU(1) gauge symmetry\nIn this section, we present a brief introduction of the\nelectrodynamics with quadratic action in the pure pho-\nton sector in the framework of SME with both Lorentz\nand U(1) gauge symmetry breaking. This represents\nan extension of the Lorentz-violating modified electro-\ndynamics with U(1) gauge invariance in [ 24] by includ-\ning the U(1) gauge symmetry-breaking operators in the\nquadratic action. In the construction of this theory, we\nalso analyze the properties of the coefficients for Lorentz-\nand U(1) gauge-violating operators.\nForourpurpose,followingthesimilarconstructionper-\nformed in [ 24], let us start with the general quadratic\naction for the Lorentz-violating electrodynamics in the\npure photon sector, which can be written as [ 24]\nS=/integraldisplay\nd4xL (1)\nwith\nL=−1\n4FµνFµν+∞/summationdisplay\nd=2Kα1α2α3...αd\n(d)Aα1∂α3...∂αdAα2,\n(2)\nwhereFµν≡∂µAν−∂νAµandKα1α2α3...αd\n(d)are con-\nstant coefficients with mass dimension 4 −d. One possi-\nbleexplanationforthe coefficients Kα1α2α3...αd\n(d)originates\nfrom non-zero vacuum expectation values to the Lorentz-\ntensorfields. Note that each term associatedwith the co-\nefficientKα1α2α3...αd\n(d)violates CPTif dis odd orpreserves\nCPT ifdis even.\nThe symmetry among indices {α3,...,α d}of tensor\n∂α3...∂αdand the use of integration by parts results in\ntwo properties of the coefficients Kα1α2α3...αd\n(d). One is the\ntotal symmetry in the d−2 indices {α3,···,αd}, another\nis the symmetry of the two indices {α1,α2}whendis\neven and antisymmetry when dis odd. Depending on the\nspecific intrinsic symmetries of the Lorentz-violating op-\nerators, the coefficients Kα1α2α3...αd\n(d)can be decomposed\ninto five representations [ 24]. When one imposes the\nconditions of the U(1) gauge invariance on the Lorentz-\nviolatingoperators,thesefiverepresentationsarereduced3\ninto two representations, one corresponds to the CPT-\neven coefficient and the other corresponds to CPT-odd\ncoefficient [ 24].\nA.Lorentz-violating electrodynamics with U(1)\ngauge invariance\nLet us first consider the Lorentz-violating electrody-\nnamics with U(1) gauge invariance [ 24]. TheU(1) gauge\ninvariance is a symmetry of the theory under the U(1)\ngauge transformation\nAµ→˜Aµ=Aµ+∂µΛ, (3)\nwhere Λ is an arbitrary function. The variation of the\naction (1) under this gauge transformation reads\nδgS=−∞/summationdisplay\nd=2/integraldisplay\nd4xKα1α2α3···αd\n(d)Λ\n×∂α3···∂αd/parenleftbigg\n∂[α1Aα2]±+1\n2∂[α1∂α2]±Λ/parenrightbigg\n= 0,\n(4)\nwhere “ + /−” corresponds to even/odd dimension,\nand the brackets [] +and []−indicate symmetrization and\nantisymmetrization respectively . Specifically, the terms\nin the big bracket of the above expression can be written\nas\n∂[α1Aα2]±+1\n2∂[α1∂α2]±Λ\n=/braceleftBigg\n∂α1Aα2+∂α2Aα1+∂α1∂α2Λ, dis even,\n∂α1Aα2−∂α2Aα1, d is odd.(5)\nTheU(1) gauge invariance requires the variation of the\naction in ( 4) vanishes. When dis even, since the two first\nindicesα1andα2are symmetric, in order to make ( 4)\nvanishes, one has to require both indices α1andα2are\nantisymmetric with one of {α3,α4,···,αd}[24]. With\nthese properties, it is straightforward to infer that all\nthe CPT-even operators with d= 2 are gauge-violating\nand the gauge invariance requires d≥4. By using these\nproperties, one canalsocountthe number ofindependent\ncomponents of the CPT-even operators, which leads to\nNF= (d+1)d(d−3). (6)\nWhendis odd, similarly, the first two indices α1andα2\nare antisymmetric, and the vanishing of ( 4) require α1\nandα2are antisymmetric with one of {α3,α4,···,αd}.\nFor this case, only operators with d≥3 are allowed and\nthe number of independent components of the CPT-odd\noperators is\nNAF=1\n2(d+1)(d−1)(d−2). (7)\nNote that in Table. I, we summarize the number of in-\ndependent components for the CPT-odd and CPT-even\noperators.B.Lorentz-violating electromagnetics with the\nviolations of U(1)gauge symmetry\nNow let us turn to consider the case with the break-\ning of the U(1) gauge symmetry. When the U(1) gauge\nsymmetry is violated, one does not require the variation\nof the action in ( 4) vanishing. For this case, one does\nnot need to impose extra conditions on the coefficients\nKα1α2α3...αd\n(d)to ensure the gauge invariance of the theory.\nFor CPT-even operators without gauge invariance, as\nwe mentioned before, while the indices {α3,α4,···,αd}\nin the CPT-even coefficients Kα1α2α3...αd\n(d)are symmetric,\nfirst two indices α1andα2are also symmetric. Simi-\nlarly, for CPT-odd operators without gauge invariance,\nthe indices {α3,α4,···,αd}are symmetric, while the\nfirst two indices α1andα2are antisymmetric. One can\nthen count the numberofindependent components ofthe\nCPT-evenand CPT-oddoperatorswith a specific dimen-\nsiond, which are summarized in Table. I.\nTABLE I: The number of independent components of\nthe Lorentz-violatingoperatorswith/without U(1) gauge\ninvariance.\ndCPTGauge Number of\ninvariance independent components\n2 even no 10\neven,\n≥4evenyes ( d+1)d(d−3)\nno2\n3(d+2)(d+1)d\nodd,\n≥3oddyes1\n2(d+1)(d−1)(d−2)\nno1\n2(d+2)(d+1)(d−1)\nFor later convenience, one can decompose the La-\ngrangian density ( 2) into two parts, the U(1) gauge in-\nvariance part and the gauge-violating part. Specifically,\nthis involves rewriting the Lorentz-breaking coefficients\nin five different forms with distinct symmetries. The spe-\ncific representation decomposition of these five forms can\nbe found in [ 24]. Then, the Lagrangian density ( 2) can4\nbe rewritten as\nL=−1\n4FανFαν\n+∞/summationdisplay\nevend=2K(1)α1α2α3...αd\n(d)Aα1∂α3...∂αdAα2\n+∞/summationdisplay\nd=3K(2)α1µνα2...αd−2\n(d)Aα1∂ν∂α2...∂αd−2Aµ\n+∞/summationdisplay\nd=3K(3)µα1να2...αd−2\n(d)Aµ∂ν∂α2...∂αd−2Aα1\n+∞/summationdisplay\nd=4K(4)µνρσα 1...αd−4\n(d)Aµ∂ρ∂σ∂α1...∂αd−4Aν\n+∞/summationdisplay\noddd=3K(5)µνρα1...αd−3\n(d)Aµ∂ρ∂α1...∂αd−3Aν,\n(8)\nwhere the five coefficients K(i)α1α2α3···αd\n(d)withi=\n1,2,3,4,5 are distinguished by their distinct symmetries\namong the indices {α1,α2,α3,···,αd}. The coefficients\nK(1)α1α2α3···αd\n(d)are all CPT-even coefficients with d≥2,\nwhiletheirindices {α1,α2,α3,···,αd}areallsymmetric.\nIt is obvious to infer from ( 4) that these coefficients vio-\nlate theU(1) gauge symmetry of the theory. For the co-\nefficients K(2)α1µνα2···αd−2\n(d), they are be either CPT-even\nor CPT-odd with d≥3. Except the two indices µand\nνare antisymmetric, the rest indices {α1,α2,···,αd−2}\nare all symmetric. Note that these coefficients are also\ngauge-violating. Similarly, for K(3)µα1να2···αd−2\n(d), they can\nalso be either CPT-even or CPT-odd with d≥3. The\ntwo indices of this coefficient µandνare antisymmetric\nandtherestindices {α1,α2,···,αd−2}areallsymmetric.\nOne can check that these coefficients also break the U(1)\ngauge symmetry of the theory. Then for the indices in\nthe coefficients K(4)µνρσα 1...αd−4\n(d), the two indices µand\nρ, and the two indices νandσ, are both antisymmet-\nric. We also note that these coefficients are symmetric\nupon interchange of the two pairs of indices ( µ,ρ) and\n(ν,σ). By inspecting the variation of the action ( 4) with\nthese coefficients, one can infer that the CPT-odd oper-\nators with K(4)µνρα1···αd−4\n(d)violates the gauge invariance,\nwhile the CPT-even ones are gauge invariance. The last\ncoefficients K(5)µνρα1...αd−3\n(d)are CPT-odd coefficient with\nd≥3, andtheir threeindices {µ,ν,ρ}areantisymmetric.\nThese coefficients preserve the U(1) gauge symmetry of\nthe theory.\nTo simplify the laterhandling of the Hamiltonian anal-\nysis of the theory, one can rewrite the Lagrange density\nof the theory in a compact form of\nL=−1\n4FανFαν+Aα1ˆK(1)α1α2Aα2+1\n2Aα1ˆK(2,3)µνα1Fνµ\n−1\n4FµρˆK(4)µρνσFνσ+1\n2ǫκµνρAµˆK(5)\nκFνρ, (9)where\nˆK(1)α1α2≡/summationdisplay\nd=evenK(1)α1α2α3...αd\n(d)∂α3...∂αd,(10)\nˆK(4)µρνσ≡∞/summationdisplay\nd=4K(4)µνρσα 1...αd−4\n(d)∂α1...∂αd−4,(11)\nˆK(5)\nκ≡ǫκµνρ\n6/summationdisplay\nd=oddK(5)µνρα1...αd−3\n(d)∂α1...∂αd−3,\n(12)\nand\nˆK(2,3)µνα1≡\n∞/summationdisplay\nd=3/bracketleftBig\nK(2)α1µνα2...αd−2\n(d)+(−1)dK(3)µα1να2...αd−2\n(d)/bracketrightBig\n×∂α2...∂αd−2. (13)\nThe indice symmetries of the operators in ( 9) are shown\nin Table. II. The indices enclosed in the same braces\n{}of the second and third columns represent symmetry\nand antisymmetry between them, respectively. {µρ,νσ}\nin the second column indicates that the corresponding\noperators are symmetry when the two pairs ( µ,ρ) and\n(ν,σ) are exchanged. The fourth column displays the\nconditions under which each class of operators appears.\nTABLE II: Symmetries of the indices of the coefficients\nof operators in ( 9).\nCoefficient Symmetry Antisymmetry d\nˆK(1)α1α2{α1,α2} ··· even,≥2\nˆK(2,3)µνα1 ··· { µ,ν} ≥ 3\nˆK(4)µρνσ{µρ,νσ} { µ,ρ},{ν,σ} ≥ 4\nˆK(5)\nκ ··· ··· odd,≥3\nIII.Hamiltonian structure of the theory\nInthissection, weperformHamiltoniananalysisonthe\nLorentz-violatingelectromagneticwith and without U(1)\ngaugeinvariancebyusingthe Dirac-Bergmannprocedure\n[40–44]. For simplicity, we will focus on the theory with\nLagrangian densities with a specific dimension d≤4 of\nLorentz/gauge-breaking operators.5\nA.d= 4\nWe start the Hamiltonian analysis for d= 4. For this\ncase, the Lagrangian density of the theory reduces to\nL(4)=−1\n4FµνFµν−1\n2Uµνρσ∂µAν∂ρAσ\n−1\n2VµνρσFνµ∂σAρ−1\n4WµρνσFµρFνσ,(14)\nwhere\nUµνρσ= 2K(1)µνρσ\n(4), (15)\nVµνρσ=K(2)ρµνσ\n(4)+K(3)µρνσ\n(4), (16)\nWµρνσ=K(4)µνρσ\n(4). (17)\nThe terms in L(4)with coefficients UµνρσandVµνρσ\nbreak the U(1) gauge symmetry. The four indices\n{µ,ν,ρ,σ}inUµνρσare total symmetric, while Vµνρσ\nareantisymmetricin {µ,ν}andsymmetricin {ρ,σ}. The\ntermswithcoefficients Wµρνσaregaugeinvariantandthe\ncorresponding indices of the coefficients Wµρνσhave the\nsame symmetric properties of the Riemann tensor, i.e.\nthe first pair {µρ}and the last pair {νσ}ofWµρνσare\nboth antisymmetric, and symmetric upon interchange of\nthe two pairs. Variation the action ( 1) with the above\nLagrangian with respect to the field Aµ, one obtains the\nequation of motion of the electromagnetic, i.e.,\n∂µ/parenleftBig\nFµν+Uµνρσ∂ρAσ+Vνµρσ∂σAρ\n+1\n2VρσνµFσρ+WρσµνFρσ/parenrightBig\n= 0.(18)\nNow to perform the Hamiltonian analysis, it is conve-\nnient to define the conjugate momentum\nπµ≡∂L(4)\n∂˙Aµ=−F0µ−U0µνρ∂νAρ−Vµ0νρ∂ρAν\n−1\n2Vνρµ0Fρν−W0µνρFνρ,(19)which makes the fundamental Poisson brackets (PB) as\n{Aµ(x),πν(y)}=δν\nµδ(x−y). (20)\nFrom the conjugate momentum, the canonical Hamilto-\nnian of the system can be expressed as\nH(4)≡˙Aµπµ−L(4). (21)\nIn Hamiltonian analysis, a significant portion of the\nwork involves computing the Poisson bracket between\nthe Hamiltonian and functions in phase space. In this\ncontext, it is important to express the Hamiltonian as a\nfunction of the conjugate momenta and coordinates. To\ndo so, let’s write the time and spatial components of the\nconjugate momentum for ( 19),\nπ0=−U00µν∂µAν−1\n2Vµν00Fνµ,(22)\nπk=DkiF0i+Nk, (23)\nwhere the matrices DkiandNkare defined as\nDki≡ −δki−Vi0k0−Vk00i−2W0k0i−U0k0i,(24)\nNk≡ −/parenleftBig1\n2Vjik0+W0kij/parenrightBig\nFij−(U0k00+Vk000)∂0A0\n−2(U0k0i+Vk00i)∂iA0−(U0kij+Vk0ij)∂iAj.\n(25)\nFrom the expression of Dki, it is easy to get that Dki=\nDik. Further analysis is required to determine the con-\nstraint structure of the system. For this purpose, we\nassume that Dkiis a non-degenerate matrix, just like in\ngauge invariance [ 27,28], such that\nF0i= (D−1)i\nk(πk−Nk). (26)\nAfter some manipulations, one can finally express the\ncanonical Hamiltonian density as a function of the field\nquantity and its conjugate momentum, namely\nH(4)=πk∂kA0+1\n2(D−1)jl(πl−Nl)/bracketleftBig\nπj−Nj+(V0j0k−Vk0j0)(D−1)k\ni(πi−Ni)−2(U0j00+Vj000)∂0A0/bracketrightBig\n+1\n2Fij/parenleftbigg1\n2Fij+1\n2WijklFkl+2Vji0k∂kA0+Vjikl∂lAk/parenrightbigg\n+/parenleftbigg1\n2Uijkl∂iAj+2U0jkl∂jA0/parenrightbigg\n∂kAl\n+2U00kl∂kA0∂lA0−1\n2U0000∂0A0∂0A0, (27)\nwhere we keep the term of ∂0A0explicitly for the reason\nthat when studying the constrained system with d= 4 in\nthe following we have to impose some constraints on the\nLorentz-breaking coefficients and one of our choice here\nis exactly U0000= 0 that can be seen as throwing awaythe term with respect to ∂0A0.\nFollowing the Dirac-Bergmann algorithm, our next\nstep is to write the total Hamiltonian density of the sys-\ntem, which is composed of the canonical Hamiltonian\ndensity and primary constraints. However, we will see6\nthat for the system in d= 4, the existence of primary\nconstraints requires certain additional conditions.\n1.Conditions for the existence of the constraints\nLet us look at what conditions the system described\nby (14) contains constraints. According to the Dirac-\nBergmann procedure, the presence of constraints in the\nsystem with Lagrangian density L(4)can be determined\nbycheckingwhethertheHessianmatrix∂2L\n∂˙Aµ∂˙Aνisdegen-\nerate. Therefore, for making the system with Lagrangian\ndensityL(4)a constrained system, it is sufficient to pro-\nvide the condition that the matrix∂2L(4)\n∂˙Aµ∂˙Aνis degenerate.\nThe Hessian matrix takes the form,\n∂2L(4)\n∂˙Aµ∂˙Aν=−/parenleftbig\nηµν−η0µδ0ν/parenrightbig\n−2W0µ0ν−U0µ0ν\n−Vµ0ν0−Vν0µ0. (28)\nWhen the Hessian matrix is non-degenerate, the system\ndoesnotpossessanyconstraintthatisnot ofourinterest.\nIt is easy to observe that for gauge invariant case since\none has\nUµνρσ= 0 =Vµνρσ, (29)\nthe Hessian matrix is identically degenerate. This indi-\ncates that the theory with U(1) gauge symmetry always\nhas constraints. However, when the U(1) gauge symme-\ntry is broken, whether the theory possesses constraints\ndepends on the specific forms of the gauge-breaking co-\nefficients UµνρσandVµνρσ.\nFor our purpose, we intend to consider the case when\nthe Hessian matrix is degenerate, either with gauge in-\nvariance or gauge breaking. Considering the complexity\nof the Lagrangian density for d= 4 case, for simplicity,\nlet us only focus on the following simple case to make the\nHessian matrix degenerate,\nU0000= 0, U000i=−Vi000. (30)\nFrom the above equation, it can be seen that this con-\nstraint condition only restricts the gauge-breaking coef-ficients and the first row and column of Hessian matrix\nthat are 0, which allows the subsequent conclusions to be\nfully applicable to gauge-invariant systems. For the dis-\ncussion clarity, similar to gauge invariance case [ 27,28],\nwe will assume that the 3 ×3 matrix∂2L(4)\n∂˙Ai∂˙Ajis non-\ndegenerate.\nThus, for d= 4 case, hereafter we will analyze the\nsystem with\nL(4)=−1\n4FµνFµν−1\n2Uµνρσ∂µAν∂ρAσ\n−1\n2VµνρσFνµ∂σAρ−1\n4WµρνσFµρFνσ,(31)\nwith conditions U0000= 0 and U000i=−Vi000.\n2.Canonical analysis\nAfter obtaining the constrained system with a La-\ngrangiandensityof( 31), thissubsectionanalyzesthecon-\nstraint structure of the system. Under the condition of\n(30), the conjugate momentum in ( 23) can be written as\nπ0=−(U00ij+Vji00)∂iAj−2U000i∂iA0,(32)\nπk=Dk\niF0i+Nk, (33)\nwhere\nNk=−(1\n2Vjik0+W0kij)Fij−2(U0k0i+Vk00i)∂iA0\n−(U0kij+Vk0ij)∂iAj. (34)\nThe rank of matrix∂2L(4)\n∂˙Aµ∂˙Aνis 3, giving rise to a unique\nprimary constraint\nφ(4)\n1=π0+(U00ij+Vji00)∂iAj+2U000i∂iA0≈0.\n(35)\nThesymbol” ≈”iscalledweakequalitysymbolwhichim-\nplies the equation only holds on the constraint surfaces\nbut not throughout phase space. With the expressions\nof (30), the canonical Hamiltonian density ( 27) also be-\ncomes\nH(4)=πk∂kA0+1\n2(D−1)jl(πl−Nl)/parenleftBig\nπj−Nj+(V0j0k−Vk0j0)(D−1)k\ni(πi−Ni)/parenrightBig\n+1\n2Fij/parenleftBig1\n2Fij+1\n2WijklFkl+2Vji0k∂kA0+Vjikl∂lAk/parenrightBig\n+/parenleftbigg1\n2Uijkl∂iAj+2U0jkl∂jA0/parenrightbigg\n∂kAl+2U00kl∂kA0∂lA0, (36)\nand the total Hamiltonian density can be written as\nH(4)T=H(4)+u(4)φ(4)\n1, (37)which gives the total Hamiltonian in the form of\nH(4)T=/integraldisplay\nH(4)T(x)d3x, (38)7\nwhereu(4)is an arbitrary Lagrangian multiplier. Note\nthat∂tA0(x) ={A0(x),H(4)T}=u(4)(x), because the\nPoisson bracket between A0(x) andH(4)(y) is zero, and\nbetween A0(x) and other terms in φ(4)\n1(y), except for\nπ(y), is also zero. This gives meaning to the coefficient\nu1(x): the time derivative of A0(x).\nFollowingthe standardDirac-Bergmannprocedure, we\nthen analyze the requirement for the preservation of the\nprimary constraint. Such a requirement is also known\nas the consistency condition of the primary constraint,\nwhich requires that the time derivative of this constraint\nalso vanishes. The consistency condition of the primary\nconstraint, i.e.,\n˙φ(4)\n1(x) ={φ(4)\n1(x),H(4)T} ≈0, (39)\ngives rise to a secondary constraint of the system,\nφ(4)\n2=∂kπk+Mmi(∂mπi−∂mNi)\n+(5U00ij+Vji00)∂i∂jA0\n+Vji0k∂kFij+2U0jkl∂j∂kAl≈0,(40)\nwhere\nMmi=1\n2/bracketleftBig\n(D−1)ij+(D−1)ji\n+2(V0l\n0k−Vk0l0)(D−1)li(D−1)k\nj/bracketrightBig\n×/parenleftbig\n3U0j0m+2Vj00m+Vjm00/parenrightbig\n,(41)\nwhich indicates that the structure of Gauss’s law is in-\nfluenced by the gauge-breaking term.\nSimilar to the primary constraint, the secondary con-\nstraintφ2≈0, being a constraint itself, also has a corre-\nsponding consistency condition, which is\n˙φ(4)\n2(x) ={φ(4)\n2(x),H(4)T} ≈0. (42)\nThis condition then leads to\nOik\n1j∂i∂k(πj−Nj)+Oijk\n2∂i∂j∂kA0+Ojilk\n3∂j∂i∂lAk\n+Oklij\n4∂k∂lFij+Tij∂i∂ju(4)≈0, (43)where\nOjilk\n3≡(δjm+Mjm)(1\n2Vmikl+Uimlk), (44)\nOklij\n4≡1\n2/parenleftbig\nδkm+Mkm/parenrightbig\n(ηliηmj+Wijlm+Vjiml),(45)\nOik\n1j≡1\n2/bracketleftBig\n(δim+Mim)/parenleftbig\nVmkl0+2W0lkm\n+U0lkm+Vl0km/parenrightbig\n+Mi\nm(Vlkm0+2W0mkl+U0mkl+Vm0kl)\n+2(Vli0k+U0ikl)/bracketrightBig\n×/bracketleftBig\n(D−1)jl+(D−1)lj\n+2(V0i10k1−Vk10i10)(D−1)i1j(D−1)k1\nl/bracketrightBig\n,\n(46)\nOijk\n2≡(δil+Mil)(Vlj0k+2U0jkl)\n+Mil(U0ljk+Vl0jk)+2U0ijk, (47)\nand\nTij≡6U00ij+(3U00jk+Vkj00+2Vk00j)Mi\nk.\n(48)\nForgaugeinvariancecase, one has Tij= 0, one can check\nthat the condition ( 43) satisfies identically. When the\nU(1) gaugesymmetry of the theory is violated, Tijare in\ngeneralnonzeros,then( 43)representsarestrictiononthe\nLagrange multiplier u(4). One special case is that Tij=\n0 for some specific combination of the gauge breaking\ncoefficients, for which a new constraint arises, which is\nφ(4)\n3=Oik\n1j(∂i∂kπj−∂i∂kNj)+Oijk\n2∂i∂j∂kA0\n+Ojilk\n3∂j∂i∂lAk+Oklij\n4∂k∂lFij≈0.(49)\nWith this constraint, repeating the above procedure, its\nconsistency condition may produce more constraints un-\nder certain conditions. It is worth mentioning that ac-\ncording to Table. 1, the number of independent coeffi-\ncients is finite, and the emergence of new constraints will\ngive a set of limiting equations about the coefficients.\nJust as in the presence of φ(4)\n3≈0,Tij= 0 gives 9\nequations between the coefficients, so that the number of\nindependent coefficients will be reduced. Detailed analy-\nsis will show that each time a new constraint is present,\nthe number of limiting equations for the coefficients will\nincrease quickly as compared to the previous constraint,\neventuallystoppingthe generationofpossibleconstraints\nat a certain step, thus making the Poisson bracket closed\nand the number of constraints limited. According to the\nanalysis in [ 45,46], such constraint may also lead to the\nunphysical halfdegreeof freedom. However, this requires\na very special choice of gauge-breaking coefficients. For\nsimplicity, we will not explore these specific cases in de-\ntail in this paper.\nBefore we go further, we would like to summarize the\nmain results we have gotten so far for the d= 4 case. For8\nboth the gauge invariance case and gauge breaking case\nwith the degenerate condition ( 30), the theory can have\none primaryconstraint φ(2)\n1and one secondaryconstraint\nφ(4)\n2.\n3.Counting the degrees of freedom\nAfter obtaining all the primary and secondary con-\nstraints of the system, let us turn to identify the first-\nand second-class constraints of the system by analyzing\nthe Poisson bracket of the constraints. In general, the\nfirst-class constraints are associated with the gauge sym-\nmetry of the theory, they are gauge generators, which\ngenerategaugetransformationsthat don’talterthephys-\nical state. The second-class constraints can not gener-\nate gauge transformations since the transformation gen-\nerated by a second-class does not preserve all the con-\nstraints, whichviolatestheconsistencycondition, butare\nimportant in the definition of Dirac bracket, which plays\na key role in the transition from classical theory to quan-\ntum theory [ 43]. Specifically, the first-class constraints\nare those constraints whose Poisson bracket with every\nconstraintvanishesweakly,otherwisearethesecond-class\nof constraint.\nThe Poissonbracketof the primaryconstraint φ(4)\n1and\nthe secondary constraint φ(4)\n2gives\n{φ(4)\n2(y),φ(4)\n1(x)}= (C1)ij∂′i∂jδ(x−y)\n+(C2)ij∂′\ni∂′\njδ(x−y).(50)\nwhere∂′denotes the partial derivative with respect to y,\nand\n(C1)ij=−/parenleftbig\nU00ji+(U00jk+Vkj00)Mik/parenrightbig\n,(51)\n(C2)ij= 2Mik(U0k0j+Vk00j)+5U00ij.(52)\nIt is easy to see that for the gauge invariance case,\nCij\n1= 0 =Cij\n2, therefore, φ(4)\n1andφ(4)\n2are both first-\nclass constraints. For the case without the gauge sym-\nmetry, the gauge breaking coefficients UµνρσandVµνρσ\nare in general nonzero, thus the theory does not possess\nany first-class constraint. In this case, φ(4)\n1andφ(4)\n2are\nboth second-class constraints. This result is also what\nwe expect since the first-class constraint can only exist\nwhen the theory has gauge symmetry, and thus a the-\nory without gauge symmetry should not have first-class\nconstraint.\nGetting all first- and second-class constraints, one can\ncount the number of physical degrees of freedom ( NDOF)\nby using the following formula [ 47],\nNDOF=1\n2(Nvar−2NOF−NOS), (53)\nwhere “Nvar” means the total number of canonical vari-\nables, “NOF” is the number of constraints of the first\nclass, and “NOS” represents the number of second-classconstraints. Thus, for the gauge invariance case, the\nnumber of degrees of freedom is\nNDOF=1\n2(8−2×2) = 2, (54)\nwhich is the same as that in the standard Maxwell’s elec-\ntrodynamics, while in the case of gauge violation, one\nhas\nNDOF=1\n2(8−2) = 3. (55)\nTherefore, the violation of the U(1) gauge symmetry in-\nduces one additional physical degree of freedom, com-\nparedtothe standardMaxwell’selectrodynamicsandthe\ntheory with gauge invariance. This additional physical\ndegree of freedom results in a third state of polarization,\ncorresponding to a new particle called longitudinal pho-\nton [48]. This representsa new electromagneticradiation\nwhich may alter the radiation spectra of a lot of sources\nwith nonzerotemperature [ 48]. However, its phenomeno-\nlogical effects in both experiments and astrophysical ob-\nservations are expected to be too small to be detected up\nto now [48–50].\nB.d= 3\nIn this subsection, let’s turn to analyze the constraint\nstructure of the theory with Lorentz/gauge-breaking op-\nerators with dimension d= 3. The Lagrangian density\nof the considered theory can be written in the form of\nL(3)=−1\n4FανFαν+1\n2SµνρAρFνµ\n+1\n2ǫκµνρ(kAF)κAµFνρ. (56)\nwhere\n(kAF)κ=1\n3!ǫκµνρK(5)µνρ\n(3), (57)\nSµνρ=K(2)ρµν\n(3)−K(3)µρν\n(3). (58)\nNote that the terms with ( kAF)κare gauge-invariantand\nthe terms with Sµνρare gauge-breaking. Variation of the\naction with respect to Aµ, one obtains field equation\nǫκνµρ(kAF)κFµρ+1\n2SµρνFρµ\n+∂µ(Fµν−SνµρAρ) = 0. (59)\nThecorrespondingconjugatemomentumofthistheory\nread,\nπµ=∂L(3)\n∂˙Aµ=−F0µ+/parenleftbig\nǫβν0µkAFβ+Sµ0ν/parenrightbig\nAν.(60)\nTo analyze the constraint structure of the theory, it is\nconvenient to write down the time and spatial compo-\nnents of the above conjugate momentum ( 60) as\nπ0= 0, (61)\nπk=−F0k+/parenleftbig\nǫjµ0kkAFj+Sk0µ/parenrightbig\nAµ.(62)9\nOne can conclude that in the specific dimension d= 3,\nthegauge-breakingandgauge-invariantLorentz-breaking\nelectrodynamics have the same form for the conjugate\nmomentum π0of the photon field A0. From these prop-\nerties, the only primary constraint of the system is\nφ(3)\n1=π0≈0. (63)\nThen, the canonical Hamiltonian density and total\nHamiltonian density are given by\nH(3)=πk˙Ak−L(3), (64)\nand\n(H(3))T=H(3)+u(3)φ(3)\n1. (65)\nHere, similar to the case of d= 4, the coefficient u(3)is\nalso the time derivative of A0.\nOnce again, the consistency condition of the primary\nconstraint gives a secondary constraint\nφ(3)\n2=˙φ(3)\n1\n=∂kπk+1\n2Fik/parenleftBig\nǫj0ik(kAF)j+Ski0/parenrightBig\n+Sk00(ǫjµ0k(kAF)j+Sk0µ)Aµ−πkSk00\n≈0. (66)\nSimilarly, this means that the form of Gauss’s law needs\nto be modified when considering the presence of the\ngauge-breaking term. The consistency condition of φ(3)\n2\nimpose a restriction on the u(3):\n˙φ(3)\n2=Sl00Sl00u(3)−/parenleftbig\nSi0k+Sik0/parenrightbig\n∂kπi\n+Sk00/parenleftbig\n2ǫjk0i(kAF)j+Si0k−Sk0i/parenrightbig\nπi\n+Si00/bracketleftbig\n2ǫl0ik(kAF)l+Si0k−Sk0i/bracketrightbig\n×[ǫjn0k(kAF)j+Sk0n]An\n+1\n2/parenleftbig\n−2Sj00ηki+Sjik)∂kFij\n−1\n2(ǫ0kij(kAF)0+Sjik)Sk00Fij\n+2Sl00(Sl0k+Slk0)∂kA0\n+Si00/bracketleftbig\n2ǫl0i\nk(kAF)l+Si0\nk−Sk0i/bracketrightbig\nSk00A0\n+/bracketleftBig\nSk00(ǫ0jik(kAF)0+Skij)\n+(Sk0i+Ski0)/bracketleftbig\nǫlj0k(kAF)l+Sk0j/bracketrightbig/bracketrightBig\n∂iAj\n≈0. (67)\nFor gauge invariance case, all the gauge breaking coef-\nficientsSµνρ= 0, and thus the above consistency con-\ndition satisfies identically. In this case, the theory only\nhas two constraints, the primary constraint φ(3)\n1and the\nsecondary constraint φ(3)\n2.\nWhen the gauge symmetry is violated, the gauge\nbreaking coefficients Sµνρare in general nonzeros. Forthis case, the above consistency condition leads to a spe-\ncific form of u(3),\nu(3)≈ −1\nSm00Sm00(68)\n×/bracketleftBigg\nSi00/bracketleftbig\n2ǫl0ik(kAF)l+Si0k−Sk0i/bracketrightbig\n×[ǫjn0k(kAF)j+Sk0n]An\n−/parenleftbig\nSi0k+Sik0/parenrightbig\n∂kπi\n+Sk00/parenleftbig\n2ǫjk0i(kAF)j+Si0k−Sk0i/parenrightbig\nπi\n+1\n2/parenleftbig\n−2Sj00ηki+Sjik)∂kFij\n−1\n2(ǫ0kij(kAF)0+Sjik)Sk00Fij\n+2Sl00(Sl0k+Slk0)∂kA0\n+Si00(2ǫl0i\nk(kAF)l+Si0\nk−Sk0i)Sk00A0\n+/bracketleftBig\nSk00(ǫ0jik(kAF)0+Skij)\n+(Sk0i+Ski0)/bracketleftbig\nǫlj0k(kAF)l+Sk0j/bracketrightbig/bracketrightBig\n∂iAj/bracketrightBigg\n.\n(69)\nThis indicates that the theory for this case does not have\nany additional constraints. Similar to the gauge invari-\nance case, it only has two constraints, one primary con-\nstraintφ(3)\n1and one secondary constraint φ(3)\n2. Here we\nwould like to mention that, under certain conditions on\nthe gauge breaking coefficients such that S00\nlSl00= 0,\nthe theorymay produceadditional constraints. However,\nthis requires a very special choice of gauge-breaking co-\nefficients. For simplicity, we will not explore this specific\ncase in detail in this paper and focus on the case with\nS00\nlSl00/ne}ationslash= 0 when gauge symmetry is violated.\nThen, let us consider the Poisson bracket of the pri-\nmary constraint φ(3)\n1and the secondary constraint φ(3)\n2,\nwhich is\n{φ(3)\n1(x),φ(3)\n2(y)}=−Sk00Sk00δ(x−y).(70)\nIt is obvious that for gauge invariance case, since Sµνρ=\n0, the abovePoissonbracketvanishes and the constraints\nφ(3)\n1andφ(3)\n2are both first-class. Thus, the number of\nthe physical degrees of freedom is\nNDOF=1\n2(8−2×2) = 2, (71)\nwhich is the same as that in the standard Maxwell’s elec-\ntrodynamics.\nFor the case with gauge violation, since in general\nSk00Sk00/ne}ationslash= 0,φ(3)\n1,φ(3)\n2arebothsecond-classconstraints.\nThus, the number of the physical degrees of freedom is\nNDOF=1\n2(8−2) = 3. (72)10\nSimilar to the case with d= 4, the violation of the U(1)\ngauge symmetry induces one extra physical degree of\nfreedom, compared to the standard Maxwell’s electro-\ndynamics and the case with gauge invariance.\nC.d= 2\nAfter completing the constraint structure analysis of a\nspecific dimension d= 3 andd= 4, let us turn to analyze\nthe structure of d= 2, which Lagrangian density is\nL(2)=−1\n4FµνFµν+1\n2UµνAµAν, (73)\nwith\nUµν= 2K(1)µν\n(2), (74)\nwhich are gauge-breaking terms. Variation of the action\nwith respect to Aµ, one obtains\nUνµAµ+∂µFµν= 0 (75)\nThe conjugate momenta now read\nπµ=∂L(2)\n∂˙Aµ=−F0µ. (76)\nOne difference from d= 3 and d= 4 is that at this point\nthe conjugate momenta πµare not affected by gauge-\nbreaking coefficients, Uµν, so they are the same as in\nMaxwell electrodynamics since the time derivative of the\nphotofield Aµonlyappearsinthe Maxwellterm, the first\nterm in ( 73). Therefore, the only primary constraint is\nφ(2)\n1=π0≈0. (77)\nThe canonical Hamiltonian density just differs by one\nterm ofUµνfrom that in Maxwell electrodynamics:\nH(2)=πk∂kA0−1\n2πkπk+1\n4FikFik−1\n2UµνAµAν.\n(78)\nThis gives the total Hamiltonian density:\nH(2)T=H(2)+u(2)φ(2)\n1. (79)\nThe influence of gauge-breaking terms are reflected in\nsecondary constraint because it originates from the Pois-\nson bracket between primary constraint and total Hamil-\ntonian which now is gauge-breaking. It reads\nφ(2)\n2=∂kπk+U0µAµ≈0. (80)\nThe consistency condition of φ2is\n˙φ(2)\n2=Uµk∂kAµ+U0k∂kA0−U0kπk+u(2)U00≈0,\n(81)which provides the constraint on u(2)whenU00/ne}ationslash= 0:\nu(2)≈(U00)−1/bracketleftbig\nU0k(πk−∂kA0)−Uµk∂kAµ/bracketrightbig\n.(82)\nNoticing the absence of π0inH(2), we get ˙A0(x) =\n{A0(x),H(2)T}=u(2)(x). As a consequence, the mean-\ning ofu(2)is the time derivative of A0. Combining ( 82),\nthe time component A0of the photon field will be deter-\nmined by the first-order differential equation.\nFinally, the Poisson bracket of φ(2)\n1andφ(2)\n2is\n{φ(2)\n1(x),φ(2)\n2(y)}=−U00δ(x−y). (83)\nWhen the gauge breaking coefficients Uµνare nonzeros,\nφ(2)\n1,φ(2)\n2areboth secondclassconstraints. Similartothe\ncases ofd= 4 andd= 3, we will not explore in detail the\ncase with U00= 0, which may induce more constraint.\nThen the number of the physical degree of freedom is\nNDOF=1\n2(8−2) = 3. (84)\nAgain, comparedto the standardMaxwellelectrodynam-\nics, since d= 2 operators always break the U(1) gauge\nsymmetry, it induces one additional physical degree of\nfreedom.\nIV.Map to several specific models\nThe Lorentz-violating electrodynamics presented in\nthis paper provide a unifying framework for describing\npossible violations of Lorentz and U(1) gauge symme-\ntries in the electromagnetic interaction. In this section,\nwe present several specific modified electrodynamics by\nwriting their actions in the form of ( 1) and summa-\nrize their Hamiltonian structure from our general anal-\nysis. We consider three specific theories, the Lorentz-\ninvariance-Violating (LIV) electrodynamics, the Carroll-\nField-Jackiw (CFJ) electrodynamics, and the Proca elec-\ntrodynamics. The first two theories only break the\nLorentz symmetry of the theory, while the last one only\nbreaksU(1) gauge symmetry.\nA.LIV electrodynamics\nWe first consider LIV electrodynamics, which is pro-\nposed in [ 27,28] and the Lagrangian density is given by\n[27,28]\nLLIV=−1\n4FµνFµν−1\n4WµνρσFµνFρσ,(85)\nwhich corresponds to the case in our model where the\ngauge-breaking coefficients UµνρσandVµνρσare set to\nbe zeros for d= 4. Using the result from ( 35) and (40),\nwe obtain the constraint structure in this case:\n(φ1)LIV=π0≈0, (86)\n(φ2)LIV=∂kπk≈0. (87)11\nTABLE III: The number of the first-class and second-class const raints, and the number of the physical degrees of\nfreedom for each case with a specific dimension d= 2,d= 3, and d= 4.\ndgauge invariance number of first-class constraint number of first-class constraint number of DOF\n2 no 0 2 3\n3yes 2 0 2\nno 0 2 3\n4yes 2 0 2\nno 0 2 3\nAnd we need verify that the consistency of ( φ2)LIVgives\nno new constraints. Noticing that Tij=Mij= 0 by\nusing (41) and (48) sinceVµνρσ=Uµνρσ= 0, then we\nonly need to check whether φ(4)\n3in (49) is identically zero\nthroughout the phase space, not just on the constraint\nsurface. In this case, the only possible non-zero coeffi-\ncients remaining in ( 49) areOik\n1jandOklij\n4. According\n(33), as a result, φ(4)\n3now changes to\nW0lik((D−1)jl+(D−1)lj)∂k∂i(πj−Nj)\n+1\n2(ηliηkj+Wijlk)∂k∂lFij. (88)\nSincethelasttwoindicesof Wijklareantisymmetric, and\ni,jinηliηkj∂k∂lare symmetric, ( 88) is equal to 0. This\nproves that ( φ2)LIVdoes not give new constraints. In\naddition, since the theory has gauge symmetry, both the\nconstraints ( φ1)LIVand (φ2)LIVare first-class, thus the\ntheory has two degrees of freedom, the same as that in\nthe standard Maxwell electrodynamics. Our result here\nis consistent with the results in LIV [ 27,28].\nB.CFJ electrodynamics\nFor CFJ electrodynamics, its Lagrangian density is\ngiven by [ 35]:\nLCFJ=−1\n4FµνFµν+1\n2ǫκµνρ(kAF)κAµFνρ,(89)\nwith the caveat that the coefficients ( kAF)κdiffers by a\nfactor of −1\n2compared to [ 35]. The Lagrangian density\n(89) can be obtained in our model by setting Sµνρ= 0 to\nzero ind= 3. Similarly, according to the results in ( 63)\nand (66), we have two constraints for this theory,\n(φ1)CFJ=π0=F00≈0, (90)\n(φ2)CFJ=∂kπk+1\n2Fikǫj0ik(kAF)j≈0,(91)\nwhere (φ1)CFJis the primary constraint and ( φ2)CFJ\nis the secondary constraint. According to ( 67), it canbe concluded that in this case, ( ˙φ2)CFJis always zero\nin phase space, which does not generate new constraints.\nThis result confirms [ 35], with the additional point to\nnote that the coefficient kAFhere differs by a factor of\n−1\n2compared to [ 35]. Note that both the ( φ1)CFJand\n(φ2)CFJare first-class and the number of degrees of free-\ndom is the same as that in the standard Maxwell electro-\ndynamics.\nC.The Proca electrodynamics\nWhend= 2, if one sets Uµν=m2ηµν, then the La-\ngrangian density L(2)will return to the case of Proca\nelectrodynamics [ 36]:\nLProca=−1\n4FανFαν+1\n2m2AµAµ,(92)\nwheremis the mass of the photon. Because of this mass\nterm, the Proca electrodynamics breaks the U(1) gauge\nsymmetry of the theory. Replacing the constant coeffi-\ncientsUµνwith the product of metric tensor ηµνandm2\nis the reason why Proca electrodynamics doesn’t break\nLorentz symmetry as the product is also a tensor. Same\nreason as d= 3,4, at this point, the theory has two con-\nstraints as well,\n(φ1)Proca=π0≈0, (93)\n(φ2)Proca=∂kπk+m2A0≈0.(94)\nAnd (φ2)Procagives no new constraints. They are\nboth second-class since {(φ1)Proca(x),(φ2)Proca(y)}=\nm2δ(x−y). Thus this theory propagates three physical\ndegrees of freedom, which is different from the two de-\ngrees of freedom in the standard Maxwell electrodynam-\nics and the cases with gauge invariance. These results\nare consistent with [ 36], with the only difference being\nthat their coefficient m2differs from ours by a factor of\n1\n2.12\nV.Summary and discussion\nIn this paper, we perform an extended analysis of\nthe modified electrodynamics with the violations of both\nLorentz symmetry and U(1) gauge symmetry in the\nframework of the Standard-Model Extension. This rep-\nresents an extension of the previous construction of the\nLorentz-violating electrodynamics with gauge invariance\n[24]. For our purpose, by following the procedure in\n[24], we construct the quadratic Lagrangian density of\nelectrodynamics by allowing the violations of both the\nLorentz and U(1) gauge symmetries. The Lorentz- and\ngauge-violating effects in the quadratic Lagrangian are\nrepresented by the new operators with specific dimen-\nsiond≤4. With the constructed quadratic Lagrangian,\nwe calculate in detail the number of independent com-\nponents of the Lorentz-violating operators at different\ndimensions with both cases of the gauge invariance and\ngauge violation cases.\nWe then perform the Hamiltonian analysis of the gen-\neral theory by considering Lorentz and gauge-breaking\noperators with dimension d≤4 as examples. Specifi-\ncally, weperformthe analysisfor d= 4,d= 3, and d= 2,\nrespectively. It is shown that the Lorentz-breaking op-\nerators with gauge invariance do not change the classes\nof the constraints of the theory and have the same num-\nber of physical degrees of freedom as that in the stan-\ndard Maxwell’s electrodynamics. When the U(1) gauge\nsymmetry-breaking operators are presented, the theo-ries in general lack first-class constraint and have one\nadditional physical degree of freedom, compared to the\nstandard Maxwell’s electrodynamics. The results of the\nHamiltonian structure and their corresponding number\nof degrees of freedom are presented in Table. III.\nFinally,wealsomapourgeneralanalysistoseveralspe-\ncific modified electrodynamics, including LIV electrody-\nnamics, CFJ electrodynamics, and Proca electrodynam-\nics. While the former two theories represent two specific\nexamples of Lorentz-violating theory with gauge invari-\nance, the latter one is a theory that breaks U(1) gauge\nsymmetrybutstillkeepsthe Lorentzsymmetry. Weshow\nthat our general results are consistent with the existing\nHamiltonian analysis in the literature for these specific\nexamples.\nAcknowledgments\nThis work is supported by the National Key Re-\nsearch and Development Program of China under Grant\nNo. 2020YFC2201503, the Zhejiang Provincial Nat-\nural Science Foundation of China under Grant No.\nLR21A050001 and No. LY20A050002, and the Na-\ntionalNaturalScienceFoundationofChinaunderGrants\nNo. 12275238 and No. 11675143. B.-F. Li is supported\nby the National Natural Science Foundation of China\n(NNSFC) with the grant No. 12005186.\n[1]G. Aad et al. (ATLAS), Observation of a\nnew particle in the search for the Standard\nModel Higgs boson with the ATLAS detec-\ntor at the LHC, Phys. Lett. B 716, 1–29(2012) ,\narXiv:1207.7214 [hep-ex] .\n[2]S. Chatrchyan et al.(CMS), Observation of a New\nBoson at a Mass of 125 GeV with the CMS Ex-\nperiment at the LHC, Phys. Lett. B 716, 30–61 (2012) ,\narXiv:1207.7235 [hep-ex] .\n[3]B. P. Abbott et al.(LIGO Scientific,Virgo), Ob-\nservation of Gravitational Waves from a Binary\nBlack Hole Merger, Phys. Rev. Lett. 116,061102 (2016) ,\narXiv:1602.03837 [gr-qc] .\n[4]V. A. Kosteleck´ y and R. Potting, CPT, strings, and\nmeson factories, Phys. Rev. D 51, 3923–3935 (1995) ,\narXiv:hep-ph/9501341 .\n[5]D.Colladay andV. A.Kosteleck´ y, CPTviolation andthe\nstandard model, Phys. Rev. D 55, 6760–6774 (1997) .\n[6]V. A. Kosteleck´ y, Gravity, lorentz violation, and the\nstandard model, Phys. Rev. D 69, 105009 (2004) .\n[7]V. A. Kosteleck´ y and N. Russell, Data Tables for Lorentz\nand CPT Violation, Rev. Mod. Phys. 83, 11–31 (2011) ,\narXiv:0801.0287 [hep-ph] .\n[8]V. A. Kosteleck´ y and M. Mewes, Lorentz-\nviolating electrodynamics and the cosmic microwave\nbackground, Phys. Rev. Lett. 99, 011601 (2007) ,\narXiv:astro-ph/0702379 .[9]V. A. Kosteleck´ y and M. Mewes, Astrophysi-\ncal Tests of Lorentz and CPT Violation with\nPhotons, Astrophys. J. Lett. 689, L1–L4 (2008) ,\narXiv:0809.2846 [astro-ph] .\n[10]P. Cabella, P. Natoli, and J. Silk, Constraints on CPT\nviolation from WMAP three year polarization data:\na wavelet analysis, Phys. Rev. D 76, 123014 (2007) ,\narXiv:0705.0810 [astro-ph] .\n[11]M. Li and X. Zhang, Cosmological CPT violating effect\non CMB polarization, Phys. Rev. D 78, 03516 (2008) ,\narXiv:0810.0403 [astro-ph] .\n[12]N. Aghanim et al.(Planck), Planck intermediate\nresults. XLIX. Parity-violation constraints from po-\nlarization data, Astron. Astrophys. 596, A110 (2016) ,\narXiv:1605.08633 [astro-ph.CO] .\n[13]G. Gubitosi, L. Pagano, G. Amelino-Camelia, A. Mel-\nchiorri, and A. Cooray, A Constraint on Planck-scale\nModifications to Electrodynamics with CMB polariza-\ntion data, JCAP08, 021,arXiv:0904.3201[astro-ph.CO] .\n[14]S. M. Lee, D. W. Kang, J.-O.Gong, D. Jeong, D.-W.\nJung, and S. C. Park, Cosmological Consequences of Ki-\nnetic Mixing between Photon and Dark Photon, (2023),\narXiv:2307.14798 [hep-ph] .\n[15]L. Caloni, S. Giardiello, M. Lembo, M. Gerbino,\nG. Gubitosi, M. Lattanzi, and L. Pagano, Probing\nLorentz-violating electrodynamics with CMB polariza-\ntion,JCAP03, 018,arXiv:2212.04867 [astro-ph.CO] .13\n[16]M. Galaverni and G. Sigl, Lorentz Violation in\nthe Photon Sector and Ultra-High Energy Cos-\nmic Rays, Phys. Rev. Lett., vol. 100, p. 021102, (2008),\narXiv:0708.1737 [astro-ph] .\n[17]V. Vasileiou, A. Jacholkowska, F. Piron, J. Bol-\nmont, C. Couturier, J. Granot, F. W. Stecker,\nJ. Cohen-Tanugi, and F. Longo, Constraints on\nLorentz invariance violation from Fermi-Large\nArea Telescope observations of gamma-ray bursts,\nPhysical Review D 87, 10.1103/physrevd.87.122001 (2013) .\n[18]Desai, Shantanu, Astrophysical and Cosmological\nSearches for Lorentz Invariance Violation. (2023),\narXiv:2303.10643 [astro-ph.CO] .\n[19]Z. Cao et al. (LHAASO), Exploring Lorentz Invariance\nViolation from Ultrahigh-Energy γRays Observed\nby LHAASO, Phys. Rev. Lett. 128, 051102 (2022) ,\narXiv:2106.12350 [astro-ph.HE] .\n[20]Pierre Auger Collaboration, Search for Ultra-high-energy\nPhotons from Gravitational Wave Sources with the\nPierre Auger Observatory, Astrophys.J. 952 (2023) 1, 91,\ne-Print: 2307.10839 [astro-ph.HE] .\n[21]Pierre Auger collaboration, Testing effects of Lorentz in-\nvariance violation in the propagation of astroparticles\nwith thePierre Auger Observatory, JCAP 01 (2022) 023 ,\narXiv:2112.06773 [astro-ph.HE] .\n[22]K. Astapov, D. Kirpichnikov, and P. Satunin,\nPhoton splitting constraint on Lorentz invari-\nance violation from Crab Nebula spectrum,\nJournal of Cosmology and Astroparticle Physics 2019 (04), 0 54???054 .\n[23]M. Chr´ etien, J. Bolmont, and A. Jacholkowska,\nConstraining photon dispersion relations from ob-\nservations of the vela pulsar with H.E.S.S (2015),\narXiv:1509.03545 [astro-ph.HE] .\n[24]V. A. Kosteleck´ y and M. Mewes, Electrodynamics\nwith lorentz-violating operators of arbitrary dimension,\nPhys. Rev. D 80, 015020 (2009) .\n[25]L. H. C. Borges and A. F. Ferrari, Exter-\nnal sources in a minimal and nonminimal\nCPT-odd Lorentz violating Maxwell electrody-\nnamics, Mod. Phys. Lett. A 37, 2250021 (2022) ,\narXiv:2202.04424 [hep-th] .\n[26]P. D. S. Silva, L. Lisboa-Santos, M. M. Ferreira, Jr.,\nand M. Schreck, Effects of CPT-odd terms of dimen-\nsions three and five on electromagnetic propagation\nin continuous matter, Phys. Rev. D 104, 116023 (2021) ,\narXiv:2109.04659 [hep-th] .\n[27]R. Casana, M. M. Ferreira,Jr,J. S. Rodrigues,\nand M. R. O. Silva, Finite Temperature CPT-\neven Electrodynamics of the Standard Model\nExtension, Phys. Rev. D 80, 085026 (2009) ,\narXiv:0907.1924 [hep-th] .\n[28]C. A. Escobar and M. A. G. Garcia, The Full CPT-\nEven Photon Sector of the Standard Model Extension\nat Finite Temperature, Phys. Rev. D 92,025034 (2015) ,\narXiv:1505.00069 [hep-th] .\n[29]R. Avila, J. R. Nascimento, A. Y. Petrov,\nC. M. Reyes, and M. Schreck, Causality, uni-\ntarity, and indefinite metric in Maxwell-Chern-\nSimons extensions, Phys. Rev. D 101, 055011 (2020) ,\narXiv:1911.12221 [hep-th] .\n[30]M. M. Ferreira, J. a. A. A. S. Reis, and\nM. Schreck, Dimensional reduction of the elec-\ntromagnetic sector of the nonminimal standardmodel extension, Phys. Rev. D 100, 095026 (2019) ,\narXiv:1905.04401 [hep-th] .\n[31]M. A. Anacleto, F. A. Brito, E. Maciel, A. Mo-\nhammadi, E. Passos, W. O. Santos, and\nJ. R. L. Santos, Lorentz-violating dimension-\nfive operator contribution to the black body\nradiation, Phys. Lett. B 785, 191–196 (2018) ,\narXiv:1806.08273 [hep-th] .\n[32]V. A. Kosteleck´ y and M. Mewes, Cosmolog-\nical constraints on Lorentz violation in elec-\ntrodynamics, Phys. Rev. Lett. 87, 251304 (2001) ,\narXiv:hep-ph/0111026 .\n[33]S. M. Carroll, G. B. Field, and R. Jackiw, Limits on a\nLorentz and Parity Violating Modification of Electrody-\nnamics,Phys. Rev. D 41, 1231 (1990) .\n[34]E. Allys, P. Peter, and Y. Rodriguez, Generalized\nProca action for an Abelian vector field, JCAP02, 004,\narXiv:1511.03101 [hep-th] .\n[35]Casana, R., Manoel M. Ferreira and Josberg\nS. Rodrigues, Lorentz-violating contributions\nof the Carroll-Field-Jackiw model to the CMB\nanisotropy, Physical Review D 78 (2008): 125013 ,\narXiv:0810.0306 [hep-th] .\n[36]Z. Molaee and A. Shirzad, Constraint structure of the\ngeneralized proca model in the lagrangian formalism\n(2023),arXiv:2307.02384[hep-th] .\n[37]D. Colladay and V. A. Kostelecky, Lorentz\nviolating extension of the standard model,\nPhys. Rev. D 58, 116002 (1998) ,arXiv:hep-ph/9809521 .\n[38]E. Svanberg, Theories with higher-order time deriva-\ntives and the Ostrogradsky ghost,Bachelor thesis (2022),\narXiv:2211.14319 [physics.class-ph] .\n[39]R. P. Woodard, Ostrogradsky’s theorem on Hamil-\ntonian instability, Scholarpedia 10,32243 (2015) ,\narXiv:1506.02210 [hep-th] .\n[40]P. G. Bergmann and J. H. M. Brunings, Non-linear\nfield theories II. canonical equations and quantization,\nRev. Mod. Phys., 21,480–487 (1949).\n[41]J. L. Anderson and P. G. Bergmann, Constraints in co-\nvariant field theories, Phys. Rev., 83,1018–1025 (1951).\n[42]P. A. M. Dirac, Generalized Hamiltonian Dynamics,\nCanadian Journal of Mathematics, 2, 129-148 (1950).\n[43]P. A. M. Dirac, The Hamiltonian Form of Field Dynam-\nics,Canadian Journal of Mathematics, 3, 1-23 (1951).\n[44]P. A. M. Dirac, Lectures on Quantum Mechanics, Dover\nPublications (2001).\n[45]X. Gao and Z.-B. Yao, Spatially covariant grav-\nity theories with two tensorial degrees of free-\ndom: The formalism, Physical Review D 101,\n10.1103/physrevd.101.064018 (2020).\n[46]X. Gao and Z.-B. Yao, Spatially covariant gravity with\nvelocity of the lapse function: the Hamiltonian analysis,\nJCAP05, 024,arXiv:1806.02811[gr-qc] .\n[47]M. Henneaux and C. Teitelboim, Quantization of gauge\nsystems, Princeton University Press, chap. 1, pp. 24, 29\n(1992).\n[48]L.-C. Tu, J. Luo, and G. T. Gillies, The mass of the pho-\nton,Reports on Progress in Physics 68, 77 - 130 (2004) .\n[49]A. S. Goldhaber and M. M. NIETO, Terrestrial\nand Extraterrestrial Limits on The Photon Mass,\nRev. Mod. Phys, 43, 277–296 (1971) .\n[50]J. C. Byrne and R. R. Burman, A photon\nrest mass and electric currents in the Galaxy,\nJ. Phys. A: Math. Nucl. Gen. 6L12 (1973) ." }, { "title": "2106.11212v1.Gagliardo_Nirenberg_inequalities_in_Lorentz_type_spaces_and_energy_equality_for_the_Navier_Stokes_system.pdf", "content": "arXiv:2106.11212v1 [math.AP] 21 Jun 2021Gagliardo-Nirenberg inequalities in Lorentz type spaces a nd\nenergy equality for the Navier-Stokes system\nYanqing Wang∗, Wei Wei†and Yulin Ye‡\nAbstract\nIn this paper, we derive some new Gagliardo-Nirenberg type inequalit ies in Lorentz\ntype spaces without restrictions on the second index of Lorentz n orms, which generalize\nalmost all known corresponding results. Our proof mainly relies on th e Bernstein in-\nequalities in Lorentz spaces, the embedding relation among various L orentz type spaces,\nand Littlewood-Paley decomposition techniques. In addition, we est ablish several novel\ncriteria in terms of the velocity or the gradient of the velocity in Lore ntz spaces for\nenergy conservation of the 3D Navier-Stokes equations. Particu larly, we improve the\nclassical Shinbrot’s condition for energy balance to allow both the sp ace-time directions\nof the velocity to be in Lorentz spaces.\nMSC(2020): 35A23, 42B35, 35L65, 76D03\nKeywords: Gagliardo-Nirenberg inequality; Lorentz spaces; energy e quality; Navier-\nStokes equations; Littlewood-Paley decomposition\n1 Introduction\n1.1 Gagliardo-Nirenberg inequality\nAn important way to study well-posedness of partial different ial equations is via the re-\nsearch of their weak solutions. It is well known that Gagliar do-Nirenberg inequality is a\nfundamental tool to improve the regularity of weak solution s, which has gained widespread\napplications such as in Hilbert’s 19th problem [15], Caffarel li-Kohn-Nirenberg theorem [8]\nfor the 3D Navier-Stokes equations and the critical quasi-g eostrophic equations [9]. The\nclassical integer version of Gagliardo-Nirenberg inequal ity is the generalization of Sobolev\nembedding theorem and was discovered independently by Gagl iardo [19] and Nirenberg [33]\nas follows: for all smooth functions uinRnwith compact support, there holds\n/ba∇dblDju/ba∇dblLp(Rn)≤C/ba∇dblDmu/ba∇dblθ\nLr(Rn)/ba∇dblu/ba∇dbl1−θ\nLq(Rn), (1.1)\n∗Department of Mathematics and Information Science, Zhengz hou University of Light Industry,\nZhengzhou, Henan 450002, P. R. China Email: wangyanqing200 56@gmail.com\n†Corresponding author. Center for Nonlinear Studies, Schoo l of Mathematics, Northwest University,\nXi’an, Shaanxi 710127, P. R. China Email: ww5998198@126.co m\n‡School of Mathematics and Statistics, Henan University, Ka ifeng, Henan 475004, P. R. China. Email:\nylye@vip.henu.edu.cn\n1wherej,mare any integers satisfying 0 ≤j 0,\n/ba∇dblf/ba∇dblLp,α(Rn)≤C/ba∇dblf/ba∇dblθ\nLq,∞(Rn)/ba∇dblΛsf/ba∇dbl1−θ\nL2(Rn)(1.5)\nwith\n1\np=θ\nq+(1−θ)/parenleftbigg1\n2−s\nn/parenrightbigg\n, n/parenleftbigg1\n2−1\np/parenrightbigg\n1 results from an application of Lemma 3.1.\nRemark 1.3.A special case of (1.6) is that\n/ba∇dblu/ba∇dblLp,p1(Rn)≤C/ba∇dblΛsu/ba∇dbl1−θ\nLr,∞(Rn)/ba∇dblu/ba∇dblθ\nLq,∞(Rn),10,\nwheren\np= (1−θ)/parenleftBign\nr−s/parenrightBig\n+nθ\nq,0<θ<1.\nThis result still extends the aforementioned inequalities (1.3)-(1.5) for Lorentz spaces.\nInspiredbythework[13], weshalldividetheproofofTheore m1.1intothreecases bythe\nrelationship between s−σandn/r. Firstly, we focus on the subcritical case s−σ n/r by setting up the following key estimate\n/ba∇dblu/ba∇dblL∞(Rn)≤C/ba∇dblu/ba∇dblθ\nLp,∞(Rn)/ba∇dblu/ba∇dbl1−θ\n˙Bsr,∞,∞,with 0 =θn\np+(1−θ)/parenleftBign\nr−s/parenrightBig\n,0<θ≤1.\nBy a similar argument used in the previous two cases, this yie lds the generalized Gagliardo-\nNirenberg inequality (1.6) under the supcritical case, whi ch concludes Theorem 1.1.\nFurthermore, it should be stated that the Bernstein inequal ities (2.13)-(2.14) together\nwith low-high frequency techniques as in [13] also guarante e the following generalized\nGagliardo-Nirenberg inequality in the framework of Besov- Lorentz spaces, which extends\nthe corresponding results in [13, 21].\nTheorem 1.2. Assume that u∈˙Bs\nr,∞,∞(Rn)∩˙B0\nq,∞,∞(Rn) with 1< q,r≤ ∞and\n0≤σmax{p,r}, and he also proved that\n/ba∇dblu/ba∇dblLq1,q2(Rn)≤C/ba∇dblu/ba∇dblp1\nq1\nLp1,p2(Rn)/vextenddouble/vextenddouble/vextenddoubleΛn\np1u/vextenddouble/vextenddouble/vextenddouble1−p1\nq1\nLp1,p2(Rn),with 14; (1.20)\n(3)v∈Lp(0,T;Lq,∞(R3)),with1\np+3\nq= 1,31, q>1,1\ns1,0max{1,1\np}.\nWe remark that even the stronger version of this condition, w ith space direction in Besov\nspaces˙B5\n2p+3\nq−3\n2q,∞(R3),still improves the results involving nonhomogeneous Besov spaces in\n[12].\nRemark 1.8.According to the boundedness of Riesz transform on Lorentz s paces,∇vin\n(1.22) can be replaced by its symmetrical part vorticity cur lvor its antisymmetric part\n1\n2(∇v−∇vT).\nRemark 1.9.To the best of authors’ knowledge, it remains an open problem to show that\nenergy equality can be derived from the following condition\nv∈L4,∞(0,T;L4,∞(R3)).\nRemark 1.10.Eventually, we would like to mention that an energy conserva tion criterion\nvia a combination of velocity and its gradient for the equati ons (1.16) in T3was recently\nestablished in [43].\nThe rest of this paper is organized as follows. In Section 2, w e recall some basic ma-\nterials of various Lorentz type spaces and present embeddin g relation among these spaces.\nThe generalized Young inequality and Bernstein inequaliti es for Lorentz spaces are also\nestablished in this section. Section 3 is devoted to the proo f of Theorem 1.1 and Theorem\n1.2. Finally, as an application of the above two theorems, we prove Theorem 1.4 in Section\n4, which gives several new criteria for energy conservation of 3D Navier-Stokes equations in\nLorentz spaces.\n62 Notations and key auxiliary lemmas\n2.1 Lorentz spaces and generalized Bernstein inequality\nThroughout this paper, we will use the summation convention on repeated indices. Cwill\ndenote positive absolute constants which may be different fro m line to line unless otherwise\nstated in this paper. a≈bmeans that C−1b≤a≤Cbfor some constant C >1.χΩ\nstands for the characteristic function of a set Ω ⊂Rn.|E|represents the n-dimensional\nLebesgue measure of a set E⊂Rn. LetMbe the Hardy-Littlewood maximal operator and\nits definition is given by\nMf(x) = sup\nr>01\n|B(r)|/integraldisplay\nB(r)|f(x−y)|dy,\nwherefis any locally integrable function on Rn, andB(r) is the open ball centered at the\norigin with radius r>0.\nNext, we present some basic facts on Lorentz spaces. Recall t hat the distribution func-\ntion of a measurable function fon Ω is the function f∗defined on [0 ,∞) by\nf∗(α) =|{x∈Ω :|f(x)|>α}|.\nThe decreasing rearrangement of fis the function f∗defined on [0 ,∞) by\nf∗(t) = inf{α>0 :f∗(α)≤t}.\nForp,q∈(0,∞], we define\n/ba∇dblf/ba∇dblLp,q(Ω)=\n\n/parenleftBig/integraldisplay∞\n0/parenleftBig\nt1\npf∗(t)/parenrightBigqdt\nt/parenrightBig1\nq,ifq<∞,\nsup\nt>0t1\npf∗(t),ifq=∞.\nFurthermore,\nLp,q(Ω) =/braceleftbig\nf:fis a measurable function on Ω and /ba∇dblf/ba∇dblLp,q(Ω)<∞/bracerightbig\n,\nwhich implies that L∞,∞=L∞,Lq,q=LqandL∞,q={0}for 00αf∗(α)1\np,ifq=∞.\nSimilarly, one can define Lorentz spaces Lp,q(0,T;X) in time for 0 < p,q≤ ∞.f∈\nLp,q(0,T;X) means that /ba∇dblf/ba∇dblLp,q(0,T;X)<∞, where\n/ba∇dblf/ba∇dblLp,q(0,T;X)=\n\n/parenleftBig\np/integraldisplay∞\n0αq|{t∈[0,T) :/ba∇dblf(t)/ba∇dblX>α}|q\npdα\nα/parenrightBig1\nq,ifq<∞,\nsup\nα>0α|{t∈[0,T) :/ba∇dblf(t)/ba∇dblX>α}|1\np,ifq=∞.\n7Note that the triangle inequality is not valid for /ba∇dbl·/ba∇dblLp,q(Rn).Another equivalent norm\nin Lorentz spaces is defined as\n/ba∇dblf/ba∇dbl∗\nLp,q(Rn)=\n\n/parenleftbigg/integraldisplay∞\n0/parenleftBig\nt1\npf∗∗(t)/parenrightBigqdt\nt/parenrightbigg1\nq\n,if 10t1\npf∗∗(t),if 10.\nIn addition, Lorentz spaces endowed with the norm /ba∇dbl·/ba∇dbl∗\nLp,qare Banach spaces, and there\nholds\n/ba∇dblf/ba∇dblLp,q(Rn)≤ /ba∇dblf/ba∇dbl∗\nLp,q(Rn)≤p\np−1/ba∇dblf/ba∇dblLp,q(Rn). (2.2)\nMost of the above statement is borrowed from [6, 10, 20].\nSubsequently, we present norm-equivalence concerning Lor entz spaces.\nLemma 2.1. Letfbe inLp,q(Rn)with10,\nwherec1andc2are positive constants depending only on n. This together with (2.1) means\n/ba∇dblf/ba∇dbl∗\nLp,q(Rn)≤C/ba∇dblMf/ba∇dblLp,q(Rn)≤C/ba∇dblf/ba∇dbl∗\nLp,q(Rn).\nThe conclusion is a straightforward consequence of the latt er and (2.2).\nRemark 2.1.Even if 0< q <1, the equivalent relation /ba∇dblMf/ba∇dblLp,q(Rn)≈/ba∇dblf/ba∇dblLp,q(Rn)still\nholds for all functions f∈Lp,q(Rn) with 1< p≤ ∞. Indeed, thanks to Marcinkiewicz’s\ninterpolation theorem for Lorentz spaces [20, Theorem 1.4. 19], it follows from the fact that\nHardy-littlewood maximal operator Mis a sublinear operator of both weak type (1 ,1) and\nstrong type ( ∞,∞) thatMis also bounded on Lp,q(Rn) for anyp∈(1,∞] andq∈(0,∞],\nwhich yields that /ba∇dblMf/ba∇dblLp,q(Rn)≤C(n,p,q)/ba∇dblf/ba∇dblLp,q(Rn)for all functions f∈Lp,q(Rn). On\nthe other hand, Lebesgue’s differentiation theorem implies t hat|f(x)| ≤ Mf(x) for almost\nallx∈Rn, which yields that f∗≤(Mf)∗and/ba∇dblf/ba∇dblLp,q(Rn)≤ /ba∇dblMf/ba∇dblLp,q(Rn)for 1 s. Indeed, here is a counterexample as\nfollows: for any f∈Lq,l(Rn) with 0< s < l ≤ ∞and 1< q <∞, it follows from\n(2.7) and (2.6) that fis a locally integrable function on Rn. Takeg=χQ, whereQ=\n{(y1,y2,...,yn)∈Rn: maxi|yi| ≤1/2}is a cube in Rn. Let (χQ)r(x) =r−nχQ(x/r) for\nallx∈Rnandr >0. Then Lebesgue’s differentiation theorem guarantees that l im\nm→∞f∗\n(χQ)1\nm(x) =f(x) for almost all x∈Rn, which together with [20, Proposition 1.4.5] implies\nthatf∗≤liminf\nm→∞/parenleftBig\nf∗(χQ)1\nm/parenrightBig∗\n. Hence we may apply Fatou’s lemma and (2.11) with r= 1\nto derive that for 0 0defined on Lp0(X) +Lp1(X)and taking values in the set of measurable functions\nonYor a linear operator defined on the set of simple functions on Xand taking values as\nbefore. Assume that for some M0,M1<∞the following (restricted) weak type estimates\nhold:\n/ba∇dblT(χA)/ba∇dblLq0,∞≤M0µ(A)1/p0,\n/ba∇dblT(χA)/ba∇dblLq1,∞≤M1µ(A)1/p1,\nfor all measurable subsets AofXwithµ(A)<∞.Fix0<θ<1and let\n1\np=1−θ\np0+θ\np1and1\nq=1−θ\nq0+θ\nq1.\nThen there exists a positive constant C,which depends only on K,p0,p1,q0,q1,randθ,such\nthat for all functions fin the domain of Tand inLp,r(X)we have\n/ba∇dblT(f)/ba∇dblLq,r≤C(M0+M1)/ba∇dblf/ba∇dblLp,r.\nNow we continue with the proof of Lemma 2.2.\n10Proof of Lemma 2.2. Since (2.6) implies Lp,l(Rn)֒→Lp,s(Rn), it suffices to prove (2.11) for\nthe case when s=l.\nTo this end, fix g∈Lr(Rn). LetT(f) =f∗g, thenTis a linear operator defined on the\nset of simple functions on Rn. For all measurable subsets AofRnwith|A|<∞, it follows\nfrom (2.6) and Young’s inequality for Lebesgue spaces that\n/ba∇dblT(χA)/ba∇dblLr,∞(Rn)≤ /ba∇dblT(χA)/ba∇dblLr(Rn)≤ /ba∇dblg/ba∇dblLr(Rn)/ba∇dblχA/ba∇dblL1(Rn)=/ba∇dblg/ba∇dblLr(Rn)|A|,\n/ba∇dblT(χA)/ba∇dblL∞,∞(Rn)=/ba∇dblT(χA)/ba∇dblL∞(Rn)≤ /ba∇dblg/ba∇dblLr(Rn)/ba∇dblχA/ba∇dblLr′(Rn)=/ba∇dblg/ba∇dblLr(Rn)|A|1/r′,\nwhere 10 depends only on r,qandl. This completes the proof.\nAs an application of this lemma, we may derive the generalize d Bernstein inequality for\nLorentz spaces as follows.\nLemma 2.4. Let a ballB={ξ∈Rn:|ξ| ≤R}with0< R <∞and an annulus C=\n{ξ∈Rn:r1≤ |ξ| ≤r2}with01results from Young\ninequality (2.11) and (2.9) for Lorentz spaces.\n11Proof.(1) Letψbe a Schwartz function on Rnsuch thatχB≤ˆψ≤χ2B. Since/hatwideu(ξ) =\nˆψ(ξ/λ)/hatwideu(ξ) when supp /hatwideu⊂λB, we have\n∂αu=i|α|F−1(ξαˆu) =i|α|F−1(ξαˆψ(ξ/λ)/hatwideu(ξ)) =λ|α|(∂αψ)λ∗u.\nHere (∂αψ)λ(x) =λn∂αψ(λx) for allx∈Rn.\nFix|α|=k. From the H¨ older inequality (2.5) for Lorentz spaces, we in fer that\n/ba∇dbl∂αu/ba∇dblL∞(Rn)≤λksup\nx∈Rn/integraldisplay\nRn|(∂αψ)λ(x−y)||u(y)|dy\n≤Cλksup\nx∈Rn/ba∇dbl(∂αψ)λ(x−·)/ba∇dbl\nLp\np−1,1(Rn)/ba∇dblu/ba∇dblLp,∞(Rn)\n=Cλk+n\np/ba∇dbl∂αψ/ba∇dbl\nLp\np−1,1(Rn)/ba∇dblu/ba∇dblLp,∞(Rn),\nwhereC=C(p)>0. Note that ∂αψ∈ S(Rn), then for any ε>0, there exists a constant\nC=C(ε,n,∂αψ)>0 such that 0 ≤(∂αψ)∗(s)≤Cs−εfor alls>0. This yields that\n/ba∇dbl∂αψ/ba∇dbl\nLp\np−1,1(Rn)=p\np−1/integraldisplay∞\n0(∂αψ)p−1\np\n∗(s)ds≤C(/integraldisplay1\n0s−1\n2ds+/integraldisplay∞\n1s−2ds)<∞,\nwhich implies (2.12).\n(2) Take 1/r= 1+1/q−1/p, then the hypotheses on the indices imply that 1 < r <\nq<∞. In light of (2.9), we see that there exists a positive consta ntC=C(p,q) such that\n/ba∇dbl∂αu/ba∇dblLq,1(Rn)=λk/ba∇dbl(∂αψ)λ∗u/ba∇dblLq,1(Rn)\n≤Cλk/ba∇dbl(∂αψ)λ/ba∇dblLr,1(Rn)/ba∇dblu/ba∇dblLp,∞(Rn)\n=Cλk+n(1−1\nr)/ba∇dbl∂αψ/ba∇dblLr,1(Rn)/ba∇dblu/ba∇dblLp,∞(Rn)\n=Cλk+n/parenleftBig\n1\np−1\nq/parenrightBig\n/ba∇dbl∂αψ/ba∇dblLr,1(Rn)/ba∇dblu/ba∇dblLp,∞(Rn).\nBy a similar argument used in the proof of (2.12), we may deriv e that/ba∇dbl∂αψ/ba∇dblLr,1(Rn)<∞.\nThis implies (2.13).\n(3) As∂αψ∈ S(Rn)⊂L1(Rn), Lemma 2.2 enable us to deduce that\n/ba∇dbl∂αu/ba∇dblLq,l(Rn)=λk/ba∇dbl(∂αψ)λ∗u/ba∇dblLq,l(Rn)\n≤Cλk/ba∇dbl(∂αψ)λ/ba∇dblL1(Rn)/ba∇dblu/ba∇dblLq,l(Rn)\n=Cλk/ba∇dbl∂αψ/ba∇dblL1(Rn)/ba∇dblu/ba∇dblLq,l(Rn),\nwhereC=C(q,l)>0. This implies (2.14).\n(4) Observe that (2.14) implies the second inequality of (2. 15), it suffices to show the\nfirst inequality in (2.15). Let ηbe a Schwartz function on Rnsuch thatχC≤ˆη≤χ˜C, where\n˜C={ξ∈Rn:r1/2≤ |ξ| ≤2r2}. It follows from suppˆ u⊂λCthat for all ξ∈Rn,\nˆu(ξ) =/summationdisplay\n|α|=k(−iξ)α\n|ξ|2kˆη(ξ/λ)(iξ)αˆu(ξ) =λ−k/summationdisplay\n|α|=k(−iξ/λ)α\n|ξ/λ|2kˆη(ξ/λ)F(∂αu)(ξ).\n12Therefore, we may write\nu=λ−k/summationdisplay\n|α|=k(gα)λ∗∂αu,\nwhere (gα)λ(x) =λngα(λx) for allx∈Rn, and\ngα=F−1/parenleftbigg(−iξ)α\n|ξ|2kˆη(ξ)/parenrightbigg\n∈ S(Rn)⊂L1(Rn).\nThis together with Lemma 2.2 yields that a constant C=C(n,q,l,k)>0 exists such that\n/ba∇dblu/ba∇dblLq,l(Rn)≤Cλ−k/summationdisplay\n|α|=k/ba∇dbl(gα)λ/ba∇dblL1(Rn)/ba∇dbl∂αu/ba∇dblLq,l(Rn)\n≤Cλ−k\n/summationdisplay\n|α|=k/ba∇dblgα/ba∇dblL1(Rn)\nsup\n|α|=k/ba∇dbl∂αu/ba∇dblLq,l(Rn).\nThis concludes (2.15) and the proof is complete.\n2.2 Besov-Lorentz spaces, Sobolev-Lorentz spaces and Trie bel-Lizorkin-\nLorentz spaces\nSdenotes the Schwartz class of rapidly decreasing functions ,S′the space of tempered\ndistributions, S′/Pthequotient spaceof tempered distributionswhich modulop olynomials.\nWe useFfor/hatwidefto denote the Fourier transform of a tempered distribution f. To define\nBesov-Lorentz spaces, we need the following dyadic unity pa rtition (see e.g. [1]). Choose\ntwo nonnegative radial functions ̺,ϕ∈C∞(Rn) be supported respectively in the ball\n{ξ∈Rn:|ξ| ≤4\n3}and the shell {ξ∈Rn:3\n4≤ |ξ| ≤8\n3}such that\n̺(ξ)+/summationdisplay\nj≥0ϕ(2−jξ) = 1,∀ξ∈Rn;/summationdisplay\nj∈Zϕ(2−jξ) = 1,∀ξ/\\e}atio\\slash= 0.\nThe nonhomogeneous dyadic blocks ∆ jare defined by\n∆ju:= 0 ifj≤ −2,∆−1u:=̺(D)u,∆ju:=ϕ/parenleftbig\n2−jD/parenrightbig\nuifj≥0, Sju:=/summationdisplay\nk≤j−1∆ku.\nThe homogeneous Littlewood-Paley operators are defined as f ollows\n∀j∈Z,˙∆jf:=ϕ(2−jD)fand˙Sjf:=/summationdisplay\nk≤j−1˙∆kf.\nThe homogeneous Besov-Lorentz space ˙Bs\np,q,r(Rn) is the set of f∈ S′(Rn)/P(Rn) such that\n/ba∇dblf/ba∇dbl˙Bsp,q,r:=/vextenddouble/vextenddouble/vextenddouble/vextenddouble/braceleftbigg\n2js/vextenddouble/vextenddouble/vextenddouble˙∆jf/vextenddouble/vextenddouble/vextenddouble\nLp,q(Rn)/bracerightbigg/vextenddouble/vextenddouble/vextenddouble/vextenddouble\nℓr(Z)<∞.\nHereℓr(Z) represents the set of sequences with summable r-th powers. The homogeneous\nSobolev-Lorentz norm /ba∇dbl·/ba∇dbl˙Hsp,p1(Rn)isdefinedas /ba∇dblf/ba∇dbl˙Hsp,p1(Rn)=/ba∇dblΛsf/ba∇dblLp,p1(Rn).Whenp1=p,\nthe Sobolev-Lorentz spaces reduce to the classical Sobolev spaces˙Hs\np(Rn).\n13Forp,q,r∈(0,∞] ands∈R,the homogeneous Triebel-Lizorkin-Lorentz space\n˙Fs\np,q,r(Rn) is defined by\n˙Fs\np,q,r(Rn) =/braceleftBig\nf∈ S′(Rn)/P(Rn) :/ba∇dblf/ba∇dbl˙Fsp,q,r<∞/bracerightBig\n.\nHere\n/ba∇dblf/ba∇dbl˙Fsp,q,r:=\n\n/ba∇dbl{∞/summationdisplay\nj=−∞(2js|˙∆jf|)r}1\nr/ba∇dblLp,q(Rn),00depends only on n,s,pandϕ.\nRemark 2.5.It is worth remarking that the nonhomogeneous embedding Hs\np,∞֒→Fs\np,∞,∞\nwas recently proved by Ko and Lee in [24] and they mentioned th at the results hold for\nthe homogeneous case. One can modify the argument in [24] to g et the homogeneous case\n˙Hs\np,∞֒→˙Fs\np,∞,∞. Here, we shall present a different proof via the Hardy-Little wood maximal\nfunction.\nProof.Since it is obvious that\n/ba∇dblf/ba∇dbl˙Bsp,∞,∞= sup\nj∈Z/ba∇dbl2js|˙∆jf|/ba∇dblLp,∞(Rn)≤ /ba∇dblsup\nj∈Z2js|˙∆jf|/ba∇dblLp,∞(Rn)=/ba∇dblf/ba∇dbl˙Fsp,∞,∞,\nwe focus on the proof\n/ba∇dblf/ba∇dbl˙Fsp,∞,∞≤C/ba∇dblf/ba∇dbl˙Hsp,∞(Rn).\nIt suffices to show that for all functions g∈Lp,∞(Rn),\n/ba∇dblsup\nj∈Z|φj∗g|/ba∇dblLp,∞(Rn)≤C/ba∇dblg/ba∇dblLp,∞(Rn),\nwhereφ= (|ξ|−sϕ(ξ))∨andφj(x) = 2jnφ(2jx) for allx∈Rn.\nTo this end, observe that φis a Schwartz function on Rnand there exists a positive\nconstantCsuch that for all x∈Rn,|φ(x)| ≤C(1+|x|)−n−1.This yields that for all j∈Z\nandx∈Rn,\n|φj∗g(x)| ≤C2jn/integraldisplay\nRn|g(x−y)|(1+|2jy|)−n−1dy\n=C∞/summationdisplay\nk=−∞2jn/integraldisplay\n2k<|y|≤2k+1|g(x−y)|(1+|2jy|)−n−1dy\n≤C∞/summationdisplay\nk=−∞2jn(1+2j+k)−n−1/integraldisplay\n2k<|y|≤2k+1|g(x−y)|dy\n≤C∞/summationdisplay\nk=−∞2(j+k)n(1+2j+k)−n−1Mg(x)\n14=C∞/summationdisplay\nk=−∞2kn(1+2k)−n−1Mg(x)\n≤CMg(x)(−1/summationdisplay\nk=−∞2kn+∞/summationdisplay\nk=02−k)\n≤CMg(x),\nwhich implies that there exists a positive constant C=C(n,φ) such that for all x∈Rn,\nsup\nj∈Z|φj∗g(x)| ≤CMg(x). (2.17)\nThen it follows from (2.3) that\n/ba∇dblsup\nj∈Z|φj∗g|/ba∇dblLp,∞(Rn)≤C/ba∇dblMg/ba∇dblLp,∞(Rn)≤C/ba∇dblg/ba∇dblLp,∞(Rn),\nwhereC >0 depends only on n,pandφ. This concludes the proof.\n3 Proof of Theorem 1.1 and 1.2\n3.1 Gagliardo-Nirenberg inequalities in Lorentz spaces\nThe goal of this subsection is to prove Theorem 1.1 involving Gagliardo-Nirenberg inequal-\nities in Lorentz spaces. Firstly, we establish three key ine qualities, which play an important\nrole in the proof of three cases in Theorem 1.1. Secondly, in v iew of a pointwise interpo-\nlation estimate for derivatives found in [1] and equivalent norms of Lorentz spaces, we get\nProposition 3.2. Finally, we are in a position to show Theore m 1.1.\nLemma 3.1. (1) Suppose that u∈˙Hs\nr,p1(Rn)with1 n/r and1< p,r≤ ∞.Then there exists a\npositive constant C=C(n,s,p,r)such that\n/ba∇dblu/ba∇dblL∞(Rn)≤C/ba∇dblu/ba∇dblθ\nLp,∞(Rn)/ba∇dblu/ba∇dbl1−θ\n˙Bsr,∞,∞,\nwhere\n0 =θn\np+(1−θ)/parenleftBign\nr−s/parenrightBig\n,0<θ≤1.\n15Remark 3.1.As a corollary of this lemma and imbedding relation in Lemma 2 .5, we have\n/ba∇dblΛσu/ba∇dblLq,l(Rn)≤C/ba∇dblΛσu/ba∇dblp\nq\nLp,∞(Rn)/ba∇dblΛσ+n\nru/ba∇dbl1−p\nq\nLr,∞(Rn)(3.2)\nwith 1n/r and 11 results from Lemma\n2.2, Young inequality (2.9) and Bernstein inequality (2.13 ) for Lorentz spaces. Due to (2.2),\nit is also essential to make the assumption that q >1 andl≥1 to ensure that the space\nLq,l(Rn) is normable.\nProof of Lemma 3.1. (1) Thanks to Fourier transform, there exists a positive con stantC=\nC(n,s) such that\nf=F−1/parenleftBig1\n|ξ|s|ξ|sˆf(ξ)/parenrightBig\n=F−1/parenleftBig1\n|ξ|s/parenrightBig\n∗Λsf=C|·|s−n∗Λsf.\nWiththehelpoftheYounginequality (2.9)inLorentzspaces andthefactthat |x|−1∈Ln,∞,\nwe see that\n/ba∇dblf/ba∇dblLp,p1(Rn)≤C/ba∇dbl|·|s−n∗Λsf/ba∇dblLp,p1(Rn)\n≤C/ba∇dbl|·|s−n/ba∇dblLn\nn−s,∞(Rn)/ba∇dblΛsf/ba∇dblLr,p1(Rn)\n≤C/ba∇dblΛsf/ba∇dblLr,p1(Rn).\n(2) By means of the low and high frequencies, it follows from ( 2.2) that\n/ba∇dblu/ba∇dblLq,l(Rn)≤ /ba∇dblu/ba∇dbl∗\nLq,l(Rn)≤ /ba∇dbl˙Sj0u/ba∇dbl∗\nLq,l(Rn)+/summationdisplay\nj≥j0/ba∇dbl˙∆ju/ba∇dbl∗\nLq,l(Rn)\n≤q\nq−1/ba∇dbl˙Sj0u/ba∇dblLq,l(Rn)+q\nq−1/summationdisplay\nj≥j0/ba∇dbl˙∆ju/ba∇dblLq,l(Rn).(3.4)\nHerej0is an integer to be chosen later. Observe that\n˙Sj0u=/summationdisplay\nk≤j0−1˙∆ku=hj0∗u,\nwherehj0(x) = 2j0nh(2j0x) for allx∈Rn, and\nh=F−1/parenleftBig/summationdisplay\nk≤−1ϕ(2−kξ)/parenrightBig\n∈ S(Rn).\nOwing to Bernstein’s inequality (2.13) and (2.6), we may app ly Lemma 2.2 to infer that\nthere exists a positive constant Cindependent of j0such that\n/ba∇dbl˙Sj0u/ba∇dblLq,l(Rn)≤C2j0n(1\np−1\nq)/ba∇dbl˙Sj0u/ba∇dblLp,∞(Rn)\n=C2j0n(1\np−1\nq)/ba∇dblhj0∗u/ba∇dblLp,∞(Rn)\n≤C2j0n(1\np−1\nq)/ba∇dblhj0/ba∇dblL1(Rn)/ba∇dblu/ba∇dblLp,∞(Rn)\n16=C2j0n(1\np−1\nq)/ba∇dblh/ba∇dblL1(Rn)/ba∇dblu/ba∇dblLp,∞(Rn).\nFor the high-frequency part, it follows from (2.4), (2.6), ( 2.13) and Lemma 2.2 that\n/summationdisplay\nj≥j0/ba∇dbl˙∆ju/ba∇dblLq,l(Rn)≤C/summationdisplay\nj≥j0/ba∇dbl˙∆ju/ba∇dbl1−α\nLp,∞(Rn)/ba∇dbl˙∆ju/ba∇dblα\nLq+r,∞(Rn)\n≤C/summationdisplay\nj≥j02jnα(1\nr−1\nq+r)/ba∇dbl˙∆ju/ba∇dblα\nLr,∞(Rn)/ba∇dblϕj∗u/ba∇dbl1−α\nLp,∞(Rn)\n≤C/summationdisplay\nj≥j02−jnα\nq+r/ba∇dblu/ba∇dblα\n˙Bsr,∞,∞/ba∇dblϕj/ba∇dbl1−α\nL1(Rn)/ba∇dblu/ba∇dbl1−α\nLp,∞(Rn)\n=C2−j0nα\nq+r/ba∇dblϕ/ba∇dbl1−α\nL1(Rn)/ba∇dblu/ba∇dbl1−α\nLp,∞(Rn)/ba∇dblu/ba∇dblα\n˙Bsr,∞,∞.\nHeres=n/r, 1/q= (1−α)/p+α/(q+r) with 0<α<1, andϕj(x) = 2jnϕ(2jx) for all\nx∈Rn. It turns out that\n/ba∇dblu/ba∇dblLq,l(Rn)≤C2j0n(1\np−1\nq)/ba∇dblu/ba∇dblLp,∞(Rn)+C2−j0nα\nq+r/ba∇dblu/ba∇dbl1−α\nLp,∞(Rn)/ba∇dblu/ba∇dblα\n˙Bsr,∞,∞,\nwhere the positive constant Cis independent of j0. Since 1/p−1/q+α/(q+r) =α/p, by\nchoosingj0such that 2j0n(1\np−1\nq)/ba∇dblu/ba∇dblLp,∞(Rn)≈2−j0nα\nq+r/ba∇dblu/ba∇dbl1−α\nLp,∞(Rn)/ba∇dblu/ba∇dblα\n˙Bsr,∞,∞, we may derive\nthat\n/ba∇dblu/ba∇dblLq,l(Rn)≤C/ba∇dblu/ba∇dblp\nq\nLp,∞(Rn)/ba∇dblu/ba∇dbl1−p\nq\n˙Bsr,∞,∞.\n(3) Ifs>n/r, we takeq=l=∞in (3.4). Using Bernstein’s inequality (2.12) and Young\ninequality for Lorentz spaces, we find that\n/ba∇dbl˙Sj0u/ba∇dblL∞(Rn)≤C2j0n\np/ba∇dblu/ba∇dblLp,∞(Rn)and/summationdisplay\nj≥j0/ba∇dbl˙∆ju/ba∇dblL∞(Rn)≤C2j0n(1\nr−s\nn)/ba∇dblu/ba∇dbl˙Bsr,∞,∞,\nwhere the positive constant Cis independent of j0. As the derivation of the above, we may\nchoosej0appropriately to conclude that\n/ba∇dblu/ba∇dblL∞(Rn)≤C2j0n\np/ba∇dblu/ba∇dblLp,∞(Rn)+C2j0n(1\nr−s\nn)/ba∇dblu/ba∇dbl˙Bsr,∞,∞\n≤C/ba∇dblu/ba∇dbl1−1\np\n1\np−1\nr+s\nn\nLp,∞(Rn)/ba∇dblu/ba∇dbl1\np\n1\np−1\nr+s\nn\n˙Bsr,∞,∞.\nThis completes the proof of this lemma.\nProposition 3.2. Assume that u∈Lq,q1(Rn)∩˙Hs\nr,r1(Rn)with1< q,r≤ ∞and1≤\nq1,r1≤ ∞. Then there holds for 0<σ0, the Sobolev embedding (3.1) yields\n/ba∇dblΛσu/ba∇dblLr∗,∞(Rn)≤C/ba∇dblΛsu/ba∇dblLr,∞(Rn), (3.7)\nwith1\nr∗=1\nr−s−σ\nn.\nIt follows from Proposition 3.2 that for 1 0.\nAccording to Proposition 3.2, we arrive at that for1\n˜p=/parenleftbig\n1−σ\ns/parenrightbig1\nq+σ\nsr=/parenleftbig\n1−σ\ns/parenrightbig/parenleftBig\n1\nq+σ\nn/parenrightBig\n,\n/ba∇dblΛσu/ba∇dblL˜p,∞(Rn)≤C/ba∇dblu/ba∇dbl1−σ\ns\nLq,∞(Rn)/ba∇dblΛsu/ba∇dblσ\ns\nLr,∞(Rn). (3.10)\nSince1\np=θ/parenleftBig\n1\nq+σ\nn/parenrightBig\nn/r.\n(III1) Ifσ>0, we conclude from Proposition 3.2 that for 1 0 andn\np−σ=θn\nq+(1−θ)(n\nr−s)\nimply that\ns−σ−n\nr+n\np>0, (3.16)\nwhich will be frequently used later. Indeed, thanks to q≤pandσ>0, we see that\nn\np−(1−θ)σ>n\np−σ=θn\nq+(1−θ)(n\nr−s)≥θn\np+(1−θ)(n\nr−s),\nthat is,\ns−σ−n\nr+n\np>0.\nThe assertion follows. Similarly, we also assert that qk2jσ/ba∇dbl˙∆ju/ba∇dblLp,1(Rn)\n≤C/summationdisplay\nj≤k2j[σ+n(1\nr−1\np)]/ba∇dbl˙∆ju/ba∇dblLr,∞(Rn)+C/summationdisplay\nj>k2−j[s−σ−n(1\nr−1\np)]2js/ba∇dbl˙∆ju/ba∇dblLr,∞(Rn)\n≤C2k[σ+n(1\nr−1\np)]\n1−2−[σ+n(1\nr−1\np)]/ba∇dblu/ba∇dbl˙B0r,∞,∞+C2−k[s−σ−n(1\nr−1\np)]\n1−2−[s−σ−n(1\nr−1\np)]/ba∇dblu/ba∇dbl˙Bsr,∞,∞,\n(3.17)\nwhere we have used σ+n(1\nr−1\np)>0 and (3.16).\nAs a consequence, by choosing the integer kappropriately such that\n2k[σ+n(1\nr−1\np)]/ba∇dblu/ba∇dbl˙B0r,∞,∞≈2−k[s−σ−n(1\nr−1\np)]/ba∇dblu/ba∇dbl˙Bsr,∞,∞, we further get\n/ba∇dblu/ba∇dbl˙Bσ\np,1,1≤C/ba∇dblu/ba∇dblθ\n˙B0r,∞,∞/ba∇dblu/ba∇dbl1−θ\n˙Bsr,∞,∞.\n(II) We turn our attention to the case qk2−j[s−σ−n(1\nr−1\np)]2js/ba∇dbl˙∆ju/ba∇dblLr,∞\n≤C2k[σ+n(1\nq−1\np)]\n1−2−[σ+n(1\nq−1\np)]/ba∇dblu/ba∇dbl˙B0q,∞,∞+C2−k[s−σ−n(1\nr−1\np)]\n1−2−[s−σ−n(1\nr−1\np)]/ba∇dblu/ba∇dbl˙Bsr,∞,∞,\nwhere we have used σ+n(1\nq−1\np)>0 and (3.16).\nHence, we conclude that\n/ba∇dblu/ba∇dbl˙Bσ\np,1,1≤C/ba∇dblu/ba∇dblθ\n˙B0q,∞,∞/ba∇dblu/ba∇dbl1−θ\n˙Bsr,∞,∞.\n(II2) We deal with the case q < p < r . By means of the interpolation characteristic\n(2.4) of Lorentz spaces, we observe that\n/ba∇dbl˙∆ju/ba∇dblLp,1(Rn)≤C/ba∇dbl˙∆ju/ba∇dblα\nLq,∞(Rn)/ba∇dbl˙∆ju/ba∇dbl1−α\nLr,∞(Rn),1\np=α\nq+1−α\nr.\nCombining this, the Bernstein inequality (2.13) in Lorentz spaces and (3.16), we know that\n/ba∇dblu/ba∇dbl˙Bσ\np,1,1=/summationdisplay\nj≤k2jσ/ba∇dbl˙∆ju/ba∇dblLp,1(Rn)+/summationdisplay\nj>k2jσ/ba∇dbl˙∆ju/ba∇dblLp,1(Rn)\n≤C/summationdisplay\nj≤k2j[σ+n(1\nq−1\np)]/ba∇dbl˙∆ju/ba∇dblLq,∞(Rn)+C/summationdisplay\nj>k2jσ/ba∇dbl˙∆ju/ba∇dblα\nLq,∞(Rn)/ba∇dbl˙∆ju/ba∇dbl1−α\nLr,∞(Rn)\n≤C2k[σ+n(1\nq−1\np)]\n1−2−[σ+n(1\nq−1\np)]/ba∇dblu/ba∇dbl˙B0q,∞,∞+C2−k[s(1−α)−σ]\n1−2−[s(1−α)−σ]/ba∇dblu/ba∇dblα\n˙B0q,∞,∞/ba∇dblu/ba∇dbl1−α\n˙Bsr,∞,∞,\n(3.19)\nwhere we have used σ+n(1\nq−1\np)>0 ands(1−α)−σ >0 which is derived from (3.18).\nChoosingtheinteger ksuch that 2k[σ+n(1\nq−1\np)]/ba∇dblu/ba∇dbl˙B0q,∞,∞≈2−k[s(1−α)−σ]/ba∇dblu/ba∇dblα\n˙B0q,∞,∞/ba∇dblu/ba∇dbl1−α\n˙Bsr,∞,∞,\nwe also have\n/ba∇dblu/ba∇dbl˙Bσ\np,1,1≤C/ba∇dblu/ba∇dblθ\n˙B0q,∞,∞/ba∇dblu/ba∇dbl1−θ\n˙Bsr,∞,∞.\n(II3) We treat the case q0 will be determined later. We derive from this, the Bernstei n inequality (2.13)\n22in Lorentz spaces and (3.16) that\n/ba∇dblu/ba∇dbl˙Bσ\np,1,1=/summationdisplay\nj≤k2jσ/ba∇dbl˙∆ju/ba∇dblLp,1(Rn)+/summationdisplay\nj>k2jσ/ba∇dbl˙∆ju/ba∇dblLp,1(Rn)\n≤C/summationdisplay\nj≤k2j[σ+n(1\nq−1\np)]/ba∇dbl˙∆ju/ba∇dblLq,∞(Rn)+C/summationdisplay\nj>k2jσ/ba∇dbl˙∆ju/ba∇dbl1−α\nLq,∞(Rn)/ba∇dbl˙∆ju/ba∇dblα\nL(1+ε)p,∞(Rn)\n≤C2k[σ+n(1\nq−1\np)]\n1−2−[σ+n(1\nq−1\np)]/ba∇dblu/ba∇dbl˙B0q,∞,∞+C/summationdisplay\nj>k2−j[sα−nεα\n(1+ε)p−σ]/ba∇dblu/ba∇dbl1−α\n˙B0q,∞,∞/ba∇dblu/ba∇dblα\n˙Bsr,∞,∞,\n(3.21)\nwhere we have used the fact that σ+n(1\nq−1\np)>0.\nDenoteδ(ε) =sα−nεα\n(1+ε)p−σ. From (3.20), we see that\nδ(ε) =ps(1+ε)(p−q)−σp[(1+ε)p−q]−εn(p−q)\np[(1+ε)p−q],\nandδ(ε) is a continuous function on a neighborhood of 0. Since δ(0)>0, there exists a\nsufficiently small ε>0 such that δ(ε)>0. It follows from (3.21) that\n/ba∇dblu/ba∇dbl˙Bσ\np,1,1≤C2k[σ+n(1\nq−1\np)]\n1−2−[σ+n(1\nq−1\np)]/ba∇dblu/ba∇dbl˙B0q,∞,∞+C2−k[sα−nεα\n(1+ε)p−σ]\n1−2−[sα−nεα\n(1+ε)p−σ]/ba∇dblu/ba∇dbl1−α\n˙B0q,∞,∞/ba∇dblu/ba∇dblα\n˙Bsr,∞,∞,\n(3.22)\nwhich also yields that\n/ba∇dblu/ba∇dbl˙Bσ\np,1,1≤C/ba∇dblu/ba∇dblθ\n˙B0q,∞,∞/ba∇dblu/ba∇dbl1−θ\n˙Bsr,∞,∞.\n(III) Finally, it remains to show (1.10) under the case that q>r.\n(III1) We first consider (1.10) under the hypothesis that rk2−j[s−σ−n(1\nr−1\np)]2js/ba∇dbl˙∆ju/ba∇dblLr,∞(Rn)\n≤C2k[σ+n(1\nq−1\np)]\n1−2−[σ+n(1\nq−1\np)]/ba∇dblu/ba∇dbl˙B0q,∞,∞+C2−k[s−σ−n(1\nr−1\np)]\n1−2−[s−σ−n(1\nr−1\np)]/ba∇dblu/ba∇dbl˙Bsr,∞,∞,\nwhere we have used σ+n(1\nq−1\np)>0 and (3.16).\nTherefore, we conclude that\n/ba∇dblu/ba∇dbl˙Bσ\np,1,1≤C/ba∇dblu/ba∇dblθ\n˙B0q,∞,∞/ba∇dblu/ba∇dbl1−θ\n˙Bsr,∞,∞.\n(III12) We need to show (1.10) under the hypothesis that r0 in this case. In the same manner as (3.20), we\nsee that\n/ba∇dbl˙∆ju/ba∇dblLp,1(Rn)≤C/ba∇dbl˙∆ju/ba∇dbl1−α\nLr,∞(Rn)/ba∇dbl˙∆ju/ba∇dblα\nL(1+ε)p,∞(Rn),1\np=1−α\nr+α\n(1+ε)p,(3.23)\n23whereε>0 will be determined later. Then we obtain\n/ba∇dblu/ba∇dbl˙Bσ\np,1,1≤C/summationdisplay\nj≤k2jσ/ba∇dbl˙∆ju/ba∇dbl1−α\nLr,∞(Rn)/ba∇dbl˙∆ju/ba∇dblα\nL(1+ε)p,∞(Rn)+C/summationdisplay\nj>k2−j[s−σ−n(1\nr−1\np)]2js/ba∇dbl˙∆ju/ba∇dblLr,∞(Rn)\n≤C/summationdisplay\nj≤k2j[σ+nεα\np(1+ε)−s(1−α)]/ba∇dblu/ba∇dblα\n˙B0q,∞,∞/ba∇dblu/ba∇dbl1−α\n˙Bsr,∞,∞+C2−k[s−σ−n(1\nr−1\np)]\n1−2−[s−σ−n(1\nr−1\np)]/ba∇dblu/ba∇dbl˙Bsr,∞,∞,\nAs the arguments in ( II3), we can choose ε>0 sufficiently small to ensure that σ+nεα\np(1+ε)−\ns(1−α)>0. This yields the desired inequality (1.10). We omit the det ails.\n(III2) Letrk2jσ/ba∇dbl˙∆ju/ba∇dblLp,1(Rn)\n≤C/summationdisplay\nj≤k2jσ/ba∇dbl˙∆ju/ba∇dblα\nLq,∞(Rn)/ba∇dbl˙∆ju/ba∇dbl1−α\nLr,∞(Rn)+C/summationdisplay\nj>k2−j[s−σ−n(1\nr−1\np)]2js/ba∇dbl˙∆ju/ba∇dblLr,∞(Rn)\n≤C2k[σ−s(1−α)]\n1−2−[σ−s(1−α)]/ba∇dblu/ba∇dblα\n˙B0q,∞,∞/ba∇dblu/ba∇dbl1−α\n˙Bsr,∞,∞+C2−k[s−σ−n(1\nr−1\np)]\n1−2−[s−σ−n(1\nr−1\np)]/ba∇dblu/ba∇dbl˙Bsr,∞,∞.\nHere we need the fact that σ−s(1−α)>0 with 1/p=α/q+(1−α)/r, which is derived\nfrom the hypothesis that1\np<(1−σ\ns)1\nq+σ\ns1\nr. Therefore, we obtain the desired estimate\n/ba∇dblu/ba∇dbl˙Bσ\np,1,1≤C/ba∇dblu/ba∇dblθ\n˙B0q,∞,∞/ba∇dblu/ba∇dbl1−θ\n˙Bsr,∞,∞.\n(III22) Now, it remains to show (1.10) with r(1−σ\ns)1\nq+σ\ns1\nr, which imply\nthat\np0 and1\np=α\nq+1−α\nr. (3.25)\nIt follows from (3.24) and1\np>(1−σ\ns)1\nq+σ\ns1\nrthat\n1\np<1\nr−s−σ\nn<1\nr,\nwhich together with p0 and 0<α<1. (3.26)\n24Making use of the Bernstein inequality (2.13) for Lorentz sp aces, (2.4), (3.25) and (3.26),\nwe arrive at\n/ba∇dblu/ba∇dbl˙Bσ\np,1,1=/summationdisplay\nj≤k2jσ/ba∇dbl˙∆ju/ba∇dblLp,1(Rn)+/summationdisplay\nj>k2jσ/ba∇dbl˙∆ju/ba∇dblLp,1(Rn)\n≤C/summationdisplay\nj≤k2j[σ−s+n(1\nr−1\np)]2js/ba∇dbl˙∆ju/ba∇dblLr,∞(Rn)+C/summationdisplay\nj>k2jσ/ba∇dbl˙∆ju/ba∇dblα\nLq,∞(Rn)/ba∇dbl˙∆ju/ba∇dbl1−α\nLr,∞(Rn)\n≤C2k[σ−s+n(1\nr−1\np)]\n1−2−[σ−s+n(1\nr−1\np)]/ba∇dblu/ba∇dbl˙Bsr,∞,∞+C2−k[s(1−α)−σ]\n1−2−[s(1−α)−σ]/ba∇dblu/ba∇dblα\n˙B0q,∞,∞/ba∇dblu/ba∇dbl1−α\n˙Bsr,∞,∞.\nWe thereby deduce the inequality\n/ba∇dblu/ba∇dbl˙Bσ\np,1,1≤C/ba∇dblu/ba∇dblθ\n˙B0q,∞,∞/ba∇dblu/ba∇dbl1−θ\n˙Bsr,∞,∞.\nThe proof of this theorem is completed.\n4 Proof of Theorem 1.4\nThissection is concernedwiththeapplication ofGagliardo -Nirenberg inequalities inLorentz\ntype spaces to the energy conservation of 3D Navier-Stokes e quations. We shall follow the\npath of [12] to prove (1.20).\nProof of Theorem 1.4. As in [11, 12], since Leray-Hopf weak solutions vsatisfy (1.16) in\nthe sense of distributions, there holds for any Q∈Z,\n1\n2/ba∇dblSQv(T)/ba∇dbl2\nL2(R3)+/integraldisplayT\n0/ba∇dbl∇SQv/ba∇dbl2\nL2(R3)ds=1\n2/ba∇dblSQv0/ba∇dbl2\nL2(R3)+/integraldisplayT\n0/integraldisplay\nTr(SQ(v⊗v)·∇SQv)dxds.\nIn order to get the energy equality, it is enough to show/integraltextT\n0/integraltext\nTr(SQ(v⊗v)·∇SQv)dxds→0\nasQ→ ∞. To this end, we recall the following estimates proved in [11 , 12]\n/integraldisplayT\n0/integraldisplay\n|Tr(SQ(v⊗v)·∇SQv)|dxds\n≤C/integraldisplayT\n0\n/summationdisplay\nkQ\n22−2|k−Q|\n3/parenleftBig/integraldisplayT\n022k/ba∇dbl∆kv/ba∇dbl2\nL2(R3)ds/parenrightBig1\n2\n≤/parenleftBig/summationdisplay\nk>Q\n22−4|k−Q|\n3/parenrightBig1\n2/parenleftBig/summationdisplay\nk>Q\n2/integraldisplayT\n022k/ba∇dbl∆kv/ba∇dbl2\nL2(R3)ds/parenrightBig1\n2\n≤/parenleftBig\n2∞/summationdisplay\nk=Q2−4|k−Q|\n3/parenrightBig1\n2/parenleftBig/summationdisplay\nk>Q\n2/integraldisplayT\n022k/ba∇dbl∆kv/ba∇dbl2\nL2(R3)ds/parenrightBig1\n2\n≤2/parenleftBig/summationdisplay\nk>Q\n2/integraldisplayT\n022k/ba∇dbl∆kv/ba∇dbl2\nL2(R3)ds/parenrightBig1\n2→0,asQ→ ∞.(4.4)\nOn the other hand, we observe that\n/summationdisplay\nk≤Q\n22−2|k−Q|\n3/parenleftBig/integraldisplayT\n022k/ba∇dbl∆kv/ba∇dbl2\nL2(R3)ds/parenrightBig1\n2\n≤/parenleftBig1\n2/parenrightBig5\n3+2Q\n3/parenleftBig/integraldisplayT\n0/ba∇dbl∆−1v/ba∇dbl2\nL2(R3)ds/parenrightBig1\n2+4\n3/summationdisplay\n0≤k≤Q\n22−2|k−Q|\n3/parenleftBig/integraldisplayT\n0/ba∇dbl∇v/ba∇dbl2\nL2(R3)ds/parenrightBig1\n2(4.5)\n26≤/parenleftBig1\n2/parenrightBig5\n3+2Q\n3/parenleftBig/integraldisplayT\n0/ba∇dblv/ba∇dbl2\nL2(R3)ds/parenrightBig1\n2+/parenleftBig1\n2/parenrightBigQ\n3−8\n3/parenleftBig/integraldisplayT\n0/ba∇dbl∇v/ba∇dbl2\nL2(R3)ds/parenrightBig1\n2→0,asQ→ ∞.\nHere we have used the fact that\n/integraldisplayT\n0/ba∇dbl∆−1v/ba∇dbl2\nL2(R3)ds=/integraldisplayT\n0/ba∇dbl̺ˆv/ba∇dbl2\nL2(R3)ds≤/integraldisplayT\n0/ba∇dblˆv/ba∇dbl2\nL2(R3)ds=/integraldisplayT\n0/ba∇dblv/ba∇dbl2\nL2(R3)ds<∞.\nCombining (4.4) and (4.5), we thereby conclude that\n/summationdisplay\nk2−2|k−Q|\n3/parenleftBig/integraldisplayT\n022k/ba∇dbl∆kv/ba∇dbl2\nL2(R3)ds/parenrightBig1\n2→0,asQ→ ∞. (4.6)\n(2) Letq>4. Before going further, we write\nIQ=:{s∈[0,T] :/ba∇dblv(s)/ba∇dblLq,∞(R3)>2Q(q−2)\nq}\nandIc\nQ=: [0,T]\\IQ.\nAccording to the interpolation inequality (2.4) and (2.11) , we infer that\n/ba∇dbl∆kv/ba∇dblL3(R3)≤C/ba∇dbl∆kv/ba∇dbl2q−6\n3(q−2)\nL2(R3)/ba∇dbl∆kv/ba∇dblq\n3(q−2)\nLq,∞(R3)\n≤C/ba∇dbl∆kv/ba∇dbl2q−6\n3(q−2)\nL2(R3)/ba∇dblv/ba∇dblq\n3(q−2)\nLq,∞(R3).(4.7)\nPlugging this into (4.1) and using the H¨ older inequality, w e observe that\n/integraldisplay\nIQ/summationdisplay\nk2k/ba∇dbl∆kv/ba∇dbl3\nL3(R3)2−2|k−Q|\n3ds\n≤C/integraldisplay\nIQ/summationdisplay\nk2k2−2|k−Q|\n3/ba∇dbl∆kv/ba∇dbl2q−6\nq−2\nL2(R3)/ba∇dblv/ba∇dblq\nq−2\nLq,∞(R3)ds\n≤C/summationdisplay\nksup\nt∈[0,T]/ba∇dbl∆kv/ba∇dbl4q−14\n3(q−2)\nL2(R3)/integraldisplay\nIQ2−2|k−Q|\n322k\n3/ba∇dbl∆kv/ba∇dbl2\n3\nL2(R3)2k\n3/ba∇dblv/ba∇dblq\nq−2\nLq,∞(R3)ds\n≤Csup\nt∈[0,T]/ba∇dblv/ba∇dbl4q−14\n3(q−2)\nL2(R3)/summationdisplay\nk2−2|k−Q|\n3/parenleftBig/integraldisplay\nIQ22k/ba∇dbl∆kv/ba∇dbl2\nL2(R3)ds/parenrightBig1\n32k\n3/parenleftBig/integraldisplay\nIQ/ba∇dblv/ba∇dbl3q\n2(q−2)\nLq,∞(R3)ds/parenrightBig2\n3.(4.8)\nThe hypothesis (1.20) enables us to obtain\nf∗(λ) =|{s∈[0,T] :/ba∇dblv(s)/ba∇dblLq,∞(R3)>λ}| ≤Cλ−2q\nq−2. (4.9)\nThis yields that\n2k\n3/parenleftBig/integraldisplay\nIQ/ba∇dblv/ba∇dbl3q\n2(q−2)\nLq,∞(R3)ds/parenrightBig2\n3=2k\n3/parenleftBig3q\n2q−4/parenrightBig2\n3/parenleftBig/integraldisplay∞\n2Q(q−2)\nqλ3q\n2(q−2)−1f∗(λ)dλ+/integraldisplay2Q(q−2)\nq\n0λ3q\n2(q−2)−1|IQ|dλ/parenrightBig2\n3\n≤C2k\n3/parenleftBig/integraldisplay∞\n2Q(q−2)\nqλ3q\n2(q−2)−1−2q\nq−2dλ+23Q\n2f∗(2Q(q−2)\nq)/parenrightBig2\n3\n≤C2k\n3/parenleftBig/integraldisplay∞\n2Q(q−2)\nqλ−q\n2(q−2)−1dλ+2−Q\n2/parenrightBig2\n3\n≤C2k−Q\n3.\n(4.10)\n27Inserting the latter inequality into (4.8), we get\n/integraldisplay\nIQ/summationdisplay\nk2k/ba∇dbl∆kv/ba∇dbl3\nL3(R3)2−2|k−Q|\n3ds\n≤C(/ba∇dblv0/ba∇dblL2(R3))/summationdisplay\nk2−|k−Q|\n3/parenleftBig/integraldisplayT\n022k/ba∇dbl∆kv/ba∇dbl2\nL2(R3)ds/parenrightBig1\n3.(4.11)\nIn the same manner as derivation of (4.6), we arrive at\n/summationdisplay\nk2−|k−Q|\n3/parenleftBig/integraldisplayT\n022k/ba∇dbl∆kv/ba∇dbl2\nL2(R3)ds/parenrightBig1\n3→0,asQ→ ∞.\nNow, it suffices to show/integraltext\nIc\nQ/summationtext\nk2k/ba∇dbl∆kv/ba∇dbl3\nL3(R3)2−2|k−Q|\n3ds→0 asQ→ ∞.\nA slight modification of deduction of (4.8) ensures that for a nyε∈(0, q−4],\n/integraldisplay\nIc\nQ/summationdisplay\nk2k/ba∇dbl∆kv/ba∇dbl3\nL3(R3)2−2|k−Q|\n3ds\n≤C/integraldisplay\nIc\nQ/summationdisplay\nk2k2−2|k−Q|\n3/ba∇dbl∆kv/ba∇dbl2q−6\nq−2\nL2(R3)/ba∇dblv/ba∇dblq\nq−2\nLq,∞(R3)ds\n≤C/summationdisplay\nksup\nt∈[0,T]/ba∇dbl∆kv/ba∇dbl2q−8−2ε\n(2+ε)(q−2)\nL2(R3)/integraldisplay\nIc\nQ2−2|k−Q|\n322(1+ε)k\n2+ε/ba∇dbl∆kv/ba∇dbl2(1+ε)\n2+ε\nL2(R3)2−εk\n2+ε/ba∇dblv/ba∇dblq\nq−2\nLq,∞(R3)ds\n≤Csup\nt∈[0,T]/ba∇dblv/ba∇dbl2q−8−2ε\n(2+ε)(q−2)\nL2(R3)/summationdisplay\nk2−2|k−Q|\n3/parenleftBig/integraldisplay\nIc\nQ22k/ba∇dbl∆kv/ba∇dbl2\nL2(R3)ds/parenrightBig1+ε\n2+ε2−εk\n2+ε/parenleftBig/integraldisplay\nIc\nQ/ba∇dblv/ba∇dblq(2+ε)\nq−2\nLq,∞(R3)ds/parenrightBig1\n2+ε.\n(4.12)\nThen it follows from (4.9) that\n2−εk\n2+ε/parenleftBig/integraldisplay\nIc\nQ/ba∇dblv/ba∇dblq(2+ε)\nq−2\nLq,∞(R3)ds/parenrightBig1\n2+ε≤2−εk\n2+ε/parenleftBigq(2+ε)\nq−2/integraldisplay2Q(q−2)\nq\n0λq(2+ε)\nq−2−1f∗(λ)dλ/parenrightBig1\n2+ε\n≤C2−εk\n2+ε/parenleftBig/integraldisplay2Q(q−2)\nq\n0λq(2+ε)\nq−2−1−2q\nq−2dλ/parenrightBig1\n2+ε\n≤C2−εk\n2+ε/parenleftBig/integraldisplay2Q(q−2)\nq\n0λqε\nq−2−1dλ/parenrightBig1\n2+ε\n≤C2ε(Q−k)\n2+ε.\nInserting the latter inequality into (4.12), we get\n/integraldisplay\nIc\nQ/summationdisplay\nk2k/ba∇dbl∆kv/ba∇dbl3\nL3(R3)2−2|k−Q|\n3ds\n≤C(/ba∇dblv0/ba∇dblL2(R3))/summationdisplay\nk2−(4−ε)|k−Q|\n3(2+ε)/parenleftBig/integraldisplayT\n022k/ba∇dbl∆kv/ba∇dbl2\nL2(R3)ds/parenrightBig1+ε\n2+ε.(4.13)\nTake the positive constant ε 3/2 that\n1/˜p= 2−3(1−α)/2−3α/q>(1−α)/2.\nThanks to the interpolation characteristic (2.4) of Lorent z spaces, (2.6) and the H¨ older\ninequality, we arrive at\n/integraldisplayT\n0/ba∇dbl∇v/ba∇dbl˜p\nL˜q(R3)ds≤C/integraldisplayT\n0/ba∇dbl∇v/ba∇dbl(1−α)˜p\nL2(R3)/ba∇dbl∇v/ba∇dblα˜p\nLq,∞(R3)ds\n≤C/parenleftBig/integraldisplayT\n0/ba∇dbl∇v/ba∇dbl2\nL2(R3)ds/parenrightBig(1−α)˜p\n2/parenleftBig/integraldisplayT\n0/ba∇dbl∇v/ba∇dblp\nLq,∞(R3)ds/parenrightBig1−(1−α)˜p\n2,\nwhere we have used the fact that\n2α˜p\n2−(1−α)˜p=p,and 0<(1−α)˜p\n2<1.\nHence, we conclude the desired energy equality from the know n result (1.18).\n(5) Note that 1 /s < p < 3,we may choose an index p1such that max {1,p}< p1<\nmin{3,sp}.Takeθ= 1−p/p1,then 0<θ<1−1/s.Let 1/p1+6/(5q1) = 1,which implies\nthat 9/50such that N(Tm) =N(Tm+1), the\nsmallest such integer is called the ascent of Tand we denote it by A(T). If no such\nintegermexists such that N(Tm) =N(Tm+1), we say that ascent of Tis infinite\ni.e.,A(T) =∞.\nDefinition 2. If there is an integer m >0such that R(Tm+1) =R(Tm), the\nsmallest such integer is called the descent of T and is denote d byD(T). If there\nis no such integer msuch that R(Tm+1) =R(Tm), we say that descent of Tis\ninfinite i.e., D(T) =∞.\nMany authors have studied the characterizations of composition a nd weighted\ncomposition operators with ascent and descent on different funct ion spaces such\nasLpspace,lpspace, Orlicz space, Lorentz space [[2], [4], [5], [6]]. Motivated\nby their work, we have discussed the characterization of composit ion operators on\nOrlicz-Lorentz spaces L(φ,ω)to have finite ascent and descent.\n2010Mathematics Subject Classification. 47B33, 47B38, 46E30.\nKey words and phrases. Ascent,composition operators, descent, Orlicz-Lorentz s paces, non-\nsingular transformation, Radon-Nikodym derivative.\nSubmitted Month Day, 20xx. Published Month Day, 20xx.\n12 NEHA BHATIA AND ANURADHA GUPTA\n2.Composition Operators on Orlicz-Lorentz Spaces\nLet(E,E,ν)beaσ-finitemeasurespaceand fbeanycomplex-valuedmeasurable\nfunction on the space.\nFors≥0, define νfthedistribution function as\nνf(s) =ν{x∈E:|f(x)|> s}\nClearly,νfis decreasing. By f∗we mean the non-increasing rearrangement off\ngiven as\nf∗(t) = inf{s >0 :µf(s)≤t}, t≥0\nf∗is a non-negative and decreasing function.\nBy a weight function ω, we mean ω:J→JwhereJ= (0∞) a non-increasing\nlocally integrable function such that/integraltext∞\n0ω(t)dt=∞.\nLetφ: [0∞)→[0∞) be a convex function such that\nφ(x) = 0⇔x= 0,\nlim\nx→∞φ(x) = +∞.\nSuch a function is called a young function . The young function is said to satisfy\nthe ∆2-condition if for some k >0,\nφ(2x)≤kφ(x),for allx >0.\nIfφsatisfies ∆ 2-condition, then we define the space Lφ,ωas\n/braceleftbigg\nf:E→Cmeasurable :/integraldisplay∞\n0φ(αf∗(t))ω(t)dt <∞for some α >0/bracerightbigg\nThe space L(φ,ω)is called an Orlicz-Lorentz space and is a Banach space with\nrespect to the Luxemburg norm given by\n/bardblf/bardblφ,ω= inf/braceleftbigg\nǫ >0 :/integraldisplay∞\n0φ(|f∗(t)|/ǫ)ω(t)dt≤1/bracerightbigg\nOrlicz-Lorentzspaceis commongeneralizationofOrliczspaceand Lo rentzspace.\nIfw(t) = 1 then it becomes an Orlicz space. If φ(x) =xp, 1≤p <∞, then it\nbecomes a Lorentz space L(p,q). LetB(L(φ,ω)) denote the Banach algebra of all\nbounded linear operators on L(φ,ω). For more details on Orlicz-Lorentz spaces, we\nrefer to [1], [7].\nLetΨ :E→Ebeameasurablenon-singulartransformation(i.e., ν(Ψ−1(S)) = 0\nwhenever ν(S) = 0 for S∈ E).\nIf Ψ is non-singular, then we say that νΨ−1is absolutely continuous with re-\nspect to ν. Hence, by Radon-Nikodym theorem, there exists a unique non-ne gative\nmeasurable function gsuch that\n(νΨ−1)(S) =/integraldisplay\nSgdν,forS∈ E.\ngis called the Radon-Nikodym derivative and is denoted bydνΨ−1\ndν. Form≥2,\nwe observe that νm\nΨ≪νm−1\nΨ≪ ···νΨ. Hence\nνm\nΨ=/integraldisplay\nSfΨdν for S ∈ E.\nDefinition 3. Measures ν1andν2are said to be equivalent if ν2≪ν1≪ν2.ASCENT AND DESCENT OF COMPOSITION OPERATORS ON ORLICZ-LORE NTZ SPACES3\nDefinition 4. A measurable transformation Ψis said to be measure preserving if\nit preserves the measure i.e., ν(Ψ−1(A)) =ν(A)for allA∈ E.\nDefinition 5. Let(E,E,ν)be a measure space. A measurable transformation Ψ :\nE→Eis said to be pre-positive if it satisfies the condition ν(Ψ−1(A))>0whenever\nν(A)>0.\nLet Ψ be a measurable transformation then the composition operat orCΨ[8]\ninduced by Ψ is given by\nCΨ=f◦Ψ,for every f∈L(φ,w).\nLet Ψ :E→Eis a non- singular measurable transformation, then Ψmis also\na non-singular measurable transformation for every non-negativ e integer mwith\nrespect to the measure ν. Thus, we can define the composition operator CΨmon\nOrlicz-Lorentz space L(φ,w), such that Cm\nΨ(fΨ) =fΨoΨm=CΨm(fΨ) for every\nmeasurable function fΨof the Orlicz-Lorentz space. Also, the measure νoΨ−mis\ndefined as\nνoΨ−m(S) =νoΨ−(m−1)(Ψ−1(S)) =νoΨ−1(Ψ−(m−1)(S))for S∈ E.\nThen\n...≪ν oΨ−(m+1)≪ν oΨ−m≪νΨ−(m−1)≪...≪ν oΨ−1≪ν.(2.1)\nIf we put νm=ν oΨ−m, then by Radon - Nikodyn theorem, there exists a\nnon-negative locally integrable function fΨmonEsatisfying\nνm(S) =/integraldisplay\nSfΨm(x)dν(x)for S∈ E. (2.2)\nHere,fΨm/parenleftbig\n=dνm\ndν/parenrightbig\nis called the Radon -Nikodym derivative of νmwith respect\ntoν.\nIn [1], the necessary and sufficient condition for the boundedness o f the com-\nposition operators on Orlicz-Lorentz spaces Lφ,ware discussed and given by the\nfollowing theorem:\nTheorem 2.1. LetΨ :E→Ebe a non-singular measurable transformation. Then\nthe composition operator CΨis bounded on Lφ,wif and only if there exists a constant\nK >0such that\nνΨ−1(A)≤Kν(A)for A∈ E.\nAlso, the measurable transformation Ψ is said to be bounded away fr om zero if\nthere exists a positive real number ǫ, such that f(x)≥ǫfor almost all x∈Swhere\nS={x∈E:f(x)/ne}ationslash= 0}.\n3.Ascent of the Composition Operators\nTheorem 3.1. LetCΨbe a composition operator on L(φ,w). Then N(Cm\nΨ) =\nL(φ,w)(Em)form≥0i.e., the kernel is the collection of all the measurable func tions\nfΨinL(φ,w)satisfying fΨ(x) = 0forx∈E/Em, whereEm={x∈E:fΨm(x) =\n0}.\nProof.Letfbe an element of L(φ,w)(Em). Then\nν{x∈E:f oΨm(x)/ne}ationslash= 0} ≤ν oΨ−m(Em) =/integraldisplay\nEmfΨm(x)dν(x) = 0.4 NEHA BHATIA AND ANURADHA GUPTA\nThis implies that f◦Ψm= 0 a.e. This gives that fΨ∈ N(Cm\nΨ). This implies that\nL(φ,w)(Em)⊆ N(Cm\nΨ). (3.1)\nAlso, let fΨ∈ N(Cm\nΨ), thenν oΨ−m{x∈E:fΨ(x)/ne}ationslash= 0}= 0.TakeF={x∈\nE/Em:fΨ(x)/ne}ationslash= 0}andG={x∈Em:fΨ(x)/ne}ationslash= 0}. From (2.2), we have\n0 =/integraldisplay\nFfΨm(x)dν(x) +/integraldisplay\nGfΨm(x)dν(x)\n=/integraldisplay\nFhΨm(x)dν(x)\n≥1\nn/integraldisplay\nFn∩Fdν\n=1\nnν(Fn∩F)\nfor each nwhereFn={x∈E/Em:fΨ>1\nn} ⊆F. SinceFn∩F=F, this gives\nthatν(Fn∩F) =ν(F) for each n. Therefore ν(F) = 0. This implies that f= 0\na.e. onE/Em. Hencef∈L(φ,w)(Em). This gives that\nN(Cm\nΨ)⊆L(φ,w)(Em). (3.2)\nFrom (3.1) and (3.2), we obtain\nN(Cm\nΨ) =L(φ,w)(Em).\n/square\nLemma 3.2. LetΨbe a non-singular measurable transformation on the measure\nspace(E,E,ν)that induces the composition operator CΨon the Orlicz-Lorentz space\nL(φ,w). ThenN(Cm\nΨ) =N(Cm+1\nΨ)if and only if νmandνm+1are equivalent.\nProof.LetCΨ∈ B(L(φ,w)). Suppose that νmandνm+1are equivalent. Then\nνm+1≪νm≪νm+1. Sofrom(2 .1), wehave νm≪νm+1≪νandνm+1≪νm≪ν,\nand hence, by the chain rule,\nfΨm(x) =dνm\ndνm+1(x).fΨm+1(x)\nfΨm+1(x) =dνm+1\ndνm(x).fΨm(x)\nThis gives that Em=Em+1. SinceN(Cm\nΨ) =L(φ,w)(Em) form≥0, therefore\nN(Cm\nΨ) =L(φ,w)(Em) =L(φ,w)(Em+1) =N(Cm+1\nΨ).\nConversely,supposethat N(Cm\nΨ) =N(Cm+1\nΨ). Thisgives Lφ,w(Em) =Lφ,w(Em+1).\nWe first claim that ν(Em/Em+1) = 0. The assumption of ν(Em/Em+1)>0 pro-\nvides a set Fn={x∈Em:fΨm+1(x)>1\nn,n∈N}of non-zero finite measure.\nCharacteristic functions χSare dense in Orlicz-Lorentz space for S∈ E. Now,\n/bardblχFn/bardbl=1\nφ−1(1\nν(Fn)).\nThus,χFn∈L(φ,w)(Em) =L(φ,w)(Em+1) i.e.,χFnvanishes outside Em+1which\ngives that Fn⊆Em+1. Therefore,\n0≤1\nn/integraldisplay\nFnχFndν≤/integraldisplay\nFnfΨm+1dν= 0.ASCENT AND DESCENT OF COMPOSITION OPERATORS ON ORLICZ-LORE NTZ SPACES5\nThis implies that ν(Fn) = 0 which is a contradiction to our assumption. Similarly,\nν(Em+1/Em) = 0. To show that νmandνm+1are equivalent, it is suffices to show\nthatνm≪νm+1. Suppose that νm+1(A) = 0 for A∈ E. This implies that for each\nsubsetBofA, we have\n/integraldisplay\nBfΨm+1dν≤/integraldisplay\nAfΨm+1dν=νm+1(A) = 0\n⇒ν{x∈A:fΨm(x)/ne}ationslash= 0,fΨm+1(x)/ne}ationslash= 0}= 0.\nAlso,ν{x∈A:fΨm(x)/ne}ationslash= 0,fΨm+1(x) = 0}= 0 as it is a subset of Em+1/Em.\nTherefore,\nνm(A) =/integraldisplay\nCfΨmdν=/integraldisplay\nDfΨmdν= 0\nwhereC={x∈A:fΨm= 0}andD={x∈A:fΨm/ne}ationslash= 0}. /square\nTheorem 3.3. LetΨbe a non-singular measurable transformation on the measure\nspaceEinducing the composition operator CΨon the Orlicz-Lorentz space L(φ,w).\nA necessary and sufficient condition for A(CΨ) =mis thatmis the least non-\nnegative integer such that the measures νmandνm+1are equivalent.\nCorollary 3.4. LetCΨ∈ B(L(φ,w)). Then the A(CΨ) = 0in each of the following\nsituations:\n(1)Ψis a measure preserving.\n(2)Ψis surjective.\nTheorem 3.3 can be restated in the following manner:\nTheorem 3.5. The necessary and sufficient condition for a composition oper ator\nCΨon Orlicz-Lorentz space L(φ,w)to haveA(CΨ) =∞is that the measures νm\nandνm+1can never be equivalent for any natural number m.\nTheorem 3.6. A sufficient condition for an operator CΨ∈ B(L(φ,w)), induced by\na measurable transformation Ψ, to have A(CΨ)<∞is thatΨis a pre-positive\nmeasurable transformation.\nProof.Let Ψ be a pre-positive measurable transformation inducing CΨonL(φ,w).\nLet us suppose that A(CΨ)≮∞, then for each natural number m, there exists\nfm∈ N(Cm\nΨ) such that\nν/braceleftbig\nx∈E:fψm◦Ψm−1(x)/ne}ationslash= 0/bracerightbig\n>0.\nSince Ψ is pre-positive, so ν(Ψ−1(Em)>0 where\nν(Ψ−1(Em) =ν({x∈E: (fΨm◦Ψm)(x)/ne}ationslash= 0}),\nwhich is a contradiction to the fact that fΨm∈ N(Cm\nΨ). Therefore, it follows that\nA(CΨ)<∞. /square\nTheorem 3.7. If the measurable transformation Ψinducing CΨ∈ B(L(φ,w)), is\npre-positive, then A(CΨ) = 0.\nTheorem 3.8. LetCΨ∈ B(L(φ,w)). If there exists a sequence {A}mof mea-\nsurable sets such that for each m,0< ν(Am)<∞,ν(Ψ−m(Am)) = 0 and\nν(Ψ−(m−1)(Am))/ne}ationslash= 0, thenA(CΨ) =∞.6 NEHA BHATIA AND ANURADHA GUPTA\nProof.Let{Am}m≥1be a sequence of measurable sets satisfying 0 < ν(Am)<∞,\n(Ψ−m(Am)) = 0 and ν(Ψ−(m−1)(Am))/ne}ationslash= 0 for each m. For each measurable set S,\nwe know that\n/bardblχS/bardbl=1\nφ−1(1\nν(S)).\nHence, for each natural number m, the characteristics function XAm∈L(φ,w)\nandχAm∈(N(Cm\nΨ)/N(Cm−1\nΨ)). Therefore, A(CΨ) =∞. /square\nTheorem 3.9. Let the measurable transformation ΨonEinducing CΨonL(φ,w)\nbe such that the image of each measurable set is measurable. I fA(CΨ)≮∞on\nL(φ,w), then there exists a sequence of subsets AmofEsuch that for all m >1,\n(1)0< ν(Am)<∞,\n(2)Am⊆Ψm−1(B)for some B∈ E,\n(3)Am/∈ {Ψm(S) :S∈ Eand ν(S)>0}.\nProof.LetA(CΨ) =∞. Then, for each positive integer m, we have N(Cm−1\nΨ)⊂\nN(Cm\nΨ). This implies that there exists fΨm∈ N(Cm\nΨ) such that fΨm/∈ N(Cm−1\nΨ).\nTakeEm=/braceleftbig\nx∈E:fΨm◦Ψm−1(x)/ne}ationslash= 0/bracerightbig\n. Clearly, ν(Em)>0 and Ψm−1(Em)\nis measurable. Now, we show that ν(Ψm−1(Em))>0. Let us suppose that\nν(Ψm−1(Em)) = 0. Since Ψ is non-singular and Em⊆Ψ−(m−1)(Ψm−1(Em)),\nwe get\nν(Em)≤ν(Ψ−(m−1)(Ψm−1(Em))) = 0,\nwhich is a contradiction. Hence, ν(Ψm−1(Em))>0.\nNow, since the measure νisσ-finite, we can choose a measurable subset Am\nof Ψm−1(Em) such that 0 < ν(Am)<∞wherex∈Amsuch that fΨm(x)/ne}ationslash= 0.\nNext, let if possible that Am= Ψm(S) for some measurable set Shaving positive\nmeasure, then\nν({x∈S:fΨm◦Ψm(x) = 0}) = 0.\nSincefΨm∈ N(Cm\nΨ),ν({x∈S: (fΨm◦Ψm)(x)/ne}ationslash= 0}) = 0. The above two sets each\nhaving measure zero mean that ν(S) = 0, this contradicts the fact that ν(S)>0.\nHence the proof. /square\nIf we put E=N,E= 2Nandνis the counting measure, then the correspond-\ning Orlicz-Lorentz space is called as Orlicz-Lorentz sequence space and is denoted\nbyl(φ,ω)[3]. The following theorem gives the characterization of the composit ion\noperators to have infinite ascent on the Orlicz-Lorentz sequence spacesl(φ,ω).\nTheorem 3.10. A necessary and sufficient condition for the composition oper ator\nCΨon the Orlicz-Lorentz sequence space l(φ,ω)induced by Ψ :N→Nto have\nA(CΨ) =∞is that there exists a sequence of distinct natural numbers /an}bracketle{tnm/an}bracketri}htsuch\nthatnm/ne}ationslash∈Ψm(N)butnm∈Ψm−1(N)for each m≥1.\nProof.The sufficient part is a direct conclusion. For the necessary part, u sing\nTheorem3.9, wegetasequenceofnonemptymeasurablesubsets AmofNsatisfying\nAm⊆Ψm−1(N) andAm/notsubseteqlΨm(N)⊆Ψm−1(N). Hence we can take a sequence of\ndistinct natural number nmsatisfying nm∈Ψm−1(N) andnm/∈Ψm(N). /squareASCENT AND DESCENT OF COMPOSITION OPERATORS ON ORLICZ-LORE NTZ SPACES7\n4.Descent of composition operators\nDefinition 6. A measure space (E,E,ν)is said to be separable if for every distinct\npair of points xandyinE, we can find disjoint positive measurable sets E1and\nE2such that x∈E1andy∈E2.\nTheorem 4.1. The composition operator CΨon the Orlicz-Lorentz space L(φ,w)\nhasD(CΨ) = 0ifCΨis bounded away from zero on its support and Ψ−1(E) =E.\nTheorem 4.2. Suppose that (E,E,ν)is a separable σ-finite measure space. A\nnecessary condition for the composition operator CΨonL(φ,w)to haveD(CΨ)<∞\nis that ˆΨmis injective for some non negative integer m, where ˆΨm= Ψ|R(Ψm):\nR(Ψm)→ R(Ψm).\nProof.Suppose that D(CΨ)<∞onL(φ,w). Let if possible foreach m, the mapping\nˆΨmis not injective, then there exist x1,x2, inEsuch that\nΨm(x1)/ne}ationslash= Ψm(x2) and\nΨm+1(x1) = Ψm+1(x2).\nAlso, since ( E,E,ν) is separable and σ-finite, therefore there exists two disjoint\nmeasurable sets E1andE2with non-zero finite measures containing x(= Ψm(x1))\nandy(= Ψm(x2)), respectively.\nHenceχE1,χE2∈L(φ,ω). Now consider the element f1ofL(φ,ω)given by\nf1=XX1−XX2.\nThenf2=Cm\nΨf1∈ R(Cm\nΨ). Butf2/∈ R(Cm+1\nΨ) because if f2=Cm+1\nΨf3for\nsomef3∈L(φ,ω)(E),\n1 =f1(x)\n=f1(Ψmx1)\n=Cm+1\nΨfΨ(x1)\n=Cm+1\nΨfΨ(x2)\n=f2(x2)\n=Cm\nΨf1(x2)\n=f1(y)\n=−1.\nThis implies that R(Cm+1\nΨ)⊂ R(Cm\nΨ) for each non negative integer m. Hence\nD(CΨ) =∞. /square\nTheorem 4.3. Suppose that (E,E,ν)is a separable σ-finite measure space. Then a\nnecessary condition for the composition operator CΨonL(φ,w), to have the descent\nat mostmis thatΨmis injective.\nCorollary 4.4. Let(E,E,ν)be the measure space such that every singleton set has\npositive measure and CΨ∈ B(L(φ,ω)). ThenD(CΨ)> mif the mapping Ψ|R(Ψm):\nR(Ψm)→ R(Ψm)is not injective.\nProof.Let the measure space ( E,E,ν) beσ-finite and separable and Ψ |R(Ψm)is not\ninjective, then (as in Theorem 4.2), there exists measurable sets E1andE2con-\ntainingx(= Ψm(x1)) andy(= Ψm(x2)) as the singleton sets x1andx2respectively,\nwe obtain that D(CΨ)> m. /square8 NEHA BHATIA AND ANURADHA GUPTA\nReferences\n[1] S.C. Arora, G. Datt, S. Verma, Multiplication and compos ition operators on Orlicz-Lorentz\nspaces,Int. J. Math. Anal. (Ruse) 1no. 25 (2007), 1227–1234.\n[2] D.S.Bajaj, G. Datt, Ascent and Descent of Composition op erators on Lorentz Spaces, Com-\nmun. Korean Math. Soc. 37(2022), no. 1, 195-205.\n[3] P. Bala, A. Gupta, N. Bhatia, Multiplication Operators o n Orlicz-Lorentz Sequence Spaces,\nInt. Journal of Math. Analysis 7(2013), no. 30, 1461-1469.\n[4] H. Chandra, P. Kumar, Ascent and Descent of Composition O perators on lpspaces,Demon-\nstratio Math. 43(2010), no. 1, 161-165.\n[5] R. K. Giri, S. Pradhan, On the properties of ascent and des cent of composition operator on\nOrlicz spaces, Math. Sci. Appl. E-Notes 5(2017), no. 1, 70-76.\n[6] R. Kumar, Ascent and Descent of Weighted Composition Ope rators on Lp-spaces, Mathe-\nmaticki Vensik ,60(2008), no. 1, 47-51.\n[7] S.J. Montgomery-Smith, Orlicz-Lorentz spaces, Procee dings of the Orlicz Memorial Confer-\nence, Oxford, Mississippi, 1991.\n[8] R. K. Singh, J. S. Manhas, Composition operators on funct ion spaces, North-Holland Mathe-\nmatics Studies, 179, North-Holland Publishing Co., Amster dam, 1993.\nNeha Bhatia\nDepartment of Mathematics, University of Delhi, Delhi 11000 7, India\nEmail address :nehaphd@yahoo.com\nAnuradha Gupta\nDepartment of Mathematics, Delhi College of Arts and Commerc e, University of Delhi,\nDelhi 110023, India\nEmail address :dishna2@yahoo.in" }, { "title": "1008.1279v1.Lorentz_Violation_and_Extended_Supersymmetry.pdf", "content": "arXiv:1008.1279v1 [hep-ph] 6 Aug 2010November 12, 2018 15:41 WSPC - Proceedings Trim Size: 9in x 6i n patsusyproc\n1\nLORENTZ VIOLATION AND EXTENDED\nSUPERSYMMETRY\nDON COLLADAY and PATRICK MCDONALD∗\nDivision of Natural Science, New College of Florida\nSarasota, FL 34234, USA\n∗E-mail: mcdonald@ncf.edu\nWe construct a collection of Lorentz violating Yang-Mills t heories exhibiting\nsupersymmetry.\n1. Introduction and background\nSymmetryhasplayedafundamentalroleintheconstructionoffield theories\npurportingto describefundamental physics.Nowhereis this more true than\nin the construction of the Minimal Supersymmetric Standard Model where\nsymmetries mixing bosonic and fermionic states lead to tightly constr ained\ntheories often exhibiting remarkable properties. This is particularly true for\nN= 4 extended supersymmetric Yang-Mills theories,1which are known to\nbe finite.\nIn this note we construct field theories which exhibit N= 4 ex-\ntended supersymmetry and Lorentz violation. To do so we combine id eas of\nBergerandKosteleck´ y2involvingsupersymmetricscalartheoriesexhibiting\nLorentz violation, and well-known constructions of extended supe rsymmet-\nric theories involving dimensional reduction (nicely described in the wo rk\nof Brink, Schwartz and Scherk1). We begin by establishing notation in the\ncontext of the standard construction.\nConsider a gauge theory involving a single fermion λand lagrangian\nL=−1\n4F2+i\n2¯λ/ne}ationslash∂λ, (1)\nwhereFis the field strength, Fµν= [Dµ,Dν]/igandDis the covariant\nderivative,\nDµ=∂µ+igAµ. (2)November 12, 2018 15:41 WSPC - Proceedings Trim Size: 9in x 6i n patsusyproc\n2\nTo simplify notation, we will first consider the abelian case. To implemen t\nsupersymmetry we introduce a supercharge, Q,satisfying\n[Pµ,Q] = 0, {Q,¯Q}= 2γµPµ, (3)\nwhereγµare the standard Dirac matrices\n{γµ,γν}= 2gµν, (4)\nand the energy momentum 4-vector Pµgenerates spacetime translations.\nThe construction of a supercharge is elegantly carried out in the co ntext\nofsuperspace. More precisely, introduce four independent anticommuting\nvariables, θ,and consider the general vector superfield\nV(x,θ) =C(x)+i¯θγ5w(x)−i\n2¯θγ5θM(x)−1\n2¯θθN(x)+1\n2¯θγ5γµθAµ\n−i¯θγ5θ¯θ[λ+i\n2/ne}ationslash∂w(x)]+1\n2(¯θθ)2(D(x)−1\n2∂µ∂µC(x)) (5)\nand theQoperator\nQ=−i∂¯θ−γµθ∂µ. (6)\nFixing an arbitrary spinor α,standard analysis3of the operator δQV=\n−i¯αQVleads to the supersymmetry transformations defining an N= 1\nsupersymmetric theory.\n2.N= 1 supersymmetry\nFollowing Berger and Kosteleck´ y,2we introduce Lorentz violation by defin-\ning a twisted derivative:\n˜∂µ=∂µ+kµν∂ν, (7)\nwherekµνis a symmetric, traceless, dimensionless tensor parametrizing\nLorentz violation. To obtain a gauge invariant theory, we also twist t he\nunderlying connection:\n˜Aµ=Aµ+kµνAν. (8)\nThese perturbations lead to a general vector superfield\n˜V(x,θ) =C(x)+i¯θγ5w(x)−i\n2¯θγ5θM(x)−1\n2¯θθN(x)+1\n2¯θγ5γµθ˜Aµ\n−i¯θγ5θ¯θ[λ+i\n2/ne}ationslash˜∂w(x)]+1\n2(¯θθ)2(D(x)−1\n2˜∂µ˜∂µC(x)) (9)\nand perturbed Qoperator\n˜Q=−i∂¯θ−γµθ˜∂µ. (10)November 12, 2018 15:41 WSPC - Proceedings Trim Size: 9in x 6i n patsusyproc\n3\nUsing Wess-Zumino gauge we obtain a lagrangian\n˜L=1\n4˜F2+i\n2¯λ/ne}ationslash˜∂λ+1\n2D2, (11)\nwhereDis an auxiliary chiral field and ˜Fis the twisted field strength,\n˜Fµν=˜∂µ˜Aν−˜∂ν˜Aµ.The twisted field strength can be written in terms of\nthe standard SME parameters: ˜F2=F2+kµναβ\nFFµνFαβ, where\nkµναβ\nF= 2(2kαµ+(k2)αµ)gβν+4(kµα+(k2)αµ)kνβ+(k2)αµ(k2)βν.(12)\nDirect calculation confirms that the action is invariant under the sup ersym-\nmetry transformations\nδ˜Aµ=−i¯αγµλ,\nδλ=i\n2σµν˜Fµνα−γ5Dα,\nδD= ¯α/ne}ationslash∂γ5λ. (13)\nThis defines an N= 1 supersymmetric theory with Lorentz violation.\n3.N= 4 supersymmetry\nTo build an N= 4 supersymmetric theory we work in 4 + 6-dimensional\nspacetime. We represent the 32 ×32 gamma matrices via Γµ=γµ⊗I8\nwhereI8is the 8×8 identity matrix and µ= 0,1,2,3,and\nΓ4= Γ14+Γ23,Γ6= Γ34+Γ12, iΓ8= Γ24+Γ13,\nΓ5= Γ24−Γ13, iΓ7= Γ14−Γ23, iΓ9= Γ34−Γ12,(14)\nwhere\nΓij=γ5⊗/parenleftbigg0ρij\nρij0/parenrightbigg\n(15)\nand the 4 ×4 matrices ρare defined by\n(ρij)kl=δikδjl−δjkδil,\n(ρij)kl=1\n2ǫijmn(ρmn)kl=ǫijkl. (16)\nWe consider the lagrangian\nL=−1\n4˜F2+i\n2¯λ˜/ne}ationslash∂λ, (17)\nwhere˜/ne}ationslash∂= Γµ(∂µ+kµν∂ν) is the twisted derivative with kµνparametrizing\nSO(1,9) violation and ˜Fis the corresponding perturbed field strength.November 12, 2018 15:41 WSPC - Proceedings Trim Size: 9in x 6i n patsusyproc\n4\nImposingboththeWeylandtheMajoranaconditionandcompactify ing,\nthe fermion λsatisfies\nλ=/parenleftbiggLχ\nR˜χ/parenrightbigg\n, (18)\nwhereLandRdenote left and right projection operators, respectively,\nthe spinor χtransforms as a 4 of SU(4) and the (independent) spinor ˜ χ\ntransforms as a ¯4 ofSU(4).\nChoosing the Lorentz violating parameters with care leads to Loren tz\nviolating extended supersymmetric theories which are easy to desc ribe. For\nexample, taking the kµνto vanish in the compactified directions leads to a\nlagrangian of the form\nL=−1\n4˜F2+i¯χ˜/ne}ationslash∂Lχ+1\n4˜∂µφij˜∂µφij, (19)\nwhere the complex scalar fields φijtransform as a 6 of SU(4) and are given\nby\nφi4=1√\n2(Ai+3+iAi+6),\nφjk=1\n2ǫjklmφlm= (φjk)∗. (20)\nThe associated action is invariant under the supersymmetry trans forma-\ntions\nδ˜Aµ=−i(¯αiγµLχi−¯χiγµLαi),\nδφij=−i√\n2(¯αjR˜χi−¯αiR˜χj+ǫijkl¯˜αkLχl),\nδLχi=i\n2σµν˜FµνLαi−√\n2γµ˜∂µφijR˜αj,\nδR˜χi=i\n2σµν˜FµνR˜αi+√\n2γµ˜∂µφijLαj. (21)\nSimilarly, choosing the kµνparameters to vanish in the spacetime direc-\ntionsµ= 0,1,2,3,we obtain a Lagrangian of the form\nL=−1\n4F2+i¯χ/ne}ationslash∂Lχ+1\n4∂µ˜φij∂µ˜φij, (22)\nwhere˜φij=φij+ Λijklφklwith the matrix Λ ijklcontaining the effect of\nLorentz violation in the compactified directions. The associated act ion is\ninvariant under the supersymmetry transformationsNovember 12, 2018 15:41 WSPC - Proceedings Trim Size: 9in x 6i n patsusyproc\n5\nδAµ=−i(¯αiγµLχi−¯χiγµLαi),\nδ˜φij=−i√\n2(¯αjR˜χi−¯αiR˜χj+ǫijkl¯˜αkLχl),\nδLχi=i\n2σµνFµνLαi−√\n2γµ∂µ˜φijR˜αj,\nδR˜χi=i\n2σµνFµνR˜αi+√\n2γµ∂µ˜φijLαj. (23)\nNote that if the scalars ˜φijare identified with physical scalars φijwe\nrestoreSU(4) symmetry and remove any Lorentz violating effects. If, how-\never, the φijcouple to other sectors, Lorentz effects may reappear in these\nsectors.\n4. Extensions and clarifications\nThe above results warrant a number of additional comments:\n•Theseconstructionscanbecarriedoutinthenonabeliancasewher e\nthey yield supersymmetric theories which exhibit Lorentz viola-\ntion.4\n•The same techniques can be applied to obtain an N= 2 supersym-\nmetric theory with Lorentz violation. The construction proceeds\nby working in 4+2-dimensional spacetime and using dimensional\nreduction.4\n•Because these constructions involve changing the structure of t he\nunderlying superalgebra,4,5the no-go results of Nibbelink and\nPospelov6do not apply.\nReferences\n1. L. Brink, J. Schwartz and J. Scherk, Nucl. Phys. B 121, 77 (1977).\n2. M. Berger and V.A. Kosteleck´ y, Phys. Rev. D. 65, 091701 (2002).\n3. M. Srednicki, Quantum Field Theory. Cambridge University Press, Cam-\nbridge, 2007.\n4. D. Colladay and P. McDonald, in preparation.\n5. D. Colladay and P. McDonald, J. Math. Phys. 43, 3554 (2002).\n6. S. Nibbelink and M. Pospelov, Phys. Rev. Lett. bf 94, 08160 1 (2005)." }, { "title": "1110.5258v1.Magnetohydrodynamics_dynamical_relaxation_of_coronal_magnetic_fields__I__Parallel_untwisted_magnetic_fields_in_2D.pdf", "content": "arXiv:1110.5258v1 [astro-ph.SR] 24 Oct 2011Astronomy& Astrophysics manuscriptno.13902 c/ci∇cleco√y∇tESO 2018\nNovember20,2018\nMHDDynamical RelaxationofCoronal Magnetic Fields\nI.ParallelUntwisted MagneticFields in 2D\nJorge Fuentes-Fern´ andez, ClareE.Parnell and Alan W.Hood\nSchool of Mathematics andStatistics,Universityof StAndr ews, NorthHaugh, StAndrews, Fife,KY16 9SS,Scotland\nPreprintonline version: November 20, 2018\nABSTRACT\nContext. For the last thirtyyears, most of the studies on the relaxati on of stressed magnetic fields in the solar environment have o nly\nconsidered the Lorentz force, neglecting plasma contribut ions, andtherefore, limitingeveryequilibrium tothat of a force-free field.\nAims.Here we begin a study of the non-resistive evolution of finite beta plasmas and their relaxation to magnetohydrostatic st ates,\nwhere magnetic forces are balanced by plasma-pressure grad ients, by using a simple 2D scenario involving a hydromagnet ic dis-\nturbance to a uniform magnetic field. The final equilibrium st ate is predicted as a function of the initial disturbances, w ith aims to\ndemonstrate what happens tothe plasma during the relaxatio n process and tosee what e ffects ithas on the finalequilibrium state.\nMethods. A set of numerical experiments are run using a full MHD code, w ith the relaxation driven by magnetoacoustic waves\ndamped by viscous e ffects. The numerical results are compared with analytical ca lculations made within the linear regime, in which\nthe whole process must remainadiabatic. Particular attent ionis paidtothe thermodynamic behaviour of the plasma duri ng the relax-\nation.\nResults.The analyticalpredictions forthe finalnonforce-freeequi librium depend onlyontheinitialperturbations andthetot alpres-\nsure of the system. It is found that these predictions hold su rprisingly well even for amplitudes of the perturbation far outside the\nlinear regime.\nConclusions. Includingtheeffectsofafiniteplasmabetainrelaxationexperimentsleadst osignificantdifferencesfromtheforce-free\ncase.\nKey words. Magnetohydrodynamics (MHD) –Sun:corona – Magnetic fields\n1. Introduction\nThe magnetic field of the solar corona is believed to evolve\nthrough a series of force-free states (Heyvaerts&Priest 19 84).\nSince the solar corona involves a low-beta plasma in which\nmagnetic forces dominate over plasma forces, this is not\nan unreasonable assumption, and so, most of the recent\nstudies on the relaxation of coronal magnetic fields (e.g.\nMackay&vanBallegooijen 2006; Ji et al. 2007; Inoueet al.\n2008; Janse &Low 2009; Milleret al. 2009; Pontinet al. 2009)\nhave been done by considering the approximation of an ex-\ntremelytenuousplasma,forwhichtheplasmapressuredoesn ot\nplay an important role, and the persistent hydromagnetic st ruc-\nturesofthe solarcoronaareassumedto bein magneticbalanc e,\nwith zeropressuregradients.On the otherhand,onlywithin the\npast few years, Ruanet al. (2008) and Gary (2009) have starte d\nto consider the reconstruction of the global coronal magnet ic\nfield including a finite Lorentz force balanced by magnetic an d\ngravityforces.\nIn addition, there are many codes available to calculate\nthose force-free fields from the observed magnetic field in th e\nphotosphere (Amariet al. 1998; Wiegelmannetal. 2006, 2008 ;\nSchrijveret al. 2006; Metcalfetal. 2008). These codes have\nbeenusedwithvaryingdegreesofsuccesstodeterminethema g-\nnetic field of solar flares and active regions (e.g. DeRosa et a l.\n2009;Schrijveretal.2008;R´ egnier2008;Wheatland& R´ eg nier\n2009).However,problemsremainwiththeseapproaches.Inp ar-\nticular, a non-linear force-free field determined from a lin e-of-\nsight photosphericmagnetic field is not unique,but is one of aninfinite number of possible solutions. This fact is well know n\nandhasbeendiscussedbyseveralauthors(seeLow2006).\nAlso, the low beta plasma assumption is only valid for the\nsolar corona, but is not valid in the photosphere, chromosph ere\nor much of the transition region where plasma e ffects become\nmore important. Furthermore, in the solar corona, the tenuo us\nplasma will be able to create a high beta in the surroundingso f\na magnetic null point (McLaughlin& Hood 2006). Moreover,\neven where the plasma has a low beta, there are still some ef-\nfectsonthe finalequilibriumofthe magneticfield that will l ead\ntoenergeticconsequencesasthefieldisrelaxingtoitsnewe qui-\nlibriumstate.\nInparticular,relaxationcaninvolvethemagneticfieldevo lv-\ning from a stressed state to a force-free field with a lower\nenergy, and for a consistent scenario where the energy can-\nnot escape the system, conservation of energy implies that t he\nlosses of magnetic energy must be converted into something\nelse. Browninget al. (2008) and Hoodet al. (2009) investiga ted\nTaylor relaxation (Taylor 1974) through a series of non-lin ear\n3Dsimulations,initiatedbyan idealMHD instability.Alth ough\nthe initial state was force-free, the final state involved a h igh\ntemperatureplasmawith a significantvalueforthe plasmabe ta.\nPlacing aside some extra contributions such as radiative lo sses,\nifasubstantialfractionofthemagneticenergyreleasedgo esinto\ntheinternalenergy,thentheplasmabetacannotbesmall.He nce,\nconsidering the behaviour of the plasma is important even if it\nhas little effect on the final magnetic equilibrium. On the other\nhand, Gary (2001) suggested the possibility that there is hi gh\nbetaplasmainthe solarcoronaaboveactiveregions.2 Jorge Fuentes-Fern´ andez, ClareE.Parnell andAlanW.Hoo d: MHD Dynamical Relaxation of Coronal Magnetic Fields\nIn thispaper,we will start byconsideringa simple scenario ,\nwithauniform2Dmagneticfield.Therewillbenocurrentshee t\nformation and no changes in the magnetic topology. The aim\nof the paper is to investigate the relaxation of the hydromag -\nnetic fluid for various di fferent values of the plasma beta. To\nperturb the system, we introducea local small enhancement ( or\ndeficit) in the plasma pressure (or in the magnetic field), un-\nderthefrozen-incondition.Thesubsequentrelaxationlea dstoa\nnew magnetohydrodynamic equilibrium in which the perturba -\ntionhasbeendissipatedbyviscousforcesactingontheflow.\nIt is worthwhile mentioning that relaxation via Ohmic dis-\nsipation, due to the e ffect of resistivity, or magnetic di ffusiv-\nity, represents a substantially di fferent problem; While viscos-\nity dissipates the plasma velocity, di ffusivity tends to eliminate\nthe electric current density, and such a relaxed state can on ly\ninvolve potential fields, which are mathematically well defi ned\nandareuniquelydeterminedbythecomponentsofthemagneti c\nfield normal to the boundaries.Furthermore,the time-scale s for\nan Ohmic relaxation in very high magnetic Reynolds number\nenvironments, such as the solar corona, are in general proba bly\nlarger than the age of the Sun itself, outwith regions with ve ry\nsmall lengthscales(seePriest 1982).\nBy considering the e ffects of the plasma pressure in the re-\nlaxation, we are facing a totally di fferent problem from that of\nthe force-free relaxation studied by many others. In our non -\nresistive MHD relaxation, the plasma displacements driven by\nthe initial pressure enhancement will carry the magnetic fie ld\nwith them, generating an electric current and a magnetic pre s-\nsure.Hence,theresultingequilibriumwillhavetoinvolve abal-\nance between the Lorentz force and the plasma-pressure grad i-\nent.\nThe effectsof includinga finite plasma beta are relevantnot\nonlyinthehighplasmabetaregionsofthesolaratmospheres uch\nas the photosphere and chromosphere, but will also be releva nt\nin the solar corona. Obvious regions where the plasma beta is\nlikely to have a significant e ffect are in the vicinities of mag-\nnetic neutral points, where the magnetic field vanishes. The se\nconfigurations will be the subject of a further paper, while i n\nthe present paper we study the plasma beta e ffects on the sim-\nplestmagneticconfiguration,auniformmagneticfield,whic his\nabsolutely general and might be compared with di fferent solar\nenvironments such as a region in a coronal prominence or part\nofacoronalloop.\nIn Section 2, we first present the setup of the two-\ndimensional linear problem. In Sections 2.1 and 2.2, we solv e\nthe equationsfor a one-dimensionalvertical and horizonta lper-\nturbation, respectively, and in 2.3 we obtain the whole solu tion\nfrom a general two-dimensional linear perturbation, produ cing\na practical and precise analytical solution to the problem. In\nSection 3, we present a set of numerical experiments run us-\ning the Lare2D code (Arberet al. 2001), and we compare these\nwiththepreviousanalyticalresults.Thee ffectsofthenon-linear\nterms are considered when the magnitude of the perturbation is\nincreased,inSection4,followedbythesummaryandsomecon -\nclusionswhichare presentedinSection5.\n2. Linear2DEquations\nThe initial set up involves a uniform magnetic field pointing in\nthe vertical y-direction, B0=B0ˆey, and a background plasma\nof a constant gas pressure, density and temperature. The ini -\ntial disturbances are supposed to be small, in order to stay i n\nthe linear regime. Expressing each quantity q(x,y,t) as the sum\nof a background constant value plus a perturbation, q(x,y,t)=q0+q1(x,y,t), and substituting those expressions into the vis-\ncous, non-resistive, two-dimensional MHD equations, with no\ngravityforceforsimplicity,andneglectingtermsinvolvi ngprod-\nucts of perturbations, we obtain the following set of linear ized\nequations,inscalarform:\n∂ρ1\n∂t=−ρ0∇·v1, (1)\nρ0∂vx\n∂t=−∂p1\n∂x−B0\nµ0∂By\n∂x+B0\nµ0∂Bx\n∂y+ρ0ν/parenleftigg\n∇2vx+1\n3∂\n∂x(∇·v1)/parenrightigg\n,(2)\nρ0∂vy\n∂t=−∂p1\n∂y+ρ0ν/parenleftigg\n∇2vy+1\n3∂\n∂y(∇·v1)/parenrightigg\n, (3)\n∂p1\n∂t=−γp0∇·v1, (4)\n∂Bx\n∂t=B0∂vx\n∂y, (5)\n∂By\n∂t=−B0∂vx\n∂x, (6)\nwhereρ0,p0andB0are the constant background plasma den-\nsity, plasma pressure and magnetic field, νis the kinematic vis-\ncosity,ρ1,p1andv1are the perturbations on plasma density,\nplasmapressureandvelocity,and vx,vy,BxandByarethexand\nycomponentsof the perturbedvelocity and perturbed magneti c\nfield, respectively. Also, the plasma pressure, p, density,ρ, and\ninternalenergy,ǫ, arerelatedbytheperfectgaslaw,foranideal\npolytropicgas,\np=ρǫ(γ−1), (7)\nwhereγistheratioofspecificheats,oftenassumedtobe5 /3for\na highlyionisedhydrogenplasma.\nIn a general scheme, viscosity would add a heating term to\nthe energy equation, but this term is second order, and so, th e\nprocessisadiabaticwithinthelinearregime,andthereisn oheat-\ning of any kind taking place. Hence, the entropy per unit mass ,\nS=p/ργ, is conserved, for each single fluid element, and for\ntheentirebox.\nFrom the conservation of entropy, a relation between the\nplasma pressure and density perturbations may be obtained,\nwithin first order (i.e. neglecting the terms involving prod ucts\nofperturbations):\np0+p1\n(ρ0+ρ1)γ=p0\nργ\n0/parenleftigg\n1+p1\np0−γρ1\nρ0/parenrightigg\n=constant.\nHence,\n∆p1\np0−γ∆ρ1\nρ0=0,\nwhere∆indicatesthedifferencebetweenfinalandinitialstateof\ntheperturbation,suchthat\n∆p1=c2\ns∆ρ1, (8)\nwherecs=/radicalbig\nγp0/ρ0is thesoundspeed.\nTo solve these equations, we first determine the solution for\na one-dimensionalperturbation, which depends only on x(Sec.\n2.1), then only on y(Sec. 2.2), and finally, using the 1D results,\nwe derive the solution for a more general 2D perturbation(Se c.\n2.3).Jorge Fuentes-Fern´ andez, ClareE.Parnell andAlanW.Hood : MHD Dynamical Relaxation of Coronal Magnetic Fields 3\n2.1. Perturbationdependentonx\nConsidera perturbationvaryingonlyinthe directionperpe ndic-\nulartothemagneticfieldlines, x.Themagneticfieldvectorwill\nhave a non-zero y-component, B1(x,t)=By(x,t)ˆey, while the\nvelocitywill have a non-zero x-component, v1(x,t)=vx(x,t)ˆex.\nEquations(1)to (6)reduceto\n∂ρ1\n∂t=−ρ0∂vx\n∂x, (9)\nρ0∂vx\n∂t=−∂pT\n∂x+ρ0ν′∂2vx\n∂x2, (10)\n∂p1\n∂t=−γp0∂vx\n∂x, (11)\n∂By\n∂t=−B0∂vx\n∂x, (12)\nwithν′=4ν/3,whereνisthe kinematicviscosity,and pTisthe\nperturbedtotalpressure,givenby\npT=p1+B0By\nµ0. (13)\nThe equation governing the final equilibrium state can be ob-\ntained using Eq. (10). At equilibrium,the time dependence d is-\nappears,andthevelocityiszero,thus,theequilibriummus thave\nconstanttotal pressure:\n∂pT\n∂x=0. (14)\nCombiningEquations(11) and (12), we get the evolutionof th e\ntotalpressureas\n1\nρ0∂pT\n∂t=−(c2\ns+c2\nA)∂vx\n∂x, (15)\nwherecsisthesoundspeed,definedabove,and cA=B0/√µ0ρ0\nistheAfv´ enspeed.BytheappropiatecombinationofEquati ons\n(10)and(15)wegetthewaveequationforfastmagneto-acous tic\nwavesforthetotal pressure:\n∂2pT\n∂t2=(c2\ns+c2\nA)∂2pT\n∂x2+ν′∂\n∂t/parenleftigg∂2pT\n∂x2/parenrightigg\n. (16)\nAssuming that the total pressure can be considered as a conti n-\nuous, periodic function, the solution of the last equation c an be\nexpressedasasuperpositionofplanewaves,suchas\npT(x,t)=Re/summationdisplay\nkϕkei(kx−ωt), (17)\nwhereeachwavenumber kcorrespondstoadi fferentoscillation\nmodeandisassociatedwitha complexfrequency ω(k)givenby\nthedispersionrelation,\nω2+ik2ν′ω−(c2\ns+c2\nA)k2=0.\nThe dispersion relation has the solution ω=a−bi, whereais\ntherealfrequencyofthewave,and bisthedampingterm:\na=k\n2/radicalig\n4(c2s+c2\nA)−k2ν′2,\nb=1\n2k2ν′.\nIn orderto havea harmonicmode,the wave number kmust sat-\nisfyk2ν′2<4(c2\ns+c2\nA), and higher modeswill be dampedwith-\nout anytypeof oscillation.On the otherhand,notice that ω=0whenk=0.Theundampedmode k=0 correspondstothe con-\nstantFouriercoefficientintheexpansionofEq.(17),whichdoes\nnotchangeintime,andisgivenbyFourieranalysisasthehom o-\ngeneous redistribution of the initial total pressure to its average\nvalue. As t→∞, it is only this constant that remains, as all the\nother terms are proportional to e−bt. Hence, this homogeneous\nredistributionisexactlytheconstantperturbedtotalpre ssurethat\ndefinesourfinal equilibriumstate:\npT(∞)=1\nLx/integraldisplay\nx/parenleftigg\np1(x,0)+B0By(x,0)\nµ0/parenrightigg\ndx, (18)\nwhereLxisthelengthofthe x-domain.\nFrom the solution of Eq. (16), and Eq. (15), we obtain an\nexpression for v1(x,t), which, after substitution into Equations\n(9),(11)and(12),gives\n∂ρ1\n∂t=1\nc2s+c2\nA∂pT\n∂t, (19)\n∂p1\n∂t=c2\ns\nc2s+c2\nA∂pT\n∂t, (20)\n∂By\n∂t=B0\nρ0(c2s+c2\nA)∂pT\n∂t. (21)\nIntegrating now from t=0 tot=∞, we obtain the perturbed\nquantitiesforthe final equilibriumstate, as functionsof t heper-\nturbedtotalpressure,to beaddedtothebackgroundvalues:\nρ1(x)=ρ1(x,0)+1\nc2s+c2\nA(pT(∞)−pT(x,0)), (22)\np1(x)=p1(x,0)+c2\ns\nc2s+c2\nA(pT(∞)−pT(x,0)), (23)\nBy(x)=By(x,0)+B0\nρ0(c2s+c2\nA)(pT(∞)−pT(x,0)).(24)\nEquations (22), (23) and (24) state that no matter how we\nset ourinitial disturbance,the final equilibriumdistribu tionsare\ncompletelydeterminedbytheinitialandthefinaltotalpres sures\nofthesystem,whicharegivenbythesolutionofthewaveequa -\ntion. Note, that also the adiabatic equation for the linear r egime\ngiveninEq.(8) issatisfied.\n2.2. Perturbationdependentony\nIn the same way as above, we can get the solution for a pertur-\nbation varying only along the field. This time, we are dealing\nwith a purely non-magneticevolution,that will lead to a hom o-\ngeneousredistribution of the plasma pressure all along the field\nlines. Theequationsto solveare the y-componentsofEquations\n(1),(3) and(4).Theequilibriumisnowgivenby\n∂p1\n∂y=0, (25)\nandthesolutionfortheperturbedquantitiesis\np1(∞)=1\nLy/integraldisplay\nyp1(y,0)dy, (26)\nρ1(y)=ρ1(y,0)+1\nc2s(p1(∞)−p1(y,0)), (27)\nwhereLyis thelengthofthe y-domain.4 Jorge Fuentes-Fern´ andez, ClareE.Parnell andAlanW.Hoo d: MHD Dynamical Relaxation of Coronal Magnetic Fields\n2.3. Two-dimensionalperturbation\nIn this section, we study the relaxation of a general individ -\nualtwo-dimensionalperturbationwithinthelinearregime .Once\nagain,settingthevelocitiestozero,wegettheequationsg overn-\ningthefinal2D equilibrium:\n∂\n∂x/parenleftigg\np1+B0By\nµ0/parenrightigg\n−B0\nµ0∂Bx\n∂y=0, (28)\n∂p1\n∂y=0. (29)\nEquation (29) tells us that the final plasma pressure cannot\ndepend on y, so the solution for the pressure must remain one-\ndimensional,asbefore.Ontheotherhand,Eq.(28)doesnoth ave\na direct interpretation, as both spatial derivatives are in volved.\nThe term p1+B0By/µ0represents the perturbed total pressure\nfrom Sec. 2.1, and the new term B0Bx/µ0represents the mag-\nnetic tension due to the curvature of the field lines, which wa s\nzerointhe1Dcases.Asusual,whenlookingforaperiodicsol u-\ntion,Fourieranalysingmakeslifemuchsimpler:Expressin gour\nvariablesas functionsof ei(kx+ly−ωt), where each pair ( k,l) repre-\nsentsonesinglemodeofoscillationintheglobaltimeevolu tion,\nEquations(28) and(29) canberewritenas\nk/parenleftigg\np1+B0By\nµ0/parenrightigg\n−lB0Bx\nµ0=0,\nlp1=0,\nwhere:\ni. The mode k=0,l=0 represents the unperturbed back-\ngroundvalues.\nii. Fork/nequal0,l=0,theequationoftheequilibriumis\n∂\n∂x(p1+B0By)=0. (30)\nThese modes only depend on x, and represent the homoge-\nneousredistributionofthetotalpressurestudiedinSec.2 .1.\niii. Fork=0,l/nequal0we get\n∂p1\n∂y=0, (31)\n∂Bx\n∂y=0. (32)\nThesemodesdonotmodify By,insteadtheysimplyremove\nboth the vertical gradients of magnetic tension and plasma\npressureasinSec.2.2.Eachofthemistreatedindividually ,\nastheyarenotcoupledintheequations.\niv. Finally,forthosemodeswith k/nequal0,l/nequal0,we get\nkBy−lBx=0,\nwhich can be combined with the solenoidal condition for\nthemagneticfield, ∇·B=0,or,withinourfouriernotation,\nkBy+lBx=0.\nFrom these equations, we can conclude that, in the final\nequilibrium, the existance of a variation of Byin thex-\ndirection is totally incompatible with a variation of Bxin\nthey-direction. Hence, the modes with both wave numbers\nkandlnon-zeromayappearinthedynamicalevolution,but\nnotinthefinal equilibriumdistributions.Therefore, with our uniform background magnetic field\npointingstraightinthevertical y-direction,thefinal equilibrium\nstate is a combination of the backgroundvalues ( k=0,l=0),\nplus the vertical non-magnetic evolution to a state with pla sma\npressure that is constant along y, and/or the smoothing of the\nhorizontal component of the magnetic field ( k=0), plus the\none-dimensional hydromagnetic evolution across the field l ines\n(l=0). Note, that the perturbed Bxin the vertical direction (i.e.\ncurved magnetic field) is not coupled with either p1orBy, so\nthe final magnetic field remains as straight lines, and Bxis not\ninvolvedinthe evolutionofthetotal pressureofthesystem .\nHence, we calculate the analytical 2D equilibrium in two\nsteps: Firstly, the non-magnetic evolution in the vertical direc-\ntion,\np∗\n1(x)=1\nLy/integraldisplay\nyp1(x,y,0)dy, (33)\nρ∗\n1(x,y)=ρ1(x,y,0)+1\nc2s(p∗\n1(x)−p1(x,y,0)), (34)\nand secondly, the hydromagneticevolution in the horizonta l di-\nrection,acrossthefield,\nρ1(x,y)=ρ∗\n1(x,y)+1\nc2s+c2\nA(pT(∞)−p∗\nT(x)), (35)\np1(x,y)=p∗\n1(x)+c2\ns\nc2s+c2\nA(pT(∞)−p∗\nT(x)), (36)\nBy(x,y)=By(x,0)+B0\nρ0(c2s+c2\nA)(pT(∞)−p∗\nT(x)).(37)\nInEquations(33)to (37)the quantitieswitha superscriptr epre-\nsent the state after the vertical evolution, with p∗\nT(x)=p∗\n1(x)+\nB0By(x,0)/µ0.\nLooking back at the equations, we see a well known result\nfrom magnetohydrostatics, namely: In equilibrium, and in t he\nabsence of gravity, the plasma pressure must be constant alo ng\nthe field lines. The constant plasma pressure in the y-direction\ngiven by Eq. (29) is aligned with the straight magnetic config u-\nration,with nomagnetictension,forthe finalequilibriums tate.\nWhen analyzing the validity of the results above for a non-\nidealexperiment,itisimportanttorememberthatEquation s(34)\nto (37) come from the linear approximation, but Eq. (33) does\nnot. Hence, we expect our analytical calculations for the pr es-\nsure to hold for much larger initial perturbations than the o nes\nfor the density. If the initial pressure disturbance is not s mall,\nbut the linear expression for the plasma pressure is still va lid,\nthe adiabatic condition (i.e. conservation of entropy), wh ich is\nmore robust than the linear calculations, gives us a better a p-\nproximationfor the final equilibriumplasma density, calcu lated\nas\nρ(x,y,t→∞)=/parenleftiggp(x,y,t→∞)ργ(x,y,t=0)\np(x,y,t=0)/parenrightigg1/γ\n.(38)\n3. NumericalExperiments\nTo investigate the validity of the analytical results, we ha ve\nusedLare2D,astaggeredLagrangian-remapcodewithuserco n-\ntrolled viscosity, to solve the full MHD equations, with the re-\nsistivity set to zero (for further details, see Arberetal. 2 001).\nThe numerical domain is a square box with a uniform grid of\n256×256 points. The background magnetic field is pointing in\nthe vertical y-directionand all the perturbationsdependon bothJorge Fuentes-Fern´ andez, ClareE.Parnell andAlanW.Hood : MHD Dynamical Relaxation of Coronal Magnetic Fields 5\nxandy. The top andbottomboundariesof the boxare periodic,\nsothatthefieldlinesarenotline-tied.Theboundariesonth eleft\nandrightsidesareclosed.Thechoiceofeitherclosedorper iodic\nside boundariesmakesno di fferenceforour experiments.There\nisneithermassnorenergyflowingacrossthe sideboundaries .\nThe numerical code uses the normalised MHD equations,\nwherethenormalisedmagneticfield,densityandlengths,\nx=Lˆx,y=Lˆy,B=BnˆB, ρ=ρnˆρ,\nimply that the normalising constants for pressure, interna l en-\nergyandcurrentdensityare\npn=B2\nn\nµ0, ǫn=B2\nn\nµ0ρnandjn=Bn\nµ0L.\nWehaveusedherethesubscript nforthenormalisingconstants,\ninstead of 0, to avoid confusion with the initial background\nvalues. The hatquantities are the dimensionless variables with\nwhich the code works. The expression for the plasma beta can\nbeobtainedfromthisnormalisationas\nβ=2ˆp\nˆB2=2ˆρˆǫ(γ−1)\nˆB2. (39)\nFig.1.Time evolution of the energies for the case where\nmax(p1)=p0=0.1B2\nn/µ0andβ=0.2,integratedoverthewhole\n2Dbox.Thefourevolutionshavebeenshiftedinthevertical axis\nbysubstractingtheconstantvalues0,0.5,0.15and0.65,re spec-\ntively, for kinetic, magnetic, internal and total energy, b ut their\nrelative amplitudes are not scaled. The final losses of inter nal\nenergyare entirelybalancedwith a netincrease ofmagnetic en-\nergy.The top figure is logarithmicin time and coversthe whol e\nrelaxation. The bottom figure is linear in time and only cover s\nthe first part of the relaxation. From this graph, we can appre -\nciate the complex oscillation periods, which are products o f the\nsumofthedifferentplanewavesthat drivetherelaxation.The initial disturbance consists of an internal energy and\npressure enhancement, leaving the initial plasma density a nd\nmagneticfieldundisturbed.Theperturbationistakenasasi ngle\ntwo-dimensionalGaussiancenteredinthe middleofthebox:\nǫ1(x,0)=aexp/parenleftigg\n−(x−b)2\n2c2/parenrightigg\nexp/parenleftigg\n−(y−b)2\n2c2/parenrightigg\n, (40)\nwithb=0.5Landc=0.05L. As there is no initial perturbation\nof the magnetic field, the initial perturbed total pressure i s just\ntheperturbedplasmapressure,andthefinalanalyticalpert urbed\ntotalpressureistheaveragevalueofthatinitialperturbe dplasma\npressuredistribution,so that\npT(∞)−p∗\nT(x)=/integraldisplay\nx/parenleftigg/integraldisplay\nyp1(x,y,0)dy/parenrightigg\ndx−/integraldisplay\nyp1(x,y,0)dy.\nFor the experiment shown here, we have chosen a rather\nlarge perturbation, of the same order as the background valu e,\ni.e.a=ǫ0, for which one may expectlinear theory notto be\napplicable.Thebackgroundplasmabetais0 .2.\nFigure1 showsthe time evolutionof the variousenergiesof\nthe system, integrated over the whole box. Kinetic energy (i n\ngreen) grows quickly from zero to its maximum value, and is\nsubsequently damped to zero in the final equilibrium. A small\nfractionoftheinternalenergyisconvertedintomagnetice nergy\nat the new equilibrium. For this particular set up, in which t he\nperturbationhasbeenintroducedintheplasmapressure,it isthe\nplasmathatlosessomeofitsinitialinternalenergy,trans feringit\ntothemagneticfield.Note,thattheamountofenergytransfe red\nisdirectlyproportionaltothemagnitudeoftheperturbati on.The\ntotal energy (i.e. the sum of the magnetic, kinetic and inter nal\nenergies)isconservedtoanaccuracyof ∼10−7.\nThe 2D contour plots of the normalised plasma pressure,\ndensityandperpendicularcurrentdensityfortheinitiala ndfinal\nequilibriumstateareshowninFig.2forthiscase.Theiniti alin-\ncreaseofplasmapressurecreatesalocalizeddecreaseinpl asma\ndensity. The displacement of the magnetic field lines is hard to\nappreciate from the field lines themselves, but is clear from the\nnon-zeroperpendicularelectric current density, jz, that exists in\nthefinal equilibrium.\nFigures3 and4showverticalcutsat x=L/2andhorizontal\ncutsaty=L/2,respectively,throughthecontourplotsinFig.2,\nfor the relevant quantities for the initial and final state. I n these\nplots, a detailed comparisonof the numerical and analytica l so-\nlutionscanbemade.Theanalyticalpredictionsfortotalpr essure\nare very good, even though the amplitude of the perturbation is\nrelatively large ( a=ǫ0) and so, strictly speaking, our analytical\napproximationshould not hold.However,the linear approxi ma-\ntionfortheplasmadensity(redcrossesin Figures3and4)do es\nnot fit well. Instead, if we use Eq. (38), a much better fit for\nthe plasmadensityis found(bluecrossesin theplots), impl ying\nthat the processis approximatelyadiabatic. Since the nume rical\nexperiments have been performed using a full MHD code that\nsolves the non-linearequations, the processis not entirel y adia-\nbatic, but must have a finite amount of viscous heating that wi ll\nbecomeimportantasthe initialperturbationisincreased.\n4. Importanceofnon-lineareffects\nTo study how non-linearity a ffects the results as the magnitude\nof the initial perturbation increases, we focus again on the total\npressure. The total pressure of the final numerical equilibr ium\nmust be constant, whether the relaxation remains in the line ar6 Jorge Fuentes-Fern´ andez, ClareE.Parnell andAlanW.Hoo d: MHD Dynamical Relaxation of Coronal Magnetic Fields\nFig.2.Two-dimensionalcontourplotsof plasma pressure (top),\nplasmadensity(center)andcurrentdensity(bottom)inthe initial\nstate (left column) and final equilibrium (right column), fo r the\nsameexperimentasFig. 1.\nFig.3.Vertical cuts for the plasma pressure (left) and plasma\ndensity (right), for the same experiment as Fig. 1 and Fig. 2.\nInitial perturbed state (dashed) is compared with the final e qui-\nlibrium,asfoundbythefullMHDnumericalsimulations(sol id)\nand predicted by the linear analysis (red crosses). For the d en-\nsity predictions, the blue crosses represent predictions f rom the\nadiabaticconditiongivenbyEq.(38).\nregimeor not.On the otherhand,the analyticaldefinitionof to-\ntal pressure given by (13) is an approximation from the linea r\nanalysis, and will become less valid as the non-linear terms be-\ncome more important. We perform a series of experiments for\nvarious plasma beta values in which the relative amplitude o f\nthe initial perturbationis changedfrom a very small value, well\nwithin the linear regime, to a large value way outside it. Usi ng\ntheseexperiments,weinvestigatehowthefinaltotalpressu rede-\npartsfromthelinearpredictionsfordi fferentbackgroundplasma\nbetavalues.\nBut first, we recall that the 2D relaxation may be separated\ninto a verticalnon-magneticevolution(vertical redistri butionof\nplasma pressure)and a horizontalevolution(horizontalre distri-\nbution of total pressure), in which the total pressure in the ver-Fig.4.Horizontalcutsfortheplasmapressure(topleft),plasma\ndensity(topright),totalpressure(bottomleft)andmagne ticfield\nstrength (bottom left), for the same experiment as Figures 1 , 2\nand3.\ntical case is notdeterminedbythe linearanalysis.Thissuggests\nthat,effectively,inordertofindasignificanterrorinthefinalto-\ntal pressure for the 2D experiment,we will need very large va l-\nues of the initial two-dimensional perturbation. Hence, th e fol-\nlowing experimentshave been made for just a one-dimensiona l\nperturbationacrossthefieldlines.Theseresultsmaybemap ped\nontothoseforourintital 2Dperturbation,usingthefollow ing:\n/parenleftiggmax(p1)\np0/parenrightigg\n2D=L/integraltext\nyexp/parenleftbig−(y−b)2/2c2/parenrightbigdy/parenleftiggmax(p1)\np0/parenrightigg\n1D.\nFigure 5 shows the relative error of the linear aproximation\nin both 1D and 2D for the total pressure, as a function of the\namplitude of the initial perturbation, for five di fferent values of\nthe plasma beta (β=0.05,β=0.1,β=0.2,β=1.3andβ=2).\nThe bottom x-axis showsthe magnitudeof the one-dimensional\nperturbation,and the top x-axis shows the magnitudeof the ini-\ntial two-dimensional perturbation before its vertical exp ansion.\nThe erroron the y-axisis calculated as the maximumdi fference\nbetween the linear prediction and the numerical results for the\ntotalpressure,\nmax(|plin\nT−pnum\nT|)\npnum\nT,\nwherepnum\nTisthe final constant total pressureobtainedfromthe\nnumerical simulations, and, for our non-magneticinitial p ertur-\nbation,\nplin\nT(x)=p0+p1(x,∞)+B2\n0\n2µ0+B0By(x,∞)\nµ0,\nwithp1(x,∞) andBy(x,∞) being the final perturbed plasma\npressureand perturbedmagneticfield fromthe numericalsim u-\nlations.\nAsβ→∞, we expect the relative error of the linear anal-\nysis to tend to zero, independently of the perturbation, as i n\nthis case, the magnetice ffects dissapear, and the initial pressure\nperturbation completely redistributes to a well defined con stant\nvalue in the whole box. On the other hand, if β≪1, then the\nmagneticfield will dominateoverthe plasma contributions, andJorge Fuentes-Fern´ andez, ClareE.Parnell andAlanW.Hood : MHD Dynamical Relaxation of Coronal Magnetic Fields 7\nFig.5.Relative error in the linear prediction of the total pres-\nsure against the magnitude of the initial pressure perturba tion,\nforfivedifferentvaluesoftheplasmabeta.Theslowgrowthrate\noftheerror(non-lineare ffects)indicatesthevalidityofthelinear\nanalysisstudiedinthispaper.\nmuch larger values for the relativeinitial perturbation will be\nneeded for strictly leaving the linear regime. These two beh av-\niorscanbeseeninFig.5,wheretheplotsforlargeplasma-be tas\ntend to a smaller error, while the plots for small plasma-bet as\ntake longer to reach significant errors, i.e. to escape from t he\nlinear regime.Furthermore,we must not forgetthat we are he re\nonlytalkingabouttheinitialbackgroundplasmabeta,so al arge\nbackgroundbeta combined with a large initial perturbation will\nmakethefinalplasmabetaevenhigher.Thus,max( p1)/p0→∞\nwill implyβ→∞for the final equilibrium, so we expect the\ncurves of the relative error of the linear analysis to turn ba ck to\nzero as the initial perturbation is greatly increased. In te rms of\nenergyconservation,asthevelocityiszeroattheinitiala ndfinal\nstates,theintegraloverthewholedomainofinternalenerg yplus\nmagneticenergymustbeconserved:If β→∞,thentheinternal\nenergyismuchlargerthanthemagneticenergy,andwilljust re-\ndistribute the plasma pressure, without transferring any e nergy\nintothemagneticfield.\nOn the contrary, the final plasma density is entirely deter-\nmined by the linear analysis, in both the vertical and the hor i-\nzontal evolutions along and across the field lines, or in a bet ter\napproximation,bytheadiabaticcondition.Hence,thenon- linear\neffectsfortheplasmadensitywill growmuchquicker,asshown\nin Fig. 6. These last numericalexperimentshave been made fo r\nthe original two-dimensional Gaussian perturbation. The e rror\nonthey-axisisgivenby\nmax(ρad−ρnum)\nρnum,\nwhereρadis the plasma density given by Eq. (38), and ρnumis\nthefinal densityobtainedwiththenumericalexperiments.\nThe relative error in the plasma density is considerably big -\nger than the relative error in pressure, and so, for only a sma ll\nchangein p1/p0inthe2Dcase,wefindalargeerrorin ρ.Asthis\nerror quickly reaches significant values, the plasma beta pl ays\nmuch less of a role for the non-lineare ffects in the plasma den-\nsity thanin theabovetotal pressure.Fig.6.Relative error of the density predicted assuming an adi-\nabatic evolution, with Eq. 38, against the magnitude of the 2 D\ninitial pressureperturbation,forthreevaluesoftheplas mabeta.\nNote, that the x-axis in this plot is to be compared with the top\nx-axisin Fig.5.\n5. Summaryand Conclusions\nWe have presentedanalyticaland numericalcalculationsfo r the\n2D magnetohydrodynamicrelaxation of an untwisted perturb ed\nmagnetic system embedded in a plasma with beta of any size,\nandfindafinalequlibriumstatewhichdi fferssubstantiallyfrom\nthe initial background configuration. The equilibrium reac hed\nis a non force-free state in which the plasma pressure gradie nts\nare balanced by the magnetic Lorentz force. For a set of speci -\nfiedboundaries,allthehydromagneticquantitiesarefully deter-\nminedbytheinitial perturbedstate.\nThe initial disturbance evolves into the final relaxed state\nby different families of magnetoacoustic waves, dissipated via\nviscousdamping.Fast magnetoacousticwavespropagateacr oss\nthe field lines in the horizontal direction, slow magnetoaco ustic\nwaves redistribute the thermodynamicquantities along the field\nin the vertical direction, and an extra contributionof slow mag-\nnetoacousticwavespropagatesalongthemagneticfieldline sin-\ntroducingamagnetictensionterm.Nevertheless,theselas t slow\nmagnetoacoustic waves dissipate the magnetic tension in su ch\na way that it is totally unimportant when determining the fi-\nnal equilibrium distributions. The vertical redistributi on of the\nplasmapressureto a homogeneousvaluedemandsthe magnetic\ntensiontodissapearcompletely,soboththeplasmapressur eand\ntotalpressureareone-dimensionalat theendofthe process .\nWithin the linear regime, the final distributions are com-\npletely independent of the viscosity, even though it is requ ired\ntopermittherelaxationtoocurr,asitistheonlydampingme ch-\nanism of the waves. An increase in the viscosity enhances the\ndiffusive term in the wave equation,and so, acceleratesthe pro-\ncess, but the final distribution is not modified. Instead, in t he\nfinal equilibrium, all the quantities are simply determined by\nthe behaviour of the final equilibrium total pressure, invol ving\nplasma andmagnetic e ffects. Hence,the final equilibriumstates\nforplasmapressureandmagneticfielddonotdi fferifthe initial\nperturbationisofthedensity,temperatureorinternalene rgy.\nFinally,byinvestigatingthelinearregime,wehavebeenab le\nto make analytical predictions for the final MHS equilibrium ,\neven when the regime is far from linear. The linear predictio ns\nremain remarkably valid even outside the linear regime, as t he8 Jorge Fuentes-Fern´ andez, ClareE.Parnell andAlanW.Hoo d: MHD Dynamical Relaxation of Coronal Magnetic Fields\ngrowth rate of the non-ideal e ffects is very small, compared to\ntheinitial perturbations.\nIn this paper, the introduction of plasma e ffects in the re-\nlaxation of hydromagnetic systems have been studied in sim-\nple schemes, producing a series of analytical predictions w hich\nhave been confirmed by our numerical results. By starting wit h\na uniform magnetic field, we have reached a state in which the\nmagnetic field itself remains almost uniform, but where some\ncurrent density has been built. Also, even in this simple con -\nfiguration, a non-neglectible amount of energy has been tran s-\nferredfromtheplasmatothemagneticfield.Theseimplicati ons\nofenergytransferduringtherelaxationprocessindicatet hatthis\nprocess will be of importance in the solar corona, specially in\nthe study of magnetic null points and their surroundings. Nu ll\npoints have been found to have a reasonable populationdensi ty\nin the Solar Corona by Longcope&Parnell (2009). In further\nstudies, we will consider more complex scenarios, introduc ing\ntwo-dimensional null points and starting to consider the im pli-\ncationsofnon-zeroplasmabetasin three-dimensionalmagn etic\nenvironments.\nReferences\nAmari, T.,Boulmezaoud, T.Z.,&Maday, Y.1998, A&A,339, 252\nArber, T. D., Longbottom, A. W., Gerrard, C. L., & Milne, A. M. 2001,\nComputational Phys.,171, 151\nBrowning,P.K.,Gerrard,C.,Hood,A.W.,Kevis,R.,&vander Linden,R.A.M.\n2008, A&A,485, 837\nDeRosa, M.L.,Schrijver, C. J.,Barnes, G.,et al. 2009, ApJ, 696, 1780\nGary, G.A.2001, Sol. Phys.,203, 71\nGary, G.A.2009, Sol. Phys.,257, 271\nHeyvaerts, J.&Priest, E.R. 1984, A&A,137, 63\nHood,A.W.,Browning,P.K.,&vanderLinden,R.A.M.2009,A& A,506,913\nInoue, S., Kusano, K., Masuda, S., et al. 2008, in Astronomic al Society of\nthe Pacific Conference Series, Vol. 397, First Results From H inode, ed.\nS.A.Matthews, J.M.Davis, &L.K.Harra, 110– +\nJanse, Å.M.& Low,B.C. 2009, ApJ,690, 1089\nJi,H.,Huang, G.,& Wang,H.2007, ApJ,660, 893\nLongcope, D.W.&Parnell, C. E.2009, Sol. Phys.,254, 51\nLow,B. C.2006, ApJ,649, 1064\nMackay, D.H.& van Ballegooijen, A.A. 2006, ApJ,641, 577\nMcLaughlin, J.A.&Hood, A.W.2006, A&A,459,641\nMetcalf, T.R.,DeRosa, M.L.,Schrijver, C. J.,et al. 2008, S ol. Phys.,247,269\nMiller, K.,Fornberg, B., Flyer, N.,&Low,B. C.2009, ApJ, 69 0,720\nPontin, D.I.,Hornig, G.,Wilmot-Smith, A.L.,&Craig, I.J. D.2009,ApJ,700,\n1449\nPriest, E.R. 1982, Solar magneto-hydrodynamics, 74P– +\nR´ egnier, S. 2008, in Astronomical Society of the Pacific Con ference Series,\nVol. 397, First Results From Hinode, ed. S. A. Matthews, J. M. Davis, &\nL.K.Harra, 75–+\nRuan, P.,Wiegelmann, T.,Inhester, B.,et al. 2008, A&A,481 , 827\nSchrijver, C.J.,DeRosa, M.L.,Metcalf, T.,etal. 2008, ApJ ,675, 1637\nSchrijver, C.J.,DeRosa, M.L.,Metcalf, T.R.,et al. 2006, S ol. Phys.,235,161\nTaylor, J.B. 1974, Phys.Rev. Lett., 33, 1139\nWheatland, M.S.& R´ egnier, S.2009, ApJ,700, L88\nWiegelmann, T.,Inhester, B.,&Sakurai, T.2006, Sol. Phys. ,233, 215\nWiegelmann, T., Thalmann, J. K., Schrijver, C. J., De Rosa, M . L., & Metcalf,\nT.R.2008, Sol. Phys.,247, 249" }, { "title": "2401.09803v1.Multithermal_apparent_damping_of_slow_waves_due_to_strands_with_a_Gaussian_temperature_distribution.pdf", "content": "Astronomy &Astrophysics manuscript no. multiT_dampingslow_for_arxiv ©ESO 2024\nJanuary 19, 2024\nMultithermal apparent damping of slow waves due to strands with\na Gaussian temperature distribution\nT. Van Doorsselaere1, S. Krishna Prasad2, V . Pant2, D. Banerjee2,3,4, and A. Hood5\n1Centre for mathematical Plasma Astrophysics, Department of Mathematics, KU Leuven, Celestijnenlaan 200B, B-3001 Leuven,\nBelgium\ne-mail: tom.vandoorsselaere@kuleuven.be\n2Aryabhatta Research Institute of Observational Sciences, Nainital, India\n3Indian Institute of Astrophysics, Koramangala, 560034, Bengaluru, India\n4Center of Excellence in Space Sciences, IISER, 741246, Kolkata, India\n5School of Mathematics and Statistics, University of St Andrews, St Andrews, Fife, KY16 9SS, UK\nReceived ; accepted\nABSTRACT\nContext. Slow waves in solar coronal loops are strongly damped. The current theory of damping by thermal conduction cannot explain\nsome observational features.\nAims. We investigate the propagation of slow waves in a coronal loop built up from strands of di fferent temperatures.\nMethods. We consider the loop to have a multithermal, Gaussian temperature distribution. The di fferent propagation speeds in di ffer-\nent strands lead to an multithermal apparent damping of the wave, similar to observational phase mixing. We use an analytical model\nto predict the damping length and propagation speed for the slow waves, including in imaging with filter telescopes.\nResults. We compare the damping length due to this multithermal apparent damping with damping due to thermal conduction and find\nthat the multithermal apparent damping is more important for shorter period slow waves. We have found the influence of instrument\nfilters on the wave’s propagation speed and damping. This allows us to compare our analytical theory to forward models of numerical\nsimulations.\nConclusions. We find that our analytical model matches the numerical simulations very well. Moreover, we o ffer an outlook for using\nthe slow wave properties to infer the loop’s thermal properties.\nKey words. Magnetohydrodynamics (MHD) – Plasmas – Waves – Methods: analytical – Methods: numerical – Sun: oscillations\n1. Introduction\nSince the turn of the century, slow waves in coronal loops are\nregularly observed through high resolution space observations\n(Berghmans & Clette 1999). These waves are seen as propagat-\ning intensity disturbances along open magnetic field or in the\nfootpoint of loops (De Moortel et al. 2002; Krishna Prasad et al.\n2012). In previous years, there was a debate on their interpreta-\ntion in terms of slow waves or periodic flows (e.g. De Moortel\net al. 2015), but for coronal loops or fans rooted in sunspots,\nthere is a consensus that these are definitely slow waves driven\nby p-mode wave leakage in the sunspot (Banerjee et al. 2021).\nSlow waves in coronal loops are observed to be very heav-\nily damped. Traditionally, it has been thought that the damp-\ning is caused by thermal conduction, after an extensive review\nof damping mechanisms by De Moortel & Hood (2003, 2004).\nStill, some of the observed properties of the damping may not be\nexplained adequately by this traditional approach. For instance,\nin closed-field regions, it was found that the damping scales\nwith the period with a positive coe fficient (Krishna Prasad et al.\n2014), which is still more-or-less compatible with the damp-\ning by thermal conduction (Mandal et al. 2016). However, for\nopen-field regions in coronal holes, it was found that the damp-\ning scales with the period with a negative coe fficient (Krishna\nPrasad et al. 2014, 2017), which is incompatible with that damp-\ning theory. Gupta (2014) finds di fferent damping behaviour inslow waves at di fferent heights. At larger heights (10-70 Mm),\nthey find shorter damping lengths for short period waves (as ex-\npected due to thermal conduction) but closer to the limb ( <10\nMm), the long period waves ( >6 min) appear to damp faster. Ad-\nditionally, Mandal et al. (2018) have shown, through a statistical\nstudy, that the damping length of slow waves in polar regions, in-\ndeed display a negative dependence on oscillation period. More-\nover, Krishna Prasad et al. (2019) find that the observed damp-\ning lengths of slow waves are much shorter than those expected\nfrom the theory of thermal conduction. Thus, it seems that other\ndamping mechanisms are also at work.\nAnother point to consider is that the propagation speeds of\nslow modes in a loop seems to depend on the filter passband of\nthe spacecraft that is used (King et al. 2003; Kiddie et al. 2012;\nUritsky et al. 2013). This, too, seems to be incompatible with a\nslow mode wave propagating through a monolithic loop, subject\nto damping by thermal conduction. It was originally attributed\nto the fact that two adjacent loops (or aligned along the line-of-\nsight) would be observed in di fferent filters. This may be correct,\nbut it leaves the question why the slow waves are so coherently\nin phase between the observed structures.\nAnother reason for disagreeing with the damping of slow\nwaves by thermal conduction should be that it was previously\nobserved that slow waves are damped with a Gaussian envelope\n(Krishna Prasad et al. 2014). This feature could not be explained\nby damping by thermal conduction, nor resonant damping of\nArticle number, page 1 of 9arXiv:2401.09803v1 [astro-ph.SR] 18 Jan 2024A&A proofs: manuscript no. multiT_dampingslow_for_arxiv\nslow waves in the cusp continuum (Yu et al. 2017b,a; Geeraerts\net al. 2022). The latter is despite the fact that resonant absorp-\ntion of kink modes in the Alfvén continuum has been convinc-\ningly shown to result in a Gaussian damping profile (Pascoe et al.\n2012, 2017, 2022). However, the mechanism at work for res-\nonant absorption in the Alfvén continuum (Hood et al. 2013)\ndoes not seem to carry over to slow waves (Hood, 2015, private\ncommunication).\nFurthermore, it was argued by Wang et al. (2015) that the\nthermal conduction coe fficient is significantly suppressed in ob-\nserved coronal loops. Despite the suppression of the thermal con-\nduction, they still find a strong damping of slow waves. In their\npaper, they argue that this damping is caused by an enhanced\ncompressive viscosity (Wang & Ofman 2019). However, here\nwe offer a suggestion that may also explain the observed strong\ndamping of slow waves, without invoking unrealistically high\nviscosity coe fficients (a factor 10 higher than normally consid-\nered, Wang & Ofman 2019). The model we propose below, does\nnot need either thermal conduction or viscosity to result in ap-\nparent damping.\nIn DC heating models for the corona, it is thought that the\nloops are built out of isolated strands (e.g. Aschwanden et al.\n2000), which are each individually heated by nanoflares after\nwhich the higher temperature is redistributed by thermal con-\nduction only along the magnetic field, but inhibited across (e.g.\nWilliams et al. 2021, and references therein). They are conse-\nquently modelled as a collection of 1D field lines and their ther-\nmodynamic evolution. This is in contrast to numerical evidence\nthat loop strands have a short life time, because of the mixing by\ntransverse waves Magyar & Van Doorsselaere (2016). The latter\nwould result in a more continuous transverse temperature profile\nperpendicular to the magnetic field (Judge 2023).\nIn models of AC heating, transverse waves lead to turbulent\nbehaviour in the loop boundary and entire cross-section (Karam-\npelas & Van Doorsselaere 2018), leading to patchy heating in\nthe cross-section, or in the turbulent layers (Van Doorsselaere\net al. 2018; Shi et al. 2021). Despite the di fferences between AC\nand DC heating models, it is safe to say that coronal loops do\nnot have a uniform temperature across their cross-section. This\nwill have a major impact on the propagation properties of slow\nwaves in those non-uniform temperature profiles. Here we will\nshow that this leads to extra apparent damping (which we call\n“multithermal apparent damping” or MAD) and di fferent propa-\ngation speeds in di fferent filter channels. In the future, this will\nallow us to infer the thermal structure of coronal loops from the\npropagation and damping behaviour. Aside from this potential\nfor probing the coronal thermal structure, nowadays slow waves\nare also considered for their ability to diagnose the coronal heat-\ning function (Kolotkov et al. 2019) through their perturbation of\nthe energy balance equation for the background corona. Thus, it\nseems that slow waves are the optimal MHD waves for seismol-\nogy of thermal e ffects in the solar corona (perhaps aside from\nthe entropy mode).\nEven though our results were derived independently, it was\npointed out in discussions at conferences that the physical e ffect\nwe consider is the same as was considered by V oitenko et al.\n(2005)1. They modelled the propagation of sound waves in a\nmulti-stranded loop with strands drawn from a uniform distri-\nbution, as seen in a top-hat shaped instrumental filter. Here we\ngo much beyond that initial description of this phenomena.\n1In fact, given their pioneering idea on this, as much as 18 years be-\nfore this manuscript, we may consider naming the multithermal appar-\nent damping as V oitenko-damping of slow waves.2. Results – theoretical models\nWe model a coronal loop as a superposition of strands, each with\ntheir own temperature and associated sound speed vs. However,\nthe model also carries over to a loop with a temperature continu-\nously varying in its cross-section. Both of these models have no\ntemperature variation along the magnetic field. In these models,\na sound wave front is launched at the footpoint. We describe the\npropagation of the sound wave in the multithermal plasma.\n2.1. Intuition\nWe take the z-direction along the uniform magnetic field, and\nonly consider the hydrodynamic behaviour along the magnetic\nfield lines (e.g. De Moortel & Hood 2003; V oitenko et al. 2005;\nMandal et al. 2016). We first think about the position zpof the\npeak perturbation on the strands. We have that\nzp=z0+vst, (1)\nwhere z0is the height of the initial excitation of the wave, in a\nstrand with sound speed vs. We consider the initial position of\nthe peak z0to be independent of the strand, mimicking a joint\nimpulsive excitation of the pulse low down in the atmosphere. In\nthis paper, we consider a loop for which the strands have a sound\nspeed that is randomly drawn from a normal distribution centred\non ¯vand standard deviation σv\nvs∼N(¯v,σ2\nv). (2)\nThe distribution of the temperature in these strands is tightly\nrelated to the heating mechanism, which is currently not well\nunderstood (e.g. Van Doorsselaere et al. 2020). Thus, this as-\nsumption of a Gaussian distribution of strand’s sound speed is\nan ad-hoc assumption in this paper. A sketch of the considered\nconfiguration is included in Fig. 1.\nWith such a Gaussian distribution of the strands, we then find\nthat the peak positions zpare also a Gaussian distribution. Fol-\nlowing well-known rules for transforming random variables in\nstatistics, we find that its distribution is\nzp∼N(z0+¯vt,t2σ2\nv). (3)\nWe then consider all pulse perturbations on each strand to have\nthe same amplitude. The line-of-sight integration over all these\nstrands will then result in an intensity variation that is modelled\nwell by this Eq. 3. This shows that the peak position distribution\n(and integrated intensity signal) propagates up with the average\nsound speed ¯ vin the loop bundle. It also shows that the peak\nposition distribution steadily widens linearly in time, because its\nstandard deviation goes as tσv.\nThe crucial realisation for understanding the multithermal\napparent damping of sound waves is that the normalising factor\nof the Gaussian distribution with a certain σis given as 1 /σ√\n2π.\nApplying this for the distribution of zp, we find that its peak value\nwill vary over time as\n1\ntσv√\n2π, (4)\nand thus the wave will have a multithermal apparent damping\nthat is proportional to t−1. Indeed, given that there is no dis-\nsipation (all the wave energyR\nvsR\nzρv2\nzdvsdzis still in the in-\nfinitely long system), the damping is only apparent, because of\nthe spreading of the wave front over time. This is because of\nthe di fferent propagation speeds in each strand, leading to an in-\ncreasing spread in z. Thus, it is very similar to the process of\nphase mixing.\nArticle number, page 2 of 9T. Van Doorsselaere et al.: Multithermal apparent damping of slow waves\n/vectorB\nz0\nFig. 1. A schematic representation of the considered configuration.\nThree magnetic strands are shown. The cyan Gaussian pulses are ex-\ncited at time t=0 at z=z0on all strands simultaneously. On each\nstrand they propagate with a di fferent speed, first to the blue line and\nthen the purple line. The resultant observed intensity, as integrated over\nthe di fferent strands, is given by the red line, which shows the multi-\nthermal apparent damping and broadening.\n2.2. Gaussian pulses\nLet us now build on this intuition to describe a system with an\ninitial Gaussian pulse in the density perturbation, as could be the\nresult of an impulsive excitation at the footpoint of the loop be-\ncause of (e.g.) granular bu ffeting or a reconnection event. We\nimagine a group of strands all simultaneously excited with a\npulse W(z,0) of the form\nW(z,0)=aexp \n−(z−z0)2\n2w2!\n(5)\nat position z0, with pulse width wand amplitude a. Because of\nthe propagation of the pulse on each individual strand, at a later\ntime, it will appear as\nW(z,t)=aexp \n−(z−(z0+vst))2\n2w2!\n, (6)\nin which vsdiffers from strand to strand. A sketch of the config-\nuration is shown in Fig. 1.\nNow we look at the integrated signal for a Gaussian strand\ndistribution for which vs∼N (¯v,σ2\nv) as before. We consider the\nintegral of the di fferent wavepackets (Eq. 6), with the distribu-\ntion of the strand’s sound speeds as weight function. So, in a\nsense, the integral is computing the intensity of a line-of-sight\nintegration through the multistranded loop with a Gaussian dis-\ntribution in the DEM (di fferential emission measure, see e.g. VanDoorsselaere et al. 2018). The emission measure is defined asR\nzn2dz(with electron density n), which linearises to 2R\nzn0n1dz\nfor background density n0and density perturbation n1. The in-\ntegral over the line-of-sight covers all strands with density n0in\nour loop model, and thus the wave perturbation W(z,t) has to be\nmultiplied with the Gaussian strand distribution. Moreover, the\nintegral over z(which traverses the entire loop system) is equiv-\nalent to integrating over all strands (in sound speed space).\nWith this reasoning, the total signal S(z,t) is then given as\nS(z,t)=Z\nvs1\nσv√\n2πexp \n−(vs−¯v)2\n2σ2v!\naexp \n−(z−(z0+vst))2\n2w2!\ndvs,\n(7)\nor\nS(z,t)=a\nσv√\n2πZ\nvsexp \n−1\n2\"(vs−¯v)2\nσ2v+(z−(z0+vst))2\nw2#!\ndvs.\n(8)\nCompleting the square, the evaluation of the integral is given as\nS(z,t)=awp\nw2+σ2vt2exp \n−(z−(z0+¯vt))2\n2(w2+σ2vt2)!\n. (9)\nThis is a signal that peaks at z0+¯vtand thus propagates with the\naverage sound speed upwards. Its Gaussian width (as a function\nofz) is given byp\nw2+σ2vt2, showing that it steadily increases\nin a hyperbolic fashion. For large t, the width increases approxi-\nmately linearly.\nAs in Subsect. 2.1, we also note here that the amplitude of\nthe peak signal (at z=z0+¯vt, i.e. co-propagating with the wave)\ndecreases steadily over time. Its decay d(t) from its initial ampli-\ntude is given as\nd(t)=wp\nw2+σ2vt2=1q\n1+σ2vt2\nw2. (10)\nFor large time t, this scales thus as t−1, recovering the results of\nSubsect. 2.1. These latter results may also be recovered consid-\nering the limit of w→0, corresponding to an initial δ-function\nperturbation.\nIn Fig. 2, we display the predicted damping envelope of\nEq. 10, and compare it to a Monte Carlo simulation of a Gaus-\nsian wave packet on di fferent strands. For the Monte Carlo\nsimulation, we have drawn 1000 vsfrom the normal distribu-\ntionN(¯v,σ2\nv), with ¯ v=152km /s (corresponding to 1MK) and\nσv=26.4km/s (for a motivation of these particular values, see\nSubsect. 3.2). The correspondence between the analytical solu-\ntion and the Monte Carlo simulation is excellent.\nWe may calculate the damping time τas the e-folding time\nof this damping profile d(t). We have then that\ne−1=d(τ)=1q\n1+σ2vτ2\nw2(11)\nresulting in\nσvτ\nw=√\ne2−1≈2.53 orτ=w\nσv√\ne2−1. (12)\nWith a substitution ∆z≡z−z0=¯vt, we may transform Eq. 10\nto a damping profile as a function of ∆z. We find that\nd(∆z)=1q\n1+σ2v∆z2\nw2¯v2. (13)\nArticle number, page 3 of 9A&A proofs: manuscript no. multiT_dampingslow_for_arxiv\nFig. 2. Comparison of Monte Carlo simulation with the analytical result.\nThe analytically predicted envelope (Eq. 10) is drawn with the dark\nblue line. The evolution of the Monte Carlo wave packet is drawn with\nthe light blue to pink colour, progressively in time. The mean sound\nspeed was taken as ¯ v=152km /s and the spread in sound speed as σv=\n26.4km/s\nMaking the same reasoning as in the derivation of the damping\ntimeτ, we may also derive the damping length Ld.\ne−1=d(Ld)=1q\n1+σ2vL2\nd\nw2¯v2orLd=w¯v\nσv√\ne2−1=¯vτ. (14)\n2.3. Driven waves\nNow we consider the case of driven, sinusoidal waves. At a cer-\ntain height z=0, a periodic driver is inserted, resulting in propa-\ngating waves asin(kz−ωt), with amplitude a, frequency ωand\nwavenumber k. The resulting intensity signal of the loop bundle\nis then given as an integral (similar to Eq. 7)\nS(z,t)=Z\nvs1\nσv√\n2πexp \n−(vs−¯v)2\n2σ2v!\nasin(kz−ωt)dvs\n=a\nσv√\n2πZ\nvsexp \n−(vs−¯v)2\n2σ2v!\nsin ωz\nvs−ωt!\ndvs, (15)\nwhere we have used the dispersion relation k=ω/vs. The latter\nintegral is not analytically solvable. We can still compare it to a\nMonte Carlo simulation with a 1000 vsdrawn from aN(¯v,σ2\nv)\ndistribution and summed up (see Fig. 3). The Monte Carlo simu-\nlation is shown with the blue line, while the full integral in Eq. 15\nis shown with the light orange line. The two lines closely match,\nand the deviation is due to the finite number of drawn vsvalues.\nFor a higher number of draws, the two lines converge.\nFurther analytical progress is possible by making a Taylor\napproximation of the denominator in the sine:\nsin ωz\nvs−ωt!\n=sin\u0012ωz\n¯v+δv−ωt\u0013\n(16)\n≈sin\u0012ωz\n¯v(1−δv\n¯v)−ωt\u0013\n. (17)\nThis approximation is valid if δv≪¯v. Since 95% of the contri-\nbution to the full integral (Eq. 15) is for |δv|≤3σv, the Taylor\napproximation is reasonably satisfied for our considered param-\neters of ¯ v=152km /s andσv=26.4km/s, for which we sub-\nsequently have|δv| ≤ 3σv=79.2km/s⪉¯v=152km /s. So,\nFig. 3. Comparison of Monte Carlo simulation of driven sine functions\n(blue line) with the full integral (light orange) and the approximations\nin Eq. 19 (green), all normalised to the starting value of 1. The expected\nGaussian damping envelope (Eq. 26) is shown in red. The time is ar-\nbitrarily chosen to be t=0, and the mean sound speed 152km /s and\nspread in sound speed σv=26.4km/s was chosen as before.\nthe assumption δv≪¯vseems to be su fficiently well satisfied in\nloops that are not too extremely multithermal (i.e. with σv≲¯v).\nThis Taylor approximation allows to rewrite S(z,t) as\nS(z,t)≈a\nσv√\n2π(\nsin\u0012ωz\n¯v−ωt\u0013Z\nδvexp \n−δv2\n2σ2v!\ncos\u0012ωzδv\n¯v2\u0013\ndδv\n−cos\u0012ωz\n¯v−ωt\u0013Z\nδvexp \n−δv2\n2σ2v!\nsin\u0012ωzδv\n¯v2\u0013\ndδv)\n.(18)\nIt turns out that the rightmost integral in this expression is exactly\n0, because its integrand is an odd function in δv. Thus, we have\nthat\nS(z,t)≈a\nσv√\n2πsin\u0012ωz\n¯v−ωt\u0013Z\nδvexp \n−δv2\n2σ2v!\ncos\u0012ωzδv\n¯v2\u0013\ndδv.\n(19)\nThe numerically calculated result of Eq. 19 is shown in Fig. 3\nwith the green line. It matches the Monte Carlo simulations\n(blue) and full integral (light orange) reasonably well. The in-\ntegral in Eq. 19 can be calculated analytically by writing it as a\ncomplex function:\nZ\nδvexp \n−δv2\n2σ2v!\ncos\u0012ωzδv\n¯v2\u0013\ndδv=ℜZ\nδvexp \n−δv2\n2σ2v+ıωzδv\n¯v2!\ndδv.\n(20)\nWe subsequently have\nℜZ\nδvexp \n−δv2\n2σ2v+ıωzδv\n¯v2!\ndδv (21)\n=ℜZ\nδvexp−δv√\n2σv−ı√\n2σvωz\n2¯v22\n−σ2\nvω2\n2¯v4z2dδv (22)\n=exp \n−σ2\nvω2\n2¯v4z2!\nℜZ\nδvexp−δv√\n2σv−ı√\n2σvωz\n2¯v22dδv\n(23)\n=√\n2πσvexp \n−σ2\nvω2\n2¯v4z2!\n. (24)\nArticle number, page 4 of 9T. Van Doorsselaere et al.: Multithermal apparent damping of slow waves\nFig. 4. Here we show the expected damping lengths (in Mm) as a func-\ntion of period (in s), for both multithermal apparent damping (Eq. 26)\nand thermal conduction (Eq. 3 in Mandal et al. 2016). The density was\ntaken to be 109cm−3, the mean temperature as 106K, and the spread in\ntemperature as σv=26.4km/s.\nInserting this in Eq. 19, we find as end result\nS(z,t)≈aexp \n−σ2\nvω2\n2¯v4z2!\nsin\u0012ωz\n¯v−ωt\u0013\n. (25)\nThe wave is thus propagating with the average sound speed, and\nhas additionally a Gaussian damping envelope with a Gaussian\ndamping length (keeping the traditional factor 2 in the denomi-\nnator, Pascoe et al. 2017)\nLG=¯v2\nσvω. (26)\nFor the values considered in this paper (¯ v=152km /s,σv=\n26.4km/s,ω=2π/180s), this reduces to a damping length\nLG=25.1Mm.\nThe formula shows that the damping length is inversely pro-\nportional to the frequency. This is a di fferent dependence than\nthe thermal conduction damping length, which is proportional to\nω−2. In Fig. 4, we compare the multithermal apparent damping\nto the damping by thermal conduction. For the latter, we have\ntaken the results in Mandal et al. (2016), and these are shown\nwith the blue line. The other, light orange line corresponds to\nEq. 26. The graph shows that for intermediate periods (i.e. be-\ntween 300s and 1000s), the damping by thermal conduction is\ncomparable. However, for shorter or longer periods, the mul-\ntithermal apparent damping becomes more significant. Caution\nis appropriate here, because the multithermal apparent damping\nhas a Gaussian damping profile which is compared in this graph\nto the exponential damping profile of the thermal conduction.\n3. Results – loops in filter images\n3.1. Influence of a finite filter in imaging observations\nLet us consider the influence of a filter on the observability and\nmultithermal apparent damping of slow waves. For a filter Fde-\nscribed by a Gaussian function in vs-space as\nF(vs)=aFexp−(vs−vF)2\n2σ2\nF (27)with amplitude aF, mean vFand widthσF, the resulting observed\nsignal (equivalent to Eq. 7) would be\nS(z,t)=Z\nvs1\nσv√\n2πexp \n−(vs−¯v)2\n2σ2v!\naFexp−(vs−vF)2\n2σ2\nF\naexp \n−(z−(z0+vst))2\n2w2!\ndvs.(28)\nThe first two Gaussian distributions may be combined by realis-\ning that\nexp \n−(vs−¯v)2\n2σ2v!\naFexp−(vs−vF)2\n2σ2\nF\n=aFexp−(vF−¯v)2\n2(σ2\nF+σ2v)exp \n−(vs−V)2\n2Σ2!\n,(29)\nwhere we have introduced the notation\n1\nΣ2=1\nσ2\nF+1\nσ2vV=σ2\nvvF+σ2\nF¯v\nσ2\nF+σ2v(30)\nfor the width Σand average Vof the resulting Gaussian. This\nmeans that the width of the Gaussian is always decreased, due to\nthe harmonic average.\nThese expressions for ΣandVmay then be inserted in Eq. 9,\nwhile also remembering to also take the extra factors of Eq. 29\nalong and incorporate them with a. This will result in\nS(z,t)=aaFw√\nw2+ Σ2t2exp−(vF−¯v)2\n2(σ2\nF+σ2v)exp \n−(z−(z0+Vt))2\n2(w2+ Σ2t2)!\n.\n(31)\nAs before, the damping (following a wave packet, at a ray of\nz=z0+Vt) will be as\nd(t)=w√\nw2+ Σ2t2. (32)\nThis is weaker damping than in the non-filtered case, because\nΣ<σ v.\nLikewise, we may also insert ΣandVforσvand ¯vrespectively\nin the Gaussian damping lengths (Eq. 26):\nLG=V2\nΣω. (33)\nIt is also possible to use the propagation speed in di fferent\nfilters to estimate the temperature spread σvand the mean tem-\nperature ¯ v. This model naturally explains the di fferent propaga-\ntion speeds in di fferent filter channels and this di fference may be\nused to measure the loop’s fundamental thermal properties and\nquantify its DEM. As in Eq. 30, we see that σFandvFare known\nfor each filter. Then the propagation speed Vmay be measured\nin different filters, allowing us to estimate ¯ vandσvthrough least\nsquares fitting.\nMoreover, the e ffective filter width σFis much larger for imag-\ning observations than spectral observations. Thus, it is to be ex-\npected that imaging observations are much more profoundly im-\npacted by the e ffect of multithermal apparent damping. Spectral\nobservations will experience very little damping from the multi-\nthermal apparent damping e ffect, given the narrowness of the ef-\nfective filter. Thus, we may use the combination and contrast be-\ntween imaging and spectral observations to disentangle the real\ndamping mechanism from the multithermal apparent damping.\nThat opens exciting prospects for seismology of thermal proper-\nties of coronal loops.\nArticle number, page 5 of 9A&A proofs: manuscript no. multiT_dampingslow_for_arxiv\n3.2. Comparison with simulations\nHere, we verify the analytical results described in the previous\nsections using a 3D MHD multithermal loop model similar to\nthe one presented in Krishna Prasad & Van Doorsselaere (2023).\nThey solve the ideal MHD equations with MPI-AMRV AC (Porth\net al. 2014), where only numerical di ffusion is present and no\nexplicit di ffusive terms. They consider a bundle of 33 vertical\nstrands with randomly assigned plasma temperatures and den-\nsities to represent a coronal loop, similar to the setup we con-\nsider in this paper. To elaborate, the plasma temperature T(num-\nber density n) for each strand is selected from a random nor-\nmal distribution whose peak value corresponds to log T=6.0\n(logn=9.2), with a standard width of 0.15 (0.10). The plasma\ntemperature outside the strands and that outside the loop are kept\nat the same value, 1 MK. The corresponding number densities\nare fixed at 5×108cm−3. The magnetic field is vertical and par-\nallel to the axis of the loop. For further details on the simulation\nsetup, we refer the interested reader to Krishna Prasad & Van\nDoorsselaere (2023).\nConsidering the peak ( µlogT=6.0) and the width ( σlogT=\n0.15) values of temperature distribution in the simulations, and\nassuming that the resulting distribution of sound speeds is su ffi-\nciently normal (so that the theory in Sec. 2 applies), we can es-\ntimate the sound speed distribution properties from the temper-\nature distribution. For that, we calculate that log T=6.0±0.15\ncorresponds to a sound speed value of vs=152+28\n−24km/s, by\ncalculating the sound speed for µlogTandµlogT±σlogTsepa-\nrately. So, in what follows we take the following parameters:\n¯v=152km/s andσv=26.4km/s. Since the period of the driver\nin the simulation was 180s, this ¯ vis expected to result in a wave-\nlength of 27.4Mm. The wavelength values λobtained in Fig. 5\nindicate a good agreement with this.\nIn their model, Krishna Prasad & Van Doorsselaere excite\nslow magneto-acoustic waves within the loop by periodically\n(period≈180 s) driving the vertical velocity ( vz) at the bottom\nboundary of the loop, with an amplitude of ≈7.6 km s−1which\nis approximately 5% of the sound speed at 1MK. This driving\namplitude was chosen to be small, because we wanted to avoid\nany damping caused by non-linear e ffects of the waves. We use\nthe same driver in this study, however, to highlight the multi-\nthermal e ffects we restrict the spatial location of the driver to\nthe positions of the strands. In other words, the amplitude of the\ndriver is zero outside the strand locations and consequently the\noscillations are restricted to the strands. Once the generated slow\nwaves start approaching the top boundary of the loop, we com-\npute the mean density and vertical velocity ( vz) across the loop\nas a function of distance along the loop to analyse how the oscil-\nlation amplitudes evolve. Fig. 5 displays the density and vzpro-\nfiles along the loop in the top and bottom panels respectively. As\ncan be seen, the oscillations appear to damp very quickly. For a\nproper quantitative assessment, we measure the damping lengths\nby fitting the data with the following damped sinusoid function\nf(z)=A0exp−z2\n2L2\ndsin 2πz\nλ+ϕ!\n+b1z+b0. (34)\nHere zis the coordinate along the loop axis, A0is the maximum\namplitude, Ldis the damping length, λis the wavelength, ϕis the\ninitial phase, and b1andb0are appropriate constants. It may be\nnoted that this function describes a Gaussian damped sine wave\nsimilar to that described in Pascoe et al. (2016). Although an\nexponential damping is generally considered for slow waves, as\ndescribed in Sec. 2.3, the multithermal apparent damping is ex-\npected to produce a Gaussian damping which justifies our choice\nFig. 5. The average density (top) and vertical velocity, vz(bottom) pro-\nfiles along the simulated multithermal loop. The solid lines represent a\nGaussian damped sinusoid fit to the data following the function given in\nEq. 34. The obtained wavelength ( λ) and damping length Ldvalues are\nlisted in the plot.\nhere. The solid light orange lines in Fig. 5 represent the obtained\nfits to the data. The damping lengths obtained from the fits are\n20±0.3 Mm, and 21.2±0.3 Mm for the density and vertical ve-\nlocity respectively. These values are within 20% of the expected\nvalue of 25Mm (see Sec. 2.3, Eq. 26), and are thus a reasonable\nmatch. The deviation may be due to (1) the approximation of\nthe full integral (Eq. 15 by Eq. 19), or (2) the “small” number\nof strands (only 33) in the simulations of Krishna Prasad & Van\nDoorsselaere (2023), insu fficient to fully cover the continuous,\nGaussian DEM which is modelled in Eq. 15 due to the finite\nsample size. Because of the chosen small driver amplitude, non-\nlinear e ffects do not play a role in the damping of these waves.\nNumerical di ffusion could play a role, but we have verified that\nthe increasing the numerical resolution has no e ffect on the mea-\nsured damping lengths.\nFor a direct comparison with observations, we also forward\nmodel the data using the FoMo code (Van Doorsselaere et al.\n2016). In particular, we generate synthetic images in the 6 coro-\nnal channels of SDO /AIA, namely, 94 Å, 131 Å, 171 Å, 193 Å,\n211 Å, and 335 Å. As described in Krishna Prasad & Van Doors-\nselaere (2023), we add appropriate data noise (following Yuan\n& Nakariakov 2012) and subsequently build time-distance maps\nto study the evolution of oscillations along the loop. The oscil-\nlations are found only in three channels, the 171 Å, 193 Å, and\n211 Å. The propagation properties are exactly the same as those\nfound in Krishna Prasad & Van Doorsselaere (2023) so we are\nnot going to discuss them in detail here. Just highlighting one\ncrucial point, Krishna Prasad & Van Doorsselaere found that the\nforward modelled propagation speeds in the 211Å and 193Å fil-\nter were very close to each other, despite their di fference in tem-\nperature. Here we explain quantitatively this phenomena: as can\nbe seen in Table 1, the predicted propagation speeds Vare indeed\nvery close to each other for these filters. We propose that the\nvariation of phase speeds in di fferent observational filters (e.g.\nKing et al. 2003) may be quantitatively explained through the\nproposed formula in Eq. 30: indeed, that equation shows that the\nobserved phase speed Vis influenced by the specific filter’s σF\nandvF.\nLet us now focus on the damping properties of the slow\nwaves. Fig. 6 displays the spatial intensity profiles at a particular\nArticle number, page 6 of 9T. Van Doorsselaere et al.: Multithermal apparent damping of slow waves\nFig. 6. The spatial intensity profiles at a particular instant along the\nloop obtained from the synthetic data corresponding to the AIA 171,Å,\n193,Å, and 211,Å filters. The solid lines represent a Gaussian damped\nsinusoid fit to the data following the function given in Eq. 34. The ob-\ntained wavelength ( λ) and damping length Ldvalues are listed in the\nplot.\ninstant along the loop for the three AIA channels. The solid lines\nin the figure correspond to the fitted profiles following Eq. 34.\nWe notice that the fitted curve in AIA 171 has a significant\ndeviation beyond a distance of 40 Mm. Imposing tighter con-\nstraints on the fitting function did not improve the results much.\nHowever, the larger uncertainty obtained on the corresponding\ndamping length should have incorporated this. The resulting as-\nsociated damping lengths are 31.8 ±1.3 Mm, 36.8±0.9 Mm, and\n34.7±0.9 Mm, for the 171 Å, 193 Å, and 211 Å channels, respec-\ntively. As may be noted these values are di fferent from those ob-\ntained for the density and velocity parameters (see Fig. 5). This\nis due to the temperature response of the observing filter, which\nalso has an influence on this multithermal apparent damping, as\ndescribed in Sec. 3.1.\nIn order to quantitatively assess the e ffect of SDO /AIA fil-\nters, we fit the temperature response curves (version 9) for each\ncoronal filter with a Gaussian function and estimate their stan-\ndard width. The temperature response curves and the fitted pro-\nfiles are plotted in Fig. 7. For the curves with multiple peaks, we\nchoose the peak that is closer to log T=6.0, which is the charac-\nteristic temperature in our simulations. In each of the panels, the\ndashed line shows the full response curve, the light orange line\ndenotes the fitted segment, and the green line shows the fitted\nfunction. The obtained widths σlogTare 0.22±0.01, 0.17±0.00,\n0.13±0.00, 0.12±0.01, 0.12±0.01, and 0.31±0.01, for the 94 Å,\n131 Å, 171 Å, 193 Å, 211 Å, and 335 Å channels, respectively.\nThese values along with the respective peak locations are listed\nin Table 1. From these fitted filter curves in log T-space, we have\ncomputed, for each filter, the corresponding peak propagation\nspeed vFfrom the sound speed of the fitted peak temperature.\nThen, we have calculated σF, as the average of sound speeds be-longing to the peak temperature plus and minus the filter peak\nwidth. Thus, we have assumed that the filter is symmetric in ve-\nlocity space. The results of these calculations are listed in table 1.\nThen we have used Eqs. 30 to compute ΣandV, and we have also\nlisted the obtained values in table 1.\nFor the density, we have calculated the predicted damping\ntime with Eq. 26. For predicting the damping times observed in\nthe filters, we have used the values for ΣandVand inserted them\nin Eq. 33. The predicted damping values are listed in table 2,\nalong with the measured damping values from the simulations.\nThe predicted damping times match reasonable well with the\nmodelled damping times (with a maximum deviation of 30%).\nAs before, we think that this deviation between the numerical\ndamping lengths and the predicted damping lengths is due to\nthe finite number of strands of which the loop consists in the\nsimulation. Thus, the number of strands is insu fficient to fill\nthe entire Gaussian DEM. In essence, there are insu fficient\nMonte Carlo realisations of the strands to completely cover the\nexpected Gaussian DEM distribution.\n4. Conclusions and discussion\nIn this paper, we have considered the apparent damping of slow\nwaves (which we call “multithermal apparent damping” (MAD)\nor “V oitenko-damping”) due to a di fferent propagation speed in\ncoronal loop strands. We have considered a superposition of δ-\nfunction impulses, Gaussian pulses or driven waves. All of these\nmodels led to the multithermal apparent damping of slow waves\ndue to observational phase mixing. We should stress that the\ndamping is indeed only apparent, and that no wave energy was\nharmed dissipated during the production of this paper. This mul-\ntithermal apparent damping of the slow waves is expected to be\nstronger than damping by thermal conduction for short periods\n(less than 200s), and comparable for longer periods. We have\nfound that the case of driven slow waves leads to a predicted\nGaussian damping profile, with a predicted damping length LG\nof\nLG=¯v2\nσvω,\nwhere ¯ vis the average sound speed in the loop, σvis the spread in\nthe sound speed, and ωis the frequency. The resulting, predicted\nvalue of the damping length matches reasonably well with the\none found in the simulations of Krishna Prasad & Van Doorsse-\nlaere (2023). The predicted damping length scales linearly with\nthe period of the wave. This is compatible with the observational\nsynthesis made by Cho et al. (2016), who observed a unified pic-\nture of solar and stellar quasi-periodic pulsations with a damp-\ning time scaling linearly with the period. Moreover, this di fferent\nscaling of the multithermal apparent damping time with period\nfrom thermal conduction may explain the di fference in damp-\ning scalings in open-field or closed-field regions (e.g. Krishna\nPrasad et al. 2014, and follow-up works), but also for di fferent\ndamping regimes at di fferent heights (Gupta 2014). These dif-\nferent damping regimes could then be associated with di fferent\nlevels of multi-thermal structuring of loops /plumes and the rel-\native importance of thermal conduction damping and multither-\nmal apparent damping.\nIn the second part of the paper, we have considered the e ffect\nof a finite filter width in imaging instruments such as SDO /AIA.\nWe have found that the finite filter has as e ffect that the waves\nhave a di fferent propagation speed Vand damping length LGin\nArticle number, page 7 of 9A&A proofs: manuscript no. multiT_dampingslow_for_arxiv\nFig. 7. SDO/AIA temperature response curves for the 6 coronal channels as listed. In each of the panels, the blue dashed line represents the full\nresponse curve, the light orange solid line represents the segment fitted with a Gaussian function, and the green solid line represents the fitted\nfunction. The obtained standard widths are listed in the plot.\nTable 1. Properties of AIA filter response curves.\nChannel name AIA 94 AIA 131 AIA 171 AIA 193 AIA 211 AIA 335\nPeak temperature µlogT(logT) 6.02±0.01 5.75±0.0 5.90±0.0 6.14±0.01 6.24±0.01 5.91±0.00\nvF(km/s) 135 179 200\nPeak Width σlogT(logT) 0.22±0.01 0.17±0.0 0.13±0.0 0.12±0.01 0.12±0.01 0.31±0.01\nσF(km/s) 20.3 24.7 27.8\nΣ(km/s) 16.1 18.0 19.1\nV(km/s) 141.6 166.1 174.9\nTable 2. Gaussian damping lengths in Mm for various quantities.\nDensity Velocity AIA 171 AIA 193 AIA 211\nNumerical model 20.0 ±0.3 21.2±0.3 31.8±1.3 36.8±0.9 34.7±0.9\nPredicted damping 25.1 35.7 43.8 45.8\neach filter. These are given by\nV=σ2\nvvF+σ2\nF¯v\nσ2\nF+σ2v,LG=V2\nΣω,1\nΣ2=1\nσ2\nF+1\nσ2v\nwhere vFis the central sound speed of the filter, and σFis the\nwidth of the filter. This explains two phenomena: (1) the ob-\nserved phase speed in di fferent filters depends on the thermal\nproperties of the loop, and (2) the damping in each filter is also\ndifferent. We have once again checked these formulas against the\ndamping in forward models of the simulations of Krishna Prasad\n& Van Doorsselaere (2023). We found that our predictions matchreasonably well with the simulated values (within 30%). We sus-\npect that the deviation is mostly caused by the small number\nof strands in the simulation, which compares to our continuous\nDEM distribution that we considered in this paper.\nWe expect that these results may be used in the future to\nperform MHD seismology (Nakariakov & Verwichte 2005) of\ncoronal loops with slow waves. With the above formulas, it is\npossible to fit the loop’s DEM properties of central temperature\n(through the average sound speed ¯ v) and spread in temperature\n(through the value of the spread in sound speed σv). These DEM\nproperties of the loops are only sensitive to the loop itself in\nwhich the slow wave propagates. This is in contrast to the cur-\nArticle number, page 8 of 9T. Van Doorsselaere et al.: Multithermal apparent damping of slow waves\nrently used method of DEM inversion (e.g. Hannah & Kontar\n2012; Cheung et al. 2015; Krishna Prasad et al. 2018), which\nis very sensitive to the careful background subtraction from the\nloop’s emission. This proposed method will at least allow to re-\nmove this sensitivity, and perhaps reveal more detailed thermal\nproperties of loops.\nAlso the combination of spectral observations with imaging ob-\nservations is an interesting avenue to consider, because the spec-\ntral observations are much less impacted by the multithermal ap-\nparent damping, and the combination of this with the imaging\nobservations would allow to disentangle physical damping from\nthe multithermal apparent damping.\nTopics for future research as a follow-up to this work would\nbe (1) to consider the e ffect of a combination of multithermal ap-\nparent damping and thermal conduction in a multistranded loop\nsystem, (2) to model the e ffect of multithermal apparent damp-\ning on standing sound waves in e.g. flaring loops (Wang 2011;\nCho et al. 2016), and (3) investigation of the usage of di fferent\nlines-of-sight from di fferent spacecraft (e.g. Solar Orbiter and\nSDO) to probe the inner multithermal structure of loops using\nmultithermal apparent damping properties.\nAcknowledgements. TVD was supported by the European Research Council\n(ERC) under the European Union’s Horizon 2020 research and innovation pro-\ngramme (grant agreement No 724326), the C1 grant TRACEspace of Internal\nFunds KU Leuven, and a Senior Research Project (G088021N) of the FWO\nVlaanderen. The research benefited greatly from discussions at ISSI. TVD would\nlike to thank Dipankar Banerjee, Vaibhav Pant and Krishna Prasad for their hos-\npitality when visiting ARIES, Nainital in spring 2023. TVD would like to thank\nRoberto Soler for referring to the V oitenko et al. (2005) paper at the AGU Chap-\nman conference in Berlin 2023. VP is supported by SERB start-up research\ngrant (File no. SRG /2022/001687). AWH acknowledges the financial support\nof the Science and Technology Facilities Council (STFC) through Consolidated\nGrants ST /S000402 /1 and ST /W001195 /1 to the University of St Andrews and\nsupport from the European Research Council (ERC) Synergy grant ‘The Whole\nSun’ (810218). The computational resources and services used in this work were\nprovided by the VSC (Flemish Supercomputer Center), funded by the Research\nFoundation Flanders (FWO) and the Flemish Government – department EWI.\nReferences\nAschwanden, M. J., Nightingale, R. W., & Alexander, D. 2000, ApJ, 541, 1059\nBanerjee, D., Krishna Prasad, S., Pant, V ., et al. 2021, Space Sci. Rev., 217, 76\nBerghmans, D. & Clette, F. 1999, Sol. Phys., 186, 207\nCheung, M. C. M., Boerner, P., Schrijver, C. J., et al. 2015, ApJ, 807, 143\nCho, I.-H., Cho, K.-S., Nakariakov, V . M., Kim, S., & Kumar, P. 2016, ApJ, 830,\n110\nDe Moortel, I., Antolin, P., & Van Doorsselaere, T. 2015, Sol. Phys., 290, 399\nDe Moortel, I. & Hood, A. W. 2003, A&A, 408, 755\nDe Moortel, I. & Hood, A. W. 2004, A&A, 415, 705\nDe Moortel, I., Hood, A. W., Ireland, J., & Walsh, R. W. 2002, Sol. Phys., 209,\n89\nGeeraerts, M., Vanmechelen, P., Van Doorsselaere, T., & Soler, R. 2022, A&A,\n661, A100\nGupta, G. R. 2014, A&A, 568, A96\nHannah, I. G. & Kontar, E. P. 2012, A&A, 539, A146\nHood, A. W., Ruderman, M., Pascoe, D. J., et al. 2013, A&A, 551, A39\nJudge, P. G. 2023, ApJ, 957, 25\nKarampelas, K. & Van Doorsselaere, T. 2018, A&A, 610, L9\nKiddie, G., De Moortel, I., Del Zanna, G., McIntosh, S. W., & Whittaker, I. 2012,\nSol. Phys., 279, 427\nKing, D. B., Nakariakov, V . M., Deluca, E. E., Golub, L., & McClements, K. G.\n2003, A&A, 404, L1\nKolotkov, D. Y ., Nakariakov, V . M., & Zavershinskii, D. I. 2019, A&A, 628,\nA133\nKrishna Prasad, S., Banerjee, D., & Van Doorsselaere, T. 2014, ApJ, 789, 118\nKrishna Prasad, S., Banerjee, D., Van Doorsselaere, T., & Singh, J. 2012, A&A,\n546, A50\nKrishna Prasad, S., Jess, D. B., & Van Doorsselaere, T. 2019, Frontiers in As-\ntronomy and Space Sciences, 6, 57\nKrishna Prasad, S., Jess, D. B., Van Doorsselaere, T., et al. 2017, ApJ, 847, 5\nKrishna Prasad, S., Raes, J. O., Van Doorsselaere, T., Magyar, N., & Jess, D. B.\n2018, ApJ, 868, 149Krishna Prasad, S. & Van Doorsselaere, T. 2023, submitted to A&A\nMagyar, N. & Van Doorsselaere, T. 2016, ApJ, 823, 82\nMandal, S., Krishna Prasad, S., & Banerjee, D. 2018, ApJ, 853, 134\nMandal, S., Magyar, N., Yuan, D., Van Doorsselaere, T., & Banerjee, D. 2016,\nApJ, 820, 13\nNakariakov, V . M. & Verwichte, E. 2005, Living Rev. Solar Phys., 2, 3\nPascoe, D. J., Anfinogentov, S., Nisticò, G., Goddard, C. R., & Nakariakov, V . M.\n2017, A&A, 600, A78\nPascoe, D. J., Goddard, C. R., Nisticò, G., Anfinogentov, S., & Nakariakov, V . M.\n2016, A&A, 585, L6\nPascoe, D. J., Hood, A. W., de Moortel, I., & Wright, A. N. 2012, A&A, 539,\nA37\nPascoe, D. J., Van Doorsselaere, T., & De Moortel, I. 2022, ApJ, 929, 101\nPorth, O., Xia, C., Hendrix, T., Moschou, S. P., & Keppens, R. 2014, ApJS, 214,\n4\nShi, M., Van Doorsselaere, T., Guo, M., et al. 2021, ApJ, 908, 233\nUritsky, V . M., Davila, J. M., Viall, N. M., & Ofman, L. 2013, ApJ, 778, 26\nVan Doorsselaere, T., Antolin, P., & Karampelas, K. 2018, A&A, 620, A65\nVan Doorsselaere, T., Antolin, P., Yuan, D., Reznikova, V ., & Magyar, N. 2016,\nFrontiers in Astronomy and Space Sciences, 3\nVan Doorsselaere, T., Srivastava, A. K., Antolin, P., et al. 2020, Space Sci. Rev.,\n216, 140\nV oitenko, Y ., Andries, J., Copil, P. D., & Goossens, M. 2005, A&A, 437, L47\nWang, T. 2011, Space Sci. Rev., 158, 397\nWang, T. & Ofman, L. 2019, ApJ, 886, 2\nWang, T., Ofman, L., Sun, X., Provornikova, E., & Davila, J. M. 2015, ApJ, 811,\nL13\nWilliams, T., Walsh, R. W., Regnier, S., & Johnston, C. D. 2021, Sol. Phys., 296,\n102\nYu, D. J., Van Doorsselaere, T., & Goossens, M. 2017a, ApJ, 850, 44\nYu, D. J., Van Doorsselaere, T., & Goossens, M. 2017b, A&A, 602, A108\nYuan, D. & Nakariakov, V . M. 2012, A&A, 543, A9\nArticle number, page 9 of 9" }, { "title": "1704.06505v1.Magnetic_field_influence_on_the_early_time_dynamics_of_heavy_ion_collisions.pdf", "content": "Magnetic \feld in\ruence on the early time dynamics of heavy-ion collisions\nMoritz Greif,1,\u0003Carsten Greiner,1and Zhe Xu2, 3\n1Institut f ur Theoretische Physik, Johann Wolfgang Goethe-Universit at,\nMax-von-Laue-Str. 1, D-60438 Frankfurt am Main, Germany\n2Department of Physics, Tsinghua University, Beijing 100084, China\n3Collaborative Innovation Center of Quantum Matter, Beijing, China\n(Dated: April 24, 2017)\nIn high energy heavy-ion collisions the magnetic \feld is very strong right after the nuclei penetrate\neach other and a non-equilibrium system of quarks and gluons builds up. Even though quarks might\nnot be very abundant initially, their dynamics must necessarily be in\ruenced by the Lorentz force.\nEmploying the 3+1d partonic cascade BAMPS we show that the circular Larmor movement of the\nquarks leads to a strong positive anisotropic \row of quarks at very soft transverse momenta. We\nexplore the regions where the e\u000bect is visible, and explicitly show how collisions damp the e\u000bect.\nAs a possible application we look at photon production from the \rowing non-equilibrium medium.\nI. INTRODUCTION.\nShortly after the collision of heavy nuclei in experi-\nments at the LHC at CERN or RHIC at BNL the en-\nergy density is high enough that a so called quark-gluon\nplasma (QGP) is formed [1{4]. Due to the high velocity\nand small distances of the nucleons passing each other at\nthe moment of the collision, the magnetic \feld for non-\ncentral collisions in the center of the reaction at, e.g., top\nRHIC/LHC energies is very high. Indeed, those \felds can\nbe expected to be among the largest \feld strengths in the\nuniverse [5]. The \felds will decay very fast, even assum-\ning a conducting medium which, by Maxwells equations,\nslows down the decay of the magnetic \rux [6]. Strong\nenough magnetic \felds can generate several novel e\u000bects,\nsuch as the chiral magnetic e\u000bect, the chiral vortical ef-\nfect [7{10], the chiral separation e\u000bect [7], or a chiral\nmagnetic wave [11, 12]. In this article, we turn to a fun-\ndamental question which has not gained much attention\nin literature: will there be directly measurable e\u000bects of\nthe electromagnetic Lorentz force in heavy-ion collisions?\nEarly attempts [13] to model a hadron gas under the in-\n\ruence of magnetic \felds did not show strong e\u000bects.\nThe authors of Ref. [14] have studied the charge de-\npendent directed \row of pions and protons in a simpli\fed\nanalytic model, taking electric and magnetic \felds into\naccount. They \fnd a very small signal, owing mainly to\nthe currents induced by electric \felds generated by the\nfast decaying magnetic \felds (Faraday e\u000bect). For very\nlowpThowever, they see a strong in\ruence of the mag-\nnetic \feld itself (dubbed as \\Hall\" e\u000bect). In particular,\nthe magnetic e\u000bect becomes important for the directed\n\row at transverse momenta pT.0:25 GeV for RHIC and\nLHC. In this paper, we will also see the growing in\ruence\nof the magnetic e\u000bect at low pT.\nHydrodynamic calculations including magnetic \felds\nare rare, and still under development [15, 16]. Recently,\nit has been found that the directed \row of charm quarks\n\u0003greif@th.physik.uni-frankfurt.deis very sensitive to the magnetic and electric \feld [17].\nFurthermore, it has been proposed, that the J=\t forma-\ntion becomes anisotropic which leads, e.g., to a sizable\nelliptic \row at high transverse momenta [18].\nWe attempt an exploratory study of the early-time\nnon-equilibrium dynamics of decon\fned quarks and glu-\nons including simple parametrizations of an external\nmagnetic \feld. We \fnd that the quark momenta ro-\ntate parallel to the event plane and develop a surprisingly\nlarge momentum anisotropy at mid- and forward rapidity\nfor very low transverse momenta. This is roughly rem-\niniscent to the Hall e\u000bect. The particle velocity stems\nmainly from the large boosts in beam direction (longitu-\ndinal expansion), whereas the magnetic \feld comes from\nthe charged spectator nucleons.\nWe see furthermore that the spectra are also slightly\nenhanced at early times and we show explicitly how col-\nlisions damp both the e\u000bect of the \row and the spectra.\nThis paper is organized as follows. In Sec. II we\nexplain the model setup, including the magnetic \feld\nparametrizations. In Sec. III we show the collisionless\nresult which is purely due to the Larmor movement of\nquarks in the magnetic \feld, before we turn in Sec. IV to\nthe result with collisions and conclude in Sec. V.\nFigure 1. Geometry of our model. The magnetic \feld is\nconstant and homogeneous.arXiv:1704.06505v1 [hep-ph] 21 Apr 20172\nII. SIMPLE MODEL WITHOUT COLLISIONS\nWe investigate how strong the e\u000bect of the Lorentz\nforce alone can be on the particle distributions. To this\nend, we outline a simple model for the heavy-ion collision,\nneglecting collisions in a \frst step. We consider the colli-\nsion of two heavy nuclei along the z-axis. For simplicity,\nwe assume the magnetic \feld ~Bto be constant and homo-\ngeneous, pointing in y-direction, ~B\u0011By~ ey. The situa-\ntion is depicted in Fig. 1. Here we neglect electrodynami-\ncal induction e\u000bects. We assume that all events are sym-\nmetric and the impact parameter points in x-direction.\nIn this geometry elliptic \row can be seen as an average\n(h\u0001i) momentum asymmetry\n(p2\nx\u0000p2\ny)=p2\nT\u000b\n\u0011v2, where\npT=q\np2x+p2y.\nA. Initial state and formation time\nIn this simple model setup we do not consider space-\ndependent e\u000bects, thus we sample only four-momenta of\nthe particles. All particles are assumed to be massless.\nThepTdistribution is sampled according to a power law,\ndN=dpT= \nn\u00001\np1\u0000n\nT;min!\np\u0000n\nT; n = 2;3;4: (1)\nWe choose a minimal value pT;min= 0:01 GeV. For all\nthe following results we assume a constant distribution\nin rapidity, y= 1=2 log(E+pz)=(E\u0000pz),\ndN=dy = const:; pz=pTsinhy: (2)\nWe \fnd that the results are not dependent on the rapid-\nity window in which we initialize the particles, as long\nas it is larger than the observed rapidity bins. For most\nstudies,\u00003< y < 3 is su\u000ecient. It is possible to use\na formation time \u0001 tf= cosh(y)=pTduring which par-\nticles are still o\u000b-shell and do not interact, but propa-\ngate freely. This formation time has been used earlier\nin transport approaches using the Minijet model for the\ninitial condition [19{21]. We can assume that the mag-\nnetic \feld will also not in\ruence the partons within their\nformation time. However, as quarks carry their electric\ncharge even o\u000b-shell, their classical interaction with mag-\nnetic \felds is arguable, and the formation time could be\nirrelevant. As this point is conceptionally uncertain, we\nshow results for both options, assuming the particle-\feld\ninteraction to be switched on immediately (no formation\ntime), or, only after \u0001 tf, respectively. In this simpli\fed\ncollisionless scenario we only initialize quarks, carrying\nthe electric charge q=e=3 orq= 2e=3, respectively.\nThe exact quark and gluon content in the early phase of\nheavy-ion collisions is under debate. It is clear, that the\nmore gluon dominated the system is, the less pronounced\nsuch electromagnetic e\u000bects will be.B. Magnetic \feld parametrizations\nThe external magnetic \feld present at t= 0+after\nthe collision is still subject to active research, and de-\npends strongly on the geometric modeling of the nuclei\nas well as the electric conductivity and also possible non-\nequilibrium e\u000bects. Common to all the results in litera-\nture is the dominant By-component, perpendicular to the\nevent plane, which is about an order of magnitude larger\nthan theBx-component, the Bz-component is nearly ab-\nsent. The authors of Ref. [22] look explicitly at \ruc-\ntuations of the direction of the magnetic \feld and \fnd\nthat for middle central collisions the \feld \ructuates less\naround the direction perpendicular to the event plane\nthan for near central or very peripheral collisions. For\nthe qualitative understanding of the dynamical e\u000bects to\nthe quark momenta, we adopt several simpli\fed scenar-\nios for the \feld strength By, and setBx=Bz= 0. In\nRef. [13] it was found that the spatial dependence over\nthe overlap region is mild, so that we restrict ourselves\nhere to a homogeneous \feld in space, parametrized as\nparam 1:eBy(t) = 4m2\n\u0019\u0002(0:3 fm=c\u0000t)\nparam 2:eBy(t) =eBt=0\ny(1 +t2=t2\nc)\u00003=2withtc=\n0:065 fm=c.\neBy [mπ2]\nt [fm/c]param 1\nparam 2\n024689 \n 0 0.1 0.2 0.3 0.4 0.5\nFigure 2. The two simple parametrizations of the homoge-\nneous magnetic \feld. Param 2 follows Ref. [23].\nParam 2 is the parametrization of the results of Ref. [23]\nas given in Ref. [5]. We use it with parameters\ncorresponding to RHIC collisions (Au+Au,psNN =\n200 GeV) at typical impact parameters of \u00188 fm\ncorresponding to 20 \u000040% centrality (see also, e.g.,\nRefs. [24, 25] for typical \feld strengths). In the very early\nstage, the medium is assumed to be gluon dominated,\nsuch that the electric conductivity can be neglected [5]\n(being roughly proportional to the sum of the electric\ncharges squared, weighted by the densities of the charge-\ncarrying species [26, 27]). The authors of Ref. [5, 23]\napproximate the total magnetic \feld thus by the exter-\nnal component produced by the charged nucleons passing3\neach other. The full solution of the Maxwell- and Boltz-\nmann equation will slow down the decay of the magnetic\n\feld, but so far, only little is known about the precise\nevolution. Parametrization 1 is an optimistic imitation\nof a strongly conducting medium, which would keep the\nmagnetic \feld present for some time. We have tried even\nhigher or longer \feld parametrizations, but for simplicity\nwe restrict ourselves to an optimistic, and a realistic one.\nC. Larmor movement\nThe magnetic \feld changes the direction of velocity of\nthe particles by the Lorentz force, ~FL=q~ v\u0002~B. In our\ngeometry, particles will move in a circle around the y-\ndirection, clockwise or anticlockwise depending on their\nchargeq. Thereby, any momentum in z-direction will\nincrease or decrease the momentum in x-direction, px!\npx+ \u0001px.\nTo analytically estimate the e\u000bect of increasing px\ncomponents, we note that by symmetry hpxi= 0;hpyi=\n0. We consider two particles with opposite pxmomentum\ncomponents, px;1=\u0000px;2as representer of the particle\nensemble. Their pymomenta are equal, and chosen in\na way, that the initial momentum asymmetry v2takes a\ngiven value. The change px!px+ \u0001pxon thev2of the\nwhole particle ensemble can then, in a simpli\fed fashion,\nbe estimated by\nv2(\u0001px)\n=1\n2 \n(px+ \u0001px)2\u0000p2\ny\n(px+ \u0001px)2+p2y+(\u0000px+ \u0001px)2\u0000p2\ny\n(\u0000px+ \u0001px)2+p2y!\n:(3)\nIn Fig. 3 we show this momentum asymmetry for three\nchoices of the initial v2, positive, zero and negative.\nClearly, for zero initial v2, the increase of \u0001 pxmust be\nlarger than pTin order to enhance the asymmetry. All\ncases show a minimum in v2for \u0001px< pT, which can\nbe strongly negative. The radius of de\rection due to the\nmagnetic \feld is\nrLarmor =p\np2x+p2z\nqBy; (4)\nand the angle of the circular movement of time tis\n\u000bLarmor =t=rLarmor . The value of \u0001 pxdepends on the\nmomentapxandpzof each particle, for a given magnetic\n\feld times its duration, Byt.\nIn this study, we do not include electromagnetic e\u000bects\nother than the Larmor movement. The reason is outlined\nin the following. The Faraday e\u000bect due to the time\ndependent magnetic \rux '=ByAthrough surface A\ngenerates an electric \feld,\n\u0000@'\n@t=I\nd~ r\u0001~E: (5)\nThis electric \feld accelerates charges in the opposite di-\nrection than the Lorentz force ~FL=qv\u0002~B. Assuming\n-1-0.5 0 0.5 1\n 0 1 2 3 4 5v2(∆px)\n∆px /pTpx=py=pT/√ 2\npx=2py\npx=py/2Figure 3. After the increase of px, the momentum asymme-\ntryv2is in all cases positive for \u0001 px> pT. The result is\nsymmetric in \u0001 px.\nfor a moment, that ~FL\u00110, the electric current due to the\nforceq~Ewill generate a magnetic \feld component Bind\ncounter balancing the decay of the \feld, ~Bind\u0018@~B=@t ,\ndepending on the electric conductivity. On top of these\ne\u000bects, the electric \felds generated by the spectators, al-\nbeit small in magnitude, has also an x-component [28],\nwhich is positive Espec\nx>0 forx > 0 and negative\nEspec\nx<0 forx < 0. All these 3 e\u000bects can cancel\nor enhance each other, and depend crucially on the as-\nsumed electric conductivity and parametrization of the\nbare spectator induced \felds. Furthermore, the calcu-\nlation of the magnetic \rux as well as the electric \felds\nwould require a full space-time dependent (propagating)\nsolution of the electromagnetic \felds. This is why we re-\nstrict ourselves to show what maximum e\u000bect on the par-\nticle dynamics is expected from the magnetic \feld only.\nIII. RESULTS OF THE COLLISIONLESS\nMODEL\nFirst we show how pT-spectra of quarks are in\ruenced\nin the early time by a magnetic \feld in Fig. 4. Here,\nfor parametrization 1 of the magnetic \feld, the spectrum\nis enhanced for 0 :02< pT=GeV<0:1 due to the Lar-\nmor turn of pz-momenta. In Fig. 3 we explained that\nthe \fnal momentum asymmetry depends strongly on the\nadditional \u0001 px. We explore which momentum space re-\ngion (regions in rapidity) is necessary to gain su\u000eciently\nlarge values of \u0001 pxfor thev2to change visibly. Here\nwe di\u000berentiate between initial quantities, and those af-\nter the circular Larmor-movement has been applied to\nthe particle. For this purpose we show in Fig. 5 the \f-\nnalv2(after the Larmor movement had been applied for\na timet) as function of initial rapidity yinitial . We split\nthis up in a soft region, for \fnal pT<0:3 GeV, where\nthe averaged v2reaches large values, and the region of\n\fnalpT>0:3 GeV, where the v2is consistent with zero.4\n1021041061081010\n 0.01 0.1 1dNquarks/(2πpTdpTdy) [GeV-2]\npT [GeV]dashed: initial|y|<0.5\neBy=4 mπ2\nt=0.3 fm/c\nno collisions\narbitrary normalisationn=2\nn=3\nn=4\nFigure 4. For three di\u000berent initial pT-distributions (power-\nlaw exponent n) we show how the spectra change after\n0:3 fm=c under the in\ruence of a magnetic \feld. Here\nwe use an arbitrary number of particles and magnetic \feld\nparametrization 1.\nThis ultrasoft pTrange can already be expected from the\nspectra, Fig. 4. Note that the saturation in Fig. 5 is due\nto the cuts in \fnalpT, which means, that, in the curve\nfor \fnalpT<0:3 GeV, the larger sinh( y), the smaller\nthe values of pTwhich contribute.\n-0.4-0.2 0 0.2 0.4 0.6 0.8 1\n-3 -2 -1 0 1 2 3final v2\nyinitialeB=4 mπ2\nt=0.3 fm/cinitial state: \nn=2, |y| < 3all final pTfinal pT > 0.3 GeV\nfinal pT < 0.3 GeV\nFigure 5. Final average v2per particle for three di\u000berent pT\nranges as function of initial rapidity yinitial of the particle.\nHere we use an initial state with power law exponent n= 2\nand magnetic \feld parametrization 1. The result for n= 3;4\nlooks very similar, only the maximal value of the \fnal v2\nincreases by up to 25%.\nClearly, momentum rapidities y > 1 are responsible\nfor \u0001px&pTand the average momentum asymmetries\nlarger than zero. The three initial pT-distributions show\nsimilar behavior, only the maximal value of v2increases\nwith increasing n. Finally we turn to the di\u000berential v2.\nUsing magnetic \feld parametrization 1, we show in Fig. 6\nthe resulting v2(pT) without the use of the formation\ntime, for mid- and forward rapidity and all three initial\nstate parametrizations. The v2can be (temporary) up\n-0.2 0 0.2 0.4 0.6 0.8 1\n 0 0.1 0.2 0.3 0.4 0.5v2(pT)\npT [GeV]eB=4 mπ2\nt=0.3 fm/c\nno formation timesn=2: |y|<0.5\n1<|y|<3\nn=3: |y|<0.5\n1<|y|<3\nn=4: |y|<0.5\n1<|y|<3Figure 6. The pT-di\u000berential v2in a collisionless toy model for\ninitial power law spectra with exponent n= 2;3;4. Here we\ndo not assign formation times. We show results for forward-\nand midrapidity with magnetic \feld parametrization 1.\n-0.2 0 0.2 0.4 0.6 0.8 1\n 0 1 2 3 4 5v2(pT)\npT [GeV]eB=4 mπ2\nt=0.3 fm/c\nformation timesn=2: |y|<0.5\n1<|y|<3\nn=3: |y|<0.5\n1<|y|<3\nFigure 7. Same as Fig. 6, but here formation times had been\nassigned. For the shown snapshot at t= 0:3 fm=c, the mini-\nmum occurring momentum at midrapidity is pT\u00190:66 GeV.\nto 80 %. It is larger for forward rapidity. In Fig. 7 we\nshow the result when the formation time was taken into\naccount. This results in deleting all relevant interactions\namong the \feld and the particles, and the v2remains\nzero.\nIV. THE EFFECT OF COLLISIONS\nNext we want to consider the e\u000bect of particle colli-\nsions. To this end we employ the 3+1-dimensional trans-\nport approach BAMPS (Boltzmann Approach to Multi-\nParton Scatterings), which solves the relativistic Boltz-\nmann equation by Monte-Carlo techniques [29, 30] for\nmassless on-shell quarks and gluons1. The Boltzmann\n1corresponding to an ideal equation of state5\n101102103104105106\n 0.01 0.1dNquarks/(2πpTdpTdy) [GeV-2]\npT [GeV]|y|<0.5eBy=4 mπ2\narbitrary normalisationinitial, n=2\nno collisions, t=0.3 fm/c\ncollisions, t=0.3 fm/c\nno field, collisions, t=0.3 fm/c\nFigure 8. Transverse momentum spectra of quarks under the\nin\ruence of a magnetic \feld in a free streaming and a col-\nlisional medium. We use an arbitrary number of particles\nand magnetic \feld parametrization 1. Particles collide with\nconstant isotropic cross sections, \u001btot= 10 mb.\nequation is ideally suited to study thermalization and\nisotropization processes [29, 31] and the electromagnetic\n\felds enter by an external force term. With the phase-\nspace distribution function fi(x;k)\u0011fi\nkfor particle\nspeciesi, the Boltzmann equation reads\nk\u0016@\n@x\u0016fi\nk+k\u0017qiF\u0016\u0017@\n@k\u0016fi\nk=NspeciesX\nj=1Cij(x\u0016;k\u0016);(6)\nwhereCijis the collision term, and qithe electric charge.\nThe \feld strength tensor F\u0016\u0017=E\u0016u\u0017\u0000E\u0017u\u0016\u0000B\u0016\u0017,\nwithB\u00160=B0\u0017= 0,Bij=\u0000\u000fijkBkandE\u0016= (0;~E),\nintroduces the electromagnetic forces to the charged par-\nticles [32]. For the BAMPS simulations we include 3\n\ravors of light quarks, antiquarks and gluons. Space is\ndiscretized in small cells with volume \u0001 Vand particles\nscatter and propagate within timesteps \u0001 t. In each cell,\nthe probability for binary/inelastic scattering is\nP22=23=\u001btot;22=23(s)\nNtestvrel\u0001t\n\u0001V; (7)\nwhere\u001btot(s) is the (in general Mandelstam sdependent)\ntotal cross section and vrelthe relative velocity. The in-\nelastic backreaction works similar. In the simplest case,\nwe employ constant and isotropic cross sections, how-\never, BAMPS features binary and radiative perturbative\nQuantum Chromo Dynamic (pQCD) cross sections (see,\ne.g., Ref. [33, 34]) and running coupling \u000bs(Q2), which\nis evaluated at the momentum transfer Q2of the respec-\ntive scattering process [35]. For the purpose of this study\nhere, there is no qualitative di\u000berence when employing\npQCD cross sections, so we restrict ourselves to constant\nand isotropic scattering. As a new feature, we include\nthe electromagnetic force, which within the Monte-Carlo\nframework reduces to the additional change of the parti-\n-0.2-0.1 0 0.1 0.2 0.3 0.4 0.5 0.6\n0.01 0.1 0.2 0.3 0.4 0.5quark v2\npT [GeV]σtot=10 mb, |y|<0.5no field, t=0.3 fm/c\nno field, t=2 fm/c \nparam 1, t=0.3 fm/c\nparam 1, t=2 fm/c \nPHENIXFigure 9. The pT-di\u000berential v2from BAMPS for the ultrasoft\npTrange with and without magnetic \feld. The initial state\nis equivalent to Sec. II A, with exponent n= 2. The initial\ngeometry is equivalent to an impact parameter of b= 8:5fm.\nWe ignore formation times here. For a rough comparison we\nshow data from PHENIX [37] (unidenti\fed charged hadrons,psNN= 200 GeV, 20% \u000060% centrality,j\u0011j<0:35).\ncle momenta (for every computational timestep) by\nd~ki= \u0001tFLorenz = \u0001tqi\u0010\n~E+~ v\u0002~B\u0011\n: (8)\nAs mentioned before, we set ~E= 0. It is clear that\nthe propagation of \felds (by retarded Li\u0013 enard-Wiechert\npotentials) generated by moving quarks would re\fne\nthe picture, this will be done in a forthcoming publi-\ncation. Nevertheless, electric currents appear by default\nin BAMPS, and the electric conductivity of the matter\nis built in naturally [26]. We use the same initial state\nin momentum space in BAMPS as in the simple model\nfrom Sec. II, and use smooth a Glauber Monte Carlo dis-\ntribution of particle positions. Here we use an impact pa-\nrameter ofb= 8:5 fm. The particle numbers are roughly\nequal to simulations performed in earlier studies using\nBAMPS for Au+Au collisions atpsNN= 200 GeV [34{\n36]. Flavors for gluons and quarks ( Nf= 3) are sampled\nrandomly with probabilities Pg= 16=52;Pq= 36=52.\nWe note that this setup is certainly rough, but it should\nsu\u000ece for our purpose of an optimistic upper estimate\nof the \\Hall current\" to the anisotropic \row. We\nshow in Fig. 8, how the spectra are a\u000bected by colli-\nsions. Here we see, that in the viscous case (including\ncollisions,\u001btot= 10 mb), the spectra in\ruenced by the\nmagnetic \feld for momenta pT&1 GeV are very close to\nthe \feld free case. Without \felds, the medium thermal-\nizes at timescales of 0 :5\u00181 fm=c (see Ref.[29]). Again,\nin the region of 0 :01.pT=GeV.0:1 the spectra are\nenhanced compared to the \feld free spectra. The green\ndashed line shows the collisionless result, which is close\nto the initial power law for larger momenta, pT&1 GeV,\nand enhanced in the soft region.\nIn Fig. 9 we show the pT-di\u000berential v2of light quarks\nfrom BAMPS, and turn the magnetic \feld on and o\u000b.6\n-0.2-0.1 0 0.1 0.2 0.3 0.4 0.5 0.6\n0.01 0.1 0.2 0.3 0.4 0.5quark v2\npT [GeV]σtot=10 mb, |y|<0.5param 1, t=0.3 fm/c\nparam 1, t=2 fm/c \nparam 2, t=0.3 fm/c\nparam 2, t=2 fm/c \nPHENIX\nFigure 10. Same as Fig. 9, but here we compare the magnetic\n\feld parametrizations 1 and 2.\nClearly, the \feld causes a strong momentum anisotropy\nbelowpT\u00180:1 GeV, but has hardly any e\u000bect above\nthis softpT-range. For comparison we plot the softest\npoints of experimentally measured unidenti\fed charged\nparticle \row from PHENIX [37]. Unfortunately they are\nstill measured at such high transverse momenta, that a\ndetection of the presented magnetic \feld e\u000bect is unlikely\nat present.\nWe see in comparison with Fig. 6, which shows the col-\nlisionless result, that the collisions damp the v2(about\n20% lower v2at aroundpT= 40\u000060 MeV). Here we\nshow results with parametrization 1, which switches o\u000b\nthe \feld at t= 0:3 fm=c. After that time, the collisions\nisotropize this initial \row, such that after t= 2 fm=c it\nis around 0:34 and the maximum is pushed to even lower\npT. In Fig. 10 we compare the e\u000bect of the two mag-\nnetic \feld parametrizations. Parametrization 2, probably\nmore realistic, has a weaker e\u000bect than parametrization\n1. The maximal \row is still about 30%, but, more impor-\ntantly, it is shifted to much lower transverse momenta.\nWe need to recall at this point, that all strong elliptic \row\nsignals appear only, when the formation time of quarks\nis neglected for the interaction among the \feld and the\nparticles. This issue must be further addressed in future.\nWe note, that we ignore for simplicity hadronization and\nthe subsequent hadron gas evolution here, nevertheless,\nelliptic \row on the order of 30% is likely to survive to\nsome degree. This remains subject for future work.\nPhotons are an ideal probe to test e\u000bects throughout\nthe spacetime evolution of the medium, such as magnetic\n\feld induced \row, as they leave the \freball nearly undis-\nturbed, once produced. In an earlier study [38] we have\nimplemented photon production in BAMPS, consistent\nwith leading order rates. Here, we make use of the 2 $2\nphoton production method from Ref. [38]. At very low\npT(where all interesting magnetic e\u000bects happen), the\nmicroscopic photon production processes for collisions of\ntwo lowpTpartons will have typical Mandelstam vari-\nables at magnitudes, where the concept of perturbative\n-0.1 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8\n0.01 0.1 0.2v2\npT [GeV]|y|<0.5, Param 1\nn=3, σtot=10 mbQuarks\nPhotonsFigure 11. Photon v2as function of pTcompared to quark v2\nfor magnetic \feld parametrization 1. Here we use an initial\nstate with power law exponent n= 3. Photons and quarks\nare evaluated at midrapidity at t= 2 fm=c.\nQCD methods is questionable ( s.\u00032\nQCD). Neverthe-\nless, we allow photons to be produced as we are mainly\ninterested in the non-equilibrium e\u000bect of photon pro-\nduction from a \rowing quark medium. In Fig. 9 we show\nthev2(pT) of produced photons with that of the quarks\nat timet= 2 fm=c. Photons are produced during the\nwhole collision, and observables are thus always space-\ntime averaged, weighted by the production yield. In the\nbeginning of the collision, the medium is dense and the\nenergydensity is high, so many photons are produced, but\n(in our simpli\fed initial state) the \row is zero. Later,\nthe photons inherit some of the \row, but their rate de-\ncreases steadily. This is the reason, why the observed\nphoton \row is smaller than the pure quark \row. Below\npT.0:1 GeV the photon \row is enhanced. It will be\nchallenging to measure such an e\u000bect, considering that\nrecent measurements of (direct) photon \row [39, 40] ex-\ntend down to pT= 0:4 GeV (PHENIX)/ pT= 1 GeV\n(ALICE).\nV. SUMMARY\nWe have shown how the Lorentz force in heavy-ion\ncollisions can a\u000bect observables. To this end, we have\nassumed two simple parametrizations for an external,\nhomogeneous magnetic \feld, which is produced by the\nfast spectator nucleons. We investigate a free streaming,\nand a viscous medium (with collisions), employing the\npartonic transport simulation BAMPS. We use a sim-\nple boost invariant initial state, assuming a power-law\nin the transverse momentum distribution, and a periph-\neral Monte-Carlo Glauber geometry of the overlap zone.\nWe have shown that the magnetic \feld will generate a\nstrong elliptic \row only at very small transverse momenta\ndue to the Larmor movement of the charged particles.\nIn this very soft region, also the transverse momentum7\nspectra are enhanced. We show that collisions will wash\nout both the enhancement of the spectra and the elliptic\n\row. However, the \row is still quite large, such that it\ncould be measured, if experiments had access to ultrasoft\ntransverse momenta. Assuming an initial formation time\nof the particles, within which the magnetic \feld can not\nact, all strong e\u000bects are deleted. The interaction of clas-\nsical \felds and unformed particles is however a di\u000ecult\ntheoretical problem and must be clari\fed further. This\nstudy can be extended in several ways. Apart from other\nobservables, electric \felds, stemming from the Faraday\nlaw, might also play a role. A logical next step would be\na realistic space dependence of the external \felds, and,\nin the long run, a full spacetime evolution of retarded\n\felds including induction e\u000bects (similar to Ref. [13]).\nWe emphasize, that the present study should only give\nan order-of-magnitude estimate of what can be expected\nfrom the magnetic Lorentz force (\\Hall e\u000bect\") for light\nquarks. Apart of the experimental challenge, there might\nbe other consequences. Especially \fnal spectra of tomo-graphic probes like photons or dileptons will inherit in-\nformation of this strongly \rowing but ultrasoft region.\nTo get an idea of this e\u000bect we have shown that photons\ninherit a fraction of the elliptic \row from the quarks at\nnearly the same ultralow transverse momenta.\nACKNOWLEDGEMENTS\nM.G. is grateful to Tsinghua university in Beijing for\ntheir hospitality and acknowledges the support from the\n\\Helmhotz Graduate School for Heavy Ion research\".\nThe authors are grateful to the Center for Scienti\fc Com-\nputing (CSC) Frankfurt for the computing resources.\nThis work was supported by the Helmholtz Interna-\ntional Center for FAIR within the framework of the\nLOEWE program launched by the State of Hesse. XZ\nwas supported by the MOST, the NSFC under Grants\nNo. 2014CB845400, No. 11275103, No. 11335005, No.\n11575092.\n[1] I. Arsene et al. (BRAHMS), Nucl. Phys. A757 , 1 (2005),\narXiv:nucl-ex/0410020.\n[2] K. Adcox et al. (PHENIX), Nucl. Phys. A757 , 184\n(2005), arXiv:nucl-ex/0410003.\n[3] B. B. Back et al. , Nucl. Phys. A757 , 28 (2005),\narXiv:nucl-ex/0410022.\n[4] J. Adams et al. (STAR), Nucl. Phys. A757 , 102 (2005),\narXiv:nucl-ex/0501009.\n[5] X.-G. Huang, Rept. Prog. Phys. 79, 076302 (2016),\narXiv:1509.04073 [nucl-th].\n[6] S. Pu, V. Roy, L. Rezzolla, and D. H. Rischke, Phys.\nRev. D 93, 074022 (2016), arXiv:1602.04953 [nucl-th].\n[7] D. E. Kharzeev, L. D. McLerran, and H. J. Warringa,\nNucl. Phys. A803 , 227 (2008), arXiv:0711.0950 [hep-ph].\n[8] D. E. Kharzeev and D. T. Son, Phys. Rev. Lett. 106,\n062301 (2011), arXiv:1010.0038 [hep-ph].\n[9] D. E. Kharzeev, Annals Phys. 325, 205 (2010),\narXiv:0911.3715 [hep-ph].\n[10] K. Fukushima, D. E. Kharzeev, and H. J. Warringa,\nPhys. Rev. D78, 074033 (2008), arXiv:0808.3382 [hep-\nph].\n[11] D. E. Kharzeev and H.-U. Yee, Phys. Rev. D83, 085007\n(2011), arXiv:1012.6026 [hep-th].\n[12] Y. Burnier, D. E. Kharzeev, J. Liao, and H.-U. Yee,\nPhys. Rev. Lett. 107, 052303 (2011), arXiv:1103.1307\n[hep-ph].\n[13] V. Voronyuk, V. D. Toneev, W. Cassing, E. L.\nBratkovskaya, V. P. Konchakovski, and S. A. Voloshin,\nPhys. Rev. C 83, 054911 (2011), arXiv:1103.4239 [nucl-\nth].\n[14] U. Gursoy, D. Kharzeev, and K. Rajagopal, Phys. Rev.\nC89, 054905 (2014), arXiv:1401.3805 [hep-ph].\n[15] V. Roy, S. Pu, L. Rezzolla, and D. Rischke, Phys. Lett.\nB750, 45 (2015), arXiv:1506.06620 [nucl-th].\n[16] G. Inghirami, L. Del Zanna, A. Beraudo, M. H. Moghad-\ndam, F. Becattini, and M. Bleicher, Eur. Phys. J. C 76,\n659 (2016), arXiv:1609.03042 [hep-ph].[17] S. K. Das, S. Plumari, S. Chatterjee, J. Alam, F. Scar-\ndina, and V. Greco, (2016), arXiv:1608.02231 [nucl-th].\n[18] X. Guo, S. Shi, N. Xu, Z. Xu, and P. Zhuang, Phys.\nLett. B751 , 215 (2015), arXiv:1502.04407 [hep-ph].\n[19] J. Upho\u000b, O. Fochler, Z. Xu, and C. Greiner, Phys. Rev.\nC84, 024908 (2011), arXiv:1104.2295 [hep-ph].\n[20] K. J. Eskola, K. Kajantie, and J. Lindfors, Nucl. Phys.\nB323 , 37 (1989).\n[21] K. Kajantie, P. V. Landsho\u000b, and J. Lindfors, Phys.\nRev. Lett. 59, 2527 (1987).\n[22] J. Bloczynski, X.-G. Huang, X. Zhang, and J. Liao,\nPhys. Lett. B718 , 1529 (2013), arXiv:1209.6594 [nucl-\nth].\n[23] W.-T. Deng and X.-G. Huang, Phys. Rev. C85, 044907\n(2012), arXiv:1201.5108 [nucl-th].\n[24] A. Bzdak and V. Skokov, Phys. Lett. B710 , 171 (2012),\narXiv:1111.1949 [hep-ph].\n[25] K. Tuchin, Phys. Rev. C88, 024911 (2013),\narXiv:1305.5806 [hep-ph].\n[26] M. Greif, I. Bouras, C. Greiner, and Z. Xu, Phys. Rev.\nD90, 094014 (2014), arXiv:1408.7049.\n[27] M. Greif, C. Greiner, and G. S. Denicol, Phys. Rev. D93,\n096012 (2016), arXiv:1602.05085 [nucl-th].\n[28] H. Li, X.-l. Sheng, and Q. Wang, Phys. Rev. C94, 044903\n(2016), arXiv:1602.02223 [nucl-th].\n[29] Z. Xu and C. Greiner, Physical Review C 71, 064901\n(2005), arXiv:0406278v2 [arXiv:hep-ph].\n[30] Z. Xu and C. Greiner, Phys. Rev. C76, 024911 (2007),\narXiv:hep-ph/0703233 [hep-ph].\n[31] P. B. Arnold, Int. J. Mod. Phys. E16, 2555 (2007),\narXiv:0708.0812 [hep-ph].\n[32] C. Cercignani and G. M. Kremer, The Relativistic\nBoltzmann Equation: Theory and Applications , Vol. 22\n(Springer Basel AG, Boston, 2002).\n[33] O. Fochler, J. Upho\u000b, Z. Xu, and C. Greiner, Phys. Rev.\nD88, 014018 (2013), arXiv:1302.5250 [hep-ph].8\n[34] J. Upho\u000b, F. Senzel, O. Fochler, C. Wesp, Z. Xu,\nand C. Greiner, Phys. Rev. Lett. 114, 112301 (2015),\narXiv:1401.1364 [hep-ph].\n[35] J. Upho\u000b, O. Fochler, Z. Xu, and C. Greiner, Phys. Lett.\nB717 , 430 (2012), arXiv:1205.4945 [hep-ph].\n[36] J. Upho\u000b, O. Fochler, Z. Xu, and C. Greiner, J. Phys.\nG42, 115106 (2015), arXiv:1408.2964 [hep-ph].[37] S. Afanasiev et al. (PHENIX), Phys. Rev. C80, 024909\n(2009), arXiv:0905.1070 [nucl-ex].\n[38] M. Greif, F. Senzel, H. Kremer, K. Zhou, C. Greiner,\nand Z. Xu, (2016), arXiv:1612.05811 [hep-ph].\n[39] A. Adare et al. (PHENIX), Phys. Rev. C94, 064901\n(2016), arXiv:1509.07758 [nucl-ex].\n[40] D. Lohner (ALICE), J. Phys. Conf. Ser. 446, 012028\n(2013), arXiv:1212.3995 [hep-ex]." }, { "title": "1009.2727v2.Modeling_laser_wakefield_accelerators_in_a_Lorentz_boosted_frame.pdf", "content": "Modeling laser wakefield accelerators in a Lorentz boosted frame.\nJ.-L. Vaya, C. G. R. Geddesa, E. Cormier-Michelb, D. P. Grotec\naLBNL, Berkeley, CA, USA\nbTech-X Corp., Boulder, CO, USA\ncLLNL, Livermore, CA, USA\nAbstract\nModeling of laser-plasma wakefield accelerators in an optimal frame of reference [1] is shown to produce orders of\nmagnitude speed-up of calculations from first principles. Obtaining these speedups requires mitigation of a high-\nfrequency instability that otherwise limits e \u000bectiveness in addition to solutions for handling data input and output\nin a relativistically boosted frame of reference. The observed high-frequency instability is mitigated using methods\nincluding an electromagnetic solver with tunable coe \u000ecients, its extension to accomodate Perfectly Matched Layers\nand Friedman’s damping algorithms, as well as an e \u000ecient large bandwidth digital filter. It is shown that choosing\nthe frame of the wake as the frame of reference allows for higher levels of filtering and damping than is possible in\nother frames for the same accuracy. Detailed testing also revealed serendipitously the existence of a singular time step\nat which the instability level is minimized, independently of numerical dispersion, thus indicating that the observed\ninstability may not be due primarily to Numerical Cerenkov as has been conjectured. The techniques developed\nfor Cerenkov mitigation prove nonetheless to be very e \u000ecient at controlling the instability. Using these techniques,\nagreement at the percentage level is demonstrated between simulations using di \u000berent frames of reference, with\nspeedups reaching two orders of magnitude for a 0.1 GeV class stages. The method then allows direct and e \u000ecient\nfull-scale modeling of deeply depleted laser-plasma stages of 10 GeV-1 TeV for the first time, verifying the scaling of\nplasma accelerators to very high energies. Over 4, 5 and 6 orders of magnitude speedup is achieved for the modeling\nof 10 GeV , 100 GeV and 1 TeV class stages, respectively.\nKeywords: laser wakefield acceleration, particle-in-cell, plasma simulation, special relativity, frame of reference,\nboosted frame\nOctober 30, 2018arXiv:1009.2727v2 [physics.acc-ph] 15 Sep 2010Contents\n1 Introduction 2\n2 Theoretical speedup dependency with the frame boost 4\n2.1 Estimated speedup for 0.1-100 GeV stages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6\n3 Input and output to and from a boosted frame simulation 6\n3.1 Input . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7\n3.1.1 Particles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7\n3.1.2 Laser . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8\n3.2 Output . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9\n4 High frequency instability and Numerical Cerenkov 9\n4.1 Wideband lowpass digital filtering . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10\n4.2 Tunable solver . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13\n4.2.1 Numerical dispersion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14\n4.2.2 Current deposition and Gauss’ Law . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17\n4.3 Friedman adjustable damping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17\n5 Application to the modeling of laser wakefield acceleration 19\n5.1 Scaled 10 GeV stages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19\n5.1.1 Using standard numerical techniques . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19\n5.1.2 E \u000bect of filtering, solver with adjustable dispersion and damping . . . . . . . . . . . . . . . . 26\n5.2 Full scale 10 GeV class stages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27\n5.2.1 Simulations in 2-1 /2D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28\n5.2.2 Simulations in 3D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28\n5.3 E \u000bects of numerical parameters on the observed instability . . . . . . . . . . . . . . . . . . . . . . . 31\n5.3.1 E \u000bects of spatial resolution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31\n5.3.2 E \u000bects of time step . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31\n5.3.3 E \u000bects of field gathering procedure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32\n5.4 Full scale 100 GeV - 1 TeV class stages . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37\n6 Conclusion and outlook 38\n7 Appendix I: One dimensional analysis of the CK solver 38\n8 Appendix II: Perfectly Matched Layer 39\n9 Acknowledgments 41\n1. Introduction\nLaser driven plasma waves o \u000ber orders of magnitude increases in accelerating gradient over standard accelerating\nstructures [2] (which are limited by electrical breakdown), thus holding the promise of much shorter particle accelera-\ntors [3]. High quality electron beams of energy up-to 1 GeV have been produced in just a few centimeters [4, 5, 6, 7],\nwith 10 GeV stages being planned as modules of a high energy collider [8].\nAs the laser propagates through a plasma, it displaces electrons while ions remain essentially static, creating a\npocket of positive charges that the displaced electrons rush to fill. The resulting coherent periodic motion of the\nelectrons oscillating around their original position creates a wake with periodic structure following the laser. The\nalternate concentration of positive and negative charges in the wake creates very intense electric fields. An electron (or\npositron) beam injected with the right phase can be accelerated by those fields to high energy in a much shorter distance\nthan is possible in conventional particle accelerators. The e \u000eciency and quality of the acceleration is governed by\n2several factors which require precise three-dimensional shaping of the plasma column, as well as the laser and particle\nbeams, and understanding of their evolution.\nComputer simulations have had a profound impact on the design and understanding of past and present exper-\niments [9], with accurate modeling of the wake formation and beam acceleration requiring fully kinetic methods\n(usually Particle-In-Cell) with large computational resources due to the wide range of space and time scales involved\n[10, 11]. For example, modeling 10 GeV stages for the LOASIS (LBNL) BELLA proposal [12] demanded as many\nas 5,000 processor hours for a one-dimension simulation on a NERSC supercomputer [13]. Various reduced models\nhave been developed to allow multidimensional simulations at manageable computational costs: fluid approximation\n[14], quasistatic approximation [15, 16, 17], laser envelope models [16], scaled parameters [18, 19]. However, the\nvarious approximations that they require result in a narrower range of applicability. As a result, even using several\nmodels concurrently does not usually provide a complete description. For example, scaled simulations of 10 GeV\nLPA stages do not capture correctly some essential transverse physics, e.g. the laser and beam betatron motion, which\ncan lead to inaccurate beam emittance (a measure of the beam quality). An envelope description can capture these\ne\u000bects correctly at full scale for the early propagation through the plasma but can fail as the laser spectrum broadens\ndue to energy depletion as it propagates further in the plasma.\nAn alternative approach allows for orders of magnitude speedup of simulations, whether at full or reduced scale,\nvia the proper choice of a reference frame moving near the speed of light in the direction of the laser [1]. It does\nso without alteration to the fundamental equations of particle motion or electrodynamics, provided that the high-\nfrequency part of the light emitted counter to the direction of propagation of the beam can be neglected. This approach\nexploits the properties of space and time dilation and contraction associated with the Lorentz transformation. As\nshown in [1], the ratio of longest to shortest space and time scales of a system of two or more components crossing at\nrelativistic velocities is not invariant under such a transformation (a laser crossing a plasma is just such a relativistic\ncrossing). Since the number of computer operations (e.g., time steps), for simulations based on formulations from first\nprinciples, is proportional to the ratio of the longest to shortest time scale of interest, it follows that such simulations\nwill eventually have di \u000berent computer runtimes, yet equivalent accuracy, depending solely upon the choice of frame\nof reference.\nThe procedure appears straightforward: identify the frame of reference which will minimize the range of space\nand/or time scales and perform the calculation in this frame. However, several practical complications arise. First,\nthe input and output data are usually known from, or compared to, experimental data. Thus, calculating in a frame\nother than the laboratory entails transformations of the data between the calculation frame and the laboratory frame.\nSecond, while the fundamental equations of electrodynamics and particle motion are written in a covariant form, the\nnumerical algorithms that are derived from them may not retain this property, and calculations in frames moving at\ndi\u000berent velocities may not be successfully conducted with the use of the exact same algorithms. For example, it\nwas shown in [20] that calculating the propagation of ultra-relativistic charged particle beams in an accelerator using\nstandard Particle-In-Cell techniques lead to large numerical errors, which were fixed by developing a new particle\npusher. The modeling of a laser plasma accelerator (LPA) stage in a boosted frame involves the fully electromagnetic\nmodeling of a plasma propagating at near the speed of light, for which Numerical Cerenkov [21, 22] is a potential\nissue. Third, electromagnetic calculations that include wave propagation will include waves propagating forward and\nbackward in any direction. For a frame of reference moving in the direction of the accelerated beam (or equivalently\nthe wake of the laser), waves emitted by the plasma in the forward direction expand while the ones emitted in the\nbackward direction contract, following the properties of the Lorentz transformation. If one is to resolve both forward\nand backward propagating waves emitted from the plasma, there is no gain in selecting a frame di \u000berent from the\nlaboratory frame. However, the physics of interest for a laser wakefield is the laser driving the wake, the wake, and\nthe accelerated beam. Backscatter is weak in the short-pulse regime, and does not interact as strongly with the beam\nas do the forward propagating waves which stay in phase for a long period. It is thus often assumed that the backward\npropagating waves can be neglected in the modeling of LPA stages. The accuracy of this assumption is shown by\ncomparison between explicit codes which include both forward and backward waves and envelope or quasistatic\ncodes which neglect backward waves [10, 19, 23].\nAfter the idea and basic scaling for performing simulations of LPA in a Lorentz boosted frame were published\nin [1], there have been several reports of the application of the technique to various regimes of LPA [13, 24, 25, 26,\n11, 27, 28]. Speedups varying between several to a few thousands were reported with various levels of accuracy in\nagreement between simulations performed in a Lorentz boosted frames and in a laboratory frame. High-frequency\n3instabilities were reported to develop in 2D or 3D calculations, that were limiting the velocity of the boosted frame\nand thus the attainable speedup [29, 27, 28].\nIn this paper, we present numerical techniques that were implemented in the Particle-In-Cell code Warp [30]\nfor mitigating the numerical Cerenkov instability, including a solver with tunable coe \u000ecients, and show that these\ntechniques are e \u000bective for suppressing the high frequency instability observed in boosted frame simulations. A\ndetailed study of the application of these techniques to the simulations of downscaled LPA stages reveals that choosing\nthe frame of the wakefield as the reference frame allows for more aggressive application of the standard techniques\nmitigating numerical Cerenkov, than is possible in laboratory frame simulations. It is shown that the instability that\ndevelops with high-boost frames is well controlled, allowing for the first time 2D and 3D simulations of LPA in the\nwakefield frame, for 100 GeV and 1 TeV class stages, achieving the maximum theoretical speedups of over 105and\n106respectively.\nThis paper is organized as follows. The theoretical speedup expected for performing the modeling of a LPA stage\nin a boosted frame is derived in Section 2. Section 3 addresses the issue of input and output of data in a boosted frame.\nHigh frequency instability issues and remedies are presented in Section 4. These techniques enable accurate modeling\nof 0.1 GeV-1 TeV LPA stages. Stage modeling results are presented in section 5, and observed speedup is contrasted\nto the theoretical speedup of section 2.\n2. Theoretical speedup dependency with the frame boost\nThe obtainable speedup is derived as an extension of the formula that was derived in [1], taking in addition into\naccount the group velocity of the laser as it traverses the plasma. In [1], the laser was assumed to propagate at the\nvelocity of light in vacuum during the entire process, which is a good approximation when the relativistic factor of\nthe frame boost \ris small compared to the relativistic factor of the laser wake \rwin the plasma. The expression is\ngeneralized here to higher values of \r, for which the actual group velocity of the wake in the plasma must be taken\ninto account. We shall show that for a 10 GeV class LPA stage, the maximum attainable speedup is above four orders\nof magnitude.\nAssuming that the simulation box is a fixed number of plasma periods long, which implies the use (which is\nstandard) of a moving window following the wake and accelerated beam, the speedup is given by the ratio of the time\ntaken by the laser pulse and the plasma to cross each other, divided by the shortest time scale of interest, that is the\nlaser period. Assuming for simplicity that the wake propagates at the group velocity of plane waves in a uniform\nplasma of density ne, the group velocity of the wake is given by\nvw=c=\fw=0BBBBB@1+!2\np\n!21CCCCCA\u00001=2\n(1)\nwhere!p=p\n(nee2)=(\u000f0me) is the plasma frequency, !=2\u0019c=\u0015is the laser frequency, \u000f0is the permittivity of\nvacuum, cis the speed of light in vacuum, and eandmeare respectively the charge and mass of the electron.\nIn the simulations presented herein, the runs are stopped when the last electron beam macro-particle exits the\nplasma, and a measure of the total time of the simulation is given by\nT=L+\u0011\u0015p\nvw\u0000vp(2)\nwhere\u0015p\u00192\u0019c=!pis the wake wavelength, Lis the plasma length, vwandvp=\fpcare respectively the velocity\nof the wake and of the plasma relative to the frame of reference, and \u0011is an adjustable parameter for taking into\naccount the fraction of the wake which exited the plasma at the end of the simulation. The numerical cost Rtscales as\nthe ratio of the total time to the shortest timescale of interest, which is the inverse of the laser frequency, and is thus\ngiven by\nRt=Tc\n\u0015=\u0010\nL+\u0011\u0015p\u0011\n\u0010\n\fw\u0000\fp\u0011\n\u0015(3)\n4In the laboratory, vp=0 and the expression simplifies to\nRlab=Tc\n\u0015=\u0010\nL+\u0011\u0015p\u0011\n\fw\u0015(4)\nIn a frame moving at \fc, the quantities become\n\u0015\u0003\np=\u0015p=\u0002\r(1\u0000\fw\f)\u0003(5)\nL\u0003=L=\r (6)\n\u0015\u0003=\r(1+\f)\u0015 (7)\n\f\u0003\nw=(\fw\u0000\f)=(1\u0000\fw\f) (8)\nv\u0003\np=\u0000\fc (9)\nT\u0003=L\u0003+\u0011\u0015\u0003\np\nv\u0003w\u0000v\u0003p(10)\nR\u0003\nt=T\u0003c\n\u0015\u0003=\u0010\nL\u0003+\u0011\u0015\u0003\np\u0011\n\u0000\f\u0003w+\f\u0001\u0015\u0003(11)\nwhere\r=1=p\n1\u0000\f2.\nThe expected speedup from performing the simulation in a boosted frame is given by the ratio of RlabandR\u0003\nt\nS=Rlab\nR\u0003\nt=(1+\f)\u0010\nL+\u0011\u0015p\u0011\n(1\u0000\f\fw)L+\u0011\u0015p(12)\nAssuming that \r<<\r w, and that\fw\u00191 (which is a valid approximation for most practical cases of interest), this\nexpression is consistent with the expression derived in [1] for the LPA case which states that R\u0003\nt=\u000bRt=(1+\f)with\n\u000b=(1\u0000\f+l=L)=(1+l=L), where lis the laser length which is generally proportional to \u0011\u0015p, and S=Rt=R\u0003\nT.\nThe linear theory predicts that for the intense lasers (a &1) typically used for acceleration, the laser depletes\nits energy over approximately the same length Ld=\u00153\np=2\u00152over which the particles dephase from the wake [2].\nAcceleration is compromised beyond Ldand in practice, the plasma length is proportional to the dephasing length, i.e.\nL=\u0018Ld. In most cases, \r2\nw>>1, which allows the approximations \fw\u00191\u0000\u00152=2\u00152\np, and L=\u0018\u00153\np=2\u00152\u0019\u0018\r2\nw\u0015p=2>>\n\u0011\u0015p, so that Eq.(12) becomes\nS=(1+\f)2\r2 \u0018\r2\nw\n\u0018\r2w+(1+\f)\r2(\u0018\f=2+2\u0011)(13)\nFor low values of \r, i.e. when\r<<\r w, Eq.(13) reduces to\nS\r<<\r w=(1+\f)2\r2(14)\nConversely, if \r!1 , Eq.(13) becomes\nS\r!1=4\n1+4\u0011=\u0018\r2\nw (15)\nFinally, in the frame of the wake, i.e. when \r=\rw, assuming that \fw\u00191, Eq.(13) gives\nS\r=\rw\u00192\n1+2\u0011=\u0018\r2\nw (16)\nSince\u0011and\u0018are of order unity, and the practical regimes of most interest satisfy \r2\nw>>1, the speedup that is\nobtained by using the frame of the wake will be near the maximum obtainable value given by Eq.(15).\nNote that without the use of a moving window, the relativistic e \u000bects that are at play in the time domain would\nalso be at play in the spatial domain, as shown in [1], and the \r2scaling would transform to \r4. In the frame of the\n5wake, there is no need of the moving window, thus simplifying the procedure, while in a frame traveling faster than\nthe wake in the laboratory, a moving window propagating in the backward direction is needed. However, the scaling\nshows that there would be very little gain in doing the latter.\n2.1. Estimated speedup for 0.1-100 GeV stages\nFormula (13) is used to estimate the speedup for the calculations of 100 MeV , 1 GeV , 10 GeV and 100 GeV class\nstages, assuming a laser wavelength \u0015=0:8\u0016m. Using parameters and scaling laws from [18], the corresponding\ninitial plasma densities neare respectively 1019cc, 1018cc, 1017cc and 1016cc, while the plasma lengths Lare 1.5 mm,\n4.74 cm, 1.5 m, and 47.4 m, with \u0018\u00191:63. For these values, the wake wavelengths \u0015pare respectively 10 :6\u0016m,\n33:4\u0016m, 106:\u0016m, 334:\u0016m, and relativistic factors \rware 13:2, 41:7, 132 and 417 :In the simulations presented in this\npaper, the beam is injected near the end of the wake period (first ”bucket”). In first approximation, the beam has\npropagated through about half a wake period to reach full acceleration, and we set \u0011\u00190:5. For a beam injected into\nthenthbucket,\u0011would be set to n\u00001=2. If positrons were considered, they would be injected half a wake period\nahead of the location of the electrons injection position for a given period, and one would have \u0011=n\u00001. For the\nparameters considered here, L\u0019\u0015p=\r2\nw, and (15) gives S\r!1\u00192\r2\nw.\nFigure 1: Speedup versus relativistic factor of the boosted frame from Eq.(13) for 100 MeV - 100 GeV LPA class stages.\nThe speedup versus the relativistic factor of the boosted frame \ris plotted in Fig. 1. As expected, for low values\nof\r, the speedup scales as (14), and asymptotes to a value slightly lower than 2 \r2\nwfor large values of \r. It is of interest\nto note that the qualitative behavior is identical to the one obtained in [1] (see Fig. 1 and accompanying analysis) in\nthe analysis of the crossing of two rigid identical beams, confirming the generality of the generic analysis presented\nin [1]. For a 100 GeV class stage, the maximum estimated speedup is as large as 300,000.\n3. Input and output to and from a boosted frame simulation\nThis section describes the procedures that have been implemented in the Particle-In-Cell framework Warp [30] to\nhandle the input and output of data between the frame of calculation and the laboratory frame. Simultaneity of events\nbetween two frames is valid only for a plane that is perpendicular to the relative motion of the frame. As a result, the\ninput/output processes involve the input of data (particles or fields) through a plane, as well as output through a series\nof planes, all of which are perpendicular to the direction of the relative velocity between the frame of calculation and\nthe other frame of choice.\n63.1. Input\n3.1.1. Particles\nParticles are launched through a plane using a technique which applies to many calculations in a boosted frame, in-\ncluding LPA, and is illustrated using the case of a positively charged particle beam propagating through a background\nof cold electrons in an assumed continuous transverse focusing system, leading to a growing transverse instability [1].\nIn the laboratory frame, the electron background is initially at rest and a moving window is used to follow the beam\nprogression. Traditionally, the beam macroparticles are initialized all at once in the window, while background elec-\ntron macroparticles are created continuously in front of the beam on a plane that is perpendicular to the beam velocity.\nIn a frame moving at some fraction of the beam velocity in the laboratory frame, the beam initial conditions at a given\ntime in the calculation frame are generally unknown and one must initialize the beam di \u000berently. However, it can be\ntaken advantage of the fact that the beam initial conditions are often known for a given plane in the laboratory, either\ndirectly, or via simple calculation or projection from the conditions at a given time. Given the position and velocity\nfx;y;z;vx;vy;vzgfor each beam macroparticle at time t=0 for a beam moving at the average velocity vb=\fbc(where\ncis the speed of light) in the laboratory, and using the standard synchronization ( z=z0=0 att=t0=0) between the\nlaboratory and the calculation frames, the procedure for transforming the beam quantities for injection in a boosted\nframe moving at velocity \fcin the laboratory is as follows (the superscript0relates to quantities known in the boosted\nframe while the superscript\u0003relates to quantities that are know at a given longitudinal position z\u0003but di \u000berent times\nof arrival):\n1. project positions at z\u0003=0 assuming ballistic propagation\nt\u0003=(z\u0000¯z)=vz (17)\nx\u0003=x\u0000vxt\u0003(18)\ny\u0003=y\u0000vyt\u0003(19)\nz\u0003=0 (20)\nthe velocity components being left unchanged,\n2. apply Lorentz transformation from laboratory frame to boosted frame\nt0\u0003=\u0000\rt\u0003(21)\nx0\u0003=x\u0003(22)\ny0\u0003=y\u0003(23)\nz0\u0003=\r\fct\u0003(24)\nv0\u0003\nx=v\u0003\nx\n\r(1\u0000\f\fb)(25)\nv0\u0003\ny=v\u0003\ny\n\r(1\u0000\f\fb)(26)\nv0\u0003\nz=v\u0003\nz\u0000\fc\n1\u0000\f\fb(27)\nwhere\r=1=p\n1\u0000\f2. With the knowledge of the time at which each beam macroparticle crosses the plane into\nconsideration, one can inject each beam macroparticle in the simulation at the appropriate location and time.\n3. synchronize macroparticles in boosted frame, obtaining their positions at a fixed t0(=0) which is before any\nparticle is injected\nz0=z0\u0003\u0000¯v0\u0003\nzt0\u0003(28)\nThis additional step is needed for setting the electrostatic or electromagnetic fields at the plane of injection.\nIn a Particle-In-Cell code, the three-dimensional fields are calculated by solving the Maxwell equations (or\nstatic approximation like Poisson, Darwin or other [20]) on a grid on which the source term is obtained from\n7the macroparticles distribution. This requires generation of a three-dimensional representation of the beam\ndistribution of macroparticles at a given time before they cross the injection plane at z0\u0003. This is accomplished\nby expanding the beam distribution longitudinally such that all macroparticles (so far known at di \u000berent times\nof arrival at the injection plane) are synchronized to the same time in the boosted frame. To keep the beam\nshape constant, the particles are ”frozen” until they cross that plane: the three velocity components and the\ntwo position components perpendicular to the boosted frame velocity are fixed, while the remaining position\ncomponent is advanced at the average beam velocity. As particles cross the plane of injection, they become\nregular ”active” particles with full 6-D dynamics.\nFigure 2: (top) Snapshot of a particle beam “frozen” (grey spheres) and “active” (colored spheres) macroparticles traversing the injection plane (red\nrectangle). (bottom) Snapshot of the beam macroparticles (colored spheres) passing through the background of electrons (dark brown streamlines)\nand the diagnostic stations (red rectangles). The electrons, the injection plane and the diagnostic stations are fixed in the laboratory plane, and are\nthus counterpropagating to the beam in a boosted frame.\nFigure 2 (top) shows a snapshot of a beam that has passed partly through the injection plane. As the frozen beam\nmacroparticles pass through the injection plane (which moves opposite to the beam in the boosted frame), they are\nconverted to “active” macroparticles. The charge or current density is accumulated from the active and the frozen\nparticles, thus ensuring that the fields at the plane of injection are consistent.\n3.1.2. Laser\nSimilarly to the particle beam, the laser is injected through a plane perpendicular to the axis of propagation of the\nlaser (by default z). The electric field E?that is to be emitted is given by the formula\nE?(x;y;t)=E0f(x;y;t)sin\u0002!t+\u001e(x;y;!)\u0003(29)\nwhere E0is the amplitude of the laser electric field, f(x;y;t)is the laser envelope, !is the laser frequency, \u001e(x;y;!)\nis a phase function to account for focusing, defocusing or injection at an angle, and tis time. By default, the laser\nenvelope is a three dimensional gaussian of the form\nf(x;y;t)=e\u0000(x2=2\u001b2\nx+y2=2\u001b2\ny+c2t2=2\u001b2\nz)(30)\nwhere\u001bx,\u001byand\u001bzare the dimensions of the laser pulse; or it can be defined arbitrarily by the user at runtime. If\n\u001e(x;y;!)=1, the laser is injected at a waist and parallel to the axis z.\n8If, for convenience, the injection plane is moving at constant velocity \fsc, the formula is modified to take the\nDoppler e \u000bect on frequency and amplitude into account and becomes\nE?(x;y;t)=(1\u0000\fs)E0f(x;y;t)sin\u0002(1\u0000\fs)!t+\u001e(x;y;!)\u0003: (31)\nThe injection of a laser of frequency !is considered for a simulation using a boosted frame moving at \fcwith\nrespect to the laboratory. Assuming that the laser is injected at a plane that is fixed in the laboratory, and thus moving\nat\fs=\u0000\fin the boosted frame, the injection in the boosted frame is given by\nE?\u0000x0;y0;t0\u0001=(1\u0000\fs)E0\n0f\u0000x0;y0;t0\u0001sin\u0002(1\u0000\fs)!0t0+\u001e\u0000x0;y0;!0\u0001\u0003(32)\n=(E0=\r)f\u0000x0;y0;t0\u0001sin\u0002!t0=\r+\u001e\u0000x0;y0;!0\u0001\u0003(33)\nsince E0\n0=E0=!0=!=1=(1+\f)\r.\nThe electric field is then converted into currents that get injected via two dual 2-D arrays of macro-particles, with\none positive and one negative macro-particle per cell in the plane of injection, whose weights and motion are governed\nbyE?(x0;y0;t0). Injecting using these dual arrays of macroparticles o \u000ber the advantages of automatically including\nthe longitudinal component which arise from emitting into a boosted frame, and to verify the discrete Gauss law\nthanks to the use of the Esirkepov current deposition scheme [31].\nThe technique implemented in Warp presents several advantage over other procedures that have been proposed\nelsewhere [13, 28]. In [28], the laser beam is initialized entirely in the computational box, leading to larger boxes\ntransversely in a boosted frame, as the Rayleigh length of the laser shortens and the overall laser pulse radius rises,\neventually o \u000bsetting the benefits of the boosted frame. The transverse broadening of the box is avoided in [13] at the\ncost of a more complicated injection scheme, requiring to launch the laser from all but one faces of the simulation\nbox. The method presented here avoids the caveat of the broadening and retains simplicity with a standard injection\ntechnique through one plane.\n3.2. Output\nSome quantities, e.g. charge, are Lorentz invariant, while others, like dimensions perpendicular to the boost\nvelocity, are the same in the laboratory frame. Those quantities are thus readily available from standard diagnostics\nin the boosted frame calculations. Quantities which do not fall in this category are recorded at a number of regularly\nspaced “stations”, immobile in the laboratory frame, at a succession of time intervals to record data history, or averaged\nover time. A visual example is given on Fig. 2 (bottom). Since the space-time locations of the diagnostic grids in the\nlaboratory frame generally do not coincide with the space-time positions of the macroparticles and grid nodes used for\nthe calculation in a boosted frame, some interpolation is performed at runtime during the data collection process. As\na complement or an alternative, selected particle or field quantities are dumped at regular interval for post-processing.\nThe choice of the methods depends on the requirements of the diagnostics and particular implementations.\n4. High frequency instability and Numerical Cerenkov\nAs reported in [27] and [28], for simulations using a boosted frame at \r\u001510\u000020 (depending on parameters),\na fast growing short wavelength instability was observed developing at the front of the plasma (see Fig. 3). The\npresence and growth rate of the instability was observed to be very sensitive to the resolution (slower growth rate at\nhigher resolution), choice of field solver, and to the amount of damping of high frequencies and smoothing of short\nwavelengths. The instability is always propagating at some angle from the longitudinal axis, and is observed in 2D\nand 3D runs but was never observed in any of the 1D runs performed by the authors. When modeling an LPA setup in\na relativistically boosted frame, the background plasma is traveling near the speed of light and it has been conjectured\n[28] that he observed instability might be caused by numerical Cerenkov. We investigate in this paper whether the\ninstability that is observed in boosted frame simulations of LPA is indeed of numerical Cerenkov type and if the cures\naimed at mitigating numerical Cerenkov are e \u000bective.\nDue to spatial and time discretization of the Maxwell equations, numerical light waves may travel faster or slower\non the computational grid than the actual speed of light in vacuum c, with the magnitude of the e \u000bect being larger at\nshort wavelength, where discretization errors are the largest. When the numerical speed is lower than c, it is possible\n9Figure 3: Snapshot of a surface plot of the longitudinal field from a 2-1 /2D simulation of a full scale 10GeV LPA in a boosted frame at \r=130\n(elevation is proportional to the magnitude of the electric field). The laser is propagating from left to right and the plasma from right to left. A fast\ngrowing short wavelength instability is developing at the front of the plasma.\nfor fast macro-particles to travel faster than the wave modes, leading to numerical Cerenkov e \u000bects that may result\nin instabilities [21, 22, 32, 33, 34]. The e \u000bect was studied analytically and numerically in detail for one-dimensional\nsystems in [32, 33]. Several solutions were proposed: smoothing the current deposited by the macro-particles [21, 32],\ndamping the electromagnetic field [34, 35, 36], solving the Maxwell equations in Fourier space [22], or using a field\nsolver with a larger stencil to provide lower numerical dispersion [34].\nSeveral of the abovementioned techniques to mitigate numerical Cerenkov and high frequency errors have been\nimplemented in Warp. All the simulations presented in this paper employed cubic splines for current deposition and\nelectromagnetic force gathering between the macro-particles and the grid [37], whose beneficial e \u000bects on standard\nLPA PIC simulations have been demonstrated in [38]. In addition, a Maxwell solver with tunable coe \u000ecients was\nimplemented, as well as a damping scheme, and filtering of the deposited current and gathered electromagnetic fields,\nwhich are described in this section. The use of Fourier based Maxwell solvers is not considered in this paper.\n4.1. Wideband lowpass digital filtering\nIt is common practice to apply digital filtering to the charge or current density in Particle-In-Cell simulations, for\nsmoothing or compensation purpose [47]. The most commonly used filter is the three points filter\n\u001ef\nj=\u000b\u001ej+(1\u0000\u000b)\u001ej\u00001+\u001ej+1\n2(34)\nwhere\u001efis the filtered quantity. This filter is called a binomial filter when \u000b=0:5. Assuming \u001e=ejkxand\n\u001ef=g(\u000b;k)ejkx, where gis the filter gain, which is function of the filtering coe \u000ecient\u000band the wavenumber k, we\nfind from (34) that\ng(\u000b;k)=\u000b+(1\u0000\u000b)cos(k\u000ex) (35)\n\u00191\u0000(1\u0000\u000b)(k\u000ex)2\n2+O\u0010\nk4\u0011\n(36)\nFornsuccessive applications of filters of coe \u000ecients\u000b1...\u000bn, the total attenuation Gis given by\nG=nY\ni=1g(\u000bi;k) (37)\n\u00191\u00000BBBBB@n\u0000nX\ni=1\u000bi1CCCCCA(k\u000ex)2\n2+O\u0010\nk4\u0011\n(38)\nIf\u000bn=n\u0000Pn\u00001\ni=1\u000bithen G\u00191+O\u0010\nk4\u0011\n, providing a sharper cuto \u000binkspace. Such step is called a compensation\nstep [47]. For the bilinear filter ( \u000b=1=2), the compensation factor is \u000bc=2\u00001=2=3=2. For a succession of n\napplications of the bilinear factor, it is \u000bc=n=2+1. The gain versus wavelength is plotted in Fig. 4 for the bilinear\n10Figure 4: Gain versus wavelength for the bilinear filter without compensation ( g=g(1=2;k)), with compensation ( g\u0001c3=2=g(1=2;k)\u0001g(3=2;k)),\nand n-pass bilinear filters with compensation ( gn\u0001c\u000bc=g(1=2;k)n\u0001g(\u000bc;k)) for n=4, 20, 50 and 80.\nfilter without compensation ( G=g(1=2;k)), with compensation ( G=g(1=2;k)\u0001g(3=2;k)), and four n-pass bilinear\nfilters with compensation ( G=g(1=2;k)n\u0001g(3=2;k)) for n=4, 20, 50 and 80.\nThe bilinear filter provides complete suppression of the signal at the grid Nyquist wavelength (twice the grid cell\nsize). Suppression of the signal at integers of the Nyquist wavelength can be obtained by using a stride sin the filter\n\u001ef\nj=\u000b\u001ej+(1\u0000\u000b)\u001ej\u0000s+\u001ej+s\n2(39)\nfor which the gain is given by\ng(s;\u000b;k)=\u000b+(1\u0000\u000b)cos(sk\u000ex) (40)\n\u00191\u0000(1\u0000\u000b)(sk\u000ex)2\n2+O\u0010\nk4\u0011\n(41)\nThe gain is plotted in Fig. 5 (top) for four passes bilinear filters with compensation ( G=g(s;1=2;k)4\u0001g(s;3=2;k))\nfor strides s =1 to 4. For a given stride, the gain is given by the gain of the bilinear filter shifted in k space, with the\npole g=0 shifted from \u0015=2=\u000exto\u0015=2s=\u000ex, with additional poles, as given by\nsk\u000ex=arccos\u0012\u000b\n\u000b\u00001\u0013\n(mod 2\u0019) (42)\nThe resulting filter is pass band between the poles, but since the poles are spread at di \u000berent integer values in k space,\na wide band low pass filter can be constructed by combining filters at di \u000berent strides. Examples are given in Fig. 5\n(bottom) for combinations of the filter with stride 1 to 4.\nThe combined filters with strides 2, 3 and 4 have nearly equivalent fall-o \u000bs in gain (in linear scale) to the 20, 50\nand 80 passes of the bilinear filter (see Fig. 6). Yet, the filters with stride need respectively 10, 15 and 15 passes of\na three-point filter while the n-pass bilinear filer need respectively 21, 51 and 81 passes, giving gains of respectively\n2.1, 3.4 and 5.4 in number of operations in favor of the filters with stride.\n11Figure 5: (top) gain for four passes bilinear filters with compensation ( Gs=g(s;1=2;k)4\u0001g(s;3=2;k)) for strides s =1 to 4 linear with (left)\nlinear ordinate (right) logarithmic ordinate; (bottom) gain for four low pass filters combining the G1toG4filters with (left) linear ordinate (right)\nlogarithmic ordinate.\nFigure 6: Comparison between filters with stride and filter s20-80 with (left) linear ordinate (right) logarithmic ordinate.\n124.2. Tunable solver\nIn [39] and [40], Cole introduced an implementation of the source-free Maxwell’s wave equations for narrow-\nband applications based on non-standard finite-di \u000berences (NSFD). In [41], Karkkainen et al adapted it for wideband\napplications. At the Courant limit for the time step and for a given set of parameters, the stencil proposed in [41]\nhas no numerical dispersion along the principal axes, provided that the cell size is the same along each dimension\n(i.e. cubic cells in 3D). The solver from [41] was modified to be consistent with the Particle-In-Cell methodology and\nimplemented in the code Warp, with the ability given to the user of setting the solver adjustable coe \u000ecients, providing\ntunability of the numerical properties of the solver to better fit the requirements of a particular application.\nThe ”Cole-Karkkainnen”’s solver [41] uses a non-standard finite di \u000berence formulation (extended stencil) of the\nMaxwell-Ampere equation. For implementation into a Particle-In-Cell code, the formulation must introduce the\nsource term into Cole-Karkkainen’s source free formulation in a consistent manner. However, modifying the NSFD\nformulation of the Maxwell-Ampere equation so that it includes the source term in a way that is consistent with the cur-\nrent deposition scheme is challenging. To circumvent this problem, Warp implementation departs from Karkkainen’s\nby applying the enlarged stencil on the Maxwell-Faraday equations, which does not contain any source term but\nis formally equivalent to the source-free Maxwell-Ampere equation. Consequently, in Warp’s implementation, the\ndiscretized Maxwell-Ampere equation is the same as in the Yee scheme, and the discretized Maxwell’s equations\nread:\n\u0001tB=\u0000r\u0003\u0002E (43)\n\u0001tE=c2r\u0002B\u0000J\n\u000f0(44)\nr\u0001E=\u001a\n\u000f0(45)\nr\u0003\u0001B=0 (46)\nwhere\u000f0is the permittivity of vacuum, and Eq. 45 and 46 not being solved explicitly but verified via appropriate\ninitial conditions and current deposition procedure. The di \u000berential operators are defined as\nr= \u0001 xˆx+ \u0001 yˆy+ \u0001 zˆz (47)\nr\u0003= \u0001\u0003\nxˆx+ \u0001\u0003\nyˆy+ \u0001\u0003\nzˆz; (48)\nthe finite di \u000berences and sums operators being\n\u0001tGjn\ni;j;k=Gjn+1=2\ni;j;k\u0000Gjn\u00001=2\ni;j;k\n\u000et(49)\n\u0001xGjn\ni;j;k=Gjn\ni+1=2;j;k\u0000Gjn\ni\u00001=2;j;k\n\u000ex(50)\n\u0001\u0003\nx=\u0010\n\u000b+\fS1\nx+\rS2\nx\u0011\n\u0001x (51)\nwith\nS1\nxGjn\ni;j;k=Gjn\ni;j+1=2;k+Gjn\ni;j\u00001=2;k\n+Gjn\ni;j;k+1=2+Gjn\ni;j;k\u00001=2 (52)\nS2\nxGjn\ni;j;k=Gjn\ni;j+1=2;k+1=2+Gjn\ni;j\u00001=2;k+1=2\n+Gjn\ni;j+1=2;k\u00001=2+Gjn\ni;j\u00001=2;k\u00001=2 (53)\nThe quantity Gis a sample vector component, \u000etand\u000exare respectively the time step and the grid cell size along\nx, while\u000b,\fand\rare constant scalars verifying \u000b+4\f+4\r=1. The operators along yandz, i.e.\u0001y,\u0001z,\u0001\u0003\ny,\u0001\u0003\nz,S1\ny,\nS1\nz,S2\ny, and S2\nz, are obtained by circular permutation of the indices.\n13In 2D, assuming the plane ( x;z), the enlarged finite operators simplify to\n\u0001\u0003\nx=\u0010\n\u000b+\fS1\nx\u0011\n\u0001x (54)\nS1\nxGjn\ni;j;k=Gjn\ni;j+1=2;k+Gjn\ni;j\u00001=2;k: (55)\nAn extension of this algorithm for non-cubic cells provided by Cowan in [43] is not considered in this paper. How-\never, all considerations given here for the solver implemented in Warp apply readily to the solver developed by Cowan.\n4.2.1. Numerical dispersion\nThe dispersion relation of the solver is given by\n0BBBB@sin!\u000et\n2\nc\u000et1CCCCA2\n=Cx0BBBBB@sinkx\u000ex\n2\n\u000ex1CCCCCA2\n+Cy0BBBBBB@sinky\u000ey\n2\n\u000ey1CCCCCCA2\n+Cz0BBBBB@sinkz\u000ez\n2\n\u000ez1CCCCCA2\n(56)\nwith\nCx=\u000b+2\f(cy+cz)+4\rcycz (57)\nCy=\u000b+2\f(cz+cx)+4\rczcx (58)\nCz=\u000b+2\f(cx+cy)+4\rcxcy (59)\nand\ncx=cos(kx\u000ex) (60)\ncy=cos\u0010\nky\u000ey\u0011\n(61)\ncz=cos(kz\u000ez) (62)\nThe Courant-Friedrichs-Lewy condition (CFL) is given by\nc\u000etc\u0014min[\u000ex;\u000ey;\u000ez;\n1=q\n(\u000b\u00004\r)maxh\n\u0014x+\u0014y;\u0014x+\u0014z;\u0014y+\u0014zi\n;\n1=q\n(\u000b\u00004\f+4\r)\u0010\n\u0014x+\u0014y+\u0014z\u0011\n] (63)\nwhere\u0014x=1=\u000ex2,\u0014y=1=\u000ey2and\u0014z=1=\u000ez2.\nAssuming cubic cells ( \u000ex=\u000ey=\u000ez), the coe \u000ecients given in [41] ( \u000b=7=12,\f=1=12 and\r=1=48) allow\nc\u000et=\u000ex, and thus no dispersion along the principal axes.\nIt is of interest to note that (56) can be rewritten\n0BBBB@sin!\u000et\n2\nc\u000et1CCCCA2\n=\u0010\ns2\nx+s2\ny+s2\nz\u0011\n+\f0\u0010\ns2\nxs2\ny+s2\nxs2\nz+s2\nys2\nz\u0011\n+\r0\u0010\ns2\nxs2\nys2\nz\u0011\n(64)\nwith sx=sin(kx\u000ex=2),sy=sin\u0010\nky\u000ey=2\u0011\n,sz=sin(kz\u000ez=2),\f0=\u00008\f\u000016\rand\r0=48\r, for which the coe \u000ecients\nfrom [41] take the nice values \f0=\u00001 and\r0=1.\nSets of possible coe \u000ecients and the corresponding CFL condition, assuming cubic cells, are given in Table 1. The\nnumerical dispersion using those coe \u000ecients are plotted in figure 7 along the principal axes and diagonals for cubic\ncells (\u000ex=\u000ey=\u000ez) and contrasted with the one of the Yee solver (all taken at each solver’s CFL time step limit). At\nthe CFL limit, the Yee algorithm o \u000bers no numerical dispersion along the 3D diagonal, but relatively large numerical\ndispersion at the Nyquist frequency along the main axes. Conversely, the Cole-Karkkainen solver (CK) o \u000bers no\nnumerical dispersion along the main axes but significant dispersion along the diagonals. The CK solver also allows\n14Yee CK CK2 CK3 CK4 CK5\n\f00\u00001\u00001=2 0\u00001=2\u00009=10\n\r00 1 1=2\u00001 0 9=10\n\u000b 1 7=12 19=24 11=12 3=4 5=8\n\f 0 1=12 1=24 1=24 1=16 3=40\n\r 0 1=48 1=96\u00001=48 0 3=160\nc\u000et=\u000ex1=p\n3 1 1=p\n2 1=p\n2p\n2=p\n3p\n5=p\n6\nTable 1: List of coe \u000ecients\nlarger time steps than the Yee solver by almost a factor of two in 3D. The solver labeled ”CK2” o \u000bers numerical\ndispersion that is intermediate between the Yee solver and the CK solver along the main axes and the 3D diagonal, but\nslightly degraded along the 2D diagonal. Conversely, while solver CK3 also o \u000bers intermediate numerical dispersion\nalong the main axes and the 3D diagonal, it o \u000bers no numerical dispersion along the 2D diagonal. Solver CK4\nimproves slightly the numerical dispersion along the main axes over CK2 and CK3 at the expense of the dispersion\nalong the diagonals. Finally, CK5 o \u000bers the highest level of isotropy. The CFL time steps of solvers CK2, 3, 4 and\n5 are intermediate between the Yee and the CK CFL time steps. This provides solvers with a range of numerical\ndispersion among which some may be more favorable with regard to the mitigation of numerical instabilities for a\ngiven application.\nTo reduce numerical dispersion to its lowest level, it is desirable to operate the CK solver as close as possible to\nthe CFL limit c\u000et=\u000ex. However, an instability (other than numerical Cerenkov) arises at the Nyquist frequency in\nsuch a case. The analysis is given in 1D in Appendix I, as well as its mitigation using digital filtering. Since for the\nCK solver, the CFL limit is independent of dimensionality, the analysis and mitigation apply readily to 2D and 3D\nsimulations.\nFor absorption of outgoing waves at the computational box boundaries, the extension of the solver to a Perfectly\nMatched Layer [44] is given in Appendix II.\n15Yee CK\nCK2 CK3\nCK4 CK5\nFigure 7: Numerical dispersion along the principal axis and diagonals for cubic cells ( \u000ex=\u000ey=\u000ez) at the Courant limit for the solver with\nadjustable numerical dispersion using the parameters from Table 1.\n164.2.2. Current deposition and Gauss’ Law\nIn most applications, it is essential to prevent accumulations of errors to the discretized Gauss’ Law. This is\naccomplished by providing a method for depositing the current from the particles to the grid which is compatible with\nthe discretized Gauss’ Law, or by providing a mechanism for ”divergence cleaning” [47, 48, 49, 50]. For the former,\nschemes which allow a deposition of the current that is exact when combined with the Yee solver is given in [51] for\nlinear form factors and in [31] for higher order form factors. Since the discretized Gauss’ Law and Maxwell-Faraday\nequation are the same in our implementation as in the Yee solver, charge conservation is readily verified using the\ncurrent deposition procedures from [51] and [31], and this was verified numerically. Hence divergence cleaning is not\nnecessary.\n4.3. Friedman adjustable damping\nThe tunable damping scheme developed by Friedman [36] was shown to be the most potent practical method for\nmitigating the numerical Cerenkov instability in [34], among the selected methods that were considered. It is readily\napplicable to the solver presented above by modifying (43) to\nBn+3=2=Bn+1=2\u0000\u000etr\u0003\u0002\"\u0012\n1+\u0012\n4\u0013\nEn+1\u00001\n2En+ 1\n2\u0000\u0012\n4!\n¯En\u00001#\n(65)\nwith\n¯En\u00001=\u0012\n1\u0000\u0012\n2\u0013\nEn+\u0012\n2¯En\u00002(66)\nwhere 0\u0014\u0012\u00141 is the damping factor. The numerical dispersion becomes\n0BBBB@sin!\u000et\n2\nc\u000et1CCCCA2\n=F\n2(67)\nwhere\nF=1\u00002\u0012sin2(!\u000et=2)\n2e\u0000i!\u000et\u0000\u0012(68)\nand\n\n2=266666664Cx0BBBBB@sinkx\u000ex\n2\n\u000ex1CCCCCA2\n+Cy0BBBBBB@sinky\u000ey\n2\n\u000ey1CCCCCCA2\n+Cz0BBBBB@sinkz\u000ez\n2\n\u000ez1CCCCCA2377777775(69)\nThe CFL is given by\nc\u000et\u0003\nc=c\u000etcr\n2+\u0012\n2+3\u0012(70)\nwhere\u000etcis the critical time step of the numerical scheme without damping ( \u0012=0), as given by (63).\nThe numerical dispersion of the Cole-Karkkainen-Friedman (CKF) solver (using the coe \u000ecients from the CK\nsolver in Table 1) is plotted in figure 8 along the principal axis and diagonals for cubic cells ( \u000ex=\u000ey=\u000ez) and\ncontrasted with the one of the Yee-Friedman (YF) solver (both taken at the Courant time step limit). The amount\nof phase error rises with the value of the damping parameter \u0012(partly due to the slightly more constraining limit on\nthe critical time step). However, it was shown in [34] that the amount of damping provided by the YF solver was\nsu\u000ecient to counteract the slight degradation of numerical dispersion with raising \u0012, reducing the numerical Cerenkov\ne\u000bects to an acceptable level for the problem that was considered. The damping is very isotropic with the CKF\nsolver but not with the YF one. The YF implementation o \u000bers a higher level of damping of the shortest wavelengths\nalong the 3D diagonals, while the CKF o \u000bers more damping along the axes, and the amount of damping along the 2D\ndiagonals are similar. In summary, the YF implementation delivers respectively the highest /lowest level of damping in\nthe direction of lowest /highest numerical dispersion, while the CKF implementation delivers a proportionally higher\nlevel of dispersion than the YF implementation along the direction of highest numerical dispersion. Thus it may be\nexpected that the CKF implementation will be more e \u000ecient in reducing numerical Cerenkov e \u000bects.\n17Yee-Friedman Cole-Karkkainen-Friedman\nFigure 8: Numerical dispersion along the principal axis and diagonals for cubic cells ( \u000ex=\u000ey=\u000ez) at the Courant limit for: (left) the Yee-Friedman\nsolver; (right) the Cole-Karkkainen-Friedman solver. The real part (phase) and the imaginary part (amplitude) are plotted respectively in the top\nand bottom rows.\n185. Application to the modeling of laser wakefield acceleration\nThis section presents applications of the methods to the modeling of 10 GeV LPA stages at full scale in 2-1 /2D\nand 3D, which has not been done fully self-consistently with other methods. It has been shown that many parameters\nof high energy LPA stages can be accurately simulated at reduced cost by simulating stages of lower energy gain, with\nhigher density and shorter acceleration distance, by scaling the physical quantities relative to the plasma wavelength,\nand this has been applied to design of 10 GeV LPA stages [18, 19]. The number of oscillations of a mismatched laser\npulse in the plasma channel however depends on stage energy and does not scale, though this e \u000bect is minimized for\na channel guided stage as considered in [18, 19]. The number of betatron oscillations of the trapped electron bunch\nwill also depend on the stage energy, and may a \u000bect quantities like the emittance of the beam. For these reasons, and\nto prove validity of scaled designs of other parameters, it is necessary to perform full scale simulations, which is only\npossible by using reduced models or simulations in the boosted frame.\nAs a benchmarking exercise, we first perform scaled simulations similar to the ones performed in [18], at a\ndensity of ne=1019cm\u00003, using various values of the boosted frame relativistic factor \rto show the accuracy and\nconvergence of the technique. These stages were shown to e \u000eciently accelerate both electrons and positrons with low\nenergy spread, and the scaled simulations predicted acceleration of hundreds of pC to 10 GeV energies using a 40\nJ laser. The accuracy of the technique is evaluated by modeling scaled stages [18, 19] at 0.1 GeV , which allows for\na detailed comparison of simulations using a reference frame ranging from the laboratory frame to the frame of the\nwake. Excellent agreement is obtained on wakefield histories on axis, beam average energy history and momentum\nspread at peak energy, with speedup over a hundred, in agreement with the theoretical estimates from Section 2. The\ndownscaled simulations are also used for an in-depth exploration of the e \u000bects of the methods presented in Sections 3\nand 4, and evaluation of which techniques are required to permit maximum \rboost while maintaining high accuracy.\nWe then apply the boosted frame technique to provide full scale simulation of high e \u000eciency quasilinear LPA stages\nat higher energy, verifying the scaling laws in the 10 GeV-1 TeV range.\n5.1. Scaled 10 GeV stages\nThe parameters were chosen to be close (though not identical) to the case where kpL=2 in [18] where kpis\nthe plasma wavenumber and Lis the laser pulse length. In the cases considered in this paper, the beam is composed\nof test particles only, with the goal of testing the fidelity of the algorithm in modeling laser propagation and wake\ngeneration. The results from simulations of LPA in a boosted frame where beam loading is present will be presented\nelsewhere. These simulations are scaled replicas of 10 GeV stages that would operate at actual densities of 1017cm\u00003\n[18, 19] and allow short run times to permit e \u000bective benchmarking between the algorithms. The main physical and\nnumerical parameters of the simulation are given in Table 2. Unless noted otherwise, in all the simulations presented\nherein, the field is gathered from the grid onto the particles directly from the Yee mesh locations, i.e. using the ’energy\nconserving’ procedure (see [47], chapter 10).\n5.1.1. Using standard numerical techniques\nThese runs were done using the standard Yee solver with no damping, and with the 4-pass stride-1 filter plus\ncompensation, similarly to the simulations reported in [18]. No signs of detrimental numerical instabilities were\nobserved at the resolutions reported here with these settings.\nThe approximate relativistic factor of the wake that is formed by the laser traveling in the plasma is given, ac-\ncording to linear theory, by \rw=2\u0019c=\u0015! pwhere!p=p\nnee2=\u000f0meis the electron plasma frequency. For the given\nparameters, \rw\u001913:2. Thus, Warp simulations were performed using reference frames moving between \r=1 (lab-\noratory frame) and 13. For a boosted frame associated with a value of \rapproaching \rwin the laboratory, the wake\nis expected to travel at low velocity in this boosted frame, and the physics to appear somewhat di \u000berent from the one\nobserved in the laboratory frame, in accordance to the properties of the Lorentz transformation. Figure 9 and 10 show\nsurface renderings of the transverse and longitudinal electric fields respectively, as the beam enters its early stage of\nacceleration by the plasma wake, from a calculation in the laboratory frame and another in the frame at \r=13. The\ntwo snapshots o \u000ber strikingly di \u000berent views of the same physical processes: in the laboratory frame, the wake is fully\nformed before the beam undergoes any significant acceleration and the imprint of the laser is clearly visible ahead of\nthe wake; while in the boosted frame calculation, the beam is accelerated as the plasma wake develops, and the laser\n19Table 2: List of parameters for scaled 10GeV class LPA stage simulation.\nbeam radius Rb 82:5 nm\nbeam length Lb 85:nm\nbeam transverse profile exp\u0010\n\u0000r2=8R2\nb\u0011\nbeam longitudinal profile exp\u0010\n\u0000z2=2L2\nb\u0011\nlaser wavelength \u0015 0:8\u0016m\nlaser length (FWHM) L 10:08\u0016m\nnormalized vector potential a0 1\nlaser longitudinal profile sin (\u0019z=L)\nplasma density on axis ne 1019cm\u00003\nplasma longitudinal profile flat\nplasma length L 1:5 mm\nplasma entrance ramp profile half sinus\nplasma entrance ramp length 4 \u0016m\nnumber of cells in x Nx 75\nnumber of cells in z Nz860 (\r=13)-1691 (\r=1)\ncell size in x \u000ex 0:65\u0016m\ncell size in z \u000ez \u0015=32\ntime step \u000et at CFL limit\nparticle deposition order cubic\n# of plasma particles /cell 1 macro-e\u0000+1 macro-p+\nimprint is not visible on the snapshot. Close examination reveals that the short spatial variations which make the laser\nimprint in front of the wake are transformed into time variations in the boosted frame of \r=13.\nHistories of the perpendicular and longitudinal electric fields recorded at a number of stations at fixed locations in\nthe laboratory o \u000ber direct comparison between the simulations in the laboratory frame ( \r=1) and boosted frames at\n\r=2, 5, 10 and 13. Figure 11 and 12 show respectively the transverse and longitudinal electric fields collected at the\npositions z=0:3 mm and z=1:05 mm (in the laboratory frame) on axis ( x=y=0). The agreement is excellent and\nconfirms that despite the apparent di \u000berences from snapshots taken from simulations in di \u000berent reference frames,\nthe same physics was recovered. This is further confirmed by the plot of the average scaled beam energy gain as a\nfunction of position in the laboratory frame, and of relative longitudinal momentum dispersion at peak energy (Fig.\n13). The small di \u000berences observed on the mean beam energy histories and on the longitudinal momentum spread are\nattributed to a lack of convergence at the resolution that was chosen. The beam was launched with the same phase in\nthe 2-1 /2D and the 3D simulations, resulting in lower energy gain in 3D, due to proportionally larger laser depletion\ne\u000bects in 3D than in 2-1 /2D.\nThe CPU time recorded as a function of the average beam position in the laboratory frame (Fig. 13-middle)\nindicates that the simulation in the frame of \r=13 took\u001925 s in 2-1 /2D and\u0019150 s in 3D versus \u00195;000 s in\n2-1/2D and\u001920;000 s in 3D in the laboratory frame, demonstrating speedups of \u0019200 in 2-1 /2D and\u0019130 in 3D,\nbetween calculations in a boosted frame at \r=13 and the laboratory frame.\nAll the simulations presented so far in this section were using the Yee solver, for which the Courant condition is\ngiven by c\u000et<\u0010\n1=\u000ex2+1=\u000ez2\u0011\u00001=2in 2D and c\u000et<\u0010\n1=\u000ex2+1=\u000ey2+1=\u000ez2\u0011\u00001=2in 3D where \u000etis the time step and\n\u000ex,\u000eyand\u000ezare the computational grid cell sizes in x,yandz. As\rwas varied, the transverse resolution was kept\nconstant, while the longitudinal resolution was kept at a constant fraction of the incident laser wavelength \u000ez=\u0010\u0015,\nsuch that in a boosted frame, \u000ez\u0003=\u0010\u0015\u0003=\u0010(1+\f)\r\u0015. As a result, the speedup becomes, when using the Yee solver\nSyee2D=S\u000ezp\n1=\u000ex2+1=\u000ez2\n\u000e\u0003zp\n1=\u000ex2+1=\u000ez\u00032(71)\n20Figure 9: Colored surface rendering of the transverse electric field from a 2-1 /2D Warp simulation of a laser wakefield acceleration stage in the\nlaboratory frame (top) and a boosted frame at \r=13 (bottom), with the beam (white) in its early phase of acceleration. The laser and the beam are\npropagating from left to right.\nin 2D and\nSyee3D=S\u000ezp\n1=\u000ex2+1=\u000ey2+1=\u000ez2\n\u000e\u0003zp\n1=\u000ex2+1=\u000ey2+1=\u000ez\u00032(72)\nin 3D where Sis given by Eq. (13).\nThe speedup versus relativistic factor of the reference frame is plotted in Fig. 14, from (13), (71) and (72),\nand contrasted with measured speedups from 1D, 2-1 /2D and 3D Warp simulations, confirming the scaling obtained\nanalytically.\n21Figure 10: Colored surface rendering of the longitudinal electric field from a 2-1 /2D Warp simulation of a laser wakefield acceleration stage in the\nlaboratory frame (top) and a boosted frame at \r=13 (bottom), with the beam (white) in its early phase of acceleration. The laser and the beam are\npropagating from left to right.\n2-1/2D 3-D\nFigure 11: History of transverse electric field at the position x=y=0,z=0:3 mm and z=1:05 mm (in the laboratory frame) from simulations in\nthe laboratory frame ( \r=1) and boosted frames at \r=2, 5, 10 and 13.\n222-1/2D 3-D\nFigure 12: History of longitudinal electric field at the position x=y=0,z=0:3 mm and z=1:05 mm (in the laboratory frame) from simulations\nin the laboratory frame ( \r=1) and boosted frames at \r=2, 5, 10 and 13.\n232-1/2D 3-D\nFigure 13: (top) Average scaled beam energy gain and (middle) CPU time, versus longitudinal position in the laboratory frame from simulations;\n(bottom) distribution of relative longitudinal momentum dispersion at peak energy, in the laboratory frame ( \r=1) and boosted frames at \r=2, 5,\n10 and 13.\n24Figure 14: Speedup versus relativistic factor of the boosted frame from Eq. (13), (71), (72), and Warp simulations.\n255.1.2. E \u000bect of filtering, solver with adjustable dispersion and damping\nThe modeling of full scale stages, which allow for higher values of \rfor the reference frame, is more prone to\nthe high frequency instability that was mentioned in a previous section, as we will show below. In anticipation of the\napplication of the method presented above to mitigate the instability, simulations of the scaled stage were conducted\nusing the Yee solver with digital filter S(1:2:4) as described above (Fig. 15), the Cole-Karkkainen solver (Fig. 16) or\nthe Yee-Friedman solver (Fig. 17).\nSmoothing with the wideband filter S(1:2:4) did not produce significant degradations for the calculation in the\nwake frame ( \r=13) but did otherwise. The calculations with the Yee solver and the Cole-Karkkainen solver gave\nidentical results, validating our implementation of the CK solver. Despite the more expensive stencil, the run with the\nCK solver was almost 40% faster, due to a time step larger byp\n2. Similarly to filtering, damping aggressively did\nnot degrade the result in the range 10 \u0014\r\u001413 but did significantly in the range 1 \u0014\r\u00145. Comparing the timings\nwith those of Fig. 13 (middle-left) shows that the smoothing and the damping added less than a factor of two of total\nruntime to the simulations.\nFigure 15: ( (left) Average scaled beam energy gain and (right) CPU time, versus longitudinal position in the laboratory frame from simulations\nin the laboratory frame ( \r=1) and boosted frames at \r=2, 5, 10 and 13, using the Yee solver with digital filter S(1:2:4) (grey cross is reference\nfrom run with filter S(1)).\nFigure 16: ( (left) Average scaled beam energy gain and (right) CPU time, versus longitudinal position in the laboratory frame from simulations in\nthe laboratory frame ( \r=1) and boosted frames at \r=2, 5, 10 and 13, using the Cole-Karkkainen solver with filter S(1) (red curve is reference\nfrom calculation with Yee solver and filter S(1)).\nThose results lead to several observations: (i) while the grid dimensions and number of cells were chosen such that\nsquare cells were obtained for \r=13, meaning a larger dispersion in the longitudinal direction with the Yee solver\nthan with the Cole-Karkkainen solver, both gave the same result. This is significant since for simulations of LPA in\nthe laboratory frame reported in the literature, the need to have nearly perfect numerical dispersion in the longitudinal\ndirection imposes a constraint on the cell aspect ratio and thus on resolution [45, 46]. This constraint is removed when\nsimulating in the frame of the wake ( \r=13\u0019\rw); (ii) damping of high frequencies with the Yee-Friedman solver\nor wideband smoothing of short wavelength have a negligible e \u000bect on accuracy for simulations in the frame of the\n26Figure 17: ( (left) Average scaled beam energy gain and (right) CPU time, versus longitudinal position in the laboratory frame from simulations in\nthe laboratory frame ( \r=1) and boosted frames at \r=2, 5, 10 and 13, using the Yee-Friedman solver with \u0012=1 (grey cross is reference from run\nwith no damping).\nwake, but degrade the accuracy very significantly for slower moving reference frames. The dependency of the e \u000bect of\ndamping and smoothing with \rboost has two causes. First, simulations with a boost \r\u0019\rwrequire fewer time steps\nthan simulations using a lower value of \r. Thus, for a given value of the damping coe \u000ecient\u0012, the integrated amount\nof damping will be lower for the simulations with \r\u0019\rw. Second, as mentioned above in the discussion of the surface\nrenderings shown in Fig. 9 and 10, a large fraction of the short wavelength content that is present in the simulations\nin the laboratory frame is transformed into time oscillations in simulations in the wake frame. Hence, filtering short\nwavelength has less e \u000bect on the physics when calculating in the wake frame than when calculating in the laboratory\nframe; (iii) the cost of using even the most aggressive damping or smoothing is low, especially considering that the\nsimulations presented here were using only two plasma macro-particles per cell.\nIn summary, calculating in a boosted frame near the frame following the wake ( \r\u0019\rw) relaxes the constraint on\nthe numerical dispersion in the direction of propagation of the laser (which is essential in simulations in the laboratory\nframe), and allows for more aggressive damping of high frequencies and smoothing of short wavelengths than is\npossible in standard laboratory frame calculations.\n5.2. Full scale 10 GeV class stages\nAs noted in [13], full scale simulations using the laboratory frame of 10 GeV stages at plasma densities of 1017\ncm\u00003are not practical on present computers in 2D and 3D. At this density, the wake relativistic factor \rw\u0019132, and\n2-1/2D and 3D simulations were done in boosted frames up to \r=130.\nFigure 18: (left) Average beam energy gain versus longitudinal position (in the laboratory frame), (right) Fourier Transform of the longitudinal\nelectric field at t =40 ps, averaged over whole domain, from 2D-1 /2 simulations of a full scale 10GeV LPA in a boosted frame at \r=130, using the\nYee solver and various digital filter kernels. Square cells ( \u000ex=\u000ez=6:5\u0016m) and the CFL time step (c \u000et=\u000ez=1=p\n2) were used.\n27Figure 19: Average beam energy gain versus longitudinal position (in the laboratory frame) from 2D-1 /2 simulations of a full scale 10GeV LPA in\na boosted frame at \r=30, 60 and 130, using the Yee solver.\n5.2.1. Simulations in 2-1 /2D\nFig. 18 shows the average beam energy gain versus longitudinal position and the averaged Fourier Transform of\nthe longitudinal electric field taken at t =40 ps, from 2D-1 /2 simulations of a full scale 10GeV LPA in a boosted frame\nat\r=130, using the Yee solver and various smoothing kernels. Fig. 19 shows the average beam energy gain versus\nlongitudinal position from simulations in boosted frames at \r=30, 60 and 130. All runs gave the same beam energy\nhistory within a few percents, and no sign of instability is detected in the Fourier transform plot of the longitudinal\nelectric field. The average energy gain peaks around 8 GeV , in agreement with the scaled simulations (see Fig. 13).\n5.2.2. Simulations in 3D\nIn 3D, all simulations at \r=130 using the Yee solver (using cubic cells and a time step at the CFL limit) developed\nthe instability and loss of the beam, regardless of the amount of filtering or damping that has been tried. The failure of\nthe 3D simulations using the Yee solver motivated use of the Cole-Karkkainen-Friedman (CKF) solver, with various\nlevels of filtering and damping. Data from 3D simulations using the CKF solver and various smoothing kernels are\nplotted in Fig. 20. Stability is attained when using a su \u000ecient level of filtering. Damping is detrimental to stability at\nlow levels (\u0012=0:1) but is beneficial at a higher level ( \u0012=0:5).\nNext, simulations using the solver coe \u000ecients CK2-5 from Table 1 were performed, with the time step set at their\nrespective CFL limit. The best results were obtained using solvers CK2 and CK3, while CK4 and CK5 did not o \u000ber\nsubstantial improvement over the CK solver. The results from the runs using CK2 and CK3 were nearly identical and\nhence only thoses from CK2 are reported in Fig. 21, which show very consistent beam energy gain histories, and no\nsign of instability in the Fourier Transform plot of the longitudinal electric field at t =40 ps (closer inspection revealed\nthat when using the lowest level of filtering S(1), a mild instability was developing but it was not a \u000becting the average\nbeam energy gain history). As shown on Fig. 22, the results at \r=30\u0000125 are in excellent agreement while the run\nat\r=130 predicts a slightly lower energy gain, all within 10 percent of the maximum energy gain predicted around\n5.7 GeV by the scaled simulations shown on Fig. 13 (top-right).\nIn summary, the full scale 6-7 GeV simulations using the frame of the wake performed in this subsection show:\n(i) 2-1 /2D simulations using the Yee solver at the CFL limit (with square cells) were free of instability; (ii) 3D\nsimulations using the CK solver developed moderately strong instabilities that were mitigated using moderate to high\nlevels of damping and /or filtering, the latter being the most e \u000bective; (iii) 3D simulations using the CK2 (or CK3)\nsolver developed very mild instabilities that were mitigated with a low level of filtering.\n28\u0012=0\n \u0012=0:1\n\u0012=0:5\nFigure 20: (left) Average beam energy gain versus longitudinal position (in the laboratory frame); (right) Fourier Transform of the longitudinal\nelectric field at t =40 ps, averaged over plane on axis perpendicular to laser polarization, from 3D simulations of a full scale 10GeV LPA in a\nboosted frame at \r=130, using the Cole-Karkkainen-Friedman solver and various smoothing kernels, with (top) no numerical damping ( \u0012=0),\n(middle) damping with \u0012=0:1 and (bottom) \u0012=0:5.\n29Figure 21: (left) Average beam energy gain versus longitudinal position (in the laboratory frame), (right) Fourier Transform of the longitudinal\nelectric field at t =40 ps, averaged over whole domain, from 3D simulations of a full scale 10GeV LPA in a boosted frame at \r=130, using the\nCK2 solver and various digital filter kernels.\nFigure 22: Average beam energy gain versus longitudinal position (in the laboratory frame) from 3D simulations of a full scale 10GeV LPA in a\nboosted frame at \r=30, 60, 120, 125 and 130, using the Yee solver ( \r=30 and 60) and the CK2 solver ( \r=120\u0000130), with digital filter S(1)\nand with the time step set by c \u000et=\u000ez=1=p\n2 for stability (see discussion below) .\n305.3. E \u000bects of numerical parameters on the observed instability\nFigure 23: Fourier Transform of the longitudinal electric field at t =40 ps, averaged over plane on axis perpendicular to laser polarization, from\n(left) 3D and (right) 2D-1 /2 simulations of a full scale 10GeV LPA in a boosted frame at \r=130, using the Yee solver and various smoothing\nkernels. The same time step at the 3D CFL limit c \u000et=\u000ex=p\n3 was used for both simulations.\nThe Fourier transform of the longitudinal electric field averaged over the whole domain at t =40 ps, from 3D\nsimulations using the Yee solver, is given in Fig. 23 (left). It is contrasted to the same data taken from 2-1 /2D\nsimulations (right). Both simulations used the same time step at the 3D CFL limit c \u000et=\u000ez=p\n3. The similarity of the\ntwo plots indicates that the degradation of the numerical dispersion that resulted from going from the 2D to the 3D\nCFL limit is the cause of the failure of the 3D runs using the Yee solver. Taking advantage of this observation, we\nstudy in this section the instability arising in 2-1 /2D simulations using a time step at the 3D CFL limit.\n5.3.1. E \u000bects of spatial resolution\nSnapshots of the longitudinal electric field at the front of the plasma taken at t=12:5 ps, and their corresponding\nFourier transform, are given in Fig. 24, from 2-1 /2D simulations using the Yee solver with the time step at the 3D\nCFL limit c \u000et=\u000ez=p\n3. Three resolutions were considered: (a) \u000ex=\u000ez=13\u0016m, (b)\u000ex=\u000ez=6:5\u0016m, and (c)\n\u000ex=\u000ez=3:25\u0016m. The amplitude of the instability is roughly inversely proportional to the resolution. For this\nconfiguration, the instability exhibits two primary modes at various relative levels, both at a fixed number of grid\ncells in the longitudinal direction, but at a fixed absolute length in the transverse direction. This indicates that the\ntransverse part of the modes is governed by the physical geometry of the problem while the longitudinal part is\ngoverned by numerical resolution.\nResults from 2-1 /2D simulation using the CK solver at the 3D CFL limit c \u000et=\u000ez=1=p\n3 at the resolution \u000ex=\n\u000ez=6:5\u0016m are given in Fig. 25. The same two modes that were observed in the plots from the equivalent simulation\nusing the Yee solver (see Fig. 24-middle), are present, and the overall amplitude of the instability is similar. These\nsimilarities on the details of the instability between the Yee and CK solvers indicate that the di \u000berences in numerical\ndispersion of the solvers do not constitute a key factor a \u000becting the instability.\n5.3.2. E \u000bects of time step\nIt is striking that all the solvers that lead to the lowest levels of instability had the same CFL time step c \u000etCFL=\n\u000ez=p\n2. For checking whether this is coincidental, simulations were performed using the CK solver, scanning the time\nstep between c \u000et=\u000ez=0:5 and c\u000et=\u000ez=1. The Fourier Transform of the longitudinal field averaged over the entire\ndomain taken at t=40 ps, is given in Fig. 26, exhibiting a sharp reduction of the instability level in a narrow band\naround c\u000et=\u000ez=p\n2. Since the numerical dispersion degrades in all directions when the time step diminishes, this\nindicates that the value of the time step value is of more importance than the numerical dispersion of the solver being\nused.\nSimulations using the Yee or the CK solver with the singular time step c\u000et=\u000ez=p\n2 were performed and produced\nlevels of instabilities that were much reduced (and delayed) compared to the 3D CFL time step (not shown). The\nsnapshot of the electric field and its Fourier Transform taken at t=49 ps are given in Fig. 27. The Fourier spectrum\nis very similar in each case, although the instability is slightly stronger with the CK solver than with the Yee solver.\nIn both cases, the instability is easily removed by using the S(1:2) filter (see Fig. 28).\n31As mentioned in the previous section, the solvers CK, CK4 and CK5, which all have a CFL time step above the\nsingular time step c \u000et=\u000ez=p\n2, produced significant levels of instability when run at their CFL limit. It was verified\nthat using those solvers in 3D at the time step c\u000et=\u000ez=p\n2 resulted in greatly reduced levels of instability. It was also\nobserved that running the Yee solver using non-cubic cells, e.g. with lower resolution transversely such as \u000ex=2\u000ezat\n\r=130, or\u000ex=2:6\u000ezat\r=50, produced the same pattern: a significant instability was present when using the CFL\ntime step and was greatly reduced by using c\u000et=\u000ez=p\n2. Hence for the suppression of the instability, the choice of the\nsolver seems to depends solely on whether its CFL condition allows stability at the special time step c\u000et=\u000ez=p\n2 for\na given grid cell aspect ratio, but not significantly on its numerical dispersion nor on the value of the grid cell aspect\nratio.\n5.3.3. E \u000bects of field gathering procedure\nThe scan of time step was repeated using the ’momentum conserving’ procedure [47] , in which the field values\nare interpolated at the grid nodes before being gathered onto the particles. The result is given in Fig. 29. With the\nmomentum conserving procedure, the level of instability is consistently high and independent of the time step. Since\nthe numerical dispersion of the solver varies substantially with the time step, this result supports the conclusion that\nthe instability may not be of numerical Cerenkov nature. The identification of the nature of the instability and the\nexplanation of the singular time step c\u000etScall for a multidimensional (no instability was observed in 1D regardless of\nthe field gathering method) analysis of the discretized Vlasov algorithm that was employed, which is left for future\nwork.\nThe results that were obtained lead to the following conclusions: (i) the time step c \u000etS=\u000ez=p\n2 consistently pro-\nduces the lowest levels of instability, regardless of dimensionality (2D vs 3D), the field solver being used, resolution,\naspect ratio of cells (within the range of the finite number of cases that were experimented); (ii) the main advantage\nof the tunable field solver resides in allowing access to the singular time step c\u000etSrather than providing improved\nnumerical dispersion, which consequently do not appear to be a primary driver of the instability; (iii) the instability is\nnot completely removed at c\u000etSand filtering is still needed, albeit at lower levels; (iv) the field gathering procedure\nis key, as the existence of a singular time step at which the instability is greatly reduced is observed using an ’energy\nconserving’ procedure, but not using a ’momentum conserving’ procedure. These results indicate that the instability\nthat is being observed may not be a type of numerical Cerenkov instability, as originally conjectured.\n32E== FFTE==\nFigure 24: (left) Snapshot of the longitudinal electric field ( E==) at the front of the plasma at t=12:5 ps; (right) Fourier Transform of the longitudinal\nelectric field, from 2-1 /2D simulations of a full scale 10GeV LPA in a boosted frame at \r=130, using the Yee solver, for (top) \u000ex=\u000ez=13\u0016m;\n(middle)\u000ex=\u000ez=6:5\u0016m; (bottom)\u000ex=\u000ez=3:25\u0016m. The time step at the 3D CFL limit c \u000et=\u000ez=p\n3 was used for all three simulations.\n33Figure 25: (left) Snapshot of the longitudinal electric field ( E==) at the front of the plasma at t=12:5 ps; (right) Fourier Transform of the longitudinal\nelectric field, from 2-1 /2D simulations of a full scale 10GeV LPA in a boosted frame at \r=130, with the CK solver, using \u000ex=\u000ez=6:5\u0016m, and\nthe time step at the 3D CFL limit c \u000et=\u000ex=p\n3.\nFigure 26: Fourier Transform of the longitudinal electric field at t =40 ps, averaged over the whole domain, from 2-1 /2D simulations of a full scale\n10GeV LPA in a boosted frame at \r=130, using the CK solver, for time steps between c\u000et=\u000ez=0:5 and c\u000et=\u000ez=1, versus\u0015=\u000ez(left) and at\n\u0015=\u000ez=4 (right).\n34Figure 27: (left) Snapshot of the longitudinal electric field ( E==) at the front of the plasma at t=49 ps; (right) Fourier Transform of the longitudinal\nelectric field, from 2-1 /2D simulations of a full scale 10GeV LPA in a boosted frame at \r=130, using\u000ex=\u000ez=6:5\u0016m, and the time step at the\n2D CFL limit c\u000et=\u000ez=p\n2, for (top) the Yee solver; (bottom) the CK solver.\nFigure 28: Snapshot of the longitudinal electric field ( E==) at the front of the plasma at t=49 ps from 2-1 /2D simulations of a full scale 10GeV\nLPA in a boosted frame at \r=130, using\u000ex=\u000ez=6:5\u0016m, and the time step at the 2D CFL limit c\u000et=\u000ez=p\n2, for (left) the Yee solver; (right) the\nCK solver. The filter S(1:2) was used to remove the instability that is visible in Fig. 27. The remaining feature is the wake.\n35Figure 29: Fourier Transform of the longitudinal electric field at t =40 ps, averaged over the whole domain, from 2-1 /2D simulations of a full\nscale 10GeV LPA in a boosted frame at \r=130, using the CK solver, for time steps between c\u000et=\u000ez=0:5 and c\u000et=\u000ez=1, using a ’momentum\nconserving’ field gathering scheme.\n365.4. Full scale 100 GeV - 1 TeV class stages\nFigure 30: Average beam energy gain versus longitudinal position (in the laboratory frame) for simulations at ne=1019cc down to 1015cc, using\nframes of reference between \r=12 and\r=1300, in 2-1 /2D (left) and 3D (right).\nUsing the knowledge acquired from the 10 GeV class study, simulations of stages in the range of 0.1 GeV-1 TeV\nwere performed in 2-1 /2D and in the range of 0.1-100 GeV in 3D. The plasma density nescales inversely to the energy\ngain, from 1019cc down to 1015cc in the 0.1 GeV-1 TeV range. These simulations used the parameters given in Table\n2 scaled appropriately, and used the high speed of the boosted simulations to allow fast-turnaround improvement of\nthe stage design [18, 19]. Scaled energy gain was increased by adjusting the phase of the beam injection behind the\nlaser by\u001812% in 3D and 7% in 2D, with respect to the results presented in the preceding section. The 5% level\ndi\u000berence between the 2D and 3D beam phases is likely due to small di \u000berences in wake structure, laser depletion,\nand the small number of betatron oscillations of the laser. To minimize beam loss, the beam dimensions were reduced\nby a factor of 3 in each dimension. Simulations showing performance of this design in 2-1 /2D were performed using\nthe Yee solver with filter S(1) for the 0.1-10 GeV runs, S(1:2) for the 100 GeV and S(1:2:3) for the 1 TeV ones. The\n3D simulations were performed using the CK2 solver with filter S(1) for the 0.1-1 GeV runs, and S(1:2) for the 10-100\nGeV ones. The average beam energy gain history is plotted in Fig. 30, scaling the 0.1-100 GeV runs to the 1 TeV\nrange in 2-1 /2D, and the 0.1-10 GeV runs to the 100 GeV range in 3D. The results exhibit an excellent agreement on\nthe peak scaled beam energy gain between 0.1-100 GeV runs, and on the scaled beam energy gain histories between\nthe 1-100 GeV runs. A higher level of smoothing was needed for the 1TeV case, explaining the deviation past 1 km.\nThis deviation is of little importance in practice, where one is mostly interested in the beam evolution up-to the peak\nenergy point. The di \u000berences at 1019on the scaled beam energy gain history can be attributed to the e \u000bects from\nhaving only a few laser oscillations per pulse.\nUsing (13), the speedup of the full scale 100 GeV class run, which used a boosted frame of \r=400 as frame of\nreference, is estimated to be over 100,000, as compared to a run using the laboratory frame. Assuming the use of a\nfew thousands of CPUs, a simulation that would require several decades to complete using standard PIC techniques\nin the laboratory frame, was completed in four hours using 2016 CPUs of the Cray system at NERSC. With the same\nanalysis, the speedup of the 2-1 /2D 1 TeV stage is estimated to be over a million.\n376. Conclusion and outlook\nThe technique proposed in [1] was applied successfully to speedup by orders of magnitude calculations of laser-\nplasma accelerators from first principles. The theoretical speedup estimate from [1] was improved, while complica-\ntions associated with the handling of input and output data between a boosted frame and the laboratory frame were\ndiscussed. Practical solutions were presented, including a technique for injecting the laser that is simpler and more\ne\u000ecient than methods proposed previously.\nControl of an instability that was limiting the speedup of such calculations in previous work is demonstrated, via\nthe use of a field solver with tunable coe \u000ecients and digital filtering. The tunable solver was shown to be compatible\nwith existing ”exact” current deposition techniques for conservation of Gauss Law, and accommodates Perfectly\nMatched Layers for e \u000ecient absorption of outgoing waves.\nExtensive testing of the methods presented for numerical Cerenkov mitigation reveals that choosing the frame of\nthe wake as the frame of reference allows for higher levels of filtering and damping than is possible in other frames with\nthe same accuracy. It also revealed that there exists a singular time step for which the level of instability is minimal,\nindependently of other numerical parameters, especially the numerical dispersion of the solver. This indicates that\nthe observed instability may not be caused by numerical Cerenkov e \u000bects. Analysis of the nature of the instability is\nunderway, but regardless of cause, the methods presented mitigate it e \u000bectively. The tunability of the field solver is\nkey in providing stability in 3D at the singular time step, which is not attainable by the standard Yee solver.\nThe use of those techniques permitted the first calculations in the optimal frame of 10 GeV , 100 GeV and 1\nTeV class stages, with speedups over 4, 5 and 6 orders of magnitude respectively over what would be required by\n”standard” laboratory frame calculations, which are impractical for such stages due to computational requirements.\nThese results show that the technique can be applied to the modeling of 10 GeV stages, and future work will\ninclude the e \u000bects of beam loading, plasma density ramps, as well as particle trapping in the near future. Future work\non the numerical methods include a comprehensive analysis of the instability and the existence of a singular time\nstep under certain conditions, as well as the local application of filtering, smoothing and /or mesh refinement [57, 58]\naround the front of the plasma, where the instability develops. The latter is expected to provide mitigation of the\ninstability while preserving accuracy in the core of the simulation.\n7. Appendix I: One dimensional analysis of the CK solver\nAlthough the most interesting applications of the CK solver require two or three dimensions, analysis of the\nmethod in one dimension reveals a potential issue when c\u000et=\u000ex. In one dimension (choosing x), Equations (43)-(44)\nreduce to\nByjn+1=2\ni+1=2=Byjn\u00001=2\ni+1=2+\u000et\n\u000ex\u0010\nEzjn\ni+1\u0000Ezjn\ni\u0011\n(73)\nEzjn+1\ni =Ezjn\ni+c2\u000et\n\u000ex\u0010\nByjn+1=2\ni+1=2\u0000Byjn+1=2\ni\u00001=2\u0011\n\u0000Jn\ni\n\u000f0(74)\nDue to uniform time discretization and linearity, the response of the system (73)-(74) to arbitrary distributions and\nevolutions of sources (i.e. macro-particles) can be written as the sum of its response to the excitation from a Heaviside\nfunction in time, at one location in the grid. Assuming a source term of the form Jjn\ni=H(t) where His the Heaviside\nfunction, and setting the time step at the Courant limit c\u000et=\u000ex, the system (73)-(74) produces a spurious ”odd-even”\noscillations at the Nyquist frequency, as shown in Fig. 31 (middle-left). If a sinusoidal signal oscillating at the Nyquist\nfrequency is added to the source term, the amplitude of the spurious oscillation grows linearly with time, as shown in\nFig. 31 (middle-right). The spurious oscillation is e \u000bectively suppressed in both cases by the application of a ”1-2-1”\nbilinear digital filter, as shown in Fig. 31 (bottom) . These types of filtering are of common use in Particle-In-Cell\ncodes, often repeated a prescribed number of times and followed by a compensation stage to avoid excessive damping\nof long wavelengths [47].\nThe impact of the spurious oscillations and the e \u000bectiveness of the bilinear filtering at suppressing it in actual\nsimulations was tested on a 1D simulation of a scaled wakefield acceleration stage. The physical and numerical\nparameters of the simulation are given in table 3. Snapshots of the transverse electric field (aligned with the laser\n38Heaviside step excitation oscillatory excitation\nFigure 31: (top) time history (in time steps) of the current source for (left) a Heaviside step (right) a heaviside step modulated by a sinusoidal\noscillation at the Nyquist frequency; (middle) response of the system of equations (73)-(74) via a snapshot of the electric field after 10 time steps,\nwithout filtering of the source term; (bottom) response of the system of equations (73)-(74) with application of bilinear digital filter of the source\nterm in space. A time step of c \u000et=\u000exwas used in all runs and scaled constants c=\u000f0=1 were assumed.\npolarization) and the plasma electron phase space, taken once the laser has propagated about half way through the\nplasma (after\u001820,000 time steps) are given in Fig. 32. Without filtering of the current density, an instability develops\nat the grid Nyquist frequency, severely disrupting the plasma wake, despite the fact that cubic splines were used to de-\nposit current from macro-particles to the grid and gather the electromagnetic field from the grid to the macro-particles.\nOne application of the bilinear filtering (without compensation) is su \u000ecient to suppress the spurious instability and\nproduce a steady and clean wake.\n8. Appendix II: Perfectly Matched Layer\nThe split form of Perfectly Matched Layer (PML) [52] framework applies readily to Eqs (43)-(44). The equations\non the component along zof the magnetic field are given by\n(\u0001t+\u001bx)Bzx=\u0000\u0001\u0003\nxEy (75)\u0010\n\u0001t+\u001by\u0011\nBzy= \u0001\u0003\nyEx (76)\n(\u0001t+\u001bx)Ey=\u0000c2\u0001x\u0010\nBzx+Bzy\u0011\n(77)\n\u0010\n\u0001t+\u001by\u0011\nEx=c2\u0001y\u0010\nBzx+Bzy\u0011\n(78)\nwhere\u001bxand\u001byare the absorbing layer coe \u000ecients along xandyrespectively. The equations for the other compo-\nnents of the magnetic field and for the electric field are obtained similarly, applying the standard di \u000berence operator\non the spatial derivatives of the electric field and the enlarged di \u000berence operator on the spatial derivatives of the\nmagnetic field. The formula to update the fields is obtained by solving the finite-di \u000berence equations or by integrating\nover one time step, giving\nBzxjn+1=2\ni+1=2;j+1=2;k=\u0018xBzxjn\u00001=2\ni+1=2;j+1=2;k\u00001\u0000\u0018x\n\u001bx\u0001\u0003\nxEyjn\ni+1=2;j+1=2;k (79)\n39Table 3: List of parameters for scaled 10GeV class LPA stage simulation.\nbeam length Lb 85 nm\nbeam peak density nb 1014cm\u00003\nbeam longitudinal profile exp\u0010\n\u0000z2=2L2\nb\u0011\nlaser wavelength \u0015 0:8\u0016m\nlaser length (FWHM) L 10:08\u0016m\nnormalized vector potential a0 1\nlaser longitudinal profile sin (\u0019z=L)\nplasma density on axis ne 1019cm\u00003\nplasma longitudinal profile flat\nplasma length L 1:5 mm\nplasma entrance ramp profile half sinus\nplasma entrance ramp length 4 \u0016m\nnumber of cells Nz 952\ncell size \u000ez\u0015=24\ntime step \u000et\u000ez=c\nparticle deposition order cubic\n# of plasma particles /cell 10\nBzyjn+1=2\ni+1=2;j+1=2;k=\u0018yBzyjn\u00001=2\ni+1=2;j+1=2;k+1\u0000\u0018y\n\u001by\u0001\u0003\nyExjn\ni+1=2;j+1=2;k (80)\nEyjn+1\ni;j+1=2;k=\u0018xEyjn\ni;j+1=2;k\u0000c21\u0000\u0018x\n\u001bx\u0001x\u0010\nBzx+Bzy\u0011\njn+1=2\ni;j+1=2;k(81)\nExjn+1\ni+1=2;j;k=\u0018yExjn\ni+1=2;j;k+c21\u0000\u0018y\n\u001by\u0001y\u0010\nBzx+Bzy\u0011\njn+1=2\ni+1=2;j;k(82)\nwhere\u0018=(1\u0000\u001b\u000et=2)=(1+\u001b\u000et=2)via direct solve, or \u0018=e\u0000\u001b\u000etvia time integration (note that in our tests, both\nimplementations gave nearly identical results).\nThe PML using the stencil given by (82) was tested and compared to the standard Yee implementation in 2D and\n3D. Fig. 33 snapshots from 2D simulations of the reflected residue from a PML layer of a pulse with amplitude given\nby the Harris function (10 \u000015\u0003cos(2\u0019ct=L)+6\u0003cos(4\u0019ct=L)\u0000cos(6\u0019ct=L))=32 where tis time, cis the speed\nof light and L=50\u000exis the pulse length in cell size units. A grid of 400x400 cells was used with \u000ex=\u000ey. The\nabsorbing layer was 8 cells deep and the dependency of the PML coe \u000ecients with the index position iin the layer\nwas\u001bi=\u001bm(i\u000ex=\u0001)nwith\u001bm=4=\u000ex,\u0001 = 5\u000exandn=2. The alternative prescription for the coe \u000ecients given in\n[53, 54], which reads \u001b\u0003\ni=\u0000\u0018i+1=2\u00001=\u0018i\u0001=\u000exwith\u0018i=e\u0000\u001bi\u000etand\u001bi=\u001bm(i\u000ex=\u0001)n, was also tested.\nFor the generic test case that has been considered, the new implementation exhibited a very low residue of re-\nflections from the PML layer, which are qualitatively and quantitatively very similar to the residue obtained with a\nstandard PML implementation. In agreement with results from [53, 54], the use of the modified coe \u000ecients\u001b\u0003led to\nan order of magnitude improvement over the use of the standard coe \u000ecients.\nThe 3D tests gave similar absorption e \u000eciency between the Yee and the new solver implementations of the PML,\nfor all the CK solver coe \u000ecients given in Table 1.\nIt was shown in [53, 54] that the e \u000eciency of the layer can be improved further for the standard PML by augment-\ning the equations with additional terms. However, a similar extension may not be readily available when using the\nCole-Karkkainen stencil and is not considered here.\n40No filtering bilinear filtering\nFigure 32: Snapshots from transverse electric field (normalized to maximum laser amplitude E0) and plasma electrons longitudinal phase space\nprojection, from a 1D simulation of a laser wakefield acceleration stage the CFL limits ( c\u000et=\u000ex) with (left) no filtering of current density; (right)\napplication of a bilinear digital filter to the current density.\n9. Acknowledgments\nWe are thankful to D. L. Bruhwiler, J. R. Cary, B. Cowan, E. Esarey, A. Friedman, C. Huang, S. F. Martins, W. B.\nMori, B. A. Shadwick, and C. B. Schroeder for insightful discussions.\n41Yee Cole-Karkkainen\n\u001b \u001b\u0003\n400 \n300 \n200 \n100 \n0 Y \n80 40 0\nX-50 -45 -40 -35 -30 dB \n-50 -45 -40 -35 -30 \n400 \n300 \n200 \n100 \n0 Y \n80 40 0\nX-50 -45 -40 -35 -30 dB \n-50 -45 -40 -35 -30 \u001b \u001b\u0003\n400 \n300 \n200 \n100 \n0 Y \n80 40 0\nX-50 -45 -40 -35 -30 dB \n-50 -45 -40 -35 -30 \n400 \n300 \n200 \n100 \n0 Y \n80 40 0\nX-50 -45 -40 -35 -30 dB \n-50 -45 -40 -35 -30 \nFigure 33: Reflected signal (in dB) from a PML layer using the Yee or the Cole-Karkkainen solver. Each simulation was run for the time step set\nat the Courant limit.\n42References\n[1] J.-L. Vay, Phys. Rev. Lett. 98(2007) 130405.\n[2] T. Tajima, J. M. Dawson, Phys. Rev. Lett. ,43(1979) 267.\n[3] E. Esarey, et al. ,Rev. Modern Phys. 81, 252 (2009) 1229.\n[4] C. G. R. Geddes, et al. ,Nature 431, 538 (2004).\n[5] S. P. D. Mangles, et al. ,Nature 431, 535 (2004).\n[6] J. Faure, et al. ,Nature 431, 541 (2004).\n[7] W. P. Leemans, et al. ,Nature Physics 2, 696 (2006).\n[8] C. B. Schroeder, et al. Proc. 13th Advanced Accelerator Concepts Workshop , Santa Cruz, CA (2008) 208.\n[9] C. G. R. Geddes, et al. , ”Laser Plasma Particle Accelerators: Large Fields for Smaller Facility Sources,” SciDAC Review 13(2009) 13.\n[10] C. G. R. Geddes et al. ,J. Phys. Conf. Series V 125(2008) 12002 /1-11.\n[11] C. Huang et al. ,J. Phys. Conf. Series 180(2009) 12005\n[12] http: //loasis.lbl.gov\n[13] D. L Bruhwiler et al. ,Proc. 13th Advanced Accelerator Concepts Workshop , Santa Cruz, CA (2008) 29.\n[14] B. A Shadwick, C. B. Schroeder, E. Esarey, Phys. Plasmas 16(2009) 056704\n[15] P. Sprangle, E. Esarey, and A. Ting, Phys. Rev. Letters 64(1990) 2011-2014.\n[16] C. Huang et al. ,J. of Comput. Phys. 217(2006) 658-679.\n[17] B. Feng, C. Huang, V . Decyk, W.B. Mori, P. Muggli, T. Katsouleas, J. Comput. Phys. 228(2009) 5340.\n[18] E. Cormier-Michel, et al. Proc. 13th Advanced Accelerator Concepts Workshop , Santa Cruz, CA (2008) 297.\n[19] C. G. R. Geddes et al. ,Proc. Particle Accelerator Conference , Vancouver, Canada (2009) WE6RFP075.\n[20] J.-L. Vay, Phys. Plasmas ,15(2008) 056701.\n[21] J. P. Boris, J. Comput. Phys. 12(1973) 131-136.\n[22] I. Haber, R. Lee, H. H. Klein, J. P. Boris, Proc. Sixth Conf. Num. Sim. Plasmas , Berkeley, CA (1973) 46-48.\n[23] B. Cowan, et al. Proc. 13th Advanced Accelerator Concepts Workshop , Santa Cruz, CA (2008) 309.\n[24] J.-L. Vay et al. ,Proc. Particle Accelerator Conference , Vancouver, Canada (2009) TU1PBI04.\n[25] S. F. Martins, Proc. Particle Accelerator Conference , Vancouver, Canada (2009) TH4GBC05.\n[26] J.-L. Vay et al. ,J. Phys. Conf. Series 180(2009) 12006\n[27] J. -L. Vay, W. M. Fawley, C. G. Geddes, E. Cormier-Michel, D. P. Grote, arXiv:0909.5603 (Sept. 2009)\n[28] S. F. Martins, R. A. Fonseca, L. O. Silva, W. Lu, W. B. Mori, Comput. Phys. Comm. 182(2010) 869-875.\n[29] D. L. Bruhwiler, Private Communication .\n[30] D. P. Grote, A. Friedman, J.-L. Vay, I. Haber, AIP Conf. Proc. 749(2005) 55.\n[31] T. Esirkepov, Comput. Phys. Comm. 135(2001) 144-53.\n[32] B. B. Godfrey, J. Comput. Phys. 15(1974) 504-521.\n[33] B. B. Godfrey, J. Comput. Phys. 19(1975) 58-76.\n[34] A. D. Greenwood, K. L. Cartwright, J. W. Luginsland, E. A. Baca, J. Comp. Phys. 201(2004) 665-684.\n[35] B. B. Godfrey, Proc. Ninth Conf. on Num. Sim. of Plasmas (1980).\n[36] A. Friedman, J. Comput. Phys. 90(1990) 292.\n[37] H. Abe, N. Sakairi, R. Itatani, H. Okuda, J. Comput. Phys. 63(1986) 247-267.\n[38] E. Cormier-Michel, B. A. Shadwick, C. G. R. Geddes, E. Esarey, C. B. Schroeder, W. P. Leemans, Phys. Rev. E 78(2008) 016404.\n[39] J. B. Cole, IEEE Trans. Microw. Theory Tech. ,45(1997) 991996.\n[40] J. B. Cole, IEEE Trans. Antennas Prop. ,50(2002) 11851191.\n[41] M. Karkkainen, E. Gjonaj, T. Lau, T. Weiland, Proc. International Computational Accelerator Physics Conference , Chamonix, France (2006).\n[42] K. S. Yee, IEEE Trans. Ant. Prop. 14(1966) 302-307\n[43] B. Cowan, Proc. 10thInternat. Comput. Accel. Phys. Conf. , San Francisco, CA (2009).\n[44] J.-P. B ´erenger, J. Comput. Phys. 114(1994) 185.\n[45] F. S. Tsung, et al. ,Phys. Plasmas 13(2006) 056708.\n[46] B. Cowan, Private communication .\n[47] C. K. Birdsall and A. B. Langdon, Plasma Physics Via Computer Simulation (Adam-Hilger, 1991).\n[48] A. B. Langdon, Comput. Phys. Comm. 70(1992) 447.\n[49] B. Marder, J. Comput. Phys. 68(1987) 48.\n[50] J.-L. Vay, C. Deutsch, Phys. Plasmas 5(1998) 1190.\n[51] J. Villasenor, O. Buneman, Comput. Phys. Comm. 69(1992) 306.\n[52] J.-P. Berenger, J. Comput. Phys. 114(1994) 185.\n[53] J.-L. Vay, J. Comput. Phys. 165(2000) 511.\n[54] J.-L. Vay, J. Comput. Phys. 183(2002) 367.\n[55] J.-L. Vay, et al. ,in preparation.\n[56] A. B. Langdon, C. K. Birdsall, Phys. Fluids 13(1970) 2115.\n[57] J.-L. Vay et al. ,Phys. Plasmas 11(2004) 2928.\n[58] J.-L. Vay, J.-C. Adam, A. H ´eron, Comput. Phys. Comm. 164(2004) 171.\n43" }, { "title": "1608.08373v1.Astroparticles_and_tests_of_Lorentz_invariance.pdf", "content": "arXiv:1608.08373v1 [hep-ph] 30 Aug 2016Proceedings of the Seventh Meeting on CPT and Lorentz Symmet ry (CPT’16), Indiana University, Bloomington, June 20-24, 20 16\n1\nAstroparticles and Tests of Lorentz Invariance\nJ.S. D´ ıaz\nInstitute for Theoretical Physics, Karlsruhe Institute of Technology\n76128 Karlsruhe, Germany\nSearches for violations of Lorentz invariance using cosmic rays, gamma rays,\nand astrophysical neutrinos and the prospects for future te sts using cosmic-ray\nshowers are presented.\n1. Introduction\nThe study of energetic particles bombarding Earth from distant as trophys-\nical sources has led to the development of a new discipline: astropar ticle\nphysics. The high energy and the long propagation distance of thes e as-\ntroparticles can serve as a sensitive tools to search for new physic s, as\nminute unconventional effects can get enhanced by the energy an d the path\nlength.\n2. Cosmic rays and gamma rays\nA simple modification of quantum electrodynamics (QED) is obtained by\nincorporating a Lorentz-violatingoperator that preserves CPT, coordinate,\nand gauge invariance in the form1\nL=−1\n4FµνFµν+ψ/bracketleftbig\nγµ(i∂µ−eAµ)−m/bracketrightbig\nψ−1\n4(kF)µνρσFµνFρσ,(1)\nwhere the first two terms correspond to conventional QED, while t he last\nterm is a dimension-four operator for Lorentz violation in the Stand ard-\nModel Extension (SME).1The nine independent components of the tensor\n(kF)µνρσthat produce nonbirefringent effects have been constrained usin g\nthe observation of high-energy cosmic rays.2,3In particular, the isotropic\nlimit is controlled by a single parameter κ, whose upper limit has been\nconstrained by the observation of high-energy cosmic rays, wher eas a lower\nlimit has been obtained from energetic gamma rays.4\nForκ>0,aneffectiverefractiveindexisproduced,whichallowsthepro-\nduction of Cherenkov radiation in vacuum. This is an efficient energy- lossProceedings of the Seventh Meeting on CPT and Lorentz Symmet ry (CPT’16), Indiana University, Bloomington, June 20-24, 20 16\n2\nmechanism for electrically charged fermions above a threshold ener gyEth.\nCosmic-ray primaries would rapidly lose energy and fall below the thre sh-\nold, so that the emission of Cherenkov photons would rapidly stop. H ence,\ncosmic rays reaching Earth will always have energies below the thres hold\nE −9×10−16(98% C.L.).4\nProspects for potential improvement on this lower limit could make us e of\ncosmic rays by dedicated studies of the development of extensive a ir show-\ners in the atmosphere, whose maximum is very sensitive to the value o f\nκ.6\n3. Astrophysical neutrinos\nLorentz-violating neutrinos and antineutrinos in the SME are effect ively\ndescribed by a 6 ×6 hamiltonian of the form7,8\nH=|ppp|/parenleftbigg10\n01/parenrightbigg\n+1\n2|ppp|/parenleftbiggm20\n0m2∗/parenrightbigg\n+1\n|ppp|/parenleftBigg\nˆaaaeff−ˆccceff−ˆgggeff+ˆHHHeff\n−ˆggg†\neff+ˆHHH†\neff−ˆaaaT\neff−ˆcccT\neff/parenrightBigg\n,(2)Proceedings of the Seventh Meeting on CPT and Lorentz Symmet ry (CPT’16), Indiana University, Bloomington, June 20-24, 20 16\n3\nwhereeachblockisa3 ×3matrixinflavorspace. Thefirsttwotermsappear\nin the Lorentz-invariant description of massive neutrinos of momen tumppp\ncharacterized by a mass-squared matrix m2. The last term includes all the\nmodifications introduced by Lorentz violation. Each matrix compone nt\ndepends in general on the neutrino momentum, direction of propag ation,\nand the relevant coefficients controlling Lorentz violation.8In particular,\ncoefficients arising from a dimension-three operator lead to CPT viola tion\nthat can mimic nonstandard interactions9and even the cosmic neutrino\nbackground.10\nThe effective hamiltonian (2) has been used for the formulation of\nnovel models describing global neutrino-oscillation data.11This hamilto-\nnian also has served for implementing generic searches of the key sig natures\nof Lorentz violation in experiments12using neutrino oscillations,13–23beta\ndecay,24,25and double beta decay.26,27\nAstrophysical neutrinos of very high energy have been observed by Ice-\nCube,28which can be used to determine stringent limits on coefficients for\nLorentz violation that modify the kinematics of neutrinos but are un ob-\nservable in oscillation experiments. These oscillation-free coefficient s can\nbe constrained using an approach similar to the one presented in the pre-\nvious section for cosmic rays and gamma rays. For isotropic operat ors\nof arbitrary dimension d, the relevant coefficients are denoted by ˚ c(d)\nof.8\nFor ˚c(d)\nof<0, an effective refractive index is produced for neutrinos. This\nallows the Cherenkov production of Zbosons, which rapidly decay into\nelectron-positron pairs. This is an efficient energy-loss mechanism f or neu-\ntrinos above a threshold energy Eth. Astrophysical neutrinos would rapidly\nlose energy and fall below the threshold, so that the Cherenkov em ission\nwould rapidly stop. Hence, high-energy astrophysical neutrinos r eaching\nEarth will always have energies below the threshold E 4.8Regarding the sensitive interferometric measurements\nof neutrino oscillations, the absence of an antineutrino component in the\nflux of solar neutrinos can be used to determine the most stringent limits\non Majorana couplings for CPT violation in the neutrino sector of the SME\nthat would produce neutrino-antineutrino oscillations.32Proceedings of the Seventh Meeting on CPT and Lorentz Symmet ry (CPT’16), Indiana University, Bloomington, June 20-24, 20 16\n4\nAcknowledgments\nThis work was supported in part by the German Research Foundatio n\n(DFG) under Grant No. KL 1103/4-1.\nReferences\n1. D. Colladay and V.A. Kosteleck´ y, Phys. Rev. D 55, 6760 (1997); Phys. Rev.\nD58, 116002 (1998).\n2. F.R. Klinkhamer and M. Risse, Phys. Rev. D 77, 016002 (2008).\n3. F.R. Klinkhamer and M. Risse, Phys. Rev. D 77, 117901 (2008).\n4. F.R. Klinkhamer and M. Schreck, Phys. Rev. D 78, 085026 (2008).\n5. J.S. D´ ıaz and F.R. Klinkhamer, Phys. Rev. D 92, 025007 (2015).\n6. J.S. D´ ıaz, F.R. Klinkhamer, and M. Risse, arXiv:1607.02 099.\n7. V.A. Kosteleck´ y and M. Mewes, Phys. Rev. D 69, 016005 (2004).\n8. V.A. Kosteleck´ y and M. Mewes, Phys. Rev. D 85, 096005 (2012).\n9. J.S. D´ ıaz, arXiv:1506.01936.\n10. J.S. D´ ıaz and F.R. Klinkhamer, Phys. Rev. D 93, 053004 (2016).\n11. V.A. Kosteleck´ y and M. Mewes, Phys. Rev. D 70, 031902 (2004); T. Katori\net al., Phys. Rev. D 74, 105009 (2006); V. Barger et al., Phys. Lett. B 653,\n267 (2007); J.S. D´ ıaz and V.A. Kosteleck´ y, Phys. Lett. B 700, 25 (2011);\nPhys. Rev. D 85, 016013 (2012).\n12.Data Tables for Lorentz and CPT Violation, V.A. Kosteleck´ y and N. Russell,\n2016 edition, arXiv:0801.0287v9.\n13. V.A. Kosteleck´ y and M. Mewes, Phys. Rev. D 70, 076002 (2004).\n14. J.S. D´ ıaz, V.A. Kosteleck´ y, and M. Mewes, Phys. Rev. D 80, 076007 (2009).\n15. L.B. Auerbach et al., Phys. Rev. D 72, 076004 (2005).\n16. P. Adamson et al., Phys. Rev. Lett. 101, 151601 (2008); Phys. Rev. Lett.\n105, 151601 (2010); Phys. Rev. D 85, 031101 (2012).\n17. R. Abbasi et al., Phys. Rev. D 82, 112003 (2010).\n18. T. Katori, Mod. Phys. Lett. A 27, 1230024 (2012).\n19. Y. Abe et al., Phys. Rev. D 86, 112009 (2012).\n20. A.A. Aguilar-Arevalo et al., Phys. Lett. B 718, 1303 (2013).\n21. B. Rebel and S. Mufson, Astropart. Phys. 48, 78 (2013).\n22. J.S. D´ ıaz et al., Phys. Lett. B 727, 412 (2013).\n23. K. Abe et al., Phys. Rev. D 91, 052003 (2015).\n24. J.S. D´ ıaz, V.A. Kosteleck´ y, and R. Lehnert, Phys. Rev. D88, 071902 (2013).\n25. J.S. D´ ıaz, Adv. High Energy Phys. 2014, 305298 (2014).\n26. J.S. D´ ıaz, Phys. Rev. D 89, 036002 (2014).\n27. J.B. Albert et al., Phys. Rev. D 93, 072001 (2016).\n28. M.G. Aartsen et al., Science 342, 1242856 (2013).\n29. J.S. D´ ıaz, V.A. Kosteleck´ y, and M. Mewes, Phys. Rev. D 89, 043005 (2014).\n30. J.S. D´ ıaz, Adv. High Energy Phys. 2014, 962410 (2014).\n31. C.A. Arg¨ uelles et al., Phys. Rev. Lett. 115, 161303 (2015).\n32. J.S. D´ ıaz and T. Schwetz, Phys. Rev. D 93, 093004 (2016)." }, { "title": "1202.3131v1.Lorentz_violating_vs_ghost_gravitons__the_example_of_Weyl_gravity.pdf", "content": "arXiv:1202.3131v1 [gr-qc] 14 Feb 2012YITP-12-3\nLorentz-violating vsghost gravitons: the example of Weyl gravity\nNathalie Deruelle,1Misao Sasaki,2Yuuiti Sendouda,3,1,2and Ahmed Youssef4\n1APC, CNRS-Universit´ e Paris 7, 75205 Paris CEDEX 13, France\n2Yukawa Institute for Theoretical Physics, Kyoto Universit y, Kyoto 606-8502, Japan\n3Graduate School of Science and Technology, Hirosaki Univer sity, Hirosaki, Aomori 036-8561, Japan\n4Institut f¨ ur Mathematik und Institut f¨ ur Physik,\nHumboldt-Universit¨ at zu Berlin, 12489 Berlin, Germany\n(Dated: November 2, 2018)\nWe show that the ghost degrees of freedom of Einstein gravity with a Weyl term can be eliminated\nby a simple mechanism that invokes local Lorentz symmetry br eaking. We demonstrate how the\nmechanism works in a cosmological setting. The presence of t he Weyl term forces a redefinition of\nthe quantum vacuum state of the tensor perturbations. As a co nsequence the amplitude of their\nspectrum blows upwhen the Lorentz-violatingscale becomes comparable tothe Hubbleradius. Such\na behaviour is in sharp contrast to what happens in standard W eyl gravity where the gravitational\nghosts smoothly damp out the spectrum of primordial gravita tional waves.\nPACS numbers: 04.50.Kd, 04.62.+v, 04.30.-w, 98.80.Cq\nI. INTRODUCTION\nFour-dimensional gravity theories based on action integrals which a re non-linear in the curvature invariants have\nbeen part of the landscape of fundamental physics since Weyl [1] in troduced them, soon after Einstein’s invention of\nGeneral Relativity.\nSuch theories possess, apart from the einsteinian ones, new degr ees of freedom (dofs) because the action contains\nterms which are non-linear in the second derivatives of the metric. I t is commonly believed that, except in f(R)\ntheories, see e.g. [2], these new dofs will render even flat spacetime unstable, because they are ghosts, that is, dofs\nwith kinetic terms with the “wrong” sign, which implies that the phase s pace of the whole system is no longer finite\nand its energy is unbounded from below. The reason why that is cons idered dangerous is that any kind of interaction\nterm in the action is expected to yield equations of motion whose solut ions are erratic, unstable and eventually diverge\n[3].\nThe simplest representative of such pathological theories is Weyl g ravity whose action is Einstein-Hilbert’s supple-\nmented by a Weyl-squared term:1\nS[gab] =1\n2κ/integraldisplay\nd4x√−g/parenleftBig\nR−γ\n2CabcdCabcd/parenrightBig\n. (1.1)\nIt was analysed by Stelle [4], who showed the existence of ghosts (no n-tachyonicif γ >0) but did not exhibit explicitly\ntheir malignancy since he worked at linear level, where the dofs do not interact.\nIn [5], we analysed linear cosmological perturbations generated dur ing inflation in Weyl gravity, restricting however\nour attention, as in [6], to the case when the Weyl length scale√γis shorter than the Hubble scale. We found that,\ndespite the fact that no interactions between the dofs were take n into account at this linear level, still the scalarmodes\ndiverge in the newtonian gauge, although they remain bounded in the comoving slicing. It therefore seems that the\nmalignancy of ghosts shows up already at linear level, when the backg round is richer than Minkowski spacetime.\nRealising a potential difficulty associated with Weyl gravity already at linear level, we present in the present paper\na drastic way out of the potential “horrors” of its ghosts. Followin g the lines of [7] and [8] in their covariant version\nof Hoˇ rava gravity [9] we introduce a scalar field that breaks local L orentz covariance (in defining a preferred time\ndirection) and couple it to the Weyl tensor in a way that eliminates its g hosts. However Lorentz-violation will imply a\nmodification of the dispersion relation of the remaining, einsteinian, d ofs. Specialising to an inflationary background\nwe study tensor perturbations and their spectrum (vector and s calar perturbations are not affected by the presence\nof our Weyl term) that we compare and contrast to the behaviour of tensor modes and their spectrum in standard\nWeyl gravity (1.1).\n1Units:c= 1;κ= 8πG,γhas dimension length2,κhas dimension length /mass,/planckover2pi1=κM2\nPl. Indices a,b,···run from 0 to 3; i,j,···\nrun from 1 to 3; metric gabwith inverse gab, determinant gand signature ( −,+,+,+).Rabcd=∂cΓabd−∂dΓabc+···with Γabcthe\nChristoffel symbols; Rbd=Rabad;R=gabRab;Gab=Rab−1\n2gabRandCabcd=Rabcd−1\n2(gacGbd−gadGbc−gbcGad+gbdGac)−\nR\n3(gacgbd−gadgbc).2\nThe organisation of the paper is as follows. In section II, we presen t our model. In section III we show, in a\ncosmological setting, how it eliminates the Weyl ghosts. Section IV d escribes the behaviour of tensor perturbations\non a de Sitter cosmological background. In section V, we quantise t hese tensor perturbations to obtain the spectrum\nof primordial gravitational waves. Section VI develops the analysis of [6] of tensor perturbations in standard Weyl\ngravity when ghost dofs are present. Section VII discusses and s ummarises our results.\nII. THE MODEL\nOur gravity model is specified by the action\nS[gab,χ] =1\n2κ/integraldisplay\nd4x√−g/parenleftbig\nR+2γCabcdCefghγaeγbfγcguduh/parenrightbig\n+Sχ[gab,χ], (2.1)\nwhere\nua≡∂aχ√−∂aχ∂aχandγab≡gab+uaub. (2.2)\nNote that the Weyl-term is conformally invariant, see appendix A, wh ere its relation to the Weyl-square scalar is also\ngiven. For our purpose the specific form of Sχ[gab,χ] for the scalar field χneed not be specified, but is requested to be\nsuch that it yields solutions in which the gradient vector ∂aχis everywhere timelike and future-directed. The vector\nfielduathen determines a preferred time direction and that necessarily imp lies that the theory breaks local Lorentz\ncovariance. Indeed, the spacetime solution is then preferentially f oliated by a family of spacelike hypersurfaces {Σχ}\non each of which χtakes a constant value. Viewed geometrically, uais the future-directed unit normal to Σ χandγab\nis the induced metric on Σ χ. The scalar field χcan be called a chronon as in [7]. Note that it can act as an inflaton\nand drive cosmological inflation.\nThe equations of motion (eoms) for the metric extremise the action with respect to perturbations of the metric and\nread\nGab−γBab=κTχ\nab, (2.3)\nwhereGabis the Einstein tensor, Babis an analogue of the Bach tensor whose expression is given in append ix B, and\nTab\nχ≡2(−g)−1/2δSχ/δgabis the energy-momentum tensor of χ. The eom for χ, on the other hand, is\n1√−gδSχ\nδχ−γ\nκ∇aWa= 0, (2.4)\nwhere∇aWa, the variational derivative with respect to χof the Weyl part of the action, is also given in appendix B.\nAs one can see from the expression of Babin appendix B, the Einstein equation (2.3) contains the derivatives of\nthe metric up to fourth order. One may hence worry that the theo ry should be pathological because higher order\nderivatives theories a priori induce ghosts.\nHowever, as a closer examination shows, see its expression in appen dix B,Babcontains only time derivatives up to\nsecond order if uais timelike. Since ghosts are induced by the presence of timederivatives higher than the second in\nthe eoms, our theory may therefore be free of ghosts. Note tha t our Weyl action does not give rise either to timelike\nderivatives higher than second in the chronon eom (2.4) because th e divergence term only contains spacelike derivative\nof the spacelike vector Wa. The reason why the field equations are not fourth order in time der ivatives is due to\nthe presence of the timelike vector ua, which allows to distinguish time and space derivatives, and thus brea ks local\nLorentz covariance.\nIII. GHOST ELIMINATION BY LORENTZ VIOLATION\nLet us first show explicitly and in a specific case that the eoms are inde ed second order in time.\nWeshallconsidertheexamplewhentheconstant χsurfacesareintrinsicallyflatandtakeaflatFriedmann-Lemaˆ ıtre-\nRobertson-Walker (FLRW) spacetime\ngabdxadxb=a2(η)(−dη2+δijdxidxj) (3.1)\nas a background solution, where ηis conformal time.3\nNote that because of conformal flatness of FLRW spacetimes the Weyl-squared term in the action does not affect\nthe background Friedmann equation for the scale factor a(η) since, then, BabandWavanish, see appendix B; its\nevolution is therefore the same as in Einstein gravity coupled to a sca lar fieldχ=χ(η).\nMetric perturbations about a flat FLRW spacetime are decomposed into scalar, vector and tensor parts as\nδgabdxadxb=a2/bracketleftbig\n−2Adη2+2(∂iB+Bi)dηdxi+(2Cδij+2∂i∂jE+∂iEj+∂jEi+hij)dxidxj/bracketrightbig\n,(3.2)\nwhere∂iBi=∂iEi= 0 and where ∂ihij=hii= 0 (all spatial indices being raised with δij). The following variables\nare gauge-invariant [10]:\nΨ≡A+(B−E′)′+H(B−E′),Φ≡C+H(B−E′),Ψi≡Bi−E′\ni, hij, (3.3)\nwhere the Hubble parameter is defined by H ≡a′/aand where hereafter a prime denotes derivative with respect to\nη.\nThe perturbation to second-order of the action of General Relat ivity with a scalar field can be expressed in terms of\ngauge-invariant perturbations of scalar, vector, and tensor ty pes [10]. The perturbation of our higher-curvature term\nis found to depend on the gauge-invariant tensor and vector varia bles,hijand Ψ i, but not on the scalar variables. It\nreads, see [5] and appendix A,\n(2)SW2[hij,Ψi] =/integraldisplay\ndηd3x√−g(1)Cabcd(1)Cefghγaeγbfγcguduh\n=/integraldisplay\ndηd3x(1)Cijk0(1)Cijk\n0|a(η)=1\n=1\n2/integraldisplay\ndηd3x/parenleftbigg\n∂kh′\nij∂kh′ij+1\n2△Ψi△Ψi/parenrightbigg\n,(3.4)\nwhere△ ≡∂i∂i.\nThis is where the role of the chronon is unveiled; it was so coupled to th e Weyl tensors that the perturbation of the\naction only contains first order time derivatives, whereas the othe r possible combinations, CabcdCefghγaeγbfγcgγdh\nandCabcdCefghγaeubufγcguduh, are quadratic in second order time derivatives, see appendix A.\nIt is clear that the vector perturbations are nondynamical, as no t ime derivatives of Ψ iappear in (3.4). Moreover\nthey are constrained to vanish just as in pure Einstein gravity.\nThe total action for the remaining, tensor, perturbations is (see [10, 11] for the obtention of the Einstein contribu-\ntion):\nST[hij] =1\n8κ/integraldisplay\ndηd3x/bracketleftbig\na2(h′\nijh′ij−∂khij∂khij)+4γ∂kh′\nij∂kh′ij/bracketrightbig\n, (3.5)\nand the corresponding eoms are\n/parenleftbigg\n1−4γ△\na2/parenrightbigg\nh′′\nij+2Hh′\nij−△hij= 0. (3.6)\nWe hence see explicitly that, asannounced, this action containsonly Einstein’s gravitonsas dofs and that the resulting\neoms are second order in time derivatives.\nLet us now check that these tensorial perturbations are not gho sts. To do so we decompose them, as usual, into\ntwo orthogonal polarisations and go to Fourier space:\nhij(η,/vector x) =/summationdisplay\nλ=1,2/integraldisplayd3k\n(2π)3/2eλ\nij(/vectork)hλ\n/vectork(η)ei/vectork·/vector x, (3.7)\nwhere the two polarisation tensors eλ\nij(/vectork), such that eλ\nij(/vectork)eij\nλ′∗(/vectork) =δλ\nλ′, are transverse and traceless, kjeλ\nij(/vectork) =\neλii(/vectork) = 0, and where the reality conditions,\neλ\nij(/vectork) =eλ\nij∗(−/vectork) and hλ\n/vectork(η) =hλ\n−/vectork∗(η), (3.8)\nare imposed, a star denoting complex conjugation. Then the action (3.5) reads, in Fourier space\nST[{hλ\n/vectork}] =1\n8κ/summationdisplay\nλ=1,2/integraldisplay\ndηd3k/bracketleftBig\n(a2+4γk2)|h′λ\n/vectork|2−a2k2|hλ\n/vectork|2/bracketrightBig\n, (3.9)4\nwherek=|/vectork|. The gravitoncan thus be viewed as a collection of non-interacting s calarfields as in General Relativity.\nThe important point to note here is that γmust be positive, otherwise the graviton modes would be tachyonic g hosts\non Minkowski spacetime when a(η) = 1. Therefore ( a2+4γk2) is always positive, so that the kinetic terms |h′λ\n/vectork|2in\n(3.9) are always positive, and the graviton never becomes a ghost o n a FLRW background.\nIV. EVOLUTION OF TENSOR PERTURBATIONS IN INFLATIONARY COSM OLOGY\nWhat is to be checked now is whether or not the Weyl term modifies dr astically the time evolution of the modes.\nThe eom in Fourier space deduced from (3.6) and (3.9) is:\nχ′′\n/vectork+Ω2\nkχ/vectork= 0 with χ/vectork(η)≡/radicalbig\na2+4γk2h/vectork(η), (4.1)\nwhere from now on we omit the index λand where the pulsation Ω k(η) is given by2\nΩ2\nk(η)≡a2/bracketleftbiggk2−H2−H′\na2+4γk2−4γk2H2\n(a2+4γk2)2/bracketrightbigg\n. (4.2)\nWe shall first discuss the time evolution of the modes on a de Sitter ba ckground. Setting\na=1\n−Hηandz≡ −kη, (4.3)\nHbeing a constant, the eom reduces to\nd2χ/vectork\ndz2+z2−2+4ǫ2z2(z2−3)\nz2(1+4ǫ2z2)2χ/vectork= 0 with χ/vectork=k√\n1+4ǫ2z2\nH zh/vectork, (4.4)\nwhere we have introduced the dimensionless parameter\nǫ≡√γH . (4.5)\nEquation (4.4) can be solved exactly in terms of hypergeometric fun ctions, see next section. It is however enlightening\nto study first the qualitative behaviour of the modes.\nAs in General Relativity, a mode with wave number kbecomes larger than the horizon at late times, when z≪1,\nthat is when its physical wavelength a/k= 1/(Hz) becomes much larger than the Hubble radius H−1. As for the\nLorentz violating regime it holds at early times when ǫz≫1, that is when the physical wavelength a/kof the mode\nkis still much shorter than the length scale√γset by the Weyl term.\nLet us first consider the case ǫ≪1, when the length scale√γon which the Lorentz-violating Weyl term operates\nis much shorter than the Hubble radius H−1. (This is the case if the Weyl correction is viewed as a low energy limit\nof some Planck scale quantum theory of gravity, inflation happening at, say, the grand unified theory (GUT) scale.)\nThe modes χ/vectork, which solve (4.4), will then go through three different regimes: the early, Lorentz violating regime,\nwhenǫz≫1; an intermediate regime, 1 ≫ǫz≫ǫwhen they have left the Lorentz violating period but are still\nsub-horizon; and the late stage, z≪1 when they have exited the Hubble scale.\nIn the late and intermediary regimes, when ǫz≪1, which are Lorentz symmetric, equation (4.4) is the same as in\nGeneral Relativity:\nd2χ/vectork\ndz2+/parenleftbigg\n1−2\nz2/parenrightbigg\nχ/vectork≃0, (4.6)\nwhose two independent solutions are well-known: after horizon cro ssing, that is for z≪1, we have χ(g)\n/vectork∝1/zand\nχ(d)\n/vectork∝z2, so that the growing tensor perturbation behaves as h/vectork∝zχ(g)\n/vectork→const., i.e., “freezes out”; and on\nsub-horizon scales z≫1 (butz≪1/ǫ) the modes χ/vectorkoscillate as sin zand cosz.\n2The analysis of the modes can also be performed in terms of cos mic time tsuch that dt=a(η)dη, in which case the eom reads\n¨fk+ω2\nkfk= 0 with fk≡a3/2/radicalbig\n1+4γ(k/a)2hk, where a dot denotes derivative with respect to cosmic time, whereH≡˙a/aand\nwhere\nω2\nk=−1\n4(H2+2˙H)+(k/a)2−(2H2+˙H)\n1+4γ(k/a)2−4γ(k/a)2H2\n[1+4γ(k/a)2]2.5\n510 50100 500z\n/Minus4/Minus2246810Χk/OverRVector/LParen1g/RParen1\n2Εz\n510 50100 500z\n/Minus4/Minus2246810Χk/OverRVector/LParen1d/RParen1\n2Εz\nFIG. 1. The growing and decaying mode functions χ(g)\n/vectork(z) andχ(d)\n/vectork(z) on a de Sitter background, with z=−kη=k/(aH), for\nǫ≡√γH≪1, that is, when the energy scale κM2\nPl/√γof the Weyl-term is much higher than the inflationary scale κM2\nPlH.\nIn the above figures, where ǫ= 0.01 and time increases from right to left, the Lorentz violati ng period ends at around z= 100\nand Hubble radius crossing occurs at around z= 1.\nIn the, early, Lorentz-violating regime, ǫz≫1,χ/vectork∝h/vectork, and the eom (4.4) reduces to\nd2χ/vectork\ndz2+χ/vectork\n4ǫ2z2≃0, (4.7)\nwhose solutions oscillate as χ/vectork∝z1/2±iν/2, withν≡/radicalbig\n1/ǫ2−1 real since ǫ <1. See figure 1.\nLet us now consider the case ǫ≫1, when the length scale on which the Lorentz-violating Weyl term op erates is\nmuchlongerthan the Hubble radius H−1. (This would be the case if the Weyl correction was emerging from so me\neffective theory operating below the inflationary, GUT, energy sca le.)\nAt late times, ǫz≪1≪ǫthe modes behave as before and as in General Relativity: Ω2\nk≃ −2/z2so thatχ(g)\n/vectork∝1/z\nandχ(d)\n/vectork∝z2. In the intermediary stage on the other hand, when 1 ≪ǫz≪ǫ, they are already in the Lorentz-\nviolating stage and since, then, Ω2\nk≃ −3/(4ǫ2z4), they behave as χ/vectork∝ze±√\n3/(2ǫz)and do not oscillate. But the\nreally important difference with the previous caseoccurs at veryea rly times when ǫz≫ǫ≫1. We still have then that\nΩ2\nk≃1/(4ǫ2z2) but the modes, instead of oscillating, behave as χ/vectork∝z1/2±¯ν/2with ¯ν=/radicalbig\n1−1/ǫ2real. Because they\nnever oscillate we can qualify these modes as “rampant”, that is, as “flourishing and spreading unchecked” (Oxford\ndictionary).\nWe have up to now approximated inflation by a de Sitter stage. In the case of slow-roll inflation, the Hubble\nparameter His no longer constant but slowly decreases with cosmic time, i.e. with de creasing z. The previous results\nthen hold when the parameter ǫ=√γHslowly decreases.\nThus, if the energy scale at which inflation starts is lower than the en ergy scale set by the Weyl term, in other\nwords if the Lorentz-violating length scale is always shorter than th e inflationary Hubble radius ( ǫ≡√γH <1), then\nall modes behave as in figure 1. If now inflation starts when the Hubb le radius is shorter than the Lorentz-violating\nscale, then the modes whose wavelengths are big enough to exit the Hubble scale when ǫis still bigger than 1 will\nnever oscillate. On the other hand shorter wavelengths modes, wh ich will still be within the Hubble radius when ǫ\ngoes through the critical value 1 will start in a non-oscillatory way at early times and then will behave qualitatively\nas in figure 1, once ǫbecomes smaller than 1.\nLet us summarise sections III and IV: we confirmed that our Loren tz-violating Weyl action (2.1), when considered\nin a cosmological setting, indeed yields second order equations of mo tion for the tensor perturbations (which are\nthe only ones to be modified by the presence of the Weyl term), see (3.6); we confirmed also that, for γ >0, the\nperturbations neverbecome ghost-likedespite the modification of the coefficient ofthe kinetic term of the gravitondue\nto the Lorentz-violatingWeyl term, see (3.9); finally we studied the perturbation modes, saw that they differ markedly\nfrom standard Einstein-de Sitter gravitons in the distant past, an d distinguished two different early time behaviours\naccording to whether the Lorentz-violating scale is shorter or long er than the Hubble scale. These modifications\nof the early time behaviour of the modes will induce a significant chang e in the spectrum of quantized primordial\ngravitational waves, as we shall now see.6\nV. QUANTISATION OF TENSOR PERTURBATIONS AND PRIMORDIAL GRA VITATIONAL WAVE\nSPECTRUM\nLet us for clarity write again the action (3.5), for the tensor pertu rbations on a FLRW background with scale factor\na(η):\nST[hij] =1\n8κ/integraldisplay\ndηd3x/bracketleftbig\na2(h′\nijh′ij−∂khij∂khij)+4γ∂kh′\nij∂kh′ij/bracketrightbig\n.\nThe momentum conjugate to hijis\nπij=1\n4κ(a2h′ij−4γ△h′ij). (5.1)\nCanonical quantisation turns hijandπijinto operators satisfying the standard commutation relation\n[ˆhij(η,/vector x1),ˆπij(η,/vector x2)] = 2i/planckover2pi1δ(/vector x1−/vector x2), (5.2)\nall other commutators being zero (the factor 2 comes from the fa ct thatˆhijis the sum of two independent dofs).\nExpanding ˆhijin Fourier modes as in (3.7), we further decompose ˆhλ\n/vectorkas\nˆhλ\n/vectork(η) = ˆaλ\n/vectorkhk(η)+ˆ¯aλ\n−/vectork¯hk(η), (5.3)\nwhere ˆaλ\n/vectorkandˆ¯aλ\n/vectorkare arbitrary operators (we introduce ˆ¯aλ\n−/vectorkfor later convenience) and where hk(η) and¯hk(η) are two\nindependent solutions, depending on k=|/vectork|only, of the second order eoms in Fourier space deduced from (3.6) , that\nis:\n/parenleftbigg\n1+4γk2\na2/parenrightbigg\nh′′\nk+2Hh′\nk+k2hk= 0,/parenleftbigg\n1+4γk2\na2/parenrightbigg\n¯h′′\nk+2H¯h′\nk+k2¯hk= 0. (5.4)\nIt follows from the eoms (5.4) that the Wronskian of hk(η) and¯hk(η),\nW(hk,¯hk)≡hk(a2+4γk2)¯h′\nk−¯hk(a2+4γk2)h′\nk, (5.5)\nis a constant ∝ne}ationslash= 0 ifhk(η) and¯hk(η) are independent solutions of (5.4). Hence\nˆhij(η,/vector x) =/summationdisplay\nλ=1,2/integraldisplayd3k\n(2π)3/2/bracketleftBig\neλ\nij(/vectork)ˆaλ\n/vectorkhk(η)ei/vectork·/vector x+eλ\nij∗(/vectork)ˆ¯aλ\n/vectork¯hk(η)e−i/vectork·/vector x/bracketrightBig\n, (5.6)\nwhere we used the reality condition on the polarisation tensors (3.8) . (Note that we have not yet implemented the\nhermiticity condition on ˆhλ\n/vectork(η).)\nPlugging the expansion (5.6) into (5.1), we find that the commutation relations (5.2) are equivalent to\n[ˆaλ\n/vectork1,ˆ¯aλ\n/vectork2] =δ(/vectork1−/vectork2) (5.7)\n(all other commutators being zero) if the Wronskian of the two inde pendent mode functions hkand¯hkis normalised\nto\nW(hk,¯hk)≡hk(a2+4γk2)¯h′\nk−¯hk(a2+4γk2)h′\nk= 4κ/planckover2pi1i. (5.8)\nIn Minkowski space we have a= 1 and the eoms (5.4) for the two independent solutions hk(η) and¯hk(η) reduce to\n¨hk+ω2\nMkhk= 0,¨¯hk+ω2\nMk¯hk= 0 with ω2\nMk=k2\n1+4γk2, (5.9)\nwhereω2\nMkis positive and where, for clarity, a dot denotes derivation with resp ect to cosmic time tsuch that\ndt=a(η)dη.\nThe “positive frequency” modes, which will define the vacuum |0∝an}b∇acket∇i}htsuch that ˆ a/vectork|0∝an}b∇acket∇i}ht= 0 and ˆ¯a†\n/vectork|0∝an}b∇acket∇i}ht= 0, are chosen\nto be\nhk=nke−iωMkt,¯hk= ¯nke+iωMkt, (5.10)7\nwhere the coefficients nkand ¯nkmust be such that the hermiticity and Wronskian conditions, see (3.8 ) and (5.8), are\nsatisfied, that is, such that\nˆaλ\n/vectorknk=ˆ¯aλ\n/vectork†¯n∗\nk, nk¯nk=2κ/planckover2pi1ωMk\nk2, (5.11)\nwhich impose (up to irrelevant constants)\nˆaλ\n/vectork=ˆ¯aλ\n/vectork†, nk= ¯n∗\nk=√2κ/planckover2pi1ωMk\nk. (5.12)\nTherefore, the choice for the modes is\nhk=¯h∗\nk=√2κ/planckover2pi1ωMk\nke−iωMkt. (5.13)\nNote for further reference that in the short wavelength limit, γk2→ ∞, the pulsation becomes independent of k:\nωMk→(4γ)−1/2, so that we have\nhk→√\nκ/planckover2pi1\nkγ1/4e−it/(2√γ). (5.14)\nNext we take accelerated cosmological expansion into account, wh ere another scale H, the Hubble parameter,\ncomes into play.\nIn the de Sitter case when a= 1/(−Hη) withH= const. and ηconformal time, the eoms for the modes as given\nin (5.4) read\n(1+4y2)d2hk\ndy2−2\nydhk\ndy+hk\nǫ2= 0,(1+4y2)d2¯hk\ndy2−2\nyd¯hk\ndy+¯hk\nǫ2= 0, (5.15)\nwhere\nǫ≡√γHandy≡ −ǫkη. (5.16)\nAs for the Wronskian normalisation condition (5.8) it becomes\nhkd¯hk\ndy−¯hkdhk\ndy=−i4κ/planckover2pi1H2\nk3ǫ3y2\n1+4y2. (5.17)\nNow, two independent solutions of (5.15) are\nh(g)(y) =1\n2F/parenleftbigg−1−iν\n4,−1+iν\n4,−1\n2;−4y2/parenrightbigg\n, h(d)(y) =32\n3y3F/parenleftbigg5+iν\n4,5−iν\n4,5\n2;−4y2/parenrightbigg\n(5.18)\nwith\nν≡/braceleftBigg/radicalbig\n1/ǫ2−1 if 0 < ǫ <1\ni/radicalbig\n1−1/ǫ2ifǫ >1. (5.19)\nNote that the hypergeometric functions introduced in (5.18) are r eal, whether νis real or imaginary, and that\nh(g)(0) =1\n2, h(d)(0) = 0. (5.20)\nIn order now to determine which combinations of h(g)andh(d)we must choose as our independent modes hkand\n¯hkwe have to study the early-time, large klimit.\nWheny→ ∞we have\nh(g)(y)→c−\n(g)(2y)1/2+iν/2+c+\n(g)(2y)1/2−iν/2, h(d)(y)→c−\n(d)(2y)1/2+iν/2+c+\n(d)(2y)1/2−iν/2(5.21)\nwith, when νis real,\nc+\n(g)=c−\n(g)∗=−√πΓ(−iν/2)\nΓ2(−1/4−iν/4), c+\n(d)=c−\n(d)∗=√πΓ(−iν/2)\nΓ2(5/4−iν/4). (5.22)\n(Whenνis imaginary then the coefficients are real but unequal.)8\nIncosmictime t, suchthat dt=adη, wehave a= eH tandH y=kǫe−H t, sothat, for νreal(i.e.for0 < γH2<1):\ny1/2+iν/2∝e−H t/2e−iωdStwith ωdS=/radicalbig\n1−γH2\n2√γ, (5.23)\nwhich, when γH2≪1, identifies with the Minkowski positive high frequency modes chose n in (5.10):\ny1/2+iν/2∝e−it/(2√γ). (5.24)\nWe recover here what we have discussed in the previous section, to wit that the early time behaviour of the modes\ndiffers drastically depending on the value of γH2: they oscillate if νis real (γH2<1), and are “rampant” if νis\nimaginary ( γH2>1). It is clear that quantisation will make (easy) sense only when mod es can qualify as “positive\nfrequency ”, which implies that they must oscillate. We shall therefore suppose from now on that\n0< γH2<1⇐⇒0< ǫ≡√γH <1⇐⇒ν≡/radicalbig\n1/ǫ2−1 real. (5.25)\nIn keeping to what we chose when the background is Minkowski spac etime, we shall hence impose that the two\nindependent solutions hkand¯hkbehave, for large y, as\nhk(y)→nk(2y)1/2+iν/2,¯hk(y)→¯nk(2y)1/2−iν/2. (5.26)\nAs for the values of the coefficients nkand ¯nkthey follow from the large ylimit of Wronskian normalisation condi-\ntion (5.17) as well as the hermiticity condition, see (3.8), which impose s, up to irrelevant constant that\nˆaλ\n/vectork=ˆ¯aλ\n/vectork†, nk= ¯n∗\nk=/radicalbigg\nκ/planckover2pi1\n2νk3ǫ3H . (5.27)\nThus, all in all, the hermitian operator ˆhijreads\nˆhij(η,/vector x) =/summationdisplay\nλ=1,2/integraldisplayd3k\n(2π)3/2/bracketleftBig\neλ\nij(/vectork)ˆaλ\n/vectorkhk(η)ei/vectork·/vector x+h.c./bracketrightBig\n, (5.28)\nwhere the modes are:\nhk=¯h∗\nk=−i/radicalbigg\nκ/planckover2pi1ν\n8k3ǫ3H(c+\n(d)h(g)−c+\n(g)h(d)), (5.29)\nwhere the functions h(g)andh(d)and the coefficients c(g)andc(d)are defined in (5.18) and (5.22) (and where we used\nthe fact that c+\n(d)c−\n(g)−c−\n(d)c+\n(g)= 2i/ν). Finally the “Bunch-Davies” vacuum |0∝an}b∇acket∇i}htis defined as usual by\nˆaλ\n/vectork|0∝an}b∇acket∇i}ht= 0. (5.30)\nThe power spectrum P(k;η) of the gravitational waves hijis now defined as\n∝an}b∇acketle{t0|ˆhij(η,/vector x1)ˆhij(η,/vector x2)|0∝an}b∇acket∇i}ht=/integraldisplay\nd3kP(k;η)\n4πk3ei/vectork·(/vector x1−/vector x2)(5.31)\nand is given by, using the commutation rule (5.7):\nP(k;η) =k3\nπ2|hk(η)|2. (5.32)\nAfter horizon crossing, that is in the limit η→0, we have hk∝k−3/2, see (5.29), and the power spectrum is scale\ninvariant just like the power spectrum of ordinary gravitational wa ves in Einstein theory on a de Sitter background.\nHowever its amplitude depends on ν=/radicalbig\n1/ǫ2−1 and is, using the late time limit of h(g)andh(d), see (5.20)\nP(k;η→0)≡2κH2\nπ2Ξ,where Ξ =cosh(πν/2) coth(πν/2)Γ2(−1/4+iν/4)Γ2(−1/4−iν/4)\n128π2ǫ3,(5.33)\nwhere we used the relations Γ( −iν/2)Γ(iν/2) = 2π/[νsinh(πν/2)] and Γ(5 /4−iν/4)Γ(5/4 + iν/4)Γ(−1/4−\niν/4)Γ(−1/4+iν/4) = 2π2/cosh2(πν/2), see figure 2.9\n0.0 0.2 0.4 0.6 0.8 1.0Ε0.51.01.52.0/CapXi\nFIG. 2. The modification factor of the power spectrum of the te nsor perturbations on a de Sitter background as a function of\nǫ=γH2.\nThis is the main result of the paper: In our ghost-free but Lorentz violating Weyl theory of gravity the spectrum\nof gravitational waves, when evaluated on a de Sitter background , is scale-invariant, as in General Relativity, but its\namplitude in the late time limit varies with ǫ≡√γH, where√γis the length scale on which the Weyl term operates\nandH−1is the Hubble radius. For small ǫ, Ξ≈1: the spectrum of primordial gravitational waves is almost the\nsame as in General Relativity [11] because the Lorentz symmetric re gime before horizon crossing lasts long (it spans\naz-interval much bigger than 1); as ǫincreases, that is as the Lorentz-violating length scale approache s the Hubble\nradius, its amplitude decreasesup to a factor 65%, before eventu ally blowing up as ǫgrowsfurther and approachesthe\ncritical value 1. For ǫ >1 the spectrum, as we saw, cannot be defined as the vacuum expec tation value of quantized\nmodes since they do not oscillate at early time. Another criterion mus t then be chosen to define it in this regime.\nInanycase, thespectrumofprimordialgravitationalwavesisgre atlymodified, iftheLorentz-violatingscalehappens\nto be of the same order of magnitude or larger than the Hubble radiu s. But, again, it must be stressed that only\nwhen the Hubble radius is considerably larger than the Weyl-correct ion characteristic length scale ( ǫ≪1) do the high\nfrequency modes become those of flat spacetime and it is only then t hat there is a natural definition of the vacuum\nstate.\nIn slow-roll inflation now, the Hubble parameter His no longer constant but becomes a slowly decreasing function\nof cosmic time. The previous results therefore still hold but H(and hence ǫ=√γHandν=/radicalbig\n1/ǫ2−1) in the\nexpression (5.33) of the spectrum may be evaluated around horizo n crossing when H(t) =k/a(t). This allows to\nexpresst, and therefore H,ǫandνin terms of k. Sinceǫ(k) is a decreasing function of k, the qualitative behaviour\nof gravitational wave spectrum P(k;t|H=k/a) is then read off from the right to left in figure 2 when kincreases. What\nhappens when ǫ(k) crosses the critical value 1 during inflation remains to be investigat ed.\nVI. PRIMORDIAL GRAVITATIONAL WAVES IN EINSTEIN PLUS PURE WE YL SQUARE GRAVITY\nIn order to compare and contrast the ghost-free but Lorentz- violating Weyl theory of gravity that we have studied\nin this paper with ordinary “ghastly”3Weyl theory, we summarise and develop here the results of [6] to ob tain the\nexact spectrum of gravitational waves on a de Sitter background in Weyl gravity.\nWhen expanded to second orderin the tensorialperturbations ar ounda FLRW background,see (3.2) for definitions,\nthe action of Weyl gravity (1.1) becomes (see see [5] or appendix A)\nS[hij] =1\n8κ/integraldisplay\ndηd3x/bracketleftbig\na2(h′\nijh′ij−∂khij∂khij)−γ(h′′\nijh′′ij−2∂kh′\nij∂kh′ij+∂klhij∂klhij)/bracketrightbig\n,(6.1)\nwhich must be compared to (3.5). (Recall, see [4], that γmust be positive to avoid the ghost dofs to be tachyonic on\nMinkowski spacetime.) Proceeding as in [6] “` a la” Ostrogradsky we in troduce a new variable Qij≡h′\nijas well as a\n3i.e. “causing great horror or fear” (Oxford dictionary)10\nLagrange multiplier λij(both of them transverse traceless) and consider the equivalent action\nS[hij,Qij,λij] =1\n8κ/integraldisplay\ndηd3x/bracketleftbig\na2(h′\nijh′ij−∂khij∂khij)\n−γ(Q′\nijQ′ij−2∂kh′\nij∂kh′ij+∂klhij∂klhij)+2λij(Qij−h′\nij)/bracketrightbig\n.(6.2)\nThe conjugate momenta are\nπij\nh=1\n4κ(a2h′ij−2γ△h′ij−λij), πij\nQ=−γ\n4κQ′ij. (6.3)\nThe eoms obtained by extremisation of the action with respect to λijandQijare\nQij=h′\nij, γQ′′\nij+λij= 0 (6.4)\nand allow to express the momenta in terms of hijalone as\nπij\nh=1\n4κ(a2h′ij−2γ△h′ij+γh′′′ij), πij\nQ=−γ\n4κh′′ij. (6.5)\nQuantization is implemented by imposing the commutation relations\n[ˆhij(η,/vector x1),ˆπij\nh(η,/vector x2)] = 2i/planckover2pi1δ(/vector x1−/vector x2),[ˆQij(η,/vector x1),ˆπij\nQ(η,/vector x2)] = 2i/planckover2pi1δ(/vector x1−/vector x2), (6.6)\nall other commutators being zero and the factor 2 coming from the fact that both hijandQijrepresent two dofs.\nWe now expand ˆhijin Fourier modes\nˆhij(η,/vector x) =/integraldisplayd3k\n(2π)3/2/bracketleftbigg/parenleftbigg/summationdisplay\nλ=1,2eλ\nij(/vectork)ˆaλ\n/vectorkh(1)\nk(η)+/summationdisplay\nλ=1,2eλ\nij(/vectork)ˆbλ\n/vectorkh(2)\nk(η)/parenrightbigg\nei/vectork·/vector x+h.c./bracketrightbigg\n. (6.7)\nPlugging this expansion into the definitions of the momenta (6.5), we fi nd that the commutation relations (6.6) are\nequivalent to imposing (omitting the index λ)\n[ˆa/vectork1,ˆa†\n/vectork2] =δ(/vectork1−/vectork2),[ˆb/vectork1,ˆb†\n/vectork2] =−δ(/vectork1−/vectork2) (6.8)\n(all other commutators vanishing) if the mode functions satisfy th e following Wronskian conditions\nh(1)\nk[(a2+2γk2)h′(1)\nk∗+γh′′′(1)\nk∗]−h(2)\nk[(a2+2γk2)h′(2)\nk∗+γh′′′(2)\nk∗]−c.c.= 4κi/planckover2pi1,\nh′(1)\nkh′′(1)\nk∗−h′(2)\nkh′′(2)\nk∗−c.c.=−4κi/planckover2pi1\nγ.(6.9)\nThe constancy of the Wronskians is ensured if the mode functions h(1)\nkandh(2)\nkand their complex conjugates are four\nindependent solutions of the eom for hijwhich reads, in Fourier space\n(a2h′\nk)′+a2k2hk+γ(h′′′′\nk+2k2h′′\nk+k4hk) = 0. (6.10)\nLet us now specialise to a de Sitter background where a= 1/(−Hη),Hbeing a constant and ηconformal time.\nThen, as shown in [6] the eom (6.10) factorises neatly: if h(1)\nk≡zµ(1)\nkandh(2)\nk≡zµ(2)\nkare taken to solve, respectively,\n(withz≡ −kη)\nd2µ(1)\nk\ndz2+/parenleftbigg\n1−2\nz2/parenrightbigg\nµ(1)\nk= 0,\nd2µ(2)\nk\ndz2+/parenleftbigg\n1+1\nγH2z2/parenrightbigg\nµ(2)\nk= 0,(6.11)\nthen, as can be shown explicitly, they satisfy the original eom (6.10) and, also, the Wronskian conditions (6.9) if\nh(1)\nkdh(1)\nk∗\ndz−c.c.=h(2)\nkdh(2)\nk∗\ndz−c.c.=−4κi/planckover2pi1z2\nγk3[2+1/(γH2)]. (6.12)11\nThe first equation (6.11) can be seen as the eom for the usual Einst ein gravitonand the second for the Weyl ghostdofs.\nWhat remains to be done is to choose “positive frequency” modes wh ich define a “Bunch-Davies” vacuum state |0∝an}b∇acket∇i}ht\nsuch that ˆ ak|0∝an}b∇acket∇i}ht=ˆbk|0∝an}b∇acket∇i}ht= 0. This is easy in this case since there is no violation of Lorentz covar iance, so that the\neoms (6.11) both reduce to the standard Minkowski eoms for mass less fields for z≫1 andz≫1/(γH2), that is in\nthe remote past when the physical wavelengths a/kof the modes are much shorter than both the Hubble radius scale\nH−1and the length scale√γset by the Weyl term. We therefore impose that for z→ ∞,\nh(1)\nk∼h(2)\nk∼/radicalbigg\n2κ\nk3H/radicalbig\n1+2γH2zeiz. (6.13)\nThe solutions of the eoms (6.11) having this early time/short wavelen gth behaviour are, up to a phase\nh(1)\nk(z) =/radicalbigg\n2κ\nk3H/radicalbig\n1+2γH2(1+iz)eiz,\nh(2)\nk(z) =/radicalbigg\n2κ\nk3H/radicalbig\n1+2γH2/radicalbiggπ\n2eiπ¯ν/2z3/2H(1)\n¯ν(z)(6.14)\nwith ¯ν≡(1/2)/radicalbig\n1−4/(γH2).\nThe power spectrum P(k;z) of the gravitational waves hijis again defined as\n∝an}b∇acketle{t0|ˆhij(η,/vector x1)ˆhij(η,/vector x2)|0∝an}b∇acket∇i}ht=/integraldisplay\nd3kP(k;η)\n4πk3ei/vectork·(/vector x1−/vector x2)(6.15)\nand is given by\nP(k;z) =k3\nπ2/parenleftBig\n|h(1)\nk(z)|2−|h(2)\nk(z)|2/parenrightBig\n=2κH2\nπ21\n1+2γH2/parenleftBig\n1+z2−π\n2z3|eiπ¯ν/2H(1)\n¯ν(z)|2/parenrightBig\n.(6.16)\nWhatever the value of γH2, that is whatever the sign of ¯ ν2=1\n4[1−4/(γH2)], the last two terms do not contribute\nto the power spectrum when z→0, that is, at late time when the modes have exited the Hubble radius, and we have\nthat\nP(k;z→0)≡2κH2\nπ2ΞW,where Ξ W=1\n1+2γH2. (6.17)\n(This generalises [6], where the spectrum was computed for small va lues of the parameter γH2only.)\nThis de Sitter spectrum of primordial gravitational waves at late tim es, obtained in Einstein plus pure Weyl-square\ngravity based on the action (1.1), and that obtained in (5.33) and fig ure 2 in the ghost-free but Lorentz-violating\nWeyl gravity theory based on the action (2.1), are very different: contrarily to Ξ, Ξ Wnever blows up and, as γH2\ntends to infinity, goes to zero as ( γH2)−1.\nVII. CONCLUSION\nWe have proposed and investigated a mechanism to eliminate the ghos t degrees of freedom of higher-curvature\ngravity theories. The mechanism invokes the gradient of a scalar fie ld which is timelike and therefore implies a\nbreakdown of Lorentz covariance. Although we only discussed her e the particular example of Weyl gravity, the\nmechanism appears generic enough to eliminate any higher-derivativ e ghosts. We expect that a canonical way of\nanalysing generic higher-curvature gravity, developed by the aut hors [12], will be useful in studies in this direction.\nWe investigated quantisation of the inflationary tensor perturbat ions in our ghost-free Weyl gravity model and\ndefined the “positive frequency” modes as those reducing to flat s pacetime, Lorentz-violating, positive frequency\nmodes. Such a modification of the quantum vacuum state may offer a n observable signature of Lorentz violation at\nshort wavelengths. Indeed, we found that in de Sitter inflation this modification of the vacuum state gives rise to an\nextra overall factor for the super-horizon power spectrum of g ravitational-wave background generated from quantum\nfluctuations.\nSince, from a phenomenological point of view, the energy scale of Lo rentz violation κM2\nPl/√γcan be as low as the\nHubble parameter during inflation we expect that inflationary cosmo logical perturbations may open a useful window\nto the physics of Lorentz violation.12\nACKNOWLEDGMENTS\nND thanks YITP for its enduring and generous hospitality. She is also grateful to Hirosaki University where this\nwork was completed. YS thanks ASC at LMU for hospitality, where pa rt of this work was done. He also thanks\nShunichiro Kinoshita for useful conversations. This work was supp orted in part by MEXT thorough Grant-in-Aid\nfor Scientific Research (A) No. 21244033 and Grant-in-Aid for Cre ative Scientific Research No. 19GS0219 (MS) and\nby JSPS through Postdoctoral Fellowship for Research Abroad an d Grant-in-Aid for JSPS Fellows (YS). This work\nwas also supported in part by MEXT through Grant-in-Aid for the Glo bal COE Program “The Next Generation of\nPhysics, Spun from Universality and Emergence” at Kyoto Universit y.\nAppendix A: Weyl-squared actions\nThe Weyl-squared action SW2=/integraltextd4x√−gCabcdCefghγaeγbfγcguduhdiscussed in the main text is one of the\nthree invariants consisting of quadratic Weyl tensor fully contrac ted with γabandua; the other two are\nSW0=/integraldisplay\nd4x√−gCabcdCefghγaeγbfγcgγdh, SW4=/integraldisplay\nd4x√−gCabcdCefghγaeubufγcguduh.(A1)\nThey are such that\n−4SW2+SW0+4SW4=/integraldisplay\nd4x√−gCabcdCabcd, (A2)\nhence the coefficient −γ/2 in (1.1) and +2 γin (2.1).\nA useful fact is that these actions are invariant under a conforma l transformation, gab= Ω2˜gab: the Weyl tensor\nis invariant, Cabcd=˜Cabcd, whereas the geometrical quantities associated with the chronon field are transformed\nasua= Ω˜uaandγab= Ω2˜γab, respectively. Then it can be checked that SW0,SW2andSW4are all conformally\ninvariant.\nThe expansion of these actions to second order in gauge-invariant perturbations around FLRW spacetimes are (see\nmain text for definitions and with W≡Ψ−Φ and [5]):\n(2)SW2=−1\n4/integraldisplay\ndηd3x(−2∂kh′\nij∂kh′ij−△Ψi△Ψi),\n(2)SW0= 4(2)SW4\n=/integraldisplay\ndηd3x/bracketleftbigg1\n4h′′\nijh′′ij+1\n4△hij△hij+1\n2∂kh′\nij∂kh′ij+1\n2∂iΨ′\nj∂iΨ′j+2\n3(△W)2/bracketrightbigg\n.(A3)\nAppendix B: Derivation of the equations of motion\nTocompute thevariationofourghost-freeWeylactionwith respe cttothe metric, it isuseful toemployanADM-like\nformalism. To begin with, we express the action as\nSW2=/integraldisplay\ndDx√−gCabcdCefghγaeγbfγcguduh≡/integraldisplay\ndDx√−gWabcWabc, (B1)\nwhere\nua≡N ∂aχ, N≡1√σ∂aχ∂aχ, γab≡gab−σuaub, σ≡uaua, Wabc≡γadγbeγcfugCdefg.(B2)\nNote that W[ab]c=Wabc,uaWabc=uaWbca= 0, and Wabb= 0. We assume σnever vanishes so that the lapse\nfunction Nremains finite. We also define the extrinsic curvature of the constant- χsurfaces by, ∇being the covariant\nderivative,\nKab≡γacγbd∇duc. (B3)\nThe Codazzi relation gives\nWabc=γadγbeγcf/bracketleftbigg\n2∇[dKe]f−2\nD−2∇[d(Ke]f−Kγe]f)/bracketrightbigg\n, (B4)13\nwhereK≡γabKab. In order to avoid copious appearances of γabandua, we introduce the following notations:\nγcdγaeγfb···∇dTe···f···→DcTa···b···, udγaeγfb···∇dTe···f···→ ∇uTa···b···, uaTa···b···→Tu···b···(B5)\nand so on. With this, we can write\nKab=Daub, Wabc= 2D[aKb]c−2\nD−2D[a(Kb]c−Kγb]c). (B6)\nIt should be noted that the tensor Wabconly contains one derivative along the direction of ua.\nLet us concentrate on the variation with respect to the metric ten sor,gab→gab+δgab. Define the variables\nA≡σ\n2uaubδgab, Ba≡γabucδgbc, Cab≡1\n2γacγbdδgcd. (B7)\nThen we have e.g.\nδgab= 2Cab+2σu(aBb)+2σAuaub, δ√−g=C+A, (B8)\nwhereC≡γabCab. We also have\nδua=Aua, δua=−Aua−Ba, δγab= 2Cab+2σu(aBb), δγab=σuaBb, δγab=−2Cab,(B9)\nand\nδKab=∇uCab+2K(acCb)c−N−1D(a(N Bb))−KabA+2σu(aKb)cBc. (B10)\nThe variation of the Weyl action with respect to the metric can henc e be computed, up to divergences, as\nδSW2=1\n2/integraldisplay\ndDx√−gBabδgab, (B11)\nwhere\nBab≡4∇u[N−1Dc(N Wc(ab))]+4N−1Dc(N Wc(ab))K−4N−1Dc(N Wcd(a)Kdb)−4N−1Dc(N Wc(a|d|)Kdb)\n−4Dd[N−1Dc(N Wcd(a)]ub)−4Dd[N−1Dc(N Wc(a|d|)]ub)+4σN−1Dc(N Wcde)Kdeuaub\n+4N−1Dc(N W(a|dc|Kb)d)+4N−1Dc(N Wcd(aKb)d)+4N−1Dc(N Wd(ab)Kcd)\n+8\nD−2Wc(ab)Dd(Kcd−Kγcd)−4WacdWbcd−2WcdaWcdb+WcdeWcdegab.\n(B12)\nIt can be shown that Babtensor is traceless and free from the trace part of Kab.\nA most important property of Babis that it contains derivatives ∇ualonguaup to second order only (in the first\nterm which is of the type ∇uDWwithW∼D∇uγ); hence the eoms remain second order in timelike derivatives\nwhenever uais timelike. Note also that the other fourth order terms in Babare all of the type DDWwhich are\nfirst order in time and third order in space when uais timelike. Because of this distinction between space and time\nderivatives Babis not Lorentz covariant.\nThe variations of the quantities with respect to χare given by\nδua=Va, δγab=−2σu(aVb)whereVa≡N γab∂bδχ. (B13)\nThe variation of the Weyl action with respect to χis then computed, up to divergence, as\nδSW2=−/integraldisplay\ndDx√−g∇aWaδχ, (B14)\nwhere\nWa≡4σWabcCbucu+2γabWcdeCbcde. (B15)\nNote that we may rewrite as ∇aWa=N−1Da(N Wa).14\n[1] H. Weyl, Sitzungsber. Preuss. Akad. Wiss. Berlin (Math. Phys.) , 465 (1918); Annalen Phys. (Leipzig) 59, 101 (1919);\nRaum - Zeit - Materie , 5th ed. (Springer-Verlag, Berlin, 1923) Chap. IV, [Englis h translation Space, Time, Matter (fourth\nedition, Dover Publications, NewYork, USA, 1952)].\n[2] A. De Felice and S. Tsujikawa, Living Rev. Rel. 13, 3 (2010), arXiv:1002.4928 [gr-qc].\n[3] A. Pais and G. E. Uhlenbeck, Phys. Rev. 79, 145 (1950); J. M. Cline, S. Jeon, and G. D. Moore,\nPhys. Rev. D70, 043543 (2004), arXiv:hep-ph/0311312 [hep-ph]; R. Woodar d, Lect. Notes Phys. 720, 403 (2007),\narXiv:astro-ph/0601672 [astro-ph]; A. V. Smilga, Nucl. Ph ys.B706, 598 (2005), arXiv:hep-th/0407231 [hep-th];\nSIGMA 5, 017 (2009), arXiv:0808.0139 [quant-ph].\n[4] K. S. Stelle, Gen. Rel. Grav. 9, 353 (1978).\n[5] N. Deruelle, M. Sasaki, Y. Sendouda, and A. Youssef, J. Co smol. Astropart. Phys. 1103, 040 (2011),\narXiv:1012.5202 [gr-qc].\n[6] T. Clunan and M. Sasaki, Class. Quant. Grav. 27, 165014 (2010), arXiv:0907.3868 [hep-th].\n[7] D. Blas, O. Pujolas, and S. Sibiryakov, J. High Energy Phy s.10, 029 (2009), arXiv:0906.3046 [hep-th].\n[8] C. Germani, A. Kehagias, and K. Sfetsos, J. High Energy Ph ys.0909, 060 (2009), arXiv:0906.1201 [hep-th].\n[9] P. Horava, Phys.Rev. D79, 084008 (2009), arXiv:0901.3775 [hep-th].\n[10] J. M. Bardeen, Phys. Rev. D22, 1882 (1980); H. Kodama and M. Sasaki, Prog. Theor. Phys. Sup pl.78, 1 (1984); V. F.\nMukhanov, H. A. Feldman, and R. H. Brandenberger, Phys. Rept .215, 203 (1992).\n[11] L. P. Grishchuk, J. Exp. Theor. Phys. 40, 409 (1975); A. A. Starobinskii, J. Exp. Theor. Phys. Lett. 30, 682 (1979).\n[12] N. Deruelle, M. Sasaki, Y. Sendouda, and D. Yamauchi, Pr og. Theor. Phys. 123, 169 (2010), arXiv:0908.0679 [hep-th]." }, { "title": "0707.3247v1.Cosmological_Particle_Creation_in_the_Presence_of_Lorentz_Violation.pdf", "content": "arXiv:0707.3247v1 [hep-ph] 22 Jul 2007Cosmological Particle Creation in the Presence of Lorentz V iolation\nE. Khajeh,∗N. Khosravi,†and H. Salehi‡\nDepartment of Physics, Shahid Beheshti University, Evin, T ehran 19839, Iran\nIn recent years, the effects of Lorentz symmetry breaking in c osmology has attracted considerable\namount of attention. In cosmological context several topic s can be affected by Lorentz violation,e.g.,\ninflationary scenario, CMB, dark energy problem and barryog enesis. In this paper we consider the\ncosmological particle creation due to Lorentz violation (L V). We consider an exactly solvable model\nfor finding the spectral properties of particle creation in a n expanding space-time exhibiting Lorentz\nviolation. In this model we calculate the spectrum and its va riations with respect to the rate and\nthe amount of space-time expansion.\nPACS numbers: 04.62.+v, 11.30.Cp, 98.80.-k, 98.80.Jk\nI. INTRODUCTION\nParticle creation in the cosmological context is one of\nthe most interesting feature of QFT in curved space time\n[1]. In this process, a quantum field propagates on an ex-\npanding space-time and the field quanta (particles) are\ngenerated through an impulse of the space-time expan-\nsion. The density and the rate of particle production\ndepend on the vigor of the expansion.\nFurthermore, it is well known that if a conformally in-\nvariant field propagates in a space-time which is confor-\nmally equivalent to Minkowski (conformal triviality) no\nparticle production occurs. Thus, we expect that mass-\nless quanta of Maxwell and Dirac fields do not arise from\nthe expansion of Universe. But if a term is added to the\nfield equation such that conformal invariance is broken\nthen these particles are created [1, 2].\nQFT in curved background is based on the hypothesis\nthatthe fieldequationsarelocallyLorentzinvariant. But\nin recent years there have been many attentions to the\npossibility of Lorentz symmetry breaking at high ener-\ngies. The initial motivation came from string theory [3],\nand more recently this symmetry breaking has been dis-\ncussed in the context of noncommutative geometry [4, 5].\nAlso, there has been evidence that this symmetry may be\nbroken in at least three different phenomena:\ni) Observation of ultra-high energy cosmic rays with\nenergies beyond the so-called GZK cutoff, EGZK⋍\n1019eV[6, 7].\nii) Events involving gamma radiation with energies be-\nyond 20TeV from distant sources such as Markarian 421\nand Markarian 501 blazers [8].\niii)Studies of the evolution of air showers produced by\nultra high-energy hadronic particles suggest that pions\nlive longer than expected [9]. These observations can\nbe explained via the breaking of Lorentz symmetry [10,\n11]. Now, if we want to consider the effect of LV in\nQFT in curved space-time, an important question for\n∗e-khajeh@sbu.ac.ir\n†n-khosravi@sbu.ac.ir\n‡h-salehi@sbu.ac.irinvestigation is: How dose Lorentz violation affect the\nparticle creation in an expanding space-time?\nThe possible effects of LV physics in inflationary cos-\nmology has been studied in [12, 13, 14, 15]. The au-\nthors found that the spectrum of fluctuations in infla-\ntionary cosmology and so the spectrum of temperature\nanisotropiesintheCosmicMicrowaveBackgroundcanbe\naffected by LV. Also it is shown that Lorentz violation\nis relevant to the dark energy problem and barryogene-\nsis [16, 17, 18, 19]. In the present work we look for the\neffects of LV in particle creation process in an expanding\nspace-time. Our interest is to study the form and prop-\nerties of particle creation spectrum of a scalar field. For\nthis purpose we choose an exactly solvable model and we\nfind the characteristics of the spectrum.\nWe benefit from the model that has been introduced\nin [20, 21] for LV to study this subject. In this model\nthe usual dispersion relation is modified by adding a\ntermα2k4. The modified dispersion relation preserves\nrotational invariance but violates the boost invariance of\nLorentzsymmetry. It has been shown that addinga term\nα2k4to the dispersion relation is equivalent to putting a\nvector field which is coupled with matter field in the La-\ngrangian of the model. This vector field can physically\nbe assumed as the four velocity of a preferred inertial\nobserver. The additional term in the Lagrangian which\nenforces the LV, breaks the conformal invariance. There-\nfore we expect that the massless field quanta are created\nduring the space-time expansion.\nThe organization is as follows: in section II we quan-\ntize the Lorentz violation model introduced in [20] on\nMinkowski space-time. In section III we define an ex-\npanding cosmological model and consider the quantiza-\ntion of the LV model on this space-time. Finally, we ob-\ntain and discuss the spectrum characteristics of created\nmassless particles in details.\nII. LORENTZ VIOLATION MODEL AND\nQUANTIZATION IN MINKOWSKI SPACE-TIME\nAs was mentioned before, in the model introduced for\nLV in [20] the dispersion relation is modified by an ad-\nditional term α2k4and the modified dispersion relation2\nreads as follows\nω2(/vectork) =/vextendsingle/vextendsingle/vextendsingle/vectork/vextendsingle/vextendsingle/vextendsingle2\n−α2/vextendsingle/vextendsingle/vextendsingle/vectork/vextendsingle/vextendsingle/vextendsingle4\n. (1)\nA Lagrangian that can lead the above dispersion relation\nfor a scalar field φis\nL=1\n2(∂µϕ∂µϕ+α2(D2ϕ)2), (2)\nwhereαis a constant (of order the Planck energy) that\nsets the scale of the Lorentz violation and D2is the spa-\ntial Laplacian, that is\nD2ϕ=−DµDµϕ=−qµν∂ν(qτ\nµ∂τϕ),(3)\nwhereqµνis the (positive definite) spatial metric orthog-\nonal to the unit timelike vector uµ\nqµν=−ηµν+uµuν, ηµνuµuν= 1.(4)\nThe vector field uµcan physically be interpreted as the\nfour velocity of a preferred inertial observer. The rest\nframe of this preferred observer may be called aether. In\nthis rest frame we have uµ=(1,0,0,0).\nFor the subsequent consideration in this chapter we\nshall take a coordinate system in which uµis constant\n(The aether is a particular example for such a coordinate\nsystem). With this assumption the equation of motion\nfor the Lagrangian (2) is\n[/square−α2qµνqγδ∂µ∂ν∂γ∂δ]ϕ= 0. (5)\nThe complex mode solution of (5) are taken as\nu/vectork∝e−ikµxµ, (6)\nwhere\nω2(/vectork) =|/vectork|2−α2qµνqγδkµkνkγkδ. (7)\nIn general the relation (7) admits imaginary frequen-\ncies. These frequencies lead to instability and unbound-\nedness [2]. It is necessary in this context to restrict our-\nselves to ordinary solutions with real frequencies. In this\ncase the mode solutions (6) become positive frequency\nmode solutions.\nWe define the scalar product for two solutions ϕ1and\nϕ2of Eq.(5) as follows\n(ϕ1,ϕ2) =−i/integraldisplay\nΣϕ1(x)(← →∂µ−α2qµνqγδ←−−→∂ν∂γ∂δ)ϕ∗\n2(x)dΣµ,\n(8)\nwheredΣµ=nµdΣ, withnµafuture-directedunit vector\northogonalto the space like hypersurface Σ and dΣ is the\nvolumeelementinΣ. ThehypersurfaceΣistakentobe a\nCauchy surface in the space-time and we can show, using\nGauss’ theorem, that the value of ( ϕ1,ϕ2) is independent\nof Σ. The notation←−−→∂ν∂γ∂δin (8) is defined by\nϕ1←−−→∂ν∂γ∂δϕ2=\n+ϕ1∂ν∂γ∂δϕ2−ϕ2∂ν∂γ∂δϕ1\n−∂νϕ1∂γ∂δϕ2+∂νϕ2∂γ∂δϕ1\n+∂γϕ1∂ν∂δϕ2−∂γϕ2∂ν∂δϕ1\n−∂δϕ1∂γ∂νϕ2+∂δϕ2∂γ∂νϕ1.(9)The ordinary modes u/vectorkare orthogonal\n(u/vectork,u/vectork′) = 0,/vectork∝negationslash=/vectork′, (10)\nand if we choose\nu/vectork= [2(2π)3(ω−2α2q0νqγδkνkγkδ)]−1\n2e−ikµxµ,(11)\nthese ordinary modes are normalized in the sense of the\nscalar product (8). An ordinary solution of Eq.(5) may\nbeexpandedin termofthe ordinarymodes(11)andtheir\ncomplex conjugates\nϕ(t,/vector x) =/summationdisplay\n/vectorka/vectorku/vectork(t,/vector x)+a†\n/vectorku∗\n/vectork(t,/vector x)ω∈R,(12)\nandthe system isquantized in the canonicalquantization\nscheme by imposing the following commutation relations\n[a/vectork,a/vectork′] = 0 [ a†\n/vectork,a†\n/vectork′] = 0 [ a/vectork,a†\n/vectork′] =δ/vectork/vectork′.(13)\nWith these commutation relations, a/vectork’s anda†\n/vectork’s are an-\nnihilation and creation operators and the vacuum of the\nLorentz violation model in Minkowski space-time is de-\nfined by\na/vectork|0M\nLV∝angb∇acket∇ight= 0, ω∈R. (14)\nThe above vacuum was defined with respect to a coor-\ndinate system in which the components of uµwere con-\nstant. We shall work in a coordinate system which corre-\nsponds to the rest frame of a preferred inertial observer\n(aether). In this case uµtakes the form (1 ,0,0,0).With\nrespect to the aether the dispersion relation (7) and the\nnormalization constant in (11) are transformed to Eq.(1)\nandω−1/2, respectively.\nItisimportanttonotethatinwhatsenseistheLorentz\ninvariance violated. The Lagrangian (2) does not ex-\nhibit the invariance under particle boost transformations\n[22, 23]. These are transformations of a physical system\n(particles or localized fields) within a fixed coordinate\nsystem. It is in this sense that Lorentz invariance is vi-\nolated. Therefore the above vacuum may not be consid-\nered as an invariant state under a particle Lorentz trans-\nformation transforming a physical system within the rest\nframe of the preferred observer (aether).\nIII. COSMOLOGICAL MODEL AND PARTICLE\nCREATION\nTo show how particle creation occurs in the cosmologi-\ncal context, we use a two-dimensional Robertson-Walker\nspace-time with the line element\nds2=dt2−a2(t)dx2, (15)\nand consider the generalization of the Lagrangian (2) to\nthis cosmological space-time, namely\nL(ϕ,uµ,λ) =1\n2√−g(gµν∇µϕ∇νϕ−ξRϕ2+α2(D2ϕ)2\n+λ(1−uµuµ)), (16)3\nwhereD2ϕis the covariant spatial Laplacian [20](the co-\nvariant analogue of (3))\nD2ϕ=−DµDµϕ=−qµν∇ν(qτ\nµ∇τϕ),(17)\nandqµν=−gµν+uµuνwithgµνcorrespondingto (15). ξ\nis a coupling constant between the scalar field and scalar\ncurvature. The vector field uµis as a non-dynamical\nvectorfieldtobespecifiedbytheconditionsofthetheory.\nBy introducing the conformal time η, defined by dη=\ndt/a(t), the metric (15) takes the form\nds2=c(η)(dη2−dx2), c(η) =a2(t).(18)\nThe Lagrange multiplier λin (16) imposes the follow-\ning constraint on uµ\ngµνuµuν= 1. (19)\nIn the context of the homogeneous and isotopic cosmo-\nlogicalmetric(18) thevectorfield uµistakenastosatisfy\nthe isotropic property of cosmological metric and equa-\ntion (19). This leads to\nuµ≡(/radicalbig\nc(η),0). (20)\nThusqµνis\nq00=q01=q10= 0and q 11=c(η).(21)\nNow setting the variation of the action S=/integraltext\nL(ϕ)d4x\nwith respect to ϕequals to zero yields the equations of\nmotion for ϕin the metric (18) as follows\n/squareϕ+ξRϕ−α2\nc2(η)∂4ϕ\n∂x4= 0. (22)/s32/s32/s67/s40 /s41\n/s61/s48/s65/s43/s66\n/s65/s45/s66/s65/s45 /s66/s65/s43 /s66\n/s65\nFIG. 1: The figure shows c(η) with respect to η. It shows\nc(η=±∞) =A±Bsoa(η=±∞) =√\nA±B. Also it\nshows the time interval ∆ η=η2−η1that the proportion\nof total expansion that occurs in this time interval is ε. We\nchooseε= 0.99.We want to quantize ϕin equation (22) when ξ= 0\n(it is the conformal coupling case in two dimension in\nabsence of the Lorentz violation term). For this purpose\nwe will obtain the mode solutions uk(x) of equation (22).\nIn order to avoid the well known ambiguities in the\nparticle concept in curved space-time [1], we suppose\nthat the space-time can be treated as asymptotically\nMinkowskian in the remote past and future. We refer\nto the remote past and future as the inandoutregions\n, respectively. We take c(η) as\nc(η) =A+Btanh(ρη), (23)\nwhereA,Bandρare some constants such that 2 Bandρ\nrepresentthe amountandthe rateofexpansionin confor-\nmal time η, respectively [see fig.(1)]. Then in the remote\npast and future the space-time becomes Minkowskian\nsince\nc(η)→A±B, η→±∞. (24)\nThis kind of conformal scale factor first introduced by\nBernard and Duncan in [24] and then has been used in\nmany works [25, 26, 27, 28] for considering the cosmolog-\nical particle creation.\nWe can solve Eq.(22) and obtain uk(x) by the method\nof separation of variables. The mode solution of this\nequation is\nuk(η,x) = (2π)−1\n2eik.xχk(η), (25)\nsuch that χk(η) satisfies the following equation\nd2\ndη2χk(η)+(k2−α2\nc(η)k4)χk(η) = 0.(26)\nThis equation can be solved in terms of hypergeometric\nfunctions. The normalized modes which behave like the\npositive frequency Minkowski space modes in the remote\npast (η−→−∞) are\nukin(η,x) = (4πωin)−1\n2exp(ikx−iω+η (27)\n−iω−\nρln[(A+B)eρη+(A−B)e−ρη])\n×F(1+iω−/ρ,iω−/ρ;1−iωin/ρ;z),\nand the modes which behave like positive frequency\nMinkowski modes in the outregion as η−→+∞are\nfound to be\nuout\nk(η,x) = (4πωin)−1\n2exp(ikx−iω+η (28)\n−iω−\nρln[(A+B)eρη+(A−B)e−ρη])\n×F(1+iω−/ρ,iω−/ρ;1−iωout/ρ;1−z),\nwhere\nz=1\n2(A+B)tanh(ρη)+1\nA+Btanh(ρη), (29)4\nand\nωin=k(1−α2\nA−Bk2)1\n2\nωout=k(1−α2\nA+Bk2)1\n2\nω±=1\n2(ωout±ωin).(30)\nThe energies ωinandωoutcan be imaginary. But\nwe limit ourselves to the modes with real frequencies\n(the ordinary modes). The ordinary modes ukin(η,x)/parenleftBig\nuout\nk(η,x)/parenrightBig\nare orthonormal in the following conserved\ninner product\n(ϕ1,ϕ2) = (31)\n−i/integraldisplay\nΣ√−gϕ1(x)(← →∂µ−α2qµνqγδ←−−→∂ν∂γ∂δ)ϕ∗\n2(x)dΣµ,\nwhere it is the generalization of inner product (8) to\ncurved space-time and←−−→∂ν∂γ∂δis defined in (9) .Thus we\nmayexpandoneordinarysolutionofEq.(22)withrespect\ntoukin(η,x)/parenleftBig\nuout\nk(η,x)/parenrightBig\nas\nϕ(t,x) =/summationtext\nkain\nkuin\nk(t,x)+a†in\nku∗in\nk(t,x)ωin∈R\nϕ(t,x) =/summationtext\nkaout\nkuout\nk(t,x)+a†out\nku∗out\nk(t,x)ωout∈R\nand the second quantization is implemented in the same\nway as the Minkowski space-time by the following com-\nmutation relations\n[ain\nk,a†in\nk′] =δkk′[aout\nk,a†\nk′out] =δkk′.(32)\nAs we mentioned before we have chosen a conformally\nflat expanding space-time which is in the inandoutre-\ngions Minkowskian. With using the Eqs.(20) and (23) we\ndefine the aether in the inandoutregions as a frame in\nwhichuµis (√\nA−B,0) and (√\nA+B,0),respectively.\nThe vacuums with respect to the aether in the inand\noutregions are defined by\nain\n/vectork|0M\nLV∝angb∇acket∇ightin= 0 ωin∈R\naout\n/vectork|0M\nLV∝angb∇acket∇ightout= 0 ωout∈R.(33)\nSuppose the quantum field in the remote past resides\nin the state|0M\nLV∝angb∇acket∇ightinand suppose we are working in the\nHeisenberg picture. Thus in the outregion, the quantum\nfield is also in the state |0M\nLV∝angb∇acket∇ightin. But this state is not\nregarded by the aether in the outregion as the physical\nvacuum, this role being reserved for the state |0M\nLV∝angb∇acket∇ightout.\nWe want to calculate the number of particles detected\nin theoutregion in the state |0M\nLV∝angb∇acket∇ightin. If we use the\nlinear transformation properties of hypergeometric func-\ntions, we expand uin\nkwith respect to uout\nkand obtain the\nfollowing relation\nukin(η,x) =αkuout\nk(η,x)+βkuout\n−k(η,x),(34)/s48/s46/s48/s48 /s48/s46/s48/s53 /s48/s46/s49/s48/s48/s46/s48/s53/s46/s48/s120/s49/s48/s45/s57\n/s32/s50\n/s40/s107/s41\n/s107/s40/s69\n/s80/s41/s40/s98/s41\n/s107\n/s109/s97/s120\nFIG. 2: The figure shows the spectrum of created particles\nwhere√\n2B+1 = 103andρf= 10−5soλρf= 0.032≪1 .\n(Noteα= 1−,ε= 0.99 andA−B= 1)\nwhere\nαk= (ωout\nωin)1\n2Γ(1−(iωin/ρ))Γ(−iωout/ρ)\nΓ(−iω+/ρ)Γ(1−iω+/ρ)\nβk= (ωout\nωin)1\n2Γ(1−(iωin/ρ))Γ(iωout/ρ)\nΓ(iω−/ρ)Γ(1+iω−/ρ).(35)\nasiswellknown, αkandβkaretheBogolubovcoefficients\nand|βk|2is equal to the number of particles per mode\nkthat is detected in the outregion (created particles),\n|βk|2=sinh2(πω−/ρ)\nsinh(πωin/ρ)sinh(πωout/ρ).(36)\nIn the above relation |βk|2is dependent on the param-\netersα,A,Bandρ. We set the first parameter αin nat-\nural units ( /planckover2pi1=c= 1) asα= 1−[29] because we assume\nthat LV occurs in the scale of Planck energy. The three\nlast parameters are related to the conformal scalar factor\nc(η) by (23). These parameters determine the spectrum\nof created particles with respect to the coordinates ( η,x)\nand metric (18) through |βk|2in (36).\nIt is instructive to relate |βk|2to the physical param-\neters in the free-fall frame which is determined by the\ncoordinates ( t,x) and the metric (15). To do this we first\nfind the analogue of the parameters A,Bandρin the\nfree-fall frame. The parameters, AandBhave a good\nmeaning in the free-fall frame because as we have shown\nin fig.(1),√\nA−Band√\nA+Bare initial and final scale\nfactors in the free fall frame, respectively. For some ad-\nvantages we put A−B= 1, it implies that the conformal\ntime is the same as the proper time in the remote past\nalso the initial scale factor is 1 and the final scale factor\nis√\n2B+1. Now let us consider the analogue of ρ. This\nparameter is proportional to the inverse of time inter-\nval of expansion with respect to the conformal time. We\nshow this by the following argument. We define an effec-\ntive time interval of expansion, ∆ η.As we have shown5\n/s48/s46/s48\n/s52/s46/s48/s120/s49/s48/s51\n/s56/s46/s48/s120/s49/s48/s51\n/s49/s46/s50/s120/s49/s48/s52/s48/s46/s48/s48/s49/s46/s50/s48/s120/s49/s48/s45/s52\n/s40/s97/s41/s32\n/s32/s32/s40/s107\n/s109/s97/s120/s41\n/s40/s50/s66/s43/s49/s41/s49/s47/s50\n/s48/s46/s48\n/s52/s46/s48/s120/s49/s48/s51\n/s56/s46/s48/s120/s49/s48/s51\n/s49/s46/s50/s120/s49/s48/s52/s48/s46/s48/s48/s46/s49/s48/s46/s50/s48/s46/s51/s48/s46/s52/s48/s46/s53\n/s40/s98/s41\n/s32/s32/s107\n/s109/s97/s120/s40/s69\n/s80/s41\n/s40/s50/s66/s43/s49/s41/s49/s47/s50\n/s48/s46/s48\n/s52/s46/s48/s120/s49/s48/s45/s53\n/s56/s46/s48/s120/s49/s48/s45/s53\n/s49/s46/s50/s120/s49/s48/s45/s52/s48/s46/s48/s48/s49/s46/s53/s48/s120/s49/s48/s45/s52\n/s32/s32/s50\n/s40/s107\n/s109/s97/s120/s41\n/s32\n/s102/s32/s40/s49/s47/s116\n/s80/s41/s40/s99/s41\n/s48/s46/s48\n/s52/s46/s48/s120/s49/s48/s45/s53\n/s56/s46/s48/s120/s49/s48/s45/s53\n/s49/s46/s50/s120/s49/s48/s45/s52/s48/s46/s48/s48/s46/s49/s48/s46/s50/s48/s46/s51/s48/s46/s52/s48/s46/s53\n/s40/s100/s41\n/s32/s32/s107\n/s109/s97/s120/s40/s69\n/s80/s41\n/s102/s32/s40/s49/s47/s116\n/s80/s41\nFIG. 3: The abovefiguresshowtheproperties ofthespectrum\nof created particles. (a) and (b) show β2(kmax) andkmax\nversus the final amount of the scale factor ,√\n2B+1, where\nρf= 10−5. (c) and (d) show β2(kmax) andkmaxversus the\nrate ofexpansionin free fall frame, ρf, where√\n2B+1 = 103.\n(Noteα= 1−,ε= 0.99 andA−B= 1)in fig.(1), ∆ η(=η2−η1) is taken as a time interval such\nthat the proportion εof the total expansion is done in\nthis interval. We set ε= 0.99 that means 99% of total\nexpansion occurs in ∆ η. By this definition one can find\n,from (23), η2=−η1= tanh−1(ε)/ρand so\nρ=2tanh−1(ε)\nη2−η1, (37)\nwhereη1andη2arethe initialandthefinaltime oftheef-\nfective expansion respectively, and tanh−1(0.99) = 2.64.\nThe Eq.(37) shows ρ∝1/∆ηand the proportional con-\nstant is of order one. This relation makes it clear that ρ\nis proportional to the inverse of the time interval of ex-\npansion so it can be interpreted as the rate of expansion.\nIn analogy to ρwe can define ρfin free falling frame\nsuch that it is the inverse of expansion time interval in\nfree falling frame, namely\nρf=1\nt(η2)−t(η1). (38)\nFor evaluating ρfin the above relation we find the rela-\ntionship between tandηfrom (23) and dt=/radicalbig\nc(η)dη.\nWe get\nt(η) =√\n2B+1\nρtanh−1/parenleftBigg/radicalbig\nB[1+tanh( ρη)]+1√2B+1/parenrightBigg\n+1\n2ρln/parenleftBigg/radicalbig\nB[1+tanh( ρη)]+1−1/radicalbig\nB[1+tanh( ρη)]+1+1/parenrightBigg\n+C,(39)\nwhereCis a constant that sets the initial values of ηand\nt. From (39) together with (38) we find ρf=ρ/λwhere\nλis a function of Bandε, namely\nλ=√\n2B+1tanh−1/parenleftBigg/radicalbig\nB(1+ε)+1√\n2B+1/parenrightBigg\n−√\n2B+1tanh−1/parenleftBigg/radicalbig\nB(1−ε)+1√2B+1/parenrightBigg\n+1\n2ln/parenleftBigg/radicalbig\nB(1+ε)+1−1/radicalbig\nB(1+ε)+1+1/parenrightBigg\n−1\n2ln/parenleftBigg/radicalbig\nB(1−ε)+1−1/radicalbig\nB(1−ε)+1+1/parenrightBigg\n. (40)\nFrom (40) we find that for sufficiently large amount of B,\nλcan be approximated by (2 B)1\n2. Substituting ρbyλρf\nin (36) we get|βk|2in term of the physical parameter ρf\nin the free falling frame. We can approximate (36) in the\nlimitλρf≫1 by\n|βk|2=(ωout−ωin)2\n4ωinωout, (41)\nwhere in the region B≫1 andk∝negationslash= 0 the above relation\ntakes the form\n|βk|2=/parenleftbig\n1−√\n1−α2k2/parenrightbig2\n4√\n1−α2k2, (42)6\nwhich implies that |βk|2is independent of λρf.\nIn the opposite limit where λρfis small the spectrum\nshows an interesting behavior. The figure(2) shows the\nspectrum of created particles for√2B+1 = 103and\nρf= 10−5soλρf= 0.032. It shows that particles are\nnot created at low and high momenta. Most particles\nare produced with momenta around kmax. In figure (3)\nwe show the properties of this spectrum. Figures (3,a)\nand (3,c) show the variations of the number of created\nparticles at momentum kmaxwith respect to the amount\n,√\n2B+1, and the rate , ρf,of the space-time expansion\nrespectively. With increasing the amount and the rate of\nexpansion the number of created particles increases. Fig-\nures (3,b) and (3,d) shows kmaxwith respect to√\n2B+1andρf. We see that with increasing the amount and the\nrate of the expansion, the particles with higher momenta\nare created.\nAcknowledgments\nE. Khajehand N. Khosravithank H. Salehi forencour-\nagements. E. Khajeh thanks S. Jalalzadeh for useful dis-\ncussionsandresearchofficeofShahidBeheshtiUniversity\nfor financial supports. The authors thank an unknown\nreferee for useful comments.\n[1] N. D. Birrell and P. C. W. Davies, ”Quantum Fields In\nCurved Space”, (Cambridge University Press) (1982).\n[2] S. A. Fulling, ”Aspects of Quantum Field Theory\nin Curved Space-Time”, (Cambridge University Press)\n(1989).\n[3] V. A. Kostelecky and S. Samuel, Phys. Rev. D 39, 683\n(1989).\n[4] S. Carroll et. al., Phys. Rev. Lett. 87, 141601 (2001).\n[5] F. Lizzi, G. Mangano, G. Miele and M. Peloso, JHEP\n0206, 049 (2002), [hep-th/0203099].\n[6] K. Greisen, Phys. Rev. Lett. 16, 748 (1966).\n[7] M. Takeda et. al., (AGASA Collab), Phys. Rev. Lett. 81,\n1163 (1998).\n[8] F. Krennrich et. al., Ap. J. 560, L45 (2001).\n[9] E.E. Antonov, et. al., Pisma ZhETF 73, 506 (2001).\n[10] H. Sato, [astro-ph/0005218].\n[11] S. R. Coleman and S. L. Glashow, Phys. Rev. D 59,\n116008 (1999), [ hep-ph/9812418].\n[12] R. H. Brandenberger and J. Martin, Int. J. Mod. Phys.\nA17 (2002) 3663-3680, [hep-th/0202142]; Phys. Rev D\n71, 023504 (2005), [hep-th/0410223].\n[13] R. Easther, B. R. Greene, W. H. Kinney, G. Shiu, Phys.\nRev. D64 (2001) 103502, [hep-th/0104102].\n[14] A. A. Starobinsky, JETP Lett. 73 (2001) 371-374,\n[astro-ph/0104043].\n[15] J. Martin, R. H. Brandenberger, Phys. Rev. D65 (2002)\n103514, [hep-th/0201189]; Phys.Rev.D68(2003) 063513,\n[hep-th/0305161].\n[16] O. Bertolami, Classical Quantum Gravity 14, 2785(1997).\n[17] O. Bertolami, R. Lehnert, R. Potting, and A. Ribeiro,\nPhys. Rev. D 69, 083513 (2004).\n[18] G. L. Alberghi, R. Casadio, and A. Tronconi,\n[hep-ph/ 0310052].\n[19] G. L. Alberghi, R. Casadio, and A. Tronconi,\n[hep-ph/ 0310052].\n[20] T. Jacobson and D. Mattingly , Phys. Rev. D 63, 041502\n(2001), [hep-th/0009052]; D. Mattingly and T. Jacobson,\n[gr-qc/0112012].\n[21] T. Jacobson and D. Mattingly, Phys. Rev. D 70, 024003\n(2004), [gr-qc/0402005].\n[22] D. Collady and V. A. Kostelecky, Phys. Rev. D 55, 6760\n(1997).\n[23] D. Collady and V. A. Kostelecky, Phys. Rev. D 58,\n116002 (1998).\n[24] C. Bernard and A. Duncan, Ann. Phys. (NY) 107 (1977)\n201.\n[25] M. R. Setare, Int.J.Theor.Phys. 43 (2004) 2237-2242,\n[hep-th/0203126].\n[26] Andrew J. Tolley, Daniel H. Wesley, Phys.Rev. D72\n(2005) 124009, [hep-th/0509151].\n[27] Alex J. Hamilton, Daniel Kabat, Maulik K. Parikh,\nJHEP 0407 (2004) 024, [hep-th/0311180].\n[28] Don Koks, B. L. Hu, Andrew Matacz, Alpan Raval,\nPhys.Rev. D56 (1997) 4905-4915, [gr-qc/9704074].\n[29] The minus in 1−means approaching 1 from the left." }, { "title": "2007.14480v5.Complete_complementarity_relations_and_its_Lorentz_invariance.pdf", "content": "Complete complementarity relations and its Lorentz invariance\nMarcos L. W. Basso1,\u0003and Jonas Maziero1,y\n1Departamento de Física, Centro de Ciências Naturais e Exatas,\nUniversidade Federal de Santa Maria, Avenida Roraima 1000,\nSanta Maria, Rio Grande do Sul, 97105-900, Brazil\nIt is well known that entanglement under Lorentz boosts is highly dependent on the boost scenario\nin question. For single particle states, a spin-momentum product state can be transformed into an\nentangled state. However, entanglement is just one of the aspects that completely characterizes\na quantum system. The other two are known as the wave-particle duality. Although the entan-\nglement entropy does not remain invariant under Lorentz boosts, and neither do the measures of\npredictability and coherence, we show here that these three measures taken together, in a complete\ncomplementarityrelation(CCR),areLorentzinvariant. Peresetal., in[Phys. Rev. Lett. 88, 230402\n(2002)], realized that even though it is possible to formally define spin in any Lorentz frame, there is\nno relationship between the observable expectation values in different Lorentz frames. Analogously,\none can, in principle, define complementary relations in any Lorentz frame, but there is no obvious\ntransformation law relating complementary relations in different frames. However, our result shows\nthatthe CCRhave thesame valuein any Lorentz frame, i.e., there’sa transformationlawconnecting\nthe complete complementarity relations. In addition, we explore relativistic scenarios for single and\ntwo particle states, which helps in understanding the exchange of different aspects of a quantum\nsystem under Lorentz boosts.\nKeywords: Complete complementarity relations; Lorentz boosts; Relativistic scenarios\nI. INTRODUCTION\nEntanglement is one of the most intriguing characteristics that turns apart the quantum world from the classical\nworld. Its fundamental importance in quantum foundations [1, 2], together with its application in several areas, such\nas quantum information and quantum computation [3–5], has made the entanglement theory achieve great progress\nin recent decades. Moreover, there has been more and more interest in how entanglement behaves under relativistic\nsettings [6]. For instance, in Ref. [7] the author considered the relativistic version of the famous Einstein-Podolsky-\nRosen experiment with massive spin-1/2 particles. Czachor argued that the degree of violation of the Bell inequality\nis dependent on the velocity of the particles, leading to implications for quantum cryptography. A few years later,\nthe authors in Refs. [8, 9] showed that the entanglement of Bell states depends on the velocity of an observer. On\nthe other hand, the authors of Ref. [10] argued that the entanglement fidelity of a Bell state remains invariant for\na Lorentz boosted observer. However, in the same year, it was demonstrated by Peres et al. [11] that the entropy\nof a single massive spin-1/2 particle does not remain invariant under Lorentz boosts. Thereafter, the behavior of\nentanglement under Lorentz boosts has been receiving a lot of attention by researchers [12–20].\nAs pointed out by Palge and Dunningham in Ref. [21], the main aspect to be noticed here is that many of these\napparently conflicting results involve systems containing different particle states and boost geometries. Therefore,\nentanglement under Lorentz boosts is highly dependent on the boost scenario in question [22]. For single particle\nstates, a spin-momentum product state can be transformed into an entangled state. Beyond that, Lorentz boosts\ncan be regarded as controlled quantum operations where momentum plays the role of the control system, whereas\nthe spin can be taken as the target qubit, as argued in Ref. [17]. This implies that Lorentz boosts perform global\ntransformations on single particle systems. As in Refs. [16, 18, 21], by using discrete momentum states, in this article\nwe discuss the fact that for a spin-momentum product state be transformed into a entangled state it needs coherence\nbetween themomentumstates. Otherwise, ifthemomentumstateiscompletelypredictable, thespin-momentumstate\nremains separable, and the Lorentz boost will at most generate superposition between the spin states. In addition,\nwe discuss similar results for the two-particle states under Lorentz boosts. As already noticed in Refs. [8, 21], the\nstate and entanglement changes of the different degrees of freedom depend considerably on the initial states involved,\nas well as on the geometry of the boost scenario. Whereas some states and geometries leave the overall entanglement\ninvariant, others create entanglement.\n\u0003Electronic address: marcoslwbasso@mail.ufsm.br\nyElectronic address: jonas.maziero@ufsm.brarXiv:2007.14480v5 [quant-ph] 31 Aug 20212\nBesides, it is known that entanglement is just one of the aspects that completely characterizes a quanton [24]. The\nother two, which also are intriguing characteristics that turn apart the quantum world from the classical world, are\nknown as the wave-particle duality. This distinguished aspect is generally captured, in a qualitative way, by Bohr’s\ncomplementarity principle [23]. For instance, in the Mach-Zehnder interferometer or in the double-slit interferometer,\nthe wave aspect is characterized by interference fringes visibility, meanwhile the particle nature is given by the which-\nway information of the path along the interferometer. A quantitative version of the wave-particle duality was first\ninvestigated by Wooters and Zurek [25], and later captured by a complementarity inequality in Refs. [26, 27]:\nP2+V2\u00141; (1)\nwherePis an a priori predictability, for which the particle aspect is inferred once the quanton is more likely to\nfollow one path than the other and it is directly related to the probability distribution given by the diagonal elements\nof the density operator as we will discuss in further section. Besides, Vis the visibility of the interference pattern.\nRecently, several steps have been taken towards the quantification of the wave-particle duality, with the establishment\nof minimal and reasonable conditions that visibility and predictability measures should satisfy [29, 30]. As well, with\nthe development of the field of quantum information, it was suggested that the quantum coherence [31] would be\na good generalization of the visibility measure [32–35]. Until now, many approaches were taken for quantifying the\nwave-particle properties of a quantum system [36–40]. As pointed out by Qian et al. [41], complementarity relations\nlike Eq. (1) do not really predict a balanced exchange between PandVsimply because the inequality permits a\ndecrease of PandVtogether, or an increase by both. It even allows the extreme case P=V= 0to occur (neither\nwave or particle) while, in an experimental setup, we still have a quanton on hands. Such a quanton cannot be\nnothing. Thus, one can see that something must be missing from Eq. (1). As noticed by Jakob and Bergou [42], this\nlack of knowledge about the system is due to entanglement, or, more generally, to quantum correlations [43]. This\nmeans that the information is being shared with another system and this kind of quantum correlation can be seen as\nresponsible for the loss of purity of each subsystem such that, for pure maximally entangled states, it is not possible\nto obtain information about the local properties of the subsystems. Hence, to completely quantify a quanton, one has\nalso to regard its correlations with other systems, such that the entire system is pure.\nIn this paper, we study how these different aspects of a quanton behave under Lorentz boosts. Even though\nentanglement entropy does not remain invariant under Lorentz boosts, and neither do measures of predictability and\ncoherence, we show that these three measures together, in what is known as a complete complementarity relation\n(CCR), are Lorentz invariant. In Ref. [11], the authors showed that, even though it is possible to formally define spin\nin any Lorentz frame, there is no relationship between the observable expectation values in different Lorentz frames.\nHere the situation is different. First, one can define complementary relations in any Lorentz frame, but there is no\nobvious transformation law relating complementary relations in different frames. However since the purity of a state\nis preserved under transformations between inertial frames, the complementary relations have the same value in any\nLorentz frame, i.e., there is a transformation law connecting the complete complementarity relations. In addition, we\nexplore several relativistic scenarios for single and two particle states, what helps in understanding the exchange of\nthese different aspects of a quanton under Lorentz boosts.\nThe organization of this article is as follows. In Sec. II, we discuss the representations of the Poincaré group in\nthe Hilbert space, as well as the Wigner’s little group, by focusing in spin- 1=2massive particles. In Sec. III, we\nobtain complete complementarity relations for multipartite pure quantum systems, and show that CCR are Lorentz\ninvariant. Thereafter, in Sec. IV, we turn to the study of the behavior of CCR in relativistic scenarios for several\nsingle and two particle states. Lastly, in Sec. V, we give our conclusions.\nII. REPRESENTATIONS OF THE POINCARÉ GROUP IN THE HILBERT SPACE\nOne of the fundamental questions when studying the relativistic formulation of the quantum theory is how quantum\nstates behave under Lorentz boosts. In the language of group theory, we are seeking to represent an element of the\nLorentz group by a unitary operator on the Hilbert space that the quantum states belongs to. More specifically, single\nparticle quantum states are classified by their transformation under the inhomogeneous Lorentz group, or Poincaré\ngroup, which consists of homogeneous Lorentz transformations \u0003and translations a[44]. For our discussion, we adopt\nthe following notation: Greek indices run over the 4-spacetime coordinate labels f0;1;2;3g; Latin indices run over the\nthree spacial coordinates labels f1;2;3g; the Minkowski metric \u0011\u0016\u0017is diagonal with elements f\u00001;1;1;1g;4-vectors\nare in un-boldfaced type while spacial vectors are represented by an arrow. For instance, the 4-momentum for a\nparticle with mass mis given byp= (p0;p1;p2;p3) = (p0;~ p), with norm p2:=p\u0016p\u0016=\u0011\u0016;\u0017p\u0017p\u0016=\u0000(p0)2+~ p2=\u0000m2,\nwhere we use natural units, i.e., c=~= 1.\nAn inertial reference frame Ois related to another inertial frame O0via a Poincaré transformation\nx0\u0016:=T(\u0003;a)x\u0017= \u0003\u0016\n\u0017x\u0017+a\u0016; (2)3\nwithx= (x0;~ x)being the coordinates of O, and similarly for O0. ThenT(\u0003;a)induces a unitary transformation on\nquantum states characterized by\nj\ti!U(\u0003;a)j\ti; (3)\nwhich satisfies the same composition rule of T(\u0003;a):\nU(\u00031;a1)U(\u00032;a1) =U(\u00031\u00032;\u00031a2+a1): (4)\nSingle particle quantum states can be denoted by jpi\nj\u001bi:=jp;\u001bi, whereplabels the 4-momenta and \u001blabels the\nspin for massive particles. The quantum states jp;\u001biare eigenvectors of the momentum operator P\u0016with eigenvalues\np\u0016, i.e.,P\u0016jp;\u001bi=p\u0016jp;\u001bi. This corresponds to a basis of plane waves and, thus, transforms under translations as\nU(I;a)jp;\u001bi:=U(a)jp;\u001bi=e\u0000ipajp;\u001bi, wherepa:=p\u0016a\u0016=p\u0016a\u0016. Meanwhile, a general Lorentz boost \u0003takes\nthe eigenvalue p\u0016!\u0003\u0016\n\u0017p\u0017, and therefore U(\u0003;0)jp;\u001bi:=U(\u0003)jp;\u001bimust be a linear combination of all states with\nmomentum \u0003p, i.e.,\nU(\u0003)jp;\u001bi=X\n\u0015D\u0015;\u001b(\u0003;p)j\u0003p;\u0015i: (5)\nAsU(\u0003)is a representation, it preserves the group structure, and imposes conditions on the values of D\u0015;\u001b. To see\nthis, let us recall that U(\u0003)leavesp2:=p\u0016p\u0016=~ p\u0001~ p\u0000E2=\u0000m2and the sign of p0=Eunchanged for a particle\nwith massm. Hence, we can use these two invariant quantities to classify states into specific classes. For each value\nofp2and for each sign(p0), it is possible to choose a ‘standard’ 4-momentum kthat identifies a specific class of\nquantum states [45]. For massive particles, we can fix the standard momentum kto be the particle’s momentum in\nthe rest frame, i.e., k= (m;0;0;0). Then, any momenta pcan be expressed in terms of the standard momentum, i.e.,\np\u0016= (L(p)k)\u0016=L(p)\u0016\n\u0017k\u0017, whereL(p)is a Lorentz transformation which depends on pand takesk!p. Therefore,\nquantum statesjp;\u001bican be defined in terms of the standard momentum state jk;\u001bi:\njp;\u001bi=U(L(p))jk;\u001bi: (6)\nNow, if we apply a Lorentz boost \u0003onjp;\u001bi, then\nU(\u0003)jp;\u001bi=U(\u0003)U(L(p))jk;\u001bi (7)\n=U(I)U(\u0003L(p))jk;\u001bi (8)\n=U(L(\u0003p)L\u00001(\u0003p))U(\u0003L(p))jk;\u001bi (9)\n=U(L(\u0003p))U(L\u00001(\u0003p)\u0003L(p))jk;\u001bi (10)\n=U(L(\u0003p))U(W(\u0003;p))jk;\u001bi; (11)\nwhereW(\u0003;p) =L\u00001(\u0003p)\u0003L(p)is called Wigner rotation, which leaves the standard momentum kinvariant, and\nonly acts on the internal degrees of freedom of jk;\u001bi:kL\u0000 !p\u0003\u0000 !\u0003pL\u00001\n\u0000\u0000\u0000!k. Hence, the final momentum in the rest\nframe is different from the original one by a Wigner rotation, i.e., U(W(\u0003;p))jk;\u001bi=P\n\u0015D\u0015;\u001b(W(\u0003;p))jk;\u0015i. On\nthe other hand, U(L(\u0003p))takesk!\u0003pwithout affecting the spin, by definition. Therefore,\nU(\u0003)jp;\u001bi=U(L(\u0003p))U(W(\u0003;p))jk;\u001bi (12)\n=U(L(\u0003p))X\n\u0015D\u0015;\u001bW(\u0003;p))jk;\u0015i (13)\n=X\n\u0015D\u0015;\u001b(W(\u0003;p))j\u0003p;\u0015i: (14)\nIt is worth mentioning that the subscripts of D\u0015;\u001b(W(\u0003;p))can be suppressed, and we can write U(\u0003)jp;\u001bi=\nj\u0003pi\nD(W(\u0003;p))j\u001bi. The set of Wigner rotations forms a group known as the little group , which is a subgroup of\nthe Poincaré group [46]. In other words, under a Lorentz transformation \u0003, the momenta pgoes to \u0003p, and the spin\ntransforms under the representation D(W(\u0003;p))of the little group W. For massive particles, the little group is the\nwell known group of rotations in three dimensions, SO(3). However, it is also known that SO(3)is homomorphic to\nSU(2), and the irreducible unitary representations of SU(2)span a Hilbert space of 2j+ 1dimensions, with j=n=2,\nwherenis an integer [47, 48]. The value of jis what we usually refer to as the spin of the massive particle. In this\narticle, we will be interested in spin- 1=2particles, hence the representation of the Wigner rotation is given by [49, 50]\nD(W(\u0003;p)) =(p0+m) cosh(!=2)I2\u00022+ (~ p\u0001^e) sinh(!=2)\u0000isinh(!=2)~ \u001b\u0001(~ p\u0002^e)p\n(p0+m)((\u0003p)0+m)(15)\n= cos\u001e\n2I2\u00022+isin\u001e\n2(~ \u001b\u0001^n); (16)4\nwithI2\u00022being the identity matrix, meanwhile ~ \u001bare the Pauli matrices, and\ncos\u001e\n2=cosh(!=2) cosh(\u000b=2) + sinh(!=2) sinh(\u000b=2)(^e\u0001^p)q\n1\n2(1 + cosh!cosh\u000b+ sinh!sinh\u000b(^e\u0001^p)); (17)\nsin\u001e\n2^n=sinh(!=2) sinh(\u000b=2)(^e\u0002^p)q\n1\n2(1 + cosh!cosh\u000b+ sinh!sinh\u000b(^e\u0001^p)); (18)\nwhere cosh\u000b=p0=m,!= tanh\u00001vis the rapidity of the boost [51], ^eis the unit vector pointing in the direction of\nthe boost,pis the 4-momenta of the particle in O, and \u0003pis the 4-momenta of the particle in O0. As an example, if\nthe momentum is in the x-direction of the referece frame Oand the boost is given in the z-axis, then\nD(W(\u0003;p)) = cos\u001e\n2I2\u00022\u0000isin\u001e\n2\u001by=\u0012\ncos\u001e\n2\u0000sin\u001e\n2\nsin\u001e\n2cos\u001e\n2;\u0013\n(19)\nwhere the Wigner angle \u001eis given by\ncos\u001e\n2=cosh(!=2) cosh(\u000b=2)q\n1\n2(1 + cosh!cosh\u000b); (20)\nsin\u001e\n2^n=sinh(!=2) sinh(\u000b=2)^yq\n1\n2(1 + cosh!cosh\u000b); (21)\ntan\u001e=sinh!sinh\u000b\ncosh!+ cosh\u000b; (22)\nwhich implies that \u001e2[0;\u0019=2]. Hence, the transformation law for the spin- 1=2particle with momentum ~ palong the\nx-axis ofOis given by\nU(\u0003)jp;0i=j\u0003pi\n(cos\u001e\n2j0i+ sin\u001e\n2j1i); (23)\nU(\u0003)jp;1i=j\u0003pi\n(\u0000sin\u001e\n2j0i+ cos\u001e\n2j1i); (24)\nwherej0imeans spin ‘up’ and j1istands for spin ‘down’ along the z-axis. Therefore, as one can see, for separable\nand completely predictable states, a Lorentz boost will only generate superposition between the possible states of the\nspin of the particle, as already noticed in Ref. [45].\nIII. THE LORENTZ INVARIANCE OF COMPLETE COMPLEMENTARITY RELATIONS\nIn Ref. [43], we developed a general framework to obtain complete complementarity relations for a subsystem that\nbelongs to an arbitrary multipartite pure quantum system, just by exploring the purity of the multipartite quantum\nsystem. To make our investigation easier, we begin by assuming that momenta can be treated as discrete variables\n[16, 18, 21]. This can be justified once we can consider narrow distributions centered around different momentum\nvalues such that it is possible to represent them by orthogonal state vectors, i.e., hpijpji=\u000ei;j. Although narrow\nmomenta are an idealization, it is a system worth studying since it helps to understand more realistic systems, and,\nalso, it is possible to approximate continuous momenta as a finite (but large) number of discrete momenta. Also,\nthroughout this article, we will consider only massive particles of spin 1=2. By doing this, we are considering a\nparticular representation of the Wigner little group. However, the result obtained in this section does not depend on\nthe particular choice of representation, once the representation is unitary.\nSo, let us consider nmassive quantons with spin 1=2in a pure state described by j\tiA1;:::;A 2n2H 1\n:::\nH 2n\nwith dimension d=dA1dA2:::dA2n, in the reference frame O. For instance, A1,A2are referred as the momentum and\nspin of the first quanton, and so on. By defining a local orthonormal basis for each degree of freedom (DOF) Am,\nfjimiAmgdm\u00001\ni=0,m= 1;:::;2n, the state of the multipartite quantum system can be written as [52]\n\u001aA1;:::;A 2n=j\tiA1;:::;A 2nh\tj=X\ni1;:::;i 2nX\nj1;:::;j 2n\u001ai1:::i2n;j1:::j2nji1;:::;i 2niA1;:::;A 2nhj1;:::;j 2nj: (25)5\nWithout loss of generality, let us consider the state of the DOF A1, which is obtained by tracing over the other\nsubsystems,\n\u001aA1=X\ni1;j1\u001aA1\ni1;j1ji1iA1hj1j=X\ni1;j1X\ni2;:::;j 2n\u001ai1i2:::i2n;j1i2:::i2nji1iA1hj1j; (26)\nforwhichtheHilbert-Schmidtquantumcoherenceandthecorrespondingpredictabilitymeasureintermsofthedensity\nmatrix elements are given by\nChs(\u001aA1) =X\ni16=j1\f\f\f\u001aA1\ni1;j1\f\f\f2\n=X\ni16=j1\f\f\f\f\f\fX\ni2;:::;i 2n\u001ai1i2:::i2n;j1i2:::i2n\f\f\f\f\f\f2\n; (27)\nPl(\u001aA1) =X\ni1(\u001aA1\ni1;i1)2\u00001=dA1=X\ni1(X\ni2;:::;i 2n\u001ai1i2:::i2n;i1i2:::i2n)2\u00001=dA1: (28)\nBesides, such predictability measure can be first defined as Pl(\u001aA1) :=Smax\nl\u0000Sl(\u001aA1diag)[40], where \u001aA1diagcor-\nresponds to the diagonal elements of \u001aA1andSl(\u001a) := 1\u0000Tr\u001a2is the linear entropy. To get more intuition about\nthis predictability measure, let us consider the projector onto the state index i1:\u0005i1:=ji1ihi1j, which can be\none of the paths of a Mach-Zehnder interferometer. Now, the uncertainty of the state i1is given by its variance\nV(\u001aA1;\u0005i1) =\n\u00052\ni1\u000b\n\u0000h\u0005i1i2=\u001aA1\ni1i1\u0000(\u001aA1\ni1i1)2such that sum of the uncertainties of all the possible states (or paths)\nis given byP\ni1V(\u001aA1;\u0005i1) = 1\u0000P\nj(\u001aA1\ni1i1)2;which represents the total uncertainty of all states. One can easily see\nthatP\njV(\u001aA1;\u0005j)is exactly the linear entropy of \u001aA1diag:Sl(\u001aA1diag) = 1\u0000Tr\u001a2\nA1diag. Thus, after repeating the\nsame experiment several times, we obtain a probability distribution given by \u001aA1\n00;:::;\u001aA1\ndA1\u00001dA1\u00001;which represents\na probability of the quanton being measured in the state j0i;:::;jdA1\u00001i. From this, it is possible to calculate the\nuncertainty about the paths through Sl(\u001adiag)such thatPl(\u001aA1diag) :=Smax\nl\u0000Sl(\u001aA1diag)offers a measure of the\ncapability to predict what outcome will be obtained in the next run of the experiment. For instance, if we obtain a\nuniform probability distribution, i.e., f\u001aA1\ni1i1= 1=dA1gdA1\u00001\ni1=0, after repeating the experiment several times, our ability\nto make a prediction is null. On the other hand, if \u001aA1diag6=IA1=dA1, whereIA1is the identity operator, then\nPhs(\u001aA1)6= 0. Therefore, it is possible to see that Eq.(28) is a way of quantifying how much the probability distri-\nbution expressed by \u001aA1diagdiffers from the uniform probability distribution. While the Hilbert-Schmidt quantum\ncoherence can be defined as Chs(\u001aA1) := min\u00132Ijj\u001aA1\u0000\u0013jj2\nhs, whereIis the set of all incoherent states (diagonal\ndensity operators), and the Hilbert-Schmidt’s norm of a matrix M2Cd\u0002dis defined askMkhs:=qP\nj;kjMjkj2. The\nminimization procedure yields Eq.(27). Therefore, the Hilbert-Schmidt quantum coherence is measuring how distant\nthe density operator \u001aA1is in comparison with its closest incoherent state, which in this case is \u001aA1diag, given the\nHilbert-Schmidt’s norm. Loosely speaking, the quantum coherence is measuring “how much” orthogonal superposi-\ntion is encoded in a given density operator \u001aA1and it is directly related to the off-diagonal elements of \u001aA1. Besides,\nwe showed in Ref. [40] that these are bona-fide measures of visibility and predictability, respectively. From these\nequations, an incomplete complementarity relation, Phs(\u001aA1) +Chs(\u001aA1)\u0014(dA1\u00001)=dA1, is obtained by exploring\nthe mixedness of \u001aA1, i.e., 1\u0000Tr\u001a2\nA1\u00150.\nNow, since \u001aA1;:::;A 2nis a pure quantum system, then 1\u0000Tr\u001a2\nA1;:::;A 2n= 0, or equivalently,\n1\u0000\u0010X\n(i1;:::;i 2n)=(j1;:::;j 2n)+X\n(i1;:::;i 2n)6=(j1;:::;j 2n)\u0011\nj\u001ai1i2:::i2n;j1j2:::j2nj2= 0; (29)\nwhere\nX\n(i1;:::;i 2n)6=(j1;:::;j 2n)=X\ni16=j1\ni2=j2\n...\ni2n=j2n+X\ni1=j1\ni26=j2\n...\ni2n=j2n+:::+X\ni1=j1\ni2=j2\n...\ni2n6=j2n+X\ni16=j1\ni26=j2\n...\ni2n=j2n+:::+X\ni16=j1\ni2=j2\n...\ni2n6=j2n+:::+X\ni16=j1\ni26=j2\n...\ni2n6=j2n: (30)\nThe purity condition (29) can be rewritten as a CCR:\nPl(\u001aA1) +Chs(\u001aA1) +Sl(\u001aA1) =dA1\u00001\ndA1; (31)6\nwithSl(\u001aA1)being the linear entropy of the subsystem A1given by\nSl(\u001aA1) :=X\ni16=j1X\n(i2;:::;i 2n)6=(j2;:::;j 2n)\u0010\nj\u001ai1i2:::i2n;j1j2:::j2nj2\u0000\u001ai1i2:::i2n;j1i2:::i2n\u001a\u0003\ni1j2:::j2n;j1j2:::j2n\u0011\n:(32)\nIt is worthwhile mentioning the CCR in Eq. (31) is a natural generalization of the complementarity relation obtained\nby Jakob and Bergou [53, 54] for bipartite pure quantum systems. More generally, E=p\n2Sl(\u001aA1), whereEis the\ngeneralized concurrence obtained in Ref. [55] for multi-particle pure states. Now, for the boosted observer O’ of Sec.\nII, the same nmassive quantons system is described by j\t\u0003iA1;:::;A 2n=U(\u0003)j\tiA1;:::;A 2n, and the density matrix of\nthe multipartite pure quantum system can be written as [56, 57]\n\u001a\u0003\nA1;:::;A 2n=j\t\u0003iA1;:::;A 2nh\t\u0003j=U(\u0003)\u001aA1;:::;A 2nUy(\u0003); (33)\nwhich implies that Tr\u0000\n\u001a\u0003\nA1;:::;A 2n\u00012= Tr(\u001aA1;:::;A 2n)2, and the whole system remains pure under the Lorentz\nboost. As we used the purity of the density matrix to obtain the complete complementarity relation, then, from\n1\u0000Tr\u0000\n\u001a\u0003\nA1;:::;A 2n\u00012= 0, we can obtain\nPl(\u001a\u0003\nA1) +Chs(\u001a\u0003\nA1) +Sl(\u001a\u0003\nA1) =dA1\u00001\ndA1: (34)\nThis proves our claim that CCR are invariant under Lorentz transformations. For continuous momenta, the result\nshowed here remains valid if applied to the discrete degrees of freedom with the continuous momenta been traced\nout, since we defined Pl;ChsandSlonly for the discrete degrees of freedom. Therefore, one can see that there is a\ntransformation law connecting the CCRs for different Lorentz frames.\nIV. RELATIVISTIC SETTINGS\nA. Single-particle system scenarios\nWe begin by considering three different single-particle states where the particle is moving in two opposing directions\nalong theyaxis and the spins are aligned with the zaxis irrespective of the direction of the boost in the reference\nframeO:\nj\ti=1p\n2(jpi+j\u0000pi)\nj0i; (35)\nj\u0004i=1p\n2(jp;0i+j\u0000p;1i); (36)\nj\bi=1\n2(jpi+j\u0000pi)\n(j0i+j1i); (37)\nwherej0imeans spin ‘up’, and j1ispin ‘down’. In addition, j\u0000piis describing the state whose spatial momentum has\nopposite direction in comparison with jpi. It is worthwhile mentioning that the states j\ti;j\biare separable states,\nwhilej\u0004iis a maximal entangled state. Moreover, the state j\tihas maximal coherence in the momentum degree of\nfreedom, and maximal predictability in the spin degree of freedom. Meanwhile j\biis maximally coherent in both\ndegrees of freedom, and j\u0004ihas no local properties. Now, let us consider an observer O’ boosted with velocity vin\na direction orthogonal to the momentum of the particle in the frame O, i.e., in the x\u0000zplane, making an angle\n\u00122[0;\u0019=2]with thex-axis. Hence, the direction of boost is given by ^e= cos\u0012^x+ sin\u0012^z, and the Wigner rotation\nfollows directly:\nD(W(\u0003;\u0006p)) = cos\u001e\n2I2\u00022+isin\u001e\n2(\u0007sin\u0012\u001bx\u0006cos\u0012\u001bz) (38)\n=\u0012\ncos\u001e\n2\u0006isin\u001e\n2cos\u0012\u0007isin\u001e\n2sin\u0012\n\u0007isin\u001e\n2sin\u0012 cos\u001e\n2\u0007isin\u001e\n2cos\u0012\u0013\n; (39)7\n0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6\n0.00.10.20.30.40.5Sl\n(a)Sl(\u001a\u0003s) =Sl(\u001a\u0003p)as a function of the\nWigner angle.\n0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6\n0.00.10.20.30.40.5Pl\n=0\n=/8\n=/4\n=/3\n=/2\n(b)Pl(\u001a\u0003s)as a function of the Wigner\nangle.\n0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6\n0.000.020.040.060.080.100.12Chs(c)Chs(\u001a\u0003s)as a function of the Wigner\nangle.\n0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6\n0.00.10.20.30.40.5Chs\n(d)Chs(\u001a\u0003p)as a function of the Wigner\nangle.\nFigure 1: The different aspects of the degrees of freedom in the state j\t\u0003ifor different values of \u0012.\nsince\u0006^p=\u0006^y. Therefore, the observer in O’ assigns in general a different state to the same system. For instance,\nthe state given by Eq. (35) in O’ is described by\nj\t\u0003i=U(\u0003)j\ti=1p\n2(j\u0003pi\nD(W(\u0003;p))j0i+j\u0000\u0003pi\nD(W(\u0003;\u0000p))j0i) (40)\n=1p\n2\u0010\nj\u0003pi[(cos\u001e\n2+isin\u001e\n2cos\u0012)j0i\u0000isin\u001e\n2sin\u0012j1i] +j\u0000\u0003pi[(cos\u001e\n2\u0000isin\u001e\n2cos\u0012)j0i+isin\u001e\n2sin\u0012j1i]\u0011\n;\n(41)\nwhich in general is an entangled state. The reduced density matrix of the spin (momentum) is obtained by tracing\nout the momentum (spin) states:\n\u001a\u0003s= Tr \u0003pj\t\u0003ih\t\u0003j=\u0012\ncos2\u001e\n2+ sin2\u001e\n2cos2\u0012\u0000sin2\u001e\n2sin\u0012cos\u0012\n\u0000sin2\u001e\n2sin\u0012cos\u0012 sin2\u001e\n2sin2\u0012\u0013\n; (42)\n\u001a\u0003p= Tr \u0003sj\t\u0003ih\t\u0003j=\u00121\n21\n2(cos\u001e+isin\u001ecos\u0012)\n1\n2(cos\u001e\u0000isin\u001ecos\u0012)1\n2\u0013\n: (43)\nIn Fig. 1, we plotted the different aspects of the degrees of freedom of the quanton for different values of \u0012. For\ninstance, if there is no boost, i.e., \u001e= 0, the state remains unchanged, regardless the direction of the boost. Also, if\nthe boost is along the x-axis,\u0012= 0, the state remains the same. Now, if the boost is along the z-axis, the entanglement\nbetween the moment and the spin of the particle increases with the increase of the Wigner angle. In exchange, the\ncoherence of the momentum and the predictability of the spin decrease with \u001e. Beyond that, for any \u0012;\u001e2[0;\u0019=2],\nthe complete complementarity relation Phs+Chs+Sl= 1=2is always satisfied.\nNow, the statej\u0004igiven by Eq. (36) is described in O0as\nj\u0004\u0003i=1p\n2\u0010\nj\u0003pi[(cos\u001e\n2+isin\u001e\n2cos\u0012)j0i\u0000isin\u001e\n2sin\u0012j1i] +j\u0000\u0003pi[isin\u001e\n2sin\u0012j0i+ (cos\u001e\n2+isin\u001e\n2cos\u0012)j1i]\u0011\n;\n(44)\nwhile the reduced density matrices are given by\n\u001a\u0003s=\u001ay\n\u0003p=\u00121\n2icos\u001e\n2sin\u001e\n2sin\u0012\n\u0000icos\u001e\n2sin\u001e\n2sin\u00121\n2\u0013\n: (45)8\n0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6\n0.00.10.20.30.40.5Sl\n=0\n=/8\n=/4\n=/3\n=/2\n(a)Sl(\u001a\u0003j),j=s;p, as a function of the\nWigner angle.\n0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6\n0.00.10.20.30.40.5Chs(b)Chs(\u001a\u0003j),j=s;p, as a function of the\nWigner angle.\nFigure 2: The different aspects of the degrees of freedom of the quanton in state the j\u0004\u0003ifor different values of \u0012.\nIn this example, by inspecting Fig. 2, if there is no boost, i.e., \u001e= 0the state remains unchanged, regardless\nthe direction of the boost. Also, if the boost is along the x-axis,\u0012= 0, the state remains the same. However, for\n\u00122(0;\u0019=2]and\u001e6= 0, there is an increase of the coherence of both degrees of freedom, in exchange of the consumption\nof the entanglement between the momenta and spin of the particle. In the extreme case where \u0012=\u0019=2and\u001e!\u0019=2,\nboth degrees of freedom have maximal coherence and the state j\u0004\u0003ibecomes separable\nj\u0004\u0003i\u001e=\u0012=\u0019=2=1\n2(j\u0003pi+ij\u0000\u0003pi)\n(j0i\u0000ij1i): (46)\nLastly, in the boosted frame O0, the statej\bigiven by Eq. (37) is described by\nj\b\u0003i=1\n2\u0010\nj\u0003pif[cos\u001e\n2+isin\u001e\n2(cos\u0012\u0000sin\u0012)]j0i+ [cos\u001e\n2\u0000isin\u001e\n2(cos\u0012+ sin\u0012)]j1ig (47)\n+j\u0000\u0003pif[cos\u001e\n2\u0000isin\u001e\n2(cos\u0012\u0000sin\u0012)]j0i+ [cos\u001e\n2+isin\u001e\n2(cos\u0012+ sin\u0012)]j1ig\u0011\n; (48)\nwith the reduced density matrices being\n\u001a\u0003s=\u00121\n21\n2(cos2\u001e\n2\u0000sin2\u001e\n2cos 2\u0012)\n1\n2(cos2\u001e\n2\u0000sin2\u001e\n2cos 2\u0012)1\n2\u0013\n; (49)\n\u001a\u0003p=\u00121\n21\n2(cos\u001e\u0000isin\u001esin\u0012)\n1\n2(cos\u001e+isin\u001esin\u0012)1\n2\u0013\n: (50)\nIn contrast with the second example, here the entanglement between momentum and spin increases with the Wigner\nangle, in exchange of the consumption of the coherence of both degrees of freedom. However, for \u0012=\u0019=2the state is\nseparable:\nj\b\u0003i\u001e=\u0019=2=1\n2(Aj\u0003pi+A\u0003j\u0000\u0003pi)\n(j0i+j1i); (51)\nwhereA=cos\u001e\n2\u0000isin\u001e\n2andA\u0003is the complex conjugate of A. The coherence and entropy of the momentum has\nthe same qualitative behavior as for the spin, which is plotted in Fig. 3.\nFrom these examples, we can see if the state of the system is separable, and has no superposition between the mo-\nmentum states in the reference frame O, then a Lorentz boost cannot generate entanglement between the momentum\nand spin degrees of freedom. This result helps to explain the reported generation of entanglement between momenta\nand spin in one of the first studies of single-particle systems carried out by Peres et al. [11], where they considered\na particle with mass mwhose momentum wave function in the rest frame is given by (~ p) = (2\u0019)\u00003=4w3=2e\u0000~ p2=2w2,\nwithwbeing the width of the wave packet. This is a Gaussian state of minimum uncertainty, but still represents a\ncontinuous superposition. Therefore, in this case it is possible to generate entanglement via a Lorentz transformation.\nB. Two-particle system scenarios\nAs showed in Sec. IVA, if the momentum states have no coherence, then it is not possible to generate entanglement\nbetween the momenta and spin degrees of freedom. In this section, we begin by discussing this issue for the two-9\n0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6\n0.00.10.20.30.40.5Sl\n(a)Sl(\u001a\u0003s)as a function of the Wigner\nangle.\n0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6\n0.000.020.040.060.080.100.12Pl=0\n=/8\n=/4\n=/3\n=/2\n(b)Pl(\u001a\u0003s)as a function of the Wigner\nangle.\n0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6\n0.00.10.20.30.40.5Chs(c)Chs(\u001a\u0003s)as a function of the Wigner\nangle.\nFigure 3: The different aspects of the spin of the quanton in the state j\b\u0003ifor different values of \u0012.\nparticle case and end this section giving two examples. Now, let us consider a two-particle state described in O\nas\nj\tiA;B=X\n\u001b;\u0015 \u001b \u0015jp;qiA;B\nj\u001b;\u0015iA;B; (52)\nwhereP\njj jj2= 1forj=\u001b;\u0015. In addition,jp;qiA;B=jpiA\njqiBdenotes the momentum state of particle A and\nB, respectively, meanwhile j\u001b;\u0015iA;B=j\u001biA\nj\u0015iBrepresents the state of the spins of the particle A and B. The state\nj\tiA;Bis separable and has no coherence between momentum states in the reference frame O. Now, in the boosted\nframeO0, we havej\t\u0003iA;B=U(\u0003)j\tiA;B, i.e.,\nj\t\u0003iA;B=X\n\u001b;\u0015 \u001b \u0015j\u0003p;\u0003qiA;B\nD(W(\u0003;p))j\u001biA\nD(W(\u0003;q))j\u0015iB(53)\n=j\u0003p;\u0003qiA;B\nX\n\u001b \u001bD(W(\u0003;p))j\u001biA\nX\n\u0015 \u0015D(W(\u0003;q))j\u0015iB; (54)\nwhich is also separable. In this case, the Wigner rotation will only change the coherences of the spin states of the\nparticles A and B. Now, let’s consider a state in Owith superposition in the momentum states of the particle A\nj\biA;B=X\np (p)jp;qiA;B\nj\u001b;\u0015iA;B; (55)\nwithP\npj (p)j2= 1. Then, a Lorentz boost can generate entanglement between the momentum and spin of the\nparticle A\nj\b\u0003iA;B=X\np \u001b(p)j\u0003piA\nD(W(\u0003;p))j\u001biA\nj\u0003qiB\nD(W(\u0003;q))j\u0015iB: (56)\nHowever, there is no entanglement between particles A and B. Similarly, if we consider that the state in Ohas\ncoherence in the momentum states of A and B, there will be no entanglement between particles A and B. To obtain\nan entangled state of the whole system in O0, we have to consider a state in Oalready entangled in the momentum\ndegrees of freedom, i.e.,\nj\u0004iA;B=X\np;q (p;q)jp;qiA;B\nj\u001b;\u0015iA;B; (57)\nwithP\np;qj (p;q)j2= 1. Hence, in boosted frame O0we have\nj\u0004\u0003iA;B=X\np;q (p;q)j\u0003p;\u0003qiA;B\nD(W(\u0003;p))j\u001biA\nD(W(\u0003;q))j\u0015iB: (58)\nFor instance, if we consider the particles moving in two opposing directions along the y axis and the spins are\naligned with the zaxis irrespective of the direction of the boost in the reference frame O,\nj\u0004iA;B=1p\n2(jp;\u0000pi+j\u0000p;pi)\nj0;0i; (59)10\n0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6\n0.00.10.20.30.40.5Sl\n(a)Sl(\u001a\u0003s)as a function of the Wigner\nangle.\n0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6\n0.00.10.20.30.40.5Pl\n=0\n=/8\n=/4\n=/3\n=/2\n(b)Pl(\u001a\u0003s)as a function of the Wigner\nangle.\n0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6\n0.000.020.040.060.080.100.12Chs(c)Chs(\u001a\u0003s)as a function of the Wigner\nangle.\n0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6\n0.00.10.20.30.40.5Chs,Sl\nChs\nSl\n(d)Chs(\u001a\u0003p\u0003p)as a function of the\nWigner angle.\nFigure 4: The different aspects of j\u0004\u0003ifor different values of \u0012.\nas before, the observer O’ is boosted with velocity vin a direction orthogonal to the momentum of the particle in the\nframeO, i.e., in the x\u0000zplane, making an angle \u00122[0;\u0019=2]with thex-axis. Then, the Wigner rotation is given by\nEq. (39) and the state in the boosted frame is described by\nj\u0004\u0003iA;B=1p\n2(j\u0003p;\u0000\u0003pi+j\u0000\u0003p;\u0003pi)\n[(cos2\u001e\n2+ sin2\u001e\n2cos2\u0012)j0;0i+ sin2\u001e\n2sin2\u0012j1;1i\n\u0000sin2\u001e\n2sin\u0012cos\u0012(j0;1i+j1;0i)] +icos\u001e\n2sin\u001e\n2sin\u0012p\n2(j\u0003p;\u0000\u0003pi\u0000j\u0000 \u0003p;\u0003pi)\n(j0;1i\u0000j1;0i);(60)\nwhich is an entangled state between all degrees of freedom. In this case, the resource consumed to generate entangle-\nment of the spins of the particles is the bipartite coherence of the reduced momentum-momentum density matrix of\nthe particles A and B\n\u001a\u0003p;\u0003p=1\n2(j\u0003p;\u0000\u0003pih\u0003p;\u0000\u0003pj+j\u0000\u0003p;\u0003pih\u0000\u0003p;\u0003pj) +1\n2(cos4\u001e\n2+ sin4\u001e\n2)(j\u0003p;\u0000\u0003pih\u0000\u0003p;\u0003pj+t:c:);(61)\nwhere t.c. stands of transpose conjugated. In Fig. 4(d), we plotted the coherence and the linear entropy of \u001a\u0003p;\u0003p\nas a function of \u001e, whereSl(\u001a\u0003p;\u0003p)is measuring the entanglement of \u001a\u0003p;\u0003pas a whole with rest of the degrees of\nfreedom. Meanwhile, the concurrence measure [58] Eof\u001a\u0003p;\u0003pdecreases monotonically with Wigner angle, which\nmeans the momentum-momentum entanglement decreases with \u001e, onceE(\u001a\u0003p\u0003p) =p\n2Chs(\u001a\u0003p\u0003p). In addition,\nFigs. 4(a), 4(b) and 4(c) represent the behavior of the different aspects of the spin of particle A. The aspects of the\nspin of particle B display similar behavior. It is worth emphasizing that it is not the entanglement between spin-spin\nthat increases, but the entanglement of the spin of one of the particles with all the other degrees of freedom. In Ref.\n[8], the authors discussed the generation of spin-spin entanglement for two particles under Lorentz boosts. Meanwhile,\nthe momentum state of particles A and B are given by \u001aA\n\u0003p=\u001aB\n\u0003p=1\n2(j\u0003pih\u0003pj+j\u0000\u0003pih\u0000\u0003pj), which implies that\nSl(\u001aj\n\u0003p) = 1=2,j=A;B, and the overall entanglement between the momentum of particle A (B) with rest of the\nsystem does not change under the Lorentz boost, even though the entanglement of momentum-momentum decreases.\nHence, in this case, the entanglement of the momentum of particle A (B) is redistributed among the others degrees\nof freedom. For \u001e= 0(no boost), just the momentum of the particles are entangled. In the limit \u001e=\u0019=2, the\nmomentum of the particles are entangled with the spins, therefore the entanglement of momentum-momentum has to\ndecrease.11\n0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6\n0.00.10.20.30.40.5Chs,Sl\nChs(s)\nSl(s)\n(a)Chs(\u001a\u0003s);Sl(\u001a\u0003s)as a function of the\nWigner angle.\n0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6\n0.00.10.20.30.40.5Chs,SlChs(pp)\nSl(pp)\n(b)Chs(\u001a\u0003p\u0003p);Sl(\u001a\u0003p\u0003p)as a function of\nthe Wigner angle.\nFigure 5: The different aspects of j\u0007\u0003i.\nIn contrast, we also can redistribute entanglement to generate coherence in the spin states of the particles A and\nB. For instance, let us consider the following two-particle states in O\nj\u0007iA;B=1p\n2(jp;0iA\nj\u0000p;1iB+j\u0000p;1iA\njp;0iB) =1p\n2(jp;\u0000piA;B\nj0;1iA;B+j\u0000p;piA;B\nj1;0iA;B);(62)\nwith the momentum of the particles along the y-axis. Now, for a boosted frame O’ along the z-axis, the Wigner\nrotation is given by Eq. (39) imposing \u0012=\u0019=2. Hence,\nj\u0007\u0003iA;B=1p\n2icos\u001e\n2sin\u001e\n2(j\u0003p;\u0000\u0003pi+j\u0000\u0003p;\u0003pi)\n(j0;0i\u0000j1;1i) +1p\n2j\u0003p;\u0000\u0003pi\n(cos2\u001e\n2j0;1i\n+ sin2\u001e\n2j1;0i) +1p\n2j\u0000\u0003p;\u0003pi\n(sin2\u001e\n2j0;1i+ cos2\u001e\n2j1;0i): (63)\nThe reduced spin density matrices of each particle are given by\n\u001aA\n\u0003s=\u001aB\n\u0003s=\u00121\n2icos\u001e\n2sin\u001e\n2\n\u0000icos\u001e\n2sin\u001e\n21\n2\u0013\n; (64)\nand\u001aA\n\u0003p=\u001aB\n\u0003p=1\n2I2\u00022, whereI2\u00022is the identity matrix. The entanglement of the spin of the particle A (B) with\nthe rest of the system decreases with the Wigner angle. In exchange, the coherence of the spin of particle A (B)\nincreases, as shown in Fig. 5(a). In addition, the entanglement of \u001a\u0003p\u0003pas a whole with the spins of the particles also\ndecreases with \u001e. From Fig. 5(b), the bipartite coherence increases, since it is related to the momentum-momentum\nentanglement of particle A and B, once E(\u001a\u0003p\u0003p) =p\n2Chs(\u001a\u0003p\u0003p), as we also can see from\n\u001a\u0003p\u0003p=1\n2(j\u0003p;\u0000\u0003pih\u0003p;\u0000\u0003pj+j\u0000\u0003p;\u0003pih\u0000\u0003p;\u0003pj) +\u0012\n2 cos2\u001e\n2sin2\u001e\n2j\u0003p;\u0000\u0003pih\u0000\u0003p;\u0003pj+t:c:\u0013\n:(65)\nHence, the entanglement of the momentum of the particle A (B) with rest of the degrees of the system remains\nthe same under the Lorentz boost, although it is shuffled around among the degrees of freedom. For instance, when\n\u001e= 0(no boost), the momentum of particle A is entangled with all the others degrees of freedom. However, in the\nlimit\u001e=\u0019=2, the momentum of the particle A is entangled just with the momentum of the particle B, since\n\f\f\u0007\u0003\u001e=\u0019=2\u000b\nA;B=1p\n2(j\u0003p;\u0000\u0003piA;B+j\u0000\u0003p;\u0003piA;B)\n1p\n2(ij0iA+j1iA)\n1p\n2(j0iB\u0000ij1iB);(66)\nandSl(\u001a\u0003p\u0003p) = 0for\u001e=\u0019=2.\nV. CONCLUSIONS\nAlthough the entanglement entropy does not remain invariant under Lorentz boosts, and neither do the measures\nof predictability and coherence, we showed in this work that these three measures taken together, in a complete12\ncomplementarity relation (CCR), are Lorentz invariant. Even though it is possible to formally define spin in any\nLorentzframe, thereisnorelationshipbetweentheobservableexpectationvaluesindifferentLorentzframes, according\nto Peres et. al. [11]. Here the situation is quite different. First, it is possible to formally define complementarity in any\nLorentz frame and, in principle, there is no relationship between the complementarity relations in different Lorentz\nframes. However, our results showed that it is possible indeed to connect complete complementarity relations in\ndifferent Lorentz frames. Therefore, we showed how the connection between the CCR defined in different Lorentzian\nframes is possible, and disclosed interesting aspects of the redistribution of quantum features for the relativistic\ndynamics of one- and two-particle states and how the role of CCRs helps to keep tracking of how this redistribution\nof quantum features is done.\nAcknowledgments\nThis work was supported by the Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES), process\n88882.427924/2019-01, and by the Instituto Nacional de Ciência e Tecnologia de Informação Quântica (INCT-IQ),\nprocess 465469/2014-0.\n[1] E. Schrödinger, Discussion of probability relations between separated systems, Math. Proc. Camb. Phil. Soc. 31, 555\n(1935).\n[2] R. Horodecki, P. Horodecki, M. Horodecki, and K. Horodecki, Quantum entanglement, Rev. Mod. Phys. 81, 865 (2009).\n[3] S. Popescu, Bell’s inequalities versus teleportation: What is nonlocality?, Phys. Rev. Lett. 72, 797 (1994).\n[4] J. Preskill, Quantum information and physics: some future directions, J. Mod. Opt. 47, 127 (2000).\n[5] P. S. D. Cavalcanti and I. Šupić, All entangled states can demonstrate nonclassical teleportation, Phys. Rev. Lett. 119,\n110501 (2017).\n[6] D. R. Terno, Introduction to relativistic quantum information, arXiv:quant-ph/0508049 (2006).\n[7] M. Czachor, Einstein-Podolsky-Rosen-Bohm experiment with relativistic massive particles, Phys. Rev. A 55, 72 (1997).\n[8] R. M. Gingrich and C. Adami, Quantum entanglement of moving bodies, Phys. Rev. Lett. 89, 270402 (2002).\n[9] H. Terashima and M. Ueda, Relativistic Einstein-Podolsky-Rosen correlation and Bell’s inequality, Int. J. Quantum Inform.\n01, 93 (2003).\n[10] P. M. Alsing and G. J. Milburn, Lorentz invariance of entanglement, Quantum Inf. Comput. 2, 487 (2002).\n[11] A. Peres, P. F. Scudo, and D. R. Terno, Quantum entropy and special relativity, Phys. Rev. Lett. 88, 230402 (2002).\n[12] R. Gingrich, A. Bergou, and C. Adami, Entangled light in moving frames, Phys. Rev. A 68, 042102 (2003).\n[13] H. Li and J. Du, Relativistic invariant quantum entanglement between the spins of moving bodies, Phys. Rev. A 68, 022108\n(2003).\n[14] Y. H. Moon, S. W. Hwang, and D. Ahn, Relativistic entanglements of Spin 1/2 particles with general momentum, Prog.\nTheor. Phys. 112, 219 (2004).\n[15] D. Lee and E. Chang-Young, Quantum entanglement under Lorentz boost, New J. Phys. 6, 67 (2004).\n[16] T. F. Jordan, A. Shaji, and E. C. G. Sudarshan, Lorentz transformations that entangle spins and entangle momenta, Phys.\nRev. A 75, 022101 (2007).\n[17] J. Dunningham, V. Palge, and V. Vedral, Entanglement and nonlocality of a single relativistic particle, Phys. Rev. A 80,\n044302 (2009).\n[18] N. Friis, R. A. Bertlmann, M. Huber, and B. C. Hiesmayr, Relativistic entanglement of two massive particles, Phys. Rev.\nA 81, 042114 (2010).\n[19] B. N. Esfahani and M. Aghaee, Relativistic entanglement for spins and momenta of a massive three-particle system, Int.\nJ. Quantum Inf. 09, 1255 (2011).\n[20] V. A. S. V. Bittencourt, A. E. Bernardini, and M. Blasone, Effects of Lorentz boosts on Dirac bispinor entanglement, J.\nPhys.: Conf. Ser. 1071, 012001 (2018).\n[21] V. Palge and J. Dunningham, Entanglement of two relativistic particles with discrete momenta, Ann. Phys. 363, 275\n(2015).\n[22] V. Palge and J. Dunningham, Generation of maximally entangled states with sub-luminal Lorentz boost, Phys. Rev. A 85,\n042322 (2012).\n[23] N. Bohr, The quantum postulate and the recent development of atomic theory, Nature 121, 580 (1928).\n[24] According with J. -M. Lévy-Leblond, the term \"quanton\" was given by M. Bunge. The usefulness of this term is that one\ncan refer to a generic quantum system without using words like particle or wave: J.-M. Lévy-Leblond, On the Nature of\nQuantons, Science and Education 12, 495 (2003).\n[25] W. K. Wootters and W. H. Zurek, Complementarity in the double-slit experiment: Quantum nonseparability and a\nquantitative statement of Bohr’s principle, Phys. Rev. D 19, 473 (1979).\n[26] B.-G. Englert, Fringe visibility and which-way information: An inequality, Phys. Rev. Lett. 77, 2154 (1996).13\n[27] D. M. Greenberger and A. Yasin, Simultaneous wave and particle knowledge in a neutron interferometer, Phys. Lett. A\n128, 391 (1988).\n[28] G. Jaeger, A. Shimony, and L. Vaidman, Two interferometric complementarities, Phys. Rev. A 51, 54 (1995).\n[29] S. Dürr, Quantitative wave-particle duality in multibeam interferometers, Phys. Rev. A 64, 042113 (2001).\n[30] B.-G. Englert, D. Kaszlikowski, L. C. Kwek, and W. H. Chee, Wave-particle duality in multi-path interferometers: General\nconcepts and three-path interferometers, Int. J. Quantum Inf. 6, 129 (2008).\n[31] T. Baumgratz, M. Cramer, and M. B. Plenio, Quantifying coherence, Phys. Rev. Lett. 113, 140401 (2014).\n[32] M. N. Bera, T. Qureshi, M. A. Siddiqui, and A. K. Pati, Duality of quantum coherence and path distinguishability, Phys.\nRev. A 92, 012118 (2015).\n[33] E. Bagan, J. A. Bergou, S. S. Cottrell, and M. Hillery, Relations between coherence and path information, Phys. Rev.\nLett. 116, 160406 (2016).\n[34] T. Qureshi, Coherence, interference and visibility, Quanta 8, 24 (2019).\n[35] S. Mishra, A. Venugopalan, and T. Qureshi, Decoherence and visibility enhancement in multi-path interference, Phys. Rev.\nA 100, 042122 (2019)\n[36] R. M. Angelo and A. D. Ribeiro, Wave-particle duality: An information-based approach, Found. Phys. 45, 1407 (2015).\n[37] P. J. Coles, Entropic framework for wave-particle duality in multipath interferometers, Phys. Rev. A 93, 062111 (2016).\n[38] E. Bagan, J. Calsamiglia, J. A. Bergou, and M. Hillery, Duality games and operational duality relations, Phys. Rev. Lett.\n120, 050402 (2018).\n[39] P. Roy and T. Qureshi, Path predictability and quantum coherence in multi-slit interference, Phys. Scr. 94, 095004 (2019),\n[40] M. L. W. Basso, D. S. S. Chrysosthemos, and J. Maziero, Quantitative wave-particle duality relations from the density\nmatrix properties, Quantum Inf. Process. 19, 254 (2020).\n[41] X.-F. Qian et al, Turning off quantum duality, Phys. Rev. Research 2, 012016 (2020).\n[42] M. Jakob and J. A. Bergou, Quantitative complementarity relations in bipartite systems: Entanglement as a physical\nreality, Opt. Comm. 283, 827 (2010).\n[43] M. L. W. Basso and J. Maziero, Complete complementarity relations for multipartite pure states. J. Phys. A: Math. Theor.\n53 465301 (2020).\n[44] S. Weinberg, The Quantum Theory of Fields I (Cambridge University Press, Cambridge, 1995).\n[45] M. Lanzagorta, Quantum Information in Gravitational Fields (Morgan & Claypool Publishers, California, 2014).\n[46] Y. Ohnuki, Unitary Representations of the Poincaré group and Relativistic Wave Equations (World Scientific, Singapore,\n1988).\n[47] R. U. Sexl and H. K. Urbantke, Relativity, Groups, Particles: Special Relativity and Relativistic Symmetry in Field and\nparticle Physics (Springer, New York, 2001).\n[48] W.-K. Tung, Group Theory in Physics (World Scientific, Philadelphia, 1985).\n[49] D. Ahn, H. Lee, Y. H. Moon, and S. W. Hwang, Relativistic entanglement and Bell’s inequality, Phys. Rev. A 67, 012103\n(2003).\n[50] F. R. Halpern, Special Relativity and Quantum Mechanics (Prentice-Hall, New Jersey, 1968).\n[51] J. A. Rhodes and M. D. Semon, Relativistic velocity space, Wigner rotation and Thomas precession, Am. J. Phys. 72,\n943-960 (2004).\n[52] J. A. Bergou and M. Hillery, Introduction to the Theory of Quantum Information Processing (Springer, New York, 2013).\n[53] M. Jakob and J. A. Bergou, Complementarity and entanglement in bipartite qudit systems, Phys. Rev. A 76, 052107\n(2007).\n[54] M. Jakob and J. A. Bergou, Generalized complementarity relations in composite quantum systems of arbitrary dimensions,\nInt. J. Mod. Phys. B 20, 1371 (2006).\n[55] V. S Bhaskara and P. K Panigrahi, Generalized concurrence measure for faithful quantification of multiparticle pure state\nentanglement using Lagrange’s identity and wedge product, Quantum Inf. Process. 16, 118 (2017).\n[56] P. Caban and J. Rembielinski, Lorentz-covariant reduced spin density matrix and EPR-Bohm correlations, Phys. Rev. A\n72, 012103 (2005).\n[57] T. Debarba and R. O. Vianna, Quantum state of a free spin-1/2 particle and the inextricable dependence of spin and\nmomentum under Lorentz transformations, Int. J. Quantum Info. 10, 1230003 (2012).\n[58] W. K. Wootters, Entanglement of formation of an arbitrary state of two qubits. Phys. Rev. Lett. 80, 2245 (1998)." }, { "title": "1906.02097v8.Extended_time_dependent_Ginzburg_Landau_theory.pdf", "content": "Extended time-dependent Ginzburg-Landau theory\nKonstantin V. Grigorishin∗\nBogolyubov Institute for Theoretical Physics of the National Academy\nof Sciences of Ukraine, 14-b Metrolohichna str. Kiev, 03143, Ukraine.\nWe formulate the gauge invariant Lorentz covariant Ginzburg-Landau theory which describes\nnonstationary regimes: relaxation of a superconducting system accompanied by eigen oscillations of\ninternal degrees of freedom (Higgs mode and Goldstone mode), and also forced oscillations under\nthe action of an external gauge \feld. The theory describes Lorentz covariant electrodynamics\nof superconductors where Anderson-Higgs mechanism occurs, at the same time the dynamics of\nconduction electrons remains non-relativistic. It is demonstrated that Goldstone oscillations cannot\nbe accompanied by oscillations of charge density and they generate the transverse \feld only. In\naddition, we consider Goldstone modes and features of Anderson-Higgs mechanism in two-band\nsuperconductors. We study dissipative processes, which are caused by movement of the normal\ncomponent of electron liquid and violate the Lorentz covariance, on the examples of the damped\noscillations of the order parameter and the skin-e\u000bect for electromagnetic waves. An experimental\nconsequence of the extended time-dependent Ginzburg-Landau theory regarding the penetration of\nthe electromagnetic \feld into a superconductor is proposed.\nPACS numbers: 74.20.De, 74.25.-q, 74.25.N-, 74.25.Nf\nKeywords: Ginzburg-Landau theory, Lorentz covariance, gauge invariance, Anderson-Higgs mechanism, Lon-\ndon penetration depth, wave skin-e\u000bect\nI. INTRODUCTION\nThe system undergoing the second-order phase transition (on example of superconductivity) is described with a\nLandau functional\nF=Z\nd3rF=Z\nd3r\u0014~2\n4m(r\t)\u0000\nr\t+\u0001\n+aj\tj2+b\n2j\tj4\u0015\n; (1)\nwhereFis density of free energy, a=\u000b(T\u0000Tc), \t =j\t(r)jei\u0012(r)is a two-component order parameter (the wave\nfunction of condensate of Cooper pairs) so that ns= 2j\tj2is density of superconducting (SC) electrons, a term\n~2\n4mjr\tj2can be understood as density of kinetic energy of Cooper pairs of mass 2 meach. Con\fguration of the \feld\n\t(r) which minimizes the functional (1) is obtained from equation:\n\u000eF\n\u000e\t+=@F\n@\t+\u0000r@F\n@(r\t+)= 0)~2\n4m\u0001\t\u0000a\t\u0000bj\tj2\t = 0: (2)\nThis con\fguration is shown in Fig.1a: they say that the \feld is a string laying in a valley of the potential aj\tj2+b\n2j\tj4.\nAt the same time, out the equilibrium the \feld \t depends on time. For small deviations from the equilibrium\nit is natural to assume [1, 2], that the time derivative @\t=@tis proportional to the variational derivative of the\nfree energy functional \u000eF=\u000e \t+which is equal to zero at the equilibrium. Thus, one can write the time-dependent\nGinzburg-Landau equation [1{10] (TDGL equation) in a form:\n~2\n4mD@\t\n@t=\u0000\u000eF\n\u000e\t+)\u001c@ \n@t=\u00182\u0001 + \u0000j j2 ; (3)\nwhere = \t=\t0is the dimensionless order parameter, \t 0=p\n\u0000a=bis an equilibrium value of the spatial homo-\ngeneous order parameter, Dis a di\u000busion coe\u000ecient of electrons in the normal state. The temperature-dependent\ncoherence length \u0018(T) =q\n~2\n4mja(T)jand the temperature-dependent relaxation time \u001c(T) =~2\n4mDja(T)jcan be found\nfrom the microscopic theory for the case of a gapless SC alloy containing a high concentration of paramagnetic im-\npurities [3, 4] and for dirty superconductors in the Ginzburg-Landau (GL) regime jTc\u0000Tj<Tcthe TDGL equation has a form\n\u001c0@\t\n@t=\u00182\u0001\t\u0000\t, where\u001c0=\u0019~\n8(T\u0000Tc)is the relaxation time for a homogeneous mode (here and further kB= 1) [6].\nIn this case the equilibrium value is h\ti= 0 but\n\t2\u000b\n6= 0. The corresponding relaxation processes are illustrated in\nFig.1b.\nThus, Eq.(3) describes relaxation of the order parameter at small deviations from the equilibrium. At the same\ntime, in the system undergoing the second-order phase transition the collective excitations can exist which are resonant\noscillations. When a continuous symmetry is spontaneously broken, there emerge two types of collective modes in\ngeneral: massive Higgs mode, which is oscillation of modulus j\tjof the order parameter, and Goldstone mode, which\nis oscillation of the phase \u0012. As it has been demonstrated in [11{13] in collisionless approximation the Higgs mode in\nSC system can exist as perturbation of amplitude of the gap \u0001 which takes the form of oscillations having a frequency\n\u0018j\u0001j. At the same time, the oscillations of order parameter damp in some time \u001c. Thus, the system is out from\nthe equilibrium then the relaxation process depends on relation between period of the eigen oscillations \u00181=!and\nthe damping time: if 1 =!\u001d\u001cthen aperiodic relaxation occurs as shown in Fig.1b and it is described with TDGL\nequation (3). In general case the parameters 1 =!and\u001ccan be in arbitrary relation, hence the relaxation process can\nhave more complicated form, for example, if 1 =!\u001c\u001cthen an oscillatory process with small damping occurs as shown\nin Fig.1c. Moreover, an external \feld can swing the system, that is the undamped oscillations occur, while heat is\nreleased (for example, electromagnetic wave, falling on a superconductor, induces Foucault currents both normal jn\nand superconducting js). Such situation is shown in Fig.1d.\nIn presence of electric 'and magnetic Apotentials the replacements\n@\n@t!@\n@t+i2e\n~';r!r\u0000i2e\nc~A (4)\nmust be done in free energy functional (1) and in Eqs.(2,3) for a gauge invariance. Moreover, the total current is a sum\nof normal current and supercurrent: j=\u001b\u0000\n\u0000r'\u00001\nc@A\n@t\u0001\n+js, where\u001bis conductivity, js=\u0000ie~\n2m(\t+r\t\u0000\tr\t+)\u00003\n2e2\nmcj\tj2A. Thus, the TDGL equations determine dynamics of both order parameter and electromagnetic \feld\nA\u0016\u0011(';A). On the other hand, equations for electromagnetic \feld should be Lorentz covariant both in vacuum and\nwithin any medium, despite dynamics of particles (medium) is non-relativistic, for example, Maxwell equations in\ndielectrics. However the light speed in the medium is less than the speed in vacuum cand is determined with dynamic\nproperties of the system. At the same time, the dissipative mechanisms give terms which violate Lorentz covariance\nsince the dissipation distinguishes a time direction, i.e., violates the time symmetry t$\u0000twhich is symmetry of the\nLorentz boost. For example, Maxwell equation in medium with conductivity \u001bhas a form:\ncurlH=1\nc@D\n@t+4\u0019\nc\u001bE: (5)\nHere, the \frst two terms are an equation from Lorentz covariant Maxwell equations, the last term4\u0019\nc\u001bEis a dissipative\npart. Depending on material and processes occurring in it, some terms in the equations can be neglected. So, in\nmetals the dissipative term dominates, for example, the strength of electrostatic \feld must be E= 0 inside metal:\nthe nonzero strength would lead to a current, meanwhile the propagation of the current is associated with energy\ndissipation according to Joule-Lenz law Q=j2=\u001b=\u001bE2and, therefore, it cannot be supported in an equilibrium\nstate by itself [14]. For not very large frequencies of electromagnetic waves (less than plasma frequency) a condition\n\u001b\n!\u001d\"(!) is satis\fed for good metals, then the conduction current \u0000\u001b1\nc@A\n@tgives main contribution in electromagnetic\nresponse that gives the skin-e\u000bect, at the same time the displacement current\"\n4\u0019@E\n@tcan be neglected [14, 15]. On the\ncontrary, in dielectrics the dissipative processes can be neglected in the \frst approximation (since conductivity is zero);\nhence, Lorentz covariant electrodynamics remains only. In bad conductors (semiconductors, weak electrolytes, weakly\nionized plasma) both conduction and displacement currents can be important. Analogously to dielectrics and metals\nelectrodynamics of superconductors must be composed of both Lorentz covariant part and dissipative part (due to\nfriction of normal electrons, damping of collective excitations and breaking of Cooper pairs). Which term is dominant\nis determined with physical conditions and processes. Unlike the normal metal phase, in the SC phase the conductivity\nessentially depends on temperature due to density of normal electrons nn:\u001b(T) =\u001cph\nme2nn(T))\u001b(T!0)!0\n(here\u001cphis the mean free time of electron caused by electron-phonon interaction), that is at low temperatures\nthe contribution of the dissipative part decreases. It should be noted that superconductivity is a thermodynamic\ne\u000bect and is not electrodynamic one (superconductor is not ideal conductor). Hence equations for electromagnetic\n\feld in superconductor are result of variation of some functional of action (or free energy functional for equilibrium\ncase):\u000eS\n\u000eA\u0016= 0. The action S[\t(r;t);A\u0016(r;t)] must be Lorentz invariant in order to ensure Lorentz covariant\nelectrodynamics of super\ruid component. The dissipative terms are introduced in the equations of motion by means\nof Rayleigh dissipation function. At the same time, the Lorentz invariance of the action should have consequences for\ndynamics of the scalar \feld \t: the collective pseudo-relativistic excitations (Higgs mode and Goldstone mode) occur.\nIn [2, 5] the general equations for the dynamic behavior of dirty superconductors in GL regime jTc\u0000Tj<< Tcare\nderived from microscopic theory. The local equilibrium approximation leads to a simple generalized TDGL equation\ndescribing relaxation of the order parameter. As indicated above, in this nonequilibrium regime the dissipative\nprocesses dominant, hence the Lorentz covariation can be neglected and a relaxation equation of type TDGL (3) is\nvalid in mean \feld approximation.\nThe dynamic extension of GL theory has been proposed in [16] as a time-dependent nonlinear Schr odinger La-\ngrangian:\nL=i~\t+@\t\n@t\u0000~2\n4mr\t+r\t\u0000V(j\tj); (6)\nwhich describes the low-frequency, long-wavelength dynamics of the pair \feld \t( r;t) for a BCS-type s-wave super-\nconductor at T= 0,Vis potential leading to spontaneously broken U(1) symmetry and is assumed to be a function\nofj\tjonly. We can see that Lagrangian (6) is Galilean invariant. In this Lagrangian the electromagnetic \feld\nA\u0016= (';A) must be included by replacement (4) based on gauge invariance, and Lagrangian of electromagnetic\n\feld\u0000\nE2\u0000H2\u0001\n=8\u0019should be added. Varying the corresponding action:\u000eS(\t;A\u0016)\n\u000eA\u0016= 0 we will \fnd equations for\nthe electromagnetic \feld A\u0016(r;t) in SC medium. Obviously, these equations will not be Lorentz covariant, since\ninitial Lagrangian (6) is not Lorentz invariant. However, as mentioned above, equations for electromagnetic \feld in\nnondissipative medium must be Lorentz covariant like the Maxwell equations in vacuum.\nIt is believed that superconductors cannot contain macroscopic electric \felds in static con\fgurations. This fact is\ndirectly based on the \frst of the London equations:\ndjs\ndt=nse2\nmE (7)\ncurljs=\u0000nse2\nmcH)\u0001H=1\n\u00152H; (8)4\nwhere\u0015=q\nmc2\n4\u0019nse2is the London penetration depth, nsis density of SC electrons, js=nsevis supercurrent density,\nandE,Hare electric and magnetic \felds respectively. Indeed, unlike normal metals, in superconductors due to zero\nresistance to support the direct current jthe presence of the electric \feld Eis not necessary. This means that in the\nstationary regimedj\ndt= 0 we have E= 0 inside superconductor. However, it should be noted, that the second London\nequation (8) is a result of minimization of free energy of the superconductor:1\n8\u0019R\u0002\nH2+\u00152(curlH)2\u0003\ndV, where the\n\frst term is energy of the magnetic \feld, the second term is kinetic energy of the supercurrent. At the same time, the\n\frst London equation is not result of minimization of the free energy of superconductor: it is suggested that motion\nof SC electrons is not accompanied with friction hence they are accelerated by electric \feld E, i.e., it is just the\nsecond Newton law. This means that the \frst London equation (7)is equation of an ideal conductor that discussed\nin Appendix A. Superconductivity is the thermodynamically steady state: con\fguration of both the electric \feld E\nand the magnetic \feld Hmust be found from minimum of some free energy functional F(\t;\t+;';A).\nFrom Eqs.(7,8) we can see that the coupling of the supercurrent density to the electromagnetic \felds through the\nLondon and GL equations is not space-time covariant : under Lorentz boost, the supercurrent density jin the presence\nof the electric \feld Eought to transform as the space components of a 4-vector whose time component would then\nplay role of some supercharge density, while the electric Eand magnetic H\felds transform as components of the two\nindex antisymmetric \feld strength tensor F\u0016\u0017. Thus, it would seem, the Lorentz covariance requires the possibility of\nelectric \felds on the same footing as magnetic ones within superconductors, with an electric penetration depth equal\nto the familiar magnetic one. In works [17, 18] it has been proposed the natural covariant extension of the GL free\nenergy (1) in terms of the Higgs Lagrangian for spontaneous U(1) gauge symmetry breaking in the vacuum (in SI\nunits):\nL=1\n2\"0c2\u0012~\n2e\u0015\u00132(\f\f\f\f\u0012\n@\u0016+i2e\n~A\u0016\u0013\n \f\f\f\f2\n\u00001\n2\u00182(j j2\u00001)2)\n\u00001\n4\"0c2F\u0016\u0017F\u0016\u0017; (9)\nwhere the order parameter is normalized to the density of electron pairs in a bulk sample (x) = \t(x)=\t0\n(\t0=p\n\u0000a=b) in the absence of any electromagnetic \feld A\u0016= ('=c;\u0000A), the tensor F\u0016\u0017=@\u0016A\u0017\u0000@\u0017A\u0016is the\nelectromagnetic \feld strength with @\u0016=@=@x\u0016being the space-time gradients for the coordinates x\u0016= (ct;r),\u0015is\nthe magnetic penetration depth. The dynamics is determined from the Lorentz invariant action S=R\nLd4xthrough\nthe variational principle provides the equations of motion for the scalar \feld and the vector \feld A\u0016respectively.\nMain result of this model is the e\u000bect of electrostatic \feld on a superconductor: the \feld Eas well as the \feld Bcan\ndestroy the SC state. Namely, the critical values of the \felds satisfy the relation:\n\u0012B\nBcm\u00132\n+\u0012E=c\nBcm\u0013\n'1; (10)\nwhereBcmis a thermodynamical magnetic \feld (at B > Bcma type-I superconductor goes into the normal state\nby the \frst-order phase transition). Thus, we can see that at electric \feld E'cBcmsuperconductivity should be\ndestroyed at absence of magnetic \feld. The analysis of experimental data presented in [19]suggests that an external\nelectric \feld does not a\u000bect signi\fcantly the superconducting state . This conclusion is in total contradiction with the\nexpected behavior based on the covariant Lagrangian (9). It has been suggested that such negative experimental\nresult can be explained with that some non-paired normal electrons could play a crucial role against the electric \feld,\nwhereas such contributions are not included in the model.\nIn the work [20] relativistically covariant London equations have been proposed, and they can be understood as\narising from the \"rigidity\" of the super\ruid wave function in a relativistically covariant microscopic theory. They\npredict that for slowly varying electric \felds, both longitudinal and transverse, and assuming no charge density in the\nsuperconductor:\nr2E=1\n\u00152E: (11)\nwhich implies that an electric \feld penetrates a distance \u0015, as a magnetic \feld does. Thus, the screening length of\nthe electric \feld in the SC state is equal to the London penetration depth, not the Thomas-Fermi screening length\n\u0015TF. The associated longitudinal dielectric function is obtained as\n\"(q;!!0) = 1 +1\n\u00152q2(12)\nunlike the dielectric function for normal metal \"(q;!!0) = 1 +1\n\u00152\nTFq2. This model remains debatable [21, 22] and\nrequires experimental veri\fcation [23, 24].5\nIn the Lorentz covariant models the Higgs mode (spectrum with energy gap) and the Goldstone mode (acoustic\nspectrum) should arise. However, unlike the Higgs mode, the Goldstone mode is unobservable that can be explained\nwith the Anderson-Higgs mechanism: oscillations of the phase \u0012are absorbed into the gauge \feld A\u0016. At the same\ntime, the above-mentioned models essentially di\u000ber from the results of the theory of gauge-invariant response of\nsuperconductors to external electromagnetic \feld presented in [25]. In this model the Coulomb interaction \"pushes\"\nthe frequency of the acoustic oscillations to the plasma frequency !p. Thus, Goldstone mode becomes unobservable\nin itself since it turns to plasma oscillations. In addition, in such model the electric \feld is screened with Thomas-\nFermi length \u0015TFlike in the normal metal phase. However, it should be noted, that the phase oscillations \u0012(r;t)\nare oscillations of the order parameter j\tjei\u0012, i.e., they are speci\fc for the SC state. At the same time, the plasma\noscillations exist unchanged both in SC phase and in the normal metal phase, i.e., they are unrelated to the SC\nordering and are not speci\fc for the SC state. Thus, if the plasma oscillations were the phase oscillations of the order\nparameter, then this would be a\u000bected them at the transition point Tcnecessarily; however, the spectrum of the\nplasma oscillations does not depend on temperature.\nProceeding from aforesaid, we are aimed to generalize the GL theory for the nonstationary regimes Fig.1(b-d):\nthe theory should describe the relaxation of the system with accounting of eigen oscillations of internal degrees of\nfreedom, and also forced oscillations under the action of external \feld. Thus, this theory includes the GL theory and\nthe TDGL equation as special cases, and we will call it as the extended TDGL theory . Moreover, this theory must be\nLorentz covariant without accounting of dissipative processes, since it includes Lorentz covariant electrodynamics, at\nthe same time the dynamics of conduction electrons remains non-relativistic. Accounting of dissipative mechanisms\nshould violate the Lorentz covariance of the theory and should lead to relaxation processes in SC system. Thus, our\npaper is organized by the following way. In Sect.II we generalize GL free energy functional to relativistic-like action\nby phenomenological approach. Using this action we study the possible eigen oscillations of the order parameter\n\t(t;r): Higgs mode and Goldstone mode. Moreover, we obtain this relativistic-like GL functional by microscopical\napproach also. In Sect.III using the gauge symmetry we formulate electrodynamics of superconductors in the sense\nof Lorentz covariant equations for 4D electromagnetic potential A\u0016\u0011(';A). The equations describe propagate of\nelectromagnetic \feld in SC medium where Anderson-Higgs mechanism (absorption of the Goldstone bosons into the\ngauge \feld A\u0016and the gaining of mass by the gauge \feld) takes place. In Sect.IV we consider features of Anderson-\nHiggs mechanism in two-band superconductors and occurrence of Leggett's mode. In Sect.V we study in\ruence of the\nmotion of normal component of electron liquid which causes the damping of oscillations both order parameter and\nelectromagnetic \feld. As example we consider the wave skin-e\u000bect, eigen electromagnetic oscillations and relaxation\nof \ructuation of the order parameter. We propose an experimental consequence of the extended TDGL theory\nregarding the penetration of the electromagnetic \feld in a superconductor. Besides we demonstrate that the London\nelectrodynamics and the TDGL equation (3) are limit cases of the extended TDGL theory.\nII. NORMAL MODES AND PSEUDO-RELATIVISTIC COLLECTIVE EXCITATIONS\nA. Phenomenological approach\nIn general case the SC order parameter \t is both spatially inhomogeneous and it can change over time: \t = \t( r;t).\nThe order parameter is a complex scalar \feld which is equivalent to two real \felds: modulus j\t(r;t)jand phase\u0012(r;t)\n(the modulus-phase representation):\n\t(r;t) =j\t(r;t)jei\u0012(r;t): (13)\nFor stationary case \t = \t( r) the free energy functional (1) exists and the steady con\fguration of the \feld \t( r)\nminimizes this functional. However for the nonstationary case \t( r;t) the minimizing procedure loses any sense.\nThus, it is necessary to \fnd an equation determining evolution of the order parameter in time. Our method for\nsolving this problem is as follows. The parameter t- the time can be turned into a coordinate t!\u001dtin some\n4D Minkowski space f\u001dt;rg, where\u001dis an parameter of dimension of speed (like the light speed) which must be\ndetermined with dynamical properties of the system . At the same time, the dynamics of conduction electrons remains\nnon-relativistic. Then the two-component scalar \feld \t( r;t) minimizes some action S(like in the relativistic \feld\ntheory [26]) in the Minkowski space:\nS=1\n\u001dZ\nL(\t;\t+)d\n; (14)\nwhereLis some Lagrangian (density of Lagrange function L=R\nLd3r),d\n\u0011\u001ddtd3ris an element of the 4D\nMinkowski space. The Lagrangian can be built by generalizing the density of free energy Fin Eq.(1) to \"relativistic\"6\ninvariant form by substitution of covariant and contravariant di\u000berential operators\ne@\u0016\u0011\u00121\n\u001d@\n@t;r\u0013\n;e@\u0016\u0011\u00121\n\u001d@\n@t;\u0000r\u0013\n(15)\ninstead the gradient operators: r\t!e@\u0016\t;r\t+!e@\u0016\t+. Then the required Lagrangian takes a form:\nL=~2\n4m\u0010\ne@\u0016\t\u0011\u0010\ne@\u0016\t+\u0011\n\u0000aj\tj2\u0000b\n2j\tj4=~2\n4m1\n\u001d2\u0012@\t\n@t\u0013\u0012@\t+\n@t\u0013\n\u0000~2\n4m(r\t)\u0000\nr\t+\u0001\n\u0000aj\tj2\u0000b\n2j\tj4\u0011T\u0000F:(16)\nHere, as we can see from Eq.(1), the spatial terms of the Lagrangian (free energy F) play role of \"potential\" energy,\nand the time term plays role of \"kinetic\" energy T. Lagrange equation for functional (14) is\ne@\u0016@L\n@(e@\u0016\t+)\u0000@L\n@\t+= 0)~2\n4me@\u0016e@\u0016\t +a\t +bj\tj2\t = 0; (17)\nwhere\ne@\u0016e@\u0016=e@\u0016e@\u0016=1\n\u001d2@2\n@t2\u0000\u0001: (18)\nSince the \feld \t is complex and Lagrangian (16) has U(1) symmetry then the conserving charge takes place. Using\nEq.(17) and its complex conjugate equation we can obtain the continuity condition in a form:\ne@\u0016j\u0016=@\u001a\n@t+ divjs= 0; (19)\nwherej\u0016is a 4D supercurrent:\nj\u0016\u0011(\u001d\u001as;\u0000js) =ie~\n2m\u0010\n\t+e@\u0016\t\u0000\te@\u0016\t+\u0011\n=\u0012\n\u001die~\n2m\u001d2\u0012\n\t+@\t\n@t\u0000\t@\t+\n@t\u0013\n;ie~\n2m\u0000\n\t+r\t\u0000\tr\t+\u0001\u0013\n=\u0012\n\u0000\u001de~\nm\u001d2j\tj2@\u0012\n@t;\u0000e~\nmj\tj2r\u0012\u0013\n: (20)\nIt should be noted that the correct de\fnition of the current j\u0016is possible only with using of gauge invariance, which\nis done in Sect.III.\nSubstituting representation (13) in the Lagrangian (16) we obtain:\nL=~2\n4me@\u0016j\tje@\u0016j\tj+~2\n4mj\tj2e@\u0016\u0012e@\u0016\u0012\u0000aj\tj2\u0000b\n2j\tj4: (21)\nLet us consider small variations of modulus of order parameter from its equilibrium value: j\tj=p\u0000a\nb+\u001e\u0011\t0+\u001e,\nwherej\u001ej\u001c\t0. Then atT 0, that is the Landau criterion for super\ruidity is satis\fed. Since the Higgs mode is oscillations of modulus of the\norder parameterj\tjwhich determines the density of SC electrons as ns= 2j\tj2atT!Tc, then these oscillations\nare accompanied by changes of SC density. At the same time, the total electron density must be n=ns+nn= const,\nbecause otherwise the oscillations of charge (plasmons) should take place (the plasmons exist in normal metal too,\nhence these oscillations are not speci\fc for SC state). Thus, the oscillation of nswhenn= const and q6= 0 can be\npresented as counter\rows of SC and normal components so that nsvs+nnvn= 0 - Fig.3. Hence the Higgs mode\ncan be considered as sound in the gas of above-condensate quasiparticles (with spectrump\nj\u0001j2+v2\nF(p\u0000pF)2) -\nthe second sound. It should be noted that since the normal component nn=n\u00002j\tj2is gas of excitations above\ncondensate of the Cooper pairs (moreover, the size of a pair is much more than average distance between electrons),\nthen separation of electrons into SC and normal is some conditionality: in reality each electron makes SC and normal\nmovements simultaneously.\nThe speed\u001dcan be found with the following way. Let us consider the long-wave limit q!0 for Higgs mode, that\nis the whole system oscillates in a phase. To change the SC density and, hence, the normal density, one Cooper pair\nmust be broken as minimum. For this the energy 2 j\u0001jmust be spent as minimum (SC energy gap). In turn, from\nEq.(25) it can see that minimal energy to excite one Higgs boson isp\n8jajm\u001d2, then\np\n8jajm\u001d2= 2j\u0001j)\u001d=vFp\n3; (27)\nwherevFis Fermi velocity, we have used j\u0001j=Tc\u0010\n8\u00192\n7\u0010(3)\u00111=2\u0010\n1\u0000T\nTc\u00111=2\n,a=6\u00192T2\nc\n7\u0010(3)\"F\u0010\nT\nTc\u00001\u0011\nfor pure supercon-\nductor [27]. Thus, the speed \u001ddoes not depend on temperature (at T!Tc) and is\u0018106m=s\u001cc. Physical sense\nof this speed will be de\fned in the next section. Using Eq.(27) we can see that a value ~!= 2j\u0001jis achieved in the\nGoldstone mode (26) at q=p\n2\n\u0018, where\u0018=q\n~2\n4mjajis a temperature-dependent coherence length. Noteworthy that,\nif we setq=1\n\u0015, then ~!(q)>2j\u0001jbe for type-I superconductors, and ~!(q)<2j\u0001jbe for type-II superconductors -\nFig.2, since \u0014\u0011\u0015\n\u0018<1p\n2for the type-I and \u0014>1p\n2for the type-II.\nFigure 2: Higgs oscillations with spectrum (25) - blue line, and Goldstone oscillations with spectrum (26) - red line. For q=p\n2\n\u0018\nwe have ~!= 2j\u0001jin the Goldstone mode. If to set q=1\n\u0015, then ~!(q)>2j\u0001jfor type-I superconductors and ~!(q)<2j\u0001j\nfor type-II superconductors. The region where the pair breaking occurs, because E > 2j\u0001j, is shaded in gray. The free Higgs\nmode lies entirely in this region, hence this mode is unstable due to decay to the above-condensate quasiparticles.\nThe dispersion relation (25) can be rewritten in the following relativistic-like form:\nE2=em2\u001d4+p2\u001d2; (28)8\nFigure 3: Higgs oscillations with spectrum (25) at q6= 0. This mode is oscillations of modulus of order parameter j\tj, the\noscillations are accompanied by changes of density of SC electrons ns= 2j\tj2. At the same time, the total electron density\nmust ben=ns+nn= const, hence the oscillations can be presented as counter\rows of SC and normal components so that\nnsvs+nnvn= 0.\nwhere\nem\u0011p\n8jajm\n\u001d=p\n2~\n\u0018\u001d/(Tc\u0000T)1=2(29)\nis the mass of a Higgs boson. The mass emdetermines the scale of spatial inhomogeneities in a superconductor:\nletE= 0 (that is we consider a stationary con\fguration: \t( t) = const) then p2=\u0000em2\u001d2)p=i~p\n2\n\u0018, hence\n\t = \t 0\u0000\u001e0eip\n~x= \t 0\u0000\u001e0e\u0000p\n2\n\u0018x(for example, proximity e\u000bect, where a superconductor occupies half-space x>0),\nthat is the order parameter is recovered on length\u0018p\n2=~\nem\u001d. It should be noticed since in the normal phase (that is at\nT >TcorH >Hcwherens= 0,nn=n) equilibrium value of the order parameter is \t = 0 then the Goldstone and\nHiggs oscillations loss any sense. In the normal phase the relaxation of the nonequilibrium order parameter \t 6= 0\nto the equilibrium one \t = 0 takes place only. In the normal phase the speed \u001dlosses physical sense and the correct\nlimit transition to the normal state in expressions, which depend on the factor c=\u001d, will be formulated in Sect.III.\nAs seen from Eqs.(25,27) the energy of Higgs boson is E\u00152j\u0001j, that is this mode exists in the free quasiparticle\ncontinuum. Hence the free Higgs oscillation decays to quasiparticles with energyp\nj\u0001j2+v2\nF(p\u0000pF)2each. Thus,\nthe free Higgs mode in a pure superconductor is unstable , hence its observation is problematical. So, investigations of\nresonant excitations in cuprate superconductors using THz pulse in [28] exhibits that in this nonlinear optical process\nthe light-induced excitation of Cooper pairs is dominated, while the collective amplitude (Higgs) \ructuations of the\nSC order parameter give, in general, a negligible contribution. At the same time, Higgs mode is a scalar excitation\nof the order parameter, distinct from charge or spin \ructuations, and thus does not couple to electromagnetic \felds\nlinearly. It makes possible to study the Higgs mode through the nonlinear light{Higgs coupling which manifests itself,\nfor example, in the third harmonic generation and pump-probe spectroscopy mediated by the Higgs mode [29]. It\nshould be noted that the impurity scattering drastically enhances the light{Higgs coupling: in the pure limit the\nHiggs-mode contribution is subleading to quasiparticles, whereas in the dirty regime it becomes comparable with or\neven larger than the quasiparticle contribution. The precise ratio between the Higgs and quasiparticle contributions\nmay depend on details of the system.\nThe Goldstone mode (24) is eddy (Foucault) currents: alternating current j=e~\nmj\tj2r\u0012(since\u0012/ei!t) generates\nalternating magnetic \feld curl H(t) =4\u0019\ncj(t), hence the electric \feld curl E=1\nc@H\n@tis induced, which drives the\neddy currents in turn. In the long wave limit ( q!0) the energy of the Goldstone mode is E= 0, that means\nthe passing of nondissipative direct current. These processes will be detail considered in Sects.III,V. Substituting the\nsupercurrent (20) in the continuity condition (19) we obtain the equation for Goldstone mode (24). Thus, the equation\nfor Goldstone mode (24) is the continuity condition for the corresponding eddy currents . In the Sect.III we will reveal\nthat Goldstone oscillations are not accompanied by the charge oscillations, i.e.,@\u001as\n@t= 0, hence these supercurrents\nare closured div js= 0.\nAt the same time, in this Section we neglected by the normal component nn=n\u00002j\tj2with corresponding friction\n\u0000m\n\u001cphv, where\u001cphis the mean free path time of electrons caused by electron-phonon interaction (or by scattering on\nlattice defects). The friction causes damping of oscillations of the order parameter and can lead to the overdamped\nregime when monotonic relaxation of a \ructuation occurs. These processes are considered in Sect.V.9\nB. Microscopic derivation\nFollowing [30] let us consider electrons in a normal metal ( T >Tc) propagating in a random \"\feld\" of thermodynamic\n\ructuations of the SC order parameter \u0001 (however h\u0001i= 0) which are static and \"smooth\" enough in space. Then\nwe can write down the following Hamiltonian for interaction of an electron with these \ructuations:\nbHint=X\nkh\n\u0001a+\nk\"a+\n\u0000k#+ \u0001a\u0000k#ak\"i\n: (30)\nCorrection to thermodynamic potential (free energy) due to (30) is\nFs\u0000Fn=\u0000j\u0001j2TX\nkX\nnG(k;\"n)G(\u0000k;\u0000\"n) +j\u0001j2TcX\nkX\nnG(k;\"n)G(\u0000k;\u0000\"n)jT=Tc\n+T\n2j\u0001j4X\nkX\nnG2(k;\"n)G2(\u0000k;\u0000\"n) +:::; (31)\nwhereG=1\ni\"n\u0000\u0018is an electron propagator for normal metal, \"n=\u0019T(2n+ 1),\u0018=~\u001dF(k\u0000kF) is energy of an\nelectron near Fermi surface. The \frst term (which is /j\u0001j2T) diverges, but it is compensated by the second term\n(which is/j\u0001j2Tc), the third term is \fnite and it can be taken at T=TcnearTc. The terms from Eq.(31) can be\nrepresented by diagrams shown in Fig.4. Calculation of the coe\u000ecients at j\u0001j2,j\u0001j4gives:\nFs\u0000Fn=V\u0017FT\u0000Tc\nTcj\u0001j2+V\u0017F7\u0010(3)\n16\u00192T2cj\u0001j4+:::; (32)\nwhere\u0017F=mkF\n2\u00192~2is density of state on Fermi surface per spin, Vis volume of the system,P\nk!V\n(2\u0019)2R\nd3k. Thus,\nwe have microscopic method of derivation of GL expansion of the free energy functional which will be generalized as\nfollows.\nLet us consider the \frst diagram in Fig.4 and let us give an additional momentum qand an additional energy\nparameter!m=\u0019T(2m+ 1) to one side of the loop as illustrated in Fig.5. This means that the order parameter will\nbe function of these parameters \u0001( q;!m), and note \u0018k+q\u0019\u0018k+~2kq\nm2wherejkj=kF. Then instead the \frst term\nin the expansion (31) we have:\n\u0000j\u0001j2TX\nkX\nnG(k+q;\"n+!m)G(\u0000k;\u0000\"n) =\u0000j\u0001j2TX\nkX\nn1\ni\"n\u0000\u0018+\u0010\ni!m\u0000~2kq\nm\u00111\n\u0000i\"n\u0000\u0018\n\u0019\u0000j \u0001j2TV\u0017FZ+1\n\u00001d\u0018X\nn1\n\"2n+\u00182+j\u0001j2TcV\u0017F\n2Z+1\n\u00001d\u0018Z1\n\u00001d(cos\u0012)X\nn\"2\nn\u0000\u00182\n(\"2n+\u00182)3\u0014\n(i!m)2+~4k2\nF\nm2q2cos2\u0012\u0015\n=\u0000j\u0001j2TV\u0017FZ+1\n\u00001d\u0018X\nn1\n\"2n+\u00182+j\u0001j2V\u0017F1\n4\u00192T2c7\u0010(3)\n4(i!m)2+j\u0001j2V\u0017F~2\u001d2\nF\n12\u00192T2c7\u0010(3)\n4q2(33)\nup to the second-order terms in the additional parameters qand!m.\nFigure 4: Diagrammatic representation of Ginzburg-Landau expansion from [30].10\nFigure 5: The \frst diagram from Fig.4 which accounts spatial and time inhomogeneities of the order parameter.\nAt \frst let us consider the expansion in the parameter qonly:\nFq=V\u0017F~2\u001d2\nF\n12\u00192T2c7\u0010(3)\n4q2j\u0001j2+V\u0017FT\u0000Tc\nTcj\u0001j2+V\u0017F7\u0010(3)\n16\u00192T2cj\u0001j4: (34)\nLet us introduce \"wave function\" of Cooper pairs which determines density of SC component ns= 2j\tj2[30, 31] so\nthat to obtain the term~2\n4mq2j\tj2(\"kinetic energy\" of Cooper pairs) instead the \frst term in Eq.(34):\n\t =(14\u0010(3)n)1=2\n4\u0019Tc\u0001; (35)\nwheren=k3\nF\n3\u00192is the total electron density. Then\nFq=V~2\n4mq2j\tj2+Vaj\tj2+Vb\n2j\tj4; (36)\nwhere\na=\u0017F(4\u0019Tc)2\n14\u0010(3)nT\u0000Tc\nTc; b =\u0017F(4\u0019Tc)2\n14\u0010(3)n2; (37)\nso thatj\tq=0j2=\u0000a\nb=nTc\u0000T\nTc)ns\nn= 2\u0010\n1\u0000T\nTc\u0011\n. The function Fqshould be understood as free energy per a\nwave-vector q. Corresponding free energy functional is\nF=X\nqFq=VX\nq\u0014~2\n4mq2j\tqj2+aj\tqj2+b\n2j\tqj4\u0015\n\u0019Z\u0014~2\n4mjr\t(r)j2+aj\t(r)j2+b\n2j\t(r)j4\u0015\nd3r; (38)\nwhere\n\t(q) =1\nVZ\n\t(r)e\u0000iqrd3r;\t(r) =X\nq\t(q)eiqr=VZ\n\t(q)eiqrd3q\n(2\u0019)3: (39)\nThus, we have standard GL expansion.\nNow let us consider the expansion with a term with ( i!m)2in Eq.(33). Then we have an expression:\neFq;!m=V1\n4m(i!m)2\n(\u001dF=p\n3)2j\tq;!mj2+V~2\n4mq2j\tq;!mj2+Vaj\tq;!mj2+Vb\n2j\tq;!mj4: (40)\nTo perform analytic continuation of this expression to the real axis we have to make a substitution i!m!~!+i\u000e\nwhere\u000e!0 [30]. Then we obtain:\neFq;!=V~2\n4m\u001d2!2j\tq;!j2+V~2\n4mq2j\tq;!j2+Vaj\tq;!j2+Vb\n2j\tq;!j4\u0011Tq+Fq: (41)11\nHere, we have noted:\n\u001d\u0011\u001dFp\n3; (42)\nthat coincides with a \"light\" speed (27). Presence of the frequency !means dependence of the order parameter on\ntime since\n\t(t) =TZ\n\t(!)e\u0000i!td!\n2\u0019;\t(!) =1\nTZ\n\t(t)ei!tdt; (43)\nwhere T\u00183p\nV=\u001d is some time interval introduced for the same dimensions of \t( !) and \t(t). The function eFq;!does\nnot have sense of free energy. From Eq.(41) we can see that the \frst term plays role of \"kinetic\" energy Tqand the\nremaining terms Fqplays role of \"potential\" energy. Therefore corresponding Lagrangian has a form:\nLq;!=Tq\u0000Fq=V~2\n4m\u001d2!2j\tq;!j2\u0000V~2\n4mq2j\tq;!j2\u0000Vaj\tq;!j2\u0000Vb\n2j\tq;!j4: (44)\nThen an action takes place (like a free energy functional (38)):\nS=T2X\nqZd!\n2\u0019Lq;!=Z\nL[\t(r;t)]d3rdt=1\n\u001dZ\nL[\t(r;t)]d\n; (45)\nwhere the Lagrangian is\nL[\t(r;t)]\u0019~2\n4m1\n\u001d2\u0012@\t\n@t\u0013\u0012@\t+\n@t\u0013\n\u0000~2\n4m(r\t)\u0000\nr\t+\u0001\n\u0000aj\tj2\u0000b\n2j\tj4\n\u0011~2\n4m\u0010\ne@\u0016\t\u0011\u0010\ne@\u0016\t+\u0011\n\u0000aj\tj2\u0000b\n2j\tj4; (46)\nwhich coincides with phenomenological Lagrangian (16) in the extended TDGL theory. Here, e@\u0016ande@\u0016are covariant\nand contravariant di\u000berential operators (15) accordingly. We can see that Lagrangian (46) is Lorentz invariant in\nsome 4D Minkowski space f\u001dt;rg, where\u001dis an parameter of dimension of speed (like the light speed) which is\ndetermined by dynamical properties of the system. At the same time, the dynamics of conduction electrons remains\nnon-relativistic.\nIII. ELECTRODYNAMICS\nA. Gauge invariance\nLet a superconductor be in electromagnetic \feld A\u0016= (';\u0000A) (or contravariant vector A\u0016= (';A)). In order to\nensure the gauge invariance the di\u000berential operation @\u0016\t must be changed as follows:\n@\u0016\t!\u0012\n@\u0016+i2e\nc~A\u0016\u0013\n\t =ei\u0012\u0014\n@\u0016j\tj+ij\tj\u0012\n@\u0016\u0012+2e\nc~A\u0016\u0013\u0015\n: (47)\nIndeed, making a gauge transformation\n\u0012=\u00120+2e\nc~\u001f; A\u0016=A0\n\u0016\u0000@\u0016\u001f=\u001a\n'='0\u00001\nc@\u001f\n@t\nA=A0+r\u001f\u001b\n; (48)\nwe have:\n@\u0016\u0012+2e\nc~A\u0016=@\u0016\u00120+2e\nc~A0\n\u0016: (49)\nHowever in Lagrangian (16) the di\u000berential operators (15) take place:\ne@\u0016\u0011\u00121\n\u001d@\n@t;r\u0013\n6=@\u0016\u0011\u00121\nc@\n@t;r\u0013\n: (50)12\nAt the same time, in order to ensure the gauge invariance of Maxwell equations the \feld A\u0016must be transformed\nwith transformation (48) only. Hence the equality (49) with the di\u000berential operator e@\u0016cannot be satis\fed:\ne@\u0016\u0012+2e\nc~A\u00166=e@\u0016\u00120+2e\nc~A0\n\u0016; (51)\ntherefore the gauge invariance of the Lagrangian is violated. This fact is consequence of that the \felds \t and A\u0016\nmove in di\u000berent Minkowski spaces: with the limit speeds \u001dandcaccordingly.\nIn order to ensure the gauge invariance of the Lagrangian (16) with electromagnetic \feld we should consider a \feld\neA\u0016=\u0010c\n\u001d';\u0000A\u0011\n\u0011(e';\u0000A): (52)\nThen from the transformation (48) we obtain the gauge transformations for the \feld eA\u0016:\n\u0012=\u00120+2e\nc~\u001f;eA\u0016=eA0\n\u0016\u0000e@\u0016\u001f=\u001a\ne'=e'0\u00001\n\u001d@\u001f\n@t\nA=A0+r\u001f\u001b\n: (53)\nAs a result we have\ne@\u0016\u0012+2e\nc~eA\u0016=e@\u0016\u00120+2e\nc~eA0\n\u0016: (54)\nIt should be noticed the following property:\ne@\u0016eA\u0016\u0011\u00121\n\u001d@e'\n@t;rA\u0013\n=\u00121\nc@'\n@t;rA\u0013\n\u0011@\u0016A\u0016: (55)\nInteraction of a charge ewith electromagnetic \feld A\u0016is described with an action\nSint=\u0000Ze\ncA\u0016dx\u0016: (56)\nTaking into account A\u0016dx\u0016='cdt\u0000Adr=e'\u001ddt\u0000Adr=eA\u0016dex\u0016and de\fning the new charge eease\nc=ee\n\u001dwe have\nSint=\u0000Ze\ncA\u0016dx\u0016=\u0000Zee\n\u001deA\u0016dex\u0016: (57)\nThus, interaction of the new charges with the new \felds does not change, for example e'=eee'. Moreover, the\nmagnetic \rux quantum does not change also: \b 0=\u0019~c\ne=\u0019~\u001d\nee. Then the action for a charged particle in the \feld is\nSp+int=Z\u0012mv2\n2+ee\n\u001dAv\u0000eee'\u0013\ndt: (58)\nTherefore equation of motion is\nmdv\ndt=\u0000ee\n\u001d@A\n@t\u0000eere'+ee\n\u001dv\u0002curlA\u0011eeeE+ee\n\u001dv\u0002H=eE+e\ncv\u0002H: (59)\nHere, the electric \feld eEis such that\neE=c\n\u001dE;eeeE=eE: (60)\nFrom the de\fnition of eEandHin Eq.(59) the \frst pair of Maxwell equations follows:\ncurleE=\u00001\n\u001d@H\n@t; (61)\ndivH= 0: (62)\nThen corresponding action for the electromagnetic \feld eA\u0016should be\nSf=\u00001\n16\u0019\u001dZ\neF\u0016\u0017eF\u0016\u0017d\n =1\n8\u0019Z\u0010\neE2\u0000H2\u0011\ndVdt; (63)13\nwhered\n =\u001ddtdV is an element of the 4D Minkowski space, eF\u0016\u0017=e@\u0016eA\u0017\u0000e@\u0017eA\u0016\u0011\u0010\neE;H\u0011\nis Faraday tensor. The\naction (57) can be written as\nSint=\u00001\n\u001dZ\ne\u001aeA\u0016dex\u0016dV=\u00001\n\u001d2Z\neA\u0016ej\u0016d\n; (64)\nwhere the 4D current is\nej\u0016=e\u001adex\u0016\ndt= (\u001de\u001a;e\u001av) =\u0010\n\u001de\u001a;ej\u0011\n; (65)\nhere the new charge and current are\ne\u001a=\u001d\nc\u001a;ej=\u001d\ncj: (66)\nThen the action describing the electromagnetic \feld and interaction of the current with the \feld is\nSint+f=Z\u0014\n\u00001\n\u001d2eA\u0016ej\u0016\u00001\n16\u0019\u001deF\u0016\u0017eF\u0016\u0017\u0015\nd\n: (67)\nVariation of the action \u000eSint+fgives the second pair of Maxwell equations:\ncurlH=1\n\u001d@eE\n@t+4\u0019\n\u001dej; (68)\ndiveE= 4\u0019e\u001a; (69)\nfrom where we can obtain conservation of charges both e\u001aand\u001ain a form:\ne@\u0016ej\u0016= 0)@e\u001a\n@t+ divej= 0)@\u001a\n@t+ divj= 0: (70)\nIn order to \fgure out the physical sense of \feld eE, chargee\u001aand speed \u001dlet us consider Maxwell equations for\nelectromagnetic \feld in a dielectric:\ncurlE=\u00001\nc@H\n@t\ndivH= 0\ncurlH=1\nc@D\n@t+4\u0019\ncj\ndivD= 4\u0019\u001af; (71)\nhereD=\"Eis electric displacement, \"is electric permittivity, \u001afis density of free charges, j=\u001afvis conduction\ncurrent. Then for a case \"= const we can write:\ncurlE=\u00001\nc@H\n@t\ndivH= 0\ncurlH=\"\nc@E\n@t+4\u0019\ncj\ndivE= 4\u0019\u001af\n\"\u0011curl (p\"E) =\u0000p\"\nc@H\n@t\ndivH= 0\ncurlH=p\"\nc@(p\"E)\n@t+4\u0019p\"\ncjp\"\ndiv (p\"E) = 4\u0019\u001afp\"\u0011curleE=\u00001\n\u001d@H\n@t\ndivH= 0\ncurlH=1\n\u001d@eE\n@t+4\u0019\n\u001dej\ndiveE= 4\u0019e\u001af; (72)\nwhere we have de\fned eE\u0011p\"E,e\u001af\u0011\u001af=p\",ej\u0011j=p\",\u001d\u0011c=p\". Comparing Eqs.(72) with Eqs.(61,62,68,69) we\nconclude that superconductor is equivalent to dielectric (in some e\u000bective sense, not in conductivity) with permittivity\n\"=c2\n\u001d2\u0018105; (73)\nand the speed\u001dis the light speed in SC medium if there were no the skin-e\u000bect and Meissner e\u000bect . We can see that\nthis permittivity is giant compared to the permittivity of true dielectrics (maximum \u0018103in segnetoelectrics). In\nvacuum\u001d=c, that is\"= 1 andeA\u0016\u0011A\u0016.\nIt should be noted that the dielectric permittivity is \"=c2=\u001d2in the long wave limit q<1=\u0018only. If frequency of\nelectromagnetic wave is such that ~!\u00152j\u0001j, then a photon can break a Cooper pair with transfer of its constituents\nin the free quasiparticle states. Hence in this area the strong absorption of the waves takes place. Thus, we can\nsuppose the permittivity \"is equal toc2=\u001d2= const in a frequency interval ~!<2j\u0001jonly, at ~!\u001d2j\u0001jwe suppose14\n\"!\"n(!), where\"n(!) is the dielectric function of normal metal. As we transit into the normal phase, where \u0001 = 0,\nthe permittivity c2=\u001d2losses sense, then \"must be replaced by the dielectric function of normal metal \"n(!) with the\nfollowing properties[14, 15]:\n\"n(!6= 0)\u001c4\u0019\u001b\n!; \"n(0) =1; (74)\nwhere the conductivity \u001bcan be assumed constant at low frequencies. The case \"(!= 0) is special and it will be\nconsider in next subsection.\nIn these terms, energy of the electric \feld Ueland density of the \rux of electromagnetic energy Selare\nUel=ED\n8\u0019=\"E2\n8\u0019=eE2\n8\u0019(75)\nSel=c\n4\u0019E\u0002H=\u001d\n4\u0019eE\u0002H; (76)\nso that div Sel=\u0000@Uel\n@t(here,Uelshould be understood as energy of the electromagnetic \feldeE2\n8\u0019+H2\n8\u0019). The\nelectromagnetic \feld tensors in a dielectric are F\u0016\u0017= (E;H) andH\u0016\u0017= (\u0000D;H) with an invariant1\n2F\u0016\u0017H\u0016\u0017=\nH2\u0000ED =H2\u0000\"E2=H2\u0000eE2. In new representation the tensors can be written in a symmetrical form:\neF\u0016\u0017=\u0010\neE;H\u0011\nandeF\u0016\u0017=\u0010\n\u0000eE;H\u0011\nwith the corresponding invariant1\n2eF\u0016\u0017eF\u0016\u0017=H2\u0000eE2=1\n2F\u0016\u0017H\u0016\u0017.\nAs a result of the above reasoning we can write the Lorentz-invariant gauge invariant Lagrangian:\nL=~2\n4m\u0012\ne@\u0016+i2ee\n\u001d~eA\u0016\u0013\n\t\u0012\ne@\u0016\u0000i2ee\n\u001d~eA\u0016\u0013\n\t+\u0000aj\tj2\u0000b\n2j\tj4\u00001\n16\u0019eF\u0016\u0017eF\u0016\u0017(77)\nfor the action\nS=1\n\u001dZ\nL(\t;\t+;eA\u0016;eA\u0016)\u001ddtd3r: (78)\nThe Lagrangian (77) is a sum of energy of spatial-time inhomogeneity (the \frst term), potential energy of the \feld \t\nand self-action (the second and third terms), and the last term is Lagrangian of the electromagnetic \feld eA\u0016which\nis a sum of the external \feld and the self-consistent internal \feld.\nB. Anderson-Higgs mechanism\nThe modulus-phase representation (13) can be considered as a local gauge U(1) transformation [26]:\n\t =j\tjei2ee\n~\u001d\u001f; (79)\nso that the gauge \feld is transformed as\neA0\n\u0016=eA\u0016+e@\u0016\u001f: (80)\nAfter transformations (79,80) the Lagrangian (77) takes a form:\nL=~2\n4m\u0012\ne@\u0016+i2ee\n~\u001deA\u0016\u0013\nj\tj\u0012\ne@\u0016\u0000i2ee\n~\u001deA\u0016\u0013\nj\tj\u0000aj\tj2\u0000b\n2j\tj4\u00001\n16\u0019eF\u0016\u0017eF\u0016\u0017\n\u0019~2\n4me@\u0016\u001ee@\u0016\u001e\u00002jaj\u001e2+1\n2a2\nb+~2\n4m\u00122ee\n~\u001d\u00132\n\t2\n0eA\u0016eA\u0016\u00001\n16\u0019eF\u0016\u0017eF\u0016\u0017; (81)\nwhere we have redesignated eA0\n\u0016!eA\u0016after the gauge transformation. We have neglected the nonlinear term\n~2\n2m\u00002ee\n~\u001d\u00012\t0\u001eeA\u0016eA\u0016which describes coupling between a Higgs boson \u001eand two photons, since in the pure limit\nthe Higgs-mode contribution is subleading to quasiparticles due to instability of the Higgs boson [29]. Compar-\ning this Lagrangian with Eq.(22) we can see that the Goldstone boson \u0012is absorbed into the gauge \feld eA\u0016, i.e.,\nAnderson-Higgs mechanism occurs. From this Lagrangian the equation for the \feld eA\u0016can be obtained as\ne@\u0016@L\n@\u0010\ne@\u0016eA\u0017\u0011\u0000@L\n@eA\u0017= 0)e@\u0016eF\u0016\u0017+1\n\u00152eA\u0017= 0; (82)15\nwhere we denoted\n\u00152\u0011m\u001d2\n8\u0019ee2\t2\n0=mc2\n8\u0019e2\t2\n0: (83)\nUsing the Lorentz gauge e@\u0016eA\u0016= 0, Eq.(82) is reduced to\ne@\u0016e@\u0016eA\u0017+1\n\u00152eA\u0017= 0)1\n\u001d2@2A\n@t2\u0000\u0001A+1\n\u00152A= 0\n1\n\u001d2@2e'\n@t2\u0000\u0001e'+1\n\u00152e'= 0: (84)\nTaking the \felds as harmonic modes A=A0ei(qr\u0000!t)ande'=e'0ei(qr\u0000!t)we obtain dispersion relation for photons\nin a superconductor:\n!2=\u001d2q2+\u001d2\n\u00152)E2=\u001d2p2+m2\nA\u001d4; (85)\nwhere\nmA=~\n\u0015\u001d/(Tc\u0000T)1=2(86)\nis mass of a photon, i.e., the Higgs mechanism takes place. It is noteworthy that the mass of a Higgs boson (29)\nem=p\n2~\n\u0018\u001dand the mass of a photon (86) are related as\nem(T)\nmA(T)=p\n2\u0014; (87)\nwhere\u0014=\u0015=\u0018is GL parameter. That is for type-I superconductors emmA.\nThe dispersion relation (85) can be rewritten in a form:\nq2=\u00001\n\u00152+!2\n\u001d2; (88)\nthat determines the penetration of electromagnetic \feld in a superconductor. Let us consider stationary case != 0,\nthenq2=\u00001\n\u00152, hence A=A0eiqx=A0e\u0000x=\u0015. We can see that Anderson-Higgs mechanism manifests itself in that the\nelectromagnetic (magnetic) \feld penetrates a superconductor in the depth \u0015(83), which is London penetration depth .\nAt the same time, for the nonstationary case !6= 0 the penetration depth is obtained as \u0015\u0010\n1\u0000\u00152!2\n\u001d2\u0011\u00001=2\n> \u0015. If\n!\u0015!c, where\n!c=\u001d\n\u0015=mA\u001d2\n~/(Tc\u0000T)1=2; (89)\nwe haveq2\u00150, that is electromagnetic \feld penetrates superconductor through entire its depth. Thus, the increase\nof the penetration depth with frequency and existence of the critical frequency (89), when the depth becomes in\fnitely\nlarge, is principal result of the extended TDGL theory . The London screening is caused by SC electrons with density\nns. However in a superconductor the normal electrons with density nnexist even at T= 0 due to impurities [30, 31],\natT6= 0 the normal component is thermally excited quasiparticles. The normal component causes absorption of\nelectromagnetic waves and the skin-e\u000bect. These processes smear the penetration e\u000bect, that will be considered in\nSect.V. Moreover, if the frequency ~!\u00152j\u0001j, then intensive absorption of the electromagnetic waves occurs due to\nthe breaking of Cooper pairs, hence in order to observe the penetration e\u000bect it must be ~!c<2j\u0001j. Using Eq.(27)\nin a form ~\u001dp\n2\n\u0018= 2j\u0001j, where\u0018=~p\n4mjajis coherence length, we can represent !cin another form:\n~!c= 2j\u0001j1p\n2\u0014: (90)\nThus, the condition ~!c<2j\u0001jcan be satis\fed in type-II superconductors only (where \u0014 > 1=p\n2). It should be\nnoted that this result has been obtained only for s-wave superconductors.16\nThe physical cause of the threshold !cis illustrated in Fig.6 and it is as follows. Electromagnetic \feld A\u0016is\nscreened in the depth \u0015by the induced supercurrent js=\u0000c\n4\u0019\u00152Aaccording to Higgs mechanism. Thus, the system\ncan respond to the external \feld A\u0016in the minimal length\u0015. For example, if we have a thin plate d\u001c\u0015, then\nmagnetic \feld penetrates it completely and does not a\u000bect its state [6, 32]. Let, at \frst, the electromagnetic wave\nwith wavelength \u0003 =2\u0019\u001d\n!\u001d\u0015fall on a superconductor. Then the depth \u0015is enough to screen this \feld. Now, let\nthe wavelength be \u0003 \u001c\u0015, then the \feld essentially changes within the length \u0015- it changes the sign as illustrated in\nFig.6. In this situation the SC system should screen the oppositely directed \felds in the length which is much less\nthan\u0015, that cannot be done. Hence the \felds with wavelength \u0003 .\u0015(that is!\u0015!c) cannot be screened by the\nsupercurrent. It should be noted that, according to our model, the speed of light in a superconductor is \u001d\u001cc, hence\nat given frequency !the wavelength in the superconductor is much less than in vacuum: \u0003 =2\u0019=\u001d\n!\u001cc\n!, precisely\nbecause of this property the screening e\u000bect can be observable: at the frequencies ~! < 2j\u0001jwe can get into the\ninterval \u0003 .\u0015. In turn, in\ruence of the Anderson-Higgs mechanism on the wave with !\u0015!cis reduced to increasing\nof the wavelength as \u0003 ( !) =\u0015\u0010\n!2\n\u001d2\u00152\u00001\u0011\u00001=2\n, so that \u0003 ( !c) =1.\nFigure 6: long-wave \feld (blue line) \u0003 \u001d\u0015and short-wave \feld (red line) \u0003 \u001c\u0015fall on surface of a superconductor. The\nlong-wave \feld can be screened by SC component in the London depth \u0015(solid blue line). The short-wave oscillation changes\nits sign within the depth \u0015(dash red line), hence the system cannot screen this \feld, because the screening supercurrent would\nhave to turn over many times within the minimal length \u0015. In\ruence of the Anderson-Higgs mechanism on this wave is reduced\nto increasing of the wavelength (solid red line). The screening by the normal component (skin e\u000bect) is not considered here.\nIf we suppose q= 0 in Eq.(85) then we obtain homogeneous oscillation (i.e., all points of a sample oscillate in a\nphase) with frequency !c(89). AtT= 0 we have\n!c(0) =\u001d\n\u0015(0)=\u001dFp\n3cr\n4\u0019e2n\nm\u0011\u001dFp\n3c!p\u001c!p; (91)\nwhere!pis the plasma frequency in metal. We can see this frequency is much lower than the ordinary plasma frequency.\nAs have been demonstrated before these oscillations are consequence of Anderson-Higgs mechanism: Goldstone boson\nis absorbed into gauge \feld and the electromagnetic oscillation mode (84) with spectrum (85) appears. As will\nbe demonstrated below Goldstone oscillations are not accompanied by longitudinal electric \feld (by oscillations of\ncharge density) and they are eddy Meissner currents which generate the transverse \feld div A= 0 only. Thus,\nthese electromagnetic oscillations are eigen oscillations of the SC system instead the phase oscillations. The critical\nfrequency!cis a minimal limit of frequencies which can propagate through the system , since for frequencies ! < !c\nwe haveq2<0. If electromagnetic wave with frequency !\u0015!cfalls on superconductor then the wave is carried by\nthese eigen oscillations hence the superconductor becomes transparent for such wave, at the same time for the wave\nwith frequencies ! Tc1;Tc2, moreover, the solution does not depend on the sign of \u000f. The sign\ndetermines the phase di\u000berence of the order parameters j\t1jei\u00121andj\t2jei\u00122:\ncos(\u00121\u0000\u00122) = 1 if\u000f<0\ncos(\u00121\u0000\u00122) =\u00001 if\u000f>0; (112)\nthat follows from Eq.(110). The case \u000f<0 corresponds to attractive interband interaction, the case \u000f>0 corresponds\nto repulsive interband interaction.\nAccording to the method described in Sect.II the two-component scalar \felds \t 1;2(r;t) should minimize an action\nSin the Minkowski space:\nS=1\n\u001dZ\nL(\t1;\t2;\t+\n1;\t+\n2)\u001ddtd3r: (113)\nThe LagrangianLis built by generalizing the density of free energy in Eq.(109) to the \"relativistic\" invariant form\nby substitution of covariant and contravariant di\u000berential operators (15) instead the gradient operator:\nL=~2\n4m1\u0010\ne@\u0016\t1\u0011\u0010\ne@\u0016\t+\n1\u0011\n+~2\n4m2\u0010\ne@\u0016\t2\u0011\u0010\ne@\u0016\t+\n2\u0011\n\u0000a1j\t1j2\u0000b1\n2j\t1j4\u0000a2j\t2j2\u0000b2\n2j\t2j4\u0000\u000f\u0000\n\t+\n1\t2+ \t 1\t+\n2\u0001\n; (114)\nwhere for \t 1and \t 2with the masses m1andm2accordingly the same speed \u001dis used. According to [38] the theory\nof a two-band superconductor can be reduced to GL theory of a single-band superconductor for equilibrium values\nof the orders parameters (time independent). In this model the orders parameters are related as \t 2=C(T)\t1at\nT!Tc;T >Tc1;Tc2, where the coe\u000ecient Cis\nC=q\na1\na2;if\u000f<0\nC=\u0000q\na1\na2;if\u000f>0: (115)\nThen the relation (27) can be written in a form:\np\n8jAjM\u001d2= 2j\u0001j; (116)\nwhereA=a1+a2C2+2\u000fC,B=b1+b2C4,M\u00001=1\nm1+C2\nm2andj\u0001j/\t0=p\n\u0000A=B is an e\u000bective gap. Obviously,\n\u001d\u0018vF1;vF2like in Eq.(27).21\nLet us consider movement of the phases only. Using the modulus-phase representation and assuming j\t1j= const\nandj\t2j= const Lagrangian (114) takes a form:\nL=~2\n4m1j\t1j2\u0010\ne@\u0016\u00121\u0011\u0010\ne@\u0016\u00121\u0011\n+~2\n4m2j\t2j2\u0010\ne@\u0016\u00122\u0011\u0010\ne@\u0016\u00122\u0011\n\u00002j\t1jj\t2j\u000fcos (\u00121\u0000\u00122) +L(j\t1j;j\t2j):(117)\nCorresponding Lagrange equations are\n~2\n4m1j\t1j2e@\u0016e@\u0016\u00121\u0000j\t1jj\t2j\u000fsin (\u00121\u0000\u00122) = 0 (118)\n~2\n4m2j\t2j2e@\u0016e@\u0016\u00122+j\t1jj\t2j\u000fsin (\u00121\u0000\u00122) = 0: (119)\nThe phases can be written in a form of oscillations:\n\u00121=\u00120\n1+Aei(qr\u0000!t)\n\u00122=\u00120\n2+Bei(qr\u0000!t); (120)\nwhere equilibrium phases \u00120\n1;2satisfy the relation (112). Substituting the phases in Eqs.(118,119) and linearizing\n(using the relation (112) in a form \u000fcos(\u00120\n1\u0000\u00120\n2) =\u0000j\u000fjand sin(\u00120\n1\u0000\u00120\n2) = 0) we obtain the following dispersion\nrelations:\n!2=q2\u001d2; (121)\nwhereinA=B, and\n(~!)2= 4j\u000fjj\t1j2m2+j\t2j2m1\nj\t1jj\t2j\u001d2+ (~q)2\u001d2; (122)\nwherein\nA\nB=\u0000m1\nm2j\t2j2\nj\t1j2: (123)\nFor symmetrical bands m1=m2\u0011mandj\t1j=j\t2jwe obtain\n(~!)2= 8j\u000fjm\u001d2+ (~q)2\u001d2; A =\u0000B: (124)\nThus, we can see that in two-band superconductors there are two modes of the phase oscillations: the common mode\noscillations with spectrum (121) like Goldstone mode (26) in single-band superconductors, and the oscillations of the\nrelative phase between two SC condensates (122,124) which can be identi\fed as Leggett's mode [39{41]. In a two\nband superconductor a current (\row) takes the following form:\nj=e~\u0012j\t1j2\nm1r\u00121+j\t2j2\nm2r\u00122\u0013\n=iei(qr\u0000!t)e~\u0012j\t1j2\nm1A+j\t2j2\nm2B\u0013\nq; (125)\nfrom where we can see that for Leggett's mode (123) the current is j= 0. Thus, unlike the Goldstone mode (121)\nwhich is the eddy current, Leggett's oscillations are not accompanied by any currents.\nLet us consider the Anderson-Higgs mechanism in a two-band superconductor. The gauge invariant form ( U(1)\u0002\nU(1) symmetry) of Lagrangian (114) is [35{38]:\nL=~2\n4m1\u0012\ne@\u0016+i2ee\n\u001d~eA\u0016\u0013\n\t1\u0012\ne@\u0016\u0000i2ee\n\u001d~eA\u0016\u0013\n\t+\n1+~2\n4m2\u0012\ne@\u0016+i2ee\n\u001d~eA\u0016\u0013\n\t2\u0012\ne@\u0016\u0000i2ee\n\u001d~eA\u0016\u0013\n\t+\n2\n\u0000a1j\t1j2\u0000b1\n2j\t1j4\u0000a2j\t2j2\u0000b2\n2j\t2j4\u0000\u000f\u0000\n\t+\n1\t2+ \t 1\t+\n2\u0001\n\u00001\n16\u0019eF\u0016\u0017eF\u0016\u0017: (126)\nAs in previous consideration the modulus-phase representation (79) can be considered as local gauge U(1) transfor-\nmation:\n\t1=j\t1jei2ee\n~\u001d\u001f1;\t2=j\t2jei2ee\n~\u001d\u001f2: (127)22\nThen the gauge \feld should be transformed as\neA0\n\u0016=eA\u0016+\u000be@\u0016\u001f1+\fe@\u0016\u001f2; (128)\nwhere\n\u000b=j\t1j2\nm1\nj\t1j2\nm1+j\t2j2\nm2; \f =j\t2j2\nm2\nj\t1j2\nm1+j\t2j2\nm2; (129)\nso that\n\u000b+\f= 1;j\t1j2\nm1\f=j\t2j2\nm2\u000b: (130)\nThe transformation (128) excludes the phases \u00121and\u00122from Lagrangian (126) individually leaving only their di\u000ber-\nence:\nL=~2\n4m1\u0012\ne@\u0016+i2ee\n\u001d~eA\u0016\u0013\nj\t1j\u0012\ne@\u0016\u0000i2ee\n\u001d~eA\u0016\u0013\nj\t1j+~2\n4m2\u0012\ne@\u0016+i2ee\n\u001d~eA\u0016\u0013\nj\t2j\u0012\ne@\u0016\u0000i2ee\n\u001d~eA\u0016\u0013\nj\t2j\n+~2\n4j\t1j2j\t2j2\nj\t1j2m2+j\t2j2m1e@\u0016(\u00121\u0000\u00122)e@\u0016(\u00121\u0000\u00122)\u00002\u000fj\t1jj\t2jcos (\u00121\u0000\u00122) +L\u0010\nj\t1j;j\t2j;eF\u0016\u0017eF\u0016\u0017\u0011\n:(131)\nThus, the gauge \feld eA\u0016absorbs the Goldstone bosons \u00121;2so that the Lagrangian becomes dependent on the di\u000berence\n\u00121\u0000\u00122only. The equation for \u00121\u0000\u00122has a form:\ne@\u0016e@\u0016(\u00121\u0000\u00122)\u00004\n~2j\t1j2m2+j\t2j2m1\nj\t1jj\t2j\u000fsin (\u00121\u0000\u00122) = 0: (132)\nThis equation is similar to the sin-Gordon equation, but the coe\u000ecient \u000fis a function of di\u000berence of the equilibrium\nphases\f\f\u00120\n1\u0000\u00120\n2\f\f= 0;\u0019according to Eq.(112). Eq.(132) can be linearized using \u000fcos(\u00120\n1\u0000\u00120\n2) =\u0000j\u000fj, sin(\u00120\n1\u0000\u00120\n2) = 0\nand small oscillations (120), that gives the following spectrum:\n(~!)2= 4j\u000fjj\t1j2m2+j\t2j2m1\nj\t1jj\t2j\u001d2+ (~q)2\u001d2;\nwhich coincides with Leggett's mode spectrum (122). Thus, we can see that in two-band superconductors the common\nmode oscillations with the spectrum (121) are absorbed into the gauge \feld eA\u0016like in single-band superconductors.\nAt the same time, the oscillations of the relative phase between two SC condensates (Leggett's mode) \"survive\" due to\nthat these oscillations are not accompanied by current j= 0 - Eqs.(123,125). Hence Leggett's mode can be observable,\nthat is con\frmed in experiment [42].\nV. DAMPING AND RELAXATION\nIn the previous sections we have considered eigen harmonic oscillations of the order parameter - Higgs mode and\nGoldstone modes absorbed into the gauge \feld except Leggett's mode (in multi-band systems). At the same time,\nmovement of the normal component is accompanied with friction and generation of heat by Joule-Lenz law Q=j2\nn=\u001b,\nwhere jn=ennvnis normal current, nn=n\u0000ns=n\u00002j\tj2is density of the normal component, vnis its speed, \u001b\nis conductivity. Thus, to support the normal movement some electric \feld Emust act, the current and the \feld are\nconnected with Ohm's law: jn=\u001bE. Then, instead the harmonic oscillations, we will have situations illustrated in\nFig.1b - relaxation, Fig.1c - damping oscillations, Fig.1d - forced (undamped) oscillations under the action of external\n\feld.\nThe energy dissipation is accounted with the Rayleigh dissipation function Rwhich determines speed of change of\nthe energy of the system WasdW=dt =\u00002R, that isQ= 2R. As a rule, the dissipative force (friction) is proportional\nto generalized velocity _ q:F=\u0000k_q(kis a friction coe\u000ecient), then R=1\n2k_q2andF=\u0000dR\nd_q. Corresponding equation\nof motion is\nd\ndt@L\n@_q\u0000@L\n@q+@R\n@_q= 0: (133)\nLet us consider several important examples using Lagrangian (77).23\nA. Wave skin-e\u000bect in superconductors\nLet monochromatic electromagnetic wave eA\u0016= (0;A) (in a gauge '= 0, div A= 0) with frequency ~! < 2j\u0001j\nfall on a superconductor perpendicularly to its surface. The wave induces eddy currents both superconducting jsand\nnormal jn. For normal electrons an equation of motion is m_v=eE(t)\u0000m\n\u001cphv, where E(t) =\u00001\nc@A\n@tis the electric\n\feld,\u001cphis the mean free path time of electrons ( \u001c\u00001\nphis frequency of collisions of electrons with phonons or with the\nlattice defects). Then equation for the normal current is\nm\nenndjn\ndt+m\nennjn\n\u001cE=eE(t): (134)\nFor the harmonic \feld E=E0ei(qr\u0000!t)and current jn=jn0ei(qr\u0000!t)we have the Ohm's law in a form jn=\n\u001b1+i!\u001cph\n1+(!\u001cph)2E, where\u001b=\u001cph\nme2nnis conductivity of the normal component. We will consider regime of normal skin-\ne\u000bect only, that is when !\u001cph\u001c1. The Joule-Lenz heat is Q=\u001b!E2, where\u001b!=\u001b\n1+(!\u001cph)2. Respectively, the\nRayleigh dissipation function is\nR=1\n2\u001b!(1 +i!\u001cph)E2=1\n2e\u001b!(1 +i!\u001cph)eE2: (135)\nHere,eE\u0011\u00001\n\u001d@A\n@t\u0011\u0000e@0eA\u0016=e@0eA\u0016, so that\u001bE2=e\u001beE2wheree\u001b=\u001d2\nc2\u001b. Unlike the Joule-Lenz law the Rayleigh\ndissipation function has both an active part and a reactive part determined by the term i!\u001cphfor agreement with the\nresults of London theory [32]. The active part determines dissipation of the electromagnetic energy, the reactive part\ndetermines the phase shift of the current jnrelatively to the \feld Edue to inertia of the system. Then an analog of\nEq.(133) using Eq.(82) can be written:\ne@\u0016@L\n@\u0010\ne@\u0016eA\u0017\u0011\u0000@L\n@eA\u0017+1\n\u001d@R\n@\u0010\ne@0eA\u0017\u0011= 0: (136)\nObviously, this equation is not Lorentz covariant due to the dissipative term. Dissipation distinguishes a time direction,\ni.e., violates the time symmetry t$\u0000twhich is symmetry for the Lorentz boost. Using the gauge '= 0, div A= 0\nEq.(136) is reduced to\n1\n\u001d2@2A\n@t2\u0000\u0001A+1\n\u00152A+4\u0019e\u001b!(1 +i!\u001cph)\n\u001d2@A\n@t= 0: (137)\nTaking the \feld as harmonic mode A=A0ei(qr\u0000!t)we obtain a dispersion relation for photons in a superconductor:\n!2=\u001d2q2+\u001d2\n\u00152!\u0000i4\u0019e\u001b!!)q2=!2\n\u001d2\u00001\n\u00152!+i4\u0019\nc2\u001b!!: (138)\nHere, we have denoted:\n1\n\u00152!\u00111\n\u00152+4\u0019\nc2\u001b!!2\u001cph=1\n\u00152\u0012\n1 +nn\nns(!\u001cph)2\n1 + (!\u001cph)2\u0013\n: (139)\nObviously, \u0015!\u0014\u0015, atT= 0 in pure metal nn= 0 takes place [30, 31] hence we have \u0015!=\u0015in this case. The\nobtained expression (138) di\u000bers from the result of London theory q2=\u00001\n\u00152!+i4\u0019\nc2\u001b!!by a term!2\n\u001d2. However, at\nsmall frequencies ~!\u001cj\u0001j,!2\n\u001d2\u001c1\n\u00152this term can be omitted, main contribution is given by the Meissner e\u000bect \u00001\n\u00152\nand by the skin-e\u000bect of the normal component i4\u0019\nc2\u001b!!. At frequencies !\u0018!c=\u001d\n\u0015the term!2\n\u001d2becomes important,\nthat is discussed in Sect.III. The term!2\n\u001d2has the following nature. The Maxwell equations for the electromagnetic\n\feld in normal metal are\n8\n<\n:curlH=4\u0019\nc\u001bE+1\nc@D\n@t\ncurlE=\u00001\nc@H\n@t9\n=\n;)\u0001H=4\u0019\u001b\nc@H\n@t+\"n(!)\nc2@2H\n@t2; (140)\nwhere D=\"nEhas been used. Taking the \feld as harmonic mode H=H0ei(qr\u0000!t)we obtain the dispersion relation\nfor photons in metal:\nq2=i4\u0019\nc2\u001b!+\"n(!)\nc2!2: (141)24\nIn good metals4\u0019\u001b\n!\u001d\"n(!) (if!6= 0), then the second term in Eq.(141) can be neglected [14, 15]. Hence we obtain\nusual expression for the skin-e\u000bect: q2=i4\u0019\nc2\u001b!. In superconductors the conductivity is determined with the normal\ncomponent \u001b=\u001cph\nme2nnwhich is\u001b!0 in pure system at small temperatures T\u001cTc, so that\u001b\n!.\"can be. Thus,\nfor SC materials the induction term1\nc@D\n@tis as important as the ohmic term \u001bE, unlike normal metals where the\ninduction term can be neglected for quasistationary \felds.\nIt should be noted that the accounting of the complex conductivity from the \frst London equation js=\u0000ic2\n4\u0019\u00152!E\n(i.e.,\u001bs=\u0000ic2\n4\u0019\u00152!) is unnecessary because the induced electric \feld \u00001\nc@A\n@tdrives the eddy currents which are\nGoldstone oscillations, which, in turn, are absorbed into the gauge \feld A. Thus, this complex conductivity has\nalready hidden in the nondissipative part of the \feld equation (137). As we have seen before, the London equations\n(106,107) are not equivalent to Eqs.(84,104,105) since the extraction of current jviolates Anderson-Higgs mechanism.\nAs a result of the extraction we obtain the \frst London equation (106) which is the second Newton law for SC\nelectrons, i.e., it is equation for an ideal conductor. Since we lose a term1\n\u001d2@2eA\n@t2then the frequency dependence\nis caused by the complex conductivity \u001bsonly, that provides the low-frequency regime such that the penetration\ndepth does not depend on frequency: L(!) =\u0015[6] (if we exclude the wave skin-e\u000bect due to normal electrons where\nL(!) =q\nc2\n2\u0019\u001b!).\nFigure 7: Schematic image of dependence of the penetration depth Lon the frequency !for di\u000berent densities of the normal\ncomponent nnand for the London theory. At !!0 the penetration depth is equal to the London depth \u0015for all regimes. If\nnn= 0 then for !\u0015!cSC material becomes transparent i.e., L!1 , however at frequencies ~!\u00152j\u0001jstrong absorption\ntakes place (the dashed lines). For large normal density (low SC density) nn.nthe result for Lis close to the London theory\n(Lweakly depends on frequency). It should be noted that for di\u000berent normal densities nnwe have di\u000berent parameters !c,\nj\u0001jand\u0015.\nFrom Eq.(138) we obtain:\nq=4s\n1\n\u00154!\u0012!2\n\u001d2\u00152!\u00001\u00132\n+\u00124\u0019\nc2\u001b!!\u00132\u0010\ncos'\n2+isin'\n2\u0011\n; (142)\nwhere\n'= arccotRe\u0000\nq2\u0001\nIm (q2)= arccot\u00001\n\u00152!\u0010\n1\u0000!2\n\u001d2\u00152\n!\u0011\n4\u0019\nc2\u001b!!; (143)\nthat corresponds to attenuation of the wave A=A0ei(qx\u0000!t)in the depth of the superconductor which occupies\nhalf-spacex>0. Then substituting q(142) in this \feld Awe obtain the penetration depth:\nL=\u0015!\n4q\u0000!2\n\u001d2\u00152!\u00001\u00012+\u00004\u0019\nc2\u00152!\u001b!!\u000121\nsin'\n2; (144)25\nin a sense A=A0e\u0000x=Lei(Re(q)x\u0000!t). Let us consider the following limit cases:\n1.!= 0. Then '=\u0019and from Eq.(144) we have L=\u0015. i.e., low-frequency \feld is screened like static \feld.\n2. The frequency is equal to the critical frequency (89): !=!c\u0011\u001d\n\u0015!(it should be noted that at nn!0 we have\n\u0015!=\u0015). In this case we have '=\u0019\n2, hence\nL=1q\n2\u0019e2\nmc2nn!c\u001cph\n1+(!c\u001cph)2: (145)\nWe can see that at nn!0 we have result of Sect.III, where the penetration depth becomes in\fnitely large\nL\u001d\u0015. Formally we can suppose \u001cph= 0, thenL!1 too, however the condition \u001cph= 0 means full blocking\nof movement of normal electrons i.e., the substance ceases to be conductor.\n3.! > !c. Then from Eq.(143) we can see 0 \u0014' <\u0019\n2. Ifnn!0 then'!0, hence from Eq.(144) we have\nL!1 , i.e., the superconductor becomes transparent for such electromagnetic waves.\n4.!0 and for dirty superconductors for T= 0 even [31] the normal electrons are present.\nInduced electric \feld E=\u00001\nc@A\n@tin such oscillations causes movement of the normal electrons according to Ohm's law\njn=\u001bEand dissipation in a form of Joule-Lenz heat Q=j2\nn=\u001b=\u001bE2occurs. Let us consider a homogeneous mode,\ni.e.,q= 0 is supposed in Eq.(138), and \u0015!=\u0015,\u001b!=\u001bare supposed for simplicity. Then we obtain an equation for\nfrequency:\n!2+i4\u0019\u001d2\nc2\u001b!\u0000\u001d2\n\u00152= 0; (147)\nwhose solution is\n2!=\u0000i4\u0019\u001d2\nc2\u001b\u0006s\n4\u001d2\n\u00152\u0000\u00124\u0019\u001d2\nc2\u001b\u00132\n: (148)\nThis solution determines oscillations of electromagnetic \feld A=A0e\u0000i!t=A0e\u0000t=\u001ce\u0000iRe(!)t, where 1=\u001c=\u0000Im(!)\nis the decay time. We can see that the oscillations are sensitive to the normal density nnvia conductivity \u001b=\u001cph\nme2nn.\nLet us consider the following limit cases:\n1.T!Tc. Then\u0015(T)!1 andnn\u0019n, hence!=\u0000i4\u0019\u001d2\nc2\u001b, that is\u001c=c2\n4\u0019\u001d2\u001band Re(!) = 0. Thus, monotonic\nattenuation of the electromagnetic excitation occurs.\n2.nn= 0, that takes place at T= 0 in pure material [31]. Then !=\u001d\n\u0015which coincides with the critical frequency\n!c(89), and Im( !) = 0. Thus, harmonic oscillations of the electromagnetic \feld in superconductor occur.\nObviously, if ~!c\u00152j\u0001j, then a quant of these oscillations breaks Cooper pair, hence intensive absorption of these\noscillations takes place. As we can see from Eq.(148) the frequency !is sensitive to the normal density nnin the\nsense that the normal electrons caused attenuation of the oscillations. The crossover between overdamped (at high\nT) and underdamped (low T) regimes can be found from Eq.(148) in a form:\nc\n\u0015(T)=2\u0019\u001d\nc\u001b(T): (149)\nIf take\u001cph\u001810\u000010c,n\u00181022cm\u00003,\u001d\u0018108cm=c,\u0015=\u0015(0)\u0012\n1\u0000\u0010\nT\nTc\u00114\u0013\u00001=2\nandnn=n\u0010\nT\nTc\u00114\n, then we can\nevaluate the temperature of crossover:T\nTc\u00180:5. However these oscillations are bulk oscillations, and, as we could see\nin previous subsection, propagation of \ructuations of electromagnetic \feld (with q6= 0) is blocked by the skin-e\u000bect\nifnn6= 0. That is any \feld, by which such eigen oscillations are tried to excite, will be re\rected from surface of a\nsuperconductor. Thus, experimental observation of such eigen electromagnetic oscillations is problematically. Only\nif normal electrons are absent nn= 0, such \ructuations can show themselves by the transparency of superconductor\nfor electromagnetic wave with frequency !c Tcthe eigen electromagnetic oscillations do not turn into relaxation of charge\ndensity with relaxation time1\n4\u0019\u001b. According to the rulec2\n\u001d2!\"n(!) (either supposing q= 0 in Eq.(141)) we have\n\u001c(T >Tc) =\"n(!)\n4\u0019\u001b, which is relaxation time for eddy current in metal.\nC. Relaxation of a \ructuation\nLet a \ructuation is formed in some region so that j\u0001(x)j>j\u0001j, wherej\u0001jis the equilibrium value - Fig.8, but\nn=nn+ns= const. Let us consider relaxation of this bubble which can be both damped oscillation and monotonous27\nrelaxation to the equilibrium - Fig.1b,c. After the \ructuation is formed the system tends to the equilibrium: the \row\nof the normal component is directed to the bubble, at the same time the \row of the super\ruid component is directed\nfrom the bubble so that the total \row is j=nsvs+nnvn= 0. Then div js= div(nsvs) =\u0000div(nnvn)\u0019\u00001\n\u0018nnvn,\nbecause the changes of super\ruid and normal components occur in the coherence length \u0018(T)/(Tc\u0000T)\u00001=2. The\ndensity of the super\ruid component is ns= 2j\tj2. The normal motion is accompanied with friction f=\u0000m\n\u001cphvn,\ncorresponding Rayleigh dissipation function is R=m\n2\u001cphnnv2\nn=1\n2\u001b\ne2v2\nn. Using the continuity equation@ns\n@t=\u0000divjs\nandnn\u0019nthe dissipation function takes a form:\nR=\u001b\ne22\u00182\nn2\u0012\n\t@\t+\n@t+ \t+@\t\n@t\u00132\n=\u001b\ne22\u00182\u001d2\nn2\u0010\n\te@0\t++ \t+e@0\t\u00112\n: (150)\nThen equation for the \feld \t is\ne@\u0016@L\n@\u0010\ne@\u0016\t+\u0011\u0000@L\n@\t++1\n\u001d@R\n@\u0010\ne@0\t+\u0011= 0)~2\n4me@\u0016e@\u0016\t +a\t +bj\tj2\t +\u001b\ne24\u00182\u001d\nn2\u0010\n\te@0\t++ \t+e@0\t\u0011\n\t = 0:(151)\nAnalogously to previously considered problem about the skin-e\u000bect this equation is not Lorentz covariant due to the\ndissipative term. Using the modulus-phase representation (13) and linearizing with help j\tj=p\u0000a\nb+\u001e\u0011\t0+\u001e\n(wherej\u001ej\u001c\t0) we obtain an equation for small deviations of the super\ruid density \u001eand for the phase \u0012accordingly:\n~2\n4me@\u0016e@\u0016\u001e+ 2jaj\u001e+\u001b\ne28\u00182\u001d\nn2\t2\n0e@0\u001e= 0; (152)\ne@\u0016e@\u0016\u0012= 0: (153)\nThe oscillations of phase remains without damping in this approximation, since, as has been demonstrated in Sect.II,\nthe Goldstone mode is the eddy supercurrent. For harmonic oscillations \u001e(r;t) =\u001e0ei(qr\u0000!t)we obtain their dispersion\nrelation in a form:\n!2\u0000q2\u001d2\u0000!2\n0+i2\r!= 0; (154)\nwhere!2\n0=8jajm\u001d2\n~2=4j\u0001j2\n~2is a characteristic frequency of the system and \r=\u001b\ne216\u00182m\u001d2\n~2n2\t2\n0is an attenuation\nconstant. We can see that !0/(Tc\u0000T)1=2, at the same time \r/\u00182\t2\n0= const at T!Tc, i.e.,!0\u001c\r. Thus, the\nregime of overdamped oscillation takes place - the \feld \t( r;t) is monotonically relaxing to the thermodynamically\nsteady state \t 0. Then evolution of the \ructuation is \u001e(t)/e\u0000t=\u001c, where\n\u001c=2\r\n!2\n0/(Tc\u0000T)\u00001(155)\nis the relaxation time for a homogeneous ( q= 0) mode in the limit !0\u001c\r. Another solution with \u001c= 1=2\rcan\nbe omitted because this mode decays much faster. The relaxation time is lifetime of the \ructuation: the closer\ntemperature to Tcthe larger size of a \ructuation \u0018(T!Tc)!1 and it lives longer \u001c(T!Tc)!1 . The new phase\nis a \ructuation of in\fnite size (\flls the entire system) and in\fnite lifetime. The relaxation time (155) corresponds to\nthe reduced equation describing the relaxation only:\nq2\u001d2+!2\n0\u0000i2\r!= 0)\u0000~2\n4m\u0001\t +a\t +bj\tj2\t + 2\r~2\n4m\u001d2@\t\n@t= 0: (156)\nThis equation can be made dimensionless using a dimensionless order parameter = \t=\t0:\n\u001c@ \n@t=\u00182\u0001 + \u0000j j2 ; (157)\nwhere\u0018(T) =q\n~2\n4mja(T)jis temperature-dependent coherence length and \u001c(T) =~2\r\n4mja(T)j\u001d2=\u00182(T)\r\n\u001d2is the\ntemperature-dependent relaxation time. Thus, due to the strong damping at T!Tcthe equation (151) is re-\nduced to Eq.(157) which is analogous to TDGL equation (3). Thus, the TDGL equation is a limit case of the extended\nTDGL theory proposed here . We can see that in consequence of the strong damping the monotonous relaxation of\n\ructuation with the relaxation time (155) takes place. Moreover, this means strong damping of the free Higgs mode,\nso that observation of these oscillations is problematical at T!Tc. On the other hand at T!0 we can have such\nsituation that !0>\rthen the Higgs mode can be more clearly observable.28\nFigure 8: relaxation of a \ructuation of modulus of the order parameter j\tj/j\u0001j. The \ructuation is accompanied by changes\nof density of SC electrons ns= 2j\tj2and normal electrons nnso thatns+nn=n= const. The relaxation is carried out by\ncounter\rows of SC and normal components so that nsvs+nnvn= 0. The dissipation is caused by the friction f=\u0000m\n\u001cphvnof\nthe normal component.\nVI. RESULTS\nIn this work we have formulated the extended TDGL theory which is generalization of the GL theory for the\nnonstationary regimes: damped eigen oscillations (including relaxation) and forced oscillations of the order parameter\n\t(r;t) under the action of external \feld. In this theory, instead the GL functional (1), we propose action (78)\nwith Lorentz invariant Lagrangian (77) for the complex scalar \feld \t = j\tjei\u0012and the gauge \feld A\u0016= (';A) in\nsome 4D Minkowski space f\u001dt;rg, where speed \u001dis determined with dynamical properties of the system. At the\nsame time, the dynamics of conduction electrons remains non-relativistic. For accounting of movement of the normal\ncomponent, which is accompanied by friction, we have used approach with the Rayleigh dissipation function which\ndetermines speed of the dissipation. This makes the theory is not Lorentz covariant since dissipation distinguishes\na time direction, i.e., violates the time symmetry t$\u0000twhich is symmetry of the Lorentz boost. Our results are\nfollows:\n1) The SC system has two types of collective excitations: with an energy gap (quasi-relativistic spectrum) E2=\nem2\u001d4+p2\u001d2(whereemis the mass of a Higgs boson, so that em\u001d2= 2j\u0001j) - free Higgs mode, and with acoustic\n(ultrarelativistic) spectrum E=p\u001d- Goldstone mode, which are oscillations of the order parameter \t( t;r). The\nlight speed \u001dis determined with dynamical properties of the system (27), and it is much less than the vacuum\nlight speed: \u001d=vF=p\n3\u001cc. The Higgs mode is oscillation of modulus of the order parameter j\t(t;r)jand it\ncan be considered as sound in the gas of above-condensate quasiparticles. It should be noted that the free Higgs\nmode in a pure superconductor is unstable due to both strong damping of these oscillations at T!Tc, so that\naperiodic relaxation takes place and decay into above-condensate quasiparticles since E(q)\u00152j\u0001j. At the same time,\nvarious manifestations of Higgs mechanism plays important role in the dynamics of SC system. The Goldstone mode\nis oscillations of the phase \u0012which are eddy current and they are absorbed into the gauge \feld A\u0016according to\nAnderson-Higgs mechanism.\n2) In two-band superconductors the Goldstone mode splits into two branches: common mode oscillations r\u00121=r\u00122\nwith the acoustic spectrum, and the oscillations of the relative phase \u00121\u0000\u00122between two SC condensates (for symmet-\nrical condensates we have r\u00121=\u0000r\u00122) with energy gap in spectrum determined by interband coupling - Leggett's\nmode. The common mode oscillations are absorbed into the gauge \feld A\u0016like in single-band superconductors, at\nthe same time Leggett's mode \"survives\" due to these oscillations are not accompanied by current. Hence Leggett's\nmode can be observable that is con\frmed in experiment [42].\n3) From the gauge invariance of Lagrangian (77) it follows that superconductor is equivalent to dielectric (in some\ne\u000bective sense, not in conductivity) with permittivity \"=c2\n\u001d2\u0018105. At the same time, inside superconductor the\npotential electric \feld is absent E=\u0000r'= 0 as consequence of boundary condition (97). Hence the dielectric\npermittivity \"=c2\n\u001d2is correct for the induced electric \feld E=\u00001\nc@A\n@tonly, and the speed \u001dis the speed of light\nin SC medium if there were no skin-e\u000bect and Meissner e\u000bect. For electrostatic \feld E=\u0000r'the permittivity is\n\"(!= 0;q= 0) =1like in metals. This fact explains experimental results in [19] which illustrate that the external\nelectric \feld does not a\u000bect signi\fcantly the SC state unlike predictions of the covariant extension of GL theory\n[17, 18], since in our model a superconductor screens electrostatic \feld by metallic mechanism in the microscopic\nThomas-Fermi length, unlike the screening of magnetic \felds by SC electrons in the London depth. Thus, unlike the\nmodels [17{20], electrostatics and magnetostatics take di\u000berent forms despite Lorentz covariance of the model. It\nshould be noted a photon with frequency ~!\u00152j\u0001jcan break a Cooper pair with transfer of its constituents in the\nfree quasiparticle states. Hence in this frequency region the strong absorption of the waves takes place. Thus, the\npermittivity \"is equal to c2=\u001d2only when 0 <~! <2j\u0001j. At ~!\u001d2j\u0001jwe can suppose \"=\"n(!), where\"n(!) is29\nthe dielectric function of normal metal.\n4) Unlike popular opinion [25] the Goldstone mode cannot be associated with the plasmon mode. We have demon-\nstrated - Eqs.(102,103), that Goldstone oscillations cannot be accompanied by oscillations of charge density, they\ngenerate the transverse \feld div A= 0 only and they are currents for which div j= 0 as result of the boundary\ncondition (97). This is expressed in that the Anderson-Higgs mechanism takes place: the oscillations of the phase \u0012\nare absorbed into the gauge \feld A\u0016, hence the pure Goldstone oscillations become unobservable. It should be noted\nthat the \fxing of the transverse gauge div A= 0 is result of the spontaneously broken gauge symmetry in SC phase.\nVariation of order parameter is impossible in distances \u0018\u0015TFand time intervals \u00181=!p, since the SC system does\nnot have reserve of the free energy to compensate the spatial-time variation energies. Therefore on these spatial-time\nscales (TF length and plasma frequency) the SC order parameter \t (or \u0001) losses any sense and a superconductor\nshould behave like ordinary metal. Thus, the plasma (Thomas-Fermi) screening and the plasma oscillations exist in\nmetal independently on its state.\n5) As result of interaction of the gauge \feld A\u0016with the scalar \feld \t with spontaneously broken U(1) symmetry a\nphoton obtains mass mA=~\n\u0015\u001din a superconductor, i.e., the Anderson-Higgs mechanism takes place, which manifests\nitself as the penetration depth L(0) =\u0015. The penetration depth depends on frequency: the depthL(!)increases\nwith frequency and such frequency exists !c=\u001d\n\u0015, that the depth becomes in\fnitely large: L(!\u0015!c) =1. This\nis principal result of the extended TDGL theory. It should be noted that ~!c<2j\u0001jfor type-II superconductors\nonly. However the normal component causes absorption of electromagnetic waves and the wave skin-e\u000bect occurs,\nas a result the penetration depth Lis very sensitive to the normal density nn, that is for observation of the e\u000bect\nL(!\u0015!c)!1 the normal electrons should be absent nn= 0. Thus, we have shown that atT= 0 pure type-\nII superconductors (in understanding of a monocrystalline sample without defects and impurities) should become\ntransparent for electromagnetic waves with frequency !from interval !c\u0014! < 2j\u0001j=~, that can be subject for\nexperimental veri\fcation , that illustrated in Fig.7. Thereby observation of this e\u000bect can be di\u000ecult. At the other\nhand, when Goldstone mode is absorbed into gauge \feld, the electromagnetic oscillation mode (84) with spectrum\n(85) appears instead the phase oscillations. These electromagnetic oscillations are eigen oscillations of the SC system\ninstead phase oscillations and they are eddy Meissner current. The critical frequency !cis a minimal limit of\nfrequencies which can propagate through the system. And besides this frequency is much lower than the ordinary\nplasma frequency: !c(T= 0) =\u001d\nc!p\u001c!p. If electromagnetic wave with frequency !\u0015!cfalls on superconductor\nthen the wave is carried by these eigen oscillations hence the superconductor becomes transparent for such wave, at\nthe same time for the wave with frequencies ! < !cthe carrier is absent hence such wave re\rects from the surface\nof superconductor. Moreover, the analogy between above-described electrodynamics of bulk superconductors and\nthe phase wave in a plane of Josephson junction exists that is presented in a Table I. Results of calculation of the\ntransparency intervals \u0017c\u0014\u0017 <2j\u0001j\nhfor pure elemental type-II superconductors niobium Nb, technetium Tcand\nvanadium VatT= 0 are presented in a Table II. The interval is in terahertz range.\n6) The London electrodynamics is a low frequency limit of the extended TDGL theory, i.e., the term1\n\u001d2@2A\n@t2can be\nneglected in the equation for the \feld (137), but the Rayleigh dissipation function (135), which gives the term \u0018\u001b@A\n@t,\nhas to be accounted. At high frequencies !\u0018!c, we should use the extended TDGL theory. At the same time, at\nthe large normal density (low SC density) nn!nthe results are close to the London theory at high frequencies too.\nMoreover, the TDGL equation (3) is a limit case of the extended TDGL theory at T!Tc: due to strong damping\nof oscillations of j\tjthe monotonous relaxation of a \ructuation to the equilibrium state takes place, the process is\ndescribed with Eq.(157) which is analogous to the TDGL equation.\nIn conclusion, it should be noted that the extended TDGL theory has a form of a scalar \feld theory with sponta-\nneously broken gauge symmetry interacting with a gauge \feld. However, the boundary condition (95) violates this\nanalogy due to the fact that, unlike \felds, a superconductor is \fnite system with some surface on which corresponding\nboundary conditions must be set by determining of current through the surface. Consequently, electrostatics and\nmagnetostatics take di\u000berent forms despite Lorentz covariance of the model. In addition, it should be noted, that\nGL theory is two-liquid approximation (the total electron density is sum of the SC density and the normal den-\nsityn=ns+nn). At the same time, in dynamics and electrodynamics of superconductors the coherency factors\nuk0uk\u0006vk0vkanduk0vk\u0006vk0ukplay some role [6]: they give corrections at temperature slightly less than Tc, i.e.,\nwhen there are many quasiparticles. At low temperatures T\u001cTc, there are few quasiparticles, hence the a\u000bect of\nthe coherency factors is negligible.30\nAppendix A: \"Derivation\" of the second London equation from the \frst London equation and vice versa\nFollowing [46] let us write the Newton equation for SC electrons (they do not experience friction) in the electric\n\feldE:mdv\ndt=eE, which can be given a form:\ndjs\ndt=nse2\nmE; (A1)\nwhere js=nsevis supercurrent, nsis density of SC electrons. Eq.(A1) is the \frst London equation (7). Making the\noperation curl for both sides of the equation and taking into account the Maxwell equation curl E=\u00001\nc@H\n@twe obtain:\nd\ndt(curljs) =\u0000nse2\nmc@H\n@t: (A2)\nIntegrating over time we obtain:\ncurljs=\u0000nse2\nmc(H\u0000H0); (A3)\nwhere H0- a constant of integration which does not depend on time, but it can be function of coordinates. Thus,\nby setting the \feld H0we set the initial condition, i.e., con\fguration of the \feld His determined by both response\nof the medium and the initial \feld . Supposing H0= 0 (a sample is introduced into magnetic \feld) we obtain the\nsecond London equation (8). However supposing H06= 0 (a sample in the normal state is in the magnetic \feld H0,\nthen it is cooled below the transition temperature Tc), we obtain the freezing of the magnetic \feld inside the sample.\nThus, the \feld H0is a constant of motion, that characterizes an ideal conductor: curl E=\u00001\nc@B\n@t)B= const\nsince we have inside the sample E=\u001aj= 0 due to ideal conductivity \u001a= 0. Therefore Eq.(A3) does not describe\nthermodynamically steady state, this equation describes the ideal conductor which pushes out or freezes magnetic\n\feld due to the electromagnetic induction and Lenz's rule.\nFrom other hand, following [6] we can di\u000berentiate the second London equation in a form js=\u0000c\n4\u0019\u00152A, where\ndivA= 0 (i.e., A=A?), with respect to time:\n@js\n@t=\u0000c\n4\u0019\u00152@A\n@t=c2\n4\u0019\u00152E=nse2\nmE: (A4)\nThus, we obtain the \frst London equation (A1) which is the second Newton law for SC electrons, i.e., it is equation\nfor an ideal conductor. At the same time, the second London equation is result of minimization of the free energy\nfunctional, i.e., it is equation for a superconductor. This contradiction is resolved in Subsection III B within the\nAnderson-Higgs mechanism. It should be noticed that in Eq.(A4) the \feld Eis transverse \feld E=E?=\u00001\nc@A?\n@t\nonly. At the same time, in Eq.(A1) the electric \feld can be longitudinal \u0000r'\u00001\nc@Ajj\n@talso (here div Ajj6= 0). Thus,\nthe \frst London equation (A1) cannot be obtain from the second London equation js=\u0000c\n4\u0019\u00152A?completely.\nAcknowledgments\nThis research was supported by theme grant of Department of physics and astronomy of NAS of Ukraine: \"Math-\nematical models of nonequilibrium processes in open systems\" 0120U100857 and by grant of National Research\nFoundation of Ukraine \"Models of nonequilibrium processes in colloidal systems\" 2020.02/0220.\n[1] A. Larkin, A. Varlamov, Theory of Fluctuations in Superconductors , Oxford U. Press (2005).\n[2] N. Kopnin, Theory of nonequilibrium superconductivity , Oxford Univ. Press, 2001\n[3] L.P. Gor'kov, G.M. Eliashberg, Soviet Phys.-JETP 27, 328 (1968)\n[4] T.J. Rieger, D.J. Scalapino, J.E. Mercereau, Phys. Rev. Lett. 27, 1787 (1971),\nhttps://doi.org/10.1103/PhysRevLett.27.1787\n[5] R.J. Watts-Tobin, Y. Kr ahenb uhl, L. Kramer, J. Low Temp. Phys. 42, 459 (1981), https://doi.org/10.1007/BF00117427\n[6] M. Tinkham, Introduction to superconductivity , McGRAW-HILL Book Company (1975).\n[7] S.N. Artemenko, A.F. Volkov , Physics-Uspekhi 22, 295 (1979), http://dx.doi.org/10.1070/PU1979v022n05ABEH005495\n[8] T.S. Alstr\u001cm, M.P. S\u001crensen, N.F. Pedersen, S. Madsen, Acta Appl. Math. 115, 63 (2011), https://doi.org/10.1007/s10440-\n010-9580-831\n[9] T.S. Larsen, Type II superconductivity studied by the Ginzburg{Landau equation , Master Thesis Department of Mathematics\nTechnical University of Denmark (2005).\n[10] K. Binder, Phys. Rev. B 8, 3423 (1973), https://doi.org/10.1103/PhysRevB.8.3423\n[11] A.F. Volkov, Sh.M. Kogan, Sov. Phys. JETP 38, 1018 (1974).\n[12] C.M. Varma, J. Low Temp. Phys., 126, 901 (2002), https://doi.org/10.1023/A:1013890507658\n[13] V.L. Vadimov, I.M. Khaymovich, A.S. Mel'nikov, Phys. Rev. B 100, 104515 (2019),\nhttps://doi.org/10.1103/PhysRevB.100.104515\n[14] L.D. Landau, E.M. Lifshitz, Electrodynamics of Continuous Media , Pergamon Press LTD. (1960).\n[15] D.R. Tilley, J. Tilley, Super\ruidity and Superconductivity , IOP Publishing Ltd. (1990).\n[16] I.J.R. Aitchison, P. Ao, D.J. Thouless, X.-M. Zhu, Phys. Rev. B 51, 6531 (1995),\nhttps://doi.org/10.1103/PhysRevB.51.6531\n[17] J. Govaerts, D. Bertrand, G. Stenuit, Supercond. Sci. Technol. 14, 463 (2001), https://doi.org/10.1088/0953-\n2048/14/7/308\n[18] D. Bertrand, J. Govaerts, G. Stenuit, Phys. C: Supercond. 369, 339 (2002), https://doi.org/10.1016/S0921-4534(01)01272-\n2\n[19] D. Bertrand, A Relativistic BCS Theory of Superconductivity. An Experimentally Motivated Study of Electric Fields in\nSuperconductors , Universit\u0013 e catholique de Louvain (2005).\n[20] J.E. Hirsch, Phys. Rev. B 69, 214515 (2004), https://doi.org/10.1103/PhysRevB.69.214515\n[21] T. Koyama, Phys. Rev. B 70, 226503 (2004), https://doi.org/10.1103/PhysRevB.70.226503\n[22] J.E. Hirsch, Phys. Rev. B 70, 226504 (2004), https://doi.org/10.1103/PhysRevB.70.226504\n[23] J.E. Hirsch, Phys. C: Supercond. 508, 21 (2015), http://dx.doi.org/10.1016/j.physc.2014.10.018\n[24] A. Peronio, F.J. Giessibl, Phys. Rev. B 94, 094503 (2016), https://doi.org/10.1103/PhysRevB.94.094503\n[25] P.I. Arseev, S.O. Loiko, N.K. Fedorov, Physics-Uspekhi 49, 1 (2006), http://dx.doi.org/10.1070/PU2006v049n01ABEH002577\n[26] M.V. Sadovskii, Quantum Field Theory , de Gruyter Studies in Mathematical Physics, 2013.\n[27] M.V. Sadovskii, Statistical Physics , de Gruyter Studies in Mathematical Physics, 2012.\n[28] T. Cea, C. Castellani, L. Benfatto, Phys. Rev. B 93, 180507(R) (2016), https://doi.org/10.1103/PhysRevB.93.180507\n[29] R. Shimano, N. Tsuji, Ann. Rev. of Cond. Matt. Phys. 11, 103 (2019), https://doi.org/10.1146/annurev-conmatphys-\n031119-050813\n[30] M.V. Sadovskii, Diagrammatics: Lectures on Selected Problems in Condensed Matter Theory , World Scienti\fc, 2006.\n[31] L.S. Levitov, A.V. Shitov, Green's Functions. Problems and Solutions (in Russian), Fizmatlit, Moscow, 2003.\n[32] V.V. Schmidt, The Physics of Superconductors. Introduction to Fundamentals and Applications , Springer, 1997\n[33] E.M. Lifshitz, L.P. Pitaevskii, Statistical Physics. Part 2: theory of condensed state , Pergamon Press LTD. (1981).\n[34] A.J. Dahm, A. Denenstein, T.F. Finnegan, D.N. Langenberg, D.J. Scalapino, Phys. Rev. Lett. 20, 859 (1968),\nhttps://doi.org/10.1103/PhysRevLett.20.859\n[35] I.N. Askerzade, Phys. C: Supercond. 397, 99 (2003), https://doi.org/10.1016/j.physc.2003.07.003\n[36] I.N. Askerzade, Physics-Uspekhi 49, 1003 (2006), https://doi.org/10.1070/PU2006v049n10ABEH006055\n[37] H. Doh, M. Sigrist, B.K. Cho, S.I. Lee, Phys. Rev. Lett. 83, 5350 (1999), https://doi.org/10.1103/PhysRevLett.83.5350\n[38] K.V. Grigorishin, Phys. Lett A. 380, 1781 (2016), https://doi.org/10.1016/j.physleta.2016.03.023\n[39] A.J. Leggett, Prog. Theor. Phys. 36, 901 (1966), https://doi.org/10.1143/PTP.36.901\n[40] S.G. Sharapov, V.P. Gusynin, H. Beck, Eur. Phys. J. B 30, 45 (2002), https://doi.org/10.1140/epjb/e2002-00356-9\n[41] T. Yanagisawa, Nov. Supercond. Mater. 1, 95 (2015), https://doi.org/10.1515/nsm-2015-0010\n[42] G. Blumberg, A. Mialitsin, B.S. Dennis, M.V. Klein, N.D. Zhigadlo, J. Karpinski, Phys. Rev. Lett. 99, 227002 (2007),\nhttps://doi.org/10.1103/PhysRevLett.99.227002\n[43] D.K. Finnemore, T.F. Stromberg, C.A. Swenson, Phys. Rev. 149, 231 (1966), https://doi.org/10.1103/PhysRev.149.231\n[44] S.T. Sekula, R. H. Kernohan, G. R. Love, Phys. Rev. 155, 364 (1967), https://doi.org/10.1103/PhysRev.155.364\n[45] S.T. Sekula, R.H. Kernohan, Phys. Rev. B 5, 904 (1972), https://doi.org/10.1103/PhysRevB.5.904\n[46] M.S. Svirskiy, The electron theory of matter (in Russian), Prosveshcheniye, Moscow, 1980." }, { "title": "2111.10599v1.Canonical_Coordinates_and_Natural_Equation_for_Lorentz_Surfaces_in___mathbb_R_3_1_.pdf", "content": "CANONICAL COORDINATES AND NATURAL EQUATION FOR\nLORENTZ SURFACES IN R3\n1\nKRASIMIR KANCHEV, OGNIAN KASSABOV, AND VELICHKA MILOUSHEVA\nAbstract. We consider Lorentz surfaces in R3\n1satisfying the condition H2\u0000K6= 0, where\nKandHare the Gauss curvature and the mean curvature, respectively, and call them\nLorentz surfaces of general type. For this class of surfaces we introduce special isotropic\ncoordinates, which we call canonical, and show that the coefficient Fof the first fundamental\nform and the mean curvature H, expressed in terms of the canonical coordinates, satisfy a\nspecial integro-differential equation which we call a natural equation of the Lorentz surfaces\nof general type. Using this natural equation we prove a fundamental theorem of Bonnet type\nfor Lorentz surfaces of general type. We consider the special cases of Lorentz surfaces of\nconstant non-zero mean curvature and minimal Lorentz surfaces. Finally, we give examples\nof Lorentz surfaces illustrating the developed theory.\nContents\n1. Introduction 1\n2. Preliminaries 3\n3. Canonical isotropic coordinates of Lorentz surfaces in R3\n1 4\n4. Natural equation of Lorentz surfaces of general type in R3\n1 7\n5. Examples 10\nReferences 12\n1.Introduction\nThe question of describing surfaces with prescribed mean or Gauss curvature in the Eu-\nclidean 3-space R3and also in the other Riemannian space forms have been subject of an\nintensive study. Especially, the geometry of spacelike or timelike surfaces in the Minkowski\n3-space R3\n1has been of wide interest. For example, a Kenmotsu-type representation formula\nfor spacelike surfaces with prescribed mean curvature was obtained by K. Akutagawa and\nS. Nishikawa in [1]. In [7], G´ alvez et al. obtained a representation for spacelike surfaces\ninR3\n1using the Gauss map and the conformal structure given by the second fundamental\nform. M. A. Magid proved that the Gauss map and the mean curvature of a timelike sur-\nface satisfy a system of partial differential equations and found a Weierstrass representation\nformula for timelike surfaces in R3\n1[13]. Timelike surfaces in R3\n1with prescribed Gauss cur-\nvature and Gauss map are studied in [2] where a Kenmotsu-type representation for such\nsurfaces is given. This representation is used to classify the complete timelike surfaces with\npositive constant Gaussian curvature in terms of harmonic diffeomorphisms between simply\nconnected Lorentz surfaces and the universal covering of the de Sitter Space.\n2010Mathematics Subject Classification. Primary 53B30; Secondary 53A35.\nKey words and phrases. Lorentz surfaces, pseudo-Euclidean space, canonical coordinates, natural\nequations.\n1arXiv:2111.10599v1 [math.DG] 20 Nov 20212 KRASIMIR KANCHEV, OGNIAN KASSABOV, AND VELICHKA MILOUSHEVA\nOn the other hand, it is known that the minimal Lorentz surfaces in R3\n1,R4\n1, andR4\n2\ncan be parametrized by special isothermal coordinates, called canonical , such that the main\ninvariants (the Gauss curvature and the normal curvature) of the surface satisfy a system\nof partial differential equations called a system of natural PDEs . The geometry of the\ncorresponding minimal surface is determined by the solution of this system of natural PDEs.\nIn [10], canonical coordinates for the class of minimal Lorentz surfaces in the Minkowski\nspaceR4\n1are introduced and the following system of natural PDEs is obtained:\n(1.1)4q\nK2+{2\u0001hln4q\nK2+{2= 2K;\n4q\nK2+{2\u0001harctan{\nK= 2{;K2+{26= 0;\nwhereKis the Gauss curvature, {is the curvature of the normal connection (the normal\ncurvature), and \u0001his the hyperbolic Laplace operator in R2\n1.\nSimilar results are obtained for minimal Lorentz surfaces in the pseudo-Euclidean space\nwith neutral metric R4\n2in [3] and [9]. The corresponding system of PDEs has the following\nform:\n(1.2)4q\f\fK2\u0000{2\f\f\u0001hln4q\f\fK2\u0000{2\f\f= 2K;\n4q\f\fK2\u0000{2\f\f\u0001hln\f\f\f\fK+{\nK\u0000{\f\f\f\f= 4{;K2\u0000{26= 0:\nThe minimal Lorentz surfaces in R3\n1can also be considered as surfaces in R4\n1orR4\n2, in\nwhich cases {= 0. So, systems (1.1) and (1.2) are reduced to one PDE:\n(1.3)p\njKj\u0001hlnp\njKj= 2K;K6= 0;\nwhich is the natural equation of minimal Lorentz surfaces in R3\n1. Of course, the results in\nthis case can be directly obtained. In [8], canonical coordinates are introduced for minimal\nLorentz surfaces in R3\n1and equations equivalent to (1.3) are derived.\nThus the following natural question arises: How to generalize the concepts of canonical\ncoordinates and natural equation for a wider class of Lorentz surfaces in R3\n1than that of the\nminimal ones? The class of Weingarten Lorentz surfaces in R3\n1with different real principal\ncurvatures (which is equivalent to H2\u0000K > 0, whereKandHare the Gauss curvature\nand the mean curvature, respectively) is considered in [14]. Canonical principal coordinates\nare introduced for this class of surfaces and a natural non-linear partial differential equation\nis derived, which is equivalent to (1.3) in the case of a minimal surface.\nIn the present paper, we propose an alternative approach. We consider Lorentz surfaces\ninR3\n1satisfyingH2\u0000K6= 0and call them surfaces of general type . We introduce special\nisotropiccoordinates(whichwecall canonical )forthesesurfacesandobtaina natural integro-\ndifferential equation . The natural equation for the class of minimal Lorentz surfaces is given\nbyp\njKj\u0000\nlnp\njKj\u0001\nuv=K;K6= 0:\nIt can be reduced to (1.3) by changing the isotropic coordinates with isothermal ones. This\nshows, that the newly obtained results for an arbitrary Lorentz surface of general type in R3\n1\ngeneralize the known results for the case of a minimal Lorentz surface.\nIn Section 2, we give some basic formulas for Lorentz surfaces in R3\n1parametrized by\narbitrary isotropic coordinates. We present the Gauss and Codazzi equations in terms of\nthese coordinates and formulate the fundamental theorem of Bonnet type.CANONICAL COORDINATES AND NATURAL EQUATION FOR LORENTZ SURFACES IN R3\n1 3\nIn Section 3, we introduce the notion of canonical isotropic coordinates for the class of\nLorentz surfaces of general type ( H2\u0000K6= 0) inR3\n1. We prove existence and uniqueness\ntheorems for these coordinates and give the relation between the canonical coordinates and\nthe natural parameters of the isotropic curves of the surface.\nIn Section 4, we consider Lorentz surfaces of general type parametrized by canonical\ncoordinates and show that the coefficient Fof the first fundamental form and the mean\ncurvatureHof such surface satisfy the following integro-differential equation\nFF uv\u0000FuFv\nF=\u0010\n\"1+Rv\nv0F(u;s)Hu(u;s)ds\u0011\u0010\n\"2+Ru\nu0F(s;v)Hv(s;v)ds\u0011\n\u0000F2H2;\n\"1=\u00061;\"2=\u00061. We call it the natural equation of the Lorentz surfaces in R3\n1and prove\na fundamental theorem of Bonnet type. We consider in detail the special cases of a Lorentz\nsurface with non-zero constant mean curvature and a minimal Lorentz surface.\nIn Section 5, we give examples of different types of Lorentz surfaces and their canonical\ncoordinates in R3\n1.\n2.Preliminaries\nLetR3\n1be the standard three-dimensional pseudo-Euclidean space in which the indefinite\ninner scalar product is given by the formula:\nha;bi=\u0000a1b1+a2b2+a3b3:\nLetM= (D;x)be a Lorentz surface in R3\n1, whereD\u001aR2andx :D!R3\n1is an immersion.\nThe coefficients of the first fundamental form of Mare denoted as usually by E;F;Gand\nL;M;N denote the coefficients of the second fundamental form. Then, the Gauss curvature\nKand the mean curvature HofMare given by the formulas (see [4, 12]):\nK=LN\u0000M2\nEG\u0000F2;H=EN\u00002FM +GL\n2(EG\u0000F2):\nIn a neighbourhood of each point of Mthere exist isotropic coordinates (u;v)such that\nE=G= 0[4]. Such parameters are also called null coordinates [11]. It can easily be seen\nthat if (u;v)and(~u;~v)are two different pairs of isotropic coordinates in a neighbourhood of\na fixed point, then they are related either by u=u(~u),v=v(~v), oru=u(~v),v=v(~u).\nFurther, we consider a surface Mparametrized by isotropic coordinates and without loss\nof generality we assume that F > 0. Then, the formulas for KandHtake the following\nform:\n(2.1) K=M2\u0000LN\nF2;H=M\nF:\nConsider the tangent vector fields X = x u;Y = x vand denote by lthe unit normal vector\nfieldl=xu\u0002xv\njxu\u0002xvjsuch thatfX;Y;lgbe a positively oriented frame field in R3\n1. SinceMis\nparametrized by isotropic coordinates, we have:\nX2= Y2= 0;l2= 1;hX;Yi=F;hX;li=hY;li= 0:\nHence, we get\nXv= Y u;hXu;Xi=hXv;Xi=hYu;Yi=hYv;Yi=hlu;li=hlv;li= 0:4 KRASIMIR KANCHEV, OGNIAN KASSABOV, AND VELICHKA MILOUSHEVA\nUsing the last equalities we obtain the following Frenet-type formulas for the frame field\nfX;Y;lg:\n(2.2)\f\f\f\f\f\f\f\f\f\fXu=Fu\nFX + Ll;\nYu= Ml;\nlu=\u0000M\nFX\u0000L\nFY;\f\f\f\f\f\f\f\f\f\fXv= Ml;\nYv=Fv\nFY +Nl;\nlv=\u0000N\nFX\u0000M\nFY:\nThe integrability conditions of (2.2), considered as a system of PDE for the triple (X;Y;l),\nimply the following Gauss equation\n(2.3)FF uv\u0000FuFv\nF=LN\u0000M2\nand the Codazzi equations\n(2.4) Lv=F\u0012M\nF\u0013\nu;Nu=F\u0012M\nF\u0013\nv:\nNote that (2.3) and (2.1) imply K=\u00001\nF(lnF)uv, which is the Gauss’s Theorema Egregium\nin the case of isotropic coordinates.\nAs it is well known, the Gauss and Codazzi equations are not only necessary, but also\nsufficient conditions for the existence of a solution to the PDE system (2.2). This gives us\na fundamental Bonnet-type theorem for Lorentz surfaces in R3\n1. The proof is analogous to\nthat of the classical theorem for surface in R3(see ([5]).\nTheorem 2.1. LetMbe a Lorentz surface in R3\n1parametrized by isotropic coordinates.\nThen, the coefficients F,L,M,Nof the first and the second fundamental form of Mgive\na solution to the Gauss and Codazzi equations (2.3)and(2.4). If ^Mis obtained fromMby\na proper motion in R3\n1, then ^Mgenerates the same solution to (2.3)and(2.4).\nConversely, if the functions F,L,M,Nsatisfy equations (2.3)and(2.4), then, at least\nlocally, there exists a unique (up to a proper motion in R3\n1) Lorentz surface Mparametrized\nby isotropic coordinates, such that the given functions are the coefficients of the first and the\nsecond fundamental form of M.\n3.Canonical isotropic coordinates of Lorentz surfaces in R3\n1\nIn the present section, we will show that the Lorentz surfaces satisfying H2\u0000K6= 0\nadmit special isotropic coordinates which we will call canonical. It follows from the Codazzi\nequations (2.4) and the second equality of (2.1) that\n(3.1) Lv=FH u;Nu=FH v:\nIntegrating the last equalities, we obtain\n(3.2)L=L(u;v 0) +Zv\nv0F(u;s)Hu(u;s)ds;N=N(u0;v) +Zu\nu0F(s;v)Hv(s;v)ds:\nNow, we will try to choose the isotropic coordinates in such a way that L(u;v 0)and\nN(u0;v)to have the simplest form.CANONICAL COORDINATES AND NATURAL EQUATION FOR LORENTZ SURFACES IN R3\n1 5\nFirst, let’s find the transformation formulas for the coefficients of the first and the second\nfundamental form under changes of the isotropic coordinates. Consider the following change\nof the isotropic coordinates (it preserves the numeration):\nu=u(~u);v=v(~v):\nThen, we have ~F=Fu0v0, which implies u0v0>0, since we have assumed at the beginning\nthat ~F > 0andF > 0. In this case, the orientation of the surface does not change, i.e. ~l=l.\nThen, for the coefficients ~F,~L,~M,~Nwe have\n(3.3) ~F=Fu0v0; ~L=Lu02; ~M=Mu0v0; ~N=Nv02:\nIn the case of changing the numeration, it is sufficient to consider only the change of\ncoordinates:\nu= ~v;v= ~u;\nsince the general case is reduced to this and the previous one. In this case, the orientation\nof the surface changes, i.e. ~l=\u0000land we have\n(3.4) ~F=F; ~L=\u0000N; ~M=\u0000M; ~N=\u0000L:\nThe transformation formulas (3.3) and (3.4) show that, if L= 0orN= 0for some\nisotropic coordinates, then ~L= 0or~N= 0for any isotropic coordinates. Further, we\nconsider surfaces satisfying L6= 0andN6= 0at least locally. It follows from (2.1) that\n(3.5) H2\u0000K=LN\nF2;\nwhich implies that the conditions L6= 0andN6= 0are equivalent to H2\u0000K6= 0.\nWe give the following definition.\nDefinition 3.1. A Lorentz surface MinR3\n1is said to be of general type , ifH2\u0000K6= 0.\nThe Lorentz surfaces of general type are naturally divided into two subclasses.\nDefinition 3.2. A Lorentz surface of general type in R3\n1is said to be of first kind (resp.\nsecond kind ), ifH2\u0000K > 0(respH2\u0000K < 0).\nRemark 3.1.SinceKandHare invariants ofM, the property of a surface to be of general\ntype, as well as the kind of the surface, are geometric – they do not depend on the local\nparametrization and are invariant under motions in R3\n1. It is known that the discriminant\nof the characteristic polynomial of the Weingarten map has the form D= 4(H2\u0000K)[12].\nHence, the surfaces of general type are those surfaces for which the Weingarten map has two\ndifferent eigenvalues. Moreover, the surfaces of first kind are those with real eigenvalues, the\nsurfaces of second kind are those with complex eigenvalues.\nNow, we will introduce special isotropic coordinates by the following\nDefinition 3.3. LetM= (D;x)be a Lorentz surface of general type in R3\n1parametrized by\nisotropic coordinates (u;v), such that F > 0, and x0= x(u0;v0)be a point ofM. We call\n(u;v)canonical coordinates with initial point x0, if the coefficients of the second fundamental\nform satisfy the conditions\n(3.6) L(u;v 0) =\"1;N(u0;v) =\"2;\nwhere\"1=\u00061and\"2=\u00061.6 KRASIMIR KANCHEV, OGNIAN KASSABOV, AND VELICHKA MILOUSHEVA\nRemark 3.2.In the general case, condition (3.6) for the coordinates to be canonical depends\non the choice of the initial point (u0;v0). If(u1;v1)is another point in the same neighbour-\nhood, from (3.2) it is obvious that L(u;v 0)6=L(u;v 1)andN(u0;v)6=N(u1;v), in general.\nIn the special case of a surface with constant mean curvature H,Lis a function of uand\nNis a function of v, because of (3.1). Hence, for the class of surfaces with constant mean\ncurvature the canonicity of the coordinates does not depend on the choice of the initial point\n(u0;v0).\nTheorem 3.1. IfM= (D;x)is a Lorentz surface of general type in R3\n1andx0is a fixed\npoint, then we can introduce canonical coordinates with initial point x0.\nProof.Letx0be a fixed point of Mand(u;v)be isotropic coordinates in a neighbourhood\nofx0such thatF > 0andx0= x(u0;v0). Consider the change of the coordinates u=u(~u)\nandv=v(~v). According to (3.3) and Definition 3.3, the new coordinates (~u;~v)are canonical\nif and only if\n(3.7) L(u;v 0)u02=~L(~u;~v0) =\u00061;N(u0;v)v02=~N(~u0;~v) =\u00061:\nThe signs in the right-hand sides of (3.7) are chosen to coincide with the signs of Land\nN, respectively. Thus we obtain ordinary differential equations for u(~u)andv(~v)whose\nsolutions have the following form:\n(3.8) ~u= ~u0+Zu\nu0p\njL(s;v 0)jds; ~v= ~v0+Zv\nv0p\njN(u0;s)jds;\nwhere (~u0;~v0)is arbitrary chosen. It follows from (3.5) that L6= 0andN6= 0, which\nimply ~u0>0and ~v0>0. Hence, equalities (3.8) define new isotropic coordinates (~u;~v)\nsatisfying the condition ~F > 0. Since (3.8) is equivalent to (3.7), then ~L(~u;~v0) =\u00061and\n~N(~u0;~v) =\u00061. So, (~u;~v)are canonical coordinates of Mwith initial point x0. \u0003\nRemark 3.3.IfMis a surface of first kind according to Definition 3.2, then (3.5) implies\nLN > 0. It follows from (3.4) that a change in the coordinates numeration leads to a change\nin the signs of LandN. Hence, surfaces of first kind admit both canonical coordinates for\nwhichL=N= 1and canonical coordinates for which L=N=\u00001.\nIfMis of second kind, then LN < 0, and hence, (3.3) and (3.4) show that the signs of\nLandNdo not change under changes of the isotropic coordinates. Therefore, the surfaces\nof second kind can be divided into two subclasses: surfaces with canonical coordinates such\nthatL= 1andN=\u00001, and surfaces with canonical coordinates such that L=\u00001and\nN= 1.\nNow, we will discuss the question of uniqueness of the canonical coordinates.\nTheorem 3.2. LetMbe a Lorentz surface of general type in R3\n1and(u;v)be canonical\ncoordinates with initial point x0=x(u0;v0)ofM. Then (~u;~v)is another pair of canonical\ncoordinates with the same initial point x0if and only if\n(3.9)u=\u000e~u+c1;\nv=\u000e~v+c2;oru=\u000e~v+c1;\nv=\u000e~u+c2;\nwhere\u000e=\u00061,c1andc2are constants.\nProof.First, we consider the case u=u(~u)andv=v(~v). Equalities (3.3) imply that (~u;~v)\nare canonical coordinates if and only if u02= 1,v02= 1, andu0v0>0. The last conditions\nare equivalent to the first pair of equalities in (3.9)\nThe caseu=u(~v)andv=v(~u)reduces to the previous one by means of (3.4). \u0003CANONICAL COORDINATES AND NATURAL EQUATION FOR LORENTZ SURFACES IN R3\n1 7\nThe meaning of the last theorem is that the canonical coordinates are uniquely determined\nup to a numeration, a sign and an additive constant.\nAt the end of this section we will characterize the canonical coordinates in terms of the\nnull curves lying on the considered surfaces. Recall that, if \u000bis a null curve ( \u000b02= 0) in\nR3\n1, then\u000b002\u00150. The null curves satisfying \u000b002>0are called non-degenerate . It is known\nthat these curves admit a parametrization such that \u000b002= 1[12]. Such parameter is known\nin the literature as a natural parameter orpseudo arc-length parameter , since it plays a role\nsimilar to the role of the arc-length parameter for non-null curves [6].\nTheorem 3.3. LetM= (D;x)be a Lorentz surface in R3\n1parametrized by isotropic coor-\ndinates (u;v), such that F > 0, and x0= x(u0;v0)be a point ofM. Then,Mis of general\ntype if and only if the null curves lying on Mare non-degenerate. The coordinates (u;v)of\nMare canonical with initial point x0if and only if they are natural parameters of the null\ncurves onMpassing through x0.\nProof.The Frenet formulas (2.2) imply x2\nuu= X2\nu=L2andx2\nvv= Y2\nv=N2. Hence, the\nu-lines andv-lines are non-degenerate if and only if L6= 0andN6= 0, which is equivalent\ntoMbeing of general type. Moreover, uis a natural parameter ( x2\nuu= 1) of the null curve\nx(u;v 0)passing through x0if and only if L2(u;v 0) = 1. Analogously, vis a natural parameter\n(x2\nvv= 1) of the null curve x(u0;v)passing through x0if and only if N2(u0;v) = 1. The\nlast conditions are equivalent to L(u;v 0) =\u00061andN(u0;v) =\u00061, which means that the\ncoordinates (u;v)ofMare canonical with initial point x0. \u0003\n4.Natural equation of Lorentz surfaces of general type in R3\n1\nIn this section, we will consider the Gauss and Codazzi equations of a Lorentz surface of\ngeneral typeM= (D;x)inR3\n1parametrized by canonical isotropic coordinates (u;v)with\ninitial point x0= x(u0;v0)2M. In such case, the coefficients of the second fundamental\nform are expressed by the coefficient Fof the first fundamental form, the mean curvature\nH, and the constants \"1and\"2(see Definition 3.3). It follows from (2.1), (3.2), and (3.6)\nthat:\n(4.1)L=\"1+Rv\nv0F(u;s)Hu(u;s)ds;M=FH;N=\"2+Ru\nu0F(s;v)Hv(s;v)ds;\nwhere\"1=\u00061;\"2=\u00061, the signs depending on the kind of the surface. Substituting these\nexpressions in the Gauss equation (2.3), we obtain:\n(4.2)FF uv\u0000FuFv\nF=\u0010\n\"1+Rv\nv0F(u;s)Hu(u;s)ds\u0011\u0010\n\"2+Ru\nu0F(s;v)Hv(s;v)ds\u0011\n\u0000F2H2:\nConsequently, FandHgive a solution to the integro-differential equation (4.2), which we\ncall the natural equation of the Lorentz surfaces of general type in R3\n1. The converse is\nalso true. Namely, the following Bonnet-type theorem holds:\nTheorem 4.1. LetM= (D;x)be a Lorentz surface of general type in R3\n1and(u;v)be\ncanonical isotropic coordinates with initial point x0= x(u0;v0)2M. Then, the coefficient\nFof the first fundamental form and the mean curvature HofMgive a solution to the natural\nequation (4.2). If ^Mis obtained fromMby a proper motion in R3\n1, then ^Mgenerates the\nsame solution to (4.2).\nConversely, let F > 0andHbe functions of (u;v)defined in a neighbourhood of (u0;v0)\nand satisfying the natural equation (4.2), where\"1=\u00061and\"2=\u00061. Then, at least locally,\nthere exists a unique (up to a proper motion in R3\n1) Lorentz surface of general type in R3\n1\ndefined byx=x(u;v)in canonical isotropic coordinates with initial point x0=x(u0;v0),8 KRASIMIR KANCHEV, OGNIAN KASSABOV, AND VELICHKA MILOUSHEVA\nsuch that the given functions FandHare the non-zero coefficient of the first fundamental\nform and the mean curvature, respectively, and the signs of the corresponding coefficients L\nandNof the second fundamental form coincide with the signs of \"1and\"2.\nProof.We have already seen that the coefficient Fand the mean curvature Hof a surface\nwith the given properties satisfy the natural equation. Now we will prove the converse.\nGiven the functions FandHsatisfying (4.2), we define functions L,M, andNby equal-\nities (4.1). Then, (4.2) implies that the quadruple F,L,M,Nsatisfies the Gauss equation\n(2.3). Differentiating (4.1) we get that the Codazzi equations (2.4) are also fulfilled. Apply-\ning Theorem 2.1 we get a Lorentz surface Mparametrized by isotopic coordinates whose\ncoefficient of the first fundamental form is the given function Fand the coefficients of the\nsecond fundamental form are the functions L,M,N, defined by (4.1). Comparing (3.2) with\n(4.1) we obtain L(u;v 0) =\"1,N(u0;v) =\"2, which means that Mis of general type and\n(u;v)are canonical isotropic coordinates. Comparing (2.1) with (4.1) we see that the mean\ncurvature ofMis the given function H. Hence, the surface Mhas the necessary properties.\nMoreover, if ^Mis another surface with the same properties, then (4.1) is also valid for ^M.\nSo,Mand ^Mhave one and the same coefficients of the first and second fundamental form.\nHence, according to Theorem 2.1, ^Mis obtained fromMby a proper motion in R3\n1.\u0003\nNow, we will consider the natural equation (4.2) in the case of a surface with constant\nmean curvature H. In this case, equation (4.2) takes the form:\n(4.3)FF uv\u0000FuFv\nF=\"1\"2\u0000F2H2;\nand (4.1) implies L=\"1,N=\"2. Then, it follows from (3.5) that\n(4.4) H2\u0000K=\"1\"2\nF2;F=1p\njH2\u0000Kj:\nIf we rewrite (4.3) in the form\n(4.5)1\nF(lnF)uv=\"1\"2\nF2\u0000H2\nthen by use of (4.4) we obtain\n(4.6)p\njH2\u0000Kj\u0000\nlnp\njH2\u0000Kj\u0001\nuv=K;H2\u0000K6= 0:\nWe call (4.6) the natural equation of constant mean curvature Lorenz surfaces in R3\n1.\nSo, we can formulate the following Bonnet-type theorem for Lorentz surfaces of constant\nnon-zero mean curvature.\nTheorem 4.2. LetM= (D;x)be a Lorentz surface of general type in R3\n1with constant non-\nzero mean curvature Hparametrized by canonical isotropic coordinates. Then, the Gauss\ncurvatureKsatisfies the natural equation (4.6). If ^Mis obtained from Mby a proper\nmotion in R3\n1, then ^Mgenerates the same solution to (4.6).\nConversely, let Hbe a non-zero constant and Kbe a function of (u;v)satisfying the natural\nequation (4.6). Then, at least locally, there exist (up to a proper motion in R3\n1) exactly two\nLorentz surfaces of general type in R3\n1parametrized by canonical isotropic coordinates, with\nthe constant Has non-zero mean curvature and the function Kas Gauss curvature.\nProof.We have already seen that the mean curvature Hand the Gauss curvature Kof a\nsurface with the given properties satisfy equation (4.6). Now we will prove the converse.\nGiven the constant Hand the function Ksatisfying (4.6), we define a function Fand\nconstants\"1=\u00061,\"2=\u00061such that equalities (4.4) hold true. Then equality (4.6) impliesCANONICAL COORDINATES AND NATURAL EQUATION FOR LORENTZ SURFACES IN R3\n1 9\n(4.5), which is equivalent to (4.3), the latter being the natural equation (4.2) in the case H\nis constant. Note that the function Fis determined uniquely by (4.4), while for the choice\nof\"1and\"2we have two different options depending on the choice of signs. This means that\naccording to Theorem 4.1 we obtain two different Lorentz surfaces M1andM2parametrized\nby canonical isotropic coordinates, whose mean curvature is the given constant Hand the\ncoefficient of the first fundamental form is the given function F. Equalities (4.4) hold true\nfor bothM1andM2, and hence, Kis determined uniquely by FandH. Consequently, the\nGauss curvature of both M1andM2is the given function K.\nThe surfaceM2cannot be obtained from M1by a proper motion in R3\n1, since the proper\nmotions preserve the signs of \"1and\"2, and the signs of these constants are different for M1\nandM2. If ^Misanothersurfacewiththesameproperties, thenequalities(4.4)holdtruealso\nfor^M. Hence,M1,M2, and ^Mhave one and the same coefficients of the first fundamental\nform and equal mean curvatures. Moreover, the constants \"1and\"2for ^Mcoincide with\nthe constants for one of the two surfaces M1orM2. So, according to Theorem 4.1, ^Mcan\nbe obtained from one of the two surfaces M1orM2by a proper motion in R3\n1. \u0003\nRemark 4.1.The two surfaces obtained in the last theorem are really different. We have\nalready seen that M2cannot be obtained from M1by a proper motion in R3\n1. Furthermore,\nM2cannot be obtained from M1by coordinate change of the form u=u(~u)andv=v(~v),\nsince such a change preserves the signs of \"1and\"2, according to (3.3). M2cannot be\nobtained fromM1also by coordinate change of the form u=u(~v)andv=v(~u), since in\nsuch case the sign of Hchanges, according to (2.1) and (3.4), but the surfaces M1andM2\nhave equal mean curvatures. Such a pair of surfaces is presented in Examples 5.4 and 5.5.\nNow we will consider the case of a surface Mwith zero mean curvature H, i.e.Mis a\nminimal surface in R3\n1. In this case, equality (4.6) takes the form:\n(4.7)p\njKj\u0000\nlnp\njKj\u0001\nuv=K;K6= 0:\nLet us point out that the Gauss curvature Kand the canonical coordinates are invariant\nunder non-proper motions in R3\n1. Hence, in this case the surface is determined uniquely by\nthe solution of (4.7) up to an arbitrary motion. We have the following Bonnet-type theorem\nfor minimal Lorentz surfaces in R3\n1.\nTheorem4.3. LetM= (D;x)be a minimal Lorentz surface of general type in R3\n1parametrized\nby canonical isotropic coordinates. Then, the Gauss curvature KofMsatisfies the natural\nequation (4.7). If ^Mis obtained fromMby a motion in R3\n1, then ^Mgenerates the same\nsolution to (4.6).\nConversely, given a function Kof(u;v)satisfying the natural equation (4.7), there exists,\nat least locally, a unique (up to a motion in R3\n1) minimal Lorentz surface of general type\nparametrized by canonical isotropic coordinates, such that its Gauss curvature is the given\nfunctionK.\nProof.TheproofissimilartotheproofofTheorem4.2, thereforewewillnotgiveitindetails,\nwe will only point out the difference. Again, given the function K, we obtain two different\nminimal surfacesM1andM2parametrized by canonical isotropic coordinates, having Kas\nthe Gauss curvature and having different signs of \"1and\"2. Let ~M1be a surface obtained\nfromM1by a non-proper motion in R3\n1. Then, the Gauss curvature Kand the signs of \"1\nand\"2of~M1are the same as those of M2. So, equalities (4.4) imply that ~M1andM2have\none and the same coefficient Fof the first fundamental form. Hence, according to Theorem\n4.1,M2is obtained from ~M1by a proper motion in R3\n1. Consequently,M2is obtained from\nM1by a non-proper motion in R3\n1. \u000310 KRASIMIR KANCHEV, OGNIAN KASSABOV, AND VELICHKA MILOUSHEVA\n5.Examples\nIn this section, we will consider examples of Lorentz surfaces of general type in R3\n1, illus-\ntrating the developed theory. First, we give an example of a minimal Lorentz surface.\nExample 5.1.Let us consider the surface in R3\n1, determined by the following parametrization:\n(5.1) x =1\n6(u3\u0000v3+ 3u\u00003v;\u0000u3+v3+ 3u\u00003v;3u2\u00003v2):\nThe coefficients of the first and the second fundamental form are:\nE=G= 0;F=1\n2(u\u0000v)2;L= 1;M= 0;N= 1:\nThe Gauss curvature and the mean curvature of the surface defined above are expressed as\nfollows:\n(5.2) K=\u00004\n(u\u0000v)4;H= 0:\nHence,thesurfacedefinedby(5.1)isaminimalLorentzsurface(ofEnneper-type)parametrized\nby isotropic coordinates. Moreover, the coordinates (u;v)are canonical, since L=N= 1.\nThe Gauss curvature Kis negative, so the surface is of first kind according to Definition 3.2.\nThe function Kgiven in (5.2) is a solution to the natural equation (4.7).\nExample 5.2.Let us consider the surface in R3\n1, defined by\n(5.3) x =1\n6(u3\u0000v3+ 3u\u00003v;\u0000u3+v3+ 3u\u00003v;3u2+ 3v2):\nThe coefficients of the first and the second fundamental form are:\nE=G= 0;F=1\n2(u+v)2;L= 1;M= 0;N=\u00001:\nThe Gauss curvature and the mean curvature are expressed as follows:\n(5.4) K=4\n(u+v)4;H= 0:\nAs in the previous example, (5.3) defines a minimal Lorentz surface of Enneper-type\nparametrized by canonical isotropic coordinates. In this example, the Gauss curvature Kis\npositive and hence, the surface is of second kind according to Definition 3.2. The function\nKgiven in (5.4) is also a solution to the natural equation (4.7).\nNow, we will give examples of surfaces with non-zero constant mean curvature.\nExample 5.3.We consider the Lorentz sphere in R3\n1parametrized by isothermal coordinates\n(t;s)as follows:\nx = (sinhtsechs;coshtsechs;tanhs):\nChanging the coordinates with isotropic ones, we obtain:\n(5.5) x = (sinh(u\u0000v) sech(u+v);cosh(u\u0000v) sech(u+v);tanh(u+v)):\nThe coefficients of the first and the second fundamental form are:\nE=G= 0;F= 2 sech2(u+v);L= 0;M= 2 sech2(u+v);N= 0:\nThe Gauss curvature and the mean curvature are given by:\nK= 1;H= 1:\nHence, the surface defined by (5.5) is a Lorentz surface with non-zero constant mean\ncurvature parametrized by isotropic coordinates. In this example, H2\u0000K= 0which meansCANONICAL COORDINATES AND NATURAL EQUATION FOR LORENTZ SURFACES IN R3\n1 11\nthat the surface is not of general type within the meaning of Definition 3.1. In this case, we\ncannot introduce canonical coordinates in the sense of Definition 3.3.\nExample 5.4.Let us consider the cylinder in R3\n1, parametrized by isothermal coordinates\n(t;s)as follows:\nx = (t;coss;sins):\nChanging the coordinates with isotropic ones, we obtain:\nx = (u\u0000v;cos(u+v);sin(u+v)):\nThe coefficients of the first and the second fundamental form are:\nE=G= 0;F= 2;L= 1;M= 1;N= 1:\nThe Gauss curvature and the mean curvature are given by:\nK= 0;H=1\n2:\nThisisanexampleofaLorentzsurfacewithnon-zeroconstantmeancurvatureparametrized\nby canonical isotropic coordinates. It corresponds to the trivial (zero) solution to equation\n(4.6).\nExample 5.5.Now, let us consider the hyperbolic Lorentz cylinder in R3\n1, parametrized by\nisothermal coordinates (t;s)as follows:\nx = (sinhs;coshs; t):\nChanging the coordinates with isotropic ones, we obtain:\nx = (sinh(u\u0000v);cosh(u\u0000v); u+v):\nThe coefficients of the first and the second fundamental form are:\nE=G= 0;F= 2;L=\u00001;M= 1;N=\u00001:\nThe Gauss curvature and the mean curvature are given by:\nK= 0;H=1\n2:\nThis is also an example of a Lorentz surface with non-zero constant mean curvature\nparametrized by canonical isotropic coordinates. It also corresponds to the trivial (zero)\nsolution to equation (4.6).\nComparing the results of the last two examples, we see that the two cylinders have equal\nconstant mean curvatures and equal Gauss curvatures. Hence, they give one and the same\nsolution to the natural equation (4.6). But there is a difference in the signs of \"1=Land\n\"2=N. These two cylinders form a pair of surfaces M1andM2as the ones described in\nthe proof of Theorem 4.2.\nFinally, we will consider a surface with non-constant mean curvature.\nExample 5.6.Let us consider the hyperbolic Lorentz cone MinR3\n1, parametrized by isother-\nmal coordinates (t;s)as follows:\nx =\u0000\net\n2sinhs;p\n3 et\n2;et\n2coshs\u0001\n:\nChanging the coordinates with isotropic ones, we obtain:\nx =\u0000\neu+v\n2sinh(u\u0000v);p\n3 eu+v\n2;eu+v\n2cosh(u\u0000v)\u0001\n:12 KRASIMIR KANCHEV, OGNIAN KASSABOV, AND VELICHKA MILOUSHEVA\nThe coefficients of the first and the second fundamental form are:\nE=G= 0;F= 2eu+v;L=p\n3\n2eu+v\n2;M=\u0000p\n3\n2eu+v\n2;N=p\n3\n2eu+v\n2:\nThe Gauss curvature and the mean curvature are given by:\nK= 0;H=\u0000p\n3\n4e\u0000u+v\n2:\nHence,inthisexample, MisaLorentzsurfacewithnon-constantmeancurvatureparametrized\nby isotropic coordinates. The coordinates (u;v)are not canonical, since L(u;v 0)6=\u00061and\nN(u0;v)6=\u00061.\nWe will introduce canonical isotropic coordinates (~u;~v)with initial point (u0;v0) = (0;0).\nUsing formulas (3.8) we get:\n~u= ~u0+ 2p\n24p\n3(eu\n4\u00001); ~v= ~v0+ 2p\n24p\n3(ev\n4\u00001):\nTo simplify the formulas we choose ~u0= ~v0= 2p\n24p\n3. Then,\n~u= 2p\n24p\n3 eu\n4; ~v= 2p\n24p\n3 ev\n4;u= 4 ln~u\n2p\n24p\n3;v= 4 ln~v\n2p\n24p\n3:\nUsing the last equalities and formulas (3.3), we express the coefficient of the first funda-\nmental form ~Fand the mean curvature ~Hin terms of the canonical coordinates (~u;~v)as\nfollows:\n(5.6) ~F=~u3~v3\n1152; ~H=\u000048p\n3\n~u2~v2:\nAccording to Theorem 4.1, the functions ~Fand ~Hgiven by (5.6) give a solution to the\nnatural equation (4.2) in the case \"1=\"2= 1.\nAcknowledgments: The third author is partially supported by the National Science\nFund, Ministry of Education and Science of Bulgaria under contract DN 12/2.\nReferences\n[1] K. Akutagawa and S. Nishikawa. The Gauss map and spacelike surfaces with prescribed\nmean curvature in Minkowski 3-space. Tohoku Math. J. (2) , 42(1):67–82, 1990. doi:\n10.2748/tmj/1178227694.\n[2] J. A. Aledo, J. M. Espinar, and J. A. G´ alvez. Timelike surfaces in the\nLorentz–Minkowski space with prescribed Gaussian curvature and Gauss map. Journal\nof Geometry and Physics , 56(8):1357–1369, 2006. doi: 10.1016/j.geomphys.2005.07.004.\n[3] Y. Aleksieva and V. Milousheva. Minimal Lorentz surfaces in pseudo-Euclidean 4-space\nwith neutral metric. Journal of Geometry and Physics , 142:240–253, 2019. doi:\n10.1016/j.geomphys.2019.04.008.\n[4] H. Anciaux. Minimal Submanifolds in Pseudo-Riemannian Geometry . World Scientific\nPublishing Co. Pte. Ltd., 2011. ISBN 978-981-4291-24-8.\n[5] M. P. do Carmo. Differential Geometry of Curves and Surfaces . Dover Publications,\nInc. Mineola, New York, 2 edition, 2016. ISBN 978-0-486-80699-0.\n[6] K. Duggal and D. H. Jin. Null curves and hypersurfaces of semi-Riemannian manifolds .\nWorld Scientific, 2007. ISBN 978-981-3106-97-0.\n[7] J. A. G´ alvez, A. Mart´ ınez, and A. Mil´ an. Complete constant Gaussian curvature sur-\nfaces in the Minkowski space and harmonic diffeomorphisms onto the hyperbolic plane.\nTohoku Math. J. (2) , 55(4):467–476, 2003. doi: 10.2748/tmj/1113247124.CANONICAL COORDINATES AND NATURAL EQUATION FOR LORENTZ SURFACES IN R3\n1 13\n[8] G. Ganchev. Canonical Weierstrass representation of minimal and maximal surfaces\nin the three-dimensional Minkowski space, 2008. URL https://arxiv.org/abs/0802.\n2632.\n[9] G. Ganchev and K. Kanchev. Canonical coordinates and natural equations for mini-\nmal time-like surfaces in R4\n2.Kodai Mathematical Journal , 43(3):524–572, 2020. doi:\n10.2996/kmj/1605063628.\n[10] G. Ganchev and V. Milousheva. Timelike surfaces with zero mean curvature in\nMinkowski 4-space. Isr. J. Math. , 196:413–433, 2013. doi: 10.1007/s11856-012-0169-y.\n[11] J. Inoguchi. Timelike surfaces of constant mean curvature in Minkowski 3-space. Tokyo\nJ. Math., 21(1):141–152, 1998. doi: 10.3836/tjm/1270041992.\n[12] R. L´ opez. Differential geometry of curves and surfaces in Lorentz-Minkowski space. Int.\nElectron. J. Geom. , 7(1):44–107, 2014. ISSN 1307-5624. doi: 10.36890/iejg.594497.\n[13] M. A. Magid. Timelike surfaces in Lorentz 3-space with prescribed mean curvature and\nGauss map. Hokkaido Math. J. , 20(3):447–464, 1991. doi: 10.14492/hokmj/1381413979.\n[14] V. Mihova and G. Ganchev. Partial differential equations of time-like Wein-\ngarten surfaces in the three-dimensional Minkowski space. Ann. Sofia Univ., Fac-\nulty Math. and Inf. , 101:143–165, 2013. URL https://www.fmi.uni-sofia.bg/en/\nannuaire-by-tome/101 .\nDepartment of Mathematics and Informatics, Todor Kableshkov University of Trans-\nport, 158 Geo Milev Str., 1574, Sofia, Bulgaria\nEmail address :kbkanchev@yahoo.com\nInstitute of Mathematics and Informatics, Bulgarian Academy of Sciences, Acad. G.\nBonchev Str. bl. 8, 1113, Sofia, Bulgaria\nEmail address :okassabov@math.bas.bg\nInstitute of Mathematics and Informatics, Bulgarian Academy of Sciences, Acad. G.\nBonchev Str. bl. 8, 1113, Sofia, Bulgaria\nEmail address :vmil@math.bas.bg" }, { "title": "1303.2057v1.Polariton_excitation_in_epsilon_near_zero_slabs__transient_trapping_of_slow_light.pdf", "content": "arXiv:1303.2057v1 [physics.optics] 8 Mar 2013Polariton excitation in epsilon-near-zero slabs: transie nt trapping of slow light\nAlessandro Ciattoni1, Andrea Marini2, Carlo Rizza1,3, Michael Scalora4and Fabio Biancalana2,5\n1Consiglio Nazionale delle Ricerche, CNR-SPIN 67100 L’Aqui la,\nItaly and Dipartimento di Fisica, Universit` a dell’Aquila , 67100 L’Aquila, Italy\n2Max Planck Institute for the Science of Light, Guenther-Sch arowsky-Straße 1, 91058 Erlangen, Germany\n3Dipartimento di Fisica e Matematica, Universit` a dell’Ins ubria, Via Valleggio 11, 22100 Como, Italy\n4Charles M. Bowden Research Center RDMR-WDS-WO,\nRDECOM, Redstone Arsenal, Alabama 35898-5000, USA and\n5School of Engineering & Physical Sciences, Heriot-Watt Uni versity, Edinburgh, EH14 4AS, United Kingdom\n(Dated: October 1, 2018)\nWe numerically investigate the propagation of a spatially l ocalized and quasi-monochromatic\nelectromagnetic pulsethroughaslab withLorentzdielectr ic response intheepsilon-near-zeroregime,\nwheretherealpartofthepermittivityvanishesatthepulse carrier frequency. Weshowthatthepulse\nis able to excite a set of virtual polariton modes supported b y the slab, the excitation undergoing\na generally slow damping due to absorption and radiation lea kage. Our numerical and analytical\napproaches indicate that in its transient dynamics the elec tromagnetic field displays the very same\nenhancement of the field component perpendicular to the slab , as in the monochromatic regime.\nThe transient trapping is inherently accompanied by a signi ficantly reduced group velocity ensuing\nfrom the small dielectric permittivity, thus providing a no vel platform for achieving control and\nmanipulation of slow light.\nI. INTRODUCTION\nPhysical mechanisms driving to slow and fast light\nhave attracted considerable attention from the scien-\ntific community in the last decades [1–3]. The inher-\nent interest in slow light comes from the long matter-\nradiation interaction time, which can lead to consider-\nable enhancement of all nonlinear processes that in turn\nmay be exploited for active functionalities [4], e.g. all-\noptical switching and modulation [5, 6]. The nonlin-\nearity may also be enhanced by reducing the effective\narea in subwavelength silicon on insulator and plasmonic\nwaveguides [7, 8], where tight confinement opens up pos-\nsibilities for miniaturized nonlinear applications [9–11].\nAlternatively, extreme nonlinear dynamics [12, 13], en-\nhanced second and third harmonic generation [14, 15]\nis predicted in epsilon-near-zero (ENZ) metamaterials,\nwhere the linear susceptibility is tailored in such a way\nthat its modulus becomes comparable with the nonlinear\ncounterpart. BoostingthenonlinearityofENZplasmonic\nchannels can also lead to active control of tunneling [16],\nswitching and bistable response [17]. ENZ metamateri-\nals have also been used for directive emission [18, 19],\ncloaking [20], energy squeezing in narrow channels [21]\nand subwavelength imaging [22, 23].\nInalloftheabovementionedmechanisms,thetrade-off\nneeded to achieve enhanced active functionalities is paid\nin terms of increased losses. As a result, the ENZ regime\none usually invokes refers to the case where the real part\nof the susceptibility becomes very small, while its imag-\ninary part remains finite. Indeed, due to the stringent\nphysical requirement of causality, Kramers-Kronig rela-\ntions impose that dispersion be inherently accompanied\nby loss and the dielectric susceptibility can not become\nrigorously null. The residual loss either limits or even\npreventsgiant enhancement ofcoherent mechanisms, e.g.in second and third harmonic generation setups [14, 15].\nRecently, in the context of surface plasmon polaritons,\na method has been proposed to overcome the loss bar-\nrier for superlensing applications by loading the effect\nof loss into the time domain [24]. In our analytical cal-\nculations, we will use a similar approach to study the\nbehavior of an electromagnetic pulse that scatters from\na slab having a Lorentz dielectric response. Indeed, by\nconsidering non-monochromatic virtual modes with com-\nplex frequency [25], it is possible to drop off the effect of\nloss on the temporal dependence of the “mode” itself. In\nthis complex frequency approach, it is possible to achieve\nthe condition where the dielectric susceptibility exactly\nvanishes. Our formalism treats the dielectric polariza-\ntion of the medium as a generic Lorentz oscillator that,\nin the epsilon-equal-to-zero condition, encompasses lon-\ngitudinal collective oscillations of both electrons (volume\nplasmons) and ions (longitudinal phonons) that can not\nbe excited by light [26]. Recently, the question whether\nor not volume plasmons can be excited ornot by classical\nlight has been revived [27–31]. Some studies on Mie ex-\ntinctionefficienciesrevealamaximumaroundthecharac-\nteristic frequency where the dielectric susceptibility van-\nishes, attributing the enhanced extinction to the excita-\ntion of volume plasmons [28, 30]. Conversely, other sim-\nilar studies identify the physical origin of the enhanced\nextinction in the excitation of leaky modes [27, 29]. The\nlatter interpretation is also supported by studies of the\nexcitation ofsurface phonon polaritons in ENZ slabs [32–\n35].\nIn this manuscript we numerically investigate and ana-\nlyticallyinterpretthe scatteringofaspatiallyandtempo-\nrally localized optical pulse from a dielectric slab in the\nENZ regime. We used a finite difference time domain\n(FDTD) algorithm to solve the full vectorial Maxwell\nequations coupled to the Lorentz oscillator equation for2\nthe dielectric polarizationofthe slab. We find that, if the\ncarrier frequency of the optical pulse matches the ENZ\ncondition, electromagnetic quasi-trapping occurs within\nthe Lorentz slab since, after the pulse has passed through\nit, an elecromagnetic-polarization (polariton) oscillation\npersists and generally slowly damps out. We demon-\nstrate that non-trivial ENZ features like the enhance-\nment of the longitudinal electric field component are still\nobservable in the time-domain. We also find that the\nabove mentioned phenomenology is not observed for op-\ntical pulses with carrier frequencies far from the ENZ\ncondition. Thus, in order to grasp the underpinning\nphysical mechanisms responsible for transient trapping\nin the ENZ regime, we analytically investigate the scat-\nteringfeaturesofthe Lorentzslabbystudying the virtual\nleaky modes of the structure. We recognize that a set of\npolariton modes with reduced transverse group velocity\n(vg≃c/100, where cis the speed of light in vacuum)\nis excited. Indeed, the plasma frequency plays the role\nof a cut-off frequency and polaritons in the ENZ regime\nare intrinsically characterized by a reduced group veloc-\nity. Thus, we are able to interpret the transient trapping\nby means of the excitation of slow polariton modes that\nare damped off due to medium absorption and radiation\nleakage in the outer medium.\nThe paper is organized as follows. In Sec.II we re-\nport the results of numerical finite difference time do-\nmain (FDTD) simulations, comparing the distinct phe-\nnomenologies occuring in ENZ and standard dielectric\nregimes. In Sec.III we analytical investigate the virtual\nleaky modes of the structure, we address their properties\nand we discuss their role in the interpretation of numeri-\ncal results are developed in section III. In Sec.IVwe draw\nour conclusions.\nII. FINITE DIFFERENCE TIME DOMAIN\nANALYSIS OF THE TIME-DOMAIN ENZ\nREGIME\nA. Pulse scattering by a Lorentz slab\nLet us consider the scattering interaction sketched in\nFig.1(a), where an electromagnetic pulse is launched\nalong the z-axis in vacuum and orthogonally impinges\non the surface of a dielectric slab. The pulse is a Trans-\nverse Magnetic (TM) excitation, with electric Ex(x,z,t),\nEz(x,z,t) and magnetic Hy(x,z,t) field components.\nThe initial profile profile of the transverse electric com-\nponent is\nEx(x,z,t= 0) =E0e−x2\nw2xe−(z−z0)2\nw2zsin/bracketleftBig¯ω\nc(z−z0)/bracketrightBig\n.(1)\nThis field is spatially confined both along the x- and\nz- axis,wxandwzbeing its transverse and longitu-\ndinal widths, respectively, it is centered at the point\n(x,z) = (0,z0) and it is longitudinally modulated with\nperiod¯λ= 2πc/¯ω. Hereafter we will focus on very\nFIG. 1: (a) Interaction geometry of the pulse colliding onto\nthe dielectric slab. (b) Real and imaginary part of the slab\ndielectric permittivity for the Lorentz parameters used in the\nFDTD analysis as a function of the wavelength.\nlong pulses such that wz≫c/(¯ω). Thus, the quasi-\nmonochromatic condition δω/¯ω≪1 is satisfied, where\n¯ωis the pulse central frequency and δω≃c/wzis the\nspectral width.\nThe dielectric slab has width L, it is centered at\n(x,z) = (0,0) and we assume that, in the presence of\nthe external electric field E, the dynamics of its dielectric\npolarization is governed by the Lorentz oscillator model\nd2P\ndt2+γdP\ndt+ω2\neP=ǫ0feE, (2)\nwhereωeis the resonant angular frequency, γis the\ndamping constant, feis the oscillator strength and ǫ0\nis the dielectric permittivity of vacuum. It is well\nknown that Eq.(2) leads, in the frequency domain, to\nthe constitutive relation ˜Dω=ǫ0ǫ(ω)˜Eωwhere˜fω=/integraltext+∞\n−∞dteiωtf(t) is the Fourier transform of f(t),D=\nǫ0E+Pis the displacement field vector and ǫ= 1 +\nfe/(ω2\ne−iγω−ω2) is the frequency dependent medium\ndielectric permittivity. The realistic model of Eq.(2) is\nparticularlyaccuratefordescribingthemediumdielectric\nresponse to fields with frequencies close to the resonant\nfrequency ωe(so that contributions due to other reso-\nnances can be neglected). Therefore, the Lorentz model\nis particularly suitable for our analysis since we are here\nconcerned with quasi-monochromatic pulses whose car-\nrier frequency ¯ ωcoincides with (or is close to) the fre-\nquency\nω0=/radicalbigg\n1\n2/bracketleftBig\n(2ω2e−γ2+fe)+/radicalbig\nf2e+γ2(γ2−4ω2e−2fe)/bracketrightBig\n(3)\nwhere the real part of the permittivity vanishes, i.e.\nRe[ǫ(ω0)] = 0, the so called epsilon-near-zero (ENZ)\nregime.\nWehaveperformedthe numericalanalysisofthepulse-\nslab collision by means of a Finite Difference Time Do-\nmain (FDTD) scheme where the polarization dynamics\nof Eq.(2) are coupled to Maxwell equations for the TM3\nFIG. 2: Results of FDTD simulation pertaining the interacti on of the pulse 0 (with ¯ ω=ω0) with the slab. The fields Ex,Ez\nand the Fourier transform F[Ex] are captured at three different time steps corresponding to the three rows. In the first and\nthe second time steps the pulse peak is just behind and beyond the slab, respectively, whereas in the third time step the pu lse\nhas completely left the slab. Note the large Ezcomponent within the slab (signature of the epsilon near zer o regime) and the\npersisting and pulse-free polariton oscillation in the thi rd time step.\nfield. Specifically, in order to isolate the relevant phe-\nnomenology characterizing the ENZ regime we have ana-\nlyzed through FDTD simulations two different situations\nwherethesamedielectricslabishitbytwospatiallyequal\npulseswith differentcarrierfrequencies ¯ ω: the first(pulse\n0) is such that ¯ ω=ω0so that it is suitable to scan the\nslab behavior in its ENZ regime; the second (pulse 1) has\n¯ω=ω1for which Re[ ǫ(ω1)]>0 so that it experiences\nstandard dielectric behavior.\nIn view of the generality and ubiquity of the Lorentz\nmodel of Eq.(2), we have chosen for our numerical sim-\nulations a realistic medium with Lorentz parameters\nωe= 3.75·1015Hz,γ= 1.50·1012sec−1andfe=\n2.25·1030sec−2in order to deal with optical pulses in the\nvisible spectrum, the resonant frequency corresponding\nto the wavelength λe= 2πc/ωe= 0.502µm. For such\nparameters Eq.(3) yields ω0= 4.03·1015Hzand we have\nsetω1= 4.23·1015Hz(for which Re[ ǫ(ω1)] = 0.42),\nthe two frequencies corresponding to the wavelengths\nλ0= 2πc/ω0= 0.466µmandλ1= 2πc/ω1= 0.445µm\nrespectively. In Fig.1(b) we plot the real and imaginary\nparts of the dielectric permittivity for the chosen Lorentz\nparameters as functions of the wavelength λ= 2πc/ω,\nindicating the carrier wavelengths λ0andλ1that char-\nacterize the two pulses. We have chosen a slab width\nL= 0.41µmto minimize the pulse propagation features\nand to effectively highlight the impact of the medium\npolarization on field dynamics. We set z0=−150µm,so that the pulse peak reaches the slab 500 fsafter it\nhas been launched. Pulse widths are chosen so that\nwx= 1µmandwz= 60µm: the former is compa-\nrable with the central wavelength of the pulse in order to\nprovide the optical beam with a non negligible longitu-\ndinal field component Ez(see below) whereas the latter\ncorresponds to a temporal width T=wz/c= 200fs\nand a spectral width δω≃1/T= 5·1012Hzso that the\npulses are in the quasi-monochromatic regime.\nB. Pulse 0scattering\nIn Fig.2 we report the main results of the FDTD sim-\nulation dealing with the interaction of the pulse 0 with\ncarrier frequency ¯ ω=ω0with the Lorentz slab repre-\nsented in the figure by the semitransparent rectangu-\nlar blocks. The first two columns of the figure con-\ntain the plots of Ex(x,z,t),Ez(x,z,t) as functions of\n(x,z) whereas the third contains the Fourier transform\nF[Ex](kx,z,t) =/integraltext+∞\n−∞dxeikxxEx(x,z,t) as a function of\n(kx,z); each row of the figure corresponds to a selected\nsimulation time step. At the first time step (first row of\nFig.(2)), t= 371fs, the pulse is fully interacting with\nthe slab (the pulse peak being about to hit the slab at\nt= 500fs) and the electromagnetic field is character-\nized by standard reflection/transmission features; in par-\nticular both the reflected and transmitted pulses have a4\nFIG. 3: Results of FDTD simulation pertaining the interacti on of the pulse 1 (with ¯ ω=ω1) with the slab. The fields Ex,Ez\nand the Fourier transform F[Ex], for comparison purposes, are captured at same time steps c onsidered in Fig.2. Note that\nbothExandF[Ex] are bell-shaped, that magnitudes of Ezwithin the slab and in vacuum are comparable and that no resid ual\npolariton oscillation has been produced by pulse passage.\ntransverse bell-shaped spatial profile and accordingly the\nFourier transform F[Ex] is peaked as well. Note however\nthat, even in this early transient stage of the interac-\ntion, within the slab the longitudinal component Ezis\ncomparable with Exand much greater than its vacuum\ncounterpart. Such an enhancement of the electric field\ncomponent perpendicular to the slab is a feature typi-\ncally associated with monochromatic ENZ regime, aris-\ningasaconsequenceofthecontinuityofthedisplacement\nfield component perpendicular to the interface [14, 15].\nTherefore this is the first evidence that the ENZ regime\ncaneffectivelybeobservedinthoroughlyrealisticLorentz\nslabs by means of an equally realistic scattering interac-\ntion configuration. The second row of Fig.2 corresponds\nto the time step t= 767fsa time when the incoming\npulse (if freely propagating) would have passed behind\nthe slab (its temporal width being 200 fs). Note that\nthe longitudinal component Ezis even greater than the\nprevious time step, testimony to the fact that the ENZ\nregime also occurs in the time-domain. The transverse\ncomponent Exshows novel spatial features, even more\nevidently displayed by its Fourier transform F[Ex] which\nis no longer bell-shaped and characterized by a complex\nmulti-structured profile. The third row of Fig.2 considers\na later time step t= 941fsmuch longer than the time\nspentbythepulsetofullytravelintotheslabandleaveit.\nAt this time step, the longitudinal component Ezisstill\nvery large within the slab and the transverse component\nExdisplays novel and unexpected features: it is symmet-ric under the reflection z→ −z, it is not transversally\nbell-shaped and its Fourier transform F[Ex] displays two\npeaks at the sides of kx= 0. Such phenomenology can\nbe interpreted only by assuming that the interaction of\npulse 0 with the slab is accompanied by the excitation\nof a polariton mode whose oscillation lasts a time much\nlonger than the pulse-slab interaction time.\nC. Pulse 1scattering\nIn order to appreciate the novelty of the above dis-\ncussed time-domain ENZ phenomenology, we now dis-\ncuss the interaction of pulse 1 with the slab, its carrier\nfrequency being associated to standard slab dielectric be-\nhavior. In Fig.3 we report the results of the FDTD sim-\nulation relative to pulse 1 and, for comparison purposes,\nwe have given Fig.3 the same structure as Fig.2 with the\nsame fields at same time steps. Remarkably, both Ex\nandF[Ex] are everywhere and always bell-shaped, while\nthe magnitude of Ezwithin the slab is comparable with\nits vacuum magnitude. At the last time step the slab\nhosts no residual polariton oscillation resulting from the\npulse passage. This is precisely the standard expected\nphenomenology of the reflection and transmission of the\npulsebyadielectricslab, andthe comparisonwiththere-\nsults of Fig.2 proves that the phenomenology it contains\nis a manifestation of the time-domain ENZ regime5\nFIG. 4: FDTD predictions about the fields ExandEz, for both pulse 0 and pulse 1 as function of ( z,t) at a fixed plane\nx=xp= 1.76µ. Pulse 0 triggers a novel mechanism of metastable light trap ping since its passage produces a strong and\ndamped polariton oscillation which is absent in the case of p ulse 1.\nD. Transient trapping in the time-domain epsilon\nnear zero regime\nIn addition to the remarkable fact that the same fea-\ntures of the monochromatic ENZ regime characterize its\ntime domain counterpart (e.g. the slab hosts a pro-\nnounced enhancement of the field Ez), the results dis-\ncussed in the previous sections also clearly reveal that\nthe scattering situation leads to the unique excitation of\na polariton mode. In order to show more explicitly such\na phenomenology, in Fig.4 we have plotted the fields Ex\nandEz, for both pulse 0 and pulse 1 as functions of ( z,t)\nat a fixed plane x=xp= 1.76µm. The evident fea-\nture that emerges is that pulse 0 (see subplots (a) and\n(b) of Fig.4) produces a strong and damped electromag-\nnetic self-oscillation persisting a time (about 5500 fs)\nmuch longer than the probing pulse duration (200 fs),\nself-oscillation which is conversely not produced by pulse\n1 (see subplots (c) and (d) of Fig.4), whose electromag-\nnetic track fades within the slab just after it has left the\nmedium (at about t= 1000fs). We conclude that, in\nthe ENZ regime, the pulse travelling through the slab\ntriggers a novel mechanism of transient light trapping.\nIII. THEORETICAL ANALYSIS OF\nTIME-DOMAIN ENZ REGIME\nA. Polariton virtual modes analysis\nFrom the above discussed phenomenology, it is evi-\ndent that a quasi-monochromatic pulse with a spectrum\ncentered at the zero of the real part of the slab permit-\ntivity excites a polariton mode that lasts a time much\nlonger than the pulse-slab interaction time. In order to\nrigorously prove this statement and gain deeper under-\nstanding of the underpinning physical mechanisms thatsupport the time-domain ENZ regime, in this section we\nanalyze the exact quasi-steady modes (virtual modes) of\nthe slab. In ouranalysiswefully takeinto accountdamp-\ning processes, which include medium absorbtion and ra-\ndiation leakage in vacuum, adopting the complex fre-\nquency approach [25]. We start our analysis from the\ncurl Maxwell equations for TM fields\n−∂Ez\n∂x+∂Ex\n∂z=−µ0∂Hy\n∂t,\n−∂Hy\n∂z=ǫ0∂Ex\n∂t+∂Px\n∂t, (4)\n∂Hy\n∂x=ǫ0∂Ez\n∂t+∂Pz\n∂t,\nwhere the polarization P(x,z,t) =Px(x,z,t)ˆex+\nPz(x,z,t)ˆezsatisfies Eq.(2) within the slab ( |z|< L/2)\nand it vanishes outside the slab ( |z|> L/2). We take the\nAnsatzAj(x,z,t) = Re/bracketleftbig\naj(z)ei(kxx−Ωt)/bracketrightbig\nfor every field\ncomponent( A=E,H,P,a=e,h,pandj=x,z), where\nkxis the (real) transverse wavevector and Ω = ω−iΓ is\nthe complex angular frequency with Γ >0 so that only\ndamping modes are considered. Owing to the mutual\ntemporal evolution of the electromagnetic field ( E,H)\nand of the polarization field P, the Ansatz effectively\namounts to considering polariton virtual modes. The\nmagnetic field can be expressed in terms of the electric\nfield components hy=−(kxez+i∂zex)/(µ0Ω) so that\nMaxwell’s equations reduce to\nez=ikx\nΩ2\nc2˜ǫ(Ω,z)−k2\nxdex\ndz, (5)\nd2ex\ndz2+/bracketleftbiggΩ2\nc2˜ǫ(Ω,z)−k2\nx/bracketrightbigg\nex= 0, (6)\nwhere ˜ǫ(Ω,z) =ǫ(Ω)θ(L/2− |z|) +θ(|z| −L/2) (θ(z)\nbeing the Heaviside step function) is the z-dependent di-\nelectric profile. It is worth stressing that the permittivity6\nǫis evaluated at the complex frequency Ω. The general\nsolution of Eqs.(5,6) is explicitly given by\nex(z) =C\n\nΘe−iΞK(z+L\n2)z <−L\n2,\neikz+Θe−ikz\neikL\n2+Θe−ikL\n2−L\n2≤z≤L\n2,\neiΞK(z−L/2)z >L\n2,\n(7)\nez(z) =C\n\nΘΞkx\nKe−iΞK(z+L\n2)z <−L\n2,\nkx\nk−eikz+Θe−ikz\neikL\n2+Θe−ikL\n2−L\n2≤z≤L\n2,\n−Ξkx\nKeiΞK(z−L\n2)z >L\n2,\nwhereK=/radicalBig\nΩ2\nc2−k2x,k=/radicalBig\nΩ2\nc2ǫ(Ω)−k2x,Cis the arbi-\ntrary mode amplitude, Θ = ±1 is a parameter that dis-\ntinguishes the symmetry of the solutions and Ξ = ±1 is\nanother parameter selecting the sign of the exponentials\nin vacuum. By construction, the modal fields in Eqs.(7)\nalready satisfy the continuity of the field component par-\nallel to the slab surface ( ex) at the interfaces x=±L/2.\nThe boundary conditions (BCs) for the continuity of the\ndisplacement field component (˜ ǫez)perpendicular to the\ninterfaces x=±L/2 yield the dispersion relation\n(Kǫ−Ξk)eikL= Θ(Kǫ+Ξk), (8)\nwhich provides the complex frequency Ω for every given\nslab thickness Land transverse wave vector kx. We have\nsolved Eq.(8) numerically and obtained the allowed Ω\ncorresponding to different values of kxusing the same\nslab thickness and Lorentz dispersive parameters of the\nslab considered in Section II. In Fig.5 we plot the results\nforthe caseΘ = −1in the complex planeΩ parametrized\nthrough the wavelength λand the damping constant Γ\n(i.e. Ω = 2 πc/λ−iΓ), using circles and stars for the\nΞ =−1 and Ξ = 1 modes, respectively, and using the\nmarker color to label the value of the corresponding kx.\nNote that the Ξ = −1 and Ξ = 1 modes belongs to two\ndifferent branches which are characterized by the fact\nthat the Ξ = −1 modes have damping constant greater\nthan the Ξ = +1 modes. This property can be easily\nunderstood by considering the z-component of the non-\noscillatory part of the Poynting vector S=E×Hfor\nz > L/2:\n/angbracketleftSz/angbracketright= Ξe−2[ΞIm(K)(z−L\n2)+Γt]1\n2ǫ0Re/parenleftbiggΩ\nK/parenrightbigg\n|C|2.(9)\nFor Ξ = −1, the energy outflows from the slab and the\ndamping of the virtual mode is more rapid since it loses\nenergy through both medium absorption and radiation\nleakage. Conversely, for Ξ = 1 the electromagnetic en-\nergyisdraggedintothe slab, thus partiallycompensating\nfor the medium absorption and consequently decreasing\nthe virtual mode damping time (note that there is also a\npoint where Γ = 0 on the Ξ = 1 branch corresponding to0.4655 0.4660 0.46650123456x 10−3Γ (fs−1)\n \n00.511.522.53Θ = −1 kx (µm−1)\n: Ξ = −1\n: Ξ = 1°\n∗\nλ (µm)ΩPδλ\nλ0\nFIG. 5: Virtual modes of the slab considered in Sec.II for\nthe symmetry Θ = −1 in the complex plane Ω = 2 πc/λ−\niΓ. Circles and stars label the Ξ = −1 and Ξ = 1 modes\nwhereas the marker color labels the corresponding kxvalue.\nThe complex frequency Ω Pis such that ǫ(ΩP) = 0. The\nthin continuous line is the temporal spectrum of the incomin g\npulse reported in arbitrary units.\nthe exact balance between medium absorption and radi-\nation drag). It is remarkable that the Ξ = −1 and Ξ = 1\nbranches intersect each other at point Ω Pforkx→0. In\nthis limit the dispersion relation of Eq.(8) (for Θ = −1)\nreduces to\neiΩP\nc√\nǫ(ΩP)=Ξ−/radicalbig\nǫ(ΩP)\nΞ+/radicalbig\nǫ(ΩP), (10)\nand is satisfied only if ǫ(ΩP) = 0. Starting from the\nLorentzmodel, itisstraightforwardtoprovethattheper-\nmittivity vanishes at Ω P=/radicalbig\nfe+ω2e−γ2/4−iγ/2that,\nfor the above used Lorentz dispersive parameters, yields\nλP= 2πc/Re(ΩP) = 0.4667µmand Γ P=−Im(ΩP) =\n8.3·10−3fs−1, precisely matching the point of Fig.5\nwhere the two branches Ξ = ±1 intersect each other. In\nturn, the plasmonic ΩPat which the permittivity van-\nishes plays a central role in the analysis of the virtual\nmodes. For the complex frequencies Ω reported in Fig.5,\n|ǫ(Ω)|<0.06 and therefore, all the obtained modes with\nsymmetry Θ = −1 imply the time-domain ENZ regime.\nIn the same portion of the complex plane Ω we have\nnumerically found no allowed modes for the symmetry\nΘ = 1. This can be grasped by expanding both sides\nof Eq.(8) in Taylor series of ǫ(since|ǫ(Ω)| ≪1); at the\nzeroth order we readily obtain e−kxL=−Θ which is not\nconsistent if Θ = 1 (and which, on the other hand, yields\nkx= 0 for Θ = −1).\nB. Interpretation of FDTD results in terms of\nvirtual polariton modes and slow-light regime\nUsually, within the standard real frequency approach,\nthe solutions for the slab modes with Ξ = −1 are dis-7\nregarded since they are considered unphysical. Indeed,\nif|kx|> ω/c, the solutions with Ξ = +1 represent con-\nfined modes propagating along the x-direction, while so-\nlutions with Ξ = −1 are unbound modes that diverge\natz→ ±∞. In addition, the introduction of the com-\nplex frequency introduces an inherent field singularity\nin the far past t→ −∞. Due to such intrinsic singu-\nlarities, it is strictly impossible to rigorously excite a\nsingle virtual mode, its global existence on the whole\nspace-time being unphysical. However, both singular-\nities occur asymptotically and therefore virtual modes\nprovide a very adequate description of the transient ENZ\nslabbehavioroccurringwithinaspatiallyboundedregion\nand through a finite time lapse. In order to prove this\nstatement and to basically provide a theoretical analyt-\nical description of the transient light trapping discussed\nin Sec.2D, in Fig.5 we have superimposed the temporal\nspectrum profile of the incoming pulse of Eq.(1) (using\nthe thin continuous line) on the complex-plane virtual\nmodal structure. Note that, due to its wavelength band-\nwidthδλ=/parenleftbig\n2πc/ω2\n0/parenrightbig\nδω= 5.8·10−4µm, the pulse spec-\ntrum centered at λ0overlaps a limited portion of the\nconsidered complex frequency plane so that, specifically,\nthe sole virtual modes with |λ−λ0|< δλ/2 are actu-\nally excited by the considered pulse 0. From Fig.5 it\nis evident that the excited virtual modes are character-\nized by the transverse wavevector kxspanning the range\n1.1µm−1< kx<2.7µm−1and that, due to the fi-\nnite bandwidth of the impinging pulse, the excited vir-\ntual modes with largest amplitude are those around the\ncentral transverse wavevector kx= 1.9µm, which cor-\nresponds to λ= 2πc/Re(Ω) close to λ0= 0.466. This\nobservation is in striking agreement with the results con-\ntained in Fig.2, where one can see that the transverse\nFourier transform of the field has, at the latest time step,\ntwo peaks centered at kx≃1.9µmandkx≃ −1.9µm\nwhose width is of the order of δkx≃0.8µm. There-\nfore, when the pulse 0 impinges onto the slab, it ex-\ncites precisely the virtual modes analytically predicted\nin Sec.IIIA, which are compatible with its spectral struc-\nture. As a further validation of this statement, note that\nthe virtual modes excited by the pulse 0 ( |λ−λ0|<\nδλ/2) have a damping constant Γ spanning the range\n1.2·10−3fs−1<Γ<4.9·10−3fs−1(see Fig.5), which\ncorresponds to the extinction time τ= 3/Γ spanning the\nrange 609 fs < τ < 2604fs. Also this prediction based\non the above virtual mode analysis is in striking agree-\nment with the FDTD results since, by looking at panels\n(a) and (b) of Fig.4, one can see that the electromag-\nnetic excitationpersists for a time ofthe orderof3000 fs,\nwhichiscompatiblewiththemaximumextinctiontimeof\nthe excited virtual modes. In addition, the spatial sym-\nmetry of the Θ = −1 virtual polariton modes matches\nthe numerical results displayed in Fig.4: the transverse\nfield component ( ex) is antisymmetricwith respectto the\nz= 0 axis, whereas the longitudinal field component ( ez)\nis symmetric.\nThe final ingredient needed to thoroughlyinterpret thetransient light trapping observed in FDTD simulations\nis related to the intrinsic slow-light nature of the phe-\nnomenon, which may be preliminarily grasped by con-\nsidering a bulk Lorentz medium, where transverse plane\nwaves satisfy the dispersion relation k(ω) = (ω/c)/radicalbig\nǫ(ω).\nNeglecting medium absorption ( γ≃0), one finds that\nthe phase velocity is vf=ω/kand the group velocity\nvg= dω/dkis\nvg(ω) =c/bracketleftBigg\n/radicalbig\nǫ(ω)+feω2\n/radicalbig\nǫ(ω)(ω2e−ω2)2/bracketrightBigg−1\n.(11)\nThus, in the ENZ regime the phase velocity diverges\nvf→ ∞whereasthe groupvelocitytends to zero vg→0.\nEven though the subwavelength Lorentz slab used in our\nFDTD simulations is not a bulk medium and absorp-\ntion has not been neglected, the rough argument above\nstillpredictsthecorrectoutcome. Indeed, bynumerically\nsolving the dispersion relation of Eq.(8) without neglect-\ninglossesonefindsthat, attheopticalwavelength λ0, the\ntransversephasevelocityofvirtualpolaritonmodesis su-\nperluminal vf=ω/kx≃10c, while the transverse group\nvelocity is extremely reduced vg= dω/dkx≃c/100.\nFor this reason, it is now clear how the virtual polari-\nton modes, once excited, do not disperse quickly in the\nx-direction and remain quasi-trappedwithin the slab ow-\ning to the tremendously reduced temporal dynamics. We\nconclude that the above described transient light trap-\nping can be fully interpreted and physically understood\nby means of slow polariton modes supported by the slab.\nC. Volume plasmons\nAlthough the above discussed numerical and analyti-\ncal analysis of the transient light trapping characteriz-\ning the time-domain ENZ regime is quite exhaustive, we\nnow discuss its connection with the purely longitudinal\nmodes, either volume plasmons (collective oscillations of\nelectrons) or volume phonons (collective oscillations of\nions), which the Lorentz medium can support. Hereafter\nwe focus on volume plasmons, considering an unbounded\nbulk Lorentzmedium where the TM electromagneticand\npolarizationdynamicsaredescribedbyEqs.(2,4). Forthe\nplane-wave Ansatz Aj(x,z,t) = Re/bracketleftbig\najei(kxx+kzz−Ωt)/bracketrightbig\n,\nwhereA=E,H,P,a=e,h,p,j=x,z,kx,kzare\nthe (real) wavevector components and Ω is the generally\ncomplex frequency one gets\nµ0Ωhy=−kxez+kzex,\nkzhy= Ωǫ0ǫ(Ω)ex, (12)\nkxhy=−Ωǫ0ǫ(Ω)ez,\nwhereǫ(Ω) is the dielectric permittivity with complex\nfrequency Ω. Volume plasmons are purely longitudinal\nelectric oscillations owing to to the collective motion of\nelectrons and are not accompanied by the generation of\nmagnetic field, a feature that for plane waves amounts8\nto the collinearity of the wave vector k=kxˆex+kzˆez\nand the electric field E=kxˆEx+kzˆEz. Therefore, im-\nposing the condition k×E= 0, i.e. −kxez+kzex= 0,\nEqs.(12) readily yield hy= 0 and ǫ(ΩP) = 0. Thus, vol-\nume plasmonsareinherentlyinvolvedin the time-domain\nENZ regime we are considering in this paper. However,\nit is worth noting that the virtual modes of the Lorentz\nslab are polaritons, entities fundamentally different from\nvolume plasmons (or volume phonons). Indeed, a vol-\nume plasmon is strictly characterized by the condition\nǫ(Ω) = 0 that implies the severe dispersion Ω = Ω Pand,\nin the presence of the slab boundaries at z=±L/2, in-\nevitably leads to the inconsistency Ez→ ∞within the\nslab unless Ez= 0 in the outer medium. This is consis-\ntent with the well-known impossibility to excite volume\nplasmons by means of light. On the other hand, from\nEq.(7) one can see that the virtual polariton mode com-\nponentEzneither vanishes outside the slab nor diverges\nwithin it. This is because for polaritons the dispersion\nrelation of Eq.(8) is not as stiff as the volume plasmon\ndispersion Ω = Ω Pand is satisfied also for Ω /negationslash= ΩP. In\naddition the volume plasmon is a purely electric oscilla-\ntion with strictly null magnetic field, whereas the consid-\nered virtual polariton modes are accompanied by a mag-\nneticfield. Inturn, eventhoughTMpolaritonmodesand\nvolume plasmons occur in the same spectral region and\nare accidentally connected by the fact that in the limit\nL >> λ the Lorentz slab is almost equivalent to a bulk\nmedium, conceptually they are very distinct entities. In\nview of this, we conclude remarking that volume plas-\nmons can not be excited by classical light and that the\nabsorption peak observed in experiments [28, 30] is dueto the excitation of virtual polariton modes, confirming\nthe results given in Refs. [27, 29].\nIV. CONCLUSIONS\nIn conclusion we have investigated both numerically\nand analytically the properties of the time-domain ENZ\nregime. Specifically we have considered a dielectric slab\nwhose polarization dynamics has been described through\nthe realistic and ubiquitous Lorentz model and we have\nanalyzed its interaction with quasi-monochromatic and\nspatially confined pulses with carrier frequencies close\nto the crossing point of the permittivity real part. The\nFDTD analysis has shown that the pulse is able to ex-\ncite a polarization-electromagnetic(polariton) oscillation\nwhich is damped and persists for a time generally longer\nthan the effective time required by the pulse for passing\nthrough the slab. The underlying nature of this excita-\ntion has been elucidated through the analysis of the slab\nvirtual modes that turn out to be located in a portion\nof the complex frequency plane close to the plasmonic\nfrequency characterizing plasmon/phonons longitudinal\nvolume excitations. Remarkably, due to this spectral\nproperty, both the group velocity and the transverse ve-\nlocity (parallel to the slab) of each virtual mode turn out\nto be very small and therefore, the time-domain ENZ\nregime be naturally regarded as a novel platform for dis-\ncussing and investigating a plethora of slow-light phe-\nnomena.\n[1] K. L. Tsakmakidis, A. D. Boardman, and O. Hess, Na-\nture Letters 450, 397 (2007).\n[2] R. W. Boyd, Journal of Modern Optics 56, 1908 (2009).\n[3] K. H. Kim, A. Husakou, and J. Herrmann, Optics Ex-\npress20, 25790 (2012).\n[4] Y. A. Vlasov, M. O’Boyle, H. F. Hamann, and S. J. Mc-\nNab, Nature 438, 65 (2005).\n[5] S. F. Mingaleev, A. E. Miroshnichenko, Y. S. Kivshar,\nand K. Busch, Physical Review E 74, 046603 (2006).\n[6] M. Bajcsy, S. Hofferberth, V. Balic, T. Peyronel,\nM. Hafezi, A. S. Zibrov, V. Vuletic, and M. D. Lukin,\nPhysical Review Letters 102, 203902 (2009).\n[7] S. V. Afshar and T. M. Monro, Optics Express 17, 2298\n(2009).\n[8] A. Marini, R. Hartley, A. V. Gorbach, and D. V.\nSkryabin, Physical Review A 84, 063839 (2011).\n[9] C. Koos, L. Jacome, C. Poulton, J. Leuthold, and\nW. Freude, Optics Express 15, 5976 (2007).\n[10] S. Palomba and L. Novotny, Physical Review Letters\n101, 056802 (2008).\n[11] M. Kauranen and A. V. Zayats, Nature Photonics 6, 737\n(2012).\n[12] A. Ciattoni, C. Rizza, and E. Palange, Physical Review\nA81, 043839 (2010).[13] A. Ciattoni, C. Rizza, and E. Palange, Physical Review\nA83, 043813 (2011).\n[14] M. A. Vincenti, D. DeCeglia, A. Ciattoni, and\nM. Scalora, Physical Review A 84, 063826 (2011).\n[15] A. Ciattoni and E. Spinozzi, Physical Review A 85,\n043806 (2012).\n[16] D. A. Powell, A. Alu, B. Edwards, A. Vakil, Y. S.\nKivshar, and N. Engheta, Physical Review B 79, 245135\n(2009).\n[17] C. Argyropoulos, P.-Y.Chen, G. DAguanno, N.Engheta,\nand A. Alu, Physical Review B 85, 045129 (2012).\n[18] S. Enoch, G. Tayeb, P. Sabouroux, N. Guerin, and\nP. Vincent, Physical Review Letters 89, 213902 (2002).\n[19] A.Alu, F.Bilotti, N.Engheta, andL.Vegni, IEEETrans.\nAntennas Propag. 54, 1632 (2006).\n[20] A. Alu and N. Engheta, Physical Review E 72, 016623\n(2005).\n[21] M. Silveirinha and N. Engheta, Physical Review Letters\n97, 157403 (2006).\n[22] A. Alu, M. G. Silveirinha, A. Salandrino, and N. En-\ngheta, Physical Review B 75, 155410 (2007).\n[23] G. Castaldi, S. Savoia, V. Galdi, A. Alu, and N. Engheta,\nPhysical Review B 86, 115123 (2012).\n[24] A. Archambault, M. Besbes, and J.-J. Greffet, Physical9\nReview Letters 109, 097405 (2012).\n[25] A. Archambault, T. V. Teperik, F. Marquier, and J. J.\nGreffet, Physical Review B 79, 195414 (2009).\n[26] N. Ashcroft and N. Mermin, Solid State Physics (Har-\ncourt College Publishers, 1976).\n[27] L. Henrard, O.Stephan, andC. Colliex, SyntheticMetal s\n103, 2502 (1999).\n[28] K. Hoeflich, U. Goesele, and S. Christiansen, Physical\nReview Letters 103, 087404 (2009).\n[29] L. Henrard, C. VanDenBem, P. Lambin, and A. Lucas,\nPhysical Review Letters 104, 149701 (2010).\n[30] K. Hoeflich, U. Goesele, and S. Christiansen, PhysicalReview Letters 104, 149702 (2010).\n[31] P. Muys, Optics Letters 37, 4928 (2012).\n[32] D. W. Berreman, Physical Review 130, 2193 (1963).\n[33] R. Ruppin and R. Englman, Rep. Prog. Phys. 33, 149\n(1970).\n[34] S. Vassant, J.-P. Hugonin, F. Marquier, and J.-J. Greffe t,\nOptics Express 20, 23971 (2012).\n[35] S. Vassant, A. Archambault, F. Marquier, F. Pardo,\nU. Gennser, A. Cavanna, J. L. Pelouard, and J. J. Gref-\nfet, Physical Review Letters 109, 237401 (2012)." }, { "title": "2109.07597v1.Penning_Trap_Searches_for_Lorentz_and_CPT_Violation.pdf", "content": "arXiv:2109.07597v1 [hep-ph] 15 Sep 2021\nArticle\nPenning-Trap Searches for Lorentz and CPT Violation\nYunhua Ding1,2,*, Teague D. Olewiler2and Mohammad Farhan Rawnak2\n1W.M. Keck Science Department, Claremont McKenna, Pitzer, a nd Scripps Colleges, Claremont, CA 91711,\nUSA; yding@kecksci.claremont.edu (Y.D.)\n2Department of Physics, Gettysburg College, Gettysburg, PA 17325, USA;\nolewte01@gettysburg.edu(T.D.O.); rawnmo01@gettysburg .edu(M.F.R.)\n*Correspondence: yding@kecksci.claremont.edu\nVersion September 17, 2021 submitted to Symmetry\nAbstract: An overview of recent progress on testing Lorentz and CPT sym metry using Penning\ntraps is presented. The theory of quantum electrodynamics w ith Lorentz-violating operators of mass\ndimensions up to six is summarized. Dominant shifts in the cy clotron and anomaly frequencies of\nthe confined particles and antiparticles due to Lorentz and C PT violation are derived. Existing\nresults of the comparisons of charge-to-mass ratios and mag netic moments involving protons,\nantiprotons, electrons, and positrons are used to constrai n various coefficients for Lorentz violation.\nKeywords: Lorentz and CPT violation; Penning trap; Standard-Model Ex tension\n1. Introduction\nPrecision experiments involving Penning traps have in rece nt years achieved impressive\nsensitivities to properties of fundamental particles. For example, the magnetic moment of electrons\nhas been measured to a record precision of 0.28 ppt [ 1]. The high precision offered by Penning-trap\nexperiments provides excellent opportunities to test fund amental symmetries. This includes the\nLorentz symmetry, one of the foundations of both general rel ativity and the Standard Model of\nparticle physics. It has been recently shown that tiny viola tions of Lorentz symmetry could naturally\narise in a fundamental theory that unifies gravity with quant um physics at the Plank scale MP∼1019\nGeV , such as string theory [ 2,3]. Since in any effective field theory, violations of CPT symm etry\nalso break Lorentz symmetry [ 4–6], testing Lorentz symmetry also includes CPT tests. In rece nt\nyears, searches for Lorentz and CPT violation in precision e xperiments have been performed across\nmany subfields of physics [ 7], including Penning-trap experiments. Here, we provide an overview\nof the recent progress on testing Lorentz and CPT violation i n Penning-trap experiments measuring\ncharge-to-mass ratios and magnetic moments of protons, ant iprotons, electrons, and positrons.\nIn the context of effective field theory, the comprehensive f ramework that describes all possible\nLorentz violation is the Standard-Model Extension (SME) [ 4,5,8]. The Lagrange density of the SME is\nconstructed from general relativity and the Standard Model by adding all possible Lorentz-violating\nterms. Each of such terms is formed from a coordinate-indepe ndent contraction of a general\nLorentz-violating operator with a corresponding coefficie nt. The subset of the SME with operators\nof mass dimensions d≤4 is called the minimal SME, which is power-counting renorma lizable.\nFor the nonminimal SME, it restricts attention to operators of mass dimensions d>4, which\nis viewed to produce higher-order effects. Study of the nonm inimal SME serves as a basis\nfor further investigations of many aspects of Lorentz and CP T violation, such the causality and\nstability [ 9,10], Lorentz-violating models in supersymmetry [ 11], noncommutative Lorentz-violating\nquantum electrodynamics [ 12–14], and the underlying pseudo-Riemann-Finsler geometry [ 15–18].\nSubmitted to Symmetry , pages 1 – 20 www.mdpi.com/journal/symmetryVersion September 17, 2021 submitted to Symmetry 2 of 20\nFor Penning-trap experiments measuring charge-to-mass ra tios and magnetic moments of\nconfined particles or antiparticles, both the minimal and no nminimal SME can produce various\nmeasurable Lorentz- and CPT-violating effects via shifts i n the cyclotron and anomaly frequencies.\nThese effects in general can depend on sidereal time and diff er between particles and antiparticles. In\nthe minimal SME, Refs. [ 19,20] present the first theoretical analysis to study Lorentz and CPT violation\nin Penning traps. An extension to the nonminimal SME by inclu ding Lorentz-violating operators\nof mass dimensions up to six was recently made in Ref. [ 21], in which analysis of the magnetic\nmoment comparisons between particles and antiparticles us ing Penning traps was also performed.\nA similar application to charge-to-ratio comparisons is pr esented in Ref. [ 22]. For the effects arising\nfrom sidereal variations due to the Earth’s rotation, the re lated discussions are given in Ref. [ 23–25].\nIn this work, we provide an overview of recent progress on sea rching for Lorentz- and\nCPT-violating signals using Penning traps and provide the m ost updated constraints on the\ncoefficients for Lorentz violation that are relevant to thes e experiments. The results provided in\nthis work are complementary to these from the studies of Lore ntz and CPT violation in experiments\ninvolving measurements of the muon’s anomalous magnetic mo ment [ 26,27], the spectroscopic\nanalysis of hydrogen, antihydrogen, and other related syst ems [ 28], and clock comparisons [ 29].\nThis work is organized as follows. We start in Section 2with the related theory, where we\npresent in Subsection 2.1the theory of Lorentz-violating quantum electrodynamics w ith operators\nof mass dimensions up to six. The perturbative energy shifts to the confined particles or antiparticles\nare obtained in Subsection 2.2using perturbation theory. Subsection 2.3gives the corresponding\nshifts in the cyclotron and anomaly frequencies due to Loren tz and CPT violation. The discussion\nof sidereal variations and rotation matrices are given in Su bsection 2.4. Then we turn in Section 3\nto experimental applications to various Penning-trap expe riments and present the constraints on the\ncoefficients for Lorentz violation. The applications to cha rge-to-mass ratio comparisons are treated in\nSubsection 3.1, with Subsection 3.1.1 focusing on the proton sector, and Subsection 3.1.2 discussing the\nelectron sector, respectively. The resulting constraints on the coefficients for Lorentz violation from\nthe reported experimental results are summarized in Table 1, Table 2, and Table 3. Subsection 3.2\ndiscusses the applications to magnetic moment comparisons , with Subsection 3.2.1 for the proton\nsector, and Subsection 3.2.2 for the electron sector, respectively. The corresponding l imits on the\ncoefficients for Lorentz violation are listed in Table 4and Table 5. The summary of this work is\ngiven in Section 4. Finally, Appendix Alists the explicit expressions of the transformations of th e\nrelevant coefficients into different frames. Throughout th is work, we follow the same notation used\nin Refs. [ 21,22] and adopt natural units with ¯ h=c=1.\n2. Theory\nIn this section, we summarize the theory developed in Ref. [ 21] of Lorentz-violating quantum\nelectrodynamics with operators of mass dimension d≤6 and derive the energy shifts for particles\nand antiparticles confined in Penning traps due to Lorentz an d CPT violation.\n2.1. Lagrange density\nFor a single Dirac fermion field ψof charge qand mass mψ, the general Lorentz-violating\nLagrange density Lψcan be constructed by adding a general Lorentz-violating op erator/hatwideQto the\nconventional Lagrange density,\nLψ=1\n2ψ(γµiDµ−mψ+/hatwideQ)ψ+H.c., (1)\nwhere Dµ= (∂µ+iqA µ)is the covariant derivative given by the minimal coupling wi thAµbeing\nthe electromagnetic four-potential. H.c. means Hermitian conjugate. The general Lorentz-violating\noperator /hatwideQin the Lagrange density ( 1) is a 4×4 spinor matrix that contains terms formed by\nthe contraction of a generic coefficient for Lorentz violati on, the covariant derivative iDµ, theVersion September 17, 2021 submitted to Symmetry 3 of 20\nantisymmetric electromagnetic field tensor Fαβ≡∂αAβ−∂βAα, and one of the 16 Dirac matrix\nbases. For example, one of the dimension-five operators invo lving the F-type coefficients for Lorentz\nviolation takes the form b(5)µαβ\nFFαβγ5γµ. For mass dimension d≤6, a full list of the relevant\ncoefficients for Lorentz violation and their properties are given in Table I in Ref. [ 21]. Note that\nthe hermiticity of the Lagrange density ( 1) indicates that the operator /hatwideQsatisfies the condition\n/hatwideQ=γ0/hatwideQ†γ0. In the free-fermion limit where Aα=0, the explicit expression of the Lagrange density\n(1) at arbitrary mass dimension has been studied in Ref. [ 30]. For the interaction case where Aα/negationslash=0,\nRef. [ 21] developed a theory for operators of mass dimensions up to si x. An extension of the theory\nto include operators of arbitrary mass dimension was recent ly given in Ref. [ 31]. Similar analysis\nhas also been performed for other SME sectors, including the se for photon [ 32], neutrino [ 33] and\ngravity [ 34].\nDue to the existence of the general operator /hatwideQin the Lagrange density ( 1), the conventional Dirac\nequation for a fermion in electromagnetic fields is modified t o\n(p·γ−mψ+/hatwideQ)ψ=0, (2)\nwhere we have chosen the momentum space for convenience, wit h the identification pα↔iDα. Given\nthe fact that no Lorentz-violating signals have been observ ed so far, any such signal must be tiny\ncompared to the energy scale of the system of interest. There fore, we can treat the corrections due to\nLorentz and CPT violation to the conventional Hamiltonian a s perturbative and apply perturbation\ntheory to obtain the dominant shifts to the energy levels of t he confined particles and antiparticles.\nFrom Equation ( 2), the exact Hamiltonian Hcan be defined as\nHψ≡p0ψ=γ0(p·γ+mψ−/hatwideQ)ψ= (H0+δH)ψ, (3)\nwhere p0is the exact energy of the system of interest, including all c ontributions from Lorentz and\nCPT violation,H0is the conventional Hamiltonian for a fermion in an electrom agnetic field, and\nδH=−γ0/hatwideQis the exact perturbative Hamiltonian.\nTo construct δH, we note that operator /hatwideQin general contains terms that are of powers of p0,\nwhich corresponds to the exact Hamiltonian Hitself. In certain simple cases, one can apply a field\nredefinition to remove the additional time derivatives and a dopt the standard procedure involving\ntime translation on wave functions to obtain the exact pertu rbative Hamiltonian δH[20]. In more\ngeneral cases, it is challenging to directly construct δHdue to the existence of powers of time\nderivatives. However, we notice that any contributions to δHdue to the exact Hamiltonian Hare at\nsecond or higher orders in the coefficients for Lorentz viola tion. Therefore, to obtain the leading-order\nresults, one can apply the following substitution [ 21,33],\nδH≈− γ0/hatwideQ|p0→E0, (4)\nwhere E0is the unperturbative eigenvalue, which can be obtained by s olving the conventional Dirac\nequation for a fermion in an electromagnetic field.\n2.2. Perturbative energy shifts\nWith the perturbative Hamiltonian δHdetermined by Equation ( 4), the shifts in the energy levels\nof a confined particle can be obtained by applying perturbati on theory,\nδEn,±=/angbracketleftχn,±|δH|χn,±/angbracketright, (5)\nwhere χn,±are the unperturbative stationary eigenstates with nbeing the level number and ±\ndenoting the spin for a positive-energy fermion. δEn,±are the perturbative corrections to the energy\nlevels due to Lorentz and CPT violation.Version September 17, 2021 submitted to Symmetry 4 of 20\nTo derive the expressions for δEn,±, we can take an idealized Penning trap where a constant\nuniform magnetic field is applied to confine the particle’s ra dial motion and a quadrupole electric\nfield provides the axial confinement. The dominant effects in the unperturbative energy levels are\ndue to the interactions between the confined particle and the magnetic field. The quadruple electric\nfield generates effects suppressed by a factor of E/B≈10−5in natural units for a typical Penning\ntrap field configuration with E≈20 kV/m and B≈5 T. Therefore, the leading-order results in\nδEn,±can be obtained by further idealizing the trap as a pure unifo rm magnetic field in which a\nquantum fermion moves. Following the above discussion, for a spin-1/2 fermion, the leading-order\nperturbative energy shifts due to Lorentz and CPT violation are found to be [ 22]\nδEw\nn,±1=/tildewidea0\nw∓σ/tildewideb3\nw−/tildewidem3\nF,wB±σ/tildewideb33\nF,wB+/parenleftbig±σ/tildewideb′3\nw−mw[/tildewidec00\nw+(/tildewidec11\nw+/tildewidec22\nw)s]/parenrightbig(2n+1∓σ)|qB|\n2m2w\n+/parenleftbig∓σ(/tildewideb311\nw+/tildewideb322\nw)−1\nmw(/tildewidec11\nw+/tildewidec22\nw)s¬/parenrightbig(2n+1)|qB|\n2, (6)\nwhere σis the charge sign of the fermion, given by q≡σ|q|. The subscript windicates the fermion\nspecies. For example, for electrons and protons, w=eand w=p, respectively. The tilde coefficients\nare defined by\n/tildewidea0\nw=a0\nw−mwc00\nw−mwe0\nw+m2\nwm(5)00\nw+m2\nwa(5)000\nw−m3\nwc(6)0000\nw−m3\nwe(6)000\nw ,\n/tildewideb3\nw=b3\nw+H12\nw−mwd30\nw−mwg120\nw+m2\nwb(5)300\nw+m2\nwH(5)1200\nw−m3\nwd(6)3000\nw−m3\nwg(6)12000\nw ,\n/tildewidem3\nF,w=m(5)12\nF,w+a(5)012\nF,w−mwc(6)0012\nF,w−mwe(6)012\nF,w,\n/tildewideb33\nF,w=b(5)312\nF,w+H(5)1212\nF,w−mwd(6)3012\nF,w−mwg(6)12012\nF,w,\n/tildewideb′3\nw=b3\nw+mw(g120\nw−g012\nw+g021\nw)−m2\nwb(5)300\nw−2m2\nw(H(5)1200\nw−H(5)0102\nw+H(5)0201\nw)\n+2m3\nwd(6)3000\nw+3m3\nw(g(6)12000\nw−g(6)01002\nw+g(6)02001\nw),\n/tildewidec00\nw=c00\nw−mwm(5)00\nw−2mwa(5)000\nw+3m2\nwc(6)0000\nw+2m2\nwe(6)000\nw , (7)\nand the “11+22\" types of tilde coefficients are defined by\n(/tildewidecjj\nw)s=cjj\nw−2mwa(5)j0j\nw+3m2\nwc(6)j00j\nw ,\n(/tildewidecjj\nw)s¬=−mwa(5)0jj\nw−mwm(5)jj\nw+3m2\nwc(6)00jj\nw+3m2\nwe(6)0jj\nw ,\n/tildewideb3jj\nw=b(5)3jj\nw+H(5)12jj\nw−3mwd(6)30jj\nw−3mwg(6)120jj\nw , (8)\nwith j=1 or 2 only. The subscripts sand s¬in the/tildewidecjj\nwtilde coefficients in the energy shifts ( 6) show\nthat(/tildewidecjj\nw)sgive both spin-dependent and spin-independent energy shif ts, while (/tildewidecjj\nw)s¬produce only\nspin-independent ones, as evident from the corresponding p roportional factors 2 n+1∓σand 2 n+1.\nWe note that the tilde coefficients /tildewidea0\nw,/tildewideb3\nw,/tildewidem3\nF,w, and/tildewideb33\nF,win the energy shifts ( 6) produce effects that\nare independent of the level number n.\nFollowing a similar analysis, the corresponding leading-o rder shifts to the energy levels of\nantifermions can be determined by using\nδEc\nn,±=/angbracketleftχc\nn,±|δHc|χc\nn,±/angbracketright, (9)\nwhere χc\nn,±are the positive-energy antifermion eigenstates, obtaine d by applying charge conjugation\non the negative-energy fermion solutions χn,±.δHcis the antifermion perturbative Hamiltonian,Version September 17, 2021 submitted to Symmetry 5 of 20\nwhich can be obtained from δHby reversing the charge sign σand changing the signs for all CPT-odd\ncoefficients in operator /hatwideQ. Using Equation ( 9), the antifermion results are given by [ 22]\nδEw\nn,±1=−/tildewidea∗0\nw±σ/tildewideb∗3\nw−/tildewidem∗3\nF,wB∓σ/tildewideb∗33\nF,wB+/parenleftbig∓σ/tildewideb′∗3\nw−mw[/tildewidec00\nw+(/tildewidec11\nw+/tildewidec22\nw)s]/parenrightbig(2n+1∓σ)|qB|\n2m2w\n+/parenleftbig±σ(/tildewideb∗311\nw+/tildewideb∗322\nw)−1\nmw(/tildewidec11\nw+/tildewidec22\nw)s¬/parenrightbig(2n+1)|qB|\n2, (10)\nwhere the tilde quantities with a star subscript are given by\n/tildewidea∗0\nw=a0\nw+mwc00\nw−mwe0\nw−m2\nwm(5)00\nw+m2\nwa(5)000\nw+m3\nwc(6)0000\nw−m3\nwe(6)000\nw ,\n/tildewideb∗3\nw=b3\nw−H12\nw+mwd30\nw−mwg120\nw+m2\nwb(5)300\nw−m2\nwH(5)1200\nw+m3\nwd(6)3000\nw−m3\nwg(6)12000\nw ,\n/tildewidem∗3\nF,w=m(5)12\nF,w−a(5)012\nF,w−mwc(6)0012\nF,w+mwe(6)012\nF,w,\n/tildewideb∗33\nF,w=b(5)312\nF,w−H(5)1212\nF,w+mwd(6)3012\nF,w−mwg(6)12012\nF,w,\n/tildewideb′∗3\nw=b3\nw+mw(g120\nw−g012\nw+g021\nw)−m2\nwb(5)300\nw+2m2\nw(H(5)1200\nw−H(5)0102\nw+H(5)0201\nw)\n−2m3\nwd(6)3000\nw+3m3\nw(g(6)12000\nw−g(6)01002\nw+g(6)02001\nw),\n/tildewidec∗00\nw=c00\nw−mwm(5)00\nw+2mwa(5)000\nw+3m2\nwc(6)0000\nw−2m2\nwe(6)000\nw , (11)\nand the corresponding “11+22\" starred tilde quantities are defined by\n(/tildewidec∗jj\nw)s=cjj\nw+2mwa(5)j0j\nw+3m2\nwc(6)j00j\nw ,\n(/tildewidec∗jj\nw)s¬=mwa(5)0jj\nw−mwm(5)jj\nw+3m2\nwc(6)00jj\nw−3m2\nwe(6)0jj\nw ,\n/tildewideb∗3jj\nw=b(5)3jj\nw−H(5)12jj\nw+3mwd(6)30jj\nw−3mwg(6)120jj\nw . (12)\nIn the result ( 10), the charge sign σis understood to be reversed for the antifermion. Comparing\nthe fermion and antifermion energy shifts results ( 6) and ( 10),δEw\nn,±1can also been obtained from\nδEw\nn,±1by reversing the charge sign σ, the spin orientation, and the signs of all CPT-odd coefficie nts\nfor Lorentz violation, as expected.\nWe remark in passing that the rotation properties of the coef ficients for Lorentz violation\nappearing in results ( 6) and ( 10) are represented by their indices. For example, the index pa ir “12\" on\nthe right sides of the definitions ( 7) and ( 11) is antisymmetric. This implies that these coefficients for\nLorentz violation transform like a single “3\" index under ro tations, while coefficients with an index\n“0\" or an index pair “00\" are invariant under rotations. The c ylindrical symmetry of the Penning\ntrap is correctly reflected by the fact that the results ( 6) and ( 10) only depend on index “0\", “3\", and\n“11+22\".\n2.3. Cyclotron and anomaly frequencies\nThe primary observables of interest in a Penning-trap exper iment are frequencies. Two key\nfrequencies are the cyclotron frequency νc≡ωc/2πand the Larmor spin-precession frequency νL≡\nωL/2π. The difference of the two frequencies gives the anomaly fre quency νL−νc=νa≡ωa/2π\n[35]. For a confined fermion of flavor win a Penning trap, the cyclotron and anomaly frequencies are\ndefined as the energy difference between the following energ y levels [ 20,21],\nωw\nc=Ew\n1,σ−Ew\n0,σ, ωw\na=Ew\n0,−σ−Ew\n1,σ. (13)\nFor an antifermion of flavor w, the corresponding definitions for the cyclotron and anomal y\nfrequencies are given by [ 20,21]\nωw\nc=Ew\n1,σ−Ew\n0,σ, ωw\na=Ew\n0,−σ−Ew\n1,σ, (14)Version September 17, 2021 submitted to Symmetry 6 of 20\nwith the understanding that the charge signs σin definitions ( 14) are reversed compared to these in\ndefinitions ( 13).\nIn a Lorentz-invariant scenario, the charge-to-mass ratio and the gfactor of a particle or an\nantiparticle confined in a Penning trap with a magnetic field s trength Bare related to the above\ncyclotron and anomaly frequencies by\n|q|\nm=ωc\nB, (15)\nand\ng\n2=ωL\nωc=1+ωa\nωc, (16)\nrespectively. For Penning-trap experiments comparing the charge-to-mass ratios or the gfactors\nbetween a particle of species wand its corresponding antiparticle w, the CPT theorem guarantees\nthat the following differences must be zero,\n(|q|/m)w\n(|q|/m)w−1=ωwc\nωwc−1=0, (17)\nand\n1\n2(gw−gw) =ωwa\nωwc−ωwa\nωwc=0, (18)\nwhere in Equation ( 17), for simplicity, we have assumed the same magnetic field is u sed in the\ncomparison. When different magnetic fields are used, the exp ression can be easily obtained by\nsubstituting ωw\nc/ωw\ncin the middle term of Equation ( 17) by(ωw\nc/Bw)(ωw\nc/Bw), where Bwand Bw\nare the strengths of the magnetic fields used to confine the par ticle and antiparticle, respectively.\nHowever, this assumption is not required to derive Equation (18) as the ratios ωw\na/ωw\ncand ωw\na/ωw\nc\ndon’t depend on the magnetic field used in the trap.\nHowever, in the presence of Lorentz and CPT violation, the pi cture changes dramatically as\nboth the cyclotron and anomaly frequencies for fermions and antifermions can be shifted due to the\nLorentz- and CPT-violating corrections to the energy level s,\nδωw\nc=δEw\n1,σ−δEw\n0,σ, δωw\na=δEw\n0,−σ−δEw\n1,σ,\nδωw\nc=δEw\n1,σ−δEw\n0,σ δωw\na=δEw\n0,−σ−δEw\n1,σ. (19)\nApplying energy shift results ( 6) and ( 10), the corrections to the cyclotron frequencies for a fermio n\nand antifermion are found to be [ 22]\nδωw\nc=/parenleftbigg1\nm2w/tildewideb′3\nw−1\nmw(/tildewidec00\nw+/tildewidec11\nw+/tildewidec22\nw)−(/tildewideb311\nw+/tildewideb322\nw)/parenrightbigg\neB,\nδωw\nc=/parenleftbigg\n−1\nm2w/tildewideb′∗3\nw−1\nmw(/tildewidec∗00\nw+/tildewidec∗11\nw+/tildewidec∗22\nw)+(/tildewideb∗311\nw+/tildewideb∗322\nw)/parenrightbigg\neB, (20)\nwhere the tilde and starred tilde coefficients are given by de finitions ( 7) and ( 11). The/tildewidecjj\nwand/tildewidec∗jj\nwtilde\ncoefficients with jtaking values of 1 or 2 are the sum of the two quantities given i n definitions ( 8)\nand ( 12),\n/tildewidecjj\nw= (/tildewidecjj\nw)s+(/tildewidecjj\nw)s¬\n=cjj\nw−2mwa(5)j0j\nw+3m2\nwc(6)j00j\nw−mwa(5)0jj\nw−mwm(5)jj\nw+3m2\nwc(6)00jj\nw+3m2\nwe(6)0jj\nw ,\n/tildewidec∗jj\nw= (/tildewidec∗jj\nw)s+(/tildewidec∗jj\nw)s¬\n=cjj\nw+2mwa(5)j0j\nw+3m2\nwc(6)j00j\nw+mwa(5)0jj\nw−mwm(5)jj\nw+3m2\nwc(6)00jj\nw−3m2\nwe(6)0jj\nw . (21)Version September 17, 2021 submitted to Symmetry 7 of 20\nFor the shifts in the anomaly frequencies, the results are fo und to be [ 21]\nδωw\na=2/tildewideb3\nw−2/tildewideb33\nF,wB,\nδωw\na=−2/tildewideb∗3\nw+2/tildewideb∗33\nF,wB, (22)\nwhere various tilde and starred tilde coefficients are given by definitions ( 7), (11), (8) and ( 12). We\nnote in passing that the shifts in the cyclotron and anomaly f requencies for an antifermion in results\n(20) and ( 22) reveals that they can be obtained from these for the fermion by changing the signs of all\nthe basic coefficients for Lorentz violation that control CP T-odd effects, as might be expected.\nResult ( 20) shows that the cyclotron frequency shifts due to Lorentz an d CPT violation for a\nfermion are different from these for its corresponding anti fermion. The same conclusion holds for\nthe anomaly frequency shifts from the result ( 22). This implies that in the presence of Lorentz and\nCPT violation, the differences ( 15) and ( 17) do not vanish in general. For the charge-to-mass ratio\ncomparisons, the result becomes\n(|q|/m)w\n(|q|/m)w−1←→ωw\nc\nωwc−1=δωw\nc−δωw\nc\nωwc, (23)\nwhere the Lorentz- and CPT-invariant pieces in the measured cyclotron frequencies are exactly\ncanceled by the CPT theorem if the same magnetic field is used. The notation←→ indicates\nthe correspondence between the experimental interpreted c harge-to-mass ratio comparison and the\nmeasured frequency difference ωwc/ωwc−1. For the gfactor comparison between a fermion and\nantifermion, the related ratio comes to be\n1\n2(gw−gw)←→ωw\na\nωwc−ωw\na\nωwc=δωw\na\nωwc−δωw\na\nωwc, (24)\nwhere again all Lorentz- and CPT-invariant contributions a re canceled out on the right side. We\nnote that in the above expression ( 24), we only keep shifts in the anomaly frequencies δωw\naand δωw\na\ndue to Lorentz and CPT violation, as contributions from the c yclotron frequencies δωwcand δωwcare\nsuppressed by factors of eB/m2\nw, as evident from results ( 20) and ( 22). Even for a comparatively large\nmagnetic field of B≈5 T in a Penning trap, these factors are at orders of eB/m2\ne≈10−9for electrons\nor positrons, and eB/m2\np≈10−16for protons or antiprotons, which can be ignored in Equation (24).\n2.4. Sidereal variations\nThe cyclotron and anomaly frequency shifts ( 20) and ( 22), which also appear in comparisons ( 23)\nand ( 24), are obtained in a particular apparatus frame xa≡(x1,x2,x3), where the positive ˆx3axis\nis aligned with the applied magnetic field in the trap. Howeve r, as Earth rotates about its axis, this\napparatus frame is not inertial. The standard canonical fra me that is adopted in the literature to\ncompare results from different experiments searching for L orentz and CPT violation is called the\nSun-centered frame XJ≡(X,Y,Z)[36,37]. In this frame, the Zaxis is defined to be aligned along\nthe Earth’s rotation axis, the Xaxis points from the Earth to the Sun, and the Yaxis completes a\nright-handed coordinate system. The time origin of this coo rdinate system is chosen to be at the\nvernal equinox 2000. The coefficients for Lorentz violation in this frame are assumed to be constants\nin time and space [ 4,8].\nTo explicitly express the relationship of the coefficients f or Lorentz violation between the\nSun-centered frame XJ≡(X,Y,Z)and the apparatus frame xa≡(x1,x2,x3), it is convenient to\nintroduce a third frame called the standard laboratory fram exj≡(x,y,z), with the zaxis pointing\nto the local zenith, the xaxis aligned with the local south, and the yaxis completing a right-handed\ncoordinate system. To relate the coordinates of these three frames, we define two rotation matrices RajVersion September 17, 2021 submitted to Symmetry 8 of 20\nand RjJ[36,37], with Rajconnecting xj≡(x,y,z)toxa≡(x1,x2,x3)byxa=Rajxj, and RjJrelating\nXJ≡(X,Y,Z)to(x,y,z)byxj=RjJXJ. The expressions for these two rotation matrices are given b y\nRjJ=\ncosχcosω⊕T⊕cosχsinω⊕T⊕−sinχ\n−sinω⊕T⊕ cosω⊕T⊕ 0\nsinχcosω⊕T⊕sinχsinω⊕T⊕ cosχ\n, (25)\nand\nRaj=\ncosγ sinγ0\n−sinγcosγ0\n0 0 1\n×\ncosβ0−sinβ\n0 1 0\nsinβ0 cos β\n×\ncosα sinα0\n−sinαcosα0\n0 0 1\n, (26)\nwhere ω⊕≈2π/(23 h 56 min )is the sidereal frequency of the Earth’s rotation, T⊕denotes the\nlocal sidereal time, χspecifies the colatitude of the laboratory, and (α,β,γ)are the Euler angles\nin the convenient “ y-convention\" of the rotation. The coordinates of the appara tus frame and the\nSun-centered frame can then be related by using the followin g expression,\nxa=Rajxj=RajRjJxJ. (27)\nThe relationship between the coefficients for Lorentz viola tion in these two frames can also be derived\nfrom this result. We note that in the case that a vertical upwa rd magnetic field is used in the trap, the\nEuler angles become (α,β,γ) = ( 0, 0, 0)and the rotation matrix Rajreduces to the identity matrix.\nThe above transformation ( 27) shows that in general the coefficients for Lorentz violatio n and\nthus the cyclotron and anomaly frequency shifts ( 20) and ( 22) determined in the apparatus frame\ndepend on the sidereal time and the geometric location of the laboratory. As a result, signals\nobserved in Earth-based experiments, including the above c omparisons ( 23) and ( 24), can oscillate\nat harmonics of the Earth’s sidereal frequency ω⊕, with the amplitudes depending on the laboratory\ncolatitude χ. To explicitly illustrate this, here we consider a Penning- trap experiment located\nat colatitude χand assume the magnetic field is aligned with the zaxis. We give two explicit\nexamples of the transformation of coefficients for Lorentz v iolation using Equation ( 27). We first\nfocus on the single-index laboratory-frame coefficient /tildewideb′3\nw, which appears in the shifts of the cyclotron\nfrequency ( 20). Applying the transformation ( 27) and taking (α,β,γ) = ( 0, 0, 0)imply\n/tildewideb′3\nw=cosω⊕T⊕/tildewideb′X\nwsinχ+sinω⊕T⊕/tildewideb′Y\nwsinχ+/tildewideb′Z\nwcosχ. (28)\nThis result relates the tilde coefficients /tildewideb′3\nwobserved in the noninertial apparatus frame to the constant\ncoefficients /tildewidebJ\nwwith J=X,Y,Zin the canonical inertial Sun-centered frame. The expressi on contains\nterms proportional to the first harmonic in the Earth’s sider eal frequency and a constant. The\ncolatitude dependence is evident from the factors sin χand cos χappearing above.\nA slightly more complicated result can also be obtained for t he sum of the two-index tilde\nquantities /tildewidec11\nw+/tildewidec22\nwin an analogous fashion. Applying the rotation ( 27) for each index with Rajbeing\nthe identity matrix yields\n/tildewidec11\nw+/tildewidec22\nw=cos 2 ω⊕T⊕/parenleftbig−1\n2(/tildewidecXX\nw−/tildewidecYY\nw)sin2χ/parenrightbig+sin 2 ω⊕T⊕(−/tildewidec(XY)\nw sin2χ)\n+cosω⊕T⊕(−/tildewidec(XZ)\nw sin 2 χ)+sinω⊕T⊕(−/tildewidec(YZ)\nw sin 2 χ)\n+1\n4(/tildewidecXX\nw+/tildewidecYY\nw)(3+cos 2 χ)+/tildewidecZZ\nwsin2χ, (29)\nwhere a parenthes on two indices of the coefficients indicate s symmetrization with a factor of 1/2. For\nexample, /tildewidec(XY)\nw= (/tildewidecXY\nw+/tildewidecYX\nw)/2. It is revealed from the result ( 29) that the sum of the tilde coefficients\n/tildewidec11\nw+/tildewidec22\nwin the apparatus frame depends on six independent quantitie s/tildewidec(JK)\nw with J,K=X,Y,ZinVersion September 17, 2021 submitted to Symmetry 9 of 20\nthe Sun-centered frame, producing up to second harmonics in the sidereal frequency of the Earth’s\nrotation. For both examples ( 28) and ( 29), if the applied magnetic field points along a generic direct ion,\ntrigonometric functions of the Euler angles ( α,β,γ) appear as well.\nAt the end of this subsection, we note that the revolution of t he Earth about the Sun can generate\nadditional types of time variations for the coefficients for Lorentz violation, such as the boost of the\nEarth relative to the Sun β⊕≈10−4, and the boost of the laboratory due to the Earth’s rotation\nβL≈10−6. However, as studied in the literature [ 27,28,38–40], such boost effects are suppressed by\none or more powers of their boost factors β⊕and βLcompared to these from rotations. Therefore, we\ntreat them as negligible effects in the present work.\n3. Experiments\nIn this section, we analyze existing Penning-trap experime nts that compare the charge-to-mass\nratios and the gfactors between protons, antiprotons, electrons, and posi trons. We use reported\nresults for the comparisons ( 23) and ( 24) from these experiments to constrain the relevant\nSun-centered frame tilde coefficients for Lorentz violatio n.\n3.1. The charge-to-mass ratios\nFor charge-to-mass ratio comparisons between a particle an d its corresponding antiparticle, the\nresult ( 23) implies that the relevant tilde coefficients for Lorentz vi olation are /tildewideb′3\nw,/tildewidec11\nw+/tildewidec22\nw,/tildewideb311\nw+/tildewideb322\nw,\n/tildewideb′∗3\nw,/tildewidec∗11\nw+/tildewidec∗22\nw, and/tildewideb∗311\nw+/tildewideb∗322\nwin the apparatus frame, given in the result ( 20). The transformations\nof these tilde coefficients into the Sun-centered frame depe nd on the field configuration in the\ntrap. For a typical Penning-trap experiment, the magnetic fi eld is oriented either horizontally or\nvertically. Ref. [ 22] presents a complete list of transformation results for the se tilde coefficients for\nthe above two field orientations. Here, for completeness, we include them in Appendix A. From\nthese transformation results, it shows that for a given ferm ion of species w, the relevant quantities\nin the Sun-centered frame that are related to the charge-to- mass comparisons in a Penning trap\nare the following 54 independent tilde coefficients, /tildewideb′J\nw,/tildewidec(JK)\nw,/tildewidebJ(KL)\nw ,/tildewideb′∗J\nw,/tildewidec∗(JK)\nw , and/tildewideb∗J(KL)\nw . In the\nfollowing subsections, we apply the reported precisions fo r the charge-to-mass ratio comparisons\nfrom Penning-trap experiments to set bounds on the relevant tilde coefficients for Lorentz violation.\n3.1.1. The proton sector\nWe start the analysis with the charge-to-mass comparisons b etween protons and antiprotons.\nIn a Penning-trap experiment located at CERN by the ATRAP col laboration, Gabrielse and his\ngroup achieved a precision of 90 ppt for the proton-antiprot on charge-to-mass ratio comparison [ 41].\nThe experiment used a trap with a vertical uniform magnetic fi eldB=5.85 T. Recently, another\nPenning-trap experiment at CERN by the BASE collaboration l ed by Ulmer improved the comparison\nto the record sensitivity of 69 ppt [ 42], by applying a horizontal magnetic field B=1.946 T which\nis oriented 60◦east of north. For the measurement of the charge-to-mass rat io of a proton, both\nexperiments used a trapped hydrogen ion (H−) as a proxy for the proton to eliminate systematic\nshifts caused by polarity switching of the trapping voltage s. The charge-to-mass ratio comparison\nbetween an antiproton and a proton is then related to that bet ween an antiproton and a hydrogen ion\nby\n(|q|/m)¯p\n(|q|/m)p−1=(|q|/m)¯p\nR(|q|/m)H−−1←→δω¯p\nc−RδωH−\nc\nRωH−\nc, (30)\nwhere R=mH−/mp=1.001089218754 is the ratio of the mass between a hydrogen io n\nand a proton [ 42], and ωH−\ncis the cyclotron frequency for the hydrogen ion, with δωH−\ncbeing\nthe corresponding shifts. To obtain δωH−\nc, one can take w=H−in the expression ( 20) and\nthe related tilde coefficients for Lorentz violation become the effective ones for a hydrogen ion.\nExpressing these effective coefficients in terms of the corr esponding fundamental coefficients forVersion September 17, 2021 submitted to Symmetry 10 of 20\nthe hydrogen ion constituents, which are the coefficients fo r electrons and protons, is challenging\ndue to nonperturbative issues including binding effects in the composite hydrogen ion. However,\nan approximation to these coefficient relations can be obtai ned by treating the wave function of the\nhydrogen ion as a product of the wave functions of a proton and two electrons. Applying perturbation\ntheory at the lowest order and ignoring the related binding e nergies, the cyclotron frequency shifts\nδωH−\ncof the hydrogen ion due to Lorentz and CPT violation can then b e approximated as the sum of\nthese for its constituents, δωH−\nc≈δωp\nc+2δωe−\nc. Substituting this into the result ( 30) yields\n(|q|/m)¯p\n(|q|/m)p−1←→δω¯p\nc−Rδωp\nc−2Rδωe−\nc\nRωH−\nc. (31)\nAs shown from the above result, the choice of using a hydrogen ion as a proxy for the proton in\nthe Penning trap provides sensitivities not only to the coef ficients for Lorentz violation in the proton\nsector, but also introduces additional sensitivities to th ese for electrons. Putting the above discussion\ntogether, the related Sun-centered frame tilde coefficient s for Lorentz violation that are sensitive to\nPenning-trap experiments comparing the charge-to-mass ra tios between protons and antiprotons are\nthe following 81 independent tilde quantities, /tildewideb′J\np,/tildewidec(JK)\np,/tildewidebJ(KL)\np ,/tildewideb′∗J\np,/tildewidec∗(JK)\np ,/tildewideb∗J(KL)\np ,/tildewideb′J\ne,/tildewidec(JK)\ne, and/tildewidebJ(KL)\ne .\nThe published results for the comparison ( 23) from both the ATRAP and the BASE experiments can\nbe adopted to set bounds on these tilde coefficients for Loren tz violation.\nFor the ATRAP experiment, the reported precision was obtain ed by analyzing the measurements\nof the cyclotron frequencies in a time-averaged way, so any o scillating effects in the difference ( 31)\naveraged out. This implies that only the constant terms that appear in the transformation results in the\nfirst half of Appendix Acan be constrained using the published precision. Applying expression ( 31)\nby taking the reported precision of 90 ppt for (|q|/m)¯p/(|q|/m)p−1 and identifying ωH−\nc=2π×89.3\nMHz given in the ATRAP experiment, the following limit can be obtained,\n|δω¯p\nc−1.001 δωp\nc−2.002 δωe−\nc|const∼<3.33×10−26GeV, (32)\nwhere the subscript “const\" indicates that only the constan t terms in the transformations results are\nrelevant to the limit. However, future sidereal-variation analysis of the measurements of the cyclotron\nfrequencies can provide constraints on the non-constant te rms that appear in the harmonics in the\ntransformation results.\nFor the experiment carried out by the BASE collaboration, si nce the magnetic field was oriented\nat 60◦east of north, both of the transformation matrices ( 25) and ( 26) are required to relate the\ntilde coefficients for Lorentz violation in the apparatus fr ame to these in the Sun-centered frame.\nFor a general horizontal magnetic field with an angle θmeasured from the local south in the\ncounterclockwise direction, the corresponding Euler angl es are found to be (α,β,γ) = ( θ,π/2, 0).\nThe second half of Appendix Alists the full expressions of the transformations for the re lated tilde\ncoefficients for Lorentz violation. The BASE experiment ana lyzed the data of the charge-to-mass ratio\ncomparisons to search for both time-averaged effects and si dereal variations in the first harmonics of\nthe Earth’s rotation frequency. Therefore, the reported re sults can be taken to set bounds on not only\nthe constant terms but also on the terms proportional to the fi rst harmonics in the transformation\nresults given in Appendix A. Using the reported 69 ppt for the time-averaged precision a nd 720 ppt\nfor the limit of the first harmonic amplitude for the comparis on (31), and taking ωH−\nc=2π×29.635\nMHz for the BASE experiment, the following limits are obtain ed,\n|δω¯p\nc−1.001 δωp\nc−2.002 δωe−\nc|const∼<8.46×10−27GeV (33)\nand\n|δω¯p\nc−1.001 δωp\nc−2.002 δωe−\nc|1st∼<8.83×10−26GeV, (34)Version September 17, 2021 submitted to Symmetry 11 of 20\nwhere the subscript “const\" in the limit ( 33) has similar meaning as the one in ( 32), while the subscript\n“1st\" in the limit ( 34) specifies only the amplitude of the first harmonics in the sid ereal variation.\nThe limit ( 32) for the ATRAP experiment and limits ( 33) and ( 34) for the BASE experiment set\nconstraints on a combination of the Sun-centered frame tild e coefficients /tildewideb′J\np,/tildewidec(JK)\np,/tildewidebJ(KL)\np ,/tildewideb′∗J\np,/tildewidec∗(JK)\np ,\n/tildewideb∗J(KL)\np ,/tildewideb′J\ne,/tildewidec(JK)\ne, and/tildewidebJ(KL)\ne . These constraints can be obtained by substituting result ( 20) in limit ( 32)\nand applying the corresponding transformations given in Ap pendix Awith χ=43.8◦for the ATRAP\nexperiment, and substituting result ( 20) in both limits ( 33) and ( 34) and identifying θ=2π/3 and\nχ=43.8◦in the transformation expressions in Appendix Afor the BASE experiment. To get some\nintuition about the scope of the constraints on the individu al component of the tilde coefficients,\nwe can take a common practice that is adopted in many subfields searching for Lorentz and CPT\nviolation [ 7] which assumes that only one individual tilde coefficient is nonzero at a time. From the\nATRAP limit ( 32), constraints on 27 independent tilde coefficients for Lore ntz violation are obtained,\nwhile from the BASE limits ( 33) and ( 34), a total of 69 independent tilde coefficients for Lorentz\nviolation are constrained. We summarize in Table 1and Table 2the constraints in the proton sector\nand the electron sector, respectively. In both tables, the fi rst column lists the individual components,\nthe second column presents the corresponding constraint on the modulus of the component, the third\ncolumn specifies the related experiment, and the related ref erence is given in the final column. We\nnote that some of the tilde coefficients for Lorentz violatio n are constrained by both the ATRAP and\nthe BASE experiments. To keep the results clean, we only keep the more stringent ones in both tables.\nTable 1. Constraints on tilde coefficients for Lorentz violation in t he proton sector using the\ncharge-to-mass ratio comparisons from the ATRAP and the BAS E experiments.\nCoefficient Constraint Experiment Reference\n|/tildewideb′Zp|,|/tildewideb′∗Zp| <1×10−10GeV ATRAP [ 41]\n|/tildewidecXXp|,|/tildewidec∗XXp| <1×10−10ATRAP [ 41]\n|/tildewidecYYp|,|/tildewidec∗YYp| <1×10−10ATRAP [ 41]\n|/tildewidecZZp|,|/tildewidec∗ZZp| <8×10−11BASE [ 42]\n|/tildewidebX(XZ)\np|,|/tildewideb∗X(XZ)\np|<2×10−10GeV−1BASE [ 42]\n|/tildewidebY(YZ)\np|,|/tildewideb∗Y(YZ)\np|<2×10−10GeV−1BASE [ 42]\n|/tildewidebZZZp|,|/tildewideb∗ZZZp|<2×10−10GeV−1BASE [ 42]\n|/tildewidebZXXp|,|/tildewideb∗ZXXp|<2×10−10GeV−1ATRAP [ 41]\n|/tildewidebZYYp|,|/tildewideb∗ZYYp|<2×10−10GeV−1ATRAP [ 41]\n|/tildewideb′Xp|,|/tildewideb′∗Xp| <7×10−10GeV BASE [ 42]\n|/tildewideb′Yp|,|/tildewideb′∗Yp| <7×10−10GeV BASE [ 42]\n|/tildewidec(XZ)\np|,|/tildewidec∗(XZ)\np|<1×10−9BASE [ 42]\n|/tildewidec(YZ)\np|,|/tildewidec∗(YZ)\np|<1×10−9BASE [ 42]\n|/tildewidebXXXp|,|/tildewideb∗XXXp|<2×10−9GeV−1BASE [ 42]\n|/tildewidebX(XY)\np|,|/tildewideb∗X(XY)\np|<2×10−9GeV−1BASE [ 42]\n|/tildewidebXYYp|,|/tildewideb∗XYYp|<1×10−9GeV−1BASE [ 42]\n|/tildewidebXZZp|,|/tildewideb∗XZZp|<9×10−10GeV−1BASE [ 42]\n|/tildewidebYXXp|,|/tildewideb∗YXXp|<1×10−9GeV−1BASE [ 42]\n|/tildewidebY(XY)\np|,|/tildewideb∗Y(XY)\np|<2×10−9GeV−1BASE [ 42]\n|/tildewidebYYYp|,|/tildewideb∗YYYp|<2×10−9GeV−1BASE [ 42]\n|/tildewidebYZZp|,|/tildewideb∗YZZp|<9×10−10GeV−1BASE [ 42]\n|/tildewidebZ(XZ)\np|,|/tildewideb∗Z(XZ)\np|<3×10−9GeV−1BASE [ 42]\n|/tildewidebZ(YZ)\np|,|/tildewideb∗Z(YZ)\np|<3×10−9GeV−1BASE [ 42]Version September 17, 2021 submitted to Symmetry 12 of 20\nTable 2. Constraints on tilde coefficients for Lorentz violation in t he electron sector using the\ncharge-to-mass ratio comparisons from the ATRAP and the BAS E experiments.\nCoefficient Constraint Experiment Reference\n|/tildewideb′Ze| <2×10−17GeV ATRAP [ 41]\n|/tildewidecXXe|<3×10−14ATRAP [ 41]\n|/tildewidecYYe| <3×10−14ATRAP [ 41]\n|/tildewidecZZe| <2×10−14BASE [ 42]\n|/tildewidebX(XZ)\ne|<1×10−10GeV−1BASE [ 42]\n|/tildewidebY(YZ)\ne|<1×10−10GeV−1BASE [ 42]\n|/tildewidebZZZe|<1×10−10GeV−1BASE [ 42]\n|/tildewidebZXXe|<9×10−11GeV−1ATRAP [ 41]\n|/tildewidebZYYe|<9×10−11GeV−1ATRAP [ 41]\n|/tildewideb′Xe| <1×10−16GeV BASE [ 42]\n|/tildewideb′Ye| <1×10−16GeV BASE [ 42]\n|/tildewidec(XZ)\ne|<3×10−13BASE [ 42]\n|/tildewidec(YZ)\ne|<3×10−13BASE [ 42]\n|/tildewidebXXXe|<1×10−9GeV−1BASE [ 42]\n|/tildewidebX(XY)\ne|<9×10−10GeV−1BASE [ 42]\n|/tildewidebXYYe|<5×10−10GeV−1BASE [ 42]\n|/tildewidebXZZe|<5×10−10GeV−1BASE [ 42]\n|/tildewidebYXXe|<5×10−10GeV−1BASE [ 42]\n|/tildewidebY(XY)\ne|<9×10−10GeV−1BASE [ 42]\n|/tildewidebYYYe|<1×10−9GeV−1BASE [ 42]\n|/tildewidebYZZe|<5×10−10GeV−1BASE [ 42]\n|/tildewidebZ(XZ)\ne|<2×10−9GeV−1BASE [ 42]\n|/tildewidebZ(YZ)\ne|<2×10−9GeV−1BASE [ 42]\n3.1.2. The electron sector\nFor the comparison of the charge-to-mass ratios between an e lectron and a positron, the current\nmost accurate result was made in an experiment at the Univers ity of Washington, with a precision at\n130 ppb [ 43]. From the comparison ( 23), this time-average result gives the following limit,\n|δωe−\nc−δωe+\nc|const∼<7.66×10−20GeV. (35)\nFollowing a similar analysis as the one used for the proton se ctor in the previous subsection, one can\nuse the colatitude and magnetic field that are relevant to thi s experiment χ=42.5◦and B=5.1 T\nupward, together with the transformation expressions give n in Appendix A, to obtain the constraints\non the tilde coefficients for Lorentz violation /tildewideb′J\ne,/tildewidec(JK)\ne,/tildewidebJ(KL)\ne ,/tildewideb′∗J\ne,/tildewidec∗(JK)\ne , and/tildewideb∗J(KL)\ne . We summarize\nthe results in Table 3, which is organized in the same way as Table 1and Table 2. We note that some of\nthe constraints of the tilde coefficients from the electron- positron charge-to-mass ratio comparison are\nnot comparable to these given in Table 2from the proton-antiproton comparison, so we don’t include\nthem in Table 3. We also note that the Penning trap experiment at the Univers ity of Washington used\na radioactive source of positrons that requires special pre cautions. Efforts in applying a safe source\nand ensuring efficient positron accumulation are currently being made at both Harvard University\nand Northwestern University [ 44,45], providing great potential for improving the current boun ds for\nthe tilde coefficients listed in Table 2and Table 3.Version September 17, 2021 submitted to Symmetry 13 of 20\nTable 3. Constraints on tilde coefficients for Lorentz violation in t he electron sector using the\ncharge-to-mass ratio comparisons from the experiment at th e University of Washington.\nCoefficient Constraint Experiment Reference\n|/tildewideb′∗Ze|<9×10−11GeV Washington [ 43]\n|/tildewidec∗XXe|<2×10−7Washington [ 43]\n|/tildewidec∗YYe|<2×10−7Washington [ 43]\n|/tildewidec∗ZZe|<3×10−7Washington [ 43]\n|/tildewideb∗X(XZ)\ne|<8×10−4GeV−1Washington [ 43]\n|/tildewideb∗Y(YZ)\ne|<8×10−4GeV−1Washington [ 43]\n|/tildewideb∗ZZZe|<8×10−4GeV−1Washington [ 43]\n|/tildewideb∗ZXXe|<4×10−4GeV−1Washington [ 43]\n|/tildewideb∗ZYYe|<4×10−4GeV−1Washington [ 43]\n3.2. The g factors and magnetic moments\nFor the gfactor and magnetic moment comparisons between particles a nd antiparticles using\nPenning traps, the result ( 24) implies that the relevant coefficients for Lorentz violati on in the\napparatus frame are these given in Equation ( 22),/tildewideb3\nw,/tildewideb∗3\nw,/tildewideb33\nF,w, and/tildewideb∗33\nF,w. For a Penning trap\nwith a vertical or horizontal magnetic field, the expression s of the transformation for these tilde\ncoefficients into the Sun-centered frame are also given in Ap pendix A. The results show that there\nare 18 independent components of the tilde coefficients for e ach fermion species w, given by /tildewidebJ\nw,/tildewideb∗J\nw,\n/tildewideb(JK)\nF,w, and/tildewideb∗(JK)\nF,w. In the subsection follows, we list the related Penning-tra p experiments and use\nthe reported comparison results to set bounds on the corresp onding tilde coefficients for Lorentz\nviolation.\n3.2.1. The proton sector\nIn the proton sector, the current best measurements of the ma gnetic moments for both a\nproton and an antiproton were achieved by the BASE collabora tion. The proton magnetic moment\nmeasurement has a sensitivity of 0.3 ppb using a Penning trap located at Mainz [ 46], improving their\nprevious best measurement [ 47] by a factor of 11. The antiproton magnetic moment measureme nt\nwas measured to a precision of 1.5 ppb with a similar Penning t rap located at CERN [ 48]. Combining\nthe reported 0.3 ppb and 1.5 ppb precisions for the time-aver aged measurements, and identifying\nωp\nc=2π×28.96 MHz and ωp\nc=2π×29.66 MHz for each experiment, comparison ( 24) yields\n|δωp\na−0.98δωp\na|const∼<9.53×10−25GeV, (36)\nwhere same subscript “const\" is used to specify only the cons tant terms in the transformation are\nrelevant to the above limit.\nThe corresponding constraints for the combinations of the t ilde coefficients in the Sun-centered\nframe can be obtained by applying the corresponding transfo rmation results given in Appendix A\nthat are related to the specific field configuration in the trap and substituting the numerical values\nof the laboratory quantities for both experiments in limit ( 36). For the proton magnetic moment\nmeasurement at Mainz, the colatitude is χ=40.0◦and the magnetic field B=1.9 T points θ=18◦\nfrom local south in the counterclockwise direction [ 46]. For the experiment measuring the antiproton\nmagnetic moment at CERN, the trap is located at a slight diffe rent colatitude χ=43.8◦and the\nmagnetic field B=1.95 T points θ=120◦in the same convention as above [ 48]. Adopting the same\nassumption that only one tilde coefficient is nonzero at a tim e, the constraint on each independent\ntilde coefficient can be obtained. We summarize them in Table 4in the same fashion as before.Version September 17, 2021 submitted to Symmetry 14 of 20\nTable 4. Constraints on tilde coefficients for Lorentz violation in t he proton sector using the gfactor\ncomparison from the BASE experiments at Mainz and CERN.\nCoefficient Constraint Experiment Reference\n|/tildewidebZp| <8×10−25GeV BASE [ 46,48]\n|/tildewideb∗Zp| <1×10−24GeV BASE [ 46,48]\n|/tildewidebXX\nF,p+/tildewidebYY\nF,p|<4×10−9GeV−1BASE [ 46,48]\n|/tildewidebZZ\nF,p| <3×10−9GeV−1BASE [ 46,48]\n|/tildewideb∗XX\nF,p+/tildewideb∗YY\nF,p|<3×10−9GeV−1BASE [ 46,48]\n|/tildewideb∗ZZ\nF,p| <1×10−8GeV−1BASE [ 46,48]\nAt the end of this subsection, we point out that a sidereal-va riation analysis of the anomaly\nfrequencies for both protons and antiprotons is currently b eing carried out by the BASE collaboration.\nThis could in principle set bounds on other components of the tilde coefficients /tildewidebJ\np,/tildewideb(JK)\nF,p,/tildewideb∗J\npand\n/tildewideb∗(JK)\nF,pthat haven’t been constrained before. The BASE collaborati on is also developing a quantum\nlogic readout system to allow more rapid measurements of the anomaly frequencies in the trap [ 49],\nwhich will offer excellent opportunities to perform a sider eal-variation analysis of the experimental\ndata, with great potential to set more stringent limits on th e tilde coefficients for Lorentz violation.\n3.2.2. The electron sector\nIn the electron sector, the comparison of the anomaly freque ncies between electrons and\npositrons were performed in a Penning-trap experiment at th e University of Washington, with a\nprecision of about 2 ppt [ 50]. The experimental data were analyzed in a time-averaged wa y to obtain\na constraint of b∼<50 rad/s using the notation given in Ref. [ 50]. Translating to the notation in this\nwork yields\n|δωe−\na−δωe+\na|const∼<2.09×10−23GeV. (37)\nTaking the experimental quantities χ=42.5◦and B=5.85 T upward, as well as the transformation\npresented in Appendix A, the limit ( 37) can be converted to constraints on the following Sun-cente red\nframe tilde coefficients, /tildewidebZ\ne,/tildewideb∗Z\ne,/tildewidebXX\nF,e+/tildewidebYY\nF,e,/tildewideb∗XX\nF,e+/tildewideb∗YY\nF,e,/tildewidebZZ\nF,e, and/tildewideb∗ZZ\nF,e.\nThe measurement of the gfactor for electrons has reached a record precision of 0.28 p pt at\nHarvard University [ 1]. A sidereal-variation analysis of the anomaly frequencie s was performed\nat the frequencies of ω⊕and 2 ω⊕, yielding the same limit on the amplitudes of both the first an d\nthe second harmonics in the oscillation, |δωea|1st/2nd∼<2π×0.05 Hz [ 24]. Converting the results in\nnatural units gives the following limits,\n|δωe−\na|1st∼<2.07×10−25GeV, (38)\nand\n|δωe−\na|2nd∼<2.07×10−25GeV. (39)\nTaking the magnetic field adopted in the experiment as B=5.36 T in the local upward direction\nwith the geometrical colatitude χ=47.6◦and applying the transformation results in Appendix A,\nconstraints on the following additional components of the t ilde coefficients, /tildewidebX\ne,/tildewidebY\ne,/tildewideb(XY)\nF,e,/tildewideb(XZ)\nF,e,/tildewideb(YZ)\nF,e,\nand/tildewidebXX\nF,e−/tildewidebYY\nF,eare obtained. The results from the University of Washington and Harvard University\nare summarized in Table 5.Version September 17, 2021 submitted to Symmetry 15 of 20\nTable 5. Constraints on tilde coefficients for Lorentz violation in t he electron sector using the gfactor\nmeasurements from the experiments at the University of Wash ington and Harvard University.\nCoefficient Constraint Experiment Reference\n|/tildewidebXe| <1×10−25GeV Harvard [ 24]\n|/tildewidebYe| <1×10−25GeV Harvard [ 24]\n|/tildewidebZe| <7×10−24GeV Washington [ 50]\n|/tildewideb∗Ze| <7×10−24GeV Washington [ 50]\n|/tildewidebXX\nF,e+/tildewidebYY\nF,e|<2×10−8GeV−1Washington [ 50]\n|/tildewidebZZ\nF,e| <8×10−9GeV−1Washington [ 50]\n|/tildewideb(XY)\nF,e| <2×10−10GeV−1Harvard [ 24]\n|/tildewideb(XZ)\nF,e| <1×10−10GeV−1Harvard [ 24]\n|/tildewideb(YZ)\nF,e| <1×10−10GeV−1Harvard [ 24]\n|/tildewideb∗XX\nF,e+/tildewideb∗YY\nF,e|<2×10−8GeV−1Washington [ 50]\n|/tildewideb∗XX\nF,e−/tildewideb∗YY\nF,e|<4×10−10GeV−1Harvard [ 24]\n|/tildewideb∗ZZ\nF,e| <8×10−9GeV−1Washington [ 50]\nThe measurement of the magnetic moment for positrons are cur rently under development at\nboth Harvard University and Northwestern University [ 44,45]. A comparison with the results for\nthat of electrons would offer opportunities to extract the c oefficients that control only CPT-odd effects\nfrom the combination of coefficients in the difference ( 24). A sidereal-variation analysis of the positron\nanomaly frequencies would offer limits on the starred tilde coefficients /tildewideb∗J\ne,/tildewideb∗(JK)\nF,eas well.\nAt the end of this subsection, we note that from Table 4and Table 5, for the 18 independent\ncomponents of the tilde coefficients for Lorentz violation t hat are relevant to the gfactor\nmeasurements in Penning-trap experiments, only 6 of them in the proton sector and 12 of them in\nthe electron sector have been constrained so far. Performin g a full sidereal-variation analysis for the\nmeasurements data would permit access to the other componen ts of the tilde coefficients.\n4. Summary\nIn this work, we provide an overview of recent progress on sea rching for Lorentz- and\nCPT-violating signals using measurements of charge-to-ma ss ratios and magnetic moments in\nPenning-trap experiments. We first revisit the theory of Lor entz-violating quantum electrodynamics\nwith operators of mass dimensions d≤6. The explicit expressions of the Lagrange density ( 1)\nare given in Ref. [ 21]. Perturbation theory is then applied to obtain the dominan t energy shifts\n(6) and ( 10) for a confine particle and antiparticle due to Lorentz and CP T violation. This leads\nto the corresponding contributions to the cyclotron and ano maly frequencies in Equations ( 20)\nand ( 22). The results are then used to relate the coefficients for Lor entz violation to the experimental\ninterpreted charge-to-mass ratio comparisons ( 23) between a particle and an antiparticle, as well as\nthe magnetic moment comparisons ( 24). The general transformation of the related coefficients fo r\nLorentz violation into different frames is performed using Equation ( 27). The explicit expressions\nrelating the coefficients in the apparatus frame to the Sun-c entered frame are given in Appendix A.\nThe results show that for the charge-to-mass ratio comparis ons between particles and antiparticles in\nPenning-trap experiments, the related coefficients for Lor entz violation in the Sun-centered frame are\n/tildewideb′J\nw,/tildewidec(JK)\nw,/tildewidebJ(KL)\nw ,/tildewideb′∗J\nw,/tildewidec∗(JK)\nw , and/tildewideb∗J(KL)\nw . For experiments involving magnetic moment comparisons,\nthe corresponding coefficients are /tildewidebJ\nw,/tildewideb∗J\nw,/tildewideb(JK)\nF,w, and/tildewideb∗(JK)\nF,w. Using published results from existing\nPenning-trap experiments, constraints on various compone nts of the coefficients for Lorentz violation\nare obtained. They are summarized in Tables 1-5. In conclusion, the high-precision measurementsVersion September 17, 2021 submitted to Symmetry 16 of 20\nand excellent coverage of the coefficients for Lorentz viola tion offered by Penning-trap experiments\nprovide strong motivations to continue the searches for pos sible Lorentz- and CPT-violating signals.\nAppendix A Transformations\nIn this Appendix, we list the explicit expressions of the tra nsformation results for the tilde\ncoefficients for Lorentz violation that are relevant to the c yclotron and anomaly frequency shifts.\nThese coefficients are /tildewideb′3w,/tildewidec11w+/tildewidec22w,/tildewideb311w+/tildewideb322w,/tildewideb3w,/tildewideb3\nF,w,/tildewideb′∗3w,/tildewidec∗11w+/tildewidec∗22w,/tildewideb∗311w+/tildewideb∗322w,/tildewideb∗3w, and/tildewideb∗3\nF,w\nin the apparatus frame.\nWe first consider Penning-trap experiments that use a vertic al upward magnetic field. Taking\nRajas the identity matrix and applying transformation ( 27) yield\n/tildewideb′3\nw=cosω⊕T⊕/tildewideb′X\nwsinχ+sinω⊕T⊕/tildewideb′Y\nwsinχ+/tildewideb′Z\nwcosχ, (A1)\n/tildewidec11\nw+/tildewidec22\nw=cos 2 ω⊕T⊕/parenleftig\n−1\n2(/tildewidecXX\nw−/tildewidecYY\nw)sin2χ/parenrightig\n+sin 2 ω⊕T⊕/parenleftig\n−/tildewidec(XY)\nw sin2χ/parenrightig\n+cosω⊕T⊕/parenleftig\n−/tildewidec(XZ)\nw sin 2 χ/parenrightig\n+sinω⊕T⊕/parenleftig\n−/tildewidec(YZ)\nw sin 2 χ/parenrightig\n+1\n4(/tildewidecXX\nw+/tildewidecYY\nw)(3+cos 2 χ)+/tildewidecZZ\nwsin2χ, (A2)\n/tildewideb311\nw+/tildewideb322\nw=cos 3 ω⊕T⊕/parenleftig\n[−1\n4(/tildewidebXXX\nw−/tildewidebXYY\nw)+1\n2/tildewidebY(XY)\nw]sin3χ/parenrightig\n+sin 3 ω⊕T⊕/parenleftig\n[−1\n2/tildewidebX(XY)\nw−1\n4(/tildewidebYXX\nw−/tildewidebYYY\nw)]sin3χ/parenrightig\n+cos 2 ω⊕T⊕/parenleftig\n[−/tildewidebX(XZ)\nw+/tildewidebY(YZ)\nw−1\n2(/tildewidebZXX\nw−/tildewidebZYY\nw)]cosχsin2χ/parenrightig\n+sin 2 ω⊕T⊕/parenleftig\n[−/tildewidebX(YZ)\nw−/tildewidebY(XZ)\nw−/tildewidebZ(XY)\nw]cosχsin2χ/parenrightig\n+cosω⊕T⊕/parenleftig\n1\n8/tildewidebXXX\nw(5+3 cos 2 χ)sinχ+1\n8/tildewidebXYY\nw(7+cos 2 χ)sinχ+/tildewidebXZZ\nw sin3χ\n−1\n2/tildewidebY(XY)\nw sin3χ−2/tildewidebZ(XZ)\nw cos2χsinχ/parenrightig\n+sinω⊕T⊕/parenleftig\n−1\n2/tildewidebX(XY)\nw sin3χ+1\n8/tildewidebYXX\nw(7+cos 2 χ)sinχ+1\n8/tildewidebYYY\nw(5+3 cos 2 χ)sinχ\n+/tildewidebYZZ\nwsin3χ−2/tildewidebZ(YZ)\nw cos2χsinχ/parenrightig\n−(/tildewidebX(XZ)\nw+/tildewidebY(YZ)\nw−/tildewidebZZZ\nw)cosχsin2χ+(/tildewidebZXX\nw+/tildewidebZYY\nw)cosχcos 2 χ, (A3)\n/tildewideb3\nw=cosω⊕T⊕/tildewidebX\nwsinχ+sinω⊕T⊕/tildewidebY\nwsinχ+/tildewidebZ\nwcosχ, (A4)\nand\n/tildewideb33\nF,w=cos 2 ω⊕T⊕/parenleftig\n1\n2(/tildewidebXX\nF,w−/tildewidebYY\nF,w)sin2χ/parenrightig\n+sin 2 ω⊕T⊕/tildewideb(XY)\nF,wsin2χ/parenrightig\n+cosω⊕T⊕/tildewideb(XZ)\nF,wsin 2 χ+sinω⊕T⊕/tildewideb(YZ)\nF,wsin 2 χ\n+1\n2(/tildewidebXX\nF,w+/tildewidebYY\nF,w−2/tildewidebZZ\nF,w)sin2χ+/tildewidebZZ\nF,w. (A5)Version September 17, 2021 submitted to Symmetry 17 of 20\nWe then switch to the case where a horizontal magnetic field is used in the trap, directing at\nan angle θfrom the local south in the counterclockwise direction. Thi s implies that the Euler angles\n(α,β,γ) = ( θ,π/2, 0). Applying the transformation ( 27) we have\n/tildewideb′3\nw=cosω⊕T⊕/parenleftig\n/tildewideb′X\nwcosθcosχ+/tildewideb′Y\nwsinθ/parenrightig\n+sinω⊕T⊕/parenleftig\n−/tildewideb′X\nwsinθ+/tildewideb′Y\nwcosθcosχ/parenrightig\n−/tildewideb′Z\nwcosθsinχ, (A6)\n/tildewidec11\nw+/tildewidec22\nw=cos 2 ω⊕T⊕/parenleftig\n1\n8(/tildewidecXX\nw−/tildewidecYY\nw)(1−3 cos 2 θ−2 cos2θcos 2 χ)−/tildewidec(XY)\nw cosχsin 2 θ/parenrightig\n+sin 2 ω⊕T⊕/parenleftig\n1\n2(/tildewidecXX\nw−/tildewidecYY\nw)cosχsin 2 θ+1\n4/tildewidec(XY)\nw(1−3 cos 2 θ−2 cos2θcos 2 χ)/parenrightig\n+cosω⊕T⊕/parenleftig\n/tildewidec(XZ)\nw cos2θsin 2 χ+/tildewidec(YZ)\nw sin 2 θsinχ/parenrightig\n+sinω⊕T⊕/parenleftig\n−/tildewidec(XZ)\nw sin 2 θsinχ+/tildewidec(YZ)\nw cos2θsin 2 χ/parenrightig\n+1\n2(/tildewidecXX\nw+/tildewidecYY\nw)(cos2θ+cos2χsin2θ+sin2χ)+/tildewidecZZ\nw(cos2χ+sin2θsin2χ), (A7)\n/tildewideb311\nw+/tildewideb322\nw\n=cos 3 ω⊕T⊕/parenleftig\n[1\n64(/tildewidebXXX\nw−/tildewidebXYY\nw)−1\n32/tildewidebY(XY)\nw][3(cosθ−5 cos 3 θ)cosχ−4 cos3θcos 3 χ]\n+[1\n16/tildewidebX(XY)\nw+1\n32(/tildewidebYXX\nw−/tildewidebYYY\nw)][3 sin θ(1−4 cos2θcos 2 χ)−5 sin 3 θ]/parenrightig\n+sin 3 ω⊕T⊕/parenleftig\n[−1\n32(/tildewidebXXX\nw−/tildewidebXYY\nw)+1\n16/tildewidebY(XY)\nw][3 sin θ(1−4 cos2θcos 2 χ)−5 sin 3 θ]\n+[1\n32/tildewidebX(XY)\nw+1\n64(/tildewidebYXX\nw−/tildewidebYYY\nw)][3(cosθ−5 cos 3 θ)cosχ−4 cos3θcos 3 χ]/parenrightig\n+cos 2 ω⊕T⊕/parenleftig\n[1\n16/tildewidebX(XZ)\nw−1\n16/tildewidebY(YZ)\nw+1\n32(/tildewidebZXX\nw−/tildewidebZYY\nw)][(5 cos 3 θ−cosθsinχ+4 cos3θsin 3 χ]\n+(/tildewidebX(YZ)\nw+/tildewidebY(XZ)\nw+/tildewidebZ(XY)\nw)sinθcos2θsin 2 χ/parenrightig\n+sin 2 ω⊕T⊕/parenleftig\n[−/tildewidebX(XZ)\nw+/tildewidebY(YZ)\nw−1\n2(/tildewidebZXX\nw−/tildewidebZYY\nw)]sinθcos2θsin 2 χ\n−(1\n16/tildewidebX(YZ)\nw+1\n16/tildewidebY(XZ)\nw+1\n16/tildewidebZ(XY)\nw)[(cosθ−5 cos 3 θ)sinχ−4 cos3θsin 3 χ]/parenrightig\n+cosω⊕T⊕/parenleftig\n1\n16/tildewidebXXX\nw cosθcosχ(−6 cos2θcos 2 χ+3 cos 2 θ+7)\n−1\n2/tildewidebX(XY)\nw sinθ(cos2θcos 2 χ+sin2θcos2χ+sin2χ)\n+1\n16/tildewidebXYY\nwcosθcosχ(−2 cos2θcos 2 χ+cos 2 θ+13)\n+/tildewidebXZZ\nw cosθcosχ(sin2θsin2χ+cos2χ)\n+1\n32/tildewidebYXX\nw[sinθ(25−4 cos2θcos 2 χ)+sin 3 θ]\n+1\n16/tildewidebY(XY)\nw cosθ[(cos 2 θ−7)cosχ−2 cos2θcos 3 χ]\n+1\n32/tildewidebYYY\nw[sinθ(11−12 cos2θcos 2 χ)+3 sin 3 θ]\n+/tildewidebYZZ\nwsinθ(sin2θsin2χ+cos2χ)\n−2/tildewidebZ(XZ)\nw cos3θsin2χcosχ−2/tildewidebZ(YZ)\nw sinθcos2θsin2χ/parenrightig\n+sinω⊕T⊕/parenleftig\n1\n32/tildewidebXXX\nw[sinθ(12 cos2θcos 2 χ−11)−3 sin 3 θ]\n+1\n16/tildewidebX(XY)\nw cosθ[(cos 2 θ−7)cosχ−2 cos2θcos 3 χ]\n+1\n32/tildewidebXYY\nw[sinθ(4 cos2θcos 2 χ−25)−sin 3 θ]Version September 17, 2021 submitted to Symmetry 18 of 20\n−/tildewidebXZZ\nw sinθ(sin2θsin2χ+cos2χ)\n+1\n16/tildewidebYXX\nw cosθcosχ(−2 cos2θcos 2 χ+cos 2 θ+13)\n+1\n2/tildewidebY(XY)\nw sinθ(cos2θcos 2 χ+sin2θcos2χ+sin2χ)\n+1\n16/tildewidebYYY\nwcosθcosχ(−6 cos2θcos 2 χ+3 cos 2 θ+7)\n+/tildewidebYZZ\nwcosθcosχ(sin2θsin2χ+cos2χ)\n+2/tildewidebZ(XZ)\nw sinθcos2θsin2χ−2/tildewidebZ(YZ)\nw cos3θsin2χcosχ/parenrightig\n+1\n8(/tildewidebX(XZ)\nw+/tildewidebY(YZ)\nw−/tildewidebZZZ\nw)cosθ(−4 cos 2 θsin3χ+5 sin χ+sin 3 χ)\n+1\n16(/tildewidebZXX\nw+/tildewidebZYY\nw)[2 cos3θsin 3 χ−cosθsinχ(3 cos 2 θ+11)], (A8)\n/tildewideb3\nw=cosω⊕T⊕/parenleftig\n/tildewidebX\nwcosθcosχ+/tildewidebY\nwsinθ/parenrightig\n+sinω⊕T⊕/parenleftig\n−/tildewidebX\nwsinθ+/tildewidebY\nwcosθcosχ/parenrightig\n−/tildewidebZ\nwcosθsinχ, (A9)\nand\n/tildewideb33\nF,w=cos 2 ω⊕T⊕/parenleftig\n1\n2(/tildewidebXX\nF,w−/tildewidebYY\nF,w)(cos2θcos2χ−sin2θ)+/tildewideb(XY)\nF,wsin 2 θcosχ/parenrightig\n+sin 2 ω⊕T⊕/parenleftig\n1\n4/tildewideb(XY)\nF,w(2 cos2θcos 2 χ+3 cos 2 θ−1)−(/tildewidebXX\nF,w−/tildewidebYY\nF,w)sin 2 θcosχ/parenrightig\n+cosω⊕T⊕/parenleftig\n−2/tildewideb(XZ)\nF,wcos2θcosχ−/tildewideb(YZ)\nF,wsin 2 θsinχ/parenrightig\n+sinω⊕T⊕/parenleftig\n/tildewideb(XZ)\nF,wsin 2 θsinχ−2/tildewideb(YZ)\nF,wcos2θcosχ/parenrightig\n+1\n2(/tildewidebXX\nF,w+/tildewidebYY\nF,w)(cos2θcos2χ+sin2θ)+/tildewidebZZ\nF,wcos2θsin2χ. (A10)\nThe corresponding transformations for the starred tilde qu antities/tildewideb′∗3\nw,/tildewidec∗11\nw+/tildewidec∗22\nw,/tildewideb∗311\nw+/tildewideb∗322\nw,\n/tildewideb∗3w, and/tildewideb33\nF,wtake the same forms as these given above, with the substituti ons of the usual tilde\ncoefficients by the starred ones.\nFunding: This work was supported in part by the Gettysburg College Res earch and Professional Development\nGrant, the Gettysburg College Cross-Disciplinary Science Institute (X-SIG), and the Keck Grant from the W.M.\nKeck Science Department at Claremont McKenna, Pitzer, and S cripps Colleges.\nAcknowledgments: The authors would like to thank Matthew Mewes for the invitat ion and Lisa Portmess for\nreading the manuscript.\nConflicts of Interest: The authors declare no conflict of interest.\nReferences\n1. Hanneke, D.; Fogwell, S.; Gabrielse, G. New measurement o f the electron magnetic moment and the fine\nstructure constant. Phys. Rev. Lett. 2008 ,100, 120801.\n2. Kostelecký, V .A.; Samuel, S. Spontaneous breaking of Lor entz symmetry in string theory. Phys. Rev. D 1989 ,\n39, 683.\n3. Kostelecký, V .A.; Potting, R. CPT and strings. Nucl. Phys. B 1991 359, 545.\n4. Colladay, D.; Kostelecký, V .A. CPT violation and the Stan dard Model. Phys. Rev. D 1997 ,55, 6760.\n5. Colladay, D.; Kostelecký, V .A. Lorentz-violating exten sion of the Standard Model. Phys. Rev. D 1998 ,58,\n116002.\n6. Greenberg, O.W. CPT violation implies violation of Loren tz invariance. Phys. Rev. Lett. 2002 ,89, 231602.\n7. Kostelecký, V .A.; Russell, N. Data tables for Lorentz and CPT violation. Rev. Mod. Phys. 2011 ,83, 11; 2021\nEdition, arXiv:0801.0287v14.\n8. Kostelecký, V .A. Gravity, Lorentz violation, and the Sta ndard Model. Phys. Rev. D 2004 ,69, 105009.Version September 17, 2021 submitted to Symmetry 19 of 20\n9. Kostelecký, V .A.; Lehnert, R. Stability, causality, and Lorentz and CPT violation. Phys. Rev. D 2001 ,63,\n065008.\n10. Drummond, I.T. Quantum field theory in a multimetric back ground. Phys. Rev. D 2013 ,88, 025009.\n11. Belich, H; Bernald, L.D.; Gaete, P; Helayël-Neto, J.A.; Leal, F.J.L. Aspects of CPT-even Lorentz-symmetry\nviolating physics in a supersymmetric scenario. Eur. Phys. J. C 2015 ,75, 291.\n12. Carroll, S.M.; Harvey, J.A.; Kostelecký, V .A.; Lane, C. D.; Okamoto, T. Noncommutative field theory and\nLorentz violation. Phys. Rev. Lett. 2001 ,87, 141601.\n13. Chaichian, M.; Sheikh-Jabbari, M.M.; Tureanu, A. Non-c ommutativity of space-time and the hydrogen atom\nspectrum. Eur. Phys. J. C 2004 ,36, 251.\n14. Hayakawa, M. Perturbative analysis on infrared aspects of noncommutative QED on R4.Phys. Lett. B 2000 ,\n478, 394.\n15. Kostelecký, V .A. Riemann-Finsler geometry and Lorentz -violating kinematics. Phys. Lett. B 2011 ,701, 137.\n16. Silva, J.E.G.; Maluf, R.V .; Almeida, C.A.S. A nonlinear dynamics for the scalar field in Randers spacetime.\nPhys. Lett. B 2017 ,766, 263.\n17. Foster, J.; Lehnert, R. Classical-physics application s for Finsler bspace. Phys. Lett. B 2015 ,746, 164.\n18. Edwards, B.; Kostelecký, V .A. Riemann-Finsler geometr y and Lorentz-violating scalar fields. Phys. Lett. B\n2018 ,786, 319.\n19. Bluhm, R.; Kostelecký, V .A.; Russell, N. Testing CPT wit h anomalous magnetic moments. Phys. Rev. Lett.\n1997 ,79, 1432.\n20. Bluhm, R.; Kostelecký, V .A.; Russell, N. CPT and Lorentz tests in Penning traps. Phys. Rev. D 1998 ,57, 3932.\n21. Ding, Y.; Kostelecký, V .A. Lorentz-violating spinor el ectrodynamics and Penning traps. Phys. Rev. D 2016 ,\n94, 056008.\n22. Ding, Y.; Rawnak, M.F. Lorentz and CPT tests with charge- to-mass ratio comparisons in Penning traps.\nPhys. Rev. D 2020 ,94, 056009.\n23. Mittleman, R.K.; Ioannou, I.I.; Dehmelt, H.G.; Russell , N. Bound on CPT and Lorentz symmetry with a\ntrapped electron. Phys. Rev. Lett. 1999 ,102, 2116.\n24. Hanneke, D. Cavity control in a single-electron quantum cyclotron: An improved measurement of the\nelectron magnetic moment. Ph.D. thesis, Harvard Universit y, 2007.\n25. Ding, Y. Lorentz and CPT tests using Penning traps. Symmetry 2019 ,11, 1220.\n26. Bluhm, R.; Kostelecký, V .A.; Lane, C.D. CPT and Lorentz t ests with muons. Phys. Rev. Lett. 2000 ,84, 1098.\n27. Gomes, A.H.; Kostelecký, V .A.; Vargas, A. Laboratory te sts of Lorentz and CPT symmetry with muons.\nPhys. Rev. D 2014 ,90, 076009.\n28. Kostelecký, V .A.; Vargas, A. Lorentz and CPT tests in hyd rogen, antihydrogen, and related systems. Phys.\nRev. D 2015 ,92, 056002.\n29. Kostelecký, V .A.; Vargas, A. Lorentz and CPT Tests with c lock-comparison experiments. Phys. Rev. D 2018 ,\n98, 036003.\n30. Kostelecký, V .A.; Mewes, M. Fermions with Lorentz-viol ating operators of arbitrary dimension. Phys. Rev.\nD2013 ,88, 096006.\n31. Li, Z.; Kostelecký, V .A. Gauge field theories with Lorent z-violating operators of arbitrary dimension.\nPhys. Rev. D 2019 ,99, 056016.\n32. Kostelecký, V .A.; Mewes, M. Electrodynamics with Loren tz-violating operators of arbitrary dimension. Phys.\nRev. D 2009 ,80, 015020.\n33. Kostelecký, V .A.; Mewes, M. Neutrinos with Lorentz-vio lating operators of arbitrary dimension. Phys. Rev.\nD2012 ,85, 096005.\n34. Kostelecký, V .A.; Mewes, M.; Testing local Lorentz inva riance with gravitational waves. Phys. Lett. B 2016 ,\n757, 510.\n35. For a review see, for example, Brown, L.S.; Gabrielse, G. Geonium theory: Physics of a single electron or ion\nin a Penning trap. Rev. Mod. Phys. 1986 ,58, 233.\n36. Kostelecký, V .A.; Lane, C.D. Constraints on Lorentz vio lation from clock-comparison experiments. Phys. Rev.\nD1999 ,60, 116010.\n37. Kostelecký, V .A.; Mewes, M. Signals for Lorentz violati on in electrodynamics. Phys. Rev. D 2002 ,66, 056005.\n38. Canè, F.; Bear, D.; Phillips, D.F.; Rosen, M.S.; Smallwo od, C.L.; Stoner, R.E.; Walsworth, R.L.; Kostelecký, V .A.\nBound on Lorentz and CPT violating boost effects for the neut ron. Phys. Rev. Lett. 2004 ,93, 230801.Version September 17, 2021 submitted to Symmetry 20 of 20\n39. Heckel, B.R.; Adelberger, E.G.; Cramer, C.E.; Cook, T.S .; Schlamminger, S.; Schmidt, U. Preferred-frame and\nCP-violation tests with polarized electrons. Phys. Rev. D 2008 ,78, 092006.\n40. Bluhm, R.; Kostelecký, V .A.; Lane, C.D.; Russell, N. Pro bing Lorentz and CPT violation with space-based\nexperiments. Phys. Rev. D 2003 ,68, 125008.\n41. Gabrielse, G.; Khabbaz, A.; Hall, D.S.; Heimann, C.; Kal inowsky, H.; Jhe, W. Precision mass spectroscopy of\nthe antiproton and proton using simultaneously trapped par ticles. Phys. Rev. Lett. 1999 ,82, 3198.\n42. Ulmer, S.; Smorra, C.; Mooser, A.; Franke, K.; Nagahama, H.; Schneider, G.; Higuchi, T.; Van Gorp S.; Blaum,\nK.; Matsuda, Y.; Quint, W.; Walz, J.; Yamazaki Y. High-preci sion comparison of the antiproton-to-proton\ncharge-to-mass ratio. Nature 2015 ,524, 196.\n43. Schwinberg, P .B.; Van Dyck Jr. R.S.; Dehmelt H.G. Trappi ng and thermalization of positrons for geonium\nspectroscopy. Phys. Lett. A 1981 ,81, 119.\n44. Hoogerheide, S.F.; Dorr, J.C.; Novitski, E.; Gabrielse , G. High efficiency positron accumulation for\nhigh-precision magnetic moment experiments. Rev. Sci. Instrum. 2015 ,86, 053301.\n45. Gabrielse, G.; Fayer, S.E.; Myers, T.G.; Fan, X. Towards an improved test of the Standard Model’s most\nprecise prediction. Atoms 2019 ,7, 45.\n46. Schneider, G.; Mooser, A.; Bohman, M.; Schön, N.; Harrin gton, J.; Higuchi, T.; Nagahama, H.; Sellner, S.;\nSmorra, C.; Blaum, K.; Matsuda, Y.; Quint, W.; Walz, J.; Ulme r, S. Double-trap measurement of the proton\nmagnetic moment at 0.3 parts per billion precision. Science 2017 ,358, 1081.\n47. Mooser, A.; Ulmer, S.; Blaum, K.; Franke, K.; Kracke, H.; Leiteritz, C.; Quint, W.; Rodegheri, C.C.; Smorra,\nC.; Walz, J. Direct high-precision measurement of the magne tic moment of the proton. Nature 2014 ,509, 596.\n48. Smorra, C.; Sellner, S.; Borchert, M.J.; Harrington, J. A.; Higuchi, T.; Nagahama, H.; Tanaka, T.; Mooser, A.;\nSchneider, G.; Bohman, M.; Blaum, K.; Matsuda, Y.; Ospelkau s, C.; Quint, W.; Walz, J.; Yamazaki, Y.; Ulmer,\nS. A parts-per-billion measurement of the antiproton magne tic moment. Nature 2017 ,550, 371.\n49. Meiners, T.; Niemann, M.; Paschke, A.G.; Borchert, M.; I del, A.; Mielke, J.; Voges, K.; Bautista-Salvador,\nA.; Lehnert, R.; Ulmer, S.; Ospelkaus, C. Towards sympathet ic laser cooling and detection of single\n(anti-)protons. in Kostelecký, V .A. ed., Proceedings of the Seventh Meeting on CPT and Lorentz Symmet ry, World\nScientific, Singapore, 2017.\n50. Dehmelt, H.G.; Mittleman, R.K.; Van Dyck R.S.; Schwinbe rg, P . Past electron-positron g−2 experiments\nyielded sharpest bound on CPT violation for point particles .Phys. Rev. Lett. 1999 ,83, 4694.\n© 2021 by the authors. Submitted to Symmetry for possible open access publication\nunder the terms and conditions of the Creative Commons Attri bution (CC BY) license\n(http://creativecommons.org/licenses/by/4.0/)." }, { "title": "2106.11782v3.Sharp_decay_rate_for_the_damped_wave_equation_with_convex_shaped_damping.pdf", "content": "arXiv:2106.11782v3 [math.AP] 5 Jan 2022SHARP DECAY RATE FOR THE DAMPED WAVE EQUATION WITH\nCONVEX-SHAPED DAMPING\nCHENMIN SUN\nAbstract. We revisit the damped wave equation on two-dimensional toru s where the damped region\ndoes not satisfy the geometric control condition. It was sho wn in [AL14] that, for sufficiently regular\ndamping, the damped wave equation is stale at a rate sufficient ly close to t−1. We show that if the\ndamping vanishes like a H¨ older function |x|β, and in addition, the boundary of the damped region is\nlocally strictly convex with positive curvature, the wave i s stable at rate t−1+2\n2β+7, which is better\nthan the known optimal decay rate t−1+1\nβ+3for strip-shaped dampings of the same H¨ older regularity.\nMoreover, we show by example that the decay rate is optimal. T his illustrates the fact that the sharp\nenergy decay rate depends not only on the order of vanishing o f the damping, but also on the shape of\nthe damped region. The main ingredient of the proof is the ave raging method (normal form reduction)\ndeveloped by Hitrik and Sj¨ ostrand ([ Hi1][Sj]).\n1.Introduction\n1.1.Background. Let(M,g)beacompact RiemannianmanifoldwiththeBeltrami-Laplac e operator\n∆. Consider the damped wave equation\n/braceleftBigg\n∂2\ntu−∆u+a(z)∂tu= 0,inR+×M,\n(u,∂tu)|t=0= (u0,u1),inM,(1.1)\nwherea(z)≥0 is the damping. The well-posedness of ( 1.1) is a consequence of the Lumer-Philips\ntheorem and the maximal dissipative property of the generat or\nL=/parenleftbigg0 Id\n∆−a(z)/parenrightbigg\n(1.2)\non the Hilbert space H:=H1(M)×L2(M). For a solution ( u,∂tu)∈H1(M)×L2(M), the energy\ndefined by\nE[u](t) :=1\n2/ba∇dbl∇u(t)/ba∇dbl2\nL2(M)+1\n2/ba∇dbl∂tu(t)/ba∇dbl2\nL2(M)\nis decreasing in time:\nd\ndtE[u](t) =−/integraldisplay\nMa|∂tu|2≤0.\nA basic question is the decay rate of the energy as t→+∞.\nIt was proved by Rauch-Taylor [ RaT] (∂M=∅) and by Bardos-Lebeau-Rauch [ BLR] (∂M/\\e}atio\\slash=∅)\nthat, for continuous damping a∈C(M), if the set ω={a>0}verifies the geometric control condition\n(GCC), then there exists α0>0 such that the uniform stabilization holds:\nE[u](t)≤E[u](0)e−α0t,∀t≥0. (1.3)\n12 CHENMIN SUN\nIf (GCC) for ω={a >0}is not satisfied, there are very few cases that the uniform sta bilization\n(1.3) holds (see [ BG17] and [Zh])1. Lebeau [ Le93] constructed examples with arbitrary slowly decaying\ninitial data in the energy space H1(M)×L2(M). Nevertheless, if the initial data is more regular,\nsay inH2(M)×H1(M), the uniform decay rate 1 /log(1 +t) holds ([ Le93]). Since then, intensive\nresearch activities focus on possible improvement of the lo garithmic decay rate for regular initial data,\nin special geometric settings.\nBeyond(GCC),thedeterminationofbetterdecay rateforspe cial manifolds Mandspecialdampings\ndepends on at least the following factors from the existing l iterature:\n(a) The dynamical properties for the geodesic flow of the unde rlying manifold M.\n(b) The dimension of the trapped rays as well as relative posi tions between trapped rays and the\nboundary∂{a>0}of the damped region.\n(c) Regularity and the vanishing properties of the damping anear∂{a>0}.\nIt is known that the energy decay rate is linked to the average d function along the geodesic flow ϕt:\nρ/ma√sto→ AT(a)(ρ) :=1\nT/integraldisplayT\n0a◦ϕt(ρ)dt, ρ∈T∗M.\nIndeed, (GCC) is equivalent to the lower bound AT(ρ)≥c0>0 for some T >0 large enough on the\nsphere bundle S∗M. Roughly speaking, when the geodesic flow is “unstable”, one may improve the\nenergy decay rate (see [ No] for more detailed explanation and references therein). As an illustration\nof (a), when Mis a compact hyperbolic surface, Jin [ Ji] shows the exponential energy decay rate for\nregular data living in H2(M)×H1(M). In this direction, we refer also [ BuC],[Ch],[CSVW],[Riv] and\nreferences therein.\nThepolynomial decay rate is theintermediate situation bet ween thelogarithmic decay rates andthe\nexponential decay rates, exhibitedinless chaotic geometr y liketheflattorusandboundeddomains(see\n[LiR][BH05][Ph07][AL14] and references therein), where the generalized geodesic fl ows are unstable.\nWe refer [ LLe],[BZu] for other situations of polynomial stabilization, where t he undamped region is a\nsubmanifold.\nWe point out that the factor (a) is almost decisive for the obs ervability (and exact controllability)\nof wave and Schr¨ odinger equations. Comparing with the obse rvability for the wave equation where\n(GCC) is the only criteria (see [ BLR] [BG97]), the stabilization problem is more complicated. Indeed,\nit was shown in [ AL14] (Theorem 2.3) that the observability for the Schr¨ odinger semigroup in some\ntimeT >0 implies automatically that the dampedwave is stable at rat et−1\n2. However, this decay rate\nis not optimal in general. On the two-dimensional torus, if t he damping function is regular enough\nand vanishing nicely, the decay rate can be very close to t−1( [BH05][AL14]). Even when the damping\nis the indicator of a vertical (or horizontal) strip, the opt imal decay rate is known to be t−2\n3(the lower\nbound was obtained by Nonnenmacher in [ AL14] and the upper bound was obtained in [ St]). These\nresults provide evidences of factors (b) and (c) mentioned p reviously. As explained in [ AL14], the\nsignificant difference to the controllability problem is that , there is no general monotonicity property\n1If the trapped rays are all grazing on the boundary ∂{a >0}, the uniform stabilization ( 1.3) can be characterized\n(see [BG17] and [Zh]) by relative positions of the grazing trapped rays and a >0.SHARP DECAY RATE FOR THE DAMPED WAVE EQUATION WITH CONVEX-SH APED DAMPING 3\nof the type: a1≤a2implies the decay rate associated to a2is better (or worse) than the decay rate\nassociated to a1.\nIn this article, we revisit the polynomial stabilization fo r wave equations on flat torus. Our main\nresult reveals that, with the same vanishingorder, the curv ature of the boundaryof the damped region\nalso affects the energy decay rate of damped wave equations.\n1.2.The main result. We concern the polynomial decay rate for ( 1.1) on the two-dimensional flat\ntorusM=T2:=R2/(2πZ)2:\n/braceleftBigg\n∂2\ntu−∆u+a(z)∂tu= 0,inR+×T2,\n(u,∂tu)|t=0= (u0,u1),inT2.(1.4)\nTo present the main result, we introduce some definitions.\nDefinition 1.1. We say that (1.4)isstable at rate t−α, if there exists C >0, such that all the\nsolutionuwith initial data (u0,u1)∈ H2:=H2(T2)×H1(T2)satisfies\n(E[u](t))1\n2≤Ct−α/ba∇dbl(u0,u1)/ba∇dblH2.\nWe say that the rate t−αis optimal, if moreover\nlimsup\nt→+∞tαsup\n0/negationslash=(u0,u1)∈H2(E[u](t))1\n2\n/ba∇dbl(u0,u1)/ba∇dblH2>0.\nNext we introduce the class of damping that we will consider:\nDefinition 1.2. Letm,k∈N,σ>0andkσ<1. The function class Dm,k,σ(Td)is defined by:\nDm,k,σ:={f∈Cm(Td) :|∂αf|/lessorsimilarα,σ|f|1−|α|σ,∀|α| ≤k}.\nNote that Dm,k,σ1⊂ Dm,k,σ2, ifσ1< σ2andkσ2<1. This class contains non-negative functions\nwhich vanish like H¨ older functions. One typical example is\na1(z) =b(z)(max{0,0.1−|z|})1\nσ∈ Dm,m,σ(Td),\nwhereσ <1\nm,b∈Cm(Td) and infTdb>0. The associated damped region is {z∈Td:a1(z)>0}=\n{z∈Td:|z|<0.1}is a disc. Another example is the strip-shaped damping a2(z) =a2(x) such that\n{z∈T2:a2(z)>0}:= (−0.1,0.1)x×Tyand for some m≥4,\ndm\ndxma2(x)≤0 nearx= 0.1 anddm\ndxma2(x)≥0 nearx=−0.1.\nIt was shown in Lemma 3.1 of [ BH05] thata2∈ Dm,m,1\nm(T2).\nFora∈ Dm,k,σ, we denote by Σ a:=∂{a(z)>0}. LetT2\nA,B:=R2/(2πA×2πB) be a general flat\ntorus defined via the covering map πA,B:R2/ma√sto→T2\nA,B.\nDefinition 1.3. An open set ω⊂T2\nA,Bis said to be locally strictly convex with positive curvature , if\nthe boundary of each component of π−1\nA,B(ω)⊂R2isC2and has strictly positive curvature, as a closed\ncurve in R2. Sometimes, we also say that the boundary is locally strictl y convex.\nOur main result is the following:4 CHENMIN SUN\nTheorem 1.1. Letβ >4,m≥10. Assume that a≥0, a∈ Dm,2,1\nβand the open set ω:={z∈T2:\na(z)>0}is locally strictly convex with positive curvature. Assume thata(z)is locally H¨ older of order\nβnear∂ω, in the sense that there exists R0>1, such that\n1\nR0dist(z,∂ω)β≤a(z)≤R0dist(z,∂ω)β,forz∈ωnear∂ω.\nThen the damped wave equation (1.4)is stable at rate t−1+2\n2β+7.Moreover, the decay rate t−1+2\n2β+7is\noptimal in the following sense: there exists a damping a0(z), satisfying all the hypothesis above with\nβ=m≥10, such that the associated damped wave equation (1.4)cannot be stable at rate t−1+2\n2β+7−ǫ,\nfor anyǫ>0.\nRemark 1.4. As a comparison, if a(z) =a(x) depends only on one direction (supported on the\nvertical strip ω) and is locally H¨ older of order βnear∂ω, the optimal stable rate is t−1+1\nβ+3(see\n[Kl][DKl]) which is worse than t−1+2\n2β+7. Our result provides examples that, with the same local\nH¨ older regularity, smaller damped regions better stabili ze the wave equation. To the best knowledge\nof the author, Theorem 1.1also provides the first example where not only the vanishing o rder of the\ndamping can affect the stable rate, but also the shape of the bou ndary of the damped region.\nω2\nω1\na1(z) = (0.1−|z|)β\n+, a2(z) = (0.5−|x|)β\n+T2=R2/(2πZ)2\nThe damping a1generates better decay rate than a2\nRemark 1.5. In the case of rectangular H¨ older regular damping a(z) =a(x), the optimal decay\nrate is shown to be t−1+1\nβ+3([Kl][DKl]) for allβ≥0. In our result, the regularity assumptions\nβ >4,m≥10 in the first part and β≥10 in the second part of Theorem 1.1might be relaxed, as\nfar as tools of microlocal analysis can still be applied (for example, the paradifferential calculus for\nlow-regularity symbols). We leave open the problem to deter mine whether the optimal decay rate\nt−1+2\n2β+7in Theorem 1.1can be extended to rough dampings (i.e. small β≥0), as well as the case\nwherea(z) is an indicator function for a strictly convex subset.\nRemark 1.6. Itwas shown in [ AL14] that, whenthe dampingsatisfies |∇a| ≤Ca1−1\nβfor large enough\nβ, then (1.4) is stable at rate t−1+4\nβ+4(Theorem 2.6 of [ AL14]). With an additional assumption on\n|∇2a| ≤Ca1−2\nβ, our proof of Theorem 1.1essentially provides an alternative proof of this rougher\nstable rate. Indeed, we need one more condition for |∇2a|only to perform the normal form reduction\nin the Section 5. Once reduced to the one-dimensional setting, we are able to apply the same argument\nof Burq-Hitrik [ BH05].\nRemark 1.7. For the reason of exhibition, we have used a contradiction ar gument and the notion of\nsemiclassical defect measures in the proof of Theorem 1.1. Comparing to the argument of [ AL14], we\ndo not make use of second semiclassical measures.SHARP DECAY RATE FOR THE DAMPED WAVE EQUATION WITH CONVEX-SH APED DAMPING 5\nFinally we give some microlocal interpretation of Theorem 1.1. It is known that the decay rate of\nthe damped wave equation is related to the time average along geodesics (see [ No]). As the damping\nhas conormal singularities at the boundary ∂ωof the damped region, the decay rate depends more\nprecisely on the reflected and transmitted energy of waves co ncentrated on trapped rays. If along\nthe conormal direction, the damping is more regular, its int eraction with transversal free waves is\nweaker (analog to the high-low frequency interaction), hen ce the transmission effect is stronger than\nthe reflection, and consequently, the decay rate is better. W hen the boundary ∂ωof the damped\nregion is convex, the average of the damping along any direct ion gains1\n2local H¨ older regularity, near\nthe vanishing points along the transversal direction (see P roposition 4.4for details). This heuristic\nindicates that, with the same local H¨ older regularity near the boundary (of the damped region), the\nconvex-shaped dampinghasbetterstablerate for( 1.4), thanstrip-shapeddampings(that areinvariant\nalong one direction).\n1.3.Resolvent estimate. The proof of Theorem 1.1relies on Borichev-Tomilov’s criteria of the\npolynomial semi-group decay rate and the corresponding res olvent estimate for Lgiven in ( 1.2):\nProposition 1.8 ([BoT10]).We have Spec (L)∩iR=∅. Then, the following statements are equivalent:\n(a)/vextenddouble/vextenddouble(iλ−L)−1/vextenddouble/vextenddouble\nL(H)≤C|λ|1\nαfor allλ∈R,|λ| ≥1;\n(b)The damped wave equation (1.1)is stable at rate t−α.\nBy Proposition 1.8and Proposition 2.4 of [ AL14], the proof of Theorem ( 1.1) can be reduced to\nthe following semiclassical resolvent estimate:\nTheorem 1.2. Letβ >4,m≥10. Assume that a≥0, a∈ Dm,2,1\nβand the open set ω:={z∈T2:\na(z)>0}is locally strictly convex with positive curvature. Assume thata(z)is locally H¨ older of order\nβnear∂ω, in the sense that there exists R0>1, such that\n1\nR0dist(z,∂ω)β≤a(z)≤R0dist(z,∂ω)β,forz∈ωnear∂ω.\nThen there exist h0∈(0,1)andC0>0, such that for all 00}is disjoint unions\nof vertical strips (αj,βj)x×Tyand{W≥0} /\\e}atio\\slash=T2. Assume moreover that for each j∈ {1,···,l},\nC1Vj(x)≤W(x)≤C2Vj(x)on(αj,βj),6 CHENMIN SUN\nwhereVj(x)>0are continuous functions on (αj,βj), satisfying\nVj(x) =\n\n(x−αj)γ, αj0such that for all 00of order\no(h2+2\n2β+5) is non-zero along finitely many closed trajectories with pe riodic directions on the phase\nspace. On the other hand, it turns out that the restriction of semiclassical measure to any periodic\ndirection is zero, which leads to a contradiction. This anal ysis follows from a second microlocalization\nprocedure and will be achieved in three major steps:\n•In Section 3, using the positive commutator method, we show that the semi classical measure\ncorresponding to the transversal high frequency part of sca le/greaterorsimilarh−1\n2−1\n2β+5is zero.\n•In Section 5, we deal with scales for transversal low frequencies. Using the averaging method,\nwe transfer quasi-modes ( uh)h>0to new quasi-modes ( vh)h>0, satisfying new equations that\ncommute with the vertical derivative. This allows us to redu ce the problem to the one-\ndimensional setting. This is the key part of the proof , for which we need several elementary\nproperties of the averaging operator, presented in Section 4.\n•In Section 6, we prove the reduced one-dimensional resolvent estimate ( Proposition 6.1).\nFurthermore, we prove in Section 7the lower bound in Theorem 1.2. At the end of this article, we add\ntwo appendices. In Appendix A, we reproduce the proof of Theo rem1.3in order to be self-contained\nand to fix some gaps in the paper of [ DKl]. In Appendix B, we review several technical results about\nthe semiclassical pseudo-differential calculus, needed in S ection5.\nAcknowledgment. The author is supported by the program: “Initiative d’excel lence Paris Seine”\nof CY Cergy-Paris Universit´ e and the ANR grant ODA (ANR-18- CE40- 0020-01). The author would\nlike to thank anonymous referee’s suggestions that help to i mprove this article.\n2.Contradiction argument and the first microlocalization\n2.1.A priori estimate and the contradiction argument. We will adapt basic conventions for\nnotations in the semiclassical analysis ([ Zw12]). Denote by Ph=−h2∆−1 and we fix the parameter\nσ=1\nβthroughout this article. We denote by δha small parameter such that δh→0 ash→0. We\ndenoteby /planckover2pi1=h1\n2δ1\n2\nhasecond semiclassical parameter. For theproofof Theorem 1.2, we fixδh=h2\n2β+5.\nTheorem 1.2is the consequence of the following key proposition:SHARP DECAY RATE FOR THE DAMPED WAVE EQUATION WITH CONVEX-SH APED DAMPING 7\nProposition 2.1. Letuhbe a sequence of quasi-modes of width h2δhwithδh=h2\n2β+5i.e.\n(Ph+iha)uh=fh=oL2(h2δh).\nThen ifuh=OL2(1), we haveuh=oL2(1).\nThe proof of Proposition 2.1will occupy the rest of this article. We argue by contradicti on. First\nwe delete all the zero elements in a given sequence of uh. Then, up to extracting a subsequence and\nrenormalization, we may assume that\n/ba∇dbluh/ba∇dblL2(T2)= 1. (2.1)\nThe following a priori estimate is simple:\nLemma 2.2. We have the following a priori estimates:\n(a)/ba∇dbla1/2uh/ba∇dblL2=o(h1\n2δ1\n2\nh) =o(/planckover2pi1).\n(b)/ba∇dblh∇uh/ba∇dbl2\nL2−/ba∇dbluh/ba∇dbl2\nL2=o(h2δh).\nProof.Multiplying the equation ( Ph+iha)uh=fhbyuh, and integrating by part, we get\n/ba∇dblh∇uh/ba∇dbl2\nL2−/ba∇dbluh/ba∇dbl2\nL2+ih(auh,uh)L2= (fh,uh)L2.\nTaking the imaginary part and the real part, we obtain (a) and (b), with respectively. /square\nSince the sequence ( uh)h>0is bounded L2(T2), there exist a subsequence, still denoted as ( uh)h>0,\nand a Radon measure µonT∗T2, such that for any symbol a∈C∞\nc(T∗T2), there holds\nlim\nh→0(Oph(a)uh,uh)L2=/a\\}b∇acketle{tµ,a/a\\}b∇acket∇i}ht. (2.2)\nBelow, we will denote µthe semiclassical defect measure associated to this subseq uence (uh)h>0. For\nthe proof of this existence of semi-classical measure, one m ay consult Chapter 5 of [ Zw12].\nLemma 2.3. we have\nsupp(µ){(z,ζ)∈T∗T2:|ζ|= 1}andµ|ω×R2= 0,\nwhereω={z∈T2:a>0}.\nProof.This property follows from the standard elliptic regularit y which only requires quasi-mode for\nPhof orderOL2(h). The damping term ihauhcan be roughly treated as an error OL2(h). For example,\none can consult Theorem 5.4 of [ Zw12] for a proof. /square\nLetϕtbe the geodesic flow on T∗T2. We recall the following invariant property of the semiclas sical\nmeasure:\nLemma 2.4. The semiclassical measure µis invariant by the flow ϕt, i.e.\nϕ∗\ntµ=µ.\nProof.This property holds true for quasi-mode of Phof orderoL2(h). From (a) of Lemma 2.2, we\nhavefh−ihauh=oL2(h). The proof then follows from a standard propagation argume nt (see for\nexample Theorem 5.5 of [ Zw12]). /square8 CHENMIN SUN\n2.2.Reducing to periodic trapped directions. Inordertoperformthefineranalysisneartrapped\nrays, we need to do a change of coordinate, following [ BZ12] and [AL14]. The spirit of the second-\nmicrolocalization here is also close to the work [ AM14].\nBy identifying T2=R2/(2πZ)2, we decompose S1as rational directions\nQ:={ζ∈S1:ζ=(p,q)/radicalbig\np2+q2,(p,q)∈Z2,gcd(p,q) = 1}\nand irrational directions R:=S1\\Q. Since the orbit of an irrational direction is dense, by Lemm a\n2.3and Lemma 2.4, we have\nµ=µ|T2×Q=/summationdisplay\nζ0∈Qµζ0.\nIt suffices to show that, for each ζ0=(p0,q0)√\np2\n0+q2\n0∈ Q, the restricted measure µζ0is zero2. Denote by Λ 0,\nthe rank 1 submodule of Z2generated by Ξ 0= (p0,q0). Denote by\nΛ⊥\n0:={ζ∈R2:ζ·Ξ0= 0}\nthe dual of the submodule Λ 0. Denote by\nT2\nΞ0:= (RΛ0/(2πΛ0))×(Λ⊥\n0/(2πZ)2∩Λ⊥\n0).\nThen we have a natural smooth covering map πΞ0:T2\nΞ0/ma√sto→T2of degreep2\n0+q2\n0. The pullback of a\n2π×2πperiodic function fsatisfies\n(π∗\nΞ0f)(X+kτ,Y+lτ) = (π∗\nΞ0f)(X,Y), k,l∈Z,(X,Y)∈R2,\nwhereτ= 2π/radicalbig\np2\n0+q2\n0.\nT2\nΞ0\nΞ0= (3,−2)\nΛ0Λ⊥\n0\nBy pulling back to the torus T2\nΞ0, we can identify the sequence ( uh)⊂L2(T2) as (π∗\nΞ0uh)⊂L2(T2\nΞ0)\nand in this new coordinate system, ζ0=Ξ0\n|Ξ0|= (0,1). The semi-classical defect measure µonT∗T2is\nthe pushforward of the semi-classical measure associated t o (π∗\nΞ0uh). Since the period of the torus T2\nΞ0\n2In fact, we only need to consider finitely many ζ0∈ Q, since when p2\n0+q2\n0is large enough, the associated periodic\ndirection is close to an irrational direction and the trajec tory will eventually enter ω.SHARP DECAY RATE FOR THE DAMPED WAVE EQUATION WITH CONVEX-SH APED DAMPING 9\nhas no influence of the analysis in the sequel3, we will still use the notation T2to stand for T2\nΞ0, the\nvariablesz= (x,y),ζ= (ξ,η) to stand for variables Z= (X,Y),Ξ onT∗T2\nΞ0, and assuming the period\nto be 2πfor simplicity. The only thing that will change is that the pr e-image of the damping π−1\nΞ0(ω)\nis now a disjoint union of p2\n0+q2\n0copies ofωonT2\nΞ0, and each component is still strictly convex with\npositive curvature. For this reason, in the hypothesis of Pr oposition 4.4, we assume that the boundary\nof{a>0}is made of disjoint unions of strictly convex curves.\n3.Analysis of the transversal high frequencies\nRecall that ζ0= (0,1), and our goal is to show that µ1ζ=ζ0= 0. In this section, we deal with the\ntransversal high frequencies of size O(/planckover2pi1−1) and use the positive commutator method to show these\nportions are propagated into the flowout of the damped region ω.\nFor the geodesic flow ϕtonT∗T2andζ∈S1, we denote by ϕt(·,ζ) :T2→T2the projection of the\nflow mapϕt. By shifting the coordinate, we may assume\nω0:=I0×Ty⊂/uniondisplay\nt∈[0,2π]ϕt(·,ζ0)(ω),\nwhere\nI0= (−σ0,σ0)⊂π1({z:a(z)≥c0>0}),for someσ0<π\n100\nandπ1:T2/ma√sto→Txthe canonical projection. Therefore, there exist ǫ0>0,c0>0 sufficiently small and\nT0>0, such that for any |ζ|= 1,z0∈ω0,|ζ−ζ0| ≤ǫ0,\n/integraldisplayT0\n0(a◦ϕt)(z0,ζ)dt≥c0>0. (3.1)\nω0\nsupp(a)ζ0= (0,1)\nTo microlocalize the solution near ζ0, we pickψ0∈C∞\nc(R) and consider u1\nh:=ψ0/parenleftbighDx\nǫ0/parenrightbig\nuh, then\n(Ph+iha)u1\nh=f1\nh:=ψ0/parenleftbighDx\nǫ0/parenrightbig\nfh−ih/bracketleftbig\nψ0/parenleftbighDx\nǫ0/parenrightbig\n,a/bracketrightbig\nuh.\nLetµ1be the semiclassical measure of u1\nh, thenµ1=|ψ/parenleftbigξ\nǫ0/parenrightbig\n|2µ.\nLemma 3.1. We have\n/ba∇dbla1\n2u1\nh/ba∇dblL2=o(h1\n2δ1\n2\nh),/ba∇dblf1\nh/ba∇dblL2=o(h2δh).\n3As we fix one periodic direction ζ0and consider the semi-classical limit h→0, one does not need to worry about\nthe fact that the period 2 π/radicalbig\np2\n0+q2\n0may be very large.10 CHENMIN SUN\nProof.It suffices to show that f1\nh=oL2(h2δh), and the first assertion follows from the same proof of\n(a) of Lemma 2.2. By the symbolic calculus,\ni/bracketleftbig\nψ0/parenleftbighDx\nǫ0/parenrightbig\n,a/bracketrightbig\n=ǫ−1\n0hOph/parenleftbig\nψ′\n0/parenleftbigξ\nǫ0/parenrightbig\n∂xa/parenrightbig\n+OL2(ǫ−2\n0h2).\nFrom the pointwise inequality for the non-negative functio na:\n|∇a(x)|2≤2/ba∇dbla/ba∇dblW2,∞a(x), (3.2)\nwe have, for some C >0,\nCa−/vextendsingle/vextendsingleǫ−1\n0ψ′\n0/parenleftbigξ\nǫ0/parenrightbig\n∂xa/vextendsingle/vextendsingle2≥0 onT∗T2.\nTherefore, by the sharp G˚ arding inequality, we get\n/vextenddouble/vextenddoubleǫ−1\n0Oph/parenleftbig\nψ′\n0/parenleftbigξ\nǫ0/parenrightbig\n∂xa/parenrightbig\nuh/vextenddouble/vextenddouble\nL2≤C|(auh,uh)L2|1\n2+Ch1\n2/ba∇dbluh/ba∇dblL2.\nTogether with (a) of Lemma 2.2, this implies that\n/vextenddouble/vextenddoubleih/bracketleftbig\nψ0/parenleftbighDx\nǫ0/parenrightbig\n,a/bracketrightbig\nuh/vextenddouble/vextenddouble\nL2=o(h5\n2) =o(h2δh).\nThe proof of Lemma 3.1is complete. /square\nRecall that /planckover2pi1=h1\n2δ1\n2\nh. Letψ∈C∞\nc(R), and consider\nvh=ψ(/planckover2pi1Dx)u1\nh, wh= (1−ψ(/planckover2pi1Dx))u1\nh. (3.3)\nIn this decomposition, whcorresponds to the transversal high frequency part, while vhcorresponds to\nthe transversal low frequency part for which will be treated in next sections. Note that\n(Ph+iha)vh=ψ(/planckover2pi1Dx)f1\nh−ih[ψ(/planckover2pi1Dx),a]u1\nh=:r1,h (3.4)\nand\n(Ph+iha)wh= (1−ψ(/planckover2pi1Dx))f1\nh+ih[ψ(/planckover2pi1Dx),a]u1\nh=:r2,h. (3.5)\nWe need to show that the commutator term h[ψ(/planckover2pi1Dx),a]u1\nhcan be viewed as the remainder:\nLemma 3.2. We have\n/ba∇dblr1,h/ba∇dblL2+/ba∇dblr2,h/ba∇dblL2=o(h2δh) =o(h/planckover2pi12).\nConsequently, from Lemma 2.2,\n/ba∇dbla1\n2vh/ba∇dblL2+/ba∇dbla1\n2wh/ba∇dblL2=o(h1\n2δ1\n2\nh) =o(/planckover2pi1).\nProof.According to the symbolic calculus,\ni[ψ(/planckover2pi1Dx),a] =/planckover2pi1Op/planckover2pi1(ψ′(ξ)∂xa)+C/planckover2pi12Op/planckover2pi1(∂2\nξψ·∂2\nxa)+OL(L2)(/planckover2pi13).\nUsing the fact that a∈ Dm,2,σandσ <1\n4, we havea1\n2∈C2\nc(T2). Applying the special symbolic\ncalculus Lemma B.3(b) withκ=∂x(a1\n2),b2=a1\n2andϕ=ψ′, we have\n1\n2Op/planckover2pi1(ψ′(ξ)∂xa) =Op/planckover2pi1(ψ′(ξ)∂x(a1\n2))a1\n2−1\ni/planckover2pi1Op/planckover2pi1(ψ′′(ξ)∂x(a1\n2)·∂x(a1\n2))+OL(L2)(/planckover2pi12).SHARP DECAY RATE FOR THE DAMPED WAVE EQUATION WITH CONVEX-SH APED DAMPING 11\nApplying Lemma B.3(a) withκ=b1=∂x(a1\n2),ϕ=ψ′′, we have\n−/planckover2pi1\niOp/planckover2pi1(ψ′′(ξ)∂x(a1\n2)·∂x(a1\n2)) =−/planckover2pi1\niOp/planckover2pi1(ψ′′(ξ)∂x(a1\n2))∂x(a1\n2)+OL(L2)(/planckover2pi12).\nSinceψis only a function of ξanda−1\n2∂xa∈L∞, we have\n/vextenddouble/vextenddoubleOp/planckover2pi1/parenleftbig\nψ′(ξ)∂xa\na1\n2/parenrightbig\na1\n2u1\nh/vextenddouble/vextenddouble\nL2(T2)=/vextenddouble/vextenddouble(a−1\n2∂xa)ψ′(hDx)(a1\n2u1\nh)/vextenddouble/vextenddouble\nL2(T2)≤C/ba∇dbla1\n2u1\nh/ba∇dblL2=o(/planckover2pi1).\nFor the term Op/planckover2pi1(∂2\nξψ∂2\nxa), since|∂2\nxa|/lessorsimilara1−2σ/lessorsimilara1\n2, by the sharp G˚ arding inequality,\nRe/parenleftbig\nOp/planckover2pi1(Ca−|ψ′′(ξ)∂2\nxa|2)u1\nh,u1\nh/parenrightbig\nL2≥ −C/planckover2pi1/ba∇dblu1\nh/ba∇dbl2\nL2.\nThus/ba∇dbl/planckover2pi12Op/planckover2pi1(ψ′′(ξ)∂2\nxa)u1\nh/ba∇dblL2=O(/planckover2pi15\n2). Therefore, by Calder´ on-Vaillancourt (Theorem B.1), we can\nwrite\nih[ψ(/planckover2pi1Dx),a] =h/planckover2pi1A/planckover2pi1a1\n2+h/planckover2pi12B/planckover2pi1∂x(a1\n2)+OL(L2)(h/planckover2pi13),\nwithA/planckover2pi1,B/planckover2pi1bounded operators on L2, uniformly in /planckover2pi1. Since|∂x(a1\n2)|/lessorsimilara1\n2−σ, by (a) of Lemma 2.2\nand the interpolation, we get\n/ba∇dblih[ψ(/planckover2pi1Dx),a]u1\nh/ba∇dblL2=o(h/planckover2pi12)+O(h/planckover2pi13).\nThis completes the proof of Lemma 3.2. /square\nRemark 3.3. Compared to [ AL14] where the damping only satisfies a∈Wk0,∞(T2) and|∇a|/lessorsimilara1−σ,\nour assumption a∈ Dm,2,σis slightly stronger, in order to ensure the commutator of th e damping\nterm is still a remainder. Indeed, here we chose /planckover2pi1=h1\n2δ1\n2\nh≪h1\n2while in [ AL14], the authors chose\n/planckover2pi1=hα,α<1\n3(see the sentence after Proposition 7.2 of [ AL14]). The use of the sharp G˚ arding as in\n[AL14] would not get o(h/planckover2pi12) for the remainders r1,h,r2,h. We also remark that a direct application of\ntheCalder´ on-Vaillancourt theorem inthesymbolic calcul us requires a1\n2∈W3,∞. Sinceweassumeonly\na∈ Dm,2,σ, (thusa1\n2∈W2,∞), we need to exploit the special structure of the commutator [ψ(/planckover2pi1Dx),a]\nand apply the special symbolic calculus Lemma B.3.\nRecall that ω0=I0×Ty.\nLemma 3.4. We have\n/ba∇dblwh1ω0/ba∇dblL2+/ba∇dblh∇wh1ω0/ba∇dblL2=O(h1\n2),ash→0.\nProof.The proof follows from the classical propagation argument, using the geometric control condi-\ntion. Take small intervals I′\n0⊂Tx,I1= (σ1,σ2)⊂Ty, such that I0⊂I′\n0andω1=I′\n0×I1⊂ {a≥δ0}\nfor someδ0>0. For any z0= (x0,y0)∈ω0, by the geometric control condition, there exist\nT1>0,δ1>0,δ2>0, and the small neighborhood U= (x0−δ1,x0+δ1)×(y0−δ1,y0+δ1) of\nz0, such that for all |ζ−ζ0| ≤ǫ0,z∈U, we have\nz+sζ∈ω1, s∈[T1−δ2,T1+δ2].\nIn particular, a(z+sζ)≥δ0. Without loss of generality, we assume that π > σ2> σ1> y0+δ1>\ny0−δ1>−π. Pick two cutoffs χ1(x),χ2(y)≥0, supported in ( x0−δ1,x0+δ1), (y0−δ1,y0+δ1) and12 CHENMIN SUN\nequal to 1 on ( x0−δ1/2,x0+δ1/2),(y0−δ1/2,y0+δ1/2), respectively. Let χ0∈C∞\nc(R) be a cutoff\nnear|ζ−ζ0| ≤ǫ0. For anys≥0, define the symbol\nbs(z,ζ) :=χ0(ζ)·(χ1⊗χ2)◦ϕ−s(z,ζ) =χ0(ζ)χ1(x−sξ)χ2(y−sη).\nω0\nω1\nUϕT0(·,ζ)(U)⊂ω1\nDirect computation yields\nd\nds(Oph(bs)wh,wh)L2=(Oph(∂sbs)wh,wh)L2=−(Oph(ζ·∇zbs)wh,wh)L2.\nIntegrating this equality from s= 0 tos=T0,\n(Oph(bT1)wh,wh)L2−(Oph(b0)wh,wh)L2=−/integraldisplayT1\n0(Oph(ζ·∇zbs)wh,wh)L2ds. (3.6)\nNote that for fixed s∈[0,T1],\ni\nh[Ph,Oph(bs)] = 2Oph(ζ·∇zbs)+OL(L2)(h),\nwe have\n(Oph(b0)wh,wh)L2=(Oph(bT1)wh,wh)L2+i\n2h/integraldisplayT1\n0([Ph,Oph(bs)]wh,wh)L2+O(h).(3.7)\nUsing the equation\nPhwh=r2,h−ihawh,\nwe have\n1\nh([Ph,Oph(bs)]wh,wh)L2=2i\nhIm(Oph(bs)wh,r2,h−ihawh)L2\n=o(h1+δ)+O(1)/ba∇dbla1\n2wh/ba∇dblL2/ba∇dbla1\n2Oph(bs)wh/ba∇dblL2\n=O(h), (3.8)\nwhere to the last step, we write\na1\n2Oph(bs) = Oph(bs)a1\n2+[a1\n2,Oph(bs)]\nand use the last assertion of Lemma 3.2, as well as the symbolic calculus.\nFinally, from the support property of b0(z) =χ0(ζ)χ1(x)χ2(y), we have\na(z)χ0(ζ)≥δ0bT1(z,ζ).\nThus by the sharp G˚ arding inequality,\n|(Oph(bT1)wh,wh)L2| ≤δ−1\n0|(Oph(a(z)χ0(ζ))wh,wh)L2|+O(h)≤Cδ−1\n0/ba∇dbla1\n2wh/ba∇dbl2\nL2+O(h).SHARP DECAY RATE FOR THE DAMPED WAVE EQUATION WITH CONVEX-SH APED DAMPING 13\nCombining this with ( 3.7),(3.8) and the last assertion of Lemma 3.2, we deduce that /ba∇dblwh1U/ba∇dblL2=\nO(h1\n2),/ba∇dblh∇wh1U/ba∇dblL2=O(h1\n2). By the partition of unity of ω0, we complete the proof of Lemma\n3.4. /square\nNow we are ready to prove the main result in this section, that is the transversal high frequency\npart is of order oL2(1):\nProposition 3.5. We have /ba∇dblwh/ba∇dblL2=O(δh) =O(/planckover2pi1h−1\n2), ash→0.\nProof.We use the positive commutator method to detect the transver sal propagation, similarly as in\n[BS19]. Recall that ω0=I0×TandI0= (−σ0,σ0),σ0<π\n100. Takeφ=φ(x)∈C∞(Tx;[0,1]) such\nthat:\nsupp(1−φ)⊂I0, φ≡0 near[−σ0\n2,σ0\n2],supp(φ′)⊂I0.\nDenote by\nX(x) := (x+π)1−π≤x<−σ0\n2+(x−π)1σ0\n2≤x<π,\nthenφ(x)X∂xis well-defined smooth vector field on T2. We now compute the inner product\n([Ph,φ(x)X∂x]wh,wh)L2\nin two ways. On the one hand, from the commutator relation\n[Ph,φ(x)X∂x] =−2(φX)′h2∂2\nx−h2(φX)′′∂x,\nwe have\n([Ph,φ(x)X∂x]wh,wh)L2≥2(φ(x)h∂xwh,h∂xwh)L2−C/ba∇dbl(φ′(x))1/2h∂xwh/ba∇dbl2\nL2\n−Ch/ba∇dblh∂xwh/ba∇dblL2/ba∇dblwh/ba∇dblL2. (3.9)\nOn the other hand, using the equation ( 3.5), we have\n([Ph,φ(x)X∂x]wh,wh)L2=(φ(x)X∂xwh,r2,h−ihawh)L2−(φ(x)X∂x(r2,h−ihawh),wh)L2\n≤C\nh(/ba∇dblh∂xwh/ba∇dblL2+h/ba∇dblwh/ba∇dblL2)/ba∇dblr2,h/ba∇dblL2+C/ba∇dbla1/2h∂xwh/ba∇dblL2/ba∇dbla1\n2wh/ba∇dblL2+C/ba∇dbla1\n2wh/ba∇dbl2\nL2.\n(3.10)\nCombining ( 3.9) and (3.10) and Lemma 3.4, we get\n/ba∇dblh∂xwh/ba∇dbl2\nL2≤C/ba∇dblh∂xwh1ω0/ba∇dbl2\nL2+C/ba∇dbla1/2h∂xwh/ba∇dbl2\nL2+C/ba∇dbla1\n2wh/ba∇dbl2\nL2+C\nh2/ba∇dblr2,h/ba∇dbl2\nL2+Ch2/ba∇dblwh/ba∇dbl2\nL2\n≤O(h)+o(hδh).\nIn particular, by the definition of wh, we have\nC(h/planckover2pi1−1)/ba∇dblwh/ba∇dblL2≤ /ba∇dblh∂xwh/ba∇dblL2≤O(h1\n2),\nand this completes the proof of Proposition 3.5. /square14 CHENMIN SUN\n4.The averaging properties of functions\nIn order to treat the transversal low frequency part vh=ψ(/planckover2pi1Dx)u1\nhof (3.3), we will adapt the\naveraging argument of Sj¨ ostrand ([ Sj]) and Hitrik ([ Hi1]) to average the operator Ph+ihaalong the\ntrapped direction (it worth also mentioning a series work of Hitrik-Sj¨ ostrand ([ HS1][HS2][HS3]) that\na related averaging procedure was used to describe the spect rum of the damped wave operator). This\namounts to understand the regularity of averaged functions . The goal of this section is to establish\nseveral properties of the averaging operator which will be u sed in Section 5for normal form reductions.\nMore importantly, we will prove the key geometric propositi on (Proposition 4.4) for convex-shaped\ndamping that is responsible for the improvement of the resol vent estimate.\nGiven a direction v∈S1, we say that visperiodic, ifv= (ξ,η) andξ,ηareQ-linearly dependent.\nOtherwise, we say that visergodic4. We define the averaging operator along v:\nf/ma√sto→ A(f)v(z) := lim\nT→∞1\nT/integraldisplayT\n0f(z+tv)dt,\nwhere the limit exists, thanks to Weyl’s equidistribution t heorem.\nLemma 4.1. Assume that v∈S1. Then for any non-negative function f∈ Dm,k,σ(T2),m≥10, the\naveraged function A(f)v∈ Dm,k,σ(T2).\nProof.First we assume that v= (ξ,η) is periodic. This implies that the orbit z/ma√sto→z+tvis periodic,\nandTvis the period. Clearly, for any f∈ Dm,k,σ,\nA(f)v(z) =1\nTv/integraldisplayTv\n0f(z+tv)dt.\nSince the function s/ma√sto→ |s|1−|α|σis concave, by Jensen’s inequality we have\n1\nTv/integraldisplayTv\n0|f(z+tv)|1−|α|σdt≤/parenleftBig1\nTv/integraldisplayTv\n0f(z+tv)dt/parenrightBig1−|α|σ\n. (4.1)\nIndeed, if/integraltextTv\n0f(z+tv)dt= 0, thenf(z+tv)≡0 for allt∈[0,Tv], and the inequality ( 4.1) is trivial.\nAssume now that X0=1\nTv/integraltextTv\n0f(z+tv)dt>0, then for any X≥0, we have\nX1−|α|σ≤X1−|α|σ\n0+(1−|α|σ)X−|α|σ\n0(X−X0).\nReplacing the inequality above by X=f(z+tv) and averaging over t∈[0,Tv], we obtain ( 4.1). Since\nby definition, |∂αf|/lessorsimilarα,σ|f|1−|α|σfor all|α| ≤k, we get\n|∂α(A(f)v)(z)|/lessorsimilarα,σ|A(f)v(z)|1−|α|σ.\nNext we assume that v= (ξ,η) is ergodic. In this case, the orbit z/ma√sto→z+tvis ergodic, then by\nWeyl’s equidistribution theorem, we have\nAv(f)(z) =/hatwidef(0) =1\n(2π)2/integraldisplay\nT2f(z′)dz′. (4.2)\n4These definitions stemmed from the fact that we are on the two- dimensional torus T2.SHARP DECAY RATE FOR THE DAMPED WAVE EQUATION WITH CONVEX-SH APED DAMPING 15\nTo prove this, by the assumption of v, we havek·v/\\e}atio\\slash= 0 for allk∈Z2\\{0}. Moreover, since f∈Cm,\nwe can write\n1\nT/integraldisplayT\n0f(z+tv)dt=1\nT/integraldisplayT\n0/summationdisplay\nk∈Z2/hatwidef(k)eik·(z+tv)dt=/hatwidef(0)+/summationdisplay\nk/negationslash=0/hatwidef(k)eik·z·eiTk·v−1\niT(k·v).\nAs|eiTk·v−1| ≤2|iTk·v|for allk∈Z2and\nlim\nT→∞eiTk·v−1\niTk·v= 0,\nby the dominated convergence theorem, we get ( 4.2), which means that Av(f) is a constant function.\nClearly,A(f)v(z)∈ Dm,k,σ(T2). The proof of Lemma 4.1is now complete. /square\nBy the triangle inequality, the following Lemma is immediat e:\nLemma 4.2. Letv∈S1. For any function fonT2, there holds\n|A(f)v| ≤ A(|f|)v.\nMoreover, if f1,f2are two non-negative functions such that f1≤f2, we have\nA(f1)v≤ A(f2)v.\nLemma 4.3. Assume that f∈ Dm,k,σ(T2)andf≥0. Denote by\nF(x,y) :=/integraldisplayy\n−π/parenleftbig\nf(x,y′)−A(f)e2(x)/parenrightbig\ndy′,−π0}is a disjoint union of strictly convex curves with positive c urvature. Assume moreover\nthat there exists R>0such that for every z∈ {a>0}nearΣa,\nR−1·dist(z,Σa)1\nσ≤a(z)≤R·dist(z,Σa)1\nσ. (4.3)\nThen for any periodic direction v∈S1, we have A(a)v∈ Dm,k,2σ\nσ+2, as a one-dimensional periodic\nfunction. Furthermore, there exists Rv>0, such that for every z∈ {A(a)v(z)>0}nearΣA(a)v, we\nhave\nR−1\nvdist(z,ΣA(a)v)1\nσ+1\n2≤ A(a)v(z)≤Rvdist(z,ΣA(a)v)1\nσ+1\n2. (4.4)\nProof.Since the vector vis periodic, we can find p0,q0∈Z, gcd(p0,q0) = 1, such that v=(p0,q0)√\np2\n0+q2\n0.\nNow we perform a change of coordinate as described in Section 2.2. Recall that we have a covering\nmapπv:T2\nv/ma√sto→T2of degreep2\n0+q2\n0that lifts every 2 πperiodic function to a 2 π/radicalbig\np2\n0+q2\n0-periodic\nfunction. Moreover, the change of coordinate system z/ma√sto→(X,Y) is given by z=Xv⊥+Yvlocally.\nAsπvis locally isometric, each component of the pre-image of the damped region π−1\nv(ω) is still\nstrictly convex with positive curvature, so the inequality (4.3) is preserved, near the boundary of each\ncomponent. Denote by τ= 2π/radicalbig\np2\n0+q2\n0. We define the averaging operator on the new torus T2\nvas\n/tildewideA(F)(X,Y) :=1\nτ/integraldisplayτ/2\n−τ/2F(X,Y+t)dt.\nthen by definition\nπ∗(A(a)v)(X,Y) =A(a)v(π(Xv⊥+Y)) =1\nτ/integraldisplayτ\n0a◦π(Xv⊥+(Y+t)v)dt=1\nτ/integraldisplayτ\n0(π∗a)(X,Y+t)dt.\nThusπ∗\nv/parenleftbig\nA(a)v/parenrightbig\n=/tildewideA(π∗\nva).Therefore, if we are able prove ( 4.4) for the lifted averaging function\n/tildewideA(π∗\nv(a)), we obtain automatically ( 4.4) by projection.\nInsummary, from theargument above, withoutloss of general ity, weassumethat v= e2andassume\nthat Ω := {a>0}hasl=l(v) connected components Ω 1,···,Ωlsuch that the boundary Σ a,jof each\nΩjhas positive curvature. We first consider the situation wher el= 1. By translation invariance, we\nmay assume that Ω 1={a1>0}is contained in the fundamental domain ( −Kπ,Kπ)x×(−Mπ,Mπ )y.\nThen the function A(a)e2can be identified as a function on Rx,\nSince Ω 1={a1>0}is strictly convex, each line PxofR2, passing through ( x,0) and parallel to e 2\ncan intersect at most 2 points of the curve Σ a,1. Consider the function x/ma√sto→P(x) := mes( Px∩Ω1).\nThis function is continuous and is supported on a single inte rvalI= (α,β)⊂(−Kπ,Kπ). Since\nA(a1)(x) = 0 ifP(x) = 0, the vanishing behavior of A(a1) is determined when xis close toαandβ.\nBelow we only analyze A(a1)(x) forx∈[α,α+ǫ), since the analysis is similar for xnearβ. First we\nobserve that Pαmust be tangent at a point z0:= (α,y0) to the curve Σ a,1. For sufficiently small ǫ>0,\nwe may parametrize the curve Σ a,1nearz0by the function x=α+g(y) withg(y0) =g′(y0) = 0 and\ng′′(y0) =c0>0, thanks to the fact that the curvature is strictly positive . Therefore, there exists a C1\ndiffeomorphism Y= Φ(y) from a neighborhood of y0to a neighborhood of Y= 0 such that Φ( y0) = 0\nandg(y) =Y2. For eachx∈(α,α+ǫ),Px∩Σa,1={(x,l−(x)),(x,l+(x))}. We have\nl+(x) = Φ−1(√\nx−α), l−(x) = Φ−1(−√\nx−α).SHARP DECAY RATE FOR THE DAMPED WAVE EQUATION WITH CONVEX-SH APED DAMPING 17\nΩ1\nxz0\nαα+ǫ(x,l−(x))(x,l−(x))Px\nAveraging improves the local H¨ older regularity\nSince near z0, we havea(z)∼dist(z,Σa)1\nσ= dist(z,Σa,1)1\nσ∼(x−(α+g(y)))1\nσ\n+. Forx∈(α,α+ǫ),\nwe have\nA(a1)(x) =1\n2Kπ/integraldisplayKπ\n−Kπa1(x,y)dy=1\n2Kπ/integraldisplayl+(x)\nl−(x)a1(x,y)dy\n=1\n2Kπ/integraldisplay√x−α\n−√x−α|x−α−Y2|1\nσ|(Φ−1)′(Y)|dY∼Φ,K,σ|x−α|1\nσ+1\n2.\nHere the implicit constant depends on Φ ,K,σ, hence it depends on the periodic direction vand the\ndampinga(z).\nFinally, since a1∈ Dm,k,σ, forj≤k,x∈(α,α+ǫ),\n|∂j\nxA(a1)(x)| ≤1\n2Kπ/integraldisplayKπ\n−Kπ|∂j1xa1(x,y)|dy/lessorsimilarK/integraldisplayl+(x)\nl−(x)a1−jσ\n1(x,y)dy\n/lessorsimilar|x−α|1−jσ\nσ+1\n2∼(A(a1)(x))1−2σj\nσ+2.\nThis implies that A(a1)∈ Dm,k,2σ\nσ+2.\nTo complete the proof of Proposition 4.4, we need to deal with the situation where l >1. In this\ncase, the supports of A(aj) may overlap. By linearity and the inequality\n|∂j′\nxA(a1+···al)| ≤l/summationdisplay\nj=1|∂j′\nxA(aj)| ≤C(a1,···,al)A(a1+···+al)1−2σj′\nσ+2,\nwe deduce that A(a)∈ Dm,k,2σ\nσ+2. It remains to show ( 4.4). We define\nS+:=/braceleftbig\nj∈ {1,···,l}:/integraldisplayx0+ǫ\nx0A(aj)(x)dx>0,∀ǫ>0/bracerightbig\n,\nS−:=/braceleftbig\nj∈ {1,···,l}:/integraldisplayx0\nx0−ǫA(aj)(x)dx>0,∀ǫ>0/bracerightbig\n.\nObserve that x0∈ΣA(a)if and only if A(aj)(x0) = 0 for all j∈ {1,···,l}andS+∪S−/\\e}atio\\slash=∅.Note that\nforj∈S±, we have\ndist(x,ΣA(a))1\n2+1\nσ= dist(x,ΣA(aj))1\n2+1\nσ∼ A(aj)(x),∀x∓x0>0 nearx0,\nandA(aj) = 0, in a neighborhood of x0, for allj /∈S+∪S−. Summing over j∈S+∪S−, we obtain\n(4.4). The proof of Proposition 4.4is now complete. /square18 CHENMIN SUN\n5.Normal form reductions\nNow we treat the transversal low frequency part vh=ψ(/planckover2pi1Dx)u1\nh, defined in ( 3.3). We want to\naverage the operator Ph+ihaalong the direction e 2= (0,1). Recall that A(a) is the averaging of a\nalong the vertical direction. Recall that vhsatisfies the equation\n(Ph+iha)vh=r1,h=oL2(h2δh) =oL2(h/planckover2pi12).\nWe will apply successively two normal form reductions. The fi rst reduction replaces ihabyihA(a),\nwith aO(h2) anti-selfadjoint remainder which cannot be absorbed dire ctly as a remainder of size\nO(h2δh). Weneedtoperformasecondnormalformreductiontoaverag etheanti-selfadjoint remainder.\n5.1.The first averaging. Throughout this section, we denote by\nA(x,y) :=/integraldisplayy\n−π(a(x,y′)−A(a)(x))dy′.\nWe will also fix a cutoff ψ1∈C∞\nc(R), supported on |η±1| ≤1\n2andψ1(η) = 1 if|η±1| ≤1\n4. We need\nthe following basic lemma for exponentials of bounded linea r operators:\nLemma 5.1. Let(Gh)00,\n5Note that A(a) commutes with ∂j\nxGh.24 CHENMIN SUN\nwe deduce that\n/ba∇dblA(a)1\n2v(1)\nh/ba∇dbl2\nL2≤1\n2/ba∇dblA(a)1\n2v(1)\nh/ba∇dbl2\nL2+Cmax{h2,(h/planckover2pi1−1)1\nσ,h2\n1+2σ}/ba∇dblv(1)\nh/ba∇dbl2\nL2+o(/planckover2pi12)\nNote thatσ<1\n4sinceβ≥4, andh3δ−1\nh=o(1), the second term on the right hand side is o(/planckover2pi12). This\ncompletes the proof of Lemma 5.5. /square\nFinally, we note that the principal symbol of −1\nh2[Gh,[Gh,Ph]] is\nq0(x,y,ξ,η) =−∂ηg∂y(2ξ∂xg+2η∂yg)+∂xg∂ξ(2ξ∂xg+2η∂yg)+∂yg∂η(2ξ∂xg+2η∂yg).\nSinceg(x,y,η) = 2b(η)A(x,y) = 2b(η)/integraltexty\n−π(a(x,y′)− A(a)(x))dy′, we deduce from Lemma 4.3, the\nfact thata,A(a)∈ Dm,2,σ,σ<1\n4, that\n|q0(x,y,ξ,η)|2≤Ca(x,y)+CA(a)(x)+q1(x,y,ξ,η),\nwhere supp( q1)∩WFm\nh(v(1)\nh) =∅. Therefore, by the sharp G˚ arding inequality,\n/ba∇dbl1\nh2[Gh,[Gh,Ph]]v(1)\nh/ba∇dblL2≤C/ba∇dbla1\n2v(1)\nh/ba∇dblL2+C/ba∇dblA(a)1\n2v(1)\nh/ba∇dblL2+O(h1\n2) =o(δh).\nHence [Gh,[Gh,Ph]]v(1)\nh=oL2(h2δh) =oL2(h/planckover2pi12).The proof of Proposition 5.2is now complete. /square\n5.2.The second averaging. In this subsection, we prove the following proposition of th e second\nnormal form reduction. Unlike the first normal form which is l ess perturbative (the operator eGhis\nnot close to the identity), we are able to make the normal form transform close to the identity, in the\nspirit of [ BZ12] (see also [ BS19],[LeS] for related applications to the Bouendi-Grushin operator s):\nProposition 5.6. There exist a real-valued symbol g1(x,y,η)inS0and the associated pseudo-differential\noperatorG1,h= Oph(g1), a function κ(x)∈W1,∞(Tx), the Fourier multiplier b(hDy), such that\nv(2)\nh:= (Id−G1,hhDx)−1v(1)\nhsatisfies the equation\n(Ph+ihA(a))v(2)\nh+ihκ(x)A(a)1\n2b(hDy)hDxv(2)\nh=r4,h=oL2(h/planckover2pi12).\nMoreover,\n/ba∇dblv(2)\nh−v(1)\nh/ba∇dblL2=O(h/planckover2pi1−1),/ba∇dblA(a)1\n2v(2)\nh/ba∇dblL2=o(/planckover2pi1).\nThe importance of the above proposition is that it makes poss ible to take the Fourier transform\ninyvariable and reduce the equation of v(2)\nhmode-by-mode to one-dimensional ordinary differential\nequations.\nTo prove Proposition 5.6, we want to average the non self-adjoint part [ h2D2\nx,Gh] in the equation\n(Ph+iA(a))v(1)\nh−[h2D2\nx,Gh]v(1)\nh=/tildewiderh=oL2(h/planckover2pi12).\nWe first identify the principal part of this lower order non-s elfadjoint part:\nLemma 5.7. We have\n[h2D2\nx,Gh]v(1)\nh=−4ih2/planckover2pi1−1(∂xA)b(hDy)/planckover2pi1Dxv(1)\nh+oL2(h/planckover2pi12).\nMoreover,\n[h2D2\nx,Gh]v(1)\nh=oL2(h2).SHARP DECAY RATE FOR THE DAMPED WAVE EQUATION WITH CONVEX-SH APED DAMPING 25\nProof.Recall that Gh=b(hDy)A+Ab(hDy), hence\n[h2D2\nx,Gh] = [[h2D2\nx,A],b(hDy)]+2b(hDy)[h2D2\nx,A].\nThe principal symbol of −1\nh2[[h2D2\nx,A],b(hDy)] is\nq2(x,y,ξ,η) =−2ξb′(η)(∂xa−A′(a)(x)).\nThus from Lemma 4.3,|∂xa|+|A′(a)| ≤Ca1\n2+CA(a)1\n2, thus\n|q2(x,y,ξ,η)|2≤Ca(x,y)+CA(a)(x)+q3(x,y,ξ,η),\nwhere supp( q3)∩WFm\nh(v1\nh) =∅. By the sharp G˚ arding inequality (Theorem B.2),\n/ba∇dbl[[h2D2\nx,A],b(hDy)]v(1)\nh/ba∇dblL2≤O(h5\n2) =o(h/planckover2pi12).\nIt remains to treat b(hDy)[h2D2\nx,A]. Note that [ h2D2\nx,A] =h2(D2\nxA)+2h(DxA)hDx.From Lemma\n5.3we may replace v(1)\nhby/tildewideψ(/planckover2pi1Dx)v(1)\nh. Therefore,\n[h2D2\nx,A]v(1)\nh=h2(D2\nxA)v(1)\nh+2h2/planckover2pi1−1b(hDy)(DxA)/planckover2pi1Dx/tildewideψ(/planckover2pi1Dx)v(1)\nh+OL2(h3).\nFrom Lemma 4.3,|Dj\nxA|/lessorsimilarA(a)1−jσ≤ A(a)1\n2forj= 1,2, we have\nh2b(hDy)(D2\nxA)v(1)\nh=oL2(h2/planckover2pi1).\nNext, we write\nhb(hDy)(DxA)hDx=h2/planckover2pi1−1[b(hDy),DxA]/planckover2pi1Dx+h2/planckover2pi1−1(DxA)b(hDy)/planckover2pi1Dx.\nBy the symbolic calculus, h2/planckover2pi1[b(hDy),DxA]/planckover2pi1Dx/tildewideψ(/planckover2pi1Dx)v(1)\nh=OL2(h3/planckover2pi1−1).This completes the proof\nof Lemma 5.7. /square\nProof of Proposition 5.7.Thanks to Lemma 5.7, we can write\n(Ph+ihA(a))v(1)\nh+ihQ1,hhDxv(1)\nh=r2,h=oL2(h/planckover2pi12),\nwhere\nQ1,h=−4(∂xA)b(hDy), b(η) =−χ(η)\n4η.\nNote that the principal symbol of Q1,his independent of ξvariable. Now we perform a second normal\nform transform. Recall that\nA(x,y) =/integraldisplayy\n−π(a(x,y′)−A(a)(x))dy′.\nConsider the ansatz v(1)\nh= (1−G1,hhDx)v(2)\nh, whereG1,hwill be chosen such that G1,hhDx=\nOL(L2)(h/planckover2pi1−1). The new quasi-modes v(2)\nhsatisfy the equation\n(Ph+ihA(a)+ihQ1,hhDx)v(2)\nh−[h2D2\ny,G1,hhDx]v(2)\nh=r3,h\nwhere\nr3,h=(1−G1,hhDx)−1r2,h+(1−G1,hhDx)−1[h2D2\nx+ihA(a)+ihQ1,hhDx,G1,hhDx]v(2)\nh\n+/parenleftbig\n(1−G1,hhDx)−1−Id/parenrightbig\n[h2D2\ny,G1,hhDx]v(2)\nh. (5.11)26 CHENMIN SUN\nNote that if G1,hhDx=OL(L2)(h/planckover2pi1−1), the operator (1 −G1,hhDx) is invertible for sufficiently small h\n(thus/planckover2pi1). In particular,\nv(2)\nh= (1−G1,hhDx)−1v(1)\nh=∞/summationdisplay\nn=0(G1,hhDx)nv(1)\nh=9/summationdisplay\nn=0(G1,hhDx)nv(1)\nh+OL2(h10/planckover2pi1−10),\nwhere the last error term is oL2(h/planckover2pi12). Since from Lemma 5.3, we may replace v(1)\nhby/tildewideψ(/planckover2pi1Dx)v(1)\nh,\nmodulo an error of OL2(hN\n2) for anyN≤m, we may also replace v(2)\nhby/tildewideψ(/planckover2pi1Dx)v(2)\nhimplicitly in the\nargument below. Therefore, with G1,h= Oph(g1)/tildewideψ(/planckover2pi1Dx), we have\n(Ph+ihA(a))v(2)\nh+ihQ1,hhDxv(2)\nh+ihOph({η2,g1})hDxv(2)\nh+h2ChhDxv(2)\nh=r3,h,\nwhereChis uniformly bounded on L2(T2). Note that the last term on the right hand size is of size\nOL2(h3/planckover2pi1−1). Now we set6\ng1(x,y,η) =2b(η)\nη/integraldisplayy\n−π/parenleftbig\n∂xA(x,y′)−A(∂xA)(x)/parenrightbig\ndy′.\nThen we have 2 η∂yg1−4(∂xA)b(η) =−4A(∂xA)(x)b(η). By the symbolic calculus, modulo an error of\nsizeOL2(h3/planckover2pi1−1), wecanreplace ih(Q1,h+Oph({η2,g1}))hDxv(2)\nhby−4ihA(∂xA)b(hDy)/tildewideψ(/planckover2pi1Dx)hDxv(2)\nh.\nTherefore, the equation of v(2)\nhbecomes\n(Ph+ihA(a)−4ihA(∂xA)b(hDy)hDx)v(2)\nh=r4,h,\nwhere\nr4,h=r3,h+OL2(h3/planckover2pi1−1) =r3,h+oL2(h/planckover2pi12).\nIt is clear that /ba∇dblv(2)\nh−v(1)\nh/ba∇dblL2=O(h/planckover2pi1−1) =o(1),and from the relation\nv(2)\nh=v(1)\nh+G1,hhDxv(1)\nh+OL2(h2/planckover2pi1−2),\nwe deduce that A(a)1\n2v(2)\nh=oL2(/planckover2pi1).\nSet\nκ:=−4A(∂xA)\nA(a)1\n2,\nto complete the proof, we need to verify that\n(i)κ∈W1,∞(Tx);\n(ii)r3,h=oL2(h/planckover2pi12).\nTo verify (i), observe that A′(f)(x) =A(∂xf)(x). By Lemma 4.2, Lemma 4.3and the fact that σ<1\n4,\nwe have\n|A(∂j\nxA)| ≤ A(|∂j\nxA|)≤CA(a)1−jσ≤CA(a)1\n2,∀j= 1,2.\nThis shows that κ,κ′are bounded.\nIt remains to prove (ii). Recall ( 5.11) and the fact that (1 −G1,hhDx)−1−Id =OL(L2)(h/planckover2pi1−1), it\nsuffices to show that\n[h2D2\nx+ihA(a)+ihQ1,hhDx,G1,hhDx]v(2)\nh=oL2(h/planckover2pi12) (5.12)\n6Recall that the support of b(η) is away from η= 0.SHARP DECAY RATE FOR THE DAMPED WAVE EQUATION WITH CONVEX-SH APED DAMPING 27\nand\n[h2D2\ny,G1,hhDx]v(2)\nh=oL2(/planckover2pi13). (5.13)\nDenote by\nA2(x,y) =/integraldisplayy\n−π/parenleftbig\n∂xA(x,y′)−A(∂xA)(x)/parenrightbig\ndy′.\nPointwise, we have\n|A2|/lessorsimilarA(|∂xA|)/lessorsimilarA(a)1−σ,|∂xA2|/lessorsimilarA(|∂2\nxA|)/lessorsimilarA(a)1−2σ\nand\n|∂yA2|/lessorsimilar|∂xA|+A(|∂xA|)/lessorsimilarA(a)1−σ,\nthanks to Lemma 4.3. Therefore,\n|∇jA2|v(2)\nh=oL2(/planckover2pi1), j= 0,1. (5.14)\nNote that by the symbolic calculus,\nih[Q1,hhDx,G1,hhDx] =ih(h/planckover2pi1−1)2[Q1,h/planckover2pi1Dx,G1,h/planckover2pi1Dx] =OL(L2)(h3/planckover2pi1−1),\nwhich isoL(L2)(h/planckover2pi12) sinceh2=o(/planckover2pi13). For the other terms, if we only apply the symbolic calculus , we\nwill gain only O(h2)+O(h3/planckover2pi1−2) for (5.12) andO(h2/planckover2pi1−1) for (5.13), which are not enough to conclude.\nWe need to open the definition of G1,h. Sinceh2=o(/planckover2pi13), it suffices to take into account the principal\npart ofG1,h. Therefore, without loss of generality, we assume that G1,h=A2(x,y)b1(hDy)/tildewideψ(/planckover2pi1Dx),\nwithb1(η) =2b(η)\nη. Note that any commutator will generate at least one more /planckover2pi1, the main contribu-\ntion of [h2D2\nx,G1,hhDx] is−2ihb1(hDy)/tildewideψ(/planckover2pi1Dx)h2D2\nx(∂xA2)v(2)\nhwhoseL2norm iso(h3/planckover2pi1−1) =o(h/planckover2pi12),\nthanks to ( 5.14). Similarly, modulo acceptable errors from the commutator s, the main contribution\nofih[A(a),G1,hhDx]v(2)\nhis\nih2/planckover2pi1−1b1(hDy)[A(a),/tildewideψ(/planckover2pi1Dx)/planckover2pi1Dx]A2v(2)\nh,\nwhoseL2norm is, thanks to ( 5.14), bounded by O(h2)/ba∇dblA2v(2)\nh/ba∇dblL2=o(h/planckover2pi12). This verifies ( 5.12). By\nthe same argument, to verify ( 5.13), we note that, modulo acceptable errors from commutators, the\nmain contribution of [ h2D2\ny,G1,hhDx]v(2)\nhis\n−ih/planckover2pi1−12h·hDyb1(hDy)/tildewideψ(/planckover2pi1Dx)/planckover2pi1Dx(∂yA2)v(2)\nh,\nwhich is of size oL2(h2) =oL2(/planckover2pi13), by (5.14). This verifies ( 5.13) and the proof of Proposition 5.6is\nnow complete. /square\n6.One-dimensional resolvent estimate\nFrom Proposition 5.6,v(2)\nhsatisfies the equation\n(Ph+ihA(a))v(2)\nh+ihκ(x)A(a)1\n2b(hDy)hDxv(2)\nh=r4,h=oL2(h/planckover2pi12), (6.1)\nand/ba∇dblv(2)\nh/ba∇dblL2=O(1),/ba∇dblA(a)1\n2v(2)\nh/ba∇dblL2=o(/planckover2pi1). In this section, we are going to show that /ba∇dblv(2)\nh/ba∇dblL2=o(1).\nSince the left hand side of ( 6.1) commutes with Dy, by taking the Fourier transform in y, are can\nreduce the analysis to a sequence of one-dimensional proble ms.28 CHENMIN SUN\n6.1.1D resolvent estimate for the H¨ older damping. In order to finish the proof of Theorem 1.2,\nit remains to prove a one-dimensional resolvent estimate. B elow we establish a slightly more general\nversion. By abusing a bit the notation, we denote by vh,E∈H2(Tx), solutions of equations\n−h2∂2\nxvh,E−Evh,E+ihW(x)vh,E+h2κh,E(x)W(x)1\n2∂xvh,E=rh,E. (6.2)\nWe assume that ( κh,E)h>0,E∈Ris a uniform bounded family in W1,∞(T;R).\nProposition 6.1. Assume that W∈ Dm,2,θ(Tx),θ≤1\n4be a non-negative function such that the set\n{W(x)>0}is a disjoint unions of finitely many intervals Ij= (αj,βj)⊂Tx,j= 1,···,land\nC−1(x−αj)1\nθ\n+≤W(x)≤C(x−αj)1\nθ\n+inIjnearαj (6.3)\nand\nC−1(βj−x)1\nθ\n+≤W(x)≤C(βj−x)1\nθ\n+inIjnearβj, (6.4)\nfor allj∈ {1,···,l}. Then there exists h0∈(0,1)andC0>0, such that for all h∈(0,h0)and all\nE∈R, the solutions vh,Eof(6.2)satisfy the uniform estimate\n/ba∇dblvh,E/ba∇dblL2≤C0h−2−θ\n2θ+1/ba∇dblrh,E/ba∇dblL2+C0h−3θ+1\n2(2θ+1)/ba∇dblW1\n2vh,E/ba∇dblL2. (6.5)\nRemark 6.2. We will reduce the proof, in the low-energy hyperbolic regim e to a known one-\ndimensional resolvent estimate (Proposition 6.9), which is the main result of [ DKl]. However, in\nthe paper of [ DKl], the final gluing argument is not clear to the author. For thi s reason as well as\nself-containedness, we will reprove Proposition 6.9(thus Theorem 1.3) in Appendix A.\nWe postpone the proof of Proposition 6.1for the moment and proceed on proving Theorem 1.2.\nLetθ=2\n2β+1andk= 2m,δ=θ\n2θ+1, and/planckover2pi1=h1+δ\n2=h1\n2δ1\n2\nh=h3θ+1\n2(2θ+1). LetW(x) =A(a)(x). By\nProposition 4.4,W∈ Dm,2,θ(Tx) andWsatisfies ( 6.3), (6.4) near the vanishing points inside the\ndamped region. Take the Fourier transform in yfor (6.1) and denote by v(2)\nh,n(x) =Fy(v(2)\nh)(x,n), we\nhave\n(−h2∂2\nx+h2n2−1+ihW(x)+h2κ(x)W(x)1\n2b(hn)∂x)v(2)\nh,n=Fyr4,h.\nRecall that /ba∇dblr4,h/ba∇dblL2(T2)=o(h/planckover2pi12) =o(h2+δ) and/ba∇dblW1\n2v(2)\nh/ba∇dblL2(T2)=o(/planckover2pi1) =o(h1+δ\n2). LetE= 1−h2n2\nandκh,E(x) =κ(x)b(hn) which is uniformly bounded in W1,∞(T) with respect to handn. Applying\nProposition 6.1for each fixed n∈Zand then taking the l2\nnnorm, by Plancherel we get\n/ba∇dblv(2)\nh/ba∇dblL2/lessorsimilarh−2−δ/ba∇dblr4,h/ba∇dblL2(T2)+h−1+δ\n2/ba∇dblW1\n2v(2)\nh/ba∇dblL2(T2)=o(1).\nRecall that from Proposition 5.2and Proposition 5.6,\nv(2)\nh= (Id−G1,hhDx)−1v(1)\nh= (Id−G1,hhDx)−1◦eGhvh.\nThus/ba∇dblvh/ba∇dblL2=o(1), and this contradicts to ( 2.1). The proof of Proposition 2.1(as well as Theorem\n1.2) is now complete.SHARP DECAY RATE FOR THE DAMPED WAVE EQUATION WITH CONVEX-SH APED DAMPING 29\nNow we prove Proposition 6.1. In what follows, we note that δ=θ\n2θ+1≤1\n6. We argue by\ncontradiction. If ( 6.5) is untrue, by normalization, we may assume that there exist sequenceshn→\n0,(En)n∈N⊂Rand (vhn,En)n∈N⊂L2, (rhn,En)n∈N⊂L2, such that\n(−h2\nn∂2\nx−En+ihnW(x)+h2\nnκhn,En(x)W(x)1\n2∂x)vhn,En=rhn,En (6.6)\nand\n/ba∇dblvhn,En/ba∇dblL2= 1,/ba∇dblrhn,En/ba∇dblL2=o(h2+δ\nn),/ba∇dblW1\n2vhn,En/ba∇dblL2=o(h1+δ\n2n). (6.7)\nIn what follows, when we use the asymptotic notations as smal loandO, we mean a limit (or bound)\nindependent of the sequences hn→0 andEn, asn→ ∞. To simplify the notation, we will sometimes\nomit the subindex nin the sequel. For a function f, sometimes we denote by f′=∂xf. Also, when\nwe write /lessorsimilar,/greaterorsimilar, the implicit bounds are independent ofhandE.\nWe record an elementary weighted energy identity which allo ws us to deal with the elliptic regime\nwhereE≪h2.\nLemma 6.3 (Weighted energy identity) .Letw∈C2(T;R), then\n/integraldisplay\nTw(x)|h∂xvh,E|2dx+/integraldisplay\nT(−1\n2h2∂2\nxw−Ew)|vh,E|2dx−h2\n2/integraldisplay\nT(wκh,EW1\n2)′|vh,E|2dx= Re/integraldisplay\nTwrh,Evh,Edx.\n(6.8)\nProof.Multiplying ( 6.6) bywvh,Eand integrating over T, taking the real part and using the relation\n(|vh,E|2)′= 2Re(v′\nh,Evh,E), we get\nRe/integraldisplay\nTh2(wvh)′v′\nh,Edx−/integraldisplay\nTEw|vh,E|2dx−h2\n2/integraldisplay\nT(wκh,EW1\n2)′|vh,E|2dx= Re/integraldisplay\nTwrh,Evh,Edx.\nTo finish the proof, we just write Re( w′vh,Ev′\nh,E) =1\n2w′(|vh,E|2)′and integrate by part. /square\nBy choosing w= 1 and using the fact that κ′\nh,Eis uniformly bounded in L∞(T) and (W1\n2)′/lessorsimilarW1\n4,\nwe have\nh2|((κh,EW1\n2)′,|vh,E|2)L2|/lessorsimilarh2/ba∇dblW1\n8vh,E/ba∇dbl2\nL2≤h2/ba∇dblvh,E/ba∇dbl3\n2\nL2/ba∇dblW1\n2vh,E/ba∇dbl1\n2\nL2=o(h9+δ\n4).\nSinceδ≤1\n6, we have:\nCorollary 6.4 (Energy identity) .There holds\n/ba∇dblh∂xvh,E/ba∇dbl2\nL2−E/ba∇dblvh,E/ba∇dbl2\nL2=o(h2+δ).\nThe proof of Proposition 6.1will be divided into several steps, according to the range of E.\n•(A) Elliptic regime E≪h2:Recall that Wis supported on disjoint intervals Ij= (αj,βj)⊂\n(−π,π),j= 1,···,l. Therefore, we are able to construct a weight w∈C2(T;R) such that\nw≥c0>0, w′′<0,in a neighborhood of T\\∪l\nj=1Ij.\nTherefore, there exists c1>0, sufficiently small, such that\n−1\n2w′′(x)−c1w>0,in a neighborhood of T\\∪l\nj=1Ij.30 CHENMIN SUN\nLemma 6.5. IfE≤c1h2, the solution vhn,Ensatisfies\n/ba∇dblvh,E/ba∇dblL2/lessorsimilarh−2/ba∇dblrh,E/ba∇dblL2+/ba∇dblW1\n2vh,E/ba∇dblL2.\nProof.SinceE≤c1h2, we have1\n2h2w′′+Ew <0 in a neighborhood of T\\∪l\nj=1Ij.Thus there exists\na compact set K⊂ ∪l\nj=1Ijsuch that\n/integraldisplay\nT(1\n2h2∂2\nxw+Ew)|vh,E|2dx≤/integraldisplay\nK(1\n2h2∂2\nxw+Ew)|vh,E|2dx/lessorsimilarh2/integraldisplay\nTW(x)|vh,E|2dx.\nThen applying Lemma 6.3, we have\nc0/ba∇dblh∂xvh,E/ba∇dbl2\nL2≤/integraldisplay\nTw(x)|h∂xvh,E|2dx\n≤Re/integraldisplay\nTwrh,Evh,Edx+h2/integraldisplay\nTW(x)|vh,E|2dx+Ch2\n2/integraldisplay\nT(wκh,EW1\n2)′|vh,E|2dx.\nNote that |(wκh,EW1\n2)′|/lessorsimilarW1\n4(x), by interpolation\n/ba∇dblW1\n8vh,E/ba∇dbl2\nL2≤ /ba∇dblW1\n2vh,E/ba∇dbl1\n2\nL2/ba∇dblvh,E/ba∇dbl3\n2\nL2\nand Young’s inequality, we deduce that\n/ba∇dblv′\nh,E/ba∇dblL2≤Ch−1/ba∇dblvh,E/ba∇dbl1\n2\nL2/ba∇dblrh,E/ba∇dbl1\n2\nL2+Cǫ/ba∇dblW1\n2vh,E/ba∇dblL2+ǫ/ba∇dblvh,E/ba∇dblL2,∀ǫ>0.\nBy the Poincar´ e-Wirtinger inequality,\n/vextenddouble/vextenddoublevh,E−/hatwidevh,E(0)/vextenddouble/vextenddouble\nL2(T)≤C/ba∇dblv′\nh,E/ba∇dblL2,\nwhere/hatwidevh,E(0) =1\n2π/integraltext\nTvh,E. Combining with the fact that/integraltext\nTW >0 and the elementary inequality\n/parenleftBig/integraldisplay\nTWdx/parenrightBig\n|/hatwidevh,E(0)|2≤C/integraldisplay\nTW(x)|vh,E(x)|2dx+C/integraldisplay\nTW(x)|vh,E(x)−/hatwidevh,E(0)|2dx,\nwe deduce that\n/ba∇dblvh,E/ba∇dblL2+/ba∇dblv′\nh,E/ba∇dblL2/lessorsimilarh−1/ba∇dblwvh,E/ba∇dbl1\n2\nL2/ba∇dblrh,E/ba∇dbl1\n2\nL2+/ba∇dblW1\n2vh,E/ba∇dblL2.\nUsingYoung’s inequality again to absorb /ba∇dblvh,E/ba∇dblL2to theleft, wecomplete theproofof Lemma 6.5./square\n•(B) High energy hyperbolic regime E > h1+δ:In this regime, we put the damping terms to\nthe right as remainders and use the estimate from the geometr ic control as a black box. Let us recall:\nLemma 6.6. LetI⊂Tbe a non-empty open set. Then there exists C=CI>0, such that for any\nv∈L2(T),f1∈L2(T),f2∈H−1(T),λ≥1, if\n(−∂2\nx−λ2)v=f1+f2,\nwe have\n/ba∇dblv/ba∇dblL2(T)≤Cλ−1/ba∇dblf1/ba∇dblL2(T)+C/ba∇dblf2/ba∇dblH−1(T)+C/ba∇dblv/ba∇dblL2(I).\nThe proof is standard and can be found, for example in [ Bu19] (Proposition 4.2). In the one-\ndimensional setting, a straightforward proof using the mul tiplier method is also available. Conse-\nquently, we have:SHARP DECAY RATE FOR THE DAMPED WAVE EQUATION WITH CONVEX-SH APED DAMPING 31\nCorollary 6.7. IfE >h1+δ, then\n/ba∇dblvh,E/ba∇dblL2(T)/lessorsimilarh−3+δ\n2/ba∇dblrh,E/ba∇dblL2(T)+h−1+δ\n2/ba∇dblWvh,E/ba∇dblL2(T)+/ba∇dblW1\n2vh,E/ba∇dblL2.\nProof.Letλ=h−1E1\n2(≥h−1−δ\n2), then\n(−∂2\nx−λ2)vh,E=h−2rh,E−ih−1Wvh,E−κh,EW1\n2∂xvh,E.\nApplying Lemma 6.6tov=vh,E,f1=h−2rh,E−ih−1Wvh,E,f2=−κh,EW1\n2v′\nh,EwithI= (α1+\nǫ0,β1−ǫ0) for someǫ0<β1−α1\n2, we get\n/ba∇dblvh,E/ba∇dblL2(T)/lessorsimilarǫ0λ−1/ba∇dblh−2rh,E−ih−1Wvh,E/ba∇dblL2(T)+/ba∇dblκh,EW1\n2v′\nh,E/ba∇dblH−1(T)+/ba∇dblvh,E/ba∇dblL2(I)\n/lessorsimilarh−3+δ\n2/ba∇dblrh,E/ba∇dblL2+h−1+δ\n2/ba∇dblWvh,E/ba∇dblL2+/ba∇dbl(κh,EW1\n2vh,E)′−(κh,EW1\n2)′vh,E/ba∇dblH−1(T).\nSince (κh,EW1\n2)′/lessorsimilarW1\n4, the last term on the right hand side is bounded by\nC/ba∇dblW1\n4vh,E/ba∇dblL2≤C/ba∇dblW1\n2vh,E/ba∇dbl1\n2\nL2/ba∇dblvh,E/ba∇dbl1\n2\nL2.\nBy Young’s inequality, we obtain the desired estimate. /square\n•(C) Low energy hyperbolic regime: c1h20}is disjoint unions\nof intervals Ij= (αj,βj),j= 1,2,···,land that for each j∈ {1,···,l},\nC1Vj(x)≤W(x)≤C2Vj(x)on(αj,βj),\nwhereVj(x)>0are continuous functions on (αj,βj)such that\nVj(x) =\n\n(x−αj)γ, αj0,c1>0,C >0, such that for all 0< h < h 0,√c1≤λ≤h−1−δ\n2and all\nsolutionsvh,λof the equation\n−v′′\nh,λ−λ2vh,λ+ih−1W(x)vh,λ=rh,λ,\nwe have\n/ba∇dblvh,λ/ba∇dblL2≤Ch−1\nγ+2/ba∇dblrh,λ/ba∇dblL2. (6.11)\nThe proof of Proposition 6.9will be given in Appendix A.\nNow applying Proposition 6.9forγ=1\nθ(then1\nγ+2=δ=θ\n2θ+1),vh,λ=vh,Eandrh,λ=h−2rh,E−\nκh,EW1\n2v′\nh,Ein our previous setting, combining with Lemma 3.2, we deduce that when c1h2≤E <\nh1+δ,\n/ba∇dblvh,E/ba∇dblL2≤Ch−2−δ/ba∇dblrh,E/ba∇dblL2+h−1+δ\n2/ba∇dblW1\n2vh,E/ba∇dblL2+h−δ/ba∇dblW1\n2vh,E/ba∇dbl1\n2\nL2/ba∇dblvh,E/ba∇dbl1\n2\nL2.(6.12)\nFinally, to get a contradiction, we denote three index sets f or three regimes:\nEA:={n:En≤c1h2\nn},EB:={n:En>h1+δ\nn},EC:={n:c1h2\nn0\nF′(0) = 1,(7.7)\nthenvh,l(x) takes the form hδαhF/parenleftbigr0−x\nhδ;θh/parenrightbig\n, withθh=h−2(γ+1)\nγ+2λ2\nh, and a constant O(1) =αh∈Cto\nbe determined in order to match the compatibility condition (7.5).\nDenote byµ0be the lowest Neumann eigenvalue of the operator −∂2\ny+xγonL2(R+) andF0the\nDirichelet trace F(x;0)|x=0. It was shown that µ0>0 (Lemma 4.1 of [ Kl]). Moreover, by using the\nimplicit function theory (Lemma 4.2 of [ Kl]), there exists a uniform constant C0>0, such that for all\n|θ| ≤µ0\n2, the (unique) solution F(y;θ) of the Neumann problem ( 7.7) satisfies\n1\nC0≤ |F(0;θ)| ≤C0. (7.8)\nIn particular, F0=F(0;0)/\\e}atio\\slash= 0.\nNow we are ready to solve ( 7.4) with the compatibility condition ( 7.5). In view of ( 7.6), we precise\nthe ansatz of λhas\nλh=π/parenleftbig\nl+1\n2/parenrightbig\nh\nr0+γhh1+δ, O(1) =γh∈C.SHARP DECAY RATE FOR THE DAMPED WAVE EQUATION WITH CONVEX-SH APED DAMPING 35\nNowαh,γhare the parameters to be determined. Plugging into ( 7.5) with\nvh,l(x) =hδαhF/parenleftbigr0−x\nhδ;θh/parenrightbig\n,\nwe obtain a system\nαhF(0;θh) = (−1)l+1h−δsin(γhr0hδ),\nαh= (−1)l/bracketleftbigπ/parenleftbig\nl+1\n2/parenrightbig\nr0+γhhδ/bracketrightbig\ncos(γhr0hδ). (7.9)\nSince|θh| ∼h2\nγ+2, by Taylor expansion of sin and cos, the leading term of γhshould beγ0:=π/parenleftbig\nl+1\n2/parenrightbig\nF0\nr2\n0\nand the leading term for αhshould beα0= (−1)lπ/parenleftbig\nl+1\n2/parenrightbig\ny0. Fix the number l, by using the implicit\nfunction theorem, the solution ( αh,γh) to (7.9) exists for 0 ≤h≪1 and\n|(αh,γh)−(α0,γ0)|/lessorsimilarhδ.\nFor the detailed argument, we refer Lemma 4.3 and Lemma 4.4 of [Kl].\nFinally, we take h=hk=r0 /radicalBig\nk2r2\n0+π2(l+1\n2)2, then 1−h2\nkk2=λ2\nhk+O(h2+δ\nk), hence\nuh(x,y) =eikyvh(x) =eiky/bracketleftbig\ncos/parenleftbigλh|x|\nh/parenrightbig\n1r0<|x|≤π+hδαhF/parenleftbigr0−|x|\nhδ;ηh/parenrightbig\n1|x|≤r0/bracketrightbig\n(7.10)\nare the desired T2quasimodes, satisfying\n−h2∆uh−uh+ihb0(x)uh=OL2(h2+δ).\nWe are going to prove more estimates on uhas well as its transformations:\nProposition 7.1. Letuhis given by (7.10)andψ∈C∞\nc(R)be a bump function supported near 0.\nThen/tildewideuh:=ψ(h1+δ\n2Dx)uhsatisfies\n(−h2∆−1+ihb0(x))/tildewideuh=OL2(h2+δ).\nMoreover, there exist uniform constants h0>0,C1>0, such that for all 00.\nWe need a Lemma:\nLemma 7.2. Letµ0be the least eigenvalue of the operator Aγ=−∂2\nx+xγonL2(R+)associated\nto the Neumann boundary condition. Then there exists a unifor m constant C >0, such that for all\n|θ| ≤µ0\n4, the solution F(x;θ)of(7.7)satisfies\n/ba∇dblF/ba∇dblH2(R+)≤C.36 CHENMIN SUN\nProof.TakeH∈C∞\nc([0,∞)) such that H(0) = 0 and H′(0) = 1. Consider /tildewideF:=F−H, then\n−/tildewideF′′+ixγ/tildewideF−θ/tildewideF=W,\nwithW=H′′−ixγH+θH. Multiplying by /tildewideFand doing the integration by part, we get/integraldisplay∞\n0|/tildewideF′(x)|2dx+i/integraldisplay∞\n0xγ|/tildewideF(x)|2dx−θ/integraldisplay∞\n0|/tildewideF(x)|2dx=/integraldisplay∞\n0W(x)/tildewideF(x)dx.\nTaking the real part and imaginary part, we get\n/ba∇dbl/tildewideF′/ba∇dbl2\nL2(R+)≤ |Reθ|/ba∇dbl/tildewideF/ba∇dbl2\nL2(R+)+/ba∇dblW/ba∇dblL2(R+)/ba∇dbl/tildewideF/ba∇dblL2(R+),\nand\n/ba∇dblxγ\n2/tildewideF/ba∇dbl2\nL2(R+)≤ |Imθ|/ba∇dbl/tildewideF/ba∇dbl2\nL2(R+)+/ba∇dblW/ba∇dblL2(R+)/ba∇dbl/tildewideF/ba∇dblL2(R+).\nAdding two inequalities above, using Aγ−µ0≥0 and the fact that |θ| ≤µ0\n4, we get\nµ0/ba∇dbl/tildewideF/ba∇dbl2\nL2(R+)≤µ0\n2/ba∇dbl/tildewideF/ba∇dbl2\nL2(R+)+2/ba∇dblW/ba∇dblL2(R+)/ba∇dbl/tildewideF/ba∇dblL2(R+).\nThis proves the boundedness of /ba∇dbl/tildewideF/ba∇dblL2(R+)+/ba∇dbl∂x/tildewideF/ba∇dblL2(R+)+/ba∇dblxγ\n2/tildewideF/ba∇dblL2(R+)in terms of /ba∇dblW/ba∇dblL2(R+).\nReplacing /tildewideFby/tildewideF′, the same argument yields the boundedness of /ba∇dbl/tildewideF′′/ba∇dblL2(R+)in terms of /ba∇dblW/ba∇dblH1(R+).\nSince theH1bound forWis uniform with respect to |θ| ≤µ0\n4, the proof of Lemma 7.2is complete.\n/square\nProof of Proposition 7.1.Denote by /planckover2pi1=h1+δ\n2as before. The fact that /tildewideuh=ψ(/planckover2pi1Dx)uhsatisfies the\nequation of quasimodes is clear from Lemma 3.2, up to changing oL2(h/planckover2pi12) there toOL2(h/planckover2pi12). We need\nto prove estimates only.\nAt the first step, we prove the same estimates (a),(b),(c) for uh. The inequality (a) is clear by the\nconstruction and Lemma 2.2. For (b), since |k| ∼1\nh, we have /ba∇dblhj∂j\nyuh/ba∇dblL2/lessorsimilar1 for allj∈N. The\nderivatives of uhinxsatisfy\n|∂j\nxuh|/lessorsimilar1+1|x|≤r0h−(j−1)δ|(∂j\nxF)/parenleftbigr0−|x|\nhδ;θh/parenrightbig\n|,\nby Lemma 7.2, we deduce that /ba∇dblhj∂j\nxuh/ba∇dblL2/lessorsimilarhj−(j−1\n2)δ, forj= 1,2. Finally, since ∂xuh=\nF′/parenleftbigr0−|x|\nhδ;θh/parenrightbig\n,on supp(b′\n0), we have\n/ba∇dblb′\n0(x)∂xuh/ba∇dblL2/lessorsimilar/vextenddouble/vextenddouble(r0−|x|)γ−1\n+F′/parenleftbigr0−|x|\nhδ;θh/parenrightbig/vextenddouble/vextenddouble\nL2/lessorsimilarhδ\n2·hδ(γ−1)=hγ−1/2\nγ+2=h1−5δ\n2≤hδ,(7.11)\nthanks toδ=1\nγ+2andγ >2. Next,\n/vextenddouble/vextenddouble/vextenddouble/parenleftBig/integraldisplayy\n−π(∂xa0(x,y′)−b′\n0(x))dy′/parenrightBig\n∂xuh/vextenddouble/vextenddouble/vextenddouble\nL2/lessorsimilar/vextenddouble/vextenddouble/vextenddouble/parenleftBig/integraldisplayπ\n−π|∂xa0(x,y′)|dy′/parenrightBig\n∂xuh/vextenddouble/vextenddouble/vextenddouble\nL2(|x|≤r0)+/ba∇dbl|b′\n0(x)|∂xuh/ba∇dblL2(|x|≤r0).\nSince for |x| ≤r0,\n/integraldisplayπ\n−π|∂xa0(x,y′)|dy′=/integraldisplay√\nr2\n0−x2\n−√\nr2\n0−x2|∂xa0(x,y′)|dy′/lessorsimilar(r0−|x|)γ−1\n+,\nthe same computation as ( 7.11) yields\n/vextenddouble/vextenddouble/vextenddouble/parenleftBig/integraldisplayy\n−π/parenleftbig\n∂ya0(x,y′)−b′\n0(x)/parenrightbig\ndy′/parenrightBig\n∂xuh/vextenddouble/vextenddouble/vextenddouble\nL2/lessorsimilarh1−5δ\n2≤hδ.SHARP DECAY RATE FOR THE DAMPED WAVE EQUATION WITH CONVEX-SH APED DAMPING 37\nAs the second step, we deal with /tildewideuh. As the Fourier multiplier ψ(/planckover2pi1Dx) commutes with derivatives,\nthe upper bounds for /ba∇dbl/tildewideuh/ba∇dblL2,/ba∇dblhj∂j\nx/tildewideuh/ba∇dblL2and/ba∇dblhj∂j\ny/tildewideuh/ba∇dblL2follow directly from the estimates for uh,\nhence (b) holds for /tildewideuh. To prove the lower bound of /ba∇dbl/tildewideuh/ba∇dblL2, we take a smooth function χ∈C∞(R)\nsuch thatχ≡1 on the support of 1 −ψ. Then we write\nuh−/tildewideuh=/parenleftbig\n∂−1\nxχ(/planckover2pi1Dx)/parenrightbig\n·∂x(1−ψ(/planckover2pi1Dx))uh.\nSince/ba∇dbl∂xuh/ba∇dblL2/lessorsimilarh−δ/2and/ba∇dbl∂−1\nxχ(/planckover2pi1Dx)/ba∇dblL(L2)/lessorsimilar/planckover2pi1, we obtain that /ba∇dbluh−/tildewideuh/ba∇dblL2/lessorsimilar/planckover2pi1h−δ=o(1).\nTherefore, the assertions (a) follows for /tildewideuh. By the commutator estimate\n/ba∇dbl[f,ψ(/planckover2pi1Dx)]/ba∇dblL(L2)/lessorsimilar/planckover2pi1,\nforf∈W1,∞and the fact that /ba∇dbl∂xuh/ba∇dblL2/lessorsimilarh−δ\n2, we deduce that\n/vextenddouble/vextenddouble/vextenddouble/parenleftBig/integraldisplayy\n−π(∂xa0(x,y′)−b′\n0(x))dy′/parenrightBig\n∂x/tildewideuh/vextenddouble/vextenddouble/vextenddouble\nL2+/ba∇dblb′\n0(x)∂x/tildewideuh/ba∇dblL2/lessorsimilar/planckover2pi1·h−δ\n2+hδ/lessorsimilarhδ.\nThe last step is to prove estimates (a),(b),(c) for /tildewidevh=e−Gh/tildewideuhwhereGh= Opw\nh(g). Sincee−Gh\nis invertible and is uniformly bounded, we get /ba∇dbl/tildewidevh/ba∇dblL2∼1. Moreover, since Ghcommutes with the\nmultiplication by functions depending only in x, we have /ba∇dblb0(x)1\n2/tildewidevh/ba∇dblL2/lessorsimilarh1+δ\n2. So (a) holds for /tildewidevh.\nTo prove (b),(c) for /tildewidevh, from the same estimates for /tildewideuh, it suffices to control the commutator terms\ninvolving the operator e−Gh. Applying the formula\n[B2,e−Gh] = 2[B,e−Gh]B+[B,[B,e−Gh]],\ntoB=h∂x,h∂yand using ( 5.2) and (5.4), we deduce that for j= 1,2,\n/ba∇dbl[hj∂j\nx,e−Gh]/tildewideuh/ba∇dblL2/lessorsimilarhj,/ba∇dbl[hj∂j\ny,e−Gh]/tildewideuh/ba∇dblL2/lessorsimilarhj,\nhence (b) follows for /tildewidevh.\nTo estimate /ba∇dbl[b′\n0(x)∂x,e−Gh]/tildewideuh/ba∇dblL2, sinceGh,esGhboth commute with b′\n0(x), by formula ( 5.4), we\nhave\nb′\n0(x)[∂x,e−Gh] = [∂x,e−Gh]b′\n0(x).\nFrom (5.2),/ba∇dbl[∂x,e−Gh]/ba∇dblL(L2)/lessorsimilar1 and|b′\n0|/lessorsimilarb1/2\n0, we deduce that\n/ba∇dbl[b′\n0(x)∂x,e−Gh]/tildewideuh/ba∇dblL2/lessorsimilarh1+δ\n2.\nFinally, to estimate /ba∇dbl[(∂xA)∂x,e−Gh]/tildewideuh/ba∇dblL2, where\nA(x,y) :=/integraldisplayy\n−π(a0(x,y′)−b0(x))dy′,\nwe write\n[(∂xA)∂x,e−Gh]/tildewideuh= [(∂xA),e−Gh]∂x/tildewideuh+(∂xA)[∂x,e−Gh]/tildewideuh.\nSince/ba∇dbl[(∂xA),e−Gh]/ba∇dblL(L2)/lessorsimilarh, together with the estimate (b) for /tildewideuh, we have\n/ba∇dbl[(∂xA),e−Gh]∂x/tildewideuh/ba∇dblL2/lessorsimilarh1−δ\n2/lessorsimilarhδ.\nFurther commuting ∂xAand [∂x,e−Gh], by (5.4), we can write\n(∂xA)[∂x,e−Gh] = [∂x,e−Gh](∂xA)+OL(L2)(h).38 CHENMIN SUN\nTherefore,\n/ba∇dbl(∂xA)[∂x,e−Gh]/tildewideuh/ba∇dblL2/lessorsimilar/ba∇dbl(∂xA)/tildewideuh/ba∇dblL2+O(h).\nBy Lemma 4.3, together with the fact that |∇a0|/lessorsimilara1\n2\n0,|b′\n0|/lessorsimilarb1/2\n0and Jensen’s inequality, we have\n|(∂xA)(x,y)|/lessorsimilarb0(x)1\n2.Therefore, we obtain that /ba∇dbl(∂xA)/tildewideuh/ba∇dblL2/lessorsimilarh1+δ\n2≤hδ, hence (c) follows.\nThe proof of Proposition 7.1is now complete. /square\n7.2.Proof of the optimality. Take quasimodes /tildewideuhin Proposition 7.1. Define /tildewidevh=e−Gh/tildewideuh, we are\ngoing to show that /tildewidevhare the desired quasimodes, satisfying\n(−h2∆−1+iha0)/tildewidevh=OL2(h2+δ),/ba∇dbl/tildewidevh/ba∇dblL2∼1. (7.12)\nThen (7.12) implies ( 7.1).\nRecall the notations b0(x) =A(a0)(x),Gh= Opw\nh(g(x,y,η)), where\ng(x,y,η) =−ψ1(η)\n2η/integraldisplayy\n−π(a0(x,y′)−b0(x))dy′.\nFirst, we prove:\nProposition 7.3. The quasimodes /tildewidevh=e−Gh/tildewideuhsatisfy the equation\n(−h2∆−1+iha0)/tildewidevh+[h2D2\nx,Gh]/tildewidevh=OL2(h/planckover2pi12),\nwhere/planckover2pi1=h1+δ\n2. Moreover, /tildewidevhverifies properties (a),(b),(c)in Proposition 7.1and\n/ba∇dbla1\n2\n0/tildewidevh/ba∇dblL2+/ba∇dblb1\n2\n0/tildewidevh/ba∇dblL2=O(/planckover2pi1).\nProof.Theproof is very similar to the proof of Proposition 5.2, but much simpler, since we have better\nestimates for /tildewidevh=e−Gh/tildewideuh, thanks to Proposition 7.1. Consider the conjugate operator\n/tildewideFh(s) :=e−sGh(Ph+ihb0(x))esGh,\nwherePh=−h2∆−1. Similar to ( 5.5), we obtain a simpler formula\n/tildewideFh(1) =e−Gh(Ph+ihb0(x))eGh=Ph+ih/parenleftbig\nb0−i\nh[h2D2\ny,Gh]/parenrightbig\n+[h2D2\nx,Gh]+1\n2[Gh,[Gh,Ph]]+OL(L2)(h2+δ). (7.13)\nHere we use the fact that [ Gh,b0(x)] = 0. As the principal symbol ofi\nh[h2D2\ny,Gh] isψ1(η)(b0−a0),\nwe have\n(Ph+iha0+[h2D2\nx,Gh])/tildewidevh=fh+OL2(h2+δ), (7.14)\nwhere\nfh=ih/parenleftbig\nb0−a0−i\nh[h2D2\ny,Gh]/parenrightbig\n/tildewidevh−1\n2[Gh,[Gh,Ph]]/tildewidevh. (7.15)\nIt remains to show that fh=OL2(h2+δ).\nFirst of all, the analogue of Lemma 5.3holds with the same proof:SHARP DECAY RATE FOR THE DAMPED WAVE EQUATION WITH CONVEX-SH APED DAMPING 39\nLemma 7.4. Let/tildewideψbe any cutoff such that /tildewideψ≡1on the support of ψthat defined /tildewideuhin Proposition\n7.1. Then\nWF10\nh(/tildewidevh)⊂WFh(/tildewideuh)\nand\n/ba∇dbl(1−/tildewideψ(/planckover2pi1Dx))/tildewidevh/ba∇dblL2=O(/planckover2pi14) =O(h2+2δ).\nSimilarly, the analogue of Lemma 5.4holds. Therefore, we have\n(Ph+iha0+[h2D2\nx,Gh])/tildewidevh+1\n2[Gh,[Gh,Ph]]/tildewidevh=OL2(h2+δ).\nNext, we claim that\n/ba∇dbla1\n2\n0/tildewidevh/ba∇dblL2=O(/planckover2pi1) =O(h1+δ\n2). (7.16)\nThis follows by mimicking the proof of Lemma 5.5with the new operator\nQh=a0−i\nh[h2D2\nx,Gh].\nIndeed, by multiplying by /tildewidevhto the equation and taking the imaginary part, we have\n(Qh/tildewidevh,/tildewidevh)L2=O(/planckover2pi12).\nHence\n/ba∇dbla1\n2\n0/tildewidevh/ba∇dbl2\nL2+i\nh([h2D2\nx,Gh]/tildewidevh,/tildewidevh) =O(/planckover2pi12).\nBy (5.8), (5.9) and (5.10), we have\n/vextendsingle/vextendsinglei\nh([h2D2\nx,Gh]/tildewidevh,/tildewidevh)/vextendsingle/vextendsingle/lessorsimilarh/ba∇dblb0(x)1\n2/tildewidevh/ba∇dblL2/ba∇dbl/tildewidevh/ba∇dblL2+h/ba∇dblb0(x)1\n2hDx/tildewidevh/ba∇dblL2/ba∇dblb0(x)1\n2−σ/tildewidevh/ba∇dblL2,\nwhereσ<1\n4. Note that by Lemma 7.4, we can write /tildewidevh=ψ(/planckover2pi1Dx)/tildewidevh+OL2(h2+2δ), hence\n/ba∇dblb0(x)1\n2hDx/tildewidevh/ba∇dblL2≤h/planckover2pi1−1/ba∇dblb0(x)1\n2/planckover2pi1Dxψ(/planckover2pi1Dx)/tildewidevh/ba∇dblL2+O(h2+2δ)≤h/planckover2pi1−1/ba∇dbl[b1\n2\n0,/planckover2pi1Dxψ(/planckover2pi1Dx)]/tildewidevh/ba∇dblL2+O(h)\n≤O(h),\nthanks to the symbolic calculus. Therefore,\n/vextendsingle/vextendsinglei\nh([h2D2\nx,Gh]/tildewidevh,/tildewidevh)/vextendsingle/vextendsingle≤O(h2).\nIn particular, we obtain ( 7.16). Finally, the estimate\n/ba∇dbl[Gh,[Gh,Ph]]/tildewidevh/ba∇dblL2=O(h2+δ)\nfollows from the same argument in the last paragraph of the pr oof of Proposition 5.2. The proof of\nProposition 7.3is now complete. /square\nIn view of Proposition 7.1and Proposition 7.3, to complete the proof of ( 7.12), it remains to show\nthat\n/ba∇dbl[h2D2\nx,Gh]/tildewidevh/ba∇dblL2=O(h/planckover2pi12). (7.17)\nBy the same argument as in the proof of Lemma 5.7, we have\n[h2D2\nx,Gh]/tildewidevh=−4ih2/planckover2pi1−1(∂xA)b(hDy)/planckover2pi1Dx/tildewidevh+OL2(h/planckover2pi12),40 CHENMIN SUN\nwhere\nA(x,y) =/integraldisplayy\n−π(a0(x,y′)−b0(x))dy′, b(hDy) =−χ(hDy)\n4hDy.\nSince/ba∇dbl[(∂xA),b(hDy)]/ba∇dblL(L2)=O(h), from (c) of Proposition 7.1, we deduce that\n/ba∇dbl4ih2/planckover2pi1−1(∂xA)b(hDy)/planckover2pi1Dx/tildewidevh/ba∇dblL2≤O(h3/planckover2pi1−1)+O(h2+δ) =O(h2+δ),\nthanks to the fact that δ<1\n3. This proves ( 7.17).\nAppendix A.Proof of Proposition 6.9\nIn this appendix, we prove Proposition 6.9. Note that the proof works also for γ= 0, thus covering\nthe main result in [ St] for the piecewise constant rectangular damping. Without l oss of generality,\nwe assume that ∪l\nj=1Ij/\\e}atio\\slash=T, otherwise, we can apply Theorem 1.7 of [ LLe] and the corresponding\nresolvent estimate h2\nγ+2is much better than ( 6.11).\nRecall the numerology: δ=1\nγ+2.To simplify the notation in the exposition, we argue by contr adic-\ntion. We assume that there exists a sequence hn→0 andλn⊂Rsuch thatc1\n2\n1≤λn≤h−1−δ\n2n, such\nthat\n−v′′\nhn,λn−λ2\nnvhn,λn+ih−1W(x)vhn,λn=rhn,λn (A.1)\nand\n/ba∇dblvhn,λn/ba∇dblL2= 1,/ba∇dblrhn,λn/ba∇dblL2=o(hδ\nn).\nFor simplicity, we will ignore the subindex nforhn,λnand write simply v=vh,λ,r=rh,λsometimes\nwithout displaying their dependence in handλ. First we record the a priori estimate, for which the\nproof is a direct consequence of integration by part (see the proof of Lemma 2.2)\nLemma A.1.\n(a)/ba∇dblv′/ba∇dbl2\nL2−λ2/ba∇dblv/ba∇dbl2\nL2=o(hδ).\n(b)/ba∇dblW1\n2v/ba∇dblL2=o(h1+δ\n2).\nPickχ∈C∞(R;[0,1]) such that χ(s)≡1 whens≤1 andχ(s)≡0 whens >2. Denote by\nIj= (αj,βj) and without loss of generality, we assume that\n−π<α1<β1≤α2<β2≤ ··· ≤αl<βl≤π.\nDefine the function\nV0(x) :=l/summationdisplay\nj=1max{0,(x−αj)(βj−x)}.\nNotethatV0(x) = (x−αj)(βj−x)whenx∈Ijforsomej∈ {1,···,l}andV0(x) = 0whenever W(x) =\n0. Defineχh(x) :=χ/parenleftBig\nV0(x)3\nh3δ/parenrightBig\n. Note that χh∈C2. Denote by Ij,h= (αj,αj+2πhδ)∪(βj−2πhδ,βj)\nforj∈ {1,···,n}and/tildewideIj,h= (αj+σhδ,αj+2πhδ)∪(βj−2πhδ,βj−σhδ). Here the constant σ<2πSHARP DECAY RATE FOR THE DAMPED WAVE EQUATION WITH CONVEX-SH APED DAMPING 41\nis chosen so that χh|Ijis constant on I\\/tildewideIj,h. Hence supp( χ′\nh), supp(χ′′\nh) are all contained in ∪l\nj=1/tildewideIj,h,\nand\n|χ′\nh|/lessorsimilarh−δl/summationdisplay\nj=11/tildewideIj,h,|χ′′\nh|/lessorsimilarh−2δ/summationdisplay\nj=11/tildewideIj,h. (A.2)\nWe decompose\nv=v1+v2, v1=χhv, v2= (1−χh)v.\nWhenγ >0,v2is supported in the damped region W/greaterorsimilarhαwhilev1is supported on W/lessorsimilarhα, where\nα=γ\nγ+2.\nLemma A.2. Forv2, we have\n/ba∇dblv2/ba∇dblL2=o(h3\n2(γ+2)).\nProof.First we assume that γ >0, then from (b) of Lemma A.1,\n/ba∇dblv2/ba∇dblL2≤h−α\n2/ba∇dblW1\n2v/ba∇dblL2≤o(h1+δ−α\n2) =o(h3\n2(γ+2)).\nWhenγ= 0,δ=1\n2, by Lemma 5.5,/ba∇dblv2/ba∇dblL2≤ /ba∇dblW1\n2v/ba∇dblL2≤o(h3\n4). /square\nIt remains to estimate v1. We see that v1solves the equation\n−v′′\n1−λ2v1+ih−1Wv1=/tildewider:=χhr−2(χ′\nhv)′+χ′′\nhv. (A.3)\nRemark A.3. IfWisC2, the choice of cutoff χhis the same as χ(h−αW). Note that the parameter\nαis chosen so that the size of ih−1Wvandχ′′\nhvare the same in L2. This choice of cutoff is more\naccurate than the choice χ(h−1W) in [BH05], in order to balance the size of ih−1Wvandχ′′\nhvcoming\nfrom the commutator on the right hand side of ( A.3).\nIfλis relatively large, we are still able to apply the estimate f rom the geometric control:\nLemma A.4. Ifλ≥h−δ\n2, we have\n/ba∇dblv1/ba∇dblL2/lessorsimilarhδ\n2/ba∇dblr/ba∇dblL2+h−1+δ\n2/ba∇dblW1\n2v/ba∇dblL2.\nProof.By Lemma 6.6, we have\n/ba∇dblv1/ba∇dblL2/lessorsimilarλ−1/ba∇dblχhr−ih−1Wv1+χ′′\nhv/ba∇dblL2+/ba∇dbl(χ′\nhv)′/ba∇dblH−1+/ba∇dblWv1/ba∇dblL2\n/lessorsimilarhδ\n2/ba∇dblr/ba∇dblL2+h−1+α+δ\n2/ba∇dblW1\n2v/ba∇dblL2+h−3δ\n2l/summationdisplay\nj=1/ba∇dblv/ba∇dblL2(/tildewideIj,h)+hα\n2/ba∇dblW1\n2v/ba∇dblL2,\nwhere we use the fact that W/lessorsimilarhαon supp(χh). SinceW∼hαon/tildewideIj,handα+2δ= 1, we have\n/ba∇dblv/ba∇dblL2(/tildewideIj,h)/lessorsimilarh−α\n2/ba∇dblW1\n2v/ba∇dblL2/lessorsimilarh−1\n2+δ/ba∇dblW1\n2v/ba∇dblL2.\nThis completes the proof of Lemma A.4. /square42 CHENMIN SUN\nIt remains to deal with the regime where c1\n2\n1≤λ < h−δ\n2, the key point is to exclude the possible\nconcentration of the energy density\ne0(x) :=|v′\n1(x)|2+λ2|v1(x)|2\nin the damped shell of size hδnear the interface where W= 0. The tool used in [ DKl] is a Morawetz\ntype inequality introduced:\nLemma A.5 (Morawetz type inequality) .LetΦ∈C0(T)be a piece-wise C1function on T, then there\nexists a uniform constant C >0, such that/integraldisplay\nTΦ′(x)e0(x)dx≤Ch−1/vextendsingle/vextendsingle/vextendsingle/integraldisplay\nTΦ(x)W(x)v1v′\n1dx/vextendsingle/vextendsingle/vextendsingle\n+C/vextendsingle/vextendsingle/vextendsingleRe/integraldisplay\nTΦ(x)v′\n1·(χhr−2(χ′\nhv)′+χ′′\nhv)dx/vextendsingle/vextendsingle/vextendsingle. (A.4)\nProof.Direct computation yields\n(Φe0)′= Φ′e0+2Re(Φv′′\n1v′\n1+λ2Φ(|v1|2)′).\nUsing the equation ( A.3), we have\n(Φe0)′=Φ′e0+λ2Φ∂x(|v1|2)−2λ2Re(Φv′\n1v1)+2h−1Re(iΦv′\n1·Wv1)\n−2Re[Φv′\n1(χhr−2(χ′\nhv)′+χ′′\nhv)].\nSince 2λ2Re(Φv′\n1v1) =λ2Φ·∂x(|v1|2), integrating the above identity, we obtain the desired est imate\n(A.4). /square\nLetǫj<βj−αj\n2,j∈ {1,···,n}. Define\nΨh(x) :=\n\nh−δ, x∈ ∪l\nj=1Ij,h;\n1, x∈ ∪l\nj=1(αh+2πhδ,αj+ǫj)∪(βj−ǫj,βj−2πhδ);\n−M, x∈ ∪l\nj=1Ij\\((αj,αj+ǫj)∪(βj−ǫj,βj));\n1, x∈T\\∪l\nj=1Ij\nwhere the constant M >0 (independent of h) is chosen such that/integraltext\nTΨh(x)dx= 0. Then the primitive\nfunction Φ h∈C0(T) is well-defined, piecewise smooth and Φ′\nh(x) = Ψh(x). Since Φ his unique up to\na constant, we choose Φ hsuch that Φ h(0) = 0, then /ba∇dblΦh/ba∇dblL∞(T)≤CM. Define\nΘ(x) = Φ′\nh(x)1Φ′\nh>0,\nsince supp( e0)⊂supp(χh), we have Φ′\nh(x)e0(x) = Θ(x)e0(x). From Lemma A.5andv′\n1=χ′\nhv+χhv′,\nwe have/integraldisplay\nTΘ(x)e0(x)dx≤Ch−1/vextendsingle/vextendsingle/vextendsingle/integraldisplay\nTW(x)(χ2\nhvv′+2χhχ′\nh|v|2)dx/vextendsingle/vextendsingle/vextendsingle\n+C/integraldisplay\nT|χhr(χ′\nhv+χhv′)|dx+C/integraldisplay\nT|χ′′\nhv(χ′\nhv+χhv′)|dx\n+C/vextendsingle/vextendsingle/vextendsingleRe/integraldisplay\nTχ′\nhvΦh(x)χhv′′dx/vextendsingle/vextendsingle/vextendsingle+C/integraldisplay\nT|χ′\nh(Φhχh)′vv′|dx+C/integraldisplay\nT|χ′\nhv(Φh(x)χ′\nhv)′|dx,\n(A.5)SHARP DECAY RATE FOR THE DAMPED WAVE EQUATION WITH CONVEX-SH APED DAMPING 43\nwhere the terms in the third line of the right side is obtained by integration by part of\n/vextendsingle/vextendsingle/vextendsingleRe/integraldisplay\nTΦh(x)v′\n1(χ′\nhv)′dx/vextendsingle/vextendsingle/vextendsingle.\nHere we keep the real part for this term in order to perform som e cancellation when replacing v′′\nby−λ2v+ih−1Wv−rlater on, after using the equation ( A.1) ofv. Denote by /tildewideIh=∪l\nj=1/tildewideIj,hand\nIh=∪l\nj=1Ij,h. Using ( A.2) and the facts that λ2≤h−δ,W∼hαon/tildewideIh, we can control terms on the\nright hand side of ( A.5) by the sum of following types:\nType I : I = h−δ/ba∇dblv1/tildewideIh/ba∇dblL2/ba∇dblr/ba∇dblL2+/ba∇dblr/ba∇dblL2/ba∇dblv′χh/ba∇dblL2;\nType II : II = h−1/integraldisplay\nTW|vv′|1/tildewideIhdx+h−1/integraldisplay\nTWχ2\nh|vv′|dx;\nType III : III = h−(1+δ)/ba∇dblW1\n2v1/tildewideIh/ba∇dbl2\nL2.\nWe analyze the type II term. By Cauchy-Schwarz,\nII/lessorsimilarh−1/ba∇dblW1\n2v1Ih/ba∇dblL2/ba∇dblW1\n21Ihv′/ba∇dblL2. (A.6)\nNote that on Ih, Θ(x) =h−δ, we have\nII/lessorsimilarh−1+α\n2/ba∇dblW1\n2v/ba∇dblL2/ba∇dbl1Ihv′/ba∇dblL2≤h−1+α\n2+δ\n2/ba∇dblW1\n2v/ba∇dblL2/parenleftBig/integraldisplay\nTΘ(x)e0(x)dx/parenrightBig1\n2.\nSinceα+2δ= 1, using Young’s inequality, we have\nII≤ǫ/integraldisplay\nTΘ(x)e0(x)dx+Cǫh−(1+δ)/ba∇dblW1\n2v/ba∇dbl2\nL2,∀ǫ>0.\nPlugging into ( A.5), we get\n/ba∇dblv1/ba∇dbl2\nL2≤/integraldisplay\nTΘ(x)e0(x)dx/lessorsimilarh−(1+δ)/ba∇dblW1\n2v/ba∇dbl2\nL2+h−δ/ba∇dblv/ba∇dblL2/ba∇dblr/ba∇dblL2+/ba∇dblr/ba∇dblL2/ba∇dblv′/ba∇dblL2.\nSinceλ≤h−δ\n2, by (a) of Lemma A.1,\n/ba∇dblv′/ba∇dbl2\nL2≤o(hδ)+λ2/ba∇dblv/ba∇dbl2\nL2≤o(h−δ).\nRecall that /ba∇dblr/ba∇dblL2=o(hδ) and/ba∇dblW1\n2v/ba∇dblL2=o(h1+δ\n2), we get\n/ba∇dblv1/ba∇dbl2\nL2/lessorsimilar/integraldisplay\nTΘ(x)e0(x)dx/lessorsimilaro(1).\nSince we have already shown that /ba∇dblv2/ba∇dblL2=o(1), this is a contradiction to our assumption that\n/ba∇dblv/ba∇dblL2= 1. The proof of Proposition 6.9is now complete.\nAppendix B.Semiclassical pseudo-differential calculus\nForm∈R, the symbol class Sm(T∗Rd) consists of smooth functions c(z,ζ) such that\n|∂α\nz∂β\nζc(z,ζ)| ≤Cα,β/a\\}b∇acketle{tζ/a\\}b∇acket∇i}htm−|α|.\nGiven a symbol c(z,ζ), we associate it with the Weyl quantization Opw\nh(c):\nf∈ S(Rd)/ma√sto→Opw\nh(f)(z) :=1\n(2πh)d/integraldisplay/integraldisplay\nR2dc/parenleftbigz+z′\n2,ζ/parenrightbig\nei(z−z′)·ζ\nhf(z′)dz′dζ.44 CHENMIN SUN\nMost of the time we will use the standard quantization Oph(c):\nOph(c)(f)(z) :=1\n(2πh)d/integraldisplay/integraldisplay\nR2dc(z,ζ)ei(z−z′)·ζ\nhf(z′)dz′dζ.\nAn important mapping property is the following theorem due t o Calder´ on-Vaillancourt:\nTheorem B.1. There exists a constant C >0such that for any function conT∗Rdwith uniformly\nbounded derivatives up to order d, we have\n/ba∇dblOph(c)/ba∇dblL(L2(Rd))+/ba∇dblOpw\nh(c)/ba∇dblL(L2(Rd))≤C/summationdisplay\n|α|,|β|≤dh|β|/ba∇dbl∂α\nz∂β\nζc/ba∇dblL∞(T∗Rd).\nWe use frequently the sharp G˚ arding inequality:\nTheorem B.2 (Sharp G˚ arding’s inequality) .Assume that c∈S0(T∗Td)andc(z,ζ)≥0for all\n(z,ζ)∈T∗Td. Then there exist C >0andh0>0such that for all 00, such that for all |α| ≤d+1,\n|∂α\nξb(x,y,ξ)| ≤Cα/a\\}b∇acketle{tξ/a\\}b∇acket∇i}ht−(|α|+µ0).\nThen the operator Thassociated with the Schwartz kernel\nKh(x,y) =1\n(2πh)d/integraldisplay\nRdb(x,y,ξ)ei(x−y)·ξ\nhdξ\nis bounded on L2(Rd), uniformly in h∈(0,1].\nProof.The proof with µ0= 1 can be found in Lemma A.1 of [ BS20]. The same argument works for\nallµ0>0. /square\nA direct consequence is the following commutator estimates for Lipschitz functions:\nCorollary B.2. Assume that κ∈W1,∞(Rd)andb∈S0(T∗Rd), then there exists C >0such that\n/ba∇dbl[κ,Oph(b)]/ba∇dblL(L2(Rd))+/ba∇dbl[κ,Opw\nh(b)]/ba∇dblL(L2(Rd))≤Ch.\nMoreover generally, for some m≥1, we have\n/ba∇dbladm\nκ(Oph(b))/ba∇dblL(L2(Rd))+/ba∇dbladm\nκ(Opw\nh(b))/ba∇dblL(L2(Rd))≤Chm.\nProof.We do the proof for the standard quantization. The same proof applied to the Weyl quantiza-\ntion. Denote by Tj= adj\nκ(Oph(b)) andKj(x,y) the Schwartz kernel of Tj. Then\nKj(x,y) = (κ(x)−κ(y))jK0(x,y),\nwhere\nK0(x,y) =1\n(2πh)d/integraldisplay\nRdb(x,ξ)ei(x−y)·ξ\nhdξSHARP DECAY RATE FOR THE DAMPED WAVE EQUATION WITH CONVEX-SH APED DAMPING 45\nis the Schwartz kernel of Oph(b). Sinceκ∈W1,∞(Rd), there exists Ψ ∈L∞(R2d;Rd), such that\nκ(x)−κ(y) = (x−y)·Ψ(x,y) =d/summationdisplay\nl=1(xl−yl)Ψ(l)(x,y),\nwhere Ψ(l)is thel-th component of Ψ. Thus by integration by part,\nKm(x,y) =/summationdisplay\nj1,···,jmhm\n(2πh)d/integraldisplay\nRdei(x−y)·ξ\nh(Dξ1···Dξmb)(x,ξ)m/productdisplay\nk=1Ψ(jk)(x,y)dξ.\nApplying Lemma B.1, we deduce that /ba∇dblTm/ba∇dblL(L2(Rd))≤Chm. /square\nWe also need a special version of symbolic calculus, used in t he proof of Lemma 3.2:\nLemma B.3. Letκ∈C1\nc(R2),ϕ=ϕ(ξ)∈C∞\nc(R)andb2∈C1\nc(R2),b2∈C2\nc(R2). Denote by\nc(x,y,ξ) =κ(x,y)ϕ(ξ). Then\n(a)/ba∇dblOph(c)b1−Oph(cb1)/ba∇dblL(L2(R2))≤C1h;\n(b)/vextenddouble/vextenddoubleOph(c)b2−Oph(cb2)−h\niOph(∂ξc·∂xb2)/vextenddouble/vextenddouble\nL(L2(R2))≤C2h2,\nwhere the constants C1depend only on /ba∇dblκ/ba∇dblW1,∞,/ba∇dblb1/ba∇dblW1,∞andC2depend only on /ba∇dblκ/ba∇dblW1,∞and\n/ba∇dblb2/ba∇dblW2,∞.\nProof.Sincecdoes not depend on η, by viewing yas a parameter as in the proof of Lemma B.1,\nit suffices to view Oph(c),bas operators acting on L2(Rx) and prove the one-dimensional estimate.\nHence we will not display the dependence in yin the analysis below.\nForj= 1,2, note that the symbol of operators Oph(c)bjare given by\ncj(x,ξ) =1\n2πh/integraldisplay/integraldisplay\nR2e−ix1·ξ1\nhc(x,ξ+ξ1)bj(x+x1)dx1dξ1=1\n2π/integraldisplay\nReixξ′c(x,ξ+hξ′)/hatwidebj(ξ′)dξ′.\nNote that the Fourier transform makes sense since the functi onbjhas compact support. By the Taylor\nexpansions up to order 1 and 2:\nc(x,ξ+hξ′) =c(x,ξ)+hξ′/integraldisplay1\n0∂ξc(x,ξ+thξ′)dt,\nc(x,ξ+hξ′) =c(x,ξ)+hξ′∂ξc(x,ξ)+h2ξ′2/integraldisplay1\n0(1−t)∂2\nξc(x,ξ+thξ′)dt,\nwe have\nc1(x,ξ) =c(x,ξ)b1(x)+h/integraldisplay1\n0Φ1,t(x,ξ)dt\nand\nc2(x,ξ) =c(x,ξ)b2(x)+h\ni∂ξc(x,ξ)∂xb2(x)+h2/integraldisplay1\n0(1−t)Φ2,t(x,ξ)dt,\nwhere\nΦ1,t(x,ξ) =1\n2πi/integraldisplay\nReixξ′∂ξc(x,ξ+thξ′)/hatwidest∂xb1(ξ′)dξ′46 CHENMIN SUN\nand\nΦ2,t(x,ξ) =−1\n2π/integraldisplay\nReixξ′∂2\nξc(x,ξ+thξ′)/hatwidest∂2xb2(ξ′)dξ′.\nTherefore, the symbol of the operator Oph(c)b1−Oph(cb1) ish/integraltext1\n0Φ1,t(x,ξ)dtand the symbol of the\noperator Oph(c)b2−Oph(cb2)−h\niOph(∂ξc∂xb2) ish2/integraltext1\n0(1−t)Φt(x,ξ)dt. It suffices to show that\noperatorsTj,twith Schwartz kernels\nKj,t(x,x′) :=1\n2πh/integraldisplay\nRei(x−x′)ξ\nhΦj,t(x,ξ)dξ, j= 1,2\nare uniformly bounded on L2(R) with respect to t∈(0,1) andh∈(0,1]. Sincec(x,ξ) =κ(x)ϕ(ξ),\nexplicit computation yields\nKj,t(x,x′) =1\nijh/hatwidestϕ(j)/parenleftbigx′−x\nh/parenrightbig\nκ(x)·(∂j\nxbj)((1−t)x+tx′), j= 1,2.\nSince|Kj,t(x,x′)| ≤C1\nh/vextendsingle/vextendsingle/hatwiderϕ′′/parenleftbigx′−x\nh/parenrightbig/vextendsingle/vextendsingle. Finally, by Young’s convolution inequality, we obtain tha t\n/ba∇dblTj,tf/ba∇dblL2(R)≤C/ba∇dblf/ba∇dblL2(R)\nforj= 1,2. This completes the proof of Lemma B.3. /square\nThe above estimates can be generalized on to symbols and func tions on compact manifolds. In\nthe special situation where the manifold is Td, we still have explicit formulas. Indeed, following\n[Zw12] (Chapter 5), for a symbol c(z,ζ) onT∗Td, by periodicity c(z+ 2πk,ξ) =c(z,ζ),k∈Zd, the\nquantization is explicitly given by\nOph(c)f(z) =/summationdisplay\nk∈ZdCkf(z), Ckf(z) :=1\n(2πh)d/integraldisplay\nRd/integraldisplay\nTdc(z,ζ)ei(z−z′+2πk)·ζ\nhf(z′)dz′dζ.\nThenCk=1Tdτ−2πkOph(a)1Tdwhereτz0f(z) :=f(z−z0). By the stationary phase analysis, for\nsymbolsc∈S0(T∗Td) and for |k|>2,\n/ba∇dblCk/ba∇dblL(L2(Td))=O(hN/a\\}b∇acketle{tk/a\\}b∇acket∇i}ht−N),∀N∈N.\nThese facts imply that Lemma B.1, Lemma B.2and Lemma B.3still hold by replacing Rd,R2,Rto\nTd,T2,T, respectively7.\nFinally, we recall the definition of the semiclassical wavef ront set, following Chapter 8 of [ Zw12].\nThesemiclassical wavefront set WFh(u) associated with a h-tempered family u= (uh)0900 0.76 \nLP773K 1.2±0.1 (1.081 ±0.002) 0.46 2.05 (2.0) 1.78 890 1.46 \nHP300K 0.9±0.1 (1.066 ±0.002) 1.32 2.01 (2.0) 3.12 -- 3.22 \n 16 \n Figure 1 \nFig. 1. (a) Layout of the in -house made VNA-based out -of-plane ferromagnetic resonance \nsetup . (b) Out -plane ferromagnetic resonance spectra recorded for the well -ordered LP673K \nsample at different temperatures 𝑓𝑓=10 GHz . \n \n \n17 \n Figure 2 \nFig. 2. (a) Magnetization vs. temperature plots measured on the CFA films with different \ndegree of atomic order. (b) Theoretically calculated magnetization vs. temperature curves for \nCFA phases with different degree of atomic order, where 50 % (100 %) Fe atoms on Heusler \nalloy 4a sites indicate B2 (L2 1) ordered phase, and the rest are intermediate B2 & L2 1 mixed \nordered phases. \n \n \n \n18 \n Figure 3 \nFig. 3. Resonance field vs. in -plane orientation of the applied magnetic field of (a) 𝑇𝑇𝑆𝑆=\n300℃ , 𝑃𝑃𝐼𝐼𝑖𝑖𝑖𝑖=75 𝑊𝑊 deposited, (b) 𝑇𝑇𝑆𝑆=400℃ , 𝑃𝑃𝐼𝐼𝑖𝑖𝑖𝑖=75 𝑊𝑊 deposited, (c) 𝑇𝑇 𝑆𝑆=500℃ , \n𝑃𝑃𝐼𝐼𝑖𝑖𝑖𝑖=75 𝑊𝑊 deposited, and (d) 𝑇𝑇 𝑆𝑆=27℃ , 𝑃𝑃𝐼𝐼𝑖𝑖𝑖𝑖=100 𝑊𝑊 deposited films. Red lines \ncorrespond to fits to the data using Eq. (1). \n \n \n \n19 \n \nFigure 4 \nFig. 4. Line-width vs. frequency of (a) 𝑇𝑇𝑆𝑆=300℃ , 𝑃𝑃𝐼𝐼𝑖𝑖𝑖𝑖=75 𝑊𝑊 deposited, (b) 𝑇𝑇𝑆𝑆=400℃ , \n𝑃𝑃𝐼𝐼𝑖𝑖𝑖𝑖=75 𝑊𝑊 deposited, (c) 𝑇𝑇 𝑆𝑆=500℃ , 𝑃𝑃𝐼𝐼𝑖𝑖𝑖𝑖=75 𝑊𝑊 deposited, and (d) 𝑇𝑇𝑆𝑆=27℃ , \n𝑃𝑃𝐼𝐼𝑖𝑖𝑖𝑖=100 𝑊𝑊 deposited samples. Red lines correspond to fits to the data to extract the \nexperimental Gilbert damping constant and inhomogeneous line -width. Respective insets \nshow the experimentally determined temperature dependent Gilbert damping constants. \n \n \n \n \n20 \n Figure 5 \nFig. 5. Frequency vs. applied field of (a) 𝑇𝑇𝑆𝑆=300℃ , 𝑃𝑃𝐼𝐼𝑖𝑖𝑖𝑖=75 𝑊𝑊 deposited, (b) 𝑇𝑇𝑆𝑆=\n400℃ , 𝑃𝑃𝐼𝐼𝑖𝑖𝑖𝑖=75 𝑊𝑊 deposited, (c) 𝑇𝑇𝑆𝑆=500℃ , 𝑃𝑃𝐼𝐼𝑖𝑖𝑖𝑖=75 𝑊𝑊 deposited, and (d) 𝑇𝑇𝑆𝑆=27℃ , \n𝑃𝑃𝐼𝐼𝑖𝑖𝑖𝑖=100 𝑊𝑊 deposited samples. Red lines correspond to Kittel’s fits to the data. Respective \ninsets show the temperature dependent effective magnetization a nd inhomogeneous line -width \nbroadening values. \n \n \n \n21 \n Figure 6 \nFig. 6. (a) Linewidth vs. frequency with and without a glass spacer between the waveguide \nand the sample. Red lines correspond to fits using Eq. (4). (b) Temperature dependent values \nof the radiative damping using Eq. (6). The lines are guide to the eye. (c) 𝛼𝛼−𝛼𝛼𝑟𝑟𝑚𝑚𝑟𝑟≈𝛼𝛼𝑒𝑒𝑟𝑟𝑟𝑟𝑒𝑒 \nvs 𝛿𝛿2. The red line corresponds to a fit using Eq. (7) to extract the value of the correction \nfactor 𝐶𝐶. (d) Temperature dependent values of eddy current dampi ng using Eq. (7). The lines \nare guide to the eye. \n \n \n \n \n22 \n Figure 7 \nFig. 7. Experimental (a) and theoretical (b) results for the temperature dependent intrinsic \nGilbert damping constant for CFA samples with different degree of atomic order . The B2 & \nL21 mixed phase corresponds to the 75 % occupancy of Fe atoms on the Heusler alloy 4a \nsites. The lines are guide to the eye. \n \n \n \n23 \n Figure 8 \nFig. 8. Total and atom -resolved spin polarized density of states plots for various \ncompositional CFA phases; (a) A2, (b) B2 and (c) L2 1. \n \n \n \n" }, { "title": "2009.11947v1.Squeezing_the_Parameter_Space_for_Lorentz_Violation_in_the_Neutrino_Sector_by_Additional_Decay_Channels.pdf", "content": "arXiv:2009.11947v1 [hep-ph] 22 Sep 2020Squeezing the Parameter Space for Lorentz Violation in\nthe Neutrino Sector by Additional Decay Channels\nUlrich D. Jentschura\nDepartment of Physics, Missouri University\nof Science and Technology, Rolla, Missouri 65409, USA; ulj@mst.edu\nMTA–DE Particle Physics Research Group,\nP.O. Box 51, H–4001 Debrecen, Hungary\nMTA Atomki, P.O. Box 51, H–4001 Debrecen, Hungary\nSeptember 28, 2020\nAbstract\nThe hypothesis of Lorentz violation in the neutrino sector has intrig ued\nscientists for the last two to three decades. A number of theoret ical argu-\nments support the emergence of such violations first and foremos t for neutrinos,\nwhich constitute the “most elusive” and “least interacting” particle s known to\nmankind. It is of obvious interest to place stringent bounds on the L orentz-\nviolating parameters in the neutrino sector. In the past, the most stringent\nbounds have been placed by calculating the probability of neutrino de cay into\na lepton pair, a process made kinematically feasible by Lorentz violatio n in\nthe neutrino sector, above a certain threshold. However, even m ore stringent\nbounds can be placed on the Lorentz-violating parameters if one ta kes into\naccount, additionally, the possibility of neutrino splitting, i.e., of neut rino de-\ncay into a neutrino of lower energy, accompanied by “neutrino-pair ˘Cerenkov\nradiation”. This process has negligible threshold and can be used to im prove\nthe bounds on Lorentz-violating parameters in the neutrino secto r. Finally, we\ntake the opportunity to discuss the relation of Lorentz and gauge symmetry\nbreaking, with a special emphasis on the theoretical models employe d in our\ncalculations.\nKeywords: Lorentz Violation, Neutrinos, Gauge Invariance , Mass Mixing, IceCube\nDetector; Physics beyond the Standard Models\n1 Introduction\nNeutrinos are the most elusive particles within the Standar d Model of Elemen-\ntary Interactions. Speculation about their tachyonic natu re started with Ref. [1],\nand has led to the development of a few interesting scenarios [2]. Within\nthe Lorentz-violating scenarios [3–13], many different tens or structures involving\nLorentz-violating parameters can be pursued. Here, we assu me that, in a preferred\n1particular (observer) Lorentz frame, an isotropic dispers ion relation of the form\nE=/radicalbig\n/vector p2v2+m2withv >1 holds. (In this article, we use physical units with\n/planckover2pi1=ǫ0=c= 1.) Formalizing the Lorentz-violating ideas, the Lorentz –Violating\nExtension of the Standard Model (SME) was developed with a st rong inspiration\nfrom string theory [14,15].\nKinematically, decay among neutrino mass eigenstates is al lowed due to their mass\ndifferences, while decay rates for “ordinary” neutrinos with in the Standard Model\nformalism (for both Dirac as well as Majorana) exceed the lif etime of Universe by\norders of magnitude. Lorentz-violating neutrinos undergo stronger decay and energy\nloss mechanisms than “ordinary” neutrinos because of their dispersion relation E=/radicalbig\n/vector p2v2+m2≈ |/vector p|v(at high energy), which makes a number of decay channels\n(without GIM suppression, see Refs. [16,17]) kinematicall y possible.\nSome remarks on the origin of modified dispersion relations o f the form E=/radicalbig\n/vector p2v2+m2≈ |/vector p|vmight be in order. Modified dispersion relations could in pri n-\nciple be induced by modified theories of gravity, and modifica tions of the Einstein\nequivalence principle. In Ref. [18], it has been suggested t hat the invariant arena\nfor nonquantum physics is a phase space rather than spacetim e, and the locality\nof an even in space-time is replaced by relative locality in which different observers\nsee different spacetimes, and the spacetimes they observe are energy and momen-\ntum dependent. This hypothesis can lead to modified dispersi on relation of the\nkind investigated here. Effects due to quantum gravity may als o induce modified\ndispersion relations [19,20]. In a different context, Lorent z breaking induced at\nthe Planck scale may also induce such relations [21–23], in t he sense of “doubly\nspecial relativity” which works on the assumption that dyna mics are governed by\ntwo observer-independent quantities, the speed of light an d an additional constant\nenergy scale, which could be the Planck energy scale (see als o Ref. [24]). The phe-\nnomenological consequences of tiny Lorentz violations, wh ich are rotationally and\ntranslationally invariant in a preferred frame, and are ren ormalizable while preserv-\ning anomaly cancellation and gauge invariance under the Sta ndard Model gauge\ngroupSU(3)⊗SU(2)⊗U(1), have been analyzed in Ref. [25].\nThere are a number of phenomena which could direct us to have a look at the neu-\ntrino sector for Lorentz violation. Namely, for example, th e early arrival of neutrinos\nfrom the1987 supernovastill inspires(some) physicists. S pecifically, undertheMont\nBlanc, in the early morning hours of February 23, 1987, a show er of neutrinos of in-\nterstellar origin arrived about 6 hours earlier then the vis ible light from the Siderius\nNuntius SN1987A supernova. This event has been recorded in R ef. [26], and it was\nasserted that such an event could happen by accident once in a bout 1000 years.\nDirect measurements of neutrino velocities have given resu lts that are consistent\nwith the speed of light within experimental uncertainty, bu t with the experimental\nresult being a littler larger than the speed of light. For exa mple, the MINOS exper-\niment [27] has measured superluminal neutrino propagation velocities which differ\nfrom the speed of light by a relative factor of (5 .1±2.9)×10−5at an energy of about\nEν≈3GeV, compatible with an earlier FERMILAB experiment [28]. Furthermore,\nneutrinos cannot be used to transmit information (at least n ot easily) because of\ntheir small interaction cross sections. Superluminality o f neutrinos would thus not\n2necessarily lead to violation of causality at a macroscopic level, as demonstrated in\nAppendix A.2 of Ref. [29]. Similar arguments have been made i n Ref. [30], where\nit was shown that problems with microcausality, in Lorentz- violating theories, are\nalleviated for small Lorentz-violating parameters and in s o-calledconcordant frames\nwhere the boost velocities are not too large. For neutrinos, corresponding prob-\nlems are further alleviated by the fact that their interacti on cross sections are small;\nhence, it becomes very hard to transport information superl uminally even if the\ndispersion relation indicates such effects (see also Appendi x A.2 of Ref. [29]).\nWe should also note that, when Lorentz invariance is violate d, superluminality does\nnot necessarily lead to problems. Under certain additional assumptions, causality\narguments related to superluminal signal propagation simp ly do not apply in this\ncase. For example, if photons themselves propagate faster t hanc(the limiting\nvelocity of massive standard fermions) in some Lorentz-vio lating theories, then, as\nlongasinformationpropagates alongorinsidethemodifiedl ightcone(as definedbya\nmodified effective metric), it can propagate faster than cwithout implying causality\nissues (see, e.g., the paragraph around Eq. (9) in Ref. [31]) . Further information\non this point can be found in Refs. [32–36]. E.g., Ref. [32] pr ovides information\non (micro)causality problems for the purely timelike case o f Maxwell-Chern-Simons\ntheory (operator of dimension 3 for photons).\nSome more explanatory remarks on gauge invariance and Loren tz violation are prob-\nably in order. One might argue that gauge invariance should b e seen as the guiding\nprinciple to make consistent a nonabelian gauge interactio n mediated by spin-one\nmassive particles through the (experimentally confirmed) H iggs mechanism, and\nthat gauge invariance should be retained at all cost, even if Lorentz symmetry is\nbroken. However, this demand overlooks two aspects. The firs t is that the origin of\nthe Lorentz-violating terms could be assumed to be commensu rate with the Planck\nscale [14,21,24], in which case it is questionable if our usu al concepts of gauge in-\nvariance could be transported without changes to the extrem e energy scales; in any\ncase, we would assume the Standard Model gauge group, at the e nergy scale relevant\nto Lorentz violation, to be replaced by a unified gauge group, e.g., SO(1,13) (see\nRefs. [37–41]). Further fundamental modifications of the ga uge principle at extreme\nenergyscalesarealsoconceivable. Thesecondandmoreimpo rtantoverlooked aspect\nis that Lorentz violation constitutes a form of gauge invariance violation. Namely,\nEinstein’s theory of general relativity isthe classical, gauged theory of gravitation,\nin which the global Lorentz symmetry is elevated to a gauged, local symmetry [42].\nThis observation offers the construction principle for the sp in connection of Dirac\nfields in curved space-times, which has been explored in rece nt monographs [42] and\npapers [43,44]. To reemphasize the point, we recall that Lor entz invariance violation\nisgauge invariance violation within General Relativity [42] .\nWe organize the paper as follows. After a consideration of th e calculation of the\nthreshold for the superluminal decay processes (Sec. 2), we outline the calculation\nof the decay and energy loss rates in Sec. 3, before discussin g an attractive scenario\nfor Lorentz violation in theneutrinosector, andits signat ures, in Sec. 4. Conclusions\nare reserved for Sec. 5.\n3e\nν(m)\niν(m)\ni\nZ0e−p2 p4\np1p3\n(a)ν(m)\nf\nν(m)\niν(m)\ni\nZ0ν(m)\nf−p2 p4\np1p3\n(b)\nFigure 1: In the lepton-pair ˘Cerenkov radiation process (a), an on-\ncoming Lorentz-violating initial neutrino mass eigenstat eν(m)\nide-\ncays, underemission ofavirtual Z0boson, intoanelectron-positron\npair. The sum of the outgoing pair momenta is p2+p4; one ob-\nserves the inverted direction of the fermionic antiparticl e line. The\narrow of time is from bottom to top. The (blue) bosonic line ca r-\nries the four-momentum q. Diagram (b) describes the neutrino-pair\n˘Cerenkov radiation process, with a final neutrino mass eigen state\nν(m)\nf.\n2 Threshold Considerations\nWe refer to the lepton-pair ˘Cerenkov radiation process (LPCR) in Fig. 1(a) and\nthe neutrino-pair ˘Cerenkov radiation process (NPCR) depicted in Fig. 1(b). In\norder to make neutrino decay kinematically possible, it is n ecessary to fulfill certain\nthreshold conditions. Let us denote the outgoing fermions i n the generic decay\nprocesses depicted in Fig. 1 by\nν→ν+f+¯f, (1)\nwith apair of amassive fermion fandits antiparticle ¯fbeingemitted in theprocess.\nEnergy-momentum conservation implies that (in the notatio n of Fig. 1)\n(p1−p3)2=q2= (p2+p4)2. (2)\nLet us first consider the case of a massive outgoing pair 2 + 4, w ith rest mass\nmf, and vanishing Lorentz-violating parameter. Threshold is reached for collinear\nemission geometry. The incoming four-momentum is p1= (E1,/vector p1), whereE1=/radicalBig\n/vector p2\n1v2\ni+m2ν≈ |/vector p1|vi, andmνis theneutrinomass. Assumingan incoming neutrino\nenergy well above its rest mass, we can do this approximation . By contrast, one has\np3= (0,/vector0), so that the total four-momentum transfer qgoes into the pair.\nA few remarkson thedispersionrelations usedinthecurrent paper, areinorder. We\nobservethat therelation E1=/radicalBig\n/vector p2\n1v2\ni+m2νis isotropic, andit inducessuperluminal\n4velocities. However, it should be noted that the modified dis persion relation applied\nto neutrino sector can introduce a rich phenomenology even i n other sectors, as\noscillations, without requiring superluminal velocities forvi>1. Many models can\nbe explored (see Refs. [5–7,9,45–48]). As explained above, we here assume a simple\nform for the modified dispersion relation, applicable at lea st in a preferred observer\nLorentz frame.\nReturning to the discussion of the threshold, we observe tha t, for collinear geometry,\none haspµ\n2=pµ\n4= (Ef,/vector pf), whereEf=/radicalBig\n/vector p2\nf+m2. Under these assumptions,\np2+p4= (2/radicalBig\n/vector p2\nf+m2,2/vector pf), so that ( p2+p4)2= 4m2\nf. The threshold condition\nbecomes\np2\n1(v2\ni−1)≥4m2\nf, p 1≈E1≥2mf/radicalBig\nv2\ni−1=2mf√δi. (3)\nHere,\nvi=/radicalbig\n1+δi. (4)\nThe threshold condition Eth= 2mf/√δihas been used extensively in Refs. [49–51].\nThe formula (3) implies that the threshold for NPCR is lower b y at a least six orders\nof magnitude as compared to LPCR.\nThe kinematic considerations are very different in the high-e nergy regime, when\nboththe incoming (decaying) particle as well as the outgoing par ticles are Lorentz\nviolating. Masses can be neglected. In this case, one has at t hresholdp2=p4=\n(Ef,/vector pf), whereEf=|/vector pf|vf, so that at threshold\np2\n1(v2\ni−1)≥4p2\nf(v2\nf−1), (5)\nDue to equipartition of the energy among each particle of the outgoing pair at\nthreshold, one has pf≈Ef=E1/2≈p1/2. In this case, the threshold condition\nreduces to\n(v2\ni−1)≥(v2\nf−1), δi>δf. (6)\nHere,vf=/radicalbig\n1+δf. Forδi=δf, no phase space is available in order to accom-\nmodate for the decay. This consideration explains why all re sults communicated in\nRef. [51] display a factor δi−δf; decay takes place from “faster” to “slower” mass\neigenstates.\n3 Outline of the Calculation\nTheunderstandingof decay processes involving Lorentz-vi olation has beenadvanced\nthrough Refs. [49–51]. Let us briefly recall elements of the d erivation given in\nRef. [51]. One particular question is how to express the deca y rate for an (initially)\nflavor-eigenstate neutrino (the electroweak Lagrangian is flavor-diagonal) in terms\nof mass eigenstates. We have, in the same obvious notation as used in Ref. [51],\nν(f)\nk=/summationdisplay\nℓUkℓν(m)\nℓ, (7)\n5with the Pontecorvo–Maki–Nakagawa–Sakata (PMNS) matrix Ukℓ. The interaction\ninteraction LWwith theZ0boson in the flavor basis is\nLW=−gw\n4 cosθW/summationdisplay\nkν(f)\nkγµ(1−γ5)ν(f)\nkZµ. (8)\nHere,gwis the weak coupling constant, and θWis the Weinberg angle. A unitary\ntransformation leads to\nL=−gw\n4 cosθW/summationdisplay\nk,ℓ,ℓ′U+\nℓkUkℓ′ν(m)\nℓγµ(1−γ5)ν(m)\nℓ′Zµ. (9)\nThe interaction with the Z0boson in the mass eigenstate basis therefore reads as\nfollows,\nL=−gw\n4 cosθW/summationdisplay\nℓν(m)\nℓγµ(1−γ5)ν(m)\nℓZµ. (10)\nIn order to model the free Lorentz-violating neutrino Lagra ngian, one introduces an\neffective metric with a tilde:\nL=/summationdisplay\nℓiν(m)\nℓγµ(1−γ5)˜gµν(vℓ)∂νν(m)\nℓ. (11)\nThe modified metric defines a modified light cone according to ˜ gµν(vℓ)kµkν= 0.\nHere,\n˜gµν(vℓ) = diag(1,−vℓ,−vℓ,−vℓ). (12)\nThe dispersion relation\nEℓ=|/vector p|vℓ (13)\nfollows as the massless limit of Eℓ=/radicalBig\n(|/vector p|vℓ)2+m2\nℓ. For neutrinos, we know that\nthemℓterms are different. So, there is reason to assume that the δℓ=/radicalBig\nv2\nℓ−1\nterms are also different among mass (flavor) eigenstates, if th ey are nonvanishing.\nOne defines parameters vintandδintby the relation\nvint=/radicalbig\n1+δint (14)\nfortheunifieddescriptionofLPCRandNPCR;theeffectivefour -fermionLagrangian\nfor the process reads as\nLint=feGF\n2√\n2ν(m)\niγλ(1−γ5)ν(m)\ni˜gλσ(vint)¯ψfγσ(cV−cAγ5)ψf.(15)\nCohen and Glashow [49] set vint= 1. (In Ref. [51], on a number of occasions, the\nparameter usedin Ref. [49] had beeninadvertently indicate d asvint= 0, which is not\nthecase. We take theopportunity to point out that of course, theparameter vint= 1\nimplies that δint= 0, which was the intended statement in Ref. [49].) Bezrukov and\nLee [50] use the parameters vint= 1 (“model I”) and vint=vi(“model II”). In\nRef. [51], the parameter vintis kept as a variable. As explained in detail in Ref. [52],\n“gauge invariance” (with respect to a restricted subgroup o f the electroweak sector)\n6can be restored if one uses the value vint=vivf. Both Cohen and Glashow [49],\nas well as Bezrukov and Lee [50], assume that δf= 0 for LPCR. The parameter fe\ncharacterizes the process:\nfe=/braceleftBigg\n1, ψ f=ν(m)\nf\n2, ψ f=e. (16)\nApproximately, one has\n(cV,cA) =/braceleftBigg\n(1,1) ψf=ν(m)\nf\n(0,−1\n2), ψ f=e. (17)\nThe characteristic matrix element is\nM=feGF\n2√\n2/bracketleftBig\n¯ui(p3)γλ(1−γ5)ui(p1)/bracketrightBig\n˜gλσ(vint)/bracketleftbig\n¯uf(p4)(cVγσ−cAγσγ5)vf(p2)/bracketrightbig\n.\n(18)\nKey to the calculation is the fact that one can split the three -particle outgoing phase\nspace\nΓ =1\n2E1/integraldisplay\ndφ3(p2,p3,p4;p1)1\nns/summationdisplay\nspins|M|2\n=1\n2E1/integraldisplayM2\nmax\nM2\nmindM2\n2πdφ2(p3,p24;p1)dφ2(p2,p4;p24)1\nns/summationdisplay\nspins|M|2.(19)\nwith appropriate limits for M2\nminandM2\nmaxbeing given as follows,\nM2\nmin=δf(|/vector p2|+|/vector p4|)2, M2\nmax=δi(|/vector p1|−|/vector p3|)2. (20)\nThe following splitting relation for the phase space is cruc ial to a simplification of\nthe integrations [for details, see Ref. [53] and Eq. (43) of R ef. [51]],\ndφ3(p2,p3,p4;p1)\n=/integraldisplaydM2\n2πd4p3\n(2π)3δ+(p2\n3−δik2\n3)d4p24\n(2π)3δ+(p2\n24−M2)(2π)4δ(4)(p1−p3−p24)\n/bracehtipupleft /bracehtipdownright/bracehtipdownleft /bracehtipupright\n=dφ2(p3,p24;p1)\n×d4p2\n(2π)3δ+(p2\n2−δfk2\n2)d4p4\n(2π)3δ+(p2\n4−δfk2\n4)(2π)4δ(4)(p24−p2−p4)\n/bracehtipupleft /bracehtipdownright/bracehtipdownleft /bracehtipupright\n=dφ2(p2,p4;p24)\n=/integraldisplaydM2\n2πdφ2(p3,p24;p1)dφ2(p2,p4;p24).(21)\nHere,p24=p2+p4, andwedenotethespatialpartofthefour-vector pµ\nias/vectorkiwithi=\n1,2,3,4, so thatpµ\ni= (Ei,/vectorki), and since Ei= (1+δi)|/vectorki|, one hasgµν˜gµα˜gνβpα\nipβ\ni=\ngµνpµ\nipν\ni−δik2\ni=p2\ni−δik2\ni[see also Eq. (40) of Ref. [51]]. This relation explains\n7the argument of some of the Dirac- δfunctions in Eq. (21). The general result for\nthe decay rate, unifying both processes depicted in Fig. 1, r eads as follows,\nΓνi→νiψf¯ψf=G2\nFk5\n1\n192π3f2\nec2\nV+c2\nA\n420ns(δi−δf)/bracketleftbigg\n(60−43σi)(δi−δf)2\n+2(50−32σi−25σf+7σiσf)(δi−δf)δf\n+7(4−3σi−3σf+2σiσf)δ2\nf+7δ2\nint/bracketrightbigg\n.(22)\nThisresultvanishesfor δi=δf(seethediscussioninSec.2). CohenandGlashow[49]\nhavens= 2 active spin states for the (initial) neutrino, while Bezr ukov and Lee [50]\ncalculate with ns= 1, implying that the authors of Ref. [50] assume that only on e\nspin state exists in nature. This assumption affects the avera ging over the initial\nquantum states involved in the process. The σparameters depend on the way in\nwhich spin polarization sums are carried out,\nσi=/braceleftbigg0,CG spin sum for νi\n1,BL spin sum for νi, σf=/braceleftbigg0,CG spin sum for ψf\n1,BL spin sum for ψf.(23)\nInRef.[49], theCohen–Glashow (CG) spinsum(“polarizatio n sum”)issimplytaken\nas the standard spin sum for massless fermions,\n/summationdisplay\nsνℓ,s⊗¯νℓ,s=pµgµνγν. (24)\nInRef. [50], theBezrukov–Lee (BL) spinsumis based on asome whatmoreadvanced\ntreatment of the eigenspinors of superluminal neutrino mas s eigenstates and reads\nas /summationdisplay\nsνℓ,s⊗¯νℓ,s=pµ˜gµν(vℓ)γν. (25)\nThe general result for the energy loss rate, applicable to bo th processes in Fig. 1,\nreads as\ndEνi→νiψf¯ψf\ndx=−G2\nFk6\n1\n192π3f2\nec2\nV+c2\nA\n672ns(δi−δf)\n×/bracketleftbigg\n(75−53σi)(δi−δf)2+(122−77σi−61σf+16σiσf)(δi−δf)δf\n+8(4−3σi−3σf+2σiσf)δ2\nf+8δ2\nint/bracketrightbigg\n.(26)\nInRef.[51], wehaveverifiedandchecked compatibility with all formulascontainedin\nRefs. [49] and [50]. This is important because it confirms tha t the model dependence\nof the results is only contained in the numerical prefactors , but not in the overall\nscaling of the results.\nAs outlined in Ref. [51], one can parameterize the results fo r NPCR as follows,\nΓνi→νiνf¯νf=bG2\nF\n192π3k5\n1,dEνi→νiνf¯νf\ndx=−b′G2\nF\n192π3k6\n1. (27)\n8For the CG spin sum, one obtains the following bparameters,\nbCG=1\n7(δi−δf)/bracketleftbigg\n(δi−δf)2+5\n3δf(δi−δf)+7\n15δ2\nf/bracketrightbigg\n, (28a)\nb′\nCG=25\n224(δi−δf)/bracketleftbigg\n(δi−δf)2+112\n75δf(δi−δf)+32\n75δ2\nf/bracketrightbigg\n.(28b)\nFor the BL spin sum, one obtains\nbBL=17\n210(δi−δf)/bracketleftbigg\n(δi−δf)2+7\n17δ2\nint/bracketrightbigg\n, (29a)\nb′\nBL=11\n168(δi−δf)/bracketleftbigg\n(δi−δf)2+4\n11δ2\nint/bracketrightbigg\n. (29b)\nTypically, one finds [51] numerical prefactors in these form ulas are larger than those\nfor LPCR by a factor of four or five. Also, NPCR has negligible t hreshold.\nIn papers of Stecker and Scully [12,54,55], the following bo und is derived for the\nLorentz-violating parameter of the electron-positron fiel d alone (watch out for a\ndifference in the conventions used for defining the δeparameter):\nδe≤1.04×10−20. (30)\nWe should stress that this bound concerns oncoming electron s (not neutrinos!) and\nhas nothing to do with the processes studied here.\nThe observation of very-high-energy neutrinos by IceCube, taking into consideration\nthe LPCR process (but not NPCR!), implies that the Lorentz-v iolating parameter\nfor neutrinos cannot be larger than (Ref. [12])\nδν≤2.0×10−20. (31)\nThis bound is based on the assumption that δeandδνare different. Colloquially\nspeaking, we can say that, if δνwere larger, then “Big Bird” (the 2PeV specimen\nfound in IceCube, see Refs. [56,57]) would have already deca yed before it arrived at\nthe IceCube detector. However, the full analysis requires M onte Carlo simulations\ninvolving astrophysical data and is much more involved [12, 54,55].\nProvidedtheLorentz-violating parametersforthedifferent neutrinomasseigenstates\naredifferent, low-energy neutrinosareaffectedbythedecayan denergylossprocesses\nconnected with NPCR, in view of a negligible threshold for NP CR. As already\nemphasized, typical numerical coefficients forNPCR areafac tor of fourorfivelarger\nthan for LPCR, depending on the model used for the spin sums. T his enhances the\nimportance of the NPCR effect. Inspired by Eq. (31), we thus con jecture here that\na full analysis of astrophysical data, using the NPCR proces s as a limiting factor for\nthe observation of high-energy neutrinos, should yield a bo und on the order of\n|δi−δf| ≤1\n51/3×2.0×10−20∼1.2×10−21, (32)\nwhere the prefactor takes into account the scaling of the effec t with theδparameter.\nSpecifically, the decay and energy loss rates typically scal e with the factor ( δi−δf)3.\nIt would be very fruitful if this conjecture were to be checke d against astrophysical\ndata in an independent investigation.\n94 An Attractive Scenario\nAt first, one might see a dilemma: Within a fully SU(2)Lgauge-invariant theory,\nwith uniform Lorentz-violating parameters over all partic le generations, one nec-\nessarily has δν=δe(see Ref. [52] for a detailed discussion), and so, the bound\nδν≤2.0×10−20given in Eq. (31) is not applicable, because the LPCR process\ndoes not exist. But then, one has to acknowledge that the boun dδe≤1.04×10−20\ngiven in Eq. (30), which is originally derived for electrons , based on other physical\nprocesses, automatically also applies to the neutrino sect or.\nSo, the dilemma is that either, one has to give up gauge invari ance and use dif-\nferent Lorentz-violating parameters for neutrinos as oppo sed to charged leptons\nwithin the same particle generation, or, if one insists on ga uge invariance, or, as-\nsume different Lorentz-violating parameters for different gen erations. Otherwise,\nthe insistence on gauge invariance would defeat part of the p urpose of looking at\nthe neutrino sector for Lorentz violation. This is because i n the latter case, for\nuniform Lorentz-violating parameters among all three gene rations, because, by as-\nsumption, the Lorentz-violating parameters for neutrinos and charged left-handed\nleptons within the same SU(2)Ldoublet are necessarily the same, the tight bounds\non Lorentz-violating parameters in the charged-fermion se ctor automatically apply\nto the neutrino sector as well1. This observation has important consequences when\nexamining the first-generation SU(2)Ldoublet, consisting of ( νe,eL). Electrons and\npositrons are stable particles, and small violations of Lor entz invariance would lead\nto violations of causality on a macroscopic level (see Appen dix A.2 of Ref. [29])2.\nConversely, if we had to carry over all restrictions on Loren tz-violating electron pa-\nrameters to the electron neutrino sector, then this would nu llify all the motivations\nlisted in Sec. 1 for investigating the first-generation neut rino sector.\nOnthecontrary, If oneaccepts thenecessity that different Lo rentz-violating parame-\nters should be used for each of the three known particle gener ations, then one needs\nto acknowledge that the parameter space for differential Lore ntz-violation among\nneutrino mass eigenstates is restricted by additional cons traints due to the NPCR\nprocess [51]. An attractive gauge-invariant scenario coul d still be found, as follows.\nNamely, one might observe that, as per the discussion in Appe ndix A.2 of Ref. [29],\ncausality violations due to Lorentz violation are less seve re for unstable particles,\nwhich decay and therefore are not amenable to the reliable tr ansport of informa-\ntion. Part of the above sketched dilemma could thus be avoide d as follows. One\nfirst observes that, as per the above argument, problems with respect to causality\nare less severe in the second-generation SU(2)Ldoublet (νµ,µL) and also in the\nthird-generation SU(2)Ldoublet (ντ,τL), which are composed entirely of unstable\nparticles. Full gauge invariance can be retained if we assum e generation-dependent\nLorentz-violating parameters δe,δµ, andδτ, for the three SU(2)Ldoublets, which\n1We here ignore the somewhat remote possibility of different L orentz violating parameters for\nthe right-handed and left-handed sectors of one and the same generation.\n2Note, also, that this statement does not hold if the limiting velocity for fermions turns out\nto be smaller as opposed to larger than the speed of light. Fur thermore, there are conceivable\nmodifications of Maxwell theory, as already discussed in Sec . 1, where causality violations are\navoided due to modified light cones [31–36].\n10could be encoded in modified Dirac matrices /tildewideγi=vfγiwithf=e,µ,τ[see Eq. (5)\nof Ref. [52]]. In the charged-fermion sector, we have nearly no mixing of mass and\ncharge eigenstates. Let us then go into the high-energy regi me where, where mass\nand flavor eigenstates, under the assumptions\nδµ,δτ>0, δµ/negationslash=δτ, δe= 0, (33)\nbecome equal. In this case, at high energy, one would have two neutrino mass eigen-\nstates, which asymptotically approach the muon neutrino an d tau neutrino flavor\neigenstates at very high energy, decay into electron-posit ron pairs and (asymptoti-\ncally) electron neutrinos, via LPCR and NPCR.\nOf course, other scenarios and flavor and mass mixing phenome nologies are also\npossible, as discussed in Sec. IV B of Ref. [51]. In general, o ne could interpret\nthe emergence of a specific predominant flavor composition of incoming super-high-\nenergy cosmic neutrinos, consistent with one, and only one, specific mass eigenstate,\nas a signature of Lorentz violation. This is because a single , defined, oncoming mass\neigenstate would be consistent with the two other mass eigen states being “faster”\nand thus decaying into the single “slow” eigenstate.\nIn all discussed scenarios, one might find a conceivable expl anation for the apparent\ncutoff in the cosmic neutrino spectrum at about 2PeV, at the ex pense of reducing\nthe allowed regime of Lorentz-violating δparameters to the range of about 10−20.\nIn our “attractive scenario”, one retains gauge invariance as outlined in Sec. 4 of\nRef. [52] and still is able to account for a super-high-energ y cutoff of the cosmic\nneutrino spectrum. Experimental confirmation or dismissal of this hypothesis will\nrequire better cosmic neutrino statistics at very high ener gies.\n5 Conclusions\nThe existence of the NPCR process [see Fig. 1(b)] reveals a ce rtain dilemma for\nLorentz-violating neutrinos (provided the Lorentz violat ing parameters indicate su-\nperluminality). Namely, under the hypothesis of a nonvanis hing Lorentz-violating\nparameterδ, given as in Eq. (4), the virtuality\nE2−p2=p2(v2−1)≈E2(v2−1) =E2δ (34)\nof a neutrino becomes large for large energy, rendering a num ber of decay pro-\ncesses kinematically possible. Conversely, based on high- energy astrophysical obser-\nvations, very strict bounds can be imposed on the Lorentz-vi olating parameters [see\nEqs. (30), (31), and (32)].\nDeep connections exist between Lorentz violation and gauge invariance. In Ref. [58],\nit is shown that spontaneous Lorentz violation can lead to an effective low-energy\nfield theory with both Lorentz-breaking as well as gauge-inv ariance breaking terms.\nAccording to Refs. [58–69], even the photon could potential ly be formulated as the\nNambu-Goldstone boson linked to spontaneous Lorentz invar iance violation. (This\nansatzwas originally formulated before electroweak unification. ) For a broader view\n11of this point, we refer to Appendix A of Ref. [52]. If one insis ts on the persistence\nof gauge invariance within the electroweak sector, then one has to acknowledge that\nbounds on Lorentz-violating parameters for charged lepton s [e.g., Eq. (30)] also ap-\nply to the neutrino sector [thus lowering the bound otherwis e given in Eq. (31) by a\nfactortwo, andfurtherrestrictingtheavailableparamete rspaceforLorentz-violating\nparameters in the neutrino sector]. Also, the assumption th atδν=δewould defeat\nthe purpose of looking at neutrinos for Lorentz violation. I f one insists on gauge in-\nvariance and still pursues the exploration of Lorentz viola tion in the neutrino sector,\nthen more sophisticated considerations are required (see S ec. 4). Namely, one could\npotentially invoke flavor-dependent differential Lorentz vi olation across generations\n(i.e., with different Lorentz-violating parameters for each generation). In this case,\nflavor and mass eigenstates would become identical in the hig h-energy limit, and\ndecay and energy loss processes could potentially contribu te to an explanation for\nthe apparent cutoff in the cosmic neutrino spectrum in the ran ge of a few PeV (see\nRefs. [56,57] and the discussion in Sec. 4).\nAcknowledgments\nThe authors acknowledges support from the National Science Foundation (Grant\nPHY–1710856) as well as insightful conversations with G. So mogyi and I. N´ andori.\nReferences\n[1] A. Chodos, A. I. Hauser, and V. A. Kosteleck´ y, The Neutrino as a Tachyon ,\nPhys. Lett. B 150, 431–435 (1985).\n[2] U. D. Jentschura, I. N´ andori, and R. Ehrlich, Calculation of the Decay\nRate of Tachyonic Neutrinos against Charged–Lepton–Pair an d Neutrino–Pair\nCerenkov Radiation , J. Phys. G 44, 105201 (2017).\n[3] D. Colladay and V. A. Kosteleck´ y, Lorentz-violating extension of the standard\nmodel, Phys. Rev. D 58, 116002 (1998).\n[4] V. A. Kosteleck´ y, Gravity, Lorentz violation, and the standard model , Phys.\nRev. D69, 105009 (2004).\n[5] V. A. Kosteleck´ y and M. Mewes, Electrodynamics with Lorentz-violating oper-\nators of arbitrary dimension , Phys. Rev. D 80, 015020 (2009).\n[6] J. S. Diaz, V. A. Kosteleck´ y, and M. Mewes, Perturbative Lorentz and CPT\nviolation for neutrino and antineutrino oscillations , Phys. Rev. D 80, 076007\n(2009).\n[7] V. A. Kosteleck´ y and M. Mewes, Neutrinos with Lorentz-violating operators of\narbitrary dimension , Phys. Rev. D 85, 096005 (2012).\n12[8] V. A. Kosteleck´ y and M. Mewes, Fermions with Lorentz-violating operators of\narbitrary dimension , Phys. Rev. D 88, 096006 (2013).\n[9] J. S. Diaz, V. A. Kosteleck´ y, and M. Mewes, Testing relativity with high-energy\nastrophysical neutrinos , Phys. Rev. D 89, 043005 (2014).\n[10] J. Diaz, Neutrinos as Probes of Lorentz Invariance , Adv. High Energy Phys.\n2014, 962410 (2014).\n[11] J. D. Tasson, What do we know about Lorentz invariance? , Rep. Prog. Phys.\n77, 062901 (2014).\n[12] F. W. Stecker, S. T. Scully, S. Liberati, and D. Mattingl y,Searching for traces\nof Planck-scale physics with high energy neutrinos , Phys. Rev. D 91, 045009\n(2015).\n[13] S. Liberati, Lorentz symmetry breaking: phenomenology and constraints , J.\nPhys. Conf. Ser. 631, 012011 (2015).\n[14] V. A. Kosteleck´ y and S. Samuel, Spontaneous breaking of Lorentz symmetry in\nstring theory , Phys. Rev. D 39, 683–685 (1989).\n[15] V. A. Kosteleck´ y and R. Potting, CPT and Strings , Nucl. Phys. B 359, 545–570\n(1991).\n[16] P. B. Pal and L. Wolfenstein, Radiative decays of massive neutrinos , Phys. Rev.\nD25, 766 (1982).\n[17] S. L. Glashow, J. Iliopoulos, and L. Maiani, Weak interactions with lepton-\nhadron symmetry , Phys. Rev. D 2, 1285 (1970).\n[18] G. Amelino-Camelia, L. Freidel, J. Kowalski-Glikman, andL. Smolin, The prin-\nciple of relative locality , Phys. Rev. D 84, 084010 (2011).\n[19] C. Pfeifer, L. Barcaroli, L. Brunkhorst, G. Gubitosi, a nd N. Loret, Hamilton\ngeometry: Phase space geometry from modified dispersion rel ations.InProceed-\nings of the 4th Marcel Grossmann Meeting on Recent Developme nts in Theoret-\nical and Experimental General Relativity, Astrophysics, a nd Relativistic Field\nTheories (MG14) , Rome, Italy, 12–18 July 2015, pp. 3929–3934.\n[20] M.D.C.Torri, V.Antonelli, andL.Miramonti, Homogeneously Modified Special\nrelativity (HMSR) , Eur. Phys. J. C 79, 808 (2019).\n[21] G. Amelino-Camelia, Relativity: Special treatment (Doubly special relativity ),\nNature (London) 418, 34–35 (2002).\n[22] J. Magueijo and L. Smolin, Gravity’s rainbow , Class. Quantum Grav. 21, 1725–\n1736 (2004).\n[23] F. Girelli, S. Liberati, and L. Sindoni, Planck-scale modified dispersion relations\nand Finsler geometry , Phys. Rev. D 75, 064015 (2007).\n13[24] J. Magueijo and L. Smolin, Lorentz Invariance with an Invariant Energy Scale ,\nPhys. Rev. Lett. 88, 190403 (2002).\n[25] S. Coleman and S. L. Glashow, High-energy tests of Lorentz invariance , Phys.\nRev. D59, 116008 (1999).\n[26] V.L. Dadykin, G.T.Zatsepin, V. B.Karchagin, P. V.Korc hagin, S.A.Mal’gin,\nO. G. Ryazhskaya, V. G. Ryasnyi, V. P. Talochkin, F. F. Khalch ukov, V. F.\nYakushev, M. Aglietta, G. Badino, G. Bologna, C. Castagnoli , A. Castellina,\nW. Fulgione, P. Galeotti, O. Saavedra, J. Trincero, and S. Ve rnetto,Detection\nof a rare event on 23 February 1987 by the neutrino radiation d etector under\nMont Blanc , JETP Lett. 45, 593–595 (1987), [Pis’ma ZhETF 45, 464 (1987)].\n[27] MINOS Collaboration, Measurement of neutrino velocity with the MINOS de-\ntectors and NuMI neutrino beam , Phys. Rev. D 76, 072005 (2007). [See also\nPhys. Rev. D 92, 052005 (2015)].\n[28] G. R. Kalbfleisch, N. Baggett, E. C. Fowler, and J. Alspec tor,Experimental\nComparison of Neutrino, Antineutrino, and Muon Velocities , Phys. Rev. Lett.\n43, 1361–1364 (1979).\n[29] U. D. Jentschura, D. Horv´ ath, S. Nagy, I. N´ andori, Z. T r´ ocs´ anyi, and B. Ujv´ ari,\nWeighing the Neutrino , Int. J. Mod. Phys. E 23, 1450004 (2014).\n[30] V. A. Kosteleck´ y and R. Lehnert, Stability, causality, and Lorentz and CPT\nviolation, Phys. Rev. D 63, 065008 (2001).\n[31] S. Dubovsky, A. Nicolis, E. Trincherini, , and G. Villad oro,Microcausality in\ncurved space-time , Phys. Rev. D 77, 084016 (2008).\n[32] C. Adam and F. R. Klinkhamer, Causality and CPT violation from an Abelian\nChern-Simons like term , Nucl. Phys. B 607, 247–267 (2001).\n[33] C. M. Reyes, Causality and stability for Lorentz-CPT violating electrody namics\nwith dimension-5 operators , Phys. Rev. D 82, 125036 (2010).\n[34] F. R. Klinkhamer and M. Schreck, Consistency of isotropic modified Maxwell\ntheory: Microcausality and unitarity , Nucl. Phys. B 848, 90–107 (2011).\n[35] M. Schreck, Analysis of the consistency of parity-odd nonbirefringent modified\nMaxwell theory , Phys. Rev. D 86, 06503 (2012).\n[36] M.Schreck, Quantum field theory based on birefringent modified Maxwell t heory,\nPhys. Rev. D 89, 085013 (2014).\n[37] N. S. Mankoc Borstnik, Unification of Spins and Charges , Int. J. Theor. Phys.\n40, 315–337 (2001).\n[38] N.S.MankocBorstnikandH.B.F.Nielsen, How to generate families of spinors ,\npreprint arXiv:hep-th/0303224.\n14[39] N. S. Mankoc Borstnik, Can the spin-charge-family theory explain baryon num-\nber non conservation? , Phys. Rev. D 91, 065004 (2015).\n[40] N. S. Mankoc Borstnik and H. B. F. Nielsen, The spin-charge-family theory of-\nfers understanding of the triangle anomalies cancellation in the standard model ,\nProgress in Physics 65, 1700046 (2016).\n[41] N. Mankoc Borstnik, in Conference on New Physics at the Large Hadron Col-\nlider, edited by H. Fritzsch (World Scientific, Singapore, 2017), pp. 161–194.\n[42] M.Bojowald, Canonical Gravity and Applications (CambridgeUniversity Press,\nCambridge, 2011).\n[43] U. D. Jentschura and J. H. Noble, Nonrelativistic Limit of the Dirac–\nSchwarzschild Hamiltonian: Gravitational Zitterbewegun g and Gravitational\nSpin–Orbit Coupling , Phys. Rev. A 88, 022121 (2013).\n[44] U. D. Jentschura, Gravitational Effects in gFactor Measurements and High–\nPrecision Spectroscopy: Limits of Einstein’s Equivalence P rinciple, Phys. Rev.\nA98, 032508 (2018).\n[45] P. Arias, J. Gamboa, J. Lopez-Sarrion, F. Mendez, and A. Das,CPT/Lorentz\ninvariance violation and neutrino oscillation , Phys. Lett. B 650, 401–406\n(2007).\n[46] S. Liberati, Tests of Lorentz invariance: a 2013 update , Class. Quantum Grav.\n30, 133001 (2013).\n[47] V. Antonelli, L. Miramonti, and M. D. C. Torri, Neutrino oscillations and\nLorentz invariance violation in a Finslerian geometrical m odel, Eur. Phys. J. C\n78, 667 (2018).\n[48] M. D. C. Torri, Neutrino Oscillations and Lorentz Invariance Violation , Uni-\nverse202006, 37 (2020).\n[49] A. G. Cohen and S. L. Glashow, Pair Creation Constrains Superluminal Neu-\ntrino Propagation , Phys. Rev. Lett. 107, 181803 (2011).\n[50] F.Bezrukov andH.M.Lee, Model dependence of the bremsstrahlung effects from\nthe superluminal neutrino at OPERA , Phys. Rev. D 85, 031901(R) (2012).\n[51] G. Somogyi, I. N´ andori, and U. D. Jentschura, Neutrino Splitting for Lorentz–\nViolating Neutrinos: Detailed Analysis , Phys. Rev. D 100, 035036 (2019).\n[52] U. D. Jentschura, I. N´ andori, and G. Somogyi, Lorentz Breaking and SU(2)L×\nU(1)YGauge Invariance for Neutrino Decays , Int. J. Mod. Phys. E 28, 1950072\n(2019).\n[53] E. Byckling and K. Kajantie, Particle Kinematics (J. Wiley & Sons, New York,\nNY, 1973).\n15[54] F. W. Stecker, Limiting superluminal electron and neutrino velocities usi ng the\n2010 Crab Nebula flare and the IceCube PeV neutrino events , Astropart. Phys.\n56, 16–18 (2014).\n[55] F. W. Stecker and S. T. Scully, Propagation of superluminal PeV IceCube neu-\ntrinos: A high energy spectral cutoff of new constraints on Lo rentz invariance\nviolation, Phys. Rev. D 90, 043012 (2014).\n[56] IceCube Collaboration, First Observation of PeV-Energy Neutrinos with Ice-\nCube, Phys. Rev. Lett. 111, 021103 (2013).\n[57] IceCube Collaboration, Observation of High-Energy Astrophysical Neutrinos in\nThree Years of IceCube Data , Phys. Rev. Lett. 113, 101101 (2014).\n[58] J. L. Chkareuli and J. G. Jejeleva, Spontaneous Lorentz violation: Non-Abelian\ngauge fields as pseudo–Goldstone vector bosons , Phys. Lett. B 659, 754–760\n(2008).\n[59] W.Heisenberg, Quantum Theory of Fields and Elementary Particles , Rev.Mod.\nPhys.29, 269–278 (1957).\n[60] J. D. Bjorken, A Dynamical Origin for the Electromagnetic Field , Ann. Phys.\n(N.Y.)24, 174–187 (1963).\n[61] I. Bialynicki-Birula, Quantum Electrodyanmics without Electromagnetic Field ,\nPhys. Rev. 130, 465–468 (1963).\n[62] T. Eguchi, New approach to collective phenomena in superconductivity models,\nPhys. Rev. D 14, 2755–2763 (1976).\n[63] J. L. Chkareuli, C. D. Froggatt, and H. B. Nielsen, Lorentz Invariance and\nOrigin of Symmetries , Phys. Rev. Lett. 87, 091601 (2001).\n[64] J. L. Chkareuli, C. D. Froggatt, and H. B. Nielsen, Spontaneously generated\ngauge invariance , Nucl. Phys. B 609, 46–60 (2001).\n[65] J. D. Bjorken, Emergent Gauge Bosons , hep-th/0111196.\n[66] A. T. Azatov and J. L. Chkareuli, Nonlinear QED and physical Lorentz invari-\nance, Phys. Rev. D 73, 065026 (2006).\n[67] J. L. Chkareuli and Z. R. Kepuladze, Nonlinear massive QED and physical\nLorentz invariance , Phys. Lett. B 644, 212–217 (2007).\n[68] J. L. Chkareuli, C. D. Froggatt, J. G. Jejeleva, and H. B. Nielsen, Constrained\ngauge fields from spontaneous Lorentz violation , Nucl. Phys. B 796, 211–223\n(2008).\n[69] J. L. Chkareuli, C. D. Froggatt, and H. B. Nielsen, Deriving Gauge Symmetry\nand Spontaneous Lorentz Violation , Nucl. Phys. B 821, 65–73 (2009).\n16" }, { "title": "2210.02114v1.Searching_for_Lorentz_violating_Signatures_from_Astrophysical_Photon_Observations.pdf", "content": "arXiv:2210.02114v1 [astro-ph.HE] 5 Oct 2022Proceedings of the Ninth Meeting on CPT and Lorentz Symmetry (CPT’22), Indiana University, Bloomington, May 17–26, 2022\n1\nSearching for Lorentz-violating Signatures from Astrophy sical\nPhoton Observations\nJun-Jie Wei1,2,3\n1Purple Mountain Observatory, Chinese Academy of Sciences,\nNanjing 210023, China\n2School of Astronomy and Space Sciences, University of Scien ce and Technology of\nChina, Hefei 230026, China\n3Guangxi Key Laboratory for Relativistic Astrophysics,\nNanning 530004, China\nAs a basic symmetry of Einstein’s theory of special relativi ty, Lorentz invari-\nance has withstood very strict tests. But there are still mot ivations for such\ntests. Firstly, many theories of quantum gravity suggest vi olations of Lorentz\ninvariance at the Planck energy scale. Secondly, even minut e deviations from\nLorentz symmetry can accumulate as particle travel across l arge distances,\nleading to detectable effects at attainable energies. Thank s to their long base-\nlines and high-energy emission, astrophysical observatio ns provide sensitive\ntests of Lorentz invariance in the photon sector. In this pap er, I briefly in-\ntroduce astrophysical methods that we adopted to search for Lorentz-violating\nsignatures, including vacuum dispersion and vacuum birefr ingence.\n1. Introduction\nAlthough experimental tests of Lorentz invariance violation (LIV) have\nbeen performed in a wide range of systems (see Ref. 1 for a compilat ion of\nresults), there are still motivations for such tests. From theore tical consid-\neration, establishing a quantum theory that includes gravity is being hailed\nas the Holy Grail of modern physics. However, theories of quantum gravity\n(QG) predict that Lorentz invariance may be violated at the Planck e nergy\nscaleEPl. From experimental feasibility, even very tiny deviations from\nLorentz symmetry can become measurable at attainable energies ≪EPl,\nsince Lorentz-violating effects gradually accumulate over large pro pagation\ndistances. In brief, both the prospect of relativity violations arisin g in a\ngrand unified theory and the feasibility of discovering LIV have attr acted\nphysicists to constantly work on experimental searches.Proceedings of the Ninth Meeting on CPT and Lorentz Symmetry (CPT’22), Indiana University, Bloomington, May 17–26, 2022\n2\nAstrophysical observations involving long baselines are very suitab le to\nsearch for LIV effects. In vacuum, the LIV-induced modifications to the\nphoton dispersion relation can produce rich and detectable astrop hysical\nphenomena. With a slight modification, the speed of light will no longer\nbe independent of frequency and polarization. If the modified dispe rsion\nrelation is related to the frequency of photons, then we may obser ver a\nfrequency-dependent vacuum dispersion of light. If the modificat ion is re-\nlated to the circular polarization state of photons, the photons wit h right-\nand left-handed polarization states have different velocities, then we may\nobserver vacuum birefringence. Vacuum dispersion can be tested by mea-\nsuringthe arrival-timedifferencesofphotonswith differentfreque nciesorig-\ninating from the same astrophysical source (see, e.g., Ref. 2). Va cuum\nbirefringence results in a frequency-dependent rotation of the p olarization\nvector of linearly polarized light, which can be tested by astrophysic alspec-\ntropolarimetric measurements (see, e.g., Refs. 3, 4).\nIn this paper, we present some recent tests of nonbirefringent L IV using\nspectral-lag transitions of gamma-ray bursts (GRBs),5,6and some searches\nfor LIV using polarization measurements of GRBs and blazars.7,8\n2. Vacuum dispersion\nVacuum dispersion is a potential LIV signature. The arrival time dela y\nbetween photons with different energies ( Eh> El) emitted simultaneously\nfrom the source at redshift zcan be derived by introducing the LIV terms\nin a Taylor series:9\n∆t=s±1+n\n2H0En\nh−En\nl\nEn\nQG,n/integraldisplayz\n0(1+z′)ndz′\n/radicalbig\nΩm(1+z′)3+ΩΛ, (1)\nwheres±=±1 is the sign of LIV, corresponding to the “subluminal”\n(s±= +1) or “superluminal” ( s±=−1) scenarios. EQGdenotes the\nQG energy scale at which LIV effects become significant, n= 1 (n= 2)\ncorrespondsto the linear (quadratic)energydependence, and( H0, Ωm, ΩΛ)\nare the cosmological parameters of the standard flat ΛCDM model.\nGRBs are ideal astrophysical phenomena that one can used to per form\ntime-of-flight tests because they are the most distant transient sources in-\nvolving a wide range of photon energies. However, a key challenge in s uch\ntests is to distinguish an intrinsic astrophysical time delay at the sou rce\nfrom a time delay induced by LIV. We proposed that GRB 160625B, th e\nburst having an apparent transition from positive to negative spec tral lags,Proceedings of the Ninth Meeting on CPT and Lorentz Symmetry (CPT’22), Indiana University, Bloomington, May 17–26, 2022\n3\nprovides a good opportunity to disentangle the intrinsic time delay pr ob-\nlem.5Spectral lag, the arrivaltime difference between high- and low-ene rgy\nphotons, is conventionally defined to be positive if the high-energy p hotons\nprecede the low-energy ones. In the subluminal case ( s±= +1), photons\nwith higher energies would arrive at the observer after those with lo wer\nones, implying a negative spectral lag due to LIV. Assuming the sour ce-\nintrinsic time lag to have a positive dependence on the photon energy ,\nand considering the contributions to the spectral lag from both th e in-\ntrinsic positive lag and LIV-related negative lag, we derived new limits o n\nlinear and quadratic leading-order Lorentz-violating vacuum disper sion by\ndirectly fitting the spectral lag behavior of GRB 160625B.Recently , similar\ntime-of-flight tests were carried out by analyzing the spectral-lag transition\nof GRB 190114C.6\n3. Vacuum birefringence\nVacuum birefringence is another potential LIV signature. For a so urce at\nredshiftz, the LIV-induced rotation angle of the polarization vector of a\nlinearly polarized wave is4\n∆φ(E)≃ηE2\n/planckover2pi1EPlH0/integraldisplayz\n0(1+z′)dz′\n/radicalbig\nΩm(1+z′)3+ΩΛ, (2)\nwhereEis the energy of the observed light and ηis a dimensionless pa-\nrameter characterizing the degree of LIV effects.\nIf the rotation angles of photons with different energies differ by mo re\nthanπ/2 over an energy band ( E1< E < E 2), significant depletion of\nthe initial polarization of the signal is expected. The detection of hig h\npolarization can therefore set upper limits on the birefringent para meterη.\nWe gave a detailed calculation on the GRB polarization evolution arising\nfrom the birefringent effect, and confirmed that the initial polariza tion is\nnot significantly depleted even if the differential rotation angle |∆φ(E2)−\n∆φ(E1)|is as large as π/2.7Applying our formulate for calculating LIV-\ninduced polarization evolution to the gamma-ray polarimetric data of 12\nGRBs, we improved existing bounds on the birefringent parameter ηby\nfactors ranging from 2 to 10.\nIf all photons in the observed energy range are assumed to be emit ted\nwith the same intrinsic polarization angle, we expect to observe vacu um\nbirefringence as an energy-dependent linear polarization vector. We tried\nto search for an energy-dependent change of the linear polarizat ion angle\nin the spectropolarimetric data of 5 blazars.8At the 2 σconfidence level,Proceedings of the Ninth Meeting on CPT and Lorentz Symmetry (CPT’22), Indiana University, Bloomington, May 17–26, 2022\n4\nthe absence of the birefringent effect was limited to be in the range o f\n−9.63×10−8< η <6.55×10−6. As might be expected, opticalpolarimetric\ndata of blazars pose less stringent constraints on ηas compared to gamma-\nray polarizations of GRBs.\n4. Summary\nGRBs are promising astrophysical sources for searching for LIV- induced\nvacuum dispersion and vacuum birefringence. Future detections o f\nextremely-high-energy emission from GRBs with LHAASO, MAGIC,\nHAWC, and the future international Cherenkov Telescope Array c ould im-\nprove the limits on LIV using vacuum dispersion (time-of-flight) test s. Fu-\nture X-ray/gamma-ray polarization measurements of GRBs with PO LAR-\nII, TSUBAME, COSI, and GRAPE could also improve the limits on LIV\nusing vacuum birefringence tests.\nAcknowledgments\nThis work is partially supported by the National Natural Science Fou nda-\ntion of China (grant Nos. 11725314 and 12041306), the Key Resea rch Pro-\ngram of Frontier Sciences (grant No. ZDBS-LY-7014) of Chinese A cademy\nof Sciences, the Major Science and Technology Project of Qinghai Province\n(2019-ZJ-A10), the Natural Science Foundation of Jiangsu Prov ince (grant\nNo. BK20221562), the China Manned Space Project (CMS-CSST-2 021-\nB11), and the Guangxi Key Laboratory for Relativistic Astrophys ics.\nReferences\n1.Data Tables for Lorentz and CPT Violation, V.A. Kosteleck´ y and N. Russell,\n2022 edition, arXiv:0801.0287v15.\n2. G. Amelino-Camelia et al., Nature 393, 763 (1998); A.A. Abdo et al., Nature\n462, 331 (2009); V. Vasileou et al., Phys. Rev. D 82, 122001 (2013).\n3. V.A. Kosteleck´ y and M. Mewes, Phys. Rev. Lett. 87, 251304 (2001); V.A.\nKosteleck´ y and M. Mewes, Phys. Rev. Lett. 110, 201601 (2013).\n4. P. Laurent et al., Phys. Rev. D 83, 121301 (2011); K. Toma et al., Phys.\nRev. Lett. 109, 241104 (2012).\n5. J.-J. Wei et al., Astrophys. J. Lett. 834, L13 (2017).\n6. S.-S. Du et al., Astrophys. J. 906, 8 (2021).\n7. J.-J. Wei, Mon. Not. Royal Astron. Soc. 485, 2401 (2019).\n8. Q.-Q. Zhou et al., Galaxies 9, 44 (2021).\n9. U. Jacob and T. Piran, J. Cosmol. Astropart. Phys. 01, 031 (2008)." }, { "title": "1109.6624v4.Superluminal_neutrino_and_spontaneous_breaking_of_Lorentz_invariance.pdf", "content": "arXiv:1109.6624v4 [hep-ph] 20 Dec 2011JETP Lett. 94 (2011) 673 arXiv:1109.6624\nSuperluminal neutrino and spontaneous breaking of Lorentz invariance\nF.R. Klinkhamer#1)and G.E. Volovik∗+1)\n#Institute for Theoretical Physics, University of Karlsruh e, Karlsruhe Institute of Technology, 76128 Karlsruhe, Ger many\n∗Low Temperature Laboratory, Aalto University P.O. Box 1510 0, FI-00076 AALTO, Finland\n+L.D. Landau Institute for Theoretical Physics, Russian Aca demy of Sciences, Kosygina 2, 119334 Moscow, Russia\nSubmitted October 10, 2011\nGenerally speaking, the existence of a superluminal neutri no can be attributed either to re-entrant Lorentz\nviolation at ultralow energy from intrinsic Lorentz violat ion at ultrahigh energy or to spontaneous breaking\nof fundamental Lorentz invariance (possibly by the formati on of a fermionic condensate). Re-entrant Lorentz\nviolation in the neutrino sector has been discussed elsewhe re. Here, the focus is on mechanisms of spontaneous\nsymmetry breaking.\nPACS: 11.30.Cp, 11.30.Qc, 14.60.St\nIt is possible that OPERA’s claimed discovery [1] of\na superluminal muon-type neutrino does not come from\nthe violation of Lorentz invariance but from unknown\nfactorsin the clock-synchronizationprocess[2]orfroma\npurely statistical effect [3]. In fact, it has been shown[4]\nthat OPERA’s claimed value ( vνµ−c)/c∼10−5is ruled\nout by the expected but unobserved energy losses from\nelectron-positron-pair emission ( νµ→νµ+e−+e+), at\nleast, as long as there exists a preferred frame from the\nLorentz violation.\nStill, the claim by OPERA has provided new impe-\ntus for the discussion on the possible sources of Lorentz\nviolation. Inordertoengageinthisdiscussion, let usas-\nsume that OPERA’s result is correct qualitatively (ex-\nistence of a superluminal muon-neutrino) even if not\nquantitatively (most likely, |vνµ−c|/c≪10−5).\nCondensed-matter physics, which possesses an ana-\nlog of Lorentz invariance (LI), now suggests several dif-\nferent scenarios of Lorentz violation (LV). Among them\nare:\n(1a) LI is not a fundamental symmetry but an approxi-\nmatesymmetrywhichemergesatlowenergiesand\nis violated at ultrahigh energies (cf. [5]).\n1)e-mail: frans.klinkhamer@kit.edu, volovik@boojum.hut. fi(1b) Intrinsic LV at ultrahigh energies gives an emer-\ngent Lorentz-invariant theory at lower energies\nbut ultimately, at or below an ultralow energy\nscale,inducesare-entrantviolationofLI(see, e.g.,\nSec. 12.4 of [6]).\n(2) LI is fundamental but broken spontaneously (see,\ne.g., [7, 8] and references therein).\nIn this Letter, we discuss the spontaneous break-\ning of Lorentz invariance (SBLI), that is, the sponta-\nneous appearance of a preferred frame in the vacuum,\nwhich can be derived from Lorentz-invariant physical\nlaws. The order parameter of SBLI can be a vector\nfieldbα(for example, the vector field of Fermi-point\nsplitting [9, 10] or an aether-type velocity field [11]), an\nemergent tetrad-type field eα\na[12, 13, 14, 15], or any\nother field which is covariant but not invariant under\nLorentz transformations.\nIf SBLI occurs only in the neutrino sector, which\ninteracts weakly with the charged-matter sector, then\nSBLI has no direct impact on this other matter (cer-\ntain indirect quantum-loop effects can be suppressed by\nnear-zero mixing angles). The non-neutrino matter es-\nsentially does not feel the existence of the preferred ref-\nerence frame. In fact, it is very well possible that SBLI\noccurs only for the neutrino field, because the other\nfermions have already experienced electroweak symme-\n12 F.R. Klinkhamer, G.E.Volovik\ntry breakingphase transitions and aretoo heavyfor any\nfurther type of symmetry breaking.\nIn condensed-matter physics, the re-entrant viola-\ntion of LI, as well as Fermi-point splitting (FPS), fol-\nlow from general topological properties of the vacuum\n(ground state) in 3–momentum space. Condensed mat-\nter provides many examples of homogeneous vacuum\nstates, which have nontrivial topology in 3–momentum\nspace [6]. Among them is a class of vacua which have\nFermi points, exceptional points in 3–momentum space\nwhere the energy of fermionic excitations is nullified.\nSuch a Fermi point (alternatively called Dirac or Weyl\npoint) has a topological invariant. The existence of\nthe Fermi point is thus protected by topology, or by\nthe combined action of topology and symmetry. The\nFermi point is robust to small perturbations of the sys-\ntem. In turn, different Fermi points may collide, an-\nnihilate, and split again, but their total topological\ncharge is conserved. In this respect, Fermi points in\n3–momentum space behave as topologically-charged ’t\nHooft–Polyakovmagnetic monopoles in real space. The\nsplitting or recombination of Fermi points represents\na topological quantum phase transition. This type of\nquantum phase transition takes place, for example, in\ngraphene, graphite, etc. (see, e.g., Figs. 4 and 6 in [16]\nfor the splitting of a degenerate Fermi point into 3 and\n4 elementary points, respectively).\nFigure 8 in the review [10] (which elaborates on the\ndiscussion of the original research paper [9]) illustrates\nthespecialroleofneutrinos. Fromthemomentum-space\ntopology of Fermi points, it follows that the phase tran-\nsition awayfrom the symmetric vacuum ofthe Standard\nModel with massless fermions may occur in two ways:\neither coinciding Fermi points with opposite topologi-\ncal charge annihilate each other, giving rise to a Dirac\nmass (the process commonly known as the Higgs mech-\nanism), orcoinciding Fermi points do not annihilate but\nsplit in momentum space, giving rise to Lorentz viola-\ntion [9]. As mentioned above, it is possible that this\nFPS process only occurs for neutrinos, since all other\nparticles have already obtained Dirac masses via the\nHiggs mechanism. After this splitting, the energy spec-tra of the left- and right-handed neutrino are given by\nthe following expressions (the small neutrino mass can\nbe neglected for the conditions relevant to the OPERA\nexperiment):\ngαβ/parenleftBig\ncpα−/tildewidebL\nα/parenrightBig/parenleftBig\ncpβ−/tildewidebL\nβ/parenrightBig\n= 0, (1a)\ngαβ/parenleftBig\ncpα−/tildewidebR\nα/parenrightBig/parenleftBig\ncpβ−/tildewidebR\nβ/parenrightBig\n= 0, (1b)\nwithcthe velocity of light in vacuo . In (1), we have\nput a tilde on the dimensional vector /tildewidebαin order to dis-\ntinguish it from the dimensionless vector bαappearing\nbelow and we allow for /tildewidebL\nα∝negationslash=/tildewidebR\nα.\nThe possible role of FPS for the (qualitative)\nOPERA result has already been discussed in [17]. Here,\nwe considertwoscenariosforthe spontaneousformation\nof a preferred reference frame. The first scenario corre-\nsponds to the appearance of a dimensionless 4–vector\n(bα) = (b0,b) in the neutrino vacuum. This 4–vector bα\ninteracts with the neutrino Dirac field in a way which\ndoes not violate the fundamental laws of special relativ-\nity:\nS=/integraldisplay\nd3xcdtψbα(−i∇α)ψ. (2)\nThis action term corresponds to a momentum-\ndependent mass term,2)M=pαbα/c, which modifies\nthe spectrum of the neutrino as follows:\npαpα≡E2−c2|p|2=/parenleftbig\ncpαbα/parenrightbig2.(3)\nLet us, for example, take bαto be a timelike vector,\nhaving (b0)2− |b|2>0 , and consider the particular\nreference frame with b= 0. Assume |b0|<1. Then,\nthe neutrino energy spectrum becomes\nE2/bracketleftbig\n1−(b0)2/bracketrightbig\n=c2|p|2, (4)\nwhich is superluminal for b0∝negationslash= 0. The same happens\nfor a spacelike vector bα. In the reference frame with\nb0= 0, the neutrino energy spectrum is given by\nE2=c2|p|2+c2(b·p)2, (5)\n2)At this stage, it is clear that the same procedure can be fol-\nlowed with a Majorana mass term ψc\nLmψLin the action density,\nsimply replacing the Majorana mass mby−ibα∇α/c.Superluminal neutrino and spontaneous breaking of Lorentz invar iance 3\nwhich is both anisotropic and superluminal for b∝negationslash= 0.\nThe 4–vector field bαmay emerge as the order pa-\nrameter of the neutrino condensate,\nbα∝gαβ/angbracketleftbig\nψ(−i∇β)ψ/angbracketrightbig\n, (6)\nin a theory with 4–fermionormulti-fermion interactions\nof the following type:\nSint=/integraldisplay\nd3xcdt f(X), (7a)\nX=−gαβ/parenleftbig\nψ∇αψ/parenrightbig /parenleftbig\nψ∇βψ/parenrightbig\n, (7b)\nwith appropriate dimensional constants entering the\nfunctionf. This scenario gives a possible realization\nof the phenomenological Coleman–Glashow model [18]\nin terms of fermionic condensates [7, 8].\nThe second scenario involves another type of neu-\ntrino condensate, which also leads to SBLI. Specifi-\ncally, this neutrino condensate gives rise to a tetrad-like\nfield [12, 13, 14, 15]:\nea\nα∝/angbracketleftbig\nψγa(−i∇α)ψ/angbracketrightbig\n. (8)\nThe induced tetrad field ea\nαmust be added to the orig-\ninal fundamental tetrad E(0)α\na= diag(−1,c, c, c), and\nthe fermionic action becomes\nS=/integraldisplay\nd3xcdt Eα\naψγa(−i∇α)ψ , (9a)\nEα\na=E(0)α\na+eα\na. (9b)\nUsing the induced tetrad field ( eα\na) = diag(b0,0,0,0)\nas an example, one obtains, for 0 < b0<1, the super-\nluminal neutrino velocity vν=c/(1−b0).\nThe tetrad-type neutrino condensate (8) may pro-\nvide a realization of the hypothetical spin-2 field dis-\ncussed in [19]. A recent paper [20] presents another\nmodel, where a scalar-field composite plays a similar\nrole as our condensate eα\na. Also related may be a geo-\nmetric model [21], based on a particular class of Finsler-\nspacetime backgrounds, which essentially modifies the\neffective metric entering particle dispersion relations.\nThiscompletesourdiscussionoftwopossible scenar-\nios of spontaneous symmetry breaking to explain a su-\nperluminal neutrino. Spontaneous breaking of Lorentzinvariance in the neutrino sector corresponds to the ap-\npearance of a preferred frame for the relevant neutrino:\nLI is violated if the neutrino momentum pαis trans-\nformed but not the vacuum field bα(orea\nα) which is\nkept at a fixed value. Still, LI remains an exact sym-\nmetry of the physical laws: the invariance holds if both\nexcitations and vacuum are transformed, that is, if the\nLorentz transformation acts simultaneously on pαand\nbα(orea\nα).\nIn these SBLI scenarios, as well as in the FPS sce-\nnario [17], the vacuum remains homogeneous, which is\nthereasonwhyconservationofenergyandmomentumis\nexact. But the energy spectrum of the neutrino is mod-\nified, which must certainly have consequences for reac-\ntions involving neutrinos. Hence, there must be experi-\nmental constraints on bαorea\nα. Alternatively, the study\nof neutrino-interaction processes may provide valuable\ninformation on mechanisms proposed to explain non-\nstandard (e.g., superluminal) propagation properties of\nthe neutrinos.\nThe advantage of the spontaneous-symmetry-\nbreaking scenario is that it stays fully within the realm\nof standard physics, which obeys special relativity. The\nmulti-fermion interaction (7) can, in principle, origi-\nnate from trans-Planckian physics, but we now have\nbounds [22, 23] indicating that Lorentz invariance holds\nfar above the Planck energy scale, i.e., ELV≫EPlanck.\nThis suggests that, if a neutrino has superluminal mo-\ntion, it can be attributed either to re-entrant Lorentz\nviolation at ultralow energy due to intrinsic (built-in)\nLorentz violation at ultrahigh trans-Planckian ener-\ngies [presumably with a re-entrance energy of order\n(EPlanck/ELV)nEPlanckforn≥1] or to spontaneous\nbreaking of fundamental Lorentz invariance [possibly\nby the formation of a fermionic condensate].\n1. T. Adam et al.[OPERA Collaboration], “Measurement\nof the neutrino velocity with the OPERA detector in\nthe CNGS beam,” arXiv:1109.4897v1.\n2. C.R. Contaldi, “The OPERA neutrino velocity result\nand the synchronisation of clocks,” arXiv:1109.6160.4 F.R. Klinkhamer, G.E.Volovik\n3. R. Alicki, “A possible statistical mechanism of\nanomalous neutrino velocity in OPERA experiment?,”\narXiv:1109.5727.\n4. A.G.CohenandS.L.Glashow, “Newconstraints onneu-\ntrino velocities,” arXiv:1109.6562.\n5. S. Chadha and H.B. Nielsen, “Lorentz invariance as\na low-energy phenomenon,” Nucl. Phys. B 217, 125\n(1983).\n6. G.E. Volovik, The Universe in a Helium Droplet ,\nClarendon Press, Oxford (2003).\n7. J.D. Bjorken, “A dynamical origin for the electromag-\nnetic field,” Annals Phys. 24, 174 (1963).\n8. A. Jenkins, “Spontaneous breaking of Lorentz\ninvariance,” Phys. Rev. D 69, 105007 (2004),\narXiv:hep-th/0311127.\n9. F.R. Klinkhamer and G.E. Volovik, “Emergent CPT vi-\nolation from the splitting of Fermi points,” Int. J. Mod.\nPhys. A 20, 2795 (2005), arXiv:hep-th/0403037.\n10. G.E. Volovik, “Quantum phase transitions from topol-\nogy in momentum space,” Lect. Notes Phys. 718, 31\n(2007), arXiv:cond-mat/0601372.\n11. T. Jacobson, “Einstein–aether gravity: A status re-\nport,” PoS QG-PH , 020 (2007), arXiv:0801.1547.\n12. K. Akama, “An attempt at pregeometry – gravity with\ncomposite metric,” Prog. Theor. Phys. 60, 1900 (1978).\n13. G.E. Volovik, “Superfluid3He–B and gravity,” Physica\nB162, 222 (1990).14. C. Wetterich, “Gravity from spinors,” Phys. Rev. D 70,\n105004 (2004), arXiv:hep-th/0307145.\n15. D. Diakonov, “Towards lattice-regularized quantum\ngravity,” arXiv:1109.0091.\n16. T.T. Heikkil¨ a and G.E. Volovik, “Fermions with cu-\nbic and quartic spectrum,” JETP Lett. 92, 681 (2010),\narXiv:1010.0393.\n17. F.R. Klinkhamer, “Superluminal muon-neutrino veloc-\nity from a Fermi-point-splitting model of Lorentz viola-\ntion,” arXiv:1109.5671.\n18. S. R. Coleman and S. L. Glashow, “Cosmic ray and neu-\ntrino tests of special relativity,” Phys. Lett. B 405, 249\n(1997), arXiv:hep-ph/9703240; S. R. Coleman and S.\nL. Glashow, “High-energy tests of Lorentz invariance,”\nPhys. Rev.D 59, 116008 (1999), arXiv:hep-ph/9812418.\n19. G. Dvali and A. Vikman, “Price for environmental\nneutrino-superluminality,” arXiv:1109.5685.\n20. A. Kehagias, “Relativistic superluminal neutrinos,”\narXiv:1109.6312.\n21. C. Pfeifer and M.N.R. Wohlfarth, “Beyond the speed of\nlight on Finsler spacetimes,” arXiv:1109.6005.\n22. O. Gagnon and G.D. Moore, “Limits on Lorentz viola-\ntion from the highest energy cosmic rays,” Phys. Rev.\nD70, 065002 (2004), arXiv:hep-ph/0404196.\n23. S. Bernadotte and F.R. Klinkhamer, “Bounds on length\nscales of classical spacetime foam models,” Phys. Rev.\nD75, 024028 (2007), arXiv:hep-ph/0610216." }, { "title": "1909.01990v1.Probing_Lorentz_Invariance_With_Top_Pair_Production_at_the_LHC_and_Future_Colliders.pdf", "content": "Proceedings of the Eighth Meeting on CPT and Lorentz Symmetry (CPT'19), Indiana University, Bloomington, May 12{16, 2019\n1\nProbing Lorentz Invariance\nWith Top Pair Production at the LHC and Future Colliders\nA. Carle, N. Chanon, and S. Perri\u0012 es\nUniversit\u0013 e de Lyon, Universit\u0013 e Claude Bernard Lyon 1,\nCNRS-IN2P3, Institut de Physique Nucl\u0013 eaire de Lyon,\nVilleurbanne 69622, France\nThis article presents prospects for Lorentz-violation searches with t\u0016tat the\nLHC and future colliders. After a short presentation of the Standard-Model\nExtension as a Lorentz-symmetry-breaking e\u000bective \feld theory, we will focus\nont\u0016tproduction. We study the impact of Lorentz violation as a function of\ncenter-of-mass energy and evaluate the sensitivity of collider experiments to\nthis signal.\n1. Introduction\nThe top-quark sector of Standard-Model Extension (SME) is weakly con-\nstrained. Since the LHC is a top factory, it provides a unique opportunity\nto search for Lorentz violation (LV). The SME is an e\u000bective \feld theory\nincluding all LV operators. Here, we consider the LV CPT-even part of the\nlagrangian modifying the top-quark kinematics:1\nLSME\u001bi\n2(cL)\u0016\u0017\u0016Qt\r\u0016$\nD\u0017Qt+i\n2(cR)\u0016\u0017\u0016Ut\r\u0016$\nD\u0017Ut; (1)\nwhereQtandUtdenote the left- and right-handed top-quark spinors, re-\nspectively. The c\u0016\u0017coe\u000ecients are constant in an inertial frame, taken to\nbe the Sun-centered frame. We aim at measuring the constant coe\u000ecients:2\nc\u0016\u0017=1\n2[(cL)\u0016\u0017+ (cR)\u0016\u0017]; d\u0016\u0017=1\n2[(cL)\u0016\u0017\u0000(cR)\u0016\u0017]: (2)\nExpressions for these coe\u000ecients in a laboratory frame on Earth will intro-\nduce a time dependence of the cross section for t\u0016tproduction owing to the\nEarth's rotation around its axis. This time dependence can be exploited to\nsearch for LV at hadron colliders.\nTo express the c\u0016\u0017coe\u000ecients in the reference frame of a hadron circular\ncollider, we need:arXiv:1909.01990v1 [hep-ph] 4 Sep 2019Proceedings of the Eighth Meeting on CPT and Lorentz Symmetry (CPT'19), Indiana University, Bloomington, May 12{16, 2019\n2\n\u000fthe latitude \u0015, i.e., the angle between the equator and the poles,\n\u000fthe azimuth \u0012,3i.e., the angle between the Greenwich tangent vec-\ntor and the clockwise ring collider tangent vector,\n\u000fthe longitude impacts only the phase of the signal because of the\nEarth's rotation around its axis, and\n\u000fthe Earth's angular velocity \n.\n2. Modulation of the t\u0016tcross section\nThe analysis aims at measuring the time dependence of the t\u0016tcross section\n\u001bSME= [1 +f(t)]\u001bSM: (3)\nA \frst analysis of this kind was performed with the D0 detector at the\nTevatron.4We use here the same benchmarks. We analyze Wilson's coe\u000e-\ncients for a couple of non-null c\u0016\u0017:cXX=\u0000cYY,cXY=cYX,cXZ=cZX\norcYZ=cZY. Each of these scenarios generates an oscillating behavior\nof the amplitude. The latitude \u0015and the azimuth \u0012a\u000bect the amplitude\nwhile the Earth's angular velocity \n a\u000bects the frequency. In the case of\ncXX=\u0000cYYandcXY=cYX,f(t) has a period of one sidereal day. On\nthe other hand, in the cXZ=cZXandcYZ=cZYcase, the amplitude has\na period of one half of a sideral day. More detailed expressions are given in\nRefs. 2, 4.\n3. Expected sensitivity\nIn this work, samples of t\u0016twith dilepton decay were generated with\nMadGraph-aMC@NLO 2.6. It was found that the amplitude of the LV\nt\u0016tsignal is increasing with the center-of-mass energy. The signal ampli-\ntude as a function of the center-of-mass energy in p{pcollisions (with CMS\nor ATLAS as the laboratory frame) increases from 0 :001 at D0 (in the\ncXY=cYX= 0:01 scenario) to 0 :045 at the LHC Run II (13 TeV) and to\n0:055 at the Future Circular Collider (FCC, 100 TeV).\nWe evaluate the expected sensitivity to the signal for each benchmark.5\nAs a consequence of the increase in luminosity, the increase in cross section,\nand the increase in the amplitude of the LV signal, we \fnd the following\nexpected sensitivities to the SME coe\u000ecient c\u0016\u0017in thecXX=\u0000cYYcase:\n\u000f\u0001c= 7\u000210\u00001: D0 (ps= 1:96 TeV,L= 5:3 fb\u00001),\n\u000f\u0001c= 1\u000210\u00003: LHC Run II (ps= 13 TeV,L= 150 fb\u00001),\n\u000f\u0001c= 2\u000210\u00004: HL-LHC (ps= 14 TeV,L= 3000 fb\u00001),Proceedings of the Eighth Meeting on CPT and Lorentz Symmetry (CPT'19), Indiana University, Bloomington, May 12{16, 2019\n3\n\u000f\u0001c= 3\u000210\u00005: HE-LHC (ps= 27 TeV,L= 15 ab\u00001),\n\u000f\u0001c= 9\u000210\u00006: FCC (ps= 100 TeV,L= 15 ab\u00001).\n4. Signal amplitude at hadron colliders\nA noticeable fact is the dependence of the signal amplitude on the lati-\ntude and azimuth of the collider experiment on Earth. This dependence\nis presented in Fig. 1. We \fnd that performing such an experiment at the\nAmplitude fSME (/g79,/g84) fXX Azimuth /g84 (in rad) CMS\n0000.20.40.60.8\n0.5 111.21.4\n1.5 -0.5 -1 -1.5123456\nATLAS \nD0 \nLatitude /g79 (in rad)\nFig. 1. Amplitude of f(\u0015;\u0012) as a function of latitude and azimuth for the XX,YY,\nandXYbenchmarks.\nLHC would increase the sensitivity to SME coe\u000ecients in the top sector\nby two orders of magnitude. Further improvements are expected at future\ncolliders.\nAcknowledgments\nThanks to Alan Kosteleck\u0013 y and Ralf Lehnert for giving me the opportunity\nto expose my work in front of a benevolent and hard-working community.\nReferences\n1. D. Colladay and V.A. Kosteleck\u0013 y, Phys. Rev. D 93, 036005 (2016).\n2. M.S. Berger, V.A. Kosteleck\u0013 y, and Z. Liu, Phys. Rev. D 93, 036005 (2016).\n3. M. Jones, Activity Report, EDMS, 322747 (2005).\n4. D0 Collaboration, V.M. Abazov et al. , Phys. Rev. Lett. 108, 261603 (2012).\n5. A. Carle, N. Chanon, and S. Perri\u0012 es, arXiv:1908.11256 [hep-ph]" }, { "title": "1412.2574v3.Bi___cal_PT___symmetry_in_nonlinearly_damped_dynamical_systems_and_tailoring___cal_PT___regions_with_position_dependent_loss_gain_profiles.pdf", "content": "arXiv:1412.2574v3 [nlin.PS] 15 Dec 2015Bi-PTsymmetry in nonlinearly damped dynamical systems and\ntailoring PTregions with position dependent loss-gain profiles\nS. Karthiga1, V.K. Chandrasekar2, M. Senthilvelan1, M. Lakshmanan1\n1Centre for Nonlinear Dynamics, School of Physics,\nBharathidasan University, Tiruchirappalli - 620 024, Tami l Nadu, India.\n2Centre for Nonlinear Science & Engineering, School of Elect rical & Electronics Engineering,\nSASTRA University, Thanjavur -613 401, Tamil Nadu, India.\nWe investigate the remarkable role of position dependent da mping in determining the parametric\nregions of symmetry breaking in nonlinear PT-symmetric systems. We illustrate the nature of\nPT-symmetry preservation and breaking with reference to a rem arkable integrable scalar nonlinear\nsystem. In the two dimensional cases of such position depend ent damped systems, we unveil the\nexistence of a class of novel bi- PT-symmetric systems which have two fold PTsymmetries. We\nanalyze the dynamics of these systems and show how symmetry b reaking occurs, that is whether the\nsymmetry breaking of the two PTsymmetries occurs in pair or occurs one by one. The addition o f\nlinear damping in these nonlinearly damped systems induces competition between the two types of\ndamping. This competition results in a PTphase transition in which the PTsymmetry is broken\nfor lower loss/gain strength and is restored by increasing t he loss/gain strength. We also show that\nby properly designing the form of the position dependent dam ping, we can tailor the PT-symmetric\nregions of the system.\nPACS numbers: 11.30.Er, 05.45.-a, 11.30.Qc\nI. INTRODUCTION\nIn recent times considerable interest has been shown\nin investigating systems which do not show parity ( P)\nand time reversal ( T) symmetries separately but which\nexhibit a combined PTsymmetry. These PT-symmetric\nsystemshaveseveralintriguingfeaturessuchaspoweros-\ncillations [1], absorption enhanced transmission [2], dou-\nble refraction, and non-reciprocity of light propagation\n[1]. Thus, these systems open up novel applications in\noptics [1], quantum optics [3, 4], solid state physics [5],\nmetamaterials [6, 7], optomechanical systems [8, 9], etc.\nThe understanding of PT-symmetric systems as non-\nisolated systems with balanced loss and gain has led to\ntheexplorationofthesesystemsinmechanicsaswellasin\nelectronics. Suchobservationsof PT-symmetricmechan-\nical and electronic systems provide the simplest ground\nto experiment on these PT-symmetric systems [10–14].\nA.Bi-PT symmetry The above oscillator based PT-\nsymmetric systems are generically constructed by cou-\npling an oscillator with linear loss to an oscillator with\nequal amount of linear gain [11–14]. Apart from the\nabove type of systems, there exists a class of interest-\ning dynamical systems with position dependent damp-\ning (or position dependent loss-gain profile) where the\namount of damping depends on its displacement. Conse-\nquently one can have PT-symmetric systems even with\na single degree of freedom. In this case, the systems are\ninvariant with respect to the PToperation defined by P:\nx→ −x,T:t→ −t, so that PT:x→ −x,t→ −twhich\nwe denote as the PT−1 operation. As the position de-\npendent damping term is found to be a nonlinear term in\nthe evolution equation, we call this damping as nonlinear\ndamping for simplicity. The main aim of this paper is to\ninvestigate the dynamics and underlying novel structuresin these systems in comparison with the standard ones.\nThe recent explorations on the damping in systems\nwith one or more atomic-scale dimensions have unveiled\nthat the damping present in these systems is strongly po-\nsition dependent [15–17]. Ref. [15] shows that this type\nof damping in mechanical resonators enhances the fig-\nure of merit of the system tremendously. In particular,\nwith this type of damping, a quality factor of 100 ,000\nhas been achieved with graphene resonators. In addi-\ntion, such systems are found to play an important role in\nmany areas of physics, biology and engineering [18] and\nthey are typically called Li´ enard systems or Li´ enard os-\ncillators. Recently, a class of chemical and biochemical\noscillations which are governed by two-variable kinetic\nequations are shown to be reducible to Li´ enard systems\nby linear transformations. As the nonlinear damping\nterm in the Li´ enard systems can act as a damping term\nor a pumping term depending on the amplitude of the\noscillation, through an internal energy source, it gives\nrise to self sustained oscillations. The above property\nenables one to understand and to control several chem-\nical and biochemical oscillations which are discussed in\n[19]. Li´ enard systems are also found to be paradigmatic\nmodels in the biological regulatory systems [20]. For ex-\nample, they have been used to model the heart and res-\npiratory systems (van-der Pol equation [21, 22]) and the\nnerve impulse (FitzHugh-Nagumo equations [23]). The\nLi´ enard equation with a cubic polynomial potential has\nbeen used to describe the isotropic turbulence [24]. One\ncan also find the appearance ofthese systems in reaction-\ndiffusion systems [25].\nConcerning the importance of the above type of non-\nlinearly damped systems, we here focus on the PT-\nsymmetriccasesofthis category. The Hamiltonian struc-\nture [26] and quantization [27, 28] of some of the nonlin-2\nearly damped PT-symmetric systems with single degree\nof freedom have been studied recently, which show in-\nteresting symmetry breaking in these systems (see also\nSection III below).\nA proper coupling of two scalar nonlinearly damped\nPT-symmetric systems can yield novel bi- PT-symmetric\nsystems which are invariant with respect to the PT −1\n(x→ −x,y→ −y,t→ −t) operation as well as with\nthePT −2 operation which is defined as PT −2:x→\n−y,y→ −x,t→ −t. Such type of studies on the\nsystems with multiple PTsymmetries is interesting, for\nexample one can see an earlier paper on such multiple\nPTsymmetriccases[29]. Inthispaper,wepointoutthat\nthe study of bi- PTsymmetries in such coupled nonlinear\ndamped systems can lead to interesting novel dynamical\nstates of PTsymmetry preserving and breaking types,\nbesides oscillation death and bistable states.\nB. Spontaneous symmetry breaking: An interesting\nmechanism that is found to arise in the PTsymmetric\nsystems is the spontaneous symmetry breaking, where\nthe system in the symmetric state transits to an asym-\nmetric state by the variation of certain parameters. In\nclassical systems, the simplest state of broken symme-\ntry is the equilibrium state which may correspond to the\nminimum of the potential but which does not possess\nall the symmetries underlying the dynamical equation.\nLetGbe the transformation under which the dynamical\nequation is invariant. Then a symmetric state u=us\ncorresponds to the state which remains invariant under\nthe transformation us=Gus. But an asymmetric or\nsymmetry broken state ua(that may also correspond to\nthe minimum of the potential) is the one that gets trans-\nformed into another asymmetric state ui=Guaunder\nthe transformation G. Here the transformed state ui\nalso corresponds to an equilibrium of the system. A typ-\nical example is the reflection symmetry in a double well\nquartic anharmonic oscillator. From a dynamical point\nof view the spontaneous breaking of symmetries is also\nmanifested in the stability nature of the fixed points and\nthe trajectories around it in the phase space and nature\nof bifurcations as a system parameter is varied, again\nas in the case of the double well quartic oscillator un-\ndergoing spontaneous P-symmetry breaking. In this pa-\nper, we also show that the above existence of symmetry\npreserving/breaking equilibrium states can be identified\nwith the existence or nonexistence of the general solu-\ntion of the initial value problem underlying the dynam-\nical system satisfying the symmetry and the system can\nadmit more general classes of solution corresponding to\nsymmetry preservation/breaking.\nA universalfeature ofthe standard PT-symmetricsys-\ntems is that the PTsymmetry is broken by increas-\ning the loss/gain strength and is restored by reducing\nit [11, 12]. In contrast to this behavior, Liang et al. [30]\nhaveobservedareverse PTphasetransitionphenomenon\nin a lattice model known as PT-symmetric Aubry-Andre\nmodel[31], inwhichthe PTsymmetryisbrokenforlower\nloss/gain strength and is restored for higher loss/gainstrength. Theyobservedthisphenomenononlywhentwo\nlattice potentials that introduce loss/gain in the system\nare applied simultaneously (which is not observed when\na single lattice potential is present). This type of inverse\nPTphase transition arises as a result of the competition\nbetween the two lattice potentials. Similarly, Mirosh-\nnichenko et al. [32] have studied the competing effect of\nlinearand nonlinearloss-gainprofilein discretenonlinear\nSchr¨ odinger system. The observation of PTrestoration\nat higher loss-gain strengths also attracted wide inter-\nests and the recent studies show that it could happen\neven through an interplay of kinematical and dynamical\nnonlocalities [33].\nC. Nonlinear damping and PTsymmetry: From a dif-\nferent point of view, in the present work, we add a linear\ndamping in addition to the nonlinear damping and study\nthe competing effects of the linear and nonlinear damp-\ning forces. With a single nonlinear damping, our sys-\ntem shows PTsymmetry breakinglike the standard PT-\nsymmetric systems, but as soon we add the linear damp-\ning to the nonlinear damping, we observe PTrestora-\ntion at higher loss/gain strength similar to the case of\nAubry-Andre model. Importantly, we illustrate that this\ncompetition among the damping terms in addition to the\npositiondependent natureofdampingaidin tailoringthe\nPTregions of the system.\nThe organization of the paper is as follows, in section\nII, we discuss the loss-gain profiles of the scalar PT-\nsymmetric and non- PT-symmetric nonlinearly damped\nsystems. In section III, we consider a specific model\nof scalar PTsymmetric nonlinear damped oscillator,\nnamely the modified Emden equation. Analyzing the\ninitial value problem of an integrable case explicitly,\nwe greatly clarify the nature of PTsymmetry preser-\nvation/breaking. In section IV, we consider a coupled\nsystem with a simple nonlinear damping h(x,˙x) =x˙x,\nwhich is also a bi- PT-symmetric system. In section V,\nin addition to the nonlinear damping, we introduce a lin-\near damping in the system and show the occurrence of\nPTrestoration at higher values of loss/gain strength. In\nsection VI, we consider a general coupled system with\nlinear and nonlinear damping and show the tailoring of\nPTregions in the system. In section VII, we summa-\nrize the results of our work. In Appendix A, we consider\nthe initial value problem of a double-well oscillator and\ndiscuss the spontaneous P-symmetrybreakingfrom solu-\ntion point of view. In Appendix B we consider non- PT\nsymmetric scalar systems. In Appendices C, D and E,\nwe have presented the eigenvalues obtained through the\nlinear stability analysis for the systems we considered.\nII. NONLINEARLY DAMPED\nSYSTEMS-REVISITED\nTo start with, we analyze the loss-gain profiles of posi-\ntion dependent scalar nonlinearly damped systems. For\nthispurpose,wefirstconsiderasystemwhichisdescribed3\nby the second order nonlinear differential equation\n¨x+h(x,˙x)+g(x) = 0./parenleftbigg\n˙=d\ndt/parenrightbigg\n(1)\nHere,h(x,˙x) =f(x)˙xis the position dependent damp-\ning which we call for simplicity as the nonlinear damping\nterm. Also, f(x) is taken as a non-constant function in\nx. The above equation can be considered as a dynamical\nsystemonitsownmerit, oftenwithanonstandardHamil-\ntonian description [26], or as a conservative nonlinear os-\ncillator perturbed by a nonlinear damping force h(x,˙x)\nwhich supplies or absorbs energy at different points in\nthe (x,˙x) phase space,\n¨x+g(x) =−h(x,˙x) =−f(x)˙x. (2)\nThekinetic andthe potential energiesofthe unperturbed\nparticle are given respectively by\nT(˙x) =1\n2˙x2;V(x) =/integraldisplay\ng(x)dx. (3)\nThusthetotalenergyoftheparticleinthepotential V(x)\nwhenh(x,˙x) = 0 is\nE=1\n2˙x2+/integraldisplay\ng(x)dx. (4)\nThe rate of change of energy of the particle is\ndE\ndt= ˙x(¨x+g(x)). (5)\nFrom Eq. (1), we can write\ndE\ndt=−˙xh(x,˙x) =−f(x)˙x2. (6)\nIf the quantitydE\ndt<0 (or ˙xh(x,˙x)>0) in a region\nin (x,˙x) phase space, then the energy is withdrawn from\nthe system for the states lying in this region and the role\nofh(x,˙x) is like a damping or loss term and ifdE\ndt>0\n(or ˙xh(x,˙x)<0), then in the corresponding region the\neffect of h(x,˙x) is like negative-damping or gain.\nThe above type of nonlinearly damped systems can\nbe classified as ( i)PT-symmetric systems and ( ii) non-\nPT-symmetric systemsdepending on the form of h(x,˙x),\nwhereas all linearly damped systems are always non- PT-\nsymmetric. Here, the PT-symmetric systems are those\nsystems that are invariant under the combined operation\nofPT(and not individual operation of PorT):x→\n−x,t→ −t. We denote this as PT −1 symmetry (in\norder to distinguish it from the additional PTsymmetry\nin two dimensional systems). Then PT −1 symmetric\nsystems belonging to (1) are those systems where h(x,˙x)\nis a nonlinear function in x, ˙xthat is odd in xas well\nas ˙x. In this article, we focus our attention towards the\nsystems with h(x,˙x) =f(x)˙x, wheref(x) andg(x) in\n(1) are odd functions. Systems of the form (1) which do\nFIG. 1: (Color online) Loss-gain profilesdE\ndtof the systems\ngiven by (a) Eq. (7), (b) Eq. (8) and (c) Eq. (9) in the ( x,˙x)\nspace: The pink shaded regions in the figures correspond to\nthe regions in whichdE\ndtis positive (or it denotes the region\nin which gain is present). Similarly, the gray shaded region s\ndenote the regions in whichdE\ndtis negative.\nnotmeetthis requirementarenon- PT-symmetric. These\nnon-PT-symmetricsystemsaretypicallyoftwotypes, ( i)\nsystems exhibiting damped oscillations and ( ii) systems\nadmitting limit cycle oscillations. In the following we\npresent specific examples of these three cases:\n1.PT-symmetric conservativesystem - Modified Em-\nden Equation (MEE)[26, 34]:\n¨x+αx˙x+βx3+ω2\n0x= 0 (7)\n2. Non-PT-symmetric damped system[35]:\n¨x+αx2˙x+βx3+ω2\n0x= 0 (8)\n3. Limit cycle oscillator (van der Pol oscillator) [36]:\n¨x+(x2−1)˙x+ω2\n0x= 0. (9)\nThe system (7) is known as the modified Emden equa-\ntion and is obviously invariant under the PT −1 opera-\ntion. The PT-symmetric nature of this system [26] and\nits quantization [27] have been studied for the specific\ncaseβ=α2\n9which admits symmetry breaking states for\nλ <0. A critical analysis of the PT- symmetry of (7)\nis given in section III. The systems given in Eqs. (8)\nand (9) are examples of non- PT-symmetric ones, as the\ndamping term in these cases are found to be even func-\ntions ofx. The system (8) admits damped oscillations,\nwhile the system (9) (the famous van der Pol oscillator)\nis found to have self sustained oscillations which is also\nnoted in Appendix B.4\nFigure 1 shows the loss-gain profiles corresponding to\nEqs. (7), (8) and (9), which are obtained by substitut-\ning the corresponding forms of f(x) in Eq. (6) . From\nthe loss-gain profile (shown in Fig.1(a)) corresponding to\nthePT −1 symmetric case (7), we can find that we have\nvarying loss along the positive x−axis and varying gain\nalong the negative x−axis. The amount of gain present\nforx <0 is balanced by the amount of loss present for\nx >0. Then from Figs. 1(b) and 1(c), we can see that\nin the case of non- PT-symmetric systems, the loss and\ngain will not be balanced. In the case of the non- PT\ndamped oscillator (8), from Fig. 1(b) we can find that\nloss is present everywhere in space. In the case of limit\ncycle oscillator (9), from Fig. 1(c), we can find that gain\nexists in the region |x|<1 and loss exists in the region\n|x|>1. This clearly shows that in this case, the amount\nof loss present in the ( x,˙x) space is not balanced by an\nequal amount of gain.\ntx\n30 15 00.7\n0\n-0.7\ntdE\ndt\n30 15 00.8\n0\n-0.8\ntx\n30 15 00.6\n0\n-0.6\ntdE\ndt\n30 15 00.07\n0\n-0.07\n-0.14\ntx\n30 15 02.5\n0\n-2.5\ntdE\ndt\n30 15 06.5\n0\n-6.5(a) (b)\n(c) (d)\n(e) (f)\nFIG. 2: (Color online) Figures ( a), (c) and (e) depict the\nsolution of Eqs. (7), (8) and (9), respectively, for two dif-\nferent initial conditions. Figures ( b), (d) and (f) show the\ncorresponding ratesdE\ndtas a function of time.\nFrom Fig. 2, we can see that in the PT-symmetric and\nlimit cycle oscillator cases, there exists periodic and selfsustained oscillations (Figs. 2(a), 2(e)), respectively, and\nin the non- PT-symmetric damped oscillator case (Fig.\n2(c)), we have damped oscillations. The corresponding\nrates of change of energydE\ndtprofiles are shown in Figs.\n2(b), 2(d) and 2(f), respectively.\nComparing the periodic oscillations (Figs. 2(a) and\n2(e)) corresponding to the PT-symmetric oscillator case\n(Eq. (7)) and the limit cycle oscillator case (Eq. (9)),\nwe can find that the PT-symmetric system takes up dif-\nferent paths for different initial conditions but the limit\ncycle oscillator for different initial conditions tends to a\nparticular path as time t→ ∞. The reason is that the\nbalanced loss-gain profile (shown in Fig. 1(a)) of the\nPT-symmetric system allows it to have multiple paths\nalong which netdE\ndtis zero. But in the case of limit cy-\ncle oscillator, Fig. 1(c) shows that the loss and gain are\nnot balanced in the ( x,˙x) space. Thus the paths along\nwhich totaldE\ndtis zero are limited in this case. Conse-\nquently the phase space of limit cycle oscillators contains\nisolated paths only.\nNow let us consider a system of coupled nonlinear\ndamped oscillators (for simplicity we consider a linear\ncoupling)\n¨x+h1(x,˙x)+h2(x,˙x)+g(x)+κy= 0,\n¨y+h1(y,˙y)−h2(y,˙y)+g(y)+κx= 0,(10)\nwhereh1(x,˙x) =f1(x)˙xandh2(x,˙x) =f2(x)˙xare the\ntwo position dependent nonlinear damping terms. Here,\nthe functions f1(x) andf2(x) are chosen to be odd and\neven functions in x, respectively, and also the function\ng(x) is chosen as odd. Consequently, the system becomes\nsymmetric with respect to the PT −2 operation (which\nis defined as PT −2:x→ −y,y→ −x,t→ −t). Now,\nby making f2(x) to be zero, the system is symmetric\nwithrespecttoboth PT −1andPT −2operations. (Here\nPT −1 corresponds to the operation x→ −x,y→ −y,\nt→ −t.) Thus the system is bi- PT-symmetric in this\ncase.\nSimilar to the scalar case, we can consider the above\nsystem as a system of two coupled oscillators\n¨x+g(x)+κy= 0,\n¨y+g(y)+κx= 0, (11)\nacted upon by additional external forces h1(x,˙x) and\nh2(x,˙x). The total energy of the system (in the absence\nof nonlinear damping) is given by\nE=1\n2˙x2+/integraldisplay\ng(x)dx+1\n2˙y2+/integraldisplay\ng(y)dy+κxy.(12)\nThe rate of change of energy in the system due to the\nweak nonlinear damping term as specified by Eq. (10) is\ngiven by\ndE\ndt=−˙x[h1(x,˙x)+h2(x,˙x)]−˙y[h1(y,˙y)−h2(y,˙y)].(13)\nThe above expression shows that similar to the scalar\ncase, thecoupledsystem(10)alsohaspositiondependent5\nloss-gain profile. Further, the question whether a non-\nstandard Hamiltonian description similar to the scalar\ncase (Section III) exists for (10) has not yet been an-\nswered in the literature as far as the knowledge of the\nauthors goes, though a class of such systems has recently\nbeen identified [37, 38].\nIII.PTSYMMETRY BREAKING IN THE\nMODIFIED EMDEN EQUATION\nThe system mentioned in Eq. (7), namely\n¨x+αx˙x+βx3+λx= 0, λ=ω2\n0,(14)\nis the simplest example for PT −1 symmetric system.\nThe reversible nature of the system has been studied and\nthis equation is used as a normal form for describing the\nsymmetry breaking bifurcation in certain reversible sys-\ntems which includes an externally injected class B laser\nsystem [39]. This x˙xtype damping has been found to ap-\npear in many chemically relevant kinetic equations [19].\nThe model is found to be useful in fluid mechanics where\nthe linearly forced isotropic turbulence [24] can be de-\nscribed in terms of a cubic Li´ enard equation which is of\nthe form similar to (14). This system is also found to\nappear in some important astrophysical phenomena and\nit occurs in the study of equilibrium configurations of a\nspherical cloud acting under the mutual attraction of its\nmolecules and is subject to the thermodynamic laws [40].\nEq. (14) is known to admit a nonstandard conservative\nHamiltonian description [34] and interesting dynamical\nproperties [35]. In particular, the specific choice β=α2\n9\nadmitsisochronousproperties[26](seebelow)andcanbe\neven quantized in momentum space, exhibiting PTsym-\nmetry and broken PTsymmetry as shown by Chithiika\nRuby et al [27] recently, see subsection IIIB below.\nA. Linear stability analysis\nLet us analyze the dynamical behavior of the system\n(14) qualitatively through a linear stability analysis. Eq.\n(14) can be rewritten as\n˙x=x1\n˙x1=−αxx1−βx3−λx. (15)\nThissystemhasatrivialequilibriumpoint E0: (x∗,x∗\n1) =\n(0,0) and a pair of non-trivial equilibrium points sym-\nmetrically positioned along x-axis about x= 0,E1,2:\n(±/radicalBig\n−λ\nβ,0) (which exist only if λ <0 orβ <0). In our\nfollowing analysis, we take β >0 and so E1,2exist only\nforλ <0. The Jacobian matrix corresponding to the\nsystem (15) is given by\nJ=/bracketleftbigg0 1\n−αx∗\n1−3βx∗2−ω2\n0−αx∗/bracketrightbigg\n(16)The eigenvalues of Jcorresponding to the equilibrium\npointE0areµ(0)\n1,2=±i√\nλ. Similarly, the eigenvalues of\nJcorresponding to E1andE2areµ(1)\n1,2=1\n2√β(−α√\n−λ±\n/radicalbig\n−λ(α2−8β)),µ(2)\n1,2=1\n2√β(α√\n−λ±/radicalbig\n−λ(α2−8β)).\nThe real part of the eigenvalues of Jassociated with\nthe above equilibrium points are given in Fig. 3. The fig-\nure shows that in the region λ >0, the equilibrium point\nE0alone exists and all the eigenvalues of E0are found to\nbe pure imaginary(or Re[µ] = 0). So in the region λ >0,\nperiodic oscillations exist in the system corresponding to\nwhichthephasetrajectoriesaroundtheequilibriumpoint\nE0preserve their structure under PToperation, where\nE0itself remains invariant: PT[E0]=E0. Thus PT-\nsymmetry is unbroken while λ >0. But by varying λ\ntoλ <0, a pair of equilibrium points ( E1andE2) with\nopposite stabilities arise, where E1is stable (as all the\neigenvalues have Re[µ]<0) while E2is unstable (as all\neigenvalues have Re[µ]>0). In this region E0becomes\na saddle (as one of the eigenvalues of E0hasRe[µ]>0\nand the other eigenvalue has Re[µ]<0 ). Under the\nPToperation, E1gets transformed to E2and vice-versa:\nPT[E1]=E2andPT[E2]=E1so that the PTsymmetry\ngets broken. Correspondingly the trajectories around E1\nget transformed to trajectories around E2and vice-versa\nunder the PToperation. Note that the above kind of\nbifurcations fall within the scope of Thom’s catastrophe\ntheory [41].\n−3 −1 1 3\nλ−3.0−1.50.01.53.0Re[µ]E0\nE1\nE2\nUnbroken PTregion BrokenPTregion\nE0(us),E1(s),E2(us) E0(ns)(s): stable\n(us): unstable\n(ns): neutrally stable\nFIG. 3: (Color online) Plot of the real part of the eigenvalue s\nofJassociated with the equilibrium point E0,E1,E2of the\nsystem (14) for the values of α= 2 and β= 1.\nTo appreciate these aspects more clearly, we plot the\nphase portraits of the system for the explicitly integrable\ncaseβ=α2\n9, obtained from the exact solutions of the\nsystem [26]. A qualitatively similar set of phase portraits\nresults for the general case β/negationslash=α2\n9, which can be drawn\nthrough a numerical analysis.6➤ ➤➤ ➤\n➤ ➤E0➤\n➤➤➤➤➤\n➤\n➤➤\nx(t)˙x(t)\n1.5 0 -1.54\n2\n0\n-2\nFIG. 4: (Color online) Phase portrait of the system (14) for\nλ= 1,α= 3 and β= 1. The green colored diamond in the\nfigure denotes the position of the neutrally stable equilibr ium\npointE0.\n➤\nE0➤ ➤➤ ➤ ➤ ➤➤➤➤➤\n➤\nx(t)˙x(t)\n2 0 -23\n0\n-3\nFIG. 5: (Color online) Phase portrait of the system (14) at\nthe bifurcation point λ= 0 with α= 3 and β= 1.\nB. The exactly integrable case: β=α2\n9\nWe consider the specific case β=α2\n9of Eq. (14),\nnamely\n¨x+αx˙x+α2\n9x3+λx= 0 (17)\nor equivalently\n˙x=y,\n˙y=−αxy−α2\n9x3−λx. (18)\nEq. (17) or (18) admits a nonstandard Lagrangian/ con-\nservative Hamiltonian description [26] with\nL=27λ3\n2α2/parenleftBigg\n1\nα˙x+α2\n3x2+3λ/parenrightBigg\n+3λ\n2α˙x−9λ2\n2α2.(19)➤ ➤➤ ➤➤➤ ➤➤➤➤➤\n➤➤➤\n➤➤\n➤\n➤➤ ➤➤ ➤E1 E2E0\nx(t)˙x(t)\n2 0 -22\n0\n-2\nFIG. 6: (Color online) Phase portrait of the system (14) for\nλ=−1,α= 3 and β= 1. The green circle denotes the sta-\nble node type equilibrium point E1, red colored triangle and\nsquare correspond to the saddle type equilibrium point ( E0)\nand unstable node type of equilibrium point E2respectively.\nThe continuous line denotes the orbits corresponding to sym -\nmetric solutions and the dashed lines corresponds to that of\nasymmetric solutions.\nThen the canonically conjugate momentum is\np=−27λ3\n2α/parenleftBigg\n1\n(α˙x+α2\n3x2+3λ)2/parenrightBigg\n+3λ\n2α,(20)\nso that the Hamiltonian H\nH=9λ2\n2/parenleftBigg\n((˙x+α\n3x2)2+λx2)\n(α˙x+α2\n3x2+3λ)2/parenrightBigg\n=9λ2\n2α2/bracketleftBigg\n2−2/parenleftbigg\n1−2αp\n3λ/parenrightbigg1\n2\n+α2x2\n9λ−2αp\n3λ−2α3x2p\n27λ2/bracketrightBigg\n(21)\nwhich is a conserved quantity and we may call it as the\n’energy’E.\nNow the exact solution of (17) for the three cases\nλ >0,λ= 0 and λ <0 are as follows [26]:\n(i) Case-1: λ >0:Here one has periodic solutions of\n(17) or (18) as\nx(t) =Asin(ω0t+δ)\n1−Aα\n3ω0cos(ω0t+δ), ω0=√\nλ(22a)\n˙x(t) =Aω0cos(ω0t+δ)\n1−Aα\n3ω0cos(ω0t+δ)−α\n3x2(t),(22b)\nwhere,Aandδare constants. Note that the solution is\nperiodic and bounded for 0 ≤A <3ω0\nα. ForA≥3ω0\nα, the\nsolution is singular and periodic. Also one can evaluate\nfrom (21) using (22) the ’energy’ in this case as\nH=E=1\n2ω2\n0A2. (23)7\n(ii) Case-2: λ= 0:One has a decaying type or front\nlike solution in this case as\nx(t) =I1+t\nαt2\n6+I1αt\n3+I2, (24a)\nand\n˙x(t) =1\nαt2\n6+I1αt\n3+I2−α\n3x2(t),(24b)\nsuch that\nH=E= 0, (25)\nwhereI1andI2are arbitrary constants.\n(iii) Case-3: λ <0:Here we have the general solution\nx(t) =3/radicalbig\n|λ|(I1e√\n|λ|t−e−√\n|λ|t)\nα(I1I2+I1e√\n|λ|t+e−√\n|λ|t)(26a)\n˙x(t) =3|λ|(I1e√\n|λ|t+e−√\n|λ|t)\nα(I1I2+I1e√\n|λ|t+e−√\n|λ|t)−α\n3x2(t).(26b)\nwith\nH=E=18|λ|2\nα21\nI1I2\n2, (27)\nwhereI1andI2are arbitrary constants.\nNow treating the nonlinear differential equation (17)\nor (18) as a dynamical system, we shall consider the so-\nlution of its initial value problem (IVP) admitting the\nPTsymmetry. Since we require the PTsymmetry to be\nvalid for the entire duration of evolution, starting from\ntheinitialreferencetimewhichmaybetakenwithoutloss\nof generality as t= 0, we require the initial values of the\ndynamical variables corresponding to a definite ’energy’\nsatisfy the PT-symmetry conditions ( x(t)→ −x(−t),\nt→ −t, ˙x(t)→˙x(−t)):\nx(0) =c1=−x(0)\n˙x(0) =c2= ˙x(0) (28)\nwherec1andc2are arbitrary constants. Then one can\nidentify two possibilities.\n(i)PT-symmetric solution:\nc1= 0, c2=c (29)\nsuch that\nPT[x(t)] =−x(−t) =x(t),\nPT[˙x(t)] = ˙x(−t) = ˙x(t),for allt≥0,(30)\n(ii)PT-asymmetric solution:\nOne can consider two distinct values\nx1(0) =c1, x2(0) =−c1, c1/negationslash= 0 (31)such that for t >0, one can have a disjoint set of two\ndisconnected solutions/trajectories for a given E:\nPT[x1(t)] =−x1(−t) =x2(t)/negationslash=x1(t),\nPT[˙x1(t)] = ˙x1(−t) = ˙x2(t)/negationslash= ˙x1(t),(32)\nand\nPT[x2(t)] =−x2(−t) =x1(t)/negationslash=x2(t),\nPT[˙x2(t)] = ˙x2(−t) = ˙x1(t)/negationslash= ˙x2(t),for allt≥0.(33)\nassociated with the same energy value E. Sincex1(t) and\nx2(t) correspondto two distinct unconnected trajectories\nin phase space but with the same ’energy’ value, they\nrepresent solutions of broken PT- symmetry.\nWe now point out explicitly the abovetype ofsolutions\nfor the system (17) or (18) in the following, depending on\nthe sign of λ. We also demonstrate in Appendix A that a\nsimilar type of consideration exists for the P-symmetric\nsystem also, for example in the case of the double well\ncubic anharmonic oscillator.\nC. Observation of symmetry breaking from the\nsolution point of view\nCase-1: λ >0:PTinvariant solutions\nConsidering the general solution (22) for λ >0, with-\nout loss of generality we consider the solution of the ini-\ntial value problem with\nx(0) = 0,˙x(0) =B=3Aω2\n0\n3ω0−Aα(34)\nwhich is itself PTinvariant. This fixes δ= 0 in the solu-\ntion (22). Then the resultant general solution (22) with\nδ= 0oftheinitialvalueproblemisfully PT-invariantfor\nallt≥0, that satisfies (30). The ’energy’ associated with\nthe solution is again E=1\n2ω2\n0A2as given in (23). Note\nthat the above solution includes the equilibrium point\nE0= (0,0) when A= 0 with the energy Etaking the\nminimum value. The corresponding phase trajectories in\n(x,˙x) space are plotted in Fig. 4 which form concentric\nclosed curves around E0as long as A <3ω0\nα, so that it\nis a centre type equilibrium point. The associated eigen-\nvalues of the equilibrium point E0are±i√\nλ(as shown\nin Sec. IIIA above). Note that for A≥3ω0\nα, the solution\nbecomes singular at finite times giving rise to open tra-\njectories in the phase space Fig. 4 but which shall show\nPTsymmetry.\nOne can also observe that the phase trajectories are\ninvariant under time translation. Consequently, the so-\nlution corresponding to any other initial condition ob-\ntainable from (22) also follows an identical phase trajec-\ntory for a given Aand so a given value of ’energy’ Eas\nit is obtained by a time translation which is an allowed\nsymmetry of the original dynamical system (14). Hence\nthese solutions may not be treated as distinct from the8\none corresponding to (34), if time translation symmetry\nis also included, along with PTsymmetry. Due to the\nreason,nosymmetrybreakingasymmetricsolutionexists\nhere.\nCase-2: λ= 0- Bifurcation point\nHere again the solutions of the initial value problem\nwithx(0) = 0, ˙x(0) =1\nI2deduced from (24) correspond-\ning toE= 0, satisfy the PTsymmetry as shown with the\nphase trajectories in Fig. 5.\nCase-3: λ <0-PTsymmetry breaking\nIn this case one can identify three distinct classes of\nsolutionsfrom the generalsolution (26) of(17) or(18) for\nλ <0, namely x0(t),x1(t) andx2(t). Among them x0(t)\nforms the symmetric solution satisfying (30) and the set\nx(t) = (x1(t),x2(t)) satisfying(32)and (33) constitutes a\nspontaneously symmetry breaking set of solutions which\nare discussed below.\n(a) Symmetric solution:\nThe explicit form of the solution satisfying the initial\nconditions x0(0) = 0, ˙ x0(0) = constant turns out to be\nthe following:\nx0(t) =3/radicalbig\n|λ|(e√\n|λ|t−e−√\n|λ|t)\nα(I2+e√\n|λ|t+e−√\n|λ|t)\n˙x0(t) =3|λ|(e√\n|λ|t+e−√\n|λ|t)\nα(I2+e√\n|λ|t+e−√\n|λ|t)−α\n3x2\n0(t),(35)\nas can be deduced from the general solution (26). Here\nI2is an arbitrary constant. Note that the solution\n(35) satisfies the PTsymmetry PT(x0(t),˙x0(t)) =\n(x0(t),˙x0(t)) and that ( x0,˙x0) = (0,0) =E0in the\nlimitI2→ ∞. Also, we observe that asymptotically,\nast→ ∞, (x0(t),˙x0(t))−→\nt→∞(3√\n|λ|\nα,0) =E1. That is\nall the nonsingular trajectories approach the fixed point\nE1, exceptE0, so that E0is a saddle.\n(b) Asymmetric solution:\nNext we have the other two distinct solutions which\nbreak the PTsymmetry. The first one is given by\nx1(t) =3/radicalbig\n|λ|(I1e√\n|λ|t−e−√\n|λ|t)\nα(−2+I1e√\n|λ|t+e−√\n|λ|t), I1<0,\n˙x1(t) =3|λ|(I1e√\n|λ|t+e−√\n|λ|t)\nα(−2+I1e√\n|λ|t+e−√\n|λ|t)−α\n3x2\n1(t).(36)\nNote that ( x1(0),˙x1(0)) = (3√\n|λ|\nα,6|λ|\nα(I1−1)) and asymp-\ntotically ( x1(∞),˙x1(∞)) = (3√\n|λ|\nα,0) =E1. Also whenI1→ ∞, (x1(0),˙x1(0)) tends to E1. Again all the non-\nsingular trajectories approach E1asymptotically.\nSimilarly, we have the other distinct set of trajectories\nx2(t) =3/radicalbig\n|λ|(e√\n|λ|t−I1e−√\n|λ|t)\nα(−2+e√\n|λ|t+I1e−√\n|λ|t), I1<0,\n˙x2(t) =3|λ|(e√\n|λ|t+I1e−√\n|λ|t)\nα(−2+e√\n|λ|t+I1e−√\n|λ|t)−α\n3x2\n2(t).(37)\nNote that ( x2(0),˙x2(0)) = ( −3√\n|λ|\nα,6|λ|\nα1\nI1−1). In\nthe limit I1→ −∞ this approaches the equilibrium\npointE2= (−3√\n|λ|\nα,0). Interestingly, these trajec-\ntories (except E2) also approach E1asymptotically:\n(x2(∞),˙x2(∞))=(3√\n|λ|\nα,0). Notethatintheaboveeach\ndistinct trajectory corresponds to the invariant ’energy’\nE=18|λ|2\nα21\nI1I2\n2.\nThe above facts are illustrated by the corresponding\nphase trajectories for the case λ <0 in Fig. 6. In the\ncase 0< I1<∞(but not equal to 1), the evolution\ncorrespondingto x1(t)andx2(t)frominitialtime t= 0to\n∞lie along the same path, the trajectory corresponding\ntox1(t) is found to be a part of the trajectory of x2(t)\n(=PT[x1(t)]) (forI1>1) as well as that of x0(t) or the\ntrajectory corresponding to x2(t) is found to be a part\nof the trajectory of x1(t) (=PT[x2(t)]) for 0 < I1<1\nas well as that of x0(t). Under time translation these\ntrajectories may be mapped onto each other and so may\nbe considered equivalent to the symmetrical trajectories\nx0(t). These are not shown explicitly in Fig. 6.\nBut the most important fact is that for I1<0,x1(t)\nandx2(t) are truely asymmetric and so break the PT\nsymmetry. Consequently the solutions x1(t) andx2(t)\ngiveriseto distinct trajectories, depending on the choices\nof the arbitrary constants I1andI2.\nThus the above detailed analysis of the completely in-\ntegrable nonlinear damped system (17) or (18) estab-\nlishes the fact that a necessary and sufficient condition\nfor the preservation of PT −1 symmetry is the existence\nof a single fixed point which is of PT −1 invariant center\ntype (that is neturally stable fixed point associated with\nimaginary eigenvalues of the linearized equation). Note\nthat this requirement demands the existence of a single\nwell potential and rules out cases like three well potential\nforPT −1 symmetry preservation. Also the origin has\nto be necessarily the fixed point for PT−1 invariance,\nx→ −x,t→ −t. The above requirement allows the ex-\nistence of PT-symmetric non-isolated periodic solutions\naroundthe fixed point correspondingto concentricclosed\ncurves as trajectories as shown in Fig.4. Otherwise the\nPTsymmetry is broken as confirmed for the λ <0 case.\nThe above discussion also confirms that the existence of\nPTsymmetric fixed point and PT-symmetric solutions\nnear it alone does not imply PTsymmetry of the full\nsystem if the fixed point is not of centre type as seen in\nthe case of λ <0. Now we can use the above criteria as\nthe basis for PTinvariance for our further studies.9\nWe also note that the above results hold good for\nthe case of standard Hamiltonian type complex classi-\ncalPTsymmetric systems also, where one can find that\nthe symmetry implies x(t) =−x∗(−t) which implies\nRe[x(t)] =xR(t) =−xR(−t), Im[x(t)] =xI(t) =xI(−t),\nRe[p(t)] =pR(t) = ˙xR(t) = ˙xR(−t) =pR(−t) and\nIm[p(t)] =pI(t) = ˙xI(t) =−˙xI(−t) =−pI(−t). Thus\nin these cases the PTpreserving fixed point will be of\nthe form ( xR(t),xI(t),pR(t),pI(t)) = (0,c1,c2,0), where\nc1andc2are arbitrary constants. The studies on the\nclassical trajectories of complex PTsymmetric systems\nshow the existence of regular periodic orbits (possibly\nwith some unbounded orbits) in the unbroken PTre-\ngions and non-periodic or open and irregular trajectories\nin the case of broken PTregions [42–44]. In addition, in\n[42, 44] one can also note that the closed orbits are cen-\ntered around the PTpreserving fixed point as discussed\nabove which confirms our results.\nIV. A BI- PT-SYMMETRIC SYSTEM\nAs a simple case of the coupled nonlinear damped sys-\ntem (10), we consider a system of coupled modified Em-\nden equations (MEE)\n¨x+αx˙x+βx3+ω2\n0x+κy= 0,\n¨y+αy˙y+βy3+ω2\n0y+κx= 0. (38)\nHere,αis the nonlinear damping coefficient, κis the\ncoupling strength and ω0is the natural frequency of the\nsystem when ω2\n0>0. However, we will also consider the\ncaseω2\n0<0 corresponding to the double well potential.\nIt is obvious that the system (38) admits a bi- PTsym-\nmetry. ( i) It is invariant under the PT −1 symmetry:\nx→ −x,y→ −yandt→ −t. Eq. (38) is also invari-\nant under ( ii)PT −2 symmetry: x→ −y,y→ −xand\nt→ −t. Note that the above two symmetries also imply\nthe symmetry x(t)→y(t).\nEq (38) can be rewritten as\n˙x=x1,\n˙x1=−αxx1−βx3−ω2\n0x−κy,\n˙y=y1,\n˙y1=−αyy1−βy3−ω2\n0y−κx. (39)\nThe above set of dynamical equations (39) admit five\nsymmetrical equilibrium points, e0,e1,e2,e3ande4:\n(i) The trivial equilibrium point e0: (x∗,x∗\n1,y∗,y∗\n1) =\n(0,0,0,0).\n(ii) A symmetric pair of non-zero equilibrium points\ne1,2: (x∗,x∗\n1,y∗,y∗\n1)=(±a∗\n1,0,∓a∗\n1,0), where a∗\n1=/radicalBig\nκ−ω2\n0\nβ.\n(iii) Another pair of symmetric non-zero equilibrium\npointse3,4: (x∗,x∗\n1,y∗,y∗\n1) = (±a∗\n2,0,±a∗\n2,0), where\na∗\n2=/radicalBig\n−κ−ω2\n0\nβ.Besides the above five fixed points there exist four more\nasymmetric fixed points which turn out to be unstable\nin the parametric range of our interest. So we do not\nconsider them in this paper further.\nA. Case: Ω =ω2\n0>0\nIn analyzing (38), we first consider the case where\nΩ =ω2\n0>0. The existence of the above mentioned\nequilibrium points in different regions in the parametric\nspace for this case is indicated in Table I (for our further\nstudies we let β >0 in Eq. (38) or (39)).\nBefore entering into the classification of unbroken and\nbrokenPTregions of the system, we note here that the\nequilibrium points are also playing a key role in iden-\ntifying symmetry breaking as shown in the scalar case\nin the previous section. In this connection, we classify\nthePT −1 andPT −2 invaraint fixed points of the sys-\ntem (38), which can be identified by looking for the fixed\npoints which satisfy PT −k[ei]=ei, wherek= 1,2 and\ni= 0,1,2,3,4. Using this, one can find that the fixed\npointe0alone isPT −1 invariant (that is PT −1[e0]=e0),\nwhile the three fixed points e0,e1ande2arePT −2 in-\nvarantandthefixedpoints e3ande4areinvariantneither\nunderPT −1 symmetry nor under PT −2 symmetry.\nGeneralizing the discussion in the previous section, we\ncan identify the following two criteria on the fixed points\nof the coupled system of the type (38) or (39) for the\ninvariance of PT −1 andPT −2 symmetries:\n(i) For the preservation of PT −1 symmetry again one\nrequires the existence of a single fixed point at the\noriginwhichisofneutrallystabletype. Therequire-\nment that for PT −1 symmetry x→ −x,y→ −y\nt→ −tdemands the exclusion of any other fixed\npointand that the originwill be the solefixed point.\n(ii) For the preservation of PT −2 symmetry which de-\nmandsx→ −y,y→ −x,t→ −t, the criterion is\nthe existence of one or more fixed points which are\nallPT −2 invariant out of which atleast one should\nbe neutrally stable type. For example, in the above\nsystem (39) as well as (47) below besides the origin\ne0, the fixed points e1ande2are also PT −2 in-\nvaraint and it is sufficient that atleast one of them\nis neutrally stable for preservation of PT −2 sym-\nmetry (see Figs. 7 and 11 below). A specific case is\nillustrated in Fig. 8 below.\n1. Linear Stability Analysis\nNow, to explore the regions in which PTsymmetries\nare found to be broken and unbroken, we first deduce\nthe Jacobian matrix obtained from the linear stability10\nκ <−ω2\n0−ω2\n0≤κ≤ω2\n0κ > ω2\n0\nΩ =ω2\n0>0e0,e3,e4 e0 e0,e1,e2\nΩ =ω2\n0= 0 e0,e3,e4 e0 e0,e1,e2\nκ < ω2\n0ω2\n0≤κ≤ −ω2\n0κ >−ω2\n0\nΩ =ω2\n0<0e0,e3,e4e0,e1,e2e0,e1,e2\ne3,e4\nTABLE I: Symmetric equilibrium points of (39) in different\nregions of the ( κ,ω0) parametric space with β >0 and Ω =\nω2\n0>0,ω2\n0= 0 and ω2\n0<0.\nanalysis of the above system. It is given by\nJ=\n0 1 0 0\nc21−αx∗−κ0\n0 0 0 1\n−κ0c43−αy∗\n, (40)\nwherec21=−αx∗\n1−3βx∗2−ω2\n0,c43=−αy∗\n1−3βy∗2−ω2\n0\nand (x∗,x∗\n1,y∗,y∗\n1) are the equilibrium points of (39).\nThe eigenvalues of the above matrix determine the dy-\nnamical behavior of the system in the neighborhood of\nthe equilibrium points qualitatively and the results will\nbe helpful in identifying the broken and unbroken PT-\nsymmetric regions of the system. In the unbroken PT\nregion, the trajectories of the system, in addition to the\nevolution equation, replicate the full symmetry of the\nsystem, while in the symmetry broken region it does not.\nIn order that the trajectories of the system to be sym-\nmetric under PToperation, it should havea non-isolated\nperiodic nature (due to the presence of the time reversal\noperator TinthePToperator). Thus,welookforthere-\ngions of the system parameters for which the equilibrium\npoint isneutrally stable , that is the eigenvalues of the Ja-\ncobian matrix corresponding to the equilibrium point are\npure imaginary. These regions give rise to unbroken PT-\nsymmetric ranges. The eigenvalues of the linear stability\nmatrixJcorresponding to different equilibrium points of\nthe system are presented in the Appendix C, where the\nranges of linear stability are also discussed.\nFixing the parameters α,ω0,βasα= 1.0,ω0= 1.0\nandβ= 1.0, Fig. 7 shows the real parts of the eigenval-\nues of the equilibrium points e0,e1,2ande3,4(given in\nAppendix C, Eqs. (C1), (C2) and (C8)) under the varia-\ntionofκ. Whenevertherealpartsofalltheeigenvaluesof\nJ(Re[µ]) corresponding to an equilibrium point become\nzero, the eigenvalues are purely imaginary and the latter\nis said to be neutrally stable. On the other hand, when\nallRe[µ]’s corresponding to an equilibrium point are less\nthen zero, it is said to be stable, while the equilibrium\npoint is unstable in all the other cases. From the forms of\nthe fixed points and the nature of their stability proper-\nties, we can identify four separateregions R1,R2,R3and\nR4in the (κ,Re[µ]) plane, as follows: (i) R1:κ <−ω2\n0,\n(ii)R2:−ω2\n0< κ < ω2\n0, (iii)R3:ω2\n0< κ < cω2\n0, wherec−3 −1 1 3 5\nκ−2−1012Re[µ]e0\ne1,2e3\ne4\nR1 R2 R3 R4\nPT-1Broken\nPT-2BrokenPT-1Unbroken\nPT-2UnbrokenPT-1\nBroken\nPT-2\nunbrokenPT −1Broken\nPT −2Broken\ne0(us),e3(s)\ne4(us)e0(ns)e0(us)\ne1,2(ns)e0,1,2(us)(s)-stable\n(us)-unstable\n(ns)-neutrally stable\nFIG. 7: (Color online) Linear stability of equilibrium poin ts\nof (39) for Ω = ω2\n0>0 given in Table. I. Real parts of\neigenvalues of Jgiven by Eq. (40) are plotted as a function\nofκfor the parameters α= 1.0,β= 1.0 andω0= 1.0.\nis given in Eq. (C5) in Appendix C, (iv) R4:κ > cω2\n0.\nNote that ω2\n0= 1.0 in Fig.3. Then, using the criteria\ndiscussed above, we can identify the following facts, as\ndepicted in Fig. 7.\n•In the region R1(denoted in Fig. 7), where κ <\n−ω2\n0=−1.0, one can see that three branches ap-\npear fore0, and a single branch appears each for e3\nande4. Among the four eigenvalues of e0(see Eq.\n(C1)), two are found to be pure imaginary, while\nthe third one has a positive real part and the other\nhas a negative real part. Thus, in the region R1,\nthere are three branches corresponding to e0. In\neach of the cases of e3ande4, all the eigenvalues\nhave the same real parts (as seen from Eq. (C8)).\nThus,e3ande4havea single branch each in Fig. 7.\nFrom the values of Re[µ] in the region R1, we can\nfind that among the equilibrium points e0,e3and\ne4, onlye3is found to be stable. The stabilization\nofe3in the region gives rise to oscillation death .\nHere oscillation death in a system of coupled os-\ncillators denotes the stabilization of the system to\na non-trivial steady state due to the interaction of\noscillators in the system. We can also note that\nthe equilibrium points e3ande4get transformed\nto one another by both PT −1 andPT −2 opera-\ntions(that is PT −1[e3]=e4,PT −2[e3]=e4andvice\nversa) and the symmetry preserving equilibrium\nstatee0(that isPT −1[e0]=e0andPT −2[e0]=e0)\nis unstable. Thus both the PT −1 andPT −2\nsymmetries are broken in this region.11\n•In the region R2, where −ω2\n0≤κ≤ω2\n0(that is\nregion−1≤κ≤1), the equilibrium points e3and\ne4disappear, and e0alone exists. The eigenvalues\nof the equilibrium point e0in this region are found\nto be pure imaginary (see also Eq. (C1)). The\nneutral stability of the symmetric state e0signals\nthat in this region R2both the PT−1 andPT −2\nsymmetries are unbroken.\n•Forκ > ω2\n0= 1, in the region R3, (defined\nby Eq. (C5)), e0loses its stability and gives\nrise to two new equilibrium points e1ande2.\nThese new equilibrium points are found to\nbe neutrally stable. Further, they also get\ntransformed to each other by PT −1 operation:\nPT−1[e1]⇒ PT − 1[(a∗\n1,0,−a∗\n1,0)]=(−a∗\n1,0,a∗\n1,0)\n=e2and similarly PT −1[e2] =e1. How-\never, the equilibrium points show invari-\nance under PT −2 operation: PT −2[e1]⇒\nPT−2[(a∗\n1,0,−a∗\n1,0)]=(a∗\n1,0,−a∗\n1,0)=e1and\nsimilarly PT −2[e2] =e2. The invariance of the\nequilibrium points e1ande2withPT −2 operation\nis also illustrated in terms of the phase portraits\nin Fig. 8 obtained by numerical analysis of (39).\nAs the fixed point preserving PT −1 symmetry\n(e0) is not of neutrally stable type and due to the\ncoexistence of PT −1 violating fixed points e1and\ne2, thePT −1 symmetry is broken in the region.\nIn the case of PT −2 symmetry, all the fixed points\n(e0,e1ande2) preserve the symmetry and also two\nof them ( e1ande2) are neutrally stable. Thus the\nPT−2 symmetry is unbroken, as demonstrated in\nFig. 8.\n•For values of κin the region R4(beyond R3), all\nthe equilibrium points e0,e1ande2are found to\nbe unstable. Thus, both the PT −1 andPT −2\nsymmetries are found to be broken in the region.\n2. Dynamics in the (κ,α)parametric space\nNext, weextendourstudyasafunctionofthedamping\nparameter αalso. Fig. 9 shows the broken and unbro-\nkenPT-symmetric regions corresponding to system (38)\nin the (κ,α) parametric space. It shows that oscillation\ndeath appears in the region κ <−1 due to the stabi-\nlization of e3as seen earlier in Fig. 7 (as can be seen\nfrom Eq. (C8) in Appendix C). Looking at the region\n−1≤κ≤1 in Fig. 9, we can observe that the cou-\npled nonlinearly damped system (38) like the scalar case\n(7) (see Sec. III), does not show any symmetry break-\ning on increasing α(see Eq. (C1) in Appendix C). This\nis in contrast to the systems with linear damping which\nshow symmetry breaking when the loss/gain strength is\nincreased [11]. As mentioned in the previous subsection,\nin this region (that is the region R2seen in Fig. 7), both\nPT −1 andPT−2 symmetries are unbroken. Increasingx(t)˙x(t)\n0.9 0 -0.90.3\n0\n-0.3\ny(t)˙y(t)\n0.9 0 -0.90.3\n0\n-0.3PT-1 PT-1(a) (b)\nPT-2\nPT-2\nFIG. 8: (Color online) Illustration of broken PT −1 and un-\nbrokenPT −2 in region R3: (a) (x−˙x) (b) (y−˙y) projections\nshow the oscillations about the equilibrium points e1ande2\nin the region R3forκ= 1.5,α= 1.0,β= 1.0 andω2\n0= 1.0\nobtained by solving Eq. (39) numerically (The trajectories\naway from e1ande2are not shown here). The filled square\nand the circle represents the position of e1ande2, respec-\ntively. By PT −1 operation on e1we transit to e2and so\nPT −1 symmetry is broken. But, on the operation of PT −2\none1the equilibrium point remains unchanged thereby the\nsymmetry remains unbroken.\nOscillation \nDeathOscillations \nabout e0Oscillations \nabout e1,2PT− 1 Broken\nPT− 2 BrokenPT− 1 Unbroken\nPT− 2 UnbrokenPT− 1 Broken \nPT− 2 UnbrokenPT− 1, PT− 2 \nBroken\nBistable\n/Minus3 0 3012\nΚΑ\nFIG. 9: (Color online) Unbroken and broken PTregions in\nthe parametric space of ( κ,α) for Ω = ω2\n0>0 = 1.0 and\nβ= 1.0. Here the light-gray shaded region denotes the re-\ngion where both PTsymmetries are unbroken and the dark-\ngray shaded region denotes unbroken PT −2 symmetric re-\ngion. The dark gray shaded regions are denoted as bistable\nregions in the sense that the equilibrium points e1ande2are\nneutrally stable in that region. The light blue shaded regio ns\ncorrespond to the oscillation death regions.\nκfurther ( κ >1), the system shows breaking of PT −112\nsymmetry (for the values of κ >1 or in the region R3in\nFig. 7) through a pitchfork bifurcation. In this region,\nPT −2 symmetry alone is unbroken. Fig. 9 shows that\nthePT −2 symmetry is unbroken only if αis small (from\nEq. (C6) in Appendix C) and it is broken for increased\nα(Note that this type of symmetry breaking at higher\nvalues of loss/gain strength is a universal feature of all\nthePT-symmetric systems [11]). On further increasing\nκ, Fig. 9 shows that the PTregions with respect to α\nget reduced.\n3. Rotating wave approximation\nIn this section, we analyze the stability of the symmet-\nric orbits centered around e0in the region R2using the\nwell known rotating wave approximation. We consider\nperiodic solutions for the system in the region R2to be\nof the form\nx(t) =R1(t)eiωt+R∗\n1(t)e−iωt,\ny(t) =R2(t)eiωt+R∗\n2(t)e−iωt, (41)\nwhereω=ω0−∆ω, and ∆ωis a small deviation. Here,\nR1(t) andR2(t) are the slowly varying amplitudes with\nrespect to a slow time variable. Substituting (41) in (38),\nand by rotating wave approximation, we obtain\n˙R1=1\n2iω(−3β|R1|2R1+(ω2−ω2\n0)R1−κR2),(42)\n˙R2=1\n2iω(−3β|R2|2R2+(ω2−ω2\n0)R2−κR1).(43)\nNow, we separate the real and imaginary parts of\nthe equation as R1=a1+ib1,R2=a2+ib2. We\nhave steady periodic solutions when ˙ ai=˙bi= 0,\ni= 1,2. Thus, the equilibrium points of the\nsystem represent steady periodic solutions. The\nsystem has five symmetric equilibrium points rep-\nresenting symmetric orbits, which are E0:(0,0,0,0),\nE1,2:(0,±b∗\n11,0,∓b∗\n11),E3,4:(0,±b∗\n22,0,±b∗\n22), where\nb∗\n11=/radicalBig\nκ+ω2−ω2\n0\n3βandb∗\n22=/radicalBig\n−κ+ω2−ω2\n0\n3β. The system\nalso has asymmetric equilibrium points, which are E5,6:\n(0,±b∗\n33,0,±2κ\n6βb∗2\n44b∗\n33),E7,8: (0,±b∗\n44,0,±2κ\n6βb∗2\n33b∗\n44),\nwhereb∗\n33=/radicalbigg\n(ω2−ω2\n0)+√\n−4κ2+(ω2−ω2\n0)2\n6βand,b∗\n44\n=/radicalbigg\n(ω2−ω2\n0)−√\n−4κ2+(ω2−ω2\n0)2\n6β. Asω=ω0−∆ωand ∆ω\nis a small deviation, ω2−ω2\n0is also small. Thus, b∗\n22,b∗\n33\nandb∗\n44cannot be real and the equilibrium points E3,4,\nE5,6will not exist. So, we confine our attention to the\nequilibrium points E0,E1andE2.\nNow, in order to investigate the stability of the above\nperiodic solutions through a linear stability analysis, we\nobtain the eigenvalue equation as Aχj=λjχj, whereχj= [ξ1η1ξ2η2]Tand\nA=\n−3βa∗\n1b∗\n1\nωc110−κ\n2ω\nc123βa∗\n1b∗\n1\nωκ\n2ω0\n0−κ\n2ω−3βa∗\n2b∗\n2\nωc21\nκ\n2ω0c223βa∗\n2b∗\n2\nω\n.(44)\nHereci1=−3β(a∗\ni2+3b∗\ni2)\n2ω+ω2−ω2\n0\n2ω,ci2=3β(3a∗\ni2+b∗\ni2)\n2ω−\nω2−ω2\n0\n2ω,i= 1,2.λj,χj(j= 1,2,3,4) are the eigenvalues\nandeigenfunctionsofthe aboveeigenvalueequation. The\neigenvalues of Acorresponding to the equilibrium point\nE0: (0,0,0,0) are\nλj=±i(−κ+(ω2−ω2\n0))\n2ω,±i(κ+(ω2−ω2\n0))\n2ω.(45)\nThe eigenvalues of Acorresponding to E1andE2are\nλj=±/radicalbigg\n−2κ2−κ(ω2−ω2\n0)\nω,0,0. (46)\nThe eigenvalues of Acorresponding to E0are found to\nbe neutrally stable always, whereas two of the eigenval-\nues associated with E1andE2are pure imaginary when\n2κ2+κ(ω2−ω2\n0)>0. Whenalltheeigenvaluesof Acorre-\nsponding to an equilibrium point are pure imaginary, the\nneutral stability of the equilibrium point will make the\noscillation with frequency ωto be modulated by a slowly\nvarying periodic amplitude. It indicates that the system\nshows beats type oscillations. As the equilibrium point\nE0is always neutrally stable, we have stable beats type\nperiodic oscillations in the complete region R2. However,\nthe equilibrium points E1,2have two of their eigenvalues\nas zero, and so one needs to include higher order correc-\ntions to conclusively decide about their stability.\nB. Case: Ω =ω2\n0= 0\nIn this case, the existence of equilibrium points for dif-\nferent values of κis demonstrated in Table. I. The eigen-\nvalues of Jwith respect to e0(Eq. (C1)) clearly show\nthat it is always unstable. The equilibrium points e1and\ne2are found to be neutrally stable for κ >0 and for the\nvalues of αspecified in (C6). The equilibrium points e3\nore4stabilize for κ <0 and give rise to oscillation death.\nC. Case Ω =ω2\n0≤0:\nNext, wewishtoshowtheunbrokenandbroken PTre-\ngions corresponding to the system (38) with Ω = ω2\n0<0\nor the double well potential case. The equilibrium points\nat different values of κfor this case are also given in Ta-\nble. I. From the table, we can note that in contrastto the\nprevious cases, in the region −ω2\n0≤κ≤ω2\n0, the equilib-\nrium points e3,4coexist with e1,2. From the results of the\nlinear stability analysis of this case (where Ω = ω2\n0<0),13\nR0\nOscillations about e1,2PT− 1 Broken\nPT− 2 Unbroken\nOscillation DeathPT− 1 Broken\nPT− 2 BrokenPT− 1, PT− 2 Broken\ne3stable\n/Minus3 0 301\nΚΑ\nFIG. 10: (Color online) Phase diagram of (38) in ( κ,α) para-\nmetric space for Ω = ω2\n0<0. Figure is plotted for Ω = −1.0,\nβ= 1.0, which shows the regions in oscillations about e1and\ne2exists (dark gray shaded region) and the region where os-\ncillation death (light blue shaded region) occurs. One can\nclearly note from the figure that the PT −1 symmetry is bro-\nkeneverywhere. Intheregion denotedby R0(Regionoutlined\nby thick black line), we can find that there exists oscillatio ns\naboute1,2and oscillation death occurs about e3, thusPT −2\nsymmetry is broken in the region. The gray shaded region\nexcluding R0region gives rise to unbroken PT −2 region.\nwe can find that the equilibrium point e0(see Eq. (C1)\nAppendix C) completely loses its stability. Thus, when\nΩ<0, as in the scalar case, PT −1 symmetry is always\nbroken. The symmetric pair of equilibrium points e1,e2\nande3,e4are still found to be stable in some regions in\nthe (κ,α) parametric space. The region in which they\nare found to be neutrally stable or stable is given by Eqs.\n(C6) and (C8) and are shown by Fig.10. From the fig-\nure, we can observe that the PT −1 symmetry is broken\neverywhere in the parametric space.\nRegarding the PT −2 symmetry, Fig. 10 shows the re-\ngionin whichthe PT −2preservingfixed points e1ande2\nare neutrally stable (gray shaded region) and the region\nin which PT −2 violating fixed point e3is stable (Light\nblue shaded regions). All the regionsin which e3is stable\nobviously correspond to the broken PT −2 region. Inter-\nestingly, in this case, there exists a region denoted by\nR0in Fig. 10, in which the stable region of e1ande2\noverlaps with the oscillation death region (stable region\nof thePT−2 violating fixed point e3). Due to such co-\nexistence, PT −2 symmetry is broken in the region R0.\nThus the PT −2 symmetry is unbroken only in the gray\nshaded region excluding R0.V. NONLINEAR PLUS LINEAR DAMPING\nNext we wish to investigate the effect of the introduc-\ntion of a linear damping on the dynamics of the non-\nlinearly damped system (38). For this purpose, let us\nintroduce the linear damping terms in addition to the\nnonlinear damping introduced in Eq. (38). Now, the\nsystem takes the form\n¨x+γ˙x+αx˙x+βx3+ω2\n0x+κy= 0,\n¨y−γ˙y+αy˙y+βy3+ω2\n0y+κx= 0,(47)\nwhereγis the linear loss/gain strength. Obviously, the\nadded linear damping term in (47) breaks the PT −1\nsymmetry. Thus the system is only symmetric with re-\nspect to the PT −2 operation. Note that the equilibrium\npoints of this system are the same as that of (38). The\nstability determining Jacobian matrix in this case be-\ncomes\nJ=\n0 1 0 0\nc21−γ−αx∗−κ0\n0 0 0 1\n−κ0c43γ−αy∗\n, (48)\nwhere,c21=−αx∗\n1−3βx∗2−ω2\n0,c43=−αy∗\n1−3βy∗2−\nω2\n0. The eigenvalues of this Jacobian matix for different\nequilibrium points are given in Appendix D. For simplic-\nity, we take β= 1 for further studies. As in Sec. III, we\nlook forPTregions of (47) for the cases Ω = ω2\n0>0 and\nΩ =ω2\n0≤0 respectively.\nA. Case: Ω =ω2\n0>0\nTo begin, we look for the PTregions of the system\nwith respect to κfor the case Ω = ω2\n0>0. By fixing all\nthe other parameters of the system as α= 1.0,γ= 0.5,\nβ= 1.0 andω2\n0= 1.0 in (47), Fig. 11 shows the plot\nof the real part of eigenvalues of Jcorresponding to the\nequilibriumpoints e0,e1,e2,e3ande4asκisvaried. It is\ndivided into seven regions S1,S2, ...,S7along the κ-axis.\nFor the system (47), PT −2 symmetry alone exists and\nthePTregions correspond to the regions in which the\nPT −2 symmetry is unbroken. The details are as follows.\n•In the region S1ofFig. 11, where κ <−ω2\n0=−1.0,\nwe can find that among the equilibrium points e0,\ne3ande4, onlye3is found to be stable which leads\nto oscillation death. As mentioned in the previous\ncase, the PT −2 symmetry is broken in this region.\n•The region corresponding to the values of κbe-\ntween−ω2\n0< κ < ω2\n0(−1< κ <1) is now divided\ninto three regions, namely S2,S3andS4. In these\nregions as mentioned in Table. I, the equilibrium\npointe0alone exists.\n(a) In the region S2, where −ω2\n0< κ≤\n−/radicalBig\n4ω4\n0−(2ω2\n0−γ2)2\n4(that is−1≤κ≤ −0.484),14\n−2 0 2 4\nκ−1.50.01.5Re[µ]S1 S2 S3S4S5 S6 S7e0 e1 e2 Broken PTregions - S1,S3,S7\nUnbroken PTregions - S2,S4,S5,S6e3 e4\ne0,4(us)\ne3(s)e0\n(ns)e0\n(us)e0\n(ns)e0\n(us)\ne1,2\n(ns)e0,1(us)\ne2(ns)e0,1,2\n(us)(s)-stable\n(us)-unstable\n(ns)-neutrally stable\nFIG. 11: (Color online) Linear stability of equilibrium poi nts of (47) for Ω = ω2\n0>0 given in Table. I. Real parts of eigenvalues\nofJgiven by Eq. (48) are plotted as a function of κfor the parameters γ= 0.5,α= 1.0,β= 1.0 andω0= 1.0.\ntx(t)\n60 30 00.3\n0\n-0.3\nty(t)\n40 20 015\n0\n-15a) b)\nFIG. 12: (Color online) Broken PTsymmetry in the region\nS3: Figures ( a) and (b) are plotted for κ= 0.01,γ= 0.2,\nα= 1.0,ω0= 1.0 andβ= 1.0 that show the time series plots\nofxandy. The damped and growing oscillations of x(t) and\ny(t) indicate that for finite values of κ, thePT −2 symmetry\nis broken.\nwe can note that the equilibrium point e0is\nfound to be neutrally stable (which can also\nbe seen from Eq. (D2)) and gives rise to an\nunbroken PTregion.\n(b) In the region S3, whereκtakes smaller values,\n−/radicalBig\n4ω4\n0−(2ω2\n0−γ2)2\n4≤κ≤/radicalBig\n4ω4\n0−(2ω2\n0−γ2)2\n4\n(that is−0.484≤κ≤0.484), we can see that\nthe equilibrium point e0loses its stability (can\nbe seen also from Eq. (D2)) and the PT −2symmetry is broken now. As this PT −2 sym-\nmetry appears because of coupling (that is,\nthePT−2 symmetry disappears when κ= 0)\nit will not be preserved for smaller values of\nκ. Figs. 12(a) and 12(b) are plotted in the\nregion, which shows the damped oscillation in\nxand grow up oscillation ywhich shows the\nunbalanced energy between the xandyoscil-\nlators.\n(c) Now increasing κ, in the region S4, for/radicalBig\n4ω4\n0−(2ω2\n0−γ2)2\n4≤κ < ω2\n0,e0again becomes\nneutrallystableandgivesrisetounbroken PT\nregion.\n•Forκ > ω2\n0= 1, there exists three regions which\nare designatedas S5,S6andS7, identified from Eq.\n(D4). In these regions, the equilibrium points e0,\ne1ande2are found to exist (see Table-I).\n(a) In the region S5, (1.0< κ <1.28), the equi-\nlibrium point e0is found to be unstable, but\ne1ande2are found to be neutrally stable (can\nbe seen also from Eq. (D4)). As these equi-\nlibrium points traces itself upon PT −2 oper-\nation (that is PT −2[e1]=e1), thePT −2 sym-\nmetry in the region is said to be unbroken.\n(b) In the region S6, (1.28< κ <4.4), in addition\ntoe0,e1also loses its stability (can be seen15\nFIG. 13: (Color online) Broken and unbroken PTregions cor-\nresponding to the system (47) in the ( κ,γ) parametric space\nfor Ω = ω2\n0>0, which is plotted for α= 1.0,ω0=β= 1.0.\nThe light blue shaded region corresponds to the oscillation\ndeath region. The light and dark gray shaded regions denote\nthe unbroken PT −2 region. The dark gray shaded region\ncorresponds to the bistable region in the sense that the equi -\nlibrium points e1ande2are neutrally stable in the region.\nalso from Eq. (D4)). But e2is still neutrally\nstable, thus the region again corresponds to\nan unbroken PTregion.\n(c) On further increasing κ, forκ >4.4, in the\nregionS7, all the equilibrium points e0,e1and\ne2become unstable. Thus PTis broken for\nhigher values of κ.\nForα= 1.0,ω0= 1.0, andβ= 1.0, the broken and\nunbroken regions in the ( κ,γ) parametric space of (47)\nare indicated in Fig. 13. By comparing Fig. 13 with\nFig. 9, we can find the appearance of oscillation death\nfor the values κ <−1 as in the previous case (38). But\nin contrast to the previous case, the oscillation death\nregime disappears with an increase of γ. By increasing\nκ, the unbroken PTregion appears in the range −ω2\n0≤\nκ≤ω2\n0(whereω0= 1.0). In this region by increasing γ,\nthe system shows symmetry breaking (see Eqs. (D1) in\nAppendix D). But in the previous case (38), we cannot\nfind this type of behavior, where the PTsymmetry is\nnever broken by increasing the loss/gain strength α(see\nFig. 9 and Eq. (C1)).\nForκ >1 (the region in which e1ande2appear), Fig.\n13 indicates that when κis smaller than ≈2.2, thePT\nsymmetry of the system is preserved for lower values of\nγand it is broken for higher values of γ. Increasing κ\nbeyond≈2.2, thePTsymmetry of the system is brokenfor lowervalues of γ, and on increasing γthePTsymme-\ntry is restored or it becomes unbroken for the values of γ\nmentionedinEq. (D6). Onfurtherincreasing γ,thesym-\nmetry is again broken. Generally, in the standard type of\nPT-symmetric systems, PTis unbroken for lower values\nofγand broken for higher values of γ. Thus, this type\nofPTrestoration with the increase of loss/gain strength\nis unusual compared to the general PT-symmetric sys-\ntems, except for the case of Aubry- Andre model with\ntwo lattice potentials [30, 31]. As mentioned in the in-\ntroduction, the latter model is a lattice model in which\nthe lattice potential is applied in such a way that each\nelement of the lattice has different amount of loss and\ngain that makes the loss and gain present in the lattice\nto be position dependent. Then, the phenomenon of PT\nrestorationat higher values of loss/gainstrength appears\nonly when two such lattice potentials are applied simul-\ntaneously. The reason for this type of PTrestoration\nis the competition between the two applied potentials\nwhich introduces loss and gain in the system [30].\nSimilarly, in our case if a single damping is present in\nthe system (38), we cannot observe such PTrestoration\nat higher loss/gain strength (see Fig. 9). But when two\nor more types of damping present in the system, as in\nthe caseof(47) (where linearand nonlineardampingsare\npresent in the system) we can observe this type of PT\nrestoration(see Fig. 13). The above point will be further\ndiscussed in detail in the next section, where we will also\nshow that by properly choosing the form of nonlinear\ndamping, we can also tailor the PTregions of the system\nin the parametric space. Fig. 13 shows that there exists\nbistable regions for finite values of γand by increasing\nthe coupling strength κthe bistable region disappears.\nB. Rotating wave approximation\nNow, we look for the stability of the periodic orbits\naboute0in the region −ω2\n0≤κ≤ω2\n0. As we did in the\nprevious case (38), we find that the amplitude equations\nare\n˙R1=1\n2iω(−iγωR1−3β|R1|2R1+(ω2−ω2\n0)R1−κR2),\n˙R2=1\n2iω(iγωR2−3β|R2|2R2+(ω2−ω2\n0)R2−κR1).(49)\nNow, separating the real and imaginary parts of the\nequation as R1=a1+ib1,R2=a2+ib2, and from the\nlinear stability analysis of the above equation, we can\nfind that the system has an equilibrium point (0 ,0,0,0),\nwhose eigenvalues are\nλ=±1\n2ω/radicalBig\n(−(κ2−γ2ω2)−(ω2−ω2\n0)2±2√c1) (50)\nwherec1= (κ2−γ2ω2)(ω2−ω2\n0)2. The equilibrium\npoints are found to be neutrally stable for −/radicalbigκ\nω≤γ\n≤/radicalbigκ\nω. The linear stability discussed in the previous\nsection tells that the equilibrium point e0can become\nneutrally stable in the region given by Eq. (D3) (see16\nS0\nPT− 2 Broken PT− 2 Broken \nOscillation DeathPT − 2 Unbroken\n/Minus3 0 3012\nΚΓ\nFIG. 14: (Color online) Broken and unbroken PTregions cor-\nresponding to the system (47) in the ( κ,γ) parametric space\nfor Ω = ω2\n0<0, which is plotted for α= 1.0, Ω = −1.0,\nβ= 1.0. The light blue and light gray shaded regions corre-\nspond to oscillation death (stable region of e3) and neutrally\nstable region of e2respectively. In the region S0, the region of\nstable region of e2coexists with the stable region of e3, thus\nPTsymmetry is broken in the region. The unbroken region\ncorresponds to the gray shaded region excluding S0.\nAppendix D) and the above stability analysis of periodic\norbits in the region shows that the oscillations are found\nto be stable only for the values of γmentioned above.\nC. Case: Ω =ω2\n0≤0\nBy taking Ω = ω2\n0≤0, the equilibrium point e0loses\nits stability (see Eq. D1). The equilibrium points e1,2\nande3,4alone are found to be stable and the stable re-\ngions of these equilibrium points are given in Appendix\nD.Similartothepreviouscase, wehaveobservedaregion\ndenoted by S0in Fig. 14, in which a neutrally stable PT\npreserving fixed point ( e2) coexists with PTviolating\nfixed points. Thus this region S0corresponds to broken\nPTregion. The gray shaded region excluding S0alone\ncorresponds to the unbroken PTregion. As in the case\nwhereΩ >0, herealso PTrestorationathigherloss/gain\noccurs.\nVI. GENERAL CASE\nIn this section, we consider a more general coupled\nPT-symmetric cubic anharmonic oscillator system with\nnonlineardamping. Here, we take the nonlinear damping\ntermh(x,˙x) to be of the form f(x)˙xso that the equationof motion will take the form\n¨x+γ˙x+(−1)nαf(x)˙x+βx3+ω2\n0x+κy= 0,\n¨y−γ˙y+αf(y)˙y+βy3+ω2\n0y+κx= 0,(51)\nwheren= 0iff(x) isanoddfunction and n= 1iff(x)is\neven. Thus the system is PT-symmetric with respect to\nthePT−2 operation. The novel bi- PT-symmetric case\narises when f(x) is odd and γ= 0. For all forms of\nf(x), the equilibrium points are found to be the same as\nthat of (38). Now through the linear stability analysis let\nus find the unbroken and broken PT-symmetric regions.\nThe Jacobian matrix corresponding to (51) is\nJ=\n0 1 0 0\nc21−γ−(−1)nαf(x∗)−κ0\n0 0 0 1\n−κ 0 c43γ−αf(y∗)\n,(52)\nwherec21=−αf′(x∗)x∗\n1−3βx∗2−ω2\n0,c43=\n−αf′(y∗)y∗\n1−3βy∗2−ω2\n0. For simplicity, we consider\nthe case of ω0= 1,β= 1. The eigenvalues of this Ja-\ncobian matrix corresponding to odd and even f(x) cases\nof the system (51) about various equilibrium points are\ngiven in Appendix E.\nA. Case: f(x)is odd\nConsidering the case where f(x) is an odd function, in\nthe region −1≤κ≤1 (see Table I), in which the equilib-\nrium point e0alone exists, the corresponding eigenvalues\nofJare the same as in (D1). In this region, we can find\nthat the eigenvalues do not depend on αbut depends on\nγ(see Eq. (D1)). The region of unbroken PTsymmetry\nis confined to\n−/radicalBig\n2−2/radicalbig\n1−κ2≤γ≤/radicalBig\n2−2/radicalbig\n1−κ2.(53)\nFrom the above, it is clear that when γ= 0 the PT\nis always unbroken for all the values of αin the region\n−1≤κ≤1. This indicates that in a purely nonlinearly\ndamped system, we cannot observeany symmetry break-\ning while varying the nonlinear damping strength ( α) in\nthis region. By varying γ, we observesymmetry breaking\nfor higher values of |γ|>/radicalbig\n2−2√\n1−κ2.\nIn the region κ >1, where the non-trivial equilibrium\npointse1ande2come into action, we will show that by\nproperly choosing the nonlinear damping we can tailor\nthePTregions. In this regime, forthe casein which f(x)\nis an odd function, the unbroken PTregion lies within\nthe range of γspecified by (see Eq. (E3) in Appendix E)\n±αf(√\nκ−1)−√a1≤γ≤ ±αf(√\nκ−1)+√a1,(54)\nwherea1= (6κ−4)−4/radicalbig\n(2κ−1)(κ−1). The presence\nof the term αf(√κ−1) in the above equation is found\nto be important. Because considering the case where\nαf(√κ−1) = 0, the PTsymmetry is unbroken for lower17\nUnbroken PT region\nBistableBroken PT region\nPT revival\n0 5 1 0 1 5 2 00123\nΚ/DoubleGamma\nUnbroken PT region\nBistable/DoubleGammac\nΑcBroken PT region\n0 1 2 30123\nΑ/DoubleGammaBroken\nPT Region\nBistableUnbroken \nPT region\n0 1 2 3012\nΚ/DoubleGammaa) b) c)\nFIG. 15: (Color online) Phase diagram in ( κ,γ) space: ( a) and (b) denote broken (white), unbroken (gray) and bistable (dark\ngray) regions with f(x) =x3andf(x) =sinx, respectively, for α= 1.5. (c): Phase diagram in ( α,γ) space corresponding to\nf(x) =sinxandκ= 1.5\nvalues of γspecified by |γ|<√a1and is broken for the\nhigher values of γspecified by |γ|>√a1. But, in the\ncasewhere αf(√κ−1)/negationslash= 0, forthe valuesof γdefined by\n0<|γ|< αf(√κ−1)−√a1, thePTsymmetryis broken\nwhile it is unbroken for the values of γdefined by (54).\nThus, here the PTsymmetry breaking occurs at lower\nvalues of γand the restoration of symmetry occurs by\nincreasing γ. We can also note that the term αf(√κ−1)\ndepends on the form of f(x), which helps in tailoring PT\nregions of the system.\nIn Fig. 15, we have presented the PTregions of the\nsystem for the cases f(x) =x3andf(x) =sinx, which\nclearlyshowthat the PTregionscanbe tailoredwith the\nsystems of the type (51) by properly choosing the form\noff(x). From Fig. 15(b), we can note that by choosing\nf(x) to be a periodic one, we can observe PTrevivals.\nIn Fig. 15(c), we have shown the PTregions of the\nsystem in the ( γ,α) parametric space corresponding to\nthef(x) =sinxcase, while the figure looks qualitatively\nthe same for f(x) =x3. The figure indicates that in-\ncreasing γ(orα) beyond a critical value, denoted as γc\n(orαc), the unbroken PTregion appears only when α\n(orγ) is also sufficiently large.\nB. Case: f(x)is even\nThe case of even f(x) can again be divided into two\nsub-cases: (i) f(0) = 0and(ii) f(0) =anonzeroconstant\nsay, 1 (For the odd f(x) case,f(0) = 0 always and so\nthere are no sub-cases.)\nCase (i) f(0) = 0:Considering the case of f(x)\nwithf(0) = 0 (Example: f(x) =x2), in the region\n−1≤κ≤1 where the equilibrium point e0alone exists\n(see Table-I), the corresponding eigenvalues of J(given\nin (52)) are found to be the same as in (D1) and the\nunbroken PTregions of the system are also the same as\nthat of (53).Case (ii) f(0) = 1:In this case, for example f(x) =\ncosxore−x2, the eigenvalues of Jare different from case\n(i) and they aregiven in (E4). In contrastto the previous\ncases, the eigenvalues of Jcorresponding to e0are found\nto depend on α, see Eq. (E4), and the PTunbroken\nregion can be given in terms of γas\nα−√a2≤γ≤α+√a2. (55)\nwherea2= 2ω2\n0−/radicalbig\n4(1−κ2). This equation indicates\nthat thePTsymmetryis found to be brokenfor valuesof\nγoutside the range specified by (55) and PTsymmetry\nbecomes unbroken by choosing γwithin the range given\nin(55). Thusthe PTsymmetryisbrokenforlowervalues\nofγ,γ < α−√a2and restored at higher γ, as in Eq.\n(55).\nAsPTis broken for γ < α−√a2, forα >0 the\nPTregions preferentially exist for γ >0 and found to\nbe scarce for γ <0. In other words, the unbroken PT\nregions are abundant, if the loss due to the linear (or\nnonlinear) damping is introduced in the x-oscillator and\nthe loss due to the nonlinear (or linear) damping is intro-\nduced in the y- oscillator. When loss (or also gain) due\nto both the linear and nonlinear damping is introduced\nin the same oscillator, the unbroken PTregions become\nscarce.\nNow considering the region ( κ >1), where the non-\ntrivial equilibrium points exist (see Table-I), the dynam-\nics corresponding to the two sub-cases (case (i) and case\n(ii)) are the same. The eigenvalues of e1ande2are given\nin (E6), which become purely imaginary in the region\nαf(√\nκ−1)−√a1≤γ≤αf(√\nκ−1)+√a1,(56)\nwherea1= (6κ−4)−4/radicalbig\n(2κ−1)(κ−1). Comparing\nthe above with the one corresponding to the f(x) odd\ncase (see Eqs. (54) and (56)), we can find that in this\ncase the unbroken PTregions are scarce for γ <0. The\npresence of the term αf(√κ−1) indicates that the PT\nrestoration can occur at higher values of loss/gain, which\nconfirms that the PTregions can be tailored by a proper\nchoice of f(x).18\nUnbroken \nPT regionBroken PT \nregionBistable\n0 1 2 30123\nΚ/DoubleGamma\nBroken PT region\nUnbroken PT region\n0 5 1 0 1 50123\nΚ/DoubleGamma\nΓc\nΑcBroken PT region\nUnbroken PT region\n0 1 2 30123\nΑ/DoubleGammaa)\nUnbroken PT region\nBroken PT region\n0 0 .5 10123\nΚ/DoubleGammab) c)\nFIG. 16: (Color online) Phase diagram in ( κ,γ) space: ( a) and (b) denote broken (white), unbroken (gray) and bistable (dark\ngray) regions with f(x) =x2andf(x) =cosx(whereα= 1.5). The inset in Fig. (b) shows the PTregions corresponding to\nthe case f(x) =cosxfor values of κbetween 0 < κ <1 in (κ,γ) space. (c): Phase diagram in ( α,γ) space corresponding to\nf(x) =cosxandκ= 1.5\nFig. 16(a) shows the PTregions of the system (51)\nfor the choice of f(x) =x2which corresponds to the sub-\ncase (i)f(0) = 0. Fig. 16(b) is plotted for f(x) =cosx,\ncorrespondingtothe sub-case(ii), namely, f(0) = 1. The\ninset in the figure clearly shows that in this system even\nforκ <1 thePTrestoration at higher loss/gainstrength\noccurs. Figs. 16(a) and 16(b) clearly show that the PT\nregions can be tailored by the proper choice of f(x). Fig.\n16(c) show the PTregions in the ( γ,α) parametric space\nfor the choice f(x) = cosx, which shows the existence of\ncritical values γcandαcabove which the PTis unbroken\nfor higher loss/gain strength.\nVII. CONCLUSION\nIn this work, we have brought out the nature of the\nnovel bi-PTsymmetry of certain nonlinear systems with\nposition dependent loss-gain profiles. We have pointed\nout that the PT-symmetric cases of this type of non-\nlinear systems with position dependent loss-gain profile\noccur even with a single degree of freedom. These scalar\nnonlinear PT-symmetric systems are also found to show\nPTsymmetry breaking. We have demonstrated the na-\nture ofPT-symmetry preservation and breaking with an\ninteresting integrable example of damped nonlinear sys-\ntem. By coupling two such scalar PT-symmetric sys-\ntems in a proper way, we have shown the existence of the\nnovel bi- PT-symmetric systems in two dimensions. We\nhavealsoillustratedthe phenomenonofsymmetrybreak-\ning of the two PTsymmetries in this bi- PT-symmetric\nsystem. When this system is acted upon by a single\nnonlinear damping, we observed that for smaller cou-\npling strengths, the coupled system shows no symme-\ntry breaking while varying nonlinear loss/gain strength,\nwhereas the coupled PT-symmetric system with a lin-\near damping [11] shows symmetry breaking by increasing\nloss/gain strength. By strengthening the coupling, this\nnonlinearlydampedsystemshowssymmetrybreakingforhigher loss/gain strength. Then, by applying the linear\ndamping in addition to the nonlinear damping in a com-\npetingway,ourresultsshowthatasinthe PT-symmetric\nAubry-Andre model, PTrestoration at higher values of\nloss/gain strength occurs. The advantage of having po-\nsition dependent nonlinear damping with a competing\nlinear damping is to help to tailor the PTregions of the\nsystem according to the needs by properly designing the\nnonlinear loss and gain profile. We have also observed\nPTrevivals in the systems which have loss and gain pe-\nriodically in space.\nAcknowledgement\nSK thanks the Department of Science and Technology\n(DST), Government of India, for providing a INSPIRE\nFellowship. The work of VKC forms part of a research\nproject sponsored by INSA Young Scientist Project. The\nwork of MS forms part of a research project sponsored\nby Department of Science and Technology, Government\nof India. The work forms part of an IRHPA project of\nML, sponsored by the Department of Science Technology\n(DST), Government of India, who is also supported by a\nDAE Raja Ramanna Fellowship.\nAppendix A: Symmetry breaking in a P- symmetric\ncubic anharmonic oscillator\nHere, we demonstrate the P-symmetry breaking in a\ncubic anharmonic oscillator through the solution of its\nIVP. Let us consider the cubic oscillator equation\n¨x+λx+βx3= 0, λ=ω2\n0. (A1)\nFor simplicity, we consider β >0 for further discussions.\nThePsymmetry breakingin such a system is well known\nin the literature. For λ >0, this system has an equilib-\nrium point e0:(0,0) and the equilibrium point is found to19\nbe neutrally stable. As P[(0,0)]=(0,0), thePsymmetry\nin the region is unbroken. By decreasing λtoλ <0,\nthe equilibrium point e0loses its stability and gives birth\nto two new neutrally stable equilibrium points which are\ne1,2:(±/radicalBig\n−λ\nβ,0). In fact e0is a saddle and e1,2are cen-\ntre type equilibrium points. But these new equilibrium\npointse1ande2do not preservesymmetry as P(e1) =e2\nand vice versa. Thus Psymmetry is broken while λ <0.\nAll the stable equilibrium points correspond to minimum\nenergy values.\nThe system is an integrable one and its exact solu-\ntion is also available in the literature [45, 46]. Now, we\ndemonstrate the above Psymmetry breaking from the\nsolution of the IVP of the system.\nHere the general solution of the system is given as\nfollows\nCase-1:λ >0:\nx(t) =Acn[Ωt+δ,k] (A2)\nwhere Ω =/radicalbig\nω2\n0+βA2, the square of the modulus\nk2=βA2\n2(ω2\n0+βA2)andδis a constant. The associated\nenergy integral is E=H=1\n2˙x2+1\n2ω2\n0x2+1\n4βx4\n=1\n2ω2\n0A2+1\n4βA4. Then considering without loss of\ngenerality the IVP, x(0) =A, ˙x(0) = 0, in order that\nPx(0) =x(0),P˙x(0) = ˙x(0)⇒A=−Awhich is\npossible only if A= 0. Further since one requires\nP[x(t)] =−x(t)⇒x(t) =−x(t). From (A2), only the\npossibility A= 0⇒x(t) = 0, ˙x(t) = 0 for all t≥0 is\nthe admissible solution of the IVP which preserves P\nsymmetry. The corresponding energy E= 0 has the\nminimum value. The excited states of the system may\nsaidtobe Psymmetricifthetimetranslationisincluded.\nCase-2:λ <0:\nOn the other hand one finds the following general so-\nlutions for the case λ <0 in Eq. (A1)\n(i) 0≤A≤/radicalBig\n|λ|\nβ:\nIn this range only the trivial solution exists.\nx(t) = 0,˙x= 0 (A3)\n(ii)/radicalBig\n|λ|\nβ≤A≤/radicalBig\n2|λ|\nβ:\nIn this region we have the following two distinct peri-\nodic solutions in the two wells\nx(t) =±Adn(Ωt+δ,k) (A4)\n˙x(t) =∓AΩk2sn(Ωt+δ,k)cn(Ωt+δ,k) (A5)\nwhereΩ2=βA2\n2andk2=2(βA2−|λ|)\nβA2, andδis aconstant.\n(iii)A≥/radicalBig\n2|λ|\nβ:\nIn this region, one has the solution\nx(t) =Acn(Ωt+δ), (A6)\n˙x(t) =−AΩsn(Ωt+δ,k)dn(Ωt+δ,k),(A7)\nΩ =/radicalbig\n−|λ|+βA2, k2=βA2\n2(−|λ|+βA2).Considering the IVP x(0) =A, ˙x(0) = 0, one again finds\nx(t) = 0, ˙x(t) = 0 is the only possible P-symmetric\nsolution, existing when A ω2\n0. The eigenvalues of (40) corresponding\ntoe1ande2are the same and they are given by\nµ(1,2)\nj=±/radicalBigg\nb1±/radicalbig\nb2\n1−b2\n2;j= 1,2,3,4,(C2)\nwhere,\nb1=\nα/radicalBigg\nκ−ω2\n0\nβ\n2\n−(6κ−4ω2\n0),(C3)\nb2= 16(2κ−ω2\n0)(κ−ω2\n0). (C4)\nFor fixed values of αandβ, these eigenvalues are found\nto be pure imaginary for the values of κin the range\nω2\n0< κ≤(α2−6β)(α2−4β)−24β2−4αβ√2β\n((α2−6β)2−32β2)ω2\n0.(C5)21\nSimilarly, for a particular value of κin the range κ >\nω2\n0, the range of values of αfor which the eigenvalues will\nbe pure imaginary is given below,\n−/radicalBigg\nβ\nκ−ω2\n0b3≤α≤/radicalBigg\nβ\nκ−ω2\n0b3 (C6)\nwhere\nb3=/radicalBig\n6κ−4ω2\n0−√\nb2 (C7)\nwith the values of κ≥ω2\n0.\nThe eigenvalues of Jcorresponding to the equilibrium\npointe3(which exists when κ < ω2\n0) are\nµ(3)\n1,2=−α/radicalbig\n−(κ+ω2\n0)±/radicalbig\n(−α2+8β)(κ+ω2\n0)\n2√β\nµ(3)\n3,4=−α/radicalbig\n−(κ+ω2\n0)±/radicalbig\n−α2(κ+ω2\n0)+8β(2κ+ω2\n0)\n2√β(C8)\nWe can find from the above equation that these eigen-\nvalues can never be pure imaginary if α/negationslash= 0. The equi-\nlibrium point e3is found to be stable and gives rise to\noscillation death when α >0. The eigenvalues of Jcor-\nresponding to e4can be obtained by simply changing\nα→ −αin Eq. (C8). One can check that its eigenvalues\ncan never be pure imaginary for α/negationslash= 0 and that they can\nbecome stable when α <0.\nAppendix D: Eigenvalues of Eq. (47)\nIn this appendix, we present the eigenvalues of Jgiven\nin(48) fortheequilibriumpointsofthesystem(47). This\nsystem has the same set of equilibrium points as that of\n(38). The eigenvalues of Jfore0are\nµ(0)\nj=±/radicalBigg\n−(2ω2\n0−γ2)±/radicalbig\n(2ω2\n0−γ2)2+4(κ2−ω4\n0)\n2.(D1)\nFor the values of γin the range −/radicalbig\n2ω2\n0< γ κ≥/radicalbigg\n4ω4\n0−(2ω2\n0−γ2)2\n4. (D2)\nFor a particular values of κin the region −ω2\n0≤κ≤\nω2\n0,e0is neutrally stable for the values of γdefined by\n−/radicalbigg\n2ω2\n0−2/radicalBig\nω4\n0−κ2≤γ≤/radicalbigg\n2ω2\n0−2/radicalBig\nω4\n0−κ2.(D3)\nFrom the above relations, one can see that the increase\ninγbeyond this range causes symmetry breaking in the\nsystem (in the region −ω2\n0≤κ≤ω2\n0).Then, the eigenvalues of Jfor the equilibrium point e1\nare\nµ(1)\nj=±/radicalBigg\nb′1±/radicalbig\n(b′2\n1−b2)\n2(D4)\nwhere\nb′\n1=\nα/radicalBigg\nκ−ω2\n0\nβ+γ\n2\n−(6κ−4ω2\n0),(D5)\nandb2is given in (C4). The equilibrium point e1exists\nonly when κ > ω2\n0, and the associated eigenvalues are\npure imaginary when\n−α/radicalBigg\nκ−ω2\n0\nβ−b3≤γ≤ −α/radicalBigg\nκ−ω2\n0\nβ+b3(D6)\nwhereb3is given in (C7). Thus, the PTsymmetry is\nunbroken in the region given above. Similarly, the eigen-\nvalues of Jwith respect to e2and the regions in which\nthey take pure imaginary eigenvalues can be obtained by\nreplacing αbe−αin (D4) and (D6).\nThen, considering the equilibrium point e3(which exist\nforκ≤ −ω2\n0), its eigenvalues are the roots of the alge-\nbraic equation\nµ(3)4+2α/radicalBigg\n−(κ+ω2\n0)\nβµ(3)3+(−α2(κ+ω2\n0)\nβ−γ2\n−(6κ+4ω2\n0))µ(3)2+α/radicalBigg\n−(κ+ω2\n0)\nβ(6κ+4ω2\n0)µ(3)\n+4(2κ−ω2\n0)(κ−ω2\n0) = 0.(D7)\nAs the coefficients of µ(3)3andµ(3)are non-zero for\nα/negationslash= 0,β/negationslash= 0, the eigenvalues of the equilibrium point\ncannot take pure imaginary values. Similarly, the eigen-\nvalue equation corresponding to the equilibrium point e4\ncan be obtained by changing α→ −αin (D7).\nAppendix E: Eigenvalues of Eq. (51)\nNowweconsiderthe generalcaseofEq. (51), wherewe\ncan choose f(x) to be an odd or an even function. In this\nsection, depending on the nature of f(x) (odd or even),\nwe have presented their corresponding eigenvalues.\n1. Case: f(x)- odd\nIn this case, the eigenvalues of Jcorresponding to the\nequilibrium point e0are found to be the same as in (D1).\nThe eigenvalues about the equilibrium point e1ande2\nare\nµ(1,2)\nj=±/radicaltp/radicalvertex/radicalvertex/radicalbt˜b(1,2)\n1±/radicalBig\n((˜b(1,2)\n1)2−b2)\n2(E1)22\nwhere\n˜b(1)\n1=\n+αf\n/radicalBigg\nκ−ω2\n0\nβ\n+γ\n2\n−(6κ−4ω2\n0),\n˜b(2)\n1=\n−αf\n/radicalBigg\nκ−ω2\n0\nβ\n+γ\n2\n−(6κ−4ω2\n0),(E2)\nandb2is as given in (C4). The regions in which the\neigenvalues of e1ande2are found to be pure imaginary\nare given respectively by\n−αf\n/radicalBigg\nκ−ω2\n0\nβ\n−b3≤γ≤ −αf\n/radicalBigg\nκ−ω2\n0\nβ\n+b3and\n+αf\n/radicalBigg\nκ−ω2\n0\nβ\n−b3≤γ≤+αf\n/radicalBigg\nκ−ω2\n0\nβ\n+b3.(E3)\nwhereb3is given in (C7).\n2. Case: f(x)- even\nConsidering the case of even f(x), the eigenvalues of\ne0are\nµ(0)\nj=±/radicalBigg\n−c±/radicalbig\nc2−4(ω4\n0−κ2)\n2. (E4)wherec= (2ω2\n0−(γ−αf(0))2). The eigenvalues in (E4)\nare found to be same as that of (D1) when f(0) = 0. In\nthe case f(0) = 1, thus the eigenvalues given in (E4) are\ndifferent from that of (D1). In contrast to the previous\ncases(Eq. (C1) and(D1)), the eigenvaluescorresponding\ntoe0are found to depend on αand the region in which\nthe eigenvalues given in (E4) take pure imaginary values\nis\nα−/radicalbigg\n2ω2\n0−/radicalBig\n4(ω4\n0−κ2)≤γ≤α+/radicalbigg\n2ω2\n0−/radicalBig\n4(ω4\n0−κ2).(E5)\nThe eigenvalues corresponding to both e1ande2are\nfound to be the same and they are\nµ(1,2)\nj=−/radicaltp/radicalvertex/radicalvertex/radicalbt˜b(2)\n1±/radicalBig\n(˜b(2)\n1)2−b2\n2. (E6)\nThe eigenvalues corresponding to both e1ande2are\nfound to be pure imaginary only when\nαf\n/radicalBigg\nκ−ω2\n0\nβ\n−b3≤γ≤αf\n/radicalBigg\nκ−ω2\n0\nβ\n+b3.(E7)\n[1] K. G. Makris, R. El-Ganainy, D. N. Christodoulides, and\nZ. H. Musslimani, Phys. Rev. Lett. 100103904 (2008).\n[2] A. Guo et al, Phys. Rev. Lett. 103093902 (2009).\n[3] J. Sheng, M.A. Miri, D. N. Christodoulides, and M.Xiao,\nPhys. Rev. A 88041803(R) (2013).\n[4] C. Hang, G.Huang, and V.V. Konotop, Phys. Rev. Lett.\n110083604 (2013).\n[5] Y. N. Joglekar, D. Scott, M. Babbey, and A. Saxena,\nPhys. Rev.A 82030103(R) (2010).\n[6] Z. Lin, H. Ramezani, T. Eichelkraut, T. Kottos, H. Cao,\nand D. N. Christodoulides, Phys. Rev. Lett. 106213901\n(2011).\n[7] N. Lazarides and G. P. Tsironis, Phys. Rev. Lett. 110\n053901 (2013).\n[8] H. Jing, S. K. ¨Ozdemir, Xin-You L¨ u, J. Zhang, L. Yang,\nand F. Nori, Phys. Rev. Lett. 113053604 (2014).\n[9] X. W. Xu Y.Liu, C.P. Sun and Y. Li, arXiv:1402.7222\n(2014).\n[10] J. Schindler, A. Li, M. C. Zheng, F. M. Ellis, and T.\nKottos, Phys. Rev. A 84040101(R) (2011).\n[11] C. M. Bender, B. K. Berntson, D. Parker, and E. Samuel,\nAm. J. Phys. 81173 (2013).\n[12] C. M. Bender, M. Gianfreda, S. K. ¨Ozdemir, B. Peng,\nand L. Yang, Phys. Rev. A 88062111 (2013).\n[13] C. M. Bender, M. Gianfreda, and S. P. Klevansky, Phys.\nRev. A 90022114 (2014).\n[14] J. Cuevas, P.G. Kevrekidis, A. Saxena, and A. Khare,\nPhys. Rev. A 88032108 (2013).[15] A. Eichler, J. Moser, J. Chaste, M. Zdrojek, I. Wilson-\nRae, and A. Bachtold, Nature Nanotechnology 6339\n(2011).\n[16] S. Zaitsev, O. Shtempluck, E. Buks, and O. Gottileb,\nNonlinear Dyn 67859 (2012).\n[17] B. Jeong, H.Cho, Min-FengYu,A.F.Vakakis,D.M.Mc-\nFarland, and L. A. Bergman, ACS Nano 78547 (2013).\n[18] A. A. Andronov, E. A. Leontovich, I. I. Gordon, and A.\nG. Maier, Qualitative theory of second-order dynamical\nsystems (Wiley, New York, 1973).\n[19] S. Ghosh and D.S. Ray, Eur. Phys. J. B 8765 (2014).\n[20] N. Glade, L. Forest, and J. Demongeot, C. R. Acad. Sci.\nParis, Ser. I 344253 (2007).\n[21] T. P. Dinh, J. Demongeot, P. Baconnier, and G.\nBenchetrit, J. Theor. Biol. 103113 (1983).\n[22] P. F. Rowat and A. I. Selverston, J. Neurophysiology 70\n1030 (1993).\n[23] R. Fitzhugh, Biophys. J. 1445 (1961).\n[24] Z. Ran, Adv. Appl, Fluid Mech. 541 (2009).\n[25] V. Ma˜ nosa, Chaos soliton Fract. 18241 (2003).\n[26] V. K. Chandrasekar, M. Senthilvelan, and M. Laksh-\nmanan, Phys. Rev. E 72066203 (2005).\n[27] V. C. Ruby, M. Senthilvelan, and M. Lakshmanan, J.\nPhys. A: Math. Theor 45382002 (2012).\n[28] B. Bagchi, S. Modak, P. K. Panigrahi, F. Ruzicka, M.\nZnojil, arXiv: 1505.07552 (2015).\n[29] A. Cavaglia, A. Fring and B. Bagchi, J. Phys. A: Math.\nTheor.44325201 (2011).23\n[30] C. H. Liang, D. D. Scott, and Y. N. Joglekar, Phys. Rev.\nA89030102(R) (2014).\n[31] C. Yuce, Phys. Lett. A 3782024 (2014).\n[32] A. E. Miroshnichenko, B. A. Malomed, and Y. S.\nKivshar, Phys. Rev. A 84012123 (2011).\n[33] M. Znojil, J.Phys.A 48195303 (2015).\n[34] V. K. Chandrasekar, M. Senthilvelan, and M. Laksh-\nmanan, J. Phys. A: Math. Theor. 404717 (2007).\n[35] V. K. Chandrasekar, M. Senthilvelan, and M. Laksh-\nmanan, J. Phys. A: Math. Gen. 374527 (2004).\n[36] S. H. Strogatz, Nonlinear Dynamics and Chaos , (Perseus\nBook Publishing, USA, 1994).\n[37] R.G.Pradeep, V.K.Chandrasekar, M. Senthilvelan, and\nM. Lakshmanan, J. Math. Phys. 50052901 (2009).\n[38] A. D. Devi, R. G. Pradeep, V. K. Chandrasekar, and M.\nLakshmanan, J. Nonlinear. Math. Phys, 2078 (2013).\n[39] A. Politi, G. L. Oppo, and R. Badii, Phys. Rev. A 334055 (1986).\n[40] J. M. Dixon and J. A. Tuszynski, Phys. Rev. A 414166\n(1990).\n[41] V. I. Arnold, V. S. Afrajmovich, Y. S. Il’yashenko, L. P.\nShil’nikov, Dynamical Systems V: Bifurcation Theory and\nCatastrophe theory , (Springer-Verlag, Berlin, 1994).\n[42] C.M. Bender, S. Boettcher, P.N. Meisinger, J. Math.\nPhys.402201 (1999).\n[43] A. Nanayakkara, J. Phys. A: Math. Gen. 374321 (2004).\n[44] A. Sinha, D. Dutta and P. Roy, Phys. Lett. A 375452\n(2011).\n[45] M. Lakshmanan, S. Rajasekar, Nonlinear Dynamics: In-\ntegrability, Chaos and Patterns , (Springer-Verlag, Berlin,\n2003).\n[46] P. M. Mathews and M. Lakshmanan, Ann. Phys. (N. Y.)\n79171 (1973)." }, { "title": "2002.12297v3.Dipole_polarizability_of_time_varying_particles.pdf", "content": "Dipole polarizability of time-varying particles\nM. S. Mirmoosa1;2,\u0003T. T. Koutserimpas1, G. A. Ptitcyn2, S. A. Tretyakov2,yand R. Fleury1y\n1Laboratory of Wave Engineering, Swiss Federal Institute of Technology in Lausanne (EPFL), CH-1015 Lausanne, Switzerland\n2Department of Electronics and Nanoengineering,\nAalto University, P.O. Box 15500, FI-00076 Aalto, Finland\nInvariance under time translation (or stationarity) is probably one of the most important assump-\ntions made when investigating electromagnetic phenomena. Breaking this assumption is expected\nto open up novel possibilities and result in exceeding conventional limitations. However, to explore\nthe \feld of time-varying electromagnetic structures, we primarily need to contemplate the funda-\nmental principles and concepts from a nonstationarity perspective. Here, we revisit one of those\nkey concepts: The polarizability of a small particle, assuming that its properties vary in time. We\ndescribe the creation of induced dipole moment by external \felds in a nonstationary, causal way,\nand introduce a complex-valued function, called temporal complex polarizability, for elucidating a\nnonstationary Hertzian dipole under time-harmonic illumination. This approach can be extended\nto any subwavelength particle exhibiting electric response. In addition, we also study the classical\nmodel of the polarizability of an oscillating electron using the equation of motion whose damping\ncoe\u000ecient and natural frequency are changing in time. Next, we theoretically derive the e\u000bective\npermittivity corresponding to time-varying media (comprising free or bound electrons, or dipolar\nmeta-atoms) and explicitly show the di\u000berences with the conventional macroscopic Drude-Lorentz\nmodel. This paper will hopefully pave the road towards better understanding of nonstationary\nscattering from small particles and homogenization of time-varying materials, metamaterials, and\nmetasurfaces.\nI. INTRODUCTION\nTemporal modulation [1] in electromagnetic systems\n(e.g., Refs. [2{11]) is an e\u000ecient technique to achieve\nexotic wave phenomena and intriguing functionalities.\nNonreciprocity and isolation [12{18], frequency con-\nversion and generation of higher-order frequency har-\nmonics [19{21], wavefront engineering [21{23], one-way\nbeam splitting [24], extreme accumulation of energy [25],\nparametric ampli\fcation [2, 3, 26, 27], and wideband\nimpedance matching [28] are some of those functionalities\nthat have been reported in the past. One possibility that\ntime modulation can provide is to instantaneously con-\ntrol the radiation from subwavelength particles [29, 30].\nThis is due to the fact that electric and magnetic dipole\nmoments, p(t) and m(t), induced in a particle under il-\nlumination, can be temporally engineered in a desired\nfashion. In general, both the geometry of the particle\nand the optical properties of the material from which the\nparticle is made can be properly modulated in time by\nan external force.\nFrom the stationary perspective, it is assumed that the\nparticle is static, and its characteristics do not vary in\ntime. As a consequence, the induced dipole moments are\nconventionally described in the frequency domain simply\nthrough the complex dyadic electric and magnetic polar-\nizabilities [31]:\np=\u000bee(!)\u0001E+\u000bem(!)\u0001H;\nm=\u000bme(!)\u0001E+\u000bmm(!)\u0001H:\n\u0003mohammad.mirmoosa@aalto.\f\nyThese authors jointly supervised this work.Here, EandHare the Fourier transforms of the ex-\nternal electric and magnetic \felds, respectively. How-\never, the above equations cannot be generally applied\nfor a dynamic particle, because the very de\fnition of\nfrequency-domain parameters is based on the assump-\ntion that the particle is stationary. In fact, concerning\na time-varying particle, we need to return to the time\ndomain, and, subsequently, revisit the description of the\ninstantaneously induced dipole moments in terms of the\ndyadic polarizabilities model. An appropriate description\nshould explicitly indicate the nonstationary characteris-\ntic of the problem, along with the linearity and memory\n(frequency dispersion). The importance of this study is\nnot limited only to the understanding and engineering\nof instantaneous radiation, but it is also important for\nthe proper characterization and realistic implementation\nof time-varying metamaterials and metasurfaces [32, 33],\nbecause they are formed by time-varying meta-atoms.\nTherefore, having a clear picture about the polarizabil-\nity of meta-atoms paves the road towards homogenization\nmodels [34, 35] taking into account nonstationarity, and\nits interplay with dispersion phenomena.\nIn this paper, we thoroughly scrutinize the concept of\npolarizability associated with a particle which is vary-\ning in time. For simplicity, we assume that the parti-\ncle has only isotropic electric response. We analytically\nstudy how to determine the polarizability of such a par-\nticle when it is located in free space. For this study,\nwe employ the Hertzian dipole model which is a conven-\ntional model to describe a stationary resonant particle\nwith electric response. Determining the polarizability,\nwe also explain the interaction of nonstationary dipoles\nwith waves in terms of the particle polarizability. Fur-\nthermore, we move forward and consider the particle asarXiv:2002.12297v3 [physics.app-ph] 5 Oct 20212\na constituent of a time-varying material. Regarding this\nscenario, we focus on the classical bound electron (as the\nparticle under study) and derive the corresponding po-\nlarizability by assuming a time-dependent damping coef-\n\fcient and natural frequency in the equation of motion.\nAccordingly, we obtain the nonstationary Drude-Lorentz\nmodel for an e\u000bective medium and show how fundamen-\ntally di\u000berent this new model is from the conventional\nmodel written for a stationary medium.\nThe paper { as a foundational step towards under-\nstanding of nonstationary scattering from small particles\nand time-varying (arti\fcial) media { is organized as fol-\nlows: In Section II, we give a fundamental description\nof polarization of an arbitrary time-varying dipolar par-\nticle as a response to the excitation \feld by using the\nconcept of electric polarizability. Since this description\n(initially inspired by the analysis and synthesis of linear\ntime-varying systems in communications and control en-\ngineering [36, 37]) is not well covered in the literature and\nis missing in the classical electrodynamics textbooks [38{\n40], it helps the reader to obtain a proper perspective and\nis essential for understanding of the other sections of the\npaper. In the same section, we additionally discuss causal\nmodels of e\u000bective material parameters of time-varying\nmedia, following the same principles as for a single time-\nvarying particle. Next, in Sections III and IV, under\nnonstationary conditions, we treat small particles and\nclassical electrons based on their corresponding polariz-\nabilities. Section IV is also devoted to e\u000bective material\nmodels of nonstationary media. Finally, Section V con-\ncludes the paper.\nII. BASIC CONCEPTS\nA. Polarizability kernel\nFor a linear and stationary subwavelength particle with\nelectric response, there is a temporally nonlocal connec-\ntion between the instantaneous electric dipole moment\np(t) and the exciting electric \feld E(t). This connection\nis described by a convolution integral as\np(t) =Z1\n0\u000b(\r)E(t\u0000\r)d\r; (1)\nwhere\u000b(\r) is a time-dependent function called electric\npolarizability kernel (here, we assume that the dipole\nand the \feld are parallel and there is no bianisotropy).\nThe above equation illustrates two notable characteris-\ntics. The \frst is that if the electric \feld is temporally\nshifted bytsh, the dipole moment will be also shifted by\nthe same time tshdue to the stationarity of the particle.\nIn other words,\np(t\u0000tsh) =Z1\n0\u000b(\r)E(t\u0000\r\u0000tsh)d\r: (2)\nThe second characteristic, associated with causality,\nstates that the instantaneous dipole moment at a certaintime depends on the \feld at that time and the evolution-\nary progress over past times.\nThe situation is very di\u000berent if the particle under\nstudy is changing in time. Causality is certainly a fun-\ndamental concept in nature which should be scrutinised\ncarefully. However, the \frst characteristic, having to do\nwith invariance with respect to translations in time, is\nnot true anymore. For interactions of nonstationary par-\nticles with \felds, a temporal shift of the electric \feld does\nnot result in an equivalent temporal shift of the induced\ndipole moment. We should use a more general linear and\ncausal relation between the induced dipole moment and\nthe exciting \feld, which we write as\np(t) =Z+1\n0\u000b(\r;t)E(t\u0000\r)d\r: (3)\nHere, in contrast to Eq. (1), the polarizability kernel \u000bis\nnot only a function of the delay time between the action\nand reaction ( \r), but it also depends on the observation\ntime (t). In other words, this formula means that at\nevery moment of time t, we deal with a di\u000berent particle\nand with a di\u000berent frequency dispersion rule (de\fned by\nthe integral kernel as a function of \r). As a consequence\nof that, the instantaneous value of the dipole moment\ndepends not only on the past and present values of the\nexciting \feld, but also on the whole history of evolution\nof the particle properties.\nBased on Eq. (3), let us discuss the physical mean-\ning of the polarizability of a nonstationary particle. If\nthe electric \feld is chosen to be the Dirac delta function\nE(t) =\u000e(t\u0000t0)u(uis a unit vector), the dipole moment\nequals\np(t) =\u000b(t\u0000t0;t)u: (4)\nIn other words, the polarizability \u000bis the impulse re-\nsponse of the dipole. As we see, in the nonstationary\nsituation, the impulse response depends, as in usual sta-\ntionary linear systems, on how much time has passed\nsince the pulse excitation was applied, but also on time\nexplicitly. This property clearly manifests the fact that\nthe particle responds di\u000berently at di\u000berent moments of\ntime.\nBeside using Eq. (3), it is sometimes convenient to ap-\nply an alternative integral form to describe the dipole\nmoment (for example, see Sections III and IV). Let us\nconsider the following independent variable: \u001c=t\u0000\r.\nBy changing variable in Eq. (3) and de\fning\nh(t;\u001c) =\u000b(\r;t)j\r=t\u0000\u001c; (5)\nwe can equivalently write\np(t) =Zt\n\u00001h(t;\u001c)E(\u001c)d\u001c: (6)\nIn this alternative representation of causal linear rela-\ntions, the\u001cvariable has the meaning of time moments\nin the past, and the chosen integration limits ensure that3\nthe induced dipole does not depend on the \feld values in\nthe future. Also, in this form, assuming delta-function\nexcitation E(\u001c) =\u000e(\u001c\u0000t0)u, we \fnd the impulse re-\nsponse in general form p(t) =h(t;t0)u. Notice that in\nthe stationary scenario, the function h(t;\u001c) dependents\nonly on the time di\u000berence between the observation time\nand a time moment in the past: h(t;\u001c) =h(t\u0000\u001c). Conse-\nquently, the integral expressed by Eq. (6) becomes simply\na convolution, and the dependency of the polarizability\nkernel on the observation time tvanishes, i.e., the polar-\nizability kernel is only written in terms of the variable\n\r:\u000b(\r;t) =h(t;t\u0000\r) =h(\r). Accordingly, we obtain\nEq. (1) which was explained in the beginning of this sec-\ntion.\nAt this point, where the functions \u000b(\r;t) andh(t;\u001c)\nand the corresponding relation between them have been\ndiscussed, it is better to build a general consensus that\nwe refer only to the function \u000b(\r;t) as the polarizability\nkernel. This agreement will help us to avoid any con-\nfusion throughout the paper. Therefore, when working\nwith Eq. (6) and the function h(t;\u001c), remember that we\nneed to do a simple algebraic manipulation and deter-\nmine\u000b(\r;t) in order to present the polarizability kernel.\nB. Temporal complex polarizability\n1. De\fnition\nLet us consider a time-harmonic excitation by a given\nelectric \feld E(t) = Reh\nE0exp(j!t)i\n. Here, E0denotes\nthe complex amplitude, !is the angular frequency, and\nRe means the real part of the expression inside the brack-\nets. The reason for choosing the time-harmonic exci-\ntation is the fact that we want to concentrate on un-\nderstanding the e\u000bects of time variations of the particle\nitself, and it is convenient to use the simplest possible\nexciting \felds. Since the particle is linear, response to\narbitrary excitation can be found using the Fourier ex-\npansion of the incident \feld. Therefore, it is logical to\ncreate a model for time-harmonic excitation.\nSince from the beginning we have assumed that the\n\feld and the dipole moment are parallel, the polarizabil-\nity kernel is a scalar value. Substituting the electric \feld\ninto Eq. (3), we \fnd that\np(t) = Reh\n\u000bT(!;t)\u0001E0exp(j!t)i\n; (7)\nwhere\n\u000bT(!;t) =Z+1\n0\u000b(\r;t) exp(\u0000j!\r)d\r: (8)\nThe instantaneous electric dipole moment is the real part\nof a complex-valued function which is multiplied by the\ncomplex amplitude of the time-harmonic electric \feld\nE0exp(j!t). This is in clear analogy with the conven-\ntional stationary case in which the instantaneous dipole\n(a))\n(b))\n(c))\n𝐩𝑡=0+∞𝛼𝛾,𝑡𝐄𝑡−𝛾𝑑𝑡)\n𝐩𝑡=Re𝛼T𝜔,𝑡𝐄0exp𝑗𝜔𝑡 )\nഥ𝐩Ω=1\n2𝜋−∞+∞ത𝛼T𝜔,Ω−𝜔ത𝐄𝜔𝑑𝜔 )ഥ𝐩𝜔,Ω=𝐄0ത𝛼T𝜔,Ω−𝜔+𝐄0∗ത𝛼T∗𝜔,−Ω−𝜔/2\n𝐩𝑡=1\n2𝜋−∞+∞𝛼T𝜔,tത𝐄𝜔exp𝑗𝜔𝑡𝑑𝜔 )FIG. 1. Schematic view of wave scattering by a time-varying\nspherical particle. We assume that, for example, the optical\nproperties of the particle is changing in time. (a){The par-\nticle under time-harmonic illumination. We use the general\nconcept of polarizability kernel \u000b(\r;t) in order to study this\nscattering problem. (b){The particle under the same time-\nharmonic illumination. However, we employ the particular\nconcept of temporal complex polarizability \u000bT(!;t) and its\nFourier transform \u000bT(!;\n) for studying the corresponding\nproblem. (c){The particle under arbitrary signal illumina-\ntion. Here, temporal complex polarizability and its Fourier\ntransform are still valid to be used.\nmoment is the real part of the complex-valued, frequency-\ndomain electric polarizability multiplied by the electric\n\feld amplitude and the time-harmonic exponential fac-\ntor. However, here, the complex function \u000bTdepends\non the time variable t. Thus, we name such function as\n\\temporal complex polarizability\". The index \\T\", re-\nminding \\temporal\", distinguishes the function \u000bTfrom\nthe polarizability kernel \u000b.\nRecall that the above de\fnition is for the case of time-\nharmonic excitation (which is also indicated in Figure 1).\nAs one can realize in accordance with Eq. (7), by using\nthis de\fnition, the complexity of the problem dramati-\ncally reduces, and the induced dipole moment is simply\ndescribed based on the temporal complex polarizability\n\u000bT(!;t). The result of such simplicity is clear, for ex-\nample, in the next section { Section III A { where we\ndiscuss the interaction of the point electric dipole with\ntime-harmonic incident electric \felds. We will see how\nthe instantaneous power exerted on the dipole and the in-\nstantaneous power scattered from the dipole are elegantly\nwritten in terms of the temporal complex polarizability\n(without making any integration). Furthermore, the im-\nportance and advantage of the aforementioned simplicity\ncan be also understood in more complicated problems\nsuch as homogenization of time-varying arti\fcial media\n(metamaterials), in which the e\u000bective macroscopic pa-\nrameters should be derived, and, subsequently, the corre-\nsponding dispersion relations need to be calculated. Re-4\ngarding those problems, therefore, it is quite reasonable\nto employ the concept of the temporal complex polariz-\nability\u000bT(!;t) instead of the polarizability kernel \u000b(\r;t)\nin the time domain. We emphasize that due to the linear-\nity, the temporal complex polarizability can be used for\nnonharmonic excitations as well by applying the Fourier\nexpansion (see Figure 1).\n2. Properties\nLet us describe some features which are inferred from\nEqs. (7) and (8). By contemplating Eq. (8), we \frstly see\nthat the temporal complex polarizability is the Fourier\ntransform of the polarizability kernel with respect to the\ntemporal variable \r. Secondly, because the functions\np(t),E(t), and\u000b(\r;t) are real-valued, based on Eq. (8)\nand for real angular frequencies, we deduce that\n\u000b\u0003\nT(!;t) =\u000bT(\u0000!;t); (9)\nin which\u0003represents the complex conjugate. Further-\nmore, similar to the stationary scenario, in Eq. (8), the\nintegration is over the positive half-axis of \r, which re-\n\rects causality of the system and indicates that the tem-\nporal complex polarizability obeys Kramers-Kronig rela-\ntions [38].\nAbout Eq. (7), it explicitly con\frms the expectations\nthat the dipole moment induced by time-harmonic \felds\nis not necessarily time-harmonic. In Ref. [29], the authors\nhave recently discussed this fact without studying the po-\nlarizability. Importantly, the temporal variations of the\ndipole moment can be in principle fully engineered (while\nthe excitation \feld is time-harmonic) only by choosing\nthe proper temporal variation of the particle modula-\ntion. Beside this property, it is instructive to take the\nFourier transform of Eq. (7). For that, this equation can\nbe rewritten as\np(t) =1\n2\"\nE0\u000bT(!;t) exp(j!t) +E\u0003\n0\u000b\u0003\nT(!;t) exp(\u0000j!t)#\n:\n(10)\nNow, by de\fning the usual Fourier transform of\nan arbitrary temporal function g(t) asg(\n) =R1\n\u00001g(t) exp(\u0000j\nt)dtand applying this operation to\nEq. (10), we obtain\np(!;\n) =E0\u000bT(!;\n\u0000!) +E\u0003\n0\u000b\u0003\nT(!;\u0000\n\u0000!)\n2:(11)\nThe \frst argument of the Fourier transform of p(t) in-\ndicates that this expression is derived for excitation by\na time-harmonic electric \feld at frequency !. For exci-\ntations at other frequencies, the time dependence of the\ninduced dipole moment p(t) will be di\u000berent. The de-\npendence on two frequency variables is intriguing: The\n\frst frequency argument, !, corresponds to the Fourier\ntransform with respect to the variable \r, and the second\nangular frequency, \n, is due to the Fourier transform withrespect to the variable t. Notice that since the dipole mo-\nment p(t) is real-valued, we have p\u0003(!;\n) = p(!;\u0000\n).\nThis relation can be readily proved by using Eq. (11). It\nis worth noting that \u000bT(!;t) does not obey this relation,\nbecause\u000bT(!;t) is not necessarily a real-valued function.\nTherefore, in general,\n\u000b\u0003\nT(!;\n)6=\u000bT(!;\u0000\n): (12)\n3. Temporal complex susceptibility and permittivity\nBefore moving to the next section, it is worth noting\nthat in analogy with the dipole moments, a similar time-\ndomain description should be used for the electric and\nmagnetic \rux densities D(t) andB(t) as linear and causal\nfunctions of E(t) and H(t). If a medium is stationary, we\nreadily write [31]\nD=\u000f(!)\u0001E+\u0018(!)\u0001H;\nB=\u0010(!)\u0001E+\u0016(!)\u0001H;(13)\nin which\u000f,\u0016,\u0018,\u0010are the frequency-domain material pa-\nrameters. However, for a medium which is not stationary\nand its properties are time-variant, we need to express the\nconstitutive relations which concurrently respect nonsta-\ntionarity and memory. In the literature, assuming a time-\nvarying dielectric isotropic medium ( \u0016=\u0018=\u0010= 0), that\nconstitutive relation is often given by (e.g., Refs. [41{43])\nD(t) =\u000f(t)E(t): (14)\nThis model of a time-varying dielectric medium is based\non a very dramatic approximation of instantaneous re-\nsponse of matter, which is not consistent with the tem-\nporal dispersion naturally present in materials. There-\nfore, more complete and rigorous de\fnitions need to be\nintroduced and applied. Indeed, for any point in space\nwe should write (for an isotropic time-varying dielectric\nmedium)\nD(t) =\u000f0E(t) +Z+1\n0\u000f0\u001f(\r;t)E(t\u0000\r)d\r; (15)\nin which\u001f(\r;t) is the electric susceptibility kernel (which\nmay depend also on spatial coordinates). In general,\ntime-varying \felds can be expressed as an inverse Fourier\ntransform\nE(t) =1\n2\u0019Z+1\n\u00001E(!) exp(j!t)d!: (16)\nSubstituting the above expression in Eq. (15), we deduce\nthat\nD(t) =\u000f0\n2\u0019Z+1\n\u00001\u000fT(!;t)E(!) exp(j!t)d!; (17)\nwhere the temporal complex relative permittivity equals\n\u000fT(!;t) = 1 +\u001fT(!;t) (18)5\nwith\n\u001fT(!;t) =Z+1\n0\u001f(\r;t) exp(\u0000j!\r)d\r: (19)\nThis is similar to the de\fnition used by N. S. Stepanov\nin Ref. [44] for describing macroscopic susceptibility of\ntime-varying plasma. Here, the indices \\T\" are used\nto discern the temporal complex functions, \u000fT(!;t) and\n\u001fT(!;t), from the relative permittivity and susceptibil-\nity kernels, respectively (index \\T\" refers to \\temporal\").\nBased on the above derivations and explanations, what\nwe unequivocally perceive is the fact that the temporal\ncomplex susceptibility and relative permittivity are gen-\neral concepts which are valid and useful even if the \feld\nis nonharmonic. Later, in the last part of Section IV,\nwe employ these important equations and de\fnitions,\nEqs. (15){(19), to calculate the e\u000bective macroscopic pa-\nrameters of time-varying materials.\nIn the end of this discussion, we should mention that in\nanalogue to Eq. (11), we can take the Fourier transform of\nthe electric \rux density given by Eq. (17). This operation\nsigni\fcantly helps, allowing us to study and solve the\nMaxwell equations in the frequency domain (\n domain).\nKeeping in mind that \u000fT(!;\n) is the Fourier transform\nof\u000fT(!;t) with respect to the time variable t, we simply\nconclude that\nD(\n) =\u000f0\n2\u0019Z+1\n\u00001\u000fT(!;\n\u0000!)E(!)d!: (20)\nTo appraise this relation, we carefully look at two par-\nticular instances. The \frst one is if the medium is im-\nmutable and stationary, but dispersive. Thus, the rela-\ntive permittivity kernel depends on only one time vari-\nable and is given by \u000f(\r;t) =\u000f(\r). Consequently, the\ntemporal complex function and its Fourier transform are\nexpressed as \u000fT(!;t) =\u000f(!) and\u000fT(!;\n) = 2\u0019\u000f(!)\u000e(\n),\nrespectively, in which \u000e(\n) is the one-dimensional Dirac\ndelta function. By plugging this result in Eq. (20) and\nusing the property thatR\nf(x)\u000e(x\u0000x0)dx=f(x0), we\nobtain D(\n) =\u000f0\u000f(\n)E(\n) which is the conventional\nexpression for modeling dispersive time-invariant media.\nFor the second case, we assume that the medium pos-\nsesses instantaneous response and varies in time. As we\nwrote earlier, this is what a multitude of research works\nhave assumed in their studies of interactions of waves\nwith time-varying media. Regarding this case, the rela-\ntive permittivity kernel is \u000f(\r;t) =\u000e(\r)\u000fI(t) which results\nin\u000fT(!;t) =\u000fI(t) and\u000fT(!;\n) =\u000fI(\n). The initial ob-\nservation, based on Eq. (17), is that D(t) =\u000f0\u000fI(t)E(t),\nand the next observation, in accordance with Eq. (20),\nexplains the fact that D(\n) is the convolution of \u000fI(\n)\nand the \feld E(\n). Both these two observations are quite\nexpected for the nondispersive model of time-varying me-\ndia.III. INDIVIDUAL TIME-VARYING PARTICLE\nIN FREE SPACE\nBased on the fundamental notions introduced and dis-\ncussed above, we can address an important question on\nhow to determine the polarizability kernel \u000b(\r;t) and\nthe temporal complex polarizability \u000bT(!;t) for a single\ntime-varying particle located in free space. To answer\nthis central question, \frst, we need to write a linear di\u000ber-\nential equation which describes the polarization dynam-\nics of time-varying particles. Studying this equation and\nusing the fundamentals explained before, we will develop\na systematic method to \fnd the particle polarizability.\nHere, let us focus on the canonical example of a non-\nstationary Hertzian dipole. The understanding of this\nbasic scatterer can be extended to small time-varying\ninclusions which have electric response. We model a\nHertzian dipole as a short wire antenna of length l, as-\nsuming that the current is uniform along the wire. The\nantenna parameters are its e\u000bective inductance L, capaci-\ntanceC, and a resistive load R(that accounts for dissipa-\ntion losses). Parameters LandCmeasure the magnetic\nand electric energies stored in the reactive \felds around\nthe dipole. Accordingly, for such a Hertzian dipole, a\nthird-order di\u000berential equation for the oscillating charge\nQis expressed as\n\u0000Zd3Q\ndt3+Ld2Q\ndt2+RdQ\ndt+1\nCQ=lE: (21)\nThis is the \\R udenberg equation\" which was initially\nwritten in 1907 for elucidating the Hertzian dipole an-\ntenna in the receiving regime [45]. On the right side, E\ndenotes the incident electric \feld at the dipole location,\nparallel to the dipole. Notice that the \frst term in the\nabove equation on the left side, which is proportional to\nthe third time derivative of the electric charge, is associ-\nated with the radiation of the dipole and is linked with\nthe radiation reaction (see, e.g., Ref. [30]). The param-\neterZ=l2\u00160=(6\u0019c) determines what R udenberg calls\n\\Strahlungswiderstand\" or radiation resistance (here, \u00160\nis the free-space permeability, and cdenotes the speed\nof light). Including this \frst term in Eq. (21), we see\nthat the R udenberg equation is in fact analogous to the\nAbraham-Lorentz equation [46] that also includes the ra-\ndiation reaction term. Since the electric dipole moment\nis the multiplication of the electric charge and the dipole\nlength, therefore, we readily modify the authentic version\nof R udenberg equation and write\n\u0000l\u00160\n6\u0019cd3p\ndt3+L\nld2p\ndt2+R\nldp\ndt+1\nlCp=lE: (22)\nTill this point, we have assumed that the Hertzian dipole\nis stationary and the resistance, inductance, and the ca-\npacitance are immutable and constant in time. To re-\nalize a nonstationary Hertzian dipole, it is su\u000ecient to\nmake those parameters time-dependent in the di\u000beren-\ntial equation. For instance, suppose that we add a time-\nvarying capacitance as an extra load which is connected6\nin series with a dynamic load resistance R(t). For such\ntime-varying dipoles, the di\u000berential equation, Eq. (22),\nis rewritten as\n\u0000l2\nL\u00160\n6\u0019cd3p\ndt3+d2p\ndt2+R(t)\nLdp\ndt+1\nLC(t)p=l2\nLE:(23)\nNote thatC(t) is the total capacitance which contains\nboth the e\u000bective capacitance due to the stored electric\nenergy around the dipole and the time-modulated load\ncapacitance. Also, note that in Eq. (23), if the \frst term\n(the radiation term) on the left side is neglected (by as-\nsuming that it is much smaller than the last term (the\nrestoring-force term)), we achieve the di\u000berential equa-\ntion that is utterly similar to the equation of motion for\nbound electrons in matter, discussed in the next section\n(see Eq. (38)). Here, however, we study a dipole in free\nspace and keep the radiation term.\nEquation (23) relates the instantaneous dipole moment\nand the excitation \feld. On the other hand, in Sec-\ntion II, we related the instantaneous dipole moment and\nthe \feld in terms of the polarizability kernel. There-\nfore, by bringing these two models together, we can \fnd\nthe polarizability kernel (and the temporal complex po-\nlarizability) in terms of the time-varying resistance and\ncapacitance. We start from Eq. (6), which is an alterna-\ntive integral form to describe the dipole moment pof a\ntime-varying particle excited by an external electric \feld\nE:p(t) =Rt\n\u00001h(t;\u001c)E(\u001c)d\u001c. Next, we need to replace\nthis description in Eq. (23). For replacing, however, we\nhave to calculate the \frst, second, and third time deriva-\ntives of the dipole moment. To \fnd these derivatives, we\nuse the chain rule and the Leibniz integral rule which say\nthat for any integrable function f(x;y), we can write\nd\ndxZb(x)\na(x)f(x;y)dy=\nf(x;b(x))db(x)\ndx\u0000f(x;a(x))da(x)\ndx+Zb(x)\na(x)@\n@xf(x;y)dy;\n(24)\nwherea(x) andb(x) denote the lower and upper limits,\nrespectively. By employing Eq. (24) and doing careful al-\ngebraic manipulations, the \frst, second, and third deriva-\ntives of the dipole moment are written in terms of h(t;\u001c)\nand the corresponding partial derivatives of h(t;\u001c) as\ndp\ndt=h(t;\u001c)j\u001c=tE+Zt\n\u00001@h(t;\u001c)\n@tE(\u001c)d\u001c;\nd2p\ndt2=\"\n2@h(t;\u001c)\n@tj\u001c=t+@h(t;\u001c)\n@\u001cj\u001c=t#\nE\n+h(t;\u001c)j\u001c=tdE\ndt+Zt\n\u00001@2h(t;\u001c)\n@t2E(\u001c)d\u001c;(25)and\nd3p\ndt3=\"\n3@2h(t;\u001c)\n@t2+ 3@2h(t;\u001c)\n@t@\u001c+@2h(t;\u001c)\n@\u001c2#\n\u001c=tE+\n\"\n3@h(t;\u001c)\n@t+ 2@h(t;\u001c)\n@\u001c#\n\u001c=tdE\ndt+h(t;\u001c)j\u001c=td2E\ndt2+\nZt\n\u00001@3h(t;\u001c)\n@t3E(\u001c)d\u001c:\n(26)\nUltimately, by substituting the results of Eqs. (25) and\n(26) into Eq. (23), we derive four expressions (one char-\nacteristic equation and three initial conditions) which de-\n\fne the function h(t;\u001c). These four expressions read\n1)R@3h(t;\u001c)\n@t3+@2h(t;\u001c)\n@t2+R\nL@h(t;\u001c)\n@t+\n1\nLC(t)h(t;\u001c) = 0;\n2) 3R@2h(t;\u001c)\n@t2j\u001c=t+ 3R@2h(t;\u001c)\n@t@\u001cj\u001c=t+\nR@2h(t;\u001c)\n@\u001c2j\u001c=t+ 2@h(t;\u001c)\n@tj\u001c=t+@h(t;\u001c)\n@\u001cj\u001c=t=l2\nL;\n3) 3@h(t;\u001c)\n@tj\u001c=t+ 2@h(t;\u001c)\n@\u001cj\u001c=t= 0;\n4)h(t;\u001c)j\u001c=t= 0;\n(27)\nin which R=\u0000(l2=L)(\u00160=6\u0019c).\nAt this point, one may ask what will happen if the\ndipole is loaded with a time-varying inductance rather\nthan a time-varying capacitance. How do the above ex-\npressions change? This is a valid and intriguing question.\nIn fact, we should \frst revise the third-order di\u000beren-\ntial equation which describes the nonstationary dipole\nconnected to time-varying inductance. For that, as one\nmay expect, we start from the primary version of the\nR udenberg equation, Eq. (21), which is based on the\nelectromotive force, and we rewrite the voltage over the\ntime-varying inductance in this equation. Subsequently,\nafter a simple modi\fcation, we express the desired di\u000ber-\nential equation that is similar to Eq. (23). Therefore, we\nhave\n\u0000l2\nL(t)\u00160\n6\u0019cd3p\ndt3+d2p\ndt2+\u0000(t)dp\ndt+1\nL(t)Cp=l2\nL(t)E;(28)\nwhere \u0000(t) = [1=L(t)][R+dL(t)=dt] is a function that is\nrelated to the time derivative of the inductance. Conse-\nquently, this function becomes a constant if the deriva-\ntive of the inductance vanishes. Here, we do not continue\nwith the derivations of the polarizability kernel because\nthe method for deriving them is quite clear for the readers\n(Eqs. (25) and (26) should be substituted into Eq. (28)).\nTherefore, we pass those derivations and calculations on\nto the interested reader.7\nThe above theory, which we developed for determining\nthe polarizability kernel, will be complemented by deriv-\ning another partial di\u000berential equation that \\directly\"\ndescribes the temporal complex polarizability when the\nincident electric \feld is time-harmonic. Similar to what\nwe did for the polarizability kernel in Eq. (23), it is su\u000e-\ncient to supersede the dipole moment by its de\fnition in\nterms of\u000bT(!;t) (i.e., p(t) = Re\u0002\n\u000bT(!;t)\u0001E0exp(j!t)\u0003\n)\nin order to derive such a partial di\u000berential equation. Af-\nter doing simpli\fcations, we write that\nR@3\u000bT(!;t)\n@3t+\u0012\n1 +j3!R\u0013@2\u000bT(!;t)\n@2t+\n\u0012R\nL+j2!\u00003!2R\u0013@\u000bT(!;t)\n@t+\n\u00121\nLC(t)\u0000!2+j!R\nL\u0000j!3R\u0013\n\u000bT(!;t) =l2\nL:(29)\nNotice that here we assumed that the nonstationary\ndipole is loaded by a time-varying capacitance. How-\never, one can repeat the same procedure when the dipole\nload is a time-varying inductance (see Eq. (28)).\nA. Dipole interaction with incident waves\nThe analytical approach for determining the polariz-\nability kernel \u000b(\r;t) and the temporal complex polariz-\nability\u000bT(!;t) of nonstationary dipoles was explained in\nthe previous part of this section. At the following step, we\nscrutinize classical interactions of nonstationary dipoles\nwith external \felds based on the notion of polarizability,\nand study instantaneous powers exerted on and radiated\nby the dipole under illumination.\n1. Instantaneous power exerted on dipole\nLet us assume that a Hertzian dipole is located at\nthe origin of the Cartesian coordinate system, directed\nalong thezaxis, and illuminated by a time-harmonic\nelectric \feld. Also, let us assume that the nonstationary\ncharacteristic of the dipole can be realized, for example,\nby loading the dipole with a time-varying lumped ele-\nment [29] or varying the dipole length in time. For the\n\frst case, in which the e\u000bective length of the Hertzian\ndipole is \fxed and the dipole is loaded by a time-varying\nlumped element, we can simply employ the concept of in-\nduced electromotive force and calculate the total instan-\ntaneous power exerted on the dipole. From the basics,\nwe know that the electric dipole moment is the multipli-\ncation of the electric charge Q(t) and the dipole length\nl:p(t) =lQ(t). Since the time derivative of the electric\ncharge is identical with the electric current, as a result,\nthe time derivative of the dipole moment becomes equal\nto the length of the dipole lmultiplied by the electric\ncurrenti(t) carried by the dipole: dp(t)=dt=l\u0001i(t). We\nstress that in this simple calculation, the length of thedipole is not changing in time, and the temporal vari-\nation of the dipole moment is only due to the tempo-\nral variation in the electric charge. On the other hand,\nthe induced electromotive force corresponding to e\u000bective\nlengthlis expressed as v(t) =l\u0001E(t), in whichE(t) is the\ncomponent of the excitation \feld parallel to the Hertzian\ndipole. Using the above two equations, representing the\ninduced electromotive force and the time derivative of\nthe dipole moment, the instantaneous power is obtained\nfromS(t) =v(t)\u0001i(t), which \fnally gives\nS(t) =E(t)\u0001dp(t)\ndt: (30)\nRegarding the second case, where the length of the dipole\nis also changing in time, and, therefore, the temporal\nvariation of the dipole moment is not only due to the\nelectric charge Q(t), but it is also due to the length vari-\nationsl(t), it may initially seem to be complicated to \fnd\nthe instantaneous power. However, in general, the elec-\ntric current density corresponding to the Hertzian dipole\nis written as J(r;t) =dp(t)\ndt\u000e3(r), and the instantaneous\npowerS(t) is expressed as\nS(t) =Z\nVJ\u0001Ed3r; (31)\nwhereVdenotes the volume occupied by the dipole.\nBased on Eq. (31) and by substituting the electric cur-\nrent density, we explicitly achieve the same expression as\ngiven by Eq. (30). Therefore, we conclude that Eq. (30)\nis valid even for the case when the dipole length is chang-\ning in time. Notice that Eq. (30) is also true for any time\nvariation of the dipole moment, including the stationary\nscenario.\nAccording to Eq. (7), the dipole moment p(t) is de-\nscribed in terms of the temporal complex polarizability\n\u000bT(!;t). Therefore, this equation can be substituted into\nEq. (30) in order to present the power S(t) in terms on\nthe polarizability \u000bT(!;t). Before we proceed and do the\nsubstitution, let us note that due to the single-frequency\nexcitation, for brevity, we can drop the \frst argument of\n\u000bT(!;t) (i.e. the angular frequency of the incident \feld).\nIn addition, for simplicity, we also de\fne the following\ncomplex-valued function:\n\u0010(t) =\u000bT(t) +1\nj!d\u000bT(t)\ndt; (32)\nwhich is associated with the temporal complex polariz-\nability and its time derivative. Now, by substituting the\ntemporal complex polarizability and using this auxiliary\nfunction de\fnition, Eq. (32), the extracted power is sim-\npli\fed to\nS(t) =E(t)\u0001Reh\nj!\u0010(t)\u0001E0exp(j!t)i\n: (33)\nWriting the real part as Re[ x] = (1=2)(x+x\u0003), \fnally,\nthe extracted power reduces to\nS(t) =\u0000!\n2Im\u0002\n\u0010(t)\u0003\njE0j2\u0000!\n2Im\u0014\n\u0010(t)E2\n0exp(j2!t)\u0015\n;\n(34)8\nin which Im[ ] denotes the imaginary part.\nLet us check this equation for the special case of a\nstationary dipole. In this case the time derivative of\n\u000bT(t) vanishes, and \u0010becomes a complex constant which\nis equal to \u000bT. As a consequence, the time-averaged\nvalue of the second term in the above equation be-\ncomes zero and the time-averaged power extracted by\nthe dipole from the incident \feld is simply Sstationary =\n\u0000!\n2Im[\u0010]jE0j2=\u0000!\n2Im[\u000bT]jE0j2, which is the same rela-\ntion as we know from the literature (see e.g. Ref. [47]).\nHere, we stress that for stationary dipoles the expression\nin Eq. (34) is also valid in the time domain.\nAnother special case is the case when \u0010(t) = 0. From\nEq. (34) we see that if \u0010(t) = 0, the extracted power S(t)\nis zero meaning that the dipole does not interact with\nthe incident \feld. According to Eq. (32), the condition\n\u0010(t) = 0 corresponds to \u000bT(t) =Ae\u0000j!twhereAis a\nconstant coe\u000ecient. Substituting this result into Eq. (7),\nwe see that the dipole moment is constant over time. In\nother words, we have a static dipole moment whose time\nderivative is zero. Consequently, there should not be any\ninteraction with the incident \feld.\nConsidering Eq. (34), it is intriguing to assume a\nperiodic function \u0010(t). This is because periodicity al-\nlows us to employ simple time averaging. Based on the\nFourier series written for a periodic function, depending\non the period and the complex Fourier coe\u000ecients, the\ntime-averaged value associated with the second term in\nEq. (34) is not zero, and it can signi\fcantly contribute to\nthe time-averaged total power. For example, it is clearly\nseen that if the period is equal to the excitation period\nT= 2\u0019=!, the second-order term in the Fourier series\nn= 2 can produce a nonzero averaged value (in contrast\nwith the stationary case, in which the average is zero).\n2. Instantaneous power radiated by dipole\nSome part of the extracted power is re-radiated to the\nbackground medium (here, free space). The instanta-\nneous power which is re-radiated by the dipole is propor-\ntional to the \frst and the third time derivatives of the\ndipole moment [29, 30]:\nSrad(t) =\u0000\u00160\n6\u0019cdp(t)\ndt\u0001d3p(t)\ndt3: (35)\nSimilarly to what we did for S(t), we substitute the tem-\nporal complex polarizability in the above equation for\nthe re-radiated power. In accordance with Eqs. (7) and\n(32), the \frst and the third time derivatives of the dipole\nmoment are given by\ndp(t)\ndt= Re\"\nj!\u0010(t)\u0001E0exp(j!t)#\n;\nd3p(t)\ndt3= Re\"\nj!\u0010d2\u0010\ndt2+j2!d\u0010\ndt\u0000!2\u0010\u0011\n\u0001E0exp(j!t)#\n:\n(36)Subsequently, by using these equations and after doing\nsome algebraic manipulations, we \fnd the re-radiated\npower (in Eq. (35)) as\nSrad(t) =\u00160!4\n12\u0019cj\u0010j2jE0j2+\n\u00160!4\n12\u0019cjE0j2Re\"\n\u0010\u00121\n!2d2\u0010\ndt2+j2\n!d\u0010\ndt\u0000\u0010\u0011E2\n0\njE0j2exp(j2!t)\n\u0000\u0010\u00121\n!2d2\u0010\ndt2+j2\n!d\u0010\ndt\u0013\u0003#\n:\n(37)\nIf the dipole is stationary, all the time derivatives in the\nabove equation become zero, and the time-averaged scat-\ntered power is simpli\fed to Sstationary\nrad=\u00160!4\n12\u0019cj\u0010j2jE0j2=\n\u00160!4\n12\u0019cj\u000bTj2jE0j2, which can be conveniently found in the\nliterature (e.g. Ref. [47]).\nTo summarize, for a nonstationary electric dipole un-\nder illumination, we \frst derived the corresponding dif-\nferential equations (with the corresponding initial con-\nditions) for obtaining the polarizability. Afterwards, we\nutilized the introduced notion of the temporal complex\npolarizability to \fnd the total instantaneous extracted\npower and the instantaneous scattered power, which are\nnot necessarily equal to each other (see Appendix B for\nmore information).\nIV. NONSTATIONARY PARTICLE AS A\nCONSTITUENT OF A TIME-VARYING\nMATERIAL\nIn contrast to the previous section, where the time-\nvarying particle is located in free space, in this section, we\nassume that the particle is a constituent of a time-varying\nmaterial, and, accordingly, we investigate the polarizabil-\nity of such a particle. This investigation is important\nfor understanding the e\u000bective macroscopic parameters\nsuch as susceptibility and permittivity of dynamic ma-\nterials. To determine the polarizability kernel and the\ntemporal complex polarizability, we need to study the\ncorresponding di\u000berential equation. Since the particle is\nimmersed in a time-varying material composed of many\nidentical particles, the radiated power is compensated by\nthe power received from other particles. This cancel-\nlation ensures that in the absence of dissipation in the\nparticles and power exchange with the external devices\nthat modulate the particles, the e\u000bective medium param-\neters correspond to a lossless material. In this case, the\norder of the di\u000berential equation is two because the ra-\ndiation reaction term that was proportional to the third\nderivative of the dipole moment vanishes.9\nA. Di\u000berential equations for the dipole moment\nand the polarizability kernel\nLet us concentrate on the classical model of a bound\nelectron as our particle under study. In this case, the\ndi\u000berential equation needed for description of the elec-\ntric dipole moment is given by the classical equation of\nmotion. Based upon this equation, we model external\ntime modulations of the system by assuming that the\ndamping coe\u000ecient \u0000 Dand the natural frequency !nare\nvarying in time. Therefore, the second-order di\u000berential\nequation describing the electron motion is expressed as\nd2x(t)\ndt2+ \u0000D(t)dx(t)\ndt+!2\nn(t)x(t) =e\nmE(t), in which mde-\nnotes the electron mass, erepresents the electron charge,\nandx(t) is the displacement. Since the dipole moment\nis the multiplication of the electron charge and the dis-\nplacement, we can consequently write\nd2p\ndt2+ \u0000D(t)dp\ndt+!2\nn(t)p=e2\nmE: (38)\nHere, it is worth noting that one can entitle Eq. (38)\nas the Lorentz equation which results in the Lorentz\nmodel for the e\u000bective macroscopic parameters of dielec-\ntric materials. This model reduces to the Drude model\nin the limit of vanishing natural frequency ( !n= 0),\nwhich means that there is no restoring force and the\nelectron is not bound. Therefore, in general, we use the\nterm \\Drude-Lorentz\" to entitle Eq. (38). By employing\nEqs. (6) and (24), we already derived the \frst and sec-\nond time derivatives of the electric dipole moment and\ndemonstrated them based on the function h(t;\u001c) and its\npartial derivatives. Thus, by inserting those derivations\n(written in Eq. (25)) into Eq. (38) (the Drude-Lorentz\nequation), we arrive to three crucial expressions which\ndetermine the polarizability:\n1)@2h(t;\u001c)\n@t2+ \u0000D(t)@h(t;\u001c)\n@t+!2\nn(t)h(t;\u001c) = 0;\n2) 2@h(t;\u001c)\n@tj\u001c=t+@h(t;\u001c)\n@\u001cj\u001c=t=e2\nm;\n3)h(t;\u001c)j\u001c=t= 0:(39)\nWe stress that the function h(t;\u001c) is not the polariz-\nability kernel. Indeed, the polarizability kernel that we\nintroduced above is \u000b(\r;t) =h(t;\u001c) when\u001c=t\u0000\r. As\na consequence, \u001c=tin the second and third expressions\nrefers to\r= 0, and Eq. (39) de\fnes two initial condi-\ntions at\r= 0 for\u000b(\r;t). Depending on the temporal\nfunctions of the damping coe\u000ecient and the natural fre-\nquency, these three expressions in Eq. (39) give a speci\fc\nfunction for the polarizability.\nAs a check, let us examine the results by considering\n\frst the stationary scenario, assuming that the damping\ncoe\u000ecient and the natural frequency are not varying in\ntime. By remembering that h(t;\u001c) =h(t\u0000\u001c) in this\nscenario and solving Eq. (39), the polarizability kernel isderived as\n\u000b(\r;t) =h(t;t\u0000\r) =\ne2\nmq\n!2n\u0000\u00002\nD\n4exp\u0000\n\u0000\u0000D\n2\r\u0001\nsin\u0000r\n!2n\u0000\u00002\nD\n4\r\u0001\n:(40)\nNotice that the full derivation is explained in the Ap-\npendix of the paper. In the above equation, as it is ex-\nplicitly seen, the polarizability kernel depends on only\none time parameter, \r. Therefore, its Fourier transform\nis only a function of the frequency and gives the known\nDrude-Lorentz dispersion in the frequency domain (e.g.,\n[40]).\nRegarding the nonstationary scenario, in which the\ndamping coe\u000ecient or the natural frequency depends on\ntime, we will give a complete example in the last part\nof this section, comprehensively calculating the polar-\nizability kernel (see Eqs. (52){(57) and the correspond-\ning explanations). Also, subsequently, we will compare\nthe obtained results with the ones known for the con-\nventional stationary scenario. In this example, we will\nassume that the damping coe\u000ecient is temporally mod-\nulated as \u0000 D(t) = 2\u0000 0=(1 + \u0000 0t) (where \u0000 0corresponds\nto the damping coe\u000ecient at t= 0) and the natural fre-\nquency is zero.\n1. Noncausal interpretation\nPrior to studying the temporal complex polarizabil-\nity, here, we would like to have a brief discussion around\ncausality of time-modulated particles. If we do not re-\nspect causality, the dipole moment can also depend on\nthe electric \feld in the future, and, therefore, p(t) =R+1\n\u00001h(t;\u001c)E(\u001c)d\u001c. Such interpretation a\u000bects strikingly\nthe results of Leibniz integral expressions. In fact, by\napplying Eq. (24), the \frst and second derivatives of the\ndipole moment are modi\fed as\ndp\ndt=Z1\n\u00001@h(t;\u001c)\n@tE(\u001c)d\u001c;\nd2p\ndt2=Z1\n\u00001@2h(t;\u001c)\n@t2E(\u001c)d\u001c:(41)\nWe can compare these equations with the expressions\nfor the causal interpretation (Eq. (25)), in order to un-\nderstand the fundamental di\u000berence between them. By\nusing Eq. (41) and considering Eq. (38), which is the\nDrude-Lorentz equation, we write\nZ1\n\u00001\"\n@2h(t;\u001c)\n@t2+ \u0000D(t)@h(t;\u001c)\n@t+!2\nn(t)h(t;\u001c)#\nE(\u001c)d\u001c\n=e2\nmE(t):\n(42)10\nAs we explicitly see from the above, since the electric\n\feld is simultaneously attending both sides of the equal-\nity and it is inside the integral on the left side, we can con-\nclude that the whole expression within the square brack-\nets should be equal to the Dirac delta function. In other\nwords,\n@2h(t;\u001c)\n@t2+ \u0000D(t)@h(t;\u001c)\n@t+!2\nn(t)h(t;\u001c) =e2\nm\u000e(\u001c\u0000t):\n(43)\nEquation (43) is the key equation for calculating the po-\nlarizability kernel describing a noncausal response. Solv-\ning this equation and \fnding a solution for the function\nh(t;\u001c) may not be straightforward (and even possible)\ndue to the presence of the Dirac delta function on the\nright side. However, if there is a nonzero solution, notice\nthat it must be a real solution in time because the func-\ntionh(t;\u001c) is indeed a real-valued function. Here, we also\npoint out an intriguing issue which in\ruences Eq. (43).\nIn the expression p(t) =R+1\n\u00001h(t;\u001c)E(\u001c)d\u001cwritten ini-\ntially, although we chose the upper limit of the integral\nas in\fnity, we can de\fnitely assume a \fnite upper limit\nand still describe a noncausal response. The only condi-\ntion is that the \fnite upper limit should be larger than\nthe observation time t, i.e. upper limit = t+\fwhere\f\nis real and \f > 0. In accordance with this assumption,\np(t) =Rt+\f\n\u00001h(t;\u001c)E(\u001c)d\u001c, subsequently, we can simply\nrewrite Eq. (43) by using the Leibniz integral rule and\nDrude-Lorentz equation. In this paper, since our focus\nis only on the causal response, we leave such derivations\nfor the interested reader.\nB. Di\u000berential equation for \fnding the temporal\ncomplex polarizability\nWe have hitherto discussed how to derive the polariz-\nability kernel of the electron based on the classical model.\nHere, our aim is to introduce a linear di\u000berential equation\nwhose solution gives the temporal complex polarizability.\nWe already know that, for time-harmonic excitation, the\ninduced electric dipole moment is expressed in terms of\n\u000bT(!;t) (see Eq. (7)). Ergo, by substituting that expres-\nsion into the Drude-Lorentz equation (see Eq. (38)), we\ncome to the desired di\u000berential equation for the temporal\ncomplex polarizability, which is written as\n@2\u000bT(!;t)\n@t2+\u0010\n\u0000D(t) +j2!\u0011@\u000bT(!;t)\n@t+\n\u0010\n!2\nn(t)\u0000!2+j!\u0000D(t)\u0011\n\u000bT(!;t) =e2\nm:\n(44)\nEquation (44) is a second-order di\u000berential equation\nwith time-dependent coe\u000ecients which allows us to \fnd\n\u000bT(!;t) for arbitrary time variations of the particle pa-\nrameters. Certainly, this equation should be comple-\nmented by initial conditions in order to obtain a speci\fcsolution for \u000bT(!;t). These initial conditions are given\nby Eq. (39) in which h(t;\u001c) and the polarizability ker-\nnel are present. Therefore, after solving Eq. (44) and\n\fnding\u000bT(!;t), one needs to \frstly make the inverse\nFourier transform to calculate the polarizability kernel\n\u000b(\r;t) and subsequently h(t;\u001c). As a result, by having\nthe function h(t;\u001c), the initial conditions expressed in\nEq. (39) (the second and third expressions) can be read-\nily checked to achieve the speci\fc solution.\nIn order to check Eq. (44), we \frst make \u0000 D(t) and\n!n(t) time-invariant. Since \u000bT(!;t) does not depend on\ntime in this case, the corresponding time derivatives in\nEq. (44) become zero, and we instantly see that the so-\nlution is the usual Lorentz dispersion rule:\n\u000bT(!;t) =e2\nm\n!2n\u0000!2+j!\u0000D: (45)\nAs a second check, we consider a time-variant \u0000 D(t) or\n!n(t). For this, let us take the same example as in the\nnext subsection of the paper, where we will assume that\nthe damping coe\u000ecient is \u0000 D(t) = 2\u0000 0=(1 + \u0000 0t) and\nthe natural frequency is zero. For this example, we ap-\nply Eq. (39) and derive the polarizability kernel which\nis given by Eq. (57) in the next subsection. Therefore,\nsince we know \u000b(\r;t), we can obtain the temporal com-\nplex polarizability \u000bT(!;t) by simply taking the Fourier\ntransform with respect to \r, see Eq. (8). After some\nalgebraic manipulations, \u000bT(!;t) is expressed as\n\u000bT(!;t) =\u0000e2\nm\u00101\n!2+j2\n!3\u00000\n1 + \u0000 0t\u0011\n: (46)\nNow, as one can expect, this expression in Eq. (46) should\nde\fnitely satisfy the second-order di\u000berential equation\nEq. (44). If we substitute \u000bT(!;t) (written above) into\nEq. (44), we observe that Eq. (44) indeed holds. Notice\nthat Eq. (46) is not calculated based on Eq. (44), but\nit is obtained by employing the comprehensive Eq. (39).\nTherefore,\u000bT(!;t) in Eq. (46) represents the general so-\nlution of the second-order di\u000berential equation, which\nconsists of both the complementary function (the so-\nlution of the homogeneous equation) and the particu-\nlar integral solution (the solution of the inhomogeneous\nequation). This general solution also respects the ini-\ntial conditions expressed in Eq. (39). From this point\nof view, such general solution can be \fnally considered\nas the unique solution to Eq. (44) with !n(t) = 0 and\n\u0000D(t) = 2\u0000 0=(1 + \u0000 0t).\nC. On the Drude-Lorentz model of time-varying\ndielectrics and plasma\nNext, we use the above theoretical results to analyse\napproximate models of e\u000bective parameters of Lorentzian\ndielectrics and electron plasma. The dipole moment of\neach electron is governed by Eq. (38), where the param-\neters may depend on time due to changing environment11\nwhere the charges are located. However, apart from\nEq. (38), the electron density (i.e. the number of elec-\ntrons per unit volume N(t)) can also depend on time.\nIn consequence, in the following, we will consider two\ndi\u000berent cases: A particular case in which only the elec-\ntron density is time variant and the parameters in the\nequation of motion (Eq. (38)) do not change in time, and\na more general case in which those parameters are also\ntime dependent in addition to the electron density. Both\naforementioned cases de\fnitely result in nonstationary\nmodels. The reason for having a discussion on the for-\nmer case is that while the polarizability kernel of the\nsingle electron should be the same as the one for the sta-\ntionary scenario, because the damping coe\u000ecient and the\nnatural frequency are assumed to be time invariant, and,\ntherefore, the polarizability kernel is only a function of\n\r(i.e.\u000b(\r;t) =\u000b(\r)), we will soon show that the corre-\nsponding e\u000bective permittivity kernel becomes a function\nof both\randt(i.e.\u000f(\r;t)).\nAccordingly, let us start from the \frst case which is\npossibly the simplest case. Since \u0000 Dand!nare constant\nin time and only the density N(t) varies, we can see this\nas a low-density approximation where we assume that\nthe electrons interact very weakly and, as a result, the\ncharacteristics of movement of a single electron do not\ndepend on the electron density. Under these assump-\ntions, the volume density of electric dipole moment or\npolarization density is written as\nP(t) =N(t)p(t) =Z+1\n0N(t)\u000b(\r;t)E(t\u0000\r)d\r=\nZ+1\n0\"0\u001f(\r;t)E(t\u0000\r)d\r;(47)\nin which the electric susceptibility kernel equals\n\u001f(\r;t) =N(t)\n\u000f0\u000b(\r;t) =\nN(t)\n\u000f0e2\nmq\n!2n\u0000\u00002\nD\n4exp\u0000\n\u0000\u0000D\n2\r\u0001\nsin\u0000r\n!2n\u0000\u00002\nD\n4\r\u0001\n:\n(48)\nHere, in Eq. (48), note that we take the polarizabil-\nity kernel from Eq. (40). Before proceeding, we draw\nthe attention to Eq. (47) which describes the polariza-\ntion density. In this equation, we simply wrote P(t)\nas the multiplication of the electric dipole moment and\nthe electron density. To do that, \frstly, we should as-\nsume that by varying the number of electrons (per unit\nvolume) in time, the time-varying material remains ho-\nmogeneous meaning that the electric susceptibility (or\npermittivity) does not depend on the position vector r:\n\u001f(r;\r;t) =\u001f(\r;t). Secondly, we should also suppose\nthat the process of changing the electron density in time\ndoes not a\u000bect the velocities of electrons. Within these\nassumptions, Eq. (47) is valid, and, in fact, we can write\nP(t) =N(t)p(t). Now, by knowing the electric suscep-\ntibility kernel from Eq. (48), we readily \fnd the relativepermittivity kernel of the e\u000bective medium as\n\u000f(\r;t) =\u000e(\r)+\nN(t)\n\u000f0e2\nmq\n!2n\u0000\u00002\nD\n4exp\u0000\n\u0000\u0000D\n2\r\u0001\nsin\u0000r\n!2n\u0000\u00002\nD\n4\r\u0001\n;\n(49)\nwhere\u000e(\r) is the one-dimensional Dirac delta function.\nSince we have the kernels from the above equations,\nEqs. (48) and (49), we can apply the Fourier transform\nEq. (19) and calculate the temporal complex relative per-\nmittivity de\fned in Eqs. (17) and (18). The result reads\n\u000fT(!;t) = 1 +!2\np(t)\n!2n\u0000!2+j\u0000D!; (50)\nin which\n!2\np(t) =e2\n\u000f0mN(t) (51)\nis the time-dependent plasma frequency. The expression\nin Eq. (50) is complex-valued and explicitly depends on\ntime indicating the nonstationarity characteristic. Sub-\nstituting!n= 0 (free-electron plasma) we arrive to the\nconventionally used expression for the e\u000bective permit-\ntivity of plasma with varying electron density, e.g. [48].\nThe only di\u000berence with the stationary case is that in the\nformula the plasma frequency explicitly depends on time.\nThe reason is due to the low-density approximation that\nwe have made. Within this approximation, the damping\ncoe\u000ecient and the natural frequency are considered to be\nconstant in time. Therefore, the polarizability kernel is\nthe same as the one written for the stationary scenario,\nand consequently the same kind of dispersion is observed.\nLet us consider a more general case when the damp-\ning coe\u000ecient and the natural frequency also change in\ntime. In this case, \u000fT(!;t) can be dramatically di\u000berent.\nSpeci\fc dependencies of the e\u000bective parameters in (50)\nare determined by the plasma structure and can be set\nas empirical parameters. As a particular example, here\nwe assume that h(t;\u001c) is a product of two functions K(t)\nandL(\u001c) which depend on single independent variables\ntand\u001c, respectively. Since h(t;t) = 0, according to the\ninitial condition in Eq. (39), one also assumes a multiplier\nin form (t\u0000\u001c)n. Thus, we consider the time-varying dis-\npersion kernel in form\nh(t;\u001c) = (t\u0000\u001c)nK(t)L(\u001c): (52)\nContemplating the second expression in Eq. (39), we \fnd\nthat\nn(t\u0000\u001c)n\u00001K(\u001c)L(\u001c) =e2=m; (53)\nin whichtand\u001cmust be equal. Due to this feature\n(t=\u001c), we can conclude that if n6= 1, the above iden-\ntity does not hold for nonzero functions of K(t) andL(\u001c).12\nTherefore, the only condition for holding the identity oc-\ncurs when nbecomes equal to unity, and, as a result,\nonly under this condition there is a nonzero solution for\nh(t;\u001c). Now, in the case of n= 1, satisfying the second\ninitial condition determines the function L(\u001c) as inversely\nproportional to K(\u001c) such that L(\u001c) =e2=[mK(\u001c)]. Us-\ning the partial di\u000berential equation (the \frst expression)\nin Eq. (39) and substituting\nh(t;\u001c) =e2\nm(t\u0000\u001c)K(t)\nK(\u001c); (54)\nwe \fnd the corresponding function K(t):\nK(t) = exp\u0010\n\u0000Z\u0000D(t)\n2dt\u0011\n; (55)\nwith an important constraint:\n!2\nn(t) =1\n4h\n\u00002\nD(t) + 2d\u0000D(t)\ndti\n: (56)\nThis equation shows that in this case the temporal vari-\nation of the natural frequency fully depends on the tem-\nporal variation of the damping coe\u000ecient.\nNext, let us assume a free-electron model so that !n=\n0. Based on Eq. (56), this assumption forces the damping\ncoe\u000ecient to vary homographically as \u0000 D(t) = 2\u0000 0=(1 +\n\u00000t), which results in K(t) = 1=(1 + \u0000 0t), according to\nEq. (55). With h(t;\u001c) = (e2=m)(t\u0000\u001c)(1+\u0000 0\u001c)=(1+\u0000 0t),\nthe electric polarizability kernel is given by\n\u000b(\r;t) =e2\nm\r\u0010\n1\u0000\u00000\r\n1 + \u0000 0t\u0011\n: (57)\nHaving the polarizability kernel from Eq. (57), and after\nsome algebraic manipulations, the relative permittivity\nkernel is expressed as\n\u000f(\r;t) =\u000e(\r) +N(t)\n\u000f0e2\nm\r\u0010\n1\u0000\u00000\r\n1 + \u0000 0t\u0011\n: (58)\nAs a sanity check, if \u0000 0= 0 andN(t) are time-invariant,\nwe obtain the conventional stationary lossless Drude\nmodel:\n\u000f(\r) =\u000e(\r) +N\n\u000f0e2\nm\r: (59)\nLet us again apply the Fourier transform (19) and \fnd\nthe temporal complex relative permittivity which corre-\nsponds to kernel (58). In accordance with the properties\nof Fourier transform, since\nZ+1\n0(\u0000j\r)nexp(\u0000j!\r)d\r=dn\nd!n\u00101\nj!\u0011\n; (60)\nwe \fnd that\n\u000fT(!;t) = 1\u0000!2\np(t)\n!2\u0000j2!2\np(t)\u00000\n!3(1 + \u0000 0t);\nor\n\u000fT(!;t) = 1\u0000!2\np(t)\n!2\u0000j!2\np(t)\n!3\u0000D(t):(61)The temporal permittivity has the imaginary part which\nis time-dependent, and is negative indicating that the\nmedium is lossy. Comparing the above expression with\nthe conventional stationary Drude model\n\u000fDrude (!) = 1\u0000!2\np\n!21\n1 + (\u0000D\n!)2\u0000j!2\np\n!3\u0000D\n1 + (\u0000D\n!)2;(62)\nwe explicitly observe how the e\u000bective permittivity of\nplasma with a time-varying damping coe\u000ecient cannot\nbe found by simply assuming that \u0000 Ddepends on time in\nthe conventional Drude formula: Time variations of the\ndamping coe\u000ecient \u0000 D(t) modi\fes the real and imagi-\nnary parts of the relative permittivity in a di\u000berent way,\nas is seen from Eq. (61).\nFinally, before \fnishing this section, we point out an\nissue about numerical simulations. The time-invariant\n(stationary) frequency-dispersive media have been thor-\noughly studied, and there are well developed time-\ndomain full-wave electromagnetic simulation tools such\nas Ansys HFSS, CST Microwave Studio, and COMSOL\nMultiphysics. However, up to our knowledge, there are\nno numerical tools that could \\properly\" simulate disper-\nsive time-varying (nonstationary) media. We hope that\nthis work will help developing such numerical methods\nwhich would allow studying electromagnetic processes in\ndispersive and time-varying media (including time-space\nmodulated metamaterials).\nV. CONCLUSIONS\nWe have theoretically studied the fundamental prin-\nciples related to the electric polarizability of arbitrary\ndipolar linear particles whose characteristics are varying\nin time due to some external force. Since at every mo-\nment of time the time-varying particle is di\u000berent, the po-\nlarizability additionally depends on the observation time\n(the one at which we measure the dipole moment). Im-\nportantly, this time-dependent polarizability is not fully\ndetermined by the particle parameters at the observation\nmoment, this polarizability is a causal-response parame-\nter which depends on the whole history of the particle.\nThis is in contrast with a stationary particle whose po-\nlarizability depends only on the delay time between the\nexcitation and observation moments.\nFor time-harmonic excitations, we demonstrated that\nthe instantaneous dipole moment is found as the real part\nof a complex-valued temporal function that is multiplied\nby the complex amplitude of the \feld and the time-\nharmonic exponential factor. This temporal response\nfunction is the Fourier transform of the polarizability ker-\nnel with respect to the delay time ( \r!!). We named\nthis function as temporal complex polarizability and ex-\nplained some of its salient properties. Importantly, using\nthe notion of temporal complex polarizability, we intro-\nduced the second Fourier transform that is with respect\nto the observation time ( t!\n). By this way, we could13\ndescribe the dipole moment completely in the frequency\ndomain.\nNext, we considered a nonstationary particle that is\nlocated in free space. By employing the R udenberg equa-\ntion (the Hertzian dipole model), we presented a method-\nical approach to determine the polarizability kernel and\nthe temporal complex polarizability. Having the polariz-\nability, we studied the classical interaction of the dipole\nwith the incident time-harmonic electromagnetic wave.\nTherefore, in terms of the temporal complex polarizabil-\nity, we contemplated the instantaneous powers that are\nextracted by the dipole and scattered from the dipole.\nFinally, we took one step forward and considered the\ndipole particle as a constituent of a time-varying ma-\nterial. This time, we focused on the classical bound\nelectron model and derived the corresponding equations\nfor \fnding the polarizability. To do that, we used the\nequation of motion (or the Drude-Lorentz equation) and\nassumed that the damping coe\u000ecient and the natural\nfrequency are varying in time. Afterwards, we moved\ntowards e\u000bective material parameters, and for particular\nexample cases, we derived the e\u000bective permittivity of the\ntime-varying medium comprising bound or free electrons.\nIt is observed that this model for describing the e\u000bective\npermittivity is signi\fcantly di\u000berent from the conven-\ntional Drude-Lorentz formula with time-dependent pa-\nrameters.\nACKNOWLEDGMENTS\nThis work was supported by the Academy of Finland\nunder grant 330260. M.S.M. wishes to acknowledge the\nsupport of Ulla Tuominen Foundation. Also, the authors\nthank V. Asadchy and A. Sihvola for their invaluable\ncomments. In addition, M.S.M. thanks X. Wang for help-\ning to prepare the \fgure of the paper.\nAppendix A: Polarizability of a classical bound\nelectron for the stationary scenario\nLet us present the rigorous derivation of Eq. (40). The\n\frst expression in Eq. (39) results in\nd2h(u)\ndu2+ \u0000Ddh(u)\ndu+!2\nnh(u) = 0!h(u) = exp(\u0000\u0000D\n2u)\n\u0002\u0010\nH1cos(r\n!2n\u0000\u00002\nD\n4u) +H2sin(r\n!2n\u0000\u00002\nD\n4u)\u0011\n:\n(A1)\nHere, there are two unknown coe\u000ecients, H1andH2,\nwhich should be determined from the other two remain-\ning expressions in Eq. (39). In principle, the second and\nthe third expressions, as mentioned, are the initial condi-\ntions for determining a speci\fc solution for h(t;\u001c). Fromthe second expression, we deduce that\ndh(u)\nduju=0=e2\nm!\u0000\u0000D\n2H1+r\n!2n\u0000\u00002\nD\n4H2=e2\nm;\n(A2)\nand from the last expression in Eq. (39), we conclude\nthat\nh(0) = 0!H1= 0: (A3)\nFinally, by combining the above results, h(t;\u001c) and the\npolarizability are given by\nh(t;\u001c) =h(t\u0000\u001c) =\ne2\nmq\n!2n\u0000\u00002\nD\n4exp\u0000\n\u0000\u0000D\n2(t\u0000\u001c)\u0001\nsin\u0000r\n!2n\u0000\u00002\nD\n4(t\u0000\u001c)\u0001\n;\n(A4)\nand\n\u000b(\r;t) =h(t;t\u0000\r) =\ne2\nmq\n!2n\u0000\u00002\nD\n4exp\u0000\n\u0000\u0000D\n2\r\u0001\nsin\u0000r\n!2n\u0000\u00002\nD\n4\r\u0001\n;(A5)\nrespectively.\nAppendix B: Inverse polarizability and\ninstantaneous reactive powers\nUtilizing the introduced notion of the temporal com-\nplex polarizability, we can \fnd the instantaneous scat-\ntered power and the total extracted power. Here, we\nremind an important relation for the imaginary part of\nthe inverse polarizability which is known for stationary\ndipoles. In the lossless regime and in the time-averaged\nperspective, by writing that Sstationary =Sstationary\nrad, we\nsimply derive the following expression:\nImh1\n\u000bTi\n=k3\n0\n6\u0019\u000f0; (B1)\nwherek0and\u000f0are the free-space wavenumber and per-\nmittivity, respectively. However, under nonstationary\nconditions, the above equation is not true, and \fnding\na similar relation is not straightforward. This is due to\nthe fact that the conservation of instantaneous power is\nnotsimplyS(t) =Srad(t) even for the stationary dipole.\nThe stored reactive electric and magnetic energy per unit\ntime should be also taken into account. In other words,\nby neglecting Ohmic losses, S(t) =Srad(t) +Sreactive (t),\nwhereSreactive (t) =Selectric (t)+Smagnetic (t). To perceive\nthis expression, as an example, let us consider a nonsta-\ntionary Hertzian dipole loaded with a reactive element\n(e.g., an inductance Lload) changing in time. By using the\ncircuit theory, we can describe the instantaneous reactive\npowers,Selectric (t) andSmagnetic (t), through the e\u000bective14\nparameters of the dipole and the temporally modulated\nreactive load. Those e\u000bective parameters, which play\nan important role in the interaction of the dipole with\nthe incident \feld, are an e\u000bective capacitance Cand an\ne\u000bective inductance L. As it can be expected, the ef-\nfective capacitance Ccorresponds to the stored electric\nenergy near to the dipole, and the stored magnetic energy\naround the dipole is measured by the e\u000bective inductance\nL. Based on this explanation, one can conclude that the\ntotal electric energy per unit time is equal to Selectric (t) =\n(Q(t)=C)i(t) in whichQ(t) is the electric charge and i(t)\nis the induced electric current in the dipole. However,\nsince we assume a time-varying inductance as the load,\nthe total magnetic energy per unit time consists of contri-butions from the e\u000bective inductance and the load. Con-\nsequently,Smagnetic (t) =L[di(t)=dt]\u0001i(t)+Sload(t), where\nSload(t) = [d(Lload(t)i(t))=dt]\u0001i(t). In the above example,\nwe stress that in this derivation the length of the dipole is\nassumed to be constant in time, and the temporal mod-\nulation comes only from the time-varying load connected\nto the dipole. However, a short dipole can be modulated\nin time in a number of di\u000berent ways, for example, the\ndipole length can vary in time or the shape of a sub-\nwavelength particle such as a lossless sphere can change\nin time. For each speci\fc case, implications of the above\nequations and expressions are indeed intriguing and need\nto be carefully investigated.\n[1] M. Faraday, On a peculiar class of acoustical \fgures; and\non certain forms assumed by groups of particles upon\nvibrating elastic surfaces, Phil. Trans. R. Soc. Lond. 121,\n299 (1831).\n[2] A. L. Cullen, A travelling-wave parametric ampli\fer, Na-\nture181, 332 (1958).\n[3] P. K. Tien, Parametric ampli\fcation and frequency mix-\ning in propagating circuits, J. Appl. Phys. 29, 1347\n(1958).\n[4] F. R. Morgenthaler, Velocity modulation of electromag-\nnetic waves, IRE Trans. Microwave Theory Tech. 6, 167\n(1958).\n[5] A. K. Kamal, A parametric device as a nonreciprocal\nelement, Proc. IRE 48, 1424 (1960).\n[6] M. R. Currie, and R. W. Gould, Coupled-cavity\ntraveling-wave parametric ampli\fers: Part I-analysis,\nProc. IRE 48, 1960 (1960).\n[7] J. C. Simon, Action of a progressive disturbance on\na guided electromagnetic wave, IRE Trans. Microwave\nTheory Tech. 8, 18 (1960).\n[8] J. R. Macdonald, and D. E. Edmondson, Exact solution\nof a time-varying capacitance problem, Proc. IRE 49,\n453 (1961).\n[9] A. Hessel, and A. A. Oliner, Wave propagation in a\nmedium with a progressive sinusoidal disturbance, IRE\nTrans. Microwave Theory Tech. 9, 337 (1961).\n[10] B. D. O. Anderson, and R. W. Newcomb, On reciprocity\nand time-variable networks, Proc. IEEE 53, 1674 (1965).\n[11] D. E. Holberg, and K. S. Kunz, Parametric properties\nof \felds in a slab of time-varying permittivity, IEEE\nTrans. Antennas Propag. 14, 183 (1966).\n[12] Z. Yu, and S. Fan, Complete optical isolation created by\nindirect interband photonic transitions, Nature Photon-\nics3, 91 (2009).\n[13] D. L. Sounas, and A. Al\u0012 u, Angular-momentum-biased\nnanorings to realize magnetic-free integrated optical iso-\nlation, ACS Photonics 1, 198 (2014).\n[14] Y. Hadad, D. L. Sounas, and A. Al\u0012 u, Space-time gradient\nmetasurfaces, Phys. Rev. B 92, 100304(R) (2015).\n[15] D. L. Sounas, and A. Al\u0012 u, Non-reciprocal photonics\nbased on time modulation, Nature Photonics 11, 774\n(2017).\n[16] T. T. Koutserimpas, and R. Fleury, Nonreciprocal gain\nin non-Hermitian time-\roquet systems, Phys. Rev. Let-ters120, 087401 (2018).\n[17] S. Taravati, N. Chamanara, and C. Caloz, Nonreciprocal\nelectromagnetic scattering from a periodically space-time\nmodulated slab and application to a quasisonic isolator,\nPhys. Rev. B 96, 165144 (2017).\n[18] X. Wang, G. Ptitcyn, V. S. Asadchy, A. D\u0013 \u0010az-Rubio,\nM. S. Mirmoosa, S. Fan, and S. A. Tretyakov, Non-\nreciprocity in bianisotropic systems with uniform time\nmodulation, Phys. Rev. Lett. 125, 266102 (2020).\n[19] M. M. Salary, S. Farazi, and H. Mosallaei, A dynami-\ncally modulated all-dielectric metasurface doublet for di-\nrectional harmonic generation and manipulation in trans-\nmission, Adv. Optical Mater. 7, 1900843 (2019).\n[20] Z. Wu, and A. Grbic, Serrodyne frequency translation\nusing time-modulated metasurfaces, IEEE Trans. An-\ntenn. Propag. 68, 1599 (2020).\n[21] M. Liu, D. A. Powell, Y. Zarate, and I. V. Shadrivov,\nHuygens' metadevices for parametric waves,\nPhys. Rev. X 8, 031077 (2018).\n[22] M. M. Salary, S. Jafar-Zanjani, and H. Mosallaei, Elec-\ntrically tunable harmonics in time-modulated metasur-\nfaces for wavefront engineering, New J. Phys. 20, 123023\n(2018).\n[23] N. Chamanara, Y. Vahabzadeh, and C. Caloz, Simulta-\nneous control of the spatial and temporal spectra of light\nwith space-time varying metasurfaces, IEEE Trans. An-\nten. Propag. 67, 2430 (2019).\n[24] S. Taravati, and A. A. Kishk, Dynamic modulation yields\none-way beam splitting, Phys. Rev. B 99, 075101 (2019).\n[25] M. S. Mirmoosa, G. A. Ptitcyn, V. S. Asadchy,\nand S. A. Tretyakov, Time-varying reactive elements\nfor extreme accumulation of electromagnetic energy,\nPhys. Rev. Applied 11, 014024 (2019).\n[26] T. T. Koutserimpas, A. Al\u0012 u, and R. Fleury, Paramet-\nric ampli\fcation and bidirectional invisibility in PT-\nsymmetric time-Floquet systems, Phys. Rev. A 97,\n013839 (2018).\n[27] T. T. Koutserimpas, and R. Fleury, Electromagnetic\nwaves in a time periodic medium with step-varying re-\nfractive index, IEEE Trans. Anten. Propag. 66, 5300\n(2018).\n[28] A. Shlivinski, and Y. Hadad, Beyond the Bode-Fano\nbound: Wideband impedance matching for short pulses\nusing temporal switching of transmission-line parame-15\nters, Phys. Rev. Letters 121, 204301 (2018).\n[29] G. Ptitcyn, M. S. Mirmoosa, and S. A. Tretyakov, Time-\nmodulated meta-atoms, Phys. Rev. Research 1, 023014\n(2019).\n[30] M. S. Mirmoosa, G. A. Ptitcyn, R. Fleury, and\nS. A. Tretyakov, Instantaneous radiation from time-\nvarying electric and magnetic dipoles, Phys. Rev. A 102,\n013503 (2020).\n[31] A. Serdyukov, I. Semchenko, S. Tretyakov, and A. Si-\nhvola, Electromagnetics of Bi-anisotropic Materials:\nTheory and Applications (Gordon and Breach Science\nPublishers, Amsterdam, 2001).\n[32] C. Caloz, and Z.-L. Deck-L\u0013 eger, Spacetime\nmetamaterials{Part I: General concepts, IEEE\nTrans. Antenn. Propag. 68, 1569 (2020).\n[33] C. Caloz, and Z.-L. Deck-L\u0013 eger, Spacetime\nmetamaterials{Part II: Theory and applications,\nIEEE Trans. Antenn. Propag. 68, 1583 (2020).\n[34] A. Sihvola, Electromagnetic Mixing Formulas and Appli-\ncations (The Institution of Electrical Engineers, London,\nUK, 1999).\n[35] V. Pacheco-Pe~ na, and N. Engheta, E\u000bective medium con-\ncept in temporal metamaterials, Nanophotonics 9, 379\n(2020).\n[36] H. D'Angelo, Linear Time-Varying Systems: Analysis\nand Synthesis (Allyn & Bacon, Boston, 1970).\n[37] L. A. Zadeh, Frequency analysis of variable networks,\nProc. IRE 38, 291 (1950).\n[38] L. D. Landau, and E. M. Lifshitz, Electrodynamics of\nContinuous Media (Pergamon, Oxford, 1960).[39] J. Schwinger, L. L. DeRaad, Jr., K. A. Milton, and W.-\nY. Tsai, Classical Electrodynamics (Perseus, Reading,\nMA, 1998).\n[40] J. D. Jackson, Classical Electrodynamics (John Wiley &\nSons, New York, 1999).\n[41] J. R. Zurita-S\u0013 anchez, P. Halevi, and J. C. Cervantes-\nGonz\u0013 alez, Re\rection and transmission of a wave incident\non a slab with a time-periodic dielectric function \u000f(t),\nPhys. Rev. A 79, 053821 (2009).\n[42] N. V. Budko, Electromagnetic radiation in a time-varying\nbackground medium, Phys. Rev. A 80, 053817 (2009).\n[43] J. S. Mart\u0013 \u0010nez-Romero, O. M. Becerra-Fuentes, and\nP. Halevi, Temporal photonic crystals with modulations\nof both permittivity and permeability, Phys. Rev. A 93,\n063813 (2016).\n[44] N. S. Stepanov, Dielectric constant of unsteady plasma,\nIzv. Vyssh. Uchebn. Zaved., Radio\fz. 19, 960 (1976).\n[45] R. R udenberg, Der Empfang elektrischer Wellen in der\ndrahtlosen Telegraphie, Annalen der Physik 330, 446\n(1908).\n[46] L. Novotny, and B. Hecht, Principles of Nano-Optics\n(Cambridge University Press, Cambridge, UK, 2012).\n[47] S. Tretyakov, Maximizing absorption and scattering by\ndipole particles, Plasmonics 9, 935 (2014).\n[48] S. G. Ohler, B. E. Gilchrist, and A. D. Gallimore, Elec-\ntromagnetic signal modi\fcation in a localized high-speed\nplasma \row: Simulations and experimental validation\nof a stationary plasma thruster, IEEE Transactions on\nPlasma Science 27, 587 (1999)." }, { "title": "2402.11606v1.Finite_frequency_normal_and_superfluid_drag_effects_in_two_component_atomic_Bose_Einstein_condensates.pdf", "content": "Finite-frequency normal and superfluid drag effects\nin two-component atomic Bose-Einstein condensates\nAzat F. Aminov,1, 2,∗Alexey A. Sokolik,3, 1,†and Yurii E. Lozovik3, 1\n1National Research University Higher School of Economics, 109028 Moscow, Russia\n2Institute of Microelectronics Technology and High Purity Materials,\nRussian Academy of Sciences, Chernogolovka 142432, Russia\n3Institute for Spectroscopy, Russian Academy of Sciences, 142190 Troitsk, Moscow, Russia\nTwo-component systems consisting of mutually interacting particles can demonstrate both in-\ntracomponent transport effects and intercomponent entrainment (or drag) effects. In the presence\nof superfluidity, the intracomponent transport is characterized by dissipative conductivity and su-\nperfluid weight in the framework of two-fluid model, and intercomponent entrainment gives rise to\nnormal and nondissipative drag effects. We present unified treatment of all these effects for spatially\nhomogeneous two-component atomic Bose-Einstein condensates based on the Bogoliubov theory,\nfocusing specifically on the drag effects. Calculating finite-frequency intra- and intercomponent\nconductivities with taking into account quasiparticle damping, we derive and numerically check an-\nalytical Drude-like approximations applicable at low frequencies, and Lorentz-like approximations\napplicable at higher frequencies in vicinity of the resonant energy of spin-to-density Bogoliubov\nquasiparticle conversion. As possible physical realizations of two-component atomic systems, we\nconsider three-dimensional Bose-Bose mixtures and closely spaced two-layered systems of magnetic\ndipolar atoms.\nI. INTRODUCTION\nUnderstanding of many-body phenomena in ultracold\natomic gases helps to shed light on the properties of con-\ndensed matter systems. For example, studying Bose-\nEinstein condensation (BEC) of ultracold atomic gases\nprovides deeper insight into physics of superconductors,\nsuperfluids, and strongly correlated systems [1–4]. One of\nsuch phenomena, which might occur in semiconducting,\nsuperconducting, and ultracold atomic systems, is drag\neffect, the transport phenomenon which reveals both\nsingle-particle and many-body physics.\nThe Coulomb drag effect in closely spaced two-layer\nsystems, which is caused by frictional entrainment of par-\nticles in one layer in response to a current in the other\nlayer, is extensively studied in solid-state electronic sys-\ntems [5]. Experimentally, this effect is detected by mea-\nsuring nonlocal transresistance between layers. In super-\nfluid or superconducting two-component systems, a non-\ndissipative counterpart of the drag, or Andreev-Bashkin\neffect, can also emerge, when superfluid or superconduct-\ning components of the constituents entrain each other\nwithout dissipation. This effect was predicted for3He-\n4He mixtures [6], superconducting systems [7], ultracold\natomic gases [8], superfluid mixtures of nucleons in the\ncores of neutron stars [9], and for superconducting layers\ninteracting with polaritons [10].\nBoth Coulomb and Andreev-Bashkin drag effects are\nconventionally studied in the DC regime (at ω= 0).\nRecently there appeared an interest in studying the AC\n(ω >0) drag effect [11]: an alternating force at nonzero\n∗afaminov@hse.ru\n†asokolik@hse.ru\nL(b) (a)z\njbja Fa\njbja FaFIG. 1. Schematic depiction of systems considered. (a) 3D\natomic mixture. Alternating force Faimposed on the con-\nstituent agives rise to both intracomponent ja∼σaaFaand\nintercomponent jb∼σabFaresponse currents. (b) Dipolar\natomic quasi-2D system. Atoms in two pancake-like Bose-\ncondensed clouds have their dipole moments aligned with the\nzaxis, and the drag effects are induced by long-range inter-\naction across the interlayer distance L.\nfrequency acts upon one component, and the alternat-\ning current of the other component is detected. This\neffect is described by the conductivity matrix σij(ω) re-\nsolved over the components i, j=a, b. The key feature\nof AC drag effects in a superfluid system is interplay of\ndissipative and non-dissipative current responses, when\nthe Coulomb drag and Andreev-Bashkin effects in their\npure DC form can be extracted from analysis of the low-\nfrequency limit of AC drag conductivity σab(ω).\nIn this paper we calculate, using many-body theory,\nthe AC mass conductivities σij(ω) of a two-component\natomic BEC at nonzero temperature, considered as a\nhomogeneous 3D mixture [see Fig. 1(a)]. We analyze\nboth intracomponent conductivities (or intraconductivi-\nties)σaa,σbb, which characterize normal and superfluid\nresponses of each component, and intercomponent con-\nductivity (or transconductivity )σab, which is responsi-arXiv:2402.11606v1 [cond-mat.quant-gas] 18 Feb 20242\nble for normal and superfluid drag effects. In contrast\nto Ref. [11], we consider generally non-symmetric two-\ncomponent system with different masses and densities of\nconstituents, and assume nonzero damping γof the Bo-\ngoliubov excitations, which can arise from interaction of\natoms with the external disordered potentials, conden-\nsate inhomogeneities, and non-condensed particles [12–\n14].\nBesides three-dimensional mixtures, we study spatially\nseparated magnetic dipolar atomic gases [see Fig. 1(b)],\nwhere the interlayer drag can appear due to long-range\ndipole-dipole interaction. Such systems are gaining pop-\nularity nowadays: BEC of dipolar atomic gases was re-\nalised in recent experiments [15], and mutual friction (i.e.\nnormal drag effect) in a non-condensed phase was de-\ntected in the two-layered geometry [16].\nWe show that at large enough temperatures the dissi-\npative and non-dissipative response currents may be of\nthe same order, leading to non-trivial phase shifts (be-\nsides 0 and π/2) between currents and driving forces.\nAt low frequencies both intra- and transconductivities\nσij(ω) are well approximated analytically by a kind of\ntwo-fluid Drude model [17, 18] with a mixture of nondis-\nsipative response (giving rise to superfluidity of each com-\nponent and Andreev-Bashkin effect between the com-\nponents) and dissipative response caused by quasipar-\nticle decay (which gives rise to a normal conductivity\nof each component and normal drag between the compo-\nnents). At higher frequencies of the order of atomic chem-\nical potentials, σij(ω) in certain conditions can reveal\nthe Lorentz-type resonance originating from interconver-\nsion between spin and density Bogoliubov quasiparti-\ncles, which is also analytically approximated. For dipolar\natoms with interlayer interaction, we predict similar be-\nhavior of transconductivity, although the resonance fre-\nquency may be tuned by changing the interlayer distance.\nThe paper is structured as follows. In Sec. II the\noutline of the theory is presented, providing the general\nexpressions for conductivity calculations and parameters\nof the atomic systems we consider. Then in Sec. III we\nderive analytic approximations for AC conductivities in\nthe Drude (low-frequency) and Lorentz (high-frequency)\nregimes, followed by Sec. IV, where the results of numer-\nical calculations are presented and compared with the\nanalytical approximations. Sec. V concludes the paper\nwith discussion. Appendices A, B, C, D present details\nof calculations.\nII. THEORY\nA. Intra- and transconductivities\nSuperfluid, dissipative and drag transport effects in a\ntwo-component system are characterized by AC intracon-\nductivities σaa(ω),σbb(ω), and transconductivity σab(ω),\nwhich relate the force Fje−iωtimposed on the component\njto the current density (or flux density of particles) in-duced in the component i:\nji(t) =σij(ω)Fje−iωt. (1)\nSuch conductivities have dimensionality of ℏ−1cm−1\n(ℏ−1) for 3D (2D) system. In experiments on ultra-\ncold atomic gases, they can be determined by measuring\nvelocities and coordinates of atoms using time-of-flight\nexpansion imaging [19] or temperature change due to\ndissipation-induced heating [20].\nIn the linear response theory, the AC conductivities\ncan be related to the retarded correlation functions of\ncurrents [11, 21]\nσij(ω) =i\nω\u0014δijni\nmi+ lim\nq→0χT\nij(q, ω)\u0015\n. (2)\nHere the first term is diamagnetic response present only\nin the intracomponent channel i=j, with niandmi\nbeing the atomic density and mass of the ith compo-\nnent; χT\nijis transverse part of the retarded paramagnetic\ncurrent response tensor. In Matsubara representation at\nnonzero temperature T, this tensor is given by\nχνη\nij(q, iω) =−1\nA1/TZ\n0dτ eiωτ\nTτjν\ni(q, τ)jη\nj(−q,0)\u000b\n,(3)\nwhere jν\ni(q, τ) is the Heisenberg-evolved (in imaginary\ntime) operator of the qth spatial harmonic of the ith\ncomponent current density along the axis ν,Ais the\nsystem volume (area) in the case of 3D (2D) geometry,\nand hereafter ℏ= 1 is assumed in the formulas. The\nretarded correlation function of currents entering Eq. (2)\ncan be obtained from the Matsubara one (3) by taking\ntransverse tensorial part over ν,η, and performing an-\nalytic continuation iω→ω+i0 from the upper half of\nthe complex plane. Since we are interested in response\nof currents on a homogeneous force Fi, we take q= 0 in\nEq. (3). Note, however, that the ω→0 limit has to be\ntaken carefully in DC conductivity calculations. It can be\nshown [22], that in order to correctly calculate the super-\nfluid drag density [23] in a system without quasiparticle\ndamping, the DC limit ω→0 has to be taken before the\nq→0 limit, although in the presence of damping these\ntwo limits commute, which will be used below.\nThe induced current can be divided into the in-phase\n(with respect to the driving force) part, which is responsi-\nble for dissipation, and π/2 phase-delayed part, which is\nnon-dissipative. The latter can be additionally divided\ninto the diamagnetic and paramagnetic contributions.\nBy this reason, imaginary part of each conductivity\nσs\nij(ω)≡Imσij(ω) =δijni\nmiω+σsp\nij(ω), (4)\nwhich consists of dia- ( δijni/mi) and paramagnetic ( σsp\nij)\nparts, will be referred to as superfluid conductivity, and\nthe real part\nσn\nij(ω)≡Reσij(ω) (5)3\nwill be referred to as normal conductivity. Note that the\ndistinction between superfluid and normal responses is\nstrictly defined only in the DC limit ω= 0 [17, 18, 24],\nwhere the 1 /ωsingularity of σs\nijindicates superfluidity\nin both inter- and intracomponent channels. In particu-\nlar, the theory of DC superfluid drag [23] deals with the\nsuperfluid drag mass density\nρdr=mamblim\nω→0ωσs\nab(ω), (6)\nand the intracomponent superfluid mass density is re-\nlated to the low-frequency divergence of the intraconduc-\ntivity\nρs\ni=m2\nilim\nω→0ωσs\nii(ω). (7)\nIn contrast, σn\nij(ω) tends to a constant in the limit ω→0,\nand its intracomponent part σn\nii(0) provides dissipative\nDC conductivity, while the intercomponent part σn\nab(0)\nis related to the DC drag coefficient, or transresistiv-\nityσn\nab(0)/[σn\naa(0)σn\nbb(0)−σn\nab(0)2], which is usually mea-\nsured in drag experiments [5].\nAtω >0, the strict distinction between superfluid and\nnormal responses becomes elusive because even in normal\nsystems both σn\nijandσs\nijare finite and nonzero. Detect-\ning the dissipative response of a normal current against\nthe superfluid background is harder task for Bose systems\nthan for conventional s-wave superconductors, where the\ndissipative part of conductivity is suppressed at ω <2∆\n[25, 26]. In contrast, Bose-condensed systems lack gap in\nthe quasiparticle spectrum, so the normal conductivity is\ngenerally nonzero at any ω >0 [26]. This is why analy-\nsis of AC conductivities σij(ω) provides unified and more\ndetailed information about both the superfluid entraint-\nment (Andreev-Bashkin) effect and normal drag effect, as\nwell as about normal (dissipative) and superfluid (nondis-\nsipative) responses of each component, than conventional\nDC calculations commonly accepted in the theories of\ndrag and superfluidity.\nAs specific examples of 3D mixtures [Fig. 1(a)], we\nconsider spinor atomic BECs: the symmetric mixtures\nof87Rb and23Na in atomic states F= 1, mF=±1\n[27], and the non-symmetric mixture of39K in the states\nF= 1, mF= 1 and F= 1, mF= 0 [28]; here mFare\nmagnetic sublevels of hyperfine state with total angular\nmomentum F. Besides, we consider the mixture of atoms\nwith different masses,174Yb-133Cs [29]. As the spatially\nseparated quasi-2D dipolar system [Fig. 1(b)], we con-\nsider pairs of parallel clouds of either52Cr or168Er atoms\nwith the long-range magnetic dipole interaction [30].\nThe realistic parameters used in numerical calculations\nare listed in the Table I. Each intra- or intercomponent\ninteraction constant gij= 2πas\nij(1/mi+ 1/mj) is related\nto the s-wave scattering length as\nij, and we neglect the\nprocesses which permit population transfer between dif-\nferent magnetic sublevels mFof the state with total angu-\nlar momentum F. In the case of spatially separated dipo-\nlar atomic clouds, we take into account both s-wave scat-\ntering and dipole-dipole interaction within each cloud,Non-dipolar atoms\nParameter87Rb23Na39K174Yb-133Cs\nas\nii(a0) 100 55 30, 100 105, 150\nas\nab(a0) 95 51 −50 −75\nTc(nK) 170 1000 150 460, 200P\niµi/2π(kHz) 2.7 10.8 1 12.7\nDipolar atoms\nParameter52Cr168Er\ndi(µB) 6 7\nas\nii(a0) 103 137\nTc(nK) 700 410\nTABLE I. Upper table: Parameters for 3D spinor mixtures,\nnamely intra- as\niiand intercomponent as\nabscattering lengths\nin the units of the Bohr radius a0≈0.529˚A, critical tem-\nperatures Tc, and sums of chemical potentials µiof the com-\nponents at T= 0 (separated by a comma for non-symmetric\nmixtures). Lower table: parameters for dipolar atoms includ-\ning magnetic dipole moments diin Bohr magneton µBunits.\nand only the dipole interaction between atoms from dif-\nferent clouds (see details in Appendix D). In accordance\nwith the recent experiment [16], the thickness of both\nclouds is assumed to be wz= 20 nm, and the distance\nbetween clouds is L= 60 nm.\nIn this paper we consider systems with large con-\ndensate fraction, when the temperature is much lower\nthan the BEC critical temperatures of both constituents,\nT≪Ti\nc, but nonzero, since we are interested in the nor-\nmal drag effect as well. For 3D homogeneous atomic\ngases, the condensate densities n0\niare found with tak-\ning into account their thermal and quantum depletions\nfrom the system of equations n0\ni(T) =ni−nnc\ni(n0\na, n0\nb, T),\ni=a, b, where nnc\niis a density of non-condensed fraction\ngiven by Eq. (A8), niis the total density of the icompo-\nnent, which is assumed to be temperature-independent\nand estimated from the experimental critical tempera-\nture as ni=ζ(3/2)\u0000\nmiTi\nc/2π\u00013/2. The sums of chemical\npotentialsP\niµi=P\nigiin0\nilisted in Table I, which pro-\nvide characteristic energy scales of excitation energies,\nare taken at zero temperature.\nB. Current response function\nIn order to calculate the current response function (3),\nwe use diagrammatic technique to express it in the single-\nloop approximation through the intra- and intercompo-\nnent matrix Green functions ˆGij. Explicit formulas for\nthe Green functions are provided in Appendix A, and\ncalculation details for the current response are given in\nAppendix B.\nIt is known [23, 31], that in a two-component BEC\ntwo types of quasiparticles emerge, which correspond to\ndensity and spin collective modes, with dispersions Ed(q)\nandEs(q), respectively [see Fig. 2(a-b)]. Calculating the4\ntotal current response function χµν\nij, we express it through\nthe response functions S(Eα, Eβ) resolved over the quasi-\nparticle branches α, β= d,s and weighted with uandv\nBogoliubov coefficients. For the transverse part of the\ncurrent response tensor (3), we obtain\nχT\nij(q= 0, iω) =X\npp2\n2Adm imjX\nα1α2s1s2s1s2\n×\u0000\nus1\niα1us2\niα2−u−s1\niα1u−s2\niα2\u0001\u0000\nus1\njα1us2\njα2−u−s1\njα1u−s2\njα2\u0001\n×S(s1Eα1, s2Eα2).(8)\nHere the sums are taken over d-dimensional momentum\np, positive and negative energy indices s1,2=±, and\nquasiparticle branches α1,2= d,s. The response function\nS(Eα, Eβ) =−TX\niωn1\n(iωn−Eα+iω)(iωn−Eβ)(9)\ncorresponds to the loop-diagram constructed from two\nMatsubara Green functions 1 /(iωn−Eα) of Bogoliubov\nquasiparticles, which have infinite lifetime. Since we aim\nto analyze both normal and superfluid drag effects, we\nought to account for their non-zero damping γby re-\nplacing the quasiparticle Green functions 1 /(iωn−Eα)\nwith the broadened onesR\ndx ρ α(x)/(iωn−x), where\nρα(x) = (γ/π)[(x−Eα)2+γ2]−1is the Lorentzian spec-\ntral function. For simplicity of the forthcoming analytical\ncalculations, we assume γto be momentum- and energy-\nindependent and to be the same for both spin and density\nmodes. In this approximation the sum over Matsubara\nfrequencies in Eq. (9) can be taken analytically:\nS(Eα, Eβ) =Z\ndxdx′ρα(x)ρβ(x′)nB(x′)−nB(x)\niω+x′−x,\n(10)\nwhere nB(x) = (ex/T−1)−1is the Bose-Einstein distribu-\ntion function. To approximate this integral, we perform\nTaylor expansion of nB(x) and nB(x′) near the max-\nimax=Eα, x′=Eβof the spectral functions. After\nthat, integration over x, x′and analytical continuation\niω→ω+i0 yield the approximate retarded S-function\nS(Eα, Eβ) =nB(Eα)−nB(Eβ)−iγ[n′\nB(Eα) +n′\nB(Eβ)]\nEα−Eβ−ω−2iγ,\n(11)\nwhich will be used in the following. Numerical verifi-\ncation of this approximation proves its accuracy in the\nconsidered parameter ranges for α̸=β. In contrast, at\nα=βthis approximation lacks quantitative accuracy\nin the Hagen-Rubens regime ω≪γ, although it pro-\nvides qualitatively correct results and becomes exact in\nthe clean DC limit γ= 0, ω→0 (assumed, e.g., in the\nsuperfluid drag calculations in Ref. [23]).\nWe will limit ourselves to the case of relatively weak\ndamping γto maintain applicability of the quasiparticle\ndescription. Similarly to the Mott-Ioffe-Regel bound [32],\nvalidity of quasiparticle description requires the mean\n(c)\n(e)\n(f)\n \n ωdd\n \n \n \n \nωsd\n \n \n \n \nωsd\n \n \n≈ μa+μb\n∝ p2(d)\n \n \nωs\n \n s(a)\n(b)Δ\nΔFIG. 2. Left panels: Bogoliubov quasiparticle dispersions\nEd,s(p) in the cases of close (a) and distant (b) atomic masses.\nGreen arrows indicate energy differences Ed(p)−Es(p) at the\nrelevant momenta p∼¯p. Right panels (c-f): excitations of\npairs of the Bogoliubov quasiparticles contributing to conduc-\ntivities.\nfree path l=ci/γof quasiparticles (with their character-\nistic velocities ci=p\nµi/mi) being larger than the mean\ninterparticle distance n1/3\ni. Expressing nithrough the Ti\nc,\nwe obtain restriction for the damping rate γ≪p\nµiTic.\nFortunately this condition allows us to consider the sys-\ntem in ballistic regime and neglect the vertex corrections,\nbecause the ballistic approximation is appropriate when-\never ¯pl > 1 [33], where ¯ pis the characteristic momentum\nof quasiparticles defined in the next section. Roughly es-\ntimating this momentum as ¯ p∼√\nmT(see Appendix C),\nwe obtain ¯ pl∼√µiT/γ. At low enough damping rate\nassumed above, we obtain ¯ plmuch larger than the ratio p\nT/Tic, which is expected to be of the order of unity\nat moderate temperatures T∼1\n3Ti\nctaken in our cal-\nculations. Thus our neglect of the vertex corrections is\nconsistent in the assumed range of parameters.\nIII. ANALYTICAL APPROXIMATIONS\nA. Contributions of quasiparticle branches\nIn this section we develop analytical approximations\nfor frequency-dependent conductivities σn\nij(ω), σs\nij(ω),\nwhich resemble the familiar Drude and Lorentz mod-\nels. Our analysis is applicable to 3D atomic mixtures\nwith short-range interactions in the temperature range\nµi< T≪Tc. Inserting the approximate S-function (11)\ninto Eq. (8) and performing momentum integration, we\nobtain\nσij(ω)≈i\nω\u001aδijni\nmi−D0\nij\n−(−1)δij\nmimj\u0002\nΛ+(ω) + Λ−(ω)\u0003\u001b\n+iD0\nij\nω+ 2iγ.(12)5\nHere the D0\nijterms describe processes where quasiparti-\ncles are scattered from one branch into the same branch\n[density-to-density and spin-to-spin, see Figs. 2(c-d)],\nand the corresponding conductivity weights are defined\nas\nD0\nij=−X\npp2\n2Adm imj[PidPjdn′\nB(Ed) +PisPjsn′\nB(Es)].\n(13)\nThe coefficients Piα, quantifying contribution of the ith\ncomponent to Bogoliubov excitation branch α, are de-\nfined by Eq. (A5). The expression (13) is the counterpart\nof conventional Landau formula for density of the normal\ncomponent [34, 35] generalized for a two-component su-\nperfluid system.\nTwo other terms Λ±(ω) depend on frequency and can-\nnot be calculated analytically, so we derive approxima-\ntions for them. The function Λ+(ω) is responsible for the\nprocesses of quasiparticle scattering with interconversion\nfrom one branch to the distinct one [spin-to-density and\nvice versa, see Fig. 2(e)]. The second function Λ−(ω)\ncorresponds to creation or annihilation of two quasipar-\nticles of different branches [Fig. 2(f)]; note that similar\nsame-branch processes are forbidden at q= 0. These\nfunctions can be written as momentum integrals\nΛ±(ω) =∞Z\n0dp\u0002\nf±(p)R±(ω, p) +f∗\n±(p)R∗\n±(−ω, p)\u0003\n,\n(14)\nwhere f±(p) are defined in Appendix B and will be called\nenvelope functions , while the resonant functions are de-\nfined as\nR±(ω, p) =1\nEd(p)∓Es(p)−ω−2iγ. (15)\nThe envelope functions f±(p) endure power-law increase\nat low momenta and decrease exponentially at Ed,s≳T\nthanks to the Bose-Einstein distribution functions, so\nthey have extrema at some momentum ¯ pwhere the quasi-\nparticle energies match the temperature. Therefore it\nis convenient to define the characteristic momentum ¯ p,\nwhose neighbourhood provides the major contribution to\nthe integral, as solution of equation Ed(¯p) +Es(¯p) = 2 T\n(see more detailed discussion in Appendix C). Thus the\ncharacteristic sum of quasiparticle energies entering R−\nis of the order of T. The characteristic difference of en-\nergies entering R+is the important energy parameter\n∆ =Ed(¯p)−Es(¯p), (16)\nwhich has a meaning of resonance frequency for quasi-\nparticle inter-conversion processes [Fig. 2(e)] giving rise\nto the Lorentz-type response at moderately high ω. De-\npending on relationship between ωand ∆, we can sepa-\nrate the Drude and Lorentz regimes.\nγ <1∞\np(b)FIG. 3. Schematic depiction of different regimes on the\nω,γplane: the Drude regime at low frequencies and the\nLorentz regime at higher frequencies. According to the val-\nues of κshown by color, we separate the Lorentz regime\ninto weak- ( κ < 1) and strong-damping ( κ > 1) cases. In-\nsets (a-d) show how f+(p),R+(p), and the total integrand\nF(p) =f±(p)R±(ω, p) +f∗\n±(p)R∗\n±(−ω, p) in Eq. (14) behave,\nby absolute value, as functions of p.\nB. Drude regime\nThe Drude regime occurs when ωis far lower than the\nresonance frequencies Ed±Esin denominators of R±.\nAccording to the estimates above, it corresponds to the\nfrequency range ω≪∆, Tshown by green shading in\nFig. 3. In this limit we assume R±(ω, p)≈R±(0, p) in\nthe integrals (14), so the functions Λ±(ω) become almost\nfrequency-independent, and we obtain the simple expres-\nsion for conductivities in the Drude regime:\nσij(ω)≈i\nω\u001aδijni\nmi−D0\nij+D+\nij+D−\nij\u001b\n+iD0\nij\nω+ 2iγ.(17)\nHere\nD±\nij=−(−1)δij\nmimjΛ±(0)\n=∓(−1)δij\nmimj∞Z\n0dpRe2f±(p)\nEd∓Es−2iγ. (18)\nNote that the terms D±\nij, being almost reals, contribute\nmainly to the nondissipative part of the conductivities\n(17), because low frequencies ωare far off-resonant from\nthe absorption processes, corresponding to these terms\nand depicted in Fig. 2(e-f). Totally, the diamagnetic\nni/miωand paramagnetic −D0\nij+D+\nij+D−\nijterms in the6\nFIG. 4. Temperature dependencies of the Drude (left panel)\nand superfluid (right panel) weights for Yb-Cs mixture at\nγ/2π= 1 Hz. Solid lines correspond to the intracomponent,\nand black dashed line to the intercomponent conductivities.\nThin red line shows the approximation Ds\nab≈πρdr/mamb,\nwhere ρdris the drag density calculated at γ= 0 [23].\nbraces of Eq. (17) do not cancel each other in the Bose-\ncondensed regime giving rise to the uncompensated i/ω\nsingularity of both intra- and transconductivities in the\nDC limit ω→0. The superfluid weights corresponding to\nsuch singularities are Ds\nij=π(δijni/mi−D0\nij+D+\nij+D−\nij),\nand the Drude weights, corresponding to the dissipative\nresponse and defined as 2R∞\n0dωReσij(ω) [22], equal to\nDn\nij=πD0\nij.\nThe example of temperature dependencies of Drude\nand superfluid weights is shown in Fig. 4 for the mass-\nimbalanced Yb-Cs mixture. As expected, Drude weights\nvanish at T= 0, when the mixture is fully in superfluid\nstate, and superfluid weights involving Cs subsystem van-\nish at T=TCs\ncwhen it becomes normal. We also notice\nthat our results at low γare in agreement with the theory\nof DC superfluid drag developed for clean systems (red\nthin line) by Fil and Shevchenko [23]. The recession of\nthe intercomponent Drude weight Dn\nabdown to zero near\nT=TCs\ncis the artefact of our one-loop approximation,\nwhich neglects more complicated diagrams contributing\nto drag in the normal state [33]. However, they can be\nneglected at low enough temperatures T≲Ti\nc[36].\nC. Lorentz regime\nThe Lorentz regime occurs when ωis close to the res-\nonant energy ∆ of the spin-to-density quasiparticle con-\nversion. In this regime only R+demonstrates a reso-\nnance behavior and becomes dominant, and the other\nfunction R−can be neglected, because Ed+Es≫Ed−Es\nat typical momentum ¯ p. Also we may notice that the\nf+(p)R+(ω, p) term in the integral (14) is dominant over\nthe off-resonant term f∗\n+(p)R∗\n+(−ω, p).\nWe subdivide the Lorentz regime into weak- and\nstrong-damping cases, depending on the dimensionless\nparameter κ= ∆p/¯p, defined as the related to ¯ pmo-\nmentum width ∆ p≈4γ/[E′\nd(p+)−E′\ns(p+)] of the reso-\nnant function R+(ω, p) around its maximum at p=p+,which characterizes both the maximum and the typi-\ncal width of the envelope function f+(p). As shown in\nFig. 3, the weak-damping case κ≪1 [Fig. 3(d)] means\nthatR+(ω, p) is very narrow along the momentum axis,\nin comparison with f+(p). In the strong-damping case\nκ≫1, as depicted in Figs. 3(a-c), the situation is oppo-\nsite. We assign the frequency region ω > µ a+µb, where\nR+is never resonant and monotonously increases (so ∆ p\nis undefined), to the strong-damping case as well, setting\nformally κ=∞in this region.\n1. Weak-damping case\nIn the weak-damping case, when R±(ω, p) is very nar-\nrow, we can bring γin its denominator to zero and find\nΛ+(ω) analytically by integration of the resulting Dirac\ndelta function in Eq. (14) to obtain\nRe Λ+(ω)≈2∞Z\n0dp f +(p)Ed(p)−Es(p)\n[Ed(p)−Es(p)]2−ω2,(19)\nIm Λ+(ω)≈ −πf+(p+)\nE′\nd(p+)−E′s(p+), (20)\nwhere p+is solution of equation Ed−Es=ωdependent\nonω, i.e. the momentum where |R+(p, ω)|attains sharp\nmaximum; we also assume γ= 0 in the expression (B6)\nforf+(p). The result (20) for the function Im Λ+(ω), re-\nlated to the dissipation spectrum Re σij(ω), may be inter-\npreted as the sharp resonant function R+(ω, p) scanning\nthe broad envelope function f+(p) when ωis changed.\nThe derivative dp+/dω = [E′\nd(p+)−E′\ns(p+)]−1deter-\nmines the scanning speed along the paxis and hence\nmagnitude of Im Λ+(ω). The resonance maximum of\nReσij(ω) is located at p+= ¯p, where the maxima of\ntwo functions R+(ω, p) and f+(p) coincide. Shape of this\nresonance depends on the ratio of atomic masses ma,mb\nof two components. As discussed in more detail in Ap-\npendix C, we outline two cases: when atomic masses are\nclose to each other (and, in particular, equal in the case of\nspin mixtures), and when they are distant. These cases\nare distinguished by how the energy difference Ed−Es\ndepends on pin the relevant range of momenta p∼¯p.\nIn case of close masses this difference is almost constant\n[Fig. 2(a)], Ed−Es≈µa+µb, thus the scanning speed\ndp+/dωis high, and the resonance is sharp (see Fig. 5\nin the next section). In the case of distant masses the\nenergy difference retains an essential momentum depen-\ndence, Ed−Es∝p2[Fig. 2(b)], so dp+/dωis low, and\nthe resonance becomes strongly smeared or vanishes com-\npletely (see results for Yb-Cs mixture in Fig. 6 below).\n2. Strong-damping case\nIn the strong-damping case, the resonant function\nR+(ω, p) is much wider than the envelope f+(p) [κ≫1,7\nsee Figs. 3(a-c)], so we approximate R+(ω, p) byR+(ω,¯p)\nto obtain\nΛ+(ω)≈\u00121\n∆−ω−2iγ+1\n∆ +ω+ 2iγ\u0013∞Z\n0dp f +(p).\n(21)\nHere we neglected the γn′\nBterms in the f+(p) function\n(B6), since their order is γ/T≪1 (however in the Drude\nregime these terms should be retained, see Appendix B).\nThe conductivity in this case is predominantly deter-\nmined by the first term in the parentheses of Eq. (21)\nwhich is resonant near ω= ∆ with the width 2 γ. The\nrole of the second non-resonant term is to red-shift and\nbroaden this resonance.\nIV. NUMERICAL CALCULATIONS\nIn this section we present numerical results for the con-\nductivities for various systems and compare them with\nanalytical approximations. The numerically calculated\nconductivities are found using Eqs. (2), (8) with the ap-\nproximation (11) for the S-function, which proves to be\nquite accurate in the considered range of parameters.\nThe analytical approximations are given by Eq. (17) in\nthe Drude regime and Eq. (12) in the Lorentz regime,\nwith Λ−= 0 and Λ+given by Eqs. (19)–(20) in the\nweak-damping case ( κ <1) and by Eq. (21) in the strong-\ndamping case ( κ >1). For each atomic mixture, we take\nthe temperature T=1\n3min[Ta\nc, Tb\nc], which is low enough\nfor the Bogoliubov approximation to be applicable yet\nstill experimentally feasible.\nIn Fig. 5 we show the trans- and intraconductivities for\nthe symmetric Rb-Rb mixture at weak [Fig. 5(a,c)] and\nstrong [Fig. 5(b,d)] damping. At low frequencies ω≪γ,\nthe superfluid conductivities σs\nijare positive and diverge\nas 1/ω(in the intracomponent channel i=jthe pos-\nitive diamagnetic part na/maωdominates the negative\nparamagnetic part σsp\naa). It is a signature of nonzero and\npositive drag (6) and superfluid (7) densities. The nor-\nmal conductivities σn\nijtend to constants in DC limit in\nconformity with traditional normal Coulomb drag effect\nand Drude theory of conductivity, although in the weak-\ndamping case [Fig. 5(a,c)] their levelling off at ω→0\nis not visible at the chosen scale, because the Drude\npeaks are much higher than the Lorentz-regime features\nwhich we are concentrating on. At higher frequencies\nnear ω= ∆ (which is ∆ ≈µa+µbwhen ma=mb),\nabsolute value of the normal conductivity σn\nijexhibits\nabsorption peak, while the superfluid transconductivity\nσs\naband paramagnetic part σsp\naain the intracomponent\nchannel change sign. Such features resemble resonant\nbehaviour of the Lorentz model [Fig. 5(c,d)], although in\nthe intracomponent channel this resonant-like behavior\nof the paramagnetic superfluid conductivity (4) is masked\nby the large and monotonously decreasing diamagnetic\nterm.\nγ/2π = 1 Hz γ/2π = 300 Hz\nγ/2π = 300 Hz γ/2π = 1 HzRb-Rb(a) (b)\n(c) (d)FIG. 5. Trans- (a,b) and intercondictivity (c,d) for Rb-Rb\nspinor mixture at weak (left panels) and strong (right panels)\ndamping γ. Solid and dashed lines show numerical calcula-\ntions, and symbols show analytical approximations for ap-\npropriate regimes depicted by the same color shadings as in\nFig. 3: Drude regime (squares, green), weak-damping Lorentz\nregime (circles, blue), and strong-damping Lorentz regime\n(triangles, red). Vertical dashed lines indicate the frequency\nω=µa+µb, which is close to the resonance frequency ∆. In\nthe intracomponent channel (c,d) the superfluid conductivity\nσs\naais separated into dia- ( na/maω) and paramagnetic ( σsp\naa)\nparts.\nAt frequencies near the resonance ∆, the conduc-\ntivities are related to each other via m2\niσsp,n\nii(ω)≈\n−mambσs,n\nab(ω). This is evident in Fig. 5 where σsp\naa(ω)≈\n−σs\nab(ω) and σn\naa(ω)≈ −σn\nab(ω) near the resonance, since\nma=mb. This feature follows from Eq. (12) if we omit\nthe terms iD0\nij/(ω+ 2iγ) and −iD0\nij/ω, whose contribu-\ntion is diminished at large frequencies.\nTo get more insight into behavior of transconductivi-\nties, in Fig. 6 we show them for equal-mass K-K, Na-Na\nmixtures and for mass-imbalanced mixture Yb-Cs. It can\nbe seen that the transconductivity of the non-symmetric\nspin mixture K-K with relatively low resonance energy ∆\nexhibits the same features as for Rb-Rb mixture: Drude\npeak at low frequencies and resonance at ω≈∆. In the\ncase of symmetric spin mixture Na-Na, the resonance\nfrequency ∆ is higher (more than 10 kHz) and presum-\nably out of reach of present experiments capabilities. For\nthe mass-imbalanced mixture Yb-Cs, the resonance fre-\nquency ∆ turns out to be much lower than µa+µb, but\nthe resonance itself is degraded in both weak- and strong-\ndamping cases by the reasons discussed in Sec. III C 1.\nIn Fig. 7 we present numerically calculated transcon-\nductivities for the pairs of dipolar atomic gases Er-Er and\nCr-Cr arranged into quasi-2D two-layered systems [see\nFig. 1(b)]. The analytical approximations are not ap-8\nYb-Csγ/2π= 1 Hz\nγ/2π= 1 HzK-K\nNa-Naγ\n/2π\u0003= 600\u0003Hz\nγ/2π= 600 Hz\nγ/2π= 1 Hz γ/2π= 100 Hz\nFIG. 6. Transconductivities of K-K, Na-Na, and Yb-Cs mix-\ntures, each calculated at two values of γ, where the weak- or\nstrong-damping cases develop in the Lorentz regime. Desig-\nnations of curves and symbols are the same as in Fig. 5.\nplied in this case due to different form of intercomponent\ndipole-dipole interaction which retains essential momen-\ntum dependence, as discussed in Appendix D. In contrast\nto 3D mixtures with short-range interactions, here we can\ntune the resonance frequency ∆ by varying the interlayer\ndistance L. This frequency, found as the maximum of the\nquasiparticle energy difference ∆ = max[ Ed(p)−Es(p)],\napproximately follows the ∆ ∝L−1trend, as shown in\nEr-Er\nγ/2π = 10 Hz\nCr-Cr\nγ/2π = 1 Hz γ/2π = 100 Hz\nγ/2π = 300 Hz\nFIG. 7. Transconductivity between spatially separated clouds\nof dipolar atoms Er-Er and Cr-Cr in the systems at weak (left\npanels) and strong (right panels) damping with L= 60 nm.\nVertical dashed lines indicate the resonance frequency ω= ∆.\nInsets show dependence of ∆ on interlayer distance L.\nthe insets in Fig. 7. The intraconductivities in this case\nare not shown, since the relation between total niand\ncondensate n0\nidensities, needed to describe partial com-\npensation of the diamagnetic term with quantitative ac-\ncuracy, is not well-defined in 2D systems in the frame-\nwork of Bogoliubov theory, and more complicated ap-\nproaches, such as quasicondensate analysis [37], should\nbe applied, which is beyond the scope of our paper.\nV. DISCUSSION\nIn this paper, we studied intra- and transconductivi-\ntiesσij(ω) of a homogeneous two-component superfluid\nBose-condensed systems at nonzero frequencies ω. We\ncalculated the conductivities in one-loop approximation\nusing the Bogoliubov theory of a two-component BEC at\nfinite temperature and with taking into account the phe-\nnomenological damping γof spin and density quasipar-\nticle modes. Two possible setups of the two-component\natomic system are considered: 3D spinor atomic mix-\ntures (Rb-Rb, K-K, Na-Na, Yb-Cs) and spatially sep-\narated two-layered systems with magnetic dipole-dipole\ninteractions (Er-Er and Cr-Cr).\nWe separate each conductivity σij(ω) into the real part\nσn\nij(ω), which is responsible for dissipative response (cur-\nrent in phase with a driving force), and imaginary part\nσs\nij(ω), which corresponds to non-dissipative response\nwith the π/2 phase delay. Our analysis shows that, at\nfrequencies much lower than the characteristic energy\ngap ∆ between spin and density quasiparticle modes, the\nconductivities are well described by the two-fluid Drude9\nmodel [17, 18] where σn\nij(ω) exhibits the Drude peak\n∝[ω2+ 4γ2]−1like the normal metallic conductivity (in\nthe intracomponent channel i=j) or normal drag ef-\nfect (in the intercomponent channel i̸=j), while σs\nij(ω)\ndemonstrates the 1 /ωsingularity indicating superfluid-\nity (at i=j) or superfluid drag effect (at i̸=j). Thus\nour theory describes dissipative conductivity, superfluid-\nity, as well as normal and superfluid drag effects on equal\nfooting.\nAt higher frequencies near ω∼∆ the dissipative\npart of conductivity σn\nij(ω) exhibits peak, while non-\ndissipative part σsp\nij(ω) changes sign, which is qualita-\ntively similar to the Lorentz model of resonant response.\nHowever in our case the resonance shape is asymmetric\nand can essentially differ depending on the damping rate\nγand whether the atomic masses in a mixture are close\nto each other or distant enough. In a symmetric mixture\nwith equal masses, ∆ is close to the sum µa+µbof atomic\nchemical potentials, and the general case is considered in\nAppendix C. For two-layered quasi-2D system of dipolar\natoms, ∆ can be tuned by varying the interlayer separa-\ntionL. For 3D mixtures [Fig.1(a)], we derive the ana-\nlytical formulas which approximate the conductivities in\nboth Drude and Lorentz regimes rather accurately.\nThe drag effects predicted in our paper can be ob-\nserved in experiments with two-component or two-\nlayered atomic BECs by detecting currents arising in re-\nsponse to an alternating force, which selectively drives\none of the components (or drives them in opposite di-\nrections). The currents can be determined by measur-\ning atomic velocities via atomic cloud imaging after trap\nrelease or by time-of-flight measurements. The driving\nforce can be imposed by magnetic field gradients [38], op-\ntical lattices [39], magnetic trap shaking [20], or sudden\ndisplacement of optical trap [16]. Such methods can pro-\nvide oscillation frequencies up to several kHz, and achiev-\nable frequency ranges are often dictated by properties of\nthe atoms themselves [26].\nLet us estimate a magnetic field gradient required to in-\nduce strong enough oscillations, which could be observed\nby standard atomic cloud imaging. Consider the Yb-\nCs atomic mixture [29], where174Yb lacks magnetic mo-\nment, so its Lande factor is zero ( gF= 0), and thus\nonly133Cs is affected by magnetic field ( gF=−0.25\n[40]). In a homogeneous system the field gradient, re-\nquired to induce oscillations of the133Cs atomic cloud\nwith the amplitude xCsand frequency ω, is|∇B|=\nmCsω2xCs/mFgFµB, where mCsis the mass of133Cs\natom, mFis the magnetic sublevel of hyperfine state\nwith angular momentum F. Assuming the detectable\namplitude xCs∼10µm and using parameters from [29],\nwe obtain |∇B|(G/cm)≈[(ω/2π)(Hz)]2×10−4. The\ngradients up to 3000 G/cm used in experiments [41] are\nsufficient to create oscillations in both Drude ( ω/2π∼\n100 Hz, |∇B| ∼1 G/cm) and Lorentz ( ω/2π∼5000 Hz,\n|∇B| ∼2500 G/cm) regimes. The presence of harmonic\ntrap alters relationship between xCsand∇B, and we may\nhope to use the mechanical resonance effects to enhancethe oscillation amplitude even more.\nFor reliable detection of the drag effects, we need to\nachieve large enough amplitude xYbof oscillating motion\nof the174Yb atomic cloud in response to the magnetic\nfield gradient force applied to133Cs atoms. The ratio of\noscillation amplitudes can be estimated as xYb/xCs=\njYbnCs/jCsnYb= (nCs/nYb)× |σab(ω)/σaa(ω)|. At\nplausibly low damping rate γ/2π= 1 Hz, we obtain\nxYb/xCs∼0.01 at ω/2π= 100 Hz and xYb/xCs∼0.05\natω/2π= 5000 Hz. Such ratios are not restrictingly\nsmall, so we may hope to detect oscillations of the pas-\nsive174Yb component at high enough oscillating force\nand large enough oscillation amplitudes xCsof the active\ncomponent. For Rb-Rb mixture (see Fig. 5) this ratio\nis generally larger: xa/xb∼0.06 at low frequencies and\nxa/xb∼0.4 at the Lorentz-like resonance.\nPlane-parallel systems of magnetic dipole atoms [15,\n16] possess several additional controllable parameters:\nthickness of the clouds wz, intercloud separation L, and\ndipole moments orientation. The theory presented in our\npaper allows us to calculate the transconductivity be-\ntween dipolar atomic clouds in the setup of Ref. [16].\nHowever, direct comparison of our calculations with re-\nsults of this experiment is hindered because atomic gases\nin Ref. [16] were not Bose-condensed, and harmonic traps\nused to hold them made the atomic clouds inhomoge-\nneous and prone to mean-field repulsion not described by\nour theory. It is of interest to extend our approach to take\ninto account the normal-state drag diagrams [5, 33, 36]\nwhich would allow to describe the AC drag in wide tem-\nperature range both below and above Tc.\nAn alternative way to infer information about trans-\nand intraconductivities can rely on measuring tempera-\nture changes after several oscillations [20]. Mutual en-\ntrainment of two components can also affect dispersions\nand damping rates of first and second sounds in the two-\ncomponent BECs, which can be detected in sound veloc-\nity measurements [42]. Our approach of conductivity cal-\nculations is aimed on homogeneous systems correspond-\ning to flat traps [42–44]. In harmonic traps the resonance\nin center-of-mass motion of atomic clouds alters the be-\nhaviour of conductivity [20], and the mean-field repulsion\neffects mimicking intrinsic interlayer conductivity can ap-\npear [29, 45–47], so the problem of mutual entrainment\nbecomes more complicated.\nTo conclude, the theory of conductivities of Bose-\ncondensed two-component systems developed in this\npaper unifies calculations of the normal drag effect,\nAndreev-Bashkin effect, as well as intracomponent DC\nconductivity and superfluid density. Investigation of fre-\nquency dependencies of the conductivity tensor allows\nus to study interplay of dissipative and nondissipative\ncurrent responses. Our approach can be generalized for\nspin conductivity calculations [11, 26] and for coupled 1D\natomic gases [4]. Besides, similar AC entrainment effects,\nboth dissipative and nondissipative, can be expected in\nFermi-atom and condensed-matter superconducting sys-\ntems.10\nACKNOWLEDGMENTS\nThe work on analytical calculation and approximation\nof conductivities was done as a part of research Project\nNo. FFUU-2024-0003 of the Institute for Spectroscopy of\nthe Russian Academy of Sciences. The work on numeri-\ncal calculations was supported by the Program of Basic\nResearch of the Higher School of Economics.\nAppendix A: Green functions\nThe Hamiltonian of homogeneous two-component\natomic system is\nˆH=X\nipϵipa†\nipaip\n+1\n2AX\nijpp′qVij(q)a†\ni,p+qa†\nj,p′−qajp′aip, (A1)\nwhere aipis the destruction operator of the atomic par-\nticle of the component i=a, bwith momentum p,\nϵip=p2/2miis atomic dispersion, and Vij(q) is the\nFourier transform of the interaction between particles i\nandj. Replacing each zero-momentum operator ai,p=0\nby square root of the number of condensate particles\n(An0\ni)1/2, we obtain the mean-field Bogoliubov Hamil-\ntonian, which can be further diagonalized by the trans-\nformation\naip=u+\nidBdp+u−\nidB†\nd,−p+u+\nisBsp+u−\nisB†\ns,−p(A2)\ninto usual formP\np(EdB†\ndpBdp+EsB†\nspBsp), where Bdp,\nBspare destruction operators of density and spin quasi-\nparticles. Their energies read\nE2\nd,s=E2\na+E2\nb\n2±s\u0012E2a−E2\nb\n2\u00132\n+ 4ϵaϵbn0an0\nb|Vab|2,\n(A3)\nandEiis the energy of Bogoliubov excitation of isolated\nith component: Ei=p\nϵi(ϵi+ 2n0\niVii). The Bogoliubov\ntransformation coefficients are\nuζ\naα=ϵa+ζEα\n2√ϵaEαp\nPaα, uζ\nbα=±ϵb+ζEα\n2√ϵbEαp\nPbα,\n(A4)\nwhere\nPiα=±4n0\nan0\nb|Vab|2\n(E2\nd−E2s)(E2α−E2\ni)(A5)\nis the positive weight fraction of the ith component in\ntheαth quasiparticle mode. The upper and lower signs\nin Eqs. (A4)–(A5) correspond, respectively, to the den-\nsity ( α= d) and spin ( α= s) modes. To approximate the\ninteraction potentials Vij(q) in the case of 3D mixtures,\nwe assume the momentum-independent contact interac-\ntions gijrelated to s-wave scattering lengths.We define the matrix Green functions in the imaginary-\ntime domain as\nˆGij(p, τ) =−⟨Tτ \naip(τ)\na†\ni,−p(τ)!\u0010\na†\njp(0)aj,−p(0)\u0011\n⟩.\n(A6)\nIn a two-component Bose-condensed system, these func-\ntions can be found from the Dyson-Beliaev equations [48],\nand in the frequency domain they can be written as com-\nbinations\nˆGij(p, iωn) =X\nα=d,sX\ns=±s\niωn−sEα \nus\niα\nu−s\niα!\u0010\nus\njαu−s\njα\u0011\n(A7)\nof positive- and negative-frequency Green functions\n1/(iωn∓Eα) of Bogoliubov quasiparticles weighted with\nthe transformation coefficients (A4).\nThe density of non-condensate fraction of the ith com-\nponent can be calculated as nnc\ni=A−1P\np̸=0⟨a†\nipaip⟩.\nUsing the Bogoliubov transformation (A2) and taking\nthe thermal averages, we obtain\nnnc\ni=1\nAX\npX\nα=d,s\u001a\n(u−\niα)2+(u+\niα)2+ (u−\niα)2\neEα/T−1\u001b\n.(A8)\nAppendix B: Current response function\nWe define the Fourier harmonic operator of current as\nji(q) =m−1\niP\np(p+1\n2q)a†\nipai,p+q. In the simplest one-\nloop approximation, which is also used by other authors\nto describe the conductivity and superfluid drag effect in\nmulti-component ballistic systems [11, 21], the transverse\npart of the current response tensor (3) in Matsubara rep-\nresentation at q= 0 reads:\nχT\nij(0, iω) =−TX\npωnp2\n2Adm imj\n×Tr [σzˆGij(p, iωn+iω)σzˆGji(p, iωn)].(B1)\nUsing here the Green functions (A7), we obtain Eqs. (8)–\n(9) for the current response function. Separating terms\nwith α1=α2andα1̸=α2, we obtain\nχT\nij(0, iω) = Υ ij(iω)−(−1)δij\nmimj\u0002\nΛ+(iω) + Λ−(iω)\u0003\n.\n(B2)\nThe function Υ ij(ω), responsible for intra-branch scat-\ntering processes [see Fig. 2(c,d)], is defined as\nΥij(iω) =X\npX\nα=d,sp2PiαPjα\n2Adm imj\n×[S(Eα, Eα) +S(−Eα,−Eα)]. (B3)\nAfter introducing the quasiparticle damping and per-\nforming analytical continuation iω→ω+i0, we obtain11\nS(Eα, Eα) = 2 iγn′\nB(Eα)/(ω+ 2iγ) from Eq. (11), so this\nfunction can be written as\nΥij(ω) =X\npiγp2\nAdm imjPidPjdn′\nB(Ed) +PisPjsn′\nB(Es)\nω+ 2iγ\n=−2iγD0\nij\nω+ 2iγ, (B4)\nwhere we defined the conductivity weight (13).\nThe functions Λ+and Λ−are defined as\nΛ±(iω) =±X\npp2\n8Adp\nPidPisPjdPjs(Ed±Es)2\nEdEs\n×X\nα=d,s[S(Eα,±E˜α) +S(−Eα,∓E˜α)],(B5)\nwhere ˜d = s and ˜ s = d. Using the identity PadPas=\nPbdPbsfor the weight factors (A5), we obtain for 3D sys-\ntems the final expression (14) with the envelope functions\nf±(p) =±p4PidPis\n8π2d(Ed±Es)2\nEdEs\n× {nB(Ed)−nB(±Es)−iγ[n′\nB(Ed) +n′\nB(Es)]}.(B6)\nIn the Lorentz regime we omit the terms γn′\nB, because\nγn′\nB∼γ/T, which is much smaller than 1 in realistic\nsystems (since 1 nK ≈2π×138 Hz, so Tc∼102−103nK\ncorresponds to ∼105Hz). However, these terms should\nbe taken into account in the Drude regime: the con-\nductivity weights (13), (18) should be calculated as ac-\ncurately as possible, because their combined contribu-\ntion to the intercomponent superfluid weight Ds\nab=\nπ(−D0\nab+D+\nab+D−\nab) can be close to zero due to almost\ncomplete canceling of intra- and inter-Bogoliubov branch\nexcitation processes. For instance, for the Rb-Rb mixture\nwith γ/2π= 300 Hz and T=1\n3Tc, the Drude weights are\nD+\nab≈1.02D0\nabandD−\nab≈0.04D0\nab, and, consequently,\nDs\nab≈0.06πD0\nab. Therefore, even small errors in calcula-\ntions of D0,±\nijcan significantly affect conductivity in the\nlow-frequency limit.\nAppendix C: Approximations for ∆\nThe envelope functions (B6) increase at low momenta,\nwhen Ed,s< T, due to the power-law factor p4, and then\nexponentially decrease at large momenta, when Ed,s>\nT, thanks to the Bose distribution functions. Therefore,\nf±(p) reach maxima near some intermediate momentum\n¯pwhere Ed,s∼T. We restrict ourselves to the case when\nEd−Es≪Ed+Esand hence Ed≈Esnear p= ¯p,\nso that we are able to formally define ¯ pas a solution of\nequation Ed(¯p) +Es(¯p) = 2 T. In the parameter range\nwe consider, when T > µ a,b, the dispersions Ed,sare\nalmost quadratic near the momentum ¯ p. Using Eq. (A3)\nin the this quadratic regime, we can approximate it as\n¯p≈p\n2Tmamb/(ma+mb).To comprehend behaviour of the integrands in\nEq. (14), we should consider Ed+EsandEd−Esin\ndenominators of the R±functions (15) near p= ¯p.\nThe sum of energies, by definition, is about Ed(¯p) +\nEs(¯p) = 2 Tnear this momentum. The difference of\nenergies, denoted as ∆ = Ed(¯p)−Es(¯p), can be es-\ntimated using quadratic approximation of dispersions\n(A3): ∆ ≈p\n(E2a−E2\nb)2+ 16r2ϵaϵbµaµb/(ϵa+ϵb), where\nr2=g2\nab/gaagbbshould be less than 1 for stability of\nthe two-component BEC [23]. The first term under the\nsquare root can be rewritten as E2\na−E2\nb= (ϵ2\na−ϵ2\nb) +\n(2ϵaµa−2ϵbµb). It is straightforward to show, that ∆ ≈\n|ϵa−ϵb|when|ϵ2\na−ϵ2\nb| ≫ϵiµiatp= ¯p, which happens\nwhen masses of atoms are distant enough: such condition\ncan be written as |ma−mb|/mamb≳µi/Tm i. Other-\nwise, when masses are close to each other, |ϵ2\na−ϵ2\nb| ≪ϵiµi,\nwe obtain ∆ ≈p\nµ2a+µ2\nb+ 2(2−r2)µaµb≈µa+µb.\nOverall, near p= ¯p, where the envelope functions\nf±(p) attain the maximum, we obtain the following esti-\nmates for sum and difference of the quasiparticle energies:\nEd+Es∼2T, E d−Es∼∆, (C1)\nwhere\n∆∼\n\nT|ma−mb|\nma+mbif|ma−ma|\nmamb≳µi\nTmi\nµa+µb if|ma−ma|\nmamb≪µi\nTmi.(C2)\nThe first and second lines in Eq. (C2) correspond to the\ncase of distant and close masses, respectively. The sym-\nmetric mixtures ma=mbare obviously related to the\nsecond case. The aforementioned condition Ed−Es≪\nEd+Es, taken at the most relevant momenta p≈¯p, re-\nduces to ∆ ≪2T. In the case of distant masses it reads\n|ma−mb| ≪ma+mb(thus implying that the mass dif-\nference in this case is bounded both above and below),\nand in the case of close masses it is µa+µb≪T(which\nis fulfilled in the parameter ranges we consider).\nAppendix D: Interactions between dipolar atoms\nIn a system of magnetic dipolar atoms, total inter-\natomic interaction\nVij(r) =gijδ(r) +Vdd\nij(r), (D1)\nconsists of conventional isotropic interaction due to short-\nrange atomic scattering gijδ(r) and long-range magnetic\ndipole-dipole interaction\nVdd\nij(r) =didj1−3 cos2θ\n|r|3, (D2)\nwhere diis a magnetic dipole moment of the ith atomic\nspecie, and θis the angle between rand magnetic dipole\nmoments of all atoms which are assumed to be directed\nalong the zaxis.12\nWe consider quasi-two-dimensional atomic clouds with\nan effective thickness wz. In this case 2D Fourier trans-\nform of the full intracomponent (i.e. in the same planar\ncloud) interaction (D1) can be approximated as [49, 50]\nVii(q) =gii(1−Ci|q|), (D3)\nwhere Ci= 2πd2\niwz/gii. This interaction potential is\nevaluated with assumption r∗q≪1, where r∗=mid2\ni\nis the characteristic range of dipole-dipole interaction.\nThis assumption is valid for the parameters used in our\ncalculations: r∗= 12 nm and 2 .6 nm for168Er and52Cr\natoms respectively is much smaller than interlayer dis-\ntance L= 60 nm, which determines the scale of inverse\nmomentum q−1.\nThe Fourier transform of interaction between particles\nin different spatially separated atomic clouds is found as\nfollows. First, we rewrite the interaction (D2) for the\ntwo-layer geometry:\nVdd\nab(ρ, z−z′) =dadbr2−3(L+z−z′)2\nr5\n×\u0012π\n2wz\u00132\ncos\u0012πz\nwz\u0013\ncos\u0012πz′\nwz\u0013\n,(D4)\nwhere ρand L+z−z′are in-plane and out-\nof-plane distances between two atoms, while r=p\n(L+z−z′)2+ρ2is the total distance; zandz′are\ntheir vertical coordinates relative to the cloud centers\nranging from −wz/2 to wz/2. The cosine functions\nmodel atomic density profiles in the z-axis direction, and\n(π/2wz)2is normalization factor. 2D Fourier transform\nof Eq. (D4) in the xyplane reads\nVdd\nab(q, z, z′) =−2πdadbqe−q(L+z−z′)\n×\u0012π\n2wz\u00132\ncos\u0012πz\nwz\u0013\ncos\u0012πz′\nwz\u0013\n.(D5)\nWe assume thinness of atomic clouds, wz≪L(which was\nachieved in the recent experiment [16]), so out-of-plane\nmomenta of interacting particles are almost unchanged\nby the interlayer interaction. Therefore we can integrate\nthe interaction (D5) over zandz′, arriving at the formula\nVdd\nab(q) =−2πdadbqe−qL\u0014cosh( qwz/2)\n1 +q2w2z/π2\u00152\n, (D6)\nwhich is used in numerical calculations for Fig. 7. The\ntwo-dimensional condensate density of each component is\nestimated as n0\niwz, where the typical three-dimensional\ndensities ni=ζ(3/2)\u0000\nmiTi\nc/2π\u00013/2are related to the\ncritical temperatures Ti\nctaken from Table I.\n[1] C.-C. Chien, S. Peotta, and M. Di Ventra, Quantum\ntransport in ultracold atoms, Nature Physics 11, 998\n(2015).\n[2] S. D. Proukakis N., Universal Themes of Bose-Einstein\nCondensation , edited by P. Littlewood (Cambridge Uni-\nversity Press, 2017).\n[3] I. Bloch, J. Dalibard, and S. Nascimb` ene, Quantum sim-\nulations with ultracold quantum gases, Nature Physics\n8, 267 (2012).\n[4] S. Mistakidis, A. Volosniev, R. Barfknecht, T. Fogarty,\nT. Busch, A. Foerster, P. Schmelcher, and N. Zinner,\nFew-body Bose gases in low dimensions — A laboratory\nfor quantum dynamics, Physics Reports 1042 , 1 (2023).\n[5] B. N. Narozhny and A. Levchenko, Coulomb drag, Rev.\nMod. Phys. 88, 025003 (2016).\n[6] A. F. Andreev and E. P. Bashkin, Three-velocity hydro-\ndynamics of superfluid solutions, Sov. Phys. JETP 42,\n164 (1975).\n[7] J.-M. Duan and S. Yip, Supercurrent drag via the\nCoulomb interaction, Phys. Rev. Lett. 70, 3647 (1993).\n[8] B. Tanatar and A. K. Das, Drag effect for a bilayer\ncharged-Bose-gas system, Phys. Rev. B 54, 13827 (1996).\n[9] M. A. Alpar, S. A. Langer, and J. A. Sauls, Rapid post-\nglitch spin-up of the superfluid core in pulsars, Astrophys.\nJ.282, 533 (1984).\n[10] A. F. Aminov, A. A. Sokolik, and Y. E. Lozovik,\nSuperconductor-polariton nondissipative drag in optical\nmicrocavity, Phys. Rev. B 108, 115415 (2023).\n[11] Y. Sekino, H. Tajima, and S. Uchino, Spin conductivity\nspectrum and spin superfluidity in a binary Bose mixture,Physical Review Research 5, 023058 (2023).\n[12] C. A. M¨ uller and C. Gaul, Condensate deformation and\nquantum depletion of Bose–Einstein condensates in ex-\nternal potentials, New Journal of Physics 14, 075025\n(2012).\n[13] B. Nagler, M. Will, S. Hiebel, S. Barbosa, J. Koch,\nM. Fleischhauer, and A. Widera, Ultracold Bose gases\nin dynamic disorder with tunable correlation time, Phys.\nRev. Lett. 128, 233601 (2022).\n[14] S. Zhou, J. Chab´ e, R. Salem, T. David, D. Groswasser,\nM. Keil, Y. Japha, and R. Folman, Phase space tomog-\nraphy of cold-atom dynamics in a weakly corrugated po-\ntential, Phys. Rev. A 90, 033620 (2014).\n[15] C. Politi, A. Trautmann, P. Ilzh¨ ofer, G. Durastante, M. J.\nMark, M. Modugno, and F. Ferlaino, Interspecies inter-\nactions in an ultracold dipolar mixture, Physical Review\nA105, 023304 (2022).\n[16] L. Du, P. Barral, M. Cantara, J. de Hond, Y.-K. Lu,\nand W. Ketterle, Atomic physics on a 50 nm scale: Re-\nalization of a bilayer system of dipolar atoms (2023),\narXiv:2302.07209.\n[17] S. D. Brorson, R. Buhleier, I. E. Trofimov, J. O. White,\nC. Ludwig, F. F. Balakirev, H.-U. Habermeier, and\nJ. Kuhl, Electrodynamics of high-temperature supercon-\nductors investigated with coherent terahertz pulse spec-\ntroscopy, Journal of the Optical Society of America B 13,\n1979 (1996).\n[18] F. Yang and M. W. Wu, Gauge-invariant microscopic ki-\nnetic theory of superconductivity in response to electro-\nmagnetic fields, Physical Review B 98, 094507 (2018).13\n[19] K. L. Lee, N. B. Jørgensen, L. J. Wacker, M. G. Skou,\nK. T. Skalmstang, J. J. Arlt, and N. P. Proukakis, Time-\nof-flight expansion of binary Bose–Einstein condensates\nat finite temperature, New Journal of Physics 20, 053004\n(2018).\n[20] I. L. Garc´ ıa, B. Darqui´ e, C. D. J. Sinclair, E. A. Curtis,\nM. Tachikawa, J. J. Hudson, and E. A. Hinds, Shaking-\ninduced dynamics of cold atoms in magnetic traps, Phys-\nical Review A 88, 043406 (2013).\n[21] D. Romito, C. Lobo, and A. Recati, Linear response\nstudy of collisionless spin drag, Phys. Rev. Research. 3,\n023196 (2021).\n[22] D. J. Scalapino, S. R. White, and S. Zhang, Insulator,\nmetal, or superconductor: The criteria, Phys. Rev. B 47,\n7995 (1993).\n[23] D. V. Fil and S. I. Shevchenko, Nondissipative drag of\nsuperflow in a two-component Bose gas, Phys. Rev. A\n72, 013616 (2005).\n[24] A. Griffin, Excitations in a Bose-condensed liquid (Cam-\nbridge University Press, 1993).\n[25] D. C. Mattis and J. Bardeen, Theory of the anomalous\nskin effect in normal and superconducting metals, Phys.\nRev.111, 412 (1958).\n[26] Y. Sekino, H. Tajima, and S. Uchino, Optical spin con-\nductivity in ultracold quantum gases, Phys. Rev. Res. 4,\n043014 (2022).\n[27] E. Fava, T. Bienaim´ e, C. Mordini, G. Colzi, C. Qu,\nS. Stringari, G. Lamporesi, and G. Ferrari, Observation\nof spin superfluidity in a Bose gas mixture, Phys. Rev.\nLett. 120, 170401 (2018).\n[28] C. R. Cabrera, L. Tanzi, J. Sanz, B. Naylor, P. Thomas,\nP. Cheiney, and L. Tarruell, Quantum liquid droplets in\na mixture of Bose-Einstein condensates, Science 359, 301\n(2018).\n[29] K. E. Wilson, A. Guttridge, I.-K. Liu, J. Segal, T. P.\nBillam, N. G. Parker, N. P. Proukakis, and S. L. Cornish,\nDynamics of a degenerate Cs-Yb mixture with attractive\ninterspecies interactions, Phys. Rev. Research 3, 033096\n(2021).\n[30] L. E. Young-S. and S. K. Adhikari, Mixing, demixing,\nand structure formation in a binary dipolar Bose-Einstein\ncondensate, Physical Review A 86, 063611 (2012).\n[31] H. Pu and N. P. Bigelow, Collective excitations, metasta-\nbility, and nonlinear response of a trapped two-species\nBose-Einstein condensate, Physical Review Letters 80,\n1134 (1998).\n[32] P. A. Lee and T. V. Ramakrishnan, Disordered electronic\nsystems, Rev. Mod. Phys. 57, 287 (1985).\n[33] A. Kamenev and Y. Oreg, Coulomb drag in normal met-\nals and superconductors: Diagrammatic approach, Phys.\nRev. B 52, 7516 (1995).\n[34] D. Landau, E. M. Lifshits, and L. P. Pitaevskii, Statistical\nPhysics (Elsevier, 1980).[35] G. Baym, The microscopic description of superfluidity,\ninMathematical Methods in Solid State and Superfluid\nTheory: Scottish Universities’ Summer School , edited by\nR. C. Clark and G. H. Derrick (Springer US, Boston,\nMA, 1968) pp. 121–156.\n[36] A. F. Aminov, A. A. Sokolik, and Y. E. Lozovik, Su-\nperfluid drag between excitonic polaritons and supercon-\nducting electron gas, Quantum 6, 787 (2022).\n[37] N. Prokof’ev and B. Svistunov, Two-dimensional weakly\ninteracting Bose gas in the fluctuation region, Phys. Rev.\nA66, 043608 (2002).\n[38] K. Merloti, R. Dubessy, L. Longchambon, A. Perrin, P.-\nE. Pottie, V. Lorent, and H. Perrin, A two-dimensional\nquantum gas in a magnetic trap, New Journal of Physics\n15, 033007 (2013).\n[39] G. Gauthier, T. A. Bell, A. B. Stilgoe, M. Baker,\nH. Rubinsztein-Dunlop, and T. W. Neely, Chapter one\n- dynamic high-resolution optical trapping of ultracold\natoms (Academic Press, 2021) pp. 1–101.\n[40] D. A. Steck, Cesium D line data, available online at http:\n//steck.us/alkalidata (revision 2.3.2, 10 September\n2023).\n[41] V. Vuletic, T. W. H¨ ansch, and C. Zimmermann, Steep\nmagnetic trap for ultra cold atoms, Europhysics Letters\n36, 349 (1996).\n[42] T. A. Hilker, L. H. Dogra, C. Eigen, J. A. Glidden, R. P.\nSmith, and Z. Hadzibabic, First and second sound in a\ncompressible 3D Bose fluid, Physical Review Letters 128,\n223601 (2022).\n[43] Z. Yan, P. B. Patel, B. Mukherjee, C. J. Vale, R. J.\nFletcher, and M. Zwierlein, Thermography of the super-\nfluid transition in a strongly interacting Fermi gas (2022),\narXiv:2212.13752.\n[44] N. Navon, R. P. Smith, and Z. Hadzibabic, Quantum\ngases in optical boxes, Nature Physics 17, 1334 (2021).\n[45] C.-C. Huang and W.-C. Wu, Center motions of nonover-\nlapping condensates coupled by long-range dipolar inter-\naction in bilayer and multilayer stacks, Phys. Rev. A 82,\n053612 (2010).\n[46] M. V. Demin, Y. E. Lozovik, and V. A. Sharapov, Bose\ncondensate drag in a system of two coupled traps, JETP\nLett. 76, 135 (2002).\n[47] N. Matveeva, A. Recati, and S. Stringari, Dipolar drag in\nbilayer harmonically trapped gases, The European Phys-\nical Journal D 65, 219 (2011).\n[48] O. I. Utesov, M. I. Baglay, and S. V. Andreev, Effective\ninteractions in a quantum Bose-Bose mixture, Physical\nReview A 97, 053617 (2018).\n[49] A. Boudjemˆ aa and G. V. Shlyapnikov, Two-dimensional\ndipolar Bose gas with the roton-maxon excitation spec-\ntrum, Physical Review A 87, 025601 (2013).\n[50] A. Boudjemˆ aa, Two-dimensional quantum droplets in\ndipolar Bose gases, New Journal of Physics 21, 093027\n(2019)." }, { "title": "1803.10925v1.Giant_resonant_nonlinear_damping_in_nanoscale_ferromagnets.pdf", "content": "Giant resonant nonlinear damping in nanoscale ferromagnets\nI. Barsukov,1,\u0003H. K. Lee,1A. A. Jara,1Y.-J. Chen,1A. M. Gon\u0018 calves,1\nC. Sha,1J. A. Katine,2R. E. Arias,3B. A. Ivanov,4, 5and I. N. Krivorotov1\n1Physics and Astronomy, University of California, Irvine, CA 92697, USA\n2Western Digital, 5600 Great Oaks Parkway, San Jose, CA 95119, USA\n3Departamento de F\u0013 \u0010sica, CEDENNA, FCFM, Universidad de Chile, Santiago, Chile\n4Institute of Magnetism, National Academy of Sciences of Ukraine, Vernadsky av. 36 B, Kyiv, 03142, Ukraine\n5National University of Science and Technology MISiS, Moscow, 119049, Russian Federation\nMagnetic damping is a key metric for emerging technologies based on magnetic nanoparticles,\nsuch as spin torque memory and high-resolution biomagnetic imaging. Despite its importance,\nunderstanding of magnetic dissipation in nanoscale ferromagnets remains elusive, and the damping\nis often treated as a phenomenological constant. Here we report the discovery of a giant frequency-\ndependent nonlinear damping that strongly alters the response of a nanoscale ferromagnet to spin\ntorque and microwave magnetic \feld. This novel damping mechanism originates from three-magnon\nscattering that is strongly enhanced by geometric con\fnement of magnons in the nanomagnet. We\nshow that the giant nonlinear damping can invert the e\u000bect of spin torque on a nanomagnet leading\nto a surprising current-induced enhancement of damping by an antidamping torque. Our work\nadvances understanding of magnetic dynamics in nanoscale ferromagnets and spin torque devices.\nI. INTRODUCTION\nNanoscale magnetic particles are the core components\nof several emerging technologies such as nonvolatile spin\ntorque memory [1], spin torque oscillators [2{7], targeted\ndrug delivery, and high-resolution biomagnetic imaging\n[8{11]. Control of magnetic damping holds the key to\nimproving the performance of many nanomagnet-based\npractical applications. In biomagnetic characterization\ntechniques such as magnetic resonance imaging [12], re-\nlaxometry [13], and magnetic particle imaging [14, 15],\nmagnetic damping a\u000bects nanoparticles relaxation times\nand image resolution. In spin torque memory and oscil-\nlators, magnetic damping determines the electrical cur-\nrent necessary for magnetic switching [1] and generation\nof auto-oscillations [16] and thereby determines energy-\ne\u000eciency of these technologies. The performance of\nnanomagnet-based microwave detectors is also directly\na\u000bected by the damping [17{19]. Despite its impor-\ntance across multiple disciplines, magnetic damping in\nnanoparticles is poorly understood and is usually mod-\neled as a phenomenological constant [6, 16].\nIn this article, we experimentally demonstrate that a\nferromagnetic nanoparticle can exhibit dynamics quali-\ntatively di\u000berent from those predicted by the constant\ndamping model. We show that nonlinear contributions\nto the damping can be unusually strong and the damp-\ning parameter itself can exhibit resonant frequency de-\npendence. Our work demonstrates that nonlinear damp-\ning in nanomagnets is qualitatively di\u000berent from that in\nbulk ferromagnets and requires a new theoretical frame-\nwork for its description. We show both experimentally\nand theoretically that such resonant nonlinear damping\noriginates from multi-magnon scattering in a magnetic\n\u0003igorb@ucr.edusystem with a discrete spectrum of magnons induced by\ngeometric con\fnement.\nWe also discover that the resonant nonlinear damping\ndramatically alters the response of a nanomagnet to spin\ntorque. Spin torque arising from injection of spin cur-\nrents polarized opposite to the direction of magnetization\nacts as negative damping [2]. We \fnd, however, that the\ne\u000bect of such antidamping spin torque is reversed, lead-\ning to an enhanced dissipation due to the nonlinear res-\nonant scattering. This counterintuitive behavior should\nhave signi\fcant impact on the operation of spin torque\nbased memory [1], oscillators [2{7] and microwave detec-\ntors [17{19].\nII. RESULTS\nA. Spin wave spectroscopy\nWe study nonlinear spin wave dynamics in nanoscale\nelliptical magnetic tunnel junctions (MTJs) that consist\nof a CoFeB free layer (FL), an MgO tunnel barrier, and a\nsynthetic antiferromagnet (SAF) pinned layer [20]. Spec-\ntral properties of the FL spin wave modes are studied in a\nvariety of MTJs with both in-plane and perpendicular-to-\nplane equilibrium orientations of the FL and SAF magne-\ntization. We observe strong resonant nonlinear damping\nin both the in-plane and the perpendicular MTJs, which\npoints to the universality of the e\u000bect.\nWe employ spin torque ferromagnetic resonance (ST-\nFMR) to measure magnetic damping of the FL spin wave\nmodes. In this technique, a microwave drive current\nIacsin(2\u0019ft) applied to the MTJ excites oscillations of\nmagnetization at the drive frequency f. The resulting\nmagnetoresistance oscillations Racsin(2\u0019ft+\u001e) generate\na direct voltage Vmix. Peaks in ST-FMR spectra Vmix(f)\narise from resonant excitation of spin wave eigenmodes\nof the MTJ [21{28]. To improve signal-to-noise ratio,arXiv:1803.10925v1 [cond-mat.mes-hall] 29 Mar 20182\n0 0.5 1 1.5 2H1H2\n36912\nField (kOe)0 0.5 1 1.5 2H2H1\n00.20.4\nField (kOe)Linewidth (GHz)Experiment\nSimulationFrequency (GHz)\n+1\n1a b\n~\nFIG. 1. Spin wave spectra in a nanoscale MTJ. (a) Normalized ST-FMR spectra h~Vmix(f)iof spin wave eigenmodes in a\nperpendicular MTJ device (Sample 1) measured as a function of out-of-plane magnetic \feld. Resonance peaks arising from\nthree low frequency modes of the MTJ free layer j0i,j1i, and j2iare observed. (b) Spectral linewidth of the quasi-uniform\nj0ispin wave mode as a function of out-of-plane magnetic \feld. Strong linewidth enhancement is observed in the resonant\nthree-magnon regime at H1andH2.\nthe magnitude of external magnetic \feld Happlied par-\nallel to the free layer magnetization is modulated, and\na \feld-derivative signal ~Vmix(f) = dVmix(f)=dHis mea-\nsured via lock-in detection technique [20]. Vmix(f) can\nthen be obtained via numerical integration (Supplemen-\ntal Material).\nFigure 1(a) shows ST-FMR spectra ~Vmix(f) measured\nas a function of out-of-plane magnetic \feld Hfor an el-\nliptical 52 nm\u000262 nm perpendicular MTJ device (Sam-\nple 1). Three spin wave eigenmodes with nearly linear\nfrequency-\feld relation fn(H) are clearly visible in the\nspectra. Micromagnetic simulations (Supplemental Ma-\nterial) reveal that these modes are three lowest frequency\nspin wave eigenmodes of the FL (Supplemental Material).\nThe lowest frequency (quasi-uniform) mode j0iis node-\nless and has spatially uniform phase. Each of the two\nhigher-order modes jni(n= 1;2) has a single node at\nthe FL center that is either perpendicular ( n= 1) or\nparallel (n= 2) to the ellipse long axis.\nThe spectral linewidth of the resonances in Fig. 1(a)\ncan be used for evaluation of the mode damping. The\nquasi-uniform mode j0iresonance visibly broadens at\ntwo magnetic \feld values: H1= 0:74 kOe (4 GHz) and\nH2= 1:34 kOe (6 GHz). Near H1, the modej1iresonance\nalso broadens and exhibits splitting, same behavior is ob-\nserved for the mode j2iatH2. At these \felds, the higher-\norder mode frequency is twice that of the quasi-uniform\nmodefn= 2f0. This shows that three-magnon con\ru-\nence [29{33] is the mechanism of the quasi-uniform mode\ndamping increase: two magnons of the quasi-uniform\nmodej0imerge into a single magnon of the higher-ordermodejni.\nThe most striking feature of the quasi-uniform mode\nresonance near H1is its split-peak shape with a local min-\nimum at the resonance frequency. Such a lineshape can-\nnot be \ft by the standard Lorentzian curve with symmet-\nric and antisymmetric components [20]. We therefore use\na double-peak \ftting function (Supplemental Material)\nto quantify the e\u000bective linewidth \u0001 f0of the resonance\npro\fle. For applied \felds su\u000eciently far from H1, the\nST-FMR curve recovers its single-peak shape and \u0001 f0\nis determined as half width of the standard Lorentzian\n\ftting function [20]. Figure 1(b) shows \u0001 f0as a function\nofHand demonstrates a large increase of the linewidth\nnear the \felds of the resonant three-magnon regime H1\nandH2. The stepwise increase of \u0001 f0nearH1is a result\nof the ST-FMR curve transition between the split-peak\nand single-peak shapes. For \felds near H2, the resonance\npro\fle broadens but does not develop a visible split-peak\nlineshape. As a result, \u0001 f0(H) is a smooth function in\nthe vicinity of H2.\nB. E\u000bect of spin torque\nIn MTJs, direct bias current Idcapplied across the\njunction exerts spin torque on the FL magnetization, act-\ning as antidamping for Idc>0 and as positive damping\nforIdc<0 [22, 34]. The antidamping spin torque in-\ncreases the amplitude of the FL spin wave modes [22, 35]\nand decreases their spectral linewidth [36]. We can em-\nploy spin torque from Idcto control the amplitude of spin3\nFIG. 2. E\u000bect of spin torque on spin wave resonance lineshape. (a)-(b) Spin wave resonance lineshapes in the nonresonant\nregime at H > H 1for di\u000berent values of direct bias current Idc. (c)-(d) Spin wave resonance lineshapes in the resonant three-\nmagnon regime at H=H1. (a), (c) Measured ST-FMR spectra (Sample 2). (b), (d) Solutions of Eqs. (3) and (4). Identical\nbias current values Idc(displayed in (a) are used in (a)-(d).\nwave eigenmodes excited in ST-FMR measurements, and\nthereby study the crossover between linear and nonlinear\nregimes of spin wave resonance.\nFigure 2 shows the dependence of ST-FMR resonance\ncurve of thej0imodeVmix(f) onIdcfor a 50 nm\u0002110 nm\nelliptical in-plane MTJ (Sample 2). For in-plane mag-\nnetic \feld values far from the three-magnon resonance\n\feldsHn, the amplitude of ST-FMR resonance curve\nVmix(f) shown in Fig. 2(a) monotonically increases with\nincreasing antidamping spin torque, as expected. At\nH=H1, the antidamping spin torque has a radically\ndi\u000berent and rather surprising e\u000bect on the resonance\ncurve. As illustrated in Fig. 2(c), increasing antidamp-\ning spin torque \frst broadens the resonance at H=H1\nand then transforms a single-peak resonance lineshape\ninto a split-peak lineshape with a local minimum at the\nresonance frequency f0. The data in Fig. 2 demonstrate\nthat the unusual split-peak lineshape of the resonance is\nonly observed when (i) the three-magnon scattering of\nthe quasi-uniform mode is allowed by the conservation of\nenergy and (ii) the amplitude of the mode is su\u000eciently\nhigh, con\frming that the observed e\u000bect is resonant and\nnonlinear in nature.\nFig. 2(c) reveals that antidamping spin torque can in-\ncrease the spectral linewidth and the e\u000bective damping\nof the quasi-uniform spin mode if the mode undergoes\nresonant three-magnon scattering. Figure 3 further illus-\ntrates this counterintuitive e\u000bect. It shows the linewidth\nof the quasi-uniform mode of a 50 nm \u0002110 nm elliptical\nin-plane MTJ (Sample 3) measured as a function of bias\ncurrent. In Fig. 3, blue symbols show the linewidth mea-\nsured at an in-plane magnetic \feld su\u000eciently far fromthe three-magnon resonance \felds Hn. At this \feld, the\nexpected quasi-linear dependence of the linewidth on Idc\nis observed for currents well below the critical current\nfor the excitation of auto-oscillatory magnetic dynamics.\nNear the critical current, the linewidth increases due to\na combination of the fold-over e\u000bect [37{39] and ther-\nmally activated switching between the large- and small-\namplitude oscillatory states of the fold-over regime [22].\nThe red symbols in Fig. 3 show the linewidth measured\nin the resonant three-magnon regime at H=H1. In con-\ntrast to the nonresonant regime, the linewidth increases\nwith increasingjIdcjfor both current polarities. Fur-\nthermore, the maximum linewidth is measured for the\nantidamping current polarity.\nIII. THEORETICAL MODEL\nNonlinear interactions among spin wave eigenmodes\nof a ferromagnet give rise to a number of spectacu-\nlar magneto-dynamic phenomena such as Suhl instabil-\nity of the uniform precession of magnetization [40, 41],\nspin wave self-focusing [42] and magnetic soliton forma-\ntion [43{45]. In bulk ferromagnets, nonlinear interac-\ntions generally couple each spin wave eigenmode to a\ncontinuum of other modes via energy- and momentum-\nconserving multi-magnon scattering [40]. This kinemat-\nically allowed scattering limits the achievable amplitude\nof spin wave modes and leads to broadening of the spin\nwave resonance. These processes lead to a resonance\nbroadening [40, 46{48] and cannot explain the observed\nsplit-peak lineshape of the resonance. In nanoscale ferro-4\nmagnets, geometric con\fnement discretizes the spin wave\nspectrum and thereby generally eliminates the kinemati-\ncally allowed multi-magnon scattering. This suppression\nof nonlinear scattering enables persistent excitation of\nspin waves with very large amplitudes [49] as observed in\nnanomagnet-based spin torque oscillators [2, 50]. Tun-\nability of the spin wave spectrum by external magnetic\n\feld, however, can lead to a resonant restoration of the\nenergy-conserving scattering [31]. The description of\nnonlinear spin wave resonance in the nanoscale ferromag-\nnet geometry therefore requires a new theoretical frame-\nwork. To derive the theory of resonant nonlinear damp-\ning in a nanomagnet, we start with a model Hamilto-\nnian that explicitly takes into account resonant nonlinear\nscattering between the quasi-uniform mode and a higher-\norder spin wave mode (in reduced units with ~\u00111):\nH=!0aya+!nbyb+\t0\n2ayayaa+\tn\n2bybybb (1)\n+( naaby+ \u0003\nnayayb)\n+\u0010\b\nexp(\u0000i!t)ay+ exp(i!t)a\t\nwhereay,aandby,bare the magnon creation and an-\nnihilation operators for the quasi-uniform mode j0iwith\nfrequency!0and for the higher-order spin wave mode\njnimode with frequency !n, respectively. The non-\nlinear mode coupling term proportional to the coupling\nstrength parameter ndescribes the annihilation of two\nj0imagnons and creation of one jnimagnon, as well as\nthe inverse process. The Hamiltonian is written in the\nresonant approximation, where small nonresonant terms\nsuch asaab,aaayare neglected. The terms proportional\nto \t 0and \t ndescribe the intrinsic nonlinear frequency\nshifts [51] of the modes j0iandjni. The last term de-\nscribes the excitation of the quasi-uniform mode by an\nexternal ac drive with the amplitude \u0010and frequency !.\nWe further de\fne classically a dissipation function Q,\nwhere\u000b0and\u000bnare the intrinsic linear damping param-\neters of the modes j0iandjni[52{54]:\nQ=day\ndtda\ndt(\u000b0+\u00110aya) +dby\ndtdb\ndt(\u000bn+\u0011nbyb) (2)\nFor generality, Eq. (2) includes intrinsic nonlinear\ndamping [16] of the modes j0iandjnidescribed by the\nnonlinearity parameters \u00110and\u0011n. However, our analy-\nsis below shows that the split-peak resonance lineshape\nis predicted by our theory even if \u00110and\u0011nare set equal\nto zero.\nEquations describing the nonlinear dynamics of the\ntwo coupled spin wave modes of the system follow from\nEq. (1) and Eq. (2):\nida\ndt=@H\n@ay+@Q\n@(day=dt)(3)\nidb\ndt=@H\n@by+@Q\n@(dby=dt)(4)\nIt can be shown (Supplemental Material) that these\nequations have a periodic solution a= \u0016aexp (\u0000i!t) and\n100 50 0 50 10000.10.20.30.40.5\nI (A)f0 (GHz)Resonant regime\nNonresonant regimeFIG. 3. E\u000bect of spin torque on linewidth. Linewidth of the\nquasi-uniform spin wave mode as a function of the applied\ndirect bias current (Sample 3): blue symbols { in the non-\nresonant regime H6=H1and red symbols { in the resonant\nthree-magnon regime H=H1. Lines are numerical \fts using\nEqs. (3) and (4).\nb=\u0016bexp (\u0000i2!t), where \u0016a,\u0016bare the complex spin wave\nmode amplitudes. For such periodic solution, Eqs. (3)\nand (4) are reduced to a set of two nonlinear algebraic\nequations for absolute values of the spin wave mode am-\nplitudesj\u0016ajandj\u0016bj, which can be solved numerically.\nSince the ST-FMR signal is proportional to j\u0016aj2(Supple-\nmental Material), the calculated j\u0016aj2(!) function can be\ndirectly compared to the measured ST-FMR resonance\nlineshape.\nWe employ the solution of Eqs. (3) and (4) to \ft the\n\feld dependence of the quasi-uniform mode linewidth in\nFig. 1(b). In this \ftting procedure, the resonance line-\nshapej\u0016aj2(!) is calculated, and its spectral linewidth\n\u0001!0is found numerically. The resonance frequencies !0\nand!nare directly determined from the ST-FMR data\nin Fig. 1(a). The intrinsic damping parameters \u000b0and\n\u000bnnearH1andH2are found from linear interpolations\nof the ST-FMR linewidths \u0001 f0and \u0001fnmeasured at\n\felds far from H1andH2. We \fnd that \u0001 !0weakly\ndepends on the nonlinearity parameters \t and \u0011, and\nthus these parameters are set to zero (Supplemental Ma-\nterial). We also \fnd that the calculated linewidth \u0001 !0\ndepends on the product of the drive amplitude \u0010and\nmode coupling strength n, but is nearly insensitive to\nthe individual values of \u0010and nas long as\u0010\u0001 n= const\n(Supplemental Material). Therefore, we use \u0010\u0001 nas a\nsingle \ftting parameter in this \ftting procedure. Solid\nline in Fig. 1(b) shows the calculated \feld dependence\nof the quasi-uniform mode linewidth on magnetic \feld.\nThe agreement of this single-parameter \ft with the ex-\nperiment is excellent.\nFigures 2(b) and 2(d) illustrate that Eqs. (3) and\n(4) not only describe the \feld dependence of ST-FMR\nlinewidth but also qualitatively reproduce the spectral5\nlineshapes of the measured ST-FMR resonances as well\nas the e\u000bect of the antidamping spin torque on the line-\nshapes. Fig. 2(b) shows the dependence of the calculated\nlineshapej\u0016aj2(!) on antidamping spin torque for a mag-\nnetic \feldHfar from the three-magnon resonance \felds\nHn. At this nonresonant \feld, increasing antidamping\nspin torque induces the fold-over of the resonance curve\n[37] without resonance peak splitting. The dependence of\nj\u0016aj2(!) on antidamping spin torque for H=H1is shown\nin Fig. 2(d). At this \feld, the resonance peak in j\u0016aj2(!)\n\frst broadens with increasing antidamping spin torque\nand then splits, in qualitative agreement with the ex-\nperimental ST-FMR data in Fig. 2(c). Our calculations\n(Supplemental Material) reveal that while the nonlinear-\nity parameters \t 0,\u00110, \tnand\u0011nhave little e\u000bect on\nthe linewidth \u0001 !0, they modify the lineshape of the res-\nonance. Given that the nonlinearity parameter values\nare not well known for the systems studied here, we do\nnot attempt to quantitatively \ft the measured ST-FMR\nlineshapes.\nEquations (3) and (4) also quantitatively explain\nthe observed dependence of the quasi-uniform mode\nlinewidth \u0001 !0on direct bias current Idc. Assuming an-\ntidamping spin torque linear in bias current [36, 55, 56]:\n\u000b0!\u000b0(1\u0000Idc=Ij0i\nc),\u000bn!\u000bn(1\u0000Idc=Ijni\nc), where\nIjni\nc>Ij0i\ncare the critical currents, we \ft the measured\nbias dependence of ST-FMR linewidth in Fig. 3 by solv-\ning Eqs. (3) and (4). The solid lines in Fig. 3 are the best\nnumerical \fts, where \u0010\u0001 nandIcare used as indepen-\ndent \ftting parameters. The rest of the parameters in\nEqs. (3) and (4) are directly determined from the experi-\nment following the procedure used for \ftting the data in\nFig. 1(b). Theoretical curves in Fig. 3 capture the main\nfeature of the data at the three-magnon resonance \feld\nH1{ increase of the linewidth with increasing antidamp-\ning spin torque.\nIV. DISCUSSION\nFurther insight into the mechanisms of the nonlinear\nspin wave resonance peak splitting and broadening by an-\ntidamping spin torque can be gained by neglecting the in-\ntrinsic nonlinearities \t nand\u0011nof the higher-order mode\njni. Setting \t n= 0 and\u0011n= 0 in Eqs. (3) and (4) allows\nus to reduce the equation of motion for the quasi-uniform\nmode amplitudej\u0016ajto the standard equation for a single-\nmode damped driven oscillator (Supplemental Material)\nwhere a constant damping parameter \u000b0is replaced by\nan e\u000bective frequency-dependent nonlinear damping pa-\nrameter\u000be\u000b\n0:\n\u000be\u000b\n0=\u000b0+\u0014\n\u00110+4\u000bn 2\nn\n(2!\u0000!n)2+ 4\u000b2n!2\u0015\nj\u0016aj2(5)and the resonance frequency is replaced by an e\u000bective\nresonance frequency:\n!e\u000b\n0=!0+\u0014\n\t0+2j nj2(2!\u0000!n)\n(2!\u0000!n)2+ 4\u000b2n!2\u0015\nj\u0016aj2(6)\nEquation (5) clearly shows that the damping parame-\nter of the quasi-uniform mode itself becomes a resonant\nfunction of the drive frequency with a maximum at half\nthe frequency of the higher order mode ( !=1\n2!n). The\namplitude and the width of this resonance in \u000be\u000b\n0(!) are\ndetermined by the intrinsic damping parameter \u000bnof\nthe higher-order mode jni. If\u000bnis su\u000eciently small,\nthe quasi-uniform mode damping is strongly enhanced\nat!=1\n2!n, which leads to a decrease of the quasi-\nuniform mode amplitude at this drive frequency. If the\ndrive frequency is shifted away from1\n2!nto either higher\nor lower values, the damping decreases, which can re-\nsult in an increase of the quasi-uniform mode amplitude\nj\u0016aj. Therefore, the amplitude of the quasi-uniform mode\nj\u0016aj(!) can exhibit a local minimum at !=1\n2!n. Due to\nits nonlinear origin, the tendency to form a local min-\nimum inj\u0016aj(!) at1\n2!nis enhanced with increasing j\u0016aj.\nSincej\u0016ajis large near the resonance frequency !0, tun-\ning!0to be equal to1\n2!ngreatly ampli\fes the e\u000bect of\nlocal minimum formation in j\u0016aj(!). This qualitative ar-\ngument based on Equation (5) explains the data in Fig. 2\n{ the split-peak nonlinear resonance of the quasi-uniform\nmode is only observed when external magnetic \feld tunes\nthe spin wave eigenmode frequencies to the three-magnon\nresonance condition !0=1\n2!n.\nEquation (6) reveals that the nonlinear frequency shift\nof the quasi-uniform mode is also a resonant function of\nthe drive frequency. In contrast to the nonlinear damping\nresonance described by Equation (5), the frequency shift\nresonance is an antisymmetric function of !\u00001\n2!n. The\nnonlinear shift is negative for ! <1\n2!nand thus causes\na fold-over towards lower frequencies while it is positive\nfor!>1\n2!ncausing fold-over towards higher frequencies.\nAt the center of the resonance pro\fle, the three-magnon\nprocess induces no frequency shift. This double-sided\nfold-over also contributes to the formation of the split-\npeak lineshape of the resonance shown in Figs. 2(c) and\n2(d) and to the linewidth broadening. As with the non-\nlinear damping resonance, the antisymmetric nonlinear\nfrequency shift and the double-sided fold-over become\ngreatly ampli\fed when the spin wave mode frequencies\nare tuned near the three-magnon resonance !0=1\n2!n.\nEquations (5) and (6) also shed light on the origin\nof the quasi-uniform mode line broadening by the an-\ntidamping spin torque. The antidamping spin torque in-\ncreases the quasi-uniform mode amplitude j\u0016ajvia transfer\nof angular momentum from spin current to the mode [57].\nSince the nonlinear damping and the nonlinear frequency\nshift are both proportional to j\u0016aj2and both contribute to\nthe line broadening, the antidamping spin torque can in-\ndeed give rise to the line broadening. Equation (5) reveals\ntwo competing e\u000bects of the antidamping spin torque on\nthe quasi-uniform mode damping parameter \u000be\u000b\n0: spin6\ntorque from Idcdecreases the linear component of the\ndamping parameter \u000b0!\u000b0(1\u0000Idc=Ij0i\nc) and increases\nthe nonlinear component via increased j\u0016aj2. Whether the\nantidamping spin torque decreases or increases the spec-\ntral linewidth of the mode depends on the system param-\neters. Our numerical solution of Eqs. (3) and (4) shown\nin Fig. 3 clearly demonstrates that the antidamping spin\ntorque can strongly increase the linewidth of the quasi-\nuniform mode when the three-magnon resonance condi-\ntion!0=1\n2!nis satis\fed. Furthermore, we \fnd that\nthe three-magnon process exhibits no threshold behav-\nior upon increasing amplitude (Supplemental Material)\nor decreasing intrinsic damping.\nThe key requirement for observation of the resonant\nnonlinear damping is the discreteness of the magnon\nspectrum imposed by geometric con\fnement in the\nnanoscale ferromagnet. The split-peak nonlinear reso-\nnance discovered in this work cannot be realized in bulk\nferromagnets because the three-magnon resonance con-\ndition in bulk is not only valid at the uniform mode\nfrequency!0=1\n2!nbut instead in a broad frequency\nrange. Owing to the magnon spectrum continuity in\nbulk, shifting the excitation frequency away from !0does\nnot suppress the three-magnon scattering of the uniform\nmode { it simply shifts it from one group of magnons to\nanother [29, 40]. Therefore, the amplitude of the uni-\nform mode does not increase when the drive frequency is\nshifted away from !0and the split-peak resonance is not\nrealized.\nWe expect that the resonant nonlinear damping dis-\ncovered in this work will have strong impact on the\nperformance of spin torque devices such as spin torque\nmagnetic memory, spin torque nanooscillators and spin\ntorque microwave detectors. Since all these devices rely\non large-amplitude oscillations of magnetization driven\nby spin torque, the amplitude limiting resulting from the\nresonant nonlinear damping is expected to have detri-\nmental e\u000bect on the device performance.V. CONCLUSIONS\nIn conclusion, our measurements demonstrate that\nmagnetic damping of spin wave modes in a nanoscale\nferromagnet has a strong nonlinear component of reso-\nnant character that appears at a discrete set of magnetic\n\felds corresponding to resonant three-magnon scattering.\nThis strong resonant nonlinearity can give rise to unusual\nspin wave resonance pro\fle with a local minimum at the\nresonance frequency in sharp contrast to the properties\nof the linear and nonlinear spin wave resonances in bulk\nferromagnets. The resonant nonlinearity has a profound\ne\u000bect on the response of the nanomagnet to spin torque.\nAntidamping spin torque, that reduces the quasi-uniform\nspin wave mode damping at magnetic \felds far from the\nresonant three-magnon regime, can strongly enhance the\ndamping in the resonant regime. This inversion of the\ne\u000bect of spin torque on magnetization dynamics by the\nresonant nonlinearity is expected to have signi\fcant im-\npact on the performance of nanoscale spin torque devices\nsuch as magnetic memory and spin torque oscillators.\nACKNOWLEDGMENTS\nThis work was supported by the National Science\nFoundation through Grants No. DMR-1610146, No.\nEFMA-1641989 and No. ECCS-1708885. We also ac-\nknowledge support by the Army Research O\u000ece through\nGrant No. W911NF-16-1-0472 and Defense Threat Re-\nduction Agency through Grant No. HDTRA1-16-1-0025.\nA. M. G. thanks CAPES Foundation, Ministry of Educa-\ntion of Brazil for \fnancial support. R.E.A acknowledges\nFinanciamiento Basal para Centros Cienti\fcos y Tec-\nnologicos de Excelencia under project FB 0807 (Chile),\nand Grant ICM P10-061-F by Fondo de Innovacion para\nla Competitividad-MINECON. B.A.I. was supported by\nthe National Academy of Sciences of Ukraine via project\n#1/17-N and by the Program of NUST \"MISiS\" (grant\nNo. K2-2017-005), implemented by a governmental de-\ncree dated 16th of March 2013, No. 211.\n[1] Luqiao Liu, Chi-Feng Pai, Y. Li, H. W. Tseng, D. C.\nRalph, and R. A. Buhrman, \\Spin-torque switching with\nthe giant spin Hall e\u000bect of tantalum,\" Science 336, 555{\n558 (2012).\n[2] S. Il. Kiselev, J.C. Sankey, I.N. Krivorotov, N.C. Emley,\nR.J. Schoelkopf, R.A. Buhrman, and D.C. Ralph, \\Mi-\ncrowave oscillations of a nanomagnet driven by a spin-\npolarized current,\" Nature 425, 380{383 (2003).\n[3] William H. Rippard, Alina M. Deac, Matthew R. Pufall,\nJustin M. Shaw, Mark W. Keller, Stephen E. Russek,\nGerrit E. W. Bauer, and Claudio Serpico, \\Spin-transfer\ndynamics in spin valves with out-of-plane magnetized\nCoNi free layers,\" Phys. Rev. B 81, 014426 (2010).[4] D. Houssameddine, U. Ebels, B. Delat, B. Rodmacq,\nI. Firastrau, F. Ponthenier, M. Brunet, C. Thirion, J.-\nP. Michel, L. Prejbeanu-Buda, M.-C. Cyrille, O. Redon,\nand B. Dieny, \\Spin-torque oscillator using a perpendic-\nular polarizer and a planar free layer,\" Nat. Mater. 6,\n447{453 (2007).\n[5] A. Houshang, E. Iacocca, P. Drrenfeld, S. R. Sani, J. ker-\nman, and R. K. Dumas, \\Spin-wave-beam driven syn-\nchronization of nanocontact spin-torque oscillators,\" Nat.\nNanotech. 11, 280{286 (2016).\n[6] Vladislav E. Demidov, Sergei Urazhdin, Henning Ul-\nrichs, Vasyl Tiberkevich, Andrei Slavin, Dietmar Baither,\nGuido Schmitz, and Sergej O. Demokritov, \\Magnetic\nnano-oscillator driven by pure spin current,\" Nat. Mater.7\n11, 1028{1031 (2012).\n[7] Ferran Maci\u0012 a, Dirk Backes, and Andrew D. Kent, \\Sta-\nble magnetic droplet solitons in spin-transfer nanocon-\ntacts,\" Nat. Nanotech. 9, 992{996 (2014).\n[8] Zhiliang Cheng, Ajlan Al Zaki, James Z. Hui, Vladimir R.\nMuzykantov, and Andrew Tsourkas, \\Multifunctional\nnanoparticles: Cost versus bene\ft of adding targeting\nand imaging capabilities,\" Science 338, 903{910 (2012).\n[9] Edward Kai-Hua Chow and Dean Ho, \\Cancer\nnanomedicine: From drug delivery to imaging,\" Science\nTransl. Med. 5, 216rv4 (2013).\n[10] Miriam Colombo, Susana Carregal-Romero, Maria F.\nCasula, Lucia Gutierrez, Maria P. Morales, Ingrid B.\nBohm, Johannes T. Heverhagen, Davide Prosperi, and\nWolfgang. J. Parak, \\Biological applications of magnetic\nnanoparticles,\" Chem. Soc. Rev. 41, 4306{4334 (2012).\n[11] Michael Bietenbeck, Anca Florian, Cornelius Faber, Udo\nSechtem, and Ali Yilmaz, \\Remote magnetic targeting\nof iron oxide nanoparticles for cardiovascular diagnosis\nand therapeutic drug delivery: where are we now?\" Intl.\nJ. Nanomed. 11, 3191 (2016).\n[12] Huilin Shao, Tae-Jong Yoon, Monty Liong, Ralph\nWeissleder, and Hakho Lee, \\Magnetic nanoparticles\nfor biomedical nmr-based diagnostics,\" Beilstein J. Nan-\notech. 1, 142 (2010).\n[13] P.W. Goodwill, A. Tamrazian, L.R. Croft, C.D. Lu, E.M.\nJohnson, R. Pidaparthi, R.M. Ferguson, A.P. Khandhar,\nK.M. Krishnan, and S.M. Conolly, \\Ferrohydrodynamic\nrelaxometry for magnetic particle imaging,\" Appl. Phys.\nLett. 98, 262502 (2011).\n[14] Patrick William Goodwill, Emine Ulku Saritas,\nLaura Rose Croft, Tyson N. Kim, Kannan M. Kr-\nishnan, David V. Scha\u000ber, and Steven M. Conolly,\n\\X-space mpi: Magnetic nanoparticles for safe medical\nimaging,\" Adv. Mater. 24, 3870{3877 (2012).\n[15] R. Matthew Ferguson, Kevin R. Minard, Amit P. Khand-\nhar, and Kannan M. Krishnan, \\Optimizing magnetite\nnanoparticles for mass sensitivity in magnetic particle\nimaging,\" Med. Phys. 38, 1619{1626 (2011).\n[16] Andrei Slavin and Vasil Tiberkevich, \\Nonlinear auto-\noscillator theory of microwave generation by spin-\npolarized current,\" IEEE Trans. Magn. 45, 1875{1918\n(2009).\n[17] Jian Zhu, J. A. Katine, Graham E. Rowlands, Yu-Jin\nChen, Zheng Duan, Juan G. Alzate, Pramey Upad-\nhyaya, Juergen Langer, Pedram Khalili Amiri, Kang L.\nWang, and Ilya N. Krivorotov, \\Voltage-induced ferro-\nmagnetic resonance in magnetic tunnel junctions,\" Phys.\nRev. Lett. 108, 197203 (2012).\n[18] S. Miwa, S. Ishibashi, H. Tomita, T. Nozaki, E. Tamura,\nK. Ando, N. Mizuochi, T. Saruya, H. Kubota,\nK. Yakushiji, T. Taniguchi, H. Imamura, A. Fukushima,\nS. Yuasa, and Y. Suzuki, \\Highly sensitive nanoscale\nspin-torque diode,\" Nat. Mater. 13, 50{56 (2014).\n[19] Bin Fang, Mario Carpentieri, Xiaojie Hao, Hongwen\nJiang, Jordan A. Katine, Ilya N. Krivorotov, Berthold\nOcker, Juergen Langer, Kang L. Wang, Baoshun Zhang,\nBruno Azzerboni, Pedram Khalili Amiri, Giovanni Finoc-\nchio, and Zhongming Zeng, \\Giant spin-torque diode\nsensitivity in the absence of bias magnetic \feld,\" Nat.\nCommun. 7, 11259 (2016).\n[20] A. M. Gon\u0018 calves, I. Barsukov, Y.-J. Chen, L. Yang, J. A.\nKatine, and I. N. Krivorotov, \\Spin torque ferromagnetic\nresonance with magnetic \feld modulation,\" Appl. Phys.Lett. 103, 172406 (2013).\n[21] A.A. Tulapurkar, Y. Suzuki, A. Fukushima, H. Kub-\nota, H. Maehara, K. Tsunekawa, D.D. Djayaprawira,\nN. Watanabe, and S. Yuasa, \\Spin-torque diode e\u000bect in\nmagnetic tunnel junctions,\" Nature 438, 339{342 (2005).\n[22] J. C. Sankey, P. M. Braganca, A. G. F. Garcia, I. N. Kriv-\norotov, R. A. Buhrman, and D. C. Ralph, \\Spin-transfer-\ndriven ferromagnetic resonance of individual nanomag-\nnets,\" Phys. Rev. Lett. 96, 227601 (2006).\n[23] Motoya Shinozaki, Eriko Hirayama, Shun Kanai, Hideo\nSato, Fumihiro Matsukura, and Hideo Ohno, \\Damp-\ning constant in a free layer in nanoscale CoFeB/MgO\nmagnetic tunnel junctions investigated by homodyne-\ndetected ferromagnetic resonance,\" Appl. Phys. Exp. 10,\n013001 (2017).\n[24] Christopher J. Safranski, Yu-Jin Chen, Ilya N. Krivo-\nrotov, and Jonathan Z. Sun, \\Material parameters of\nperpendicularly magnetized tunnel junctions from spin\ntorque ferromagnetic resonance techniques,\" Appl. Phys.\nLett. 109, 132408 (2016).\n[25] Michael Harder, Yongsheng Gui, and Can-Ming Hu,\n\\Electrical detection of magnetization dynamics via spin\nrecti\fcation e\u000bects,\" Phys. Rep. 661, 1{59 (2016).\n[26] O. Mosendz, J. E. Pearson, F. Y. Fradin, G. E. W. Bauer,\nS. D. Bader, and A. Ho\u000bmann, \\Quantifying spin Hall\nangles from spin pumping: Experiments and theory,\"\nPhys. Rev. Lett. 104, 046601 (2010).\n[27] C. Liu, Y. Boyko, C. C. Geppert, K. D. Christie,\nG. Stecklein, S. J. Patel, C. J. Palmstrm, and P. A.\nCrowell, \\Electrical detection of ferromagnetic resonance\nin ferromagnet/n-GaAs heterostructures by tunneling\nanisotropic magnetoresistance,\" Appl. Phys. Lett. 105,\n212401 (2014).\n[28] Shinji Miwa, Junji Fujimoto, Philipp Risius, Kohei\nNawaoka, Minori Goto, and Yoshishige Suzuki, \\Strong\nbias e\u000bect on voltage-driven torque at epitaxial Fe-MgO\ninterface,\" Phys. Rev. X 7, 031018 (2017).\n[29] C\u0013 esar L. Ord\u0013 o~ nez-Romero, Boris A. Kalinikos, Pavol\nKrivosik, Wei Tong, Pavel Kabos, and Carl E. Pat-\nton, \\Three-magnon splitting and con\ruence processes\nfor spin-wave excitations in yttrium iron garnet \flms:\nWave vector selective Brillouin light scattering measure-\nments and analysis,\" Phys. Rev. B 79, 144428 (2009).\n[30] H. Schultheiss, X. Janssens, M. van Kampen, F. Ciub-\notaru, S. J. Hermsdoerfer, B. Obry, A. Laraoui, A. A.\nSerga, L. Lagae, A. N. Slavin, B. Leven, and B. Hille-\nbrands, \\Direct current control of three magnon scat-\ntering processes in spin-valve nanocontacts,\" Phys. Rev.\nLett. 103, 157202 (2009).\n[31] C. T. Boone, J. A. Katine, J. R. Childress, V. Tiberke-\nvich, A. Slavin, J. Zhu, X. Cheng, and I. N. Krivorotov,\n\\Resonant nonlinear damping of quantized spin waves\nin ferromagnetic nanowires: A spin torque ferromagnetic\nresonance study,\" Phys. Rev. Lett. 103, 167601 (2009).\n[32] Hidekazu Kurebayashi, Oleksandr Dzyapko, Vladislav E.\nDemidov, Dong Fang, Andrew J. Ferguson, and\nSergej O. Demokritov, \\Controlled enhancement of spin-\ncurrent emission by three-magnon splitting,\" Nat. Mater.\n10, 660{664 (2011).\n[33] R. N. Costa Filho, M. G. Cottam, and G. A. Farias,\n\\Microscopic theory of dipole-exchange spin waves in fer-\nromagnetic \flms: Linear and nonlinear processes,\" Phys.\nRev. B 62, 6545{6560 (2000).8\n[34] Alina M Deac, Akio Fukushima, Hitoshi Kubota, Hi-\nroki Maehara, Yoshishige Suzuki, Shinji Yuasa, Yoshinori\nNagamine, Koji Tsunekawa, David D Djayaprawira, and\nNaoki Watanabe, \\Bias-driven high-power microwave\nemission from MgO-based tunnel magnetoresistance de-\nvices,\" Nat. Phys. 4, 803{809 (2008).\n[35] V. E. Demidov, S. Urazhdin, E. R. J. Edwards, M. D.\nStiles, R. D. McMichael, and S. O. Demokritov, \\Con-\ntrol of magnetic \ructuations by spin current,\" Phys. Rev.\nLett. 107, 107204 (2011).\n[36] G. D. Fuchs, J. C. Sankey, V. S. Pribiag, L. Qian,\nP. M. Braganca, A. G. F. Garcia, E. M. Ryan, Zhi-\nPan Li, O. Ozatay, D. C. Ralph, and R. A. Buhrman,\n\\Spin-torque ferromagnetic resonance measurements of\ndamping in nanomagnets,\" Appl. Phys. Lett. 91, 062507\n(2007).\n[37] G.A. Melkov, D.V. Slobodianiuk, V.S. Tiberkevich,\nG. de Loubens, O. Klein, and A.N. Slavin, \\Nonlinear\nferromagnetic resonance in nanostructures having dis-\ncrete spectrum of spin-wave modes,\" IEEE Magn. Lett.\n4, 4000504 (2013).\n[38] M. Helsen, A. Gangwar, J. De Clercq, A. Vansteenkiste,\nM. Weigand, C. H. Back, and B. Van Waeyenberge,\n\\Non-linear radial spinwave modes in thin magnetic\ndisks,\" Appl. Phys. Lett. 106, 032405 (2015).\n[39] Jan Podbielski, Detlef Heitmann, and Dirk Grundler,\n\\Microwave-assisted switching of microscopic rings: Cor-\nrelation between nonlinear spin dynamics and critical mi-\ncrowave \felds,\" Phys. Rev. Lett. 99, 207202 (2007).\n[40] H. Suhl, \\The theory of ferromagnetic resonance at high\nsignal powers,\" J. Phys. Chem. Solids 1, 209{227 (1957).\n[41] Hans G Bauer, Peter Majchrak, Torsten Kachel, Chris-\ntian H Back, and Georg Woltersdorf, \\Nonlinear spin-\nwave excitations at low magnetic bias \felds,\" Nat. Com-\nmun. 6, 8274 (2015).\n[42] M. Bauer, O. B uttner, S. O. Demokritov, B. Hillebrands,\nV. Grimalsky, Yu. Rapoport, and A. N. Slavin, \\Obser-\nvation of spatiotemporal self-focusing of spin waves in\nmagnetic \flms,\" Phys. Rev. Lett. 81, 3769{3772 (1998).\n[43] Arnold Markovich Kosevich, B.A. Ivanov, and A.S.\nKovalev, \\Magnetic solitons,\" Phys. Rep. 194, 117{238\n(1990).\n[44] A. N. Slavin and I. V. Rojdestvenski, \\Bright and dark\nspin wave envelope solitons in magnetic \flms,\" IEEE\nTrans. Magn. 30, 37{45 (1994).\n[45] Mingzhong Wu, Boris A. Kalinikos, and Carl E. Patton,\n\\Self-generation of chaotic solitary spin wave pulses in\nmagnetic \flm active feedback rings,\" Phys. Rev. Lett.\n95, 237202 (2005).\n[46] V. V. Naletov, G. de Loubens, V. Charbois, O. Klein,\nV. S. Tiberkevich, and A. N. Slavin, \\Ferromagnetic\nresonance spectroscopy of parametric magnons excited\nby a four-wave process,\" Phys. Rev. B 75, 140405 (2007).\n[47] Y. Khivintsev, Bijoy Kuanr, T. J. Fal, M. Haftel, R. E.\nCamley, Z. Celinski, and D. L. Mills, \\Nonlinear fer-\nromagnetic resonance in permalloy \flms: A nonmono-\ntonic power-dependent frequency shift,\" Phys. Rev. B 81,\n054436 (2010).\n[48] Bivas Rana, Yasuhiro Fukuma, Katsuya Miura, Hiro-\nmasa Takahashi, and YoshiChika Otani, \\E\u000bect of ex-\ncitation power on voltage induced local magnetization\ndynamics in an ultrathin CoFeB \flm,\" Sci. Rep. 7, 2318\n(2017).[49] M. d'Aquino, A. Quercia, V. Scalera, S. Perna,\nG. Bertotti, I. D. Mayergoyz, and C. Serpico, \\Ana-\nlytical treatment of nonlinear ferromagnetic resonance in\nnanomagnets,\" IEEE Trans. Magn. 53, 4301005 (2017).\n[50] I. N. Krivorotov, N. C. Emley, R. A. Buhrman, and D. C.\nRalph, \\Time-domain studies of very-large-angle mag-\nnetization dynamics excited by spin transfer torques,\"\nPhys. Rev. B 77, 054440 (2008).\n[51] Feng Guo, Lyubov M. Belova, and Robert D. McMichael,\n\\Nonlinear ferromagnetic resonance shift in submicron\npermalloy ellipses,\" Phys. Rev. B 91, 064426 (2015).\n[52] Hans T. Nembach, Justin M. Shaw, Carl T. Boone, and\nT. J. Silva, \\Mode- and size-dependent Landau-Lifshitz\ndamping in magnetic nanostructures: Evidence for non-\nlocal damping,\" Phys. Rev. Lett. 110, 117201 (2013).\n[53] Y. Li and W.E. Bailey, \\Wave-number-dependent Gilbert\ndamping in metallic ferromagnets,\" Phys. Rev. Lett. 116,\n117602 (2016).\n[54] Koji Sekiguchi, Keisuke Yamada, Soo-Man Seo, Kyung-\nJin Lee, Daichi Chiba, Kensuke Kobayashi, and Teruo\nOno, \\Time-domain measurement of current-induced\nspin wave dynamics,\" Phys. Rev. Lett. 108, 017203\n(2012).\n[55] V. Lauer, D. A. Bozhko, T. Brcher, P. Pirro, V. I.\nVasyuchka, A. A. Serga, M. B. Jung\reisch, M. Agrawal,\nYu. V. Kobljanskyj, G. A. Melkov, C. Dubs, B. Hille-\nbrands, and A. V. Chumak, \\Spin-transfer torque based\ndamping control of parametrically excited spin waves in\na magnetic insulator,\" Appl. Phys. Lett. 108, 012402\n(2016).\n[56] Chi Zhang, Yong Pu, Sergei A. Manuilov, Shane P.\nWhite, Michael R. Page, Erick C. Blomberg, Denis V.\nPelekhov, and P. C. Hammel, \\Engineering the spec-\ntrum of dipole \feld-localized spin-wave modes to enable\nspin-torque antidamping,\" Phys. Rev. Appl. 7, 054019\n(2017).\n[57] S. M. Rezende, F. M. de Aguiar, and A. Azevedo,\n\\Magnon excitation by spin-polarized direct currents\nin magnetic nanostructures,\" Phys. Rev. B 73, 094402\n(2006).Supplemental Material:\nGiant resonant nonlinear damping in nanoscale ferromagnets\nI. Barsukov,1H. K. Lee,1A. A. Jara,1Y.-J. Chen,1A. M. Gon¸ calves,1C.\nSha,1J. A. Katine,2R. E. Arias,3B. A. Ivanov,4, 5and I. N. Krivorotov1\n1Physics and Astronomy, University of California, Irvine, CA 92697, USA\n2Western Digital, 5600 Great Oaks Parkway, San Jose, CA 95119, USA\n3Departamento de F´ ısica, CEDENNA, FCFM, Universidad de Chile, Santiago, Chile\n4Institute of Magnetism, National Academy of Sciences of Ukraine, Vernadsky av. 36 B, Kyiv, 03142, Ukraine\n5National University of Science and Technology MISiS, Moscow, 119049, Russian Federation\nI. METHODS\nA. Linewidth evaluation\nAll measurements presented were carried out with magnetic field applied along the easy axis of the MTJ devices so\nthat the magnetic moments of the free and pinned layers are collinear to each other. In this geometry, the ST-FMR sig-\nnals are dominated by photo-resistance contribution and are proportional to the square of the transverse component of\nthe dynamic magnetization magnetization [1], which allows us to directly compare calculated |a|2(ω) resonance curves\nto measured ST-FMR resonance curves ˜Vmix(f) and toVmix(f) approximated by numerical integration/integraltext˜Vmix(f)df.\nWhenVmix(f) and|a|2(ω) are single-peak curves, they are fit to a sum of symmetric and antisymmetric Lorentzian\ncurves with identical central frequencies and linewidth parameters as described in Ref. [2], and the spectral linewidth\nis determined as half-width at the half-maximum of the symmetric Lorentzian curve.\nIn order to quantify the linewidth of the split-peak resonance profile, we introduce a fitting function that is a sum\nof two Lorentzian curves with different central frequencies separated by δf. The half width of the resonance profile\n∆f0is then defined as the average of the half widths of the two Lorentzians plus δf/2.\nSupplemental Figure 1. Spatial profiles of spin wave eigenmodes. Normalized amplitude and phase of the three lowest frequency\nspin wave eigenmodes of the MTJ free layer, given by micromagnetic simulations.\nB. Micromagnetic simulations\nMicromagnetic simulations were performed using OOMMF software [3, 4]. To account for all magnetic interactions\nin the MTJ, a three dimensional model was employed with three ferromagnetic layers: free, SAF top and SAF bottom.\nWe use material parameters obtained from the measurements and/or their accepted literature values (see Ref. [2] for2\nthe MTJ structure and fabrication details). Magnetization dynamics is excited by a combined pulse of spin torque\nand Oersted field, resulting from a sinc-shaped spatially uniform current pulse. The spatial profile of the Oersted\nfield corresponds to that of a long wire with elliptical cross section. The direction of the spin torque vector acting\non the free layer is determined by the magnetization orientation of the SAF top layer. The spectrum of spin wave\neigenmodes is obtained via fast Fourier transform (FFT) of the time dependent components of the layers’ magnetic\nmoment. Spatial mapping of the resulting Fourier amplitude and phase at a given frequency provides the mode\nprofiles (Supplmental Fig. 1). The observed excitations are confirmed to be spin wave modes localized to the free\nlayer. SAF modes are found at much higher frequencies than the free layer modes, and their frequencies are found to\nbe incommensurable to the free layer quasi-uniform mode frequency [5].\nII. SOLUTION OF THE EQUATIONS OF MOTION\nThe Hamiltonian equations of motion describing the coupled dissipative dynamics of the quasi-uniform ( a) and the\nhigher-order ( b) spin wave modes are:\nida\ndt=∂H\n∂a†+∂Q\n∂(da†/dt)(1)\nidb\ndt=∂H\n∂b†+∂Q\n∂(db†/dt)(2)\nwhereHis the Hamiltonian of the system and Qis the dissipation function, given by:\nH=ω0a†a+ωnb†b+1\n2Ψ0a†a†aa+1\n2Ψnb†b†bb+ (ψ∗\nnaab†+ψna†a†b) +ζ{exp(−iωt)a†+ exp(iωt)a} (3)\nQ=da†\ndtda\ndt(α0+η0a†a) +db†\ndtdb\ndt(αn+ηnb†b) (4)\nBy using Eq. (3) and Eq. (4) in Eq. (1) and Eq. (2), the Hamiltonian equations can be written as:\nida\ndt−(α0+η0a†a)da\ndt=ω0a+ 2ψna†b+ Ψ 0a†aa+ζexp(−iωt) (5)\nidb\ndt−(αn+ηnb†b)db\ndt=ωnb+ψ∗\nnaa+ Ψnb†bb (6)\nUsing a periodic ansatz a= ¯aexp(−iωt) andb=¯bexp(−2iωt) in Eq. (5) and Eq. (6), where ¯ aand¯bare complex\namplitudes, reduces the Hamiltonian equations to a set of two algebraic equation for the complex amplitudes:\n/parenleftbig\nω−ω0−Ψ0|¯a|2+i(α0+η0|¯a|2)ω/parenrightbig\n¯a−2ψn¯a∗¯b=ζ (7)\n/parenleftbig\n2ω−ωn−Ψn|¯b|2+ 2i(αn+ηn|¯b|2)ω/parenrightbig¯b=ψ∗\nn¯a2(8)\nWe solve Eq. (8) for ¯band multiply the numerator and denominator of this expression by the complex conjugate of\nthe denominator:\n¯b=ψ∗\nn¯a2/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig\n−i2(αn+ηn|¯b|2)ω\n/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2(9)\nthen we multiply Eq. (9) by2ψn¯a∗\n¯aand evaluate the real and imaginary parts.\n/Rfractur/bracketleftbigg2ψn¯a∗¯b\n¯a/bracketrightbigg\n=|ψn|2|¯a|22/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig\n/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2(10)\n/Ifractur/bracketleftbigg2ψn¯a∗¯b\n¯a/bracketrightbigg\n=|ψn|2|¯a|2−4(αn+ηn|¯b|2)ω\n/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2(11)\nBy taking the modulus of Eq. (8), we obtain:\n|¯a|2=|¯b|\n|ψn|/radicalBig/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2 (12)3\nUsing Eq. (12) in Eqs. (10-11), we derive:\n/Rfractur/bracketleftbigg2ψn¯a∗¯b\n¯a/bracketrightbigg\n=2/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig\n|ψn||¯b|/radicalBig/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2(13)\n/Ifractur/bracketleftbigg2ψn¯a∗¯b\n¯a/bracketrightbigg\n=−4(αn+ηn|¯b|2)ω|ψn||¯b|/radicalBig/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2(14)\nTaking the modulus squared of Eq. (7):\n/braceleftBigg/parenleftbigg\nω−ω0−Ψ0|¯a|2−/Rfractur/bracketleftbigg2ψn¯a∗¯b\n¯a/bracketrightbigg/parenrightbigg2\n+/parenleftbigg\n(α0+η0|¯a|2)ω−/Ifractur/bracketleftbigg2ψn¯a∗¯b\n¯a/bracketrightbigg/parenrightbigg2/bracerightBigg\n|¯a|2=ζ2(15)\nand using Equations (12)–(14) in Eq. (15) gives us an algebraic equation for the absolute value of the higher order\nmode amplitude|¯b|:\n\n\n\nω−ω0−Ψ0|¯b|\n|ψn|/radicalBig/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2−2/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig\n|ψn||¯b|/radicalBig/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2\n2\n+\n\n/parenleftbigg\nα0+η0|¯b|\n|ψn|/radicalBig/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2/parenrightbigg\nω−−4(αn+ηn|¯b|2)ω|ψn||¯b|/radicalBig/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2\n2\n\n×\n|¯b|\n|ψn|/radicalBig/parenleftbig\n2ω−ωn−Ψn|¯b|2/parenrightbig2+ 4(αn+ηn|¯b|2)2ω2=ζ2\n(16)\nAfter numerically solving Eq. (16) for |¯b|, and using it in Eq. (12), we can calculate the amplitude of the quasi-uniform\nmode|¯a|.\nIII. EFFECTS OF THE DRIVE AMPLITUDE AND INTRINSIC NONLINEARITITES\nTo understand the impact of the intrinsic nonlinearity parameters (Ψ 0, Ψn,η0,ηn) on the quasi-uniform spin wave\nmode resonance, we plot the numerical solution of Eq. (16) in Supplemental Figure 2. Each panel of this figure shows\na reference lineshape of the resonance calculated with all intrinsic nonlinearity parameters set to zero (red curve) and\na lineshape calculated with one of the intrinsic nonlinearity parameter different from zero (blue curve). This figure\nreveals that increasing η0decreases the mode amplitude and slightly increases the linewidth. Increasing ηndecreases\nthe degree of the double-peak lineshape splitting. Increasing Ψ nincreases the lineshape asymmetry. Increasing Ψ 0\nincreases lineshape asymmetry and induces fold-over.\nSupplemental Figure 3 shows the linewidth as a function of the drive amplitude for three scenarios, where the\nintrinsic nonlinearities Ψ 0, Ψn,η0,ηnare set to zero for simplicity. If the coupling parameter is zero, ψn= 0, the\nlinewidth does not depend on the drive amplitude, as expected for a single-mode linear oscillator. The second case\ndemonstrates that the linewidth remains constant when the product ψn·ζis constant. For a constant non-zero\ncoupling parameter, the linewidth shows an increase with the drive amplitude. This observation allows us to employ a\nsingle fitting parameter ( ψn·ζ) to fit the data in Fig. 1b. This conjecture can be confirmed analytically by introducing\na normalized spin wave amplitude ˆ a=ψn¯a, which allows us to rewrite Eq. (16) omitting all intrinsic nonlinearities\ninto the following form:\nω/bracketleftbigg\n1 +iα0+i4αn|ˆa|2\n(2ω−ωn)2+ 4α2nω2/bracketrightbigg\nˆa−ω0ˆa−2(2ω−ωn)\n(2ω−ωn)2+ 4α2nω2|ˆa|2ˆa=ψnζ (17)\nThis equation describes an effective single-mode nonlinear oscillator with renormalized excitation amplitude ψnζ.4\n2 2.5 3 3.500.010.02\nFrequency (GHz)0\nη0\n2 2.5 3 3.500.010.02\nFrequency (GHz)|a|20\nηn\n2 2.5 3 3.500.010.02\nFrequency (GHz)0\nΨ0\n2 2.5 3 3.500.010.02\nFrequency (GHz)0\nΨn|a|2|a|2|a|2a b\nc d\nSupplemental Figure 2. Effect of intrinsic nonlinearities on the quasi-uniform spin wave resonance lineshape. Spectral lineshape\nof the quasi-uniform spin wave mode resonance |¯a|2(ω) at the three-magnon resonance condition 2 ω0=ωncalculated by\nnumerically solving Eq. (16). The red curve is a reference lineshape calculated with all intrinsic nonlinearity parameters\n(η0,ηn,Ψ0,Ψn) set to zero. The blue lineshape in each panel is calculated with one of the intrinsic nonlinearity parameters set\nto a non-zero value: (a) η0= 1.325·10−24J, (b)ηn= 3.313·10−24J, (c) Ψ 0= 1.325·10−24J, (d) Ψ n= 1.325·10−23J. Other\nparameters employed in the calculation are: ω0= 2π·2.63 GHz,ωn= 2π·5.26 GHz;α0= 0.02662,αn= 0.03042 atIdc= 0;\nψn·ζ=h2·0.006 GHz2, wherehis the Planck constant.\n0 0.05 0.1 0.15 0.20.150.20.25\nζ h-1 (GHz)∆f0 (GHz)\n(i)(ii)(iii)\nSupplemental Figure 3. Effect of the drive amplitude on linewidth in the resonant three-magnon regime. Calculated linewidth\nof the quasi-uniform spin wave mode as a function of the drive amplitude ζfor different values of the mode coupling parameter\nψn. (i) Green: ψn= 0, (ii) red: variable ψnwith a constraint ψn·ζ=h2·0.006 GHz2, and (iii) blue: ψn=h·0.1 GHz. All\nintrinsic nonlinearity parameters: Ψ 0, Ψn,η0andηnare set to zero. his the Planck constant. Other parameters employed in\nthe calculation are: ω0= 2π·2.63 GHz,ωn= 2π·5.26 GHz;α0= 0.02662 andαn= 0.03042 atIdc= 0.5\nIV. EFFECTIVE SINGLE-MODE NONLINEAR OSCILLATOR APPROXIMATION\nIf we neglect intrinsic nonlinearities Ψ nandηnof the higher order spin wave mode, Eq. (16) can be reduced to a\ncubic equation for ¯ aand solved analytically. This approximation allows us to obtain several important qualitative\ninsights into the properties of the resonant nonlinear damping of the quasi-uniform mode. By setting Ψ n= 0 and\nηn= 0 in Eq. (8), we obtain an exact solution for ¯b:\n¯b=ψ∗\nn¯a2\n2ω(1 +iαn)−ωn(18)\nUsing this result, we reduce Eq. (16) to a cubic algebraic equation for ¯ a:\nω/bracketleftbigg\n1 +i(α0+η0|¯a|2) +i4|ψn|2αn|¯a|2\n(2ω−ωn)2+ 4α2nω2/bracketrightbigg\n¯a−ω0¯a−/bracketleftbigg\nΨ0+2|ψn|2(2ω−ωn)\n(2ω−ωn)2+ 4α2nω2/bracketrightbigg\n|¯a|2¯a=ζ (19)\nThis equation describes the amplitude ¯ aof an effective single-mode nonlinear oscillator.\nIt is evident from Eq. (19) that the frequency of the quasi-uniform mode experiences a nonlinear shift:\nωeff\n0=ω0+/bracketleftbigg\nΨ0+2|ψn|2(2ω−ωn)\n(2ω−ωn)2+ 4α2nω2/bracketrightbigg\n|¯a|2(20)\nThe nonlinear frequency shift has a well-pronounced antisymmetric resonant character near the resonance frequency\nωn/2, that arises from the resonant three-magnon scattering.\nFurther, it is clear from Eq. (19) that the effective damping of the quasi-uniform mode also acquires a term arising\nfrom the three-magnon interaction:\nαeff\n0=α0+/bracketleftbigg\nη0+4|ψn|2αn\n(2ω−ωn)2+ 4α2nω2/bracketrightbigg\n|¯a|2(21)\nThe last term describes a resonant enhancement of the nonlinear damping by three-magnon scattering near the\nresonance frequency ωn/2. Strikingly, the magnitude of the resonant damping enhancement at ωn/2 increases when\nthe intrinsic damping of the higher order mode αndecreases. In the limit αn→0, the effective damping becomes\nαeff\n0→α0+/bracketleftbigg\nη0+2π|ψn|2\nωδ(2ω−ωn)/bracketrightbigg\n|¯a|2(22)\nwhereδis Dirac’s delta function. Equation (21) suggests that the effective damping of the quasi-uniform mode αeff\n0\ncan increase with increasing antidamping spin torque applied to the nanomagnet. Indeed, the antidamping spin torque\ntends to increase the amplitude [6] of the quasi-uniform mode |¯a|and decrease the intrinsic damping parameter of the\nhigher order mode αn→αn(1−Idc/I|n/angbracketright\nc), both enhancing the nonlinear damping term in Eq. (19). For a sufficiently\nlarge mode coupling parameter ψn, the enhancement of the nonlinear damping term by the antidamping spin torque\ncan exceed the reduction of the linear damping parameter α0→α0(1−Idc/I|0/angbracketright\nc) by the torque, leading to an increase\nofαeff\n0byIdc>0 and broadening of the quasi-uniform mode resonance by the antidamping spin torque. This scenario\nis indeed realized in the MTJ devices studied here as demonstrated by the data and calculations in Fig. 3.\nV. MODE COUPLING PARAMETER\nIn this Supplementary Note, we discuss how the coupling parameter between the spin wave modes, ψnin Eq. (3),\ncan be calculated. We consider a very thin, magnetically soft ferromagnetic disk with elliptical cross section, that\nis magnetized in-plane. Within a classical micromagnetic model, we include Zeeman, dipolar and exchange terms in\nthe free energy. An applied field Halong thexdirection (long axis of the ellipse) magnetizes the sample to a nearly\nuniform state. Through a classical Holstein-Primakoff transformation [7] we introduce variables c(/vector x,t) andc∗(/vector x,t) to\ndescribe the magnetization such that the magnetization magnitude is conserved:\nmx= 1−cc∗, m +=c√2−cc∗, m−=c∗√2−cc∗, (23)\nwhere/vector m=/vectorM/Ms, andm±≡mz±imy. Approximating the exchange energy to the fourth order in candc∗, the\nnormalized free energy of the disk, U≡E/4πM2\ns, is given by\nU/similarequal−hx/integraldisplay\n(1−cc∗) dV+ (lex)2/integraldisplay/bracketleftbigg\n/vector∇c·/vector∇c∗+1\n4c2(/vector∇c∗)2+1\n4c∗2(/vector∇c)2/bracketrightbigg\ndV−1\n2/integraldisplay\ndV/vectorhD(/vector m)·/vector m, (24)6\nwithhx≡H/4πMs,lex≡/radicalbig\nA/2πM2sis the exchange length, and /vectorhD(/vector m) =/vectorHD(/vector m)/4πMsis the normalized\ndemagnetizing field. The Landau-Lifshitz equations of motion in the new variables are: i˙c=δU/δc∗,i˙c∗=−δU/δc\nwitht/prime= 4πMs|γ|t.\nAssuming the normal modes involved in three magnon scattering dominate the magnetization dynamics, the free\nenergy in Eq. (24) can be written in terms of amplitudes of these modes, by expressing cin terms of aandb:\nc(/vector x,t)/similarequala(t)f(/vector x) +a∗(t)g(/vector x) +b(t)p(/vector x) +b∗(t)q(/vector x) (25)\nThe functions f,g,p,q can be determined from calculating the linear modes of oscillation of the sample. The terms of\nthe free energy proportional to aab∗anda∗a∗bdescribe the three-magnon process and the magnitude of these terms\ngives the coupling parameter ψn.\nIf the magnetization state is approximated as exactly uniform, the dipolar energy for a very thin film may be\napproximated as UD=m2\nz/2 = (c+c∗)2(1−cc∗/2), and in this case all three-magnon terms are zero. However, when\nthe effects due to the sample edges (such as spatial inhomogeneity of the demagnetization field and edge roughness)\nare taken into account, the equilibrium magnetization configuration is generally nonuniform. In this case, there are\nnon-zero three-magnon terms in the free energy expression. An explicit calculation of the corresponding overlap\nintegrals is necessary for a quantitative prediction of ψn. Refs. [8, 9] show such extensive calculations for circular disks\nand include explicit expressions for the exchange and dipolar energies.\n[1] Michael Harder, Yongsheng Gui, and Can-Ming Hu, “Electrical detection of magnetization dynamics via spin rectification\neffects,” Phys. Rep. 661, 1–59 (2016).\n[2] A. M. Gon¸ calves, I. Barsukov, Y.-J. Chen, L. Yang, J. A. Katine, and I. N. Krivorotov, “Spin torque ferromagnetic\nresonance with magnetic field modulation,” Appl. Phys. Lett. 103, 172406 (2013).\n[3] M. J. Donahue and D. G. Porter, OOMMF User’s Guide (National Institute of Standards and Technology, Gaithersburg,\nMD, 1999).\n[4] Robert D. McMichael and Mark D. Stiles, “Magnetic normal modes of nanoelements,” J. Appl. Phys. 97, 10J901 (2005).\n[5] P. S. Keatley, V. V. Kruglyak, A. Neudert, R. J. Hicken, V. D. Poimanov, J. R. Childress, and J. A. Katine, “Resonant\nenhancement of damping within the free layer of a microscale magnetic tunnel valve,” J. Appl. Phys. 117, 17B301 (2015).\n[6] J. C. Sankey, P. M. Braganca, A. G. F. Garcia, I. N. Krivorotov, R. A. Buhrman, and D. C. Ralph, “Spin-transfer-driven\nferromagnetic resonance of individual nanomagnets,” Phys. Rev. Lett. 96, 227601 (2006).\n[7] T. Holstein and H. Primakoff, “Field dependence of the intrinsic domain magnetization of a ferromagnet,” Phys. Rev. 58,\n1098–1113 (1940).\n[8] D. Mancilla-Almonacid and R. E. Arias, “Instabilities of spin torque driven auto-oscillations of a ferromagnetic disk mag-\nnetized in plane,” Phys. Rev. B 93, 224416 (2016).\n[9] D. Mancilla-Almonacid and R. E. Arias, “Spin-wave modes in ferromagnetic nanodisks, their excitation via alternating\ncurrents and fields, and auto-oscillations,” Phys. Rev. B 95, 214424 (2017)." }, { "title": "1903.10135v1.Distributed_Inter_Area_Oscillation_Damping_Control_for_Power_Systems_by_Using_Wind_Generators_and_Load_Aggregators.pdf", "content": "arXiv:1903.10135v1 [math.OC] 25 Mar 20191\nDistributed Inter-Area Oscillation Damping Control\nfor Power Systems by Using Wind Generators and\nLoad Aggregators\nZhiyuan Tang, Yue Song, Tao Liu, and David J. Hill, Life Fellow, IEEE\nAbstract —This paper investigates the potential of wind turbine\ngenerators (WTGs) and load aggregators (LAs) to provide sup -\nplementary damping control services for low frequency inte r-area\noscillations (LFOs) through the additional distributed da mping\ncontrol units (DCUs) proposed in their controllers. In orde r\nto provide a scalable methodology for the increasing number\nof WTGs and LAs, a novel distributed control framework is\nproposed to coordinate damping controllers. Firstly, a dis tributed\nalgorithm is designed to reconstruct the system Jacobian ma trix\nfor each damping bus (buses with damping controllers). Thus ,\nthe critical LFO can be identified locally at each damping bus by\napplying eigen-analysis to the obtained system Jacobian ma trix.\nThen, if the damping ratio of the critical LFO is less than a\npreset threshold, the control parameters of DCUs will be tun ed\nin a distributed and coordinated manner to improve the dampi ng\nratio and minimize the total control cost at the same time. Th e\nproposed control framework is tested in a modified IEEE 39-bu s\ntest system. The simulation results with and without the pro posed\ncontrol framework are compared to demonstrate the effectiv eness\nof the proposed framework.\nIndex Terms —Low frequency oscillation, load-side control,\nwind generator, distributed control\nI. I NTRODUCTION\nLow frequency inter-area oscillations (LFOs) have always\nbeen a matter of concern to power system operators due to\ntheir potential threats to the power system stability [1]. W ith\nthe development of the electricity market and growing power\ndemand, future power systems will become more stressed\nand operate closer to their stability limits, which highlig hts\nthe need to improve the damping ratio of LFOs and prevent\nsustained oscillations that can result in serious conseque nces\nsuch as system separations or even large-area blackouts [1] .\nThe power system stabilizers (PSSs) installed on conven-\ntional synchronous machines are the most important compon-\nents to improve system damping against LFOs. However, the\nincreasing penetration of wind power limits the availabili ty\nof PSSs to provide sufficient damping against LFOs. For one\nthing, the displacement of conventional synchronous gener -\nators with wind turbine generators (WTGs) may reduce the\nThis work was fully supported by the Research Grants Council of the\nHong Kong Special Administrative Region under the Theme-ba sed Research\nScheme through Project No. T23-701/14-N.\nZ. Tang, Y . Song, and T. Liu are with the Department of Electri cal and\nElectronic Engineering, The University of Hong Kong, Hong K ong (email:\nzytan@eee.hku.hk; yuesong@eee.hku.hk; taoliu@eee.hku. hk).\nD. J. Hill is with the Department of Electrical and Electroni c Engineering,\nThe University of Hong Kong, Hong Kong. He is also with the Sch ool of\nElectrical and Information Engineering, The University of Sydney, NSW 2006,\nAustralia (email: dhill@eee.hku.hk; david.hill@sydney. edu.au).damping ratio of inter-area modes by the reconfiguration of\nline power flows, reduction of system inertia, and interacti on\nof converter controls with power system dynamics [2]. For\nanother thing, once the conventional synchronous machines\nare replaced by WTGs, the associated PSSs are also removed\nfrom the system with no replacement controllers for WTGs to\nprovide damping control services. Thus, if no new alternati ve\ncontrollers are developed to provide supplementary dampin g\ncontrol services, insufficient system controls may jeopard ize\nthe system security and stability. To solve this issue, in th is\npaper, we are looking for solutions from both the generation\nand load sides.\nFor the generation side, we utilize the converter interface d\nWTGs which can provide damping torques for LFOs by\nquickly adjusting their active power outputs though a prope r\ncontrol of electronic devices that interface them with the\ngrid [3], [4]. For the load side, the option of using highly\ndistributed controllable loads (demand control) is appeal ing.\nDue to properties such as instantaneous responses and spati al\ndistributions, demand control has gained a lot of attention\n[5]–[7]. In particular, demand control has been utilized to\naccomplish important system support tasks such as frequenc y\ncontrol [5], voltage control [6], and small-disturbance an gle\nstability enhancement [7]. However, the ability of demand\ncontrol to provide supplementary damping control services\nagaist LFOs has not been thoroughly investigated yet. In thi s\npaper, the load aggregators (LAs) will be coordinated with\nWTGs to provide damping torques against LFOs through the\nadditional distributed damping control units (DCUs) devel oped\nin their controllers.\nIn the literature, numerous methods have been proposed\nto coordinate traditional damping controllers (e.g. PSS) [ 8]–\n[10] and new damping controllers (e.g. FACTS and HVDC)\n[11]–[13]. Approaches based on robust control theories and\nlinear matrix inequalities have been utilized to deal with t he\nuncertainties of operating conditions [9]–[11]. For examp le,\nin [10], the synthesis of the controller is formulated as a\nmixedH2/H∞output feedback control problem with regional\npole placement that is resolved through a linear matrix in-\nequality approach. However, such a robust controller desig n\nmethod is too conservative and unable to incorporate all\nsystem constraints (e.g. hard limits on the control signals ).\nApproaches based on model predictive control have been\nutilized to incorporate all system constraints. For exampl e,\nthe authors of [12] propose a model predictive control based\nHVDC supplementary controller which can incorporate plant2\nconstraints explicitly. Unfortunately, the model used in s uch a\nmethod is developed at a pre-given operating point, and henc e,\nthe obtained controller cannot directly guarantee robustn ess\naround the other operating points. Approaches based on fuzz y\nlogic have been utilized to handle the variations of operati ng\npoints [13]. For example, a fuzzy logic adaptive control uni t is\nproposed in [13] to adjust control gains for different opera ting\npoints. However, this fuzzy logic based method becomes very\ncomplicated when the number of damping controllers becomes\nlarge. Moreover, all the methods mentioned above are carrie d\nout in a centralized manner that lacks scalability and flexib ility,\ni.e., a new damping controller is added into the original con trol\nsystem, the whole control law need to be redesigned.\nTo overcome the drawbacks of the abovementioned meth-\nods, in this paper, a novel distributed control framework is\nproposed to coordinate damping controllers, which can be\nimplemented by local measurements and limited communic-\nations between neighboring buses. The proposed distribute d\ncontrol framework consists of two modules: a critical LFO\nidentification module and a controller parameters tuning mo d-\nule where the communication network used in each module\nis different. The critical LFO identification module aims at\nreconstructing the system Jacobian matrix for each damping\nbus (a bus with damping controller) in a distributed manner\nwhere the communication network used covers all buses in\nthe system. Thus, the critical LFO (the LFO with the least\ndamping ratio) can be identified locally at each damping bus\nby applying eigen-analysis to the obtained system Jacobian\nmatrix. Further, if the damping ratio of the critical LFO is\nless than a preset threshold, the parameters of DCUs will be\ntuned in a distributed manner to improve the damping ratio\nof the critical LFO and minimize the total control cost at the\nsame time where the communication network used only covers\nthose damping buses. The contributions of this paper are lis ted\nbelow:\n•A novel two-step communication based distributed con-\ntrol framework is proposed to coordinate LAs and WTGs.\nThe proposed control method can survive one-point fail-\nure in the communication network and is suitable in\npractice for its scalability.\n•In the critical LFO identification module, based on struc-\ntural properties of the original power grid, a distributed\ncalculation algorithm is developed to recover the Jacobian\nmatrices for each damping bus.\n•In the controller parameters tuning module, based on the\neigenvalue sensitivities, a controller tuning problem is\nformulated and solved in a distributed manner.\nThe rest of the paper is organized as follows. Section II\nintroduces the DCU and the power system model to be stud-\nied. The proposed distributed control framework is explici tly\npresented in Section III. Section IV presents a case study by\nusing a modified IEEE 39-bus test system. Conclusions are\ngiven in Section V .\nNotations\nDenoteRandCas the set of real numbers and complex\nnumbers, respectively. An m-dimensional vector is denoted\nFigure 1. The control block diagram of the proposed DCU.\nasx= [xi]∈Rm. The transpose of a vector or a matrix\nis defined as (·)T. The notation Im∈Rm×mdenotes the\nidentity matrix, 0is a zero vector or matrix with an appropriate\ndimention, and ei∈Rpdenotes the vector with the ithentry\nbeing one and others being zeros. The notation |x|(∠x) takes\nthe modulus (angle) of a complex number x∈C. The notation\nV(A)means converting the matrix A= [a1,...,ap]∈Rm×p\nwithai∈Rmwithi= 1,...,p to a vector, i.e., V(A) =\n[aT\n1,...,aT\np]T∈Rmp. The symbols /bardbl · /bardbl and/bardbl· /bardbl∞denote\nthel2andl∞norms for a vector, respectively.\nII. N ETWORK DESCRIPTION\nIn this section, we firstly introduce the DCU proposed for\neach damping controller. Then, the power system network to\nbe studied is introduced, which will be used to design the\ncontrol framework in Section III.\nA. Distributed Damping Control Unit\nFig. 1 shows the block diagram of the proposed DCU\nwhich mimics the structure of PSS. The input is the local bus\nvoltage angle θi, and the output is posciwhich is added to the\nreference active power demand of the WTG or LA to provide\nsupplementary damping control services. The mathematical\nmodel of the ithDCU is given by\n˙x1i=−1\nTwi(Kiθi+x1i)\n˙x2i=1\nT2i/parenleftbigg\n(1−T1i\nT2i)(Kiθi+x1i)−x2i/parenrightbigg\n˙x3i=1\nT4i/parenleftbigg\n(1−T3i\nT4i)/parenleftbigg\nx2i+/parenleftbiggT1i\nT2i(Kiθi+x1i)/parenrightbigg/parenrightbigg\n−x3i/parenrightbigg\nposci=x3i+T3i\nT4i/parenleftbigg\nx2i+T1i\nT2i(Kiθi+x1i)/parenrightbigg\n.\n(1)\nThe dynamics can be written in a compact form as ˙xCi=\nfCi(xCi,θi)wherexCi= [x1i,x2i,x3i]Tis the supplement-\nary state variables, Kiis the gain, Twiis the wash-out time\nconstant,T1i,T2i,T3i, andT4iare time constants for lead-lag\ncompensation. In the proposed control framework, Ki,T1i,\nT2i,T3i, andT4iwill be tuned to improve the damping ratio\nof the critical LFO.\nB. Power system network\nConsider a connected power system consisting of Nbuses\nwithNGsynchronous generators (SGs), NWWTGs,NL\nloads, andNTtransfer buses where N=NG+NW+NL+NT.\nThe SG (WTG or load) bus refers to a bus that connects a\nSG (WTG or load) only. The transfer bus is a bus with no3\ngeneration or demand. We number the SG buses as VG=\n{1,...,N G}, WTG buses as VW={NG+1,...,N G+NW},\nload buses as VL={NG+NW+1,...,N G+NW+NL},\nand transfer buses as VT={NG+NW+NL+1,...,N}.\n1) SG model: To highlight the effectiveness of the proposed\ndamping controllers, PSSs are not included in the SG models.\nWith the 4th-order two-axis synchronous machine model and\nIEEE standard exciter model (IEEET1), the mathematical\nmodel of the ithSG is written as:\n˙xGi=fGi(xGi,θi,vi)\npGi=gpGi(xGi,θi,vi)\nqGi=gqGi(xGi,θi,vi), i∈ VG(2)\nwhere the state variable xGiis defined as xGi=\n[e′\nqi,e′\ndi,δi,ωi,xmi,xr1i,xr2i,xfi]T;e′\nqiande′\ndiare transient\nd-axis and q-axis voltages, respectively; δiandωiare the\nrotor angle and speed, respectively; xmi,xr1i,xr2iandxfi\nare the state variables corresponding to the IEEET1 exciter .\nThe algebraic variables are the local bus voltage angle θiand\nmagnitudevi. The active and reactive power injections of the\nithSG bus are denoted as pGiandqGi, respectively. The\ndetailed descriptions of nonlinear functions fGi,gpGi,gqGi\ncan be found in [14], which is given in the Appendix A for\nself-completeness.\n2) WTG model: Fully rated converter WTGs are adopted,\nwhich employ the configuration of a synchronous machine\nwith a permanent magnet rotor [15]. Normally, the controlle r\nof WTG gives a reference active power demand to optimize\nthe wind energy capture based on the measured rotor speed\n(see the lower branch in Fig. 2). In this paper, two additiona l\ncontrol units are added into the original WTG’s controller t o\nadapt the active power reference set point, i.e., the primar y\nfrequency support unit proposed in [16] (see the upper branc h\nin Fig. 2) and the DCU (see the middle branch in Fig. 2). The\nmathematical model of the ithWTG is written as:\n˙θi=ωi\n˙xWi=fWi(xWi,ωi,θi,vi)\npWi=gpWi(xWi,ωi,θi,vi)\nqWi=gqWi(xWi,ωi,θi,vi), i∈ VW(3)\nwhereωiis the local bus frequency. The state variable\nxWi= [ωmi,θpi,isqi,icdi,xT\nCi]Twhereωmiis the rotor\nspeed;θpiis the pitch angle used for maximum power control;\niqsiis the generator stator quadrature current used for active\npower/speed control; and icdiis the converter direct current\nused for reactive power/voltage control; xCi= [x1i,x2i,x3i]T\nare state variables corresponding to the DCU. The active and\nreactive power injections of the ithWTG bus are denoted\naspWiandqWi, respectively. The detailed descriptions of\nnonlinear functions fWi,gpWi,gqWiare given in the Appendix\nB.\n3) Load model: The active power of each load pLiis\ndivided into two parts: the controllable part di=posci(xLi,θi)\n(referred to as LA in this paper) and static voltage frequenc y\ndependent part, whereas the reactive power of each load qLi\nis assumed to be static voltage frequency dependent. With th e\nFigure 2. Control block diagram of the controller for WTG.\nadditional DCU, the mathematical model of the ithload bus\nis given as follows:\n˙θi=ωi\n˙xLi=fLi(xLi,θi)\npLi=poi(vi)αi(1+kpfi(ωi−ω0))+di(xLi,θi)\n:=gpLi(xLi,ωi,θi,vi)\nqLi=qoi(vi)βi(1+kqfi(ωi−ω0))\n:=gqLi(ωi,vi), i∈ VL(4)\nwhere the state variable xLi= [x1i,x2i,x3i]Tcorresponds to\nthe DCU;poiandqoiare the nominal values; αiandβiare\nvoltage coefficients; kpfiandkqfiare frequency coefficients.\n4) Transfer bus: As transfer buses have no generations or\nloads, theithtransfer bus is simply modeled as:\npTi= 0, qTi= 0, i∈ VT (5)\nwherepTiandqTiare the active and reactive power injections,\nrespectively.\n5) Network power flows: The network power flows are\nrepresented by the usual set of algebraic power flow equation s,\nwhich are used to couple all buses power injection equations\nmentioned above. For the ithbus in the system, the power\nflow equations are given as:\n0 =−pinj\ni+viN/summationdisplay\nj=1vj(Gijcosθij+Bijsinθij)\n0 =−qinj\ni+viN/summationdisplay\nj=1vj(Gijsinθij−Bijcosθij), i∈ V(6)\nwhereGijandBijare the real and imaginary parts of Yij\nwhich is the (i,j)entry of the admittance matrix Y; the nota-\ntionθijis the short for θi−θj; the setV=VG∪VW∪VL∪VT;\npinj\niandqinj\niare injected active and reactive power of the ith\nbus, respectively. In particular, for SG buses, pinj\ni=pGiand\nqinj\ni=qGi; for WTG buses, pinj\ni=pWiandqinj\ni=qWi; for\nload buses, pinj\ni=−pLiandqinj\ni=−qLi; and for transfer\nbuses,pinj\ni=pTiandqinj\ni=qTi.\n6) Overall system: Combining (2)-(6), the overall system\ncan be expressed as differential-algebraic equations:\n˙x=f(x,y)\n0=h(x,y)(7)\nwhere the vector x= [xT\nG,θT\nW,xT\nW,θT\nL,xT\nL]Tand the\nvectory= [θT\nG,ωT\nW,ωT\nL,θT\nT,vT\nG,vT\nW,vT\nL,vT\nT]T;xk=\n[xT\nk1,...,xT\nki,...,xT\nkNk]T, i∈ Vk,k∈ {G,W,L};θk=4\n[θi]∈RNk, i∈ Vk,k∈ {G,W,L,T };vk= [vi]∈\nRNk, i∈ Vk,k∈ {G,W,L,T };ωk= [ωi]∈RNk, i∈ Vk,\nk∈ {W,L}. The nonlinear functions fandhrepresent\nthe system dynamics and network power flow equations,\nrespectively.\nLinearizing system (7) gives the following linear model:\n/bracketleftbigg∆˙x\n0/bracketrightbigg\n=/bracketleftbiggAsBs\nCsDs/bracketrightbigg/bracketleftbigg∆x\n∆y/bracketrightbigg\n(8)\nwhere the detailed expressions of the matrices As,Bs,Cs,\nandDsare given in the Appendix C. Assuming Dsis nonsin-\ngular (it is a common assumption adopted in the literature [1 7])\nand eliminating ∆yin (8) gives:\n∆˙x=Ar∆x (9)\nwhereAr=As−BsD−1\nsCs∈RNA×NAwithNA= 8NG+\n8NW+4NL.\nIII. D ISTRIBUTED CONTROL FRAMEWORK\nIn this section, the critical LFO identification module and\ncontroller parameters tuning module that form the proposed\ndistributed control framework will be introduced in detail s.\nA. Critical LFO identification module\nAs mentioned earlier, this module aims at identifying the\ncritical LFO for each damping bus in a distributed manner.\nIt is known that the critical LFO can be investigated by\napplying eigenvalue analysis based on the global informati on,\ni.e., the system Jacobian matrix Arin (9) which is usually\nobtained in a centralized manner [1]. However, in this paper ,\nwe will reconstruct the matrix Arfor each damping bus in a\ndistributed manner by revealing the structure properties o f the\npower grid contained in the matrices As,Bs,Cs, andDs.\nBy performing an elementary column operation, the\nmatricesAs,Bs,Cs, andDscan be reformulated as:\n/bracketleftbiggAsBs\nCsDs/bracketrightbigg\n=/bracketleftbiggK1K2\nK3Jpf/bracketrightbigg\nT (10)\nwhere the matrix Tis the elementary column operator and Jpf\nis the power flow Jacobian matrix. The detailed expressions\nof the matrices T,K1,K2,K3, andJpfare given in the\nAppendix C.\nThrough the matrix transform (10), we can see that the\nmatricesAs,Bs,Cs, andDscan be reconstructed by all\nJacobian matrices K∧\n∨(refer to (52) in the Appendix C for\ndetails),Jpf, identity matrices, and T. For identity matrices\nandT, since they are constant, they can be broadcasted or\nstored at each damping bus in advance. For all Jacobian\nmatricesK∧\n∨(all are block diagonal matrices) and Jpf, we\nadopt the distributed algorithm proposed in [17] that has to tal\n2Nsteps to calculate their elements, where the communication\nnetwork used covers all buses in the system and has the same\ntopology as the physical grid. The communication network ca n\nbe described by the undirected graph G1={V,E}, whereV\nis the set of nodes (buses) and E ⊆ V×V represents the set of\nedges (branches). The set of neighbors of node iis represent\nbyNi={j∈ V: (j,i)∈ E} with cardinality |Ni|=Di.We assume that 1) each bus knows the parameters of its local\nmachine (or load) and lines connecting it; 2) each damping\nbus knows the model structure of SG, WTG, and load; 3)\neach bus knows its own bus number and total number of\nbusesN; 4) each bus in the system has the capability of local\nmeasurement, storing data, processing data, communicatin g\nwith its neighbors, and calculation; and 5) communication\ndelays are negligible.\nAt each step, bus i, i∈ V has four columns of data for\ncommunication, denoted as γa\ni,̟a\ni,γb\ni,̟b\ni∈R2N. The\ndata update process is designed as follows:\n[Xa(τ+1),Xb(τ+1)] =Jpf[Xa(τ),Xb(τ)] (11)\nwhereXa(τ),Xb(τ)∈R2N×2Nare the data matrices at the\nτthstep iteration with the definitions as follows:\nXa(τ) = [γa\n1(τ),...,γa\nN(τ),̟a\n1(τ),...,̟a\nN(τ)]T\nXb(τ) = [γb\n1(τ),...,γb\nN(τ),̟b\n1(τ),...,̟b\nN(τ)]T(12)\nwhich are initialized by Xa(0) =I2NandXb(0) =\n[γb\n1(0),...,γb\nN(0),̟b\n1(0),...,̟b\nN(0)]T. The vectors γb\ni(0)\nand̟b\ni(0)assigned to the ithbus satisfies:\n[γb\ni(0);̟b\ni(0)] = [V(KfGixGi);V(KfGi\nθi);V(KfGivi);V(KhpGixGi);\nV(KhqGixGi);0], i∈ VG;\n[γb\ni(0);̟b\ni(0)] = [V(KfWi\nθi);V(KfWixWi);V(KfWiωi);V(KfWivi);\nV(KhpWixWi);V(KhpWiωi);V(KhqWixWi);\nV(KhqWiωi);0], i∈ VW;\n[γb\ni(0);̟b\ni(0)] = [V(KfLi\nθi);V(KfLixLi);V(KhpLixLi);V(KhpLiωi);\nV(KhqLiωi);0], i∈ VL;\n[γb\ni(0);̟b\ni(0)] = [0], i∈ VT.\n(13)\nThe designed update process (11) can be realized in a dis-\ntributed manner via the communication network G1={V,E}\nmentioned earlier since\n1) the initial values of vectors γa\ni(0),̟a\ni(0),γb\ni(0), and\n̟b\ni(0)can be assigned locally for each bus i, because\ni) the vectors γa\ni(0),̟a\ni(0)can be assigned locally as\neach bus knows its own bus number and ii) the elements\nof vectorsγb\ni(0),̟b\ni(0)can be calculated based on local\nmeasurements θi,vi,pinj\niandqinj\ni[14], [15];\n2) for each sub-matrix Jhp\nθ,Jhp\nv,Jhq\nθ,Jhq\nvofJpf(see\n(52) and (53) in Appendix C for details), the nonzero\nelements of the ithrow are functions of measurements\nof busiand its neighboring bus j∈ Ni[17], [18].\nDuring the update process, at each step τ,0< τ≤2N,\neach damping bus i, i∈ VW∪ VLstores its own data and\ndata from its neighboring buses (which can be realized via\ncommunication links between neighboring buses). Thus, the\nwhole distributed algorithm is expressed as:\n[Xa(τ+1),Xb(τ+1)] =Jpf[Xa(τ),Xb(τ)]\n[ξa\ni(τ),ξb\ni(τ)] =Si[Xa(τ),Xb(τ)], i∈ VW∪VL(14)\nwhere the matrix Si= [ei,ej,eN+i,eN+j]T∈R2(Di+1)×2N,\nj∈ Niselects the rows with respect to the damping bus iand5\nits neighboring buses j, j∈ Ni;ξa\ni(τ),ξb\ni(τ)∈R2(Di+1)×2N\ndenote the data collected by the damping bus i. We assume\nthe discrete-time system (14) is observable, which usually\nholds in practice [17], i.e., rank(Oi) = 2NwhereOi∈\nR4(Di+1)N×2Nis defined as\nOi= [ST\ni,(SiJpf)T,...,(SiJ2N−1\npf)T]T. (15)\nAfter the update process (14), each damping bus i, i∈\nVW∪VLcan recoverJpfandXb(0)via the data it collected\nξa\ni(τ)andξb\ni(τ), τ= 0,1,...,2N. For simplicity, we define\nthe following data matrices:\nΞa\ni1= [ξa\ni(0)T,...,ξa\ni(2N−1)T]T∈R4(Di+1)N×2N\nΞa\ni2= [ξa\ni(1)T,...,ξa\ni(2N)T]T∈R4(Di+1)N×2N\nΞa\ni= [Ξa\ni1T,Ξa\ni2T]T∈R8(Di+1)N×2N\nΞb\ni1= [ξb\ni(0)T,...,ξb\ni(2N−1)T]T∈R4(Di+1)N×2N.(16)\nThe singular value decomposition of Ξa\niis also needed, which\nis given as:\nΞa\ni= [˜Uξi,˜U0\nξi]/bracketleftbigg\nΣξi\n0/bracketrightbigg\n˜VT\nξi=˜UξiΣξi˜VT\nξi(17)\nwhereΣξi,˜Vξi∈R2N×2N,˜Uξi∈R8(Di+1)N×2N,˜U0\nξi∈\nR8(Di+1)N×(8(Di+1)N−2N). Based on the matrices given in\n(16) and (17), each damping bus i, i∈ VW∪VLcan recover\nJpfandXb(0)by the following equations:\nJpf= (˜UT\nξi1Ξa\ni1)−1Θi˜UT\nξi1Ξa\ni1 (18a)\nXb(0) = (Ξa\ni1)†Ξb\ni1 (18b)\nwhere˜Uξi1,˜Uξi2∈R4(Di+1)N×2Nare sub-matrices of ˜Uξi\nwith˜Uξi= [˜UT\nξi1,˜UT\nξi2]T,Θi= (˜UT\nξi1˜Uξi2)(˜UT\nξi1˜Uξi1)−1∈\nR2N×2N, and the superscript †denotes the Moore-Penrose\ninverse. The mathematical proof of (18a)-(18b) can be found\nin [17].\nAs mentioned earlier, each damping bus is assumed to\nknow the model structures of SG, WTG, and load. Thus, each\ndamping bus can identify the type of bus i(i.e, SG, WTG,\nload, or transfer bus) based on the γb\ni(0)and̟b\ni(0)ofXb(0)\nobtained, and hence can recover all K∧\n∨Jacobian matrices\ninK1,K2, andK3of (10) fromXb(0)obtained based on\n(13). Combined with Jpfobtained, each damping bus can\nreconstructArby (9) and (10). Therefore, the critical LFO\ncan be calculated by applying eigenvalue analysis to Arat\neach damping bus.\nRemark 1: In the proposed update process (14), we assume\nthat the sum of the length of all vectorized K∧\n∨matrices related\nto each type of bus (i.e., SG, WTG, load, or transfer bus) is le ss\nthan the length of the data vectors [γb\ni;̟b\ni], i∈ V assigned\nfor each type of bus that is 4N(refer to (13) for details).\nIf there exist one type of bus whose sum of the length of\nall vectorized K∧\n∨matrices is more than 4N, additional data\nvectorsγc\ni,̟c\ni∈R2Nare assigned for each bus to form the\nadditional data matrix Xc∈R2N×2N. For the type of bus\nwhose sum of the length of all the vectorized K∧\n∨matrices is\nmore than 4N,[γc\ni(0);̟c\ni(0)] is initialized by the remaining\nelements. For the other types of buses whose sum of the length\nof the vectorized K∧\n∨matrices is less than 4N,[γc\ni(0);̟c\ni(0)]\nFigure 3. The closed-loop representation of the system.\nis initialized by zeros. The additional data matrix Xc(0)can\nbe recovered by each damping bus via the same way as the\ndata matrixXb(0)is recovered.\nB. Controller parameters tuning module\nIn order to guarantee an adequate stability margin, the\ndamping ratio ςcof the critical LFO λc=σc+jωcshould\nsatisfyςc≥ς⋆whereςc=−σc/|λc|andς⋆>0is the preset\nthreshold. Once the damping ratio of the critical LFO is less\nthanς⋆, the parameters of each DCU (i.e., Ki,T1i,T2i,T3i,\nandT4i) of each damping bus will be tuned coordinately to\nimprove the damping ratio of the critical LFO.\nWithout loss of generality, we firstly study the impact of\nthe parameter changes of the ith, i∈ VW∪ VLDCU (i.e.,\nthe DCU of the bus NG+i) onλc. For analysis purposes,\nthe system model (8) is rewritten in the following form by\nreordering the variables of xin (8):\n/bracketleftbigg\n∆˙˜xi\n0/bracketrightbigg\n=/bracketleftbigg˜Asi˜Bsi\n˜Csi˜Dsi/bracketrightbigg/bracketleftbigg∆˜xi\n∆y/bracketrightbigg\n(19)\nwhere∆˜xi= [∆xT\ni,∆xT\nCi]T,xi∈RNA−3includes all state\nvariables inxexceptxCi∈R3that is the corresponding state\nof theithDCU;˜Asi=T−1\niAsTi(hereTi∈RNA×NAis\ninvertable which is the corresponding elementary row opera tor\nsuch thatx=Ti˜xi);˜Bsi=T−1\niBs;˜Csi=CsTi; and\n˜Dsi=Ds. Then the system model (19) can be written in the\nclosed-loop form [19]. In the closed-loop form, the system\nmodel is partitioned into two subsystems. For subsystem 1,\nwhich does not depend on parameters of the ithDCU, we\nhave the following state space description:\n/bracketleftbigg∆˙xi\n0/bracketrightbigg\n=/bracketleftbiggAiBi\nCiDi/bracketrightbigg/bracketleftbigg∆xi\n∆y/bracketrightbigg\n+/bracketleftbiggEi\nFi/bracketrightbigg\n∆ui.(20)\nwhereui=posciis the output of the ithDCU. Assuming Di\nis nonsingular and eliminating ∆yin (20) gives:\n∆˙xi=Asi∆xi+Bsi∆ui; ∆θi=Csi∆xi (21)\nwhereAsi=Ai−BiD−1\niCi∈R(NA−3)×(NA−3),Bsi=\nEi−BiD−1\niFi∈RNA−3andCT\nsi∈RNA−3. For subsystem\n2, which only depends on the parameters of the ithDCU, we\nhave the following state space description:\n/bracketleftbigg∆˙xCi\n∆ui/bracketrightbigg\n=/bracketleftbiggACiBCi\nCCiDCi/bracketrightbigg/bracketleftbigg∆xCi\n∆θi/bracketrightbigg\n. (22)\nwhereACi,BCi,CCi, andDCican be easily obtained from\n(1). A transfer function description for (22) is given as:\nFi(s,Ki) =CCi(sI−ACi)−1BCi+DCi (23)6\nwhereKi∈Ris the gain factor in the ithDCU model. Based\non (21) and (23), the schematic diagram of the closed-loop\nform is given in Fig. 3.\nThen the sensitivity of λcwith respect to Kiof the transfer\nfunctionFi(s,Ki)is given by [19]:\n∂λc\n∂Ki=Ri∂Fi(s,Ki)\n∂Ki/vextendsingle/vextendsingle/vextendsingle/vextendsingle\ns=λc(24)\nwhereRi=CsiφsiψT\nsiBsi∈Cis the residue with respect to\nthe critical eigenvalue λc;φsi∈RNA−3andψsi∈RNA−3\nare the right and left eigenvectors of λc, respectively. Here,\nφsi(ψsi) consists of the first NA−3elements ofφi∈RNA\n(ψi∈RNA) which is the right (left) eigenvector of λcwith\nrespect to ˜Arithat is obtained by eliminating ∆yin (19), i.e.,\n˜Ari=˜Asi−˜Bsi˜D−1\nsi˜Csi=T−1\niArTi. (25)\nCombining (24) and the transfer function of DCU given in\nFig. 1, the sensitivity of λcwith respect to Kibecomes:\nsi=∂λc\n∂Ki=Ri·10λc\n1+10λc·1+T1iλc\n1+T2iλc·1+T3iλc\n1+T4iλc.(26)\nHere, the wash-out time constant of each DCU is assumed to\nbe 10, i.e.,Twi= 10 .\nIt follows from (26) that the tuning process of DCUs\ncan be split into two parts: 1) tuning parameters of lead-lag\ncompensation of the ithDCU such that ∠si= 180◦; and then\n2) tuning gain factors Kiof all DCUs such that λcmoves\nto the desired location, i.e.,/summationtextNW+NL\ni=1|si|∆Ki≥∆ℜ(λc)⋆\nwhere∆ℜ(λc)⋆=ς⋆|ωc|//radicalbig\n1−(ς⋆)2+σcis the expected\nreal part change of λc. For part 1), the parameters of T1i,T2i,\nT3i, andT4ican be calculated by [20]\n\n\nαi= (1+sin( ∠Ki)/2)/(1−sin(∠Ki)/2)\nT1i=T3i= (√αi)/ωc\nT2i=T4i= 1/(√αiωc)(27)\nwhere∠Ki= 180◦−∠Ri. For part 2), the gain factor change\n∆Kiis calculated by solving the following optimization\nproblem:\nminNW+NL/summationdisplay\ni=1ci (28)\ns.t.NW+NL/summationdisplay\ni=1|si|∆Ki≥∆ℜ(λc)⋆(29)\n∆Kmin\ni≤∆Ki≤∆Kmax\ni,i= 1,...,N W+NL\n(30)\nwhere∆Kmin\ni and∆Kmax\ni are the lower and upper bounds\non the gain factor of the ithDCU, respectively. To account for\nthe damping controller adjustments, in this work, we introd uce\na simple quadratic cost function for the ithdamping bus\nwhich has been widely used in the literature (e.g., [12]), i. e.,\nci=πi∆K2\niandπi>0is the cost parameter assigned for\ntheithdamping bus. The objective (28) is to minimize the\ntotal control cost. For convenience, the convex optimizati on\nproblem (28)-(30) is rewritten in a compact form as:\nmin\n∆KNW+NL/summationdisplay\ni=1ci(∆Ki)s.t. g(∆K)≤0,∆Ki∈∆ˆKi(31)where∆K= [∆K1,...,∆KNW+NL]Tdenotes the gain\nfactor changes of NW+NLDCUs;g(∆K)≤0represents the\nglobal constraint in (29); ∆ˆKirepresents the local constraint\nin (30).\nAs mentioned earlier, in this module, the proposed two-part\ntuning process will be realized in a distributed manner. For\nthe first part tuning process, it is realized locally as the Ri\nrequired of the ithdamping bus can be obtained locally. It\nfollows from (24) that Rican be calculated by ˜Ari,Bsi, and\nCsi. ForCsi, based on (21), it can be easily obtained as the\nithdamping bus knows the order of variables in xi. For˜Ari,\nit can be calculated by (25) as Tiis known locally and Ar\nhas been obtained in the critical LFO identification module\nfor each damping bus. For Bsi, based on (19)-(22), we have\n˜Asi=/bracketleftbiggAi+EiDCiCsiEiCCi\nBCiCsiACi/bracketrightbigg\n,˜Bsi=/bracketleftbiggBi\n0/bracketrightbigg\n˜Csi=/bracketleftbig\nCi+FiDCiCsiFiCCi/bracketrightbig\n,˜Dsi=Di.\n(32)\nTheBsican be obtained by (21) locally as: 1) matrices ACi,\nBCi,CCi, andDCiis known locally, 2) according to (19),\nmatrices ˜Asi,˜Bsi,˜Csi, and˜Dsican be calculated based\nonAs,Bs,Cs, andDswhich have been obtained by each\ndamping bus in the critical LFO identification module, and 3)\nCsican be obtained locally, then based on the matrix relations\nin (32), matrices Ai,Bi,Ci,Di,Ei, andFican be calculated.\nFor the second part tuning process, in order to solve the\nconvex optimization problem (28)-(30) in a distributed man ner,\nwe decompose the Lagrange function of (31) into a sum of\nNW+NLlocal Lagrange functions where each of them is\nassigned to a damping bus:\nL(∆K,µ) =NW+NL/summationdisplay\ni=1Li(∆K,µ) (33)\nwhereLi(∆K,µ) =ci(∆Ki) +µg(∆K), scalarµis the\nLagrange multiplier for g(∆K)≤0in (31).\nInspired by (33), based on the distributed Lagrangian\nprimal-dual sub-gradient algorithm proposed in [21], a dis -\ntributed algorithm is designed to update the decision varia bles\n∆Kand Lagrangian multiplier µvia communication between\nneighboring damping buses. The communication network used\nonly covers damping buses and is allowed to have a different\ntopology from the physical grid, which can be described by\nthe undirected graph G2={V2,E2,W}, whereV2=VW∪VL,\nE2⊆ V2× V2, andW={wij} ∈R(NW+NL)×(NW+NL). If\n(i,j)∈ E2,i/ne}ationslash=j, thenwij=wji>0and/summationtextNW+NL\nj=1,i/negationslash=jwij<1;\notherwise,wij=wji= 0. We define the diagonal entry wii\nof the matrix Waswii= 1−/summationtextNW+NL\nj=1,i/negationslash=jwij. In the proposed\ndistributed algorithm, the following assumptions are adop ted:\n1) The function gin (31) is known to all damping buses.\n2) The topology of the communication network G2is\nundirected and connected, and communication delays\nare negligible.\nFor assumption 1), since As,Bs,Cs, andDshave been ob-\ntained by each damping bus via the critical LFO identificatio n\nmodule, then all sensitivities siin functiongcan be calculated7\nlocally for each damping bus via the same method used for\ncalculatingRiin the first part tuning process.\nBased on the abovementioned assumptions, the update pro-\ncess of decision variables ∆Kand Lagrangian multiplier µis\nexpressed as follows:\n∆Ki(τ+1) =P∆ˆKi[∆¯Ki(τ)−ς(τ)DLi,∆¯Ki(τ)]\nµi(τ+1) =PˆUi[¯µi(τ)+ς(τ)DLi,¯µi(τ)](34)\nwhere∆Ki∈RNW+NLandµi∈Rare the information\ndata assigned for the ithdamping bus. We use ∆¯Ki(τ) =/summationtextNW+NL\nj=1wij∆Kj(r)and¯µi(r) =/summationtextNW+NL\nj=1wijµj(r)for\nshort. At each time τ+ 1, theithdamping bus calculates\nvectorsDLi,∆¯Ki=∂Li/∂(∆¯Ki)andDLi,¯µi=∂Li/∂¯µi\nin the gradient direction of its local Li. Combined with\ninformation received from its neighboring buses ∆¯Ki(τ)and\n¯µi(τ), theithdamping bus updates its own decision variables\n∆Ki(τ+ 1) andµi(τ+ 1) by taking a projection onto its\nlocal constraint ∆ˆKiandˆUi={µi≥0}, respectively. Here,\nthe projection operator P∆ˆKiis defined by the definition\nofP∆ˆKi[¯x] = argminx∈∆ˆKi/bardbl¯x−x/bardbl, where¯xis a given\nvector. The projection operator PˆUiis defined in the same way\nasP∆ˆKi. The diminishing step size is ς(r)which satisfies\nlimr→+∞ς(r) = 0 ,/summationtext+∞\nr=0ς(r) = +∞, and/summationtext+∞\nr=0ς(r)2<\n+∞. It has been proven in [21] that for a convex optimization\nproblem, the proposed distributed algorithm will asymptot ic-\nally converge to a pair of primal-dual optimal solutions (i. e.,\nlimτ→∞∆Ki(τ) = ∆K∗, i= 1,...,N W+NLwhere\n∆K∗= [∆K∗\n1,···,∆K∗\nNW+NL]Tis the optimal solution)\nunder the Slater’s condition, assumptions 1) and 2) mention ed\nabove. In our case, the optimization problem (28)-(30) is a\nconvex optimization program whose global optimal solution s\ncan be solved in a distributed way via the algorithm (34).\nIt is worth mentioning that, different damping buses have\ndifferent geometric controllability/obserbility measur es (COs)\nof the critical LFO λcunder different operating conditions\n[10]. The definition of the CO of the ithdamping bus is given\nasCOi=|ψT\nsiBsi|\n||ψsi||||Bsi||·|Csiφsi|\n||Csi||||φsi||which can be calculated\nlocally. In the proposed two-part tuning process, only the\ndamping buses with high COs participate the tuning process.\nIn other words, if the CO of the ithdamping bus satisfies\nCOi< CO⋆whereCO⋆is a threshold, then this damping\nbus does not participate the first part tuning process and the\nsecond part tuning process by setting ∆Kmin\ni= ∆Kmax\ni= 0\nin (30) locally.\nIV. C ASE STUDY\nIn this section, the modified IEEE 39-bus test system used\nfor simulation is introduced firstly. Then, the simulation r esults\nand explanations will be presented.\nA. Test system\nFig. 4 shows the modified IEEE 39-bus test system that is\nused to demonstrate the proposed distributed control frame -\nwork. In the modified 39-bus system, the SG at bus 10 is\nreplaced by a FRC-WTG with the same size of maximum\npower generation. All buses are renumbered according to the\nFigure 4. The modified IEEE 39-bus system\nrules described in Section II-B, i.e., NG= 9 ,NW= 1 ,\nNL= 17 , andNT= 12 . The damping buses considered\nare 1 WTG bus and 17 load buses. The model and system\nparameters are taken from [22]. For model parameters that\nare not provided in [22], we use the default values of models\ngiven in the library developed in PSAT/MATLAB [23].\nFor the communication network G2used in the control\nparameters tuning module, each edge is assigned a weight\nwhich can be calculated by a simplified computational method :\nwij=1\n1+max{˜Di,˜Dj}, i∈ V2, j∈˜Ni (35)\nwhere˜Nidefines the set of adjacent damping buses of the ith\ndamping bus with the definition of ˜Ni={j∈ V2: (j,i)∈ E2}\nand cardinality |˜Ni|=˜Di.\nB. Simulation results\nFollowing the procedure described in Section III-A, the cri t-\nical LFO identified by each damping bus is −0.0476±j1.7311\nwith the damping ratio ςc= 0.0275 (the preset threshold\nς⋆= 0.1) and oscillation frequency equal to 0.275Hz, where\nG2-G9 oscillate against G1 (see Fig. 6(a)). Then, the contro ller\nparameter tuning module is activated to tune the correspond ing\nparameters as described in Section III-B. The price paramet ers\nneeded for the optimization problem (28)-(30) are given in\nTable I. In this case study, for simplicity, we assume the\ngain limits for DCUs are the same (i.e. ∆Kmin=−10\nand∆Kmax= 60 ). As mentioned in Section III-B, only the\ndamping buses with high COs participate the tuning process.\nFig. 4 shows the COs of all 18 damping buses, and the\nthresholdCO⋆= 10−4. It follows from Fig. 4 that buses\n10, 17, 20, and 21 participate in the parameter tuning proces s.\nThe obtained optimal gain factor changes are also given in\nTable I. The Table II compares the original λc, expectedλc,\nand the new λc. It can be seen from Table II that the critical\nLFO is stabilized as desired.\nTo illustrate the effectiveness of the proposed distribute d\ncontrol framework, we investigate the variation of rotor an gle\nof G1 after a three-phase fault before and after the proposed\ntuning process. The three-phase fault happens at 1 s for 0.1\nseconds on bus 25. From Fig. 6(b) we can see that the system8\n10111213141516171819202122232425262700.511.5x 10−4\nBus numberControllability/obserbility (CO) measures\nFigure 5. CO measures of all damping buses\nTable I\nTHE PARAMETERS FOR DAMPING CONTROLLERS\nBus no. π∆K⋆Bus no. π∆K⋆\n10 0.8692 37 19 0.8524 0\n11 0.9566 0 20 0.9367 60\n12 0.7578 0 21 0.7306 33\n13 1.2769 0 22 0.9391 0\n14 0.8650 0 23 0.7993 0\n15 1.3035 0 24 1.1363 0\n16 1.3578 0 25 1.0443 0\n17 0.9937 42 26 0.7898 0\n18 1.0715 0 27 0.9862 0\nTable II\nORIGINAL ,EXPECTED ,AND NEW λc\nOriginalλc Expected λc Newλc\n−0.0476±j1.7311−0.1749±j1.7311−0.1785±j1.7340\n(a) Mode shape (b) Rotor angle response\n0 10 20 30-0.148-0.146-0.144-0.142-0.14-0.138-0.136-0.134-0.132\nTime (s)Rotor angle of G1 (p.u.)Before tuning\nAfter tuning\n0.51\n30\n21060\n24090\n270120\n300150\n330180 0G1G2-G9\nFigure 6. (a) Compass plot of relative mode shape (ref. G6) an d (b) rotor\nangle of G1 responses to a three-phase fault on bus 25\nperformance is improved with the presence of the proposed\ndistributed control framework.\nV. C ONCLUSION\nIn this paper, WTGs and LAs have been coordinated\nto provide damping torques for the critical low frequency\noscillation by adapting their active power generations and\nconsumptions, respectively. In order to provide a scalable\ncontrol framework for the increasing number of WTGs andLAs, a novel distributed control framework has been propose d\nwhich consists of a critical LFO identification module and a\ncontroller parameters tuning module. The simulation resul ts\nhave shown that the proposed distributed control framework\nis feasible and effective.\nREFERENCES\n[1] P. Kundur, Power System Stability and Control . New York: McGraw-\nHill, 1994.\n[2] M. Garmroodi, D. J. Hill, G. Verbic, and J. Ma, “Impact of t ie-line\npower on inter-area modes with increased penetration of win d power,”\nIEEE Trans. Power Syst., vol. 31, no. 4, pp. 3051-3059, Jul. 2016.\n[3] M. Singh, A. J. Allen, E. Muljadi, V . Gevorgian, Y . Zhang, and S.\nSantoso, “Interarea oscillation damping controls for wind power plants,”\nIEEE Trans. Sustain. Energy, vol. 6, no. 3, pp. 967-975, Jul. 2015.\n[4] A. E. Leon and J. A. Solsona, “Power oscillation damping i mprovement\nby adding multiple wind farms to wide-area coordinating con trols,” IEEE\nTrans. Power Syst., vol. 29, no. 3, pp. 1356-1364, May 2014.\n[5] T. Liu, D. J. Hill, and C. Zhang, “Non-disruptive load-si de control for\nfrequency regulation in power systems,” IEEE Trans. Smart Grid, vol. 7,\nno. 4, pp. 2142-2153, Jul. 2016.\n[6] Z. Tang, D. J. Hill, T. Liu, and H. Ma, “Hierarchical volta ge control of\nweak subtransmission networks with high penetration of win d power,”\nIEEE Trans. Power Syst., vol. 33, no. 1, pp. 187-197, Jan. 2018.\n[7] B. Ramanathan and V . Vittal, “Small-disturbance angle s tability en-\nhancement through direct load control part I-framework dev elopment,”\nIEEE Trans. Power Syst., vol. 21, no. 2, pp. 773-781, May 2006.\n[8] X. Zhang, C. Lu, S. Liu, and X. Wang, “A review on wide-area\ndamping control to restrain inter-area low frequency oscil lation for large-\nscale power systems with increasing renewable generation, ”Renew. and\nSustain. Energy Rev., vol. 57, pp. 45-58, May. 2016.\n[9] R. A. Ramos, A. C. Martins, and N. G. Bretas, “An improved m ethodo-\nlogy for the design of power system damping controllers,” IEEE Trans.\nPower Syst., vol. 20, no. 4, pp. 1938-1945, Nov. 2005.\n[10] Y . Zhang and A. Bose, “Design of wide-area damping contr ollers for\ninterarea oscillations,” IEEE Trans. Power Syst., vol. 23, no. 3, pp. 1136-\n1143, Aug. 2008.\n[11] J. Deng, C. Li, and X. P. Zhang, “Coordinated design of mu ltiple robust\nFACTS damping controllers: A BMI-based sequential approac h with\nmulti-model systems,” IEEE Trans. Power Syst., vol. 30, no. 6, pp. 3150-\n3159, Nov. 2015.\n[12] S. P. Azad, R. Iravani, and J. E. Tate, “Damping inter-ar ea oscillations\nbased on a model predictive control (MPC) HVDC supplementar y\ncontroller,” IEEE Trans. Power Syst., vol. 28, no. 3, pp. 3174-3183, Aug.\n2013.\n[13] D. Z. Fang, X. Yang, T. S. Chung, and K. P. Wong, “Adaptive fuzzy-logic\nSVC damping controller using strategy of oscillation energ y descent,”\nIEEE Trans. Power Syst., vol. 19, no. 3, pp. 1414-1421, Aug. 2004.\n[14] P. W. Sauer and M. A. Pai, Power System Dynamics and Stability . Upper\nSaddle River, New Jersey: Prentice Hall, 1998.\n[15] A. Perdana, “Dynamic models of wind turbines,” Ph.D. di ssertation,\nChalmers Univ. Technol., Gotteborg, Sweden, 2008.\n[16] J. Morren, S. W. De Haan, W. L. Kling, and J. A. Ferreira, “ Wind\nturbines emulating inertia and supporting primary frequen cy control,”\nIEEE Trans. Power Syst., vol. 21, no. 1, pp. 433-434, Feb. 2006.\n[17] Y . Song, D. J. Hill, T. Liu, and Y . Zheng, “A distributed f ramework\nfor stability evaluation and enhancement of inverter-base d microgrids,”\nIEEE Trans. Smart Grid, vol. 8, no. 6, pp. 3020-3034, Nov. 2017.\n[18] Z. Tang, D. J. Hill, and T. Liu, “Fully distributed volta ge control in\nsubtransmission networks via virtual power plants,” in Proc. IEEE Int.\nConf. on Smart Grid Commun., pp. 193-198, 2016.\n[19] F. L. Pagola, I. J. Perez-Arriaga, and G. C. Verghese, “O n sensitivities,\nresidues and participations: applications to oscillatory stability analysis\nand control,” IEEE Trans. Power Syst., vol. 4, no. 1, pp. 278-285, Feb.\n1989.\n[20] L. Rouco and F. L. Pagola, “An eigenvalue sensitivity ap proach to\nlocation and controller design of controllable series capa citors for\ndamping power system oscillations,” IEEE Trans. Power Syst., vol. 12,\nno. 4, pp. 1660-1666, Nov. 1997.\n[21] M. H. Zhu and S. Mart´ ınez, “On distributed convex optim ization under\ninequality and equality constraints,” IEEE Trans. Autom. Control, vol.\n57, no. 1, pp. 151-164, Jan. 2012.9\n[22] “IEEE PES Task Force on Benchmark Systems for Stability Controls,”\nI. A. Hisken, 2013 [online]. Available: http://eioc.pnnl. gov/benchmark/\nieeess/index.htm\n[23] F. Milano, “An open source power system analysis toolbo x,”IEEE Trans.\nPower Syst., vol. 20, no. 3, pp. 1199-1206, Aug. 2005.\nAPPENDIX\nA. SG model\nWith4th-order two-axis synchronous machine model\nand IEEE standard exciter model (IEEET1), the resulting\ndifferential-algebraic equations for the ithSG bus are given\nas:\n1) Differential equations:\n˙e′\nqi=1\nT′\ndoi/parenleftbig\n−e′\nqi−(xdi−x′\ndi)idi+vfi/parenrightbig\n˙e′\ndi=1\nT′\nqoi/parenleftbig\n−e′\ndi−(xqi−x′\nqi)iqi/parenrightbig\n˙δi=ω0(ωi−1)\n˙ωi=1\nMi/parenleftbig\npmi−e′\ndiidi−e′\nqiiqi−(x′\nqi−x′\ndi)idiiqi\n−Di(ωi−1))\n˙vmi=1\nTri(vi−vmi)\n˙vr1i=1\nTai/parenleftbigg\nKai(vrefi−vmi−vr2i−Kfi\nTfivfi)−vr1i/parenrightbigg\n˙vr2i=−1\nTfi(Kfi\nTfivfi+vr2i)\n˙vfi=−1\nTei(vfi(Kei+Sei(vfi)−vri))\n(36)\nwhereω0is the base frequency, T′\ndoiandT′\nqoi;xdiandxqi;\nx′\ndiandx′\nqi;idiandiqiare the d-axis and q-axis transient\ntime constant; reactance; transient reactance; current, r espect-\nively;pmi,Di, andMiare the mechanical power, damping\ncoefficient, and moment of inertia, respectively; vfiandvrefi\nare the field and reference voltages, respectively; Tri,Tai,Tfi,\nandTeiare measurement, amplifier, stabilizer, and field circuit\ntime constants, respectively; Kai,Kfi, andKeiare amplifier,\nstabilizer, and field circuit gains, respectively; Seiis the ceiling\nfunction.\n2) Algebraic equations: The stator algebraic equations are\ngiven as:\npGi=idivisin(δi−θi)+iqivicos(δi−θi)\nqGi=idivicos(δi−θi)−iqivisin(δi−θi).(37)\nIn order to express network voltages in the polar form, idiand\niqiin (36) and (37) are expressed in terms of state variables\nxGiand algebraic variables vi,θi:\n/bracketleftbiggidi\niqi/bracketrightbigg\n=/bracketleftbiggrsi−x′\nqi\nx′\ndirsi/bracketrightbigg−1/bracketleftbigge′\ndi−visin(δi−θi)\ne′\nqi−vicos(δi−θi)/bracketrightbigg\n(38)\nwherersiis the stator resistance. Substitution of (38) into (36)\nand (37) gives\n˙xGi=fGi(xGi,θi,vi)\npGi=gpGi(xGi,θi,vi)\nqGi=gqGi(xGi,θi,vi), i∈ VG(39)B. WTG model\nThe model of a WTG includes models of the direct drive\nsynchronous generator (DDSG), controller, and converter.\n1) DDSG model: As the stator and rotor flux dynamics\nare fast in comparison with grid dynamics and the converter\ncontrols decoupled the generator from the grid, the steady- state\nelectrical equations of DDSG are assumed. The differential\nand algebraic equations for DDSG of the ithWTG are given\nas:\n˙ωmi=1\n2Hmi(τmi−τei)\npsi=vsdiisdi+vsqiisqi\nqsi=vsqiisdi−vsdiisqi(40)\nwith\nτmi=pwi(θpi)\nωmi\nτei=ψsdiisqi−ψsqiisdi\nvsdi=−rsiisdi−ωmiψsqi\nvsqi=−rsiisqi+ωmiψsdi\nψsdi=−xsdiisdi+ψpmi\nψsqi=−xsqiisqi(41)\nwhereHmiis the rotor inertia; pwi(θpi)is the mechanical\npower which is the function of pitch angle θpi;τmiandτei\nare the mechanical and electrical torques, respectively; vsdi\nandvsqi;isdiandisqi;xsdiandxsqi;ψsdiandψsqiare stator\nd-axis and q-axis voltages; currents; reactances; and fluxe s,\nrespectively; psiandqsiare produced active and reactive\npower, respectively; rsiis the stator resistance; ψpmi is the\npermanent magnet flux of rotor. Assuming the power factor\nequal to 1 (permanent magnet rotor), the reactive power outp ut\nof the DDSG equals zero, i.e., qsi= 0. Substituting (41) into\n(40) and expressing isqiwithisdibased onqsi= 0 in (40)\ngives:\n˙ωmi=fmi(ωmi,θpi,isqi)\npsi=gspi(isqi).(42)\n2) Controller: The model of the controller includes models\nof the pitch angle control unit, primary frequency control u nit,\nand the DDCU. For pitch angle control unit, its dynamic is\ndescribed by the differential equation:\n˙θpi=1\nTpi(Kpiφi(ωmi−ωmrefi)−θpi) (43)\nwhereKpi,ωmrefi , andTpiare pitch control gain, reference\nrotor speed, and pitch control time constant, respectively ;φiis\na function which allows varying the pitch angle set point onl y\nwhen the difference ωmi−ωmrefi exceeds a predefined value\n±∆ωmi. For the primary frequency control unit, its control is\ngiven as:\npfi=Kfi(ωi−ω0) (44)10\nwithωi=˙θiwhereω0is the nominal frequency and Kfiis\nthe control gain. For the DDCU, its model is already given in\nSection II-A and repeated here for completeness:\n˙x1i=−1\nTwi(Kiθi+x1i)\n˙x2i=1\nT2i/parenleftbigg\n(1−T1i\nT2i)(Kiθi+x1i)−x2i/parenrightbigg\n˙x3i=1\nT4i/parenleftbigg\n(1−T3i\nT4i)/parenleftbigg\nx2i+/parenleftbiggT1i\nT2i(Kiθi+x1i)/parenrightbigg/parenrightbigg\n−x3i/parenrightbigg\nposci=x3i+T3i\nT4i/parenleftbigg\nx2i+T1i\nT2i(Kiθi+x1i)/parenrightbigg\n.\n(45)\n3) Converter model: Converter dynamics are highly simpli-\nfied as they are fast in comparison with the electromechanica l\ntransients. Thus, the converter are modeled as an ideal curr ent\nsource where isqiandidciare state variables and are used for\nthe active power/speed control and the reactive power/volt age\ncontrol, respectively. The differential equations for the con-\nverter of the ithWTG are given as:\n˙isqi=1\nTpri(isqrefi−isqi)\n˙idci=1\nTVi((vrefi−vi)−icdi)(46)\nwith\nisqrefi=prefi(ωmi)+pfi+posci\nωmi(ψpmi−xsdiisdi)(47)\nwhereisqrefi is the reference current, prefi(ωmi)is the power-\nspeed characteristic which roughly optimizes the wind ener gy\ncapture and is calculated by based on current rotor speed ωmi.\nThe active and reactive power injected into the grid from the\nconverter are given as:\npci=vcdiicdi+vcqiicqi\nqci=vcqiicdi−vcdiicqi(48)\nwherevcdi=−visinθiandvcqi=vicosθi.\nAssuming a lossless converter, the outputs of the WTG\nbecome\npWi=pci=psi\nqWi=vi/parenleftbigg\nicdicosθi+sinθi(psi+viicdisinθi)\nvicosθi/parenrightbigg\n.(49)\nSubstituting pfiin (44) andposciin (45) into (47), combining\n(42), (43), (45), (46), and (49) gives\n˙θi=ωi\n˙xWi=fWi(xWi,ωi,θi,vi)\npWi=gpWi(xWi,ωi,θi,vi)\nqWi=gqWi(xWi,ωi,θi,vi), i∈ VW.(50)\nC. Matrices\n1) MatricesAs,Bs,Cs, andDs:Refer to (51) on next\npage for the detailed definition, where the notation K∧\n∨(J∧\n∨)\nexpresses Jacobian matrix of the ∨in the subscript with respect\nto the∧in the superscript. It should be noted that all the\nJacobian matrices K∧\n∨are block diagonal matrices.2) MatricesK1,K2,K3, andJpf:Refer to (52) on next\npage for the detailed definition. All the Jacobian matrices J∧\n∨\nin (52) form the power flow Jacobian matrix Jpf∈R2N×2N\nwhere\nJpf=/bracketleftBigg\nJhp\nθJhp\nv\nJhq\nθJhq\nv/bracketrightBigg\n. (53)\n3) Elementary column operator matrix T:\n\nI8NG0 0 0 0 0 0 0 0 0 0 0 0\n0 0 0 0 0 0 INW0 0 0 0 0 0\n0 0I7NW0 0 0 0 0 0 0 0 0 0\n0 0 0 0 0 0 0 INL0 0 0 0 0\n0 0 0 0 I3NL0 0 0 0 0 0 0 0\n0 0 0 0 0 ING0 0 0 0 0 0 0\n0INW0 0 0 0 0 0 0 0 0 0 0\n0 0 0 INL0 0 0 0 0 0 0 0 0\n0 0 0 0 0 0 0 0 INT0 0 0 0\n0 0 0 0 0 0 0 0 0 ING0 0 0\n0 0 0 0 0 0 0 0 0 0 INW0 0\n0 0 0 0 0 0 0 0 0 0 0 INL0\n0 0 0 0 0 0 0 0 0 0 0 0 INT\n\n(54)11\n/bracketleftbiggAsBs\nCsDs/bracketrightbigg\n=\nKfGxG0 0 0 0 KfG\nθG0 0 0 KfGvG0 0 0\n0 0 0 0 0 0 INW0 0 0 0 0 0\n0KfW\nθWKfWxW0 0 0 KfWωW0 0 0 KfWvW0 0\n0 0 0 0 0 0 0 INL0 0 0 0 0\n0 0 0 KfL\nθLKfLxL0 0 0 0 0 0 0 0\nKhpGxGJhpG\nθW0JhpG\nθL0JhpG\nθG0 0JhpG\nθTJhpGvGJhpGvWJhpGvLJhpGvT\n0JhpW\nθWKhpWxWJhpW\nθL0JhpW\nθGKhpWωW0JhpW\nθTJhpWvGJhpWvWJhpWvLJhpWvT\n0JhpL\nθW0JhpL\nθLKhpLxLJhpL\nθG0KhpLωLJhpL\nθTJhpLvGJhpLvWJhpLvLJhpLvT\nKhqGxGJhqG\nθW0JhqG\nθL0JhqG\nθG0 0JhqG\nθTJhqGvGJhqGvWJhqGvLJhqGvT\n0JhqW\nθWKhqWxWJhqW\nθL0JhqW\nθGKhqWωW0JhqW\nθTJhqWvGJhqWvWJhqWvLJhqWvT\n0JhqW\nθW0JhqL\nθL0JhqL\nθG0KhqLωLJhqL\nθTJhqLvGJhqLvWJhqLvLJhqLvT\n(51)\n/bracketleftbiggK1K2\nK3Jpf/bracketrightbigg\n=\nKfGxG0 0 0 0 KfG\nθG0 0 0 KfGvG0 0 0\n0INW0 0 0 0 0 0 0 0 0 0 0\n0KfWωWKfWxW0 0 0 KfW\nθW0 0 0 KfWvW0 0\n0 0 0 INL0 0 0 0 0 0 0 0 0\n0 0 0 0 KfLxL0 0KfL\nθL0 0 0 0 0\nKhpGxG0 0 0 0 JhpG\nθGJhpG\nθWJhpG\nθLJhpG\nθTJhpGvGJhpGvWJhpGvLJhpGvT\n0KhpWωWKhpWxW0 0JhpW\nθGJhpW\nθWJhpW\nθLJhpW\nθTJhpWvGJhpWvWJhpWvLJhpWvT\n0 0 0 KhpLωLKhpLxLJhpL\nθGJhpL\nθWJhpL\nθLJhpL\nθTJhpLvGJhpLvWJhpLvLJhpLvL\nKhqGxG0 0 0 0 JhqG\nθGJhqG\nθWJhqG\nθLJhqG\nθTJhqGvGJhqGvWJhqGvLJhqGvT\n0KhqWωWKhqWxW0 0JhqW\nθGJhqW\nθWJhqW\nθLJhqW\nθTJhqWvGJhqWvWJhqWvLJhqWvT\n0 0 0 KhqLωL0JhqL\nθGJhqL\nθWJhqL\nθLJhqL\nθTJhqLvGJhqLvWJhqLvLJhqLvT\n(52)" }, { "title": "2108.07542v3.Application_of_Herglotz_s_Variational_Principle_to_Electromagnetic_Systems_with_Dissipation.pdf", "content": "Application of Herglotz’s Variational Principle to\nElectromagnetic Systems with Dissipation\naJordi Gaset∗,bAdrià Marín-Salvador†\naEscuela Superior de Ingeniería y Tecnología, Universidad Internacional de La Rioja, Spain.\nbMathematical Institute, University of Oxford, United Kingdom.\nDecember 15, 2021\nAbstract\nThis work applies the contact formalism of classical mechanics and classical field theory, introduced\nby Herglotz and later developed in the context of contact geometry, to describe electromagnetic systems\nwith dissipation. In particular, we study an electron in a non-perfect conductor and a variation of the\ncyclotron radiation. In order to apply the contact formalism to a system governed by the Lorentz force,\nit is necessary to generalize the classical electromagnetic gauge and add a new term in the Lagrangian.\nWe also apply the k-contact theory for classical fields to model the behaviour of electromagnetic fields\nthemselves under external damping. In particular, we show how the theory describes the evolution\nof electromagnetic fields in media under some circumstances. The corresponding Poynting theorem is\nderived. We discuss its applicability to the Lorentz dipole model and to a highly resistive dielectric.\nKeywords: Contact geometry, Lagrangian systems, dissipation, field theories, Maxwell equations, electro-\nmagnetic gauge.\nMSC 2020 codes: 37K58, 37L05, 53D10, 35Q53\nContents\n1 Introduction 2\n2 Contact and k-contact Lagrangian systems 4\n2.1 k-contact Lagrangian systems . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6\n∗jordi.gaset@unir.net( ORCID:0000-0001-8796-3149).\n†adria.marin@st-hughs.ox.ac.uk (ORCID: 0000-0001-8054-1576).\n1arXiv:2108.07542v3 [physics.class-ph] 14 Dec 20213 Study of Particles under a Lorentz Force 8\n3.1 Symplectic Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8\n3.1.1 The Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8\n3.1.2 Classical Gauge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9\n3.2 Contact Formulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10\n3.2.1 A First Attempt at the Equations of Motion . . . . . . . . . . . . . . . . . . . . . . 10\n3.2.2 Generalized Gauge . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11\n3.2.3 The Equations of Motion Revisited . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12\n3.2.4 Equivalent Lagrangians . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13\n3.2.5 Example. Electron in a non-Perfect Conductor . . . . . . . . . . . . . . . . . . . . . 14\n3.2.6 Example. Particle in a Magnetic Field . . . . . . . . . . . . . . . . . . . . . . . . . . 16\n4 Electromagnetic Fields 18\n4.1 Covariant Formulation of Classical Electromagnetism . . . . . . . . . . . . . . . . . . . . . . 18\n4.2k-contact Formulation for Electromagnetic Fields . . . . . . . . . . . . . . . . . . . . . . . . 20\n4.3 Generalized Poynting’s Theorem . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21\n4.4k-contact Maxwell’s equations as Maxwell’s equations in matter . . . . . . . . . . . . . . . . 22\n4.5 Example. The Lorentz Dipole Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22\n4.6 Example. Highly Resistive Dielectric . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23\n4.7 Gauge Invariance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24\n5 Conclusions 24\n1 Introduction\nAround 1930, G. Herglotz introduced a variational method to model mechanical systems with dissipation\n[24,25]. Herglotz allowed the Lagrangian of the system to depend on the action itself, and obtained a\nset of generalized Euler-Lagrange equations that happened to be useful when considering some dissipative\nphenomena. The framework of such description was later discovered to be contact geometry [13,19].\nRecently, there has been a renewed interest in contact geometry due to its success in modelling several\nsystems. Some of these applications include thermodynamics [2,21,31,33,34], statistical mechanics [3],\ngeometric optics [1], hydrodynamics [14], circuit theory [20] and control theory [8,35].\nIn parallel, contact geometry has been expanded with new structures and frameworks. The Lagrangian\nformulation and its symmetries have been studied [9–11,15,17], and several generalizations have been\ndeveloped: the higher-order case [5], the Hamilton-Jacobi theory [7] and the unified formalism [6].\nRecently, a higher-dimensional analogue of contact geometry has been proposed, called k-contact ge-\nometry, which can be applied to model field theories with dissipation [16,18]. This description generalizes\nprevious attempts to define an action principle for action-dependant Lagrangian field densities [27].\nThe present work applies the contact formalism of classical mechanics and classical field theory to\nelectromagnetism. This has, to the best of our knowledge, not been done explicitly in literature, apart from\n2a small example in [27]. Electromagnetism provides an interesting context in which to apply the theory. On\nthe one hand, the study of a particle under a Lorentz force allows for the use of the contact formalism of\nclassical mechanics. On the other, one can focus on the study of the evolution of the electromagnetic vector\nfields themselves applying the k-contact geometry theory. In addition, one can think of multiple examples\nin which a particle moves under the influence of a Lorentz force in the presence of external dissipation\n[32, Ch. 6]. It is also known how electromagnetic fields can be damped when in a medium [22, Sec. 11.5].\nIt should be clear that contact geometry does not provide a description of all dissipative phenomena, but\nrather simply produces a set of equations of motion from a Lagrangian function that is action-dependant.\nThus, not all dissipative systems can be explained via this theory and its use is not constrained to such\nmodels. This work explores to what extent dissipation in electromagnetic systems can be modelled using\ncontact formalism.\nDuring this work, some objects will be referred to as dissipative although they actually represent a break\nin energy conservation, which might be an increase or a decrease of the energy. It will we specified when\nthis is the case and when we actually refer to energy loss.\nDriven by previous successful applications of contact geometry in dissipation [10,16–18,27], we have\ndecided to produce the equations of motion from a Lagrangian function which is the classical symplectic\nLagrangian plus a linear term in the action. We allow the linear term to be tuned a posteriori in order to\nfit the experimental results in every particular example.\nThe currently developed contact techniques for classical mechanics and classical field theory only apply\nto autonomous Lagrangians. That is, in the classical mechanical case, the Lagrangians cannot depend\nexplicitly on time, and in the k-contact framework, the Lagrangian densities cannot depend explicitly on the\ncomponents of the spacetime. This is the reason for which during this work it will be asked that the objects\nappearing in the Lagrangian do not have explicit terms in such components. However, using variational\nprinciples, one can show that the produced equations of motion are still valid in the non-autonomous case.\nFuture research is needed in order to develop the mathematical tools to describe such Lagrangians using\nthe contact formalism.\nThisworkisstructuredasfollows. Section2presentsamathematicalintroductiontoLagrangiancontact\nsystems and Lagrangian k-contact systems. In Section 3, we apply the contact formalism of classical\nmechanics to study particles under a Lorentz force. When producing the equations of motion from the\nclassical Lagrangian of a Lorentz force, one finds that they are not gauge invariant. This can be solved\nby slightly changing the Lagrangian and introducing a new characterization of the electromagnetic gauge,\nwhich reduces to the classical one for non-dissipative systems. The theory is applied to two particular\nexamples, an electron in a non-perfect conductor and a particle in a magnetic field with damping. Finally,\nin section 4 we apply the k-contact theory to electromagnetic fields. Apart of the dissipative term, we\nconsider a particular metric to model lineal materials. In some cases, the classical Lagrangian density in\nvacuum can model electromagnetic fields in media when adding a linear term in the action. We derive\nsufficient and necessary conditions that a system must satisfy so that the theory applies. Finally, the\napplicability of the produced equations of motion is discussed in two real-life examples.\n32 Contact and k-contact Lagrangian systems\nLagrangian contact systems are an special case of contact systems, which we will introduce briefly. The\nreader can find a more detailed exposition in [10,11,17].\nConsider the manifold TQ\u0002R, whereQis anndimensional manifold which represents the configuration\nspace of the system, with local coordinates (qi;vi;s). The canonical endomorphism Sand the Liouville\nvector field \u0001ofTQextend toTQ\u0002Rin the usual way. Their local expressions are\nS=@\n@vi\ndqi;\n\u0001 =vi@\n@vi:\nA Lagrangian is a function L:TQ\u0002R!R. The associated contact form is \u0011L=ds\u0000S\u0003(dL), with local\nexpression\n\u0011L=ds\u0000@L\n@vidqi:\nThe Reeb vector field of \u0011Lis\nRL=@\n@s\u0000Wij@2L\n@vj@s@\n@vi\nwhere (Wij)is the inverse of the Hessian of L. The associated Lagrangian energy is EL= \u0001(L)\u0000L. Its\nlocal expression is\nEL=vi@L\n@vi\u0000L:\nThen, wehavethecontactsystem (\u0011L;EL). Itssolutionsareintegralcurvesofavectorfield X2X(TQ\u0002R),\nwhich is a SODE, and satisfies the generalized Euler-Lagrange equations\n\u0013Xd\u0011L=dEL\u0000\u0000\nLRLEL\u0001\n\u0011L;\n\u0013X\u0011L=\u0000EL:\nFor a holonomic curve \u001b(t) = (qi(t);vi(t);s(t)), these equations take the local expression\n_vj@2L\n@vi@vj+vj@2L\n@vi@qj+ _s@2L\n@vi@s\u0000@L\n@qi=@L\n@vi@L\n@s;\n_s=L: (1)\nHence,scan be interpreted as the action of the system. We are therefore modelling systems in which\nthe Lagrangian depends on the action itself.\nThe generalized Euler-Lagrange equations can also be obtained via a variational method, as seen in\n[10, Section 5], which was introduced by G. Herglotz in [24]. Let L2C1(TQ\u0002R)be a Lagrangian\nfunction and consider two points x;y2Q. Consider [0;1]\u0012Rand let us denote by Sthe space of all\nsmooth curves \u000b:[0;1]!Qsuch that\u000b(0) =xand\u000b(1) =y.\nLet us denote by C1([0;1]!X)the set of all smooth mappings from [0;1]to a manifold X. One can\ndefine the functional\nZ:C1([0;1]!Q)!C1([0;1]!R)\n\u00187! Z (\u0018);\n4whereZ(\u0018)is the unique solution to\n8\n><\n>:dZ(\u0018)(t)\ndt=L(\u0018(t);_\u0018(t);Z(\u0018)(t))\nZ(\u0018)(0) = 0:(2)\nAfter the previous discussion on the problem and particularly Equation (1), it is clear why one would\ndefine such an operator Z: given a curve \u0018, its imageZ(\u0018)is the action of the system associated to the\npath\u0018. Hence, the physically realisable path between xandyis the curve \u00182Sthat minimizes\nZ(\u0018)(1);\nthe action at the endpoint, see [10, Thm. 2]. Note that, from Equation (2), we find\nZ(\u0018)(1) =Z1\n0L(\u0018(t);_\u0018(t);Z(\u0018)(t))dt: (3)\nLet\u0011(t) =\u0000\n\u00111(t);\u00112(t);\u00113(t)\u0001\n2C1([0;1]!Q)such that\u0011(0) =\u0011(1) = 0and considerZ(\u0018+\"\u0011)(1).\nThe variational problem (3) implies that\n0 =dZ(\u0018+\"\u0011)(1)\nd\"\f\f\f\n\"=0=Z1\n0\u0010@L\n@qi\u0011i+@L\n@vi_\u0011i\u0011\ndt+Z1\n0@L\n@sdZ(\u0018+\"\u0011)(t)\nd\"\f\f\f\n\"=0dt: (4)\nLet us now define \u0010(r) =dZ(\u0018+\"\u0011)(r)\nd\"j\"=0andA(t) =@L\n@qi\u0011i+@L\n@vi_\u0011i. Note that, for r2(0;1],\n\u0010(r) =Zr\n0A(t)dt+Zr\n0@L\n@s\u0010(t)dt; (5)\nwhich implies\nd\u0010\ndr(r) =A(r) +@L\n@s\u0010(r):\nHence, solving the ODE with the initial condition \u0010(0) = 0,\n\u0010(r) = exp\u0010Zr\n0@L\n@s(\u0012)d\u0012\u0011\n\u0001Zr\n0exp\u0010\n\u0000Zt\n0@L\n@s(\u0012)d\u0012\u0011\nA(t)dt:\nLetusimposethat \u0010(1) = 0, whichisequivalenttoimposingthattheactionreachesarelativeextremum.\nSince the first factor cannot vanish, this reads\n0 =Z1\n0exp\u0010\n\u0000Zt\n0@L\n@s(\u0012)d\u0012\u0011\nA(t)dt=Z1\n0exp\u0010\n\u0000Zt\n0@L\n@s(\u0012)d\u0012\u0011\u0010@L\n@qi\u0011i+@L\n@vi_\u0011i\u0011\ndt;\nand integrating by parts the second term\n0 =Z1\n0exp\u0010\n\u0000Zt\n0@L\n@s(\u0012)d\u0012\u0011@L\n@qi\u0011idt+Z1\n0exp\u0010\n\u0000Zt\n0@L\n@s(\u0012)d\u0012\u0011@L\n@vi_\u0011idt=\n=Z1\n0exp\u0010\n\u0000Zt\n0@L\n@s(\u0012)d\u0012\u0011@L\n@qi\u0011idt\u0000Z1\n0exp\u0010\n\u0000Zt\n0@L\n@s(\u0012)d\u0012\u0011hd\ndt\u0010@L\n@vi\u0011\n\u0000@L\n@s@L\n@vii\n\u0011idt=\n=Z1\n0exp\u0010\n\u0000Zt\n0@L\n@s(\u0012)d\u0012\u0011h@L\n@qi\u0000d\ndt\u0010@L\n@vi\u0011\n+@L\n@s@L\n@vii\n\u0011idt;\nwhich, by the fundamental lemma of calculus of variations [26, Lemm. 1.1.1], implies the generalized\nEuler-Lagrange equations\n@L\n@qi\u0000d\ndt\u0010@L\n@vi\u0011\n+@L\n@s@L\n@vi= 0:\n52.1 k-contact Lagrangian systems\nThe k-contact structure was introduced in [16] in order to generalize contact mechanics to field theories.\nThe Lagrangian formalism, which we use in Section 4, was developed in [18]. In this section the principle\nelements of this formalism are stated. Moreover, we will give a variational formulation of the k-contact\nEuler-Lagrange equations.\nThe Lagrangian k-contact formalism of a system with an k-dimensional configuration space Qover ank-\ndimensional space-time takes place in \bkTQ\u0002Rk. The bundle\bkTQoverQis the Whitney sum of kcopies\nof the tangent bundle, each one representing the derivative of the coordinate over the different directions\nof space-time. Moreover, kdissipative variables are considered. Natural coordinates on \bkTQ\u0002Rkare\n(qi;qi\n\u0016;s\u0016), where 1\u0014i\u0014nand1\u0014\u0016\u0014k.\nA Lagrangian is a function L:\bkTQ\u0002Rk!R. On this work we only consider a particular class of\nLagrangians, those with linear dependence on s\u0016with constant coefficients.\nThe k-contact structure is formed by the k1-forms\n\u0012\u0016\nL=ds\u0016\u0000@L\n@qi\u0016dqi:\nThe Reeb vector fields are a set of kvector fields such that\n\u0013R\u000b\u0012\f=\u000e\f\n\u000b;\u0013R\u000bd\u0012\f= 0: (6)\nFor the particular class of Lagrangians we will consider in this work, they can be chosen to be R\u0016=@\n@s\u0016.\nFinally, the Lagrangian energy is given by\nEL=@L\n@qi\u0016qi\n\u0016\u0000L:\nThe solutions are holonomic functions \u001b:Rk!\bkTQ\u0002Rk, which are integrable sections of a k-\ndimensional distribution. This distribution can be described by kvector fields (X\u0016)and a section \u001bis\nintegral of (X\u0016)if\nT\u001b\u000e@\n@x\u0016=X\u0016\u000e\u001b:\nThe k-contact Euler-Lagrange equations are\n\u0013X\u0016d\u0012\u0016\nL=dEL\u0000\u0000\nLR\u0016EL\u0001\n\u0012\u0016\nL;\n\u0013X\u0016\u0012\u0016\nL=\u0000EL:(7)\nFor a holonomic function \u001b= (qi(x\u0016);qi\n\u0016(x\u0016);s\u0016(x\u0016), these equations take the local expression\n@qj\n\u0017\n@x\u0016@2L\n@qi\u0016@qj\n\u0017+qj\n\u0016@2L\n@qi\u0016@qj+@s\u0017\n@x\u0016@2L\n@qi\u0016@s\u0017\u0000@L\n@qi=@L\n@qi\u0016@L\n@s\u0016;\n@s\u0016\n@x\u0016=L:\nThe Lagrangian k-contact formalism presented in [18] is developed for regular Lagrangians. Unfortunately,\nthe Lagrangian used in section 4 is singular. Nevertheless, one can try to follow all the steps described\nabove, but there is one problem: Equations (6) do not have a unique solution for singular Lagrangians. To\n6circumvent this issue, in section 4 we will use an idea introduced in [10] for the mechanical case: to proof\nthat Equations (7) are independent of the solution of (6) chosen.\nJust like in the classical contact formulation, it is possible to derive the k-contact Euler-Lagrange\nequations for fields as a result of a variational principle. Let L(qi;qi\n\u0016;s\u0016)be a Lagrangian function on\n\bkTQ\u0002Rk. A field on Qis a smooth map\n\t:\n\u001aRk! Q\u0002Rk\nx\u0017!(\t\u0016(x\u0017);\ts\u0016(x\u0017)\u0011s\u0016(x\u0017))\nfrom an open subset \n, and the action related to such field is defined as\nS(\t) =Z\n\nL(\t\u0016;@\u0017\t\u0016;s\u0016)d4x:\nThen, the equations of motion can be obtained by minimizing the action S(\t)with respect to \tunder the\nconstraint\n@\u0016s\u0016=L(\t\u0016;@\u0017\t\u0016;s\u0016):\nBy the Lagrange multiplier Theorem for Banach spaces [29, Thm. 9.3.1], extremizing Swith the above\nconstraint is equivalent to extremizing the following function with respect to \t,\nf(\t\u0016;@\u0017\t\u0016;s\u0016;@\u0017s\u0016;\u0015) =Z\n\nh\nL(\t\u0016;@\u0017\t\u0016;s\u0016)\u0000\u0015(x\u000b)\u0000\n@\u0016s\u0016\u0000L(\t\u0016;@\u0017\t\u0016;s\u0016)\u0001i\nd4x\n=Z\n\nh\n@\u0016s\u0016\u0000\u0015(x\u000b)\u0000\n@\u0016s\u0016\u0000L(\t\u0016;@\u0017\t\u0016;s\u0016)\u0001i\nd4x;\nwhere\u0015:\n!Ris a smooth function which we call Lagrange multiplier function. The Euler-Lagrange\nequations for fread8\n>>>>>><\n>>>>>>:@\u0017\u0010\n\u0015(x\u000b)@L\n@(@\u0017\t\u0016)\u0011\n=\u0015(x\u000b)@L\n@\t\u0016\n\u0000@\u0016\u0015(x\u000b) =\u0015(x\u000b)@L\n@s\u0016\n0 =@\u0016s\u0016\u0000L(\t\u0016;@\u0017\t\u0016;s\u0016);\nwhere the last equation comes from imposing that the Euler-Lagrange equations are also satisfied for \u0015.\nExpanding the first equation,\n\u0015(x\u000b)@L\n@\t\u0016=@L\n@(@\u0017\t\u0016)@\u0017\u0015(x\u000b) +\u0015(x\u000b)@\u0017\u0010@L\n@(@\u0017\t\u0016)\u0011\n=\u0000\u0015(x\u000b)@L\n@(@\u0017\t\u0016)@L\n@s\u0017+\u0015(x\u000b)@\u0017\u0010@L\n@(@\u0017\t\u0016)\u0011\n;\nwhich, dividing by \u0015(x\u000b), implies\n@\u0017\u0010@L\n@(@\u0017\t\u0016)\u0011\n\u0000@L\n@\t\u0016=@L\n@(@\u0017\t\u0016)@L\n@s\u0017;\nthek-contact Euler-Lagrange equations for fields.\n73 Study of Particles under a Lorentz Force\n3.1 Symplectic Formulation\n3.1.1 The Equations of Motion\nLetQ\u0012R3be an open subset of R3on which a particle of mass mand charge kmoves under the influence\nof the electromagnetic force\nF=k(E+v\u0002B)\ninduced by an electric field Eand a magnetic field Bwhich are time independent. Let q= (q1;q2;q3)be\ncoordinates defined on Q. Assume, in addition, that the electromagnetic potentials\n\u001e:Q!R\nq7!\u001e(q)A:Q!R3\nq7!A(q)\nthat define EandBare also time independent. Recall that Aand\u001eare such that\nE=\u0000r\u001eand B=r\u0002A:\nThe Lagrangian associated with the Lorentz force is\nL=m\n2v\u0001v+kA\u0001v\u0000k\u001e:\nThe Lagrangian energy function and the Lagrangian symplectic form read\nEL=@L\n@v\u0001v\u0000L=\u0000\nmv+kA\u0001\n\u0001v\u0000L=m\n2v\u0001v+k\u001e\nand\n!L=\u00003X\ni=1d\u0010@L\n@vi\u0011\n^dqi=\u00003X\ni=1d(mvi+kAi)^dqi=m3X\ni=1dqi^dvi+k3X\ni=1\u0010X\nj6=i@Ai\n@qjdqi^dqj\u0011\n:\nThe dynamics of the Lagrangian dynamical system (TQ;!L;EL)are encoded in the vector field XL2\nX(TQ)solution to\n\u0013XL!L=dEL: (8)\nOne can compute\ndEL=k@\u001e\n@q\u0001dq+mv\u0001dv:\nIf we write\nXL=3X\ni=1Qi@\n@qi+3X\ni=1Vi@\n@vi;\nthen Equation (8) reads\nm3X\ni=1Qidvi+3X\ni=1 \n\u0000mVi+kX\nj6=i\u0010@Aj\n@qi\u0000@Ai\n@qj\u0011\nQj!\ndqi=k3X\ni=1@\u001e\n@qidqi+m3X\ni=1vidvi:(9)\n8Since all the differentials are linearly independent, the previous equation implies\n8\n>><\n>>:Qi=vi\nmVi=kX\nj6=i\u0010@Aj\n@qi\u0000@Ai\n@qj\u0011\nQj\u0000k@\u001e\n@qi:(10)\n(11)\nHence, the time evolution of the particle in configuration space is given by (q(t);v(t))with _q(t) =v(t),\nand\nm_v(t) =kv(t)\u0002(r\u0002A)\u0000kr\u001e; (12)\nwhich are, of course, the equations of motion obtained by using the Euler-Lagrange equations. Note that\nwe findm_v=kE+kv\u0002B, which are the equations of motion obtained by applying Newton’s second law\nto the Lorentz force.\n3.1.2 Classical Gauge\nThe choice of vector and scalar potentials at the beginning of Section 3.1.1 is not unique. Indeed, let\nf2C1(Q\u0002R)be a smooth function on Q\u0002R, where the Rcoordinate depicts time, and let\n\u001e0=\u001e\u0000@f\n@tand A0=A+rf:\nAssume that the primed potentials are also time independent, which is equivalent to imposing\n@2f\n@t2=@2f\n@t@qi= 0:\nThen,\nr\u001e0=r\u001e\u0000r@f\n@t=r\u001e\u00003X\ni=1@2f\n@qi@t=r\u001e\nand\nr\u0002A0=r\u0002A+r\u0002(rf) =r\u0002A;\nand hence the induced electric and magnetic fields remain unchanged. This freedom in choosing the vector\nand scalar potentials is known as gauge freedom, and the choice of a particular pair (\u001e;A)is called a gauge\nfixing or a choice of gauge. It is believed that the gauge is not measurable since, as deduced in Section\n3.1.1, the equations of motion only depend on the observable fields EandB.\nAlthough the equations of motion remain unchanged when the gauge changes, the Lagrangian does not.\nIndeed, let us compute\nL0=m\n2v\u0001v+kA0\u0001v\u0000k\u001e0=L+k(rf)\u0001v+k@f\n@t=L+d(kf)\ndt:\nIt is known that the addition of a full time derivative to the Lagrangian does not change the extrema\nof the action and hence it does not change the equations of motion. However, it is important to note that\nthe Lagrangian and the action themselves do depend on the choice of gauge.\n93.2 Contact Formulation\n3.2.1 A First Attempt at the Equations of Motion\nConsider again a particle of mass mand charge kmoving in an open subset Q\u0012R3under the influence\nof a Lorentz force F=k(E+v\u0002B). Assume EandBare time independent and let (\u001e;A)be a choice\nof gauge for the electric and magnetic fields such that the potentials are also time independent. Driven\nby previous successful applications of contact formalism to mechanical systems with dissipation [10,17], we\npropose the following contact Lagrangian on TQ\u0002R:\nL=m\n2v\u0001v+kA\u0001v\u0000k\u001e\u0000\rs\nfor some\r2R.\nThe Lagrangian energy associated to this system is\nEL=@L\n@v\u0001v\u0000L= (mv+kA)\u0001v\u0000L=m\n2v\u0001v+k\u001e+\rs;\nand the canonical contact form equals\n\u0011L=ds\u0000@L\n@v\u0001dq=ds\u0000mv\u0001dq\u0000kA\u0001dq:\nSince all the second derivatives@2L\n@s@vivanish, the Reeb vector field is simply R=@\n@s. It is clear that the\ndirectional derivative of ELin the direction of@\n@sis the constant \r. Thus, the generalized Euler-Lagrange\nequations simplify to\n\u0013XLd\u0011L=dEL\u0000\r\u0011L\n\u0013XL\u0011L=\u0000EL(13)\n(14)\nfor an unknown vector field XL2X(TQ\u0002R). Let us compute the differentials\ndEL=kr\u001e\u0001dq+mv\u0001dv+\rds\nand\nd\u0011L=m3X\ni=1dqi^dvi+k3X\ni;j=1@Aj\n@qidqj^dqi:\nIf we write the unknown vector field XLin coordinates as\nXL=3X\ni=1Qi@\n@qi+3X\ni=1Vi@\n@vi+S@\n@s;\nthen Equation (14) reads\nS\u0000@L\n@v\u0001Q=\u0000@L\n@v\u0001v+L\nand if we impose that the system is holonomic, that is Qi= _qi=vi, then the solution curve (q(t);v(t);s(t))\nwill satisfy\n_s=L;\nand the Lagrangian is interpreted to depend on the actionR\nLdtitself.\n10If we now let _vi=Vi, Equation (13) implies\n\u0000m_v\u0001dq+k3X\ni=1X\nj6=i\u0010@Aj\n@qi\u0000@Ai\n@qj\u0011\nQjdqi=kr\u001e\u0001dq+\r(mv+kA)\u0001dq;\nand given the linear independence of the 1-forms dqi,\nm_v(t) =\u0000k(r\u0002A)\u0002v(t)\u0000kr\u001e\u0000\r(mv(t) +kA): (15)\nNote that we find the same two terms as in Equation (12), together with a dissipative term in velocities\n\u0000\rmv(t)and an interaction term between the dissipation and the vector potential, \u0000\rkA.\nThe obtained equations of motion are not invariant under the classical gauge defined in Section 3.1.2.\nIndeed, if we make a change of gauge\n\u001e0=\u001e\u0000@f\n@tand A0=A+rf\nbetween time independent potentials, the first two terms remain invariant, but the interaction term does\nnot, and hence then the equations of motion become\nm_v(t) =\u0000k(r\u0002A)\u0002v(t)\u0000kr\u001e\u0000\r(mv(t) +kA)\u0000\rkrf:\nNotethatachangeof gauge withasmoothfunction fintroducesafulltimederivativeintheLagrangian,\nas discussed in Section 3.1.2. Hence, we observe that, in the contact framework, adding a total time\nderivative does not, in general, preserve the equations of motion. The Lorentz force provides an example\nof how producing equivalent Lagrangians in the contact framework differs from the symplectic case.\nSince we want the equations of motion to be gauge invariant, and the gauge to be non-observable, we will\nintroduce a new characterization of the gauge and propose a new contact Lagrangian in the next section.\n3.2.2 Generalized Gauge\nIn this section we introduce a new characterization of the classical gauge that generalizes the one defined\nin Section 3.1.2. Instead of a pair (\u001e;A), a choice of gauge will now be a triplet\n(\u001e;A;f);\nwhere (\u001e;A)is a classical gauge and f2C1(Q)is a smooth function on Q. As discussed in previous\nsections, only time independent scalar and vector potentials are considered. Note that, since fis a function\nonQ, it is time independent.\nWe will define two choices of gauge (\u001e;A;f);(\u001e0;A0;f0)to be related by the gauge if there exists a\nsmooth function g2C1(Q\u0002R)such that\n8\n>>>><\n>>>>:\u001e0=\u001e\u0000@g\n@t\nA0=A+rg\nf0=f\u0000g:\nFor both gauges to be time independent, it will be necessary and sufficient that@g\n@t= 0. Note that this\nis equivalent to g2C1(Q)and that it implies that the scalar potential remains unchanged. Note also that\n11the relation defined on the choices of gauge is an equivalence relation, and hence it produces equivalence\nclasses of choices of gauge. Let [(\u001e;A;f)]denote the equivalence class of the choice of gauge (\u001e;A;f).\nWe claim that, in each class [(\u001e;A;f)], there exists a unique choice of gauge of the type (\u001e0;A0;0).\nIndeed,\n[(\u001e;A;f)] = [(\u001e\u0000@f\n@t;A+rf;f\u0000f)] = [(\u001e\u0000@f\n@t;A+rf;0)];\nand if [(\u001e;A;0)] = [(\u001e0;A0;0)]the function g2C1(Q)that relates them must satisfy 0 = 0\u0000g, and hence\nmust beg= 0. Then,\u001e=\u001e0andA=A0,\nWe propose, for a choice of gauge (\u001e;A;f), the contact Lagrangian\nL=m\n2v\u0001v+kA\u0001v+k@f\n@q\u0001v\u0000k\u001e\u0000\rs; (16)\nwhere we explicitly make use of the function f2C1(Q)of the gauge.\n3.2.3 The Equations of Motion Revisited\nAssume a particle of mass mand charge kis moving in an open subset Q\u0012R3under the influence of an\nelectromagnetic force. Let (\u001e;A;f)be a choice of gauge as defined in Section 3.2.2, and let the contact\nLagrangian of the system be the one defined in Equation (16),\nL=m\n2v\u0001v+kA\u0001v+k@f\n@q\u0001v\u0000k\u001e\u0000\rs: (17)\nThe Lagrangian energy density associated with this Lagrangian is\nEL=@L\n@v\u0001v\u0000L=m\n2v\u0001v+k\u001e+\rs;\nand the contact 1-form is\n\u0011L=ds\u0000@L\n@v\u0001dq=ds\u0000mv\u0001dq\u0000kA\u0001dq\u0000k@f\n@q\u0001dq:\nJustlikeinSection3.2, theReebvectorfieldis RL=@\n@s, andthedirectionalderivativeoftheenergywith\nrespect to the Reeb vector field is LRLEL=@EL\n@s=\r. Hence, the generalized Euler-Lagrange equations\nread\n\u0013XLd\u0011L=dEL\u0000\r\u0011L\n\u0013XL\u0011L=\u0000EL:(18)\n(19)\nA straightforward computation gives\ndEL=mv\u0001dv+kr\u001e\u0001dq+\rds\nand\nd\u0011L=m3X\ni=1dqi^dvi+k3X\ni;j=1\u0010@Aj\n@qi\u0011\ndqj^dqi:\nSince the only difference with respect to Equations (13) and (14) is the extra term \u0000krf\u0001dqin the\ncontact form, the equations of motion are\n_s=L\n12and\nm_v(t) =\u0000k(r\u0002A)\u0002v(t)\u0000kr\u001e\u0000\r\u0000\nmv(t) +kA\u0001\n\u0000\rkrf; (20)\nonce we have imposed that the system is holonomic, i.e. _q=v.\nWe claim that Equation (20) is now independent of the choice of gauge. Indeed, if we take g2C1(Q)\nand let 8\n>>><\n>>>:\u001e0=\u001e\nA0=A+rg\nf0=f\u0000g;\nEquation (20) becomes\nm_v(t) =\u0000k(r\u0002A)\u0002v(t)\u0000kr\u001e\u0000\r\u0000\nmv(t) +kA\u0001\n\u0000\rkrg\u0000\rkrf+\rkrg=\n=\u0000k(r\u0002A)\u0002v(t)\u0000kr\u001e\u0000\r\u0000\nmv(t) +kA\u0001\n\u0000\rkrf;\nand thus, it remains unchanged. Actually, note that the Lagrangian (16) is itself invariant under a change\nof gauge, unlike in the symplectic case.\nIn this new proposed framework, the observable fields are\n8\n>>><\n>>>:E=\u0000r\u001e\nR=A+rf\nB=r\u0002A=r\u0002R(21)\nfor a choice of gauge (\u001e;A;f), and the equations of motion are\nm_v=kv(t)\u0002B+kE\u0000\rmv(t)\u0000\rkR;\nwhere all three fields are invariant under the gauge. Note that knowing Rallows us to know Aup to a\ngradient, which is exactly the same freedom for Awhen knowing B. In addition, the generalized moment\nof a particle under the proposed Lagrangian is\n@L\n@v=mv+kA+krf=mv+kR;\nwhich is also an observable, whilst in the symplectic case the generalized moment is not an observable.\n3.2.4 Equivalent Lagrangians\nIn order to have a more complete understanding of the gauge freedom of the Lagrangian, let us analyze its\nequivalent Lagrangians.\nTwo Lagrangians are equivalent if they lead to the same solutions. In symplectic mechanics, two La-\ngrangians that differ by a total derivative are equivalent. In contact mechanics, one can construct equivalent\nLagrangians by considering transformations for the svariable, which can be thought of as constructing new\nactions with the same critical points. One can generalize the symplectic result to the contact setting by\nconsidering transformations of the form s!\u0010(qi;s) =s+h(qi).\nFor a Lagrangian Land a transformation \u0010, the corresponding equivalent Lagrangian is given by:\n\u0016L(x;q;v;\u0010 ) =L(x;q;v;s ) +rh\u0001v:\n13Notice that the action variable for the new Lagrangian is \u0010. Considering the Lagrangian (17) we are\ninterested in, for any function h(qi)we have an equivalent Lagrangian\n\u0016L(x;q;v;\u0010 ) =m\n2v\u0001v+kA\u0001v+krf\u0001v\u0000k\u001e\u0000\r(\u0010\u0000h) +rh\u0001v:\nSettingh=kfwe have a particularly interesting equivalent Lagrangian:\n\u0016L(x;q;v;\u0010 ) =m\n2v\u0001v+kA\u0001v\u0000k\u001e\u0000\r(\u0010+kf); (22)\nwhich is another gauge invariant realization for a particle moving with dissipation under a Lorentz force\ngiven by (\u001e;A;f). Indeed, performing a gauge transformation given by a function g, we obtain\n\u0016\u0016L(x;q;v;\u0010 ) =m\n2v\u0001v+kA\u0001v+krg\u0001v\u0000k\u001e\u0000\r(\u0010\u0000kg+kf);\nwhich is an equivalent Lagrangian to (22) given by the transformation s!s+kg.\nWhen performing a gauge transformation on the Lagrangian (22) we obtain an equivalent Lagrangian,\nas in the symplectic case. On the other hand, a gauge transformation leaves (17) invariant, with no need\nto invoke equivalence theorems. Notice that this is also the case when we recover the symplectic case\nsetting\r= 0. The triplet description of the electromagnetic gauge (\u001e;A;f), together with the term\nkA\u0001v+krf\u0001v\u0000k\u001ein the Lagrangian gives us a variational description of the Lorentz force where the\nLagrangian is invariant under gauge transformations (in both the classical and the contact settings).\n3.2.5 Example. Electron in a non-Perfect Conductor\nWe shall now apply the formalism discussed in Section 3.2.3 to describe the motion of an electron in a\nnon-perfect conductor.\nConsider a sufficiently large but finite cylindrical non-perfect conductor of length Land cross-section\nA. Let\u001bdenote the conductivity of the material. Assume a voltage \u0001Vis applied between the ends of the\nconductor, which is known to generate an electric field of constant magnitude\nE=\u0000\u0001V\nL\nin the longitudinal direction of the conductor. We take this direction to correspond to be the xcoordinate.\nAssume no magnetic fields intervene in the problem. Hence, one can take the vector potential to be A= 0\nand the gauge to be\n(\u001e;A;f) =\u0010\u0001V\nLx;0;0\u0011\nto describe the problem. Indeed, the observables for this choice as described in Equation (21) are\nE=\u0000\u0001V\nLexand B=R= 0;\nwhere exdenotes the unit vector in the xdirection. The equations of motion (20) read\n_v(t) =\u0000\u0016\u0001V\nL\u0000\rv(t) (23)\n14in thexdirection, where \u0016:=ke\nme<0, forkeandmethe charge and mass of the electron respectively. If\nwe assume the electron starts at rest at x= 0, the previous ODE can be solved for the velocity to find\nv(t) =\u0016\u0001V\n\rL\u0010\ne\u0000\rt\u00001\u0011\n; (24)\nand Equation (24) can be integrated to obtain\nx(t) =\u0000\u0016\u0001V\n\r2L(\rt+e\u0000\rt\u00001):\nNote that the electron reaches a limit velocity\nvlim= lim\nt!1v(t) =\u0000\u0016\u0001V\n\rL>0: (25)\nGiven that the conductor is ohmic, the limit velocity (or drift velocity) of an electron in the conductor can\nbe shown to be [12, p. 187]\nvlim=\u0000mmol\u001b\u0001V\n\u001akenL;\nwheremmolis the molecular mass of the conductor, \u001ais the density of the conductor and nthe number\nof free electrons per molecule of conductor. Hence, one can impose the limit velocity in (25) is the drift\nvelocity of the electron to find\n\r=\u0016ke\u001an\nmmol\u001b:\nIf we take copper as the material the conductor is made of, one finds that \r\u00194\u00011013s\u00001and hence\nthe limit velocity is achieved at the order of t\u001910\u000013s. This means that the electron has travelled\napproximately 3\u000110\u000016m. Thus, the assumption that Lis large enough so that the limit velocity is\nachieved is physically realizable.\nThe energy of the electron in the described electric field is known to be,\nE=me\n2v2+ke\u0001V\nLx\nand hence it dissipates at a rate\ndE\ndt=mev(t) _v(t) +ke\u0001V\nLv(t)\n=\u0000ke\u0016(\u0001V)2\n\rL2\u0010\ne\u0000\rt\u00001\u0011\ne\u0000\rt+ke\u0016(\u0001V)2\n\rL2\u0010\ne\u0000\rt\u00001\u0011\n=\n=\u0000ke\u0016(\u0001V)2\n\rL2\u0010\ne\u0000\rt\u00001\u00112\n;\nwhich, ast!1, when the drift velocity is achieved, becomes\ndE\ndt=\u0000ke\u0016(\u0001V)2\n\rL2=\u0000mmol\u001b(\u0001V)2\nn\u001aL2: (26)\nAssume there is no interaction between the electrons in the conductor, which implies that all of them\ndissipate energy at the same rate when they reach the drift velocity. The number of electrons in the\nconductor is given in the variables of the problem by\nne=nMT\nmmol=n\u001aAL\nmmol;\n15whereMTdenotes the total mass of the conductor. Hence, the total energy dissipation rate within the\nconductor when all electrons have achieved the drift velocity is\ndET\ndt=nedE\ndt=\u0000A\u001b\nL(\u0001V)2;\nwhich is precisely Joule’s heating law.\n3.2.6 Example. Particle in a Magnetic Field\nConsider a particle of charge kand massm, in a magnetic field B= (0;0;B). A choice of vector potential\nforBis\nR=A=B\n2(\u0000y;x;0);\ntakingf= 0, and if we assume no electric fields are present, the equations of motion (20) read\n8\n><\n>:_vx=\u0016Bvy\u0000\rvx+\r\u0016\n2By\n_vy=\u0000\u0016Bvx\u0000\rvy\u0000\r\u0016\n2Bx;\nwhere\u0016:=k\nm. This system of equations is not easy to discuss for arbitrary values of \r. However, if we\ndefine the energy of the particle as its kinetic energy, we see that\ndE\ndt=m\n2d\ndt(v2\nx+v2\ny) =m(vx_vx+vy_vy) =m\u0010\nvx(\u0016Bvy\u0000\rvx+\r\u0016\n2By) +vy(\u0000\u0016Bvx\u0000\rvy\u0000\r\u0016\n2Bx)\u0011\n=\u0000\rm(v2\nx+v2\ny) +m\r\u0016B\n2(vxy\u0000xvy) =\u0000\rmv2\u0000\r\u0016\n2B\u0001L:\nIt is known that a charged particle spinning in a magnetic field which is perpendicular to its velocity\nexperiences a dissipation of its energy due to the emission of radiation [30]. This effect is known as cyclotron\nradiation. The frequency of a particle of mass memitting cyclotron radiation in the classical limit is\nf=kB\n2\u0019m;\nand the energy dissipated due to the cyclotron radiation satisfies [28, Eq. 18.8]\ndE\ndt=\u0000\u001btB2v2\nc\u00160;\nwhere\u001bt=8\u0019\n3\u0010\nk2\n4\u0019\u000f0mc2\u00112\nis the Thomson total cross-section.\nIn order to argue about the nature of the motion of the particle, let us assume that the cross terms\n\r\u0016\n2Bxand\r\u0016\n2Bycan be neglected. Then, the system reads\n(_vx=\u0016Bvy\u0000\rvx\n_vy=\u0000\u0016Bvx\u0000\rvy;\nand can be solved by 8\n<\n:vx(t) =e\u0000\rt\u0000\nvx0cos(\u0016Bt) +vy0sin(\u0016Bt)\u0001\nvy(t) =e\u0000\rt\u0000\nvy0cos(\u0016Bt)\u0000vx0sin(\u0016Bt)\u0001\n;\nfrom which we deduce that the particle describes a decreasing spiral in the plane if \r >0. Note that the\nfrequency of the motion is exactly the frequency of a particle emitting cyclotron radiation.\n16Let the energy of the particle be\nE(t) =mv2(t)\n2=mv2\nx+v2\ny\n2=me\u00002\rt\n2v2\n0;\nwherev2\n0=v2\nx0+v2\ny0is the square of its initial velocity. Then, the energy is dissipated at a rate\ndE\ndt=\u0000m\re\u00002\rtv2\n0=\u0000m\rv2;\nand we only obtain the term in the cyclotron dissipation. Let us impose that the dissipated energy of the\nmodel is precisely the dissipation in the cyclotron radiation. Then,\n\r=\u001btB2\nc\u00160m>0: (27)\nIfwetaketheparticletobeanelectron, then \u0016\u0019\u00001:76\u00011011C=kg. IfweletB= 1T, then\r\u00190:1938s\u00001.\nLet usreturn now to thegeneral case, consideringthe cross-terms. Letus assume we canfix \ras inEquation\n(27) in order to estimate the solutions of the system. Assume that 0<\r\u001cj\u0016Bj. If we let\u0011=x+iy, then\n\u0011=\u0000_\u0011\u0000\ni\u0016B+\r\u0001\n\u0000i\r\u0016B\n2;\nand the eigenvalues of the characteristic polynomial are\n8\n>>><\n>>>:\u0000\r\u0000i\u0016B\n2+i\n2p\n\u00162B2\u0000\r2\u0019\u0000\r\n2\u0000i\r2\n4\u0016B\n\u0000\r\u0000i\u0016B\n2\u0000i\n2p\n\u00162B2\u0000\r2\u0019\u0000\r\n2\u0000i\u0016B+i\r2\n4\u0016B:\nThe frequency of the motion is altered in the order O(\r2\n\u0016B). In addition, if 0< \r\u001cj\u0016Bj, the particle\nwill describe a slightly perturbed decreasing spiral motion for small times. The term\n\u0000\r\u0016\n2B\u0001L\nis always positive, but it can be seen that, when 0<\r\u001cj\u0016Bj, it is smaller in norm than the dissipative\nterm due to the cyclotron radiation. Hence, the particle does indeed lose energy, but at a lower rate than in\nthe cyclotron. We are modelling the small-time behaviour of a particle inside a magnetic field that dissipates\nenergy due to the emission of a cyclotron radiation that has been altered by the interaction between the\nexternal magnetic field and the angular momentum of the particle itself.\n17Figure 1: Position (left) and energy (right) of an electron starting at the origin with velocity \u00001:76\u0001\n1011m=seybetweent= 0sandt= 10s. Darker color represents larger times. Note that the energy of the\nelectron follows an exponential pattern plus an oscillating fashion of period \u00185\u000110\u000011sand amplitude of\nthe order of 10\u000018J.\n4 Electromagnetic Fields\n4.1 Covariant Formulation of Classical Electromagnetism\nIn Section 4.1, we introduce the main objects and tools of the covariant formulation of classical elec-\ntromagnetism, so that our work is self-contained. Throughout Section 4, Einstein’s summation conven-\ntion will be used unless specifically mentioned, and the considered metric in Minkowski space will be\n\u0011\u000b\u0016=\u0011\u000b\u0016=diag(1;\u00001;\u00001;\u00001).\nRecall that the four-displacement tensor is defined as x\u0016= (ct;x;y;z ), wherecis the speed of light in\nvacuum, and the covariant four-gradient is @\u0016=@\n@\u0016=\u0010\n1\nc@\n@t;r\u0011\n. IfC\u000bis a four-tensor, we will use the\nnotation@\u0016C\u000b=C\u000b;\u0016.\nThe main object of this formulation of electromagnetism is the covariant antisymmetric tensor\nF\u000b\u0016=0\nBBBBBB@0\u0000Ex=c\u0000Ey=c\u0000Ez=c\nEx=c 0\u0000BzBy\nEy=c Bz 0\u0000Bx\nEz=c\u0000ByBx 01\nCCCCCCA(28)\nfor a pair of electric and magnetic vector fields E= (Ex;Ey;Ez);B= (Bx;By;Bz). The tensor F\u000b\u0016is\nknown as the electromagnetic tensor. If \u001eandAare a choice of scalar and vector potentials of Eand\nBunder the classical gauge defined in Section 3.1.2, the electromagnetic four-potential is defined to be\nA\u000b=\u0010\n\u001e\nc;A\u0011\n, which satisfies F\u000b\u0016=@\u000bA\u0016\u0000@\u0016A\u000b=A\u0016;\u000b\u0000A\u000b;\u0016.\nFinally, an electric charge density \u001aand an electric current density jdefine the tensor J\u000b= (c\u001a;j),\nwhich is known as the four-current.\nThe four Maxwell’s equations in vacuum in vector notation reduce to two tensor equations. The first,\n18known as the Gauss-Faraday law, reads\n@\u0016\u00101\n2\u000f\u0016\u000b\f\u001cF\f\u001c\u0011\n= 0; (29)\nwhere\u000f\u0016\u000b\f\u001cis the Levi-Civita tensor, and it comes from the fact that @\u0016F\u0017\u0015+@\u0017F\u0015\u0016+@\u0015F\u0016\u0017= 0.\nThe Gauss-Faraday law is the same in vacuum and in media, and thus it will not be central in our\ndiscussion. The other Maxwell’s equation is known as de Gauss-Ampère law and reads\n@\u0016F\u0016\u000b=\u00160J\u000b; (30)\nwhere\u00160is the magnetic permeability of vacuum.\nThe Lagrangian density for classical electromagnetism is defined to be\nL=\u00001\n4\u00160F\u000b\u0016F\u000b\u0016\u0000A\u000bJ\u000b; (31)\nand it derives the Gauss-Ampère law via the Euler-Lagrange equation for fields [4, Ch. 1.10]. The electro-\nmagnetic energy density is defined as u=\u000f0\n2E\u0001E+1\n2\u00160B\u0001B, and the energy flux density is given by the\nPoynting’s vector field S=E\u0002B\n\u00160.\nWhen considering electromagnetic fields in matter, the polarization density and magnetization density\nvector fields PandMencode the response of the medium to the incoming electric and magnetic vector\nfields, see [12]. The electric displacement vector is then defined as D=\u000f0E+P, where\u000f0is the electric\npermittivity of vacuum, and the magnetic intensity is H=1\n\u00160B\u0000M.\nThese quantities can be absorbed into the antisymmetric magnetisation-polarisation tensor\nM\u000b\u0016=0\nBBBBBB@0Pxc Pyc Pzc\n\u0000Pxc 0\u0000MzMy\n\u0000Pyc Mz 0\u0000Mx\n\u0000Pzc\u0000MyMx 01\nCCCCCCA(32)\nand the antisymmetric electromagnetic displacement tensor\nD\u000b\u0016=0\nBBBBBB@0\u0000Dxc\u0000Dyc\u0000Dzc\nDxc 0\u0000HzHy\nDyc Hz 0\u0000Hx\nDzc\u0000HyHx 01\nCCCCCCA; (33)\nwhich are related to the electromagnetic tensor via\nD\u000b\u0016=1\n\u00160F\u000b\u0016\u0000M\u000b\u0016: (34)\nLet us also recall that the bound current in a material is defined as\nJ\u000b\nbd:= (c\u001abd;jbd):=@\u000bM\u000b\u0016=\u0000\n\u0000cr\u0001P;@P\n@t+r\u0002M\u0001\n: (35)\nAs discussed above, the Gauss-Faraday law (29) does not change when considering fields in matter.\nHowever, the Gauss-Ampère law (30) becomes\n@\u0016D\u0016\u000b=J\u000b; (36)\n19which is known as the Gauss-Ampère law in matter. The electromagnetic energy density in a material\nbecomesumatter =1\n2\u0000\nE\u0001D+B\u0001H\u0001\nand the Poynting’s vector is Smatter =E\u0002H.\n4.2 k-contact Formulation for Electromagnetic Fields\nIn the current section, we develop a theory that allows us to model linear and dissipative electromagnetic\nsystems, as well as systems with contributions from both. The linear features are obtained by replacing the\nMinkowski metric of the spacetime with a diagonal metric that encodes information of the material, whilst\nthe dissipative facet is modelled by introducing a linear term in the action to the Lagrangian density. Our\ndiscussion generalises the results derived in [27].\nRecall that a linear material is such that there exist constants \u001feand\u001fmfor which\n8\n><\n>:P=\u000f0\u001feE\nM=1\n\u00160\u001fm\n1 +\u001fm:=\u001fm\n\u0016B:(37)\nThe Gauss-Ampère law for a linear material can be obtained from the electromagnetic Lagrangian\ndensity in vacuum where the Minkowski metric has been replaced. For a particular four-current J\u000b:U\u001a\nM4!Q\u001aR4, define\nL=\u00001\n4\u00160g\u000b\u0016g\f\u0017F\u0016\u0017F\u000b\f\u0000A\u000bJ\u000b2C1\u00104M\ni=1TQ\u0011\n;\nwith the symmetric bilinear form\ng=g\u0016\u0017=1p1 +\u001fmdiag\u0010\n(1 +\u001fe)(1 +\u001fm);\u00001;\u00001;\u00001\u0011\nonU.\nThe Euler-Lagrange equations for fields imply g\u0016\u001bg\u000b\u001c@\u0016F\u001b\u001c=\u00160J\u000b, which reads\n8\n>><\n>>:\u00160c\u001a=\u00160J0=g\u0016\u001bg0\u001c@\u0016F\u001b\u001c=g00gii@iFi0=1 +\u001fe\ncr\u0001E() r\u0001 D=\u001a\n\u00160ji=\u00160Ji=giig\u0016\u001b@\u0016F\u001bi=giig00\nc2@Ei\n@t+ (gii)2\u0000\nr\u0002B\u0001\ni()j=\u0000@D\n@t+r\u0002H;\nthe non-geometric Maxwell’s equations for a material satisfying (37).\nWe now add a linear term in the action to this Lagrangian density in order to model a larger subset of\nmaterials. Let\nL=\u00001\n4\u00160g\u000b\u0016g\f\u0017F\u0016\u0017F\u000b\f\u0000A\u000bJ\u000b\u0000\r\u000bs\u000b2C1\u00104M\ni=1TQ\u0002R\u0011\n:\nFollowing the discussion in [18], the Lagrangian energy density defined by Lis\nEL=A\u0016;\u000b@L\n@A\u0016;\u000b\u0000L=1\n\u00160g\u0016\u0017g\u000b\fA\u0016;\u000bF\u0017\f+1\n4\u00160g\u0016\u0017g\u000b\fF\f\u0017F\u000b\u0016+A\u000bJ\u000b+\r\u000bs\u000b\nand gives rise to four contact 1-forms\u0012\u000b\nL2\n\u00104L\ni=1TQ\u0002R4\u0011\ndefined by\n\u0012\u000b\nL=ds\u000b\u0000@L\n@A\u0016;\u000bdA\u0016=ds\u000b+1\n\u00160g\u000b\fg\u0016\u0017F\f\u0017dA\u0016:\n20The usual Reeb vector fields of the form R\u000b=@\n@s\u000bfullfill the conditions\n\u0013R\u000b\u0012\f=\u000e\f\n\u000b;\u0013R\u000bd\u0012\f= 0;\nbut they are not unique. We can construct all the solutions of the previous equations by adding a general\nterm of the form R\u000b=R\u000b+F\u0016\u0017;\u000b@\n@A\u0016;\u0017withF\u0016\u0017;\u000b\u0000F\u0017\u0016;\u000b = 0. Fortunately, \u0013(R\u000b)dEL=\r\u000bfor any\npossible (antisymetric) F\u0016\u0017;\u000b, therefore the choice of Reeb vector fields does not change the equations. From\nnow on we will use R\u000b=@\n@s\u000b.\nThe equations of motion for a k-vector field X\u000b\n\u0013X\u000bd\u0012\u000b\nL=dEL\u0000\r\u000b\u0012\u000b\nL\n\u0013X\u000b\u0012\u000b\nL=\u0000EL\nhence imply\n@\u000bs\u000b=L\nF\u001c\f=A\f;\u001c\u0000A\u001c;\f\n\u00160J\u0016=g\u0017\u001bg\u0016\u001c\u0010\n@\u0017F\u001b\u001c+\r\u0017F\u001b\u001c\u0011\n:(38)\n(39)\n(40)\nLetting\r\u0016= (\r\nc;\r\r\r), Equation (40) reads in vector notation as\n(1 +\u001fe)\u0000\nr\u0001E+\r\r\r\u0001E\u0001\n=\u001a\n\u000f0\n\u0000(1 +\u001fe)\u000f0\u0000@E\n@t+\rE\u0001\n+1\n1 +\u001fm\u0000\nr\u0002B\n\u00160+\r\r\r\u0002B\n\u00160\u0001\n=j;(41)\n(42)\nwhich we refer to as the k\u0000contact Maxwell’s equations. These reduce to the Gauss-Ampère law for linear\nmaterials (or vacuum) when \r\u0017= 0.\n4.3 Generalized Poynting’s Theorem\nWe next derive a generalized Poynting’s Theorem for the obtained k\u0000contact Maxwell equations in order\nto discuss when the modelled systems are dissipative. Dot multiplying Equation (42) by the vector field E\nand using Faraday’s law of induction ( r\u0002E=\u0000@B\n@t), we find\nE\u0001j=\u0000(1 +\u001fe)\u000f0\u00101\n2@E\u0001E\n@t+\rE\u0001E\u0011\n\u00001\n1 +\u001fm\u00101\n2\u00160@B\u0001B\n@t+r\u0001S+\r\r\r\u0001S\u0011\n;\nand integrating over a volume Vwe obtain\nZ\nV@u\n@tdV+Z\nS=@VS\u0001^ndS\n=\u0000Z\nVE\u0001jdV\u0000Z\nVE\u0001\u0010\n\u001fe\u000f0@E\n@t+ (1 +\u001fe)\u000f0\rE+\u001fm\n1 +\u001fmr\u0002B\n\u00160\u00001\n1 +\u001fm\r\r\r\u0002B\n\u00160\u0011\ndV;\nwhich we refer to as the generalised Poynting’s theorem. In order to shed some light on the utility of this\nresult, let us consider the case in which \u001fe=\u001fm= 0. Then, the generalised Poynting’s theorem reads\nZ\nV@u\n@tdV+Z\nS=@VS\u0001^ndS=\u0000Z\nVE\u0001jdV\u0000Z\nVE\u0001\u0010\n\u000f0\rE\u0000\r\r\r\u0002B\n\u00160\u0011\ndV;\nand the energy of the electromagnetic fields is being dissipated by both the real current jand by a virtual\ncurrent j\r:=\u000f0\rE\u0000\r\r\r\u0002B\n\u00160. Note that E\u0001j\r=\u000f0\rE\u0001E+\r\r\r\u0001S. Whenever the vector \r\r\rvanishes and \r >0,\nthen the volume integralZ\nVE\u0001j\rdV=\u000f0\rZ\nVE\u0001EdV\n21is positive, and thus the modelled system is indeed dissipative. This discussion agrees with that made in\n[27]. In general, the virtual current is j\r=\u001fe\u000f0@E\n@t+ (1 +\u001fe)\u000f0\rE+\u001fm\n1+\u001fmr\u0002B\n\u00160\u00001\n1+\u001fm\r\r\r\u0002B\n\u00160, and\nE\u0001j\r=1\n2@\n@t\u0010\n\u001fe\u000f0E\u0001E\u0000\u001fm\n1 +\u001fm1\n\u00160B\u0001B\u0011\n+ (1 +\u001fe)E\u0001E+\u001fm\n1 +\u001fmr\u0001S+1\n1 +\u001fm\r\r\r\u0001S:\nOne can also argue as follows. Following the constitutive relations for DandHin a linear material, let\n~D:= (1+\u001fe)\u000f0Eand~H:=1\n1+\u001fm1\n\u00160B. Then, define ~u:=1\n2\u0010\nE\u0001~D+B\u0001~H\u0011\n=1\n2\u0010\n(1+\u001fe)\u000f0E\u0001E+1\n1+\u001fm1\n\u00160B\u0001B\u0011\nand~S:=E\u0002~H=1\n1+\u001fmS. Then, the generalized Poynting’s theorem implies\nZ\nV@~u\n@tdV+Z\nS=@V~S\u0001^ndS=\u0000Z\nVE\u0001jdV\u0000Z\nVE\u0001\u0010\n\r~D\u0000\r\r\r\u0002~H\u0011\ndV;\nand the virtual current dissipating energy from the system becomes ~j\r=\r~D\u0000\r\r\r\u0002~H.\n4.4 k-contact Maxwell’s equations as Maxwell’s equations in matter\nAs discussed at the end of Section 4.2, the k-contact Maxwell’s equations are precisely Maxwell’s equations\nfor a linear material or vacuum when \r\u0017= 0. In general, imposing that the obtained equations of motion\nare precisely the Gauss-Ampère law is equivalent to the existence of constants \u001feand\u001fmand a four tensor\n\r\u0017for which the identity g\u0017\u001bg\u0016\u001c\u0010\n@\u0017F\u001b\u001c+\r\u0017F\u001b\u001c\u0011\n=\u00160\u0011\u0017\u000b\u0011\u0016\f@\u0017D\u000b\fis satisfied. Looking at the k-contact\nMaxwell’s equations, it is sufficient that the material satisfies\nr\u0001P=\u001fe\u000f0r\u0001E+ (1 +\u001fe)\u000f0\r\r\r\u0001E\n@P\n@t=\u001fe\u000f0@E\n@t+ (1 +\u001fe)\u000f0\rE\nr\u0002M=\u001fm\n1 +\u001fmr\u0002B\n\u00160\u00001\n1 +\u001fm\r\r\r\u0002B\n\u00160(43)\nfor certain constants \u001fe;\u001fm;\r2Rand\r\r\r2R3. Note that it is necessary that the incoming electric field\nfulfils\r\r\r\u0001@E\n@t=\r0r\u0001E.\nNote that whenever these relations are satisfied, and hence our framework models electromagnetism\nin matter, the bound current jbdinduced by the material is precisely the virtual current appearing in the\ngeneralized Poynting’s theorem in Section 4.3, that is\njbd=@P\n@t+r\u0002M=\u001fe\u000f0@E\n@t+ (1 +\u001fe)\u000f0\rE+\u001fm\n1 +\u001fmr\u0002B\n\u00160\u00001\n1 +\u001fm\r\r\r\u0002B\n\u00160=j\r:\nHence, our framework models dissipative electromagnetic systems whenever the bound current jbd=j\r\nis a dissipative term, as seen in the generalized Poynting’s Theorem.\n4.5 Example. The Lorentz Dipole Model\nThe Lorentz dipole models the interaction between an oscillating electric field and an electron [23, Ch. 3].\nThe model assumes the electron behaves like a harmonic oscillator attached to the nucleus by a hypothetical\nspring of constant C, and that the oscillations are driven by the electric field. The source of damping is not\nspecified but comes from a drag force ~F=\u0000\u0016~ veon the electron. Let mbe the mass of the electron and\nqe>0the absolute value of it’s charge.\n22The solution to the equations of motion can be found assuming the incoming field to be of the form\nE(r;t) =E0eik\u0001r\u0000i!t, where kis called the wave vector and !is the frequency. One can derive the\nconstitutive relation of the polarization P=\u000f0(\u001f0\u0000i\u001f00)Eof a material formed of Nsuch atoms that do\nnot interact with each other, where the electric susceptibility becomes a complex quantity with\n\u001f0=!2\np(!2\n0\u0000!2)\n(!2\n0\u0000!2)2+ (!=\u001c)2\u001f00=!2\np!=\u001c\n(!2\n0\u0000!2)2+ (!=\u001c)2;\nwhere!0=q\nC\nmis the natural frequency of the electron-nucleus system and !2\np=Nqe\n\u000f0m. We have also\ndefined\u001c=m\n\u0016. The complex part of the susceptibility models the absorption of light by the material, and\nhence the dissipation of electromagnetic energy of the incoming field.\nNote that@P\n@t=\u000f0(\u001f0\u0000i\u001f00)@E\n@t=\u000f0\u001f0@E\n@t\u0000\u000f0\u001f00!E\nr\u0001P=\u000f0(\u001f0\u0000i\u001f00)r\u0001E=\u000f0\u001f0r\u0001E+\u000f0\u001f00k\u0001E;\nand since no magnetic effects are considered, it is sufficient to define \u001fe=\u001f0,\u001fm= 0and\r=\u0000\u001f00\n1+\u001f0!,\n\r\r\r=\u001f00\n1+\u001f0kfor Equations (43) to be satisfied.\n4.6 Example. Highly Resistive Dielectric\nConsider a parallel capacitor filled with a dielectric. The induced vector field is E=\u0001V\ndez, where \u0001V\nis the constant applied voltage difference, dis the distance between the plates and ezis the unit vector\northogonal to the plates. Assume no external electric or magnetic fields intervene and that the dielectric is\nnon-magnetic.\nAssume that the dielectric is composed of spherical molecules with a high resistivity. These are of radius\nRand separated a distance s\u001dR. The electric field inside the dielectric away from the centre of the spheres\nis essentially constant and equal to\u0001V\ndez.\nDue to the high resistivity of the molecules, the accumulation of charge at the poles as a response\nto the applied electric field is described as a rate, rather than a magnitude, which is proportional to the\nexternal electric field, see [22, Sec. 11.5]. Thus, we expect@P\n@tto be proportional to E, contrary to when\nthe spheres are perfect conductors, which makes the dielectric linear and hence Pproportional to E. Thus,\nthe polarization constitutive law for a short initial period of time reads\n@P\n@t=\u000bE;\nwhere\u000bis a positive real constant dependant only on the properties and geometry of the spheres.\nHence, one can take \u001fe=\u001fm= 0and\r=\u000b\n\u000f0c. Also, if no interaction between the spheres is considered,\nr\u0001D= 0andr\u0001P= 0. Also, since the magnetization of the molecules is zero, r\u0002M= 0and if we let\n\r\r\r= 0,\n0 =\r\r\r\u0001E=r\u0001P= 0\nand\n0 =\u0000\r\r\r\u0002B=\u00160r\u0002M= 0\n23and hence Equations (43) are satisfied. Since \r\r\r= 0and\r0>0, the bound current is a source of dissipation,\nas discussed in Section 4.3. This coincides with the conclusions drawn in [22, Sec. 11.5] by means of physical\narguments.\n4.7 Gauge Invariance\nIn the current section we show how the developed k-contact theory for electromagnetic fields is invariant\nunder the classical gauge theory for fields.\nLetQ=R4be the space in which the four tensor A\u0016takes its values, and consider a general function\nf:Q!R. The change of gauge given by fis\nA0\n\u0016=A\u0016\u0000@\u0016f:\nLet us consider the case in which the external current J\u000bvanishes. Let the primed variables denote the\nobjects in the new gauge, and the unprimed variables denote the objects in the old gauge. Note first that\nF0\n\u000b\u0016=@\u000bA0\n\u0016\u0000@\u0016A0\n\u000b=F\u000b\u0016\u0000@\u000b\u0016f+@\u000b\u0016f=F\u000b\u0016;\nand henceF\u000b\u0016is invariant under the gauge. Thus, the Lagrangian\nL=\u00001\n4\u00160g\u000b\u0016g\f\u0017F\u0016\u0017F\u000b\f\u0000\r\u000bs\u000b\nis invariant under the gauge. Also, the energy satisfies\nE0\nL=1\n\u00160g\u0016\u0017g\u000b\fA0\n\u0016;\u000bF0\n\u0017\f+1\n4\u00160g\u0016\u0017g\u000b\fF0\n\f\u0017F0\n\u000b\u0016+\r\u000bs\u000b=EL\u00001\n\u00160g\u0016\u0017g\u000b\fF\u0017\f@\u000b\u0016f\nwhere the last equality follows from the fact that the tensor @\u000b\u0016fis symmetric while F\u000b\u0016is antisymmetric.\nIt is also clear that the equations of motion\n0 =g\u0017\u001bg\u0016\u001c\u0010\n@\u0017F\u001b\u001c+\r\u0017F\u001b\u001c\u0011\nare invariant under the gauge.\nThus, ourk-contact theory and all of the objects involved are invariant under the classical gauge theory\nfor electromagnetic fields.\n5 Conclusions\nThe work presented here contributes to the understanding of how contact geometry can be applied in classi-\ncal mechanics and classical field theory to describe systems with dissipation and damping. A lot of research\nhas been done lately to build a general theory, discovering conservation and dissipation theorems and ana-\nlogues to classical results such as Noether’s theorem [10,16–18], but no extensive studies on applications to\ncurrent theories and real-life examples have been made apart from small examples in [10,16–18,27].\nThe present work focuses on how contact formalism can be applied to electromagnetic systems with\ndissipation. It has been studied how the techniques introduced by G. Herglotz in 1930, and recently\nformalized using the language of contact geometry, can be applied to particles under the influence of a\n24Lorentz force and external damping. Moreover, a recent theory developed in [16,18], which generalizes\ncontact formalism to study classical field theory, has been used to discuss electromagnetic fields themselves\nwhen being dissipated by external phenomena.\nThe contact formalism of classical mechanics deals with systems whose Lagrangian function or density\ndepends on the action itself. Driven by previous successful applications of the theory, in the present work\nthe Lagrangians under study have been taken to be the classical Lagrangians of electromagnetism (the\nLagrangian of the Lorentz force and the Lagrangian density of electromagnetic fields) plus a linear term in\nthe action.\nWhen applying the theory to particles under the Lorentz force, the produced equations of motion\nhave turned out to be not invariant under the current gauge theory. This has driven us to propose a\nnew gauge approach for electromagnetism which generalizes the classical one and reduces to the currently\naccepted theory for non-dissipative symplectic systems. It has also been necessary to slightly change the\nLagrangian of the Lorentz force, adding an extra term which produces the same equations of motion in\nthe non-dissipative case but that is key in the contact formalism. All of these generalizations make a new\nmathematical observable vector field appear, which can be seen to vanish from the equations of motion\nin the symplectic case. Moreover, in this new paradigm, the generalized momentum of a particle under a\nLorentz force becomes an observable, whilst in the symplectic theory of electromagnetism the generalized\nmomenta are not gauge-invariant.\nThe developed approach for particles under a Lorentz force in the presence of external damping has been\napplied to two real-life scenarios. Firstly, our theory is able to model the behaviour of an electron inside\na non-perfect conductor. When imposing that the limit velocity of the particle agrees with the current\ntheory, Joules’ heating law is recovered. It has also been discussed how our description can be applied to a\nparticle in a magnetic field with dissipation.\nWhen considering vector fields, we have produced a new set of Gauss-Ampère equations which reduce\nto the classical Gauss-Ampère equation when omitting the dissipation parameters. In the general case,\nwe have showed how our theory models electromagnetic fields in matter for a certain type of systems. It\nhas been argued how these systems might have in common that the bound four-current generated by the\nmaterial is precisely a dissipative term. Further research is needed in order to better understand such\nsystems and characterize them further. Our theory has been seen to be able to describe a particular regime\nof the Lorentz dipole model.\nThis work pretends to be a first approach towards the use of contact geometry techniques to model\ndissipation in electromagnetism. It should be further studied whether the generalizations needed to make\nthe theory gauge invariant for the Lorentz force can actually be seen as generalizations of the current\ntheory of electromagnetism. For fields, more research is needed to better understand how the modelled\nsystems behave and find whether other variations of the Lagrangian can describe different systems. It is\nalso necessary to produce more and more relevant examples of how our equations can be applied to real\nproblems in Physics and reproduce current data and predict new phenomena. Furthermore, the contact\nframework needs to be further developed so that it admits non-autonomous Lagrangians which depend\nexplicitly on time or the components of spacetime, as these arise naturally in many theories in Physics.\n25References\n[1] V. I. Arnold, Singularities of caustics and wave fronts , Kluwer Academic Publishers, Dordrecht, Netherlands Boston,\n1990.\n[2] A. Bravetti, Contact geometry and thermodynamics , International Journal of Geometric Methods in Modern Physics 16\n(2019), no. 1, 1940003.\n[3] A. Bravetti and D. Tapias, Thermostat algorithm for generating target ensembles , Physical Review E 93(2016feb), no. 2,\n22139.\n[4] S. Carroll, Spacetime and geometry: an introduction to general relativity , Addison Wesley, San Francisco, 2004.\n[5] M. de León, J. Gaset, M. Lainz Valcázar, M. C. Muñoz-Lecanda, and N. Román-Roy, Higher-order contact mechanics ,\nAnnals of Physics 425(2021), no. 34, 168396.\n[6] M. de León, J. Gaset, M. Lainz Valcázar, X. Rivas, and N. Román-Roy, Unified Lagrangian-Hamiltonian Formalism for\nContact Systems , Fortschritte der Physik 68(2020), 2000045.\n[7] M. de León, M. Lainz Valcázar, and Á. Muñiz-Brea, The Hamilton–Jacobi theory for contact Hamiltonian systems (2021),\n1–33 pp., available at 2103.17017 .\n[8] M. de León, M. Lainz Valcázar, and M. C. Muñoz-Lecanda, Optimal control, contact dynamics and Herglotz variational\nproblem (2020), available at 2006.14326 .\n[9] M. de León and M. Lainz Valcázar, Contact hamiltonian systems , Journal of Mathematical Physics 60(2019oct), no. 10,\n102902.\n[10] M. de León and M. Lainz Valcázar, Singular lagrangians and precontact hamiltonian systems , International Journal of\nGeometric Methods in Modern Physics 16(2019oct), no. 10, 1950158.\n[11] M. de León and M. Lainz Valcázar, Infinitesimal symmetries in contact hamiltonian systems , Journal of Geometry and\nPhysics 153(2020jul), 103651.\n[12] D. J. Morin E. M. Purcell, Electricity and magnetism , 3rd ed., Cambridge University Press, 2013.\n[13] J. B. Etnyre, Introductory lectures on contact geometry (2001nov), available at math/0111118 .\n[14] J. B. Etnyre and R. Ghrist, Contact topology and hydrodynamics: I. beltrami fields and the seifert conjecture , Nonlinearity\n13(2000feb), no. 2, 441–458.\n[15] J. Gaset, A contact geometry approach to symmetries in systems with dissipation (2020), 1–5 pp., available at 2003.03137 .\n[16] J. Gaset, X. Gràcia, M. C. Muñoz-Lecanda, X. Rivas, and N. Román-Roy, A contact geometry framework for field theories\nwith dissipation , Annals of Physics 414(2020mar), 168092.\n[17] J. Gaset, X. Gràcia, M. C. Muñoz-Lecanda, X. Rivas, and N. Román-Roy, New contributions to the hamiltonian and la-\ngrangian contact formalisms for dissipative mechanical systems and their symmetries , International Journal of Geometric\nMethods in Modern Physics 17(2020may), no. 06, 2050090.\n[18] J. Gaset, X. Gràcia, M. C. Muñoz-Lecanda, X. Rivas, and N. Román-Roy, A k-contact lagrangian formulation for\nnonconservative field theories , Reports on Mathematical Physics 87(2021jun), no. 3, 347–368.\n[19] H. Geiges, An introduction to contact topology , Cambridge University Press, 2008.\n[20] S. Goto, Contact geometric descriptions of vector fields on dually flat spaces and their applications in electric circuit\nmodels and nonequilibrium statistical mechanics , Journal of Mathematical Physics 57(2016oct), no. 10, 102702.\n[21] S. Grillo and E. Padrón, Extended Hamilton–Jacobi theory, contact manifolds, and integrability by quadratures , Journal\nof Mathematical Physics 61(2020jan), no. 1, 012901.\n[22] H. Haus, Electromagnetic fields and energy , Prentice Hall, Englewood Cliffs, New Jersey, 1989.\n[23] E. Hecht, Optics, Pearson, Boston, Massachusetts, 2017.\n[24] G. Herglotz, Berührengstransformationen , Lectures at the University of Göttingen (1930).\n26[25] G. Herglotz, Vorlesungen über die mechanik der kontinua , Teubner-Archiv zur Mathematik, Springer Vienna, Vienna,\n1985.\n[26] X. Li-Jost J. R. Jost, Calculus of variations , Cambridge University Press, 2007.\n[27] M. J. Lazo, J. Paiva, J. T. S. Amaral, and G. S. F. Frederico, An action principle for action-dependent lagrangians:\nToward an action principle to non-conservative systems , Journal of Mathematical Physics 59(2018mar), no. 3, 032902.\n[28] M. S. Longair, Stars, the galaxy and the interstellar medium , High energy astrophysics, vol. 2, Cambridge University\nPress, 1994.\n[29] D. Luenberger, Optimization by vector space methods , Wiley, New York, 1968.\n[30] B. Monreal, Single-electron cyclotron radiation , Physics Today 69(2016jan), no. 1, 70–71.\n[31] J. D. Mrugala R.and Nulton, J. C. Schön, and P. Salamon, Contact structure in thermodynamic theory , Reports on\nMathematical Physics 29(1991feb), no. 1, 109–121.\n[32] M. Razavy, Classical and quantum dissipative systems , World Scientific, 2016.\n[33] A. A. Simoes, M. de León, M. Lainz Valcázar, and D. Martín de Diego, Contact geometry for simple thermodynamical\nsystems with friction ,ProceedingsoftheRoyalSocietyA:Mathematical,PhysicalandEngineeringSciences 476(2020sep),\nno. 2241, 20200244.\n[34] A. A. Simoes, D. Martín de Diego, M. Lainz Valcázar, and M. de León, The geometry of some thermodynamic systems ,\narXiv (2020), available at 2012.07404 .\n[35] P. J. Vassiliou, Contact geometry and its application to control , Advances in the theory of control, signals and systems\nwith physical modeling, 2010, pp. 225–237.\n27" }, { "title": "1601.00069v2.Atmospheric_Circulation_of_Hot_Jupiters__Dayside_Nightside_Temperature_Differences.pdf", "content": "Submitted to The Astrophysical Journal\nPreprint typeset using L ATEX style emulateapj v. 5/2/11\nATMOSPHERIC CIRCULATION OF HOT JUPITERS: DAYSIDE-NIGHTSIDE TEMPERATURE\nDIFFERENCES\nThaddeus D. Komacek1and Adam P. Showman1\n1Department of Planetary Sciences, University of Arizona, Tucson, AZ, 85721\ntkomacek@lpl.arizona.edu\nSubmitted to The Astrophysical Journal\nABSTRACT\nThe full-phase infrared light curves of low-eccentricity hot Jupiters show a trend of increasing\ndayside-to-nightside brightness temperature di\u000berence with increasing equilibrium temperature. Here\nwe present a three-dimensional model that explains this relationship, in order to shed insight on\nthe processes that control heat redistribution in tidally-locked planetary atmospheres. This three-\ndimensional model combines predictive analytic theory for the atmospheric circulation and dayside-\nnightside temperature di\u000berences over a range of equilibrium temperature, atmospheric composition,\nand potential frictional drag strengths with numerical solutions of the circulation that verify this\nanalytic theory. This analytic theory shows that the longitudinal propagation of waves mediates\ndayside-nightside temperature di\u000berences in hot Jupiter atmospheres, analogous to the wave adjust-\nment mechanism that regulates the thermal structure in Earth's tropics. These waves can be damped\nin hot Jupiter atmospheres by either radiative cooling or potential frictional drag. This frictional drag\nwould likely be caused by Lorentz forces in a partially ionized atmosphere threaded by a background\nmagnetic \feld, and would increase in strength with increasing temperature. Additionally, the ampli-\ntude of radiative heating and cooling increases with increasing temperature, and hence both radiative\nheating/cooling and frictional drag damp waves more e\u000eciently with increasing equilibrium tempera-\nture. Radiative heating and cooling play the largest role in controlling dayside-nightside temperature\ntemperature di\u000berences in both our analytic theory and numerical simulations, with frictional drag\nonly being important if it is stronger than the Coriolis force. As a result, dayside-nightside temper-\nature di\u000berences in hot Jupiter atmospheres increase with increasing stellar irradiation and decrease\nwith increasing pressure.\nSubject headings: hydrodynamics - methods: numerical - methods: analytical - planets and satel-\nlites: gaseous planets - planets and satellites: atmospheres - planets and satel-\nlites: individual (HD 189733b, HD 209458b, WASP-43b, HD 149026b, WASP-14b,\nWASP-19b, HAT-P-7b, WASP-18b, WASP-12b)\n1.INTRODUCTION\nHot Jupiters, gas giant exoplanets with small semi-\nmajor axes and equilibrium temperatures exceeding\n1000 K, are the best characterized class of exoplanets to\ndate. Since the \frst transit observations of HD 209458b\n(Henry et al. 2000; Charbonneau et al. 2000), infrared\n(IR) phase curves have been obtained for a variety of\nobjects (e.g. Knutson et al. 2007; Cowan et al. 2007;\nBorucki et al. 2009; Knutson et al. 2009a,b; Cowan et al.\n2012; Knutson et al. 2012; Demory et al. 2013; Maxted\net al. 2013; Stevenson et al. 2014; Zellem et al. 2014;\nWong et al. 2015a). Such phase curves allow for the\nconstruction of longitudinally resolved maps of surface\nbrightness. These maps exhibit a wide diversity, showing\nthat|across the class of hot Jupiters|the fractional dif-\nference between dayside and nightside \rux varies drasti-\ncally from planet to planet. Figure 1 shows the fractional\ndi\u000berence between dayside and nightside brightness tem-\nperatures as a function of equilibrium temperature for\nthe nine low-eccentricity transiting hot Jupiters with\nfull-phase IR light curve observations. This fractional\ndi\u000berence in dayside-nightside brightness temperature,\nAobs, has a value of zero when hot Jupiters are longitu-\ndinally isothermal and unity when the nightside has ef-\nfectively no emitted \rux relative to the dayside. As seenin Figure 1, the fractional dayside-nightside tempera-\nture di\u000berence increases with increasing equilibrium tem-\nperature. The correlation between fractional dayside-\nnightside temperature di\u000berences and stellar irradiation\nshown in Figure 1 has also been found by Cowan & Agol\n(2011), Perez-Becker & Showman (2013), and Schwartz\n& Cowan (2015).\nMotivated by these observations, a variety of groups\nhave performed three-dimensional (3D) numerical sim-\nulations of the atmospheric circulation of hot Jupiters\n(e.g. Showman & Guillot 2002; Cooper & Showman 2005;\nMenou & Rauscher 2009; Showman et al. 2009; Thrastar-\nson & Cho 2010; Heng et al. 2011b,a; Perna et al. 2012;\nRauscher & Menou 2012a; Dobbs-Dixon & Agol 2013;\nMayne et al. 2014; Showman et al. 2015). These general\ncirculation models (GCMs) generally exhibit day-night\ntemperature di\u000berences ranging from \u0018200{1000 K (de-\npending on model details) and fast winds that can ex-\nceed several km s\u00001. When such models include realis-\ntic non-grey radiative transfer, they allow estimates of\nday-night temperature and IR \rux di\u000berences that can\nbe quantitatively compared to phase curve observations.\nSuch comparisons are currently the most detailed for HD\n189733b, HD 209458b, and WASP-43b because of the ex-\ntensive datasets available for these \\benchmark\" planets\n(Showman et al. 2009; Zellem et al. 2014; Kataria et al.arXiv:1601.00069v2 [astro-ph.EP] 18 Feb 20162 T.D. Komacek & A.P. Showman\n1000 1500 2000 2500\nEquilibrium Temperature (Kelvin)0.00.20.40.60.81.0Aobs=(Tb,day−Tb,night)/Tb,dayHD 189733b\nWASP-43b\nHD 209458b\nHD 149026b\nWASP-14b\nWASP-19b\nHAT-P-7b\nWASP-18b\nWASP-12b\nFig. 1.| Fractional dayside to nightside brightness temper-\nature di\u000berences Aobsvs. global-average equilibrium tempera-\nture from observations of transiting, low-eccentricity hot Jupiters.\nHere we de\fne the global-average equilibrium temperature, Teq=\n[F?=(4\u001b)]1=4, whereF?is the incoming stellar \rux to the planet\nand\u001bis the Stefan-Boltzmann constant. Solid points are from\nthe full-phase observations of Knutson et al. (2007, 2009b, 2012)\nfor HD 189733b, Cross\feld et al. (2012); Zellem et al. (2014) for\nHD 209458b, Knutson et al. (2009a) for HD 149026b, Wong et al.\n(2015a) for WASP-14b, Wong et al. (2015b) for WASP-19b and\nHAT-P-7b, and Cowan et al. (2012) for WASP-12b. The error bars\nfor WASP-43b (Stevenson et al. 2014) and WASP-18b (Nymeyer\net al. 2011; Maxted et al. 2013) show the lower limit on Aobsfrom\nthe nightside \rux upper limits (and hence fractional temperature\ndi\u000berence lower limits). See Appendix A for the data and method\nutilized to make this \fgure. There is a clear trend of increasing\nAobswith increasing equilibrium temperature, and hence dayside-\nnightside temperature di\u000berences at the photosphere are greater\nfor planets that receive more incident \rux.\n2015).\nDespite the proliferation of GCM investigations, our\nunderstanding of the underlying dynamical mechanisms\ncontrolling the day-night temperature di\u000berences of hot\nJupiters is still in its infancy. It is crucial to empha-\nsize that, in and of themselves, GCM simulations do\nnot automatically imply an understanding: the under-\nlying dynamics is often su\u000eciently complex that careful\ndiagnostics and a hierarchy of simpli\fed models are of-\nten necessary (e.g., Held 2005; Showman et al. 2010).\nThe ultimate goal is not simply matching observations\nbut also understanding physical mechanisms and con-\nstructing a predictive theory that can quantitatively ex-\nplain the day-night temperature di\u000berences, horizontal\nand vertical wind speeds, and other aspects of the circu-\nlation under speci\fed external forcing conditions. Taking\na step toward such a predictive theory is the primary goal\nof this paper.\nThe question of what controls the day-night temper-\nature di\u000berence in hot Jupiter atmospheres has been a\nsubject of intense interest for many years. Most stud-\nies have postulated that the day-night temperature dif-\nferences are controlled by a competition between radi-\nation and atmospheric dynamics|speci\fcally, the ten-\ndency of the strong dayside heating and nightside cool-\ning to create horizontal temperature di\u000berences, and the\ntendency of the atmospheric circulation to regulate those\ntemperature di\u000berences by transporting thermal energy\nfrom day to night. Describing this competition using a\ntimescale comparison, Showman & Guillot (2002) \frst\nsuggested that hot Jupiters would exhibit small frac-\ntional dayside-nightside temperature di\u000berences when\n\u001cadv\u001c\u001cradand large fractional dayside-nightside tem-perature di\u000berences when \u001cadv\u001d\u001crad. Here,\u001cadvis\nthe characteristic timescale for the circulation to advect\nair parcels horizontally over a hemisphere, and \u001cradis\nthe timescale over which radiation modi\fes the thermal\nstructure (e.g., the timescale to relax toward the local ra-\ndiative equilibrium temperature). Since then, numerous\nauthors have invoked this timescale comparison to de-\nscribe how the day-night temperature di\u000berences should\ndepend on pressure, atmospheric opacity, stellar irradia-\ntion, and other factors (e.g., Cooper & Showman 2005;\nShowman et al. 2008b, 2009; Fortney et al. 2008; Lewis\net al. 2010; Rauscher & Menou 2010; Cowan & Agol 2011;\nMenou 2012; Perna et al. 2012; Ginzburg & Sari 2015).\nOne would expect that hot Jupiters have short advec-\ntive timescales due to their fast zonal winds. This is\nobservationally evident from phase curves from tidally-\nlocked planets with \u001cadv\u0018\u001crad, which consistently show\na peak in brightness just before secondary eclipse. This\nindicates that the point of highest emitted \rux (\\hot\nspot\") is eastward of the substellar point (the point of\npeak absorbed \rux), due to downwind advection from a\nsuperrotating1equatorial jet (Showman & Polvani 2011).\nThe full-phase observations of HD 189733b (Knutson\net al. 2007), HD 209458b (Zellem et al. 2014), and\nWASP-43b (Stevenson et al. 2014) show hot spot o\u000bsets.\nThese o\u000bsets agree with those predicted from correspond-\ning circulation models (Showman et al. 2009; Kataria\net al. 2015). Hence, we have observational con\frmation\nthat hot Jupiters have fast ( \u0018kilometers/second) zonal\nwinds.\nAs discussed above, fast zonal winds are a robust fea-\nture of hot Jupiter general circulation models (GCMs),\nand notably also exist when the model hot Jupiter has\na large eccentricity (Lewis et al. 2010; Kataria et al.\n2013). However, these circulation models show a range\nof dayside-nightside temperature di\u000berences, showing no\nclear trend with wind speeds. Perna et al. (2012) ex-\namined how heat redistribution is a\u000bected by incident\nstellar \rux, showing that the nightside/dayside \rux ra-\ntio decreases with increasing incident stellar \rux. This\ntrend is akin to the observational trend shown in Fig-\nure 1. This trend was explained in Perna et al. (2012) by\ncalculating the ratio of \u001cadv=\u001crad, which increases with\nincident stellar \rux in their models.\nAs pointed out by Perez-Becker & Showman (2013),\nthere exist several issues with the idea that a timescale\ncomparison between \u001cadvand\u001cradgoverns the amplitude\nof the day-night temperature di\u000berences. First, although\nit is physically motivated, this timescale comparison has\nnever been derived rigorously from the equations of mo-\ntion; as such, it has always constituted an ad-hoc albeit\nplausible hypothesis, as opposed to a theoretical result.\nSecond, the comparison between \u001cadvand\u001craddoes not\ninclude any obvious role for other timescales that are\nimportant, including those for planetary rotation, hori-\nzontal and vertical wave propagation, and frictional drag\n(if any). These processes in\ruence the circulation and\nthus one might expect the timescale comparison to de-\npend on them. Third, the comparison is not predictive|\n\u001cadvdepends on the horizontal wind speeds, which are\n1Superrotation occurs where the zonal-mean atmospheric circu-\nlation of a planet has a greater angular momentum per unit mass\nthan the planet itself at the equator.Dayside-Nightside Temperature Di\u000berences in Hot Jupiter Atmospheres 3\nonly known a posteriori . Hence, it is only possible to\nevaluate the comparison between advective and radia-\ntive timescales if one already has a numerical model (or\ntheory) for the atmospheric circulation, which is neces-\nsarily related to other relevant timescales governing the\ncirculation.\nTo show how other timescales play a large role in de-\ntermining the atmospheric circulation, consider the equa-\ntorial regions of Earth. On Earth, horizontal tempera-\nture gradients in the tropics are weak, and the radia-\ntive cooling to space that occurs in Earth's troposphere\nis balanced primarily by vertical advection rather than\nhorizontal advection|a balance known as the weak tem-\nperature gradient (WTG) regime (Polvani & Sobel 2001;\nSobel et al. 2001; Sobel 2002). This vertical advection\ntimescale is related to the e\u000ecacy of lateral wave prop-\nagation. When gravity, Rossby, or Kelvin waves propa-\ngate, they induce vertical motion, which locally advects\nthe air parcels upward or downward. If these waves are\nable to propagate away, and rotation plays only a mod-\nest role, then this wave-adjustment tends to leave behind\na state with \rat isentropes|which is equivalent to eras-\ning the horizontal temperature di\u000berences (Bretherton\n& Smolarkiewicz 1989; Showman et al. 2013b). Given\nthat adjustment of isentropes due to propagating waves\nis known to occur on planetary scales on both Earth\n(Matsuno 1966; Gill 1980) and exoplanets (Showman &\nPolvani 2010, 2011; Tsai et al. 2014), Perez-Becker &\nShowman (2013) suggested that wave adjustment like-\nwise acts to lessen horizontal temperature di\u000berences in\nhot Jupiter atmospheres.\nIn hot Jupiter atmospheres, planetary-scale Kelvin and\nRossby waves are generated by the large gradient in\nradiative heating from dayside to nightside (Showman\n& Polvani 2011). Unlike in Earth's atmosphere, these\nwaves exhibit a meridional half-width stretching nearly\nfrom equator to pole, as the Rossby deformation radius\nis approximately equal to the planetary radius (Show-\nman & Guillot 2002). These waves cause horizontal con-\nvergence/divergence that forces vertical motion through\nmass continuity. This vertical motion moves isentropes\nvertically, and if the waves are not damped this leads to a\n\fnal state with \rat isentropes. However, if these Kelvin\nand Rossby waves cannot propagate (i.e. are damped),\nthen this process cannot occur. Hence, the ability of\nwave adjustment processes to lessen horizontal temper-\nature gradients can be weakened by damping of propa-\ngating waves. As shown in Figure 1, damping processes\nthat increase day-night temperature di\u000berences seem to\nincrease in e\u000ecacy with increasing equilibrium temper-\nature. The most natural damping process is radiative\ncooling, which should increase in e\u000eciency with the cube\nof equilibrium temperature (Showman & Guillot 2002).\nAdditionally, frictional drag on the atmosphere can re-\nduce the ability of wave adjustment to reduce longitudi-\nnal temperature gradients. This drag could either be due\nto turbulence (Li & Goodman 2010; Youdin & Mitchell\n2010) or the Lorentz force in a partially ionized atmo-\nsphere threaded by a dipole magnetic \feld (Batygin &\nStevenson 2010; Perna et al. 2010; Menou 2012; Batygin\net al. 2013; Rauscher & Menou 2013; Rogers & Showman\n2014; Rogers & Komacek 2014). Both of these processes\nshould increase fractional day-night temperature di\u000ber-\nences with increasing equilibrium temperature, helpingexplain Figure 1. However, it is not obvious a priori\nwhether radiative e\u000bects or drag should more e\u000eciently\ndamp wave adjustment in hot Jupiter atmospheres.\nTo understand the mechanisms controlling day-night\ntemperature di\u000berences|including their dependence on\nradiative and frictional e\u000bects|Perez-Becker & Show-\nman (2013) introduced a shallow-water model with a\nsingle active layer representing the atmosphere, which\noverlies a deeper layer, representing the interior, whose\ndynamics are \fxed and assumed to be quiescent. Perez-\nBecker & Showman (2013) performed numerical simula-\ntions over a broad range of drag and radiative timescales,\nwhich generally showed that strong radiation and fric-\ntional drag tend to promote larger day-night temperature\ndi\u000berences. They then compared their model results to\nderived analytic theory and found good agreement be-\ntween the expected fractional dayside-nightside temper-\nature di\u000berences and model results, albeit with minor ef-\nfects not captured in the theory. This theory showed rig-\norously that wave adjustment allows for reduced horizon-\ntal temperature di\u000berences in hot Jupiter atmospheres.\nThey also showed that horizontal wave propagation is\nmainly damped by radiative e\u000bects, with potential drag\nplaying a secondary, but crucial, role. As a result, they\nfound that the strength of radiative heating/cooling is\nthe main governor of dayside-nightside temperature dif-\nferences in hot Jupiter atmospheres. Nevertheless, be-\ncause the model is essentially two-dimensional (in longi-\ntude and latitude) and lacks a vertical coordinate, ques-\ntions exist about how the results would carry over to a\nfully three-dimensional atmosphere.\nHere we extend the work of Perez-Becker & Show-\nman (2013) to fully three-dimensional atmospheres. The\nuse of the full three-dimensional primitive equations en-\nables us to present a predictive analytic understanding\nof dayside-nightside temperature di\u000berences and wind\nspeeds that can be directly compared to observable quan-\ntities. Our analytic theory is accompanied by numerical\nmodels which span a greater range of radiative forcing\nand drag parameter space than Perez-Becker & Showman\n(2013), enabling quantitative validation of these analytic\nresults. We keep the radiative forcing simple in order to\npromote a physical understanding. Despite this simpli\f-\ncation, we emphasize that this is the \frst fully predictive\nanalytic theory for the day-night temperature di\u000berences\nof hot Jupiter atmospheres in three dimensions.\nThis paper is organized as follows. In Section 2, we de-\nscribe our methods, model setup, and parameter space\nexplored. We discuss the results of our numerical param-\neter study of dayside-nightside temperature di\u000berences in\nSection 3. In Section 4, we develop our theory in order\nto facilitate a comparison to numerical results in Sec-\ntion 5. In Section 6, we explore the implications of our\nmodel results in the context of previous observations and\ntheoretical work, and express conclusions in Section 7.\n2.MODEL\nWe adopt the same physical model for both our nu-\nmerical and analytic solutions. GCMs with accurate ra-\ndiative transfer have proven essential for detailed com-\nparison with observations (Showman et al. 2009, 2013a;\nKataria et al. 2015); however, our goal here is to promote\nanalytic tractability and a clean environment in which to\nunderstand dynamical mechanisms, and so we drive the4 T.D. Komacek & A.P. Showman\ncirculation using a simpli\fed Newtonian heating/cooling\nscheme (e.g. Showman & Guillot 2002; Cooper & Show-\nman 2005). This enables us to systematically vary the\ndayside-nightside thermal forcing and control the rate at\nwhich temperature relaxes to a \fxed radiative equilib-\nrium pro\fle. We also incorporate a drag term in the\nequations to investigate how day-night temperature dif-\nferences are modi\fed by the combined e\u000bects of atmo-\nspheric friction and di\u000berential stellar irradiation. Our\nmodel setup is nearly identical to that in Liu & Showman\n(2013).\n2.1. Dynamical equations\nWe solve the horizontal momentum, vertical momen-\ntum, continuity, energy equation, and ideal gas equation\nof state (i.e. the hydrostatic primitive equations), which,\nin pressure coordinates, are:\ndv\ndt+f^k\u0002v+r\b =Fdrag+DS; (1)\n@\b\n@p+1\n\u001a= 0; (2)\nr\u0001v+@!\n@p= 0; (3)\nTd(ln\u0012)\ndt=dT\ndt\u0000!\n\u001acp=q\ncp+ES; (4)\np=\u001aRT: (5)\nWe use the following symbols: pressure p, density\u001a, tem-\nperatureT, speci\fc heat at constant pressure cp, spe-\nci\fc gas constant R, potential temperature2\u0012, horizon-\ntal velocity (on isobars) v, horizontal gradient on isobars\nr, vertical velocity in pressure coordinates !=dp=dt ,\ngeopotential \b = gz, Coriolis parameter f= 2\nsin\u001e\n(with \n planetary rotation rate, here equivalent to or-\nbital angular frequency, and \u001elatitude), and speci\fc\nheating rate q. In this coordinate system, the total\n(material) derivative is d=dt =@=@t +v\u0001r+!@=@p .\nFdragrepresents a drag term that we use to represent\nmissing physics (Rauscher & Menou 2012b), for exam-\nple drag due to turbulent mixing (Li & Goodman 2010;\nYoudin & Mitchell 2010), or the Lorentz force (Perna\net al. 2010; Rauscher & Menou 2013). The terms DS\nandESrepresent a standard fourth-order Shapiro \flter,\nwhich smooths grid-scale variations while minimally af-\nfecting the \row at larger scales, and thereby helps to\nmaintain numerical stability in our numerical integra-\ntions. Because the Shapiro \flter terms do not a\u000bect the\nglobal structure of our equilibrated numerical solutions,\nthey are negligible in comparison to the other terms in\nthe equations at the near-global scales captured in our\nanalytic theory. As a result, we neglect them from our\nanalytic solutions.\n2Potential temperature is de\fned as \u0012=T(p=p0)\u0014, where\n\u0014=R=cp, which is here assumed constant. The potential temper-\nature is the temperature that an air parcel would have if brought\nadiabatically to an atmospheric reference pressure p0. We choose\na reference pressure of p0= 1 bar, but the solution is independent\nof the value of p0chosen.2.2. Thermal forcing and frictional drag\nWe represent the radiative heating and cooling using\na Newtonian heating/cooling scheme, which relaxes the\ntemperature toward a prescribed radiative equilibrium\ntemperature, Teq, over a speci\fed radiative timescale\n\u001crad:\nq\ncp=Teq(\u0015;\u001e;p )\u0000T(\u0015;\u001e;p;t )\n\u001crad(p): (6)\nIn this scheme, Teqdepends on longitude \u0015, latitude\u001e,\nand pressure, while, for simplicity, \u001cradvaries only with\npressure. The radiative equilibrium pro\fle is set to be\nhot on the dayside and cold on the nightside:\nTeq(\u0015;\u001e;p ) =\u001aTnight;eq(p) + \u0001Teq(p)cos\u0015cos\u001edayside;\nTnight;eq(p) nightside :\n(7)\nHere\u0015is longitude, Tnight;eq(p) is the radiative\nequilibrium temperature pro\fle on the nightside and\nTnight;eq(p) + \u0001Teq(p) is that at the substellar point. To\nacquire the nightside heating pro\fle, Tnight;eq, we take\nthe temperature pro\fle of HD 209458b from Iro et al.\n(2005) and subtract our chosen \u0001 Teq(p)=2. We specify\n\u0001Teqas in Liu & Showman (2013), setting it to be a\nconstant \u0001 Teq;topat pressures less than peq;top, zero at\npressures greater than pbot, and varying linearly with log\npressure in between:\n\u0001Teq(p) =8\n>><\n>>:\u0001Teq;top p

p bot:\n(8)\nAs in Liu & Showman (2013), we assume peq;top= 10\u00003\nbars andpbot= 10 bars. However, we vary \u0001 Teq;top\nfrom 1000\u00000:001 Kelvin, ranging from highly nonlinear\nto linear numerical solutions.\nRadiative transfer calculations show that the radiative\ntime constant is long at depth and short aloft (Iro et al.\n2005; Showman et al. 2008a). To capture this behav-\nior, we adopt the same functional form for \u001cradas Liu &\nShowman (2013): \u001cradis set to a large constant \u001crad;bot\nat pressures greater than pbot, a (generally smaller) con-\nstant\u001crad;topat pressures less than prad;top, and varies\ncontinuously in between:\n\u001crad(p) =8\n>><\n>>:\u001crad;top p

p bot;\n(9)\nwith\n\u000b=ln(\u001crad;top=\u001crad;bot)\nln(prad;top=pbot): (10)\nHere, as in Liu & Showman (2013), we set prad;top=\n10\u00002bars andprad;bot= 10 bars. Note that the pres-\nsures above which \u0001 Teqand\u001cradare \fxed to a constant\nvalue at the top of the domain are di\u000berent, in order\nto be fully consistent with the model setup of Liu &\nShowman (2013). The model is set up such that the cir-\nculation forced by Newtonian heating/cooling has three-\ndimensional temperature and wind distributions that areDayside-Nightside Temperature Di\u000berences in Hot Jupiter Atmospheres 5\n103\n104\n105\n106\n107\nτrad (sec)10-4\n10-3\n10-2\n10-1\n100\n101\n102Pressure (bar)\n103104105106107108\nτdrag (sec)10-810-710-610-510-410-3kv (sec−1)\nFig. 2.| Radiative forcing and drag pro\fles used in the numerical model. Left: Radiative timescale vs. pressure, for all assumed\n\u001crad;top= 103\u0000107sec. Right: Drag timescale as a function of pressure, for each of assumed spatially constant \u001cdrag= 103\u00001 sec. The\ndrag constant kv=\u001c\u00001\ndragis shown on the upper x-axis.\nsimilar to results from simulations driven by radiative\ntransfer. This motivated the choices of prad;topand\npeq;top, along with the values of other \fxed parameters\nin the model.\nThe various \u001crad-pressure pro\fles used in our models\n(for di\u000berent assumed \u001crad;top) are shown on the left hand\nside of Figure 2. We choose \u001crad;bot= 107sec, which\nis long compared to relevant dynamical and rotational\ntimescales but short enough to allow us to readily inte-\ngrate to equilibrium. For the purposes of our study, we\nvary\u001crad;topfrom 103\u0000107sec, corresponding to a range\nof radiative forcing.\nWe introduce a linear drag in the horizontal momen-\ntum equation, given by\nFdrag=\u0000kv(p)v; (11)\nwherekv(p) is a pressure-dependent drag coe\u000ecient.\nThis drag has two components:\n\u000fFirst, we wish to examine how forces that crudely\nparameterize Lorentz forces a\u000bect the day-night\ntemperature di\u000berences. This could be represented\nwith a drag coe\u000ecient that depends on longitude,\nlatitude, and pressure and, moreover, di\u000bers in all\nthree dimensions (e.g. Perna et al. 2010; Rauscher\n& Menou 2013). However, the Lorentz force should\ndepend strongly on the ionization fraction and\ntherefore the local temperature, requiring full nu-\nmerical magnetohydrodynamic solutions to exam-\nine the e\u000bects of magnetic \\drag\" in detail, e.g.\nBatygin et al. (2013); Rogers & Komacek (2014);\nRogers & Showman (2014). For simplicity and\nanalytic tractability, we represent this component\nwith a spatially constant drag timescale \u001cdrag, cor-responding to a drag coe\u000ecient \u001c\u00001\ndrag. We system-\natically explore \u001cdrag values (in sec) of 103, 104,\n105, 106, 107, and1. The latter corresponds to\nthe drag-free limit. Such a scheme was already ex-\nplored by Showman et al. (2013a).\n\u000fSecond, following Liu & Showman (2013), we in-\ntroduce a \\basal\" drag at the bottom of the do-\nmain, which crudely parameterizes interactions be-\ntween the vigorous atmospheric circulation and\na relatively quiescent planetary interior. For\nthis component, the drag coe\u000ecient is zero at\npressures less than pdrag;topand is\u001c\u00001\ndrag;bot(p\u0000\npdrag;top)=(pdrag;bot\u0000pdrag;top) at pressures greater\nthanpdrag;top, wherepdrag;botis the mean pres-\nsure at the bottom of the domain (200 bars) and\npdrag;topis the lowest pressure where this basal\ndrag component is applied. Thus, the drag coef-\n\fcient varies from \u001c\u00001\ndrag;botat the bottom of the\ndomain to zero at a pressure pdrag;top; this scheme\nis similar to that in Held & Suarez (1994). We\ntakepdrag;top= 10 bars, and set \u001cdrag;bot= 10 days.\nThus, basal drag acts only at pressures greater than\n10 bars and has a minimum characteristic timescale\nof 10 days at the bottom of the domain, increas-\ning to in\fnity (meaning zero drag) at pressures less\nthan 10 bars. We emphasize that the precise value\nis not critical. Changing the drag time constant at\nthe base to 100 days, for example (corresponding\nto weaker drag) would lead to slightly faster wind\nspeeds at the base of the model, and would require\nlonger integration times to reach equilibrium, but\nwould not qualitatively change our results.6 T.D. Komacek & A.P. Showman\nTo combine the two drag schemes, we simply set the drag\ncoe\u000ecient to be the smaller of the two individual drag\ncoe\u000ecients at each individual pressure, leading to a \fnal\nfunctional form for the drag coe\u000ecient:\nkv(p) = max\u0014\n\u001c\u00001\ndrag;\u001c\u00001\ndrag;bot(p\u0000pdrag;top)\n(pdrag;bot\u0000pdrag;top)\u0015\n(12)\nThe righthand side of Figure 2 shows the various \u001cdrag(p)\npro\fles used in our models. Corresponding values for\nkv(p) are given along the top axis.\nWe adopt planetary parameters ( cp,R, \n,g,R)\nrelevant for HD 209458b. This includes speci\fc heat\ncp= 1:3\u0002104J kg\u00001K\u00001, speci\fc gas constant R=\n3700 J kg\u00001K\u00001, rotation rate \n = 2 :078\u000210\u00005s\u00001,\ngravityg= 9:36 m s\u00002, and planetary radius a= 9:437\u0002\n107m. Though we use parameters relevant for a given\nhot Jupiter, our results are not sensitive to the precise\nvalues used. Moreover, we emphasize that the qual-\nitative model behavior should not be overly sensitive\nto numerical parameters such as the precise values of\nprad;top,peq;top,pbot, and so on. Modifying the values of\nthese parameters over some reasonable range will change\nthe precise details of the height-dependence of the day-\nnight temperature di\u000berence and wind speeds but will\nnot change the qualitative behavior or the dynamical\nmechanisms we seek to uncover. We would \fnd simi-\nlar behavior regardless of the speci\fc parameters used,\nas long as they are appropriate for a typical hot Jupiter.\n2.3. Numerical details\nOur numerical integrations are performed using the\nMITgcm (Adcroft et al. 2004) to solve the equations\ndescribed above on a cubed-sphere grid. The horizon-\ntal resolution is C32, which is roughly equal to a global\nresolution of 128 \u000264 in longitude and latitude. There\nare 40 vertical levels, with the bottom 39 levels evenly\nspaced in log-pressure between 0.2 mbars and 200 bars,\nand a top layer that extends from 0.2 mbars to zero pres-\nsure. Models performed at resolutions as high as C128\n(corresponding to a global resolution of 512 \u0002256) by\nLiu & Showman (2013) behave very similar to their C32\ncounterparts, indicating that C32 is su\u000ecient for current\npurposes. All models are integrated to statistical equi-\nlibrium. We integrate the model from a state of rest with\nthe temperatures set to the Iro et al. (2005) temperature-\npressure pro\fle. Note that this system does not exhibit\nsensitivity to initial conditions (Liu & Showman 2013).\nFor the most weakly nonlinear runs described in Sec-\ntion 3.1, reaching equilibration required 25 ;000 Earth\ndays of model integration time. However, our full grid of\nsimulations varying radiative and drag timescales with\na \fxed equilibrium day-night temperature di\u000berence re-\nquired .5;000 days of integration.\n3.NUMERICAL RESULTS\n3.1. Parameter space exploration\nGiven the forcing and drag prescriptions speci\fed in\nSection 2 and the planetary parameters for a typical hot\nJupiter, the problem we investigate is one governed by\nthree parameters|\u0001 Teq;top,\u001crad;top, and\u001cdrag. Our goal\nis to thoroughly explore a broad, two-dimensional grid of\nGCM simulations varying \u001crad;topand\u001cdragover a wide\n10−310−210−110010110210310−410−310−210−1100101102103104\n∆ Teq,top (K)Urms (m/s)\n \nτdrag = 105 secτdrag = basalFig. 3.| Root-mean-square (RMS) horizontal wind speed at a\npressure of 80 mbars plotted against equilibrium dayside-nightside\ntemperature di\u000berences \u0001 Teq;top. These day-night temperature\ndi\u000berences set the forcing amplitude of the circulation. When\n\u0001Teq;topis small, the RMS velocities respond linearly to forcing,\nand when \u0001 Teq;topis large, the RMS velocities respond nonlin-\nearly. The transition from nonlinear to linear response occurs at\ndi\u000berent \u0001Teq;topdepending on whether or not spatially constant\ndrag is applied. When there is not spatially constant drag (blue\nopen circles) and only basal drag is applied, the transition occurs at\n\u0001Teq;top\u00180:1 Kelvin. When \u001cdrag = 105sec (red \flled circles),\nthe transition occurs at \u0001 Teq;top\u001810 Kelvin. For comparison,\na linear relationship between Urmsand \u0001Teq;topis shown by the\nblack line.\nrange. Here we \frst explore the role of \u0001 Teqso that we\nmay make appropriate choices about the values of \u0001 Teq\nto use in the full grid.\nThe day-night radiative-equilibrium temperature con-\ntrast \u0001Teqrepresents the amplitude of the imposed ra-\ndiative forcing, and controls the amplitude of the result-\ning \row. In the low-amplitude limit (\u0001 Teq!0), the\nwind speeds and temperature perturbations are weak,\nand the nonlinear terms in the dynamical equations\nshould become small compared to the linear terms. Thus,\nin this limit, the solutions should behave in a mathemat-\nically linear3manner: the spatial structure of the circu-\nlation should become independent of forcing amplitude,\nand the amplitude of the circulation|that is, the wind\nspeeds and day-night temperature di\u000berences|should\nvary linearly with forcing amplitude. On the other hand,\nat very high forcing amplitudes (large \u0001 Teq), the wind\nspeeds and temperature di\u000berences are large, and the so-\nlutions behave nonlinearly.\nTherefore, we \frst performed a parameter sweep of\n\u0001Teqto understand the transition between linear and\nnonlinear forcing regimes, and to determine the value of\n\u0001Teqat which this transition occurs. For this param-\neter sweep, we performed a sequence of models varying\n\u0001Teq;topfrom 1000 to 0 :001 Kelvin4. We did one such\nsweep using \u001crad;top= 104s and\u001cdrag =1(meaning\nbasal-drag only), and another such sweep using \u001crad;top=\n3By this we mean that as \u0001 Teq!0, the solutions of the full\nnonlinear problem should converge toward the mathematical solu-\ntions to versions of Equations (1){(4) that are linearized around a\nstate with zero wind and the background T(p) pro\fle.\n4Speci\fcally, we tested values of \u0001 Teq;top =\n1000;500;200;100;10;1;0:1;0:01;and 0:001K.Dayside-Nightside Temperature Di\u000berences in Hot Jupiter Atmospheres 7\n⌧rad,top(sec)\u0000Teq= 1000KBasal103104105106107\n \nTemperature (K)9001000110012001300140015001600107106105104103\nlongitude (degrees)\nlatitude (degrees)\n−150−100−50050100150−80−60−40−20020406080Basal103104105106107107106105104103longitude (degrees)\nlatitude (degrees)−90090−45045\nFig. 4.| Maps of temperature (colors) and wind (vectors) for suite of 30 GCM simulations varying \u001crad;topand\u001cdragwith \u0001Teq;top= 1000\nKelvin. All maps are taken from the 80-mbar statistical steady-state end point of an individual model run. All plots share a color scheme\nfor temperature but have independent overplotting of horizontal wind vectors. The substellar point is located at 0\u000e;0\u000ein each plot, with\nthe lower right plot displaying latitude & longitude axes.\n104s and\u001cdrag= 105s. These sweeps verify that we are\nindeed in the linear limit (where variables such as wind\nspeed and temperature respond linearly to forcing) at\n\u0001Teq;top.0:1 Kelvin for any \u001cdrag. Figure 3 shows\nhow the root-mean-square (RMS) horizontal wind speed\nvaries with \u0001 Teq;topfor these two parameter sweeps.\nHere, the RMS horizontal wind speed, Urms, is de\fned\nat a given pressure as:\nUrms(p) =rR\n(u2+v2)dA\nA: (13)\nHere the integral is taken over the globe, with Athe hori-\nzontal area of the globe and u;vthe zonal and meridional\nvelocities at a given pressure level, respectively. It is no-table that the linear limit is reached at di\u000berent \u0001 Teq;top\nvalues depending on the strength of the drag applied.\nWith a spatially constant \u001cdrag= 105sec, the models\nrespond linearly to forcing at \u0001 Teq;top.10 Kelvin, and\nare very nearly linear throughout the range of \u0001 Teq;top\nconsidered. This causes the Urms-\u0001Teq;toprelationship\nfor\u001cdrag= 105sec to be visually indistinguishable from\na linear slope in Figure 3. However, without a spa-\ntially constant drag, the linear limit is not reached until\n\u0001Teq;top.0:1 Kelvin, and the dynamics are nonlinear\nfor \u0001Teq;top&1K.\nThe main grid presented involves a parameter study\nmimicking that of Perez-Becker & Showman (2013), but\nusing the 3D primitive equations. We do so in order\nto understand mechanisms behind hot Jupiter dayside-8 T.D. Komacek & A.P. Showman\n\u0000Teq=0.001KBasal103104105106107\n \nTemperature (K)1285.63541285.63551285.63561285.63571285.63581285.63591285.636107106105104103\nlongitude (degrees)\nlatitude (degrees)\n−150−100−50050100150−80−60−40−20020406080\nBasal103104105106107107106105104103longitude (degrees)\nlatitude (degrees)−90090−45045⌧rad,top(sec)\nFig. 5.| Same as Figure 4, except with \u0001 Teq;top= 0:001 Kelvin.\nnightside temperature di\u000berences with the full system of\nnonlinear primitive equations. We varied \u001crad;topfrom\n103\u0000107sec and\u001cdrag from 103\u00001 sec (the range\nof timescales displayed in Figure 2), extending an or-\nder of magnitude lower in \u001cdrag than Perez-Becker &\nShowman (2013). We ran this suite of models for both\n\u0001Teq;top= 1000 Kelvin (nonlinear regime, see Figure\n4) and \u0001Teq;top= 0:001 Kelvin (linear regime, see Fig-\nure 5) to better understand the mechanisms controlling\ndayside-nightside temperature di\u000berences.\n3.2. Description of atmospheric circulation over a wide\nrange of radiative and frictional timescales\nFigure 4 shows latitude-longitude maps of tempera-\nture (with overplotted wind vectors) at a pressure of\n80 mbars for the entire suite of models performed at\n\u0001Teq;top= 1000 Kelvin. These are the statisticallysteady-state end points of 30 separate model simulations,\nwith\u001crad;topvarying from 103\u0000107sec and\u001cdragranging\nfrom 103\u00001 sec. All plots have the same temperature\ncolorscale for inter-comparison.\nOne can identify distinct regimes in this \u001crad;topand\n\u001cdragspace. First, there is the nominal hot Jupiter regime\nwith\u001cdrag=1and\u001crad;top.105sec, which has been\nstudied extensively in previous work. This regime has a\nstrong zonal jet which manifests as equatorial superrota-\ntion. However, there is no zonal jet when \u001cdrag.105sec.\nHence, there is a regime transition in the atmospheric\ncirculation at \u001cdrag\u0018106sec between a strong coher-\nent zonal jet and weak or absent zonal jets. Addition-\nally, there are two separate regimes of dayside-nightside\ntemperature di\u000berences as \u001crad;topvaries. When \u001crad;top\nis short (103\u0000104sec), the dayside-nightside temper-Dayside-Nightside Temperature Di\u000berences in Hot Jupiter Atmospheres 9\nature di\u000berences are large. When \u001crad;top&106sec,\ndayside-nightside temperature di\u000berences are small on\nlatitude circles. However, \u001crad;topis not the only control\non dayside-nightside temperature di\u000berences. If atmo-\nspheric friction is strong, with \u001cdrag.104sec, dayside-\nnightside temperature di\u000berences are large unless \u001crad;top\nis extremely long.\nEquivalent model inter-comparison to Figure 4 for the\n\u0001Teq;top= 0:001 Kelvin case is presented in Figure 5.\nThe same major trends apparent in the \u0001 Teq;top=\n1000 Kelvin results are seen here in the linear limit. That\nis,\u001crad;topis still the key parameter control on dayside-\nnightside temperature di\u000berences. If \u001crad;topis small,\ndayside-nightside temperature di\u000berences are large, and\nif\u001crad;topis large, dayside-nightside di\u000berences are small.\nThis general trend is modi\fed slightly by atmospheric\nfriction|if \u001cdrag.104sec, drag plays a role in deter-\nmining the dayside-nightside temperature di\u000berences.\nA key di\u000berence between simulations at high and low\n\u0001Teqis that, under conditions of short \u001cradand weak\ndrag (upper-left quadrant of Figures 4{5), the maximum\ntemperatures occur on the equator in the nonlinear limit,\nbut they occur in midlatitudes in the linear limit. This\ncan be understood by considering the force balances.\nNamely, because the Coriolis force goes to zero at the\nequator, the only force that can balance the pressure\ngradient at the equator is advection. Advection is a non-\nlinear term that scales with the square of wind speed,\nand hence this force balance is inherently nonlinear. As\na result, the advection term and pressure gradient force\nboth weaken drastically at the equator in the linear limit,\nand the nonlinear balance cannot hold. This causes the\nequator to be nearly longitudinally isothermal in the lin-\near limit, rather than having large day-night temperature\ndi\u000berences as in the nonlinear case. This phenomenon\nwas already noted by Showman & Polvani (2011) and\nPerez-Becker & Showman (2013) in one-layer shallow wa-\nter models, and Figure 5 represents an extension of it to\nthe 3D system.\nAnother di\u000berence in the temperature maps between\nthe nonlinear and linear limit is the orientation of the\nphase tilts in the long \u001cdrag, short\u001cradupper left quad-\nrant of Figure 5. These phase tilts are the exact op-\nposite of those that are needed to drive superrotation.\nThe linear dynamics in the weak-drag limit causing these\ntilts has been examined in detail by Showman & Polvani\n(2011), see their Appendix C. They showed analytically\nthat in the limit of long \u001cdrag, the standing Rossby waves\ndevelop phase tilts at low latitudes that are northeast-\nto-southwest in the northern hemisphere and southeast-\nto-northwest in the southern hemisphere. This is the\nopposite of the orientation needed to transport eastward\nmomentum to equatorial regions and drive equatorial su-\nperrotation. Hence, the upper left quadrant of Figure 5\nshows key distinctions from the same quadrant in Fig-\nure 4. In the linear limit, there is no superrotation, and\nthe ratio of characteristic dayside and nightside temper-\natures at the equator is much smaller than in the case\nwith large forcing amplitude.\nOur model grids in Figures 4 and 5 exhibit a striking re-\nsemblance to the equivalent grids from the shallow-water\nmodels of Perez-Becker & Showman (2013, see their Fig-\nures 3 and 4). This gives us con\fdence that the same\nmechanisms determining the day-night temperature dif-ferences in their one-layer models are at work in the full\n3D system. Additionally, by extending \u001cdragone order\nof magnitude shorter, we have reached the parameter\nregime where drag can cause increased day-night tem-\nperature di\u000berences even when \u001cradis extremely long, al-\nlowing a more robust comparison to theory in this limit.\nDespite the distinctions between the nonlinear and lin-\near limits discussed above, both grids show similar over-\nall parameter dependences on radiative forcing and fric-\ntional drag. As radiative forcing becomes stronger (i.e.,\n\u001crad;topbecomes shorter), day-night temperature di\u000ber-\nences increase, with drag only playing a role if it is ex-\ntremely strong. Additionally, drag is the key factor to\nquell the zonal jet. The fact that the same general trends\nin day-night temperature di\u000berences occur in both the\nnonlinear and linear limit suggests that the same quali-\ntative mechanisms are controlling day-night temperature\ndi\u000berences in both cases, although nonlinearities will of\ncourse introduce quantitative di\u000berences at su\u000eciently\nlarge \u0001Teq. This makes it likely that a simple analytic\ntheory can explain the trends seen in Figures 4 and 5. We\ndevelop such a theory in Section 4, and continue in Sec-\ntion 5 to compare our results to the numerical solutions\npresented in this section.\n4.THEORY\n4.1. Pressure-dependent theory\nWe seek approximate analytic solutions to the problem\nposed in Sections 2{3. Speci\fcally, here we present solu-\ntions for the pressure-dependent day-night temperature\ndi\u000berence and the characteristic horizontal and vertical\nwind speeds as a function of the external control param-\neters (\u0001Teq;top,\u001crad;top,\u001cdrag, and the planetary param-\neters). In this theory, we do not distinguish variations\nin longitude from variations in latitude. As a result, we\nassume that the day-to-night and equator-to-pole tem-\nperature di\u000berences are comparable. Most GCM studies\nproduce relatively steady hemispheric-mean circulation\npatterns (e.g., Showman et al. 2009; Liu & Showman\n2013), including our own simulations shown in Section\n3, and so we seek steady solutions to the primitive equa-\ntions (1){(5). For convenience, we here cast these in\nlog-pressure coordinates (Andrews et al. 1987; Holton &\nHakim 2013):\nv\u0001rv+w?@v\n@z?+fk\u0002v=\u0000r\b\u0000v\n\u001cdrag; (14)\n@\b\n@z?=RT; (15)\nr\u0001v+ez?@(e\u0000z?w?)\n@z?= 0; (16)\nv\u0001rT+w?N2H2\nR=Teq\u0000T\n\u001crad: (17)\nEquation (14) is the horizontal momentum equation,\nEquation (15) the vertical momentum equation (hydro-\nstatic balance), Equation (16) the continuity equation,\nand Equation (17) the thermodynamic energy equation.\nIn this coordinate system, z?is de\fned as\nz?\u0011\u0000lnp\np00; (18)10 T.D. Komacek & A.P. Showman\nwithp00a reference pressure, and the vertical velocity\nw?\u0011dz?=dt, which has units of scale heights per sec-\nond (such that 1 =w?is the time needed for air to \row\nvertically over a scale height). Nis the Brunt-Vaisala fre-\nquency and H=RT=g is the scale height. This equation\nset is equivalent to the steady-state version of Equations\n(1-5). Note that drag is explicitly set on the horizontal\ncomponents of velocity. We use the steady-state system\nhere in order to facilitate comparison with our models,\nwhich themselves are run to steady-state with kinetic en-\nergy equilibration.\nGiven the set of Equations (14){(17) above, we can now\nutilize scaling to give approximate solutions for compar-\nison to both our fully nonlinear (high \u0001 Teq) and linear\n(low \u0001Teq) numerical solutions. Here, we step systemat-\nically through the equations, starting with the continuity\nequation, then the thermodynamic energy equation, and\n\fnally the momentum equations.\n4.1.1. Continuity equation\nFirst, consider the continuity equation (16). The scal-\ning for the \frst term on the left hand side is subtle.\nWhen the Rossby number Ro &1, we expect that\nr\u0001v\u0018U=L, whereUis a characteristic horizontal ve-\nlocity andLa characteristic lengthscale of the circula-\ntion. However, when Ro .1, geostrophy holds and in\nprinciple we could have r\u0001v\u001c U=L. For a purely\ngeostrophic \row, r\u0001v\u0011\u0000\fv=f , withvmeridional ve-\nlocity (see Showman et al. 2010, Eq. 32). On a sphere,\n\f= 2\ncos\u001e=a, and hence \f=f = cot\u001e=a. As a result,\nr\u0001v\u0019Ucot\u001e=a. For hot Jupiters, L\u0018a, and hence it\nturns out that for geostrophic \row not too close to the\npole thatr\u0001v\u0018U=L. Hence, the scaling r\u0001v\u0018U=L\nholds throughout the circulation regimes considered here.\nNow consider the second term on the left side of (16).\nExpanding out the derivative yields \u0000w?+@w?=@z?. The\nterm@w?=@z?scales asw?=\u0001z?, where \u0001z?is the ver-\ntical distance (in scale heights) over which w?varies by\nits own magnitude. Thus, one could write\nU\nL\u0018max\u0014\nw?;w?\n\u0001z?\u0015\n: (19)\nPrevious GCM studies suggest that w?maintains co-\nherency of values over several scale heights (e.g. Par-\nmentier et al. 2013), suggesting that \u0001 z?is several (i.e.,\ngreater than one), in which case the \frst term on the\nright dominates. De\fning an alternate characteristic ver-\ntical velocityW=Hw?, which gives the approximate\nvertical velocity in m s\u00001, the continuity equation be-\ncomes simply\nU\nL\u0018W\nH: (20)\n4.1.2. Thermodynamic energy equation\nWe next consider the thermodynamic energy equation,\nwhich contains the sole term that drives the circulation\n(i.e., radiative heating and cooling). The quantity Teq\u0000T\non the rightmost side of Equation (17) represents the\nlocal di\u000berence between the radiative-equilibrium and\nactual temperature. This di\u000berence varies spatially in\nvalue and sign, as it is typically positive on the dayside\nand negative on the nightside. Here we seek an expres-\nsion for its characteristic magnitude. We note that if\nLongitude \nTemperature Day Night Night \n∆T \n∆Teq \n|Teq-T|night |Teq-T|day \nActual temperature Radiative equilibrium temperature Fig. 6.| Simpli\fed diagram of the model, schematically display-\ning longitudinal pro\fles of actual and radiative equilibrium temper-\nature. We show this schematic to help explain Equation (22). The\ndi\u000berence between the characteristic actual and radiative equilib-\nrium temperature di\u000berences from dayside to nightside, \u0001 Teq\u0000\u0001T,\nis approximately equal to the sum of the characteristic di\u000berences\non each hemisphere, jTeq\u0000Tjday+jTeq\u0000Tjnight.\njTeq\u0000Tjglobalis de\fned as\njTeq\u0000Tjglobal\u0011jTeq\u0000Tjday+jTeq\u0000Tjnight;(21)\nwhere the di\u000berences on the right hand side are charac-\nteristic di\u000berences for the appropriate hemisphere, one\ncan write\n\u0001Teq\u0000\u0001T\u0018jTeq\u0000Tjglobal: (22)\nIn Equation (22), \u0001 Tand \u0001Teqare de\fned to be the\ncharacteristic di\u000berence between the dayside and night-\nside temperature and equilibrium temperature pro\fles,\nrespectively. Figure 6 shows visually this approximate\nequality between \u0001 Teq\u0000\u0001TandjTeq\u0000Tjglobal.\nWith this formalism for characteristic dayside-\nnightside temperature di\u000berences, we can write an ap-\nproximate version of Equation (17) as:\n\u0001Teq\u0000\u0001T\n\u001crad\u0018max\u0014U\u0001T\nL;WN2H\nR\u0015\n: (23)\nThe quantitiesU,W, \u0001T,H, andNare implicitly func-\ntions of pressure. For the analyses that follow, we use a\nvalue ofLapproximately equal to the planetary radius.\nWhat is the relative importance of the two terms on\nthe right-hand side of (23)? Using Equation (20) and the\nde\fnition of Brunt-Vaisala frequency, the second term\ncan be expressed as U=Ltimes\u000eTstrat, where\u000eTstrat is\nthe di\u000berence between the actual and adiabatic tempera-\nture gradients integrated vertically over a scale height|\nor, equivalently, can approximately be thought of as\nthe change in potential temperature over a scale height.\nThus, the \frst term (horizontal entropy advection) dom-\ninates over the second term (vertical entropy advection)\nonly if the day-night temperature (or potential temper-\nature) di\u000berence exceeds the vertical change in potential\ntemperature over a scale height. For a highly strati\fed\ntemperature pro\fle like those expected in the observable\natmospheres of hot Jupiters, the vertical change in po-\ntential temperature over a scale height is a signi\fcant\nfraction of the temperature itself.5\n5For a vertically isothermal temperature pro\fle, \u000eTstrat =Dayside-Nightside Temperature Di\u000berences in Hot Jupiter Atmospheres 11\nThus, one would expect that vertical entropy advec-\ntion dominates over horizontal entropy advection unless\nthe fractional day-night temperature di\u000berence is close to\nunity. This is just the weak temperature gradient regime\nmentioned in the Introduction. Considering this WTG\nbalance, we have6\n\u0001Teq\u0000\u0001T\n\u001crad\u0018WN2H\nR: (24)\nGiven our full solutions, we will show in Section 5.4.2 that\nhorizontal entropy advection is indeed smaller than ver-\ntical entropy advection in a hemospheric-averaged sense,\ndemonstrating the validity of (24).\n4.1.3. Hydrostatic balance\nHydrostatic balance relates the geopotential to the\ntemperature, and thus we can use it to relate the\nday-night temperature di\u000berence, \u0001 T, to the day-night\ngeopotential di\u000berence (alternatively pressure gradient)\non isobars. Hydrostatic balance implies that \u000e\b = R\nRTdlnp, where here \u000e\b is a vertical geopotential di\u000ber-\nence. Then, consider evaluating \u000e\b on the dayside and\nnightside, for two vertical air columns sharing the same\nvalue of \b at their base, at points separated by a hori-\nzontal distanceL. Given that these two points have the\nsame \b at the bottom isobar, we can di\u000berence \u000e\b at\nthese locations to solve for the geopotential change from\ndayside to nightside. Doing so, we \fnd the horizontal\ngeopotential di\u000berence7:\n\u0001\b\u0019RZpbot\np\u0001Tdlnp0: (25)\nIn Equation (25), both \u0001\b and \u0001 Tare functions of pres-\nsure. Here we only have to integrate to the pressure level\nat which our prescribed equilibrium dayside-nightside\ntemperature di\u000berence \u0001 Teqgoes to zero, which is la-\nbeledpbot.\nGiven that di\u000berences in scalar quantities from day-\nside to nightside are a function of pressure in our model,\nwe can use our Newtonian cooling parameterizations as\na guide to the form this pressure-dependence will take.\nHence, we take a form of \u0001 Tsimilar to that of \u0001 Teq,\nfocusing only on the region with pressure dependence:\n\u0001T= \u0001Ttopln(p=pbot)\nln(peq;top=pbot): (26)\nNow, integrating Equation (25) and dropping a factor of\ntwo, we \fnd\n\u0001\b\u0019R\u0001Tln\u0012pbot\np\u0013\n: (27)\n4.1.4. Momentum equation\nNext we analyze the approximate horizontal momen-\ntum equation, where the pressure gradient force driving\ngH=cp=RT=cp, and ifR=cp= 2=7 as appropriate for an H 2\natmosphere, then \u000eTstrat\u0019400 K for a typical hot Jupiter with a\ntemperature of 1500 K.\n6This balance was \frst considered for hot Jupiters by Showman\n& Guillot (2002, Eq. 22).\n7Note that this is essentially the hypsometric equation (e.g.\nHolton & Hakim 2013 pp. 19 or Wallace & Hobbs 2004 pp. 69-72).the circulation can be balanced by horizontal advection\n(in regions where the Rossby number Ro \u0011U=fL\u001d 1),\nvertical advection, the Coriolis force (Ro \u001c1), or drag:\nr\b\u0018max\u0014U2\nL;UW\nH;fU;U\n\u001cdrag\u0015\n: (28)\nNote that, given our continuity equation (20), the hor-\nizontal and vertical momentum advection terms U2=L\nandUW=Hare identical. As a result, we expect that\nthe \fnal solutions for \u0001 T(p),U(p), andW(p) should be\nthe same if horizontal momentum advection balances the\npressure gradient force as for the case when vertical mo-\nmentum advection balances it.\nExpressingr\b as \u0001\b=L, and relating \u0001\b to \u0001 Tus-\ning (27), we evaluate Equation (28) for every possible\ndominant term on the right hand side, now with explicit\npressure-dependence. This leads to approximate hori-\nzontal velocities that scale as\nU(p)\u00188\n>>><\n>>>:R\u0001T(p)\u0001lnp \u001cdrag(p)\nLDrag\nR\u0001T(p)\u0001lnp\nLfCoriolis\np\nR\u0001T(p)\u0001lnp Advection;\n(29)\nwhere we de\fne \u0001 ln p= ln(pbot=p) as the di\u000berence in\nlog pressure from the deep pressure pbot= 10 bars, where\nthe day-night forcing goes to zero, to some lower pressure\nof interest. As foreshadowed above, the same solution,\ngiven by the \fnal expression in (29), is attained when\neither horizontal or vertical momentum advection bal-\nances the pressure gradient force. The expressions for\nUin Equation (29) for advection and Coriolis force bal-\nancing pressure gradient match those of Showman et al.\n(2010) (see their Equations 48 and 49). Hence, we recap-\nture within our idealized model the previous expectations\nfor characteristic horizontal wind speeds on hot Jupiters\nbased on simple force balance. Note that these are ex-\npressions for characteristic horizontal wind speed, which\nrepresents average day-night \row and hence is not the\nzonal-mean zonal wind associated with superrotation.\nInvoking the relationship between UandWfrom Equa-\ntion (20), we can now write expressions for the vertical\nwind speed as a function of \u0001 T:\nW(p)\u00188\n>>>>><\n>>>>>:R\u0001T(p)H(p)\u0001lnp \u001cdrag(p)\nL2Drag\nR\u0001T(p)H(p)\u0001lnp\nL2fCoriolis\nH(p)\nLp\nR\u0001T(p)\u0001lnp Advection:\n(30)\n4.2. Full Solution for fractional dayside-nightside\ntemperature di\u000berences and wind speeds\nTo obtain our \fnal solution, we simply combine Equa-\ntion (24) and (30) to yield expressions for \u0001 T(p) and12 T.D. Komacek & A.P. Showman\nW(p) as a function of control parameters:\n\u0001T(p)\n\u0001Teq(p)\u00188\n>>>>>>><\n>>>>>>>:\u0012\n1 +\u001crad(p)\u001cdrag(p)\n\u001c2wave(p)\u0001lnp\u0013\u00001\nDrag\n\u0012\n1 +\u001crad(p)\nf\u001c2wave(p)\u0001lnp\u0013\u00001\nCoriolis\np\n\r(p) + 4\u0001Teq(p)\u0000p\n\r(p)p\n\r(p) + 4\u0001Teq(p) +p\n\r(p)Advection;\n(31)\nand\nW\u00188\n>>>>>>>>>>>>>>>>>>>>>>>>><\n>>>>>>>>>>>>>>>>>>>>>>>>>:RH(p)\nL2\u001c2\nwave(p)\n\u001crad(p)\u0001Teq(p)\n\u0002\"\n1\u0000\u0012\n1 +\u001crad(p)\u001cdrag(p)\n\u001c2wave(p)\u0001lnp\u0013\u00001#\nDrag\nRH(p)\nL2\u001c2\nwave(p)\n\u001crad(p)\u0001Teq(p)\n\u0002\"\n1\u0000\u0012\n1 +\u001crad(p)\nf\u001c2wave(p)\u0001lnp\u0013\u00001#\nCoriolis\nRH(p)\nL2\u001c2\nwave(p)\n\u001crad(p)\u0001Teq(p)\n\u0002\"\n1\u0000p\n\r(p) + 4\u0001Teq(p)\u0000p\n\r(p)p\n\r(p) + 4\u0001Teq(p) +p\n\r(p)#\nAdvection:\n(32)\nThe solution forU(p) is simply that for W(p) timesL=H.\nIn these solutions, we have de\fned \r(p) =\n\u001c2\nrad(p)L2\u0001lnp=(\u001c4\nwave(p)R). Moreover, we have substi-\ntuted a Kelvin wave propagation timescale \u001cwave(p) =\nL=(N(p)H(p)), as in Showman et al. (2013a). This wave\npropagation timescale is derived by taking the long ver-\ntical wavelength limit of the Kelvin wave dispersion rela-\ntionship, giving the fastest phase and group propagation\nspeedsc= 2NH. Hence, the wave propagation time\nacross a hemisphere is approximately L=(NH).\nOur drag and Coriolis-dominated equations are the\nsame expressions as in Perez-Becker & Showman (2013),\nbut the advection-dominated cases di\u000ber. Note that as in\nPerez-Becker & Showman (2013), the dayside-nightside\ntemperature (thickness in their case) di\u000berences for the\ndrag and Coriolis-dominated regimes are equal if \u001cdrag\u0018\n1=f. The solutions for \u0001 T=\u0001Teqin the drag and Corio-\nlis cases are independent of the forcing amplitude \u0001 Teq,\nwhereas the solution in the advection case depends on\n\u0001Teq, as one would expect given the nonlinear nature of\nthe momentum balance.\nWith these \fnal solutions for the dayside-nightside\ntemperature di\u000berence and characteristic wind speeds,\nwe can compare our theory to the GCM results in Sec-\ntion 5. However, we \frst must diagnose which solution\nis appropriate for any given combination of parameters.\n4.3. Momentum equation regime\nTo determine which regime is appropriate in our solu-\ntions from Section 4.2 for given choices of control pa-\nrameters (\u001crad;top,\u001cdrag, and \u0001Teq;top), we perform a\ncomparison of the magnitudes of the drag, Coriolis, and\nmomentum advection terms. We do this by comparingthe relative amplitues of their characteristic timescales:\n\u001cdrag, 1=f, and\u001cadv. Using Equations (20) and (32), we\ncan write\u001cadvas\n\u001cadv(p) =L2\u001crad(p)\nR\u001c2wave(p)\u0001Teq(p)\n\u0002\"\n1\u0000p\n\r(p) + 4\u0001Teq(p)\u0000p\n\r(p)p\n\r(p) + 4\u0001Teq(p) +p\n\r(p)#\u00001\n:\n(33)\nNow, we can write the condition that determines which\nregime the solution is in. If \u001cdrag><\n>>:\u0012\n1 +\u001crad(p)\u001cdrag(p)\n\u001c2wave(p)\u0001lnp\u0013\u00001\n\u001cdragf\u00001:\n(34)\nIn the regime where Ro >1, the advection term has a\ngreater magnitude than the Coriolis force. This is rele-\nvant to the equatorial regions of any given planet, where\nf!0 due to the latitudinal dependence of the Coriolis\nforce. Additionally, if advection is very strong (or rota-\ntion very slow), this regime could extend to the mid-to-\nhigh latitudes. The three-term force balance in regions\nwhere Ro>1 is hence between drag, advection, and the\nday-night pressure gradient force. Using the same rea-\nsoning as above, we can write the fractional day-night\n8Alternatively, one can calculate the Rossby number from the\nadvective timescale, as we take fas an external parameter and\nU=L=\u001c\u00001\nadv. Hence, one can calculate the Rossby number (as a\nfunction of latitude) as Ro( \u001e) = [f(\u001e)\u001cadv]\u00001, where\u001cadvis given\nby Equation (33).Dayside-Nightside Temperature Di\u000berences in Hot Jupiter Atmospheres 13\ntemperature di\u000berence at the equator Aeq(p):\nAeq(p)\u00188\n>>>>>>><\n>>>>>>>:\u0012\n1 +\u001crad(p)\u001cdrag(p)\n\u001c2wave(p)\u0001lnp\u0013\u00001\n\u001cdrag(p)<\u001cadv(p)\np\n\r(p) + 4\u0001Teq(p)\u0000p\n\r(p)p\n\r(p) + 4\u0001Teq(p) +p\n\r(p)\u001cdrag(p)>\u001cadv(p);\n(35)\nwhere\u001cadvis evaluated from Eq. (33).\n4.4. Transition from low to high day-night temperature\ndi\u000berences\nFrom the expressions above for the fractional day-night\ntemperature di\u000berence, it is clear that a comparison be-\ntween wave propagation timescales and other relevant\ntimescales (radiative, drag, Coriolis) controls the ampli-\ntude of the day-night temperature di\u000berence. Here, we\nuse our theory to obtain timescale comparisons for the\ntransition from small to large day-night temperature dif-\nference (relative to that in radiative equilibrium).\nUsing the expression for overall Afrom Equation (34),\nthe transition between low fractional dayside-nightside\ntemperature di\u000berences ( A!0) and high fractional\ndayside-nightside temperature di\u000berences ( A!1) oc-\ncurs when9:\n\u001cwave(p)\u0018(q\n\u001cdrag(p)\u001crad(p)\u0001lnp \u001c drag(p)f\u00001:\n(36)\nThis timescale comparison is valid in regions where Ro <\n1, which occurs at nearly all latitudes (except those bor-\ndering the equator, where Ais given by Equation 35).\nWhen\u001cwave is smaller than the expression on the right-\nhand side of Equation (36), the day-night temperature\ndi\u000berences are small, while if it is larger the day-night\ntemperature di\u000berences are necessarily large. It is no-\ntable that the transition between low and high day-night\ntemperature di\u000berence always involves a comparison be-\ntween wave and radiative timescales, no matter what\nregime the system is in. Meanwhile, drag is only im-\nportant if\u001cdragp\nf\u00001\u001crad(p)\u0001lnp. If frictional drag has a charac-\nteristic timescale shorter than 1 =f, the transition is\ninstead governed by a similar comparison between\n\u001cwave andp\n\u001cdrag(p)\u001crad(p)\u0001lnp. The theory also\ncovers the situation where both drag and Corio-\nlis forces are weak compared to advective forces, a\nsituation which should be relevant to planets that\nrotate especially slowly.\n5. The same theory used to predict dayside-nightside\ntemperature di\u000berences can also be inverted to cal-\nculate characteristic horizontal and vertical veloc-\nities from phase curve information. Horizontal\nvelocities may be estimated given only the rota-\ntion rate and dayside-nightside \rux di\u000berences on\na given tidally locked planet. Similarly, vertical\nvelocities may be calculated with the additional\nknowledge of atmospheric composition and tem-\nperature at a given pressure. This velocity scaling\nis a powerful tool to gain information on atmo-\nspheric circulation from data attainable through\nphase curves and may be extended to investigate\nproperties of the circulation for the suite of tidally\nlocked exoplanets without a boundary layer.\nThis research was supported by NASA Origins grant\nNNX12AI79G to APS. TDK acknowledges support from\nNASA headquarters under the NASA Earth and Space\nScience Fellowship Program Grant PLANET14F-0038.\nWe thank the anonymous referee for useful comments.We also thank Josh Lothringer, Xianyu Tan, and Xi\nZhang for helpful comments on the manuscript. Re-\nsources supporting this work were provided by the\nNASA High-End Computing (HEC) Program through\nthe NASA Advanced Supercomputing (NAS) Division at\nAmes Research Center.\nAPPENDIX\nA.OBSERVED FRACTIONAL DAYSIDE-NIGHTSIDE\nBRIGHTNESS TEMPERATURE DIFFERENCES\nTable 1 displays the data utilized to calculate the\nAobs\u0000Teqpoints shown in Figure 1, along with appropri-\nate references. To compile this data, we collected either\nthe error-weighted or photometric dayside brightness\ntemperature and dayside-nightside brightness tempera-\nture di\u000berences. If no error-weighted values were given,\nwe calculated each brightness temperature as an arith-\nmetic mean of that provided at each wavelength. Specif-\nically, we computed arithmetic means for HD 189733b,\nWASP-43b, and WASP-18b, with the other transiting\nplanets either having an error-weighted value provided\nor only one photometric wavelength with available data.\nFor both WASP-19b and HAT-P-7b, we utilized the\n4:5\u0016m day-night brightness temperature di\u000berence, as\nonly upper limits on the nightside brightness tempera-\ntures at 3:6\u0016m were available. Note that though HAT-\nP-7b seems to have a day-night brightness temperature\ndi\u000berence somewhat below that expected, the lower limit\nonAobsat 3:6\u0016m is 0:483 (Wong et al. 2015b), 33 :4%\nlarger than the value of Aobsat 4:5\u0016m.\nTo compute a lower limit on Aobsfor WASP-43b, we\nutilized the 1 \u001bupper limit on the nightside bolometric\n\rux from Stevenson et al. (2014). We used a similar\nmethod for WASP-18b, averaging the lower limits on\nAobsfrom the 3:6\u0016m and 4:5\u0016m nightside 1 \u001b\rux up-\nper limits from Maxted et al. (2013). This enabled us to\ncalculate the lower limit of Aobs\u0011Aobs;llas a function\nof dayside and nightside \ruxes as\nAobs;ll=F1=4\nday\u0000F1=4\nnight;max\nF1=4\nday: (A1)\nIn Equation (A1) above, Fdayis the dayside bolometric\n\rux andFnight;maxis the upper limit on the nightside\nbolometric \rux.\nREFERENCES\nAdcroft, A., Hill, C., Campin, J., Marshall, J., & Heimbach, P.\n2004, Monthly Weather Review, 132, 2845\nAndrews, D., Holton, J., & Leovy, C. 1987, International\nGeophysics Series, Vol. 40, Middle Atmosphere Dynamics, ed.\nR. Dmowska & J. Holton (Academic Press)\nBatygin, K., Stanley, S., & Stevenson, D. 2013, The Astrophysical\nJournal, 776, 53\nBatygin, K. & Stevenson, D. 2010, The Astrophysical Journal\nLetters, 714, L238\nBorucki, W., Koch, D., Jenkins, J., Sasselov, D., Gilliland, R.,\nBatalha, N., Latham, D., Caldwell, D., Basri, G., Brown, T.,\nChristensen-Dalsgaard, J., Cochran, W., Devore, E., Dunham,\nD., Dupree, A., Gautier, T., Geary, J., Gould, A., Howell, S.,\nKjeldsen, H., Lissauer, J., Marcy, G., Meibom, S., Morrison,\nD., & Tarter, J. 2009, Science, 325, 709\nBretherton, C. & Smolarkiewicz, P. 1989, Journal of the\nAtmospheric Sciences, 46, 740Charbonneau, D., Brown, T., Latham, D., & Mayor, M. 2000,\nThe Astrophysical Journal, 529, L45\nCooper, C. & Showman, A. 2005, The Astrophysical Journal\nLetters, 629, L45\nCowan, N. & Agol, E. 2011, The Astrophysical Journal, 759, 54\nCowan, N., Agol, E., & Charbonneau, D. 2007, Monthly Notices\nof the Royal Astronomical Society, 379, 641\nCowan, N., Machalek, P., Croll, B., Shekhtman, L., Burrows, A.,\nDeming, D., Greene, T., & Hora, J. 2012, The Astrophysical\nJournal, 747, 82\nCross\feld, I., Knutson, H., Fortney, J., Showman, A., Cowan, N.,\n& Deming, D. 2012, The Astrophysical Journal, 752, 81\nDemory, B., Wit, J. D., Lewis, N., Fortney, J., Zsom, A., Seager,\nS., Knutson, H., Heng, K., Madhusudhan, N., Gillon, M.,\nBarclay, T., Desert, J., Parmentier, V., & Cowan, N. 2013, The\nAstrophysical Journal Letters, 776, L25\nDobbs-Dixon, I. & Agol, E. 2013, Monthly Notices of the Royal\nAstronomical Society, 435, 3159Dayside-Nightside Temperature Di\u000berences in Hot Jupiter Atmospheres 23\nPlanet Teq(K) DaysideTb(K) DaysideTb- Nightside Tb(K)Aobs References\nHD 189733b 1202:8 1250\u000613 313\u000623 0:250\u00060:018 Knutson et al. (2007, 2009b, 2012)\nWASP-43b 1378:0 1725\u00066 N/A \u00150:54 Stevenson et al. (2014)\nHD 209458b 1449:5 1499\u000615 527\u000646 0:352\u00060:031 Cross\feld et al. (2012); Zellem et al. (2014)\nHD 149026b 1677:8 1440\u0006150 480\u0006140 0:333\u00060:103 Knutson et al. (2009a)\nWASP-14b 1869:2 2302\u000638 1081\u000682:3 0:470\u00060:037 Wong et al. (2015a)\nWASP-19b 2070:2 2357\u000664 1227\u0006248 0:521\u00060:106 Wong et al. (2015b)\nHAT-P-7b 2230:0 2682\u000649 972\u0006187 0:362\u00060:070 Wong et al. (2015b)\nWASP-18b 2403:3 3153\u000651 N/A \u00150:40 Nymeyer et al. (2011); Maxted et al. (2013)\nWASP-12b 2589:7 2928\u000697 1945\u0006223 0:664\u00060:079 Cowan et al. (2012)\nTABLE 1\nCompilation of data used to make Figure 1, along with appropriate references.\nFortney, J. J., Lodders, K., Marley, M. S., & Freedman, R. S.\n2008, ApJ, 678, 1419\nGill, A. 1980, Quarterly Journal of the Royal Meteorological\nSociety, 106, 447\nGinzburg, S. & Sari, R. 2015, arXiv 1511.00135\nHeld, I. 2005, Bull. Amer. Meteorological Soc., 86, 1609\nHeld, I. & Suarez, M. 1994, Bulletin of the American\nMeteorological Society, 75, 1825\nHeng, K., Frierson, D., & Phillips, P. 2011a, Monthly Notices of\nthe Royal Astronomical Society, 418, 2669\nHeng, K., Menou, K., & Phillips, P. 2011b, Monthly Notices of\nthe Royal Astronomical Society, 413, 2380\nHenry, G., Marcy, G., Butler, R., & Vogt, S. 2000, The\nAstrophysical Journal, 529, L41\nHolton, J. & Hakim, G. 2013, An introduction to dynamic\nmeteorology, 5th edn. (Academic Press)\nIro, N., B\u0013 ezard, B., & Guillot, T. 2005, Astronomy and\nAstrophysics, 436, 719\nKataria, T., Showman, A., Fortney, J., Stevenson, K., Line, M.,\nKriedberg, L., Bean, J., & Desert, J. 2015, The Astrophysical\nJournal, 801, 86\nKataria, T., Showman, A., Lewis, N., Fortney, J., Marley, M., &\nFreedman, R. 2013, The Astrophysical Journal, 767, 76\nKnutson, H., Charbonneau, D., Allen, L., Fortney, J., Agol, E.,\nCowan, N., Showman, A., Cooper, C., & Megeath, S. 2007,\nNature, 447, 183\nKnutson, H., Charbonneau, D., Cowan, N., Fortney, J.,\nShowman, A., Agol, E., & Henry, G. 2009a, The Astrophysical\nJournal, 703, 769\nKnutson, H., Charbonneau, D., Cowan, N., Fortney, J.,\nShowman, A., Agol, E., Henry, G., Everett, M., & Allen, L.\n2009b, The Astrophysical Journal, 690, 822\nKnutson, H., Lewis, N., Fortney, J., Burrows, A., Showman, A.,\nCowan, N., Agol, E., Aigrain, S., Charbonneau, D., Deming,\nD., Desert, J., Henry, G., Langton, J., & Laughlin, G. 2012,\nThe Astrophysical Journal, 754, 22\nKoll, D. & Abbot, D. 2015, The Astrophysical Journal, 802, 21\nLewis, N., Showman, A., Fortney, J., Marley, M., Freedman, R.,\n& Lodders, K. 2010, The Astrophysical Journal, 720, 344\nLi, J. & Goodman, J. 2010, The Astrophysical Journal, 725, 1146\nLiu, B. & Showman, A. 2013, The Astrophysical Journal, 770, 42\nMatsuno, T. 1966, Journal of the Meteorological Society of\nJapan, 44, 25\nMaxted, P., Anderson, D., Doyle, A., Gillon, M., Harrington, J.,\nIro, N., Jehin, E., Lafreniere, D., Smalley, B., & Southworth, J.\n2013, Monthly Notices of the Royal Astronomical Society, 428,\n2645\nMayne, N., Bara\u000be, I., Acreman, D., Smith, C., Browning, M.,\nAmundsen, D., Wood, N., Thuburn, J., & Jackson, D. 2014,\nAstronomy and Astrophysics, 561, A1\nMenou, K. 2012, The Astrophysical Journal, 745, 138\nMenou, K. & Rauscher, E. 2009, The Astrophysical Journal, 700,\n887\nNymeyer, S., Harrington, J., Hardy, R., Stevenson, K., Campo,\nC., Madhusudhan, N., Collier-Cameron, A., Loredo, T., Blecic,\nJ., Bowman, W., Britt, C., Cubillos, P., Hellier, C., Gillon, M.,\nMaxted, P., Hebb, L., Wheatley, P., Pollaco, D., & Anderson,\nD. 2011, The Astrophysical Journal, 742, 35\nParmentier, V., Showman, A., & Lian, Y. 2013, Astronomy and\nAstrophysics, 558, A91Perez-Becker, D. & Showman, A. 2013, The Astrophysical\nJournal, 776, 134\nPerna, R., Heng, K., & Pont, F. 2012, The Astrophysical Journal,\n751, 59\nPerna, R., Menou, K., & Rauscher, E. 2010, The Astrophysical\nJournal, 719, 1421\nPolvani, L. & Sobel, A. 2001, Journal of the Atmospheric\nSciences, 59, 1744\nRauscher, E. & Menou, K. 2010, The Astrophysical Journal, 714,\n1334\n|. 2012a, The Astrophysical Journal, 750, 96\n|. 2012b, The Astrophysical Journal, 745, 78\n|. 2013, The Astrophysical Journal, 764, 103\nRogers, T. & Komacek, T. 2014, The Astrophysical Journal, 794,\n132\nRogers, T. & Showman, A. 2014, The Astrophysical Journal\nLetters, 782, L4\nSchwartz, J. & Cowan, N. 2015, Monthly Notices of the Royal\nAstronomical Society, 449, 4192\nShowman, A., Cho, J., & Menou, K. Exoplanets, ed. S. Seager\n(Tucson, AZ: University of Arizona Press)\nShowman, A., Cooper, C., Fortney, J., & Marley, M. 2008a, The\nAstrophysical Journal, 682, 559\nShowman, A., Fortney, J., Lewis, N., & Shabram, M. 2013a, The\nAstrophysical Journal, 762, 24\nShowman, A., Fortney, J., Lian, Y., Marley, M., Freedman, R.,\nKnutson, H., & Charbonneau, D. 2009, The Astrophysical\nJournal, 699, 564\nShowman, A. & Guillot, T. 2002, Astronomy and Astrophysics,\n385, 166\nShowman, A., Lewis, N., & Fortney, J. 2015, The Astrophysical\nJournal, 801, 95\nShowman, A., Menou, K., & Cho, J. 2008b, Extreme Solar\nSystems ASP Conference Series, 398, 419\nShowman, A. & Polvani, L. 2010, Geophysical Research Letters,\n37, L18811\n|. 2011, The Astrophysical Journal, 738, 71\nShowman, A., Wordsworth, R., Merlis, T., & Kaspi, Y.\nComparative Climatology of Terrestrial Planets, ed.\nS. Mackwell, A. Simon-Miller, J. Harder, & M. Bullock\n(Tucson, AZ: University of Arizona Press)\nSobel, A. 2002, Chaos, 12, 451\nSobel, A., Nilsson, J., & Polvani, L. 2001, Journal of the\nAtmospheric Sciences, 58, 3650\nStevenson, K., Desert, J., Line, M., Bean, J., Fortney, J.,\nShowman, A., Kataria, T., Kriedberg, L., McCullough, P.,\nHenry, G., Charbonneau, D., Burrows, A., Seager, S.,\nMadhusudhan, N., Williamson, M., & Homeier, D. 2014,\nScience, 346, 838\nThrastarson, H. & Cho, J. 2010, The Astrophysical Journal, 716,\n144\nTsai, S., Dobbs-Dixon, I., & Gu, P. 2014, The Astrophysical\nJournal, 793, 141\nWallace, J. & Hobbs, P. 2004, Atmospheric Science: An\nIntroductory Survey, 2nd edn. (San Diego: Academic Press)\nWong, I., Knutson, H., Lewis, N., Burrows, A., Agol, E., Cowan,\nN., Deming, D., Desert, J., Fortney, J., Fulton, B., Howard, A.,\nKataria, T., Langton, J., Laughlin, G., Schwartz, J., Showman,\nA., & Todorov, K. 2015a, The Astrophysical Journal, 811, 12224 T.D. Komacek & A.P. Showman\nWong, I., Knutson, H., Lewis, N., Kataria, T., Burrows, A.,\nFortney, J., Schwartz, J., Shporer, A., Agol, E., Cowan, N.,\nDeming, D., Desert, J., Fulton, B., Howard, A., Langton, J.,\nLaughlin, G., Showman, A., & Todorov, K. 2015b, arXiv\n1512.09342\nWordsworth, R. 2015, The Astrophysical Journal, 806, 180Youdin, A. & Mitchell, J. 2010, The Astrophysical Journal, 721,\n1113\nZellem, R., Lewis, N., Knutson, H., Gri\u000eth, C., Showman, A.,\nFortney, J., Cowan, N., Agol, E., Burrows, A., Charbonneau,\nD., Deming, D., Laughlin, G., & Langton, J. 2014, The\nAstrophysical Journal, 790, 53" }, { "title": "1702.00637v1.Exponential_stability_for_a_coupled_system_of_damped_undamped_plate_equations.pdf", "content": "arXiv:1702.00637v1 [math.AP] 2 Feb 2017EXPONENTIAL STABILITY FOR A COUPLED SYSTEM OF\nDAMPED-UNDAMPED PLATE EQUATIONS\nROBERT DENK AND FELIX KAMMERLANDER\nAbstract. We consider the transmission problem for a coupled system of\nundamped and structurally damped plate equations in two suffi ciently smooth\nand bounded subdomains. It is shown that, independently of t he size of the\ndamped part, the damping is strong enough to produce uniform exponential\ndecay of the energy of the coupled system.\n1.Introduction\nIn this paper, we investigate a coupled system of linear plate equatio ns where an\nundamped plate and a structurally damped plate are coupled throug h transmission\nconditions. From the point of view of applications, there is a connect ion to the\nsuppression of vibration of elastic structures which is a main topic in m aterial\nscience. The undamped plate equation can be seen as a linear model f or vibrating\nstiff objects where the potential energy is related to curvature- like terms, resulting\nin the bi-Laplacianoperatoras the main elastic operator, see, e.g., [ 12], Chapter 12.\nFor the purely undamped plate, we have no energy dissipation, and t he governing\nsemigroup is unitary. The model of structural damping is widely used to describe\nsmoothing effects and loss of energy (cf. [ 20] for a discussion of the model). Here,\nwe consider the damping term which has order two in the spatial varia bles, so it is\nof half order of the leading elastic term, see also [ 8] and [9] for the analysis of the\nstructurally damped plate equation.\nFrom a theoretical point of view, the resulting system can be seen a s a trans-\nmission problem of mixed type: While the structurally damped plate equ ation is\nof parabolic nature, the undamped part is of dissipative nature. Be low we will see\nthat the damping is strong enough (independent of the size of the d amped part) to\nobtain exponential stability for the semigroup of the coupled syste m. The analog\nresult for a coupled system of thermoelastic / elastic plates was obt ained in [ 16].\nThe question of analyticity of the semigroup for a coupled thermoela stic plate /\nplate system is discussed in [ 10]. In [14], a plate / plate transmission problem with\ndamping only on a part of the boundary with resulting polynomial deca y was stud-\nied, see also [ 4] for the proof of exponential stability for a boundary stabilized pla te\n/ plate transmission problem. Transmission problems of plate / plate t ype can also\nbe seen as an equation with coefficients having jumps, cf. [ 13].\nIn the system we consider the damping effect acting only through th e transmis-\nsion interface. Closely related is the question of boundary damping, see, e.g., [ 17]\nor [22]. In the literature, there are many results on coupled systems of p late / wave\ntype (cf. [ 3] and the references therein). In particular, in [ 6] and [7], the exponen-\ntial stability for an abstract wave equation coupled with a plate-like e quation on\nthe boundary is studied. To our knowledge, the undamped / struct urally damped\nplate system has not yet been studied in literature.\nDate: February 2, 2017.\n2010Mathematics Subject Classification. 74K20; 74H40; 35B40; 35Q74.\nKey words and phrases. Plate equation, transmission problem, exponential stabil ity.\n12 ROBERT DENK AND FELIX KAMMERLANDER\nLet Ω⊂Rnbe a bounded domain with boundary Γ 1:=∂Ω, and let Ω 2⊂Ω be a\nnon-empty bounded domain satisfying Ω2⊂Ω. We set Γ := ∂Ω2and Ω 1:= Ω\\Ω2.\nThen, Γ is the common interface (transmission interface) between Ω1and Ω 2, and\n∂Ω1=∂Ω∪Γ (see Figure 1for the geometrical situation). All domains are assumed\nto be of class C4. For technical reasons, we assume n≤4, including the physically\nmost relevant cases n= 1 and n= 2. Let νdenote the outer unit normal on Γ 1.\nOn Γ, we choose νto be the outer unit normal with respect to Ω 2. Thus,νis the\ninner unit normal vector on Γ with respect to Ω 1, see Figure 1. Note that, apart\nfrom the smoothness, we do not impose a geometrical condition on t he domains.\nν\nν\nΩ2Ω1\nΓ1\nΓ\nFigure 1. The set Ω = Ω 1∪Γ∪Ω2.\nWe consider a transmission problem for thin plates where the plate in Ω 2is\nundamped and the material in Ω 1is structurally damped. More precisely, we are\nlooking for solutions ui: Ωi→Cof the system\n∂2\ntu1+∆2u1−ρ∆∂tu1= 0 in (0 ,∞)×Ω1, (1-1)\n∂2\ntu2+∆2u2= 0 in (0 ,∞)×Ω2 (1-2)\nwithclamped boundary conditions\nu1=∂νu1= 0 on Γ 1 (1-3)\nHere,ρ∈(0,∞) is the damping factor. The transmission conditions on Γ are given\nby\nu1=u2, (1-4)\n∂νu1=∂νu2, (1-5)\n∆u1= ∆u2, (1-6)\n−ρ∂ν∂tu1+∂ν∆u1=∂ν∆u2. (1-7)\nThe problem is completed by the initial conditions\nu1(0,·) =u0\n1, ∂tu1(0,·) =u1\n1in Ω1, (1-8)\nu2(0,·) =u0\n2, ∂tu2(0,·) =u1\n2in Ω2. (1-9)\nThe energy of the system ( 1-1)-(1-9) is defined as\nE(t) :=1\n2/integraldisplay\nΩ1|∂tu1|2+|∆u1|2dx+1\n2/integraldisplay\nΩ2|∂tu2|2+|∆u2|2dx. (1-10)\nIf (u1,u2) is a solution, integration by parts yields the estimate\nd\ndtE(t) =−ρ/ba∇dbl∇∂tu1/ba∇dbl2\nL2(Ω1)≤0. (1-11)\nNote that ui=∂νui= 0 on Γ iimplies∂tui=∂ν∂tui= 0 on Γ ifori= 1,2.The\nestimate shows that the energy of the transmission problem is decr easing in time\nand the dissipation is caused by the damped part u1.EXPONENTIAL STABILITY FOR A COUPLED SYSTEM 3\nOur main result, Theorem 4.5below, states that the damping in Ω 1is strong\nenough to achieve exponential decrease of the energy, i.e. there exist constants\nC,κ >0 such that\nE(t)≤CE(0)e−κt\nholds for all t≥0.To prove this, we first study the resolvent and the spectrum of\nthe first-ordersystem related to( 1-1)–(1-7) in Section 2. In the proofofexponential\nstability, we also need an a priori estimate on the damped part which is obtained in\nSection 3 with the help of the interpolation-extrapolation scales of B anach spaces.\nFinally, the results from Section 2 and 3 are used to prove the main re sult on\nexponential stability in Section 4.\n2.The spectrum of the first-order system\nSettingU:= (u1,u2,v1,v2)⊤withvj:=∂tuj, we rewrite the transmission prob-\nlem (1-1)-(1-9) as\n∂tU(t)−AU(t) = 0 (t >0), U(0) =U0 (2-1)\nwhere the operator Aacts in form of the matrix\nA(D) :=\n0 0 1 0\n0 0 0 1\n−∆20ρ∆ 0\n0−∆20 0\n.\nAs the basic space for the first two components ( u1,u2), we will choose\nX(Ω) :=/braceleftbig\n(u1,u2)∈H2(Ω1)×H2(Ω2) :u1=∂νu1= 0 on Γ 1,\nu1=u2on Γ, ∂νu1=∂νu2on Γ/bracerightbig\n.\nRemark 2.1. a) Let (u1,u2)∈H2(Ω1)×H2(Ω2). Then the conditions u1=u2,\n∂νu1=∂νu2on Γ are equivalent to u:=χΩ1u1+χΩ2u2∈H2(Ω), where χΩjstands\nfor the characteristic function of Ω j, i.e.χΩj(x) = 1 for x∈ΩjandχΩj(x) = 0\nelse. Therefore, we have\nX(Ω) ={(u|Ω1,u|Ω2) :u∈H2\n0(Ω)}.\nIn the following, we will several times use the identification of ( u1,u2) andu.\nb) The norm in X(Ω) is defined as\n/ba∇dbl(u1,u2)/ba∇dblX(Ω):=/parenleftig\n/ba∇dbl∆u1/ba∇dbl2\nL2(Ω1)+/ba∇dbl∆u2/ba∇dbl2\nL2(Ω2)/parenrightig1/2\n.\nNotethatthisnormisequivalenttothestandardnorm( /ba∇dblu1/ba∇dbl2\nH2(Ω1)+/ba∇dblu2/ba∇dbl2\nH2(Ω2))1/2.\nIn fact, due to the invertibilityofthe Dirichlet Laplacianin Ω, the norm s/ba∇dbl∆u/ba∇dblL2(Ω)\nand/ba∇dblu/ba∇dblH2(Ω)are equivalent on the space H2(Ω)∩H1\n0(Ω). As H2\n0(Ω) is a closed\nsubspace of H2(Ω)∩H1\n0(Ω), these norms are also equivalent on H2\n0(Ω), and now\nthe assertion follows from part a) (see also [ 11], Proposition 2.1 and Proposition\n2.2).\nWe say that the transmission conditions ( 1-6) and (1-7) are weakly satisfied if\n/a\\}b∇acketle{t∆2u1−ρ∆v1,ϕ1/a\\}b∇acket∇i}htL2(Ω1)+/a\\}b∇acketle{t∆2u2,ϕ2/a\\}b∇acket∇i}htL2(Ω2)\n=/a\\}b∇acketle{t∆u1,∆ϕ1/a\\}b∇acket∇i}htL2(Ω1)+/a\\}b∇acketle{t∆u2,∆ϕ2/a\\}b∇acket∇i}htL2(Ω2)+ρ/a\\}b∇acketle{t∇v1,∇ϕ1/a\\}b∇acket∇i}htL2(Ω1)(2-2)\nholds for all ( ϕ1,ϕ2)∈X(Ω). Let\nH:=X(Ω)×L2(Ω1)×L2(Ω2).\nThen we define the operator A:H ⊃D(A)→ Hby4 ROBERT DENK AND FELIX KAMMERLANDER\nD(A) :=/braceleftbig\n(u1,u2,v1,v2)∈X(Ω)×X(Ω) : ∆2u1∈L2(Ω1),∆2u2∈L2(Ω2),\n(1-6) and (1-7) are weakly satisfied/bracerightbig\nandAU:=A(D)U(U∈D(A)).\nWe will see in Lemma 2.4below that the functions in D(A) are sufficiently\nsmooth and the transmission conditions hold in the sense of traces.\nTheorem 2.2. The operator Ais the generator of a C0-semigroup of contractions\non the Hilbert space H.Therefore, for all U0∈D(A)the Cauchy problem (2-1)has\na unique classical solution U∈C1([0,∞),H)withU(t)∈D(A)for allt≥0.\nProof.By the definition of D(A) and the weak transmission conditions ( 2-2), it is\nimmediately seen that\nRe/a\\}b∇acketle{tAU,U/a\\}b∇acket∇i}htH=−ρ/ba∇dbl∇v1/ba∇dbl2\nL2(Ω1)(U∈D(A)).\nHence,Ais dissipative. We want to show that 1 −Ais surjective. For this, let\nF= (f1,f2,g1,g2)⊤∈ H.We have to find U= (u1,u2,v1,v2)⊤∈D(A) satisfying\nu1−v1=f1,\nu2−v2=f2,\nv1+∆2u1−ρ∆v1=g1,\nv2+∆2u2=g2.\nPlugging in vi=ui−fifori= 1,2 in the third and fourth equation yields that we\nhave to solve\nu1+∆2u1−ρ∆u1=g1+f1−ρ∆f1, (2-3)\nu2+∆2u2=g2+f2 (2-4)\nas equalities in L2(Ω1) andL2(Ω2),respectively.\nWe define the continuous sesquilinear form B:X(Ω)×X(Ω)→Cby\nB(u,ϕ) =/a\\}b∇acketle{tu1,ϕ1/a\\}b∇acket∇i}htL2(Ω1)+/a\\}b∇acketle{t∆u1,∆ϕ1/a\\}b∇acket∇i}htL2(Ω2)+ρ/a\\}b∇acketle{t∇u1,∇ϕ1/a\\}b∇acket∇i}htL2(Ω1)\n+/a\\}b∇acketle{tu2,ϕ2/a\\}b∇acket∇i}htL2(Ω2)+/a\\}b∇acketle{t∆u2,∆ϕ2/a\\}b∇acket∇i}htL2(Ω2)\nforu= (u1,u2),ϕ= (ϕ1,ϕ2)∈X(Ω). Since\nReB(u,u)≥ /ba∇dbl(u1,u2)/ba∇dbl2\nX(Ω)(u∈X(Ω)),\nBis coercive. Obviously, the mapping Λ: X(Ω)→Cdefined by\nΛ(ϕ) :=/a\\}b∇acketle{tg1+f1,ϕ1/a\\}b∇acket∇i}htL2(Ω1)+ρ/a\\}b∇acketle{t∇f1,∇ϕ1/a\\}b∇acket∇i}htL2(Ω1)+/a\\}b∇acketle{tg2+f2,ϕ2/a\\}b∇acket∇i}htL2(Ω2)\nforϕ= (ϕ1,ϕ2)∈X(Ω) is linear and continuous. By the theorem of Lax-Milgram,\nthere exists a unique u= (u1,u2)∈X(Ω) such that B(u,ϕ) = Λ(ϕ) holds for all\nϕ∈X(Ω). In particular, choosing ( ϕ1,ϕ2)∈C∞\n0(Ω1)×C∞\n0(Ω2)⊂X(Ω), we see\nthat (2-3) and (2-4) hold in the sense of distributions in Ω 1and Ω 2, respectively.\nAs the right-hand side of ( 2-3) belongs to L2(Ω1), the same holds for the left-hand\nside. Due to u1∈H2(Ω1), this yields ∆2u1∈L2(Ω1). In the same way, we see\nthat (2-4) holds as equality in L2(Ω2) and that ∆2u2∈L2(Ω2).\nSetv1:=u1−f1andv2:=u2−f2. By (2-3)–(2-4), we have\n∆2u1−ρ∆v1=−u1+g1+f1,\n∆2u2=−u2+g2+f2.(2-5)\nLetϕ= (ϕ1,ϕ2)∈X(Ω). Then, because of ( 2-5) andB(u,ϕ) = Λ(ϕ), we get\n/a\\}b∇acketle{t∆2u1−ρ∆v1,ϕ1/a\\}b∇acket∇i}htL2(Ω1)+/a\\}b∇acketle{t∆2u2,ϕ2/a\\}b∇acket∇i}htL2(Ω2)EXPONENTIAL STABILITY FOR A COUPLED SYSTEM 5\n=/a\\}b∇acketle{t−u1+g1+f1,ϕ1/a\\}b∇acket∇i}htL2(Ω1)+/a\\}b∇acketle{t−u2+g2+f2,ϕ2/a\\}b∇acket∇i}htL2(Ω2)\n=/a\\}b∇acketle{t∆u1,∆ϕ1/a\\}b∇acket∇i}htL2(Ω1)+/a\\}b∇acketle{t∆u2,∆ϕ2/a\\}b∇acket∇i}htL2(Ω2)+ρ/a\\}b∇acketle{t∇v1,∇ϕ1/a\\}b∇acket∇i}htL2(Ω1).\nTherefore, the weak transmissionconditions ( 2-2) aresatisfied. Altogether, we have\nseen that U:= (u1,u2,v1,v2)⊤belongs to D(A). Because of ( 2-3)–(2-4) and the\ndefinition of v1,v2, we also have (1 −A)U=F. Therefore, 1 −Ais surjective which\nimplies that Ais densely defined (see [ 18], Theorem 4.6). An application of the\nLumer-Phillips theorem now yields the statement of the theorem. /square\nRemark 2.3. In the same way as in the previous proof, one can show that the\noperator Ais continuously invertible, i.e. 0 belongs to the resolvent set ρ(A). To\nshow this, we now have to consider\n∆2u1=g1−ρ∆f1, (2-6)\n∆2u2=g2 (2-7)\ninstead of ( 2-3)–(2-4). The sesquilinear form Band the functional Λ are now\ndefined by B(u,ϕ) =/a\\}b∇acketle{tu,ϕ/a\\}b∇acket∇i}htX(Ω)and\nΛ(ϕ) :=/a\\}b∇acketle{tg1,ϕ1/a\\}b∇acket∇i}htL2(Ω1)+ρ/a\\}b∇acketle{t∇f1,∇ϕ1/a\\}b∇acket∇i}htL2(Ω1)+/a\\}b∇acketle{tg2,ϕ2/a\\}b∇acket∇i}htL2(Ω2)\nforu= (u1,u2),ϕ= (ϕ1,ϕ2)∈X(Ω).\nAs before, we see that there exists a unique solution u= (u1,u2)∈X(Ω)\nsatisfying B(u,ϕ) = Λ(ϕ) for all ϕ∈X(Ω). Moreover, setting vj:=−fj, the\nvectorU:= (u1,u2,v1,v2)⊤belongs to D(A) and satisfies −AU=F.\nOn the other hand, if /tildewideU∈D(A) solves−A/tildewideU=F, thenB(/tildewideu,ϕ) = Λ(ϕ) holds for\nallϕ∈X(Ω) due to the definition of D(A) and the weak transmission conditions.\nTherefore, U=/tildewideU, andA:D(A)→ His a bijection. Since Ais the generator of a\nC0-semigroup, Ais closed and the continuity of A−1:H → Hfollows. Therefore,\n0∈ρ(A).\nLemma 2.4. a) The domain of Ais given by\nD(A) =/braceleftbig\n(u1,u2,v1,v2)∈/parenleftbig\nH4(Ω1)×H4(Ω2)/parenrightbig\n∩X(Ω)×X(Ω) :\n∆u1= ∆u2onΓ,−ρ∂νv1+∂ν∆u1=∂ν∆u2onΓ/bracerightbig\n.(2-8)\nHere, the equalities on Γcan be understood as equalities in the trace spaces H3/2(Γ)\nandH1/2(Γ), respectively.\nb) The operator Ahas compact resolvent and, consequently, discrete spectru m.\nProof.a) Let/tildewideU∈D(A) andF= (f1,f2,g1,g2)⊤:=−A/tildewideU∈ H. To show the\nstatement, we construct a strong solution Uof−AU=Fbelonging to the right-\nhand side of ( 2-8) and show that U=/tildewideU. So we consider\n∆2u1=g1−ρ∆f1, (2-9)\n∆2u2=g2 (2-10)\ninL2(Ω1)×L2(Ω2) with boundary conditions\nu1=∂νu1= 0 on Γ 1 (2-11)\nand transmission conditions\nu1−u2= 0, (2-12)\n∂νu1−∂νu2= 0, (2-13)\n∂2\nνu1−∂2\nνu2= 0, (2-14)\n∂3\nνu1−∂3\nνu2=−ρ∂νf1. (2-15)6 ROBERT DENK AND FELIX KAMMERLANDER\nConcerningthehigher-ordertransmissionconditions( 2-14)and(2-15), notethatfor\nall(u1,u2)∈H4(Ω1)×H4(Ω2)satisfying ( 2-12) and(2-13) alltangentialderivatives\nofu1−u2and∂νu1−∂νu2alongΓdisappear. Therefore, forsuch uthe transmission\nconditions ( 2-14)–(2-15) are equivalent to the conditions\n∆u1−∆u2= 0,\n∂ν∆u1−∂ν∆u2=−ρ∂νf1.\nDefine the operator B:L2(Ω)⊃D(B)→L2(Ω) byD(B) :=H4(Ω)∩H2\n0(Ω)\nandBw:= ∆2w. Then, Bis a selfadjoint operator with 0 ∈ρ(B). To construct\na strong solution of the transmission problem ( 2-9)–(2-15), we first eliminate the\ninhomogeneity on the right-hand side of ( 2-15). By [21], Section 4.7.1, p. 330, the\nmapping\nRh:=/parenleftbig\nh|∂Ω1,∂νh|∂Ω1,∂2\nνh|∂Ω1,∂3\nνh|∂Ω1/parenrightbig⊤\nis a retraction from H4(Ω1) onto/producttext3\nj=0H4−1/2−j(∂Ω1).Therefore, there exists a\nfunction h∈H4(Ω1) such that\nRh= (0,0,0,−χΓρ∂νf1)⊤.\nHere again χΓstands for the characteristic function of Γ. We define w:=B−1G∈\nH4(Ω)∩H2\n0(Ω),where\nG=χΩ1(g1−ρ∆f1−∆2h)+χΩ2g2∈L2(Ω).\nFinally, we set u1:=w|Ω1+handu2:=w|Ω2.Then,u= (u1,u2)∈H4(Ω1)×\nH4(Ω2) satisfies the strong transmission problem ( 2-9)–(2-15). Therefore, U:=\n(u1,u2,v1,v2)⊤withvj:=−fjbelongs to the right-hand side of ( 2-8) and solves\n−A(D)U=F.\nOn the other hand, using integration by parts and the fact that usolves the\nstrong transmission problem, we see that Usatisfies the weak transmission condi-\ntions (2-2). Therefore, Ubelongs to D(A) and solves −AU=F. By Remark 2.3,\nthis solution is unique which implies U=/tildewideU.\nb) Due to a), we have\nD(A)⊂/parenleftbig\nH4(Ω1)×H4(Ω2)/parenrightbig\n∩X(Ω)×X(Ω).\nBy the Rellich-Kondrachov theorem, the space on the right-hand s ide is compactly\nembedded into H. Therefore, A−1is compact, and the spectrum of Ais discrete.\n/square\nWe already know that the spectrum of Ais discrete and that 0 is no eigenvalue.\nIn fact, there are no purely imaginary eigenvalues of A, as the next result shows.\nTheorem 2.5. The imaginary axis is a subset of the resolvent set of A,i.e.iR⊂\nρ(A).\nProof.Assume that U= (u1,u2,v1,v2)⊤∈D(A) satisfies ( −iλ+A)U= 0 with\nλ∈R\\{0}. Thenvj= iλujforj= 1,2, and (u1,u2) satisfies\n−∆2u1+iλρ∆u1+λ2u1= 0 in Ω 1, (2-16)\n−∆2u2+λ2u2= 0 in Ω 2 (2-17)\nwith boundary conditions u1=∂νu1= 0 on Γ 1and transmission conditions\nu1=u2,\n∂νu1=∂νu2,\n∆u1= ∆u2,\n−iλρ∂νu1+∂ν∆u1=∂ν∆u2EXPONENTIAL STABILITY FOR A COUPLED SYSTEM 7\non the common interface Γ .\nWe will show that ( u1,u2) = 0.We multiply ( 2-16) and (2-17) withu1andu2,\nrespectively. Summing up and performing an integration by parts yie lds\n−/ba∇dbl∆u1/ba∇dbl2\nL2(Ω1)−iλρ/ba∇dbl∇u1/ba∇dbl2\nL2(Ω1)+λ2/ba∇dblu1/ba∇dbl2\nL2(Ω1)\n−/ba∇dbl∆u2/ba∇dbl2\nL2(Ω2)+λ2/ba∇dblu2/ba∇dbl2\nL2(Ω2)= 0.\nHere we have used the boundary conditions as well as the transmiss ion conditions\non Γ.Considering only the imaginary part we get /ba∇dbl∇u1/ba∇dblL2(Ω1)= 0.Together with\nu1|Γ1= 0 we obtain u1= 0.Therefore, u2satisfies the boundary value problem\n−∆2u2+λ2u2= 0 in Ω 2, (2-18)\nu2=∂νu2= ∆u2=∂ν∆u2= 0 on Γ = ∂Ω2. (2-19)\nBecause of ( 2-19), the trivial extension /tildewideu2by zero to Rnbelongs to H4(Rn) and\nsatisfies ∆2/tildewideu2=λ2/tildewideu2inRn. As ∆2inL2(Rn) has no eigenvalues, this implies\n/tildewideu2= 0 and therefore u2= 0. Altogether we have seen U= 0. /square\nThe last results already implies strong stability of the semigroup ( T(t))t≥0gen-\nerated by A, i.e., for any U0∈ Hwe have /ba∇dblT(t)U0/ba∇dblH→0 (t→ ∞) (see [5],\nTheorem 2.4). We will see in Section 4 that Tis even exponentially stable.\n3.A priori estimates for the damped plate equation\nFor the proof of exponential stability of the coupled damped–unda mped plate\nequation, we need some a priori estimates for the damped part. Fo r this, we\nwill apply the theory of interpolation-extrapolation scales due to Am ann (see [ 2],\nChapter V).\nThroughout this section, let U⊂Rnbe a bounded C4-domain. We define the\noperator Ain the space H2\n0(U)×L2(U) by\nD(A) := (H4(U)∩H2\n0(U))×H2\n0(U),\nA:=/parenleftbigg0 1\n−∆2ρ∆/parenrightbigg\n.(3-1)\nIt was shown in [ 8], Proposition 3.1 (see also [ 9], Theorem 5.1) that Agenerates an\nanalytic exponentially stable C0-semigroup in H2\n0(U)×L2(U). To extrapolate this\nresult to spaces of negative regularity, we need to determine the a djoint operator A′\nconsidered in the dual spaces. In the following, /a\\}b∇acketle{t·,·/a\\}b∇acket∇i}htX′×Xdenotes the dual pairing\nin a Banach space X. We begin with a small observation on the bi-Laplacian\noperator.\nRemark 3.1. Under the above assumptions on U, the operator ∆2:H2\n0(U)→\nH−2(U) is an isomorphism. In fact, we have the coercive estimate\n/a\\}b∇acketle{t∆2u,u/a\\}b∇acket∇i}htH−2(U)×H2\n0(U)=/ba∇dbl∆u/ba∇dbl2\nL2(U)≥C/ba∇dblu/ba∇dbl2\nH2(U)(u∈H2\n0(U)).\nHere the last inequality holds by elliptic regularity and invertibility of the Dirichlet\nLaplacian ∆ D:H2(U)∩H1\n0(U)→L2(U). Now an application of the Lax-Milgram\ntheorem yields the invertibility of ∆2:H2\n0(U)→H−2(U).\nLemma 3.2. The adjoint operator A′ofAis given by\nA′:H−2(U)×L2(U)⊃D(A′) :=L2(U)×H2\n0(U)→H−2(U)×L2(U),\nA′:=/parenleftbigg0−∆2\n1ρ∆/parenrightbigg\n.8 ROBERT DENK AND FELIX KAMMERLANDER\nProof.We define E:=H2\n0(U)×L2(U) and/tildewideD:=L2(U)×H2\n0(U)⊂E′,where\nE′=H−2(U)×L2(U).Then, for all v= (v1,v2)∈/tildewideDand\nu= (u1,u2)∈D(A) =/parenleftbig\nH4(U)∩H2\n0(U)/parenrightbig\n×H2\n0(U),\nintegration by parts and the definition of distributional derivatives yield\nv(Au) =/a\\}b∇acketle{tv1,u2/a\\}b∇acket∇i}htL2(U)+/a\\}b∇acketle{tv2,−∆2u1/a\\}b∇acket∇i}htL2(U)+/a\\}b∇acketle{tv2,ρ∆u2/a\\}b∇acket∇i}htL2(U)\n=/a\\}b∇acketle{tv1,u2/a\\}b∇acket∇i}htL2(U)+/a\\}b∇acketle{t−∆v2,∆u1/a\\}b∇acket∇i}htL2(U)+/a\\}b∇acketle{tρ∆v2,u2/a\\}b∇acket∇i}htL2(U)\n=/a\\}b∇acketle{t−∆2v2,u1/a\\}b∇acket∇i}htH−2(U)×H2\n0(U)+/a\\}b∇acketle{tv1,u2/a\\}b∇acket∇i}htL2(U)+/a\\}b∇acketle{tρ∆v2,u2/a\\}b∇acket∇i}htL2(U)\n=w1(u1)+w2(u2),\nwithw1:=−∆2v2∈H−2(U) andw2:=v1+ρ∆v2∈L2(U).Therefore, we set\n/tildewideA:=/parenleftbigg\n0−∆2\n1ρ∆/parenrightbigg\nwithD(/tildewideA) :=/tildewideD.With this definition, we have v(Au) = (/tildewideAv)(u) for allu∈D(A)\nand allv∈/tildewideD.Moreover, for all v∈/tildewideDthe mapping [ u/mapsto→v(Au)]:D(A)→Cis\ncontinuous with respect to /ba∇dbl·/ba∇dblE.Hence, we have /tildewideA⊂A′.\nLetu∈D(A) andv∈E′.Then\nv(Au) =/a\\}b∇acketle{tv1,u2/a\\}b∇acket∇i}htH−2(U)×H2\n0(U)+/a\\}b∇acketle{tv2,−∆2u1/a\\}b∇acket∇i}htL2(U)+/a\\}b∇acketle{tv2,ρ∆u2/a\\}b∇acket∇i}htL2(U).(3-2)\nNow, let v∈D(A′).Then, the mapping [ u/mapsto→v(Au)]:D(A)→Ccan be extended\nto a linear, continuous mapping from EtoC.In particular, considering\n|/a\\}b∇acketle{tv2,∆2u1/a\\}b∇acket∇i}htL2(U)|=|v(A(u1,0))| ≤C/ba∇dbl(u1,0)/ba∇dblE=C/ba∇dblu1/ba∇dblH2(U)\nforu1∈H4(U)∩H2\n0(U),it holds that\nϕ:H4(U)∩H2\n0(U)→C, u1/mapsto→ϕ(u1) :=/a\\}b∇acketle{tv2,∆2u1/a\\}b∇acket∇i}htL2(U)(3-3)\nis continuous with respect to /ba∇dbl·/ba∇dblH2(U).By Remark 3.1,\n∆2:H2\n0(U)→H−2(U) (3-4)\nis an isomorphism. Therefore, ( 3-3) and (3-4) imply that\n/bracketleftbig\n/tildewideu1/mapsto→ϕ/parenleftbig\n(∆2)−1/tildewideu1/parenrightbig\n=/a\\}b∇acketle{tv2,/tildewideu1/a\\}b∇acket∇i}htL2(U)/bracketrightbig\n:L2(U)→C\nis continuousconsideredasamapping from( L2(U),/ba∇dbl·/ba∇dblH−2(U)) toC.By the density\nofL2(U)⊂H−2(U), there exists a unique continuous extension\n/tildewideϕ∈/parenleftbig\nH−2(U)/parenrightbig′=H2\n0(U)\nof this mapping. Together with\n/a\\}b∇acketle{t/tildewideϕ,/tildewideu1/a\\}b∇acket∇i}htH2\n0(U)×H−2(U)=/a\\}b∇acketle{tv2,/tildewideu1/a\\}b∇acket∇i}htH2\n0(U)×H−2(U)\nfor/tildewideu1∈L2(U),we deduce v2=/tildewideϕ∈H2\n0(U).\nThe fact that v2∈H2\n0(U) implies that the last term in ( 3-2),\n/bracketleftig\nu2/mapsto→ /a\\}b∇acketle{tv2,ρ∆u2/a\\}b∇acket∇i}htL2(U)=/a\\}b∇acketle{tv2,ρ∆u2/a\\}b∇acket∇i}htH2\n0(U)×H−2(U)/bracketrightig\n:H2\n0(U)→C,\nis continuous on L2(U). Since ( 3-2) needs to be continuous, by setting u1= 0 it\nfollows that also the first term/bracketleftig\nu2/mapsto→ /a\\}b∇acketle{tv1,u2/a\\}b∇acket∇i}htH−2(U)×H2\n0(U)/bracketrightig\n:H2\n0(U)→C\ncan be extended continuously to L2(U),which means v1∈L2(U).\nWe have shown that v∈D(A′) implies v2∈H2\n0(U) andv1∈L2(U),i.e.v∈/tildewideD.\nHence, we obtain /tildewideD=D(A′) and therefore /tildewideA=A′. /squareEXPONENTIAL STABILITY FOR A COUPLED SYSTEM 9\nIn the following,\nΣϕ={z∈C\\{0}:|arg(z)|< ϕ}\ndenotes the open sector in C.\nTheorem 3.3. There exists a constant C0>0such that for any λ∈ρ(A)⊃\nΣπ/2\\{0}and any F∈H2\n0(U)×L2(U)the unique solution u= (u1,v1)∈D(A)of\n(λ−A)u=F∈H2\n0(U)×L2(U) (3-5)\nsatisfies the estimate\n/ba∇dblu/ba∇dblH2+θ(U)×Hθ(U)≤C0/ba∇dblF/ba∇dblHθ(U)×H−2+θ(U)(θ∈[0,2]). (3-6)\nIn particular, for u= (u1,v1)∈D(A)solving\n(λ−A)u=/parenleftbigg0\nf/parenrightbigg\nwithf∈L2(U)we obtain the estimate\n/ba∇dblu1/ba∇dblH2+θ(U)≤C0/ba∇dblf/ba∇dblH−2+θ(U)(θ∈[0,2]). (3-7)\nProof.By [8], Proposition 3.1, Ais the generator of an analytic, exponentially sta-\nble, strongly continuous semigroup on H2\n0(U)×L2(U).Therefore, ( 3-5) is uniquely\nsolvable, and we have the uniform resolvent estimate\n/ba∇dblu/ba∇dblH4(U)×H2(U)≤C1/ba∇dblF/ba∇dblH2(U)×L2(U) (3-8)\nwith some constant C1>0 independent of Fandλ.\nLetA♯:=A′be the adjoint operator of Aand set\nE0:=H2\n0(U)×L2(U),\nE1:=D(A) =/parenleftbig\nH4(U)∩H2\n0(U)/parenrightbig\n×H2\n0(U),\nE♯\n0:=E′\n0=H−2(U)×L2(U),\nE♯\n1:=D(A♯).\nObviously, E0is reflexive and E1is dense in E0.SinceAis the generator of an\nanalytic C0-semigroup on E0with domain E1, in symbols A∈ H(E1,E0),by [2],\np. 13, Proposition 1.2.3, the same holds true for A♯onE♯\n0with domain E♯\n1,i.e.\nA♯∈ H(E♯\n1,E♯\n0).\nHence, we can define the interpolation-extrapolation scales {(Aα,Eα) :α∈R}and\nits dual scale {(A♯\nα,E♯\nα) :α∈R}.Then, Theorem 1.5.12 in [ 2] states that Eαis\nreflexive and we have\n(Eα)′=E♯\n−αand (Aα)′=A♯\n−α\nfor allα∈R.Moreover, by [ 1], Theorem 6.1 and [ 2], Theorem 2.1.3it holds that Aα\nandA♯\nαare generators of analytic C0-semigroups in Eαwith domain Eα+1andE♯\nα\nwith domain E♯\nα+1for allα∈R,respectively. Again, we write Aα∈ H(Eα+1,Eα)\nandA♯\nα∈ H(E♯\nα+1,E♯\nα).\nInparticular, A−1isthegeneratorofananalytic C0-semigroupon E−1with domain\nE0.By [2], Theorem 2.1.3, λ−A−1is an isomorphism from E0toE−1and we have\n/ba∇dbl(µ−A−1)−1/ba∇dblL(E−1,E0)≤C/ba∇dbl(µ−A)−1/ba∇dblL(E0,E1)≤C′\nfor allµ∈ρ(A) with a constant C′independent of µ.By Lemma 3.2, the space\nE−1equals\nE−1= (E−1)′′=/parenleftig\nE♯\n1/parenrightig′\n= (D(A′))′=/parenleftbig\nL2(U)×H2\n0(U)/parenrightbig′=L2(U)×H−2(U).10 ROBERT DENK AND FELIX KAMMERLANDER\nTherefore, there exists a constant C2>0 such that\n/ba∇dblu/ba∇dblH2(U)×L2(U)≤C2/ba∇dblF/ba∇dblL2(U)×H−2(U). (3-9)\nNow the inequality ( 3-6) follows by (real) interpolation between ( 3-8) and (3-9)\nwithC0:= max{C1,C2}.\nConsidering the particular case F=/parenleftbig0\nf/parenrightbig\nand only the first component of u, we\nobtain (3-7). /square\n4.Exponential stability\nIn this section, we continue the analysis of the coupled system ( 1-1)–(1-2). We\nwill estimates the resolvent ( A−iλ)−1of the corresponding first-order system on\nthe imaginaryaxisfor λ∈Rwith|λ|large. Bya result due to Pr¨ uss([ 19], Corollary\n4), uniform boundedness of the resolvent on the imaginary axis implie s exponential\nstability of the semigrroup.\nWe start with some identities which will be useful for our estimates. I n the\nfollowing, we will shortly write xfor the identity function x/mapsto→x. For vectors\ny,z∈Cnwe sety·z:=/summationtextn\nj=1yjzj(note that this is not the scalar product in Cn).\nLemma 4.1. LetU⊂Rnbe aC4-domain, let w∈H4(U), and let ν:∂U→Rn\nbe the outer unit normal vector. Then,\n2Re/integraldisplay\nU(x·∇w)∆2wdx= (4−n)/ba∇dbl∆w/ba∇dbl2\nL2(U)+/integraldisplay\n∂U(x·ν)|∆w|2dS\n+2Re/integraldisplay\n∂U/bracketleftig\n(x·∇w)∂ν∆w−∆w∂ν(x·∇w)/bracketrightig\ndS.\nProof.This follows by straightforward calculation from the divergence the orem,\napplied to the vector field\nV:=|∆w|2x+2(x·∇w)∇∆w−2∆w∇(x·∇w).\nNote that\ndivV=n|∆w|2+x·∇|∆w|2+2(x·∇w)∆2w−2∆w∆(x·∇w)\nand Re(div V) = 2Re( x·∇w)∆2w+(n−4)|∆w|2. A more general variant of the\nstatement is also known as Rellich’s identity, see, e.g., [ 15], Proposition 2.2, or [ 14],\np. 238. /square\nLemma 4.2. LetU⊂Rnbe aC4-domain, and let w∈H4(U)be a solution of\n−∆2w+λ2w=zwithλ∈Randz∈L2(U). Then we have\nnλ2/ba∇dblw/ba∇dbl2\nL2(U)+(4−n)/ba∇dbl∆w/ba∇dbl2\nL2(U)+/integraldisplay\n∂U(x·ν)|∆w|2dS\n=−2Re/integraldisplay\nU(x·∇w)zdx+λ2/integraldisplay\n∂U(x·ν)|w|2dS\n−2Re/integraldisplay\n∂U/bracketleftig\n(x·∇w)∂ν∆w−∆w∂ν(x·∇w)/bracketrightig\ndS.\nProof.Applying the divergence theorem to the vector field |w|2xand taking the\nreal part, we obtain\n2Re/integraldisplay\nU(x·∇w)wdx=−n/ba∇dblw/ba∇dbl2\nL2(U)+/integraldisplay\n∂U(x·ν)|w|2dS.\nFrom this and ∆2w=λ2w−zwe get\n2Re/integraldisplay\nU(x·∇w)∆2wdx=−2Re/integraldisplay\nU(x·∇w)zdx−nλ2/ba∇dblw/ba∇dbl2\nL2(U)EXPONENTIAL STABILITY FOR A COUPLED SYSTEM 11\n+λ2/integraldisplay\n∂U(x·ν)|w|2dS.\nPlugging this into the statement of Lemma 4.1, the assertion follows. /square\nThe following result can be found, e.g., in [ 14], Proof of Theorem 2.2.\nLemma 4.3. LetU⊂Rnbe aC3-domain, and let S⊂∂Ube a nontrivial part\nof the boundary. Then, for every w∈H3(U)withw=∂νw= 0onSwe have\n∂ν(x·∇w) = (x·ν)∆wonS.\nIn the next step, we considerthe resolventequation ( −iλ+A)U=Ffor a partic-\nular right-hand side F= (0,0,0,g2)⊤with inhomogeneous transmission conditions.\nMore precisely, we consider\n−iλu1+v1= 0 in Ω 1, (4-1)\n−∆2u1+ρ∆v1−iλv1= 0 in Ω 1, (4-2)\n−iλu2+v2= 0 in Ω 2, (4-3)\n−∆2u2−iλv2=g2in Ω2 (4-4)\nwith transmission conditions\n∆u1= ∆u2,\n−iλρ∂νu1+∂ν∆u1=∂ν∆u2+iλρ∂νw1/bracerightbigg\non Γ. (4-5)\nThefollowingaprioriestimatewillbethecrucialstepfortheproofo fexponential\nstability.\nProposition 4.4. Letw1∈H4(Ω1)andg2∈L2(Ω2)be given. Then, there exists\nλ0>0and a constant C >0(only depending on n,ρ,δ0andλ0) such that for\nany solution U= (u1,u2,v1,v2)⊤∈X(Ω)×X(Ω)withui∈H4(Ωi)fori= 1,2of\n(4-1)–(4-5)the estimate\n/ba∇dblU/ba∇dblH≤C/parenleftbig\n/ba∇dblg2/ba∇dblL2(Ω2)+|λ|/ba∇dbl∂νw1/ba∇dblL2(Γ)/parenrightbig\n(λ∈R,|λ|> λ0)\nholds.\nIn the following proof, we will use a generic constant Cindependent of λ,U,and\nF. Moreover, an estimate of the form /ba∇dbl·/ba∇dbl ≤ε/ba∇dbl·/ba∇dbl1+Cε/ba∇dbl·/ba∇dbl2has to be understood\nin the sense that for every small ε >0 there exists a constant Cε>0 such that\nthe inequality holds. Again Cεdenotes a generic constant. Note that all constants\nmay depend on ρ.\nProof.We have to estimate\n/ba∇dblU/ba∇dblH=/parenleftig\n/ba∇dbl∆u1/ba∇dbl2\nL2(Ω1)+/ba∇dbl∆u2/ba∇dbl2\nL2(Ω2)+/ba∇dblv1/ba∇dbl2\nL2(Ω1)+/ba∇dblv2/ba∇dbl2\nL2(Ω2)/parenrightig1/2\n.\nThe proof is done in several steps.\n(i)Estimate of v1.Letλ∈Rwith|λ| ≫1 andU= (u1,u2,v1,v2)∈X(Ω)×\nX(Ω) be a solution of ( 4-1)–(4-5). Hence, ( u1,u2) is a solution of\n−∆2u1+iλρ∆u1+λ2u1= 0, (4-6)\n−∆2u2+λ2u2=g2 (4-7)\nin Ω1×Ω2satisfying the transmission conditions ( 4-5). By the definition of X(Ω),\nwe have u1=∂νu1= 0 on Γ 1. In order to show the assertion of the theorem, we\nneed to establish an estimate of the form\n/ba∇dblU/ba∇dbl2\nH≤ε/ba∇dblU/ba∇dbl2\nH+Cε/parenleftig\n/ba∇dblg2/ba∇dbl2\nL2(Ω2)+|λ|2/ba∇dbl∂νw1/ba∇dbl2\nL2(Γ)/parenrightig12 ROBERT DENK AND FELIX KAMMERLANDER\nSimilar to the proof of the dissipativity of Ain Theorem 2.2, we obtain\nRe/a\\}b∇acketle{tF,U/a\\}b∇acket∇i}htH= Re/a\\}b∇acketle{tAU,U/a\\}b∇acket∇i}htH=−ρ/ba∇dbl∇v1/ba∇dbl2\nL2(Ω1)−Re/integraldisplay\nΓiλρv1∂νw1dS.\nTherefore, Poincar´ e and Young’s inequality yield\n/ba∇dblv1/ba∇dbl2\nH1(Ω1)≤C/parenleftbig\n/ba∇dblg2/ba∇dblL2(Ω2)/ba∇dblU/ba∇dblH+|λ|/ba∇dbl∂νw1/ba∇dblL2(Γ)/ba∇dblv1/ba∇dblH1(Ω1)/parenrightbig\n≤ε/parenleftig\n/ba∇dblU/ba∇dbl2\nH+/ba∇dblv1/ba∇dbl2\nH1(Ω1)/parenrightig\n+Cε/parenleftig\n/ba∇dblg2/ba∇dbl2\nL2(Ω2)+|λ|2/ba∇dbl∂νw1/ba∇dbl2\nL2(Γ)/parenrightig\n,\nthat is\n/ba∇dblv1/ba∇dbl2\nH1(Ω1)≤ε/ba∇dblU/ba∇dbl2\nH+Cε/parenleftig\n/ba∇dblg2/ba∇dbl2\nL2(Ω2)+|λ|2/ba∇dbl∂νw1/ba∇dbl2\nL2(Γ)/parenrightig\n.(4-8)\nTogether with ( 4-1) this implies\n/ba∇dblu1/ba∇dbl2\nH1(Ω1)≤ε|λ|−2/ba∇dblU/ba∇dbl2\nH+Cε/parenleftig\n|λ|−2/ba∇dblg2/ba∇dbl2\nL2(Ω2)+/ba∇dbl∂νw1/ba∇dbl2\nL2(Γ)/parenrightig\n.(4-9)\n(ii)Estimate of ∆u1and∆u2.We multiply ( 4-6) by−u1and (4-7) by−u2.\nIntegration by parts and summing up yields\n/ba∇dbl∆u1/ba∇dbl2\nL2(Ω1)+/ba∇dbl∆u2/ba∇dbl2\nL2(Ω2)+iλρ/ba∇dbl∇u1/ba∇dbl2\nL2(Ω2)\n=λ2/parenleftig\n/ba∇dblu1/ba∇dbl2\nL2(Ω1)+/ba∇dblu2/ba∇dbl2\nL2(Ω2)/parenrightig\n−/a\\}b∇acketle{tg2,u2/a\\}b∇acket∇i}htL2(Ω2)+iλρ/integraldisplay\nΓu1∂νw1dS,\nwhere we have used the transmission conditions ( 4-5) andu1=∂νu1= 0 on Γ 1.\nTaking the real part and observing vj=iλuj, we see that\n/ba∇dbl∆u1/ba∇dbl2\nL2(Ω1)+/ba∇dbl∆u2/ba∇dbl2\nL2(Ω2)≤ |λ|2/parenleftbig\n/ba∇dblu1/ba∇dbl2\nL2(Ω1)+/ba∇dblu2/ba∇dbl2\nL2(Ω2)/parenrightbig\n+/ba∇dblg2/ba∇dblL2(Ω2)/ba∇dblu2/ba∇dblL2(Ω2)+|λ|ρ/ba∇dblu1/ba∇dblL2(Γ)/ba∇dbl∂νw1/ba∇dblL2(Γ)\n=/parenleftbig\n/ba∇dblv1/ba∇dbl2\nL2(Ω1)+/ba∇dblv2/ba∇dbl2\nL2(Ω2)/parenrightbig\n+/ba∇dblg2/ba∇dblL2(Ω2)/ba∇dblu2/ba∇dblL2(Ω2)+ρ/ba∇dblv1/ba∇dblL2(Γ)/ba∇dbl∂νw1/ba∇dblL2(Γ).(4-10)\nAssuming |λ| ≥1, we get with the trace theorem and Young’s inequality\n/ba∇dblg2/ba∇dblL2(Ω2)/ba∇dblu2/ba∇dblL2(Ω2)≤1\n2/ba∇dblg2/ba∇dbl2\nL2(Ω1)+1\n2/ba∇dblv2/ba∇dbl2\nL2(Ω2),\nρ/ba∇dblv1/ba∇dblL2(Γ)/ba∇dbl∂νw1/ba∇dblL2(Γ)≤ρ\n2/ba∇dblv1/ba∇dbl2\nH1(Ω1)+ρ\n2/ba∇dbl∂νw1/ba∇dbl2\nL2(Γ).\nInserting this into ( 4-10) yields\n/ba∇dbl∆u1/ba∇dbl2\nL2(Ω1)+/ba∇dbl∆u2/ba∇dbl2\nL2(Ω2)≤(1+ρ\n2)/ba∇dblv1/ba∇dbl2\nH1(Ω1)+/ba∇dblv2/ba∇dbl2\nL2(Ω2)\n+/ba∇dblg2/ba∇dbl2\nL2(Ω1)+ρ\n2/ba∇dbl∂νw1/ba∇dbl2\nL2(Γ)\n≤ε/ba∇dblU/ba∇dbl2\nH+C/ba∇dblv2/ba∇dbl2\nL2(Ω2)\n+Cε/parenleftbig\n/ba∇dblg2/ba∇dbl2\nL2(Ω2)+|λ|2/ba∇dbl∂νw1/ba∇dbl2\nL2(Γ)/parenrightbig\n. (4-11)\nHere, in the last step we estimated /ba∇dblv1/ba∇dblH1(Ω1)due to (4-8).\n(iii)Estimate of v2.We apply Lemma 4.2inU= Ω2withw=u2andz=g2\nand obtain, noting iλu2=v2,\nn/ba∇dblv2/ba∇dbl2\nL2(Ω2)=−(4−n)/ba∇dbl∆u2/ba∇dbl2\nL2(Ω2)−/integraldisplay\nΓ(x·ν)|∆u2|2dS\n−2Re/integraldisplay\nΩ2(x·∇u2)g2dx+λ2/integraldisplay\nΓ(x·ν)|u2|2dS\n−2Re/integraldisplay\nΓ/bracketleftig\n(x·∇u2)∂ν∆u2−∆u2∂ν(x·∇u2)/bracketrightig\ndS. (4-12)\nIn the same way, we apply Lemma 4.2inU= Ω1withw=u1andz=−ρ∆v1.\nHere we remark that −∆2u1+λ2u1=−ρ∆v1by (4-1) and (4-2). Moreover, theEXPONENTIAL STABILITY FOR A COUPLED SYSTEM 13\nnormal vector νis the outer normal at the part Γ 1of the boundary ∂Ω1, butνis\nthe inner normal at the part Γ of ∂Ω1. We obtain\nn/ba∇dblv1/ba∇dbl2\nL2(Ω1)=−(4−n)/ba∇dbl∆u1/ba∇dbl2\nL2(Ω1)−/integraldisplay\nΓ1(x·ν)|∆u1|2dx\n+/integraldisplay\nΓ(x·ν)|∆u1|2dx+2Re/integraldisplay\nΩ1(x·∇u1)ρ∆v1dx\n+λ2/integraldisplay\nΓ1(x·ν)|u1|2dS−λ2/integraldisplay\nΓ(x·ν)|u1|2dS\n−2Re/integraldisplay\nΓ1/bracketleftig\n(x·∇u1)∂ν∆u1−∆u1∂ν(x·∇u1)/bracketrightig\ndS\n+2Re/integraldisplay\nΓ/bracketleftig\n(x·∇u1)∂ν∆u1−∆u1∂ν(x·∇u1)/bracketrightig\ndS. (4-13)\nDue to the condition ( u1,u2)∈X(Ω) and the transmission conditions ( 4-5), we\nhave\nu1=u2,∇u1=∇u2,∆u1= ∆u2, ∂ν∆u2=∂ν∆u1−iλρ∂ν(u1+w1) on Γ.\n(4-14)\nLet/tildewideu2∈H4(Ω) be a regular extension of u2to Ω, and define ϕ:=u1−/tildewideu2|Ω1∈\nH4(Ω1). Thenϕ=∂νϕ= 0 on Γ, and an application of Lemma 4.3yields\n∂ν(x·∇ϕ) = (x·ν)∆ϕ= (x·ν)(∆u1−∆u2) = 0 on Γ\nwhich gives\n∂ν(x·∇u1) =∂ν(x·∇u2) on Γ. (4-15)\nMoreover, with Lemma 4.3again we get\nu1= 0,∇u1= 0, ∂ν(x·∇u1) = (x·ν)∆u1on Γ1. (4-16)\nAdding ( 4-12) and (4-13) and taking into account ( 4-14)–(4-16), we obtain\nn/parenleftbig\n/ba∇dblv1/ba∇dbl2\nL2(Ω1)+/ba∇dblv2/ba∇dbl2\nL2(Ω2)/parenrightbig\n=−(4−n)/parenleftbig\n/ba∇dbl∆u1/ba∇dbl2\nL2(Ω1)+/ba∇dbl∆u2/ba∇dbl2\nL2(Ω2)/parenrightbig\n+2Re/bracketleftig\niλρ/integraldisplay\nΩ1(x·∇u1)∆u1dx/bracketrightig\n−2Re/bracketleftig/integraldisplay\nΩ2(x·∇u2)g2dx/bracketrightig\n+2Re/bracketleftbig\niλρ/integraldisplay\nΓ(x·∇u1)∂ν(u1+w1)dS/bracketrightig\n+/integraldisplay\nΓ1(x·ν)|∆u1|2dS.\nTherefore,\n/ba∇dblv2/ba∇dbl2\nL2(Ω2)≤C/bracketleftig\n|λ|/ba∇dblu1/ba∇dblH1(Ω1)/ba∇dbl∆u1/ba∇dblL2(Ω1)+/ba∇dbl∇u2/ba∇dblL2(Ω2)/ba∇dblg2/ba∇dblL2(Ω2)\n+|λ|/ba∇dblu1/ba∇dbl2\nH1(Γ)+|λ|/ba∇dblu1/ba∇dblH1(Γ)/ba∇dbl∂νw1/ba∇dblL2(Γ)+/ba∇dblu1/ba∇dbl2\nH2(Γ1)/bracketrightig\n.(4-17)\nWe estimate the first four terms on the right-hand side of ( 4-17) while the last term\nwill be treated in part (iv) of this proof.\nThe first term in ( 4-17) can be estimated by ( 4-8),/ba∇dbl∆u1/ba∇dblL2(Ω1)≤ /ba∇dblU/ba∇dblHand\nYoung’s inequality. We obtain\n|λ|/ba∇dblu1/ba∇dblH1(Ω1)/ba∇dbl∆u1/ba∇dblL2(Ω1)≤/parenleftig\nε/ba∇dblU/ba∇dblH+Cε/parenleftbig\n/ba∇dblg2/ba∇dblL2(Ω2)+|λ|/ba∇dbl∂νw1/ba∇dblL2(Γ)/parenrightbig/parenrightig\n/ba∇dblU/ba∇dblH\n≤ε/ba∇dblU/ba∇dbl2\nH+Cε/parenleftbig\n/ba∇dblg2/ba∇dblL2(Ω2)+|λ|/ba∇dbl∂νw1/ba∇dblL2(Γ)/parenrightbig2.(4-18)\nFor the second term in ( 4-17), we apply Green’s formula, using u1=u2and\n∂νu1=∂νu2on Γ to see that\n/ba∇dbl∇u2/ba∇dbl2\nL2(Ω2)≤ /ba∇dblu2/ba∇dblL2(Ω2)/ba∇dbl∆u2/ba∇dblL2(Ω2)+/ba∇dblu1/ba∇dblL2(Γ)/ba∇dbl∇u1/ba∇dblL2(Γ)\n≤1\n2/ba∇dblu2/ba∇dbl2\nL2(Ω2)+1\n2/ba∇dbl∆u2/ba∇dbl2\nL2(Ω2)+C/ba∇dblu1/ba∇dbl2\nH2(Ω1)14 ROBERT DENK AND FELIX KAMMERLANDER\n≤C/parenleftbig\n/ba∇dblu1/ba∇dbl2\nH2(Ω1)+/ba∇dblu2/ba∇dbl2\nH2(Ω2)/parenrightbig\n≤C/ba∇dblU/ba∇dbl2\nH.\nFor the last inequality, we have applied Remark 2.1b). Therefore, the second term\nin (4-17) can be estimated by\n/ba∇dbl∇u2/ba∇dblL2(Ω2)/ba∇dblg2/ba∇dblL2(Ω2)≤ε/ba∇dblU/ba∇dbl2\nH+Cε/ba∇dblg2/ba∇dbl2\nL2(Ω2).\nFor the third term in ( 4-17) we use interpolation to see that\n|λ|/ba∇dblu1/ba∇dbl2\nH1(Γ)≤C|λ|/ba∇dblu1/ba∇dbl2\nH3/2(Ω1)≤C|λ|/ba∇dblu1/ba∇dblH1(Ω1)/ba∇dblu1/ba∇dblH2(Ω1)\n≤ε/ba∇dblU/ba∇dbl2\nH+Cε/parenleftbig\n/ba∇dblg2/ba∇dbl2\nL2(Ω2)+|λ|2/ba∇dbl∂νw1/ba∇dbl2\nL2(Γ)/parenrightbig\n.(4-19)\nSimilarly, for the fourth term in ( 4-17) we write\n|λ|/ba∇dblu1/ba∇dblH1(Γ)/ba∇dbl∂νw1/ba∇dblL2(Γ)≤1\n2/ba∇dblu1/ba∇dbl2\nH1(Γ)+1\n2|λ|2/ba∇dbl∂νw1/ba∇dbl2\nL2(Γ).\nAs we have |λ| ≥1, this again can be estimated by the right-hand side of ( 4-19).\nAltogether, we obtain\n/ba∇dblv2/ba∇dbl2\nL2(Ω2)≤ε/ba∇dblU/ba∇dbl2\nH+Cε/parenleftbig\n/ba∇dblg2/ba∇dbl2\nL2(Ω2)+|λ|2/ba∇dbl∂νw1/ba∇dbl2\nL2(Γ)/parenrightbig\n+C/ba∇dblu1/ba∇dbl2\nH2(Γ1).(4-20)\n(iv)Estimate of u1onΓ1.The only term still left is /ba∇dblu1/ba∇dblH2(Γ1). We introduce\na cut-off function χ∈C∞(Ω1),0≤χ≤1,satisfying χ= 1 in a neighbourhood of\nΓ1andχ= 0 in a neighbourhood of the transmission interface Γ .Now, set\nz1:=χu1, z2:= iλz1, z:= (z1,z2)⊤.\nThen, since u1is a solution of ( 4-6),zsatisfies\n(−iλ+A)z=/parenleftbigg0\n/tildewidef/parenrightbigg\n,\nwhereAis defined as in ( 3-1) and\n/tildewidef= (−∆2χ)u1−2(∇∆χ)·∇u1−∆χ∆u1−2∆(∇χ·∇u1)\n−∆χ∆u1−2∇χ·∇∆u1+iλρ((∆χ)u1+2∇χ·∇u1)∈L2(Ω1)\n=B3(D,χ)u1+iλρ((∆χ)u1+2∇χ·∇u1)\nwith aλ-independent differential operator B3(D,χ) of order3 with coefficients only\nconsisting of derivatives of the C∞-function χ.Hence,\nB3(D,χ)∈L(H3/2(Ω1),H−3/2(Ω1)).\nFrom Theorem 3.3withθ=1\n2, we obtain\n/ba∇dblu1/ba∇dblH2(Γ1)=/ba∇dblz1/ba∇dblH2(Γ1)≤C/ba∇dblz1/ba∇dblH5/2(Ω1)≤C/ba∇dbl/tildewidef/ba∇dblH−3/2(Ω1)\n≤C/parenleftbig\n/ba∇dblB3(D,χ)u1/ba∇dblH−3/2(Ω1)+|λ|/ba∇dblu1/ba∇dblH1(Ω1)/parenrightbig\n≤C/parenleftbig\n/ba∇dblu1/ba∇dblH3/2(Ω1)+|λ|/ba∇dblu1/ba∇dblH1(Ω1)/parenrightbig\n≤C/parenleftbig\n|λ|1/2/ba∇dblu1/ba∇dblH3/2(Ω1)+/ba∇dblv1/ba∇dblH1(Ω1)/parenrightbig\n.\nNow, (4-8) and (4-19) yield\n/ba∇dblu1/ba∇dbl2\nH2(Γ1)≤ε/ba∇dblU/ba∇dbl2\nH+Cε/parenleftbig\n/ba∇dblg2/ba∇dbl2\nL2(Ω2)+|λ|2/ba∇dbl∂νw1/ba∇dbl2\nL2(Γ)/parenrightbig\n.(4-21)\nThe assertion of the Proposition now follows from ( 4-8), (4-11), (4-20), and\n(4-21). /squareEXPONENTIAL STABILITY FOR A COUPLED SYSTEM 15\nTheorem 4.5. There exists a constant C=C(ρ)>0such that\n/ba∇dbl(−iλ+A)−1/ba∇dblL(H)≤C(λ∈R\\{0},|λ|> λ0)\nfor some λ0>0.Consequently, the C0-semigroup (T(t))t≥0generated by Ais\nexponentially stable, i.e. there exist constants M >0andκ >0such that\n/ba∇dblT(t)U0/ba∇dblH≤Me−κt/ba∇dblU0/ba∇dblH(t≥0)\nholds for all U0∈ H.\nProof.Letλ∈Rwith|λ| ≫1 and let F= (f1,f2,g1,g2)∈ H.Furthermore, let\nU= (u1,u2,v1,v2)∈D(A) be the unique solution of\n(−iλ+A)U=F,\ni.e.Usatisfies\n−iλu1+v1=f1in Ω1,\n−∆2u1+ρ∆v1−iλv1=f2in Ω1,\n−iλu2+v2=g1in Ω2,\n−∆2u2−iλv2=g2in Ω2.\nIn order to show the assertion, we will subtract the solution Wof a structurally\ndamped plate equation with clamped boundary conditions on the whole domain\nΩ fromU.For this difference we will be able to use the a-priori estimate from\nProposition 4.4, whereas for Wan appropriate estimate is known.\nRecall the definition of the operator Afrom (3-1) and define\n/tildewiderW= (w,z)∈D(A) =/parenleftbig\nH4(Ω)∩H2\n0(Ω)/parenrightbig\n×H2\n0(Ω)\nby\n/tildewiderW:= (−iλ+A)−1/parenleftbiggχ1f1+χ2f2\nχ1g1+χ2g2/parenrightbigg\n,\nwhereχiis the characteristic function on Ω ifori= 1,2.Sinceχ1f1+χ2f2∈H2\n0(Ω)\nandχ1g1+χ2g2∈L2(Ω) due to the definition of H,by Theorem 3.3,/tildewiderWis well-\ndefined. In the following, we denote the restrictions of the compon ents of/tildewiderWby\nwi:=w|Ωiandzi:=z|Ωifori= 1,2.Finally, we set\nW:= (w1,w2,z1,z2)∈X(Ω)×X(Ω).\nNote that ui∈H4(Ωi) fori= 1,2.With this definitions, we obtain that the\ndifference U−Wsatisfies\n(−iλ+A)(U−W) =F−(−iλ+A)W= (0,0,0,/tildewideg2)⊤(4-22)\nwith/tildewideg2=ρ∆z2∈L2(Ω2),subject to the transmission conditions\n/braceleftigg\n∆(u1−w1) = ∆(u2−w2),\n−iλρ∂ν(u1−w1)+∂ν∆(u1−w1) =∂ν∆(u2−w2)+iλρ∂νw1.\nThanks to Proposition 4.4, we have\n/ba∇dblU−W/ba∇dblH≤C/parenleftbig\n/ba∇dbl/tildewideg2/ba∇dblL2(Ω2)+|λ|/ba∇dbl∂νw1/ba∇dblL2(Γ)/parenrightbig\n. (4-23)\nAn application of Theorem 3.3withU= Ω and θ= 2 gives\n/ba∇dbl/tildewideg2/ba∇dblL2(Ω2)=ρ/ba∇dbl∆z2/ba∇dblL2(Ω2)≤C/ba∇dblz2/ba∇dblH2(Ω2)≤C/ba∇dbl/tildewiderW/ba∇dblH4(Ω)×H2(Ω)\n≤C/vextenddouble/vextenddouble/vextenddouble/parenleftbiggχ1f1+χ2f2\nχ1g1+χ2g2/parenrightbigg/vextenddouble/vextenddouble/vextenddouble\nH2(Ω)×L2(Ω)≤C/ba∇dblF/ba∇dblH.16 ROBERT DENK AND FELIX KAMMERLANDER\nSinceAis the generator of a bounded, analytic C0-semigroup on H2\n0(Ω)×L2(Ω)\nby Theorem 3.3, we see that\n|λ|/ba∇dbl∂νw1/ba∇dblL2(Γ)≤C|λ|/ba∇dblw1/ba∇dblH2(Ω1)≤C/ba∇dblF/ba∇dblH.\nTherefore, ( 4-23) yields\n/ba∇dblU−W/ba∇dblH≤C/ba∇dblF/ba∇dblH.\nInvoking Theorem 3.3again, we deduce\n/ba∇dblU/ba∇dblH≤ /ba∇dblU−W/ba∇dblH+/ba∇dblW/ba∇dblH≤C/ba∇dblF/ba∇dblH\nwith a constant C=C(ρ)>0.This proves the theorem. /square\nReferences\n[1] H. Amann. Parabolic evolution equations in interpolati on and extrapolation spaces. J. Funct.\nAnal., 78(2):233–270, 1988.\n[2] H. Amann. Linear and quasilinear parabolic problems. Vol. I , volume 89 of Monographs in\nMathematics . Birkh¨ auser Boston, Inc., Boston, MA, 1995.\n[3] K. Ammariand S. Nicaise. Stabilization of a transmissio nwave/plate equation. J. Differential\nEquations , 249(3):707–727, 2010.\n[4] K. Ammari and G. Vodev. Boundary stabilization of the tra nsmission problem for the\nBernoulli-Euler plate equation. Cubo, 11(5):39–49, 2009.\n[5] W.Arendtand C.J.K.Batty. Tauberian theorems and stabi lity ofone-parameter semigroups.\nTrans. Amer. Math. Soc. , 306(2):837–852, 1988.\n[6] G. Avalos. The exponential stability of a coupled hyperb olic/parabolic system arising in\nstructural acoustics. Abstr. Appl. Anal. , 1(2):203–217, 1996.\n[7] G. Avalos and I. Lasiecka. The strong stability of a semig roup arising from a coupled hyper-\nbolic/parabolic system. Semigroup Forum , 57(2):278–292, 1998.\n[8] S. P. Chen and R. Triggiani. Proof of extensions of two con jectures on structural damping\nfor elastic systems. Pacific J. Math. , 136(1):15–55, 1989.\n[9] R. Denk and R. Schnaubelt. A structurally damped plate eq uation with Dirichlet–Neumann\nboundary conditions. Journal of Differential Equations , 259(4):1323–1353, 2015.\n[10] H. D. Fern´ andez Sare and J. E. Mu˜ noz Rivera. Analytici ty of transmission problem to ther-\nmoelastic plates. Quart. Appl. Math. , 69(1):1–13, 2011.\n[11] F. Hassine. Logarithmic stabilization of the Euler-Be rnoulli transmission plate equation with\nlocally distributed Kelvin-Voigt damping. arXiv preprint arXiv:1403.0356 , 2014.\n[12] R. Leis. Initial-boundary value problems in mathematical physics . B. G. Teubner, Stuttgart;\nJohn Wiley & Sons, Ltd., Chichester, 1986.\n[13] W. Liu and G. H. Williams. Exact controllability for pro blems of transmission of the plate\nequation with lower-order terms. Quart. Appl. Math. , 58(1):37–68, 2000.\n[14] S. Mansouri. Boundary stabilization of coupled plate e quations. Palest. J. Math. , 2(2):233–\n242, 2013.\n[15] E. Mitidieri. A Rellich type identity and applications .Comm. Partial Differential Equations ,\n18(1-2):125–151, 1993.\n[16] J. E. Mu˜ noz Rivera and H. Portillo Oquendo. A transmiss ion problem for thermoelastic\nplates.Quart. Appl. Math. , 62(2):273–293, 2004.\n[17] M. I. Mustafa and G. A. Abusharkh. Plate equations with v iscoelastic boundary damping.\nIndag. Math. (N.S.) , 26(2):307–323, 2015.\n[18] A. Pazy. Semigroups of linear operators and applications to partial differential equations ,\nvolume 44. Springer Science & Business Media, 2012.\n[19] J. Pr¨ uss. On the spectrum of C0-semigroups. Trans. Amer. Math. Soc. , 284(2):847–857, 1984.\n[20] D. L. Russell. Mathematical models for the elastic beam and their control-theoretic implica-\ntions. In Semigroups, theory and applications, Vol. II (Trieste, 198 4), volume 152 of Pitman\nRes. Notes Math. Ser. , pages 177–216. Longman Sci. Tech., Harlow, 1986.\n[21] H. Triebel. Interpolation Theory, Function Spaces, Differential Opera tors. North Holland,\n1978.\n[22] J. C. Vila Bravo and J. E. Mu˜ noz Rivera. The transmissio n problem to thermoelastic plate\nof hyperbolic type. IMA J. Appl. Math. , 74(6):950–962, 2009.EXPONENTIAL STABILITY FOR A COUPLED SYSTEM 17\nUniversit ¨at Konstanz, Fachbereich f ¨ur Mathematik und Statistik, 78457 Konstanz,\nGermany\nE-mail address :robert.denk@uni-konstanz.de\nUniversit ¨at Konstanz, Fachbereich f ¨ur Mathematik und Statistik, 78457 Konstanz,\nGermany\nE-mail address :felix.kammerlander@uni-konstanz.de" }, { "title": "2010.05614v2.Decays_rates_for_Kelvin_Voigt_damped_wave_equations_II__the_geometric_control_condition.pdf", "content": "DECAY RATES FOR KELVIN-VOIGT DAMPED WAVE EQUATIONS II: THE\nGEOMETRIC CONTROL CONDITION\nNICOLAS BURQ AND CHENMIN SUN\nAbstract. We study in this article decay rates for Kelvin-Voigt damped wave equations under a geometric\ncontrol condition. When the damping coe\u000ecient is su\u000eciently smooth ( C1vanishing nicely, see (1.3)) we show\nthat exponential decay follows from geometric control conditions (see [5, 12] for similar results under stronger\nassumptions on the damping function).\n1.Introduction\nIn this paper we investigate decay rates for Kelvin-Voigt damped wave equations under geometric control\nconditions. We work in a smooth bounded domain \n \u001aRdand consider the following equation\n(1.1)8\n><\n>:(@2\nt\u0000\u0001)u\u0000div(a(x)rx@tu) = 0\nujt=0=u02H1\n0(\n); @tujt=0=u12L2(\n)\nuj@\n= 0\nwith a non negative damping term a(x). The solution can be written as\n(1.2) U(t) =\u0012u\n@tu\u0013\n=eAt\u0012u0\nu1\u0013\n;\nwhere the generator Aof the semi-group is given by\nA=\u0012\n0 1\n\u0001 divar\u0013\u0012\nu0\nu1\u0013\n;\nwith domain\nD(A) =f(u0;u1)2H1\n0\u0002L2; \u0001u0+ divaru12L2;u12H1\n0g:\nThe energy of solutions\nE(u)(t) =Z\n\n(jrxuj2+j@tuj2)dx\nsatis\fes\nE((u0;u1))(t)\u0000E((u0;u1))(0) =\u0000Zt\n0Z\n\na(x)jrx@tuj2(s;x)ds:\nOur purpose here is to show that if the damping ais su\u000eciently smooth, the exponential decay rate holds,\ndropping some unnecessary assumptions on the behaviour of the damping term where it becomes positive in\nprevious works [5]. Namely we shall assume a(x)>0 isC1(\n) and satisfy the regularity hypothesis\njraj6Ca1\n2: (1.3)\nOur main result is\nTheorem 1. Assume that \nis a compact Riemannian manifold with smooth boundary. Let a2C1(\n)be a\nnonnegative function satisfying (1.3) , such that the following geometric control condition is satis\fed:\n\u000fThere exists \u000e > 0such that all rays of geometric optics (straight lines) re\recting on the boundary\naccording to the laws of geometric optics eventually reach the set !\u000e=fx2\n :a(x)>\u000egin \fnite time.\n1arXiv:2010.05614v2 [math.AP] 19 Mar 20212 N. BURQ AND C-M. SUN\nThen there exists \u000b>0, such that for all t>0and every (u0;u1)2H1\n0(\n)\u0002L2(\n), the energy of solution u(t)\nof(1.1) with initial data (u0;u1)satis\fes\nE[u](t)6e\u0000\u000btE[u](0):\nTo prove this result, we \frst reduce it very classicaly in Section 2 to resolvent estimates. Since the low\nfrequency estimates are true, we are reduced to the high frequency regime. The proof relies on resolvent\nestimates which are proved through a contradiction argument that we establish in Section 2. In Section 3 we\nprove a priori estimates for our sequences. The main task then is to prove a propagation invariance for these\nmeasures. A main di\u000eculty to overcome is that it is not possible to put the damping term in the r.h.s. of the\nequation (1.1) and treat it as a perturbation . Instead we have to keep it on the left hand side and revisit the\nproof of the propagation property from [7]. In Section 4, we introduce the geometric tools necessary to tackle the\nboundary value problem and de\fne semi-classical measures associated to our sequences. In Section 5 we prove\nthe interior propagation result for our measures. Finally, in Section 6, we \fnish the proof of the contradiction\nargument by establishing the invariance of the semi-classical measures we de\fned up to the boundary. Here the\nproof uses crucially the main result in [7, Th\u0013 eor\u0012 eme 1].\nRemark 1.1. Throughout this note, we shall prove that some operators of the type P\u0000\u0015Id,\u00152R(resp.\n\u00152iR) are invertible with estimates on the inverse. All these operators share the feature that they have\ncompact resolvent, i.e. 9z02C; (P\u0000z0)\u00001exists and is compact (or it will be possible to reduce the question\nto this situation). As a consequence, since\n(P\u0000\u0015) = (P\u0000z0)\u00001(Id + (z0\u0000\u0015))\u00001);\nand (Id + (z0\u0000\u0015)\u00001) is Fredholm with index 0, to show that ( P\u0000\u0015) is invertible with inverse bounded in norm\nbyA, it is enough to bound the solutions of ( P\u0000\u0015)u=fand prove\n(P\u0000\u0015)u=f)kukL26AkfkL2:\nRemark 1.2. Assume that ais the restriction to \n of a nonnegative C2(Rd) function. Then the hypothesis (1.3)\nis satis\fed.\nProof. It is enough to prove (1.3) for \n = Rd,a2C2(\n). Letx02Rdand denote by z0=ra(x0) From\nTaylor's formula, we have for any s2R, there exists \u00122(0;1), such that\na(x0+sz0) =a(x0) +sjz0j2+s2\n2a00(x0+\u0012sz0)(z0;z0)>0\nSince this polynomial in sis non negative, we deduce tat its discriminant is non positive\njz0j4\u00002ka00k1jz0j2a(z0)60)jrxa(x0)jj262ka00k1a(z0):\nNotice that in the above lemma, the condition cannot be relaxed to a2C2(\n);a>0. Indeed, consider the\nfollowing example: \n = B(0;1) anda(x) = 1\u0000jxj2forjxj61. Then obviously a2C2(\n),a>0 , but on the\nboundary,rxa6= 0, whilea= 0. \u0003\nAcknowledgment. The \frst author is supported by Institut Universitaire de France and ANR grant ISDEEC,\nANR-16-CE40-0013. The second author is supported by the postdoc programe: \\Initiative d'Excellence Paris\nSeine\" of CY Cergy-Paris Universit\u0013 e and ANR grant ODA (ANR-18-CE40- 0020-01).\n2.Contradiction argument\nIt is well known that decay estimates for the evolution semi-group follow from resolvent estimates [1, 2, 4].\nHere we shall need the classical (see e.g. [6, Proposition A.1])\nTheorem 2. The exponential decay of the Kelvin Voigt semi-group is equivalent to the following resolvent\nestimate: There exists Csuch that for all \u00152R, the operator (A\u0000i\u0015)is invertible from D(A)toHand its\ninverse satis\fes\n(2.1) k(A\u0000i\u0015)\u00001kL(H)6CKELVIN-VOIGT DAMPING 3\nLet us \frst recall that\n(2.2) ( A\u0000i\u0015)\u0012u\nv\u0013\n=\u0012f\ng\u0013\n,\u001a\u0000i\u0015u+v=f\n\u0001u+ divarxv\u0000i\u0015v=g\nFrom [8, Section 2], we have the following low frequencies estimates of the resolvent of the operator A:\nProposition 2.1. Assume that a2L1is non negative a>0and non trivialR\n\na(x)dx > 0. Then for any\nM > 0, there exists C > 0such that for all \u00152R;j\u0015j6M, the operatorA\u0000i\u0015is invertible from D(A)toH\nwith estimate\n(2.3) k(A\u0000i\u0015)\u00001kL(H)6C:\nAs a consequence, to prove Theorem 1 it is enough to study the high frequency regime \u0015!+1and prove\nProposition 2.2. Assume that a2C1(\n)is a nonnegative function satisfying (1.3) . Then under the geometric\ncontrol condition, there exists \u00030>0such that for any j\u0015j>\u00030we have\nk(A\u0000i\u0015)\u00001kL(H)6C:\nBy standard argument, we can reduce the proof of Proposition 2.2 to a semi-classical estimate. We denote\nby 0 0, such that for all 0